content
stringlengths 1
15.9M
|
---|
\section{Introduction}\label{sec:intruction}
Irradiation introduces mobile defects such as self-interstitial atom (SIA) clusters and vacancy clusters in metals. These defects will interact with each other and with the pre-existing microstructure, mainly dislocations, thereby inducing nanostructural changes that will affect the mechanical properties of the material. To fully understand the evolution of the defect populations over time, the rates of these reactions needs to be correctly assessed.
Object Kinetic Monte Carlo (OKMC) is a stochastic simulation method, where the dynamic behaviour of all objects, such as SIA or vacancy clusters, is described by pre-defined probabilities. It is a well-suited technique for simulation of the evolutions of radiation induced defects in iron alloys (See \textit{e.g.} \cite{jansson2013simulation}). The OKMC has been shown to be equivalent to rate theory calculations \cite{stoller2008mean}. It has the advantage of going beyond the mean-field approximation and taking explicitly all spatial correlations, except the elastic interactions, into account.
In mean-field approaches, the rate at which a mobile defect interacts with a cluster or dislocation of a given shape and size, acting as a sink, is given by the sink strength, $k^2$. This is proportional to the inverse square of the average distance covered by the defects before the interaction, which is normally absorption or clustering. The sink strength is \textit{a priori} affected by the shape and size of the sinks, their number density, their type and orientation. Also, the migration regime of the defects will have an impact: defects that migrate in a 1D fashion are less likely to interact with sinks than defects that migrate in 3D or in a fashion between fully 1D and 3D.
Analytical expressions for different sink shapes, such as spheres, toroids and dislocations have been theoretically obtained in the case of 3D migrating defects and a number of them is reviewed by \textit{e.g.} F.A. Nichols in \cite{nichols1978estimation}. Barashev \textit{et al}. derived expressions for fully 1D migrating defects in the case of spherical absorbers, dislocations and grain boundaries \cite{barashev2001reaction}. The 3D to 1D transition has been studied by Trinkaus \textit{et. al.}, who also proposed a master curve for the transition \cite{trinkaus20021d,trinkaus2004reaction}. Malerba \textit{et. al.} \cite{malerba2007object} showed that OKMC calculations of the sink strength for spherical sinks and grain boundaries show good agreement with analytical expressions used in rate theory and that also the 3D to 1D defect migration regime transition with spherical absorbers can be reproduced using OKMC. These results simultaneously corroborate the theory and show the equivalence between OKMC and
mean-field approaches.
In this work, we extend the study by Malerba \textit{et. al.} \cite{malerba2007object}, where spherical absorbers were considered, to also calculate the sink strength for dislocation lines and toroidal absorbers, the latter corresponding to dislocation loops, and compare the results with analytical expressions available from rate theory. The structure of this paper is as follows: In section \ref{sec:methods}, we describe our methodology. In Sec. \ref{sec:dislocations} we compare the sink strength of dislocations obtained by OKMC to rate theory expressions, in the limits of 3D and 1D migration, and in Sec. \ref{sec:loops}, we do the same with toroidal absorbers. Finally, in Sec. \ref{sec:transition}, we study the transition of the migration regime from 3D to 1D of defects absorbed by dislocations and toroidal sinks and compare with the theoretical master curve. The discussion and conclusions are found in Sec. \ref{sec:discussion} and \ref{sec:conclusions}, respectively.
\section{Computation method}\label{sec:methods}
We have estimated the sink strength of straight dislocations and dislocation loops using the same methods as in \cite{malerba2007object} (where spherical absorbers were considered), which is here briefly recalled. For our calculation we use the OKMC code LAKIMOCA \cite{domain2004simulation}. The probabilities for migration jumps of defects in the simulations are given in terms of Arrhenius frequencies for thermally activated events, $\Gamma_i = \nu_i\exp \left(\frac{-E_{a,i}}{k_B T}\right)$, where $\nu_i$ is the attempt frequency, $E_{a,i}$ the activation energy for the process, $k_B$ Boltzmann's constant and $T$ the temperature. Events are randomly chosen according to their probability, following the Monte Carlo algorithm \cite{metropolis1953equation}. The simulated time is increased according to the resident time algorithm \cite{young1966monte} with $\Delta t = 1/\left(\sum_{i=1}^{N_{int}} \Gamma_i + \sum_{j=1}^{N_{ext}}P_j\right)$, where $N_{int}$ is the number of internal events such as defect jumps and
$N_{ext}$ the number of external events, such as cascades or Frenkel pair creation, with $P_j$ being the probabilities for the external events. In the long term, this equation substitutes $\Delta t' = -\ln{u}\Delta t$ , which is fully exact by including the stochasticity due to the Poisson distribution \cite{bortz1975new}. $u$ is here a uniform random number between 1 and 0 .
Objects, such as defects or clusters, in the model are described as geometrical objects, such as spheres, toroids or cylinders. Reactions between objects take place when these overlap geometrically. Reactions could be annihilations or clustering. In this study, the only reaction considered is absorption, where the reacting defect will not change the volume of the absorbing objects.
The straight dislocation was simulated as an immobile cylinder-shaped sink whose two opposite faces touch the faces of a non-cubic simulation box. No defect can ever impinge on the cylinder from one of the faces, as no defect is allowed to be outside the box, so this is an effective way to simulate an infinitely long straight dislocation. However, since periodic boundary conditions are applied in all direction, we are in practice simulating a regular array of infinitely long straight dislocations: this must be taken into account to rationalize the results and certainly when choosing the theoretical expression to which the simulation results are to be compared. (On the other hand, as discussed by Brailsford and Bullough \cite{brailsford1981theory}, there is no real theoretical expression for the sink strength of an array of dislocations significantly different from a lattice and for real random arrays the $Z$ factor of proportionality with the dislocation density should be regarded as an empirical parameter.)
The simulation box with the dislocation is pictorially represented in Fig \ref{1.jpg}. Different dislocation densities and capture radii, $r_d$, (as in Fig \ref{1.jpg}) were explored. With reference to Fig. \ref{1.jpg}, the dislocation density, $\rho_d$, was changed by varying $l_y$ and $l_z$: $\rho_d = (l_y l_z)^{-1}$, for a fixed $l_x=100a_0$ (since iron is taken as materials of reference, the underlying lattice is body-centred-cubic (bcc) and the lattice parameter is $a_0=2.87$ Å), with $l_x \neq l_y\neq l_z$. Dislocation densities were varied between $10^{14}$ and $10^{15}$ m$^{-2}$, capture radii, $r_d$, between 2 and 9 nm (smaller densities, down to $5\cdot10^{13}$ m$^{-2}$, were also considered in the 3D limit only; simulations with smaller radii, 0.5 and 1 nm, for $\rho_d=10^{14}$ m$^{-2}$ did not provide sufficient statistics to be fully acceptable in the 1D limit). The simulation temperature was arbitrarily set to 573 K, but it does not have any influence on the sink strength calculation. .
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{okmc_dislocation.pdf}
\caption{Upper side: Pictorial representation of the non-cubic simulation box and the cylinder mimicking the sink effect of a straight dislocation. Lower side: view normal to $x$ axis of the simulation box and image boxes corresponding to applying periodic boundary conditions: the image dislocations effectively create a regular array.}
\label{1.jpg}
\end{figure}
Dislocation loops are simulated by immobile sinks of toroidal shape, depicted in Fig. \ref{fig:torus}, and characterized by the major radius, $R$, the minor radius $r_t$ and their orientation with four possible [111] Burgers vectors. Toroidal sinks with the same $R$ and $r$ were randomly distributed in the simulation box with random orientations for a given number density. The simulation box size used for the loop sink strength calculations were $300\times350\times400a_0^3$, except in Sec. \ref{sec:orientation} and Sec. \ref{sec:transition}, were smaller box sizes of $150\times200\times250a_0^3$ and $250\times300\times350a_0^3$, respectively, were used in the loop cases. A simulation temperature of 373 K was arbitrarily used for all loop cases.
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{torus_sage.pdf}
\caption{Depiction of a toroid with the major radius $R$ and the minor radius $r_t$.}
\label{fig:torus}
\end{figure}
The sink strength is obtained as \cite{malerba2007object}
\begin{equation}\label{eq:sink_strength}
k^2 = \frac{2n}{d^2_j\langle n_j\rangle},
\end{equation}
where $\langle n_j \rangle$ is the average number of jumps performed by the defect, introduced one by one in the simulation. $n$ is the dimensionality of the motion and $d_j$ is the jump distance, defined in the bcc lattice as the first nearest neighbour distance,
\begin{equation}
d_j = \frac{\sqrt{3}}{2}a_0.
\end{equation}
In the simulations, 3D to 1D migrating defects were introduced at a random position one by one.
The sink strength is then calculated using the average number of jumps needed to be absorbed by a sink, $\langle n_j\rangle$, and Eq. \eqref{eq:sink_strength}. The defects have a spherical capture radius of $r_0=2.5$ Å. The attempt frequency used is $8.071\cdot10^{13}$ s$^{-1}$ and the migration energy 0.31 eV, corresponding to a single SIA \cite{jansson2013simulation,anento2010atomistic,takaki1983resistivity}. Clustering of defects was explicitly forbidden.
For the dislocation case, the simulation was stopped when one of the two following conditions was fulfilled: (a) 30000 defects had been introduced and disappeared at the sinks (in practice, somewhat less than this, because the defects that happened to be created inside the sink are disregarded); (b) the sum of the number of jumps taken by all the defects introduced in the box equals $10^{15}$. Generally the first condition dictated the end of the simulation, but for small volume fractions in the 1D case sometimes the second one overruled. For the loop case, the simulation was stopped when 10000 defects had been introduced and absorbed by the sinks.
The 3D to 1D regime transition was explored by increasing the number of 1D jumps after which the defect changes direction from 1 (effective 3D migration) to $10^{10}$--$10^{11}$ (effective 1D migration). In addition for the straight dislocation case, the 1D limit was also simulated by requiring an energy of change of direction of 2 eV at 573 K, which corresponds in practice to no change of direction in the course of the whole simulation, at the temperatures considered (the effectiveness of this way of operating to reach the 1D limit, using a non-cubic box, is demonstrated and discussed in \cite{malerba2007object}).
\section{Results}\label{sec:results}
\subsection{Straight dislocations}\label{sec:dislocations}
\subsubsection{3D limit for dislocations}
Fig. \ref{2.jpg} shows the ``cloud'' of sink strength points hitherto obtained from the simulation in the 3D limit, in the above-specified ranges of capture radii and dislocation densities, plotted versus volume fraction. Patterns according to which the data points can be grouped are clearly recognisable: they correspond to the same capture radius (indicated by the same colour) or the same dislocation density. On the same figure the points calculated for the same dislocation density and capture radius using different approximations, as found in \cite{nichols1978estimation} (\textit{Cf.} Table \ref{table:dislocation_eq}), are indicated. All the theoretical expressions are for a regular array of parallel dislocations. It can be seen that, while for small volume fractions all expressions are acceptably valid, for large volume fractions only the expression deduced by Wiedersich (\textit{Cf.} Table \ref{table:dislocation_eq})
remains valid and in agreement with the simulation results. The expressions obtained by solving either Laplace or Poisson equations underestimate the sink strength, while the approximated expression proposed by Nichols in \cite{nichols1978estimation} (\textit{Cf.} Table \ref{table:dislocation_eq}) exhibits a singularity in the range of volume fractions of interest and starts diverging above a certain volume fraction.
\begin{table}
\centering
\caption{Analytical expressions for the sink strength of dislocation lines. Here $\rho = r_d\sqrt{\pi\rho_d}$ \cite{nichols1978estimation}.}
\label{table:dislocation_eq}
\begin{tabular*}{\columnwidth}{@{\extracolsep{\fill}} l c p{0.7\columnwidth}}
\toprule
& $k^2_{d,3}$\\
\midrule
\vspace{0.5cm} Wiedersich \cite{wiedersich1972theory} &
$\frac{2\pi\rho_d(1-\rho^2)}{\ln\left(\frac{1}{\rho}\right) - \frac{3}{4} + \frac{1}{4}\rho^2(4-\rho^2)}$\\
\vspace{0.5cm} Laplace &
$\frac{2\pi\rho_d}{\ln\left(\frac{1}{\rho}\right)}$\\
\vspace{0.5cm} Poisson &
$\frac{2\pi\rho_d}{\ln\left(\frac{1}{\rho}\right)-\frac{1}{2}+\frac{1}{2}\rho^2}$\\
Nichols &
$\frac{2\pi\rho_d}{\ln\left(\frac{1}{\rho}\right)-\frac{3}{4}}$\\
\bottomrule
\end{tabular*}
\end{table}
In Fig. \ref{3.jpg} the comparison between OKMC simulation and different theoretical predictions is made directly on a 45\textdegree{} straight line. As can be seen, in the case of the Wiedersich expression (\textit{Cf.} Table \ref{table:dislocation_eq}), the points fall perfectly on the line, so we can state that there is perfect correspondence between simulation and theory.
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{lorenzo_dislocation_3d_fv.pdf}
\caption{OKMC dislocation sink strengths in the 3D limit versus volume fraction. The capture radius, $r_d$, is varied between 1.25 and 9.25 nm. The dislocation densities are specified in the text. Points from different theoretical expressions are superposed with the corresponding colours. Only the expression of Wiedersich (\textit{Cf.} Table \ref{table:dislocation_eq}) matches all simulation points.}
\label{2.jpg}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{lorenzo_dislocation_3d_comp.pdf}
\caption{Comparison between simulation values and different theoretical values for the sink strength of a regular array of parallel dislocations in the 3D limit.}
\label{3.jpg}
\end{figure}
\subsubsection{1D limit for dislocations}
Figs \ref{4.jpg} and \ref{5.jpg} are the equivalent to Figs. \ref{2.jpg} and \ref{3.jpg}, comparing the simulation results with theory using the ``cloud'' and the 45\textdegree{} line representation, respectively, for 1D migrating defects. The only theoretical expression used in this case for comparison is the one proposed by A. Barashev \textit{et al.} in \cite{barashev2001reaction}:
\begin{equation}\label{eq:barashev}
k_{d,1}^2 = 3\cdot2(\pi r_d \rho^\ast)^2,
\end{equation}
where the factor 3 (which does not appear in the mentioned reference) comes from the fact of using here the 3D diffusion coefficient (in order to consistently trace the transition between regimes) and $\rho^\ast$ is defined as ``the mean number of dislocations lines intersecting a unit area (surface density)''. In the present case of regular array of parallel dislocations, $\rho^\ast=\rho_d$: there is indeed only one surface crossed by dislocations and the surface density is (with reference to Fig. \ref{1.jpg}) $\rho^\ast = (l_y l_z)^{-1}$, which is coincident with the dislocation density. This equality has been therefore used to apply Eq.. \eqref{eq:barashev} and to compare it with the simulation results in the figures. It can be seen that the comparison is acceptable, particularly
in Fig. \ref{5.jpg}, although it is certainly not as good as in the 3D case. At the same time, in Fig. \ref{4.jpg} the patterns according to which the data points should be grouped by capture radii or dislocation densities are less easy to spot, a sign of larger scatter. We shall address later on the problem of establishing where this scatter and the less good agreement between theory and simulation in this 1D case may come from.
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{lorenzo_dislocation_1d_fv.pdf}
\caption{OKMC dislocation sink strength in the 1D limit versus volume fraction in the range of capture radii, $r_d$, and dislocation densities specified in the text, compared with the theoretical points obtained from Eq. \eqref{eq:barashev}.}
\label{4.jpg}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{lorenzo_dislocation_1d_comp.pdf}
\caption{Comparison between theory, Eq. \eqref{eq:barashev}, and simulation values for the sink strength of a regular array of parallel dislocations in the 1D limit. The colours correspond to different capture radii, $r_d$ = 0.75--9.26 nm, and are the same as in Fig. \ref{4.jpg}.}
\label{5.jpg}
\end{figure}
\subsection{Loops}\label{sec:loops}
If the rate of annihilation at sinks is defined as $Dck^2$, where $D$ is the 3D-diffusion coefficients, $c$ the defect concentration and $k^2$ the sink strength, the theoretical expression of sink strength for toroidal absorbers can be derived from the work of F.A. Nichols \cite{nichols1978estimation} as:
\begin{equation}\label{eq:theory_torids}
k^2_{t,3} = \frac{4 \pi^2 n (R^2 -r_t^2)^{1/2} }{ \ln(8 R / r_t) },
\end{equation}
where $R$ is the major radius and $r_t$ the minor radius, as shown in Fig. \ref{fig:torus}, and $n$ is the number density of toroidal sinks: $R \gtrsim 3r_t$ is the condition for the applicability of Eq. \eqref{eq:theory_torids}.
For comparison, the sink strength of spherical absorbers with a radius $r_s$ and a density $n$ for 3D migrating defects is given by \cite{malerba2007object,brailsford1981theory,brailsford1979effect}:
\begin{equation}\label{eq:theory_spheres}
k^2_{s,3} = 4\pi n r_s \left(1+ r_s\sqrt{4\pi n r_s}\right).
\end{equation}
For 1D migrating defects, the sink strength is given by
\begin{equation}\label{eq:theory_spheres_1d}
k^2_{s,1} = 6(\pi r_s^2n)^2.
\end{equation}
In this section we will calculate the sink strength for a population of different number densities, $n$, different $r$ and $R$ values and compare with the theory for toroidal and spherical sinks.
\subsubsection{The dependence on the sink number density}
We fix the minor radius as $r_t = 5$ Å and use different major radii from $R = 3r_t$ to $5r_t$, as well as different sink densities. All SIA defects migrated fully in 3D. The results are shown in Fig.\ref{fig:P20100824-1.pdf} as a function of the number density and compared to the theoretical expression for toroidal absorbers, Eq. \eqref{eq:theory_torids}. In Fig. \ref{fig:P20100824-1_Vfrac.pdf}, the OKMC results are plotted as functions of the volume fraction and compared to the theory. It is clear that the theory is not valid for large volume fractions, although for densities of loops typically encountered in materials the theoretical expression is perfectly suitable. Interestingly, both theory and OKMC results collapse onto essentially the same curve, revealing that the volume fraction is the key variable, irrespective of the ratio between radii. Eq. \eqref{eq:theory_torids} can be indeed rewritten as
\begin{equation}
k^2_{t,3} = \frac{2f_V \sqrt{1-\left(\frac{r_t}{R}\right)^2}}{r_t^2\ln\left(\frac{8R}{r_t}\right)},
\end{equation}
where the volume fraction, $f_V$, is the dominant parameter, compared to $R$, as $r_t$ is constant.
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{P20100824-1.pdf}
\caption{Sink strength of loops for different ratios between radii when the major one increases, as a function of number density: both OKMC calculations and theoretical expressions collapse onto essentially the same curve.}
\label{fig:P20100824-1.pdf}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{P20100824-1_Vfrac.pdf}
\caption{The effect of the loop number density in terms of volume fraction and compared to the theory for toroidal sinks}
\label{fig:P20100824-1_Vfrac.pdf}
\end{figure}
\subsubsection{The dependence on the major radius}
We fixed the minor radius at $r_t = 5$ Å and the sink density at $n = 1$ appm ($n=8.46\cdot10^{22}$ m$^{-3}$) or 100 appm ($n=8.46\cdot10^{24}$ m$^{-3}$). The major radii are varied between $R=0.5r_t$ and $R=15r_t$. In Fig \ref{fig:P20100824-2.pdf}, the results are compared to the theory for sink strength of toroids, Eq. \eqref{eq:theory_torids} and spheres, Eq. \eqref{eq:theory_spheres}, where the spherical radius is calculated as $r_s = R+r_t$. In Fig. \ref{fig:P20100824-2_Vfrac.pdf}, the results are plotted versus volume fraction. The theoretical sink strength for toroids differs more at larger volume fractions. The sink strength is better described by the theory for spheres when $R\sim r$.
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{P20100824-2.pdf}
\caption{Sink strength of loops for two different number densities (distinguished by different colours), as a function of the ratio between radii when the major one increases: OKMC calculations are compared to theory for both loops and spheres.}
\label{fig:P20100824-2.pdf}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{P20100824-2_Vfrac.pdf}
\caption{Sink strength of loops for two different number densities (distinguished by different colours) and different ratios between radii, as a function of the volume fraction: OKMC calculations are compared to theory for both loops and spheres.}
\label{fig:P20100824-2_Vfrac.pdf}
\end{figure}
\subsubsection{The dependence on the minor radius}
We fixed the major radius at $R=12.5$ Å and the minor radius was varied from $r_t=5$ Å to 13 Å (with $r_t>R$, the toroids become increasingly spherical). The resulting sink strengths are shown for different sink densities in Fig. \ref{fig:P20100824-3.pdf} and in terms of volume fraction in Fig. \ref{fig:P20100824-3_Vfrac.pdf}. The sink strengths are again compared with the theory for toroidal and spherical sinks, Eqs. \eqref{eq:theory_torids} and \eqref{eq:theory_spheres}, respectively. The toroidal theory works best when $r_t$ is small and the spherical theory works best when $r_t$ is large, compared to $R$. The ratio of the radii are here in fact below the constraint of the toroidal theory, which requires $R \gtrsim 3r_t$. Nevertheless, good agreement is reached already at a ratio of 2.5 for the two lowest number densities.
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{P20100824-3.pdf}
\caption{Sink strength of loops for different number densities (distinguished by different colours), as a function of the ratio between radii when the minor one decreases. Dots are OKMC calculations, lines theoretical values for toroidal sinks, dashed lines theoretical values for spherical sinks.}
\label{fig:P20100824-3.pdf}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{P20100824-3_Vfrac.pdf}
\caption{Effect of the minor radius in terms of volume fraction. Dots are our OKMC calculations, lines theoretical values for toroidal sinks and dashed lines the theoretical values for spherical sinks. Different colours correspond to different number densities.}
\label{fig:P20100824-3_Vfrac.pdf}
\end{figure}
\subsubsection{1D limit for loops}\label{sec:orientation}
We studied the effect on the sink strength of loops when the loops were oriented randomly in four different orientations, compared to having all loops in the same orientation. We used $R=15$ Å, $r_t=5$ Å and $n=10^{24}$ m$^{-3}$. The SIA defects were made to migrate in 1D.
As a result, we got that the sink strength with all sink loops parallel is 0.000279159 nm$^{-2}$ and 0.000310821 nm$^{-2}$ when the loops are randomly oriented; a difference of about 11 \%, smaller than the difference with respect to theoretical expressions. To our knowledge, no theoretical expressions for the sink strength with toroidal sinks and 1D migrating defects exists, so we could not compare results in this case.
Fig. \ref{R20121211_1D_Vfrac.pdf} shows the results of the sink strength of loops for 1D migrating defects as a function of volume fraction for different ratios between radii: $R$ is varied between $0.5r_t$ and $3.5r_t$. Larger $R$ gives larger volume fraction. The loops were randomly oriented. As no analytical expression exist for toroids with 1D migrating defects, we compare the result with the analytical expression for spherical absorbers, Eq. \eqref{eq:theory_spheres_1d}, and with the theory for dislocations with 1D migrating defects. For dislocations, we use Eq. \eqref{eq:barashev} with $\rho^\ast = 1/2\rho_d$, which correspond to random orientation of dislocations \cite{barashev2001reaction} and should be a good approximation for dislocation loops. Using $\rho_d = 2n\pi R$ and $r_d = r_t$, we get from Eq. \eqref{eq:barashev} an approximative expression for the sink strength for loops with 1D migrating defects:
\begin{equation}\label{eq:toroids_1d}
k_{t,1}^2 = 6(n\pi^2 r_t R)^2
\end{equation}
(The 3D diffusion coefficient is again used; for 1D, divide by 3.) For spherical sinks, we calculate the spherical radius as $r_s = R+r_t$. For comparison, we also include simulation results with 3D migrating defects and the corresponding analytical expression for toroidal sinks, Eq. \eqref{eq:theory_torids}, and dislocations (The Wiedersich equation in Table \ref{table:dislocation_eq}). In the 1D case, we get good agreement with the theory for dislocations and also fair agreement using the theory for spheres. The dislocation comparison does, however, not work as well with 3D migrating defects and we only get fair agreement. The data is in this case, as expected, better described by the theory for toroidal sinks, Eq. \eqref{eq:theory_torids}. The migration regime plays a significantly larger role than the shape of the sinks.
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{R20121211_1D_Vfrac.pdf}
\caption{Effect of the major radius in terms of volume fraction for toroidal sinks with 1D migrating defects. $R$ is varied between $0.5r_t$ and $3.5r_t$. The results are compared to the analytical expression for dislocations, using Eq. \eqref{eq:toroids_1d}, and spherical absorbers with 1D migrating defects, Eq. \eqref{eq:theory_spheres_1d}. For comparison, we also include simulation results with 3D migrating defects and the corresponding analytical expression for toroids, Eq. \eqref{eq:theory_torids}, and dislocations (Wiedersich from Table \ref{table:dislocation_eq}).}
\label{R20121211_1D_Vfrac.pdf}
\end{figure}
\subsection{The transition from 3D to 1D migration regimes}\label{sec:transition}
The sink strength of an absorber of any shape will depend on the migration regime of the defects it absorbs. Purely 3D migrating defects are more easily absorbed than purely 1D migrating defects and intermediate regimes correspond to intermediate sink strengths, as seen for spherical absorbers in \cite{malerba2007object}. The transition between 1D and 3D migration of the defects has been theoretically investigated by Trinkaus \textit{et al.} in \cite{trinkaus2004reaction}. They proposed a simple master curve which has been shown to agree very well with OKMC simulations of spherical absorbers \cite{malerba2007object}. This master curve depends on two variables $x$ and $y$, defined as
\begin{align}
x^2 &= \frac{
\frac{\delta f^2(\delta) l_{ch}^2 k_3^4}{12k_1^2} + 1}
{\frac{l_{ch}^2 k_1^4}{12k_1^2} +1}
\left( \frac{k_3^2}{k_1^2}\right)
\left(\frac{k_3^2}{k_1^2}-1\right) , \label{eq:dislo_x}\\
y &= \frac{k^2}{k_1^2},\label{eq:dislo_y}
\end{align}
where $k_3^2$ is the sink strength in the 3D limit, $k_1^2$ the sink strength in the 1D limit, $k^2$ the sink strength for a given $l_{ch}=d_j\sqrt{n_{ch}}$, the distance travelled in 1D before change of direction, with $n_{ch}$ being the number of jumps before change of direction. $\delta = D_{tr}/D_{lo}$ is the ratio between transversal and longitudinal diffusion; here $\delta f^2(\delta) \sim 0$, as transversal motion is not accounted for in our simulations. All sink strengths in the equation refer of course to the same choice of capture radius and dislocation density. The master curve may then be expressed as \cite{trinkaus2004reaction}
\begin{equation}\label{eq:dislo_master}
y = \frac{1}{2}\left(1+\sqrt{1+4x^2}\right)
\end{equation}
Note that in this definition, the master curve differs slightly from the older and less general definition in \cite{trinkaus20021d}, which was used for the spherical absorbers in \cite{malerba2007object}. That master curve is equivalent to
\begin{equation}
y'=\frac{1}{2}\left(1+\sqrt{1+\frac{4}{x^2}\left(1-\frac{k_1^2}{k_3^2}\right)}\right)
\end{equation}
with the definition of $x^2$ as in Eq. \eqref{eq:dislo_x}. The difference between how the simulation data satisfy either master curve will only depend on the factor $(1-k_1^2/k_3^2)$, which is essentially equal to 1.
\subsubsection{Dislocations}
In Fig. \ref{6.jpg} the 3D to 1D transition of the sink strength of dislocations is shown for a few capture radii (from 2.25 to 6.25 nm), for a dislocation density of $10^{14}$ or $10^{15}$ m$^{-2}$, as a function of the length before change of direction, $l_{ch}$ (more curves could be shown, but they would not add anything qualitatively different). If this figure is compared with Fig. 5 in the published paper on spherical absorbers \cite{malerba2007object}, it can be clearly seen that the transition from 3D to 1D regime seems to occur, in the case of the dislocation, in a much more abrupt way: for a length before change of direction of $\sim$250 nm the 1D regime is already reached, whereas in the case of spherical absorbers a length of at least one order of magnitude larger was needed.
This more abrupt transition is reflected in the corresponding master curve representation, given in Fig. \ref{7.jpg} for all the conditions (capture radii and dislocation densities) hitherto studied. In this figure, the simulation results were used to calculate $x$ and $y$ as defined by Eqs. \eqref{eq:dislo_x} and \eqref{eq:dislo_y}, respectively.
It can be seen that, as much as we can say based on the data points we have collected, the master curve expressed as in Eq. \eqref{eq:dislo_master} is reproduced by the data points only in the 1D and 3D limit regions, with, in addition, a tendency to underestimate $y$ in the 3D region, while, again, the 3D to 1D transition is more abrupt than the theoretical master curve predicts.
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{lorenzo_dislocation_lch.pdf}
\caption{3D to 1D transition as a function of length before change of direction for a few dislocation capture radii (2.25 to 6.25 nm), with a dislocation line density of $10^{15}$ m$^{-2}$}
\label{6.jpg}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{lorenzo_dislocation_master_mmm2_df0.pdf}
\caption{3D to 1D transition in master curve representation for dislocation lines: simulation data points elaborated according to Eqs. \eqref{eq:dislo_x} and \eqref{eq:dislo_y} and the master curve as given in \eqref{eq:dislo_master}. Different dislocation densities [m$^{-2}$] and $r_d$ [nm] have been used. Only representative cases are labelled.}
\label{7.jpg}
\end{figure}
\subsubsection{Loops}
In Fig. \ref{fig:R20121211.pdf}, the 3D to 1D transition for toroidal absorbers is plotted for different major radii, $R$, \textit{i.e.} different volume fractions, as a function of the distance before change of direction, $l_{ch}$. The sink number density was $n=8.38\cdot10^{23}$ m$^{-3}$. The 1D limit was reached slightly faster for larger $R$ (higher volume fractions), possibly because the loop in this case tends to resemble more and more a dislocation line. With $l_{ch} = 10^{-5}$ m, all cases have reached the 1D limit.
In Fig. \ref{fig:R20121211_master.pdf}, the same data is plotted in the master curve representation and compared to the theory, Eq. \eqref{eq:dislo_master}. The simulation data shows in this case excellent agreement with the theory. The trends are similar to the ones seen for spherical absorbers in \cite{malerba2007object} and thus also less abrupt transition than for dislocations.
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{R20121211.pdf}
\caption{The transition from 3D to 1D migration regimes for different toroidal major radii.}
\label{fig:R20121211.pdf}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{R20121211_master.pdf}
\caption{The master curve for different toroidal major radii, compared to the master curve, Eq. \eqref{eq:dislo_master}.}
\label{fig:R20121211_master.pdf}
\end{figure}
\section{Discussion}\label{sec:discussion}
\subsection{Dislocation 1D limit}
The 45\textdegree{} plot of Fig. \ref{5.jpg} comparing theory and simulation in the 1D limit appears not to be as satisfactory as the same plot for the 3D case (Fig. \ref{3.jpg}), but from there it is difficult to identify any systematic deviation. At first sight, this less good agreement could be attributed to lack of statistics, knowing that it is more difficult to have proper statistics in the 1D case. After closer inspection, however, it appears that the deviation is probably systematic. Namely, for low volume fractions the simulation values tend to be smaller than the theoretical one, while they are larger for high volume fractions. This can be seen in Fig. \ref{8.jpg}, where the percentual error
\begin{equation}
e(\%) = 100\cdot \frac{k_{simul.}^2-k_{theory}^2}{k_{theory}^2}
\end{equation}
versus sink volume fraction, $f_V$, and the sum of the errors up to the given volume fraction, are plotted. We see that $e(\%)$ is tendentially negative for low $f_V$ and positive for high $f_V$. A sort of critical $f_V$ value, at which the trend is reversed can be identified around 0.02.
\begin{figure}
\centering
\includegraphics[width=\columnwidth]{lorenzo_dislocation_1d_error.pdf}
\caption{Percentual error committed by taking the simulation result for the sink strength of dislocations in the 1D case, as compared to the reference, chosen to be the theoretical value from Eq. \eqref{eq:barashev}.}
\label{8.jpg}
\end{figure}
Something similar, though much less spectacular, can be observed, with hindsight, also in the case of the spherical absorbers (see \cite{malerba2007object}). However, in that case the simulation always gave a higher value than the theory and this could be understood in terms of cutting out the tail of very long distances before absorption, given that the time in the simulation is limited and the statistics not very good. This is particularly true for small volume fractions. Since discrepancies were observed also for large volume fractions, however, we might tentatively explain it, in the case of spherical absorbers, with the fact that the simulation may include some degree of order in the distribution of absorbers, as a consequence of the periodic boundary conditions which is of course not included in the theoretical expression, that is strictly valid for a random distribution of absorbers.
In the present case, \textit{i.e.} dislocation lines, however, for low volume fractions the simulation provides smaller sink strength than the theory, so the statistical explanation (elimination of tails of small values) does not hold any more. Indeed in all cases the number of absorbed defects was well in excess of $10^4$, thereby guaranteeing that the values are converged. In addition, in this case we positively know that the simulation is done in an ordered array of dislocations. At the same time, in the theoretical expression this is reflected in the choice of $\rho^\ast = \rho_d$. We therefore surmise that the theoretical expression might in this case need to be completed with higher order terms that slightly modulate the expression as given in Eq. \eqref{eq:barashev}.
\subsection{Loops}
The sink strengths, calculated for toroidal sinks, are in good agreement with the theoretical expression for toroidal sinks, Eq. \eqref{eq:theory_torids}, when the volume fractions are low enough, $f_V\lesssim10^{-3}$, and the minor ($r_t$) and major ($R$) toroidal radii are not too small or large, compared to each other. Indeed, we get good agreement when $R/r_t>2$ at small volume fractions. When the $r_t$ and $R$ are of comparable sizes ($R/r<2$), the sink strengths are better described by the theory for spherical absorbers, Eq. \eqref{eq:theory_spheres}.
The orientation of the toroidal sinks makes a difference if the defects are migrating in 1D. With 3D migrating defects, the orientations play an insignificant role. Experimentally, we can expect the loops to be oriented randomly, perhaps with the exception of mechanically strained materials.
The sink strength for toroidal sinks with 1D migrating defects can be well described at low densities by the theory for dislocations in random orientations, Eq. \eqref{eq:barashev}. The agreement is only fair with 3D migrating defects and it also remains an open question what happens at higher densities.
\subsection{The master curve}
The above-discussed discrepancy for dislocations does not influence the master curve representation, since simulation data in it are compared to other simulation data in order to define $x$ and $y$ (Eqs. \eqref{eq:dislo_x} and \eqref{eq:dislo_y}), so for example in the 1D limit the ratio that gives $y$ (Eq. \eqref{eq:dislo_y}) will tend to unity in any case. The master curve coming out of the simulation, however, has a somewhat different shape from the theoretical one, related to the more abrupt 3D to 1D transition than in the spherical absorber case. In this case the fact that the dislocations in the simulation form an ordered array might play a role, as the master curve expression does not include any variable that takes explicitly into account the presence of order; thus the discrepancy might simply be due to this and to the need to compare to a master curve obtained for a regular array of parallel dislocations.
For the toroidal absorbers, the agreement between simulations and theory is excellent and the values clearly reach the 1D limit, as also was the case for dislocations. The transition also follows the master curve perfectly. The toroidal shapes, as compared to spherical shape, thus, do not alter the validity of the master curve.
\section{Conclusions}\label{sec:conclusions}
There is excellent agreement for both dislocations and toroidal absorbers in the 3D limit between sink strengths estimated statistically in an OKMC simulation and sink strengths obtained using the theoretical expressions available from the literature.
There is, however, only fair agreement for dislocations in the 1D limit between simulation and theory, the reasons of the discrepancy being not the lack of statistics or the inadequacy of the simulation box size. The actual origin of the discrepancy is not established.
The sink strength of toroids approaches the theoretical prediction for spherical absorbers when $r_t \rightarrow R$. Good agreement with theory is reached when the minor radius $r_t$ is small, the major radius between $2r_t$ and $15r_t$, and the volume fraction is low. With 1D migrating defects, the sink strength of toroids is well described at low densities by theory if the loops are considered as dislocations with random orientations.
The master curve is reproduced with excellent agreement using toroidal absorbers and to a good extent also correctly reproduced by the simulation data for the regular array of dislocations. However, the transition between 3D and 1D regime is clearly faster than theoretically predicted in the case of an array of parallel dislocations.
Overall, in any case, we see that theory and OKMC simulations are in mutual agreement for both dislocations and toroidal shaped sinks, such as SIA loops, provided that the volume fractions are small. The master curve representation is globally confirmed by OKMC simulations independently of the shape of the absorbers.
\section*{Acknowledgement}
This work was carried out as part of the PERFORM60 project of the 7th Euratom
Framework Programme, partially supported by the European Commission, Grant
agreement number FP7-232612. Discussions with S. Golubov and A. Barashev were of fundamental importance to interpret the results of the present work.
\bibliographystyle{model1a-num-names}
|
\section{Introduction}
In this paper, we examine fundamental upper limits of the light
trapping efficiency of a class of one-dimensional gratings consisting
of amorphous silicon (a-Si) and silver as a reflecting substrate. We focus
uniquely on the {\it light-trapping properties} of a class of lamellar
surface relief
gratings for which we can solve Maxwell's equations in the continuum
limit, i.e. without discretization in physical space, and use
analytical expressions for the electromagnetic field as a function of
the spatial coordinates. The main numerical work consists in the
solution of algebraic eigenvalue problems, and at the end of solving a
system of linear equations. Everything else can be expressed analytically.
In order to obtain upper limits for light-trapping efficiency we use an
ideal absorption-free reflector made from a hypothetical metal,
idealised silver (Ag$^*$) whose dielectric permittivity is real and
identical to the real part of the permittivity of silver
$\epsilon_{Ag}(\lambda)$. We also employ an ideal
antireflection coating (AR$^*$) that suppresses reflection in the
whole wavelength range of a-Si absorption. The reason for this choice
of back reflector and antireflection coating, is that we wish to focus
our attention uniquely on the light-trapping efficiency and want to
make sure that all the photons in the reflected light have had a
chance of being absorbed in the semiconductor.
We wish to point out that the absorption in a real metal will not only lead
to a reduction of absorption
in the photoelectrically active region of the semiconductor, but also
affect the number of photons leaving the cell and in this way reduce the
intensity of reflected light. Our results present the basis for predicting the
improvement possible by reducing the parasitic absorption in the
metallic reflector.
In the future, we may hope that losses by absorption in the metal
reflector and reflection at the front surface will be reduced more
than presently possible or even will be overcome by some novel methods.
Such exact limits have traditionally played an important role,
the best known example being the Carnot limit of thermodynamic
engines, as well as in photovoltaic converters, e.g. the detailed
balance limit established by Shockley and Queisser
\cite{Shockley:1961lr}, or the famous
limit obtained for random scatterers by Yablonovitch and Cody
\cite{Yablonovitch:1982lr}. These upper limits serve as a yardstick against which
progress can be measured and which indicate what might become
possible
In the context of light-trapping induced by diffraction gratings, we are
not able to present upper limits of such general validity. They have
been the subject of recent work by
\cite{Yu12102010,haug2011resonances,Naqavi2011optimal1d}. They are
typically made within mode-coupling theory using assumptions about
possible couplings between the modes that are consistent with general
principles, e.g. detailed balance.
Here, we wish to present limits of light-trapping efficiency for concrete
diffraction gratings without making assumptions about the couplings
between modes, but obtaining them from exact calculations which are
possible for the classes of diffraction gratings with lamellar
geometry chosen for the present study as well as generalizations of
those with 2-dimensional periodicity.
All work on light-trapping is motivated by the need to increase light
absorption in solar cells with indirect
bandgap semiconductors. The use of diffraction gratings
to improve absorption in solar cells is well known, and has been
described in detail in the review article by S. Mokkapati and K.R.
Catchpole\cite{mokkapati2012nanophotonic}.
We present an optical study of diffraction structures with a-Si
as semiconductor material primarily for practical reasons.
In particular, silicon is abundant, and methods for vapour deposition
of a-Si are well developed. Unlike thin solar cells
using a direct band-gap semiconductor that absorbs very well for all
photon energies above the band-gap, a-Si becomes a weak
absorber in the near-infrared, and thus long optical
path lengths within the material are required. On the other hand,
electronic properties of the semiconductor such as recombination losses
set upper limits on layer thickness, and a compromise between absorption
and these electronic properties is usually found at around $200$
nm layer thickness for a-Si solar cells. All results
for the spectrally integrated absorption $A_{int}$ quoted in this
paper, represent spectrally integrated values weighted with the solar AM1.5
spectral photon density. Note that the number of absorbed photons
is the relevant quantity, when the quantum
effciency is limited to one electron-hole pair per absorbed photon.
Our method for solving Maxwell's equations follows reference \cite{morf1995exponentially}.
The method described therein solves the diffraction problem for a
lamellar gratings by first solving the Helmholtz eigenvalue equation for
each layer. For each layer, there is a set of eigenvalues and
eigenfunctions that form the basis in which the electromagnetic field
is expanded. Continuity of the field and its derivative normal to the
layer, see below, allows to relate expansion coefficients between
different layers, and at the bottom and top of the structure there exist
boundary conditions that define a unique solution to the system of
linear equations linking all expansion parameters.
The eigenmodes obtained with this method are numerically exact, but
of course finite in number.
Following the strategy outlined in reference \cite{morf1995exponentially},
the boundary value problem can be solved for arbitrary grating depth
and arbitrary number of eigenmodes without encountering numerical
instability. However, the finite number of eigenmodes
taken into account implies that the resulting calculation is still approximate.
Interfaces with sharp optical boundaries with discontinuities in the
permittivities are inherent to the problem and lead to the excitation
of an infinite number of evanescent modes with very short decay length
in the perpendicular direction. Therefore high spatial resolution or,
as in our case, a large number of modes need to be taken into account.
The case of the polarisation with magnetic field parallel to the grating
grooves (H polarisation) is particularly slowly converging as the
H field has discontinuities in its normal derivative at interfaces
between different materials. The case of E polarisation has better
convergence properties as the first derivative of the E field is continuous,
only its second derivative being discontinuous. These discontinuities
of first (H-polarisation) or second (E-polarisation) derivatives are particularly large
at interfaces between metals and semiconductor, in our case between
silver and a-Si. This difficulty needs to be addressed
in particular in the spectral range where the absorptivity of the
semiconductor becomes weak, i.e. where light-trapping is most needed.
It is this problem that we address in the present work.
\begin{figure}[ht]
\centering
\includegraphics[width=6.6cm]{figure1.pdf}
\vskip -2mm
\caption{Grating structure with two grating layers: a) Fully symmetric,
b) Mirror axis in upper layer shifted.
Mirror symmetry in individual layers remains.}
\label{sym-antisym}
\end{figure}
We limit our studies to perpendicular incidence and
we only treat the absorption in the semiconductor layer, and note
that the parasitic absorption in the metal layer will be discussed
in a forthcoming publication.
Furthermore, we confine our discussion
to structures with two interfaces per layer, leading to a mirror symmetry
in each grating layer, as shown in Figure \ref{sym-antisym}. We benefit
from this symmetry property to separate the eigenfunction space into
symmetric and antisymmetric subspaces in each layer. As long as the
entire structure retains the same mirror plane, then symmetric eigenfunctions
do not couple to antisymmetric eigenfunctions. Since the zero order
light incident perpendicular to the device belongs to the symmetric
class, only symmetric eigenfunctions will couple to the incident light.
As a coupling is induced between symmetric and antisymmetric eigenfunctions
in different layers by asymmetric structures of the type shown in
the lower panel of Figure \ref{sym-antisym}, it is interesting to
study whether asymmetric structures lead to improved light trapping,
cf. Reference \cite{heine1995submicrometer}.
The simplifications made possible by geometrical symmetry and
perpendicular incidence allow us to present results corresponding to
resolutions that are a factor of two larger. We present our
calculations with varying resolutions to provide a measure of the
precision of such calculations and to assess the continuum limit.
In section 2, we outline the problem and our approach for its solution.
In section 3.1 we examine the absorption behavior in the semiconductor
layer in a one-layer grating with rectangular geometry. Specifically,
we look at a structure with an absorption free metallic substrate
and an absorbing semiconductor layer with an antireflective cover.
In particular we describe how the absorption extends to larger wavelengths
as a function of grating depth, and we find an optimum depth for maximum
absorption. This optimum depth is almost the same for both polarisations.
In Section 3.2 we investigate multi-layer staircase gratings that
mimic sinusoidal gratings. Here we observe again that the optimum
thickness for the grating structures does not differ much between
the two polarisations. We note that with increasing number of grating
layers for the same depth and structure, the spectra differ in some
parts, while the integrated absorption $A_{int}$ is almost independent of the
number of grating layers included. The absorption depends little on
the period $\Lambda$ of the structure for both a rectangular and
a stair-case sine grating.
In Section 3.3, we investigate asymmetric structures. In the simplest
case, they consist of two symmetric layers with interfaces between
the metal and the semiconductor. As layers are moved horizontally
with respect to one another, symmetry of the entire structure is broken,
as shown in Figure \ref{sym-antisym} b). As long as the symmetry
axes for the two layers coincide, only the symmetric eigenfunctions
are populated in each layer. But as soon as the symmetry is broken,
symmetric eigenfunctions couple to antisymmetric eigenfunctions of different layers, and vice versa, such that all modes are populated. We investigate how these additional modes
excited by the coupling between symmetric and antisymmetric eigenfunctions
may increase absorption. Note that these symmetry considerations apply
only to the case of perpendicular incidence
The paper closes with a discussion of the observed effects and an
outlook.
\section{Method}
In this section we first give a brief recapitulation of the calculational
method \cite{morf1995exponentially}. Then we discuss the specific
changes necessary for the present calculations.
\paragraph*{Setup}
We split the diffraction problem into regions I-III, as shown in Figure
\ref{coordinates}, using the polarisation convention denoted there.
We denote $F(x,y)$ as either the electric field in the y-direction
$E_{y}(x,y)$ for E parallel polarisation or similarly the magnetic
field in the y-direction $H_{y}(x,y)$ for H parallel polarisation.
In region I, we study the case of perpendicular incidence, that is
the incident field can be written as
\begin{equation}
F^{inc}(x,z)=1\times\exp(-ik_{0}z),\label{F1}
\end{equation}
where $k_0=2\pi/\lambda$ and for regions I and III, the outgoing light can be written as
a superposition of plane waves,
\begin{eqnarray}
F^{I}(x,z) & = & \sum_{n=-M}^{M}a_{n}^{+(I)}\exp(i(\alpha_{n}x+\chi_nz)),\\
F^{III}(x,z) & = & \sum_{n=-M}^{M}a_{n}^{-(III)}\exp(i(\alpha_{n}x-\chi_nz)),
\end{eqnarray}
where
\begin{equation}
\alpha_{n}=\frac{2\pi}{\Lambda}n,\qquad\textrm{and}\qquad\chi_{n}=\sqrt{\epsilon^{I,III}k_{0}^{\,2}-\alpha_{n}^{\;2}}.
\end{equation}
Region II contains the diffraction grating consisting of $q$ layers,
with thickness $h_{j}=z_{j+1}-z_{j}$, where $1\le j\le q$ refers
to the layer position within the grating.
\begin{figure}[ht]
\centering
\includegraphics[width=7cm]{figure2}
\vskip -3mm
\caption{\label{coordinates} General problem setup and polarisation convention.}
\end{figure}
\paragraph*{Grating Problem}
We briefly sketch our method for solving the grating problem,
developed by one of us \cite{morf1995exponentially}.
As we confine our study to perpendicular incidence, we note that the
fields $E$ and $H$ do not depend on y. Within Region II, we briefly
note that because of the lamellar structure, inside each layer the
permittivity $\epsilon(x,z)$ does not depend on the $z$. Thus, a
separation of variables becomes possible and the field $F(x,z)$ can
be written as a sum of products $X_{n}(x)\times Z_{n}(z)$, the functions
$X_{n}(x)$ is an eigenfunction Helmholtz equation which needs to
be solved within each grating layer. Because $\epsilon(x)$ is piecewise
constant, the Helmholtz equation can be written piecewise, too. For
simplicity, we limit ourselves to two domains of different permittivity:
\begin{eqnarray}
\partial_{x}^{2}\, X_{n}^{(j,1)}(x)+k_{0}^{\;2}\epsilon_{1}^{(j)}X_{n}^{(j,1)}(x) & = & \mu_{n}^{(j)2}X_{n}^{(j,1)}(x)\label{eq:helmholtz1}\\
\partial_{x}^{\;2}X_{n}^{(j,2)}(x)+k_{0}^{\;2}\epsilon_{2}^{(j)}X_{n}^{(j,2)}(x) & = & \mu_{n}^{(j)2}X_{n}^{(j,2)}(x)\label{eq:helmholtz2}
\end{eqnarray}
Here, equation \ref{eq:helmholtz1} refers to domain 1 and equation
\ref{eq:helmholtz2} refers to domain 2 and $\mu_{n}^{(j)^2}$ is $n$-th
eigenvalue of the Helmholtz equation in layer $j$ and
$X_{n}^{(j,1)}(x)$ and $X_{n}^{(j,2)}(x)$ are the domain-wise parts of the
the corresponding $n$-th eigenfunction $X_{n}^{(j)}(x)$
They need to fulfill a set of boundary conditions
at interfaces $x_{i}$ in addition to the eigenvalue equations,
\begin{eqnarray}
X_{n}^{(j,1)}(x_{i}^{-}) & = & X_{n}^{(j,2)}(x_{i}^{+})\\
\frac{1}{\epsilon_{1}^{(j)}}\partial_{x}X_{n}^{(j,1)}(x)\mid_{x=x_{i}-\delta} & = & \frac{1}{\epsilon_{2}^{(j)}}\partial_{x}X_{n}^{(j,2)}(x)\mid_{x=x_{i}+\delta}
\end{eqnarray}
for $H$ parallel polarised light in the limit $\delta\rightarrow0$,
and likewise for $E$ parallel polarised light by replacing $\epsilon_{1}$
and $\epsilon_{2}$ by $1$.
The electromagnetic field $F^{(j)}(x,z)$ can then be expanded in terms
of the eigenfunctions $X_n^{(j)}(x)$
as a sum of modes propagating in between the adjacent layers,
\begin{eqnarray}
F^{(j)}(x,z)& = &\sum_{n=1}^{N}\left(a_{n}^{(j)-}/A_n^{(j)}(z)+
a_{n}^{(j)+}A_n^{(j)}(z)\right)X_{n}^{(j)}(x),\nonumber \\
A_n^{(j)}(z)& = &\exp(i\mu_{n}^{(j)}(z-z^{(j)}).\label{eq:planewave}
\end{eqnarray}
where $\mu_{n}^{(j)}$ is the square root
of the eigenvalue $\mu_{n}^{(j)2}$,
with the sign convention that the imaginary part of $\mu_n^{(j)}$ be
positive if it is non-zero. This sign definition ensures that the
amplitudes $a^{+}$ ($a^{-}$) refer to eigenmodes
that as functions of $z$ are exponentially decreasing (increasing),
respectively.
The Helmholtz equation together with the sets of basis functions and
the additional constraint of boundary conditions between the domains
define the numerical eigenvalue problem for each grating layer $j$.
This method allows the calculation of a freely selectable number of eigenvalues
and eigenfunctions to a precision limited only by machine number size:
The electromagnetic field is infinitely many times differentiable
inside each domain of constant permittivity, a fact that can be verified
easily by repeated differentiation of the Helmholtz equation. If the
functions $X^{(j,k)}(x)$ are expanded in terms of polynomials $P_{m}^{(j,k)}(x)$
that are orthogonal for $x$ inside the domain $D^{(j,k)}$, one obtains
exponential convergence in the number $M$ of included basis functions
for eigenvalues and eigenfunctions. Thus, within
a grating layer, our eigenmodes are obtained to the same precision
as by analytic calculation following references by Botten et
al. \cite{botten1981dielectric,botten1981finitely,botten1981highly}.
Note that the symmetry at perpendicular incidence together with the
symmetry within the grating layer results in either purely symmetric
or antisymmetric functions. If the basis functions $P_{m}^{(j,k)}(x)$
are also either symmetric( antisymmetric) for $m$ even(odd) or odd
$m$), half the number of coefficients defining the eigenfunctions
$X^{(j,k)}(x)$ will vanish, and therefore the symmetric eigenvalue
problem can be formulated either for symmetric or antisymmetric case
in terms of the $M/2$ non-zero variables and requires only
$\frac{1}{8}$ of the numerical effort as it
scales with the number of variables to the third power. For entirely
symmetric structures, this property carries over to the solution of
the radiation condition boundary value problem.
At the interface $z=z_{0}$ between grating layers, the field $F=E_{y}$
or $F=H_{y}$ has to satisfy the following continuity equations
\begin{eqnarray}
F(x,z_{0}+\delta) & \equiv & F(x,\, z_{0}-\delta)\\
(\frac{\partial_{z}F(x,z)}{\epsilon(x,z)})\mid_{z=z_0+\delta}& \equiv &(\frac{\partial_{z}F(x,z))} {\epsilon(x,z)})\mid_{z=z_0-\delta}
\end{eqnarray}
identically for $0\le x\le\Lambda$ and for $\delta\rightarrow0$,
for H-polarisation and likewise for E-polarisation dropping the term
$1/\epsilon(x,z)$. Inserting the expansion (\ref{eq:planewave})
leads to linear equations between the expansion coefficients $a_{n}^{(j)\pm}$
of the upper layer $j$ and $a_{n}^{(j-1)\pm}$ of the layer $j-1$
underneath. The fact that eigenmodes in the upper (lower) layer have
discontinuities at positions $x_{j}$ where the eigenmodes in the
lower (upper) layer are analytic, leads to a Gibbs phenomenon at $x=x_{j}$,
when eigenfunctions in the lower layer are expanded in terms of eigenfunctions
in the upper layer. This gives rise to slow convergence and thus a
large number of modes have to be included if the optical behaviour
at such structures needs to be known accurately. This is the case
even if the eigenmodes are known analytically. For H polarisation
the convergence is substantially slower than for E polarisation because
both the second and the first derivative in the direction perpendicular
to the interface are discontinuous, whereas for E polarisation only
the second derivative is discontinuous.
The substantial reduction in modes by fully exploiting
symmetry and the stable implementation of the boundary value problem
following \cite{morf1995exponentially} allows us to achieve the high
resolution necessary to obtain reliable absorption spectra for such
structures.
\paragraph*{Anti-Reflective Coating}
All the structures in this paper are calculated with an idealised
artificial antireflective coating that almost completely eliminates
reflection of the incident light at the top interface of the structure
structure. We choose this setting, because we wish
to describe an optimum limit for the absorption $A_{int}$ within the semiconductor
layer, while the light is incident from outside. The antireflective
coating used in our calculations consists of a homogeneous layer
with an artificial material AR{*} whose refractive index is chosen
wavelength dependent $n^{*}(\lambda)$=$\sqrt{\Re n_{a-Si}(\lambda)}$,
where $n_{a-Si}$ is the refractive index of a-Si in
the layer below, and the $\Re$ symbol stands for the real part. We
take the antireflection coating to have a thickness corresponding
to $\frac{\lambda}{4n^{*}(\lambda)}$. Of course this is very
artificial. However there exist realizations of antireflective
structures which lead to similarly low reflexion
losses \cite{heine1995submicrometer},\cite{heine1996coatedsubmicron}.
The broadband antireflection structure described therein consists
of a rectangular grating with an additional coating with a lower refractive
index than the grating below.
\section{Results}
\subsection{rectangular gratings\label{sub:rectangular-gratings}}
Here, we look at gratings consisting of only one grating layer. Furthermore,
we use an idealised, absorption free silver reflector whose permittivity
$\epsilon(\lambda)$ is given by the real part of the permittivity
of bulk silver, cf. Appendix.
\begin{figure}[ht]
\centering
\includegraphics[width=7cm]{figure3}
\vskip -2mm
\caption{Depth dependence of absorption $A_{int}$ for rectangular grating with
period $\Lambda=441$ nm for maximum absorption of unpolarised
light and an a-Si layer with thickness $h_{mean}=200$ nm.}
\label{depthdependence-rect}
\vskip -3mm
\end{figure}
The rectangular gratings are constructed like the ones shown in Figure
\ref{sym-antisym}, with a width of the metallic domain $b_{3}=0.5\Lambda$, and
$h_{2}=0$ and $h_{mean}=200$ nm, which stays constant while varying the
grating layer thickness. Thus, the amount of semiconductor material
is kept fixed. We also limit ourselves to grating depths that are at most
about one third of the semiconductor layer thickness. Figure \ref{depthdependence-rect}
shows the depth dependence of the absorption $A_{int}$ in a rectangular structure
at the optimum period for the first absorption maximum of unpolarised
light. In Figure \ref{depthdependence-rect} and the following similar
figures, an integrated absorption of $100$\% corresponds to the absorption of all
photons in the spectral range of $350$ nm up to $770$ nm, weighted
with an AM-1.5 spectrum. We limit the spectral range of our studies
to $550-770$ nm, where light trapping is most helpful, and the contribution
of this region to the total absorption is limited to $62.6$\% for
infinite thickness.
Figure \ref{depthdependence-rect} shows that the maximum absorption $A_{int}$
for H parallel polarisation in the limited range of $550-770$ nm is almost $44$\%, whereas for E parallel
polarisation it is $\approx 37$\%, and the optimum depth for both polarisations
is close to $40$ nm. The figure also shows that the E parallel polarisation
is fully converged, whereas for H polarisation the value is between
$44.0$\% and $43.2$\%. To these values one may add a contribution
of 38\% for full absorption in the wavelength range from 350-550nm,
if reflection is suppressed. Thus, the total absorption $A_{int}$
amounts to 62\% for an ideal planar reflector, and if an ideal
reflector in the shape of this rectangular grating is used, it is
75\% for E polarisation and between 81.2 and 82.0 percent for H polarisation.
\begin{figure}[h!]
\centering
\includegraphics[width=6.6cm]{figure4}
\vskip -2mm
\caption{The absorption spectrum of the rectangular grating with $h_{mean}=200$
nm, a period $\Lambda=401$ nm, and $36$ nm grating depth.}
\label{spectrum-rect}
\vskip -2mm
\end{figure}
\begin{figure}[ht]
\includegraphics[clip,width=7cm]{figure5}
\vskip -3mm
\caption{Absorption peaks in E-polarisation converged,
in H-polarisation they shift with N, keeping the integrated
absorption fixed within 0.2\%.}
\label{spectrum-zoom}
\vskip -4mm
\end{figure}
The spectrum in Figure \ref{spectrum-rect} of the absorption shows
that the grating leads to a spectral shift of the absorption edge,
and the H parallel polarisation shows significantly more absorption,
especially in the infrared region. We observe that even for weakly
absorbing semiconductor layers, the strong resonances can lead to
a noticeable increase in absorption. Note that Figure \ref{spectrum-rect}
shows that the region between $700$ and $750$ nm is highly sensitive.
This behaviour is further investigated in Figure \ref{spectrum-zoom},
which shows spectral mode dependence of high order calculations. Note
how mostly the resonances move to higher wavelengths when increasing
the number of modes for the calculation, such as the one at $708$
nm. The integrated absorption values for H-polarisation vary little
for different mode numbers, even for small values of $N$, as can be
seen in Figure \ref{depthdependence-rect}, as has been mentioned above.
Due to the slow convergence we are unable to present absorption
spectra for H polarisation in the limit of large $N$.
On the contrary, for E polarisation our spectra are well converged.
Figure \ref{spectrum-zoom} also shows how the background of the absorption
spectrum, i.e. the broad absorption below the peaks clearly increases:
It may be taken as $\approx 0.3$ near the $708$ nm resonance, and
still around $0.2$ for the $728$ nm resonance. Such values would
demand much greater thicknesses $d$ of a-Si, namely
$d\approx700$ nm for the absorption value 0.3 at $\lambda=700$ nm
and even $d\approx1200$ nm for the value 0.2 at $\lambda=730$ nm,
if instead of the grating a planar reflector is used. Thus, the effective
thickness of the a-Si layer is increased by a factor
of 3.5 at $\lambda=700$ nm and about 6 at $\lambda=1200$ nm.
\begin{figure}[ht]
\centering
\includegraphics[width=6.6cm]{figure6}
\vskip -3mm
\caption{Spectrum of strongly excited modes $2$, $4$ and $5$ and of mode $1$ for comparison.}
\label{spectrum-eigenmodes}
\vskip -3mm
\end{figure}
In the following, we take a closer look at the eigenmodes of the rectangular
grating for which we have shown the spectrum in Figure \ref{spectrum-zoom}.
Figure \ref{spectrum-eigenmodes} shows the amplitudes
$a_{k}$ for the eigenmodes of the Helmholtz equation, sorted according
to increasing imaginary part of the eigenvalue implying shorter decay
length in the z-direction. The amplitudes are
shown at the interface between the grating layer and the metallic
reflector, and they result from solving the multilayer boundary value
problem, with an incident amplitude $a_{0}^{-}=1$, as defined in equation (\ref{F1}). Interestingly,
the second and fourth eigenmodes are almost 60 times as much excited as
the incident light, corresponding to a very high photon density near
the bottom interface. We expect this excitation to become much weaker
if losses in the metallic reflector are included. The fourth
and fifth eigenmodes switch order near $\lambda=727$ nm.
\begin{figure}
\centering
\includegraphics[width=7cm]{figure7}
\caption{The strongly excited $2^{nd}$ and $4^{th}$
eigen-functions at $\lambda=729$ nm. a) real, b) imaginary part.}
\label{fig:spatial-eigenmodes}
\vskip -3mm
\end{figure}
In Figures \ref{fig:spatial-eigenmodes}, we show the spatial behaviour
of the strongly excited eigenmodes. Note how both modes display a
sharp discontinuity at the interface between the two materials, as
required by the boundary conditions. The structure of these modes
shows clearly why the discontinuity of the first derivative gives
rise to a strong Gibbs phenomenon leading to the excitation of evanescent
waves of high order and consequently slow convergence as a function
of the number of modes included in the expansion of the electromagnetic
field. The comparatively small imaginary part in these eigenfunctions
is the result of the rather weak absorptivity of a-Si
at a wavelength of 729nm. We also note that a Fourier expansion of
these eigenfunctions will not allow to correctly compute the absorption
due to the 4-th eigenmode that mainly occurs in the vicinity of the
interface. Here, the derivative of the imaginary part shows a dramatic
discontinuity. Thus a Fourier expansion will converge only slowly in
its vicinity. In
view of the very large amplitude of this mode, the exact treatment of
the boundary conditions at the a-Si-AG$^*$ interface as implented in
our work is of particular relevance.
\subsection{Mimicking sinusoidal gratings\label{sub:Mimicking-sinusoidal-gratings}}
Here, we want to study more complex structures, in particular we are
interested in structures that mimic a sinusoidal grating. For that,
we want an optimised grating structure, which is not defined by a
number of parameters, but uniquely defined by a single parameter.
This should facilitate changing the structure depth without altering
the grating characteristics. To that end, we construct our grating
structure as follows: We fix the layer thickness $h_{j}$ and interface
positions $x_j$ by requiring that in each layer the excess area of the
lamellar approximation above the sine cancels the missing area
underneath it, and that the sum of excess areas is minimal and equal
to the sum of missing areas. This minimization leads to a unique
solution for $x_j$ and $h_j$ depending only on the number of layers
$n$, which we will refer to as pseudo-sine. The amount of
semiconductor material is then identical
for sine and pseudo-sine, see Figure \ref{pseudo-sinus-sketch}.
\begin{figure
\centering
\includegraphics[width=6.5cm]{figure8}
\vskip -2mm
\caption{Illustration of 3 (a) and 5 (b) layer staircase gratings mimicking a sinusoidal grating.}
\label{pseudo-sinus-sketch}
\end{figure}
Figure \ref{perioddependence-rect} shows the integrated absorption $A_{int}$
for the interval between $550$ nm and $770$ nm, and how the first
absorption maximum for both rectangular and pseudo-sine gratings only
depends weakly on the period $\Lambda$. In particular, we can see
that the 5-step pseudo-sine offers more absorption, and that in both
cases, the H polarisation is more readily absorbed. In Figure \ref{perioddependence-rect},
the depth of $36$ and $72$ nm for the rectangular and for the pseudo-sine
structure respectively are chosen close to the first absorption maximum
for each structure type at around $40$ and $70$ nm as can be seen
in Figures \ref{depthdependence-rect} and \ref{depthdep-sine}.
\begin{figure}[ht]
\centering
\includegraphics[width=6.3cm]{figure9}
\caption{Grating period dependence of integrated absorption:
rectangular grating with depth $h=36$ nm vs. a $5$-step
sine grating with $h_{tot}=\sum_j{ h_j}=72$ nm total depth.
At these depth values, integrated absorption is close to maximum.
a-Si layer thickness $h_{mean}=200$ nm.}
\label{perioddependence-rect}
\end{figure}
Figure \ref{depthdep-sine} shows that the integrated absorption
depends weakly on the number of modes used for the calculation, and
in particular that for the E parallel polarisation, $24$ modes are
sufficient to achieve convergence, whereas similar curves for
H parallel polarisation are already in good agreement, but still varying
for deep structures. It also shows a significant difference in optimum
depth between polarisations, unlike for rectangular gratings, and
that H parallel polarised light can be absorbed more effectively.
When further comparing the depth dependence in Figure \ref{depthdep-sine}
with Figure \ref{depthdependence-rect}, note that the optimum depth
for a pseudo-sine structure is almost twice the optimum depth of the
corresponding rectangular structure, and the optimum for the pseudo
sine structure is much flatter with respect to varying depths.
\begin{figure}[ht]
\centering
\includegraphics[width=6.8cm]{figure10}
\caption{Depth dependence for a 5 step pseudo-sine grating with period $\Lambda=421$
nm.}
\label{depthdep-sine}
\end{figure}
The spectra in Figure \ref{spectrum-sine-E} and \ref{spectrum-sine-H}
show that the pseudo-sine gratings offer more absorption over the
integrated spectral range when compared to the rectangular gratings,
both by increasing the background absorption in the red and near
infrared, and by exciting additional resonances.
\begin{figure
\centering
\includegraphics[width=6.8cm]{figure11}
\caption{Spectrum of 3 and 5-step pseudo-sine vs rectangular and flat reference, $N=36$ modes for E parallel incidence}
\label{spectrum-sine-E}
\end{figure}
\begin{figure
\centering
\includegraphics[width=6.8cm]{figure12}
\caption{Spectra for H parallel incident light, same gratings as in Figure \ref{spectrum-sine-E}}
\label{spectrum-sine-H}
\vskip -4mm
\end{figure}
We note that the background absorption of the pseudo-sine grating
without taking any peaks into account takes values of about 0.6
at $\lambda=680$ nm and about 0.4 at $\lambda=730$ nm. A device
with a planar reflector would have to have an a-Si thickness
of about $1000$ nm at $680$ nm wavelength, and even $2700$ nm at
$730$ nm wavelength, i.e. 5 and 13 times the actual thickness of the
device with the pseudo-sine grating.
For the pseudo-sine, the optimum parameters in terms of depth and
period for the two polarisations are further apart when compared to
a rectangular grating. The spectral comparison of 3-step and 5-step
pseudo-sine gratings with period $401$ nm and depth $72$ nm in Figure
\ref{spectrum-sine-E} reveals that the spectra are almost identical
for the E parallel polarisation. Thus the absorption spectra are almost
independent of the number of layers chosen for the pseudo-sine approximation.
This shows the validity of our geometrical argument for the unique
choice of the pseudo-sine approximation. On the other hand, the same
spectral comparison for the H polarisation in Figure \ref{spectrum-sine-H}
shows differing spectra for the 3 and 5-step pseudo-sine gratings,
but they do share common spectral features, explaining the good convergence
of the integrated absorption. Because both the mode number and the
number of layers included are very limited, the pseudo-sine spectrum
is not converged and must not be confused with the spectrum for a
true sine grating, for which other methods such as the one pioneered
by Chandezon et al. \cite{chandezon1980original} will be required.
The 5-step pseudo-sine grating shows a maximum integrated absorption
in H polarisation of about 52\%, while for unpolarised light it is
given by the mean value from E- and H-polarisation and is about 45\%.
These values include the spectral range from 550-770nm. Adding
the contribution of 38\% from the wavelength range 350-550 nm, where
light-trapping is not needed, we obtain a total of 90\% for H
polarisation and 83\% for unpolarised light.
\subsection{asymmetric gratings\label{sub:asymmetric-gratings}}
In this section we show the benefits of asymmetry in
the grating structure, as shown in Figure \ref{sym-antisym}. All these
results refer to the case of E-polarisation. The calculations for
H-polarisation have not yet achieved the reliability that we require in
this work. Further developments will be necessary to overcome these
difficulties.
In Figure \ref{Depth-asy}
we have calculated the depth dependence of a two-step pseudo-sine
grating with period $\Lambda=401$ nm where one layer is shifted by
$\triangle_{a}$ from the symmetry axis, using 39 modes which is sufficient
to display the high resolution limit behaviour. Figure \ref{Depth-asy}
shows that the asymmetry improves the absorption from around $37$\% maximum to around $41$\%,
and that the benefit of the asymmetric shift saturates if $\Delta_a$ is increased beyond $0.1\Lambda$.
\begin{figure}[ht]
\centering
\includegraphics[width=6.8cm]{figure13}
\vskip -2mm
\caption{Depth dependence of absorption $A_{int}$ in E-polarisation, asymmetric grating, $39$ modes.}
\label{Depth-asy}
\end{figure}
\begin{figure
\centering
\includegraphics[width=6.8cm]{figure14}
\vskip -2mm
\caption{Spectra of asymmetric and rectangular gratings for a grating depth of 60 nm.}
\label{Spectrum-Asy-a}
\end{figure}
\begin{figure}[ht]
\centering
\includegraphics[width=6.8cm]{figure15}
\vskip -2mm
\caption{Spectra asymmetric and rectangular gratings for a grating depth of 42 nm.}
\label{Spectrum-Asy-b}
\label{Spectrum-Asy}
\end{figure}
\begin{figure
\centering
\includegraphics[width=6.8cm]{figure16}
\vskip -2mm
\caption{A comparison of the spectra of asymmetric structures
for E parallel polarisation}
\label{asy-modes}
\end{figure}
Figure \ref{Spectrum-Asy-a} shows that for a structure with depth
$60$ nm, the asymmetry not increases the background absorption
into the red and near infrared, but also introduces additional resonances
around $670$ nm and around $730$ nm, which correspond to the excitation
of two antisymmetric eigenmodes of the grating. For comparison, the
spectrum at a depth of $42$ nm in Figure \ref{Spectrum-Asy-b} shows
absorption resonances at the same wavelengths. To the maximum absorption
of $41$\% observed for E-polarisation with this asymmetric grating
integrated over $550\le\lambda\le 770$ nm we can add the absorption
of 38\% from $350\le\lambda\le 550$ nm, where light-trapping is not needed,
and obtain a total absorption of 79\%, quoted in the Abstract. Finally,
Figure \ref{asy-modes} shows the absorption in E-polarisation of a grating
with period $401$nm and depth $66$ nm calculated with $N=25$ and 35 modes. Despite
large asymmetry, the spectra have well converged.
\section{Discussion}
We have investigated how much light can be trapped in a layer of a-Si
with thickness 200 nm, which is typically used for a-Si solar cells, on top
a metallic surface relief grating with one-dimensional (1-d) lamellar
geometry. In order to obtain upper
limits for the absorption, we have made the idealizations of a
loss-free metallic reflector and a perfect antireflection coating.
We have used a modal method with exact eigenfunctions
to calculate accurate absorption spectra of such gratings.
The absorption spectra for E polarisation of a sine-shaped diffraction grating
approximated by lamellar gratings with 3 and 5 layers turn out almost identical,
which is an interesting result. For such sine-like gratings, we have found
that the limit of absorption is $A_{int}$ is about 90\% for H-polarisation and
about 75\% for E-polarisation, compared to $62$\%
for a planar reflector. We have also shown that an
asymmetry in the grating structure generates additional peaks in
the absorption spectrum and can thus increase the absorption.
In Section 3.1, we have shown that {\bf exact} calculation of eigenmodes
is particularly important as significant absorption occurs near
interfaces where derivatives of the fields are
discontinuous, see Figure \ref{fig:spatial-eigenmodes}.
While we have shown that in all cases studied, light-trapping in
H-polarisation is significantly more effective than in E-polarisation,
calculating fully converged spectra for H-polarisation remains
difficult. The results presented in this paper have been made possible
by making full use of the symmetry properties of the structures studied.
One may hope that two-dimensional surface
relief gratings of similar lamellar geometry will allow unpolarised
light to be absorbed as effectively as H-polarised light is absorbed
with the 1-d gratings, reported here.
As already pointed out, our
results are idealised in the sense that absorption
in the metallic layer is suppressed. We expect that the peak
height of absorption resonances will be
strongly reduced by experimentally available metallic reflectors.
On the other hand, one may hope that the background absorption below the peaks
that is observed for all studied gratings, will still contribute significantly.
\section*{Acknowledgments}
We acknowledge the financial support of this work by the Swiss Federal
Office of Energy, and by the Paul Scherrer Institute. We also
like to thank F.J. Haug and H.P.Herzig for helpful discussions and
advice.
\section*{References}
\bibliographystyle{iopart-num}
|
\section{Introduction}
The multiple-access relay channel (MARC) is a multiuser network in which several sources communicate with a single destination with the help of a relay \cite{Kramer:2005}, \cite{Sankar:07}.
This model represents cooperative uplink communication in wireless networks.
In this work, we study the lossless transmission of arbitrarily correlated sources over MARCs, in which both the relay and the destination have access to side information correlated with the sources.
It is well known \cite{Shannon:48} that a source can be reliably transmitted over a memoryless point-to-point (PtP) channel, if its entropy is less than the channel capacity. Conversely, if the source entropy is larger than the channel capacity, then reliable transmission is not possible. Therefore, for memoryless PtP channels, a separate design of the source and channel codes achieves the optimal end-to-end performance.
However, the optimality of separate designs does not generalize to multiuser networks \cite{Cover:80}, \cite{GunduzErkip:09}, \cite{Murin:ISIT12}.
Since the MARC combines both the multiple access channel (MAC) and the relay channel models,
and since separate source-channel coding is not optimal for MAC with correlated sources \cite{Cover:80}, we conclude that separate designs are not optimal for MARCs.
Therefore, it is important to develop methods for joint source-channel coding (JSCC) for this network.
In this work we derive separate sets of sufficient and necessary conditions, which are not necessarily tight.
In deriving our sufficiency conditions we focus on cooperation schemes based on the decode-and-forward (DF) protocol, such that the sequences of both sources are decoded at the relay.
Accordingly, transmission to both the relay and the destination can benefit from joint design of the source and channel codes.
\vspace{-0.25cm}
\subsection{Prior Work}
\vspace{-0.1cm}
The MARC has received a lot of attention in recent years, especially from a channel coding perspective.
In \cite{Kramer:2005}, Kramer et al. derived an achievable rate region for the MARC with independent messages, using a coding scheme based on DF relaying, regular encoding, successive decoding at the relay, and backward decoding at the destination.
In \cite{Sankar:07} it was shown that for the MARC, in contrast to the relay channel, DF schemes with different decoding techniques at the destination yield different rate regions. Specifically, backward decoding can support a larger rate region than sliding window decoding. Another DF-based coding scheme, which uses offset encoding, successive decoding at the relay and sliding window decoding at the destination, was presented in \cite{Sankar:07}. This scheme was shown to be at least as good as sliding window decoding. Moreover, this scheme achieves the corner points of the backward decoding rate region, but with a smaller delay.
While the focus of \cite{Kramer:2005} and \cite{Sankar:07} was mainly on achievable rate regions, outer bounds on the capacity region of MARCs were derived in \cite{KramerMandayam:04}.
More recently, in \cite{Tandon:CISS:11}, Tandon and Poor derived the capacity region of two classes of MARCs, which include a primitive relay assisting the transmitters through an orthogonal finite-capacity link to the destination.
While the works \cite{Kramer:2005}, \cite{Sankar:07}, \cite{KramerMandayam:04} and \cite{Tandon:CISS:11} considered channel coding for MARCs, in \cite{Murin:IT11} we studied source-channel coding for MARCs with correlated sources.
In \cite{Murin:ISIT12} we presented an explicit example in which separate source and channel code design is suboptimal for this model.
The suboptimality of separate source and channel coding for multiuser scenario was first shown by Shannon in \cite{Shannon:61} by considering the transmission of correlated sources over a two-way channel.
Lossless transmission of correlated sources over relay channels with correlated side information was studied in \cite{ErkipGunduz:07}, \cite{Gunduz:12}, \cite{ElGamalCioffi:07} and \cite{Sefidgaran:ISIT:09}. Specifically, in \cite{ErkipGunduz:07} G\"und\"uz and Erkip proposed a DF based achievability scheme and showed that separation is optimal for physically degraded relay channels as well as for cooperative relay-broadcast channels. This work was later extended to multiple relay networks in \cite{Gunduz:12}.
The relay channel with arbitrarily correlated sources, in which one of the sources is available at the transmitter while the other is known at the relay, and the destination is interested in a lossless reconstruction of both sources, was considered in \cite{SmithVishwanath:04}, \cite{SalehkalaibarAref:ISIT11} and \cite{SalehkalaibarAref:ISIT12}.
The work \cite{SmithVishwanath:04} used block Markov irregular encoding with list decoding (based on \cite{CoverG:79}), at both the relay and the destination, to characterize sufficient conditions for reliable transmission using a separation-based source-channel code. The works \cite{SalehkalaibarAref:ISIT11} and \cite{SalehkalaibarAref:ISIT12} used block Markov regular encoding with backward decoding, in which the relay partially decodes the sequence transmitted from the transmitter prior to sending both its own source sequence and the cooperation information to the destination.
As shown in \cite{Murin:ISIT12}, source-channel separation is suboptimal for general MARCS. Therefore, optimal performance require employing a joint source-channel code.
An important technique for JSCC is the correlation preserving mapping (CPM) technique in which the channel codewords are correlated with the source sequences. CPM was introduced in \cite{Cover:80} in which it was used to obtain single-letter sufficiency conditions for reliable transmission of discrete, memoryless (DM) arbitrarily correlated sources over a MAC. CPM typically enlarges the set of feasible input distribution, thereby enlarging the set of sources which can be reliably transmitted compared to separate source and channel coding.
The CPM technique of \cite{Cover:80} was extended to source coding with side information for MACs in \cite{Ahlswede:83}, to broadcast channels with correlated sources in \cite{HanCosta:87} (with a correction in \cite{KramerNair:09}), and to the transmission of correlated sources over interference channels (ICs) in \cite{SalehiKurtas:93}. However, when the sources are independent, the region obtained from \cite{SalehiKurtas:93} does not specialize to the Han and Kobayashi (HK) region of \cite{HK:81}. Sufficient conditions for reliable transmission, based on the CPM technique, which specialize to the HK region were derived in \cite{LiuChen:2011}.
The transmission of independent sources over ICs with correlated receiver side information was studied in \cite{GunduzLiu:Rev}, where it was shown that separation is optimal when each receiver has access to side information correlated only with its own desired source. When each receiver has access to side information correlated only with the interfering transmitter's source, \cite{GunduzLiu:Rev} provided sufficient conditions for reliable transmission based on the CPM technique together with the HK superposition encoding and partial interference cancellation.
Although CPM implements JSCC, in \cite{Dueck:81} Dueck observed that the sufficiency conditions derived in \cite{Cover:80} are not necessary. Therefore, in this work, in addition to sufficient conditions, necessary conditions are considered as well.
Observe that the feasible joint distributions of the sources and the respective channel inputs for the MAC (and for the MARC), must satisfy a Markov relationship which reflects the fact that the channel inputs at the transmitters are correlated {\em only via the correlation of the sources}.
In \cite{Cover:80}, in addition to the single-letter sufficient conditions, multi-letter necessary and sufficient conditions, which account for the above constraint, were also presented.
However, as noted in \cite{Cover:80}, these conditions are based on $n$-letter mutual information expressions, and thereby not computable.
The work \cite{Mitran:11} followed the lines of \cite{Cover:80}, and established necessary conditions for reliable transmission of correlated sources over DM MARCs, which are based on $n$-letter expressions. Furthermore, \cite{Mitran:11} showed that in some cases source-channel separation is optimal and the $n$-letter expressions specialize to single-letter expressions.
In contrast to \cite{Cover:80}, in \cite{Kang:2011} Kang and Ulukus used the above constraint to derive a new set of {\em single-letter} necessary conditions for reliable transmission of correlated sources over a MAC.
\vspace{-0.3cm}
\subsection{Main Contributions}
\vspace{-0.1cm}
This work has a number of important contributions:
\begin{enumerate}
\item
We derive a novel JSCC achievable scheme for MARCs. The scheme uses CPM for encoding information from the sources to {\em both} the relay and the destination. The relay, on the other hand, uses SW source coding\footnote{Throughout this work we refer to separate source-channel coding (i.e., a source code followed by a channel code) as encoding using SW source coding.} for forwarding its cooperation information. Therefore, the sources and the relay send {\em different} types of information to the destination: the sources send source-channel codewords, while the relay sends binning information (SW bin indices). This is in contrast to the schemes of \cite[Thm. 1, Thm. 2]{Murin:ISIT12}, and to \cite{SalehkalaibarAref:ISIT11}, in which the same type of information is sent to the destination from the sources as well as from the relay (either SW bin indices or source-channel codewords).
The new scheme uses the DF strategy with successive decoding at the relay and simultaneous backward decoding of both cooperation information and source sequences at the destination.
This scheme achieves {\em the best known results for all previously characterized} special cases.
\item
We show that, similarly to the capacity analysis for MARCs, also for JSCC simultaneous backward decoding of the cooperation information and source sequences at the destination, outperforms sequential backward decoding at the destination. We also show that simultaneous backward decoding at the destination outperforms the scheme derived in \cite[Thm. 1]{Murin:ISIT12}.
Additionally, we show that there are cases in which simultaneous backward decoding at the destination strictly outperform the schemes derived in \cite{Murin:ISIT12}. This is proved through an explicit analysis of the error probability for a specific MARC model.
\item
We derive three new sets of single-letter necessary conditions for reliable transmission of correlated sources over DM MARCs. The first set of conditions is a ``MAC-type" bound, considering the cut around the sources and the relay, while the other two sets are ``broadcast-type" bounds, derived using the cut around the destination and the relay.
The new sets of necessary conditions are shown to be at least as tight as previously known conditions, and in some scenarios, the new sets are strictly tighter than known conditions.
\end{enumerate}
The rest of this paper is organized as follows: in Section \ref{sec:Preliminaries} we introduce the notations and the channel model. In Section \ref{sec:prevSchemes} we briefly review the existing schemes and give motivation for a new JSCC scheme. In Section \ref{sec:MixedJointAchiev} we present the new achievability scheme and derive it's corresponding set of sufficiency conditions. In Section \ref{sec:Comparison} a comparison between the existing schemes and the new scheme is presented. Necessary conditions are presented in Section \ref{sec:NecessaryConditions}, and concluding remarks are provided in Section \ref{sec:conclusions}.
\vspace{-0.15cm}
\section{Preliminaries} \label{sec:Preliminaries}
\vspace{-0.2cm}
\subsection{Notations}
\vspace{-0.15cm}
\input{Notations.tex}
\vspace{-0.3cm}
\subsection{System Model} \label{subsec:model}
\vspace{-0.1cm}
The MARC consists of two transmitters (sources), a receiver (destination) and a relay.
Transmitter $i$ observes the source sequence $S_i^n$, for $i=1,2$.
The receiver is interested in a lossless reconstruction of the source sequences observed by the two transmitters, and the objective of the relay is to help the transmitters and the receiver in reconstructing the source sequences.
The relay and the receiver each observes its own side information, denoted by $W_3^n$ and $W^n$, respectively, correlated with the source sequences.
Figure \ref{fig:MARCsideInfo} depicts the MARC with side information scenario.
\begin{figure}[ht]
\vspace{-0.3cm}
\centering
\captionsetup{font=small}
\scalebox{0.5}{\includegraphics{MARC_SideInfo.eps}}
\caption{The multiple-access relay channel with correlated side information.
$(\hat{S}^n_{1}, \hat{S}^n_{2})$ are the reconstructions at the destination.}
\label{fig:MARCsideInfo}
\vspace{-0.65cm}
\end{figure}
The sources and the side information sequences, $\{ S_{1,k},S_{2,k},W_{k},W_{3,k} \}_{k=1}^{n}$, are
arbitrarily correlated at each sample index $k$, according to the joint distribution $p(s_1,s_2,w,w_3)$ defined over a
finite alphabet $\mathcal{S}_1 \times \mathcal{S}_2 \times \mathcal{W} \times \mathcal{W}_3$, and independent across different sample indices $k$.
This joint distribution is known at all nodes.
For transmission, a DM MARC with inputs $X_i \in \mathcal{X}_i, i=1,2,3$,
and outputs $Y, Y_3$ over finite output alphabets $\mathcal{Y},\mathcal{Y}_3$, respectively, is available.
The MARC is causal and memoryless in the sense of \cite{Massey:90}:
\vspace{-0.2cm}
\begin{equation}
p(y_{k},y_{3,k}|y^{k-1},y_3^{k-1},x_1^k,x_2^k,x_3^k, s_1^n, s_2^n, w_3^n, w^n) = p(y_{k},y_{3,k}|x_{1,k},x_{2,k},x_{3,k}), \quad k=1,2,\dots,n.
\label{eq:MARCchanDist}
\end{equation}
\vspace{-0.3cm}
\begin{MyDefinition}
\label{def:MABRCcodeDef}
\textnormal{ A {\em source-channel code} for the MARC with correlated side information consists of two encoding functions at the transmitters,
\vspace{-0.2cm}
\begin{equation}
f_i^{(n)} : \mathcal{S}_i^n \mapsto \mathcal{X}_i^n, \quad i=1,2,
\label{eq:MARC_encFunc}
\vspace{-0.2cm}
\end{equation}
a set of causal encoding functions at the relay, $\{ f_{3,k}^{(n)} \}_{k=1}^n$, such that
\vspace{-0.1cm}
\begin{equation}
x_{3,k} = f_{3,k}^{(n)}(y_{3,1}^{k-1},w_{3,1}^n), \quad k=1,2,\dots,n,
\label{eq:MARC_relayEncFunc}
\vspace{-0.3cm}
\end{equation}
and a decoding function at the destination
\vspace{-0.25cm}
\begin{eqnarray}
g^{(n)} &:& \mathcal{Y}^n \times \mathcal{W}^n \mapsto \mathcal{S}_1^n \times \mathcal{S}_2^n.
\label{eq:MABRC_decFunc}
\end{eqnarray}
}
\end{MyDefinition}
\vspace{-0.15cm}
\begin{MyDefinition}
\label{def:MARCpErr}
\textnormal{Let $\hat{S}_i^n, i=1,2$, denote the reconstruction of $S_i^n, i=1,2,$ respectively, at the receiver, i.e., $(\hat{S}_1^n,\hat{S}_2^n) = g^{(n)}(Y^n, W^n)$. The {\em average probability of error}, $P_e^{(n)}$, of a source-channel code for the MARC is defined as:
\vspace{-0.15cm}
\begin{eqnarray}
P_e^{(n)} & \triangleq & \Pr \Big((\hat{S}_1^n,\hat{S}_2^n) \neq (S_1^n,S_2^n) \Big) .
\label{eq:MARC_pErr}
\vspace{-0.15cm}
\end{eqnarray}}
\end{MyDefinition}
\vspace{-0.15cm}
\begin{MyDefinition}
\textnormal{
The sources $S_1$ and $S_2$ can be \emph{reliably transmitted} over the MARC with side information if there exists a sequence of source-channel codes such that $P_e^{(n)} \rightarrow 0$ as $n \rightarrow \infty$.
}
\end{MyDefinition}
\vspace{-0.4cm}
\subsection{The Primitive Semi-Orthogonal MARC} \label{subsec:PrimitiveSOMARC}
\vspace{-0.1cm}
The DM semi-orthogonal MARC (SOMARC) is a MARC in which the relay-destination link is orthogonal to the channels from the sources to the relay and the destination. Let $Y_R$ denote the signal received at the destination due to the relay channel input $X_3$, and $Y_S$ denote the signal received at the destination due to the transmission of $X_1$ and $X_2$. The conditional distribution function of the SOMARC is:
\vspace{-0.25cm}
\begin{equation}
p(y_R,y_S,y_3|x_1,x_2,x_3)=p(y_R|x_3)p(y_S,y_3|x_1,x_2).
\label{eq:SOMARC_chain}
\end{equation}
\vspace{-0.15cm}
A special case of the SOMARC, called the primitive SOMARC (PSOMARC), was considered by Tandon and Poor in \cite{Tandon:CISS:11}. In this channel the relay-destination link $X_3-Y_R$ is replaced with a finite-capacity link whose capacity is $C_3$. This model is depicted in Figure \ref{fig:PrimitiveSOMARC}.
Observe that in the PSOMARC setup there is no side-information at either the relay or destination.
\begin{figure}[ht]
\vspace{-0.1cm}
\centering
\captionsetup{font=small}
\scalebox{0.50}{\includegraphics{PrimitiveSOMARC.eps}}
\caption{Primitive semi-orthogonal multiple-access relay channel (PSOMARC).}
\label{fig:PrimitiveSOMARC}
\vspace{-0.5cm}
\end{figure}
\vspace{-0.4cm}
\subsection{Implementing JSCC via CPM} \label{subsec:CPMTechnique}
\vspace{-0.1cm}
JSCC is implemented via CPM by generating the channel inputs (codewords) statistically dependent with the source sequences, thus, the channel codewords ``preserve" some of the correlation exhibited among the sources.
For example, if two sources $(S_1,S_2)$ are to be transmitted over a MAC with channel inputs $(X_1,X_2)$, then the CPM encoded channel codewords are generated according to ${ \prod_{k=1}^{n}{p(x_{1,k}|s_{1,k})}}$.
The main benefit of the CPM technique is enlarging the set of possible joint input distributions, thereby improving the performance compared to separately constructing the source code and the channel code.
For an illustrative example we refer the reader to the example presented in \cite[pg. 649]{Cover:80}, which demonstrates the sub-optimality of separate source-channel coding, compared to the CPM technique, for the transmission of correlated sources over a DM MAC.
\vspace{-0.2cm}
\section{Previous Schemes and Motivation for a New Scheme} \label{sec:prevSchemes}
Before introducing the new coding scheme we motivate our work by briefly reviewing the two sets of sufficient conditions for reliable transmission of correlated sources over DM MARCs derived in \cite{Murin:ISIT12} and in \cite{Murin:IT11}.
\vspace{-0.2cm}
\subsection{Previously Derived Joint Source-Channel Coding Schemes for DM MARCs} \label{subsec:ExistingSchemes}
\vspace{-0.15cm}
In \cite{Murin:ISIT12} two JSCC schemes for reliable transmission of correlated sources over DM MARCs were derived. The corresponding sufficient conditions are as follows:
\begin{theorem}
\thmlabel{thm:jointCond}
(\cite[Thm. 1]{Murin:ISIT12}) A source pair $(S_1,S_2)$ can be reliably transmitted over a DM MARC with relay and receiver side information as defined in Section \ref{subsec:model} if,
\vspace{-0.2cm}
\begin{subequations} \label{bnd:Joint}
\begin{eqnarray}
H(S_1|S_2,W_3) &<& I(X_1;Y_3|S_2, V_1, X_2, X_3, W_3) \label{bnd:Joint_rly_S1} \\
H(S_2|S_1,W_3) &<& I(X_2;Y_3|S_1, V_2, X_1, X_3, W_3) \label{bnd:Joint_rly_S2} \\
H(S_1,S_2|W_3) &<& I(X_1,X_2;Y_3|V_1, V_2, X_3, W_3) \label{bnd:Joint_rly_S1S2} \\
H(S_1|S_2,W) &<& I(X_1,X_3;Y|S_1, V_2, X_2) \label{bnd:Joint_dst_S1} \\
H(S_2|S_1,W) &<& I(X_2,X_3;Y|S_2, V_1, X_1) \label{bnd:Joint_dst_S2} \\
H(S_1,S_2|W) &<& I(X_1,X_2,X_3;Y|S_1,S_2), \label{bnd:Joint_dst_S1S2}
\end{eqnarray}
\end{subequations}
\vspace{-0.15cm}
\noindent are satisfied for some joint distribution that factorizes as:
\vspace{-0.2cm}
\begin{align}
& p(s_1,s_2,w_3,w) p(v_1) p(x_1|s_1,v_1) p(v_2) p(x_2|s_2,v_2) p(x_3|v_1,v_2) p(y_3,y|x_1,x_2,x_3).
\label{eq:JntJointDist}
\vspace{-0.2cm}
\end{align}
\end{theorem}
\begin{theorem}
\thmlabel{thm:jointCondFlip}
(\cite[Thm. 2]{Murin:ISIT12}) A source pair $(S_1,S_2)$ can be reliably transmitted over a DM MARC with relay and receiver side information as defined in Section \ref{subsec:model} if,
\vspace{-0.2cm}
\begin{subequations} \label{bnd:JointFlip}
\begin{eqnarray}
H(S_1|S_2,W_3) &<& I(X_1;Y_3|S_1, X_2, X_3) \label{bnd:JointFlip_rly_S1} \\
H(S_2|S_1,W_3) &<& I(X_2;Y_3|S_2, X_1, X_3) \label{bnd:JointFlip_rly_S2} \\
H(S_1,S_2|W_3) &<& I(X_1,X_2;Y_3|S_1, S_2, X_3 ) \label{bnd:JointFlip_rly_S1S2} \\
H(S_1|S_2,W) &<& I(X_1,X_3;Y|S_2, X_2, W) \label{bnd:JointFlip_dst_S1} \\
H(S_2|S_1,W) &<& I(X_2,X_3;Y|S_1, X_1, W) \label{bnd:JointFlip_dst_S2} \\
H(S_1,S_2|W) &<& I(X_1,X_2,X_3;Y| W), \label{bnd:JointFlip_dst_S1S2}
\end{eqnarray}
\end{subequations}
\vspace{-0.15cm}
\noindent are satisfied for some joint distribution that factorizes as:
\vspace{-0.2cm}
\begin{align}
& p(s_1,s_2,w_3,w) p(x_1|s_1)p(x_2|s_2) p(x_3|s_1,s_2) p(y_3,y|x_1,x_2,x_3).
\label{eq:JntFlipJointDist}
\vspace{-0.2cm}
\end{align}
\vspace{-0.3cm}
\end{theorem}
\vspace{-0.5cm}
\begin{remark} \label{rem:oldSchemesComp}
%
\Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip} differ in both the decoding constraints and the admissible joint distribution chains, i.e., \eqref{eq:JntJointDist} and \eqref{eq:JntFlipJointDist}.
The main difference between \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip} is the target nodes for CPM and SW coding:
In \Thmref{thm:jointCond}, CPM is used for encoding information from the transmitters to the relay and SW coding is used for encoding information cooperatively from the transmitters and the relay to the destination.
Thus, in \Thmref{thm:jointCond} the cooperation between the relay and the transmitters is based on the binning information.
The RVs $V_1$ and $V_2$ in \Thmref{thm:jointCond} carry the bin indices of the SW source code.
In \Thmref{thm:jointCondFlip}, SW coding is used for encoding information from the transmitters to the relay and CPM is used for cooperatively encoding information to the destination.
Thus, in \Thmref{thm:jointCondFlip} the cooperation between the transmitters and the relay is based on the sources $S_1$ and $S_2$.
Recall that in \cite{Cover:80} it was shown that separate source and channel coding is generally suboptimal for transmitting correlated sources over MACs. Thus, it follows that the relay decoding constraints of \Thmref{thm:jointCond} are generally looser compared to the relay decoding constraints of \Thmref{thm:jointCondFlip}. Using similar reasoning we conclude that the destination decoding constraints of \Thmref{thm:jointCondFlip} are looser compared to the destination decoding constraints of \Thmref{thm:jointCond} (as long as coordination is possible, see \cite[Remark 18]{Murin:IT11}).
\end{remark}
\begin{remark}
The work \cite{SalehkalaibarAref:ISIT11} considered JSCC for the relay channel, in which one of the sources is available at the transmitter while the other is known at the relay.
The authors presented a transmission scheme similar to \Thmref{thm:jointCondFlip}, where CPM is utilized to transmit the sources from the transmitters to the destination while the relay applies binning for cooperation.
\end{remark}
\begin{remark} \label{rem:MABRC}
In the multiple-access broadcast relay channel (MABRC) \cite{Murin:IT11}, the relay also wants to reconstruct the sources in a lossless fashion. This channel model is depicted in Figure \ref{fig:MABRCsideInfo}. As both \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip} use the DF protocol, the conditions of \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip} are also sufficient conditions for reliable transmission over the MABRC.
\vspace{-0.15cm}
\begin{figure}[ht]
\centering
\captionsetup{font=small}
\scalebox{0.49}{\includegraphics{MABRC_SideInfo.eps}}
\vspace{-0.05cm}
\caption{The multiple-access broadcast relay channel with correlated side information.
$(\tilde{S}^n_{1}, \tilde{S}^n_{2})$ are the reconstructions at the relay, and $(\hat{S}^n_{1}, \hat{S}^n_{2})$ are the reconstructions at the destination.}
\label{fig:MABRCsideInfo}
\vspace{-0.6cm}
\end{figure}
\end{remark}
\vspace{-0.55cm}
\subsection{The Motivation for a New JSCC Scheme} \label{subsec:motivation}
\vspace{-0.1cm}
{\bf Motivating observation 1:}
As stated in Remark \ref{rem:oldSchemesComp}, the achievability schemes of \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip} use different combinations of the CPM technique with a SW source code paired with a channel code.
The achievability scheme of \Thmref{thm:jointCond} uses SW source coding for cooperatively encoding information from the transmitters and the relay to the destination while CPM is used for encoding information from the transmitters to the relay. In \Thmref{thm:jointCondFlip}, CPM is used for cooperatively encoding information from the transmitters and the relay to the destination while SW source coding is used for encoding information from the transmitters to the relay.
Since CPM can generally support the transmission of sources with higher entropies compared to separate source-channel coding, a natural question that arises is {\em whether the CPM technique can be used for simultaneously encoding information to both the relay and the destination}.
{\bf Motivating observation 2:}
It was observed in \cite{CoverG:79} that for the relay channel, when decoding at the relay does not constrain the rate, DF as implemented in \cite[Thm. 1]{CoverG:79} is capacity achieving . It follows that cooperation based on binning is optimal in this case.\footnote{We note that in the channel coding problem for the relay channel, other schemes, e.g. the regular encoding schemes of \cite{Carleial:82}, \cite{Willems:82}, achieve the DF-rate without binning, but these schemes are not directly applicable for this scenario, see also \cite{Murin:IT11}.} This raises the question {\em whether it is possible to construct a scheme that combines CPM from the sources to the destination with binning from the relay to the destination, and how does such a scheme compare with \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip}}.
{\bf Motivating observation 3:} The cooperative relay-broadcast channel (CRBC) model is a special case of the MABRC obtained by setting $\mathcal{S}_2 \mspace{-4mu} = \mspace{-4mu} \mathcal{X}_2 \mspace{-4mu} = \mspace{-4mu} \phi$, such that there is a single transmitter \cite{ErkipGunduz:07}.
Figure \ref{fig:CBRCsideInfo} depicts the CRBC model.
\begin{figure}[h]
\vspace{-0.35cm}
\centering
\captionsetup{font=small}
\scalebox{0.49}{\includegraphics{CBRC_SideInfo.eps}}
\vspace{-0.1cm}
\caption{The cooperative relay broadcast channel. $\tilde{S}_1^n$ and $\hat{S}_1^n$ are the reconstructions of the source sequence, $S_1^n$, at the relay and the destination, respectively.}
\label{fig:CBRCsideInfo}
\vspace{-0.5cm}
\end{figure}
\noindent For this channel model \cite{ErkipGunduz:07} presented the following necessary and sufficient conditions:
\begin{MyProposition} (\cite[Thm. 3.1]{ErkipGunduz:07})
A source $S_1$ can be reliably transmitted over a DM CRBC with relay and receiver side information if:
\vspace{-0.25cm}
\begin{subequations} \label{eq:relayJoint}
\begin{eqnarray}
H(S_1|W_3) &<& I(X_1;Y_3|X_3) \label{eq:RelayJnt_cond} \\
H(S_1|W) &<& I(X_1,X_3;Y), \label{eq:DestJnt_cond}
\end{eqnarray}
\end{subequations}
\vspace{-0.15cm}
\noindent for some input distribution $p(s_1, w_3, w)p(x_1, x_3)$. Conversely, if a source $S_1$ can be reliably transmitted over the CRBC then the conditions in \eqref{eq:RelayJnt_cond} and \eqref{eq:DestJnt_cond} are satisfied with $<$ replaced by $\leq$ for some input distribution $p(s_1, w_3, w)p(x_1, x_3)$.
\end{MyProposition}
In \cite[Remark 6]{Murin:ISIT12} it is shown that for a CRBC, the conditions of \Thmref{thm:jointCond} can be specialized to the conditions of \cite[Thm. 3.1]{ErkipGunduz:07}, while the conditions obtained from \Thmref{thm:jointCondFlip} are generally more restrictive. The reason is that when specializing \Thmref{thm:jointCondFlip} to the case of a single transmitter, the set of joint distributions of the source and relay channel inputs which satisfy \eqref{eq:JntFlipJointDist} does not exhaust the entire space of joint distributions, and in particular, does not include the optimal distribution according to \cite[Thm. 3.1]{ErkipGunduz:07}. We conclude that the downside of using CPM for encoding information to the destination, as implemented in \Thmref{thm:jointCondFlip}, is that it restricts the set of admissible joint distributions; thereby constrains the achievable coordination between the sources and the relay when cooperating to send information to the destination. This leads to the question {\em whether it is possible to construct a scheme in which CPM is used for encoding information to the destination, while the constraints on the source-relay coordination imposed by the distribution chain \eqref{eq:JntFlipJointDist} are relaxed or entirely removed}.
\noindent In the next section a new JSCC scheme is derived which gives affirmative answers to the above three questions.
\vspace{-0.1cm}
\section{A New Joint Source-Channel Coding Scheme} \label{sec:MixedJointAchiev}
\vspace{-0.1cm}
We now present a new set of sufficient conditions for reliable transmission of correlated sources over DM MARCs with side information. The achievability scheme (\Thmref{thm:jointCond_NewSimult}) is based on DF at the relay, and uses CPM for encoding information to {\em both} the relay and the destination and successive decoding at the relay. Cooperation in the new scheme is based on {\em binning implemented via SW source coding.}
The decoding method applied at the destination in the new scheme is simultaneous backward decoding of the cooperation information and the transmitted source sequences.
By combining cooperation based on binning with CPM for encoding information to the destination, the constraints on the distribution chain imposed by the scheme of \Thmref{thm:jointCondFlip} are removed.
%
Note that in the schemes implemented in \Thmref{thm:jointCond} and in \Thmref{thm:jointCondFlip} the same type of information is sent to the destination from both the relay and from the sources, while in the new scheme implemented in \Thmref{thm:jointCond_NewSimult} {\em different types} of information are sent to the destination from the relay and from the sources. This is illustrated in Figure \ref{fig:cpm_binning_combinations}. It can be observed that in \Thmref{thm:jointCond} (Figure \ref{fig:cpmRelay}) both the relay and the sources send bin indices to the destination, while in \Thmref{thm:jointCondFlip} (Figure \ref{fig:cpmDest}) both the relay and the sources send source-channel codewords. However, this is not the case in \Thmref{thm:jointCond_NewSimult} (Figure \ref{fig:cpmNew}), in which the relay sends bin indices while the sources send source-channel codewords.
\begin{figure}[ht]
\vspace{-1.0cm}
\begin{center}
\captionsetup{font=small}
\subfloat[]{\scalebox{0.31}{\includegraphics{CPM_to_Relay.eps}} \label{fig:cpmRelay}}
\subfloat[]{\scalebox{0.31}{\includegraphics{CPM_to_Dest.eps}} \label{fig:cpmDest}}
\subfloat[]{\scalebox{0.31}{\includegraphics{CPM_to_Dest_Relay_Binning.eps}} \label{fig:cpmNew}}
\vspace{-0.15cm}
\caption{Types of information sent to the destination in the schemes of (a) \Thmref{thm:jointCond}; (b) \Thmref{thm:jointCondFlip}; and (c) the new proposed scheme of \Thmref{thm:jointCond_NewSimult}. Solid arrows indicate bin indices, while dashed arrows indicate source-channel codewords.}
\vspace{-1.2cm}
\label{fig:cpm_binning_combinations}
\end{center}
\end{figure}
\vspace{-0.25cm}
\subsection{Sufficient Conditions for Simultaneous Backward Decoding at the Destination} \label{subsec:newScheme}
\vspace{-0.1cm}
Using simultaneous backward decoding the following sufficient conditions are obtained:
\begin{theorem}
\thmlabel{thm:jointCond_NewSimult}
A source pair $(S_1,S_2)$ can be reliably transmitted over a DM MARC with relay and receiver side information as defined in Section \ref{subsec:model} if the conditions
\vspace{-0.2cm}
\begin{subequations} \label{bnd:JointNewSimult}
\begin{align}
H(S_1|S_2,W_3) &< I(X_1;Y_3|S_2, V_1, X_2, X_3, W_3) \label{bnd:JointNewSimult_rly_S1} \\
H(S_2|S_1,W_3) &< I(X_2;Y_3|S_1, V_2, X_1, X_3, W_3) \label{bnd:JointNewSimult_rly_S2} \\
H(S_1,S_2|W_3) &< I(X_1,X_2;Y_3| V_1, V_2, X_3, W_3) \label{bnd:JointNewSimult_rly_S1S2} \\
H(S_1|S_2,W) &< \min \Big\{I(X_1,X_3;Y|S_2,V_2,X_2,W), \nonumber \\
& \qquad \qquad I(X_1,X_3;Y|S_1,V_2,X_2) + I(X_1;Y|S_2,V_1,X_2,X_3,W) \Big\} \label{bnd:JointNewSimult_dst_S1}
\end{align}
\begin{align}
H(S_2|S_1,W) &< \min \Big\{I(X_2,X_3;Y|S_1,V_1,X_1,W), \nonumber \\
& \qquad \qquad I(X_2,X_3;Y|S_2,V_1,X_1) + I(X_2;Y|S_1,V_2,X_1,X_3,W) \Big\} \label{bnd:JointNewSimult_dst_S2} \\
H(S_1,S_2|W) &< I(X_1,X_2,X_3;Y|W), \label{bnd:JointNewSimult_dst_S1S2}
\end{align}
\end{subequations}
\vspace{-0.2cm}
\noindent are satisfied for some joint distribution that factorizes as"
\vspace{-0.25cm}
\begin{align}
& p(s_1,s_2,w_3,w)p(v_1)p(x_1|s_1,v_1) p(v_2)p(x_2|s_2,v_2)p(x_3|v_1,v_2)p(y_3,y|x_1,x_2,x_3).
\label{eq:JntNewJointDistSimult}
\end{align}
\end{theorem}
\vspace{-0.2cm}
\begin{proof}
The proof is given in Appendix \ref{annex:jointNewSimultProof}.
\end{proof}
\vspace{-0.3cm}
\subsection{Discussion} \label{subsec:jointDiscussion}
\vspace{-0.15cm}
\begin{remark} \label{rem:SameDist}
The achievability schemes of \Thmref{thm:jointCond} and \Thmref{thm:jointCond_NewSimult} require the same joint distribution (cf. equations \eqref{eq:JntJointDist} and \eqref{eq:JntNewJointDistSimult}).
\end{remark}
\vspace{-0.15cm}
\begin{remark} \label{rem:decConstraints}
Conditions \eqref{bnd:JointNewSimult_rly_S1}--\eqref{bnd:JointNewSimult_rly_S1S2} in \Thmref{thm:jointCond_NewSimult} are constraints due to decoding at the relay, while conditions \eqref{bnd:JointNewSimult_dst_S1}--\eqref{bnd:JointNewSimult_dst_S1S2} are decoding constraints at the destination.
Note that the decoding constraints at the relay in \Thmref{thm:jointCond_NewSimult} are identical to \eqref{bnd:Joint_rly_S1}--\eqref{bnd:Joint_rly_S1S2} in \Thmref{thm:jointCond}.
\end{remark}
\begin{remark}
Note that as \Thmref{thm:jointCond_NewSimult} uses the DF scheme, the conditions of \Thmref{thm:jointCond_NewSimult} are also sufficient conditions for reliable transmission over the MABRC.
\end{remark}
%
\begin{remark} \label{rem:ExpressionsMeaning}
In \Thmref{thm:jointCond_NewSimult}, $V_1^n$ and $V_2^n$ represent the binning information for $S_1^n$ and $S_2^n$, respectively.
Consider \Thmref{thm:jointCond_NewSimult} which uses simultaneous backward decoding: condition \eqref{bnd:JointNewSimult_dst_S1} can be written as follows:
\vspace{-0.15cm}
\begin{align}
H(S_1|S_2,W) &< I(X_1;Y|S_2,V_1,X_2,X_3,W) + \nonumber \\
& \qquad \qquad \min \big\{ I(V_1, X_3;Y|S_2,V_2,X_2,W),I(X_1,X_3;Y|S_1,V_2,X_2) \big\}.
\label{eq:SimultIndivInterp}
\end{align}
\vspace{-0.15cm}
\noindent On the right-hand side (RHS) of \eqref{eq:SimultIndivInterp}, the mutual information expression $I(X_1;Y|S_2,V_1,X_2,X_3,W)$ represents the available rate for encoding information on the {\em source sequence} $S_1^n$, in excess of the bin index conveyed by the sequence $V_1^n$. This is because $S_2$, $V_1$, $X_2$, $X_3$ and $W$ are known.
The expression $I(V_1, X_3;Y|S_2,V_2,X_2,W)$ represents the rate of binning information on $S_1$ that can be utilized at the destination.
Also the expression $I(X_1,X_3;Y|S_1,V_2,X_2)$, as $S_1$ and $V_2$ are known, represents the rate for sending the bin index of the source sequence $S_1$, cooperatively from Transmitter 1 and the relay to the destination.
The reason for the two possible binning rates is that $I(V_1, X_3;Y|S_2,V_2,X_2,W)$ represents the maximal rate increase that can be achieved due to the binning information available on the current message in the backward decoding scheme, while $I(X_1,X_3;Y|S_1,V_2,X_2)$ represents the maximal rate for decoding the binning information for the next step in the backward decoding scheme.
Therefore, decoding via simultaneous backward decoding results in two constraints on the binning rate.
\end{remark}
%
%
%
\begin{remark} \label{rem:MAC}
\Thmref{thm:jointCond_NewSimult} can be specialized to the MAC with correlated sources by letting $\mathcal{V}_1= \mathcal{V}_2= \mathcal{X}_3= \mathcal{W}=\phi$. For this setting the conditions \eqref{bnd:JointNewSimult_dst_S1}--\eqref{bnd:JointNewSimult_dst_S1S2} specialize to the ones in \cite[Eqn. (12)]{Cover:80} with $Y$ as the destination.
Similarly, the MABRC, under $\mathcal{V}_1= \mathcal{V}_2= \mathcal{X}_3= \mathcal{W}_3= \mathcal{W}=\phi$, specializes to the compound MAC \cite[Section VI]{GunduzErkip:09}, and \Thmref{thm:jointCond_NewSimult} specializes to \cite[Thm. 6.1]{GunduzErkip:09}.
We conclude that \Thmref{thm:jointCond_NewSimult} implements a {\em CPM encoding for both the relay and the destination}. This is in contrast to the previous results of \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip} in which CPM is used for encoding information {\em either} to the relay {\em or} to the destination.
\end{remark}
\begin{remark} \label{rem:CRBC}
The CRBC model with correlated relay and destination side information can be obtained as a special case of the MABRC model by letting $\mathcal{X}_2=\mathcal{S}_2=\phi$. The sufficient conditions for the CRBC given in \cite[Thm. 3.1]{ErkipGunduz:07} can also be obtained from \Thmref{thm:jointCond_NewSimult} by letting $V_1=X_3$, $\mathcal{S}_2 =\mathcal{X}_2 =\mathcal{V}_2=\phi$, and considering an input distribution independent of the sources.
This is in contrast to \Thmref{thm:jointCondFlip} which specializes to more restrictive conditions (see Subsection \ref{subsec:motivation}).
We conclude that \Thmref{thm:jointCond_NewSimult} allows more flexibility in the achievable coordination between the sources and the relay compared to \Thmref{thm:jointCondFlip}.
\end{remark}
\begin{remark}
Using successive backward decoding at the destination the following sufficient conditions are obtained:
\vspace{-0.15cm}
\begin{MyProposition} \label{prop:jointCond_New}
A source pair $(S_1,S_2)$ can be transmitted reliably over a DM MARC with relay and receiver side information as defined in Section \ref{subsec:model} if,
\vspace{-0.15cm}
\begin{subequations} \label{bnd:JointNew}
\begin{align}
H(S_1|S_2,W_3) &< I(X_1;Y_3|S_2, V_1, X_2, X_3, W_3) \label{bnd:JointNew_rly_S1} \\
H(S_2|S_1,W_3) &< I(X_2;Y_3|S_1, V_2, X_1, X_3, W_3) \label{bnd:JointNew_rly_S2} \\
H(S_1,S_2|W_3) &< I(X_1,X_2;Y_3| V_1, V_2, X_3, W_3) \label{bnd:JointNew_rly_S1S2} \\
H(S_1|S_2,W) & < I(X_1;Y|S_2,V_1,X_2,X_3, W) + I(V_1,X_3;Y|V_2,W) \label{bnd:JointNew_dst_S1} \\
H(S_2|S_1,W) & < I(X_2;Y|S_1,V_2,X_1,X_3,W) + I(V_2,X_3;Y|V_1,W) \label{bnd:JointNew_dst_S2} \\
H(S_1,S_2|W) & < I(X_1,X_2;Y|V_1, V_2, X_3,W) + I(V_1,V_2,X_3;Y|W), \label{bnd:JointNew_dst_S1S2}
\end{align}
\end{subequations}
\vspace{-0.2cm}
\noindent are satisfied for some joint distribution that factorizes as:
\vspace{-0.25cm}
\begin{align}
& p(s_1,s_2,w_3,w)p(v_1)p(x_1|s_1,v_1) p(v_2)p(x_2|s_2,v_2)p(x_3|v_1,v_2)p(y_3,y|x_1,x_2,x_3).
\label{eq:JntNewJointDist}
\end{align}
\end{MyProposition}
\vspace{-0.2cm}
\begin{proof}
The proof is given in Appendix \ref{annex:jointNewProof}.
\end{proof}
\end{remark}
\begin{remark}
As the scheme of \Thmref{thm:jointCond_NewSimult} applies simultaneous backward decoding at the destination, then the source vectors and the binning information are {\em jointly} decoded (see Appendix \ref{subsec:jointNewProofDecoding}). On the other hand, the scheme of Prop. \ref{prop:jointCond_New} applies successive backward decoding at the destination, thus, first the binning information is decoded, and then, the source vectors are decoded (see Appendix \ref{annex:jointNewProof_decoding}).
Since in the latter scheme decoding the binning information uses only part of the available information, the sufficient conditions obtained for the scheme of Prop. \ref{prop:jointCond_New} are more restrictive than those obtained for the scheme of \Thmref{thm:jointCond_NewSimult} This is rigorously shown in the following section.
\end{remark}
\vspace{-0.3cm}
\section{Comparison of the Different Achievability Schemes} \label{sec:Comparison}
\vspace{-0.1cm}
We now present a detailed comparison of the sufficient conditions established by \Thmref{thm:jointCond_NewSimult}, \Thmref{thm:jointCond}, \Thmref{thm:jointCondFlip} and Prop. \ref{prop:jointCond_New}.
Specifically, we show the following:
\begin{itemize}
\item
In Subsection \ref{subsec:CompareCorrSources} we show that for correlated sources and side information the scheme of \Thmref{thm:jointCond_NewSimult} outperforms the schemes of \Thmref{thm:jointCond} and Prop. \ref{prop:jointCond_New}.
%
\item In Subsection \ref{subsec:ExampMixedOutperform} we show that there are scenarios for which the scheme of \Thmref{thm:jointCond_NewSimult} strictly outperforms the schemes of \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip}.
\end{itemize}
\vspace{-0.4cm}
\subsection{Correlated Sources and Side Information} \label{subsec:CompareCorrSources}
\vspace{-0.15cm}
We now compare \Thmref{thm:jointCond}, \Thmref{thm:jointCond_NewSimult} and Prop. \ref{prop:jointCond_New} for the general input distributions \eqref{eq:JntJointDist}, \eqref{eq:JntNewJointDistSimult} and \eqref{eq:JntNewJointDist}. As stated in Remark \ref{rem:decConstraints}, the decoding constraints at the relay in \Thmref{thm:jointCond_NewSimult} are identical to the decoding constraints at the relay in \Thmref{thm:jointCond} and Prop. \ref{prop:jointCond_New}.
Therefore, in the following we compare only the decoding constraints at the destination. The conclusion is summarized in the following proposition:
\begin{MyProposition} \label{prop:CorrSources}
The scheme of \Thmref{thm:jointCond_NewSimult} is at least as good as the schemes of \Thmref{thm:jointCond} and Prop. \ref{prop:jointCond_New}.
\end{MyProposition}
\vspace{-0.2cm}
\begin{proof}
The proof is given in Appendix \ref{annex:ProofPropCorrSources}.
\end{proof}
\begin{remark}
We emphasize that Prop. \ref{prop:CorrSources} implies that the superiority of the scheme of \Thmref{thm:jointCond_NewSimult} over the scheme of \Thmref{thm:jointCond} and the scheme of Prop. \ref{prop:jointCond_New} holds in general.
\end{remark}
Proposition \ref{prop:CorrSources} implies that
for JSCC for MARCs, simultaneous backward decoding outperforms sequential backward decoding.
For the case of separate source and channel codes, \cite[Thm. 1]{Murin:IT11} presented a separation-based achievability scheme subject to the input distribution:
\vspace{-0.25cm}
\begin{eqnarray}
p(s_1,s_2,w_3,w,v_1,v_2,x_1,x_2,x_3) = p(s_1,s_2,w_3,w)p(v_1)p(x_1|v_1)p(v_2)p(x_2|v_2)p(x_3|v_1,v_2).
\label{eq:sepDist}
\end{eqnarray}
In this case, we have $p(x_i|s_i,v_i)=p(x_i|v_i), i=1,2$, the joint distributions in \eqref{eq:JntJointDist} and \eqref{eq:JntNewJointDistSimult} specialize to the one in \eqref{eq:sepDist}, and the sufficient conditions of \Thmref{thm:jointCond} and \Thmref{thm:jointCond_NewSimult} specialize to the conditions of \cite[Thm. 1]{Murin:IT11}.
\begin{remark}
When the source and side information sequences are independent, that is $p(s_1,s_2,w_3,w)=$ $p(s_1)p(s_2)$ $p(w_3)p(w)$, the joint distributions in \eqref{eq:JntNewJointDistSimult} and \eqref{eq:JntNewJointDist} specialize to $p(s_1)p(s_2)p(w_3)p(w)p(v_1)p(x_1|v_1)p(v_2)$ $p(x_2|v_2)$ $p(x_3|v_1,v_2)$.
%
\noindent In this case, the conditions of Prop. \ref{prop:jointCond_New} specialize to the conditions obtained for sending independent messages over the MARC using sliding-window decoding at the destination \cite[Section III.B]{Sankar:07}, while the conditions of \Thmref{thm:jointCond_NewSimult} specialize to the conditions obtained for sending independent messages over the MARC using backward decoding at the destination \cite[Section III.A]{Sankar:07}.\footnote{The same observation holds when the side information is not present. This follows since when the side information is independent of the sources then it cannot help in decoding the sources. Thus, we can set $\mathcal{W} = \mathcal{W}_3 = \phi$.}
\end{remark}
%
%
%
%
%
%
\vspace{-0.35cm}
\subsection{Mixed JSCC Can Strictly Outperform the Schemes of \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip}} \label{subsec:ExampMixedOutperform}
\vspace{-0.1cm}
Recall Remark \ref{rem:SameDist}, which states that the underlying input distributions of \Thmref{thm:jointCond_NewSimult} and \Thmref{thm:jointCond} are identical, while the underlying input distribution for \Thmref{thm:jointCondFlip} is different. Here, we present a comparison of all three schemes for a special case in which the two input distribution chains are the same.
In this example the sources can be reliably transmitted by using the scheme of \Thmref{thm:jointCond_NewSimult}, while reliable transmission is not possible via the schemes of \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip}.
Consider a PSOMARC, defined by $\mathcal{X}_1 = \mathcal{X}_2 = \{0,1\}, \mathcal{Y}_3 = \{0,1,2\}, \mathcal{Y}_S = \{0,1\}$.
\noindent Let $C_3 = 1$, and consider the deterministic channel mapping $(X_1,X_2) \mapsto (Y_3,Y_S)$ specified in Table \ref{tab:PrimitiveSOMARC}.
\vspace{-0.25cm}
\begin{table}[h]
\begin{center}
\begin{tabular}[t]{|c|c|c|c|c|}
\hline
$(X_1, X_2)$ & $(0,0)$ & $(0,1)$ & $(1,0)$ & $(1,1)$ \\
\hline
$ Y_3$ & 0 & 1 & 1 & 2 \\
\hline
$ Y_S$ & 0 & 0 & 1 & 1 \\
\hline
\end{tabular}
\vspace{-0.1cm}
\captionsetup{font=small}
\caption{A deterministic channel mapping $(X_1,X_2) \mapsto (Y_3,Y_S)$ for the PSOMARC. \label{tab:PrimitiveSOMARC}}
\vspace{-0.8cm}
\end{center}
\end{table}
\vspace{-0.1cm}
\noindent The sources $(S_1,S_2)$ are defined over the sets $\mathcal{S}_1=\mathcal{S}_2=\{0,1\}$ with the joint distribution specified in Table \ref{tab:SourceDist}.
\begin{table}[h!]
\begin{center}
\begin{tabular}[t]{|c|c|c|}
\hline
$S_1$ \textbackslash $S_2$ & 0 & 1 \\
\hline
0 & 1/3 & 1/3 \\
\hline
1 & 0 & 1/3 \\
\hline
\end{tabular}
\vspace{-0.1cm}
\captionsetup{font=small}
\caption{The joint distribution of $(S_1,S_2)$. The entry in the $j^{\text{th}}$ row and $m^{\text{th}}$ column, $j,m = 0,1$, corresponds to $\Pr \left((S_1,S_2) = (j,m) \right)$. \label{tab:SourceDist}}
\vspace{-0.8cm}
\end{center}
\end{table}
\vspace{-0.2cm}
These sources can be reliably transmitted by letting $X_1 = S_1$ and $X_2 = S_2$. The probability of decoding error at the relay is zero since there is a one-to-one mapping between the channel inputs from the sources and the channel output at the relay. The probability of decoding error at the destination
can be made arbitrarily small by using the fact that each channel output at the destination corresponds only to two possible pairs of channel inputs. This ambiguity can be resolved using the relay-destination link whose capacity is 1 bit per channel use.
Next, consider the transmission
via the schemes of \Thmref{thm:jointCond}, \Thmref{thm:jointCondFlip} and \Thmref{thm:jointCond_NewSimult}.
%
%
For transmission via the schemes of \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip} we have the following proposition:
\begin{MyProposition} \label{prop:NotFeasible}
The sources defined in Table \ref{tab:SourceDist} {\em cannot} be reliably transmitted over the PSOMARC defined in Table \ref{tab:PrimitiveSOMARC}, by using the schemes of \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip}.
\end{MyProposition}
\begin{IEEEproof}
First we make the following claim:
\begin{claim} \label{clm:lowerPerr}
If an inequality sign in the conditions of \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip} is {\em reversed}, then reliable transmission is not possible with the corresponding schemes.
\end{claim}
\noindent {\em Proof sketch:} The average probability of error for decoding the sources transmitted via the scheme of \Thmref{thm:jointCond} can be {\em lower bounded} by using the properties of jointly typical sequences, \cite[Ch. 6.3]{YeungBook}. This can be done by following arguments similar to those used in \cite[Appendix B.D]{Murin:IT11}, but instead of upper bounding the different quantities in the calculation of the probability of error, we apply lower bounds, see the left-hand side (LHS) of \cite[Eqns. (6.106)--(6.108)]{YeungBook}.
In particular it follows that if conditions \eqref{bnd:Joint} hold with {\em opposite strict inequality}, e.g., $H(S_1|S_2,W_3) > I(X_1;Y_3|S_2, V_1, X_2, X_3, W_3)$, see \eqref{bnd:Joint_rly_S1}, then reliable transmission is not possible {\em via the scheme of \Thmref{thm:jointCond}}.
These arguments also apply to \Thmref{thm:jointCondFlip}, that is, if conditions \eqref{bnd:JointFlip} hold with {\em opposite strict inequality}, e.g., $H(S_1|S_2,W_3) > I(X_1;Y_3|S_1, X_2, X_3)$ , see \eqref{bnd:JointFlip_rly_S1}, then reliable transmission is not possible {\em via the scheme of \Thmref{thm:jointCondFlip}}.
In Appendix \ref{annex:proofNotFeasible} we show that indeed evaluating both \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip} for the example in this section, some conditions in \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip} hold with opposite strict inequality to what is required by the theorems.
This shows that reliable transmission of the sources is not possible via the schemes of \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip}.
\end{IEEEproof}
%
%
In contrast to \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip}, we have the following proposition for \Thmref{thm:jointCond_NewSimult}:
\begin{MyProposition} \label{prop:feasible}
The sources defined in Table \ref{tab:SourceDist} can be reliably transmitted over the PSOMARC specified in Table \ref{tab:PrimitiveSOMARC}, by using the scheme of \Thmref{thm:jointCond_NewSimult}.
\end{MyProposition}
\begin{IEEEproof}
Conditions \eqref{bnd:JointNewSimult} can be specialized to the PSOMARC
by letting $\mathcal{V}_1=\mathcal{V}_2=\mathcal{W}_3=\mathcal{W}=\phi$ and $I(X_3;Y_R)=C_3$. In particular, a specialization of the conditions of \Thmref{thm:jointCond_NewSimult} which involve $H(S_1,S_2)$, i.e. \eqref{bnd:JointNewSimult_rly_S1S2} and \eqref{bnd:JointNewSimult_dst_S1S2}, gives the following condition:
\vspace{-0.2cm}
\begin{align}
H(S_1,S_2) &< \min \{I(X_1, X_2;Y_3), I(X_1, X_2;Y_S) + C_3 \}, \label{bnd:PrimitiveSOMARC_JointNew_Sum}
\end{align}
\vspace{-0.2cm}
\noindent where the joint distribution \eqref{eq:JntNewJointDistSimult} specializes to $p(s_1,s_2)p(x_1|s_1)p(x_2|s_2)p(y_3,y_S|x_1,x_2)$. Next, note that for the sources defined in Table \ref{tab:SourceDist} we have $H(S_1,S_2) = \log_2 3$. Moreover, as $|\mathcal{Y}_3|=3,|\mathcal{Y}_S|=2$ and $C_3=1$, the RHS of \eqref{bnd:PrimitiveSOMARC_JointNew_Sum} is upper bounded by $\log_2 3$, thus, the LHS of \eqref{bnd:PrimitiveSOMARC_JointNew_Sum} equals to the RHS of \eqref{bnd:PrimitiveSOMARC_JointNew_Sum}. However, as condition \eqref{bnd:PrimitiveSOMARC_JointNew_Sum} requires strict inequality, {\em the conditions provided in the statement of \Thmref{thm:jointCond_NewSimult} do not imply that reliable transmission is possible} in the present example.
Note that this case is different than the case of Prop. \ref{prop:NotFeasible}, see Remark \ref{rem:possibleVsimpossible} below.
In Appendix \ref{annex:proofFeasible} we specify an explicit p.m.f $p(x_i|s_i),i=1,2$, for which we show, through an explicit calculation of the probability of decoding error, that reliable transmission
is possible via the scheme of \Thmref{thm:jointCond_NewSimult}.
\end{IEEEproof}
\begin{remark} \label{rem:possibleVsimpossible}
The case of Prop. \ref{prop:feasible} is different than the case of Prop. \ref{prop:NotFeasible}.
In the case of Prop. \ref{prop:feasible} we have an equality between the LHS and RHS,\footnote{Conditions \eqref{bnd:JointNewSimult}, specialized to the PSOMARC, evaluated by setting $p(x_i|s_i), i=1,2$, to be the deterministic distribution $p(x_i|s_i)=\delta(x_i-s_i)$, where $\delta(x)$ is the Kronecker Delta function, hold with an {\em equality}.}
while for Prop. \ref{prop:NotFeasible}, evaluating the conditions of \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip} we show that the inequality sign is reversed compared to what is required by the theorems. Then, in the proof of Prop. \ref{prop:NotFeasible} we show that such reversal implies that reliable transmission is impossible (see Appendix \ref{annex:proofNotFeasible}). Since in the case of Prop. \ref{prop:feasible} we have an equality between the LHS and the RHS quantities, we examine the situation in more detail in Appendix~\ref{annex:proofFeasible}.
\end{remark}
\vspace{-0.1cm}
\section{Necessary Conditions for Reliable Transmission of Correlated Sources Over DM MARCs} \label{sec:NecessaryConditions}
In this section three sets of necessary conditions for reliable transmission of correlated sources over DM MARCs with side information are derived.
These new converse results are based on the fact that only certain joint input distributions $p(x_1,x_2)$ can be achieved.
Observe that from Def. \ref{def:MABRCcodeDef} it follows that valid channel input distributions must obey the Markov chain:
\vspace{-0.15cm}
\begin{equation}
X_1 \leftrightarrow S_1^n \leftrightarrow S_2^n \leftrightarrow X_2.
\label{eq:MarkovChain}
\end{equation}
\vspace{-0.15cm}
\noindent In the following we use the technique introduced by Kang and Ulukus in \cite{Kang:2011} to constrain the achievable joint input distributions to take into account \eqref{eq:MarkovChain}. We start by reviewing some basic definitions and results from \cite{Kang:2011} and \cite{Witsenhausen:75}.
\vspace{-0.3cm}
\subsection{Definitions and Known Results}
\begin{MyDefinition} \label{def:maxCorr}
({\em Maximal correlation}, \cite[Sec. 2]{Witsenhausen:75})
The maximal correlation between the RVs $X$ and $Y$ is defined as $\rho^{\ast}_{XY} \triangleq \sup \mathds{E} \left\{ f(X) g(Y) \right\}$, where the supremum is taken over $f: \mathcal{X} \mapsto \mathfrak{R}, g: \mathcal{Y} \mapsto \mathfrak{R}$, s.t $\mathds{E} \left\{ f(X) \right\} = \mathds{E} \left\{ g(Y) \right\} = 0$, $\mathds{E} \left\{ f^2 (X) \right\} = \mathds{E} \left\{ g^2(Y) \right\} = 1$, and with the convention that the supremum over the empty set equals to 0. The conditional maximal correlation $\rho^{\ast}_{XY|z}$ is defined similarly.
\end{MyDefinition}
\begin{MyDefinition} \label{def:KangDefs}
({\em Matrix notation for probability distributions}, \cite[Eqn. (6)]{Kang:2011})
%
%
%
Let $X \in \mathcal{X}$, and $Y \in \mathcal{Y}$, be two discrete random variables with finite cardinalities. The joint probability distribution matrix $\mathds{P}_{XY}$ is defined as $\mathds{P}_{XY}(i,j) \triangleq \Pr \left( X\mspace{-2mu}=\mspace{-2mu}x_i, Y\mspace{-2mu}=\mspace{-2mu}y_j \right), i=1,2,\dots,|\mathcal{X}|, j=1,2,\dots,|\mathcal{Y}|$.
The marginal distribution matrix of an RV $X$ is defined as the diagonal matrix $\mathds{P}_X$ such that $\mathds{P}_{X}(i,i) = \Pr \left( X=x_i \right), x_i \in \mathcal{X}$; $\quad \mathds{P}_{X}(i,j)=0, \quad i \neq j$. This marginal distribution can also be represented in a vector form denoted by $\mathbf{p}_X$. The $i$'th element of $\mathbf{p}_X$ is $\mathbf{p}_X(i) \triangleq \Pr \left( X=x_i \right)$.
The conditional joint probability distribution matrix $\mathds{P}_{XY|z}$ is defined similarly.
\end{MyDefinition}
\begin{MyDefinition} \label{def:spectralRep}
({\em Spectral representation}, \cite[Eqns. (12)--(13)]{Kang:2011})
We define the matrix $\tilde{\mathds{P}}_{XY}$ as $\tilde{\mathds{P}}_{XY} \triangleq \mathds{P}^{- \frac{1}{2}}_X \mathds{P}_{XY} \mathds{P}^{- \frac{1}{2}}_Y$,
\noindent and the vector $\tilde{\mathbf{p}}_X$ as $\tilde{\mathbf{p}}_X = \mathbf{p}_X^{\frac{1}{2}}$,
\noindent where $\mathbf{p}_X^{\frac{1}{2}}$ stands for an element-wise square root of $\mathbf{p}_X$. The conditional distributions $\tilde{\mathds{P}}_{XY|z}$ and $\tilde{\mathbf{p}}_{X|y}$ are defined similarly.
\end{MyDefinition}
Note that not every matrix $\tilde{\mathds{P}}_{XY}$ can correspond to a given joint distribution matrix $\mathds{P}_{XY}$. This is because a valid joint distribution matrix $\mathds{P}_{XY}$ must have all its elements to be nonnegative and add to 1. \cite[Thm. 1]{Kang:2011} gives a necessary and sufficient condition for $\tilde{\mathds{P}}_{XY}$ to correspond to a joint distribution matrix $\mathds{P}_{XY}$:
\begin{theoremA}
(\cite[Thm. 1]{Kang:2011}) Let $\mathds{P}_X$ and $\mathds{P}_Y$ be a pair of marginal distributions. A nonnegative matrix $\mathds{P}_{XY}$ is a joint distribution matrix with marginal distributions $\mathds{P}_X$ and $\mathds{P}_Y$ if and only if the singular value decomposition (SVD) of the corresponding nonnegative matrix $\tilde{\mathds{P}}_{XY}$
%
\noindent satisfies:
\vspace{-0.15cm}
\begin{equation}
\tilde{\mathds{P}}_{XY} = \mathds{M} \mathds{D} \mathds{N}^{T} = \mathbf{p}^{\frac{1}{2}}_X \left( \mathbf{p}^{\frac{1}{2}}_Y \right)^T + \sum_{i=2}^{l}{\sigma_i \boldsymbol{\mu}_i \boldsymbol{\nu}^T_i},
\label{eq:KangThm1}
\end{equation}
\vspace{-0.15cm}
\noindent where $l = \min \{ |\mathcal{X}|, |\mathcal{Y}| \}$, $\mathds{M} \triangleq [\boldsymbol{\mu}_1, \boldsymbol{\mu}_2, \dots \boldsymbol{\mu}_l]$ and $\mathds{N} \triangleq [\boldsymbol{\nu}_1, \boldsymbol{\nu}_2, \dots \boldsymbol{\nu}_l]$ are two matrices such that $\mathds{M}^T \mathds{M} = \mathds{I}$ and $\mathds{N}^T \mathds{N}=\mathds{I}$, and $\mathds{D} \triangleq \text{diag}[\sigma_1, \sigma_2, \dots, \sigma_l]$\footnote{We use $\mathds{D} = \text{diag}[\mathbf{a}]$ to denote a rectangular matrix $\mathds{D}$ s.t $\mathds{D}_{i,i} = a_i, \mathds{D}_{i,j} = 0, \forall i \ne j$.
}; $\boldsymbol{\mu}_1 = \mathbf{p}^{\frac{1}{2}}_X, \boldsymbol{\nu}_1 = \mathbf{p}^{\frac{1}{2}}_Y$, and $\sigma_1 = 1 \ge \sigma_2 \ge \dots \ge \sigma_l \ge 0$. That is, all the singular values of $\tilde{\mathds{P}}_{XY}$ are non-negative and smaller than or equal to 1.
We sometime denote $\sigma_i = \sigma_i(\tilde{\mathds{P}}_{X Y})$ to explicitly indicate the matrix for which the singular value is computed.
The largest singular value of $\tilde{\mathds{P}}_{XY}$ is 1, and its corresponding left and right singular vectors are $\mathbf{p}^{\frac{1}{2}}_X$ and $\mathbf{p}^{\frac{1}{2}}_Y$.
\end{theoremA}
Next, we define the set of all possible conditional distributions $p(x_1,\mspace{-1mu} x_2|s_{1,1}, \mspace{-1mu} s_{2,1} \mspace{-1mu} )$ satisfying the Markov chain~\eqref{eq:MarkovChain}:
\begin{align}
\mathcal{B}_{X_1 X_2|S_1 S_2} \triangleq \begin{Bmatrix*}[l] & p_{X_1,X_2|S_1,S_2}(x_1,x_2|s_{1,1}, s_{2,1}): \nonumber \\
& \exists n \in \mathfrak{N}^{+}, p_{X_1|S_1^n}(x_1|s_1^n), p_{X_2|S_2^n}(x_2|s_2^n) \nonumber \\
& \text{s.t. } \forall (x_1,x_2,s_{1,1},s_{2,1}) \in \mathcal{X}_1 \times \mathcal{X}_2 \times \mathcal{S}_1 \times \mathcal{S}_2, \nonumber \\
& p_{X_1,X_2|S_1,S_2}(x_1,x_2|s_{1,1}, s_{2,1}) = \nonumber \\
& \quad \frac{1}{p_{S_1,S_2}(s_{1,1}, s_{2,1})} {\displaystyle \sum_{\substack{s_{1,2}^n \in \mathcal{S}_1^{n-1} \\ s_{2,2}^n \in \mathcal{S}_2^{n-1}}}{p_{X_1|S_1^n}(x_1|s_1^n)p_{X_2|S_2^n}(x_2|s_2^n)p_{S_1^n,S_2^n}(s_1^n, s_2^n)}} \end{Bmatrix*},
\label{eq:B_Set_Def}
\end{align}
\noindent where ${p_{S_1^n,S_2^n}(s_1^n, s_2^n) = \prod_{k=1}^{n}{p_{S_1,S_2}(s_{1,k}, s_{2,k})}}$.
Note that as $n$ can be arbitrarily large, the set of all conditional distributions $p_{X_1|S_1^n}(x_1|s_1^n)$ and $p_{X_2|S_2^n}(x_2|s_2^n)$, for all positive integers $n$, is countably infinite. Therefore, we are interested in a characterization of the $n$-letter Markov chain \eqref{eq:MarkovChain} via a set which has a {\em bounded and finite cardinality}.
In order to achieve this, we first note that as $p_{S_1,S_2}(s_{1,1},s_{2,1})$ is given, $p_{X_1,X_2}(x_1,x_2), p_{X_1,X_2|S_1}(x_1,x_2|s_{1,1})$ and $p_{X_1,X_2|S_2}(x_1,x_2|s_{2,1})$ are all uniquely determined by $p_{X_1,X_2|S_1,S_2}(x_1,x_2|s_{1,1}, s_{2,1})$.
Furthermore, in \cite[Sec. 4]{Witsenhausen:75} it is shown that $\sigma_2(\tilde{\mathds{P}}_{X_1 X_2}) = \rho^{\ast}_{X_1 X_2}$.
Therefore, $\rho^{\ast}_{X_1 X_2}$, $\rho^{\ast}_{X_1 X_2| s_{1,1}}, \rho^{\ast}_{X_1 X_2| s_{2,1}}$ and $\rho^{\ast}_{X_1 X_2| s_{1,1}, s_{2,1}}$
are all functions of $p_{X_1,X_2|S_1,S_2}(x_1,x_2|s_{1,1}, s_{2,1})$ for a given $p_{S_1,S_2}(s_{1,1}, s_{2,1})$.
The following theorem characterizes constraints on these maximal correlations, and thereby gives a necessary condition for the $n$-letter Markov chain~\eqref{eq:MarkovChain}:\footnote{Here we present a simplified version of \cite[Thm. 4]{Kang:2011}.}
\begin{theoremA}
(\cite[Thm. 4]{Kang:2011}) Let $(S_1^n,S_2^n)$ be a pair of length-$n$ independent and identically distributed (i.i.d.) sequences such that $p_{S_{1,k},S_{2,k}}(a,b) = p_{S_{1},S_{2}}(a,b), \forall (a,b) \in \mathcal{S}_1 \times \mathcal{S}_2, \forall k \in \{1,2,\dots,n \}$, and let the variables $X_1$ and $X_2$ satisfy the Markov chain \eqref{eq:MarkovChain}. Let $S_{1,k}$ and $S_{2,j}$ be arbitrary elements of $\mathbf{S}_{1,1}^n$ and $\mathbf{S}_{2,1}^n$, respectively, that is, $k,j \in \left\{1,2,\dots,n \right\}$, then
\vspace{-0.15cm}
\begin{equation}
\rho^{\ast}_{X_1 X_2| s_{1,k}, s_{2,k}} \le \rho^{\ast}_{S_1 S_2}.
\end{equation}
\end{theoremA}
\vspace{-0.15cm}
Now, we define the set $\mathcal{B}_{X_1 X_2|S_1 S_2}'$ as follows:
\begin{align}
\mathcal{B}_{X_1 X_2|S_1 S_2}' \triangleq \begin{Bmatrix*}[l] & p_{X_1,X_2|S_1,S_2}(x_1,x_2|s_{1,1}, s_{2,1}): \nonumber \\
& \forall (s_{1,1}, s_{2,1}) \in \mathcal{S}_1 \times \mathcal{S}_2 \\
& \rho^{\ast}_{X_1 X_2} \le \rho^{\ast}_{S_1 S_2}, \nonumber \\
& \rho^{\ast}_{X_1 X_2| s_{1,1}} \le \rho^{\ast}_{S_1 S_2}, \nonumber \\
& \rho^{\ast}_{X_1 X_2| s_{2,1}} \le \rho^{\ast}_{S_1 S_2}, \nonumber \\
& \rho^{\ast}_{X_1 X_2| s_{1,1}, s_{2,1}} \le \rho^{\ast}_{S_1 S_2} \end{Bmatrix*}.
\end{align}
\noindent Note that by \cite[Thm. 4]{Kang:2011} the set $\mathcal{B}_{X_1 X_2|S_1 S_2}'$ is invariant to the symbol index, that is, $s_{1,1}$ and $s_{2,1}$ can be replaced by $s_{1,k}$ and $s_{2,k}$ for any $k \in \{2,3,\dots,n \}$.
\noindent Since \cite[Thm. 4]{Kang:2011} gives necessary conditions for the $n$-letter Markov chain \eqref{eq:MarkovChain}, it follows that $\mathcal{B}_{X_1 X_2|S_1 S_2} \subseteq \mathcal{B}_{X_1 X_2|S_1 S_2}'$.
Furthermore, the set $\mathcal{B}_{X_1 X_2|S_1 S_2}'$ is characterized by the singular values\footnote{Recall that $\sigma_2(\tilde{\mathds{P}}_{X_1 X_2}) = \rho^{\ast}_{X_1 X_2}$.} of the matrices $\tilde{\mathds{P}}_{X_1 X_2}, \tilde{\mathds{P}}_{X_1 X_2| s_{1,1}}, \tilde{\mathds{P}}_{X_1 X_2| s_{2,1}}$ and $\tilde{\mathds{P}}_{X_1 X_2| s_{1,1},s_{2,1}}$. Therefore, while the set $\mathcal{B}_{X_1 X_2|S_1 S_2}$ has countably infinite dimensions, the set $\mathcal{B}_{X_1 X_2|S_1 S_2}'$ has finite and bounded dimensions.
\vspace{-0.25cm}
\subsection{A MAC Bound}
Next, we derive a new set of necessary conditions which is a reminiscent of the so-called ``MAC bound'' for the relay channel, \cite[Ch. 16]{KimElGamal:12}, that takes into account \eqref{eq:MarkovChain}.
\begin{theorem}
\thmlabel{thm:OuterMarkov}
Any source pair $(S_1,S_2)$ that can be reliably transmitted over the DM MARC with receiver side information $W$, as defined in Section \ref{subsec:model}, must satisfy the constraints:
\vspace{-0.15cm}
\begin{subequations} \label{bnd:outr_markov_dst}
\begin{eqnarray}
H(S_1|S_2,W) &\leq& I(X_1,X_3;Y|S_2,X_2,W,Q) \label{bnd:outr_markov_dst_S1} \\
H(S_2|S_1,W) &\leq& I(X_2,X_3;Y|S_1,X_1,W,Q) \label{bnd:outr_markov_dst_S2} \\
H(S_1,S_2|W) &\leq& I(X_1,X_2,X_3;Y|W,Q), \label{bnd:outr_markov_dst_S1S2}
\end{eqnarray}
\end{subequations}
\vspace{-0.15cm}
\noindent for a joint distribution that factorizes as:
\vspace{-0.15cm}
\begin{align}
p(q,s_1,s_2,w,x_1,x_2,x_3,y) = p(q)p(s_1,s_2,w)p(x_1,x_2|s_1,s_2,q)p(x_3|x_1,x_2,s_1,s_2,q)p(y|x_1,x_2,x_3),
\label{eq:MarkovBound_dist}
\end{align}
\vspace{-0.15cm}
\noindent with $\left| \mathcal{Q} \right| \leq 4$, and for every $q \in \mathcal{Q}$, it follows that:
\vspace{-0.15cm}
\begin{align}
p(x_1,x_2|s_1,s_2,Q=q) & \in \mathcal{B}_{X_1 X_2|S_1 S_2} \subseteq \mathcal{B}_{X_1 X_2|S_1 S_2}'.
\label{eq:MarkovBound_dist_Bset}
\end{align}
\end{theorem}
\vspace{-0.2cm}
\begin{IEEEproof}
The proof is given in Appendix \ref{subsec:MarkovBound_proof}.
\end{IEEEproof}
\begin{remark}
This bound does not include $W_3$ because decoding is done based only on the information available at the destination, while the relay channel input is allowed to depend on $X_1,X_2,S_1$ and $S_2$. Therefore, $W_3$ does not add any useful information for generating the relay channel input.
\end{remark}
%
%
\vspace{-0.4cm}
\subsection{Broadcast Bounds} \label{sec:MAC_BC_bound}
The next two new sets of necessary conditions are a reminiscent of the so-called ``broadcast bound'' for the relay channel, \cite[Ch. 16]{KimElGamal:12}.
\begin{MyProposition}
\label{prop:VGeneral}
Any source pair $(S_1,S_2)$ that can be reliably transmitted over the DM MARC with relay side information $W_3$ and receiver side information $W$, as defined in Section \ref{subsec:model}, must satisfy the constraints:
\vspace{-0.2cm}
\begin{subequations} \label{bnd:V_general_dst}
\begin{align}
H(S_1|S_2,W,W_3) & \leq I(X_1;Y,Y_3|S_2,X_2,W,V) \label{bnd:outr_V_dst_S1} \\
H(S_2|S_1,W,W_3) & \leq I(X_2;Y,Y_3|S_1,X_1,W,V) \label{bnd:outr_V_dst_S2} \\
H(S_1,S_2|W,W_3) & \leq I(X_1,X_2;Y,Y_3|W,V), \label{bnd:outr_V_dst_S1S2}
\end{align}
\end{subequations}
\vspace{-0.15cm}
\noindent for some joint distribution of the form:
\vspace{-0.15cm}
\begin{align}
& p(v,s_1,s_2,w,w_3,x_1,x_2,x_3,y,y_3) = p(v,s_1,s_2,w,w_3)p(x_1,x_2|s_1,s_2,v)p(x_3|v)p(y,y_3|x_1,x_2,x_3), \label{eq:BCbound_dist}
\end{align}
\vspace{-0.15cm}
\noindent with $\left| \mathcal{V} \right| \leq 4$.
\end{MyProposition}
\vspace{-0.2cm}
\begin{IEEEproof}
The proof is given in Appendix \ref{annex:ProofOuterV}.
\end{IEEEproof}
\begin{remark} \label{rem:outerVImprove}
In Prop. \ref{prop:VGeneral} we did not place restrictions on $p(x_1,x_2|s_1,s_2)$ as in \Thmref{thm:OuterMarkov}. This is because \cite[Thm. 4]{Kang:2011} requires $(S_1^n, S_2^n)$ to be a pair of {\em i.i.d sequences} of length $n$. However, in the proof of Prop. \ref{prop:VGeneral} $V^n$ is {\em not} an {\em i.i.d sequence}, and therefore $(S_1^n, S_2^n, V^n)$ is {\em not} a triplet of {\em i.i.d sequences}. Hence, it is not possible to use the approach of \cite{Kang:2011} to tighten Prop. \ref{prop:VGeneral}. It is possible, however, to establish a different set of ``broadcast-type" necessary conditions which benefits from the results of \cite{Kang:2011}. This is stated in \Thmref{thm:BCMarkov}.
\end{remark}
\begin{theorem}
\thmlabel{thm:BCMarkov}
Any source pair $(S_1,S_2)$ that can be reliably transmitted over the DM MARC with relay side information $W_3$ and receiver side information $W$, as defined in Section \ref{subsec:model}, must satisfy the constraints:
\vspace{-0.15cm}
\begin{subequations} \label{bnd:BC_markov_dst}
\begin{eqnarray}
H(S_1|S_2,W,W_3) &\leq& I(X_1;Y,Y_3|S_2,X_2,X_3,W,Q) \label{bnd:BC_markov_dst_S1} \\
H(S_2|S_1,W,W_3) &\leq& I(X_2;Y,Y_3|S_1,X_1,X_3,W,Q) \label{bnd:BC_markov_dst_S2} \\
H(S_1,S_2|W,W_3) &\leq& I(X_1,X_2;Y, Y_3|X_3,W,Q), \label{bnd:BC_markov_dst_S1S2}
\end{eqnarray}
\end{subequations}
\vspace{-0.15cm}
\noindent for a joint distribution that factorizes as:
\vspace{-0.2cm}
\begin{align}
& p(q,s_1,s_2,w,w_3,x_1,x_2,x_3,y,y_3) = \nonumber \\
& \qquad p(q)p(s_1,s_2,w,w_3)p(x_1,x_2|s_1,s_2,q)p(x_3|x_1,x_2,w_3,q)p(y,y_3|x_1,x_2,x_3),
\label{eq:BCMarkovBound_dist}
\end{align}
\vspace{-0.2cm}
\noindent with $\left| \mathcal{Q} \right| \leq 4$, and for every $q \in \mathcal{Q}$, it follows that:
\vspace{-0.2cm}
\begin{align}
p(x_1,x_2|s_1,s_2,Q=q) & \in \mathcal{B}_{X_1 X_2|S_1 S_2} \subseteq \mathcal{B}_{X_1 X_2|S_1 S_2}',
\label{eq:BCMarkovBound_dist_Bset}
\end{align}
\end{theorem}
\vspace{-0.2cm}
\begin{IEEEproof}
The proof follows similar arguments to the proofs of \Thmref{thm:OuterMarkov} and Prop. \ref{prop:VGeneral}, thus, it is omitted here.
\end{IEEEproof}
\vspace{-0.4cm}
\subsection{Discussion}
%
\begin{remark}
Note that the side information may affect the corresponding chain, see e.g., \Thmref{thm:BCMarkov}.
\end{remark}
\begin{remark}
For independent sources ($p(s_1,s_2)=p(s_1)p(s_2)$) and $\mathcal{W} = \mathcal{W}_3 = \phi$, a combination of \Thmref{thm:OuterMarkov} and \Thmref{thm:BCMarkov} specializes to the cut-set bound for the MARC derived in \cite[Thm. 1]{KramerMandayam:04}. To see this, note that in this case the RHSs of \eqref{bnd:BC_markov_dst} are identical to the first term in the RHS of \cite[Eqn. (7)]{KramerMandayam:04}, while the RHSs of \eqref{bnd:outr_markov_dst} are identical to the second term in the RHS of \cite[Eqn. (7)]{KramerMandayam:04}, for $G=\{1\}, \{2\}, \{1,2\}$, respectively. Furthermore, we have that \eqref{eq:MarkovBound_dist} and \eqref{eq:BCMarkovBound_dist} are the same. Next, note that for independent sources, $\rho^{\ast}_{S_1 S_2} = 0$, which implies that $\rho^{\ast}_{X_1 X_2}= \rho^{\ast}_{X_1 X_2| s_{1,1}} = \rho^{\ast}_{X_1 X_2| s_{2,1}} = \rho^{\ast}_{X_1 X_2| s_{1,1}, s_{2,1}} = 0$. Therefore, $X_1$ and $X_2$ are independent and conditions \eqref{eq:MarkovBound_dist_Bset} and \eqref{eq:BCMarkovBound_dist_Bset} are satisfied for any $p_{S_1,S_2}(s_1,s_2)=p_{S_1}(s_1)p_{S_2}(s_2)$.
Finally, letting $R_1 \triangleq H(S_1), R_2 \triangleq H(S_2)$ implies that $H(S_1,S_2)=R_1+R_2$, and therefore for independent sources the combination of \Thmref{thm:OuterMarkov} and \Thmref{thm:BCMarkov} coincides with \cite[Eqn. (7)]{KramerMandayam:04}.
\end{remark}
\begin{remark}
For Gaussian MARCs subject to i.i.d phase fading, and for the channel inputs that maximize the achievable region at the destination obtained via DF, the achievable region at the destination is a subset of the corresponding achievable region at the relay (i.e., decoding at the relay does not constrain the rate to the destination). In this case, \Thmref{thm:OuterMarkov} specializes to \cite[Prop. 1]{Murin:ISWCS11}.\footnote{In \cite[Thm. 4]{Murin:IT11} we showed that for Gaussian MARCs subject to i.i.d phase fading, when decoding at the relay does not constrain the rate to the destination, then source-channel separation is optimal.} From \cite[Thm. 8]{Kramer:2005} it follows that in this case mutually independent channel inputs simultaneously maximize the RHSs of \cite[Eqns. (3)]{Murin:ISWCS11}. Additionally, note that for mutually independent channel inputs, Eqns. \eqref{bnd:outr_markov_dst} coincide with \cite[Eqns. (3)]{Murin:ISWCS11}. Lastly we observe that the mutual independence of the channel inputs implies that $\rho^{\ast}_{X_1 X_2}= \rho^{\ast}_{X_1 X_2| s_{1,1}} = \rho^{\ast}_{X_1 X_2| s_{2,1}} = \rho^{\ast}_{X_1 X_2| s_{1,1}, s_{2,1}} = 0$, thus \eqref{eq:MarkovBound_dist_Bset} is satisfied for any joint distribution of the sources.
\end{remark}
\begin{remark}
When specialized to the MAC with correlated sources \Thmref{thm:OuterMarkov} and \Thmref{thm:BCMarkov} coincide and both are tighter than Prop. \ref{prop:VGeneral}.
Setting $\mathcal{X}_3 = \mathcal{Y}_3 = \mathcal{W}_3 = \phi$,
the expressions in \eqref{bnd:outr_markov_dst}, \eqref{bnd:V_general_dst} and \eqref{bnd:BC_markov_dst} become identical.
However, note that in \eqref{eq:BCbound_dist} a general joint distribution $p(v,s_1,s_2,w)$ is considered, while in \eqref{eq:MarkovBound_dist} and \eqref{eq:BCMarkovBound_dist} $Q \independent (S_1,S_2,W)$.
Moreover, the required Markov chain of \eqref{eq:MarkovChain} is not accounted for by the chain of Prop. \ref{prop:VGeneral}, contrary to \Thmref{thm:OuterMarkov} and \Thmref{thm:BCMarkov}.
Therefore, we conclude that when specialized to the MAC scenario, \Thmref{thm:OuterMarkov} and \Thmref{thm:BCMarkov} give the same bound which is tighter then the one in Prop. \ref{prop:VGeneral}.
Setting $\mathcal{X}_3 = \mathcal{Y}_3 = \mathcal{W}_3 = \phi$ as well as $\mathcal{W} = \phi$, specializes our model to the MAC with no side information at the receiver. For this model, both \Thmref{thm:OuterMarkov} and \Thmref{thm:BCMarkov} specialize to \cite[Thm. 7]{Kang:2011}, which establishes necessary conditions for the MAC with correlated sources.
\end{remark}
%
%
%
%
%
\vspace{-0.4cm}
\subsection{Numerical Examples}
\vspace{-0.1cm}
We now demonstrate the improvement of \Thmref{thm:OuterMarkov} and \Thmref{thm:BCMarkov} upon the cut-set bound of \cite[Ch. 18.1]{KimElGamal:12}.
In order to simplify the arguments, we consider a scenario with no side information $\mathcal{W}=\mathcal{W}_3=\phi$, and focus on the bound on $H(S_1,S_2)$.
In the following, we consider explicit PSOMARC and sources for which we show that the cut-set bound fails to indicate whether reliable transmission of the sources over the channel is possible, while a relaxed version of our outer bounds do indicate that reliable transmission of the sources over the channel is impossible.
Consider the PSOMARC defined by $\mathcal{X}_1 = \mathcal{X}_2 = \mathcal{Y}_3 = \mathcal{Y}_S = \{0,1\}$, the channel transition probabilities detailed in Tables \ref{tab:TranProbY3} and \ref{tab:TranProbY}, and let $C_3 = 0.1$.
\begin{table}[h]
\vspace{-0.25cm}
\begin{center}
\begin{tabular}[t]{|c|c|c|c|c|}
\hline
$Y_3$ \textbackslash $(X_1,X_2)$ & (0,0) & (0,1) & (1,0) & (1,1) \\
\hline
0 & 0.87 & 0.25 & 0.51 & 0.24 \\
\hline
1 & 0.13 & 0.75 & 0.49 & 0.76 \\
\hline
\end{tabular}
\captionsetup{font=small}
\caption{The transition probability $(X_1,X_2) \mapsto Y_3$. \label{tab:TranProbY3}}
\vspace{-0.6cm}
\end{center}
\end{table}
\begin{table}[h]
\vspace{-0.5cm}
\begin{center}
\begin{tabular}[t]{|c|c|c|c|c|}
\hline
$Y$ \textbackslash $(X_1,X_2)$ & (0,0) & (0,1) & (1,0) & (1,1) \\
\hline
0 & 0.23 & 0.19 & 0.65 & 0.91 \\
\hline
1 & 0.77 & 0.81 & 0.35 & 0.09 \\
\hline
\end{tabular}
\captionsetup{font=small}
\caption{The transition probability $(X_1,X_2) \mapsto Y$. \label{tab:TranProbY}}
\vspace{-1cm}
\end{center}
\end{table}
\noindent Next, consider the cut-set bound for the sum-rate of the PSOMARC, \cite[Eqn. (9)]{Tandon:CISS:11}. When evaluated for the PSOMARC defined in Tables \ref{tab:TranProbY3}, \ref{tab:TranProbY} the necessary conditions of \cite[Eqn. (9)]{Tandon:CISS:11} yield:
\vspace{-0.1cm}
\begin{equation}
H(S_1,S_2) \le \text{I}_{\text{cut-set}} \triangleq \max_{p(x_1,x_2)} \Big\{ I(X_1,X_2;Y_S) + \min \big\{ C_3, I(X_1,X_2;Y_3|Y_S) \big\} \Big\} \approx 0.516.\footnote{Note that the cut-set bound in \eqref{eq:Example_cutset} depends only on the channel transition probabilities and {\em not} on the joint distribution of the sources.}
\label{eq:Example_cutset}
\end{equation}
\vspace{-0.1cm}
\noindent The ~maximum ~in ~\eqref{eq:Example_cutset} ~is ~achieved ~by ~$\Pr \left((X_1,X_2)=(0,0) \right) \approx 0.1$, $\Pr \left((X_1,X_2)=(0,1) \right) \approx 0.39$, $\Pr \left((X_1,X_2)=(1,0) \right) \approx 0$, $\Pr \left((X_1,X_2)=(1,1) \right) \approx 0.51$. This and the following optimizations are done numerically using an exhaustive search over all relevant parameters with a step size of 0.01 in each variable.
Next, we consider the combination of the relaxed versions of \eqref{bnd:outr_markov_dst_S1S2} and \eqref{bnd:BC_markov_dst_S1S2}, with $\mathcal{W}=\mathcal{W}_3=\phi$, specialized to the PSOMARC:
\vspace{-0.25cm}
\begin{equation}
H(S_1,S_2) \le \text{I}_{\text{new}} \triangleq \max_{p(x_1,x_2): \rho^{\ast}_{X_1 X_2} \le \rho^{\ast}_{S_1 S_2}} \Big\{ I(X_1,X_2;Y_S) + \min \big\{ C_3, I(X_1,X_2;Y_3|Y_S) \big\} \Big\}.
\label{eq:Example_markov}
\end{equation}
\vspace{-0.05cm}
\noindent Note that \eqref{eq:Example_markov} is less restrictive than \eqref{bnd:outr_markov_dst_S1S2} and \eqref{bnd:BC_markov_dst_S1S2}, as the maximization in \eqref{eq:Example_markov} includes only the restriction due to $\tilde{\mathds{P}}_{X_1 X_2}$, while the restrictions due to the conditional distributions $\tilde{\mathds{P}}_{X_1 X_2|S_1}, \tilde{\mathds{P}}_{X_1 X_2|S_2}$ and $\tilde{\mathds{P}}_{X_1 X_2|S_1,S_2}$ are ignored.
Finally, we recall the sum-rate condition of \Thmref{thm:jointCond_NewSimult} stated in \eqref{bnd:PrimitiveSOMARC_JointNew_Sum} obtained by combining \eqref{bnd:JointNewSimult_rly_S1S2} and \eqref{bnd:JointNewSimult_dst_S1S2} and specializing the expressions to the PSOAMRC:
\vspace{-0.15cm}
\begin{equation}
H(S_1,S_2) < \text{I}_{\text{suff}} \triangleq \max_{ p(s_1,s_2)p(x_1|s_1)p(x_2|s_2) } \min \big\{I(X_1,X_2;Y_3), I(X_1,X_2;Y_S) + C_3 \big\}.
\label{eq:Example_suff}
\end{equation}
\vspace{-0.15cm}
Let $(S_1,S_2)$ be a pair of sources such that $\mathcal{S}_1 = \mathcal{S}_2 = \{0,1\}$, and their joint distribution is given in Table \ref{tab:SourceDist_p2}.
\begin{table}[h!]
\vspace{-0.25cm}
\begin{center}
\begin{tabular}[t]{|c|c|c|}
\hline
$S_1$ \textbackslash $S_2$ & 0 & 1 \\
\hline
0 & 0 & 0.04 \\
\hline
1 & 0.045 & 0.915 \\
\hline
\end{tabular}
\captionsetup{font=small}
\caption{
The joint distribution $p(s_1,s_2)$.
\label{tab:SourceDist_p2}}
\vspace{-1cm}
\end{center}
\end{table}
\noindent For this joint distribution we evaluate $H(S_1,S_2) \approx 0.504$, therefore, the cut-set necessary condition \eqref{eq:Example_cutset} does not indicate whether these sources can be transmitted reliably or not.
Furthermore, for the joint distribution given in Table \ref{tab:SourceDist_p2}, the RHS of \eqref{eq:Example_suff} is evaluated as $\text{I}_{\text{suff}} \approx 0.274$.
This value is achieved by $\Pr \left(X_1=0|S_1=0 \right) \approx 0$, $\Pr \left(X_1=0|S_1=1 \right) \approx 1$, $\Pr \left(X_1=1|S_1=0 \right) \approx 0.84$, $\Pr \left(X_1=1|S_1=1 \right) \approx 0.16$, $\Pr \left(X_2=0|S_2=0 \right) \approx 0.98$, $\Pr \left(X_2=0|S_2=1 \right) \approx 0.02$, $\Pr \left(X_2=1|S_2=0 \right) \approx 0.49$, $\Pr \left(X_2=1|S_2=1 \right) \approx 0.51$. Thus, the scheme of \Thmref{thm:jointCond_NewSimult} cannot transmit these sources reliably since condition \eqref{eq:Example_suff} is not satisfied.
In contrast to \eqref{eq:Example_cutset}, which is larger than $H(S_1,S_2)$, for the joint distribution given in Table \ref{tab:SourceDist_p2} we have $\text{I}_{\text{new}} \approx 0.485$. This value is
achieved by $\Pr \left((X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(0,0) \right) \approx 0.08$, $\Pr \left((X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(0,1) \right) \approx 0.41$, $\Pr \left((X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(1,0) \right) \approx 0.07$, $\Pr \left((X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(1,1) \right) \approx 0.44$.
Hence, our new necessary condition \eqref{eq:Example_markov}, explicitly indicates that reliable transmission of these sources is impossible.
%
This demonstrates the improvement of \Thmref{thm:OuterMarkov} and \Thmref{thm:BCMarkov} upon the cut-set bound.
\begin{remark}
This numerical example {\em does not follow immediately from the results of Kang and Ulukus for the MAC}, detailed in \cite[Subsection III.C]{Kang:2011}. To see this, consider the PSOMARC and sources as defined in Tables \ref{tab:TranProbY3}, \ref{tab:TranProbY} and \ref{tab:SourceDist_p2}, and let $C_3 = 0.2$ (instead of $0.1$).
Here, \eqref{eq:Example_cutset} is evaluated as $\text{I}_{\text{cut-set}} \approx 0.600$\footnote{This value was found via an exhaustive search over over all $p(x_1,x_2)$ and can be achieved by $\Pr \left((X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(0,0) \right) \approx 0.26$, $\Pr \left((X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(0,1) \right) \approx 0.24$, $\Pr \left((X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(1,0) \right) \approx 0$, $\Pr \left((X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(1,1) \right) \approx 0.5$.}, while \eqref{eq:Example_markov} is evaluated as $\text{I}_{\text{new}} \approx 0.514$\footnote{This value was found via an exhaustive search over over all $p(x_1,x_2)$ s.t $\rho^{\ast}_{X_1 X_2} \le \rho^{\ast}_{S_1 S_2}$, and can be achieved by $\Pr \left( (X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(0,0) \right) \approx 0.2$, $\Pr \left( (X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(0,1) \right) \approx 0.36$, $\Pr \left( (X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(1,0) \right) \approx 0.14$, $\Pr \left( (X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(1,1) \right) \approx 0.3$.}. Moreover, recall that $H(S_1,S_2) \approx 0.504$. Hence, for $C_3 = 0.2$, \eqref{eq:Example_markov} does not indicate whether reliable transmission of the sources is possible, while for $C_3 = 0.1$, \eqref{eq:Example_markov} explicitly indicates that reliable transmission is impossible. Observe that the necessary conditions are affected by the presence of the relay. Also note that the cut-set conditions \eqref{eq:Example_cutset} does not indicate whether reliable transmission is possible or not, for either value of $C_3$.
\end{remark}
\begin{remark}
In the above numerical example we assume that side information is not present. To see the effect of side information at the relay on \eqref{eq:Example_markov} consider the PSOMARC and sources as defined in Tables \ref{tab:TranProbY3}, \ref{tab:TranProbY} and \ref{tab:SourceDist_p2}, and let $C_3 = 0.5$. Here, $I(X_1,X_2;Y_2|Y_S) \approx 0.185, I(X_1,X_2;Y_S) \approx 0.329$ and $\text{I}_{\text{new}} \approx 0.514$\footnote{These value were found via an exhaustive search over over all $p(x_1,x_2)$ s.t $\rho^{\ast}_{X_1 X_2} \le \rho^{\ast}_{S_1 S_2}$, and can be achieved by $\Pr \left( (X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(0,0) \right) \approx 0.04$, $\Pr \left( (X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(0,1) \right) \approx 0.46$, $\Pr \left( (X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(1,0) \right) \approx 0.03$, $\Pr \left( (X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(1,1) \right) \approx 0.47$.}. Therefore, in this case $I(X_1,X_2;Y_2|Y_S)$ is the dominant term in the minimization on the RHS of \eqref{eq:Example_markov}. Now, let $W_3 = (S_1,S_2)$, which makes \eqref{bnd:BC_markov_dst_S1S2} redundant.\footnote{When $W_3 = (S_1,S_2)$ the chains \eqref{eq:MarkovBound_dist} and \eqref{eq:BCMarkovBound_dist} are the same, and $H(S_1,S_2|W,W_3)=0$.} In this case, the RHS of \eqref{eq:Example_markov} becomes ${\displaystyle \mspace{-15mu} \max_{\mspace{15mu} p(x_1,x_2): \rho^{\ast}_{X_1 X_2} \le \rho^{\ast}_{S_1 S_2}} \mspace{-15mu} I(X_1,X_2;Y_S) + C_3}$,~and we have $\text{I}_{\text{new}} \approx 0.919$\footnote{This value is found via an exhaustive search over over all $p(x_1,x_2)$ s.t $\rho^{\ast}_{X_1 X_2} \le \rho^{\ast}_{S_1 S_2}$, and can be achieved by $\Pr \left( (X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(0,0) \right) \approx 0.01$, $\Pr \left( (X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(0,1) \right) \approx 0.47$, $\Pr \left( (X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(1,0) \right) \approx 0.01$, $\Pr \left( (X_1,X_2)\mspace{-2mu}=\mspace{-2mu}(1,1) \right) \approx 0.51$.}. To conclude, in this case, the presence of side information at the relay significantly enlarges~$\text{I}_{\text{new}}$.
\end{remark}
\begin{remark}
We note that the necessary conditions presented in \Thmref{thm:OuterMarkov} and \Thmref{thm:BCMarkov} are not tight in general. For instance, consider the PSOMARC specified in Table \ref{tab:PrimitiveSOMARC} with $C_3 = 1$, and the pair of sources defined in Table \ref{tab:SourceDist}. Prop. \ref{prop:feasible} implies that the sources defined in Table \ref{tab:SourceDist} can be reliably transmitted over this PSOMARC by using the scheme of \Thmref{thm:jointCond_NewSimult}. Here, the maximal sum-rate sufficient condition which is evaluated using \eqref{eq:Example_suff} is $\text{I}_{\text{suff}} = \log_2 3$.
For this combination of sources and channel, the sum-rate necessary condition due to the cut-set bound is evaluated via \eqref{eq:Example_cutset} as $\text{I}_{\text{cut-set}} = 2$, which is achieved by setting $\Pr \left((X_1,X_2)=(0,0) \right) = \Pr \left((X_1,X_2)=(0,1) \right) = \Pr \left((X_1,X_2)=(1,0) \right) = \Pr \left((X_1,X_2)=(1,1) \right) = 0.25$. Furthermore, using the same $p_{X_1,X_2}(x_1,x_2)$ we also evaluate the newly derived sum-rate necessary condition (from either \Thmref{thm:OuterMarkov} or \Thmref{thm:BCMarkov}) via \eqref{eq:Example_markov} as $\text{I}_{\text{new}} = 2$. Thus, for this combination of channel and sources
the RHSs of \eqref{eq:Example_cutset} and \eqref{eq:Example_markov} are strictly larger than the RHS of \eqref{eq:Example_suff}.
On the other hand, there are sources and channels for which $\text{I}_{\text{cut-set}} = \text{I}_{\text{new}} = \text{I}_{\text{suff}}$. As an example, consider a PSOMARC, defined by $\mathcal{X}_1 = \mathcal{X}_2 = \{0,1,2\}, \mathcal{Y}_3 = \{0,1,2,3,4,5\}$ and $\mathcal{Y}_S = \{0,1,2\}$.
Let $C_3 = 1$, and consider the deterministic channel mapping $(X_1,X_2) \mapsto (Y_3,Y_S)$ specified in Table~\ref{tab:PrimitiveSOMARC_Ex2}.
\vspace{-0.1cm}
\begin{table}[h]
\begin{center}
\begin{tabular}[t]{|c|c|c|c|c|c|c|}
\hline
$(X_1, X_2)$ & $(0,0)$ & $(1,1)$ & $(1,2)$ & $(2,0)$ & $(2,2)$ & \mbox{Otherwise} \\
\hline
$ Y_3$ & 0 & 2 & 3 & 4 & 5 & 1 \\
\hline
$ Y_S$ & 0 & 2 & 1 & 2 & 0 & 1 \\
\hline
\end{tabular}
\captionsetup{font=small}
\caption{A deterministic channel mapping $(X_1,X_2) \mapsto (Y_3,Y_S)$ for the PSOMARC. \label{tab:PrimitiveSOMARC_Ex2}}
\vspace{-0.75cm}
\end{center}
\end{table}
\noindent The sources $(S_1,S_2)$ are defined over the sets $\mathcal{S}_1 \mspace{-3mu} = \mspace{-3mu} \mathcal{S}_2 \mspace{-3mu} = \mspace{-3mu} \{0,1,2\}$ with the joint distribution specified in Table \ref{tab:SourceDist_Ex2}.
\vspace{-0.15cm}
\begin{table}[h]
\begin{center}
\begin{tabular}[t]{|c|c|c|c|}
\hline
$S_1$ \textbackslash $S_2$ & 0 & 1 & 2 \\
\hline
0 & 1/6 & 1/6 & 0 \\
\hline
1 & 0 & 1/6 & 1/6 \\
\hline
2 & 1/6 & 0 & 1/6 \\
\hline
\end{tabular}
\captionsetup{font=small}
\caption{The joint distribution of $(S_1,S_2)$. The entry in the $j^{\text{th}}$ row and $m^{\text{th}}$ column, $j,m = 0,1,2$, corresponds to $\Pr \left((S_1,S_2) = (j,m) \right)$. \label{tab:SourceDist_Ex2}}
\vspace{-0.75cm}
\end{center}
\end{table}
\noindent Following the arguments presented in Appendix E, it can be shown that, using the scheme of \Thmref{thm:jointCond_NewSimult} the sources defined in Table \ref{tab:SourceDist_Ex2} can be reliably transmitted over the PSOMARC defined in Table \ref{tab:PrimitiveSOMARC_Ex2}, with $C_3 = 1$. In particular, we have $H(S_1,S_2) = \text{I}_{\text{suff}} = \log_2 6$ (note that since $|\mathcal{Y}_3| = 6$, it follows from \eqref{eq:Example_suff} that $\text{I}_{\text{suff}} \le \log_2 6$).
For the channel mapping specified in Table \ref{tab:PrimitiveSOMARC_Ex2}, we also have $\text{I}_{\text{new}} \le \log_2 6$ and $\text{I}_{\text{cut-set}} \le \log_2 6$. This follows from the fact that $|\mathcal{Y}_S| = 3$ and from the fact that $C_3=1$. In fact, $\text{I}_{\text{cut-set}} = \text{I}_{\text{new}} = \log_2 6$ is obtained by setting $p(x_1,x_2) = p(s_1,s_2)$.
Hence, for this combination of channel and sources the RHSs of \eqref{eq:Example_cutset}, \eqref{eq:Example_markov} and \eqref{eq:Example_suff} coincide and tightness in sum-rate is achieved.
Furthermore, for every $C_3 \ge 1$ we obtain $\text{I}_{\text{new}} = \text{I}_{\text{suff}}$.
To understand this equality, first recall from the above discussion that $\text{I}_{\text{suff}} \le \log_2 6$ with equality obtained with the assignment $p(x_1,x_2) = p(s_1,s_2)$. For evaluating $\text{I}_{\text{new}}$, we recall the expression for $\text{I}_{\text{new}}$ given by \eqref{eq:Example_markov}, repeated here for ease of reference:
\vspace{-0.15cm}
\begin{equation*}
\text{I}_{\text{new}} = \max_{p(x_1,x_2): \rho^{\ast}_{X_1 X_2} \le \rho^{\ast}_{S_1 S_2}} \Big\{ I(X_1,X_2;Y_S) + \min \big\{ C_3, I(X_1,X_2;Y_3|Y_S) \big\} \Big\}.
\end{equation*}
Now, since $|\mathcal{Y}_S| = 3$ we have that $I(X_1,X_2;Y_S) \le \log_2 3$, and this is achieved with equality by the assignment $p(x_1,x_2) = p(s_1,s_2)$. For $I(X_1,X_2;Y_3|Y_S)$ we write:
\vspace{-0.15cm}
\begin{equation*}
I(X_1,X_2;Y_3|Y_S) \stackrel{(a)}{=} H(Y_3|Y_S) \stackrel{(b)}{\le} 1,
\end{equation*}
\vspace{-0.15cm}
\noindent where (a) follows from the the fact that in the considered PSOMARC the mapping from $(X_1,X_2)$ to $Y_3$ is deterministic, and (b) follows from the fact that for every possible value of $Y_S$ there are only two possible values of $Y_3$. An equality in (b) is achieved with the assignment $p(x_1,x_2) = p(s_1,s_2)$. Hence, for $C_3 \ge 1$ the active term in the minimization on the RHS of \eqref{eq:Example_markov} is $I(X_1,X_2;Y_3|Y_S)$, and we have $\text{I}_{\text{new}} = \text{I}_{\text{suff}}$, both maximized with the assignment $p(x_1,x_2) = p(s_1,s_2)$.
Finally, note that if $C_3 < 1$ then the necessary conditions \eqref{eq:Example_cutset} and \eqref{eq:Example_markov} are not satisfied.
\end{remark}
%
%
%
%
%
%
%
%
\vspace{-0.3cm}
\section{Conclusions} \label{sec:conclusions}
\vspace{-0.1cm}
In this work we studied JSCC for lossless transmission of correlated sources over DM MARCs.
We derived a new DF-based JSCC scheme which uses the CPM technique for encoding the correlated source sequences for transmission to both the relay and the destination, while SW source coding is used for cooperation between the sources and the relay. This combination allows removing the constraints on the distribution chain required by a previously derived scheme which used CPM to the destination \cite[Thm. 2]{Murin:ISIT12} (quoted as \Thmref{thm:jointCondFlip} in this manuscript).
The new scheme of \Thmref{thm:jointCond_NewSimult} applies simultaneous backward decoding at the destination to simultaneously decode both source sequences and the cooperation information. As the scheme implements CPM-based encoding of the source sequences at the transmitters, both the relay and the destination benefit from the joint source-channel encoding. This is in contrast to the JSCC schemes derived in \cite{Murin:ISIT12} (quoted as \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip} in this manuscript), in which either the relay or the destination benefits from the CPM encoding, but not both simultaneously.
We then provided a detailed comparison of the new scheme of \Thmref{thm:jointCond_NewSimult} with the two JSCC schemes of \cite{Murin:ISIT12} and with the scheme of Prop. \ref{prop:jointCond_New} which apply sequential decoding of the source sequences and the cooperation information at the destination.
We showed that the scheme of \Thmref{thm:jointCond_NewSimult} is better than the scheme derived in \cite[Thm. 1]{Murin:ISIT12} and the scheme of Prop. \ref{prop:jointCond_New}.
We also showed that there are cases in which the scheme of \Thmref{thm:jointCond_NewSimult} strictly outperforms the schemes of \Thmref{thm:jointCond} and \Thmref{thm:jointCondFlip}.
However, we cannot show that the new scheme of \Thmref{thm:jointCond_NewSimult} is universally better than the scheme of \cite[Thm. 2]{Murin:ISIT12}. This follows from the different admissible joint distributions (see Remarks \ref{rem:oldSchemesComp} and~\ref{rem:SameDist}).
Finally, we derived three different sets of necessary conditions for reliable transmission of correlated sources over DM MARCs. We also showed that the newly derived sets are at least as tight as previously known results.
One of the new sets is in the spirit of the ``MAC bound" for the classic relay channel, while the other two sets are in the spirit of the ``broadcast bound" for the relay channel. Two of the new sets use the Markov relationship between the sources and the channel inputs to restrict the set of feasible distributions.
\appendices
\numberwithin{equation}{section}
\numberwithin{MyProposition}{section}
%
%
%
\vspace{-0.2cm}
\section{Proof of Theorem \thmref{thm:jointCond_NewSimult}} \label{annex:jointNewSimultProof}
\vspace{-0.2cm}
\subsection{Codebook Construction} \label{subsecsec:Thm3_codeconst}
\vspace{-0.1cm}
\begin{itemize}
\item
For each $i=1,2$, consider a set of $2^{nR_i}$ bins and let $\mathcal{\msgBig}_i \triangleq \{1,2,\dots,2^{nR_i}\}, i=1,2$, be the corresponding set of bin indices. For $i=1,2$, assign every $\mathbf{s}_i \in \mathcal{S}_i^n$ to one of the $2^{nR_i}$ bins independently according to a uniform distribution over the bin indices. Denote this assignment by $f_i: \mathcal{S}_i^n \mapsto \mathcal{\msgBig}_i, i=1,2$.
\item
For $i=1,2$, generate $2^{nR_i}$ codewords $\mathbf{v}_i(u_i), u_i \in \mathcal{\msgBig}_i $, by choosing the letters $v_{i,k}(u_i), k = 1,2,\dots,n$,
independently according to the p.m.f $p_{V_i}(v_{i,k}(u_i))$.
For each pair $(\mathbf{s}_i, u_i) \in \mathcal{S}_i^n \times \mathcal{\msgBig}_i, i=1,2$, generate one codeword $\mathbf{x}_i(\mathbf{s}_i, u_i)$ by choosing the letters $x_{i,k}(\mathbf{s}_i, u_i)$ independently according to the p.m.f $p_{X_i|S_i,V_i}(x_{i,k}|s_{i,k},v_{i,k}(u_i))$, $k=1,2,\dots, n$. Finally, generate one relay codeword $\mathbf{x}_3(u_1,u_2)$ for each pair $(u_1, u_2) \in \mathcal{\msgBig}_1 \times \mathcal{\msgBig}_2$, by choosing the letters $x_{3,k}(u_1,u_2)$ independently according to the p.m.f $p_{X_3|V_1,V_2}(x_{3,k}|v_{1,k}(u_1),v_{2,k}(u_2))$, $k=1,2,\dots, n$.
\end{itemize}
\vspace{-0.25cm}
\subsection{Encoding} \label{subsecsec:Thm3_enc}
\vspace{-0.1cm}
Consider two source sequences each of length $Bn$, $s^{Bn}_{i,1} \in \mathcal{S}^{Bn}_i, i=1,2$. Partition each sequence into $B$ length-$n$ subsequences, $\mathbf{s}_{i,b} \in \mathcal{S}_i^n, b=1,2,\dots,B$. Similarly partition the side information sequences $w_{3,1}^{Bn}$ and $w^{Bn}$ into $B$ length-$n$ subsequences $\mathbf{w}_{3,b} \in \mathcal{W}_3^n, \mathbf{w}_{b} \in \mathcal{W}^n, b=1,2,\dots,B$, respectively. A total of $Bn$ source samples is transmitted over $B+1$ blocks, such that at each block $n$ channel symbols are transmitted.
At block $1$, transmitter $i, i=1,2$, transmits the channel codeword $\mathbf{x}_i(\mathbf{s}_{i,1}, 1)$.
At block $b, b=2,3,\dots,B$, transmitter $i$ transmits the channel codeword $\mathbf{x}_i(\mathbf{s}_{i,b}, u_{i,b-1})$, where $u_{i,b-1} = f_i(\mathbf{s}_{i,b-1}) \in \mathcal{\msgBig}_i$ is the bin index of source vector $\mathbf{s}_{i,b-1}$.
Let $(\mathbf{a}_1, \mathbf{a}_2) \in \mathcal{S}_1^n \times \mathcal{S}_2^n$ be two sequences generated according to $p(\mathbf{a}_1,\mathbf{a}_2) = \prod_{k=1}^{n}{p_{S_1,S_2}(a_{1,k}, a_{2,k})}$. These sequences are known to all nodes. At block $B+1$, transmitter $i, i=1,2$, transmits $\mathbf{x}_i(\mathbf{a}_i,u_{i,B})$.
At block $b=1$, the relay transmits $\mathbf{x}_3(1,1)$.
Assume that at block $b, b=2,3,\dots,B,B+1$, the relay has the estimates $(\tilde{\mathbf{s}}_{1,b-1}, \tilde{\mathbf{s}}_{2,b-1})$ of $(\mathbf{s}_{1,b-1}, \mathbf{s}_{2,b-1})$. It then finds the corresponding bin indices $\tilde{u}_{i,b-1} = f_i(\tilde{\mathbf{s}}_{i,b-1}) \in \mathcal{\msgBig}_i, i=1,2$, and transmits the channel codeword $\mathbf{x}_3(\tilde{u}_{1,b-1},\tilde{u}_{2,b-1})$ at time $b$.
\vspace{-0.4cm}
\subsection{Decoding} \label{subsec:jointNewProofDecoding}
\vspace{-0.1cm}
The relay decodes the source sequences sequentially. At the end of channel block $b$ the relay decodes $\mathbf{s}_{i,b}, i=1,2$, as follows:
Using the estimates $(\tilde{\msg}_{1,b-1},\tilde{\msg}_{2,b-1})$, the received signal $\mathbf{y}_{3,b}$ and the side information $\mathbf{w}_{3,b}$, the relay decodes $(\mathbf{s}_{1,b}, \mathbf{s}_{2,b})$ by looking for a unique pair $(\tilde{\svec}_{1}, \tilde{\svec}_{2}) \in \mathcal{S}_1^n \times \mathcal{S}_2^n$ such that:
\vspace{-0.2cm}
\begin{align}
& \big(\tilde{\svec}_{1}, \tilde{\svec}_{2}, \mathbf{v}_1(\tilde{\msg}_{1,b-1}), \mathbf{v}_2(\tilde{\msg}_{2,b-1}), \mathbf{x}_1(\tilde{\svec}_{1}, \tilde{\msg}_{1,b-1}), \mathbf{x}_2(\tilde{\svec}_{2}, \tilde{\msg}_{2,b-1}), \mathbf{x}_3(\tilde{\msg}_{1,b-1}, \tilde{\msg}_{2,b-1}), \mathbf{w}_{3,b}, \mathbf{y}_{3,b}\big) \in A_{\epsilon}^{*(n)}.
\label{eq:RelayJntNewDecType}
\end{align}
\vspace{-0.2cm}
Decoding at the destination is done via simultaneous backward decoding.
Let $\boldsymbol{\alpha} \in \mathcal{W}^n$ be an i.i.d sequence such that each letter $\alpha_k$ is selected independently according to $p_{W|S_1,S_2}(\alpha_k | a_{1,k}, a_{2,k}), k=1,2,\dots,n$.
The destination node waits until the end of channel block $B+1$. It first tries to decode $(u_{1,B},u_{2,B})$ using the received signal at channel block $B+1$, $\mathbf{y}_{b+1}$, and using $\mathbf{a}_1, \mathbf{a}_2$, and $\boldsymbol{\alpha}$.
Going backwards from the last channel block to the first, we assume that at block $b$ the destination has estimates $(\hat{\msg}_{1,b},\hat{\msg}_{2,b})$ of $(u_{1,b},u_{2,b})$.
The destination simultaneously decodes $(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, u_{1,b-1},u_{2,b-1})$ based on the received signal $\mathbf{y}_{b}$, and the side information $\mathbf{w}_{b}$, by looking for a unique combination $(\hat{\svec}_{1},\hat{\svec}_{2}, \hat{\msg}_{1},\hat{\msg}_{2})\in \mathcal{S}_1^n \times \mathcal{S}_2^n \times \mathcal{\msgBig}_1 \times \mathcal{\msgBig}_2$ such that:
\vspace{-0.25cm}
\begin{align}
& \big( \hat{\svec}_{1}, \hat{\svec}_{2}, \mathbf{v}_1(\hat{\msg}_{1}), \mathbf{v}_2(\hat{\msg}_{2}), \mathbf{x}_1(\hat{\svec}_{1}, \hat{\msg}_{1}), \mathbf{x}_2(\hat{\svec}_{2}, \hat{\msg}_{2}), \mathbf{x}_3(\hat{\msg}_{1}, \hat{\msg}_{2}), \mathbf{w}_{b}, \mathbf{y}_{b} \big) \in A_{\epsilon}^{*(n)},
\label{eq:DestJntNewSimultDecTypeList}
\end{align}
\vspace{-0.25cm}
\noindent and $f_1(\hat{\svec}_{1})=\hat{\msg}_{1,b}, f_2(\hat{\svec}_{2,b})=\hat{\msg}_{2}$. Denote the decoded variables by $(\hat{\svec}_{1,b},\hat{\svec}_{2,b}, \hat{\msg}_{1,b-1},\hat{\msg}_{2,b-1})$.
\vspace{-0.4cm}
\subsection{Error Probability Analysis} \label{subsec:jointNewProofErrorAnalysis}
\vspace{-0.1cm}
\textbf{Relay error probability:} The relay error probability analysis follows the same arguments as the relay error probability analysis detailed in \cite[Appendix B]{Murin:IT11}.
\textbf{Destination error probability:} The average probability of error in decoding at the destination at block $b$, $\bar{P}_{\mbox{\scriptsize dest},b}^{(n)}$, is defined by:
\vspace{-0.45cm}
\begin{align}
\bar{P}_{\mbox{\scriptsize dest},b}^{(n)} & \triangleq \Pr\big((\hat{\mathbf{S}}_{1,b},\hat{\mathbf{S}}_{2,b}) \ne (\mathbf{S}_{1,b},\mathbf{S}_{2_b})\big). \nonumber
\end{align}
\vspace{-0.2cm}
Due to backward decoding, the pair of source sequences sent at time $b$ is decoded after the pair at time $b+1$ is decoded. Let $ \mathcal{F}_b \triangleq \big\{\big(\hat{\mathbf{S}}_{1,b}, \hat{\mathbf{S}}_{2,b}, \hat{U}_{1,b-1}, \hat{U}_{2,b-1}\big) \ne \big(\mathbf{S}_{1,b}, \mathbf{S}_{2,b}, U_{1,b-1}, U_{2,b-1} \big)\big\}$. Then, as in \cite[Eqn. (40)]{CoverG:79}, we write:
\vspace{-0.2cm}
\begin{equation}
\bar{P}_{\mbox{\scriptsize dest}}^{(n)}\le \sum_{b=1}^B\Pr\big(\mathcal{F}_b\cap\mathcal{F}_{b+1}^c\big).\footnote{As stated in Subsection \ref{subsecsec:Thm3_enc}, at block $B+1$, source terminal $i$ transmits $\mathbf{x}_i(\mathbf{a}_i,u_{i,B})$, where $\mathbf{a}_i \in \mathcal{S}_i^n, i=1,2$, is known to all nodes. Therefore, at block $B+1$ we define $\mathcal{F}_{B+1} \triangleq \big\{\big( \hat{U}_{1,B}, \hat{U}_{2,B}\big) \ne \big(U_{1,B}, U_{2,B} \big)\big\}$.}
\end{equation}
\vspace{-0.15cm}
\noindent Let $\epsilon_0, \epsilon_1, \dots, \epsilon_8$ be positive numbers such that $\epsilon_0 \ge \epsilon_1 > \epsilon, \epsilon_m > \epsilon$ and $\epsilon_m \rightarrow 0$ as $\epsilon \rightarrow 0$, for $ m=0,1,\dots,8$.
\noindent Now, define two error events at block $b$:
\vspace{-0.15cm}
\begin{itemize}
\item Joint-typicality fails:
\vspace{-0.3cm}
\begin{align*}
\mathcal{E}_{1,b} & \triangleq \Big\{\big(\mathbf{S}_{1,b}, \mathbf{S}_{2,b}, \mathbf{V}_1(U_{1,b-1}), \mathbf{V}_2(U_{2,b-1}),\mathbf{X}_1(\mathbf{S}_{1,b},U_{1,b-1}), \\
& \qquad \qquad \mathbf{X}_2(\mathbf{S}_{2,b},U_{2,b-1}), \mathbf{X}_3(U_{1,b-1},U_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big)\notinA_{\epsilon}^{*(n)}\Big\}.
\end{align*}
\vspace{-0.25cm}
\item Simultaneous decoding of the bin indices (for the next step) and the source sequences fails:
\vspace{-0.3cm}
\begin{align*}
\mathcal{E}_{2,b} & \triangleq \Big\{\exists \big(\hat{\svec}_1, \hat{\svec}_2, \hat{u}_1, \hat{u}_2\big)\in\mathcal{S}_1^n\times\mathcal{S}_2^n\times\mathcal{U}_1\times\mathcal{U}_2, \\
& \qquad \qquad \big(\hat{\svec}_1, \hat{\svec}_2, \hat{u}_1, \hat{u}_2\big) \ne \big(\mathbf{S}_{1,b}, \mathbf{S}_{2,b}, U_{1,b-1}, U_{2,b-1} \big), f_1(\tilde{\svec}_1) = \hat{U}_{1,b}, f_2(\tilde{\svec}_2) = \hat{U}_{2,b},\\
& \qquad \qquad \qquad \big(\hat{\svec}_{1}, \hat{\svec}_{2}, \hat{\mathbf{V}}_1(\hat{u}_{1}), \hat{\mathbf{V}}_2(\hat{u}_{2}), \hat{\mathbf{X}}_1(\hat{\svec}_{1},\hat{u}_{1}), \hat{\mathbf{X}}_2(\hat{\svec}_{2},\hat{u}_{2}), \hat{\mathbf{X}}_3(\hat{u}_{1},\hat{u}_{2}), \mathbf{W}_b,\mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big\}.
\end{align*}
\end{itemize}
\vspace{-0.3cm}
\noindent Then, $\mathcal{F}_b = \mathcal{E}_{1,b}\cup\mathcal{E}_{2,b}$, and we bound:
\vspace{-0.3cm}
\[
\Pr\big(\mathcal{F}_b\cap\mathcal{F}_{b+1}^c\big) \le \Pr\big(\mathcal{E}_{1,b}\cup\mathcal{E}_{2,b}\big|\mathcal{F}_{b+1}^c\big) = \Pr\big(\mathcal{E}_{1,b}\big|\mathcal{F}_{b+1}^c\big) + \Pr\big(\mathcal{E}_{2,b}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big).
\]
\vspace{-0.25cm}
\noindent By ~applying ~the ~properties ~of ~strong ~typicality, ~\cite[Theorem 6.9]{YeungBook} ~we ~have ~that ~for ~$n$ ~sufficiently ~large, $\Pr\Big(\mathcal{E}_{1,b}\Big|\mathcal{F}_{b+1}^c\Big)\le \epsilon$.
For bounding $\Pr\big(\mathcal{E}_{2,b}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$ we consider the following error events:
\begin{flalign*}
& \mathcal{E}_{2,b}^{(1)} \triangleq \Big\{ \exists \hat{u}_1 \in \mathcal{U}_1, \hat{u}_1 \ne U_{1,b-1}, \big(\mathbf{S}_{1,b}, \mathbf{S}_{2,b}, \hat{\mathbf{V}}_1(\hat{u}_1), \mathbf{V}_2(U_{2,b-1}), & \\
& \mspace{80mu} \hat{\mathbf{X}}_1(\mathbf{S}_{1,b},\hat{u}_1), \mathbf{X}_2(\mathbf{S}_{2,b},U_{2,b-1}), \tilde{\mathbf{X}}_3(\hat{u}_1,U_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b \big)\inA_{\epsilon}^{*(n)}\Big\}. & \\
& \mathcal{E}_{2,b}^{(2)} \triangleq \Big\{ \exists \hat{u}_2 \in \mathcal{U}_2, \hat{u}_2 \ne U_{2,b-1}, \big(\mathbf{S}_{1,b}, \mathbf{S}_{2,b}, \mathbf{V}_1(U_{1,b-1}), \hat{\mathbf{V}}_2(\hat{u}_2), & \\
& \mspace{80mu} \mathbf{X}_1(\mathbf{S}_{1,b},U_{1,b-1}), \hat{\mathbf{X}}_2(\mathbf{S}_{2,b},\hat{u}_2), \hat{\mathbf{X}}_3(U_{1,b-1},\hat{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b \big)\inA_{\epsilon}^{*(n)}\Big\}. & \\
&\mathcal{E}_{2,b}^{(3)} \triangleq \Big\{ \exists \hat{u}_1 \in \mathcal{U}_1, \hat{u}_1 \ne U_{1,b-1}, \exists \hat{u}_2 \in \mathcal{U}_2, \hat{u}_2 \ne U_{2,b-1},& \\
& \mspace{80mu} \big(\mathbf{S}_{1,b}, \mathbf{S}_{2,b}, \hat{\mathbf{V}}_1(\hat{u}_1), \hat{\mathbf{V}}_2(\hat{u}_2), \hat{\mathbf{X}}_1(\mathbf{S}_{1,b},\hat{u}_1), \hat{\mathbf{X}}_2(\mathbf{S}_{2,b},\hat{u}_2), \hat{\mathbf{X}}_3(\hat{u}_1,\hat{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b \big)\inA_{\epsilon}^{*(n)}\Big\}.& \\
&\mathcal{E}_{2,b}^{(4)} \triangleq \Big\{ \exists \hat{\svec}_1 \in \mathcal{S}_1^n, \hat{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\hat{\svec}_1) = \hat{U}_{1,b}, \big(\hat{\svec}_1,\mathbf{S}_{2,b}, \mathbf{V}_1(U_{1,b-1}), \mathbf{V}_2(U_{2,b-1}), & \\
& \mspace{80mu} \hat{\mathbf{X}}_1(\hat{\svec}_1,U_{1,b-1}), \mathbf{X}_2(\mathbf{S}_{2,b},U_{2,b-1}), \mathbf{X}_3(U_{1,b-1}, U_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big\}. & \\
&\mathcal{E}_{2,b}^{(5)} \triangleq \Big\{ \exists \hat{\svec}_2 \in \mathcal{S}_2^n, \hat{\svec}_2 \ne \mathbf{S}_{2,b}, f_2(\hat{\svec}_2) = \hat{U}_{2,b}, \big(\mathbf{S}_{1,b}, \hat{\svec}_{2}, \mathbf{V}_1(U_{1,b-1}), \mathbf{V}_2(U_{2,b-1}), & \\
& \mspace{80mu} \mathbf{X}_1(\mathbf{S}_{1,b},U_{1,b-1}), \hat{\mathbf{X}}_2(\tilde{\svec}_{2},U_{2,b-1}), \mathbf{X}_3(U_{1,b-1}, U_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big\}. & \\
&\mathcal{E}_{2,b}^{(6)} \triangleq \Big\{ \exists (\hat{\svec}_1, \hat{\svec}_2) \in \mathcal{S}_1^n\times\mathcal{S}_2^n, \hat{\svec}_1 \ne \mathbf{S}_{1,b}, \hat{\svec}_2 \ne \mathbf{S}_{2,b}, f_1(\hat{\svec}_1) = \hat{U}_{1,b}, f_2(\hat{\svec}_2) = \hat{U}_{2,b}, \big(\hat{\svec}_{1}, \hat{\svec}_{2}, \mathbf{V}_1(U_{1,b-1}), & \\
& \mspace{80mu} \mathbf{V}_2(U_{2,b-1}), \hat{\mathbf{X}}_1(\hat{\svec}_{1},U_{1,b-1}), \hat{\mathbf{X}}_2(\hat{\svec}_{2},U_{2,b-1}), \mathbf{X}_3(U_{1,b-1}, U_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big\}. &
\end{flalign*}
\begin{flalign*}
&\mathcal{E}_{2,b}^{(7)} \triangleq \Big\{ \exists \hat{\svec}_1 \in \mathcal{S}_1^n, \hat{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\hat{\svec}_1) = \hat{U}_{1,b}, \exists \hat{u}_1\in\mathcal{U}_1, \hat{u}_1 \ne U_{1,b-1}, \big(\hat{\svec}_{1}, \mathbf{S}_{2,b}, \hat{\mathbf{V}}_1(\hat{u}_{1}), & \\
& \mspace{80mu} \mathbf{V}_2(U_{2,b-1}), \hat{\mathbf{X}}_1(\hat{\svec}_{1},\hat{u}_{1}), \mathbf{X}_2(\mathbf{S}_{2,b},U_{2,b-1}), \hat{\mathbf{X}}_3(\hat{u}_{1}, U_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big\}. & \\
&\mathcal{E}_{2,b}^{(8)} \triangleq \Big\{ \exists \hat{\svec}_2 \in \mathcal{S}_2^n, \hat{\svec}_2 \ne \mathbf{S}_{2,b}, f_2(\hat{\svec}_2) = \hat{U}_{2,b}, \exists \hat{u}_2\in\mathcal{U}_2, \hat{u}_2 \ne U_{2,b-1}, \big(\mathbf{S}_{1,b}, \hat{\svec}_{2}, \mathbf{V}_1(U_{1,b-1}), & \\
& \mspace{80mu} \hat{\mathbf{V}}_2(\hat{u}_{2}), \mathbf{X}_1(\mathbf{S}_{1,b},U_{1,b-1}), \hat{\mathbf{X}}_2(\hat{\svec}_{2},\hat{u}_{2}), \hat{\mathbf{X}}_3(U_{1,b-1}, \hat{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big\}. & \\
&\mathcal{E}_{2,b}^{(9)} \triangleq \Big\{ \exists \hat{\svec}_1 \in \mathcal{S}_1^n, \hat{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\hat{\svec}_1) = \hat{U}_{1,b}, \exists \hat{u}_2\in\mathcal{U}_2, \hat{u}_2\ne U_{2,b-1}, \big(\hat{\svec}_{1}, \mathbf{S}_{2,b}, \mathbf{V}_1(U_{1,b-1}), & \\
& \mspace{80mu} \hat{\mathbf{V}}_2(\hat{u}_{2}), \hat{\mathbf{X}}_1(\hat{\svec}_{1},U_{1,b-1}), \hat{\mathbf{X}}_2(\mathbf{S}_{2,b},\hat{u}_{2}), \hat{\mathbf{X}}_3(U_{1,b-1}, \hat{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big\}. & \\
&\mathcal{E}_{2,b}^{(10)} \triangleq \Big\{ \exists \hat{\svec}_2 \in \mathcal{S}_2^n, \hat{\svec}_2 \ne \mathbf{S}_{2,b}, f_2(\hat{\svec}_2) = \hat{U}_{2,b}, \exists \hat{u}_1\in\mathcal{U}_1, \hat{u}_1\ne U_{1,b-1}, \big(\mathbf{S}_{1,b}, \hat{\svec}_{2}, \hat{\mathbf{V}}_1(\hat{u}_{1}), & \\
& \mspace{80mu} \mathbf{V}_2(U_{2,b-1}), \hat{\mathbf{X}}_1(\mathbf{S}_{1,b},\hat{u}_{1}), \hat{\mathbf{X}}_2(\hat{\svec}_{2},U_{2,b-1}), \hat{\mathbf{X}}_3(\hat{u}_{1}, U_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big\}. & \\
&\mathcal{E}_{2,b}^{(11)} \triangleq \Big\{ \exists \hat{\svec}_1 \in \mathcal{S}_1^n, \hat{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\hat{\svec}_1) = \hat{U}_{1,b}, \exists \hat{u}_1\in\mathcal{U}_1, \hat{u}_1\ne U_{1,b-1}, \exists \hat{u}_2\in\mathcal{U}_2, \hat{u}_2\ne U_{2,b-1}, & \\
& \mspace{120mu} \big(\hat{\svec}_{1}, \mathbf{S}_{2,b}, \hat{\mathbf{V}}_1(\hat{u}_{1}), \hat{\mathbf{V}}_2(\hat{u}_{2}), \hat{\mathbf{X}}_1(\hat{\svec}_{1},\hat{u}_{1}), \hat{\mathbf{X}}_2(\mathbf{S}_{2,b},\hat{u}_{2}), \hat{\mathbf{X}}_3(\hat{u}_{1}, \hat{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big\}. & \\
&\mathcal{E}_{2,b}^{(12)} \triangleq \Big\{ \exists \hat{\svec}_2 \in \mathcal{S}_2^n, \hat{\svec}_2 \ne \mathbf{S}_{2,b}, f_2(\hat{\svec}_2) = \hat{U}_{2,b}, \exists \hat{u}_1\in\mathcal{U}_1, \hat{u}_1\ne U_{1,b-1}, \exists \hat{u}_2\in\mathcal{U}_2, \hat{u}_2\ne U_{2,b-1}, & \\
& \mspace{80mu} \big(\mathbf{S}_{1,b}, \hat{\svec}_{2}, \hat{\mathbf{V}}_1(\hat{u}_{1}), \hat{\mathbf{V}}_2(\hat{u}_{2}), \hat{\mathbf{X}}_1(\mathbf{S}_{1,b},\hat{u}_{1}), \hat{\mathbf{X}}_2(\hat{\svec}_{2},\hat{u}_{2}), \hat{\mathbf{X}}_3(\hat{u}_{1}, \hat{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big\}. & \\
&\mathcal{E}_{2,b}^{(13)} \triangleq \Big\{ \exists \hat{\svec}_1 \in \mathcal{S}_1^n, \hat{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\hat{\svec}_1) = \hat{U}_{1,b}, \exists \hat{\svec}_2 \in \mathcal{S}_2^n, \hat{\svec}_2 \ne \mathbf{S}_{2,b}, f_2(\hat{\svec}_2) = \hat{U}_{2,b}, \exists \hat{u}_1\in\mathcal{U}_1, \hat{u}_1\ne U_{1,b-1}, & \\
& \mspace{80mu} \big(\hat{\svec}_{1}, \hat{\svec}_{2}, \hat{\mathbf{V}}_1(\hat{u}_{1}), \mathbf{V}_2(U_{2,b-1}), \hat{\mathbf{X}}_1(\hat{\svec}_{1},\hat{u}_{1}), \hat{\mathbf{X}}_2(\hat{\svec}_{2},U_{2,b-1}), \hat{\mathbf{X}}_3(\hat{u}_{1}, U_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big\}. & \\
&\mathcal{E}_{2,b}^{(14)} \triangleq \Big\{ \exists \hat{\svec}_1 \in \mathcal{S}_1^n, \hat{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\hat{\svec}_1) = \hat{U}_{1,b}, \exists \hat{\svec}_2 \in \mathcal{S}_2^n, \hat{\svec}_2 \ne \mathbf{S}_{2,b}, f_2(\hat{\svec}_2) = \hat{U}_{2,b}, \exists \hat{u}_2\in\mathcal{U}_2, \hat{u}_2\ne U_{2,b-1}, & \\
& \mspace{80mu} \big(\hat{\svec}_{1}, \hat{\svec}_{2}, \mathbf{V}_1(U_{1,b-1}), \hat{\mathbf{V}}_2(\hat{u}_{2}), \hat{\mathbf{X}}_1(\hat{\svec}_{1},U_{1,b-1}), \hat{\mathbf{X}}_2(\hat{\svec}_{2},\hat{u}_{2}), \hat{\mathbf{X}}_3(U_{1,b-1}, \hat{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big\}. & \\
&\mathcal{E}_{2,b}^{(15)} \triangleq \Big\{ \exists \hat{\svec}_1 \in \mathcal{S}_1^n, \hat{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\hat{\svec}_1) = \hat{U}_{1,b}, \exists \hat{\svec}_2 \in \mathcal{S}_2^n, \hat{\svec}_2 \ne \mathbf{S}_{2,b}, f_2(\hat{\svec}_2) = \hat{U}_{2,b}, & \\
& \mspace{80mu} \exists \hat{u}_1\in\mathcal{U}_1, \hat{u}_1\ne U_{1,b-1}, \exists \hat{u}_2\in\mathcal{U}_2, \hat{u}_2\ne U_{2,b-1}, & \\
& \mspace{80mu} \big(\hat{\svec}_{1}, \hat{\svec}_{2}, \hat{\mathbf{V}}_1(\hat{u}_{1}), \hat{\mathbf{V}}_2(\hat{u}_{2}), \hat{\mathbf{X}}_1(\hat{\svec}_{1},\hat{u}_{1}), \hat{\mathbf{X}}_2(\hat{\svec}_{2},\hat{u}_{2}), \hat{\mathbf{X}}_3(\hat{u}_{1}, \hat{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big\}. &
\end{flalign*}
%
%
%
%
%
\ifthenelse{\boolean{SquizFlag}}{}{\begin{comment}}
\vspace{-0.15cm}
Following the same arguments as in the error probability analysis detailed in \cite[Appendix B, Eqns. (B.37)--(B.45)]{Murin:IT11}, we have that the probability $\Pr\big(\mathcal{E}_{2,b}^{(m)}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$ can be made arbitrarily small for $m=1,2,3$, by increasing the block length $n$, if the following conditions are satisfied correspondingly:
\vspace{-0.3cm}
\begin{subequations} \label{eq:ErrAnalysisCondSimult_1-3}
\begin{align}
R_1 & < I(X_1, X_3 ; Y | S_1, V_2, X_2 ) - 2\epsilon_2 \label{eq:ErrAnalysisCondSimult_1} \\
R_2 & < I(X_2, X_3 ; Y | S_2, V_1, X_1 ) - 2\epsilon_2 \label{eq:ErrAnalysisCondSimult_2} \\
R_1 + R_2 & < I(X_1,X_2,X_3;Y|S_1,S_2) - 2\epsilon_2. \label{eq:ErrAnalysisCondSimult_3}
\end{align}
\end{subequations}
\vspace{-0.1cm}
The bounds for $\Pr\big(\mathcal{E}_{2,b}^{(m)}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big), 4 \le m \le 15$, follow similar arguments. We demonstrate the technique for $m=7$.
We begin by writing:
\vspace{-0.1cm}
\begin{align*}
\Pr & \big(\mathcal{E}_{2,b}^{(7)}|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big) \\
& = \Pr\Big( \exists \hat{\svec}_1 \in \mathcal{S}_1^n, \hat{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\hat{\svec}_1) = \hat{U}_{1,b}, \exists \hat{u}_1\in\mathcal{U}_1, \hat{u}_1\ne U_{1,b-1}, \big(\hat{\svec}_{1}, \mathbf{S}_{2,b}, \hat{\mathbf{V}}_1(\hat{u}_{1}), \mathbf{V}_2(U_{2,b-1}), \\
& \qquad \qquad \hat{\mathbf{X}}_1(\hat{\svec}_{1},\hat{u}_{1}), \mathbf{X}_2(\mathbf{S}_{2,b},U_{2,b-1}), \hat{\mathbf{X}}_3(\hat{u}_{1}, U_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big) \\
%
& = \sum_{\hat{u}_{1,b}\in \mathcal{U}_1, u_{1,b-1}\in\mathcal{U}_1, u_{2,b-1}\in\mathcal{U}_2} \mspace{-40mu}
p_{U_1}(\hat{u}_{1,b}) p_{U_1 U_2}(u_{1,b-1},u_{2,b-1})\times \nonumber\\
& \mspace{100mu} \Pr\Big( \exists \hat{\svec}_1 \in \mathcal{S}_1^n, \hat{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\hat{\svec}_1) = \hat{u}_{1,b}, \exists \hat{u}_1\in\mathcal{U}_1, \hat{u}_1\ne u_{1,b-1}, \big(\hat{\svec}_{1}, \mathbf{S}_{2,b}, \hat{\mathbf{V}}_1(\hat{u}_{1}), \mathbf{V}_2(u_{2,b-1}), \\
& \qquad \qquad \qquad \qquad \hat{\mathbf{X}}_1(\hat{\svec}_{1},\hat{u}_{1}), \mathbf{X}_2(\mathbf{S}_{2,b},u_{2,b-1}), \hat{\mathbf{X}}_3(\hat{u}_{1}, u_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big).
\end{align*}
\vspace{-0.3cm}
\noindent We now bound:
\vspace{-0.25cm}
\begin{align*}
& \Pr\Big( \exists \hat{\svec}_1 \in \mathcal{S}_1^n, \hat{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\hat{\svec}_1) = \hat{u}_{1,b}, \exists \hat{u}_1\in\mathcal{U}_1, \hat{u}_1\ne u_{1,b-1}, \big(\hat{\svec}_{1}, \mathbf{S}_{2,b}, \hat{\mathbf{V}}_1(\hat{u}_{1}), \mathbf{V}_2(u_{2,b-1}), \\
& \qquad \quad \hat{\mathbf{X}}_1(\hat{\svec}_{1},\hat{u}_{1}), \mathbf{X}_2(\mathbf{S}_{2,b},u_{2,b-1}), \hat{\mathbf{X}}_3(\hat{u}_{1}, u_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big) \\
& \stackrel{(a)}{=} \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-120mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{50mu} \sum_{\substack{\hat{u}_1\in\mathcal{U}_1,\\ \hat{u}_1 \ne u_{1,b-1}}} \mspace{-100mu} \sum_{ \mspace{120mu} \substack{\hat{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big), \\ \hat{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-170mu} \Pr\Big( f_1(\hat{\svec}_1) = \hat{u}_{1,b}, \big(\hat{\svec}_{1}, \hat{\mathbf{V}}_1(\hat{u}_{1}), \hat{\mathbf{X}}_1(\hat{\svec}_{1},\hat{u}_{1}), \hat{\mathbf{X}}_3(\hat{u}_{1},u_{2,b-1})\big) \in \\
& \mspace{240mu} A_{\epsilon}^{*(n)}\big(S_1,V_1,X_1,X_3\big|\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big) \Big) \\
& \stackrel{(b)}{=} \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-120mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{50mu} \sum_{\substack{\hat{u}_1\in\mathcal{U}_1,\\ \hat{u}_1 \ne u_{1,b-1}}} \mspace{-100mu} \sum_{ \mspace{120mu} \substack{\hat{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big), \\ \hat{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-170mu} 2^{-nR_1} \Pr\Big( \big(\hat{\mathbf{V}}_1(\hat{u}_{1}), \hat{\mathbf{X}}_1(\hat{\svec}_{1},\hat{u}_{1}), \hat{\mathbf{X}}_3(\hat{u}_{1},u_{2,b-1})\big) \in \\
& \mspace{240mu} A_{\epsilon}^{*(n)}\big(V_1,X_1,X_3\big|\hat{\svec}_{1}, \mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big) \Big)
\end{align*}
\vspace{-0.2cm}
\noindent where (a) follows from the conditioning on $\mathcal{E}_{b,1}^c$ which implies that the sequences at block $b$ are jointly typical, and from consistency of strong typicality \cite[Theorem 6.7]{YeungBook}: Let $\mathbf{z}_b \triangleq \big(\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}),$ $ \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}),\mathbf{w}_b,\mathbf{y}_b \big)$.
~By ~\cite[Eqn. (6.110)]{YeungBook}, ~when ~$\mathbf{z}_b\inA_{\epsilon}^{*(n)}(S_2,V_2,X_2,W,Y)$, the conditionally typical set $A_{\epsilon}^{*(n)}(S_1,V_1,X_1,X_3|\mathbf{z}_b)$ is defined as:
\vspace{-0.2cm}
\[
A_{\epsilon}^{*(n)}(S_1,V_1,X_1,X_3|\mathbf{z}_b) \triangleq \big\{(\hat{\svec}_1,\hat{\vvec}_1,\hat{\xvec}_1,\hat{\xvec}_3)\inA_{\epsilon}^{*(n)}(S_1,V_1,X_1,X_3): (\hat{\svec}_1,\hat{\vvec}_1,\hat{\xvec}_1,\hat{\xvec}_3,\mathbf{z}_b)\inA_{\epsilon}^{*(n)}\big\}.
\]
\vspace{-0.25cm}
\noindent Next, note that due to consistency
\vspace{-0.3cm}
\begin{align*}
(\hat{\svec}_1,\hat{\vvec}_1,\hat{\xvec}_1,\hat{\xvec}_3,\mathbf{z}_b)\inA_{\epsilon}^{*(n)} \quad \Rightarrow \quad (\hat{\svec}_1,\mathbf{z}_b)\inA_{\epsilon}^{*(n)},
\end{align*}
\vspace{-0.2cm}
\noindent hence if $\hat{\svec}_1\notinA_{\epsilon}^{*(n)}(S_1|\mathbf{z}_b)$, then $(\hat{\svec}_1,\hat{\vvec}_1,\hat{\xvec}_1,\hat{\xvec}_3)\notinA_{\epsilon}^{*(n)}(S_1,V_1,X_1,X_3|\mathbf{z}_b)$, and we therefore can restrict the summation over $\hat{\svec}_1$ to the set $A_{\epsilon}^{*(n)}(S_1|\mathbf{z}_b)$.
Step (b) follows as when $\hat{\svec}_1 \mspace{-3mu} \in \mspace{-3mu} A_{\epsilon}^{*(n)} \mspace{-2mu} \big( \mspace{-2mu} S_1 \mspace{-2mu} \big|\mathbf{z}_b \big)$,
then joint typicality is achieved when:
\vspace{-0.2cm}
\[
\big( \hat{\mathbf{V}}_1(\hat{u}_{1}), \hat{\mathbf{X}}_1(\hat{\svec}_{1},\hat{u}_{1}), \hat{\mathbf{X}}_3(\hat{u}_{1},u_{2,b-1})\big) \in A_{\epsilon}^{*(n)}\big(V_1,X_1,X_3\big|\hat{\svec}_{1}, \mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big).
\]
\vspace{-0.2cm}
Next, we bound
\vspace{-0.2cm}
\begin{align*}
\Pr & \Big( \big(\hat{\mathbf{V}}_1(\hat{u}_{1}), \hat{\mathbf{X}}_1(\hat{\svec}_{1},\hat{u}_{1}), \hat{\mathbf{X}}_3(\hat{u}_{1},u_{2,b-1})\big) \in A_{\epsilon}^{*(n)}\big(V_1,X_1,X_3\big|\hat{\svec}_{1}, \mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\Big)\\
& = \mspace{-20mu} \sum_{ \substack{\big( \hat{\vvec}_1(\hat{u}_{1}), \hat{\xvec}_1(\hat{\svec}_{1},\hat{u}_{1}), \hat{\xvec}_3(\hat{u}_{1},u_{2,b-1})\big) \in\\
A_{\epsilon}^{*(n)}\big(V_1,X_1,X_3\big|\hat{\svec}_{1}, \mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)}} \mspace{-160mu}
p\big( \hat{\vvec}_1(\hat{u}_{1}), \hat{\xvec}_1(\hat{\svec}_{1},\hat{u}_{1}), \hat{\xvec}_3(\hat{u}_{1},u_{2,b-1})\big| \hat{\svec}_{1}, \mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big)\\
& \stackrel{(a)}{\le} 2^{n\big(H(V_1,X_1,X_3|S_1,S_2,V_2,X_2,W,Y)+\epsilon_0\big)} 2^{-n\big(H(V_1,X_1,X_3|S_1,V_2) -\epsilon_1 \big)},
\end{align*}
\vspace{-0.2cm}
\noindent where (a) follows from the properties of conditionally typical sequences, \cite[Theorem 6.9]{YeungBook} and \cite[Theorem 6.10]{YeungBook}.
Thus, we have:
\vspace{-0.2cm}
\begin{align*}
\Pr & \big(\mathcal{E}_{2,b}^{(7)} \big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big) \\
& \le \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-120mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b ) \times \\
& \mspace{50mu} \sum_{\substack{\hat{u}_1\in\mathcal{U}_1,\\ \hat{u}_1 \ne u_{1,b-1}}} \mspace{-100mu} \sum_{ \mspace{120mu} \substack{\hat{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big), \\ \hat{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-170mu} 2^{-nR_1}
2^{n\big(H(V_1,X_1,X_3|S_1,S_2,V_2,X_2,W,Y)+\epsilon_0\big)} 2^{-n\big(H(V_1,X_1,X_3|S_1,V_2) -\epsilon_1 \big)}\\
& = 2^{n\big(H(S_1,V_1,X_1,X_3|S_{2},V_2,X_2,W,Y)+2\epsilon_0\big)} 2^{-n\big(H(V_1,X_1,X_3|S_1,V_2) -\epsilon_1 \big)},
\end{align*}
\vspace{-0.2cm}
\noindent which implies that in order to get an arbitrarily small probability of error as $n$ increases, it must hold that:
\vspace{-0.2cm}
\begin{align*}
H(S_1,V_1,X_1,X_3|S_{2},V_2,X_2,W,Y) - H(V_1,X_1,X_3|S_1,V_2) + 3\epsilon_0 < 0.
\end{align*}
\vspace{-0.2cm}
\noindent Note that $H(S_1,V_1,X_1,X_3|S_{2},V_2,X_2,W,Y) - H(V_1,X_1,X_3|S_1,V_2)$ can also be written as
\vspace{-0.2cm}
\begin{align*}
& H(S_1,V_1,X_1,X_3|S_{2},V_2,X_2,W,Y) - H(V_1,X_1,X_3|S_1,V_2) \\
%
& \qquad = H(S_1,V_1,X_1,X_3|S_2, V_2,X_2,W, Y ) - H(S_1, V_1,X_1,X_3|V_2) + H(S_1|V_2) \\
%
& \qquad \stackrel{(a)}{=} H(S_1) - I(S_1,V_1,X_1,X_3;S_2,X_2,W, Y| V_2 ) \\
%
& \qquad = H(S_1) - I(S_1,V_1,X_1,X_3;S_2,X_2,W| V_2 ) - I(S_1,V_1,X_1,X_3;Y| S_2, V_2,X_2,W) \\
%
& \qquad = H(S_1) - I(S_1;S_2,X_2,W| V_2 ) - I(V_1,X_1,X_3;S_2,X_2,W|S_1, V_2 ) - I(S_1,V_1,X_1,X_3;Y| S_2, V_2,X_2,W) \\
%
& \qquad \stackrel{(b)}{=} H(S_1) - H(S_1|V_2) + H(S_1|S_2, V_2,X_2,W) - I(S_1,V_1,X_1,X_3;Y| S_2, V_2,X_2,W) \\
%
& \qquad \stackrel{(c)}{=} H(S_1|S_2,W) - I(X_1,X_3;Y| S_2, V_2,X_2,W),
\end{align*}
\vspace{-0.15cm}
\noindent where (a) follows form the independence $S_1$ and $V_2$; (b) follows from the Markov relationship $(S_2,X_2,W) \leftrightarrow (S_1,V_2) \leftrightarrow (V_1,X_1,X_3)$; and (c) follows from the Markov relationship $(V_2,X_2) \leftrightarrow (S_2,W) \leftrightarrow S_1$ and from the Markov relationship $(S_1,V_1) \leftrightarrow (S_2,V_2,X_1,X_2,X_3,W) \leftrightarrow Y$. Therefore, we conclude that as long as:
\vspace{-0.2cm}
\begin{equation}
H(S_1|S_2,W) < I(X_1,X_3;Y|S_2,V_2,X_2,W) - 3\epsilon_0,
\label{eq:ErrAnalysisCondSimult_s1u1}
\end{equation}
\vspace{-0.2cm}
\noindent then $\Pr\big(\mathcal{E}_{2,b}^{(7)}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$ can be made arbitrarily small by taking $n$ large enough.
Using similar arguments we can show that $\Pr\big(\mathcal{E}_{2,b}^{(m)}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big), m=4,5,6,8,9...,15$, can be made arbitrarily small by taking $n$ large enough, if the following conditions are satisfied correspondingly:
\vspace{-0.2cm}
\begin{subequations} \label{eq:ErrAnalysisCondSimult_8-15}
\begin{align}
H(S_1|S_2,W) & < I(X_1;Y|S_2,V_1,X_2,X_3,W) +R_1 - 3\epsilon_3 \label{eq:ErrAnalysisCondSimult_s1} \\
H(S_2|S_1,W) & < I(X_2;Y|S_1,V_2,X_1,X_3,W) +R_2 - 3\epsilon_3 \label{eq:ErrAnalysisCondSimult_s2} \\
H(S_1,S_2|W) & < I(X_1,X_2;Y|V_1,V_2,X_3,W) + R_1 + R_2 - 3\epsilon_3, \label{eq:ErrAnalysisCondSimult_s1s2} \\
H(S_2|S_1,W) & < I(X_2,X_3;Y|S_1,V_1,X_1,W) - 3\epsilon_4 \label{eq:ErrAnalysisCondSimult_8} \\
R_2 + H(S_1|S_2,W) & < I(X_1,X_2,X_3;Y|S_2,V_1,W) + R_1 - 3\epsilon_5 \label{eq:ErrAnalysisCondSimult_9} \\
R_1 + H(S_2|S_1,W) & < I(X_1,X_2,X_3;Y|S_1,V_2,W) + R_2 - 3\epsilon_5 \label{eq:ErrAnalysisCondSimult_10} \\
R_2 + H(S_1|S_2,W) & < I(X_1,X_2,X_3;Y|S_2,W) - 3\epsilon_6 \label{eq:ErrAnalysisCondSimult_11} \\
R_1 + H(S_2|S_1,W) & < I(X_1,X_2,X_3;Y|S_1,W) - 3\epsilon_6 \label{eq:ErrAnalysisCondSimult_12} \\
H(S_1,S_2|W) & < I(X_1,X_2,X_3;Y|V_2,W) + R_2 - 3\epsilon_{7} \label{eq:ErrAnalysisCondSimult_13} \\
H(S_1,S_2|W) & < I(X_1,X_2,X_3;Y|V_1,W) + R_1 - 3\epsilon_{7} \label{eq:ErrAnalysisCondSimult_14} \\
H(S_1,S_2|W) & < I(X_1,X_2,X_3;Y|W) - 3\epsilon_{8}. \label{eq:ErrAnalysisCondSimult_15}
\end{align}
\end{subequations}
%
%
%
%
%
%
%
\vspace{-0.2cm}
\noindent Now, define $\epsilon' = \max \{\epsilon_0, \epsilon_1, \dots, \epsilon_8 \}$, then it follows that constraints \eqref{eq:ErrAnalysisCondSimult_1-3}--\eqref{eq:ErrAnalysisCondSimult_8-15} hold with $\epsilon_k, k=0,1,\dots,8$, replaced by $\epsilon'$.
Finally, by using Fourier-Motzkin algorithm to eliminate $R_1$ and $R_2$ from the constraints \eqref{eq:ErrAnalysisCondSimult_1-3}--\eqref{eq:ErrAnalysisCondSimult_8-15}, we obtain \eqref{bnd:JointNewSimult_dst_S1}--\eqref{bnd:JointNewSimult_dst_S1S2}.
\ifthenelse{\boolean{SquizFlag}}{}{\end{comment}}
\ifthenelse{\boolean{SquizFlag}}{\begin{comment}}{}
\subsubsection{Bounding $\Pr\big(\mathcal{E}_{u_1}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$}
Writing explicitly we obtain
\begin{align*}
\Pr & \Big( \exists \tilde{u}_1 \in \mathcal{U}_1, \tilde{u}_1 \ne U_{1,b-1}, \big(\mathbf{S}_{1,b}, \mathbf{S}_{2,b}, \tilde{\mathbf{V}}_1(\tilde{u}_1), \mathbf{V}_2(U_{2,b-1}), \tilde{\mathbf{X}}_1(\mathbf{S}_{1,b},\tilde{u}_1), \mathbf{X}_2(\mathbf{S}_{2,b},U_{2,b-1}),\\
& \qquad \tilde{\mathbf{X}}_3(\tilde{u}_1,U_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b \big)\inA_{\epsilon}^{*(n)} \Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big)\\
& = \sum_{\substack{u_{1,b-1}\in\mathcal{U}_1 \\ u_{2,b-1}\in\mathcal{U}_2}} p_{U_1 U_2}(u_{1,b-1},u_{2,b-1}) \Pr\Big( \exists \tilde{u}_1 \in \mathcal{U}_1, \tilde{u}_1 \ne u_{1,b-1}, \big(\mathbf{S}_{1,b}, \mathbf{S}_{2,b}, \tilde{\mathbf{V}}_1(\tilde{u}_1), \mathbf{V}_2(u_{2,b-1}), \\
& \mspace{170mu} \tilde{\mathbf{X}}_1(\mathbf{S}_{1,b},\tilde{u}_1), \mathbf{X}_2(\mathbf{S}_{2,b},u_{2,b-1}), \tilde{\mathbf{X}}_3(\tilde{u}_1,u_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b \big)\inA_{\epsilon}^{*(n)} \Big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\Big).
\end{align*}
\noindent We now bound
\begin{align*}
& \Pr\Big( \exists \tilde{u}_1 \in \mathcal{U}_1, \tilde{u}_1 \ne u_{1,b-1}, \big(\mathbf{S}_{1,b}, \mathbf{S}_{2,b}, \tilde{\mathbf{V}}_1(\tilde{u}_1), \mathbf{V}_2(u_{2,b-1}), \\
& \mspace{50mu} \tilde{\mathbf{X}}_1(\mathbf{S}_{1,b},\tilde{u}_1), \mathbf{X}_2(\mathbf{S}_{2,b},u_{2,b-1}), \tilde{\mathbf{X}}_3(\tilde{u}_1, u_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b \big)\inA_{\epsilon}^{*(n)} \Big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c \Big)\\
& = \sum_{\substack{\big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}),\\\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}),\mathbf{w}_b,\mathbf{y}_b\big)\inA_{\epsilon}^{*(n)}} } \mspace{-50mu} p\big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}),\mathbf{w}_b,\mathbf{y}_b\big) \times \\
& \mspace{110mu} \Pr\Big( \exists \tilde{u}_1 \in \mathcal{U}_1, \tilde{u}_1 \ne u_{1,b-1}, \big(\tilde{\mathbf{V}}_1(\tilde{u}_1), \tilde{\mathbf{X}}_1(\mathbf{s}_{1,b},\tilde{u}_1), \tilde{\mathbf{X}}_3(\tilde{u}_1, u_{2,b-1}) \big) \in \\
& \mspace{200mu} A_{\epsilon}^{*(n)}(V_1, X_1, X_3 |\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}) , \mathbf{w}_b, \mathbf{y}_b ) \Big)\\
& = \sum_{\substack{\big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}),\\\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}),\mathbf{w}_b,\mathbf{y}_b\big)\inA_{\epsilon}^{*(n)}} } \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1\ne u_{1,b-1}}}
p\big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}),\mathbf{w}_b,\mathbf{y}_b\big) \times \\
& \mspace{110mu} \Pr\Big(\big(\tilde{\mathbf{V}}_1(\tilde{u}_1), \tilde{\mathbf{X}}_1(\mathbf{s}_{1,b},\tilde{u}_1), \tilde{\mathbf{X}}_3(\tilde{u}_1, u_{2,b-1}) \big) \in \\
& \mspace{150mu} A_{\epsilon}^{*(n)}(V_1, X_1, X_3 |\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}) , \mathbf{w}_b, \mathbf{y}_b ) \Big)\\
& = \sum_{\substack{\big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}),\\ \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}),\mathbf{w}_b,\mathbf{y}_b\big)\inA_{\epsilon}^{*(n)}} } \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1\ne u_{1,b-1}}}
p\big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}),\mathbf{w}_b,\mathbf{y}_b\big) \times \\
& \mspace{-50mu} \sum_{\mspace{50mu}\substack{\big(\tilde{\mathbf{v}}_1(\tilde{u}_1), \tilde{\mathbf{x}}_1(\mathbf{s}_{1,b},\tilde{u}_1), \tilde{\mathbf{x}}_3(\tilde{u}_1, u_{2,b-1})\big)\in \\ A_{\epsilon}^{*(n)}(V_1, X_1, X_3 |\mathbf{s}_{1,b}, \mathbf{s}_{2,b},
\mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}) , \mathbf{w}_b, \mathbf{y}_b ) } } \mspace{-180mu} p\big( \tilde{\mathbf{v}}_1(\tilde{u}_1), \tilde{\mathbf{x}}_1(\mathbf{s}_{1,b},\tilde{u}_1),
\tilde{\mathbf{x}}_3(\tilde{u}_1, u_{2,b-1}) |\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}) , \mathbf{w}_b, \mathbf{y}_b \big) \\
& \stackrel{(a)}{ = } \sum_{\substack{\big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}),\\\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}),\mathbf{w}_b,\mathbf{y}_b\big)\inA_{\epsilon}^{*(n)}} } \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1\ne u_{1,b-1}}}
p\big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}),\mathbf{w}_b,\mathbf{y}_b\big) \times \\
& \qquad \qquad \qquad \sum_{\substack{\big(\tilde{\mathbf{v}}_1(\tilde{u}_1), \tilde{\mathbf{x}}_1(\mathbf{s}_{1,b},\tilde{u}_1), \tilde{\mathbf{x}}_3(\tilde{u}_1, u_{2,b-1})\big)\inA_{\epsilon}^{*(n)}(V_1, X_1, X_3 |\mathbf{s}_{1,b}, \\ \mathbf{s}_{2,b},
\mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}) , \mathbf{w}_b, \mathbf{y}_b ) } } \mspace{-150mu} p\big( \tilde{\mathbf{v}}_1(\tilde{u}_1), \tilde{\mathbf{x}}_1(\mathbf{s}_{1,b},\tilde{u}_1),
\tilde{\mathbf{x}}_3(\tilde{u}_1, u_{2,b-1}) |\mathbf{s}_{1,b}, \mathbf{v}_2(u_{2,b-1})\big) \\
& \stackrel{(b)}{ \le } \sum_{\substack{\big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}),\\\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}),\mathbf{w}_b,\mathbf{y}_b\big)\inA_{\epsilon}^{*(n)}} } \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1\ne u_{1,b-1}}}
p\big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}),\mathbf{w}_b,\mathbf{y}_b\big) \times \\
& \qquad 2^{n\big(H(V_1, X_1, X_3 |S_1, S_2, V_2, X_2, W, Y ) + \epsilon_5 \big) } 2^{-n\big(H(V_1, X_1, X_3 |S_1, V_2) -\epsilon_5\big)}\\
& \le 2^{nR_1} 2^{n\big(H(V_1, X_1, X_3 |S_1, S_2, V_2, X_2, W, Y ) + \epsilon_5 \big) } 2^{-n\big(H(V_1, X_1, X_3 |S_1, V_2) -\epsilon_5\big)}\\
& = 2^{-n\big(I(X_1, X_3 ; Y | S_1, V_2, X_2 ) - R_1 + 2\epsilon_5 \big) }.
\end{align*}
\noindent
To see (a) we write
\begin{align*}
& p\big( \tilde{\mathbf{v}}_1(\tilde{u}_1), \tilde{\mathbf{x}}_1(\mathbf{s}_{1,b},\tilde{u}_1), \tilde{\mathbf{x}}_3(\tilde{u}_1, u_{2,b-1}), \mathbf{v}_1(u_{1,b-1}), \mathbf{x}_1(\mathbf{s}_{1,b},u_{1,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}) |\mathbf{s}_{1,b}, \mathbf{v}_2(u_{2,b-1}) \big)\\
& \quad =\frac{1}{p\big(\mathbf{s}_{1,b}, \mathbf{v}_2(u_{2,b-1})\big)} p\big( \tilde{\mathbf{v}}_1(\tilde{u}_1), \tilde{\mathbf{x}}_1(\mathbf{s}_{1,b},\tilde{u}_1), \tilde{\mathbf{x}}_3(\tilde{u}_1, u_{2,b-1}), \mathbf{v}_1(u_{1,b-1}), \\
& \mspace{220mu} \mathbf{x}_1(\mathbf{s}_{1,b},u_{1,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}),\mathbf{s}_{1,b}, \mathbf{v}_2(u_{2,b-1}) \big)\\
& \quad = \frac{1}{p\big(\mathbf{s}_{1,b}, \mathbf{v}_2(u_{2,b-1})\big)} \prod_{i=1}^n p\big(s_{1,b,i}\big) p\big( \tilde{v}_{1,i}(\tilde{u}_1)\big)p\big(v_{1,i}(u_{1,b-1})\big)p\big(v_{2,i}(u_{2,b-1})\big)
p\big(\tilde{x}_{1,i}(s_{1,b,i},\tilde{u}_1)| s_{1,b,i}, \tilde{v}_{1,i}(\tilde{u}_1) \big) \times\\
& \qquad \qquad \qquad \qquad\qquad \qquad p\big( \tilde{x}_{3,i}(\tilde{u}_1, u_{2,b-1}) | \tilde{v}_{1,i}(\tilde{u}_1), v_{2,i}(u_{2,b-1})\big) p\big( x_{1,i}(s_{1,b,i},u_{1,b-1})\big| s_{1,b,i}, v_{1,i}(u_{1,b-1}) \big)\times\\
& \qquad \qquad \qquad\qquad \qquad\qquad \qquad p\big(x_{3,i}(u_{1,b-1}, u_{2,b-1})\big|
v_{1,i}(u_{1,b-1}), v_{2,i}(u_{2,b-1}) \big)\\
& \quad = \prod_{i=1}^n p\big( \tilde{v}_{1,i}(\tilde{u}_1)\big)
p\big(\tilde{x}_{1,i}(s_{1,b,i},\tilde{u}_1)| s_{1,b,i}, \tilde{v}_{1,i}(\tilde{u}_1) \big) \times\\
& \qquad \qquad \quad p\big( \tilde{x}_{3,i}(\tilde{u}_1, u_{2,b-1}) | \tilde{v}_{1,i}(\tilde{u}_1), v_{2,i}(u_{2,b-1})\big)
p\big(v_{1,i}(u_{1,b-1})\big) \times \\
& \qquad \qquad \qquad p\big( x_{1,i}(s_{1,b,i},u_{1,b-1})\big| s_{1,b,i}, v_{1,i}(u_{1,b-1}) \big) p\big(x_{3,i}(u_{1,b-1}, u_{2,b-1})\big|
v_{1,i}(u_{1,b-1}), v_{2,i}(u_{2,b-1}) \big),
\end{align*}
\noindent which implies $(V_1,X_1,X_3) \leftrightarrow (S_1, V_2) \leftrightarrow (\tilde{V}_1, \tilde{X}_1, \tilde{X}_3)$. Finally (b) follows from the properties of conditional strongly typical sets and joint strongly typical sets, \cite[Theorem 6.9]{YeungBook} and \cite[Theorem 6.10]{YeungBook}.
We conclude that the probability $\Pr\big(\mathcal{E}_{u_1}\big|\mathcal{E}_{1,b}^c\cap F_{b+1}^c\big)$ can be made arbitrarily small by increasing the block length as long as
\begin{equation}
R_1 < I(X_1, X_3 ; Y | S_1, V_2, X_2 ) - 2\epsilon_5.
\label{eq:ErrAnalysisCondSimult_u1}
\end{equation}
Following similar arguments it can be shown that $\Pr\big(\mathcal{E}_{u_2}\big|\mathcal{E}_{1,b}^c\cap F_{b+1}^c\big)$ can be made arbitrarily small by taking $n$ large enough as long as
\begin{equation}
R_2 < I(X_2, X_3 ; Y | S_2, V_1, X_1 ) - 2\epsilon_5.
\label{eq:ErrAnalysisCondSimult_u2}
\end{equation}
and that $\Pr\big(\mathcal{E}_{u_1,u_2}\big|\mathcal{E}_{1,b}^c\cap F_{b+1}^c\big)$ can be made arbitrarily small by taking $n$ large enough as long as
\begin{equation}
R_1 + R_2 < I(X_1,X_2,X_3;Y|S_1,S_2) - 2\epsilon_5.
\label{eq:ErrAnalysisCondSimult_u1u2}
\end{equation}
\subsubsection{Bounding $\Pr\big(\mathcal{E}_{\mathbf{s}_1}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$}
We begin by writing
\begin{align}
\Pr & \Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = U_{1,b}, \big(\tilde{\svec}_1, \mathbf{S}_{2,b}, \mathbf{V}_1(U_{1,b-1}), \mathbf{V}_2(U_{2,b-1}), \tilde{\mathbf{X}}_1(\tilde{\svec}_1,U_{1,b-1}), \nonumber\\
& \qquad \quad \mathbf{X}_2(\mathbf{S}_{2,b},U_{2,b-1}), \mathbf{X}_3(U_{1,b-1}, U_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)} \Big| \mathcal{E}_{1,b}^c\cap\mathcal{F}_{b+1}^c\Big)\nonumber\\
& = \sum_{u_{1,b}\in \mathcal{U}_1, u_{1,b-1}\in\mathcal{U}_1, u_{2,b-1}\in\mathcal{U}_2}
p_{U_1}(u_{1,b}) p_{U_1 U_2}(u_{1,b-1},u_{2,b-1})\times \nonumber\\
& \mspace{100mu} \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = u_{1,b}, \big(\tilde{\svec}_1, \mathbf{S}_{2,b}, \mathbf{V}_1(u_{1,b-1}), \mathbf{V}_2(u_{2,b-1}), \nonumber\\
& \mspace{160mu} \tilde{\mathbf{X}}_1(\tilde{\svec}_1,u_{1,b-1}), \mathbf{X}_2(\mathbf{S}_{2,b},u_{2,b-1}), \mathbf{X}_3(u_{1,b-1}, u_{2,b-1})\big) \in A_{\epsilon}^{*(n)} \Big| \mathcal{E}_{1,b}^c\cap\mathcal{F}_{b+1}^c\Big)\nonumber.
\end{align}
\noindent We now bound
\begin{align}
& \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = u_{1,b}, \big(\tilde{\svec}_1, \mathbf{S}_{2,b}, \mathbf{V}_1(u_{1,b-1}), \mathbf{V}_2(u_{2,b-1}), \nonumber \\
& \qquad \qquad \qquad \tilde{\mathbf{X}}_1(\tilde{\svec}_1,u_{1,b-1}), \mathbf{X}_2(\mathbf{S}_{2,b},u_{2,b-1}), \mathbf{X}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)} \Big| \mathcal{E}_{1,b}^c\cap\mathcal{F}_{b+1}^c\Big) \nonumber \\
& \stackrel{(a)}{=} \mspace{-20mu} \sum_{\substack{\big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \mathbf{v}_2(u_{2,b-1}), \\ \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\inA_{\epsilon}^{*(n)}}}
\mspace{-120mu }p\big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big)\times\nonumber\\
& \qquad \qquad \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{s}_{1,b}, f_1(\tilde{\svec}_1) = u_{1,b}, \big(\tilde{\svec}_1, \tilde{\mathbf{X}}_1(\tilde{\svec}_1,u_{1,b-1}) \big) \in A_{\epsilon}^{*(n)}\big(S_1,X_1 \big| \mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \nonumber\\
& \qquad \qquad \qquad \qquad \mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b, \mathbf{y}_b \big) \Big)\nonumber\\
& \stackrel{(b)}{=} \mspace{-20mu} \sum_{\substack{\big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}), \mathbf{v}_2(u_{2,b-1}), \\ \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\inA_{\epsilon}^{*(n)}}}
\mspace{-120mu }p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big)\times\nonumber\\
& \qquad \quad \sum_{\substack{\tilde{\svec}_1\in A_{\epsilon}^{*(n)}\big(S_1\big| \mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}),
\mathbf{v}_2(u_{2,b-1}),\\ \quad\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b, \mathbf{y}_b \big), \tilde{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-120mu} \Pr\Big( f_1(\tilde{\svec}_1) = u_{1,b}, \big(\tilde{\svec}_1, \tilde{\mathbf{X}}_1(\tilde{\svec}_1,u_{1,b-1}) \big) \in A_{\epsilon}^{*(n)}\big(S_1,X_1 \big| \mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \nonumber\\
& \mspace{300mu} \mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b, \mathbf{y}_b \big) \Big)\nonumber\\
& \stackrel{(c)}{=} \mspace{-20mu} \sum_{\substack{\big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}), \mathbf{v}_2(u_{2,b-1}), \\ \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\inA_{\epsilon}^{*(n)}}}
\mspace{-120mu }p\big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big)\times\nonumber\\
& \qquad \sum_{\substack{\tilde{\svec}_1\in A_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}),
\mathbf{v}_2(u_{2,b-1}),\\ \quad\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b, \mathbf{y}_b \big), \tilde{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-120mu} 2^{-nR_1} \Pr\Big( \tilde{\mathbf{X}}_1(\tilde{\svec}_1,u_{1,b-1}) \in A_{\epsilon}^{*(n)}\big(X_1 \big|\tilde{\svec}_1,\mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \nonumber\\
\label{eqn:derivation_mutual_inform_1}
& \mspace{340mu} \mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b, \mathbf{y}_b \big) \Big),
\end{align}
\noindent where (a) follows due to conditioning on $\mathcal{E}_{b,1}^c$ which implies that the sequences at block $b$ are jointly typical; (b) follows as if
\[
\tilde{\svec}_1\notin A_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}),
\mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b, \mathbf{y}_b \big),
\]
then
\[
\big(\tilde{\svec}_1, \tilde{\mathbf{X}}_1(\tilde{\svec}_1,u_{1,b-1}) \big) \notin A_{\epsilon}^{*(n)}\big(S_1,X_1 \big| \mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b, \mathbf{y}_b \big),
\]
see also argument leading to \eqref{eq:Consistency_step}. Finally, (c) follows as when $\tilde{\svec}_1\in A_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}),
\mathbf{v}_2(u_{2,b-1}),$ $\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b, \mathbf{y}_b \big)$, then the joint conditional typicality of $\tilde{\svec}_1$ and $\tilde{\mathbf{X}}_1(\tilde{\svec}_1,u_{1,b-1})$ is achieved when
\[
\tilde{\mathbf{X}}_1(\tilde{\svec}_1,u_{1,b-1}) \in A_{\epsilon}^{*(n)}\big(X_1 \big|\tilde{\svec}_1, \mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b, \mathbf{y}_b \big),
\]
see also the argument leading to \eqref{eq:Summing_only_over_x_step}.
Next, we evaluate
\begin{align*}
\Pr & \Big( \tilde{\mathbf{X}}_1(\tilde{\svec}_1,u_{1,b-1}) \in A_{\epsilon}^{*(n)}\big(X_1 \big|\tilde{\svec}_1, \mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b, \mathbf{y}_b \big) \Big)\\
& \stackrel{(a)}{=} \sum_{\substack{\tilde{\mathbf{x}}_1(\tilde{\svec}_1,u_{1,b-1}) \in A_{\epsilon}^{*(n)}\big(X_1 \big|\tilde{\svec}_1, \mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \\ \mathbf{v}_2(u_{2,b-1}),\mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{x}_3(u_{1,b-1}, u_{2,b-1}), \mathbf{w}_b, \mathbf{y}_b \big)}} \mspace{-130mu}
p(\tilde{\mathbf{x}}_1(\tilde{\svec}_1,u_{1,b-1}) |\tilde{\svec}_1, \mathbf{v}_1(u_{1,b-1}))\\
& \stackrel{(b)}{\le} 2^{n\big(H(X_1|S_1,S_2,V_1,V_2,X_2,X_3,W,Y)+\epsilon_6 \big)} 2^{-n\big(H(X_1|S_1,V_1)-\epsilon_6 \big)}
\end{align*}
\noindent To show (a) we write
\begin{align*}
p & \big(\tilde{\mathbf{x}}_1(\tilde{\svec}_1,u_{1,b-1}), \mathbf{x}_{1}(\mathbf{s}_{1,b},u_{1,b-1}) \big|\tilde{\svec}_1, \mathbf{s}_{1,b}, \mathbf{v}_1(u_{1,b-1}) \big)\\
& = \frac{1}{p\big(\tilde{\svec}_1, \mathbf{s}_{1,b}, \mathbf{v}_1(u_{1,b-1}) \big)}p\big(\tilde{\mathbf{x}}_1(\tilde{\svec}_1,u_{1,b-1}), \mathbf{x}_{1}(\mathbf{s}_{1,b},u_{1,b-1}) , \tilde{\svec}_1, \mathbf{s}_{1,b}, \mathbf{v}_1(u_{1,b-1}) \big)\\
& = \frac{1}{p\big(\tilde{\svec}_1, \mathbf{s}_{1,b}, \mathbf{v}_1(u_{1,b-1}) \big)}\prod_{i=1}^n p\big(\tilde{x}_{1,i}(\tilde{s}_{1,i},u_{1,b-1}), x_{1,i}(s_{1,b,i},u_{1,b-1}) , \tilde{s}_{1,i}, s_{1,b,i}, v_{1,i}(u_{1,b-1}) \big)\\
& = \frac{1}{p\big(\tilde{\svec}_1, \mathbf{s}_{1,b}, \mathbf{v}_1(u_{1,b-1}) \big)}\prod_{i=1}^n p\big(\tilde{s}_{1,i}, s_{1,b,i}, v_{1,i}(u_{1,b-1}) \big)p\big(\tilde{x}_{1,i}(\tilde{s}_{1,i},u_{1,b-1}) \big| \tilde{s}_{1,i}, v_{1,i}(u_{1,b-1}) \big) \times\\
& \qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad p\big( x_{1,i}(s_{1,b,i},u_{1,b-1}) | s_{1,b,i}, v_{1,i}(u_{1,b-1}) \big)\\
& = \prod_{i=1}^n p\big(\tilde{x}_{1,i}(\tilde{s}_{1,i},u_{1,b-1}) \big| \tilde{s}_{1,i}, v_{1,i}(u_{1,b-1}) \big) \times\prod_{i=1}^n p\big( x_{1,i}(s_{1,b,i},u_{1,b-1}) | s_{1,b,i}, v_{1,i}(u_{1,b-1}) \big),
\end{align*}
\noindent which implies that $\mathbf{Y}_b \independent \tilde{\mathbf{X}}_1(\tilde{\svec}_1,u_{1,b-1})$, when $\tilde{\svec}_1 \ne \mathbf{s}_{1,b}$ and $\mathbf{v}_1(u_{1,b-1})$ are given. Step (b) follows from the properties of conditional strongly typical sets, \cite[Theorem 6.9]{YeungBook}
and \cite[Theorem 6.10]{YeungBook}.
\noindent Therefore, plugging this back into \eqref{eqn:derivation_mutual_inform_1} we obtain
\begin{align*}
\Pr\big(\mathcal{E}_{\mathbf{s}_1}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big) & \le 2^{n\big(H(S_1|S_2,V_1,V_2,X_2,X_3,W,Y)+\epsilon_6\big)} 2^{-nR_1} \times \\
& \qquad 2^{n\big(H(X_1|S_1,S_2,V_1,V_2,X_2,X_3,W,Y)+\epsilon_6 \big)} 2^{-n\big(H(X_1|S_1,V_1)-\epsilon_6 \big)}\\
%
& \stackrel{(a)}{=} 2^{-n\big(R_1 - H(S_1|S_2,W) + I(X_1;Y|S_2,V_1,X_2,X_3,W)-3\epsilon_6\big)}
\end{align*}
\noindent where (a) follows from the same steps leading to \eqref{proof:last step_R_u1}. We conclude that the probability $\Pr\big(\mathcal{E}_{\mathbf{s}_1}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$
goes to zero asymptotically as the block length increases,
as long as $ R_1 - H(S_1|S_2,W) + I(X_1;Y|S_2,V_1,X_2,X_3,W) - \epsilon_6>0$, or, equivalently,
\begin{equation}
H(S_1|S_2,W) < I(X_1;Y|S_2,V_1,X_2,X_3,W) +R_1 - 3\epsilon_6.
\label{eq:ErrAnalysisCondSimult_s1}
\end{equation}
Following similar arguments it can be shown that $\Pr\big(\mathcal{E}_{\mathbf{s}_2}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$ can be made arbitrarily small by taking $n$ large enough as long as
\begin{equation}
H(S_2|S_1,W) < I(X_2;Y|S_1,V_2,X_1,X_3,W) +R_2 - 3\epsilon_6.
\label{eq:ErrAnalysisCondSimult_s2}
\end{equation}
and that $\Pr\big(\mathcal{E}_{\mathbf{s}_1,\mathbf{s}_2}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$ can be made arbitrarily small by taking $n$ large enough as long as
\begin{equation}
H(S_1,S_2|W) < I(X_1,X_2;Y|V_1,V_2,X_3,W) + R_1 + R_2 - 3\epsilon_6.
\label{eq:ErrAnalysisCondSimult_s1s2}
\end{equation}
\subsubsection{Bounding $\Pr\big(\mathcal{E}_{\mathbf{s}_1,u_2}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$}
Begin by writing
\begin{align*}
\Pr & \big(\mathcal{E}_{\mathbf{s}_1,u_2}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big) \\
& = \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = U_{1,b}, \exists \tilde{u}_2\in\mathcal{U}_2, \tilde{u}_2\ne U_{2,b-1}, \big(\tilde{\svec}_{1}, \mathbf{S}_{2,b}, \mathbf{V}_1(U_{1,b-1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \\
& \qquad \qquad \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},U_{1,b-1}), \tilde{\mathbf{X}}_2(\mathbf{S}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(U_{1,b-1}, \tilde{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big)\\
& = \sum_{u_{1,b}\in \mathcal{U}_1, u_{1,b-1}\in\mathcal{U}_1, u_{2,b-1}\in\mathcal{U}_2} \mspace{-60mu}
p_{U_1}(u_{1,b}) p_{U_1 U_2}(u_{1,b-1},u_{2,b-1})\times \nonumber\\
& \mspace{100mu} \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = u_{1,b}, \exists \tilde{u}_2\in\mathcal{U}_2, \tilde{u}_2\ne u_{2,b-1}, \big(\tilde{\svec}_{1}, \mathbf{S}_{2,b}, \mathbf{V}_1(u_{1,b-1}), \\
& \qquad \qquad \qquad \qquad \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},u_{1,b-1}), \tilde{\mathbf{X}}_2(\mathbf{S}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(u_{1,b-1}, \tilde{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big)
\end{align*}
\noindent We now bound
\begin{align*}
& \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = u_{1,b}, \exists \tilde{u}_2\in\mathcal{U}_2, \tilde{u}_2\ne u_{2,b-1}, \big(\tilde{\svec}_{1}, \mathbf{S}_{2,b}, \mathbf{V}_1(u_{1,b-1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \\
& \qquad \qquad \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},u_{1,b-1}), \tilde{\mathbf{X}}_2(\mathbf{S}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(u_{1,b-1}, \tilde{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big)\\
& \stackrel{(a)}{=} \sum_{\substack{\big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-90mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_2\in\mathcal{U}_2,\\ \tilde{u}_2 \ne u_{2,b-1}}} \mspace{-100mu} \sum_{\mspace{120mu} \substack{\tilde{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big| \mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \mathbf{w}_b, \mathbf{y}_b\big) \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-130mu} \Pr\Big( f_1(\tilde{\svec}_1) = u_{1,b}, \big(\tilde{\svec}_{1}, \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},u_{1,b-1}), \tilde{\mathbf{X}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(u_{1,b-1}, \tilde{u}_{2})\big) \in \\
& \mspace{280mu} A_{\epsilon}^{*(n)}\big(S_1,V_2,X_1,X_2,X_3\big|\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}),\mathbf{w}_b, \mathbf{y}_b\big)\Big)\\
& = \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-90mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}), \mathbf{w}_b,\mathbf{y}_b) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_2\in\mathcal{U}_2,\\ \tilde{u}_2 \ne u_{2,b-1}}} \mspace{-100mu} \sum_{\mspace{120mu} \substack{\tilde{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big| \mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \mathbf{w}_b, \mathbf{y}_b\big) \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-130mu} 2^{-nR_1} \Pr\Big(\big(\tilde{\svec}_{1}, \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},u_{1,b-1}),\tilde{\mathbf{X}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(u_{1,b-1}, \tilde{u}_{2})\big) \in \\
& \mspace{320mu} A_{\epsilon}^{*(n)}\big(S_1,V_2,X_1,X_2,X_3\big|\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}),\mathbf{w}_b, \mathbf{y}_b\big) \Big)\\
& \stackrel{(b)}{=} \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-90mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}), \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_2\in\mathcal{U}_2,\\ \tilde{u}_2 \ne u_{2,b-1}}} \mspace{-100mu} \sum_{\mspace{120mu} \substack{\tilde{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \mathbf{w}_b, \mathbf{y}_b\big) \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-120mu} 2^{-nR_1} \Pr\Big(\big(\tilde{\mathbf{V}}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},u_{1,b-1}), \tilde{\mathbf{X}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(u_{1,b-1}, \tilde{u}_{2})\big) \in \\
& \mspace{320mu} A_{\epsilon}^{*(n)}\big(V_2,X_1,X_2,X_3\big|\tilde{\svec}_{1},\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}),\mathbf{w}_b, \mathbf{y}_b\big)\Big)
\end{align*}
\noindent where (a) follows from the conditioning on $\mathcal{E}_{1,b}^c$ which implies that the sequences at block $b$ are jointly typical, and as if
\[
\tilde{\svec}_1\notin A_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}), \mathbf{w}_b, \mathbf{y}_b\big),
\]
then
\[
\big(\tilde{\svec}_{1}, \tilde{\mathbf{V}}_2(\tilde{u}_{2}),\tilde{\mathbf{X}}_1(\tilde{\svec}_{1},u_{1,b-1}), \tilde{\mathbf{X}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(u_{1,b-1}, \tilde{u}_{2})\big) \notin
A_{\epsilon}^{*(n)}\big(S_1,V_2,X_1,X_2,X_3\big|\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}),\mathbf{w}_b, \mathbf{y}_b\big),
\]
see also argument leading to \eqref{eq:Consistency_step}. Step (c) follows as when $\tilde{\svec}_1\in A_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}), \mathbf{w}_b, \mathbf{y}_b\big)$,
then joint typicality is achieved when
\[
\big( \tilde{\mathbf{V}}_2(\tilde{u}_{2}),\tilde{\mathbf{X}}_1(\tilde{\svec}_{1},u_{1,b-1}),\tilde{\mathbf{X}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(u_{1,b-1}, \tilde{u}_{2})\big) \in A_{\epsilon}^{*(n)}\big(V_2,X_1,X_2,X_3\big|\tilde{\svec}_{1},\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}),\mathbf{w}_b, \mathbf{y}_b\big).
\]
Next we bound
\begin{align*}
\Pr & \Big( \big(\tilde{\mathbf{V}}_2(\tilde{u}_{2}),\tilde{\mathbf{X}}_1(\tilde{\svec}_{1},u_{1,b-1}), \tilde{\mathbf{X}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(u_{1,b-1}, \tilde{u}_{2})\big) \in
A_{\epsilon}^{*(n)}\big(V_2,X_1,X_2,X_3\big|\tilde{\svec}_1,\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}),\mathbf{w}_b, \mathbf{y}_b\big)\Big)\\
& = \sum_{ \substack{\big(\tilde{\mathbf{v}}_2(\tilde{u}_{2}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},u_{1,b-1}), \tilde{\mathbf{x}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{x}}_3(u_{1,b-1}, \tilde{u}_{2})\big) \in\\
A_{\epsilon}^{*(n)}\big(V_2,X_1,X_2,X_3\big|\tilde{\svec}_1,\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}),\mathbf{w}_b, \mathbf{y}_b\big)}} \mspace{-130mu}
p\big( \tilde{\mathbf{v}}_2(\tilde{u}_{2}),\tilde{\mathbf{x}}_1(\tilde{\svec}_{1},u_{1,b-1}),\tilde{\mathbf{x}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{x}}_3(u_{1,b-1}, \tilde{u}_{2})\big| \tilde{\svec}_1,\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}),\mathbf{w}_b, \mathbf{y}_b \big)\\
& = \sum_{ \substack{\big(\tilde{\mathbf{v}}_2(\tilde{u}_{2}),\tilde{\mathbf{x}}_1(\tilde{\svec}_{1},u_{1,b-1}),\tilde{\mathbf{x}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{x}}_3(u_{1,b-1}, \tilde{u}_{2})\big) \in\\
A_{\epsilon}^{*(n)}\big(V_2,X_1,X_2,X_3\big|\tilde{\svec}_1,\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}),\mathbf{w}_b, \mathbf{y}_b\big)}} \mspace{-130mu}
p\big( \tilde{\mathbf{v}}_2(\tilde{u}_{2}),\tilde{\mathbf{x}}_1(\tilde{\svec}_{1},u_{1,b-1}), \tilde{\mathbf{x}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{x}}_3(u_{1,b-1}, \tilde{u}_{2})\big| \tilde{\svec}_1,\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1})\big)\\
& \le 2^{n\big(H(V_2,X_1,X_2,X_3|S_1,S_2,V_1,W,Y)+\epsilon_7\big)} 2^{-n\big(H(V_2,X_1,X_2,X_3|S_1,S_2,V_1) -\epsilon_7 \big)}.
\end{align*}
\noindent Thus, we have
\begin{align*}
\Pr & \big(\mathcal{E}_{\mathbf{s}_1,u_2}\big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\big) \\
& \le \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-80mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{v}_1(u_{1,b-1}), \mathbf{w}_b,\mathbf{y}_b \big)\times\\
& \qquad \sum_{\substack{\tilde{u}_2\in\mathcal{U}_2,\\ \tilde{u}_2 \ne u_{2,b-1}}} \sum_{\substack{\tilde{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}), \mathbf{w}_b, \mathbf{y}_b\big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-90mu} 2^{-nR_1}
2^{n\big(H(V_2,X_1,X_2,X_3|S_1,S_2,V_1,W,Y)+\epsilon_7\big)} 2^{-n\big(H(V_2,X_1,X_2,X_3|S_1,S_2,V_1) -\epsilon_7 \big)}\\
& \le \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-80mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_1(u_{1,b-1}), \mathbf{w}_b,\mathbf{y}_b|\mathcal{E}_{1,b}^c\big)\times\\
& \qquad 2^{nR_2} 2^{n\big(H(S_1|S_{2},V_1, W, Y)+\epsilon_7\big) }2^{-nR_1}
2^{n\big(H(V_2,X_1,X_2,X_3|S_1,S_2,V_1,W,Y)+\epsilon_7\big)} 2^{-n\big(H(V_2,X_1,X_2,X_3|S_1,S_2,V_1) -\epsilon_7 \big)}\\
& \le 2^{n(R_2-R_1)} 2^{n\big(H(S_1,V_2,X_1,X_2,X_3|S_2,V_1,W,Y)+2\epsilon_7\big)} 2^{-n\big(H(V_2,X_1,X_2,X_3|S_1,S_2,V_1) -\epsilon_7 \big)},
\end{align*}
\noindent which implies that in order to get an arbitrarily small probability of error as $n$ increases, it must hold that
\begin{eqnarray*}
& & R_2 - R_1 + H(S_1,V_2,X_1,X_1,X_3|S_2,V_1,W,Y) - H(V_2,X_1,X_2,X_3|S_1,S_2,V_1) + 3\epsilon_7 < 0.
\end{eqnarray*}
Now consider
\begin{align*}
& H(S_1,V_2,X_1,X_1,X_3|S_2,V_1,W,Y) - H(V_2,X_1,X_2,X_3|S_1,S_2,V_1) \\
& \quad = H(S_1,V_2,X_1,X_1,X_3|S_2,V_1,W,Y) - H(S_1,V_2,X_1,X_2,X_3|S_2,V_1) + H(S_1|S_2,V_1) \\
& \quad = H(S_1|S_2,V_1) - I(S_1,V_2,X_1,X_2,X_3;W,Y|S_2,V_1) \\
& \quad = H(S_1|S_2) - I(S_1,V_2,X_1,X_2,X_3;W|S_2,V_1) - I(S_1,V_2,X_1,X_2,X_3;Y|S_2,V_1,W) \\
& \quad \stackrel{(a)}{=} H(S_1|S_2) - I(S_1,V_2,X_1,X_2,X_3;W|S_2,V_1) - I(X_1,X_2,X_3;Y|S_2,V_1,W) \\
& \quad \stackrel{(b)}{=} H(S_1|S_2) - I(S_1;W|S_2) - I(X_1,X_2,X_3;Y|S_2,V_1,W) \\
& \quad = H(S_1|S_2,W) - I(X_1,X_2,X_3;Y|S_2,V_1,W),
\end{align*}
\noindent where (a) follows from the Markov chain $(S_1,V_2) \leftrightarrow (X_1,X_2,X_3) \leftrightarrow Y$ for given $(S_2,V_1,W)$; and (b) follows from the Markov chain $(V_1,V_2,X_1,X_2,X_3) \leftrightarrow (S_1,S_2) \leftrightarrow W$.
We conclude that as long as
\begin{equation}
R_2 + H(S_1|S_2,W) < I(X_1,X_2,X_3;Y|S_2,V_1,W) + R_1 - 3\epsilon_7,
\label{eq:ErrAnalysisCondSimult_s1u2}
\end{equation}
$\Pr\big(\mathcal{E}_{\mathbf{s}_1,u_2}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$ can be made arbitrarily small by taking $n$ large enough.
Using similar arguments we can show that as long as
\begin{equation}
R_1 + H(S_2|S_1,W) < I(X_1,X_2,X_3;Y|S_1,V_2,W) + R_2 - 3\epsilon_7,
\label{eq:ErrAnalysisCondSimult_s2u1}
\end{equation}
then $\Pr\big(\mathcal{E}_{\mathbf{s}_2,u_1}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$ can be made arbitrarily small by taking $n$ large enough.
\subsubsection{Bounding $\Pr\big(\mathcal{E}_{\mathbf{s}_1,u_1}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$}
Begin by writing
\begin{align*}
\Pr & \big(\mathcal{E}_{\mathbf{s}_1,u_1}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big) \\
& = \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = U_{1,b}, \exists \tilde{u}_1\in\mathcal{U}_1, \tilde{u}_1\ne U_{1,b-1}, \big(\tilde{\svec}_{1}, \mathbf{S}_{2,b}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \mathbf{V}_2(U_{2,b-1}), \\
& \qquad \qquad \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \mathbf{X}_2(\mathbf{S}_{2,b},U_{2,b-1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1}, U_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big)\\
& = \sum_{u_{1,b}\in \mathcal{U}_1, u_{1,b-1}\in\mathcal{U}_1, u_{2,b-1}\in\mathcal{U}_2} \mspace{-40mu}
p_{U_1}(u_{1,b}) p_{U_1 U_2}(u_{1,b-1},u_{2,b-1})\times \nonumber\\
& \mspace{100mu} \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = u_{1,b}, \exists \tilde{u}_1\in\mathcal{U}_1, \tilde{u}_1\ne u_{1,b-1}, \big(\tilde{\svec}_{1}, \mathbf{S}_{2,b}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \mathbf{V}_2(u_{2,b-1}), \\
& \qquad \qquad \qquad \qquad \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \mathbf{X}_2(\mathbf{S}_{2,b},u_{2,b-1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1}, u_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big)
\end{align*}
\noindent We now bound
\begin{align*}
& \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = u_{1,b}, \exists \tilde{u}_1\in\mathcal{U}_1, \tilde{u}_1\ne u_{1,b-1}, \big(\tilde{\svec}_{1}, \mathbf{S}_{2,b}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \mathbf{V}_2(u_{2,b-1}), \\
& \qquad \quad \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \mathbf{X}_2(\mathbf{S}_{2,b},u_{2,b-1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1}, u_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big) \\
& \stackrel{(a)}{=} \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-120mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1 \ne u_{1,b-1}}} \mspace{-100mu} \sum_{ \mspace{120mu} \substack{\tilde{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-170mu} \Pr\Big( f_1(\tilde{\svec}_1) = u_{1,b}, \big(\tilde{\svec}_{1}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},u_{2,b-1})\big) \in \\
& \mspace{240mu} A_{\epsilon}^{*(n)}\big(S_1,V_1,X_1,X_3\big|\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big) \Big)\\
& = \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-120mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1 \ne u_{1,b-1}}} \mspace{-100mu} \sum_{ \mspace{120mu} \substack{\tilde{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-170mu} 2^{-nR_1} \Pr\Big( \big(\tilde{\svec}_{1}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},u_{2,b-1})\big) \in \\
& \mspace{240mu} A_{\epsilon}^{*(n)}\big(S_1,V_1,X_1,X_3\big|\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big) \Big)\\
& \stackrel{(b)}{=} \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-120mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1 \ne u_{1,b-1}}} \mspace{-100mu} \sum_{ \mspace{120mu} \substack{\tilde{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-170mu} 2^{-nR_1} \Pr\Big( \big(\tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},u_{2,b-1})\big) \in \\
& \mspace{240mu} A_{\epsilon}^{*(n)}\big(V_1,X_1,X_3\big|\tilde{\svec}_{1}, \mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big) \Big)
\end{align*}
\noindent where (a) follows from the conditioning on $\mathcal{E}_{b,1}^c$ which implies that the sequences at block $b$ are jointly typical, and as if
\[
\tilde{\svec}_1\notin A_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big),
\]
then
\[
\big(\tilde{\svec}_{1}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},u_{2,b-1})\big) \notin A_{\epsilon}^{*(n)}\big(S_1,V_1,X_1,X_3\big|\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big),
\]
see also argument leading to \eqref{eq:Consistency_step}. Step (b) follows as when $\tilde{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big)$,
then joint typicality is achieved when
\[
\big( \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},u_{2,b-1})\big) \in A_{\epsilon}^{*(n)}\big(V_1,X_1,X_3\big|\tilde{\svec}_{1}, \mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big).
\]
Next we bound
\begin{align*}
\Pr & \Big( \big(\tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},u_{2,b-1})\big) \in A_{\epsilon}^{*(n)}\big(V_1,X_1,X_3\big|\tilde{\svec}_{1}, \mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\Big)\\
& = \mspace{-20mu} \sum_{ \substack{\big( \tilde{\mathbf{v}}_1(\tilde{u}_{1}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{x}}_3(\tilde{u}_{1},u_{2,b-1})\big) \in\\
A_{\epsilon}^{*(n)}\big(V_1,X_1,X_3\big|\tilde{\svec}_{1}, \mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)}} \mspace{-160mu}
p\big( \tilde{\mathbf{v}}_1(\tilde{u}_{1}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{x}}_3(\tilde{u}_{1},u_{2,b-1})\big| \tilde{\svec}_{1}, \mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big)\\
& = \mspace{-20mu} \sum_{ \substack{\big(\tilde{\mathbf{v}}_1(\tilde{u}_{1}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{x}}_3(\tilde{u}_{1},u_{2,b-1})\big) \in\\
A_{\epsilon}^{*(n)}\big(V_1,X_1,X_3\big|\tilde{\svec}_{1}, \mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)}} \mspace{-160mu}
p\big( \tilde{\mathbf{v}}_1(\tilde{u}_{1}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{x}}_3(\tilde{u}_{1},u_{2,b-1})\big| \tilde{\svec}_{1},\mathbf{v}_2(u_{2,b-1})\big)\\
& \le 2^{n\big(H(V_1,X_1,X_3|S_1,S_2,V_2,X_2,W,Y)+\epsilon_8\big)} 2^{-n\big(H(V_1,X_1,X_3|S_1,V_2) -\epsilon_8 \big)}.
\end{align*}
\noindent Thus, we have
\begin{align*}
\Pr & \big(\mathcal{E}_{\mathbf{s}_1,u_1}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big) \\
& \le \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-120mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b ) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1 \ne u_{1,b-1}}} \mspace{-100mu} \sum_{ \mspace{120mu} \substack{\tilde{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-170mu} 2^{-nR_1}
2^{n\big(H(V_1,X_1,X_3|S_1,S_2,V_2,X_2,W,Y)+\epsilon_8\big)} 2^{-n\big(H(V_1,X_1,X_3|S_1,V_2) -\epsilon_8 \big)}\\
& \le \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-120mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b},\mathbf{v}_2(u_{2,b-1}), \mathbf{x}_2(\mathbf{s}_{2,b},u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big)\times\\
& \qquad 2^{nR_1} 2^{n\big(H(S_1|S_{2},V_2,X_2,W,Y)+\epsilon_8\big) }2^{-nR_1}
2^{n\big(H(V_1,X_1,X_3|S_1,S_2,V_2,X_2,W,Y)+\epsilon_8\big)} 2^{-n\big(H(V_1,X_1,X_3|S_1,V_2) -\epsilon_8 \big)}\\
& = 2^{n\big(H(S_1,V_1,X_1,X_3|S_{2},V_2,X_2,W,Y)+2\epsilon_8\big)} 2^{-n\big(H(V_1,X_1,X_3|S_1,V_2) -\epsilon_8 \big)},
\end{align*}
\noindent which implies that in order to get an arbitrarily small probability of error as $n$ increases, it must hold that
\begin{align*}
H(S_1,V_1,X_1,X_3|S_{2},V_2,X_2,W,Y) - H(V_1,X_1,X_3|S_1,V_2) + 3\epsilon_8 < 0.
\end{align*}
Now consider
\begin{align*}
& H(S_1,V_1,X_1,X_3|S_2,V_2,X_2,W,Y) - H(V_1,X_1,X_3|S_1,V_2) \\
& \quad = H(S_1,V_1,X_1,X_3|S_2,V_2,X_2,W,Y) - H(S_1,V_1,X_1,X_3|V_2) + H(S_1|V_2) \\
& \quad \stackrel{(a)}{=} H(S_1) - I(S_1,V_1,X_1,X_3;S_2,X_2,W,Y|V_2) \\
& \quad = H(S_1) - I(S_1,V_1,X_1,X_3;S_2,X_2,W|V_2) - I(S_1,V_1,X_1,X_3;Y|S_2,V_2,X_2,W) \\
& \quad = H(S_1) - I(S_1;S_2,X_2,W|V_2) - I(V_1,X_1,X_3;S_2,X_2,W|S_1,V_2) \\
& \qquad - I(S_1,V_1,X_1,X_3;Y|S_2,V_2,X_2,W) \\
& \quad \stackrel{(b)}{=} H(S_1) - H(S_1|V_2) + H(S_1|S_2,V_2,X_2,W) - I(S_1,V_1,X_1,X_3;Y|S_2,V_2,X_2,W) \\
& \quad \stackrel{(c)}{=} H(S_1|S_1,W) - I(X_1,X_3;Y|S_2,V_2,X_2,W)
\end{align*}
\noindent where (a) follows form the independence $S_1$ and $V_2$; (b) follows from the Markov relationship $(S_2,X_2,W) \leftrightarrow (S_1,V_2) \leftrightarrow (V_1,X_1,X_3)$; and (c) follows from the Markov relationship $(V_2,X_2) \leftrightarrow (S_2,W) \leftrightarrow S_1$, and from the Markov relationship $(S_1,V_1) \leftrightarrow (S_2,V_2,X_1,X_2,X_3,W) \leftrightarrow Y$.
We conclude that as long as
\begin{equation}
H(S_1|S_2,W) < I(X_1,X_3;Y|S_2,V_2,X_2,W) - 3\epsilon_8,
\label{eq:ErrAnalysisCondSimult_s1u1}
\end{equation}
$\Pr\big(\mathcal{E}_{\mathbf{s}_1,u_1}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$ can be made arbitrarily small by taking $n$ large enough.
By similar arguments we can show that as long as
\begin{equation}
H(S_2|S_1,W) < I(X_2,X_3;Y|S_1,V_1,X_1,W) - 3\epsilon_8,
\label{eq:ErrAnalysisCondSimult_s2u2}
\end{equation}
then $\Pr\big(\mathcal{E}_{\mathbf{s}_2,u_2}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$ can be made arbitrarily small by taking $n$ large enough.
\subsubsection{Bounding $\Pr\big(\mathcal{E}_{\mathbf{s}_1,u_1,u_2}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$}
Begin by writing
\begin{align*}
\Pr & \big(\mathcal{E}_{\mathbf{s}_1,u_1, u_2}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big) \\
& = \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = U_{1,b}, \exists \tilde{u}_1\in\mathcal{U}_1, \tilde{u}_1\ne U_{1,b-1}, \exists \tilde{u}_2\in\mathcal{U}_2, \tilde{u}_2\ne U_{2,b-1}, \\
& \qquad \qquad \big(\tilde{\svec}_{1}, \mathbf{S}_{2,b}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\mathbf{S}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1}, \tilde{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big)\\
& = \sum_{u_{1,b}\in \mathcal{U}_1, u_{1,b-1}\in\mathcal{U}_1, u_{2,b-1}\in\mathcal{U}_2} \mspace{-40mu}
p_{U_1}(u_{1,b}) p_{U_1 U_2}(u_{1,b-1},u_{2,b-1})\times \nonumber\\
& \qquad \quad \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = u_{1,b}, \exists \tilde{u}_1\in\mathcal{U}_1, \tilde{u}_1\ne u_{1,b-1}, \exists \tilde{u}_2\in\mathcal{U}_2, \tilde{u}_2\ne u_{2,b-1}, \\
& \qquad \qquad \quad \big(\tilde{\svec}_{1}, \mathbf{S}_{2,b}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\mathbf{S}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},\tilde{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big)
\end{align*}
\noindent We now bound
\begin{align*}
& \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = u_{1,b}, \exists \tilde{u}_1\in\mathcal{U}_1, \tilde{u}_1\ne u_{1,b-1}, \exists \tilde{u}_2\in\mathcal{U}_2, \tilde{u}_2\ne u_{2,b-1}, \\
& \qquad \big(\tilde{\svec}_{1}, \mathbf{S}_{2,b}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\mathbf{S}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},\tilde{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big) \\
& \stackrel{(a)}{=} \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-50mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1 \ne u_{1,b-1}}} \sum_{\substack{\tilde{u}_2\in\mathcal{U}_2,\\ \tilde{u}_2 \ne u_{2,b-1}}} \sum_{ \substack{\tilde{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b \big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-55mu} \Pr\Big( f_1(\tilde{\svec}_1) = u_{1,b}, \big(\tilde{\svec}_{1}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}),\tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \\
& \mspace{220mu} \tilde{\mathbf{X}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},\tilde{u}_{2})\big) \in A_{\epsilon}^{*(n)}\big(S_1,V_1,V_2,X_1,X_2,X_3\big|\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big) \Big) \\
& = \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-50mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1 \ne u_{1,b-1}}} \sum_{\substack{\tilde{u}_2\in\mathcal{U}_2,\\ \tilde{u}_2 \ne u_{2,b-1}}} \sum_{\substack{\tilde{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b \big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-55mu} 2^{-nR_1} \Pr\Big( \big(\tilde{\svec}_{1}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \\
& \mspace{220mu} \tilde{\mathbf{X}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},\tilde{u}_{2})\big) \in A_{\epsilon}^{*(n)}\big(S_1,V_1,V_2,X_1,X_2,X_3\big|\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big) \Big)
\end{align*}
\begin{align*}
& \stackrel{(b)}{=} \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-50mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1 \ne u_{1,b-1}}} \sum_{\substack{\tilde{u}_2\in\mathcal{U}_2,\\ \tilde{u}_2 \ne u_{2,b-1}}} \sum_{\substack{\tilde{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b \big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-55mu} 2^{-nR_1} \Pr\Big( \big( \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \\
& \mspace{220mu} \tilde{\mathbf{X}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},\tilde{u}_{2})\big) \in A_{\epsilon}^{*(n)}\big(V_1,V_2,X_1,X_2,X_3\big|\tilde{\svec}_{1},\mathbf{s}_{2,b}, \mathbf{w}_b, \mathbf{y}_b \big) \Big)
\end{align*}
\noindent where (a) follows from the conditioning on $\mathcal{E}_{b,1}^c$ which implies that the sequences at block $b$ are jointly typical, and as if $ \tilde{\svec}_1\notin A_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b \big)$, then
\[
\big(\tilde{\svec}_{1}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}),\tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},\tilde{u}_{2})\big) \notin A_{\epsilon}^{*(n)}\big(S_1,V_1,V_2,X_1,X_2,X_3\big|\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big),
\]
see also argument leading to \eqref{eq:Consistency_step}. Step (b) follows as when $\tilde{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b \big)$,
then joint typicality is achieved when
\[
\big( \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}),\tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},\tilde{u}_{2}) \big) \in A_{\epsilon}^{*(n)}\big(V_1,V_2,X_1,X_2,X_3\big|\tilde{\svec}_{1}, \mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big).
\]
Next we bound
\begin{align*}
\Pr & \Big( \big( \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},\tilde{u}_{2}) \big) \in A_{\epsilon}^{*(n)}\big(V_1,V_2,X_1,X_2,X_3\big|\tilde{\svec}_{1}, \mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big)\Big)\\
& = \mspace{-150mu} \sum_{\mspace{170mu} \substack{\big(\tilde{\mathbf{v}}_1(\tilde{u}_{1}), \tilde{\mathbf{v}}_2(\tilde{u}_{2}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{x}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{x}}_3(\tilde{u}_{1},\tilde{u}_{2})\big) \in
A_{\epsilon}^{*(n)}\big(V_1,V_2,X_1,X_2,X_3\big|\tilde{\svec}_{1}, \mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big)}} \mspace{-320mu} p\big( \tilde{\mathbf{v}}_1(\tilde{u}_{1}), \tilde{\mathbf{v}}_2(\tilde{u}_{2}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{x}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{x}}_3(\tilde{u}_{1},\tilde{u}_{2})\big| \tilde{\svec}_{1}, \mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b \big)\\
& = \mspace{-150mu} \sum_{\mspace{170mu} \substack{\big(\tilde{\mathbf{v}}_1(\tilde{u}_{1}), \tilde{\mathbf{v}}_2(\tilde{u}_{2}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{x}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{x}}_3(\tilde{u}_{1},\tilde{u}_{2})\big) \in A_{\epsilon}^{*(n)}\big(V_1,V_2,X_1,X_2,X_3\big|\tilde{\svec}_{1}, \mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big)}} \mspace{-320mu}
p\big( \tilde{\mathbf{v}}_1(\tilde{u}_{1}), \tilde{\mathbf{v}}_2(\tilde{u}_{2}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{x}}_2(\mathbf{s}_{2,b},\tilde{u}_{2}), \tilde{\mathbf{x}}_3(\tilde{u}_{1},\tilde{u}_{2})\big| \tilde{\svec}_{1}, \mathbf{s}_{2,b} \big)\\
& \le 2^{n\big(H(V_1,V_2,X_1,X_2,X_3|S_1,S_2,W,Y)+\epsilon_9\big)} 2^{-n\big(H(V_1,V_2,X_1,X_2,X_3|S_1,S_2) -\epsilon_9 \big)}.
\end{align*}
\noindent Thus, we have
\begin{align*}
\Pr & \big(\mathcal{E}_{\mathbf{s}_1,u_1,u_2}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big) \\
& \le \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-50mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1 \ne u_{1,b-1}}} \sum_{\substack{\tilde{u}_2\in\mathcal{U}_2,\\ \tilde{u}_2 \ne u_{2,b-1}}} \mspace{-70mu} \sum_{ \mspace{90mu} \substack{\tilde{\svec}_1\inA_{\epsilon}^{*(n)}\big(S_1\big|\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b \big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}}} \mspace{-100mu} 2^{-nR_1} 2^{n\big(H(V_1,V_2,X_1,X_2,X_3|S_1,S_2,W,Y)+\epsilon_9\big)} 2^{-n\big(H(V_1,V_2,X_1,X_2,X_3|S_1,S_2) -\epsilon_9 \big)} \\
& \le \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-50mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b \big) \times\\
& \qquad 2^{nR_1} 2^{nR_2} 2^{n\big(H(S_1|S_2,W,Y)+\epsilon_9\big) }2^{-nR_1}
2^{n\big(H(V_1,V_2,X_1,X_2,X_3|S_1,S_2,W,Y)+\epsilon_9\big)} 2^{-n\big(H(V_1,V_2,X_1,X_2,X_3|S_1,S_2) -\epsilon_9 \big)}\\
& = 2^{nR_2} 2^{n\big(H(S_1,V_1,V_2,X_1,X_2,X_3|S_2,W,Y)+2\epsilon_9\big)} 2^{-n\big(H(V_1,V_2,X_1,X_2,X_3|S_1,S_2) -\epsilon_9 \big)},
\end{align*}
\noindent which implies that in order to get an arbitrarily small probability of error as $n$ increases, it must hold that
\begin{align*}
R_2 + H(S_1,V_1,V_2,X_1,X_2,X_3|S_2,W,Y) - H(V_1,V_2,X_1,X_2,X_3|S_1,S_2) + 3\epsilon_9 < 0.
\end{align*}
Now consider
\begin{align*}
& H(S_1,V_1,V_2,X_1,X_2,X_3|S_2,W,Y) - H(V_1,V_2,X_1,X_2,X_3|S_1,S_2) \\
& \quad = H(S_1,V_1,V_2,X_1,X_2,X_3|S_2,W,Y) - H(S_1,V_1,V_2,X_1,X_2,X_3|S_2) + H(S_1|S_2) \\
& \quad = H(S_1|S_2) - I(S_1,V_1,V_2,X_1,X_2,X_3;W,Y|S_2) \\
& \quad = H(S_1|S_2) - I(S_1,V_1,V_2,X_1,X_2,X_3;W|S_2) - I(S_1,V_1,V_2,X_1,X_2,X_3;Y|S_2,W) \\
& \quad \stackrel{(a)}{=} H(S_1|S_2) - I(S_1,V_1,V_2,X_1,X_2,X_3;W|S_2) - I(X_1,X_2,X_3;Y|S_2,W) \\
& \quad \stackrel{(b)}{=} H(S_1|S_2) - I(S_1;W|S_2) - I(X_1,X_2,X_3;Y|S_2,W) \\
& \quad = H(S_1|S_2,W) - I(X_1,X_2,X_3;Y|S_2,W),
\end{align*}
\noindent where (a) follows from the Markov chain $(S_1,V_1,V_2) \leftrightarrow (X_1,X_2,X_3) \leftrightarrow Y$ for given $(W,S_2)$; and (b) follows from the Markov chain $(V_1,V_2,X_1,X_2,X_3) \leftrightarrow (S_1,S_2) \leftrightarrow W$.
We conclude that as long as
\begin{equation}
R_2 + H(S_1|S_2,W) < I(X_1,X_2,X_3;Y|S_2,W) - 3\epsilon_9,
\label{eq:ErrAnalysisCondSimult_s1u1u2}
\end{equation}
$\Pr\big(\mathcal{E}_{\mathbf{s}_1,u_1,u_2}\big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\big)$ can be made arbitrarily small by taking $n$ large enough.
By similar arguments we can show that as long as
\begin{equation}
R_1 + H(S_2|S_1,W) < I(X_1,X_2,X_3;Y|S_1,W) - 3\epsilon_9,
\label{eq:ErrAnalysisCondSimult_s2u2u2}
\end{equation}
then $\Pr\big(\mathcal{E}_{\mathbf{s}_2,u_1,u_2}\big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\big)$ can be made arbitrarily small by taking $n$ large enough.
\subsubsection{Bounding $\Pr\big(\mathcal{E}_{\mathbf{s}_1, \mathbf{s}_2, u_1}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$}
Begin by writing
\begin{align*}
\Pr & \big(\mathcal{E}_{\mathbf{s}_1,\mathbf{s}_2,u_1}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big) \\
& = \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = U_{1,b}, \exists \tilde{\svec}_2 \in \mathcal{S}_2^n, \tilde{\svec}_2 \ne \mathbf{S}_{2,b}, f_2(\tilde{\svec}_2) = U_{2,b}, \exists \tilde{u}_1 \in \mathcal{U}_1, \tilde{u}_1\ne U_{1,b-1}, \\
& \qquad \big(\tilde{\svec}_{1}, \tilde{\svec}_{2},\tilde{\mathbf{V}}_1(\tilde{u}_{1}), \mathbf{V}_2(U_{2,b-1}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},U_{2,b-1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1}, U_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big)\\
& = \sum_{u_{1,b}\in \mathcal{U}_1, u_{2,b}\in \mathcal{U}_2, u_{1,b-1}\in\mathcal{U}_1, u_{2,b-1}\in\mathcal{U}_2} \mspace{-40mu}
p_{U_1 U_2}(u_{1,b},u_{2,b}) p_{U_1 U_2}(u_{1,b-1},u_{2,b-1})\times \nonumber\\
& \qquad \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = u_{1,b}, \exists \tilde{\svec}_2 \in \mathcal{S}_2^n, \tilde{\svec}_2 \ne \mathbf{S}_{2,b}, f_2(\tilde{\svec}_2) = u_{2,b}, \exists \tilde{u}_1 \in \mathcal{U}_1, \tilde{u}_1\ne u_{1,b-1}, \\
& \qquad \big(\tilde{\svec}_{1}, \tilde{\svec}_{2},\tilde{\mathbf{V}}_1(\tilde{u}_{1}), \mathbf{V}_2(U_{2,b-1}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},U_{2,b-1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1}, U_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big)
\end{align*}
\noindent We now bound
\begin{align*}
& \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = u_{1,b}, \exists \tilde{\svec}_2 \in \mathcal{S}_2^n, \tilde{\svec}_2 \ne \mathbf{S}_{2,b}, f_2(\tilde{\svec}_2) = u_{2,b}, \exists \tilde{u}_1 \in \mathcal{U}_1, \tilde{u}_1\ne u_{1,b-1}, \\
& \qquad \big(\tilde{\svec}_{1}, \tilde{\svec}_{2},\tilde{\mathbf{V}}_1(\tilde{u}_{1}), \mathbf{V}_2(U_{2,b-1}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},U_{2,b-1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1}, U_{2,b-1}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big) \\
& \stackrel{(a)}{=} \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-60mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1 \ne u_{1,b-1}}} \sum_{\substack{\big( \tilde{\svec}_1, \tilde{\svec}_2 \big)\inA_{\epsilon}^{*(n)}\big(S_1,S_2\big| \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}, \tilde{\svec}_2\ne\mathbf{s}_{2,b}}} \mspace{-110mu} \Pr\Big( f_1(\tilde{\svec}_1) = u_{1,b}, f_2(\tilde{\svec}_2) = u_{2,b}, \\
& \mspace{170mu} \big(\tilde{\svec}_{1}, \tilde{\svec}_{2}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},u_{2,b-1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},u_{2,b-1})\big) \in \\
& \mspace{185mu} A_{\epsilon}^{*(n)}\big(S_1,S_2,V_1,X_1,X_2,X_3\big| \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big) \Big)\\
& = \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-60mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1 \ne u_{1,b-1}}} \sum_{\substack{\big( \tilde{\svec}_1, \tilde{\svec}_2 \big)\inA_{\epsilon}^{*(n)}\big(S_1,S_2\big| \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}, \tilde{\svec}_2\ne\mathbf{s}_{2,b}}} \mspace{-115mu} 2^{-nR_1} 2^{-nR_2} \times \\
& \mspace{120mu} \Pr\Big( \big(\tilde{\svec}_{1}, \tilde{\svec}_{2}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},u_{2,b-1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},u_{2,b-1})\big) \in \\
& \mspace{170mu} A_{\epsilon}^{*(n)}\big(S_1,S_2,V_1,X_1,X_2,X_3\big| \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big) \Big)\\
& \stackrel{(b)}{=} \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-60mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big) \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1 \ne u_{1,b-1}}} \sum_{\substack{\big( \tilde{\svec}_1, \tilde{\svec}_2 \big)\inA_{\epsilon}^{*(n)}\big(S_1,S_2\big| \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}, \tilde{\svec}_2\ne\mathbf{s}_{2,b}}} \mspace{-115mu} 2^{-nR_1} 2^{-nR_2} \times\\
& \mspace{120mu} \Pr\Big( \big( \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},u_{2,b-1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},u_{2,b-1})\big) \in \\
& \mspace{170mu} A_{\epsilon}^{*(n)}\big(V_1,X_1,X_2,X_3\big| \tilde{\svec}_{1}, \tilde{\svec}_{2}, \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big) \Big)
\end{align*}
\noindent where (a) follows from the conditioning on $\mathcal{E}_{b,1}^c$ which implies that the sequences at block $b$ are jointly typical, and as if $ (\tilde{\svec}_1, \tilde{\svec}_2)\notin A_{\epsilon}^{*(n)}\big(S_1,S_2\big|\mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big)$, then
\[
\big(\tilde{\svec}_{1}, \tilde{\svec}_{2},\tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},u_{2,b-1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},u_{2,b-1})\big) \notin A_{\epsilon}^{*(n)}\big(S_1, S_2,V_1, X_1,X_2,X_3\big|\mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big),
\]
see also argument leading to \eqref{eq:Consistency_step}. Step (b) follows as when $(\tilde{\svec}_1, \tilde{\svec}_2) \in A_{\epsilon}^{*(n)}\big(S_1,S_2\big|\mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big)$,
then joint typicality is achieved when
\[
\big( \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},u_{2,b-1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},u_{2,b-1})\big) \in A_{\epsilon}^{*(n)}\big(V_1,X_1,X_2,X_3\big| \tilde{\svec}_{1}, \tilde{\svec}_{2}, \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big).
\]
Next we bound
\begin{align*}
\Pr & \Big( \big( \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},u_{2,b-1}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},u_{2,b-1})\big) \in A_{\epsilon}^{*(n)}\big(V_1,X_1,X_2,X_3\big| \tilde{\svec}_{1}, \tilde{\svec}_{2}, \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big) \Big)\\
& = \sum_{ \substack{\big( \tilde{\mathbf{v}}_1(\tilde{u}_{1}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{x}}_2(\tilde{\svec}_{2},u_{2,b-1}), \tilde{\mathbf{x}}_3(\tilde{u}_{1},u_{2,b-1}) \big) \in\\
A_{\epsilon}^{*(n)}\big(V_1,X_1,X_2,X_3\big| \tilde{\svec}_{1}, \tilde{\svec}_{2}, \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)}} \mspace{-120mu}
p\big( \tilde{\mathbf{v}}_1(\tilde{u}_{1}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{x}}_2(\tilde{\svec}_{2},u_{2,b-1}), \tilde{\mathbf{x}}_3(\tilde{u}_{1},u_{2,b-1})\big| \tilde{\svec}_{1}, \tilde{\svec}_{2}, \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big)\\
& = \sum_{ \substack{\big( \tilde{\mathbf{v}}_1(\tilde{u}_{1}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{x}}_2(\tilde{\svec}_{2},u_{2,b-1}), \tilde{\mathbf{x}}_3(\tilde{u}_{1},u_{2,b-1}) \big) \in\\
A_{\epsilon}^{*(n)}\big(V_1,X_1,X_2,X_3\big| \tilde{\svec}_{1}, \tilde{\svec}_{2}, \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)}} \mspace{-120mu}
p\big( \tilde{\mathbf{v}}_1(\tilde{u}_{1}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{x}}_2(\tilde{\svec}_{2},u_{2,b-1}), \tilde{\mathbf{x}}_3(\tilde{u}_{1},u_{2,b-1})\big| \tilde{\svec}_{1}, \tilde{\svec}_{2}, \mathbf{v}_2(u_{2,b-1}) \big)\\
& \le 2^{n\big(H(V_1,X_1,X_2,X_3|S_1,S_2,V_2,W,Y)+\epsilon_{10}\big)} 2^{-n\big(H(V_1,X_1,X_2,X_3|S_1,S_2,V_2) -\epsilon_{10} \big)}.
\end{align*}
\noindent Thus, we have
\begin{align*}
\Pr & \big(\mathcal{E}_{\mathbf{s}_1,\mathbf{s}_2,u_1}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big) \\
& \le \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-60mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{30mu} \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1 \ne u_{1,b-1}}} \mspace{-120mu} \sum_{ \mspace{140mu} \substack{\big( \tilde{\svec}_1, \tilde{\svec}_2 \big)\inA_{\epsilon}^{*(n)}\big(S_1,S_2\big| \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}, \tilde{\svec}_2\ne\mathbf{s}_{2,b}}} \mspace{-160mu} 2^{-nR_1} 2^{-nR_2} 2^{n\big(H(V_1,X_1,X_2,X_3|S_1,S_2,V_2,W,Y)+\epsilon_{10}\big)} 2^{-n\big(H(V_1,X_1,X_2,X_3|S_1,S_2,V_2) -\epsilon_{10} \big)} \\
& \le \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-60mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{v}_2(u_{2,b-1}), \mathbf{w}_b,\mathbf{y}_b \big) \cdot 2^{nR_1} 2^{n\big(H(S_1,S_2|V_2,W,Y)+\epsilon_{10}\big) } \times\\
& \qquad 2^{-nR_1} 2^{-nR_2}
2^{n\big(H(V_1,X_1,X_2,X_3|S_1,S_2,V_2,W,Y)+\epsilon_{10}\big)} 2^{-n\big(H(V_1,X_1,X_2,X_3|S_1,S_2,V_2) -\epsilon_{10} \big)} \\
& = 2^{-nR_2} 2^{n\big(H(S_1,S_2,V_1,X_1,X_2,X_3|V_2,W,Y)+2\epsilon_{10}\big)} 2^{-n\big(H(V_1,X_1,X_2,X_3|S_1,S_2,V_2) -\epsilon_{10} \big)},
\end{align*}
\noindent which implies that in order to get an arbitrarily small probability of error as $n$ increases, it must hold that
\begin{align*}
H(S_1,S_2,V_1,X_1,X_2,X_3|V_2,W,Y) - H(V_1,X_1,X_2,X_3|S_1,S_2,V_2) - R_2 + 3\epsilon_{10} < 0.
\end{align*}
Now consider
\begin{align*}
& H(S_1,S_2,V_1,X_1,X_2,X_3|V_2,W,Y) - H(V_1,X_1,X_2,X_3|S_1,S_2,V_2) \\
& \quad = H(S_1,S_2,V_1,X_1,X_2,X_3|V_2,W,Y) - H(S_1,S_2,V_1,X_1,X_2,X_3|V_2) + H(S_1,S_2|V_2) \\
& \quad \stackrel{(a)}{=} H(S_1,S_2) - I(S_1,S_2,V_1,X_1,X_2,X_3;W,Y|V_2) \\
& \quad = H(S_1,S_2) - I(S_1,S_2,V_1,X_1,X_2,X_3;W|V_2) - I(S_1,S_2,V_1,X_1,X_2,X_3;Y|V_2,W) \\
& \quad \stackrel{(b)}{=} H(S_1,S_2) - I(S_1,S_2,V_1,X_1,X_2,X_3;W|V_2) - I(X_1,X_2,X_3;Y|V_2,W) \\
& \quad \stackrel{(c)}{=} H(S_1,S_2) - I(S_1,S_2;W) - I(X_1,X_2,X_3;Y|V_2,W) \\
& \quad = H(S_1,S_2|W) - I(X_1,X_2,X_3;Y|V_2,W),
\end{align*}
\noindent where (a) follows form the independence $(S_1,S_2)$ and $V_2$; (b) follows from the Markov chain $(S_1,S_2,V_1) \leftrightarrow (X_1,X_2,X_3) \leftrightarrow Y$ for given $(V_2,W)$; and (c) follows from the Markov chain $(V_1,V_2,X_1,X_2,X_3) \leftrightarrow (S_1,S_2) \leftrightarrow W$.
We conclude that as long as
\begin{equation}
H(S_1,S_2|W) < I(X_1,X_2,X_3;Y|V_2,W) + R_2 - 3\epsilon_{10},
\label{eq:ErrAnalysisCondSimult_s1s2u1}
\end{equation}
$\Pr\big(\mathcal{E}_{\mathbf{s}_1,\mathbf{s}_2,u_1}\big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\big)$ can be made arbitrarily small by taking $n$ large enough.
By similar arguments we can show that as long as
\begin{equation}
H(S_1,S_2|W) < I(X_1,X_2,X_3;Y|V_1,W) + R_1 - 3\epsilon_{10},
\label{eq:ErrAnalysisCondSimult_s1s2u2}
\end{equation}
then $\Pr\big(\mathcal{E}_{\mathbf{s}_1,\mathbf{s}_2,u_2}\big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\big)$ can be made arbitrarily small by taking $n$ large enough.
\subsubsection{Bounding $\Pr\big(\mathcal{E}_{\mathbf{s}_1, \mathbf{s}_2, u_1, u_2}\big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\big)$}
Begin by writing
\begin{align*}
\Pr & \big(\mathcal{E}_{\mathbf{s}_1,\mathbf{s}_2,u_1,u_2}\big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\big) \\
& = \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = U_{1,b}, \exists \tilde{\svec}_2 \in \mathcal{S}_2^n, \tilde{\svec}_2 \ne \mathbf{S}_{2,b}, f_2(\tilde{\svec}_2) = U_{2,b}, \\
& \qquad \qquad \exists \tilde{u}_1\in\mathcal{U}_1, \tilde{u}_1\ne U_{1,b-1}, \exists \tilde{u}_2\in\mathcal{U}_2, \tilde{u}_2\ne U_{2,b-1}, \big(\tilde{\svec}_{1}, \tilde{\svec}_{2}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \mathbf{V}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \\
& \qquad \qquad \qquad \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1}, \tilde{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big)\\
& = \sum_{u_{1,b}\in \mathcal{U}_1, u_{2,b}\in \mathcal{U}_2, u_{1,b-1}\in\mathcal{U}_1, u_{2,b-1}\in\mathcal{U}_2} \mspace{-40mu}
p_{U_1 U_2}(u_{1,b},u_{2,b}) p_{U_1 U_2}(u_{1,b-1},u_{2,b-1})\times \nonumber\\
& \qquad \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = u_{1,b}, \exists \tilde{\svec}_2 \in \mathcal{S}_2^n, \tilde{\svec}_2 \ne \mathbf{S}_{2,b}, f_2(\tilde{\svec}_2) = u_{2,b}, \\
& \qquad \qquad \exists \tilde{u}_1\in\mathcal{U}_1, \tilde{u}_1\ne u_{1,b-1}, \exists \tilde{u}_2\in\mathcal{U}_2, \tilde{u}_2\ne u_{2,b-1}, \big(\tilde{\svec}_{1}, \tilde{\svec}_{2}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \mathbf{V}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \\
& \qquad \qquad \qquad \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1}, \tilde{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big)
\end{align*}
\noindent We now bound
\begin{align*}
& \Pr\Big( \exists \tilde{\svec}_1 \in \mathcal{S}_1^n, \tilde{\svec}_1 \ne \mathbf{S}_{1,b}, f_1(\tilde{\svec}_1) = u_{1,b}, \exists \tilde{\svec}_2 \in \mathcal{S}_2^n, \tilde{\svec}_2 \ne \mathbf{S}_{2,b}, f_2(\tilde{\svec}_2) = u_{2,b}, \\
& \qquad \qquad \exists \tilde{u}_1\in\mathcal{U}_1, \tilde{u}_1\ne u_{1,b-1}, \exists \tilde{u}_2\in\mathcal{U}_2, \tilde{u}_2\ne u_{2,b-1}, \big(\tilde{\svec}_{1}, \tilde{\svec}_{2}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \mathbf{V}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \\
& \qquad \qquad \qquad \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1}, \tilde{u}_{2}), \mathbf{W}_b, \mathbf{Y}_b\big) \in A_{\epsilon}^{*(n)}\Big|\mathcal{E}_{1,b}^c \cap \mathcal{F}_{b+1}^c\Big) \\
& \stackrel{(a)}{=} \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-60mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1 \ne u_{1,b-1}}} \sum_{\substack{\tilde{u}_2\in\mathcal{U}_2,\\ \tilde{u}_2 \ne u_{2,b-1}}} \sum_{\substack{\big( \tilde{\svec}_1, \tilde{\svec}_2 \big)\inA_{\epsilon}^{*(n)}\big(S_1,S_2\big| \mathbf{w}_b,\mathbf{y}_b \big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}, \tilde{\svec}_2\ne\mathbf{s}_{2,b}}} \mspace{-70mu} \Pr\Big( f_1(\tilde{\svec}_1) = u_{1,b}, f_2(\tilde{\svec}_2) = u_{2,b}, \\
& \mspace{170mu} \big(\tilde{\svec}_{1}, \tilde{\svec}_{2}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},\tilde{u}_{2})\big) \in \\
& \mspace{200mu} A_{\epsilon}^{*(n)}\big(S_1,S_2,V_1,V_2,X_1,X_2,X_3\big| \mathbf{w}_b,\mathbf{y}_b\big) \Big)
\end{align*}
\begin{align*}
& = \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-60mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1 \ne u_{1,b-1}}} \sum_{\substack{\tilde{u}_2\in\mathcal{U}_2,\\ \tilde{u}_2 \ne u_{2,b-1}}} \sum_{\substack{\big( \tilde{\svec}_1, \tilde{\svec}_2 \big)\inA_{\epsilon}^{*(n)}\big(S_1,S_2\big| \mathbf{w}_b,\mathbf{y}_b \big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}, \tilde{\svec}_2\ne\mathbf{s}_{2,b}}} \mspace{-70mu} 2^{-nR_1} 2^{-nR_2} \times \\
& \mspace{170mu} \Pr\Big( \big(\tilde{\svec}_{1}, \tilde{\svec}_{2}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},\tilde{u}_{2}) \big) \in \\
& \mspace{210mu} A_{\epsilon}^{*(n)}\big(S_1,S_2,V_1,V_2,X_1,X_2,X_3\big| \mathbf{w}_b,\mathbf{y}_b\big) \Big)\\
& \stackrel{(b)}{=} \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-60mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1 \ne u_{1,b-1}}} \sum_{\substack{\tilde{u}_2\in\mathcal{U}_2,\\ \tilde{u}_2 \ne u_{2,b-1}}} \sum_{\substack{\big( \tilde{\svec}_1, \tilde{\svec}_2 \big)\inA_{\epsilon}^{*(n)}\big(S_1,S_2\big| \mathbf{w}_b,\mathbf{y}_b \big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}, \tilde{\svec}_2\ne\mathbf{s}_{2,b}}} \mspace{-70mu} 2^{-nR_1} 2^{-nR_2} \times \\
& \mspace{170mu} \Pr\Big( \big( \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},\tilde{u}_{2}) \big) \in \\
& \mspace{210mu} A_{\epsilon}^{*(n)}\big(V_1,V_2,X_1,X_2,X_3\big| \tilde{\svec}_1,\tilde{\svec}_2, \mathbf{w}_b,\mathbf{y}_b\big)\Big)
\end{align*}
\noindent where (a) follows from the conditioning on $\mathcal{E}_{b,1}^c$ which implies that the sequences at block $b$ are jointly typical, and as if $ (\tilde{\svec}_1, \tilde{\svec}_2)\notin A_{\epsilon}^{*(n)}\big(S_1,S_2\big|\mathbf{w}_b,\mathbf{y}_b \big)$, then
\[
\big(\tilde{\svec}_{1}, \tilde{\svec}_{2}, \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},\tilde{u}_{2})\big) \notin A_{\epsilon}^{*(n)}\big(S_1, S_2, V_1, V_2, X_1, X_2,X_3\big| \mathbf{w}_b,\mathbf{y}_b\big),
\]
see also argument leading to \eqref{eq:Consistency_step}. Step (b) follows as when $(\tilde{\svec}_1, \tilde{\svec}_2)\in A_{\epsilon}^{*(n)}\big(S_1,S_2\big| \mathbf{w}_b,\mathbf{y}_b \big)$,
then joint typicality is achieved when
\[
\big( \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},\tilde{u}_{2}) \big) \in A_{\epsilon}^{*(n)}\big(V_1,V_2,X_1,X_2,X_3\big| \tilde{\svec}_1,\tilde{\svec}_2, \mathbf{w}_b,\mathbf{y}_b\big).
\]
Next we bound
\begin{align*}
\Pr & \Big( \big( \tilde{\mathbf{V}}_1(\tilde{u}_{1}), \tilde{\mathbf{V}}_2(\tilde{u}_{2}), \tilde{\mathbf{X}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{X}}_2(\tilde{\svec}_{2},\tilde{u}_{2}), \tilde{\mathbf{X}}_3(\tilde{u}_{1},\tilde{u}_{2}) \big) \in A_{\epsilon}^{*(n)}\big(V_1,V_2,X_1,X_2,X_3\big| \tilde{\svec}_1,\tilde{\svec}_2, \mathbf{w}_b,\mathbf{y}_b\big) \Big)\\
& = \sum_{ \substack{\big(\tilde{\mathbf{v}}_1(\tilde{u}_{1}), \tilde{\mathbf{v}}_2(\tilde{u}_{2}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{x}}_2(\tilde{\svec}_{2},\tilde{u}_{2}), \tilde{\mathbf{x}}_3(\tilde{u}_{1},\tilde{u}_{2}) \big) \in\\
A_{\epsilon}^{*(n)}\big(V_1,V_2,X_1,X_2,X_3\big| \tilde{\svec}_{1}, \tilde{\svec}_{2}, \mathbf{w}_b,\mathbf{y}_b\big)}} \mspace{-120mu}
p\big( \tilde{\mathbf{v}}_1(\tilde{u}_{1}), \tilde{\mathbf{v}}_2(\tilde{u}_{2}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{x}}_2(\tilde{\svec}_{2},\tilde{u}_{2}), \tilde{\mathbf{x}}_3(\tilde{u}_{1},\tilde{u}_{2})\big| \tilde{\svec}_{1}, \tilde{\svec}_{2}, \mathbf{w}_b,\mathbf{y}_b \big)\\
& = \sum_{ \substack{\big(\tilde{\mathbf{v}}_1(\tilde{u}_{1}), \tilde{\mathbf{v}}_2(\tilde{u}_{2}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{x}}_2(\tilde{\svec}_{2},\tilde{u}_{2}), \tilde{\mathbf{x}}_3(\tilde{u}_{1},\tilde{u}_{2}) \big) \in\\
A_{\epsilon}^{*(n)}\big(V_1,V_2,X_1,X_2,X_3\big| \tilde{\svec}_{1}, \tilde{\svec}_{2}, \mathbf{w}_b,\mathbf{y}_b\big)}} \mspace{-120mu}
p\big( \tilde{\mathbf{v}}_1(\tilde{u}_{1}), \tilde{\mathbf{v}}_2(\tilde{u}_{2}), \tilde{\mathbf{x}}_1(\tilde{\svec}_{1},\tilde{u}_{1}), \tilde{\mathbf{x}}_2(\tilde{\svec}_{2},\tilde{u}_{2}), \tilde{\mathbf{x}}_3(\tilde{u}_{1},\tilde{u}_{2})\big| \tilde{\svec}_{1}, \tilde{\svec}_{2} \big)\\
& \le 2^{n\big(H(V_1,V_2,X_1,X_2,X_3|S_1,S_2,W,Y)+\epsilon_{11}\big)} 2^{-n\big(H(V_1,V_2,X_1,X_2,X_3|S_1,S_2) -\epsilon_{11} \big)}.
\end{align*}
\noindent Thus, we have
\begin{align*}
\Pr & \big(\mathcal{E}_{\mathbf{s}_1,\mathbf{s}_2,u_1,u_2}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big) \\
& \le \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-60mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b \big) \times \\
& \mspace{50mu} \sum_{\substack{\tilde{u}_1\in\mathcal{U}_1,\\ \tilde{u}_1 \ne u_{1,b-1}}} \sum_{\substack{\tilde{u}_2\in\mathcal{U}_2,\\ \tilde{u}_2 \ne u_{2,b-1}}} \sum_{\substack{\big( \tilde{\svec}_1, \tilde{\svec}_2 \big)\inA_{\epsilon}^{*(n)}\big(S_1,S_2\big| \mathbf{w}_b,\mathbf{y}_b \big), \\ \tilde{\svec}_1\ne\mathbf{s}_{1,b}, \tilde{\svec}_2\ne\mathbf{s}_{2,b}}} \mspace{-70mu} 2^{-nR_1} 2^{-nR_2} \times \\
& \mspace{100mu} 2^{n\big(H(V_1,V_2,X_1,X_2,X_3|S_1,S_2,W,Y)+\epsilon_{11}\big)} 2^{-n\big(H(V_1,V_2,X_1,X_2,X_3|S_1,S_2) -\epsilon_{11} \big)} \\
& \le \sum_{\substack{\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b\big)\in A_{\epsilon}^{*(n)}}} \mspace{-60mu} p\big(\mathbf{s}_{1,b},\mathbf{s}_{2,b}, \mathbf{w}_b,\mathbf{y}_b \big) \cdot 2^{nR_1} 2^{nR_2} 2^{n\big(H(S_1,S_2|W,Y)+\epsilon_{11}\big) } \times\\
& \mspace{100mu} 2^{-nR_1} 2^{-nR_2} 2^{n\big(H(V_1,V_2,X_1,X_2,X_3|S_1,S_2,W,Y)+\epsilon_{11}\big)} 2^{-n\big(H(V_1,V_2,X_1,X_2,X_3|S_1,S_2) -\epsilon_{11} \big)} \\
& = 2^{n\big(H(S_1,S_2,V_1,V_2,X_1,X_2,X_3|W,Y)+2\epsilon_{11}\big)} 2^{-n\big(H(V_1,V_2,X_1,X_2,X_3|S_1,S_2) -\epsilon_{11} \big)},
\end{align*}
\noindent which implies that in order to get an arbitrarily small probability of error as $n$ increases, it must hold that
\begin{align*}
H(S_1,S_2,V_1,V_2,X_1,X_2,X_3|W,Y) - H(V_1,V_2,X_1,X_2,X_3|S_1,S_2) + 3\epsilon_{11} < 0.
\end{align*}
Now consider
\begin{align*}
& H(S_1,S_2,V_1,V_2,X_1,X_2,X_3|W,Y) - H(V_1,V_2,X_1,X_2,X_3|S_1,S_2) \\
& \quad = H(S_1,S_2,V_1,V_2,X_1,X_2,X_3|W,Y) - H(S_1,S_2,V_1,V_2,X_1,X_2,X_3) + H(S_1,S_2) \\
& \quad = H(S_1,S_2) - I(S_1,S_2,V_1,V_2,X_1,X_2,X_3;W,Y) \\
& \quad = H(S_1,S_2) - I(S_1,S_2,V_1,V_2,X_1,X_2,X_3;W) - I(S_1,S_2,V_1,V_2,X_1,X_2,X_3;Y|W) \\
& \quad \stackrel{(a)}{=} H(S_1,S_2) - I(S_1,S_2,V_1,V_2,X_1,X_2,X_3;W) - I(X_1,X_2,X_3;Y|W) \\
& \quad \stackrel{(b)}{=} H(S_1,S_2) - I(S_1,S_2;W) - I(X_1,X_2,X_3;Y|W) \\
& \quad = H(S_1,S_2|W) - I(X_1,X_2,X_3;Y|W),
\end{align*}
\noindent where (a) follows from the Markov chain $(S_1,S_2,V_1,V_2) \leftrightarrow (X_1,X_2,X_3) \leftrightarrow Y$ for given $W$; and (b) follows from the Markov chain $(V_1,V_2,X_1,X_2,X_3) \leftrightarrow (S_1,S_2) \leftrightarrow W$.
We conclude that as long as
\begin{equation}
H(S_1,S_2|W) < I(X_1,X_2,X_3;Y|W) - 3\epsilon_{11},
\label{eq:ErrAnalysisCondSimult_s1s2u1u2}
\end{equation}
$\Pr\big(\mathcal{E}_{\mathbf{s}_1,\mathbf{s}_2,u_1,u_2}\big|\mathcal{E}_{1,b}^c\cap \mathcal{F}_{b+1}^c\big)$ can be made arbitrarily small by taking $n$ large enough.
\subsection{Simplifying the Decoding Constraints}
Combining the constraints for reliable decoding detailed in \eqref{eq:ErrAnalysisCondSimult_u1}--\eqref{eq:ErrAnalysisCondSimult_s1s2u1u2} results in the following set of constraints
\begin{subequations} \label{eq:DestNewSimultBounds_all}
\begin{align}
R_1 & < I(X_1,X_3;Y|S_1,V_2,X_2) \label{eq:DestNewSimultBounds_a} \\
H(S_1|S_2,W) - R_1 & < I(X_1;Y|S_2,V_1,X_2,X_3,W) \label{eq:DestNewSimultBounds_b} \\
H(S_1|S_2,W) & < I(X_1, X_3;Y|S_2, V_2, X_2,W) \label{eq:DestNewSimultBounds_c} \\
R_2 & < I(X_2,X_3;Y|S_2,V_1,X_1) \label{eq:DestNewSimultBounds_d} \\
H(S_2|S_1,W) - R_2 & < I(X_2;Y|S_1,V_2,X_1,X_3,W) \label{eq:DestNewSimultBounds_e} \\
H(S_2|S_1,W) & < I(X_2, X_3;Y|S_1,V_1,X_1,W) \label{eq:DestNewSimultBounds_f} \\
R_1 + R_2 & < I(X_1, X_2, X_3;Y|S_1,S_2) \label{eq:DestNewSimultBounds_g} \\
H(S_2|S_1,W) + R_1 - R_2 & < I(X_1, X_2, X_3;Y|S_1, V_2,W) \label{eq:DestNewSimultBounds_h} \\
H(S_1|S_2,W) - R_1 + R_2 & < I(X_1, X_2, X_3;Y|S_2, V_1,W) \label{eq:DestNewSimultBounds_i} \\
H(S_1,S_2|W) - R_1 - R_2 & < I(X_1, X_2;Y|V_1,V_2,X_3,W) \label{eq:DestNewSimultBounds_j} \\
H(S_1|S_2,W) + R_2 & < I(X_1, X_2, X_3;Y|S_2,W) \label{eq:DestNewSimultBounds_k} \\
H(S_1,S_2|W) - R_2 & < I(X_1, X_2, X_3;Y|V_2,W) \label{eq:DestNewSimultBounds_l} \\
H(S_2|S_1,W) + R_1 & < I(X_1, X_2, X_3;Y|S_1,W) \label{eq:DestNewSimultBounds_m} \\
H(S_1,S_2|W) - R_1 & < I(X_1, X_2, X_3;Y|V_1,W) \label{eq:DestNewSimultBounds_n} \\
H(S_1,S_2|W) & < I(X_1, X_2, X_3;Y|W). \label{eq:DestNewSimultBounds_o}
\end{align}
\end{subequations}
\noindent By using Fourier-Motzkin algorithm to eliminate $R_1$ and $R_2$ from \eqref{eq:DestNewSimultBounds_all} we get \eqref{bnd:JointNewSimult}.
\ifthenelse{\boolean{SquizFlag}}{\end{comment}}{}
\section{Proof of Proposition \ref{prop:jointCond_New}} \label{annex:jointNewProof}
\vspace{-0.2cm}
\subsection{Codebook Construction and Encoding}
\vspace{-0.1cm}
The codebook construction and encoding are identical to \Thmref{thm:jointCond_NewSimult}, see Appendix \ref{annex:jointNewSimultProof}.
\vspace{-0.3cm}
\subsection{Decoding} \label{annex:jointNewProof_decoding}
\vspace{-0.1cm}
Decoding at the relay is identical to \Thmref{thm:jointCond_NewSimult}, see Appendix \ref{annex:jointNewSimultProof}. Decoding at the destination is done using successive backward decoding.
Let $\boldsymbol{\alpha} \in \mathcal{W}^n$ be an i.i.d sequence such that each letter $\alpha_k$ is selected independently according to $p_{W|S_1,S_2}(\alpha_k | a_{1,k}, a_{2,k}), k=1,2,\dots,n$.
The destination node waits until the end of channel block $B+1$. It first tries to decode $(u_{1,B},u_{2,B})$ using the received signal at channel block $B+1$, $\mathbf{y}_{B+1}$, and $\boldsymbol{\alpha}$.
Going backwards from the last channel block to the first, the destination has the estimates $(\hat{\msg}_{1,b},\hat{\msg}_{2,b})$ of $(u_{1,b},u_{2,b})$ when decoding at block $b$.
Now, for decoding at block $b$ the destination first recovers the bin indices $\hat{\msg}_{i,b-1}, i=1,2$, corresponding to $\mathbf{s}_{i,b-1}$, based on its received signal $\mathbf{y}_{b}$ and the side information $\mathbf{w}_{b}$. This is done by looking for a unique pair $(\hat{\msg}_{1}, \hat{\msg}_{2}) \in \mathcal{\msgBig}_1 \times \mathcal{\msgBig}_2$ such that:
\vspace{-0.2cm}
\begin{align}
& \big(\mathbf{v}_1(\hat{\msg}_{1}), \mathbf{v}_2(\hat{\msg}_{2}), \mathbf{x}_3(\hat{\msg}_{1}, \hat{\msg}_{2}), \mathbf{w}_{b}, \mathbf{y}_{b}\big) \in A_{\epsilon}^{*(n)}.
\label{eq:DestJntNewDecTypeIndices}
\end{align}
\vspace{-0.2cm}
\noindent Denote the decoded indices by $(\hat{\msg}_{1,b-1},\hat{\msg}_{2,b-1})$. Next, the destination decodes $\left(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}\right)$ by looking for a unique pair $\left(\hat{\svec}_{1}, \hat{\svec}_{2}\right)$ such that:
\vspace{-0.2cm}
\begin{align}
& \big( \hat{\svec}_{1}, \hat{\svec}_{2}, \mathbf{v}_1(\hat{\msg}_{1,b-1}), \mathbf{v}_2(\hat{\msg}_{2,b-1}), \mathbf{x}_1(\hat{\svec}_{1}, \hat{\msg}_{1,b-1}), \mathbf{x}_2(\hat{\svec}_{2}, \hat{\msg}_{2,b-1}), \mathbf{x}_3(\hat{\msg}_{1,b-1}, \hat{\msg}_{2,b-1}), \mathbf{w}_{b}, \mathbf{y}_{b} \big) \in A_{\epsilon}^{*(n)},
\label{eq:DestJntNewDecTypeList}
\end{align}
\vspace{-0.2cm}
\noindent and $f_1(\hat{\svec}_{1})=\hat{\msg}_{1,b}, f_2(\hat{\svec}_{2})=\hat{\msg}_{2,b}$. Denote the decoded sequences with $\left(\hat{\svec}_{1,b}, \hat{\svec}_{2,b}\right)$.
\vspace{-0.3cm}
\subsection{Error Probability Analysis}
\vspace{-0.1cm}
Following arguments similar to those in Appendix \ref{subsec:jointNewProofErrorAnalysis} it can be shown that decoding the source sequences at the relay can be done reliably as long as \eqref{bnd:JointNew_rly_S1}--\eqref{bnd:JointNew_rly_S1S2} hold, and decoding the source sequences at the destination can be done reliably as long as \eqref{bnd:JointNew_dst_S1}--\eqref{bnd:JointNew_dst_S1S2} hold.
\vspace{-0.3cm}
\section{Proof of Proposition \ref{prop:CorrSources}} \label{annex:ProofPropCorrSources}
\vspace{-0.2cm}
\subsection{\Thmref{thm:jointCond_NewSimult} Vs. \Thmref{thm:jointCond}}
\vspace{-0.2cm}
First we compare \eqref{bnd:JointNewSimult_dst_S1} and \eqref{bnd:Joint_dst_S1}. The first term on the RHS of \eqref{bnd:JointNewSimult_dst_S1} can be written as:
\vspace{-0.2cm}
\begin{align}
I(X_1,X_3;Y|S_2,V_2,X_2,W) & \stackrel{(a)}{=} I(S_1;Y|S_2,V_2,X_2,W) + I(X_1,X_3;Y|S_1,V_2,X_2) \nonumber \\
& \ge I(X_1,X_3;Y|S_1,V_2,X_2),
\end{align}
\vspace{-0.2cm}
\noindent where (a) follows from the Markov chains $S_1 \leftrightarrow\ (S_2,V_2,X_1,X_2,X_3,W) \leftrightarrow Y$, $(S_2,W) \leftrightarrow (S_1,V_2,X_2) \leftrightarrow Y$ and from the chain rule for mutual information. From the non-negativity of mutual information it follows that the second term on the RHS of \eqref{bnd:JointNewSimult_dst_S1}, $I(X_1,X_3;Y|S_1,V_2,X_2) + I(X_1;Y|S_2,V_1,X_2,X_3,W)$ is greater than or equal to $I(X_1,X_3;Y|S_1,V_2,X_2)$. As the LHSs of \eqref{bnd:JointNewSimult_dst_S1} and \eqref{bnd:Joint_dst_S1} are the same, we conclude that \eqref{bnd:JointNewSimult_dst_S1} is less restrictive than \eqref{bnd:Joint_dst_S1}. Using similar arguments it also follows that \eqref{bnd:JointNewSimult_dst_S2} is less restrictive than \eqref{bnd:Joint_dst_S2}.
Next, compare \eqref{bnd:JointNewSimult_dst_S1S2} and~\eqref{bnd:Joint_dst_S1S2}:
\vspace{-0.2cm}
\begin{align}
I(X_1,X_2,X_3;Y|W) & \ge I(X_1,X_2,X_3;Y|S_1,S_2), \label{eq:compare_Simult_Thm1_f}
\end{align}
\vspace{-0.2cm}
\noindent where \eqref{eq:compare_Simult_Thm1_f} follows from the Markov chain $(S_1,S_2) \leftrightarrow (X_1,X_2,X_3,W) \leftrightarrow Y$, and from the non-negativity of mutual information. As the LHSs of \eqref{bnd:JointNewSimult_dst_S1S2} and \eqref{bnd:Joint_dst_S1S2} are the same, we conclude that \eqref{bnd:JointNewSimult_dst_S1S2} is less restrictive than \eqref{bnd:Joint_dst_S1S2}.
In conclusion: \Thmref{thm:jointCond_NewSimult} is at least as good as \Thmref{thm:jointCond}.
\vspace{-0.4cm}
\subsection{\Thmref{thm:jointCond_NewSimult} Vs. Prop. \ref{prop:jointCond_New}}
\vspace{-0.2cm}
First consider \eqref{bnd:JointNewSimult_dst_S1} and \eqref{bnd:JointNew_dst_S1}. We begin with the first term on the RHS of \eqref{bnd:JointNewSimult_dst_S1}:
\vspace{-0.2cm}
\begin{align}
& I(X_1,X_3;Y|S_2,V_2,X_2,W) - I(X_1;Y|S_2,V_1,X_2,X_3,W) - I(V_1,X_3;Y|W, V_2) \nonumber \\
& \qquad \stackrel{(a)}{=} I(V_1,X_3;Y|S_2,V_2,X_2,W) - I(V_1,X_3;Y|W, V_2) \nonumber \\
& \qquad \stackrel{(b)}{=} I(V_1,X_3; S_2,X_2|V_2,W,Y) \ge 0,
\end{align}
\vspace{-0.2cm}
\noindent where (a) follows from the chain rule for mutual information; and (b) follows from the Markov relationship $(S_2,X_2) \leftrightarrow (V_2,W) \leftrightarrow (V_1,X_3)$. Next, consider the second term on the RHS of \eqref{bnd:JointNewSimult_dst_S1}:
\vspace{-0.2cm}
\begin{align}
& I(X_1,X_3;Y|S_1,V_2,X_2) + I(X_1;Y|S_2,V_1,X_2,X_3,W) - I(X_1;Y|S_2,V_1,X_2,X_3,W) - I(V_1,X_3;Y|V_2,W) \nonumber \\
& \qquad \stackrel{(a)}{=} I(X_1;Y|S_1,V_2,X_1,X_2,W) + I(V_1,X_3;Y|S_1,V_2,X_2,W) - I(V_1,X_3;Y|W,V_2) \nonumber \\
& \qquad \stackrel{(b)}{=} I(X_1;Y|S_1,V_2,X_1,X_2,W) + I(V_1,X_3;S_1,X_2|V_2,W,Y) \ge 0,
\end{align}
\vspace{-0.2cm}
\noindent where (a) follows from the chain rule for mutual information; and (b) follows from the Markov relationship $(S_1,X_2) \leftrightarrow (V_2,W) \leftrightarrow (V_1,X_3)$. As the LHS of \eqref{bnd:JointNewSimult_dst_S1} and \eqref{bnd:JointNew_dst_S1} is the same, we conclude that \eqref{bnd:JointNewSimult_dst_S1} is less restrictive than \eqref{bnd:JointNew_dst_S1}. Using similar arguments it follows that \eqref{bnd:JointNewSimult_dst_S2} is less restrictive than \eqref{bnd:JointNew_dst_S2}. For the expressions involving $H(S_1,S_2|W)$, note that the RHS of \eqref{bnd:JointNewSimult_dst_S1S2} equals to the RHS of \eqref{bnd:JointNew_dst_S1S2}.
Therefore, we conclude that \Thmref{thm:jointCond_NewSimult} is at least as good as Prop. \ref{prop:jointCond_New}.
%
\vspace{-0.45cm}
\section{Proof of Proposition \ref{prop:NotFeasible}} \label{annex:proofNotFeasible}
\vspace{-0.2cm}
It is enough to show that if at least one of the conditions in \eqref{bnd:Joint} holds with opposite strict inequality, then reliable transmission is not possible via the scheme of \Thmref{thm:jointCond}. The same statement holds for \eqref{bnd:JointFlip} and \Thmref{thm:jointCondFlip}. Furthermore, note that for the deterministic PSOMARC specified in Table \ref{tab:PrimitiveSOMARC}, and for the pair of correlated sources specified in Table \ref{tab:SourceDist}, reliable transmission to the destination requires assistance from the relay. To see this note that $H(S_1,S_2)= \log_2 3$, while $|\mathcal{Y}_S|=2$, which implies that the sources cannot be decoded at the destination without the help of the relay.
In Appendix \ref{annex:proofNotFeasible_thm1} we show that when the scheme of \Thmref{thm:jointCond} is used, if the sources can be decoded at the relay then they {\em cannot} be decoded at the destination, i.e., condition \eqref{bnd:Joint_dst_S1S2} holds with strict inequality. In Appendix \ref{annex:proofNotFeasible_thm2} we show that when the scheme of \Thmref{thm:jointCondFlip} is used, then the sources cannot be decoded at the relay, i.e., condition \eqref{bnd:JointFlip_rly_S1S2} holds with strict inequality.
\vspace{-0.45cm}
\subsection{Transmission Using the Scheme of Theorem \thmref{thm:jointCond}} \label{annex:proofNotFeasible_thm1}
\vspace{-0.15cm}
We begin with specializing the conditions of \Thmref{thm:jointCond} in \eqref{bnd:Joint_rly_S1}--\eqref{bnd:Joint_dst_S1S2} to the PSOMARC by letting $\mathcal{W}_3=\mathcal{W}=\phi$ and $I(X_3;Y_R)=C_3$. From the orthogonality of the relay-destination link it follows that the scheme of \Thmref{thm:jointCond} is optimized by letting $\mathcal{V}_1=\mathcal{V}_2=\phi$.
This fact and the resulting sufficient conditions are stated in the following proposition:
\begin{MyProposition}
The sufficient conditions of \Thmref{thm:jointCond} in \eqref{bnd:Joint_rly_S1}--\eqref{bnd:Joint_dst_S1S2}, specialized to the PSOMARC, are optimized by letting $\mathcal{V}_1=\mathcal{V}_2=\phi$. The resulting conditions are:
\vspace{-0.2cm}
\begin{subequations} \label{bnd:Thm7}
\begin{align}
H(S_1|S_2) &< \min \{I(X_1;Y_3|S_2, X_2), I(X_1;Y_S|S_1,X_2) + C_3 \} \label{bnd:Thm7_rly_S1} \\
H(S_2|S_1) &< \min \{I(X_2;Y_3|S_1, X_1), I(X_2;Y_S|S_2,X_1) + C_3 \} \label{bnd:Thm7_rly_S2} \\
H(S_1,S_2) &< \min \{I(X_1, X_2;Y_3), I(X_1, X_2;Y_S|S_1,S_2) + C_3 \}, \label{bnd:Thm7_rly_S1S2}
\end{align}
\end{subequations}
\vspace{-0.15cm}
\noindent subject to a joint distribution that factorizes as
\vspace{-0.2cm}
\begin{align}\label{eq:PrimSOMARC_JointDist}
p(s_1,s_2)p(x_1|s_1)p(x_2|s_2)p(y_3,y_S|x_1,x_2).
\end{align}
\end{MyProposition}
\vspace{-0.15cm}
\begin{IEEEproof}
We begin with the constraints due to decoding at the relay given by \eqref{bnd:Joint_rly_S1}--\eqref{bnd:Joint_rly_S1S2}. For the RHS of condition \eqref{bnd:Joint_rly_S1} (with $\mathcal{W}_3=\phi$) we write:
\begin{subequations} \label{eq:ExtraThm3RlySum}
\vspace{-0.2cm}
\begin{align}
I(X_1;Y_3|S_2, V_1,X_2,X_3) & \stackrel{(a)}{=} H(Y_3|S_2, V_1,X_2, X_3) - H(Y_3|S_2, X_1,X_2) \nonumber \\
%
& \stackrel{(b)}{\le} H(Y_3|S_2,X_2) - H(Y_3|S_2,X_1,X_2) \nonumber \\
%
& = I(X_1;Y_3|S_2,X_2), \label{eq:ExtraThm3RlySingle}
\end{align}
\vspace{-0.15cm}
\noindent where (a) follows from the definition of the PSOMARC which implies that the
Markov chain $(V_1, X_3) \leftrightarrow (S_2, X_1,X_2) \leftrightarrow Y_3$ holds; and (b) follows from the fact the conditioning reduces entropy. Similarly, for the RHS of conditions \eqref{bnd:Joint_rly_S2}--\eqref{bnd:Joint_rly_S1S2} we have:
\vspace{-0.2cm}
\begin{align}
I(X_2;Y_3|S_1, V_2,X_1, X_3) & \le I(X_2;Y_3|S_1,X_1) \\
I(X_1,X_2;Y_3|V_1,V_2,X_3) & \le I(X_1,X_2;Y_3).
\end{align}
\end{subequations}
\vspace{-0.15cm}
Next, consider the constraints due to decoding at the destination given by \eqref{bnd:Joint_dst_S1}--\eqref{bnd:Joint_dst_S1S2}, and recall that for the PSOMARC the channel output at the destination, $Y$, is replaced by the pair of channel outputs $(Y_R,Y_S)$. For the RHS of \eqref{bnd:Joint_dst_S1} we write:
\vspace{-0.25cm}
\begin{subequations} \label{eq:ExtraThm4DstSingle_2_sum}
\begin{align}
& I(X_1,X_3;Y_R,Y_S|S_1,V_2,X_2) \nonumber \\
%
& \qquad = I(X_1;Y_R,Y_S|S_1,V_2,X_2) + I(X_3;Y_R|S_1,V_2,X_1,X_2) + I(X_3;Y_S|S_1,V_2,X_1,X_2,Y_R) \nonumber \\
%
& \qquad \stackrel{(a)}{=} I(X_1;Y_S|S_1,V_2,X_2) + I(X_1;Y_R|S_1,V_2,X_2,Y_S) + I(X_3;Y_R|S_1,V_2,X_1,X_2) \nonumber \\
%
& \qquad = I(X_1;Y_S|S_1,V_2,X_2) + H(Y_R|S_1,V_2,X_2,Y_S) - H(Y_R|S_1,V_2,X_1,X_2,Y_S) \nonumber \\
& \qquad \qquad + H(Y_R|S_1,V_2,X_1,X_2) - H(Y_R|S_1,V_2,X_1,X_2,X_3) \nonumber \\
%
& \qquad \stackrel{(b)}{=} I(X_1;Y_S|S_1,V_2,X_2) + H(Y_R|S_1,V_2,X_2,Y_S) - H(Y_R|X_3) \nonumber \\
%
& \qquad \stackrel{(c)}{\le} I(X_1;Y_S|S_1,X_2) + I(X_3;Y_R), \label{eq:ExtraThm4DstSingle_1}
\end{align}
\vspace{-0.2cm}
\noindent where (a) follows from the fact that $Y_S$ is uniquely determined by $X_1$ and $X_2$, and therefore it follows that $I(X_3;Y_S|S_1,V_2,X_1,X_2,Y_R)=0$; (b) follows from the Markov chain $Y_S \leftrightarrow (S_1,V_2,X_1,X_2) \leftrightarrow Y_R$ (which directly follows from the definition of the conditional distribution function of the SOMARC: $p(y_R,y_S,y_3|x_1,x_2,x_3)=p(y_R|x_3)p(y_S,y_3|x_1,x_2)$), and from the Markov chain $(S_1,V_2,X_1,X_2) \leftrightarrow X_3 \leftrightarrow Y_R$; and (c) follows from the arguments leading to \eqref{eq:ExtraThm3RlySingle} and from the fact that conditioning reduces entropy. Similarly, for the RHS of conditions \eqref{bnd:Joint_dst_S2}--\eqref{bnd:Joint_dst_S1S2} we have:
\vspace{-0.25cm}
\begin{align}
I(X_2,X_3;Y_R,Y_S|S_2,V_1,X_1) & \le I(X_2;Y_S|S_2,X_1) + I(X_3;Y_R) \\
%
I(X_1,X_2,X_3;Y_R,Y_S|S_1,S_2) & \le I(X_1,X_2;Y_S|S_1,S_2) + I(X_3;Y_R).
\end{align}
\end{subequations}
\vspace{-0.15cm}
\noindent Finally, substituting $I(X_3;Y_R)=C_3$ in \eqref{eq:ExtraThm4DstSingle_2_sum} and combining with \eqref{eq:ExtraThm3RlySum}, we obtain the RHSs of conditions \eqref{bnd:Thm7}.
Note that conditions \eqref{bnd:Thm7} are subject to the chain:
\vspace{-0.2cm}
\begin{equation*}
p(s_1,s_2,v_1,v_2,x_1,x_2,y_3,y_s) = p(s_1,s_2)p(v_1)p(x_1|s_1,v_1)p(v_2)p(x_2|s_2,v_2)p(y_3,y_S|x_1,x_2).
\end{equation*}
\vspace{-0.2cm}
\noindent Furthermore, as \eqref{bnd:Thm7} is independent of $(V_1,V_2)$ then the resulting chain is:
\vspace{-0.2cm}
\begin{equation}
\sum_{(v_1,v_2) \in \mathcal{V}_1 \times \mathcal{V}_2} p(s_1,s_2,v_1,v_2,x_1,x_2,y_3,y_s) \mspace{-3mu} = \mspace{-3mu} p(s_1,s_2)p(x_1|s_1)p(x_2|s_2)p(y_3,y_S|x_1,x_2). \label{eq:chainSum}
\end{equation}
\vspace{-0.15cm}
\noindent Lastly, note that the upper bounds \eqref{eq:ExtraThm3RlySum}--\eqref{eq:ExtraThm4DstSingle_2_sum}, subject to the chain \eqref{eq:chainSum}, are obtained by letting $\mathcal{V}_1=\mathcal{V}_2=\phi$ in \eqref{bnd:Joint} and \eqref{eq:JntJointDist}. Thus, $\mathcal{V}_1=\mathcal{V}_2=\phi$ maximizes the sufficient conditions of \Thmref{thm:jointCond}.
\end{IEEEproof}
Next, note that the LHS of condition \eqref{bnd:Thm7_rly_S1S2}, evaluated for the sources defined in Table \ref{tab:SourceDist}, equals $\log_2 3$ bits. Therefore, for successfully transmitting $S_1$ and $S_2$ we must have that the RHS of \eqref{bnd:Thm7_rly_S1S2} is greater than (or equals to) $\log_2 3$.
Now, consider the RHS of condition \eqref{bnd:Thm7_rly_S1S2} for these sources and the PSOMARC defined in Table \ref{tab:PrimitiveSOMARC}: finding the maximum of $I(X_1, X_2;Y_3)$ over all $p(x_1|s_1)p(x_2|s_2)$ we have:
\vspace{-0.1cm}
\begin{equation}
\max_{p(x_1|s_1)p(x_2|s_2)} I(X_1,X_2;Y_3) = \max_{p(x_1|s_1)p(x_2|s_2)} H(Y_3),
\label{eq:Thm1Maximization}
\end{equation}
\vspace{-0.1cm}
\noindent which follows as the channel from $(X_1,X_2)$ to $Y_3$ is deterministic.
As $|\mathcal{Y}_3|=3$, it follows that ${\displaystyle \max_{p(x_1|s_1)p(x_2|s_2)} \mspace{-12mu} H(Y_3)= }$ ${\log_2 3}$ if and only if $\Pr \{Y_3 = j\}=1/3, j=0,1,2$. This requires that $\Pr \{(X_1,X_2) = (0,0) \} = \Pr \{(X_1,X_2) = (1,1) \} = 1/3$ and $\Pr \{((X_1,X_2) = (0,1)) \cup ((X_1,X_2) = (1,0)) \} = 1/3$.
Since the sources distribution is given, $\Pr \{(X_1,X_2) = (i,j) \}$ depends only on $p(x_1|s_1)p(x_2|s_2)$, which consists of four unknowns. This corresponds to an algebraic equations system with three equations, four unknowns, and the constraint that all the variables are in the range $[0,1]$. The two possible solutions of this system, solved using {\tt Mathematica}\footnote{Let $p_{i,j} \triangleq \Pr\{X_i = j | S_i = 0 \}, i,j=0,1$. The following algebraic equations system is solved:
\vspace{-0.15cm}
\begin{align*}
\mathtt{Solve[} & \mathtt{p_{00} \cdot p_{10} + p_{00} \cdot p_{11} + p_{01} \cdot p_{11} == 1 \&\& (1 - p_{00}) \cdot (1 - p_{10}) + (1 - p_{00}) \cdot (1 - p_{11}) + (1 - p_{01}) \cdot (1 - p_{11}) == 1 \&\&} \\
& \mathtt{p_{00} \cdot (1 - p_{10}) + p_{00} \cdot (1 - p_{11}) + p_{01} \cdot (1 - p_{11}) + (1 - p_{00}) \cdot p_{10} + (1 - p_{00}) \cdot p_{11} + (1 - p_{01}) \cdot p_{11} == 1 \&\&} \\
& \mathtt{0 <= p_{00} <= 1 \&\& 0 <= p_{01} <= 1 \&\& 0 <= p_{10} <= 1 \&\& 0 <= p_{11} <= 1, \{p_{00}, p_{01}, p_{10}, p_{11}\}]},
\end{align*}
\vspace{-0.2cm}
to obtain $\mathtt{\{\{p_{00} = 0, p_{01} = 1, p_{10} = 0, p_{11} = 1\}, \{p_{00} = 1, p_{01} = 0, p_{10} = 1, p_{11} = 0\}\}}$.
}, are deterministic mappings from $s_i$ to $x_i$.\footnote{This is also validated via an exhaustive search.}
The expression $I(X_1, X_2;Y_S|S_1,S_2) + C_3$, evaluated using each of these conditional distributions, equals $1$ bit.
Therefore, the RHS of condition \eqref{bnd:Thm7_rly_S1S2}, when evaluated using
these conditional distributions, is strictly smaller than $\log_2 3$.
This implies that for these sources and PSOMARC, condition \eqref{bnd:Thm7_rly_S1S2} holds with opposite strict inequality, and we conclude that reliable transmission via the scheme of \Thmref{thm:jointCond} is impossible.
%
%
\vspace{-0.25cm}
\subsection{Transmission Using the Scheme of Theorem \thmref{thm:jointCondFlip}} \label{annex:proofNotFeasible_thm2}
\vspace{-0.10cm}
Specializing the conditions of \Thmref{thm:jointCondFlip} in \eqref{bnd:JointFlip_rly_S1}--\eqref{bnd:JointFlip_dst_S1S2} to the PSOMARC by letting $\mathcal{W}_3=\mathcal{W}=\phi$ and $I(X_3;Y_R)=C_3$, results in the following sufficient conditions:
\ifthenelse{\boolean{SquizFlag}}{\begin{comment}}{}
\footnote{These conditions are derived in Appendix \ref{subannex:PSOMARC_CondsDeriveThm2}.}
\ifthenelse{\boolean{SquizFlag}}{\end{comment}}{}
\vspace{-0.25cm}
\begin{subequations} \label{bnd:Thm8}
\begin{align}
H(S_1|S_2) &< \min \{I(X_1;Y_3|S_1, X_2), I(X_1;Y_S|S_2,X_2) + C_3 \} \label{bnd:Thm8_rly_S1} \\
H(S_2|S_1) &< \min \{I(X_2;Y_3|S_2, X_1), I(X_2;Y_S|S_1,X_1) + C_3 \} \label{bnd:Thm8_rly_S2} \\
H(S_1,S_2) &< \min \{I(X_1, X_2;Y_3|S_1,S_2), I(X_1, X_2;Y_S) + C_3 \}, \label{bnd:Thm8_rly_S1S2}
\end{align}
\end{subequations}
\vspace{-0.25cm}
\noindent subject to the input distribution \eqref{eq:PrimSOMARC_JointDist}.
%
%
\noindent Consider maximizing the mutual information expression $I(X_1, X_2;Y_3|S_1,S_2)$ on the RHS of condition \eqref{bnd:Thm8_rly_S1S2} for the considered sources and PSOMARC, over all $p(x_1|s_1)p(x_2|s_2)$:
\vspace{-0.2cm}
\begin{align}
&\mspace{-15mu} \max_{p(x_1|s_1)p(x_2|s_2)} I(X_1,X_2;Y_3|S_1,S_2) \nonumber \\
& \stackrel{(a)}{=} \max_{p(x_1|s_1)p(x_2|s_2)} H(Y_3|S_1,S_2) \nonumber \\
& \stackrel{(b)}{=} \max_{p(x_1|s_1)p(x_2|s_2)} \sum_{\substack{(\tilde{s}_1,\tilde{s}_2) \in \mathcal{S}_1 \times \mathcal{S}_2, \\ p(\tilde{s}_1,\tilde{s}_2) \ne 0}}{ p(\tilde{s}_1,\tilde{s}_2) \cdot H\big(Y_3|(S_1,S_2)=(\tilde{s}_1,\tilde{s}_2)\big) } \nonumber \\
& \stackrel{(c)}{\le} \frac{1}{6} \cdot \sum_{\substack{(\tilde{s}_1,\tilde{s}_2) \in \mathcal{S}_1 \times \mathcal{S}_2, \\ p(\tilde{s}_1,\tilde{s}_2) \ne 0}}{ \max_{p(x_1|\tilde{s}_1)p(x_2|\tilde{s}_2)} \left\{ - \sum_{y_3 \in \mathcal{Y}_3}{p(y_3|\tilde{s}_1,\tilde{s}_2) \cdot \log_2 p(y_3|\tilde{s}_1,\tilde{s}_2)} \right\} } \nonumber \\
& = \frac{1}{6} \cdot \sum_{\substack{(\tilde{s}_1,\tilde{s}_2) \in \mathcal{S}_1 \times \mathcal{S}_2, \\ p(\tilde{s}_1,\tilde{s}_2) \ne 0}}{ \max_{p(x_1|\tilde{s}_1)p(x_2|\tilde{s}_2)}} \left\{ - \mspace{-10mu} \sum_{y_3 \in \mathcal{Y}_3}{\mspace{-35mu} \sum_{\mspace{50mu} (x_1,x_2) \in \mathcal{X}_1 \times \mathcal{X}_2 } \mspace{-50mu} p(y_3,x_1,x_2|\tilde{s}_1,\tilde{s}_2) \cdot \log_2 \left( \mspace{-35mu} \sum_{\mspace{50mu} (x_1,x_2) \in \mathcal{X}_1 \times \mathcal{X}_2 } \mspace{-50mu} p(y_3,x_1,x_2|\tilde{s}_1,\tilde{s}_2) \right) } \right\} \nonumber
\end{align}
\begin{align}
& \stackrel{(d)}{=} \frac{1}{6} \cdot \sum_{\substack{(\tilde{s}_1,\tilde{s}_2) \in \mathcal{S}_1 \times \mathcal{S}_2, \\ p(\tilde{s}_1,\tilde{s}_2) \ne 0}}{ \max_{p(x_1|\tilde{s}_1)p(x_2|\tilde{s}_2)}} \nonumber \\
& \mspace{50mu} \left\{ - \mspace{-10mu} \sum_{y_3 \in \mathcal{Y}_3}{\mspace{-40mu} \sum_{ \mspace{50mu} (x_1,x_2) \in \mathcal{X}_1 \times \mathcal{X}_2 } \mspace{-50mu} p(x_1|\tilde{s}_1)p(x_2|\tilde{s}_2) p(y_3|x_1,x_2) \cdot \log_2 \left( \mspace{-40mu} \sum_{\mspace{50mu} (x_1,x_2) \in \mathcal{X}_1 \times \mathcal{X}_2 } \mspace{-50mu} p(x_1|\tilde{s}_1)p(x_2|\tilde{s}_2) p(y_3|x_1,x_2) \right) } \right\} \nonumber \\
& \stackrel{(e)}{=} \frac{1}{6} \cdot \mspace{-50mu} \sum_{\mspace{50mu} \substack{(\tilde{s}_1,\tilde{s}_2) \in \mathcal{S}_1 \times \mathcal{S}_2, \\ p(\tilde{s}_1,\tilde{s}_2) \ne 0}}{\mspace{-7mu} \max_{p(x_1)p(x_2)}} \left\{ - \mspace{-10mu} \sum_{y_3 \in \mathcal{Y}_3}{\mspace{-40mu} \sum_{\mspace{50mu} (x_1,x_2) \in \mathcal{X}_1 \times \mathcal{X}_2 } \mspace{-50mu} p(x_1)p(x_2)p(y_3|x_1,x_2) \cdot \log_2 \left( \mspace{-40mu} \sum_{\mspace{50mu} (x_1,x_2) \in \mathcal{X}_1 \times \mathcal{X}_2 } \mspace{-50mu} p(x_1)p(x_2)p(y_3|x_1,x_2) \right)} \right\} \nonumber \\
& = \max_{p(x_1)p(x_2)} H(Y_3) \nonumber \\
& \stackrel{(f)}{=} 1.5,
\end{align}
\vspace{-0.25cm}
\noindent where (a) follows from the fact that $Y_3$ is a deterministic function of $(X_1,X_2)$; (b) follows from the definition of conditional entropy; (c) follows from the joint distribution of the sources in Table \ref{tab:SourceDist} and the fact that the maximum of a sum is less than the sum of the maximum of the summands; (d) follows from the Markov chain $(S_1,S_2)-(X_1,X_2)-Y_3$; (e) follows from the fact that since $\tilde{s}_1$ and $\tilde{s}_2$ appear only in the conditioning of the conditional distributions $p(x_1|\tilde{s}_1),p(x_2|\tilde{s}_2)$, the maximizing $p(x_1|\tilde{s}_1)p(x_2|\tilde{s}_2)$ is the same for any pair $(\tilde{s}_1,\tilde{s}_2)$. Thus, the maximizing $p(x_1|\tilde{s}_1)p(x_2|\tilde{s}_2)$ is independent of the value of $(\tilde{s}_1,\tilde{s}_2)$;
finally, (f) follows from \cite{Cover:80}.
Recall that $H(S_1,S_2) = \log_2 3$ bits. Thus, $H(S_1,S_2) > \max_{p(x_1|s_1)p(x_2|s_2)} I(X_1,X_2;Y_3|S_1,S_2)$,
%
\noindent and \eqref{bnd:Thm8_rly_S1S2} holds with strict opposite inequality. Therefore we conclude that reliable transmission via the scheme of \Thmref{thm:jointCondFlip} is impossible.
This concludes the proof of Prop. \ref{prop:NotFeasible}.
\vspace{-0.3cm}
\section{Proof of Proposition \ref{prop:feasible}} \label{annex:proofFeasible}
%
%
Here, instead of specializing the conditions of \Thmref{thm:jointCond_NewSimult} to the PSOAMRC, we analyze the decoding rules of \Thmref{thm:jointCond_NewSimult} given in \eqref{eq:RelayJntNewDecType}--\eqref{eq:DestJntNewSimultDecTypeList} for a specific $p(x_i|s_i),i=1,2$.
Let $p(x_i|s_i), i=1,2$, be the deterministic distribution $p(x_i|s_i)=\delta(x_i-s_i)$,
where $\delta(x)$ is the Kronecker Delta function,
and set $\mathcal{V}_1=\mathcal{V}_2=\phi$. Hence, there is no superposition encoding at the sources, and the cooperation between the sources and the relay is based only on the codeword transmitted by the relay.
\vspace{-0.4cm}
\subsection{Encoding at the Relay}
Let $\mathcal{Q} \triangleq \{ 1,2,\dots, 2^n\}$, and let $f_3: (\mathbf{s}_1,\mathbf{s}_2) \mapsto \mathcal{Q}$, be the encoding function at the relay.
At block $b=1$, the relay transmits the codeword $1$.
Assume that at block $b, b=2,3,\dots,B,B+1$, the relay has the estimates $(\tilde{\mathbf{s}}_{1,b-1}, \tilde{\mathbf{s}}_{2,b-1})$ of $(\mathbf{s}_{1,b-1}, \mathbf{s}_{2,b-1})$. Then, at time $b$, the relay transmits the channel codeword $q_{b-1} = f_3(\tilde{\svec}_{1,b-1},\tilde{\svec}_{2,b-1}), q_{b-1} \in \mathcal{Q}$.
\vspace{-0.4cm}
\subsection{Decoding at the Relay}
\vspace{-0.2cm}
\subsubsection{Decoding rule} For the mapping defined in Table \ref{tab:PrimitiveSOMARC} and the specified $p(x_i|s_i)$, the relay decoding rule \eqref{eq:RelayJntNewDecType} is specialized to the following decoding rule: {\slshape the relay decodes $(\mathbf{s}_{1,b}, \mathbf{s}_{2,b})$ by looking for a unique pair $(\tilde{\svec}_{1}, \tilde{\svec}_{2}) \in \mathcal{S}_1^n \times \mathcal{S}_2^n$ such that $\big(\tilde{\svec}_{1}, \tilde{\svec}_{2}, \mathbf{y}_{3,b}\big) \in A_{\epsilon}^{*(n)}$}. Denote the decoded sequences by $(\tilde{\svec}_{1,b}, \tilde{\svec}_{2,b})$.
\subsubsection{Error probability analysis} Let $\mathcal{E}_r \triangleq \left\{ \left( \tilde{\mathbf{S}}_{1,b}, \tilde{\mathbf{S}}_{2,b} \right) \ne \left( \mathbf{S}_{1,b}, \mathbf{S}_{2,b} \right) \right\}$. The average probability of error for decoding at the relay at block $b$, $\bar{P}_{r,b}^{(n)}$, is defined as:
\vspace{-0.2cm}
\begin{align}
\bar{P}_{r,b}^{(n)} & \triangleq \sum_{(\mathbf{s}_{1,b},\mathbf{s}_{2,b}) \in \mathcal{S}_1^n \times \mathcal{S}_2^n}{\mspace{-24mu} p(\mathbf{s}_{1,b},\mathbf{s}_{2,b})} \Pr \Big( \mathcal{E}_r | \mathbf{s}_{1,b},\mathbf{s}_{2,b} \Big) \nonumber \\
&\leq \sum_{(\mathbf{s}_{1,b},\mathbf{s}_{2,b}) \notin A_{\epsilon}^{*(n)}(S_1,S_2)}{\mspace{-54mu} p(\mathbf{s}_{1,b},\mathbf{s}_{2,b})} + \sum_{(\mathbf{s}_{1,b},\mathbf{s}_{2,b}) \in A_{\epsilon}^{*(n)}(S_1,S_2)}{\mspace{-54mu} p(\mathbf{s}_{1,b},\mathbf{s}_{2,b})} \Pr \Big( \mathcal{E}_r | (\mathbf{s}_{1,b},\mathbf{s}_{2,b})\in A_{\epsilon}^{*(n)} \Big). \label{eq:RelayDecErrProbDef_PrimSOMARC}
\end{align}
\noindent From \cite[Thm. 6.9]{YeungBook} the first sum in \eqref{eq:RelayDecErrProbDef_PrimSOMARC} can be bounded by $\epsilon$. Next, by the union bound we write:
\vspace{-0.2cm}
\begin{align}
\Pr \Big( \mathcal{E}_r | (\mathbf{s}_{1,b},\mathbf{s}_{2,b})\in A_{\epsilon}^{*(n)} \Big) & \leq \Pr \Big( \big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{Y}_{3,b}\big) \notin A_{\epsilon}^{*(n)} | (\mathbf{s}_{1,b},\mathbf{s}_{2,b})\in A_{\epsilon}^{*(n)} \Big) \nonumber \\
& \quad + \Pr \Big( \exists (\tilde{\svec}_{1}, \tilde{\svec}_{2}) \neq (\mathbf{s}_{1,b}, \mathbf{s}_{2,b}): (\tilde{\svec}_{1}, \tilde{\svec}_{2}, \mathbf{Y}_{3,b}\big) \in A_{\epsilon}^{*(n)} | (\mathbf{s}_{1,b},\mathbf{s}_{2,b})\in A_{\epsilon}^{*(n)} \Big). \label{eq:PrimSOMARC_relayErrorProb}
\end{align}
\vspace{-0.2cm}
\noindent For the specified $p(x_i|s_i), i=1,2$, and the channel mapping defined in Table \ref{tab:PrimitiveSOMARC}, $Y_3$ is a deterministic function of the sources $S_1$ and $S_2$. Moreover, there is one-to-one mapping between the source pairs $(S_1,S_2)$ and $Y_3$. Hence, for each possible source pair $(S_1,S_2)$ there is a unique value of $Y_3$, and we conclude that:
\vspace{-0.2cm}
\begin{equation}
\Pr \Big( \big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{Y}_{3,b}\big) \notin A_{\epsilon}^{*(n)} | (\mathbf{s}_{1,b},\mathbf{s}_{2,b})\in A_{\epsilon}^{*(n)} \Big) = 0.
\label{eq:PrimSOMARC_RelayGoodSources}
\end{equation}
\vspace{-0.2cm}
\noindent From the one-to-one mapping between the source pairs $(S_1,S_2)$ and $Y_3$, and from the definition of strong typicality, \cite[Ch. 6.1]{YeungBook}, it follows that:
\vspace{-0.2cm}
\begin{align}
\Pr \Big( \exists (\tilde{\svec}_{1}, \tilde{\svec}_{2}) \neq (\mathbf{s}_{1,b}, \mathbf{s}_{2,b}): (\tilde{\svec}_{1}, \tilde{\svec}_{2}, \mathbf{Y}_{3,b}\big) \in A_{\epsilon}^{*(n)} | (\mathbf{s}_{1,b},\mathbf{s}_{2,b})\in A_{\epsilon}^{*(n)} \Big) = 0.
\label{eq:PrimSOMARC_RelayBadSources}
\end{align}
\noindent Combining \eqref{eq:PrimSOMARC_relayErrorProb}--\eqref{eq:PrimSOMARC_RelayBadSources} yields $\bar{P}_{r,b}^{(n)} \leq \epsilon$ for sufficiently large $n$. We conclude that the sources of Table \ref{tab:SourceDist} can be reliably transmitted over the channel to the relay.
\vspace{-0.4cm}
\subsection{Decoding at the Destination}
\vspace{-0.2cm}
\subsubsection{Decoding rule}
Recall that $q_{b}$ is available at the destination assuming the relay correctly decoded the source sequences.
The destination decoding rule of \Thmref{thm:jointCond_NewSimult}, see \eqref{eq:DestJntNewSimultDecTypeList}, is specialized to the following decoding rule:\footnote{This follows from the fact that the relay's information is transmitted via an orthogonal link.} {\slshape the destination decodes $(\mathbf{s}_{1,b}, \mathbf{s}_{2,b})$, by looking for a unique pair $(\hat{\svec}_{1}, \hat{\svec}_{2})\in \mathcal{S}_1^n \times \mathcal{S}_2^n$ such that $\big(\hat{\svec}_{1}, \hat{\svec}_{2}, \mathbf{y}_{S,b}\big) \in A_{\epsilon}^{*(n)}$ and $f_3(\hat{\svec}_{1}, \hat{\svec}_{2})=q_{b}$}. Denote the decoded sequences by $(\hat{\svec}_{1,b}, \hat{\svec}_{2,b})$.
\subsubsection{Error probability analysis} Let $\mathcal{E}_d \triangleq \left\{ \big( \hat{\mathbf{S}}_{1,b}, \hat{\mathbf{S}}_{2,b} \big) \ne \left( \mathbf{S}_{1,b}, \mathbf{S}_{2,b} \right) \right\}$. Following the same arguments that led to \eqref{eq:RelayDecErrProbDef_PrimSOMARC}, the average probability of decoding error at the destination at block $b$, $\bar{P}_{d,b}^{(n)}$ can be upper bounded as:
\vspace{-0.25cm}
\begin{align}
\bar{P}_{d,b}^{(n)} & \leq \epsilon + \sum_{(\mathbf{s}_{1,b},\mathbf{s}_{2,b}) \in A_{\epsilon}^{*(n)}}{\mspace{-30mu} p(\mathbf{s}_{1,b},\mathbf{s}_{2,b})} \Pr \Big( \mathcal{E}_d \big| (\mathbf{s}_{1,b},\mathbf{s}_{2,b}) \in A_{\epsilon}^{*(n)} \Big). \label{eq:DestDecErrProbDef_PrimSOMARC}
\end{align}
\vspace{-0.1cm}
\noindent Using the union bound $\Pr \Big( \mathcal{E}_d \big| (\mathbf{s}_{1,b},\mathbf{s}_{2,b}) \in A_{\epsilon}^{*(n)} \Big)$ can be upper bounded by:
\vspace{-0.2cm}
\begin{align}
& \Pr \Big( \big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{Y}_{S,b}\big) \notin A_{\epsilon}^{*(n)} \big| (\mathbf{s}_{1,b},\mathbf{s}_{2,b}) \in A_{\epsilon}^{*(n)} \Big) + \nonumber \\
& \qquad \Pr \Big( \exists (\hat{\svec}_{1}, \hat{\svec}_{2}) \neq (\mathbf{s}_{1,b}, \mathbf{s}_{2,b}): \big\{(\hat{\svec}_{1}, \hat{\svec}_{2}, \mathbf{Y}_{S,b}\big) \in A_{\epsilon}^{*(n)} \big\} \cap \big\{ f_3(\hat{\svec}_{1}, \hat{\svec}_{2})=q_{b}\big\} \big| (\mathbf{s}_{1,b},\mathbf{s}_{2,b}) \in A_{\epsilon}^{*(n)} \Big). \label{eq:PrimSOMARC_destErrorProb}
\end{align}
\vspace{-0.25cm}
\noindent Since $x_i=s_i, i=1,2$, and $Y_S$ is a deterministic function of $(X_1,X_2)$ then as $(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}) \in A_{\epsilon}^{*(n)}$ it follows that $(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{Y}_{S,b}) \in A_{\epsilon}^{*(n)}$, thus
\vspace{-0.25cm}
\begin{equation}
\Pr \Big( \big(\mathbf{s}_{1,b}, \mathbf{s}_{2,b}, \mathbf{Y}_{S,b}\big) \notin A_{\epsilon}^{*(n)} \big| ((\mathbf{s}_{1,b},\mathbf{s}_{2,b}) \in A_{\epsilon}^{*(n)} \Big) = 0.
\label{eq:PrimSOMARC_DestGoodSources}
\end{equation}
\vspace{-0.2cm}
\noindent The channel to the destination does not provide a one-to-one mapping between the pair $(S_1,S_2)$ and $Y_S$.
\noindent Let $\theta(y_S)$ denote the inverse mapping from the channel output $Y_S$ to the sources, e.g., $\theta(0) = \{ (0,0), (0,1)\}$.
From \cite[Def. 6.6]{YeungBook} it follows that if $\big(\mathbf{s}_{1}, \mathbf{s}_{2}, \mathbf{Y}_{S}\big) \in A_{\epsilon}^{*(n)}$ then:
\vspace{-0.2cm}
\begin{align}
\forall y_{S,k}: (s_{1,k}, s_{2,k}) \in \theta(y_{S,k}), \quad k=1,2,\dots,n.
\label{eq:inverseMapping}
\end{align}
\vspace{-0.2cm}
\noindent Furthermore, $\forall y_S\in\mathcal{Y}_S: \left\| \theta(y_S) \right\| = 2$. Therefore, by mapping the two elements of $\theta(y_S)$ into different symbols transmitted from the relay we can guarantee that the condition $f_3(\hat{\svec}_{1},\hat{\svec}_{2})=q_{b}$ holds only for the transmitted source sequences.\footnote{From the fact that $\forall y\in\mathcal{Y}: \left\| \theta(y) \right\| = 2$ it follows that resolving the ambiguity in $\theta(y)$ requires 1 bit per source pair, and therefore, this information can be transmitted from the relay via the relay-destination link with capacity $C_3=1$ bit.
}
\noindent Hence, we conclude that the combination of the codeword transmitted by the relay and $\mathbf{Y}_S$ uniquely identifies the transmitted source pair. Thus,
\vspace{-0.2cm}
\begin{align}
& \mspace{-5mu} \Pr \Big( \exists (\hat{\svec}_{1}, \hat{\svec}_{2}) \neq (\mathbf{s}_{1,b}, \mathbf{s}_{2,b}): \big\{(\hat{\svec}_{1}, \hat{\svec}_{2}, \mathbf{Y}_{S,b}\big) \in A_{\epsilon}^{*(n)}\big\} \cap \big\{ f_3(\hat{\svec}_{1}, \hat{\svec}_{2})=q_{b}\big\} \Big| (\mathbf{s}_{1,b},\mathbf{s}_{2,b}) \in A_{\epsilon}^{*(n)} \Big) = 0. \label{eq:PrimSOMARC_destBadSources}
\end{align}
\vspace{-0.2cm}
\noindent Combining \eqref{eq:DestDecErrProbDef_PrimSOMARC}--\eqref{eq:PrimSOMARC_destBadSources} yields $\bar{P}_{d,b}^{(n)} \leq \epsilon$ for $n$ large enough. We conclude that the sources of Table \ref{tab:SourceDist} can be reliably transmitted over the channel to the destination.
\ifthenelse{\boolean{SquizFlag}}{\begin{comment}}{}
\section{Conditional Distributions Maximizing \eqref{eq:Thm1Maximization}} \label{annex:ProofPropInfeasibleTables}
Table \ref{tab:CondMaxDistributions} details the conditional distributions which maximize \eqref{eq:Thm1Maximization}.
\begin{figure}[ht]
\begin{center}
\subfloat[$p_1(x_1|s_1)$]{\scalebox{0.75}{
\begin{tabular}{|c|c|c|c|}
\hline
$S_1$ \textbackslash $S_2$ & 0 & 1 & 2 \\
\hline
0 & 0 & 0 & 1 \\
\hline
1 & 0 & 1 & 0 \\
\hline
2 & 1 & 0 & 0 \\
\hline
\end{tabular}}
}
\quad
\subfloat[$p_1(x_2|s_2)$]{\scalebox{0.75}{
\begin{tabular}{|c|c|c|c|}
\hline
$S_1$ \textbackslash $S_2$ & 0 & 1 & 2 \\
\hline
0 & 1 & 0 & 0 \\
\hline
1 & 0 & 0 & 1 \\
\hline
2 & 0 & 1 & 0 \\
\hline
\end{tabular}}
}
\quad
\subfloat[$p_2(x_1|s_1)$]{\scalebox{0.75}{
\begin{tabular}{|c|c|c|c|}
\hline
$S_1$ \textbackslash $S_2$ & 0 & 1 & 2 \\
\hline
0 & 0 & 1 & 0 \\
\hline
1 & 0 & 0 & 1 \\
\hline
2 & 1 & 0 & 0 \\
\hline
\end{tabular}}
}
\quad
\subfloat[$p_2(x_2|s_2)$]{\scalebox{0.75}{
\begin{tabular}{|c|c|c|c|}
\hline
$S_1$ \textbackslash $S_2$ & 0 & 1 & 2 \\
\hline
0 & 0 & 1 & 0 \\
\hline
1 & 0 & 0 & 1 \\
\hline
2 & 1 & 0 & 0 \\
\hline
\end{tabular}}
}
\subfloat[$p_3(x_1|s_1)$]{\scalebox{0.75}{
\begin{tabular}{|c|c|c|c|}
\hline
$S_1$ \textbackslash $S_2$ & 0 & 1 & 2 \\
\hline
0 & 0 & 0 & 1 \\
\hline
1 & 1 & 0 & 0 \\
\hline
2 & 0 & 1 & 0 \\
\hline
\end{tabular}}
}
\quad
\subfloat[$p_3(x_2|s_2)$]{\scalebox{0.75}{
\begin{tabular}{|c|c|c|c|}
\hline
$S_1$ \textbackslash $S_2$ & 0 & 1 & 2 \\
\hline
0 & 0 & 0 & 1 \\
\hline
1 & 1 & 0 & 0 \\
\hline
2 & 0 & 1 & 0 \\
\hline
\end{tabular}}
}
\quad
\subfloat[$p_4(x_1|s_1)$]{\scalebox{0.75}{
\begin{tabular}{|c|c|c|c|}
\hline
$S_1$ \textbackslash $S_2$ & 0 & 1 & 2 \\
\hline
0 & 0 & 1 & 0 \\
\hline
1 & 1 & 0 & 0 \\
\hline
2 & 0 & 0 & 1 \\
\hline
\end{tabular}}
}
\quad
\subfloat[$p_4(x_2|s_2)$]{\scalebox{0.75}{
\begin{tabular}{|c|c|c|c|}
\hline
$S_1$ \textbackslash $S_2$ & 0 & 1 & 2 \\
\hline
0 & 0 & 0 & 1 \\
\hline
1 & 0 & 1 & 0 \\
\hline
2 & 1 & 0 & 0 \\
\hline
\end{tabular}
}}
\subfloat[$p_5(x_1|s_1)$]{\scalebox{0.75}{
\begin{tabular}{|c|c|c|c|}
\hline
$S_1$ \textbackslash $S_2$ & 0 & 1 & 2 \\
\hline
0 & 1 & 0 & 0 \\
\hline
1 & 0 & 0 & 1 \\
\hline
2 & 0 & 1 & 0 \\
\hline
\end{tabular}}
}
\quad
\subfloat[$p_5(x_2|s_2)$]{\scalebox{0.75}{
\begin{tabular}{|c|c|c|c|}
\hline
$S_1$ \textbackslash $S_2$ & 0 & 1 & 2 \\
\hline
0 & 0 & 1 & 0 \\
\hline
1 & 1 & 0 & 0 \\
\hline
2 & 0 & 0 & 1 \\
\hline
\end{tabular}}
}
\quad
\subfloat[$p_6(x_1|s_1)$]{\scalebox{0.75}{
\begin{tabular}{|c|c|c|c|}
\hline
$S_1$ \textbackslash $S_2$ & 0 & 1 & 2 \\
\hline
0 & 1 & 0 & 0 \\
\hline
1 & 0 & 1 & 0 \\
\hline
2 & 1 & 0 & 1 \\
\hline
\end{tabular}}
}
\quad
\subfloat[$p_6(x_2|s_2)$]{\scalebox{0.75}{
\begin{tabular}{|c|c|c|c|}
\hline
$S_1$ \textbackslash $S_2$ & 0 & 1 & 2 \\
\hline
0 & 1 & 0 & 0 \\
\hline
1 & 0 & 1 & 0 \\
\hline
2 & 0 & 0 & 1 \\
\hline
\end{tabular}}
}
\caption{Conditional distributions maximizing \eqref{eq:Thm1Maximization}: $p_j(x_1|s_1),p_j(x_2|s_2), j=1,2,\dots,6$. \label{tab:CondMaxDistributions}}
\vspace{-0.8cm}
\end{center}
\end{figure}
\ifthenelse{\boolean{SquizFlag}}{\end{comment}}{}
\vspace{-0.2cm}
\section{Proofs of Theorem \thmref{thm:OuterMarkov} and Proposition \ref{prop:VGeneral}} \label{annex:MAC_BC_proofs}
\vspace{-0.25cm}
\subsection{Proof of \Thmref{thm:OuterMarkov}} \label{subsec:MarkovBound_proof}
\vspace{-0.15cm}
Assume a sequence of encoders $f_i^{(n)}, i=1,2,3$, and decoders $g^{(n)}$ is specified such that $P_e^{(n)} \rightarrow 0$ as $n \rightarrow \infty$. Fano's inequality \cite[Ch. 2.8]{YeungBook}, in the context of the current scenario, states that:
\vspace{-0.25cm}
\begin{eqnarray}
H(S_1^n, S_2^n| \hat{S}_{1}^n, \hat{S}_{2}^n) \leq 1 + n P_e^{(n)} \log_2 \left| \mathcal{S}_1 \times \mathcal{S}_2 \right|\triangleq n \gamma(P_e^{(n)}),
\label{eq:FanoDest}
\end{eqnarray}
\vspace{-0.2cm}
\noindent where $\gamma(x)$ is a non-negative function that approaches $\frac{1}{n}$ as $x \rightarrow 0$.
We also obtain:
\vspace{-0.2cm}
\begin{align}
H(S_1^n, S_2^n| \hat{S}_{1}^n, \hat{S}_{2}^n) & \stackrel{(a)}{\ge} H(S_1^n, S_2^n |W^n, Y^n ) \stackrel{(b)}{\geq} H(S_1^n |S_2^n, W^n, Y^n) \label{eq:generalDestEntropyBasic},
\end{align}
\vspace{-0.2cm}
\noindent where (a) follows from the fact that conditioning reduces entropy, and from the fact that $(\hat{S}_{1}^n, \hat{S}_{2}^n)$ is a deterministic function of $(Y^n,W^n)$; (b) follows from non-negativity of the entropy function for discrete sources.
Constraint \eqref{bnd:outr_markov_dst_S1} is a consequence of the following chain of inequalities:
\vspace{-0.15cm}
\begin{align}
& \sum_{k=1}^{n}{I(X_{1,k}, X_{3,k};Y_{k}| S_{2,k}, X_{2,k}, W_{k})} \nonumber \\
& \qquad \qquad \stackrel{(a)}{=} \sum_{k=1}^{n}{\Big[H(Y_{k}|S_{2,k}, X_{2,k}, W_{k})} - H\big(Y_{k}|S_1^n, S_2^n, X_{1,1}^{k}, X_{2,1}^{k}, X_{3,1}^{k}, W^n, W_{3,1}^n, Y^{k-1}, Y_{3,1}^{k-1}\big)\Big] \nonumber \\
& \qquad \qquad \stackrel{(b)}{\geq} \sum_{k=1}^{n}{\Big[H(Y_{k}|S_2^n, X_{2,k}, W^n, Y^{k-1})} - H(Y_{k}|S_1^n, S_2^n, W^n, W_{3,1}^n, Y^{k-1})\Big] \nonumber \\
& \qquad \qquad \stackrel{(c)}{=} I(S_1^n, W_{3,1}^n; Y^n|S_2^n, W^n) \nonumber \\
& \qquad \qquad \stackrel{(d)}{\geq} H(S_1^n | S_2^n, W^n) - H(S_1^n |S_2^n, W^n, Y^n) \nonumber \\
& \qquad \qquad \stackrel{(e)}{\geq} nH(S_1 | S_2, W) - n\gamma(P_e^{(n)}) \label{eq:generalDestSingleChain},
\end{align}
\vspace{-0.2cm}
\noindent where (a) follows from the memoryless channel assumption (see \eqref{eq:MARCchanDist}) and the causal Markov relation $(S_1^n, S_2^n, W^n,$ $W_{3,1}^n) \leftrightarrow (X_{1,1}^{k}, X_{2,1}^{k}, X_{3,1}^{k}, Y^{k-1}, Y_{3,1}^{k-1}) \leftrightarrow Y_{k}$ (see \cite{Massey:90});
(b) follows from the fact that conditioning reduces entropy;
(c) follows from the fact that $X_{2,k}$ is a deterministic function of $S_2^n$;
(d) follows from the non-negativity of the mutual information;
and (e) follows from the memoryless sources and side information assumption and from \eqref{eq:FanoDest}--\eqref{eq:generalDestEntropyBasic}.
Following arguments similar to those that led to \eqref{eq:generalDestSingleChain} we obtain:
\vspace{-0.2cm}
\begin{subequations} \label{eq:generalDestChain}
\begin{align}
H(S_2|S_1, W) &\leq \frac{1}{n} \sum_{k=1}^{n}{I(X_{2,k}, X_{3,k};Y_{k} | S_{1,k}, X_{1,k}, W_{k})} + \gamma(P_e^{(n)}) \label{eq:generalDestSingleChain2} \\
H(S_1, S_2| W) & \leq \frac{1}{n} \sum_{k=1}^{n}{I(X_{1,k},X_{2,k}, X_{3,k};Y_{k}| W_{k})} + \gamma(P_e^{(n)}). \label{eq:generalDestJointChain12}
\end{align}
\end{subequations}
\vspace{-0.1cm}
\noindent Note that the following three expressions, $I(X_{1,k}, X_{3,k};Y_{k}| S_{2,k}, X_{2,k}, W_{k})$, $I(X_{2,k}, X_{3,k};Y_{k} | S_{1,k}, X_{1,k}, W_{k})$, and $I(X_{1,k},X_{2,k}, X_{3,k};Y_{k}| W_{k})$, depend on the marginal conditional distribution:
\vspace{-0.2cm}
\begin{equation*}
p(x_{1,k},x_{2,k},x_{3,k}|s_{1,k},s_{2,k}) = p(x_{1,k},x_{2,k}|s_{1,k},s_{2,k})p(x_{3,k}|s_{1,k},s_{2,k},x_{1,k},x_{2,k}),
\end{equation*}
\vspace{-0.2cm}
\noindent and on $p(s_{1,k},s_{2,k}, w_k)$ and $p(y_k|x_{1,k},x_{2,k},x_{2,k})$. Moreover, note that $X_{1,k}$ is a function of $S_1^n$ while $X_{2,k}$ is a function of $S_2^n$, and therefore the Markov chain in \eqref{eq:MarkovChain} holds. Thus, it follows that:
\vspace{-0.2cm}
\begin{align}
p(x_{1,k},x_{2,k}|s_{1,k},s_{2,k}) \in \mathcal{B}_{X_1 X_2|S_1 S_2} \subseteq \mathcal{B}_{X_1 X_2|S_1 S_2}'. \label{eq:probLargerSet}
\end{align}
\vspace{-0.2cm}
\noindent Next, we introduce the time-sharing random variable $Q$ uniformly distributed over $\{ 1,2,\dots,n \}$ and independent of all other random variables. We can write the following:
\vspace{-0.15cm}
\begin{align}
\frac{1}{n} \sum_{k=1}^{n}{I(X_{1,k}, X_{3,k};Y_{k} | S_{2,k},X_{2,k}, W_{k})} & = I(X_{1,Q}, X_{3,Q};Y_{Q} | S_{2,Q}, X_{2,Q}, W_{Q}, Q) \nonumber \\
& = I(X_{1}, X_{3};Y | S_{2}, X_{2}, W, Q),
\label{eq:MarkovBoundConcavity}
\end{align}
\vspace{-0.2cm}
\noindent where $X_1 \triangleq X_{1,Q}$, $X_2 \triangleq X_{2,Q}$, $X_3 \triangleq X_{3,Q}$, $Y \triangleq Y_{Q}$, $S_2 \triangleq S_{2,Q}$ and $W \triangleq W_{Q}$. Furthermore, since for all values of $q$ we have $p(x_{1,q},x_{2,q}|s_{1,q},s_{2,q},Q=k) = p(x_{1,k},x_{2,k}|s_{1,k},s_{2,k})$ which satisfies \eqref{eq:probLargerSet}, then we have that for $k=1,2,\dots,n$ it holds that:
\vspace{-0.3cm}
\begin{align}
p(x_{1,q},x_{2,q}|s_{1,q},s_{2,q}, Q=k) \in \mathcal{B}_{X_1 X_2|S_1 S_2}'.
\end{align}
\vspace{-0.2cm}
\noindent Finally, note that for all $k$, the expressions and structural constraints on the distribution chain are identical. Thus, repeating the steps leading to \eqref{eq:MarkovBoundConcavity} for \eqref{eq:generalDestSingleChain2} and \eqref{eq:generalDestJointChain12}, and taking the limit $n \mspace{-4mu} \rightarrow \mspace{-4mu} \infty$, leads to the constraints in~\eqref{bnd:outr_markov_dst}.
%
\vspace{-0.25cm}
\subsection{Proof of Proposition \ref{prop:VGeneral}} \label{annex:ProofOuterV}
\vspace{-0.1cm}
First, define the auxiliary RV $V_k \triangleq (W_{3,1}^n, Y_{3,1}^{k-1}), k=1,2,\dots,n$.
\noindent Constraint \eqref{bnd:outr_V_dst_S1} is a consequence of the following chain of inequalities:
\vspace{-0.2cm}
\begin{align}
& \sum_{k=1}^{n}{I(X_{1,k};Y_{k}, Y_{3,k}| S_{2,k}, X_{2,k}, W_{k}, V_{k})} \nonumber \\
& \qquad \stackrel{(a)}{=} \sum_{k=1}^{n}{\Big[H(Y_{k}, Y_{3,k}|S_{2,k}, X_{2,k}, W_{k}, W_{3,1}^n , Y_{3,1}^{k-1})} \nonumber \\
& \qquad \qquad \qquad - H(Y_{k}, Y_{3,k}|S_{2,k}, X_{1,1}^{k}, X_{2,1}^{k}, X_{3,1}^{k}, W_{k}, W_{3,1}^n, Y^{k-1}, Y_{3,1}^{k-1})\Big] \nonumber \\
& \qquad \stackrel{(b)}{\geq} \sum_{k=1}^{n}{\Big[H(Y_{k}, Y_{3,k}|S_2^n, X_{2,k}, Y^{k-1}, W^n, W_{3,1}^n, Y_{3,1}^{k-1})} \nonumber \\
& \qquad \qquad \qquad - H(Y_{k}, Y_{3,k}|S_1^n, S_2^n, X_{1,1}^{k}, X_{2,1}^{k}, X_{3,1}^{k}, W^n, W_{3,1}^n, Y^{k-1}, Y_{3,1}^{k-1} )\Big] \nonumber
\end{align}
\begin{align}
& \qquad \stackrel{(c)}{\geq} \sum_{k=1}^{n}{\Big[H(Y_{k}, Y_{3,k}| S_2^n, W^n,W_{3,1}^n, Y^{k-1}, Y_{3,1}^{k-1})} - H(Y_{k}, Y_{3,k}|S_1^n, S_2^n, W^n, W_{3,1}^n, Y^{k-1},Y_{3,1}^{k-1})\Big] \nonumber \\
& \qquad \geq H(S_1^n|S_2^n, W^n,W_{3,1}^n ) - H(S_1^n|S_2^n, W^n, W_{3,1}^n, Y^n) \nonumber \\
& \qquad \stackrel{(d)}{\geq} nH(S_1 | S_2, W, W_3) - n\gamma(P_e^{(n)}), \label{eq:VDestSingleChain}
\end{align}
\vspace{-0.25cm}
\noindent where (a) follows from the definition of $V_k$, the fact that $X_{3,1}^k$ is a deterministic function of $(W_{3,1}^n,Y_{3,1}^{k-1})$ and from the memoryless channel assumption, see \eqref{eq:MARCchanDist};
(b) follows from the fact that conditioning reduces entropy and, \cite{Massey:90};
(c) follows from the fact that $X_{2,k}$ is a deterministic function of $S_2^n$, and from the property that conditioning reduces entropy;
(d) follows again from the fact that conditioning reduces entropy, the memoryless sources and side information assumption, and \eqref{eq:FanoDest}--\eqref{eq:generalDestEntropyBasic}.
Following arguments similar to those that led to \eqref{eq:VDestSingleChain} we can also show that:
\vspace{-0.15cm}
\begin{subequations} \label{eq:VDestJointChain}
\begin{align}
H(S_2 | S_1, W, W_3) & \leq \frac{1}{n} \sum_{k=1}^{n}{I(X_{2,k};Y_{k}, Y_{3,k}| S_{1,k}, X_{1,k}, W_{k}, V_{k})} + \gamma(P_e^{(n)}) \label{eq:VDestJointChain_S2} \\
H(S_1,S_2| W, W_3) & \leq \frac{1}{n} \sum_{k=1}^{n}{I(X_{1,k}, X_{2,k};Y_{k}, Y_{3,k}|W_{k}, V_{k})} + \gamma(P_e^{(n)}) \label{eq:VDestJointChain_Sum}.
\end{align}
\end{subequations}
\vspace{-0.1cm}
\noindent Next, we define the time-sharing random variable $Q$ uniformly distributed over $\{ 1,2,\dots,n \}$ and independent of all other random variables. We can write the following:
\vspace{-0.15cm}
\begin{align}
\frac{1}{n} \sum_{k=1}^{n}{I(X_{1,k};Y_{k}, Y_{3,k} | S_{2,k},X_{2,k}, W_{k}, V_{k})}
& = I(X_{1,Q};Y_{Q}, Y_{3,Q}| S_{2,Q}, X_{2,Q}, W_{Q}, V_{Q}, Q) \nonumber \\
& = I(X_{1};Y, Y_{3}|S_{2}, X_{2}, W, V),
\label{eq:VBoundConcavity}
\end{align}
\vspace{-0.2cm}
\noindent where $X_1 \triangleq X_{1,Q}$, $X_2 \triangleq X_{2,Q}$, $Y \triangleq Y_{Q}$, $Y_3 \triangleq Y_{3,Q}$, $S_2 \triangleq S_{2,Q}$, $W \triangleq W_Q$ and $V \triangleq (V_Q, Q)$. Since $(X_{1,k}, X_{2,k})$ and $X_{3,k}$ are independent given $(S_{1,k},S_{2,k},V_k)$, for $\bar{v} = (v,k)$ we have:
\vspace{-0.2cm}
\begin{align}
& \Pr \big( X_1=x_1,X_2=x_2,X_3=x_3|S_1=s_1, S_2=s_2,V=\bar{v} \big) \nonumber \\
& \qquad = \Pr \big( X_1=x_1,X_2=x_2|S_1=s_1, S_2=s_2, V=\bar{v} \big) \Pr \big(X_3=x_3|V=\bar{v} \big).
\label{eq:VouterDist}
\end{align}
\vspace{-0.2cm}
\noindent Hence, the probability distribution is of the form given in \eqref{eq:BCbound_dist}.
Finally, repeating the steps leading to \eqref{eq:VBoundConcavity} for \eqref{eq:VDestJointChain_S2} and \eqref{eq:VDestJointChain_Sum}, and taking the limit $n \rightarrow \infty$, leads to the constraints in \eqref{bnd:V_general_dst}.
\ifthenelse{\boolean{SquizFlag}}{\begin{comment}}{}
\section{Detailed Derivations} \label{annex:extra}
\subsection{Optimizing the sufficient conditions of \Thmref{thm:jointCond_New} and \Thmref{thm:jointCond_NewSimult} for the PSOMARC} \label{subannex:PSOMARC_CondsDeriveThm34}
In this subsection we show that conditions \eqref{bnd:JointNew} as well as conditions \eqref{bnd:JointNewSimult} are optimized by letting $\mathcal{V}_1 = \mathcal{V}_2 = \phi$. Moreover, by recalling that for the PSOMARC $\mathcal{W}_3 = \mathcal{W} = \phi$, and $I(X_3;Y_R)=C_3$, we obtain conditions \eqref{bnd:PrimitiveSOMARC_JointNew}. We begin with \Thmref{thm:jointCond_New}.
\subsubsection{\Thmref{thm:jointCond_New}}
First consider the RHS of condition \eqref{bnd:JointNew_rly_S1} (with $\mathcal{W}_3 = \phi$):
\begin{align}
I(X_1;Y_3|S_2, V_1,X_2,X_3) & \stackrel{(a)}{=} H(Y_3|S_2, V_1,X_2, X_3) - H(Y_3|S_2, X_1,X_2) \nonumber \\
& \stackrel{(b)}{\le} H(Y_3|S_2,X_2) - H(Y_3|S_2,X_1,X_2) \nonumber \\
& = I(X_1;Y_3|S_2,X_2), \label{eq:ExtraThm3RlySingle}
\end{align}
\noindent where (a) follows from from the Markov chain $(V_1, X_3) \leftrightarrow (S_2, X_1,X_2) \leftrightarrow Y_3$ and (b) follows from the fact the conditioning reduces entropy. Similarly, for the RHS of conditions \eqref{bnd:JointNew_rly_S2}--\eqref{bnd:JointNew_rly_S1S2} we have
\begin{subequations} \label{eq:ExtraThm3RlySum}
\begin{align}
I(X_2;Y_3|S_1, V_2,X_1, X_3) & \le I(X_2;Y_3|S_1,X_1) \\
I(X_1,X_2;Y_3|V_1,V_2,X_3) & \le I(X_1,X_2;Y_3).
\end{align}
\end{subequations}
\noindent Next, we consider the RHS of condition \eqref{bnd:JointNew_dst_S1} (with $\mathcal{W} = \phi$):
\begin{align}
& I(X_1;Y_S,Y_R|S_2,V_1,X_2,X_3) + I(V_1,X_3;Y_R|V_2) \nonumber \\
& \qquad = I(X_1;Y_S|S_2,V_1,X_2,X_3) + I(X_1;Y_R|S_2,V_1,X_2,X_3,Y_S) + I(V_1,X_3;Y_R|V_2) \nonumber \\
& \qquad \stackrel{(a)}{=} I(X_1;Y_S|S_2,V_1,X_2,X_3) + I(V_1,X_3;Y_R|V_2) \nonumber \\
& \qquad \stackrel{(b)}{\le} I(X_1;Y_S|S_2,X_2) + I(V_1,X_3;Y_R|V_2) \nonumber \\
& \qquad \stackrel{(c)}{\le} I(X_1;Y_S|S_2,X_2) + I(X_3;Y_R), \label{eq:ExtraThm3DstSingle}
\end{align}
\noindent where (a) follows from the Markov chain $(S_2,V_1,X_1,X_2,Y_S) \leftrightarrow X_3 \leftrightarrow Y_R$; (b) follows from arguments similar to those leading to \eqref{eq:ExtraThm3RlySingle}; and (c) follows from the Markov chain $(V_1,V_2) \leftrightarrow X_3 \leftrightarrow Y_R$ and from the fact that conditioning reduces entropy. Similarly, for the RHS of conditions \eqref{bnd:JointNew_dst_S2}--\eqref{bnd:JointNew_dst_S1S2} we have
\begin{subequations} \label{eq:ExtraThm3DstSum}
\begin{align}
I(X_2;Y_S,Y_R|S_1,V_2,X_1,X_3) + I(V_2,X_3;Y_R|V_1) & \le I(X_2;Y_S|S_1,X_1) + I(X_3;Y_R) \\
I(X_1,X_2;Y_S,Y_R|V_1, V_2, X_3) + I(V_1,V_2,X_3;Y_R) & \le I(X_1,X_2;Y_S) + I(X_3;Y_R).
\end{align}
\end{subequations}
\noindent By combining \eqref{eq:ExtraThm3RlySingle}--\eqref{eq:ExtraThm3DstSum} we get the RHSs of \eqref{bnd:PrimitiveSOMARC_JointNew_rly_S1}--\eqref{bnd:PrimitiveSOMARC_JointNew_rly_S1S2}. Finally, we note that conditions \eqref{bnd:PrimitiveSOMARC_JointNew_rly_S1}--\eqref{bnd:PrimitiveSOMARC_JointNew_rly_S1S2} are sufficient conditions for reliable transmission over PSOMARC. This follows from \cite[Eqns. (12)]{Cover:80} and from the orthogonality of the relay-destination link.
\subsubsection{\Thmref{thm:jointCond_NewSimult}}
First we recall that the relay decoding constraints of \Thmref{thm:jointCond_New} and \Thmref{thm:jointCond_NewSimult} are identical. Therefore, in the following we consider only the destination decoding constraints, that is \eqref{bnd:JointNewSimult_dst_S1}--\eqref{bnd:JointNewSimult_dst_last}. We begin with an upper bound for the first term on the RHS of condition \eqref{bnd:JointNewSimult_dst_S1}:
\begin{align}
& I(X_1,X_3;Y_R,Y_S|S_2,V_2,X_2) \nonumber \\
& \qquad = I(X_1;Y_R,Y_S|S_2,V_2,X_2) + I(X_3;Y_R|S_2,V_2,X_1,X_2) + I(X_3;Y_S|S_2,V_2,X_1,X_2,Y_R) \nonumber \\
& \qquad \stackrel{(a)}{=} I(X_1;Y_S|S_2,V_2,X_2) + I(X_1;Y_R|S_2,V_2,X_2,Y_S) + I(X_3;Y_R|S_2,V_2,X_1,X_2) \nonumber \\
& \qquad = I(X_1;Y_S|S_2,V_2,X_2) + H(Y_R|S_2,V_2,X_2,Y_S) - H(Y_R|S_2,V_2,X_1,X_2,Y_S) \nonumber \\
& \qquad \qquad + H(Y_R|S_2,V_2,X_1,X_2) - H(Y_R|S_2,V_2,X_1,X_2,X_3) \nonumber \\
& \qquad \stackrel{(b)}{=} I(X_1;Y_S|S_2,V_2,X_2) + H(Y_R|S_2,V_2,X_2,Y_S) - H(Y_R|X_3) \nonumber \\
& \qquad \stackrel{(c)}{\le} I(X_1;Y_S|S_2,X_2) + I(X_3;Y_R), \label{eq:ExtraThm4DstSingle_1}
\end{align}
\noindent where (a) follows from the Markov chain $(S_2,V_2,X_3,Y_R) \leftrightarrow (X_1,X_2) \leftrightarrow Y_S$; (b) follows from the Markov chain $Y_S \leftrightarrow (S_2,V_2,X_1,X_2) \leftrightarrow Y_R$ and from the Markov chain $(S_2,V_2,X_1,X_2) \leftrightarrow X_3 \leftrightarrow Y_R$; and (c) follows from the arguments leading to \eqref{eq:ExtraThm3RlySingle} and from the fact the conditioning reduces entropy. Next, we consider the second term on the RHS of condition \eqref{bnd:JointNewSimult_dst_S1}:
\begin{align}
& I(X_1,X_3;Y_R,Y_S|S_1,X_2,V_2) + I(X_1;Y_R,Y_S|S_2,V_1,X_2,X_3) \nonumber \\
& \qquad \stackrel{(a)}{\le} I(X_1,X_3;Y_R,Y_S|S_1,X_2,V_2) + I(X_1;Y_S|S_2,X_2) \nonumber \\
& \qquad \stackrel{(b)}{\le} I(X_1;Y_S|S_1,X_2) + I(X_3;Y_R) + I(X_1;Y_S|S_2,X_2), \label{eq:ExtraThm4DstSingle_2}
\end{align}
\noindent where (a) follows from the arguments leading to \eqref{eq:ExtraThm3DstSingle}; and (b) follows from the arguments leading to \eqref{eq:ExtraThm4DstSingle_1}. In addition, we note that \eqref{eq:ExtraThm4DstSingle_2} is redundant due to \eqref{eq:ExtraThm4DstSingle_1}. Similarly, for the RHS of \eqref{bnd:JointNewSimult_dst_S2}--\eqref{bnd:JointNewSimult_dst_S1S2} we have
\begin{subequations}
\begin{align}
& \min \Big\{I(X_2,X_3;Y_R,Y_S|S_1,V_1,X_1), \nonumber \\
& \qquad I(X_2,X_3;Y_R,Y_S|S_2,X_1,V_1) + I(X_2;Y_R,Y_S|S_1,V_2,X_1,X_3) \Big\} \le I(X_2;Y_S|S_1,X_1) + I(X_3;Y_R),
\end{align}
and
\begin{align}
I(X_1,X_2,X_3;Y) \le I(X_1,X_2;Y_S) + I(X_3;Y_R).
\end{align}
\end{subequations}
\noindent Finally, we consider the RHS of \eqref{bnd:JointNewSimult_dst_last}, and in particular the first term:
\begin{align}
& I(X_1;Y_R,Y_S|S_2,V_1,X_2,X_3) + I(X_1,X_2,X_3;Y_R,Y_S|S_1) \nonumber \\
& \qquad \stackrel{(a)}{\le} I(X_1;Y_S|S_2,X_2) + I(X_1,X_2,X_3;Y_R,Y_S|S_1) \nonumber \\
& \qquad = I(X_1;Y_S|S_2,X_2) + I(X_1,X_2;Y_S|S_1) + I(X_1,X_2;Y_R|S_1,Y_S) \nonumber \\
& \qquad \qquad + I(X_3;Y_R|S_1,X_1,X_2) + I(X_3;Y_S|S_1,X_1,X_2,Y_R) \nonumber \\
& \qquad \stackrel{(b)}{=} I(X_1;Y_S|S_2,X_2) + I(X_1,X_2;Y_S|S_1) + I(X_1,X_2;Y_R|S_1,Y_S) + I(X_3;Y_R|S_1,X_1,X_2) \nonumber \\
& \qquad \stackrel{(c)}{\le} I(X_1;Y_S|S_2,X_2) + I(X_1,X_2;Y_S|S_1) + I(X_3;Y_R) \nonumber \\
& \qquad = I(X_1;Y_S|S_2,X_2)+ I(X_2;Y_S|S_1,X_1) + I(X_1;Y_S|S_1) + I(X_3;Y_R), \label{eq:ExtraThm4Dstsum_2}
\end{align}
\noindent where (a) follows from the arguments leading to \eqref{eq:ExtraThm3DstSingle}; (b) follows from the Markov chain $X_3 \leftrightarrow (S_1,X_1,X_2,Y_R) \leftrightarrow Y_S $; and (c) follows from arguments similar to those leading to \eqref{eq:ExtraThm4DstSingle_1}. Similarly, for the second term on the RHS of \eqref{bnd:JointNewSimult_dst_last} we have:
\begin{align}
& I(X_1;Y_R,Y_S|S_2,V_1,X_2,X_3) + I(X_1,X_2,X_3;Y_R,Y_S|S_1) \nonumber \\
& \qquad \le I(X_1;Y_S|S_2,X_2)+ I(X_2;Y_S|S_1,X_1) + I(X_2;Y_S|S_2) + I(X_3;Y_R). \label{eq:ExtraThm4DstSum_2}
\end{align}
By combining \eqref{eq:ExtraThm4DstSingle_1}--\eqref{eq:ExtraThm4DstSum_2} we have the RHSs of \eqref{bnd:PrimitiveSOMARC_JointNew}. Finally, we note that conditions \eqref{bnd:PrimitiveSOMARC_JointNew} are sufficient conditions for reliable transmission over PSOMARC. This follows from \cite[Eqns. (12)]{Cover:80} and from the orthogonality of the relay-destination link.
\subsection{The LHSs and RHSs of \eqref{bnd:PrimitiveSOMARC_JointNew} are equal} \label{subannex:extraEquality34Derive}
First, we show that the LHSs and RHSs of conditions \eqref{bnd:PrimitiveSOMARC_JointNew_rly_S1}--\eqref{bnd:PrimitiveSOMARC_JointNew_rly_S1S2} are the same for $p(x_i|s_i)=\delta(x_i-s_i)$, and then we show that condition \eqref{bnd:PrimitiveSOMARC_JointNew_rly_S1_plus_S2} is redundant.
For the sources defined in Table \ref{tab:SourceDist} we have that $H(S_1|S_2)=H(S_2|S_1) = 1$ bit, and $H(S_1,S_2) = \log_2 6$ bits. Next, consider the RHS of \eqref{bnd:PrimitiveSOMARC_JointNew_rly_S1}, evaluated using $p(x_i|s_i)=\delta(x_i-s_i)$:
\begin{align}
\min \{I(X_1;Y_3|S_2, X_2), I(X_1;Y_S|S_2, X_2) + C_3 \} & = \min \{I(S_1;Y_3|S_2, S_2), I(S_1;Y_S|S_2, S_2) + C_3 \} \nonumber \\
& = \min \{I(S_1;Y_3|S_2), I(S_1;Y_S|S_2) + C_3 \} \nonumber \\
& \stackrel{(a)}{=} \min \{H(Y_3|S_2), H(Y_S|S_2) + C_3 \} \nonumber \\
& \stackrel{(b)}{=} \min \{1, 2 \} = 1. \label{eq:ExtraEqualityDeriveIndividual}
\end{align}
\noindent where (a) follows from the fact that $Y_3$ and $Y$ are deterministic functions of $(X_1, X_2)$ (and therefore, in the case of $p(x_i|s_i)=\delta(x_i-s_i)$, deterministic functions of $(S_1, S_2)$); and (b) follows from direct calculation of $H(Y_3|S_2), H(Y|S_2)$ and from the fact that $C_3 = 1$. Following similar arguments we can also show that the RHS of \eqref{bnd:PrimitiveSOMARC_JointNew_rly_S2} equals 1 bit. Next consider the RHS of \eqref{bnd:PrimitiveSOMARC_JointNew_rly_S1S2}, evaluated using $p(x_i|s_i)=\delta(x_i-s_i)$. Using the same arguments that led to \eqref{eq:ExtraEqualityDeriveIndividual} we can show that
\begin{align}
\min \{I(X_1, X_2;Y_3), I(X_1, X_2;Y_S) + C_3 \} & = \min \{H(Y_3), H(Y_S) + C_3 \} \nonumber \\
& = \min \{\log_2 6, \log_2 3 + 1 \} = \log_2 6.
\end{align}
Finally, consider the RHS of condition \eqref{bnd:PrimitiveSOMARC_JointNew_rly_S1_plus_S2}, evaluated using $p(x_i|s_i)=\delta(x_i-s_i)$:
\begin{align}
& I(X_1;Y_S|S_2, X_2) + I(X_2;Y_S|S_1, X_1) + C_3 + \min \big\{ I(X_1;Y_S|S_1), I(X_2;Y_S|S_2) \big\} \nonumber \\
& \qquad = H(Y_S|S_2) + H(Y_S|S_1) + C_3 + \min \big\{ H(Y_S|S_1) - H(Y_S|S_1,S_1), H(Y_S|S_2) - H(Y_S|S_2,S_2) \big\} \nonumber \\
& \qquad = 1 + 1 + 1 + 0 = 3.
\end{align}
\noindent On the other hand, From \eqref{eq:ExtraEqualityDeriveIndividual} it follows that $H(S_1|S_2)+H(S_2|S_1) = 2$ bits. Therefore, we conclude that condition \eqref{bnd:PrimitiveSOMARC_JointNew_rly_S1_plus_S2}, when evaluated using $p(x_i|s_i)=\delta(x_i-s_i)$, is redundant.
\subsection{Optimizing the sufficient conditions of \Thmref{thm:jointCond} for the PSOMARC} \label{subannex:PSOMARC_CondsDeriveThm1}
In this subsection we show that conditions \eqref{bnd:Joint} are optimized by letting $\mathcal{V}_1 = \mathcal{V}_2 = \phi$. Moreover, by recalling that for the PSOMARC $\mathcal{W}_3 = \mathcal{W} = \phi$, and $I(X_3;Y_R)=C_3$, we obtain conditions \eqref{bnd:Thm7}.
First we recall that the relay decoding constraints of \Thmref{thm:jointCond_New} and \Thmref{thm:jointCond} are identical. Therefore, in the following we consider only the destination decoding constraints, that is \eqref{bnd:Joint_dst_S1}--\eqref{bnd:Joint_dst_S1S2}. Following arguments similar to those leading \eqref{eq:ExtraThm4DstSingle_2} and \eqref{eq:ExtraThm4Dstsum_2} we have that
\begin{subequations}
\begin{align}
I(X_1,X_3;Y|S_1, V_2, X_2) & \le I(X_1;Y_S|S_1,X_2) + I(X_3;Y_R) \\
I(X_2,X_3;Y|S_2,V _1, X_1) & \le I(X_2;Y_S|S_2,X_1) + I(X_3;Y_R) \\
I(X_1,X_2,X_3;Y|S_1,S_2) & \le I(X_1,X_2;Y_S|S_1,S_2) + I(X_3;Y_R).
\end{align}
\end{subequations}
\noindent Therefore we obtained the RHSs of \eqref{bnd:Thm7}. Finally, we note that conditions \eqref{bnd:Thm7} are sufficient conditions for reliable transmission over PSOMARC. This follows from \cite[Eqns. (12)]{Cover:80} and from the orthogonality of the relay-destination link.
\subsection{Optimizing the sufficient conditions of \Thmref{thm:jointCondFlip} for the PSOMARC} \label{subannex:PSOMARC_CondsDeriveThm2}
In this subsection we show that conditions \eqref{bnd:JointFlip} are optimized by letting $\mathcal{V}_1 = \mathcal{V}_2 = \phi$. Moreover, by recalling that for the PSOMARC $\mathcal{W}_3 = \mathcal{W} = \phi$, and $I(X_3;Y_R)=C_3$, we obtain conditions \eqref{bnd:Thm8}. We begin with the RHS of \eqref{bnd:JointFlip_rly_S1}:
\begin{align}
I(X_1;Y_3|S_1,X_2,X_3) & = H(Y_3|S_1,X_2,X_3) - H(Y_3|S_1,X_1,X_2,X_3) \nonumber \\
& \stackrel{(a)}{\le} I(X_1;Y_3|S_1,X_2),
\end{align}
\noindent where (a) follows from the Markov chain $X_3 \leftrightarrow (S_1,X_1,X_2) \leftrightarrow Y_3$, and from the fact that conditioning reduces entropy. Following similar arguments we have:
\begin{subequations}
\begin{align}
I(X_2;Y_3|S_2, X_1, X_3) & \le I(X_2;Y_3|S_2, X_1) \\
I(X_1,X_2;Y_3|S_1, S_2, X_3 ) & \le I(X_1,X_2;Y_3|S_1, S_2).
\end{align}
\end{subequations}
\noindent Next, we consider the RHS of \eqref{bnd:JointFlip_dst_S1}. Following arguments similar to those leading to \eqref{eq:ExtraThm4DstSingle_1} we have:
\begin{subequations}
\begin{align}
I(X_1,X_3;Y_R,Y_S|S_2, X_2) & \le I(X_1;Y_S|S_2,X_2) + I(X_3;Y_R) \\
I(X_2,X_3;Y_R,Y_S|S_1, X_1) & \le I(X_2;Y_S|S_1,X_1) + I(X_3;Y_R) \\
I(X_1,X_2,X_3;Y) & \le I(X_1,X_2;Y_S) + I(X_3;Y_R).
\end{align}
\end{subequations}
\noindent Therefore we obtained the RHSs of \eqref{bnd:Thm8}. Finally, we note that conditions \eqref{bnd:Thm8} are sufficient conditions for reliable transmission over PSOMARC. This follows from \cite[Eqns. (12)]{Cover:80} and from the orthogonality of the relay-destination link.
\ifthenelse{\boolean{SquizFlag}}{\end{comment}}{}
\vspace{-0.3cm}
|
\section{INTRODUCTION}
\label{sec:intro}
The Atacama Large Millimeter/submillimeter Array (ALMA)
is a major new astronomical observatory. It started operation in 2011 and
was officially inaugurated in 2013. ALMA consists of an array of fifty 12-m antennas
and an additional compact array (ACA) of twelve 7-m and four 12-m antennas to enhance
ALMA's ability to image extended targets. Among the 12-m antennas of the interferometric array,
baseline lengths up to 16~km will be achieved.
The ALMA project is an international collaboration between Europe, East Asia, and
North America in cooperation with the Republic of Chile.
The official project website for scientists is the {\it ALMA Science Portal}
{\tt http://www.almascience.org}. A more detailed description can be found
in the proceedings of a parallel session at this conference \cite{cox} and references therein.
ALMA has been performing ``Early Science'' observations with a subset of its antennas
since 2011. The many capabilities are gradually commissioned.
In the first two observing proposal cycles (Cycles 0 and 1), ALMA offered
receivers that covered four separate atmospheric windows
(bands 3, 6, 7, and 9)
in the range between 84~GHz to 720~GHz (3~mm to 0.42~mm). In subsequent cycles,
more receiver bands will become available.
Cycle 0 observations resulted in 78 refereed publications so far.
Cycle 1 has ended on 31 May 2014, i.e. shortly before the beginning of this conference.
Its science impact cannot fully be measured, yet.
In Cycle 1, typically 30 of the 50 antennas of the 12-m array were used for science observations.
From the ACA, typically 8 antennas were available. The last of
the 66 antennas was brought up to the observatory site in March 2014 and test observations with
more than 50 antennas have already taken place.
When the full set of antennas is in operation, the average daily science data volume
generated by ALMA is expected to be roughly one TByte.
All ALMA data is stored in the {\it ALMA Science Data Model} (ASDM) format
which in its present implementation is a collection of XML and binary MIME
encoded files.
When a team of scientists (led by the principle investigator, the ``PI'') proposes an
observation with ALMA, they are not applying for a certain amount of observation time
but they are asking for the achievement of a {\it Science Goal}.
This is a sensitivity and observation setup requirement which the observatory pledges to meet
if the proposal is accepted.
The observations belonging to a certain proposal are regarded to be completed when
the Science Goal(s) defined in the proposal are achieved. In order to confirm that
achievement, a full calibration and at least partial imaging of the science data is necessary.
This analysis is carried out as part of the ALMA {\it Quality Assurance} (QA).
The ALMA operations are based on {\it service observing}.
This means firstly that the scientists who proposed the observations
and who will have the proprietary rights on the data for the first year after they were taken,
are not required (nor permitted) to be present during the observation.
Instead, the observations are carried out using dynamic scheduling, i.e. matching
the requirements set in the proposal with the present
status of the observatory, the atmospheric conditions, and the local siderial time.
It is furthermore the ALMA project's ambition to make the observatory available to all
astronomers, not only sub-mm radio astronomy experts.
The observatory therefore provides {\it service data analysis}. As part of the QA work,
personnel at the Joint ALMA Observatory in Chile (JAO) and the three ALMA Regional Centers (ARCs)
perform detailed calibration and imaging on all observations. The PI obtains the resulting
calibration and imaging scripts and imaging products \cite{qa2products}.
In the simplest case, if the PI is content with the imaging products, he/she can
proceed to a publication based on the obtained imaging products
without worrying about learning the use of the ALMA data analysis software.
For more complex observing projects, the PI may need to put more work
into at least the imaging, sometimes also the calibration. In any case, the standard ALMA imaging
products and calibration tables will be archived and made available to the public (together with
the raw data and all analysis scripts) as soon as the PI's proprietary time ends.
The main software tool in the processing of ALMA data is CASA (``Common Astronomy Software Applications''),
a package which is developed together by all ALMA partners under NRAO management.
This article summarises the design of the ALMA service data analysis
and QA infrastructure and describes how CASA is employed in this context.
Detailed descriptions of many of the processes mentioned here can be found
in the ALMA Technical Handbook \cite{handbook2} which is updated for every observing cycle.
Also, there are several other related articles to be found in the different
proceedings volumes of this conference. See in particular Ref.\citenum{schnee2014}.
\section{The ALMA Pipelines}
To achieve high data quality, ALMA employs three data analysis pipelines at different stages
of the observing process:
\begin{enumerate}
\item The Telescope Calibration pipeline (TelCal)
\item The Quicklook Display pipeline
\item The offline data analysis (including the Science Pipeline, simply called ``The Pipeline'')
\end{enumerate}
The first two of these are running during and immediately after the observations (online).
They produce measurements of the characteristic parameters of the observatory
for the {\it Observatory State and Calibration Database} and permit
the Astronomer on Duty (AoD) to monitor the data quality.
Due to the real-time processing constrains and the high data volume, TelCal and Quicklook in many cases
cannot work on the data at full spectral resolution nor perform detailed high-resolution imaging.
Instead they mostly work on averaged data.
The full performance of ALMA is achieved in the offline analysis which is (trivially) parallelised in several
ways and therefore able to handle the instrument's native temporal, spectral, and spatial resolution.
This offline analysis is based on the CASA software package which is described in the next section.
\section{CASA technical summary}
The offline data analysis needs a software package
capable of handling all features of the ALMA data including the high spectral resolution and sensitivity
and the large data volume. In 2003, the {\it CASA} package was selected
by the project as the ALMA data reduction software.
The structure of CASA was already described in Refs. \citenum{mcmullin_2007} and \citenum{petry_2012}. It has not changed significantly since 2012. So only
a brief summary is given here. See also Ref. \citenum{ott_2014}.
CASA (Common Astronomy Software Applications) is a general package for all
radio-astronomical data analysis. The development team consists of scientists
and software engineers from NRAO, ESO, and NAOJ. The main focus is on the
support for the ALMA and the Carl Jansky VLA observatories.
The CASA management, the project scientist, and the subsystem scientists from the ALMA
and JVLA side are all based at NRAO. At ESO and NAOJ so-called cognizant leads
accompany the development for the ALMA side.
Documentation is available from the CASA homepage {\tt http://casa.nrao.edu}.
The lower-level CASA functionality is implemented using C++.
The fundamental libraries of CASA are collected in a sub-package named {\tt casacore}
which extends the C++ Standard Template Library
providing methods for the handling of files, physical and astronomical
quantities, coordinate systems and reference frames, advanced mathematical
operations, and more. {\tt casacore} is developed in collaboration with ASTRON
and ATNF and is documented on googlecode at {\tt http://code.google.com/p/casacore/}.
The CASA team presently still keeps its own copy of a {\tt casacore} repository
with regular mergers to the googlecode version. It is planned to end this
duplication soon and converge to a single version.
{\tt casacore} with its rich and well-tested set of basic C++ classes is
useful for any astronomical software application, not only for radio astronomy.
Based on {\tt casacore}, CASA implements in a separate sub-package the functionality
needed specifically for radio astronomy. This includes
calibration algorithms based on the Measurement Equation \cite{hamakeretal_1996},
interferometric imaging algorithms, a powerful viewer tool to visualise imaging results,
image analysis algorithms, and a simulator for radio interferometers.
In addition, a complete set of algorithms for the analysis of single-dish (non-interferometric)
radio data is provided by including the ASAP package \cite{asap} with special extensions.
In a third sub-package, the user interface and high-level analysis functionality
is implemented by binding the functionality implemented in C++ to the {\it Python} scripting language
and adding a layer of high-level ``tasks'' which are Python scripts implementing common
procedures used in interferometric and single-dish radio data analysis.
Finally, the employment of the {\it iPython} package creates the command line user
interface which gives CASA the look and feel which users of modern data analysis packages like, e.g., MatLab or IDL
are used to.
The supported computer platforms of CASA in its latest version 4.2.2 are the common Linux
distributions (64 bit only) and Mac OSX 10.7 and 10.8. The binary release and the source code are
available under GNU Public License (see {\tt http://casa.nrao.edu}).
\section{ALMA Quality Assurance}
The goal of ALMA Quality Assurance (QA) is to deliver to the PI a
reliable final data product that has reached the desired control
parameters outlined in the science goals and that is calibrated to the
desired accuracy and free of calibration or imaging artifacts.
As mentioned in the introduction, ALMA follows a paradigm of ``Science-goal-oriented service data analysis''.
This means the PI defines science goals in the proposal aided by a specially developed ALMA software
package, the Observing Tool (OT).
This formalised proposal defines so-called ``Scheduling Blocks'' (SBs).
An SB is the prototype of an atomic (ca. 0.5 h - 1 h) observation to reach a specific science goal.
SBs are stored in a special XML format which can be directly understood by the observation scheduler.
An execution of an SB, i.e. the actual observation, is called an ``Exec Block'' (EB).
By executing one SB several times, many EBs can be produced and hence sensitivity accumulated.
SBs which depend on each other because their data analysis has to be carried out together,
form an ``Observation Unit Set'' (OUS). The way ALMA projects are set up at the moment,
typically every OUS only contains one SB.
The number of required EBs to reach the science goal of the OUS is estimated by the OT in
an analytic calculation by the so-called sensitivity calculator which takes into account detailed
models of the observatory, the atmosphere, and the calibration sources.
It is the task of the QA to verify that each EB was properly executed and that
the OUS (after the nominal number of good EBs is reached) does indeed achieve the science goal.
ALMA QA consists of 3 (+1) steps (see also figure \ref{fig:almadataflow}):
\begin{figure}
\begin{center}
\begin{tabular}{c}
\includegraphics[width=13.5cm]{almadataflow.eps}
\end{tabular}
\end{center}
\caption[almadataflow]
{ \label{fig:almadataflow}
Schematic overview of the ALMA data flow indicating the various QA stages.
For simplicity, the detailed data flow to and from Quicklook, QA0, and QA1
has been omitted. All these stages access the archive.
}
\end{figure}
\begin{description}
\item[QA0:] Immediate checks of data quality at the time of the observation or shortly after:
Atmosphere, Antennas, Front-Ends, Connectivity, Back-Ends
\item[QA1:] Monitor slowly varying array performance parameters (Arrays, Antennas, Calibration Sources)
and also rapid changes which affect the entire array
\item[QA2:] The completion of an OUS triggers QA2 to confirm that the Science Goal was met.
If not, additional EBs will need to be obtained. QA2 implies full calibration and
generation of official science products. After QA2 is complete (Science Goal met),
the data is delivered to the PI.
\item[QA3:] If there are problems found with the data after the delivery, the PI or the ALMA contact
scientist of the PI can file a problem report via the ALMA helpdesk which triggers QA3.
The problem is examined and a solution searched for. Re-reduction of the data may be performed,
possibly replacement of products in the archive.
\end{description}
In the following we concentrate on the QA2 step.
For every OUS which has reached the number of planned executions passing QA0 and QA1,
QA2 proceeds to perform a full calibration.
The resulting CASA calibration tables and scripts for the OUS are archived.
Furthermore, imaging is performed (a) to the point where it can be decided whether
each Science Goal was met and (b) to obtain a standard set of science products
for archiving, i.e. for later archival research (see Ref. \citenum{qa2products}).
Ultimately, QA2 on all data from standard observing is supposed to be performed
by the fully automated Pipeline as mentioned above. The Pipeline is still under
development and a first version of it is ready to be put in service at the beginning
of Cycle 2. Until the Pipeline can process data from all possible observing modes,
two analysis paths will coexist in QA2:
\begin{enumerate}
\item Semi-automatic processing (calibration and imaging) using the so-called Script Generator
\item Automated calibration with the Pipeline followed by semi-automatic imaging
\end{enumerate}
\subsection{ALMA QA2 with the Script Generator}
Before a fully automated data analysis pipeline can be commissioned, the manual data analysis
has to be understood. The necessary expertise to design the data analysis procedures
grows as the operators gain experience with the instrument. On the other hand, as a new
observatory starts operation, resources for software development are stretched thin and,
what is more, there is strong pressure to produce results quickly and open the observatory
to the community for actual science observations.
A naive, linear approach which waits for the completion of the observatory construction
and commissioning before detailed work on data analysis automation is started,
leads to intolerable delays in the development of the analysis pipeline.
Already the task of processing test observations is overwhelming if not aided by sufficient
software tools. On the other hand, if the development is started too early,
the lack of experience with the real system can lead to wasted efforts and the need to
redesign.
The solution is to {\it gradually} automate the data analysis as new observing modes become
available during the completion of the observatory commissioning.
The gradual automation is achieved by having a basic toolkit for analysis
from which a prototype pipeline can be built for each observing mode.
Once the analysts agree that the prototype meets all requirements, it can be
transferred into the final, fully automated Pipeline. In this way the Pipeline
acquires more and more capabilities and can take over more and more of
the QA2 work.
For ALMA, the role of the basic analysis toolkit is played by CASA with its
scripting language Python.
A large team of ALMA scientists worked together to develop the initial
best practices to perform a robust standard calibration of ALMA Cycle 0 data.
Based on these, the pipeline prototyping could simply have been performed
by developing separate CASA scripts
for each observing mode. However, the number of possible combinations of
setups was regarded as too large, such that an automated assistant to the analyst
was thought to be necessary.
This additional software tool is called the Script Generator.
It was first devised and implemented by Eric Villard (JAO) and is still
maintained by him.
\begin{figure}
\begin{center}
\begin{tabular}{c}
\includegraphics[width=14cm]{qa2-w-scriptgen.eps}
\end{tabular}
\end{center}
\caption[qa2-w-scriptgen]
{ \label{fig:qa2-w-scriptgen}
Illustration of the script generator principle. With the help of the Script Generator,
the data analyst generates first a draft analysis script. The Script Generator
parses the data to be analysed and implements best practices and previous
analysis experience. The analyst then manually finalises the script adapting
it (if necessary) to unforeseen features of the data. }
\end{figure}
The Script Generator (see figure \ref{fig:qa2-w-scriptgen}) is itself a CASA script.
It takes as only input the dataset to be analysed.
Based on the properties of the dataset, it selects the appropriate analysis path
and creates a new CASA draft analysis script specific for the dataset.
This draft script can in some cases already be final.
In other cases, the analyst will have to make small modifications to the script
until it accommodates all special features of the dataset.
These concern mostly shortcomings of the data caused by hardware problems.
In addition to providing analysis script drafts for the calibration,
the present ALMA Script Generator also provides scripts (a) for combining
several EBs into one final dataset ready for imaging and (b) for the
imaging itself.
The script generator approach has proven to be highly effective.
The analysts save time on repetitive tasks but keep full control as each
step of the analysis is readily visible and can quickly be modified.
The analysis does not become a black box which is blindly trusted.
Solutions to newly encountered problems and improvements can easily be fed back.
The mature Script Generator helps new team members to learn quickly
while saving time also for very experienced analysts who are enabled
to concentrate on the new and difficult aspects of the data.
As confidence increases, the automation can be increased step by step.
\subsection{ALMA QA2 with the Pipeline}
Also the fully automated ALMA Pipeline is based on CASA and is planned
to eventually become part of it as a set of additional, ALMA-specific
high-level functionality ``tasks''.
It is meant to ultimately process both interferometry and single-dish data
in a fully automated, data-driven way including the imaging.
While it is mainly meant to be run in a high-performance computing environment,
the present design foresees that any CASA user will be able to run it
on a standard, reasonably powerful workstation with sufficient disk space
to hold the raw data (typically up to 1 TB for a simple project) and
the intermediate and final data products (up to 3 TB).
As described in the previous section, the Pipeline development is fed
by the experience gained in manual and Script-Generator-assisted analysis.
This experience is encoded in so-called ``heuristics''. The Pipeline
(and also the Script Generator) can be regarded as an expert system
for ALMA data analysis.
Pipeline commissioning and verification is performed by comparing the
results with those obtained from the Script-generator-assisted analysis.
The first actual use of the Pipeline for QA2 will take place in
ALMA Cycle 2 in the second half of 2014. It will be employed to perform
the calibration for datasets from well understood observing modes.
Imaging will initially still be done only with Script Generator assistance
and (like all other analysis) gradually be moved over into the Pipeline.
\subsection{ALMA QA2 statistics in Cycle 1}
At the time of writing, ALMA Cycle 1 is nearly complete.
The QA2 effort was significant.
So far, the (at any time) ca. 40 QA2 analysts at the Joint ALMA Observatory (JAO) and the
three ALMA Regional Centres (ARCs) in East Asia, Europe, and North America
have processed more than 980 h of observations in Cycle 1 since January 2013
and delivered more than 188 Observing Unit Sets to their PIs.
The typical (median) analysis time in Script-Generator-assisted QA2 is 3 to 4 working days
per Execution Block. This includes data transfer and packaging.
Only very few cases of QA3 have occurred so far.
The observatory has received enthusiastic feedback from the PIs concerning data quality.
\section{SUMMARY}
ALMA is working well but commissioning will still have to continue
interleaved with science observations for a number of cycles until all
initially planned capabilities are available on the full set of antennas.
The service data analysis model has proven to be viable and flexible
enough to cope with the varying conditions.
PIs are excited about the achieved data quality.
The overall QA2 system is running better and better but will remain labour-intensive
until the fully automated Pipeline has implemented heuristics for at least most of the
planned observing modes.
The work spent on QA2, however, is not wasted but creates the necessary
knowledge and expertise on ALMA data analysis in a broad base of
the ALMA personnel.
|
\section{Introduction}
Many charmonium-like resonances have been discovered in the past, revealing a spectrum too rich to interpret in terms of conventional mesons expected from potential models~\cite{NRPM}. In several cases, it has not been possible to assign a spin-parity value to the resonance. Some of them have been extensively investigated as possible candidates for non-conventional mesons, such as tetraquarks, glueballs, or hybrids~\cite{XYZrev}.
In a search for exotic states, the CDF experiment studied the decay $B^+ \rightarrow J/\psi \phi K^+$~\cite{conj}, where $J/\psi \rightarrow \mu^+ \mu^-$ and $\phi(1020) \rightarrow K^+ K^-$, claiming the observation of a resonance labeled the $X(4140)$ decaying to $J/\psi \phi$~\cite{kai}. They found evidence in the same decay mode for another resonance, labeled as the $X(4270)$~\cite{kai-bis}. Recently, the LHCb experiment studied the decay $B^+ \rightarrow J/\psi \phi K^+$ in $pp$ collisions at 7~TeV, with a data sample more than three times larger than that of CDF, and set an upper limit (UL) incompatible with the CDF result~\cite{LHCb}. The D0 and the CMS experiments more recently made studies of the same decay channel, leading to different conclusions~\cite{cms, d0} than the LHCb experiment. In this work we study the rare decays $B^{+} \rightarrow J/\psi K^+K^- K^{+}$, $B^{0} \rightarrow J/\psi K^+K^- K_S^0$ and search for possible resonant states in the $J/\psi \phi$ mass spectrum. We also search for the decay $B^0 \rightarrow J/\psi \phi$, which is expected to proceed mainly via a Cabibbo-suppressed and color-suppressed transition $\bar b d \rightarrow \bar c c \bar d d$. The absence of a signal would indicate that the required rescattering of $\bar d d$ into $\bar s s$ is very small.
This paper is organized as follows. In Sec. II we describe the detector and data selection and in Sec. III we report the branching-fraction (BF) measurements. Section IV is devoted to the resonance search, while Sec. V summarizes the results.
\section{The \mbox{\slshape B\kern -0.1em{\smaller A}\kern-0.1emB\kern-0.1em{\smaller A\kern -0.2em R }} detector and data selection}
We make use of the data set collected by the \mbox{\slshape B\kern -0.1em{\smaller A}\kern-0.1emB\kern-0.1em{\smaller A\kern -0.2em R }}\ detector at the PEP-II $e^{+}e^{-}$ storage rings operating at the $\Y4S$ resonance. The integrated luminosity for this analysis is 422.5 fb$^{-1}$, which corresponds to the production of 469 million $B\overline{B}$ pairs~\cite{lumi}.
The \mbox{\slshape B\kern -0.1em{\smaller A}\kern-0.1emB\kern-0.1em{\smaller A\kern -0.2em R }}\ detector is described in detail elsewhere~\cite{luminew}. We mention here only the components of the detector that are used in the present analysis. Charged particles are detected and their momenta measured with a combination of a cylindrical drift chamber (DCH) and a silicon vertex tracker (SVT), both operating within the 1.5 T magnetic field of a superconducting solenoid. Information from a ring-imaging Cherenkov detector (DIRC) is combined with specific ionization measurements from the SVT and DCH to identify charged kaon and pion candidates. The efficiency for kaon identification is 90\% while the rate for a pion being misidentified as a kaon is 2\%. For low transverse momentum kaon candidates that do not reach the DIRC, particle identification relies only on the energy loss measurement, so that the transverse momentum spectrum of identified kaons extends down to 150 \mbox{MeV/$c$}\xspace. Electrons are identified using information provided by a CsI(Tl) electromagnetic calorimeter (EMC), in combination with that from the SVT and DCH, while muons are identified in the Instrumented Flux Return (IFR). This is the outermost subdetector, in which muon/pion discrimination is performed. Photons are detected, and their energies measured with the EMC.
For each signal event candidate, we first reconstruct the \ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace by geometrically constraining to a common vertex a pair of oppositely charged tracks, identified as either electrons or muons, and apply a loose requirement that the $\chi^2$ fit probability exceed 0.1\%. For $\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace \rightarrow \ensuremath{e^+e^-}\xspace$ we use bremsstrahlung energy-loss recovery: if an electron-associated photon cluster is found in the EMC, its three-momentum vector is incorporated into the calculation of the invariant mass $m_{e^{+}e^{-}}$. The vertex fit for a \ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace candidate includes a constraint to the nominal \ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace mass value~\cite{PDG}.
For $B^+ \rightarrow J/\psi K^+ K^- K^+$ candidates, we combine the $J/\psi$ candidate with three loosely identified kaons and require a vertex-fit probability larger than 0.1\%. Similarly, for $B^0 \rightarrow J/\psi K^- K^+ \ensuremath{K^0_{\scriptscriptstyle S}}\xspace$ candidates, we combine the $\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace$ and $\ensuremath{K^0_{\scriptscriptstyle S}}\xspace$ with two loosely identified kaons and require a vertex-fit probability larger than 0.1\%.
A $\ensuremath{K^0_{\scriptscriptstyle S}}\xspace$ candidate is formed by geometrically constraining a pair of oppositely charged tracks to a common vertex, with $\chi^2$ fit probability larger than 0.1\%. The pion mass is assigned to the tracks without particle-identification (PID) requirements. The three-momenta of the two pions are then added and the \ensuremath{K^0_{\scriptscriptstyle S}}\xspace energy is computed using the nominal \ensuremath{K^0_{\scriptscriptstyle S}}\xspace mass. We require the \ensuremath{K^0_{\scriptscriptstyle S}}\xspace flight length significance with respect to the $B^0$ vertex to be larger than 3$\sigma$.
We further select $B$ meson candidates using the energy difference \mbox{$\Delta E$}\xspace $\equiv E^{\ast}_B-\sqrt{s}/2$ in the center-of-mass frame and the beam-energy-substituted mass defined as $\mbox{$m_{\rm ES}$}\xspace \equiv \sqrt{((s/2+\vec{p}_i\cdot\vec{p}_B)/E_i)^2-\vec{p}_B^{\,2}}$, where ($E_i,\vec{p}_i$) is the initial state \ensuremath{e^+e^-}\xspace four-momentum vector in the laboratory frame and $\sqrt{s}$ is the center-of-mass energy. In the above expressions $E^{\ast}_B$ is the $B$ meson candidate energy in the center-of-mass frame, and $\vec{p}_B$ is its laboratory frame momentum.
When multiple candidates are present, the combination with the smallest \mbox{$\Delta E$}\xspace is chosen. We find that, after requiring \mbox{$m_{\rm ES}$}\xspace$>5.2$ \mbox{GeV/$c^{\rm 2}$}\xspace, the fraction of events having multiple candidates is 1.3\% for $B^+$ and 8.6\% for $B^0$. From simulation, we find that 99.6$\%$ of the time we choose the correct candidate.
The final selection requires $\left | \mbox{$\Delta E$}\xspace \right|<30$ \mbox{MeV}\xspace and $\left | \mbox{$\Delta E$}\xspace\right|<25$ \mbox{MeV}\xspace for $B^+$ and $B^0$ decays, respectively; the additional selection criterion \mbox{$m_{\rm ES}$}\xspace$>5.2$ \mbox{GeV/$c^{\rm 2}$}\xspace is required for the calculation of the BFs, while \mbox{$m_{\rm ES}$}\xspace$>5.27$ \mbox{GeV/$c^{\rm 2}$}\xspace is applied to select the signal region for the analysis of the invariant mass systems.
\section{Branching Fractions}
\begin{figure*}[ht]
\begin{center}
\mbox{
\scalebox{0.28}{\includegraphics{./mes.eps}}
\scalebox{0.28}{\includegraphics{./Fig2v24.eps}}}
\mbox{
\scalebox{0.28}{\includegraphics{./de2.eps}}
\scalebox{0.28}{\includegraphics{./Fig4v25.eps}}}
\caption{\label{fig1} The \mbox{$m_{\rm ES}$}\xspace distributions for (a) $B^+ \rightarrow J/\psi K^+ K^- K^+$ and (b) $B^0 \rightarrow J/\psi K^- K^+ \ensuremath{K^0_{\scriptscriptstyle S}}\xspace$, for the \mbox{$\Delta E$}\xspace regions indicated in the text. The \mbox{$\Delta E$}\xspace distributions for \mbox{$m_{\rm ES}$}\xspace$>5.27$ \mbox{GeV/$c^{\rm 2}$}\xspace are shown for (c) $B^+ \rightarrow J/\psi K^+ K^- K^+$ and (d) $B^0 \rightarrow J/\psi K^- K^+ \ensuremath{K^0_{\scriptscriptstyle S}}\xspace$. The continuous (red) curve represents the signal plus background, while the dotted (blue) curve represents the fitted background. Vertical (blue) lines indicate the selected signal regions.}
\end{center}
\end{figure*}
\begin{figure*}[ht]
\begin{center}
\mbox{
\scalebox{0.28}{\includegraphics{Fig5v25.eps}}
\scalebox{0.28}{\includegraphics{Fig6v23.eps}}
}
\mbox{
\scalebox{0.28}{\includegraphics{./Fig7v23.eps}}
\scalebox{0.28}{\includegraphics{./Fig8v23.eps}}
}
\mbox{
\scalebox{0.28}{\includegraphics{./Fig9v23.eps}}
\scalebox{0.28}{\includegraphics{./Fig10v23.eps}}}
\caption{\label{fig2} (a) The $K^+K^-$ mass spectrum, (c) \mbox{$m_{\rm ES}$}\xspace, and (e) \mbox{$\Delta E$}\xspace distribution for $B^+ \rightarrow J/\psi \phi K^+$. (b) The $K^+K^-$ mass spectrum, (d) \mbox{$m_{\rm ES}$}\xspace, and (f) \mbox{$\Delta E$}\xspace distribution for $B^{0} \rightarrow J/\psi \phi \ensuremath{K^0_{\scriptscriptstyle S}}\xspace$. The dots are the data points, the shaded (yellow) distributions are obtained from the \mbox{$\Delta E$}\xspace sidebands. Vertical (blue) lines indicate the selected signal regions. In (a) and (b) the \mbox{$m_{\rm ES}$}\xspace and \mbox{$\Delta E$}\xspace selection criteria described in Sec. II have been applied.}
\end{center}
\end{figure*}
\begin{figure*}[ht]
\begin{center}
\mbox{
\scalebox{0.28}{\includegraphics{./Mes2body.eps}}
\scalebox{0.28}{\includegraphics{./de2body.eps}}
}
\caption{\label{fig3} (a) The \mbox{$m_{\rm ES}$}\xspace and (b) \mbox{$\Delta E$}\xspace distribution for $B^0 \rightarrow J/\psi \phi$ event candidates. The curves in (a) and (b) are the result of the fits described in the text.}
\end{center}
\end{figure*}
\begin{table*}[!htb]
\caption{Event yields, efficiencies ($\epsilon$) and BF measurements ({\ensuremath{\cal B}\xspace}) for the different decay modes. For channels involving \ensuremath{K^0_{\scriptscriptstyle S}}\xspace, the yields and efficiencies refer to $\ensuremath{K^0_{\scriptscriptstyle S}}\xspace \ensuremath{\rightarrow}\xspace \ensuremath{\pi^+}\xspace \ensuremath{\pi^-}\xspace$, the BF includes the corrections for $K^0_S \rightarrow \pi^0 \pi^0$ and $K^0_L$ decay. The $B^0 \rightarrow J/\psi \phi$ UL at 90$\%$ c.l. is listed at the end of the table.}
\begin{center}
\begin{tabular}{lrcccccl}
\hline\hline
$B$ channel & Event& $\epsilon$ ($\%$)& Corrected &{\ensuremath{\cal B}\xspace}\ ($\times 10^{-5}$) \\
& yield& & yield & \\
\hline
\noalign{\vskip2pt}
$B^+ \rightarrow J/\psi K^+ K^- K^+$ &$~~$290$\pm$22&$~~~$15.08$\pm$0.04&~~1923$\pm$146~&~~3.37$\pm$0.25$\pm$0.14 \cr
$B^+ \rightarrow J/\psi \phi K^+$ &$~~$189$\pm$14&$~~~$13.54$\pm$0.04 &~~1396$\pm$103 &~~5.00$\pm$0.37$\pm$0.15\cr
$B^0 \rightarrow J/\psi K^+ K^- K^0$ &$~~$68$\pm$13 &$~~~$10.35$\pm$0.04 &~~~~657$\pm$126&~~3.49$\pm$0.67$\pm$0.15\cr
$B^0 \rightarrow J/\psi \phi K^0$ &41$\pm$~7 &$~~~$10.10$\pm$0.04 &~~~406$\pm$69& ~~4.43$\pm$0.76$\pm$0.19\cr
$B^0 \rightarrow J/\psi \phi$ &~~~~6 $\pm$ 4 &$~~~$31.12$\pm$0.07 &~~~~19~$\pm$13&~$<0.101$ \cr
\hline
\end{tabular}
\end{center}
\label{tab:tab1}
\end{table*}
\begin{table*}[!htb]
\caption{Systematic uncertainty contributions ($\%$) to the evaluation of the BFs.}
\begin{center}
\begin{tabular}{lcccccccc}
\hline\hline
Source&$B^+ \rightarrow J/\psi K^+ K^- K^+$ ~& ~~$B^+ \rightarrow J/\psi \phi K^+$~ &~~$B^0 \rightarrow J/\psi K^- K^+ \ensuremath{K^0_{\scriptscriptstyle S}}\xspace$~&~~$B^0 \rightarrow J/\psi \phi \ensuremath{K^0_{\scriptscriptstyle S}}\xspace$~&~~$B^0 \rightarrow J/\psi \phi$\\
\hline
\noalign{\vskip2pt}
$B \overline B$ counting & 0.6 & 0.6 & 0.6 & 0.6 & 0.6\cr
Efficiency & 0.04 & 0.04 & 0.04 & 0.04 & 0.07\cr
Tracking & 0.9 & 0.9 & 1.2 &1.2 & 0.7\cr
\ensuremath{K^0_{\scriptscriptstyle S}}\xspace & $-$ & $-$ & 1.7 & 1.7 & $-$\cr
Secondary BFs & 0.08 &0.5 &0.1 &0.5 & 0.5\cr
Decay model & $-$ &0.4 & $-$ &0.9 & 1.0\cr
$\rm pdfs$ & 3.0 &0.7 &2.0 & 0.5 & 1.0 \cr
PID & 2.5 &2.5 &3.0 &3.0 & 2.0 \cr
\hline
Total contribution&4.1&3.0&4.2&4.4&2.7 \cr
\hline
\end{tabular}
\end{center}
\label{tab2}
\end{table*}
\begin{figure*}[ht]
\begin{center}
\mbox{
\scalebox{0.28}{\includegraphics{./DalitzBch.eps}} \hspace{5mm}
\scalebox{0.30}{\includegraphics{./DalitzB0.eps}} \hspace{-5mm}
}
\caption{\label{fig5} Efficiency distribution on the Dalitz plot for (a) $B^+ \rightarrow J/\psi \phi K^+$ and (b) $B^0 \rightarrow J/\psi \phi \ensuremath{K^0_{\scriptscriptstyle S}}\xspace$.}
\end{center}
\end{figure*}
\begin{figure*}[ht]
\begin{center}
\mbox{
\scalebox{0.28}{\includegraphics{./Fig11v25.eps}}
\scalebox{0.28}{\includegraphics{./Fig12v24.eps}}
}
\caption{\label{fig4} Invariant mass distribution $J/\psi K^+ K^- $ for (a) $B^+ \rightarrow J/\psi K^+ K^- K^+$ and (b) $B^0 \rightarrow J/\psi K^+ K^- \ensuremath{K^0_{\scriptscriptstyle S}}\xspace$.
The shaded (yellow) histogram on each figure indicates the background estimated from the \mbox{$\Delta E$}\xspace~ sidebands.}
\end{center}
\end{figure*}
Figure~\ref{fig1} ~shows ~the ~\mbox{$m_{\rm ES}$}\xspace ~~distributions ~for ~(a) $B^+ \rightarrow J/\psi K^+ K^- K^+$ and (b) $B^0 \rightarrow J/\psi K^- K^+ \ensuremath{K^0_{\scriptscriptstyle S}}\xspace$ candidates after having applied the \mbox{$\Delta E$}\xspace selections described in Sec.~II, while the corresponding \mbox{$\Delta E$}\xspace distributions are shown in Fig.~\ref{fig1}(c) and Fig.~\ref{fig1}(d), respectively, for \mbox{$m_{\rm ES}$}\xspace$>5.27$ \mbox{GeV/$c^{\rm 2}$}\xspace. Figure~\ref{fig2} shows the $K^+K^-$ invariant mass distribution in the region $m_{K^+ K^-} <$1.1 \mbox{GeV/$c^{\rm 2}$}\xspace for (a) $B^+$ and (b) $B^0$ candidates. A clean $\phi(1020)$ signal is present in both mass spectra. The background contributions, estimated from the $\mbox{$\Delta E$}\xspace$ sidebands in the range $\rm{40}<|\mbox{$\Delta E$}\xspace|<\rm{70~ MeV}$, are shown as shaded histograms in Fig.~\ref{fig2}(a) and Fig.~\ref{fig2}(b) and are seen to be small. In the following we have ignored the presence of possible additional S-wave contributions in the $\phi(1020)$ signal region.
We select the $\phi(1020)$ signal region to be in the mass range [1.004$-$1.034] \mbox{GeV/$c^{\rm 2}$}\xspace.
Figure~\ref{fig2} shows the \mbox{$m_{\rm ES}$}\xspace distribution for (c) $B^+ \rightarrow J/\psi \phi K^+$ and (d) $B^0 \rightarrow J/\psi \phi \ensuremath{K^0_{\scriptscriptstyle S}}\xspace$ candidates, respectively, for events in the $\phi$ mass region, which satisfy the \mbox{$\Delta E$}\xspace selection criteria.
Figures~\ref{fig2}(e) and~\ref{fig2}(f) show the \mbox{$\Delta E$}\xspace distribution for \mbox{$m_{\rm ES}$}\xspace$>5.27$ \mbox{GeV/$c^{\rm 2}$}\xspace, when requiring the $\ensuremath{K^+}\xspace \ensuremath{K^-}\xspace$ invariant mass to be in the $\phi(1020)$ signal region. The distributions of Fig.~\ref{fig2}(c) and Fig.~\ref{fig2}(e) contain 212 events in the \mbox{$m_{\rm ES}$}\xspace and $\mbox{$\Delta E$}\xspace$ signal region, with an estimated background of 23 events. Similarly, those of Fig.~\ref{fig2}(d) and Fig.~\ref{fig2}(f) contain 50 events, with an estimated background of 9 events.
We search for the decay $B^0 \rightarrow J/\psi \phi$ by constraining a fitted \ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace and two loosely identified kaon candidates to a common vertex. Possible backgrounds originating from the decay $B^0 \rightarrow J/\psi K^{0*}(892)$, $K^{0*}(892) \rightarrow K^- \pi^+$, and from the channel $B^0 \rightarrow J/\psi K_1(1270)$, $K_1(1270) \rightarrow K^- \pi^+ \pi^{0}$ are found consistent with zero, after applying a dedicated selection as described in Sec.~II and Sec.~III. Figure~\ref{fig3} shows the corresponding \mbox{$m_{\rm ES}$}\xspace and \mbox{$\Delta E$}\xspace distributions. We do not observe a significant signal for this decay mode.
For Figs. 1-3 an unbinned maximum likelihood fit to each \mbox{$m_{\rm ES}$}\xspace distribution is performed to determine the yield and obtain a BF measurement~\cite{formula}.
We use the sum of two functions to parametrize the \mbox{$m_{\rm ES}$}\xspace distribution; a Gaussian function describes the signal, and an ARGUS function~\cite{argus} the background. A study of the \mbox{$\Delta E$}\xspace~sidebands did not show the presence of peaking backgrounds. Table~\ref{tab:tab1} summarizes the fitted yields obtained.
As a validation test, we fit the \mbox{$\Delta E$}\xspace distributions shown in Figs. 1-3, using a double-Gaussian model for the signal and a linear function for the background, and we obtain yields consistent with those from the \mbox{$m_{\rm ES}$}\xspace fits.
The signals in Fig.~\ref{fig1}, corresponding to the $B^+ \rightarrow J/\psi K^+ K^- K^+$ and the $B^0 \rightarrow J/\psi K^+ K^- \ensuremath{K^0_{\scriptscriptstyle S}}\xspace$ decays, yield 14.4$\sigma$ and 5.5$\sigma$ significance, respectively. Those in Fig.~\ref{fig2}, which restrict the invariant mass $m_{K^+K^-}$ to the signal region of the $\phi(1020)$ meson, are observed with significance 16.1$\sigma$ and 5.6$\sigma$, respectively. In this paper the statistical significance of the peaks is evaluated as $\sqrt{-2 ln(L_{0}/L_{\rm max})}$, where $L_{\rm max}$ and $L_{0}$ represent the maximum likelihood values with the fitted signal yield and with the signal yield fixed to zero, respectively.
We estimate the efficiency for the different channels using Monte Carlo (MC) simulations. For each channel we perform full detector simulations where $B$ mesons decay uniformly over the available phase space (PHSP). These simulated events are then reconstructed and analyzed as are the real data. These MC simulations are also used to validate the analysis procedure and the BF extractions.
Table~\ref{tab:tab1} reports the resulting integrated efficiencies for the
different channels, and the efficiency-corrected yields. The efficiency is computed in two different ways. For $B^{+} \rightarrow J/\psi \phi K^{+}$ and $B^{0} \rightarrow J/\psi \phi \ensuremath{K^0_{\scriptscriptstyle S}}\xspace$ we make use of a
Dalitz-plot-dependent efficiency, where each event is weighted by the inverse of the efficiency evaluated in the appropriate cell of the Dalitz plot shown in Fig.~\ref{fig5}. This approach is particularly important because of the lower efficiency observed at low $J/\psi \phi$ invariant mass, where the spectrum deviates from pure PHSP behavior. For the $\phi$ channels, the ``Corrected yield'' values in Table I are obtained as sums of inverse Dalitz-plot efficiencies for events in the $\phi$ signal regions with background-subtraction taken into account as described in Sec. IV. The events in the $\phi$ signal region account for about 65$\%$ of the data in the four-body final states. There is no evidence of structure in the remaining $\sim$35$\%$ of these events, and so they are corrected according to their average efficiency obtained from MC simulation of four-body PHSP samples. For these channels, $B^{+} \rightarrow J/\psi K^+ K^- K^{+}$ and $B^{0} \rightarrow J/\psi K^+ K^- \ensuremath{K^0_{\scriptscriptstyle S}}\xspace$, the PHSP corrected yield is added to the $\phi$ signal region corrected yield to obtain the ``Corrected yield'' values in lines 1 and 3 of Table I. The efficiency values in the third column of Table I correspond to ``Event yield'' divided by ``Corrected yield''.
Systematic uncertainties affecting the BF measurements are listed in Table~\ref{tab2}. The evaluation of the integrated luminosity
is performed using the method of $B \overline B$ counting~\cite{luminew}, and we assign a uniform 0.6\% uncertainty to all the final states. The uncertainty on the efficiency evaluation related to the size of the MC simulations is negligible with respect to the other contributions. The systematic uncertainty on the reconstruction efficiency of charged-particle tracks is estimated from the comparison of data samples and full detector simulations for well-chosen decay modes. In a similar way we obtain a 1.7\% systematic uncertainty in the reconstruction of \ensuremath{K^0_{\scriptscriptstyle S}}\xspace meson decays. In the case of the $B^0 \rightarrow \ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace \phi K_S^{0}$ and $B^{+} \rightarrow \ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace \phi K^{+}$ decay modes, since the \ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace and the $\phi$ are vector states, we compute the efficiency also under the assumption that the two vector mesons are transversely or longitudinally polarized. We consider the uncertainties related to the choice of the probability density functions ($\rm pdf$) in the fit procedure, by varying fixed parameters by $\pm$1$\sigma$ in their uncertainties. We also evaluate the efficiency variations for different charged-particle-track PID. All uncertainties are added in quadrature. We note that the BF for $B^+ \rightarrow J/\psi \phi K^+$ and that for $B^0 \rightarrow J/\psi \phi K^0$ are in agreement with their previous \mbox{\slshape B\kern -0.1em{\smaller A}\kern-0.1emB\kern-0.1em{\smaller A\kern -0.2em R }} measurements~\cite{oldpaper}, which already dominate the PDG average values~\cite{PDG}, but now we obtain more than four times better precision. The combination of these decay modes was observed first by the CLEO Collaboration~\cite{cleo}. Our BF value for the decay $B^+ \rightarrow J/\psi K^+ K^- K^+$ is the first measurement. For the decay $B^0 \rightarrow J/\psi K^+ K^- K^{0}$, the LHCb Collaboration has obtained a BF value ($2.02 \pm 0.43 \pm 0.17 \pm 0.08$)$\times 10^{-5}$~\cite{lhcb_3k}, which is consistent with our result.
We estimate an upper limit (UL) at 90$\%$ confidence level (c.l.) for the BF of the decay $B^0 \rightarrow J/\psi \phi$. The signal yield obtained from the fit to the \mbox{$m_{\rm ES}$}\xspace distribution is 6$\pm$4 events (Fig.~\ref{fig3}(a)), corresponding to an UL at 90$\%$ c.l. of 14 events. The Feldman-Cousins method~\cite{FC} is used to evaluate ULs on BFs. Ensembles of pseudo-experiments are generated according to the $\rm pdfs$ for a given signal yield (10000 sets of signal and background events), and fits are performed. We obtain an UL on the $B^0\ensuremath{\rightarrow}\xspace J/\psi\phi$ BF of 1.01$\times$10$^{-6}$. The Belle Collaboration reported a limit of $0.94 \times 10^{-6}$~\cite{belle_jpsiphi}, while a recent analysis from the LHCb Collaboration lowers this limit to $1.9 \times 10^{-7}$~\cite{lhcb_jpsiphi}.
We compute the ratios
\begin{equation}
R_{+} = \frac{{\cal B}(B^+ \rightarrow J/\psi K^+ K^- K^+)}{{\cal B}(B^+ \rightarrow J/\psi \phi K^+)} = 0.67 \pm 0.07 \pm 0.03
\end{equation}
and
\begin{equation}
R_0 = \frac{{\cal B}(B^0 \rightarrow J/\psi K^+ K^- K^0)}{{\cal B}(B^0 \rightarrow J/\psi \phi K^0)} = 0.79 \pm 0.20 \pm 0.05 ,
\end{equation}
and they are consistent with being equal within the uncertainties. We also compute the ratios
\begin{equation}
R_{\phi} = \frac{{\cal B}(B^0 \rightarrow J/\psi \phi K^0)}{{\cal B}(B^+ \rightarrow J/\psi \phi K^+)} = 0.89 \pm 0.17 \pm 0.04
\end{equation}
and
\begin{equation}
R_{2K} = \frac{{\cal B}(B^0 \rightarrow J/\psi K^+ K^- K^0)}{{\cal B}(B^+ \rightarrow J/\psi K^+ K^- K^+)} = 1.04 \pm 0.21 \pm 0.06.
\end{equation}
On the basis of the simplest relevant color-suppressed spectator quark model diagrams (e.g. Fig.1 of Ref.~\cite{cleo}), it would be expected that $R_+ = R_0$ and $R_{\phi} \sim R_{2K} \sim$ 1. Our measured values of these ratios are consistent with these expectations.
\begin{table*}
\caption{Results of the fits to the $B \ensuremath{\rightarrow}\xspace \ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace \phi K$ Dalitz plots. For each fit, the table gives the fit fraction for each resonance, and the 2D and 1D $\chi^2$ values. The fractions are corrected for the background component.}
\label{tab:tab3}
\begin{center}
\vskip -0.2cm
\begin{tabular}{llcccc}
\hline
Channel & Fit & $f_{X(4140)}$(\%) & $f_{X(4270)}$(\%) & 2D $\chi^2/\nu$ & 1D
$\chi^2/\nu$ \cr
\hline
$B^+$ & A & 9.2 $\pm$ 3.3 & 10.6 $\pm$ 4.8 & 12.7/12 & 6.5/20 \cr
& B & 9.2 $\pm$ 2.9 & 0. & 17.4/13 & 15.0/17 \cr
& C & 0. & 10.0 $\pm$ 4.8 & 20.7/13 & 19.3/19 \cr
& D & 0. & 0. & 26.4/14 & 34.2/18 \cr
\hline
$B^0 + B^+ $ & A & 7.3 $\pm$ 3.8 & 12.0 $\pm$ 4.9 & 8.5/12 & 15.9/19 \cr
\hline
\end{tabular}
\end{center}
\end{table*}
\section{Search for Resonance Production}
\begin{figure*}[ht]
\begin{center}
\mbox{
\scalebox{0.28}{\includegraphics{./Fig15v23.eps}}
\scalebox{0.28}{\includegraphics{./Fig16v23.eps}}
\scalebox{0.28}{\includegraphics{./Fig17v23.eps}}
}
\caption{Dalitz plot projections for $B^+ \rightarrow J/\psi \phi K^+$ on (a) $m^2_{\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace \phi}$, (b) $m^2_{\phi K^+}$, and (c) $m^2_{\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace K^+}$. The continuous (red) curves are the results from fit model A performed including the $X(4140)$ and $X(4270)$ resonances. The dashed (blue) curve in (a) indicates the projection for fit model D, with no resonances. The shaded (yellow) histograms indicate the background estimated from the \mbox{$\Delta E$}\xspace sidebands.}
\label{fig:fig_fit}
\end{center}
\end{figure*}
\begin{figure*}[ht]
\begin{center}
\mbox{
\scalebox{0.28}{\includegraphics{./Fig18v23.eps}}
\scalebox{0.28}{\includegraphics{./Fig19v23.eps}}
\scalebox{0.28}{\includegraphics{./Fig20v23.eps}}
}
\caption{Projections on the $J/\psi \phi$ mass spectrum from the Dalitz plot fit with the $X(4140)$ and the $X(4270)$ resonances for the (a) $B^+$, (b) $B^0$, and (c) combined $B^+$ and $B^0$ data samples. The continuous (red) curves result from the fit; the dashed (blue) curve in (a) indicates the projection for fit model D, with no resonances. The shaded (yellow) histograms show the background contributions estimated from the \mbox{$\Delta E$}\xspace sidebands.}
\label{fig:fig_fit_t}
\end{center}
\end{figure*}
\begin{figure*}[ht]
\begin{center}
\mbox{
\scalebox{0.28}{\includegraphics{./Fig8a.eps}}
\scalebox{0.28}{\includegraphics{./Fig8b.eps}}
\scalebox{0.28}{\includegraphics{./Fig8c.eps}}
}
\caption{(a) Average efficiency distribution as a function of $J/\psi \phi$ invariant mass for $B^+ \rightarrow J/\psi \phi K^+$. (b) Efficiency-corrected $J/\psi \phi$ mass spectrum for the combined $B^+$ and $B^0$ samples. The curve is the result from fit model A described in the text. The shaded (yellow) histogram represents the efficiency-corrected background contribution. (c) Efficiency-corrected and background-subtracted $J/\psi \phi$ mass spectrum for the combined $B^+$ and $B^0$ samples.}
\label{fig:fig8}
\end{center}
\end{figure*}
We plot in Fig.~\ref{fig4}(a) the $\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace \ensuremath{K^+}\xspace \ensuremath{K^-}\xspace$ mass distribution for $B^+ \rightarrow J/\psi K^+ K^- K^+$ and in Fig.~\ref{fig4}(b) that for $B^0 \rightarrow J/\psi K^- K^+ \ensuremath{K^0_{\scriptscriptstyle S}}\xspace$; the signal regions are defined by the \mbox{$\Delta E$}\xspace selections indicated in Sec. II and \mbox{$m_{\rm ES}$}\xspace$>5.27$ \mbox{GeV/$c^{\rm 2}$}\xspace. No prominent structure is observed in both mass spectra.
We select events in the $\phi$ signal regions and search for the resonant states reported by the CDF Collaboration in the
$J/\psi \phi$ mass spectrum~\cite{kai-bis}. The mass and the width values are fixed to
$m$=4143.4 \mbox{MeV/$c^{\rm 2}$}\xspace and $\Gamma$=15.3 \mbox{MeV}\xspace for the $X(4140)$, and $m$=4274.4 \mbox{MeV/$c^{\rm 2}$}\xspace and $\Gamma$= 32.3 \mbox{MeV}\xspace for the $X(4270)$ resonance.
We evaluate the mass resolution using MC simulations and obtain 2 \mbox{MeV/$c^{\rm 2}$}\xspace resolution in the mass region between 4100 \mbox{MeV/$c^{\rm 2}$}\xspace and 4300 \mbox{MeV/$c^{\rm 2}$}\xspace. Therefore resolution effects can be ignored because they are much smaller than the widths of the resonances under consideration.
We estimate the efficiency on each quasi-three-body Dalitz plot as the ratio between the reconstructed and generated distributions, where the values are generated according to PHSP. Figure~\ref{fig5} shows the resulting distributions evaluated over the $m^2_{\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace \phi} ~vs~ m^2_{\phi K}$ plane for the charged (a) and neutral (b) $B$ decay, respectively. The lower efficiency at low $J/\psi \phi$ mass is due to the lower reconstruction efficiency for low kaon momentum in the laboratory frame, as a result of energy loss in the beampipe and SVT material.
We test the agreement between data and MC by using a full MC simulation where the $B^+ \rightarrow J/\psi \phi K^+$ and $B^0 \rightarrow J/\psi \phi \ensuremath{K^0_{\scriptscriptstyle S}}\xspace$ decays are included with known branching fractions. We repeat the entire analysis
on these simulated data and find good agreement between generated and reconstructed branching fractions.
Resolution effects are small and are computed using MC simulations. We obtain
average values of 2.9 MeV for ($J/\psi \phi$) and 2.2 MeV for ($J/\psi K$). These small values do not produce bias in the evaluation of the efficiency and the measurement of the branching fractions.
To search for the two resonances in the $J/\psi \phi$ mass distributions, we perform unbinned maximum likelihood fits to the $B\rightarrow J/\psi \phi K$ decay Dalitz plots.
We model the resonances using S-wave relativistic Breit-Wigner (BW) functions with parameters fixed to the CDF values.
The non-resonant contributions are represented by a constant term, and no interference is allowed between the fit components.
We estimate the background contributions from the \mbox{$\Delta E$}\xspace sidebands, find them to be small and consistent with a PHSP behavior, and so in the fits they are incorporated into the non-resonant PHSP term. The decay of a pseudoscalar meson to two vector states may contain high spin contributions which could generate non-uniform angular distributions. However, due to the limited data sample we do not include such angular terms, and assume that the resonances decay isotropically. The amplitudes are normalized using PHSP MC generated events with $B$ parameters obtained from the fits to the data. The fit functions are weighted by the the two-dimensional efficiency computed on the Dalitz plots.
We perform fits separately for the charged $B^+$ sample and the combined $B^+$ and \ensuremath{B^0}\xspace samples. Due to the very limited statistics of the $B^0$ sample we do not perform a separate fit, but instead subtract the fit result for the $B^+$ sample from that for the combined $B^+$ and $B^0$ sample.
In this case we make use of the two different efficiencies for the two channels. In the MC simulation performed, we make use of a weighted mean of the two efficiencies evaluated on the respective Dalitz plots.
Table~III summarizes the results of the fits.
We report the background-corrected fit fractions for the two resonances, $f_{X(4140)}$ and
$f_{X(4270)}$, the two-dimensional (2D) $\chi^2$ computed on the Dalitz plot, and the one-dimensional (1D) $\chi^2$ computed on the $\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace \phi$ mass projection.
For this purpose, we use an adaptive binning method, and divide the Dalitz plot into a number of cells in such a way that the minimum expected population per cell is not smaller than 7. We generate MC simulations weighted by the efficiency and by the results from the fits. These are normalized to the event yield in data, using the same bin definitions.
We then compute the $\chi^2 = \sum_{i=1}^{N_{\rm cells}} (N^i_{\rm obs}-N^i_{\rm exp})^2/N^i_{\rm exp}$ where $N^i_{\rm obs}$ and $N^i_{\rm exp}$ are the data and MC simulation event yields, respectively. Indicating with $n$ the number of free parameters, corresponding to the number of resonances included in the fit, the number of degrees of freedom is $\nu=N_{\rm cells}-n$. In computing the 1D $\chi^2$ we rebin the $\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace \phi$ mass projection into 25 bins, again with at least 7 entries per bin.
We perform the fits using models with two resonances (labeled as model A), one resonance (models B and C), and no resonances (model D). The fit projections for fit A are displayed in Fig.~\ref{fig:fig_fit}, showing enhancements with a statistical significance smaller than 3.2$\sigma$ for all fit models. All models provide a reasonably good description of the data, with $\chi^2$ probability larger than 1$\%$.
We estimate systematic uncertainties on the fractions by varying the mass and the width values for both resonances within their uncertainties. The results shown in Table~\ref{tab:tab3} are corrected by the fraction of background estimated in each sample. This results in correction factors of 1.12 and 1.21 for the $B^+$ and the $B^0$ channels, respectively. We obtain the following background-corrected fractions for $B^+$:
\begin{equation}
f_{X(4140)} = (9.2 \pm 3.3 \pm 4.7)\%, \ f_{X(4270)} = (10.6 \pm 4.8 \pm 7.1) \%.
\end{equation}
Combining statistical and systematic uncertainties in quadrature, we obtain significances of 1.6 and 1.2$\sigma$ for the $X(4140)$ and the $X(4270)$, respectively.
Using the Feldman-Cousins method~\cite{FC}, we obtain the ULs at 90\% c.l.:
\begin{eqnarray}
{\ensuremath{\cal B}\xspace}(B^+ \rightarrow X(4140)K^+)\times {\ensuremath{\cal B}\xspace}(X(4140) \rightarrow \ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace \phi)/ \nonumber\\
{\ensuremath{\cal B}\xspace}(B^+ \rightarrow \ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace \phi K^+) < 0.133
\end{eqnarray}
\begin{eqnarray}
{\ensuremath{\cal B}\xspace}(B^+ \rightarrow X(4270)K^+)\times {\ensuremath{\cal B}\xspace}(X(4270)\rightarrow \ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace \phi)/ \nonumber\\
{\ensuremath{\cal B}\xspace}(B^+ \rightarrow \ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace \phi K^+) < 0.181.
\end{eqnarray}
The Feldman-Cousin intervals are evaluated as explained in Ref.~\cite{FC} and in Sec. III.
The $X(4140)$ limit may be compared with the CDF measurement of $0.149\pm 0.039\pm 0.024$~\cite{kai} and the LHCb limit of 0.07~\cite{LHCb}. The $X(4270)$ limit may be compared with the LHCb limit of 0.08.
The fit projections on the $J/\psi \phi$ mass spectrum using fit model A with two resonances are shown in Fig.~\ref{fig:fig_fit_t}(a) for $B^+$, in Fig.~\ref{fig:fig_fit_t}(b) for $B^0$, and in Fig.~\ref{fig:fig_fit_t}(c) for the combined $B^+$ and $B^0$ sample. The fit results are summarized in Table~\ref{tab:tab3}.
The central values of mass and width of the two resonances are also fixed to the values recently published by the CMS Collaboration~\cite{cms}. In this case we obtain, for the $B^+$ data, the following background-corrected fractions:
\begin{equation}
f_{X(4140)} = (13.2 \pm 3.8 \pm 6.8)\%, \ f_{X(4270)} = (10.9 \pm 5.2 \pm 7.3) \%.
\end{equation}
These values are consistent within the uncertainties with those obtained in Eq.~(5).
For comparison, CMS reported a fraction of $0.10 \pm 0.03$ for the X(4140), which is compatible with the CDF, the LHCb and our value within the uncertainties; CMS could not determine reliably the significance of the second structure X(4270) due to possible reflections of two-body decays.
Figure~\ref{fig:fig8}(a) shows the efficiency as a function of the $J/\psi \phi$ mass, obtained from a PHSP simulation of the $B^+ \ensuremath{\rightarrow}\xspace J/\psi \phi K^+$ Dalitz plot.
We observe a decrease of the efficiency in the $J/\psi \phi$ threshold region, as already observed in Fig.~\ref{fig5}.
Figure~\ref{fig:fig8}(b) shows the efficiency-corrected $J/\psi \phi$ mass spectrum for the combined $B^+$ and $B^0$ samples. To obtain this spectrum, we weight each event
by the inverse of the efficiency evaluated on the respective $B^+$ and $B^0$ Dalitz plots. The curve is the result from fit model A. The background contribution (shown shaded) is estimated from the \mbox{$\Delta E$}\xspace sidebands, and has also been corrected for efficiency. However, a few background events fall outside the efficiency Dalitz plots, and to these we assign the same efficiency as for $B$ signal events.
Finally, Fig.~\ref{fig:fig8}(c) shows the efficiency-corrected and background-subtracted $J/\psi \phi$ mass spectrum for the combined $B^+$ and $B^0$ samples.
\section{Summary}
In ~summary, ~we perform ~a study of the decays $B^{+,0} \rightarrow J/\psi K^+ K^- K^{+,0}$ and $B^{+,0} \rightarrow J/\psi \phi K^{+,0}$, and for the latter obtain much-improved BF measurements. For $B^{+} \rightarrow J/\psi K^+ K^- K^{+}$ this is the first measurement. We search for resonance production in the $\ensuremath{{J\mskip -3mu/\mskip -2mu\psi\mskip 2mu}}\xspace \phi$ mass spectrum and obtain significances below 2$\sigma$ for both the $X(4140)$ and the $X(4270)$ resonances, with systematic uncertainties taken into account.
Limits on the product Branching Ratio values for these resonances are obtained. We find that the hypothesis that the events are distributed uniformly on the Dalitz plot gives a poorer description of the data.
We also search for $B^0 \rightarrow J/\psi \phi$ and derive an UL on the BF for this decay mode, which is in agreement with theoretical expectations.
\section{Acknowledgements}
We are grateful for the
extraordinary contributions of our PEP-II\ colleagues in
achieving the excellent luminosity and machine conditions
that have made this work possible.
The success of this project also relies critically on the
expertise and dedication of the computing organizations that
support \mbox{\slshape B\kern -0.1em{\smaller A}\kern-0.1emB\kern-0.1em{\smaller A\kern -0.2em R }}.
The collaborating institutions wish to thank
SLAC for its support and the kind hospitality extended to them.
This work is supported by the
US Department of Energy
and National Science Foundation, the
Natural Sciences and Engineering Research Council (Canada),
the Commissariat \`a l'Energie Atomique and
Institut National de Physique Nucl\'eaire et de Physique des Particules
(France), the
Bundesministerium f\"ur Bildung und Forschung and
Deutsche Forschungsgemeinschaft
(Germany), the
Istituto Nazionale di Fisica Nucleare (Italy),
the Foundation for Fundamental Research on Matter (The Netherlands),
the Research Council of Norway, the
Ministry of Education and Science of the Russian Federation,
Ministerio de Ciencia e Innovaci\'on (Spain), and the
Science and Technology Facilities Council (United Kingdom).
Individuals have received support from
the Marie-Curie IEF program (European Union), the A. P. Sloan Foundation (USA)
and the Binational Science Foundation (USA-Israel).
|
\section{Introduction}
Inflation provides elegant solutions to several theoretical problems of the standard
big bang cosmology such as the horizon and flatness problems \cite{Guth:1980zm,Kazanas:1980tx},
and the slow-roll inflation paradigm~\cite{Linde:1981mu, Albrecht:1982wi}
successfully explains observations of cosmic microwave background (CMB)
and large-scale structure.
Particularly interesting is the so called large-field inflation that can generate
a sizable tensor-to-scalar ratio, $r$, within the reach of the on-going and planned CMB
experiments. Among various large-field inflation models, the simplest one is the quadratic chaotic inflation model
proposed by Linde long time ago~\cite{Linde:1983gd}.
The Planck satellite observed the CMB temperature and polarization anisotropy
with unprecedented accuracy. Planck released the first data with a series of papers in March 2013~\cite{Ade:2013rta}, providing tight constraints on the scalar spectral index $n_s$ and the tensor-to-scalar ratio $r$.
Soon after the Planck first data release,
we proposed a polynomial chaotic inflation in supergravity (SUGRA)~\cite{Nakayama:2013jka}
as an extension of Refs.~\cite{Kawasaki:2000yn,Kallosh:2010ug}.
We showed in Ref.~\cite{Nakayama:2013jka} that the predicted values of $n_s$ and $r$ can cover almost
entire region allowed by
the Planck data. The inflaton dynamics was further studied in a more general set-up in Ref.~\cite{Nakayama:2013txa}.
The polynomial chaotic inflation has gained momentum recently, especially after the BICEP2
collaboration claimed a detection of primordial B-mode polarization~\cite{Ade:2014xna}. The dynamics of
the polynomial chaotic inflation and its variation have been studied in Refs.~\cite{Linde:2014nna,Kobayashi:2014jga,Kallosh:2014xwa,Nakayama:2014hga}.
In Ref.~\cite{Linde:2014nna}, several issues were raised concerning our inflation model: 1) the real component
of the inflaton field acquires a non-zero vacuum
expectation value which depends on the inflaton field, and so, we will have no longer the simple single-field inflation;
2) the potential is not exactly polynomial as a result of 1); 3) the kinetic term of the fields will be
non-canonical and non-diagonal; 4) there is an extra minimum.
The purpose of the present letter is to study our polynomial chaotic inflation model in detail,
in answer to the above issues.
Our short answer is that, in any practical sense, the inflaton potential
is very well approximated by a polynomial potential for the parameters of our interest, and
in particular, the spectral index and the tensor-to-scalar ratio can be estimated by single-field approximation.
The kinetic terms can be easily diagonalized and canonically normalized by only slightly rotating the
field basis. There is an extra minimum which however is located outside of the validity region of our inflation model, and it does not affect the inflaton dynamics significantly.
The typical change of the field basis is so small that the predicted
values of $(n_s, r)$ remain almost unchanged.
This justifies our analysis of the polynomial chaotic inflation in SUGRA in Ref.~\cite{Nakayama:2013jka}.
Therefore, our model is a concrete realization of the polynomial chaotic inflation in SUGRA\footnote{
It was already mentioned by Linde in Ref.~\cite{Linde:2014nna} that the above issues may not be a big problem
in our polynomial chaotic inflation model~\cite{Nakayama:2013jka}.
}, for which the predicted values of the spectral index and the tensor-to-scalar ratio can cover
almost entire region allowed by Planck.
\\
The central issue in building successful chaotic inflation models in SUGRA
is how to have a good control of the inflaton potential over super-Planckian field ranges.
A simple prescription was given in the paper~\cite{Kawasaki:2000yn}, where they
introduce a shift symmetry on the inflaton field $\phi$ along its imaginary component:
\begin{eqnarray}
\label{ss}
\phi \to \phi + i C,
\end{eqnarray}
where $C$ is a real transformation parameter.
The K\"ahler potential takes the following form~\cite{Kawasaki:2000yn}
\begin{equation}
K = \frac{1}{2}(\phi+\phi^\dagger)^2 + |X|^2 + \cdots
\end{equation}
which satisfies the shift symmetry, whereas
it is explicitly broken by the superpotential of the form,
\begin{equation}
W = mX \phi,
\end{equation}
where $X$ is a singlet chiral superfield with $R$-charge $2$. The introduction of $X$ is crucial for avoiding a negative inflaton
potential at large field values of $\phi$. The scalar potential in SUGRA is given by
\begin{equation}
V = e^K \left[ K^{i\bar j}(D_iW)(D_{\bar j}\bar W) - 3|W|^2 \right],
\end{equation}
where $D_iW \equiv \partial_i W + K_i W$ and $K^{i\bar j} = K_{i\bar j}^{-1}$.
Here and in what follows, we adopt the Planck units in which $M_P\simeq 2.4\times 10^{18}$\,GeV is set to be unity.
The inflaton potential is generated by the small shift symmetry breaking
superpotential, and it is given by
\begin{equation}
V = \frac{1}{2}m^2\varphi^2,
\end{equation}
where $\varphi\equiv \sqrt{2}{\rm Im}(\phi)$, and both $X$ and ${\rm Re}[\phi]$ are stabilized
at the origin.
The approximate shift symmetry ensures the flatness of the potential along the imaginary
component ${\rm Im}(\phi)$ beyond the Planck scale.
Now we move on to the polynomial chaotic inflation model~\cite{Nakayama:2013jka,Nakayama:2013txa}.
We consider the following K\"ahler potential satisfying the shift symmetry (\ref{ss}),
\begin{equation}
K = \frac{1}{2}(\phi + \phi^\dagger)^2 + |X|^2 - c_\phi |X|^2(\phi + \phi^\dagger)^2 - c_X|X|^4 + \cdots
\end{equation}
where $c_\phi$ and $c_X$ are constants of order unity, the dots represent higher
order terms, and a linear term of $\phi + \phi^\dag$ is dropped, since it does not affect the inflaton
dynamics~\cite{Nakayama:2013txa}. We introduce shift symmetry breaking terms in the superpotential
as~\footnote{In Ref.~\cite{Nakayama:2014hga}, we proposed an extension to include multiple $X$ fields.
For instance we can consider $W = X_1 f(\phi) + X_2 g(\phi)$ to induce the scalar potential
$V = e^K \left(|f(\phi)|^2 + |g(\phi)|^2 \right)$. By taking e.g. $f(\phi) \propto \phi$ and $g(\phi) \propto \phi^2$,
one can realize a polynomial chaotic inflation in supergravity without a cross term.
}
\bea
W = m X (\phi + k_2 \phi^2 + k_3 \phi^3 + \cdots),
\label{Wmany}
\end{eqnarray}
where $k_i$ represents the numerical coefficient of higher order terms.
See Ref.~\cite{Harigaya:2014qza} for the case with shift symmetry breaking terms in the K\"ahler potential.
To be concrete, we focus on the case where the first two terms in the superpotential
make the dominant contribution to the inflaton dynamics:
\begin{equation}
W = m X \left(\phi + \lambda e^{i \theta} \phi^2 \right), \label{W}
\end{equation}
where we have defined $\lambda \equiv |k_2|$ and $\theta \equiv {\rm arg}[k_2]$, and
we assume $\lambda = {\cal O}(0.1)$~\cite{Nakayama:2013jka,Nakayama:2013txa}\footnote{The cut-off scale one order of magnitude larger
than the Planck scale can be
understood as follows.
Suppose that the shift symmetry is broken by various Planck-suppressed shift symmetry breaking terms.
Then, if the kinetic term coefficient happens to be enhanced by by a factor of ${\cal O}(10-100)$,
all the higher order terms are suppressed when they are expressed in terms of the canonically normalized field.
Such an enhancement may be realized if there are many singlet scalars whose kinetic term coefficients are
subject to a certain random distribution~\cite{Nakayama:2014hga}. }. See also Appendix of this letter
for another case.
First let us see that $X$ is stabilized at the origin $X = 0$ during inflation.
This is because $X$ obtains an inflaton-dependent mass term as
\begin{equation}
V \supset 4c_X |X|^2 m^2 |\phi + \lambda e^{i \theta} \phi^2|^2 \simeq 12c_X H^2 |X|^2,
\end{equation}
where $H$ denotes the Hubble parameter during inflation.
Therefore, for $c_X \gtrsim \mathcal O(0.1)$, $X$ obtains a mass of order of the Hubble scale
and it is stabilized at the origin during inflation.\footnote{
By including the constant term in the superpotential $W_0 = m_{3/2}$, where $m_{3/2}$ denotes the gravitino mass,
the minimum of $X$ during inflation slightly is deviated from $X=0$. Such a shift is safely neglected
as long as $m_{3/2}$ is much smaller than the inflaton mass.
}
Thus in the following analysis we take $X=0$.
Note that the inflaton field is canonically normalized for $X=0$.
\begin{figure}
\begin{center}
\includegraphics[scale=1.5]{pot_app_p1q2.eps}
\caption{
Schematic picture for the scalar potential (\ref{pol}).
}
\label{fig:pot12}
\end{center}
\end{figure}
Let us decompose the scalar field $\phi$ as
\begin{equation}
\phi = \frac{1}{\sqrt{2}}(\chi + i\varphi),
\end{equation}
where $\chi$ and $\varphi$ are real and imaginary components, respectively.
As noted above, $\varphi$ can develop a field value much larger than the Planck scale because of the shift symmetry.
On the other hand, $\chi$ obtains a Hubble-induced mass and stabilized at sub-Planckian field values $|\chi| \ll 1$.
In Refs.~\cite{Nakayama:2013jka,Nakayama:2013txa} we approximated $\chi \approx 0$ and obtained the inflaton potential
as
\begin{equation}
V \;\simeq\; V_{\rm inf}(\varphi) = \frac{1}{2}m^2 \varphi^2
\left( 1 - \sqrt{2} \lambda \sin \theta\, \varphi + \frac{\lambda^2}{2} \varphi^2 \right).
\label{pol}
\end{equation}
The potential shape is shown in Fig.~\ref{fig:pot12}.
One can see that the inflaton potential changes its form as one varies $\theta$. Therefore, as long as the approximation is
valid, the polynomial chaotic inflation can be realized by the first two terms in (\ref{Wmany}).
In Ref.~\cite{Kallosh:2010ug} the superpotential was extended to be the form
of $W = X f(\phi)$, where $f(\phi)$ is an arbitrary holomorphic function, and it was shown that the real component
of $\phi$ can be stabilized at the origin for a certain class of $f(\phi)$, where the coefficients are either purely
real or imaginary depending on the definition of the shift symmetry. In this case,
the polynomial chaotic inflation can be realized for a certain combination of three terms
in the superpotential~\cite{Kallosh:2014xwa}.
The above inflaton potential (\ref{pol}) is slightly modified once one takes account of the fact that
the real component $\chi$ acquires a vacuum expectation value which depends on $\varphi$,
as pointed out in Ref.~\cite{Linde:2014nna}.
To see this, we expand the full SUGRA potential in $\chi$.
Then we obtain
\bea
V&=&V_{\rm inf}(\varphi) \left(1+(2c_\phi+1)\chi^2\right)+ \frac{1}{\sqrt{2}}\lambda\cos\theta m^2\varphi^2\chi
+ \frac{1}{2} \lambda^2 m^2\varphi^2\chi^2+ \cdots.
\label{Vhigher}
\end{eqnarray}
Thus $\chi$ obtains a mass of order Hubble scale and it is stabilized at
\bea
\chi_{\rm min} \;\approx\; - \frac{\sqrt{2} \lambda \cos\theta}{2 c_\phi+1} \frac{1}{1-\sqrt{2} \lambda \sin \theta \varphi + \frac{\lambda^2}{2} \varphi^2}
\label{vsugra}
\end{eqnarray}
during inflation. It is the $\varphi$-dependence of $\chi_{\rm min}$ that modifies the inflationary path and the inflaton
potential, because the constant part of $\chi_{\rm min}$ simply modifies the coefficients of the polynomial potentials
as can be seen from (\ref{Vhigher}).
In fact,
it is easy to see that the $\varphi$-dependence of $\chi_{\rm min}$ is rather suppressed:
\bea
\frac{\partial \chi_{\rm min}}{\partial \varphi} = {\cal O}(\lambda^2),
\end{eqnarray}
which becomes even smaller for $\varphi \gtrsim \lambda^{-1}$.
For $\lambda = {\cal O}(0.1)$, therefore, the modification of the inflationary path as well as to the inflaton potential
is at most of order ${\cal O}(1)$\% level. The corrections to $n_s$ and $r$ are expected to be of a similar
order.
The contour of the inflaton potential in the complex $\phi$ plane is shown in Fig.~\ref{fig:pot}.
It is seen that there are two global minima at
\begin{equation}
\phi=0,~~~\phi=-\lambda^{-1} e^{-i\theta},
\label{min}
\end{equation}
and the potential is deformed asymmetrically. Note that the second
minimum is super-Planckian along the real component for $\lambda = {\cal O}(0.1)$ and a general value of $\theta$,
and therefore it is outside the validity region of our inflation model. We draw the contour
only for visualization purpose, simply by extrapolating the SUGRA potential to $|\chi| \gg 1$.
In any case, there is an exponential potential barrier between these two minima, and the effect of deformation is not significant.
Also for $c_\phi > 0$, $\chi$ becomes heavier and the inflationary trajectory becomes closer to $\chi = 0$.
For the reasons stated above, we expect that we can approximately set $\chi \simeq 0$ during inflation
as in our previous study~\cite{Nakayama:2013jka,Nakayama:2013txa}. Next we study the inflaton dynamics
numerically to show this explicitly.
\begin{figure}
\begin{center}
\includegraphics[scale=1.2]{pot_p1q2_xi2e-1_witharrow.eps}
\caption{
Contours of the scalar potential in Planck unit for the model (\ref{W}), $\theta=\pi/3$ and $\lambda= 0.2$.
We have taken $m=10^{-5}$ and $c_\phi = 0$. The inflaton path is well approximated by $\chi \simeq 0$. The arrow shows the second minimum in (\ref{min}), which is located outside the validity region of our model. See the text for details.
}
\label{fig:pot}
\end{center}
\end{figure}
In order to see how large is the effect of the deformation of the inflaton potential on the predicted values of $n_s$ and $r$,
we have performed numerical calculation using the full SUGRA potential.
We have solved the two field inflaton dynamics $\chi$ and $\varphi$ and identify the inflaton direction $\tilde\varphi$ as a mixture of
$\chi$ and $\varphi$ as $\tilde\varphi = c \varphi + s \chi$ where~\cite{Gordon:2000hv}
\begin{equation}
c = \frac{V_\varphi}{\sqrt{V_\varphi^2+V_\chi^2}},~~~s=\frac{V_\chi}{\sqrt{V_\varphi^2+V_\chi^2}}.
\end{equation}
with subscript $\varphi$ and $\chi$ being the derivative with respect to it.
The scalar spectral index and the tensor-to-scalar ratio is obtained from $n_s = 1-6\epsilon + 2\eta$
and $r=16\epsilon$ where
\begin{equation}
\epsilon = \frac{1}{2}\left(\frac{V'}{V}\right)^2,~~\eta = \frac{V''}{V},
\end{equation}
with prime denoting the derivative with respect to $\tilde\varphi$:
they are given by $V' = cV_\varphi + sV_\chi$, $V'' = c^2 V_{\varphi\varphi} + 2sc V_{\varphi\chi} + s^2 V_{\chi\chi}$.
They are evaluated at the point where the e-folding number is $N_e$.
In the numerical analysis, we take $N_e = 60$.
\begin{figure}[t!]
\begin{center}
\includegraphics[scale=1.8]{p1q2.eps}
\vskip 1cm
\includegraphics[scale=1.8]{p1q2_c1.eps}
\caption{
Comparison of $(n_s, r)$ between full SUGRA result and approximate result for $\theta=\pi/3, 3\pi/8$ and $\pi/2$
for the model (\ref{W}).
We have taken $c_\phi=0$ (upper panel) and $c_\phi=1$ (lower panel).
}
\label{fig:ns}
\end{center}
\end{figure}
The result is shown by (red) solid lines in Fig.~\ref{fig:ns} for $\theta=\pi/3, 3\pi/8, \pi/2$
and $c_\phi=0$ (upper panel) and $c_\phi=1$ (lower panel). We have varied $\lambda$ in the range of
$\lambda=0 \sim 0.2$. The Planck normalization on the density perturbation is imposed.
For comparison, we have also plotted the result for the approximate case where $\chi$ is set to be zero ((black) dashed lines).
For $c_\phi=0$ (upper panel), the results based on the full SUGRA potential agree well with the one based
on the single-field approximation.
The discrepancy between these two results are actually small:
the change of $(n_s,r)$ can be absorbed by small change of $\theta$.
This is because the $\varphi$-dependence of $\chi_{\rm min}$ stretches the inflaton potential by a small amount, which effectively amounts to shifting the parameters $\lambda$ and $\theta$ slightly.
The lower panel is the same plot but for $c_\phi=1$.
As can be clearly seen, the full SUGRA results almost coincide with those of polynomial potential (\ref{pol}).
In any case, even if we take account of the full SUGRA potential, the predicted values of $(n_s, r)$ of our model
cover the almost entire region allowed by Planck, as in the case where the inflaton potential is approximated by a polynomial.
Considering the uncertainties on $N_e$ and observational errors of $n_s$ and $r$, the single-field approximation
with a polynomial potential is sufficient to estimate the prediction of $n_s$ and $r$. We have also estimated $n_s$ and $r$ based on $\delta N$-formalism solving the two field dynamics numerically, and obtained consistent results.
\section*{Acknowledgments}
We thank Andrei Linde for pointing out the issues of the polynomial chaotic inflation model.
This work was supported by the Grant-in-Aid for Scientific Research on
Innovative Areas (No.23104008 [FT], No.26104009 [TTY]), Scientific Research (A) (No.26247042 [KN, FT]),
Scientific Research (B) (No.26287039 [FT, TTY]),
Young Scientists (B) (No.26800121 [KN], No.24740135) [FT]), and Inoue Foundation for Science [FT].
This work was also supported by World Premier International Center Initiative (WPI Program), MEXT, Japan.
|
\section{#1}\let\thesection\oldthesection}
\newtheorem{Def}[Theorem]{Definition}
\newtheorem{notation}[Theorem]{Notation}
\newtheorem{say}[Theorem]{}
\newtheorem{remarks}[Theorem]{Remarks}
\theoremstyle{remark}
\newtheorem{claim}{Claim}
\newtheorem{Remark}[Theorem]{Remark}
\newtheorem{Example}[Theorem]{Example}
\newtheorem{construction}[Theorem]{Construction}
\DeclareFontFamily{OT1}{pzc}{}
\DeclareFontShape{OT1}{pzc}{m}{it}
{<-> s * [1.1] pzcmi7t}{}
\DeclareMathAlphabet{\mathpzc}{OT1}{pzc}
{m}{it}
\title[A boundary divisor in the moduli space of stable quintic surfaces]
{A boundary divisor in the moduli space of\\ stable quintic surfaces}
\author{Julie Rana}
\begin{document}
\maketitle
\begin{abstract}
We give a bound on which singularities may appear on Koll\'ar--Shepherd-Barron--Alexeev stable surfaces for a wide range of topological invariants and use this result to describe all stable numerical quintic surfaces (KSBA-stable surfaces with $K^2=\chi=5$) whose unique non Du Val singularity is a Wahl singularity. We then extend the deformation theory of Horikawa in~\cite{horikawa1975} to the log setting in order to describe the boundary divisor of the moduli space $\overline{\mathcal{M}}_{5,5}$ corresponding to these surfaces. Quintic surfaces are the simplest examples of surfaces of general type and the question of describing their moduli is a long-standing question in algebraic geometry.
\end{abstract}
\section{Introduction}\label{Introduction}
Let $\mathcal{M}_{K^2, \chi}$ be the moduli space of minimal surfaces of general type, and $\overline{\mathcal{M}}_{K^2, \chi}$ its KSBA compactification~\cite{ksb1988, alexeev1994b}. Here stable surfaces are surfaces with ample canonical class and at most semi log canonical singularities. The moduli spaces $\overline{\mathcal{M}}_{K^2, \chi}$ are complicated; they may have many connected components~\cite{catanese1986} and arbitrary singularities~\cite{vakil2006}. Recently there has been substantial interest in describing singular stable surfaces explicitly, as a means to understanding the structure of the moduli spaces themselves. We are especially interested in those singularities with a one-parameter $\mathbb{Q}$-Gorenstein smoothing, as these may, in the absence of obstructions, give a divisor in the boundary of the moduli space, corresponding to equisingular deformations of the singularity (see, for example,~\cite{hacking2012}).
An important type of semi log canonical singularity is the cyclic quotient singularity. Those cyclic quotient singularities which admit a one-parameter smoothing are of type $\frac{1}{n^2}(1, na-1)$ where $a$ and $n$ are relatively prime~\cite{ksb1988}. We refer to these as Wahl singularities~\cite{wahl1981}.
In consideration of the above observation, we focus on surfaces whose unique non Du Val (or ADE) singularity is a Wahl singularity. We begin Section~\ref{wahlsings} with the following simple observation.
\begin{Lemma}\label{kawamata1} Let $W$ be a stable surface whose unique non Du Val singularity is a Wahl singularity, and let $X$ be its minimal resolution. Let $S$ be the minimal model of $X$, obtained by contracting all $(-1)$ curves on $S$. If $K_{S}$ is big and nef, then $K_W^2>K_{S}^2$.
\end{Lemma}
We remark that Lemma~\ref{kawamata1} is similar to a result of Kawamata~\cite[2.4, 4.6]{kawamata1992}, but in his case the surface $W$ must be the central fiber of a $\mathbb{Q}$-Gorenstein degeneration whose generic fiber is a smooth connected surface. In this paper, we study the case where the difference $K_W^2-K_{S}^2$ is as small as possible: What happens when $K_W^2=K_{S}^2+1$?
As with all cyclic quotient singularities, the minimal resolution of a Wahl singularity
consists of a string of exceptional curves with negative self-intersections. If this string contains $r$ exceptional curves, then we say that the singularity itself has length $r$. It is tempting to try to prove restrictions on the types of Wahl singularities that may appear on a given surface by bounding their lengths. This is possible; in~\cite{lee1999}, Y. Lee shows that if $W$ is a surface of general type whose unique non-Du Val singularity is a Wahl singularity of length $r$, then $r\le 400 (K_{S}^2)^4$, where $S$ is the minimal model of the minimal resolution of $W$. The following result greatly improves Lee's bound, although it applies only to those surfaces for which $K_W^2=K_{S}^2+1$.
\begin{Theorem}\label{r is two - 1} Let $W$ be a surface with a unique Wahl singularity $p$ of length $r$ and at most Du Val singularities elsewhere, and let $S$ be the minimal model of the minimal resolution of $W$. If $K_W$ and $K_{S}$ are big and nef and if $K_{S}^2=K_W^2-1$, then $r=1$ or $2$. That is, $p$ is a $\frac{1}{4}(1,1)$ or $\frac{1}{9}(1,2)$ singularity.
\end{Theorem}
Using Horikawa's descriptions of surfaces lying on the Noether line~\cite{horikawa1976}, we improve the result further for surfaces near it:
\begin{Theorem}\label{r is one - 1} With the same hypotheses as in Theorem~\ref{r is two - 1}, assume moreover that $K_W^2=2p_g-3$. If $S$ is of general type then $p$ is a $\frac{1}{4}(1,1)$ singularity. Moreover, if $p$ is a $\frac{1}{4}(1,1)$ singularity and $K_W^2>3$, then $S$ is of general type.
\end{Theorem}
Beginning in Section~\ref{classification section}, we apply this result to the moduli space $\mathcal{M}_{5,5}$ of numerical quintic surfaces, or minimal surfaces with $K^2=\chi=5$. This moduli space was described by Horikawa in~\cite{horikawa1975}, and is a union of two $40$-dimensional irreducible components meeting, transversally at a general point, in a 39-dimensional irreducible variety. Figure~\ref{picture of M55-1} gives a schematic diagram of $\mathcal{M}_{5,5}$. Each component parametrizes \emph{smooth} surfaces with $K^2=\chi=5$, although surfaces in components IIa and IIb are not quintic surfaces in the usual sense.
\begin{figure}[h!]
\centering
\includegraphics[scale=.3]{moduli_space_of_quintics.jpg}
\caption{On the left, a visualization of $\mathcal{M}_{5,5}$. Components I and IIa are 40-dimensional; IIb is 39-dimensional. On the right, how to obtain a numerical quintic surface of type IIa or IIb from double covers of $\mathbb{P}^1\times\mathbb{P}^1$, or $\mathbb{F}_2$, respectively. \label{picture of M55-1}}
\end{figure}
Theorem~\ref{r is one - 1}, together with Horikawa's description of surfaces with small $K^2$~\cite{horikawa1976}, suggests that it is possible to describe all stable surfaces lying one above the Noether line whose unique non Du Val singularity is a $\frac{1}{4}(1,1)$ singularity. We do this for the case of stable numerical quintic surfaces by looking at the minimal resolution $X$ of a stable numerical quintic surface. In particular, we prove that the surface $X$, which contains a rational curve of self-intersection $-4$, arises from the double cover of a smooth or nodal quadric, with branch locus intersecting a given curve in one of a few specified ways.
We note here three such constructions, which we refer to as surfaces of types 1, 2a, and 2b, which will be essential in describing the divisor in $\overline{\mathcal{M}}_{5,5}$ corresponding to surfaces whose unique non Du Val singularity is a $\frac{1}{4}(1,1)$ singularity. The minimal resolution of a surface of type 1 is a double cover of $\mathbb{P}^1\times\mathbb{P}^1$, branched over a sextic intersecting a given diagonal tangentially at $6$ points. The preimage of the diagonal is two $(-4)$-curves, intersecting at 6 points. Contracting one of these $(-4)$-curves gives a stable numerical quintic surface of type 1. The minimal resolutions of type 2a (respectively, 2b) surfaces are themselves minimal resolutions of double covers of $\mathbb{P}^1\times\mathbb{P}^1$ (respectively, a quadric cone), the branch curve of which is a sextic $B$ intersecting a given ruling at two nodes of $B$ and transversally at two other points.
The minimal models of the stable numerical quintic surfaces we study all arise from double covers of a smooth or nodal quadric surface. Thus, our approach is to describe equisingular deformations of these surfaces by deforming the quadric, together with its branch locus, in such a way that the $(-4)$-curve on $X$ is preserved. In doing so, we are met with an interesting difficulty: the $(-4)$-curve may break on the special fiber. The hope is that one may avoid this by performing a sequence of flops, but this is not immediate. We use representation theoretic tools to prove that such a sequence does exist in a number of important cases.
For instance, we have the following general result, which we use to describe the closure of the locus of type 1 surfaces in $\overline{\mathcal{M}}_{5,5}$.
\begin{Theorem}\label{simultaneous resolution for even intersection - 1}Let $Z$ be a smooth surface, $B$ a divisor on $Z$ with at most Du Val singularities, and $D$ a smooth irreducible divisor on $Z$. Let $(\mathcal{Z,B,D})$ be a family of triples over the unit disk in $\mathbb{C}$, with special fiber $(Z, B, D)$, and such that the divisors $\mathcal{D}_t$ and $\mathcal{B}_t$ are reduced, irreducible and smooth for $t\neq 0$. Suppose that at each point $p\in \mathcal{D}_t\cap \mathcal{B}_t$ over the general fiber, the local intersection $(\mathcal{D}_t\cdot\mathcal{B}_t)_p$ is even. Let $f:\mathcal{Y}\rightarrow \mathcal{Z}$ be the double cover branched over $\mathcal{B}$. Then there exists, after a possible finite base change, a simultaneous resolution of singularities $\psi:\mathcal{X}\rightarrow \mathcal{Y}$ such that the closure of one of the two components of $\psi^{-1}(f^{-1}(\mathcal{D}))_t$ over the general fiber has irreducible special fiber.
\end{Theorem}
In Section~\ref{wahldiv}, we explore the deformation theory of surfaces of types 1 and 2a, as these surfaces correspond to $39$-dimensional loci in $\overline{\mathcal{M}}_{5,5}$. To begin with, we describe explicit $\mathbb{Q}$-Gorenstein smoothings of type 1, 2a, and 2b surfaces to numerical quintic surfaces, showing that these loci lie on the boundary of the components of type I, IIa, and IIb of $\mathcal{M}_{5,5}$, respectively. Note that the smoothings of types 2a and 2b which we describe are simple extensions of an example of Friedman found in~\cite{friedman1983}. For surfaces of types 1 and 2a, we then prove vanishing of the cohomology groups in which obstructions to deformations of these surfaces lie, and conclude that the closures of these loci are smooth Cartier divisors in $\overline{\mathcal{M}}_{5,5}$ at their general points. This implies that $\overline{\mathcal{M}}_{5,5}$ is smooth generically along these divisors.
Section~\ref{2b} begins with a proof that obstructions to deformations of 2b surfaces do not vanish. Understanding the deformations of such surfaces proves to be the key to our full description of the divisor in $\overline{\mathcal{M}}_{5,5}$ corresponding to stable numerical quintic surfaces whose unique non Du Val singularity is a $\frac{1}{4}(1,1)$ singularity. Indeed, our study of these surfaces, together with the description of the closures of the 1 and 2a loci, and Horikawa's description of $\mathcal{M}_{5,5}$, allows us to prove the following theorem.
\begin{Theorem}\label{Wahl divisor} The locus of stable numerical quintic surfaces whose unique non Du Val singularity is a $\frac{1}{4}(1,1)$ singularity forms a divisor in $\overline{\mathcal{M}}_{5,5}$ which consists of two 39-dimensional components $\bar{1}$ and $\overline{\mbox{2a}}$ meeting, transversally at a general point, in a 38-dimensional component $\overline{\mbox{2b}}$. These components are the closures of the loci of the surfaces of types 1, 2a, and 2b described above. This divisor is smooth and Cartier at general points of the $\bar{1}$ and $\overline{\mbox{2a}}$ components, and is Cartier at general points of the $\overline{\mbox{2b}}$ component. Moreover, the types $\bar{1}$, $\overline{\mbox{2a}}$, and $\overline{\mbox{2b}}$ components belong to the closures of the components in $\mathcal{M}_{5,5}$ of types I, IIa, and IIb, respectively.
\end{Theorem}
We remark that Theorem~\ref{Wahl divisor} answers a question Friedman posed in~\cite{friedman1983}, specifically that of explicitly describing deformations of 2b surfaces.
The proof of Theorem~\ref{Wahl divisor} is outlined at the beginning of Section~\ref{2b}. We note here a few key facts which we prove in Section~\ref{2b}. The first is that the space of obstructions to $\mathbb{Q}$-Gorenstein deformations of a 2b surface is one-dimensional, and so the moduli space of $\mathbb{Q}$-Gorenstein deformations of 2b surfaces is a hypersurface singularity. Together with our description of the closures of the loci of types 1 and 2a surfaces and Horikawa's description of $\mathcal{M}_{5,5}$, this implies that it is enough to understand the equisingular deformations of a generic 2b surface. To describe these deformations, we locate a subfunctor of the functor of $\mathbb{Q}$-Gorenstein deformations of 2b surfaces, corresponding to deformations of covers. These deformations are unobstructed, so the there is a smooth component in the moduli space of equisingular deformations of a 2b surface. This observation implies that it is enough to show that the second order part of the Kuranishi function, given by the Schouten bracket, does not vanish and is not a square. We describe this bracket by extending the deformation theory of Horikawa in~\cite{horikawa1975}.
\noindent \emph{Acknowledgments.} This paper is a revision of my thesis. I am especially grateful to my advisor, Jenia Tevelev, for his guidance and support throughout. I would also like to thank Eduardo Cattani, Stephen Coughlan, Paul Hacking, and Radu Laza for many helpful discussions.
\section{Restrictions on singularities}\label{wahlsings}
We give bounds on which Wahl singularities may appear on a stable surface with limited invariants.
The two-dimensional quotient singularities which admit $\mathbb{Q}$-Gorenstein smoothings are called T-singularities, and are those cyclic quotient singularities of the form $\frac{1}{dn^2}(1, dna-1)$ where $a$ and $n$ are coprime \cite{ksb1988}. Those which admit only a one-parameter $\mathbb{Q}$-Gorenstein smoothing are T-singularities with $d=1$. They were studied first by Wahl \cite{wahl1981} and so are called \emph{Wahl singularities}.
The minimal resolution of a surface with a Wahl singularity of the form $\frac{1}{n^2}(1, na-1)$ contains a string of exceptional curves $C_1,\ldots, C_r$ such that
$$ \label{dotsofcs}
C_i\cdot C_j = \left\{ \begin{array}{rl}
1 &\mbox{ if $i=j\pm1$} \\
-b_i & \mbox{if $i=j$}\\
0 &\mbox{ otherwise}
\end{array} \right.
$$
where $[b_1,\cdots, b_r]$ is the Hirzebruch-Jung continued fraction expansion of $\frac{n^2}{na-1}$. We say that the T-string $C_1,\ldots, C_r$ and the singularity corresponding to it have \emph{length $r$}.
The T-string of a Wahl singularity has an especially useful iterative description by Wahl.
\begin{Prop}\label{wahltype}\cite{wahl1981} The cyclic quotient singularity $\frac{1}{4}(1,1)$ is a Wahl singularity of length 1 with $b_1=4$. Moreover, every Wahl singularity has a T-string $C_1,\ldots, C_r$ where $[b_1,\cdots, b_r]$ is one of the following types:
\begin{itemize}
\item[i)] if $[b_1,\ldots, b_{r-1}]$ is a Wahl singularity then
$$[2,b_1,\ldots,b_{r-1}+1]$$
and
$$[b_1+1,b_2,\ldots, b_{r-1}, 2]$$
are also Wahl singularities and
\item[ii)] The T-string of any Wahl singularity may be found by starting with the resolution $[4]$ and iterating the steps described in $i)$.
\end{itemize}
\end{Prop}
Because they are quotient singularities, Wahl singularities are log terminal~\cite[4.7]{kollarmori}.
Thus, if $W$ contains a unique Wahl singularity and is otherwise smooth, and if $X$ is its minimal resolution containing the T-string $C_1,\ldots, C_r$, then we can write
$$K_X=\phi^*K_W+\sum_{i=1}^ra_iC_i$$
where $-1<a_i<0$. There is a very simple relationship between $K_X^2$ and $K_W^2$, also discovered by Wahl.
\begin{Lemma}\label{X W and r}\cite{wahl1981} Let $W$ be a surface with a unique Wahl singularity of length $r$ and at most Du Val singularities otherwise. Let $X$ be is the minimal resolution of $W$. Then $K_X^2=K_W^2-r$.
\end{Lemma}
To describe the possible Wahl singularities which may occur on a surface with given invariants, one might hope to bound $r$ in terms of $K_W^2$ and $K_S^2$, where $S$ is the minimal model of $X$. The best known bound to date was discovered by Y. Lee.
\begin{Theorem}\label{lee's bound}\cite[Th. 23]{lee1999} Suppose $W$ is a surface of general type with a unique Wahl singularity
of length $r$. Let $X$ be its minimal resolution and $S$ the minimal model of $X$. If $K_S$ is ample then $r\le400(K_S^2)^4$.
\end{Theorem}
We prove a much stronger bound, at the cost of restricting to a smaller class of surfaces.
Let $W$ be a surface with a unique Wahl singularity of length $r$ and possibly Du Val singularities, let $\psi:X\rightarrow W$ be its minimal resolution, and $\pi: X\rightarrow S$ be the minimal model of $X$ as in Figure~\ref{firstmaps}.
\begin{figure} [h]
\centering
\[
\xymatrix{
&X \ar[ld]_\phi \ar[rd]^\pi \\
W& &S}
\]
\caption{The surfaces $W$, $X$, and $S$.\label{firstmaps}}
\end{figure}
If $\pi$ contracts $n$ $(-1)$-curves, then $K_X^2=K_S^2+n$. By Lemma~\ref{X W and r}, we have $K_X^2=K_W^2-r$. We bound $r$ by investigating the relationship between $n$ and $r$. The following Lemma shows that if $K_W$ and $K_S$ are big and nef, then $r>n$; that is, $K_W^2>K_S^2$.
\begin{Lemma}\label{kawamata}
If $K_W$ and $K_S$ are big and nef then $K_W^2>K_S^2$.
\end{Lemma}
\begin{proof}
Let $W$ be a surface with a unique Wahl singularity of type $\frac{1}{n^2}(1,na-1)$ at $p$ and at most Du Val singularities elsewhere. Since resolving the Du Val singularities on $W$ does not affect $K_W^2$ and nefness of $K_W$, we can assume without loss of generality that $W$ is smooth away from $p$. Choose $m>0$ such that $n | m$. Then $mK_W$ is Cartier.
Since $K_S$ and $K_W$ are big and nef, we have
$$h^i(S, mK_S)=h^i(S, (m-1)K_S+K_S)=0$$
and
$$h^i(W, mK_W)=h^i(W, (m-1)K_W+K_W)=0$$
for $i>0$ by the Kawamata--Viehweg vanishing theorem.
In particular,
$$\chi(S, mK_S) =h^0(S, mK_S) \textrm{ and } \chi(W, mK_W) =h^0(W, mK_W).$$
We claim that
$$h^0(W, mK_W)>h^0(X, mK_X)=h^0(S, mK_S)$$
for $m$ sufficiently large.
To see this, write
$$K_X=\phi^*(K_W)+\sum_{i}a_iC_i,$$
where $-1<a_i<0$, because $p$ is log terminal. Choose $m$ sufficiently large and divisible so that the denominators of the $a_i$ divide $m$ for all $i$. Then
$$\phi^*(mK_W)=mK_X+C,$$
where $C=-m\sum_{i}a_iC_i$ is an effective Cartier divisor. Consider the restriction exact sequence
$$0\rightarrow \mathcal{O}(mK_X)\rightarrow\mathcal{O}(\phi^*(mK_W))\rightarrow \mathcal{O}_C\rightarrow 0.$$
To show that $h^0(W, mK_W)>h^0(X, mK_X)$, it suffices to show that the induced map
$$H^0(X, \phi^*(mK_W))\rightarrow H^0(C, \mathcal{O}_C)$$
is nonzero. By the Kawamata-Shokurov base point free theorem,
we can choose a section $s$ of $mK_W$, for $m$ sufficiently large and divisible, such that $s(p)\neq 0$. Thus, the map is indeed nonzero.
Since $p$ has index $n$, the divisor $mK_W$ is Cartier and the usual Riemann--Roch Theorem holds~\cite{reidypg}.
Thus,
\begin{eqnarray*}
\chi(W, \mathcal{O}_W) +\frac{m(m-1)}{2}K_W^2 &=&\chi(W, mK_W)= h^0(W, mK_W)\\
&>& h^0(S, mK_S)= \chi(S, mK_S)\\
&= & \chi(S, \mathcal{O}_S) +\frac{m(m-1)}{2}K_S^2.
\end{eqnarray*}
Since $\psi$ is the resolution of a rational singularity, we have
$$\chi(W, \mathcal{O}_W)=\chi(X, \mathcal{O}_X)=\chi(S, \mathcal{O}_S),$$
and so $K_W^2>K_S^2$ as we wished to show.
\end{proof}
\begin{Remark} Kawamata makes a similar statement and argument, but requires that $W$ be the central fiber of a $\mathbb{Q}$-Gorenstein degeneration $\mathcal{X}\rightarrow \Delta$ whose generic fiber is a smooth connected surface~\cite[2.4, 4.6]{kawamata1992}.
\end{Remark}
Because it is difficult to give a useful bound on $r$ without any assumptions on $n$, we begin by restricting to the case that $K_W^2=K_S^2+1$. We will then use Noether's inequality together with Lemma~\ref{kawamata} to show that this holds in the case that $W$ is a stable numerical quintic surface.
\begin{Theorem}\label{r is two} Suppose $W$ is a surface with a unique Wahl singularity $p$ of length $r$ and at most Du Val singularities elsewhere. Let $X$ be its minimal resolution, and $\pi:X\rightarrow S$ the minimal model of $X$ as in Figure~\ref{firstmaps}. If $K_W$ and $K_S$ are big and nef, and if $K_W^2=K_S^2+1$, then $p$ is a $\frac{1}{4}(1,1)$, or $\frac{1}{9}(1,2)=\frac{1}{9}(1,5)$ singularity.
\end{Theorem}
\begin{Remark} Although we do not have a specific example, the assumption that $K_S$ is of general type is likely essential. In~\cite{lee-park2011}, Y. Lee and J. Park give an infinite family of examples of $\mathbb{Q}$-Gorenstein degenerations of minimal surfaces of general type with $K^2=2p_g-4$ to surfaces that contain \emph{two} Wahl singularities of type $\frac{1}{(n-2)^2}(1, n-3)$.The central fibers of the minimal resolutions of these families are minimal elliptic surfaces.
\end{Remark}
The proof of Theorem~\ref{r is two} requires two lemmas, but we begin with some notation.
Let us write $\pi$ as a composition of birational maps, each of which contracts a single (-1)-curve to a point $x_j\in X_j$:
$$
\xymatrix{X=X_n\ar[r]^{\pi_n} &X_{n-1}\ar[r]^{\pi_{n-1}}&\cdots\ar[r]^{\pi_2}&X_1\ar[r]^{\pi_1}&X_0=S}
$$
For $j\in\{1,\ldots,n\}$, let $F_j=\pi_{j}^{-1}(x_{j-1})\subset X_{j}$ be the (-1)-curve on $X_{j-1}$ obtained by blowing up the smooth point $x_{j-1}\in X_{j-1}$. Let
$$E_j=(\pi_j\circ\pi_{j+1}\circ\cdots\circ\pi_{n})^{-1}(x_{j-1})\subset X.$$
We call each $E_j$ an ``exceptional divisor" of $\pi$. With this notation, we can write
$$K_X=\pi^*(K_S)+\sum_{i=1}^nE_j.$$
We note that because the maps $\pi_i$ are birational, the self-intersection of $E_j$ is $(-1)$ and $E_i\cdot E_j=0$ for $i\ne j$. We have $E_n=F$ for some $(-1)$-curve $F$. Moreover, each $E_j$ contains at least one $(-1)$-curve and $E_j$ is not necessarily reduced, but its reduction is a tree of rational curves. Finally, each $E_j$ contains no loops of curves and pairs of curves in $E_j$ intersect at most once.
\begin{Lemma}\label{technical lemma 1} $\sum_{j=1}^n\sum_{i=1}^rE_j\cdot C_i\le r$.
\end{Lemma}
\begin{proof}
By adjunction
$$K_X\cdot \sum_{i=1}^rC_i=\sum_{i=1}^r(b_i-2).$$
It is easy to see by induction using Proposition~\ref{wahltype} that
\begin{equation}\label{claim1}\sum_{i=1}^r(b_i-2)= r+1.\end{equation}
Since $K_S$ is nef, we have
$$\pi^*K_S\cdot \sum_{i=1}^rC_i\ge 1.$$
Therefore,
\begin{equation}\label{KXwithCs}
K_X\cdot \sum_{i=1}^rC_i=\sum_{i=1}^r(\pi^*K_S+\sum_{j=1}^nE_j)\cdot C_i\ge 1+\sum_{i=1}^r\sum_{j=1}^nE_j\cdot C_i
\end{equation}
and so
$$\sum_{i=1}^r\sum_{j=1}^nE_j\cdot C_i\le K_X\cdot \sum_{i=1}^rC_i-1.$$
Combining this with Equation~\eqref{claim1} gives
\begin{equation*}
\sum_{i=1}^{r}\sum_{j=1}^{n}E_j\cdot C_i\le \sum_{i=1}^{r}C_i\cdot K_X-1=\sum_{i=1}^{r}(b_i-2)-1=r.
\end{equation*}
\end{proof}
\begin{Lemma}\label{technical lemma 2} $\sum_{i=1}^r\sum_{j=1}^{n}E_j\cdot C_i\ge2n.$
\end{Lemma}
\begin{proof}
The claim is obvious for $n=0$. Fix an exceptional divisor $E=E_j$ for some $j$ and a curve $C=C_i$ for some $i$. If $C\subset E$, then $C\cdot E_j=-1$ if and only if
$$(\pi_j\circ\pi_{j+1}\circ\cdots\circ\pi_n)(C)=x_j$$
and
$$(\pi_{j+1}\circ\pi_{j+2}\circ\cdots\circ\pi_n)(C)=F_j.$$
Otherwise, $C\cdot E_j=0$.
Thus, $\sum_{i=1}^rC_i\cdot E\ge-1$. Since we want $\sum_{i=1}^rC_i\cdot E\ge 2$, it suffices to show that there are at least three points of intersection (counted with multiplicity) among curves in the T-string which are not in $E$ and curves in $E$.
Given a T-string $\mathcal{C}$ containing curves $C_1,\ldots, C_r$, let
$$
\xymatrix{\bullet\ar@{-}[r]&\bullet\ar@{-}[r]&\cdots\ar@{-}[r]&\bullet\ar@{-}[r]&\bullet}
$$
be the dual graph of the T-string, where the $i^{\textrm{th}}$ vertex corresponds to the curve $C_i$. If $C_i\subset E$, we replace the $i^{\textrm{th}}$ vertex in the above graph by a box, and denote the resulting graph by $\Gamma_{E}$. For instance, if $\Gamma_{E}$ is
$$
\xymatrix{\Box\ar@{-}[r]&\bullet\ar@{-}[r]&\Box\ar@{-}[r]&\bullet\ar@{-}[r]&\Box}
$$
then there are at least 4 points of intersection among curves in $\mathcal{C}\backslash E$ and curves in $E$. With this notation we can immediately see that if there are less than 3 such intersections then $\Gamma_{E}$ must have one of the following forms:
\begin{enumerate}
\item[1)]\label{badone}
$$
\xymatrix{\bullet\ar@{-}[r]&\cdots\ar@{-}[r]&\bullet\ar@{-}[r]&\Box\ar@{-}[r]&\cdots\ar@{-}[r]&\Box\ar@{-}[r]&\bullet\ar@{-}[r]&\cdots\ar@{-}[r]&\bullet}
$$
\item[2)]\label{badtwo}$$
\xymatrix{\Box\ar@{-}[r]&\cdots\ar@{-}[r]&\Box \ar@{-}[r]&\bullet\ar@{-}[r]&\cdots\ar@{-}[r]&\bullet\ar@{-}[r] &\Box\ar@{-}[r]&\cdots\ar@{-}[r]&\Box}
$$
\item[3)]\label{badthree}$$
\xymatrix{\Box\ar@{-}[r]&\cdots\ar@{-}[r]&\Box \ar@{-}[r]&\bullet\ar@{-}[r]&\cdots\ar@{-}[r]&\bullet}
$$
\end{enumerate}
Since $n\ge 1$, there is a (-1)-curve $F$ in $E$.
Because $C_i^2<-1$ for all $i$, we also have that $C_i\cdot F\ge0$ for each $i$. We claim moreover that $\phi^*K_W\cdot F>0$. Suppose for a contradiction that $\phi^*K_W\cdot F\le0$. Since $K_W$ is nef, this implies that $\phi^*K_W\cdot F=0$. The surface $W$ is a resolution of Du Val singularities on a stable surface $W'$. Let $\theta: W\rightarrow W'$ be the resolution of Du Val singularities. Since $K_{W'}$ is ample, this implies that $F$ is contracted by $\theta$. But then $F$ is a $(-2)$ curve, a contradiction.
Writing $K_X=\phi^*K_W+\sum_{i=1}^ra_iC_i$, we have
\begin{equation*}
\sum_{i=1}^rC_i\cdot F\ge-\sum_{i=1}^ra_iC_i\cdot F=\phi^*K_W\cdot F-K_X\cdot F=\phi^*K_W\cdot F+1>1.
\end{equation*}
In particular,
\begin{equation}\label{F dot C}\sum_{i=1}^rC_i\cdot F \ge 2.
\end{equation}
Thus $F$ intersects at least two of the curves $C_i$, or one curve $C_i$ with multiplicity at least two. Moreover, if a curve $C_i$ intersecting $F$ is contained in $E$, then $\pi_{k+1}\circ\cdots\circ\pi_n(C_i)=F_k$ for some $k$. Thus, $\pi_{k+1}\circ\cdots\circ\pi_n(C_i)$ is a smooth curve and so $C_i\cdot F=1$.
Because $E$ does not contain loops of curves, we see that in Cases 1 and 3 the curve $F$ must intersect at least one $C_i$ which is not in $E$. In Case 1, this gives our third point of intersection. In Case 3 it gives a second.
We now have only to deal with Cases 2 and 3, for both of which we now have $$\sum_{i=1}^rC_i\cdot E\ge1.$$ Suppose there are $k$ exceptional curves $E$ such that $\sum_{i=1}^rC_i\cdot E=1$. We claim that $k=0$.
Suppose for a contradiction that $k>0$.
By the above argument and Lemma~\ref{technical lemma 1} we have
$$r\ge\sum_{j=1}^n\sum_{i=1}^rE_j\cdot C_i\ge 2(n-k) +k =2n-k.$$
Since $r=n+1$, we have that $k\ge n-1$. On the other hand, since $E_n=F$ is a single $(-1)$-curve, we have $k\le n-1$. Thus, $k=n-1=r-2$.
In particular, this implies that $r\ge 3$, that all but two curves in $\mathcal{C}$ are contained in exceptional divisors, and that all exceptional divisors other than $E_n$ satisfy $\sum_{i=1}^rC_i\cdot E=1$. This means that there is only one (-1) curve which must therefore be contained in all of the exceptional divisors.
Let us begin with Case 2. If the (-1)-curve $F$ intersects both a bullet and a box in $\Gamma_{E_1}$, then since $\Gamma_{E_i}$ is obtained from $\Gamma_{E_1}$ by replacing some boxes with bullets, this gives the third intersection point for all $E_i$. So we can assume that it intersects two boxes as in Figure~\ref{string1}.
\begin{figure}[h]
\centering
\includegraphics{string1.png}
\caption{$\Gamma_{E_1}$. The curved line along the bottom represents the (-1)-curve $F$.}\label{string1}
\end{figure}
Every exceptional divisor $E_j$ other than $E_n=F$ satisfies $\sum_{i=1}^rC_i\cdot E_j=1$ and must be a subset of $E_1$. Each $E_j$ also contains $F$, so the only possibility is that $F$ intersects $C_1$ and $C_r$.
However, by~\cite[3.2]{kawamata1992} we have $a_1+a_r=-1$,
so
$$-1=K_X\cdot F=(\phi^*K_W+\sum_{i=1}^ra_iC_i)\cdot F=\phi^*K_W\cdot F-1.$$
Therefore, $K_W\cdot\phi(F)=0$. Since $\phi(F)$ has positive arithmetic genus and $K_W$ is nef, this is a contradiction.
The final case to consider is Case 3. Here $\Gamma_{E_1}$ must be of the form:
\begin{figure}[h]
\centering
\includegraphics{string2.png}
\end{figure}
\noindent where the curved line along the bottom represents the $(-1)$-curve $F$. Here, $E_1$ is a chain of curves with a $(-1)$-curve at the end. Contracting $F$ under $\pi_n$ gives another $(-1)$-curve, and so $C_r$ is necessarily a $(-2)$-curve. Contracting $\pi_n(C_r)$ under $\pi_{n-1}$ must also give a $(-1)$-curve, so that $C_{r-1}$ must also be a $(-2)$-curve. Continuing in this way, we see that $E_1$ must consist of $n-1$ $(-2)$-curves and a $(-1)$-curve $F$. Thus, $\mathcal{C}$ must correspond to the Wahl singularity with T-string $[r+3, 2,\ldots, 2]$ or $[r, 5, 2, \ldots, 2]$.
Suppose first that $b_2=2$. Then using the fact that $K_S$ is nef and that $C_1\cdot F\ge 1$, we have
$$0=K_X\cdot C_{2}=\pi^*K_S\cdot C_{2}+\sum_{j=1}^nE_j\cdot C_{2}\ge \pi^*K_S\cdot C_2+1\ge 1$$
and we have a contradiction.
The only Wahl singularity left to consider is that with Hirzebruch-Jung continued fraction $[r, 5, 2, \ldots, 2 ]$. In this case, $\Gamma_{E_1}$ together with $F$ is the graph shown in Figure~\ref{string3}.
\begin{figure}[h]
\centering
\includegraphics{string3.png}
\caption{The remaining possibility for $\Gamma_{E_1}$. \label{string3}}
\end{figure}
\noindent Since $K_X\cdot C_2=3$ and $C_2\cdot \sum_{j=1}^nE_j\ge n$ we have
\begin{equation*}
0\le \pi^*K_S\cdot C_2= (K_X-\sum_{j=1}^nE_j)\cdot C_2= 3-\sum_{i=1}^nE_j\cdot C_2= 3-n.
\end{equation*}
This gives $n\le 3$, and so $r\le 4$. If $r=3$, then $C_2^2=-5$. The image $\pi(C_2)$ has self-intersection $0$ and arithmetic genus $1$. Therefore, by adjunction $K_S\cdot \pi(C_2)=0$, contradicting the fact that $K_S$ is big and nef.
Similarly, if $r=4$ then the $\pi(C_2)$ has self-intersection $1$ and arithmetic genus $1$. By adjunction, we have $K_S\cdot \pi(C_2)=-1$, contradicting the fact that $K_S$ is nef.
Since all possibilities lead to a contradiction, we conclude that $k=0$.
\end{proof}
We can now prove Theorem~\ref{r is two}.
\begin{proof}[Proof of Theorem~\ref{r is two}]
We must show that $r\le 2$. By Lemma~\ref{technical lemma 1} we have
$$\sum_{j=1}^n\sum_{i=1}^rE_j\cdot C_i\le r.$$
On the other hand, Lemma~\ref{technical lemma 2} tells us that
$$\sum_{j=1}^n\sum_{i=1}^rE_j\cdot C_i\ge 2n.$$
Since $n=r-1$, we have that $r\le 2$, so $p$ is a $\frac{1}{4}(1,1)$ or $\frac{1}{9}(1,2)$ singularity.
\end{proof}
Now suppose that $W$ be a stable surface whose unique non Du Val singularity is a Wahl singularity $p$ of length $r$. Let $\phi: X\rightarrow W$ be the minimal resolution of $W$, and let $\pi: X\rightarrow S$ be the minimal model of $W$, which is obtained from $X$ by contracting $n$ $(-1)$-curves.
\begin{Theorem}\label{r is one} Suppose that $K_W$ is big and nef and satisfies $K_W^2=2p_g-3$.
If $S$ is of general type then $p$ is a $\frac{1}{4}(1,1)$ singularity. Moreover, if $p$ is a $\frac{1}{4}(1,1)$ singularity and $K_W^2>3$, then $S$ is of general type.
\end{Theorem}
Noether's inequality (that for surfaces $S$ of general type, we have $K_S^2\ge 2p_g-4$) implies the following corollary of Lemma~\ref{kawamata}.
\begin{Cor}\label{kawamata corollary}
If the surface $W$ satisfies $K_W^2=2p_g-4$, then $S$ is not of general type.
\end{Cor}
The significance of the equality $K_W^2=2p_g-3$ in Theorem~\ref{r is one} is that such surfaces lie one above the ``Noether line" $K_W^2=2p_g-4$. That is, $K_W^2$ is the smallest it can be and still have $S$ be of general type.
For the proof of Theorem~\ref{r is one}, we recall Horikawa's description of minimal surfaces of general type with $K^2=2p_g-4$ in ~\cite{horikawa1976}. For $d\ge 0$, the Hirzebruch surface $\mathbb{F}_d$ is the $\mathbb{P}^1$-bundle over $\mathbb{P}^1$ whose zero section $\Delta_0$ has self-intersection $-d$. We denote by $\Gamma$ a generic fiber of $\mathbb{F}_d$ and note that $\mathbb{F}_0=\mathbb{P}^1\times\mathbb{P}^1$.
\begin{Theorem}\label{horikawa}\cite{horikawa1976} Let $S$ be a minimal algebraic surface with $K^2=2p_g-4$ for $p_g\ge 3$.
Then $S$ is the minimal resolution of one of either:
\begin{enumerate}
\item ($K^2=2$) a double cover of $\mathbb{P}^2$ branched over a curve of degree 8,
\item ($K^2=8$) a double cover of $\mathbb{P}^2$ branched over a curve of degree 10,
\item a double cover of $\mathbb{F}_d$, where $p_g\ge \max( d+4, 2d-2)$ and $p_g-d$ is even, branched over $B\sim 6\Delta_0+(p_g+3d+2)\Gamma$, or
\item ($K^2=4, 6,$ or $8$) a double cover of the Hirzebruch surface $\mathbb{F}_{p_g-2}$ branched over $B\sim6\Delta_0+(4p_g-4)\Gamma$.
\end{enumerate}
In each case, the branch curve has at most ADE singularities.
\end{Theorem}
We call a surface as in Theorem~\ref{horikawa} a \emph{Horikawa surface}. These surfaces are key to the proof of Theorem~\ref{r is one}.
\begin{proof}[Proof of Theorem~\ref{r is one}]
By taking a resolution of Du Val singularities $W'\rightarrow W$, we can assume that $W$ has no Du Val singularities. We first show that if $p$ is a $\frac{1}{4}(1,1)$ singularity and $K_W^2\ge3$, then $S$ is of general type. Since $K_W^2\ge 3$ and $K_W^2=2p_g-3$, we have $p_g\ge 3$. Because $p$ has length $1$, we have $K_X^2=K_W^2-1=2p_g-4\ge 2$. Thus, $K_S^2\ge K_X^2\ge 2$. By the Enriques-Kodaira classification, $S$ is of general type.
Now suppose that $S$ is of general type. Then $S$ satisfies Noether's inequality $K_S^2\ge 2p_g-4$. On the other hand, by Lemma~\ref{kawamata}, we have $K_S^2<K_W^2=2p_g-3$. Therefore $K_S^2=2p_g-4$. Since the maps $\pi$ and $\phi$ in Figure~\ref{firstmaps} do not affect the invariants $p_g$ and $q$, the surface $S$ must be a Horikawa surface. Furthermore, we have that $K_W^2=K_S^2-1$, so by Lemma~\ref{r is two}, the only possible Wahl singularities on $W$ have length 1 or 2.
If $p\in W$ is a Wahl singularity of length 2, then the resolution of $p$ in $X$ is a T-string $\{C_1, C_2\}$ where, without loss of generality, $C_1^2=-2$ and $C_2^2=-5$. Since $K_X^2=K_W^2-2=K_S^2-1$, the surface $X$ is the blowup of $S$ in a single point. Let $E$ be the exceptional curve of $\pi$. We have:
\begin{equation}\label{KX in terms of W}
K_X=\phi^*K_W-\frac{1}{3}C_1-\frac{2}{3}C_2
\end{equation}
\begin{equation}\label{KX in terms of S}
K_X=\pi^*K_S+E
\end{equation}
We multiply Equation~\eqref{KX in terms of S} with $C_1$ and $C_2$ and use that $K_S$ is nef to find that $E\cdot C_1=0$ and $E\cdot C_2\le 3$. On the other hand, if we multiply Equation~\eqref{KX in terms of W} with $E$ and use that $K_W$ is nef, we see that $E \cdot C_2 \ge 2$.
If $E\cdot C_2=3$, then $\pi^*K_S\cdot C_2 =0$, so $K_S\cdot \pi(C_2)=0$. Since $K_S$ is bif and nef, the only possibility is that $\pi(C_2)$ is a $(-2)$-curve. But $\pi(C_2)$ is singular, so this is not possible.
Now suppose that $E\cdot C_2 =2$. Then $K_S\cdot \pi(C_2)=1$ and $\pi(C_2)^2=-1$. This implies that $\pi(C_2)$ is a nodal or cuspidal cubic. We will use the fact that $S$ is a Horikawa surface to show that in fact such a curve cannot exist on $S$.
By Theorem~\ref{horikawa}, the surface $S$ is the minimal resolution of a surface $Y$ with at most Du Val singularities, which is in turn a double cover of $Z$ where $Z$ is either $\mathbb{P}^2$ or a Hirzebruch surface $\mathbb{F}_d$. Let $\psi: S\rightarrow Y$ be the minimal resolution of $Y$ and $f:Y\rightarrow Z$ the double cover branched over a curve $B$. See Figure~\ref{secondmaps}.
\begin{figure}[h]
\centering
\[
\xymatrix{
&X \ar[ld]_\phi \ar[rd]^\pi \\
W& &S \ar[r]^\psi & Y\ar[r]^f&Z}
\]
\caption{The surfaces $W$, $X$, $S$, $Y$ and $Z$ and their corresponding maps. Here, $Z$ is either $\mathbb{P}^2$ or $\mathbb{F}_d$ for some $d$. \label{secondmaps}}
\end{figure}
We must consider four cases, corresponding to the cases in Theorem~\ref{horikawa}. Let $C=\pi(C_2)$, and let $D=f(\psi(C))$ be the image of $C$ on $Z$.
Case I. ($K^2=2$) Suppose that $Z=\mathbb{P}^2$ and $B\sim 8H$, where $H$ is a hyperplane class. Then
$K_Y=f^*(-3H+4H)=f^*(H)$, so
\begin{equation*}
1=K_S\cdot C = \psi^*K_Y\cdot C= K_Y\cdot \psi(C)= f^*(H)\cdot \psi(C).
\end{equation*}
Since $f^*H\cdot \psi(C)$ is odd, this implies that $f^*H\cdot \psi(C)=H\cdot D=1$, so $D\sim H$.
But then $\psi^*(f^*(f(\psi(C))))$ is a union of smooth curves meeting transversally, one component of which is $C$, whereas $C$ is singular.
Case II. ($K^2=8$) If $Z=\mathbb{P}^2$ and $B\sim|10H|$, then $K_Y=f^*(2H)$. In particular $K_Y\cdot F$ is even for any $F$. However, $K_Y\cdot \psi(C)= K_S\cdot C=1$, so this case is impossible.
Case III. Suppose that $Z=\mathbb{F}_d$ and $B\sim|6\Delta_0+(p_g+3d+2)\Gamma|$ where $p_g\ge \max(d+4, 2d-2)$ and $p_g-d$ is even. Then $$K_Y=f^*\left(\Delta_0+\frac{p_g+d-2}{2}\Gamma\right).$$
We know $K_Y\cdot\psi( C)=1$, so if $f(\psi(C))\sim (a\Delta_0+b\Gamma)$ where $a$ and $b$ are nonnegative, then
$$a\frac{m-d-1}{2}+b=1.$$
Since $p_g\ge d+4$ and $f(\psi(C))$ is irreducible, there are two possibilities: $f(\psi(C))\sim \Delta_0$ or $f(\psi(C))\sim \Gamma$. But then in either case, $\psi^{*}(f^{*}(f(\psi(C))))$ is a union of smooth curves meeting transversally, with $C$ as one of the components, whereas $C$ is singular. Therefore, this case is impossible.
Case IV. Suppose that $Z=\mathbb{F}_{p_g-2}$ and $B\sim 6\Delta_0+4(p_g-1)\Gamma$. In this case, $K_Y=f^*(\Delta_0+(p_g-2)\Gamma)$. If $f(\psi(C))\sim (a\Delta_0+b\Gamma)$, where $a$ and $b$ are nonnegative, then intersecting $f(\psi(C))$ with $\Delta_0+(p_g-2)\Gamma$ implies that $b=1$. Since $f(C)$ is irreducible, we have that $a=0$, and so $f(\psi(C))\sim \Gamma$. But again, $\psi^{*}(f^{*}(f(\psi(C))))$ is a union of smooth curves meeting transversally, with $C$ as one of the components, whereas $C$ is singular, and we have a contradiction.
Therefore the only possible length Wahl singularity on $W$ has length 1, so is a $\frac{1}{4}(1,1)$ singularity.
\end{proof}
\section{Stable numerical quintic surfaces with a unique $\frac{1}{4}(1,1)$ singularity}\label{classification section}
A stable numerical quintic surface $W$ is a stable surface with $K^2=5$, $p_g=4$ and $q=0$. We classify all stable numerical quintic surfaces $W$ whose unique non Du Val singularity is a $\frac{1}{4}(1,1)$ singularity. By Theorem~\ref{r is one}, the minimal resolution $\phi: X\rightarrow W$ is a minimal surface such that $K_X^2=K_W^2=4$, $p_g=4$ and $q=0$, so $X$ is a Horikawa surface. Moreover, $X$ contains a $(-4)$-curve $C$, the exceptional divisor of $\phi$. On the other hand, given a Horikawa surface with $K^2=p_g=4$ and $q=0$ and containing a $(-4)$-curve, we can contract $C$ to obtain a stable numerical quintic surface with a unique $\frac{1}{4}(1,1)$ singularity. Thus, the classification of surfaces such as $W$ becomes a question of classifying all Horikawa surfaces with $K^2=p_g=4$ and $q=0$ that contain a $(-4)$-curve.
Theorem~\ref{r is one} suggests that in order to describe surfaces $W$ ``one above the Noether line'' whose unique non Du Val singularity is a $\frac{1}{4}(1,1)$ singularity, we might instead describe pairs $(X, C)$, where $X$ is a Horikawa surface and $C$ is a $(-4)$-curve contained in $X$. Because Horikawa surfaces are all described as minimal resolutions of double covers $f: Y\rightarrow Z$, we can attempt to ``find'' a $(-4)$ curve on a Horikawa surface by describing how such a $(-4)$ curve must arise from a curve on $Z$ intersecting the branch locus in a certain way.
We begin in~\ref{Double covers} with some notation and basic results about double covers. We then use these results in~\ref{classification of 1/4(1,1)} to describe all stable numerical quintic surfaces whose unique non Du Val singularity is a $\frac{1}{4}(1,1)$ singularity. In~\ref{dimension counts}, we count dimensions of a number of loci in $\overline{\mathcal{M}}_{5,5}$ of such surfaces, and continue in~\ref{1 and 2a closure} to prove that every stable numerical quintic surface whose unique non Du Val singularity is a $\frac{1}{4}(1,1)$ singularity lies in the closure of one of two distinguished loci.
\subsection{Double covers}\label{Double covers}
Let $f: Y\rightarrow Z$ be a double cover of a smooth surface $Z$ branched over a curve $B$ with at most ADE singularities, and let $\psi: X\rightarrow Y$ be the minimal model of $Y$, obtained by resolving all Du Val singularities on $Y$. Then by~\cite[Lemma 5]{horikawa1975}, the surface $X$ is the double cover of a smooth surface $\tilde{Z}$ with smooth branch locus $B'$ obtained as follows:
Let $p=p_0$ be a singular point of $B=B_0$ and let $\sigma_1:Z_1\rightarrow Z=Z_0$ be the blowup of $Z$ at $p$. Let $E_1$ be the exceptional divisor of $\sigma_1$, and let $B_1'=\sigma^*(B)-2E_1$. Define $f_1: Y_1\rightarrow Z_1$ to be the double cover of $Z_1$ branched over $B_1'$. Then there exists a map $\psi_1: Y_1\rightarrow Z_1$ such that the following diagram is commutative.
\[\xymatrix{
Y_1\ar[r]^{f_1} \ar[d]_{\psi_1} & Z_1 \ar[d]^{\sigma_1}\\
Y\ar[r]^f& Z
}\]
If $B_1'$ is smooth, then $Y_1$ is smooth and so we can take $B'=B_1'$, $X=Y_1$, $\tilde{Z}=Z$ and $\tilde{f}=f_1$. Otherwise, repeat the process, taking $p$ to be a singularity of $B_1'$. In this way, we obtain a map $\sigma: \tilde{Z}\rightarrow Z=Z_0$ which is a composition of maps $\sigma_1\circ\cdots\circ\sigma_m$ where $\sigma_i: Z_i\rightarrow Z_{i-1}$ is the blowup of a single smooth point $p_{i-1}\in Z_{i-1}$, where $p_{i-1}$ is a singular point of $B_i'=\sigma_{i}^*(B_{i-1}')-2E_i$.
We remark that the resolution given is not necessarily the log resolution of $B$, because we consider singularities of the curves $B_i'=\sigma_{i}^*(B_{i-1}')-2E_i$, as opposed to non-nodal singularities of the preimage of $B$.
Now suppose that $D$ is a smooth curve contained in $Z$, and let $\tilde{D}$ be the proper transform of $D$ under the map $\sigma$. We denote by $(B\cdot D)_p$ the local intersection multiplicity of $B$ and $D$ at $p\in B\cap D$. If $p\in B\cap D$ is an ADE singularity of $B$, let $D_i$ be the proper transform of $D$ under $\sigma_{1}\circ\cdots\circ\sigma_{i}$, and let $q_i$ be the point of $D_i$ such that $\sigma_{1}\circ\cdots\circ\sigma_{i}(q_i)=p$. Then for some $l>0$ we can rearrange the blowups so that $q_j=p_j$ for $j\le l$ and $q_j\neq p_j$ for $j>l$. That is, $l$ is the smallest integer for which either $B_{l}'$ is smooth at $q_{l}$ or $B_l'$ does not contain $q_l$. In addition, all maps $\sigma_{l+1}, \ldots, \sigma_{m}$ blowup points away from $q_l\in D_l$, so that
$$(B'\cdot \tilde{D})_{q}=(B_l'\cdot D_l)_{q_l}.$$
We call $l$ the \emph{separation number} of $p$ and note that $l$ depends on both the singularity of $B$ at $p$ as well as how the branches of $B$ at $p$ intersect $D$.
We state here three lemmas, the proofs of which are almost immediate, which will be useful in Theorem~\ref{classify} below.
\begin{Lemma}\label{B dot D changes} Suppose that $p\in B\cap D$ is an ADE singularity of $B$ and that $D$ is smooth. Then $(B_1' \cdot D_1)_{q_1}=(B\cdot D)_p-2$. In particular, if $l$ is the separation number of $p$, then $(B'\cdot \tilde{D})_q=(B\cdot D)_p-2l$.
\end{Lemma}
\begin{proof} We have
$$(B_1'\cdot D_1)_{q_1}=((\sigma^*B-2E_1)\cdot (\sigma^*D-E)) = (B\cdot D)_p-2,$$
as desired.
\end{proof}
\begin{Lemma}\label{D in B} Suppose that the branch locus $B$ of $f$ is reducible and contains an irreducible smooth curve $D$. Let $\bar{B}=B - D$ and let $p$ be a point of $D\cap \bar{B}$. Let $\bar{B}_1=B_1'-D_1$. Then $(\bar{B}_1\cdot D_1)_{q_1}=(\bar{B}\cdot D)_p-1$. In particular, the separation number of $p$ is equal to the local intersection $(\bar{B}\cdot D)_p$.
\end{Lemma}
\begin{proof} Since $D$ is smooth and $B$ has ADE singularities, any singularity of $B$ has either $2$ or $3$ branches at $p$, of which $D$ is locally a smooth one. If $B$ has two branches at $p$, then $p$ is either an $A_n$ singularity of $B$ for $n$ odd, a $D_n$ singularity of $B$ for $n$ odd, or an $E_7$ singularity of $B$. If $B$ has $3$ branches at $p$, then $p$ is a $D_n$ singularity of $B$ for $n$ even.
In each case, $B_1'=\sigma_1^*(\bar{B})-E_1 +\sigma_1^*(D)-E_1$. Since
$$((\sigma_1^*\bar{B}-E_1) \cdot (\sigma_1^*D-E_1))_{q_1}=(\bar{B}\cdot D)_{p}-1,$$
we have obtained the desired result.
\end{proof}
Lemma~\ref{D in B} says in particular that if $\bar{B}\cap D$ consists of $k$ distinct points with separation numbers $l_1,\ldots, l_k$, then
$$\tilde{D}^2=D^2-\sum_{i=1}^kl_i=D^2-(\bar{B}\cdot D).$$
Given $g(y)=y^k(a_k+a_{k+1}y+\textrm{h.o.t.})\in\mathbb{C}[[y]]$, where $a_k\in \mathbb{C}^*$, we call $k$ the \emph{minimal degree} of $g(y)$, and take $k=\infty$ if $g(y)=0$.
\begin{Lemma}\label{B dot D even} Suppose that $p\in B\cap D$ is an $E_8$ singularity of $B$. Then $(B\cdot D)_p$ is either $3$ or $5$.
\end{Lemma}
\begin{proof}
Note that $B$ is unibranched and has multiplicity $3$ at $p$. Thus, if the tangent cone of $B$ at $p$ is transversal to $D$, then $(B\cdot D)_p=3$. On the other hand, if the tangent cone of $B$ at $p$ is tangent to $D$, then choose coordinates on $Z$ so that $B$ has local equation $x^3+y^5$. Then $D$ is locally given by $x-f(y)$ where $f(y)$ has minimal degree $k\ge 2$.Then $(B\cdot D)_p$ is the minimal degree of $f(y)^3+y^5$. Since $f(y)$ has minimal degree at least $2$, this implies that $(B\cdot D)_p=5$.
\end{proof}
\subsection{The classification}\label{classification of 1/4(1,1)}
We continue to use the notation of subsection~\ref{Double covers}.
Let $\Gamma$ be a fiber of $Z$ and $\Delta$ an irreducible curve in the linear system $|\mathcal{O}_{\mathbb{F}_0}(1,1)|$ on $\mathbb{F}_0$ or $|\Delta_0+2\Gamma|$ on $\mathbb{F}_2$.
\begin{Theorem} \label{classify}
There is a one-to-one correspondence between stable numerical quintic surfaces with at most Du Val singularities and a unique $\frac{1}{4}(1,1)$ singularity, and triples $(Z, B, D)$, where $Z=\mathbb{F}_d$ for $d=0$ or $2$, $B\sim 6\Delta$ has at most ADE singularities, and $D\sim \Gamma$ or $D\sim \Delta$ intersects $B$ as follows:
\begin{enumerate}
\item $D\sim \Gamma$, there exists $p\in D\cap B$ such that $(B\cdot D)_p$ is odd, and $B$ has either $1$ or $2$ singularities along $D$ and intersects $D$ transversally elsewhere. Moreover,
\begin{enumerate}
\item if two singularities of $B$ are contained in $D$, then each singularity $p$ has separation number $1$, and either $(B\cdot D)_p=2$
or $(B\cdot D)_p=3$.
\item if one singularity $p$ of $B$ is contained in $D$, then $p$ has separation number $2$, and either $(B\cdot D)_p=4$ or $(B\cdot D)_p=5$.
\end{enumerate}
Figures~\ref{distinct points}, \ref{same points 4}, and \ref{same points 5} show all possible ways $B$ and $D$ may intersect in this case.
\item $D\sim \Delta$, $D\not \subset B$, and for all $p\in D\cap B$, $(B\cdot D)_p$ is even.
Figure~\ref{even case resolution} details the possible ways $B$ and $D$ may intersect in this case.
\item $D\sim \Delta$ and $D\subset B$.
\end{enumerate}
\end{Theorem}
\begin{proof}
Suppose that $W$ is a stable numerical quintic surface whose unique non Du Val singularity is a $\frac{1}{4}(1,1)$ singularity and let $X$ be its minimal resolution. Then $X$ is a Horikawa surface with $K^2=p_g=4$ and $q=0$, containing a $(-4)$-curve $C$. Let $\hat{\psi}: X\rightarrow \hat{Y}$ be the canonical model of $X$, so that $\hat{Y}$ has at most Du Val singularities. By~\cite{horikawa1976}, $\hat{Y}$ is a double cover of a smooth or singular quadric $\hat{Z}$, with branch locus away from any singularity of $\hat{Z}$. We resolve both $A_1$ singularities of $\hat{Y}$ lying over the singularity of $\hat{Z}$. Then there exists a map $\psi: X\rightarrow Y$, where $Y$ is the double cover $f: Y\rightarrow Z$ of $Z$, where $Z=\mathbb{F}_2$ or $\mathbb{F}_0$, branched over $B\sim 6\Delta$ with at most ADE singularities~\cite{horikawa1976}. We claim that the curve $D=\psi(f(C))$ is linearly equivalent to either $\Delta$ or $\Gamma$.
The canonical class $K_Z$ of $Z$ is linearly equivalent to $-2\Delta$. Let $L$ be a divisor such that $B\sim 2L$. Then since $f$ is a double cover, the canonical class $K_Y$ is given by $f^*(K_Z+L)=f^*(\Delta)$. Thus, $K_Y\cdot f^*D=2\Delta\cdot D$.
Let $\bar{C}=\psi(C)\subset Y$. If $D$ is not contained in the branch locus $B$, then $f^*(D)$ is either a union of two curves $\bar{C}$ and $\bar{C}'$ or $f^*D=\bar{C}$, depending upon how the curve $D$ intersects the branch locus $B$. More precisely, $f^*(D)=\bar{C}+\bar{C}'$ if and only if the multiplicity of $B$ and $D$ is even at each point of intersection. We consider the three cases, $f^*(D)=\bar{C}$, $f^*(D)=\bar{C}+\bar{C}'$, and $D\subset B$, separately.
\textbf{Case I.} Suppose that there exists $p\in D\cap B$ such that $(B\cdot D)_p$ is odd. Then $f^*(D)=\bar{C}$ and we have
$$2\Delta\cdot D=K_Y\cdot f^*(D)=K_Y\cdot \bar{C}= 2,$$
so $\Delta\cdot D=1$. Since $C$ is irreducible the curve $D$ is also irreducible. Thus, $D\sim\Gamma$. Note that $B\cdot D=6$.
On the other hand, since $\tilde{f}$ is the double cover of a smooth surface and $C^2=-4$, the curve $\tilde{f}(C)$ is a $(-2)$-curve $\tilde{D}$ on $\tilde{Z}$. Since $\tilde{D}$ has genus $0$ and $\tilde{f}$ is a double cover, the Riemann--Hurwitz formula gives $B'\cdot\tilde{D}=2$. Because $C$ is smooth, the branch divisor $B'$ intersects $\tilde{D}$ transversally. Commutativity of the diagram
\[\xymatrix{
X\ar[r]^{\tilde{f}} \ar[d]_{\psi} & \tilde{Z} \ar[d]^{\sigma}\\
Y\ar[r]^f& Z
}\]
implies that $\sigma(\tilde{D})=D$. Noting that $D^2=0$ and $\tilde{D}^2=-2$ we see that the map $\sigma$ blows up exactly two points $p_1$ and $p_2$ on $D$, which may be infinitely near.
Suppose that $p_1$ and $p_2$ are distinct, and let $p=p_1$. Then $p$ has separation number $1$. Moreover, because $C$ is smooth, either $B'$ intersects $\tilde{D}$ transversally at $q$, or $B'$ and $D$ do not intersect at $q$. That is, $(B'\cdot \tilde{D})_q=0$ or $1$. By Lemma~\ref{B dot D changes}, this implies that $(B\cdot D)_p=2$ or $3$. Conversely, if $(B\cdot D)_p=2$ or $3$, then since $B$ is singular, $p$ has separation number $1$.
If $(B\cdot D)_p=2$, then $p$ is an $A_n$ singularity of $B$, and any branches of $B$ at $p$ intersect $D$ transversally. See Figures~\ref{distinct points}(a) and~\ref{distinct points}(b) for the local intersection of $B$ and $D$.
Now suppose that $(B\cdot D)_p=3$. If $p$ is an $A_n$ singularity of $B$ for $n$ odd, then since $(B\cdot D)_p=3$, one branch of $B$ intersects $D$ transversally at $p$ while the other intersects $D$ at $p$ with multiplicity $2$. For $n >1$, both branches of $B$ are tangent to each other, so this is not possible. Thus, $p$ is an $A_1$ singularity of $B$ and $B$ intersects $D$ at $p$ as in Figure~\ref{distinct points}(c).
If $p$ is an $A_n$ singularity for $n$ even, then the tangent cone of $B$ at $p$ is tangent to $D$. Choose local coordinates on $Z$ so that $B$ has local equation $x^2-y^{n+1}$ and $D$ has local equation $x-f(y)$ where $f(y)$ has minimal degree $k\ge 2$. Then $(B\cdot D)_p=3$ if and only if $n=2$. In this case, the proper transform $B_1$ of $B$ is smooth and transversal to $D$, as desired. See Figure~\ref{distinct points}(d) for the local picture.
If $B$ has a $D_n$, $E_6$, $E_7$ or $E_8$ singularity at $p$, then since $(B\cdot D)_p=3$, the tangent cone of each branch of $B$ at $p$ must be transversal to $D$. The local intersection of $B$ and $D$ is shown in Figure~\ref{distinct points}(e), (f), (g), (h), and (i).
Figure~\ref{distinct points} summarizes all possible singularities of $B$ along $D$ that may occur if $\sigma$ blows up two distinct points, as well as how the exceptional curves on $\tilde{Z}$ intersect $\tilde{D}$ and the branch divisor $B'$ of $\tilde{f}$.
\begin{figure}
\centering
\includegraphics[scale=.5]{DistinctPointswithResolution.png}
\caption{On the left, the possible singularities of $B$ along $D\sim \Gamma$ if $p_1\neq p_2$. In each case, the vertical line represents the curve $D$. On the right, the curve $\tilde{D}$, dashed, together with the exceptional divisor of $\sigma$ and the proper transform of $B$. The solid concave down curves denote exceptional divisors. The branch locus of $\tilde{f}$ is denoted by solid bold curves. \label{distinct points}}
\end{figure}
We now consider the case where $p_1$ and $p_2$ are infinitely near. Denote by $p$ the center of the the blowup. Then $p$ has separation number $2$. Moreover, because the curves $\tilde{D}$ and $B'$ are transversal at $q$, we have $(B\cdot D)_p=4$ or $5$. We show that these two properties (that is, $(B\cdot D)_p=4$ or $5$ and $p$ having separation number $2$) hold if and only if $B$ and $D$ intersect at $p$ in one of the ways listed. By Lemma~\ref{B dot D changes}, if $(B\cdot D)_p=4$ or $5$ then $p$ has separation number at most $2$. Thus, it is enough to list all possible intersections with $(B\cdot D)_p=4$ or $5$ and $B_1'$ singular at $q_1$. We note that for $p$ to have separation number greater than 1, at least one branch of $B$ at $p$ must have tangent cone tangent to $D$.
Our method is to consider the possible singularities of $B$ case-by-case, choosing local coordinates at $p$ so that the equation of $B$ is a standard form (for instance $x^2-y^{n+1}$ for an $A_n$ singularity and $y(x^2-y^{n-2})$ for a $D_n$ singularity). In these coordinates, $D$ has local equation $x-f(y)$ or $y-g(x)$ where $f(y)$ (or $g(x)$) has minimal degree $k\ge 2$, and a simple case-by-case calculation tells us which values of $k$ and $n$ will give $(B\cdot D)_p=4$ or $5$. We then determine which of these values will give $B_1'$ singular at $q_1$.
Consider the case $(B\cdot D)_p=4$. Suppose that $p$ is an $A_n$ singularity of $B$ for $n$ odd. If $n=1$, then $B_1$ is smooth, so $p$ has separation number $1$. If $n>1$, then
$(B\cdot D)_p=4$ if and only if
\begin{enumerate}
\item[(1)] $n=3$ and $k>2$ (Figure~\ref{same points 4}(a)),
\item[(2)] $n=3$ and $f(y)=ay^2+\mbox{ h.o.t.}$ for $a\neq1$ (Figure~\ref{same points 4}(b)), or
\item[(3)] $n>3$ and $k=2$ (Figure~\ref{same points 4}(c)).
\end{enumerate}
Note that in each case, $B_1$ is singular at $q_1$, so $p$ has separation number $2$.
If $p$ is an $A_n$ singularity of $B$ for $n$ even, then
$(B\cdot D)_p=4$ if and only if $k=2$ and $n>2$. Since $n>2$, $B_1'$ is singular at $q_1$. Figure~\ref{same points 4}(d) shows the local intersection of $B$ and $D$.
Suppose that $p$ is a $D_n$ singularity of $B$ for $n$ odd and that the singular branch of $B$ at $p$ has tangent cone parallel to $D$. Since the smooth branch is transversal to $D$, the singular branch interesects $D$ with multiplicty $3$.
Since $n$ is odd and $(B\cdot D)_p$ is even, we have $(B\cdot D)_p =n-1=4$. Thus, $p$ is a $D_5$ singularity of $B$. See Figure~\ref{same points 4}(e) for a visualization of how $B$ and $D$ intersect at $p$. Note that $B_1'$ is singular at $q_1$ as desired.
Now suppose that $p$ is a $D_n$ singularity of $B$ for odd $n$ such that the smooth branch of $B$ at $p$ is tangent to $D$.
Then $(B\cdot D)_p$ is the minimum of $k+2$ or $k(n-1)$. Since $k>1$ and $n\ge 5$, this implies that $k=2$. The curve $B_1'$ is singular at $q_1$ as desired. See Figure~\ref{same points 4}(f) for the local intersection of $B$ and $D$ at $p$.
If $p$ is a $D_n$ singularity of $B$ for $n$ even, then two branches of $B$ at $p$ are transversal to $D$ and the third is tangent to $D$ with multiplicity $2$. For $n\ge 6$, two branches of $B$ have the same tangent cone, so the branch locus $B$ intersects $D$ at $p$ as in Figure~\ref{same points 4}(g). The local picture for $n=4$ is similar.
If $p$ is an $E_6$ singularity of $B$, then we can choose coordinates so that $x^3-y^4$ is the local equation of $B$ at $p$ and the local equation of $D$ at $p$ is $x-f(y)$, where $f(y)$ has minimal degree $k\ge 2$. We quickly see that $(B\cdot D)_p=4$ as desired. The intersection of $B$ and $D$ at $p$ is shown in Figure~\ref{same points 4}(h).
If $p$ is an $E_7$ singularity of $B$, then choose coordinates so that $B$ is locally given by $x(x^2-y^3)$ and $D$ has local equation $x-f(y)$, where $f(y)$ has minimal degree $k\ge 2$. Then $(B\cdot D)_p$ is the minimum of $3k$ and $k+3$. But we require $(B\cdot D)_p=4$, and since $k\ge 2$, this is impossible.
By Lemma~\ref{B dot D even}, $p$ is not an $E_8$ singularity.
See Figure~\ref{same points 4} for a summary of the ways in which $B$ and $D$ intersect at $p$ if $p_1=p_2$ and $(B\cdot D)_p=4$.
\begin{figure}
\centering
\includegraphics[scale=.6]{SamePoints4withResolution.png}
\caption{On left, the possible singularities of $B$ along $D$ if $p_1= p_2$ and $(B\cdot D)_p=4$. In each case, the dashed line represents $D$. On the right, the curve $\tilde{D}\sim \Gamma$, dashed, together with the exceptional divisor of $\sigma$. The solid concave down curves denote exceptional divisors. The branch locus of $\tilde{f}$ is denoted by solid bold curves. \label{same points 4}}
\end{figure}
We move on to the case $(B\cdot D)_p=5$.
Suppose that $p$ is an $A_n$ singularity of $B$ where $n$ is odd. If $n=1$, then the singularity of $B$ at $p$ is resolved after a single blowup, so we can assume that $n>1$. Choose coordinates so that the local equation of $B$ at $p$ is $x^2-y^{n+1}$ and the local equation of $D$ at $p$ is $x-f(y)$ where $f(y)=a_ky^k+a_{k+1}y^{k+1}+\mbox{ h.o.t.}$ for some $k\ge 2$. Then in order to have $(B\cdot D)_p$ odd, we must have $n+1=2k$ and $a_k=1$. Thus $(B\cdot D)_p=5=2k+1=n+2$, so $k=2$ and $n=3$. The intersection of $B$ and $D$ at $p$ is shown in Figure~\ref{same points 5}(a).
If $p$ is an $A_n$ singularity of $B$ where $n$ is even, then
$(B\cdot D)_p$ is the minimum of $2k$ and $n+1$. Thus $(B\cdot D)_p=5$ if and only if $n=4$ and $k\ge 3$. In this case, $B_1$ has an $A_2$ singularity at $q_1$.
See Figure~\ref{same points 5}(b) for the local picture.
Next, suppose that $p$ is a $D_n$ singularity of $B$ where $n$ is odd and that the tangent cone of the singular branch $S$ is tangent to $D$ at $p$. Then we have $(S\cdot D)_p=4$. Using the same analysis as in previous cases, we
see that this case occurs if and only if $n\ge 5$ and $k=2$.
See Figure~\ref{same points 5}(c) for the local picture.
If $p$ is a $D_n$ singularity of $B$ for $n$ odd such that the singular branch of $B$ at $p$ has tangent cone transversal to $D$, then the smooth branch is tangent to $D$ at $p$ with multiplicity $3$. See Figure~\ref{same points 5}(d) for the local picture.
If $p$ is a $D_n$ singularity of $B$ where $n$ is even, then
either two branches of $B$ are tangent to $D$ at $p$ with multiplicity $2$ each and the third is transversal, or two are transversal to $D$ and the third is tangent to $D$ with multiplicity $3$. In the former case, $p$ is a $D_6$ singularity and $B$ intersects $D$ at $p$ as in Figure~\ref{same points 5}(e).
In the latter case, $n$ has no further restrictions and the local intersection is shown in Figure~\ref{same points 5}(f).
We showed above that if $B$ has an $E_6$ singularity at $p$ such that the tangent cone of $B$ at $p$ is tangent to $D$, then $(B\cdot D)_p=4$, so we need not consider the singularity in this case.
If $p$ is an $E_7$ singularity of $B$, then the same analysis as above shows that $D$ is tangent to the tangent cone of $B$ at $p$ with multiplicity $2$.
See Figure~\ref{same points 5}(g) for the local picture.
Finally, suppose that $p$ is an $E_8$ singularity of $B$. An analysis of the local equations of $B$ and $D$ as above shows that as long as the tangent cone of $B$ at $p$ is tangent to $D$, we will have $(B\cdot D)_p=5$. In this case, the proper transform $B_1$ of $B$ has a cusp at $q_1$.
See Figure~\ref{same points 5}(h) for the local picture of $B$ and $D$ at $p$.
See Figure~\ref{same points 5} for a summary of the ways in which $B$ and $D$ intersect at $p$ if $p_1=p_2$ and $(B\cdot D)_p=5$.
This completes our discussion of Case I.
\begin{figure}
\centering
\includegraphics[scale=.6]{SamePoints5withResolution.png}
\caption{On left, the possible singularities of $B$ along $D$ if $p_1= p_2$ and $(B\cdot D)_p=5$. In each case, the dashed line represents $D$. On the right, the curve $\tilde{D}\sim \Gamma$, dashed, together with the exceptional divisor of $\sigma$. The solid concave down curves denote exceptional divisors. The branch locus of $\tilde{f}$ is denoted by solid bold curves. \label{same points 5}}
\end{figure}
\textbf{Case II.} Suppose that $ D\not\subset B$ and $f^*(D)=\bar{C}+\bar{C}'$. Then for each point $p$ of $B\cap D$ the multiplicity $(B\cdot D)_p$ is even, $\bar{C}$ and $\bar{C}'$ are isomorphic, and we have
$$\Delta\cdot D=\frac{1}{2}K_Y\cdot f^*D=\frac{1}{2}K_Y\cdot (\bar{C}+\bar{C}')= 2.$$
Suppose that on $\mathbb{F}_2$, we have $D\sim a\Delta_0+b\Gamma$, where $a$ and $b$ are nonnegative.
Then $D\cdot\Delta=b$, so that $b=2$. Multiplying $a\Delta_0+2\Gamma$ by $\Delta_0$, we see that in order for a divisor in the linear system $a\Delta_0+2\Gamma$ to be irreducible, we must have $a=1$. Thus, $D\sim \Delta_0+2\Gamma=\Delta$. A similar calculation on $\mathbb{P}^1\times\mathbb{P}^1$ shows that in either case $D\sim \Delta$.
We now show that if $D$ is an irreducible curve in the linear system $\Delta$ such that at each point $p\in D\cap B$ we have $(B\cdot D)_p$ even, then $\tilde{f}^{-1}(\tilde{D})$ is a union of two $(-4)$-curves $C$ and $C'$.
Suppose that $p_1, \ldots p_j$ are the singular points of $B$ lying on $D$. Let $l_i$ be the separation number of $p_i$. Then
\begin{equation*}
(C+C')^2=2\tilde{D}^2=2\left(D^2-\sum_{i=1}^jl_i\right)=2\left(2-\sum_{i=1}^jl_i\right),
\end{equation*}
where the second equality follows from Lemma~\ref{B dot D changes}.
On the other hand
\begin{equation*}
(C+C')^2=2C^2+2C\cdot C'=2C^2+B'\cdot \tilde{D}=2C^2+12-\sum_{i=1}^j2l_i=2C^2+2\left(6-\sum_{i=1}^jl_i\right),
\end{equation*}
where we again use Lemma~\ref{B dot D changes}.
Thus,
$$C^2+6-\sum_{i=1}^jl_i=2-\sum_{i=1}^jl_i,$$
so $C^2=-4$ as desired.
By Lemma~\ref{B dot D changes}, a singularity $p$ of $B$ on $D$ may be an $A_n$, $D_n$, $E_6$, or $E_7$ singularity, as long as the branches of $B$ intersect $D$ in such a way that the multiplicity of $B$ and $D$ at $p$ is even. By considering the local equations of each type of ADE singularity, we can determine all possible ways $B$ and $D$ may intersect, and also have both branches of $\tilde{f}^*(\tilde{D})$ over $p$ smooth. As an example, suppose that $B$ has a $D_n$ singularity at $p\in D$. We can choose local coordinates around $p$ so that the local equation of $B$ is $y(x^2+y^{n-2})$. Since $(B\cdot D)_p$ is even, the curve $D$ must be tangent to one of the tangent cones of $B$ at $p$. We can then write the equation of $D$ at $p$ as either $x-f(y)$ or $y-g(x)$ where $f(y)$ (or $g(x)$) has minimal degree $k\ge 2$. Then $(B\cdot D)_p$ is the minimal degree of either $y(f(x)^2+y^{n-2})$ or $g(x)(x^2+(g(x))^{n-2}$. Using the fact that $(B\cdot D)_p$ is even, we can then determine the possibilities for $k$ and for each such $k$ resolve the singularity of $B$ as described above. This tells us in particular how $\tilde{D}$ intersects the exceptional locus of $\sigma$. The same analysis may be used for the other ADE singularities. See Figure~\ref{even case resolution} for the list of possible singularities as well as how the curve $\tilde{D}$ intersects the exceptional locus of $\sigma$. We remark that this analysis is more general in that it depends only on the fact that both branches of $\tilde{f}^*(\tilde{D})$ over $p$ are smooth, and not on the self-intersection of $\tilde{D}$.
\begin{figure}
\includegraphics[scale=.5]{EvenIntersection.png}
\caption{On the left, all possible intersections of $B$ and $D$ if $(B\cdot D)_p$ is even. Here, $k$ denotes the multiplicity with which $D$ intersects the tangent cone of the given branch of $B$. On the right, the resolution of $B$ together with the exceptional curves and the proper transform of $D$ (dashed). The solid concave down curves denote exceptional divisors of $\sigma$. The branch locus of $\tilde{f}$ is denoted by solid bold curves. \label{even case resolution}}
\end{figure}
\textbf{Case III.} If $C\subset R$ then $f^*(D)=2C$, and so
$$2\Delta\cdot D=K_Y\cdot f^*D=K_Y\cdot 2D=4.$$
Since $D$ is irreducible, we must have $D\sim \Delta$. The fact that $D\subset B$ implies that $B= D + \bar{B}$ where $\bar{B}$ is in the linear system
$|5\Delta|$, so $D\cdot \bar{B}=10$. By Lemma~\ref{D in B},
we have
\begin{equation*}
\tilde{D}^2 = D^2-(\bar{B}\cdot D)= -8
\end{equation*}
as desired.
The generic $\bar{B}$ intersects $D$ intersect in 10 distinct points and so the double cover $Y$ has 10 $A_1$ singularities.
\end{proof}
\subsection{Dimension counts}\label{dimension counts}
We begin by showing that every stable numerical quintic surface $W$ whose unique non Du Val singularity is a $\frac{1}{4}(1,1)$ singularity corresponds, up to isomorphism, to a unique triple $(Z, B, D)$, where $Z$ is a quadric surface, $B\subset Z$ is a divisor in $|6\Delta|$, and $D\subset Z$ is the image of the $(-4)$-curve on the minimal resolution of $W$. This observation allows us to count the dimensions of a number of important loci in $\overline{\mathcal{M}}_{5,5}$.
\begin{Lemma}\label{triples}
Suppose that $X$ and $X'$ are the minimal resolutions of stable numerical quintic surfaces $W$ and $W'$, each of which has a unique $\frac{1}{4}(1,1)$ singularity and no other non Du Val singularities. Let $C$ and $C'$ be the $(-4)$-curves on $X$ and $X'$, respectively. Let $[W]$ and $[W']$ be the points of $\overline{\mathcal{M}}_{5,5}$ corresponding to $W$ and $W'$, respectively. The following are equivalent:
\begin{enumerate}
\item[1)] $[W]=[W']$.
\item[2)] There is an isomorphism $\theta: X\rightarrow X'$ such that $\theta(C)=C'$.
\item[3)] The triples $(Z, B, D)$ and $(Z', B', D')$ corresponding to $X$ and $X'$ are isomorphic; that is, there is an isomorphism $\eta: Z\rightarrow Z'$ such that $\eta(B)=B'$ and $\eta(D)=D'$.
\end{enumerate}
\end{Lemma}
\begin{proof}
1) $\iff$ 2) is clear.
3) $\Rightarrow$ 2) follows by construction of $X$ and $X'$ from the triples given. For 2) $\Rightarrow$ 3), suppose that $\theta: X\rightarrow X'$ is an isomorphism such that $\theta(C)=C'$.
Let $Y$ and $Y'$ be the canonical models of $X$ and $X'$, respectively, and denote by $\bar{C}$ and $\bar{C'}$ the images of $C$ and $C'$, respectively. Then the isomorphism $\theta$ induces an isomorphism of $Y$ sending $Y$ to $Y'$ and $\bar{C}$ to $\bar{C'}$. The map $\phi_{K_Y}$ is a double cover $f:Y\rightarrow Z$, where $Z$ is either a quadric cone or a smooth quadric. Thus, the isomorphism $\theta$ induces an isomorphism $\eta:Z\rightarrow Z'$. Moreover, if $(Z, B,D)$ and $(Z', B', D')$ are the triples corresponding to $X$ and $X'$ under the correspondence of Theorem~\ref{classify}, then since $\theta(C)=C'$, we have $\eta(B)=B'$ and $\eta(D)=D'$, so the triples are isomorphic.
\end{proof}
Let $p$ be a $\frac{1}{4}(1,1)$ singularity on a stable numerical quintic surface $W$, let $X$ be its minimal resolution, and let $C$ denote the $(-4)$ curve on $X$. We call $W$ a surface of type
\begin{itemize}
\item 1 if $Z=\mathbb{F}_0$, $D\sim \Delta$ and $B$ and $D$ intersect as in Figure~\ref{BandD}(d).
\item 1' if $Z=\mathbb{F}_0$, $D\sim \Delta$ and $B$ and $D$ intersect as in Figure~\ref{BandD}(e).
\item 1'' if $Z=\mathbb{F}_2$, $D\sim \Delta$ and $B$ and $D$ intersect as in Figure~\ref{BandD}(d).
\item 1''' if $Z=\mathbb{F}_0$, $D\sim \Delta$, there exists $p\in B\cap D$ with $(B\cdot D)_p=4$, and $B$ and $D$ intersect as in Figure~\ref{BandD}(f).
\item 2a if $Z=\mathbb{F}_0$, $D$ is a fiber, and $B$ and $D$ intersect as in Figure~\ref{BandD}(a).
\item 2a' if $Z=\mathbb{F}_0$, $D$ is a fiber and $B$ and $D$ intersect as in Figure~\ref{BandD}(b).
\item 2a'' if $Z=\mathbb{F}_0$, $D$ is a fiber, $B$ has an $A_2$ singularity along $D$ and $B$ and $D$ intersect as in Figure~\ref{BandD}(c).
\item 2b if $Z=\mathbb{F}_2$, $D$ is a fiber, and $B$ and $D$ intersect as in Figure~\ref{BandD}(a).
\end{itemize}
\begin{figure}[h]
\centering
\includegraphics[scale=.5]{BandD.png}
\caption{Six ways $B$ and $D$ may intersect.\label{BandD}}
\end{figure}
\begin{Lemma}\label{counting} The stable numerical quintic surfaces of types 1 and 2a correspond to $39$-dimensional loci in $\overline{\mathcal{M}}_{5,5}$. Those of types 1', 1'', 1''', 2a', 2a'', and 2b correspond to $38$-dimensional loci in $\overline{\mathcal{M}}_{5,5}$. All other types of stable numerical quintic surfaces with a unique $\frac{1}{4}(1,1)$ singularity correspond to loci of dimension less than 38.
\end{Lemma}
\begin{proof}
Lemma~\ref{triples} implies that each triple $(Z,B,D)$ of Theorem~\ref{classify} corresponds to a unique stable numerical quintic surface, up to automorphisms of $Z$. We count the dimension of such triples in the given cases. The main difficulty is to check that requiring that the branch divisor obtain different types of singularities at different points imposes independent conditions on $B$.
To create a triple $(Z, B, D)$:
\noindent 1. Fix a smooth or singular quadric $Z$.
\noindent 2. Choose a divisor $D\sim\Delta$ or $D\sim \Gamma$. Then by Riemann-Roch, since $K_Z=-2\Delta$ and $\Delta^2=2$, we have $h^0(Z, \mathcal{O}(D))=4$ if $D\sim \Delta$ and $h^0(Z, \mathcal{O}(D))=2$ if $D\sim \Gamma$. Projectivizing gives a 3-dimensional space of choices if $D\sim \Delta$ and a $1$-dimensional space if $D\sim \Gamma$.
\noindent 3. Choose $k$ points on $D$ (through which $B$ will eventually pass). Since $D\simeq \mathbb{P}^1$, we have a $k$-dimensional space of choices for these points. (If $Z$ is a cone, we can choose the $k$ points so that none of them are the singularity of $Z$.)
\noindent 4. Choose a divisor $B$:
\begin{enumerate}
\item[4a.] To obtain $D\not\subset B$, choose $B\sim6\Delta$. Again by Riemann-Roch, $h^0(Z, \mathcal{O}(B))=49$. Projectiving gives a 48 dimensional space of possible branch curves $B$.
\item[4b.] To obtain $D\subset B$, choose $B'\sim 5\Delta$. By Riemann-Roch, $h^0(Z, \mathcal{O}(B'))=36$. Projectivizing gives a $35$-dimensional space of possible branch curves $B'$. By abuse of notation, take $B=B'$ (and note that the resulting triple will be of the form $(Z, B'+D, D)$, or with our abuse of notation, $(Z, B+D, D)$).
\end{enumerate}
\noindent 5. Consider the restriction exact sequence
$$0\rightarrow \mathcal{O}_Z(B-D)\rightarrow\mathcal{O}_Z(B)\rightarrow\mathcal{O}_D(B)\rightarrow 0.$$
By Kodaira vanishing, $H^1(Z, \mathcal{O}_Z(B-D))=0$. Thus, the map
$$H^0(Z, \mathcal{O}_Z(B))\rightarrow H^0(D,\mathcal{O}_D(B))$$
is surjective, and so we can find a curve $B\in |6\Delta|$ (or $B'\in |5\Delta|$) such that the restriction of $B$ to $D$ passes through any $m$ points on $D$, counted with multiplicities, where $m=(B\cdot D)$.
Thus, the requirement that $B$ pass through the given $m$ points, counted with multiplicities, is a codimension $m$ condition.
\noindent 6. The group of automorphisms of $Z$ is $6$-dimensional if $Z$ is smooth and $7$-dimensional if $Z$ is a cone. Thus, modding out by automorphisms of $Z$ is either a codimension $6$ condition or codimension $7$ condition.
Triples $(Z, B, D)$ where $D\subset B$ give a locus of dimension at most $3+10+35-10-6=32$, so we can assume for the rest of the proof that $D\not\subset B$.
\noindent 7. So far there is no guarantee that the most general $B$ is smooth at any given point, nor is it immediate that imposing the condition that $B$ obtain a certain mild singularity at a given point does not impose conditions on $B$ at the other $k-1$ points. Provided the multiplicity at each point is small enough, the fact that these conditions \emph{are} linearly independent follows from the fact that $B$ is sufficiently big. That is, for $n\le 5$, the divisor $B-nD$ is big and nef, so the cohomology group $H^1(Z, \mathcal{O}_Z(B-nD))$ is zero by Kodaira vanishing. Thus, the map
$$H^0(Z, \mathcal{O}_Z(B))\rightarrow H^0(D,\mathcal{O}_{nD}(B))$$
induced by the restriction $\mathcal{O}_Z(B)\rightarrow\mathcal{O}_{nD}(B)$ is surjective. This means that we can choose $B$ in such a way that we can require the degree $1, 2,\ldots, n-1$ parts of the ``Taylor expansion" of its equation
$$s|_{nD}=s_0+s_1d+s_2d^2+...$$
to be of any form we desire, where $d\in H^0(Z, \mathcal{O}_Z(D))$ is the equation of $D$ and $s_i\in H^0(D, \mathcal{O}_{D}(B-iD))$.
Suppose we want to impose the condition that $B$ acquires a node at a given point $p$ for which $(B\cdot D)_p=2$. This is equivalent to requiring that the linear term in its Taylor expansion vanish at $p$, and that the discriminant of the quadratic term be non-vanishing at $p$. Therefore, this condition has expected codimension $1$. Since this is a requirement on the degree $1$ and $2$ parts of the Taylor expansion, taking $n=3$ implies that the requirements that $B$ be either smooth or obtain at most a node at each of its points are linearly independent conditions. That is, the condition that $B$ acquire a node at a point with multiplicity $2$ is indeed a codimension $1$ condition.
Similarly, the requirement that $B$ acquire an $A_2$ singularity at a point $p$ for which $(B\cdot D)_p=2$ is equivalent to requiring that the linear term in its Taylor expansion vanish at $p$, the discriminant of the quadratic term also vanish at $p$, and the cubic term be nonvanishing. Since this is a requirement on the part of the Taylor expansion of degrees $1$, $2$, and $3$, taking $n=4$ implies that $B$ acquiring an $A_2$ singularity at the desired point is a codimension $2$ condition.
Requiring $B$ to have a node at a point $p$ for which $(B\cdot D)_p=3$ is equivalent to forcing the linear term in its Taylor expansion to vanish at $p$, and the coefficient of one monomial in the quadratic term to vanish at $p$. Again, this is a requirement on the degree $2$ part of the Taylor expansion, so taking $n=3$ implies that this is a codimension $2$ condition that does not impose conditions on the other points of $B\cap D$.
Let $l$ be the dimension of the set of triples such that $|B\cap D|=k$ (set theoretically). Then
$$ l=
\begin{cases}
33+k & \mbox{if } D\in|\Delta| \mbox{ on } \mathbb{F}_0\\
32+k & \mbox{if } D\in|\Delta| \mbox{ on } \mathbb{F}_2\\
37+k & \mbox{if } D\in|\Gamma| \mbox{ on } \mathbb{F}_0\\
36+k & \mbox{if } D\in|\Gamma| \mbox{ on } \mathbb{F}_2\\
\end{cases}.
$$
Thus, if $m$ is the codimension of the set of triples such that $B$ has prescribed singularities, then in order for the set of such triples to have dimension $38$ or $39$, we have
\begin{eqnarray*}
m= k-6 \mbox{ or } m=k-5 & \mbox{if } D\in|\Delta| \mbox{ on } \mathbb{F}_0\\
m=k-6 \mbox{\hspace{1cm}} & \mbox{if } D\in|\Delta| \mbox{ on } \mathbb{F}_2\\
m=k-1 \mbox{ or } m=k-2 & \mbox{if } D\in|\Gamma| \mbox{ on } \mathbb{F}_0\\
m=k-2 \mbox{ or } m=k-3 & \mbox{if } D\in|\Gamma| \mbox{ on } \mathbb{F}_2
\end{eqnarray*}.
In particular, we see that if $D\sim \Delta$, then since $k\le 6$, we have $m=0$ or $1$. If $D\sim \Gamma$, then since $k\le4$, we have $m\le 3$.
For instance, the dimension of the locus of type 1 surfaces is $3+6+48-12-6=39$, and of type 1' surfaces is $3+6+48-12-1-6=38$.
The dimension of the locus of type 2a surfaces is $1+4+48-6-1-1-6=39$, and of type 2b surfaces is $1+4+48-6-1-1-7=38$.
Working through each of the remaining possibilities in Theorem~\ref{classify} gives the desired result.
\end{proof}
\subsection{The closures of the loci of type 1 and 2a surfaces}\label{1 and 2a closure}
The goal of this subsection is to prove the following.
\begin{Theorem}\label{types 1 and 2a are main} Let $W$ be a stable numerical quintic surface corresponding to the triple $(Z, B, D)$, and let $[W]$ denote its corresponding point in $\overline{\mathcal{M}}_{5,5}$. If $D\sim \Gamma$, then $[W]$ is in the closure of the locus of 2a surfaces. If $D\sim \Delta$, then $[W]$ is in the closure of the locus of surfaces of type 1. Thus, the closures of the loci of surfaces of types 1 and 2a contain all surfaces whose unique non Du Val singularity is a $\frac{1}{4}(1,1)$ singularity.
\end{Theorem}
We begin by showing that if $W$ is a surface corresponding to the triple $(Z, B, D)$ for $D\sim \Delta$ (respectively, $D\sim\Gamma$) then there is a family of triples $(\mathcal{Z}, \mathcal{B},\mathcal{D})$ with special fiber $(Z, B, D)$ whose general fiber is a triple corresponding to a surface of type 1 (respectively, 2a). We then take a double cover of $\mathcal{Z}$ branched over $\mathcal{B}$, followed (after a possible finite base change) by a simultaneous resolution of Du Val singularities. Ideally, this would give the minimal resolution of the desired family of stable numerical quintic surfaces, and then contracting a section of $(-4)$ curves would give the family itself. However, we may first need to do a sequence of flops in order to guarantee existence of a family of $(-4)$-curves with irreducible special fiber. The bulk of the work is showing that such a sequence exists.
\begin{Lemma}\label{triples connect} Let $(Z, B, D)$ be a triple corresponding to a stable numerical quintic surface. If $D\sim \Delta$ (respectively, $D\sim\Gamma$), then $(Z,B,D)$ is the special fiber of a family of triples $(\mathcal{Z}, \mathcal{B}, \mathcal{D})$ with general fiber a triple corresponding to a stable numerical quintic surface of type 1 (respectively, 2a).
\end{Lemma}
\begin{proof}
Suppose that $D\sim \Delta$ and consider the restriction sequence
$$0\rightarrow \mathcal{O}_Z(6\Delta-D)\rightarrow \mathcal{O}_Z(6\Delta)\rightarrow\mathcal{O}_D(12)\rightarrow 0.$$
Since $H^1(Z, \mathcal{O}_Z(6\Delta-D))=0$, the induced map
$$r: H^0(Z, \mathcal{O}_Z(6\Delta))\rightarrow H^0(D, \mathcal{O}_D(12))$$
is surjective.
Let $T\subset H^0(Z, \mathcal{O}_Z(6\Delta))$ be the locus of effective divisors with at most Du Val singularities, and let $U\subset V= H^0(D,\mathcal{O}_D(6))$ be the locus of effective divisors consisting of six distinct points. Note that $U$ is open in $V$ and $T$ is open in $H^0(Z, \mathcal{O}_Z(6\Delta))$. Let $V\rightarrow H^0(D, \mathcal{O}_D(12))$ be the map sending a section to its square. Projectivizing, we obtain an injective map $i: \mathbb{P}(V)\rightarrow \mathbb{P}(H^0(D, \mathcal{O}_D(12)))$ which sends an effective divisor $E$ to the divisor $2E$. Being a projective map, the map $i$ is a closed. Therefore,
$$\overline{r^{-1}(i(U))\cap T}\cap T=r^{-1}(i(V))\cap T.$$
That is, the locus of branch divisors with at most Du Val singularities that intersect $D$ in any six points, with even multiplicities, is the closure of the locus of branch divisors with at most Du Val singularities that intersect $D$ in six distinct points with multiplicity two at each point.
A similar argument holds for $D\sim \Gamma$, where we consider instead the restriction map
$$\mathcal{O}_Z(6\Delta)\rightarrow\mathcal{O}_D(6),$$
take $U\subset H^0(D, \mathcal{O}_D(6))$ to be the locus of effective divisors consisting of four distinct points, two of which have multiplicity two, and take $i$ to be the identity map.
\end{proof}
Suppose that $p$ is a Du Val singularity of $B$ on $D$, and let $\sum E_i \subset X$ be the exceptional divisor of $\phi$ over $p$. Let $Q=\sum \mathbb{Z}E_i$ denote the corresponding root lattice of type $A$, $D$, or $E$, with positive-definite inner product $\circ$. Note that $\circ$ is the negative of the intersection product $\cdot$ on the Picard group of $X$. That is,
$$E_i\cdot E_j =-E_i\circ E_j.$$
Let $P=\textrm{Hom}(Q, \mathbb{Z})=\sum \mathbb{Q}\omega_i$ be the corresponding weight lattice, where $\omega_i$ are the fundamental weights, satisfying $\omega_i(E_j)=\delta_{ij}$.
By means of the inner product, we identify $P$ with a lattice in $Q\otimes\mathbb{Q}$, and in this way think of $Q$ as a sublattice of $P$. In particular, we may extend in a unique way the inner product on $Q$ to $P$.
Given any effective divisor $C\in \textrm{Pic}(X)$, we can associate to $C$ the weight $\omega_C$ satisfying
$$\omega_C(E_i)=C\cdot E_i$$
for all $i$.
As an example, if $C$ is a divisor such that for some $j$ we have $C\cdot E_j = 1$ and $C\cdot E_i=0$ for $i\neq j$, then under this correspondence, we have $\omega_C=\omega_j$.
We will use the classification of Theorem~\ref{classify} to describe all ways the $(-4)$-curve $C\subset X$ intersects the exceptional divisor $\sum E_i$. We label the dual graphs of the exceptional divisor of a Du Val singularity as follows:
\begin{center}
\begin{tikzpicture}
\draw (0,0) -- (1,0);
\draw (2,0) -- (2.7,0);
\draw (1,0) -- (2,0);
\draw (3.3, 0) -- (4,0);
\draw (4,0) -- (5,0);
\node[label={1}] at (0,0) {$\bullet$};
\node at (1,0) {$\bullet$};
\node at (2,0) {$\bullet$};
\node at (4,0) {$\bullet$};
\node[label={$n$}] at (5,0) {$\bullet$};
\node at (3,0) {$\cdots$};
\node at (-.5,0) {$A_{n}$};
\draw (7,0) -- (9.7,0);
\draw (10.3, 0) -- (11,0);
\draw (11,0) -- (12,-.5);
\draw (11,0) -- (12,.5);
\node[label={1}] at (7,0) {$\bullet$};
\node at (8,0) {$\bullet$};
\node at (9,0) {$\bullet$};
\node [label={$n-2$}]at (11,0) {$\bullet$};
\node[label={$n$}] at (12,-.5) {$\bullet$};
\node[label={$n-1$}] at (12,.5) {$\bullet$};
\node at (10,0) {$\cdots$};
\node at (6.5,0) {$D_{n}$};
\end{tikzpicture}
\begin{tikzpicture}
\draw (0,0) -- (4,0);
\draw (2,0) -- (2,1);
\node[label={1}] at (0,0) {$\bullet$};
\node at (1,0) {$\bullet$};
\node at (2,0) {$\bullet$};
\node[label={6}] at (2,1) {$\bullet$};
\node at (3,0) {$\bullet$};
\node[label={5}] at (4,0) {$\bullet$};
\node at (-.5,0) {$E_6$};
\draw (7,0) -- (12,0);
\draw (9,0) -- (9,1);
\node[label={1}] at (7,0) {$\bullet$};
\node at (8,0) {$\bullet$};
\node at (9,0) {$\bullet$};
\node[label={7}] at (9,1) {$\bullet$};
\node at (10,0) {$\bullet$};
\node at (11,0) {$\bullet$};
\node[label={6}] at (12,0) {$\bullet$};
\node at (6.5,0) {$E_7$};
\end{tikzpicture}
\begin{tikzpicture}
\draw (0,0) -- (6,0);
\draw (2,0) -- (2,1);
\node[label={1}] at (0,0) {$\bullet$};
\node at (1,0) {$\bullet$};
\node at (2,0) {$\bullet$};
\node at (3,0) {$\bullet$};
\node[label={8}] at (2,1) {$\bullet$};
\node at (4,0) {$\bullet$};
\node at (5,0) {$\bullet$};
\node[label={7}] at (6,0) {$\bullet$};
\node at (-1,0) {$E_8$};
\end{tikzpicture}
\end{center}
It will be useful to have a description of the root systems $A_n$, $D_n$, and $E_n$.
For $A_n$, let $e_0, \ldots, e_n$ be an orthonormal basis of $\mathbb{R}^{n+1}$. A basis of simple roots $E_1, \ldots, E_n$ of $A_n$ is given by $E_i=e_{i-1}-e_{i}$.
For $D_n$, we let $e_1, \ldots, e_n$ be an orthonormal basis of $\mathbb{R}^n$. A basis of simple roots is given by $E_i=e_{i}-e_{i+1}$ for $i =1, \ldots n-1$ and $E_n=e_{n-1}+e_n$.
To describe the root system of $E_n$, let $S$ be a Del Pezzo surface given by the blowup of $\mathbb{P}^2$ in $n$ points in general position. Then the root lattice $E_n$ can be described as the orthogonal complement of the canonical class $K_S$ in $\textrm{Pic}(S)$, after a sign change. Let $h, e_1, \ldots e_n$ be a basis of $\textrm{Pic}(S)$, where $e_i$ are classes of the exceptional curves on $S$, and $h$ is the pullback of a hyperplane section. We note that $K_S=-3h+\sum_{i=1}^n e_i$. With respect to the pairing $\circ$ satisfying $e_i\circ e_j=\delta_{ij}$, $e_i\circ h=0$, and $h^2=-1$ (the negative of the intersection pairing on $S$), we can describe the simple roots of $E_n$ as
$$E_i=e_i-e_{i+1} \textrm{ for } i=1,\ldots, n-1$$
$$E_n=h-e_1-e_2-e_3.$$
Our main tool for proving the existence of the desired sequence of flops is the following.
\begin{Lemma}\label{why flopping helps}
Let $\mathcal{X}$ be a family of surfaces over the unit disk in $\mathbb{C}$ with special fiber $X$. Let $\mathcal{D}\subset \mathcal{X}$ be a divisor whose restriction to $X$ is an effective sum of curves $\mathcal{D}|_X=C+\sum a_iE_i\subset X$, where the $E_i$ are distinct $(-2)$-curves,
and $C\cdot E_i\ge 0$ for all $i$. Suppose there exists $j$ such that $\mathcal{D}\cdot E_j <0$. Let $\phi: \mathcal{X}'\rightarrow \mathcal{X}$ be the flop of $E_j$. Then
$$(\phi^*\mathcal{D})|_X=C+a_j'E_j +\sum_{i\neq j} a_i E_i $$
where $a_j>a_j'\ge 0$. That is, flopping the curve $E_j$ results in a divisor $\phi^*\mathcal{D}$ with at least one fewer $(-2)$-curve on the special fiber.
\end{Lemma}
\begin{proof} Suppose that $a=\mathcal{D}\cdot E_j$ is negative. Flopping $E_j$ does not change the coefficients $a_i$ for $i\neq j$, and so
$$(\phi^*\mathcal{D})|_X=(C+a_j'E_j+\sum_{i\neq j} a_i E_i)$$
for some nonnegative integer $a_j'$. Moreover, because $\mathcal{D}\cdot E_j<0$, flopping $E_j$ gives $\phi^*\mathcal{D}\cdot E_j> 0$. Thus
$$(C+a_j'E_j+\sum_{i\neq j} a_i E_i)\cdot E_j = a-(-2a_j)+(-2a_j')> 0,$$
so
$$a+2(a_j-a_j')> 0.$$
Since $a<0$, we have that $a_j> a_j'$.
\end{proof}
\begin{Lemma}\label{the odd case}
Let $(\mathcal{Z}, \mathcal{B},\mathcal{D})$ be a family of triples over the unit disk in $\mathbb{C}$, with general fiber corresponding to a stable numerical quintic surface of type 2a or 2b and such that $D=\mathcal{D}_0$ is a ruling. Let $\mathcal{Y}$ be the double cover of $\mathcal{Z}$ branched over $\mathcal{B}$. Then there exists, after a possible finite base change, a simultaneous resolution of singularities $\mathcal{X}\rightarrow\mathcal{Y}$ such the closure of the $(-4)$-curve on the general fiber is irreducible.
\end{Lemma}
\begin{proof}
Let $(Z, B, D)$ be a triple corresponding to a stable numerical quintic surface, and suppose that $D$ is a ruling. Let $(\mathcal{Z}, \mathcal{B}, \mathcal{D})$ be a family of triples whose general fiber is of type 2a or 2b, and whose special fiber is the triple $(Z, B, D)$. Consider the maps
\[\xymatrix{
\mathcal{X} \ar[r]^{\psi} & \mathcal{Y}\ar[r]^f& \mathcal{Z}
}\]
where $f:\mathcal{Y}\rightarrow \mathcal{Z}$ is the double cover branched over $\mathcal{B}$, and $\psi:\mathcal{X}\rightarrow \mathcal{Y}$ is any simultaneous resolution of Du Val singularities, which exists after a possible finite base change. Let $\mathcal{C}$ be the closure of the $(-4)$-curve on the general fiber of $\mathcal{X}$. Recall that $\mathcal{B}$ has two nodes lying on the general fiber of $\mathcal{D}$, and thus the general fiber of $\mathcal{X}$ contains two $(-2)$ curves intersecting the general fiber of $\mathcal{C}$. Let $\mathcal{E}_1$ and $\mathcal{E}_2$ be the closures of these $(-2)$-curves.
We claim that there exists a sequence of flops of $\mathcal{X}$ so that $\mathcal{C}$ has irreducible special fiber.
Let $C$ be the $(-4)$-curve on $X=\mathcal{X}_0$ and let $\sum E_i\subset X$ exceptional divisor over the singularities of $B$ on $D$. Then
$$\psi^{-1}(f^{-1}(\mathcal{D}))_0=C+\sum a_i E_i +\sum b_i E_i +\sum c_i E_i,$$
where
$$C+\sum a_i E_i=\mathcal{C}_0, \; \; \; \sum b_iE_i=(\mathcal{E}_1)_0, \; \; \; \sum c_iE_i=(\mathcal{E}_2)_0,$$
and the $a_i, b_i, c_i$ are nonnegative integers. Since the general fibers of $\mathcal{E}_1$ and $\mathcal{E}_2$ are disjoint $(-2)$-curves, the divisors $\sum b_iE_i$ and $\sum c_i E_i$
satisfy $(\sum b_iE_i)\cdot(\sum c_i E_i)=0$ and $(\sum b_iE_i)^2=(\sum c_i E_i)^2=-2$. Moreover
\begin{equation}\label{C + stuff is weight 0}
(C+\sum a_i E_i +\sum b_i E_i +\sum c_i E_i)\cdot E_j =0 \textrm{ for all $j$.}
\end{equation}
We pass to the weight lattice $\Lambda$ of the corresponding singularity (or singularities) on $Y=\mathcal{Y}_0$. Note that if $B$ has one singularity on $D$, then $\Lambda$ is irreducible. But if $B$ has two singularities, then the lattice $\Lambda$ is a direct sum $P\oplus P'$ of irreducible lattices, each corresponding to a singularity of $B$ on $D$. In either case, Equation~\eqref{C + stuff is weight 0} implies that the weight $$\omega_C-\sum a_i E_i -\sum b_i E_i -\sum c_i E_i\in \Lambda$$ is equal to $0$. Thus, as weights, we have
$$\omega_C-\sum a_i E_i =\sum b_i E_i +\sum c_i E_i.$$
We prove that if $(C+\sum a_i E_i)\cdot E_j \ge 0$ for all $j$, then $a_i=0$ for all $i$. Together with this, Lemma~\ref{why flopping helps} implies that if $a_i\neq 0$ for some $i$, then there exists a flop of $\mathcal{X}$ after which $\mathcal{C}_0$ will contain at least one fewer $(-2)$-curve.
Suppose that for all $j$, we have $(C+\sum a_i E_i)\cdot E_j \ge 0$. Passing again to the weight lattice, this is the statement that
$$\left(\omega_C-\sum a_i E_i\circ E_j\right)\ge 0$$
for all $j$. The last inequality is equivalent to the statement that the weight
$$\omega_C-\sum a_i E_i=\sum b_i E_i +\sum c_i E_i$$
is dominant. We therefore have
\begin{eqnarray*}
\omega_C^2&=&\left(\sum b_i E_i +\sum c_i E_i\right)^2+2\left(\sum b_i E_i +\sum c_i E_i\right)\circ\left(\sum a_i E_i\right) +\left(\sum a_i E_i\right)^2\\
&=& 4 +2\left(\sum b_i E_i +\sum c_i E_i\right)\circ\left(\sum a_i E_i\right) +\left(\sum a_i E_i\right)^2\\
&\ge& 4,
\end{eqnarray*}
where equality holds if and only if $a_i=0$ for all $i$. Thus, our proof is complete once we show that $\omega_C^2=4$ independent of the singularity (or singularities) of $B$ on $D$.
By Theorem~\ref{classify}, and referring to Figures~\ref{distinct points}, \ref{same points 4}, and~\ref{same points 5}, we have the following table detailing how $C$ intersects the curves $E_i$. We write $\omega_C|_P$ to denote the weight $\omega_C$ restricted to the irreducible lattice $P$ corresponding to the singularity of $B$, and note that if $\Lambda=P$ is irreducible (i.e., if $B$ has only one singularity on $D$), then $\omega_C|_{\Lambda}=\omega_C|_{P}=\omega_C$. Otherwise, $\omega_C=\omega_C|_{P}+\omega_C|_{P'}$ where $\omega_C|_{P}$ and $\omega_C|_{P'}$ are orthogonal weights in $\Lambda$, each corresponding to a singularity of $B$ on $D$.
\begin{table}[h!]\label{odd case table}
\begin{tabular}{c|c|c|c}
$(B\cdot D)_p$ & Singularity & Nonzero intersections $C\cdot E_i$ & $\omega_C|_P$\\
\hline
2 & $A_n$ & $C\cdot E_1=C\cdot E_n=1$ & $e_0-e_n$\\
\hline
3 & $A_1$ & $C\cdot E_1=1$ & $e_0-e_1$\\
3 & $A_2$ & $C\cdot E_1=C\cdot E_2 =1$&$e_0-e_2$ \\
3 & $D_n$ & $C\cdot E_2=1$ & $e_1+e_2$\\
3 & $E_6$ & $C\cdot E_6 =1$ & $-2h+\sum_{i=1}^6e_i$ \\
3 & $E_7$ & $C\cdot E_1=1$ & $2h-\sum_{i=1}^7e_i$\\
3 & $E_8$ & $C\cdot E_7=1$ & $3h-2e_8-\sum_{i=1}^7e_i$\\
\hline
4 & $A_n$, $n\ge4$ & $C\cdot E_2=C\cdot E_{n-1}=1$ & $e_0+e_1-e_{n-1}-e_n$ \\
4 & $D_5$ & $C\cdot E_4=C\cdot E_{5}=1$ & $e_1+e_2+e_3+e_4$\\
4 & $D_n$ & $C\cdot E_1=2$ & $2e_1$\\
4 & $E_6$ & $C\cdot E_1=C\cdot E_{5}=1$ & $2h+2e_1+\sum_{i=2}^5$ \\
\hline
5 & $A_3$ & $C\cdot E_2=1$ & $e_0+e_1-e_2-e_3$\\
5 & $A_4$ & $C\cdot E_2=C\cdot E_3=1$ & $e_0+e_1-e_3-e_4$ \\
5 & $D_n$, $n\ge 6$ & $C\cdot E_4=1$ & $e_1+e_2+e_3+e_4$\\
5 & $D_n$ & $C\cdot E_1=2$ & $2e_1$\\
5& $E_7$ & $C\cdot E_5=1$ & $3h-2e_6-2e_7-\sum_{i=1}^5e_i$\\
5 & $E_8$ & $C\cdot E_1=1$ & $5h-e_1-2\sum_{i=2}^8e_i$\\
\end{tabular}
\end{table}
One quickly checks that if $(B\cdot D)_p$ is $2$ or $3$, then $(\omega_C|_P)^2=2$ and if $(B\cdot D)_p$ is $4$ or $5$, then $(\omega_C|_P)^2=4$. In the former case, $\omega_C$ is the sum of two perpendicular vectors $\omega_C|_P$ and $\omega_C|_{P'}$, and thus
$$\omega_C^2=(\omega_C|_P + \omega_C|_{P'})^2=4.$$ This completes the proof.
\end{proof}
\begin{Remark}
The proof of Lemma~\ref{the odd case} relies on the fact that in each case, the weight $\omega_C$ is dominant with square $4$. It is interesting to note that the vectors listed in Table~\ref{odd case table} are in fact all vectors with square $4$ or $2$ corresponding to dominant weights, up to the action of the corresponding Weyl group.
\end{Remark}
In the case where $D\sim \Delta$ is a diagonal, have a more general statement. The proof will require the following.
\begin{Lemma}\label{from rep theory}~\cite[VIII, 7.3]{bourbaki} Let $\omega$ be a weight and suppose that $\omega-\sum b_j E_j$ is dominant. Then the weight $\omega-\sum b_j E_j$ is a (dominant) weight of the irreducible representation $V_{\omega}$ of highest weight $\omega$ of the corresponding Lie algebra.
\end{Lemma}
A weight $\omega$ is \emph{minuscule} if it is the only dominant weight of $V_\omega$.
\begin{Theorem}\label{simultaneous resolution for even intersection} Let $Z$ be a smooth surface, $B$ a divisor on $Z$ with at most Du Val singularities, and $D$ a smooth irreducible divisor on $Z$. Let $(\mathcal{Z,B,D})$ be a family of triples over the unit disk in $\mathbb{C}$, with special fiber $(Z, B, D)$, and such that the divisors $\mathcal{D}_t$ and $\mathcal{B}_t$ are reduced, irreducible and smooth for $t\neq 0$. Suppose that at each point $p\in \mathcal{D}_t\cap \mathcal{B}_t$ over the general fiber, the local intersection $(\mathcal{D}_t\cdot\mathcal{B}_t)_p$ is even. Let $f:\mathcal{Y}\rightarrow \mathcal{Z}$ be the double cover branched over $\mathcal{B}$. Then there exists, after a possible finite base change, a simultaneous resolution of singularities $\psi:\mathcal{X}\rightarrow \mathcal{Y}$ such that the closure of one of the two components of $\psi^{-1}(f^{-1}(\mathcal{D}))_t$ over the general fiber has irreducible special fiber.
\end{Theorem}
\begin{proof}
We can work locally, so we assume that $B$ has a unique singularity at $p\in B\cap D$. Let
$f:\mathcal{Y}\rightarrow \mathcal{Z}$ be the double cover branched over $\mathcal{B}$ and $\psi:\mathcal{X}\rightarrow\mathcal{Y}$ a simultaneous resolution of Du Val singularities of $\mathcal{Y}$, which exists after a possible finite base change. Let $\mathcal{C}_1$ and $\mathcal{C}_2$ be the closures in $\mathcal{X}$ of the two components of $\psi^{-1}(f^{-1}(\mathcal{D}))$ over the general fiber.
Note that
$${f}^{-1}(\psi^{-1}(\mathcal{D}))_0=\left(C_1+\sum a_i E_i\right) +\left(C_2+\sum b_i E_i\right)$$
where $\sum E_i\subset X$ is the exceptional divisor over $p$, $C_1+\sum a_i E_i$ is the special fiber of $\mathcal{C}_1$, $C_2+\sum b_i E_i$ is the special fiber of $\mathcal{C}_2$,
and the $a_i$ and $b_i$ are nonnegative integers.
If the special fiber of either $\mathcal{C}_1$ or $\mathcal{C}_2$ is irreducible, that is, if either all $a_i$ or all $b_i$ are zero, then we are done. Otherwise, we claim that there exists a sequence of flops $\phi:\tilde{\mathcal{X}}\rightarrow\mathcal{X}$ so that $\phi^*(\mathcal{C}_1)$ has irreducible special fiber. As in the proof of Lemma~\ref{the odd case}, we show that if $(C_1+\sum a_i E_i)\cdot E_j\ge0$ for all $j$, then $a_i=0$ for all $i$. We then apply Lemma~\ref{why flopping helps} to prove the claim.
Suppose that $(C_1+\sum a_i E_i)\cdot E_j\ge0$ for all $j$. Passing to the weight lattice, this is equivalent to the statement that the corresponding weight is dominant. By Lemma~\ref{from rep theory}, we have that $\omega=\omega_{C_1}-\sum a_i E_i$ is a dominant weight of the irreducible representation $V_{\omega_{C_1}}$. If $\omega_{C_1}$ is minuscule, then this implies that $a_i=0$ for all $i$ and we are done.
We refer the reader to Figure~\ref{even case resolution}, which details all possible ways that $B$ and $D$ may intersect at $p$ together with how $\tilde{D}$ intersects the exceptional curves on $\tilde{Z}$ and the branch divisor of $\tilde{f}$. This gives the following table of weights $\omega_{C_1}$, depending on the singularity of $B$ at $p$.
\begin{center}
\begin{tabular}{c|c}
Singularity & $\omega_{C_1}$\\
\hline
$A_n$ & $\omega_i$ for any $i$ \\
$D_n$ & $\omega_1, \omega_{n-1}, \omega_n$ \\
$E_6$ & $\omega_1, \omega_5$ \\
$E_7$ & $\omega_6$
\end{tabular}
\end{center}
One can check that these are all minuscule fundamental weights.
\end{proof}
\begin{Remark} The proof of Theorem~\ref{simultaneous resolution for even intersection} relies on the definition of a minuscule fundamental weight. It is interesting to note that the table at the end of the proof contains all minuscule fundamental weights of the simple Lie algebras.
\end{Remark}
We can now prove the main theorem of this section.
\begin{proof}[Proof of Theorem~\ref{types 1 and 2a are main}]
Let $W$ be a stable numerical quintic surface corresponding to the triple $(Z, B, D)$, and let $[W]$ denote its corresponding point in $\overline{\mathcal{M}}_{5,5}$. By Lemma~\ref{triples connect}, there exists a family of triples $(\mathcal{Z, B, D})$ over the unit disk in $\mathbb{C}$ such that $(\mathcal{Z, B, D})_0=(Z, B, D)$ and if $D\sim \Gamma$ (respectively, $D\sim \Delta$), then the general fiber of $(\mathcal{Z, B, D})$ is a triple corresponding to a stable numerical quintic surface of type 2a (respectively, type 1).
Let $\mathcal{Y}$ be the double cover of $\mathcal{Z}$ branched over $\mathcal{D}$, and let $\tilde{\mathcal{X}}\rightarrow \mathcal{Y}$ be a simultaneous resolution of singularities of $\mathcal{Z}$, which exists after a possible finite base change. Let $\mathcal{C}$ be the closure of (one of the) $(-4)$-curve(s) on the general fiber of $\tilde{\mathcal{X}}$. By Lemma~\ref{the odd case} and Theorem~\ref{simultaneous resolution for even intersection}, there exists a sequence of flops $\phi:\mathcal{X}\rightarrow\tilde{\mathcal{X}}$ so that $\phi^*(\mathcal{C})$ has irreducible special fiber. Contracting $\phi^*(\mathcal{C})$ results in the desired family of stable numerical quintic surfaces.
\end{proof}
\section{Deformations of surfaces of types 1 and 2a}\label{wahldiv}
We describe the components of $\overline{\mathcal{M}}_{5,5}$ corresponding to surfaces of types 1 and 2a and show that their closures are generically Cartier divisors in the boundary of the type I and IIa components of $\mathcal{M}_{5,5}$. In Lemma~\ref{counting}, we showed that these components are both $39$-dimensional.
In~\ref{families}, we construct explicit $\mathbb{Q}$-Gorenstein families of numerical quintic surfaces to show that these components are in the boundary of the respective components on $\mathcal{M}_{5,5}$. In~\ref{sheaf calculations}, we prove a number of technical results which we use in~\ref{1 and 2a} to show that the closures $\bar{1}$ and $\overline{\mbox{2a}}$ of these components are generically Cartier divisors. In particular, we show that the cohomology groups controlling obstructions to $\mathbb{Q}$-Gorenstein deformations of surfaces of types 1 and 2a vanish.
\subsection{Families of stable quintic surfaces}\label{families}
\subsubsection{Type 1}\label{Type 1}
We describe a family of quintic surfaces degenerating to a stable numerical quintic surface of type 1.
\begin{Theorem}\label{quintic example}
Consider the family $(\mathcal{X}, T)$ of surfaces
$$S_t=\{q^2l+tqf+t^2g=0\}\subset \mathbb{P}^3 \times T_t$$
where $T$ is the unit disk in $\mathbb{C}$ and $f$ and $g$ are forms of degrees $3$, and $5$, respectively, such that
\begin{itemize}
\item $S_t$ is a smooth quintic surface for $t\in T^*$
\item $S_0$ is the union of a smooth double quadric $Z$ given by $q=0$ and a plane $L$ given by $l=0$ intersecting transversally.
\end{itemize}
For general $f$ and $g$, the KSBA stable limit of the family $(\mathcal{X},T)$ is a stable numerical quintic surface of type 1. Conversely, any stable numerical quintic surface of type $1$ is the stable limit of such a family.
\end{Theorem}
\begin{proof} The singular locus of $\mathcal{X}$ is the surface $Z$, so $\mathcal{X}$ is not normal. To compute the stable limit we first normalize the family. After normalization and an extremal contraction, we will see that the family of surfaces obtained has reduced special fiber and ample canonical class.
Let $\nu:\mathcal{X}^{\nu}\rightarrow \mathcal{X}$ be the normalization of $\mathcal{X}$. We determine the structure of $\mathcal{X}^{\nu}$. First note that the normalization is an isomorphism away from $Z$.
Let $U$ be a complex analytic neighborhood in $\mathcal{X}$ of a point $p \in Z$. Then on $U$, we can write
$$q|_U=q_1+q_2, \;\; l|_U=l_0+l_1, \;\; f|_U=\sum_{i=0}^3 f_{i}, \;\; g|_U=\sum_{i=0}^5 g_{i}$$
where the subscripts indicate the degree of each term in linear coordinates centered at $p$. Giving $t$ weight 1, we can write the equation of $\mathcal{X}\cap U$ as
$$q_1^2l_0+tq_1f_{0}+t^2g_{0}+\textrm{higher order terms}.$$
Let $B\subset Z$ be the ``discriminant curve" given by $\{f^2-4lg=0\}\subset Z\cap U$. If $p\not \in B$, then the equation of $\mathcal{X}\cap U$ factors into the product of two linear terms which are not equal. That is, $(p\in\mathcal{X})$ is locally analytically isomorphic to a threefold $\mathcal{Y}=(xy=0)\subset \mathbb{A}^4$. Thus, over the open set $Z\backslash B $, the special fiber $\mathcal{X}_0^\nu$ is an unramified double cover of $Z\backslash B$.
Now consider a point $p\in B$. The equation of $\mathcal{X}\cap U$ may be written locally analytically as
$$
h=\begin{cases}
(q+\frac{1}{2}f_{0} t)^2+\textrm{h.o.t.} & \mbox{if } p\not\in L\\
t^2+\textrm{h.o.t.} & \mbox{if } p\in L
\end{cases}
$$
Thus, in order to determine the structure of $\mathcal{X}^{\nu}$ near $p$, we must consider the degree three part of $h$:
$$h_3=q_1^2l_1+2q_1q_2l_0+tq_1f_{1}+tq_2f_{0}+t^2g_{1}.$$
Suppose first that $p\not\in L$. Then we may assume that $l_0=1$ and complete the square in the first few terms of $h$:
\begin{eqnarray*}
h&=&(q_1+\frac{1}{2}tf_{0})^2 + 2q_2(q_1+\frac{1}{2}tf_{0}) +q_1^2l_1+tq_1f_{1} + t^2g_{1} +\textrm{h.o.t.}\\
&=&(q_1+\frac{1}{2}tf_{0} +q_2)^2 +q_1^2l_1+tq_1f_{1} +t^2g_{1}+\textrm{h.o.t.}
\end{eqnarray*}
Let $y=q_1+\frac{1}{2}tf_{0}$ and note that $y$ is a linear form. This last equation now becomes
$$h=(y +q_2)^2 +y^2\alpha+yt\beta +t^2\gamma+\textrm{h.o.t.}$$
where
$$\alpha=l_1,$$
$$\beta=f_{1}-l_1f_{0},$$ and
$$\gamma=g_{1}-\frac{1}{2}f_{0}(f_{1}+\frac{1}{2}l_1)$$ are linear forms. Finally we can rewrite this as
\begin{eqnarray*}
h&=&(y +q_2)^2 +(y+q_2)(y\alpha+t\beta) -q_2(y\alpha+t\beta)+t^2\gamma + \textrm{h.o.t.}\\
&=&[(y+q_2)+\frac{1}{2}(y\alpha+t\beta)]^2+t^2\gamma+ \textrm{h.o.t.}\\
&=& z^2+t^2\gamma +\textrm{h.o.t.}
\end{eqnarray*}
where $z$ is a linear form. Thus, in a complex analytic neighborhood of any point $p\in B\backslash L$, the threefold $\mathcal{X}$ is locally analytically isomorphic to the threefold $\mathcal{Y}=\{ z^2-t^2\gamma=0\}\subset\mathbb{A}^4_{\gamma, t, z, s}$ which is the product of $\mathbb{A}^1$ with the Whitney umbrella, or pinch point. The normalization of $\mathcal{Y}$ is $\mathbb{A}^3_{u,v,w}$ with normalization map $(u,v,w)\mapsto(u^2, v, uv, w)$.
Here, the quadric $Z$ corresponds to the locus $(z=t=0)\subset\mathcal{Y}$, so away from $L$, the normalization $\mathcal{X}_0^{\nu}$ of $\mathcal{X}_0$ is the double cover of the smooth quadric $Z$, ramified along the discriminant curve $B$.
Since $B\subset Z$ is a curve of degree 12, the surface $\mathcal{X}_0^{\nu}$ is the double cover of $Z\simeq \mathbb{P}^1\times\mathbb{P}^1$,
ramified along a divisor in the linear system $|6\Delta|$.
Next, we consider a point $p\in L$. We begin by assuming that $p\in L\cap Z\backslash B$. Then $l_0=0$ and $f_{0}\neq 0$, so we can assume that $f_{0}=1$ and we have
$$h=tq_1+t^2g_{0}+q_1^2l_1+tq_1f_{1}+tq_2+t^2g_{1}+ \textrm{ h.o.t.}.$$
By choosing $g$ sufficiently general, we can assume that $g_{0}\neq 0$ and so take $g_{0}=1$.
Thus, $h$ factors as
\begin{eqnarray*}
h&=& tq_1+t^2+q_1^2l_1+tq_1f_{1}+tq_2+t^2g_{1}+ \textrm{ h.o.t.}\\
&=& (t+q_1l_1+ \textrm{ h.o.t.} )\cdot(t+q_1 -q_1l_1+\textrm{ h.o.t.})
\end{eqnarray*}
The linear term of each factor is unique up to multiplication by a nonzero constant. In particular, the second factor does not vanish identically along $L$. Since $h(p)=0$ the first term must vanish along $L$. Thus, the normalization of $(p\in \mathcal{X})$ is an unramified double cover of $Z\backslash B$, of which one component (the component corresponding to the first factor of $g$ above) contains the entire proper transform of $L\backslash B$.
For the six points $p\in L\cap B$, we have $l_0=0$ and $f_{0}=0$. We suppose first that $p$ is a smooth point of $B$. Then we can assume that $g_{0}=1$ and so we can write the local equation of $\mathcal{X}$ as
$$h=t^2+tq_1f_{1}+q_1^2l_1+t^2g_{1}+\textrm{ h.o.t.}$$
Completing the square gives
$$h=(t+\frac{1}{2}q_1f_{1})^2+q_1^2l_1+t^2g_{1}+\textrm{ h.o.t.}$$
Let $\alpha=t+\frac{1}{2}q_1f_{1}$ and note that we can write $t=\alpha-\frac{1}{2}q_1f_{1}$. Then $h$ can be rewritten in terms of $\alpha$ as
\begin{eqnarray*}
h&=&\alpha^2+q_1^2l_1+(\alpha-\frac{1}{2}q_1f_{1})^2g_{1}+\textrm{ h.o.t.}\\
&=& \alpha^2(1+g_{1})+q_1^2l_1+\textrm{ h.o.t.}\\
&=& y^2+q_1^2l_1+\textrm{ h.o.t.}
\end{eqnarray*}
Thus, the threefold $\mathcal{X}$ is again locally analytically isomorphic to the threefold $\mathcal{Y}=\{y^2-x^2z=0\}\subset\mathbb{A}^4_{x,y,z, s}$ which is the product of $\mathbb{A}^1$ with the Whitney umbrella. The normalization of $\mathcal{Y}$ is $\mathbb{A}^3_{u,v,w}$ with normalization map $(u,v,w)\mapsto(u, uv, v^2, w)$. In the coordinates of $\mathbb{A}^4_{x,y,z,s}$ the plane $L$ corresponds to the plane $P=(z=y=0)\subset \mathcal{Y}$. Because the normalization is an isomorphism over this locus, we have $P^{\nu}$ is the plane given by $v=0$. The quadric $Z$ corresponds to the locus $(x=y=0)\subset \mathcal{Y}$, which under the normalization becomes the plane $u=0$. Thus, we see that the proper transforms $L^{\nu}$ and $Z^{\nu}$ of $L$ and $Z$ intersect transversally after the normalization.
The plane $L$ intersects the quadric $Z$ in a conic $D$. Thus, for general $q$, $l$ and $B$, the curve $D=L\cap Z$ intersects the locus $B\cap Z$ tangentially at 6 points. Taking the double cover of $Z$ branched over $B$ gives a smooth surface $\tilde{W}$ with a smooth $(-4)$-curve $C$ given by the intersection of the plane $L$ with the surface $\mathcal{X}_0^{\nu}$.
We now show that an extremal contraction of $L^{\nu}$ results in a family of surfaces with ample canonical class. The canonical class $K_{X_0}$ is given by $K_{\mathcal{X}^{\nu}}|_{X_0}$. Since $K_{\mathcal{X_0}^{\nu}}|_{\tilde{W}}= K_{\tilde{W}}+C$ and
$$K_{\mathcal{X}^{\nu}}|_{L}=K_{L}+C\sim -2H+H\sim -H,$$
we see that $L\subset \mathcal{X}^{\nu}$ can be contracted and that the surface $W$ obtained after contracting $C\subset \tilde{W}$ gives the stable limit. Note moreover that $C$ is a $(-4)$-curve on $\tilde{W}$, so this contraction produces a $\frac{1}{4}(1,1)$ singularity on $W$. By construction, the stable limit of the family is a stable numerical quintic surface $W$ with a $\frac{1}{4}(1,1)$ singularity of type 1 if $B$ is smooth, and of type 1' if $B$ has a node on $L\cap Z$.
We claim that any stable numerical quintic surface of type 1 may be obtained as the stable limit of such a family. By Lemma~\ref{triples}, it suffices to show that given any triple $(Z, B, D)$ (where $Z$ is a fixed smooth quadric, $B\sim 6\Delta$ and $D\sim\Delta$ are smooth, and such that $B$ intersects $D$ with multiplicity $2$ at $6$ points) we can find a family of the desired form whose stable limit is a stable numerical quintic surface $W$ corresponding to $(Z, B, D)$ under the correspondence of Theorem~\ref{classify}.
Fix such a triple. Then $Z$ is isomorphic to a smooth quadric in $\mathbb{P}^3$ given by $q=0$. Let $l$ be the equation of the hyperplane $L$ in $\mathbb{P}^3$ such that $L\cap Z=D$. We claim that $B$ is also given by $V\cap Z$, where $V$ is a hypersurface of degree $6$ in $\mathbb{P}^3$. To see this, let $H$ be a general hyperplane section of $\mathbb{P}^3$ and consider the exact sequence
$$0\rightarrow\mathcal{O}_{\mathbb{P}^3}(-Z+6H)\rightarrow \mathcal{O}_{\mathbb{P}^3}(6H)\rightarrow\mathcal{O}_Z(6H)\rightarrow 0.$$
Since $H^1(\mathbb{P}^3, \mathcal{O}_{\mathbb{P}^3}(-Z+6H))=H^1(\mathbb{P}^3, \mathcal{O}_{\mathbb{P}^3}(4H))=0$, we see that global sections of $\mathcal{O}_{\mathbb{P}^3}(6H)$ surject onto global sections of $\mathcal{O}_Z(6H)$. Noting that $\mathcal{O}_Z(6H)\simeq\mathcal{O}_Z(6\Delta)$, this implies that the element $B\in |6\Delta|$ can be lifted to a hypersurface $V$ of degree $6$ in $\mathbb{P}^3$, proving the claim.
Next consider the exact sequence
$$0\rightarrow \mathcal{O}_Z(V-L)\rightarrow\mathcal{O}_Z(V)\rightarrow \mathcal{O}_{Z\cap L}(V)\rightarrow 0.$$
Since $B$ intersects $D$ at $6$ points with multiplicity $2$ each, this implies that the equation of $V|_{L}$ is of the form $f^2$, where the six points of $B\cap D$ are given by $f=q=0$. Therefore $V$ can be chosen to have equation $f^2-lg$, where $g$ is a general form of degree $5$. Then taking
$$S_t=\{q^2l+tqf+t^2g=0\}\subset\Delta_t\times \mathbb{P}^3$$
gives the desired family.
\end{proof}
\begin{Remark} We remark that the family given in Theorem~\ref{quintic example} is one case of a more general degeneration of Castelnuovo surfaces (minimal surfaces of general type with $K^2=2p_g-7$ whose canonical maps are birational onto their images) described by Ashikaga and Konno~\cite[2.3]{ashikaga-konno1991}. Indeed, the family they give is the minimal resolution of the $\mathbb{Q}$-Gorenstein family we described in the proof of Theorem~\ref{quintic example}.
\end{Remark}
\subsubsection{Types 2a and 2b}\label{Types 2a and 2b}
Friedman~\cite{friedman1983} constructed a family of stable numerical quintic surfaces with general fiber a numerical quintic surface of type IIb and special fiber a stable numerical quintic surface of type 2b. His construction easily generalizes to give a family of stable numerical quintic surfaces whose general fiber is a numerical quintic surface of type IIa and with special fiber a stable numerical quintic surface of type 2a.
\begin{Theorem}\label{Friedman}\cite{friedman1983} There is a $\mathbb{Q}$-Gorenstein deformation $\mathcal{X}\rightarrow T$, where $T$ is the unit disk in $\mathbb{C}$, with general fiber $\mathcal{X}_t$, $t\neq 0$, a smooth numerical quintic surface of type IIa (respectively, IIb) and special fiber $\mathcal{X}_0$ a stable numerical quintic surface with a $\frac{1}{4}(1,1)$ singularity of type 2a (respectively, 2b).
Furthermore, this deformation induces a versal local $\mathbb{Q}$-Gorenstein deformation of a $\frac{1}{4}(1,1)$ singularity.
\end{Theorem}
We omit the proof of Theorem~\ref{Friedman}, but make two important observations. The first is that Friedman's construction is a degeneration of a family of IIb surfaces to a 2b surface. The construction of a family of IIa surfaces degenerating to a 2a surface is similar. For details, see \cite{mythesis}.
Secondly, we remark that Friedman's family induces a versal local $\mathbb{Q}$-Gorenstein deformation of the $\frac{1}{4}(1,1)$ singularity on the special fiber. To see this, note that if
$(p\in W)$ is a germ of a $\frac{1}{4}(1,1)$ singularity, then $(p\in W)$ is analytically isomorphic to the singularity
$$(xy=z^2)\subset \frac{1}{2}(1,1,1).$$
Moreover, any deformation of $(p\in X)$ is analytically isomorphic to a deformation of the form
$$(xy=z^2+t^{\alpha})\subset \frac{1}{2}(1,1,1)\times\mathbb{A}^1_t,$$
for some integer $\alpha>0$ called the \emph{axial multiplicity} of the deformation. The resolution of the total space of such a deformation consists of two components intersecting with multiplicity $\alpha$. A versal local $\mathbb{Q}$-Gorenstein deformation of $(p\in X)$ has axial multiplicity $1$; that is, its resolution consists of two components meeting transversally.
The special fiber of Friedman's family consists of two components meeting transversally and therefore induces a versal local $\mathbb{Q}$-Gorenstein deformation of the $\frac{1}{4}(1,1)$ singularity.
\begin{Remark} In~\cite[Corollary 1.2]{friedman1983}, Friedman uses Horikawa's description of the moduli space $\mathcal{M}_{5,5}$ to deduce the existence of a $\mathbb{Q}$-Gorenstein family $\tilde{\mathcal{X}}\rightarrow T$ of smooth quintic surfaces whose special fiber is an ``accordion'' of surfaces $V\cup W_1 \cup W_2 \cup \cdots \cup W_n$ where $V$ is the minimal resolution of a stable quintic surface of type 2b, $W_1, \ldots, W_{n-1}$ are copies of $\mathbb{F}_4$, and $W_n$ is a copy of $\mathbb{P}^2$, intersecting transversally as in Figure~\ref{non versal smoothing}.
\begin{figure}[h!]
\centering
\includegraphics{nonversal.png}
\caption{\label{non versal smoothing}}
\end{figure}
Here, the canonical class $K_{\tilde{\mathcal{X}}}$ is not ample, and the stable limit of $\tilde{\mathcal{X}}\rightarrow T$ is obtained by contracting the surfaces $W_1,..., W_n$. We now recognize the resulting special fiber as a stable numerical quintic surface of type 2b. Thus, Friedman's family is a $\mathbb{Q}$-Gorenstein smoothing of a 2b surface to a quintic surface. This family gives a local deformation of the $\frac{1}{4}(1,1)$ singularity with axial multiplicity $n$, so unless $n=1$, the induced deformation is not versal. Theorem~\ref{the best} in Section~\ref{main proof} implies that if $W$ is a 2b surface, then there exists a $\mathbb{Q}$-Gorenstein smoothing of $W$ to a quintic surface with $n=1$.
Friedman also raises the question of describing deformations of 2b surfaces explicitly. Theorem~\ref{the best} answers this question.
\end{Remark}
\subsection{Some sheaf calculations}\label{sheaf calculations}
Let $X$ be a smooth surface and $D=\sum_{i=1}^k D_i$ a divisor in $X$ with simple normal crossings (in particular, each component divisor $D_i$ is smooth). Let $\Omega_X^1(\log D)$ denote the sheaf of logarithmic differentials. This sits in the short exact sequence of sheaves
$$0\rightarrow \Omega_X^1\rightarrow \Omega_X^1(\log D)\rightarrow \bigoplus_{i=1}^k\mathcal{O}_{D_i} \rightarrow 0$$
where the map $\Omega_X^1(\log D)\rightarrow \bigoplus_{i=1}^k\mathcal{O}_{D_i}$ is the residue map.
Now let $W$ be a surface whose only non Du Val singularity is a Wahl singularity and let $X$ be its minimal resolution. If $C$ is the exceptional divisor on $X$, then one can show that obstructions to $\mathbb{Q}$-Gorenstein deformations of $W$ lie in the cohomology group $H^2(X, T_X(\log C))$~\cite{lee-park2007}. Thus, if $H^2(X, T_X(\log C))=0$, then the locus of such surfaces is generically smooth in $\overline{\mathcal{M}}_{K^2, \chi}$. The calculations of $H^2(X, T_X(\log C))$ in Theorems~\ref{vanishing for type 1}, \ref{vanishing for type 2a}, and~\ref{obstruction} require the following lemmas.
\begin{Lemma}\label{blowup of two points} Let $\sigma: Y\rightarrow Z$ be the blowup of a smooth surface at a point $p$ lying in the smooth locus of a divisor $D\subset Z$ with normal crossings. Let $\tilde{D}\subset Y$ be the proper transform of $D$. Then $\sigma_*\Omega_Y^1(\log \tilde{D})=\Omega_Z^1(\log D)\otimes \mathfrak{M}_{p}$, where $\mathfrak{M}_{p}$ is the ideal sheaf of $p$ on $Z$.
\end{Lemma}
\begin{proof}
It suffices to show the equality in a neighborhood of the exceptional divisor $E$. Let $V\subset Z$ be a coordinate neighborhood around $p$. Choose coordinates $(z,w)$ on $V$ so that $p$ is at the origin and the local equation of $D$ is $z$. Then $\sigma^{-1}(V)$ is covered by two neighborhoods $U_1$ and $U_2$.
Choose coordinates $(x,y)$ on $U_1$ so that $\sigma(x,y)=(x,xy)$ and the local equation of $E\cap U_1$ is $x$. Note that $\tilde D$ does not appear in $U_1$. Let coordinates on $U_2$ be $(u,v)$ so that $\sigma(u,v)=(uv, v)$. On $U_2$, the local equation of $E$ is $v$ and the local equation of $\tilde D$ is $u$. See Figure~\ref{blowup picture}.
\begin{figure}
\centering
\includegraphics[scale=.8]{blowup.png}
\caption{The map $\sigma$.\label{blowup picture}}
\end{figure}
On $U_1$, we have
\begin{eqnarray}
\Omega_Y^1(\log \tilde D )(U_1)&=&\displaystyle{\left\{f \left(z,\frac{w}{z}\right)dz+g\left(z,\frac{w}{z}\right)d\left(\frac{w}{z}\right)\, \middle \vert \, f, g \in \mathcal{O}_{Z}(V)\right\}}\\
&=&\left\{\left[f\left(z,\frac{w}{z}\right)-\frac{w}{z^2}g\left(z, \frac{w}{z}\right)\right]dz+\frac{1}{z}g\left(z,\frac{w}{z}\right)dw\, \middle \vert \, f, g \in \mathcal{O}_{Z}(V)\right\}.\label{coh on Z}
\end{eqnarray}
On $U_2$
\begin{eqnarray*}
\Omega_Y^1(\log \tilde D)(U_2)&=&\left\{p\left(\frac{z}{w},w\right)\frac{d\left(\frac{z}{w}\right)}{\frac{z}{w}}+q\left(\frac{z}{w},w\right)dw\, \middle \vert \, p, q \in \mathcal{O}_{Z}(V)\right\}\\
&=&\left\{\frac{1}{z}p\left(\frac{z}{w},w\right)dz+\left[q\left(\frac{z}{w},w\right)-\frac{1}{w}p\left(\frac{z}{w},w\right)\right]dw\, \middle \vert \, p, q \in \mathcal{O}_{Z}(V)\right\}.
\end{eqnarray*}
These sections glue to a section of $\sigma_*(\Omega_Y^1(\log \tilde{D}))$ over $V$ if coefficients of $dz$ and $dw$ are equal:
\begin{equation}\label{coeff of dw}
\frac{1}{z}g\left(z,\frac{w}{z}\right)=q\left(\frac{z}{w},w\right)-\frac{1}{w}p\left(\frac{z}{w},w\right)
\end{equation}
\begin{equation}\label{coeff of dz}
\frac{1}{z}p\left(\frac{z}{w},w\right)=f\left(z,\frac{w}{z}\right)-\frac{w}{z^2}g\left(z, \frac{w}{z}\right)
\end{equation}
Replacing $\frac{1}{z}g(z,\frac{w}{z})$ in Equation~\eqref{coeff of dz} with its equivalent expression coming from Equation~\eqref{coeff of dw} yields the equality
$$f\left(z, \frac{w}{z}\right)=\frac{w}{z}q\left(\frac{z}{w},w\right).$$
From this last expression, we see that
$$f\left(z, \frac{w}{z}\right)=\frac{1}{z}f'(z,w)$$
and
$$q\left(\frac{z}{w},w\right)=\frac{1}{w}f'(z,w)$$
where $f'(z,w)$ is a polynomial with $f'(0,0)=0$. Plugging these into Equation~\eqref{coeff of dw} and multiplying through by $zw$ gives
$$wg\left(z,\frac{w}{z}\right)=z\left(f'\left(z,w\right)-p\left(\frac{z}{w},w\right)\right).$$
Since the right hand side is a polynomial in $z$, we can write $g\left(z, \frac{w}{z}\right)=zg'(z,w)$ for some polynomial $g'$ with $g'(0,0)=0$, and rewrite the above equality as
$$wg'(z,w)=f'(z,w)-p\left(\frac{z}{w},w\right).$$
Therefore, $p\left(\frac{z}{w},w\right)=wg'(z,w)-f'(z,w)$. We now have expressions for $f$, $g$, $p$, and $q$ as polynomials in $z$ and $w$, which we can use in Equation~\eqref{coh on Z}. This gives us
\begin{eqnarray*}
\sigma_*(\Omega_Y^1(\log \tilde{D}))(V)&=&\left\{\left[\frac{1}{z}f'(z,w)-\frac{w}{z}g'(z, w)\right]dz+ \, g'(z,w)dw\, \middle \vert \, f', g' \in \mathcal{O}_{Z}(V)\right\}\\
&=& \left\{f'(z,w)\frac{dz}{z} + g'(z,w)dw\, \middle \vert \, f', g' \in \mathcal{O}_{Z}(V)\right\}
\end{eqnarray*}
where the only restrictions on $f'(z,w)$ and $g'(z,w)$ are that neither has a constant term; that is, they both lie in the maximal ideal $\mathfrak{M}_p=(z,w)\subset \mathcal{O}_Z(V)\simeq\mathbb{C}[z,w]$. Thus,
$$\sigma_*(\Omega_Y^1(\log \tilde D))=\Omega_Z^1(\log D)\otimes \mathfrak{M}_{p}.$$
\end{proof}
\begin{Lemma}\label{decomposition into eigenspaces} Let $f:X\rightarrow Y$ be a double cover of a smooth surface $Y$, and let $B$ denote its smooth branch divisor. Let $C=f^{-1}(D)$ be the preimage of a smooth curve $D$ on $Y$, and suppose that $D$ intersects $B$ transversally. Then
$$f_*(\Omega_X^1(\log C))=\Omega_Y^1(\log D)\oplus\Omega_Y^1((\log D+B)(-L))$$
and
$$f_*(T_X(\log C))=T_Y(\log(D+B))\oplus T_Y(\log D)(-L)$$
where $B\sim 2L$.
Moreover, these decompositions break the sheaves into their invariant and anti-invariant subspace under the action of $\mathbb{Z}/2\mathbb{Z}$ by deck transformations.
\end{Lemma}
\begin{Remark} Lemma~\ref{decomposition into eigenspaces} is an extension of the double cover version of~\cite[Lemma 4.2]{pardini1991} to the log tangent sheaf.
\end{Remark}
\begin{proof}
In order to compute $f_*\Omega_X^1(\log C)$, note that it admits an action of $\mathbb{Z}/2\mathbb{Z}$ via deck transformations, so we can decompose it into its invariant and anti-invariant eigenspaces.
Let $V$ be an open neighborhood of $p\in D \cap B$ and choose coordinates $(z,w)$ on $V$ so that $p$ is at the origin and the local equation of $D$ is $z$ and the local equation of $B$ is $w$. Then we have an open neighborhood $U$ of $f^{-1}(p)$ with local coordinates $(x,y)$ so that $f(x,y)=(x, y^2)$. Note that the ramification locus $R$ of $f$ has local equation $y$ and the curve $C$ on $X$ has local equation $x$. See Figure~\ref{double cover picture}.
\begin{figure}[!h]
\centering
\includegraphics[scale=.8]{doublecover.png}
\caption{The map $f$.\label{double cover picture}}
\end{figure}
On $U$ we have
$$\Omega_X^1(\log C)(U)=\left<\frac{dx}{x},\,dy\right>_{\mathcal{O}_X(U)}.$$
Noting that $\mathcal{O}_Y(V)\simeq \mathbb{C}[x,y^2]$, we have
$$
f_*(\Omega_X^1(\log C))(V)=
\left<\frac{dx}{x},\,y\frac{dx}{x},\,dy, \,y dy \right>_{\mathcal{O}_Y(V)}
$$
The action of $\mathbb{Z}/2\mathbb{Z}$ sends $(x,y)$ to $(x,-y)$. Therefore the invariant subspace of $f_*(\Omega_X^1(\log C))(V)$ is
$$
f_*(\Omega_X^1(\log C))_+(V)=\left<\frac{dx}{x}, ydy\right>_{\mathcal{O}_Y(V)}=\left<\frac{dz}{z}, dw\right>_{ \mathcal{O}_Y(V)} = \Omega_Y^1(\log D)(V).
$$
The anti-invariant subspace of $f_*(\Omega_X^1(\log C))(V)$ is
$$
f_*(\Omega_X^1(\log C))_-(V)=\left< y\frac{dx}{x}, dy\right>_{\mathcal{O}_Y(V)}=
=y\left<\frac{dz}{z}, \frac{dw}{w}\right>_{\mathcal{O}_Y(V)}\\
=\Omega_Y^1((\log D+B)(-L))(V).
$$
One checks easily that these modules extend to the expected sheaves over all of $Y$.
The proof for the log tangent bundle is similar.
\end{proof}
\subsection{Smooth boundary components of $\overline{\mathcal{M}}_{5,5}$}\label{1 and 2a}
We show that loci corresponding to surfaces of type 1 and 2a give generically smooth loci in the moduli space $\overline{\mathcal{M}}_{5,5}$. In both cases, we obtain this result by proving the vanishing of the cohomology group in which obstructions to $\mathbb{Q}$-Gorenstein deformations lie. By Theorem~\ref{counting}, the type 1 and 2a loci are $39$-dimensional, so we conclude that the closure of the 1 and 2a loci are generically smooth Cartier divisors in $\overline{\mathcal{M}}_{5,5}$.
\subsubsection{The type 1 component}
For this subsection, let $W$ be a stable numerical quintic surface of type 1 or 1'' and denote by $X$ its minimal resolution. Let $f:X\rightarrow Z$ be the double cover, where $Z= \mathbb{P}^1\times\mathbb{P}^1$ or $\mathbb{F}_2$, and $f$ is branched over a smooth curve $B\sim 6\Delta$, tangent to $D\sim \Delta$ at six points. Then $f^*(D)=C_1+C_2$ and the curves $C_1$ and $C_2$ are $(-4)$ curves on $X$. Let $R=f^*B$ denote the ramification locus of $f$, and let $L\subset Z$ be a curve such that $B\sim 2L$.
In order to show that deformations of $W$ are unobstructed, it suffices to show that $H^2(W, T_W)=0$. Equivalently, as described above, we show that $H^2(X,T_X(\log (C_1)))=0$.
\begin{Theorem}\label{vanishing for type 1} Let $X$ be the minimal resolution of a stable numerical quintic surface of type 1 or 1'', and let $C_1$ and $C_2$ be the $(-4)$-curves on $X$. Then $H^2(X,T_X(\log (C_1)))=0$.
\end{Theorem}
\begin{proof}
By Serre duality, it is enough to show that $H^2(X,\Omega_X^1(\log (C_1))(K))=0,$ where $K=K_X$.
The double cover $f:S\rightarrow Z$ gives rise to an action of $\mathbb{Z}/2\mathbb{Z}$ on $H^0(X, \Omega_X^1(\log(C_1+C_2))(K))$ via deck transformations. To begin with, we prove that the anti-invariant subspace
$H^0(X, \Omega_X^1\log(C_1+C_2)(K))_-$
vanishes.
By the projection formula, noting that $K\sim f^*(\Delta)$, we have
$$f_*(\Omega_X^1\log(C_1+C_2)(K))=(f_*\Omega_X^1\log(C_1+C_2))(\Delta).$$
We claim that $$f_*\Omega_X^1(\log(C_1+C_2))_-\subset \Omega_{Z}(\log B)(-2\Delta).$$
To compute $f_*(\Omega_X^1\log(C_1+C_2))_-$, we need only consider a point in $C_1\cap C_2\cap R$. Indeed, suppose that $U$ is a neighborhood of $p\in X$ such that $U\cap C_1\cap C_2\cap R=\emptyset$, and let $V$ denote the image of $U$ under $f$. By Lemma~\ref{decomposition into eigenspaces}, we have
\begin{eqnarray*}
f_*(\Omega_X^1\log(C_1+C_2))_-(V)&=&\Omega_{Z}^1(\log (B+D))(-3\Delta)(V)\\
&\subset& \Omega_{Z}^1(\log B)(-2\Delta)(V),
\end{eqnarray*}
because $D\cap V=\emptyset$.
Now let $U$ be an open subset of $X$ containing $p\in C_1\cap C_2\cap R$, and let $V$ an open neighborhood of $f(p)$. Choose coordinates $(x,y)$ on $U$ so that $p$ is at the origin and the local equation of $R$ is $y$. We can then choose coordinates $(w,z)$ on $V$ such that the local equation of $B$ is $z$ and the local equation of $D$ is $z-w^2$. Then the local equations of $C_1$ and $C_2$ are $y-x$ and $y+x$. With these coordinates, the cover $f$ is given by the function $(x,y)\mapsto (x, y^2)$. See Figure~\ref{map to Z}.
\begin{figure}[!h]
\centering
\includegraphics[scale=.7]{MapToZ.png}
\caption{The map $f$. \label{map to Z}}
\end{figure}
The $\mathcal{O}_X(U)$-module $\Omega_X^1\log(C_1+C_2)(U)$ is generated by $\left\{\frac{d(y-x)}{y-x},\,\frac{d(y+x)}{y+x}\right\}$. As a module over $\mathcal{O}_Y(V)$, we have that $f_* \Omega_X^1\log(C_1+C_2)(V)$ is generated by
$$\left\{\frac{d(y-x)}{y-x}, d(y-x), \,\frac{d(y+x)}{y+x}, \, d(y+x)\right\}$$
Since the action of $\mathbb{Z}/2\mathbb{Z}$ sends $y$ to $-y$, we see quickly that the anti-invariant submodule is generated as an $\mathcal{O}_Y(V)$--module by
\begin{eqnarray*}
\left\{\frac{d(y-x)}{y-x}+\frac{d(y+x)}{y+x}, \, dy\right\}
&\subset&\left\{\frac{1}{y^2-x^2}(-2ydx+2xdy), \, \frac{1}{y^2-x^2} dy\right\}\\
&=&\frac{y}{z-w^2}\left\{-2dw, \, \frac{dz}{z}\right\}
\end{eqnarray*}
This last module we recognize as $\Omega_{Z}^1(\log B)(-3\Delta+D)(V)=\Omega_{Z}^1(\log B)(-2\Delta)(V)$.
Thus,
$$f_*\Omega^1_X(\log(C_1+C_2))_-\subset\Omega_{Z}^1(\log (B))(-2\Delta).$$
By the projection formula, using that $K\sim f^*\Delta$, we have
$$f_*\Omega^1_X(\log(C_1+C_2))(K)_-\subset \Omega_{Z}^1(\log B)(-\Delta).$$
To show that $H^0(Z, \Omega_{Z}^1(\log B)(-\Delta))=0$, consider the exact sequence
$$0\rightarrow \Omega_{Z}^1\rightarrow \Omega_{Z}^1(\log B)\rightarrow \mathcal{O}_B\rightarrow 0$$
where $\Omega_{Z}^1(\log B)\rightarrow \mathcal{O}_B$ is the residue map. Twisting by $-\Delta$ gives the exact sequence
$$0\rightarrow \Omega_{Z}^1(-\Delta)\rightarrow \Omega_{Z}^1(\log B)(-\Delta)\rightarrow \mathcal{O}_B(-\Delta)\rightarrow 0.$$
Looking at the corresponding long exact sequence in cohomology, it remains to show that $H^0(Z, \Omega_{Z}^1(-\Delta))=0$ and $H^0(B, \mathcal{O}_B(-\Delta))=0$. Both of these are obvious, the first because $H^0(Z, \Omega_{Z}^1(-\Delta))\subset H^0(Z, \Omega_{Z}^1)=0$ and the second because $-\Delta\cdot B=-12<0$. Thus,
$$H^0(X, \Omega_X^1\log(C_1+C_2)(K))_-=0,$$
as we wished to show.
Now consider a one-form $\alpha\in\Omega_X^1(\log C_1)(K)$. Since
$$\Omega_X^1(\log C_1)(K)\subset\Omega_X^1(\log C_1+C_2)(K)$$
and the latter sheaf has no anti-invariant part, the one-form $\alpha$ must be invariant. But the action of $\mathbb{Z}/2\mathbb{Z}$ on cohomology interchanges $C_1$ and $C_2$, so $\alpha$ must not have a pole along $C_1$.
Thus,
$$\alpha\in H^0(X, \Omega_X^1(K))\simeq H^2(X, T_X)^{\vee}.$$
By Horikawa~\cite{horikawa1976}, we have $H^2(X, T_X)=0$, and so $\alpha=0$, completing the proof.
\end{proof}
In Section~\ref{dimension counts}, we showed that the locus of stable quintic surfaces of type 1 is $39$-dimensional, so Theorems~\ref{quintic example} and~\ref{vanishing for type 1} imply the following:
\begin{Cor}\label{1 is divisor} The closure of the locus of surfaces of type 1 is a generically smooth Cartier divisor in $\overline{\mathcal{M}}_{5,5}$, lying in the closure of the type I component of $\mathcal{M}_{5,5}$.
\end{Cor}
\subsubsection{The 2a component}\label{2a component}
Let $W$ be a stable numerical quintic surface of type 2a, 2a', or 2a'' and let $S$ denote its minimal resolution. Then there is a map $\tilde{f}: X \rightarrow \tilde{Z}$, which is the double cover of the blowup of $Z=\mathbb{P}^1\times\mathbb{P}^1$ in two points $p$ and $q$ lying on a fiber $D$. The branch locus $\tilde{B}$ of $\tilde{f}$ is the proper transform of an irreducible curve $B \sim 6\Delta$ which has either a node or an $A_2$ singularity at each of $p$ and $q$ and is smooth elsewhere. Denote by $\Gamma_1$ and $\Gamma_2$ generic rulings of $\tilde{Z}$ so that $\Gamma_2\sim \tilde{D}+E_1+E_2$, where $\tilde{D}$ is the proper transform of $D\subset Z$.
\begin{Theorem}\label{vanishing for type 2a} Let $W$ be a stable numerical quintic surface of type 2a, 2a', or 2a'', let $X$ be its minimal resolution and $C$ the $(-4)$-curve on $X$. Then $H^2(X, T_X(\log C)))=0.$
\end{Theorem}
We begin with a lemma.
\begin{Lemma}\label{another lemma for type 2a} $H^0(\tilde{Z}, \Omega_{\tilde{Z}}^1(\log\tilde{D}+\tilde{B})(K_{\tilde{Z}}))=0.$
\end{Lemma}
\begin{proof}
We have the following exact sequence of sheaves on $\tilde{Z}$:
$$0\rightarrow \Omega_{\tilde{Z}}^1\rightarrow \Omega_{\tilde{Z}}^1(\log(\tilde{D}+\tilde{ B}))\rightarrow\mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}}\rightarrow 0$$
where $\Omega_{\tilde{Z}}^1(\log{\tilde{D}+\tilde B})\rightarrow\mathcal{O}_{\tilde{D}+\tilde{B}}$ is the residue map. Twisting by $K_{\tilde{Z}}$ gives the exact sequence
\begin{equation}\label{SES with K}
0\rightarrow \Omega_{\tilde{Z}}^1(K_{\tilde{Z}})\rightarrow \Omega_{\tilde{Z}}^1(\log{\tilde{D}+\tilde B})(K_{\tilde{Z}})\rightarrow(\mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}})(K_{\tilde{Z}})\rightarrow 0.
\end{equation}
Note that
\begin{equation}
K_{\tilde{Z}}=\sigma^*(K_{\mathbb{P}^1\times\mathbb{P}^1})+E_1+E_1=-2\Gamma_1-2\Gamma_2+E_1+E_2\sim -2\Gamma_1-2\tilde{D}-E_1-E_2,
\end{equation}
and so $-K_{\tilde{Z}}$ is effective. Thus $H^0(\tilde{Z}, \Omega_{\tilde{Z}}^1(K_{\tilde{Z}}))\subset H^0(\tilde{Z}, \Omega_{\tilde{Z}}^1)$. Since the irregularity of ${\tilde{Z}}$ is zero, we have
$H^0({\tilde{Z}}, \Omega_{\tilde{Z}}^1)(K_{\tilde{Z}})=0$. Moreover, noting that $\sigma^*(B)=\tilde{B}+2E_1+2E_2$ and $\sigma^*(K_Z)=K_{\tilde{Z}}-E_1-E_2$, we have
$$K_{\tilde{Z}}\cdot\tilde{B}=-24<0$$
and
$$K_{\tilde{Z}}\cdot \tilde{D}=0.$$
Therefore $H^0(\tilde{Z}, (\mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}})(K_{\tilde{Z}}))=\mathbb{C}$, so the cohomology group
$$H^0(\tilde{Z}, \Omega_{\tilde{Z}}^1(\log(\tilde{D}+\tilde{B}))(K_{\tilde{Z}}))$$
is $0$ if and only if the connecting homomorphism
$$\delta:H^0(\tilde{Z}, (\mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}})(K_{\tilde{Z}}))\rightarrow H^1(\tilde{Z}, \Omega_{\tilde{Z}}^1(K_{\tilde{Z}}))$$
is injective.
Since $-K_{\tilde{Z}}$ is effective, we have a section $s\in H^0(\tilde{Z}, \Omega_{\tilde{Z}}^1(-K_{\tilde{Z}}))$, so we have a map from the short exact sequence~\eqref{SES with K} to the short exact sequence
$$0\rightarrow \Omega_{\tilde{Z}}^1\rightarrow \Omega_{\tilde{Z}}^1(\log(\tilde{D}+\tilde{B}))\rightarrow\mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}}\rightarrow 0.$$
where the map is given by tensoring with $s$.
The connecting homomorphism
$$\delta_2: H^0(\tilde{Z}, \mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}})\rightarrow H^1(\tilde{Z}, \Omega_{\tilde{Z}}^1)$$
of the corresponding short exact sequence
is the first Chern class map. That is, if $1_{\tilde{D}}$ and $1_{\tilde{B}}$ are generators of $H^0(\tilde{Z}, \mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}})$, then $\delta_2(1_{\tilde{D}})=c_1(\tilde{D})$ and $\delta_2(1_{\tilde{B}})=c_1(B)$. Thus, the map $\delta_2$ is injective if and only if the curves $\tilde{D}$ and $\tilde{B}$ are linearly independent in the Picard group of $\tilde{Z}$. Recalling that Pic$(\tilde{Z})$ is generated by $\Gamma_1$, $\Gamma_2$, $E_1$ and $E_2$, and that $\tilde{B}\sim 6\Gamma_1+6\Gamma_2-2E_1-2E_2$ and $\tilde{D}\sim \Gamma_2-E_1-E_2$, we see that the two divisors are indeed linearly independent.
Thus, we have a diagram
\[\xymatrix{
H^0(\tilde{Z}, (\mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}})(K_{\tilde{Z}}))\ar[r]^{\delta} \ar[d]_{\otimes s} & H^1(\tilde{Z}, \Omega_{\tilde{Z}}^1(K_{\tilde{Z}})) \ar[d]^{\otimes s}\\
H^0(\tilde{Z}, \mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{
\tilde{B}})\ar[r]^{\delta_2}& H^1(\tilde{Z}, \Omega_{\tilde{Z}}^1)
}\]
where the bottom arrow is injective. We see that $\delta$ is injective as long as the map on the left is injective. But this map simply takes a section of $(\mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}})(K_{\tilde{Z}})$ and multiplies it by $s$. Since $s\neq0$, the map is injective.
\end{proof}
\begin{proof}[Proof of Theorem~\ref{vanishing for type 2a}]
We show that $H^2(X, T_X(\log C))=0$, where $X$ is the minimal resolution of $W$ and $C$ is the $(-4)$-curve on $X$. By Serre duality, it is enough to show that $H^0(X, \Omega_X^1(\log C)(K_X))=0$. Recall that $C=f^*\tilde{D}$ and $K_X=f^*(K_Y+\tilde{L})$. By the projection formula
$$f_*(\Omega_X^1(\log C)(K_X))=(f_*\Omega_X^1(\log C))\otimes (K_Y+\tilde{L}).$$
Together with Lemma~\ref{decomposition into eigenspaces}, this gives
$$
f_*(\Omega_X^1(\log C)(K_S))=\Omega_Y^1(\log\tilde{D})(K_Y+\tilde{L})\oplus\Omega_Y^1(\log\tilde{D}+\tilde{B})(K_Y).
$$
By Lemma~\ref{another lemma for type 2a}, we have $H^0(Y, \Omega_Y^1(\log\tilde{D}+\tilde{B})(K_Y))=0$.
It remains to show that $H^0(Y, \Omega_Y^1(\log\tilde{D})(K_Y+\tilde{L}))=0$, which we do via the projection formula.
By Lemma~\ref{blowup of two points}, we have $\sigma_*\Omega_Y^1(\log\tilde{D})=\Omega_Z^1(\log D)\otimes \mathfrak{M}_{p,q}$, where $\mathfrak{M}_{p,q}$ is the ideal sheaf of $p$ and $q$ which are the centers of $\sigma$. Noting that $(K_Y+\tilde{L})=f^*(\Delta)$, the projection formula gives
\begin{eqnarray*}
\sigma_*(\Omega_Y^1(\log\tilde{D})(K_Y+\tilde{L})) &=&(\Omega_{\mathbb{P}^1\times\mathbb{P}^1}^1(\log D)\otimes \mathfrak{M}_{p,q})\otimes \mathcal{O}(\Delta)\\
&=&[(p_1^*\Omega_{\mathbb{P}^1}^1(\log D)\otimes \mathfrak{M}_{p,q})\oplus(p_2^*\Omega_{\mathbb{P}^1}^1\otimes\mathfrak{M}_{p,q})]\otimes \mathcal{O}(\Delta)\\
&=&(\mathcal{O}(0,1)\otimes\mathfrak{M}_{p,q})\oplus(\mathcal{O}(1,-1)\otimes \mathfrak{M}_{p,q}).
\end{eqnarray*}
We have $H^0(\mathbb{P}^1\times\mathbb{P}^1, \mathcal{O}(1,-1)\otimes \mathfrak{M}_{p,q})=0$, because $H^0(\mathbb{P}^1\times\mathbb{P}^1, \mathcal{O}(a,b))=0$ for $a<0$ or $b<0$. And $H^0(\mathbb{P}^1\times\mathbb{P}^1, \mathcal{O}(0,1)\otimes\mathfrak{M}_{p,q})=0$, since $p$ and $q$ lie on $D\in |1,0|$.
\end{proof}
By Theorem~\ref{counting}, the locus of 2a surfaces is $39$-dimensional. Moreover, Theorem~\ref{Friedman} shows that every 2a surfaces may be obtained as the stable limit of a family of numerical quintic surfaces of type IIa. Together with Theorem~\ref{vanishing for type 2a}, this implies the following
\begin{Cor}\label{2a is divisor} The closure of the locus of surfaces of type 2a is a generically smooth Cartier divisor in $\overline{\mathcal{M}}_{5,5}$, lying in the closure of the type IIa component of $\mathcal{M}_{5,5}.$
\end{Cor}
\section{Deformations of 2b surfaces}\label{2b}
We study the versal $\mathbb{Q}$-Gorenstein deformation space $\textrm{Def}^{QG}(W)$~\cite{hacking2004} where $W$ is a general 2b surface. All deformation functors considered are functors of Artinian rings. However, because $W$ is a stable surface, we often abuse notation and view $\textrm{Def}^{QG}(W)$ as an analytic germ of a point $[W]$ in the KSBA moduli space. The same notational ambiguity applies to other deformation functors we consider which admit a moduli space. This enables us to study the moduli space $\overline{\mathcal{M}}$ using analytic methods of Horikawa~\cite{horikawa1975, horikawa1976}. The main theorem is
\begin{Theorem}\label{the best} The locus of stable numerical quintic surfaces whose unique non Du Val singularity is a $\frac{1}{4}(1,1)$ singularity forms a divisor in $\overline{\mathcal{M}}_{5,5}$ which consists of two 39-dimensional components $\bar{1}$ and $\overline{\mbox{2a}}$ meeting, transversally at a general point, in a 38-dimensional component $\overline{\mbox{2b}}$. This divisor is Cartier at general points of the $\bar{1}$, $\overline{\mbox{2a}}$, and $\overline{\mbox{2b}}$ components. These components are the closures of the loci of 1, 2a, and 2b surfaces described at the beginning of Section~\ref{dimension counts}. Moreover, the type $\bar{1}$, $\overline{\mbox{2a}}$, and $\overline{\mbox{2b}}$ components belong to the closure of the components in $\mathcal{M}_{5,5}$ of types I, IIa, and IIb, respectively.
\end{Theorem}
The proof will consist of several pieces. Theorems~\ref{vanishing for type 1} and~\ref{vanishing for type 2a} showed that obstructions to deformations of surfaces of types 1 and 2a vanish, and so the closures of their corresponding $39$-dimensional loci in $\overline{\mathcal{M}}_{5,5}$ are generically smooth Cartier divisors. In Theorem~\ref{obstruction}, we show that deformations of 2b surfaces are obstructed and that the obstruction space is one-dimensional. This implies that the space $\textrm{Def}^{QG}(W)$ of $\mathbb{Q}$-Gorenstein deformations of a generic 2b surface $W$ is a hypersurface singularity. We show that the space of equisingular $\mathbb{Q}$-Gorenstein deformations of $W$ consists of irreducible components $\bar{1}$ and $\overline{\mbox{2a}}$ meeting, transversally at a general point, in the $\overline{\mbox{2b}}$ component. Together with Horikawa's description of $\mathcal{M}_{5,5}$, and the smoothings described in Theorems~\ref{quintic example} and~\ref{Friedman}, this implies that the space $\textrm{Def}^{QG}(W)$ has two irreducible components.
By Theorem~\ref{Friedman}, there exists a $\mathbb{Q}$-Gorenstein smoothing of a 2b surface to a numerical quintic surface of type IIb which induces a versal deformation of the singularity. Therefore, the map $\textrm{Def}^{QG}(W)\rightarrow\textrm{Def}^{QG}_{\mbox{loc}}(p)$ to local $\mathbb{Q}$-Gorenstein deformations of the $\frac{1}{4}(1,1)$ singularity $(p\in W)$ is surjective. This latter space is one-dimensional, and since $\textrm{Def}^{QG}(W)$ has two irreducible components, the space $\textrm{Def}^{QG}(W)$ is analytically isomorphic to $\textrm{Def}^{QG}_{\mbox{e.s.}}(W)\times\mathbb{A}^1$.
The key to Theorem~\ref{the best} is the proof of the fact that the space $\textrm{Def}^{QG}_{\mbox{e.s.}}(W)$ of equisingular $\mathbb{Q}$-Gorenstein deformations of a general 2b surface $W$ consists of two irreducible components meeting transversally at a general point. This space is isomorphic to the deformation space of pairs $\textrm{Def}(X,C)$, where $X$ is the minimal resolution of $W$, containing $(-4)$-curve $C$.
In~\ref{equisingular}, we describe a subfunctor of the deformation functor of pairs $\mathpzc{Def}(X,C)$, and show that this subfunctor has no obstructions. This will imply that the space $\textrm{Def}(X,C)$ contains a smooth component corresponding to the 2a locus. Thus, to prove Theorem~\ref{the best},
it suffices to show that the degree two part of the Kuranishi map, given by the Schouten bracket, is nonzero and not a square. Horikawa makes a similar argument in~\cite{horikawa1975} and~\cite{horikawa1976}. In~\ref{technicalities} and~\ref{main proof}, we extend his work to the log setting.
We use the following notation throughout this section. Let $X$ be the minimal resolution of a surface of type 2b. We recall the construction of $X$. Let $\sigma: \tilde{\mathbb{F}}_2 \rightarrow \mathbb{F}_2$ be the blowup of $\mathbb{F}_2$ in two distinct points $p$ and $q$ lying on a fiber $D$. Denote by $\tilde{D}$ and $\Gamma$ the proper transforms of $D$ and a generic fiber, respectively, and let $E_1$ and $E_2$ be the exceptional divisors of $\sigma$. By abuse of notation, we denote by $\Delta_0$ the proper transform of the negative section $\Delta_0$ on $\mathbb{F}_2$. Let $B$ be a reduced, irreducible divisor in the linear system $|6\Delta_0+12\Gamma|$ on $\mathbb{F}_2$ with simple nodes at $p$ and $q$ and no other singularities. Let $\tilde{B}$ be its proper transform and note that $\tilde{B}\sim 2\tilde{L}$ for some smooth divisor $L$ on $\tilde{\mathbb{F}}_2$. Then $X$ is given by the double cover $f: X\rightarrow \tilde{\mathbb{F}}_2$ branched over $\tilde{B}$. The curve $C$ given by $f^*(\tilde{D})$ is the $(-4)$-curve on $X$. Moreover $X$ contains four $(-2)$-curves: $F_1$ and $F_2$ mapping to $\Delta_0$, and $\bar{E}_1$ and $\bar{E}_2$ mapping to $E_1$ and $E_2$, respectively. We denote by $\pi:\mathbb{F}_2\rightarrow \mathbb{P}^1$ and $g:X\rightarrow \mathbb{P}^1$ the projection maps to $\mathbb{P}^1$.
\subsection{The obstruction}\label{the obstruction}
To begin with, we show that the obstruction space is one-dimensional.
\begin{Theorem}\label{obstruction} Let $X$ be the minimal resolution of a 2b surface, and let $C$ denote the $(-4)$-curve on $X$. Then $H^2(X, T_X(\log C))=\mathbb{C}$.
\end{Theorem}
The proof of Theorem~\ref{obstruction} requires two lemmas.
\begin{Lemma}\label{coh vanishing on the blowup of F2} Let $Z=\mathbb{F}_2$ and $\tilde{Z}$ the blowup of $Z$ in $p$ and $q$. Then $H^0(\tilde{Z}, \Omega^1_{\tilde{Z}}(\log(\tilde{D}+\tilde{B}+\Delta_0))(K_{\tilde{Z}}))=0$.
\end{Lemma}
\begin{proof}
The proof is very similar to that of Lemma~\ref{another lemma for type 2a}.
We have the following exact sequence of sheaves on $\tilde{Z}$:
$$0\rightarrow\Omega_{\tilde{Z}}^1(K_{\tilde{Z}})\rightarrow\Omega_{\tilde{Z}}^1(\log(\tilde{D}+\tilde{B}+\Delta_0))(K_{\tilde{Z}})\rightarrow (\mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}}\oplus\mathcal{O}_{\Delta_0})(K_{\tilde{Z}})\rightarrow 0.$$
where $\Omega_{\tilde{Z}}^1(\log{\tilde{D}+\tilde B})\rightarrow\mathcal{O}_{\tilde{D}+\tilde{B}}$ is the residue map. Twisting by $K_{\tilde{Z}}$ gives the exact sequence
\begin{equation}\label{SES with K for F2}
0\rightarrow \Omega_{\tilde{Z}}^1(K_{\tilde{Z}})\rightarrow \Omega_{\tilde{Z}}^1(\log{\tilde{D}+\tilde B+\Delta_0})(K_{\tilde{Z}})\rightarrow(\mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}}\oplus\mathcal{O}_{\Delta_0})(K_{\tilde{Z}})\rightarrow 0.
\end{equation}
Note that
\begin{equation*}
K_{\tilde{Z}}=\sigma^*(K_{\mathbb{F}_2})+E_1+E_1\sim -2\Delta_0-4\tilde{D}-3E_1-3E_2,
\end{equation*}
and so $-K_{\tilde{Z}}$ is effective. Thus $H^0(\tilde{Z}, \Omega_{\tilde{Z}}^1(K_{\tilde{Z}}))\subset H^0(\tilde{Z}, \Omega_{\tilde{Z}}^1)$. Since the irregularity of ${\tilde{Z}}$ is zero, we have
$H^0({\tilde{Z}}, \Omega_{\tilde{Z}}^1)(K_{\tilde{Z}})=0$. Moreover, because $\sigma^*(B)=\tilde{B}+2E_1+2E_2$ and $\sigma^*(K_Z)=K_{\tilde{Z}}-E_1-E_2$, we have $K_{\tilde{Z}}\cdot\tilde{B}=-24<0$, $K_{\tilde{Z}}\cdot \tilde{D}=0$,
and
$K_{\tilde{Z}}\cdot \Delta_0=0.$
Therefore $$H^0(\tilde{Z}, (\mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}}\oplus\mathcal{O}_{\Delta_0})(K_{\tilde{Z}}))=\mathbb{C}^2,$$ so the cohomology group
$$H^0(\tilde{Z}, \Omega_{\tilde{Z}}^1(\log(\tilde{D}+\tilde{B}+\Delta_0))(K_{\tilde{Z}}))$$
is $0$ if and only if the connecting homomorphism
$$\delta:H^0(\tilde{Z}, (\mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}}\oplus\mathcal{O}_{\Delta_0})(K_{\tilde{Z}}))\rightarrow H^1(\tilde{Z}, \Omega_{\tilde{Z}}^1(K_{\tilde{Z}}))$$
is injective.
Since $-K_{\tilde{Z}}$ is effective, we have a section $s\in H^0(\tilde{Z}, \Omega_{\tilde{Z}}^1(-K_{\tilde{Z}}))$, so we have a map from the short exact sequence~\eqref{SES with K for F2} to the short exact sequence
$$0\rightarrow \Omega_{\tilde{Z}}^1\rightarrow \Omega_{\tilde{Z}}^1(\log(\tilde{D}+\tilde{B}+\Delta_0))\rightarrow\mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}\oplus\mathcal{O}_{\Delta_0}}\rightarrow 0$$
given by tensoring with $s$.
The connecting homomorphism
$$\delta_2: H^0(\tilde{Z}, \mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}}\oplus\mathcal{O}_{\Delta_0})\rightarrow H^1(\tilde{Z}, \Omega_{\tilde{Z}}^1)$$
of the corresponding short exact sequence
is the first Chern class map. That is, if $1_{\tilde{D}}$, $1_{\tilde{B}}$, and $1_{\Delta_0}$ are generators of $H^0(\tilde{Z}, \mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}}\oplus\mathcal{O}_{\Delta_0})$, then $\delta_2(1_{\tilde{D}})=c_1(\tilde{D})$, $\delta_2(1_{\tilde{B}})=c_1(\tilde{B})$, and $\delta_2(1_{\Delta_0})=c_1(\Delta_0)$. Thus, the map $\delta_2$ is injective if and only if the curves $\tilde{D}$, $\tilde{B}$, and $\Delta_0$ are linearly independent in the Picard group of $\tilde{Z}$. Recalling that Pic$(\tilde{Z})$ is generated by $\Delta_0$, $\Gamma$, $E_1$ and $E_2$, that $\tilde{B}\sim 6\Delta_0+12\Gamma-2E_1-2E_2$ and that $\tilde{D}\sim \Gamma-E_1-E_2$, we see that the three divisors are indeed linearly independent.
Thus, we have a diagram
\[\xymatrix{
H^0(\tilde{Z}, (\mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}}\oplus\mathcal{O}_{\Delta_0})(K_{\tilde{Z}}))\ar[r]^{\delta} \ar[d]_{\otimes s} & H^1(\tilde{Z}, \Omega_{\tilde{Z}}^1(K_{\tilde{Z}})) \ar[d]^{\otimes s}\\
H^0(\tilde{Z}, \mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{
\tilde{B}}\oplus\mathcal{O}_{\Delta_0})\ar[r]^{\delta_2}& H^1(\tilde{Z}, \Omega_{\tilde{Z}}^1)
}\]
where the bottom arrow is injective. We see that $\delta$ is injective as long as the map on the left is injective. But this map simply takes a section of $(\mathcal{O}_{\tilde{D}}\oplus\mathcal{O}_{\tilde{B}})(K_{\tilde{Z}})$ and multiplies it by $s$. Since $s\neq0$, the map is injective.
\end{proof}
\begin{Lemma}\label{where the obstruction happens} $H^0(\tilde{\mathbb{F}}_2,\Omega_{\tilde{\mathbb{F}}_2}^1(\log \tilde{D})(K_{\tilde{\mathbb{F}}_2}+\tilde{L}))=\mathbb{C}.$
\end{Lemma}
\begin{proof}
By the projection formula we have
\begin{eqnarray*}
\sigma_*(\Omega_{\tilde{\mathbb{F}}_2}^1(\log \tilde{D})(K_{\tilde{\mathbb{F}}_2}+\tilde{L}))&=&\sigma_*(\Omega_{\tilde{\mathbb{F}}_2}^1(\log \tilde{D}))(K_{\mathbb{F}_2}+L)\\
&=& \sigma_*(\Omega_{\tilde{\mathbb{F}}_2}^1(\log \tilde{D}))\otimes\mathcal{O}(\Delta_0+2\Gamma).
\end{eqnarray*}
Lemma~\ref{blowup of two points} gives
$$\sigma_*(\Omega_{\tilde{\mathbb{F}}_2}^1(\log \tilde{D}))=\Omega^1_{\mathbb{F}_2}(\log D)\otimes\mathfrak{M}_{p,q}.$$
Thus,
$$ \sigma_*(\Omega_{\tilde{\mathbb{F}}_2}^1(\log \tilde{D})(K_{\tilde{\mathbb{F}}_2}+\tilde{L}))=\Omega^1_{\mathbb{F}_2}(\log D)\otimes\mathfrak{M}_{p,q}\otimes\mathcal{O}(\Delta_0+2\Gamma).$$
Let $\pi:\mathbb{F}_2\rightarrow \mathbb{P}^1$ be the projection map, and suppose that $\pi(D)=a$. We have the short exact sequence
$$0\rightarrow\pi^*\Omega_{\mathbb{P}^1}^1(\log a)\rightarrow\Omega_{\mathbb{F}_2}^1(\log D)\rightarrow\mathcal{O}_{\mathbb{F}_2}(-2\Delta_0-2\Gamma)\rightarrow 0.$$
The sheaf $\mathcal{O}_{\mathbb{F}_2}(-2\Delta_0-2\Gamma)$ is free, so $\textrm{Tor}_1(\mathfrak{M}_{p,q}\otimes \mathcal{O}(\Delta_0+2\Gamma), \mathcal{O}_{\mathbb{F}_2}(-2\Delta_0-2\Gamma))=0$.
Thus, tensoring $\mathfrak{M}_{p,q}\otimes \mathcal{O}(\Delta_0+2\Gamma)$ with the above short exact sequence yields the new short exact sequence
\begin{eqnarray*}
0\rightarrow\pi^*\Omega_{\mathbb{P}^1}^1(\log a)\otimes \mathfrak{M}_{p,q}\otimes \mathcal{O}(\Delta_0+2\Gamma)&\rightarrow&\Omega_{\mathbb{F}_2}^1(\log D)\otimes \mathfrak{M}_{p,q}\otimes \mathcal{O}(\Delta_0+2\Gamma)\\
&\rightarrow& \mathcal{O}_{\mathbb{F}_2}(-\Delta_0)\otimes\mathfrak{M}_{p,q}\rightarrow 0.
\end{eqnarray*}
Since $\mathbb{F}_2$ is projective, the sheaf $\mathcal{O}_{\mathbb{F}_2}(-\Delta_0)\otimes\mathfrak{M}_{p,q}$ has no global holomorphic sections, and so
$$H^0(\mathbb{F}_2, \pi^*\Omega_{\mathbb{P}^1}^1(\log a)\otimes \mathfrak{M}_{p,q}\otimes \mathcal{O}(\Delta_0+2\Gamma))\cong H^0(\mathbb{F}_2,\Omega_{\mathbb{F}_2}^1(\log D)\otimes \mathfrak{M}_{p,q}\otimes \mathcal{O}(\Delta_0+2\Gamma)).$$
The sheaf $\Omega^1_{\mathbb{P}^1}$ is isomorphic to $\mathcal{O}_{\mathbb{P}^1}(-2)$, so by the projection formula we have
$$H^0(\mathbb{F}_2, \pi^*\Omega_{\mathbb{P}^1}^1(\log a)\otimes \mathfrak{M}_{p,q}\otimes \mathcal{O}(\Delta_0+2\Gamma))=H^0(\mathbb{F}_2, \mathcal{O}_{\mathbb{F}_2}(\Delta_0+\Gamma)\otimes\mathfrak{M}_{p,q}).$$
The divisor $\Delta_0$ satisfies $\Delta_0\cdot (\Delta_0+\Gamma)=-1,$
so that $\Delta_0$ is a fixed part of the linear system $|\Delta_0+\Gamma|$. Since $D$ is the only fiber containing both $p$ and $q$, this implies that
$$H^0(\mathbb{F}_2, \mathcal{O}_{\mathbb{F}_2}(\Delta_0+\Gamma)\otimes\mathfrak{M}_{p,q})=\mathbb{C},$$
completing the proof.
\end{proof}
We can now prove the theorem.
\begin{proof}[Proof of Theorem~\ref{obstruction}] By Serre duality, it suffices to show that $$H^0(X, \Omega_X^1(\log C)(K_X))=\mathbb{C}.$$
By Lemma~\ref{decomposition into eigenspaces} and the projection formula, we have
$$f_*(\Omega_X^1(\log C)(K_X))=\Omega_{\tilde{\mathbb{F}}_2}^1(\log \tilde{D})(K_{\tilde{\mathbb{F}}_2}+\tilde{L})\oplus\Omega^1_{\tilde{\mathbb{F}}_2}(\log(\tilde{D}+\tilde{B}))(K_{\tilde{\mathbb{F}}_2}).$$
By Lemma~\ref{where the obstruction happens}, we have $H^0(\Omega_{\tilde{\mathbb{F}}_2}^1(\log \tilde{D})(K_{\tilde{\mathbb{F}}_2}+\tilde{L}))=\mathbb{C}.$ Moreover,
$$\Omega^1_{\tilde{\mathbb{F}}_2}(\log(\tilde{D}+\tilde{B}))(K_{\tilde{\mathbb{F}}_2})\subset \Omega^1_{\tilde{\mathbb{F}}_2}(\log(\tilde{D}+\tilde{B}+\Delta_0))(K_{\tilde{\mathbb{F}}_2}).$$
Thus, $H^0(\Omega^1_{\tilde{\mathbb{F}}_2}(\log(\tilde{D}+\tilde{B}))(K_{\tilde{\mathbb{F}}_2}))=0$ by Lemma~\ref{coh vanishing on the blowup of F2}.
\end{proof}
\subsection{Deformations of pairs and the equisingular locus}\label{equisingular}
Let $f:X\rightarrow Y$ be the double cover of a smooth surface $Y$ branched over a smooth curve $B$. Define $\textrm{Def}_{X\rightarrow Y}$ to be the space of deformations of $X$ that are double covers of deformations of $Y$. The group $\mathbb{Z}/2\mathbb{Z}$ acts on $X$ by deck transformations, and the sheaf $f_*T_X$ decomposes into invariant and anti-invariant subspaces as
$$f_*T_X\simeq T_Y(\log B) \oplus T_Y(-L),$$
where $2L\sim B$~\cite{pardini1991}.
\begin{Theorem}\label{cvs}~\cite{cynk-vanstraten2006} Via the decomposition of $f_*T_X$ into its invariant and anti-invariant subspaces, the deformation space $\textrm{Def}(X\rightarrow Y)$ of double covers of deformations of $Y$ may be identified with the deformation space $\textrm{Def}(Y, B)$ of deformations of pairs, where $B$ is the branch divisor of $f$.
\end{Theorem}
The proof of Theorem~\ref{cvs} involves identifying the space of first order infinitesimal deformations of double covers of deformations of $Y$ with the anti-invariant subspace $H^1_+(X, T_X)\subset H^1(X, T_X)$. Then using the decomposition of $f_*(T_X)$ above, this space is isomorphic to $H^1(Y, T_Y(\log B))$.
Using Lemma~\ref{decomposition into eigenspaces}, the same analysis works in the presence of the curves $C\subset X$ and $D\subset Y$, as long as $D$ intersects $B$ transversally. More explicitly, define $\mathpzc{Def}_{(X, C)\rightarrow(Y,D)}$ to be the functor of Artinian local rings which associates to an Artinian local ring $A$ the set of isomorphism classes of deformations over $A$ of squares
$$\xymatrix{
X \ar[r] & Y \\
C \ar@{^{(}->}[u]\ar[r]& D \ar@{^{(}->}[u]
}$$
where the top and bottom maps are double covers and the left and right maps are embeddings of the smooth curves $C$ and $D$ into $X$ and $Y$, respectively.
Then the functor $\mathpzc{Def}_{(X, C)\rightarrow(Y,D)}$ may be identified with the functor $\mathpzc{Def}_{(Y, B, D)}$ of deformations of triples. The space of first-order infinitesimal deformations of triples $(Y, B, D)$ is therefore $H^1_+(X, T_X(\log C))$. By Lemma~\ref{decomposition into eigenspaces}, we have
$$H^1_+(X, T_X(\log C))\simeq H^1(Y, T_Y(\log (B+D))).$$
There is a forgetful map $\alpha:\mathpzc{Def}_{(X, C)\rightarrow(Y,D)}\rightarrow\mathpzc{Def}_{(X,C)}$. This map is an analytic embedding, because the differential
$$d\alpha: H^1(Y, T_Y(\log (B+D)))\rightarrow H^1_+(X, T_X(\log C))\subset H^1(X, T_X(\log C))$$
is an isomorphism onto its image.
Suppose now that $W$ is a stable numerical quintic surface of type 2b, $X$ its minimal resolution, and $C$ the $(-4)$ curve on $X$. Then we have the commutative square
$$\xymatrix{
X \ar[r]^{\tilde{f}} & \tilde{Z} \\
C \ar@{^{(}->}[u]\ar[r]& \tilde{D} \ar@{^{(}->}[u]
}$$
where $\tilde{f}$ is the double cover of $\tilde{Z}$, where $\tilde{Z}$ is the blowup of $Z=\mathbb{F}_2$ in two points lying on a fiber $D$. The branch curve of $\tilde{f}$ is a smooth curve $\tilde{B}$, which intersects the proper transform $\tilde{D}$ of $D$ transversally.
Deformations of this square can be identified with deformations of the triple $(\tilde{Z}, B, \tilde{D})$. The following lemma shows that in this case, the image of $\alpha$ is a neighborhood of $[W]$ in the $\overline{\mbox{2a}}$ component of $\overline{\mathcal{M}}_{5,5}$. We note that Lemma~\ref{coh vanishing on the blowup of F2} implies that there are no obstructions, so the image of $\alpha$ is smooth.
\begin{Theorem}\label{2a to 2b come from double covers} Let $\mathcal{W}\rightarrow T$ be a stable family whose fibers are all 2a or 2b surfaces and $\mathcal{X}\rightarrow \mathcal{W}$ be its simultaneous minimal resolution over $T$, which exists by~\cite[Theorem 7.68]{kollarmori}. Then there exists a double cover $j:\mathcal{X}\rightarrow \tilde{\mathcal{Z}}$ of schemes smooth over $T$, where $\tilde{\mathcal{Z}}$ is a smooth family of Hirzebruch surfaces of type $\mathbb{F}_2$ or $\mathbb{F}_0$ blown up at two points on a fiber.
\end{Theorem}
\begin{proof}
Let $\psi:\mathcal{X}\rightarrow \mathcal{Y}$ be the canonical model of $\mathcal{X}$ over $T$. Then the canonical map given by the linear system $|\omega_{\mathcal{Y}/T}|$ is a double cover $f: \mathcal{Y}\rightarrow \mathcal{Z}$ over $T$, where fibers of $\mathcal{Z}\rightarrow T$ are either smooth or singular quadrics. Let $\mathcal{B}$ denote the branch divisor of $f$ and suppose that $\mathcal{Z}_{t_0}$ is singular for some $t_0\in T$. Then because the fibers of $\mathcal{X}\rightarrow T$ are 2a or 2b surfaces, the branch divisor $\mathcal{B}_{t_0}$ of the map $f|_{t_0}$ is disjoint from the node in $\mathcal{Z}_{t_0}$.
Let $\sigma_1: \mathcal{Z}_1\rightarrow \mathcal{Z}$ be a simultaneous resolution of singularities of $\mathcal{Z}$. Then the simultaneous resolutions $\sigma_1$ and $\psi$ are locally analytically isomorphic in a neighborhood of each singularity of $\mathcal{Z}$, because the branch divisor $\mathcal{B}$ does not intersect the singularities of $\mathcal{Z}$. Thus, no finite base change of $T$ is required to construct $\mathcal{Z}_1$. Letting $\mathcal{Y}_1$ denote the double cover $f_1$ of $\mathcal{Z}_1$ branched over the preimage $\mathcal{B}_1$ of $\mathcal{B}$, there is a map $\psi_1:\mathcal{Y}_1\rightarrow\mathcal{Y}$ such that the following diagram is commutative:
\[\xymatrix{
\mathcal{Y}_1\ar[r]^{f_1} \ar[d]_{\psi_1} & \mathcal{Z}_1 \ar[d]^{\sigma_1}\\
\mathcal{Y}\ar[r]^f& \mathcal{Z}
}\]
Now let $\mathcal{B}_1$ denote the preimage of $\mathcal{B}$ under $\sigma_1$. On each fiber, $\mathcal{B}_1$ has two $A_1$ singularities. Let $X$ denote the double section of $\mathcal{B}_1\rightarrow T$ passing through these singularities, and let $\sigma_2:\tilde{\mathcal{Z}}\rightarrow \mathcal{Z}_1$ be the blowup of $\mathcal{Z}_1$ in $X$. Then $\sigma_2$ is a simultaneous embedded resolution of singularities of $\mathcal{B}_1$. Thus, no finite base change of $T$ is required to construct $\mathcal{Z}$, and the map $j:\mathcal{X}\rightarrow\tilde{\mathcal{Z}}$ defined by $\sigma_1\circ\sigma_2\circ j=f\circ\psi$ is a double cover over $T$.
\end{proof}
Thus, the space of equisingular deformations of $W$ contains a smooth $39$-dimensional component corresponding to the closure of the 2a locus in $\overline{\mathcal{M}}_{5,5}$.
\subsection{Three technical lemmas}\label{technicalities}
Our goal is to describe the degree two part of the Schouten bracket. We use a method similar to~\cite{horikawa1975, horikawa1976}. In this section, we prove three technical lemmas, analogous to Lemmas 24, 29, and 31 in~\cite{horikawa1975}.
We recall the definition of the Kuranishi deformation space in more generality. Suppose that $X$ is a smooth surface, and let $\textrm{Def}(X)$ be the space of deformations of $X$. The tangent space to $\textrm{Def}(X)$, that is the space of first order infinitesimal deformations of $X$, is isomorphic via the Kodaira--Spencer map to the cohomology group $H^1(X, T_X)$. Let $\rho_1,\ldots, \rho_{n}$ be a basis of $H^1(X, T_X)$, and let $t_1,\ldots t_n$ be a dual basis. Then $\textrm{Def}(X)$ is locally analytically isomorphic to a subspace of $\mathbb{C}^{n}$ with coordinates $t_1,\ldots, t_{n}$, and is given by the kernel of the Kuranishi map $k: H^1(X, T_X)\rightarrow H^2(X, T_X)$, which is a certain infinite series in $t_1,\ldots t_n$. Catanese's article \cite{cataneseguide} gives an excellent exposition of the construction of the Kuranishi map. For us, the important part is that the degree two part of the Kuranishi map is given by the Schouten bracket, which we now describe.
The Schouten bracket is the bilinear map
$$[,]: H^1(X, T_X)\otimes H^1(X, T_X)\rightarrow H^2(X, T_X)$$
defined as the composition of the cup product
$\cup: H^1(X, T_X)\otimes H^1(X, T_X) \rightarrow H^2(X, T_X\otimes T_X)$ followed by the Lie bracket $H^2(X, T_X\otimes T_X)\rightarrow H^2(X, T_X)$.
If $S_{\rho}$ is the infinitesimal first order deformation corresponding, via the Kodaira--Spencer map, to $\rho\in H^1(X, T_X)$, then $[\rho, \rho]$ is the cohomology class corresponding to the obstruction to extending the deformation $S_{\rho}$ to the second order.
\begin{Lemma}\label{zeta is surjective} The map
$$\zeta_*:H^1(X, T_X(\log C))\rightarrow H^1(F_1\amalg F_2, \mathcal{N}_{F_1\amalg F_2})$$
induced by the surjection $T_X|_{F_1\amalg F_2}\rightarrow \mathcal{N}_{F_1\amalg F_2}$ is surjective.
\end{Lemma}
\begin{proof}
It suffices to show that
$$H^1(\tilde{\mathbb{F}}_2, f_*(T_X(\log C)))\rightarrow H^1(\Delta_0, f_*(\mathcal{N}_{F_1\amalg F_2}))$$
is surjective. To do this, recall that the surface $\tilde{\mathbb{F}}_2$ admits an action of $\mathbb{Z}/2\mathbb{Z}$ via deck transformations. By Lemma~\ref{decomposition into eigenspaces}, the sheaf $f_*(T_X(\log C))$ decomposes into invariant and anti-invariant eigenspaces as
\begin{eqnarray*}
f_*(T_X(\log C))_+= T_{\tilde{\mathbb{F}}_2}(\log (\tilde{D}+\tilde{B})) & \textrm{ and }& f_*(T_X(\log C))_-= T_{\tilde{\mathbb{F}}_2}(\log \tilde{D})\otimes \mathcal{O}(-\tilde{L}).
\end{eqnarray*}
We have a similar decomposition of $f_*(\mathcal{N}_{F_1\amalg F_2})$ as follows. By the projection formula, we have
$$f_*(\mathcal{N}_{F_1\amalg F_2})=f_*(f^*(\mathcal{N}_{\Delta_0}))= \mathcal{N}_{\Delta_0}\otimes(\mathcal{O}_{\tilde{\mathbb{F}}_2} \oplus \mathcal{O}_{\tilde{\mathbb{F}}_2}(-\tilde{L})).$$
Thus,
\begin{eqnarray*}
f_*(\mathcal{N}_{F_1\amalg F_2}))_+=\mathcal{N}_{\Delta_0} & \textrm{ and } & f_*(\mathcal{N}_{F_1\amalg F_2}))_-=\mathcal{N}_{\Delta_0}\otimes\mathcal{O}(-\tilde{L}) \simeq \mathcal{N}_{\Delta_0}.
\end{eqnarray*}
We show that the maps
$$\zeta_+:H^1(\tilde{\mathbb{F}}_2, T_{\tilde{\mathbb{F}}_2}(\log (\tilde{D}+\tilde{B})))\rightarrow H^1(\Delta_0, \mathcal{N}_{\Delta_0}) $$
and
$$\zeta_-:H^1(\tilde{\mathbb{F}}_2, T_{\tilde{\mathbb{F}}_2}(\log \tilde{D})\otimes \mathcal{O}(-\tilde{L}))\rightarrow H^1(\Delta_0, \mathcal{N}_{\Delta_0})$$
are surjective.
To show the first, we have the exact sequence
$$0\rightarrow T_{\tilde{\mathbb{F}}_2}(\log(\Delta_0+\tilde{D}+\tilde{B}))\rightarrow T_{\tilde{\mathbb{F}}_2}(\log(\tilde{D}+\tilde{B}))\rightarrow \mathcal{N}_{\Delta_0}\rightarrow 0$$
and so it suffices to show that $H^2(\tilde{\mathbb{F}}_2, T_{\tilde{\mathbb{F}}_2}(\log(\Delta_0+\tilde{D}+\tilde{B})))=0$. By Serre duality, this is equivalent to the vanishing of $H^0(\tilde{\mathbb{F}}_2, \Omega^1_{\tilde{\mathbb{F}}_2}(\log(\Delta_0+\tilde{D}+\tilde{B}))\otimes\mathcal{O}(K))$. This is the statement of Lemma~\ref{coh vanishing on the blowup of F2}.
For the second, note that we have the exact sequence
$$0\rightarrow T_{\tilde{\mathbb{F}}_2}(\log\tilde{D}+\Delta_0)\otimes \mathcal{O}(-\tilde{L})\rightarrow T_{\tilde{\mathbb{F}}_2}(\log\tilde{D})\otimes \mathcal{O}(-\tilde{L})\rightarrow \mathcal{N}_{\Delta_0}\rightarrow 0.$$
By Lemma~\ref{where the obstruction happens}, we have $H^2(T_{\tilde{\mathbb{F}}_2}(\log\tilde{D})\otimes \mathcal{O}(-\tilde{L}))=\mathbb{C}$. Moreover, $H^2(\mathcal{N}_{\Delta_0})=0$, and thus the map $\zeta_-$ is surjective as long as $H^2(\tilde{\mathbb{F}}_2, T_{\tilde{\mathbb{F}}_2}(\log \tilde{D}+\Delta_0)\otimes\mathcal{O}(-\tilde{L}))=\mathbb{C}$. Equivalently, we show that $H^0(\tilde{\mathbb{F}}_2, \Omega^1_{\tilde{\mathbb{F}}_2}(\log \tilde{D}+\Delta_0)\otimes\mathcal{O}(K+\tilde{L}))=\mathbb{C}$.
By Lemma~\ref{blowup of two points} and the projection formula, we have
$$\sigma_*\Omega^1_{\tilde{\mathbb{F}}_2}(\log (\tilde{D}+\Delta_0))\otimes\mathcal{O}(K+\tilde{L})=\Omega^1_{\mathbb{F}_2}(\log(D+\Delta_0))\otimes\mathcal{O}(\Delta)\otimes\mathfrak{M}_{p,q}.$$
So we now want
$$H^0(\mathbb{F}_2, \Omega^1_{\mathbb{F}_2}(\log(D+\Delta_0))\otimes\mathcal{O}(\Delta)\otimes\mathfrak{M}_{p,q})=\mathbb{C}.$$
We claim that the sheaf $T_{\mathbb{F}_2}(\log(D+\Delta_0))$ fits into an exact sequence as
$$0\rightarrow\mathcal{O}(G)\rightarrow T_{\mathbb{F}_2}(\log(D+\Delta_0))\rightarrow \pi^*T_{\mathbb{P}^1}(-a)\rightarrow 0$$
where $G$ is a divisor on $\mathbb{F}_2$ and $\pi(D)=a\in\mathbb{P}^1$. To see this, note first that $\pi^*T_{\mathbb{P}^1}(-a)\simeq \mathcal{O}_{\mathbb{F}_2}(D)$. Let $U\subset \mathbb{F}_2$ be an open neighborhood of the point $0\in D\cap \Delta_0$ with coordinates $(x,y)$ so that $D$ has local equation $x$ and $\Delta_0$ has local equation $y$. Then the map
$$T_{\mathbb{F}_2}(\log(D+\Delta_0))\rightarrow \mathcal{O}_{\mathbb{F}_2}(D)$$
is locally given by
$$x\frac{\partial}{\partial x} \mapsto x\textrm{\hspace{1cm} } y\frac{\partial}{\partial y}\mapsto 0.$$
Thus the map is surjective. Since $T_{\mathbb{F}_2}(\log(D+\Delta_0))$ is a vector bundle of rank two and $\mathcal{O}_{\mathbb{F}_2}(D)$ is a line bundle, the kernel of the map $T_{\mathbb{F}_2}(\log(D+\Delta_0))\rightarrow \mathcal{O}_{\mathbb{F}_2}(D)$ is a vector bundle of rank one. All such vector bundles are given by $\mathcal{O}_{\mathbb{F}_2}(G)$ for some divisor $G$ on $\mathbb{F}_2$. We find $G$ by calculating the Chern class of $T_{\mathbb{F}_2}(\log(D+\Delta_0))$.
The determinant line bundle $\bigwedge^2T_{\mathbb{F}_2}(\log(D+\Delta_0))$ is given by $-\mathcal{O}(-K_{\mathbb{F}_2}-D-\Delta_0)=\mathcal{O}(\Delta_0+3\Gamma)$, so $c_1(T_{\mathbb{F}_2}(\log( D+\Delta_0)))=\Delta_0+3\Gamma$.
Thus, $G = \Delta_0+2\Gamma$.
Dualizing the above exact sequence and tensoring with $\mathcal{O}(\Delta)\otimes \mathfrak{M}_{p,q}$ results in the exact sequence
$$0\rightarrow \mathcal{O}(\Delta_0+\Gamma)\otimes \mathfrak{M}_{p,q}\rightarrow \Omega^1_{\mathbb{F}_2}(\log D+\Delta_0)\otimes \mathcal{O}(\Delta)\otimes \mathfrak{M}_{p,q}\rightarrow \mathcal{O}_{\mathbb{F}_2}\otimes \mathfrak{M}_{p,q}\rightarrow 0.$$
The sheaf on the right has no global sections, since the only section of $\mathcal{O}_{\mathbb{F}_2}$ vanishing at $p$ and $q$ is zero. Moreover, since $\Delta_0\cdot (\Delta_0+\Gamma)=-1$ every divisor in the linear system $|\Delta_0+\Gamma|$ is a union of two divisors $\Delta_0$ and $\Gamma$. Since there is only one such divisor passing through $p$ and $q$, namely the divisor $\Delta_0+D$, we have
$$ H^0(\mathbb{F}_2, \Omega^1_{\mathbb{F}_2}(\log D+\Delta_0)\otimes \mathcal{O}(\Delta)\otimes \mathfrak{M}_{p,q})\simeq H^0(\mathbb{F}_2, \mathcal{O}(\Delta_0+D)\otimes \mathfrak{M}_{p,q} )=\mathbb{C},$$
as we wished to show.
\end{proof}
A key ingredient of Horikawa's description in~\cite{horikawa1976} is a map
$$\gamma: H^1(X, T_X)\rightarrow H^0(G, \mathcal{O}(K_X|_{G}) ),$$
where $K_X=2G+F$ and $G$ is a generic fiber of the map $g: X\rightarrow \mathbb{P}^1$.
\begin{Lemma}\label{ideal} Let $X$ be a smooth surface with a surjective map $g: X\rightarrow \mathbb{P}^1$ such that $g_*\mathcal{O}_X=\mathcal{O}_{\mathbb{P}^1}$ and let $G$ denote a generic fiber of $g$. Suppose that $K_X=2G+F$ for some smooth divisor $F$ on $X$ such that $G\not\subset F$ and let
$$\zeta_*:H^1(X, T_X)\rightarrow H^1(F, \mathcal{N}_{F})$$
be the map induced by the surjection $T_X|_{F}\rightarrow \mathcal{N}_{F}$.
If the irregularity $q(X)=0$, $h^1(X, \mathcal{O}(G))=0$,
and $h^0(F, \mathcal{O}(K-G)|_{F})=0$, then there is a map
$\gamma: H^1(X, T_X)\rightarrow H^0(F, \mathcal{O}(K_X|_{F}) )$, defined in~\cite[4.8]{mythesis}
with the property that $\textrm{Ker }\gamma=\textrm{Ker }\zeta_*$.
\end{Lemma}
\begin{proof}
This is a generalization of the map defined in~\cite{horikawa1976}, Section 5, after the proof of Lemma 24. See~\cite{mythesis} for details.
\end{proof}
\begin{Lemma}\label{zero bracket gives restriction} With the same hypotheses as Lemma~\ref{ideal}, if $[\rho, \rho]=0$ then $(\gamma(\rho))^2$ is in the image of the restriction map $H^0(X, \mathcal{O}(2K))\rightarrow H^0(F, \mathcal{O}(2K|_{F}))$.
\end{Lemma}
\begin{proof}
The proof is a generalization of that of Lemma 31 in~\cite{horikawa1975}. See~\cite{mythesis} for details.
\end{proof}
\subsection{Proof of the main theorem}\label{main proof}
We describe the space $\textrm{Def}^{QG}(W)$ of $\mathbb{Q}$-Gorenstein deformations of a general 2b surface $W$.
\begin{Lemma}\label{coh works well}\cite[Lemma 6.3]{horikawa1976} Let $X$ be the minimal resolution of a surface of type 2b. Then $h^1(X, \mathcal{O}(G))=2$ and $h^1(X, \mathcal{O}(G+F_1+F_2))=0$.
\end{Lemma}
\begin{proof}Horikawa proves this in the case that $X$ is a double cover of $\mathbb{F}_2$ with a smooth branch divisor. The proof uses Riemann-Roch and Serre duality together with the fact that the canonical divisor on $X$ is given by $K_X=2G+F_1+F_2$. Because it only relies on numerical characteristics of $X$, $G$, $F_1$ and $F_2$, Horikawa's proof works in our case as well.
\end{proof}
By Lemma~\ref{coh works well}, we can define the map $\gamma$ as in Lemma~\ref{ideal}, where $F=F_1+F_2$. By abuse of notation, we let
$$\gamma: H^1(X, T_X(\log C))\rightarrow H^0(F_1\amalg F_2, \mathcal{O}_{F_1\amalg F_2})$$
be the restriction of this map to $H^1(X, T_X(\log C))$. We note that this map is the restriction to $H^1(X, T_X(\log C))\subset H^1(X, T_X)$ of the corresponding map defined in~\cite{horikawa1976} under the assumption that the branch locus is smooth.
\begin{proof}[Proof of Theorem~\ref{the best}]
The deformation space $\textrm{Def}^{QG, \mbox{e.s}}(W)$ is locally analytically isomorphic to the zero-set of the Kuranishi map
$$k: H^1(X, T_X(\log C))\rightarrow H^2(X, T_X(\log C))=\mathbb{C}.$$ Choose a basis $\rho_1,\rho_2,\ldots, \rho_{40}$ of $H^1(X, T_X(\log C))$.
Let $t_1, t_2,\ldots, t_{40}$ be the dual basis. A priori, the Kuranishi map is some power series in $t_1,\ldots, t_{40}$. However in this case, Theorems~\ref{2a to 2b come from double covers} and~\ref{counting} imply that $\textrm{Def}^{QG, \mbox{e.s}}(W)$ contains a smooth $39$-dimensional subspace corresponding to deformations of a 2b surface to a 2a surface.
This implies that if we choose a basis $\rho_1,\rho_2,\ldots, \rho_{40}$ of $H^1(X, T_X(\log C))$ such that $\rho_1\in H^1_-(X,T_X(\log C))$ and $\rho_i\in H^1_+(X, T_X(\log C))$ for $i>2$, then the corresponding dual basis has the property that the Kuranishi function factors into at least two terms, one of which has linear term $t_1$. To show that $\textrm{Def}^{QG, \mbox{e.s}}(W)$ is locally a product of two smooth $39$-dimensional components meeting transversally in a 38-dimensional component, it therefore suffices to show that the degree two part of the Kuranishi map is nonzero and not a square. The degree two part is given by the Schouten bracket, defined above.
We restrict the Schouten bracket $[,]$ to $H^1(X, T_X(\log C))\otimes H^1(X, T_X(\log C))$. We claim that the Lie bracket $H^2(X, T_X(\log C)\otimes T_X(\log C))\rightarrow H^2(X, T_X)$ has image in $H^2(X, T_X(\log C))$. Let $\{U_i\}$ be a sufficiently fine open covering of $X$ and let $U_{ijk}=U_i\cap U_j\cap U_k$. Let $\rho$ be an element of $H^2(X, T_X(\log C)\otimes T_X(\log C))$, represented by a 2-cocycle $\{\rho_{ij}\otimes\rho_{jk}\}$, where $\{\rho_{ij}\}$ is a 1-cocycle with coefficients in $T_X(\log C)$. Then $\rho_{ij}$ is a vector field that fixes the ideal sheaf of the $C$. Thus, $\rho_{ij}\otimes\rho_{jk}$ also fixes the ideal sheaf of $C$. Therefore the Lie bracket
$[\rho_{ij}, \rho_{jk}]$ gives a vector field on $U_{ijk}$ which also fixes the ideal sheaf of $C$. Thus, the form $[,]$ gives a $2$-cocycle with coefficients in $T_X(\log C)$; that is
$$[,]: H^1(X, T_X(\log C))\otimes H^1(X, T_X(\log C))\rightarrow H^2(X, T_X(\log C))\simeq \mathbb{C}.$$
Because $\rho_1,\ldots, \rho_{40}$ and $t_1,\ldots t_{40}$ are dual bases, the degree two part of the Kuranishi map $k$ is given by
$$\sum_{1\le i,j\le 40}[\rho_i,\rho_j]t_it_j.$$
Moreover, because $k$ factors into a product, one term of which has linear term $t_1$, we have that $[\rho_i, \rho_j]=0$ for $2\le i, j\le 40$. It therefore suffices to show that $[\rho_1, \rho_1]=0$ and $[\rho_1, \rho_i]$ is nonzero for some $i>1$.
Recall that $K_X=2G+F_1+F_2$ and consider the exact sequence
$$0\rightarrow \mathcal{O}_X(2G)\rightarrow \mathcal{O}_X(K)\rightarrow \mathcal{O}_{F_1}\oplus\mathcal{O}_{F_2}\rightarrow 0.$$
On $X$, we have $p_g=4$ and $h^0(2G)=3$, so the image of the map $r: H^0(X, \mathcal{O}_X(K))\rightarrow H^0(X, \mathcal{O}_{F_1}\oplus\mathcal{O}_{F_2})$ is one-dimensional.
Moreover, the image of $r$ is contained in the ``diagonal" in $H^0(X, \mathcal{O}_{F_1}\oplus\mathcal{O}_{F_2})\simeq \mathbb{C}^2$. That is, if $X$ is a nonzero global section of $\mathcal{O}_{F_1}\oplus\mathcal{O}_{F_2}$ in the image of $r$, then $s|_{F_1}\neq 0$ and $s|_{F_2}\neq 0$. More precisely, we have the commutative diagram below, where the arrow on the left is an isomorphism and the inclusion of $H^0(\tilde{\mathbb{F}_2}, \Delta_0)$ into $H^0(X, \mathcal{O}_{F_1}\oplus\mathcal{O}_{F_2})$ sends a section to the section of $\mathcal{O}_{F_1}\oplus\mathcal{O}_{F_2}$ whose restrictions to $F_1$ and $F_2$ are equal.
$$
\xymatrix{ H^0(X, \mathcal{O}_X(K))\ar[r]^r& H^0(X, \mathcal{O}_{F_1}\oplus\mathcal{O}_{F_2})\\
H^0(\tilde{\mathbb{F}}_2, \Delta_0+2\Gamma)\ar@{^{(}->}[u]_{\simeq}\ar[r]&H^0(\tilde{\mathbb{F}}_2, \Delta_0)\ar@{^{(}->}[u]
}
$$
By Lemmas~\ref{ideal}, \ref{coh works well} and~\ref{zeta is surjective}, the map $\gamma: H^1(X, T_X(\log C))\rightarrow H^0(X,\mathcal{O}_{F_1}\oplus\mathcal{O}_{F_2})$ is surjective. Thus, we can choose $\rho\in H^1(X, T_X(\log C))$ such that $\gamma(\rho)\neq 0$, and $\gamma(\rho)^2$ is not in the image of $r$. But then $\gamma(\rho)^2$ is not a restriction of an element of $H^0(X, \mathcal{O}(2K))$, so by Lemma~\ref{zero bracket gives restriction}, we conclude that $[\rho,\rho]\neq0$. Thus, the Schouten bracket
$$[,]:H^1(X, T_X(\log C))\times H^1(X,T_X( \log C))\rightarrow H^2(X, T_X(\log C))\simeq \mathbb{C}$$ is surjective.
Because it is locally given by the composition of the cup product followed by the Lie bracket of vector fields, the Schouten bracket is $\mathbb{Z}/2\mathbb{Z}$-equivariant under the action of $\mathbb{Z}/2\mathbb{Z}$ by deck transformations. By Lemma~\ref{obstruction}, the invariant part of $H^2(X, T_X(\log C))$ is zero, and so $[\rho_i,\rho_j]$ is nonzero if and only if $[\rho_i,\rho_j]$ is anti-invariant under the action of $\mathbb{Z}/2\mathbb{Z}$. Suppose that $\rho\otimes\eta$ is an element of $H^1(X,T_X(\log C))\otimes H^1(X, T_X(\log C))$, where $\rho$ and $\eta$ are either both invariant or both anti-invariant. Then $[\rho, \eta]$ is invariant, that is $[\rho,\eta]\in H_+^2(X, T_X(\log C))$. By Lemma~\ref{obstruction}, this space is zero, so $[\rho,\eta]=0$. Thus, by choice of basis, $[\rho_i, \rho_i]=0$ for all $i$; in particular, $[\rho_1, \rho_1]=0$.
Now suppose that $\rho\in H^1_+(X, T_X(\log C))$ is invariant and $\eta\in H^1_-(X, T_X(\log C))$ is anti-invariant. Then $[\rho,\eta]\in H^2_-(X, T_X(\log C))$ is anti-invariant. Since $[, ]$ is surjective, there exists, by choice of basis, $i>1$ such that $[\rho_1, \rho_i]\neq0$, completing the proof.
\end{proof}
\bibliographystyle{alpha}
|
\section{Introduction}
Let $k$ be an arbitrary field and $\kx= k[x_1,\ldots,x_n]$ and $\kt
= k[t_1,\ldots,t_m]$ two polynomial rings over $k$. A {\it binomial}
$f$ in $\kx$ is a difference of two monomials. Let $\mB
=\{b_1,\ldots,b_n\}$ be a set of nonzero vectors in $\N^m$; each
vector $b_i = (b_{i1},\ldots,b_{im})$ corresponds to a monomial
$\t^{b_i} = t_1^{b_{i1}} \cdots t_m^{b_{im}}$ in $\kt$.
\medskip The {\it toric ideal} determined by $\mB$ is the kernel of
the homomorphism of $k$-algebras $$\varphi\colon \kx \rightarrow
\kt;\ x_i\longmapsto \t^{b_i}$$ and is denoted by
$\IB$. By \cite[Corollary~4.3]{Sturm}, $\IB$ is an $\mB$-{\it homogeneous binomial\/} ideal, i.e.,
if one sets the $\mB$-degree of a monomial $\x^{\alpha} \in \kx$ as
${\rm deg}_{\mB}(\x^{\alpha}) := \alpha_1 b_1 + \cdots + \alpha_n
b_n \in \N^m$, and says that a polynomial $f \in k[x]$ is
$\mB$-homogeneous if its monomials have the same $\mB$-degree, then
$\IB$ is generated by $\mB$-homogeneous binomials. According to
\cite[Lemma 4.2]{Sturm}, the height of $\IB$ is ${\rm ht}(\IB) =
n-{\rm rk}(\Z \mB)$, where ${\rm rk}(\Z \mB)$ denotes the rank of
the subgroup of $\Z^m$ generated by $\mB$. The ideal $\IB$ is a {\it
complete intersection} if $\mu(\IB) = {\rm ht}(\IB)$, where
$\mu(\IB)$ denotes the minimal number of generators of $\IB$.
Equivalently, $\IB$ is a complete intersection if there exists a set
of $s = n-{\rm rk}(\Z \mB)$ $\mathcal B$-homogeneous binomials
$g_1,\ldots,g_{s}$ such that $\IA=(g_1,\ldots,g_{s}).$
\smallskip Let $\mA = \{d_1,\ldots,d_n\}$ be a subset of the positive
integers and consider the {\it affine monomial curve} $\mathcal C$
given parametrically by $x_1 = t^{d_1},\ldots,x_n = t^{d_n}$, i.e.,
$$
\mC = \{(t^{d_1},\ldots,t^{d_n})\in\mathbb{A}_k^n\,\vert\, t\in
k\}\,.
$$
By \cite[Corollary~7.1.12]{monalg}, if $k$ is an infinite field the
ideal $I(\mathcal C)$ of polynomials vanishing on $\mathcal C$ is
equal to $\IA$, the toric ideal determined by $\mA$. $\IA$ is called
the {\it toric ideal of} $\mathcal C$.
\smallskip Set $d := {\rm max}(\mA)$ and consider $\omA := \{a_1, \ldots, a_{n-1}, a_n, a_{n+1}\} \subset \N^2,$
where $a_i := (d_i,d - d_i)$ for all $i \in \{1,\ldots,n\}$ and
$a_{n+1} := (0,d)$, and the {\it projective monomial curve $\omC$}
obtained as the projective closure of $\mC$, which is given
parametrically by $x_1 = t^{d_1} u^{d-d_1},\ldots,x_n = t^{d_n} u
^{d-d_n},x_{n + 1} = u^d$, i.e.,
$$
\omC = \{(t^{d_1}u^{d-d_1}:\cdots:t^{d_n}u^{d - d_n}:u^d)\in
\P_k^n\,\vert\, (t:u) \in \P_k^1\}\,.
$$
\noindent Again by \cite[Corollary~7.1.12]{monalg}, if $k$ is an
infinite field, then $I(\omC) = \oIA$. The toric ideal $\oIA$ is
homogeneous, indeed $\oIA$ is the homogenization of $\IA$ with
respect to the variable $x_{n+1}$. $\oIA$ is called the {\it toric
ideal of} $\omC$.
\smallskip
Both $\IA$ and $\oIA$ have height $n-1$; thus $\IA$ (resp. $\oIA$)
is a {\it complete intersection\/} if there exists a system of
$\mA$-homogeneous (resp. homogeneous) binomials $g_1,\ldots,g_{n-1}$
such that $\IA=(g_1,\ldots,g_{n-1})$ (resp. $\oIA =
(g_1,\ldots,g_{n-1}) ).$ Clearly, $\IA$ is a complete intersection
whenever $\oIA$ is, the converse is not true in general.
\smallskip
The aim of this work is to provide families of positive integers
$\mA = \{d_1,\ldots,d_n\}$ such that either the toric ideal $\IA$ or
$\oIA$ is a complete intersection. The starting point are the papers
by Garc\'{\i}a-S\'anchez and Rosales \cite{GR99}, Maloo and Sengupta
\cite{MS} and Fel \cite{Fel}.
\smallskip
In the first one, the authors prove that if $\mA$ is a set of
consecutive positive integers, then $\IA$ is a complete intersection
if and only if $n = 2$ or $n = 3$ and $d_1$ is even. In the second
one, the authors obtain the same characterization when $\mA$ is an
arithmetic sequence provided $\gcd(\mA) = 1$. In Theorem \ref{gen}
we generalize this result to {\it generalized arithmetic sequences}.
We recall that $\mA$ is a generalized arithmetic sequence if there
exists $h \in \Z^+$ such that $\{h d_1,d_2,\ldots,d_n\}$ is an
increasing arithmetic sequence.
\smallskip
Maloo and Sengupta also study the case in which $\mA$ is an
almost-arithmetic sequence, i.e., $\mA \setminus \{d_n\}$ is an
arithmetic sequence and $d_n \in \Z^+$, and prove that $n \leq 4$
provided
$\IA$ is a complete intersection. In Theorem \ref{almost} we go
further and characterize when $\IA$ is a complete intersection whenever $\mA
\setminus \{d_n\}$ is a generalized arithmetic sequence and $d_n \in
\Z^+$, where $n \geq 4$.
\smallskip In the third one, the author provides certain conditions so
that the semigroup generated by $\mA = \{d_1,d_2,d_3\}$ where
$\gcd\{d_1,d_2,d_3\} = 1$ and $d_1,d_2,d_3$ are members of the
Fibonacci or Lucas sequence is symmetric. Recall that for a set $\mA
= \{d_1,\ldots,d_n\}$ with $\gcd(\mA) = 1$, setting $\mS := \sum_{i
= 1}^n \N\, d_i$, then the complement $\mS$ in $\N$ is finite and
the largest integer not belonging to $\mS$ is called the {\it
Frobenius number of $\mS$} and denoted by $\g(\mS)$. Moreover, the
semigroup $\mS$ is {\it symmetric} if for every $d \in \Z$, either
$d \in \mS$ or $\g(\mS) - d \in \mS$. It is a classical result due
to Herzog \cite[Theorem 3.10]{Herzog} that whenever $\mA =
\{d_1,d_2,d_3\} \subset \Z^+$ with $\gcd\{d_1,d_2,d_2\} = 1$, then
$\mS$ is symmetric if and only if $\IA$ is a complete intersection.
In this work we characterize the complete intersection property
for $\IA$ when $\mA$ is a certain subset of either the
$(p,q)$-Fibonacci sequence or the $(p,q)$-Lucas sequence, where $p,q
\in \Z^+$ are relatively prime. We recall that the $(p,q)$-Fibonacci
sequence, denoted by $\{F_n\}_{n\in \N}$, is defined as follows
\begin{center}$F_0 = 0,\,F_1 = 1$ and $F_{n+2} = p\, F_{n+1} + q\, F_n$ for every $n \geq 0$,
\end{center}
and the $(p,q)$-Lucas sequence, denoted by $\{L_n\}_{n \in \N}$, is
defined as \begin{center} $L_0 = 2,\,L_1 = p$ and $L_{n+2} = p\,
L_{n+1} + q\, L_n$ for every $n \geq 0$.
\end{center}
These sequences are natural generalizations of the usual Fibonacci
and Lucas sequences, now called $(1,1)$-Fibonacci sequence and
$(1,1)$-Lucas sequence. In Theorems \ref{fibonacci} and \ref{lucas}
we characterize when $\IA$ is a complete intersection, being $\mA =
\{d_1,\ldots,d_n\}$ with:
\begin{enumerate}
\item[(a)] $d_i = F_{e_i}$, where
$\{e_1,\ldots,e_n\}$ is a generalized arithmetic sequence, and
\item[(b)] $d_i = L_{e_i}$, where
$\{e_1,\ldots,e_n\}$ is an arithmetic sequence.
\end{enumerate}
\smallskip
Moreover, we characterize algorithmically when the toric ideal of a
projective monomial curve is a complete intersection with Algorithm
CI-projective-monomial-curve (see Table
\ref{algoritmocurvamonomialproyectiva}). Using this algorithm we are
able to characterize in Theorems \ref{genproy}, \ref{almostN=4proy},
\ref{fibonacciproy} and \ref{lucasproy}, which are the projective
versions of Theorems \ref{gen}, \ref{almost}, \ref{fibonacci} and
\ref{lucas} respectively, when $\oIA$ is a complete intersection
when $\mA$ belongs to any of the four families already described.
\smallskip
Whenever $\IA$ or $\oIA$ is a complete intersection, we also obtain
a minimal set of generators of the toric ideal. Furthermore, when
$\IA$ is a complete intersection, using the formula described in
\cite[Remark 11]{D} and \cite[Remark 4.5]{BGRV}, the Frobenius
number $\g(\mathcal S)$ of the numerical semigroup $\mathcal S :=
\sum_{i = 1}^n \N \, (d_i / e)$, where $e := \gcd(\mA)$, is also
provided. Indeed, the formula asserts that if $\IA =
(g_1,\ldots,g_{n-1})$ where $g_i$ is $\mA$-homogeneous for all $i
\in \{1,\ldots,n-1\}$, then $ \g(\mathcal S) =
\left(\sum_{i=1}^{n-1} {\rm deg}_{\mA}(g_i) - \sum_{i = 1}^n d_i
\right) / e$.
\smallskip
The results obtained in this work arise as consequences of our
papers \cite{BGS, BGsimplicial} and some new results concerning the
toric ideal of the curves. In \cite{BGS} we exploited the
combinatorial-arithmetical structure of complete intersections given
by the existence of a certain {\it binary tree labeled by}
$\{d_1,\ldots,d_n\}$ stated in \cite[Theorem~4.3]{BGRV} to obtain an
algorithm that determines whether the toric ideal of an affine
monomial curve is a complete intersection. In \cite{BGsimplicial} we
provided some new results concerning complete intersection toric
ideals in general and apply them in order to obtain algorithms that
characterize when either a simplicial toric ideal or a homogeneous
simplicial toric ideal is a complete intersection. Since both $\IA$
and $\oIA$ are simplicial toric ideals and, moreover, $\oIA$ is
homogeneous, the algorithms obtained in \cite{BGsimplicial} apply to
them. These algorithms have been implemented in ANSI C programming
language and also in the distributed library {\tt cisimplicial.lib}
\cite{BGlib2} of {\sc Singular} \cite{DGPS}.
\section{Complete intersection toric ideals associated to affine and projective monomial curves}
This section is devoted to present some new results concerning the
complete intersection property for toric ideals associated to either
affine or projective monomial curves. The results of this section
arise after applying some results of \cite{BGsimplicial} to the
context of affine and projective monomial curves. The main results
of this section are namely Proposition \ref{n-1dist} and Theorem
\ref{coralg}. On one hand, Proposition \ref{n-1dist} provides, under
certain hypothesis, a necessary and sufficient condition for the
toric ideal of an affine monomial curve to be a complete
intersection. This result will be useful in Sections $3, 4$ and $5$.
On the other hand, Theorem \ref{coralg} is a particularization of
\cite[Corollary 5]{BGsimplicial} for toric ideals associated to
projective monomial curves. This result, together with Remark
\ref{notaproyred}, yields Algorithm CI-projective-monomial-curve of
Table \ref{algoritmocurvamonomialproyectiva}, an algorithm for
checking whether the toric ideal of a projective monomial curve is a
complete intersection. This algorithm will be useful in Section $6$.
\medskip
In order to present the new results, we begin our explanation by
briefly describing some results of \cite{BGsimplicial}. For every
set $\mB = \{b_1,\ldots,b_n\}$ of nonzero vectors of $\N^m$, we have
the following results.
\begin{Lemma}\cite[Lemmas 2.1 and 2.2]{BGsimplicial}\label{facilred} \begin{itemize} \item If $b_i \notin \sum_{j \in \{1,\ldots,n\} \atop j \neq i} \Q\, b_j$, then $\IB =
I_{\mB \setminus \{b_i\}} \cdot \kx$. Moreover, $\IB$ is a complete
intersection $\Longleftrightarrow \, I_{\mB \setminus \{b_i\}}$ so
is.
\item If $b_i = \sum_{j \in \{1,\ldots,n\} \atop j \neq i} \alpha_j
b_j \in \sum_{j \in \{1,\ldots,n\} \atop j \neq i} \N\, b_j$, then
$\IB = I_{\mB \setminus \{b_i\}} \cdot \kx + (x_i - \prod_{j \in
\{1,\ldots,n\} \atop j \neq i} x_j^{\alpha_j})$. Moreover, $\IB$ is
a complete intersection $\Longleftrightarrow \, I_{\mB \setminus
\{b_i\}}$ so is.
\end{itemize}
\end{Lemma}
\bigskip Whenever $b_i \in \sum_{j \in \{1,\ldots,n\} \atop j\neq i}
\Q\, b_j$ we denote $B_i := {\rm min}\{B \in \Z^+\, \vert \, B b_i
\in \sum_{j \in \{1,\ldots,n\} \atop j \neq i} \Z b_j\}$ and have
the following result.
\begin{Proposition}\cite[Proposition 2.3]{BGsimplicial}\label{red}
Assume that $b_i \in \sum_{j \in \{1,\ldots,n\} \atop j\neq i} \Q\,
b_j$ for some $i \in \{1,\ldots,n\}$ and set $\mB' :=
\{b_1,\ldots,B_i b_i, \ldots, b_n\}$ and $\rho: \kx \longrightarrow
\kx$ the morphism induced by $\rho(x_i) = x_i^{B_i}$, $\rho(x_j) =
x_j$ for every $j \neq i$. Then, $\IB = \rho(I_{\mB'}) \cdot \kx$.
Moreover, $\IB$ is a complete intersection $\Longleftrightarrow$
$I_{\mB'}$ is a complete intersection.
\end{Proposition}
Applying Lemma \ref{facilred} and Proposition \ref{red} iteratively,
we can associate to $\mB$ a unique subset $\mB_{red} \subset \N^m$
which can be either empty or satisfies that $\mB_{red} =
\{b_1^{\,\prime},\ldots,b_r^{\,\prime}\}$, where $r \leq n$ and $
b_i^{\,\prime} \in \sum_{j \in \{1,\ldots,r\}\atop j \neq i} \Z\,
b_j^{\,\prime} \setminus \sum_{j \in \{1,\ldots,r\} \atop j \neq i}
\N\, b_j^{\,\prime}$ for all $i \in \{1,\ldots,r\}$. As a
consequence of this construction we have the following result.
\begin{Theorem}\cite[Theorem 2.5]{BGsimplicial} \label{TheoremReduccion}$\IB$ is a complete intersection $\Longleftrightarrow$ either
$\mB_{red} = \emptyset$ or $I_{\mB_{red}}$ is a complete
intersection.
\end{Theorem}
Let $\mA = \{d_1,\ldots,d_n\}$ be a set of $n \geq 2$ positive
integers, in this setting we have that $B_i = {\rm min}\{B \in \Z^+
\, \vert \, B d_i \in \sum_{j \in \{1,\ldots,n\} \atop j \neq i} \Z
d_j\} = \gcd(\mA \setminus \{d_i\})\, /\, \gcd(\mA)$ for all $i \in
\{1,\ldots,n\}$. For $n = 3$ we have the following result, which is
essentially a rewriting of a classical Herzog's result \cite{Herzog}
(see also \cite[Proposition 3]{Watanabe}).
\begin{Proposition}\label{n=3}If $n = 3$, $\IA$ is a complete intersection
$\Longleftrightarrow\ \mA_{red} = \emptyset$.
\end{Proposition}
For $n > 3$ the same characterization does not hold. Nevertheless,
under certain hypothesis, we prove in Proposition \ref{n-1dist} an
analogous characterization for $\IA$ to be a complete intersection.
To present this result we define
\begin{center}$m_i := {\rm min}\left\{b \in \Z^+ \, \vert \, b d_i
\in \sum_{j \in \{1,\ldots,n\} \atop j \neq i} \N d_j\right\}$ for
every $i \in \{1,\ldots,n\}$,
\end{center} and say that a binomial $f \in \IA$ is {\it critical with
respect to $x_i$} if $f = x_i^{m_i} - \prod_{j \in \{1,\ldots,n\}
\atop j \neq i} x_j^{\alpha_j}$. Critical binomials were introduced
by Eliahou in \cite{E} and later studied by Alc\'antar and
Villarreal in \cite{AV}
\begin{Lemma}\label{lemacritico}Let $f_1,\ldots,f_t$ be critical binomials with respect to
$x_{i_1},\ldots,x_{i_t}$ respectively, where $1 \leq i_1 < \cdots <
i_t \leq n$. If $m_i d_i \neq m_j d_j$ for every $1 \leq i < j \leq
t,$ then there exists a set of binomials $\frak B$ minimally
generating $\IA$ such that $f_1,\ldots,f_t \in \frak B.$
\end{Lemma}
\begin{demo}Let $\{g_1,\ldots,g_{n-1}\}$ be a set of $\mA$-homogeneous binomials
generating $\IA$. Let $D_1,\ldots, D_{n-1}$ be the $\mA$-degrees of
$g_1,\ldots,g_{n-1}$ respectively and suppose that $f_1 \in \IA$ is
a critical binomial with respect to $x_1$, thus $f_1 = x_1^{m_1} -
\prod_{j \in \{2,\ldots,n\}} x_j^{\alpha_j}$ for some
$\alpha_2,\ldots,\alpha_n \in \N$. Hence $f_1 =
q_{1}g_1+\ldots+q_{n-1}g_{n-1}$\,, where $q_{k}\in \kx$ is an
$\mA$-homogeneous polynomial of degree $m_1 a_1 - D_k\geq 0$ when
$q_{k}\neq 0$ for all $k\in \{1,\ldots,n-1\}$.
In particular, there exists $k \in \{1,\ldots,n-1\}$ such that $q_k \neq 0$ and the image of
$g_{k}$ under the evaluation morphism which sends $x_j$ to $0$ for
all $j \neq 1$ is equal to $x_1^{D_{k}/d_1}$, we assume that $k =
1$. Thus \,$g_{1}=x_1^{D_1/d_1}-\x^{\,\beta}$\,, where
$\x^{\,\beta}$ is a monomial of $\mA$-degree $D_1$ which does not
involve the variable $x_1$, and hence $D_1 \in \Z^+ d_1\cap
\sum_{j\in\{2,\ldots,n\}} \N\,d_j$. By the definition of $m_1$ we
get the equality $m_1 a_1 =D_1$, which implies that $q_1 \in k$ and
$\{f_1,g_2,\ldots,g_{n-1}\}$ is a minimal set of generators of
$\IA$. Iterating this argument, we get the result. \QED
\end{demo}
\begin{Proposition}\label{n-1dist} If $n-1$ integers from $m_1 d_1,\ldots,m_n
d_n$ are different, then $\IA$ is a complete intersection
$\Longleftrightarrow$ $\mA_{red} = \emptyset$.
\end{Proposition}
\begin{demo}$(\Leftarrow)$ Follows from Theorem
\ref{TheoremReduccion}. $(\Rightarrow)$ Assume that $m_1 d_1,\ldots,
m_{n-1} d_{n-1}$ are all different, by Lemma \ref{lemacritico} we
get that $\IA = (f_{1},\ldots,f_{n-1})$ where $f_i$ is a critical
binomial with respect to $x_i$ for every $1 \leq i \leq n-1$. We
claim that there exists $j \in \{1,\ldots,n-1\}$ such that $x_j$
does not appear in $f_k$ for all $k \in \{1,\ldots,n-1\} \setminus
\{j\}$. Suppose this claim is false, then we consider the simple
directed graph with vertex set $\{1,\ldots,n-1\}$ and arc set
$\{(j,k)\, \vert\, 1 \leq j,\,k \leq n-1,\ j \neq k$ and $x_j$
appears in $f_k\}$; since the out-degree of every vertex is greater
or equal to one, there is a cycle in the graph. Suppose
that the cycle is $(1,2,\ldots,k,1)$ with $k \leq n-1$, this means that
$(f_{1},\ldots,f_{k}) \subset (x_{1},\ldots,x_{k})$, so $\IA
\subsetneq H := (x_{1},\ldots,x_{k},f_{k+1},\ldots,f_{n-1})$ but
this is not possible because $\IA$ is prime and $n-1 = {\rm ht}(\IA)
< {\rm ht}(H) \leq n-1$.
Thus there exists $i \in \{1,\ldots,n-1\}$ such that $x_i$ only
appears in $f_i$, suppose $i = 1$. Now we write $\gamma_i := m_i e_i
- \sum_{j \in \{1,\ldots,n\} \atop j \neq i} \alpha_{i,j} e_j \in
\Z^n$ for all $i \in \{1,\ldots,n-1\}$. By \cite[Proposition
2.3]{ElVi}, $\{\gamma_{1},\ldots,\gamma_{n-1}\}$ is a $\Z$-basis for
the kernel of the homomorphism $\tau: \Z^n \longrightarrow \Z$
induced by $\tau(e_j) = d_j$. By definition, $B_1 d_1 = \sum_{j \in
\{2,\ldots,n\}} \beta_j d_j$ for some $\beta_j \in \Z$, so take
$\delta := B_1 e_1 - \sum_{j \in \{2,\ldots,n\}} \beta_j e_j \in
{\rm ker}(\tau)$. Consequently $m_1$ divides $B_1$ and by definition
$B_1$ divides $m_1$, so $B_1 d_1 = m_1 d_1 \in \sum_{j \in
\{2,\ldots,n\}} \N d_j$. Now we have that $\mA_{red} = (\mA
\setminus \{d_1\})_{red}$, and by Lemma \ref{facilred} and
Proposition \ref{red} it follows that $I_{\mA \setminus \{d_1\}}$ is
a complete intersection minimally generated by
$\{f_{2},\ldots,f_{n-1}\}$; repeating the same argument we conclude
that $\mA_{red} = \emptyset$. \end{demo} \QED
\bigskip Concerning the case of projective monomial curves,
we denote $d := {\rm max}(\mA)$ and $\omA = \{a_1,\ldots,a_{n+1}\}$
where $a_i = (d_i, d - d_i)$ for every $i \in \{1,\ldots,n\}$ and
$a_{n+1} = (0,d)$. Since toric ideals associated to projective
monomial curves is a subfamily of homogeneous simplicial toric
ideals, the following result, which is a particular case of
\cite[Corollary 3.4]{BGsimplicial}, holds.
\begin{Theorem}\label{coralg} $\oIA$ is a complete intersection $\Longleftrightarrow$ $\omA_{red} = \emptyset$.
\end{Theorem}
Moreover, the computation of $\omA_{red}$ is simpler than in the
general case if we take into account the following properties which
are easy to prove.
\begin{Remark}\label{notaproyred}
\begin{itemize}
\item[{\rm (1)}] $B_i = \gcd (\mA \setminus
\{d_i\})\, /\, \gcd(\mA)$ for all $i \in \{1,\ldots,n\}$.
\item[{\rm (2)}] $\forall i \in \{1,\ldots,n\}: B_i a_i \in \sum_{j \in \{1,\ldots,n+1\} \atop j \neq i} \N a_j
\Longleftrightarrow B_i d_i = \sum_{j \in \{1,\ldots,n\}\atop j \neq
i} \alpha_j d_j \in \sum_{j \in \{1,\ldots,n\} \atop j \neq i} \N
d_j,$ where $\sum_{j \in \{1,\ldots,n\} \atop j \neq i} \alpha_j
\leq B_i.$ Thus, checking whether $B_i a_i \in \sum_{j \in
\{1,\ldots,n\} \atop j \neq i} \N a_j$ is reduced to determining if
an integer belongs to a subsemigroup of $\N$ with an extra
condition.
\item[{\rm (3)}] If $i = n+1$ or $d_i = d = {\rm
max}(\mA)$, then $B_i a_i \notin \sum_{j \in \{1,\ldots,n\} \atop j
\neq i} \N a_j$.
\end{itemize}
\end{Remark}
In Table \ref{algoritmocurvamonomialproyectiva} we show Algorithm
CI-projective-monomial-curve, an algorithm which receives as input a
set $\mA = \{d_1,\ldots,d_n\} \subset \Z^+$ and determines whether
$\oIA$ is a complete intersection. This algorithm is essentially
obtained in \cite[Theorem 3.6]{CN}.
\begin{table}[!htb]
\centering
\begin{tabular}{|p{11.5cm}|}
\hline
\begin{center}
{\bf Algorithm CI-projective-monomial-curve}
\end{center}
\vspace{-.3cm} $$\begin{array}{ll}
\ Entrada: \mA = \{d_1,\ldots,d_n\} \subset \Z^+ \\
\ Salida:\mbox{\ {\sc True} o {\sc False}}
\end{array}$$
\begin{center}
\begin{algorithmic}
\STATE $d := {\rm max}\{d_1,\ldots,d_n\}$
\REPEAT
\STATE $\mB := \mA$
\FORALL {$d_i \in \mA \setminus \{d\}$}
\vspace{.2cm}
\STATE $B_i := \gcd(\mA \setminus
\{d_i\})\, / \, \gcd(\mA)$
\vspace{.2cm}
\IF {$B_i\, d_i = \sum_{d_j \in \mA \, \atop j \neq i} \alpha_j\,
d_j \in \sum _{d_j \in \mA \, \atop j \neq i} \N\, d_j$ and $\sum
\alpha_j \leq B_i$}
\vspace{.2cm}
\STATE $\mA :=
\mA \setminus \{d_i\}$
\ENDIF
\ENDFOR
\UNTIL $(\mA = \{d\})$ OR $(\mA = \mB)$
\IF {$\mA = \{d\}$}
\RETURN {\sc True}
\ENDIF
\RETURN {\sc False}
\end{algorithmic}
\end{center}
\\
\hline
\end{tabular}
\caption{Algorithm {\bf CI-projective-monomial-curve}}
\label{algoritmocurvamonomialproyectiva}
\end{table}
\section{Complete intersections and generalized arithmetic sequences}
\smallskip In this section we deal with the cases in which
either $\mA$ is a generalized arithmetic sequence, or $\mA \setminus
\{d_n\}$ is a generalized arithmetic sequence and $d_n$ is a
positive integer. We denote by $\mS$ the semigroup spanned by $\mA$
and assume that {\bf\mathversion{bold}$\mS$ is a numerical
semigroup}, i.e., $\gcd(\mA) = 1$, and that
{\bf\mathversion{bold}$\mS$ is minimally generated by $\mA$}. Note
that if $\mA$ is a generalized arithmetic sequence and $\gcd(\mA) =
1$, one can easily check that $\mA$ is a minimal set of generators
of $\mS$ if and only if $n \leq a_1$.
\smallskip As we mentioned in the introduction, Maloo and Sengupta
proved in \cite{MS} that whenever $\mA \setminus \{d_n\}$ is an
arithmetic sequence and $n \geq 5$, then $\IA$ is not a complete
intersection. To prove this they used the description of a minimal
set of generators of $\IA$ when $\mA \setminus \{d_n\}$ is an
arithmetic sequence obtained by Patil and Singh \cite{PatilSingh}
(see also \cite{Patil} for a shorter proof of the same result). Here
we present a generalization of Maloo and Sengupta's result which
does not require to obtain a description of a minimal set of
generators of $\IA$. More precisely, we prove that whenever $\mA
\setminus \{d_n\}$ is a generalized arithmetic sequence and $n \geq
5$, then $\IA$ is not a complete intersection. In order to prove
this result, we first introduce two results.
\begin{Lemma}\label{noIC}Let $\mA = \{d_1,\ldots,d_n\}$ be a subset of
$\Z^+$. If there exist $1 \leq i < j \leq n$ such that:
\begin{itemize}
\item $m_i d_i \neq m_j d_j$ and
\item $m_i d_i = \sum_{k \in \{1,\ldots,n\} \atop k \neq i} \alpha_k d_k,\, m_j d_j = \sum_{k \in \{1,\ldots,n\} \atop k \neq j} \beta_k d_k$
with $\alpha_k,\beta_k \in \N,\, \alpha_j \neq 0$ and $\beta_i \neq
0$,
\end{itemize} then $\IA$ is not a complete intersection.
\end{Lemma}
\begin{demo}Assume that $\IA$ is a complete intersection and set
$f_i := x_i^{m_i} - \prod_{k \neq i} x_k^{\alpha_k}$ and $f_j :=
x_j^{m_j} - \prod_{k \neq j} x_k^{\beta_k}$. By Lemma
\ref{lemacritico} there exist some binomials $g_3,\ldots,g_{n-1} \in
\IA$ such that $\IA = (f_i, f_j, g_3,\ldots,g_{n-1})$. Therefore
$\IA \subsetneq J := (x_i,x_j,g_3,\ldots,g_{n-1})$, but this is not
possible because $\IA$ is a prime ideal and $n-1 = {\rm ht}(\IA) <
{\rm ht}(J) \leq n-1$. \QED
\end{demo}
\begin{Proposition}\label{casiaritm}If $n \geq 4$ and $\mA$ contains a generalized arithmetic
sequence with $4$ elements, then $\IA$ is not a complete
intersection.
\end{Proposition}
\begin{demo}Suppose that $\{d_1, d_2, d_3, d_4\} \subset \mA$ is a generalized arithmetic
sequence, i.e., there exists $h \in \Z^+$ such that $\{h d_1, d_2,
d_3, d_4\}$ is an arithmetic sequence. Since $\mA$ is a minimal set
of generators of $\mS$, we have that $m_i > 1$ for all $1 \leq i
\leq n$. Moreover, the equalities $2 d_2 = h d_1 + d_3$ and $2 d_3 =
d_2 + d_4$ prove that $m_2 = m_3 = 2$, hence $\IA$ is not a complete
intersection by Lemma \ref{noIC}. \QED
\end{demo}
\smallskip It is worth pointing out that from \cite[Theorem 2.5]{HKV} one can deduce
a weaker version of Proposition \ref{casiaritm} which states that if
$\mA$ contains an arithmetic sequence with $5$ elements, then $\IA$
is not a complete intersection.
\smallskip From Proposition \ref{casiaritm} one directly derives the following
two corollaries.
\begin{Corollary}\label{casiaritgen5}If $\mA \setminus \{d_n\}$ is a generalized arithmetic sequence and
$n \geq 5$, then $\IA$ is not a complete intersection.
\end{Corollary}
\begin{Corollary}\label{aritgenmenor4}If $\mA$ is a generalized arithmetic sequence and
$n \geq 4$, then $\IA$ is not a complete intersection.
\end{Corollary}
Now we can proceed with the characterizations:
\begin{Theorem} \label{gen}Let $\mA$ be a generalized arithmetic sequence with $n \geq 3$.
Then, $\IA$ is a complete intersection $\Longleftrightarrow \, n =
3$ and $d_1$ is even.
\end{Theorem}
\begin{demo}By Corollary \ref{aritgenmenor4} it only remains to study when $n = 3$.
Let $h \in \Z^+$ be such that $\{h a_1, a_2, a_3\}$ is an arithmetic
sequence. Since $\gcd(\mA)= 1$, denoting $d := d_3 - d_2$ we have
that $d_2 = h d_1 + d$, $d_3 = h d_1 + 2 d$ and $\gcd\{d_1,d\} = 1$.
We separate two cases, if $d_1$ is even, then $B_2 d_2 = 2 d_2 = h
d_1 + d_3 \in \N \{d_1, d_3\}$ and $\mA_{red} = \emptyset$, thus by
Proposition
\ref{n=3}, $\IA$ is a complete intersection. If $d_1$ is odd, then
\begin{itemize}
\item $B_1 d_1 = \gcd\{h, d\}\, d_1 < d_2 < d_3$, thus $B_1 d_1 \not\in \N \{d_2, d_3\},$
\item $B_2 d_2 = d_2 \not\in \N \{d_1, d_3\}$ and
\item $B_3 d_3 = d_3 \not\in \N \{d_1, d_2\}$.
\end{itemize}
So, $\mA_{red} = \{B_1 d_1, d_2, d_3\}$ and again by Proposition
\ref{n=3} we conclude that $\IA$ is not a complete intersection.
\QED
\end{demo}
\medskip
\begin{Remark} Furthermore, whenever $\IA$ is a complete intersection, i.e.,
when $\{d_1,d_2,d_3\}$ is a generalized arithmetic sequence and
$d_1$ is even, we get the following additional information $($see
{\rm Lemma \ref{facilred}} and {\rm Proposition \ref{red})}:
\begin{itemize}
\item $\IA = \left( x_2^2\, -\, x_1^h\, x_3,\,
x_1^{ d_3 / 2}\, -\, x_3^{d_1 / 2} \right)$, with $h \in \Z^+$ such
that $\{h d_1, d_2, d_3\}$ is an arithmetic sequence, i.e., $h = (2
d_2 - d_3) / d_1$.
\item $\g(\mS) = d_1 d_3 / 2 - d_1 +\, d_2\, -\, d_3.$
\end{itemize}
The general formula for the Frobenius number of the semigroup $\N
\mA$ when $\mA$ is a generalized arithmetic sequence can be found in
{\rm \cite[Theorem 3.3.4]{RA}}.
\end{Remark}
\smallskip As a direct consequence of Theorem \ref{gen} we get the already cited results:
\begin{Corollary} \cite[Theorem 3.5]{MS} \label{aritm}
Let $\mA$ be an arithmetic sequence. Then, $\IA$ is a complete
intersection $\Longleftrightarrow$ $n = 2$ or $n = 3$ and $d_1$ is
even.
\end{Corollary}
\begin{Corollary} \cite[Corollary~9]{GR99}
Let $\mA$ be a set of consecutive integers. Then, $\IA$ is a
complete intersection $\Longleftrightarrow$ $n = 2$ or $n = 3$ and
$d_1$ is even.
\end{Corollary}
Concerning the case where $\mA \setminus \{d_n\}$ is a generalized
arithmetic sequence with $n \geq 4$, we have the following:
\begin{Theorem} \label{almost}
Let $\mA \setminus \{d_n\}$ be a generalized arithmetic sequence
with $n \geq 4$. Then, $\IA$ is a complete intersection
$\Longleftrightarrow$ $n = 4$ and one of the following holds:
\begin{enumerate}
\item[{\rm 1.}] $d_1 / \gcd\{d_1,d_2\}$ is even and $\gcd\{d_1,d_2\} d_4 \in \N
\{d_1,d_2,d_3\}$, or
\item[{\rm 2.}] $d_1, d_4$ are even and $\mC_{red} = \emptyset$ with $\mC =
\{d_1,d_3,d_4\}$\end{enumerate}
\end{Theorem}
\begin{demo}By Corollary \ref{casiaritgen5} it only remains to study when $n = 4$.
By Proposition \ref{red} we have that $\IA$ is a complete
intersection if and only if $I_{\mA'}$ so is, where $\mA' :=
\{d_1,d_2,d_3,B_4 d_4\}$. Moreover, $I_{\mA'} = \IB$ where $\mB :=
\{d_1',d_2',d_3',d_4'\}$ with $d_i' = d_i / B_4$ for $1 \leq i \leq
3$ and $d_4' = d_4$.
We separate two cases, if $d_4' \in \N \{d_1',d_2',d_3'\}$, then by
Lemma \ref{facilred}, $\IA$ is a complete intersection if and only
if $I_{\mB \setminus \{d_4'\}}$ so is. Since $\mB \setminus \{d_4'\}
= \{d_1', d_2', d_3'\}$ is a generalized arithmetic sequence and
$\gcd\{d_1',d_2',d_3'\} = 1$, by Theorem \ref{gen} we get that $\IA$
is a complete intersection if and only if $d_1'$ is even. Assume now
that $d_4' \notin \N \{d_1',d_2',d_3'\}$ and set $m_i := {\rm
min}\{b \in \Z^+ \, \vert \, b d_i' \in \N (\mB \setminus
\{d_i'\})\}$ for all $1 \leq i \leq 4$, then $m_i \geq 2$ for every
$1 \leq i \leq 4$, in particular $m_2 = 2$. Let us study the
possible values of $m_1$, if $m_1 d_1' = m_2 d_2'$ we set $d_5' :=
\gcd\{d_1',d_2'\} = 1$ and $m_3' := {\rm min}\{b \in \Z^+ \, \vert
\, b d_3' \in \N \{d_4',d_5'\}\} = 1$ and by \cite[Proposition
2.1]{BGS} we have that $\IB$ is not a complete intersection because
$m_3 \neq m_3'$. If $m_1 d_1' = m_3 d_3'$, then necessarily $d_1'$
is even; otherwise $m_1 d_1' = \lcm\{d_1',d_3'\} = d_1' d_3' > d_1'
d_2'$, which contradicts the definition of $m_1$. Then we set $d_5'
:= \gcd\{d_1',d_3'\} = 2$ and $m_i' := {\rm min}\{b \in \Z^+ \,
\vert \, b d_i' \in \sum_{j \in \{2,4,5\} \atop j \neq i} \N d_j\}$
for $i = 2,4$, and observe that $m_2' = 1$ o $m_4' = 1$. Indeed, if
$d_2'$ is even then $m_2' = 1$, if $d_4'$ is even then $m_4' = 1$
and if both are odd, then $m_4' = 1$ if $d_2' < d_4'$ , or $m_2'= 1$
otherwise. Again by \cite[Proposition 2.1]{BGS} we have that $\IB$
is not a complete intersection. If $m_1 d_1' \notin \{m_2 d_2', m_3
d_3'\}$ then $m_1 d_1',\, m_2 d_2'$ and $m_3 d_3'$ are all different
and by Proposition \ref{n-1dist}, $\IB$ is a complete intersection
if and only if $\mB_{red} = \emptyset$. Set $B_i' := \gcd(\mB
\setminus \{d_i'\})$ for $1 \leq i \leq 4$, then $B_3' = B_4' = 1$
and $B_i' d_i' \notin \N (\mB \setminus \{d_i'\})$ for $i = 3,4$.
Let $d', h' \in \Z^+$ be such that $d_2' = h' d_1' + d'$ and $d_3' =
d_2' + d'$. If both $d_1'$ and $d_4'$ are even, then
$B_2' = 2$ and $2 d_2' = h' d_1'
+ d_3'$, thus $\mB_{red} = \emptyset$ if and only if $\mC_{red} =
\emptyset$, where $\mC = \{d_1,d_3,d_4\}$. In case $d_1'$ or $d_4'$
is odd, we have that $B_2' = 1$ and $B_2' d_2' \notin \N
\{d_1',d_3',d_4'\}$. Concerning $B_1'$, we have that $B_1' =
\gcd\{h',d',d_4'\}$; since $B_1' d_1' \mid h' d_1' < d_2' < d_3'$
then $B_1' d_1' \in \N \{d_2',d_3',d_4'\}$ if and only if $d_4' \mid
B_1' d_1'$. If $d_4' \nmid B_1' d_1'$ then we can conclude that
$\mB_{red} \not= \emptyset$ and $\IA$ is not a complete
intersection. Otherwise, $\mB_{red} = \mB'_{red}$ with $\mB' :=
\{d_2',d_3',d_4'\}$, but $\{d_4', d_2', d_3'\}$ is a generalized
arithmetic sequence and $d_4'$ is odd, then by Proposition \ref{n=3}
and Theorem \ref{gen}, $\mB'_{red} \not= \emptyset$. \QED
\end{demo}
\medskip
\begin{Remark} Furthermore, whenever $\IA$ is a complete intersection, if we let $h$ be the integer such that
$\{h d_1, d_2, d_3\}$ is an arithmetic sequence, i.e., $h = (2d_2 - d_3) / d_1$, we have the following
results $($see {\rm Lemma \ref{facilred}} and {\rm Proposition \ref{red})}:
\begin{enumerate} \item If $d_1 / \gcd\{d_1,d_2\}$ is even,
$\gcd\{d_1,d_2\} d_4 \in \N \{d_1,d_2,d_3\}$ and we take $\beta_1,
\beta_2, \beta_3 \in \N$ such that $b d_4 = \beta_1 d_1 + \beta_2
d_2 + \beta_3 d_3$, where $b := \gcd\{d_1,d_2\}$, then
\begin{itemize} \item $\IA = \left( x_4^{b} -
x_1^{\beta_1} x_2^{\beta_2} x_3^{\beta_3},\,
x_2^2 - x_1^h x_3,\, x_1^{d_3 / 2 b}
- x_3^{d_1 / 2 b} \right)$; and
\item $\g(\mathcal S) = (d_1 d_3\, /\, 2 b) - d_1 + d_2 - d_3 + (b - 1) d_4.$
\end{itemize}
\vskip.2cm \item If both $d_1$ and $d_4$ are even, $\mC_{red} =
\emptyset$ with $\mC := \{d_1,d_3,d_4\}$ and we denote by $\mathcal
S^{\prime}$ the numerical semigroup $\N (d_1 / 2) + \N (d_3 / 2) +
\N (d_4 / 2)$, then
\begin{itemize} \item $\IA = \left( x_2^2 - x_1^h x_3 \right) + I_{\mathcal C} \cdot
k[x_1,x_2,x_3,x_4]$; and
\item $\g(\mathcal S) = 2\, \g(\mathcal S^{\prime}) + d_2$.
\end{itemize}
\end{enumerate}
\end{Remark}
\section{Complete intersections in Fibonacci sequences.}
Given $p,q \in \Z^+$ with $\gcd\{p,q\} = 1$, our next aim is to
characterize when $\IA$ is a complete intersection where $\mA =
\{d_1,\ldots,d_n\} \subset \Z^+$ with $d_i = F_{e_i}$ for $1 \leq i
\leq n$, $\{F_n\}$ denotes the $(p,q)$-Fibonacci sequence and
$\{e_1,\ldots,e_n\}$ is a generalized arithmetic sequence. In order
to achieve the characterization we first introduce some basic
properties of linear second order recurrence sequences, focusing
specially on those properties of the $(p,q)$-Fibonacci sequence and
also in the Lucas one. Some of these properties are straight
generalizations of those in \cite{Vaj}, some others can be found in
\cite[Section 5]{HP} and the rest can be easily proved.
\smallskip We denote by $[a]_2$ the $2$-valuation of $a \in \Z^+$, i.e.,
$[a]_2 := {\rm max}\{t \in \N \, : \, 2^t|a\}$, and have the
following result:
\begin{Lemma} \cite[Theorems $\bar{f}$, $\bar{\ell}$ y $\bar{f} \bar{\ell}$ ]{HP} \label{gcdfl}
Let $a,b$ be positive integers and set $d := \gcd\{a,b\}$. Then,
\begin{itemize} \item $\gcd\{F_a,F_b\} = F_d.$
\item $\gcd\{L_a,L_b\} = \left\{ \begin{array}{cl}
L_d &$ if $ [a]_2 = [b]_2 \\
2 &$ if $ [a]_2 \neq [b]_2$, $p,\,q$ are odd and $3|d \\ 2 &$ if
$[a]_2 \neq [b]_2$, $p$ is even and $q$ is odd $\\
1&$ otherwise$
\end{array} \right.$
\item $\gcd\{L_a,F_b\} = \left\{ \begin{array}{cl}
L_d &$ if $[a]_2 < [b]_2 \\
2 &$ if $[a]_2 \geq [b]_2$, $p,\,q$ are odd and $3|d \\ 2 &$ if
$[a]_2 \geq [b]_2$, $p,\, b$
are even and $q$ is odd $\\
1 &$ otherwise $
\end{array} \right.$ \end{itemize}
\end{Lemma}
Observing Lemma \ref{gcdfl} and that every $(p,q)$-Fibonacci
sequence is strictly increasing except for $p = 1$ (because $F_1 =
F_2$), and that every $(p,q)$-Lucas sequence is strictly increasing
except for $p \in \{1,2\}$ (because $L_0 \geq L_1$), we get the
following divisibility properties.
\begin{Corollary}\label{divisible}Let $a,\, b$ be two positive integers, we have the following
properties: \begin{itemize}
\item[{\rm (1)}] If $a \mid b$ then $F_a
\mid F_b$.
\item[{\rm (2)}] If $F_a \mid F_b$ then $a \mid b$, unless if $p = b = 1$ and $a = 2$.
\item[{\rm (3)}] If $a \geq 2$, then $L_a \mid L_b$ if and only if $b / a$ is odd.
\item[{\rm (4)}] If $b$ is even, then $\gcd\{L_a,
L_{a+b}\} = \gcd\{L_a,F_b\}$.
\end{itemize}
\end{Corollary}
\medskip
Denote by $\{U_n\}_{n \in \N}$ any sequence satisfying that $U_{n+2}
= p\, U_{n+1} + q\, U_n$ for all $n \geq 2$ and $U_0,\, U_1 \in \N$
are not both null. In the following result we provide some
properties of theses sequences that we will use in the sequel, all
of them are easy to prove. Property $(4)$ can be found in
\cite[Propositions 5.1 and 5.2]{HP}, properties $(1)$ and $(3)$ are
straight generalizations of the corresponding results for $p = q =
1$ one can find in \cite{Vaj} and $(2)$ can be easily proved.
\begin{Lemma}\label{propfib}Let $a,b,c,d,e$ be positive integers.
Then,
\begin{itemize}
\item[{\rm (1)}] $U_{a+b} = F_a\,U_{b+1} + q\, F_{a-1}\,U_b$.
\item[{\rm (2)}] If $b \in \N \{a_1,\ldots,a_k\}$ where
$a_1,\ldots,a_k \in \Z^+$, then $F_b \in \N
\{F_{a_1},\ldots,F_{a_k}\}$.
\item[{\rm (3)}] $F_a U_b - F_c U_d = (-1)^e q^e (F_{a-e}
U_{b-e} - F_{c-e} U_{d-e})$ if $a + b = c + d$ and $e \leq {\rm
min}\{a,b,c,d\}$.
\item[{\rm (4)}] $L_a = F_{2a} / F_a = F_{a+1} + q F_{a-1}$.
\item[{\rm (5)}] $U_{a+2b} + (-1)^b q^{b} U_a = L_b U_{a+b}$.
\end{itemize}
\end{Lemma}
With these basic properties, the following inequalities are easy to
prove.
\begin{Corollary}\label{propfib2}Let $a,b,c,d$ be positive integers. The following
inequalities hold:
\begin{itemize}
\item[{\rm (1)}] $q^b U_a \leq U_{a+2b}$ and equality holds if and only if $a = U_1 = 0$, $b = 1$.
\item[{\rm (2)}] $L_a < F_{a+2}$.
\item[{\rm (3)}] $U_{a+b-2} < F_a U_b < U_{a+b-1} \ $ if $\ a, b \geq 2.$
\item[{\rm (4)}] $F_a U_b < F_c U_d \ $ if $\ a + b < c + d$.
\item[{\rm (5)}] If $a < c$, $a < d$ and $a + b = c + d$, then $F_a U_b
< F_c U_d\ $ if and only if $a$ is even.
\item[{\rm (6)}] $L_{a+b-1} < L_a L_b < {\rm min}\{L_{a+b+1}, 2\,L_{a+b}\}$.
\item[{\rm (7)}] If $a \leq b$, then $L_a L_b < L_{a+b}$ if $a$ is odd and
$L_a L_b > L_{a+b}$ if $a$ is even.
\end{itemize}
\end{Corollary}
Let $d_i$ denote the $e_i$-th term of the $(p,q)$-Fibonacci
sequence, where $\{e_1,\ldots,e_n\}$ is a generalized arithmetic
sequence. This is, there exist $h, a, d \in \Z^+$ such that $d_1 :=
F_{a}$ and $d_i := F_{ha + (i-1)d}$ for all $i \geq 2$. As we have
mentioned, we aim at characterizing when $\IA$ is a complete
intersection in terms of the values of $p,q,n,h,a,d$. This objective
is achieved with the following result.
\begin{Theorem} \label{fibonacci}Let $p, q \in \Z^+$ be two relatively prime integers and let
$\{F_n\}_{n \in \N}$ be the $(p,q)$-Fibonacci sequence. Let $\mA =
\{d_1,\ldots,d_n\}$ be the set with $d_1 := F_a$ and $d_i := F_{ha +
(i-1)d}$ for all $i \in \{2,\ldots,n\}$ where $h, a, d \in \Z^+$ and
$n \geq 3$. Then, $\IA$ is a complete intersection
$\Longleftrightarrow$ one the following holds:
\begin{itemize}
\item[{\rm (a)}] $d$ is odd,
\item[{\rm (b)}] $d \geq a$,
\item[{\rm (c)}] $a = 2d$,
\item[{\rm (d)}] $\gcd\{a,d\} = a-d$ and $a$ is odd, or
\item[{\rm (e)}] $n = 3$ and $2d \mid a$.
\end{itemize}
\end{Theorem}
To prove this theorem we use two previous results, namely Lemma
\ref{pertenece} and Proposition \ref{fib}. The first one
characterizes when $d_3 \in \N \{d_1,d_2\}$ and also proves that
whenever $d_3 \in \N \{d_1,d_2\}$, then $d_i \in \N \{d_1,d_2\}$ for
all $i \geq 3$, which particularly implies that $\IA$ is a complete
intersection. The second one characterizes when $\IA$ is a complete
intersection whenever $d_3 \notin \N \{d_1,d_2\}$.
\begin{Lemma} \label{pertenece} $d_3 \in \N \{d_1, d_2\}$
$\Longleftrightarrow$ $d$ is odd or $F_{2d} \geq \lcm\{d_1,\,F_d\}$.
Moreover, if $d_3 \in \N \{d_1, d_2\}$, then $d_i \in \N
\{d_1,d_2\}$ for all $i \geq 3$.
\end{Lemma}
\begin{demo}If $d$ is odd, by $(5)$ in Lemma \ref{propfib} we get that $d_3 = (q^{d}
F_{ha} / F_a) d_1 + L_d d_2 \in \N \{d_1,d_2\}$ and that $d_{i+2} =
L_d d_{i+1} + q^d d_i$ for all $i \geq 2$, thus $d_i \in \N
\{d_1,d_2\}$ for all $i \geq 3$. Suppose that $d$ is even, set $e :=
\gcd\{d_1,d_2\} = F_{\gcd\{a,d\}}$ and consider the numerical
semigroup $\mathcal S := \N \left\{d_1 / e, d_2 / e \right\}$, its
Frobenius number is $\g(\mathcal S) = ((d_1 d_2) / e - d_1 - d_2) /
e$ (see, e.g. \cite[Theorem 2.1.1]{RA}). If $d \geq a$, then $e
\g(\mathcal S) < (d_1 d_2) / e \leq d_1 d_2 \leq F_d\, d_2 < d_i$
for all $i \geq 3$ and we conclude that
$d_i \in \N \{d_1,d_2\}$. Note that in this case $F_{2d}
> d_1 F_d \geq \lcm \{d_1,F_d\}$. Suppose now that $a >
d$ and let us study the existence of solutions $(x,y) \in \N^2$ to
the linear diophantine equation
\begin{equation}\label{eq} d_1 x + d_2 y = d_3.
\end{equation} By Lemma \ref{propfib} we know that $\left(-q^{d}\, \frac{F_{ha}}{d_1} ,\,
L_d \right) \in \Z^2$ is an integer solution to the equation; thus
the set of integral solutions is
$$\left\{ \left( -q^d\, \frac{F_{ha}}{d_1} + \lambda
\frac{d_2}{e},\, L_d - \lambda \frac{d_1}{e} \right) ,\, \lambda \in
\Z \right\}.$$ We claim that $-q^d \frac{F_{ha}}{d_1}+ \frac{d_2}{e}
> 0$; indeed since $a > d$, then $e = F_{\gcd\{a,d\}} \leq F_{a-d}$
and $q^d\, e\, F_{ha} = q^{d / 2}\, e\, q^{d / 2}\, F_{ha} < q^{d
/ 2}\, F_{a-d}\, d_2 < d_1\, d_2$. Therefore (\ref{eq}) has a
nonnegative integer solution if and only if $d_1 / e \leq L_d =
F_{2d} / F_d$, which is equivalent to $F_{2d} \geq \lcm\{d_1,
F_d\}$. Finally we have that
$$\g(\mathcal T)\,e \leq \frac{d_1\, d_2}{e} \leq L_d \, d_2 <
F_{d+2}\,d_2 < F_{ha+2d+1} < d_i {\rm \ for\ all\ } i \geq 4$$ and
$d_i \in \N \{d_1,d_2\}$ for all $i \geq 4$. \QED
\end{demo}
\vskip.5cm Now we study when $\IA$ is a complete intersection
provided $d_3 \notin \N \{d_1,d_2\}$.
\begin{Lemma} \label{fibn=4}
If $2d \mid a$ and $d_3 \not\in \N \{d_1,d_2\}$, then $d_4 \not\in
\N \{d_1,d_2,d_3\}$.
\end{Lemma}
\begin{demo}Assume that there exist $\alpha_1, \alpha_2, \alpha_3
\in \N$ such that $d_4 = \alpha_1 d_1 + \alpha_2 d_2 + \alpha_3
d_3$. Since $d_4 = -q^d d_2 + L_d d_3$, we have that $\beta d_3 =
\alpha_1 d_1 + (\alpha_2 + q^d) d_2$ where $\beta := L_d -
\alpha_3$ and the equation $\beta d_3 = x d_1 + y d_2$ has a
solution $(x,y) \in \N \times \N$. Moreover, the equality $\beta d_3
= - \beta q^d (F_{ha} / d_1) d_1 + \beta L_d d_2$ yields that the
set of integer solutions to this equation is
$$\left\{ \left(-\beta\, q^d \frac{F_{ha}}{d_1} + \lambda \,
\frac{d_2}{F_d},\, \beta\, L_d - \lambda \, \frac{d_1}{F_d} \right),
\, \lambda \in \Z \right\}.$$ Nevertheless, for all $\lambda
> 0$ we have that
$$\beta\, L_d - \lambda \, \frac{d_1}{F_d} \leq (L_d)^2 -
\frac{d_1}{F_d} \leq (L_d)^2 - \frac{F_{4d}}{F_d} = (L_d)^2 - L_{2d}
L_d < 0;$$ and we can conclude that there is no solution $(x,y) \in
\N \times \N$, a contradiction. \QED
\end{demo}
\begin{Proposition}\label{fib}If $d_3 \not\in \N \{d_1,d_2\}$, then
$\IA$ is a complete intersection $\Longleftrightarrow$ $n = 3$ and
$2d\,|\,a$.
\end{Proposition}
\begin{demo}By Lemma \ref{pertenece}, $d$ is even
and by Lemma \ref{propfib} we have that $d_3 + q^d F_{ha} = L_d d_2$
and $d_{i+1} + q^d d_{i-1} = L_d d_i$ for $3 \leq i \leq n-1$, which
implies that $m_i \leq L_d$ for all $i \in \{2,\ldots,n-1\}$. We
claim that $m_2 d_2,\ldots,m_n d_n$ are all different. Indeed if we
assume that there exist $i,j: 2 \leq i < j \leq n$ such that $m_i
d_i = m_j d_j$, then $m_i d_i \leq L_d d_i = d_{i+1} + q^d d_{i-1} <
2\, d_{i+1}$; which implies that $j = i+1$ and $m_i d_i = d_{i+1}$
and $d_i \mid d_{i+1}$, but this is not possible because $ha +
(i-1)d \nmid ha + id$. Hence by Proposition \ref{n-1dist}, $\IA$ is
a complete intersection if and only if $\mA_{red} = \emptyset$.
Let $\mA'$ be the minimal set of generators of $\N \mA$, then there
exists $k \in \N$ and $4 \leq i_1 < \cdots < i_k \leq n$ such that
$\mA' = \{d_1,d_2,d_3,d_{i_1},\ldots,d_{i_k}\}$. We set
$$B_j' := \frac{\gcd(\mA' \setminus \{d_j\})}{ \gcd(\mA')} {\rm \
for \ all \ } j \in \{1,2,3,i_1,\ldots,i_k\}$$ and have that $B_j' =
1$ for all $j \in \{3,i_1,\ldots,i_k\}$ because $\gcd\{d_1,d_2\} =
\gcd(\mA')$ and $B_1' = \frac{F_{\gcd\{ha,d\}}}{\gcd(\mA')}$.
Moreover $B_1' d_1 \not \in \N (\mA' \setminus \{d_1\})$ because
$B_1' d_1 \leq F_d d_1 < d_i$ for all $i \geq 2$. If $[a]_2 \leq
[d]_2$ or there exists $j \in \{1,\ldots,k\}$ such that $i_j$ is
even, we also have that $B_2' = 1$, this implies that $\mA_{red}
\not= \emptyset$ and $\IA$ is not a complete intersection. If $[a]_2
> [d]_2$ and $i_j$ is odd for all $j \in \{1,\ldots,k\}$, then
$\gcd\{a,2d\} = 2 \gcd\{a,d\}$ and $B_2 = L_{\gcd\{a,d\}}$. Suppose
first that $\gcd\{a,d\} < d$, then $B_2 d_2 \leq L_{d-1} d_2 < L_{ha
+ 2d - 2} < d_i$ for all $i \in \{3,\ldots,n\}$ and we claim that
$d_1 \nmid B_2\, d_2$; otherwise we take $\alpha_1,\alpha_2 \in \Z$
such that $d_3 = \alpha_1 d_1 + \alpha_2 d_2$, then $B_2 \mid
\alpha_2$ and we get that $d_1 \mid d_3$, a contradiction. Hence
$\mA_{red} \not= \emptyset$ and $\IA$ is not a complete
intersection. Finally assume that $[a]_2
> [d]_2$ and that $\gcd\{a,d\} = d$ or, equivalently, that $2 d \mid a$;
if $n \geq 4$, by Lemma \ref{fibn=4} we have that $d_4 \not\in \N
\{d_1,d_2,d_3\}$, then $i_1 = 4$ and we are in the previous case; if
$n = 3$, then $L_d\, d_2 = q^d\, F_{ha} + d_3 \in \N d_1 + \N d_3$,
$\mA_{red} = \emptyset$ and by Proposition \ref{n=3} we conclude
that $\IA$ is a complete intersection. \QED
\end{demo}
\bigskip
\noindent {\it Proof of Theorem \ref{fibonacci}.} As we proved in
Lemma \ref{pertenece}, $d_3 \in \N \{d_1,d_2\}$ if and only if $d$
is odd or $F_{2d} \geq \lcm\{d_1, F_d\}$. Moreover, $F_{2d} \geq
\lcm\{d_1, F_d\}$ $\Longleftrightarrow$ $F_{2d} F_{\gcd\{a,d\}} \geq
d_1 F_d$, and by (4) and (5) in Corollary \ref{propfib2} this is
equivalent to $\gcd\{a,d\} > a - d$ or $a = 2d$ or $\gcd\{a,d\} = a
- d$ and $a - d$ is odd. Furthermore, $\gcd\{a,d\}
> a - d$ if and only if $d \geq a$. So, we have that $d_3 \in
\N \{d_1,d_2\}$ if and only if $d$ is odd, $d \geq a$, $a = 2d$ or
$\gcd\{a,d\} = a - d$ and $a$ is odd. In this situation we also have
that $d_i \in \N \{d_1,d_2\}$ for all $i \in \{3,\ldots,n\}$ and by
Lemma \ref{facilred} we conclude that $\IA$ is a complete
intersection. If $d_3 \not\in \N \{d_1,d_2\}$, the result follows
from Proposition \ref{fib}.\QED
\bigskip It is worth to mention that the characterization obtained in Theorem \ref{fibonacci}
does not depend on the values of $p,\,q$ or $h$, but only on those
of $a,\,d$ y $n$.
\medskip
\begin{Remark} Denoting $e := F_{\gcd\{a,d\}}$ and $\mathcal S :=
\sum_{i = 1}^n \N \, (d_i / e)$, whenever $\IA$ is a complete
intersection, we get the following additional information $($see
{\rm Lemma \ref{facilred}} and {\rm Proposition \ref{red})}:
\begin{enumerate} \item If $d$ is odd. Then,
\begin{itemize} \item $\IA = \left( x_1^{d_2 / e} - x_2^{d_1 / e},
\, x_3 - x_1^{q^d F_{ha} / d_1} x_2^{L_d}, x_4 - x_2^{q^d}
x_{3}^{L_d},\ldots,\, x_n - x_{n-2}^{q^d} x_{n-1}^{L_d} \right)$; y
\item $g\left( \mathcal S \right) =
\frac{1}{e} \, \left( d_1 d_2 / e - d_1 - d_2 \right)$.
\end{itemize}
\vskip.3cm
\item If $d \geq a$, or $a = 2d$, or $\gcd\{a,d\} = a - d$ and $a$ is odd. Then,
\begin{itemize} \item $\IA= \left( x_1^{d_2 / e} - x_2^{d_1 / e},
x_3 - x_1^{b_{3,1}} x_2^{b_{3,2}},\, x_4 - x_1^{b_{4,1}}
x_2^{b_{4,2}}, \ldots,x_n - x_1^{b_{n,1}} x_{2}^{b_{n,2}}\right)$,
\\ where
$b_{3,1},\ldots,b_{n,2} \in \Z^+$ satisfy that $b_{i,1}\, d_1 +
b_{i,2}\, d_2 = d_i$ for all $i \in \{3,\ldots,n\}$; and
\item $g\left( \mathcal S \right) =
\frac{1}{e} \, \left( d_1 d_2 / e - d_1 - d_2 \right)$.
\end{itemize}
\vskip.3cm
\item If $n=3$\,, $2d \mid a$\,, $a \neq 2d$
and $d$ is even. Then $e = F_d$ and
\begin{itemize}
\item $\IA = \left( x_1^{d_3 / F_{2d}} - x_3^{d_1 / F_{2d}},
\, x_2^{L_d} - x_1^{q^d F_{ha} / d_1} x_3 \right)$; and
\item $g \left( \mathcal S \right) = \frac{1}{F_d} \left( d_1\, d_3 / F_{2d}
- d_1 + \left( L_d - 1 \right) d_2 - d_3 \right)$.
\end{itemize}
\end{enumerate}
\end{Remark}
\section{Complete intersections in Lucas sequences.}
Let $p,q \in \Z^+$ be two relatively prime intgers, in this section
$d_i$ denotes the $e_i$-th term of the $(p,q)$-Lucas sequence, where
$\{e_i\}_{1 \leq i \leq n}$ is an arithmetic sequence. This is,
there exist $a, d \in \Z^+$ such that $d_i := L_{a + (i-1)d}$ for
all $i \geq 1$ and we aim at characterizing when $\IA$ is a complete
intersection in terms of the values of $p, q, n, a$ and $d$. This
objective is achieved in Theorem \ref{lucas}. Recall that $[a]_2$
(respect. $[d]_2$) denotes the $2$-valuation of $a$ (respect. $d$).
\begin{Theorem} \label{lucas}Let $p,\,q$ be two relatively prime positive integers and let
$\{L_n\}_{n \in \N}$ be the $(p,q)$-Lucas sequence. Let $\mA =
\{d_1,\ldots,d_n\}$ be the set with $d_i := L_{a + (i-1)d}$ for all
$i \in \{1,\ldots,n\}$ where $a,\,d\in \Z^+$ and $n \geq 3$. Then,
$\IA$ is a complete intersection $\Longleftrightarrow$ one of the
following holds:
\begin{itemize}
\item[{\rm (a)}] $d$ is odd,
\item[{\rm (b)}] $d \geq a$,
\item[{\rm (c)}] $\gcd\{a,d\} = a-d$ and $[d]_2 > [a]_2 \geq 1$,
\item[{\rm (d)}] $n = 3$, $p$ and $a/d$ are odd and $q$ is even,
o
\item[{\rm (e)}] $n = 3$, $3 \nmid d$ and $p,\,q$ and $a / d$ are odd.
\end{itemize}
\end{Theorem}
To prove this theorem we have followed a completely analogous scheme
to the one we used for Theorem \ref{fibonacci} but taking into
account the $(p,q)$-Lucas sequence properties shown in Lemmas
\ref{gcdfl} and \ref{propfib} and Corollaries \ref{divisible} and
\ref{propfib2}. For a sake of brevity we are not including here the
proof of Theorem \ref{lucas}, nevertheless we will state Lemma
\ref{pertenecelucas} and Proposition \ref{luc}, that are,
respectively, the Lucas versions of Lemma \ref{pertenece} and
Proposition \ref{fib} which we have used to prove Theorem
\ref{lucas}. Lemma \ref{pertenecelucas} provides a characterization
of when $d_3 \in \N \{d_1,d_2\}$ and states that whenever $d_3 \in
\N \{d_1,d_2\}$, then $d_i \in \N \{d_1,d_2\}$ for all $i \geq 3$,
which, in particular, implies that $\IA$ is a complete intersection.
Proposition \ref{luc} characterizes when $\IA$ is a complete
intersection whenever $d_3 \notin \N \{d_1,d_2\}$.
\begin{Lemma}\label{pertenecelucas} $d_3 \in \N \{d_1,
d_2\}$ if and only if $d$ is odd or $F_{2d} \geq \lcm\{d_1,F_d\}$.
Moreover, whenever $d_3 \in \N \{d_1,d_2\}$, then $d_i \in \N \{d_1,
d_2\}$ for all $i \geq 3$.
\end{Lemma}
\vskip.5cm Now we study when $\IA$ is not a complete intersection
provided $d_3 \notin \N \{d_1,d_2\}$.
\begin{Lemma} \label{lucasn=4}
If $a / d$ is odd, $\gcd\{d_1, d_2\} = 1$ and $d_3 \not\in \N \{d_1,
d_2\}$, then $d_4 \not\in \N \{d_1, d_2, d_3\}$.
\end{Lemma}
\begin{Proposition}\label{luc}If $d_3 \not\in \N \{d_1,d_2\}$,
then $\IA$ is a complete intersection $\Longleftrightarrow$ $n =
3$, $a/d$ is odd and $\gcd\{d_1,d_2\} = 1$.
\end{Proposition}
\medskip
\begin{Remark} Denoting $e := \gcd(\mA)$ and $\mathcal S :=
\sum_{i = 1}^n \N \, (d_i / e)$, whenever $\IA$ is a complete
intersection, we get the following additional information $($see
{\rm Lemma \ref{facilred}} and {\rm Proposition \ref{red})}:
\begin{enumerate}
\item If $d$ is odd. Then,
\begin{itemize}
\item $\IA = \left( x_1^{d_2 / e} - x_2^{d_1 / e},
\, x_3 - x_1^{q^d} x_2^{L_d}, \ldots,\, x_n - x_{n-2}^{q^d}\,
x_{n-1}^{L_d} \right)$; and
\item $g\left( \mathcal S\right) =
\frac{1}{e} \, \left( d_1 d_2 / e - d_1 - d_2 \right)$.
\end{itemize}
\vskip.3cm
\item If $d \geq a$, or $\gcd\{a,d\} = a-d$ and $[d]_2 > [a]_2 \geq 1$. Then,
\begin{itemize}
\item $\IA = \left(x_1^{d_2 / e} - x_2^{d_1 / e},
x_3 - x_1^{b_{3,1}} x_2^{b_{3,2}},\, x_4 - x_1^{b_{4,1}}
x_2^{b_{4,2}}, \ldots,x_n - x_1^{b_{n,1}} x_{2}^{b_{n,2}}\right)$,
with $b_{i,1}, b_{i,2} \in \N$ such that $b_{i,1} d_1 + b_{i,2} d_2
= d_i$ for all $i \in \{3,\ldots,n\}$; and
\item
$g\left( \mathcal S \right) = \frac{1}{e} \, \left( d_1 d_2 / e -
d_1 - d_2 \right)$.
\end{itemize}
\vskip.3cm
\item If $n=3$, $d$ is even, $a / d$ and $p$ are odd and, either $q$ is even, or $3
\nmid d$. Then $e = 1$ and
\begin{itemize}
\item $\IA = \left( x_1^{d_3 / L_d} - x_3^{d_1 / L_d},
\, x_2^{L_d} - x_1^{q^d}\, x_3 \right)$; and
\item $\g(\mathcal S) = d_1 d_3 / L_d - d_1 + (L_d -1)d_2 - d_3$.
\end{itemize}
\end{enumerate}
\end{Remark}
\medskip
Theorem \ref{lucas} depends on the values of $a, d, n$ and on the
parity of $p$ and $q$, in contrast to the corresponding result for
the Fibonacci sequence where the values of $p,q$ and even of $h$ do
not play any role in the result (see Theorem \ref{fibonacci}). This
dependence of "all" initial values gives the insight that a more
general result where the complete intersection property for $\IA$ is
characterized, where $d_i = L_{e_i}$ with $\{e_1,\ldots,e_n\}$ a
generalized arithmetic sequence, the value of $h$ such that $\{h
e_1,e_2,\ldots,e_n\}$ is an arithmetic sequence is relevant. The
following example shows the relevance of the value of $h \in \Z^+$,
and gives a taste of the difficulty that might have the more general
case in which $e_1,\ldots,e_n$ is a generalized arithmetic sequence.
\begin{example} Let $\mA_h$ be the set $\{L_{5}, L_{h 5 + d}, L_{h 5 + 2d}\},$
where $\{L_n\}_{n \in \N}$ is the $(1,1)$-Lucas sequence and $h \in
\Z^+$. For $h = 1$ and $h = 3$, the sets $\mA_1 = \{L_5 = 11, L_6 =
18, L_7 = 29\}$ and $\mA_3 = \{L_5 = 11, L_{16} = 2207, L_{17} =
3571\}$ determine complete intersection toric ideals. Nevertheless,
for $h = 2$ and $h = 4$, the sets $\mA_2 = \{L_5 = 11, L_{11} = 199,
L_{12} = 322\}$ and $\mA_4 = \{L_5 = 11 , L_{21} = 24476, L_{22} =
39603\}$ determine two non complete intersection toric ideals.
\end{example}
\section{Complete intersections in certain projective monomial curves}
In this section we denote $\mA = \{d_1,\ldots,d_n\} \subset \Z^+$
and $d := {\rm max} \{d_1,\ldots,d_n\}$ and consider $$\omA :=
\{a_1, \ldots, a_{n-1}, a_n, a_{n+1}\} \subset \N^2,$$ where $a_i :=
(d_i,d - d_i)$ for every $i \in \{1,\ldots,n\}$ and $a_{n+1} :=
(0,d)$.
\smallskip The objective of this section is to characterize when $\oIA$ is a complete intersection
in terms of the set $\mA$, where $\mA$ belongs to any of the
families studied in the previous sections. For this purpose we use
Algorithm CI-projective-monomial-curve of Table
\ref{algoritmocurvamonomialproyectiva}.
\medskip We begin studying when $\oIA$ is a complete intersection, where either $\mA$
or $\mA \setminus \{d_n\}$ is a generalized arithmetic sequence and
we assume without loss of generality that
{\mathversion{bold}$\gcd(\mA) = 1$}.
\begin{Theorem}\label{genproy} Let $\mA = \{d_1,\ldots,d_n\}$ be a
generalized arithmetic sequence with $n \geq 3$. Then, $\oIA$ is a
complete intersection $\Longleftrightarrow$ $n = 3$, $\mA$ is an
arithmetic sequence and $d_1$ is even.
\end{Theorem}
\begin{demo}$(\Rightarrow)$ Set $h \in \Z^+$ such that $\{h d_1, d_2, \ldots, d_n\}$ is
an arithmetic sequence. For every $i \in \{3,\ldots,n-1\}$ we have
that $B_i = 1$ and $B_1 = \gcd\{d_1,h\}$, thus $B_1 d_1 \leq h d_1 <
d_2 < \cdots < d_n$ and $B_1 a_1 \notin \sum_{j=2}^{n+1} \N a_j$.
If $n \geq 4$ or $d_1$ is odd, it follows that $B_2 = 1$ and
$\omA_{red} \not= \emptyset$, thus $\oIA$ is not a complete
intersection by Theorem \ref{coralg}. Suppose now that $n = 3$ and
$d_1$ even, then $B_2 = 2$. If $h \geq 2$, then $B_2 d_2 \not=
\alpha_1 d_1 + \alpha_3 d_3$ with $\alpha_1,\alpha_3 \in \N$ and
$\alpha_1 + \alpha_3 \leq 2$ and, again by Theorem \ref{coralg}, we
have that $\oIA$ is not a complete intersection.
$(\Leftarrow)$ We have that $B_2 = 2$ and $2 a_2 = a_1 + a_3$, then
by Theorem \ref{coralg}, $\oIA$ is a complete intersection. \QED
\end{demo}
\medskip
\begin{Remark} Whenever $\oIA$ is a complete intersection, i.e., when $n = 3$, $\mA$
is an arithmetic sequence and $d_1$ is even, we get the following
minimal set of generators of the toric ideal {\rm (}see {\rm Lemma
\ref{facilred}} and {\rm Proposition \ref{red})}:
$$\oIA = \left( x_2^2\, -\, x_1 x_3,\, x_1^{ d_3 / 2}\, -\,
x_3^{d_1 / 2} x_4^{d_2 - d_1} \right).$$
\end{Remark}
\bigskip
Concerning when $\mA \setminus \{d_n\}$ is a generalized arithmetic
sequence, we can assume without loss of generality that
$\{d_n,d_2,\ldots,d_{n-1}\}$ is not an arithmetic sequence.
Otherwise we take $d_1' := d_n$, $d_i' = d_i$ for all $2 \leq i \leq
n-1$ and $d_n' = d_1$ and have that $\mA \setminus \{d_n'\} =
\{d_1',\ldots,d_{n-1}'\}$ is an arithmetic sequence.
\begin{Theorem} \label{almostN=4proy}
Let $\mA = \{d_1,\ldots,d_n\}$ be a set such that $\mA \setminus
\{d_n\}$ is a generalized arithmetic sequence with $n \geq 4$. Then,
$\oIA$ is a complete intersection $\Longleftrightarrow$ $n = 4$,
$\{d_1,d_2,d_3\}$ is an arithmetic sequence and one of the following
holds:
\begin{enumerate}
\item[{\rm 1.}] $d_1\, /\, \gcd\{d_1,d_2\}$ is even
and $\gcd\{d_1,d_2\}\, d_4 = \beta_1 d_1 + \beta_2 d_2 + \beta_3
d_3$ with $\beta_1 + \beta_2 + \beta_3 \leq \gcd\{d_1,d_2\}$, or
\item[{\rm 2.}] $d_1, d_4$ are even and
${\mathcal C}^{\star}_{red} = \emptyset$ with $\mathcal C =
\{d_1,d_3,d_4\}$.
\end{enumerate}
\end{Theorem}
\begin{demo}Let $h \in \Z^+$ be such that $\{hd_1,d_2,\ldots,d_{n-1}\}$ is an arithmetic sequence.
We divide the proof in two parts, if $B_n a_n \in \sum_{j \in
\{1,\ldots,n-1,n+1\}} \N a_j$ or equivalently if $B_n d_n = \sum_{j
= 1}^{n-1} \beta_j d_j$ with $\sum_{j = 1}^{n-1} \beta_j \leq B_n$
(see Remark \ref{notaproyred}), then by Proposition \ref{red} and
Lemma \ref{facilred} it follows that $\oIA$ is a complete
intersection if and only if lo is $I_{\omA \setminus \{a_n\}}$. It
is easy to check that $I_{\omA \setminus \{a_n\}} = I_{\omA_1}$ with
$\mA_1 = \{d_1/B_n, \ldots, d_{n-1} / B_n\}$. Moreover $\gcd(\mA_1)
= 1$ and $\mA_1$ is a generalized arithmetic sequence, then by
Theorem \ref{genproy} we conclude that $I_{\omA_1}$ is a complete
intersection if and only if $n = 4$, $d_1 / B_4$ is even and $\{d_1,
d_2, d_3\}$ is an arithmetic sequence; note also that $B_4 =
\gcd\{d_1,d_2\}$.
Suppose now that $B_n a_n \notin \sum_{j \in \{1,\ldots,n-1,n+1\}}
\N a_j$. We have that $B_i = 1$ for $i \in \{3,\ldots,n-1\}$ and
then $B_i a_i \notin \sum_{j \in \{1,\ldots,n+1\} \atop j \neq i} \N
a_j$. Let us study the values of $B_1$ and $B_2$. Set $h,r \in
\Z^+$ such that $d_i = h d_1 + (i-1)r$ for all $i \in
\{2,\ldots,n-1\}$, we have that $B_1 = \gcd\{h, r, d_n\}$, and then
$B_1 d_1 < d_2 < \cdots < d_{n-1}$, so $B_1 d_1 \in \N
\{d_2,\ldots,d_n\}$ if and only if $d_n \mid B_1 d_1$. However, if
$d_n \mid B_1 d_1$, by Proposition \ref{red} and Lemma
\ref{facilred} we have that $\oIA$ is a complete intersection if and
only if $\oIB$ so is, where $\mB := \{d_2, \ldots, d_n\}$, but $d_n
\mid B_1 d_1 \mid h d_1$ and $\mB$ is a generalized arithmetic
sequence; therefore, by Theorem \ref{genproy} this implies that $n =
4$ and $\{d_4, d_2, d_3\}$ is an arithmetic sequence, what we had
assumed not to happen. Concerning $B_2$, if $n \geq 5$, $d_1$ is odd
or $d_4$ is odd then $B_2 = 1$ and $B_2 a_2 \notin \sum_{j \in
\{1,\ldots,n+1\} \atop j \neq 2} \N a_j$. Otherwise, i.e., if $n =
4$ and $d_1$ and $d_4$ are even, we have that $B_2 = 2$ and set $\mC
:= \{d_1,d_3,d_4\}$. Moreover, $2 a_2 \in \N \{a_1,a_3,a_4,a_5\}$ if
and only if $2 d_2 = \alpha_1 d_1 + \alpha_3 d_3 + \alpha_4 d_4$
with $\alpha_1 + \alpha_3 + \alpha_4 \leq 2$, and this can only
happen if:
\begin{itemize}
\item[{\rm (a)}] $2 d_2 = d_1 + d_3$,
\item[{\rm (b)}] $2 d_2 = d_1 + d_4$, or
\item[{\rm (c)}] $2 d_2 = d_3 + d_4$.
\end{itemize}
If (a) holds we have that $\{d_1, d_2, d_3\}$ is an arithmetic
sequence and $\oIA$ is a complete intersection if and only if so is
$I_{\mC^{\star}}$. In (b) we have that $2 d_2 = h d_1 + d_3 = d_1 +
d_4$ and we deduce that $(h-1) d_1 + d_3 = d_4$, thus
one derives that ${\mathcal C}^{\star}_{red} \not=
\emptyset$ and $\oIA$ is not a complete intersection. Indeed,
denoting $B_i' := {\rm min}\{b \in \Z^+ \, \vert \, b a_i \in
\sum_{j \in \{1,3,4\} \atop j \neq i} \Z a_j\}$ for all $i \in
\{1,3,4\}$, from the previous equality it follows that $B_3' = B_4'
= 1$ and that $B_1' \leq h-1$, therefore $B_1' d_1 \notin \N
\{d_3,d_4\}$, due to $B_1' d_1 < d_3 < d_4$. In (c) we have that
$\{d_4,d_2,d_3\}$ is an arithmetic sequence, which we had assumed to
not to happen.
In the rest of cases we get that $B_i a_i \notin \N (\omA \setminus
\{a_i\})$ for all $i \in \{1,\ldots,n\}$, and then $\omA_{red} \not=
\emptyset$ and $\oIA$ is not a complete intersection. \QED
\end{demo}
\begin{Remark} Moreover, whenever $\oIA$ is a complete intersection, we get the following
minimal sets of generators of $\oIA$ depending on the cases
$($see {\rm Lemma \ref{facilred}} and
{\rm Proposition \ref{red}}$)$:
\begin{enumerate} \item Set $b = \gcd\{d_1,d_2\}$, if $n = 4$, $\{d_1,d_2,d_3\}$ is an arithmetic sequence, $d_1 / b$ is even and $b d_4 =
\beta_1 d_1 + \beta_2 d_2 + \beta_3 d_3$ with $\beta_1 + \beta_2 +
\beta_3 \leq b$, then
$$\oIA = \left( x_4^{b} - x_1^{\beta_1} x_2^{\beta_2} x_3^{\beta_3}
x_5^{b - \sum \beta_i},\,
x_2^2 - x_1 x_3,\, x_1^{d_3 / 2 b}
- x_3^{d_1 / 2 b} x_5^{(d_2 - d_1) / b} \right).$$
\item If $n = 4$, $\{d_1,d_2,d_3\}$ is an arithmetic sequence, $d_1, d_4$ are even
and ${\mathcal C}^{\star}_{red} = \emptyset$ with $\mathcal C =
\{d_1,d_3,d_4\}$, then
$$\oIA = \left( x_2^2 - x_1 x_3 \right) + I_{\mathcal C^{\star}} \cdot
k[x_1,x_2,x_3,x_4,x_5].$$
\end{enumerate}
\end{Remark}
\bigskip We finish this section studying when $\oIA$ is a complete intersection in the following families, where $\mA =
\{d_1,\ldots,d_n\}$, $n \geq 3$, $p,q \in \Z^+$ are relatively prime
and either
\begin{itemize}
\item $\mA$ consists of terms the $(p,q)$-Fibonacci sequence whose indices are
a generalized arithmetic sequence, i.e., there exist $h, a, d \in
\Z^+$ such that $d_1 = F_{a}$, $d_i = F_{ha + (i-1)d}$ for all $i
\geq 2$, or
\item $\mA$ consists of terms of the $(p,q)$-Lucas sequence whose indices are an arithmetic
sequence, i.e., there exist $a, d \in \Z^+$ such that $d_1 = L_{a}$,
$d_i = L_{ha + (i-1)d}$ for all $i \geq 2$.
\end{itemize}
In both cases we characterize when $\oIA$ is a complete intersection
by means of the input data; which are the values of $p, q, n, a$ and
$d$ (and $h$ for the Fibonacci sequence).
\begin{Theorem} \label{fibonacciproy}Let $p, q \in \Z^+$ be relatively prime and
let $\{F_n\}_{n \in \N}$ be the $(p,q)$-Fibonacci sequence. Set $\mA
= \{d_1,\ldots,d_n\}$ with $d_1 := F_a$ y $d_i := F_{ha + (i-1)d}$
for all $i \in \{2,\ldots,n\}$ where $h,\,a,\,d\in \Z^+$ and $n \geq
3$. Then, $\oIA$ is a complete intersection $\Longleftrightarrow$ $n
= 3$, $h = 1$, $d$ is even and $2d \mid a$.
\end{Theorem}
\begin{demo}We begin by observing that $B_i = 1$ for all $i \geq 3$
(see Lemma \ref{gcdfl}) and $B_1 = F_{\gcd\{ha,d\}} /
F_{\gcd\{a,d\}}$ and thus $B_1 d_1 \leq F_d d_1 < d_2 < \cdots <
d_n$, which implies that $B_1 a_1 \notin \sum_{j = 2}^{n+1} \N a_j$.
Let us study the possible values of $B_2$. If $n \geq 4$ or $[d]_2
\geq [a]_2$, then $B_2 = 1$, and if $n = 3$ and $[d]_2 < [a]_2$ then
$B_2 = F_{\gcd\{a,2d\}} / F_{\gcd\{a,d\}} = L_{\gcd\{a,d\}}$. If $d
\nmid a$, then $B_2 d_2 \leq L_{d-1} d_2 < L_{ha+2d-2} < d_3$ and
$B_2 d_2 \in \N d_1 + \N d_3$ if and only if $B_2 d_2 = \alpha d_1$,
but in this case $\alpha
> B_2$ and, by Remark \ref{notaproyred}, $B_2 a_2 \notin \N a_1 + \N a_3 + \N a_4$.
Suppose now that $d \mid a$, if $d$ is odd then $L_d d_2 = - q^d
(F_{ha}/d_1) d_1 + d_3 < d_3$ and again we have that $B_2 d_2 \in
\N d_1 + \N d_3$ if and only if $B_2 d_2 = \alpha d_1$, and thus
$B_2 a_2 \notin \N a_1 + \N a_3 + \N a_4$. Finally if $d$ is even we
have the inequality $B_2 d_2 = q^d (F_{ha}/d_1) d_1 + d_3 < 2 d_3$,
hence whenever $B_2 d_2 = \alpha_1 d_1 + \alpha_3 d_3$ with
$\alpha_1,\alpha_3 \in \N$, then $\alpha_3 < 2$; and if $\alpha_3 =
0$, then $\alpha_1 > B_2$. As a consequence, by Remark
\ref{notaproyred}, if follows that $B_2 a_2 \in \N a_1 + \N a_3 + \N
a_4$ if and only if $q^d (F_{ha} / d_1) + 1 \leq L_d$. If $h
> 1$, then $F_{ha}/d_1 \geq F_{2a}/ F_a = L_a \geq L_d$ and if $h =
1$ it follows that $L_d = F_{d+1} + q F_{d-1} > q^{d/2}\, F_1 + q\,
q^{d/2 - 1}\, F_1 = q^d$, and the result follows. \qed
\end{demo}
\medskip
\begin{Remark}Whenever $\oIA$ is a complete intersection, i.e., when $\mA = \{d_1,d_2,d_3\}$ with $d_1 = F_a$\,,
$d_2 = F_{a+d}$\,, $d_3 = F_{a+2d}$\,, $d$ is even and $2d \mid a$,
we get the following minimal set of generators of $\oIA$ $($see {\rm
Lemma \ref{facilred}} and {\rm Proposition \ref{red}}$):$
\begin{center}
$\oIA = \left(x_1^{ d_3 / F_{2d}} - x_3^{d_1 / F_{2d}}\, x_4^{(d_3
- d_1) / F_{2d}},\, x_2^{L_d} - x_1^{\,q^d}\, x_3\, x_4^{\, L_d -
q^d - 1} \right)$. \end{center}
\end{Remark}
\vspace{.4cm}
\begin{Theorem} \label{lucasproy}Let $p, q \in \Z^+$ be relatively prime and
let $\{L_n\}_{n \in \N}$ be the $(p,q)$-Fibonacci sequence. Set $\mA
= \{d_1,\ldots,d_n\}$ with $d_i := L_{a + (i-1)d}$ for all $i \in
\{1,\ldots,n\}$ where $a,\,d\in \Z^+$ and $n \geq 3$. Then, $\oIA$
is a complete intersection $\Longleftrightarrow$ $n = 3$, $d$ is
even, $p$ and $a/d$ are odd and, either $q$ is even or $3 \nmid d$
\end{Theorem}
\begin{demo}We begin by observing that $B_i = 1$ for all $i \geq 3$
(see Lemma \ref{gcdfl}). Let us study the possible values of
$B_2$, if $n \geq 4$ or $[d]_2
\neq [a]_2$, then $B_2 = 1$, and if $n = 3$ and $[d]_2 = [a]_2$ then
$B_2 = \gcd\{d_1,d_3\}\, / \, \gcd(\mA) = L_\gcd\{a,d\} \, /\,
\gcd(\mA)$ and $\gcd(\mA) \in \{1,2\}$. If $\gcd(\mA) = 2$ or
$\gcd\{a,d\} < d$, then $B_2 d_2 < d_3$ and thus $B_2 a_2 \notin \N
a_1 + \N a_3 + \N a_4$. Otherwise, if $\gcd(\mA) = 1$ and
$\gcd\{a,d\} = d$, then $B_2 d_2 = L_d d_2 = (-1)^d q^d d_1 + d_3$,
if $d$ is odd then $L_2 d_2 < d_3$ and again we have that $B_2 a_2
\notin \N a_1 + \N a_3 + \N a_4$; otherwise $B_2 d_2 = q^d d_1 +
d_3$ and $q^d + 1 \leq L_d$. From here we deduce that $\omA_{red} =
\emptyset$ if and only if $n = 3$, $[d]_2 = [a]_2$, $\gcd\{a,d\}=
d$, $\gcd\{d_1,d_2,d_3\} = 1$ and $d$ is even, by Lemma \ref{gcdfl}
the result follows. \qed
\end{demo}
\medskip
\begin{Remark} Whenever $\oIA$ is a complete intersection, i.e., when $\mA = \{d_1,d_2,d_3\}$ with $d_1 = L_a$\,,
$d_2 = L_{a+d}$\,, $d_3 = L_{a+2d}$\,, $d$ is even, $p$ and $a/d$
are odd and, either $q$ is even or $3 \nmid d$, we get the following
minimal set of generators of $\oIA$ $($see {\rm Lemma
\ref{facilred}} and {\rm Proposition \ref{red}}$):$
\begin{center}
$\oIA = \left( x_1^{d_3 / L_d} - x_3^{d_1 / L_d}\, x_4^{(d_3 - d_1)
/ L_{d}}, \, x_2^{L_d} - x_1^{q^d}\, x_3 \, x_4^{\, L_d - q^d - 1}
\right)$. \end{center}
\end{Remark}
\bibliographystyle{plain}
|
\section{Introduction}
Consider the systems
\begin{align}\label{original}
\left\{\begin{array}{cl}
-\Delta u_{i} = f_{i}(u) & \text{in } \mathbb{R}^n, \\
u_{i} > 0 & \text{in } \mathbb{R}^n, \\
\end{array}
\right.
\end{align}
where $i = 1, 2,\cdots, L$. Denote $\mathbb{R}^{L}_{+} = \{u\in\mathbb{R}^L|u_i>0, \text{ for } i=1,\cdots,L\}$, and throughout this article $f = (f_1, f_2,\cdots, f_L)$
is assumed continuous in $\overline{\mathbb{R}^{L}_{+}}$
and locally Lipschitz continuous in $\mathbb{R}^{L}_{+}$.
The existence and nonexistence of positive solution of this type of systems has long been studied.
An important example (after reducing the order) is the Hardy-Littlewood-Sobolev type of system,
\begin{equation} \label{HLS}
\left\{ \begin{aligned}
(-\triangle)^k u &= v^p & \text{in } \mathbb{R}^n,\\
(-\triangle)^k v &= u^q & \text{in } \mathbb{R}^n,
\end{aligned} \right.
\end{equation}
of which the Lane-Emden system is a special case (k=1).
The Lane-Emden system possesses positive solutions in critical and supercritical cases,
i.e. $\frac{1}{p+1}+\frac{1}{q+1}\leq 1-\frac{2}{n}$
(cf. \cite{SZ-1998});
meanwhile, the system admits no radial positive solution in subcritical cases, i.e. $\frac{1}{p+1}+\frac{1}{q+1}> 1-\frac{2}{n}$,
by Mitidieri \cite{Mitidieri1996}.
The Lane-Emden conjecture says that the system admits no positive solution in subcritical cases.
In \cite{CL2009-DCDS}, W. Chen and C. Li show that under an integrability condition, \eqref{HLS} with $k=1$ is equivalent to a system of integral equations of which all positive solutions are radial.
This, together with Mitidieri's result, partially solves the Lane-Emden conjecture.
Also, Souplet solves the 4-dimension case of Lane-Emden conjecture in \cite{Souplet2009}.
For more Liouville type of theorems, we refer readers to \cite{BrezisNirenberg1983, CGS, SG1981, NS} and the reference therein.
Traditionally, one can use a concentrated-compactness argument (cf. \cite{Lions1985}) to show existence of solution to critical cases of Lane-Emden systems.
In \cite{LiarXiv13}, C. Li introduces a new approach to obtain existence of radial positive solution for system \eqref{original} in critical and supercritical cases, which connects shooting method with degree theory and surprisingly relates the question to the non-existence of solution to a corresponding Dirichlet problem.
For recent development of degree approach to shooting method, reader can check \cite{LiVillavert2013, Villavert2014}.
The conditions Li places on $f$ include that
$f$ needs to be positive (i.e. $f_i$ is positive for all $i=1,\cdots,L$)
and satisfies a non-degenerate condition.
Although this covers a broad class of problems including critical and supercritical cases of \eqref{HLS}, there are cases of nonlinear Shr\"odinger type of systems that $f$ does not need to be always positive.
In this article, we are going to replace the condition $f$ being positive by $\sum f_i \geq 0$
where each $f_i$ can change sign.
Since we are looking for positive radial solution $u(x)=u(|x|)$ of \eqref{original}, the problem is equivalent to looking for global positive solution to the following ODE system,
\begin{align}\label{ode}
\left\{ \begin{array}{cl}
u_{i}^{''}(r) + \dfrac{n-1}{r}u_{i}^{'}(r) = -f_{i}(u(r)) \\
u_{i}^{'}(0) = 0, \ u_{i}(0) = \alpha_i \text{ for } i = 1,2,\cdots, L.
\end{array}\right.
\end{align}
where $\alpha = (\alpha_1, \alpha_2, \cdots, \alpha_L)$
is positive (i.e. each $\alpha_i> 0$) initial value for $u$.
By classical ODE theory, this initial value problem \eqref{ode} has a unique solution $u(r,\alpha)$
for $r$ in some maximum interval.
Let $r_{\alpha} := \inf_{r \geq 0}\{r\in\mathbb{R} | u(r,\alpha)\in \partial\mathbb{R}^{L}_{+}\}$
(by $\partial\mathbb{R}^{L}_{+}$ we mean the boundary of $\mathbb{R}^{L}_{+}$, which sometimes we call ``the wall"),
so $r=r_{\alpha}$ is where $u(r,\alpha)$ touches the wall for the first time.
There are two cases, case 1: $r_{\alpha}=\infty$, $u$ never hits wall,
then $u(r,\alpha)$ is a radial positive solution to \eqref{original}, and we are done; case 2: $r_{\alpha}<\infty$, $u$ hits wall in finite time, and
we will show that the existence of solution of \eqref{original} surprisingly depends on the non-existence of solution to its corresponding Dirichlet problem \eqref{Dirichlet} in the below,
\begin{align}\label{Dirichlet}
\left\{\begin{array}{cl}
-\Delta u_{i} = f_{i}(u) & \text{in } B_R, \\
u_{i} > 0 & \text{in } B_R, \\
u_i=0 & \text{on } \partial B_R,
\end{array}
\right.
\end{align}
where $B_R=B_R(0)$ for any $R>0$ and $i = 1,2,\cdots, L$.
It is well known that the nonexistence of solution
to this critical and supercritical Dirichlet problems \eqref{Dirichlet}
can be derived by Rellich-Poho\v{z}aev type of identities.
In fact, it belongs to a general variational problem, where
a ball of any radius can be replaced by a bounded star-shaped domain.
For this problem, we refer the reader to Pucci and Serrin's paper \cite{PS-Ind}.
In the last part of this paper, we will show the local nonexistence to prove global existence
to some example systems.
So, here is our main result,
\begin{theorem}\label{existencetheorem}
Given the nonexistence of solution to system \eqref{Dirichlet} for all $R>0$, the system \eqref{original} admits a radially symmetric solution of class $C^{2,\alpha}(\mathbb{R}^{n})$ with $0<\alpha<1$, if $f= (f_1(u), f_2(u),\cdots, f_L(u)):\mathbb{R}^L\rightarrow \mathbb{R}^L$ satisfies the following assumptions:
\begin{enumerate}
\item $f$ is continuous in $\overline{\mathbb{R}^{L}_{+}}$
and locally Lipschitz continuous in $\mathbb{R}^{L}_{+}$, and furthermore,
\begin{align}\label{decayAssumption}
\sum_{i=1}^L f_i(u) \geq 0 \text{ in } \mathbb{R}_+^L;
\end{align}
\item If $\alpha \in \partial\mathbb{R}^{L}_{+}$ and $\alpha\neq 0$, i.e., for some permutation $(i_1,\cdots,i_L)$, $\alpha_{i_1}=\cdots=\alpha_{i_m}=0$, $\alpha_{i_{m+1}},\cdots, \alpha_{i_L}>0$ where $m$ is an integer in $(0,L)$, then $\exists \delta_0 = \delta_0(\alpha) > 0$ such that for $\beta \in \mathbb{R}^{L}_{+}$ and $|\beta-\alpha|<\delta_0$,
\begin{align}\label{ControlInequality}
\sum_{j=m+1}^L |f_{i_j}(\beta)| \leq C(\alpha) \sum_{j=1}^m f_{i_j}(\beta),
\end{align}
where $C$ is a non-negative constant that depends only on $\alpha$.
\end{enumerate}
\end{theorem}
\begin{remark}\label{assumptionRemark}
Notice that under assumptions above, sign-changing source term $f$ is allowed. It is known that if $f_i$'s stay positive, there are many nice properties that we can use, c.f. \cite{LiarXiv13, LiVillavert2013, Villavert2014}. If $f_i$ changes sign, those properties are lost, which leads to estimates failing. Our work here is to derive dynamic estimate \eqref{dynamicEstimate1} and \eqref{dynamicEstimate2} under assumptions above, such that the degree theory approach to shooting method is applicable to show existence of solution with sign-changing $f$.
\end{remark}
The paper is organized as follows.
In section 2, we prove the main result.
In section 3, we will show nonexistence of solution to the corresponding Dirichlet problems
of some example systems.
\section{Proof of main result}
In this section, we will first define a target map and prove its continuity. Then we apply degree theory to prove theorem \ref{existencetheorem}.
\subsection{Target map}
For any real number $a>0$, let $\Sigma_a=\{\alpha\in\overline{\mathbb{R}^{L}_{+}}|\sum_{i=1}^L \alpha_i=a\}$, and $B_a= \{ \alpha \in \partial \mathbb{R}_+^L {\bf |}
\sum_{i=1}^L \alpha_i \leq a\}$. Recall that for positive $\alpha$ (i.e. every $\alpha_i>0$) we define $r_{\alpha} = \inf_{r \geq 0}\{r\in\mathbb{R} | u(r,\alpha)\in \partial\mathbb{R}^{L}_{+}\}$. As mentioned before, we can assume $r_{\alpha}<\infty$ since if $r_{\alpha}=\infty$ we get a solution to \eqref{original}. Then we define a target map on a initial data as following,
\begin{definition}\label{psi}
Let $u(r,\alpha)$ be a solution to \eqref{ode} with initial value $\alpha\in\Sigma_a$, we define a map
$\psi : \Sigma_a \rightarrow \partial\mathbb{R}^{L}_{+}$, such that
\begin{align}
\psi(\alpha) = \left\{\begin{array}{ll}
u(r_{\alpha},\alpha) & \alpha\in\mathbb{R}^{L}_{+}, \\
\alpha & \alpha\in\partial\mathbb{R}^{L}_{+}.
\end{array}
\right.
\end{align}
\begin{comment}
\begin{align*}
\left\{ \begin{array}{lcl}
\psi(\alpha) = u(r_{\alpha}), & \ \text{if} \ r_{\alpha} < \infty \\
\psi(\alpha) = \lim_{r \rightarrow \infty} u(r), & \ \text{otherwise}
\end{array}\right.
\end{align*}
\end{comment}
\end{definition}
Here we sketch Li's degree theory approach for shooting method
as follows.
Fix any real number $a>0$,
and assume that for any initial value $\alpha\in\Sigma_a$
no global positive solution to \eqref{ode} exists
(i.e. $r_{\alpha}<\infty$), so we can define a target map. Hence,
{\bf step 1}, we show that, under some suitable assumptions on $f$,
the target map $\psi$ is continuous from $\Sigma_a$ to $B_a$;
{\bf step 2}, by degree theory we show that $\psi$ is onto,
therefore $\exists \alpha_0\in\Sigma_a$ such that $\psi(\alpha_0)=u(r_{\alpha_0},\alpha_0)=0$;
{\bf step 3}, note that by assumption $r_{\alpha_0}<\infty$, $u(r,\alpha_0)$
for $r\in[0,r_{\alpha_0}]$ is a solution to the Dirichlet problem \eqref{Dirichlet},
which makes a contradiction
if we assume that system \eqref{Dirichlet} admits no solution for any $R>0$.
\begin{comment}
\begin{remark}
The nonexistence result to the Dirichlet problem \eqref{Dirichlet} on a ball of any radius
or more generally
on a bounded star-shaped domain is obtained in many cases.
Such as Pucci-Serrin's results in \cite{PS-Ind},
by applying Poho\v{z}aev type of identity one finds that under certain assumptions on $f$,
\eqref{Dirichlet} admits no solution for all $R>0$.
In section 3 by showing nonexistence to their corresponding Dirichlet problem.
\end{remark}
\end{comment}
In what follows, we assume \eqref{ode} admits no global positive solution, i.e. $r_{\alpha}<\infty$. We first show that under our assumptions \eqref{decayAssumption} and \eqref{ControlInequality} on $f$, the behavior of $u$ turns regular, such that $\psi$ is continuous.
\begin{lemma}\label{continuityLemma}
For any real number $a>0$, let
\begin{align*}
\Sigma_a=\{\alpha\in\overline{\mathbb{R}^{L}_{+}}|\sum_{i=1}^L \alpha_i=a\} \text{ and } B_a= \{ \alpha \in \partial \mathbb{R}_+^L | \sum_{i=1}^L \alpha_i \leq a\}.
\end{align*}
The target map $\psi$ defined in definition \ref{psi} is a continuous map from $\Sigma_a$ to $B_a$ if $f$ satisfies assumptions \eqref{decayAssumption} and \eqref{ControlInequality}.
\end{lemma}
Proof.
To see that $\psi$ maps $\Sigma_a$ to $B_a$, we need to notice that by assumption \eqref{decayAssumption} $\Sigma_{i=1}^L f_i\geq 0$, so we solve from the ODE system \eqref{ode} and get
\begin{align*}
\Sigma_{i=1}^L u_i(r,\alpha) &= \Sigma_{i=1}^L \alpha_i - \Sigma_{i=1}^L\int_0^r\int_0^s (\frac{\tau}{s})^{n-1} f_i(u(\tau)) d\tau ds \\
&\leq \Sigma_{i=1}^L \alpha_i,
\end{align*}
for $r\in[0,r_\alpha]$. Therefore, $\psi(\alpha)\in B_a$.
Next, we will show that $\psi$ is continuous on $\Sigma_a$. Fix any $\overline{\alpha}\in\Sigma_a$, then $\overline{\alpha}$ lies on the boundary of $\mathbb{R}^{L}_+$ or in $\mathbb{R}^{L}_+$ ($\overline{\alpha}\neq 0$ since $a>0$), and we will prove for these two cases.
Case 1. $\overline{\alpha} \in \partial\mathbb{R}^{L}_+$.
Suppose $\overline{\alpha}_{i_1}=\cdots=\overline{\alpha}_{i_m}=0$, and $\overline{\alpha}_{i_{m+1}},\cdots, \overline{\alpha}_{i_L}>0$, for some integer $m$ that $0<m<L$.
By the second assumption \eqref{ControlInequality}, $\exists \delta_0>0$, such that for $\alpha \in \mathbb{R}^{L}_{+}$ satisfying $|\alpha-\overline{\alpha}|<\delta_0$ we have
\begin{align}\label{environmentControl}
\sum_{j=m+1}^L |f_{i_j}(\alpha)| \leq C(\overline{\alpha}) \sum_{1\leq j\leq m} f_{i_j}(\alpha).
\end{align}
Notice that we can choose $C=C(\overline{\alpha})\geq 1$ in \eqref{environmentControl} if $\sum_{1\leq j\leq m} f_{i_j}(\alpha)\geq 0$. Indeed,
if $C=0$, the term on the left of the inequality \eqref{environmentControl} is zero, and due to the first assumption \eqref{decayAssumption}, $\sum_{1\leq j\leq m} f_{i_j}(\alpha)\geq 0$. If $C>0$, then $\sum_{1\leq j\leq m} f_{i_j}(\alpha)\geq 0$ obviously. So, we choose $C\geq 1$.
For this $\delta_0$ and $C$, we {\bf{claim}} that:
\emph{For $\alpha\in\Sigma_a$, fix any $\delta < \frac{\delta_0}{2(3+L)C}$, and if $|\alpha-\overline{\alpha}|\leq \delta$, then for $r\in[0,r_{\alpha}]$
\begin{align}\label{dynamicEstimate1}
|u(r,\alpha)-\overline{\alpha}| \leq 2(3+L)C\delta < \delta_0.
\end{align}
}
As we will see in the following proof, \eqref{dynamicEstimate1} is a dynamic estimate,
in the sense that if $|u(r,\alpha)-\overline{\alpha}|<\delta_0$
with $r\in[0, a_1)\subset[0,r_{\alpha}]$ for some $a_1>0$,
then by \eqref{ControlInequality} we have
\begin{align}
\sum_{j=m+1}^L |f_{i_j}(u(r,\alpha)| \leq C(\overline{\alpha}) \sum_{1\leq j\leq m} f_{i_j}(u(r,\alpha)),
\end{align}
and this control enable us to push the range of $r$ in
$|u(r,\alpha)-\overline{\alpha}|<\delta_0$
further than $a_1$ and up to $r_{\alpha}$.
Suppose the claim not true, then there exists $\alpha_0 \in\mathbb{R}^{L}_{+}$
satisfying $|\alpha_0-\overline{\alpha}|\leq\delta$ and $a_1\in(0,r_{\alpha_0})$ such that the equality of \eqref{dynamicEstimate1} holds at $r=a_1$ for the first time, i.e.
\begin{align}\label{firstTimeEquality}
|u(r,\alpha_0)-\overline{\alpha}|\left\{\begin{array}{ll}
< 2(3+L)C\delta, & \text{if } r<a_1, \\
= 2(3+L)C\delta, & \text{if } r=a_1.
\end{array}
\right.
\end{align}
For $r\in(0,a_1)$ we have,
\begin{align}
|u(r,\alpha_0)-\overline{\alpha}|
&\leq |u(r,\alpha_0)-\alpha_0| + |\alpha_0-\overline{\alpha}|\\
&\leq \sum_{j=1}^m|u_{i_j}(r,\alpha_0)-\alpha_{0i_j}|+\sum_{j=m+1}^L|u_{i_j}(r,\alpha_0)-\alpha_{0i_j}| + |\alpha_0-\overline{\alpha}|. \label{threeTerms}
\end{align}
So, to estimate the second term of \eqref{threeTerms}, we solve from \eqref{ode} and get
\begin{align*}
\sum_{j=m+1}^L|u_{i_j}(r,\alpha_0)-\alpha_{0i_j}|
&= \sum_{j=m+1}^L|\int_0^r\int_0^s (\frac{\tau}{s})^{n-1} f_{i_j}(u(\tau)) d\tau ds| \\
\end{align*}
Notice that since $C\geq 1$, $2(3+L)C\delta<\delta_0$, therefore assumption \eqref{ControlInequality} can be applied on $u(r,\alpha_0)$ for $r\in(0,a_1)$,
\begin{align*}
\sum_{j=m+1}^L|u_{i_j}(r,\alpha_0)-\alpha_{0i_j}|
&\leq \sum_{j=m+1}^L \int_0^r\int_0^s |(\frac{\tau}{s})^{n-1} f_{i_j}(u(\tau))| d\tau ds \\
&\leq C \int_0^r\int_0^s (\frac{\tau}{s})^{n-1} \sum_{j=1}^m f_{i_j}(u(\tau)) d\tau ds \\
&= C \sum_{j=1}^m (\alpha_{0i_j}-u_{i_j}(r,\alpha_0)),
\end{align*}
The first term of \eqref{threeTerms} can be estimated by
\begin{align*}
\sum_{j=1}^m|u_{i_j}(r,\alpha_0)-\alpha_{0i_j}|
&= \sum_{j=1}^m(\alpha_{0i_j}-u_{i_j}(r,\alpha_0))^+ + \sum_{j=1}^m(\alpha_{0i_j}-u_{i_j}(r,\alpha_0))^- \\
&\leq 2\sum_{j=1}^m(\alpha_{0i_j}-u_{i_j}(r,\alpha_0))^+.
\end{align*}
To see the inequality above, one needs to notice that due to \eqref{ControlInequality}, $\sum_{j=1}^m f_{i_j}(u)\geq 0$, so $\sum_{j=1}^m u_{i_j}(r,\alpha_0)$ is monotone decreasing on $[0,r_{\alpha_0}]$. So,
\begin{align*}
\sum_{j=1}^m(\alpha_{0i_j} - u_{i_j}(r,\alpha_0))
= \sum_{j=1}^m(\alpha_{0i_j}-u_{i_j}(r,\alpha_0))^+ - \sum_{j=1}^m(\alpha_{0i_j}-u_{i_j}(r,\alpha_0))^- \geq 0.
\end{align*}
The last term of \eqref{threeTerms} is bounded by $\delta$, and notice that $u_i(r,\alpha_0)>0$, $i=1,\cdots, L$ for $r\in(0,r_{\alpha_0})$, so we get
\begin{align*}
|u(r,\alpha_0)-\overline{\alpha}|
&\leq 2\sum_{j=1}^m(\alpha_{0i_j}-u_{i_j}(r,\alpha_0))^+ + C \sum_{j=1}^m (\alpha_{0i_j}-u_{i_j}(r,\alpha_0)) + \delta\\
&\leq (2+C) \sum_{j=1}^m \alpha_{0i_j} + \delta \\
&\leq (2+C)L|\alpha_0-\overline{\alpha}| + \delta \\
&\leq (3+C)L\delta,
\end{align*}
where $C$ is the same $C$ in \eqref{ControlInequality}. Then we get a contradiction with \eqref{firstTimeEquality} by taking $r=a_1$ in the above. Hence the claim is proved.
\begin{comment}
Now, for any $\delta\in(0,\delta_1)$ and any $\alpha\in\mathbb{R}_+^L$ satisfying $|\alpha-\overline{\alpha}|\leq \delta$, the claim ensures that if we substitute $\delta_1$ with $\delta$ the argument above is all eligible for $r\in[0,r_{\overline{\alpha}}]$. Namely,
\begin{align}
|u(r,\alpha)-\overline{\alpha}|\leq (3+C)L\delta
\end{align}
for $r\in[0,r_{\alpha}]$. So,
\begin{align}
|\psi(\alpha)-\overline{\alpha}|=|u(r_{\alpha},\alpha)-\overline{\alpha}|\leq (3+C)L\delta
\end{align}
\end{comment}
Notice that the claim and estimate \eqref{dynamicEstimate1} implies the continuity of $\psi$ at $\overline{\alpha}$, therefore, we have proved case 1.
Case 2. $\overline{\alpha} \in \mathbb{R}^{L}_+$.
As above we assume $r_{\overline{\alpha}}<\infty$, and obviously $r_{\overline{\alpha}}>0$. Let's assume at $r=r_{\overline{\alpha}}$, for some integer $m$ that $0< m\leq L$ ($m>0$ because $\psi(\overline{\alpha})=u(r_{\overline{\alpha}},\overline{\alpha})\in\partial\mathbb{R}^L_+$, i.e. $u$ touches wall at $r=r_{\overline{\alpha}}$), so suppose $u_{i_1}(r_{\overline{\alpha}},\overline{\alpha})=\cdots=u_{i_m}(r_{\overline{\alpha}},\overline{\alpha})=0$, and $u_{i_{m+1}}(r_{\overline{\alpha}},\overline{\alpha}),\cdots,u_{i_L}(r_{\overline{\alpha}},\overline{\alpha})>0$. Let $\overline{\omega}(r)=\sum_{j=1}^m u_{i_j}(r,\overline{\alpha})$, and we claim that $\overline{\omega}'(r_{\overline{\alpha}})<0$.
By the continuity of $u(r,\overline{\alpha})$ with respect to $r$, $\exists \delta_1>0$ such that for $r\in(r_{\overline{\alpha}}-\delta_1,r_{\overline{\alpha}}]$,
\begin{align}
|u(r,\overline{\alpha})-\psi(\overline{\alpha})|<\delta_0.
\end{align}
If $m<L$, the assumption \eqref{ControlInequality} is taking effect, and therefore
$\sum_{j=1}^m f_{i_j} \geq 0$ for $r\in(r_{\overline{\alpha}}-\delta_1,r_{\overline{\alpha}}]$;
if $m=L$ then by the assumption \eqref{decayAssumption},
we also have $\sum_{j=1}^m f_{i_j} \geq 0$ for $r\in(r_{\overline{\alpha}}-\delta_1,r_{\overline{\alpha}}]$.
So, in $(r_{\overline{\alpha}}-\delta_1,r_{\overline{\alpha}}]$
\begin{align}\label{dynamicEstimate2}
-\frac{1}{r^{n-1}}(r^{n-1}\overline{\omega}'(r))'=\sum_{j=1}^m f_{i_j} \geq 0.
\end{align}
Also, since $\overline{\omega}(r)>0$ for $r<r_{\overline{\alpha}}$ and $\overline{\omega}(r_{\overline{\alpha}})=0$,
there must exist $r_0\in(r_{\overline{\alpha}}-\delta_1,r_{\overline{\alpha}})$, such that $\overline{\omega}'(r_0)<0$.
So, for $r\in[r_0,r_{\overline{\alpha}}]$,
\begin{align}
\overline{\omega}'(r) = (\frac{r_0}{r})^{n-1}\overline{\omega}'(r_0)
-\int_{r_0}^r(\frac{\tau}{r})^{n-1}\sum_{j=1}^m f_{i_j}(u(\tau))d\tau<0.
\end{align}
This proves the claim $\overline{\omega}'(r_{\overline{\alpha}})<0$.
Then there exists $l_0\in\{1,\cdots, m\}$
such that $u_{i_{l_0}}'(r_{\overline{\alpha}},\overline{\alpha})<0$.
Therefore, combining with the fact $u_{i_{l_0}}(r_{\overline{\alpha}},\overline{\alpha})=0$, we see $u_{i_{l_0}}$ crosses the wall with a non-zero slope, i.e. there is a transversality at $r=r_{\overline{\alpha}}$, and by classical ODE stability theory (the continuous dependence on initial value) $\psi$ is continuous at $\overline{\alpha}$. $\Box$
\subsection{Application of degree theory}
Now, let's recall some results in degree theory (in particular, the treatment modified by P.Lax, c.f. \cite{Nirenberg2001}). Consider $C^{\infty}$ oriented manifolds $X_0,Y$ of dimension $n$ (all manifolds are assumed to be paracompact) and an open subset $X\subseteq X_0$ with compact closure. For convenience, write $dy^1\wedge\cdots\wedge dy^n = dy$. Then for a $C^1$ map $\phi:X\rightarrow Y$, the degree is defined as following:
\begin{definition}\label{degree}
Let $\mu=f(y)dy$ be a $C^{\infty}$ $n$-form with support contained in a coordinate patch $\Omega$ of $y_0$ and lying in $Y\setminus\{\phi(\partial X)\}$ such that $\int_Y \mu=1$; set
\begin{align}
deg(\phi,X,y_0)=\int_X \mu\circ\phi.
\end{align}
\end{definition}
Here are some properties of degree which we will refer to in our proof,
\begin{proposition}
For $y_1$ close to $y_0$, $deg(\phi,X,y_0)=deg(\phi,X,y_1)$.
\end{proposition}
It follows that the degree of a mapping is constant on any connected component $C$ of $Y\setminus \{\phi(\partial X)\}$, and we can write degree as $deg(\phi,X,C)$.
\begin{proposition}\label{ontoProperty}
If $y_0\notin\phi(\overline{X})$, then $deg(\phi,X,y_0)=0$.
\end{proposition}
As a result, if $deg(\phi,X,y_0)\neq0$, then $y_0\in\phi(\overline{X})$.
An important property of degree is that, the notion can be extended to maps $\phi$ which are merely continuous, since we can approximate such $\phi$ by $C^1$ maps $\phi_n$ (see Property 1.5.3 in \cite{Nirenberg2001}). Also, degree is homotopy invariant, which enables us to define degree for continuous map $\eta:\partial X\rightarrow \mathbb{R}^n\setminus y_0$ (see Property 1.5.4 in \cite{Nirenberg2001}), which leads to the following theorem (see Property 1.5.5 in \cite{Nirenberg2001}),
\begin{theorem}\label{homotopyInvariance}
$deg(\eta, X, y_0)$ depends only on the homotopy class of $\eta: \partial X\rightarrow \mathbb{R}^n\setminus y_0$.
\end{theorem}
Now we are prepared to prove the existence of solution to \eqref{original}.\\
\begin{comment}
\begin{proposition}[Homotopy Invariance]
Consider a one parameter family of maps $\phi_t(x):\overline{X}\times\[0,1\]\rightarrow Y$, continuous on $\overline{X}\times\[0,1\]$ and $C^1(X)$ for each $t\in\[0,1\]$. Suppose for all $t$, $y_0\notin\phi_t(\partial X)$, then $deg(\phi_t,X,y_0)$ is independent of t.
\end{proposition}
\end{comment}
\noindent{\bf Proof of theorem \ref{existencetheorem}. }
For any fixed real number $a>0$, assume that \eqref{ode} admits no global positive solution with any initial value $\alpha\in\Sigma_a$, so $r_{\alpha}<\infty$, and then we can define a target map $\psi$ by \eqref{psi}.
Recall that $\Sigma_a=\{\alpha\in\overline{\mathbb{R}^{L}_{+}}|\sum_{i=1}^L \alpha_i=a\}$ and $B_a= \{ \alpha \in \partial \mathbb{R}_+^L {\bf |}
\sum_{i=1,\cdots, L} \alpha_i \leq a\}$, and by lemma \ref{continuityLemma} $\psi$ is a continuous maps from $ \Sigma_a \longrightarrow B_a$.
Let $\pi(\alpha)=\alpha+\frac{1}{L}(a- \displaystyle\sum_{i=1,\cdots, L} \alpha_i)(1,\cdots,1): B_a \longrightarrow \Sigma_a$,
then $\pi$ is continuous with a continuous inverse $\pi^{-1}(\alpha)=\alpha-(\displaystyle\min_{i=1,\cdots, L} \alpha_i)(1,\cdots,1): \Sigma_a \longrightarrow B_a$.
The map: $\phi=\pi \circ \psi: \Sigma_a \longrightarrow \Sigma_a$ is continuous and $\phi(\alpha)=\alpha$
on $\partial\Sigma_a$. Let $\eta=id$ (the identity map) and $X=\Sigma_a\setminus \partial\Sigma_a$,
and then by theorem \ref{homotopyInvariance} we have $deg(\phi, X, \alpha)=deg(\eta, X, \alpha)=1$ for any
$\alpha \in \Sigma_a\setminus\partial\Sigma_a$. By property \ref{ontoProperty}, $\phi$ is onto, which implies that $\psi$ is also onto. this shows that there exists an $\alpha_0 \in \Sigma_a$ such that $\psi(\alpha_0)=0$.
Since we assume that system \eqref{Dirichlet} admits no solution, $r_{\alpha_0}$ corresponding to this $\alpha_0$ cannot be finite, a contradiction. This completes the proof of theorem \ref{existencetheorem}. $\Box$
\section{Examples}
One of the simplest systems is that $f\equiv 0$ in \eqref{original}, then $u\equiv C$ for some constant vector $C$ is a solution. Let us point out that since $f$ is not positive, this trivial system is not included in previous results (cf. \cite{LiarXiv13, LiVillavert2013, Villavert2014}), but it is included in our cases. In this section, we will show the existence of solution to some non-trivial systems. In the view of theorem \ref{existencetheorem}, we only need to show their corresponding Dirichlet problems admit no solution, and verify its source term satisfy our assumptions in theorem \ref{existencetheorem}.
\subsection{Sign-changing source terms}
Here we give a simple but non-trivial example of sign-changing source terms system, and we believe there are many other non-linear Schr\"odinger type of systems with sign-changing source terms which our method can be applied to.
Consider the following system,
\begin{align}\label{signChanging}
\left\{\begin{array}{ll}
-\Delta u = v^p-u^p, \\
-\Delta v = u^p, \\
\ \ u,v > 0, \\
\end{array}
\right.\text{in } \mathbb{R}^n,
\end{align}
and its corresponding Dirichlet problem,
\begin{align}\label{signChangingDirichlet}
\left\{\begin{array}{ll}
-\Delta u = v^p-u^p & \text{in } B, \\
-\Delta v = u^p & \text{in } B, \\
\ \ u,v > 0 & \text{in } B, \\
\ \ u,v=0 & \text{on } \partial B,\\
\end{array}
\right.
\end{align}
where $B=B_R(0)\subset \mathbb{R}^n$ for any $R>0$. We have
\begin{theorem}\label{signChangingExistenceThm}
If $p\geq\frac{n+2}{n-2}$, then \eqref{signChanging} admits radial positive solution.
\end{theorem}
Again the proof relies on the non-existence of solution to \eqref{signChangingDirichlet}. The non-existence is obtained by computing Rellich-Poho\v{z}aev type identity.
\begin{lemma}\label{signChangingDirichletNonEx}
If $p\geq\frac{n+2}{n-2}$, then system \eqref{signChangingDirichlet} admits no solution for any $R>0$.
\end{lemma}
Proof. Suppose there exists a positive solution $(u,v)$.
{\bf{Step 1.}} Claim the following identity,
\begin{align}\label{pohozaevCross}
\int_B \Delta u(x\cdot \nabla v) + \Delta v(x\cdot\nabla u) - (n-2)\nabla u\cdot \nabla v dx= \int_{\partial B} (x\cdot \nu)(\frac{\partial u}{\partial \nu}\frac{\partial v}{\partial \nu}) d\sigma>0,
\end{align}
where $\nu$ is the outward normal. The calculation of the above identity usually goes as
\begin{align*}
\int_B \Delta u(x\cdot \nabla v) dx &= \int_B \text{div} (\nabla u (x\cdot \nabla v)) - \nabla u\cdot \nabla(x\cdot \nabla v) dx \\
&= \int_{\partial B} (\nabla u \cdot \nu)(x\cdot \nabla v) d\sigma- \int_B ( \nabla u \cdot \nabla v + x_j \partial_i u \partial^2_{ij} v) dx.
\end{align*}
Then do the same to the second integrand in \eqref{pohozaevCross} and sum it with the above one to get
\begin{align*}
&\int_B \Delta u(x\cdot \nabla v) + \Delta v(x\cdot\nabla u) dx \\
&= \int_{\partial B} \frac{\partial u}{\partial \nu}(x\cdot \nabla v) +\frac{\partial v}{\partial \nu}(x\cdot \nabla u) d\sigma - \int_B (2 \nabla u \cdot \nabla v + x_j \partial_i u \partial^2_{ij} v+x_j \partial_i v \partial^2_{ij} u) dx \\
&= \int_{\partial B} \frac{\partial u}{\partial \nu}(x\cdot \nabla v) +\frac{\partial v}{\partial \nu}(x\cdot \nabla u) d\sigma - \int_B (2 \nabla u \cdot \nabla v + x\cdot\nabla(\nabla u\cdot\nabla v)) dx \\
&= \int_{\partial B} \frac{\partial u}{\partial \nu}(x\cdot \nabla v) +\frac{\partial v}{\partial \nu}(x\cdot \nabla u) d\sigma - \int_B 2 \nabla u \cdot \nabla v dx - \int_{\partial B} x\cdot \nu (\nabla u\cdot\nabla v)) d\sigma + n\int_B \nabla u\cdot\nabla v dx.
\end{align*}
Notice the fact that $x=|x|\nu$ on $\partial B$, and $\nabla u = \frac{\partial u}{\partial \nu} \nu$ and $\nabla v = \frac{\partial v}{\partial \nu} \nu$ due to $u,v=0$ on $\partial B$, and after rearrangement we have the identity \eqref{pohozaevCross}. Also, by Hopf's Lemma $\frac{\partial u}{\partial \nu}<0$ and $\frac{\partial v}{\partial \nu}<0$, so we have RHS of \eqref{pohozaevCross} $>0$.
Now, since
\begin{align*}
-\int_B v\Delta u dx = \int_B \nabla u\cdot \nabla v dx = -\int_B u\Delta v dx,
\end{align*}
for $\theta\in[0,1]$ we have
\begin{align}\label{energyCross}
\int_B \nabla u\cdot \nabla v dx = - \int_B \theta v\Delta u + (1-\theta) u\Delta vdx.
\end{align}
Also, notice that in \eqref{pohozaevCross} if we let $u=v$, then
\begin{align}\label{pohozaevSingle}
\int_B \Delta u (x\cdot \nabla u) dx= \int_B \frac{n-2}{2}|\nabla u|^2 dx + \frac 1 2 \int_{\partial B} x\cdot \nu |\nabla u|^2 d\sigma>0.
\end{align}
{\bf{Step 2.}} Merge the source terms into \eqref{pohozaevCross} (we replace $-u^p$ in the first source term by $\Delta v$, i.e. $-\Delta u=v^p+\Delta v$ and $-\Delta v = u^p$), and by \eqref{energyCross} we have
\begin{align*}
& \text{LHS of \eqref{pohozaevCross}} = \int_B -(v^p+\Delta v)(x\cdot \nabla v) - u^p(x\cdot \nabla u) - (n-2)\nabla u\cdot \nabla v dx \\
&=\int_B -x\cdot \nabla\left( \frac{v^{p+1}}{p+1}+\frac{u^{p+1}}{p+1}\right) - \Delta v (x\cdot \nabla v)
+ (n-2)\left( \theta v\Delta u + (1-\theta) u\Delta v\right) dx \\
&=\int_B -x\cdot \nabla\left( \frac{v^{p+1}}{p+1}+\frac{u^{p+1}}{p+1}\right) - \Delta v (x\cdot \nabla v)
+ (n-2)\left( -\theta (v^{p+1}+v\Delta v) - (1-\theta) u^{p+1}\right) dx \\
&= \int_B \left\lbrace \left( \frac {n} {p+1} - (n-2)\theta\right) v^{p+1} + \left( \frac {n} {p+1} - (n-2)(1-\theta)\right) u^{p+1} \right\rbrace dx \\
&\quad - \int_B \Delta v (x\cdot \nabla v) + (n-2)\theta v\Delta v dx
\end{align*}
So, by \eqref{pohozaevSingle}, \eqref{pohozaevCross} becomes
\begin{align*}
&\int_B \left\lbrace \left( \frac {n} {p+1} - (n-2)\theta\right) v^{p+1} + \left( \frac {n} {p+1} - (n-2)(1-\theta)\right) u^{p+1} \right\rbrace dx \\
&= \int_B \Delta v (x\cdot \nabla v) + (n-2)\theta v\Delta v dx +\int_{\partial B} (x\cdot \nu)(\frac{\partial u}{\partial \nu}\frac{\partial v}{\partial \nu}) d\sigma \\
&= \int_B \frac{n-2}{2}|\nabla v|^2 dx+\frac 1 2 \int_{\partial B} x\cdot \nu |\nabla u|^2 d\sigma
-\int_B\theta(n-2)|\nabla v|^2 dx
+ \int_{\partial B} (x\cdot \nu)(\frac{\partial u}{\partial \nu}\frac{\partial v}{\partial \nu}) d\sigma \\
&= \int_B (1-2\theta)\frac{n-2}{2}|\nabla v|^2 dx+\frac 1 2 \int_{\partial B} x\cdot \nu |\nabla u|^2 d\sigma
+ \int_{\partial B} (x\cdot \nu)(\frac{\partial u}{\partial \nu}\frac{\partial v}{\partial \nu}) d\sigma.
\end{align*}
Take $\theta=\frac 1 2$, and we have
\begin{align}\label{pohozaevEx1}
\int_B \left( \frac {n} {p+1} - \frac{n-2}{2}\right) \left( v^{p+1} + u^{p+1}\right) dx
= \frac 1 2 \int_{\partial B} x\cdot \nu |\nabla u|^2 d\sigma
+ \int_{\partial B} (x\cdot \nu)(\frac{\partial u}{\partial \nu}\frac{\partial v}{\partial \nu}) d\sigma >0.
\end{align}
So, by assumption $p\geq \frac{n+2}{n-2}$ the LHS of the above identity $\leq 0$, a contradiction.
$\Box$ \\
\noindent {\bf{Proof of Theorem \ref{signChangingExistenceThm}.}}
Directly we see that \eqref{signChanging} satisfies our first main assumption \eqref{decayAssumption}.
To see \eqref{ControlInequality} is satisfied, just notice that if $u=0$ then $|f_2|=0\leq f_1=v^p$, and if $v=0$ then $|f_1|=u^p\leq f_2$. Then for a neighborhood of such $(u,v)$ (i.e. $uv=0$), \eqref{ControlInequality} holds.
So, combined with Lemma \ref{signChangingDirichletNonEx}, Theorem \ref{signChangingExistenceThm} follows from Theorem \ref{existencetheorem}. \\
$\Box$
\begin{remark}
Actually system \eqref{signChanging} can be solved as follows. Suppose $-\Delta w = w^p$, let $u=\lambda w$ and $v=\nu w$, then we can find suitable $\lambda,\nu$ such that $(u,v)$ is a solution.
So, to manifest a nontrivial application of Theorem \ref{existencetheorem} we consider a very similar system,
\begin{align}\label{example2}
\left\{\begin{array}{ll}
-\Delta u = v^p +v^q -u^p, \\
-\Delta v = u^p, \\
\ \ u,v > 0, \\
\end{array}
\right. \text{in } \mathbb{R}^n,
\end{align}
where $p\neq q$ and $p,q\geq \frac{n+2}{n-2}$.
To show this system admits a solution, we only need to show the nonexistence of solution to corresponding Dirichlet problem (source terms satisfy assumptions \eqref{decayAssumption} and \eqref{ControlInequality}, see similar proof of Theorem \ref{signChangingExistenceThm}). Then the first step is exactly the same as the proof of Lemma \ref{signChangingDirichletNonEx}. In the second step, we use $-\Delta u = v^p+v^q+\Delta v$ and $-\Delta v=u^p$ to merge the source term, and taking $\theta=\frac 1 2$ \eqref{pohozaevEx1} becomes
\begin{align}
\begin{array}{cc}
\int_B \left( \frac {n} {p+1} - \frac{n-2}{2}\right) \left( v^{p+1} + u^{p+1}\right) +\left( \frac{n}{q+1} -\frac{n-2}{2}\right)v^{q+1} dx \quad\\
= \quad \frac 1 2 \int_{\partial B} x\cdot \nu |\nabla u|^2 d\sigma
+ \int_{\partial B} (x\cdot \nu)(\frac{\partial u}{\partial \nu}\frac{\partial v}{\partial \nu}) d\sigma >0.
\end{array}
\end{align}
Then the nonexistence to Dirichlet problem follows.
\end{remark}
\subsection{Conservative source terms}
In this section, we consider a system that $f$ has a potential function $F$, i.e. $f=\nabla F$, and $F(0)=0$.
{\bf{Type I.}} Consider the following system,s
\begin{align}\label{systemPotential}
\left\{\begin{array}{cl}
-\Delta u_{i} = \frac{\partial F}{\partial u_i}, \\
u_{i} > 0 ,\\
\end{array}
\right. \text{in } \mathbb{R}^n,
\end{align}
where $i=1,\ldots,L$.
The corresponding Dirichlet problem is
\begin{align}\label{systemPotentialDirichlet}
\left\{\begin{array}{cl}
-\Delta u_{i} = \frac{\partial F}{\partial u_i} & \text{in } B, \\
u_{i} > 0 & \text{in } B, \\
u_i = 0 &\text{on } \partial B,
\end{array}
\right.
\end{align}
where $i=1,\ldots,L$, and $B=B_R(0)\subset \mathbb{R}^n$ for any $R>0$.
In \cite{PS-Ind}, Pucci and Serrin have showed that for a general variational problem,
\begin{align*}
\int_{\Omega}\mathcal{F}(x,u,Du)dx = 0,
\end{align*}
there exists Poho\v{z}aev type of identity that can give
a sufficient condition on the nonexistence of solution to
Dirichlet problem on a bounded star-shaped domain.
In this case, the system \eqref{systemPotential} corresponds to
a vector-valued extremal of the variational problem with
\begin{align*}
\mathcal{F}(x,u,Du) = \frac{1}{2} \sum_{k=1}^{L} |p^k|^2 - F(u),
\end{align*}
where $p^k=Du_k$. So, theorem 6 in \cite{PS-Ind} leads to the following result (we will also give a simple proof for completion),
\begin{lemma}\label{PucciSerrin}
If for $u\neq 0$,
\begin{align}\label{potentialCondition}
\frac{n-2}{2}u^k \frac{\partial F}{\partial u^k} - nF(u) > 0,
\end{align}
then system \eqref{systemPotentialDirichlet} with $F(0)=0$ admits no nontrivial solution $u$.
\end{lemma}
Proof. Suppose there exists a solution $u=(u_1,\cdots, u_L)$.
By a calculation the same as the first step of the proof of Lemma \ref{signChangingDirichletNonEx} we get identities similar to \eqref{pohozaevSingle}, for $ i=1,\cdots,L$,
\begin{align*}
\int_B \Delta u_i (x\cdot \nabla u_i) dx &- \int_B \frac{n-2}{2}|\nabla u_i|^2 dx = \frac 1 2 \int_{\partial B} x\cdot \nu |\nabla u_i|^2 d\sigma>0.
\end{align*}
Sum the LHS of the above identities up and get
\begin{align*}
0 &< \int_B -\frac{\partial F}{\partial u_i} (x\cdot \nabla u_i) + \frac{n-2}{2} u_i \Delta u_i dx \\
&= \int_B -x\cdot \nabla F(u(x)) - \frac{n-2}{2} u_i \frac{\partial F}{\partial u_i} dx\\
&= \int_B n F- \frac{n-2}{2} u_i \frac{\partial F}{\partial u_i} dx,
\end{align*}
which contradicts to \eqref{potentialCondition}.
$\Box$
Therefore, it follows from Theorem \ref{existencetheorem} and Lemma \ref{PucciSerrin} that
\begin{theorem}\label{conservativeEx}
For system \eqref{systemPotential}, if $f$ satisfies the assumptions \eqref{decayAssumption} and \eqref{ControlInequality}, and additionally if $f=\nabla F$, where $F(0)=0$, and $F$ satisfies \eqref{potentialCondition} for $u\neq 0$, then system \eqref{systemPotential} admits a radially symmetric solution of class $C^{2}(\mathbb{R}^{n})$.
\end{theorem}
Here is an example system of {\bf{Type I}}:
Let the potential function be
\begin{align}
F(u,v)=-(u-v)^2+v^{p-1}u+u^{p-1}v, \text{with } p\geq \frac{2n}{n-2},
\end{align}
and then
\begin{align}\label{example3}
\left\{\begin{array}{ll}
-\Delta u = F_u = -2(u-v) + v^{p-1} + (p-1)u^{p-2}v, \\
-\Delta v = F_v = 2(u-v) + (p-1)v^{p-2}u + u^{p-1}, \\
\ \ u,v > 0, \\
\end{array}
\right. \text{in } \mathbb{R}^n.
\end{align}
We can verify that $F$ satisfies the condition of Theorem \ref{conservativeEx}. $F(0,0)=0$, and $F_u+F_v \geq 0$ (\eqref{decayAssumption} is satisfied). Let $u=0$ then $F_u = 2v+v^{p-1}\geq |F_v|=2v$, similarly if $v=0$, then $F_v =2u+u^{p-1}\geq |F_u|=2u$, so we can see that \eqref{ControlInequality} is satisfied in a neighborhood. Last, direct computation shows that
\begin{align*}
\frac{n-2}{2}(uF_u+vF_v)-nF &= \frac{n-2}{2}(-2(u-v)^2 + puv^{p-1} + pu^{p-1}v) -n(-(u-v)^2+v^{p-1}u+u^{p-1}v) \\
&\geq 2(u-v)^2 >0,
\end{align*}
so \eqref{potentialCondition} is satisfied. Also, notice that $F_u$ and $F_v$ are sign-changing functions, for example, let $v=0$, then $F_u<0$ and let $u=0$ then $F_u>0$.
{\bf{Type II.}}
In the following example, we give another class of systems with potential type of source term.
Consider the following system, in $\mathbb{R}^n$,
\begin{align}\label{systemPotential2}
\left\{\begin{array}{cl}
-\Delta u = f_1=\frac{\partial F}{\partial v}, \\
-\Delta v = f_2 = \frac{\partial F}{\partial u}, \\
u,v > 0 ,\\
\end{array}
\right.
\end{align}
and corresponding Dirichlet problem
\begin{align}\label{systemPotentialDirichlet2}
\left\{\begin{array}{cl}
-\Delta u = \frac{\partial F}{\partial v}, \text{in } B \\
-\Delta v = \frac{\partial F}{\partial u}, \text{in } B \\
u,v = 0 , \text{on } \partial B\\
\end{array}
\right.
\end{align}
where $B=B_R(0)\subset \mathbb{R}^n$ for any $R>0$.
Similarly we have
\begin{lemma}\label{PucciSerrin2}
If for $(u,v)\neq (0,0)$,
\begin{align}\label{potentialCondition2}
\frac{n-2}{2}\left( u \frac{\partial F}{\partial u} + v \frac{\partial F}{\partial v}\right) - nF(u,v) > 0,
\end{align}
then system \eqref{systemPotentialDirichlet2} with $F(0)=0$ admits no nontrivial solution $u$.
\end{lemma}
Then it follows from Theorem \ref{existencetheorem} and Lemma \ref{PucciSerrin2} that
\begin{theorem}\label{conservativeEx2}
For system \eqref{systemPotential2}, if $f_1=F_v$ and $f_2=F_u$ satisfies the assumptions \eqref{decayAssumption} and \eqref{ControlInequality}, and additionally if $F(0)=0$, and $F$ satisfies \eqref{potentialCondition2} for $(u,v)\neq (0,0)$, then system \eqref{systemPotential2} admits a radially symmetric solution of class $C^{2}(\mathbb{R}^{n})$.
\end{theorem}
{\bf{Proof of Lemma \ref{PucciSerrin2}.}}
Suppose there exists a solution $(u,v)$.
Then we follow the proof of Lemma \ref{signChangingDirichletNonEx}. The first step is the same, and we merge the source terms to \eqref{pohozaevCross} and get,
\begin{align*}
0 &<\int_B \Delta u(x\cdot \nabla v) + \Delta v(x\cdot\nabla u) - (n-2)\nabla u\cdot \nabla v dx \\
&= \int_B -F_v(x\cdot \nabla v) - F_u(x\cdot\nabla u)+ (n-2)(\theta v\Delta u + (1-\theta) u\Delta v)dx \\
&= \int_B -x\cdot \nabla F -(n-2)(\theta vF_v + (1-\theta) uF_u) dx\\
&= \int_B nF-(n-2)(\theta vF_v + (1-\theta) uF_u) dx,
\end{align*}
which contradicts to \eqref{potentialCondition2} when taking $\theta=\frac 1 2$.
$\Box$
Here is an example system of {\bf{Type II}}:
Let the potential function be
\begin{align}
F(u,v)=-(u-v)^2+u^p+v^p, \text{with } p\geq \frac{2n}{n-2},
\end{align}
and
\begin{align}
\left\{\begin{array}{ll}
-\Delta u = F_v = -2(u-v) + pv^{p-1}, \\
-\Delta v = F_u = 2(u-v) + pu^{p-1}, \\
\ \ u,v > 0, \\
\end{array}
\right. \text{in } \mathbb{R}^n.
\end{align}
We verify that $F$ satisfies conditions in Theorem \ref{conservativeEx2}. $F(0)=0$, and $F_u+F_v \geq 0$ (\eqref{decayAssumption} is satisfied). Let $u=0$ then $F_v = 2v+v^{p-1}\geq |F_u|=2v$, similarly if $v=0$, then $F_u =2u+u^{p-1}\geq |F_v|=2u$, so we can see that \eqref{ControlInequality} is satisfied in a neighborhood. Last, direct computation shows that
\begin{align*}
\frac{n-2}{2}(uF_u+vF_v)-nF &= \frac{n-2}{2}(-2(u-v)^2 + pu^p + pv^p) -n(-(u-v)^2+v^p+u^p) \\
&\geq 2(u-v)^2 >0,
\end{align*}
so \eqref{potentialCondition2} is satisfied. Also, notice that $F_u$ and $F_v$ are sign-changing functions, for example, let $v=0$, then there exists sufficiently small $u$ such that $F_u<0$, and let $u=0$ then there exists sufficiently small $v$ such that $F_v>0$.
|
\section{Numerical relativity}
\label{s:nr}
\subsection{A brief history}
Einstein's 1915 theory of general relativity has revolutionised the way
we think about gravitation.
Its radical difference from other field theories lies in the fact that its
equations govern the geometry of spacetime itself, as opposed to most other
theories where fields evolve on an unchanging background geometry.
The geometry of spacetime is determined by its matter content through
Einstein's field equations.
In turn, matter moves along geodesics of this spacetime manifold.
To put it simply, \emph{gravitation is geometry.}
Through observations such as the perihelion shift of Mercury,
the bending of light in the gravitational field of the sun,
the gravitational redshift, and the decrease of the orbital period of
binary pulsars consistent with the loss of energy due to emission of
gravitational radiation (Hulse \& Taylor, Nobel prize 1993), general relativity
is by now one of the most accurately verified physical theories.
Nevertheless, most of these observations only test the validity of the theory
in the weak-field limit.
Almost a century after Einstein's discovery, still relatively little is known
about the full implications of the theory in the nonlinear regime.
Aside from these astrophysical questions, there are several problems in
mathematical relativity that remain unanswered.
Two of the most important ones are the question of black hole stability and
the cosmic censorship conjecture.
Even though widely expected to be true, it was only in 1993 that Christodoulou
and Klainerman were able to prove in a voluminous work \cite{Christodoulou1993}
that flat (Minkowski) spacetime is nonlinearly stable.
Despite some recent progress, a similar theorem for the general stationary
vacuum black hole, the Kerr solution, is still lacking.
This is of central importance as black holes are believed to be ubiquitous
in the universe.
A different conjecture, first put forward by Penrose in 1969 \cite{Penrose1969}
and termed \emph{cosmic censorship}, concerns the global behaviour of solutions.
The Einstein equations are known to form singularities from quite general
initial data \cite{HawkingEllis}.
The (weak) cosmic censorship conjecture states that (very roughly) any
singularities formed from generic initial data lie inside an event horizon,
i.e.~they are causally disconnected from (invisible to) far-away observers.
So far there is no general proof of this conjecture, which has important
consequences on the determinism of the theory.
Why then do Einstein's equations pose such tremendous difficulties to the
mathematician?
Despite their elegant geometric origin, they turn out
to be a complicated system of coupled nonlinear second-order partial
differential equations (PDEs).
Exact solutions are generally only known under strong simplifying assumptions
such as the existence of spacetime symmetries.
Small perturbations of known solutions can be studied by linearising
the field equations.
One approach to studying the behaviour of more general solutions is the use
of numerical approximations.
Due to the complexity of the equations involved, this requires powerful
computers, and as a result \emph{numerical relativity} is a relatively
young field of research:
it started around 1964 with pioneering work by Hahn \&
Lindquist \cite{Hahn1964}, who studied the head-on collision of two black
holes.
Since then the field has had a history of several breakthroughs as well as long
periods of struggle. (Excellent recent textbooks on the subject are for example
\cite{Baumgarte2010,Alcubierre2008}.)
Arguably one of the most important achievements made through numerical
simulations is the discovery of critical phenomena in gravitational collapse
by Choptuik in 1993 \cite{Choptuik1993}.
This was triggered by a question posed by a mathematical relativist
(Christodoulou):
consider a family of initial data corresponding to compact matter configurations
with one parameter, such that for small values of the parameter the
configuration will disperse to leave flat spacetime behind, whereas
for large values it will collapse to form a black hole.
What happens at the threshold between the two outcomes?
Choptuik investigated this using sophisticated numerical methods
(most importantly, adaptive mesh refinement) and observed phenomena
similar to thermodynamic phase transitions, including power-law scaling
of the black hole mass in supercritical evolutions and a universal,
self-similar critical solution.
The majority of researchers in numerical relativity focused on what was
regarded as the most important outstanding problem in numerical relativity,
the collision of two orbiting black holes.
Black holes being the simplest objects in general relativity, this is the
obvious analogue of the two-body problem in Newtonian gravity.
The problem received so much attention because binary black hole collisions
are widely considered to be the strongest sources of gravitational waves,
which are hoped to be detected directly in the near future by several
earth-based detectors already in operation, a planned space-based detector
(eLISA) that has just been approved by the European Space Agency
to be launched in 2034, and alternative observational methods such as pulsar
timing arrays.
There is thus a strong need for models of gravitational waveforms from
astrophysical events to be used for matched filtering in gravitational
wave data analysis.
Despite much effort spent on the binary black hole problem,
it was not until 2005 that the final breakthrough was made and the first
complete simulations of the inspiral, merger and ringdown of a black hole
binary were presented almost simultaneously by three different groups
\cite{Pretorius2005a,Campanelli2006,Baker2006}.
By now such simulations have almost become routine.
Wider regions of the parameter space have been explored, matter has been
included (binary neutron stars or neutron star/black hole binaries) and more
complicated physics is being added.
These ``numerical laboratories'' serve as substitutes for experiments on
astronomical scales---an interesting philosophical shift of paradigm.
\subsection{The Cauchy problem for the Einstein equations}
In order to understand why the numerical solution of Einstein's equations poses
such difficulties, let us consider the general structure of these equations.
Spacetime is described by a smooth four-dimensional manifold $M$ with a smooth
Lorentzian metric $g_{ab}$.\footnote{Throughout we use abstract index notation,
whereby $g_{ab}$ represents the $0 \choose 2$ tensor field $g$ on $M$.
Indices $a,b, \ldots$ range over $0,1,2,3$.
The notation in this chapter has been streamlined to be self-consistent;
it differs from the notation used in some of the following chapters.}
The Einstein equations are
\begin{equation}
\label{e:einstein}
G_{ab} = \kappa T_{ab}.
\end{equation}
Here $G_{ab} = R_{ab} - \textstyle \frac{1}{2} R g_{ab}$ is the Einstein tensor, $R_{ab}$ is
the Ricci tensor and $R$ the scalar curvature.
These are evaluated with respect to the Levi-Civita connection compatible
with $g_{ab}$.
On the right-hand side, $T_{ab}$ is the energy-momentum tensor describing the
matter content of spacetime, and $\kappa$ is a constant.
For the time being we may assume vacuum, $T_{ab} = 0$.
Equation \eref{e:einstein} is to be solved for the metric $g_{ab}$;
it thus forms a system of second-order, quasi-linear PDEs.
A key property of \eref{e:einstein} is its invariance under arbitrary smooth
transformations of the spacetime coordinates $x^a$, a principle often referred
to as general covariance.
However, in order to solve the equations numerically, one needs to pick a
particular coordinate chart in order to obtain a definite set of PDEs.
This is most often done using the Cauchy or initial-value formulation of
general
relativity.\footnote{A different approach is the characteristic formulation;
see \cite{WinicourLRR} for a review and also section \ref{s:matching}
in this chapter.}
For this one picks a time coordinate $t := x^0$ and considers a foliation
of spacetime into the slices $\Sigma(t)$ of constant time $t$.
Indices $i,j,\ldots$ from the middle of the alphabet will be used to denote the
spatial coordinates $x^i$, $i=1,2,3$.
The Einstein equations \eref{e:einstein} split into two different classes.
The equations for which both indices are spatial ($ab=ij$) are found to
contain second time derivatives of the metric; these six equations are
called \emph{evolution equations}.
The equations for which one index is temporal (say $a=0$) are found to contain
no second time derivatives of the metric; these four equations are therefore
called \emph{constraint equations}.
The constraint equations are preserved under the time evolution in the sense
that if the constraints vanish at one instant of time then the evolution
equations imply that their time derivatives vanish as well.
This is a consequence of the contracted Bianchi identities
\[
\nabla^b G_{ab} = 0,
\]
where $\nabla$ denotes the covariant derivative compatible with $g_{ab}$.
While this is true on the analytical level, numerical simulations have long
been plagued by exponentially growing constraint violations.
Only relatively recently has this problem been cured (see below in
section \ref{s:gh}).
On an initial spacelike hypersurface $\Sigma_0$ corresponding to $t=0$,
we specify initial data for $g_{ab}$ and $\partial_t g_{ab}$ satisfying
the constraint equations.
(Constructing such data is itself a highly nontrivial problem, see
\cite{CookLRR} for a review.)
The evolution equations are then integrated forward in time in order to obtain
$g_{ab}$ for $t>0$.
There is a slight problem though: we have ten unknowns $g_{ab}$ but only
six evolution equations.
At this point general covariance comes into play: fixing the coordinates
allows us to impose four conditions on the components of $g_{ab}$, the
so-called coordinate or \emph{gauge} conditions.
Thus we really only have six free components of the metric that are evolved
by the six evolution equations.
In a numerical simulation it is difficult if not impossible to fix the
spacetime coordinates \emph{a priori} as one does not usually know what
spacetime a given set of initial data will evolve to.
Instead one ties the coordinates to the dynamical fields, hoping that the
coordinates that are thus being constructed ``on the fly'' will have desirable
properties (e.g., avoidance of singularities).
Depending on how this is done, the final set of PDEs one obtains may take on
very different forms.
In fact, the Cauchy problem may be well posed or ill posed!
In the following subsection we briefly review the two formulations of the
Einstein equations that are most often used in numerical relativity and,
in fact, in the present thesis.
\subsection{Formulations of the Einstein equations}
\label{s:formulations}
\subsubsection{Generalised harmonic coordinates}
\label{s:gh}
One way to fix the spacetime coordinates is to impose a wave equation
on each of the coordinates $x^a$:\footnote{Note the d'Alembert operator is
meant to act on each of the coordinates as scalar functions here.}
\[
\label{e:ghgauge}
\Box x^a \equiv g^{bc} \nabla_b \nabla_c (x^a) = -g^{bc} \Gamma^a{}_{bc} = H^a,
\]
where $\Gamma^a{}_{bc}$ denotes the Christoffel symbols of the Levi-Civita
connection.
Such coordinates are called \emph{(generalised) harmonic}.
The source functions $H^a$ on the right-hand side may depend on the
coordinates $x^a$ and on the metric $g_{ab}$ but not on derivatives of the
metric.
With this gauge condition the vacuum Einstein equations can be written as
\[
\label{e:gheinstein}
g^{cd} \partial_c \partial_d \, g_{ab} =
- \nabla_{a} H_{b} - \nabla_{b} H_{a} + 2 g^{cd} g^{ef} (\partial_e g_{ca}
\partial_f g_{db} - \Gamma_{ace} \Gamma_{bdf}),
\]
i.e. the principal part of the equation becomes the d'Alembert operator
associated with the metric.
Hence the system of PDEs is symmetric hyperbolic, a fact that was used
by Four\`es-Bruhat in her celebrated proof of the well-posedness of the Cauchy
problem for the Einstein equations \cite{FouresBruhat1952}.
Yet it was only much later that harmonic coordinates made their way into
numerical relativity.
Pretorius' 2005 breakthrough binary black hole
simulations \cite{Pretorius2005a} were based on this system.
A crucial ingredient was a new method to control the growth of constraint
violations.
In the generalised harmonic formulation, the role of the constraints is
taken on by the quantities
\[
\mathcal{C}^a := g^{bc} \Gamma^a{}_{bc} + H^a,
\]
which must vanish for a solution to the Einstein equations because of the
gauge condition \eref{e:ghgauge}.
The evolution equation \eref{e:gheinstein} implies the following evolution
equation for the constraints:
\[
\label{e:ghconstrevol}
\nabla^b \nabla_b \mathcal{C}_a + \mathcal{C}^b \nabla_{(a} \mathcal{C}_{b)}
= 0.
\]
A linear stability analysis of this equation shows that not all
modes decay, and they may be amplified due to the nonlinearity of the equation.
The key idea now is that we are still free to add multiples of the constraints
$\mathcal{C}_a$ to \eref{e:gheinstein} because these vanish for a solution
to Einstein's equations.
Such terms will not affect the principal part of \eref{e:gheinstein} because the
constraints contain only first derivatives of the metric.
Adding constraints to \eref{e:gheinstein} will modify the constraint evolution
equation \eref{e:ghconstrevol}.
In \cite{Gundlach2005} a particular combination of such \emph{constraint damping
terms} was devised such that on the linear level all non-constant modes of the
modified constraint evolution equation decay.
The generalised harmonic formulation of the Einstein equations forms the basis
of the first part of this thesis (chapters II--IV).
More precisely, we use a first-order reduction (with respect to time and
spatial derivatives) of \eref{e:gheinstein}
developed by the Caltech-Cornell numerical relativity collaboration.
Details of this reduction can be found in \cite{Lindblom2006} and in
\cite{Rinne2006}.
\subsubsection{ADM formulation}
\label{s:adm}
Before the introduction of generalised harmonic coordinates in numerical
relativity, most numerical work was based on the $3+1$ or \emph{ADM formulation}
of the Einstein equations originally developed by Arnowitt, Deser and Misner
in 1962 with a view towards quantising gravity (\cite{Arnowitt1962}; see
also \cite{York1979}).
In this framework one decomposes the vector field $\partial/\partial t$
associated with the time coordinate $t$ into a part normal to the hypersurface
$\Sigma(t)$ of constant $t$ and a part tangential to it:
\[
\left( \frac{\partial}{\partial t} \right)^a
= \beta^a + \alpha n^a,
\]
where $n^a$ denotes the unit timelike normal to $\Sigma(t)$, $\alpha$ is the
lapse function and $\beta^i$ the shift vector\footnote{Since $\beta^a$ is
tangential to $\Sigma(t)$, it has only three nonvanishing components,
hence we write it as $\beta^i$.}(figure \ref{f:adm}).
\begin{figure}
\centering
\input adm.pdf_t
\caption{\label{f:adm}
$3+1$ decomposition with unit timelike normal $n^a$, lapse function $\alpha$
and shift vector $\beta^i$.
}
\end{figure}
The spacetime metric takes the form
\[
g = -\alpha^2 \rmd t^2 + \gamma_{ij}(\rmd x^i + \beta^i \rmd t)(\rmd x^j
+ \beta^j \rmd t),
\]
where $\gamma_{ij}$ is the spatial metric (first fundamental form) induced on
$\Sigma(t)$.
We also need to introduce the extrinsic curvature (second fundamental form)
\[
\label{e:lngammaij}
K_{ij} = -\textstyle \frac{1}{2} \mathcal{L}_n \gamma_{ij},
\]
where $\mathcal{L}$ denotes the Lie derivative,
$\mathcal{L}_n = \alpha^{-1} ( \partial_t - \mathcal{L}_\beta )$.
Equation \eref{e:lngammaij} can be regarded as an evolution equation for
$\gamma_{ij}$.
The vacuum Einstein equations imply an evolution equation for $K_{ij}$,
\[
\label{e:lnkij}
\mathcal{L}_n K_{ij} = - \alpha^{-1} D_iD_j\alpha + \mathcal{R}_{ij}
- 2K_{ik}K^k{}_j + K_{ij} K,
\]
where $D$ denotes the covariant derivative compatible with
$\gamma_{ij}$, $\mathcal{R}_{ij}$ is the Ricci tensor of $\gamma_{ij}$,
and $K = \gamma^{ij} K_{ij}$.
The constraint equations take the form
\begin{eqnarray}
\label{e:hamcons}
\mathcal{H} &:=& \mathcal{R} + K^2 - K_{ij} K^{ij} = 0,\\
\label{e:momcons}
\mathcal{M}^j &:=& D_i (K^{ij} - \gamma^{ij} K) = 0,
\end{eqnarray}
where $\mathcal{R}$ is the scalar curvature of $\gamma_{ij}$.
It was only realised in the numerical relativity community in the 1990s that
for fixed lapse and shift, the ADM evolution equations \eref{e:lngammaij} and
\eref{e:lnkij} are only weakly hyperbolic and hence the initial value problem
is ill posed (see \cite{ReulaLRR} for a review of hyperbolicity for the
Einstein equations).
One way to cure this is to add multiples of the constraints, especially the
momentum constraint \eref{e:momcons}, to \eref{e:lnkij}.
This was the essential trick that led to the formulation of Baumgarte,
Shapiro, Shibata and Nakamura \emph{(BSSN)} \cite{Shibata1995,Baumgarte1998},
which in addition to the
generalised harmonic formulation has become one of the two standard
formulations used in binary black hole simulations.
A different approach, taken in the second part of this thesis, is the use
of elliptic gauge conditions.
As a condition on the spacetime slicing we shall require the mean curvature
$K$ of the slices to be a spacetime constant.
Apart from its geometric appeal, this will furnish the desired asymptotic
behaviour of the slices (see section \ref{s:outer_scri}).
Such slices also have good singularity avoidance properties as the mean
curvature controls the time evolution of the spatial volume element
$\sqrt{\det \gamma_{ij}}$.
Preservation of the constant mean curvature (CMC) condition under the time
evolution leads to an elliptic equation for the lapse function $\alpha$.
The spatial coordinates will be required to be \emph{spatially} harmonic,
i.e.,
\[
\label{e:shgauge}
\Delta x^i \equiv \gamma^{jk} D_j D_k (x^i) =
-\gamma^{jk} \; {}^{(3)} \Gamma^i{}_{jk} = H^i,
\]
where the $H^i$ are fixed functions of the spatial coordinates
(cf.~\eref{e:ghgauge}; now ${}^{(3)} \Gamma^i{}_{jk}$ refers to the
Christoffel symbols of $\gamma_{ij}$).
Taking a time derivative of \eref{e:shgauge} results in an elliptic equation
for the shift vector $\beta^i$.
It has been shown at least in the spatially compact case that the ADM
system with these elliptic gauge conditions (CMC slicing and spatially harmonic
gauge) has a well-posed initial value problem \cite{Andersson2003}.
The price to pay is that we need to solve elliptic equations at each time
step of the numerical evolution, which is generally more computationally
expensive than solving hyperbolic equations.
As mentioned earlier, due to general covariance, there is a redundancy in
Einstein's equations that allows one to solve only the evolution
equations\footnote{The constraints always need to be solved at the initial
time.}; the constraints will be preserved under the time evolution.
(Of course one still needs to check that violations of the
constraints remain small during a numerical evolution.)
This approach is referred to as \emph{free evolution}.
A different approach, which we shall adopt in the second part of this thesis,
is \emph{constrained evolution}, whereby the constraints \eref{e:hamcons}
and \eref{e:momcons} are solved explicitly in lieu of some of the evolution
equations.
This will give us better control of the asymptotic behaviour of the fields;
constrained evolution schemes are also often found to be more stable in highly
nonlinear gravitational collapse simulations.
Of course the constraints add to the number of elliptic equations
to be solved at each time step.
\subsection{Numerical methods}
Once we have decided on a particular formulation of the Einstein equations,
the question arises which numerical methods should be used to solve this
system of PDEs.
Here we briefly review the two methods that are most often used in
numerical relativity: pseudo-spectral methods and finite-difference methods.
These methods work well for smooth solutions, which is the case for the vacuum
Einstein equations and also for most radiative forms of matter (e.g., scalar,
electromagnetic or Yang-Mills fields).
For matter that may form discontinuities, e.g.~perfect fluids, these methods
are generally not suitable.
In this case finite-volume methods are normally used for the matter evolution
equations.
\subsubsection{Pseudo-spectral methods}
\label{s:ps}
The basic idea of spectral methods is an expansion of the numerical
approximation $u(x)$ in a known set of basis functions $u_n(x)$,
here in one dimension for simplicity:
\[
\label{e:spectral}
u(x) = \sum_{n=0}^N a_n u_n(x).
\]
The $u_n(x)$ usually belong to a complete orthonormal set of functions.
In the spherical topology that is most often encountered in numerical
relativity,
one usually expands in Chebyshev polynomials in the radial direction and
spherical harmonics in the angular directions.
Hereby the radial direction is often divided into a few subdomains and an
expansion of the form \eref{e:spectral} is used in each of the subdomains.
Derivatives of $u(x)$ can be computed \emph{exactly} within the approximation
\eref{e:spectral} using the known derivatives of the basis functions.
In order to compute nonlinear terms, \emph{pseudo-}spectral methods
evaluate the approximation $u(x)$ at a discrete set of collocation points
$x_i$, usually the Gauss- or Gauss-Lobatto points of the numerical quadrature
associated with the basis functions.
Nonlinear terms are evaluated at these collocation points and thereafter
the spectral expansion coefficients $a_n$ of the result are computed.
For smooth solutions, pseudo-spectral methods converge exponentially with the
number $N$ of expansion coefficients.
Hence $N$ is usually taken to be quite small, $N\lesssim 50$.
For larger $N$ roundoff errors quickly spoil any further gain in accuracy.
\subsubsection{Finite-difference methods}
\label{s:fd}
Finite-difference methods are based on an expansion of the solution in
a (finite) Taylor series.
Derivatives are replaced with difference quotients, e.g. for a one-dimensional
uniform grid with spacing $h$:
\[
(u')_i = \frac{1}{2h} (u_{i+1} - u_{i-1}) + \Or(h^2),
\]
where $u_i := u(x_i)$.
Near a boundary, one-sided operators are often used, e.g. for a right boundary
at $x=x_N$:
\[
(u')_N = \frac{1}{2h} (3 u_N - 4 u_{N-1} + u_{N-2}) + \Or(h^2).
\]
The above are examples of second-order accurate finite difference operators;
in the second part of this thesis we will work with fourth-order accurate
finite differences.
A subtle point is the treatment of coordinate singularities, e.g. for
axisymmetric spacetimes on the axis of symmetry $\rho = 0$ in cylindrical
polar coordinates $\rho, z, \phi$.
For this we use a staggered grid, where the first grid point is at $x_1 = h/2$,
and we add a \emph{ghost point} at $x_0 = -h/2$.
(One ghost point suffices for second-order accurate finite differences;
two are needed for fourth-order accuracy.)
The evolved fields are either even or odd with respect to $\rho$.
For an even function $u$ we set $u_0 = u_1$, whereas for an odd function we set
$u_0 = - u_1$.
This allows us to use centred finite difference operators at all interior
points $i\geqslant 1$.
\subsubsection{Multigrid for elliptic equations}
\label{s:mg}
For the constrained evolution schemes considered in the second part of this
thesis, elliptic equations need to be solved at each time step and hence
an efficient elliptic solver is needed.
The matrices arising from finite-difference approximations to elliptic
equations are sparse.
Standard relaxation method such as Gauss-Seidel relaxation are efficient
in damping short-wavelength components of the numerical error.
The slow convergence for longer wavelengths can be accelerated by using
a hierarchy of grids with increasingly coarser grid spacings, between which
the numerical approximation is transferred:
the multigrid method (\cite{Brandt1977}; an excellent concise introduction
is \cite{BriggsMG}).
We use the Full Approximation Storage variant of the algorithm in order
to treat nonlinearities in the equations directly, combined with a
nonlinear Gauss-Seidel relaxation.
\subsubsection{Time integration}
\label{s:mol}
A framework often used in numerical relativity is the method of lines:
the equations are first discretised in space and then regarded as a large
system of ordinary differential equations (ODEs) in time, one at each
grid/collocation point.
Standard ODE methods (e.g.~Runge-Kutta) can be used to integrate these ODEs
forward in time.
Some care must be taken in order to insure stability of the method,
in addition to the usual Courant-Friedrichs-Lewy condition on the timestep.
Finite-difference methods typically require artificial
Kreiss-Oliger \cite{Kreiss1989,Kreiss1995} dissipation for stability in the
context of the method of lines.
It is important to note though
that these extra terms are below the level of the truncation error.
Pseudo-spectral methods often suffer from aliasing arising from the pointwise
evaluation of nonlinear terms.
This can be cured by some form of spectral filtering \cite{Boyd2001}.
An example is Orszag's Two-Thirds rule, whereby the upper third of the
expansion coefficients is set to zero prior to evaluation of nonlinear terms.
\section{The outer boundary problem for isolated systems}
\label{s:outer}
A common task one faces in numerical relativity is the modelling of an
\emph{isolated system}, i.e.~a compact self-gravitating object, e.g.~a star,
surrounded by an asymptotically flat spacetime.
Here asymptotically flat means in a very loose sense that the spacetime metric
approaches the Minkowski metric in the limit of infinite distance from the
source.
It should be stressed that this picture is an idealisation: of course the
universe is full of compact objects, and whether the universe
is asymptotically flat is a matter of debate.
Nevertheless, if we are only interested in the dominant contribution of one
particular distant object to, say, the gravitational radiation observed on the
earth, then it is often a good approximation to surround this object by
an asymptotically flat vacuum spacetime and to consider ourselves to be at
infinite distance from the source.
The problem then arises to model an asymptotically flat spacetime of infinite
extent with finite computational resources, and this is the main subject of
this thesis.
\subsection{Conformal infinity}
\label{s:confinf}
In order to illustrate the various approaches to this problem, it is convenient
to adopt Penrose's idea of \emph{conformal compactification} \cite{Penrose1965}.
We write the spacetime metric as a conformal factor times a conformally
related metric:
\[
\label{e:confdecomp}
g_{ab} = \Omega^{-2} \tilde g_{ab}.
\]
Now we map the spacetime coordinates to a compact region such that $\Omega$
vanishes at the boundary, and $\tilde g_{ab}$ is everywhere finite when evaluated
in components with respect to the compactified coordinates.
As an example, consider Minkowski spacetime
\[
g = -\rmd t^2 + \rmd r^2 + r^2 \sigma,
\]
where $\sigma := \rmd \theta^2 + \sin^2\theta\, \rmd \phi^2$ is the round
metric on the unit sphere.
Performing the coordinate transformations
\[
\fl u=t-r, \quad v=t+r, \quad p = \arctan u, \quad q = \arctan v, \quad
T = p+q, \quad R=q-p,
\]
the metric can be written in the form \eref{e:confdecomp} with
\[
\label{e:confmink}
\Omega = 2 \cos p \, \cos q, \qquad
\tilde g = -\rmd T^2 + \rmd R^2 + (\sin^2 R) \, \sigma.
\]
Hence Minkowski spacetime is conformally related to the manifold
$\mathbb{R} \times S^3$ with standard metric.
However we obtain only part of this ``Einstein cylinder'':
the ranges of the compactified coordinates are
\[
\fl -\textstyle \frac{\pi}{2} < p \leqslant q < \textstyle \frac{\pi}{2}
\quad \Rightarrow \quad
-\pi < T < \pi, \quad 0 \leqslant R < \pi, \quad T+R < \pi, \quad T-R > -\pi.
\]
\begin{figure}
\centering
\input penrose.pdf_t
\caption{\label{f:penrose}
Penrose diagram of Minkowski spacetime.
}
\end{figure}
The resulting \emph{Penrose diagram} is shown in figure \ref{f:penrose}.
Since the mapping is conformal, light rays, i.e.~null geodesics,
propagate at $45$ degrees in the $T,R$ plane, just as they did
in the original $t,r$ coordinates.
An analysis of the asymptotic behaviour of geodesics leads to the following
results.
Future-directed timelike geodesics approach the point $(T,R)=(\pi,0)$,
which is therefore called \emph{future timelike infinity} $i^+$.
Similarly, past-directed timelike geodesics approach $(T,R)=(-\pi,0)$,
\emph{past timelike infinity} $i^-$.
Future-directed null geodesics approach the surface $T+R=\pi$,
\emph{future null infinity} $\mathrsfs{I}^+$ (``Scri$+$'').
Past-directed null geodesics approach $T-R=-\pi$, \emph{past null infinity}
$\mathrsfs{I}^-$.
Finally, spacelike geodesics approach $(T,R)=(0,\pi)$, \emph{spacelike
infinity} $i^0$.
Note that the conformal factor $\Omega$ in \eref{e:confmink} vanishes
at $\mathrsfs{I}^\pm$.
We refer the reader to \cite{FrauendienerLRR} for an in-depth discussion
of conformal infinity.
Similar Penrose diagrams can be drawn for other spacetimes.
New features can arise, e.g.~singularities and event horizons in black hole
spacetimes.
For our purposes at this point, we are mainly interested in the asymptotic
region, in particular spacelike infinity and null infinity,
which is common to all asymptotically flat spacetimes.
Hence Minkowski spacetime will serve us as a representative example of an
asymptotically flat spacetime.
\subsection{Initial-boundary evolution}
\label{s:outer_ibvp}
\begin{figure}
\centering
\input cauchy.pdf_t
\caption{\label{f:cauchy}
Cauchy evolution (blue lines) with artificial timelike boundary (red line).
Shown is the Penrose diagram of Minkowski spacetime with a
source (brown region) of radiation (yellow arrows).
}
\end{figure}
The standard method for numerical evolutions of asymptotically flat spacetimes
is to foliate spacetime by spacelike hypersurfaces all approaching spacelike
infinity, drawn in blue in figure \ref{f:cauchy}, with initial data specified
on an initial slice.
Consider a sequence of signals propagating at the speed of light,
symbolised by the diagonal yellow lines in figure \ref{f:cauchy}.
Since all spatial slices approach $i^0$, these signals can never leave
the slices.
Suppose we wanted to compactify the slices by mapping $i^0$ to a finite
spatial coordinate location.
Then an outgoing wave would appear increasingly ``blue-shifted'' (i.e.
with decreasing wavelength) with respect to the compactified coordinates,
and would ultimately fail to be resolved on the numerical grid.
Thus compactifying towards spacelike infinity is normally not a good idea.
(In \cite{Rinne2007} we assess the numerical performance of this approach,
among others.)
For these reasons one usually truncates the spatial slices at a finite
distance.
This introduces an artificial timelike boundary, the red line in
figure \ref{f:cauchy}.
Boundary conditions must be imposed there so as to obtain a well-posed
initial-boundary value problem.
These boundary condtions are not arbitrary because the constraint equations
must hold on each individual slice.
Furthermore, ideally one would like the solution on the truncated domain to be
identical with the solution on the unbounded domain.
Spurious reflections of gravitational radiation should be avoided.
Such boundary conditions are called \emph{transparent} or \emph{absorbing}.
The first part of this thesis will be devoted to the analysis and numerical
implementation of these questions, and will be summarised in section
\ref{s:ibvp} below.
For a comprehensive review article of this field of research
see \cite{SarbachLRR}.
There is a fundamental problem with this approach:
in general relativity, gravitational radiation is only well defined
at future null infinity $\mathrsfs{I}^+$.
This is the result of the seminal work by Bondi, Sachs and coworkers in a
series of papers from the 1960s \cite{Bondi1962}.
At a finite distance a ``local flux of gravitational radiation'' cannot be
defined in the full nonlinear theory.
This is only meaningful if one linearises about a given background spacetime,
e.g.~Minkowski or more generally, Schwarzschild or Kerr spacetime.
Any absorbing boundary conditions imposed at a finite distance can therefore
only be approximate.
\subsection{Cauchy-perturbative and Cauchy-characteristic matching}
\label{s:matching}
One approach is to match the fully nonlinear evolution in the interior
to an outer module that solves the \emph{linearised} Einstein equations.
Gauge-invariant treatments of gravitational perturbations exist that require
the solution of a scalar master equation, one for each pair $(\ell,m)$ with
respect to a spherical harmonic expansion of the gravitational field.
These scalars are functions of $t$ and $r$ only so it is relatively inexpensive
computationally to move the outer boundary to a very large distance.
Some more details of this method are discussed in section \ref{s:absorbing}.
It should be stressed that the linearised equations are still solved on
Cauchy slices approaching spacelike infinity $i^0$.
\begin{figure}
\centering
\input ccm.pdf_t
\caption{\label{f:ccm}
Cauchy-characteristic matching. An inner Cauchy foliation (blue) is matched
to an outer characteristic foliation (green).
}
\end{figure}
A different approach is to attach to the truncated spacelike foliation
a \emph{characteristic} foliation extending to future null infinity
$\mathrsfs{I}^+$.
This is represented by the green lines in figure \ref{f:ccm}.
The ``blue-shift problem'' mentioned above does not apply to these
null slices and hence it is straightforward to compactify them.
The difficult part of this method is the matching that needs to be done
at the artificial boundary.
So far this has been successfully implemented for \emph{a posteriori}
characteristic extraction, whereby one first carries out a Cauchy evolution
with boundary and then, in a post-processing step, reads out boundary data
for the subsequent characteristic evolution.
For this to work reliably, one needs to make sure that the artificial boundary
is placed sufficiently far out so that any inaccuracies emanating from it do
not reach the extraction surface, which is rather wasteful.
So far the ultimate task of doing the matching ``on the fly'' while the
Cauchy evolution is still running has not been fully accomplished.
We refer to \cite{WinicourLRR} for a review of the
Cauchy-characteristic matching approach.
The reader might wonder why one does not get rid of the spatial foliation
altogether and extend the characteristic slices all the way to the centre.
The reason is that null geodesic congruences, to which these slices are tied,
are generally ill behaved in strong-field regions:
they tend to form \emph{caustics}, which lead to coordinate singularities.
This caveat does not apply to situations with a high degree of symmetry,
e.g.~spherical symmetry, where characteristic evolution has indeed been
successfully used since the early days of numerical relativity.
\subsection{Hyperboloidal evolution}
\label{s:outer_scri}
Yet another approach, taken in the second part of this thesis, is to foliate
spacetime by \emph{hyperboloidal} surfaces (figure \ref{f:hyp}).
These are spacelike but approach future null infinity rather than spacelike
infinity.
An example are the standard hyperboloids in Minkowski spacetime,
\[
t = \sqrt{r^2 + \left( \frac{3}{K} \right)^2},
\]
where the constant $K$ turns out to be the mean curvature of the slices.
Such constant mean curvature surfaces can be constructed in more general
spacetimes, and will be used in the second part of this thesis.
However other choices of hyperboloidal surfaces are possible.
\begin{figure}
\centering
\input hyp.pdf_t
\caption{\label{f:hyp}
Hyperboloidal evolution.
}
\end{figure}
The hyperboloidal initial value problem consists in specifying initial data
on an initial hyperboloidal surface and evolving them to the future.
Note that hyperboloidal surfaces are only partial (future) Cauchy surfaces.
We will follow Penrose's idea and work with a conformally related metric
in a compactified coordinate system.
Unfortunately, the Einstein equations as such are not conformally invariant,
and as a result develop terms that are formally singular at $\mathrsfs{I}^+$.
Dealing with these terms is the main challenge in \cite{Moncrief2009}.
\section{Cauchy evolution with artificial timelike boundary}
\label{s:ibvp}
This section summarises my work on initial-boundary value problems
for the Einstein equations, represented by the three papers
\cite{Rinne2006}--\cite{Rinne2008b} in the first part of this thesis.
My interest in this topic arose during my time as a postdoc in the Caltech
group, who had just developed a first-order reduction \cite{Lindblom2006}
of the generalised harmonic formulation of the Einstein equations
(section \ref{s:gh}).
They had proposed on physical grounds a set of boundary conditions that seemed
to work well in numerical simulations, and they were now interested in
proving that these boundary conditions actually rendered the initial-boundary
value problem well posed.
\subsection{Well posedness \cite{Rinne2006}}
\label{s:wp}
The generalised harmonic formulation is convenient from a mathematical point
of view because it is essentially a system of nonlinear wave equations,
and the initial-boundary value problem for such equations is relatively well
understood.
However several complications arise in the Einstein case.
For simplicity, let us consider the scalar wave equation (with a source $F$),
\[
u_{tt} = u_{xx} + u_{yy} + u_{zz} + F
\]
on the half-space
\[
x \geqslant 0, \quad -\infty < y < \infty, \quad -\infty < z < \infty
\]
with boundary conditions
\[
\label{e:1storderbcs}
\alpha u_t = u_x + \beta_1 u_y + \beta_2 u_z + \alpha q \quad \mathrm{at}\;x=0,
\]
where $\alpha>0$ is a constant and $q$ are boundary data.
The initial data are
\[
u = f_1, \quad u_t = f_2 \quad \mathrm{at} \; t=0.
\]
One should think of $u$ as representing the individual components of the metric
in the generalised harmonic formulation of the Einstein equations.
For $\beta_1 = \beta_2 = 0$ the boundary conditions are maximally dissipative.
Defining the energy
\[
E := \norms{u_t} + \norms{u_x} + \norms{u_y} + \norms{u_z},
\]
it is straightforward to obtain an estimate of the form
\[
\label{e:strongwp}
\fl \int_0^t \norms{\mathbf{u}(s)} \rmd s
+ \int_0^t \normbs{\mathbf{u}(s)} \rmd s \
\; \leqslant \; K_T \left( \norms{\mathbf{f}} + \int_0^t \norms{F(s)} \rmd s
+ \int_0^t \normbs{q(s)} \rmd s \right)
\]
for every finite time interval $0\leqslant t \leqslant T$ with a constant
$K_T$ that is independent of $F$, $f_1$, $f_2$ and $q$.
Here $\norm{\cdot}$ and $\normb{\cdot}$ denote the
$L_2$ norms over the half-space and boundary, respectively, and
we have defined the vectors $\mathbf{u} := (u, u_t, u_x, u_y, u_z)$ and
$\mathbf{f} := (f_1, f_2, f_{1x}, f_{1y}, f_{1z})$.
The initial-boundary value problem is said to be \emph{strongly well posed}.
As already mentioned in section \ref{s:outer_ibvp}, boundary conditions for
Einstein's equations must be compatible with the constraint equations on the
$t=\mathrm{const}$ hypersurfaces.
The constraints satisfy a nonlinear wave equation of their
own \eref{e:ghconstrevol}.
The simplest constraint-preserving boundary condition one could imagine is
\[
\label{e:dirichletcpbc}
\mathcal{C}_a \overset{\wedge}{=} 0,
\]
where $\overset{\wedge}{=}$ denotes equality at the boundary.
This condition is of first order w.r.t.~derivatives of the metric, i.e.~of the
form \eref{e:1storderbcs}, but unfortunately with $\beta_1, \beta_2 \neq 0$,
i.e.~not maximally dissipative.
Later Kreiss and collaborators managed to prove strong well posedness
for a set of boundary conditions including \eref{e:dirichletcpbc}
using energy methods with a non-standard choice of energy
norm \cite{Kreiss2007}.
Still, the boundary conditions \eref{e:1storderbcs} are too restrictive in many
respects.
The constraint-preserving boundary conditions \eref{e:dirichletcpbc} are
a Dirichlet condition for a wave equation \eref{e:ghconstrevol}.
Consequently, any constraint violations generated in the interior will be
reflected off the boundary.
Better behaved boundary conditions can be obtained by requiring the incoming
characteristic fields of \eref{e:ghconstrevol} to vanish at the boundary
so that the constraint violations will leave the domain.
However, this will involve first derivatives of the $\mathcal{C}_a$ and hence
second derivatives of the metric.
More seriously, absorbing boundary conditions will also involve second
(or higher) derivatives of the metric.
This is because gravitational radiation is encoded in the Weyl tensor
$C_{abcd}$ (the tracefree part of the Riemann curvature tensor), which contains
second derivatives of the metric.
In \cite{Rinne2006} we use as a ``physical'' boundary condition the vanishing
of a particular projection of the Weyl tensor, the Newman-Penrose scalar
\[
\label{e:Psi0}
\Psi_0 = -C_{abcd} l^a m^b l^c m^d.
\]
Here the vectors on the right-hand side are part of a Newman-Penrose tetrad
$(l^a, k^a, m^a, \bar m^a)$,
where $l^a$ and $k^a$ are outgoing and ingoing real null vectors satisfying
$l^a k_a = -1$, $m^a$ is a complex spatial null vector orthogonal to
$l^a$ and $k^a$, and $\bar m^a$ is its complex conjugate, with
$m^a \bar m_a = 1$.
$\Psi_0$ can be regarded as an approximation to the incoming gravitational
radiation.
For boundary conditions of higher derivative order than \eref{e:1storderbcs},
the energy method can no longer be applied.
Instead, pseudo-differential techniques can be used.
For the time being we assume the source terms $F$ and initial data $f_1$, $f_2$
vanish.
The idea is to perform a Fourier-Laplace transform and write the solution as
a superposition of modes
\[
\label{e:fltransform}
u(t,x,y,z) = \tilde u(x) \exp [st + \rmi (\omega_y y + \omega_z z)]
\]
with $s\in \mathbb{C}$ and $\omega_y, \omega_z \in \mathbb{R}$.
Suppose the homogeneous problem with vanishing boundary data ($q=0$)
admits a solution with $\Re s > 0$.
Then we obtain another solution by multiplying the exponent in
\eref{e:fltransform} with any real number.
Hence the initial-boundary value problem cannot be well posed because the
growth of the solution cannot be controlled.
As reviewed in \cite{Rinne2006}, this condition amounts to showing that a
certain complex determinant does not have any zeros $s$ with $\Re s > 0$,
the \emph{determinant condition}.
What remains to be shown is that for the inhomogeneous problem, the solution
can be bounded in terms of the boundary data.
This turns out to be possible only if the zeros of the determinant have
strictly negative real part, the \emph{Kreiss condition}.
If it holds then one obtains an estimate
\[
\int_0^t \norms{\mathbf{u}(s)} \rmd s \leqslant K_T \int_0^t \normbs{q(s)}
\rmd s
\]
and the system is said to be \emph{boundary stable}.
The main result of \cite{Rinne2006} is that this condition holds for the given
first-order reduction of the generalised harmonic Einstein equations and the
given boundary conditions.
A stronger estimate that includes the source terms $F$ on the right-hand side
(cf.~\eref{e:strongwp}) is referred to as \emph{strong well posedness in the
generalised sense}.
In addition to boundary stability this requires the construction of a
symmetriser \cite{Kreiss2006}.
For the first-order reduction of the generalised harmonic Einstein equations
used in \cite{Rinne2006} it was not clear how to construct such a symmetriser.
What we do show though is that the Kreiss condition rules out
so-called weak instabilities with polynomial time dependence.
Later in \cite{Ruiz2007} strong well posedness in the generalised sense was
proved for the original second-order form of the equations, avoiding
complications arising from the first-order reduction.
From the theory of pseudo-differential operators it follows that strong well
posedness in the generalised sense carries over to systems with variable
coefficients and quasi-linear systems such as the Einstein equations.
Lacking a full proof of strong well posedness, we perform numerical
experiments in order to probe the stability of the system.
The numerical implementation uses pseudo-spectral methods as described in
section \ref{s:ps}.
The boundary conditions are implemented via a projection method,
which modifies the evolution equations at the boundary by eliminating
(derivatives of) the incoming fields using the boundary conditions.
We perform \emph{robust stability tests}, whereby small random noise is
injected in the initial data and source terms.
The background solution is taken to be either Minkowski spacetime on a spatial
domain with topology $T^2 \times \mathbb{R}$ or Schwarzschild spacetime on
$S^2 \times \mathbb{R}$.
These experiments show no signs of instabilities and strongly support the
claim that the system is well posed.
The expected instability for a deliberately chosen set of ill-posed
boundary conditions is also reproduced.
\subsection{Numerical comparisons \cite{Rinne2007}}
Having constructed a set of stable (and most likely well-posed) boundary
conditions for the Einstein equations in generalised harmonic gauge,
the question arises how well these boundary conditions perform numerically
compared to other choices.
Perfect boundary conditions would produce a solution on the truncated domain
that agrees with the solution on the unbounded domain restricted to the
truncated region.
We can use this principle in order to assess the boundary conditions in the
following way.
First we compute a \emph{reference solution} on a very large domain.
Because of the finite speed of propagation for hyperbolic PDEs, we can choose
the boundary to be sufficiently far out so that any inaccuracies emanating
from it remain out of causal contact with the interior region where comparisons
will be made.
Next we perform an evolution with the same initial data on a domain that is
truncated at a much smaller distance, where the boundary conditions are
imposed that are to be assessed.
Finally we compare the solution on the truncated domain with the reference
solution.
The test problem chosen in \cite{Rinne2007} is a Schwarzschild black hole with
an outgoing gravitational wave perturbation.
The background spacetime is written in Kerr-Schild coordinates,
which penetrate the event horizon at $r=2M$.
We can remove the interior of the black hole from the computational domain
by placing an \emph{excision} boundary just inside the event horizon.
At this interior boundary all characteristics leave the domain so that no
boundary conditions are required.
The gravitational wave perturbation is taken to be an exact solution
of the linearised (about flat space) Einstein equations with quadrupolar
($\ell=2$) angular dependence \cite{Teukolsky1982}.
The wave is taken to be outgoing initially, with a Gaussian profile.
Of course the constraints must be solved in order to obtain a valid set of
initial data for the Einstein equations.
The numerical implementation uses the same pseudo-spectral methods as
in \cite{Rinne2006} and as described above in section \ref{s:ps}.
Once the wave reaches the outer boundary, the imperfect boundary conditions
will generate reflections, which propagate into the interior.
In order to assess the amount of reflections, we evaluate the following
quantities.
\begin{enumerate}
\item The difference $\Delta \mathcal{U}$ between the test solution and the
reference solution of all components of the metric and their first
derivatives, in a suitable norm (see \cite{Rinne2007} for details).
It should be stressed that $\Delta \mathcal{U}$ is coordinate dependent so
it will measure how well the solutions agree
\emph{in the given coordinates}.
While ``gauge reflections'' have no physical meaning, they do matter from
a numerical point of view as one does not want to waste resolution
on short-wavelength features that merely correspond to a coordinate
transformation.
\item Violations of the constraints $\mathcal{C}$, again in a suitable norm.
This quantity tests how well the boundary conditions preserve the
constraints.
\item The difference between the test solution and the reference solution
of the outgoing gravitational radiation as measured by the Newman-Penrose
scalar (cf. \eref{e:Psi0})
\[
\Psi_4 = -C_{abcd} k^a \bar m^b k^c \bar m^d.
\]
This quantity is computed on a sphere close to the outer boundary
of the truncated domain (a procedure often referred to as
\emph{wave extraction}) and compared to the reference resolution.
From a physical point of view it is important to understand how the accuracy
of the extracted waveform is affected by the choice of boundary conditions.
\end{enumerate}
The benchmark set of boundary conditions used in \cite{Rinne2007} are the
boundary conditions constructed and analysed in \cite{Rinne2006},
with one small modification:
for the components of the metric that can be loosely identified as the
``gauge degrees of freedom'', a slightly different boundary condition is
used that differs from the original one only in a lower-order term.
With this extra term the gauge boundary condition is exactly absorbing
for a spherical ($\ell=0$) gauge wave.
This small modification is found to lead to a substantial reduction of the
coordinate-dependent difference $\Delta \mathcal{U}$, whereas the constraints
$\mathcal{C}$ and physical radiation $\Delta \Psi_4$ are of course unaffected.
In the following we summarise the various alternate boundary conditions that
are investigated in \cite{Rinne2007}, along with their numerical performance.
\begin{enumerate}
\item \emph{Freezing the incoming fields.}
In this approach the time derivatives of all the incoming fields
are required to vanish at the outer boundary.
While these boundary conditions render the initial-boundary problem
well posed, they are neither constraint preserving nor absorbing.
For increasing numerical resolution the quantity $\mathcal{C}$ is seen
to converge to a nonzero function.
This demonstrates that one does in fact not obtain a solution to
Einstein's equations with these simple-minded boundary conditions.
\item \emph{Sommerfeld boundary conditions.}
This type of condition is often used in numerical relativity simulations
based on the BSSN system and corresponds to imposing
\[
(\partial_r + \partial_r + r^{-1}) (g_{ab} - \eta_{ab}) \overset{\wedge}{=} 0
\]
on all components of the metric at the outer boundary, where $\eta_{ab}$
is the flat (Minkowski) metric.
The numerical performance is similar to the boundary conditions described
above, with slightly reduced constraint violations.
\item \emph{Kreiss-Winicour boundary conditions.}
These conditions, proposed in \cite{Kreiss2006}, consist in requiring the
harmonic constraints to vanish at the boundary,
equation \eref{e:dirichletcpbc} above.
We compute the remaining incoming characteristic fields from the
Schwarzschild background solution.
Although we expected this condition to be more reflective for constraint
violations, we do not find any indications for this numerically.
Apparently the constraint damping terms in our formulation are very
effective in reducing any constraint violations before they reach the
boundary.
However we do see larger errors in the physical quantities
$\Psi_4$ than with our benchmark boundary conditions, which include
the condition $\Psi_0 \overset{\wedge}{=} 0$.
\item \emph{Spatial compactification.} This approach is not technically
a boundary condition; instead we compactify the spatial domain towards
spatial infinity (see the discussion at the beginning of section
\ref{s:outer_ibvp} above).
A certain form of spectral filtering is applied in order to damp the
outgoing waves as they become increasingly ``blue-shifted''.
This turns out to work quite well as far as constraint violations are
concerned, however the errors in $\Psi_4$ are significantly larger than
with our benchmark boundary conditions.
\item \emph{Sponge layers.}
This method, often used in the context of spectral methods, adds
artificial damping terms to the evolution equations that are only active
in a region close to the outer boundary, schematically:
\[
\partial_t u = \ldots - \gamma(r) (u-u_0),
\]
where $u_0$ refers to the background solution and the function
$\gamma(r)$ is non-negligible only close to the outer boundary.
This method is found to lead to a small amount of constraint violations
and to considerable errors in the outgoing radiation $\Psi_4$.
\end{enumerate}
In summary, our boundary conditions outperform all the alternate methods
considered here.
We can even compare the reflection coefficient $\Psi_0 / \Psi_4$
with the prediction from linearised theory and
find good agreement with our simulations.
\subsection{Absorbing boundary conditions \cite{Rinne2008b}}
\label{s:absorbing}
The boundary conditions used in \cite{Rinne2006,Rinne2007} included a condition
on the vanishing of the Newman-Penrose scalar $\Psi_0$, which can be regarded
as an approximation to the outgoing gravitational radiation.
Using this condition was found to significantly reduce spurious reflections
of gravitational radiation.
It turns out that one can do better: there is a hierarchy of absorbing
boundary conditions of the form
\[
\label{e:B-S}
[r^2(\partial_t + \partial_r)]^{L-1} (r^5 \Psi_0) \overset{\wedge}{=} 0.
\]
Here $L$ refers to an expansion of the gravitational field in spherical
harmonics.
The boundary condition \eref{e:B-S} is perfectly absorbing for linearised
gravitational waves on a flat background spacetimes for all spherical harmonic
modes $\ell \leqslant L$.
For $L=1$ we recover our original condition $\Psi_0 \overset{\wedge}{=} 0$.
The boundary conditions \eref{e:B-S} were first suggested by Buchman and
Sarbach \cite{Buchman2006}.
They considered the linearised Bianchi equations, which describe the
propagation of gravitational radiation and in vacuum take the form
\[
\nabla^a C_{abcd} = 0,
\]
where $C_{abcd}$ is the Weyl tensor.
By expanding the fields in spherical harmonics and constructing exact solutions
to the linearised equations, the conditions \eref{e:B-S} were designed to
eliminate the ingoing solutions.
Later Buchman and Sarbach generalised their method to a Schwarzschild
background \cite{Buchman2007}.
In \cite{Rinne2008b} we reformulate the boundary conditions in a way that is
both conceptually more straightfoward and more amenable to numerical
implementation.
Gravitational perturbations can be described by the gauge-invariant
Regge-Wheeler-Zerilli (RWZ) scalars $\Phi_{\ell m}^{(\pm)}$ (see \cite{Sarbach2001}
and references therein).
These are complex quantities, one for each spherical harmonic index $(\ell,m)$
and for two parities: even ($+$) and odd ($-$).
On a flat background, they obey the master equation
\[
\label{e:RWZ}
\left[ \partial_t^2 - \partial_r^2 + \frac{\ell(\ell+1)}{r^2} \right]
\Phi_{\ell m}^{(\pm)} = 0.
\]
This equation is known as the Euler-Poisson-Darboux equation; it is of course
just the scalar wave equation in disguise.
The general outgoing and ingoing solutions have the form
\[
\Phi^{(\pm) \, \mathrm{out}}_{\ell m}(t,r)
= \sum_{j=0}^{\ell} \frac{f^{(\pm)}_{j\ell m}(t-r)}{r^j}, \qquad
\Phi^{(\pm) \, \mathrm{in}}_{\ell m}(t,r)
= \sum_{j=0}^{\ell} \frac{g^{(\pm)}_{j\ell m}(t+r)}{r^j}.
\]
The precise form of the functions $f^{(\pm)}_{j\ell m}$ and $g^{(\pm)}_{j\ell m}$
does not matter here.
The key observation is that
\[
\label{e:phibc}
B_L \Phi_{\ell m}^{(\pm)\, \mathrm{out}} :=
[r^2(\partial_t + \partial_r)]^{L+1} \Phi_{\ell m}^{(\pm)\, \mathrm{out}} = 0
\]
provided that $L \geqslant \ell$.
Using $B_L \Phi_{\ell m}^{(\pm)} \overset{\wedge}{=} 0$ as a boundary condition will therefore
eliminate the ingoing solutions for all $\ell \leqslant L$.
These are nothing but the well-known boundary conditions of Bayliss and
Turkel \cite{Bayliss1980} for the scalar wave equation.
It is straightforward to relate them to conditions on the Newman-Penrose
scalar $\Psi_0$ and recover \eref{e:B-S}.
Equation \eref{e:phibc} contains higher derivatives, which are difficult
to treat numerically.
In \cite{Rinne2008b} we address this by introducing a set of auxiliary variables so
that \eref{e:phibc} can be written as a system of ODEs intrinsic to the
boundary.
So far we have only considered the RWZ equation
\eref{e:RWZ}.
What we would really like is a set of boundary conditions for the Einstein
equations, in the generalised harmonic formulation already used in the
previous work.
Our algorithm thus consists in three steps:
\begin{enumerate}
\item extraction of the RWZ scalars from the spacetime metric at the
boundary,
\item evolution of the system of ODEs for the auxiliary variables
that implements the desired absorbing boundary condition,
\item construction of boundary data for certain incoming characteristic fields
of the Einstein equations from the auxiliary variables.
\end{enumerate}
In \cite{Rinne2008b} we describe each of these steps in detail.
Step (iii) can also be used as a recipe for Cauchy-perturbative matching
(section \ref{s:matching})
in the context of the generalised harmonic formulation of the Einstein
equations, as we could equally well take the boundary data from an outer
module that evolves the RWZ equations directly.
We also remark that strong well posedness in the generalised sense
(see section \ref{s:ibvp}) was proved in \cite{Ruiz2007} for the original
second-order form of the Einstein equations in harmonic gauge with the
new higher-order absorbing boundary conditions as well.
From the numerical point of view, an expansion of the fields in spherical
harmonics is required.
This fits well with our pseudo-spectral method, which already uses spherical
harmonics as the angular basis functions.
However some slightly intricate transformations between different
representations of tensor spherical harmonics need to be carried out
(see the appendix of \cite{Rinne2008b}).
In order to test our numerical implementation, we evolve initial data
corresponding to outgoing solutions of the linearised Einstein equations
with fixed spherical harmonic dependence $(\ell, m)$.
For $\ell=2$ these were derived in \cite{Teukolsky1982}.
In \cite{Rinne2008c} I constructed analogous solutions for arbitrary $\ell$.
We evolve these initial data on a truncated spherical domain using our new
absorbing boundary conditions.
During the evolution we extract the RWZ scalars at the boundary and compare
with the analytical solutions.
Since we evolve the full nonlinear Einstein equations, whereas the analytical
solutions are only valid to linear order, we perform evolutions with
different amplitudes of the initial data and check that any quantities
that should vanish at the linear level decay (at least) quadratically
with amplitude.
Using this method we show for our numerical evolutions in \cite{Rinne2008b} that our
boundary conditions $B_L$ are indeed perfectly absorbing for all
$\ell \leqslant L$.
While the boundary conditions do not eliminate incoming modes with $\ell > L$,
they reduce their amplitude significantly.
We compute the expected reflection coefficient analytically in linearised
theory and find good agreement with our numerical evolutions.
For instance, the $\ell=3$ incoming mode is suppressed by a factor of about
$100$ when the $L=2$ absorbing boundary condition is used as compared with
$L=1$, which corresponds to the old $\Psi_0 \overset{\wedge}{=} 0$
condition\footnote{Here we have taken the radius of the outer boundary to be
twice the wavelength.}.
This demonstrates the dramatic improvement achieved by these higher-order
absorbing boundary conditions.
\section{Hyperboloidal evolution to future null infinity}
\label{s:scri}
Much progress has been made with initial-boundary value problems for the
Einstein equations: well-posed formulations have been derived, particularly
in the context of generalised harmonic gauge, and improved absorbing boundary
conditions have been constructed and implemented.
The fundamental problem remains however that in the full nonlinear theory
of general relativity, boundary conditions imposed at a finite distance can
never be perfectly transparent in the sense that the solution on the truncated
domain agrees with the solution on the unbounded domain.
The absorbing boundary conditions considered in \cite{Rinne2008b}
rely on the validity of the linear approximation about a given background
spacetime (Minkowski in our case).
For this reason I became interested in hyperboloidal evolution, which aims to
place the outer boundary of the computational domain at future null infinity
$\mathrsfs{I}^+$, the only physically meaningful (conformal) boundary of spacetime.
This is the topic of the second part of this
thesis \cite{Moncrief2009}--\cite{Rinne2013}.
\subsection{Regularity at future null infinity \cite{Moncrief2009}}
Most approaches to hyperboloidal evolution are based on Penrose's idea of
a conformal transformation of the spacetime metric combined with a
compactifying coordinate transformation, as described in
section \ref{s:confinf} above.
Unfortunately the Ricci tensor is not conformally invariant and as a result
the Einstein equations contain inverse powers of the conformal factor,
which are singular at $\mathrsfs{I}^+$.
In the early 1980s Friedrich \cite{Friedrich1983a} developed an elegant
solution to this problem by constructing a symmetric hyperbolic system of
PDEs that contained the Einstein equations but also evolution equations for
the Weyl curvature.
Remarkably, his equations are regular everywhere, including at $\mathrsfs{I}^+$.
They are however rather complicated, which may explain why they have not made
their way into mainstream numerical relativity, despite a burst of activity
in the late 1990s (see \cite{Husa2003} for a review).
Recently \cite{Beyer2012}
there has been a renewed numerical interest in these equations,
especially concerning an extension \cite{Friedrich1998}
of Friedrich's original formulation that is able to address the intricate
issues that arise where null infinity meets spacelike infinity.
Here we follow a different approach, proposed by Moncrief, that aims to
tackle the (formally) singular terms in the Einstein equations directly.
We wanted to develop a system that is simpler
than Friedrich's regular conformal field equations and more similar to other
formulations already used by the numerical relativity community.
We work with an ADM-like formulation with elliptic gauge conditions:
constant mean curvature slicing and spatially harmonic coordinates, as
described in section \ref{s:adm} above.
In the spatially compact case the Cauchy problem for
these equations was shown to be well posed \cite{Andersson2003};
therefore we expect this formulation to be well behaved in our case as well,
although a formal proof of well posedness of the hyperboloidal initial value
problem with conformal boundary at $\mathrsfs{I}^+$ is still lacking.
As expected we find that both the constraints and the evolution equations
contain terms involving inverse powers of the conformal factor $\Omega$,
which become singular at $\mathrsfs{I}^+$.
This is not so much of a concern for the constraint equations,
as one can always multiply the entire equation by a suitably high power
of $\Omega$ before solving it,
but for the evolution equations it seems at first sight that the
right-hand sides are singular so that stable evolution near $\mathrsfs{I}^+$ cannot be
expected.
However in \cite{Moncrief2009} we show that the formally singular terms can actually
be evaluated explicitly at $\mathrsfs{I}^+$ in a completely regular way provided
the constraints hold.
On a given hyperboloidal slice, we expand all the fields in finite Taylor
series in $r$ near $\mathrsfs{I}^+$, in adapted coordinates so that $\mathrsfs{I}^+$ corresponds
to an $r=\mathrm{const}$ surface.
Thanks to the degeneracy of the elliptic constraint equations at $\mathrsfs{I}^+$,
we are able to evaluate the first few radial derivatives of the fields
at $\mathrsfs{I}^+$ by inserting the Taylor expansions into the constraints.
More precisely, we obtain the first three radial derivatives of $\Omega$,
the zeroth and first radial derivative of the components $\pi^{\tr \, ri}$ of the
ADM momentum (directly related to the tracefree part of the extrinsic
curvature),
and the first two radial derivatives of the conformal lapse
function $\tilde \alpha = \Omega\alpha$.
With this information we are able to evaluate the formally singular terms
in the evolution equation for $\pi^{\tr \, ij}$ and show they are regular
at $\mathrsfs{I}^+$, subject to one additional condition:
the vanishing of the shear of the null geodesic congruence that forms $\mathrsfs{I}^+$.
This condition had already been found in \cite{Andersson1992}.
In \cite{Moncrief2009} we show in addition that it is preserved under the time evolution
in the sense that if the shear vanishes at one instant of time, then its time
derivative vanishes as well.
It is important to note that we only use \emph{finite} Taylor series at $\mathrsfs{I}^+$.
We do not assume that the fields are smooth there.
The constraint equations give us just enough information about the first few
derivatives of the fields at $\mathrsfs{I}^+$ so that we can evaluate
the formally singular terms in the evolution equations.
In fact, it appears that in general, the Taylor expansion already breaks down
at the next order and a polylogarithmic term needs to be
included \cite{Bardeen2012}.
It could be that this is an artefact of CMC slicing.
Whether the polylogarithmic terms can be avoided in a different slicing
is an interesting open question.
There is a different, more straightforward way of deriving regular evolution
equations at $\mathrsfs{I}^+$ by assuming that the conformal Weyl tensor vanishes
at $\mathrsfs{I}^+$.
This \emph{Penrose regularity} implicitly assumes though that the conformal
metric is $C^3$ up to the boundary, a slightly stronger requirement than
what we needed for our original analysis.
This different regularisation technique is also explored in \cite{Moncrief2009}.
\subsection{Axisymmetric reduction and numerical implementation
\cite{Rinne2010}}
In this section we describe the first successful numerical implementation
of the formulation developed in \cite{Moncrief2009}.
Since our regularity analysis at $\mathrsfs{I}^+$ relied crucially on the satisfaction
of the constraint equations, we expect having to solve the constraints
explicitly at each timestep (constrained evolution).
This is computationally expensive and hence in this first application we
assume that spacetime is axisymmetric.
This reduces the number of effective spatial dimensions from three to two.
Unlike spherical symmetry, it is still compatible with gravitational radiation,
and we expect all the difficulties in the non-symmetric case already to be
present in axisymmetry as far as numerical stability at $\mathrsfs{I}^+$ is concerned.
I had already developed a constrained axisymmetric evolution
scheme on maximal Cauchy slices with timelike
boundary \cite{RinnePhD,Rinne2008a} and hence it was obvious to try and adapt
it to CMC slices with conformal boundary at $\mathrsfs{I}^+$.
Spherical polar coordinates $(t,r,\theta,\phi)$ are used so that the Killing
vector is $\partial/\partial\phi$, which in addition is assumed to be
hypersurface orthogonal.
The spatial gauge condition used here differs from the spatial harmonic gauge
of \cite{Moncrief2009}.
The conformal spatial metric $\tilde \gamma_{ij}$, which is related to the
physical spatial metric via $\gamma_{ij} = \Omega^{-2} \tilde \gamma_{ij}$,
is taken to have the form
\[
\label{e:quasiisotropic}
\tilde \gamma = \rme^{2\eta\sin\theta} (\rmd r^2 + r^2 \rmd \theta^2)
+ r^2 \sin^2\theta \, \rmd\phi^2.
\]
This is known as \emph{quasi-isotropic gauge} and is chosen here because it
reduces the degrees of freedom in the conformal spatial metric to just one
function $\eta(t,r,\theta)$.
Preservation of this gauge condition in time yields a system of elliptic
equations for the shift vector $\beta^i$.
In addition we need to solve the CMC slicing condition for the conformal lapse
$\tilde \alpha$ and
the Hamiltonian constraint for the conformal factor $\Omega$.
There are evolution equations for the function $\eta$ and for the three
components of the tracefree part of the extrinsic curvature.
Even though the spatial gauge condition is different, the regularity analysis
at $\mathrsfs{I}^+$ carries through as in \cite{Moncrief2009} and we obtain manifestly regular
forms of the evolution equations at $\mathrsfs{I}^+$.
We have experimented with two slightly different versions, one derived
directly from the constraint equations using Taylor expansions,
the other by assuming the somewhat stronger Penrose regularity mentioned
in the previous subsection.
Numerically both appear to work equally well.
The numerical implementation is based on the finite-difference technique
(section \ref{s:fd}) with fourth-order accurate finite-difference operators.
The outermost radial gridpoint is placed right at $\mathrsfs{I}^+$.
Here the regularised versions of the evolution equations are used, with
one-sided finite differences.
Already one further grid point in we have no choice but to use the full,
formally singular evolution equations.
Remarkably, this appears to be stable, provided the constraints are solved
at each substep of the Runge-Kutta time integration scheme.
We provide a heuristic explanation for the success of the method by observing
that the evolution equations contain terms that tend to push the solution
towards the values dictated by the regularity conditions.
Some care needs to be taken when solving the elliptic equations using multigrid.
Since the equations degenerate at $\mathrsfs{I}^+$, it is not surprising that a
straightforward pointwise Gauss-Seidel relaxation fails to converge.
Instead, we use a radial line relaxation (with a direct one-dimensional solver)
and then perform Gauss-Seidel iterations in the angular direction.
As a first test problem we consider a Schwarzschild black hole.
The metric on CMC slices is known in closed form; we just need to compactify the
radial coordinate, which requires the numerical solution of one ODE.
An inner excision boundary is placed just inside the event horizon.
We are able to evolve initial data taken from this metric for times
$t \sim 1000 M$ ($M$ being the black hole mass) without any signs of
instability and with the expected fourth-order convergence as the numerical
resolution is increased.
Next we include a gravitational wave perturbation by specifying free initial
data for the function $\eta$ in \eref{e:quasiisotropic}, which vanishes for the
unperturbed Schwarzschild spacetime.
We can read out the gravitational radiation at $\mathrsfs{I}^+$ by computing the
gauge-invariant Bondi news function \cite{Bondi1962}, which can be computed
directly from the conformal spacetime Ricci tensor,
\[
N = \bar m^a \bar m^b \tilde R_{ab},
\]
where the Newman-Penrose tetrad used must have the property that
$m^a$ is tangential to $\mathrsfs{I}^+$, i.e. $m^a \partial_a \Omega = 0$.
We observe the expected quasi-normal mode radiation generated by the perturbed
black hole (which essentially acts as a damped harmonic oscillator):
\[
N_\ell \propto \rme^{-\kappa_\ell t} \sin (\omega_l t + \phi_\ell),
\]
where $\ell$ refers to the index of an expansion in spherical harmonics.
For the small perturbation we use ($\sim 10^{-4}$),
the values of $\kappa_\ell$ and $\omega_\ell$ fitted from our numerical
evolution are in good agreement with the semi-analytic results from linear
perturbation theory.
At later times, when the quasi-normal mode radiation has decayed, one
expects a power-law \emph{tail} (often referred to as
\emph{Price's law} \cite{Price1972})
\[
N_\ell \propto t^{-p_\ell}.
\]
At the numerical resolutions that we are able to afford, we cannot see
this tail yet---as runs with two different resolutions demonstrate,
the solution has not converged yet.
The algorithm will need to be speeded up in order to study these
subtle phenomena.
\subsection{Including matter; numerical evolutions in spherical
symmetry \cite{Rinne2013}}
Resolving late-time power-law tails of gravitational and matter fields on
black hole backgrounds is a very demanding problem.
With the current axisymmetric implementation of our hyperboloidal evolution
scheme we were unable to provide sufficiently high resolution.
In order to test if our method is capable to study tails in principle,
we decided to take one step back and impose spherical symmetry.
Due to Birkhoff's theorem, spherically symmetric vacuum spacetimes are
necessarily static: they are isometric to the Schwarzschild solution.
Thus in order to have non-trivial dynamics in spherical symmetry, matter
needs to be included.
How to deal with matter in the context of hyperboloidal evolution based on
conformal compactification is an interesting problem in its own right,
and so we investigated this quite generally, without any spacetime symmetries
at first.
We need to impose the condition that the energy-momentum tensor be
tracefree,
\[
\label{e:Ttf}
g^{ab} T_{ab} = 0.
\]
Under this assumption the energy-momentum conservation equations, which
constitute the evolution equations for the matter fields, are conformally
invariant:
if we define a conformally related energy-momentum tensor
$\tilde T_{ab} := \Omega^{-2} T_{ab}$ then standard energy-momentum conservation
$g^{ab} \nabla_a T_{bc} = 0$ implies that
\[
\label{e:ConfEMC}
\tilde g^{ab} \tilde \nabla_a \tilde T_{bc} = 0,
\]
where $\tilde \nabla$ is the connection compatible with the conformal spacetime
metric $\tilde g_{ab}$.
Without the condition \eref{e:Ttf}, the equations \eref{e:ConfEMC} contain
an additional term that is singular at $\mathrsfs{I}^+$.
Condition \eref{e:Ttf} is generally satisfied for ``radiative'' forms of
matter such as a (conformally coupled) massless scalar field, Maxwell or
Yang-Mills fields.
It is not satisfied e.g.~for a general perfect fluid.
However if the support of the matter remains compact during the evolution
then one needs not to worry about the singular terms at $\mathrsfs{I}^+$.
We work out the matter evolution equations and matter source terms in the
Einstein equations explicitly for two examples:
a conformally coupled scalar field and Yang-Mills fields.
The Einstein-scalar field equations arise from varying the action
\[
S = \int ( \textstyle \frac{1}{2\kappa} R - \textstyle \frac{1}{2} g^{ab} \phi_{,a} \phi_{,b}
- \textstyle \frac{1}{12} R \phi^2 ) \sqrt{-g} \, \rmd^4 x.
\]
The last term is referred to as \emph{conformal coupling} and leads to a
conformally invariant evolution equation for the scalar field $\phi$:
\[
\Box \phi - \textstyle \frac{1}{6} R \phi = 0 \quad \Leftrightarrow \quad
\tilde \Box \tilde \phi - \textstyle \frac{1}{6} \tilde R \tilde \phi = 0,
\]
where $\Box$ is the d'Alembert operator, as above a tilde refers to the
conformal spacetime metric, and we have introduced a rescaled scalar field
$\tilde \phi := \Omega^{-1} \phi$.
Yang-Mills theory can be regarded as a nonlinear generalisation of
electromagnetism to non-abelian gauge groups.
Its fundamental field is a vector potential or connection $A_a^{(\alpha)}$.
The upper index refers to the gauge group, which we will take to be
SU(2), so Greek indices range over $1,2,3$ here.
The associated field strength tensor is
\[
\label{e:ymf}
F_{ab}^{(\alpha)} = \partial_a A_b^{(\alpha)} - \partial_b A_a^{(\alpha)}
+ f^{\alpha\beta\gamma} A_a^{(\beta)} A_b^{(\gamma)}.
\]
Note the last term, which is absent in electromagnetism.
The symbol $f^{\alpha\beta\gamma} = g \, [\alpha\beta\gamma]$ is totally
antisymmetric, where $[123] := +1$ and $g$ is a dimensionful coupling constant.
Repeated Greek indices are summed over.
The Yang-Mills field equations are given by
\[
\nabla_a F^{ab\,(\alpha)} + f^{\alpha\beta\gamma} A_a^{(\beta)} F^{ab\,(\gamma)} = 0.
\]
They have the convenient property to be conformally invariant and hence
we may adorn all quantities in the above equations with tildes
and work directly in the conformal spacetime.
When performing the $3+1$ decomposition, the Yang-Mills equations split into
a constraint and an evolution equation.
After this general discussion and examples of matter models we reduce the
equations to spherical symmetry.
Isotropic spatial coordinates are chosen so that the conformal spatial metric
is flat.
The tracefree part of the extrinsic curvature has only one free component
in this case.
Unlike in \cite{Rinne2010}, we solve the momentum constraint for it,
rather than its formally singular evolution equation.
(This is only possible in spherical symmetry.)
While the reduction to spherical symmetry is straightforward for the Einstein
and scalar field equations, it is not so obvious for the Yang-Mills fields.
The most general spherically symmetric (conformal) Yang-Mills connection has
the form
\[
\label{e:yma}
\tilde A^{i(\alpha)} = [\alpha ij] x^j F + (x^\alpha x^i - r^2 \delta^{\alpha i})H
+ \delta^{\alpha i} L, \quad \tilde A_0^{(\alpha)} = G x^\alpha,
\]
where $F,H,L$ and $G$ are functions of $t$ and $r$ only.
In most previous numerical studies only the potential $F$ was included;
we present for the first time evolutions with fully general spherically
symmetric Yang-Mills fields.
Our numerical method is very similar to the one of \cite{Rinne2010}.
Since there is only one spatial dimension now, the constraint equations
are ODEs, which we solve using a direct band-diagonal solver combined with an
outer Newton-Raphson iteration to address the nonlinearity.
The initial data are chosen to be either Minkowski or Schwarzschild spacetime
(in CMC slicing) with an approximately ingoing matter perturbation (scalar
field or Yang-Mills).
On the flat background we are able to take the amplitude to be large enough
so that a black hole forms during the evolution, and to continue the
evolution after excising its interior.
With the increased numerical resolution that is possible in the spherically
symmetric case, we can now see the tails and measure their decay exponents.
The results are in good agreement with previous numerical work.
This includes two studies that also used hyperboloidal
evolution \cite{Puerrer2005,Puerrer2009}, however in coordinates that
are not horizon-penetrating so that gravitational collapse could not be
studied.
A general property of power-law tails is that the decay at $\mathrsfs{I}^+$ is slower
than at a finite distance.
(It would be impossible to see this with a code based on Cauchy evolution
with artificial timelike boundary!)
This causes the solution to resemble a ``boundary layer'' at late times
and the runtime of the simulation at fixed resolution is limited (though
sufficient in our case to obtain reliable results).
One feature we find that does not seem to have been noticed before is that
in the Yang-Mills case, the electric field (a component of the field strength
tensor \eref{e:ymf}) has a slower decay rate at $\mathrsfs{I}^+$ ($\sim t^{-1}$) than
the connection ($\sim t^{-2}$).
Furthermore, for the general spherically symmetric connection \eref{e:yma}
we find some interesting gauge dynamics:
while all components of the energy-momentum tensor decay so that a vacuum
solution is approached, the components of the connection approach a constant
or even time-periodic solution in some cases.
We explain this behaviour by deriving the most general form of the
spherically symmetric vacuum solutions to the Einstein-Yang-Mills system.
\section{Conclusions and outlook}
\label{s:concl}
This thesis is concerned with analytical and numerical approaches to treating
the far field of asymptotically flat spacetimes satisfying the Einstein
equations.
We focus on two different approaches: Cauchy evolution with artificial
timelike boundary (part 1) and hyperboloidal evolution to future null infinity
(part 2).
In the first part, we prove a necessary condition (boundary stability)
for well posedness of the initial-boundary value problem for a first-order
reduction of the Einstein equations in generalised harmonic gauge with
constraint-preserving boundary conditions.
These include a condition on the Weyl tensor component
$\Psi_0$, which can be regarded as a first approximation to the incoming
gravitational radiation.
Numerical stability tests further demonstrate the robustness of the boundary
conditions.
Next we assess the numerical performance of various other boundary conditions
and alternate approaches such as compactification to spacelike
infinity or sponge layers by comparing the solution on the truncated domain
with a reference solution on a much larger domain.
In all cases our boundary conditions are found to be superior.
Finally we formulate and implement a hierarchy of higher-order absorbing
boundary conditions that improve on the original $\Psi_0 \overset{\wedge}{=} 0$ condition.
Our approach is based on the Regge-Wheeler-Zerilli scalars, and we show how
it can be interfaced with the generalised harmonic formulation of the Einstein
equations.
In the second part, we work with a constrained ADM-like formulation of the
Einstein equations on constant mean curvature slices extending to future
null infinity.
Upon a conformal transformation of the metric, the Einstein equations develop
terms that are formally singular at future null infinity $\mathrsfs{I}^+$.
However, we show explicitly how these terms can be evaluated at $\mathrsfs{I}^+$
in a completely regular way.
Based on this idea we present a first numerical implementation for
vacuum axisymmetric spacetimes.
Long-term stable evolutions of a gravitationally perturbed Schwarzschild black
hole are obtained and the Bondi news function, which describes the outgoing
gravitational radiation in a gauge-invariant way, is evaluated at $\mathrsfs{I}^+$.
Finally we extend our formulation to include matter with trace-free
energy-momentum tensor.
Scalar and Yang-Mills fields are coupled to the Einstein equations and
evolved numerically in spherical symmetry.
This includes spacetimes that form a black hole from regular initial data.
We study the power-law decay (``tail'') of the matter fields at late times,
both at $\mathrsfs{I}^+$ and at a finite distance.
There are a number of ways in which the research presented in this thesis can
be extended.
We discuss both parts separately.
Concerning Cauchy evolution with artifical boundary, it would of course be
nice to complete the proof of strong well posedness in the generalised sense
for the particular first-order reduction of the Einstein equations in
generalised harmonic gauge and boundary conditions we used.
However, given that there is already a proof for the original second-order
system and that the boundary conditions appear to be very robust numerically,
there is currently not so much interest in this question.
Our implementation of absorbing boundary conditions could be generalised by
allowing for a Schwarzschild rather than flat background spacetime.
In general however, the numerical relativity community seems to be quite
happy with their current codes and seem to be reluctant to invest much
effort in improved boundary conditions.
This may well change once gravitational wave astronomy has advanced to a
stage that even more accurate simulations are required.
Certainly from the current point of view, hyperboloidal evolution appears to
be a much cleaner solution to the outer boundary problem.
Our axisymmetric numerical implementation demonstrates that stable numerical
evolutions based on our approach can be achieved,
however the code will need to be speeded up in order to be useful in
practice, especially in the case without symmetries.
For instance, one could try to solve the constraints explicitly only in a
neighbourhood of $\mathrsfs{I}^+$ and use free evolution in the interior.
We also intend to generalise our formulation to more general gauge conditions,
as we do not believe the particular gauge we used (constant mean curvature
slicing and spatially harmonic coordinates) was essential for the regularity
analysis at $\mathrsfs{I}^+$.
Hyperboloidal evolution should have interesting applications whenever
global properties of spacetime are to be investigated.
An example is cosmic censorship, as one can now check whether null
geodesics manage to escape to future null infinity.
\section*{Acknowledgments}
I would like to thank my Habilitation committee at FU Berlin
(Ralf Kornhuber, Klaus Ecker, Konrad Polthier, Theodora Bourni, Florian
Litzinger) and the referees for their time and effort, and for making me
welcome at the Department.
Further thanks go to my colleagues at the Albert Einstein Institute and my
collaborators over the past years who contributed to this work.
Support through a Heisenberg Fellowship and grant RI 2246/2 of the German
Research Foundation (DFG) is gratefully acknowledged.
\section*{References}
\input{chap1.bbl}
\end{document}
|
\section{Introduction}
\label{sec:introduction}
Digital communication systems in wireless channels have been studied extensively (refer to \cite{SimonBook}, \cite{ProakisBook}, and references therein). Besides the well-investigated Rayleigh, Rician, and Nakagami-$m$ channels, the lognormal fading is widely recognized as an important channel model because it fits empirical fading measurements well in many transmission scenarios of practical interest \cite{Lotse1992}. For examples, the lognormal fading can characterize the shadowing effects in outdoor RF communications \cite{SimonBook} and describe indoor radio propagation environments. The lognormal fading is also suitable for describing ultra-wideband channels \cite{Molisch2003} and characterizing optical wireless communication (OWC) links in clear sky over several hundred meters \cite{EJLee2004}.
On the other hand, performance analysis of digital communication systems over the lognormal channels remains challenging and mathematically intractable. The error rate analysis of such systems typically involves numerical integration \cite{Zhu2002} or approximation \cite{Liu2003}. In \cite{R.1}, the authors studied the achievable finite diversity order of the lognormal channels in relatively low signal-to-noise ratio (SNR) regimes. However, their result is not valid when the SNR is asymptotically large since the diversity order for the lognormal channels is widely known to be $\infty$. Due to this fact, the asymptotic theory \cite{ZhengdaoWang2003} cannot be applied directly to the lognormal channels.
Binary phase-shift keying (BPSK) is an important benchmark modulation scheme whose bit-error rate (BER) performance over fading channels is well-known \cite{SimonBook}, \cite{ProakisBook}. For coherent demodulation, BPSK requires channel state information (CSI) at the receiver in order to achieve
its optimal performance. When the required CSI cannot be tracked accurately, binary differential
phase-shift keying (DPSK) is an attractive alternative with inferior BER performance. It is
well-known that DPSK is 3 dB worse than BPSK in Rayleigh channels in high SNR regimes \cite{Ekanayake1990}.
One fundamental question we aim to answer is what the performance loss of DPSK with respect to (w.r.t.) BPSK will be in the lognormal channel at high SNR.
In this letter, we prove that in the lognormal fading DPSK and BPSK have the same asymptotic BER performance, and differential quaternary phase-shift keying (DQPSK) is exactly $2.32$ dB worse than quaternary phase-shift keying (QPSK) in high SNR regimes.
\section{General Asymptotic Relative BER Analysis}
\label{sec-analysis}
For a digital communication system over fading channels, the SNR can be expressed as $\gamma = \overline{\gamma}I^2$ where $\overline{\gamma}$ is the average SNR\footnote{In RF communications, $\overline{\gamma}$ is the average SNR. In OWC with direct detection, $\overline{\gamma}$ is known as the electrical SNR \cite{Zhu2002}.} and $I$ is the channel gain. The asymptotic BER $P_b^{\infty}$ as $\overline{\gamma}\to\infty$ can be obtained by averaging the conditional BER $P_b(\gamma)$ over the probability density function (pdf) of $\gamma$, and it is approximated by \cite {ZhengdaoWang2003}
\begin{equation}
f(\gamma) = \frac {c \gamma^{t}}{\overline{\gamma}^{t+1}} + o\left({\gamma^{t}}\right)
\label{equ-GeneralPDFofSNR}
\end{equation}
where a function $f(x)$ is $o(g(x))$ if $\lim\limits_{x\to 0} f(x)/g(x) = 0$. Equation \eqref{equ-GeneralPDFofSNR} can be obtained from the Taylor series expansion of the SNR pdf where the parameters $c$ and $t$ are finite constants. For BPSK with conditional BER $P_{b,\mbox{\tiny BPSK}}(\gamma) = Q\left(\sqrt{2{\gamma}} \right)$ where $Q(x)=\int_{x}^\infty e^{-z^2/2}/\sqrt{2\pi}dz$ is the Gaussian $Q$-function, the asymptotic BER is \cite {ZhengdaoWang2003}
\begin{equation}
\label{equ-SERforCoherent}
P_{b,\mbox{\tiny BPSK}}^{\infty} = \frac{c \Gamma\left(t+\frac{3}{2}\right)}{2\sqrt{\pi}\left(t+1\right){\overline{\gamma}}^{t+1}} + o\left(\frac {1} {\overline{\gamma}^{t+1}}\right)
\end{equation}
where $\Gamma(\cdot)$ is the Gamma function. For DPSK with conditional BER $P_{b,\mbox{\tiny DPSK}}(\gamma) = 0.5\exp(-{\gamma})$, the asymptotic BER is \cite{Li2012}
\begin{equation}
\label{equ-SERforNonCoherent}
P_{b,\mbox{\tiny DPSK}}^{\infty} = \frac{c \Gamma\left(t+2\right)} {2(t+1){\overline{\gamma}}^{t+1}}
+ o\left(\frac {1} {\overline{\gamma}^{t+1}}\right).
\end{equation}
Using \eqref{equ-SERforCoherent} and \eqref{equ-SERforNonCoherent}, we derive the asymptotic SNR penalty factor of DPSK w.r.t. BPSK as
\begin{equation}
\label{SNR-BPSKDPSK}
{\mbox{SNR}}_{\mbox{\tiny DPSK-BPSK}}^{\infty} = \frac{10}{t+1}\log\left(\frac{\sqrt{\pi}\Gamma(t+2)}{\Gamma\left(t+\frac{3}{2}\right)}\right) \mbox{dB}
\end{equation}
where $\log(\cdot)$ is the log function with base $10$.
Since QPSK and BPSK have the same BER performance \cite{ProakisBook}, we have $P_{b,\mbox{\tiny QPSK}}^{\infty} = P_{b,\mbox{\tiny BPSK}}^{\infty}$. For Gray coded DQPSK, the conditional BER is given as \cite[Eq. (8.86)]{SimonBook}
\begin{equation}
\label{P_4DPSK}
P_{b,\mbox{\tiny{DQPSK}}}(\gamma) = F\left(\frac{\pi}{4}, \gamma\right) -F\left(\frac{5\pi}{4}, \gamma\right)
\end{equation}
where
\begin{equation}
\label{F_function}
F(\psi, \gamma) = \frac{\sin\psi}{2\pi}\int_{0}^{\pi/2}\frac{\exp\left(-2\gamma(1-\cos\psi\cos\theta)\right)}{1-\cos\psi\cos\theta} d\theta.
\end{equation}
Using \eqref{equ-GeneralPDFofSNR}, \eqref{P_4DPSK}, and \eqref{F_function}, we obtain the asymptotic BER as
\begin{equation}
\label{P_4DPSK_asy}
P_{b,\mbox{\tiny{DQPSK}}}^{\infty} = \frac{c\Gamma\left(t+2\right)\left[g\left(t,\frac{\pi}{4}\right)+g\left(t,\frac{5\pi}{4}\right)\right]}{2^{t+2}\sqrt{2}\pi(t+1)\overline{\gamma}^{t+1}} + o\left(\frac {1} {\overline{\gamma}^{t+1}}\right)
\end{equation}
where $g(t,\psi) \triangleq \int_0^{\pi/2}\frac{1}{(1-\cos\psi\cos\theta)^{t+2}}d\theta$. Using \eqref{equ-SERforCoherent} and \eqref{P_4DPSK_asy}, we derive the asymptotic SNR penalty factor of DQPSK w.r.t. QPSK as
\begin{equation}
\label{SNR-QPSKDQPSK}
{\mbox{SNR}}_{\mbox{\tiny DQPSK-QPSK}}^{\infty}\hspace{-0.06cm} = \hspace{-0.06cm}\frac{10}{t+1}\hspace{-0.06cm}\log\hspace{-0.06cm}\left(\hspace{-0.04cm}\frac{\Gamma(t+2)\left[g\left(t,\frac{\pi}{4}\right)+g\left(t,\frac{5\pi}{4}\right)\right]}{2^{t+1}\sqrt{2\pi}\Gamma\left(t+\frac{3}{2}\right)}\hspace{-0.04cm}\right) \mbox{dB}.
\end{equation}
Equations \eqref{SNR-BPSKDPSK} and \eqref{SNR-QPSKDQPSK} are general results for BPSK/DPSK and QPSK/DQPSK over fading channels when the receiver SNR pdf can be expressed as \eqref{equ-GeneralPDFofSNR}. From both equations, we observe that the SNR penalty factors are only functions of $t$.
\section{Asymptotic Relative BER Analysis for the Lognormal Fading Channels}
\label{sec:performanceanalysis}
\subsection{Single-Input Single-Output System}
\label{subsec:SISO}
In a lognormal fading environment, the channel gain $I = \exp(X)$ where $X$ is a Gaussian random variable (RV) with mean $\mu$ and variance $\sigma^2$. Thus, $I$ follows a lognormal distribution. To facilitate our
analysis, we normalize the mean of $I$ (i.e., $E[I] = \exp(\mu + \sigma^2/2) = 1$ and thus $\mu = -\sigma^2/2$) and obtain the pdf of $I$ as
\begin{equation}
\label{lognormalPDF}
f_{LN}(I) = \frac{1}{\sqrt{2\pi}\sigma I}\exp\left(-\frac{(\ln I+\sigma^2/2)^2}{2\sigma^2}\right).
\end{equation}
The parameter $\sigma$ is associated with the severity of fading. For OWC applications, the typical value of $\sigma$ is less than $0.5$ and it can be as low as $0.02$ \cite{Zhu2002}.
In order to evaluate the average BER performance, we require pdf of the receiver SNR. This pdf can be obtained by changing variables in \eqref{lognormalPDF} as
\begin{align}
\label{SNRPDF}
f_{LN}(\gamma)
&= \frac{1}{2\sqrt{2\pi}\sigma \gamma}\exp\left(-\frac{(\ln \gamma+\sigma^2-\ln\overline{\gamma})^2}{8\sigma^2}\right)
\end{align}
which is another lognormal pdf. Our goal is to study the asymptotic relative BER performance between coherent and differentially coherent modulation schemes over the lognormal channels. While it is infeasible to perform asymptotic analysis directly on the lognormal channels, we introduce the lognormal-Nakagami fading as an auxiliary channel model where the receiver SNR follows a lognormal-Gamma distribution. Unlike the lognormal pdf, the lognormal-Gamma pdf has a Taylor series expansion at the origin so that one can perform asymptotic analysis on such a channel. It can be shown that the diversity order of the lognormal-Nakagami channel is the parameter $m$. As $m$ approaches $\infty$, the lognormal-Gamma pdf approaches that of a lognormal RV. Therefore, we can study the asymptotic relative BER performance between coherent and differentially coherent modulation schemes over the lognormal-Nakagami channels and obtain the limiting results for the lognormal channels.
For the lognormal-Nakagami channels, the receiver SNR $\gamma$ follows a lognormal-Gamma distribution with pdf
\begin{equation}
\begin{split}
\label{SNRPDFGamma}
f_{LG}(\gamma) =& \int_0^{\infty}\frac{m^m\gamma^{m-1}}{\Omega^m\Gamma(m)}\exp\left(-\frac{m\gamma}{\Omega}\right) \\
&\hspace{-0.8cm}\times\frac{1}{2\sqrt{2\pi}\sigma \Omega}\exp\left(-\frac{(\ln \Omega+\sigma^2-\ln\overline{\gamma})^2}{8\sigma^2}\right)d\Omega
\end{split}
\end{equation}
where $m$ is the Nakagami-$m$ parameter and $\Omega$ is the second moment of a Nakagami-$m$ RV. When $m\to\infty$, we have
\begin{equation}
\label{limit}
\lim_{m\to\infty}\frac{m^m\gamma^{m-1}}{\Omega^m\Gamma(m)}\exp\left(-\frac{m\gamma}{\Omega}\right) = \delta\left(\frac{\gamma}{\Omega}-1\right)
\end{equation}
where $\delta(\cdot)$ is the Dirac delta function. Applying \eqref{limit} to \eqref{SNRPDFGamma}, we obtain
\begin{equation}
\lim_{m\to\infty}f_{LG}(\gamma) = \frac{1}{2\sqrt{2\pi}\sigma \gamma}\exp\left(-\frac{(\ln \gamma+\sigma^2-\ln\overline{\gamma})^2}{8\sigma^2}\right)
\end{equation}
which, as expected, is the lognormal pdf given in \eqref{SNRPDF}.
When $\gamma\to 0^+$, we can show that
\begin{equation}
\begin{split}
\label{SNRPDFGamma_re}
f_{LG}(\gamma) =& \frac{m^m{\gamma}^{m-1}}{2\sqrt{2\pi}\sigma\Gamma(m)}\int_0^{\infty}\frac{1}{\Omega^{m+1}} \\
&\hspace{-0.9cm}\times\exp\left(-\frac{(\ln \Omega+\sigma^2-\ln \overline{\gamma})^2}{8\sigma^2}\right)d\Omega + o\left({\gamma^{m-1}}\right).
\end{split}
\end{equation}
Comparing \eqref{SNRPDFGamma_re} with \eqref{equ-GeneralPDFofSNR}, we obtain $t=m-1$ so that the diversity order of the lognormal-Nakagami channels is $m$. Substituting $t=m-1$ into \eqref{SNR-BPSKDPSK}, we obtain the asymptotic relative BER performance between DPSK and BPSK over the lognormal-Nakagami channels as
\begin{equation}
\label{SNR_LG}
{\mbox{SNR}}_{\mbox{\tiny DPSK-BPSK}}^{\infty}(m) = \frac{10}{m}\log\left(\frac{\sqrt{\pi}\Gamma(m+1)}{\Gamma\left(m+\frac{1}{2}\right)}\right) \mbox{dB}.
\end{equation}
When $m\to\infty$, it follows that\footnote{Mathematically, the $0$ dB asymptotic BER performance loss of DPSK w.r.t. BPSK in \eqref{SNR_LG_infty} can also be obtained by letting $t\to\infty$ in \eqref{SNR-BPSKDPSK}. However, such manipulation is not permissible because the asymptotic analysis theory requires $t$ be finite.}
\begin{equation}
\label{SNR_LG_infty}
\lim_{m\to\infty}\hspace{-0.03cm}{\mbox{SNR}}_{\mbox{\tiny DPSK-BPSK}}^{\infty}(m)\hspace{-0.05cm} =\hspace{-0.09cm} \lim_{m\to\infty}\hspace{-0.04cm}\frac{10}{m}\log\hspace{-0.04cm}\left(\hspace{-0.03cm}\frac{\sqrt{\pi}\Gamma(m+1)}{\Gamma\left(m+\frac{1}{2}\right)}\hspace{-0.03cm}\right)\hspace{-0.04cm}=0\ \mbox{dB}
\end{equation}
which indicates that the BER of DPSK over the lognormal fading channels will approach that of BPSK in high SNR regimes. In another word, there is no substantial benefit to choose BPSK over DPSK when operating on the lognormal fading channels in asymptotically high SNR regimes.
Substituting $t=m-1$ into \eqref{SNR-QPSKDQPSK}, we obtain the asymptotic BER performance loss of DQPSK w.r.t. QPSK over the lognormal-Nakagami channels as
\begin{equation}
\label{SNR_QPSK_LG}
\begin{split}
{\mbox{SNR}}_{\mbox{\tiny DQPSK-QPSK}}^{\infty}(m) &=\left[ \frac{10}{m}\log\left(\frac{\Gamma(m+1)}{\sqrt{2\pi}\Gamma\left(m+\frac{1}{2}\right)}\right) -10\log(2)\right. \\
&\hspace{-1.2cm} \left.+\frac{10}{m}\log\left(g\left(m-1,\frac{\pi}{4}\right)+g\left(m-1,\frac{5\pi}{4}\right)\right)\right]\;\mbox{dB}.
\end{split}
\end{equation}
From the definition of $g(t,\psi)$, we have
\begin{equation}
g\left(m-1,{5\pi}/{4}\right)=\int_0^{\pi/2}{\left(1+{\sqrt{2}}\cos\theta/{2}\right)^{-(m+1)}}d\theta.
\end{equation}
It can be shown that $\lim\limits_{m\to\infty}g\left(m-1,{5\pi}/{4}\right) = 0$ when $\theta\in[0, \pi/2]$ so that the term $g\left(m-1,{5\pi}/{4}\right)$ in \eqref{SNR_QPSK_LG} is negligible when $m\to\infty$. Defining $h(\theta)\triangleq{1}/({1-{\sqrt{2}}\cos\theta/{2}})$, we have
\begin{equation}
\begin{split}
&\lim_{m\to\infty}\frac{10}{m}\log\left(g\left(m-1,\frac{\pi}{4}\right)\right) \\
&\hspace{0.8cm}= {10}\log\left(\lim_{m\to\infty}\left[\int_0^{\pi/2}{\left(h(\theta)\right)^{m+1}}d\theta\right]^{\frac{1}{m}}\right) \\
&\hspace{0.8cm}= {10}\log\left(\lim_{m\to\infty}\left[\int_0^{\pi/2}{\left(h(\theta)\right)^{m}}d\theta\right]^{\frac{1}{m}}\right) \\
&\hspace{0.8cm}=10\log\left(\left\| h \right\|_{\infty}\right)
\end{split}
\end{equation}
where $\left\| h \right\|_{\infty}$ is the uniform norm of the function $h(\theta),\,\theta\in[0, \pi/2]$. Since $h(\theta)$ is a continuous function on $[0, \pi/2]$, we have $\left\| h \right\|_{\infty}= \max \{h(\theta): \theta\in[0, \pi/2]\} = 2+\sqrt{2}$. It follows that the asymptotic BER performance loss of DQPSK w.r.t. QPSK over the lognormal channels is
\begin{equation}
\label{SNR_QPSK_LN}
\begin{split}
\lim_{m\to\infty}{\mbox{SNR}}_{\mbox{\tiny DQPSK-QPSK}}^{\infty}(m) &= 10\log\left(2+\sqrt{2}\right) - 10\log(2) \\
&= 2.32\; \mbox{dB}.
\end{split}
\end{equation}
\subsection{Selection Combining Diversity System}
\label{subsec:SC}
The asymptotic relative performance predictions in Section \ref{subsec:SISO} may not occur in practical SNR range.
However, diversity technique can facilitate the convergence of the relative BER performance in practical SNR range
to the asymptotic predictions. We will show such effect of diversity technique by studying an $L$-branch selection
combining (SC) system over independent and identically distributed (i.i.d.) lognormal channels. The pdf of
the SC output SNR $\gamma_{sc}$ is
\begin{equation}
\label{SNRPDF_SC}
\begin{split}
f_{LN}(\gamma_{sc}) =& Lf_{LN}(\gamma_{sc})\left[F_{LN}(\gamma_{sc})\right]^{L-1}
\end{split}
\end{equation}
where $F_{LN}(\gamma)$ is the cumulative distribution function of $\gamma$, and it can be obtained from \eqref{SNRPDF} as
\begin{equation}
\label{SNRCDF}
F_{LN}(\gamma) = 1-Q\left(\frac{\ln \gamma+\sigma^2-\ln\overline{\gamma}}{2\sigma}\right).
\end{equation}
Therefore, one can obtain the average BER of such a system by substituting \eqref{SNRPDF_SC}
and \eqref{SNRCDF} into $P_b = \int_0^\infty P_b(\gamma_{sc})f(\gamma_{sc})d\gamma_{sc}$.
Since the diversity order of an $L$-branch SC system is $L\varphi$ assuming $\varphi$ is the diversity order of a single-input single-output system \cite{Li2012}, the diversity parameter of such an $L$-branch system over the lognormal-Nakagami channels is $t=mL-1$. Substituting $t=mL-1$ into \eqref{SNR-BPSKDPSK} and letting $m\to\infty$, we obtain the asymptotic BER performance loss of DPSK w.r.t. BPSK over the lognormal channels as $0$ dB. Similarly, we obtain the asymptotic BER performance loss of a DQPSK system w.r.t. a QPSK system over the lognormal channels as $2.32$ dB by substituting $t=mL-1$ into \eqref{SNR-QPSKDQPSK} and letting $m\to\infty$.
\subsection{Multiple-Input Multiple-Output System}
The analysis in Section \ref{subsec:SC} can be extended to an $M\times N$ multiple-input multiple-output (MIMO) system because the maximum achievable diversity order of such a system is $MN\varphi$ \cite{Zheng2003TCOM}. It is straightforward to obtain that the diversity parameter $t=mMN-1$ for an $M\times N$ system over the lognormal-Nakagami channels. Therefore, using \eqref{SNR-BPSKDPSK} we obtain the asymptotic relative BER performance between an MIMO system using BPSK and that using DPSK as $0$ dB over the lognormal channels. The relative performance is still $2.32$ dB between an MIMO system using QPSK and that using DQPSK over the lognormal channels. However, MIMO technique will facilitate the convergence of the relative BER performance in practical SNR range to the asymptotic predictions.
\begin{figure}
\begin{center}
\includegraphics[width=0.95\linewidth, draft=false]{Fig1_BERBPSK.eps}
\caption{BERs of subcarrier intensity modulated OWC systems using BPSK and DPSK over the lognormal fading channels.}
\label{Fig_BERLognormal}
\end{center}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=0.95\linewidth, draft=false]{Fig2_BERQPSK.eps}
\caption{BERs of subcarrier intensity modulated OWC systems using QPSK and DQPSK over the lognormal fading channels.}
\label{Fig_BERQPSK_LN}
\end{center}
\end{figure}
\section{Numerical Results}
\label{sec:numerical-results}
In Fig. \ref{Fig_BERLognormal}, we present BER curves of subcarrier OWC systems using BPSK and DPSK
over the lognormal channels with different $\sigma$ values. A close examination of Fig. \ref{Fig_BERLognormal}
indicates that the BER performance loss of subcarrier DPSK w.r.t. subcarrier BPSK is $0.7$ dB
at the BER level of $10^{-10}$ when $\sigma=0.2$. When $\sigma=0.05$, this performance loss is
reduced to $0.2$ dB at the BER level of $10^{-30}$. When the SNR value is asymptotically large,
this performance loss will vanish as predicted by \eqref{SNR_LG_infty}. In Fig. \ref{Fig_BERLognormal} the exact BER curves are obtained via numerical integration, and they are validated by Monte Carlo simulations.
From the numerical results presented in Fig. \ref{Fig_BERLognormal}, we observe that the $0$ dB asymptotic prediction will not occur in practical SNR regimes. In order to investigate the performance loss in practical SNR range, through numerical investigation, we present in Table \ref{table1} the observed BER performance loss values at different BER levels in the lognormal channels. From Table \ref{table1}, we observe that the BER performance loss depends on the parameter $\sigma$ and it decreases in weaker lognormal channels at a given BER level. We also observe a clear trend that the performance loss decreases when the SNR increases and it will approach the $0$ dB prediction when the SNR value is asymptotically large. For example, the performance loss is as low as $0.63$ dB at SNR level around $14$ dB when $\sigma=0.1$, and it further decreases to $0.48$ dB at SNR level around $13$ dB when $\sigma=0.05$. We conclude that this BER performance loss is small in practical SNR range although it cannot achieve the $0$ dB asymptotic
prediction.
Figure \ref{Fig_BERQPSK_LN} presents BER curves of subcarrier OWC systems using QPSK and DQPSK over the lognormal channels with different $\sigma$ values. A close examination of Fig. \ref{Fig_BERQPSK_LN} indicates that the BER performance loss of subcarrier DQPSK w.r.t. subcarrier QPSK is $2.4$ dB at the BER level of $10^{-10}$ when $\sigma=0.2$ and this performance loss reduces to $2.3$ dB when $\sigma=0.05$. These numerical results agree with the asymptotic prediction of $2.32$ dB from \eqref{SNR_QPSK_LN}.
In order to demonstrate the effect of diversity technique on the relative performance in practical SNR range, through numerical investigation, we present in Table \ref{table2} the required SNR range to achieve a fixed performance gap of $0.5$ dB for different $L$ values in i.i.d. lognormal channels. From Table \ref{table2}, we observe that in order to achieve a fixed performance gap of $0.5$ dB for a given $\sigma$, the required SNR range depends on the number of branch $L$, and these SNR values will decrease when $L$ increases. When $\sigma =0.2$ and $L=4$, the $0.5$ dB performance gap can be achieved in a practical SNR range of $13.6-14.1$ dB. A clear trend can be observed from Table \ref{table2} that diversity technique can facilitate the convergence of the BERs of BPSK and DPSK at high SNR. When the SNR value is sufficiently large, this performance loss is expected to vanish for an $L$-branch SC system.
\begin{table}
\renewcommand{\arraystretch}{1.3}
\caption{Performance loss of DPSK with respect to BPSK in {\normalfont dB} at different BER levels in the lognormal channels
\label{table1}
\centering
\begin{tabular}{|c|c|c|c|c|c|c|}
\hline
\hline
\backslashbox{$\sigma$}{BER} & $10^{-2}$ & $10^{-4}$ & $10^{-6}$ & $10^{-8}$ & $10^{-10}$ \\
\hline
\hline
$0.5$ &$2.29$ & $1.65$ & $1.38 $ & $1.24$ & $1.10$ \\
$0.2$ &$1.84$ & $1.56$ & $0.89 $ & $0.79$ & $0.68$ \\
$0.1$ &$1.67$ & $0.98$ & $0.75 $ & $0.63$ & $0.54$ \\
$0.05$ &$1.61$ & $0.92 $ & $0.66 $ & $0.55$ & $0.48$ \\
\hline
\hline
\end{tabular}
\end{table}
\begin{table}
\renewcommand{\arraystretch}{1.3}
\caption{The required SNR ({\normalfont dB}) range to achieve a fixed performance gap of $0.5$ {\normalfont dB} for different $L$ values}
\label{table2}
\centering
\begin{tabular}{|c|c|c|c|c|c|}
\hline
\hline
\backslashbox{$\sigma$}{$L$} & $1$ & $3$ & $4$ & $5$ \\
\hline
\hline
$0.2$ &$27.5-28.0$ & $16.8-17.3$ & $13.6-14.1 $ & $12.8-13.3$ \\
$0.1$ &$14.6-15.1$ & $11.9-12.4$ & $11.7-12.2 $ & $11.5-12.0$ \\
\hline
\hline
\end{tabular}
\end{table}
\section{Conclusions}
\label{sect:conclusion}
Asymptotic analyses show that the performance advantage of BPSK over DPSK in the lognormal channels vanishes in asymptotically large SNR regimes. Under this operating condition, DPSK is a preferred modulation scheme in weak lognormal channels because it does not require CSI estimation and tracking. The performance advantage of QPSK over DQPSK in the lognormal channels is quantified to be $2.32$ dB in large SNR regimes. We also demonstrated that diversity technique can facilitate the convergence of the relative BER performance in practical SNR range to the asymptotic predictions.
\bibliographystyle{IEEEtran}
|
\subsection{Continuum limit(s)}
From the first gap equation (\ref{eq:saddle1}) we get the expansion
\begin{equation}
\Omega_0=I_0+\omega_1 a+\omega_2 a^2+\dots
\label{Omegaexpand}
\end{equation}
with
\begin{equation}
\omega_1=\frac{zx_1(z)}{\ell}.
\end{equation}
Let us denote the analogous expansion coefficients for the infinite volume
theory by $\tilde\omega_i$. The leading coefficient is:
\begin{equation}
{\rm SYM:\ }\tilde\omega_1=-\frac{m_R}{4\pi},\qquad\quad
{\rm BRO:\ }\tilde\omega_1=w_R.
\end{equation}
In addition, eq.~(\ref{eq:effective_coupling}) leads to
\begin{equation}
U^{\prime\prime}(I_0) =\left(\frac{u_{\rm eff}}{12} - U^{\prime\prime\prime}(I_0)\tilde\omega_1\right)a\ .
\end{equation}
From the second gap equation (\ref{eq:saddle2}) we see that
\begin{equation}
\frac{R}{2}=3+\frac{M^2}{2} -\left( U^\prime(\Omega_0) - \frac{R}{2}\right)
\label{3R}
\end{equation}
has to be volume-independent. Although we do not need the actual value
of couplings in the potential $U(S)$, their volume independence gives enough information
to calculate the mass gap in finite volume.
The volume independence leads to the equation,
\begin{eqnarray}
\frac{U^{\prime\prime\prime}(I_0)}{2}(2\tilde\omega_1\omega_1-\omega_1^2) -\frac{u_{\rm eff}}{12}\omega_1 +\frac{z^2}{2\ell^2}&=&
\frac{U^{\prime\prime\prime}(I_0)}{2}\tilde\omega_1^2 -\frac{u_{\rm eff}}{12}\tilde\omega_1 +\frac{m_R^2}{2},
\end{eqnarray}
where the right-hand side is volume-independent, so is the left-hand side.
It is interesting to see that $U^{\prime\prime\prime}(I_0)$ dependence appears in the equation for finite volume qualities, while such a $U^{\prime\prime\prime}(I_0)$ dependence shows up only in the 6-pt vertex in the infinite volume, which is $1/N$ suppressed compared to the 4-pt vertex, as already discussed before.
For simplicity of the analysis, we set $U^{\prime\prime\prime}(I_0)=0$ in the remaining of this appendix.
As before, the case I is obtained from the case II in the $u_{\rm eff}\to\infty$ limit.
We concentrate on the generic case II and we write (with $U^{\prime\prime\prime}(I_0)=0$)
\begin{equation}
\frac{zx_1(z)}{6\ell}=\frac{w_R}{6}-\frac{m_R}{24\pi}+\frac{1}{u_{\rm eff}}
\left(\frac{z^2}{\ell^2}-m_R^2\right).
\end{equation}
This formula is valid for both phases if we note that $w_R=0$ in the symmetric
phase and $m_R=0$ in the broken phase. We now list the equation determining the
finite volume mass gap in all cases.
\begin{itemize}
\item case IA
$L_m=-4\pi zx_1(z)$
\item case IB
$x_1(z)=0$
\item IIA
$z^2/L_u^2-zx_1(z)/(6L_u)=\alpha^2+\alpha/24\pi$ \ \ \ \ or\ \ \ \
$\alpha z^2/L_m^2-zx_1(z)/(6L_m)=\alpha+1/24\pi$
\item IIB
$L_u=6z/x_1(z)$
\item IIC
$zx_1(z)=\displaystyle L_w+\frac{\beta z^2}{8 L_w}$
\item IC
$zx_1(z)=L_w$
\end{itemize}
From these formulas we find the following qualitative behaviours.
\begin{itemize}
\item case IA
$z$ goes from $z_*$ to $\infty$ as $L_m$ goes from $0$ to $\infty$.
For small $L_m$, $z\simeq z_*$ so the model is UV conformal, as found
before. For large $L_m$, $z\approx L_m$.
\item case IB
$z=z_*$ constant, the model is conformal.
\item case IIA
$z$ moves from $0$ to $\infty$ as $L_u$ (or $L_m$) changes from
$0$ to $\infty$, the model is UV AF. For small $L_u$,
$z\approx(L_u/12)^{1/3}$ and for large $L_u$, $z\approx L_m=\alpha L_u$.
\item case IIB
$z$ moves from $0$ to $z_*$ as $L_u$ is changed from $0$ to $\infty$, it is
UV AF and IR conformal. For small $L_u$ it also behaves as
$z\approx(L_u/12)^{1/3}$.
\item case IIC
For $0<L_w <z_1\sqrt{\beta/8}$, $z(L_w)$ is monotonically
increasing from $0$ to $z_1$, where $z_1$ is defined by
$x_1(z_1)=\sqrt{\beta/2}$. For small $L_w$,
$z\approx(4L_w/\beta)^{1/3}$. This model is also UV AF.
For $L_w>z_1\sqrt{\beta/8}$, $z(L_w)$ is monotonically decreasing
and for large $L_w$, $z\approx1/(2 L_w)$.
\item case IC
$z(L_w)$ is monotonically decreasing, $z(0)=z_*$, the model is UV
conformal. For large $L_w$, $z\approx1/(2L_w)$.
\end{itemize}
\subsection{Finite volume coupling}
Now we can define the finite volume running coupling $g_{FV}$ by
\begin{equation}
g_{FV}=\frac{48(z/z_*)^3}{p+(1-p)(z/z_*)^2},
\end{equation}
where $p$ is some constant. It is normalized to $48$ for the conformal points
and has qualitatively the same behaviour as $g_4$ and also the corresponding
beta function
\begin{equation}
\beta_{FV}(g_{FV})=-L_u \frac{\partial}{\partial L_u}g_{FV}(L_u)
\end{equation}
shows walking behaviour for small $\alpha$.
Fig.~\ref{fig:FV} gives $\beta_{FV}(g_{FV})$ as a function of $g_{FV}$ at three values of $b$.
Like the beta function of the 4-pt coupling in the text, the "walking" behavior in the finite volume coupling becomes more visible, as $b$ decreases.
\begin{figure}[tbh]
\begin{center}
\scalebox{0.4}{\includegraphics{beta_FV.pdf}}
\end{center}
\caption{The beta function $\beta_{FV}(g_{FV})$ as a function of the finite volume coupling $g_{FV}$ at $b\equiv 96\alpha=0.119$(magenta line), 0.0119 (red line) and 0.00119 (blue line).
}
\label{fig:FV}
\end{figure}
\subsection{Continuum limits in the symmetric phase}
We first consider the SYM case. Defining the renormalized pion mass $m_R$ as $M=m_Ra$ ($m_R\ge 0$) with the lattice spacing $a$ and expanding $U^\prime$ around $I_0$ at $O(a^2)$, we obtain with eq.~(\ref{eq:effective_coupling})
\begin{eqnarray}
Z m_R^2 + \frac{u_{\rm eff}}{24\pi} m_R -r_2 &=& O(a),
\end{eqnarray}
where we define
\begin{eqnarray}
Z&\equiv&1 - \frac{U^{\prime\prime\prime}(I_0)}{16\pi^2} \le 1,
\end{eqnarray}
and make an additional tuning such that
$
2U^\prime(I_0) -6 = r_2 a^2
$ .
We here restrict ourselves to the case that the effective 6-pt coupling $U^{\prime\prime\prime}(I_0)$ is non-negative.
As long as $r_2 \ge 0$,
we obtain
\begin{equation}
m_R = \frac{\sqrt{u_{\rm eff}^2 + 4 (24\pi)^2Z r_2}-u_{\rm eff}}{48\pi Z},
\end{equation}
and $r_2=0$ corresponds to the massless theory, while
in the nonlinear $\sigma$ model limit we have
\begin{equation}
m_R = \frac{24\pi r_2}{u_{\rm eff}}
\label{eq:special}
\end{equation}
if $r_2 =O(u_{\rm eff})$.
Using $m_R$, the renormalized 2-pt function in the momentum space is given by
\begin{equation}
\gamma_\pi^{(2)}({\bf k}) =\frac{1}{{\bf k}^2+m_R^2} .
\end{equation}
Instead of the dimensional parameter $r_2$, we use the renormalize pion mass $m_R$, in addition to the dimensional coupling $u_{\rm eff}$, in order to specify the continuum limit. Then $r_2$ is expressed by others as
\begin{eqnarray}
r_2 &=& Z m_R^2\left( 1 + \frac{1}{24\pi Z \alpha }\right) \ge 0
\end{eqnarray}
in the symmetric phase,
where we introduce the ratio $\alpha = m_R /u_{\rm eff}$.
We call a massive theory ($m_R\not =0$) case A and a massless theory ($m_R=0$) case B, while the nonlinear $\sigma$ model limit is named as case I and otherwise as case II, so that there are four different continuum limits, IA, IB, IIA, IIB.
We next define the renormalized 4-pt vertex function in the symmetric phase as
\begin{eqnarray}
\gamma^{(4)}_\pi ({\bf k}) &=& \frac{\Gamma_\Pi^{(4)}({\bf K} = {\bf k} a)}{ a}\rightarrow
-\frac{1}{3}\frac{u_{\rm eff}} {\displaystyle 1+\frac{u_{\rm eff}}{24\pi\vert{\bf k}\vert} \arctan\left(\frac{\vert{\bf k}\vert}{2 m_R}\right)},
\end{eqnarray}
where we use the effective coupling defined in eq.~(\ref{eq:effective_coupling}).
We define a line of constant physics (LCP) by a curve on which $ \alpha \equiv m_R/u_{\rm eff} $ is kept constant even at finite lattice spacing. Taking $G=G_{k\ge 4}=0$ for simplicity, the LCP is determined by
\begin{equation}
\frac{R}{2}=3 +\frac{\alpha^2 U^2}{2} -\frac{U}{12} I _\infty ( \alpha U) .
\end{equation}
In Fig.~\ref{fig:LCP}, we draw this LCP at several values of $\alpha$ in the $(U, R)$ plane.
\begin{figure}[tbh]
\begin{center}
\scalebox{0.4}{\includegraphics{LCP_new.pdf}}
\end{center}
\caption{Lines of constant physics for $G=G_{k\ge 4}=0$. The dashed black line is the critical line. In the symmetric phase (above the critical line)
the red, green and blue curves are lines of constant $\alpha=0.05,0.1,0.2$ respectively. In the broken phase
the red, green and blue curves are lines of constant $r=0.5,0.25,0.1$
respectively.}
\label{fig:LCP}
\end{figure}
\subsection{Continuum limits in the broken phase}
In the broken phase, using the saddle point equation in the $H\rightarrow 0$ limit, we obtain
the renormalized condensate $\sigma_R^2 =\Sigma^2/a$ in the continuum limit as
\begin{eqnarray}
\sigma_R^2\equiv w_R&=& \left\{
\begin{array}{ll}
\displaystyle \frac{ u_{\rm eff}}{12 Z_B}\left(1-\sqrt{1+ r_2 Z_B \left(\frac{12}{u_{\rm eff}}\right)^2}\right), & Z_B\equiv U^{\prime\prime\prime}(I_0) > 0 \\
\displaystyle -6\frac{ r_2}{u_{\rm eff}} & Z_B =0
\\
\end{array}
\right. .
\end{eqnarray}
Therefore $r_2 $ must be negative in the broken phase.
We call the broken phase the case C, and there are IC and IIC depending on $u_{\rm eff}$.
The renormalized 4-pt vertex function in the continuum limit is given by
\begin{eqnarray}
\gamma^{(4)}_\pi ({\bf k} )&=& -\frac{1}{3} \frac{u_{\rm eff}}{\displaystyle 1+\frac{u_{\rm eff}}{48\vert{\bf k}\vert} +\frac{w_R u_{\rm eff}}{3 {\bf k}^2}} ,
\end{eqnarray}
which is related to
the $\pi\pi\sigma$ 3-pt vertex function in the continuum limit as
\begin{equation}
\gamma^{(3)}_{\pi\pi\sigma} ({\bf k} ) \equiv \frac{\Gamma^{(3)}_{\Pi\Pi\Sigma}({\bf K}={\bf k}a)}{a^{3/2}}
= \sqrt{w_R} \gamma^{(4)}_\pi ({\bf k}).
\end{equation}
Finally the continuum limit of the connected 2-pt function for $\sigma$ is given by
\begin{eqnarray}
\gamma^{(2)}_\sigma({\bf k}) &=& a^2 \Gamma^{(2)}_{\Sigma}({\bf K}={\bf k}a) =\frac{\displaystyle 1+\frac{u_{\rm eff}}{48 \vert {\bf k}\vert}}{\displaystyle {\bf k}^2 +\frac{u_{\rm eff}}{48}\vert {\bf k}\vert +\frac{w_R u_{\rm eff}}{3}} ,
\end{eqnarray}
which has a pole at
\begin{equation}
\vert {\bf k}\vert = \gamma ( -1 \pm \sqrt{1-64\beta} ), \quad \gamma=\frac{u_{\rm eff}}{96}, \quad \beta = 48\frac{w_R}{u_{\rm eff}} ,
\end{equation}
while the pion is massless in this phase.
If we interpret $\vert {\bf k}\vert$ as $\sqrt{-{\rm k}_0^2+{\rm k}_1^2+{\rm k}_2^2}$ and taking ${\rm k}_0= m_\sigma - i\Gamma_\sigma$, ${\rm k}_1={\rm k}_2=0$, we obtain
\begin{equation}
-{\rm k}_0^2 = -(m_\sigma^2-\Gamma_\sigma^2) + 2im_\sigma\Gamma_\sigma =
2 \gamma^2 \left[ 1 - 32\beta \mp\sqrt{1-64\beta}\right] .
\end{equation}
If $0 \le \beta < 1/64$, $-k_0^2$ is real and positive, so that there is no pole in the propagator.
On the other hand, if $1/64 \le \beta$, we have the $\sigma$ resonance, whose mass and width are given by
\begin{equation}
m_\sigma^2 = \gamma^2 (64\beta -1), \quad \Gamma_\sigma^2 = \gamma^2 .
\end{equation}
In Fig.~\ref{fig:LCP} we have also plotted lines of constant $r\equiv\Gamma_\sigma/m_\sigma$ (the ratio of
resonance width over mass) for the lattice regularization with $G=G_{k\ge 4}=0$.
For technical details we refer to Appendix~\ref{app_resonance}.
At $U\sim10$ on the lines for $r =0.5, 0.25, 0.1,$
the values of the lattice mass $M_\sigma$ are $\sim 0.20,0.41,1.0$. Lattice artifacts for the ratio $\Sigma^2/M_\sigma$ are rather
small along these lines (in the region plotted),
only deviating maximally a few percent from the continuum limit
$\omega_R/m_\sigma=(r+1/r)/32$.
\subsection{6-pt vertex function for $\pi$ fields}
We can also calculate the 6-pt vertex function for $\pi$ fields, by considering the third order of quantum fluctuation
in eq.~(\ref{eq:expansion}) as
\begin{equation}
U^{\prime\prime\prime}(\Omega_0)\sum_{\msvec n}\frac{\tilde\Omega_{\msvec n}^3}{3!}-\frac{4i}{3}
\sum_{\msvec{n}_1,\msvec{n}_2,\msvec{n}_3}G_{\msvec{n}_1{\msvec n}_2}G_{\msvec{n}_2{\msvec n}_3}
G_{\msvec{n}_3{\msvec n}_1}\tilde\Lambda_{{\msvec n}_1}\tilde\Lambda_{{\msvec n}_2}\tilde\Lambda_{{\msvec n}_3}
\end{equation}
at the leading order of the large $N$ expansion.
The connected 6-pt function for $\pi$ in the continuum limit is $O(1/N^2)$, and is given in the momentum space as
\begin{eqnarray}
\delta^{(3)}(\sum_{i=1}^6{\bf k}_i )\prod_{i=1}^6 \gamma_\pi^{(2)}({\bf k}_i)&\times&\frac{1}{N^2}
\left[\left\{
\gamma_\pi^{(6),{\rm 1PI}}({\bf k}_{12},{\bf k}_{34})\delta_{i_1i_2}\delta_{i_3i_4}\delta_{i_5i_6} + \mbox{14 perms}\right\} \right. \nonumber \\
&+& \left.\left\{ \gamma_\pi^{(6),{\rm 1PR} }({\bf k}_{12},{\bf k}_{123},{\bf k}_{56})\delta_{i_1i_2}\delta_{i_3i_4}\delta_{i_5i_6} + \mbox{89 perms} \right\} \right]
\end{eqnarray}
where ${\bf k}_{ij}={\bf k}_i+{\bf k}_j$, ${\bf k}_{ijk}={\bf k}_{ij}+{\bf k}_k$, and the 1-particle irreducible(1PI) 6-pt vertex function is given by
\begin{eqnarray}
\gamma_\pi^{(6),1PI}({\bf k},{\bf p} ) &=& \gamma_\pi^{(4)}({\bf k}) \gamma_\pi^{(4)}({\bf p}) \gamma_\pi^{(4)}({\bf k}+{\bf p})\left[\frac{U^{\prime\prime\prime}(I_0)}{(u_{\rm eff}/6)^3}+T({\bf k},-{\bf p})\right], \label{eq:1PI}\\
T({\bf k},-{\bf p}) &=& \int\frac{d^3 q}{(2\pi)^3}\gamma_\pi^{(2)}({\bf q}) \gamma_\pi^{(2)}({\bf q}+{\bf k})
\gamma_\pi^{(2)}({\bf q}-{\bf p})\nonumber \\
&=& \frac{1}{16\pi}\int_0^1dx \int_0^1dy \frac{ y}{\left[m_R^2 + f(x,y,{\bf k,p})\right]^{3/2}}
\end{eqnarray}
with $f(x,y,{\bf k,p})={\bf k}^2 xy(1-xy) +{\bf p}^2 y(1-y) +2 {\bf k}\cdot {\bf p}xy(1-y)$, while the 1-particle reducible (1PR) 6-pt vertex function becomes
\begin{eqnarray}
\gamma_\pi^{(6),1PR}({\bf k},{\bf p},{\bf q}) &=& \gamma_\pi^{(4)}({\bf k}) \gamma_\pi^{(2)}({\bf p})\gamma_\pi^{(4)}({\bf q}) .
\label{eq:1PR}
\end{eqnarray}
Note that the term in eq.~(\ref{eq:1PR}) and the second term in eq.~(\ref{eq:1PI}) are generated by the 4-pt vertex.
Since the theory is super-renormalizable without bare 6-pt coupling, however, they do not generate any UV divergences once $m_R$ (or $w_R$) and $u_{\rm eff}$ are made finite in the continuum limit. Therefore the effective 6-pt coupling $U^{\prime\prime\prime}(I_0)$ is not required to renormalize the theory, so that it can take an arbitrary non-negative value including zero.
The contribution from $U^{\prime\prime\prime\prime}(I_0)$ to the 8-pt vertex function, on the other hand, vanishes in the continuum limit, showing that it corresponds to the non-renormalizable coupling in the continuum theory.
\subsection{Beta function}
We take the usual definition
\begin{equation}
\beta_4(g_4)={\mathcal E}\frac{\partial}{\partial{\mathcal E}}g_4({\mathcal E}).
\end{equation}
Let us first consider the symmetric case. The beta function is always negative
here (except for the conformal case IB,
where it is identically vanishing). For IA we have
\begin{equation}
\beta_4(g_4)=-\frac{g_4^2}{48\pi}\sin(48\pi/g_4),
\label{betaIA}
\end{equation}
while for IIB
\begin{equation}
\beta_4(g_4)=-g_4+\frac{g_4^2}{48}.
\label{betaIIB}
\end{equation}
For the generic case IIA the beta function can be implicitly given by first
solving
\begin{equation}
\frac{2}{\pi}\arctan\xi+b\xi=\frac{48}{g_4}
\end{equation}
for the variable $\xi$, where $b=96\alpha$. The beta function is then
\begin{equation}
\beta_4(g_4)=-g_4+\frac{g_4^2}{24\pi}\left(\arctan\xi-\frac{\xi}{1+\xi^2}
\right).
\label{betaIIA}
\end{equation}
For small coupling we have the expansion
\begin{equation}
\beta_4(g_4)=-g_4+\frac{g_4^2}{48}-\frac{\alpha g_4^3}{6\pi}+\dots
\end{equation}
It is interesting to note that (\ref{betaIIB}) are the first two terms of
the weak coupling expansion of the beta function $\beta_4$ for any
$\alpha$ in case IIA.
For this generic case we can calculate the beta function numerically. It
depends on the parameter $b$. Fig.~\ref{fig:beta_cont} shows the beta function for
$b=96\alpha=$0.119, 0.0119 and 0.00119.
They nicely show the walking behaviour for small $\alpha$:
the beta function is close to (\ref{betaIIB}) for $0<g_4<48$,
and to (\ref{betaIA}) for $48<g_4<\infty$.
We note that it is necessary to go extremely close to the conformal
point $\alpha=0$ to be able to observe \lq\lq walking" of the
beta function.
\begin{figure}[tbh]
\begin{center}
\scalebox{0.38}{\includegraphics{beta_cont.pdf}}
\end{center}
\caption{The beta function in the symmetric phase for $b=$ 0.119 (black), 0.0119 (blue) and 0.00119 (red).
}
\label{fig:beta_cont}
\end{figure}
In the broken phase (in the generic case IIC) the beta function is
a double valued function of the coupling:
\begin{equation}
\beta_4(g_4)=\pm\frac{g_4}{48}\sqrt{(48-g_4)^2-64\beta g_4^2}
=\pm\left\{g_4-\frac{g_4^2}{48}-\frac{\beta g_4^3}{72}+\dots\right\}.
\end{equation}
The coupling has finite range:
\begin{equation}
0\leq g_4\leq\frac{48}{1+8\sqrt{\beta}}
\end{equation}
and the beta function is
\begin{equation}
{\rm positive\ for\ \ \ }0<{\mathcal E}<\frac{4w_R}{\sqrt{\beta}},\qquad
{\rm negative\ for\ \ \ }\frac{4w_R}{\sqrt{\beta}}<{\mathcal E}.
\end{equation}
For case IC the beta function is positive:
\begin{equation}
\beta_4(g_4)=g_4-\frac{g_4^2}{48}.
\label{betaIC}
\end{equation}
\subsection{Lattice artifact and finite size corrections to the "walking" behavior of the beta function}
In the case IIA, the running coupling in the continuum limit on the infinite volume is given by
\begin{eqnarray}
g_4({\cal E}) &=& \frac{1}{\displaystyle x+\frac{1}{24\pi}\arctan\left(\frac{x}{2 \alpha}\right)}
\end{eqnarray}
where $x= {{\cal E}}/{u_{\rm eff}}$.
In this subsection, we consider an effect of non-zero $a$ and finite volume to the beta function.
For simplicity of analysis, we set $G=G_{k\ge 4}=0$, and then introduce the lattice spacing through $U$ as $U = u_{\rm eff} a$. The running coupling thus becomes
\begin{eqnarray}
g_4^{\rm lat}({\bf K}) &=& \frac{1}{\displaystyle x+\frac{x}{6} U J({\bf K} )}
\end{eqnarray}
where
\begin{eqnarray}
U J({\bf K} ) &=& \frac{U}{L_0L_1L_2} \sum_{l_{0,1,2}=0}^{L_{0,1,2}-1}\frac{1}{\left[ \widehat{\bf Q}^2 +\alpha^2 U^2\right] \left[\left(\widehat{{\bf K}+{\bf Q}}\right)^2 + \alpha^2 U^2\right]}\nonumber \\
&=& \prod_{i=0}^2\frac{1}{ L_iU}
\sum_{l_{0,1,2}=0}^{L_{0,1,2}-1}
\frac{1}{\left[ \widehat{\bf q}^2 +\alpha^2\right] \left[ \left(\widehat{{\bf k}+{\bf q}}\right)^2 + \alpha^2\right]}, \\
{\bf Q} &=& 2\pi\left(\frac{l_0}{L_o},\frac{l_1}{L_1},\frac{l_2}{L_2}\right) ={\bf q} U, \quad
{\bf K} = 2\pi\left(\frac{n_0}{L_0},\frac{n_1}{L_1},\frac{n_2}{L_2}\right) ={\bf k} U
\end{eqnarray}
with ${\bf k}^2 = x^2$.
In Fig.~\ref{fig:beta_latt_FV}, we compare the beta function of $ g_4^{\rm lat}$ with that of $g_4$ as a function of $g_4$ in the symmetric phase at $b=0.0119$, where the "walking" behavior is clearly seen in the continuum limit (the black line). We take ${\bf K} = (K_0,0,0)$ with $ 0\le K_0 \le \pi$ in this calculation.
In the figure, the solid lines show the behavior of the beta function in the infinite volume at finite lattice spacing, corresponding to $m_R a =$ 0.005 (magenta), 0.05 (red), 0.1 (green) and 0.2 (blue), where $U=96 \, m_R a/ b$.
As $a$ increases, the lattice beta function deviates from its continuum one, in particular at small $g_4$, where the energy scale $x$ becomes large due to the asymptotic freedom.
We however still can observe the "walking behavior around $g_4\simeq 48$.
In the finite volume case, we take $L_0=L_1=L_2=L$ for simplicity.
Instead of the derivative, we use the symmetric difference of the discrete energy $x$ in the finite volume to define the lattice beta function.
In the figure, symbols represent the beta function at $L =$ 30 (diamonds), 40 (squares) and 80 (circles)
at $m_R a =$ 0.1 (green) while $L = $15 (diamonds), 20 (squares) and 40 (circles)
at $m_R a =$ 0.2 (blue), which give $m_R a L = $ 3,4 and 8 for both cases.
We observe that the finite size effect to the beta function is rather mild, except at strong coupling in the low energy region.
\begin{figure}[tbh]
\begin{center}
\scalebox{0.5}{\includegraphics{beta_latt_FV.pdf}}
\end{center}
\caption{The beta function in the symmetric phase at $b=0.0119$ for several values of lattice spacings and the volume. For further details please consult the text.
}
\label{fig:beta_latt_FV}
\end{figure}
\subsection{Large $N$ saddle point expansion}
Preparing the large $N$ expansion, we insert into the path integral
\begin{equation}
1\equiv\int\left[{\cal D}\Omega\right] \prod_{\msvec n}\delta\left(\Omega_{\msvec n}
-\frac{1}{N}\Phi^2_{\msvec n}\right)={\cal N}_1
\int\left[{\cal D}\Lambda\right]\left[{\cal D}\Omega\right]
\exp\left\{-N\sum_{\msvec n}i\Lambda_{\msvec n}\left(\frac{1}{N}
\Phi^2_{\msvec n}-\Omega_{\msvec n}\right)\right\},
\end{equation}
where ${\cal N}_1=\displaystyle \left(\frac{N}{2\pi} \right)^{V}$ with $V=L_0L_1L_2$, so that $\Phi$-integral becomes Gaussian and can be performed. We then obtain
\begin{equation}
Z(J)={\cal N}_2 \int\left[{\cal D}\Lambda\right]\left[{\cal D}\Omega\right]
\exp[-N S_{\rm eff}(\Lambda,\Omega,J)],\quad
{\cal N}_2 ={\cal N}_1 (2\pi)^{N V /2}
\label{master2}
\end{equation}
where the effective action is given by
\begin{equation}
S_{\rm eff}(\Lambda,\Omega, J)=\sum_{\msvec n}\left[ U(\Omega_{\msvec n}) -i\Lambda_{\msvec n}\Omega_{\msvec n}\right]-\frac{1}{2N}\sum_{{\msvec n}{\msvec m},i}
J^i_{\msvec n} \left(D^{-1}[\Lambda]\right)_{\msvec {nm}} J^i_{\msvec m} +\frac{1}{2}{\rm Tr} \ln D[\Lambda] ,
\label{Seff}
\end{equation}
and the matrix $D[\Lambda_{\msvec n}] $ here acts on an arbitrary vector $F$ as
\begin{eqnarray}
\left( D[\Lambda] F \right)_{\msvec n} &=& 2(i\Lambda_{\msvec n}-3) F_{\msvec n} -(\nabla^2 F)_{\msvec n}, \quad
(\nabla^2 F)_{\msvec n}\equiv \sum_\mu\left(F_{\msvec n +\hat\mu} + F_{\msvec n - \hat\mu} -2 F_{\msvec n}\right).
\label{Mmatrix}
\end{eqnarray}
The large $N$ limit corresponds to the saddle point, determined by the saddle point equations,
\begin{eqnarray}
\frac{\partial S_{\rm eff}(\Lambda,\Omega, J_0)}{i\partial \Lambda_{\msvec n}} &=&
-\Omega_{\msvec n} + \left(D^{-1}[\Lambda] \right)_{\msvec n \msvec n} + H^2 \left(\sum_{\msvec m} \left(D^{-1}[\Lambda] \right)_{\msvec {mn}}\right)^2 = 0 ,
\label{eq:lambda}\\
\frac{\partial S_{\rm eff}(\Lambda,\Omega, J_0)}{\partial \Omega_{\msvec n}} &=& U^\prime(\Omega_{\msvec n} ) -i\Lambda_{\msvec n} = 0.
\label{eq:rho}
\end{eqnarray}
To obtain a solution to the second equation, we have to take $i \Lambda_{\msvec n}$ real at the saddle point.
Assuming a translation invariant solution, we thus take
\begin{equation}
i \Lambda_{\msvec n}=i\Lambda_0\equiv 3+\frac{M^2}{2}= {\rm const.} \quad (M\ge 0), \qquad\quad \Omega_{\msvec n}=\Omega_0={\rm const.} ,
\label{spoint}
\end{equation}
with which eqs. (\ref{eq:lambda}) and (\ref{eq:rho}) become
\begin{eqnarray}
\frac{H^2}{M^4} + I(M)&=& \Omega_0, \label{eq:saddle1} \\
3+\frac{M^2}{2} &=& U^\prime(\Omega_0),
\label{eq:saddle2}
\end{eqnarray}
where
\begin{equation}
I(M) = \frac{1}{V}\sum_{\bf K} \frac{1}{\hat{\bf K}^2 + M^2},
\end{equation}
with
\begin{equation}
\hat{\bf K}^2 =\sum_\nu \hat{\rm K}_\nu^2,\quad \hat{\rm K}_\nu = 2\sin\frac{{\rm K}_\nu}{2} , \quad {\bf K}=2\pi\left(\frac{l_0}{L_0},\frac{l_1}{L_1},\frac{l_2}{L_2} \right) ,
\end{equation}
and $l_\mu$'s are integers which satisfy $0\le l_\mu\le L_\mu-1$.
Later we will solve these equations explicitly in detail.
As is well known \cite{MoZi,DaKeNe, DaKeNe2, KeNe}, it is possible to systematically organize
the large $N$ expansion as perturbation theory around the saddle point, by introducing the quantum fluctuation
as
\begin{eqnarray}
\Lambda_{\msvec n} &=& \Lambda_0 +\tilde \Lambda_{\msvec n} , \quad
\Omega_{\msvec n} = \Omega_0 + \tilde \Omega_{\msvec n},
\end{eqnarray}
together with $J_{\msvec n}^i = J_0^i +\tilde J_{\msvec n}^i$.
In this paper we mainly consider up to 4-pt functions in the leading large $N$ expansion
and for this purpose it is sufficient to consider Gaussian fluctuations of fields. (A brief discussion on the 6-pt function will be presented later.)
Using the saddle point equations (\ref{eq:lambda}) and (\ref{eq:rho}),
we expand the effective action as
\begin{eqnarray}
S_{\rm eff}(\Lambda,\Omega, J) &=& S_{\rm eff}(\Lambda_0,\Omega_0, J)
+ \sum_{\msvec n}\left[T_{\msvec n}(\tilde J) i\tilde \Lambda_{\msvec n} +\frac{1}{2}U^{\prime\prime}(\Omega_0) \tilde \Omega_{\msvec n}^2 -i\tilde\Lambda_{\msvec n}\tilde \Omega_{\msvec n}\right]\nonumber \\
&+& \sum_{\msvec {nm}}\frac{1}{2} S_{\msvec {nm}}(J) \tilde \Lambda_{\msvec n} \tilde \Lambda_{\msvec m}+\dots,
\label{eq:expansion}
\end{eqnarray}
where
\begin{eqnarray}
T_{\msvec n}(\tilde J) &=& \frac{1}{N}\sum_i \left[\left( G\, \tilde J^i\right)_{\msvec n} \right]^2
+\frac{2 H}{\sqrt{N} M^2} \left(G\, \tilde J^N\right)_{\msvec n}, \\
S_{\msvec {nm}}(J) &=& 2 G_{\msvec{mn}}^2+\frac{4}{N} \sum_i \left( G\, J^i \right)_{\msvec n} G_{\msvec{nm}} \left( G\, J^i \right)_{\msvec m}
\end{eqnarray}
with
\begin{eqnarray}
G_{\msvec{nm}} &\equiv& (D^{-1}[\Lambda_0])_{\msvec{nm}}= \frac{1}{V}\sum_{\bf{K}}
G({\bf K}) {\rm e}^{i{\bf K} (\msvec{n}-\msvec{m}) } , \quad
G({\bf K}) = \frac{1} {\hat{\bf K}^2 +M^2} .
\end{eqnarray}
We then perform the Gaussian integral for $\tilde \Omega_{\msvec n}$ and $\tilde\Lambda_{\msvec n}$, and finally obtain
\begin{eqnarray}
Z(J) &\simeq& {\cal N}_3 \exp[ - N W(J) ]
\label{eq:master}
\end{eqnarray}
where
\begin{eqnarray}
W(J) &=& S_{\rm eff}(\Lambda_0,\Omega_0,J) +\frac{1}{2}\sum_{\msvec{nm}} T_{\msvec n}(\tilde J) K^{-1}_{\msvec {nm}}(J)T_{\msvec m}(\tilde J) +\frac{1}{2N}{\rm Tr}\, \ln K(J) ,
\end{eqnarray}
with
\begin{eqnarray}
K_{\msvec{nm}}(J) &=& S_{\msvec{nm}}(J) +\frac{\delta_{\msvec{nm}}}{U^{\prime\prime}(\Omega_0)}, \\
S_{\rm eff}(\Lambda_0,\Omega_0,J) &=& V\left[ U(\Omega_0) - U^\prime(\Omega_0)\Omega_0 \right] -\frac{1}{2}{\rm Tr} \ln G -\frac{1}{2N} \sum_{\msvec{nm},i} J_{\msvec n}^i G_{\msvec{nm}} J_{\msvec m}^i ,
\\
{\cal N}_3 &=& {\cal N}_2\left(\frac{2\pi}{N\sqrt{U^{\prime\prime}(\Omega_0)}}\right)^{V} =\left(\frac{(2\pi)^N}{U^{\prime\prime}(\Omega_0)}\right)^{V/2} .
\end{eqnarray}
\subsection{VEV and Correlation functions}
We will call the first $N-1$ components of $\Phi$ "pions'' and
the $N^{\rm th}$ component "sigma", and denote $\Pi_{\msvec n}^i = \Phi^i_{\msvec n}$ for $i=1,2,\cdots, N-1$ and $\Sigma_{\msvec n} =\Phi^N_{\msvec n}$, respectively.
Taking derivatives of $W(J)$ in eq. (\ref{eq:master}) with respect to the source $J$ and setting $ J\rightarrow J_0$,
we can calculate the leading large $N$ contributions to the connected part of arbitrary $n$-point functions.
We start with the vacuum expectation value (VEV) of $\Sigma_{\msvec n}$ as
\begin{eqnarray}
\sqrt{N}\Sigma &\equiv& \langle \Sigma_{\msvec n}\rangle=\left. \frac{-N\partial W(J)}{\partial J^N_{\msvec n}}\right\vert_{J\rightarrow J_0} = \sqrt{N} H \sum_{\msvec{m}}G_{\msvec{nm}} -\frac{1}{2}{\rm Tr}\, K^{-1}(J_0)\left. \frac{\partial K(J)}{\partial J_{\msvec n}}\right\vert_{J\rightarrow J_0}, \nonumber \\
&=& \sqrt{N}\frac{H}{M^2}\left[1 -\frac{4}{N} \sum_{\msvec {lm}} K^{-1}_{\msvec{lm}}(J_0) G_{\msvec{ml}}G_{\msvec{mn}}\right]
= \sqrt{N}\frac{H}{M^2}\left[1 + O\left(\frac{1}{N}\right)\right] ,
\end{eqnarray}
where
\begin{eqnarray}
K_{\msvec{n m}}(J_0)&=&\frac{1}{V}\sum_{\bf K} {\rm e}^{i{\bf K}(\msvec{n}-\msvec{m})}\mathcal{K}({\bf K}),\\
\mathcal{K}({\bf K})&=&\frac{1}{U^{\prime\prime}(\Omega_0)}+2J({\bf K})+\frac{4H^2}{M^4}G({\bf K}), \\
J({\bf K}) &=& \frac{1}{V}\sum_{\bf Q} G({\bf Q}) G({\bf K}+{\bf Q}) .
\end{eqnarray}
The connected 2-pt function for $\Pi$ is given by
\begin{eqnarray}
\langle \Pi_{\msvec n}^i \Pi_{\msvec m}^j \rangle_c &\equiv& \left. \frac{-N\partial^2 W(J)}{\partial J^i_{\msvec n}\partial J_{\msvec m}^j}\right\vert_{J\rightarrow J_0} = \delta^{ij}\left[
G_{\msvec{nm}} -\frac{4}{N} \sum_{\msvec{ls}} K^{-1}_{\msvec{sl}}(J_0) G_{\msvec{nl}}G_{\msvec{ls}} G_{\msvec{sm}}\right]\nonumber \\
&=& \delta^{ij}\left[\frac{1}{V}\sum_{\bf K} G({\bf K}) {\rm e}^{i{\bf K}(\msvec{n}-\msvec{m})}
+ O\left(\frac{1}{N}\right)
\right] ,
\end{eqnarray}
so that the pole mass of pions in lattice units becomes
\begin{equation}
M_\pi = 2 \sinh^{-1} \frac{M}{2}.
\end{equation}
The connected 2-pt function for $\Sigma$, on the other hand, can be obtained as
\begin{eqnarray}
\langle \Sigma_{\msvec n} \Sigma_{\msvec m} \rangle_c &\equiv& \left. \frac{-N\partial^2 W(J)}{\partial J^N_{\msvec n}\partial J^N_{\msvec m}}\right\vert_{J\rightarrow J_0} =
G_{\msvec{nm}} -\frac{4H^2}{M^4} \sum_{\msvec{ls}}G_{\msvec{nl}}K^{-1}_{\msvec{ls}}(J_0) G_{\msvec{sm}}
+ O\left(\frac{1}{N}\right),\nonumber \\
&=&\frac{1}{V} \sum_{\bf K} \Gamma^{(2)}_\Sigma({\bf K}) {\rm e}^{i{\bf K}(\msvec{n}-\msvec{m})}
+ O\left(\frac{1}{N}\right)
\end{eqnarray}
where
\begin{eqnarray}
\Gamma^{(2)}_\Sigma({\bf K}) &=&\frac{G(\mvec{K})}{\mathcal{K}(\mvec{K})}
\left[\mathcal{K}(\mvec{K})-\frac{4H^2}{M^4}G(\mvec{K})\right] .
\end{eqnarray}
The connected 4-pt function for $\Pi$ is evaluated as\footnote{Note that the lattice delta function we use here is periodic.}
\begin{eqnarray}
&&\langle \Pi_{\msvec{n}_1}^{i_1} \Pi_{\msvec{n}_2}^{i_2} \Pi_{\msvec{n}_3}^{i_3} \Pi_{\msvec{n}_4}^{i_4} \rangle_c \equiv \left. \frac{-N\partial^4 W(J)}{\partial J^{i_1}_{\msvec{n}_1}\partial J^{i_2}_{\msvec{n}_2}\partial J^{i_3}_{\msvec{n}_3}\partial J^{i_4}_{\msvec{n}_4}}
\right\vert_{J\rightarrow J_0}\nonumber \\
&=& \frac{1}{V^3}\sum_{{\bf K}_1,{\bf K}_2,{\bf K}_3,{\bf K}_4}{\rm e}^{i({\bf K}_1\msvec{n}_1+{\bf K}_2\msvec{n}_2+{\bf K}_3\msvec{n}_3+{\bf K}_4\msvec{n}_4)}
\delta^{(3)}({\bf K}_1+{\bf K}_2+{\bf K}_3+{\bf K}_4) \frac{1}{N}\nonumber \\
&\times& G({\bf K}_1)G({\bf K}_2)G({\bf K}_3)G({\bf K}_4)
\left[ \delta^{i_1i_2}\delta^{i_3i_4}\left\{ \Gamma^{(4)}_{\Pi}({\bf K}_1+{\bf K}_2) +O\left(\frac{1}{N}\right) \right\}+ 2{\rm \ perms\ } \right],\nonumber \\
\end{eqnarray}
where
\begin{eqnarray}
\Gamma^{(4)}_{\Pi}({\bf K}) &=& -4\mathcal{K}^{-1}({\bf K}).
\end{eqnarray}
Similarly the connected 3-pt function for $\Pi\Pi\Sigma$ is given by
\begin{eqnarray}
&&\langle \Pi_{\msvec{n}_1}^{i_1} \Pi_{\msvec{n}_2}^{i_2} \Sigma_{\msvec{n}_3} \rangle_c \equiv \left. \frac{-N\partial^3 W(J)}{\partial J^{i_1}_{\msvec{n}_1}\partial J^{i_2}_{\msvec{n}_2}\partial J^{N}_{\msvec{n}_3} } \right\vert_{J\rightarrow J_0}
=\frac{1}{V^2}\sum_{{\bf K}_1,{\bf K}_2,{\bf K}_3}{\rm e}^{i({\bf K}_1\msvec{n}_1+{\bf K}_2\msvec{n}_2+{\bf K}_3\msvec{n}_3)}
\nonumber \\
&\times& \delta^{(3)}({\bf K}_1+{\bf K}_2+{\bf K}_3)
\frac{1}{\sqrt{N}}G({\bf K}_1)G({\bf K}_2)G({\bf K}_3)
\left[ \delta^{i_1i_2}\Gamma_{\Pi\Pi\Sigma}^{(3)}({\bf K}_3) + O\left(\frac{1}{N}\right) \right]
\end{eqnarray}
where
\begin{eqnarray}
\Gamma^{(3)}_{\Pi\Pi\Sigma}({\bf K}) &=& \frac{H}{M^2}\Gamma^{(4)}_{\Pi}({\bf K}) =\Sigma\, \Gamma^{(4)}_{\Pi}({\bf K}) .
\end{eqnarray}
\subsection{Scattering amplitude}
The pion scattering amplitude in the large $N$ limit is given by
\begin{eqnarray}
T^{i_1i_2,i_3i_4}({\bf k}_1,{\bf k}_2\ \vert\ {\bf k}_3, {\bf k}_4) &\equiv &
\lim_{{\bf k}_{1,2,3,4}\rightarrow
\mbox{on-shell, } \sum_i {\bf k}_i = 0}
\nonumber \\
&&
\frac{1}{N-1}
\left[\delta^{i_1i_2}\delta^{i_3i_4} \gamma_\pi^{(4)}({\bf k}_1+{\bf k}_2) + \mbox{ 2 perms}\right] ,
\end{eqnarray}
where on-shell momenta in the center of mass system are given by ${\bf k}_1 =(iE_p,\vec p)$, ${\bf k}_2 =(iE_p,-\vec p)$, ${\bf k}_3 =(-iE_q,\vec q)$, ${\bf k}_4 =(-iE_q,-\vec q)$ with $E_p^2 =\vec{p^2} +m_R^2=E_q^2 =\vec{q^2} +m_R^2$, $\vec p=(p_1,p_2)$ and $\vec q=(q_1,q_2)$.
Explicitly we have
\begin{eqnarray}
T^{i_1i_2,i_3i_4}({\bf k}_1,{\bf k}_2\ \vert\ {\bf k}_3, {\bf k}_4) &= &
\frac{1}{N-1}\left[\delta^{i_1i_2}\delta^{i_3i_4} \gamma_\pi^{(4)}(2iE_p,\vec 0)\right. \nonumber \\
&+&\left.
\delta^{i_1i_3}\delta^{i_2i_4} \gamma_\pi^{(4)}(0,\vec p+\vec q)
+\delta^{i_1i_4}\delta^{i_3i_2} \gamma_\pi^{(4)}(0,\vec p-\vec q)
\right] .
\end{eqnarray}
In terms of the "isospin" decomposition for $N-1$ pions
\begin{equation}
T^{i_1i_2,i_3i_4}({\bf k}_1,{\bf k}_2\ \vert\ {\bf k}_3, {\bf k}_4) =\sum_{I=0}^2 Q_I^{i_1i_2,i_3i_4}\, T_I(\vec p,\vec q)
\end{equation}
with projectors
\begin{eqnarray}
Q_0^{i_1i_2,i_3i_4} &=& \frac{1}{N-1}\delta^{i_1i_2}\delta^{i_3i_4}\,,
\\
Q_1^{i_1i_2,i_3i_4}&=&\frac12\left(\delta^{i_1i_3}\delta^{i_2i_4}-\delta^{i_1i_4}\delta^{i_2i_3}\right)\,,
\\
Q_2^{i_1i_2,i_3i_4}&=&\frac12\left(\delta^{i_1i_3}\delta^{i_2i_4}+\delta^{i_1i_4}\delta^{i_2i_3}\right)
-\frac{1}{N-1}\delta^{i_1i_2}\delta^{i_3i_4}\,,
\end{eqnarray}
we obtain
\begin{eqnarray}
T_0 (\vec p,\vec q) &=& \gamma_\pi^{(4)}(2iE_p,\vec 0)+\frac{1}{N-1} \left[\gamma_\pi^{(4)}(0,\vec p+\vec q)
+ \gamma_\pi^{(4)}(0,\vec p-\vec q) \right] , \\
T_1 (\vec p,\vec q) &=& \frac{1}{N-1} \left[ \gamma_\pi^{(4)}(0,\vec p+\vec q)
- \gamma_\pi^{(4)}(0,\vec p-\vec q) \right] , \\
T_2 (\vec p,\vec q) &=& \frac{1}{N-1} \left[ \gamma_\pi^{(4)}(0,\vec p+\vec q)
+ \gamma_\pi^{(4)}(0,\vec p-\vec q) \right] .
\end{eqnarray}
Therefore, in the large $N$ limit, we have
\begin{eqnarray}
T_0 (\vec p,\vec q) &=& \lim_{\varepsilon\rightarrow 0}\gamma_\pi^{(4)}((i-\varepsilon)W,\vec 0), \qquad
T_1 (\vec p,\vec q) =
T_2 (\vec p,\vec q) = 0 ,
\end{eqnarray}
where $W= 2 E_p$.
Using the integral formula in the continuum limit
\begin{eqnarray}
\lim_{a\rightarrow 0} \lim_{\varepsilon\rightarrow 0} a J\left( (i-\varepsilon)W a,\vec 0\right) &=&\frac{1}{4\pi W}{\rm arccoth} \left(\frac{W}{2m_R}\right) + i\frac{1}{8W},
\end{eqnarray}
we have
\begin{eqnarray}
T_0 (\vec p,\vec q) &=& -\frac{1}{X+iY},
\label{eq:amp}
\end{eqnarray}
where
\begin{eqnarray}
X&=&\left\{
\begin{array}{ll}
\displaystyle \frac{3}{u_{\rm eff}}+\frac{1}{8\pi W}{\rm arccoth} \left(\frac{W}{2m_R}\right), & \mbox { SYM } \\
& \\
\displaystyle \frac{3}{u_{\rm eff}}-\frac{w_R}{W^2}, & \mbox{ BRO}
\end{array}
\right. ,
\qquad
Y= \frac{1}{16W} .
\end{eqnarray}
\subsection{Unitarity and scattering phase shift}
The scattering amplitude $T_0$ in eq.~(\ref{eq:amp}) satisfies unitarity
\begin{eqnarray}
i\left[T_0 -T_0^\dagger\right] (\vec p, \vec q)&=& -\frac{1}{2W} \int \frac{{\rm d}^2 k}{(2\pi)^2}\frac{1}{2E_k}(2\pi)\delta(W-2E_k) T_0(\vec p, \vec k) T_0^\dagger(\vec k, \vec q),
\end{eqnarray}
where $E_k^2=\vec{k^2} + m_R^2$.
Therefore, $T_0$ can be expressed as
\begin{equation}
T_0(\vec p,\vec q) = 16 W {\rm e}^{i\delta_0(W)}\sin \delta_0(W)
\end{equation}
where $\delta_0(W)$ is the scattering phase shift for the $I=0$ channel, so that
we obtain
\begin{eqnarray}
\cot \delta_0(W) &=& -\frac{X}{Y} =\left\{
\begin{array}{ll}
\displaystyle -\frac{48W}{u_{\rm eff}}-\frac{2}{\pi}{\rm arccoth} \left(\frac{W}{2m_R}\right), & \mbox{ SYM } \\
\\
\displaystyle -\frac{48W}{u_{\rm eff}}+\frac{16w_R}{W}, & \mbox{ BRO } \\
\end{array}
\right. .
\end{eqnarray}
Fig.~\ref{fig:ScatteringPhase} shows behaviors of $\delta_0(W)$ as a function of $W$ in the symmetric phase, which clearly reflect behaviors of the running coupling in the various continuum limits:
In the case IA given by the dashed line, $\delta_0(W)$ starts from 0 at $W=0$ and monotonically approaches $-\pi/2$ as $W$ increases. $\delta_0(W) = -\pi/2$ is the value for the conformal theory, as shown by the solid magenta line corresponding to the case IB. This shows that the case IA is UV conformal.
The behavior of $\delta_0(W)$ in the general case IIA depends on the mass, namely on the parameter $b=96\alpha = 96 m_R/u_{\rm eff}$. At relatively large $m_R$ ($b=1.19$) denoted by the solid black line, $\delta_0(W)$ first decreases from 0 as $W$ increases but, at some value of $W$, it starts to increase toward 0, showing that the IIA case is asymptotically free in the UV. If we decrease $m_R$ ( $b=0.00119$ ), $\delta_0(W)$ rapidly decreases from 0 to $-\pi/2$ (the conformal value) near $W=0$ and gradually increases toward 0 for increasing $W$, as shown by the solid red line. In the case of the massless limit, the case IIB, $\delta_0(W) = -\pi/2$ at $W=0$ and monotonically increases toward 0, showing the theory is IR conformal and UV asymptotically free. The IIA case with small mass such as $b=0.00119$ is nearly conformal, and therefore has "walking" coupling.
This shows that, although the running coupling is not a physical observable and depends on how it is defined, it captures some properties of the scattering phase shift. In other words, it opens a possibility to identify a nearly conformal theory (or the walking coupling) unambiguously from the physical observable, the scattering phase shift.
\begin{figure}[tbh]
\begin{center}
\scalebox{0.4}{\includegraphics{ScatteringPhase.pdf}}
\end{center}
\caption{Scattering phase shift $\delta_0(W)$ as a function of $W/u_{\rm eff}$ for
IA(dashed line), IB(solid magenta line), IIA with $\b=96\alpha=1.19$ (solid black line), IIA with $b=0.00119$ (solid red line) and IIB (solid blue line).
}
\label{fig:ScatteringPhase}
\end{figure}
\section{Introduction}
\label{intro}
\input{intro_walk}
\section{Large $N$ expansion of the lattice model}
\label{largen}
\input{largen}
\section{Renormalization and continuum limits}
\label{cont}
\input{cont}
\section{Running four-point coupling}
\label{g4}
\input{g4}
\section{Scattering phase shifts}
\label{scatt}
\input{scatt}
\section{Summary}
\label{summary}
\input{summary_walk}
\section*{Acknowledgement}
S. A. is supported in part by the Grant-in-Aid of the Japanese Ministry of Education, Sciences and Technology, Sports and Culture (MEXT) for Scientific Research (No. 25287046) and by MEXT Strategic Program for Innovative Research (SPIRE) Field 5 and Joint Institute for Computational Fundamental Science (JICFuS).
This investigation has also been supported by the European Union
and the State of Hungary and co-financed by the European Social Fund in the framework
of TAMOP-4.2.4.A/ 2-11/1-2012-0001 ${}^\prime$National Excellence Program$^\prime$.
S.A would like to thank the Wigner Research Center for Physics for its kind hospitality during his stay for this research project.
S. A and J. B. would like to thank the Max-Planck-Institut f\"ur Physik for its kind hospitality during their stay for this research project.
|
\section{Introduction}
Let $G$ be a graph with adjacency matrix $A$ whose eigenvalues are $\lambda_{1} \geq \lambda_{2} \geq \ldots \geq \lambda_{n}$. The energy of $G$ is then defined as:
$$
E=\sum_{i=1}^{n}{|\lambda_{i}|}.
$$
This concept has been first introduced and intensively studied in the context of mathematical chemistry but in the last 15 years it has garnered a lot of attention from graph theorists as well. For overviews of the subject we refer the reader to the recent book \cite{Energy} by Li, Shi, and Gutman and to the earlier surveys by Gutman \cite{Gutman_survey} and Brualdi \cite{Brualdi_energy}.
Our aim in this paper is to contribute a new lower bound for the energy, obtained by revisiting the original approach of one of the early pioneers, McClelland and bringing a discrete variant of the well-known Gr\"{u}ss integral inequality to bear on it.
\subsection{Notation, terminology, and some standard facts}
Throughout the paper we shall assume that the graph $G$ has $n$ vertices and $m$ edges. We shall denote by $t$ the smallest absolute value of an eigenvalue of $G$, that is: $t=\min\{|\lambda_{i}|\}$. The graph will be called \emph{singular} if $t=0$ and \emph{non-singular} otherwise.
By the Perron-Frobenius theorem it is known that $|\lambda_{1}| \geq |\lambda_{i}|$ for any $i$. We also have the following well-known fact:
$$
\sum_{i=1}^{n}{\lambda_{i}^{2}}=\Tr{A^{2}}=2m.
$$
\section{Some known results}\label{sec:sur}
In this brief section we do not attempt to provide an exhaustive survey, but rather to indicate the main lower bounds on energy that are present in the literature, so that the reader can compare them with the new result we shall derive.
In the halcyon days of graph energy McClelland \cite{McCle71} obtained the following bounds:
\begin{thm}\cite{McCle71}\label{thm:mcc}
$$\sqrt{2m+n(n-1)|A|^{\frac{2}{n}}} \leq E \leq \sqrt{2mn}.$$
\end{thm}
A different lower bound has been given by Caporossi \emph{et al.} \cite{Caporossi_etal99}:
\begin{thm}\cite{Caporossi_etal99}\label{thm:caporossi}
$$E \geq 2\sqrt{m}.$$
\end{thm}
Clearly, for singular graphs Theorem \ref{thm:caporossi} is better than the lower bound of Theorem \ref{thm:mcc}.
\section{Statement of the new results and some discussion}\label{sec:main}
Our first main result is:
\begin{thm}\label{thm:main}
\begin{equation}\label{eq:main}
E \geq \frac{2m+n\lambda_{1}t}{\lambda_{1}+t}.
\end{equation}
\end{thm}
Since the right-hand side of \eqref{eq:main} is a non-decreasing function of $t$, we can deduce:
\begin{cor}\label{cor:nice}
$$E \geq \frac{2m}{\lambda_{1}}.$$
\end{cor}
\begin{rmrk}
Equality is attained in Corollary \ref{cor:nice} for the complete bipartite graphs $K_{p,q}$ with $E=2\sqrt{pq}$ and $m=2pq,\lambda_{1}=\sqrt{pq}$. I am grateful to Dr. Clive Elphick for this observation. It would be an interesting problem to try to find other graphs - if there are any - which attain equality.
\end{rmrk}
Since a $d$-regular graph has $2m=nd$ and $\lambda_{1}=d$, Corollary \ref{cor:nice} implies the following result by Gutman \emph{et al.} \cite{Gutman_etal07}:
\begin{thm}\cite{Gutman_etal07}
Let $G$ be an $r$-regular graph, $r>0$. Then $$E \geq n.$$
\end{thm}
Another class of graphs for which Corollary \ref{cor:nice} improves upon known results is that of triangle-free graphs. It is known \cite{Nosal} that for them $\lambda_{1} \leq \sqrt{m}$ and therefore Corollary \ref{cor:nice} is better than Theorem \ref{thm:caporossi} over this class.
It is also possible to deduce, using the arithmetic-geometric means inequality, another consequence of Theorem \ref{thm:main} which is not very strong but forms a nice counterpart to Theorem \ref{thm:mcc}:
\begin{cor}
$$
E \geq \sqrt{2mn} \cdot \sqrt{\frac{4\lambda_{1}t}{(\lambda_{1}+t)^{2}}}.
$$
\end{cor}
As we shall see, our approach will enable us to give an even stronger bound than \eqref{eq:main}. To state it we need to single out the smallest \emph{non-zero} eigenvalue of the graph:
$$
t_{nz}=\min\{|\lambda_{i}| \Big{\vert} \lambda_{i} \neq 0\}.
$$
\begin{thm}\label{thm:better}
Let $r=rank(A)$. Then
$$
E \geq \frac{2m+r\lambda_{1}t_{nz}}{\lambda_{1}+t_{nz}}.
$$
\end{thm}
We defer the proofs of Theorems \ref{thm:main} and \ref{thm:better} till Section \ref{sec:proof}, while in the next section we set up the engine of the proof.
\section{The Gr\"{u}ss inequality}
Chebyshev's classic inequality says that if $f,g:[a,b] \rightarrow \mathbb{R}$ are integrable functions, either both increasing or both decreasing, then
\begin{equation}\label{eq:cheb}
\int_{a}^{b}{f(x)g(x)dx} \geq \frac{1}{b-a}\int_{a}^{b}{f(x)dx}\int_{a}^{b}{g(x)dx}.
\end{equation}
This elegant inequality has been generalized and extended in many ways. We refer to \cite[Chapters IX--X]{MPF_book} for a survey of some of these developments.
In 1935 Gr\"{u}ss proved the following result:
\begin{thm}\cite{Gru35}\label{thm:gruss}
Let $f,g:[a,b] \rightarrow \mathbb{R}$ be integrable functions such that
\begin{equation}\label{eq:gruss_cond}
\phi \leq f(x) \leq \Phi \textit{ and } \gamma \leq g(x) \leq \Gamma \textit{ for all } x \in [a,b].
\end{equation}
Then
$$
\Big|\frac{1}{b-a}\int_{a}^{b}{f(x)g(x)dx} - \frac{1}{b-a}\int_{a}^{b}{f(x)dx}\frac{1}{b-a}\int_{a}^{b}{g(x)dx}\Big| \leq \frac{1}{4}(\Phi-\phi)(\Gamma-\gamma).
$$
\end{thm}
Now let us state the abstract formulation due to Dragomir \cite{Dra99}, as it permits an easy derivation of the discrete variant we need: Let $(X,(\cdot))$ be a real inner product space and let $e \in X, ||e||=1$. The \emph{Chebyshev functional} on $X$ is defined as:
$$
\forall x,y \in X \quad T(x,y)=\langle x,y \rangle -\langle x,e \rangle \langle y,e \rangle.
$$
Taking the product with $e$ is the operation of taking a "mean". For any $z \in X$ we denote $A(z)=\langle x,e \rangle$.
\begin{thm}\cite{Dra99}\label{thm:drag1}
Let $x,y \in X$ be such that there exist $\phi,\gamma,\Phi,\Gamma \in \mathbb{R}$ so that the conditions
\begin{equation}\label{eq:gruss_abs}
\langle \Phi e-x,x-\phi e \rangle \geq 0 \textit{ and } \langle \Gamma e-y,y-\gamma e \rangle \geq 0
\end{equation}
hold. Then
$$
|T(x,y)| \leq \frac{1}{4}|\Phi-\phi||\Gamma-\gamma|.
$$
\end{thm}
Note that \eqref{eq:gruss_abs} reduces to \eqref{eq:gruss_cond} for the inner product $\langle f,g \rangle=\frac{1}{b-a}\int_{a}^{b}f(x)g(x)dx$.
We shall need a stronger version of Theorem \ref{thm:drag1}, which is implicit in Dragomir's proof:
\begin{thm}\label{thm:drag2}
Under the assumptions of Theorem \ref{thm:drag1},
$$
|T(x,y)| \leq \sqrt{|\Phi-A(x)||A(x)-\phi||\Gamma-A(y)||A(y)-\gamma|}.
$$
\end{thm}
Now let us equip $X=\mathbb{R}^{n}$ with the inner product $\langle x,y\rangle=\frac{1}{n}\sum_{i=1}^{n}{x_{i}y_{i}}$ and record the following consequence of Theorem \ref{thm:drag2}:
\begin{thm}\label{thm:drag3}
Let $x,y \in \mathbb{R}^{n}$ and let $A(x)=\frac{1}{n}\sum_{i=1}^{n}{x_{i}},A(y)=\frac{1}{n}\sum_{i=1}^{n}{y_{i}}$.
If $\phi \leq x_{i} \leq \Phi$ and $\gamma \leq y_{i} \leq \Gamma$, then
$$
\Big|\frac{1}{n}\sum_{i=1}^{n}{x_{i}y_{i}}-\frac{1}{n^{2}}\sum_{i=1}^{n}{x_{i}}\sum_{i=1}^{n}{y_{i}}\Big| \leq \sqrt{(\Phi-A(x))(A(x)-\phi)(\Gamma-A(y))(A(y)-\gamma)}.
$$
\end{thm}
\section{Proofs for Section \ref{sec:main}}\label{sec:proof}
The following observation goes back to McClelland \cite{McCle71}:
$$
E^{2}=\sum_{i=1}^{n}{|\lambda_{i}|^{2}}+\sum_{i \neq j}{|\lambda_{i}||\lambda_{j}|}=2m+\sum_{i \neq j}{|\lambda_{i}||\lambda_{j}|}.
$$
Let us denote $P=\sum_{i \neq j}{|\lambda_{i}||\lambda_{j}|}$. It is clear that estimating $E$ is equivalent to estimating $P$. In \cite{Caporossi_etal99} the bound $P \geq 2m$ was observed, leading to the claim of Theorem \ref{thm:caporossi}. We show here a different approach to bounding $P$, based on representing it as an inner product.
Let $x_{i}=|\lambda_{i}|$ and $y_{i}=E-|\lambda_{i}|$ for $1 \leq i \leq n$. Then it is easy to see that
$$
P=\sum_{i=1}^{n}{x_{i}y_{i}}.
$$
Observe that $$t \leq x_{i} \leq \lambda_{1} \textit{ and } E-\lambda_{1} \leq y_{i} \leq E-t.$$ Also, $$A(x)=\frac{E}{n} \textit{ and } A(y)=\frac{(n-1)E}{n}.$$
Now we apply Theorem \ref{thm:drag3} to obtain:
$$
\Big|\frac{P}{n}-\frac{(n-1)E^{2}}{n^{2}}\Big| \leq \sqrt{(\lambda_{1}-\frac{E}{n})(\frac{E}{n}-t)(\frac{E}{n}-t)(\lambda_{1}-\frac{E}{n})}=(\lambda_{1}-\frac{E}{n})(\frac{E}{n}-t).
$$
Therefore:
$$
P \geq n\Big(\frac{(n-1)E^{2}}{n^{2}}-(\lambda_{1}-\frac{E}{n})(\frac{E}{n}-t)\Big)=E^{2}+n\lambda_{1}t-(\lambda_{1}+t)E.
$$
This implies:
$$
E^{2}=2m+P \geq 2m+E^{2}+n\lambda_{1}t-(\lambda_{1}+t)E
$$
which immediately leads to \eqref{eq:main} upon trivial re-arrangement. This concludes the proof of Theorem \ref{thm:main}. \qed
\medskip
To prove Theorem \ref{thm:better} it is only necessary to observe that zero eigenvalues of $A$ correspond to zero entries in the vector $x$. Delete them and the corresponding entries from $y$ to obtain shorter vectors $x^{'},y^{'} \in \mathbb{R}^{r}$ which satisfy:
$$
P=\sum_{i=1}^{n}{x^{'}_{i}y^{'}_{i}},
$$
$$t_{nz} \leq x_{i} \leq \lambda_{1} \quad \quad E-\lambda_{1} \leq y_{i} \leq E-t_{nz}$$ and $$A(x)=\frac{E}{r} \quad \quad A(y)=\frac{(r-1)E}{r}.$$
Therefore the arguments given before work the same way, with $t$ replaced by $t_{nz}$ and $n$ replaced by $r$. \qed
\section{Two conjectures by Elphick}
Dr. Clive Elphick has communicated to me two very interesting conjectures engendered by the results reported here. To state them let us introduce two measures of the irregularity of a graph, studied in the paper \cite{ElpWoc13} by Elphick and Wocjan. Let $d_{1},d_{2},\ldots,d_{n}$ be the vertex degress of $G$ and let $d$ be the average degree. Define:
$$
\epsilon=\frac{n\sum_{i \sim j}{\sqrt{d_{i}d_{j}}}}{2m^{2}}
$$
and
$$
\beta=\frac{\lambda_{1}}{d}=\frac{\lambda_{1}n}{2m}.
$$
\smallskip
It is known that $\beta \geq \epsilon \geq 1$ (cf. \cite[p. 53]{ElpWoc13}). We can now state the conjectures.
\begin{conj}
Let $G$ be a connected graph. Then
$$
E \geq \frac{n}{\epsilon}.
$$
\end{conj}
Since $\frac{2m}{\lambda_{1}}=\frac{n}{\beta}$, this would be an improvement of Corollary \ref{cor:nice}.
\begin{conj}
Let $G$ be a connected graph. Then
$$
E \leq \frac{2m}{\sqrt{\lambda_{1}}}.
$$
\end{conj}
Since $\frac{2m}{\sqrt{\lambda_{1}}} \leq \sqrt{\frac{2mn}{\beta}} \leq \sqrt{2mn}$ this would an improvement over the upper bound of Theorem \ref{thm:mcc}.
Both conjectures have been verified for all $11117$ graphs on eight vertices using a computer. The connectedness assumption is essential for both conjectures.
\bibliographystyle{abbrv}
|
\section{Introduction.}\label{intro}
\section{Introduction}\label{s:introduction}
A variation of this model was studied in filtering theory by Kalman
and Bucy (1961) \cite{Kalman1961} and Zakai (1969) \cite{Zakai1969}.
They analyze a more general model where a decision maker observes a
function of a diffusion process with an additional noise, which is
formulated as a Brownian motion. They provide equations that the
posterior or the unnormalized posterior distribution at time $t$
satisfies. Bandit problems, first described in Robbins (1952)
\cite{Robbins}, are a mathematical model for studying the trade
between exploration and exploitation. In its simplest formulation, a
decision maker (DM) faces $N$ slot machines (called \emph{arms}) and
has to choose one of them at each time instance. Each slot machine
delivers a reward when and only when chosen. The reward's
distribution of each slot machine is drawn according to an unknown
distribution, which itself is drawn according to a known probability
distribution from a set of known distributions. The DM's goal is to
maximize his total discounted payoff. The trade-off that the DM
faces at each stage is between exploiting the information that he
already has, that is, choosing the arm that looks optimal according
to his information, and exploring the arms, that is, choosing a
suboptimal arm to improve his information about its payoff
distribution. A good strategy for the DM will involve phases of
exploration and phases of exploitation. In exploration phases the DM
samples the rewards of the various machines and learns their
rewards' distributions. In exploitation phases the DM samples the
machine whose reward's distribution so far is best until evidence
shows that its reward's distribution is not as good as expected.
Bandit problems have been applied to various areas, like economics,
control, statistics, and learning; see, e.g., Chernoff (1972)
\cite{Chernoff}, Rothschild (1974) \cite{Rothschild}, Weitzman
(1979) \cite{Weitzman}, Roberts and Weitzman (1981) \cite{Roberts},
Lai and Robbins (1984) \cite{Lai}, Bolton and Harris (1999)
\cite{Bolton}, Moscarini and Squintani (2010) \cite{Moscarini},
Keller, Rady, and Cripps (2005) \cite{Keller2005}, Bergemann and
V\"{a}lim\"{a}ki (2006) \cite{Bergemann}, Besanko and Wu (2008)
\cite{Besanko}, and Klein and Rady (2011) \cite{Klein}.
Gittins and Jones (1979) \cite{Gittins} proved that in discrete time
the optimal strategy of the DM has a particularly simple form: at
every period the DM calculates for each arm an index, which is a
real number, based on past rewards of that arm, and chooses the arm
with the highest index. It turns out that to calculate the index of
an arm it is sufficient to consider an auxiliary problem with two
arms: an arm for which the index is calculated and an arm that
yields a constant payoff. The former arm is termed the \emph{risky}
arm, because its payoff distribution is not known, while the latter
is termed the \emph{safe} arm. The literature therefore focuses on
such problems, called two-armed bandit problems.\footnote{In the
literature these problems are also called one-armed bandit
problems.}
Once the optimality of the index strategy is guaranteed, one looks
for the relation between the data of the problems and the index.
Explicit formulas for the index when the payoff is one of two
distributions that have a simple form have been established in the
literature. Berry and Friestedt (1985) \cite{Berry} provide the
solution to the problem in several cases, e.g., in discrete time
when the payoff distribution is one of two Bernoulli distributions,
and in continuous time when the payoff distribution is one of two
Brownian motions. By studying the dynamic programming equation that
describes the problem in continuous time, Keller, Rady, and Cripps
(2005) \cite{Keller2005} and Keller and Rady (2010)
\cite{Keller2010} provided an explicit form for the index when time
is continuous and the payoff's distribution
is Poisson
\footnote{These authors also studied the strategic setup in which
several DMs have the same set of arms and their arms' payoff
distributions are the same (and unknown), and they compared the
cooperative solution to the non-cooperative solution.}
In the present paper we study two-armed bandit problems in
continuous time and provide an explicit solution when the payoff
distribution of the risky arm is one of two L\'{e}vy processes. We
assume that one distribution, called {\em High}, dominates the
other, called {\em Low}, in a strong sense (see Assumption
\ref{assumption} below). To eliminate trivial cases, we assume that
the expected payoff generated by the safe arm is lower than the
expected payoff generated by the High distribution, and higher than
the expected payoff generated by the Low distribution.
In discrete time, under these assumptions the optimal strategy is a
cut-off strategy: the DM keeps experimenting as long as the
posterior belief that the distribution is High is higher than some
cut-off point, and, once the posterior probability that the
distribution is High falls below the cut-off point, the DM switches
to the safe arm forever. We extend this result to our setup, and
prove that when the two payoff distributions are L\'{e}vy processes
that satisfy several requirements, the optimal strategy is a cut-off
strategy. Moreover, we provide an explicit expression for the
cut-off point in terms of the data of the problem. When
particularized to the models studied by Bolton and Harris (1999)
\cite{Bolton}, Keller, Rady, and Cripps (2005) \cite{Keller2005},
and Keller and Rady (2010) \cite{Keller2010} our expression reduces
to the expressions that they obtained.
Apart from unifying previous results, our characterization shows
that the special form of the optimal payoff derived by Bolton and
Harris (1999) \cite{Bolton} and Keller, Rady, and Cripps (2005)
\cite{Keller2005} is valid in a general setup: the optimal payoff is
the sum of the expected payoff, if no information is available, and
an option value that measures the expected gain from the ability to
experiment. It also shows that the data of the problem can be
divided into information-relevant parameters and payoff-relevant
parameters; the information-relevant parameters can be summarized in
a single real number, and the payoff-relevant parameters are the
expectations of the processes that contribute to the DM's payoff.
Finally, the characterization allows one to derive comparative
statics on the optimal cut-off and payoff. For example, as the
discount rate increases, or the signals become less informative, the
cut-off point increases and the DM's optimal payoff decreases.
The rest of the paper is organized as follows. In Section \ref{s:the
model} we present the model, the types of strategies that we allow,
and the assumptions that the payoff process should satisfy. In
Section \ref{s:posterior} we define the process of posterior belief
and we develop its infinitesimal generator. In Section \ref{s:value}
we present the value function, and in Section \ref{s:HJB} we present
the Hamilton--Jacobi--Bellman (HJB) equation. The main result, which
characterizes the optimal strategy and the optimal payoff of the DM,
is formulated and proved in Section \ref{s:main}. The Appendix
contains the proofs of several results that are needed in the paper.
\section{The model}\label{s:the model}
\subsection{Reminder about L\'{e}vy processes}\label{s:reminder}
L\'{e}vy processes are the continuous-time analog of discrete-time
random walks with i.i.d.~increments. A \emph{L\'{e}vy process}
$X=(X(t))_{t\geq 0}$ is a continuous-time stochastic process that
(a) starts at the origin: $X(0)=0$, (b) admits c\`{a}dl\`{a}g
modification,\footnote{That is, it is continuous from the right, and
has limits from the left: for every $t_0$, the limit $X(t_0
-):=\underset{t\nearrow t_0}{\lim}X(t)$ exists a.s. and $X(t_0)
=\underset{t\searrow t_0}{\lim}X(t)$. } and (c) has stationary
independent increments. Examples of L\'{e}vy processes are a
Brownian motion, a Poisson process, and a compound Poisson process.
Let $(X(t))$ be a L\'{e}vy process. For every Borel measurable set
$A\subseteq \mathbb{R}\backslash \{0\}$, and every $t\geq 0$, let
the Poisson random measure $N(t ,A)$ be the number of jumps of
$(X(t))$ in the time interval $[0,t]$ with jump size in $A$:
\[
N(t,A)=\sharp\{0\leq s \leq t \mid \Delta X(t): = X(s) - X(s-) \in
A\}.
\]
By Applebaum (2004) \cite{Applebaum}, one can define a Borel measure
$\nu$ on $\mathcal{B}(\mathbb{R}\backslash\{0\})$ by
\[\nu (A) := E[N(1,A)] = \int N(1,A)(\omega)dP(\omega),\] where $(\Omega,P)$ is the underlying probability space.
The measure $\nu (A)$ is called the \emph{L\'{e}vy measure} of
$(X(t))$, or the \emph{intensity measure} associated with $(X(t))$.
We now present the L\'{e}vy--It\={o} decomposition of L\'{e}vy
processes. Let $(X(t))$ be a L\'{e}vy process; then there exists a
constant $b\in \mathbb{R}$, a Brownian motion $\sigma Z(t)$ with
standard deviation $\sigma$, and an independent Poisson random
measure $N_{\nu}(t,dh)$ with the associated L\'{e}vy measure $\nu$
such that, for each $t\geq 0 $,
\begin{equation}\notag
X(t) = b t +\sigma Z(t) + \underset{h>|1|}{\int} hN_{\nu}(t,dh) +
\underset{h \leq |1|}{\int} h\widetilde{N}_{\nu}(t,dh),
\end{equation}
where $\widetilde{N}_{\nu} (t,A) := N_{\nu}(t,A)-t\nu (A)$ is the
\emph{compensated Poisson random measure}. This representation is
called the \emph{L\'{e}vy--It\={o} decomposition} of the L\'{e}vy
process $(X(t))$. Thus, a L\'{e}vy processes is characterized by the
triplet $\langle b,\sigma, \nu \rangle$.
If the L\'{e}vy process has finite expectation for each $t$, that
is, $E|X(t)|<\infty$ for all $t\geq 0 $, then the L\'{e}vy process
can be represented as
\begin{equation}\notag
X(t) = \mu t +\sigma Z(t) + \underset{\mathbb{R}\backslash \{0\}
}{\int} h\widetilde{N}_{\nu}(t,dh) ;
\end{equation}
that is, $X(t)$ can be represented as the sum of a linear drift, a
Brownian motion, and an independent purely discontinuous
martingale\footnote{A \textit{purely discontinuous process} is a
process that is orthogonal to all continuous local martingales. For
details, see Jacod and Shiryaev (1987, Ch.~I, Definition 4.11)
\cite{Jacod}.} (see Sato (1999, Theorem 25.3) \cite{Sato1987}).
\begin{rem} Even though the process $(X(t))$ has finite expectation, it is possible that
\[
E\left[\underset{\mathbb{R}\backslash \{0\}}{\int}
|h|N_{\nu}(t,dh)\right]= \infty,
\]
which means that the expectation of the sum of the jumps of $X(t)$
in any time interval is infinite.
\end{rem}
\subsection{L\'{e}vy bandits}\label{s:finite}
A DM operates a two-armed bandit machine in continuous time, with a
safe arm that yields a constant payoff $\varrho$, and a risky arm
that yields a stochastic payoff $(X(t))$ that depends on its type
$\theta$. The risky arm can be of two types, High or Low. With
probability $p_0=p$ the arm's type is High, and with probability
$1-p$ it is Low. If the type is High (resp. Low) we set $\theta = 1$
(resp. $0$). The process $(X(t))$ is a L\'{e}vy process with the
triplet $\langle\mu_\theta,\sigma,\nu_\theta\rangle$; that is, the
L\'{e}vy--It\={o} decomposition of $(X (t))$ is $X(t) = \mu_\theta t
+ \sigma_\theta Z(t) + \underset{\mathbb{R}\backslash \{0\} }{\int}
h\widetilde{N}_{\nu_\theta}(t,dh) $. Formally, for
$\theta\in\{0,1\}$, let $(X_\theta (t))$ be a L\'{e}vy process with
triplet $\langle\mu_\theta,\sigma,\nu_\theta\rangle$ and let
$\theta$ be an independent Bernoulli random variable with parameter
$p$. The process $(X(t))$ is defined to be $(X_\theta(t))$.
We denote by $P_p$ the probability measure over the space of
realized paths that corresponds to this description.
From now on, unless mentioned otherwise, all the expectations are
taken under the probability measure $P_p$.
\subsection{Strategies}\label{s:strategies}
We adopt the concept of continuous-time strategies first introduced
by Mandelbaum, Shepp, and Vanderbei (1990) \cite{Mandelbaum}. An
\emph{allocation strategy} $\KK= \{ \KK (t)\mid t\in[0,\infty)\}$ is
a nonnegative stochastic process $\KK(t)=(\KR(t),\KS(t))$ that
satisfies
\begin{align}\tag{K1}\label{T1}
&\KR(0) = \KS(0) = 0, \ \textrm{and}\ (\KR(t)) \;\textrm{and}\
(\KS(t))\ \textrm{are nondecreasing processes},\\\tag{K2}\label{T2}
&\KR(t)+\KS(t)=t, \;\; t\in[0,\infty), and \\\tag{K3}\label{T3}
&\{\KR(t)\leq s\} \in \F_s^X, \; \; t,s\in[0,\infty),\notag
\end{align}
where $\F_s^X$ is the sigma-algebra generated by the process
$(X(t))_{t\leq s}$. The interpretation of an allocation process is
that the quantity $\KR(t)$ (resp.~$\KS(t)$) is the time that the DM
devotes to the risky arm (resp.~safe arm) during the time interval
$[0,t)$. The process $(\KK(t))$ is basically a two-parameter time
change of the two-dimensional process $(X(t),\varrho t)$.
Below we will define a stochastic integral with respect to
$(X_\theta(t))$, and therefore we assume throughout that both
L\'{e}vy processes $ (X_{1} (t))$ and $(X_{0} (t))$ have finite
quadratic variation, that is, $E[X_\theta^2(t)]<\infty $ for every
$t\geq 0$ and each $\theta\in\{0,1\}$. It follows that the processes
$(X_\theta(t))$ have finite expectation.
\begin{assumption}$\\$\label{finite_quadratic}
A1. $E[X^2_\theta (1)]=\mu_\theta^2 +\sigma^2 +{\int}
h^2\nu_\theta(dh)<\infty.$
\end{assumption}
For every $(t,p)\in [0,\infty)\times[0,1]$, every real-valued
function $S:\mathbb{R}\rightarrow \mathbb{R}$, and every pair of
Markov processes $(H_1(t))$ and $(H_2(t))$ with respect to the
filtration $(\F_t^X)_{t\geq 0}$ under both $P_0$ and $P_1$ such that
$E \left[S(\int_t^{\infty} H_1(s) dH_2(s))\mid \theta \right]$ are
well defined for both $\theta\in\{0,1\}$, we define the following
expectation operator:
\begin{align}
\notag E^{t,p} \left[S\left(\int_t^{\infty} H_1(s)
dH_2(s)\right)\right]:&= \left.p E \left[S\left(\int_t^{\infty}
H_1(s) dH_2(s)\right)\right| \theta =
1\right]\\
\label{Etp} &+\left.(1-p) E \left[S\left(\int_t^{\infty} H_1(s)
dH_2(s)\right)\right| \theta = 0\right].\\\notag
\end{align}
Using this notation, the expected discounted payoff from time $t$
onwards under allocation strategy $\KK$ when the prior belief at
time $t$ is $p_t =p$ can be expressed as
\begin{align}\label{VKtp}
V_{\KK}(t,p) &:= E^{t,p} \left[\int_t^{\infty} r
e^{-rs}dY(\KK(s))\right],
\\\notag
\end{align}
where $Y(\KK(s)):=X(\KR(s))+\varrho \KS(s)$. The goal of the DM is
to maximize $V_{\KK}(0,p)$. Let
\begin{align}\label{defineU}
U(t,p) := \underset{\KK}{\sup}V_{\KK}(t,p)\\\notag
\end{align}
be the maximal payoff the DM can achieve from time $t$ onwards,
given that the prior belief at time $t$ is $p_t =p$. As we show in
Theorem \ref{Up1} below, under proper assumptions the DM has an
optimal strategy, so in fact the supremum in Eq. (\ref{defineU}) is
achieved. Moreover, we give explicit expressions for both the
optimal strategy and the optimal value function $U(t,p)$.
\begin{rem}\label{key}
By Conditions (\ref{T1}) and (\ref{T2}), $\KR$ and $\KS$ are
Lipschitz and thus absolutely continuous. Therefore, there exists a
two-dimensional stochastic process $\KK'(t)=\dfrac{d\KK}{dt}(t) =
(\KR'(t), \KS'(t))$ such that $\KK(t)=\int_0^t \KK'(s) ds$. To
simplify notation we denote $\K(t):=\KR (t)$, and $\key(t)
:=\K_R'(t)$. Hence, $\KK(t)=(\K(t),t-\K(t))$ and
$\KK'(t)=(\key(t),1-\key(t))$. The process $(\key(t))$ may be
interpreted as follows: At each time instance $t$, the DM chooses
$\key(t)$ (resp. $ 1 - \key(t)$), the proportion of time in the
interval $[t,t+dt)$ that is devoted to the risky arm (resp. the safe
arm). The process $(\key(t))$ will be treated as a stochastic
control parameter of the process $(X(t))$. Denote by $\F_{\K(t)}$
the sigma-algebra generated by $(X(\K(s)))_{s\leq t}$.
\end{rem}
\begin{defn}
An \emph{admissible control strategy} $(\key(t,\omega))$ is any
predictable process such that $0\leq \key\leq 1$ with probability
$1$, and such that the process $\K(t)=\int_0^t \key(s) ds$
satisfies\footnote{Since $0\leq \key\leq 1$, it follows that $\K(t)$
satisfies Conditions (\ref{T1}) and (\ref{T2}) as well.} Condition
(\ref{T3}). Denote by $\Upsilon$ the set of all admissible control
strategies.
\end{defn}
In the sequel we will not distinguish between the allocation
strategy $(\K(t))$ and the corresponding admissible control strategy
$(\key(t))$.
\subsection{Assumptions}
If the DM could deduce the type of the risky arm by observing the
payoff of the risky arm in an infinitesimal time interval, then an
almost-optimal strategy is to start at time $0$ with the risky arm,
and switch at time $\delta$ to the safe arm if the type of the risky
arm is Low, where $\delta>0$ is a small real number. Throughout the
paper we make the following assumption, which implies that the DM
cannot distinguish between the two types in any infinitesimal time.
\begin{assumption}\label{assumption}$\\$
A2. $\sigma_{\high}=\sigma_{\low}$.\\
A3. $|\nu_1(\mathbb{R}\setminus \{0\}) - \nu_0(\mathbb{R}\setminus \{0\})| < \infty$.\\
A4. $|\int h (\nu_1 (dh) - \nu_0 (dh))|<\infty$.\\
\end{assumption}
Assumption A2 states that the Brownian motion component of both the
High type and the Low type have the same standard deviation. By
Revuz and Yor (1999, Ch.~I, Theorem 2.7) \cite{Revuz} the realized
path reveals the standard deviation and therefore if Assumption A2
does not hold then the DM can distinguish between the arms in any
infinitesimal time interval. Assumption A3 states that the
difference between the L\'{e}vy measures is finite and Assumption A4
states that the difference between the expectation of the jump part
of the processes is finite. Otherwise, by comparing the jump part of
the processes, the DM could distinguish between the arms in any
infinitesimal time interval.
We also need the following assumption, which states that the High
type is better then the Low type in a strong sense.
\begin{assumption}\label{assumption}$\\$
A5. $\mu_0<\varrho<\mu_1$.\\
A6. For every $A\in\mathcal{B}(\mathbb{R}\setminus \{0\}),\; \; \vvi
(A)\leq\vi (A)$.
\end{assumption}
Assumption A5 merely says that the High (resp. Low) type provides
higher (resp. lower) expected payoff than the safe arm. Assumption
A6 is less innocuous; it requires that the L\'{e}vy measure of the
High type dominates the L\'{e}vy measure of the Low type in a strong
sense. Roughly, jumps of any size $h$, both positive and negative,
occur more often (or at the same rate) under the High type than
under the Low type. A consequence of this assumption is that jumps
always provide good news, and (weakly) increase the posterior
probability of the High type.
\begin{rem}\label{moments}
Although we require that the zeroth and first moments of $(\nu_1
(dh) - \nu_0 (dh))$ are finite (Assumptions A3 and A4), this
requirement is not made for moments of higher order, since, by
Assumption \ref{finite_quadratic}, $\int_{\mathbb{R}\setminus \{0\}}
h^2\nu_\theta(dh)\leq \mu_\theta^2+\sigma^2 +
\int_{\mathbb{R}\setminus \{0\}}h^2\nu_\theta(dh) =
E[X^2_\theta(1)]<\infty$, for $\theta\in \{0,1\}$.
\end{rem}
\section{The posterior belief}
\label{s:posterior}
\subsection{Motivation}
At time $t=0$ the type $\theta$ is chosen randomly with $P(\theta=1)
= 1 - P(\theta=0) = p$. The DM does not observe $\theta$, but he
knows the prior $p$ and observes the controlled process
$(X(\K(t)))$. Let $p_t:=P(\theta=1\mid \F_{\K(t)})$ be the posterior
belief at time $t$ that the risky arm's type is High under the
allocation strategy $(\KK(t))$. The following proposition asserts
that the payoff $V_\KK (t,p)$ can be expressed solely by
the\footnote{$ p_{t-}$ is the posterior belief at time $K(t)-$.}
posterior process $(p_{t-})$ and the allocation strategy $(\KK(t))$.
This representation motivates the investigation of the posterior
process.
\begin{prop}\label{replacep}
For every allocation strategy $\KK$,
\begin{align}\notag
V_\KK(t,p) &= E^{t,p} \left[\int_t^{\infty} r e^{-rs} [(\mu_1 p_{s-}
+\mu_0 (1-p_{s-}) )\key(s) + \varrho(1-\key(s))] ds\right].
\\\notag
\end{align}
\end{prop}
\begin{proof}
We will prove the following series of equations, which proves the
claim:
\begin{align}\label{11}
E^{t,p}&\left[\int_t^\infty re^{-rs}dY(\KK(s))\right]= E^{t,p}\left[
\underset{x\rightarrow \infty}{\lim}\int_t^x
re^{-rs}dY(\KK(s))\right] \\\label{12} &= \underset{x\rightarrow
\infty}{\lim}E^{t,p}\left[\int_t^x re^{-rs}dY(\KK(s))\right]
\\\label{13} &= \underset{x\rightarrow \infty}{\lim}E^{t,p}\left[\int_t^x
re^{-rs}[(\mu_1 p_{s-} +\mu_0 (1-p_{s-}) )\key(s) +
\varrho(1-\key(s))] ds\right]\\\label{14} &=
E^{t,p}\left[\underset{x\rightarrow \infty}{\lim}\int_t^x
re^{-rs}[(\mu_1 p_{s-} +\mu_0 (1-p_{s-}) )\key(s) +
\varrho(1-\key(s))] ds\right]\\\label{15} &=
E^{t,p}\left[\int_t^\infty re^{-rs}[(\mu_1 p_{s-} +\mu_0 (1-p_{s-})
)\key(s) + \varrho(1-\key(s))] ds\right].\\\notag
\end{align}
Eqs.~(\ref{11}) and (\ref{15}) hold by the definition of the
improper integral. Let $\textit{[}X(\K(s))\textit{]}$ be the
quadratic variation of the time-changed process $(X(\K(s)))$. From
the It\={o} isometry and Kobayashi (2011, pages 797--799)
\cite{Kobayashi}, it follows that
\begin{align}\notag
&E^{t,p} \left[\int_t^\infty re^{-rs}dX(\K(s)) - \int_t^x
re^{-rs}dX(\K(s)) \right]^2\\\notag &= E^{t,p} \left[\int_x^\infty
re^{-rs}dX(\K(s)) \right]^2 = E^{t,p} \left[\int_x^\infty
(re^{-rs})^2 d\textit{[}X(\K(s))\textit{]} \right] \\\notag & =
E\left.\left[\int_x^\infty (re^{-rs})^2
d\textsl{[}X(\K(s))\textsl{]} \right|\theta = 1 \right] p + E
\left.\left[\int_x^\infty (re^{-rs})^2 d\textit{[}X(\K(s))\textit{]}
\right|\theta = 0 \right] (1-p) \\\notag &=E
\left.\left[\int_x^\infty (re^{-rs})^2 c_1 d\K(s) \right|\theta = 1
\right] p + E \left.\left[\int_x^\infty (re^{-rs})^2 c_0 d\K(s)
\right|\theta = 0 \right] (1-p), \\\notag
\end{align}
where $c_\theta =E[X^2_\theta(1)]= \mu^2_\theta +\sigma^2 +
\int_{\mathbb{R}\setminus \{0\}}h^2\nu_\theta(dh)$, for
$\theta\in\{0,1\}$. Hence, $\int_t^x re^{-rs}dX(\K(s))$ convergence
to $\int_t^\infty re^{-rs}dX(\K(s))$ in $L^2$ and Eq.~(\ref{12})
follows. Eq.~(\ref{13}) follows from Corollary \ref{martingales}
(part C1) in the appendix. Eq.~(\ref{14}) follows from the dominated
convergence theorem, since for every $x\geq t$,
\begin{align}\notag
\left| \int_t^x re^{-rs}[(\mu_1 p_{s-} +\mu_0 (1-p_{s-}) )\key(s) +
\varrho(1-\key(s))] ds\right|\leq \max\{|\mu_0|, |\mu_1|\} .
\end{align}
\end{proof}
\subsection{Formal definition of the posterior belief}\label{s:formal_definition}
An elegant formulation of the Bayesian belief updating process was
presented by Shiryaev (1978, Ch.~4.2) \cite{Shiryaev} to a model in
which the observed process is a Brownian motion with unknown drift
and extended later to a model in which the observed process is a
Poisson process with unknown rate in Peskir and Shiryaev (2000)
\cite{Peskir2000}.\footnote{Similar work has been done in the
disorder problem; see, e.g., Shiryaev (1978) \cite{Shiryaev}, Peskir
and Shiryaev (2002) \cite{Peskir2002}, and Gapeev (2005)
\cite{Gapeev}. } We follow this formulation and extend it to the
time-changed L\'{e}vy process. For every $p\in[0,1]$, the
probability measure $P_p$ satisfies $ P_p = p P_1+(1-p)P_0.$ An
important auxiliary process is the Radon--Nikodym density, given by
\begin{align}\notag
\varphi_t := \frac{d(P_0\mid\F_{\K(t)})}{d(P_1\mid \F_{\K(t)})},
\;\;t\in[0,\infty).
\end{align}
\begin{lem}
For every $t\in[0,\infty)$,
\begin{align}\notag
p_t = \frac{p}{p +(1-p)\varphi_t}.
\\\notag
\end{align}
\end{lem}
\begin{proof}
Define the following Radon--Nikodym density process
\begin{align}\notag
\pi_t = p\frac{d(P_1\mid\F_{\K(t)})}{d(P_p\mid\F_{\K(t)})},
\;\;t\in[0,\infty),
\end{align}
where $P_p(\cdot\mid\F_{\K(t)}) = p P_1(\cdot\mid\F_{\K(t)})+(1 - p)
P_0(\cdot\mid\F_{\K(t)})$. From the definition of $(\varphi_t)$ it
follows that $\pi_t = \frac{p}{p +(1-p)\varphi_t}$. Therefore, it is
left to prove that $p_t=\pi_t$ for every $t\in[0,\infty)$. Let
$A\in\F_{\K(s)}$ where $s\geq t$. The following series of equations
yields that $p_t=\pi_t$ for every $t\in[0,\infty)$:
\begin{align}\label{31}
E^p[\chi_A p_s|\F_{K(t)}] &= E^p[\chi_A
E^p[\chi_{\{\theta=1\}}|\F_{K(s)}]|\F_{K(t)}] \\\label{32}& =
E^p[\chi_{A\cap\{\theta=1\}} |\F_{K(t)}] \\\label{33}& = p
E^1[\chi_{A} |\F_{K(t)}] \\\label{34}& = E^p[\chi_{A}\pi_s
|\F_{K(t)}],
\end{align}
where $\chi_A=1$ if $A$ is satisfied and zero otherwise.
Eq.~(\ref{31}) follows from the definition of $p_t$. Eq.~(\ref{32})
follows since $s\geq t$, and, therefore,
$\F_{K(s)}\supseteq\F_{K(t)}$. Eq.~(\ref{33}) follows from the
definition of the probability measure $P_p$, and Eq.~(\ref{34})
follows from the property of the Radon--Nikodym density $\pi_t$.
\end{proof}
By Jacod and Shiryaev (1987, Ch.~III, Theorems 3.24 and 5.19)
\cite{Jacod}, the process $(\varphi_t)$ admits the following
representation:
\begin{align}\notag
\varphi_t =\exp \left\{ \beta\sigma Z(\K(t))+ (\vg-\vvg
-\frac{1}{2}\beta^2 \sigma^2) \K(t) + \underset{\mathbb{R}\setminus
\{0\}}{\int}\ln \left(\frac{\nu_0}{\nu_1}(h)\right)N(\K(t),dh)
\right\},\notag
\end{align}
where $\beta := \frac{\mu_0-\mu_1-\underset{\mathbb{R}\setminus
\{0\}}{\int}h(\nu_0-\nu_1)(dh)}{\sigma^2}$ and $\vg-\vvg :=
\underset{\mathbb{R}\setminus \{0\}}{\int}(\nu_1(dh)-\nu_0(dh))$. By
Assumption A6, $\vg-\vvg$ is finite and the Radon--Nikodym
derivative $\frac{\nu_0}{\nu_1}(h)$ exists.\footnote{To ensure the
existence of the Radon--Nikodym derivative one does not need the
full power of Assumption A6. Its full power will be used for the
proof of Theorem \ref{Up1}.}
\begin{rem}\label{Markovian}
1. Let $B_\infty\in \mathcal{B}(\mathbb{R}\setminus \{0\})$ be a
maximal set (up to $\nu_1 $-measure zero) such that
$\nu_1(B_\infty)\geq 0=\nu_0 (B_\infty)$. Occurrence of a jump from
$B_\infty$ indicates that the risky arm is High.
By definition, $\varphi_t=0$ after such a jump and therefore $p_t=1$.\\
2. By ignoring jumps from $B_\infty$, $(\ln(\varphi_t))$ is a
L\'{e}vy process\footnote{However, it is not a L\'{e}vy process
under $P_p$ for $0<p<1$, since it is not time-homogeneous.} with
time change $(\K(t))$, under both $P_0$ and $P_1$ with respect to
the filtration generated by $(\varphi_t)$, which coincides with
$(\F_{\K(t)})$. From the one-to-one correspondence between
$\varphi_t$ and $p_t$ it follows that $p_t$ is a Markov process.
Therefore, our optimal control problem falls in the scope of optimal
control of Markov processes. Hence, we can limit the allocation
strategies to \emph{Markovian allocation strategies}, or
equivalently to \emph{Markovian control strategies}, which we define
as follows.
\end{rem}
\begin{defn}
A control strategy $(\key(t,\omega))$ is \emph{Markovian} if it
depends solely on the Markovian process $(t,p_{t-})$. That is,
$\key(t,\omega) = \key(t,p_{t-})$. Denote by $\Upsilon_M$ the set of
all Markovian control strategies.
\end{defn}
\begin{rem}
A convenient way to understand the ``Girsanov style'' process
$(\varphi_t)$ is to examine the process $(X(t))$. We may assume that
\begin{align}\notag
&X(t)=\mu_1 t+\sigma Z(t) + \underset{\mathbb{R}\setminus
\{0\}}{\int}h \widetilde{N}_{\nu_1}(t,dh),
\end{align}
where, under $P_1$, $(Z(t))$ is a Brownian motion and the last term
is a purely discontinuous martingale. By the definition of $\beta$,
the same process can be represented as
\begin{align}\notag
&X(t)=\mu_0 t +\sigma (\beta\sigma t +Z(t)) +
\underset{\mathbb{R}\setminus \{0\}}{\int} h
\widetilde{N}_{\nu_0}(t,dh).
\end{align}
Under $P_0$ the process $(\beta\sigma t +Z(t))$ is a Brownian motion
and the last term is a purely discontinuous martingale. For
details, see Jacod and Shiryaev (1987, Ch.~III) \cite{Jacod}.
\end{rem}
\subsection{The infinitesimal operator}
An important tool in the proofs is the \emph{infinitesimal operator}
of the process $(t,p_t)$ with respect to the Markovian control
strategy $\key$, which we will calculate in this section. The
infinitesimal operator (or infinitesimal generator) of a stochastic
process is the stochastic analog of a partial derivative (see
{\O}ksendal, 2000). In this section we calculate the infinitesimal
operator of the process $(t,p_t)$ with respect to the Markovian
control strategy $\key$, which we will use in the proof of Theorem
\ref{Up1}. By It\={o}'s formula (see, e.g., Kobayashi (2011), pages
797--799) \cite{Kobayashi}, the posterior process $(p_t)$ solves the
following stochastic differential equation:
\begin{align}\label{dP1}
dp_t = & ~[\beta^2\sigma^2(1-p_{t-})^2
p_{t-}-(\vg-\vvg)p_{t-}(1-p_{t-})]d\K(t) \\\notag
& -p_{t-}(1-p_{t-})\beta\sigma dZ(\K(t)) \\\notag
& +p_{t-}(1-p_{t-})\underset{h\in{\mathbb{R}\setminus \{0\}}}{\int}
\frac{1-\frac{\nu_0}{\nu_1}(h)}{p_{t-}+\frac{\nu_0}{\nu_1}(h)(1-p_{t-})}
N(d\K(t),dh) \\\notag
= & ~ p_{t-}(1-p_{t-})\left[ -\beta dM(\K(t))-(\vg-\vvg)d\K(t) \right.\\\notag
& + \left.\underset{h\in{\mathbb{R}\setminus \{0\}}}{\int} \frac{1-\frac{\nu_0}{\nu_1}(h)}{p_{t-}+\frac{\nu_0}{\nu_1}(h)(1-p_{t-})}
N(d\K(t),dh) \right] ,\notag
\end{align}
where
\begin{align}\notag
M(\K(t)) &= X(\K(t)) -\underset{\mathbb{R}\setminus \{0\}}{\int} h
\widetilde{N}_{\nu_0}(\K(t),dh) - \mu_0\K(t)+\beta\sigma^2 \int_0^t
p_{s-} dK(s)\notag
\end{align}
is a martingale under $P_p$ with respect to $\F_{\K(t)}$; see
Corollary \ref{martingales} (part C2) in the appendix.
The first term on the right-hand side of ~(\ref{dP1}),
$-p_{t-}(1-p_{t-})\beta dM(\K(t))$, is the contribution of the
continuous part of the payoff process to the change in the belief,
while the second term, $- p_{t-}(1-p_{t-})(\vg-\vvg)d\K(t)$, is the
contribution of the fact that no jump occurred. This latter
contribution is negative due to Assumption A6. If a jump of size $h$
occurs during the interval $[t,t+dt)$, then the contribution of the
jump is $\Ph_t - p_{t-}$, where, $\Ph_t:=
\frac{p_{t-}\nu_1(dh)}{p_{t-}\nu_1(dh) + (1-p_{t-} ) \nu_0(dh)}$ is
the Bayesian update of the probability that the risky arm is High
given that a jump of size $h$ occurs. By Assumption A6, for every
$0< p<1 $ we have $P_p( p_t < \Ph_t)=1$.
To calculate the infinitesimal operator of the process $(t,p_t)$
with respect to the Markovian control strategy $\key$ we apply
It\={o}'s formula\footnote{$C^{1,2}$ is the set of all functions
$f:[0,\infty)\times[0,1]\rightarrow \mathbb{R}$, which are $C^1$ in
their first coordinate, and $C^2$ in their second coordinate.} for
$f(t,p)\in C^{1,2}([0,\infty)\times[0,1])$ and obtain
\begin{align}\label{ito_for_f}
f(t,p_t)=& f(0,p_0) + \int_0^t {f_t(s,p_{s-})ds}
+ \int_0^t {f_p (s,p_{s-})dp_s}\\\notag
&+ \frac{1}{2}\int_0^t {f_{pp}(s,p_{s-}) p_{s-}^2(1-p_{s-})^2\beta^2\sigma^2d\K(s)}\\\notag
&+\sum_{s\leq t} \left[ f(s,p_s)- f(s,p_{s-})-f_p(s,p_{s-}) \Delta(p_s) \right]\\\notag
=& f(0,p_0) + \int_0^t {f_t(s,p_{s-})ds}\\\notag
&- \int_0^t {f_p (s,p_{s-})[ (\vg-\vvg)p_{s-}(1-p_{s-}) ]d\K(s)}\\\notag
&+ \frac{1}{2}\int_0^t {f_{pp}(s,p_{s-}) p_{s-}^2(1-p_{s-})^2\beta^2\sigma^2d\K(s)}\\\notag
&+ \int_{s=0}^t \int_{h\in{\mathbb{R}\setminus \{0\}}} \left(f\left(s,\frac{p_{s-}}{p_{s-}+(1-p_{s-})\frac{\nu_0}{\nu_1}(h)}\right) -
f(s,p_{s-})\right)\\\notag
\cdot &(p_{s-}\nu_1 (dh) + (1-p_{s-})\nu_0 (dh))d\K(s) \\\notag
&- \int_0^t {f_{p}(s,p_{s-}) p_{s-}(1-p_{s-})\beta dM(\K(s))}\\\notag
&+ \underset{{s=0}}{\int^t} \int_{h\in{\mathbb{R}\setminus \{0\}}} \left(f\left(s,\frac{p_{s-}}{p_{s-}+(1-p_{s-})\frac{\nu_0}{\nu_1}(h)}\right) - f(s,p_{s-})\right) \\\notag
\cdot & [N(d\K(s),dh) - (p_{s-}\nu_1 (dh) + (1-p_{s-})\nu_0 (dh))d\K(s) ].\\\notag
\end{align}
The fifth and sixth terms on the right-hand side of
Eq.~(\ref{ito_for_f}) are stochastic integrals with respect to
martingales and therefore they are local martingales (see Jacod and
Shiryaev (1987, Ch.~I, Theorem 4.40) \cite{Jacod}). The seventh term
is a stochastic integral with respect to a compensated random
measure, as will be shown in Corollary \ref{martingales} (parts C2
and C3) in the appendix. Therefore, it is a local martingale (see
Jacod and Shiryaev (1987, Ch.~II, Theorem 1.8) \cite{Jacod}). Hence,
by taking expectations of both sides it follows that the
infinitesimal operator of the process $(t,p_t)$ with respect to the
Markovian control strategy $(\key(t,p))$ is
\begin{align}\notag
(\mathbb{L}^\key f)(t,p) =& f_t(t,p)
-(\vg-\vvg)p(1-p)f_p (t,p)\key(t,p)
+\frac{1}{2} \beta^2\sigma^2 f_{pp}(t,p) p^2(1-p)^2\key(t,p)\\\notag
&+\int_{\mathbb{R}\setminus \{0\}} { \left(f\left(t,\frac{p}{p+(1-p)\frac{\nu_0}{\nu_1}(h)}\right) - f(t,p)\right)
(p \nu_1 (dh) + (1-p)\nu_0 (dh))\key(t,p) }.\\\notag
\end{align}
When $f$ is a function of $p$ only, we will use the same notation
for the infinitesimal operator of the process $(p_t)$ with respect
to the time-homogeneous Markovian control strategy $\key(p)$.
Specifically,
\begin{align}\label{Lkp}
(\mathbb{L}^\key f)(p) =& -(\vg-\vvg)p(1-p)f' (p)\key(p)
+\frac{1}{2} \beta^2\sigma^2 f''(p) p^2(1-p)^2\key(p)\\\notag
&+\int_{\mathbb{R}\setminus \{0\}} { \left(f\left(\frac{p}{p+(1-p)\frac{\nu_0}{\nu_1}(h)}\right) - f(p)\right)
(p \nu_1 (dh) + (1-p)\nu_0 (dh))\key(p) }.\\\notag
\end{align}
\section{The value function}
\label{s:value} In the next section we will introduce the
Hamilton--Jacobi--Bellman (HJB) for our problem. The value function
$U(t,p)$ is not $C^2$ in its second coordinate, and therefore we
need to formalize the optimal problem differently. Additionally to
the Markovian control strategy $\key(t,p)$, we will add an
artificial stopping time $\tau$ to the new strategy space. This new
form will help us solve the HJB although $U(t,p)$ is not $C^2$. We
start with a few basic properties of the value function $U(t,p)$.
\begin{prop}\label{convex}
For every fixed $t\geq 0$, the function $p\mapsto U(t,p)$ is
monotone, nondecreasing, convex, and continuous.
\end{prop}
\begin{proof}
Fix for a moment an allocation strategy $\KK$. By Definition
\ref{Etp} and Eq.~(\ref{VKtp}) the expected discounted payoff from
time $t$ onwards under strategy $\KK$ when $p_t =p$ is
\begin{align}\notag
V_\KK(t,p) &= E^{t,p} \left[\int_t^{\infty} r e^{-rs}
dY(\KK(s))\right]\\\notag
&= \left.p E\left[\int_t^{\infty} r e^{-rs} dY(\KK(s))\right| \theta =
1\right]+\left.(1-p) E \left[\int_t^{\infty} r e^{-rs}
dY(\KK(s))\right| \theta = 0\right].\\\notag
\end{align}
For every fixed $t\geq 0$ the function $p\mapsto V_{\KK}(t,p)$ is
linear. Therefore $U(t,p)$, as the supremum of linear functions, is
convex. By always choosing the safe arm, the DM can achieve at least
$e^{-rt}\varrho$, and by always choosing the risky arm the DM can
achieve at least $ e^{-rt} (p \mu_1 + (1-p) \mu_0) $. Since
$U(t,0)=e^{-rt}\varrho$ and $U(t,1) = e^{-rt}\mu_1$, the convexity
of $U(t,p)$ implies that the function $p \mapsto U(t,p)$ is
continuous and nondecreasing in $p$.
\end{proof}
It follows from Proposition \ref{convex} that for every fixed $t\geq
0$ there is a time-dependent cut-off $p^*_t$ in $[0,1]$ such that
$U(t,p)=\varrho$ if $p\leq p^*_t$ and $U(t,p)>\varrho$ otherwise. It
follows that for every fixed $t$ the strategy $\key (t,p)\equiv 0$
that always chooses the safe arm is optimal for prior beliefs in
$[0,p^*_t]$. By this conclusion, Proposition \ref{replacep}, and
Remark \ref{key} we deduce that the optimal problem (\ref{defineU})
can be reduced to a combined optimal stopping and stochastic control
problem as follows:
\begin{align}\label{UwithW}
U(t,p) = \underset{ t\leq \tau ,\; \key\in\Upsilon_M }{\sup}
E^{t,p} \left[\int_t^\tau r e^{-rs} W(p_{s-},\key(s,p_{s-}))ds +
\varrho e^{-r\tau}\right],
\end{align}
where $W(p,l):=(\mu_1 p +\mu_0 (1-p))l + \varrho(1-l)$ is the
instantaneous payoff given the posterior $p$, using the Markovian
control $l$. This representation of the value function will help us
solve the HJB equation. Denote the continuation region to be
\begin{align}\notag
D:=\{(t,p) \mid U(t,p)> \varrho e^{-rt}\}.
\end{align}
This is the region where the optimal action of the DM is to continue
(that is, $k(t,p)>0$, and $\tau>t$). The next lemma shows that the
region $D$ is invariant with respect to $t$. This means that the
optimal stopping time $\tau$ (whenever it exists) does not
depend\footnote{In fact, we will show in Theorem \ref{Up1} that an
optimal stopping time and an optimal control strategy do exist and
the optimal control $\key$ is also time-homogeneous; that is, $\key$
does not depend on $t$, and therefore the allocation strategy $\KK$
does not depend on $t$ either. } on $t$.
\begin{lem}
For every $t\geq 0$ and every $p\in[0,1]$ one has
$U(t,p)=e^{-rt}U(0,p)$. In particular, $(t,p)\in D$ if and only if
$(s,p)\in D$, for every $t,s\geq 0$ and every $p\in[0,1]$.
\end{lem}
\begin{proof}
The first claim follows from the following list of equalities:
\begin{align}\label{Utp_U0p}
U(t,p) &= \underset{ t\leq \tau ,\; 0 \leq \key \leq 1 }{\sup}
E^{t,p} \left[\int_t^\tau r e^{-rs} W(p_{s-},\key(s,p_{s-}))ds +
\varrho e^{-r\tau}\right]
\\\notag
&= \underset{ 0\leq {\tilde{\tau}} ,\; 0 \leq \key \leq 1 }{\sup}
E \left[\int_0^{\tilde{\tau}} r e^{-r(t+u)}
W(p_{u-}^p,\key(t+u,p_{u-}^p))du + \varrho
e^{-r(t+{\tilde{\tau}})}\right]
\\\notag
&= e^{-rt}\underset{ 0\leq \tilde{\tau} ,\; 0 \leq \key \leq 1 }{\sup}
E^{0,p} \left[\int_0^{\tilde{\tau}} r e^{-ru}
W(p_{u-},\key(t+u,p_{u-}))du + \varrho e^{-r{\tilde{\tau}}}\right]
\\\notag
&=e^{-rt}U(0,p),\\\notag
\end{align}
where the second equality follows from the Markovian property of
$p_t$ (see Remark \ref{Markovian}).
\end{proof}
This lemma yields that the cut-off $p^*_t$ discussed earlier is
independent of $t$. We therefore denote it by $p^*$.
\section{The HJB equation}
\label{s:HJB} The following proposition introduces the HJB equation
for our problem.
\begin{prop}
Let $F\in C^{1}[0,1] $ be a function that satisfies
\begin{align}
F(p)\geq \varrho \;\text{for every}\;p\in[0,1].\label{Fp_varrho}
\end{align}
Define the continuation region of $F$ by
\begin{align}
C:=\{ p\in[0,1] \mid F(p)> \varrho \}.\label{continuation_region}
\end{align}
Suppose that
\begin{align}\label{C_D}
&[0,\infty)\times C=D.\\\label{C2} &F\in C^2([0,1] \backslash
\partial C) \;\; \text{with locally bounded derivatives near} \;
\partial C.\\\label{HJB_leq} &\mathbb{L}^\key F(p)+rW(p,\key(p))-rF(p)\leq
0 \; \text{on} \; [0,1]\backslash\partial C \\\notag &\text{for all
} k\in \Upsilon_M,\; \text{and all}\; p\in[0,1].\\\label{HJB_equal}
&\text{There is a control $\key^*$ for which the inequality in
Eq.~(\ref{HJB_leq})}\\\notag &\text{ holds with equality.}\\\notag
\end{align}
Then, $\key^*$ is the optimal control, $\tau_D:=\inf \{t\geq 0 \mid
F(p_t)\not\in C\}$ is the optimal stopping time, and
$U(t,p)=e^{-rt}F(p)$.
\end{prop}
Conditions (\ref{HJB_leq}) and (\ref{HJB_equal}) represent the HJB
equation in our model.
\begin{rem}
The function $F$ need not be $C^2$ at the boundary of $C$. This is
due to the representation of $U(t,p)$ in Eq.~(\ref{UwithW}) as a
combined optimal stopping and stochastic control problem. This issue
will be further discussed in the proof.
\end{rem}
\begin{proof}
Define $J(t,p):=e^{-rt}F(p)$. Then for every $(t,p)\in
[0,\infty)\times([0,1]\setminus\partial C)$,
\begin{align}\label{LJLF}
\mathbb{L}^kJ(s,p)=-re^{-rt}F(p) + e^{-rt}\mathbb{L}^kF(p).\\\notag
\end{align}
By Eq.~(\ref{C2}), $J(t,p)\in
C^{1,2}([0,\infty)\times([0,1]\setminus\partial C))$ and
$J_{pp}(s,p)$ is bounded near $[0,\infty)\times\partial C$.
Therefore, there exists a sequence $\{J^n\}_{n\geq 1}\subseteq
C^{1,2}(D)$ such that
\begin{align}\notag
J^n\rightarrow J,\;\;J^n_t\rightarrow J_t,\;\;J^n_p\rightarrow
J_p,\;\;J^n_{pp}\rightarrow J_{pp}
\end{align}
uniformly on every compact subset of
$[0,\infty)\times([0,1]\setminus\partial C)$ as $n$ goes to infinity
(see {\O}ksendal (2000, Theorem C.1) \cite{Oksendal}). Denote by
$L(t)$ the sum of the last three terms on the right-hand side of
Eq.~(\ref{ito_for_f}). The process $(L(t))$, as the sum of local
martingales, is a local martingale. Let $(\delta_m)$ be a sequence
of increasing (a.s.) stopping times that diverge (a.s.), such that
$L(\delta_m \wedge t)$ is\footnote{$a\wedge b:= \min{\{a,b\}}$.} a
martingale for every $m$. Let $\tau$ be an arbitrary stopping time
and define $\tau_m:=\tau\wedge m \wedge\delta_m$. We will prove the
following series of equations:
\begin{align}\notag
E&^{0,p}\left[ e^{-r\tau_m}F(p_{\tau_m}) \right]-F(p)\\\label{21} &=
E^{0,p}\left[ J(\tau_m,p_{\tau_m}) \right]- J(0,p)
\\\label{22}
&=
\underset{n\rightarrow\infty}{\lim}\left(E^{0,p}\left[J^n(\tau_m,p_{\tau_m})
\right]- J^n(0,p)\right)\\ \label{23}&=
\underset{n\rightarrow\infty}{\lim}E^{0,p}\left[\int_0^{\tau_m}\mathbb{L}^kJ^n(s,p_{s-})ds
\right]\\\label{24} &=
\underset{n\rightarrow\infty}{\lim}E^{0,p}\left[\int_0^{\tau_m}\mathbb{L}^kJ^n(s,p_{s-})\chi_{\{
p_{s-}\not\in\partial C\}}ds \right] \\ \label{25} &=
E^{0,p}\left[\int_0^{\tau_m}\mathbb{L}^kJ(s,p_{s-})\chi_{\{
p_{s-}\not\in\partial C\}}ds \right]\\ \label{26} &= E^{0,p}\left[
\int_0^{\tau_m}{e^{-rs}\left(-rF(p_{s-})+
\mathbb{L}^kF(p_{s-})\right)\chi_{\{ p_{s-}\not\in\partial C\}}ds}
\right]\\ \label{27} &{\leq} - E^{0,p}\left[
\int_0^{\tau_m}e^{-rs}W(p_{s-},\key(p_{s-})) \chi_{\{
p_{s-}\not\in\partial C\}}ds \right]\\ \label{28} &= - E^{0,p}\left[
\int_0^{\tau_m}e^{-rs}W(p_{s-},\key(p_{s-})) ds \right],\\\notag
\end{align}
where, $\chi_A=1$ if $A$ is satisfied and zero otherwise.
Eq.~(\ref{21}) follows from the definition of $J(t,p)$.
Eqs.~(\ref{22}) and (\ref{25}) follow from the choice of the
sequence $J^n$. Eq.~(\ref{23}) follows from the definition of the
infinitesimal operator. By condition (\ref{C_D}), the boundary
$\partial C$ of C is a single point (specifically, it is the cut-off
point $\p1*$). Therefore, Eqs.~(\ref{24}) and (\ref{28}) follow from
Lemma \ref{spends} in the appendix.
Eq.~(\ref{26}) follows from Eq.~(\ref{LJLF}), and inequality
(\ref{27}) follows from (\ref{HJB_leq}).
By Eq. (\ref{Fp_varrho}) and the series of
Eqs.~(\ref{21})--(\ref{28}) we obtain
\begin{align}\notag
&E^{0,p}\left[ \int_0^{\tau_m}{e^{-rs}W(p_{s-},\key(p_{s-})) ds} +
\varrho e^{-r\tau_m} \right] \\\notag &\leq E^{0,p}\left[
\int_0^{\tau_m}{e^{-rs}W(p_{s-},\key(p_{s-})) ds} + e^{-r\tau_m}
F(p_{\tau_m}) \right]\leq F(p).
\end{align}
By taking $m\rightarrow\infty$ we deduce that
\begin{align}\label{F_geq_W}
E^{0,p}\left[ \int_0^{\tau}{e^{-rs}W(p_{s-},\key(p_{s-})) ds} +
\varrho e^{-r\tau} \right]\leq F(p).
\end{align}
The left-hand side of Eq.~(\ref{F_geq_W}) is the payoff of the DM
using the stopping time $\tau$ and the stationary Markovian control
strategy $k$. By taking the supremum in Eq.~(\ref{F_geq_W}) it
follows that $U(0,p)\leq F(p)$ for every $0\leq p\leq 1$. To prove
the opposite inequality, apply the argument above to the stationary
Markovian control strategy $\key^*=\key^*(p_{s-})$ and the stopping
time $\tau_D$, so that the inequality in Eq.~(\ref{27}) is replaced
by an equality. By taking the limit $m\rightarrow\infty$ and by the
definition of $D$ we obtain
\begin{align}\notag
U(0,p)&\geq E^{0,p}\left[
\int_0^{\tau_D}{e^{-rs}W(p_{s-},\key^*(p_{s-})) ds} + e^{-r\tau_D}
\varrho \right]\\\notag &= E^{0,p}\left[
\int_0^{\tau_D}{e^{-rs}W(p_{s-},\key^*(p_{s-})) ds} + e^{-r\tau_D}
F(p_{\tau_D}) \right]= F(p).
\end{align}
\end{proof}
\section{The optimal strategy}
\label{s:main} In this section we present our main result that
states that there is a unique optimal allocation strategy, and that
it is a cut-off strategy. The theorem also provides the exact
cut-off point and the corresponding expected payoff in terms of the
data of the problem. Let $\alpha^*$ be the unique solution in
$(0,\infty)$ of the equation $f(\alpha)=0$, where
\begin{equation}\label{eta}
f(\alpha) := \int\left(\left(\frac{\nu_0}{\nu_1}(h)\right)^{\alpha}
-1\right)\vvh + \alpha (\vg-\vvg) +\frac{1}{2}(\alpha+1)\alpha
\left(\dfrac{\mu_1-\mu_0}{\sigma}\right)^2 - r = 0.
\end{equation}
The existence and the uniqueness of such a solution are proved in
Lemma \ref{alpha} in the appendix.
\begin{thm}\label{Up1}
Denote $p^*:= \frac{\alpha^*
(\varrho-\mu_0)}{(\alpha^*+1)(\mu_1-\varrho) +
\alpha^*(\varrho-\mu_0)}$. Under Assumptions A1--A6 there is a
unique optimal strategy $\key^*$ that is time-homogeneous and is
given by
\begin{align}\label{kstar}
&\key^*=
\begin{cases}
0 &\text{if $p \leq p^*$}, \\
1 &\text{if $p > p^*$}.
\end{cases}
\end{align}
The expected payoff under $\key^*$ is
\begin{equation}\label{E:Up1}
U(0,p) = V_{\key^*} (0,p) =
\begin{cases}
\varrho &\text{if $p \leq p^*$}, \\
p\mu_1 +(1-p)\mu_0 +C_{\alpha^*} (1-p)(\frac{1-p}{p})^{\alpha^*}
&\text{if $p
> p^*$},
\end{cases}
\end{equation}
$\\$
where $C_{\alpha^*} = \frac{\varrho-\mu_0 - p^* (\mu_1 -\mu_0)}{(1-p^*) \left( \frac{1-p^*}{p^*}\right)^{\alpha^*}}.$\\
\end{thm}
We now discuss the relation between Theorem \ref{Up1} and the
results of Bolton and Harris (1999) \cite{Bolton}, Keller, Rady, and
Cripps (2005) \cite{Keller2005}, and Keller and Rady (2010)
\cite{Keller2010}. The expected payoff from the risky arm if no
information is available is
\begin{align}\notag
\int_0^\infty re^{-rs} E[X(s)]ds =\int_0^\infty re^{-rs} [p\mu_1
+(1-p)\mu_0]s ds = p\mu_1 +(1-p)\mu_0 .
\end{align}
One can verify that the strategy $\key^* \equiv 1$ and the function
$F(p) =p\mu_1 +(1-p)\mu_0$ satisfy Condition (\ref{HJB_equal}), and
following the results of Bolton and Harris (1999) \cite{Bolton} and
Keller, Rady, and Cripps (2005) \cite{Keller2005}, one can ``guess"
that a function of the form $C(1-p)(\frac{1-p}{p})^{\alpha}$
satisfies Condition (\ref{HJB_equal}) as well. This leads to the
form of the optimal payoff that appears in Eq.~(\ref{E:Up1}). The
function $C_{\alpha^*} (1-p)(\frac{1-p}{p})^{\alpha^*}$ is the
option value for the ability to switch to the safe arm. The
parameters of the payoff processes that determine the cut-off point
$p^*$ and the optimal payoff $U(0,p)$ are the expected payoffs
$\mu_1$ and $\mu_0$. In Bolton and Harris (1999) \cite{Bolton} the
only component in the risky arm is the Brownian motion with drift.
Therefore, $\nu_i \equiv 0$, so that $ \alpha^* =
(-1+\sqrt{1+8r\sigma^2 / (\mu_1 -\mu_0)^2})/2$.
In Keller, Rady, and Cripps (2005) \cite{Keller2005}, the risky arm
is either the constant zero (Low type, so that $\vvi \equiv 0$), or
it yields a payoff $\bar{h}$ according to a Poisson process of rate
$\lambda$ (High type). If the risky arm is High, then the only
component in the L\'{e}vy--It\={o} decomposition is the purely
discontinuous component, and $\vi (\bar{h}) = \lambda$ and zero
otherwise. Therefore, $\mu_1 = \lambda \bar{h}, \; \mu_0 = 0,$ and
$\alpha^* = r/\lambda$.
In Keller and Rady (2010) \cite{Keller2010}, the risky arm yields a
payoff $\bar{h}$ according to a Poisson process. For the High type,
the Poisson process rate is $\lambda_{\high}$, and for the Low type
the rate is $\lambda_{\low}$, where
$\lambda_{\low}<\lambda_{\high}$. The only component in the
L\'{e}vy--It\={o} decomposition is the purely discontinuous
component, and $\nu_i (\bar{h})=\lambda_i$ and zero otherwise.
Therefore, $\mu_1 = \lambda_{\high} \bar{h}, \; \mu_0 =
\lambda_{\low} \bar{h},$ and $\alpha^*$ is the unique solution of
the equation
\begin{equation}\notag
f(\alpha) :=
\lambda_{\low}\left(\frac{\lambda_{\low}}{\lambda_{\high}}\right)^{\alpha}
+ \alpha (\lambda_{\high}-\lambda_{\low})-\lambda_{\low} - r = 0.
\end{equation}\\
\begin{proof}[\textbf{Proof of Theorem \ref{Up1}}]
Let $\p1*$, $\alpha^*$, and $C_\alpha^*$ be the parameters that were
defined in the theorem. Define the cut-off strategy $\key^*$
associated with the cut-off $\p1*$ by
\begin{equation}\key^* :=
\begin{cases}
0 &\text{if $p \leq p^*$}, \\
1 &\text{if $p > p^*$},
\end{cases}
\end{equation}
$\\$ and the function $F$ by
\begin{equation}\label{F1}
F(p) :=
\begin{cases}
\varrho &\text{if $p \leq p^*$}, \\
F_1(p) &\text{if $p >
p^*$},
\end{cases}
\end{equation}
$\\$ where $F_1(p):= p\mu_1 +(1-p)\mu_0 +C_{\alpha^*}
(1-p)(\frac{1-p}{p})^{\alpha^*}$. We will show that the function
$F(p)$ and the cut-off strategy $\key^*$ are the optimal payoff and
the optimal Markovian control strategy respectively. To this end, we
verify that conditions (\ref{Fp_varrho})--(\ref{HJB_equal}) are
satisfied for $F(p)$ and $\key^*$. Conditions
(\ref{Fp_varrho})--(\ref{C2}) can be easily verified for the
function $F(p)$. To verify conditions (\ref{HJB_leq}) and
(\ref{HJB_equal}) we need the full power of Assumption A6. To prove
that $F(p)$ satisfies Condition (\ref{HJB_leq}) we check separately
the cases $p\leq \p1*$ and $p>\p1*$. Fix a Markovian control
strategy $k\in\Upsilon_M$ and $p\leq\p1*$. Then,
\begin{align}\notag
\mathbb{L}^\key & F(p)+rW(p,\key(p))-rF(p) \\\label{41} =&
-(\vg-\vvg)p(1-p)F'(p)\key(p) +\frac{1}{2} \beta^2\sigma^2 F''(p)
p^2(1-p)^2\key(p)\\\notag &+ \underset{\mathbb{R}\setminus
\{0\}}{\int} {
\left(F\left(\frac{p}{p+(1-p)\frac{\nu_0}{\nu_1}(h)}\right) -
F(p)\right) (p \nu_1 (dh) + (1-p)\nu_0 (dh))\key(p)}\\\notag &
+r[(\mu_1 p +\mu_0
(1-p))\key(p)+\varrho(1-\key(p))]-rF(p)\\\label{42}=&
\underset{G}{\int} {
\left(F_1\left(\frac{p}{p+(1-p)\frac{\nu_0}{\nu_1}(h)}\right) -
\varrho\right) (p \nu_1 (dh) + (1-p)\nu_0 (dh))\key(p)}\\\notag &
+r[(\mu_1 p +\mu_0
(1-p))\key(p)+\varrho(1-\key(p))]-r\varrho\\\label{43} \leq
&\underset{G}{\int} {
\left(F_1\left(\frac{\p1*}{\p1*+(1-\p1*)\frac{\nu_0}{\nu_1}(h)}\right)
- \varrho\right) (\p1* \nu_1 (dh) + (1-\p1*)\nu_0
(dh))\key(p)}\\\notag & +r[(\mu_1 \p1* +\mu_0
(1-\p1*))\key(p)+\varrho(1-\key(p))]-r\varrho\\\label{44} \leq &
\underset{\mathbb{R}\setminus \{0\}}{\int} {
\left(F_1\left(\frac{\p1*}{\p1*+(1-\p1*)\frac{\nu_0}{\nu_1}(h)}\right)
- \varrho\right) (\p1* \nu_1 (dh) + (1-\p1*)\nu_0
(dh))\key(p)}\\\notag & +r[(\mu_1 \p1* +\mu_0
(1-\p1*))\key(p)+\varrho(1-\key(p))]-r\varrho\\\label{45} = &\, 0,
\end{align}
where $G=\left\{ h \mid \frac{p}{p+(1-p)\frac{\nu_0}{\nu_1}(h) }>
\p1*\right\}$. Eq.~(\ref{41}) follows from Eq.~(\ref{Lkp}).
Eq.~(\ref{42}) follows since for $p<\p1*$ we have $F(p)=\varrho$,
$F'(p)=F''(p)=0$, and for $p>\p1*$ we have $F(p)=F_1(p)$. Inequality
(\ref{43}) follows from the monotonicity of $F_1$ and Assumptions A5
and A6. Inequality ~(\ref{44}) follows since $F_1(q)>\varrho$ for
every $q>\p1*$. Eq.~(\ref{45}) is satisfied for every $\key(p)$ by
the definition of $\p1*$ and $F_1$.
Fix a Markovian control strategy $k\in\Upsilon_M$ and $p > \p1*$.
Then $F(p)=F_1(p)$ and
$F\left(\frac{p}{p+(1-p)\frac{\nu_0}{\nu_1}(h)}\right)
=F_1\left(\frac{p}{p+(1-p)\frac{\nu_0}{\nu_1}(h)}\right) $, since by
Assumption A6, for every $0\leq p<1$ we have
$P_p\left(\frac{p}{p+(1-p)\frac{\nu_0}{\nu_1}(h)}
> p\right) = 1$. Therefore,
\begin{align}\notag
\mathbb{L}^\key & F(p)+rW(p,\key(p))-rF(p) \\\label{51} =&
-(\vg-\vvg)p(1-p)F'(p)\key(p) +\frac{1}{2} \beta^2\sigma^2 F''(p)
p^2(1-p)^2\key(p)\\\notag &+ \underset{\mathbb{R}\setminus
\{0\}}{\int} {
\left(F\left(\frac{p}{p+(1-p)\frac{\nu_0}{\nu_1}(h)}\right) -
F(p)\right) (p \nu_1 (dh) + (1-p)\nu_0 (dh))\key(p)}\\\notag &
+r[(\mu_1 p +\mu_0
(1-p))\key(p)+\varrho(1-\key(p))]-rF(p)\\\label{52} =&
-(\vg-\vvg)p(1-p)F_1'(p)\key(p) +\frac{1}{2} \beta^2\sigma^2
F_1''(p) p^2(1-p)^2\key(p)\\\notag &+ \underset{\mathbb{R}\setminus
\{0\}}{\int} {
\left(F_1\left(\frac{p}{p+(1-p)\frac{\nu_0}{\nu_1}(h)}\right) -
F_1(p)\right) (p \nu_1 (dh) + (1-p)\nu_0 (dh))\key(p)}\\\notag &
+r[(\mu_1 p +\mu_0
(1-p))\key(p)+\varrho(1-\key(p))]-rF_1(p)\\\label{53} \leq & \, 0,
\end{align}
where the last inequality is satisfied for every $\key(p)$ by the
definition of $F_1$.
By the definition of the Markovian control strategy $\key^*$, for
every $p\leq \p1*$ we have $\key^*(p)=0$ and therefore
Eqs.~(\ref{41})--(\ref{45}) hold with equality, and for every
$p>\p1*$ we have $\key^*(p)=1$ and therefore
Eqs.~(\ref{51})--(\ref{53}) hold with equality by the definition of
$F(p)$. This proves condition (\ref{HJB_equal}).
Without Assumption A6 there may be a set $B$ that satisfies
$\nu_0(B)>0$, such that for every $h\in B$ and every $0\leq p<1$ one
has $P_p\left(\frac{p}{p+(1-p)\frac{\nu_0}{\nu_1}(h)} < p\right) =
1$. Thus, for every $p\in\left[p^*,
\frac{p^*}{p^*+(1-p^*)\frac{\nu_0}{\nu_1}(h)} \right)$ we need to
substitute $F_1(p)$ for $F(p)$, and $\varrho$ for
$F\left(\frac{p}{p+(1-p)\frac{\nu_0}{\nu_1}(h)}\right)$. This
problem has a higher level of complexity and it is not clear how to
approach it using the tools introduced here.
\begin{rem}
Since the process $(p_t)$ has no negative jumps, it enters the
interval $[0,\p1*]$ continuously. Therefore, we expect the value
function to be $C^1$ at the cut-off point $\p1*$. In a model where
the process $(p_t)$ has negative jumps, it can enter the interval
$[0,\p1*]$ with a jump. In this case we expect that the value
function will not be $C^1$ at the cut-off point $\p1*$. For simple
cases of L\'{e}vy processes (such as when the High (resp. Low) type
is a jump process with height $h_1$ and rate $\lambda_1$ (resp.
height $h_0$ and rate $\lambda_0$), where
$h_1\lambda_1>\varrho>h_0\lambda_0$ and $\lambda_1<\lambda_0$, so,
in particular, Assumption A3 fails) the method introduced in Peskir
and Shiryaev (2000) \cite{Peskir2000} may be useful to characterize
the optimal strategy and the value function. In the general setup, a
sample path method may be helpful to approximate the value function
via iterations (see Dayanik and Sezer (2006) \cite{Dayanik}). This
investigation is left for future research.
\end{rem}
\end{proof}
\subsection{Comparative statics}\label{Comparative Statics}
The explicit forms of the cut-off point $p^*$ and the value function
$U$ allow us to derive simple comparative statics of these
quantities. As is well known, a DM who plays optimally switches to
the safe arm later than a myopic DM, and indeed $p^*$ is smaller
than the myopic cut-off point
$p^m:=\frac{\varrho-\mu_0}{\mu_1-\mu_0}$.
Note that the cut-off point $p^*$ is an increasing function of
$\alpha^*$. As can be expected, $\alpha^*$ (and therefore also
$p^*$) increases with the discount rate $r$ and with $\vvi (dh)$,
and decreases with $\vi (dh)$ and with $\mu_1-\mu_0$. That is, the
DM switches to the safe arm earlier in the game as the discount rate
increases, as jumps provide less information, or as the difference
between the drifts of the two types increases.\footnote{Moreover,
$\alpha^*(r=0) = 0$ and $\alpha^*(r=\infty) = \infty$. }
Furthermore, as long as $p>p^*$ the value function $p \mapsto
U(0,p)$ decreases in $\alpha^*$. Thus, decreasing the discount rate,
increasing the informativeness of the
jumps, or increasing the difference between the drifts is beneficial to the DM.\\\\
\subsection{Generalization}
In our model the parameters $\mu_0$ and $\mu_1$ have two roles. By
the definition of the L\'{e}vy process $(X(t))$ they play the role
of the unknown drift. In Eq.~(\ref{UwithW}) they determine the
instantaneous expected payoff. Here we separate these two roles;
that is, we assume that the parameters that determine the
instantaneous expected payoff are not $\mu_0$ and $\mu_1$, but
rather two other parameters, $g_0$ and $g_1$ respectively. Formally,
in the definition of $W(p,l)$ in Eq.~(\ref{UwithW}) we substitute
$\mu_0$ and $\mu_1$ with other parameters, $g_0$ and $g_1$, and
observe the change in the optimal strategy and the optimal payoff.
This formulation allows us to separate the information-relevant
parameters from the payoff-relevant parameters. It also supplies an
optimal strategy and an optimal payoff in a model where the DM
receives a general linear function of the process $(X(t))$.
If we replace $W(p,l)=(\mu_1 p +\mu_0 (1-p))l+\varrho(1-l)$ with
$\widehat{W}(p,l)=(g_1 p+ g_0 (1-p))l + \varrho(1-l),$ where $g_0$
and $g_1$ are constants that satisfy $g_1>\varrho>g_0$, then the
solution of the optimization problem
\begin{align}\notag
\widehat{U}(t,p) = \underset{ t\leq \tau ,\; \key\in\Upsilon_M
}{\sup} E^{t,p} \left[\int_t^\tau r e^{-rs}
\widehat{W}(p_{s-},\key(s,p_{s-}))ds + \varrho e^{-r\tau}\right]
\end{align}
has a similar form to the one given in Theorem \ref{Up1}. Denote
$\widehat{p}^*:= \frac{\widehat{\alpha}^*
(\varrho-g_0)}{(\widehat{\alpha}^*+1)(g_1-\varrho) +
\widehat{\alpha}^*(\varrho-g_0)}$, where $\widehat{\alpha}^* =
\alpha^*$. Under Assumptions A1--A6, there is a unique optimal
strategy that is time-homogeneous and is given by
$$\widehat{\key}^* =
\begin{cases}
0 &\text{if $p \leq \widehat{p}^*$}, \\
1 &\text{if $p > \widehat{p}^*$}.
\end{cases}
$$
The expected payoff under $\widehat{\key}^*$ is
\begin{equation}\notag
\widehat{U}(0,p) = \widehat{V}_{\widehat{\key}^*} (0,p) =
\begin{cases}
\varrho &\text{if $p \leq \widehat{p}^*$}, \\
pg_1 +(1-p)g_0 +\widehat{C}_{\alpha^*}
(1-p)(\frac{1-p}{p})^{\alpha^*} &\text{if $p
> \widehat{p}^*$},
\end{cases}
\end{equation}
$\\$ where $\widehat{C}_{\alpha^*} = \frac{\varrho-g_0 -
\widehat{p}^* (g_1 -g_0)}{(1-\widehat{p}^*)
\left( \frac{1-\widehat{p}^*}{\widehat{p}^*}\right)^{\alpha^*}}.$\\
The significance of this result is that it separates the
information-relevant parameters of the model from the
payoff-relevant parameters. The quantity $\alpha^*$ summarizes all
the information-relevant parameters, whereas $g_1$ and $g_0$ are the
only payoff-relevant parameters. For beliefs above the cut-off, the
optimal payoff is the sum of the expected payoff if the DM always
continues, $pg_1 +(1-p)g_0$, and the option value of
experimentation, which is given by $\widehat{C}_{\alpha^*}
(1-p)(\frac{1-p}{p})^{\alpha^*}$.
\section{APPENDIX}
The following lemma states that a time-changed martingale under an
allocation strategy remains a martingale.
\begin{lem}\label{lem:time_change}
Let $(M(t))$ be a martingale with respect to $\F_t$, and let
$(\KK(t))$ be an allocation strategy that satisfies (\ref{T1}),
(\ref{T2}), and (\ref{T3}). Then $(M(\K(t)))$ is a martingale with
respect to $\F_{\K(t)}$.
\end{lem}
\begin{proof}
Fix $0\leq s\leq t$. Then $\K(s)$ and $\K(t)$ are bounded stopping
times with $\K(s)\leq \K(t)$. The \textit{optional stopping theorem}
implies that $M(K(s)) = E[M(K(t))\mid\F_{\K(s)}]$, and therefore,
$(M(\K(t)))$ is indeed an $(\F_{\K(t)}, P)$-martingale.
\end{proof}
The following lemma presents the (predictable) compensator of a
process under a change of measure; see Jacod and Shiryaev (1987,
Ch.~I, Theorem 3.18) \cite{Jacod}.
\begin{lem}\label{compensator}
Under the notations of Section \ref{s:formal_definition}, let
$(H(t))$ be a stochastic process, and let $\theta$ be an independent
Bernoulli random variable with parameter $p$, such that given
$\theta$, the process $(H(t)-a_\theta t)$ is a martingale with
respect to $\F_t^H$ under $P_\theta$. Let $\tilde{p}_t := P_p
\{\theta=1|\F_t^H\}$ be the posterior belief that $\theta=1$ given
$(H(s))_{s\leq t}$ under the probability measure $P_p$. Then the
process $\left(\int_0^t (\tilde{p}_{s-}a_1 + (1-\tilde{p}_{s-})a_0)
ds\right)$ is the (predictable) compensator of the process $(H(t))$
with respect to $\F_t^H$ under the probability measure $P_p$.
\end{lem}
\begin{proof}
Plainly we have
\begin{align}\notag
&H(t)-\int_0^t (\tilde{p}_{s-}a_1 + (1-\tilde{p}_{s-})a_0) ds
\\\notag &= \theta(H(t)-a_1 t) + (1-\theta)(H(t)-a_0 t)+\int_0^t
(\theta - \tilde{p}_{s-})(a_1-a_0)ds.
\end{align}
Fix $0\leq u\leq t$. The expectation with respect to $P_p$ is
\begin{align}\notag
&E\left.\left[ H(t) - \int_0^{t} (\tilde{p}_{s-}a_1 +
(1-\tilde{p}_{s-})a_0) ds - H(u) + \int_0^{u} (\tilde{p}_{s-}a_1 +
(1-\tilde{p}_{s-})a_0) ds \right| \F_{u}^H\right]\\\notag &=
E\left.\left[ H(t) -H(u) - \int_u^{t} (\tilde{p}_{s-}a_1 +
(1-\tilde{p}_{s-})a_0) ds \right| \F_{u}^H\right]\\\notag &=
E\left.\left[\theta(H(t)-a_1 t - H(u)+a_1 u )) + (1-\theta)(H(t)-a_0
t - H(u)+a_0 u)\right| \F_{u}^H\right]\\\notag &\ \ +
E\left.\left[\int_u^{t} (\theta - \tilde{p}_{s-})(a_1-a_0)ds\right|
\F_{u}^H\right]\\\notag &= E\left.\left[E\left.\left[\theta(H(t)-a_1
t - H(u)+a_1 u ) + (1-\theta)(H(t)-a_0 t - H(u)+a_0 u )\right|
\theta,\F_{u}^H\right]\right| \F_{u}^H\right]\\\notag &\ \ +
E\left.\left[\int_u^{t} E\left.\left[(\theta -
\tilde{p}_{s-})\right| \F_{s-}^H\right](a_1-a_0)ds\right|
\F_{u}^H\right] =0.\\\notag
\end{align}
It follows that the process $\left(H(t)-\int_0^t (\tilde{p}_{s-}a_1
+ (1-\tilde{p}_{s-})a_0) ds\right)$ is a martingale with respect to
$\F_t^H$ under the probability measure $P_p$. Therefore, the
predictable process $\left(\int_0^t (\tilde{p}_{s-}a_1 +
(1-\tilde{p}_{s-})a_0) ds\right)$ is the (predictable) compensator
of the process $(H(t))$ with respect to $\F_t^H$ under the
probability measure $P_p$.
\end{proof}
Lemmas \ref{lem:time_change} and \ref{compensator} yield the
following corollary:
\begin{cor}\label{martingales}
$\\$
C1. The (predictable) compensator of the process $(X(\K(t)))$ is \\
$\left(\int_0^t (p_{s-}\mu_1 + (1-p_{s-})\mu_0) d\K(s)\right).$ \\
C2. The (predictable) compensator of the process is\\
$\left(X(\K(t)) -\int_{\mathbb{R}\setminus \{0\}} h
\widetilde{N}_{\nu_0}(\K(t),dh) - \mu_0\K(t)\right)$
is $\left(- \beta\sigma^2 \int_0^t p_{s-} dK(s)\right).$ \\
C3. The (predictable) compensator of the random measure $
N(d\K(t),dh)$ is $(p_{s-} \nu_1(dh) + (1-p_{s-}) \nu_0(dh))
d\K(s)$; see Jacod and Shiryaev (1987, Ch.~II, Theorem 1.8)
\cite{Jacod}.
\end{cor}
The following lemma states the the posterior process $(p_t)$ spends
zero time at any given positive contour-line lower than $1$.
\begin{lem}\label{spends}
For every $t\geq 0$, $p\in [0,1]$, and every $0<\delta<1$,
$$E^{0,p}\left[\int_0^t \chi_{\{p_{s-}=\delta\}}ds\right]=0.$$
\end{lem}
\begin{proof}
\begin{align}\notag
&E^{0,p}\left[\int_0^t \chi_{\{p_{s-}=\delta\}}ds\right] =
E^{0,p}\left[\int_0^t \chi_{\{p_s=\delta\}}ds\right] \\\notag &=
E\left.\left[\int_0^t \chi_{\{p_s=\delta\}}ds\right| \theta=1\right]
p + E\left.\left[\int_0^t
\chi_{\{p_s=\delta\}}ds\right|\theta=0\right] (1-p) \\\notag
&=E\left.\left[\int_0^t
\chi_{\{\ln\left(\frac{1-p_s}{p_s}\right)=\ln\left(\frac{1-\delta}{\delta}\right)\}}ds\right|\theta=1\right]
p + E\left.\left[\int_0^t
\chi_{\{\ln\left(\frac{1-p_s}{p_s}\right)=\ln\left(\frac{1-\delta}{\delta}\right)\}}ds\right|\theta=0\right]
(1-p)=0.\\\notag
\end{align}
The last equation follows since, as long as jumps from $B_\infty$ do
not appear (see Remark \ref{key}), the process
$\left(\ln\left(\frac{1-p_s}{p_s}\right)\right)$ is a time change of
a L\'{e}vy process whose jump process part has finite variation and
has no positive jumps, given the type $\theta$ (See Bertoin (1996,
Ch.~V, Theorem 1) \cite{Bertoin}). In case a jump from $B_\infty$
appears, from that time onwards the posterior process $(p_t)$
remains at level $1$.
\end{proof}
The following lemma assures that $\alpha^*$ is well defined.
\begin{lem}\label{alpha}
Eq.~(\ref{eta}) admits a unique solution in the interval $(0,\infty)$.\\
\end{lem}
\begin{proof}[\textbf{Proof}]
$\newline$ The function $f$ is a continuous function that satisfies
$f(0)<0$ and $f(\infty) = \infty$. To show that $f(\alpha)=0$ has a
unique solution, it is therefore sufficient to prove that $f$ is
increasing in $\alpha$. Note that (if $\sigma\neq 0$ then)
$\frac{1}{2}(\alpha+1)\alpha
\left(\frac{\mu_1-\mu_0}{\sigma}\right)^2 - r $ is increasing in
$\alpha$. It remains to prove that if $\vg-\vvg>0$, i.e., $\vi
(\mathbb{R}\backslash\{0\})-\vvi(\mathbb{R}\backslash\{0\})>0$, then
$\int_{\mathbb{R}\setminus \{0\}}
\left(\left(\frac{\nu_0}{\nu_1}(h)\right)^{\alpha}-1\right) \vvh+
\alpha (\vg-\vvg)$ is increasing in $\alpha$. Since
\begin{align}\notag
& \int_{\mathbb{R}\setminus
\{0\}}\left[\left(\left(\frac{\nu_0}{\nu_1}(h)\right)^{\alpha}-1\right)\vvh
+ \alpha (\vh-\vvh) \right]\\\notag &=\int_{\mathbb{R}\setminus
\{0\}}\left[\left(\frac{\nu_0}{\nu_1}(h)\right)\left(\left(\frac{\nu_0}{\nu_1}(h)\right)^{\alpha}-1\right)
+ \alpha \left(1-\frac{\nu_0}{\nu_1}(h)\right) \right]\vh
\end{align}
and $$\underset{ \{ h\mid\frac{\nu_0}{\nu_1}(h)=1 \}
}{\int}\left[\left(\frac{\nu_0}{\nu_1}(h)\right)\left(\left(\frac{\nu_0}{\nu_1}(h)\right)^{\alpha}-1\right)
+ \alpha \left(1-\frac{\nu_0}{\nu_1}(h)\right) \right]\vh=0$$ it is
sufficient to prove that for $\vi$-a.e. $h\in \{ h\mid
\frac{\nu_0}{\nu_1}(h)\neq 1 \} $,
$$g_h(\alpha) = \left(\frac{\nu_0}{\nu_1}(h)\right)\left(\left(\frac{\nu_0}{\nu_1}(h)\right)^{\alpha}-1\right) + \alpha \left(1-\frac{\nu_0}{\nu_1}(h)\right) $$ is increasing\footnote{In our model $\frac{\nu_0}{\nu_1}(h)\neq 1$ is actually $\frac{\nu_0}{\nu_1}(h)< 1$ $\vi$-a.s.
Yet, the proof works in the more general case of inequality.} in
$\alpha$. Now,
\begin{align}\notag
g_h'(\alpha) &= \left(\frac{\nu_0}{\nu_1}(h)\right)^{\alpha+1}
\ln\left(\frac{\nu_0}{\nu_1}(h) \right)+
\left(1-\frac{\nu_0}{\nu_1}(h)\right)
\\\notag
&> \left(\frac{\nu_0}{\nu_1}(h)\right)
\ln\left(\frac{\nu_0}{\nu_1}(h) \right)+
\left(1-\frac{\nu_0}{\nu_1}(h)\right) > 0,
\end{align}
where the first inequality holds since $\alpha>0$ and the second
inequality holds since $x\ln(x) +1-x > 0$ for every $x\neq 1$.
Therefore, $g_h(\alpha)$ is increasing, as desired.
\end{proof}
\textbf{Acknowledgments.}
We thank Sven Rady for useful comments on an earlier version of the
paper. We also would like to thank an anonymous referee for
his\textbackslash her helpful comments and suggestions for
improvements to an earlier version of this paper. This research was
supported in part by Israel Science Foundation grant $\sharp
212/09$, and by the Google Inter-university Center for Electronic
Markets and Auctions.
\bibliographystyle{plain}
|
\section*{Introduction}
This paper investigates the following minimization problem: given $\mu$, $\nu$ two - smooth enough - probability densities on $\mathbb{R}^d$ with $\mu$ supported in a domain $\Omega$, we study
$$ \inf\left\{ J_\varepsilon(T) \,:\, T_\#\mu = \nu \right\} \qquad \text{where} \qquad J_\varepsilon(T) = \int_\Omega |T(x)-x|\, \text{\textnormal{d}}\mu(x) + \varepsilon\int_\Omega |DT|^2 \, \text{\textnormal{d}} x. $$
Here $\varepsilon$ is a vanishing positive parameter, $|\cdot|$ is the usual Euclidean norm on $\mathbb{R}^d$ or $M_d(\mathbb{R})$, $DT$ denotes the Jacobian matrix of the vector-valued map~$T$ and $T_\#\mu$ is the image measure of $\mu$ by the map $T$ (defined by $T_\#\mu (B)= \mu(T^{-1}(B))$ for any Borel set $B \subset \mathbb{R}^d$). We aim to understand the behavior of the functional $J_\varepsilon$ in the sense of $\Gamma$-convergence and of characterize the limits of the minimizers $T_\varepsilon$.
If we do not take in to account this gradient penalization, we recover the classical optimal transport problem originally proposed by Monge in the 18th century \cite{monge}. For this problem, the particular constraint $T_\#\mu=\nu$ makes the existence of minimizers quite difficult to obtain by the direct method of the calculus of variations; when the Euclidean distance is replaced with nicer functions (usually strictly convex with respect to the difference $x-T(x)$), strong progresses have been realized by Kantorovich in the '40s \cite{kant, kant2} and Brenier in the late '80s \cite{Brenier polar}.
In the Monge's case of the Euclidean distance, the existence results have been shown more recently by several techniques: we just mention the first approach by Sudakov \cite{s} (later completed by Ambrosio \cite{a}) and the differential equations methods by Evans and Gangbo \cite{eg}, for the case of the Euclidean norm. More recently, this has been generalized to uniformly convex norms by approximation \cite{cfmc,tw}; finally, \cite{cdp} generalizes this result for any generic norm in $\mathbb{R}^d$. We refer to \cite{vil} for a complete overview of the optimal transport theory, and, again, the lecture notes \cite{a} for the particular case of the Monge problem.
Adding a Sobolev-like penalization is very natural in many applications, for instance in image processing, when these transport maps could model transformations in the space of colors, which are then required to avoid abrupt variations and discontinuities (see \cite{fprpa}). Also, these problems appear in mechanics (see \cite{gm}) and computer-science problems \cite{lps} , where one looks for maps with minimal distance distortion (if possible, local isometries with prescribed image measure).
However, in this paper we want to concentrate on the mathematical properties of this penalization. Notice first that this precisely allows to obtain very quickly the existence of optimal maps, since we get more compactness and we can this time use the direct method of the calculus of variations (see Prop.\@~\ref{directmethod} below; actually, this is trickier in the case where the source measure is not regular, {\it cf.}\@~\cite{l} and \cite[Chapter 1]{lphd}; this requires to use the theory developed in \cite{bbs}). Multiplying this penalization with a vanishing parameter $\varepsilon$ is motivated by the particular structure of the Monge problem. It is known that the minimizers of $J_0$ (that we will denote by $J$ in the following) are not unique and are exactly the transport maps from $\mu$ to $\nu$ which also send almost any source point $x$ to a point $T(x)$ belonging to the same {\it transport ray} as $x$ (see below the precise definition). Among these transport maps, selection results are particularly useful, and approximating with strictly convex transport costs $c_\varepsilon(x,y)=|x-y|+\varepsilon|x-y|^2$ brings to the {\it monotone transport} ({\it i.e.}~the unique transport map which is non-decreasing along each transport ray). Some regularity properties are known about it (continuity in the case of regular measures with disjoint and convex supports in the plane \cite{fgp}, uniform estimates on an approximating sequence under more general assumptions \cite{lsw}). The question is to know which of these transport map is selected by the approximation through the gradient penalization that we propose here. This is a natural question, and, due to its higher regularity as other maps, one could wonder if it is again the monotone map in this case.
For this, we analyze the behavior of $J_\varepsilon$ when $\varepsilon$ vanishes in the sense of the $\Gamma$-convergence (see \cite{braides} for the definitions and well-known results about this notion). The convergence of $J_\varepsilon$ to the transport energy $J$ is straightforward but requires a density result of the set of Sobolev maps $T \in H^1(\Omega)$ sending $\mu$ to $\nu$ among the set of transport maps. This result looks natural and can of course be used in other contexts, but was surprisingly not so much investigated so far:\smallskip
\begin{theostar} Let $\Omega$, $\Omega'$ be two bounded open star-shaped subsets of $\mathbb{R}^d$ with sufficiently smooth boundaries. Let $\mu \in \mathcal{P}(\Omega)$, $\nu \in \mathcal{P}(\Omega')$ be two probability measures, both absolutely continuous with respect to the Lebesgue measure, with densities $f$, $g$; assume that $f$, $g$ belong to $C^{0,\alpha}(\bar{\Omega})$, $C^{0,\alpha}(\bar{\Omega'})$ for some $\alpha > 0$ and are bounded from above and below by positive constants. Then the set
$$ \left\{ T \in \operatorname{Lip}(\Omega) \,:\, T_\#\mu = \nu \right\}$$
is non-empty, and is a dense subset of the set
$$ \left\{ T : \Omega \to \Omega' \,:\, T_\#\mu = \nu \right\} $$
endowed with the norm $||\cdot||_{L^2(\Omega)}$.
As a consequence, under these assumptions on $\Omega$, $\Omega'$, $\mu$, $\nu$, we have $J_\varepsilon \xrightarrow{\Gamma} J$ as $\varepsilon \to 0$. \end{theostar}
Notice that the above result would be completely satisfactory if the assumptions on the measures in order to get density of Lipschitz maps were the same as to get existence of at least one such a map. The assumptions that we used are likely not to be sharp, but are the most natural one if one wants to guarantee the existence of $C^{1,\alpha}$ transport maps (and it is typical in regularity theory that Lipschitz regularity is not easy to provide, whereas H\"older results work better; notice on the contrary that a well-established $L^p$ theory is not available in this framework). Anyway, from the proof that we give in section 2, it is clear that we are not using much more than the simple existence of Lipschitz maps.
Now a natural question concerns the behavior of the remainder with respect to the order $\varepsilon$. Namely, if we denote by $W_1$ the optimal value of the Monge problem ({\it i.e.=.}\@~the Wasserstein distance between $\mu_1$ and $\mu_2$), we need to consider $\frac{J_\varepsilon-W_1}{\varepsilon}$. The result is the most natural that we expect (this time with an almost a trivial proof)
\begin{equation} \frac{J_\varepsilon(T_\varepsilon)-W_1}{\varepsilon} \xrightarrow{\Gamma} \mathcal H \label{ordereps} \end{equation}
$$ \text{where} \qquad \mathcal H(T) = \left\{ \begin{array}{l} \displaystyle\int_\Omega |DT|^2 \qquad \text{if } T \in \O_1(\mu,\nu)\cap H^1(\Omega) \\[2mm] +\infty \qquad \text{otherwise} \end{array} \right. $$
where we have denoted by $\O_1(\mu,\nu)$ the set of optimal maps for the Monge problem. Notice that this result allows immediately to build some examples of measures $\mu$, $\nu$ for which the minimal value of the function $\mathcal H$ is not attained by the monotone transport map from $\mu$ to $\nu$, thanks to suitable analysis of the minimization of the $H^1$-norm among the set of transport maps on the real line (which has been very partially treated in \cite{ls}).
The convergence \eqref{ordereps} gives immediately a first order approximation (meaning $\inf J_\varepsilon = W_1 + \varepsilon \inf \mathcal H + o(\varepsilon)$) provided that $\inf \mathcal H \neq +\infty$, {\it i.e.}~when there exists at least one map which minimizes the Monge cost and belongs to the Sobolev space $H^1(\Omega)$. If no such map exists, the question arises to investigate the order of convergence of $\inf J_\varepsilon$ to $W_1$, the $\Gamma$-limit of the ``rest'' (provided that we are able to correctly define this rest), and which map is selected as $\varepsilon$ goest to~$0$.
The rest of this paper is devoted to the complete study of this case in a particular example, which has very interesting properties in these issues. We take as source domain the quarter disk in the plane $\Omega = \left\{ x = (r,\theta) \,:\, 0< r < 1, 0 < \theta < \pi/2 \right\}$ and as target domain an annulus located between two regular curves in polar coordinates $\Omega'= \left\{ x = (r,\theta) \,:\, R_1(\theta)< r < R_2(\theta), 0 < \theta < \pi/2 \right\}$; we endow these domains with two regular densities $f$, $g$ satisfying the condition:
$$ \text{for any }\theta, \quad \int_0^1 f(r,\theta) \,r\, \text{\textnormal{d}} r = \int_{R_1(\theta)}^{R_2(\theta)} g(r,\theta) \, r \, \text{\textnormal{d}} r $$
which means that the mass (with respect to $f$) of the segment with angle $\theta$ joining the origin to the quarter unit circle is equal to the mass (with respect to $g$) of the segment with same angle joining the two boundaries of $\Omega'$. Under these assumptions, the transport rays are the lines starting from the origin and the optimal transport maps for the Monge problem send each $x \in \Omega$ onto a point $T(x)$ with same angle; in other words, any optimal $T$ is written as
$$ T(x) = \phi(x)\frac{x}{|x|} $$
for some scalar function $\phi$ with $\inf \phi>0$. It then follows from an elementary computation that, for such a $T$, $\int_\Omega |DT|^2=+\infty$; thus, this is precisely a framework where $\mathcal{H}=+\infty$ and we cannot have $\inf J_\varepsilon = W_1+O(\varepsilon)$. Indeed, we obtain a logarithmic term, namely
\begin{equation} \inf J_\varepsilon = W_1 + \frac{K}{3}\varepsilon|\log\varepsilon|+O(\varepsilon) \quad \text{where} \quad K = \inf\left\{ \int_0^{\pi/2} (\phi^2+\phi'^2) \,:\, R_1(\theta) \leq \phi(\theta) \leq R_2(\theta) \right\} \label{K} \end{equation}
which means that $K$ is the smallest ``cost'' (in the sense of the $H^1$-norm on the interval $[0,\pi/2]$) of a curve in polar coordinates belonging to the target domain $\Omega'$. In order to justify, a formal computation suggests that, if $T_\varepsilon \to T$ for an optimal $T$,
$$ J_\varepsilon(T_\varepsilon) = W_1 + \frac{1}{3}||\phi(0,\cdot)||_{H^1(0,\pi/2)}^2\varepsilon|\log\varepsilon| + O(\varepsilon) $$
if $T(x)=\phi(r,\theta)\dfrac{x}{|x|}$. In other words, some new phenomena appear:
\begin{itemize}
\item the approximation of the minimal value of the Monge problem has the order $\varepsilon|\log\varepsilon|$;
\item the main ``rest'' (namely, $(J_\varepsilon(T_\varepsilon)-W_1)/(\varepsilon|\log\varepsilon|)$) involves only the behavior of $T$ around the origin, which is, in this case, the only crossing point of all the transport rays (and also a common singular point for any element of $\O_1(\mu,\nu)$);
\item it could be the case that this rest is not minimized (among the transport maps from $\mu$ to $\nu$) by the monotone transport map (if the function which realizes the infimum in \eqref{K}, that we call $\Phi$ in that follows, differs from the lower boundary $R_1$).
\end{itemize}
These remarks are consequences of the following main result of this paper:\smallskip
\begin{theostar} With the above notations, let us assume that $R_1$, $R_2$ are Lipschitz functions of $\theta$ and $f$, $g$ are both Lipschitz and bounded from above and from below by positive constants on $\bar{\Omega}$, $\bar{\Omega'}$. We denote by
$$ F_\varepsilon = \frac 1\varepsilon\left(J_\varepsilon - W_1 - \frac{K}{3}\varepsilon|\log\varepsilon|\right) $$
and by $F$ the function of $T$ which is equal to $+\infty$ when $T$ does not belong to $\O_1(\mu,\nu)$ and such that, for $T(x)=\phi(r,\theta)\dfrac{x}{|x|}$ belonging to $\O_1(\mu,\nu)$, we have
$$ F(T) = \int_0^1 \frac{||\phi(r,\cdot)||_{H^1(0,\pi/2)}^2-K}{r} \, \text{\textnormal{d}} r + \int_0^1 ||\partial_r\phi(r,\cdot)||_{L^2(0,\pi/2)}^2 \, r \, \text{\textnormal{d}} r $$
\begin{enumerate}
\item For any family of maps $(T_\varepsilon)_\varepsilon$ such that $(F_\varepsilon(T_\varepsilon))_\varepsilon$ is bounded, there exists a sequence $\varepsilon_k \to 0$ and a map $T$ such that $T_{\varepsilon_k} \to T$ in~$L^2(\Omega)$.
\item There exists a constant $C$, depending only on the domains $\Omega$, $\Omega'$ and on the measures $f$, $g$, so that, for any family of maps $(T_\varepsilon)_{\varepsilon>0}$ with $T_\varepsilon \to T$ as $\varepsilon \to 0$ in $L^2(\Omega)$, we~have
$$ \liminf\limits_{\varepsilon \to 0} F_\varepsilon(T_\varepsilon) \geq F(T)-C $$
\item Moreover, there exists at least one family $(T_\varepsilon)_{\varepsilon > 0}$ such that $(F_\varepsilon(T_\varepsilon))_\varepsilon$ is indeed bounded. \end{enumerate} \end{theostar}
Notice that this result does not give precisely the $\Gamma$-limit of the functional $F_\varepsilon$ but only a lower bound on the $\Gamma$-liminf. However, it is enough to obtain some important consequences on the behavior of the approximation. Indeed:
\begin{itemize}
\item the fact that there exists a family $(T_\varepsilon)_\varepsilon$ such that $(F_\varepsilon(T_\varepsilon))_\varepsilon$ is bounded implies that this is the case when we take a sequence $T_\varepsilon$ of minimizers (and in this case we necessary have $T_\varepsilon \to T$ with $F(T)<+\infty$);
\item it is possible to prove that $||\phi(r,\cdot)||_{H^1}^2-K \geq ||\phi(r,\cdot)-\Phi||_{L^2}^2$ (see Lemma~\ref{phimoinsphi} below). This implies that with the above notations, if $F(T) < +\infty$, then $\phi(r,\cdot)$ is a continuous function of the variable $r$ (and valued in the space $L^2([0, \pi/2])$) and we have $\phi(0,\cdot) = \Phi$, the optimizer for the Sobolev norm among the curves which are valued in $\Omega'$.
\end{itemize}
The fact that the $\varepsilon|\log\varepsilon|$ term in the energy is due to the blow-up of the Sobolev norm at a single point suggests a formal but deep analogy with the Ginzburg-Landau theory (see for instance \cite{bbh}), where we look at the minimization of
$$ u \mapsto \frac{1}{\varepsilon^2} \int_\Omega (1-|u|^2) + \int_\Omega |\nabla u|^2$$
with Neumann boundary conditions. Here two terms are contradictory in the functional: the first one suggests to select unit vector fields parallel to the boundary, and the second requires $H^1$-regularity, which is impossible since the previous point creates a vortex. In our case, up to a $90^\circ$ rotation, the situation is similar. The two contradictory phenomenas are the fact that $T$ has to preserve the transport rays (to optimize the Monge problem) and that it has a finite Sobolev norm; this also leads to the creation of an explosion (a vortex, rotated by $90^\circ$; here the origin, which is sent to a whole curve belonging to the target domain). The excess of order $\varepsilon|\log\varepsilon|$ is a common feature of the two problems (but the analogy is likely to stop here).
\paragraph{Plan of the paper.} Section 1 collects general notations and well-known facts about the Monge problem, basic notions of $\Gamma$-convergence and some useful and elementary results about the optimal transport with gradient term. In Section 2, we prove the density of the Sobolev transport maps among the transport maps and state the results of $\Gamma$-convergence with order $0$ and with order $1$ of $(J_\varepsilon)_\varepsilon$ for generic (regular) domains and measures. In Section 3, we study precisely the example of $\varepsilon|\log\varepsilon|$ approximation in the above framework; the main results and its interpretations are given in Paragraph 3.3, following a formal computation that we present in Paragraph 3.2. Section 4 is completely devoted to the rigorous proof of the main result of this paper (Theorem \ref{PRESQUEGAMMA}).
\paragraph{Acknowledgements.} This work has mainly been developed during some visits of the second and third authors to the Universit\`a di Pisa, which have been possible thanks to the support of ANR project ANR-12-BS01-0014-01 GEOMETRYA and the PGMO project MACRO, of EDF and Fondation Math\'ematique Jacques Hadamard. This material also is based upon work supported by the National Science Foundation under Grant No.\@ 0932078~000, while the second author was in residence at the Mathematical Science Research Institute in Berkeley, California, during the Fall term 2013.
The research of the first author is part of the project 2010A2TFX2 {\it``Calcolo delle Variazioni''}, financed by the Italian Ministry of Research
and is partially financed by the {\it ``Fondi di ricerca di ateneo''} of the University of Pisa.
\section{Preliminary notions}
\subsection{Known facts about the Monge problem}
In this section, we recall some well-known facts and useful tools about the optimal transportation problem with the Monge cost $c(x,y)=|x-y|$, where $|\cdot|$ is the Euclidean norm on $\mathbb{R}^d$.
Let $\Omega$, $\Omega'$ be two bounded open set on $\mathbb{R}^d$, $\mu \in \mathcal{P}(\Omega)$, $\nu \in \mathcal{P}(\Omega')$ and assume that $\mu$ is absolutely continuous with respect to the Lebesgue measure with density $f$. Then we set
$$ W_1(\mu,\nu) = \min\left\{ \int_\Omega |T(x)-x| \, \text{\textnormal{d}} \mu(x) \;:\; T : \Omega \to \mathbb{R}^d,\, T_\#\mu = \nu \right\} $$
the minimal value of the Monge transport cost from $\mu$ to $\nu$, and
$$ \O_1(\mu,\nu) = \left\{ T : \Omega \to \mathbb{R}^d \;:\; T_\# \mu = \nu \; \text{and} \; \int_\Omega |T(x)-x|\, \text{\textnormal{d}}\mu(x) = W_1(\mu,\nu) \right\} $$
the set of optimal transport maps for the Monge cost (if there is no ambiguity, we will simply use the notations $W_1$ and $\O_1$). Notice that the above Monge problem is a particular issue of its Kantorovich formulation, i.e.
$$\min\left\{ \int_{\Omega\times\Omega'} |y-x| \, \text{\textnormal{d}} \gamma(x,y) \;:\; \gamma\in \mathcal{P}(\Omega\times\Omega'): (\pi_x)_\#\gamma=\mu, \, (\pi_y)_\#\gamma= \nu \right\},$$
and the minimal value of this problem coincides with $W_1(\mu,\nu)$ as well.
\begin{theo}[Duality formula for the Monge problem] We have the equality
$$ W_1(\mu,\nu) = \sup\left\{ \int_{\Omega'} u(y)\, \text{\textnormal{d}}\nu(y) - \int_\Omega u(x)\, \text{\textnormal{d}}\mu(x) : u \in \operatorname{Lip}_1(\mathbb{R}^d) \right\} $$
where $\operatorname{Lip}_1(\mathbb{R}^d)$ denotes the set of $1$-Lipschitz functions $\mathbb{R}^d \to \mathbb{R}$. Such optimal Lipschitz functions are called \textnormal{Kantorovich potentials}. Moreover, the Kantorovich potential is unique provided that $\Omega$ is a connected set and that the set $\{f>0\}\setminus\Omega'$
is a dense subset of $\Omega$. \end{theo}
As a direct consequence of the duality formula, if $T:\Omega \to \mathbb{R}^d$ sends $\mu$ to $\nu$ and $u \in \operatorname{Lip}_1(\mathbb{R}^d)$, we have the equivalence:
$$ \left\{ \begin{array}{l} T \in \O_1(\mu,\nu) \\ u \text{ is a Kantorovich potential} \end{array} \right. \iff \text{for } \mu\text{-a.e. } x\in \Omega,\; u(T(x))-u(x) = |T(x)-x| $$
We now introduce the following crucial notion of {\it transport ray}:
\begin{defi}[Transport rays] Let $u$ be a Kantorovich potential and $x \in \Omega$, $y \in \Omega'$. Then:
\begin{itemize}
\item the open oriented segment $(x,y)$ is called \textnormal{transport ray} if $u(y)-u(x)=|y-x|$;
\item the closed oriented segment $[x,y]$ is called \textnormal{maximal transport ray} if any point of $(x,y)$ is contained into at least one transport ray with same orientation as $[x,y]$, and if $[x,y]$ is not strictly included in any segment with the same property.
\end{itemize} \end{defi}
\begin{prop}[Geometric properties of transport rays] The set of maximal transport rays does not depend on the choice of the Kantorovich potential $u$ and only depends on the source and target measures~$\mu$ and $\nu$. Moreover:
\begin{itemize}
\item any intersection point of two different maximal transport rays is an endpoint of these both maximal transport rays;
\item the set of the endpoints of all the maximal transport rays is Lebesgue-negligible.
\end{itemize} \end{prop}
These notions allow to prove the existence and to characterize the optimal transports:
\begin{prop}[Existence and characterization of optimal transport maps] The solutions of the Monge problem exist and are not unique; precisely, a map $T$ sending $\mu$ to $\nu$ is optimal if and only if:
\begin{itemize}
\item for a.e.~$x \in \Omega$, $T(x)$ belongs to the same maximal transport ray as $x$;
\item the oriented segment $[x,T(x)]$ has the same orientation as this transport ray.
\end{itemize} \end{prop}
We finish by recalling that, among all maps in $\O_1(\mu,\nu)$, there is a special one which has received much attention so far, and by quickly reviewing its properties.
\begin{prop}[The monotone map and a secondary variational problem] If $\mu$ is absolutely continuous, there exists a unique transport map $T$ from $\mu$ to $\nu$ such that, for each maximal transport ray $S$, $T$ is non-decreasing from the segment $S \cap \Omega$ to the segment $S \cap \Omega'$ (meaning that if $x,x' \in \Omega$ belong to the same transport ray, then $[x,x']$ and $[T(x),T(x')]$ have the same orientation).Moreover, $T$ solves the problem
$$ \inf\left\{ \int_{\Omega} |T(x)-x|^2 \, \text{\textnormal{d}}\mu(x) : T \in \O_1(\mu,\nu) \right\}.$$
\end{prop}
Notice that this solution is itself obtained as limit of minimizers of a perturbed variational problem, namely
$$ \inf\left\{ \int_\Omega |T(x)-x|\, \text{\textnormal{d}}\mu(x) + \varepsilon\,\int_\Omega |T(x)-x|^2\, \text{\textnormal{d}}\mu(x) \,:\, T_\#\mu = \nu \right\} $$
This very transport map is probably one of the most natural in $\O_1(\mu,\nu)$ and one of the most regular. In particular, under some assumptions on the densities $f$, $g$ and their supports $\Omega$, $\Omega'$ (convex and disjoint supports in the plane and continuous and bounded by above and below densities), this transport has also be shown to be continuous \cite{fgp}, and \cite{lsw} gives also some regularity results for the minimizer $T_\varepsilon$ of an approximated problem where $c$ is replaced with $c_\varepsilon(x,y) = \sqrt{\varepsilon^2+|x-y|^2}$ (local uniform bounds on the eigenvalues of the Jacobian matrix $T_\varepsilon$).
\subsection{Tools for optimal transport with gradient penalization}
\paragraph{Existence of solutions.} We begin by showing the existence of solutions for the penalized problem with an elementary proof:
\begin{prop} \label{directmethod} Let $\Omega$ be a bounded open set of $\mathbb{R}^d$ with Lipschitz boundary. Let $\mu \in \mathcal{P}(\Omega)$ be absolutely continuous
with density $f$, and $\nu \in \mathcal{P}(\mathbb{R}^d)$ such that the set
$$ \left\{ T \in H^1(\Omega) \,:\, T_\#\mu = \nu \right\} $$
is non-empty. Then for any $\varepsilon > 0$, the problem
\begin{equation} \inf\left\{ \int_\Omega |T(x)-x| f(x) \, \text{\textnormal{d}} x + \varepsilon \int_\Omega |DT(x)|^2 \, \text{\textnormal{d}} x \,:\, T \in H^1(\Omega), \, T_\#\mu=\nu \right\} \label{otwithgrad} \end{equation}
admits at least one solution.\end{prop}
\begin{proof} We use the direct method of the calculus of variations. Let $(T_n)_n$ be a minimizing sequence; since this sequence is bounded in $H^1(\Omega)$, it admits, up to a subsequence, a limit $T$ for the strong convergence in $L^2(\Omega)$; moreover, we also can assume $T_n(x)\to T(x)$ for a.e.~$x\in \Omega$, which also implies that the convergence holds $\mu$-a.e.. This implies that $T$ still satisfies the constraint on the image measure (since, for a continuous and bounded function $\phi$, the a.e.~convergence provides $\int_\Omega \phi\circ T_n \, \text{\textnormal{d}}\mu \to \int_\Omega \phi\circ T \, \text{\textnormal{d}}\mu$) and is thus admissible for \eqref{otwithgrad}. On the other hand, it is clear that the functional that we are trying to minimize is lower semi-continuous with respect to the weak convergence in $H^1(\Omega)$. This achieves the proof. \end{proof}
This existence result can be of course adapted to any functional with form $\int_\Omega L(x,T(x),DT(x)) \, \text{\textnormal{d}}\mu(x)$ under natural assumptions on the Lagrangian $L$ (continuity with respect to the two first variables, convexity and coercivity with respect to the third one).
\paragraph{Existence of smooth transport maps and Dacorogna-Moser's result.} In the above existence theorem, the existence of at least one admissible map $T$ was an assumption. If the measures are regular enough, it can be seen as a consequence of the classical regularity results of the optimal transport maps for the quadratic cost \cite{caffarelli1,caffarelli2,caffarelli3,dpf}. We recall here another result which provides the existence of a regular diffeomorphism which sends a given measure onto another one, which will be used several times in the paper. This result is due to Dacorogna and Moser \cite{dm} (this construction is nowadays regularly used in optimal transport, starting from the proof by Evans and Gangbo, \cite{eg}; it is also an important tool for the equivalence between different models of transportation, see \cite{dacmosFil}; notice that the transport that they consider is in some sense optimal for a sort of congested transport cost, as pointed out years ago by Brenier in \cite{brenier}). The version of their result that we will use is the following:
\begin{theo} \label{appdacmos} Let $U \subset \mathbb{R}^d$ be a bounded open set with $C^{3,\alpha}$ boundary $\partial U$. Let $f_1$, $f_2$ be two positive Lipschitz functions on $\bar{U}$ such that
$$ \int_U f_1 = \int_U f_2 $$
Then there exists a Lipschitz diffeomorphism $T : \bar{U} \to \bar{U}$ satisfying
$$\left\{
\begin{array}{ll}
\det \nabla T (x) = \dfrac{f_1(x)}{f_2(T(x))}, & x \in U \\[2mm]
T(x) = x, & x \in \partial U
\end{array}
\right. $$
Moreover, the Lipschitz constant of $T$ is bounded by a constant depending only on $U$, on the Lipschitz constants and on the lower bounds of $f_1$ and $f_2$ .
\end{theo}
Notice that the equation satisfied by $T$ exactly means (as $T$ is Lipschitz and one-to-one) that it sends the measures with density $f_1$ onto the measure with density $f_2$. The result of the original paper \cite[Theorem 1']{dm} deals with more general assumptions on the density $f_1$ ($f \in C^{k+3,\alpha}(\bar{U})$ and the result is a $C^{k+1,\alpha}$ diffeomorphism) but only considered $f_2=1$. We gave here a formulation better suitable for our needs, easy to obtain from the original theorem. We do not claim this result to be ``sharp'', but it is sufficient in the case which we are interested in (see paragraph \ref{construction}).
\paragraph{The one-dimensional case.} We finish by giving some very partial results about the optimal transport problem with gradient term on the real line. First, we recall the classical result about the one-dimensional optimal transportation problem (we refer for instance to \cite{sln} for the proof):
\begin{prop} Let $I$ be a bounded interval of $\mathbb{R}$ and $\mu \in \mathcal{P}(I)$ be atom-less and $\nu \in \mathcal{P}(\mathbb{R})$. Then there exists a unique map $T:I\to \mathbb{R}$ which is non-decreasing and sends $\mu$ onto $\nu$. Moreover, if $h$ is a convex function $\mathbb{R}\to\mathbb{R}$, then $T$ solves the minimization problem
$$ \inf\left\{ \int_\Omega h(T(x)-x)\, \text{\textnormal{d}}\mu(x) \,:\, T_\#\mu = \nu \right\} $$
with uniqueness provided that $h$ is strictly convex. \end{prop}
Now we state the results concerning the optimality of this monotone map $T$ for transport problems involving derivatives:
\begin{prop} Let $I$ be a bounded interval of $\mathbb{R}$. Let $\mu$ be a finite measure on $I$ with positive density $f$, and $\nu$ a positive measure on $\mathbb{R}$ with same mass. Let $h$ be a convex and non-decreasing function $\mathbb{R}\to\mathbb{R}$. Then, if $f$ is constant on $I$ (and $\mu$ is therefore the uniform measure on $I$), the monotone transport map from $\mu$ to $\nu$ is optimal for the problem
$$ \inf\left\{ \int_\Omega h(|T'(x)|) \, \text{\textnormal{d}}\mu(x) \,:\, T \in W^{1,1}(I),\, T_\#\mu = \nu \right\}. $$
The same stays true if $f$ is not constant but satisfies $\dfrac{\inf f}{\sup f} \geq \dfrac{1}{2}$.
On the other hand, the same is not true for arbitrary $f$, if $\dfrac{\inf f}{\sup f}$ is too small. Indeed, define $U$ on $[0,1]$ as being the triangle function
$$ U(x) =\begin{cases}2x& \text{if } 0 \leq x \leq \dfrac{1}{2}, \\2-2x & \text{if } \dfrac{1}{2} \leq x \leq 1 \end{cases} $$
and, for $\alpha > 0$, take $f=\alpha$ on $\left[0,\frac{1}{4}\right]\cup\left[\frac{3}{4},1\right]$ and $f=1$ elsewhere, thus obtaining a measure $\mu_\alpha$. If $T_\alpha$ is the unique non-decreasing map sending $\mu_\alpha$ onto $U_\#\mu_\alpha$, then
$$ \int_0^1 |U'|^2 < \int_0^1 |T_\alpha'|^2 \qquad \text{for } \alpha \text{ large enough}.$$
\end{prop}
We refer to \cite{ls} and \cite[Chapter 2]{lphd} for proofs and details (and notice that this construction has been used to find counter-examples to the optimality of the monotone transport for other variational problem in \cite{lps}).
\subsection{Definitions and basic results of $\Gamma$-convergence}
We finish this preliminary section by the tools of $\Gamma$-convergence that we will use throughout this paper. All the details can be found, for instance, in the classical Braides's book \cite{braides}. In that follows, $(X,d)$ is a metric space.
\begin{defi} Let $(F_n)_n$ be a sequence of functions $X \mapsto \bar\mathbb{R}$. We say that $(F_n)_n$ $\Gamma$-converges to $F$, and we write $F_n \xrightarrow[n]{\Gamma} F$ if, for any $x \in X$, we have
\begin{itemize}
\item for any sequence $(x_n)_n$ of $X$ converging to $x$,
$$ \liminf\limits_n F_n(x_n) \geq F(x) \qquad \text{($\Gamma$-liminf inequality);}$$
\item there exists a sequence $(x_n)_n$ converging to $x$ and such that
$$ \limsup\limits_n F_n(x_n) \leq F(x) \qquad \text{($\Gamma$-limsup inequality).} $$
\end{itemize} \end{defi}
This definition is actually equivalent to the following equalities for any $x \in X$:
$$ F(x) = \inf\left\{ \liminf\limits_n F_n(x_n) : x_n \to x \right\} = \inf\left\{ \limsup\limits_n F_n(x_n) : x_n \to x \right\} $$
The function $x \mapsto \inf\left\{ \liminf\limits_n F_n(x_n) : x_n \to x \right\}$ is called $\Gamma$-liminf of the sequence $(F_n)_n$, and the other one its $\Gamma$-limsup. A useful result is the following (which, for instance, implies that a constant sequence of functions does not $\Gamma$-converge to itself in general):
\begin{prop} The $\Gamma$-liminf and the $\Gamma$-limsup of a sequence of functions $(F_n)_n$ are both lower semi-continuous on $X$. \end{prop}
The main interest of $\Gamma$-convergence is its consequences in terms of convergence of minima:
\begin{theo} \label{convminima} Let $(F_n)_n$ be a sequence of functions $X \to \bar\mathbb{R}$ and assume that $F_n \xrightarrow[n]{\Gamma} F$. Assume moreover that there exists a compact and non-empty subset $K$ of $X$ such that
$$ \forall n\in N, \; \inf_X F_n = \inf_K F_n $$
(we say that $(F_n)_n$ is equi-mildly coercive on $X$). Then $F$ admits a minimum on $X$ and the sequence $(\inf_X F_n)_n$ converges to $\min F$. Moreover, if $(x_n)_n$ is a sequence of $X$ such that
$$ \lim_n F_n(x_n) = \lim_n (\inf_X F_n) $$
and if $(x_{\phi(n)})_n$ is a subsequence of $(x_n)_n$ having a limit $x$, then $ F(x) = \inf_X F $. \end{theo}
We finish with the following result, which allows to focus on the $\Gamma$-limsup inequality only on a dense subset of $X$ under some assumptions:
\begin{prop} \label{gammadense} Let $(F_n)_n$ be a sequence of functionals and $F$ be a functional $X \to \bar\mathbb{R}$. Assume that there exists a dense subset $Y \subset X$ such that:
\begin{itemize}
\item for any $x \in X$, there exists a sequence $(x_n)_n$ of $Y$ such that $x_n \to x$ and $F(x_n) \to F(x)$;
\item the $\Gamma$-limsup inequality holds for any $x \in Y$.
\end{itemize}
Then it holds for any $x$ belonging to the whole $X$. \end{prop}
\section{Generalities and density of Sobolev transport maps}
In that follows, we consider two regular enough domains $\Omega$, $\Omega'$ and two measures $\mu \in \mathcal{P}(\Omega)$, $\nu \in \mathcal{P}(\Omega')$ with positive and bounded from below densities $f$, $g$; we assume moreover that the class of maps $T \in H^1(\Omega)$ sending $\mu$ onto $\nu$ is non-empty (for instance, the assumptions of the Dacorogna-Moser's result are enough). The functional that we will study is defined, for $\varepsilon > 0$, by
$$ J_\varepsilon \,:\, T \mapsto \int_\Omega |T(x)-x|\, \text{\textnormal{d}}\mu(x) + \varepsilon \int_\Omega |DT(x)|^2 \, \text{\textnormal{d}} x $$
and we denote by $J$ the corresponding functional when $\varepsilon =0$ (which is thus the classical Monge's transport energy); moreover, we extend $J_\varepsilon$, $J$ to the whole $L^2(\Omega)$ by setting $J_\varepsilon(T)=J(T)=+\infty$ for a map $T$ which is not a transport map from $\mu$ to $\nu$.
As usual in transport theory, we consider as a setting for our variational problems the set of transport plans $\gamma$ which are probabilities on the product space $\Omega\times\Omega'$ with given marginals $(\pi_x)_\#\gamma=\mu$ and $(\pi_y)_\#\gamma=\nu$ and all the $\Gamma$-limits that we consider in that follows are considered with respect to the weak convergence of plans as probability measures. However, due to our choices of the functionals that we minimize, most of the transport plan that we consider will be actually induced by transport maps, i.e. $\gamma_T = (\text{id}\times T)_\#\mu$ with $T_\#\mu=\nu$. These maps are valued in $\Omega'$, which is bounded, and are hence bounded. We could also consider different noions of convergence, in particular based on the pointwise convergence of these plans, and we will actually do it often. For simplicity, we will use the convergence in $L^2(\Omega;\Omega')$ (but, since these functions are bounded, this is equivalent to any other $L^p$ convergence with $p<\infty$). As the following lemma (which will be also technically useful later) shows, this convergence is equivalent to the weak convergence in the sense of measures of the transport plans:
\begin{lem} \label{plansandmaps} Assume that $\Omega$, $\Omega'$ are compact domains and $\mu$ is a finite non-negative measure on $\Omega$. Let $(T_n)_n$ be a sequence of maps $\Omega \to \Omega'$. Assume that there exists a map $T$ such that $\gamma_{T_n} \rightharpoonup \gamma_T$ in the weak sense of measures. Then $T_n \to T$ in $L^2_\mu(\Omega)$.
Conversely, if $T_n \to T$ in $L^2_\mu(\Omega)$, then we have $\gamma_{T_{n}} \rightharpoonup \gamma_T$
in the weak sense of measures. \end{lem}
\begin{proof} If $\gamma_{T_n} \rightharpoonup \gamma_T$ and $\phi \in C_b(\Omega)$ is a vector-valued function, we have
$$\int_\Omega \phi(x) \cdot T_n(x) \, \text{\textnormal{d}}\mu(x) = \int_{\Omega\times \Omega'} \phi(x) \cdot y \, \text{\textnormal{d}}\gamma_{T_n}(x,y) \to \int_{\Omega\times \Omega'}\phi(x)\cdot y \, \text{\textnormal{d}}\gamma_T(x,y) = \int_{\Omega} \phi(x) \cdot T(x) \, \text{\textnormal{d}}\mu(x) $$
which proves that $T_n \rightharpoonup T$ weakly in $L^2_\mu(\Omega)$. On the other hand,
$$ \int_\Omega |T_n(x)|^2 \, \text{\textnormal{d}}\mu(x) = \int_{\Omega\times\Omega'} |y|^2 \, \text{\textnormal{d}}\gamma_{T_n}(x,y) \to \int_{\Omega\times\Omega'} |y|^2\, \text{\textnormal{d}}\gamma_T(x,y) = \int_\Omega |T(x)|^2 \, \text{\textnormal{d}}\mu(x) $$
$$ \text{thus} \qquad ||T_n||_{L^2_\mu} \to ||T||_{L^2_\mu} $$
and the convergence $T_n \to T$ is actually strong.
Conversely, assume that $T_n \to T$ in $L^2_\mu(\Omega)$ and let $(n_k)_k$ be such that the convergence $T_{n_k}(x) \to T(x)$ holds for a.e.\@~$x\in\Omega$, then for any $\phi \in C_b(\Omega\times\Omega')$ we have
$$ \int_{\Omega\times\Omega'} \phi(x,y) \, \text{\textnormal{d}}\gamma_{T_{n_k}}(x,y) = \int_\Omega \phi(x,T_{n_k}(x)) \, \text{\textnormal{d}}\mu(x) \to \int_\Omega \phi(x,T(x))\, \text{\textnormal{d}}\mu(x) = \int_{\Omega\times\Omega'} \phi(x,y) \, \text{\textnormal{d}}\gamma_T(x,y). $$
This proves $\gamma_{T_{n_k}} \rightharpoonup \gamma_T$, but, the limit being independent of the subsequence we easily get full convergence of the whole sequence.
\end{proof}
Since the set of transport plans between $\mu$ to $\nu$ is compact for the weak topology in the set of measures on $\Omega\times\Omega'$, a consequence of lemma \ref{plansandmaps} is that the equi-coercivity needed in Theorem~\ref{convminima} will be satisfied in all the $\Gamma$-convergence results that follows. Therefore, we will not focus on it anymore and still will consider that these results imply the convergence of minima and of minimizers.
\subsection{Statements of the zeroth and first order $\Gamma$-convergences}
\paragraph{Zeroth order $\Gamma$-limit.} The first step is to check that $J_\varepsilon \xrightarrow{\Gamma} J$. Here we must consider that $J_\varepsilon$ is extended to transport plan by setting $+\infty$ on those transport plans which are not of the form $\gamma=\gamma_T$ for $T\in H^1$, and that $J$ is defined as usual as $J(\gamma)=\int |x-y|\, \text{\textnormal{d}}\gamma$ for transport plans. This $\Gamma$-onvergence actually requires a non-trivial result which states that the set of Sobolev transport maps is a dense subset of the set of transport maps from $\mu$ to $\nu$. We prove this result in the next paragraph (see theorem \ref{denssobolev} below) for H\"older densities and a large class of domains given by the following definition:
\begin{defi} We call \textnormal{Lipschitz polar domain} any open bounded subset $\Omega$ having form
$$ \Omega = \left\{ x \in \mathbb{R}^d \,:\, |x-x_0| < \gamma\left( \frac{x-x_0}{|x-x_0|}\right) \right\} $$
for some $x_0 \in \Omega$ and a Lipschitz function $\gamma:S^{d-1}\to (0,+\infty)$. In particular, such a domain $\Omega$ is star-shaped with Lipschitz boundary. \end{defi}
\begin{prop}[Zeroth order $\Gamma$-limit] \label{zerogammalim} Assume that $\Omega$, $\Omega'$ are both Lipschitz polar domains and that $f$, $g$ are both $C^{0,\alpha}$ and bounded from below. Then $J_\varepsilon \xrightarrow{\Gamma} J$ as $\varepsilon \to 0$. \end{prop}
\begin{proof} The $\Gamma$-liminf inequality is trivial (we have $J_\varepsilon \geq J$ by definition, and $J$ is continuous for the weak convergence of plans), and the $\Gamma$-limsup inequality is a direct consequence of the Prop.\@ \ref{gammadense} and of the density of the set of Sobolev transports for the $L^2$-convergence. \end{proof}
\paragraph{First order $\Gamma$-limit.} We state it as follows, with this time a short proof:
\begin{prop}[First order $\Gamma$-limit] \label{gammalim} Assume simply that $f$, $g$ are bounded from below on $\Omega$, $\Omega'$ and that $\Omega$ has Lipschitz boundary. Then the functional $\dfrac{J_\varepsilon-W_1}{\varepsilon}$ $\Gamma$-converges, when $\varepsilon\to 0$, to
$$ \mathcal H : T \mapsto\begin{cases}\int_\Omega |DT(x)|^2 \, \text{\textnormal{d}} x & \text{if } T \in \O_1(\mu,\nu)\cap H^1(\Omega) \\ +\infty & \text{otherwise}, \end{cases}$$
where, again, $\mathcal H$ is extend to plans which are not induced by maps by $+\infty$. \end{prop}
\begin{proof}
{\it $\Gamma$-limsup inequality.}
If $T \in \O_1(\mu,\nu)$, then by choosing $T_\varepsilon = T$ for any $\varepsilon$ we obtain automatically $\dfrac{J_\varepsilon(T_\varepsilon)-W_1}{\varepsilon} = \int_\Omega |DT|^2$ for each $\varepsilon$. It remains to show that if $T \notin \O_1(\mu,\nu)$, then we have $\dfrac{J_\varepsilon(T_\varepsilon)-W_1}{\varepsilon} \to +\infty$ for any sequence $(T_\varepsilon)_\varepsilon$ converging to $T$; but since the map
$$ T \mapsto \int_\Omega |T(x)-x| \, \text{\textnormal{d}}\mu(x) $$
is continuous for the $L^2$-convergence, we have for such a $(T_\varepsilon)_\varepsilon$
$$ \liminf\limits_\varepsilon \frac{J_\varepsilon(T_\varepsilon)-W_1}{\varepsilon} \geq \liminf\limits_\varepsilon\frac{1}{\varepsilon} \left(\int_\Omega |T(x)-x| \, \text{\textnormal{d}}\mu(x) - W_1\right) $$
which is $+\infty$ since $T$ is not optimal for the Monge problem.
{\it $\Gamma$-liminf inequality.} We can concentrate on sequence of maps $T_\varepsilon$ with equibounded values for $\dfrac{J_\varepsilon(T_\varepsilon)-W_1}{\varepsilon}$, which provides a bound on $\int_\Omega |DT_{\varepsilon}|^2 $.
Assuming the liminf to be finite, from
$$C\geq \frac{J_{\varepsilon}(T_{\varepsilon})-W_1}{\varepsilon} = \frac{1}{\varepsilon} \left(\int_\Omega |T_{\varepsilon}(x)-x|\, \text{\textnormal{d}}\mu(x)-W_1\right) + \int_\Omega |DT_{\varepsilon}|^2 \geq \int_\Omega |DT_{\varepsilon}|^2 $$
we deduce as above that $T$ must belong to $ \O_1(\mu,\nu)$ and, since the last term is lower semi-continuous with respect to the weak convergence in $H^1(\Omega)$ (which is guaranteed up to subsequences since $(J_{\varepsilon_k}(T_{\varepsilon_k}))_k$ is bounded) we get the inequality we look for.\end{proof}
Since the $\Gamma$-convergence implies the convergence of minima, we then have
$$ \inf J_\varepsilon = \inf J + \varepsilon \inf \mathcal H + o(\varepsilon) = W_1 + \varepsilon \inf \mathcal H + o(\varepsilon)$$
provided that the infimum is finite, which means that there exists at least one transport map $T$ also belonging to the Sobolev space $H^1(\Omega)$. In the converse case, and under the assumptions of the zeroth order $\Gamma$-convergence, we have
$$ \inf J_\varepsilon \to W_1 \qquad \text{and} \qquad \frac{\inf J_\varepsilon-W_1}{\varepsilon} \to +\infty $$
which means that the lowest order of convergence of $\inf J_\varepsilon$ to $J$ is smaller as $\varepsilon$. The study of a precise example where this order is $\varepsilon|\log\varepsilon|$ is the object of the section \ref{sectionepslogeps}.
\paragraph{What about the selected map ?} The first-order $\Gamma$-convergence and the basic properties of $\Gamma$-limits imply that, if $T_\varepsilon$ minimizes $J_\varepsilon$, then $T_\varepsilon \to T$ which minimizes the Sobolev norm among the set $\O_1(\mu,\nu)$ of optimal transport maps from $\mu$ to $\nu$. This gives a selection principle, via secondary variational problem (minimizing something in the class of minimizers), in the same spirit of what we presented for the monotone transport map along each transport ray. A natural question is to find which is this new ``special'' selected map, and whether it can coincide with the monotone one. Thanks to the non-optimality results of this map for the Sobolev cost on the real line, the answer is that they are in general different. We can look at the following explicit counter-example (where we have however $\O_1(\mu,\nu)\cap H^1(\Omega) \neq \emptyset$).
Let us set $ \Omega = (0,1)^2$, $\Omega' = (2,3)\times(0,1)$ in $\mathbb{R}^2$. Let $F$, $G$ be two probability densities on the real line, supported in $(0,1)$ and $(2,3)$ respectively; we now consider the densities defined by
$$ f(x_1,x_2)=F(x_1) \qquad \text{and} \qquad g(x_1,x_2)=G(x_1)$$
Then, if $t$ is a transport map from $F$ to $G$ on the real line and $T(x_1,x_2) = (t(x_1),x_2)$, it is easy to check that $T$ sends the density $f$ onto $g$ and if $u(x_1,x_2) = x_1$ we have
$$ |T(x)-x| = |t(x_1)-x_1| = t(x_1)-x_1 = u(T(x_1,x_2))-u(x_1,x_2) $$
This proves that $T$ is optimal and $u$, which is of course 1-Lipschitz, is a Kantorovich potential, so that the maximal transport rays are exactly the segments $[0,3]\times\{x_2\}$, $0 < x_2 < 1$. As a consequence,
$$ T \in \O_1(\mu,\nu) \qquad \iff \qquad T(x_1,x_2) = (t(x_1,x_2),x_2) \quad \text{with} \quad \begin{array}{c} t(\cdot,x_2)_\# F = G \\ \text{for a.e. } x_2 \end{array} $$
In particular, the monotone transport map along the maximal transport rays is $x \mapsto (t(x_1),x_2)$, where $t$ is the non-decreasing transport map from $F$ to $G$ on the real line. For this transport map $T$, we have
$$ \int_\Omega |DT(x)|^2 \, \text{\textnormal{d}} x = \int_0^1 t'(x_1)^2 \, \text{\textnormal{d}} x_1 $$
Now, one can choose $F$, $G$ such that the solution of
\begin{equation} \inf\left\{ \int_0^1 U'(x_1)^2 \, \text{\textnormal{d}} x_1 \;:\; U_\#F = G \right\} \label{t1d} \end{equation}
is not attained by the increasing transport map from $F$, to $G$. Thus, if $\tilde{t}$ minimizes \eqref{t1d} and $\tilde{T}(x_1,x_2) = \tilde{t}(x_1)$, we have
$$ \int_\Omega |D\tilde{T}(x)|^2 \, \text{\textnormal{d}} x = \int_0^1 \tilde{t}'(x_1)^2 \, \text{\textnormal{d}} x_1 < \int_0^1 t'(x_1)^2 \, \text{\textnormal{d}} x_1 = \int_\Omega |DT(x)|^2 \, \text{\textnormal{d}} x $$
and $\tilde{T}$ is also an optimal transport map for the Monge problem.
\subsection{Proof of the density of Lipschitz transports}
In this section, we prove that, under natural assumptions (domains which are diffeomorphics to the unit ball, H\"older and bounded from above and below densities), the transport maps from $\mu$ to $\nu$ which are Lipschitz continuous form a dense subset of the set of transport maps from $\mu$ to $\nu$. Notice that, for the sake of the applications to the zero-th order $\Gamma$-convergence of the previous section, we only need the density of those maps belonging to the Sobolev space $H^1(\Omega)$; also, we needed density in the set of plans, but since it is well known that transport maps are dense in the set of transport plans, for simplicity we will prove density in the set of maps. This result, that we need in this paper to study the $\Gamma$-convergence of the functional $J_\varepsilon$, could of course be useful for other aims; however, it was, to the best of our knowledge, not investigated until now.
\begin{theo} \label{denssobolev} Assume that $\Omega$, $\Omega'$ are both Lipschitz polar domains and that $f$, $g$ are both $C^{0,\alpha}$ and bounded from below. Then the set
$$ \left\{ T \in \operatorname{Lip}(\Omega) \,:\, T_\#\mu = \nu \right\}$$
is non-empty, and is a dense subset of the set
$$ \left\{ T : \Omega \to \Omega' \,:\, T_\#\mu = \nu \right\} $$
endowed with the norm $||\cdot||_{L^2(\Omega)}$. \end{theo}
Notice that, exactly as for the density of transport maps inside the set of transport plans, the best possible result would be to show the density under the same assumptions guaranteeing the existence of at least such a map (for the density of maps into plans the assumption is that $\mu$ must be atomless, which is the same as for the existence of at least a map $T$). Here we are not so far since the already known results about regularity of some transport maps (by Caffarelli {\it et al.} or Dacorogna-Moser) deal with at least H\"older densities. Also, one can see from the proof below that we do not use much more of the existence of Lipschitz maps. Concerning the assumption on the domains, this is used to send them onto the unit ball in a Lipschiz-diffeomorphic way (in order to get sufficiently regular measures), as shown in the following lemma which is used several times:
\begin{lem} \label{lemdiffeo} If $U$ is a Lipschitz polar domain with
$$ U = \left\{ x \in \mathbb{R}^d \,:\, |x-x_0| < \gamma\left( \frac{x-x_0}{|x-x_0|}\right) \right\} $$
then there exists a map $\alpha:U\to B(0,1)$ such that:
\begin{itemize}
\item $\alpha$ is a bi-Lipschitz diffeomorphism from $U$ to $B(0,1)$;
\item $\det D\alpha$ is Lipschitz and bounded from below (thus, $\det D\alpha^{-1}$ is also Lipschitz);
\item for any $x \neq x_0$, $\dfrac{\alpha(x)}{|\alpha(x)|} = \dfrac{x-x_0}{|x-x_0|}$
\end{itemize} \end{lem}
(We will not use the third property in the proof of Theorem \ref{denssobolev}, but it will be useful later in section~\ref{construction}).
\begin{proof} Up to a translation and a dilation, we can assume $x_0=0$ and $\gamma \leq 1$ on $S^{d-1}$. Now we se
$$ \alpha(x) = \begin{cases} x \qquad \text{if } |x| \leq \dfrac 12 \gamma\left(\dfrac{x}{|x|}\right) \\ \lambda(x)\dfrac{x}{|x|} \qquad \text{otherwise} \end{cases}. $$
for a suitable choice of the function $\lambda: U \to [0,+\infty)$. Assuming that $\lambda$ is Lipschitz, we compute $ D\alpha$ on the region $\left\{|x|> \dfrac 12 \gamma(x/|x|) \right\}$. Here we have $D\alpha = x \otimes \nabla\left(\dfrac{\lambda(x)}{x}\right) + \dfrac{\lambda}{|x|} I_d $ thus, in an orthonormal basis whose first vector is $e = \dfrac{x}{|x|}$,
$$ D\alpha = \left(\begin{array}{llll}
|x| \partial_e \left(\dfrac{\lambda(x)}{|x|}\right) & |x| \partial_{e_2} \left(\dfrac{\lambda(x)}{|x|}\right) & \dots & |x|\left(\partial_{e_n} \dfrac{\lambda(x)}{|x|}\right) \\
0 & 0 & \dots 0 \\
\vdots & \vdots & \dots & \vdots \\
0 & 0 & \dots 0 \\
\end{array} \right) + \frac{\lambda}{|x|} I_n
$$
$$ \text{and} \qquad \det D\alpha = \left(|x|\partial_e \left(\dfrac{\lambda(x)}{|x|}\right)+\frac{\lambda(x)}{|x|}\right) \left(\frac{\lambda(x)}{|x|}\right)^{n-1} $$
We write $\lambda = \lambda(r,e)$ where $r = |x|$, which leads to
$$ \det D\alpha = \left(r \partial_r\left(\frac{\lambda}{r}\right)+\frac{\lambda}{r}\right)\left(\frac{\lambda}{r}\right)^{n-1} = \partial_r\lambda \left(\frac{\lambda}{r}\right)^{n-1}$$
At this time, we see that the following conditions on $\lambda$ allow to conclude:
\begin{itemize}
\item $\lambda$ is Lipschitz on the domain $\left\{ \frac{\gamma(x/|x|)}{2} \leq |x| \leq \gamma(x/|x|) \right\}$;
\item if we fix $e \in S^{d-1}$, then $\lambda(\cdot,e)$ is increasing on the interval $\left[\frac{\gamma(e)}{2},\gamma(e)\right]$ ;
\item $\lambda\left(\frac{\gamma(e)}{2},e\right) = \frac{\gamma(e)}{2}$ and $\lambda(\gamma(e),e) = 1$;
\item $\partial_r\lambda\left(\frac{\lambda }{r}\right)^{n-1}$ is Lipschitz and is equal to $1$ for $r=\frac{\gamma(e)}{2}$, which means $\partial_r\lambda\left(\frac{\gamma(e)}{2},e\right) = 2^{n-1}$.
\end{itemize}
To satisfy these conditions, it is enough to choose for $\lambda(\cdot,e)$ a second-degree polynomial function with prescribed values at $\frac{\gamma(e)}{2}$, $\gamma(e)$ and prescribed first derivative at $\frac{\gamma(e)}{2}$.
\end{proof}
\begin{proof}[Proof of Theorem \ref{denssobolev}] First, we assume $\Omega = \Omega' = \left[-\frac{1}{2},\frac{1}{2}\right]^d$ but we will also identify (for convolution purposes, see below) the cube $\Omega$ with a Torus. We will generalize later to any pair of Lipschitz star-shaped domains. Thus, let $T:\Omega\to\Omega$ sending $f$ to $g$, and let us build a sequence $(T_n)_n$ of Lipschitz maps such that
$$ T_n \xrightarrow[n\to+\infty]{L^2(\Omega)} T \qquad \text{and} \qquad \forall n \in \mathbb{N}, \; (T_n)_\#\mu = \nu $$
{\it Step 1: regularization of the transport plan.} We denote by $\gamma$ the transport plan associated to $T$; let us recall that it is defined by
$$ \iint_{\Omega^2} \phi(x,y) \, \text{\textnormal{d}}\gamma(x,y) = \int_{\Omega} \phi(x,T(x)) f(x) \, \text{\textnormal{d}} x $$
for any continuous and bounded function $\phi$ on $\Omega\times\Omega'$. We use the disintegration of measures (see Theorem 5.3.1 in \cite{ags}) to write
$$ \iint_{\Omega^2} \phi \, \text{\textnormal{d}}\gamma = \int_{\Omega} \left( \int_{\Omega} \phi(x,y) \, \text{\textnormal{d}}\gamma_x(y) \right) \, \text{\textnormal{d}}\mu(x) $$
for a family of measures $(\gamma_x)_{x\in \Omega}$ on $\Omega$. Now we take a family of periodic convolution kernels $(\rho_k)_k$,
we define, for each $x$, the measure $\gamma^k_x$ on $\Omega$ by $\gamma^k_x = \rho_k\star \gamma_x$, where the convolution product is taken in the distributional sense on the torus $\Omega'$, and set $\, \text{\textnormal{d}}\gamma_k(x,y) = \, \text{\textnormal{d}}\gamma^k_x(y) \otimes \, \text{\textnormal{d}}\mu(x)$; the fact that $(\rho_k)_k$ is an approximate identity guarantees that $\gamma_k \rightharpoonup \gamma$ for the weak convergence of measures.
On the other hand, notice that
\begin{align*} \iint_{\Omega^2} \phi(x,y) \, \text{\textnormal{d}}\gamma_k(x,y) & = \iiint_{\Omega^3} \phi(x,y) \rho_k(y-z) \, \text{\textnormal{d}}\gamma_x(z) f(x) \, \text{\textnormal{d}} x \, \text{\textnormal{d}} y
\\ & = \int_\Omega \left(\int_{\Omega^2} \phi(x,y) \rho_k(y-z) \, \text{\textnormal{d}}\gamma(x,z) \right) \, \text{\textnormal{d}} y
\\ & = \iint_{\Omega^2} \phi(x,y) \rho_k(y-T(x)) f(x) \, \text{\textnormal{d}} y \, \text{\textnormal{d}} x \end{align*}
for any test function $\phi$. Thus, $\gamma_k$ is absolutely continuous with respect to the Lebesgue measure on $\Omega^2$ and its density is $\rho_k(y-T(x)) f(x)$. In particular, this density is smooth with respect to $y$ and bounded from below by a positive constant $c_k$. Moreover, if we denote by $\nu_k$ the second marginal of $\gamma_k$, then $\nu_k$ has also a Lipschitz and bounded from below density.
{\it Step 2: construction of a transport map corresponding to the regularized transport plan.} The goal of this step is to build, for each $k$, a family of regular maps $(T^k_h)_h$ each sending the first marginal of $\gamma_k$ onto the second one and such that, for a suitable diagonal extraction, $ T^k_{h_k} \to T $ for the $L^2$-convergence.
We fix $k \in \mathbb{N}$ and a dyadic number $h$, and decompose $\Omega = \bigcup\limits_i Q_i$ into a union of cubes $Q_i$, each with size-length~$2h$, via a regular grid. For each $i$, we set
$$ \mu_i = \mu|_{Q_i} \qquad \text{and} \qquad \nu_i^k = (\pi_2)_\#(\gamma_k|_{(Q_i\times\Omega)}) $$
Notice that $\gamma_k$ has $\mu$ as first marginal, thus we have also $\mu_i = (\pi_1)_\#(\gamma_k|_{(Q_i\times\Omega)})$; in particular, $\mu_i$ (which is a measure on the cube $Q_i$) and $\nu_i^k$ (which is a measure on the whole cube $\Omega$) have both same mass. Moreover, $\mu_i$ still has $f$ as density and the density of $\nu_i^k$ is
$$ y \mapsto \int_{Q_i} \rho_k(T(x)-y)f(x) \, \text{\textnormal{d}} x $$
which is smooth and bounded from below by a positive constant. We finish this step by building a map $U_{i,h}^k : Q_i \to \Omega$ such that
\begin{itemize}
\item $U_{i,h}^k$ is Lipschitz continuous;
\item $U_{i,h}^k(x) = x$ for $x \in \partial Q_i$ (this is allowed since the source $Q_i$ is included in the target $\Omega$);
\item $U_{i,h}^k$ sends $\mu_i$ onto $\nu_i^k$.
\end{itemize}
Let us notice that the two first points guarantee that the global map defined by $T_h^k(x)=U_{i,h}^k(x)$
(which depends on the radius $h$ and the original integer $k$) for $x \in Q_i$ will be globally Lipschitz on $\Omega$ and the third point implies that $T_h^k$ sends globally the measure $\mu$ onto the measure $\nu_k=(\pi_2)_\#\gamma_k$ (which is equal to the sum of the $\nu_i^k$).
Now we explain very briefly the construction of such a $U_{i,h}^k$. For the sake of simplicity, we give the details only in the case where $Q_i$ is not an extremal cube of $\Omega$, {\it i.e.}~where $\partial Q_i \cap \partial \Omega$ is empty, and claim that the extreme case can be treated in a similar way.
We denote by $x_i$ the center of $Q_i$ which has radius $h$ (for the infinite norm in $\mathbb{R}^d$), and by $\delta_1$ and $\delta_2$ the two positive numbers such that
$$ \mu(\{x\,:\, \delta_1 \leq ||x-x_i||_\infty \leq h \}) = \mu(\{x\,:\, \delta_2 \leq ||x-x_i||_\infty \leq \delta_1 \}) = \frac{1}{2}\nu_i^k(\Omega\setminus Q_i) $$
(such positive numbers exist by the intermediate value theorem since $\delta \mapsto \mu(B_\infty(x_i,\delta))$ is continuous). Now we consider:
\begin{itemize}
\item a first Lipschitz map $V_1$ from $\{x: \delta_1 \leq ||x-x_i||_\infty \leq h \}$ to $\Omega\setminus Q_i$, with $(V_1)_\#\mu_i = \frac{1}{2}\nu_i^k (\Omega \setminus Q_i)$
and such that $V_1(x) = x$ for $||x-x_i|| = h$;
\item a second Lipschitz map $V_2$ from $\{x: \delta_2 \leq ||x-x_i||_\infty \leq \delta_1 \}$ to $\Omega\setminus Q_i$, such that $(V_2)_\#\mu_i = \frac{1}{2}\nu_i^k (\Omega \setminus Q_i)$ and $V_2(x) = V_1(x)$ for $||x-x_i||=\delta_1$;
\item a third Lipschitz map $V_3$ from $\{0: ||x-x_i||_\infty \leq \delta_2 \}$ to $Q_i$, such that $V_3(x)=V_2(x)$ on the boundary and $(V_3)_\#\mu = \nu_i^k (Q_i)$.
\end{itemize}
and we set $U_{i,h}^k(x) = V_1(x)$ if $ \delta_1 \leq ||x-x_i||_\infty \leq h$, $V_2(x)$ if $ \delta_2 \leq ||x-x_i||_\infty \leq \delta_1$ and $V_3(x)$ if $||x-x_i||_\infty \leq \delta_2$. The constraints on $(V_1)_\#\mu$, $(V_2)_\#\mu$, $(V_3)_\#\mu$ imply that $U_{i,h}^k$ defined on the whole $Q_i$ sends globally $\mu|_{Q_i}$ onto $\nu_i^k$, and the constraints on the boundary on each domain guarantee that $U_{i,h}^k$ is Lipschitz on $Q_i$. We construct $V_1$, $V_2$, $V_3$ using Dacorogna-Moser's result (a simple adaptation is needed, since the source and target domains are not the same here, but are diffeomorphic to each other: either they are both cubes, or they both have the form ``large cube minus small cube'').
In particular, $\gamma_k$
and $\gamma_{T_h^k}$ give both same mass to each $Q_i\times \Omega$ which have radius $h$. Now if $u$ is a $1$-Lipschitz function $\Omega\times\Omega' \to \mathbb{R}$, we compute
$$ \int_{\Omega^2} u(x,y) \, \text{\textnormal{d}}\gamma_{T_h^k}(x,y) - \int_{\Omega^2} u(x,y) \, \text{\textnormal{d}}\gamma_k(x,y) = \sum\limits_i \int_{Q_i\times\Omega} u(x,y) (\, \text{\textnormal{d}}\gamma_{T_h^k}-\, \text{\textnormal{d}}\gamma_k)(x,y) $$
Since $\gamma_k$ and $\gamma_{T_h^k}$ have both the same second marginal on each $Q_i\times \Omega$,
$$ \int_{Q_i\times\Omega} u(x_i,y) (\, \text{\textnormal{d}}\gamma_{T_h^k}(x,y)-\, \text{\textnormal{d}}\gamma_k(x,y)) = 0$$
where, for each $i$, $x_i$ denotes the center of $Q_i$. Thus
\begin{multline*} \int_{\Omega^2} u(x,y) \, \text{\textnormal{d}}\gamma_{T_h^k}(x,y) - \int_{\Omega^2} u(x,y) \, \text{\textnormal{d}}\gamma_k(x,y) = \sum\limits_i \int_{Q_i\times\Omega} (u(x,y)-u(x_i,y)) (\, \text{\textnormal{d}}\gamma_{T_h^k}-\, \text{\textnormal{d}}\gamma_k)(x,y)
\\\leq \sum\limits_i \int_{Q_i\times\Omega} |x-x_i| (\, \text{\textnormal{d}}\gamma_{T_h^k}-\, \text{\textnormal{d}}\gamma_k)(x,y) \end{multline*}
where in the last inequality we used the fact that $u$ is supposed to be $1$-Lipschitz. We deduce
$$ \int_{\Omega^2} u(x,y) \, \text{\textnormal{d}}\gamma_{T_h^k}(x,y) - \int_{\Omega^2} u(x,y) \, \text{\textnormal{d}}\gamma_k(x,y) \leq h \iint_{\Omega^2} (\, \text{\textnormal{d}}\gamma_{T_h^k}-\, \text{\textnormal{d}}\gamma_k) \leq 2h $$
This inequality holds for any $1$-Lipschitz function $u:\Omega^2\to\mathbb{R}$. Thus, if $W_1$ denotes the classical Wasserstein distance (on $\Omega^2$: this is just a technical trick to metrize the weak convergence of plans), we deduce $W_1(\gamma_k,\gamma_{T_h^k}) \leq h$ for any $k$. Since we know that $\gamma_k \to \gamma_T$ as $k\to+\infty$, we also have $W_1(\gamma_k,\gamma_T) \to 0$ and if we set $h = 1/{2^k}$ we have $W_1(\gamma_{T_h^k},\gamma_T) \to 0$. Thanks to the lemma \ref{plansandmaps}, we deduce $T_h^k \to T$ for the $L^2$-norm in $\Omega$.
{\it Step 3: rearranging $T_h^k$.}
It remains to compose each map $T_h^k$ with a transport map $U_k$ from $\nu_k$ to $\nu$, which will ensure the obtained map to send $\mu$ onto $\nu$ as well. We use the following classical result of Caffarelli (\cite{vil}, Theorem 4.14): if
\begin{itemize}
\item $X$, $Y$ are two bounded open sets of $\mathbb{R}^d$, uniformly convex and with $C^2$ boundary;
\item $f \in C^{0,\alpha}(\bar{X})$, $g \in C^{0,\alpha}(\bar{Y})$ are two probability densities
\end{itemize}
then the optimal transport map from $f$ to $g$ belongs to $C^{1,\alpha}(\bar{X})$. The assumption on the source domain is of course not satisfied here, thus we use the lemma \ref{lemdiffeo} to get a bi-Lipschitz map $\alpha:\Omega\to B(0,1)$ such that $\det D\alpha$ is also Lipschitz. We then denote by
$$ \bar{g} = \alpha_\# g \qquad \text{and} \qquad \bar{g}_k = \alpha_\# g_k $$
where $g_k$ is the density of $\nu_k$ (we know that $g_k$ is Lipschitz and bounded from below). The fact that $\det D\alpha$ is Lipschitz and bounded guarantees that $\bar{g}$ is H\"older and $\bar{g}_k$ is Lipschitz as well.
We know that $g_k \rightharpoonup g$ in the weak sense of measures, thus $\bar{g}_k \rightharpoonup \bar{g}$ also, and
$$ W_2(\bar{g}_k,\bar{g}) = \int_{B(0,1)} |U_k(x)-x|^2 \bar{g}_k(x) \, \text{\textnormal{d}} x \to 0 $$
where $U_k$ is the optimal transport map from $\bar{g}_k$ to $\bar{g}$ for the quadratic cost; the Caffarelli's regularity result guarantees $U_k$ to belong to $C^{1,\alpha}(\bar{B}(0,1))$, and we deduce from the convergence $W_2(\bar{g}_k,{g}) \to 0$ that $U_k(x) \to x$ for almost any $x \in B(0,1)$. Now if we consider
$$ \tilde{T}_k = \alpha^{-1}\circ U_k \circ \alpha \circ T_h^k $$
then we can check that $\tilde{T}_k \to T$ in $L^2(\Omega)$, and we have $(\tilde{T}_k)_\#\mu = \nu$ for each $k$ by construction. The proof is complete in the case $\Omega=\Omega'=\left[-\frac{1}{2},\frac{1}{2}\right]^d$.
{\it Step 4: generalization to any pair of domains.}
If $\Omega$, $\Omega'$ are now two star-shaped domains with Lipschitz boundaries and $T$ is a transport maps from $\mu$ to $\nu$, we consider two Lipschitz diffeomorphisms $\alpha_1:\Omega\to[0,1]^d$, $\alpha_2:\Omega'\to[0,1]^d$ as in the Lemma \ref{lemdiffeo}. The regularity of $\det D\alpha_1$, $\det D\alpha_2$ guarantees that the image measures $(\alpha_1)_\#\mu$, $(\alpha_2)_\#\nu$ have both $C^{0,\alpha}$ regularity, thus we are able to find a sequence $(U_k)_k$ of transport maps from $(\alpha_1)_\#\mu$ to $(\alpha_2)_\#\nu$ converging for the $L^2([0,1]^d)$-norm to $(\alpha_2)\circ T \circ (\alpha_1)^{-1}$, and it is now easy to check that $T_k = (\alpha_2)^{-1}\circ U_k\circ\alpha_1$ sends $\mu$ to $\nu$ and converges to $T$ as well \end{proof}
\section{An example of approximation of order $\varepsilon|\log\varepsilon|$ \label{sectionepslogeps}}
As we said in the above section, the convergence $\frac{J_\varepsilon-W_1}{\varepsilon} \to \mathcal H$ allows to know the behavior of $\inf J_\varepsilon$ and of any family $(T_\varepsilon)_\varepsilon$ of minimizers in the case where $\inf\mathcal H<+\infty$. This paragraph is devoted to the study of an example where this assumption fails ({\it i.e.}~where any optimal transport map for the Monge problem is not Sobolev).
\subsection{Notations and first remarks}
In the rest of this paper, we set $d = 2$ and we will denote by $(r,\theta)$ the usual system of polar coordinates in $\mathbb{R}^2$. Our source and target domains will be respectively
$$ \Omega = \left\{ x = (r,\theta) : 0 < r < 1, 0 < \theta < \frac{\pi}{2} \right\} $$
$$ \text{and} \qquad \Omega' = \left\{ x = (r,\theta) : R_1(\theta) < r < R_2(\theta), 0 < \theta < \frac{\pi}{2} \right\} $$
where $R_1$, $R_2$ are two Lipschitz functions $\left[0,\frac{\pi}{2}\right] \to (0,+\infty)$ with $\inf R_1 > 1$ and $\inf R_2 > \sup R_1$. We also suppose $R_1$ to be such that $R_1'$ is a function of bounded variation
Notice that we can choose $R_1$, $R_2$ such that the target domain $\Omega'$ is convex (for instance if the curves $r=R_1(\theta)$, $r=R_2(\theta)$ are actually two lines in the quarter plane). We assume that $f$ and $g$ are two Lipschitz densities on $\bar{\Omega}$, $\bar{\Omega'}$, bounded from above and below by positive constants, with the following hypothesis:
\begin{equation} \forall \theta \in \left(0,\frac{\pi}{2}\right), \, \int_0^1 f(r,\theta) \, r \, \text{\textnormal{d}} r = \int_{R_1(\theta)}^{R_2(\theta)} g(r,\theta) \, r \, \text{\textnormal{d}} r \label{egdensity} \end{equation}
which means that, for any $\theta$, the mass (with respect to $f$) of the segment joining the origin to the boundary of $\Omega$ and with angle $\theta$ is equal to the mass (with respect to $g$) of the segment with same angle joining the ``above'' and ``below'' boundaries of $\Omega'$ ({\it i.e.}~the curves $r=R_1(\theta)$ and $r=R_2(\theta)$).
Then the structure of the optimal maps for the Monge cost is given by the following:
\begin{prop} \label{prop1} Under these assumptions on $\Omega$, $\Omega'$, $f$, $g$, the Euclidean norm is a Kantorovich potential and the maximal transport rays are the segments joining $0$ to $(R_2(\theta),\theta)$. Consequently,
$$ T \in \O_1(\mu,\nu) \quad \iff \quad \begin{array}{c} T_\# f = g \\[1mm] \text{and } \; T(x) = \phi(x) \dfrac{x}{|x|} \end{array} $$
for some function $\phi \in \Omega \to (\inf R_1,+\infty)$. \end{prop}
\begin{proof} For $\theta \in \left(0,\frac{\pi}{2}\right)$, we denote by $t(\cdot,\theta)$ a one-dimensional transport map from the measure $r \mapsto rf(r,\theta)$ to the measure $r \mapsto rg(r,\theta)$ (such a transport map exists since these two measures have same mass thanks to the equality \eqref{egdensity}).
It is then easy to check that the map
$$ T : x = (r,\theta) \in \Omega \to t(r,\theta) \frac{x}{|x|} $$
is a transport map from $f$ to $g$,
and if we set $u=|\cdot|$, $u$ is of course $1$-Lipschitz and
$$ u(T(x))-u(x) = \left|t(r,\theta)\frac{x}{|x|}\right|-|x| = (|t(r,\theta)|-|x|) \frac{x}{|x|} =
\left|(t(r,\theta)-|x|)\frac{x}{|x|}\right|
= |T(x)-x| \qquad \text{for any } x\in \Omega. $$
We deduce that that $u = |\cdot|$ is a Kantorovich potential. Consequently, a segment $[x,y]$ is a transport ray if and only if
$$ u(y)-u(x) = |y-x| \qquad i.e. \qquad |y-x|=|y|-|x|$$
Thus, we have $y = \lambda x$ for a positive $\lambda$. In particular, $y$ and $x$ belong to the same line passing through the origin. In other words, the transport rays are included in the lines passing through the origin. Moreover, a transport map $T$ belongs to $\O_1(\mu,\nu)$ if and only if, for a.e.~$x~\in\Omega$, $ |T(x)-x| = |T(x)|-|x|$ which again means that $T(x) = \phi(x)\dfrac{x}{|x|}$ for some positive function $\phi$. \end{proof}
\begin{co} Under the above assumptions on $\Omega$, $\Omega'$, $\mu$, $\nu$, we have $ \O_1(\mu,\nu) \cap H^1(\Omega) = \emptyset $ \end{co}
\begin{proof} Let $T \in \O_1(\mu,\nu)$. By Prop.~\ref{prop1}, we have $T(x) = \phi(r,\theta)\dfrac{x}{|x|}$ for $x = (r,\theta)$, where $\phi$ is a real-valued function; the fact that $T$ sends $\mu$ onto $\nu$ implies $T(x) \in \Omega'$ for any $x$, thus $\phi$ is bounded from below on $\Omega$ by the lower bound of $R_1$. We now compute the Jacobian matrix of $T$. Denoting by $x^\perp$ the image of $x$ by the rotation with angle $\pi/2$, we have in the basis $\left(\frac{x}{|x|},\frac{x^\perp}{|x|}\right)$:
$$ DT(x) = \frac{x}{|x|}\otimes\nabla \phi(x)-\frac{\phi(x)}{|x|}I_d = \left( \begin{array}{cc} \partial_r\phi & 0 \\[2mm] \dfrac{\partial_\theta\phi}{r} & \dfrac{\phi}{r} \end{array} \right) $$
$$\text{thus} \qquad \int_\Omega |DT(x)|^2 \, \text{\textnormal{d}} x \geq \int_0^1\int_0^{\frac{\pi}{2}} \frac{\phi(r,\theta)^2}{r^2} \, r \, \text{\textnormal{d}}\theta\text{d}r \geq \int_0^{1} \frac{\pi}{2}\left( \inf R_1\right)^2 \frac{\text{d}r}{r} = +\infty \qedhere $$\end{proof}
\subsection{Heuristics}
In this paragraph, we give a preliminary example of analysis of the behavior of $J_\varepsilon(T_\varepsilon)$ when $\varepsilon \to 0$ and $T_\varepsilon$ approaches an optimal map $T$ for the Monge problem; this will not lead directly to a rigorous proof of the general result, but gives an idea of which quantities will appear.
Assume that $T \in \O_1(\mu,\nu)$ with $T(x) = \phi(r,\theta) \dfrac{x}{|x|}$, and let us build an approximation $(T_\varepsilon)_\varepsilon$ defined~by
$$ T_\varepsilon(x) = \left\{ \begin{array}{ll} S(x) & \text{if } x \in \Omega_\delta \\ T(x) & \text{otherwise} \end{array}\right. $$
where $\delta$ will be fixed (depending on $\varepsilon$), $\Omega_\delta = B(0,\delta) \cap \Omega$ and $S$ will be build to send $(f\cdot\mathcal{L}^d)|_{\Omega_\delta}$ onto the same image measure that the original $T$ has on $\Omega_\delta$. Notice that $S$ depends indeed both on $\delta$ and on $\varepsilon$, but we omit this dependance. In this case, we have
$$ J_\varepsilon(T_\varepsilon) - W_1 = \int_\Omega |T_\varepsilon(x)-x| f(x) \, \text{\textnormal{d}} x - W_1 + \int_{\Omega_\delta} |DS|^2 + \int_{\Omega\setminus \Omega_\delta} |DT|^2 $$
Since $T$ is optimal for the Monge problem and coincides with $T_\varepsilon$ outside of $\Omega_\delta$, we have
$$ \int_{\Omega} |T_\varepsilon(x)-x|f(x)\, \text{\textnormal{d}} x - W_1 = \int_\Omega (|T_\varepsilon(x)-x|-|T(x)-x|) f(x) \, \text{\textnormal{d}} x = \int_{\Omega_\delta}(|S(x)-x|-|T(x)-x|) f(x) \, \text{\textnormal{d}} x $$
We now claim that
$$ \int_{\Omega_\delta} |T(x)-x| f(x) \, \text{\textnormal{d}} x = \int_{\Omega_\delta} (|S(x)|-|x|) f(x) \, \text{\textnormal{d}} x $$
Indeed, we still have the equality $|T(x)-x| = |T(x)|-|x|$, and the image measures of $(f\cdot\mathcal{L}^d)|_{\Omega_\delta}$ by $T$ and $S$ are the same.
As a consequence,
$$ \int_{\Omega} |T_\varepsilon(x)-x|f(x)\, \text{\textnormal{d}} x - W_1 = \int_{\Omega_\delta} (|S(x)-x|-|S(x)|+|x|) f(x) \, \text{\textnormal{d}} x $$
and, by the triangle inequality, $|S(x)-x|-|S(x)|+|x| \leq |(S(x)-x)-S(x)|+|x| \leq 2|x|$ so that
$$ \int_{\Omega} |T_\varepsilon(x)-x|f(x)\, \text{\textnormal{d}} x - W_1 \leq \int_{\Omega_\delta} 2|x| \, \text{\textnormal{d}} x \leq 2\int_0^{\frac{\pi}{2}} \int_0^\delta r^2 \, \text{\textnormal{d}} r \, \text{\textnormal{d}} \theta \leq \frac{\pi\delta^3}{3} $$
In order to estimate the norm of the Jacobian matrix $DT$ outside of $\Omega_\delta$, we recall that, in the basis~$\left(\dfrac{x}{|x|},\dfrac{x}{|x|}^{\!\perp}\right)$,
$$ DT(x) = \left( \begin{array}{cc} \partial_r\phi & 0 \\[2mm] \dfrac{\partial_\theta\phi}{r} & \dfrac{\phi}{r} \end{array} \right) $$
$$ \text{so that} \qquad \int_{\Omega\setminus \Omega_\delta} |DT(x)|^2 \, \text{\textnormal{d}} x = \int_\delta^1 \int_0^{\frac{\pi}{2}} \left( \frac{\phi(r,\theta)^2+\partial_\theta\phi(r,\theta)^2}{r} + r \partial_r\phi(r,\theta)^2 \right) \, \text{\textnormal{d}} \theta \, \text{\textnormal{d}} r $$
Now we note that, for $\theta \in (0,\pi/2)$, the one-dimensional map $\phi(\cdot,\theta)$ sends the density $rf(\cdot,\theta)$ onto the density $rg(r,\theta)$. The first density is bounded from above (but vanishes around $r=0$), and the second one from below (also from above). Thus, one can assume that $\partial_r\phi(\cdot,\theta)$ is bounded (by the way, if $T$ is the monotone map along the transport rays, we also have $\partial_r\phi(r,\theta) \to 0$ as $r \to 0$), so that
$$ \int_0^{\frac{\pi}{2}}\int_\delta^1 \partial_r\phi(r,\theta)^2 r \, \text{\textnormal{d}} r \, \text{\textnormal{d}} \theta \leq C $$
where $C$ is a constant independent of $\delta$. On the other hand,
$$ \int_\delta^1 \int_0^{\frac{\pi}{2}} \frac{\phi(r,\theta)^2+\partial_\theta\phi(r,\theta)^2}{r} \, \text{\textnormal{d}} \theta \, \text{\textnormal{d}} r = \int_\delta^1 ||\phi(r,\cdot)||_{H^1(0,\pi/2)}^2 \,\frac{\text{d}r}{r} $$
In this last integral, we make the change of variable $r = \delta^t$, which gives
$$ \int_\delta^1 \int_0^{\frac{\pi}{2}} \frac{\phi(r,\theta)^2+\partial_\theta\phi(r,\theta)^2}{r} \, \text{\textnormal{d}} \theta \, \text{\textnormal{d}} r = |\log\delta| \int_0^1 ||\phi(\delta^t,\cdot)||_{H^1(0,\pi/2)}^2 \, \text{\textnormal{d}} t $$
We moreover assume that $||\phi(\delta^t,\cdot)||_{H^1(0,\pi/2)}^2 \to ||\phi(0,\cdot)||_{H^1(0,\pi/2)}^2$ as $\delta \to 0$. This leads to
$$ \varepsilon\, \int_{\Omega\setminus \Omega_\delta} |DT|^2 \mathop{\sim}\limits_{\varepsilon,\delta \to 0} \varepsilon|\log\delta| ||\phi(0,\cdot)||_{H^1(0,\pi/2)}^2 + O(\delta^3+\varepsilon)$$
It remains to estimate the $L^2$-norm of the Jacobian matrix of $S$ on $\Omega_\delta$. We recall that $S$ has to be built so that $T_\varepsilon$, defined on the whole $\Omega$, is still a transport map from $\mu$ to $\nu$ with finite Sobolev norm; thus, the map $S$, defined on $\Omega_\delta$, must send $\Omega_\delta$ onto its original image $S(\Omega_\delta)$ in a regular way and keep the constraint on the image measures:
$$ S_\#(\mu|_{\Omega_\delta}) = T_\#(\mu|_{\Omega_\delta}) $$
Moreover, the regularity of the global map $T_\varepsilon$ implies a compatibility condition at the boundary:
$$ S(x) = T(x) \qquad \text{for } |x|=\delta $$
Thanks to the Dacorogna-Moser's result, we are indeed able to build such a map $S$. Yet, the diameter of $\Omega_\delta$ is $\sqrt{2}\delta$; on the other hand, $T(\Omega_\delta)$ contains the whole curve $\theta \mapsto \phi(0,\theta)$, so that its diameter is bounded from below by a positive constant independent of $\delta$. Thus, the best estimate that one can expect~is
$$ \operatorname{Lip} S \leq \frac{C}{\delta} $$
For a reasonable transport map $T$, one can show that such a map $S$ can be found with moreover $S(x)=T(x)$ for $|x|=\delta$ (see the paragraph \ref{construction} below). In this case, the global map $T_\varepsilon$ still sends $\mu$ to $\nu$ and we~have
$$ \int_{\Omega_\delta} |DT|^2 \leq \int_{\Omega_\delta} \left(\frac{C}{\delta}\right)^2 \leq C^2 $$
Finally,
$$ J_\varepsilon(T_\varepsilon)-W_1 = \varepsilon|\log\delta| ||\phi(0,\cdot)||_{H^1(0,\pi/2)} + O(\delta^3+\varepsilon) $$
If we choose $\delta = \varepsilon^{1/3}$, we obtain
$$ J_\varepsilon(T_\varepsilon) = W_1 + \varepsilon|\log\varepsilon| \frac{1}{3}||\phi(0,\cdot)||^2_{H^1(0,\pi/2)} + O(\varepsilon) $$
In particular:
\begin{itemize}
\item the first order of convergence of $J_\varepsilon(T_\varepsilon)$ to $W_1$ is not anymore $\varepsilon$, but $\varepsilon|\log\varepsilon|$;
\item the first significative term only involves the behavior of $\phi$ around $0$, which is the crossing point of all the transport rays (and also the singularity point of the measure restricted to any transport ray, since it vanishes at $0$). This suggests that $T_\varepsilon \to T$, where $T$ at $r=0$ minimizes $||\phi(r,\cdot)||_{H^1}$.
\end{itemize}
\subsection{Main result and consequences \label{parpresquegamma}}
The analysis in the above paragraph suggests to introduce the minimal value of $||\phi(0,\cdot)||_{H^1(0,\pi/2)}$ among the functions $\phi$ such that
$$ x \mapsto \phi(r,\theta)\frac{x}{|x|} $$
is a transport map from $\mu$ to $\nu$. In particular, for such a $\phi$ and for any $x \in \Omega$, the point $\phi(r,\theta) \dfrac{x}{|x|}$ still belongs to the target domain $\Omega'$; thus, its value at $r=0$ verifies
$$ \text{for a.e.}~\theta \in (0,\pi/2), \quad R_1(\theta) \leq \phi(0,\theta) \leq R_2(\theta) $$
We will thus set
$$ K = \min\left\{ \int_0^{\frac{\pi}{2}} (\phi(\theta)^2+\phi'(\theta)^2) \, \text{\textnormal{d}}\theta \,:\, \phi \in H^1\left(0,\frac{\pi}{2}\right), \, R_1(\theta) \leq \phi(\theta) \leq R_2(\theta) \right\} $$
and, following the ideas of the last sub-section,
$$ F_\varepsilon : T \mapsto \frac{1}{\varepsilon} \left(J_\varepsilon(T)-W_1-\frac{K}{3}\varepsilon|\log\varepsilon|\right) $$
Notice here that the assumption $\inf R_2 > \sup R_1$ allows to remove the constraint $\phi(\theta) \leq R_2(\theta)$ in the minimization problem defining $K$ (indeed, if $\phi > R_2$ on a subset of $(0,\pi/2)$, one can replace it with $\min(\phi,\inf R_2)$ on this subset and this will decrease its Sobolev norm): in other words, we have
$$ K = \min\left\{ \int_0^{\frac{\pi}{2}} (\phi(\theta)^2+\phi'(\theta)^2) \, \text{\textnormal{d}}\theta \,:\, \phi \in H^1\left(0,\frac{\pi}{2}\right), \, R_1(\theta) \leq \phi(\theta) \right\} $$
The main result of this paper is the following:
\begin{theo} \label{PRESQUEGAMMA} Let us denote by
$$ G : \phi \in H^1(0,\pi/2) \mapsto ||\phi||_{H^1(0,\pi/2)}^2-K $$
and
$$ F(T) = \left\{ \begin{array}{l} +\infty \qquad \text{if } T \notin \O_1(\mu,\nu) \\[2mm] \displaystyle\int_0^1 G(\phi(r,\cdot)) \dfrac{\text{\textnormal{d}}r}{r} + \displaystyle\int_0^1 ||\partial_r\phi(r,\cdot)||_{L^2}^2 \, r \, \text{\textnormal{d}} r \qquad \text{if } T \in \O_1(\mu,\nu), T(x) = \phi(r,\theta)\dfrac{x}{|x|} \end{array} \right. $$
\begin{enumerate}
\item For any family of maps $(T_\varepsilon)_\varepsilon$ such that $(F_\varepsilon(T_\varepsilon))_\varepsilon$ is bounded, there exists a sequence $\varepsilon_k \to 0$ and a map $T$ such that $T_{\varepsilon_k} \to T$ in~$L^2(\Omega)$.
\item There exists a constant $C$, depending only on the domains $\Omega$, $\Omega'$ and of the measures $f$, $g$, so that, for any family of maps $(T_\varepsilon)_{\varepsilon>0}$ with $T_\varepsilon \to T$ as $\varepsilon \to 0$ in $L^2(\Omega)$, we~have
$$ \liminf\limits_{\varepsilon \to 0} F_\varepsilon(T_\varepsilon) \geq F(T)-C $$
\item Moreover, there exists at least one family $(T_\varepsilon)_{\varepsilon > 0}$ such that $(F_\varepsilon(T_\varepsilon))_\varepsilon$ is indeed bounded. \end{enumerate} \end{theo}
Notice
that we have not stated here a true $\Gamma$-convergence result, but we only provide an estimate on the $\Gamma$-liminf, and the existence of a sequence with equibounded energy. We conjecture indeed that the $\Gamma$-limit of the sequence $F_\varepsilon$ is of the form $F-C$ for a suitable constant $C$ depending on the shape of $\Omega'$ and $f(0)$ (again, the main important region is that around $x=0$ in $\Omega$, which must be sent on the curve~$\Phi$); precisely, the constant $C$ that we expect is
$$ C = \lim\limits_{\lambda \to +\infty} \left( \inf\left\{ \lambda^3 f(0) \int_{\Omega} \left(|y|-y\cdot \frac{S(y)}{|S(y)|}\right) \, \text{\textnormal{d}} y +\int_\Omega|DS(y)|^2 \, \text{\textnormal{d}} y - K \log\lambda \,:\, S : \Omega_1 \to \Omega' \right\} \right) $$
with may be some additional constraint in the minimization problem for each $\lambda$ (see below a justification at the end of the paragraph \ref{gammaliminfest}).
However, we do not prove it here and we only use the estimate that we are really able to prove to get some consequences on the minima and the minimizers of $F_\varepsilon$.
\paragraph{Consequences on the minimal value of $J_\varepsilon$.} If we apply the Theorem~\ref{PRESQUEGAMMA} to a sequence $(T_\varepsilon)_\varepsilon$ where each $T_\varepsilon$ minimizes $J_\varepsilon$ (which is equivalent with minimizing $F_\varepsilon$), we obtain that the sequence $(F_\varepsilon(T_\varepsilon))_\varepsilon$ is bounded and
$$ \inf J_\varepsilon = F_\varepsilon(T_\varepsilon) = W_1(\mu,\nu)+\frac{K}{3}\varepsilon|\log\varepsilon|+O(\varepsilon) $$
which is both the order in $\varepsilon$ and the constant $K$ which appeared at the end of the above paragraph. Notice that a full knowledge of the $\Gamma$-limit would allow to compute the constant in the term $O(\varepsilon)$.
\paragraph{Consequences on the behavior of $(T_\varepsilon)_\varepsilon$.} We first note the following:
\begin{lem} \label{phimoinsphi} The problem which defines the constant $K$, namely
\begin{equation} \inf\left\{ \int_0^{\frac{\pi}{2}} (\phi(\theta)^2+\phi'(\theta)^2) \, \text{\textnormal{d}}\theta \,:\, \phi \in H^1\left(0,\frac{\pi}{2}\right), \, R_1(\theta) \leq \phi(\theta) \leq R_2(\theta) \right\} \, , \label{pbphimoinsphi} \end{equation}
admits a unique minimizer $\Phi$. Moreover, if $\phi \in H^1(0,\pi/2)$ verifies $R_1 \leq \phi$, then
$$G(\phi) \geq ||\phi-\Phi||_{H^1}^2 $$ \end{lem}
\begin{proof} Let us denote by $\mathcal{C}$ the set of functions $\phi \in H^1(0,\pi/2)$ verifying the constraint $R_1 \leq \phi$ on $(0,\pi/2)$. Notice that $\mathcal{C}$ is a convex closed subset of $H^1(0,\pi/2)$, so that \eqref{pbphimoinsphi} admits as well a unique minimizer $\Phi$, which is the orthogonal projection of $0$ onto $\mathcal{C}$ in the Hilbert space $H^1(0,\pi/2)$. This implies
\begin{equation} \forall \phi \in \mathcal{C}, \, \langle \Phi,\phi-\Phi \rangle \geq 0 \label{projh1} \end{equation}
If now $\phi \in \mathcal{C}$, then
$$ G(\phi) - ||\phi-\Phi||_{H^1}^2 = ||\phi||_{H^1}^2-||\Phi||_{H^1}^2- ||\phi-\Phi||_{H^1}^2 = 2\, \langle \Phi,\phi-\Phi \rangle $$
which is non-negative thanks to the inequality \eqref{projh1}. \end{proof}
As a consequence, we obtain:
\begin{prop} \label{PhiL2fort} Let $T \in \O_1(\mu,\nu)$, $T(x) = \phi(r,\theta)\dfrac{x}{|x|}$, such that $F(T) < +\infty$. Then $r \mapsto \phi(r,\cdot)$ is continous from $[0,1]$ to $L^2(0,\pi/2)$, and we have $\phi(0,\cdot)=\Phi$. \end{prop}
This, combined with Theorem~\ref{PRESQUEGAMMA}, implies that that if $(T_\varepsilon)_\varepsilon$ has an equi-bounded energy (meaning that $F_\varepsilon(T_\varepsilon)$ is uniformly bounded in $\varepsilon$), then it has, up to a subsequence, a limit $T = \phi(r,\theta)\frac{x}{|x|}$ where $\phi(r,\theta)$ is continuous with respect to $r$ and has $\Phi(\theta)$ as limit as $r\to 0$. In other words, $T$ sends $0$ onto the curve $r = \Phi(\theta)$ which has the best $H^1$-norm among the curves with values in the target domain $\Omega'$. This is in particular true if each $T_\varepsilon$ minimizes $J_\varepsilon$ (thus $F_\varepsilon$).
\begin{proof}[Proof of Prop.~\ref{PhiL2fort}] The assumption on $T$ implies that the integrals
$$\displaystyle\int_0^1 ||\partial_r\phi(r,\cdot)||_{L^2(0,\pi/2)}^2 \, r \, \text{\textnormal{d}} r \qquad \text{and} \qquad \displaystyle\int_0^1 ||\partial_r\phi(r,\cdot)||_{L^2(0,\pi/2)}^2 \, r \, \text{\textnormal{d}} r$$
are both controlled by some finite constant $A$. Now we have for $\theta \in (0,\pi/2)$:
$$ |\phi(r_1,\theta)-\phi(r_2,\theta)| = \int_{r_1}^{r_2} \partial_r\phi(r,\theta) \, \text{\textnormal{d}} r \leq \left(\int_{r_1}^{r_2} \partial_r\phi(r,\theta)^2 \, r \, \text{\textnormal{d}} r \right)^{1/2} \, \left(\int_{r_1}^{r_2} \frac{\text{d}r}{r} \right)^{1/2} $$
$$\text{thus} \qquad \int_0^{\pi/2} |\phi(r_1,\theta)-\phi(r_2,\theta)|^2 \, \text{\textnormal{d}}\theta \leq \left(\int_{r_1}^{r_2} ||\partial_r\phi(r,\cdot)||_{L^2(0,\pi/2)}^2 \, r \, \text{\textnormal{d}} r\right) \left( \int_{r_1}^{r_2} \frac{\text{d}r}{r} \right) $$
\begin{equation} \text{so that} \qquad ||\phi(r_1,\cdot)-\phi(r_2,\cdot)||_{L^2(0,\pi/2)}^2 \, \text{\textnormal{d}}\theta \leq A\log\frac{r_2}{r_1} \label{loglip} \end{equation}
This proves the continuity of $r \mapsto \phi(r,\cdot)$.
On the other hand, thanks to the Lemma~\ref{phimoinsphi}, we have
$$ A \geq \int_0^1 G(\phi(r,\cdot)) \frac{\text{d}r}{r} \geq \int_0^1 ||\phi(r,\cdot)-\Phi||_{L^2(0,\pi/2)}^2 \frac{\text{d}r}{r} $$
By setting $r = e^{-t}$, we obtain
$$ A \geq \int_0^{+\infty} ||\phi(e^{-t},\cdot)-\Phi||_{L^2}^2 \, \text{\textnormal{d}} t $$
But, for $t_1 < t_2 \in (0,+\infty)$, we have
$$ \begin{array}{ll} ||\phi(e^{-t_1},\cdot)-\Phi||_{L^2}^2-||\phi(e^{-t_2},\cdot)-\Phi||_{L^2}^2| & = |\langle \phi(e^{-t_1},\cdot)-\phi(e^{-t^2},\cdot), \Phi \rangle_{L^2}| \\[1mm]
& \leq ||\phi(e^{-t_1},\cdot)-\phi(e^{-t_2},\cdot)||\,||\Phi|| \\[1mm]
& \leq A\log\dfrac{e^{-t_2}}{e^{-t_1}} = A(t_2-t_1) \end{array} $$
where the last inequality comes from \eqref{loglip}. Thus, the function $t \mapsto ||\phi(e^{-t},\cdot)-\Phi||_{L^2}^2$ is Lipschitz and belongs to $L^1(0,+\infty)$. This implies that it vanishes at $+\infty$, so that $\phi(r,\cdot) \to \Phi$ in $L^2(0,\pi/2)$ as $r \to 0$. \end{proof}
\section{Proof of Theorem \ref{PRESQUEGAMMA}}
\subsection{$\Gamma$-liminf estimate \label{gammaliminfest}}
Let $(T_\varepsilon)_\varepsilon$ be a family of maps and let us begin by writing precisely the expression of $F_\varepsilon(T_\varepsilon)$ for any $\varepsilon > 0$ and any transport map $T_\varepsilon$. We have
$$F_\varepsilon(T_\varepsilon) = \frac{1}{\varepsilon} \left(\int_\Omega|T_\varepsilon(x)-x|f(x)\, \text{\textnormal{d}} x - W_1\right) + \int_\Omega |DT_\varepsilon(x)|^2\, \text{\textnormal{d}} x - \frac{K}{3}|\log\varepsilon| $$
We set $\delta = \varepsilon^{1/3}$ and notice that
$$ \frac{K}{3}|\log\varepsilon| = K|\log\delta| = K\int_\delta^1 \frac{\text{d}r}{r} $$
We decompose $T_\varepsilon$ into radial and tangential components:
$$ T_\varepsilon(x) = \phi_\varepsilon(r,\theta)\frac{x}{|x|}+\psi_\varepsilon(r,\theta)\frac{x}{|x|}^{\!\perp} $$
and compute
$$ DT_\varepsilon = \frac{x}{|x|}\otimes\nabla\phi_\varepsilon(x)+\frac{\phi_\varepsilon(x)}{|x|}\left(I_d-\frac{x}{|x|}\otimes\frac{x}{|x|}\right)+\frac{x}{|x|}^{\!\perp}\otimes\nabla\psi_\varepsilon(x)+\frac{\psi_\varepsilon(x)}{|x|}\left(R-\frac{x}{|x|}^{\!\perp}\otimes\frac{x}{|x|}\right) $$
where $R$ denotes the rotation with angle $\pi/2$ and we still set $x^\perp = Rx$. Thus, the matrix of $DT_\varepsilon$ in the basis~$\left(\dfrac{x}{|x|},\dfrac{x}{|x|}^{\!\perp}\right)$~is
$$ DT_\varepsilon(x) = \left(\begin{array}{cc} \partial_r\phi_\varepsilon & \partial_r\psi_\varepsilon \\[2mm] \dfrac{\partial_\theta\phi_\varepsilon-\psi_\varepsilon}{r} & \dfrac{\phi_\varepsilon+\partial_\theta\psi_\varepsilon}{r} \end{array} \right) $$
so that
$$ |DT_\varepsilon|^2 = \partial_r\phi_\varepsilon^2+\partial_r\psi_\varepsilon^2+\frac{(\partial_\theta\phi_\varepsilon-\psi_\varepsilon)^2}{r^2}+\frac{(\phi_\varepsilon+\partial_\theta\psi_\varepsilon)^2}{r^2} $$
We obtain
$$ \int_\Omega |DT_\varepsilon|^2 = \int_{\Omega_\delta} |DT_\varepsilon|^2 + \int_\delta^1 (||\partial_r\phi_\varepsilon(r,\cdot)||_{L^2}^2+||\partial_r\psi_\varepsilon||_{L^2}^2) \, r \, \text{\textnormal{d}} r + \int_\delta^1 (||\partial_\theta\phi_\varepsilon-\psi_\varepsilon||_{L^2}^2+||\phi_\varepsilon+\partial_\theta\psi_\varepsilon||_{L^2}^2) \,\frac{\text{d}r}{r} $$
On the other hand, we already know that
$$ \int_{\Omega} |T_\varepsilon(x)-x|f(x)\, \text{\textnormal{d}} x - W_1 = \int_{\Omega} (|T_\varepsilon(x)-x|-|T_\varepsilon(x)|+|x|) f(x) \, \text{\textnormal{d}} x $$
Finally, the complete expression of $F_\varepsilon$ is the following:
\begin{multline*} F_\varepsilon(T_\varepsilon) = \frac{1}{\varepsilon} \int_\Omega (|T_\varepsilon(x)-x|-|T_\varepsilon(x)|+|x|) f(x) \, \text{\textnormal{d}} x + \int_{\Omega_\delta} |DT_\varepsilon|^2 \\ + \int_\delta^1 (||(\partial_\theta\phi_\varepsilon-\psi_\varepsilon)(r,\cdot)||_{L^2}^2+||(\phi_\varepsilon+\partial_\theta\psi_\varepsilon)(r,\cdot)||_{L^2}^2 - K) \,\frac{\text{d}r}{r} + \int_\delta^1 (||\partial_r\phi_\varepsilon(r,\cdot)||_{L^2}^2+||\partial_r\psi_\varepsilon||_{L^2}^2) \, r \, \text{\textnormal{d}} r \end{multline*}
thus, if we denote by $H(\phi,\psi) = \displaystyle\int_0^{\pi/2} (\phi'(\theta)-\psi(\theta))^2+(\phi(\theta)+\psi'(\theta))^2 \, \text{\textnormal{d}}\theta$ for $\phi,\psi \in H^1(0,\pi/2)$, we have
\begin{multline} F_\varepsilon(T_\varepsilon) = \frac{1}{\varepsilon} \int_\Omega (|T_\varepsilon(x)-x|-|T_\varepsilon(x)|+|x|) f(x) \, \text{\textnormal{d}} x + \int_{\Omega_\delta} |DT_\varepsilon|^2 \\ + \int_\delta^1 (H(\phi_\varepsilon(r,\cdot),\psi_\varepsilon(r,\cdot)) - K) \,\frac{\text{d}r}{r} + \int_\delta^1 (||\partial_r\phi_\varepsilon(r,\cdot)||_{L^2}^2+||\partial_r\psi_\varepsilon||_{L^2}^2) \, r \, \text{\textnormal{d}} r \label{fepsteps} \end{multline}
The following lemma collects some properties of the function $H$.
\begin{lem} We recall that
$$ H : (\phi,\psi) \in H^1(0,\pi/2)\times H^1(0,\pi/2) \mapsto ||\phi'-\psi||_{L^2}^2+||\phi+\psi'||_{L^2}^2 $$
for $\phi,\psi \in H^1(0,\pi/2)$. Then:
\begin{itemize}
\item The function $H$ is lower semi-continuous with respect to the strong $L^2$-convergence.
\item Assume that $(\phi,\psi)$ satisfies, for any $\theta$,
$$ \phi(\theta)\hat{x}(\theta)+\psi(\theta)\hat{x}^\perp(\theta) \in \Omega' $$
where $\hat{x}(\theta) = (\cos\theta,\sin\theta)$. We denote by $\tilde{\phi}(\theta) = \max(\phi(\theta),R_1(\theta))$. Then we have the inequality
\begin{equation} H(\phi,\psi) \geq K+\frac{1}{2}||\tilde{\phi}-\Phi||_{L^2}^2-B||\psi||_{L^2(0,\pi/2)}^{2/3} \label{deuxtiers} \end{equation}
for some positive constant $B$ which only depends on $\Omega'$.
\end{itemize} \end{lem}
\begin{proof} {\it Step 1: the semi-continuity of $H$.} We take a sequence $(\phi_n,\psi_n)_n$ converging to some $(\phi,\psi)$ for the $L^2$-norm. Up to subsequences, we can assume that
$$ \liminf\limits_{n\to+\infty} H(\phi_n,\psi_n) = \lim\limits_{n\to+\infty} H(\phi_n,\psi_n) $$
and we also assume that $(H(\phi_n,\psi_n))_n$ is bounded. Now we remark that
$$ H(\phi_n,\psi_n) = ||\phi_n'-\psi_n||_{L^2}^2+||\phi_n+\psi_n'||_{L^2}^2 \geq (||\phi_n'||_{L^2}-||\psi_n||_{L^2})^2+(||\phi_n'||_{L^2}-||\psi_n||_{L^2}^2)^2 $$
$$ \text{thus}\qquad ||\phi'_n||_{L^2} \leq \sqrt{H(\phi_n,\psi_n)}+||\psi_n||_{L^2} \qquad \text{and} \qquad ||\psi'_n||_{L^2} \leq \sqrt{H(\phi_n,\psi_n)}+||\phi_n||_{L^2}$$
We deduce that $(\phi_n)_n$, $(\psi_n)_n$ are bounded in $H^1(0,\pi/2)$ so that the convergence $(\phi_n,\psi_n)\to(\phi,\psi)$ actually holds, up to a subsequence, weakly in $H^1(0,\pi/2)$. Now the convexity of $(\phi,\phi',\psi,\psi')\mapsto(\phi'-\psi)^2+(\phi+\psi')^2$ implies that $H$ is lower semi-continuous with respect to the weak convergence in $H^1(0,\pi/2)$, which allows to conclude.
Now we pass to the proof of the inequality \eqref{deuxtiers}. We begin by a kind of ``sub-lemma'' which will be useful several times in the proof.
{\it Step 2: preliminary estimates.} We recall that $\tilde{\phi}=\max(R_1,\phi)$ and denote by $h=\tilde{\phi}-\phi \geq 0$. Now we claim that:
\begin{itemize}
\item for any $t\in(0,\pi/2)$, we have the inequality
\begin{equation} 0\leq h(t)\leq B_1|\psi(t)| \label{hpsi} \end{equation}
for some constant $B_1$ depending only on $\Omega'$.
\item we have the inequality
\begin{equation} \left|\langle \tilde{\phi},h \rangle_{H^1}\right| \leq B_2||h||_\infty \label{estimscal} \end{equation}
for some constant $B_2$ depending only on $\Omega'$;
\item the above both inequalities lead to the estimate
\begin{equation} ||\phi||_{H^1}^2 \geq K+||\tilde{\phi}-\Phi||_{H^1}^2 - B_3||\psi||_\infty \label{phih1} \end{equation}
for some constant $B_3$ depending only on $\Omega'$.
\end{itemize}
First, we remark that the constraint $\phi(\theta)\hat{x}(\theta)+\psi(\theta)\hat{x}^\perp(\theta) \in \Omega'$
implies
$$ R_1(\theta')^2 < \phi(\theta)^2+\psi(\theta)^2 < R_2(\theta')^2 \qquad \text{where} \qquad \theta'=\theta+\arcsin\dfrac{\psi(\theta)}{\sqrt{\phi(\theta)^2+\psi(\theta)^2}} $$
Thus, we have
$$ \begin{array}{ll} h(\theta) & = R_1(\theta)-\phi(\theta) \\[1mm]
& = R_1(\theta)-R_1(\theta')+R_1(\theta')-\phi(\theta) \\[1mm]
& \leq (\operatorname{Lip} R_1) |\theta-\theta'| + \sqrt{\phi^2(\theta)+\psi^2(\theta)}-\phi(\theta) \\[1mm]
& \leq (\operatorname{Lip} R_1) \arcsin\dfrac{|\psi(\theta)|}{R_1(\theta')} + |\psi(\theta)| \\[1mm]
& \leq \left(\dfrac{\pi}{2}\dfrac{\operatorname{Lip} R_1}{\inf R_1} +1 \right) \, |\psi(\theta)| \end{array}$$
which is \eqref{hpsi} with $B_1 = \dfrac{\pi}{2}\dfrac{\operatorname{Lip} R_1}{\inf R_1} +1$.
Second, we recall that $h=(R_1-\phi)^+$, thus $\phi+h=R_1$ on any point where $h \neq 0$. This leads to
$$ \left|\int_0^{\pi/2} (\phi+h)h\right| = \left| \int_0^{\pi/2} R_1 h \right| \leq \frac{\pi}{2} (\sup R_1) ||h||_\infty $$
$$ \text{and} \qquad \left|\int_0^{\pi/2} (\phi+h)'h'\right| = \left|\int_0^{\pi/2} R_1'h'\right| = \left| [R_1'h]_0^{\pi/2}-\int_0^{\pi/2} R_1''h\right| \leq (2\sup R_1'+||R_1''||_1) ||h||_\infty $$
We get \eqref{estimscal} with $B_2=\left(\dfrac{\pi}{2}\sup R_1 + 2\operatorname{Lip} R_1 + ||R_1''||_{L^1}\right)$.
Third, we write
\begin{equation} ||\phi||_{H^1}^2 = ||\tilde{\phi}||_{H^1}^2 + ||h||_{H^1}^2+2\langle\tilde{\phi},h\rangle \label{phiphitildeh} \end{equation}
Since $\tilde{\phi} \geq R_1$ on $(0,\pi/2)$ and thanks to the Lemma \ref{phimoinsphi}, we have $||\tilde{\phi}||_{H^1}^2 \geq ||\tilde{\phi}-\Phi||_{H^1}^2+K$. On the other hand, by using \eqref{hpsi} and \eqref{estimscal}, we have
$$ \langle\tilde{\phi},h\rangle \geq -B_2||h||_\infty \geq -B_1B_2||\psi||_\infty $$
We insert into \eqref{phiphitildeh} and skip $||h||_{H^1}^2$ since it is nonnegative to get
$$ ||\phi||_{H^1}^2 \geq K+||\tilde{\phi}-\Phi||_{H^1}^2 - 2B_1B_2||\psi||_\infty $$
thus \eqref{phih1} holds with $B_3=2B_1B_2$.
{\it Step 3: the inequality \eqref{deuxtiers} holds if $||\phi'||_{L^2}$ is large enough.} We start from
$$ H(\phi,\psi) = ||\phi||_{H^1}^2+||\psi||_{H^1}^2 + 2 \int_0^{\pi/2} \phi\psi'-2\int_0^{\pi/2}\psi\phi'
= ||\phi||_{H^1}^2+||\psi||_{H^1}^2-4\int_0^{\pi/2} \psi\phi' - 2[\phi\psi]_{0}^{\pi/2} $$
First, the condition on $(\phi,\psi)$ implies that $||\phi||_\infty,||\psi||_\infty \leq \sup{R_2}$ so that
$$ |[\phi\psi]_{0}^{\pi/2}| \leq 2(\sup{R_2})^2 $$
On the other hand,
$$ \left|\int_0^{\pi/2} \psi\phi'\right| \leq ||\psi||_\infty \sqrt{\frac{\pi}{2}} ||\phi'||_{L^2} \leq \sup{R_2} \sqrt{\frac{\pi}{2}} ||\phi'||_{L^2} $$
This leads to
$$ H(\phi,\psi) \geq ||\phi||_{H^1}^2+||\psi||_{H^1}^2 - 4\sup{R_2} \sqrt{\frac{\pi}{2}} ||\phi'||_{L^2} - 4(\sup{R_2})^2 $$
$$ \geq \frac{1}{2}||\phi||_{H^1}^2+\left(\frac{1}{2}||\phi'||_{L^2}^2-4\sup{R_2}\sqrt{\frac{\pi}{2}} ||\phi'||_{L^2} - 4(\sup{R_2})^2\right) $$
By using \eqref{phih1}, we obtain
$$ H(\phi,\psi) \geq \frac{1}{2}(K+||\tilde{\phi}-\Phi||_{H^1}^2-B_3||\psi||_\infty)+\left(\frac{1}{2}||\phi'||_{L^2}^2-4\sup{R_2}\sqrt{\frac{\pi}{2}} ||\phi'||_{L^2} - 4(\sup{R_2})^2\right)$$
and, since $|\psi| \leq \sqrt{\phi^2+\psi^2} \leq R_1$, we have
$$ H(\phi,\psi) \geq \frac{1}{2}(K+||\tilde{\phi}-\Phi||_{H^1})+\left(\frac{1}{2}||\phi'||_{L^2}^2-4\sup{R_2}\sqrt{\frac{\pi}{2}} ||\phi'||_{L^2} - \left(4(\sup{R_2})^2+\frac{B_3}{2}\sup R_2\right)\right) $$
The announced estimate \eqref{deuxtiers} holds provided that the term in brackets is greater that $\dfrac{K}{2}$, which is true provided that $||\phi'||_{L^2} \geq B_4$ where $B_4$ is the largest root of the polynom
$$ \frac{1}{2}X^2-4\sup~R_2\sqrt{\frac{\pi}{2}}X-\left(4(\sup{R_2})^2+\frac{B_3}{2}\sup R_2+\frac{K}{2}\right) $$
and $B_4$ only depends of $\Omega'$.
{\it Step 4: case $||\phi'||_{L^2} \leq B_4$.} In this case, we still have
$$ H(\phi,\psi) = ||\phi||_{H^1}^2+||\psi||_{H^1}^2 -4\int_0^{\pi/2} \psi\phi' - 2[\phi\psi]_{0}^{\pi/2} $$
$$ \text{with} \qquad \left|\int_0^{\pi/2} \psi\phi'\right| \leq ||\psi||_{\infty}\sqrt{\frac{\pi}{2}} ||\phi'||_{L^2} \leq \sqrt{\frac{\pi}{2}} B_4 ||\psi||_\infty $$
$$ \text{and} \qquad |[\phi\psi]_{0}^{\pi/2}| \leq 2||\phi||_\infty||\psi||_\infty \leq 2\sup{R_2} ||\psi||_\infty $$
This leads to
$$ H(\phi,\psi) \geq ||\phi||_{H^1}^2+||\psi||_{H^1}^2 - \left(\sqrt{\frac{\pi}{2}} B_4+2\sup{R_2}\right)||\psi||_\infty $$
$$ \geq K+||\tilde{\phi}-\Phi||_{H^1}^2+||\psi||_{H^1}^2-B_5||\psi||_\infty $$
where we have again used \eqref{phih1} and set $B_5=\left(\sqrt{\frac{\pi}{2}} B_4+2\sup{R_2}\right)+B_3$, which only depends on $\Omega'$.
It now remains to estimate $||\psi||_{H^1}^2-B_5||\psi||_\infty$ from below with $-||\psi||_{L^2}^{2/3}$. The condition on $(\phi,\psi)$ implies that $\psi(0) \geq 0$ and $\psi(\pi/2) \leq 0$, so that there exists $t_0$ such that $\psi(t_0) = 0$. We then have
$$ \psi^2(t) = \int_{t_0}^t \dfrac{\text{\textnormal{d}}}{\text{\textnormal{d}t}} (\psi^2) = \int_{t_0}^t 2\psi\psi' \leq 2||\psi||_{L^2}||\psi'||_{L^2} \quad \text{thus} \quad ||\psi||_\infty \leq \sqrt{2} \sqrt{||\psi||_{L^2}||\psi'||_{L^2}} $$
We use the Young inequality
$$ab \leq \frac{(\varepsilon a)^p}{p}+\frac{(b/\varepsilon)^q}{q} \qquad \text{for } \frac{1}{p}+\frac{1}{q} = 1 \text{ and } \varepsilon > 0$$
with $p=4$, $q=4/3$, $a = \sqrt{||\psi'||_{L^2}}$ and $b = \sqrt{||\psi||_{L^2}}$, to get
$$ ||\psi||_\infty \leq \frac{\sqrt{2}\varepsilon^4}{4} ||\psi'||_{L^2}^2 + \frac{3\sqrt{2}}{4\varepsilon^{4/3}} ||\psi||_{L^2}^{2/3} $$
We deduce
$$ ||\psi'||_{H^1}^2-B_5||\psi||_\infty \geq \left(1-\frac{\sqrt{2}B_5\varepsilon^4}{4}\right)||\psi'||_{L^2}-\frac{3\sqrt{2}B_5}{4\varepsilon^{4/3}}||\psi||_{L^2}^{2/3} $$
By choosing $\varepsilon$ such that $\dfrac{\sqrt{2}B_5\varepsilon^4}{4} = 1$, we obtain
$$||\psi||_{H^1}^2 -B_5||\psi||_\infty \geq -B||\psi||_{L^2}^{2/3} $$
where $B = \dfrac{3\sqrt{2}B_5}{4\varepsilon^{4/3}}$ only depends on $\Omega'$ and $K$. This achieves the proof.\end{proof}
We also will need the following estimate on the first term of the expression \eqref{fepsteps}.
\begin{lem} Let $T$ be a transport map from $\mu$ to $\nu$. We write
$$ T(x) = \phi(x) \frac{x}{|x|} + \psi(x) \frac{x}{|x|}^{\!\perp} $$
Then, for a.e.~$x$,
\begin{equation} |T(x)-x|-|T(x)|+|x| \geq A |x| \psi^2(x) \label{xpsi2} \end{equation}
for some constant $A$ which only depends on $\Omega'$.\end{lem}
\begin{proof} We compute:
$$ |T(x)-x|-|T(x)|+|x| = \frac{|T(x)-x|^2-|T(x)|^2}{|T(x)-x|+|T(x)|} + |x| $$
We have $|T(x)-x|^2 = (\phi(x)-|x|)^2+\psi(x)^2$ and $|T|^2=\phi^2+\psi^2$, so that
$$ |T(x)-x|-|T(x)|+|x| = \frac{|x|^2-2|x|\phi(x)}{|T(x)-x|+|T(x)|}+|x| = |x|\,\frac{|x|+|T(x)-x|+|T(x)|-2\phi(x)}{|T(x)-x|+|T(x)|}$$
We remark that
$$ |T(x)-x|-\phi(x)+|x| = \sqrt{(\phi(x)-|x|)^2+\psi(x)^2}-(\phi(x)-|x|) \geq 0 $$
$$ \text{thus} \quad |T(x)-x|-|T(x)|+|x| \geq |x| \frac{|T(x)|-\phi(x)}{|T(x)-x|+|T(x)|} = |x| \frac{|T(x)|^2-\phi(x)^2}{(|T(x)-x|+|T(x)|)(|T(x)|+\phi(x))} $$
Since $x \in \Omega$ and $T(x) \in \Omega'$, we have
$$ |T(x)-x|+|T(x)| \leq 2|T(x)|+|x| \leq 2\sup{R_2} + 1 $$
$$ \text{and} \qquad |T(x)|+\phi(x) \leq 2|T(x)| \leq 2\sup{R_2} $$
On the other hand, $|T(x)|^2-\phi(x)^2 = \psi(x)^2$. This leads to the result with $ A = \dfrac{1}{(2\sup{R_2}+1)(2\sup{R_2})} $ \end{proof}
The estimate \eqref{xpsi2} leads to
$$ \int_{\Omega}(|T_\varepsilon(x)-x|-|T_\varepsilon(x)|+|x|) f(x) \, \text{\textnormal{d}} x \geq A\inf{f}\, \int_0^1 ||\psi_\varepsilon(r,\cdot)||_{L^2}^2 \, r \, \text{\textnormal{d}} r $$
and the estimate \eqref{deuxtiers} to
$$ \int_\delta^1 (H(\phi_\varepsilon(r,\cdot),\psi_\varepsilon(r,\cdot))-K) \frac{\text{d}r}{r} \geq \int_\delta^1 \left(-B||\psi_\varepsilon(r,\cdot)||_{L^2}^{2/3}+\frac{1}{2}||\tilde{\phi}_\varepsilon(r,\cdot)-\Phi||_{H^1}^2\right) \, \frac{\text{d}r}{r} $$
where we again have set $\tilde{\phi}_\varepsilon=\max(R_1,\phi_\varepsilon)$. By inserting into \eqref{fepsteps}, we have
\begin{multline} F_\varepsilon(T_\varepsilon) \geq \frac{A\inf{f}}{\varepsilon} \int_0^1 ||\psi_\varepsilon(r,\cdot)||_{L^2}^2 \, r^2 \, \text{\textnormal{d}} r - B\int_\delta^1||\psi_\varepsilon(r,\cdot)||_{L^2}^{2/3} \frac{\text{d}r}{r}
\\ + \frac{1}{2}\,\int_\delta^1 (H(\phi_\varepsilon(r,\cdot),\psi_\varepsilon(r,\cdot))-K+B||\psi_\varepsilon(r,\cdot)||_{L^2}^{2/3}) \,\frac{\text{d}r}{r} + \int_\delta^1 ||\partial_r\phi_\varepsilon(r,\cdot)||_{L^2}^2 \, r \, \text{\textnormal{d}} r \label{fepstepsbis} \end{multline}
Let us denote by $X_\varepsilon = \dfrac{1}{\varepsilon} \displaystyle\int_0^1 ||\psi_\varepsilon(r,\cdot)||_{L^2}^2 r^2 \, \text{\textnormal{d}} r$. By the H\"older inequality applied with respect to the measure with density $1/r$ on $(\delta,1)$, we have
$$ \int_\delta^1||\psi_\varepsilon(r,\cdot)||_{L^2}^{2/3} \,\frac{\text{d}r}{r} = \int_\delta^1(||\psi_\varepsilon(r,\cdot)||_{L^2}^2 r^3)^{1/3} \, \frac{1}{r} \, \frac{\text{d}r}{r} \leq \left(\int_\delta^1 ||\psi_\varepsilon||_{L^2}^2 r^3 \,\frac{\text{d}r}{r} \right)^{1/3} \, \left(\int_\delta^1 \frac{1}{r^{4/3}} \,\frac{\text{d}r}{r} \right)^{3/4} $$
$$ \text{with} \qquad \int_\delta^1 ||\psi_\varepsilon||_{L^2}^2 r^3 \,\frac{\text{d}r}{r} \leq \varepsilon X_\varepsilon \quad \text{and} \quad \int_\delta^1 \frac{1}{r^{4/3}} \,\frac{\text{d}r}{r} = \frac{3}{4}\left(\frac{1}{\delta^{4/3}}-1\right) \leq \frac{3}{2\delta^{4/3}} $$
which leads to
$$ \int_\delta^1||\psi_\varepsilon(r,\cdot)||_{L^2}^{2/3} \,\frac{\text{d}r}{r} \leq (\varepsilon X_\varepsilon)^{1/3} \left(\frac{3}{2\delta^{4/3}}\right)^{3/4} = \sqrt{\frac{3\sqrt{3}}{2\sqrt{2}}}\, X_\varepsilon^{1/3} $$
since $\delta = \varepsilon^{1/3}$. We insert into \eqref{fepstepsbis} to obtain
\begin{multline} F_\varepsilon(T_\varepsilon) \geq (A\inf{f}) X_\varepsilon -B' \,X_\varepsilon^{1/3} \\
+ \int_\delta^1 (H(\phi_\varepsilon(r,\cdot),\psi_\varepsilon(r,\cdot))-K+B||\psi_\varepsilon(r,\cdot)||_{L^2}^{2/3}) \,\frac{\text{d}r}{r} + \int_\delta^1 ||\partial_r\phi_\varepsilon(r,\cdot)||_{L^2}^2 \, r \, \text{\textnormal{d}} r \label{fepsteps2} \end{multline}
where $B' = \sqrt{\frac{3\sqrt{3}}{2\sqrt{2}}} \, B$.
Let us assume that $(F_\varepsilon(T_\varepsilon))_\varepsilon$ is bounded by a positive constant $M$. This implies that $(X_\varepsilon)_\varepsilon$ is bounded by some constant $M'$ (otherwise the term $(A\inf{f}) X_\varepsilon -B' \,X_\varepsilon^{1/3}$ would tend to $+\infty$, and the other term is positive), thus
$$ \int_0^1 ||\psi_\varepsilon(r,\cdot)||_{L^2}^2 r^2 \, \text{\textnormal{d}} r \leq M'\varepsilon $$
and $\psi_\varepsilon \to 0$ a.e.~on $\Omega$. Since $(X_\varepsilon)_\varepsilon$ and $(F_\varepsilon(T_\varepsilon))_\varepsilon$ are bounded, \eqref{fepsteps2} provides
$$ \int_\delta^1 (H(\phi_\varepsilon(r,\cdot),\psi_\varepsilon(r,\cdot)-K+B||\psi_\varepsilon(r,\cdot)||_{L^2}^{2/3}) \,\frac{\text{d}r}{r} + \int_\delta^1 ||\partial_r\phi_\varepsilon(r,\cdot)||_{L^2}^2 \, r \, \text{\textnormal{d}} r \leq M'' $$
for some constant $M''$ which does not depend on $\varepsilon$. We now use the estimate \eqref{deuxtiers} to get
$$ \frac{1}{2}\int_\delta^1 ||\tilde{\phi}_\varepsilon-\Phi||_{H^1}^2\, \frac{\text{d}r}{r} + \int_\delta^1 ||\partial_r\phi_\varepsilon(r,\cdot)||_{L^2}^2 \, r \, \text{\textnormal{d}} r \leq M'' $$
We thus have a $L^2_{loc}$-loc bound $\partial_r\phi_\varepsilon$ and on $\partial_\theta\tilde{\phi}_\varepsilon$, but since $\tilde{\phi}(r,\theta)=\max(R_1(\theta),\phi(r,\theta))$, the bound on $\partial_r\phi_\varepsilon$ implies a bound on $\partial_r\tilde{\phi}_\varepsilon$. Thus, the family $(\tilde{\phi}_\varepsilon)_\varepsilon$ is bounded in $H^1_{loc}(\Omega)$ and there exists $\varepsilon_k \to 0$ and $\tilde{\phi}$ such that $\tilde{\phi}_\varepsilon \to \tilde{\phi}$ a.e.~on $\Omega$. But we recall that the estimation \eqref{hpsi} still holds and provides
$$ |\phi_{\varepsilon_k}-\tilde{\phi}_{\varepsilon_k}| \leq B_1|\psi_\varepsilon| \to 0 $$
which leads $\phi_{\varepsilon_k} \to \tilde{\phi}$ a.e.~on $\Omega$. If we set now $T(x)=\tilde{\phi}\dfrac{x}{|x|}$, we have proved that $T_{\varepsilon_k} \to T$ a.e.~on~$\Omega$. This convergence also holds in $L^2(\Omega)$ since $|T_\varepsilon| \leq \sup R_2$ for any $\varepsilon$. This proves the first statement of the Theorem~\ref{PRESQUEGAMMA}.
Assume now that $(T_\varepsilon)_\varepsilon$ is a family of transport maps converging to some $T$ for the $L^2$-norm on $\Omega$. We deduce from \eqref{fepsteps2} that, if we set $C=- \inf\limits_{X>0} (A\inf{f}X^3-B'X)$, which only depends on $\Omega'$, we have
$$ F_\varepsilon(T_\varepsilon) \geq -C+\int_\delta^1 (H(\phi_\varepsilon(r,\cdot),\psi_\varepsilon(r,\cdot))-K+B||\psi_\varepsilon(r,\cdot)||_{L^2}^{2/3}) \,\frac{\text{d}r}{r} + \int_\delta^1 ||\partial_r\phi_\varepsilon(r,\cdot)||_{L^2}^2 \, r \, \text{\textnormal{d}} r $$
Assuming that $(F_\varepsilon(T_\varepsilon))_\varepsilon$ is bounded, the above computations give a $H^1$-loc bound for $(\phi_\varepsilon)_\varepsilon$, thus
$$ \liminf\limits_{\varepsilon \to 0} \int_\delta^1 ||\partial_r\phi_\varepsilon(r,\cdot)||_{L^2}^2 \, r \, \text{\textnormal{d}} r \geq \int_0^1 ||\partial_r\phi(r,\cdot)||_{L^2}^2 \, r \, \text{\textnormal{d}} r $$
since this functional is lower semi-continuous for the weak convergence in $H^1(\Omega)$. On the other hand, the estimate \eqref{deuxtiers} shows also that
$$ H(\phi_\varepsilon(r,\cdot),\psi_\varepsilon(r,\cdot))-K+B||\psi_\varepsilon(r,\cdot)||_{L^2}^{2/3} \geq 0 $$
for any $\varepsilon$ and $r$, thus we can apply the Fatou lemma since the semi-continuity of $H$ provides
$$ \liminf\limits_{k \to +\infty} (H(\phi_\varepsilon(r,\cdot),\psi_\varepsilon(r,\cdot))-K+B||\psi_\varepsilon(r,\cdot)||_{L^2}^{2/3}) \geq G(\phi(r,\cdot)) $$
for a.e.~$r\in(0,1)$. Thus
$$ \liminf_{\varepsilon \to 0} \geq -C+\int_0^1 G(\phi(r,\cdot)) \,\frac{\text{d}r}{r} + \int_0^1 ||\partial_r\phi(r,\cdot)||_{L^2}^2 \, r \, \text{\textnormal{d}} r $$
as announced.
\paragraph{Remark.} If we choose to set $\delta = \lambda\varepsilon^{1/3}$, where $\lambda$ is to be fixed (and possibly sent to $0$ or $\infty$, or taken depending on $\varepsilon$), the precise expression~\eqref{fepsteps} becomes
\begin{multline} F_\varepsilon(T_\varepsilon) = \frac{1}{\varepsilon} \int_{\Omega_\delta} (|T_\varepsilon(x)-x|-|T(x)|+|x|) f(x) \, \text{\textnormal{d}} x + \int_{\Omega_\delta} |DT_\varepsilon|^2 - K\log\lambda \\
+ \frac{1}{\varepsilon} \int_{\Omega\setminus\Omega_\delta} (|T_\varepsilon(x)-x|-|T(x)|+|x|) f(x) \, \text{\textnormal{d}} x - B\int_\delta^1 ||\psi_\varepsilon(r,\cdot)||_{L^2}^{2/3} \, r \, \text{\textnormal{d}} r \\
+ \int_\delta^1 (H(\phi_\varepsilon(r,\cdot),\psi_\varepsilon(r,\cdot))-K+B||\psi_\varepsilon(r,\cdot)||_{L^2}^{2/3}) \frac{\text{d}r}{r}+\int_\delta^1 ||\partial_r\phi_\varepsilon(r,\cdot)||_{L^2}^2 \, r \, \text{\textnormal{d}} r \label{fepstepster} \end{multline}
By using the above estimates \eqref{deuxtiers} and \eqref{xpsi2}, we get that the second line is this time bounded from below~by
$$ A (\inf f) X_\varepsilon - \frac{B'}{\lambda} (X_\varepsilon)^{1/3} $$
which is itself bounded from below by $-C_\lambda = \inf\{ A(\inf f )X-\frac{B'}{\lambda} X^{1/3}\}$, which goes to $0$ as $\lambda \to +\infty$. Thus, these terms disappear if we allow $\lambda\to +\infty$. Let us now compute the first line of \eqref{fepstepster}:
\begin{multline*} \frac{1}{\varepsilon} \int_{\Omega_\delta} (|T_\varepsilon(x)-x|-|T(x)|+|x|) f(x) \, \text{\textnormal{d}} x + \int_{\Omega_\delta} |DT_\varepsilon|^2 - K\log\lambda
\\ = \frac{\delta^2}{\varepsilon} \int_{\Omega} (|U_\varepsilon(y)-\delta y|-|U_\varepsilon(y)|+|\delta y|) f(\delta y) \, \text{\textnormal{d}} y + \int_\Omega |DU_\varepsilon|^2 -K\log\lambda \end{multline*}
where we have set $x=\delta y$ and $U_\varepsilon(y) = T_\varepsilon(\delta y)$. Now we use the following expansion
$$ |U_\varepsilon(y)-\delta y|-|U_\varepsilon(y)| = \delta \left(|y|-y\cdot \frac{U_\varepsilon(y)}{|U_\varepsilon(y)|}\right)+o(\delta^2) $$
and recall that $\delta = \lambda \varepsilon^{1/3}$, to get
\begin{multline*} \frac{\delta^2}{\varepsilon} \int_{\Omega} |U_\varepsilon(y)-\delta y|-|U_\varepsilon(y)|+|\delta y| f(\delta y) \, \text{\textnormal{d}} y + \int_\Omega |DU_\varepsilon|^2 -K\log\lambda \\
= \int_\Omega \lambda^3\left(|y|-y\cdot \frac{U_\varepsilon(y)}{|U_\varepsilon(y)|}\right) f(\delta y) \, \text{\textnormal{d}} y + \int_\Omega |DU_\varepsilon|^2-K\log\lambda + o\left(\lambda^4\varepsilon^{4/3}\right) \end{multline*}
We can bound this term from below (up to the rest with order $\lambda^4\varepsilon^{1/3}$, that we can make going to $0$ with a good choice of $\lambda$, and another rest taking account of $f(\delta y) \to f(0)$) with
$$ - C'_\lambda = \inf\left\{ f(0) \int_\Omega \lambda^3\left(|y|-y\cdot \frac{U(y)}{|U(y)|}\right)\, \text{\textnormal{d}} y + \int_\Omega |DU|^2-K\log\lambda \right\}$$
where the constraint on $U$ has to be precised. For a good choice of $\lambda$, as $\varepsilon \to 0$, we skip the first constant $-C_\lambda$ and get, since the third line of \eqref{fepstepster} is lower semi-continuous,
$$ \liminf\limits_{\varepsilon \to 0} F_\varepsilon \geq F(T)-\lim_{\lambda\to+\infty} C'_\lambda $$
which corresponds to the result that we conjectured in order to make more precise the Theorem \ref{PRESQUEGAMMA}.
\subsection{Construction of family of transport maps with equi-bounded energy \label{construction}}
The last point of the proof of Theorem \ref{PRESQUEGAMMA} consists in building a family of maps $(T_\varepsilon)_\varepsilon$ such that $(F_\varepsilon(T_\varepsilon))_\varepsilon$ is as well bounded. The sketch of the proof is the following: we start from a transport map $T = \phi \, \dfrac{x}{|x|}$ with $\phi(0,\cdot) = \Phi$ (that we call ``the original $T$'' in the following), assume that $\phi$ is regular except around the origin, and modify $T$ only on $\Omega_\delta$.
\paragraph{Step 1: construction of the original transport map.} We set $T(x) = \phi(r,\theta) \dfrac{x}{|x|}$, where $\phi$ is built as follows:
\begin{itemize}
\item $\phi(0,\theta) = \Phi(\theta)$, and $\phi(\cdot,\theta)$ is increasing and sends the one-dimensional measure $\mu_\theta$ (the starting measure $\mu$ concentrated on the transport ray with angle $\theta$) onto $\nu_\theta/2$ (where $\nu_\theta$ is the target measure on the same transport ray), until the radius $\rho_1$ such that $\phi(\rho_1,\theta) = R_2(\theta)$;
\item starting from this radius $\rho_1$, $\phi(\cdot,\theta)$ is decreasing with the same source and target measure, until the radius $\rho_2$ such that, again, $\phi(\rho_2,\theta) = \Phi(\theta)$. Therefore, on the interval $(\rho_1,\rho_2)$, $\phi(\cdot,\theta)$ sends $\mu_\theta$ onto $\nu_\theta|_{(\Phi(\theta),R_2(\theta))}$;
\item on the last interval (if it is non-empty, which corresponds to $\Phi(\theta) > R_1(\theta)$), $\phi$ is still decreasing and sends $\mu_\theta$ onto $\nu_\theta|_{(R_1(\theta),\Phi(\theta)}$
\end{itemize}
Precisely, we fix $\theta$ and the expressions of $\mu_\theta$, $\nu_\theta$ are
$$ \, \text{\textnormal{d}}\mu_\theta(r) = rf(r,\theta) \, \text{\textnormal{d}} r \qquad \text{and} \qquad \, \text{\textnormal{d}}\nu_\theta(r) = rg(r,\theta) \, \text{\textnormal{d}} r $$
which have both same mass on $(0,1)$ and $(R_1(\theta),R_2(\theta))$ respectively. Now we define successively $\rho_1(\theta)$ and $\rho_2(\theta)$ by
$$ \int_{0}^{\rho_1(\theta)} \, \text{\textnormal{d}}\mu_\theta = \int_{\Phi(\theta)}^{R_2(\theta)} \frac{1}{2}\, \text{\textnormal{d}}\nu_\theta \qquad \text{and} \qquad \int_{\rho_1(\theta)}^{\rho_2(\theta)} \, \text{\textnormal{d}}\mu_\theta = \int_{\Phi(\theta)}^{R_2(\theta)} \frac{1}{2}\, \text{\textnormal{d}}\nu_\theta$$
which are proper definitions thanks to the intermediate value theorem, and imply
$$ \int_{\rho_2(\theta)}^1 \, \text{\textnormal{d}}\mu_\theta = \int_0^{\Phi(\theta)} \, \text{\textnormal{d}}\nu_\theta $$
Thus, we have the equality between masses:
$$ \mu_\theta(0,\rho_1(\theta)) = \mu_\theta(\rho_1(\theta),\rho_2(\theta)) = \frac{1}{2}\nu_\theta(\Phi(\theta),1) \quad \text{and} \quad \mu_\theta(\rho_2(\theta),1) = \nu_\theta(0,\Phi(\theta)) $$
and the measures $\mu_\theta$, $\nu_\theta$ are absolutely continuous on these intervals. We now define the function $\phi(\cdot,\theta)$ as being:
\begin{itemize}
\item on the interval $(0,\rho_1(\theta))$, the unique increasing map $(0,\rho_1(\theta) \to (\Phi(\theta),1)$ sending $\mu_\theta$ onto $\frac{1}{2}\nu_\theta$;
\item on the interval $(\rho_1(\theta),\rho_2(\theta))$, the unique decreasing map $(\rho_1(\theta),\rho_2(\theta)) \to (\Phi(\theta),1)$ sending $\mu_\theta$ onto $\frac{1}{2}\nu_\theta$;
\item on the interval $(\rho_2(\theta),1)$ (if this interval is not empty), the unique decreasing map $(\rho_2(\theta),1) \to (0,\Phi(\theta))$ sending $\mu_\theta$ to $\nu_\theta$
\end{itemize}
It is easy to check that $\phi(\cdot,\theta)$, defined on the whole $(0,1)$, sends globally $\mu_\theta$ onto $\nu_\theta$. As a consequence, the two-dimensional valued function
$$ T : x=(r,\theta) \in \Omega \mapsto \phi(r,\theta)\frac{x}{|x|} \in \Omega' $$
is a transport map from $\mu$ to $\nu$.
\paragraph{Step 2: estimates on $\phi$ around the origin.} As above, we set $\delta=\varepsilon^{1/3}$ and we aim to modify $T$ only on $\Omega_\delta=\Omega \cap B(0,\delta)$. In order to obtain as well a transport map from $\mu$ to $\nu$, we have to guarantee that the new map $S$ sends the domain $\Omega_\delta$ on its image $T(\Omega_\delta)$, with the constraint of image measure. For this, the following estimates on the original transport $T$ will be useful:
\begin{prop} The above function $\phi$ has Lipschitz regularity on $(0,1)\times(0,\pi/2)$. Moreover, there exists some positive constants $c,C$ depending only on $\Omega'$, $f$, $g$ such that
\begin{equation} cr^2 \leq \phi(r,\theta)-\Phi(\theta) \leq C r^2 \label{alphadelta1} \end{equation}
for any $r$ small enough and $\theta \in (0,\pi/2)$, and
\begin{equation} \operatorname{Lip} (\phi(r,\cdot)-\Phi) \leq C r^2 \label{alphadelta2} \end{equation}
for any $r \in (0,1)$. \end{prop}
\begin{proof} The Lipschitz regularity of $\phi$ is actually a consequence of its definition and of the inverse function theorem. First, let us recall that $\rho_1(\theta)$ is defined by
$$ \tilde{F}(\rho_1(\theta),\theta) = \frac{1}{2}(G(R_2(\theta),\theta)-\tilde{G}(\Phi(\theta),\theta)) $$
$$ \text{where} \qquad \tilde{F}(R,\theta) = \int_0^R \, \text{\textnormal{d}}\mu_\theta = \int_0^R rf(r,\theta)\, \text{\textnormal{d}} r \qquad \text{and} \qquad \tilde{G}(R,\theta) = \int_{R_1(\theta)}^R \, \text{\textnormal{d}}\nu_\theta = \int_{R_1(\theta)}^R r g(r,\theta)\, \text{\textnormal{d}} r $$
Notice that we have necessary $\rho_1(\theta) \geq \delta_0 > 0$, where $\delta_0$ verifies, for instance,
$$ \int_{B(0,\delta_0)\cap\Omega} \, \text{\textnormal{d}}\mu = \int_{\sup R_1 \leq r \leq \inf R_2} \, \text{\textnormal{d}}\nu $$
(such a $\delta_0$ exists since $\delta \mapsto \mu(B(0,\delta)\cap\Omega)$ is continuous and increasing). Thus, $\tilde{F}$ is $C^1$ with respect to its first variable and its derivative is Lipschitz and bounded from below by $\delta_0 \inf f$, and the same holds for $\tilde{G}$; moreover, $\tilde{F}$ and $\tilde{G}$ are both Lipschitz with respect to $\theta$. As a consequence of the inverse function theorem, $\rho_1$ is well-defined and Lipschitz. The same reasoning shows that $\rho_2$ is well-defined and Lipschitz.
Then we know that $\phi$ is defined if $0 \leq r \leq \rho_1(\theta)$ by
$$\int_0^r \, \text{\textnormal{d}}\mu_\theta = \int_{\Phi(\theta)}^{\phi(r,\theta)} \, \text{\textnormal{d}}\nu_\theta \qquad \text{\it i.e.} \qquad \tilde{F}(r,\theta) = \tilde{G}(\phi(r,\theta),\theta)-G(\Phi(\theta),\theta) $$
Again, by the inverse function theorem, $\phi$ is $C^1$ with respect to $r$ and Lipschitz with respect to $\theta$. The same reasoning can be applied on the intervals $(\rho_1(\theta),\rho_2(\theta))$ and, if it is non-empty, $(\rho_2(\theta),1)$. We get that the function $\phi$ is Lipschitz with respect to its both variables $(r,\theta)$.
Notice also the following estimate on $\phi(r,\theta)$ for $r$ small enough: we have $\int_0^r \, \text{\textnormal{d}}\mu_\theta = \int_{\Phi(\theta)}^{\phi(r,\theta)} \, \text{\textnormal{d}}\nu_\theta$, with the inequalities
$$ r^2 \inf f \leq \int_0^r \, \text{\textnormal{d}}\mu_\theta \leq r^2 \sup f \qquad \text{and} \qquad (\inf g) (\phi(r,\theta)-\Phi(\theta)) \leq \int_{\Phi(\theta)}^{\phi(r,\theta)} \, \text{\textnormal{d}}\nu_\theta \leq (\sup g) (\phi(r,\theta)-\Phi(\theta)) $$
This leads immediately to \eqref{alphadelta1}. On the other hand, we know that $\phi(r,\theta)$ has Lipschitz regularity with respect to $\theta$ and
\begin{equation} \int_{R_1(\theta)}^{\phi(R,\theta)} rg(r,\theta) \, \text{\textnormal{d}} r = \int_0^R rf(r,\theta)\, \text{\textnormal{d}} r \label{phi} \end{equation}
If $\theta$ is such that the equality \eqref{phi} is differentiable with respect to $\theta$ (which is true for a.e.~$\theta$ since all the considered functions are at least Lipschitz), we get
$$ \partial_2\phi(R,\theta)\phi(R,\theta)G(\phi(R,\theta),\theta)-R_1'(\theta)R_1(\theta)G(\phi(R,\theta),\theta) = \int_0^R r\partial_2 f(r,\theta) \, \text{\textnormal{d}} r $$
thus
$$ \partial_\theta(\phi(R,\theta)-R_1(\theta))\phi(R,\theta)G(\phi(R,\theta),\theta)+R_1'(\theta)(\phi(R,\theta)-R_1(\theta))G(\phi(R,\theta),\theta) = \int_0^R r\partial_2 f(r,\theta) \, \text{\textnormal{d}} r $$
and
$$ |\partial_\theta(\phi(R,\theta)-R_1(\theta))| \leq \frac{(\operatorname{Lip} R_1)(\sup G)}{(\inf R_1)(\inf G)}|\phi(R,\theta)-R_1(\theta)|+\frac{R^2}{2}\operatorname{Lip} F $$
Thanks to the inequality \eqref{alphadelta1}, we get
$$ |\partial_\theta(\phi(R,\theta)-R_1(\theta))| \leq CR^2 $$
for some constant $C$ depending only on $f$, $g$, $\Omega'$, which proves \eqref{alphadelta2}. \end{proof}
\paragraph{Step 3: perturbation of the optimal $T$.} In that follows, we denote by
$$ \Omega_\delta = \Omega_1 \cap B(0,\delta) = \{x=(r,\theta) : 0 < r < \delta \text{ and } 0 < \theta < \pi/2\} $$
$$ \text{and} \qquad \Omega'_\delta = T(\Omega_\delta) = \left\{x=(r,\theta) : \Phi(\theta) < r < \phi(\delta,\theta) \text{ and } 0 < \theta < \pi/2 \right\} $$
We now denote by:
\begin{itemize}
\item $S_1 : x \in \Omega_\delta \mapsto \dfrac{x}{\delta} \in \Omega_1$ and $f_\delta(x) = f(\delta x)$. Notice that $f_\delta = \dfrac{1}{\delta^2} (S_1)_\# (f|_{\Omega_\delta})$;
\item $\Omega_2$ is the rectangle $(0,1) \times (0,\pi/2)$ and
$$ S_2 : (\lambda,\theta) \in \Omega_2 \mapsto x = (\Phi(\theta)+\lambda(\phi(\delta,\theta)-\Phi(\theta)),\theta) \in \Omega_\delta' $$
where $x$ is here written in polar coordinates. We also denote by $g_\delta = \dfrac{1}{\delta^2} \left((S_2)^{-1}\,_\# g|_{\Omega_\delta'}\right)$.
\end{itemize}
As above, $S_1$ and $S_2$ actually depend on $\delta$ but we omit the index $\delta$ for the sake of simplicity of notations. We have of course $\inf f \leq f_\delta(x) \leq \sup f$ for any $x \in \Omega_1$, and $\operatorname{Lip} f_\delta \leq \delta \operatorname{Lip} f$. On the other hand, since $S_2$ is Lipschitz and one-to-one, the Monge-Amp\`ere equation provides
$$ \det DS_2(\lambda,\theta) = \frac{g_\delta(\lambda,\theta)}{g(S_2(\lambda,\theta))} $$
\begin{equation} \text{thus} \qquad g_\delta(\lambda,\theta) = (\phi(\delta,\theta)-\Phi(\theta)) \frac{1}{\delta^2} g(S_2(\lambda,\theta)) \label{gdelta} \end{equation}
We deduce from \eqref{alphadelta1} and \eqref{gdelta} that $c \leq g_\delta \leq C$ for $c,C$ positive and independent of $\delta$. Moreover, \eqref{gdelta} provides
$$ \operatorname{Lip} g_\delta \leq \sup (\phi_\delta-\phi) \frac{1}{\delta^2} \operatorname{Lip} (g\circ S_2) + \operatorname{Lip}(\phi_\delta-\Phi) \frac{1}{\delta^2} \sup g $$
Again, thanks to \eqref{alphadelta1} and \eqref{gdelta} and using that $\operatorname{Lip} S_2$ is uniformly bounded with respect to $\delta$, we get $\operatorname{Lip} g_\delta \leq C$ independent of~$\delta$.
We would like to use the Theorem \ref{appdacmos}, but this result works {\it a priori} only to connect two measures defined on a same domain. Thus, an intermediate step consists in sending $\Omega_\delta$, $\Omega'_\delta$ onto the unit ball, which can be done (with the regularity that we need) thanks to the Lemma~\ref{lemdiffeo}. Moreover, let us notice that $\Omega_1$, $\Omega_2$ are infinitely diffeomorphic to the following domains (which are obtained directly thanks to rotations/translations/dilations):
\begin{center}
\begin{tikzpicture}[xmin=-2,xmax=10,ymin=-2,ymax=-2]
\draw (0,0) circle (2cm) ;
\fill[gray!40] (0.414213562,-1.414213562) arc (-45:45:2) -- (-1,0) -- cycle ;
\draw [->] (-2.5,0) -- (2.5,0) ;
\draw [->] (0,-2.5) -- (0,2.5) ;
\draw[dashed] (-1,0) -- (0.732050808,1) ;
\draw[dashed] (0,0) -- (0.732050808,1) ;
\draw (0.414213562,1.414213562) arc (45:-45:2) ;
\draw (-0.4,0) arc (0:30:0.6) ;
\draw (-0.3,0) node[above]{$\theta$};
\draw (0.4,0) arc (0:53.7939689:0.4) ;
\draw (0.5,0) node[above]{$\tilde\theta$};
\draw (6,0) circle (2cm) ;
\fill[gray!40] (5,-1.414213562) rectangle (6.414213562,1.414213562) ;
\draw (6.414213562,1.414213562) -- (6.414213562,-1.414213562) ;
\draw [->] (3.5,0) -- (8.5,0) ;
\draw [->] (6,-2.5) -- (6,2.5) ;
\draw[dashed] (6,0) -- (6.414213562,0.565826248) -- (6,0.565826248) ;
\draw (6.4,0) arc (0:53.7939689:0.4) ;
\draw (6.7,0) node[above]{$\tilde\theta$};
\draw (5.6,0.3) node[above]{$\lambda(\theta)$};
\end{tikzpicture}
\end{center}
In the above picture, the upper right and lower right corners of the second domain are equal to the upper right and lower right corners of the first domain. Since the maps provided by the Lemma~\ref{lemdiffeo} from these two domains to the unit disk both preserve angles (with respect to the origin, see the angle $\tilde\theta$ on the picture), if we denote by $\alpha$, $\beta$ the corresponding maps starting from $\Omega_1$, $\Omega_2$, we have
$$ \beta^{-1}\circ\alpha(x) = (1,\lambda(\theta)) \qquad \text{if} \qquad |x|=1 \text{ and } x \text{ has } \theta \text{ for angle.} $$
where $\lambda$ is a bi-Lipschitz map from the interval $(0,\pi/2)$ onto $(0,1)$; by composing the second coordinate with $\lambda^{-1}$, one can actually assume that $\beta^{-1}\circ\alpha(x) = (1,\theta)$. On the other hand, we deduce from the regularity of $\det D\alpha$, $\det D\beta$ that, if
$$ \bar{f_\delta} = \alpha_\# f_\delta \qquad \text{and} \qquad \bar{g_\delta} = \beta_\# g_\delta $$
then $\bar{f_\delta}$, $\bar{g_\delta}$ have also infimum, supremum and Lipschitz constant bounded independently of $\delta$. Now the Dacorogna-Moser result provides the existence of a bi-Lipschitz diffeomorphism $u_\delta$, whose Lipschitz constant is bounded independently of $\delta$, sending the density $\bar{f_\delta}$ onto $\bar{g_\delta}$, and with $u_\delta(x) = x$ for $|x| = 1$.
Finally, we consider
$$ S = S_2 \circ \beta^{-1} \circ u_\delta \circ \alpha \circ S_1 $$
If $|x| = \delta$, we have $|S_1(x)| = 1$, thus $\alpha(S(x)) \in \partial B(0,1)$ and
$$ \beta^{-1}(u_\delta(\alpha(S_1(x)))) = \beta^{-1}(\alpha(S_1(x))) $$
so that $\beta^{-1}(u_\delta(\alpha(S_1(x)))) = (1,\theta)$, where $\theta$ is the angle of $x$. Consequently, $ S_2(\beta^{-1}(u_\delta(\alpha(S_1(x))))) $ has $(\phi(\delta,\theta),\theta)$ as polar coordinates, so it is equal to $T(x)$.
To summarize, $S$ and $T$ coincide on the line $\{|x|=\delta\}$, $S$ is Lipschitz on $\Omega_\delta$ and, thanks to the estimates on $\phi$, $T$ is Lipschitz on $\Omega\setminus \Omega_\delta$. This implies that $T_\varepsilon$ is globally Lipschitz on $\Omega$, thus it belongs to $H^1(\Omega)$. Moreover, we have
$$ \operatorname{Lip} u_\delta \leq C, \quad \operatorname{Lip} S_2 \leq \operatorname{Lip} \Phi \quad \text{and } \operatorname{Lip} S_1 = \frac{1}{\delta} $$
$$ \text{thus} \qquad \operatorname{Lip} S \leq \frac{C}{\delta} $$
for some constant $C$ which does not depend on $\delta$.
\paragraph{Step 4: estimates on $F_\varepsilon(T_\varepsilon)$.} We restart from the expression \eqref{fepsteps}, and use the facts that $\psi_\varepsilon = 0$ and that $T_\varepsilon=T$ outside of $\Omega_\delta$:
\begin{multline*} F_\varepsilon(T_\varepsilon) = \frac{1}{\varepsilon} \int_{\Omega_\delta} (|S(x)-x|-|S(x)|+|x|) f(x) \, \text{\textnormal{d}} x + \int_{\Omega_\delta} |DS(x)|^2 \, \text{\textnormal{d}} x \\
+ \int_\delta^1 (||\phi(r,\cdot)||^2_{L^2}+||\partial_\theta\phi(r,\cdot)||^2_{L^2}-K)\,\frac{\text{d}r}{r} + \int_\delta^1 ||\partial_r\phi(r,\cdot)||^2_{L^2} \, r \, \text{\textnormal{d}} r \end{multline*}
We still have $|S(x)-x|-|S(x)|+|x| \leq 2|x|$ and $|DS(x)| \leq C/\delta$, so that
$$ \frac{1}{\varepsilon} \int_{\Omega_\delta} (|S(x)-x|-|S(x)|+|x|) f(x) \, \text{\textnormal{d}} x + \int_{\Omega_\delta} |DS(x)|^2 \, \text{\textnormal{d}} x \leq \frac{\pi \sup f}{3} \frac{\delta^3}{\varepsilon} + \frac{C^2\pi}{4} $$
which is bounded since $\delta = \varepsilon^{1/3}$. On the other hand,
$$ ||\phi(r,\cdot)||_{L^2}^2+||\partial_\theta\phi(r,\cdot)||_{L^2}^2-K = (||\phi(r,\cdot)||_{L^2}^2-||\Phi||_{L^2}^2)+(||\partial_\theta\phi(r,\cdot)||_{L^2}^2-||\Phi'||_{L^2}^2) $$
$$ = \langle \phi(r,\cdot)-\Phi,\phi(r,\cdot)+\Phi \rangle_{L^2} + \langle \partial_\theta\phi(r,\cdot)-\Phi',\partial_\theta\phi(r,\cdot)+\Phi' \rangle_{L^2} $$
$$ \leq ||\phi(r,\cdot)-\Phi'||_{L^1} (||\phi(r,\cdot)||_\infty+||\Phi||_\infty) + ||\partial_\theta\phi(r,\cdot)-\Phi'||_{L^1} (\operatorname{Lip} \phi+\operatorname{Lip} \Phi) $$
Since $\Phi$, $\phi(r,\cdot)$ are valued in $\Omega'$, their $L^\infty$-norm are controlled by $\sup R_2$. By combining this and the estimates \eqref{alphadelta1} and \eqref{alphadelta2}, we obtain
$$ 0 \leq ||\phi(r,\cdot)||_{L^2}^2+||\partial_\theta\phi(r,\cdot)||_{L^2}^2-K \leq Cr^2 $$
for $r$ small enough (and where $C$ does not depend on $r$). On the other hand, we know that $\phi$, $\Phi$ and their derivatives are globally bounded on $(0,1)\times(0,\pi/2)$. This proves that
$$ \int_0^1 (||\phi(r,\cdot)||^2_{L^2}+||\partial_\theta\phi(r,\cdot)||^2_{L^2}-K)\,\frac{\text{d}r}{r} < +\infty $$
$$ \text{and} \qquad \int_0^1 ||\partial_r\phi(r,\cdot)||^2_{L^2} \, r \, \text{\textnormal{d}} r < +\infty $$
and we conclude that $(F_\varepsilon(T_\varepsilon))_\varepsilon$ is bounded as well.
|
\section{Introduction}
When gravitational waves (GWs) are detected by second-generation detectors (such as advanced LIGO (aLIGO)~\cite{ligo,Abramovici:1992ah,Abbott:2007kv}, advanced Virgo (aVirgo)~\cite{virgo,Giazotto:1988gw}, and KAGRA~\cite{KAGRA2}) some time in the next few years, one of the most exciting prospects is using these signals to test General Relativity (GR) in the very strong-field, \emph{dynamical and non-linear} regime~\cite{Yunes:2013dva}. There has been much work done on constraining departures from GR dynamics with Solar System and binary pulsar observations, and quite strong bounds have been placed on deviations from Einstein's theory in certain regimes. In particular, the strength of dipole radiation from some types of scalar-tensor (ST) theories is already tightly constrained by observations of the rate of decay of the orbital period of binary pulsars~\cite{Psaltis:2005ai, stairs, damour-taylor, kramer-wex, Freire:2012mg}.
There remains, however, a class of ST theories that can escape these constraints. For theories in this class, initially proposed by Ref.~\cite{Damour:1992we,Damour:1993hw}, the scalar charge of a compact object is dependent upon the gravitational binding energy (or compactness) of the object itself (\emph{spontaneous scalarization})~\cite{Damour:1992we,Damour:1993hw}, and, if the object is in a binary, upon the orbital binding energy of the binary system (\emph{dynamical scalarization}), a phenomenon discovered in Ref.~\cite{Barausse:2012da}. Additionally, once a star acquires a scalar charge, it can scalarize its binary companion (\emph{induced scalarization})~\cite{Damour:1992we,Damour:1993hw,Damour:1996ke,Barausse:2012da,Palenzuela:2013hsa}.
These three types of scalarization can be understood in analogy with magnetization~\cite{Damour:1996ke}. Induced scalarization~\cite{Damour:1992we,Damour:1993hw,Damour:1996ke,Barausse:2012da,Palenzuela:2013hsa} is similar to what occurs in paramagnetism, where the individual magnetic moments of a large collection of atoms align themselves in the presence of a strong, external magnetic field. This alignment induces an overall magnetization of the collection of atoms. In the case of a neutron star (NS) in the presence of an external scalar field, e.g.~one supported by its binary companion, the star can develop a scalar field of its own.
Similarly, spontaneous~\cite{Damour:1992we,Damour:1993hw,Damour:1996ke} and dynamical~\cite{Barausse:2012da,Palenzuela:2013hsa} scalarization can both be understood in the context of spontaneous magnetization. In this phenomenon, a collection of unaligned magnetic moments will spontaneously align themselves in some direction as the temperature is lowered past a critical point, even in the absence of an external field. This occurs because, when the temperature is low enough, a new energy minimum appears -- a broken symmetry state that is associated with a non-zero net magnetization. A similar second-order phase transition occurs as either the compactness of an individual body or the absolute magnitude of the binding energy of a binary system reaches a large enough value. When this happens, the effective potential of the scalar field changes and a new, spontaneously broken minimum appears. This forces the scalar field to ``roll down'' to a non-zero expectation value.
The compactness or binding energy at which this phase transition occurs is a function of the coupling constants of the ST theory.
However, the compactness (or the binding energy) of a system, be it an individual NS or a binary, is also a function of the NS's equation of state (EoS).
As a result, whether a systems scalarizes or not also depends on the EoS.
We will discuss these theories in more detail in Sec.~\ref{sec:theory}, but it suffices here to say that scalarization (spontaneous, induced or dynamical) only occurs in these theories when a particular coupling constant, $\beta_{\mbox{\tiny ST}}$, is sufficiently large and negative (for a fixed EoS, NS compactness, and orbital separation). More negative values of this parameter result in scalarization for systems with smaller compactnesses and larger orbital separations~\cite{Damour:1992we,Barausse:2012da,Palenzuela:2013hsa}.
A little-appreciated problem exists for these theories if one wishes $\beta_{{\mbox{\tiny ST}}}$ to be negative, such that scalarization can occur. Reference~\cite{Damour:1996ke,PhysRevD.48.3436,PhysRevLett.70.2217} showed that in a cosmological evolution, $\beta_{{\mbox{\tiny ST}}} > 0$ forces the ST theory to approach GR exponentially, i.e.~GR is an attractor in the theory phase space, and thus the parameterized post-Newtonian (ppN) parameters are exponentially close to their GR values. However, by this same argument, we show in Appendix~\ref{AppA} that
a cosmological evolution with $\beta_{{\mbox{\tiny ST}}} < 0$ makes GR a repeller, forcing ppN parameters in this theory to deviate from their GR values.
A more rigorous, nonlinear analysis linking cosmological scales to galactic and eventually Solar System ones would be useful to draw definitive conclusions,
but based on these results, the requirement
$\beta_{{\mbox{\tiny ST}}} < 0$, which would enable scalarization, seems incompatible with Solar System experiments.
Regardless, these problems might be avoidable if an external potential (see e.g. Ref.~\cite{cosmoST}) is included, with a minimum at small values of the scalar field.
Neglecting the above problems, as done regularly in the
literature~\cite{Damour:1992we,Damour:1993hw,kramer-wex,Freire:2012mg,
Barausse:2012da,Palenzuela:2013hsa,Shibata:2013pra}, we can study the effects of
scalarization on astrophysical observations of binary systems, such as binary
pulsars, and then use these observations to constrain ST theories. If the binary
components support a scalar field, then an unequal mass binary will decay faster
than in GR due to the emission of dipolar radiation by the scalar
field~\cite{Damour:1992we,Damour:1993hw}. Such a decay is stringently
constrained by binary pulsar observations~\cite{kramer-wex,Freire:2012mg}, which
can place strong bounds on the existence of dipole radiation, scalarization, and
the magnitude of $\beta_{{\mbox{\tiny ST}}}$. Using observations of a pulsar-white dwarf
binary (J1738+0333) and a single-polytrope EoS with polytropic index $\Gamma
=2.34$, Ref.~\cite{Freire:2012mg} has constrained $\beta_{{\mbox{\tiny ST}}} \gtrsim -4.75$.
Using an APR4 (soft) and an H4 (stiff) EoS, Ref.~\cite{
Shibata:2013pra} used the binary pulsar observations of Ref.~\cite{Freire:2012mg} to constrain $\beta_{{\mbox{\tiny ST}}} \gtrsim -4.5$ and $\beta_{{\mbox{\tiny ST}}} \gtrsim -5$ respectively.
Clearly, constraints on $\beta_{{\mbox{\tiny ST}}}$ depend on the NS EoS. For a given EoS, only certain sufficiently large, negative values of $\beta_{{\mbox{\tiny ST}}}$ produce spontaneous scalarization. The bounds quoted above are roughly the least negative values of $\beta_{{\mbox{\tiny ST}}}$ that allow for scalarization, given a particular EoS. For example, Ref.~\cite{Shibata:2013pra} showed that for the pulsar in Ref.~\cite{Freire:2012mg} with mass $\approx 1.46 M_{\odot}$, an APR4 EoS requires $\beta_{{\mbox{\tiny ST}}} < -4.5$ for scalarization to occur, while an H4 EoS requires $\beta_{{\mbox{\tiny ST}}} < -5$, precisely the constraints quoted in Ref.~\cite{Shibata:2013pra}. Thus, binary pulsar observations constrain a region in the $\beta_{{\mbox{\tiny ST}}}$--EoS space, which means that bounds on $\beta_{{\mbox{\tiny ST}}}$ can be weakened if one considers stiffer EoSs that lead to less compact stars (larger radius given a fixed mass). Additionally, binary pulsars that have been observed thus far are very widely separated (typical separation larger than $10^{5}$ km), and so their orbital binding energy is not large enough to activate dynamical scalarization.
The GWs emitted during the late inspiral of NS binaries may also allow for constraints on scalarization, as these waves are produced when the orbital binding energy is very large and, like binary pulsars, GW observations are extremely sensitive to the orbital decay rate. In this paper, we investigate such an idea by
\begin{enumerate}
\item[(i)] constructing an analytical model of the GWs emitted during late NS inspirals in ST theories,
\item[(ii)] calculating the response function of interferometric detectors to such waves in the time- and frequency-domains through the stationary phase approximation (SPA), and
\item[(iii)] carrying out a detailed, Bayesian parameter estimation and model-selection study, assuming a GW detection with second-generation detectors (in the currently-planned configuration).
\end{enumerate}
One of our main results is the following:
\begin{itemize}[leftmargin=0.2in,rightmargin=0.2in]
\item[] {\bf{GWs emitted in the late inspiral of NS binaries can be used to place constraints on ST theories that are comparable to binary pulsar ones, provided at least one binary component is sufficiently compact to be already spontaneously scalarized by the time the emitted GWs enter the detectors' sensitivity band.}}
\end{itemize}
Second-generation detectors, like aLIGO, aVirgo, and KAGRA, will be sensitive to GWs emitted by binaries with orbital frequencies above $\sim5$ Hz. If a NS in a binary is sufficiently compact, for a given EoS and value of $\beta_{{\mbox{\tiny ST}}}$, to be already spontaneously scalarized by the time the binary crosses this frequency threshold, then a GW observation consistent with GR can be used to rule out the existence of dipole radiation and constrain the magnitude of $\beta_{{\mbox{\tiny ST}}}$. For example, given a polytropic EoS and a NS binary with (gravitational) masses $(1.4074,1.7415) M_{\odot}$, a single GW observation consistent with GR at a signal-to-noise ratio (SNR) of 15 would allow us to place the constraint $\beta_{{\mbox{\tiny ST}}} \gtrsim -4.5$. This constraint is similar in strength to that obtained with current binary pulsar observations and polytropic EoSs~\cite{Freire:2012mg}, but it is complementary in that it derives from
sampling the dynamical and non-linear regime of the gravitational interaction.
Another main result of this paper is the following:
\begin{itemize}[leftmargin=0.2in,rightmargin=0.2in]
\item[] {\bf{Dynamical scalarization in the late inspiral of NS binaries will be difficult to constrain with GWs, unless the system scalarizes at a low enough orbital frequency (i.e.~large enough orbital separation) that a sufficient amount of SNR is accumulated while the ST modifications are active.}}
\end{itemize}
Reference~\cite{Sampson:2013jpa} proved that an abrupt activation of dipolar
radiation can only be observed with GWs emitted by binaries if it occurs at an
orbital frequency well below 50Hz, assuming an aLIGO type detector and an SNR
of 10. Such a threshold frequency corresponds to an orbital separation larger
than roughly $35 m$, where $m$ is the total mass of the NS binary, or
alternatively, $\approx 150$ km for a binary with total mass $3 M_{\odot}$.
Dynamical scalarization at such large separations only happens for quite specific, finely tuned binary systems
given
realistic NS EoSs and values of $\beta_{{\mbox{\tiny ST}}}$ not already ruled out by binary
pulsar observations~\cite{Freire:2012mg}. In this analysis, only three systems,
out of a total of $80$ examined, underwent dynamical scalarization at orbital frequencies
below $100$ Hz. Of these three, only one exhibited ST effects that were strong
enough to detect with aLIGO at an SNR of 15.
We additionally analyze whether another signature of dynamical scalarization -- an early plunge once the scalar field activates -- is detectable with aLIGO-like detectors. We find that:
\begin{itemize}[leftmargin=0.2in,rightmargin=0.2in]
\item[] {\bf{Early plunge and merger due to dynamical scalarization is detectable for certain frequency ranges using model-independent templates.}}
\end{itemize}
We modeled the early plunge through a \emph{toy model}, a Heaviside truncation of the GW Fourier
amplitude. For this truncation to be detectable, we find that it must occur between orbital
frequencies of roughly $45$ Hz and $150$ Hz, assuming an SNR 15 if no truncation was present.
For plunges that occur at lower frequencies, not enough signal
would be present in the detector's sensitivity band to lead to a detection in
the first place. For plunges that occur at higher frequencies, the detector's
noise is already large enough that such a plunge is difficult to detect.
The analysis described above required a Markov-Chain Monte-Carlo (MCMC) mapping of the likelihood surface to obtain posteriors and Bayes factors (BFs) between a GR and a non-GR model, given a scalarized GW injection, all of which is computationally expensive. Thus, another important result of this paper is the following:
\begin{itemize}[leftmargin=0.2in,rightmargin=0.2in]
\item[] {\bf{We re-derive a computationally inexpensive, data analysis measure to determine whether modified gravity effects are detectable, which we call the \emph{the effective cycles of phase}. For the first time, we connect this quantity directly to the Bayes factor. We find that roughly $4$ cycles are needed for a modified gravity effect to be detectable at SNR 15.}}
\end{itemize}
First derived in \cite{PhysRevD.78.124020}, the effective cycles are defined as a certain noise-weighted integral of the GW
amplitude and the phase difference between a GR and a non-GR inspiral. This
measure is inspired by the \emph{useful cycles of phase}, introduced in
Ref.~\cite{Damour:2000gg}, but it differs from this quantity in that the BF can
be estimated \emph{analytically} in terms of effective
cycles. Therefore, the effective cycles are directly connected to an important
data analysis measure of detectability in model hypothesis testing.
The effective cycles are a noise-weighted dephasing measure that is ideal for studying the distinguishability between models. A non-noise-weighted dephasing is not a good measure, in spite of being often used in theoretical studies, e.g.~\cite{Baiotti:2011am,Yagi:2013sva,Gerosa:2014kta}. The effective cycles take into account the fact that, for departures from GR to be detectable, they must occur in a frequency range in which sufficient SNR is accumulated. This is particularly difficult for GR departures that only become significant at kHz frequencies.
Another question that arose while we carried out this analysis was whether
custom-made waveforms are necessary to detect the effects of ST gravity, or
whether model-independent templates are sufficient. This question, of course, is
SNR dependent, and as explained above, dynamical scalarization is not easily
measurable with the SNRs expected in advanced detectors, except in a few cases.
Moreover, it is very difficult to develop
analytical templates that accurately model the effects of dynamical
scalarization. We therefore consider the detectability of spontaneous/induced
scalarization effects with custom-made versus model-independent templates, and
obtain the following result:
\begin{itemize}[leftmargin=0.2in,rightmargin=0.2in]
\item[] {\bf{More complicated, model-dependent templates are found to be comparably efficient at detecting GR deviations as the simplest, model-independent waveforms one can construct.}}
\end{itemize}
As a proxy for model-independent templates, we use the parameterized post-Einsteinian (ppE) approach of Ref.~\cite{PPE}. The simplest version of such waveforms include only leading post-Newtonian (PN) order modifications to GR in the waveform; thus, they do not contain the precise PN sequence of terms that a custom-made ST theory template (or signal) would possess. This additional structure should allow custom-made templates to be more effective at detecting GR deviations present in the signals they are designed to capture. The enhancement leads to an increased detection efficiency at very high SNR, but negligible effects for signals with SNR's we expect to observe. Additionally, we find that more complicated ppE templates, such as those that include a step-function in the phase to activate a modified gravity effect above a certain frequency, are as effective as the simplest ppE models at detecting ST-type GR deviations. These results are consistent with those in Ref.~\cite{Sampson:2013lpa}.
The rest of this paper presents the details of the results described above and is organized as follows. In Sec.~\ref{sec:theory}, we give an introduction to the non-GR models of interest in this paper. Next, in Sec.~\ref{sec:waveforms}, we describe in detail the methods we used to develop the waveforms tailored to these models. Then, in Secs.~\ref{sec:detect} and~\ref{sec:BF}, we address the question of detectability using effective cycles and a Bayesian analysis respectively. In the final section of this paper, Sec.~\ref{sec:conclusions}, we conclude and point to future research. Throughout this paper, Latin letters refer to spatial indices, Greek letters refer to space-time indices, and we use geometric units in which $G = c = 1$. Additionally, all masses quoted refer to gravitational mass and, unless otherwise specified,
we employ the term ``spontaneously scalarized'' binary (or signal) to
refer to binaries (or signals arising from binaries) whose components undergo spontaneous or induced scalarization. We do so because a necessary condition for induced scalarization to happen is the presence of at least
one spontaneously scalarized star. The term ``dynamically scalarized'' binary (or signal) will
denote binaries (or signals arising from binaries) that undergo dynamical scalarization.
\section{Introduction to Scalar-Tensor Theories \label{sec:theory}}
In this section, we briefly describe the theory we will study. Initially presented in Ref.~\cite{Damour:1992we,Damour:1993hw}, this theory has recently been revisited in the context of compact binary inspirals and mergers in Ref.~\cite{Barausse:2012da,Palenzuela:2013hsa,Shibata:2013pra}. Here, we review the basics of this theory, following mainly the presentation of Ref.~\cite{Palenzuela:2013hsa}.
\subsection{Basics \label{subsec:basics}}
Generic ST theories are defined by the Jordan-frame action
\be
S = \int d^{4}x \frac{\sqrt{-g}}{2 \kappa} \left[ \phi R - \frac{\omega(\phi)}{\phi} \partial_{\mu} \phi \partial^{\mu} \phi\right] + S_{M}[\chi,g_{\mu\nu}]\,,
\label{eq:Jordan-S}
\ee
where $\kappa = 8 \pi G$, $\phi$ is a scalar field, $R$ is the Ricci scalar associated with the Jordan-frame metric, $g_{\mu \nu}$, and $\chi$
are additional matter degrees of freedom that couple directly to the metric.
The function, $\omega(\phi)$, defines the particular ST theory in play (in some cases there is also a potential function, $V(\phi)$, but here this potential is set to zero). For example, Fierz-Jordan-Brans-Dicke (FJBD) theory~\cite{fierz,jordan,BD} is defined by this action with $\omega(\phi) = \omega_{{\mbox{\tiny BD}}} = {\rm{const}}$. In this paper, we will consider the class of theories studied in~\cite{Damour:1992we,Damour:1993hw}, which are defined by the action of Eq.~\eqref{eq:Jordan-S} with
\be
\label{eq:omega-of-phi}
\omega(\phi) = - \frac{3}{2} - \frac{\kappa}{4 \beta \log {\phi}}\,,
\ee
where $\beta$ is a dimensional constant, related to the dimensionless coupling constant of the theory, $\beta_{\mbox{\tiny ST}}$, by $\beta = (4 \pi G) \, \beta_{\mbox{\tiny ST}} $. The asymptotic value of $\phi$ at spatial infinity, together with the value of $\beta_{\mbox{\tiny ST}}$, controls the magnitude of the modifications to GR.
The Jordan-frame action of Eq.~\eqref{eq:Jordan-S} can be rewritten in the ``Einstein frame'' via
\begin{multline}
S = \int d^{4}x \sqrt{-g^{E}} \left(\frac{R^{E}}{2 \kappa} - \frac{1}{2} g^{\mu \nu}_{E} \partial_{\mu} \psi \partial_{\nu} \psi \right) \\+ S_{M}\left[\chi,g_{\mu \nu}^{E}/\phi(\psi)\right]\,,
\end{multline}
where the Einstein-frame metric is related to the Jordan-frame one via $g_{\mu \nu}^{E} = \phi \; g_{\mu \nu}$. The Einstein-frame scalar field $\psi$ is related to its Jordan-frame counterpart via
\be
\left(\frac{d \log \phi}{d\psi}\right)^{2} = \frac{2 \kappa}{3 + 2 \omega(\phi)}\,.
\label{eq:phi-of-psi-DE}
\ee
In the theories of interest in this paper, this differential equation can be solved to obtain
\be\label{eqphi}
\phi = \exp\left(-\beta \psi^{2}\right)\,,
\ee
choosing $\psi=0$ when $\phi = 1$. Using this equation in Eq.~\eqref{eq:phi-of-psi-DE}, one finds
\be
\psi[\omega(\phi)] = \frac{1}{2 |\beta|} \left[\frac{2 \kappa}{3 + 2 \omega(\phi)} \right]^{1/2}\,.
\ee
In the Einstein frame, the field equations for the metric and the equations of motion for the scalar field are
\begin{align}\label{mod_eins}
G_{\mu \nu}^{E} &= \kappa \left(T_{\mu \nu}^{\psi} + T_{\mu \nu}^{M,E}\right)\,,
\\
\label{eq:scalar-field-eq}
\square^{E} \psi &=- \beta \psi T^{M,E}\,,
\end{align}
where we have defined
\begin{align}
T_{\mu \nu}^{\psi} &= \partial_{\mu} \psi \partial_{\nu} \psi - \frac{1}{2} g_{\mu \nu}^{E} g^{\alpha \beta}_{E} \partial_{\alpha} \psi \partial_{\beta} \psi\,,
\end{align}
and $T^{M,E}$ is the Einstein-frame trace of $T_{\mu \nu}^{M,E}=T_{\mu \nu}^{M}/\phi$, the matter stress-energy tensor in the Einstein frame.
Because of the field redefinition to Einstein-frame variables, the stress-energy tensor conservation $\nabla_\mu T_M^{\mu\nu}=0$ becomes
\be
\nabla_{\mu}^{E} T^{\mu \nu}_{E} = \beta \psi T_{E} g_{E}^{\mu \nu} \partial_{\mu} \psi\,,
\label{eq:SEP}
\ee
which also follows from the field equations \eqref{mod_eins} and \eqref{eq:scalar-field-eq}. This equation implies, in particular, that test particles (i.e.~point particles with negligible mass) do \textit{not} follow geodesics of the Einstein frame metric $g^E_{\mu\nu}$, although they do follow geodesics of the
original Jordan-frame metric $g_{\mu\nu}$. This means that the weak equivalence principle (i.e., the universality of free fall for
weakly gravitating bodies) is satisfied in these theories. Nevertheless, as will become clearer in the next section, the strong version
of the equivalence principle (i.e.~the universality of free fall for strongly gravitating bodies) is \textit{not} satisfied in
scalar tensor theories. This is because of the presence of ``scalar charges'' for strongly gravitating bodies, whose appearance can be ultimately traced to
the right-hand side of Eq.~\eqref{eq:SEP} having a non-zero value.
Solar System tracking of the Cassini spacecraft implies the constraint $w_{{\mbox{\tiny BD}}} > w_{\rm Cassini}\equiv 4 \times 10^{4}$~\cite{will-living,Bertotti:2003rm}
on FJBD theory. Because the class of ST theories that we consider reduces to FJBD in the Solar System (with $\omega_{{\mbox{\tiny BD}}}=\omega_{0}$ related to the the asymptotic value of $\psi$, which we denote by $\psi_0$, via Eqs.~\eqref{eq:omega-of-phi} and \eqref{eqphi}), this translates into the bound
\be
\psi_0 \equiv \frac{1}{2 |\beta|} \left(\frac{2 \kappa}{3 + 2 \omega_{0}} \right)^{1/2} \lesssim 1.26\times 10^{-2} \frac{G^{1/2}}{|\beta|}\,.
\ee
Binary pulsar observations of the orbital decay rate also constrain the theory, since dipolar radiation would greatly accelerate the inspiral~\cite{Damour:1998jk,Freire:2012mg}. These observations require that $\beta_{\mbox{\tiny ST}} \gtrsim -4.75$~\cite{Freire:2012mg}, but as discussed in the introduction, this constraint depends on which EoS is used~\cite{Shibata:2013pra}.
\subsection{Time-Domain Scalar Charges \label{subsec:alphas}}
Given a binary system that consists of two bodies of masses $m_1$ and $m_2$, the evolution of the GWs emitted by this binary will depend on the scalar charges of the two bodies, $\alpha_{1}$ and $\alpha_{2}$. In turn, these charges depend on $\beta_{{\mbox{\tiny ST}}} \equiv \beta/(4 \pi G)$, but also on the NS compactness $C$ and, in a binary in quasi-circular motion, on the (magnitude of their) orbital velocity $v$ (equivalently, their separation or their orbital frequency). It is therefore necessary to have a good analytic representation of these charges as a function of $\beta_{{\mbox{\tiny ST}}}$, $C$, and $v$, in order to construct accurate SPA templates.
The scalar charges are defined by~\cite{Damour:1992we}
\be
\alpha_{A} = -\frac{1}{\sqrt{4 \pi G}} \frac{\partial \ln m_{A}^{E}(\psi)}{\partial {\psi}}\,,
\ee
where $m_{A}(\psi)$ is the mass parameter that enters the point-particle action in the Einstein frame. A related quantity, the {\emph{sensitivity}}, $s_{A}$, can be similarly defined by~\cite{1975ApJ...196L..59E}
\be
s_{A} = \frac{\partial \ln m_{A}(\phi)}{\partial \ln \phi}\,,
\ee
where $m_{A}(\phi)$ is the mass parameter that enters the point-particle action in the Jordan frame. These derivatives are to be taken by keeping the baryonic mass fixed, and the two masses are related via $m_{A}^{E} = m_{A} \phi(\psi)^{-1/2}$. The sensitivities and the scalar charges are then related by~\cite{Damour:1995kt,Mirshekari:2013vb,Palenzuela:2013hsa}
\be
\alpha_{A} = - \frac{2 s_{A} - 1}{\sqrt{3 + 2 \omega_{0}}}\,,
\ee
where $\omega_{0}$ must be greater than $4 \times 10^{4}$ due to the Cassini bound~\cite{will-living,Bertotti:2003rm}.
In an FJBD theory with a given $\omega_{\rm BD}$, the scalar charges (and thus the sensitivities) are parameters determined exclusively by the compact object's EoSs. Will and Zaglauer~\cite{Will:1989sk} found that in this theory the sensitivities are in the interval $(0.1,0.3)$ for NSs, becoming $0.5$ in the black hole limit. Of course, the scalar charges in this theory are much smaller than the sensitivities, as the former are suppressed by $\approx\omega_{0}^{-1/2}$ relative to the latter. Fixing the EoS, the sensitivities depend only on the mass of the object. Thus, since NS masses are expected to be in the range $(1,2.5) M_{\odot}$, NS binaries have $s_{1} \approx s_{2}$, and dipole radiation is suppressed~\cite{Will:1989sk}. Such suppression explains why it would be difficult to observe or constrain dipole radiation from GWs emitted during binary NS inspirals within FJBD theory.
In the ST theories of Ref.~\cite{Damour:1992we,Damour:1993hw}, however, the scalar charges can be spontaneously/dynamically generated. In this process of scalarization, the charges can be excited once the gravitational energy of the system exceeds a certain threshold. In isolation, this energy is simply proportional to the NS compactness, $C = M_{ }/R_{ }$, the ratio of the NS mass to its radius. When in a binary, this energy is not only due to the individual compactnesses, but also to the binding energy of the system, which scales as $m_1 m_2/r_{12}$, with $(m_1,m_2)$ the NS masses, and $r_{12}$ the orbital separation.
The behavior of the scalar charges during the inspiral and plunge of a NS binary has so far only been calculated semi-analytically in Ref.~\cite{Palenzuela:2013hsa}
for a polytropic EoS with exponent $\Gamma=2$ and maximum NS gravitational mass of $1.8 M_{\odot}$, and these results have been validated by comparing the binary's orbital evolution to the fully non-linear simulations of Ref.~\cite{Barausse:2012da}. Using the results of Ref.~\cite{Palenzuela:2013hsa}, Fig.~\ref{fig:alpha} shows the scalar charge for the more massive star in a NS binary system with $(m_{1},m_{2}) = (1.4074,1.7415) M_{\odot}$ as a function of the dominant GW frequency $f$ (twice the orbital frequency for a quasi-circular binary) up to contact for ST theories with $\beta_{\mbox{\tiny ST}} = -3.0, -3.25, -3.5$ and $-4.5$.
\begin{figure}[ht]
\includegraphics[clip=true,angle=-90,width=0.48\textwidth]{chargefits.eps}
\caption{\label{fig:alpha} (Color Online) The upper panel shows the scalar charges for a $1.7415 M_\odot$ NS with a $1.4074 M_\odot$ companion, for various values of $\beta_{\mbox{\tiny ST}}$ and as a function of GW frequency. The actual data is shown with symbols, and the fits to the data by the different line styles. In the lower panel, we plot the percentage error between the data and the fits. Observe that this error never exceeds $\approx 7\%$.}
\end{figure}
As noted, the construction of an SPA waveform in ST theories will require a parameterization of these scalar charges as a function of $\beta_{{\mbox{\tiny ST}}}$, $m_A$, and $v$. We desire an analytic expression for the scalar charges because, when calculating the Fourier transform of the GW phase in the SPA, we need to analytically compute {\emph{indefinite}} integrals of functions of these charges. Thus motivated, we will use the following fitting function:
\be
\label{eq:alpha-fit-poly}
\alpha_{A} = \sum_{i=0}^{i_{\rm max}} a_{i}^{(A)} \; v^{i}\,,
\ee
where $v \equiv (\pi m f)^{1/3}$, with $f$ the dominant GW frequency (twice the orbital frequency) and the coefficients $a_{i}^{(A)}$ are functions of $(m_A,m,\beta_{\mbox{\tiny ST}})$. We further expand these coefficients as polynomials:
\be a_{i}^{(A)} =\sum_{j=0}^{j_{\rm max}} \sum_{k=0}^{k_{\rm max}} \sum_{\ell=1}^{\ell_{\rm max}}
(a_{i j k \ell}^{(A)}) \left(-\beta_{\mbox{\tiny ST}}\right)^{\ell} m_{A}^{j} m^{k}\,
\ee
The more terms kept in the sum, of course, the more closely the function approximates the numerical data. We find empirically that the choice $ i_{\rm max} = l_{\rm max} = 2$ and $j_{\rm max} = k_{\rm max} = 3$ suffices for our purposes, which leads to $2 \times 3 \times 3 = 18$ fitting coefficients at each PN order.
We use the fitting function described above to fit for the scalar charges as a function of $v$. First, we use the data from Ref.~\cite{Palenzuela:2013hsa} to numerically construct $\alpha_{A}$ as a function of $v$, from a GW frequency of $10$ Hz (the beginning of the aLIGO sensitivity band) up to contact, with a fine discretization, for 37 systems with different masses $(m_{1},m_{2})$ and values of $\beta_{\mbox{\tiny ST}}$. Each of these data sets is slightly noisy in the low-frequency regime due to small numerical errors, so before proceeding, we smooth each of them with a moving average algorithm using the nearest ten neighbors.
When carrying out the fits, we do not use the full domain of the data (from $f=10 \; {\rm{Hz}}$ to contact), but rather restrict attention to the low velocity regime. As found in Ref.~\cite{Blanchet:2009sd,Blanchet:2010zd,Blanchet:2010cx}, when fitting numerical data to a PN function, the high velocity regime should not be included in the numerical data. This is because this regime would contaminate the fitting coefficients, as they attempt to capture both the low and high velocity behavior of the function. Moreover, one should not use a very high PN order fitting function, as the high PN order terms would contaminate the low-PN order ones. For these reasons, we choose $i_{\rm max} = 2$ and fit in the region $(10,800) \; {\rm{Hz}}$, which we found empirically to yield robust results, as we will show below.
Finally, in doing the fits, we have neglected scalar charges that undergo dynamical scalarization (i.e.~charges that are close to zero when the binary enters the LIGO band at an orbital frequency of 5 Hz, and then grow very quickly later in the inspiral) for the following reason. The SPA calculation requires that we bivariately Taylor expand all quantities in both $v \ll 1$ \emph{and} in the GR deviation parameter, which in the ST case is $\alpha_{A}$. Although $\alpha_{A} \ll 1$ during the inspiral, $d\alpha_{A}/d \ln{v} > 1$ at a certain frequency, around $f = 300 \; {\rm{Hz}}$ for the cases considered in Fig.~\ref{fig:D-scalar}. Thus, during dynamical scalarization, one should not expand in derivatives of
$\alpha_{A}$, which makes any analytic treatment very difficult. Third, a polynomial in velocity is ill-suited for representing scalar charges that undergo dynamical scalarization. If one insisted on using such a fitting function, then the coefficients $a_{i}^{(A)}$ would grow by factors of $10^{3}$ with increasing $i$. Such a highly divergent behavior of the fitting coefficients renders any subsequent PN expansion useless.
\begin{figure}[th]
\includegraphics[clip=true,width=8.5cm]{D-scalar-charges-new.eps}
\caption{\label{fig:D-scalar} (Color Online) Logarithmic derivative of the scalar charges with respect to velocity as a function of GW frequency for a NS binary with masses $(m_1,m_{2})= (1.4074,1.7415) M_\odot$ computed with the numerical results presented
in Ref.~\cite{Palenzuela:2013hsa}. Observe that the derivative of the charge in the $\beta_{\mbox{\tiny ST}} = -4.25$ case rises very rapidly around $f = 300 \; {\rm{Hz}}$, rapidly exceeding unity. On the other hand, the derivative in the $\beta_{\mbox{\tiny ST}} = -4.5$ case remains smaller than unity during the entire inspiral.}
\end{figure}
Given this, we fit Eq.~\eqref{eq:alpha-fit-poly} independently to each set of clean data corresponding to a \emph{particular} value of $\beta_{\mbox{\tiny ST}}$ rather than fitting all the data with the same set of coefficients at the same time. We proceed this
way because the scalar charges for larger values of $|\beta_{\mbox{\tiny ST}}|$ (i.e., $\beta_{\mbox{\tiny ST}} = -4.5$) are much larger than, for example, the charges at $\beta_{\mbox{\tiny ST}}=-3.5$. This, paired with our lack of data for values of $|\beta_{{\mbox{\tiny ST}}}| < 3.0$ and the sparseness of the data in $(m_{1},m_{2})$ space, causes our algorithm to generate poor fits for small values of $\beta_{\mbox{\tiny ST}}$. In particular, the fitting function obtained by simultaneously fitting all the data does not monotonically approach zero as $\beta_{\mbox{\tiny ST}} \to 0$. We therefore arrive at a set of $n=18$ coefficients at each PN order for each different value of $\beta_{\mbox{\tiny ST}}$. Figure~\ref{fig:alpha} shows the fitted function and the numerical data for several values of $\beta_{\mbox{\tiny ST}}$, where note the error in the fit never exceeds $7 \%$ in the frequency window $[10:10^{3}]$ Hz.
\section{Developing Waveforms for Scalar-Tensor Theories \label{sec:waveforms}}
In this section, we describe how to construct the time-domain response of the aLIGO detectors to impinging GWs. We use the restricted PN approximation, following mostly the presentation in Ref.~\cite{Yunes:2009yz,Chatziioannou:2012rf}. We then construct the frequency domain, SPA waveforms from these time-domain functions, and discuss their regimes of validity.
\subsection{Time-Domain Response}
\label{sec:time-domain-response}
The response of a GW interferometer to a signal can be computed by investigating how the metric perturbation affects the geodesic deviation. Doing so, one finds~\cite{Kidder:1995zr,poisson2014gravity}
\be
\label{eq:t-dom-resp}
h(t) = F_{+} h_{+} + F_{\times} h_{\times} + F_{{\mbox{\tiny b}}} h_{{\mbox{\tiny b}}} + F_{{\mbox{\tiny L}}} h_{{\mbox{\tiny L}}} + F_{{\mbox{\tiny se}}} h_{{\mbox{\tiny se}}} + F_{{\mbox{\tiny sn}}} h_{{\mbox{\tiny sn}}}\,.
\ee
The quantities $(F_{+},F_{\times},F_{{\mbox{\tiny b}}},F_{{\mbox{\tiny L}}},F_{{\mbox{\tiny se}}},F_{{\mbox{\tiny sn}}})$ are beam or angular pattern functions that depend on the geometry of the detector [see e.g.~Eqs.(2)--(7) in Ref.~\cite{Chatziioannou:2012rf}]. The quantities $(h_{+},h_{\times},h_{{\mbox{\tiny b}}},h_{{\mbox{\tiny L}}},h_{{\mbox{\tiny se}}},h_{{\mbox{\tiny sn}}})$ are the six possible GW polarizations in a generic theory of gravity: the plus-mode, the cross-mode, the scalar ``breathing'' mode, the scalar ``longitudinal'' mode, and the two vector modes.
Recall that GR, being a massless spin-2 theory, has only two propagating degreees of freedom.
The waveform polarizations can be computed from the trace-reversed metric perturbation in the far-zone (at a distance much greater than a GW wavelength from the center of mass of the binary) via
\begin{align}
h_{+} &= e^{+}_{ij} \left( P^{i}_{m} P^{j}_{l} \bar{h}^{ml} - \frac{1}{2} P^{ij} P_{ml} \bar{h}^{ml} \right)\,,
\\
h_{\times} &= e^{\times}_{ij} \left( P^{i}_{m} P^{j}_{l} \bar{h}^{ml} - \frac{1}{2} P^{ij} P_{ml} \bar{h}^{ml}\right)\,,
\\
h_{{\mbox{\tiny b}}} &= \frac{1}{2} \left(\hat{N}_{jk} \bar{h}^{jk} - \bar{h}^{00} \right)\,,
\\
h_{{\mbox{\tiny L}}} &= \hat{N}_{jk} \bar{h}^{jk} + \bar{h}^{00} - 2 \hat{N}_{j} \bar{h}^{0j}\,,
\\
h_{{\mbox{\tiny sn}}} &= e_{k}^{x} \left[P^{k}_{j} \left(\hat{N}_{i} \bar{h}^{ij} - \bar{h}^{0j} \right)\right]\,,
\\
h_{{\mbox{\tiny se}}} &= e_{k}^{y} \left[P^{k}_{j} \left(\hat{N}_{i} \bar{h}^{ij} - \bar{h}^{0j} \right)\right]\,,
\end{align}
where we use multi-index notation $\hat{N}^{l_1 \ldots l_n} = \hat{N}^{l_1} \cdots
\hat{N}^{l_n}$. Here, $\hat{N}^{i}$ is a unit vector pointing from the source
to the detector, $P_{ij} = \delta_{ij} - \hat{N}_{ij}$ is a projection operator
orthogonal to $\hat{N}^{i}$, and $e_{i}^{x}$ and $e^{y}_{i}$ are basis vectors
orthogonal to $\hat{N}^{i}$. If we choose a coordinate system defined by the
triad $(\hat{i}^{i},\hat{j}^{i},\hat{k}^{i}$), such that $\hat{N}^{i}$ is given
by
\be
\hat{N}^{i} = \sin{\iota} \; \hat{j}^{i} + \cos{\iota} \; \hat{k}^{i}\,,
\ee
then
\begin{align}
e_{x}^{i} &= - \hat{i}^{i}\,,
\\
e_{y}^{i} &= \cos{\iota} \; \hat{j}^{i} - \sin{\iota} \; \hat{k}^{i}\,,
\\
e^{+}_{ij} &= \frac{1}{2} \left(e^{y}_{i} e^{y}_{j} - e^{x}_{i} e^{x}_{j}\right)\,,
\\
e^{+}_{ij} &= \frac{1}{2} \left(e^{x}_{i} e^{y}_{j} + e^{y}_{i} e^{x}_{j}\right)\,,
\end{align}
where $\iota$ is the inclination angle, such that $\hat{L}^{i} \hat{N}_{i} = \cos{\iota}$, with $\hat{L}^{i}$ the unit orbital angular momentum vector.
The ST theories of Ref.~\cite{Damour:1992we,Damour:1993hw} have, in principle, three propagating GW modes: the two transverse-traceless modes ($h_{+}$ and $h_{\times}$) and the breathing mode ($h_{{\mbox{\tiny b}}}$). In the Eardley, {\emph{et al}} classification~\cite{Eardleyprd}, these modes correspond to the excitation of the Newman-Penrose scalars $\Psi_{4}$ and $\Phi_{22}$. One can show~\cite{Barausse:2012da}, however, that $\Phi_{22}$ is proportional to $\psi_{0}$ (or alternatively to $\omega_{0}^{-1/2}$) to leading order in a $\psi_{0} \ll 1$ expansion; $\Psi_{4}$ is, of course, independent of $\psi_{0}$ to leading order. Thus, the effect of $h_{{\mbox{\tiny b}}}$ in the response $h(t)$ is subleading in $\psi_{0}$ relative to the effect of $h_{+}$ and $h_{\times}$~\cite{Barausse:2012da}. In fact, this is exactly the same as in standard FJBD theory~\cite{TEGP,Will:1994fb}. Even if this were not the case and the response to the breathing mode were not suppressed as $1/\sqrt{\omega_{0}}$, the detectability of this mode would require
a network of detectors~\cite{Nishizawa:2009bf,Chatziioannou:2012rf,Hayama:2012au,Nishizawa:2013eqa,Horava:2014mra}. Given this, we will neglect the breathing mode in the response function.
\subsection{Time-Domain Waveform \\ and the Restricted PN Approximation}
As explained above, the interferometer response is dominated by the plus- and cross-polarized metric perturbations, $h_{+}(t)$ and $h_{\times}(t)$, which must be obtained by solving the modified field equations in the PN approximation. In the far-zone, these waves can be written as the following PN sum:
\be
\label{eq:def-h+x}
h_{+,\times}(t) = \frac{2 G \eta m}{D_{L}} x(t) \sum_{p=0}^{+\infty} x(t)^{p/2} H_{+,\times}^{(p/2)}(t)\,,
\ee
where $D_{L}$ is the (luminosity) distance from the source to the detector, $\eta = m_{1} m_{2}/m^{2}$ is the symmetric mass ratio, and recall that $m= m_{1} + m_{2}$ is the total mass, and $x(t) = [2 \pi G_{\rm eff} m F(t)]^{2/3}$ is the (time-dependent) PN expansion parameter of leading ${\cal{O}}(1/c^{2})$, with $F(t)$ the orbital frequency. Notice that $x$ depends on $G_{\rm eff} \equiv G (1 + \alpha_{1} \alpha_{2})$, because in these theories the Newtonian force of attraction between two bodies, and thus, Kepler's third law of orbital motion, takes on the usual Newtonian expression, but with the replacement $G \to G_{\rm eff}$~\cite{Damour:1992we,Will:1989sk}. Recall that $\alpha_{1}$ and $\alpha_{2}$ are the scalar charges of NSs, which depend on the internal structure of the bodies and on the gravitational binding energy of the system, as discussed in Sec.~\ref{subsec:alphas}.
The time-functions $H_{+,\times}^{(p/2)}$ can always be written in terms of an amplitude ($A_{+,\times}^{(p/2,n)}(t)$ or $B_{+,\times}^{(p/2,n)}$(t)) and the binary's orbital phase $\phi(t)$:
\be
H_{+,\times}^{(p/2)} = \sum_{n=0}^{\infty} \left(A_{+,\times}^{(p/2,n)} \cos{n \phi} + B_{+,\times}^{(p/2,n)} \sin{n \phi} \right)\,.
\ee
To leading PN order (keeping only the $p=0$ term), the oscillating part of Eq.~\eqref{eq:def-h+x} reduces to
\begin{align}
\label{eq:restrictedPN-hp}
h_{+}(t) &= 2 {\cal{A}}(t) \left(1 + \cos^{2}{\iota}\right) \cos{2 \phi(t)}\,,
\\
\label{eq:restrictedPN-hc}
h_{\times}(t) &= -4 {\cal{A}}(t) \cos{\iota} \sin{2 \phi(t)}\,.
\end{align}
Here we have defined the time-dependent amplitude
\be
{\cal{A}}(t) = - \frac{{\cal{M}}}{D_{L}} [2 \pi G m F(t)]^{2/3}\,,
\ee
with ${\cal{M}} = \eta^{3/5} m$ the chirp mass, and $\iota$ is the inclination. Notice that this is functionally exactly the same result as in GR, because, as discussed in Sec.~\ref{sec:time-domain-response}, ST corrections to the amplitude scale as $\psi_{0}$, and are thus subleading.
The \emph{restricted PN approximation} consists of approximating the
time-domain waveform by the leading PN order terms in the waveform amplitudes,
without restricting the PN order in the waveform phase. That is, one keeps only the
leading, $p=0$ term in Eq.~\eqref{eq:def-h+x}, thus obtaining
Eqs.~\eqref{eq:restrictedPN-hp} and~\eqref{eq:restrictedPN-hc}, but as many PN
terms as one wishes in the orbital phase $\phi(t)$. Such an
approximation is reasonable in a first analysis because interferometric
detectors are much more sensitive to the phase of the response than the
amplitude.
The plus- and cross-polarized metric perturbations in ST theories, and thus the time-domain interferometric response, are different from those predicted by GR mainly because of the temporal evolution of the orbital phase. The phase can be obtained by integrating the expression
\be
\label{eq:Fdot}
\ddot{\phi} = 2\pi \dot{F} = 2 \pi \frac{dE_{b}}{dt}
\left(\frac{dE_{b}}{dF}\right)^{-1}
\ee
twice, where $E_{b}$ is the gravitational binding energy.
By the balance law, $d E_{b}/dt$ must equal (minus) the luminosity ${\cal{L}}$ of all propagating degrees of freedom in the far-zone. In GR, the only energy loss is due to the emission of GWs, but in ST theories one must also account for dipolar radiation induced by the propagation of the scalar mode:
\be
\label{eq-luminosity}
{\cal{L}} = \frac{G}{3} \eta^{2} \left(\alpha_{1} - \alpha_{2}\right)^{2} x^{4} + \frac{32}{5} G \eta^{2} x^{5}\,,
\ee
where the first term is due to dipole radiation and the second due to quadrupolar radiation. We then see that ST theories modify the evolution of the orbital phase, which is precisely the component of the interferometric response that detectors are most sensitive to.
\subsection{Fourier Response \\ and the Stationary Phase Approximation}
In GW data analysis, one uses the Fourier transform of the response function
(the Fourier response), instead of the time-domain response.
In computing the frequency- or time-domain response,
one can effectively neglect the time-dependence of
the beam pattern functions. These are induced by the rotation and motion of the
Earth, which occurs on a time-scale much longer than the duration of the GW
signal, which is typically less than 20 minutes for the signals that fall in the sensitivity band of
second-generation, ground-based detectors, such as aLIGO, aVirgo, and KAGRA.
Thus, the Fourier response can be written
as
\be
\tilde{h}(f) = \int_{-\infty}^{\infty} dt \, e^{2 \pi i f t} \left[F_{+} h_{+}(t) + F_{\times} h_{\times}(t) \right]\,,
\label{eq:F-response-def}
\ee
where $(F_{+},F_{\times})$ are effectively constant.
Using the restricted PN approximation [Eqs.~\eqref{eq:restrictedPN-hp} and~\eqref{eq:restrictedPN-hc}], we can rewrite the Fourier response of Eq.~\eqref{eq:F-response-def} as
\be
\tilde{h}(f) = \int_{-\infty}^{\infty} dt \, e^{2 \pi i f t} {\cal{A}}(t) \left[Q_{C} \cos 2 \phi(t) + Q_{S} \sin 2 \phi(t)\right]\,,
\label{eq:F-response-new}
\ee
where we have defined the cosine- and sine-projected beam-pattern functions
\begin{align}
Q_{C} &= 2 \left(1 + \cos^{2}{\iota}\right) \cos{2 \Psi} F_{+} - 4 \cos{\iota} \sin{2 \Psi} F_{\times}\,,
\\
Q_{S} &= 2 \left(1 + \cos^{2}{\iota}\right) \sin{2 \Psi} F_{+} + 4 \cos{\iota} \cos{2 \Psi} F_{\times}\,,
\end{align}
with $\Psi$ the polarization angle. For binary systems that are in a quasi-circular orbit and whose binary components are not spinning, the inclination angle and $Q_{C,S}$ are all constant. The quasi-circular and non-spinning approximations are sufficient for our analysis because most NS binaries are expected to have circularized and spun-down by the time they enter the sensitivity band of second-generation, ground-based detectors.
The integral in Eq.~\eqref{eq:F-response-new} that defines the Fourier response falls in the class of generalized Fourier integrals. When the integrands have a stationary point, namely a time $t_{0}$ at which $\dot{\phi}(t_{0}) = \pi f$, the integral can be approximated via the method of steepest descent, which to leading order reduces to the SPA~\cite{Bender,Yunes:2009yz}. For this approximation to hold, the amplitude of the integrand must be slowly-varying, while the phase must be rapidly oscillating, such that the integral is non-vanishing only in a small neighborhood around the stationary point.
Within the SPA, the integral in Eq.~\eqref{eq:F-response-new} reduces to
\be
\tilde{h}(f) = \tilde{A} f^{-7/6} e^{i \Psi(f)}\,,
\ee
where $f$ is the Fourier or GW frequency (twice the orbital frequency), the Fourier amplitude is a constant given by
\be
\tilde{A} \equiv - \left(\frac{5}{384} \right)^{1/2} \pi^{-2/3} \frac{{\cal{M}}^{5/6}}{D_{L}} \left(Q_{C} + i Q_{S}\right)\,,
\ee
and the Fourier phase must be computed from~\cite{cutlerflanagan,Yunes:2009yz}
\be
\label{eq:SPA-phase-def}
\Psi(f) = 2 \pi \int^{f/2} dF' \left(2 - \frac{f}{F'}\right) \frac{F'}{\dot{F}(F')}\,.
\ee
This expression makes it clear that the dominant ST modifications to the Fourier response are due to modifications to the rate of change of the orbital frequency.
\subsection{SPA Templates in Scalar-Tensor Theory \label{sec:SPAtemp}}
We now follow the analysis of the previous subsection and provide explicit formulas for the Fourier transform in the SPA of restricted PN waveforms in ST theories to 1PN order. We first focus on the binding energy and its rate of change. The former is presented in Eq.~$(6.4)$ of Ref.~\cite{Mirshekari:2013vb} in terms of the individual masses and velocities of the system. One can transform the binding energy to relative coordinates, through the mapping in Eqs.~$(6.9)$--$(6.11)$ of the same paper. The rate of change of the orbital binding energy is equal to the energy flux carried by all propagating degrees of freedom by the balance law. This quantity is presented in Eqs.~$(6.16)$ and $(6.17)$ of Ref.~\cite{Mirshekari:2013vb} in relative coordinates.
The ST theory-modified version of Kepler's third law of motion can be computed by solving $|a^{i}| = r_{12} \omega^2$ in a PN expansion for $r_{12}$, the relative orbital separation, using also that $|v_{12}^{i}| = r_{12} \omega$, where recall that we are considering only binaries in quasi-circular orbits. The relative acceleration $a^{i}$ is given in Eq.~$(1.4)$ and $(1.5)$ of Ref.~\cite{Mirshekari:2013vb}, where $\omega$ is the orbital angular velocity. We find that
\begin{align}
r_{12} &= \frac{m}{x} \left\{\left(1 + \frac{1}{3} \alpha_{1} \alpha_{2}\right) + x \left[\left(\frac{\eta}{3} - 1\right)
\right. \right.
\nonumber \\
&\left. \left.
+ \frac{1}{3} \alpha_{1} \alpha_{2} \left(\eta - 1\right)\right] + {\cal{O}}(x^{2},\alpha_{A}^{4}) \right\} \,,
\end{align}
where we have set $\psi_{0} = 0$, as this quantity is constrained to be less than $10^{-2}$ by current observations.
With the above modified version of Kepler's third law, one can rewrite the binding energy and its rate of change as a function of the PN expansion parameter. Using the definition of Eq.~\eqref{eq:Fdot}, the rate of change of the orbital frequency is then
\begin{align}
\dot{F} &= \frac{1}{2} \frac{\eta}{\pi m^{2}} x^{9/2} \left(\alpha_{1} - \alpha_{2}\right)^{2}
\nonumber \\
&+ \frac{48}{5} \frac{\eta}{\pi m^{2}} x^{11/2} \left\{1 + \frac{1}{576} \left[ \left(-15 - 35 \eta\right) \alpha_{1}^{2}
\right. \right.
\nonumber \\
&\left. \left.
+ \frac{1}{576} \left(35 \eta + 63 \right) \alpha_{1} \alpha_{2} +
\alpha_{1} \alpha_{1}' \left( \frac{5}{144} \nu - \frac{5}{48} \right)
\right. \right.
\nonumber \\
&\left. \left.
- \alpha_{1} \alpha_{2}' \left( \frac{5}{144} \nu + \frac{43}{48}\right)
+ 1 \to 2 \right] \right\} + {\cal{O}}(x^{6},\alpha_{A}^{4})\,,
\end{align}
where $\alpha_{A}' := d\alpha_{A}/d\ln{v}$.
The first term in this expression corresponds to dipolar radiation, while the second one is the usual quadrupolar radiation term of GR. Notice that both are corrected by terms proportional to the scalar charges, due to the modification to Kepler's third law.
With all of this at hand, we can now compute the Fourier phase in the SPA through Eq.~\eqref{eq:SPA-phase-def}. In Sec.~\ref{subsec:alphas}, we fitted $\alpha_A$ through Eq.~\eqref{eq:alpha-fit-poly}, which can be rewritten as
\be
\alpha_A = c_A + d_A v + \ldots,
\ee
identifying $a_{0}^{(A)}$ with $c_{A}$ and $a_{1}^{(A)}$ with $d_{A}$, which are themselves functions of $m_A$, $m$, and $\beta_{\mbox{\tiny ST}}$. Finally, we find
\begin{align}
\Psi(f) &= 2 \pi f t_{c} - \phi_{c} - \frac{\pi}{4} + \frac{3}{128 \eta v^{5}}
\left[- \frac{5}{168} \left(c_{1} - c_{2}\right)^{2} v^{-2}
\right.
\nonumber \\
&- \left. \frac{8400}{109489} (c_1 - c_2)(d_1 - d_2)v^{-1} + \ldots \right]\, .
\label{eq:SPAphase}
\end{align}
We see that the main modifications to the Fourier response in the SPA is due to the scalar charges, which induce a dipole correction to leading order. We have checked that the dipole term is exactly what one obtains in FJBD theory, when one rewrites $c_{1,2}$ in terms of the sensitivities $s_{1,2}$.
The ST-modified Fourier response described above is only valid for systems of \emph{unequal} mass. In the case of equal-mass binaries, the $-1$PN and the $-0.5$PN terms in Eq.~\eqref{eq:SPAphase} vanish, because then $s_{1} = s_{2}$. In such a system, the lowest PN order correction to the Fourier phase enters at $0$PN order. Henceforth, we only use the ST-modified SPA for analysis of binaries whose component masses are unequal.
\subsection{Comparison of SPA Phase to Numerical Phase \label{sec:comp}}
We can now validate the SPA model of the previous subsection by comparing it to the Fourier transform of the time-domain numerical solutions for the GWs presented in Ref.~\cite{Palenzuela:2013hsa}. Before we do so, we define the measure we will use for such a validation: the dephasing $\Delta \Psi$. This quantity is defined as
\be
\label{eq:dephasing}
\Delta \Psi_i = {\rm{min}}_{t_{c},\phi_{c}} \left| \Psi_{{\mbox{\tiny GR}}} - \Psi_{{\mbox{\tiny ST}}}\right|\,,
\ee
where $\Psi_{{\mbox{\tiny GR}}}$ is the Fourier phase in GR, $\Psi_{{\mbox{\tiny ST}}}$ is the
Fourier phase in ST theories, and the subindex $i$ labels four different strategies we employ to evaluate $\Psi_{{\mbox{\tiny GR}}}$ and $\Psi_{{\mbox{\tiny ST}}}$ as described below.
This measure requires minimization over time and phase of
coalescence, i.e.~over a constant time and phase shift. Such a
minimization is required to compare different template families,
e.g.~a time-domain waveform to a frequency-domain one.
In order to validate the SPA model of the previous subsection, we will evaluate the dephasing of Eq.~\eqref{eq:dephasing} in four different ways:
\begin{enumerate}
\item {\bf{Numerical}}: $\Psi_{{\mbox{\tiny GR}}}$ and $\Psi_{{\mbox{\tiny ST}}}$ are given by the Fourier transforms of the time-domain GW data of Ref.~\cite{Palenzuela:2013hsa}, minimizing the difference over phase and time of coalescence.
\item {\bf{-1 PN}}: $\Psi_{{\mbox{\tiny ST}}}$ is given by Eq.~\eqref{eq:SPAphase}, keeping only the leading PN term (the $-1$PN term).
\item {\bf{-0.5 PN}}: $\Psi_{{\mbox{\tiny ST}}}$ is given by Eq.~\eqref{eq:SPAphase}, keeping the leading and first subleading PN terms (the $-1$ and the $-0.5$PN terms).
\item {\bf{0 PN}}: $\Psi_{{\mbox{\tiny ST}}}$ is given by Eq.~\eqref{eq:SPAphase}, keeping terms up to Newtonian order (the $-1$, $-0.5$ and $0$PN terms).
\end{enumerate}
In all $N$PN cases, $\Psi_{{\mbox{\tiny GR}}}$ is the GR Fourier phase in the SPA, for example given in Ref.~\cite{Blanchet:2001ax,PPE}. By comparing the numerical dephasing to the $N$PN ones, we will validate the SPA templates constructed in the previous subsection.
Such a comparison is carried out in Fig.~\ref{fig:dephasing}. The solid line is
the numerical dephasing, while the dotted lines are the $N$PN dephasings in the
SPA. Observe that the SPA dephasing at $-0.5$PN order is a better approximation than keeping the dephasing to $0$PN
or $-1$PN order, relative to the numerical dephasing.
We will therefore truncate the fit to the scalar charges at $-0.5$PN
order for the rest of this paper (i.e, we keep only two ST PN corrections) and model the ST SPA templates by Eq.~\eqref{eq:SPAphase},
without ST higher-order PN terms.
\begin{figure}[ht]
\includegraphics[clip=true,width=0.45\textwidth]{SPA.eps}
\caption{\label{fig:dephasing} (Color Online) Dephasing between a GR and ST signal as a function of GW frequency, calculated using both the numerical Fourier phases (solid) and the SPA Fourier phases (dotted curves) for a binary with masses $(1.4074,1.7415) M_{\odot}$ and $\beta_{{\mbox{\tiny ST}}} = -4.5$. The ST SPA phases were calculated to various PN orders. }
\end{figure}
\subsection{ppE Waveforms}
Before proceeding with a detailed data analysis study, we map the SPA templates of Sec.~\ref{sec:SPAtemp} to the ppE templates constructed in Ref.~\cite{PPE}. The ppE family is a theory-independent set of templates, designed to capture model-independent deviations from GR with GW observations. This family is constructed by introducing parameterized modifications to both the binding energy and the energy balance equations of GR. Both types of modifications lead to changes in the GW phase, which can be described by introducing a small number of new parameters into the GW template.
The full GW waveform for the coalescence of two compact objects is typically split into three sections: inspiral, merger, and ringdown. ppE templates have been developed for all three phases, but in this paper we are interested only in the inspiral portion. The latter can be defined as the part of the waveform that is generated before the two bodies plunge into each other. The definition of the end of inspiral is somewhat arbitrary, but we follow typical conventions and define the transition from inspiral to merger as occurring at the innermost stable circular orbit of the system in center of mass coordinates (or alternatively at contact). In any case, the merger of NS binaries occurs at the very high frequency end of aLIGO's sensitivity band, almost outside of it altogether.
The simplest, quadrupole ppE inspiral templates have the form
\be
\tilde{h} (f) = \tilde{h}^{{\mbox{\tiny GR}}}\cdot (1+\alpha_{{\mbox{\tiny ppE}}} u^{a}) e^{i\beta_{{\mbox{\tiny ppE}}} u^{b}}, \quad\quad u = (\pi \mathcal{M} f)^{1/3},
\label{eq:ppEtemp}
\ee
where $\tilde{h}^{{\mbox{\tiny GR}}}$ is the Fourier response in GR. These simple ppE waveforms contain an additional amplitude term, $\alpha_{{\mbox{\tiny ppE}}} u^{a}$, and an additional phase term, $\beta_{{\mbox{\tiny ppE}}} u^{b}$, relative to GR. We refer to $\alpha_{{\mbox{\tiny ppE}}} $ and $\beta_{{\mbox{\tiny ppE}}} $ as the \emph{strength} parameters of the ppE deviations, and to $a$ and $b$ as the \emph{exponent} parameters.
These ppE waveforms cover all known inspiral waveforms from specific alternative theories of gravity \cite{cornish-PPE} that are analytic in the frequency evolution of the GWs. Some specific examples are discussed in Ref.~\cite{Yunes:2013dva}. They can not, however, perfectly match the signals generated by the theories of interest in this paper. For the case of induced/spontaneous scalarization, Sec.~\ref{sec:comp} showed that one needs \emph{two} ST PN corrections to the SPA phase (i.e. a $-1$PN and a $-0.5$PN term), while the simplest ppE model includes only one. For the dynamically scalarized case, the corrections to the Fourier response are very abrupt (almost a step function), which cannot be captured with a single PN term. In spite of this, we will see that the ppE model is quite adequate at detecting a ST deviation, provided the deviation is strong enough to be detectable in the first case.
\section{Detectability of Scalar-Tensor Deviations Through an Effective Cycles Approach \label{sec:detect}}
In this section, we carry out the first of a two-part data analysis investigation to determine the detectability of ST deviations in GWs emitted during the late inspiral of NS binaries. We first construct a new, computationally inexpensive data analysis measure to determine when a GR deviation is sufficiently loud for detection with aLIGO-type detectors. We then use this measure on ST signals and ppE signals to estimate their detectability.
\subsection{Useful and Effective Cycles of Phase \label{sec:usecyc}}
Model hypothesis testing, i.e.~the determination of whether model A or B is better supported by some data, usually requires a detailed Bayesian analysis through MCMC techniques that map the likelihood surface and the posterior distributions of template parameters in order to calculate BFs\footnote{The BF, assuming equal priors for the two competing theories, is the odds that one theory is favored by the data over another theory. For instance, a BF of 100 in favor of GR means that there is a 100:1 ``betting odds'' that GR is the correct theory given the data. In this paper, we are considering only nested models. For example, GR is recovered from ppE templates when the strength parameter $\beta_{\mbox{\tiny ppE}} = 0$. In this case, the BF can be calculated from the Savage-Dickey density ratio, which compares the prior weight at $\beta_{\mbox{\tiny ppE}} =0$ to the posterior weight at that value. The BF is then calculated via $BF = p(\beta_{\mbox{\tiny ppE}} = 0 | d)/p(\beta_{\mbox{\tiny ppE}} = 0)$. If there is more posterior weight at this point than prior weight,
the model selection process favors GR.}. Such studies are computationally expensive, and it is therefore desirable to construct a simple and computationally inexpensive measure for accomplishing similar goals. The construction of this measure is the topic of this section.
We first describe a quantity that has been used in the literature as a stand-in for the importance of a particular GW phase term~\cite{Damour:2000gg}: the \emph{useful cycles}, $\mathcal{N}_u$. This quantity is defined in \cite{Damour:2000gg} as
\begin{multline}
\mathcal{N}_u = \left(\int_{F_{\mbox{\tiny min}}} ^{F_{\tiny \mbox{max}}} d\ln f \frac{a^2(f)}{S_n(f)} \frac{d\phi}{2\pi df}\right) \\\times \left( \int_{F_{\mbox{\tiny min}}} ^{F_{\tiny \mbox{max}}} d\ln f \frac{a^2(f)}{f S_n(f)}\right)^{-1},
\label{Eq:usecyc}
\end{multline}
which is essentially a noise-weighted measure of the total number of cycles of phase due to any particular term in the phase evolution. In Eq.~\eqref{Eq:usecyc}, $a(f)$ is defined by $|\tilde{h}(f)|^2 = \tilde{A}^2(f) =N(f) a^2(f)/f^2$,
with $N(f) = (1/2\pi)(d\phi/d\ln F)= F^2/(dF/dt)$. The expression for $\mathcal{N}_u$ can be re-expressed in terms of the characteristic strain $h_c(f) = \sqrt{f} \tilde{A}(f)$ as
\be
\mathcal{N}_u = {\rm SNR}^2 \left( \int_{F_{\mbox{\tiny min}}} ^{F_{\tiny \mbox{max}}} \frac{h_c^2(f)}{S_n(f)} \frac{1}{N(f)}\; d\ln f \right)^{-1}.
\label{Eq:usecyc2}
\ee
Thus, the number of useful cycles is equal to the harmonic mean of $N(f)$, with a weighting factor equal to the SNR squared per logarithmic frequency interval, $\Delta {\rm SNR}^2(f)=h_c^2(f)/S_n(f)$.
The difference in the number of useful cycles between waveform models is sometimes used as a proxy for the detectability of the difference in the models (see eg. Ref.~\cite{Baiotti:2011am,Yagi:2013sva,Gerosa:2014kta}). One must be very careful when doing this for two reasons. The first is made clear by re-writing Eq.~\eqref{Eq:usecyc} in the form of Eq.~\eqref{Eq:usecyc2}. This re-casting of the useful cycles shows that it is not permissible to simply replace $N(f)$ with $\Delta N(f)$, where $\Delta N(f)$ is the change that is introduced by a particular modification to the phase. In order to calculate $\mathcal{N}_u$ due to a change in the phase evolution, it is necessary to calculate both $\mathcal{N}_u$ from the original phase and from the changed phase, and then take the difference. This is not a problem, \emph{per se} - it is simply an issue that must be kept in mind when calculating $\mathcal{N}_u$ for a particular phase term.
A larger issue with $\mathcal{N}_u$ as a measure of detectability is the murkiness of its connection with quantities such as the BF, which are directly related to model selection. The logarithm of the BF, as derived in Ref.~\cite{cornish-PPE}, satisfies
\be
\log {\rm BF} \sim \frac{1}{2}(1-{\rm FF}^2) {\rm SNR^2} + {\cal{O}}[(1 - {\rm{FF}}^{2})^{2}]\,.
\ee
We can use the following expression for the fitting factor, ${\rm FF}$, given two waveforms with the same amplitude $\tilde{A}(f)$, but with phases that differ by $\Delta \Psi(f)$:
\be
{\rm FF} ={\rm SNR^{-2}} \, \max_{\lambda^{a}} \left(\int \frac{h_c^2(f) \cos(\Delta\Psi(f))}{S_n(f)} \; d\ln f\right)\,,
\ee
where the maximization is done over all system parameters $\lambda^{a}$. In the limit ${\rm FF} \sim 1$, i.e.~for small deviations from GR, these expressions can be combined to give
\begin{align}
\log {\rm BF} &\sim \frac{1}{2} \min_{\lambda^{a}} \int \frac{h_c^2(f) \Delta\Psi^2(f)}{S_n(f)}
d\ln f + {\cal{O}}(\Delta \Psi^{4})\;.
\label{Eq:bf}
\end{align}
Given the above expression for the BF, a natural definition for a computationally inexpensive data analysis measure presents itself, the \emph{effective cycles of phase}:
\be
\mathcal{N}_e = \min_{\Delta t, \Delta \phi} \left[\frac {1}{2\pi{\rm SNR}}
\left( \int \frac{h_c^2(f) \Delta \Phi^2(f)}{S_n(f)} \; d\ln f
\right)^{1/2}\right]\,,
\label{Eq:EffCyc}
\ee
where $\Delta \Phi \equiv \Delta \Psi(f) + 2 \pi f \Delta t - \Delta \phi$,
where $\Delta t$ and $\Delta \phi$ are an arbitrary time
and phase shift respectively. As before, the dephasing is
$\Delta \Psi = \Psi_{1}(f) - \Psi_{2}(f)$, where in our case $\Psi_{1}$ will be
the Fourier phase of a GR signal
and $\Psi_{2}$ the phase of a non-GR signal, e.g.~for a ppE waveform, $\Delta
\Psi(f) = \beta_{\mbox{\tiny ppE}} u^b$.
We define ${\cal{N}}_{e}$ with a $(\Delta t, \Delta \phi)$-minimization
because the non-GR terms induce a modification in the frequency and phase
evolutions that renders meaningless a direct comparison between time-shift and
phase-shift parameters in waveforms living in different theories. Notice, however,
that we do not minimize over \emph{all} parameters, as would be required to relate
${\cal{N}}_{e}$ to the BF. A full minimization procedure is costly and it would involve
an MCMC analysis in general, while the minimization with respect to only $(\Delta t, \Delta \phi)$ is inexpensive.
We can now see that this quantity is directly related to model selection through the BF:
\be
\log {\rm BF} \sim 2\pi^{2}{\rm{SNR} }^2 \min_{\lambda^{a}} {\cal{N}}_{e}^{2}\,.
\label{Eq:bf-with-N}
\ee
The effective cycles, $\mathcal{N}_e$, as defined in Eq.~\eqref{Eq:EffCyc}, i.e.~minimized over $(\Delta t, \Delta \phi)$
only, give an \emph{upper limit} to the BF. The fully minimized ${\cal{N}}_{e}$ will in general be smaller than Eq.~\eqref{Eq:EffCyc} due to covariances between system parameters. This means that $\mathcal{N}_e$ is not a perfect proxy for detectability. That is, if $\mathcal{N}_e$ due to a particular term in the phase is large, this may or may not mean that the term is detectable. However, if $\mathcal{N}_e$ due to a particular phase term is small, this \emph{does} indicate that the term will \emph{not} be detectable.
The above is an alternative means of deriving the quantity first derived in \cite{PhysRevD.78.124020}, which is there referred to as the distinguishability/measurability. Up to some rearranging of various factors, this quantity and the effective cycles of phase are the same. The connection between this quantity and the Bayes factor, though, is a new result.
How are the useful and effective cycles related? From the definitions of $\mathcal{N}_u$ and $\mathcal{N}_e$, it is clear that the former gives the difference in the harmonic mean of the number of cycles, while the latter gives the root-mean-square difference in the number of cycles. This is an important difference - $\mathcal{N}_e$ is directly related to the BF in the small deformation limit. For certain signals, however, the difference can be shown to be a GR-modification dependent constant factor. To see this, consider inspiral GWs in the PN approximation both in GR and in ppE form. The orbital phase $\phi(f) = 2 \pi f t(f) - \Psi(f) - \pi/4$ is a power series in $v = (\pi M f)^{1/3}$, just like the Fourier phase $\Psi(f)$ in both GR and ppE theory. The logarithmic derivative, $d\phi/d\ln f$, preserves the structure of such a power series, and so the PN series for $\Psi(f)$ and for $N(f)$ differ only by $b$-dependent, order unity factors in each of the coefficients. The relation between $\mathcal{N}_e$ and $\mathcal{N}_u$ is shown in Fig.~\ref{fig:NuvNe} and derived in Appendix~\ref{AppB}. Of course, the useful and effective cycles defined here can be computed given \emph{any} form of phase evolution, and thus, they are not restricted to phases in the PN approximation.
\begin{figure}[ht]
\includegraphics[clip=true,angle=-90,width=0.45\textwidth]{NusevNeff.eps}
\caption{\label{fig:NuvNe} (Color Online) The ratio of effective cycles to useful cycles, calculated as a function of $b$, for two fixed values of $\beta_{\mbox{\tiny ppE}}$.}
\end{figure}
\subsection{Useful and Effective Cycles as a Measure of Detectability of General non-GR Effects}
Now that we have introduced the concept of effective cycles, we will use it to determine when a particular GR deviation is detectable. In particular, we will determine the number of useful and effective cycles that are needed for ppE deviations to lead to a BF that favors the ppE model.
We inject ppE signals with varying strength $\beta_{{\mbox{\tiny ppE}}}$ parameters and exponent $b$ parameters and then
\begin{itemize}
\item[(i)] calculate $\mathcal{N}_u$ and $\mathcal{N}_e$ due to the ppE terms relative to a GR signal ($\beta_{{\mbox{\tiny ppE}}}=0$) and;
\item[(ii)] run an MCMC analysis to calculate the BF between a ppE model and the GR model.
\end{itemize}
The second item requires the choice of a prior range for each ppE strength parameter, which we choose to be $|\beta_{\mbox{\tiny ppE}}| \le 5\times10^{-5}$ for $b=-7$, $|\beta_{\mbox{\tiny ppE}}|\le 5\times 10^{-4}$ for $b=-6$, $|\beta_{\mbox{\tiny ppE}}| \le 5$ for $b=-5$ and $b=-4$. These prior ranges were derived by examining the results in Ref.~\cite{cornish-PPE} and requiring that the deviations be detectable given a GW signal with SNR $\approx 10$.
Every injection in this study consists of a $(1.4074,1.7415) M_\odot$, NS/NS, non-spinning binary, with zero inclination angle, and with SNR $\approx 15$, which corresponds to a luminosity distance $D_L \approx 50$ Mpc. We choose the zero-detuned, high-power spectral noise density projected for aLIGO~\cite{LIGOnoise}, stopping all integrations at $1000$Hz. This frequency is lower than the GW frequency at which the NSs touch each other, which is approximately between $1250$ and $2000$ Hz, depending on the NS EoS.
\begin{figure*}[ht]
\includegraphics[clip=true,angle=-90,width=0.475\textwidth]{BFvbetappE_panel1.eps}
\includegraphics[clip=true,angle=-90,width=0.475\textwidth]{BFvbetappE_panel2.eps}
\caption{\label{fig:BFNuse} (Color Online) BF in favor of a modification to GR versus the number of effective cycles (left) and useful (right) cycles induced by that modification for different types of ppE corrections. For the $b=-4$ cases, the line labeled (-) corresponds to a negative value for $\beta_{\mbox{\tiny ppE}}$, and the line labeled (+) corresponds to a positive value; all other injections had positive values for $\beta_{\mbox{\tiny ppE}}$. Observe that the $b$ dependence of the relationship between BF and $\mathcal{N}_u$ is the opposite of what one would expect.}
\end{figure*}
Figure~\ref{fig:BFNuse} shows the BF on a logarithmic scale versus the absolute value of the number of effective (left panel) and useful (right panel) cycles introduced by signals with different ppE exponent parameters, starting at $b=-7$ (a $-1$PN term) and going up to $b=-4$ (a $0.5$PN term). Recall that ST theories lead to modifications at $-1$PN order and higher and that a BF larger than $1$ indicates the data supports the non-GR model. Observe that detectability occurs when $\mathcal{N}_u$ is between $0.1$ and $1$ and when ${\cal{N}}_{e}$ is between $2$ and $4$ cycles of phase for most of the ppE injections. This is not true, however, for the $b=-5$ case, because of the almost perfect correlation between chirp mass $\mathcal{M}$ and $\beta_{\mbox{\tiny ppE}}$ when $b=-5$: a straight line in the two-dimensional ${\cal{M}}_{c}$--$\beta_{{\mbox{\tiny ppE}}}$ plan similar to Fig.~$6$ in~\cite{Sampson:2013lpa}. Because of this, a deviation from GR at the $0$PN (Newtonian) level would have to be very large to be detectable, as previously noted in Ref.~\cite{cornish-PPE}.
Observe also in Fig.~\ref{fig:BFNuse} the difference in detectability for phase terms that accumulate either positive or negative cycles of phase. As stated, Fig.~\ref{fig:BFNuse} shows the \emph{absolute value} of the useful or the effective cycles of phase introduced by each ppE term. For injections made with a positive $\beta_{\mbox{\tiny ppE}}$, the actual sign of the cycles of phase is negative. We used this type of injection for most of the lines in this figure, but for the $b = -4$ case, we injected both positive and negative values of $\beta_{\mbox{\tiny ppE}}$. We find that the positive values are detected more easily than the negative values.
This effect can be understood by examining the posterior distribution of
$\beta_{\mbox{\tiny ppE}}$ at $b=-4$, shown in Fig.~\ref{fig:47post} together with the
posterior of $\beta_{\mbox{\tiny ppE}}$ at $b=-7$. Observe that the posteriors are not
symmetric about $\beta_{\mbox{\tiny ppE}} = 0$. In the case of $b=-4$, injecting a positive
value of $\beta_{\mbox{\tiny ppE}}$ results in no posterior weight at $\beta_{\mbox{\tiny ppE}} = 0$ for
much smaller values of $|\beta_{\mbox{\tiny ppE}}|$ than injecting a negative value. The
asymmetry in the posterior distributions can again be understood by noting that
$\beta_{\mbox{\tiny ppE}}$ is correlated with the mass parameters, which are, of course,
forced to be positive.
\begin{figure*}[ht]
\includegraphics[clip=true,angle=-90,width=0.48\textwidth]{betappE47post_panel1.eps}
\includegraphics[clip=true,angle=-90,width=0.48\textwidth]{betappE47post_panel2.eps}
\caption{\label{fig:47post} (Color Online) Posterior distributions for $\beta_{\mbox{\tiny ppE}}$ for a ppE search template with $b=-4$ (left) and $b=-7$ (right), given a GR injection. Observe that these distributions are not symmetric about zero, which explains why changes to the GW phase of different signs are detectable at different levels.}
\end{figure*}
Figure~\ref{fig:BFNuse} also shows that the threshold in $\mathcal{N}_{e}$ for detectability (e.g.~the value of $\mathcal{N}_{e}$ at which the BF equals 10) is lower for terms that are of very high PN order. Comparing the left and right panels of this figure, observe that the $\mathcal{N}_{u}$ threshold exhibits the opposite behavior. The $\mathcal{N}_{e}$ threshold behavior is what one would expect because high PN order terms have small covariances with the ppE strength parameters and the system parameters. It is therefore very difficult for a GR waveform to match the phasing of these types of injections, while the opposite is true for low PN order ppE effects. This fact illustrates the advantage of using $\mathcal{N}_e$ instead of $\mathcal{N}_u$ as a measure of detectability: the $\log({\rm{BF}})$ as a function of $\mathcal{N}_e$ exhibits the expected behavior as $b$ changes, but $\log({\rm{BF}})$ as a function of $\mathcal{N}_u$ shows the opposite behavior.
So far, the discussion has assumed a fixed SNR of $15$, but clearly the detectability of a GR deviation depends on the SNR of the signal. To understand this, we performed the same analysis as above but with $b=-3$ fixed and signals of differing SNRs. The results are plotted in Fig.~\ref{fig:BFvSNRppE}, where observe that the number of effective cycles necessary for a detection of a modification to GR scales approximately as the SNR squared, as expected from Eq.~\eqref{Eq:bf-with-N}. That equation was derived assuming small deviations from GR, but here we see that regardless of the value of the ppE exponent, Eq.~\eqref{Eq:bf-with-N} is still approximately satisfied.
\begin{figure}[ht]
\includegraphics[clip=true,angle=-90,width=0.48\textwidth]{PhivlogBetaSNR_1panel.eps}
\caption{\label{fig:BFvSNRppE} (Color Online) The BF in favor of a deviation from GR as a function of effective cycles of phase scaled by their SNR for ppE injections with $b=-3$. Observe the almost linear relation between log(BF) and $\mathcal{N}_e^2 {\rm{SNR}}^{2}$, as predicted in Eq.~\eqref{Eq:bf-with-N}.}
\end{figure}
\subsection{Useful and Effective Cycles as a Measure of Detectability of ST Effects}
\label{sec:useful-effective-ST}
We now apply what we have learned about effective cycles to ST theories. First,
we must discuss exactly which signals we will analyze in detail. We employ data
from Ref.~\cite{Palenzuela:2013hsa}, which consists of 80
GW signals from NS binaries with constituent masses ranging from approximately $1.4$ to
$1.7M_\odot$, and with $\beta_{\mbox{\tiny ST}}$ ranging from $-4.5$ to $-3.0$. Within this
set, and for the polytropic EoS used
in Ref.~\cite{Barausse:2012da,Palenzuela:2013hsa}, some systems undergo spontaneous/induced scalarization,
while others dynamically scalarize. As already mentioned, we refer to signals
from systems whose components undergo spontaneous/induced scalarization as ``spontaneously scalarized signals.'' We do so because a necessary condition for induced scalarization to happen is the presence of at least
one spontaneously scalarized star. The term ``dynamically scalarized signals'' will
denote those from dynamically scalarized binaries.
The strength and type of scalarization that occurs depends sensitively on the constituent masses and the
value of $\beta_{{\mbox{\tiny ST}}}$.
We now wish to calculate the effective cycles $\mathcal{N}_e$ induced by the ST
corrections to investigate their detectability. For the spontaneously scalarized
cases, we can compute $\mathcal{N}_e$ with the SPA waveforms, using for $\Delta
\Psi(f)$ the terms in square brackets of Eq.~\eqref{eq:SPAphase}. The integrated SPA
dephasing is then minimized over $(\Delta t,\Delta \phi)$, as explained in Eq.~\eqref{Eq:EffCyc}.
For the dynamically scalarized cases, we compute $\mathcal{N}_e$ with the Fourier transform
of the numerical time-domain data of Ref.~\cite{Palenzuela:2013hsa},
using for $\Delta \Psi(f)$ the difference in the Fourier phases of a GR and a ST numerical signal.
We then again introduce parameters $(\Delta t,\Delta \phi)$ and minimize the integrated dephasing with
respect to them to define ${\cal{N}}_{e}$, as explained in Eq.~\eqref{Eq:EffCyc}.
\begin{figure*}[ht]
\includegraphics[clip=false,angle=-90,width=0.48\textwidth]{PhievbetaST_full.eps}
\includegraphics[clip=false,angle=-90,width=0.48\textwidth]{Nevfstar}
\caption{\label{fig:BFvbetaST} (Color Online) Effective number of cycles generated by a ST modification to GW signals as a function of $\beta_{\mbox{\tiny ST}}$ for all ST cases (left) and as a function of GW frequency of scalarization in the dynamical scalarization cases (right). Different curves correspond to systems with different masses, and the different points in the right panel correspond to different values of $\beta_{{\mbox{\tiny ST}}}$ increasing to the right. We model the GW phase through the SPA in the spontaneously scalarized cases, while we use the Fourier transform of numerical data in the dynamically scalarized cases. The shaded region corresponds roughly to the number of effective cycles necessary for BFs between $1$ and $10$, as estimated from Fig.~\ref{fig:BFNuse}. Observe that the number of effective cycles is below the detectability region for all but the $\beta_{\mbox{\tiny ST}} \lesssim -4.5$ cases. Note also that the only dynamically scalarized cases that are detectable are those in which the scalar field activates at small frequencies. For the equal-mass binaries that undergo spontaneous scalarization, the SPA is not applicable for calculating $\mathcal{N}_e$. We therefore used masses that were nearly, but not identically, equal for an approximate calculation.}
\end{figure*}
The left panel of Fig.~\ref{fig:BFvbetaST} shows the effective cycles as a function of $\beta_{{\mbox{\tiny ST}}}$ for both spontaneously and dynamically scalarized cases, with the latter labeled with an upside-down triangle. Different line styles correspond to systems with different masses. The shaded region corresponds to the region where one would expect BFs of between $1$ and $10$, given the results of Fig.~\ref{fig:BFNuse}. Modifications that lead to effective cycles above this shaded region may then be detectable with an aLIGO instrument.
Several features of this figure are worth discussing in more detail.
First, observe that almost all of the detectable cases correspond to spontaneously scalarized systems. For these systems, dipole radiation is the dominant GR modification, a $-1$PN order effect that is proportional to the \emph{difference} of scalar charges of the two bodies (see e.g.~Eq.~\eqref{eq:SPAphase}). For equal-mass binaries, this dipolar effect vanishes identically, and the dominant modification enters at Newtonian, $0$PN order. Such a GR modification, however, is strongly degenerate with the chirp mass, as shown in Fig.~\ref{fig:BFNuse}, and thus, it is difficult to detect. For this reason, spontaneously scalarized systems with larger mass differences lead to larger values of $\mathcal{N}_e$ and are easier to detect.
Another interesting feature of Fig.~\ref{fig:BFvbetaST} is that, within the cases analyzed, only a handful of dynamically scalarized systems seem detectable: the $(m_{1},m_{2}) = (1.6,1.6) M_{\odot}$ with $\beta_{{\mbox{\tiny ST}}} = -4.5$ system, the $(m_{1},m_{2}) = (1.7,1.7) M_{\odot}$ with $\beta_{{\mbox{\tiny ST}}} = -4.25$ system and the $(m_{1},m_{2}) = (1.5,1.6) M_{\odot}$ with $\beta_{{\mbox{\tiny ST}}} = -4.5$ systems. One of the key differences between these cases and all others is that they scalarize at relatively low GW frequency. Figure~\ref{fig:fstarvbeta} shows the approximate GW frequency at which the scalar field
activates in a dynamically scalarized binary, as a function of $\beta_{{\mbox{\tiny ST}}}$. Observe that as $|\beta_{{\mbox{\tiny ST}}}|$ becomes smaller, or as the total mass of the binary decreases, dynamical scalarization occurs at higher and higher frequencies. For these three cases the scalar field activates at roughly $80$, $120$, and $180$ Hz respectively, while in all other cases dynamical scalarization occurs at higher GW frequency.
\begin{figure}[ht]
\includegraphics[clip=false,angle=-90,width=0.48\textwidth]{fstarvbeta.eps}
\caption{\label{fig:fstarvbeta} (Color Online) Approximate GW frequency in Hz at which the scalar field activates in a dynamically scalarized binary. Observe that the higher $\beta_{{\mbox{\tiny ST}}}$ and the lower the total mass of the system, the higher the frequency of activation }
\end{figure}
The reason, then, that the three dynamically scalarized cases discussed above appear detectable is that detectability of a \emph{sudden} non-GR effect, i.e.~one that turns on rapidly, correlates strongly with the GW frequency at which this turn on occurs. The right panel of Fig.~\ref{fig:BFvbetaST} shows this correlation through $\mathcal{N}_{e}$ as a function of the GW frequency of dynamical scalarization. Notice that the lower the GW frequency of activation, the larger the number of effective cycles, and thus, the easier it would be to detect such a GR modification.
Reference~\cite{Sampson:2013jpa} first observed this phenomenon by studying ppE-type GR modifications that turn on suddenly. Their conclusion was that for aLIGO to detect such non-GR effects at SNRs of $12$, the modification had to turn on at a GW frequency lower than roughly 100 Hz (see Fig.~$4$ in~\cite{Sampson:2013jpa}). The ST modifications we study here are of a different PN order and at higher SNR than the non-GR ppE signals considered in Ref.~\cite{Sampson:2013jpa}, but we still see similar behavior: dynamical scalarization is only detectable when it activates below $\approx 200$ Hz at SNR 15.
Summing up, if one detected a GW signal that is consistent with GR (i.e.~lacking any dynamical, induced or
spontaneous scalarization effects), one should be able to constrain $\beta_{{\mbox{\tiny ST}}} \lesssim -4.25$,
given the data that we studied in this paper.
This statement, of course, is EoS dependent, and thus,
strictly applicable only to NSs with the polytropic EoS used here. In principle, NS binaries
with a different EoS could scalarize
at a different GW frequency. With some variability depending on the EoS, though, it remains true that one must have $\beta_{\mbox{\tiny ST}} \lesssim -4.25$ in order for scalarization
(either spontaneous/induced or dynamical) to occur at all.
\begin{table*}[!]
\begin{tabular}{c||c|c|l}
\hline
Case & Masses (in $M_\odot$) & Theory & Type of Scalarization \& Radiation \\
\hline\hline
a & $(1.4074,1.4074)$ & $\tilde{\beta}=(-4.5, -4.2$)
& Dynamical Scalarization \\
& & & Quadrupolar, No Dipolar Radiation
\\
\hline
b & $(1.4074,1.5145)$ & $\tilde{\beta}=(-4.5, -4.2$)
& Dynamical Scalarization \\
& & & Quadrupolar \& Dipolar Radiation
\\
\hline
c & $(1.4074,1.6441)$ & $\tilde{\beta}=(-4.5, -4.2$)
& Dynamical/Induced Scalarization \\
& & & Quadrupolar \& Dipolar Radiation
\\
\hline
d & $(1.4074,1.7415)$ & $\tilde{\beta}=(-4.5, -4.2$)
& Dynamical/Induced Sc. for the lower-mass star \\
& & & Spontaneous Scalarization in the higher-mass star \\
& & & Quadrupolar \& Dipolar Radiation
\\
\hline
e & $(1.5145,1.5145)$ & $\tilde{\beta}=(-4.5, -4.2$)
& Dynamical Scalarization \\
& & & Quadrupolar, No Dipolar Radiation
\\
\hline
f & $(1.5145,1.6441)$ & $\tilde{\beta}=(-4.5, -4.2$)
& Dynamical Scalarization \\
& & & Quadrupolar \& Dipolar Radiation
\\
\hline
g & $(1.5145,1.7415)$ & $\tilde{\beta}=(-4.5, -4.2$)
& Dynamical/Induced Sc. for the lower-mass star \\
& & & Spontaneous Scalarization in the higher-mass star \\
& & & Quadrupolar \& Dipolar Radiation
\\
\hline
h & $(1.6441,1.6441)$ & $\tilde{\beta}=(-4.5, -4.2$)
& Dynamical/Induced Scalarization \\
& & & Quadrupolar, No Dipolar Radiation
\\
\hline
i & $(1.6441,1.7415)$ & $\tilde{\beta}=(-4.5, -4.2$)
& Spontaneous Scalarization for $\tilde{\beta} = -4.5$ \\
& & & Dynamical/Induced Scalarization for $\tilde{\beta} = -4.2$ \\
& & & Quadrupolar \& Dipolar Radiation
\\
\hline
j & $(1.7415,1.7415)$ & $\tilde{\beta}=(-4.5, -4.2$)
& Spontaneous Scalarization for $\tilde{\beta} = -4.5$ \\
& & & Dynamical/Induced Scalarization for $\tilde{\beta} = -4.2$ \\
& & & Quadrupolar, No Dipolar Radiation
\\
\hline\hline
\end{tabular}
\caption{System parameters for the $\beta_{\mbox{\tiny ST}} = -4.5$ and $\beta_{\mbox{\tiny ST}} = -4.25$ cases discussed in this section.}
\end{table*}
\section{Detectability of Scalar-Tensor Deviations: Bayesian Model Selection}
\label{sec:BF}
In this section, we carry out the second part of our data analysis investigation to determine the detectability of ST deviations in GWs emitted during the inspiral of NS binaries. We perform a full Bayesian analysis study, separating the spontaneously scalarized cases from the dynamically scalarized ones. Such a study will allow us to confirm the expectations derived using effective cycles in the previous section.
To test these expectations, we inject ST GW signals at an SNR $\approx 15$ produced by
\begin{itemize}
\item[(a)] {\bf{Spontaneously scalarized NS binaries}}, with spontaneous/induced scalarization occurring \emph{before} GWs enter the detector's sensitivity band;
\item[(b)] {\bf{Dynamically scalarized NS binaries}}, with dynamical scalarization occurring \emph{during} the inspiral, at GW frequencies in the detector's sensitivity band.
\end{itemize}
As a case study for spontaneously scalarized injections, we consider a $(1.4074,1.7415) M_{\odot}$ binary with $\beta_{{\mbox{\tiny ST}}} \ge -4.25$. We expect these signals to lead to the largest spontaneous scalarization effects, as one can see in the left panel of Fig.~\ref{fig:BFvbetaST} at $\beta_{{\mbox{\tiny ST}}} = -4.5$. However, for $\beta_{{\mbox{\tiny ST}}} > -4.25$, we do not expect these effects to be detectable, since they lead to a very small number of effective cycles. Note that for $\beta_{{\mbox{\tiny ST}}} \to -4.25$, this system is a case that exhibits dynamical scalarization.
For dynamically scalarized injections, we consider several different systems. First, we study a $(1.4074,1.7415) M_{\odot}$ binary at $\beta_{{\mbox{\tiny ST}}} = -4.25$, since this is the limiting case of the spontaneously scalarized sequence discussed in the previous paragraph. We then study a $(1.6441,1.6441) M_{\odot}$ binary at $\beta_{{\mbox{\tiny ST}}} = -4.5$, a $(1.7415,1.7415) M_{\odot}$ binary at $\beta_{{\mbox{\tiny ST}}} = -4.25$, and a $(1.5145,1.6441) M_{\odot}$ binary at $\beta_{{\mbox{\tiny ST}}} = -4.5$, as these are the dynamically scalarized signals that look the most detectable, given the left panel of Fig.~\ref{fig:BFvbetaST}.
Spontaneously and dynamically scalarized injections are modeled differently. For the former, we use the SPA scheme of Eq.~\eqref{eq:SPAphase}. For the latter, we first compute the discrete Fourier transform of the numerical data of Ref.~\cite{Palenzuela:2013hsa}. We then take the difference of the Fourier phase between a numerical ST signal and a GR signal, and finally add this phase difference to a GR SPA signal.
We recover these injections with four different types of template families:
\begin{enumerate}
\item[(i)] {\bf{Simple ppE templates}}, constructed with a single $\beta_{{\mbox{\tiny ppE}}}$ parameter and ppE exponent $b=-7$;
\item[(ii)] {\bf{ST SPA templates}}, constructed from the results presented in Sec.~\ref{sec:SPAtemp};
\item[(iii)] {\bf{2-parameter ppE templates}}, an augmented ppE template family that uses two ppE terms, one with exponent $b=-7$ and one with $b=-6$~\cite{Sampson:2013lpa};
\item[(iv)] {\bf{ppE$_\theta$ templates}}, another augmented ppE template family with a single ppE exponent $b = -7$ but with $\beta_{{\mbox{\tiny ST}}} \to \Theta(f-f^*) \beta_{\mbox{\tiny ppE}}$, where $\Theta(\cdot)$ is a step-function and the threshold frequency, $f^*$, is a new ppE parameter~\cite{Sampson:2013lpa}.
\end{enumerate}
The simple ppE template family fixes the ppE exponent to $b=-7$, as this corresponds to the leading-order ST correction to the SPA phase for unequal mass systems where dipolar radiation is present. The ST SPA templates, of course, are the same templates as the model used for the spontaneously scalarized injections, and thus, by construction, we expect these to be the best templates for extracting ST modifications of this type. The 2-parameter ppE template family is also able to achieve a perfect match with spontaneously scalarized injected signals, but it includes two free parameters, rather than one. The ppE$_{\theta}$ template family allows the non-GR terms in the phase to ``turn-on'' at a particular threshold frequency, which is well-suited to dynamically scalarized injections.
A salient feature of the semi-analytical results of Ref.~\cite{Palenzuela:2013hsa} (and of the full general-relativistic simulations of Ref.~\cite{Barausse:2012da}) is the binary's early plunge due to the activation of dynamical scalarization. Such a feature is present in the dynamically scalarized injections we consider, but given the limited number of data sets, we cannot study its detectability in sufficient detail. In order to study whether such rapid termination of the inspiral is detectable, we will consider an additional type of injection and template:
\begin{enumerate}
\item[(v)] {\bf{Heaviside signal}} of the form $\tilde{h}_{{\mbox{\tiny GR}}}(f) \Theta(f^{*} -
f)$\,,
\end{enumerate}
where we will vary $f^{*}$ within $(40,10^{3})$ Hz. Given such injections and templates, we then study the range of $f^{*}$ that leads to early terminations that can be detected as a non-GR effect. We will explain in Sec.~\ref{dyn-sca-subsec} why we choose to work with such a toy-model.
All the template models considered above are nested. This means that for a
certain choice of non-GR parameters, the templates reduce exactly to GR. For
example, when $\beta_{{\mbox{\tiny ST}}} = 0$, the ST SPA templates reduce exactly to the GR
SPA templates. When this is the case, one can use the Savage-Dickey density
ratio to calculate the BF~\cite{dickey1971weighted}. Finding this ratio requires
the calculation of the posterior at the nested value of the non-GR parameter.
The posterior is calculated with MCMC techniques well-developed in previous
studies~\cite{TysonThesis,cornish-PPE,Sampson:2013jpa,Sampson:2013lpa}. We again
use the zero-detuned, high-power aLIGO noise curve, assuming a single detector
and truncating all integrals at $f = 1000$ Hz.
\subsection{Spontaneously Scalarized Signals}
We first consider spontaneously scalarized signals and compute the BFs between GR and the first two types of templates described above [(i) and (ii)]. Figure~\ref{fig:BFSPA} shows the BFs as a function of the injected $\beta_{\mbox{\tiny ST}}$, keeping $(m_{1},m_{2})=(1.4074,1.7415) M_{\odot}$ fixed and varying $\beta_{{\mbox{\tiny ST}}}$ with the constraint $\beta_{{\mbox{\tiny ST}}} \ge -4.25$, so as to consider only spontaneously scalarized signals. Because there are no BFs larger than one, this figure shows that spontaneously scalarized ST deviations from GR are not detectable for any of the cases shown. This is consistent with our expectations from the previous subsection. The BFs for the 2-parameter ppE templates are lower than those shown in Fig.~\ref{fig:BFSPA} for the simple ppE templates because of the Occam penalty for more complicated models, which we will discuss later on in this subsection.
\begin{figure}[ht]
\includegraphics[clip=true,angle=-90,width=0.45\textwidth]{BFSPAinj.eps}
\caption{\label{fig:BFSPA} (Color Online) The BF in favor of a non-GR signal, calculated by injecting ST signals with SNR of 15, and recovering using both simple ppE templates, and 2-parameter ppE templates. A BF above 1 indicates the data prefers the non-GR model. As expected, neither template is able to detect the non-GR deviations.}
\end{figure}
The BF for the spontaneously scalarized, $\beta_{\mbox{\tiny ST}} = -4.5$ case is not included in Fig.~\ref{fig:BFSPA} because there is so little posterior weight at $\beta_{\mbox{\tiny ppE}} = 0$ that a calculation of the BF using the Savage-Dickey density ratio is poorly defined. That is, the Savage-Dickey density essentially diverges due to poor exploration of the $\beta_{{\mbox{\tiny ppE}}} = 0$ region. Figure~\ref{fig:beta45} shows the posterior distribution for $\beta_{\mbox{\tiny ppE}}$, which illustrates this point and, as expected, indicates that the spontaneously scalarized binary with $\beta_{\mbox{\tiny ST}} = -4.5$ is easily detectable, for the polytropic EoS models considered here.
\begin{figure*}[ht]
\includegraphics[clip=true,angle=-90,width=0.45\textwidth]{betaST-45post_panel1.eps}
\includegraphics[clip=true,angle=-90,width=0.45\textwidth]{betaST-45post_panel2.eps}
\caption{\label{fig:beta45} (Color Online) Red (solid) lines: the posterior distributions for $\beta_{\mbox{\tiny ppE}}$ (left panel) and $\beta_{\mbox{\tiny ST}}$ (right panel). Blue (dashed) lines: the prior density for each of these parameters. There is essentially zero posterior weight at $\beta_{{\mbox{\tiny ST}}/{\mbox{\tiny ppE}}} =0$, indicating a large preference for the non-GR model from both template families.}
\end{figure*}
One may worry that our inability to detect spontaneously scalarized binaries when $\beta \geq -4.25$ is somehow a consequence of using ppE templates. To prove that this is not the case, we next explore the extraction of such spontaneously scalarized signals using both ppE$_\theta$ and custom-made, SPA templates. The ppE$_\theta$ templates should have more freedom to fit these signals, and the SPA templates can fit them perfectly. Because of the nature of these waveforms and the weakness of the GR deviation in the injections, the non-GR parameters in both cases are essentially unconstrained within their prior ranges. For the SPA templates, the non-GR parameter is $\beta_{\mbox{\tiny ST}}$, and its prior range is $|\beta_{\mbox{\tiny ST}}| < 5$. This means that the BF is approximately equal to one in both cases, independent of the injected parameters. Figure~\ref{fig:betathetapost} shows the prior and posterior distributions generated using the ppE$_\theta$ and the SPA templates, for a spontaneously scalarized binary with $\beta_{\mbox{\tiny ST}} = -3.5$. These distributions are identical for other injected values of $\beta_{\mbox{\tiny ST}}$, barring $\beta_{\mbox{\tiny ST}} \leq -4.5$, which is again easily extractable as a non-GR signal with either type of template.
\begin{figure*}[ht]
\includegraphics[clip=true,angle=-90,width=0.45\textwidth]{SPApostbetaposttheta_panel1.eps}
\includegraphics[clip=true,angle=-90,width=0.45\textwidth]{SPApostbetaposttheta_panel2.eps}
\caption{\label{fig:betathetapost} (Color Online) Posterior distributions for $\beta_{\mbox{\tiny ST}}$ (right) and for $\beta_{\mbox{\tiny ppE}}$ (left), recovered from an SNR 15 injection with $\beta_{\mbox{\tiny ST}} = -3.5$. Also plotted in blue (dashed) is the prior density for both parameters. The parameters are essentially unconstrained within their prior range, leading to a BF $\approx 1$ for both models, indicating no preference between this model and GR.}
\end{figure*}
We expect, though, that custom-made templates should perform better than ppE templates at extracting non-GR modifications when the custom-made templates match the signal. The extent to which this is true depends on the strength of the non-GR modification and on the loudness of the signal. Let us first consider very strong ST modifications, which we have already shown to be detectable with simple ppE templates, i.e.~a $\beta_{\mbox{\tiny ST}} = -4.5$, spontaneously scalarized ST signal, with masses of $(1.4074,1.7415) M_\odot$ and SNRs of 12, 10, 8, and 6. When the SNR is 12, 10, or 8, there is \emph{no} difference in the ability of simple ppE or SPA templates to discern the presence of a ST effect. When the SNR drops below 8, the signal is not detectable in the first place, using either type of template. Thus, custom-made templates and model-independent templates are equally good at detecting this type of GR deviations.
But what about GR deviations that are weaker, and thus, more difficult to
detect? Surely, in this case one expects custom-made templates to be more
effective at discerning such deviations. To explore this question, we inject a
$\beta_{{\mbox{\tiny ST}}}=-3.5$, spontaneously scalarized ST signal, with masses
$(1.4074,1.7415) M_\odot$ and \emph{very} high SNRs (so that the non-GR
modifications are detectable). We then recover these signals using both the SPA
templates and the simple ppE templates. For the former, we again use the prior
range on $\beta_{\mbox{\tiny ST}}$ of $|\beta_{\mbox{\tiny ST}}| \le 5$, while for the latter the prior
range on $\beta_{{\mbox{\tiny ppE}}}$ is a bit tricker. We could use the same prior range on
$\beta_{{\mbox{\tiny ppE}}}$ as in the previous subsection, but this was motivated from a
study of signals at ${\rm{SNR}} \approx 20$. The bounds on $\beta_{\mbox{\tiny ppE}}$ for the
extremely high SNR signals we are studying in this subsection should be much
stronger. We estimate the latter by relating the $-1$PN coefficient from our SPA
waveform to the ppE strength parameter when $b=-7$. This leads to a prior range
on $\beta_{\mbox{\tiny ppE}}$ of $|\beta_{\mbox{\tiny ppE}}| \le 1.035\times
10^{-10}$. We then calculate the BF for the simple ppE
model using both the full and the more restricted prior range on $\beta_{{\mbox{\tiny ppE}}}$.
Figure~\ref{fig:SPASNR} shows the BF as a function of the SNR between GR and either the SPA or the simple ppE templates, using both prior ranges for the ppE templates. Notice that the SPA templates detect the modifications at a much lower SNR than the ppE templates using the full prior, and at an SNR approximately half the value necessary for the ppE templates with the restricted prior. Low SNR, however, is a relative term. The SPA templates detect the GR modifications at SNRs $\approx 6 \times 10^{5}$, which corresponds to a ridiculous luminosity distance of $\approx 10^{3}$ pc (essentially inside the Milky Way).
\begin{figure}[ht]
\includegraphics[clip=true,angle=-90,width=0.45\textwidth]{SPAvSNR.eps}
\caption{\label{fig:SPASNR} (Color Online) BFs in favor of modified gravity as a function of SNR, calculated by injecting a signal with $\beta_{\mbox{\tiny ST}} = -3.5$. A BF over 1 indicates a preference for a non-GR theory of gravity. The red (solid) line shows the BFs calculated using SPA templates, and the blue (dashed) line shows those calculated from ppE templates.}
\end{figure}
Although we expected that the custom-made SPA templates would be more effective than the generic, ppE templates at extracting signals (at sufficiently high SNR),
it is still worth studying the reason behind this expected result. To do so, we examine the prior and posterior distributions of the two non-GR parameters, $\beta_{\mbox{\tiny ST}}$ and $\beta_{\mbox{\tiny ppE}}$, and the \emph{Occam penalty} that arises from each parameter. The Occam penalty is a built-in feature of Bayesian analysis, which causes simple models to be favored over more complicated ones. That is, models with fewer parameters are preferred to models with extra parameters, all else being fixed\footnote{Consider two nested models: $\mathcal{M}_1$ which is parameterized by a single parameter, $\theta$, and $\mathcal{M}_0$ which is unparameterized, i.e.~has $\theta=\theta_0$ where $\theta_0$ is a constant. If the likelihood function for $\mathcal{M}_1$ is a Gaussian, then ${\rm{BF}}_{1,0} \propto (\delta \theta)/\Delta \theta$, where $\delta \theta$ is the characteristic width of the
posterior in $\theta$, and $\Delta \theta$ is the prior range of $\theta$~\cite{TysonThesis}. Thus, if the value of $\theta$ is entirely unconstrained by the data, there is no penalty for an extra parameter. On the other hand, if $\theta$ is very tightly constrained, there is a large penalty.}.
\begin{figure}[ht]
\includegraphics[clip=true,width= 0.48\textwidth]{highSNRpost.eps}
\caption{\label{fig:highSNRpost} (Color Online) Posterior distributions for $\beta_{\mbox{\tiny ST}}$ (top panel) and $\beta_{\mbox{\tiny ppE}}$ with the full prior (middle panel) and restricted prior (bottom panel), generated by recovering an SNR $600,000$ signal with $\beta_{\mbox{\tiny ppE}} = -3.5$. Also plotted are the prior densities, although this is not visible in the middle panel. The posterior for $\beta_{\mbox{\tiny ppE}}$ is so highly constrained compared to the full prior that it is nearly impossible to use this model for the detection of a ST modification to gravity. }
\end{figure}
This brings us to an explanation of the results in Fig.~\ref{fig:SPASNR}. Figure~\ref{fig:highSNRpost} shows the posterior distributions for $\beta_{\mbox{\tiny ST}}$ and $\beta_{\mbox{\tiny ppE}}$, using both the full and the restricted priors, for a signal with SNR of $6 \times 10^{5}$, plotted over the entire prior range. The prior distribution is plotted in all three cases, although it is only visible for $\beta_{\mbox{\tiny ST}}$ and for $\beta_{\mbox{\tiny ppE}}$ with the restricted prior range. Notice that, where the posterior for $\beta_{\mbox{\tiny ST}}$ and $\beta_{\mbox{\tiny ppE}}$ in the restricted case have some weight over most of their entire prior ranges, $\beta_{\mbox{\tiny ppE}}$ for the full prior range is hugely constrained - so constrained that its posterior distribution looks like a delta function. This means that there is a very large Occam penalty disfavoring this model, and it will take a signal of extremely high SNR to overcome this penalty. As seen from the results in this section, a tighter prior range leads to a smaller Occam penalty, and thus a larger BF in favor of the GR deviation.
\subsection{Dynamically Scalarized Signals}
\label{dyn-sca-subsec}
Section~\ref{sec:useful-effective-ST} hinted that dynamical scalarization is much more difficult to detect than spontaneous scalarization. In fact, Fig.~\ref{fig:BFvbetaST} shows that the number of effective cycles accrued in dynamically scalarized signals is rather low in general. An example of this is the dynamically scalarized, $\beta_{{\mbox{\tiny ST}}} = -4.25$ case for a binary with masses $(1.4074,1.7415) M_{\odot}$. Indeed, a Bayesian analysis of this signal shows that such a ST effect is not detectable with simple ppE or 2-parameter ppE templates.
One way to understand this is by considering the amount of SNR that is accrued in the signal after scalarization has set in. Table~\ref{table:SNR} lists the percentage of the total $\rm{SNR}^2$ contained in the signal \emph{before} at least one of the NSs has become scalarized, for signals with $\beta_{\mbox{\tiny ST}} = -4.25$. The frequency at which either NS becomes scalarized can be easily extracted by looking at the behavior of the scalar charges as a function of orbital frequency (see Fig.~\ref{fig:fstarvbeta}). From this table, it is clear that there is very little SNR accumulated over the portion of the signal in which scalarization effects are important. Moreover, this SNR accumulation does not occur in the
frequency region in which the instrument is most sensitive.
\begin{table}[ht]
\centering
\begin{tabular}{c c c }
\hline\hline
Mass & \quad Freq.~[Hz] \quad & \quad $\% \rm{SNR}^2$ \\ [0.5ex]
\hline
$(1.4074,1.7415) M_\odot$ & 314 & 98.1 \\
$(1.5145,1.7415) M_\odot$ & 301 & 94.25 \\
$(1.6441,1.7415) M_\odot$ & 286 & 89.6 \\ [1ex]
\hline
\end{tabular}
\caption{\label{table:SNR} The percentage SNR squared accrued before dynamical scalarization has begun, for signals with $\beta_{\mbox{\tiny ST}} = -4.25$ and SNR $\approx 15$. The first column gives the mass of the system, the second the approximate frequency at which dynamical scalarization begins, and the third the percentage of the $\rm{SNR}^2$ accumulated prior to scalarization. Note that for all cases most of the $\rm{SNR}^2$ of the signal is amassed before scalarization becomes significant. }
\end{table}
The reason why the $\beta_{{\mbox{\tiny ST}}} = -4.25$, dynamically scalarized case cannot be easily detected is precisely that the frequency at which ST modifications become noticeable is rather large. But this frequency is, of course, a function of the masses of the binary and the EoS used. Recall that dynamical scalarization sets in when the energy of the system, roughly speaking the linear combination of the NS compactnesses and the (absolute value of the) gravitational binding energy, exceeds a certain threshold. Therefore, one can imagine a NS binary whose masses and radii are such that the NS compactnesses are very close to exceeding the energy threshold, and thus, dynamical scalarization can set in at very low frequencies, as shown in Fig.~\ref{fig:fstarvbeta}.
From the effective cycle study illustrated in Fig.~\ref{fig:BFvbetaST}, it appears that the dynamically scalarized binaries that are most easy to detect are those described at the beginning of Sec.~\ref{sec:BF}: a $(1.6441,1.6441) M_{\odot}$ binary at $\beta_{{\mbox{\tiny ST}}} = -4.5$, a $(1.7415,1.7415) M_{\odot}$ binary at $\beta_{{\mbox{\tiny ST}}} = -4.25$, and a $(1.5145,1.6441) M_{\odot}$ binary at $\beta_{{\mbox{\tiny ST}}} = -4.5$. Let us first consider a signal described by the first of these sets of parameters and extract it with a simple ppE template. Such a signal dynamically scalarizes at the lowest frequency of all systems considered (at roughly 80 Hz). This system is indeed detectable, leading to a very large BF and a $\beta_{{\mbox{\tiny ppE}}}$ posterior that is similar to that shown in the left panel of Fig.~\ref{fig:beta45} for a spontaneously scalarized, $\beta_{{\mbox{\tiny ST}}} = -4.5$ binary with masses $(1.4074,1.7415) M_\odot$. The width of the $\beta_{{\mbox{\tiny ppE}}}$ posterior, however, is roughly one order of magnitude larger than in the spontaneously scalarized case, with a variance of $\sigma_{1.6,1.6} = 3.7 \times 10^{-5}$ for the former and $\sigma_{1.4,1.7} = 5 \times 10^{-6}$ for the latter, indicating that the BF in this case is smaller, as expected.
We can now repeat this analysis for the other two binaries that we expect may be detectable given Fig~\ref{fig:BFvbetaST}. For both cases, we find that $BF \approx 3$, obtained from the Savage-Dickey ratio. This is again in accordance with expectations from Fig~\ref{fig:BFvbetaST}. For the $\beta_{\mbox{\tiny ST}} = -4.25$ system, the number of effective cycles indicates a marginal detection, which is precisely what we find in this Bayesian analysis. For the $\beta_{\mbox{\tiny ST}} = -4.5$ system, the number of effective cycles plotted in Fig~\ref{fig:BFvbetaST} suggests the possibility of detection; however, recall that the numbers shown in this figure are upper limits. In this case, our Bayesian analysis indicates that this upper limit is higher (by a factor of $\approx 2$) than the actual number of effective cycles induced by dynamical scalarization of the binary components.
As already mentioned, our analysis thus far has not focused on one important feature of dynamically scalarized signals: the early plunge of the NS binary. That is, once the GW frequency has exceeded the threshold for dynamical scalarization to set in, the NS binary inspirals for a few more cycles, but then plunges and merges soon after. This occurs much earlier than in GR (see e.g.~Fig.~$10$ and~$15$ in~\cite{Palenzuela:2013hsa}). Of course, after the NSs have merged, either a hypermassive NS forms, with a rotating bar that emits GWs at kHz frequencies, or a BH forms, thus cutting out GW emission exponentially through ringdown. The precise form of the waveform during this merger and ringdown phase will depend strongly on the NS equation of state.
We study in an \emph{approximate} fashion whether an early plunge can be detected in a generic modified gravity theory by considering a set of Heaviside signal injections, i.e.~GR waveforms for which the Fourier amplitude is multiplied by a Heaviside function with
argument $f^{*}_{{\mbox{\tiny inj}}} - f$, as we vary the injection cutoff frequency $f^{*}_{{\mbox{\tiny inj}}} \in (40,10^{3})$Hz. In dynamically scalarized systems, however, the transition from inspiral to early plunge and then merger is \emph{smooth}, while Heaviside templates are clearly not. Therefore, it is obvious that the latter are inappropriate templates to extract realistic dynamical scalarization signals. However, they are good and simple toy-models to study whether an early plunge (and thus an early termination) of the signal could be detected as a non-GR effect in data analysis. Since a smooth transition will be less noticeable than a sharp Heaviside transition, the use of Heaviside templates could be thought of as conservative, i.e.~if a GR deviation cannot be observed with such an abrupt termination, it certainly will not be detectable if the transition is smooth.
We carry out such a study in the following way. We place all such systems at $D_{L} =
30.75$ Mpc such that the recovered SNR is approximately $15$ when $f^*_{{\mbox{\tiny inj}}}=100 $ Hz
and $30$ when $f^{*}_{{\mbox{\tiny inj}}}=1000$ Hz. For that value of $f^{*}_{{\mbox{\tiny inj}}}$, the Heaviside signal
is thus similar to those we have been analyzing throughout this paper and also
those studied in Ref.~\cite{Sampson:2013lpa}. We then extract such injections with
templates that exactly match the signal, but with $f^{*}$ included as a template parameter to search over,
as well as with simple ppE templates with $b=-4$. This value of the exponent parameter is chosen because of its strong correlation with the total mass, which is the parameter that determines the cutoff frequency in GR. As explored in Ref~\cite{Sampson:2013jpa}, the specific value of $b$ that is chosen has little impact on the analysis.
One may worry that approximating the early plunge in this abrupt way may mask the detectability of non-GR effects, as the cycles that are effectively thrown out by this sort of study would contain these effects. Because there is so little SNR contained in those final few cycles of inspiral, however, this should not be an issue. For the systems studied in this paper, there are thousands of orbital cycles before scalarization is activated, and only tens of orbital cycles afterwards. Additionally, for all but a few cases, the orbital cycles that are affected by scalarization occur at a frequency in which the detectors are not very sensitive. These two effects combined mean that there is very little information being discarded by abruptly terminating the waveforms once scalarization has occurred.
Clearly, the earlier the binary plunges (or, in our case, the lower the injection cutoff frequency, $f^*_{{\mbox{\tiny inj}}}$), the fewer GW cycles the signal will contain in the sensitivity band of the detector. This then translates to a smaller recovered SNR. Thus, in order to detect such a signal at all, we must either be fortunate enough to detect systems that are sufficiently nearby, or fortunate enough to detect enough events such that their stacked SNR is large. Table~\ref{table:SNR-new} shows the luminosity distance required such that the SNR recovered equals 8 for different termination frequencies $f^{*}_{{\mbox{\tiny inj}}}$. This table also shows what the SNR would have been at such luminosity distances, if the signal did not terminate at $f^{*}_{{\mbox{\tiny inj}}}$, but rather continued to $1000$Hz.
\begin{table}[ht]
\centering
\begin{tabular}{c| c c c c c}
\hline\hline
$f^{*}_{{\mbox{\tiny inj}}}$ [Hz] & $100$ & $65$ & $51$ & $44$ & $39$ \\ [0.5ex]
\hline
$D_{L}$ [Mpc] & $60$ & $30$ & $18$ & $12$ & $9$ \\ [0.5ex]
\hline
${\rm{SNR}}$ & $15$ & $30$ & $50$ & $75$ & $100$ \\ [0.5ex]
\hline \hline
\end{tabular}
\caption{\label{table:SNR-new} Luminosity distance to the source such that the SNR recovered (up to a threshold frequency $f^{*}_{{\mbox{\tiny inj}}}$) is equal to 8. The third row shows the total SNR that would be recovered at the given luminosity distances if $f^{*}_{{\mbox{\tiny inj}}} = 1000$ Hz. Note how rapidly the distance to the source has to be decreased as the threshold frequency is decreased.}
\end{table}
Figure~\ref{fig:ppEthetaGRtheta} shows the posterior distributions for the recovered values of $f^*$ using Heaviside templates, and the posteriors for the recovered values of $\beta_{\mbox{\tiny ppE}}$ using simple ppE templates. All injections focus on a $(1.6,1.6) M_\odot$ NS binary, with the same polarization angle and sky position as all other injections in this paper, and $D_L = 30.75$ Mpc. The prior range on the search parameter $f^*$ is uniform between $0$ to $1000$ Hz for the Heaviside templates, and thus, the prior density is $1/1000 = 0.001$. Recall that the prior range on the search parameter $\beta_{{\mbox{\tiny ppE}}}$ is also uniform with range $-5$ to $5$ for the simple ppE templates, and thus, the prior density is $1/10 = 0.1$ in this case.
The interpretation of the posterior distributions for the recovered values of $f^*$ as tests of GR is somewhat subtle, because it is not entirely clear what the ``GR value'' for $f^*$ should be. In principle, the inspiral should end when the NSs begin to plunge, and certainly by the time the stars have come into contact. The GW frequency of the latter, $f_{\rm cont}$, depends both on the component masses and the EoS; for a $(1.6,1.6) M_\odot$ binary it is between $1250$ and $2050$ Hz, depending on the NS radius. As a simple and practical measure of the plunge, one could choose the GW frequency at the innermost stable circular orbit (ISCO) of a test particle in a Schwarzschild spacetime: $f_{\mbox{\tiny ISCO}} = 6^{-3/2}/(\pi M)$; for a $(1.6,1.6) M_\odot$ NS binary, $f_{\mbox{\tiny ISCO}} = 1354$ Hz. This is a suitable measure for the beginning frequency of plunge in GR, provided the NS is compact enough such that its contact frequency is above $f_{{\mbox{\tiny ISCO}}}$. For the systems we study here, $f_{\rm cont}>f_{{\mbox{\tiny ISCO}}}>1000$ Hz, where recall that the latter is the highest frequency of integration in all cross-correlations. We can therefore take the GR value of $f^{*}$ to be $1000$ Hz, and calculate the BF by comparing posterior and prior densities at this point.
The right panels of Fig.~\ref{fig:ppEthetaGRtheta} show that the Heaviside templates are able to distinguish a deviation from GR provided the injected cutoff frequency $f^{*}_{{\mbox{\tiny inj}}}$ is above $\approx 50$ Hz and below $\approx 400$ Hz. These panels present the posterior distributions for the parameter $f^*$, recovered using Heaviside templates on injections with various values of $f^*_{{\mbox{\tiny inj}}}$. The ability of Heaviside templates to distinguish GR deviations can be established by computing the BFs for each of these panels through the Savage-Dickey density ratio (recall that the BF is the ratio of the posterior to the prior density at the GR value of the search parameter, $f^*=1000$ Hz in the Heaviside template case). For instance, the ${\rm{BF}} \approx 1$ when $f^*_{\mbox{\tiny inj}} = 50$ Hz, while the ${\rm{BF}} \approx 5$ when $f^*_{{\mbox{\tiny inj}}} = 400$ Hz, a marginal detection of a GR deviation.
The left panels of Fig.~\ref{fig:ppEthetaGRtheta} show that the ppE templates can also distinguish Heaviside-type deviations from GR, but this time provided $f^{*}_{{\mbox{\tiny inj}}}$ is above $\approx 75$ Hz and below $\approx 400$ Hz. These panels present the posterior distributions for $\beta_{\mbox{\tiny ppE}}$, recovered using simple ppE template with $b=-4$ on Heaviside injections with various values of $f^{*}_{{\mbox{\tiny inj}}}$. The BF can still be computed through the Savage-Dickey density ratio, except that now GR is recovered when the value of $\beta_{{\mbox{\tiny ppE}}}$ is zero, and recall that the prior density is $0.1$. For instance, the ${\rm{BF}} \approx 1$ when $f^*_{\mbox{\tiny inj}} = 75$ Hz, while the BF is clearly much larger than unity when $f^{*}_{{\mbox{\tiny inj}}} = 100$ Hz.
Why is detectability of a GR deviation difficult for very low or very large $f^{*}_{{\mbox{\tiny inj}}}$? For very low injected cutoff frequencies (e.g.~below $50$Hz for the Heaviside templates and $75$ Hz for the ppE templates), the analysis fails to detect a signal altogether -- a reasonable result, considering that these very low injected cutoff frequencies drop the total recovered SNR of the signal below 8. For very high injected cutoff frequencies (e.g.~above $400$ Hz), the deviation from GR occurs too far outside of the detector's most sensitive band to be noticeable. Put another way, not enough SNR is accrued while the GR deviation is active.
Although the choice of a signal at $D_{L} = 30.75$ Mpc leads to a recovered SNR $\approx 30$ when $f^*_{{\mbox{\tiny inj}}}=1000$ Hz, a reasonable choice for comparison to other results in this paper, systems at such a close distance are not very likely. A more reasonable expectation is a system at twice that luminosity distance, $D_L \approx 61.5$ Mpc, such that the signal that \emph{would} have total SNR of $\sim15$ if $f^{*}_{{\mbox{\tiny inj}}}=1000$ Hz, but which, if subject to the early plunges analyzed here, is in fact a signal with lower recovered SNR. When we inject Heaviside signals at such a $D_{L}$, so that the SNR $\sim 15$ if $f^{*}_{{\mbox{\tiny inj}}} = 1000$ Hz, we find results qualitatively similar to those described in the previous paragraph, but with a narrower detectable injected cutoff frequency range. The high injected cut-off frequency for detectability drops to $\approx 300$ Hz, and the low cut-off frequency rises to $\approx 90$ Hz. This result is in accordance with expectations.
\begin{figure*}[ht]
\includegraphics[clip=true,width=0.45\textwidth]{ppEthetapost.eps}
\includegraphics[clip=true,width=0.45\textwidth]{GRthetapost.eps}
\caption{\label{fig:ppEthetaGRtheta} (Color Online) Left panels: posterior distributions for $\beta_{\mbox{\tiny ppE}}$, recovered by extracting a Heaviside template injection with injected cut-off frequency $f^*_{{\mbox{\tiny inj}}}$, using simple ppE templates. The only case that shows a strong preference for the non-GR model is for $f^*_{{\mbox{\tiny inj}}} = 100$ Hz. Right panels: posterior distributions for the recovered $f^*$, generated by extracting a Heaviside injection with templates of the same family and different injected cut-off frequencies $f^{*}_{{\mbox{\tiny inj}}}$. Here, the cases with $f^*_{{\mbox{\tiny inj}}}=100$ Hz and $f^*_{{\mbox{\tiny inj}}}=75$ Hz are distinguishable from GR.}
\end{figure*}
The ability to detect the early plunge of a binary system may have an important implication in terms of the detectability of dynamical scalarization. As we have seen in Fig.~\ref{fig:fstarvbeta}, higher mass NS binaries can dynamically scalarize at frequencies below $400$Hz for $\beta_{{\mbox{\tiny ST}}} \approx -4$. This then suggests that the inclusion of a early plunge, the merger and the post-merger phase in a data analysis study may allow for constraints on ST theories at such values of $\beta_{{\mbox{\tiny ST}}}$. Notice, in particular, that such constraints are stronger than those obtained when including only the inspiral phase.
One must be very careful, however, when extrapolating results and promptly concluding that the inclusion of the plunge and merger portions of these signals will increase their distinguishability. First, there are very few cycles in the post-plunge phase, and thus very little SNR accumulated after the plunge. Second, and perhaps more importantly, the post-plunge phase will be strongly affected by the NS equation of state. Degeneracies between equation-of-state effects and non-GR effects imply that the post-plunge phase of NSs may not be as useful as a means to test GR. Of course, a more detailed analysis is required to derive solid conclusions.
\section{Conclusions}
\label{sec:conclusions}
In this paper, we sought to answer one overarching question: can deviations from GR in GW signals that are caused by a certain class of ST theories be detected with aLIGO-type instruments? We find that this is the case, irrespective of whether the deviation arises from spontaneously/induced or dynamically scalarized NSs. These projected constraints will be complementary and at least comparable to current binary pulsar ones.
Not all spontaneous/induced and dynamical scalarization effects, however, are easily detectable. For a spontaneous/induced scalarization to be detectable, the scalar field anchored on each NS must be large enough, and the masses sufficiently dissimilar, that dipole radiation becomes important. This occurs, for example, for values of $\beta_{{\mbox{\tiny ST}}} \approx -4.5$ and a NS binary with masses $(1.4,1.7) M_{\odot}$, for the simple polytropic EoS that we consider in this paper (with a different EoS potentially changing these values). For a dynamically scalarized effect to be detectable in aLIGO, such scalarization must occur at a sufficiently small GW frequency so that enough GW cycles are affected in a frequency region that detectors are sufficiently sensitive to. This occurs, for example, for values of $\beta_{{\mbox{\tiny ST}}} \approx -4.5$ and a NS binary with masses $(1.6,1.6) M_{\odot}$, which scalarizes at a GW frequency of roughly 80 Hz
(again with some variability of these values depending on the EoS).
We additionally investigated the detectability of another effect associated with dynamical scalarization: the early plunge that would be induced in binaries that undergo such scalarization. We found that the sudden cutoff of a GW signal at frequencies below those expected in signals described by GR is detectable using both simple ppE templates and GR templates with a Heaviside function, for certain ranges of cutoff frequencies. Such a plunge, for example, could occur for values of $\beta_{{\mbox{\tiny ST}}} = -4$ and a NS binary with masses $(1.7,1.7) M_{\odot}$. Of course, an early plunge cannot be accurately modeled through a Heaviside function, and so a more careful data analysis study that includes inspiral-merger-post-merger signals is needed.
In the process of reaching these answers, we explored a
measure that can help estimate whether different types of phase effects are
detectable with GWs: the effective cycles of phase. We found that for a wide
variety of different non-GR signals, the modification to the GR phase needs to
lead to roughly 4 cycles of effective phase to be detectable with an aLIGO
detector at SNR 15. We further found that such detectability is independent of whether one
uses custom-made ST templates or model-independent templates to search for
deviations at the expected SNRs of aLIGO-type detectors.
A final question that we considered was whether custom-made, theory-specific templates are more useful at detecting deviations from GR than the model-independent ppE template family. We answered this question by looking at both extremely high SNR signals with ST signatures that were undetectable at low SNR, and by looking at low SNR signals with ST signatures that were easily detectable at reasonable SNR. In both cases, we found that ppE templates perform almost as well as custom-made, SPA templates at distinguishing non-GR signals from GR ones.
Future work could concentrate on extensions of the analysis presented here. One interesting extension would be to repeat this work for NSs with realistic EoSs. The work in Ref.~\cite{Barausse:2012da,Palenzuela:2013hsa}, which we used in this paper exclusively, used a polytropic EoS, but this is easily generalizable to more realistic EoSs (c.f.~Ref.~\cite{Shibata:2013pra}). One may find EoSs and masses that do not lead to spontaneous/induced scalarization for binary pulsars, yet lead to spontaneous/induced scalarization for systems that can be detected with GWs. Such systems would thus evade binary pulsar constraints and yet potentially lead to detectable deviations with aLIGO. Again, it is important to emphasize that stiffer/softer EoSs will lead to qualitatively (and quantitatively) different behavior. A very stiff EoS could perhaps support more compact stars that spontaneously scalarize at lower frequencies. A systematic study of these effects is left for future consideration.
Another interesting analysis would be to study whether one can find EoSs or masses for which dynamical scalarization sets in at very low frequency, e.g.~close to 10Hz. Reference~\cite{Sampson:2013jpa} estimated that for an abrupt GR modification to be detectable with an aLIGO-like detector at SNR 10, such a modification would have to start at a GW frequency below 100 Hz. Most cases of dynamical scalarization that we explored occur at $\sim$200Hz or higher, but there are some instances in which scalarization occurs at a lower frequency, and it is possible that different mass/EoS combinations would produce more of these scenarios.
Other extensions may include adding more complexity to the signals and the templates, through the inclusion of the merger and post-merger phases, as well as the inclusion of spin~\cite{Klein:2013qda,Chatziioannou:2013dza,Chatziioannou:2014bma,Chatziioannou:2014coa} and eccentricity~\cite{Yunes:2009yz} effects.
Recall that with respect to second-generation, ground-based detectors it is common to regard the
NS merger and post-merger phases as unimportant for testing GR as they occur at kHz frequencies where such detectors
are least sensitive.
Dynamically scalarized NSs, however, could plunge at much lower frequencies, and such effects may be detectable. References~\cite{Chatziioannou:2014bma,Chatziioannou:2014coa} showed that the inclusion of spin in NS binaries can have a large effect on parameter estimation. Similar conclusions were arrived at when including more complexity in GW signals to test GR (see e.g.~\cite{Stavridis:2010zz,Yagi:2009zm,Yagi:2009zz}). Also, Ref.~\cite{Palenzuela:2013hsa} showed that eccentric NS binaries in ST theories can give rise to scalarization/descalarization phenomena that may affect the binary's orbital evolution (effectively decreasing the eccentricity faster than in GR) at sufficiently low frequencies to be detected.
Another interesting avenue for future work would be to repeat the analysis of this paper but with a non-template based search algorithm that may be more sensitive to the post-merger phase. In our analysis, we used a template-based search of inspiral signals, neglecting the post-merger phase that occurs at high GW frequencies where aLIGO's sensitivity will be weaker. Recently, however, Ref.~\cite{Clark:2014wua} used a template-free burst algorithm to show that the merger phase may be sufficiently detectable by aLIGO to discern between a prompt collapse scenario and the formation of a hypermassive NS, if the event occurs at $\approx 10 \; {\rm{Mpc}}$. Dynamical scalarization will modify the post-merger phase, also leading to either prompt collapse or hypermassive NS formation, depending on the masses of the binaries. Such effects, however, will probably be somewhat degenerate with the EoS, and thus, it is unclear whether merger modifications will be strong enough to detect a GR deviation.
One final way to study the robustness of our conclusions would be to consider GW detection with a network of GW detectors, with second-generation detectors and noise tuning, with a large number of second-generation detections and stacking~\cite{Berti:2011jz,DelPozzo:2013ala}, or with future GW detectors, such as the Einstein Telescope~\cite{et}. All of these could lead to much lower noise at GW frequencies around 100 Hz, which would then have multiple effects. First, better sensitivity at 100Hz should push the threshold frequency at which GR deviations can be detected to higher frequencies. Second, better sensitivity overall should lead to individual detections with higher SNR and to a higher number of detections per year. Combining all of this, and perhaps through the use of stacking, one may be able to detect dynamical scalarization for lower values of $|\beta_{{\mbox{\tiny ST}}}|$. One should keep in mind, however, that the direct detection of the additional breathing mode will be extremely hard, even when detecting GWs with multiple instruments, since in ST theories that pass Solar System tests, the interaction of such modes with a detector is suppressed by $\psi_0$~\cite{Barausse:2012da}, which is constrained to $\lesssim10^{-2}$ because of the Cassini bound.
Finally, let us address the relationship between our work and that reported in Ref.~\cite{Taniguchi:2014fqa}, which appeared after the submission of this paper. One may be led to believe that the conclusion arrive at in this paper and those of Ref.~\cite{Taniguchi:2014fqa} are not in agreement, with respect to the detectability of scalarization effects with aLIGO. Reference~\cite{Taniguchi:2014fqa} first calculates the total number of cycles of phase that are accumulated in a GW signal due to the presence of scalarization effects. They then note that this number is larger than or comparable to the total number of cycles of phase that will be accumulated due to EoS effects within GR. Since the latter may be detectable with next generation detectors~\cite{Damour:2012yf,Wade:2014vqa,DelPozzo:2013ala}, they then argue that scalarization effects may also be detectable. This conclusion is in fact in \emph{agreement} with our findings because EoS effects can be measured only provided the SNR is sufficiently high (roughly above 30). Our analysis used SNRs in the tens, as expected from the first few years of detection; for such signals, non-GR effects are not detectable. Reference~\cite{Taniguchi:2014fqa} also finds that scalarization can be detected when it occurs at GW frequencies larger than roughly 130 Hz, which is in perfect agreement with our findings (once one converts orbital to GW frequency).
Although Ref.~\cite{Taniguchi:2014fqa} does not claim that the total cycles of phase due to a particular effect can directly tell us about detectability, it is possible to misread the conclusions of this paper to indicate that they can. We therefore emphasize again here that the total cycles of phase have no analytic connection to the Bayes factor and, in fact, fail to account for parameter covariances. This affects detectability in two main ways. One is obvious - a non-GR phase term that has large correlations with other system parameters will be easy to fit using GR templates. This will make the effect very difficult to discern, as illustrated in Fig.~\ref{fig:BFNuse}. The other, less obvious consequence is that although effects that enter at high PN order accumulate more slowly, they may be detectable when they lead to smaller numbers of phase cycles than those that enter at low PN order. This is again directly due to parameter covariances. In the case of spontaneous or dynamical scalarization, non-GR effects that enter at low PN order will have high covariances with system parameters, such as the chirp mass. Higher PN order effects will have weaker covariances, but they lead to much weaker effects, with most of the total dephasing accumulating from the low PN order terms that have large covariances. Thus, the use of total cycles of phase to claim detectability of non-GR effects is not appropriate and one should really either use the effective cycles discussed in this paper or a full Bayesian analysis.
Also, Ref.~\cite{Taniguchi:2014fqa} (at least in its first arXiv version) states that the simulations of Ref.~\cite{Barausse:2012da} ``misread'' the output of the LORENE code used to determine the simulation initial data, and claims that the gravitational masses given in Ref.~\cite{Barausse:2012da} are incorrect. This is certainly not the case and the gravitational masses we give in Ref.~\cite{Barausse:2012da} are exactly those of the LORENE output. As it is well known, there is no unambiguous way of defining individual gravitational masses for a tight binary system within General Relativity. We provided the readers of Ref.~\cite{Barausse:2012da} with the exact LORENE-given gravitational masses in order to clearly identify our initial data and ensure our results will be reproducible by others. An
alternative, which is the one followed in Ref.~\cite{Taniguchi:2014fqa}, would have been to give the gravitational masses of the stars in isolation, but this may have caused unnecessary confusion, as the gravitation masses are never used in Ref.~\cite{Barausse:2012da} except for identifying the initial data.
Reference~\cite{Taniguchi:2014fqa} also raises question on the dynamics of the most massive case presented in Ref.~\cite{Barausse:2012da} and used here. Their concern is related to our choice of initial data for such already scalarized case, and is about whether this could cause an earlier plunge. Such concern was already addressed in Ref.~\cite{Palenzuela:2013hsa}, where we \textit{(i)} show the results of simulations with large initial separations (e.g. Figs.~4 and 5, where no plunge is present at large separations) and \textit{(ii)} validate our simulations with an enhanced PN model.
Moreover, we stress here that the initial data used in Ref.~\cite{Barausse:2012da} for the low-mass case with dynamical scalarization are exact (because in the absence of spontaneous scalarization, the ST initial data are the same as in GR)
\acknowledgments
We thank Gilles Esposito-Farese and Gabriela Gonzalez for
insightful conversations.
NY acknowledges support from NSF grant PHY-1114374 and the NSF CAREER
Award PHY-1250636, as well as support provided by the National
Aeronautics and Space Administration from grant NNX11AI49G, under
sub-award 00001944. LS and NJC acknowledge support from NSF grant PHY-1306702.
EB acknowledges support from
the European Union's Seventh Framework Programme (FP7/PEOPLE-2011-CIG)
through the Marie Curie Career Integration Grant GALFORMBHS PCIG11-GA-2012-321608.
AK is supported by NSF CAREER Grant No. PHY-1055103.
LL acknowledges support by NSERC through a Discovery Grant and CIFAR. LL thanks
the Institut d'Astrophysique de Paris and the ILP LABEX (ANR-10-LABX-63),
for hospitality during a visit supported through the Investissements d'Avenir programme under
reference ANR-11-IDEX-0004-02.
Research at Perimeter Institute is supported through Industry Canada
and by the Province of Ontario through the Ministry of Research and Innovation.
|
\section{Introduction}
In recent years, topology has become an important tool in classifying the phases of matter\cite{reviews}.
Although the study of topological phases started with the discovery of the quantum Hall
and fractional quantum Hall phases \cite{fqhereviews}
in the eighties, it gained momentum with the discovery of the
time-reversal invariant topological insulators\cite{TI} a few years ago. Topological insulators are classified in
terms of their bulk band-structure and by now, there has been a complete classification of
free fermion topological insulators in the presence of disorder\cite{ludwig,kitaev}. All these phases
have topologically non-trivial momentum space structure, are insulating in the bulk and
have metallic surface states.
It has been generally assumed that it is the gap in the bulk electronic spectrum which makes the
topologically non-trivial ground state with its surface states, stable, and unable to
decay to the topologically trivial phase. However, more recently it has been shown that it is possible to have non-trivial momentum space topology even for gapless
fermionic systems. One such recently identified system is the
Weyl semi-metal phase\cite{vishwanath, balents} which
has isolated gapless points (weyl nodes) in the bulk spectrum, where exactly 2 bands touch.
The low energy behaviour close to these points is given by a Weyl hamiltonian of fixed chirality.
The Weyl nodes are topologically protected, because a gap cannot be opened unless two nodes of opposite chirality are coupled.
The band structure shows unusual surface states called Fermi arcs\cite{vishwanath}, which
has led to many interesting work\cite{many}.
The topological
response of the phase has been argued to be a realization of the Adler-Bell-Jackiw anomaly\cite{anomaly}
in condensed matter systems. There are several recent reviews\cite{wsmreviews} which have
discussed various interesting properties of Weyl semimetals.
The introduction of superconductivity in topological insulators, characterises a new exotic phase, the topological superconductor,
whose inherent particle-hole symmetry leads to
surface states that support Majorana fermions.
The superconductor doped TI $\text{Cu}_x\text{Bi}_2\text{Se}_3$\cite{hor}
was theoretically predicted\cite{fuberg, sato} to be a TI and experimental evidence was obtained\cite{ando} using
point-contact spectroscopy to detect the itinerant massless Majorana state on the surface.
The introduction of superconductivity via the proximity effect\cite{fukane}
has also led to considerable work on topological
insulator-superconductor hybrid
junctions\cite{stanescu,sitthison,tanaka,linder} with special attention to the surface states that develop between them.
Proximity with an $s$-wave superconductor was shown to lead to a significant renormalization
of the parameters in the effective model for surface states. It was also shown that
when the fermi surface is
close to the surface Dirac cone vertex, the electrons exhibit $s$-wave pairing, but away from the
vertex, the triplet component increases in amplitude. A full symmetry classification
of all the induced pairings for proximity to $s$-wave, $p$-wave and $d$-wave superconductors
was also studied\cite{balatsky,balatsky2} and it was shown that the different induced pairing amplitudes
modify the density of states at the interface significantly\cite{tanakarev,balatskyrev}.
\begin{figure}
\includegraphics[width=0.45\textwidth]{Fig1.pdf}
\caption{(Color online) Schematic of our setup - an $s$-wave superconductor is coupled with the Weyl semi-metal/topological insulator system through proximity. Cooper pairs from the superconductor diffuseninto the bulk of the Weyl semi-metal/topological insulator giving rise to induced superconductivity.}\label{geometry}
\end{figure}
Similarly, one might expect that the introduction of superconductivity in the
Weyl semi-metal would also lead to new phenomena. A heterostructure of topological insulators
and $s$-wave superconductors was studied by Meng and Balents\cite{mengbalents} who showed
that superconductivity split the Weyl modes into Boguliobov-Weyl modes. By studying vortices
in some of these phases, characterised by different number of Weyl modes, they found zero-energy
Majorana modes under some conditions. Cho {\it et al}\cite{choetal} studied superconducting states
of doped inversion symmetric Weyl semi-metals and showed that the finite momentum FFLO pairing
state is energetically favoured over the even parity BCS state.
Recently, Lee {\it et al}\cite{klee} studied the proximity effect in topological insulators, when the
chemical potential is close to, but not in the bulk gap. They found that the superconducting gap penetrates
the bulk and is observable at the naked surface, opposite to the one in proximity to the superconductor.
However, there has been no systematic study of proximity induced superconductivity in Weyl semi-metals,
which is the main focus of this paper. Since our model also includes the topological insulator
phase for some region of parameter space, we also provide results for proximity induced
superconductivity for topological
insulators in
our model for comparison. Our model consists of a 3D topological insulator, converted to a Weyl semi-metal, by
including either parity breaking terms or time-reversal breaking terms or both. Time-reversal breaking
leads to Weyl nodes at the same energy and surface states, which form a Fermi arc between the nodes,
whereas parity breaking leads to Weyl nodes
at different energies and no Fermi arcs.
The dispersion of the Fermi arcs or surface states is flat in a particular direction and chiral in other direction
and points along the Fermi arc can be understood as the edge states of a two dimensional Chern insulator.
When inversion symmetry is broken however, the Weyl nodes
are separated in energy and we do not get surface states, distinct from the
bulk. Since we are mainly interested in the new physics coming from the Fermi arc states,
unless otherwise specified, in this paper we always consider Weyl semi-metal phase induced by having a time reversal breaking perturbation.
Superconductivity is then
induced in the semi-metal by coupling it to an $s$-wave superconductor on one of its surfaces.
We compute the self-energy of the topological insulator/semi-metal electrons by integrating out the superconductor
degrees of freedom and use the imaginary part of the Green's function to compute the local density of states (LDOS).
We find that the superconductor induces coherence peaks on the LDOS of the
electrons on a few layers close to the interface and we contrast the behavior of the LDOS on
different layers for the TI and the WSM. For the TI, we find the reduction in the LDOS close to
$\omega=0$ which is the hallmark of the gap formation, whereas for the WSM, we find that the enhancement of the density
of states without a superconductor (the hallmark of the flat band) split into two bands with a reduction
of the density of states at $\omega=0$. We study in detail the band structure of the surface states
and find that the surface state of the TI is completely gapped by the proximity effect, whereas
the surface states of the WSM get split and acquire a small gap,
but the Weyl nodes remain unaffected.
Thus the surface band acquires a superconducting gap in a TI, but
no true superconducting gap is induced in WSM~\cite{mengbalents}
We also study the behavior of the induced pairing amplitudes (singlet, triplet,
intra-orbital and inter-orbital) as a function of the various parameters in the theory.
As shown in Ref.~\onlinecite{balatsky}, the induced pairing amplitudes in $\text{Bi}_2\text{Se}_3$ - type
materials with tetragonal symmetry are classified in terms of the irreducible representations $\Gamma$ of the $D_{4h}$ group.
Since we are only considering proximity with an $s$-wave conductor in this paper, we are only interested in representations with total
angular momentum $J_z=0$.
We find that the induced pairings fall of exponentially fast away from the interface both
in the TI and in the WSM. But they are not very sensitive to other parameters such as the
chemical potential and the time-reversal breaking parameters. The induced pairings increase as a function
of the superconducting pairing amplitude of the superconductor and the coupling of the superconductor to the TI/WSM.
It is also perhaps worth mentioning that both in the singlet and triplet amplitudes, the symmetries of the
two largest amplitudes reverse between the TI and the WSM, with inter-orbital pairings being larger
in the TI and intra-orbital pairings being larger in the WSM.
\begin{figure}
\includegraphics[width=0.45\textwidth]{Fig2.pdf}
\caption{(Color online) A typical phase diagram of our model system. The Weyl semi-metal (WSM) phase appears at the strong topological insulator (STI)/ normal insulator (NI) ($\epsilon = 6t$) and strong topological insulator (STI)/ weak topological insulator (WTI) ($\epsilon = 2t$) boundaries with broken time reversal/ parity perturbations. The WSM phase extends with increasing perturbations (blue/filled region). Parameters used here are $\lambda_{z} = \lambda_{\text{SO}}$, and $\bs{b} = (0.6\lambda_{\text{SO}} ,0,0)$. }
\label{fig:phasediagram1}
\end{figure}
\section{Model System}
We start with a simple tight-binding four-band lattice model for the topological insulator (TI) in three dimensions (3DTI), which can describe strong and weak topological insulators,
Weyl semi-metals and ordinary insulators depending on the
parameters of the model.
The $\text{Bi}_2\text{Se}_3$ family of 3DTI, have an effective
description in terms of the Hamiltonian given by
$H_0 = H_{\text{C}}+ H_{\text{SO}}$ \cite{vazifehfranz} with
\begin{align}\label{eq:h0}
H_{\text{C}}&=\epsilon \sum_{\bs r} \psi^{\dagger}_{\bs r}\tau_x\psi_{\bs r} - t\sum_{\langle \bs r, \bs r' \rangle}\psi^{\dagger}_{\bs r}\tau_x\psi_{\bs r'} + ~\text{h.c.} \nonumber \\
{\rm and} ~~~ H_{\text{SO}} &= i\lambda_{\text{SO}}\sum_{\bs r} \psi^{\dagger}_{\bs r}\tau_z\left( \sigma_x\psi_{\bs r + \bs y} - \s_y\psi_{\bs r + \bs x} \right) \nonumber \\
&+i\lambda_z\sum_{\bs r}\psi^{\dagger}_{\bs r}\tau_y\psi_{\bs r + \bs z} + ~\text{h.c.}
\end{align}
where $\psi_{\bs r}$ is the fermion operator in TI region. $\bs r, \bs r'$ refer to site indices in all three dimensions (in the TI region)
and $\bs r + \bs x$ refers to the nearest neighbour of site at $\bs r$ in $x$ direction (similarly for $y$ and $z$ directions).
Here $z$ is taken as the growth direction and $\bs{\s}$ and $\bs{\tau}$ denote Pauli matrices in spin and parity (orbital) space respectively.
$\epsilon$ and $t$ denote the on-site and nearest neighbour hopping amplitudes.
$\lambda_{\mathrm{SO}}$ and $\lambda_{z}$ are the (possibly anisotropic) spin-orbit (SO) interaction strengths
in the $x$-$y$ plane and in the $z$ direction respectively.
The topological invariants for the 3DTI, $\nu_{\mu} = (\nu_0; \nu_1, \nu_2, \nu_3)$ can be computed easily
(due to parity invariance \cite{fukane}) and are given by
\begin{align}
(-1)^{\nu_0} &= \text{sgn}\left[ (\epsilon - 6t) (\epsilon + 6t) (\epsilon - 2t)^3 (\epsilon + 2t)^3 \right], \non \\
(-1)^{\nu_i} &= \text{sgn}\left[ (\epsilon + 6t) (\epsilon - 2t) (\epsilon + 2t)^2 \right], \nonumber
\end{align}
for $i = 1, 2, 3$. This implies that we have the following phases:
\begin{align} \begin{array}{ccc}
|\epsilon| > |6t| & \nu_{\mu} = (0;0,0,0) & \text{Ordinary Insulator} \\
|6t| > |\epsilon| > |2t| & \nu_{0} =1 & \text{Strong TI} \\
|2t| > |\epsilon| > 0 & \nu_{0} = (0;1,1,1) & \text{Weak TI}\end{array}
\end{align}
At the boundaries of the topological phase transitions (at $\epsilon \approx \pm 6t, \pm2t$),
the bulk gap closes and the effective hamiltonian is a massless Dirac hamiltonian.
By introducing either parity (inversion) or time reversal (TR) symmetry breaking perturbations
to the Hamiltonian $H_0$, the Dirac node can be split in 2 weyl nodes separated in energy or momentum respectively.
Thus the Hamiltonian for the WSM is given by
$H_W = H_0 + H_{\text {E}}$ where
\begin{equation}\label{eq:he}
H_{\text{E}} = \sum_{\bs r} \psi^{\dagger}_{\bs r} \left( b_0\tau_y\s_z - b_x\tau_x\s_x + b_y\tau_x\s_y + b_z\s_z\right)\psi_{\bs r} ~.
\end{equation}
Here $b_0$ and ${\bs b}$ are parameters that break inversion and TR symmetry
respectively \cite{vazifehfranz}.
Note that the Dirac node mentioned here is the 3D Dirac node that occurs in the bulk spectrum at the phase transition between the normal and
topological insulator and should not be confused with the 2D Dirac nodes, which occur in the surface spectrum of the TI phase. The phase diagram of the different phases in this model is given in Fig.~\ref{fig:phasediagram1}.
\begin{figure}[ht]
\includegraphics[width=0.48\textwidth]{Fig3.pdf}
\caption{(Color online) The dispersion for a WSM with 2 Weyl nodes at $\pm b_x /\lambda_z$ is shown along $k_x$ and $k_y$.
The parameters used are $\bs b =(0.50t,0,0), \lambda_{SO}=\lambda_z=0.50t$. The dashed (red) lines denote both the surface bands. Note that the surface states at opposite ends have opposite chirality.}
\label{dispersion}
\end{figure}
\begin{figure*}[tb]
\includegraphics[width=1.\textwidth]{Fig4.pdf}
\caption{This panel shows the LDOS
for the TI and the WSM with the top row without coupling to the superconductor and the bottom row with coupling to the superconductor.
(a) and (b) show the LDOS
for the TI integrated over all momenta without and with coupling to the superconductor respectively.
(c) and (d) show the same for the WSM. (e) and (f) show the LDOS at $k_y=0$ summed over all $k_x$
for the WSM without and with coupling to the superconductor respectively.
The LDOS for the different layers are vertically offset for visibility.
The parameters used are $\lambda_{SO} = \lambda_z = 0.5t$, $\Delta = 0.7t$, $\epsilon = 4t$ (for TI) or $\epsilon = 6t$ (for WSM) and $\bs b = (0.5t,0,0)$ for WSM.}
\label{coherencepeak}
\end{figure*}
For finite systems, both TI and WSM phases give rise to surface states. For strong topological insulators,
surface states exist on each surface as mid-gap states\cite{TI}, whereas in weak topological insulators,
surface states arise only on particular surfaces depending on the values of $\nu_i$ ($i=1,2,3$)\cite{weakTI}.
For Weyl semi-metals (within this model), if only inversion symmetry is broken,
the Weyl nodes are separated in energy and there are no surface states
separable from the bulk states. Surface states arise only when TR symmetry is broken. As an example, if the TR symmetry is broken by ${\bs{b}} = b_x \hat{x}$, then
the Weyl nodes occur with a separation of $b_x/\lambda_z \hat{x}$ in momentum space.
In this case, we find surface states exist on the surfaces parallel to the $x-y$ and $x-z$ plane (not on the third plane).
For a large enough system, the dispersion of surface states is flat between the two Weyl nodes along the $k_x$ direction, and is linear
along $k_y$ (on $x-y$ plane) or $k_z$ (on $x-z$ plane).
The surface states exist only between the two Weyl nodes and
the states on opposite surfaces have opposite chiralities, as illustrated in Fig.~\ref{dispersion}. These are the Fermi arc states.
Fermi arcs can be understood as the edge states of a Chern insulator that exists for each value of $k_x$ between the Weyl nodes.
It may also be worth noting that the WSM formed near $\epsilon \sim 6t$ and $\epsilon \sim 2t$ have their chiralities
reversed for the top and bottom edges.
For the rest of this paper, we concentrate on the TR broken Weyl semi-metal, with $b_0=0$ and ${\bs{b}} \ne 0$,
since we are interested in the proximity
effect of the superconductor on the surface, i.e., on the Fermi arc states.
\section{Coupling to the superconductor}
We now couple one of the surfaces of the WSM to an $s$-wave superconductor as shown in Fig.~\ref{geometry}. The bulk Hamiltonian of the $s$-wave superconductor is given by
\begin{align}\label{eq:hs}
H_\text{S} &= \epsilon_\text{sc} \sum_{\bs R,\sigma} \Phi_{\bs R,\sigma}^{\dagger}\Phi_{\bs R,\sigma} -
t_\text{sc} \sum_{\langle \bs R, \bs R' \rangle,\sigma} \Phi_{\bs R,\sigma}^{\dagger}\Phi_{\bs R',\sigma} \nonumber \\
&+ \sum_{\bs R} \Delta \Phi_{\bs R,\uparrow}^{\dagger}\Phi_{\bs R,\downarrow}^{\dagger} + ~\text{h.c.}~,
\end{align}
where $\Phi$ is the fermion operator in the superconductor. $\bs R, \bs R'$ refer to site index in all three directions (in the superconducting region).
$\epsilon_\text{sc}$ and $t_\text{sc}$ denote the on-site energy and nearest-neighbour hopping amplitudes in the superconductor respectively.
The coupling is a tunnelling term between the top layer of the superconductor and the bottom layer of the WSM -
\begin{align}\label{eq:ht}
H_{\text{T}} = \sum_{\bs r_c, \tau, \sigma} \tilde{t}_{\tau} \psi_{\bs r_c,\tau,\sigma}^{\dagger}\Phi_{\bs r_c+\bs z,\sigma}+ ~\text{h.c.}~,
\end{align}
where, $\bs r_c$ denotes the sites in the last layer of the WSM (perpendicular to the interface) and $\bs r_c + \bs z$ denotes the first layer of
the superconductor.
$\tau$ denotes the orbital in the WSM and $\sigma$ is the spin.
$\tilde{t}_\tau$ is the tunnelling amplitude which can be different for the two orbitals.
In this work, we have assumed that $t_\tau$ is same for both orbitals for simplicity.
Using different tunnelling amplitudes for the two orbitals will change the results quantitatively, but not qualitatively.
For detailed results as a function of the ratio of the tunnelling amplitudes for the TI, see \onlinecite{balatsky}.
\begin{figure}[h]
\includegraphics[width=0.48\textwidth]{Fig5.pdf}
\caption{(Color online) (a) and (b) show the exponential parameter $\alpha$ of the decay (in the unit of 1/lattice spacing in $z$) for various cases in the TI/WSM,
as a function of
the chemical potential $\mu/t$. They are
denoted by the black and blue (gray) solid lines for $s$-wave intra and inter--orbital amplitudes respectively, and the black line for $p$-wave intra--orbital amplitude. In the inset of (b), we show a typical exponential decay of the pairing in the bulk. The case shown here is the decay of intra-orbital $s$-wave pairing in
the WSM. (c) and (d) show the almost flat behavior of the pairing amplitudes $F^{a \pm}$ as defined in Eq.~(\ref{eq:F}) at $z=0$ as a function of the
chemical potential.
The various parameters used are $\lambda_{\text{SO}}=\lambda_z=0.5 t$, $\Delta=0.7t$, $\lambda_S = 0.3t$, $\epsilon = 6t, ~ \bs{b}=(0.5t,0,0)$ for WSM and $\epsilon = 4t, ~ \bs{b}=(0,0,0)$ for TI.}
\label{fig:withz}
\end{figure}
\begin{figure}[h!t]
\includegraphics[width=0.4\textwidth]{Fig6.pdf}
\caption{(Color online) The pairing amplitudes as a function of (a) the time-reversal
breaking parameter $b_x/t$ (b) $\epsilon/t$. The different pairing amplitudes are denoted by the black and blue (gray) solid lines as $s$-wave intra--orbital and inter--orbital amplitudes respectively, and the black line as $p$-wave intra--orbital amplitude. Note the crossing of the black and blue lines (crossing of intra and inter orbital pairings) as a function of $\epsilon/t$.
Various parameters used are $\lambda_{\text{SO}}=\lambda_z=0.5 t$, $\Delta=0.7t$, $\lambda_S = 0.3t$. For (a), we have used $\epsilon = 6t$, and $\bs{b}=(0,0,0)$ for (b).}
\label{WSMparameter}
\end{figure}
The Nambu basis for the fermions in the WSM, denoted by $\tilde{\Psi}^{\dagger}_{\bs r}$ is given by
\begin{align}
\left(
\psi^{\dagger}_{\bs r, \uparrow, 1}, \psi^{\dagger}_{\bs r, \downarrow, 1},\psi^{\dagger}_{\bs r, \uparrow, 2}, \psi^{\dagger}_{\bs r, \downarrow, 2}, \psi_{\bs r, \downarrow, 1},
-\psi_{\bs r, \uparrow, 1},
\psi_{\bs r, \downarrow, 2}, -\psi_{\bs r, \uparrow, 2} \right) \label{eq:nambu}
\end{align}
where $\uparrow,\downarrow$ refer to the spin and $1,2$ refer to the orbitals.
Since the Hamiltonian is quadratic in the superconductor degrees of freedom $\Phi$, we can integrate
them out and compute an effective action for the WSM.
Following the analysis for the self-energy in Ref.\cite{simonbena}, we decouple the superconductor and TI degrees of freedom and define the
(Nambu-Gorkov) Green's function $G(\omega)$ for the WSM. The derivation is sketched in Appendix A. The Green's function is -
\begin{align} G(\omega) = \left[(\omega + i\delta)\mathbf{I} - H_W - \Sigma(\omega)\right]^{-1}.
\end{align}
where the self-energy is -
\begin{align}\label{eq:GF}
\Sigma_{\bs r} (\omega) = \delta_{\bs r, \bs r_c} \frac{\pi N(0) (\tilde{t})^2}{\sqrt{\Delta^2 - \omega^2}}
\left[ \omega \mathbf{I}_{\eta} - \Delta \eta^x \right] \left(\mathbf{I}_{\tau} + \tau^x \right)
\mathbf{I}_{\sigma}
\end{align}
with $\bs{r}_c$ denoting the sites in the last layer of WSM.
Note that the $G$ at each site is an $8\times 8$ matrix comprising of the spin, orbital and particle-hole pseudo-spin
subspaces. Here, we have used only the local (on-site) component of the self energy.
This approximation usually works very well and we shall justify this in the last section by comparing our
results with this approximation to the result obtained using exact diagonalisation.
\begin{figure*}[tb]
\includegraphics[width=1.\textwidth]{Fig7.pdf}
\caption{(Color online) We depict the pairing amplitudes $F^{a \pm}$ as defined in Eq.~(\ref{eq:F}) at $z=0$ for the intra- and inter-orbital $s$-wave and intra-orbital $p$-wave
channels as a function of the various parameters in the model. The black and blue (gray) solid lines denote intra- and inter-orbital pairing for WSM respectively, whereas, the black and blue (gray) dashed lines mark intra and inter-orbital pairing for TI respectively. (a) and (b) show the increase in both the $s$-wave and $p$-wave pairing channels as a function of the coupling $\lambda_S$ between the superconductor and the WSM/TI. (c) and (d) show the increase as a function of the superconducting $s$-wave pairing amplitude $\Delta$ in the superconductor. Various parameters used, which are not varied, are $\lambda_{\text{SO}}=\lambda_z=0.5 t$, $\Delta=0.7t$, $\lambda_S = 0.3t$, $\epsilon = 6t, ~ \bs{b}=(0.5t,0,0)$ for WSM and $\epsilon = 4t, ~ \bs{b}=(0,0,0)$ for TI.}
\label{pairingmag}
\end{figure*}
\section{LDOS and the pairing amplitude}
In this section, we use the Green's function to obtain the local density of states (LDOS) and the induced pairing amplitude both in the TI and WSM phases and discuss the dependence of the pairing on the various parameters of the model.
Using the Green's function that we derived in the previous section, we can compute the local density of states (LDOS) in the TI/WSM
using
\begin{align}
D(\omega,\bs{r}) &= -\frac{1}{\pi} \sum_{\sigma,\tau}{\rm Im} G^{\sigma\tau}_{\bs{r}\bs{r}}~,
\end{align}
$\sigma$ and $\tau$ being the spin and orbital index. As we increase the coupling to the superconductivity, the LDOS shows the appearance of coherence peaks in the band structure as a signature of proximity induced superconductivity. This is shown in the panel in Fig.~\ref{coherencepeak}
where we have plotted the LDOS at different values of $z$, the
layer index, both for the TI and the WSM. Panels (a) and (b) show the
LDOS as a function of energy (summed over all momenta) for the TI. Note the Dirac spectrum feature of the surface states for the TI
(layers $z=0$ and $z=9$). We can also clearly see the dip in the density of states and the appearance of the coherence peaks in the first 2 layers.
The coherence peaks are not sharp because we are at zero doping. (For the TI, the sharpness of the coherence peaks increases with
the doping, due to the increase in density of states. For WSM, we have checked that there is no significant change when the doping
is increased, because there is a significant density of states at the Weyl node even at zero doping). Panels (c) through (f) are for the Weyl semi-metal, with (c) and (d) being the LDOS summed over all momenta and (e) and (f) being the LDOS at $k_y=0$ summed over all $k_x$. First, we note the absence of the Dirac spectrum feature in the edge states. Instead, there is peak in the DOS at $\omega =0$ (and $k_y=0$) which is the signature of the flat band in the absence of coupling to the superconductor. (This feature gets lost when all $k_y$ is summed over, which is
why we have also chosen to show the band-structure without summing over $k_y$.) With the coupling to the superconductor,
the single peak splits into two with a small gap. The effect of the proximity of the superconductor on the surface states will be studied in greater detail in the next section.
Following \cite{sigrist}, we now define the different induced pairing amplitudes.
Assuming translation invariance in $x$ and $y$ directions,
we go to momentum space $\bs k = (k_x, k_y)$ in 2 directions. For each momentum ${\bs k}$ and each $z$ coordinate,
the Green's function can be written as
\begin{align}
G_{z,\bs{k}}(\omega) &=\left(\begin{array}{cc}
h_{z}(\omega, \bs{k}) & \bar{\Delta}_z(\omega,\bs{k}) \\
\bar{\Delta}^{*}_z(\omega,\bs{k}) & h'_{z}(\omega, {\bs k})
\end{array}\right)~.
\end{align}
The $4\times 4$ pairing matrix $\bar{\Delta}_{z;\sigma\tau,\sigma'\tau'}(\omega,{\bs{k}}) $ is related to the
pairing amplitudes as
\begin{align}
\bar{\Delta}_{z;\sigma\tau,\sigma'\tau'}(\omega,{\bs{k}}) =& \int_0^{\infty} \frac{dt}{2\pi}\langle c_{z,{-\bs{k}},\sigma,\tau}(t)
c_{z,{\bs{k}},\sigma',\tau'}(0)\rangle e^{i\omega t} \nonumber \\
=& \int_0^{\infty} \frac{dt}{2\pi}{\hat\Delta}_{z;\sigma\sigma'\tau\tau'}({\bs k}, t)e^{i\omega t},
\end{align}
where the last equality defines ${\hat\Delta}_{z;\sigma\sigma'\tau\tau'}({\bs k},t)$.
We only consider the equal time ($t=0$) pairing amplitudes from here on.
It is useful to form combinations of the pairing amplitudes
which are even $(+)$ or odd $(-)$ under exchange of orbital index as - \\
\begin{align}
\hat{\Delta}^{i \pm}_{z;\sigma\sigma'} ({\bs{k}}) =& \hat{\Delta}_{z;\sigma\sigma', 11}({\bs k}) \pm \hat{\Delta}_{z;\sigma\sigma', 22}({\bs k}) \nonumber \\
\hat{\Delta}^{I \pm}_{z;\sigma' \sigma} ({\bs{k}}) =& \hat{\Delta}_{z;\sigma\sigma', 12}({\bs k}) \pm \hat{\Delta}_{z;\sigma\sigma', 21}({\bs k})
\end{align}
where the superscript $i$ or $I$ on the LHS refers to intra-orbital and inter-orbital pairings.
Each of these $\hat{\Delta}$ is a $2 \times 2$ matrix in spin space and can be written as a sum of singlet and triplet components -
\begin{align}
\hat{\Delta} (\bs k) = i \sigma^y \psi({\bs k}) + i \sigma^y (\mathbf{d}({\bs k}) \cdot \boldsymbol{\sigma} )
\end{align}
where due to Fermi statistics, $\psi({\bs k})$ is even and $\mathbf{d}({\bs k})$ is odd under $ {\bs k} \rightarrow -{\bs k}$ for $i\pm$ and $I+$ pairings,
whereas, $\psi({\bs k})$ is odd and $\mathbf{d}({\bs k})$ is even under $ {\bs k} \rightarrow -{\bs k}$ for $I-$ pairings. Then, the even and odd intra-orbital and even interorbital amplitudes have the usual s-wave spin singlet and p-wave spin-triplet pairing
while the odd interorbital amplitude has a p-wave spin singlet and s-wave spin triplet pairing. Reference \cite{balatsky} found that an s-wave superconductor does not induce even frequency odd interorbital pairings in TI (although a p-wave superconductor can do so), however,
it does induce odd frequency odd interorbital pairings. In this work, since we are only interested in equal time correlations,
we ignore the even frequency odd interorbital and all odd frequency pairing amplitudes.
The Hamiltonian of $\text{Bi}_2\text{Se}_3$ - type
material considered here has tetragonal symmetry (since it is written on a cubic lattice) and the induced pairings
are classified in terms of the irreducible representations $\Gamma$ of the $D_{4h}$ group \cite{balatsky}.
Thus, $\psi({\bs k})$ and $\mathbf{d}({\bs k})$ must have a
functional dependence on ${\bs k}$, which forms an irreducible representation of $D_{4h}$.
Since we are only considering proximity with an $s$-wave superconductor in this paper and assuming that the angular momentum in the
${\hat z}$ direction is conserved in the
tunneling process, there are only three relevant representations - $A_{1g}$, $A_{1u}$ and $A_{2u}$.
Upto linear order in ${\bs k}$ (and taking $k_z = 0$) we have,
$\psi({\bs k}) = 1$ for $A_{1g}$, $ \mathbf{d}({\bs k}) = (k_x, k_y, 0) $ for $A_{1u}$ and $ \mathbf{d}({\bs k}) = (k_y, -k_x, 0) $ for $A_{2u}$.
On a square lattice, we may replace ${\bs k}^2$ by $1 - \cos ({\bs k})$ and terms linear in ${\bs k}$ by $\sin ({\bs k})$.
We can then classify the $ \hat{\Delta} ({\bs k}) $ found numerically,
by finding their inner product with the basis functions. For this, we define
\begin{align}
F^{a \pm}_{\sigma \sigma'} &= \frac{1}{2 N_{\bs{k}}} \sum_{{\bs k}} S^{*}_{\sigma \sigma'} ({\bs{k}}) \hat{\Delta}^{a \pm}_{\sigma' \sigma} ({\bs {k}})
\end{align}
where, $S_{\sigma \sigma'} ({\bs k})$ is one of the basis functions given above. The superscript $a=i,I$ refers to intra and inter-orbital pairing
respectively. Then the inner product is
\begin{align}\label{eq:F}
F^{a \pm} = \sum_{\sigma \sigma'} F^{a \pm}_{\sigma \sigma'}~.
\end{align}
We find that the spin singlet amplitudes have a dominant component with $A_{1g}$ pairing as expected, but the triplet components have $A_{2u}$ pairing and not $A_{1u}$ in both TI and WSM\cite{note1}. This is due to the form of the spin-momentum locked low energy Dirac surface state which enforces
the vanishing of the $A_{1u}$ triplet amplitude\cite{balatsky}.
We note that for the spin singlet, the odd intra-orbital pairing is lower by two orders of magnitude compared to the even orbital pairings.
Thus, only s-wave spin singlets with even orbital pairings are dominant.
For spin triplet, the even orbital pairings are lower by two orders of magnitude with respect to the odd intra-orbital pairing.
Thus, a p-wave spin triplet with odd intra-orbital pairing is dominantly induced.
Hence, for both the TI and the WSM, we only display the behavior of the following three amplitudes - spin singlet even intra- and inter-orbital
pairing and spin triplet odd intra-orbital pairing.
In both the TI and the WSM, the pairing amplitudes fall off exponentially in the bulk\cite{stanescu,klee}. The falloff can be numerically fit
to an exponential $F\propto e^{-\alpha z}$, where the direction $z$ is perpendicular to the surface of contact to the superconductor. In Fig.~\ref{fig:withz}(a),(b)we show how $\alpha$ varies for the various pairing amplitudes as a function of the chemical potential $\mu$ for TI and the WSM. We note that $\alpha$ starts to decrease as we increase $\mu$ in the case of TI, both for $s$ and $p$ wave amplitudes, which means a greater penetration~\cite{klee}.
For the case of WSM, the $s$-wave pairing has a decreased penetration with increasing $\mu$ in contrast to the TI, whereas the $p$-wave pairing
has mildly increasing penetration.
In Fig.~\ref{fig:withz}(c),(d) we compare how the various pairing amplitudes at $z=0$ for the TI and the WSM change as a function of
the chemical potential $\mu/t$. We note that the spin singlet amplitudes are higher than the spin triplet amplitudes in all cases. There is also
not much variation between the TI and the WSM as far as the the spin singlet amplitudes are concerned.
But the spin triplet amplitudes have substantially larger variation between the TI and the WSM.
We have also studied the behavior of these pairing amplitudes in both the TI and the WSM as a function of various other parameters.
The pairing amplitudes remain flat with time-reversal symmetry breaking perturbations $b_x$, as shown in Fig.~\ref{WSMparameter}(a).
In Fig.~\ref{WSMparameter}(b) $\epsilon/t$ parameterizes the flow from the TI to the WSM, which shows a switching from inter to intra--orbital pairing as one moves from
the TI to the WSM. We show the pairing amplitudes in Fig.~\ref{pairingmag} at $z=0$ as a function of the coupling
to the superconductor $\lambda_S = \pi N(0)({\tilde t})^2$ and
as a function of the the superconducting pairing amplitude $\Delta$ and in general, we see that all the pairing amplitudes increase as the parameters
increase.
\section{Surface states \& comparison with exact diagonalization}
\begin{figure}
\includegraphics[width=0.45\textwidth]{Fig8.pdf}
\caption{(Color online) Comparison of the Green's function technique that we used with exact diagonalization results for the effect of the proximity induced superconductivity in the surface bands of the WSM with momenta (a) $k_x$ and (b) $k_y$ where the Weyl nodes lie along $k_x$. The blue (gray) high density lines are the modified bands in the system with proximity to the superconductor obtained from the LDOS at $z=0$, while the red (darker) solid line is the surface band at $z=0$ via exact diagonalization. Note that we have only depicted the surface band at $z=0$ and have suppressed the
other surface state. The induced gap vanishes at the Weyl nodes for a large enough system size, but the surface band splits. The
various parameters that have been used for the LDOS are $\lambda_{\text{SO}}=\lambda_z=0.5 t$, $\Delta=0.7t$, $\lambda_S = 0.9t$, $\epsilon = 6t, ~ \bs{b}=(0.5t,0,0)$ and the number of sites in $z$ has been taken to be 20. We have used $k_y = 0$ for (a) and $k_x = 0$ for (b).}
\label{fig:surface}
\end{figure}
Finally, we discuss the effect of superconducting proximity on the surface states and we compare the results from the Green's function method with an exact diagonalization. After computing the self-energy of the electrons in the WSM due to the proximity effect as in Eq.~(\ref{eq:GF}), we can construct the effective band-structure by solving
\begin{align}\label{eq:eigval}
\mathrm{Det}\left[ H_{\mathrm{W}} + \Sigma(\omega) - \omega \right] = 0~,
\end{align}
which, in turn, is the equivalent of finding the peaks in the LDOS of the system. For the case of the TI, the surface state acquires an induced pairing which gaps the surface band completely in agreement with earlier results\cite{stanescu}, whereas for the surface state of the WSM, the induced gap is
much smaller and actually vanishes at the Weyl nodes. This is shown in Fig.~\ref{fig:surface}(a) and (b), where the effective
band structure of the WSM
in proximity with a superconductor is plotted
as a function of $k_x$ and $k_y$ respectively. The same band structure has also been plotted using exact
diagonalisation. Note that Fig.~\ref{fig:surface} is only the bandstructure at the surface in proximity to the superconductor. This is why the other surface state,
with opposite chirality is not visible here.
The first point to note is that there is no qualitative difference in the band structure using the Green's function technique
and using exact diagonalisation. This clearly justifies the approximation of using only the on-site or local component in
the self-energy.
Also, note that in contrast to the dispersion in Fig.~\ref{dispersion} without the proximity effect, we see here that the proximity
to the superconductor has split the flat band into two, giving rise to a small anisotropic gap. However, the states at the Weyl nodes are not gapped.
This is not unexpected because the $s$-wave
superconducting correlations couples the electrons at one node of a certain chirality to holes at the other node of the {\it same} chirality (because
the two nodes have opposite chirality, but the holes and electrons also have opposite chirality), hence, no gap can open up\cite{mengbalents}.
It is also of interest to consider the band structure as a function
of $k_y$ as shown in Fig.~\ref{fig:surface}(b). The edge states of the Chern insulator for each fixed value of $k_x$ between the nodes,
are now split by the proximity effect into two edge states, each carrying {\it half} the Chern number of the original edge state.
We also compare the LDOS computation with an exact diagonalization and as can be seen in the Figure, the results match quite well.
\section{Discussion and conclusions}
In summary, we have provided a detailed study of proximity induced superconductivity in Weyl
semimetals. We have focused on proximity of $s$-wave superconductor in the current work,
though a similar analysis can also be made for $p$-wave and $d$-wave superconductors.
We find that despite the presence of bulk metallic states in the WSM, the induced pairing
remains confined to a few layers close to the interface and in fact, falls off exponentially
fast away from the interface. We note that the $s$-wave superconductor induces both $s$-wave
and $p$-wave pairing, but the induced $p$-wave pairing is always smaller than the dominant
$s$-wave pairing. We also find that increasing the chemical potential does not increase either
the penetration into the bulk, or the ratio between the $p$-wave and the $s$ wave amplitudes
significantly. Both $s$-wave and $p$-wave components of the induced pairing can, however,
be increased by increasing the pairing amplitude in the superconductor or by increasing the
coupling to the superconductor.
\section*{Acknowledgments} U.K, S.P and S.R would like to thank J. D. Sau for useful discussions. Computational work for this study was carried out at the cluster computing facility in the Harish-Chandra Research Institute (http://www.hri.res.in/cluster). A.K was supported by the College of Arts and Science at Indiana University, Bloomington. Further funding was provided by the Offices of the Vice President for Research and the Vice Provost for Research at Indiana University through the Faculty Research Support Program. We would also like to thank the anonymous referee for comments, which were helpful in improving the clarity and presentation of the work.
\vspace{2cm}
|
\section{Introduction}\label{sec1}
In our work, we investigate 2-person zero-sum stochastic differential
games which dynamics\ are defined by a doubly controlled stochastic
differential equation (SDE)
\begin{eqnarray}
\label{0a}
dX_{s}^{t,x;u,v}&=&b
\bigl(s,X_{s}^{t,x;u,v},u_{s},v_{s}
\bigr)\,ds\nonumber\\
&&{}+ \sigma \bigl(s,X_{s}^{t,x;u,v},u_{s},v_{s}
\bigr)\,dB_s,\qquad s\in[t,T],
\\
X_{t}^{t,x;u,v}&=&x\in R^d,\nonumber
\end{eqnarray}
driven by a Brownian motion $B$, and endowed with pay-off functionals
defined through a doubly controlled backward stochastic differential
equation (BSDE) (see Section~\ref{sec2} for details) which, in the classical
case, reduces
to
\begin{equation}
\label{1a} I(t,x;u,v)=E \biggl[\Phi \bigl(X^{t,x;u,
v}_T \bigr)+
\int_t^Tf \bigl(s,X^{t,x;u,v}_s,u_s,v_s
\bigr)\,ds \biggr]
\end{equation}
[see (\ref{DPP-aaa})]. The initial data $(t,x)$ of the game belong to
$[0,T]\times R^d$, and the control processes $u=(u_s)$ and $v=(v_s)$
used by Players~1 and~2, take their values in compact
metric spaces $U$ and $V$, respectively. While the objective of Player
1 is
to maximize the pay-off $I(t,x;u,v)$, that of Player 2 is to minimize it:
Indeed, for Player 2 $I(t,x;u,v)$ represents a cost functional.
However, apart
from rather strong assumptions on the coefficients, for example, that
of independence of the controls $(u, v)$ and of strict
ellipticity for the diffusion coefficient $\sigma\sigma^*(t,x)\ge
\alpha
\cdot I_{R^d}, (t,x)\in[0,T]\times R^d$, for some $\alpha>0$ (refer to
Hamadene, Lepeltier, and Peng~\cite{HLP}), if one wants
to have a dynamic programming principle (DPP) the players can, in
general, not play a game of
the type ``control against control''; they can play, for instance, games
of the type ``nonanticipative strategy against control'' (see, e.g.,
\cite{FS,Buckdahn-Li-2008}) or games of the type ``NAD-strategy
against NAD-strategy'', where NAD stands for nonanticipativity
with delay (see, e.g., \cite{BCR04} and \cite{BCQ11}).
However, a central question in the theory of 2-person zero-sum stochastic
differential games is that of sufficient conditions, under which the
game admits a value, that is, under which the lower and the upper value
functions of the stochastic differential game coincide. In the literature,
since the famous works by Isaacs \cite{I} for the case of deterministic
differential
games and that by Fleming and Souganidis \cite{FS} for stochastic
differential games (see also \cite{FH}), various authors have shown
the equality between the lower
and the upper value functions under the so-called Isaacs condition.
Let us be more precise: Generalizing the pioneering paper on
stochastic differential games by Fleming and Souganidis \cite{FS},
Buckdahn and Li \cite{Buckdahn-Li-2008}, and also Buckdahn, Cardaliaguet
and Quincampoix \cite{BCQ11}, associated the dynamics (\ref{0a}) with
nonlinear cost functionals defined through a BSDE, which was first
introduced by Pardoux and Peng~\cite{PaPe1}:
\begin{equation}
\label{3a} \hspace*{14pt}\cases{
-dY^{t,x; u, v}_s = f
\bigl(s,X^{t,x;u,v}_s,Y^{t,x;u,v}_s,Z^{t,x;u,v}_s,
u_s,v_s \bigr)\,ds -Z^{t,x;u,v}_s
\,dB_s,& \vspace*{2pt}
\cr
Y^{t,x; u, v}_T = \Phi
\bigl(X^{t,x; u, v}_T \bigr),\qquad s\in[t,T].}\hspace*{-20pt}
\end{equation}
They considered as pay-off functional the random variable
(measurable with respect to the information available before the
beginning of the game)
\begin{equation}
J(t,x;u,v)=Y_t^{t,x; u,v},
\end{equation}
and the lower and the upper value functions for the game
over the time interval $[t,T]$ were introduced, respectively, by putting
\begin{eqnarray}
W(t,x)&:=& \esssup_{\alpha} \essinf_{\beta} J(t,x;
\alpha,\beta),
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
U(t,x)&:=& \essinf_{\beta}\esssup_{\alpha}J(t,x; \alpha,\beta)
\qquad (t,x)\in[0,T]\times R^d,
\end{eqnarray}
where $\alpha$ runs the NAD-strategies for Player 1 and $\beta$ those
for Player 2. Given such a couple of admissible NAD-strategies, the
cost functional $J(t,x;\alpha,\beta)$ is defined through the unique
couple of admissible controls $(u,v)$ satisfying $\alpha(v)=u,
\beta(u)=v$, by putting $J(t,x;\alpha,\beta)=J(t,x;u,v)$ (e.g., refer
to~\cite{BCQ11}). We emphasize
that in the above definition the classical case, where $f(s,x,y,z,u,v)=
f(s,x,u,v)$ is independent of $(y,z)$, can be obtained by replacing
$J(t,x;\alpha,\beta)$ by $E[J(t,x;\alpha,\beta)]=I(t,x;\alpha
,\beta)$
[see (\ref{1a})] and the essential supremum and the essential infimum
over a family of
random variables by the supremum and the infimum, respectively; this
does not change the upper and the lower value functions (see
Remark 3.4, \cite{Buckdahn-Li-2008}). The authors showed that, for the
Hamiltonians
\begin{eqnarray}
H(t,x,y,p,A,u,v)&=&\frac12 \operatorname{tr} \bigl(\sigma
\sigma^*(t,x,u,v)A \bigr) +b(t,x,u,v)p
\nonumber\\
& &{}+f \bigl(t,x,y,p\sigma(t,x,u,v),u,v \bigr),
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
H^-(t,x,y,p,A)&=&\sup_{u\in U}\inf
_{v\in V}H(t,x,y,p,A,u,v),
\\
H^+(t,x,y,p,A)&=&\inf_{v\in V}\sup_{u\in U}H(t,x,y,p,A,u,v),\nonumber
\end{eqnarray}
$(t,x,y,p,A)\in[0,T]\times R^d\times R\times R^d\times S^d$ ($S^d$ denotes the space of symmetric real matrices of the size
$d\times d$), $W$ and $U$ are the unique viscosity solutions of the
following Hamilton--Jacobi--Bellman--Isaacs (HJBI) equations in the class of
continuous functions with polynomial growth, respectively:
\begin{eqnarray}
\qquad\frac{\partial}{\partial t}W(t,x)+H^- \bigl(t,x, \bigl(W, \nabla
W,D^2W \bigr) (t,x) \bigr)&=&0,\qquad W(T,x)=\Phi(x),
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
\frac{\partial}{\partial t}U(t,x)+H^+ \bigl(t,x, \bigl(U,\nabla
U,D^2U \bigr) (t,x) \bigr)&=&0,\qquad U(T,x)=\Phi(x).
\end{eqnarray}
Isaacs condition
says that
\begin{eqnarray}
H^-(t,x,y,p,A)=H^+(t,x,y,p,A)
\nonumber
\\[-8pt]
\\[-8pt]
\eqntext{(t,x,y,p,A)\in[0,T] \times R^d\times
R\times R^d\times S^d,}
\end{eqnarray}
and under it the both above PDEs\vadjust{\goodbreak} coincide and the uniqueness
of the solution implies that $W(t,x)=U(t,x), (t,x)\in[0,T]\times R^d$,
that is, the game has a value.
But how to get a value, when Isaacs condition is not assumed?
Recently, in \cite{Buckdahn-Li_Quincampoix-2011} the authors
studied deterministic differential games without assuming Isaacs
condition. They considered an adequate notion of mixed strategies
related with a suitable randomization, and were thus able to prove
that such defined upper and lower value functions coincide, and that
this value function defined through mixed strategies
satisfies a Hamilton--Jacobi--Isaacs equation. We also refer to the works
of Chentsov, Krasovskii and Subbotin for the existence of the value
of deterministic differential games \cite{KRSU,SUC}: They studied the
problems of deterministic
differential games without Isaacs condition through positional
strategies but with techniques which differ from those in
\cite{Buckdahn-Li_Quincampoix-2011}. To the authors' best knowledge,
there does not exist any work on the existence of the value
of stochastic differential games without assuming Isaacs condition, it
has been an open problem until now. However, there are also different
recent works studying stochastic differential games without Isaacs'
condition, but without the objective to show the existence of a value
of the game. For instance, Krylov \cite{K1,K2} studied regularity
properties and the dynamic programming principle for the upper value
function of a stochastic differential game over a domain, by starting
from the Isaacs equation; for this he used the idea of \'{S}wi\c{e}ch
\cite{SA} that the viscosity solutions of nondegenerate Isaacs
equations have some regularity properties which can be used for the approach.
In the present work, our objective is to solve this open problem, that
is, to extend the results of
\cite{Buckdahn-Li_Quincampoix-2011} from deterministic
differential games without Isaacs condition to stochastic
differential games. Since this work was heavily inspired by \cite
{Buckdahn-Li_Quincampoix-2011}, we consider the game of the type
``NAD-straegies against
NAD-strategies''. The delay of the nonanticipative strategies
is defined through a partition $\pi=\{0=t_0<t<t_1<\cdots<
t_n=T\}$ of the time interval $[0,T]$. The underlying stochastic
controls for the both players are randomized along the
partition $\pi$ by a hazard which is independent of the
governing Brownian motion, and knowing all information
available at the left time point $t_{j-1}$ of the subintervals
generated by $\pi$, the controls of Players~1 and~2 are
conditionally independent over $[t_{j-1},t_j)$.
While the dynamics are defined by (\ref{0a}), the BSDE defining the pay-off
functional has to take into account that, first, the controls
of the both players are randomized by a hazard independent of the
governing Brownian motion, and second, the both players make
the randomization of their controls conditionally independent of
each other and reveal the information related with only
at the end of each subinterval generated by the partition $\pi$.
This has as consequence that the BSDE has to be considered under
a filtration $\widetilde{\mathbb{F}}^\pi$ which is smaller than the
filtration $\mathbb{F}^\pi$ (but larger than
the Brownian one) for the dynamics
(\ref{0a}); see BSDE (\ref{BSDE}).
With the help of the cost functional defined through our BSDE
we introduce the lower and the upper value functions along a partition
$\pi$, $W^\pi$ and
$U^\pi$. For these, a priori, random fields we prove that they are
deterministic and satisfy along the partition $\pi$, at its points,
the dynamic programming principle. This dynamic programming principle
combined with Peng's BSDE method, refer to Peng~\cite{Pe1}, which we
have to redevelop for our settings here is crucial for the proof that
$W^\pi$ and $U^\pi$ converge uniformly on compacts, as the mesh of
$\pi$ tends to zero, and their limit $V$, the so-called value in
mixed strategies can be characterized as the unique viscosity solution
of the Hamilton--Jacobi--Bellman--Isaacs equation
\begin{eqnarray}
\label{PDEa}
\hspace*{20pt}\frac{\partial}{\partial t}V(t,x)+\sup_{\mu\in{\cal P}(U)}
\inf_{\nu\in{\cal P}(V)}H \bigl(t,x, \bigl(V,\nabla V,D^2V \bigr)
(t,x),\mu,\nu \bigr) &=&0,
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
V(T,x)&=&\Phi(x),
\end{eqnarray}
where
\begin{eqnarray}
& &H(t,x,y,p,A,\mu,\nu)
\nonumber
\\
&&\qquad=\int_{U\times V} \biggl(\frac12 \operatorname{tr} \bigl(
\sigma \sigma^*(t,x,u,v)A \bigr)+b(t,x,u,v)p
\\
& &\hspace*{58pt}\qquad\quad{}+f \bigl(t,x,y,p\sigma(t,x,u,v),u,v \bigr) \biggr)\mu\otimes
\nu(du\,dv),
\nonumber
\end{eqnarray}
$(t,x,y,p,A)\in[0,T]\times R^d\times R\times R^d\times S^d$. Here
${\cal P}(U)$ denotes the space of all probability measures
on $U$, ${\cal P}(V)$ all on $V$. Since both control state spaces
$U$ and $V$ are supposed to be compact and metric, ${\cal P}(U)$
and ${\cal P}(V)$ are convex and compact, and from the bi-linearity
of $H(t,x,y,p,A,\mu,\nu)$ in $(\mu,\nu)$ we have that for PDE
(\ref{PDEa}) the following Isaacs condition is automatically satisfied:
\begin{eqnarray}
&&\sup_{\mu\in{\cal P}(U)}\inf_{\nu\in{\cal P}(V)}
H(t,x,y,p,A,\mu,\nu)
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&&\qquad=\inf_{\nu\in{\cal P}(V)}\sup_{\mu\in{\cal P}(U)}
H(t,x,y,p,A,\mu, \nu).
\end{eqnarray}
Of course, PDE (\ref{PDEa}) could have been also derived by considering
weak controls, that is, controls with values in ${\cal P}(U)$ and
${\cal P}(U)$,
but our objective has been to work with controls taking values in $U$
and $V$,
respectively, even for the price of a randomization.
Let us point out that the fact that, in our approach, the dynamics and
the BSDE have to be studied under different filtration, means that
unlike in \cite{Buckdahn-Li-2008} and \cite{BCQ11} we are not anymore in
a Markovian framework here for our BSDE. This requires new
approaches, not only for the redevelopment of Peng's BSDE method~\cite{Pe1}
in our settings (Section~\ref{sec4}), but also for the proof that the upper
and the lower value functions are deterministic and H\"{o}lder
continuous with respect to the time parameter.
Let us explain the organization of the paper. In Section~\ref{sec2}, we
introduce the settings for our stochastic differential games,
we define for both players the space of admissible controls
along a partition $\pi$ as well as the notion of NAD-strategies
with respect to $\pi$. Moreover, we introduce the dynamics, the
pay-off functional defined through a BSDE, as well as the upper and
the lower value functions $W^{\pi}$ and $U^{\pi}$ along $\pi$.
In Section~\ref{sec3}, we study properties of $W^{\pi}$ and $U^{\pi}$. We
show, in particular, that they are deterministic continuous
functions which, with respect to the points of the partition
$\pi$, satisfy the dynamic programming principle. In Section~\ref{sec4}, finally, it is shown that, as the mesh of $\pi$ tends to zero,
$W^{\pi}$ and $U^{\pi}$ converge uniformly on compacts to the unique
viscosity solution of the associated Hamilton--Jacobi--Bellman--Isaacs
equation. For this, Peng's BSDE method is redeveloped for our
settings.
\section{Preliminaries. Settings of the stochastic differential games}\label{sec2}
Let us begin with introducing the probability space
\[
(\Omega_1,{\cal F}_1,P_1):= \bigl(
\bigl(R^2 \bigr)^{\mathbb{N}},{\cal B} \bigl(R^2
\bigr)^{\otimes\mathbb{N}}, Q_2^{\otimes\mathbb{N}} \bigr),
\]
where $Q_2$ denotes the two-dimensional standard Normal
distribution on the real plane $R^2$ endowed with its Borel
$\sigma$-field ${\cal B}(R^2)$, and $\mathbb{N}$ is the set of all
positive integers. Then, by the above definition,
$\Omega_1=(R^2)^{\mathbb{N}}$ is the space of all $R^2$-valued
sequences $\rho=(\rho_j=(\rho_{j,1},\rho_{j,2}))_{j\ge1}$, and
${\cal F}_1={\cal B} (R^2)^{\otimes\mathbb{N}}$ is the product Borel
$\sigma$-field taken over the sequence of $\sigma$-fields, which all
elements coincide with ${\cal B} (R^2)$, and
$P_1=Q_2^{\otimes\mathbb{N}}$ is the product measure over
$(\Omega_1,{\cal F}_1)$. Let us denote the coordinate mappings on
$\Omega_1$ by $\zeta_j=(\zeta_{j,1},\zeta_{j,2})\dvtx \Omega
_1\rightarrow
R^2$, $j\ge1$:
\[
\zeta_j(\rho)= \bigl(\zeta_{j,1}(\rho),
\zeta_{j,2}(\rho) \bigr)= (\rho_{j,1},\rho_{j,2}),
\qquad \rho= \bigl((\rho_{j,1},\rho_{j,2})
\bigr)_{j\ge1} \in\Omega_1.
\]
We observe that ${\cal F}_1$ coincides with the smallest
$\sigma$-field on $\Omega_1$, with respect to which all coordinate
mappings $\zeta_j,j\ge1$, are measurable.
However, for the study of our stochastic differential games we also
need the classical Wiener space $(\Omega_2,{\cal F}_2,P_2)$, where
$\Omega_2$ is the set of all continuous functions from $[0,T]$ with
values in $R^d$ and starting from zero, endowed with the supremum
norm [i.e., $\Omega_2=C_0( [0,T];R^d)$], and ${\cal F}_2$ is the Borel
$\sigma$-field on $\Omega_2$ completed with respect to the Wiener
measure $P_2$ under which the coordinate process
$B_t(\omega')=\omega'(t), t\in[0,T], \omega'\in\Omega_2$, is a
Brownian motion.
Let us denote by $(\Omega,{\cal F},P)$ the product probability space
\[
(\Omega,{\cal F},P)=(\Omega_1,{\cal F}_1,P_1)
\otimes (\Omega_2, {\cal F}_2,P_2),
\]
which we complete with respect to the probability measure
$P$, and let us extend the coordinate mappings $\zeta$ and $B$ in a
canonical way from $\Omega_1$ and $\Omega_2$, respectively, to~$\Omega$:
\begin{eqnarray}
\zeta_j(\omega):=\zeta_j(\rho),\qquad B_t(
\omega):=B_t\bigl(\omega'\bigr)=\omega'(t),
\nonumber\\
\eqntext{\omega=\bigl(\rho,\omega'\bigr)\in\Omega=\Omega_1
\times\Omega_2, j\ge1, t\in[0,T].}
\end{eqnarray}
Let us now introduce the filtration with which we work on our
probability space $(\Omega,{\cal F},P)$. By $\mathbb{F}^B=({\cal
F}^B_t)_{t\in[0,T]}$ we denote the filtration generated by the
Brownian motion $B$ and completed by all $P$-null sets. In addition
to the filtration $\mathbb{F}^B$, we also need larger ones, defined
along a partition $\pi=\{0=t_0<t_1<\cdots<t_n=T\}$ of the interval
$[0,T]$. Given such a partition $\pi$, we define $\mathbb{F}^{\pi,i}=
({\cal F}^{\pi,i}_t)_{t\in[0,T]}$, with
\[
{\cal F}^{\pi,i}_t={\cal F}^B_t\vee
\sigma \bigl\{\zeta_\ell= (\zeta_{\ell,1}, \zeta_{\ell,2})
(1\le\ell\le j-1), \zeta_{j,i} \bigr\},
\]
$t\in[t_{j-1},t_j)$, $1\le j\le n, i=1,2$, and we put ${\cal F}^{\pi,i}_T={\cal F}^{\pi,i}_{T-},
i=1,2$. Notice that, for $j=1$, that is, on the time interval
$[t_0,t_1)$, by convention, ${\cal F}_t^{\pi,i}={\cal F}_t^B\vee
\sigma\{
\zeta_{1,i}\}, i=1, 2$. We shall also introduce the filtration
$\mathbb{F}^{\pi}=\mathbb{F}^{\pi,1}\vee\mathbb{F}^{\pi
,2}=({\cal
F}^{\pi}_t={\cal F}^{\pi,1}_t\vee{\cal F}^{\pi,2}_t)_{t\in[0,T]}$,
and we remark that, for $t\in[t_{j-1},t_j)$,
\[
{\cal F}^{\pi}_t={\cal F}^B_t\vee{
\cal H}_j\qquad \mbox{where } {\cal H}_j: = \sigma
\bigl\{\zeta_\ell= (\zeta_{\ell,1}, \zeta_{\ell,2}) (1\le
\ell\le j) \bigr\}.
\]
Finally, we will also need a smaller filtration,
$\widetilde{\mathbb{F}}^\pi=(\widetilde{\cal
F}^{\pi}_t)_{t\in[0,T]}$ with $\widetilde{\cal F}^{\pi}_t:={\cal
F}^B_t\vee{\cal H}_{j-1}$, for $t\in[t_{j-1},t_j), 1\le j\le n$.
Observe that, for all $t\in[t_{j-1},t_j)$, knowing $\widetilde{\cal
F}^\pi_t={\cal F}_t^B\vee{\cal H}_{j-1}$, the $\sigma$-fields ${\cal
F}^{\pi,1}_t$ and ${\cal F}^{\pi,2}_t$ are conditionally
independent.
Let us consider two compact metric spaces $U$ and $V$ as control
state spaces used by the Players 1 and 2, respectively. By ${\cal
P}(U)$ and ${\cal P}(V)$, we denote the space of all probability
measures over $U$ and $V$, endowed with its Borel
$\sigma$-field ${\mathcal{B}}(U)$ and ${\mathcal{B}}(V)$, respectively.
We also observe that it is an immediate consequence
of Skorohod's Representation theorem that the set ${\cal P}(U)$
[resp., ${\cal P}(V)$] coincides with the set of the laws of all
$U$-valued (resp., $V$-valued) random variables defined over
$([0,1], {\cal B}([0,1]),\lambda_1)$ [$\lambda_1$ denotes the
Lebesgue measure on $([0,1], {\cal B}([0,1]))$]. But this latter set
coincides with that of the laws of all random variables defined over
$(R, {\cal B}(R),Q_1)$, where $Q_1$ denotes the standard Normal
distribution over $(R, {\cal B}(R))$. Indeed, denoting by
\[
\Phi_{0,1}(x)=\frac{1}{\sqrt{2\pi}}\int_{-\infty}^x
\exp \biggl\{- \frac{y^2}{2} \biggr\}\,dy,\qquad x\in R,
\]
we have that, for any random variable $\xi$ over $([0,1],
{\cal B}([0,1]),\lambda_1)$, the law of $\xi$ with respect to
$\lambda_1$ coincides with that of $\xi(\Phi_{0,1}(\cdot))\dvtx R\rightarrow
R$ under
$Q_1$. A~consequence is that
\[
{\cal P}(U)=\bigl\{P_\xi\dvtx \xi \mbox{ is } U\mbox{-valued random
variable over } \bigl(\Omega,\sigma\{\zeta_{j,1}\}, P\bigr) \bigr\}
\]
and
\[
{\cal P}(V)= \bigl\{P_\xi\dvtx \xi \mbox{ is } V\mbox{-valued random
variable over } \bigl(\Omega,\sigma\{\zeta_{j,2}\}, P\bigr) \bigr\}
\]
for all $j\ge1.$
Let us now introduce the admissible controls for both players along a
given partition $\pi=\{0=t_0<t_1<\cdots<t_n=T\}$ of the time interval
$[0,T]$.
\begin{definition}[(Admissible controls)]\label{adm.control} Given a
partition $\pi$ of the time interval $[0,T]$ and an initial time
$t\in[0,T]$, the space of admissible controls along the partition
$\pi$ for Player 1 for a game over the time interval $[t,T]$ is the
totality of all $U$-valued $\mathbb{F}^{\pi,1}$-predictable
processes $u=(u_s)_{s\in[t,T]}$ defined over the probability space
$(\Omega, {\cal F},P)$; it is denoted by ${\cal U}_{t,T}^\pi$. For
Player 2 the space of admissible controls along the partition $\pi$
${\cal V}_{t,T}^\pi$ is defined similarly: It is the
collection of all $V$-valued $\mathbb{F}^{\pi,2}$-predictable
processes $v=(v_s)_{s\in[t,T]}$ defined over $(\Omega,{\cal F},P)$.
\end{definition}
After having introduced the spaces of admissible controls, we describe
now the dynamics of our stochastic differential games. For this, we
consider the coefficients
\[
b\dvtx [0,T]\times R^d\times U\times V\rightarrow R^d
\quad\mbox{and}\quad\sigma\dvtx [0,T]\times R^d\times U\times V
\rightarrow R^{d\times d}
\]
which we suppose throughout our work to be bounded,
jointly continuous and Lipschitz in $x\in R^d$, uniformly with
respect to $(t,u,v)\in[0,T]\times U\times V$. Let $\pi$ be a
partition of the time interval $[0,T]$. Then, given arbitrary
initial data $t\in[0,T]$ and $\vartheta\in L^2(\Omega,{\cal F}_t^\pi,
P;R^d)$ as well as admissible control processes $u\in{\cal U}_{t,
T}^\pi$ and $v\in{\cal V}_{t,T}^\pi$, we consider the SDE
\begin{eqnarray}
\label{SDE}
dX_{s}^{t,\vartheta;u,v}&=&b
\bigl(s,X_{s}^{t,\vartheta;u,v},u_{s},v_{s}
\bigr)\,ds+ \sigma \bigl(s,X_{s}^{t,\vartheta;u,v},u_{s},v_{s}
\bigr)\,dB_s
\nonumber
\\[-8pt]
\\[-8pt]
\eqntext{s\in[t,T], X_{t}^{t,\vartheta;u,v}=\vartheta.}
\end{eqnarray}
Under our assumptions on the coefficients $b$ and $\sigma$,
this SDE has a unique strong solution $X^{t,\vartheta;u,v}=(X_{s}^{t,
\vartheta;u,v})_{s\in[t,T]}$ in the space of $R^d$-valued,
$\mathbb{F}^\pi$-adapted continuous processes. Moreover, we have the
following estimates which are by now standard.
For all $p\ge2$, there exists some constant $C_p\in R$ (only
depending on p, on the Lipschitz constants and the bounds of $b$ and
$\sigma$) such that, for all partitions $\pi$ of $[0,T]$, for all
$t\in[0,T], \vartheta,\vartheta'\in L^2(\Omega,{\cal F}_t^\pi,P;
R^d)$ and
all $u\in{\cal U}^{\pi}_{t,T},v\in{\cal V}^{\pi}_{t,T}$, it holds,
$P$-a.s.,
\begin{eqnarray}
E \Bigl[\sup_{s\in[t,T]}\bigl|X^{t,\vartheta; u, v}_s
-X^{t,\vartheta';u,v}_s\bigr|^p| {{\mathcal{F}}_t^\pi}
\Bigr] & \leq& C_p\bigl|\vartheta-\vartheta'\bigr|^p,
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
E \Bigl[ \sup_{s\in[t,T]} \bigl|X^{t,\vartheta;u,v}_s\bigr|^p|{{
\mathcal{F}}_t^\pi } \Bigr] & \leq& C_p
\bigl(1+|\vartheta|^p \bigr).
\end{eqnarray}
Let us now come to the pay-off functional which we associate with
the above dynamics of our game. The pay-off functional is a nonlinear,
recursive one, that is, we define it through a backward
stochastic differential equation. For this, we consider the
terminal pay-off function $\Phi\dvtx R^d\rightarrow R$ which we suppose
to be bounded and Lipschitz, as well as the running pay-off function
$f\dvtx [0,T]\times R^d\times R\times R^d\times U\times V\rightarrow R$
which we assume to be jointly continuous and such that
\begin{longlist}[(iii)]
\item[(i)] $f(t,x,y,z,u,v)$ is Lipschitz in $(x,y,z)\in R^d\times
R\times R^d$, uniformly in $(s,u,v)\in[0,T]\times U\times V$;
\item[(ii)] $f(t,x,y,z,u,v)$ is uniformly continuous on $[0,T]\times R^d
\times R\times\overline{B}_K(0)\times U\times V$, for all $K>0$,
where $\overline{B}_K(0)$ denotes the closed ball in $R^d$ centered at $0$
with diameter $K$;
\item[(iii)] $(t,x,y,u,v)\rightarrow f(t,x,y,0,u,v)$ is bounded.
\end{longlist}
Given a partition $\pi$ of the interval $[0,T]$, initial data
$t\in[0,T], \vartheta\in L^2(\Omega,{\cal F}^\pi_t,\break P;R^d)$ and
admissible controls $u\in{\cal U}^{\pi}_{t,T}, v\in{\cal
V}^{\pi}_{t,T}$, we consider the following BSDE governed by the
solution $X^{t,\vartheta;u,v}$ of SDE (\ref{SDE}):
\begin{equation}
\label{BSDE} \cases{dY^{t,\vartheta; u, v}_s = -E \bigl[f
\bigl(s,X^{t,\vartheta; u, v}_s, Y^{t,
\vartheta; u, v}_s,
Z^{t,\vartheta; u, v}_s, u_s, v_s \bigr) |
\widetilde{\cal F}^\pi_s \bigr] \,ds \vspace*{2pt}\cr
\hspace*{54pt}{}+Z^{t,\vartheta;u, v}_s
\,dB_s+dM^{t,\vartheta; u, v}_s, \vspace*{2pt}
\cr
Y^{t,\vartheta; u, v}_T = E \bigl[\Phi \bigl(X^{t,\vartheta; u, v}_T
\bigr) | \widetilde{\cal F}^\pi_{T} \bigr],}
\end{equation}
where $(E[\gamma_s | \widetilde{\cal
F}^\pi_s])_{s\in[0,T]}$ is understood as
$\widetilde{\mathbb{F}}^\pi$-optional projection of integrable,
measurable processes $\gamma=(\gamma_s)_{s\in[0,T]}$.
We say that $(Y^{t,\vartheta; u, v},Z^{t,\vartheta; u, v},
M^{t,\vartheta; u, v})$ is a solution of this BSDE, if
\begin{longlist}[(iii)]
\item[(i)] $Y^{t,\vartheta; u, v}\in{\cal
S}^2_{\widetilde{\mathbb{F}}^\pi}(t,T;R)$, that is, $Y^{t,\vartheta
; u,
v}=(Y^{t,\vartheta; u, v}_s)_{s\in[t,T]}$ is an
${\widetilde{\mathbb{F}}^\pi}$-adapted c\`{a}dl\`{a}g process which
is square integrable: $E [\sup_{s\in[t,T]}|Y^{t,\vartheta; u,
v}_s|^2 ]<+\infty$;
\item[(ii)] $Z^{t,\vartheta; u, v}\in L_{\widetilde{\mathbb{F}}^\pi}^2(t,T;
R^d)$, that is, $Z^{t,\vartheta; u, v}=(Z^{t,\vartheta; u,
v}_s)_{s\in[t,T]}$
is an $R^d$-valued, ${\widetilde{\mathbb{F}}^\pi}$-predictable
process such that $E [\int_t^T|Z^{t,\vartheta; u,
v}_s|^2\,ds ]<+\infty$;
\item[(iii)] $M^{t,\vartheta; u, v}\in{\cal M}_{\widetilde{\mathbb{F}}^\pi}^2
(t,T;R)$, that is, $M^{t,\vartheta; u, v}=(M^{t,\vartheta; u,
v}_s)_{s\in[t,T]}$ is a square integrable
${\widetilde{\mathbb{F}}^\pi}$-martingale with $M^{t,\vartheta; u,
v}_t=0$. Moreover, $M^{t,\vartheta; u, v}$ is supposed to be orthogonal
to the driving Brownian motion $B$, that is, their joint quadratic
variation process satisfies $[B,M^{t,\vartheta; u, v}]_s=0,
s\in[t,T]$. For the proof of the existence and the uniqueness of the
solution of such BSDE (\ref{BSDE}) it is similar to the classical case,
see also \cite{CFS} and references inside.
\end{longlist}
We have to emphasize here that since the filtration
${\widetilde{\mathbb{F}}^\pi}$ is not the Brownian one, but contains
it strictly, we cannot expect to have a solution of the above BSDE
with vanishing $M^{t,\vartheta; u, v}$. It is by now well known that,
under our assumptions on the coefficients $f$ and $\Phi$, a BSDE of
the above type has a unique solution $(Y^{t,\vartheta; u,
v},Z^{t,\vartheta;
u, v},M^{t,\vartheta; u, v})$. Moreover, considering the special form of
the filtration ${\widetilde{\mathbb{F}}^\pi}$, we can characterize
this solution as follows.
\begin{remark}\label{remark_bsde}We first observe that on each of the
subintervals $[t_{j-1},t_j)$, $
1\le j\le n$, formed by the partition $\pi=\{0=t_0<t_1<\cdots
<t_n=T\}$, the filtration $\widetilde{\mathbb{F}}^\pi$ coincides
with the Brownian one $({\cal F}^B_s)_{s\in[t_{j-1},t_j)}$ augmented
by the independent $\sigma$-field ${\cal H}_{j-1}$. Hence, on the
interval $[t_{j-1},t_j)$ we have the martingale representation
property for random variables from $L^2(\Omega,\widetilde{\cal
F}_{t_{j}-}^\pi,P)$ with respect to the
$\widetilde{\mathbb{F}}^\pi$-Brownian motion $B$. This has as
consequence that BSDE (\ref{BSDE}) can be solved over the time
intervals $[t_{j-1},t_j)$ with $dM^{t,\vartheta;u,v}_s=0, s\in[t_{j-1},
t_j)$. However, for this $Y^{t,\vartheta;u,v}_{t_{j}-}$ has to be determined
by backward iteration. In order to compute $Y^{t,\vartheta;u,v}_{t_{n}-}$,
we determine from BSDE (\ref{BSDE}) the jump of the c\`{a}dl\`{a}g
process $Y^{t,\vartheta; u, v}$ at time $t_{n}$:
\begin{eqnarray}
\triangle Y^{t,\vartheta; u, v}_{t_{n}} \bigl(:=Y^{t, \vartheta;
u, v}_{t_{n}}-Y^{t,\vartheta; u, v}_{t_{n}-}
\bigr)&=& \triangle M^{t,\vartheta;u, v}_{t_{n}}\nonumber\\
\eqntext{\mbox{that is } Y^{t,\vartheta; u, v}_{t_{n}-}=Y^{t,\vartheta; u, v}_{t_{n}}- \triangle
M^{t,\vartheta; u, v}_{t_{n}}.}
\end{eqnarray}
Taking into account that $ M^{t,\vartheta; u, v}$ is an
$\widetilde{\mathbb{F}}^\pi$-martingale, this yields
\[
Y^{t,\vartheta; u, v}_{t_{n}-}=E \bigl[Y^{t,\vartheta; u,
v}_{t_{n}} |
\widetilde{\cal F}_{t_{n}-}^\pi \bigr]\quad\mbox{and}\quad
\triangle M^{t,\vartheta; u, v}_{t_{n}}=Y^{t,\vartheta; u,
v}_{t_{n}}-E
\bigl[Y^{t,\vartheta; u, v}_{t_{n}} | \widetilde{\cal F}_{t_{n}-}^\pi
\bigr].
\]
Having now $Y^{t,\vartheta; u, v}_{t_{n}-}\in L^2(\Omega,
\widetilde{\cal F}_{t_{n}-}^\pi,P)$, we can consider BSDE
(\ref{BSDE}) over the time interval $[t_{n-1},t_{n})$ like a
classical one, with $dM^{t,\vartheta;u,v}_s=0,
s\in[t_{n-1},t_{n})$. By slving this BSDE over $[t_{n-1},t_{n})$,
we get, in particular, $Y^{t,\vartheta; u, v}_{t_{n-1}}$.
Iterating this argument, we see that
\[
Y^{t,\vartheta; u, v}_{t_{j}-}=E \bigl[Y^{t,\vartheta; u,
v}_{t_{j}} |
\widetilde{\cal F}_{t_{j}-}^\pi \bigr]\quad\mbox{and}\quad
\triangle M^{t,\vartheta; u,
v}_{t_{j}}=Y^{t,\vartheta; u, v}_{t_{j}}-E
\bigl[Y^{t,\vartheta; u,
v}_{t_{j}} | \widetilde{\cal F}_{t_{j}-}^\pi
\bigr],
\]
for all $t_j> t$, and $M^{t,\vartheta; u, v}$ is constant in
the intervals $[t_{j-1}\vee t,
t_j)$, $1\le j\le n$.
\end{remark}
\begin{remark}\label{classical} In the classical case, where the running
payoff function
$f(s,x,y,z,u,v)$ does not depend on $y$ and on $z$, the solution
$Y^{t,\vartheta;u,v}$ of BSDE~(\ref{BSDE}) takes the simple,
well-known form
\begin{eqnarray}
Y^{t,\vartheta;u,v}_s=E \biggl[\Phi \bigl(X^{t,\vartheta;u,
v}_T
\bigr)+\int_s^Tf \bigl(r,X^{t,\vartheta;u,v}_r,u_r,v_r
\bigr)\,dr | \widetilde {\cal F}_s^\pi \biggr], \nonumber\\
\eqntext{s
\in[t,T], x\in R^d.}
\end{eqnarray}
From standard estimates for BSDEs of the type of equation
(\ref{BSDE}) we get, for all $p\ge2$, the existence of some
constant $C_p$ depending only $p$ and on the Lipschitz constants and
the bounds of the coefficients, such that, for all partitions $\pi$,
all initial\vadjust{\goodbreak} data $t\in[0,T], \vartheta,\vartheta'\in L^2(\Omega
,{\cal
F}^\pi_t,P;R^d)$ and all $u\in{\cal U}_{t,T},v\in{\cal V}_{t,T}$ it
holds, $P$-a.s.,
\begin{eqnarray}
\label{BSDE-estimates-1}
\mathrm{(i)} &&\quad\bigl |Y^{t,\vartheta; u, v}_s
\bigr| \le C_p,\qquad s\in[t,T];
\nonumber
\\
\mathrm{(ii)}&&\quad E \biggl[ \biggl(\int_t^T
\bigl|Z^{t,\vartheta; u, v}_s \bigr|^2\,ds \biggr)^{p/2} \Big|
\widetilde{\cal F}_t^\pi \biggr]\le C_p;
\nonumber\\
\mathrm{(iii)}&&\quad E \biggl[\sup_{s\in[t,T]}
\bigl|Y^{t,\vartheta; u, v}_s- Y^{t,\vartheta'; u,v}_s \bigr|^p\\
&&\hspace*{22pt}{}+ \biggl(\int_t^T \bigl|Z^{t,\vartheta; u, v}_s-Z^{t,\vartheta'; u,
v}_s\bigr|^2
\,ds \biggr)^{p/2} \Big| \widetilde{\cal F}_t^\pi
\biggr]\nonumber\\
&&\quad\qquad \le C_p E \bigl[ \bigl|\vartheta-\vartheta'
\bigr|^p | \widetilde{\cal F}_t^\pi \bigr];
\nonumber
\end{eqnarray}
from where, in particular, for some constant $C\in R$,
\begin{eqnarray}
\label{BSDE-estimates-2}
\mathrm{(i)}&&\quad \bigl|Y^{t,\vartheta; u, v}_t
\bigr| \le C,\qquad P\mbox{-a.s.};
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
\mathrm{(ii)}&&\quad \bigl|Y^{t,\vartheta; u, v}_t-Y^{t,\vartheta'; u, v}_t
\bigr| \le C \bigl(E \bigl[ \bigl|\vartheta-\vartheta' \bigr|^2 |
\widetilde{\cal F}_t^\pi \bigr] \bigr)^{1/2},
\qquad P\mbox{-a.s.}
\end{eqnarray}
For a game, in which the both Players 1 and 2 play along a partition
$\pi$ over a time interval $[t,T]$ and use the admissible controls
$u\in{\cal U}_{t,T}^\pi$ and $v\in{\cal V}_{t,T}^\pi$, we consider
the following pay-off functional:
\[
J^\pi(t,x;u,v)=Y^{t,x; u, v}_t,\qquad (t,x)\in[0,T]
\times R^d, (u,v)\in{\cal U}_{t,T}^\pi\times{
\cal V}_{t,T}^\pi.
\]
However, if we want to study the stochastic differential
game in a general frame, we can not consider games of the type
``control against control'', but we shall study games with
nonanticipative strategies with delay; for a more detailed
discussion the reader is referred to, for example, \cite{BCQ11}.
\end{remark}
Let us introduce the notion of nonanticipative strategies with
delay (NAD-strategies). They differ from the definitions given in
\cite{BCR04} and in \cite{BCQ11} and follow rather the spirit of the
definition given in \cite{Buckdahn-Li_Quincampoix-2011}, but now
extended to the stochastic case.
\begin{definition}[(NAD-strategies along the partition
$\pi$)] \label{NAD}
Let $\pi=\{0=t_0<t_1<\cdots<t_n=T\}$ $(n\ge1)$ an arbitrary partition
of the time interval $[0,T]$ and $t\in[0,T]$. We say that a mapping
$\beta\dvtx
{\cal U}_{t,T}^\pi\longrightarrow{\cal V}_{t,T}^\pi$ is an NAD-strategy
for Player~2 for the game over the time interval $[t,T]$ along the partition
$\pi$, if:
\begin{longlist}[(ii)]
\item[(i)] For all ${\widetilde{\mathbb{F}}^\pi}$-stopping times
$\tau\dvtx \Omega\rightarrow\pi=\{t_0,t_1,\ldots,t_n\}$ it holds:
Whenever two controls $u,u'\in{\cal U}_{t,T}^\pi$ coincide $ds\,dP$-a.e.
on the stochastic interval $[[t,\tau]]$, then also
$\beta(u)_s=\beta(u')_s, ds\,dP$-a.e. on $[[t,\tau]]$.
\item[(ii)] For all $0\le j\le n-1$, it holds that, whenever two
controls $u,u'\in{\cal U}_{t,T}^\pi$ coincide $ds\,dP$-a.e. on $[t,t_j]
\times\Omega$, then also $\beta(u)_s=\beta(u')_s, ds\,dP$-a.e. on $[t,
t_{j+1}]\times\Omega$.\vadjust{\goodbreak}
\end{longlist}
The set of all NAD-strategies for Player 2 over $[t,T]$ along
the partition $\pi$ is denoted by ${\cal B}_{t,T}^\pi$.
In an obvious symmetric way we define for Player 1 his set ${\cal
A}_{t, T}^\pi$ of NAD-strategies $\alpha\dvtx {\cal V}_{t,T}^\pi
\longrightarrow{\cal U}_{t,T}^\pi$ over the interval $[t,T]$ along
the partition $\pi$.
\end{definition}
Unlike the definitions in \cite{BCR04} and \cite{BCQ11}, the
delays for which we have this NAD-property (ii) in the above
definition is not considered as arbitrarily small for a given
partition $\pi$, but they are defined by the partition $\pi$. But,
however, in what follows we will study our game as the mesh of the
partition $\pi$ tends to zero.
The following result is crucial and it links our games defined
through a couple of admissible controls with those defined through
NAD-strategies.
\begin{lemma}\label{controls-NAD-strategies}
Let $\pi$ be any partition of the interval $[0,T]$ and $t\in[0,T]$.
Then, for all couples of NAD-strategies $(\alpha,\beta)\in{\cal
A}_{t,T}^\pi\times{\cal B}_{t,T}^\pi$, there is a unique couple of
admissible controls $(u,v)\in{\cal U}_{t,T}^\pi\times{\cal
V}_{t,T}^\pi$ such that $\alpha(v)=u$ and $\beta(u)=v$, $ds\,dP$-a.e.
on $[t,T]\times\Omega$.
\end{lemma}
In the above cited references \cite{BCR04,BCQ11} and
\cite{Buckdahn-Li_Quincampoix-2011} different definitions of
NAD-strategies were given, but the idea of the proof of the above
lemma remains similar. However, let us give it for the convenience of
the reader.
\begin{pf} Let $\pi=\{0=t_0<t_1<\cdots<t_n=T\}$ be a partition of
the interval $[0,T]$, and $(\alpha,\beta) \in\mathcal{A}_{t,T}^{\pi}
\times\mathcal{B}_{t, T}^{\pi}$. Let $t\in[t_i,t_{i+1})$. Then, due
to our definition of NAD strategies, $\alpha(v),\beta(u)$
restricted to $[t,t_{i+1}]$ depend only on $v\in\mathcal{V}_{t,
T}^{\pi}$ and $u\in\mathcal{U}_{t, T}^{\pi}$ restricted to the
interval $[t,t_i]$. But this interval is empty or a singleton, so
that $\alpha(v),\beta(u)$ restricted to the $[t,t_{i+1}]$ do not
depend on $v$ and $u$, respectively. Thus, putting for arbitrary
$u^0\in\mathcal{U}_{t, T}^{\pi},v^0\in\mathcal{V}_{t,T}^{\pi}$,
$u^1:=\alpha(v^0),v^1:=\beta(u^0)$, we get
\[
\alpha\bigl(v^1\bigr)=u^1, \qquad \beta
\bigl(u^1\bigr)=v^1\qquad ds\,dP\mbox{-a.s. on }
[t,t_{i+1}].
\]
Let us suppose now that we have constructed, for $j\ge2$,
$(u^{j-1}, v^{j-1})\in\mathcal{U}_{t, T}^{\pi}\times\mathcal{V}_{t,
t_{l}}^{\Pi}$ such that $\alpha(v^{j-1})=u^{j-1}$ and
$\beta(u^{j-1})=v^{j-1}$, $ds\,dP$-a.s. on $[t,t_{i+j-1}]$. Then we
set $u^{j}:=\alpha(v^{j-1}), v^{j}:=\beta(u^{j-1})$, and,
obviously, $(u^{j},v^{j})\in\mathcal{U}_{t,
T}^{\pi}\times\mathcal{V}_{t, T}^{\pi}$ is such that
$(u^{j},v^{j})=(u^{j-1},v^{j-1})$, $ds\,dP$-a.s. on $[t,t_{i+j-1}]$.
Thus, because of the NAD property [see Definition \ref{NAD}(ii)] of
$\alpha,\beta$, $u^{j}=\alpha(v^{j}), v^{j}=\beta(u^{j})$,
$ds\,dP$-a.s. on $[t,t_{i+j}]$. Consequently, iterating this argument
we obtain the existence of a couple $(u,v)\in{\cal
U}_{t,T}^\pi\times{\cal V}_{t,T}^\pi$ which satisfies the statement
of the lemma. Its uniqueness is an immediate consequence of the
above construction.
\end{pf}
Given a couple of NAD-strategies $(\alpha,\beta)\in{\cal A}_{t,
T}^\pi\times{\cal B}_{t,T}^\pi$ of the both players, the above lemma
allows to define the corresponding dynamics and the corresponding
pay-off functional through those of the associated\vadjust{\goodbreak} admissible
control processes. More precisely, for $(u,v)\in{\cal
U}_{t,T}^\pi\times{\cal V}_{t,T}^\pi$ such that $\alpha(v)=u$ and
$\beta(u)=v$, $ds\,dP$-a.e. on $[t,T]\times\Omega$, we define, for all
$\vartheta\in L^2(\Omega,{\cal F}_t^\pi,P;R^d)$ and $x\in R^d$,
\begin{eqnarray*}
X^{t,\vartheta; \alpha,\beta}&:=&X^{t,\vartheta; u, v}, \\
\bigl(Y^{t,\vartheta; \alpha,\beta},Z^{t,\vartheta; \alpha,\beta
},M^{t,\vartheta;
\alpha,\beta}
\bigr)&:=&\bigl(Y^{t,\vartheta; u, v},Z^{t,\vartheta; u,
v},M^{t,\vartheta; u, v}\bigr),
\\
J^\pi(t,x;\alpha,\beta)&:=&J^\pi(t,x;u,v).
\end{eqnarray*}
After the above preliminary discussion, we are now able to introduce
the upper and the lower value functions for the game over the time
interval $[t,T]$ along a partition $\pi$. We define the \textit{lower
value function along a partition} $\pi$~as
\begin{equation}
W^\pi(t,x):= \esssup_{\alpha\in{\mathcal
{A}}_{t,T}^\pi} \essinf_{\beta\in{\cal{B}}_{t,T}^\pi}
J^\pi(t,x;\alpha,\beta)
\end{equation}
and the \textit{upper one} as follows:
\begin{equation}
U^\pi(t,x):= \essinf_{\beta\in{\cal
{B}}_{t,T}^\pi} \esssup_{\alpha\in{\cal{A}^\pi}_{t,T}}J^\pi(t,x;
\alpha,\beta).
\end{equation}
Let us emphasize that the above lower and the upper value functions are
defined as a combination of essential supremum and essential infimum
over a bounded family of $\widetilde{\cal F}_t^\pi$-measurable
random variables $J^\pi(t,x;\alpha,\beta)$. Indeed, due to
(\ref{BSDE-estimates-2})(i),
\[
\bigl|J^\pi(t,x;\alpha,\beta) \bigr|= \bigl|Y^{t,\vartheta; \alpha
,\beta}_t \bigr| \le C,
\qquad P\mbox{-a.s., for all }(\alpha,\beta)\in{\cal A}_{t,T}^\pi
\times{\cal B}_{t,T}^\pi.
\]
Consequently, with the definitions of the essential infimum
and the essential supremum over families of random variables, given in
\cite{D} and \cite{DS} (see also \cite{KS2} for a more detailed discussion),
the upper and the lower value functions $W^\pi(t,x)$ and
$U^\pi(t,x)$ are, a priori, themselves also bounded,
$\widetilde{\cal F}_t^\pi$-measurable random variables. But,
combining arguments from \cite{Buckdahn-Li-2008} and
\cite{Buckdahn-Li_Quincampoix-2011}, we will be able to prove that
they are deterministic. However, for this proof we will have first
to establish a dynamic programming principle.
Let us finish this section with the following estimates for the
lower and the upper value functions, which are an immediate
consequence of the corresponding uniform estimates
(\ref{BSDE-estimates-2}) for the solution of BSDE (\ref{BSDE}).\vspace*{-3pt}
\begin{lemma}\label{estimates-W,U}
Under our standard assumptions on the coefficients $b,\sigma,f$ and
$\Phi$ there exists a constant $L\in R$ such that, for all partitions
$\pi$ of $[0,T]$ and all $t\in[0,T], x,x'\in R^d$,
\begin{eqnarray}
\label{estimates_W,U}
\mathrm{(i)}&&\quad
\bigl|W^{\pi} (t,x) \bigr|+ \bigl|U^{\pi}(t,x) \bigr|\le L,
\nonumber
\\
\mathrm{(ii)} &&\quad \bigl|W^{\pi}(t,x)-W^{\pi}
\bigl(t,x' \bigr)\bigr |+ \bigl|U^{\pi}(t,x)-U^{\pi}
\bigl(t,x' \bigr) \bigr| \leq L\bigl |x-x' \bigr|,\\[-4pt] \eqntext{P
\mbox{-a.s.}}
\end{eqnarray}
\end{lemma}
\section{Lower and upper value functions along a partition}\label{sec3}
This section is devoted to the study of properties of the lower
and the upper value functions $W^\pi$ and $U^\pi$ defined along a
partition $\pi$\vadjust{\goodbreak} of the interval $[0,T]$. The main objectives in this
section are to prove that both functions, characterized in the
preceding section as random fields, are in fact deterministic, and
they satisfy a dynamic programming principle along the
partition $\pi$.\vspace*{-2pt}
\begin{theorem}\label{W,U deterministic} For any partition $\pi$ of
the interval $[0,T]$ and for all $(t,x)\in[0,T]\times R^d$, we have
$W^\pi(t,x)= E[W^\pi(t,x)], U^\pi(t,x)=E[U^\pi(t,x)]$, $P$-a.s.\vspace*{-2pt}
\end{theorem}
\begin{remark} A consequence of this theorem is that, by identifying
$W^\pi(t, x) :=E[W^\pi(t,x)], U^\pi(t,x):=E[U^\pi(t,x)],
(t,x)\in[0,T]\times R^d$, the lower and the upper value functions along
a partition $\pi$
$W^\pi$ and $U^\pi$ can be regarded as deterministic functions.\vspace*{-2pt}
\end{remark}
The proof of the above theorem is strongly inspired by that of
Proposition~3.1 in~\cite{Buckdahn-Li-2008} and uses heavily the
structure of our underlying probability space $(\Omega,{\cal F},P)$.
We only give the proof for $W^\pi(t,x)$, for some arbitrarily fixed
$(t,x)\in[0,T]\times R^d$. The proof for $U^\pi(t,x)$ is analogous and
won't be given here.
Let the partition $\pi$ of the interval $[0,T]$ be of the form $\pi=
\{0=t_0<t_1<\cdots<t_n=T\}$ and let $1\le j\le n$ be such that $t\in
[t_{j-1},t_{j})$. Recalling that $W^\pi(t,x)$ is an $\widetilde{\cal
F}_t^\pi$-measurable random variable, it follows from the definition
of the $\sigma$-field $\widetilde{\cal F}_t^\pi$ that, $W^\pi(t,x)$ $P$-a.s.
coincides with a measurable functional
$W^\pi(t,x)(\zeta^{(j-1)},B^{(t)})$ of $\zeta^{(j-1)}=(\zeta
_1,\ldots,
\zeta_{j-1})$ of the first $j-1$ components of the coordinate
process $\zeta= (\zeta_\ell)_{\ell\ge1}$ on $\Omega_1$ and the
Brownian motion $B^{(t)}= (B_s)_{s\in[0,t]}$ defined over $\Omega_2$
and restricted to the time interval $[0,t]$.\looseness=-1
Let $H_t$ be the Cameron--Martin space of all absolutely continuous
functions $h\in C([0,T];R^d)$ which derivative $\dot{h}$ is square
integrable and satisfies $\dot{h}_s=0, ds$-a.e. on $[t,T]$, and let
us denote by $\Omega_1^{(j-1)}$ the set of all sequences
$\rho=(\rho_\ell= (\rho_{\ell,1},\rho_{\ell,2}))_{\ell\ge
1}\in\Omega_1$, such that $\rho_\ell=0, \ell\ge j$. Given any
$(a,h)\in\Omega_1^{(j-1)} \times H_t$, we define the transformation
$\tau_{a,h}\dvtx \Omega\rightarrow\Omega$ by putting
$\tau_{a,h}(\rho,\omega'):=(\rho+a,\omega'+h)
(=((\rho_\ell+a_\ell)_{\ell\ge1},\omega'+h) )$,
$(\rho,\omega')\in\Omega=\Omega_1\times\Omega_2$. Such defined
transformation is bijective, $\tau_{a,h}^{-1}=\tau_{-a,-h}$,
$(a,h)\in\Omega_1^{(j-1)} \times H_t$, and its law $P\circ
[\tau_{a,h}]^{-1}$ is equivalent to $P$. Indeed, the law $P\circ
[\tau_{a,h}]^{-1}$ has with respect to $P$ the density
\[
L_{a,h}=\exp \biggl\{\langle a,\zeta\rangle+\int_0^t
\dot{h}_s\,dB_s- \frac12 \biggl(|a|^2+\int
_0^t|\dot{h}_s|^2\,ds
\biggr) \biggr\},
\]
where
\begin{eqnarray*}
\langle a,\zeta\rangle&:=&\sum_{\ell\ge1}a_\ell
\zeta_\ell= \sum_{\ell=1}^{j-1}a_\ell
\zeta_\ell \biggl(=\sum_{1\le\ell\le
j-1,i=1,2}a_{\ell,i}
\zeta_{\ell,i} \biggr)\quad \mbox{and}
\\
|a|^2&=&\sum_{\ell\ge1}|a_\ell|^2=
\sum_{\ell=1}^{j-1} |a_\ell|^2
\biggl(=\sum_{1\le\ell\le j-1,i=1,2}|a_{\ell,i}|^2
\biggr),
\end{eqnarray*}
$a=(a_\ell= (a_{\ell,1},a_{\ell,2}))_{\ell\ge1}\in
\Omega
_1^{(j-1)}$. We observe that the density $L_{a,h}$ is $\widetilde{\cal
F}_t^{\pi}$-measurable and belongs to $L^p(\Omega,{\cal F},P)$, for
all $p\ge1$.
The following lemma is essential for the proof that $W(t,x)$ is
\mbox{deterministic}.\vspace*{-2pt}
\begin{lemma}\label{xi_deterministic}
Let $\xi\in L^0(\Omega,\widetilde{{\cal F}}_t^\pi,P)$ be a random
variable which, for all
$(a,h)\in\Omega_1^{(j-1)}\times H_t$, is invariant
with respect to all transformations $\tau_{a,h}\dvtx \Omega\rightarrow
\Omega$,
that is, $\xi\circ\tau_{a,h}=\xi$, $P$-a.s. Then, there exists some
deterministic real number $c\in R$, such that $\xi=c, P$-a.s.\vspace*{-2pt}
\end{lemma}
\begin{pf} Let $\xi\in L^0(\Omega,\widetilde{\cal F}_t^\pi,P)$ be invariant
with respect to all transformations
$\tau_{a,h}\dvtx \Omega\rightarrow\Omega$,
$(a,h)\in\Omega_1^{(j-1)}\times H_t$. Then, for all $(a,h)\in
\Omega_1^{(j-1)} \times H_t$ and all bounded Borel functions
$g\dvtx R\rightarrow R$,
\begin{eqnarray}
&&E \bigl[g(\xi) \bigr]\nonumber \hspace*{-35pt}\\
&&\qquad= E \bigl[g(\xi\circ \tau_{a,h})
\bigr]
\\
&& \qquad = E \biggl[g(\xi)\exp \biggl\{\langle a,\zeta\rangle +\int
_0^t\dot{h}_s\,dB_s
\biggr\} \biggr] \cdot\exp \biggl\{-\frac12 \biggl(|a|^2+\int
_0^t|\dot {h}_s|^2\,ds
\biggr) \biggr\},\nonumber\hspace*{-35pt}
\end{eqnarray}
that is,
\begin{eqnarray}
& &E \Biggl[g(\xi)\exp \Biggl\{\sum_{\ell=1}^{j-1}a_\ell
\zeta_\ell + \int_0^t
\dot{h}_s \,dB_s \Biggr\} \Biggr]
\nonumber
\\
&&\qquad = E \bigl[g(\xi) \bigr]\cdot\exp \biggl\{\frac12 \biggl(|a|^2+
\int_0^t| \dot{h}_s|^2
\,ds \biggr) \biggr\}
\\
&&\qquad = E \bigl[g(\xi) \bigr]\cdot E \Biggl[\exp \Biggl\{\sum
_{\ell=1}^{j-1}a_\ell \zeta_\ell+
\int_0^t\dot{h}_s
\,dB_s \Biggr\} \Biggr]
\nonumber
\end{eqnarray}
for all $a_\ell\in R^2, 1\le\ell\le j-1$, and all $h\in H_t$, from
where we deduce that $\xi$ is independent of
$(\zeta^{(j-1)}=(\zeta_1,\ldots,\zeta_{j-1}), B^{(t)}=(B_s)_{s
\in[0,t]})$ and, hence also of $\widetilde{\cal F}_t^\pi=\sigma\{
\zeta^{(j-1)},B^{(t)}\}$. But this means that $\xi$ as an $\widetilde
{{\cal F}}_t^\pi$-measurable random variable is independent of itself. The
statement of the lemma follows now easily.\vspace*{-2pt}
\end{pf}
\begin{pf*}{Proof of Theorem \ref{W,U deterministic}} In order to be
able to
conclude our theorem form the above lemma, we only have to show that
the random variable $W^\pi(t,x)$ is invariant with respect to the
transformations $\tau_{a,h}\dvtx \Omega\rightarrow\Omega$, for all
$(a,h)\in\Omega_1^{(j-1)} \times H_t$. For showing this, we fix
arbitrarily $(a,h)\in\Omega_1^{(j-1)} \times H_t$ and we proceed in
an analogous spirit as that in the proof of Proposition 3.1 in
\cite{Buckdahn-Li-2008}. But, however, the framework is different
here.\vadjust{\goodbreak}
\textit{Step} 1. Given a couple of admissible controls
$(u,v)\in{\cal U}_{t,T}^\pi\times{\cal V}_{t,T}^\pi$, we notice
that also the transformed couple
$(u\circ\tau_{a,h},v\circ\tau_{a,h})$ belongs to ${\cal
U}_{t,T}^\pi\times{\cal V}_{t,T}^\pi$. Indeed, having
$t\in[t_{j-1},t_{j})$,
\begin{eqnarray*}
u_s&=&u_j \bigl(s,(\zeta_1,\ldots,
\zeta_{j-1},\zeta_{j,1}, B_{{\cdot\wedge s}})
\bigr)I_{[t,t_j)}(s)
\\[-2pt]
&&{}+\sum_{\ell=j+1}^nu_\ell
\bigl(s,(\zeta_1,\ldots, \zeta_{\ell-1},\zeta_{\ell,1},B_{{\cdot\wedge s}})
\bigr)I_{[t_{\ell
-1},t_\ell)} (s)\qquad ds\,dP \mbox{-a.e.},
\end{eqnarray*}
for measurable functionals $u_\ell, 1\le\ell\le n$, the
transformed control process $u\circ\tau_{a,h}$ takes the form
\begin{eqnarray}
&& u_s\circ\tau_{a,h} \nonumber\\[-2pt]
&&\qquad= u_j
\bigl(s,(\zeta_1+a_1,\ldots, \zeta_{j-1}+a_{j-1},
\zeta_{j,1},B_{{\cdot\wedge s}}+h_{\cdot\wedge t}) \bigr)I_{[t,t_j)}(s)
\nonumber
\\[-9pt]
\\[-9pt]
\nonumber
&&\qquad\quad{}+ \sum_{\ell=j+1}^nu_\ell
\bigl(s,(\zeta_1+a_1,\ldots, \zeta_{j-1}+a_{j-1},
\zeta_j,\ldots,\zeta_{\ell-1},\zeta_{\ell,1},\nonumber\\[-2pt]
&&\hspace*{220pt}\qquad{} B_{{\cdot\wedge s}}+h_{\cdot\wedge t}) \bigr)I_{[t_{\ell-1},t_\ell)}
(s),\nonumber
\end{eqnarray}
$ds\,dP\mbox{-a.e.,}$ from where we see that also
$u\circ\tau_{a,h}$ is an admissible control for Player 1; the
symmetric argument shows that $v\circ\tau_{a,h} \in{\cal
V}_{t,T}^\pi$. Applying now the transformation to the forward
equation (\ref{SDE}) and taking into account that the increments of
the Brownian motion after $t$ are not changed by the transformation:
$(B_s-B_t)\circ\tau_{a,h} =B_s-B_t, s\in[t,T]$ (Indeed, recall
that $\dot{h}_s=0, ds$-a.e. on $[t,T]$), we obtain from the
uniqueness of the solution of SDE (\ref{SDE}) that
$X^{t,x;u,v}_s\circ\tau_{a,h}=X^{t,x;u(\tau_{a,h}),v(\tau_{a,h})}_s,
s\in[t,T]$, $P$-a.s. Let us now apply the transformation
$\tau_{a,h}$ to BSDE (\ref{BSDE}). With the argument already used for
its application to the forward SDE we see that BSDE (\ref{BSDE})
becomes
\begin{eqnarray}
\label{transformed_BSDE}
&&dY^{t,x; u, v}_s
\circ\tau_{a,h}\nonumber\\
&&\qquad=-E \bigl[f \bigl(s,X^{t,x; u(\tau_{a,h}),
v(\tau_{a,h})}_s,
Y^{t,x; u, v}_s\circ \tau_{a,h}, Z^{t,x; u, v}_s
\circ\tau_{a,h},
\nonumber
\\
& &\qquad\hspace*{164pt}{} u_s(\tau_{a,h}), v_s(
\tau_{a,h}) \bigr) | \widetilde{\cal F}^\pi_s
\bigr] \,ds
\\
&&\hspace*{10pt}\qquad{}+Z^{t,x;u, v}_s\circ \tau_{a,h}
\,dB_s+dM^{t,x; u,v}_s\circ \tau_{a,h},
\nonumber\\
&&Y^{t,x; u, v}_T\circ\tau_{a,h} = E \bigl[\Phi
\bigl(X^{t,x; u(\tau_{a,h}),
v(\tau_{a,h})}_T \bigr) | \widetilde{\cal F}^\pi_{T}
\bigr].
\nonumber
\end{eqnarray}
We remark that
(i) $(Y^{t,x; u, v}\circ\tau_{a,h},Z^{t,x; u, v}\circ\tau_{a,h})
\in{\cal S}^2_{\widetilde{\mathbb{F}}^\pi}(t,T;R)\times
L_{\widetilde{\mathbb{F}}^\pi}^2(t,\break T; R^d)$. Indeed, the
$\widetilde{\mathbb{F}}^\pi$-adaptedness of the transformed process
can be proved directly, and the square integrability follows from
standard $L^p$-estimates for the solutions of BSDEs:
\begin{eqnarray*}
&&E \biggl[\sup_{s\in[t,T]} \bigl|Y^{t,x; u,
v}_s\circ
\tau_{a,h} \bigr|^2+\int_t^T\bigl|Z^{t,x; u, v}_s
\circ\tau _{a,h}\bigr|^2\,ds \biggr]
\\
&&\qquad= E \biggl[ \biggl(\sup_{s\in[t,T]}\bigl|Y^{t,x; u,
v}_s\bigr|^2+
\int_t^T\bigl|Z^{t,x; u, v}_s\bigr|^2
\,ds \biggr)L_{a,h} \biggr]
\\
&&\qquad\le C \bigl(E\bigl[L_{a,h}^2\bigr]
\bigr)^{1/2} \biggl(E \biggl[\sup_{s\in[t,T]}
\bigl|Y^{t,x; u,v}_s\bigr|^4+ \biggl(\int
_t^T\bigl|Z^{t,x; u, v}_s\bigr|^2
\,ds \biggr)^2 \biggr] \biggr)^{1/2}
\\
&&\qquad< +\infty.
\end{eqnarray*}
On the other hand, the fact $L_{a,h}\in
L^2(\Omega,\widetilde{\cal F}_t^\pi,P)$ has as consequence that also
the transformed $(\widetilde{\mathbb{F}}^\pi,P)$-martingale
$M^{t,x;u,v}\circ\tau_{a,h}=(M^{t,x;u,v}_s\circ\tau_{a,h})_{s\in[t,T]}$
is again an $(\widetilde{\mathbb{F}}^\pi,P)$-martingale. Indeed, for
$t\le s\le T$ and $\xi\in L^\infty(\Omega,\widetilde{\cal
F}_s^\pi,P)$, also $\xi\circ\tau_{-a,-h}\in L^\infty(\Omega,
\widetilde{\cal F}_s^\pi,P)$, and
\begin{eqnarray}
&& E \bigl[ \bigl(M^{t,x;u,v}_T-M^{t,x;u,v}_s
\bigr)\circ\tau_{a,h} \cdot\xi \bigr]
\nonumber
\\
&&\qquad = E \bigl[ \bigl(M^{t,x;u,v}_T-M^{t,x;u,v}_s
\bigr) \cdot\xi\circ\tau_{-a,-h}\cdot L_{a,h} \bigr]
\\
&&\qquad = E \bigl[E \bigl[M^{t,x;u,v}_T-M^{t,x;u,v}_s
| \widetilde{\cal F}_s^\pi \bigr] \cdot\xi\circ
\tau_{-a,-h} L_{a,h} \bigr]=0.
\nonumber
\end{eqnarray}
Consequently, $M^{t,x;u,v}\circ\tau_{a,h}$ is an
$(\widetilde{\mathbb{F}}^\pi,P)$-martingale; its square
integrability follows from an argument similar to that for $(Y^{t,x;
u, v}\circ\tau_{a,h},Z^{t,x; u, v}\circ\tau_{a,h})$, (recall the
explicit representation of $M^{t,x;u,v}$ in terms of $Y^{t,x;u,v}$,
which implies the $L^p$-integrability of $M^{t,x;u,v}$ for all $p\ge
1.$) and its orthogonality to $B$ stems from the fact that it is a
pure jump martingale.
This shows that $(Y^{t,x; u, v}\circ\tau_{a,h},Z^{t,x; u, v}\circ
\tau_{a,h},
M^{t,x;u,v}\circ\tau_{a,h})$ is a solution of BSDE (\ref{BSDE}) with
the couple
of admissible controls $(u(\tau_{a,h}),v(\tau_{a,h}))$. From the
uniqueness of the
solution of this BSDE it then follows that
\begin{eqnarray}
& & \bigl(Y^{t,x; u, v}\circ \tau_{a,h},Z^{t,x; u, v}\circ
\tau_{a,h}, M^{t,x;u,v} \circ\tau_{a,h} \bigr)
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&&\qquad = \bigl(Y^{t,x; u(\tau_{a,h}), v(\tau_{a,h})},Z^{t,x;
u(\tau_{a,h}), v(\tau_{a,h})},M^{t,x;u(\tau_{a,h}),v(\tau_{a,h})}
\bigr),
\end{eqnarray}
and, in particular, it follows that
\[
J^\pi(t,x;u,v)\circ\tau_{a,h}=J^\pi
\bigl(t,x;u(\tau _{a,h}),v(\tau_{a,h})\bigr),\qquad P
\mbox{-a.s.}
\]
\textit{Step} 2. Let us translate in this step the result of
step 1 to couples of NAD strategies. For $\beta\in{\cal B}_{t,T}^\pi$
we define
$\beta_{a,h}(u):=\beta(u(\tau_{-a,-h}))(\tau_{a,h}), u\in{\cal
U}_{t,T}^\pi$.
For such defined mapping $\beta_{a,h}\dvtx {\cal U}_{t,T}^\pi\rightarrow
{\cal V}_{t,T}^\pi$ it can be verified in a straight-forward manner
that it
belongs to ${\cal B}_{t,T}^\pi$. We also observe that $(\beta
_{-a,-h})_{a,h}=\beta$.
A symmetric definition allows to introduce $\alpha_{a,h}\in{\cal
A}_{t,T}^\pi$, for
$\alpha\in{\cal A}_{t,T}^\pi$ and to get $(\alpha
_{-a,-h})_{a,h}=\alpha$.
Given a couple of NAD-strategies $(\alpha,\beta)\in{\cal
A}_{t,T}^\pi
\times{\cal
B}_{t,T}^\pi$, let us denote by $(u,v)\in{\cal U}_{t,T}^\pi\times
{\cal
V}_{t,T}^\pi$ the couple of admissible controls associated with
through Lemma
\ref{controls-NAD-strategies}. Then
\begin{eqnarray*}
\alpha_{a,h} \bigl(v(\tau_{a,h}) \bigr)&=&\alpha(v) (
\tau_{a,h})=u(\tau_{a,h}) \quad\mbox{and}
\\
\beta_{a,h} \bigl(u(\tau_{a,h}) \bigr)&=&\beta(u) (
\tau_{a,h})=v(\tau_{a,h}).
\end{eqnarray*}
Consequently, the couple $(u(\tau_{a,h}),v(\tau_{a,h}))\in
{\cal
U}_{t,T}^\pi\times{\cal V}_{t,T}^\pi$ is associated with $(\alpha
_{a,h},\beta_{a,h})$
through Lemma \ref{controls-NAD-strategies}, and from step 1 we get
\begin{eqnarray}
J^\pi(t,x;\alpha,\beta)\circ \tau_{a,h} & = &
J^\pi(t,x;u,v)\circ \tau_{a,h}= J^\pi
\bigl(t,x;u(\tau_{a,h}),v(\tau_{a,h}) \bigr)
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
& = & J^\pi(t,x;\alpha_{a,h},
\beta_{a,h}),\qquad P \mbox{-a.s.}
\end{eqnarray}
\textit{Step} 3. Using the definition of the esssup and
the essinf over a family of random variables as well as the fact
that the transformation $\tau_{a,h}$ is invertible and its law
$P\circ[\tau_{a,h}]^{-1}$ is equivalent to $P$, we show that
\begin{eqnarray}
&&W^\pi(t,x)\circ\tau_{a,h}\nonumber\\
&&\qquad = \bigl(
\esssup_{\alpha\in
{\mathcal{A}}_{t,T}^\pi} \essinf_{\beta\in{\cal{B}}_{t,T}^\pi} J^\pi(t,x;\alpha,\beta
) \bigr)\circ\tau_{a,h}
\\
&&\qquad = \esssup_{\alpha\in{\mathcal{A}}_{t,T}^\pi} \essinf_{\beta\in{\cal{B}}_{t,T}^\pi}
\bigl(J^\pi (t,x;\alpha,\beta)\circ \tau_{a,h} \bigr),\qquad
P \mbox{-a.s.}\nonumber
\end{eqnarray}
Consequently, by combining the results of the previous steps
and by considering that,
thanks to step 2, $\{\alpha_{a,h}, \alpha\in{\cal A}_{t,T}^\pi\}
={\cal
A}_{t,T}^\pi$
and $\{\beta_{a,h}, \beta\in{\cal B}_{t,T}^\pi\}={\cal B}_{t,T}^\pi$,
we obtain
\begin{eqnarray}
W^\pi(t,x)\circ\tau_{a,h} & = &
\esssup_{\alpha\in{\mathcal
{A}}_{t,T}^\pi} \essinf_{\beta\in{\cal{B}}_{t,T}^\pi} \bigl(J^\pi (t,x;
\alpha,\beta)\circ \tau_{a,h} \bigr)
\nonumber
\\
& = & \esssup_{\alpha\in{\mathcal{A}}_{t,T}^\pi} \essinf_{\beta\in{\cal{B}}_{t,T}^\pi} J^\pi(t,x;
\alpha _{a,h},\beta_{a,h})
\\
& = & W^\pi(t,x),\qquad P\mbox{-a.s.}
\nonumber
\end{eqnarray}
By combining this result with Lemma \ref{xi_deterministic}, we complete
the proof.
\end{pf*}
As an immediate consequence of Lemma \ref{estimates-W,U} and the above
result that
the lower and the upper value functions along a partition are
deterministic, we have
the following result.
\begin{lemma}\label{estimates+W,U} There exists a constant $L\in R$
which does not
depend on the partition $\pi$ of the interval $[0,T]$, such that, for
all $t\in[0,T]$,
$x,x'\in R^d$,
\begin{eqnarray}
\mathrm{(i)}&&\quad \bigl|W^{\pi} (t,x) \bigr|+ \bigl|U^{\pi}(t,x)
\bigr| \le L,
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
\mathrm{(ii)} &&\quad\bigl|W^{\pi}(t,x)-W^{\pi}
\bigl(t,x' \bigr) \bigr|+ \bigl|U^{\pi}(t,x)-U^{\pi}
\bigl(t,x' \bigr)\bigr | \leq L \bigl|x-x' \bigr|.
\end{eqnarray}
\end{lemma}
After having proved that the lower and the upper value functions along
a partition
$\pi$ are deterministic, our objective is now to show that, with
respect to the
points of the partition they satisfy the DPP. A key role will be played
here by the notion of backward stochastic semigroup,
introduced by Peng in \cite{Pe1}.
Given a partition $\pi=\{0=t_0<t_1<\cdots<t_n=T\}$ of the interval $[0,T]$,
initial data $(t,x)\in[0,T)\times R^d$, a positive $\delta<T-t$ and a
couple of\vadjust{\goodbreak}
admissible control processes $(u,v)\in{\cal U}_{t,t+\delta}^\pi
\times
{\cal V}_{t,
t+\delta}^\pi$ as well as a random variable $\eta\in L^2(\Omega
,{\cal
F}^\pi_{t+\delta},
P)$, we define the backward stochastic semigroup
\[
G^{t,x;u,v}_{s,t+\delta}(\eta):=\overline{Y}_s^{u,v},
\qquad s\in[t,t+\delta],
\]
through the BSDE with time horizon $t+\delta$,
\begin{equation}
\label{semigroup} \cases{d \overline{Y}^{u, v}_s = -E
\bigl[f \bigl(s,X^{t,\vartheta; u, v}_s, \overline{Y}^{u, v}_s,
\overline{Z}^{u, v}_s, u_s, v_s
\bigr) | \widetilde{\cal F}^\pi_s \bigr] \,ds\vspace*{2pt}\cr
\hspace*{39pt}{} +
\overline{Z}^{u, v}_s \,dB_s+d
\overline{M}^{u,
v}_s, \vspace*{2pt}
\cr
\overline{Y}^{u, v}_T
= E \bigl[\eta | \widetilde{\cal F}^\pi_{t+\delta} \bigr],}
\end{equation}
and its unique solution $(\overline{Y}^{u, v}, \overline
{Z}^{u, v},
\overline{M}^{u, v})\in{\cal S}^2_{\widetilde{\mathbb{F}}^\pi
}(t,t+\delta;R)
\times L_{\widetilde{\mathbb{F}}^\pi}^2(t,t+\delta;R^d)\times
{\cal M}_{\widetilde{\mathbb{F}}^\pi}^2(t,t+\delta;R)$ with
$[B,\overline{M}^{u, v}]_s=0, s\in[t,T]$ and $\overline{M}^{u, v}_t=0$,
where $X^{t,\vartheta; u, v}$ is the solution of SDE (\ref{SDE}).
From the discussion made in the frame of Remark \ref{remark_bsde} it becomes
clear that if, for some point $t_j$ of the partition $\pi=\{
0=t_0<t_1<\cdots
<t_n=T\}$, $t_{j-1}\le t<t+\delta=t_j$ and $\eta$ is
$\widetilde{\cal F}^\pi_{t_j-}$-measurable, then $\overline{M}^{u, v}_s=0,
s\in[t,t_j]$.
The properties of the backward stochastic semigroup follow directly
from those of the BSDE through which it is defined, so that we won't discuss
separately here (refer to \cite{Pe1}, or \cite{Buckdahn-Li-2008}). The
notion of backward stochastic semigroup now allows to
study the DPP along a partition $\pi$ of the
time interval $[0,T]$.
\begin{theorem}\label{DPP} Let $\pi=\{0=t_0<t_1<\cdots<t_n=T\}$ be a
partition of the interval $[0,T]$, and let $t\in[t_i,t_{i+1})$ and
$x\in R^d$. Then, for all $i+1\le j\le n$, $P$-a.s.,
\begin{eqnarray}
\label{DPP-aa}
W^{\pi}(t,x) &=& \esssup_{\alpha\in
\mathcal{A}^{\pi}_{t,t_j}}
\essinf_{\beta\in
\mathcal{B}^{\pi}_{t,t_j}}G^{t,x;\alpha,\beta}_{t,t_j} \bigl(W^{\pi
}
\bigl(t_j,X_{t_j}^{t,x;
\alpha,\beta} \bigr) \bigr),
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
U^{\pi}(t,x) &=& \essinf_{\beta\in\mathcal{B}^\pi_{t,t_j}}\esssup
_{\alpha
\in\mathcal{A}^\pi_{t,t_j}} G^{t,x;\alpha,\beta}_{t,t_j} \bigl(U^{\pi}
\bigl(t_j,X_{t_j}^{t,x;\alpha,\beta} \bigr) \bigr).
\end{eqnarray}
\end{theorem}
\begin{remark} The space ${\cal U}^\pi_{t,t_j}$ of admissible controls
for Player 1 for games over the time interval $[t,t_j]$ along the
partition $\pi$ is defined as the set of all control processes
$u\in{\cal U}^\pi_{t,T}$ restricted to the time interval $[t,t_j]$; the
space ${\cal V}^\pi_{t,t_j}$ of admissible controls for Player 2 is
defined analogously. The NAD-strategies for Player~2,
$\beta\in\mathcal{B}^\pi_{t,t_j}\dvtx {\cal U}^\pi_{t,t_j}\rightarrow
{\cal
V}^\pi_{t,t_j}$, are defined in the same manner as the NAD-strategies in
$\mathcal{B}^\pi_{t,T}$, with the only difference that we consider
$t_j$ instead $T=t_n$ as terminal horizon. The same is done in the
definition of the set $\mathcal{A}^\pi_{t,t_j}$ of NAD-strategies
for Player~1.
\end{remark}
The proof split into two lemmas for the lower value
function along the partition~$\pi$; it is similar for the upper value
function along the partition $\pi$. Let us fix
arbitrarily a partition\vadjust{\goodbreak} $\pi=\{0=t_0<t_1<\cdots<t_n=T\}$ of the interval
$[0,T]$, and let $t\in[t_i,t_{i+1})$,
$i+1\le j\le n$ and $x\in R^d$. We put
\[
\widetilde{W}_{t_j}^{\pi}(t,x) =\esssup_{\alpha\in\mathcal{A}^\pi_{t,t_j}}
\essinf_{\beta\in\mathcal{B}^\pi_{t,t_j}}G^{t,x;\alpha,
\beta}_{t,t_j}\bigl(W^{\pi}
\bigl(t_j,X_{t_j}^{t,x;\alpha,\beta}\bigr)\bigr).
\]
Obviously, $ \widetilde{W}_{t_j}^{\pi}(t,x)$
is a bounded, $\widetilde{\cal F}_t^\pi$-measurable random variable.
\begin{lemma}\label{DPP-1} Under the standard assumptions, we
have made on the coefficients it holds that $
\widetilde{W}_{t_j}^{\pi}(t,x)\le W^{\pi}(t,x)$, $P$-a.s.
\end{lemma}
\begin{pf}
\textit{Step} 1. Let us fix an arbitrary
$\varepsilon>0$. Then, we can find $\alpha_1^\varepsilon\in
{\cal A}_{t,t_j}^\pi$ such that
\[
\widetilde{W}_{t_j}^{\pi}(t,x)\le \essinf_{\beta\in\mathcal{B}^\pi_{t,t_j}}G^{t,x;\alpha
_1^\varepsilon,
\beta}_{t,t_j}
\bigl(W^{\pi}\bigl(t_j,X_{t_j}^{t,x;\alpha_1^\varepsilon,\beta}
\bigr)\bigr)+ \varepsilon,\qquad P\mbox{-a.s.}
\]
In order to verify this latter relation, we put
\[
I(\alpha): =\essinf_{\beta\in\mathcal{B}^\pi_{t,t_j}}G^{t,x;\alpha,\beta}_{t,t_j}
\bigl(W^{\pi}\bigl(t_j,X_{t_j}^{t,x;\alpha,\beta}
\bigr)\bigr),\qquad \alpha\in{\cal A}_{t,t_j}^\pi,
\]
and we note that, due to the
properties of the essential supremum over a family of random
variables, there is some sequence $(\alpha^k)_{k\ge1}\subset{\cal
A}_{t,t_j}^\pi$ such that
\[
\widetilde{W}_{t_j}^{\pi}(t,x)=\esssup_{\alpha\in
\mathcal{A}^\pi_{t,t_j}}I(
\alpha)=\sup_{k\ge1}I\bigl(\alpha^k\bigr),\qquad P
\mbox{-a.s.}
\]
Thus, putting $\triangle_k:=\{\widetilde{W}_{t_j}^{\pi}
(t,x)\le I(\alpha^k)+\varepsilon, \widetilde{W}_{t_j}^{\pi}(t,x)
> I(\alpha^\ell)+\varepsilon (1\le\ell\le k-1)\}\in\widetilde
{\cal F}_t^\pi$, $k\ge1$, we define a partition of $\Omega$,
and putting
\[
\alpha^\varepsilon_1(\cdot):=\sum_{k\ge1}
I_{\triangle_k}\alpha^k(\cdot)\dvtx {\cal V}_{t,t_j}^\pi
\rightarrow {\cal U}_{t,t_j}^\pi,
\]
we check easily that\vspace*{-1pt} $\alpha^\varepsilon_1$ is an
NAD-strategy in ${\cal A}_{t,t_j}^\pi$ and that
$\widetilde{W}_{t_j}^{\pi}(t,x)\le\sum_{k\ge1}
I_{\triangle_k}I(\alpha^k)+\varepsilon\le\sum_{k\ge1}
I_{\triangle_k}G^{t,x;\alpha^k,\beta_1}_{t,t_j}(W^{\pi}
(t_j,X_{t_j}^{t,x;\alpha^k,\beta_1}))+\varepsilon$, P-a.s., for all
$\beta_1\in{\cal B}_{t,t_j}^\pi$. Given an
arbitrary $\beta_1\in{\cal B}_{t,t_j}^\pi$, we let $(u^k,v^k)\in
{\cal U}_{t,t_j}^\pi\times{\cal V}_{t,t_j}^\pi$ be such that
$\alpha^k(v^k)=u^k, \beta_1(u^k)=v^k, ds\,dP$-a.e. on
$[t,t_j]\times\Omega$, and we introduce $(u_1,v_1):=\sum_{k\ge
1}I_{\triangle_k}(u^k,v^k)\in{\cal U}_{t,t_j}^\pi\times{\cal
V}_{t,t_j}^\pi$. Then, since for the
$\widetilde{\mathbb{F}}^\pi$-stopping time
$\tau_k=t_jI_{\triangle_k}+tI_{\triangle_k^c}$ the processes $u_1$
and $u^k$ coincide, $ds\,dP$-a.e. on $[[t,\tau_k]]$, also
$\beta_1(u^k)=\beta_1(u_1)$, $ds\,dP$-a.e. on $[[t,\tau_k]]$. Thus,
\[
\beta_1(u_1)=\sum_{k\ge1}I_{\triangle_k}
\beta_1 \bigl(u^k \bigr)= \sum
_{k\ge
1}I_{\triangle_k}v^k=v_1,
\qquad ds\,dP\mbox{-a.e. on} [t,t_j]\times\Omega,
\]
and with a symmetric argument we also have
\[
\alpha^\varepsilon_1(v_1)=\sum
_{k\ge1}I_{\triangle_k}\alpha^k(v_1)
=\sum_{k\ge1}I_{\triangle_k}\alpha^k
\bigl(v^k \bigr)=u_1,\qquad ds\,dP\mbox{-a.e. on }
[t,t_j]\times\Omega.
\]
This shows that the couple $(u_1,v_1)\in{\cal U}_{t,
t_j}^\pi\times{\cal V}_{t,t_j}^\pi$ is associated with
$(\alpha^\varepsilon_1,\beta_1)\in{\cal A}_{t,t_j}^\pi\times{\cal
B}_{t,t_j}^\pi$ by Lemma \ref{controls-NAD-strategies}.
Consequently, from the uniqueness of the solution of SDE (\ref{SDE})
we conclude with a standard argument that
\begin{eqnarray}
\sum_{k\ge1}I_{\triangle_k}X^{t,x;\alpha^k,\beta_1}=\sum
_{k\ge1} I_{\triangle_k}X^{t,x;u^k,v^k}=X^{t,x;u_1,v_1}=
X^{t,x;\alpha_1^\varepsilon,\beta_1} \nonumber\\
\eqntext{\mbox{on } [t,t_j], P\mbox{-a.s.}}
\end{eqnarray}
Similarly, using now the uniqueness of the solution of
BSDE defining the backward stochastic semigroup, we show that
\[
\sum_{k\ge1}I_{\triangle_k} \bigl(
\widetilde{Y}^{t,x;\alpha^k,\beta_1}, \widetilde{Z}^{t,x;\alpha^k,\beta_1},\widetilde{M}^{t,x;
\alpha^k,\beta_1}
\bigr)= \bigl(\widetilde{Y}^{t,x;\alpha_1^\varepsilon,
\beta_1},\widetilde{Z}^{t,x;\alpha_1^\varepsilon,
\beta_1},
\widetilde{M}^{t,x;\alpha_1^\varepsilon,\beta_1} \bigr),
\]
and recalling the definition of the backward stochastic
semigroup, we see that
\[
\sum_{k\ge1}I_{\triangle_k}G^{t,x;\alpha^k,\beta_1}_{t,t_j}
\bigl(W^{\pi} \bigl(t_j,X_{t_j}^{t,x;\alpha^k,\beta_1}
\bigr) \bigr)= G^{t,x;\alpha^\varepsilon_1,\beta_1}_{t,t_j} \bigl(W^{\pi}
\bigl(t_j, X_{t_j}^{t,x;\alpha^\varepsilon_1,\beta_1} \bigr) \bigr).
\]
Consequently, for all $\beta_1\in{\cal B}_{t,t_j}^\pi$,
\begin{eqnarray}
\widetilde{W}_{t_j}^{\pi}(t,x)& \le& \sum
_{k\ge1} I_{\triangle_k}I \bigl(
\alpha^k \bigr)+\varepsilon
\nonumber
\\
&\le& \sum_{k\ge1}I_{\triangle_k}
G^{t,x;\alpha^k,\beta_1}_{t,t_j} \bigl(W^{\pi} \bigl(t_j,X_{t_j}^{t,x;
\alpha^k,\beta_1}
\bigr) \bigr)+\varepsilon
\\
& = & G^{t,x;\alpha^\varepsilon_1,\beta_1}_{t,t_j} \bigl(W^{\pi}
\bigl(t_j, X_{t_j}^{t,x;\alpha^\varepsilon_1,\beta_1} \bigr) \bigr)+
\varepsilon,\qquad P\mbox{-a.s.}
\nonumber
\end{eqnarray}
Let us make now a special choice of $\beta_1\in{\cal
B}_{t,t_j}^\pi$. Given an arbitrary $\beta\in{\cal B}_{t,T}^\pi$ and
any $u_2\in{\cal U}_{t_j,T}^\pi$, we define for any $u_1\in{\cal
U}_{t,t_j}^\pi$ the process $u_1\oplus u_2:= u_1I_{[t,t_j]}+
u_2I_{(t_j,T]}\in{\cal U}_{t,T}^\pi$, and we put
\[
\beta_1(u_1):=\beta(u_1\oplus
u_2)_{|[t,t_j]},\qquad u_1\in{\cal
U}_{t,t_j}^\pi,
\]
the restriction of $\beta(u_1\oplus u_2)$ to the time
interval $[t,t_j]$. It can be easily verified that such defined
mapping $\beta_1\dvtx {\cal U}_{t,t_j}^\pi\rightarrow{\cal
V}_{t,t_j}^\pi$ belongs to ${\cal B}_{t, t_j}^\pi$, and thanks to
its nonanticipativity property it does not depend on the special
choice of $u_2$. Let us denote by
$(u_1^\varepsilon,v_1^\varepsilon)\in{\cal U}_{t,t_j}^\pi\times
{\cal V}_{t,t_j}^\pi$ the unique couple of control processes
associated with $(\alpha_1^\varepsilon,\beta_1)$ through Lemma
\ref{controls-NAD-strategies}.
\textit{Step} 2. After having proven in step 1 that
\[
\widetilde{W}_{t_j}^{\pi}(t,x)\le G^{t,x;
\alpha^\varepsilon_1,\beta_1}_{t,t_j}
\bigl(W^{\pi}\bigl(t_j,X_{t_j}^{t,
x;\alpha^\varepsilon_1,\beta_1}
\bigr)\bigr)+ \varepsilon,\qquad P\mbox{-a.s.},
\]
let us now estimate the expression $W^{\pi}(t_j,X_{t_j}^{t,
x;\alpha^\varepsilon_1,\beta_1})$ to which the backward stochastic
semigroup is applied at the right-hand side of the above estimate.
For this we consider a Borel partition ${\cal O}_k,k\ge1$, of $R^d$,
consisting of nonempty Borel sets ${\cal O}_k$ with diameter less
or equal to $\varepsilon$, and we fix arbitrarily in each of this
sets ${\cal O}_k$ an element $x_k$. With the arguments already
developed in step 1 we show that, for every $k\ge1$, there is
some $\alpha_2^k\in{\cal A}_{t_j,T}^\pi$ such that
\begin{eqnarray*}
W^{\pi}(t_j,x_k)&=&
\esssup_{\alpha_2 \in{\mathcal{A}}_{t_j,
T}^\pi}\essinf_{\beta_2 \in{\cal{B}}_{t_j,T}^\pi} J^{\pi}(t_j,x_k;
\alpha_2,\beta_2)
\\
&\le& \essinf_{\beta_2 \in{\cal{B}}_{t_j,T}^\pi} J^{\pi}\bigl(t_j,x_k;
\alpha_2^k,\beta_2\bigr)+\varepsilon, \qquad
P\mbox{-a.s.},
\end{eqnarray*}
and putting $\alpha_2^\varepsilon(\cdot):=\sum_{k\ge1}I
\{X^{t,x;\alpha_1^\varepsilon,\beta_1}_{t_j}\in{\cal O}_k\}
\alpha_2^k(\cdot)\dvtx {\cal V}_{t_j,T}^\pi\rightarrow{\cal U}_{t_j,T}^\pi$
we obtain an NAD-strategy from ${\cal A}_{t_j,T}^\pi$. Indeed, the
sets $\{X^{t,x;\alpha_1^\varepsilon,\beta_1}_{t_j}\in{\cal O}_k\}$,
$k\ge1$, forming a partition of $\Omega$, belong to
\[
{\cal F}^\pi_{t_j-}={\cal F}^B_{t_j}
\vee{\cal H}_{j}= \widetilde{\cal F}^\pi_{t_j}.
\]
(We remark that the relation ${\cal F}^\pi_{s-}=
\widetilde{\cal F}^\pi_{s}$ only holds for points of the partition
$\pi$; this is also the reason, why we do not have a DPP which
does not use the points of the partition $\pi$). Thus, by combining
the arguments developed in step 1 with the Lipschitz property of
$W^{\pi}(t_j,\cdot)$ and $J^{\pi}(t_j,\cdot; \alpha, \beta)$ we can show that,
for all $\beta_2\in{\cal
B}_{t_j,T}^\pi$,
\begin{eqnarray}
\label{W_1}
&&W^{\pi} \bigl(t_j,X^{t,x;\alpha_1^\varepsilon,\beta_1}_{t_j}
\bigr)\nonumber\\
&&\qquad\le\sum_{k\ge1}I \bigl\{X^{t,x;\alpha_1^\varepsilon, \beta_1}_{t_j}
\in{\cal O}_k \bigr\}W^{\pi}(t_j,x_k)+L
\varepsilon
\nonumber
\\
&&\qquad \le \sum_{k\ge1}I \bigl\{X^{t,x;\alpha_1^\varepsilon,\beta_1
}_{t_j}
\in{\cal O}_k \bigr\}J^{\pi} \bigl(t_j,x_k;
\alpha_2^k,\beta_2 \bigr)+ (L+1)\varepsilon
\\
&&\qquad \le \sum_{k\ge1}I \bigl
\{X^{t,x;\alpha_1^\varepsilon, \beta_1}_{t_j} \in{\cal O}_k \bigr
\}J^{\pi} \bigl(t_j,X^{t,x;\alpha_1^\varepsilon, \beta_1}_{t_j};
\alpha_2^k,\beta_2 \bigr)+ (2L+1)\varepsilon
\nonumber\\
&&\qquad = J^{\pi} \bigl(t_j,X^{t,x;\alpha_1^\varepsilon, \beta
_1}_{t_j};
\alpha _2^\varepsilon, \beta_2 \bigr)+ (2L+1)
\varepsilon,\qquad P\mbox{-a.s.}
\nonumber
\end{eqnarray}
For our arbitrarily chosen $\beta\in{\cal B}_{t,T}^\pi$ we
put $\beta_2^\varepsilon(u_2):=\beta(u_1^\varepsilon\oplus
u_2)_{|[t_j,T]}\in{\cal V}_{t_j,T}^\pi$, $u_2\in{\cal
U}_{t_j,T}^\pi$. Obviously, $\beta_2^\varepsilon\in{\cal
B}_{t_j,T}^\pi$. Let us denote by
$(u_2^\varepsilon,v_2^\varepsilon)\in{\cal
U}_{t_j,T}^\pi\times{\cal V}_{t_j,T}^\pi$ the unique couple of
control processes associated with
$(\alpha_2^\varepsilon,\beta_2^\varepsilon)$ through Lemma
\ref{controls-NAD-strategies}. Then, defining
$\alpha^\varepsilon\in{\cal A}_{t,T}^\pi$ by setting
\[
\alpha^\varepsilon(v):=\alpha^\varepsilon_1(v_{|[t,t_j]})
\oplus\alpha^\varepsilon_2(v_{|(t_j,T]}),\qquad v\in{\cal
V}_{t,T}^\pi,
\]
we see that, for $(u^\varepsilon,v^\varepsilon
):=(u^\varepsilon_1
\oplus u^\varepsilon_2, v^\varepsilon_1\oplus v^\varepsilon_2)\in
{\cal U}_{t,T}^\pi\times{\cal V}_{t,T}^\pi$,
\begin{eqnarray*}
\alpha^\varepsilon\bigl(v^\varepsilon\bigr)&=&\alpha^\varepsilon
_1\bigl(v^\varepsilon_1\bigr) \oplus
\alpha^\varepsilon_2\bigl(v^\varepsilon_2\bigr)
=u^\varepsilon_1\oplus u^\varepsilon_2=u^\varepsilon,
\\
\beta^\varepsilon\bigl(u^\varepsilon\bigr)&=&\beta^\varepsilon
\bigl(u^\varepsilon _1\oplus u^\varepsilon_2
\bigr)=\beta_1\bigl(u^\varepsilon_1\bigr) \oplus
\beta^\varepsilon_2 \bigl(u^\varepsilon_2
\bigr)=v^\varepsilon_1\oplus v^\varepsilon
_2=v^\varepsilon.
\end{eqnarray*}
Consequently, with the choice $\beta_2=\beta_2^\varepsilon$,
we have
\begin{eqnarray}
W^{\pi} \bigl(t_j,X^{t,x;\alpha_1^\varepsilon,\beta_1}_{t_j}
\bigr) &\le &J^{\pi} \bigl(t_j,X^{t,x;\alpha_1^\varepsilon, \beta_1}_{t_j};
\alpha _2^\varepsilon, \beta_2^\varepsilon
\bigr)+ (2L+1)\varepsilon
\nonumber
\\
& = & J^{\pi} \bigl(t_j,X^{t,x;u_1^\varepsilon,v_1^\varepsilon
}_{t_j};u_2^\varepsilon,
v_2^\varepsilon \bigr)+ (2L+1)\varepsilon\nonumber \\
&=& Y_{t_j}^{t_j,X^{t,x;u_1^\varepsilon,v_1^\varepsilon}_{t_j};
u_2^\varepsilon,v_2^\varepsilon}+
(2L+1)\varepsilon
\\
& = & Y_{t_j}^{t_j,X^{t,x;u^\varepsilon,v^\varepsilon}_{t_j};
u^\varepsilon,v^\varepsilon}+ (2L+1)\varepsilon\nonumber\\
& =& Y_{t_j}^{t,x;u^\varepsilon,v^\varepsilon}+
(2L+1)\varepsilon,\qquad P\mbox{-a.s.}
\nonumber
\end{eqnarray}
Indeed, the fact that
$X^{t,x;u^\varepsilon_1,v^\varepsilon_1}_{t_j}$ is ${\cal F}^\pi_{t_j-}
=\widetilde{\cal F}^\pi_{t_j}$-measurable, allows to substitute
this random variable at the place of $x'$ in the BSDE for $(Y^{t_j,x';
u^\varepsilon_1,v^\varepsilon_1}_s,Z^{t_j,x';u^\varepsilon_1,
v^\varepsilon_1}_s$, $ M^{t_j,x';u^\varepsilon_1,
v^\varepsilon_1}_s)_{s\in[t_j,T]}$. The uniqueness of the solution of the
resulting BSDE then yields $ Y^{t_j,X^{t,x;u^\varepsilon,
v^\varepsilon};u^\varepsilon,v^\varepsilon}_s=Y_s^{t,x;u^\varepsilon,
v^\varepsilon}, s\in[t_j,T]$.
Combining the above result with that of step 1, and taking into account
the monotonicity and the Lipschitz properties of the backward stochastic
semigroup, which are a direct consequence of the corresponding properties
of the solutions of BSDEs (the proof of them is similar to the
classical case (e.g., refer to Peng \cite{Pe1}), also refer to \cite
{CFS}) we obtain
\begin{eqnarray}
\label{(2.14)}
\widetilde{W}_{t_j}^{\pi}(t,x) &
\le& G^{t,x;
\alpha^\varepsilon_1,\beta_1}_{t,t_j} \bigl(W^{\pi} \bigl(t_j,X_{t_j}^{t,
x;\alpha^\varepsilon_1,\beta_1}
\bigr) \bigr)+ \varepsilon
\nonumber
\\
& \le& G^{t,x;\alpha^\varepsilon_1,\beta_1}_{t,t_j} \bigl(Y_{t_j}^{t,x;u^\varepsilon,v^\varepsilon}+
(2L+1)\varepsilon \bigr) + \varepsilon
\nonumber\\
& \le& G^{t,x;u^\varepsilon_1,v^\varepsilon_1}_{t,t_j} \bigl(Y_{t_j}^{t,x;u^\varepsilon,v^\varepsilon}
\bigr)+ C\varepsilon
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
& = & G^{t,x;u^\varepsilon,v^\varepsilon}_{t,t_j} \bigl(Y_{t_j}^{t,x;
u^\varepsilon,v^\varepsilon}
\bigr)+ C\varepsilon
\nonumber\\
& = & Y_{t}^{t,x;u^\varepsilon,v^\varepsilon}+ C\varepsilon\nonumber
\\
& = & J^{\pi} \bigl(t,x; \alpha^\varepsilon,\beta \bigr)+ C
\varepsilon,\qquad P\mbox{-a.s., for all } \beta\in{\cal B}_{t,T}^\pi.
\nonumber
\end{eqnarray}
Therefore,
\begin{eqnarray}
\widetilde{W}_{t_j}^{\pi}(t,x)&\le&\esssup_{\alpha
\in
\mathcal{A}^\pi_{t,T}}
\essinf_{\beta\in
\mathcal{B}^\pi_{t,T}}J^{\pi}(t,x;\alpha,\beta)+C\varepsilon
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&=&
W^{\pi}(t,x)+C\varepsilon, \qquad P\mbox{-a.s.},
\end{eqnarray}
and considering the
arbitrariness of the choice of $\varepsilon>0$ we can conclude the
proof.
\end{pf}
In order to complete the proof of the DPP, we need still the
following lemma.
\begin{lemma}\label{DPP-2} Under our standard assumptions it holds
that $ \widetilde{W}_{t_j}^{\pi}(t,x)\ge W^{\pi}(t,x)$, $P$-a.s.
\end{lemma}
\begin{pf} The proof of this lemma uses mainly arguments which
have been already developed in the frame of the proof of the preceding
lemma. For this reason, we give here rather a sketch than
a detailed proof.
Let us begin with fixing an arbitrary $\alpha\in{\cal A}_{t,T}^\pi$.
Given any $v_2\in{\cal V}_{t_j,T}^\pi$ we define $\alpha_1\in{\cal
A}_{t,t_j}^\pi$ by setting $\alpha_1(v_1):=\alpha(v_1\oplus
v_2)_{|[t,t_j]}\in{\cal U}_{t,t_j}^\pi$, for $v_1\in{\cal
V}_{t,t_j}^\pi$. Thanks to the nonanticipativity property of the
elements of ${\cal A}_{t,t_j}^\pi$, $\alpha_1$ does not depend on
the particular choice of $v_2$. From the definition of
$ \widetilde{W}_{t_j}^{\pi}(t,x)$, it follows that
\[
\widetilde{W}_{t_j}^{\pi}(t,x)\ge \essinf_{\beta_1\in\mathcal{B}^\pi_{t,t_j}}G^{t,x;\alpha_1,
\beta_1}_{t,t_j}
\bigl(W^{\pi} \bigl(t_j,X_{t_j}^{t,x;\alpha_1,\beta_1}
\bigr) \bigr),
\]
$P$-a.s., for all $\alpha_1\in\mathcal{A}^\pi_{t,t_j}$, and
from the argument developed in step 1 of the proof of Lemma \ref{DPP-1} we know
that, for an arbitrarily given $\varepsilon>0$ there
exists $\beta_1^\varepsilon\in\mathcal{B}^\pi_{t,t_j}$ (depending
on $\alpha_1\in{\cal A}_{t,t_j}^\pi$) such that
\[
\widetilde{W}_{t_j}^{\pi}(t,x)\ge G^{t,x;
\alpha_1,\beta_1^\varepsilon}_{t,t_j}
\bigl(W^{\pi} \bigl(t_j,X_{t_j}^{t,x;
\alpha_1,\beta_1^\varepsilon}
\bigr) \bigr)-\varepsilon,\qquad P\mbox{-a.s.}
\]
In analogy to step 2 of the proof of Lemma \ref{DPP-1}, we estimate
the expression
$W^{\pi}(t_j,X_{t_j}^{t,x;\alpha_1,\beta_1^\varepsilon})$ to which the
backward stochastic semigroup is applied in the above estimate. For
this, we let $(u_1^\varepsilon,v_1^\varepsilon)\in{\cal
U}_{t,t_j}^\pi\times{\cal V}_{t,t_j}^\pi$ be the unique control
couple associated with $(\alpha_1,\beta_1^\varepsilon)$ through
Lemma \ref{controls-NAD-strategies}, and we define
$\alpha_2^\varepsilon(v_2):=\alpha(v_1^\varepsilon\oplus
v_2)_{[t_j,T]}, v_2\in{\cal V}_{t_j,T}^\pi$. Such defined
mapping $\alpha_2^\varepsilon\dvtx {\cal V}_{t_j,T}^\pi\rightarrow{\cal
U}_{t_j,T}^\pi$ belongs to ${\cal A}_{t_j,T}^\pi$, and using an
adaptation of the argument with the Borel partition ${\cal O}_k, k\ge1$, of $R^d$, from step 2 of the proof of Lemma \ref{DPP-1},
which leads to (\ref{W_1}), we construct an NAD-strategy
$\beta_2^\varepsilon\in{\cal B}_{t_j,T}^\pi$ such that
\begin{eqnarray}
W^{\pi} \bigl(t_j,X_{t_j}^{t,x;\alpha_1,\beta_1^\varepsilon}
\bigr) & \ge& \essinf_{\beta_2\in
\mathcal{B}^\pi_{t_j,T}}J^{\pi} \bigl(t_j,X_{t_j}^{t,x;\alpha_1,
\beta_1^\varepsilon};
\alpha_2^\varepsilon,\beta_2 \bigr)
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
& \ge& J^{\pi} \bigl(t_j,X_{t_j}^{t,x;\alpha_1,\beta_1^\varepsilon};
\alpha_2^\varepsilon,\beta_2^\varepsilon \bigr)-
\varepsilon, \qquad P\mbox{-a.s.}
\end{eqnarray}
Letting $(u_2^\varepsilon,v_2^\varepsilon)\in{\cal
U}_{t_j, T}^\pi\times{\cal V}_{t_j,T}^\pi$ be the unique control
couple associated with $(\alpha_2^\varepsilon,\beta_2^\varepsilon)$
through Lemma \ref{controls-NAD-strategies}, we observe that, for
$\beta^\varepsilon\in{\cal B}_{t,T}^\pi$ defined by the relation
$\beta^\varepsilon(u):=\beta^\varepsilon_1(u_{|[t,t_j]})\oplus
\beta^\varepsilon_2(u_{|(t_j,T]}), u\in{\cal U}_{t,T}^\pi$, we
have the couple of controls $u^\varepsilon:=u^\varepsilon_1\oplus
u^\varepsilon_2\in{\cal U}_{t,T}^\pi$,
$v^\varepsilon:=v^\varepsilon_1\oplus v^\varepsilon_2\in{\cal
V}_{t,T}^\pi$ associated with $(\alpha,\beta^\varepsilon)$ through
Lemma~\ref{controls-NAD-strategies}:
\begin{eqnarray*}
\alpha\bigl(v^\varepsilon\bigr)&=&\alpha\bigl(v_1^\varepsilon
\oplus v_2^\varepsilon\bigr) =\alpha_1
\bigl(v_1^\varepsilon\bigr)\oplus\alpha_2^\varepsilon
\bigl(v_2^\varepsilon\bigr) =u_1^\varepsilon
\oplus u_2^\varepsilon=u^\varepsilon,
\\
\beta^\varepsilon\bigl(u^\varepsilon\bigr)&=&\beta^\varepsilon
_1\bigl(u_1^\varepsilon\bigr) \oplus
\beta^\varepsilon_2\bigl(u_2^\varepsilon
\bigr)=v_1^\varepsilon\oplus v_2^\varepsilon=v^\varepsilon.
\end{eqnarray*}
Consequently, thanks to the monotonicity and Lipschitz
properties of the backward stochastic semigroup, we have
\begin{eqnarray}
\label{(2.17)}
\widetilde{W}_{t_j}^{\pi}(t,x) &
\ge& G^{t,x;\alpha_1,\beta_1^\varepsilon}_{t,t_j} \bigl(W^{\pi} \bigl(t_j,X_{t_j}^{t,x;\alpha_1,\beta_1^\varepsilon}
\bigr) \bigr)-\varepsilon
\nonumber
\\[-2pt]
& \ge& G^{t,x;\alpha_1,\beta_1^\varepsilon}_{t,t_j} \bigl(J^{\pi}
\bigl(t_j, X_{t_j}^{t,x;\alpha_1,\beta_1^\varepsilon}; \alpha_2^\varepsilon,
\beta_2^\varepsilon \bigr)-\varepsilon \bigr)-\varepsilon
\\[-2pt]
& \ge& G^{t,x;u_1^\varepsilon,v_1^\varepsilon}_{t,t_j} \bigl(Y_{t_j}^{t_j,
X_{t_j}^{t,x;u_1^\varepsilon,v_1^\varepsilon}; u_2^\varepsilon,
v_2^\varepsilon}
\bigr)-C\varepsilon
\nonumber
\\[-2pt]
& = & G^{t,x;u^\varepsilon,v^\varepsilon}_{t,t_j} \bigl(Y_{t_j}^{t_j,
X_{t_j}^{t,x;u^\varepsilon,v^\varepsilon}; u^\varepsilon,
v^\varepsilon}
\bigr)-C\varepsilon
\nonumber\\[-2pt]
& = & G^{t,x;u^\varepsilon,v^\varepsilon}_{t,t_j} \bigl(Y_{t_j}^{t,x;
u^\varepsilon, v^\varepsilon}
\bigr)-C\varepsilon
\nonumber
\\[-2pt]
& = & Y_{t}^{t,x; u^\varepsilon, v^\varepsilon}-C\varepsilon
\nonumber
\\[-2pt]
& = & Y_{t}^{t,x; \alpha,\beta^\varepsilon}-C\varepsilon,\qquad P\mbox{-a.s.}
\nonumber
\end{eqnarray}
We take in the latter estimate first the essential infimum
over $\beta\in{\cal B}_{t,T}^\pi$, and then the essential supremum
over all $\alpha\in{\cal A}_{t,T}^\pi$. Thus, by considering the
arbitrariness of $\varepsilon>0$, we get the statement of the
lemma.\vspace*{-2pt}
\end{pf}
As a consequence of the proof of the DPP, we get the following proposition.\vspace*{-2pt}
\begin{proposition}\label{proposition3.1} Under our standard
assumptions, for all $(t,x)\in[0,T]
\times R^d$, it holds
\begin{eqnarray}
\label{DPP-a}
W^{\pi}(t,x) &=& \sup_{\alpha\in
\mathcal{A}^\pi_{t,T}}
\inf_{\beta\in
\mathcal{B}^\pi_{t,T}}E \bigl[J^{\pi}(t,x; \alpha,\beta) \bigr],
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
U^{\pi}(t,x) &=& \inf_{\beta\in\mathcal{B}^\pi_{t,T}}\sup
_{\alpha
\in\mathcal{A}^\pi_{t,T}}E \bigl[J^{\pi}(t,x;\alpha,\beta) \bigr].\vspace*{-2pt}
\end{eqnarray}
\end{proposition}
By combining the above lemma with Remark \ref{classical}, we get the following
result under the classical assumption of a running payoff function not depending
on $(y,z)$:\vspace*{-2pt}
\begin{corollary} Let us suppose in addition to our standard
assumptions that
the coefficient $f(s,x,y,z,u,v)$ does not depend on $(y,z)$. Then, for all
$(t,x)\in[0,T]\times R^d$,
\begin{eqnarray}
\label{DPP-aaa}
W^{\pi}(t,x) &=& \sup
_{\alpha\in
\mathcal{A}^\pi_{t,T}} \inf_{\beta\in
\mathcal{B}^\pi_{t,T}}E \biggl[\Phi
\bigl(X^{t,x;u,
v}_T \bigr)\nonumber\\[-3pt]
&&\hspace*{98pt}{}+\int_t^Tf
\bigl(s,X^{t,x;u,v}_s,u_s,v_s \bigr)
\,ds \biggr],
\nonumber
\\[-9pt]
\\[-9pt]
\nonumber
U^{\pi}(t,x) &=& \inf_{\beta\in\mathcal{B}^\pi_{t,T}}\sup_{\alpha
\in\mathcal{A}^\pi_{t,T}}E
\biggl[\Phi \bigl(X^{t,x;u,
v}_T \bigr)\\[-3pt]
&&\hspace*{98pt}{}+\int_t^Tf
\bigl(s,X^{t,x;u,v}_s,u_s,v_s \bigr)
\,ds \biggr].\nonumber\vspace*{-2pt}
\end{eqnarray}
\end{corollary}
Now we prove the above Proposition \ref{proposition3.1}.\vspace*{-2pt}
\begin{pf} Let\vspace*{1pt} $(t,x)\in[0,T)\times R^d$, and $t_j\in\pi$ be such that
$t_j\le t<t_{j+1}$. As we have shown in the proof of
the DPP that $\widetilde{W}^\pi_{t_j}(t,x)$ and $W^\pi(t,x)$ coincide,
we see
from (\ref{(2.14)}) that, for every $\varepsilon>0$, there exists
$\alpha^\varepsilon\in{\cal A}_{t,T}^\pi$ such that, for all $\beta
\in
{\cal B}_{t,T}^\pi$,
\[
W^\pi(t,x)\le J^{\pi} \bigl(t,x;\alpha^\varepsilon,
\beta \bigr)+\varepsilon,\qquad P\mbox{-a.s.}
\]
Consequently, taking into account that $W^\pi(t,x)$ is deterministic,
we get $W^\pi(t, x)\le E[J^{\pi}(t,x;\alpha^\varepsilon,\beta
)]+\varepsilon$. By taking
first the infimum over all $\beta\in{\cal B}_{t,T}^\pi$ and after
the supremum
over $\alpha\in{\cal A}_{t,T}^\pi$, we obtain
\[
W^{\pi}(t,x)\le\sup_{\alpha\in\mathcal{A}^\pi_{t,T}} \inf_{\beta\in\mathcal{B}^\pi_{t,T}}E
\bigl[J^{\pi}(t,x;\alpha,\beta) \bigr].
\]
To get the converse relation, we observe that, due to (\ref{(2.17)}),
for every $\varepsilon>0$ and all $\alpha\in{\cal A}_{t,T}^\pi$, there
exists some
$\beta^\varepsilon\in{\cal B}_{t,T}^\pi$ such that
\[
W^\pi(t,x)\ge J^{\pi}\bigl(t,x;\alpha,\beta^\varepsilon
\bigr)-\varepsilon,\qquad P\mbox{-a.s.}
\]
By taking the expectation on both sides of this inequality,
after the
infimum with respect to $\beta^\varepsilon\in{\cal B}_{t,T}^\pi$ and,
at the end, the
supremum over $\alpha\in{\cal A}_{t,T}^\pi$, we obtain that
\[
W^{\pi}(t,x)\ge\sup_{\alpha\in\mathcal{A}^\pi_{t,T}} \inf_{\beta\in\mathcal{B}^\pi_{t,T}}E
\bigl[J^{\pi}(t,x;\alpha,\beta) \bigr].
\]
This proves the statement for $W^{\pi}(t,x)$; that for
$U^{\pi
}(t,x)$ can be proved
similarly.
\end{pf}
At the end of this section, let us still consider the H\"{o}lder
continuity of the lower and the upper value functions along the
partition with respect to the time.
\begin{proposition}\label{Hoelder}
Under our standard assumptions there exists a constant~$C$ which is
independent of
the underlying partition $\pi$ of the interval $[0,T]$, such that
\begin{eqnarray}
\bigl|W^\pi(t,x)-W^\pi(s,x) \bigr|+ \bigl|U^\pi(t,x)-U^\pi(s,x)
\bigr| \le C|t-s|^{1/2},
\nonumber
\\[-8pt]
\\[-8pt]
\eqntext{s,t\in[0,T], x\in R^d.}
\end{eqnarray}
\end{proposition}
\begin{pf} We restrict ourselves to the proof for $W^\pi$; that for
$U^\pi$ is analogous.\vadjust{\goodbreak}
\textit{Step} 1. Given a partition $\pi$ of the interval $[0,T]$,
let us suppose that $0\le t< s\le T$ and fix arbitrarily $\varepsilon>0$.
From the proof of Proposition \ref{proposition3.1}, we know
that there exists $\alpha^\varepsilon\in{\cal A}_{t,T}^\pi$ such that,
for all
$\beta\in{\cal B}_{t,T}^\pi$,
\begin{equation}
\label{a1}W^\pi(t,x)\le E \bigl[J^{\pi} \bigl(t,x;\alpha
^\varepsilon,\beta \bigr) \bigr]+\varepsilon.
\end{equation}
For any fixed $v^0\in V$ we let $v^0_1:=v^0I_{[t,s)}$. Then, for
$v_2\in
{\cal V}_{s,
T}^\pi$, $v^0_1\oplus v_2:=v^0I_{[t,s)}+v_2I_{[s,T]}\in{\cal
V}_{t,T}^\pi$, and
$\widetilde{\alpha}^\varepsilon(v_2):=\alpha^\varepsilon
(v^0_1\oplus
v_2)_{|[s,T]}
\in{\cal U}_{s,T}^\pi$. Moreover, it can be easily checked that such defined
mapping $\widetilde{\alpha}^\varepsilon$ belongs to ${\cal
A}_{s,T}^\pi
$. Again from
the proof of Proposition \ref{proposition3.1}, it follows that there is
$\widetilde{\beta}^\varepsilon\in{\cal B}_{s,T}^\pi$ such that
\begin{equation}
\label{a2}W^\pi(s,x)\ge E \bigl[J^{\pi} \bigl(s,x;
\widetilde { \alpha }^\varepsilon, \widetilde{\beta}^\varepsilon \bigr)
\bigr]- \varepsilon.
\end{equation}
Let $(u_2^\varepsilon,v_2^\varepsilon)\in{\cal U}^\pi
_{s,T}\times
{\cal V}^\pi_{s,T}$ be associated with $(\widetilde{\alpha
}^\varepsilon,
\widetilde{\beta}^\varepsilon)$ through Lemma \ref{controls-NAD-strategies}:
$\widetilde{\alpha}^\varepsilon(v_2^\varepsilon)=u_2^\varepsilon,
\widetilde{\beta}^\varepsilon(u_2^\varepsilon)=v_2^\varepsilon$,
$ds\,dP$-a.e.
on $[s,T]\times\Omega$.
On the other hand, let us define $\beta^\varepsilon(u):=
v_1^0\oplus\widetilde{\beta}^\varepsilon(u_{|[s,T]}), u\in{\cal U}_{t,
T}^\pi$. Obviously, $\beta^\varepsilon\in{\cal B}_{t,T}^\pi$.
Putting $u^\varepsilon:=\alpha^\varepsilon(v_1^0\oplus
v_2^\varepsilon
)\in
{\cal U}_{t,T}^\pi$, we deduce from the fact $u^\varepsilon_{|[s,T]}
=\alpha^\varepsilon(v_1^0\oplus v_2^\varepsilon)_{|[s,T]}=
\widetilde{\alpha}^\varepsilon(v_2^\varepsilon)=u_2^\varepsilon$,
that $(u^\varepsilon,v^\varepsilon:=v_1^0\oplus v_2^\varepsilon)
\in{\cal U}_{t,T}^\pi\times{\cal V}_{t,T}^\pi$ satisfies
\begin{eqnarray}
\alpha^\varepsilon \bigl(v^\varepsilon \bigr)&=&u^\varepsilon
\quad\mbox{and }
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
\beta^\varepsilon \bigl(u^\varepsilon \bigr)&=&v_1^0
\oplus\widetilde{\beta }^\varepsilon \bigl(u_2^\varepsilon
\bigr)=v_1^0\oplus v_2^\varepsilon=v^\varepsilon,
\end{eqnarray}
over the interval $[t,T]$, while over the smaller interval
$[s,T]$ it holds
\begin{eqnarray}
\widetilde{\alpha}^\varepsilon \bigl(v^\varepsilon_{|[s,T]}
\bigr) &=&\widetilde{\alpha}^\varepsilon \bigl(v_2^\varepsilon
\bigr)=u_2^\varepsilon= u^\varepsilon_{|[s,T]}\quad
\mbox{and }
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
\widetilde{\beta}^\varepsilon \bigl(u^\varepsilon_{|[s,T]}
\bigr) &=&\widetilde{\beta}^\varepsilon \bigl(u^\varepsilon_2
\bigr)=v^\varepsilon_2= v^\varepsilon_{|[s,T]}.
\end{eqnarray}
Consequently, from the relation (\ref{a1}) and (\ref{a2}) it
follows that
\begin{eqnarray}
\label{b1}
W^\pi(t,x) &\le& E \bigl[J^{\pi}
\bigl(t,x;u^\varepsilon,v^\varepsilon \bigr) \bigr]+ \varepsilon,
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
W^\pi(s,x) &\ge& E \bigl[J^{\pi}
\bigl(s,x;u^\varepsilon,v^\varepsilon \bigr) \bigr]-\varepsilon,
\end{eqnarray}
from where
\begin{eqnarray}
\label{b2}
& & W^\pi(t,x)-W^\pi(s,x)
\nonumber
\\
&& \qquad\le E \bigl[J^{\pi} \bigl(t,x;u^\varepsilon,v^\varepsilon
\bigr) -J^{\pi} \bigl(s,x;u^\varepsilon,v^\varepsilon \bigr)
\bigr]+2 \varepsilon
\\
&&\qquad \le E \bigl[ \bigl|Y_s^{t,x;u^\varepsilon,v^\varepsilon}- Y_s^{s,x;u^\varepsilon,v^\varepsilon}
\bigr| \bigr]+ \bigl|E \bigl[Y_t^{t,x;
u^\varepsilon,v^\varepsilon}-Y_s^{t,x;u^\varepsilon,v^\varepsilon}
\bigr] \bigr| +2\varepsilon.
\nonumber
\end{eqnarray}
We emphasize that, if $s\notin\pi$, unlike the classical
Markovian case we do not have here that $Y_s^{t,x;u^\varepsilon,
v^\varepsilon}=Y_s^{s,X_s^{t,x;u^\varepsilon,
v^\varepsilon};u^\varepsilon,v^\varepsilon}=J^{\pi
}(s,X_s^{t,x;u^\varepsilon,
v^\varepsilon};u^\varepsilon,v^\varepsilon)$. Indeed, here, if $s\in
(t_{j-1},t_j)$, then $X_s^{t,x;u^\varepsilon,
v^\varepsilon}$ is ${\cal F}^\pi_{s-}$-measurable, where
${\cal F}^\pi_{s-}={\cal F}^B_s\vee{\cal H}_j\supsetneqq{\cal F}^B_s
\vee{\cal H}_{j-1}=\widetilde{\cal F}^\pi_{s-}$, where the BSDE is
considered with respect to the filtration $\widetilde{\mathbb{F}}^\pi$.
However, from the both BSDEs
\begin{equation}
\cases{
dY^{t,x; u, v}_r = -E \bigl[f
\bigl(r,X^{t,x; u, v}_r, Y^{t,
x; u, v}_r,
Z^{t,x; u, v}_r,u_r, v_r \bigr) |
\widetilde{\cal F}^\pi _r \bigr]\,dr
\vspace*{2pt}\cr
\hspace*{52pt}{}+Z^{t,x;u, v}_r \,dB_r+dM^{t,x; u, v}_r,
\vspace*{2pt}
\cr
Y^{t,x; u, v}_T = E \bigl[\Phi
\bigl(X^{t,x; u, v}_T \bigr) | \widetilde{\cal F}^\pi_{T-}
\bigr] }
\end{equation}
and
\begin{equation}
\cases{
dY^{s,x; u, v}_r = -E \bigl[f
\bigl(r,X^{s,x; u, v}_r, Y^{s,
x; u, v}_r,
Z^{s,x; u, v}_r, u_r, v_r \bigr) |
\widetilde{\cal F}^\pi _r \bigr]\,dr\vspace*{2pt}\cr
\hspace*{52pt}{}+Z^{s,x;u, v}_r \,dB_r+dM^{s,x; u, v}_r,
\vspace*{2pt}
\cr
Y^{s,x; u, v}_T = E \bigl[\Phi
\bigl(X^{s,x; u, v}_T \bigr) | \widetilde{\cal
F}^\pi_{T-} \bigr],}
\end{equation}
both studied over the time interval $[s,T]$, we deduce with
standard BSDE estimates that (or, refer to \cite{CFS})
\begin{eqnarray}
\label{aa1}
&&E \bigl[ \bigl|Y_s^{t,x;u^\varepsilon,v^\varepsilon}-
Y_s^{s,x;u^\varepsilon,v^\varepsilon} \bigr|^2 \bigr]
\nonumber
\\
&&\qquad \le CE \Bigl[\sup_{r\in[s,T]} \bigl|X^{t,x; u^\varepsilon, v^\varepsilon}_r-X^{s,x; u^\varepsilon,
v^\varepsilon}_r
\bigr|^2 \Bigr]
\\
&&\qquad \le CE \bigl[ \bigl|X^{t,x; u^\varepsilon, v^\varepsilon}_s-x \bigr|^2
\bigr]\le C(s-t)
\nonumber
\end{eqnarray}
(Recall that the coefficients $\sigma$ and $b$ are bounded
and Lipschitz).
Thus, from BSDE (\ref{BSDE}), the boundedness of $f(s,x,y,0,u,v)$,
the Lipschitz continuity of $f(s,x,y,z,u,v)$ in $z$ as well as
(\ref{BSDE-estimates-1}),
\begin{eqnarray}
\label{b3}
& & W^\pi(t,x)-W^\pi(s,x)
\nonumber
\\
&&\qquad \le E \bigl[ \bigl|Y_s^{t,x;u^\varepsilon,v^\varepsilon}- Y_s^{s,x;u^\varepsilon,v^\varepsilon}
\bigr| \bigr]+ \bigl|E \bigl[Y_t^{t,x;
u^\varepsilon,v^\varepsilon}-Y_s^{t,x;u^\varepsilon,v^\varepsilon}
\bigr] \bigr| +2\varepsilon
\nonumber
\\
&&\qquad \le C(s-t)^{1/2}+ 2\varepsilon
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
& &\qquad\quad{}+ (s-t)^{1/2} \biggl(E \biggl[\int_t^s
\bigl|f \bigl(r,X_r^{t,x;u^\varepsilon
,v^\varepsilon}, Y_r^{t,x;u^\varepsilon,v^\varepsilon},Z_r^{t,x;u^\varepsilon,
v^\varepsilon},\\
&&\hspace*{253pt}{}u^\varepsilon_r,v^\varepsilon_r
\bigr) \bigr|^2\,dr \biggr] \biggr)^{1/2}
\nonumber
\\
&&\qquad \le C|s-t|^{1/2}+2\varepsilon,
\nonumber
\end{eqnarray}
for some constant $C$ not depending on $\pi$ and on
$\varepsilon$. Thus, in virtue of the arbitrariness of $\varepsilon>0$
we have
\[
W^\pi(t,x)-W^\pi(s,x)\le C|s-t|^{1/2}.
\]
\textit{Step} 2. Now, for the same partition
$\pi$, and the case $0\le t< s\le T$, we make a lower
estimate for $W^\pi(t,x)-W^\pi(s,x)$. For this we notice that, for
arbitrarily given $\varepsilon>0$ we can find
$\widetilde{\alpha}^\varepsilon\in{\cal A}_{s,T}^\pi$ such that,
for all
$\widetilde{\beta}\in{\cal B}_{s,T}^\pi$,
\begin{equation}
\label{c1}W^\pi(s,x)\le E \bigl[J^\pi \bigl(s,x;
\widetilde {\alpha }^\varepsilon, \widetilde{\beta} \bigr) \bigr]+\varepsilon.
\end{equation}
For any fixed $u^0\in U$ we put $u^0_1:=u^0I_{[t,s)}$, and we define
$\alpha^\varepsilon\in{\cal A}_{t,T}^\pi$ by setting
$\alpha^\varepsilon(v):=u_1^0\oplus\widetilde{\alpha}^\varepsilon
(v_{|[s,T]}), v\in{\cal V}_{t,T}^\pi$. Let $\beta^\varepsilon\in
{\cal B}_{t,T}^\pi$ such that
\begin{equation}
\label{c2}W^\pi(t,x)\ge E \bigl[J^\pi \bigl(t,x;\alpha
^\varepsilon, \beta^\varepsilon \bigr) \bigr]-\varepsilon,
\end{equation}
and let $(u^\varepsilon,v^\varepsilon)\in{\cal
U}_{t,T}^\pi
\times{\cal V}_{t,T}^\pi$ be associated with $(\alpha^\varepsilon,
\beta^\varepsilon)$ through Lemma \ref{controls-NAD-strategies}. On
the other hand, by defining $\widetilde{\beta}^\varepsilon\in
{\cal B}_{s,T}^\pi$ by putting $\widetilde{\beta}^\varepsilon(u_2)=
\beta^\varepsilon(u_1^0\oplus u_2)_{|[s,T]}, u_2\in{\cal
U}_{s,T}^\pi
$, it can be easily verified
that $(u^\varepsilon_{|[s,T]},v^\varepsilon_{|[s,T]})\in{\cal
U}_{s,T}^\pi
\times{\cal V}_{s,T}^\pi$ is associated with
$(\widetilde{\alpha}^\varepsilon,\widetilde{\beta}^\varepsilon)$
in the
sense of Lemma \ref{controls-NAD-strategies}. Consequently,
\begin{eqnarray}
\label{b1}
W^\pi(s,x) &\le& E \bigl[J^\pi
\bigl(s,x;u^\varepsilon,v^\varepsilon \bigr) \bigr]+ \varepsilon,
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
W^\pi(t,x) &\ge& E \bigl[J^\pi
\bigl(t,x;u^\varepsilon,v^\varepsilon \bigr) \bigr]-\varepsilon,
\end{eqnarray}
and we can proceed now in analogy to step 1 to
deduce that
\[
W^\pi(t,x)-W^\pi(s,x)\ge-C|s-t|^{1/2}.
\]
Combining this result with that of step 1 we complete the proof.
\end{pf}
\section{Value in mixed strategies and associated HJB--Isaacs equation}\label{sec4}
The objective of this section is to study the limit of the lower and
the upper value functions $W^\pi$ and $U^\pi$ along a partition
$\pi$, when the mesh of the partition $\pi$ tends to zero, and to
show that both $W^\pi$ and $U^\pi$ converge uniformly on compacts to
the same limit function $V$ which is the unique viscosity solution
of the following Hamilton--Jacobi--Bellman--Isaac equation
\begin{equation}
\label{HJBI}\quad\hspace*{8pt}\cases{
\displaystyle\frac{\partial}{\partial t}V(t,x)+H \bigl(t,x,
\bigl(V,DV,D^2V \bigr) (t,x) \bigr)=0&\quad $(t,x)\in[0,T)\times
R^d$, \vspace*{2pt}
\cr
V(T,x)=\Phi(x),&\quad $x\in R^d$,
}\hspace*{-6pt}
\end{equation}
with Hamiltonian
\begin{eqnarray}
&&H(t,x,y,p,A) \nonumber\\
&&\qquad= \sup_{\mu\in{\cal P}(U)} \inf
_{\nu\in{\cal P}(V)}
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&&\quad\qquad{}\times\int_{U\times V} \biggl(
\frac{1}{2}\operatorname{tr} \bigl(\sigma\sigma^{T}(t,x, u, v)A \bigr)+ b(t, x, u, v)p
\\
&&\hspace*{102pt}{}+ f \bigl(t, x,y,p\cdot\sigma(t,x,u, v),u, v \bigr)
\biggr)\mu\otimes \nu(du\,dv),\nonumber
\end{eqnarray}
$(t,x,y,p,A)\in[0,T]\times R^d\times R\times R^d\times S^d$,
where $S^d$ denotes the space of symmetric matrices from
$R^{d\times d}$. For this we need the following supplementary assumption
which is coherent with our standard assumptions on the coefficients
$\sigma,b$ and~$f$.
\begin{condition}\label{condition} We suppose that either
\begin{itemize}
\item $\sigma(s,x,u,v)=\sigma(s,x), (s,x,u,v)
\in[0,T]\times R^d\times U\times V$ is independent of the controls;
or
\item $f(s,x,y,z,u,v)$ is linear in $z$:
\[
f(s,x,y,z,u,v)=f_0(s,x,y,u,v)+f_1(s)z,
\]
$(s,x,y,z,u,v)\in[0,T]\times R^d\times R\times R^d\times
U\times V$, where $f_0=(f_0(s,x,y,u,v))\dvtx\break [0,T]\times R^d\times R\times
U\times V \rightarrow R$ bounded, jointly continuous and
Lipschitz in $(x,y)$, uniformly with respect to $(s,u,v)$, and
$f_1\dvtx [0,T]\rightarrow R^d$ is continuous.\vspace*{-2pt}
\end{itemize}
\end{condition}
More precisely, we have the following
theorem.\vspace*{-2pt}
\begin{theorem}\label{main result} Under our standard assumptions on the
coefficients $\sigma,b,f$ and $\Phi$ as well as Condition \ref{condition},
we have the existence of a bounded, continuous function $V\dvtx [0,T]\times
R^d\rightarrow R$ such that, for every sequence of partitions $\pi_n$,
$n\ge1$, of the interval $[0,T]$ with mesh $|\pi_n|\rightarrow0$, as
$n\rightarrow+\infty$, $W^{\pi_n}\rightarrow V$, and $U^{\pi_n}
\rightarrow V$, uniformly on compacts, as $n\rightarrow+\infty$.
Moreover, $V$ is the viscosity solution of PDE~(\ref{HJBI}), unique in the
class of continuous functions with polynomial growth.\vspace*{-2pt}
\end{theorem}
For the convenience of the reader, we recall briefly the definition of
a viscosity solution, which we give directly for PDE (\ref{HJBI}). The reader
interested in a more detailed description of the concept of viscosity solution
is referred to the overview paper by Crandall, Ishii and Lions \cite{CIL}.\vspace*{-2pt}
\begin{definition}\label{de4.1} A function $V\in C([0,T]\times R^d)$ is said to be:
(i) a viscosity subsolution of PDE (\ref{HJBI}), if, first,
$V(T,x)\le
\Phi(x), x\in R^d$, and if, second, for any $(t,x)\in[0,T)\times R^d$
and any test function $\varphi\in C^{1,2}([0,T]\times R^d)$ such that
$V-\varphi$ achieves a local maximum at $(t,x)$, it holds
\begin{equation}
\frac{\partial}{\partial t}\varphi (t,x)+H \bigl(t,x, \bigl(\varphi, \nabla
\varphi,D^2\varphi \bigr) (t,x) \bigr)\ge0;
\end{equation}
(ii) a viscosity supersolution of PDE (\ref{HJBI}), if, first,
$V(T,x)\ge
\Phi(x), x\in R^d$, and if, second, for any $(t,x)\in[0,T)\times R^d$
and any test function $\varphi\in C^{1,2}([0,T]\times R^d)$ such that
$V-\varphi$ achieves a local minimum at $(t,x)$, it holds
\begin{equation}
\frac{\partial}{\partial t}\varphi (t,x)+H \bigl(t,x, \bigl(\varphi, \nabla
\varphi,D^2\varphi \bigr) (t,x) \bigr)\le0;
\end{equation}
(iii) a viscosity solution of (\ref{HJBI}) if it is both a viscosity
sub- but also a viscosity supersolution of (\ref{HJBI}).\vspace*{-2pt}
\end{definition}
\begin{remark}Let us point out that in Definition \ref{de4.1} the
space\break
$C^{1,2}([0,T]\times R^d)$ of the test functions can be replaced by any
subspace containing $C^\infty([0,T]\times R^d)$,\vadjust{\goodbreak} as long as one can
show the uniqueness with the help of $C^\infty$-test functions, as, for
instance, done in \cite{CIL}. Thus, our uniqueness results allows to
restrict to a class of test functions, more adapted for our
computations, the space $C^{3}([0,T]\times R^d)$ of functions which are
three times continuous differentiable with respect to $(t,x)$. On the
other hand, taking into account the uniform boundedness of the
functions $W^\pi, U^\pi$ and, hence, also of $V$, the standard
argument of changing a test function $\varphi\in C^{3}([0,T]\times
R^d)$ such that $V-\varphi$ achieves a local extremum at $(t,x)$, at
the exterior of a small ball around $(t,x)$, allows to consider only
test functions $\varphi\in C^{3}_{\ell,b}([0,T]\times R^d)$, that is,
$C^{3}$-functions with bounded derivatives of orders 1, 2 and 3 (and
which themselves have, consequently, a linear growth).
\end{remark}
Following the arguments developed, for example, in Str\"{o}mberg \cite
{ST} Theorem~5, we have the following comparison principle.
\begin{proposition}\label{comparison principle} Let us suppose our
standard assumptions on the coefficients $\sigma,b,f$ and $\Phi$, and
let $V_1, V_2\dvtx [0,T]\times R^d\rightarrow R$ be continuous functions having a growth
not exceeding that of $\exp\{\gamma|x|\}$, for some $\gamma>0$.
Then, if $V_1$ is a viscosity subsolution and $V_2$ a viscosity
supersolution of (\ref{HJBI}), we have $V_1(t,x)\le V_2(t,x), (t,x)
\in[0,T]\times R^d$.
\end{proposition}
\begin{remark}Let us emphasize that the condition of exponential growth is
optimal for the uniqueness of the continuous viscosity solution, as
long as $\sigma$ is bounded; this is the case due to our assumptions.
However, the assumption of bounded coefficients and so, in particular,
that of $\sigma$, has been imposed in order to simplify our argument.
Our approach can be extended without major difficulties to coefficients
$\sigma$ of linear growth. In this case the class of continuous
functions $V$ within which one has the uniqueness of the viscosity
solution is smaller than that of the above Proposition \ref{comparison principle}; it's that
of $V$ such that, for some $\gamma>0$,
\begin{eqnarray}
\lim_{|x|\rightarrow+\infty}V(t,x)\exp \bigl\{-\gamma \bigl(\log \bigl(|x|+1 \bigr)
\bigr)^2 \bigr\} =0
\nonumber
\\[-8pt]
\\[-8pt]
\eqntext{\mbox{uniformly with respect to }t\in[0,T];}
\end{eqnarray}
see, for example, \cite{Buckdahn-Li-2008}.
\end{remark}
As a direct consequence of this comparison principle, we have
the following corollary.
\begin{corollary} PDE (\ref{HJBI}) has at most one continuous viscosity
solution $V\dvtx [0,T]\times R^d\rightarrow R$ with exponential growth,
that is, satisfying the condition that, for suitable $\gamma>0$,
\begin{eqnarray}
\lim_{|x|\rightarrow+\infty}V(t,x)\exp\{-\gamma |x|\} =0
\nonumber
\\[-8pt]
\\[-8pt]
\eqntext{\mbox{uniformly with respect to }t\in[0,T].}
\end{eqnarray}
In particular, uniqueness holds within the class of continuous functions
with polynomial growth.\vspace*{-2pt}
\end{corollary}
All what follows will be devoted to the proof of Theorem \ref{main result}.
The proof will be given through a sequel of auxiliary results.
Let us begin by choosing an arbitrary sequence of partitions $\pi_n:=
\{0=t_0^n<t_1^n<\cdots<t_{N_n}^n=T\}$, $n\ge1$, of the interval $[0,T]$
such that $|\pi_n|:=\sup_{1\le i\le N_n}(t_i-t_{i-1})\rightarrow0$,
as $n
\rightarrow+\infty$. Then, from Lemma \ref{Hoelder} and Proposition
\ref{Hoelder}, we see that the family of functions $(W^{\pi_n},U^{\pi_n}),
n\ge1$, is uniformly Lipschitz in $x$, uniformly with respect to $t$,
and H\"{o}lder continuous in $t$, uniformly with respect to $x$. Consequently,
the following result follows from the Arzel\`{a}--Ascoli theorem
combined with
a standard diagonalization argument.\vspace*{-2pt}
\begin{lemma}\label{limit_W_pi_n} There exists a subsequence of partitions,
which we denote again
by $(\pi_n)_{n\ge1}$, as well as bounded continuous functions $W, U\dvtx [0,T]
\times R^d\rightarrow R$ such that $(W^{\pi_n},U^{\pi_n})\rightarrow(W,U)$,
uniformly on compacts in $[0,T]\times R^d$. Moreover,
\begin{equation}
\label{estimate_W,U}\qquad \bigl|W(t,x)-W \bigl(t',x'
\bigr) \bigr|+ \bigl|U(t,x)-U \bigl(t',x' \bigr) \bigr|\le C \bigl(
\bigl|t-t' \bigr|^{1/2}+ \bigl|x-x'\bigr | \bigr),
\end{equation}
$(t,x), (t',x')\in[0,T]\times R^d$, where $C$ is a constant
which does not depend on the choice of the
sequence of partitions $\pi_n, n\ge1$.\vspace*{-2pt}
\end{lemma}
Although the functions $W,U$ given by the above lemma depend a priori
on the
choice of the sequence of partitions $\pi_n,n\ge1$, as well as on the
subsequence with respect to which $(W^{\pi_n},U^{\pi_n})$ converges,
we will
show later that $W,U$ are universal and coincide even.
Inspired by the approach in \cite{Buckdahn-Li-2008} we put,
for some arbitrarily chosen but fixed $\varphi\in C^3_{\ell,b} ([0,T]
\times
{\mathbb{R}}^d)$,
\begin{eqnarray}
\qquad F(s,x,y,z,u, v)&=&f \bigl(s, x, y+\varphi(s,x), z+
D\varphi(s,x)\cdot
\sigma (s,x,u, v),u, v \bigr)
\nonumber
\\[-9pt]
\\[-9pt]
\nonumber
& &{}+ {\cal L}\varphi(s,x,u,v),
\end{eqnarray}
$(s,x,y,z,u, v)\in[0,T] \times{\mathbb{R}}^d \times
{\mathbb{R}} \times{\mathbb{R}}^d \times U \times V$, where
\begin{eqnarray}
&&{\cal L}\varphi(s,x,u,v)
\nonumber
\\[-9pt]
\\[-9pt]
\nonumber
&&\qquad:=\frac{\partial}{\partial
s}\varphi (s,x) + \frac{1}{2}
\operatorname{tr} \bigl(\sigma\sigma^{T}(s,x, u, v)D^2
\varphi \bigr)+ D \varphi\cdot b(s,x,u, v).
\end{eqnarray}
Let us now fix arbitrarily $(t,x)\in[0,T)\times R^d$. Given an arbitrary
partition $\pi=\{0=t_0<t_1<\cdots<t_n=T\}$, we let $1\le j\le n$
be such that $t<t_j$. Let us investigate the following
BSDE defined on the interval
$[t,t_j] \dvtx $
\begin{eqnarray}
\label{BSDE1appr}
&&dY^{1,u,v}_s = -E \bigl[{
\cal L}\varphi \bigl(s,X^{t,x;u,v}_s,u_s,v_s
\bigr)|\widetilde{\cal F}^\pi_{s} \bigr]\,ds\nonumber\\[-2pt]
&&\hspace*{44pt}{}-E \bigl[f
\bigl(s,X^{t,x;u,v}_s, Y^{1,u,v}_s +E
\bigl[\varphi \bigl(s,X^{t,x;u,v}_s \bigr)|\widetilde{\cal
F}^\pi _{s} \bigr],Z^{1,u,v}_s\nonumber\\
&&\hspace*{67pt}{}+E
\bigl[\nabla\varphi \bigl(s,X^{t,
x;u,v}_s \bigr)\sigma
\bigl(s,X^{t,x;u,v}_s,u_s,v_s \bigr) |
\widetilde{\cal F}^\pi _{s} \bigr], u_s,v_s
\bigr)|\widetilde{\cal F}^\pi_{s} \bigr]\,ds
\\
&&\hspace*{44pt}{}+Z^{1,u,v}_s\,dB_s+dM^{1,u,v}_s,\nonumber\\
\eqntext{ Y^{1,u,v}_{t_j} = 0, M^{1,u,v} \mbox{ martingale orthogonal to } B,
M^{1,u,v}_t=0,}
\end{eqnarray}
where the process $X^{t,x;u,v}$ is the unique solution of SDE
(\ref{SDE}) and $(u,v)\in{\cal U}_{t, t_j}^\pi\times{\cal V}_{t,
t_j}^\pi$.
It can be easily verified that (or, refer to \cite{CFS}), under our
standard assumptions on the coefficients
$\sigma,b$ and $f$, the above BSDE has a unique solution $(Y^{1,u,v},Z^{1,u,
v},M^{1,u,v})$ over the time interval $[t,t_j]$.
We have the following relation between the solution $Y^{1,u,v}$
and the backward stochastic semigroup $G^{t,x;u,v}_{s,t_j} [\varphi(t_j,
X^{t,x;u,v}_{t_j})]$:\vspace*{-2pt}
\begin{lemma}\label{lemma_BSDE1} For every $s\in[t,t_j]$, it holds
\begin{equation}\qquad
Y^{1,u,v}_s = G^{t,x;u,v}_{s,t_j} \bigl[
\varphi \bigl(t_j,X^{t,x;u,v}_{t_j} \bigr) \bigr] -E
\bigl[\varphi \bigl(s,X^{t,x;u,v}_s \bigr) | \widetilde{\cal
F}^\pi_{s} \bigr], \qquad P\mbox{-a.s.,}
\end{equation}
and in particular, for $s=t$,
\begin{equation}
Y^{1,u,v}_t = G^{t,x;u,v}_{t,t_j} \bigl[\varphi
\bigl(t_j,X^{t,x;u,v}_{t_j} \bigr) \bigr] -
\varphi(t,x), \qquad P\mbox{-a.s.}\vspace*{-2pt}
\end{equation}
\end{lemma}
\begin{pf} Recall that $G^{t,x;u,v}_{s,t_j}[\varphi(t_j, X^{t,x;
u,v}_{t_j})]$ is defined through the BSDE
\begin{equation}
\label{ab1}\qquad \cases{
dY^{u,v}_s = -E \bigl[f
\bigl(s,X^{t,x;u,v}_s, Y^{u,v}_s,Z^{u,v}_s,u_s,v_s
\bigr) |\widetilde{\cal F}_s^\pi \bigr]\,ds
\vspace*{2pt}\cr
\hspace*{38pt}{}+Z^{u,v}_s \,dB_s +dM^{u,v}_s,
\vspace*{2pt}
\cr
Y^{u,v}_{t_j} = E \bigl[\varphi
\bigl(t_j,X^{t,x;u,v}_{t_j} \bigr) | \widetilde{\cal
F}^\pi_{t_j} \bigr],\qquad s \in[t,t_j],
\vspace*{2pt}
\cr
M^{u,v} \mbox{ square integrable martingale, orthogonal
to }B, M^{u,v}_t=0, }
\end{equation}
by the relation:
\begin{equation}
G^{t,x;u,v}_{s,t_j} \bigl[\varphi \bigl(t_j,X^{t,x;u,v}_{t_j}
\bigr) \bigr] =Y^{u,v}_s, \qquad s\in[t,t_j].
\end{equation}
We notice that, since $X^{t,x;u,v}$ is $\mathbb{F}^\pi$-adapted, we have
\begin{equation}
E \bigl[\varphi \bigl(s,X^{t,x;u,v}_s \bigr) | \widetilde{\cal
F}^\pi_{s} \bigr]= E \bigl[\varphi \bigl(s,X^{t,x;u,v}_s
\bigr) | {\cal F}^B_{T}\vee{\cal H}_{\ell-1}
\bigr],
\end{equation}
$s\in[t\vee t_{\ell-1},t\vee t_\ell), 1\le\ell\le j$. Hence,
with the help of the It\^{o} formula we obtain on each
interval $[t\vee t_{\ell-1},t\vee t_\ell), 1\le\ell\le j$,
\begin{eqnarray}
&&dE \bigl[\varphi \bigl(s,X^{t,x;u,v}_s \bigr) |
\widetilde{\cal F}^\pi_{s} \bigr]\nonumber\\
&&\qquad =E \bigl[{\cal L}
\varphi \bigl(s,X^{t,x;u,v}_s,u_s,v_s
\bigr) | \widetilde{\cal F}^\pi_{s} \bigr]\,ds
\\
& &\qquad\quad {}+E \bigl[\nabla\varphi \bigl(s,X^{t,
x;u,v}_s
\bigr)\sigma \bigl(s,X^{t,x;u,v}_s,u_s,v_s
\bigr) | \widetilde{\cal F}^\pi _{s} \bigr]
\,dB_s.\nonumber
\end{eqnarray}
Let us put
\[
M_s:=\sum_{\ell\dvtx t<t_\ell\le s}\triangle E \bigl[
\varphi \bigl(t_\ell, X^{t,x;u,v}_{t_\ell} \bigr)|
\widetilde{\cal F}^\pi_{t_\ell} \bigr],\qquad s
\in[t,t_j],
\]
with
\[
\triangle E\bigl[\varphi\bigl(t_\ell, X^{t,x;u,v}_{t_\ell}
\bigr)|\widetilde{\cal F}^\pi_{t_\ell}\bigr]= E\bigl[\varphi
\bigl(t_\ell,X^{t,x;u,v}_{t_\ell}\bigr)|\widetilde{\cal
F}^\pi _{t_\ell}\bigr]- E\bigl[\varphi\bigl(t_\ell,X^{t,x;u,v}_{t_\ell}
\bigr)|\widetilde{\cal F}^\pi _{t_\ell-}\bigr].
\]
Obviously, $M$ is a pure jump martingale with respect to the
filtration $\mathbb{F}^B$ and, hence, orthogonal to $B$, and
\begin{eqnarray}
& &dE \bigl[\varphi \bigl(s,X^{t,x;u,v}_s \bigr) |
\widetilde{\cal F}^\pi_{s} \bigr]
\nonumber
\\
&&\qquad= E \bigl[{\cal L}\varphi \bigl(s,X^{t,x;u,v}_s,u_s,v_s
\bigr) | \widetilde{\cal F}^\pi_{s} \bigr]\,ds
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
& &\qquad\quad{} +E \bigl[\nabla\varphi \bigl(s,X^{t,
x;u,v}_s
\bigr)\sigma \bigl(s,X^{t,x;u,v}_s,u_s,v_s
\bigr) | \widetilde{\cal F}^\pi _{s} \bigr]
\,dB_s+dM_s, \\
\eqntext{s\in[t,t_j].}
\end{eqnarray}
Consequently, $(Y^{u,v}_s-E[\varphi
(s,X^{t,x;u,v}_s)|\widetilde{\cal F}^\pi_{s}],Z^{u,v}_s-
E [\nabla\varphi(s,X^{t,x;u,v}_s)\sigma(s, X^{t,x;u,v}_s,\break u_s,v_s)
|\widetilde{\cal F}^\pi_{s} ],
M^{u,v}_s-M_s), t\le s\le t_j$, is a solution
of BSDE (\ref{BSDE1appr}). From its uniqueness, we can conclude the
statement of the lemma.
\end{pf}
Let us now simplify the preceding BSDE (\ref{BSDE1appr}) by replacing the
process $X^{t,x;u,v}$ by its initial value $x$. Then BSDE (\ref{BSDE1appr})
takes the form
\begin{equation}
\label{BSDE2appr}\qquad \cases{
dY^{2,u,v}_s = -E
\bigl[F \bigl(s,x,Y^{2,u,v}_s,Z^{2,u,v}_s,u_s,
v_s \bigr)| \widetilde{\cal F}^\pi_s \bigr]
\,ds\vspace*{2pt}\cr
\hspace*{45pt}{}+Z^{2,u,v}_s \,dB_s+dM_s^{2,u,v},
\vspace*{2pt}
\cr
Y^{2,u,v}_{t_j} = 0, \qquad s
\in[t,t_j], \vspace*{2pt}
\cr
M^{u,v} \mbox{ square
integrable martingale, orthogonal to }B, M^{u,v}_t=0,}
\end{equation}
where $(u,v)\in{\cal U}_{t,t_j}^\pi\times{\cal
V}_{t,t_j}^\pi
$. As in
the discussion of BSDE (\ref{BSDE1appr}) we see that the above BSDE has
a unique solution. From the BSDEs (\ref{BSDE1appr}) and
(\ref{BSDE2appr}), we have
the following lemma.
\begin{lemma}\label{lemma_BSDE2} For every $(u,v)\in{\cal
U}_{t,t_j}^\pi\times{\cal V}_{t,
t_j}^\pi$ we have
\begin{equation}
\bigl|Y^{1,u,v}_t-Y^{2,u,v}_t \bigr| \leq
C(t_j-t)^{{3}/{2}},\qquad P\mbox{-a.s.},
\end{equation}
where C is independent of the control processes $u$ and $v$, but also
independent of the
partition $\pi$.
\end{lemma}
\begin{pf} Let $(u,v)\in{\cal U}_{t,t_j}^\pi\times{\cal V}_{t,
t_j}^\pi$. Then, for all $s\in[t,t_j]$, thanks to Condition~\ref{condition},
\begin{eqnarray}
& & E \bigl[{\cal L}\varphi(s,x,u_s,v_s)
\nonumber\\
&&\hspace*{12pt}{}+f \bigl(s,x,y+\varphi(s,x),z+ E \bigl[\nabla\varphi(s,x) \sigma(s,x,u_s,v_s)|
\widetilde{\cal F}^\pi_s \bigr],u_s,
v_s \bigr)|\widetilde{\cal F}^\pi_s \bigr]
\nonumber
\\
&&\qquad= E \bigl[{\cal L}\varphi(s,x,u_s,v_s)\\
&&\hspace*{44pt}{} +f
\bigl(s,x,y+\varphi(s,x),z+ \nabla\varphi(s,x)\sigma(s,x,u_s,v_s),u_s,
v_s \bigr)|\widetilde{\cal F}^\pi_s \bigr]
\nonumber\\
& &\qquad = E \bigl[F(s,x,y,z,u_s,v_s)|\widetilde{\cal
F}^\pi_s \bigr],\qquad P\mbox{-a.s.}
\nonumber
\end{eqnarray}
Consequently, we have to compare the solution of BSDE
(\ref{BSDE1appr})
\begin{eqnarray}
&&dY^{1,u,v}_s\nonumber\\
&&\qquad =-E \bigl[{\cal L}\varphi
\bigl(s,X^{t,x;u,v}_s,u_s,v_s \bigr)\nonumber\\
&&\qquad\hspace*{30pt}{}+f
\bigl(s,X^{t,x;u,v}_s,Y^{1,u,v}_s+E \bigl[
\varphi \bigl(s,X^{t,x;u,v}_s \bigr)|\widetilde{\cal
F}^\pi_{s} \bigr], Z^{1,u,v}_s
\\
&&\qquad\hspace*{51pt}{}+E
\bigl[\nabla\varphi \bigl(s,X^{t,
x;u,v}_s \bigr)\sigma
\bigl(s,X^{t,x;u,v}_s,u_s,v_s \bigr) |
\widetilde{\cal F}^\pi _{s} \bigr], u_s,v_s
\bigr)|\widetilde{\cal F}^\pi_{s} \bigr]\,ds
\nonumber\\
&&\hspace*{10pt}\qquad{}+Z^{1,u,v}_s\,dB_s+dM^{1,u,v}_s,\qquad
Y^{1,u,v}_{t_j} =0,\nonumber
\end{eqnarray}
with that of BSDE (\ref{BSDE2appr}) which can be rewritten as
\begin{eqnarray}
dY^{2,u,v}_s & = & -E \bigl[{\cal L}
\varphi(s,x,u_s,v_s)\nonumber\\
&&\hspace*{19pt}{}+f \bigl(s,x,Y^{2,u,v}_s
+E \bigl[\varphi(s,x)|\widetilde{\cal F}^\pi_{s} \bigr],
Z^{2,u,v}_s
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&&\hspace*{41pt}{} +E \bigl[\nabla\varphi(s,x)
\sigma(s,x,u_s,v_s) |\widetilde{\cal
F}^\pi_{s} \bigr], u_s,v_s
\bigr)|\widetilde{\cal F}^\pi_{s} \bigr]\,ds
\\
&&{}+Z^{2,u,v}_s\,dB_s+ dM^{2,u,v}_s,
\qquad
Y^{2,u,v}_{t_j} = 0,\nonumber
\end{eqnarray}
and from BSDE standard estimates we deduce
\begin{eqnarray}
&& \bigl|Y^{1,u,v}_t-Y^{2,u,v}_t
\bigr|^2+E \biggl[\int_t^{t_j}
\bigl|Z^{1,u,v}_r-Z^{2,u,v}_r
\bigr|^2\,dr\Big| \widetilde{\cal F}_t^\pi \biggr]\nonumber\\
&&\quad{} +E
\biggl[\sum_{\ell\le j; t<t_\ell} \bigl|\triangle M^{1,u,v}_{t_\ell}-
\triangle M^{2,u,v}_{t_\ell
} \bigr|^2\Big| \widetilde {\cal
F}_t^\pi \biggr]
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&&\qquad \le CE \biggl[ \biggl(\int_t^{t_j}\bigl|X^{t,x;u,v}_r-
x\bigr|\,dr \biggr)^2\Big|\widetilde{\cal F}_t^\pi
\biggr]
\\
&&\qquad \le C(t_j-t)^3,\qquad P\mbox{-a.s.},\nonumber
\end{eqnarray}
where the constant $C$ depends only on the boundedness
and Lipschitz constants of the coefficients and the derivatives of
$\varphi$, but not on $j$ nor the considered partition~$\pi$.
\end{pf}
Let us now state the following crucial lemma which, although
inspired by Lemma 4.3 in \cite{Buckdahn-Li-2008}, differs heavily
because of the different framework studied here.
\begin{lemma}\label{lemma_BSDE3} Let $Y^0=(Y^0_s)_{s\in[t,t_j]}$ denote
the unique solution of
the following ordinary backward differential equation:
\begin{equation}
\cases{- {\dot{Y}}^0_s = F_0
\bigl(s,x,Y^0_s,0 \bigr), &\quad $s \in[t,t_j]$,
\vspace*{2pt}
\cr
Y^0_{t_j} = 0, }
\end{equation}
where, for $ (s,y,z)\in[t,t_j]\times R\times R^d$,
\begin{eqnarray}
\label{qq}
F_0(s,x,y,z)&: =& \sup_{\mu\in{\cal P}(U)}
\Bigl(\inf_{v\in V}F(s,x,y,z,\mu,v) \Bigr)
\nonumber
\\
& = & \sup_{\mu\in{\cal P}(U)} \Bigl(\inf_{\nu\in{\cal
P}(V)}F(s,x,y,z,
\mu,\nu) \Bigr) \\
&&{}\times\Bigl(=\inf_{\nu\in{\cal P}(V)}\sup_{\mu\in{\cal
P}(U)}F(s,x,y,z,
\mu,\nu) \Bigr).\nonumber
\end{eqnarray}
Then, for all $s\in[t,t_j]$, $P$-a.s.,
\begin{equation}
Y^0_s = \esssup_{u \in{\cal U}_{t, t_j}^\pi}
\essinf_{v \in
{\cal V}_{t,t_j}^\pi} Y^{2,u,v}_s
= \essinf_{v \in
{\cal V}_{t,t_j}^\pi}\esssup_{u \in{\cal U}_{t, t_j}^\pi}Y^{2,u,v}_s.
\end{equation}
\end{lemma}
\begin{pf}\textit{Step} 1. Given $(u,v)\in{\cal U}_{t,t_j}^\pi
\times{\cal V}_{t,t_j}^\pi$, let
$(Y^{2,u,v}, Z^{2,u,v}, M^{2,u,v})$ be the unique solution of BSDE
(\ref{BSDE2appr}). We recall that, for all $s\in[t\vee t_{\ell-1},
t_\ell), (1\le\ell\le j)$, $(Y^{2,u,v}_s, Z^{2,u,v}_s,M^{2,u,v}_s)$
is $\widetilde{\cal F}^\pi_s(={\cal F}_s^B\vee{\cal H}_{\ell
-1})$-measurable,
$u_s$ is ${\cal
F}^{\pi,1}_s(={\cal F}_s^B\vee{\cal H}_{\ell-1}\vee\sigma\{\zeta
_{\ell,
1}\})$-measurable and $v_s$ is ${\cal F}^{\pi,2}_s(={\cal
F}_s^B\vee{\cal H}_{\ell-1}\vee\sigma\{\zeta_{\ell, 2}\})$-measurable.
Consequently, knowing $\widetilde{\cal F}^\pi_s$, $u_s$ and $v_s$
are conditionally independent, and defining
\begin{eqnarray}
\mu_s^u(A):=P\bigl\{u_s\in A|
\widetilde{\cal F}^\pi_s\bigr\} , \qquad
\nu_s^v(B):=P\bigl\{v_s\in B | \widetilde{
\cal F}^\pi_s\bigr\}, \nonumber\\
\eqntext{A\in{\cal B} (U), B
\in{\cal B}(V),}
\end{eqnarray}
we have
\begin{eqnarray}
& & E \bigl[F \bigl(s,x,Y^{2,u,v}_s
,Z^{2,u,v}_s,u_s, v_s \bigr) |
\widetilde{\cal F}^\pi_s \bigr]
\nonumber
\\
&&\qquad = F \bigl(s,x,Y^{2,u,v}_s
,Z^{2,u,v}_s,\mu^u_s,
\nu^v_s \bigr) \\
&&\qquad\quad{}\times\biggl(:= \int_{U\times V}F
\bigl(s,x,Y^{2,u,v}_s ,Z^{2,u,v}_s,u',
v' \bigr) \mu^u_s\otimes
\nu^v_s \bigl(du'\,dv'
\bigr) \biggr).\nonumber
\end{eqnarray}
Indeed, this relation can be easily checked by considering
first instead of $F(s,x,Y^{2,u,v}_s,Z^{2,u,v}_s, u_s, v_s)$
integrands of the form $\xi_s f_1(u_s)f_2(v_s)$, $\xi_s\in L^\infty
(\Omega$, $\widetilde{\cal F}^\pi_s,P)$ and $f_1,f_2$
bounded Borel functions over $U$ and $V$, respectively, and applying
later a Monotonic Class theorem.
Hence, with the notation
$ F(s,x,y,z,\mu,v):=\int_UF(s,x,y,z,u',v)\mu(du')$, $
\mu\in{\cal P}(U)$, and with putting
\[
F_1(s,x,y,z,\mu):=\inf_{v\in V} F(s,x,y,z,\mu,v)
\Bigl(=\inf_{\nu\in{\cal P}(V)}F(s,x,y,z,\mu,\nu ) \Bigr),
\]
$(s,y,z,\mu)\in[0,T]\times R\times R^d\times{\cal P}(U)$,
we obtain
\begin{eqnarray}
&& E \bigl[F \bigl(s,x,Y^{2,u,v}_s
,Z^{2,u,v}_s,u_s, v_s \bigr) |
\widetilde{\cal F}^\pi_s \bigr]
\nonumber\\
&&\qquad= \int_{V} F \bigl(s,x,Y^{2,u,v}_s,Z^{2,u,v}_s,
\mu_s^u, v' \bigr)\nu^v_s
\bigl(dv' \bigr)
\\
& &\qquad\ge F_1 \bigl(s,x,Y^{2,u,v}_s,Z^{2,u,v}_s,
\mu_s^u \bigr), \qquad ds\,dP\mbox{-a.e.}\nonumber
\end{eqnarray}
Consequently, denoting by $(Y^{3,u},Z^{3,u},M^{3,u})\in{\cal
S}^2_{\widetilde{\mathbb{F}}^\pi}(t,t_j; R)\times
L_{\widetilde{\mathbb{F}}^\pi}^2(t,t_j;\break R^d)\times
{\cal M}^2_{\widetilde{\mathbb{F}}^\pi}(t,t_j;R)$ the unique solution
of the
BSDE
\begin{equation}
\label{BSDE3appr}\qquad \cases{
dY^{3,u}_s =
-F_1 \bigl(s,x,Y^{3,u}_s,Z^{3,u}_s,
\mu^u_s \bigr)\,ds\vspace*{2pt}\cr
\hspace*{38pt}{} +Z^{3,u}_s
\,dB_s+dM^{3,u}_s,\qquad s
\in[t,t_j], \vspace*{2pt}
\cr
Y^{3,u}_{t_j} = 0,
\vspace*{2pt}
\cr
M^{3,u} \mbox{ square integrable martingale, orthogonal
to }B, M^{3,u}_t=0,}
\end{equation}
we deduce from the comparison theorem for BSDEs (refer to
\cite{CFS}, for classical case it can be referred to \cite{Pe1}, or
\cite{Buckdahn-Li-2008}) that $Y^{2,u,v}_s
\ge Y^{3,u}_s, s\in[t,t_j]$, $P$-a.s., for all $v\in{\cal V}^\pi_{t,t_j}$.
For this, we observe that $F_1(s,x,y,z,\mu)$ is a jointly continuous
function over $[0,T]\times R^d\times R\times R^d\times{\cal P}(U)$,
which is
Lipschitz in $(y,z)$, uniformly with respect to $(s,x,\mu)$. Thus, taking
into account the arbitrariness of $v\in{\cal V}_{t,t_j}^\pi$, we deduce
\begin{equation}
\label{BSDE3_inequ} Y^{3,u}_s\le
\essinf_{v\in
{\cal
V}_{t,t_j}^\pi}Y^{2,u,v}_s, \qquad P\mbox{-a.s, } s
\in[t,t_j].
\end{equation}
Let us show that we have even equality in the above inequality.
For this we observe that, since the function $F$ is continuous over $[t,t_j]
\times R^d\times R\times R^d\times{\cal P}(U)\times V$, there exists a Borel
measurable function $v^*\dvtx [t,t_j]\times R\times R^d\times{\cal
P}(U)\rightarrow
V$ such that
\[
F_1(s,x,y,z,\mu)=\inf_{v\in V}F(s,x,y,z,\mu,v)= F
\bigl(s,x,y,z,\mu, v^*(s,y,z,\mu)\bigr),
\]
$(s,y,z,\mu)\in[t,t_j]\times R\times R^d\times{\cal P}(U)$.
With the help of this measurable function, we introduce the control
process $v_s^*:=v^*(s,Y_s^{3,u},Z_s^{3,u},\mu_s^u), s\in[t,t_j]$. We notice
that $v^*=(v^*_s)_{s\in[t,t_j]}$ belongs to ${\cal V}_{t,t_j}^\pi$ and
is even
$\widetilde{\mathbb{F}}^\pi$-adapted. Thus,
\begin{eqnarray}
&&E \bigl[F \bigl(s,x,Y^{3,u}_s
,Z^{3,u}_s,u_s, v_s^* \bigr) |
\widetilde{\cal F}^\pi_s \bigr]\nonumber\\
&&\qquad = F
\bigl(s,x,Y^{3,u}_s,Z^{3,u}_s,
\mu^u_s, v_s^* \bigr)
\\
&&\qquad = F_1 \bigl(s,x,Y^{3,u}_s
,Z^{3,u}_s,\mu^u_s \bigr), \qquad ds
\,dP \mbox{-a.e.},\nonumber
\end{eqnarray}
from where we see that $(Y^{3,u},Z^{3,u},M^{3,u})$ is a
solution of BSDE
(\ref{BSDE2appr}) driven by the couple $(u,v^*)\in{\cal
U}_{t,t_j}^\pi
\times
{\cal V}_{t,t_j}^\pi$ of admissible controls. Consequently, the
uniqueness of the
solution of BSDE (\ref{BSDE2appr}) yields that $Y^{2,u,v^*}_s=Y^{3,u}_s,
s\in[t,t_j]$, and from~(\ref{BSDE3_inequ}) we obtain:
\begin{equation}
\label{BSDE3-1_inequ} Y^{3,u}_s=
\essinf_{v\in
{\cal
V}_{t,t_j}^\pi}Y^{2,u,v}_s, \qquad P\mbox{-a.s, } s
\in[t,t_j], u\in{\cal U}_{t,t_j}^\pi.
\end{equation}
\textit{Step} 2. We begin with showing the latter relation
in (\ref{qq}). For this end we remark that, for all $(s,y,z)$, the
function $(\mu,\nu)\rightarrow F(s,x,y,z,\mu,\nu)=\int_{U}\int_{ V}
F(s,x,y,z,u,v)\nu(dv)\mu(du)$, $(\mu,\nu)\in{\cal P}(U)\times
{\cal P}(V)$,
is bi-linear and,\break hence, concave-convex in $(\mu,\nu)$ belonging
to the cross product ${\cal P}(U)\times{\cal P}(V)$ of two convex compact
spaces. Consequently, this mapping admits a saddle point, and it
follows in
particular that the order of $\sup_{\mu\in{\cal P}(U)}$ and
$\inf_{\nu\in{\cal P}(V)}$ is exchangeable without changing the value
of $F_0(s,x,y,z)$.
Let us now consider an arbitrary $u\in{\cal U}_{t,t_j}^\pi$. From the
definition of the function $F_0(s,x,y,z)$ and that of
$F_1(s,x,y,z,\mu)$, we have
\begin{eqnarray}
&&F_0(s,x,y,z)\nonumber\\
&&\qquad = \sup_{\mu\in{\cal P}(U)}F_1(s,x,y,z,
\mu)
\\
&&\qquad \ge F_1 \bigl(s,x,y,z,\mu^u_s \bigr),\qquad
(s,y,z) \in[t,t_j]\times R\times R^d, u\in{\cal
U}_{t,t_j}^\pi.\nonumber
\end{eqnarray}
Consequently, since $(Y^0,0)$ can be regarded as the solution
of the BSDE
\[
dY^0_s=-F_0 \bigl(s,x,Y^0_s,0
\bigr)\,ds+0\cdot \,dB_s,\qquad s\in[t,t_j],
Y^0_{t_j}=0,
\]
we get from the comparison theorem for BSDEs that $Y^0_s
\ge Y^{3,u}_s, s\in[t,t_j], P$-a.s. Hence, in view of the
arbitrariness of the choice of $u\in{\cal U}_{t,t_j}^\pi$, it follows that
\begin{equation}
\label{Y_0} Y^{0}_s\ge\esssup_{u \in{\cal U}_{t,
t_j}^\pi}Y^{3,u}_s,
\qquad P\mbox{-a.s.}, s\in[t,t_j].
\end{equation}
It remains to prove that we have even equality in this latter
relation. For this end, we notice that thanks to the uniform continuity of
the function $(s,y,\mu)\rightarrow F_1(s,x,y,0,\mu)$ over
$[t,t_j]\times R
\times{\cal P}(U)$ [we note that $x$ in $F_1(s,x,y$, $0,\mu)$ is fixed],
and the compactness of ${\cal P}(U)$ endowed with the topology
generated by
the weak convergence, we have the existence of a Borel measurable selection
$\mu^*=(\mu^*(s,y))\dvtx [t,t_j]\times R\rightarrow{\cal P}(U)$
such that
\[
F_0(s,x,y,0)=F_1 \bigl(s,x,y,0,\mu^*(s,y) \bigr), \qquad (s,y)
\in[t,t_j]\times R.
\]
Again from the uniform continuity of $(s,y,\mu)\rightarrow
F_1(s,x,y,0,\mu)$, we get that, for
arbitrarily given $\varepsilon>0$ there is some
$\delta(=\delta_\varepsilon)>0$ such that $|F_1(s,x,\break y,0,\mu)-F_1(s',x,y',
0,\mu)|\le\varepsilon$, for all $\mu\in{\cal P}(U)$ and all $(s,y),
(s',y')$ with $|(s,y)-(s',y')|\le\delta$. Let $(\Delta_\ell)_{\ell
\ge1}$ be a Borel partition of the set $[t,t_j]\times R$, composed of
nonempty sets $\Delta_\ell$ with diameter less than or equal to
$\delta$.
For every $\ell\ge1$, let us fix arbitrarily an element $(s_\ell
,y_\ell)$
of $\Delta_\ell$, and let us put $\mu_\ell:=\mu^*(s_\ell,y_\ell)$.
Moreover, let us consider an independent sequence of random variables
$\xi_\ell\in L^0(\Omega,\sigma\{\zeta_{j,1}\},P;U)$ such that,
for all $\ell\ge1$, the law $P\circ[\xi_\ell]^{-1}$ coincides with~$\mu_\ell$.
With the above introduced quantities, we define the control process
\[
u_s^*:=\sum_{\ell\ge1}I \bigl\{
\bigl(s,Y^0_s \bigr)\in \Delta_\ell \bigr\}
\cdot \xi_\ell, \qquad s\in[t,t_j].
\]
Such defined process belongs, obviously, to ${\cal U}_{t,
t_j}^\pi$. Moreover, we observe that, for all $s\in[t,t_j]$, $u^*_s$ is
$\sigma\{\zeta_{j,1}\}$-measurable and, consequently, independent of
$\widetilde{\cal F}_s^\pi$. Hence, for all $A\in{\cal B}(U)$,
\begin{eqnarray}
\mu^{u^*}_s(A) & = & P \bigl
\{u^*_s\in A | \widetilde{\cal F}_s^\pi
\bigr\}= P \bigl\{u^*_s\in A \bigr\}
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
& = & \sum_{\ell\ge1}I \bigl\{
\bigl(s,Y^0_s \bigr)\in\Delta_\ell \bigr\}
\mu_\ell(A).
\end{eqnarray}
It follows that $\mu^{u^*}_s= \sum_{\ell\ge1}I\{(s,Y^0_s)
\in\Delta_\ell\}\mu_\ell$. Hence, due to our choice of the partition
$\Delta_\ell, \ell\ge1$,
\begin{eqnarray}
\label{F_0_F_1}
F_0 \bigl(s,x,Y_s^0,0 \bigr) & \le&
\varepsilon+\sum_{\ell\ge1} I \bigl\{
\bigl(s,Y^0_s \bigr)\in\Delta_\ell \bigr
\}F_0(s_\ell,x,y_\ell,0)
\nonumber\\
& = & \varepsilon+\sum_{\ell\ge1} I \bigl\{
\bigl(s,Y^0_s \bigr)\in\Delta_\ell \bigr
\}F_1(s_\ell,x,y_\ell,0,\mu_\ell)
\\
& = & \varepsilon+\sum_{\ell\ge1} I \bigl\{
\bigl(s,Y^0_s \bigr)\in\Delta_\ell \bigr
\}F_1 \bigl(s_\ell,x,y_\ell,0,
\mu_s^{u^*} \bigr)
\nonumber\\
& \le& 2\varepsilon+F_1 \bigl(s,x,Y^0_s,0,
\mu_s^{u^*} \bigr), \qquad s\in[t,t_j].\nonumber
\end{eqnarray}
Let us compare now $Y^0$ with the solution $(Y^{3,u^*},Z^{3,u^*})$ of BSDE
(\ref{BSDE3appr}) controlled by $u^*\in{\cal U}_{t,t_j}^\pi$. Obviously,
\begin{eqnarray*}
d \bigl(Y^0_s-Y^{3,u^*}_s \bigr)&=&-
\bigl(F_0 \bigl(s,x,Y^0_s,0
\bigr)-F_1 \bigl(s,x,Y^{3,u^*}_s,
Z^{3,u^*}_s,\mu_s^{u^*} \bigr) \bigr)
\,ds\\
&&{}-Z^{3,u^*}_s\,dB_s-dM^{3,u^*}_s,
\end{eqnarray*}
$s\in[t,t_j], Y^0_{t_j}-Y^{3,u^*}_{t_j}=0$,
and from the It\^{o} formula,
\begin{eqnarray}
& &d \bigl( \bigl(Y^0_s-Y^{3,u^*}_s
\bigr)^+ \bigr)^2
\nonumber\\
&&\qquad = -2 \bigl(Y^0_s-Y^{3,u^*}_s
\bigr)^+ \bigl(F_0 \bigl(s,x,Y^0_s,0
\bigr)-F_1 \bigl(s,x,Y^{3,u^*}_s,Z^{3,u^*}_s,
\mu_s^{u^*} \bigr) \bigr)\,ds
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
& &\qquad\quad{} +\bigl|Z^{3,u^*}_s\bigr|^2I \bigl
\{Y^0_s-Y^{3,u^*}_s>0 \bigr\}\,ds -
2 \bigl(Y^0_s-Y^{3,u^*}_s
\bigr)^+Z^{3,u^*}_s\,dB_s
\\
& & \qquad\quad{}+ I \bigl\{Y^0_s-Y^{3,u^*}_s>0
\bigr\}\,d \bigl[M^{3,u^*} \bigr]_s -2 \bigl(Y^0_s-Y^{3,u^*}_s
\bigr)^+dM^{3,u^*}_s,\nonumber
\end{eqnarray}
and from standard estimates combined with (\ref{F_0_F_1})
we get
\begin{eqnarray}
& & \bigl( \bigl(Y^0_s-Y^{3,u^*}_s
\bigr)^+ \bigr)^2+E \biggl[\int_s^{t_j}
\bigl|Z^{3,u^*}_r\bigr|^2I \bigl\{Y^0_r-Y^{3,u^*}_r>0
\bigr\}\,dr \nonumber\\
&&\hspace*{101pt}{}+\int_{(s,t_j]} I \bigl\{Y^0_r-Y^{3,u^*}_r>0
\bigr\}\,d \bigl[M^{3,u^*} \bigr]_r \Big| \widetilde{\cal
F}_s^\pi \biggr]
\nonumber\\
&&\qquad = 2 E \biggl[\int_s^{t_j}
\bigl(Y^0_r-Y^{3,u^*}_r \bigr)^+
\bigl(F_0 \bigl(r,x,Y^0_r,0 \bigr)
\nonumber\\
&&\hspace*{69pt}{}-F_1 \bigl(r,x,Y^{3,u^*}_r,Z^{3,u^*}_r,
\mu_r^{u^*} \bigr) \bigr)\,dr\Big | \widetilde{\cal
F}_s^\pi \biggr]
\nonumber\\
&&\qquad \le 2 E \biggl[\int_s^{t_j}
\bigl(Y^0_r-Y^{3,u^*}_r \bigr)^+
\bigl(2\varepsilon+F_1 \bigl(r,x,Y^0_r,0,
\mu_r^{u^*} \bigr)\\
&&\hspace*{82pt}{} -F_1 \bigl(r,x,Y^{3,u^*}_r,
Z^{3,u^*}_r,\mu_r^{u^*} \bigr)
\bigr)\,dr \Big| \widetilde{\cal F}_s^\pi \biggr]
\nonumber\\
&&\qquad \le 2 E \biggl[\int_s^{t_j}
\bigl(Y^0_r-Y^{3,u^*}_r \bigr)^ +
\bigl(2\varepsilon+C\bigl|Y^0_r-Y^{3,u^*}_r\bigr|+C\bigl|Z^{3,u^*}_r\bigr|
\bigr)\,dr\Big | \widetilde{\cal F}_s^\pi \biggr]
\nonumber\\
&& \qquad \le\varepsilon^2+CE \biggl[\int_s^{t_j}
\bigl( \bigl(Y^0_r- Y^{3,u^*}_r
\bigr)^+ \bigr)^2\,dr\Big | \widetilde{\cal F}_s^\pi
\biggr] \nonumber\\
&&\qquad\quad{}+\frac12 E \biggl[\int_s^{t_j}\bigl|Z^{3,u^*}_r\bigr|^2I
\bigl\{ Y^0_r-Y^{3,u^*}_r>0 \bigr
\}\,dr \Big| \widetilde{\cal F}_s^\pi \biggr].\nonumber
\end{eqnarray}
Hence, from Gronwall's lemma, we see that,
for some constant $C$ independent of $\varepsilon$,
$(Y^0_s-Y^{3,u^*}_s)^+\le C\varepsilon$, $s\in[t,t_j]$, that is,
\[
Y^0_s\le Y^{3,u^*}_s+C\varepsilon,
s\in[t,t_j], \qquad P\mbox{-a.s.}
\]
This latter relation together with (\ref{Y_0}) yields
\[
Y^{0}_s=\esssup_{u \in{\cal U}_{t,
t_j}^\pi}Y^{3,u}_s,\qquad
P\mbox{-a.s.}, s\in[t,t_j].
\]
Recalling the result of step 1 we can conclude the first relation
of the lemma. The second one follows by a symmetric argument.
\end{pf}
After the above auxiliary lemmas, we are now able to
characterize the functions~$W$ and $U$ introduced by Lemma
\ref{limit_W_pi_n} as viscosity solution of PDE (\ref{HJBI}).
\begin{lemma}\label{viscosity solution}
The functions $W, U\dvtx [0,T]\times R^d\rightarrow R$ coincide and solve
PDE~(\ref{HJBI}) in viscosity sense.
\end{lemma}
\begin{pf} \textit{Step} 1. Let us show in this step that
the function $W$ introduced in Lemma \ref{limit_W_pi_n} as the
uniform limit on compacts of a suitable sequence of lower value
functions $W^{\pi_n}, n\ge1$, is a viscosity supersolution of
(\ref{HJBI}).
For this, we fix arbitrarily $(t,x)\in[0,T)\times R^d$ and we let
$\varphi\in C_{\ell,b}^{3}([0,T]\times R^d)$ be such that $W-\varphi
\ge W(t,x)-\varphi(t,x)=0$ on $[0,T)\times R^d$. Let $\rho>0$ be
arbitrarily small and $K>0$ sufficiently large. Since $W^{\pi_n},
n\ge1$, converges uniformly on compacts to $W$, there is some
$n_{\rho,K}\ge1$ such that, for all $n\ge n_{\rho,K}$, $|W(s,x')
-W^{\pi_n}(s,x')|\le\rho$, for every $(s,x')\in[0,T]\times R^d$ with
$|x'-x|\le K$. Then it follows from the DPP (Theorem \ref{DPP}) that,
for all $n\ge n_{\rho,K}$ and every $t_j^n\in\pi_n$ with $t<t_j^n\le T$,\vspace*{1pt}
\begin{eqnarray}
\varphi(t,x)+\rho &=& W(t,x)+\rho\nonumber\\[1pt]
&\ge &W^{\pi_n}(t,x)
\\[1pt]
&=&\esssup_{\alpha\in\mathcal{A}^{\pi_n}_{t,t_j^n}} \essinf _{\beta\in
\mathcal{B}^{\pi_n}_{t,t_j^n}}G^{t,x;\alpha,\beta}_{t,t_j^n}
\bigl(W^{\pi_n} \bigl(t_j^n,X_{t_j^n}^{t,x;\alpha,\beta}
\bigr) \bigr).\nonumber\vspace*{1pt}
\end{eqnarray}
On the other hand, taking into account that the functions
$W^{\pi_n}, n\ge1$, are bounded, uniformly with respect to $n\ge1$ and $W$ is bounded,
we have, for some constant $C_0$ (independent of $n$),\vspace*{1pt}
\begin{eqnarray}
&& W^{\pi_n} \bigl(t_j^n,X_{t_j^n}^{t,x;\alpha,\beta}
\bigr) \nonumber\\[1pt]
&&\qquad \ge W \bigl(t_j^n,X_{t_j^n}^{t,
x;\alpha,\beta}
\bigr)-\rho-2C_0I \bigl\{\bigl|X_{t_j^n}^{t,x;\alpha,\beta}-x\bigr|>K \bigr
\}
\\[1pt]
&&\qquad \ge \varphi \bigl(t_j^n,X_{t_j^n}^{t,x;\alpha,
\beta}
\bigr)-\rho-2C_0I \bigl\{\bigl|X_{t_j^n}^{t,x;\alpha,\beta}-x\bigr|>K \bigr
\},\nonumber\vspace*{1pt}
\end{eqnarray}
for all $\alpha\in\mathcal{A}^{\pi_n}_{t,t_j^n}, \beta
\in
\mathcal{B}^{\pi_n}_{t,t_j^n}$, and from the comparison theorem as well
as BSDE standard estimates (refer to \cite{CFS}) applied
to the BSDE defining our backward stochastic semigroup we obtain\vspace*{1pt}
\begin{eqnarray}
\varphi(t,x)+\rho&\ge&\esssup_{\alpha\in\mathcal{A}^{\pi_n}_{t,t_j^n}}
\essinf_{\beta\in\mathcal{B}^{\pi_n}_{t,t_j^n}}G^{t,x;\alpha
,\beta}_{t,
t_j^n}
\bigl(W^{\pi_n} \bigl(t_j^n,X_{t_j^n}^{t,x;\alpha,\beta}
\bigr) \bigr)
\nonumber\\[1pt]
& \ge&\esssup_{\alpha\in\mathcal{A}^{\pi_n}_{t,t_j^n}} \essinf _{\beta
\in
\mathcal{B}^{\pi_n}_{t,t_j^n}}\nonumber\\[1pt]
&&{}\times G^{t,x;\alpha,\beta
}_{t,t_j^n}
\bigl(\varphi \bigl(t_j^n, X_{t_j^n}^{t,x;\alpha,\beta}
\bigr)-\rho -2C_0I \bigl\{\bigl|X_{t_j^n}^{t,x;\alpha,
\beta}-x\bigr|>K
\bigr\} \bigr)
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
& \ge&\esssup_{\alpha\in\mathcal{A}^{\pi_n}_{t,t_j^n}} \essinf _{\beta
\in
\mathcal{B}^{\pi_n}_{t,t_j^n}} G^{t,x;\alpha,\beta
}_{t,t_j^n}
\bigl(\varphi \bigl(t_j^n, X_{t_j^n}^{t,x;\alpha,\beta}
\bigr) \bigr) \nonumber\\[1pt]
&&{}-\esssup_{\alpha\in\mathcal{A}^{\pi_n}_{t,t_j^n},\beta\in
\mathcal{B}^{\pi_n}_{t,t_j^n}}\nonumber\\[1pt]
&&{}\times L \bigl(E \bigl[ \bigl(
\rho+2C_0I \bigl\{ \bigl|X_{t_j^n}^{t,x;\alpha,
\beta}-x\bigr|>K \bigr\}
\bigr)^2|\widetilde{\cal F}_t^{\pi_n} \bigr]
\bigr)^{1/2},\nonumber
\end{eqnarray}
where the constant $L$ depends only on the coefficient $f$.
However, since
\begin{eqnarray}
& & E \bigl[ \bigl(\rho+2C_0I \bigl\{\bigl|X_{t_j^n}^{t,x;\alpha,
\beta}-x\bigr|>K
\bigr\} \bigr)^2|\widetilde{ \cal F}_t^{\pi_n}
\bigr]
\nonumber\\
&&\qquad\le 2\rho^2+8C_0^2\frac{1}{K^2}E
\bigl[\bigl|X_{t_j^n}^{t,x;\alpha,
\beta}-x\bigr|^2|\widetilde{\cal
F}_t^{\pi_n} \bigr]
\\
&&\qquad\le 2\rho^2+\frac{C}{K^2}\qquad (\alpha,\beta)\in
\mathcal{A}^{\pi_n}_{t,
t_j^n}\times\mathcal{B}^{\pi_n}_{t,t_j^n},
n\ge1\nonumber
\end{eqnarray}
(Recall that the coefficients $\sigma$ and $b$ of the
dynamics of
the game are bounded), we get for $K:=1/\rho$, for all $n\ge n_\rho:=
n_{\rho,K}$,
\begin{equation}
\qquad
\varphi(t,x)+C\rho\ge\esssup_{\alpha\in\mathcal{A}^{\pi_n}_{t,t_j^n}} \essinf_{\beta\in\mathcal{B}^{\pi_n}_{t,t_j^n}}G^{t,x;\alpha
,\beta}_{t,
t_j^n}
\bigl(\varphi \bigl(t_j^n,X_{t_j^n}^{t,x;\alpha,\beta}
\bigr) \bigr),
\end{equation}
where $C\in R$ is a constant independent of $\rho$, $n$ and $t_j^n$.
From the latter estimate, we deduce with the help of Lemmas
\ref{lemma_BSDE1} and \ref{lemma_BSDE2} that
\begin{eqnarray}
C\rho&\ge&\esssup_{\alpha\in\mathcal{A}^{\pi_n}_{t,t_j^n}} \essinf_{\beta\in\mathcal{B}^{\pi_n}_{t,t_j^n}}
\bigl(G^{t,x;\alpha
,\beta}_{t,
t_j^n} \bigl(\varphi \bigl(t_j^n,X_{t_j^n}^{t,x;\alpha,\beta}
\bigr) \bigr)-\varphi (t,x) \bigr)
\nonumber\\
&= & \esssup_{\alpha\in\mathcal{A}^{\pi_n}_{t,t_j^n}} \essinf_{\beta\in\mathcal{B}^{\pi_n}_{t,t_j^n}}Y_t^{1,\alpha
,\beta}
\\
&\ge& \esssup_{\alpha\in\mathcal{A}^{\pi_n}_{t,t_j^n}} \essinf_{\beta\in\mathcal{B}^{\pi_n}_{t,t_j^n}}Y_t^{2,\alpha
,\beta}
-C \bigl(t_j^n-t \bigr)^{3/2},\qquad P\mbox{-a.s.}\nonumber
\end{eqnarray}
Of course, as before, the quantities $Y_t^{1,\alpha,\beta},
Y_t^{2,\alpha,\beta}$ have to be understood as $Y_t^{1,u,v},Y_t^{2,u,v}$
for $(u,v)\in{\cal U}^{\pi_n}_{t,t_j^n}\times{\cal V}^{\pi_n}_{t,t_j^n}$
associated with $(\alpha,\beta)\in{\cal A}^{\pi_n}_{t,t_j^n}\times
{\cal
B}^{\pi_n}_{t,t_j^n}$ through Lemma \ref{controls-NAD-strategies}.
Moreover, they are defined by Lemmas \ref{lemma_BSDE1} and
\ref{lemma_BSDE2} for $t_j=t_j^n$, that is, they depend on the choice of
$t_j^n\in\pi_n$ and so, in particular, $n\ge n_\rho$.
Obviously, since ${\cal U}^{\pi_n}_{t,t_j^n}$ can be regarded as a subset
of ${\cal A}^{\pi_n}_{t,t_j^n}$ by identifying $u\in
{\cal U}^{\pi_n}_{t,t_j^n}$ with the NAD strategy $\alpha^u(v):=u,
v\in
{\cal V}^{\pi_n}_{t,t_j^n}$,
\begin{eqnarray}
\label{e1}
C\rho+C \bigl(t_j^n-t
\bigr)^{3/2} &\ge& \esssup_{\alpha\in\mathcal{A}^{\pi
_n}_{t,t_j^n}} \essinf_{\beta\in\mathcal{B}^{\pi_n}_{t,t_j^n}}Y_t^{2,\alpha
,\beta}
\nonumber\\
&\ge& \esssup_{u\in\mathcal{U}^{\pi_n}_{t,t_j^n}} \essinf_{\beta\in\mathcal{B}^{\pi_n}_{t,t_j^n}}Y_t^{2,u,\beta
(u)}
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&\ge& \esssup_{u\in\mathcal{U}^{\pi_n}_{t,t_j^n}} \essinf_{v\in\mathcal{V}^{\pi_n}_{t,t_j^n}}Y_t^{2,u,v}
\\
& = & Y^0_t,\qquad P\mbox{-a.s.}, n\ge n_\rho,\nonumber
\end{eqnarray}
where the latter equality was stated in Lemma \ref{lemma_BSDE3}.
Remark that here, of course, $Y^0$ is defined by Lemma \ref{lemma_BSDE3}
for $t_j^n$. Since
\[
Y^0_s=\int_s^{t_j^n}F_0
\bigl(r,x,Y_r^0,0 \bigr)\,dr, \qquad s\in
\bigl[t,t_j^n \bigr]
\]
and $F_0(r,x,y,0)$ is bounded, continuous, and Lipschitz in
$y$, uniformly with respect to $r$, it follows that
$|Y^0_s|\le C(t_j^n-t), s\in[t,t_j^n]$, and
\begin{eqnarray}
\label{e2}
\frac{1}{t_j^n-t}{Y^0_t}&= &
\frac{1}{t_j^n-t} \int_t^{t_j^n}F_0
\bigl(r,x,Y_r^0,0 \bigr)\,dr
\nonumber\\
&\ge&\frac{1}{t_j^n-t}\int_t^{t_j^n}
\bigl(F_0(r,x,0,0)-L\bigl|Y_r^0\bigr| \bigr)\,dr
\\
&\ge&\frac{1}{t_j^n-t}\int_t^{t_j^n}F_0(r,x,0,0)
\,dr-C \bigl(t_j^n-t \bigr).\nonumber
\end{eqnarray}
Let $\rho\le(T-t)^{3/2}$. Since the mesh $|\pi_n|$ of the
partition $\pi_n$ converges to zero as $n\rightarrow+\infty$, we can find
for $n\ge n_\rho$ large enough some $t_j^n\in\pi_n, t_j^n>t$, such that
$(t_j^n-t)^{3/2}/2\le\rho\le(t_j^n-t)^{3/2}$. Consequently, for
$n\ge n_\rho$ large enough we can conclude from (\ref{e1}) and (\ref{e2})
that
\[
C \bigl(t_j^n-t \bigr)^{1/2}\ge
\frac{1}{t_j^n-t}{Y^0_t} \ge\frac{1}{t_j^n-t}\int
_t^{t_j^n}F_0(r,x,0,0)\,dr-C
\bigl(t_j^n-t \bigr).
\]
Thus, taking the limit as $\rho\rightarrow0$ (and, hence,
$n\rightarrow+\infty$ and $t_j^n-t\rightarrow0$), we obtain $F_0(t,x,0,0)
\le0$. But recalling the definition of $F_0$ from Lemma \ref{lemma_BSDE3},
we see that
\begin{eqnarray}
0&\ge& F_0(t,x,0,0)=\sup_{\mu\in{\cal P}(U)}\inf
_{\nu\in{\cal P}(V)} F(t,x,y,z,\mu,\nu)
\nonumber\\
& =& \sup_{\mu\in{\cal P}(U)}\inf_{\nu\in{\cal P}(V)}\nonumber\\
&&{}\times\int
_{U\times V} \biggl( \frac{\partial}{\partial t}\varphi(t,x) +
\frac{1}{2}\operatorname{tr} \bigl(\sigma \sigma^{T} (t,x, u, v)D^2
\varphi \bigr)
\nonumber
\\[-8pt]
\\[-8pt]
\nonumber
&&\hspace*{44pt}{}+ D\varphi.b(t, x, u, v)\\
&&\hspace*{44pt}{}+ f \bigl(t, x,\varphi(t,x),D\varphi
(t,x)\cdot\sigma (t,x,u, v),u, v \bigr) \biggr)\mu\otimes\nu(du\,dv)
\nonumber\\
& = &\frac{\partial}{\partial t}\varphi(t,x)+H \bigl(t,x, \bigl(\varphi, D
\varphi,D^2\varphi \bigr) (t,x) \bigr).\nonumber
\end{eqnarray}
Therefore, $W$ is a viscosity supersolution of PDE (\ref{HJBI}).
\textit{Step} 2. With an argument symmetric to that developed
in step 1 we show that $U$ is a viscosity subsolution of PDE (\ref{HJBI}).
Since both $W$ and $U$ are bounded continuous solutions, $W$ is a viscosity
supersolution and $U$ is a viscosity subsolution of (\ref{HJBI}), it follows
from the comparison principle (Proposition \ref{comparison principle})
that $W\ge U$ on $[0,T]\times R^d$. On the other hand, $(W,U)$ is the pointwise
limit over the sequence $(W^{\pi_n},U^{\pi_n}), n\ge1$, where the lower
value function $W^{\pi_n}$ along the partition $\pi_n$
is less than or equal to the upper one $U^{\pi_n}$, for all $n\ge1$.
Consequently, $W$ and $U$ coincide, and both are viscosity solutions of
PDE (\ref{HJBI}). Again from the comparison principle it follows that
this viscosity solution $W=U=V$ is the unique one inside the class of
continuous unions with at most polynomial growth.
\end{pf}
The above lemma allows now to prove Theorem \ref{main result}.
\begin{pf} From our above discussion, we have seen that for any arbitrary
sequence of partitions $\pi_n, n\ge1$, with $|\pi_n|\rightarrow0$, as
$n\rightarrow+\infty$, there is a subsequence which, abusing
notation, we
have also denoted by $\pi_n, n\ge1$, such that $W^{\pi_n}$ as well as
$U^{\pi_n}$ converge uniformly on compacts to the unique viscosity
solution $V$ of PDE (\ref{HJBI}) (uniqueness in the class of continuous
functions with polynomial growth); see Lemma \ref{viscosity solution}.
Consequently, the limit $V$ does not depend on the special choice of the
sequence of partitions $\pi_n, n\ge1$. Consequently, $W^{\pi_n}$ as
well as $U^{\pi_n}$ converge uniformly on compacts to the unique viscosity
solution $V$, for all sequence of partitions $\pi_n, n\ge1$ with mesh
$|\pi_n|\rightarrow0$, as $n\rightarrow+\infty$. The proof is complete.
\end{pf}
\section*{Acknowledgements}
The authors would like to thank the anonymous Associate Editor and the
anonymous referee for their valuable comments and suggestions from
which the manuscript greatly has benefited.
|
\section{Introduction}\label{sec:intro}
The geodesic deviation equation (GDE), or Jacobi equation, introduced into the canon of basic concepts in general relativity by Synge and Schild \cite{Schild,Synge}, describes the tidal acceleration between neighboring freely-falling observers, and plays an important role in interpreting the curvature of spacetime.
The GDE governs the evolution of a deviation vector field, or Jacobi field, $\xi^\alpha(\tau)$ along a fiducial geodesic worldline $y(\tau)$, which [somehow $\ldots$] specifies a neighboring geodesic worldline $z(\tau)$.
In elementary textbook treatments \cite{Weinberg,MTW}, one defines the coordinate-basis components of the deviation vector $\xi^\alpha$, at linear order, to be the coordinate differences between points on the worldline $z$ and the fiducial \mbox{geodesic $y$,}
\begin{equation}\label{xixlinear}
\xi^\alpha(\tau)=z^\alpha(\tau)-y^\alpha(\tau)+O(\xi^2),
\end{equation}
where $\tau$ is an affine parameter along both $y$ and $z$. It follows, via manipulations of connection coefficients, that the neighboring worldline $z$ will be a geodesic, to linear order, if the vector $\xi^\alpha$ satisfies the well-known linear GDE:
\begin{equation}\label{gde}
\ddot\xi^\alpha+{R^{\alpha}}_{\beta\gamma\delta}\,u^\beta u^\delta \xi^\gamma=O(\xi,\dot\xi)^2,
\end{equation}
where $u^\alpha=dy^\alpha/d\tau$ is the tangent to the geodesic $y(\tau)$, the Riemann tensor is evaluated along $y(\tau)$, dots denote covariant parameter derivatives $D/d\tau$, and $O(\xi,\dot\xi)^n$ here denotes terms with $n$ or more factors of $\xi^\alpha$ and/or $\dot\xi^\alpha$.
More sophisticated textbook treatments \cite{Synge,Wald} define the deviation vector in the following covariant manner. Consider a one-parameter family of geodesics $z(\tau,\epsilon)$, where $\epsilon$ labels the geodesics, and $\tau$ is an affine parameter along each geodesic. Taking $y(\tau)=z(\tau,0)$ as the fiducial geodesic, one can define deviation vectors $\bar\xi^\alpha$ along $y$ to be the tangents to the curves $\tau=\mathrm{constant}$:
\begin{equation}\label{de}
\bar\xi^\alpha(\tau)=\left.\frac{\partial z^\alpha(\tau,\epsilon)}{\partial\epsilon}\right|_{\epsilon=0}.
\end{equation}
It follows, after applying $D^2/d\tau^2$ and using the Ricci identity, that $\ddot{\bar\xi}^\alpha+{R^\alpha}_{\beta\gamma\delta}u^\beta u^\delta \bar\xi^\gamma=0$, like Eq.~(\ref{gde}) but exact. To connect with the previous paragraph, we can identify, for a given $\epsilon\gtrsim0$, $z(\tau)=z(\tau,\epsilon)$ and $\xi^\alpha=\epsilon\bar\xi^\alpha+O(\epsilon^2)$.
The generalization of the GDE to higher orders \mbox{in $\xi^\alpha$} seems to have been first considered by Hodgkinson \cite{Hodgkinson}, who, in a nutshell, extended the coordinate-based construction of Eq.~(\ref{xixlinear}) to second and third orders, using manipulations of connection coefficients to derive $O(\xi,\dot\xi)^2$ and $O(\xi,\dot\xi)^3$ corrections to the GDE (\ref{gde}). A more covariant derivation of the same second-order corrections was subsequently given by Bazanski \cite{Bazanski,Bazanski2}, who, in a nutshell, extended the relation (\ref{de}) by adding to it a term proportional to the second covariant $\epsilon$-derivative of $z(\tau,\epsilon)$.
The basic geometric construction underlying both authors' work, defining the $y$-$\xi^\alpha$-$z$ relationship to all orders, is to reach the point $z$ on the neighboring worldline by following the affinely parametrized geodesic issuing from the point $y$ on the fiducial geodesic with initial tangent vector $\xi^\alpha$ for an affine parameter interval of 1; i.e.,~$z$ is the result of the exponential map of $\xi^\alpha$ at $y$ \cite{Aleksandrov}. This relationship can be simply and usefully stated in the language of covariant bitensors---developed by Synge \cite{Synge} and DeWitt and Brehme \cite{DeWittBrehme}, succinctly reviewed e.g.~by Poisson et al.~\cite{MPP}, and briefly described below---as
\begin{equation}\label{link}
\xi^\alpha(\tau)=-\sigma^\alpha\big(y(\tau),z(\tau)\big),
\end{equation}
where $\sigma^\alpha=\nabla^\alpha \sigma$ is the covariant derivative at $y$ of Synge's world function $\sigma(y,z)$. The link (\ref{link}) between the bitensor formalism and the higher-order geodesic deviations of Hodgkinson and Bazanski seems to have been first noted and employed by Aleksandrov and Piragas \cite{Aleksandrov} in their later (considerably streamlined) rederivation of the second- and third-order GDEs.
\phantom{yo}
\begin{figure}[h]
\includegraphics[scale=.5]{xi_intro.pdf}
\label{fig:sigmamu}
\caption{The point $z$ on the neighboring worldline is the exponential map of the deviation vector $\xi^\alpha$ at the point $y$ on the fiducial geodesic.}
\end{figure}
\newpage
One focus of this paper is to expound upon the insights of Ref.~\cite{Aleksandrov}, that one can use bitensor methods to formulate a generalization of the GDE valid to all orders in the deviation, finding its explicit form as a covariant expansion in the deviation vector. By combining these ideas with the `semi-recursive approach' to bitensor expansions---developed through Refs.~\cite{Avramidi_thesis,Avramidi_book,Decanini,Barry_thesis,Ottewill} for applications to radiation reaction and quantum fields in curved spacetime and quantum gravity---we derive efficient recursion relations which generate the expansion of the GDE, using them to reproduce the results of Refs.~\cite{Hodgkinson,Bazanski,Bazanski2,Aleksandrov} for the second- and third-order GDEs and to produce the fourth-order GDE. We also rederive results concerning the generalized Jacobi equations of Refs.~\cite{Hodgkinson,Mashhoon75,Mashhoon77,LiNi,Ciufolini,Ciufolini2,Chicone02,Chicone06a,Chicone06b,Perlick} which linearize in the deviation $\xi^\alpha$ but work to all orders in the relative velocity $\dot\xi^\alpha$, and results concerning the description of accelerated neighboring worldlines in terms of covariant deviation vectors, as discussed e.g.~by Refs.~\cite{Synge,Manoff}.
The results for the higher-order GDE are most compactly summarized by the corresponding action functional $S[\xi]=\int d\tau\, \mathcal L(\xi,\dot\xi)$ for the deviation vector, previously derived through second order [$O(\xi,\dot\xi)^3$ in the \mbox{action}] in Refs.~\cite{Bazanski2,Aleksandrov}, and derived below through fourth order [$O(\xi,\dot\xi)^5$]; the Lagrangian reads
\begin{align} \label{S}
\mathcal L=&\;
\tfrac{1}{2}\dot\xi^2-\tfrac{1}{2}R_{\xi u\xi u}
\\\nonumber&-\tfrac{2}{3}R_{\xi\dot\xi\xi u}-\tfrac{1}{6}R_{\xi u\xi u;\xi}
\\\nonumber&
-\tfrac{1}{6}R_{\xi\dot\xi\xi\dot\xi}-\tfrac{1}{4}R_{\xi\dot\xi\xi u;\xi}-\tfrac{1}{24}R_{\xi u\xi u;\xi\xi}-\tfrac{1}{6}R_{\xi u\xi \alpha} {R^{\alpha}}_{\xi\xi u}
\\\nonumber&
-\tfrac{1}{12} R_{\xi\dot\xi\xi\dot\xi;\xi}-\tfrac{1}{15}R_{\xi\dot\xi\xi u;\xi\xi}-\tfrac{1}{120}R_{\xi u\xi u;\xi\xi\xi}
\\\nonumber&
\quad\quad\quad-\big(\tfrac{2}{15}R_{\xi\dot\xi\xi\alpha}+\tfrac{7}{60}R_{\xi u\xi\alpha;\xi}\big){R^\alpha}_{\xi\xi u}\,+\,O(\xi,\dot\xi)^6,
\end{align}
where, e.g., $R_{\xi\dot\xi\xi u;\xi\xi}=\xi^\alpha \dot\xi^\beta\xi^\gamma u^\delta \xi^\epsilon \xi^\zeta \nabla_\zeta \nabla_\epsilon R_{\alpha\beta\gamma\delta}$. The Lagrangian (\ref{S}) and the resultant GDE given by Eqs.~(\ref{gde_exact}, \ref{LJIexp4}) below (in a sense\footnote{\label{foot:isoc}The results (\ref{S}), (\ref{solution}), and (\ref{Larb}) apply to the case of `isochronous correspondence,' which requires $\tau$ to be an affine parameter along both the fiducial geodesic $y(\tau)$ and the neighboring worldline $z(\tau)$ defined by Eq.~(\ref{link}); this choice leads to the simplest forms for the GDE and its action.
\\
$\phantom{abc}$Another useful choice to fix the parametrization of $z(\tau)$ [in the timelike case] is the `normal correspondence,' in which $\xi$ is constrained to be orthogonal to $u$, so that the (parallel-transported tetrad) components of $\xi$, along with $\tau$, correspond Fermi normal coordinates based on the fiducial geodesic (see e.g.~Ref.~\cite{MPP} Sec.~9). The Lagrangian for the normal correspondence is simply \mbox{$\mathcal L_\mathrm{norm.}=\sqrt{1+2\mathcal L_\mathrm{isoc.}}-1$}, where $\mathcal L_\mathrm{isoc.}$ is the $\mathcal L$ of Eq.~(\ref{S}); the Lagrangians agree to $O(\xi,\dot\xi)^3$ but differ at higher orders.
\\
$\phantom{abc}$Results for the normal correspondence, as well as for completely generic correspondences/parametrizations are discussed in Appendix \ref{app:gen}, while the main body of the text restricts attention to the isochronous correspondence.}) encapsulate all of the geometric corrections to geodesic deviation considered in Refs.~\cite{Hodgkinson,Bazanski,Bazanski2,Aleksandrov,Mashhoon75,Mashhoon77,LiNi,Ciufolini,Ciufolini2,Chicone02,Chicone06a,Chicone06b,Perlick}.
A second focus of this paper is to review Dixon's construction of the general solution to the linear GDE in terms of fundamental bitensors \cite{Dixon,Dixon1,Dixon3}, and to generalize the construction to find the solution to the second-order GDE. We show that, given the deviation vector $\xi^\alpha$ and its derivative $\dot\xi^\alpha$ at an initial point $y$ on the fiducial geodesic, the solution for the deviation vector $\xi^{\alpha'}$ at a second point $y'$ on the fiducial geodesic, a finite affine parameter interval $\tau$ along from $y$, can be written as
\begin{align}\label{solution}
\xi^{\alpha'}&={K^{\alpha'}}_\beta\xi^\beta+\tau {H^{\alpha'}}_\beta \dot\xi^\beta
\\\nonumber
&\phantom{yo}+\tfrac{1}{2}\left({L^{\alpha'}}_{\beta\gamma}+\tau {H^{\alpha'}}_\alpha{R^\alpha}_{\beta\gamma\delta}u^\delta\right)\xi^\beta\xi^\gamma
\\\nonumber&\phantom{yo}+\tau{J^{\alpha'}}_{\beta\gamma}\xi^\beta\dot\xi^\gamma+\tfrac{1}{2} \tau^2 {I^{\alpha'}}_{\beta\gamma} \dot\xi^\beta\dot\xi^\gamma+O(\xi,\dot\xi)^3,
\end{align}
where the `Jacobi propagators' in Dixon's linear solution (the first line) are given in terms of the second derivatives of the world function by
\begin{align}
{K^{\alpha'}}_\beta(y,y')&={H^{\alpha'}}_\gamma(y,y')\;{\sigma^\gamma}_\beta(y,y'),
\nonumber\\
{H^{\alpha'}}_\beta(y,y')&=-\Big({\sigma^\beta}_{\alpha'}(y,y')\Big)^{-1},
\end{align}
and the bitensors ${L^{\alpha'}}_{\beta\gamma}$, ${J^{\alpha'}}_{\beta\gamma}$, and ${I^{\alpha'}}_{\beta\gamma}$ in the second-order solution (the last two lines) are given by
\begin{align}
{L^{\alpha'}}_{\beta\gamma}&={K^{\alpha'}}_{\beta;\gamma}+{K^{\alpha'}}_{\beta;\gamma'}{K^{\gamma'}}_{\gamma}\,,
\\\nonumber
{J^{\alpha'}}_{\beta\gamma}&={K^{\alpha'}}_{\beta;\gamma'}{H^{\gamma'}}_{\gamma}\,,
\\
{I^{\alpha'}}_{\beta\gamma}&={H^{\alpha'}}_{\beta;\gamma'}{H^{\gamma'}}_{\gamma}\,.\label{LJIintro}
\end{align}
We will see below that the bitensors ${K^{\alpha'}}_\beta$, ${H^{\alpha'}}_\beta$, \mbox{${L^{\alpha'}}_{\beta\gamma}$, ${J^{\alpha'}}_{\beta\gamma}$, and ${I^{\alpha'}}_{\beta\gamma}$} appearing in the solution to the second-order GDE are the same bitensors needed to express the GDE and its action to all orders in the deviation. \{The exact Lagrangian and GDE can be written as
\begin{multline}
\mathcal L=\tfrac{1}{2}{K^\mu}_\beta K_{\mu\gamma}u^\beta u^\gamma+{K^\mu}_\beta H_{\mu\gamma}u^\beta\dot\xi^\gamma+\tfrac{1}{2}{H^\mu}_{\beta} H_{\mu\gamma}\dot\xi^\beta \dot\xi^\gamma
\\
\Rightarrow\;0={H^\mu}_\alpha \ddot\xi^\alpha+{L^\mu}_{\beta\gamma}u^\beta u^\gamma
+2{J^\mu}_{\beta\gamma}u^\beta \dot\xi^\gamma+{I^\mu}_{\beta\gamma}\dot\xi^\beta \dot\xi^\gamma,\phantom{\Big|}
\label{Larb}
\end{multline}
where the bitensors are functions of $(y,z)$ [or of $(y,\xi)$], and $\mu$ is an index at $z$.\}
In this paper, we define these bitensors in terms of derivatives of the world function $\sigma(y,y')$, and with these definitions they would seem to inherit the limited domain over which $\sigma(y,y')$ and its derivatives are well-defined, requiring the points $y$ and $y'$ to be connected by a unique geodesic segment (to be in each other's normal convex neighborhood (NCN)). In a companion paper \cite{companion}, we discuss how these and other important bitensors can be alternately defined in terms of the horizontal and vertical covariant derivatives of the exponential map \cite{Dixon,Schattner}, thereby extending the domain of validity of those bitensors and of the solution (\ref{solution}) beyond the NCN, as well as simplifying key expressions and derivations.
\phantom{yo}
We begin in Section \ref{sec:bitensors} with a quick overview of some basic ingredients of the bitensor formalism, including Synge's world function and its derivatives, coincidence limits, the parallel propagator, and covariant expansions near coincidence.
Section \ref{sec:linear} uses those ingredients to derive, with little effort, the usual linear GDE and its action principle, along with related results which are also applicable to accelerated neighboring worldlines.
Section \ref{sec:Dixon} derives Dixon's solution to the linear GDE and discusses some properties of the Jacobi propagators.
Section \ref{sec:arb} derives forms of the GDE and its action (and of all the other results given at linear order in Sec.~\ref{sec:linear}) which are valid to all orders in the deviation. We discuss the expansion of these results first to $O(\xi,\dot\xi)^n$ [as in `Bazanski's equation' at $O(\xi,\dot\xi)^2$] and then to $O(\xi^n)$ but to all orders in $\dot\xi$ [as in the `generalized Jacobi equation' at $O(\xi)$], and we mention various applications of these improved descriptions of geodesic deviation \cite{Mashhoon75,Mashhoon77,LiNi,Ciufolini,Ciufolini2,Chicone02,Chicone06a,Chicone06b,Perlick,Tammelo,Tammelo2,Baskaran,Kerner,vanHolten,Colistete,Colistete2,Koekoek,Koekoek2}.
Appendix \ref{app:gen} extends the analysis of the main text, which considers only the isochronous correspondence, to treat generic correspondences/parametrizations and the normal correspondence.$^{\ref{foot:isoc}}$
Appendix \ref{app:ntlo} extends the derivation of the general solution to the GDE to second order in the deviation.
Appendix \ref{app:semi} summarizes results from the semi-recursive/transport-equation approach which efficiently generate high-order expansions of fundamental bitensors.
\section{Bitensors}\label{sec:bitensors}
Bitensors are generalizations of ordinary spacetime tensors which depend on not one but two spacetime points and have a tensor character at each point. The classic references on the topic are the textbook by Synge \cite{Synge} and the article by DeWitt and Brehme addressing electromagnetic radiation reaction in curved spacetime \cite{DeWittBrehme}. This section briefly reviews some essential bitensor concepts, borrowing heavily from the thorough review by Poisson et al.~\cite{MPP}.
To distinguish tensor indices referring to a first point $y$ from those referring to a second point $z$, we use indices $\alpha$, $\beta$, $\gamma$, $\delta$ at $y$ and indices $\mu$, $\nu$, $\rho$ at $z$. For example, ${K^\mu}_\alpha(y,z)$ is a vector at $z$ and a 1-form at $y$. (Later, we will add a third point $y'$ with indices $\alpha'$, $\beta'$, $\gamma'$, $\delta'$ and a fourth point $z'$ with indices $\mu'$, $\nu'$, $\rho'$.)
Covariant differentiation can be performed at each of the points $y$ and $z$, denoted by $\nabla_\alpha$ and $\nabla_\mu$ or by semicolons, as in ${K^\nu}_{\beta;\alpha}=\nabla_\alpha {K^\nu}_\beta$ and ${K^\nu}_{\beta;\mu\rho}=\nabla_\rho \nabla_\mu {K^\nu}_\beta$. For any bitensor field, covariant derivatives at $y$ commute with those at $z$. For the special case of derivatives of Synge's world function $\sigma(y,z)$, we drop the semicolons, as in ${\sigma^\mu}_{\alpha\beta}=\nabla_\beta\nabla_\alpha\nabla^\mu\sigma$.
\subsection{Synge's world function and its derivatives}\label{sec:Synge}
Consider two points $y$ and $z$ which are connected by a unique geodesic segment $\Gamma$. The function $\sigma(y,z)$ which gives half the squared geodesic interval along $\Gamma$,
\begin{equation}\nonumber
\sigma(y,z)=\left\{\begin{array}{cccc}-\tfrac{1}{2}(\textrm{proper time})^2&\quad&&\textrm{timelike}\\0&\quad&&\textrm{null}
\\\tfrac{1}{2}(\textrm{proper distance})^2&\quad&&\textrm{spacelike}\end{array}\right.
\end{equation}
is known as Sygne's world function. The world function is a biscalar which is symmetric in its arguments. It is in a sense the fundamental bitensor.
A basic property of $\sigma(y,z)$ is that its first covariant derivatives with respect to $y$ and $z$, $\sigma^\alpha=\nabla^\alpha\sigma$ and \mbox{$\sigma^\mu=\nabla^\mu\sigma$},
yield vectors which are tangent to $\Gamma$ at $y$ and $z$, and their norms are the interval $\sqrt{|2\sigma|}$. If we parametrize $\Gamma$ as $x=x(\lambda)$, where $\lambda$ is any affine parameter along $\Gamma$, with $x(\lambda_0)=y$ and $x(\lambda_0+\Delta\lambda)=z$, then
\begin{equation}\label{sigmau}
\sigma^\alpha=-\Delta\lambda\, t^\alpha,\quad \sigma^\mu=\Delta\lambda\,t^\mu,\quad
\sigma =\tfrac{1}{2} (\Delta\lambda)^2\, t^2,\quad
\end{equation}
where $t^\alpha= dx^\alpha/d\lambda(\lambda_0)$ and $t^\mu= dx^\mu/d\lambda(\lambda_0+\Delta\lambda)$ are the tangents to $\Gamma$ at $y$ and $z$ (and $t^2=t^\alpha t_\alpha=t^\mu t_\mu$, because the tangent is parallel-transported along $\Gamma$), as illustrated in Fig.~\ref{fig:sigmamu}.
\begin{figure}[h]
\includegraphics[scale=.48]{tmu.pdf}
\caption{The geodesic segment $\Gamma$ linking the points $y$ and $z$. The tangents, $t^\alpha$ at $y$ and $t^\mu$ at $z$, point in the same direction along $\Gamma$ (here from $y$ to $z$, for $\Delta\lambda>0$). The world-function derivatives, $\sigma^\alpha$ at $y$ and $\sigma^\mu$ at $z$, both point outward from $\Gamma$.
}\label{fig:sigmamu}
\end{figure}
Rearranging the relations (\ref{sigmau}) yields
\begin{equation}\label{sigma2}
\sigma^\alpha\sigma_\alpha=2\sigma=\sigma^\mu\sigma_\mu.
\end{equation}
These differential equations, along with the condition $\sigma\to0$ as $z\to y$, completely define the world function. Differentiating Eqs.~(\ref{sigma2}) yields
\begin{equation}\label{sigmageod}
\sigma^\beta{\sigma^\alpha}_\beta=\sigma^\alpha,\qquad
\sigma^\nu{\sigma^\mu}_\nu=\sigma^\mu,
\end{equation}
which are geodesic equations for the vector fields $\sigma^\alpha$ (with $z$ fixed) and $\sigma^\mu$ (with $y$ fixed). Other useful identities can be obtained from further differentiation of Eqs.~(\ref{sigma2}), as in Sec.~4.1 of Ref.~\cite{MPP}.
Note that the point $y$ and the vector $\sigma^\alpha$ at $y$ determine the point $z$: one reaches $z$ by traveling a parameter interval of 1 along the affinely parametrized geodesic issuing from $y$ with initial tangent $-\sigma^\alpha$; i.e., $z$ is the result of the exponential map of $-\sigma^\alpha$ at $y$. The vector $\sigma^\alpha$ (or more appropriately $-\sigma^\alpha$) at $y$ is like a curved-spacetime, covariant version of a displacement vector from $y$ to $z$.
\subsection{Coincidence limits}
The coincidence limit of a bitensor $T_{\alpha\mu\ldots}(y,z)$ is the ordinary tensor at $y$ obtained from the limit $y\to z$, denoted by $[T_{\alpha\mu\ldots}](y)$, where all indices are to be interpreted as indices at $y$. We assume that all such limits are independent of the path by which $z$ approaches $y$, which will be true for all of the bitensors discussed in this paper as long as the spacetime is sufficiently smooth.
The coincidence limits of the derivatives of the world function in a smooth spacetime follow the pattern$^{\ref{foot_R2}}$
\begin{gather}
[\sigma]=0
\nonumber\\
\qquad[\sigma_\alpha]=0,\qquad[\sigma_\mu]=0,\qquad
\nonumber\\
[{\sigma^\alpha}_\beta]={\delta^\alpha}_\beta,\quad
[{\sigma^\mu}_\nu]={\delta^\mu}_\nu,\quad
[{\sigma^\alpha}_\mu]=-{\delta^\alpha}_\mu,
\nonumber\\
[\sigma_{(3)}]=0,
\nonumber\\
[\sigma_{(4)}]\sim \mathrm{Riem},
\nonumber\\
[\sigma_{(n>4)}]\sim \nabla^{n-4}\mathrm{Riem},\label{ddsigma_cl}
\end{gather}
where $\sigma_{(n)}$ stands for any $n$th derivative of $\sigma$ (with derivatives at $y$ and/or $z$) and the last two right-hand sides represent various sums of permutations of the Riemann tensor and its covariant derivatives (evaluated at $y$). The derivations and results through $n=4$ can be found in Sec.~4 of Ref.~\cite{MPP}, and some higher-order results are given e.g.~in Refs.~\cite{Christensen1,companion}.
\subsection{The parallel propagator}
Given two points $y$ and $z$ linked by a unique geodesic segment $\Gamma$, and given a vector $A^\alpha$ at $y$, consider parallel transporting the vector along $\Gamma$, to obtain the vector $A^\mu$ at $z$. This is a linear map,
$$
A^{\mu}\,=\,{g^\mu}_\alpha(y,z)\, A^\alpha,
$$
from vectors at $y$ to vectors at $z$,
which defines the bitensor ${g^\mu}_\alpha(y,z)$ known as the parallel propagator.
Its inverse ${g^\alpha}_{\mu}$, satisfying ${g^\mu}_\alpha {g^\alpha}_\nu={\delta^\mu}_\nu$ and ${g^\mu}_\alpha {g^\beta}_\mu={\delta^\beta}_\alpha$, is simply ${g^\alpha}_\mu=g_{\mu\nu}g^{\alpha\beta}{g^\nu}_\beta$, so that ordering and raising or lowering of indices do not really matter. The parallel propagator can also be defined by the differential equations
\begin{equation}\label{pp}
\sigma^\beta {g^\mu}_{\alpha;\beta}=0=\sigma^{\nu} {g^\mu}_{\alpha;\nu},
\end{equation}
along with the first of the coincidence limits in the pattern$^{\ref{foot_R2}}$
\begin{gather}
[{g^\mu}_{\alpha}]={\delta^\mu}_{\alpha},
\nonumber\\
[ {g^\mu}_{\alpha;\beta}]=0,\qquad[ {g^\mu}_{\alpha;\nu}]=0,
\nonumber\\
[ {g^\mu}_{\alpha;(2)}]\sim\mathrm{Riem},
\nonumber\\
[ {g^\mu}_{\alpha;(n>2)}]\sim\nabla^{n-2}\mathrm{Riem}.\label{dpp_cl}
\end{gather}
The coincidence limits through $n=2$ are derived in Sec.~5 of Ref.~\cite{MPP}, and some higher order results can be found in Refs.~\cite{Christensen1,Christensen2,companion}.
\subsection{Expansions near coincidence}\label{sec:exp}
We saw in Sec.~\ref{sec:Synge} how the vector $(-)\sigma^\alpha$ at $y$ is like a covariant version of a displacement vector from $y$ to $z$. The covariant expansion of bitensors near coincidence (as $z\to y$) expands in powers of the vector $\sigma^\alpha(y,z)$ in analogy to how an ordinary Taylor expansion in flat space expands in powers of a coordinate displacement vector.
A smooth bitensor with indices only at $y$, like $T_{\alpha\beta}(y,z)$, can be expanded as
\begin{equation}\label{Texp}
T_{\alpha\beta}(y,z)=A_{\alpha\beta}+A_{\alpha\beta\gamma}\sigma^\gamma+\tfrac{1}{2}A_{\alpha\beta\gamma\delta}\sigma^\gamma\sigma^\delta+O(\sigma^\alpha)^3,
\end{equation}
where the $A$'s are ordinary tensors at $y$, determined by coincidence limits and derivatives thereof of $T$ and its derivatives (e.g.~$A_{\alpha\beta}=[T_{\alpha\beta}]$). For bitensors with indices at $z$, one factor of the parallel propagator is needed for each such index, to turn an expansion of the form (\ref{Texp}) into a bitensor of the proper index structure; for example,
$$
S_{\alpha\mu\nu}(y,z)={g^\beta}_\mu{g^\gamma}_\nu\Big(B_{\alpha\beta\gamma}+B_{\alpha\beta\gamma\delta}\sigma^\delta+O(\sigma^\alpha)^2\Big),
$$
where the $B$'s are ordinary tensors at $y$, determined from $S$ via derivatives and coincidence limits. Details of the standard (most straightforward) procedures to find the expansion coefficients, which work well at low orders, can be found in Sec.~6 of Ref.~\cite{MPP}. At higher orders, it becomes increasingly advantageous to employ the semi-recursive (or transport-equation) approach to bitensor expansions, developed by Refs.~\cite{Avramidi_thesis,Avramidi_book,Decanini,Barry_thesis,Ottewill} and utilized below in Appendix \ref{app:semi}.
The bitensors $\sigma^\alpha$ and ${g^\mu}_\alpha$ serve as the basic ingredients of coavriant expansions and cannot be expanded themselves (and similarly for $\sigma$ and $\sigma^\mu$). The lowest-order (most important) expansions are those of the second derivatives of the world function and the first derivatives of the parallel propagator, which follow the patterns\footnote{\label{foot_R2}At the higher orders in Eqs.~(\ref{ddsigma_cl}), (\ref{dpp_cl}), and (\ref{patt}), terms with multiple factors of the Riemann tensor and its derivatives get mixed into the coincidence limits and expansion coefficients, according to the patterns
$$
\nabla^2\mathrm{Riem}\sim (\mathrm{Riem})^2,\quad \nabla^3 \mathrm{Riem}\sim\mathrm{Riem}\cdot\nabla\mathrm{Riem},\quad\ldots
$$
which are seen when one commutes covariant derivatives of the Riemann tensor. See Eqs.~(\ref{LJIexp4}) for examples of these patterns.}
\begin{align}
\sigma_{(2)}&\;\sim\;\pm\delta+\mathrm{Riem}\cdot\sigma^2+\ldots+\nabla^{n-2}\mathrm{Riem}\cdot\sigma^{n}+\ldots
\nonumber\\\label{patt}
{g^\mu}_{\alpha;(1)}&\;\sim\; \mathrm{Riem}\cdot\sigma^1+\ldots+\nabla^{n-1}\mathrm{Riem}\cdot\sigma^n+\ldots
\end{align}
where $\sigma^n$ represents $n$ factors of the vector $\sigma^\alpha$, and we have omitted the overall factors of the parallel propagator on the right-hand sides. The patterns for the expansions of higher derivatives of these bitensors follow from a naive application of $\nabla \sigma^1\sim1$.
Some explicit low-order expansions illustrating the patterns (\ref{patt})---those used in the following section to derive the linear GDE and its action---are the world-function second derivatives,
\begin{eqnarray}\label{ddsigma}
{\sigma^\alpha}_\beta&=&{\delta^\alpha}_\beta-\tfrac{1}{3}{R^\alpha}_{\gamma\beta\delta}\sigma^\gamma\sigma^\delta+O(\sigma^\alpha)^3,
\\\nonumber
{\sigma^\alpha}_\mu&=&-{g^\beta}_\mu \Big({\delta^\alpha}_\beta+\tfrac{1}{6}{R^\alpha}_{\gamma\beta\delta}\sigma^\gamma\sigma^\delta+O(\sigma^\alpha)^3\Big),
\end{eqnarray}
and the parallel-propagator first derivatives,
\begin{eqnarray}\label{dg}
{g^\mu}_{\alpha;\beta}&=&\tfrac{1}{2}{g^\mu}_\gamma{R^\gamma}_{\alpha\beta\delta}\sigma^\delta+O(\sigma^\alpha)^2,
\\\nonumber
{g^\mu}_{\alpha;\nu}&=&\tfrac{1}{2}{g^\mu}_\gamma {g^\beta}_\nu {R^\gamma}_{\alpha\beta\delta}\sigma^\delta+O(\sigma^\alpha)^2.
\end{eqnarray}
\section{The linear GDE and its action}\label{sec:linear}
This section applies bitensor methods to derive the usual leading-order GDE and the action principle from which it follows. Those results easily follow from two more fundamental results, which apply to both geodesic and non-geodesic (accelerated) worldlines $z(\tau)$, namely, the expressions for the tangent and acceleration vectors of the worldline $z(\tau)$ in terms of the deviation vector field $\xi^\alpha(\tau)$ along the fiducial geodesic $y(\tau)$. The strategy here is to expand in powers of the deviation vector as soon as possible and manipulate expanded expressions. We will revisit these derivations in Sec.~\ref{sec:arb}, where the strategy will be to find relations valid to all orders in the deviation and then expand the final results.
Given a fiducial geodesic $y(\tau)$ and a vector field $\xi^\alpha(\tau)$ along it, we can specify a neighboring worldline $z(\tau)$ in an exact and covariant manner by the relation
\begin{equation}\label{neighbor}
\xi^\alpha(\tau)=-\sigma^\alpha\big(y(\tau),z(\tau)\big),
\end{equation}
so that $z(\tau)$ is the exponential map of $\xi^\alpha(\tau)$ at $y(\tau)$, as illustrated in Fig.~\ref{fig:xi}. We denote the tangent to the fiducial geodesic by $u^\alpha= dy^\alpha/d\tau$ and the tangent to the neighboring worldline by $v^\mu=dz^\mu/d\tau$.
\begin{figure}[h]
\includegraphics[scale=.5]{xialpha.pdf}
\caption{The point $z(\tau)$ on the neighboring worldline is the exponential map of the deviation vector $\xi^\alpha(\tau)$ at the point $y(\tau)$ on the fiducial geodesic.
}\label{fig:xi}
\end{figure}
We choose $\tau$ to be an affine parameter along the fiducial geodesic $y(\tau)$. For a generic vector field $\xi^\alpha(\tau)$, this would imply that the worldline $z(\tau)$ defined by Eq.~(\ref{neighbor}) is non-affinely parametrized. However, we will restrict attention at first (until Appendix \ref{app:gen}) to the case where $\tau$ is also an affine parameter along $z(\tau)$. This choice of `isochronous correspondence' [along with Eq.~(\ref{neighbor})] does not uniquely specify a vector field $\xi^\alpha$ given the geodesic $y$ and the worldline $z$, but it specifies a two-parameter family of vector fields related by affine reparametrizations, $\tau\to A\tau+B$, of $z(\tau)$. We will see below and in Appendix \ref{app:gen} that the `normal correspondence,' in which $\xi$ is constrained to be orthogonal to $u$, is a special case of the isochronous correspondence, for geodesic worldlines $z(\tau)$, to linear order in $\xi$ and $\dot\xi$ (but not at higher orders, and not for accelerated worldlines even at linear order). Note also that (except where otherwise indicated) our results are equally applicable to timelike, null, and spacelike geodesics.
We use dots to denote covariant $\tau$-derivatives $D/d\tau$; for example, the affinely parameterized geodesic equation for $y(\tau)$ reads $\dot u^\alpha=0$. We also employ shorthands such as $u^2=u_\alpha u^\alpha$ and $u\cdot\xi=u_\alpha\xi^\alpha$ for squares and contractions of vectors, and ${R^\alpha}_{\beta \xi u}={R^\alpha}_{\beta\gamma\delta}\xi^\gamma u^\delta$, etc., for contractions of vectors with the Riemann tensor (or its derivatives).
Now, consider acting on Eq.~(\ref{neighbor}), $\xi^\alpha=-\sigma^\alpha(y,z)$, with a total covariant $\tau$-derivative:
\begin{eqnarray}\label{zdot1}
\dot\xi^\alpha&=&\frac{D}{d\tau}\xi^\alpha=-\frac{D}{d\tau}\sigma^\alpha=-\left(u^\beta\nabla_\beta+v^\mu\nabla_\mu\right)\sigma^\alpha
\nonumber\\
&=&-u^\beta{\sigma^\alpha}_{\beta} -v^\mu{\sigma^\alpha}_\mu,
\end{eqnarray}
where we have hidden the $(y,z)$-dependence of the world-function derivatives and the $(\tau)$-dependence of $y$, $z$, $u$, $v$, $\xi$, and $\dot\xi$. The second derivatives of the world function in Eq.~(\ref{zdot1}) can be covariantly expanded in powers of the deviation vector $\xi^\alpha$, using the coincidence expansion as $z\to y$ in powers of $\sigma^\alpha(y,z)=-\xi^\alpha$, as in Sec.~\ref{sec:exp}. Inserting Eqs.~(\ref{ddsigma}), with $\sigma^\alpha=-\xi^\alpha$, into Eq.~(\ref{zdot1}) yields
\begin{eqnarray}
\dot\xi^\alpha&=&-u^\beta\left({\delta^\alpha}_\beta-\tfrac{1}{3}{R^\alpha}_{\xi\beta\xi}\right)
\\\nonumber
&&+\,v^\mu\, {g^\beta}_\mu \left({\delta^\alpha}_\beta+\tfrac{1}{6}{R^\alpha}_{\xi\beta\xi}\right)+O(\xi^3).
\end{eqnarray}
This equation can be solved for $v^\mu$ by working perturbatively in $\xi$ [using the zeroth-order solution $v^\mu={g^\mu}_\alpha (u^\alpha+\dot\xi^\alpha)+O(\xi)$ in the final term], yielding
\begin{eqnarray}\label{zdot_exp}
v^\mu&=&{g^\mu}_\alpha\left(u^\alpha+\dot\xi^\alpha-\tfrac{1}{2}{R^\alpha}_{\xi u\xi}\right)
\nonumber\\&&+\,O(\xi^3)+\dot\xi\cdot O(\xi^2),
\end{eqnarray}
which expresses the tangent vector $v^\mu$ of the neighboring worldline (which need not be a geodesic) in terms of tensors along the fiducial geodesic ($u$, $\xi$, $\dot\xi$, and the Riemann tensor) and the parallel propagator
${g^\mu}_\alpha(y,z)$.
We can find an action functional $S[\xi]$ which yields the GDE by expanding the well-known action $S[z]$ which yields geodesic motion for the worldline $z(\tau)$. The geodesic equation in affine parametrization, $\dot v^\mu=0$, follows from the action $S[z]=\tfrac{1}{2}\int d\tau\, v^2$. From Eq.~(\ref{zdot_exp}) and $g_{\mu\nu}{g^\mu}_\alpha{g^\nu}_\beta=g_{\alpha\beta}$, the square of the tangent vector is
\begin{eqnarray}\label{v2}
v^2&=&u^2+2u\cdot\dot\xi+\dot\xi^2-R_{\xi u\xi u}+O(\xi,\dot\xi)^3.
\end{eqnarray}
Since $u^2$ is a constant and $u\cdot\dot\xi$ is a total derivative (both because $\dot u^\alpha=0$), we can drop those terms from the action, and we obtain
\begin{eqnarray}\label{xiSg}
S_\mathrm{isoc.}[\xi]&=&\int d\tau \Big[\tfrac{1}{2}\dot\xi^2-\tfrac{1}{2}R_{\xi u\xi u} +O(\xi,\dot\xi)^3 \Big].
\end{eqnarray}
Varying this action with respect to $\xi$, using standard techniques, yields the linear GDE:\footnote{The analysis leading to the usual linear GDE (\ref{gde_lin}) requires one to expand in both the `deviation' $\xi$ and the `relative velocity' $\dot\xi$, taking $O(\xi)\sim O(\dot\xi)$, and working to order $O(\xi,\dot\xi)$, meaning dropping all terms with two or more factors of $\xi$ and/or $\dot\xi$. When $z(\tau)$ is a geodesic, Eq.~(\ref{gde_lin}) implies $\ddot\xi=O(\xi,\dot\xi)$. However, when $z(\tau)$ is accelerated, generically, $\ddot\xi=O(\xi,\dot\xi)^0$, which means that one must be careful in differentiating expanded expressions [as in Eq.~(\ref{amu})] because in that case $\tfrac{D}{d\tau} [O(\xi,\dot\xi)^n]=O(\xi,\dot\xi)^{n-1}$.}
\begin{equation}\label{gde_lin}
\ddot\xi^\alpha+{R^\alpha}_{u\xi u}=O(\xi,\dot\xi)^2.
\end{equation}
Note that the quantity $v^2$ of Eq.~(\ref{v2}) is a constant of the motion, because $z(\tau)$ is affinely parametrized. At $O(\xi,\dot\xi)$, this tells us $D/d\tau(u\cdot\dot\xi)=u\cdot\ddot\xi=O(\xi,\dot\xi)^2$, which also directly follows from the GDE (\ref{gde_lin}). This implies that $u\cdot\xi=A\tau+B$ is the general solution for the component of $\xi$ parallel to $u$ in the isochronous correspondence at linear order, and the choice $u\cdot\xi=0$ defining the normal correspondence is a consistent special case thereof.
We can also directly calculate the neighboring worldline's normalized acceleration vector $a^\mu$, which in affine parametrization is simply $a_\mathrm{isoc.}^\mu=\dot v^\mu$ (up to a constant rescaling). Using Eq.~(\ref{zdot_exp}), and treating $\ddot\xi$ as $O(\xi,\dot\xi)^0$ [as is the case for a generic (accelerated) worldline $z(\tau)$], we have
\begin{eqnarray}
a_\mathrm{isoc.}^\mu&=&\frac{D}{d\tau}\Big[{g^\mu}_\alpha\big(u^\alpha+\dot\xi^\alpha\big)
+O(\xi^2)+O(\xi,\dot\xi)^3\Big]
\nonumber\\\label{amu}
&=&{g^\mu}_\alpha\ddot\xi^\alpha+(u^\alpha+\dot\xi^\alpha)\frac{D}{d\tau} {g^\mu}_\alpha
+O(\xi,\dot\xi)^2.
\end{eqnarray}
The $\tau$-derivative of the parallel propagator is given by
$$
\frac{D}{d\tau}{g^\mu}_\alpha=u^\beta{g^\mu}_{\alpha;\beta}+v^\nu {g^\mu}_{\alpha;\nu}
=-{g^\mu}_\beta{R^\beta}_{\alpha u\xi}+O(\xi,\dot\xi)^2,\label{Dg}
$$
where the second equality has used $v^\nu={g^\nu}_\beta u^\beta+O(\xi,\dot\xi)$ from Eq.~(\ref{zdot_exp}) and the expansions (\ref{dg}) of the parallel-propagator derivatives with $\sigma^\alpha=-\xi^\alpha$. Plugging this into Eq.~(\ref{amu}) yields the expansion
\begin{equation}\label{amu2}
a_\mathrm{isoc.}^\mu={g^\mu}_\alpha\Big(\ddot\xi^\alpha+{R^\alpha}_{u\xi u}\Big)
+O(\xi,\dot\xi)^2
\end{equation}
for the acceleration of the neighboring worldline $z(\tau)$. The geodesic equation $a^\mu=0$ for the worldline then implies the GDE (\ref{gde_lin}) for the deviation vector (because ${g^\mu}_\alpha$ is invertible).
\section{The general solution to the linear GDE}\label{sec:Dixon}
As shown by Dixon \cite{Dixon,Dixon1,Dixon3}, one can write the general solution to the usual linear GDE [the linear GDE (\ref{gde_lin}) for the isochronous (or normal) correspondence] in terms of fundamental bitensors. While Dixon's original derivation relied on the definition (\ref{de}) of the deviation vector in terms of a one-parameter family of geodesics, this section presents a derivation based on the exponential map definition (\ref{neighbor}). This construction more easily generalizes to second order in the deviation, which we return to consider in Appendix~\ref{app:ntlo}.
Consider two points, $y$ with indices $\alpha$, $\beta$, etc., and $y'$ with indices $\alpha'$, $\beta'$, etc., linked by a unique geodesic segment which will serve as our fiducial geodesic. Taking $y$ as a fixed base point, while $y'$ moves along the fiducial geodesic, consider an affine parametrization of the geodesic given by $y'(\tau)$, with $y'(0)=y$, with the tangent $u^{\alpha'}(\tau)=dy^{\alpha'}/d\tau$ at $y'$, and with the tangent $u^\alpha=u^{\alpha'}(0)$ at $y$. As in Eqs.~(\ref{sigmau}), the tangent $u^\alpha$ at $y$ satisfies
\begin{equation}\label{rel1}
\tau u^\alpha=-\sigma^\alpha\big(y,y'(\tau)\big).
\end{equation}
\begin{figure}[h]
\includegraphics[scale=.5]{primes.pdf}
\caption{The deviation vector $\xi^\alpha$ and its derivative $\dot\xi^\alpha$ at the initial point $y$ on the fiducial geodesic determine the solution for the deviation vector $\xi^{\alpha'}$ at the second point $y'$.
}\label{fig:primes}
\end{figure}
As shown in the previous section, a vector field $\xi^{\alpha'}(\tau)$ along $y'(\tau)$ will satisfy the GDE (\ref{gde_lin}) if the neighboring worldline $z'(\tau)$ defined by
\begin{equation}\label{zpt}
\xi^{\alpha'}(\tau)=-\sigma^{\alpha'}\big(y'(\tau),z'(\tau)\big)
\end{equation}
is an affinely parametrized geodesic. We will use indices $\mu'$, $\nu'$, etc.~at the moving points $z'(\tau)$ on the neighboring geodesic, and indices $\mu$, $\nu$, etc.~at the base point \mbox{$z=z'(0)$.} Because Eq.~(\ref{gde_lin}) is second-order in $\tau$-derivatives, a solution for $\xi^{\alpha'}(\tau)$ is uniquely determined by the initial vector $\xi^\alpha=\xi^{\alpha'}(0)$ at $y$, satisfying
\begin{equation}\label{rel3}
\xi^\alpha=-\sigma^\alpha(y,z),
\end{equation}
and by the initial derivative $\dot\xi^\alpha=\dot\xi^{\alpha'}(0)$ at $y$.
Denoting the tangent to the neighboring geodesic by $v^{\mu'}(\tau)=dz^{\mu'}/d\tau$, its initial value $v^\mu=v^{\mu'}(0)$ at $z$ must satisfy
\begin{equation}\label{v_satisfy}
\sigma^\mu(z,z')=-\tau v^\mu= -\tau {g^\mu}_\beta\Big[u^\beta+\dot\xi^\beta+O(\xi,\dot\xi)^2\Big],
\end{equation}
where the first equality follows from Eqs.~(\ref{sigmau}) and the second equality follows from Eq.~(\ref{zdot_exp}).
Consider expanding the function $\sigma^\mu(z,z')$ as $z\to y$ with $z'$ held fixed. The expansion reads
\begin{align}\label{exp1}
\sigma^\mu(z,z')=&\;{g^\mu}_\beta(y,z)\Big[\sigma^\beta(y,z')-{\sigma^\beta}_\alpha(y,z')\,\sigma^\alpha(y,z)\nonumber\\
&+O\big(\sigma^\alpha(y,z)\big)^2\Big],
\end{align}
as can be verified by taking the coincidence limit as $z\to y$ of this equation and its covariant derivative with respect to $z$, using Eqs.~(\ref{ddsigma_cl}), (\ref{pp}), and (\ref{dpp_cl}). The functions $\sigma^\beta(y,z')$ and ${\sigma^\beta}_\alpha(y,z')$ appearing here can be similarly expanded as $z'\to y'$ with $y$ held fixed, with the results
\begin{align}\label{exp2}
\sigma^\beta(y,z')&=\sigma^\beta(y,y')-{\sigma^\beta}_{\alpha'}(y,y')\,\sigma^{\alpha'}(y',z')\nonumber\\&\quad+O\big(\sigma^{\alpha'}(y',z')\big)^2,
\\\nonumber
{\sigma^\beta}_\alpha(y,z')&={\sigma^\beta}_\alpha(y,y')+O\big(\sigma^{\alpha'}(y',z')\big).
\end{align}
Inserting the expansions (\ref{exp1}) and (\ref{exp2}) into Eq.~(\ref{v_satisfy}), using the relations (\ref{rel1}), (\ref{zpt}), and (\ref{rel3}), factoring out ${g^\mu}_\beta$, and simplifying yields
$$
-\tau\,\dot\xi^\beta={\sigma^\beta}_{\alpha'}(y,y')\,\xi^{\alpha'}+{\sigma^\beta}_\alpha(y,y')\,\xi^\alpha+O(\xi,\dot\xi)^2.
$$
If the matrix ${\sigma^\beta}_{\alpha'}(y,y')$ is invertible, we can solve this equation to find the desired solution for the deviation vector $\xi^{\alpha'}$ at $y'$ in terms of $\xi^\beta$ and $\dot\xi^\beta$ at $y$:
\begin{equation}\label{solution_lin}
\xi^{\alpha'}=\,{K^{\alpha'}}_\beta\,\xi^\beta\,+\,\tau\, {H^{\alpha'}}_\beta \,\dot\xi^\beta\,+\,O(\xi,\dot\xi)^2
\end{equation}
where the `Jacobi propagators' are given by
\begin{eqnarray}\label{Kdef}
{K^{\alpha'}}_\beta(y,y')&=&{H^{\alpha'}}_\gamma(y,y')\;{\sigma^\gamma}_\beta(y,y'),
\\\label{Hdef}
{H^{\alpha'}}_\beta(y,y')&=&-\Big({\sigma^\beta}_{\alpha'}(y,y')\Big)^{-1},
\end{eqnarray}
with ${}^{-1}$ denoting the matrix inverse. The matrices $-{\sigma^\beta}_{\alpha'}$ and ${H^{\alpha'}}_\beta$ are those denoted ${D^\beta}_{\alpha'}$ and ${D^{-1\,\alpha'}}_\beta$ by DeWitt and Brehme \cite{DeWittBrehme}, and we have adopted the notation ${K^{\alpha'}}_\beta$ and ${H^{\alpha'}}_\beta$ for the Jacobi propagators from Dixon \cite{Dixon}.
As discussed e.g.~by Refs.~\cite{DeWittBrehme,Harte2}, the matrix $-{\sigma^\beta}_{\alpha'}$ will be invertible as long as the points $y$ and $y'$ are connected by a unique geodesic segment (as long as $y'$ is in the NCN of $y$). When multiple geodesics connect $y$ to $y'$, generally forming `caustic surfaces,' $-{\sigma^\beta}_{\alpha'}$ becomes divergent (because $\sigma^\beta$ becomes multiple-valued), while its inverse ${H^{\alpha'}}_\beta$ is regular but has one or more zero eigenvalues. It turns out that the Jacobi propagators ${K^{\alpha'}}_\beta$ and ${H^{\alpha'}}_\beta$, unlike the world-function derivatives ${\sigma^\beta}_{\alpha'}$ and ${\sigma^\beta}_\alpha$, can be extended in a unique and well-defined way beyond the NCN.\footnote{\label{exp_foot}One way to extend the definitions of the Jacobi propagators beyond the NCN is to identify them as the horizontal and vertical covariant derivatives of the exponential map. This extends their domain of validity, in a nutshell, because the exponential map is well-defined and single-valued over the entire tangent bundle.
Let $\exp(y,\psi)$ denote the exponential map, giving the point (with coordinates $\exp^{\alpha'}$) determined by a point $y$ and a vector $\psi^\alpha$ at $y$, satisfying $\sigma^\alpha\big(y,\exp(y,\psi)\big)=-\psi^\alpha$ in the NCN; in the context of the above discussion, $\psi^\alpha=\tau u^\alpha$ and $\exp(y,\psi)=y'$; then
\begin{align}
{K^{\alpha'}}_\beta(y,y')&=\nabla_{\beta*} \exp^{\alpha'}(y,\psi)=\left(\frac{\partial}{\partial y^\beta}-\psi^\gamma\Gamma^\delta_{\beta\gamma}\frac{\partial}{\partial\psi^\delta}\right)\exp^{\alpha'},
\nonumber\\
{H^{\alpha'}}_\beta(y,y')&=\nabla_{*\beta} \exp^{\alpha'}(y,\psi)=\frac{\partial}{\partial\psi^\beta}\exp^{\alpha'},
\end{align}
where $\nabla_{\beta*}$ and $\nabla_{*\beta}$ denote the horizontal and vertical covariant derivatives of functions on the tangent bundle, introduced by Dixon \cite{Dixon}. The vertical derivative $\nabla_{*\beta}$ differentiates with respect to the vector $\psi$ at a fixed point $y$, while the horizontal derivative $\nabla_{\beta*}$ differentiates with respect to the point $y$ while parallel-transporting the vector $\psi$ along with it. We have identified (bitensor) functions of two points $(y,y')$ with functions of a point and a vector at that point $(y,\psi)$, via $(y,y')\to\big(y,\exp(y,\psi)\big)$, which implies
\begin{equation}
\nabla_{\beta*}=\nabla_\beta+{K^{\beta'}}_\beta \nabla_{\beta'},\quad \nabla_{*\beta}={H^{\beta'}}_\beta\nabla_{\beta'},
\end{equation}
where $\nabla_\beta$ and $\nabla_{\beta'}$ are the usual covariant derivatives with respect to $y$ and $\exp$. The identity (\ref{commute}) is then seen to be equivalent to the commutativity of the horizontal and vertical covariant derivatives: $$\nabla_{*\gamma}{K^{\alpha'}}_\beta=\nabla_{*\gamma}\nabla_{\beta*} \exp^{\alpha'}=\nabla_{\beta*}\nabla_{*\gamma} \exp^{\alpha'}=\nabla_{\beta*}{H^{\alpha'}}_\gamma.$$ These concepts are motivated and clarified in the companion paper \cite{companion}.} Considered as functions of the affine parameter $\tau$ along the geodesic $y'(\tau)$, ${K^{\alpha'}}_\beta(y,y'(\tau))$ and ${H^{\alpha'}}_\beta(y,y'(\tau))$ are the unique solutions to the differential equations
\begin{align}
\frac{D^2}{d\tau^2}{K^{\alpha'}}_\beta&={R^{\alpha'}}_{\gamma'\beta'\delta'}u^{\gamma'}u^{\delta'}{K^{\beta'}}_\beta,
\\\nonumber
\frac{D^2}{d\tau^2}\left(\tau{H^{\alpha'}}_\beta\right)&={R^{\alpha'}}_{\gamma'\beta'\delta'}u^{\gamma'}u^{\delta'}\left(\tau{H^{\beta'}}_\beta\right),
\end{align}
with the initial conditions
\begin{align*}
\left[{K^{\alpha'}}_\beta\right]=\left[{H^{\alpha'}}_\beta\right]={\delta^{\alpha'}}_\beta,
\\
\left[\frac{D}{d\tau}{K^{\alpha'}}_\beta\right]=\left[\frac{D}{d\tau}{H^{\alpha'}}_\beta\right]=0,
\end{align*}
where the $y'\to y$ coincidence limits (denoted with brackets) correspond to $\tau\to0$. These equations can be deduced directly from the solution (\ref{solution_lin}) to the linear GDE, or (with some effort) by differentiating the definitions (\ref{Kdef}) and (\ref{Hdef}) and using properties of the world-function derivatives. Another identity which follows from Eqs.~(\ref{Kdef}) and (\ref{Hdef}) and will be useful in Sec.~\ref{sec:arb} is
\begin{equation}\label{commute}
{K^{\alpha'}}_{\beta;\gamma'} {H^{\gamma'}}_\gamma = {H^{\alpha'}}_{\gamma;\beta}+{H^{\alpha'}}_{\gamma;\beta'} {K^{\beta'}}_\beta.
\end{equation}
Because the normal correspondence ($u\cdot\xi=0$) is a special case of the isochronous correspondence at this order, the solution (\ref{solution_lin}) also applies to the normal case. It is instructive to note the identities
\begin{align}
\sigma^\alpha {K^{\alpha'}}_\alpha&=-\sigma^{\alpha'},\quad\quad\qq \sigma_{\alpha'} {K^{\alpha'}}_\alpha=-\sigma_\alpha,
\nonumber\\\label{uu}
\Rightarrow\quad u^\alpha {K^{\alpha'}}_\alpha&=u^{\alpha'},\quad\quad\qq\phantom{-} u_{\alpha'} {K^{\alpha'}}_\alpha=u_\alpha,
\end{align}
as well as all four equations with $K\to H$, which follow from Eqs.~(\ref{Kdef}) and (\ref{Hdef}) and derivatives of Eq.~(\ref{sigma2}). The identities (\ref{uu}) imply that projection orthogonal to the tangent commutes with propagation by $K$ or $H$; i.e.
$$
{K^{\alpha'}}_\beta {P^{\beta}}_\alpha={P^{\alpha'}}_{\beta'} {K^{\beta'}}_\alpha,\quad\quad {H^{\alpha'}}_\beta {P^{\beta}}_\alpha={P^{\alpha'}}_{\beta'} {H^{\beta'}}_\alpha,
$$
where ${P^\alpha}_\beta={\delta^\alpha}_\beta+u^\alpha u_\beta$ and ${P^{\alpha'}}_{\beta'}={\delta^{\alpha'}}_{\beta'}+u^{\alpha'} u_{\beta'}$ are the the tensors which project orthogonal to $u^\alpha$ and $u^{\alpha'}$ (in the timelike case). This means that, if the initial deviation vector $\xi^\alpha$ and the initial derivative $\dot\xi^\alpha$ are orthogonal to $u^\alpha$ at $y$, then the final deviation vector $\xi^{\alpha'}$, the solution to the GDE given by Eq.~(\ref{solution_lin}), will be orthogonal to $u^{\alpha'}$ at $y'$.
\section{The GDE and its action to arbitrary orders}\label{sec:arb}
We now return to consider forms of the GDE and of all the other results of Sec.~\ref{sec:linear} which are valid to all orders in the deviation. We use all of the same notation and conventions from Sec.~\ref{sec:linear}, and our starting point once again is the exponential map relation,
\begin{equation}\label{exp_arb}
\xi^\alpha(\tau)=-\sigma^\alpha\big(y(\tau),z(\tau)\big),
\end{equation}
which, given a vector field $\xi^\alpha(\tau)$ along an affinely parametrized fiducial geodesic $y(\tau)$ with tangent $u^\alpha=dy^\alpha/d\tau$, specifies a neighboring worldline $z(\tau)$ with tangent $v^\mu=dz^\mu/d\tau$, as illustrated in Fig.~\ref{fig:xi} above. For now, we place no restrictions on the vector field $\xi^\alpha(\tau)$ [except that its norm or extent be small enough to keep $z$ in the NCN of $y$], so that the neighboring worldline $z(\tau)$ is generically non-geodesic and non-affinely parametrized.
We begin the derivation just as in Sec.~\ref{sec:linear}, by acting on the exponential map relation (\ref{exp_arb}) with a total covariant $\tau$-derivative, yielding the first of Eqs.~(\ref{v_arb}) below. Solving (exactly) for the neighbor's tangent $v^\mu$, using the definitions (\ref{Kdef}) and (\ref{Hdef}) of the Jacobi propagators ${K^\mu}_\beta(y,z)$ and ${H^\mu}_\beta(y,z)$ in terms of the world-function derivatives ${\sigma^\alpha}_\beta(y,z)$ and ${\sigma^\alpha}_\mu(y,z)$, yields the second equation:
\newpage
\begin{widetext}
\begin{equation}\label{v_arb}
\dot\xi^\alpha=-{\sigma^\alpha}_\beta u^\beta -{\sigma^\alpha}_\mu v^\mu
\quad\quad\qq\quad\quad\qq\Leftrightarrow\quad\quad\qq\quad\quad\qq
v^\mu={K^\mu}_\beta u^\beta + {H^\mu}_\beta \dot\xi^\beta.
\end{equation}
Acting on each of these relations with another $\tau$-derivative yields
\begin{align*}
\quad\quad\qq
\begin{split}
\ddot\xi^\alpha=&\;-{\sigma^\alpha}_{\beta\gamma}u^\beta u^\gamma - 2{\sigma^\alpha}_{\beta\mu}u^\beta v^\mu
\\
&\;- {\sigma^\alpha}_{\mu\nu}v^\mu v^\nu -{\sigma^\alpha}_\mu \dot v^\mu
\end{split}
&\quad\quad\qq\Leftrightarrow\quad\quad\qq&
\begin{split}
\dot v^\mu=&\;\big({K^\mu}_{\beta;\gamma}u^\gamma+{K^\mu}_{\beta;\nu}v^\nu\big) u^\beta
\\
&\;+ \big({H^\mu}_{\beta;\gamma}u^\gamma+{H^\mu}_{\beta;\nu}v^\nu\big) \dot\xi^\beta+{H^\mu}_\beta \ddot\xi^\beta.
\end{split}
\end{align*}
Taking either of these relations, substituting from Eq.~(\ref{v_arb}) for $v^\mu$ (but not for $\dot v^\mu$), and solving for $\dot v^\mu$, one finds
\begin{equation}\label{vdot_arb}
\dot v^\mu\,=\,{H^\mu}_\alpha \ddot\xi^\alpha\, + \,{L^\mu}_{\beta\gamma}u^\beta u^\gamma
\,+\,2{J^\mu}_{\beta\gamma}u^\beta \dot\xi^\gamma\,+\,{I^\mu}_{\beta\gamma}\dot\xi^\beta \dot\xi^\gamma\,,
\end{equation}
where the expressions for the bitensors ${L^\mu}_{\beta\gamma}(y,z)$, ${J^\mu}_{\beta\gamma}(y,z)$, and ${I^\mu}_{\beta\gamma}(y,z)$ obtained from the two derivations are
\begin{align}
\begin{split}
{L^\mu}_{\beta\gamma}&={H^\mu}_\alpha\Big({\sigma^\alpha}_{\beta\gamma}+2{\sigma^\alpha}_{\nu(\beta}{K^\nu}_{\gamma)}+{\sigma^\alpha}_{\nu\rho}{K^\nu}_\beta{K^\rho}_\gamma\Big)
\\
{J^\mu}_{\beta\gamma}&={H^\mu}_\alpha {H^\nu}_\gamma \Big({\sigma^\alpha}_{\beta\nu}+{\sigma^\alpha}_{\rho\nu}{K^\rho}_\beta\Big)
\\
{I^\mu}_{\beta\gamma}&={H^\mu}_\alpha {H^\nu}_\beta {H^\rho}_\gamma {\sigma^\alpha}_{\nu\rho}\phantom{\Big(}
\end{split}
&\quad\quad\Leftrightarrow\quad\quad&
\begin{split}
{L^\mu}_{\beta\gamma}&={K^\mu}_{\beta;\gamma}+{K^\mu}_{\beta;\nu}{K^\nu}_{\gamma}\phantom{\Big(}
\\
{J^\mu}_{\beta\gamma}&={K^\mu}_{\beta;\nu}{H^\nu}_{\gamma}\phantom{\Big(}
\\
{I^\mu}_{\beta\gamma}&={H^\mu}_{\beta;\nu}{H^\nu}_{\gamma}\phantom{\Big(}
\end{split},\label{LJI}
\end{align}
\end{widetext}
and we have used the identity (\ref{commute}) in the expression for ${J^\mu}_{\beta\gamma}$ on the right.\footnote{\label{sojp_foot} The bitensors of Eqs.~(\ref{LJI}) can be identified as the second horizontal/vertical derivatives of the exponential map $\exp(y,\xi)=z$:
\begin{align}\label{ddX}
{L^\mu}_{\beta\gamma}&=\nabla_{\gamma*}\nabla_{\beta*} \exp^\mu,\nonumber
\\
{J^\mu}_{\beta\gamma}&=\nabla_{*\gamma}\nabla_{\beta*} \exp^\mu,
\\
{I^\mu}_{\beta\gamma}&=\nabla_{*\gamma}\nabla_{*\beta} \exp^\mu=\nabla_{*\beta}\nabla_{*\gamma} \exp^\mu.\nonumber
\end{align}
(See Footnote \ref{exp_foot} and Ref.~\cite{companion}.)
The bitensor ${I^\mu}_{\beta\gamma}$ is symmetric in $\beta$ and $\gamma$, as can be seen from the expression on the left in Eqs.~(\ref{LJI}), while ${J^\mu}_{\beta\gamma}$ and ${L^\mu}_{\beta\gamma}$ are not. While the derivation of Eq.~(\ref{vdot_arb}) determines only the $\beta\gamma$-symmetric part of ${L^\mu}_{\beta\gamma}$, the non-symmetric definitions given in Eqs.~(\ref{LJI}) have been chosen to match the ${L^\mu}_{\beta\gamma}$ of Eqs.~(\ref{ddX}).}\
The important results (\ref{v_arb}b) and (\ref{vdot_arb}) express the tangent $v^\mu$ to $z(\tau)$ and its covariant $\tau$-derivative $\dot v^\mu$ in terms of the vectors $u^\alpha$, $\dot\xi^\alpha$, and $\ddot\xi^\alpha$ along $y(\tau)$ and the bitensors $K$, $H$, $L$, $J$, and $I$. These bitensors, functions of $(y,z)$, harbor the dependence on the (undifferentiated) deviation vector $\xi^\alpha=-\sigma^\alpha(y,z)$, and they can be covariantly expanded in powers of $\xi$.
Before expanding, we can write out the `exact' forms of the GDE and its action. In the isochronous correspondence, when the worldline $z(\tau)$ defined by Eq.~(\ref{exp_arb}) is affinely parametrized, its acceleration vector is $a^\mu_\mathrm{isoc.}=\dot v^\mu$ (up to a constant rescaling), and thus, from Eq.~(\ref{vdot_arb}), the geodesic equation for $z(\tau)$, or the GDE for $\xi(\tau)$, reads
\begin{align}\label{gde_exact}
0&=a_\mathrm{isoc.}^\mu=\dot v^\mu
\\\nonumber
&={H^\mu}_\alpha \ddot\xi^\alpha + {L^\mu}_{\beta\gamma}u^\beta u^\gamma
+2{J^\mu}_{\beta\gamma}u^\beta \dot\xi^\gamma+{I^\mu}_{\beta\gamma}\dot\xi^\beta \dot\xi^\gamma.
\end{align}
This equation, along with Eqs.~(\ref{LJI}), is equivalent to Eqs.~(6) of Aleksandrov and Piragas \cite{Aleksandrov}.
The action principle which yields the exact GDE (\ref{gde_exact}) can be found from the action $S[z]=\tfrac{1}{2}\int d\tau\,v^2$ for affinely parametrized geodesic motion; using Eq.~(\ref{v_arb}), we have
\begin{align}\label{S_exact}
&S_\mathrm{isoc.}[\xi]=\int d\tau\, \mathcal L_\mathrm{isoc.}(\xi,\dot\xi)=\tfrac{1}{2}\int d\tau\, v^2,
\\\nonumber
&v^2={K^\mu}_\beta K_{\mu\gamma}u^\beta u^\gamma+2{K^\mu}_\beta H_{\mu\gamma}u^\beta\dot\xi^\gamma+{H^\mu}_{\beta} H_{\mu\gamma}\dot\xi^\beta \dot\xi^\gamma.
\end{align}
The Lagrangian is a constant of the motion, because $z(\tau)$ is affinely parametrized.
The exact results (\ref{gde_exact}) and (\ref{S_exact}) can be consistently expanded in two ways: to $O(\xi,\dot\xi)^n$, or to $O(\xi^n)$ but to all orders in $\dot\xi$. The expansions of the bitensors $K$, $H$, $L$, $J$, and $I$, to the orders needed to derive Bazanski's equation at $O(\xi,\dot\xi)^2$ and the generalized Jacobi equation at $O(\xi)$ (and their actions) are given by
\begin{align}\nonumber
{K^\mu}_\beta=&\;{g^\mu}_\alpha\bigg[{\delta^\alpha}_\beta-\tfrac{1}{2}{R^\alpha}_{\xi\beta\xi}-\tfrac{1}{6}{R^\alpha}_{\xi\beta\xi;\xi}
+O(\xi^4)\bigg],
\\\nonumber
{H^\mu}_\beta=&\;{g^\mu}_\alpha\bigg[{\delta^\alpha}_\beta-\tfrac{1}{6}{R^\alpha}_{\xi\beta\xi}
+O(\xi^3)\bigg],
\\\nonumber
{L^\mu}_{\beta\gamma}=&\;{g^\mu}_\alpha\bigg[{R^\alpha}_{\beta\xi\gamma}+\tfrac{1}{2}\Big({R^\alpha}_{\beta\xi\gamma;\xi}-{R^\alpha}_{\xi\beta\xi;\gamma}\Big)
+O(\xi^3)\bigg],
\\\nonumber
{J^\mu}_{\beta\gamma}=&\;{g^\mu}_\alpha\bigg[{R^\alpha}_{\gamma\xi\beta}
+O(\xi^2)\bigg],
\\
{I^\mu}_{\beta\gamma}=&\;{g^\mu}_\alpha\bigg[-\tfrac{2}{3}{R^\alpha}_{(\beta\gamma)\xi}
+O(\xi^2)\bigg].
\label{LJIexp}
\end{align}
Recursion relations which generate these expansions to all orders are presented in Appendix \ref{app:semi}, along with explicit results through the orders necessary to derive the $O(\xi,\dot\xi)^4$ GDE and its action.
The following two subsections discuss some details of the two expansion schemes and mention some physical applications of Bazanski's equation and the generalized Jacobi equation found in the literature.
\subsection{Exapnding to $O(\xi,\dot\xi)^n$; \\Bazanski's equation $(n=2)$}
The $O(\xi,\dot\xi)^n$ expansion scheme treats both the deviation $\xi$ and the relative velocity $\dot\xi$ as small parameters, and assumes furthermore that they are of the same order of smallness. More precisely, it assumes
$
\dot\xi\sim\xi/\mathcal R\ll 1,
$
where $\mathcal R$ measures the spacetime's radius of curvature. This is the approximation that leads to the usual linear GDE (\ref{gde_lin}) at leading order.
The explicit GDE valid to $O(\xi,\dot\xi)^n$ is found by taking the exact GDE (\ref{gde_exact}), inserting the expansions of ${H^\mu}_\alpha$ and ${L^\mu}_{\beta\gamma}$ to $O(\xi^n)$, of ${J^\mu}_{\beta\gamma}$ to $O(\xi^{n-1})$, and of ${I^\mu}_{\beta\gamma}$ to $O(\xi^{n-2})$, and peeling off an overall factor of ${g^\mu}_\alpha$.
At $O(\xi,\dot\xi)^2$, using the expansions (\ref{LJIexp}), we obtain `Bazanski's equation' \cite{Hodgkinson,Bazanski,Bazanski2,Aleksandrov} in the isochronous correspondence:
\begin{equation}\label{Beq}
\ddot\xi^\alpha+{R^\alpha}_{u\xi u}+R\indices{_{\xi u}^\alpha_{(\xi;u)}}+2{R^\alpha}_{\dot\xi\xi u}=O(\xi,\dot\xi)^3
\end{equation}
We will see in Appendix \ref{app:gen} that this equation also applies to the normal correspondence if it is projected orthogonal to $u$ on the $\alpha$ index. For geodesic worldlines $z(\tau)$, through $O(\xi,\dot\xi)^2$, the normal correspondence is a special case of the isochronous correspondence; this is not true at $O(\xi,\dot\xi)^3$ and higher. The second-order GDE (\ref{Beq}) was first derived by Hodgkinson \cite{Hodgkinson}, and the derivation was later improved by Bazanski \cite{Bazanski,Bazanski2} and Aleksandrov and Piragas \cite{Aleksandrov}.
The GDE through $O(\xi,\dot\xi)^3$, given by Eqs.~(\ref{gde_exact}, \ref{LJIexp4}), was also first derived in Ref.~\cite{Hodgkinson} and later rederived in Ref.~\cite{Aleksandrov}; Eqs.~(\ref{gde_exact}, \ref{LJIexp4}) also present for the first time the $O(\xi,\dot\xi)^4$ GDE. While all of these results are specialized to the isochronous correspondence, we discuss their generalizations to arbitrary correspondences and their specializations to the normal correspondence in Appendix \ref{app:gen}.
The action functional which yields the $O(\xi,\dot\xi)^n$ GDE can be found by expanding the exact action (\ref{S_exact}) to $O(\xi,\dot\xi)^{n+1}$. From Eqs.~(\ref{LJIexp}), the Lagrangian for the isochronous Bazanski's equation (\ref{Beq}) is
\begin{equation}\label{BS}
\mathcal L_\mathrm{isoc.}=\tfrac{1}{2}\dot\xi^2-\tfrac{1}{2}R_{\xi u\xi u}
-\tfrac{2}{3}R_{\xi\dot\xi\xi u}-\tfrac{1}{6}R_{\xi u\xi u;\xi}+O(\xi,\dot\xi)^4.
\end{equation}
The Lagrangian through $O(\xi,\dot\xi)^5$ was quoted as Eq.~(\ref{S}) in the introduction.
Bazanski's equation has found several fruitful physical applications. An early example is the analysis of Tammelo \cite{Tammelo,Tammelo2}, which studied the effects of second-order geodesic deviations in gravitational wave detectors (focusing on Weber-type detectors), emphasizing the fact that second-order effects lead to a net longitudinal force, or pressure, on the detector. In another application to gravitational wave detectors, Baskaran and Grishchuk \cite{Baskaran} showed that the original data analysis scheme for LIGO, which modeled the detector via the linear GDE, could be off by as much as 10\% because it ignored the second-order ``magnetic component of the gravitational force,'' arising from the fourth term in Eq.~(\ref{Beq}).
Another set of applications of Bazanski's equation is represented by the works of Kerner, van Holten, Colistete, and collaborators \cite{Kerner,vanHolten,Colistete,Colistete2,Koekoek,Koekoek2}. These works have used higher-order geodesic deviations to generate analytic approximations for mildly eccentric orbits in the Schwarzschild and Kerr spacetimes, terming them ``relativistic epicycles,'' and have applied these methods to derive analytic approximations for the leading-order gravitational waveforms from small bodies in nearly circular orbits around large black holes.
\subsection{Expanding to $O(\xi^n)$;\\ the generalized Jacobi equation $(n=1)$}
The $O(\xi^n)$ expansion scheme assumes small deviations, $\xi/\mathcal R\ll 1$, but allows arbitrary relative velocities, with $\dot\xi\sim 1$ corresponding to relativistic relative velocities. At $O(\xi)$, this approximation leads to the `generalized Jacobi equation' (GJE), first derived by Hodgkinson \cite{Hodgkinson} and further developed by Mashhoon and others \cite{Mashhoon75,Mashhoon77,LiNi,Ciufolini,Ciufolini2,Chicone02,Chicone06a,Chicone06b,Perlick}.
The GJE in the isochronous correspondence can be found by taking the exact GDE (\ref{gde_exact}) and inserting the expansions (\ref{LJIexp}), all to $O(\xi)$, yielding
\begin{equation}\label{GJE_isoc}
\ddot\xi^\alpha+{R^\alpha}_{u\xi u}+2{R^\alpha}_{\dot\xi\xi u}+\tfrac{2}{3}{R^\alpha}_{\dot\xi\xi\dot\xi}=O(\xi^2).
\end{equation}
This equation can be generalized to higher orders in $\xi$ by using the higher order expansions in Eqs.~(\ref{LJIexp4}) below. The GJE in the normal correspondence (the form in which it is usually used) differs slightly from Eq.~(\ref{GJE_isoc}) and is given by Eq.~(\ref{GJEnorm}) below.
The action yielding Eq.~(\ref{GJE_isoc}) results from inserting the expansions (\ref{LJIexp}) to $O(\xi^2)$ into the exact action (\ref{S_exact}); the resultant Lagrangian is
\begin{equation}\label{GJES_isoc}
\mathcal L_\mathrm{isoc.}=\tfrac{1}{2}\dot\xi^2-\tfrac{1}{2}R_{\xi u\xi u}
-\tfrac{2}{3}R_{\xi\dot\xi\xi u}-\tfrac{1}{6}R_{\xi\dot\xi\xi\dot\xi}+O(\xi)^3.
\end{equation}
Higher orders can be generated from the expansions (\ref{LJIexp4}), and the Lagrangian for the normal correspondence is given by Eq.~(\ref{GJEnorm}).
The GJE has been applied to a number of astrophysical problems, most notably by Mashhoon and Chicone \cite{Mashhoon75, Mashhoon77,Chicone02,Chicone06a,Chicone06b}. In Refs.~\cite{Mashhoon75, Mashhoon77}, Mashhoon used the GJE to analyze the tidal dynamics of an extended body in a gravitational field, and to estimate the tidal gravitational radiation emitted by such a body, focusing on orbiting bodies in the Kerr spacetime. Continuing this line of investigation, Chicone and Mashhoon \cite{Chicone02,Chicone06a,Chicone06b} applied the GJE to collections of test particles in the Kerr spacetime, discussing applications to astrophysical relativistic jets, and to tidal dynamics in plane-wave spacetimes and the de Sitter and G\"odel spacetimes.
Other applications of the GJE include Li and Ni's analysis of the coupling of inertial and gravitational effects for accelerated and rotating observers \cite{LiNi}, Ciufolini and Demianski's prescriptions for measuring spacetime curvature using the GJE \cite{Ciufolini,Ciufolini2}, and Perlick's application of the GJE to null goedesics in plane-wave spacetimes and the Schwarzschild spacetime \cite{Perlick}.
\acknowledgements
The author would like to thank \'Eanna Flanagan, David Nichols, Leo Stein, and Barry Wardell for many helpful discussions, and to acknowledge support from
NSF grants PHY-1068541 and PHY-1404105.
|
\section{Game Setting}
We adopt a typical networked coordination game~\cite{ellison1993learning} setting to model the influences between behaviors of individuals. Consider an undirected graph $G=(V,E)$, in which the nodes are the individuals in the population, and edges denote that they are friends whose behaviors would influence each other. This is a simple model of a \emph{social network}.
We will consider a well concerned situation~\cite{ellison1993learning,kreindler2013rapid},
where each node has two behaviors: the \emph{new} behavior (innovation) labeled as $A$, and the \emph{old} behavior (status quo) labeled as $B$.
Each pair of adjacent individuals $(v,w)$ would receive a payoff from each other according to the following rules:
\begin{itemize}
\item if both $v$ and $w$ have the same new behavior $A$, they each receive a payoff $\alpha$, where $\alpha$ is a positive real number.
\item if both $v$ and $w$ have the same old behaviors $B$, they each receive a payoff $\beta$, where $\beta$ is a positive real number and $\beta\leq\alpha$.
\item if $v$ and $w$ have different behaviors, they receive payoff $\gamma$, where $0\leq \gamma\leq \beta$,
which can be interpreted as the compatibility between the innovation and the status quo.
\end{itemize}
The utility of each individual is the sum of payoffs she received from all her neighbors according to the aforementioned rules.
This can then be intuitively formulated as a game $\Gamma(u_1,u_2,\cdots,u_p)$ for $p\geq2$ players (individuals).
Then the set of players is $V=\{v_1,v_2,\cdots,v_p\}$, the set of strategies (behaviors) of the player $v_i$ is $\mathbb{S}_i=\{A,B\}$, and the payoff function of the player $i$ is $u_i: \mathbb{S}\rightarrow \mathbb{R}$, where $\mathbb{S}=\mathbb{S}_1\times \mathbb{S}_2\times\cdots\times \mathbb{S}_p$ is the set of strategy profiles, and $\mathbb{R}$ denotes the set of real numbers.
To define the $u_i$, we first define a weight $W(s_i,s_j)$ of each edge $(v_i,v_j)\in E$ as a function of the used strategies $(s_i,s_j)$ of its two endpoints.
$W(s_i,s_j)$ can be defined as the following payoff matrix:
\vspace{0.1in}
\setlength{\tabcolsep}{12pt}
\begin{tabular}{lll}
&A&B\\
A&$\alpha$&$\gamma$\\
B&$\gamma$&$\beta$\\
\end{tabular}
\vspace{0.05in}
Then given a strategy vector, $\mathbf{s}\in\mathbb{S}$, the total payoff of each player $u_i(\mathbf{s})$ and the social welfare $SW(\mathbf{s})$ are defined as follows:
\begin{eqnarray}
u_i(\mathbf{s})&=&\sum_{j,(v_i,v_j)\in E}W(s_i,s_j),\notag\\
SW(\mathbf{s})&=&\sum_{i,v_i\in V}u_i.\notag
\end{eqnarray}
\section{Upper Bound for Price of Anarchy}\label{sec:poabound}
In this section, we will show a tight upper bound of the PoA of the game $\Gamma$. We first define three types of edges in the graph as follows:
\begin{definition}
The edge $(v_i,v_j)$ linking two players using the same strategy $A$, \emph{i.e}\onedot} \def\Ie{\emph{I.e}\onedot, weighted $\alpha$, is called $A$-edge denoted as $e_a$.
The edge $(v_i,v_j)$ linking two players using the same strategy $B$, \emph{i.e}\onedot} \def\Ie{\emph{I.e}\onedot, weighted $\beta$, is called $B$-edge denoted as $e_b$.
The edge $(v_i,v_j)$ linking two players using the different strategies, \emph{i.e}\onedot} \def\Ie{\emph{I.e}\onedot, weighted $\gamma$, is called $C$-edge denoted as $e_c$.
\end{definition}
\begin{definition}
Then the total payoff of each player and the social welfare can be defined with respect to three types of edges:
\begin{eqnarray}
u_i(n_i^{e_a},n_i^{e_b},n_i^{e_c})&=&n_i^{e_a}\alpha+n_i^{e_b}\beta+n_i^{e_c}\gamma,\notag\\
SW(n^{e_a},n^{e_b},n^{e_c})&=&2n^{e_a}\alpha+2n^{e_b}\beta+2n^{e_c}\gamma,\notag
\end{eqnarray}
where $n_i^{e_a},n_i^{e_b},n_i^{e_c}$ denote the number of its incident $A$-edges, $B$-edges, $C$-edges, and $n^{e_a}, n^{e_b}, n^{e_c}$ denote the total number of $A$-edges, $B$-edges, $C$-edges in the graph: $n^{e_a}=\frac{1}{2}\sum_in_i^{e_a},n^{e_b}=\frac{1}{2}\sum_in_i^{e_b},n^{e_c}=\frac{1}{2}\sum_in_i^{e_c}$. We define the tuple $(n_i^{e_a},n_i^{e_b},n_i^{e_c})$ as $\mathbf{n}_i$ and its space is defined as $\mathbb{N}_i$, correspondingly, $\mathbf{n}=(n^{e_a},n^{e_b},n^{e_c})\in\mathbb{N}$. Each term times $2$ in the social welfare is because each edge is computed twice for its two incident nodes.
\end{definition}
\begin{definition}
\begin{equation}
PoA=\frac{\max_{\mathbf{n}\in\mathbb{N}}SW(\mathbf{n})}{\min_{\mathbf{n}\in\mathbb{N}^e}SW(\mathbf{n})},\notag
\end{equation}
where $\mathbb{N}^e$ is the space of $\mathbf{n}$ in NEs.
\end{definition}
It is easy to compute that the optimal value $SW(\mathbf{opt})$ of the social welfare $\max_{\mathbf{n}\in\mathbb{N}}SW(\mathbf{n})=2n^e\alpha$, where $n^e$ is the total number of edges, when the weights of all edges are equal to $\alpha$, since $\alpha$ is the maximum value of the weight.
Then the problem is to consider the social welfare for all NEs. It is challenging since there could be many NEs in this game. We discuss them respectively in terms of edge types.
For a state $\mathbf{n}$, define the following quotient:
\begin{definition}
\begin{equation*}
r(\mathbf{n})=r(n^{e_a},n^{e_b},n^{e_c})=\frac{2(n^{e_a}+n^{e_b}+n^{e_c})\alpha}{2n^{e_a}\alpha+2n^{e_b}\beta+2n^{e_c}\gamma}
\end{equation*}
\end{definition}
Then
$$PoA = \max_{\mathbf{n}\in \mathbb{N}^e}(r(n^{e_a},n^{e_b},n^{e_c}))$$
We begin bounding this quotient by decomposing $G$ into two subgraphs.
\begin{definition}
Define $\Psi = (H_\Phi, H_\Lambda)$ be a decomposition of $G$,
where the edge set $E(H_\Phi)$ and $E(H_\Lambda)$ form a partition of $E(G)$.
$H_\Phi$ is defined to be an edge-induced subgraph of $G$ where $E(H_\Phi)$ contains
all $C$-edges \textbf{and} those who share endpoints to $C$-edges.
Consequently, the remaining edges constitute $H_\Lambda$ where
$E(H_\Lambda) \triangleq E(G)-E(H_\Phi)$.
\label{def:decomp}
\end{definition}
Note under this decomposition,
some vertices occur in both $H_\Phi$ and $H_\Lambda$,
which is innocuous in our analysis since the payoffs reside on edges rather than vertices.
One can view that nodes are duplicated as needed during the decomposition.
\begin{definition}
For a Nash equilibrium ${\bf n}=(n^{e_a}, n^{e_b}, n^{e_c})$ in $G$,
denote the corresponding states in $H_\Phi$ and $H_\Lambda$
as ${\mathbf{n}_\Phi}=(n_\Phi^{e_a}, n_\Phi^{e_b}, n_\Phi^{e_c})$
and ${\mathbf{n}_\Lambda}=(n_\Lambda^{e_a}, n_\Lambda^{e_b}, n_\Lambda^{e_c})$
where $n_\Phi^{e_a}+n_\Lambda^{e_a} = n^{e_a}$,
$n_\Phi^{e_b}+n_\Lambda^{e_b} = n^{e_b}$,
and $n_\Phi^{e_c}+n_\Lambda^{e_c} = n^{e_c}$.
\end{definition}
From Definition~\ref{def:decomp},
for any state $\mathbf{n}$ and the corresponding $\mathbf{n}_\Lambda$ and $\mathbf{n}_\Phi$,
it satisfies the following properties.
i) There is no $C$-edge in $\mathbf{n}_\Lambda$, i.e. $n_\Lambda^{e_c}=0$.
ii) Every edge in $\mathbf{n}_\Phi$ either is a $C$-edge or shares an endpoint to a $C$-edge.
\begin{lemma}
For any NE $\bf n$ in $G$,
the corresponding $\mathbf{n}_\Phi$ and $\mathbf{n}_\Lambda$ are also NEs in $H_\Phi$ and $H_\Lambda$.
\label{lem:NEdecomp}
\end{lemma}
\begin{proof}
$n_\Lambda$ is apparently an NE in $H_\Lambda$ since no $C$-edges exist in $H_\Lambda$,
which implies that within each connected component in $H_\Lambda$,
all players play the same strategy.
And for $n_\Phi$, consider a vertex $u\in V(H_\Phi)$:
\begin{description}
\item[If $u$ is the endpoint of an $C$-edge:]
Since all of its neighbors in $G$ are in $H_\Phi$ by definition,
it must still satisfy the NE condition, since all players in $G$ satisfy the NE condition.
\item[If $u$ is not the endpoint of any $C$-edges:]
All of its neighbors play the same strategy as $u$ does,
which gives $u$ no incentive to deviate.
\end{description}
$\qed$
\end{proof}
\begin{theorem}
For any graph $G$, any Nash Equilibrium ${\bf n}=(n^{e_a}, n^{e_b}, n^{e_c})$,
and a decomposition $\Psi=(H_\Phi,H_\Lambda)$
\begin{equation*}
r(\mathbf{n}) \leq \max(r(\mathbf{n}_\Phi), r(\mathbf{n}_\Lambda))
\end{equation*}
\label{thm:decomp}
\end{theorem}
\begin{proof}
\begin{align*}
r(\mathbf{n}) &= \frac{2(n^{e_a}+n^{e_b}+n^{e_c})\alpha}{2n^{e_a}\alpha+2 n^{e_b}\beta + 2n^{e_c}\gamma}\\
&= \frac{2(n_\Phi^{e_a}+n_\Phi^{e_b}+n_\Phi^{e_c}+n_\Lambda^{e_a}+n_\Lambda^{e_b}+n_\Lambda^{e_c})\alpha}{2(n_\Phi^{e_a}+n_\Lambda^{e_a})\alpha+2(n_\Phi^{e_b}+n_\Lambda^{e_b})\beta+2(n_\Phi^{e_c}+n_\Lambda^{e_c})\gamma}\\
&= \frac{2(n_\Phi^{e_a}+n_\Phi^{e_b}+n_\Phi^{e_c})\alpha+2(n_\Lambda^{e_a}+n_\Lambda^{e_b}+n_\Lambda^{e_c})\alpha}{(2n_\Phi^{e_a}\alpha+2 n_\Phi^{e_b}\beta + 2n_\Phi^{e_c}\gamma) + (2n_\Lambda^{e_a}\alpha+2 n_\Lambda^{e_b}\beta + 2n_\Lambda^{e_c}\gamma)}\\
&\leq \max(\frac{2(n_\Phi^{e_a}+n_\Phi^{e_b}+n_\Phi^{e_c})\alpha}{2n_\Phi^{e_a}\alpha+2 n_\Phi^{e_b}\beta + 2n_\Phi^{e_c}\gamma},\frac{2(n_\Lambda^{e_a}+n_\Lambda^{e_b}+n_\Lambda^{e_c})\alpha}{2n_\Lambda^{e_a}\alpha+2 n_\Lambda^{e_b}\beta + 2n_\Lambda^{e_c}\gamma})\\
&= \max(r(\mathbf{n}_\Phi), r(\mathbf{n}_\Lambda))
\end{align*}
$\qed$
\end{proof}
We begin with bounding $r(\mathbf{n}_\Lambda)$.
\begin{theorem}\label{thm:hlambda}
For any NE $\mathbf{n}$, the corresponding $\mathbf{n}_\Lambda$ under the decomposition satisfies:
$$r(\mathbf{n}_\Lambda) \leq \frac{\alpha}{\beta}$$
\end{theorem}
\begin{proof}
Assume $\mathbf{n}_\Lambda=(n_\Lambda^{e_a},n_\Lambda^{e_b},0)$,
\begin{equation}
r(\mathbf{n}_\Lambda) = \frac{2(n_\Lambda^{e_a}+n_\Lambda^{e_b})\alpha}{2n_\Lambda^{e_a}\alpha+2n_\Lambda^{e_b}\beta}\leq\frac{\alpha}{\beta}.\notag
\end{equation}
$\qed$
\end{proof}
Before we move on to the discussion of $H_\Phi$,
we first generalize $\mathbf{n},\mathbf{n}_i$ to non-negative rational numbers rather than natural numbers.
\begin{remark}
Since the quotient $r(\mathbf{n})$ remains constant while scaling the state $\mathbf{n}$ by a factor,
it suffices to attain a fractional solution $\mathbf{n}$ whose $n^{e_a}$, $n^{e_b}$ and $n^{e_c}$ are non-negative rational numbers in order to find the worst NE compared to the optimal social welfare.
\end{remark}
For instance,
if we conclude that the state $\mathbf{n}=(\frac{1}{2},\frac{1}{2},1)$ maximizes $r(\mathbf{n})$ among all (fractional) Nash equilibria,
we could scale $\mathbf{n}$ to a integral state $\mathbf{n}'=(1,1,2)$,
while remaining the same quotient
$r(\mathbf{n}') = \frac{4\alpha}{\alpha+\beta+2\gamma} = r(\mathbf{n})$.
Therefore, $\mathbf{n}'$ is also an worst NE.
\\
Now we consider an NE $\mathbf{n}_\Phi=(n_\Phi^{e_a}, n_\Phi^{e_b}, n_\Phi^{e_c})$ in $H_\Phi$
as well as the corresponding quotient $r(\mathbf{n}_\Phi)$.
\begin{remark}\label{rm:par}
\begin{eqnarray}\label{eq:deca}
\frac{\partial r(n_{\Phi}^{e_a},n_{\Phi}^{e_b},n_{\Phi}^{e_c})}{\partial n_{\Phi}^{e_a}}&=&\alpha\frac{n_\Phi^{e_b}(\beta-\alpha)+n_\Phi^{e_c}(\gamma-\alpha)}{(n_\Phi^{e_a}\alpha + n_\Phi^{e_b}\beta+n_\Phi^{e_c}\gamma)^2}\leq0\\
\label{eqn:prpy}
\frac{\partial r(n_{\Phi}^{e_a},n_{\Phi}^{e_b},n_{\Phi}^{e_c})}{\partial n_{\Phi}^{e_b}}&=&\alpha\frac{n_\Phi^{e_a}(\alpha-\beta)+n_\Phi^{e_c}(\gamma-\beta)}{(n_\Phi^{e_a}\alpha + n_\Phi^{e_b}\beta+n_\Phi^{e_c}\gamma)^2}\\
\label{eqn:prpz}
\frac{\partial r(n_{\Phi}^{e_a},n_{\Phi}^{e_b},n_{\Phi}^{e_c})}{\partial n_{\Phi}^{e_c}}&=&\alpha\frac{n_\Phi^{e_a}(\alpha-\gamma)+n_\Phi^{e_b}(\beta-\gamma)}{(n_\Phi^{e_a}\alpha + n_\Phi^{e_b}\beta+n_\Phi^{e_c}\gamma)^2}\geq 0
\end{eqnarray}
\end{remark}
\begin{definition}
For one $C$-edge, we call one of its endpoints using strategy $A$ as $A$-player and the other endpoint using strategy $B$ as $B$-player.
\end{definition}
We have the following properties:
\begin{lemma}\label{lem:phip}
\begin{eqnarray}
n_{\Phi}^{e_a}&\geq&\frac{\beta-\gamma}{2(\alpha-\gamma)}n_\Phi^{e_c}\notag\\
n_{\Phi}^{e_b}&\geq&\frac{\alpha-\gamma}{2(\beta-\gamma)}n_\Phi^{e_c}\notag
\end{eqnarray}
\end{lemma}
\begin{proof}
We provide a proof for $n_{\Phi}^{e_a}$,
which can be symmetrically applied to $n_{\Phi}^{e_b}$.
Denote the set of all $A$-players in $H_\Phi$ as $\mathscr{A}$.
Since $\mathbf{n}_\Phi$ is an NE,
\begin{align*}
\forall i \in \mathscr{A}: \ u_i(n_i^{e_a},0,n_i^{e_c})&\geq u_i(0,n_i^{e_c},n_i^{e_a})\\
\Rightarrow \quad\quad n_i^{e_a} \alpha + n_i^{e_c}\gamma &\geq n_i^{e_a}\gamma + n_i^{e_c}\beta\\
\Rightarrow \quad\quad\ \ n_i^{e_a} &\geq \frac{\beta-\gamma}{\alpha-\gamma}n_i^{e_c}
\end{align*}
\begin{align*}
n_\Phi^{e_a} \geq \frac{1}{2} \sum_{i\in\mathscr{A}} n_i^{e_a}
\geq \frac{\beta-\gamma}{2(\alpha-\gamma)}\sum_{i\in\mathscr{A}} n_i^{e_c}
= \frac{\beta-\gamma}{2(\alpha-\gamma)}n_\Phi^{e_c}
\end{align*}
The first inequality holds because
each $A$-edge can be shared by at most two $A$-players,
while the last equation holds because no $C$-edges can be shared between $A$-players.
$\qed$
\end{proof}
We then bound $r(\mathbf{n}_\Phi)$:
\begin{theorem}\label{thm:hphi}
For any NE $\mathbf{n}$, the corresponding $\mathbf{n}_\Phi$ under the decomposition satisfies:
\begin{equation}
r(\mathbf{n}_\Phi)\leq \frac{\alpha(\alpha+\beta-2\gamma)}{\alpha\beta-\gamma^2}\notag
\end{equation}
\end{theorem}
\begin{proof}
From Remark~\ref{rm:par}, we know that $r(n_{\Phi}^{e_a},n_{\Phi}^{e_b},n_{\Phi}^{e_c})$ is monotone decreasing with $n_\Phi^{e_a}$.
Substituting $n_\Phi^{e_a}$ with its lower bound in Lemma~\ref{lem:phip}:
\begin{align*}
r(n_{\Phi}^{e_a},n_{\Phi}^{e_b},n_{\Phi}^{e_c}) \leq r(\frac{\beta-\gamma}{2(\alpha-\gamma)}n_\Phi^{e_c}, n_\Phi^{e_b}, n_\Phi^{e_c})
\end{align*}
From Equation~\ref{eqn:prpy} of Remark~\ref{rm:par},
\begin{align*}
\frac{\partial r(\frac{\beta-\gamma}{2(\alpha-\gamma)}n_\Phi^{e_c},n_{\Phi}^{e_b},n_{\Phi}^{e_c})}{\partial n_{\Phi}^{e_b}}
&=\alpha\frac{(\frac{\beta-\gamma}{2(\alpha-\gamma)}n_\Phi^{e_c})(\alpha-\beta)+n_\Phi^{e_c}(\gamma-\beta)}{(n_\Phi^{e_a}\alpha + n_\Phi^{e_b}\beta+n_\Phi^{e_c}\gamma)^2}\\
&=\frac{\alpha}{Z^2}\cdot \frac{(\beta-\gamma)(2\gamma-\alpha-\beta)}{2(\alpha-\gamma)} \leq 0
\end{align*}
where $Z^2$ is a positive normalizing factor.
Therefore, $r(\frac{\beta-\gamma}{2(\alpha-\gamma)}n_\Phi^{e_c},n_{\Phi}^{e_b},n_{\Phi}^{e_c})$
is monotone decreasing with respect to $n_\Phi^{e_b}$.
Consequently,
\begin{align*}
r(n_{\Phi}^{e_a},n_{\Phi}^{e_b},n_{\Phi}^{e_c})
&\leq r(\frac{\beta-\gamma}{2(\alpha-\gamma)}n_\Phi^{e_c}, n_\Phi^{e_b}, n_\Phi^{e_c})\\
&\leq r(\frac{\beta-\gamma}{2(\alpha-\gamma)}n_\Phi^{e_c}, \frac{\alpha-\gamma}{2(\beta-\gamma)}n_\Phi^{e_c}, n_\Phi^{e_c})\\
&= \frac{\alpha(\frac{\beta-\gamma}{2(\alpha-\gamma)}n_\Phi^{e_c}+\frac{\alpha-\gamma}{2(\beta-\gamma)}n_\Phi^{e_c}+n_\Phi^{e_c})}{\frac{\beta-\gamma}{2(\alpha-\gamma)}\alpha n_\Phi^{e_c}+\frac{\alpha-\gamma}{2(\beta-\gamma)} \beta n_\Phi^{e_c}+ \gamma n_\Phi^{e_c}}\\
&= \frac{\alpha(\alpha+\beta-2\gamma)}{\alpha\beta-\gamma^2}
\end{align*}
$\qed$
\end{proof}
Our Main Theorem follows combining Theorem.~\ref{thm:decomp}, Theorem.~\ref{thm:hlambda} and Theorem.~\ref{thm:hphi}.
\begin{theorem}[Main Theorem]
For any given $\alpha$, $\beta$, and $\gamma$:
\begin{equation}
PoA = \max_{\mathbf{n}\in \mathbb{N}^e} (r(\mathbf{n})) \leq \frac{\alpha(\alpha+\beta-2\gamma)}{\alpha\beta-\gamma^2} \notag
\end{equation}
\label{thm:mainthm}
\end{theorem}
To provide more insights into how bad an NE can be compared to the optimal social welfare,
we now present some discussions on the upper bound given in Theorem~\ref{thm:mainthm}.
\begin{remark}
$\frac{\alpha(\alpha+\beta-2\gamma)}{\alpha\beta-\gamma^2}$
is monotone decreasing with respect to $\gamma$.
\label{rem:dec-gamma}
\end{remark}
\begin{proof}
Denote $\frac{\alpha(\alpha+\beta-2\gamma)}{\alpha\beta-\gamma^2}$ as $p(\alpha,\beta,\gamma)$.
\begin{align*}
\frac{\partial p(\alpha,\beta,\gamma)}{\partial \gamma} = -\frac{2\alpha(\alpha-\gamma)(\beta-\gamma)}{(\alpha\beta-\gamma^2)^2} \leq 0
\end{align*}
\qed
\end{proof}
Remark~\ref{rem:dec-gamma} matches the intuition that Nash equilibria get better as more compatibility is introduced.
As the compatibility between the innovation and the status quo increases,
the switching barrier preventing the users adopting the innovation diminishes,
which encourages more users to adopt the innovation,
hence improving the social welfare at an NE.
Extremely, when $\gamma=\beta$, which means the innovation provides perfect compatibility with the status quo, the upper bound in Theorem~\ref{thm:mainthm} reduces to $\frac{\alpha}{\beta}$.
In this case, there are no $C$-edges, and the worst NE is the one in which all users taking $B$ strategy (the status quo),
yielding a Price of Anarchy of exactly $\frac{\alpha}{\beta}$.
On the other hand, the case where no compatibility exists ($\gamma=0$),
which is a commonly studied case in many previous works,
has the worst Price of Anarchy,
as stated in Corollary~\ref{cor:worstPoA}.
\begin{corollary}
For any given $\alpha$ and $\beta$:
\begin{equation*}
PoA \leq \frac{\alpha}{\beta} + 1
\end{equation*}
regardless of the value of $\gamma$.
\label{cor:worstPoA}
\end{corollary}
\begin{proof}
From Remark~\ref{rem:dec-gamma},
for any fixed $\alpha$ and $\beta$,
the worst PoA is achieved when $\gamma=0$.
This corollary follows when substituting $\gamma$ with $0$ in Theorem~\ref{thm:mainthm}.
\qed
\end{proof}
\section{Tightness of the PoA Upper Bound}
\begin{proposition}
The upper bound of PoA given in Theorem~\ref{thm:mainthm} is tight for any given $\alpha$,$\beta$ and $\gamma$.
\label{pro:tightness}
\end{proposition}
From the proof in Section~\ref{sec:poabound},
it can be deduced that the PoA upper bound can be achieved when $H_\Lambda$ is empty
and
\begin{align*}
n_{\Phi}^{e_a}&= \frac{\beta-\gamma}{2(\alpha-\gamma)}n_\Phi^{e_c}\\
n_{\Phi}^{e_b}&= \frac{\alpha-\gamma}{2(\beta-\gamma)}n_\Phi^{e_c}
\end{align*}
The intuition of yielding such a Nash equilibrium is to ensure each $A$-edge is shared by two $A$-players and symmetrically each $B$-edge is shared by two $B$-players,
which could minimize the number of $A$-edges and $B$-edges for a given number of $C$-edges,
resulting a worst NE.
There is, however, a caveat when constructing the worst NE for a given tuple of ($\alpha,\beta,\gamma$)
that the lower bound in Lemma~\ref{lem:phip} is not always achievable for any graph,
which is elaborated in Lemma~\ref{lem:phip2}, a stronger version of Lemma~\ref{lem:phip}.
\begin{lemma}
\begin{eqnarray}
n_{\Phi}^{e_a}&\geq&\max(\frac{\beta-\gamma}{\alpha-\gamma}n_\Phi^{e_c} - \frac{1}{2}n_\Phi^{e_c}(n_\Phi^{e_c}-1), \frac{\beta-\gamma}{2(\alpha-\gamma)}n_\Phi^{e_c})\notag\\
n_{\Phi}^{e_b}&\geq&\max(\frac{\alpha-\gamma}{\beta-\gamma}n_\Phi^{e_c}-\frac{1}{2}n_\Phi^{e_c}(n_\Phi^{e_c}-1),\frac{\alpha-\gamma}{2(\beta-\gamma)}n_\Phi^{e_c})\notag
\end{eqnarray}
\label{lem:phip2}
\end{lemma}
\begin{proof}
We only provide a proof for $n_{\Phi}^{e_a} \geq \frac{\beta-\gamma}{(\alpha-\gamma)}n_\Phi^{e_c} - \frac{1}{2}n_\Phi^{e_c}(n_\Phi^{e_c}-1)$,
which can be symmetrically applied to $n_{\Phi}^{e_b}$.
And the other part has been proved in Lemma~\ref{lem:phip}.
Similar to Lemma~\ref{lem:phip}, denote the set of all $A$-players in $H_\Phi$ as $\mathscr{A}$.
From Lemma~\ref{lem:phip}:
\begin{align*}
\forall i \in \mathscr{A}: n_i^{e_a} &\geq \frac{\beta-\gamma}{\alpha-\gamma}n_i^{e_c}
\end{align*}
Since each $C$-edge has one $A$-player,
there are at most $n_\Phi^{e_c}$ $A$-players.
And since there is at most one edge between each pair of players:
\begin{equation*}
n_\Phi^{e_a} \geq \sum_{i\in \mathscr{A}} n_i^{e_a} - \frac{n_\Phi^{e_c}(n_\Phi^{e_c}-1)}{2}
\geq\frac{\beta-\gamma}{\alpha-\gamma}n_\Phi^{e_c}-\frac{n_\Phi^{e_c}(n_\Phi^{e_c}-1)}{2}
\end{equation*}
\qed
\end{proof}
Therefore, in order to construct an NE where
$n_{\Phi}^{e_a}= \frac{\beta-\gamma}{2(\alpha-\gamma)}n_\Phi^{e_c}$
and $n_{\Phi}^{e_b}= \frac{\alpha-\gamma}{2(\beta-\gamma)}n_\Phi^{e_c}$,
it is necessary that
\begin{align*}
\frac{\beta-\gamma}{\alpha-\gamma}n_\Phi^{e_c} - \frac{1}{2}n_\Phi^{e_c}(n_\Phi^{e_c}-1) &\leq \frac{\beta-\gamma}{2(\alpha-\gamma)}n_\Phi^{e_c}
\end{align*}
and
\begin{align*}
\frac{\alpha-\gamma}{\beta-\gamma}n_\Phi^{e_c}-\frac{1}{2}n_\Phi^{e_c}(n_\Phi^{e_c}-1) &\leq \frac{\alpha-\gamma}{2(\beta-\gamma)}n_\Phi^{e_c}
\end{align*}
Solving these two inequalities:
\begin{equation}
n_\Phi^{e_c} \geq \frac{\alpha-\gamma}{\beta-\gamma} + 1
\end{equation}
The intuition for Lemma~\ref{lem:phip2} is that to share a given number of $A$-edges or $B$-edges,
enough $A$-players or $B$-players are needed.
For instance, if there are $2$ $B$-players in total, then at most $1$ $B$-edge can be shared between them.
To achieve an NE, a specific number of $B$-edges are needed.
It is hence necessary to have enough $C$-edges to produce enough $B$-players to ensure that each of these $B$-edges can be shared by two $B$-players.
Now a Nash equilibrium $\mathbf{n}=(n^{e_a},n^{e_b},n^{e_c})$ yielding the PoA bound shown in Theorem~\ref{thm:mainthm} can be constructed as follows:
\begin{itemize}
\item First construct a fractional NE $\mathbf{n}$
\begin{itemize}
\item Construct $n^{e_c} = \frac{\alpha-\gamma}{\beta-\gamma}+1$ $C$-edges.
\item For each $A$-player, create $\frac{\beta-\gamma}{\alpha-\gamma}$ half-$A$-edges, and $\frac{\alpha-\gamma}{\beta-\gamma}$ half-$B$-edges.
\item Group those half-edges into pairs,
yielding $n^{e_a} = \frac{\beta-\gamma}{2(\alpha-\gamma)}n^{e_c}$ $A$-edges
and $n^{e_b} =\frac{\alpha-\gamma}{2(\beta-\gamma)}n^{e_c}$ $B$-edges.
\end{itemize}
\item Then scale $\mathbf{n}$ to an integral solution
\end{itemize}
From the argument in Lemma~\ref{lem:phip} and Lemma~\ref{lem:phip2},
the resulting state
$$\mathbf{n}=(\frac{\beta-\gamma}{2(\alpha-\gamma)}n^{e_c}, \frac{\alpha-\gamma}{2(\beta-\gamma)}n^{e_c}, n^{e_c})$$
is an NE, where $n^{e_c}=\frac{\alpha-\gamma}{\beta-\gamma}+1$.
Therefore,
\begin{align*}
\frac{SW(OPT)}{SW(\mathbf{n})} &= \frac{2\alpha(n^{e_a}+n^{e_b}+n^{e_c})}{2\alpha n^{e_a}+2\beta n^{e_b}+2\gamma n^{e_c}}\\
&= \frac{\alpha(\frac{\beta-\gamma}{2(\alpha-\gamma)}n^{e_c} + \frac{\alpha-\gamma}{2(\beta-\gamma)}n^{e_c} + n^{e_c})}{\alpha \frac{\beta-\gamma}{2(\alpha-\gamma)}n^{e_c} + \beta \frac{\alpha-\gamma}{2(\beta-\gamma)}n^{e_c} + \gamma n^{e_c}}\\
&= \frac{\alpha(\frac{\beta-\gamma}{2(\alpha-\gamma)} + \frac{\alpha-\gamma}{2(\beta-\gamma)} + 1)}{\alpha \frac{\beta-\gamma}{2(\alpha-\gamma)} + \beta \frac{\alpha-\gamma}{2(\beta-\gamma)} + \gamma }\\
&= \frac{\alpha(\alpha+\beta-2\gamma)}{\alpha\beta-\gamma^2}
\end{align*}
To better illustrate how this can be done,
we give two concrete examples in the case of $(\alpha=1,\beta=1,\gamma=0)$ and $(\alpha=3,\beta=2,\gamma=1)$.
For $(\alpha=1,\beta=1,\gamma=0)$,
PoA has an upper bound of $\frac{1(1+1-0)}{1\cdot1-0} = 2$.
We can construct a fractional NE $\mathbf{n}=(\frac{1}{2},\frac{1}{2},1)$ following the aforementioned mechanism.
When scaling up, we get $\mathbf{n}=(1,1,2)$.
Figure~\ref{fig:poa-k1} shows such an NE in which the social welfare
$SW = \alpha+\beta+2\gamma = 2$,
while the optimal SW is $OPT = (1+1+2)\alpha = 4 = 2 SW$.
\begin{figure}[h]
\centering
\includegraphics[width=0.25\textwidth]{./materials/PoA-k1.eps}
\caption{An NE with $SW=\frac{1}{2}OPT$ for $(\alpha=1,\beta=1,\gamma=0)$}
\label{fig:poa-k1}
\end{figure}
For $(\alpha=3,\beta=2,\gamma=1)$ with $PoA\leq \frac{3(3+2-2)}{3\cdot2-1} = \frac{9}{5}$,
$n^{e_c}=\frac{\alpha-\gamma}{\beta-\gamma}+1 = 3$.
Then the constructed fractional NE $\mathbf{n}=(\frac{3}{4},3,3)$.
We get $\mathbf{n}=(1,4,4)$.
And a graph depicting such a state is shown in Figure~\ref{fig:poa-k2}.
The state shown is a NE with social welfare $SW=2\alpha+8\beta+8\gamma=30$.
And the optimal SW is $OPT=18\alpha=54= \frac{9}{5} SW$.
\begin{figure}[h]
\centering
\includegraphics[width=0.4\textwidth]{./materials/PoA-k2.eps}
\caption{An NE with $SW=\frac{5}{9}OPT$ for $(\alpha=3,\beta=2,\gamma=1)$}
\label{fig:poa-k2}
\end{figure}
\section{Introduction}
There have been intensive studies on social networks recently that range over many different facets of a social network,
including microscopic and macroscopic structures~\cite{backstrom2014romantic,ugander2013subgraph}, evolution and dynamics~\cite{leskovec2008microscopic,kempe2013selection}, among others.
Particularly, it attracts a lot of interest to study how information or influence spreads in a social network.
The diffusion of information or behavior within a network is ubiquitous and has profound significance to study.
The procedure of how a new technology emerges and prevails,
the adoption of a political stance among the population,
the dissemination of a new social convention,
all of these processes,
ranging from how smartphones prevailed in such an astounding speed,
to how the convention of \textit{RT} was adopted across Twitter,
can be modeled as a cascading behavior in a social network.
Back in 1970s,
Granovetter~\cite{granovetter1978threshold} started to formulate mathematical models for the spreading process of collective behavior.
Together with a lineage of later works,
they aimed at predicting the success or failure of the diffusion.
Granovetter proposed a threshold-based model,
which has been adopted and extended since then.
The key idea is that there is a global threshold indicating the necessary proportion of neighbors of a given person adopting the innovation in order to convince that given person to conform.
One can thus investigate problems such as how should we select initial nodes of the new behavior to maximize its diffusion~\cite{kempe2003maximizing,kempe2005influential}.
Some more relevant work tackles this problem from a game-theoretic perspective,
in which each player has a set of actions to choose from and her utility depends on the interaction between her and her neighbors.
Ellison~\cite{ellison1993learning} studied the simple case where all players form a chain and each player has two actions (the innovation and the status quo).
He showed that the threshold $\frac{1}{2}$ suffice to ensure the new action will spread to the entire graph.
Some later work extended the research to lattices~\cite{blume1995statistical}.
Morris~\cite{morris2000contagion} discussed the case of general graphs.
He focused mainly on the global threshold as well, such as for which thresholds would there be a Nash equilibrium where both actions are simultaneously played.
These game theoretic models could result in a similar threshold based contagion,
but adopted a different perspective and focused on different aspects.
For instance, a threshold-based model is often ready to provide an algorithm to select proper nodes to start the innovation in order to maximize the diffusion,
while a game-theoretic model studies how individuals interact which does not necessarily have this feature.
\\
Recently there is another lineage of work~\cite{goyal2012competitive,he2013price} that studies the Price of Anarchy of the competitive cascade games.
A major distinction is that they take a different perspective in terms of game setting,
in which competition is the central concern.
For instance, Apple and Samsung are marketing their smartphones to the users.
In this competitive setting, the players are no longer the users in the graph,
but are the companies aiming at maximizing the adoption of their products.
In our example, players are Apple and Samsung,
and their strategies are the selection of the initial adoption of their products (seeds).
Under this setting, Goyal and Kearns~\cite{goyal2012competitive} showed that the two-player competitive cascade game
has a PoA upper-bounded by $4$.
And in~\cite{he2013price}, a tighter and more general bound is proposed
that the PoA has an upper-bound of $2$ for an arbitrary number of players.
\\
Different from this competitive setting,
our work will stick to the traditional game settings where users are modeled as players,
in order to study how users react to the diffusion to an innovation spread through the network.
We mainly focus on the stable states of the network, Nash Equilibria in the game, rather than the result or the speed of diffusion in the network ~\cite{kempe2003maximizing,kreindler2013rapid} to better understand the relation between Nash equilibria and the optimal social welfare.
We now present our main results in a networked coordination game to model the influences between behaviors of individuals in social networks and how bad a Nash Equilibrium (NE) can get in these games.
Some of the questions that can be answered in this paper include:
\begin{itemize}
\item Is there an NE worse than the one that all players take the status quo?
\item If there is, what is the worst NE and how bad can it be?
\end{itemize}
Specifically, we show a \emph{tight upper bound} of the Price of Anarchy (PoA) of the game.
It is challenging because there could be numerous NEs in this game due to the heterogeneity of the network topologies.
We show that there can be an NE worse than the one all taking the status quo,
but only by a small margin.
Furthermore, we study the effect of introducing compatibility between the two strategies,
and conclude that as the compatibility increases,
the upper bound of PoA gets lower,
which matches the intuition that the compatibility will diminish the transferring barrier and make it easier to switch to the innovation,
thus resulting in lower PoA.
\input{poa}
\section{Conclusion}
In this paper,
we provide a tight upper bound for the Price of Anarchy of the cascading behavior in social networks,
which has long been used to model a variety of social behaviors such as the diffusion of innovation,
the adoption of novel conduct, etc.
We discuss the cascading behavior in a networked coordination game setting
in which nodes are individuals in the social network and edges denote relationship that influence can be exerted over.
Two behaviors (the innovation A and the status quo B) exist in the game,
and if two adjacent individuals adopt the same behavior,
they will both receive the corresponding payoff depending on which behavior they adopt,
where the payoff of A is inherently higher than (or equal to) B.
However, if they do not play coordinately,
they will receive a less payoff which can be interpreted as the compatibility between the two behaviors.
This game is known to have numerous Nash equilibria depending on the payoffs as well as on the network topology.
Previous work seldom quantitatively address how bad a Nash equilibrium can be.
To the best of our knowledge,
even the question that whether the PoA is bounded for a given setting of payoffs remained elusive until this paper.
In this work we showed that the Price of Anarchy can be slightly worse than the case where all players take B.
However, it is pretty close to the worst Nash equilibrium,
which even in the worst case that the compatibility between A and B is $0$,
still has a PoA upper bound of
$\frac{\alpha}{\beta}+1$,
where $\alpha$ and $\beta$ are the payoffs of A and B respectively.
In the future,
it would be desirable to see whether the PoA upper bound can be generalized to the game with heterogeneous payoffs instead of a universal payoff for all edges.
One may also consider a more sophisticated way of introducing compatibility into this model such as in~\cite{immorlica2007role} where an extra strategy to adopt both A and B with an additional cost is introduced.
It would be interesting to know whether and how the PoA upper bound can be adapted to those models.
\section{Acknowledgements}
We thank \'Eva Tardos and Jon Kleinberg for valuable comments and inspirational discussions on Algorithmic Game Theory and Social Cascading Behavior.
|
\section*{Introduction}
This paper is devoted to the analysis of second order mean field games systems with a local coupling. The general form of these systems is:
\begin{equation}\label{MFG}
\left\{\begin{array}{cl}
(i)&- \partial_t \phi -A_{ij}\partial_{ij}\phi+H(x,D\phi) =f(x,m(x,t))\\
(ii) & \partial_t m-\partial_{ij}( A_{ij}m)-{\rm div} (mD_pH(x,D\phi))=0\\
(iii)& m(0)=m_0, \; \phi(x,T)=\phi_T(x)
\end{array}\right.
\end{equation}
where $A:\mathbb R^d\to \mathbb R^{d\times d}$ is symmetric and nonnegative, the Hamiltonian $H:\mathbb R^d\times \mathbb R^d\to\mathbb R$ is convex in the second variable, the coupling
$f:\mathbb R^d\times [0,+\infty)\to [0,+\infty)$ is increasing with respect to the second variable, $m_0$ is a probability density and $\phi_T:\mathbb R^d\to\mathbb R$ is a given function. The functions $H$ and $f$, and the matrix $A$, could as well depend on time, but since this does not give any additional difficulty, we will avoid it just to simplify notations.
Mean field game systems (MFG systems) have been introduced simultaneously by Lasry-Lions \cite{LL06cr1, LL06cr2, LL07mf, LLperso} and Huang-Caines-Malham\'e \cite{HCMieeeAC06} to describe Nash equilibria in differential games with infinitely many players. The first unknown $\phi=\phi(t,x)$ is the value function of an optimal control problem of a typical small player. In this control problem, the dynamics is given by the controlled stochastic differential equation
$$
dX_s= v_s ds + \Sigma(X_s)dB_s,
$$
where $(v_s)$ is the control, $(B_s)$ is a Brownian motion and $\Sigma\Sigma^T=A$. The cost is given by
$$
\mathbb E\left[ \int_0^T H^*(X_s, -v_s)+ f(X_s, m(s,X_s))\ ds+ \phi_T(X_T)\right]
$$
For each time $t\in [0,T]$ the quantity $m(t,x)$ denotes the density of population of small players at position $x$. In the control problem the term involving $f$ formalizes the fact that the cost of the player depends on this density $m$. As $\phi$ is the value function of this control problem, the optimal control of a typical small player is formally given by the feedback $(t,x)\to -D_pH(x,D\phi(t,x))$. Hence
the second equation \eqref{MFG}-(ii) is the Kolmogorov equation of the process $(X_s)$ when the small player plays in an optimal way. By the mean field approach, this equation also describes the evolution of the whole population density as all players play in an optimal way.
MFG systems with uniformly parabolic diffusions---typically $A_{ij}\partial_{ij}\phi=\Delta \phi$---have been the object of several contributions, either by PDE methods (see, e.g., \cite{CLLP, LL06cr1, LL06cr2, LL07mf, LLperso, GPSM1, GPSM2, Po}) or by stochastic techniques (see, e.g., \cite{CaDe2, HCMieeeAC06}): in this setting one often expects the solutions to be smooth, at least if the coupling is nonlocal and regularizing or if it has a ``small growth". The case of local couplings with an arbitrary growth has been discussed in \cite{CLLP} for purely quadratic hamiltonians (i.e. $H=|D\phi|^2$), in which case solutions are proved to be smooth, and in \cite{Po} for general hamiltonians, by proving existence and uniqueness of weak solutions.
Here we concentrate on degenerate parabolic equations. In this case the usual fixed point techniques used to prove the existence of solutions in the uniformly parabolic setting break down by lack of regularity. One then has to rely on convex optimization methods: this idea, which goes back to the analysis of some optimal transport problems (see \cite{bb,CCN}), has already been used to study first order MFG systems (i.e., $A\equiv 0$): see \cite{c1, cg, Gra14}. However it was not clear in these papers wether the weak solution was stable with respect to viscous approximation, i.e., if we could obtain weak solutions of the first order MFG systems by passing to the limit in uniformly parabolic ones. This issue has partially motivated our study.
In this paper we show the existence and uniqueness of a weak solution for the degenerate mean field game system \eqref{MFG} as well as the stability of solutions with respect to perturbation of the data: this includes of course stability by viscous approximation.
Concerning existence and uniqueness of solutions, the paper improves the existing results in two directions. First we consider non uniformly parabolic second order MFG systems, which have never been considered before. The introduction of second order derivatives induces several issues: in particular, in contrast with the first order equations, we do not expect the function $\phi$ to be BV (as in \cite{cg,Gra14}), which obliges us to be very careful about trace properties. Secondly---and this is new even for first order MFG systems---we drop a restriction between the growth condition of $H$ and the growth condition of $f$, restriction which was mandatory in the previous papers: see \cite{c1, cg}. To overcome the difficulty, we provide new {\it integral estimates} for subsolutions of Hamilton-Jacobi equations with unbounded right-hand side (Theorems \ref{int-reg} and \ref{thm:estiHJ}). We think that these results are of independent interest.
With these estimates in hand, the structure of proof for the existence and uniqueness follows roughly the lines already developed in \cite{c1, cg, CCN, Gra14}: basically it amounts to show that the MFG system can be viewed as an optimality condition for two convex problems, the first one being an optimal control of Hamilton-Jacobi equation, the second one an optimal control problem for the Fokker-Planck equation (see section \ref{sec:2opti} for details). A byproduct of this approach is the stability of weak solutions with respect to the data (Theorem \ref{thm:stab}), which can be obtained by $\Gamma-$convergence techniques.
The paper is organized as follows. First we introduce the notation and assumptions needed throughout the paper (section \ref{sec:notass}). Then (section \ref{sec:esti}) we give our new estimates for subsolutions of Hamilton-Jacobi equations with a superlinear growth in the gradient variable and an unbounded right-hand side. In section \ref{sec:2opti}, we introduce the two optimal control problems and show that they are in duality while in section \ref{sec:optiHJ} we show that the optimal control problem for the Hamilton-Jacobi equation has a ``relaxed solution." Section \ref{sec:exuniq} is devoted to the analysis of the MFG system (existence, uniqueness and characterization). In the last section we discuss the stability of solutions.
\bigskip
{\bf Acknowledgement: }
This work has been partially supported by the Commission of the
European Communities under the 7-th Framework Programme Marie
Curie Initial Training Networks Project SADCO,
FP7-PEOPLE-2010-ITN, No 264735, by the French National Research Agency
ANR-10-BLAN 0112 and ANR-12-BS01-0008-01 and by the Italian Indam Gnampa project 2013 \lq\lq Modelli di campo medio nelle dinamiche di popolazioni e giochi differenziali\rq\rq.
\section{Notations and assumptions}\label{sec:notass}
\noindent {\bf Notations :} We denote by
$\lg x, y\rg$ the Euclidean scalar product of two vectors $x,y\in\mathbb R^d$ and by
$|x|$ the Euclidean norm of $x$. We use conventions on repeated indices: for instance, if $a,b\in \mathbb R^d$, we often write $a_ib_i$ for the scalar product $\lg a,b\rg$. More generally, if $A$ and $B$ are two square symmetric matrices of size $d\times d$, we write $A_{ij}B_{ij}$ for ${\rm Tr}(AB)$.
To avoid further difficulties arising from boundary issues, we work in the flat $d-$dimen\-sional torus $\mathbb{T}^d=\mathbb R^d\backslash \mathbb Z^d$. We denote by $P(\mathbb{T}^d)$ the set of Borel probability measures over $\mathbb{T}^d$. It is endowed with the weak convergence. For $k,n\in\mathbb N$ and $T>0$, we denote by ${\mathcal C}^k([0,T]\times \mathbb{T}^d, \mathbb R^n)$ the space of maps $\phi=\phi(t,x)$ of class ${\mathcal C}^k$ in time and space with values in $\mathbb R^n$. For $p\in [1,\infty]$ and $T>0$, we denote by $L^p(\mathbb{T}^d)$ and $L^p((0,T)\times \mathbb{T}^d)$ the set of $p-$integrable maps over $\mathbb{T}^d$ and $[0,T]\times \mathbb{T}^d$ respectively. We often abbreviate $L^p(\mathbb{T}^d)$ and $L^p((0,T)\times \mathbb{T}^d)$ into $L^p$. We denote by $\|f\|_p$ the $L^p-$norm of a map $f\in L^p$. \\
\noindent {\bf Assumptions:} We now collect the assumptions on the coupling $f$, the Hamiltonian $H$ and the initial and terminal conditions $m_0$ and $\phi_T$. These conditions are supposed to hold throughout the paper.
\begin{itemize}
\item[(H1)] (Condition on the coupling) the coupling $f:\mathbb{T}^d\times [0,+\infty)\to \mathbb R$ is continuous in both variables, increasing with respect to the second variable $m$,
and there exist $q>1$ and $C_1$ such that
\begin{equation}\label{Hypf}
\frac{1}{C_1}|m|^{q-1}-C_1\leq f(x,m) \leq C_1 |m|^{q-1}+C_1 \qquad \forall m\geq 0 \;.
\end{equation}
Moreover we ask the following normalization condition to hold:
\begin{equation}\label{Hypf(0,m)=0}
f(x,0)=0 \qquad \forall x\in \mathbb{T}^d\;.
\end{equation}
We denote by $p$ the conjugate of $q$: $1/p+1/q=1$.
\item[(H2)] (Conditions on the Hamiltonian) The Hamiltonian $H:\mathbb{T}^d\times \mathbb R^d\to\mathbb R$ is continuous in both variables, convex and differentiable in the second variable, with $D_pH$ continuous in both variables, and has a superlinear growth in the gradient variable: there exist $r>1$ and $C_2>0$ such that
\begin{equation}\label{HypGrowthH}
\frac{1}{rC_2} |\xi|^{r} -C_2\leq H(x,\xi) \leq \frac{C_2}{r}|\xi|^r+C_2\qquad \forall (x,\xi)\in \mathbb{T}^d\times \mathbb R^d\;.
\end{equation}
We note for later use that the Fenchel conjugate $H^*$ of $H$ with respect to the second variable is continuous and satisfies similar inequalities
\begin{equation}\label{HypHstar}
\frac{1}{r'C_2} |\xi|^{r'} -C_2\leq H^*(x,\xi) \leq \frac{C_2}{r'}|\xi|^{r'}+C_2\qquad \forall (x,\xi)\in \mathbb{T}^d\times \mathbb R^d\;,
\end{equation}
where $r'$ is the conjugate of $r$: $\displaystyle\frac{1}{r}+\frac{1}{r'}=1$.
\item[(H3)] (Conditions on $A$) there exists a Lipschitz continuous map $\Sigma:\mathbb{T}^d\to \mathbb R^{d\times D}$ such that $\Sigma\Sigma^T=A$ : let $C_3$ be a constant such that
\begin{equation}\label{HypA}
|\Sigma(x)-\Sigma(y)|\leq C_3 |x-y| \qquad \forall x,y\in \mathbb{T}^d,
\end{equation}
Moreover we suppose that
\begin{equation}\label{CondCroiss}
{\rm either }\; r\geq p\qquad {\rm or}\qquad A\equiv 0.
\end{equation}
We recall that $p$ is the conjugate of $q$.
\item[(H4)] (Conditions on the initial and terminal conditions) $\phi_T:\mathbb{T}^d\to \mathbb R$ is of class ${\mathcal C}^2$, while $m_0:\mathbb{T}^d\to \mathbb R$ is a $C^1$ positive density (namely $m_0>0$ and $\displaystyle \int_{\mathbb{T}^d} m_0dx=1$).
\end{itemize}
Condition \eqref{Hypf(0,m)=0} is just a normalization condition, which we may assume without loss of generality.
Indeed, if all the conditions (H1)$\dots$(H4) but \eqref{Hypf(0,m)=0} hold, then one just needs to replace $f(x,m)$ by $f(x,m)-f(x,0)$ and $H(x,p)$ by $H(x,p)-f(x,0)$: the new $H$ and $f$ still satisfy the above conditions (H1)$\dots$(H4) with
\eqref{Hypf(0,m)=0}.
Let us set
$$\displaystyle F(x,m)=
\left\{\begin{array}{ll}
\displaystyle \int_0^m f(x,\tau)d\tau & {\rm if }\; m\geq 0\\
+\infty & {\rm otherwise}
\end{array}\right.
$$
Then $F$ is continuous on $\mathbb{T}^d\times (0,+\infty)$, differentiable and strictly convex in $m$ and satisfies
\begin{equation}\label{HypGrowthF}
\frac{1}{qC_1}|m|^{q}-C_1\leq F(x,m) \leq \frac{C_1}{q}|m|^{q}+C_1 \qquad \forall m\geq 0
\end{equation}
(changing the constant $C_1$ if necessary).
Let $F^*$ be the Fenchel conjugate of $F$ with respect to the second variable. Note that $F^*(x,a)=0$ for $a\leq 0$ because $F(x,m)$ is nonnegative and equal to $+\infty$ for $m<0$. Moreover,
\begin{equation}\label{HypGrowthFstar}
\frac{1}{pC_1}|a|^{p}-C_1\leq F^*(x,a) \leq \frac{C_1}{p}|a|^{p}+C_1 \qquad \forall a\geq 0\;.
\end{equation}
\section{Basic estimates on solutions of Hamilton-Jacobi equations}\label{sec:estimates}\label{sec:esti}
In this section we prove estimates in Lebesgue spaces for subsolutions of Hamilton-Jacobi equations of the form
\begin{equation}\label{eq:HJ}
\left\{\begin{array}{cl}
(i)&- \partial_t \phi -A_{ij}(x)\partial_{ij}\phi+H(x,D\phi) \leq \alpha(t,x)\\
(ii)& \phi(x,T)\leq\phi_T(x)
\end{array}\right.
\end{equation}
in terms of Lebesgue norms of $\alpha$ and $\phi_T$. We assume that \eqref{HypGrowthH} and \eqref{HypA} hold, and \eqref{eq:HJ} is understood in the sense of distributions. This means that $D\phi\in L^r$ and, for any nonnegative test function $\zeta\in C^\infty_c((0,T]\times \mathbb{T}^d)$,
$$
-\int_{\mathbb{T}^d}\zeta(T)\phi_T+ \int_{0}^{T}\int_{\mathbb{T}^d} \phi\partial_t\zeta+ \lg D\zeta, AD\phi\rg + \zeta(\partial_iA_{ij}\partial_j\phi+ H(x,D\phi)) \leq \int_{0}^{T}\int_{\mathbb{T}^d} \alpha \zeta.
$$
The estimates will be a consequence of the divergence structure of second order terms.
\begin{Theorem}\label{int-reg}
Assume that $\phi\in L^r((0,T);W^{1,r}(\mathbb{T}^d))$ is a nonnegative function satisfying, in distributional sense,
\begin{equation}\label{ineq}
\left\{\begin{array}{cl}
(i)&- \partial_t \phi - \partial_i \left( A_{ij}(x)\partial_{j}\phi\right)+ c_0\, |D\phi|^r \leq \alpha(t,x)\\
(ii)& \phi(x,T)\leq\phi_T(x)
\end{array}\right.
\end{equation}
for some nonnegative, bounded Lipschitz matrix $A_{ij}$, and some $r>1$, $c_0>0$, $\alpha \in \pelle p$ and $\phi_T\in \elle \infty$.
Then, there exists a constant $C=C(p,d,r,c_0, T, \|\alpha\|_{L^p((0,T)\times \mathbb{T}^d)}, \|\phi_T\|_{\elle \eta})$ such that
$$
\|\phi\|_{L^\infty((0,T), L^{\eta}(\mathbb{T}^d))}+ \|\phi\|_{L^\gamma((0,T)\times \mathbb{T}^d)}\leq C
$$
where $\eta=\frac{d( r(p-1)+1)}{d-r(p-1)}$ and $\gamma = \frac{rp(1+d)}{d-r(p-1)}$ if $p<1+\frac dr$ and $\eta=\gamma=+\infty$ if $p>1+\frac dr$.
\end{Theorem}
We note for later use that $\gamma>r$.
\begin{proof}
Up to a rescaling, we may assume that $c_0=1$. We first claim that, for any real function $g\in W^{1,\infty}(\mathbb R)$ which is nondecreasing, and nonnegative in $\mathbb R_+$, we have
\begin{equation}\label{chain}
\int_{\mathbb{T}^d} G(\phi(\tau)) \, dx + \int_\tau^T \int_{\mathbb{T}^d} |D\phi |^r\, g(\phi)\,dxdt \leq \int_\tau^T \int_{\mathbb{T}^d} \alpha\, g(\phi)\, dxdt+ \int_{\mathbb{T}^d} G(\phi_T)\, dx
\end{equation}
for a.e. $\tau\in (0,T)$, where $G(r)=\int_0^r g(s)\,ds$.
There are several possible ways to justify \eqref{chain}, one is to use regularization.
We first extend $\phi$ to $(0,T+1]\times \mathbb{T}^d$ by defining $\phi= \phi_T$ on $[T,T+1]$. Then it still holds in the sense of distributions
\begin{equation}\label{extineq}
- \partial_t \phi - \partial_i \left( \tilde A_{ij}(t,x)\partial_{j}\phi\right)+ |D\phi|^r\,\chi_{(0,T)} \leq \tilde \alpha(t,x)
\end{equation}
where $\tilde A_{ij}(t,x)= A_{ij}(x)\chi_{(0,T)}(t)$ and $\tilde \alpha(t,x)= \alpha (t,x)\chi_{(0,T)}(t)$.
Let $\xi$ be a standard convolution kernel in $(t,x)$ defined on $\mathbb R^{d+1}$ and $\xi^\epsilon(t,x)=\xi((t,x)/\epsilon)/(\epsilon)^{d+1}$, $\xi^\epsilon\geq0,\ \int_{\mathbb R^{d+1}}\xi^\epsilon(t,x) dtdx=1$, for all $\epsilon>0$. Let $\phi_\epsilon= \xi^\epsilon\star \phi$ and $\alpha_\epsilon=\xi^\epsilon\star \tilde \alpha$. Then $\phi_\epsilon, \alpha_\epsilon$ are $C^\infty$ and converge to $\phi,\alpha$ in their respective Lebesgue spaces.
Convolving $\xi^\epsilon$ with (\ref{extineq}) we obtain on $(0,T+1)\times \mathbb{T}^d$:
$$
- \partial_t \phi_\epsilon - \partial_i \left( \tilde A_{ij}(t,x)\partial_{j}\phi_\epsilon\right)+ |\xi_\epsilon\star(D\tilde \phi)|^r \leq \alpha_\epsilon(t,x)+R_\epsilon
$$
where $D\tilde \phi= D\phi\chi_{(0,T)}$ and we used the fact that $D\phi\mapsto |D\phi|^r$ is convex, and where $R_\epsilon= -\partial_i \left( \tilde A_{ij}(t,x)\partial_{j}\phi_\epsilon\right)+\xi_\epsilon\star \partial_i \left( \tilde A_{ij}(t,x)\partial_{j}\phi\right)$.
Using the notation (cf. \cite{DL89})
$$
[\xi^\epsilon,c](f) := \xi^\epsilon \star (cf) - c(\xi^\epsilon \star f)
$$
we can rewrite $R_\epsilon$ as $R_\epsilon= [\xi_\epsilon, \partial_i \tilde A_{ij}](\partial_j \phi)+[\xi_\epsilon, \tilde A_{ij}\partial _i](\partial_j \phi)$.
Invoking \cite[Lemma II.1]{DL89}, we have that $R_\epsilon \to 0$ in $L^r$, since $D\phi \in L^r$ and $ A_{ij} $ is Lipschitz.
Multiplying by $g(\phi_\epsilon)$ and integrating over $[\tau,T+\varepsilon]\times \mathbb{T}^d$, for $\tau\in (0,T)$, it follows
\begin{align*} &\int_{\mathbb{T}^d}G(\phi_\epsilon(\tau))dx-\int_{\mathbb{T}^d}G(\phi_\epsilon(T+\varepsilon))dx +\int_\tau^T\int_{\mathbb{T}^d} A_{ij}(x)\partial_{j}\phi_\epsilon g'(\phi_\epsilon)\partial _i \phi_\epsilon dxdt\\ & + \int_\tau^{T+\varepsilon}\int_{\mathbb{T}^d} |\xi_\epsilon\star(D\tilde \phi)|^r g(\phi_\epsilon)dx dt \leq \int_\tau^{T+\varepsilon}\int_{\mathbb{T}^d} g(\phi_\epsilon)\alpha_\epsilon(t,x) dx dt +\int_\tau^{T+\varepsilon}\int_{\mathbb{T}^d} g(\phi_\epsilon)R_\epsilon dxdt\ .
\end{align*}
Since $\int_\tau^T\int_{\mathbb{T}^d} A_{ij}(x)\partial_{j}\phi_\epsilon g'(\phi_\epsilon)\partial _i \phi_\epsilon dxdt\geq 0$, and since $\phi_\epsilon(T+\varepsilon)= \xi^\epsilon\star\phi_T$, we obtain
\begin{align*} &\int_{\mathbb{T}^d}G(\phi_\epsilon(\tau))dx-\int_{\mathbb{T}^d}G(\xi^\epsilon\star\phi_T)dx + \int_\tau^{T+\varepsilon}\int_{\mathbb{T}^d} |\xi_\epsilon\star(D\tilde \phi)|^r g(\phi_\epsilon)dx dt \\ & \leq \int_\tau^{T+\varepsilon}\int_{\mathbb{T}^d} g(\phi_\epsilon)\alpha_\epsilon(t,x) dx dt +\int_\tau^{T+\varepsilon}\int_{\mathbb{T}^d} g(\phi_\epsilon)R_\epsilon dxdt\ .
\end{align*}
Since $g$ is bounded, while $R_\epsilon$ and $\alpha_\varepsilon$ converge in $\pelle r$ and in $\pelle p$ respectively, we can pass to the limit as $\epsilon$ goes to zero and for almost every $\tau$ we get (\ref{chain}).
\vskip0.3em
Now we proceed with the desired estimate. First we observe that, up to replacing $g(r)$ with $g(r\wedge k)$, we can assume that $\phi$ is bounded and that $g$ may be any $C^1$ function. In particular, we take $g(\phi)= \phi^{(\sigma-1) r}$ for $\sigma > 1$, obtaining
\begin{align*}
& \frac1{(\sigma-1)r+1}\int_{\mathbb{T}^d} \phi(\tau)^{(\sigma-1)r+1} \, dx + \frac1{\sigma^r}\int_\tau^T \int_{\mathbb{T}^d} |D\phi^\sigma |^r\, dxdt
\\
& \leq \int_\tau^T \int_{\mathbb{T}^d} \alpha\, \phi^{(\sigma-1)r}\, dxdt + \frac1{(\sigma-1)r+1}\int_{\mathbb{T}^d} \phi_T^{(\sigma-1)r+1}\, dx.
\end{align*}
Let us denote henceforth by $c$ possibly different constants only depending on $r$, $\sigma$, $d$ and $T$. By arbitrariness of $\tau$, the previous inequality implies
\begin{align*}
& \| \phi^\sigma\|_{\limitate{\frac{(\sigma-1)r+1}\sigma}}^{\frac{(\sigma-1)r+1}\sigma} + \|D\phi^\sigma\|_{\pelle r}^r
\\
& \leq c \int_0^T \int_{\mathbb{T}^d} \alpha\, \phi^{(\sigma-1)r}\, dxdt + c \int_{\mathbb{T}^d} \phi_T^{(\sigma-1)r+1}\, dx.
\end{align*}
On the other hand, by interpolation we have (see e.g. \cite[Proposition 3.1, Chapter 1]{DiB})
\begin{equation}\label{dibe}
\| v\|_{\pelle q}^q \leq c\, \| v\|_{\limitate \eta}^{\frac{\eta r}d} \|Dv\|_{\pelle r}^r \qquad \hbox{where $q=r \,\frac{d+\eta}d$}
\end{equation}
for any $v\in L^r((0,T);W^{1,r}(\mathbb{T}^d))$ such that $\int_{\T^d} v(t) \, dx=0$ a.e. in $(0,T)$. So we deduce that
\begin{align*}
& \| \phi^\sigma\|_{\pelle q}^q \leq c \left\{\int_0^T \int_{\mathbb{T}^d} \alpha\, \phi^{(\sigma-1)r}\, dxdt + \int_{\mathbb{T}^d} \phi_T^{(\sigma-1)r+1}\, dx\right\}^{1+\frac rd}
\\
& \qquad + c \int_0^T \left(\int_{\T^d} \phi(t)^\sigma\, dx \right)^q\, dt
\end{align*}
for $\eta=\frac{(\sigma-1)r+1}\sigma$ and $q= r\,\frac{\eta+d}d$. We choose $\sigma$ such that
$$
\sigma q= (\sigma-1)r p'
$$
and therefore, by H\"older inequality, we conclude
\begin{align*}
& \| \phi^\sigma\|_{\pelle q}^q \leq c \, \| \alpha\|_{\pelle p}^{1+\frac rd} \| \phi^\sigma\|_{\pelle q}^{\frac q{p'}(1+\frac rd)} + c \left(\int_{\mathbb{T}^d} \phi_T^{(\sigma-1)r+1}\, dx\right)^{1+\frac rd}
\\
& \qquad + c \int_0^T \left(\int_{\T^d} \phi(t)^\sigma\, dx \right)^q\, dt\,.
\end{align*}
Since $ \int_{\T^d} \phi(t)\, dx$ is estimated in terms of $\|\alpha\|_{\pelle1}$ and $\|\phi_T\|_{\elle1}$, last term can be absorbed into the left-hand side up to a constant $C= C( \|\alpha\|_{\pelle1}, \|\phi_T\|_{\elle1})$.
Moreover, since $p<1+\frac dr$, we have $\frac q{p'}(1+\frac rd)<q$. Hence we end up with an estimate
$$
\| \phi^\sigma\|_{\pelle q}^q \leq C( \|\alpha\|_{\pelle p}, \|\phi_T\|_{\elle{(\sigma-1)r+1} })\,.
$$
Computing the value of $\sigma$ in terms of $r$ and $p$ we get
$$
q\sigma= \frac{rp(1+d)}{d-r(p-1)}\qquad \hbox{and}\quad (\sigma-1)r+1=\frac{d( r(p-1)+1)}{d-r(p-1)}
$$
so the first part of the Theorem is proved.
\vskip0.3em
Finally, we prove the $L^\infty$ estimate by using a strategy which goes back to \cite{St}. To this purpose, we replace $\phi$ with $\phi-k$ and use \eqref{chain} with $g(s)= (s_+)^{r'}$; for any $k\geq \|\phi_T\|_{L^\infty(\mathbb{T}^d)}$ we obtain
$$
\int_{\mathbb{T}^d} [(\phi-k)_+(\tau)]^{\sigma+1} \, dx + \int_\tau^T \int_{\mathbb{T}^d} |D(\phi-k)_+^\sigma |^r\, dxdt
\leq \int_\tau^T \int_{\mathbb{T}^d} \alpha\, (\phi-k)_+^{\sigma}\, dxdt
$$
with $\sigma=r'$. Using as before the embedding \eqref{dibe} we get
\begin{align*}
& \|(\phi-k)_+^\sigma\|_{\pelle q}^q\leq c \left\{ \int_0^T \int_{\mathbb{T}^d} \alpha\, (\phi-k)_+^{\sigma}\, dxdt \right\}^{1+\frac rd}+ c \int_0^T \left(\int_{\T^d} (\phi-k)_+^\sigma\, dx \right)^q\, dt
\\
& \qquad \leq c\left\{ \int_0^T \int_{\mathbb{T}^d} \alpha\, (\phi-k)_+^{\sigma}\, dxdt \right\}^{1+\frac rd}+
c \int_0^T |\{ x\, : \phi(t)>k\}|^{q-1}\int_{\T^d} (\phi-k)_+^{\sigma\, q}\,dxdt\,,
\end{align*}
where, using that $\sigma=r'$, we have
\begin{equation}\label{q}
q= r \, \frac{\frac{\sigma+1}\sigma+d}d= \frac rd (d+2-\frac1r)\,.
\end{equation}
Notice that $|\{ x\, : \phi(t)>k\}|$ is uniformly small provided $k$ is large, only depending on $\|\alpha\|_{L^1}$ and $\|\phi_T\|_{L^1}$. Therefore, absorbing last term in the left-hand side we deduce
$$
\|(\phi-k)_+^\sigma\|_{\pelle q}^q\leq c \left\{ \int_0^T \int_{\mathbb{T}^d} \alpha\, (\phi-k)_+^{\sigma}\, dxdt \right\}^{1+\frac rd}
$$
for some $c=c(\|\alpha\|_{\pelle1}, \|\phi_T\|_{\elle 1})$. One can check that, since $r>1$, \eqref{q} implies $q>1+\frac rd$ and, in particular, $\frac1q+\frac1p<1$. Thus, by H\"older inequality we get
$$
\|(\phi-k)_+^\sigma\|_{\pelle q}^q\leq c \, \|(\phi-k)_+^\sigma\|_{\pelle q}^{1+\frac rd} \, \|\alpha\|_{\pelle p}^{1+\frac rd}\, |A_k|^{(1-\frac1q-\frac1p)(1+\frac rd)} \,,
$$
where $A_k:= \{(t,x)\,: \phi(t,x)>k\}$. Since, for any $h>k$ we have
$$
\int_0^T\int_{\T^d} (\phi-k)_+^{\sigma q}\,dxdt \geq |A_h| (h-k)^{\sigma q}\,,
$$
we end up with the inequality
$$
|A_h|^{1-\frac1q(1+\frac rd)} (h-k)^{\sigma\, q- \sigma(1+\frac rd)}\leq \|(\phi-k)_+^\sigma\|_{\pelle q}^{q-(1+\frac rd)} \leq c \, \|\alpha\|_{\pelle p}^{1+\frac rd}\, |A_k|^{(1-\frac1q-\frac1p)(1+\frac rd)}
$$
which means that
$$
|A_h|\leq C\, \frac{|A_k|^{\beta}}{(h-k)^\delta}\qquad \forall h>k\geq \|\phi_T\|_{L^\infty(\mathbb{T}^d)}
$$
for some $C=C(\|\alpha\|_{\pelle p})$, some $\delta>0$ and with $\beta= \frac{(1-\frac1q-\frac1p)(1+\frac rd)}{1-\frac1q(1+\frac rd)}$. One can check that $\beta>1$ since $p>1+\frac dr$. Therefore, by a classical iteration lemma (see e.g. \cite{St}), it follows that $|A_{k_0}|=0$ for some (explicit) $k_0>0$, which in particular implies the desired bound in terms of $\|\alpha\|_{\pelle p}$ and $\|\phi_T\|_{L^\infty(\mathbb{T}^d)}$.
\end{proof}
\begin{Remark}{\rm
The assumption that $\phi$ is nonnegative can be dropped and in this case the estimates are given on $\phi_+$; indeed, if $\phi$ satisfies \rife{ineq}, then $\phi_+$ also does. This can be seen in the previous proof by taking $g=g(r_+)$, with $g(0)=0$.
Let us also stress that the Lipschitz continuity of the matrix $A$ was only used to recover the estimate from the distributional formulation (namely, to be sure that $\phi$ is limit of solutions of smooth approximating problems). The constant $C$ of the estimate, however, does not depend on $A$ in any way; in particular, the estimate will hold uniformly for any viscous approximation to possibly less regular matrices. }
\end{Remark}
As a corollary, we deduce the following result for problem \eqref{eq:HJ}.
\begin{Theorem} \label{thm:estiHJ}
Assume that \eqref{HypGrowthH} and \eqref{HypA} hold true and let $\phi$ satisfy \eqref{eq:HJ} with $\alpha\in \pelle p$, $\phi_T\in \elle\infty$. Then, $\phi_+$ satisfies the estimates of Theorem \ref{int-reg}. In particular, if $\phi$ is bounded below, we have
$$
\|\phi\|_{L^\infty((0,T), L^{\eta}(\mathbb{T}^d))}+ \|\phi\|_{L^\gamma((0,T)\times \mathbb{T}^d)}\leq C
$$
where $\eta=\frac{d( r(p-1)+1)}{d-r(p-1)}$ and $\gamma = \frac{rp(1+d)}{d-r(p-1)}$ if $p<1+\frac dr$ and $\eta=\gamma=+\infty$ if $p>1+\frac dr$, with a constant $C$ depending on $T, p,d,r, C_2, C_3$ (appearing in \eqref{HypGrowthH} and \eqref{HypA}) and on $\|\alpha\|_{L^p((0,T)\times \mathbb{T}^d)}, \|\phi_T\|_{\elle \eta}$ and $\|\phi_-\|_{L^\infty(\mathbb{T}^d)}$.
\end{Theorem}
\section{Two optimization problems}\label{sec:2opti}
Mean field games systems with local coupling can be studied as an optimality condition between two problems in duality.
The first optimization problem is described as follows: let us denote by $\mathcal K_0$ the set of maps $\phi\in {\mathcal C}^2([0,T]\times \mathbb{T}^d)$ such that $\phi(T,x)= \phi_T(x)$ and define, on $\mathcal K_0$, the functional
\begin{equation}\label{DefmathcalA}
{\mathcal A}(\phi)= \int_0^T\int_{\mathbb{T}^d} F^*\left(x,-\partial_t\phi(t,x)-A_{ij}\partial_{ij}\phi+H(x,D\phi(t,x)) \right)\ dxdt - \int_{\mathbb{T}^d} \phi(0,x)dm_0(x).
\end{equation}
Then the problem consists in optimizing
\begin{equation}\label{PB:dual2}
\inf_{\phi\in \mathcal K_0} \mathcal A(\phi)\;.
\end{equation}
For the second optimization problem, let $\mathcal K_1$ be the set of pairs $(m, w)\in L^1((0,T)\times \mathbb{T}^d) \times L^1((0,T)\times \mathbb{T}^d,\mathbb R^d)$ such that $m(t,x)\geq 0$ a.e., with $\displaystyle \int_{\mathbb{T}^d}m(t,x)dx=1$ for a.e. $t\in (0,T)$, and which satisfy in the sense of distributions the continuity equation
\begin{equation}\label{conteq}
\partial_t m-\partial_{ij}(A_{ij}(x)m)+{\rm div} (w)=0\; {\rm in}\; (0,T)\times \mathbb{T}^d, \qquad m(0)=m_0.
\end{equation}
On the set $\mathcal K_1$, let us define the following functional
$$
{\mathcal B}(m,w)= \int_0^T\int_{\mathbb{T}^d} m(t,x) H^*\left(x, -\frac{w(t,x)}{m(t,x)}\right)+ F(x,m(t,x)) \ dxdt + \int_{\mathbb{T}^d} \phi_T(x)m(T,x)dx
$$
where, for $m(t,x)=0$, we impose that
$$
m(t,x) H^*\left(x, -\frac{w(t,x)}{m(t,x)}\right)=\left\{\begin{array}{ll}
+\infty & {\rm if }\; w(t,x)\neq 0\\
0 & {\rm if }\; w(t,x)=0
\end{array}\right..
$$
Since $H^*$ and $F$ are bounded from below and $m\geq 0$ a.e., the first integral in ${\mathcal B}(m,w)$ is well defined in $\mathbb R\cup\{+\infty\}$. In order to give a meaning to the last integral $\ \int_{\mathbb{T}^d} \phi_T(x)m(T,x)dx\ $ we proceed as follows: let us define$\ \displaystyle v(t,x)= -\frac{w(t,x)}{m(t,x)}\ $ if $m(t,x)>0$ and $\ v(t,x)=0\ $ otherwise.
Thanks to the growth of $H^*$ (implied by \eqref{HypHstar}, which follows from (H2)), ${\mathcal B}(m,w)$ is infinite if $m|v|^{r'}\notin L^{1}(dxdt)$. Therefore, we can assume without loss of generality that $m|v|^{r'}\in L^{1}(dxdt)$, or, equivalently, that $v\in L^{r'}(m\ dxdt)$. In this case equation \eqref{conteq} can be rewritten as a Kolmogorov equation
\begin{equation}\label{conteq12}
\partial_t m-\partial _{ij}(A_{ij}(x)m)-{\rm div} (mv)=0\; {\rm in}\; (0,T)\times \mathbb{T}^d, \qquad m(0)=m_0.
\end{equation}
\begin{Lemma} \label{lem:m-cts}
The map $t\mapsto m(t)$ is H\"older continuous a.e. for the weak* topology of $P(\mathbb{T}^d)$.
\end{Lemma}
This Lemma implies, in particular, that the measure $m(t)$ is defined for any $t$, therefore the second integral term in the definition of ${\mathcal B}(m,w)$ is well defined.
For the sake of completeness, we give the proof here.
\begin{proof} We first extend the pairs $(m,w)$ to $[-1,T]\times \mathbb{T}^d$ by defining $m=m_0$ on $[-1,0]$ and $w(s,x)=0$ for $(s,x)\in (-1,0)\times \mathbb{T}^d$. Note that $\partial_t m-\partial_{ij}(\tilde A_{ij}(t,x)m)+{\rm div}(w)=0$ holds in the sense of distributions on $(-1,T)\times \mathbb{T}^d$, where $\tilde A_{ij}(t,x)=A_{ij}(x)$ if $t\in (0,T)$ and $\tilde A_{ij}(t,x)=0$ otherwise. Let $\xi$ be a standard convolution kernel in $(t,x)$, a support compact on $\mathbb R^{d+1}$ and $\xi^\epsilon(t,x)=\xi((t,x)/\epsilon)/(\epsilon)^{d+1}$, $\xi^\epsilon\geq0,\ \int_{\mathbb R^{d+1}}\xi^\epsilon(t,x) dtdx=1$, for all $\epsilon>0$. Let $m_\epsilon= \xi^\epsilon\star m$ and $w_\epsilon=\xi^\epsilon\star w$. Then $m_\epsilon, w_\epsilon$ are $C^\infty$ and $\int_{\mathbb{T}^d}m_\epsilon(t,x)dx=1$ for all $t\in(0,T)$ and $\epsilon>0$ small enough.
Convolving $\xi^\epsilon$ with (\ref{conteq}), we obtain
$$
\partial_t m_\epsilon-\partial_{ij}(\xi^\epsilon*(\tilde A_{ij}(t,x) m))+{\rm div} (w_\epsilon)=0\; {\rm in}\; (-1/2,T)\times \mathbb{T}^d,
$$
with
$$
m_\epsilon(-1/2,x)=\int_\mathbb R\int_{\mathbb{T}^d} \xi^\epsilon(s, x-y) m_0(y)dyds.
$$
The equation can be rewritten as
\begin{equation}\label{dual-eps}
\partial_t m_\epsilon-\partial_{ij}( \tilde A^\epsilon_{ij}(t,x)m_\epsilon))-{\rm div} (m_\epsilon v_\epsilon)=0\; {\rm in}\; (-1/2,T)\times \mathbb{T}^d
\end{equation}
where
$ \tilde A^\epsilon_{ij}=\frac{\xi^\epsilon\star(\tilde A_{ij}m)}{m_\epsilon}$ and $v_\epsilon=-\frac{w_\epsilon}{m_\epsilon}$.
Let us consider the following stochastic differential equations defined for all $\epsilon>0$
\begin{equation}\label{eqstoch}
\left\{\begin{array}{ll} dX^\epsilon_t=v_\epsilon(t,X^\epsilon_t)dt+\Sigma_\epsilon(X^\epsilon_t) dB^\epsilon_t & t\in[-1/2,T]\\ X^\epsilon_{-1/2}=Z_{-1/2}^\epsilon \end{array}\right.,
\end{equation}
where $dB^\epsilon_t$ is a standard $d$-dimensional Brownian motion over some probability space $(\Omega,\mathcal A,\mathbb P)$, $\Sigma_\epsilon\Sigma_\epsilon^T=\tilde A^\epsilon$, and the initial condition $Z^\epsilon_{-1/2}\in L^1(\mathbb{T}^d)$ is random, independent of $(B^\epsilon_t)$ and with law $m_\epsilon(-1/2,\cdot)$.
For all $\epsilon>0$, the vector field $v_\epsilon$ is continuous, uniformly Lipschitz continuous in space and bounded. Therefore, there exists a unique solution to (\ref{eqstoch}). Moreover, as a consequence of Ito's formula, we have that, if the density $\mathcal L (Z^\epsilon_0)=\xi^\epsilon\star m_0$, then $m_\epsilon(t)=\mathcal L(X^\epsilon_t)$ solves (\ref{dual-eps}) in the sense of distributions.
Let $\bf d_1$ be the Kantorovich-Rubinstein distance on $P(\mathbb{T}^d)$ and $\gamma_\epsilon\in\Pi(m_\epsilon(t),m_\epsilon(s))$ the law of the pair $(X^\epsilon_t, X^\epsilon_s)$ for $0\leq s<t\leq T$, where $\Pi(m_\epsilon(t),m_\epsilon(s))$ is the set of Borel probability measures $\mu$ on $\mathbb{T}^d\times\mathbb{T}^d$ such that $\mu(A\times\mathbb{T}^d)=m_\epsilon(t,A)$ and $\mu(\mathbb{T}^d\times A)=m_\epsilon(s,A)$ for any Borel set $A\in \mathbb{T}^d$. We have
$$
{\bf d_1 }(m_\epsilon(t),m_\epsilon(s))\leq \int_{\mathbb{T}^d\times\mathbb{T}^d}|x-y| d\gamma_\epsilon(x,y)= \mathbb E[|X^\epsilon_t-X^\epsilon_s|].
$$
Moreover,
\begin{align*}
\mathbb E [|X^\epsilon_t-X^\epsilon_s|]&\leq \mathbb E[\int_s^t |v_\epsilon(\tau,X^\epsilon_\tau)|d\tau]+\mathbb E\left[\left|\int_s^t \Sigma_\epsilon(X^\epsilon_\tau)dB_\tau\right|\right]\\
&\leq \int_s^t\int_{\mathbb{T}^d}|v_\epsilon(\tau,x)| m_\epsilon(\tau,x)dx d\tau+\left(\mathbb E\left[\int_s^t \Sigma_\epsilon\Sigma_\epsilon^*(X^\epsilon_\tau)d\tau\right]\right)^{1/2}\\
&\leq \int_s^t\int_{\mathbb{T}^d}|v_\epsilon(\tau,x)| m_\epsilon(\tau,x)dx d\tau+\| A\|_\infty C |t-s|^\frac{1}{2}.
\end{align*}
Recalling the definition of $v_\epsilon$, we have that $m_\epsilon |v_\epsilon|^{r'}= \frac{|w_\epsilon|^{r'}}{m_\epsilon^{r'-1}}$ belongs to $L^1([0,T]\times\mathbb{T}^d)$ for all $\epsilon>0$. Indeed, the function $(m,w)\mapsto \frac{|w|^{r'}}{m^{r'-1}}$ is convex and $\frac{|w|^{r'}}{m^{r'-1}}$ belongs to $L^1([0,T]\times\mathbb{T}^d)$. Thus
$$
\int_{0}^{T }\int_{\mathbb{T}^d} \frac{|\xi^\epsilon \star w|^{r'}}{(\xi^\epsilon\star m)^{r'-1}} dx d\tau\leq\int_{0}^{T} \int_{\mathbb{T}^d} \xi^\epsilon\star \left(\frac{|w|^{r'}}{m^{r'-1}}\right) dx d\tau\leq \left\|\frac{|w|^{r'}}{m^{r'-1}} \right\|_1.$$
Therefore, using H\"older inequality,
\begin{align*}
{\bf d_1 }(m_\epsilon(t),m_\epsilon(s))
&\leq \left(\int_s^t\int_{\mathbb{T}^d}|v_\epsilon(\tau,x)| ^{r'}m_\epsilon(\tau,x)dx d\tau\right)^{\frac{1}{r'}}\left(\int_s^t\int_{\mathbb{T}^d} m_{\epsilon}(\tau,x)dxd\tau\right)^\frac{1}{r}+\|A\|_\infty C |t-s|^\frac{1}{2}\\
&\leq \left\|\frac{|w|^{r'}}{m^{r'-1}} \right\|_1^{\frac{1}{r'}} |t-s|^{\frac{1}{r}}+\|A\|_\infty |t-s|^\frac{1}{2}.
\end{align*}
Letting $\epsilon\to 0$ we have $m_\epsilon\to m$ in $L^1([0,T]\times\mathbb{T}^d)$ and for a.e. $\tau \in [0,T],$ $ m_\epsilon(\tau )\to m(\tau)$ in $L^1(\mathbb{T}^d)$, moreover for a.e. $0\leq s<t\leq T$
$$
\lim_{\epsilon\to 0 } {\bf d_1 }(m_\epsilon(t),m_\epsilon(s))={\bf d_1 }(m(t),m(s)).
$$
Thus for a.e. $0\leq s<t\leq T$
$$
{\bf d_1 }(m(t),m(s))\leq C |t-s|^{\frac{1}{r}} + \|A\|_\infty |t-s|^\frac{1}{2}.
$$
\end{proof}
The second optimal control problem is the following:
\begin{equation}\label{Pb:mw2}
\inf_{(m,w)\in \mathcal K_1} \mathcal B(m,w)\;.
\end{equation}
\begin{Lemma}\label{Lem:dualite} We have
$$
\inf_{\phi\in \mathcal K_0}{\mathcal A}(\phi) = - \min_{(m,w)\in \mathcal K_1} {\mathcal B}(m,w).
$$
Moreover, the minimum in the right-hand side is achieved by a unique pair $(m,w)\in \mathcal K_1$ satisfying $(m,w)\in L^q((0,T)\times \mathbb{T}^d)\times L^{\frac{r'q}{r'+q-1}}((0,T)\times \mathbb{T}^d)$.
\end{Lemma}
\begin{Remark}{\rm Note that $\frac{r'q}{r'+q-1}>1$ because $r'>1$ and $q>1$.
}\end{Remark}
\begin{proof}
The strategy of proof---which is very close to the corresponding one in \cite{c1, cg, CCN}---consists in applying the Fenchel-Rockafellar duality theorem (cf. e.g., \cite{ET}). In order to do so, it is better to reformulate the first optimization problem \eqref{PB:dual2} in a more suitable form.
Let $E_0= {\mathcal C}^2([0,T]\times \mathbb{T}^d)$ and $E_1= {\mathcal C}^0([0,T]\times \mathbb{T}^d, \mathbb R)\times {\mathcal C}^0([0,T]\times \mathbb{T}^d, \mathbb R^d)$. We define on $E_0$ the functional
$$
{\mathcal F}(\phi)= -\int_{\mathbb{T}^d} m_0(x)\phi(0,x)dx +\chi_S(\phi),
$$
where $\chi_S$ is the characteristic function of the set $S=\{\phi\in E_0, \; \phi(T, \cdot)= \phi_T\}$, i.e., $\chi_S(\phi)=0$ if $\phi\in S$ and $+\infty$ otherwise. For $(a,b)\in E_1$, we define
$$
{\mathcal G} (a,b)= \int_0^T\int_{\mathbb{T}^d} F^*(x,-a(t,x)+H(x,b(t,x)) )\ dxdt \;.
$$
The functional ${\mathcal F}$ is convex and lower semi-continuous on $E_0$ while ${\mathcal G}$ is convex and continuous on $E_1$. Let $\Lambda:E_0\to E_1$ be the bounded linear operator defined by $\displaystyle \Lambda (\phi)= (\partial_t\phi+A_{ij}\partial_{ij}\phi , D\phi)$. We can observe that
$$
\inf_{\phi\in \mathcal K_0} \mathcal A(\phi)= \inf_{\phi\in E_0} \left\{ {\mathcal F}(\phi)+{\mathcal G}(\Lambda(\phi))\right\}.
$$
It is easy to verify that the qualification hypothesis, that ensures the stability of the above optimization problem, holds. Indeed, there is a map $\phi$ such that ${\mathcal F}(\phi)<+\infty$ and
such that ${\mathcal G} $ is continuous at $\Lambda (\phi)$: it is enough to take $\phi(t,x)=\phi_T(x)$.
Therefore we can apply the Fenchel-Rockafellar duality theorem, which states that
$$
\inf_{\phi\in E_0} \left\{ {\mathcal F}(\phi)+{\mathcal G}(\Lambda(\phi))\right\}
=
\max_{ (m,w)\in E_1'} \left\{ - {\mathcal F}^*(\Lambda^*(m,w))-{\mathcal G}^*(-(m,w))\right\}
$$
where $E_1'$ is the dual space of $E_1$, i.e., the set of vector valued Radon measures $(m, w)$ over $[0,T]\times \mathbb{T}^d$ with values in $\mathbb R\times \mathbb R^d$, $E_0'$ is the dual space of $E_0$, $\Lambda^*: E'_1\to E'_0$ is the dual operator of $\Lambda$ and ${\mathcal F}^*$ and ${\mathcal G}^*$ are the convex conjugates of ${\mathcal F}$ and ${\mathcal G}$ respectively. By a direct computation we have
$$
{\mathcal F}^*(\Lambda^*(m,w))
= \left\{ \begin{array}{ll}
\displaystyle \int_{\mathbb{T}^d} \phi_T(x)dm(T,x) & {\rm if }\; \partial_t m-\partial_{ij}(A_{ij}m)+{\rm div} (w)=0, \; m(0)=m_0\\
+\infty & {\rm otherwise}
\end{array}\right.
$$
where the equation $\ \partial_t m-\partial_{ij}(A_{ij}m)+{\rm div} (w)=0, \; m(0)=m_0\ $ holds in the sense of distributions.
Following \cite{c1}, we have ${\mathcal G}^*(m,w)=+\infty$ if $(m,w)\notin L^1$ and, if $(m,w)\in L^1$,
$$
{\mathcal G}^*(m,w)= \int_0^T\int_{\mathbb{T}^d} K^*(x,m(t,x),w(t,x))dtdx ,
$$
where $$K^*(x,m,w)=\left\{\begin{array}{ll}F(x,-m)-mH^*(x,-\frac{w}{m})& \text{ if } m<0 \\
0& \text{ if } m=0,w=0\\
+\infty & \text{ otherwhise} \end{array}\right.$$ is the convex conjugate of $$ K (x,a,b)=F^*(x,-a+H(x,b))\quad \forall (x,a,b)\in\mathbb{T}^d\times\mathbb R\times\mathbb R^d.$$
Therefore
$$
\begin{array}{l}
\displaystyle \max_{ (m,w)\in E_1'} \left\{ - {\mathcal F}^*(\Lambda^*(m,w))-{\mathcal G^*}(-(m,w))\right\}\\
\qquad \qquad \qquad \displaystyle = \max \left\{ \int_0^T\int_{\mathbb{T}^d} -F(x,m ) -m H^*( x, -\frac{w}{m})\ dtdx -\int_{\mathbb{T}^d} \phi_T(x) m(T,x)\ dx \right\}
\end{array}
$$
where the last maximum is taken over the $L^1$ maps $(m,w)$ such that $m\geq 0$ a.e. and
$$
\partial_t m-\partial_{ij}(A_{ij}m)+{\rm div} (w)=0,\; m(0)=m_0\,
$$
holds in the sense of distributions.
Since $\displaystyle\ \int_{\mathbb{T}^d} m_0=1\ $, it follows that $\displaystyle\ \int_{\mathbb{T}^d} m(t)=1\ $ for any $t\in [0,T]$. Thus the pair $(m,w)$ belongs to the set $\mathcal K_1$ and the first part of the statement is proved.
Take now an optimal $(m, w)\in \mathcal K_1$ in the above system. Observe that due to optimality we have $w(t,x)=0$ for all $(t,x)\in [0,T]\times\mathbb{T}^d$ such that $m(t,x)=0$. The growth conditions \eqref{HypGrowthH} and \eqref{HypGrowthF} imply
$$
\begin{array}{rl}
C \; \geq & \displaystyle \int_0^T\int_{\mathbb{T}^d} F(x,m ) +m H^*( x, -\frac{w}{m})\ dtdx +\int_{\mathbb{T}^d} \phi_T(x) m(T,x)\ dx \\
\geq & \displaystyle \int_0^T\int_{\mathbb{T}^d} \left(\frac{1}{ C}|m|^{q}+
\frac{m}{ C } \left|\frac{w}{m}\right|^{r'} - C(m+1) \right) dxdt -\|\phi_T\|_\infty .
\end{array}
$$
Therefore $m\in L^q$. Moreover, by H\"older inequality, we also have
$$
\int_0^T\int_{\mathbb{T}^d} |w|^{\frac{r'q}{r'+q-1}} = \int\int_{\{m>0\}} |w|^{\frac{r'q}{r'+q-1}} \leq \|m\|_q^{\frac{r'-1}{r'+q-1}} \left(\int\int_{\{m>0\}} \frac{|w|^{r'}}{m^{r'-1}} \right)^{\frac{q}{r'+q-1}} \leq C
$$
so that $w\in L^{\frac{r'q}{r'+q-1}}$. Finally, a minimizer to \eqref{Pb:mw2} should be unique, because the set $\mathcal K_1$ is convex and the maps $F(x,\cdot)$ and $H^*(x,\cdot)$ are strictly convex: thus $m$ is unique and so is $\displaystyle \frac{w}{m}$ in $\{m>0\}$. As $w=0$ in $\{m=0\}$, uniqueness of $w$ follows as well.
\end{proof}
\section{Analysis of the optimal control of the HJ equation}\label{sec:optiHJ}
In general, we do not expect problem \eqref{PB:dual2} to have a solution. In this section we exhibit a relaxation for \eqref{PB:dual2} (Proposition \ref{prop:valeursegales}) and show that this relaxed problem has at least one solution (Proposition \ref{Prop:existence}).
\subsection{The relaxed problem}
Recall that the exponents $\eta>1$ and $\gamma>1$ are defined in Theorem \ref{thm:estiHJ}.
Let ${\mathcal K}$ be the set of pairs $(\phi,\alpha)\in L^\gamma((0,T)\times \mathbb{T}^d)\times L^p((0,T)\times \mathbb{T}^d)$ such that $D\phi\in L^r((0,T)\times \mathbb{T}^d)$ and which satisfy in the sense of distributions
\begin{equation}\label{ineq:phi}
-\partial_t \phi-A_{ij}(x)\partial_{ij}\phi+H(x,D\phi) \leq \alpha, \qquad \phi(T,\cdot)\leq \phi_T
\end{equation}
(for the precise meaning of the inequality, see the beginning of Section \ref{sec:esti}).
The following statement explains that $\phi$ has a ``trace" in a weak sense.
\begin{Lemma}\label{lem:intem0phiBV} Let $(\phi,\alpha)\in {\mathcal K}$. Then, for any Lipschitz continuous map $\zeta: \mathbb{T}^d\to \mathbb R$, the map $\displaystyle t\to \int_{\mathbb{T}^d} \zeta(x)\phi(t,x)dx$ has a BV representative on $[0,T]$. Moreover, if we denote by $\displaystyle \int_{\mathbb{T}^d} \zeta(x)\phi(t^+,x)dx$ its right limit at $t\in [0,T)$, then
the map $\displaystyle \zeta\to \int_{\mathbb{T}^d} \zeta(x)\phi(t^+,x)dx$ is continuous in $L^{\eta'}(\mathbb{T}^d)$.
\end{Lemma}
As a consequence, for any nonnegative $C^1$ map $\vartheta:[0,T]\times \mathbb{T}^d\to \mathbb R$, one can write the integration by parts formula: for any $0\leq t_1\leq t_2\leq T$,
$$
-\left[\int_{\mathbb{T}^d}\vartheta\phi\right]_{t_1}^{t_2} + \int_{t_1}^{t_2}\int_{\mathbb{T}^d} \phi\partial_t\vartheta+ \lg D\vartheta, AD\phi\rg + \vartheta(\partial_iA_{ij}\partial_j\phi+ H(x,D\phi)) \leq \int_{t_1}^{t_2}\int_{\mathbb{T}^d} \alpha \vartheta.
$$
\begin{proof}[Proof of Lemma \ref{lem:intem0phiBV}.] One easily checks that, for any Lipschitz continuous, nonnegative map $\zeta: \mathbb{T}^d\to \mathbb R$,
$$
- \frac{d}{dt} \int_{\mathbb{T}^d} \zeta \phi(t)+ \int_{\mathbb{T}^d} \lg D\zeta,AD\phi(t)\rg +\zeta (\partial_iA_{ij}\partial_j\phi+ H(x,D\phi) -\alpha ) \leq 0,
$$
holds in the sense of distributions. As the second integral is in $L^1((0,T))$, the map $t\to \int_{\mathbb{T}^d} \zeta \phi(t)$ is BV. If now $\zeta$ is Lipschitz continuous and changes sign, one can write $\zeta=\zeta^+-\zeta^-$ and the map $t\to \int_{\mathbb{T}^d} \zeta \phi(t)= \int_{\mathbb{T}^d} \zeta^+ \phi(t)- \int_{\mathbb{T}^d} \zeta^- \phi(t)$ is still BV. The continuity with respect to $\zeta$ comes from the $L^\infty((0,T), L^\eta(\mathbb{T}^d))$ estimate on $\phi$ given in Theorem \ref{thm:estiHJ}.
\end{proof}
We extend the functional $\mathcal A$ to ${\mathcal K}$ by setting
$$
\mathcal A(\phi,\alpha) = \int_0^T\int_{\mathbb{T}^d} F^*(x,\alpha(x,t))\ dxdt - \int_{\mathbb{T}^d} \phi(x,0)m_0(x)\ dx\qquad \forall (\phi, \alpha)\in {\mathcal K}.
$$
The next proposition explains that the problem
\begin{equation}\label{PB:dual-relaxed}
\inf_{(\phi, \alpha)\in {\mathcal K}} \mathcal A(\phi,\alpha)
\end{equation}
is the relaxed problem of \eqref{PB:dual2}. For this we first note that
\begin{equation}\label{eq:positif}
\inf_{(\phi, \alpha)\in {\mathcal K}} \mathcal A(\phi,\alpha)= \inf_{(\phi, \alpha)\in {\mathcal K}, \ \alpha\geq0\ {\rm a.e.}} \mathcal A(\phi,\alpha)
\end{equation}
because one can always replace $\alpha$ by $\alpha\vee0$ since $F^*(x,\alpha)= 0$ for $\alpha\leq0$.
\begin{Proposition}\label{prop:valeursegales} We have
$$
\inf_{\phi\in \mathcal K_0} \mathcal A(\phi)= \inf_{(\phi, \alpha)\in {\mathcal K}} \mathcal A(\phi,\alpha).
$$
\end{Proposition}
The proof requires the following inequality:
\begin{Lemma}\label{lem:phialphamw} Let $(\phi,\alpha)\in {\mathcal K}$ and $(m,w)\in {\mathcal K}_1$. Assume that $mH^*(\cdot, -w/m)\in L^1((0,T)\times \mathbb{T}^d)$ and $m \in \pelle q$. Then
\begin{equation}\label{ineq:ineqfirst}
\left[ \int_{\mathbb{T}^d} m \phi\right]_t^{T} + \int_t^{T}\int_{\mathbb{T}^d} m \left(\alpha + H^*(x,-\frac{w}{m})\right) \; \geq \; 0
\end{equation}
and
$$
\left[ \int_{\mathbb{T}^d} m \phi\right]_0^{t} + \int_0^{t}\int_{\mathbb{T}^d} m \left(\alpha + H^*(x,-\frac{w}{m})\right) \; \geq \; 0.
$$
Moreover, if equality holds in the inequality \eqref{ineq:ineqfirst} for $t=0$, then $w= -mD_pH(x,D\phi)$ a.e.
\end{Lemma}
\begin{proof} We first extend the pairs $(m,w)$ to $[-1,T+1]\times \mathbb{T}^d$ by defining $m=m_0$ on $[-1,0]$, $m= m(T)$ on $[T,T+1]$ and $w(s,x)=0$ for $(s,x)\in (-1,0)\cup (T,T+1)\times \mathbb{T}^d$. Note that $\partial_t m-\partial_{ij}(\tilde A_{ij}(t,x)m)+{\rm div}(w)=0$ on $(-1,T+1)\times \mathbb{T}^d$, where $\tilde A_{ij}(t,x)=A_{ij}(x)$ if $t\in (0,T)$ and $\tilde A_{ij}(t,x)=0$ otherwise.
Let $\xi^\epsilon= \xi^\epsilon(t,x)$ be a smooth convolution kernel with support in $B_\epsilon$; we smoothen the pair $(m,w)$ in a standard way into $(m_\epsilon,w_\epsilon)$.
Then $(m_\epsilon,w_\epsilon)$ solves
\begin{equation}
\label{eq:mep}
\partial_t m_\epsilon-\partial_{ij} (\tilde A_{ij} m_\epsilon)+{\rm div}(w_\epsilon)= \partial_i R_\epsilon \qquad {\rm in}\; (-1/2, T+1/2)
\end{equation}
in the sense of distributions, where
\begin{equation}
R_\epsilon := [\xi^\epsilon,\partial_j \tilde A_{ij}](m) + [\xi^\epsilon,\tilde A_{ij}\partial_j](m).
\end{equation}
Here we use again the commutator notation (cf. \cite{DL89})
\begin{equation} \label{eq:commutators}
[\xi^\epsilon,c](f) := \xi^\epsilon \star (cf) - c(\xi^\epsilon \star f)\,.
\end{equation}
Invoking \cite[Lemma II.1]{DL89}, we have that $R_\epsilon \to 0$ in $L^q$, since $m \in L^q$ and $\tilde A_{ij} \in W^{1,\infty}$.
Let us fix time $t\in (0,T)$ at which $\phi(t^+)=\phi(t^-)= \phi(t)$ in $L^\gamma(\mathbb{T})$ and $m_\epsilon(t)$ converges to $m(t)$.
By the inequality satisfied by $(\phi,\alpha)$, we have
$$
\displaystyle \int_t^{T}\int_{\mathbb{T}^d} \phi \partial_t m_\epsilon +\partial_i \phi \partial_{j}( \tilde A_{ij}m_\epsilon)+ m_\epsilon H(x,D\phi) + \int_{\mathbb{T}^d} m_\epsilon(t) \phi(t) - m_\epsilon(T)\phi_T\\
\leq \int_t^{T}\int_{\mathbb{T}^d} \alpha m_\epsilon\,.
$$
By \eqref{eq:mep} we have
$$
\int_t^{T}\int_{\mathbb{T}^d} \phi \partial_t m_\epsilon +\partial_i \phi \partial_{j}( \tilde A_{ij}m_\epsilon)= \int_t^{T}\int_{\mathbb{T}^d} -\partial_i \phi R_\epsilon + \lg D\phi, w_\epsilon\rg.
$$
On the other hand, by convexity of $H$,
\begin{equation}\label{ineq:ineqCV}
\begin{array}{rl}
\displaystyle \int_t^{T}\int_{\mathbb{T}^d} - m_\epsilon H^*(x,-\frac{w_\epsilon}{m_\epsilon}) \;
\leq & \displaystyle \int_t^{T}\int_{\mathbb{T}^d} \lg w_\epsilon, D\phi\rg + m_\epsilon H(x,D\phi) \,.
\end{array}
\end{equation}
Collecting the above (in)equalities we obtain
$$
\int_{\mathbb{T}^d} m_\epsilon(t) \phi(t)\leq \int_{\T^d} m_\epsilon(T)\phi_T + \int_t^T \int_{\mathbb{T}^d} m_\epsilon (\alpha+ H^*(x,-\frac{w_\epsilon}{m_\epsilon})) + \partial_j \phi R_\epsilon\, .
$$
By assumption \eqref{CondCroiss} which states that $r\geq p$, and since $D\phi \in L^r$, we have $\displaystyle \iint \partial_j \phi R_\epsilon\to 0$ as $\epsilon\to 0$.
Following the proof of Lemma 2.7 in \cite{cg} we have
$$
\int_t^{T}\int_{\mathbb{T}^d} - m_\epsilon H^*(x,-\frac{w_\epsilon}{m_\epsilon}) \to \int_t^{T}\int_{\mathbb{T}^d} - m H^*(x,-\frac{w}{m}) \qquad
\mbox{\rm as $\epsilon\to 0$. }
$$
The continuity of $t\to m(t)$ in $P(\mathbb{T}^d)$ given by Lemma \ref{lem:m-cts} implies the convergence
$$
\int_{\mathbb{T}^d} m_\epsilon(T) \phi_T\to \int_{\mathbb{T}^d} m(T) \phi_T\,.
$$
Recalling that $\phi$ is bounded below, we finally get by Fatou's Lemma the inequality
$$
-\|\phi_-\|_\infty+ \int_{\mathbb{T}^d} m(t) (\phi(t)+ \|\phi_-\|_\infty) \leq \int_{\T^d} m(T)\phi_T + \int_t^T \int_{\T^d} m(\alpha+H^*(x, -\frac{w}{m})),
$$
which implies that $m(t) \phi(t)$ is integrable with
$$
\int_{\T^d} m(t) \phi(t)\leq \int_{\T^d} m(T)\phi_T + \int_t^T \int_{\T^d} m(\alpha+H^*(x, -\frac{w}{m}))
$$
We can argue similarly in the time interval $[0,t]$ using that $\int_{\T^d} m_\epsilon(t) \phi(t) \to \int_{\T^d} m(t) \phi(t)$; this is certainly true, up to a subsequence, for a.e. $t$, because $m_\epsilon \phi$ strongly converges in $\pelle1$ since $\phi\in \pelle \gamma$, $m\in \pelle q$ and $\gamma\geq p$. We obtain then
$$
\int_{\mathbb{T}^d} m_0 \phi(0)\leq \int_{\T^d} m(t)\phi(t)+ \int_0^t \int_{\mathbb{T}^d} m (\alpha+ H^*(x,-\frac{w}{m})) .
$$
Let us assume finally that the following equality holds:
$$
\left[ \int_{\mathbb{T}^d} m \phi\right]_0^{T} + \int_0^{T}\int_{\mathbb{T}^d} m \left(\alpha + H^*(x,-\frac{w}{m})\right) \; = \; 0.
$$
Then there is an equality in inequality \eqref{ineq:ineqfirst} for almost all $t$.
Fix such a $t\in (0,T)$ and let
$$
E_\sigma(t):= \left\{(s,y)\;, \; s\in [t,T], \; m (H^*(y,-\frac{w}{m}) + H(x,D\phi)) \geq - \lg w, D\phi\rg +\sigma\right\}.
$$ If $|E_\sigma(t)|>0$, then
for $\epsilon>0$ small enough, the set
$$
E_{\epsilon,\sigma}(t):= \{(s,y)\;, \; s\in [t,T], \; m_\epsilon (H^*(y,-\frac{w_\epsilon}{m_\epsilon}) + H(x,D\phi)) \geq - \lg w_\epsilon, D\phi\rg +\sigma/2\}
$$ has a measure larger than $|E_{\sigma}(t)|/2$. Coming back to inequality \eqref{ineq:ineqCV}, we have
$$
\begin{array}{rl}
\displaystyle \int_t^{T}\int_{\mathbb{T}^d} - m_\epsilon H^*(x,-\frac{w_\epsilon}{m_\epsilon}) \;
\leq & \displaystyle \int_t^{T}\int_{\mathbb{T}^d} \lg w_\epsilon, D\phi\rg + m_\epsilon H(x,D\phi) -|E_{\sigma}(t)|\sigma/4
\end{array}
$$
Then inequality \eqref{ineq:ineqfirst} becomes
$$
\int_{\T^d} m(t) \phi(t)\leq \int_{\T^d} m(T)\phi_T + \int_t^T \int_{\T^d} m(\alpha+H^*(x, -\frac{w}{m})) -|E_{\sigma}(t)|\sigma/4,
$$
which contradicts the fact that there is an equality in \eqref{ineq:ineqfirst}. So $|E_\sigma(t)|=0$ for any $\sigma$ and for a.e. $t$, which shows that
$m (H^*(y,-\frac{w}{m}) +H(x,D\phi)) =- \lg w, D\phi\rg$ a.e. Thus $w= -m D_pH(x,D\phi)$ holds a.e. in $\{m>0\}$ and, as $w=0$ in $\{m=0\}$, a.e. in $(0,T)\times \mathbb{T}^d$.
\end{proof}
\begin{proof}[Proof of Proposition \ref{prop:valeursegales}] We follow the argument developed by Graber in \cite{Gra14}.
Inequality $\ \displaystyle \inf_{\phi\in \mathcal K_0} \mathcal A(\phi)\geq \inf_{(\phi, \alpha)\in {\mathcal K}} \mathcal A(\phi,\alpha)\ $ being obvious, let us check the reverse one. Let $(\phi,\alpha)\in {\mathcal K}$. For any $(m, w)\in {\mathcal K}_1$ with $mH^*(\cdot, -\frac{w}{m})\in L^1$, we have, by Lemma \ref{lem:phialphamw},
$$
\begin{array}{rl}
\displaystyle \mathcal A(\phi,\alpha) \; \geq & \displaystyle \int_0^T \int_{\mathbb{T}^d} \alpha m - F(m) - \int_{\T^d} m_0\phi(0) \\
\geq & \displaystyle \int_0^T \int_{\T^d} -m H^*(x, -\frac{w}{m})- F(m)- \int_{\T^d} m(T)\phi_T = -\mathcal B(m,w)
\end{array}
$$
Taking the sup with respect to $(m,w)$ in the right-hand side we obtain thanks to Lemma \ref{Lem:dualite}:
$$
\mathcal A(\phi,\alpha) \; \geq \; - \inf_{(m, w)\in {\mathcal K}_1} \mathcal B(m,w) = \inf_{\phi\in \mathcal K_0} \mathcal A(\phi).
$$
\end{proof}
\subsection{Existence of a solution for the relaxed problem}
The next proposition explains the interest of considering the relaxed problem \eqref{PB:dual-relaxed} instead of the original one \eqref{PB:dual2}.
\begin{Proposition}\label{Prop:existence} The relaxed problem \eqref{PB:dual-relaxed} has at least one solution $(\phi, \alpha) \in {\mathcal K}$ which is bounded below by a constant depending on $\|\phi_T\|_{C^2}$, on $\|A_{ij}\|_{C^0}$ and on $\|H(\cdot, D\phi_T)\|_\infty$.
\end{Proposition}
\begin{proof} We start with the construction of a suitable minimizing sequence. Let $(\tilde \phi_n)$ be a minimizing sequence for problem \eqref{PB:dual2} and let us set
\begin{equation}\label{defalphan}
\alpha_n(t,x)= \max\{0\ ; \ -\partial_t\tilde \phi_n(t,x) -A_{ij}\partial_{ij}\tilde \phi_n(t,x)+H(x,D\tilde \phi_n(t,x))\}.
\end{equation}
By Proposition \ref{prop:valeursegales} and the fact that $F^*(x,\alpha)=0$ if $\alpha\leq0$, the pair $(\phi_n, \alpha_n)$ is also a minimizing sequence of \eqref{PB:dual-relaxed}. Let $\psi$ be the unique viscosity solution to
$$
-\partial_t \psi-A_{ij}(x)\partial_{ij}\psi+H(x,D\psi)= 0, \qquad \psi(T,\cdot)= \phi_T.
$$
As $\phi_T$ is $C^2$, $\psi(t,x)\geq \tilde \phi_T(x)-C(T-t)$, where the constant $C$ depends on $\|\phi_T\|_{C^2}$, on $\|A_{ij}\|_{C^0}$ and on $\|H(\cdot, D\phi_T)\|_\infty$. Let $\phi_n$ be the (continuous) viscosity solution to
\begin{equation}\label{eqphin}
-\partial_t \phi_n-A_{ij}(x)\partial_{ij}\phi_n+H(x,D\phi_n)\leq \alpha_n, \qquad \psi(T,\cdot)\leq \phi_T.
\end{equation}
By comparison, $\phi_n\geq \tilde \phi_n\vee \psi$. As $H$ is convex, \eqref{eqphin} holds in the sense of distributions (see \cite{Ish95}). Therefore the sequence $(\phi_n,\alpha_n)$ is still minimizing, with the following bound below for $(\phi_n)$:
\begin{equation}\label{unifboundbelow}
\phi_n(t, x)\geq \phi_T(x)-C(T-t).
\end{equation}
\noindent {\bf Step 1:} We claim that $(\alpha_n)$ is bounded in $L^p((0,T)\times \mathbb{T}^d)$. For this, we integrate \eqref{eqphin} against $m_0$ on $(0,T)\times \mathbb{T}^d$
$$
\int_{\mathbb{T}^d} \phi_n(0)m_0 +\int_0^T\int_{\mathbb{T}^d} \partial_i m_0A_{ij}\partial_j \phi_n + (\partial_j A_{ij}) m_0\partial_j \phi_n +m_0H(x,D\phi_n) \leq \int_0^T \int_{\T^d} m_0\alpha_n+\int_{\T^d} \phi_Tm_0.
$$
As $(1/C_0)\leq m_0\leq C_0$ for some $C_0>0$, $\|Dm_0\|_\infty<+\infty$ and $H$ is coercive, we get
\begin{equation}\label{phi(0)Dphi}
\int_{\mathbb{T}^d} \phi_n(0)m_0 +\frac{1}{C} \int_0^T\int_{\T^d} |D\phi_n|^r \leq C_0\|\alpha_n\|_p+C.
\end{equation}
On the other hand, as $(\phi_n)$ is a minimizing sequence and $F^*$ is coercive,
$$
\frac{1}{C} \|\alpha_n\|_p^p -\int_{\T^d} \phi_n(0)m_0 \leq \int_0^T \int_{\T^d} F^*(x,\alpha_n)-\int_{\T^d} \phi_n(0)m_0+C \leq C.
$$
Adding the previous inequalities, we get
$$
\frac{1}{C} \|\alpha_n\|_p^p+\frac{1}{C} \int_0^T\int_{\T^d} |D\phi_n|^r \leq C_0\|\alpha_n\|_p+C,
$$
so that $(\alpha_n)$ is bounded in $L^p((0,T)\times \mathbb{T}^d)$ while $(D\phi_n)$ is bounded in $L^r$.\\
\noindent{\bf Step 2:} We show here that $(\phi_n,\alpha_n)$ has a limit. As $(\alpha_n)$ is bounded in $L^p$ and $(\phi_n)$ is uniformly bounded below thanks to \eqref{unifboundbelow}, Theorem \ref{thm:estiHJ} implies that $(\phi_n)$ is bounded in $L^\gamma$. So we can assume with loss of generality that $\alpha_n \rightharpoonup \bar\alpha$ in $L^p$, $\phi_n \rightharpoonup \bar \phi$ in $L^\gamma$ and $D\phi_n \rightharpoonup D\bar\phi$ in $L^r$ where, in view of the convexity of $H$, the pair $(\bar\phi,\bar\alpha)$ belongs to ${\mathcal K}$. \\
\noindent{\bf Step 3:} We now prove that $(\bar\phi,\bar\alpha)$ is a minimizer. By weak lower semicontinuity arguments, we have
$$
\liminf_n \int_0^T\int_{\T^d} F^*(x,\alpha_n)\geq \int_0^T\int_{\T^d} F^*(x,\bar\alpha).
$$
Let $\displaystyle \zeta_n(t)= \int_{\T^d} m_0\phi_n(t)$ and $\displaystyle \bar \zeta(t)= \int_{\T^d} m_0\bar \phi(t)$. Then $(\zeta_n)$ converges weak* to $\bar\zeta$ in $L^\infty$ thanks to Theorem \ref{thm:estiHJ}. As
$$
- \frac{d}{dt} \zeta_n(t)+ \int_{\mathbb{T}^d} \lg Dm_0,AD\phi_n(t)\rg +m_0 (\partial_iA_{ij}\partial_j\phi_n+ H(x,D\phi_n) -\alpha_n ) \leq 0,
$$
we also have by coercivity of $H$ and thanks to the bound on $(\alpha_n)$:
$$
\zeta_n(0) -Ct^{\frac1{p'}} \leq \zeta_n(t) \qquad \forall t\in [0,T].
$$
Letting $n\to+\infty$:
$$
\limsup_n \zeta_n(0) -Ct^{\frac1{p'}} \leq\bar \zeta(t) \qquad a.e. \ t\in [0,T],
$$
so that $\displaystyle \limsup_n \zeta_n(0) \leq \int_{\T^d} m_0\bar\phi(0)$. Hence
$$
\liminf_n \int_0^T\int_{\T^d} F^*(x,\alpha_n)-\int_{\T^d} m_0\phi_n(0) \geq \int_0^T\int_{\T^d} F^*(x,\bar \alpha)-\int_{\T^d} m_0\bar\phi(0)
$$
and $(\bar\phi,\bar\alpha)$ is a minimum.
\end{proof}
\begin{Remark}\label{rem:CS}{\rm If $r>2$ and $p>1+d/r$, then by \cite{CS} the sequence $(\phi_n)$ built at the beginning of the proof is uniformly H\"{o}lder continuous. Hence so is $\phi$.
}\end{Remark}
\section{Existence and uniqueness of a solution for the MFG system}\label{sec:exuniq}
In this section we show that the MFG system \eqref{MFG} has a unique weak solution and prove the stability of this solution with respect to the data.
\subsection{Definition of weak solutions}
The variational method described above provides weak solutions for the MFG system. By a weak solution, we mean the following:
\begin{Definition}\label{def:weaksolMFG} We say that a pair $(\phi,m)\in L^\gamma((0,T)\times\mathbb{T}^d) \times L^q((0,T)\times\mathbb{T}^d)$ is a weak solution to \eqref{MFG} if
\begin{itemize}
\item[(i)] the following integrability conditions hold:
$$\displaystyle D\phi\in L^r, \; \displaystyle mH^*(\cdot, D_pH(\cdot,D\phi))\in L^1
\quad{\rm and }\quad m D_pH(\cdot,D\phi))\in L^1 .
$$
\item[(ii)] Equation \eqref{MFG}-(i) holds in the following sense: inequality
\begin{equation}\label{eq:distrib}
\displaystyle \quad -\partial_t \phi-\partial_i (A_{ij}(x)\partial_{j}\phi) + (\partial_i A_{ij})\partial_j \phi+H(x,D\phi)\leq f(x,m) \quad {\rm in }\; (0,T)\times \mathbb{T}^d,
\end{equation}
with $\phi(T,\cdot)\leq \phi_T$, holds in the sense of distributions,
\item[(iii)] Equation \eqref{MFG}-(ii) holds:
\begin{equation}\label{eqcontdef}
\displaystyle \quad \partial_t m-\partial_{ij} (A_{ij}(x)m)-{\rm div}(m D_pH(x,D\phi)))= 0\ {\rm in }\; (0,T)\times \mathbb{T}^d, \quad m(0)=m_0
\end{equation}
in the sense of distributions,
\item[(iv)] The following equality holds:
\begin{equation}\label{defcondsup} \begin{array}{l}
\displaystyle \int_0^T\int_{\T^d} m(t,x)\left(f(x,m(t,x))+ H^*(x, D_pH(x,D\phi)(t,x)) \right)dxdt\\
\displaystyle\qquad \qquad \qquad \qquad \qquad \qquad \qquad + \int_{\T^d} m(T,x)\phi_T(x)-m_0(x)\phi(0,x)dx=0.
\end{array}\end{equation}
\end{itemize}
\end{Definition}
Our main result is the following existence and uniqueness theorem:
\begin{Theorem}\label{theo:mainex} There exists a weak solution $( \phi, m)$ to the MFG system \eqref{MFG}. Moreover this solution is unique in the following sense: if $( \phi, m)$ and $(\phi', m')$ are two solutions, then $m= m'$ a.e. and $ \phi=\phi'$ in $\{ m>0\}$.
Finally, there exists a solution which is bounded below by a constant depending on $\|\phi_T\|_{C^2}$, on $\|A_{ij}\|_{C^0}$ and on $\|H(\cdot, D\phi_T)\|_\infty$.
\end{Theorem}
\begin{Remark}{\rm Under the assumptions of Remark \ref{rem:CS}, i.e., if $r>2$ and $p>1+d/r$, the $\phi$-component of the solution is locally H\"{o}lder continuous.
}\end{Remark}
\subsection{Existence of a weak solution}
The first step towards the proof of Theorem \ref{theo:mainex} consists in showing a one-to-one equivalence between solutions of the MFG system and the two optimizations problems \eqref{Pb:mw2} and \eqref{PB:dual-relaxed}.
\begin{Theorem}\label{theo:main}
Let $(\bar m,\bar w)\in \mathcal K_1$ be a minimizer of \eqref{Pb:mw2} and $(\bar \phi,\bar \alpha)\in \mathcal K$ be a minimizer of \eqref{PB:dual-relaxed}. Then $(\bar \phi, \bar m)$ is a weak solution of the mean field games system \eqref{MFG} and $\bar w= -\bar mD_pH(\cdot,D\bar \phi)$ while $\bar\alpha= f(\cdot,\bar m)$ a.e..
Conversely, any weak solution $(\bar \phi,\bar m)$ of \eqref{MFG} is such that the pair $(\bar m,-\bar mD_pH(\cdot,D\bar\phi))$ is the minimizer of \eqref{Pb:mw2} while $(\bar \phi, f(\cdot,\bar m))$ is a minimizer of \eqref{PB:dual-relaxed}.
\end{Theorem}
\begin{proof}
Let $(\bar m,\bar w) \in \s{K}_1$ be a minimizer of Problem \eqref{Pb:mw2} and $(\bar \phi,\bar \alpha) \in \s{K}$ be a minimizer of Problem \eqref{PB:dual-relaxed}.
Due to Lemma \ref{Lem:dualite} and Proposition \ref{prop:valeursegales}, we have
\begin{equation*}
\int_0^T \int_{\bb{T}^d} F^*(x,\bar \alpha) + F(x,\bar m) + \bar mH^*\left(x,-\frac{\bar w}{\bar m}\right) dxdt + \int_{\bb{T}^d} \phi_T \bar m(T) - \bar \phi(0)m_0 dx = 0.
\end{equation*}
We show that $\bar \alpha = f(x,\bar m)$. Indeed, by convexity of $F$,
\begin{equation} \label{eq:F_Fstar_geq}
F^*(x,\bar \alpha(t,x)) + F(x,\bar m(t,x)) - \bar \alpha(t,x)\bar m(t,x) \geq 0,
\end{equation}
hence
$$
\int_0^T \int_{\bb{T}^d} \bar \alpha(t,x)\bar m(t,x)+ \bar mH^*\left(x,-\frac{\bar w}{\bar m}\right) dxdt + \int_{\bb{T}^d} \phi_T \bar m(T) -\bar \phi(0)m_0 dx \leq 0.
$$
Thanks to Lemma \ref{lem:phialphamw}, the above inequality is in fact an equality, $\bar w=-\bar mD_pH(\cdot,D\bar \phi)$ a.e. and the equality holds almost everywhere in Equation (\ref{eq:F_Fstar_geq}). Therefore,
\begin{equation} \label{eq:alpha_equals_f}
\bar \alpha(t,x) = f(x,\bar m(t,x))
\end{equation}
almost everywhere and (\ref{defcondsup}) holds:
\begin{equation*} \label{eq:ibp_equality}
\int_0^T \int_{\bb{T}^d} f \bar m + \bar mH^*\left(x,-\frac{\bar w}{\bar m}\right) dxdt + \int_{\bb{T}^d} \phi_T \bar m(T) - \bar \phi(0)m_0 dx = 0.
\end{equation*}
In particular $\bar mH^*(\cdot, D_pH(\cdot,D\bar \phi))\in L^1$.
Moreover, since $(\bar \phi,\bar \alpha) \in \s{K}$ and Equation (\ref{eq:alpha_equals_f}) holds, we have $-\partial_t\bar \phi -A_{ij}\partial_{ij}\bar \phi+ H(x,D\bar \phi) \leq f(x,\bar m)$ in the sense of distributions and $\bar \phi(T)\leq \phi_T$.
Furthermore, since $(\bar \phi,\bar \alpha) \in \s{K}$ and $\bar w=-\bar mD_pH(\cdot,D\bar \phi)$, we have that $\bar mD_pH(\cdot,D\bar \phi)\in L^1$ and (\ref{eqcontdef}) holds in the sense of distributions.
Therefore $(\bar \phi, \bar m)$ is a solution in the sense of Definition \ref{def:weaksolMFG}.
Suppose now that $(\bar \phi,\bar m)$ is a weak solution of (\ref{MFG}) as in Definition \ref{def:weaksolMFG}. Set $\bar w = -\bar mD_p H(\cdot,D\bar \phi)$ and $\bar \alpha(t,x) = f(x,\bar m(t,x))$. By definition of weak solution $\bar w,\bar \alpha \in L^1$, $\bar m\in L^q$ and $\bar \phi\in L^\gamma$. Moreover, since $f$ is increasing in $\bar m$ and $\bar m\in L^q$, the growth condition (\ref{HypGrowthF}) implies that $\bar \alpha \in L^p$.
Therefore $(\bar m,\bar w) \in \s{K}_1$ and $(\bar \phi,\bar \alpha) \in \s{K}$.
It remains to show that $(\bar \phi,\bar \alpha)$ minimizes $\s{A}$ and $(\bar m,\bar w)$ minimizes $\s{B}$.
Let $(\bar \phi',\bar \alpha') \in \s{K}$. By the convexity of $F$ in the second variable, we have
\begin{align*}
\s{A}(\bar \phi',\bar \alpha') &= \int_0^T \int_{\bb{T}^d} F^*(x,\bar \alpha'(t,x))dx dt - \int_{\bb{T}^d} \bar \phi'(0,x)m_0(x)dx\\
&\geq \int_0^T \int_{\bb{T}^d} F^*(x,\bar \alpha(t,x)) + \partial_\alpha F^*(x,\bar \alpha(t,x))(\bar \alpha'(t,x) - \bar \alpha(t,x)) dx dt - \int_{\bb{T}^d}\bar \phi'(0,x)m_0(x)dx\\
&\geq \int_0^T \int_{\bb{T}^d} F^*(x,\bar \alpha(t,x)) +\bar m(t,x)(\bar \alpha'(t,x) - \bar \alpha(t,x)) dx dt - \int_{\bb{T}^d} \bar \phi'(0,x)m_0(x)dx,\\
&\geq \s{A}(\bar \phi,\bar \alpha) + \int_0^T \int_{\bb{T}^d} \bar m(t,x)(\bar \alpha'(t,x) -\bar \alpha(t,x)) dx dt + \int_{\bb{T}^d}(\bar \phi(0,x)-\bar \phi'(0,x))m_0(x)dx.
\end{align*}
Due to Equation (\ref{defcondsup}) and Lemma \ref{lem:phialphamw} applied to $(\bar \phi',\bar \alpha')$ and $(\bar m,\bar w)$ we have
\begin{align*}
&\int_0^T \int_{\bb{T}^d} \bar m(t,x)(\bar \alpha'(t,x) - \bar \alpha(t,x)) dx dt + \int_{\bb{T}^d}(\bar \phi(0,x)- \bar \phi'(0,x))m_0(x)dx=\\ &\int_0^T \int_{\bb{T}^d} \bar m(t,x)\bar \alpha'(t,x) +\bar m(t,x) H^*(x,-\frac{\bar w(t,x)}{\bar m(t,x)}) dx dt + \int_{\bb{T}^d} \phi_T(x)\bar m(T,x)-\bar \phi'(0,x)m_0(x)dx\geq 0.
\end{align*}
Hence,
$$
\s{A}(\bar \phi',\bar \alpha') \geq \s{A}(\bar \phi,\bar \alpha),
$$
and $(\bar \phi,\bar \alpha)$ is a minimizer of $\s{A}$.
The argument for $(\bar m,\bar w)$ is similar. Let $(\bar m',\bar w')$ minimize $\s{B}$. Then because $F$ is convex in the second variable, we have
\begin{align*}
\s{B}(\bar m',\bar w') &= \int_{\bb{T}^d} \phi_T \bar m'(T) + \iint \bar m'H^*\left(x,-\frac{\bar w'}{\bar m'}\right) + F(x,\bar m')\\
&\geq \int_{\bb{T}^d} \phi_T \bar m'(T) + \iint \bar m'H^*\left(x,-\frac{\bar w'}{\bar m'}\right) + F(x,\bar m) + f(x,\bar m)(\bar m'-\bar m)\\
&= \int_{\bb{T}^d} \phi_T \bar m'(T) + \iint \bar m'H^*\left(x,-\frac{\bar w'}{\bar m'}\right) + F(x,\bar m) + \bar \alpha(\bar m'-\bar m)\\
&=\s{B}(\bar m,\bar w)+ \int_{\bb{T}^d} \phi_T \bar m'(T)-m_0\bar \phi(0) + \iint \bar m'H^*\left(x,-\frac{\bar w'}{\bar m'}\right) + \bar \alpha \bar m'\\
&\geq \s{B}(\bar m,\bar w).
\end{align*}
Here we used Equation (\ref{defcondsup}) in the next to last line, and we applied Lemma \ref{lem:phialphamw} to $(\bar \phi,\bar \alpha)$ and $(\bar m',\bar w')$ in the last line. Therefore $(\bar m,\bar w)$ is a minimizer of $\s{B}$.
\end{proof}
\subsection{Uniqueness of the weak solution}\label{subsec:uniq}
\begin{proof}[Proof of Theorem \ref{theo:mainex} (uniqueness part)]
Let $(\bar \phi,\bar m)$ be a weak solution to \eqref{MFG}. In view of Theorem \ref{theo:main}, the pair $(\bar m,-\bar mD_pH(\cdot,D\bar \phi))$ is the minimizer of \eqref{Pb:mw2} while $(\bar \phi, f(\cdot,\bar m))$ is a solution of \eqref{PB:dual-relaxed}. In particular, $\bar m$ is unique because of the uniqueness of the solution of \eqref{Pb:mw2}.
Let now $( \phi_1,\bar m)$ and $( \phi_2,\bar m)$ be two weak solutions of \eqref{MFG}, and set $\bar \alpha = f(\cdot,\bar m)$.
Let $\bar \phi = \phi_1\vee\phi_2$.
Assume for now that $\bar \phi$ is a subsolution of (\ref{ineq:phi}) in the sense of distributions.
Then $(\bar \phi,\bar \alpha) \in \s{K}$, and so because $-\int \bar \phi(0)m_0 \leq - \int \phi_1(0)m_0$ we have that $(\bar \phi,\bar \alpha)$ is also a solution of \eqref{PB:dual-relaxed}.
Indeed, one deduces from Lemma \ref{lem:phialphamw} that for a.e. $t\in [0,T]$, $(\bar \phi,\bar \alpha)$ and $(\phi_1,\bar \alpha)$, are both minimizers of the problem
$$
\inf_{(\phi, \alpha)\in {\mathcal K}} \int_t^T\int_{\T^d} F^*(x,\alpha)-\int_{\T^d} m(t)\phi(t).
$$
In particular, $\displaystyle \int_{\T^d}\bar m(t)\bar \phi(t) = \int_{\T^d} \bar m(t)\phi_1(t)$. As $\phi_1 \leq\bar \phi$, this implies that $\phi_1 =\bar \phi$ a.e. in $\{\bar m>0\}$.
The same argument, replacing $\phi_1$ with $\phi_2$, shows that $\phi_2 =\bar \phi$ a.e. in $\{\bar m>0\}$, and uniqueness is proved.
The main work to be shown is that $\bar \phi = \phi_1\vee \phi_2$ is indeed a subsolution of (\ref{ineq:phi}) in the sense of distributions, i.e.
\begin{equation}
\label{eq:max-subsolution}
-\int_{\mathbb{T}^d}\zeta(T)\phi_T + \int_0^T\int_{\mathbb{T}^d}\bar \phi\partial_t\zeta+ \lg D\zeta, AD\bar \phi\rg + \zeta(\partial_iA_{ij}\partial_j\bar \phi +H(x,D\bar \phi)) \leq \int_0^T\int_{\mathbb{T}^d} \bar \alpha \zeta
\end{equation}
for any nonnegative smooth map $\zeta$ with support in $(0,T]\times \mathbb{T}^d$.
Let $\epsilon > 0$.
Introduce the following translation and extension of $(\phi_k,\bar \alpha), k=1,2$:
\begin{equation}
\tilde{\phi}_k(t,x) = \left\{
\begin{array}{ll}
\phi_k(t+2\epsilon,x) & \text{if}~t \in [-2\epsilon,T-2\epsilon)\\
\phi_T(x) & \text{if}~t\in [T-2\epsilon,T+2\epsilon]
\end{array}\right.
\end{equation}
and
\begin{equation}
\tilde{\alpha}(t,x) = \left\{
\begin{array}{ll}
\bar \alpha(t+2\epsilon,x) & \text{if}~t \in [-2\epsilon,T-2\epsilon)\\
\lambda & \text{if}~t\in [T-2\epsilon,T+2\epsilon]
\end{array}\right.
\end{equation}
where $\lambda = \max_x H(x,D\phi_T(x)) + A_{ij}(x) \partial_{ij} \phi_T(x)$.
Then we have that
\begin{equation}
-\partial_t \tilde{\phi}_k - A_{ij}\partial_{ij}\tilde{\phi}_k + H(x,D\tilde{\phi}_k) \leq \tilde{\alpha}
\end{equation}
in the sense of distributions on $(-2\epsilon,T+2\epsilon) \times \bb{T}^d$.
For now we will fix a smooth vector field $\psi$ on $[0,T] \times \bb{T}^d$.
Notice that
\begin{equation}
-\partial_t \tilde{\phi}_k - A_{ij}\partial_{ij}\tilde{\phi}_k + \psi \cdot D\tilde{\phi}_k \leq \tilde{\alpha} + H^*(x,\psi)
\end{equation}
in the sense of distributions on $(-2\epsilon,T+2\epsilon) \times \bb{T}^d$.
Let $\xi^1$ be a smooth convolution kernel in $\bb{R}^{d+1}$ with support in the unit ball, with $\xi^1 \geq 0$ and $\int \xi^1 = 1$.
Then define the standard mollifier sequence $\xi^\epsilon(t,x) = \epsilon^{-d-1}\xi^1((t,x)/\epsilon)$.
Set $\phi_k^\epsilon = \xi^\epsilon \star \tilde{\phi}_k$ and $\alpha^\epsilon = \xi^\epsilon \star \tilde \alpha$.
By taking the convolution we have, in a pointwise sense,
\begin{equation} \label{eq:subsol+commutators}
-\partial_t \phi_k^\epsilon - A_{ij}\partial_{ij}\phi_k^\epsilon + \psi \cdot D\phi_k^\epsilon \leq \alpha^\epsilon + \xi^\epsilon \star H^*(\cdot,\psi) + R_\epsilon^k - S_\epsilon^k
\end{equation}
on $[0,T] \times \bb{T}^d$, where
\begin{equation}
R_\epsilon^k := [\xi^\epsilon,A_{ij}\partial_{j}](\partial_i \tilde{\phi}_k), ~~ S_\epsilon^k := [\xi^\epsilon,\psi ](D\tilde{\phi}_k).
\end{equation}
Here we use the same commutator notation as in (\ref{eq:commutators}).
Invoking \cite[Lemma II.1]{DL89}, we have that $R_\epsilon^k$ and $S_\epsilon^k$, $k=1,2$ are smooth functions which converge to zero in $L^{r}$, since $A_{ij} \in W^{1,\infty}$ is given and $\psi$ may also be chosen in $W^{1,\infty}$.
Define $R_\epsilon := \max\{R_\epsilon^1 - S_\epsilon^1,R_\epsilon^2 - S_\epsilon^2\}$.
This, too, converges to zero in $L^r$.
Moreover, for $k=1,2$
\begin{equation} \label{eq:subsol+error}
-\partial_t \phi_k^\epsilon - A_{ij}\partial_{ij}\phi_k^\epsilon + \psi \cdot D\phi_k^\epsilon \leq \alpha^\epsilon + \xi^\epsilon \star H^*(\cdot,\psi) + R_\epsilon
\end{equation}
holds in a pointwise sense, hence also in a viscosity sense.
By standard results, (\ref{eq:subsol+error}) holds also for $\phi^\epsilon := \phi^\epsilon_1\vee \phi^\epsilon_2$ in a viscosity sense.
The result of \cite{Ish95} implies that it also holds in the sense of distributions, that is, for any smooth map $\zeta$ with support in $(0,T] \times \bb{T}^d$ we have
\begin{multline} \label{eq:subsolindist+error}
-\int_{\mathbb{T}^d}\zeta(T)\phi^\epsilon(T) + \int_0^T\int_{\mathbb{T}^d} \phi^\epsilon\partial_t\zeta+ \lg D\zeta, AD\phi^\epsilon\rg + \zeta(\partial_iA_{ij}\partial_j\phi^\epsilon +D\phi^\epsilon \cdot \psi) \\
\leq \int_0^T\int_{\mathbb{T}^d} \zeta(\alpha^\epsilon + \xi^\epsilon \star H^*(\cdot,\psi) + R_\epsilon).
\end{multline}
By construction, $\phi^\epsilon(T) = \phi_T$ for all $\epsilon > 0$.
Observe that $\phi^\epsilon \to\bar \phi$ in $L^\gamma$ and $D\phi^\epsilon \to D\bar \phi$ in $L^r$, as these sequences are only slight adaptations of classical convolutions of $\bar \phi$ and $D\bar \phi$.
Finally, note that $\alpha^\epsilon \to\bar \alpha$ in $L^p$, while $\xi^\epsilon \star H^*(\cdot,\psi) \to H^*(\cdot,\psi)$ uniformly.
Letting $\epsilon \to 0+$, we are left with
\begin{equation} \label{eq:0-estimate}
-\int_{\mathbb{T}^d}\zeta(T)\phi_T + \int_0^T\int_{\mathbb{T}^d} \bar \phi\partial_t\zeta+ \lg D\zeta, AD\bar \phi\rg + \zeta(\partial_iA_{ij}\partial_j\bar \phi +\psi \cdot D\bar \phi)
\leq \int_0^T\int_{\mathbb{T}^d} \zeta(\bar \alpha + H^*(\cdot,\psi)).
\end{equation}
Now since $\psi$ is an arbitrary smooth vector field, we may take a sequence that approximates $\partial_p H(x,D\bar \phi)$ in $L^{r'}$.
By the convexity of $H(x,\cdot)$ this yields (\ref{eq:max-subsolution}), as desired.
\end{proof}
\subsection{Stability}
We now consider the stability of solutions with respect to the data $A$, $H$ and $f$ and the data $m_0$ and $\phi_T$. More precisely, assume that $(A^n)$, $(H^n)$, $(f^n)$ $m_0^n$ and $\phi^n_T$ satisfy conditions (H1)$\dots$(H4) uniformly with respect to $n$ and converge to $A$, $H$, $f$, $m_0$ and $\phi_T$ locally uniformly.
\begin{Theorem}\label{thm:stab} Let $(\phi^n,m^n)$ be a weak solution of \eqref{MFG} associated with $A^n$, $H^n$, $f^n$ and with the initial and terminal conditions $m_0^n$ and $\phi_T^n$. Assume also that the sequence $\phi^n$ is uniformly bounded below. Then $(m^n)$ converges strongly to $m$ in $L^q$ while $\phi^n$ converges weakly and up to a subsequence to a map $\bar \phi$ in $L^\gamma$, where the pair $(\bar \phi,\bar m)$ is a weak solution to \eqref{MFG}.
\end{Theorem}
Note that the existence of a solution $(\phi^n,m^n)$, such that $\phi^n$ is bounded by below, is ensured by Theorem \ref{theo:mainex}.
The result is a simple consequence of Theorem \ref{theo:main} and of the $\Gamma-$convergence of the corresponding variational problems.
\begin{proof} Let us set $w^n=-m^nD_pH_n(\cdot,D\phi^n))$ and $\alpha^n= f(\cdot,m^n)$. According to the second part of Theorem \ref{theo:main}, the pair $(m^n,w^n)$ is a minimizer of problem \eqref{Pb:mw2} associated with $A^n$, $H^n$, $f^n$, $m_0^n$ and $\phi_T^n$, while the pair $(\phi^n,\alpha^n)$ is a minimizer of problem \eqref{PB:dual-relaxed} associated with the same data. Using the estimates established for the proof of Proposition \ref{prop:valeursegales}, we have
\begin{equation}\label{boundmnwn}
\|m^n\|_{L^q}+ \|w^n\|_{L^{\frac{r'q}{r'+q-1}}} \leq C.
\end{equation}
By lower semi-continuity of the functional $\mathcal B$, $(m^n,w^n)$ converge weakly up to a subsequence to to the minimum $(\bar m, \bar w)$ of the problem \eqref{Pb:mw2} associated with $A$, $H$, $f$, $m_0$ and $\phi_T$. The limit problem being strictly convex, the convergence actually holds
strongly in $L^q\times L^{\frac{r'q}{r'+q-1}}$.
Then the growth condition \eqref{Hypf} on $f$ implies that the sequence $(\alpha^n=f^n(\cdot,m^n))$ converges in $L^p$ to $\bar \alpha:= f(\cdot,\bar m)$.
As $(\phi^n)$ is uniformly bounded below, Theorem \ref{thm:estiHJ} implies that
$$
\|\phi^n\|_{L^\infty((0,T), L^{\eta}(\mathbb{T}^d))}+ \|\phi^n\|_{L^\gamma((0,T)\times \mathbb{T}^d)}\leq C.
$$
The end of the proof follows closely the argument in Proposition \ref{Prop:existence}: $(D\phi^n)$ is bounded in $L^r$, so that, up to a subsequence, $(\phi^n)$ converges weakly to some $\bar \phi$ in $L^\gamma$ while $D\phi^n$ converges weakly to $D\phi$ in $L^r$ where $(\bar \phi,\bar \alpha)$ belongs to ${\mathcal K}$. Moreover $(\bar \phi,\bar \alpha)$ is a minimizer of the relaxed problem \eqref{PB:dual-relaxed}. Theorem \ref{theo:main} then states that the pair $(\bar \phi,\bar m)$ is a solution to the MFG problem \eqref{MFG}.
\end{proof}
|
\section{Introduction}
The black hole perturbation theory has been a powerful tool
to investigate the stability of the
black hole, the quasi-normal modes,
and the gravitational waves produced by matters such like
compact starts orbiting around the hole, and so on.
For the Schwarzschild case, the first order metric perturbation is described
by the Regge--Wheeler--Zerilli formalism \cite{RW1957, Z1970},
which relies on the spherical symmetry of the black hole space-time.
The Regge--Wheeler and the Zerilli equation are the single, decoupled equation for
the odd and even {parity} modes, respectively, and the master equations are reduced to
radial ordinary differential equations by using the Fourier-harmonic expansion.
On the other hand, for the Kerr case, it is well-known that there is no such a formalism
for the metric perturbation.
Instead, the perturbation of the Weyl scalars, $\psi_0$ and $\psi_4$, are described
by the Teukolksy equation with the spin-weight $s=\pm 2$.
One method to compute the metric perturbation of Kerr space-time is to solve the coupled
partial differential equations numerically.
The other method is to construct the metric perturbation from
the perturbation of $\psi_0$ and $\psi_4$ obtained from the Teukolsky equation.
Such a method was {proposed} first
by Chrzanowski \cite{c} and Cohen and Kegeles \cite{kc, ck}
(See also \cite{wald1978, Stewart1979}),
{and thus} is called the CCK formalism.
In this method, a radiation gauge is used to calculate the metric perturbation.
After these works,
however, there were very little development of the CCK formalism for a long time.
New {developments were started} about a decade ago
by Lousto and Whiting \cite{LoustoWhiting} and Ori \cite{ori03}.
These were motivated by the necessity to compute
the gravitational self-force on the point particle orbiting around a Kerr black hole.
Such situations are called EMRI (extreme mass ratio inspiral), and
are one of the most important sources
{of the gravitational wave} for the future space laser interferometers
such as eLISA \cite{eLISA}, DECIGO \cite{SNK, DECIGO} and BBO \cite{BBO}.
A first explicit computation of the metric perturbation by using the CCK formalism
was done by Yunes and Gonz$\acute{\rm a}$lez \cite{YunesGonzalez}
in which the vacuum perturbation was considered.
Keidl, Friedman, and Wiseman \cite{kfw} were the first to find
the explicit metric perturbation {produced} by a point particle, using the CCK formalism.
They considered a system which consists of a Schwarzschild black hole and a static point mass, as a toy model.
The metric perturbation is obtained straightforwardly for the multipole modes of $l\geq 2$.
They obtained lower modes of $l=0, 1$ by considering the regularity of the metric.
A singularity, however, remained along a radial line which connect
the position of the particle and either the infinity or the black hole horizon.
The presence of the singularity
was previously discussed by Wald \cite{wald1973} and by Barack and Ori \cite{BarackOri2001}.
Keidl, Shah, Friedman, Kim and Price \cite{ksf+10, skf+11, sfk12}
{further developed}
the formalism to calculate the self-force by using the CCK formalism.
In \cite{sfk12}, they reported the numerical corrections of gauge invariants
of a particle in circular orbit {around} a Kerr black hole.
For the calculation of the gravitational self{-}force
on the particle,
it is important to complete the metric perturbation by adding the lower modes in an appropriate gauge.
The $l\geq 2$ modes are calculated in a radiation gauge,
and the effects of lower modes are added in, what they call, the Kerr gauge.
Recently, Pound, Merlin, and Barack \cite{pmb} discussed prescriptions for calculating the self-force
from completed metric perturbations.
With this prescription,
once we obtain the metric perturbation which is constructed using {a} radiation gauge
and completed
with lower modes appropriately, it is possible to transform its gauge into a local Lorenz gauge.
The regularized self-force can then be calculated by using the standard mode-sum method.
In this paper, we consider the metric perturbation of a rotating circular mass ring around a Schwarzschild black hole,
in order to understand the problems {in constructing}
the metric perturbation by using the CCK formalism.
Especially, we discuss the problem of the completion of the metric perturbation with lower multipole modes.
Of course, this is a first step toward the calculation of the metric perturbation produced by a orbiting particle.
But this problem is simpler than that of an orbiting particle,
since the ring is circular and rotates with a constant angular velocity,
and the problem becomes stationary and axisymmetric.
Nevertheless, this problem is more complicated than \cite{kfw}
in that
both the mass and angular momentum perturbation are involved.
This paper is organized as follows.
The first step is to obtain the perturbed Weyl scalars $\psi_0$ and $\psi_4$
by solving the Teukolsky equation which is discussed in Section \ref{section:TeuEq}.
Next in Section \ref{section:CCK}, we describe the CCK formalism in a general form.
In Section \ref{section:IRG}, the Hertz potential is obtained from $\psi_0$ and $\psi_4$.
In Section \ref{section:PsiP}, we briefly discuss the gravitational fields computed
from the Hertz potential which contains only $l\geq 2$ modes, and show the presence of the
singularities in the gravitational fields.
In Section \ref{section:HertzH}, we obtain the Hertz potential
of $l=0, 1$ modes by considering the continuity of the gravitational field,
and obtain the metric perturbation from the completed Hertz potential.
Section \ref{section:summary} is devoted to summary and discussion.
\section{Solutions of the Teukolsky equation}
\label{section:TeuEq}
In this section we analytically derive $\psi_0$ and $\psi_4$.
The details of the derivation {are}
given in Appendix \ref{section:NP} and \ref{section:solTeukolsky}.
Here, we only give the outline and the main results which are used in the
subsequent sections.
The Schwarzschild metric is given as
\begin{equation}
\begin{split}
\dev s^2 = -\frac{\Delta}{r^2}\dev t^2+\frac{r^2}{\Delta}\dev r^2+r^2(\dev\theta^2+\sin^2\theta \dev\phi^2)~,
\end{split}
\label{eq:schwametric}
\end{equation}
where $\Delta =r^2-2Mr$~.
Five complex Weyl scalars are defined as
\begin{equation}
\begin{split}
&\Psi_0 = +C_{abcd}l^a m^b l^c m^d~,\\
&\Psi_1 = +C_{abcd}l^a {n}^b l^c {m}^d~,\\
&\Psi_2 = +C_{abcd}l^a m^b \overline{m}^c n^d~,\\
&\Psi_3 = +C_{abcd}l^a n^b \overline{m}^c n^d~,\\
&\Psi_4 = +C_{abcd}n^a \overline{m}^b n^c \overline{m}^d~,
\label{weylscalar-def}
\end{split}
\end{equation}
{where $C_{abcd}$ is the Weyl tensor, and}
\sout{Here,}
$l^a, {n}^b, {m}^d$ are the Kinnersley tetrad defined in Appendix \ref{section:NP}.
The overline $\overline{m}$ denotes the complex conjugate of $m$.
Note that we adopt the $-$$+$$+$$+$ signature which is different from that of
Newman and Penrose \cite{np62} and Teukolsky \cite{teu}.
Because of it, although
the sign of above Weyl scalars are opposite from those by Newman and Penrose \cite{np62}
and Teukolsky \cite{teu}, the Teukolsky equations are left unchanged.
In the case of Schwarzschild metric, nonzero Weyl scalar is $\Psi_2$.
\begin{equation}
\Psi_2
= -\frac{M}{r^3}~.
\end{equation}
The corresponding perturbed Weyl scalars are denoted by $\psi_0,~\psi_1,\ldots,~\psi_4$.
We consider the perturbation of the Schwarzschild metric induced by a rotating ring
which is composed by a set of point masses in a circular, geodesic orbit
on the equatorial plane.
The energy-momentum tensor of the ring is written as
\begin{equation}
\begin{split}
T^{ab} =& \sekibun{}{}{\phi'}\frac{\mass u^a u^b}{u^tr_0{}^2}\delta(r-r_0)\delta(\cos\theta)\delta(\phi-\phi')\\
=& \frac{\mass u^a u^b}{u^tr_0{}^2}\delta(r-r_0)\delta(\cos\theta)~,
\label{Tring}
\end{split}
\end{equation}
where $r_0$ is the radius of the ring, and
$u^a = u^t\left( (\partial_t)^a+\Omega(\partial_\phi)^a \right)$
{is the four-velocity of the ring}.
The angular velocity $\Omega$ and $u^t$ {are} given as
\begin{equation}
\Omega=\sqrt{\frac{M}{r_0{}^3}}~,~~~~u^t = \sqrt{\frac{r_0}{r_0-3M}}~.
\end{equation}
The rest mass of the ring becomes $2\pi \mass ~(\ll M)$.
Since our {perturbed} space-time is independent from $t$ and $\phi$,
it is sufficient to consider the case of $\omega=0$ and the $m=0$ mode of
the spin-weighted spherical harmonics ${}_sY_{lm}(\theta,\phi)$.
We expand $\psi_0$ as
\begin{equation}
\begin{split}
&\psi_0(r,\theta) = \sum_{l=2}^{\infty}R^{(2)}_l(r)~{}_{2}Y_{l}(\theta)~.
\label{T2l0}
\end{split}
\end{equation}
The Teukolsky equation for $\psi_0$ is given as
\begin{equation}
\begin{split}
\left[
\frac{1}{r^2\Delta^2}\bibun{}{r}{} \left(\Delta^3\bibun{}{r}{} \right)-\frac{(l-2)(l+3)}{r^2}
\right]
R^{(2)}_l = -4\pi T^{(2)}_l
~{.}
\label{dokei0}
\end{split}
\end{equation}
We also expand $\psi_4$ as
\begin{equation}
\begin{split}
& \rho^{-4}\psi_4(r,\theta) = \sum_{l=2}^{\infty}R^{(-2)}_l(r)~{}_{-2}Y_{l}(\theta)~.
\label{T-2l4}
\end{split}
\end{equation}
The Teukolsky equation for $\psi_4$ is given as
\begin{equation}
\begin{split}
\left[
\frac{\Delta^2}{r^2}\bibun{}{r}{} \left(\frac{1}{\Delta}\bibun{}{r}{} \right)-\frac{(l+2)(l-1)}{r^2}
\right]
R^{(-2)}_l = -4\pi T^{(-2)}_l
~{.}
\label{dokei4}
\end{split}
\end{equation}
Here we defined $_sY_l(\theta)$ as
\begin{equation}
_sY_l(\theta)\equiv{}_sY_{l0}(\theta,0)~.
\end{equation}
The source terms $T^{(2)}_l$ and $T^{(-2)}_l$ are given as
\begin{equation}
\begin{split}
T^{(2)}_l
= +2\pi & \frac{1}{r^4} {\mass u^t r_0{}^2}\frac{1}{r^2}\delta(r-r_0)
\\&\times
\sqrt{(l+2)(l-1)(l+1)l} ~{}_{0}Y_l(\pi/2)\\
-2\im \cdot 2\pi & \frac{1}{r^4} {\mass u^t\Omega r_0{}^3} \henbibun{}{r}{}\frac{1}{r} \delta(r-r_0)
\\&\times
\sqrt{(l+2)(l-1)} {}_{1}Y_l(\pi/2)\\
-2\pi & \frac{1}{r^4} {\mass u^t\Omega^2 r_0{}^4}r^2\henbibun{}{r}{}\frac{1}{r^2} \henbibun{}{r}{} \delta(r-r_0)
\\&\times
{}_{2}Y_l(\pi/2)~,
\end{split}
\end{equation}
\begin{equation}
\begin{split}
T^{(-2)}_l
= +2\pi & \frac{\Delta^2}{4r^4} {\mass u^t r_0{}^2}\frac{1}{r^2}\delta(r-r_0)
\\&\times
\sqrt{(l+2)(l-1)(l+1)l} ~{}_{0}Y_l(\pi/2)\\
+2\im \cdot 2\pi & \frac{\Delta^2}{4r^4} {\mass u^t\Omega r_0{}^3} \henbibun{}{r}{}\frac{1}{r} \delta(r-r_0)
\\&\times
\sqrt{(l+2)(l-1)} {}_{-1}Y_l(\pi/2)\\
-2\pi & \frac{\Delta^2}{4r^4} {\mass u^t\Omega^2 r_0{}^4}r^2\henbibun{}{r}{}\frac{1}{r^2} \henbibun{}{r}{} \delta(r-r_0)
\\&\times
{}_{-2}Y_l(\pi/2)~.
\end{split}
\end{equation}
A simple relation $\frac{\Delta^2}{4}T^{(2)}_l(r)=T^{(-2)}_l(r)$ holds because of the symmetries.
The Teukolsky equations for $\psi_0$ and $\psi_4$ above are solved by {using the} Green's function,
and we obtain
\begin{equation}
\begin{split}
R_l^{(2)} =
&+\frac{4}{M\Delta} \frac{4\pi^2 \mass u^t~{}_{0}Y_l(\pi/2)}{\sqrt{(l+2)(l+1)l(l-1)}}
\\&~~~ \times
\left( -\frac{\Delta_0}{2r_0{}^2} P_l^2\left(x_0^{<}\right)Q_l^2\left(x_0^{>}\right) \right)
\\%
& -\im \frac{4}{M\Delta} \frac{8\pi^2 \mass u^t \Omega r_0{}^2 ~{}_{-1}Y_l(\pi/2)}{\sqrt{(l+2)(l-1)}(l+1)l}
\\&~~~ \times
\left( -\bibun{}{r_0}{} \frac{\Delta_0}{2r_0{}^2} P_l^2\left(x_0^{<}\right)Q_l^2\left(x_0^{>}\right) \right)
\\%
& -\frac{4}{M\Delta} \frac{4\pi^2 \mass u^t \Omega^2 r_0{}^4 {}_{-2}Y_l(\pi/2)}{{(l+2)(l-1)(l+1)l}}
\\&~~~ \times
\left( -\bibun{}{r_0}{} \frac{1}{r_0{}^2}
\bibun{}{r_0}{} r_0{}^2 \frac{\Delta_0}{2r_0{}^2}
P_l^2\left(x_0^{<}\right)Q_l^2\left(x_0^{>}\right) \right) ~,
\end{split}
\label{eq:TeuR2}
\end{equation}
\begin{equation}
\begin{split}
R_l^{(-2)} =
&+\frac{\Delta}{M} \frac{4\pi^2 \mass u^t~{}_{0}Y_l(\pi/2)}{\sqrt{(l+2)(l+1)l(l-1)}}
\\&~~~ \times
\left( -\frac{\Delta_0}{2r_0{}^2} P_l^2\left(x_0^{<}\right)Q_l^2\left(x_0^{>}\right) \right)
\\%
& -\im \frac{\Delta}{M} \frac{8\pi^2 \mass u^t \Omega r_0{}^2 ~{}_{-1}Y_l(\pi/2)}{\sqrt{(l+2)(l-1)}(l+1)l}
\\&~~~ \times
\left( -\bibun{}{r_0}{} \frac{\Delta_0}{2r_0{}^2} P_l^2\left(x_0^{<}\right)Q_l^2\left(x_0^{>}\right) \right)
\\%
& -\frac{\Delta}{M}\frac{4\pi^2 \mass u^t \Omega^2 r_0{}^4 {}_{-2}Y_l(\pi/2)}{{(l+2)(l-1)(l+1)l}}
\\&~~~ \times
\left( -\bibun{}{r_0}{} \frac{1}{r_0{}^2}
\bibun{}{r_0}{} r_0{}^2 \frac{\Delta_0}{2r_0{}^2}
P_l^2\left(x_0^{<}\right)Q_l^2\left(x_0^{>}\right) \right) ~,
\end{split}
\label{eq:TeuR2m}
\end{equation}
where
\begin{equation}
{\Delta_0 \equiv r_0{}^2 -2Mr_0~,}
\end{equation}
\begin{equation}
x_0^{<}\equiv \frac{{\rm min}(r,r_0)-M}{M}~,~~~~~x_0^{>}\equiv \frac{{\rm max}(r,r_0)-M}{M}~.
\end{equation}
These two radial functions are related as $\frac{\Delta^2}{4}R_l^{(2)}(r) = R_l^{(-2)}(r)$.
With this relation, together with the fact $_2Y_l(\theta) ={}_{-2}Y_l(\theta)$,
we find that $\psi_0$ and $\psi_4$ are related in a very simple equation,
\begin{equation}
\psi_4 = \frac{\Delta^2}{4r^4} \psi_0~.
\end{equation}
Note that this relation {holds} because of the symmetries of our space-time.
We also find that because the matter is present on the equatorial plane,
and ${}_sY_l(\theta)$ is evaluated only at $\theta=\pi/2$, we have
\begin{equation*}
\Re\left(R_l^{(\pm2)}(r)\right)=0~~~{\rm for}~{\rm odd}~l
{~,}
\end{equation*}
and
\begin{equation*}
\Im\left(R_l^{(\pm2)}(r)\right)=0~~~{\rm for}~{\rm even}~l
{~.}
\end{equation*}
Therefore, the real part of $\psi_0$ and $\psi_4$ is symmetric
about the equatorial plane and the imaginary part is antisymmetric.
\begin{equation}
\begin{split}
& \Re(\psi_{0/4}(r, \pi-\theta)) = \Re(\psi_{0/4}(r, \theta))~,\\
& \Im(\psi_{0/4}(r, \pi-\theta)) = -\Im(\psi_{0/4}(r, \theta))~.
\label{oddeven}
\end{split}
\end{equation}
In Fig.~\ref{fig:TeuSolution},
We show the radial dependence of $\psi_0$ and $\psi_4$,
with fixed angular coordinate $\theta=\pi/4$.
Note that $\psi_0$ and $\psi_4$ are smooth at the sphere, $r=r_0$,
except for $\theta=\pi/2$, where the energy-momentum tensor vanishes.
\begin{figure*}[htbp]
\begin{center}
\includegraphics[width=8cm]{./Repsi0.eps}
\includegraphics[width=8cm]{./Repsi4.eps}
\\
\includegraphics[width=8cm]{./Impsi0.eps}
\includegraphics[width=8cm]{./Impsi4.eps}
\end{center}
\caption{Radial dependence of $\psi_0$ and $\psi_4$ obtained by solving the Teukolsky equation.
The real parts of $\psi_0(r,\theta=\pi/4)$ (top left) and $\psi_4(r,\theta=\pi/4)$ (top right),
and the imaginary parts of $\psi_0(r,\theta=\pi/4)$ (bottom left) and $\psi_4(r,\theta=\pi/4)$ (bottom right)
are shown.
The radius of the ring is $r_0=10M$, and {$m = M/100$}.
We see the smoothness at $r=r_0$.}
\label{fig:TeuSolution}
\end{figure*}
\section{Construction of the perturbed gravitational fields}
\label{section:Construction}
{Chrzanowski \cite{c} and Cohen and Kegeles \cite{kc} introduced a formalism }
to compute the perturbed
metric in a ``radiation gauge'' from Teukolsky valuables $\psi_0$ and $\psi_4$.
{In this section, we describe how we can use the CCK formalism to calculate the perturbed gravitational
fields produced by the rotating ring. }
\subsection{The CCK formalism}
\label{section:CCK}
{In the CCK formalism, the Hertz potential $\Psi$, which is a solution of the homogeneous Teukolsky equation,
is introduced. The perturbed metric is obtained by differentiating the Hertz potential.
In order to obtain the relation between the Hertz potential and the perturbed metric,
two kinds of gauge conditions are used.
They are called ``Ingoing Radiation Gauge'' (IRG) and ``Outgoing Radiation Gauge'' (ORG).}
The IRG is defined by the conditions $h_{ab}l^b=h^a{}_a=0$.
The perturbed metric $h_{ab}$ in IRG is related to the Hertz potential as
\begin{equation}
\begin{split}
h_{ab} &= -\big[ l_a l_b(\overline{\pmb\delta} +2\beta)(\overline{\pmb\delta} +4\beta)\overline{\Psi}\\
&~~~~~ -2l_{(a}\overline{m}_{b)}(\pmb D +\rho)(\overline{\pmb\delta} +4\beta)\overline{\Psi} \\
&~~~~~~~~~~ +\overline{m}_a\overline{m}_b(\pmb D -\rho) (\pmb D +3\rho)\overline{\Psi} \big] \\
&~~~ + \left[ {\rm c.c.} \right]~,
\label{irg}
\end{split}
\end{equation}
where $\left[ {\rm c.c.} \right]$ represents the complex conjugate of the first term.
The bold greek characters are derivative operators associated
with the tetrad defined in Appendix \ref{section:NP}.
The Hertz potential $\Psi$ in IRG satisfies the source-free Teukolsky equation
{with $s=-2$. }
\begin{equation}
(\pmb{\Delta} +\mu +2\gamma)({\pmb D} +3\rho)\Psi -3\Psi_2\Psi
= (\overline{\pmb \delta} -2\beta) ({\pmb \delta} +4\beta) \Psi~.
\label{IRGhomo1}
\end{equation}
Equivalently, this equation is written as
\begin{equation}
(\pmb{\Delta} -2\mu +2\gamma)\pmb{D} \Psi +3\rho\partial_t \Psi
= (\overline{\pmb{\delta}} -2\beta)(\pmb{\delta} +4\beta)\Psi~.
\label{IRGhomo}
\end{equation}
By using \eref{irg} and \eref{IRGhomo1},
the relations between the perturbed Weyl scalars and the Hertz potential are obtained as
\cite{ref:KFWtypo}
\begin{subequations}
\label{weyl-hertz}
\begin{align}
\psi_0 &= \frac{1}{2} \pmb D^4\overline{\Psi}~, \label{eq:psi0Hertz1} \\
\psi_1 &= \frac{1}{2} \pmb D^3(\overline{\pmb\delta} +4\beta)\overline{\Psi}~,\\
\psi_2 &= \frac{1}{2} \pmb D^2(\overline{\pmb\delta} +2\beta)(\overline{\pmb\delta} +4\beta)\overline{\Psi}~,\\
\psi_3 &= \frac{1}{2} \pmb D \overline{\pmb\delta}(\overline{\pmb\delta} +2\beta)(\overline{\pmb\delta} +4\beta)\overline{\Psi} + 3\gamma\pmb{D}\rho(\pmb{\delta} +4\beta)\Psi~, \label{eq:psi3Hertz}
\\
\psi_4 &= \frac{1}{2}(\overline{\pmb\delta} -2\beta)\overline{\pmb\delta}(\overline{\pmb\delta} +2\beta)(\overline{\pmb\delta} +4\beta)\overline{\Psi} -3\gamma\rho^2 \partial_t\Psi~. \label{eq:psi4Hertz1}
\end{align}
\end{subequations}
On the other hand, ORG is defined by the conditions $h_{ab}n^b=h^a{}_a=0$.
The perturbed metric $h_{ab}^{\rm ORG}$ {is related to the Hertz potential as}
\begin{equation}
\begin{split}
h_{ab}^{\rm ORG} &= -\Big[ n_a n_b\left(-\frac{2r^2}{\Delta}\right)^2 (\overline{\pmb\delta} +2\beta)(\overline{\pmb\delta} +4\beta)\frac{\Delta^2}{4} {\Psi}\\
&~~~~~ -2 n_{(a}\overline{m}_{b)}\left(-\frac{2r^2}{\Delta}\right)(\pmb D +\rho)(\overline{\pmb\delta} +4\beta)\frac{\Delta^2}{4} {\Psi} \\
&~~~~~~~~~~ +\overline{m}_a\overline{m}_b(\pmb D -\rho) (\pmb D +3\rho)\frac{\Delta^2}{4} {\Psi} \Big] \\
&~~~+\left[ {\rm c.c.} \right]~.
\label{org}
\end{split}
\end{equation}
The Hertz potential $\Psi$ in ORG satisfies the source-free Teukolsky equation {with $s=2$}.
\begin{equation}
\begin{split}
(\tilde{\pmb\Delta} +\mu +2\gamma)&(\tilde{\pmb{D}} +3\rho)\frac{\Delta^2}{4}\Psi
-3\Psi_2\frac{\Delta^2}{4}\Psi
\\&~~~~~
= (\pmb{\delta} -2\beta)(\overline{\pmb{\delta}} +4\beta)\frac{\Delta^2}{4}\Psi~.
\label{HertzhomoORG}
\end{split}
\end{equation}
Equivalently, this equation is written as
\begin{equation}
\begin{split}
(\tilde{\pmb\Delta} -2\mu +2\gamma)\tilde{\pmb D} \frac{\Delta^2}{4}\Psi &
- 3\rho\partial_t \frac{\Delta^2}{4}\Psi \\
&= (\pmb\delta -2\beta)(\overline{\pmb\delta} +4\beta)\frac{\Delta^2}{4}\Psi~.
\end{split}
\end{equation}
By using \eref{org} and \eref{HertzhomoORG},
the relations between the perturbed Weyl scalars and the Hertz potential are {obtained} as
\begin{subequations}
\label{weyl-hertz-org}
\begin{align}
\left(-\frac{2r^2}{\Delta}\right)^2
\psi_4
&= \frac{1}{2} \pmb D^4\frac{\Delta^2}{4}\overline{\Psi}~, \label{eq:psi0Heltz2} \\
\left(-\frac{2r^2}{\Delta}\right)
\psi_3
&= \frac{1}{2} \pmb D^3(\overline{\pmb\delta} +4\beta)\frac{\Delta^2}{4}\overline{\Psi}~,\\
\psi_2
&= \frac{1}{2} \pmb D^2(\overline{\pmb\delta} +2\beta)(\overline{\pmb\delta} +4\beta)\frac{\Delta^2}{4}\overline{\Psi}~,\\
\left(-\frac{\Delta}{2r^2}\right)
\psi_1
&= \frac{1}{2} \pmb D \overline{\pmb\delta}(\overline{\pmb\delta} +2\beta)(\overline{\pmb\delta} +4\beta)\frac{\Delta^2}{4}\overline{\Psi}
\nonumber\\&~~~~~~~~~~
+ 3\gamma\pmb{D}\rho(\pmb{\delta} +4\beta) \frac{\Delta^2}{4}\Psi~,
\\
\left(-\frac{\Delta}{2r^2}\right)^2
\psi_0
&= \frac{1}{2}({\pmb\delta} -2\beta){\pmb\delta}({\pmb\delta} +2\beta)({\pmb\delta} +4\beta)\frac{\Delta^2}{4}\overline{\Psi}
\nonumber\\&~~~~~~~~~~
+3\gamma\rho^2 \partial_t \frac{\Delta^2}{4}\Psi~. \label{eq:psi4Heltz2}
\end{align}
\end{subequations}
Whichever gauge we choose, we look for the Hertz potential that satisfies
{the relations to $\psi_0$ and $\psi_4$,
Eqs. \eref{eq:psi0Hertz1} and \eref{eq:psi4Hertz1},
or Eqs. \eref{eq:psi0Heltz2} and \eref{eq:psi4Heltz2}.}
\subsection{The Hertz potential and the metric perturbation in IRG}
\label{section:IRG}
In this paper, we use IRG to construct the perturbed gravitational fields.
From \eref{weyl-hertz}, the relations between Teukolsky valuables and the Hertz potential become
\begin{eqnarray}
&&\psi_0 =
\frac{1}{2}\left(\frac{\partial}{\partial r}\right)^4\overline{\Psi}~,
\label{psi0-hertz}
\\
&&\psi_4 =
\frac{1}{2}\frac{1}{4r^4}\sin^2\theta \left(\frac{\partial}{\partial \cos\theta}\right)^4 \sin^2\theta \overline{\Psi} ~.
\label{psi4-hertz}
\end{eqnarray}
{Here,} we used the fact that the ring and the black hole are {stationary} and axisymmetric.
Our task is to find Hertz potential which satisfies \eref{psi0-hertz}, \eref{psi4-hertz} and \eref{IRGhomo}.
{By} substituting the solution of the Teukolsky {equation,}
\begin{equation*}
\psi_4 = \frac{1}{r^4} \sum_{l=2}^\infty R_l^{(-2)}(r)_{-2}Y_l(\theta)
\end{equation*}
into \eref{psi4-hertz}, {we obtain}
\begin{equation}
\sum_{l=2}^{\infty} 8 R_l^{(-2)} \frac{_{-2}Y_l(\theta)}{\sin^2\theta}
= \left(\frac{\partial}{\partial \cos\theta}\right)^4 \sin^2\theta \overline{\Psi}~{.}
\end{equation}
From \eref{spinup}, we can obtain the following relation
\begin{equation}
\left(\frac{\partial }{\partial \cos\theta}\right)^4 \frac{_{-2}Y_l(\theta)}{\sin^2\theta}
= \frac{1}{\sin^2\theta} \frac{_2Y_l(\theta)}{(l+2)(l-1)(l+1)l}
~{.}
\end{equation}
By using this relation, $\Psi$ can be integrated as
\begin{equation}
\begin{split}
\overline{\Psi}(r,\theta) &= \overline{\Psi_{\rm P}}+\overline{\Psi_{\rm H}},
\label{hertzPH}
\end{split}
\end{equation}
where
\begin{equation}
\overline{\Psi_{\rm P}} \equiv \sum_{l=2}^{\infty} \frac{8{R_l^{(-2)}(r) _{2}Y_l(\theta)}}{(l+2)(l-1)(l+1)l} ~,
\label{PsiP}
\end{equation}
\begin{equation}
\begin{split}
\overline{\Psi_{\rm H}} &\equiv \frac{2A}{\sin^2\theta} \bigg( \frac{a(r)}{6}\cos^3\theta +\frac{b(r)}{2}\cos^2\theta
\\&~~~~~~~~~~~~~~~~~~~~~~~~~
+ c(r)\cos\theta +d(r) \bigg)~,
\end{split}
\label{PsiHabcd}
\end{equation}
and where $\overline{\Psi}$ is the complex conjugate of $\Psi$.
$a(r)$, $b(r)$, $c(r)$, and $d(r)$ are arbitrary functions and $A$ is a constant defined as
\begin{equation}
A\equiv \frac{\mass}{r_0\sqrt{\Delta_0}}~.
\end{equation}
Here, $\Psi_{\rm P}$ and $\Psi_{\rm H}$ are the particular solution and the homogeneous solution
of the equation \eref{psi4-hertz}{,} respectively.
The particular solution $\Psi_{\rm P}$ satisfies \eref{psi0-hertz} and \eref{IRGhomo} in the region,
$r\neq r_0$. The reason is as follows.
From the Teukolsky--Starobinsky relation, we obtain
\begin{equation}
\left(\frac{\partial}{\partial r}\right)^4
\frac{R_l^{(-2)}(r)}{(l+2)(l-1)(l+1)l}
= \frac{1}{4}R_l^{(2)}(r)
~{.}
\end{equation}
By using this, we can obtain $\psi_0$ by substituting $\Psi_{\rm P}$ into \eref{psi0-hertz}.
Further, since $R_l^{(-2)}(r)$ is the solution of the radial Teukolsky equation with the source term
consisting of a circular rotating ring, it satisfies the homogeneous Teukolsky equation
in the region, $r\neq r_0$.
Thus, it is clear that the particular solution $\Psi_{\rm P}$ of the form \eref{PsiP} satisfies
\eref{IRGhomo} in the region, $r\neq r_0$.
It is now shown that $\Psi_{\rm P}$ is a Hertz potential that satisfies \eref{psi0-hertz}, \eref{psi4-hertz}, and \eref{IRGhomo}
everywhere except for the region, $r=r_0$.
$\Psi_{\rm P}$ is not only singular at $r=r_0$, but also does not include lower modes
($l=0, 1$).
The monopole perturbation and the dipole perturbation of the space{-}time are considered
to be included in the ``homogeneous solution'' part $\Psi_{\rm H}$.
We can obtain constraints on the functions $a(r)$, $b(r)$, $c(r)$, and $d(r)$ in $\Psi_{\rm H}$
from \eref{IRGhomo}.
By substituting $\Psi_{\rm H}$ into \eref{IRGhomo}, we obtain
\begin{equation}
(\pmb{\Delta} -2\mu +2\gamma)\pmb{D} \Psi_{\rm H}
= (\overline{\pmb{\delta}} -2\beta)(\pmb{\delta} +4\beta)\Psi_{\rm H}~.
\end{equation}
This condition implies that each of $a(r)$, $b(r)$, $c(r)$, and $d(r)$ must be in the following forms.
\begin{equation}
\begin{split}
& a(r)=a_1 r^2 (r-3M)+a_2~,\\
& b(r)=b_1 r^2+b_2(r-M)~,\\
& c(r)=-\frac{a_1}{2}(r^2+4M^2)(r-M)-\frac{a_2}{2}
\\&~~~~~~~~~~~~~~~
+c_1 r^2+c_2(r-M)~,\\
& d(r)=\frac{b_1}{2}r^2 +\frac{b_2}{2}r +d_1 r^2(r-3M)+d_2~.
\label{abcd}
\end{split}
\end{equation}
Here $a_1$, $a_2$, etc. are arbitrary complex constants.
Then, the right-hand side of \eref{psi0-hertz} vanishes when we substitute $\Psi_{\rm H}$ with constrains \eref{abcd}.
Thus, $\Psi_{\rm H}$ with \eref{abcd} is a homogeneous solution of \eref{psi0-hertz} and \eref{psi4-hertz},
and satisfies \eref{IRGhomo}.
It is known \cite{ori03} that
the Hertz potential that globally satisfies \eref{psi0-hertz},
\eref{psi4-hertz}, and \eref{IRGhomo} simultaneously does not exist because of the presence of matter (the ring).
{Thus, we need to give up the global regularity of the solution.
We find that we can obtain a solution which is smooth at $r=r_0$ if we abandon the smoothness of
the Hertz potential at $(r\geq r_0, ~\theta=\pi/2)$.
We also find that in order to obtain the smoothness at $r=r_0$,
we need to include the contribution from the lower modes ($l=0, 1$).
We show that this can be done by choosing eight complex parameters, $a_1$, $a_2$, etc.,
appropriately, and making the Hertz potential $\Psi=\Psi_{\rm P}+\Psi_{\rm H}$ satisfy \eref{psi0-hertz},
\eref{psi4-hertz}, and \eref{IRGhomo} everywhere except for the region $(r\geq r_0,~\theta=\pi/2)$.}
\subsection{Fields corresponding to $\Psi_{\rm P}$}
\label{section:PsiP}
\begin{figure*}[ht]
\begin{center}
\includegraphics[width=5.6cm]{./Repsi1P.eps}
\includegraphics[width=5.6cm]{./Repsi2P.eps}
\includegraphics[width=5.6cm]{./Repsi3P.eps}
\end{center}
\caption{Radial dependenc{e} of the real parts of $\psi_1$ (left), $\psi_2$ (center), and $\psi_3$ (right) {derived from} $\Psi_{\rm P}$ at $\theta=\pi/4$. The radius of the ring is $r_0=10M$. They are discontinuous at $(r=r_0, ~\theta=\pi/4)$.}
\label{fig:Repsi123P}
\end{figure*}
\begin{figure*}[htb]
\begin{center}
\includegraphics[width=5.6cm]{./Impsi1P.eps}
\includegraphics[width=5.6cm]{./Impsi2P.eps}
\includegraphics[width=5.6cm]{./Impsi3P.eps}
\end{center}
\caption{Radial dependenc{e} of the imaginary parts of $\psi_1$ (left), $\psi_2$ (center), and $\psi_3$ (right) {derived from} $\Psi_{\rm P}$ at $\theta=\pi/4$. The radius of the ring is $r_0=10M$. They are discontinuous at $(r=r_0, ~\theta=\pi/4)$.}
\label{fig:Impsi123P}
\end{figure*}
{Here, we demonstrate the behavior of the Weyl scalars associated with $\Psi_{\rm P}$.
We introduce a notation like $\psi_1^{\rm P}$ which means that
it is calculated by substituting $\Psi=\Psi_{\rm P}$ into the equation for $\psi_1$ in \eref{weyl-hertz}.
In Figs. \ref{fig:Repsi123P} and \ref{fig:Impsi123P}, we show the radial dependence of the real and
imaginary parts of $\psi_1^{\rm P}, \psi_2^{\rm P}$ and $\psi_3^{\rm P}$ at $\theta=\pi/4$. }
{As discussed in the previous section, }
$\psi_0^{\rm P}$ agree with the Teukolsky solution $\psi_0$,
therefore the graph is the same as Fig. \ref{fig:TeuSolution}.
{Other Weyl scalars, }
$\psi_1^{\rm P}$, $\psi_2^{\rm P}$, and $\psi_3^{\rm P}$, have discontinuity on the
surface of sphere at radius $r=r_0$, although there is no matter field
on the surface $(r_0,~\theta\neq \pi/2)$.
{It is also apparent that the }
perturbed metric $h_{\mu\nu}^{\rm P}$ {calculated from $\Psi_{\rm P}$}
is not smooth on the surface of
the sphere, too.
\begin{figure*}[ht]
\begin{center}
\includegraphics[width=5.6cm]{./Impsi1.eps}
\includegraphics[width=5.6cm]{./Impsi2.eps}
\includegraphics[width=5.6cm]{./Impsi3.eps}
\end{center}
\caption{Radial dependence of the imaginary parts of $\psi_1$ (left), $\psi_2$ (center), and $\psi_3$ (right) {derived from} $\Psi_{\rm P}+\Psi_{\rm H}$ at $\theta=\pi/4$. The radius of the ring is $r_0=10M$. {It is clear that they are continuous at $r=r_0$.}}
\label{fig:Impsi123}
\end{figure*}
\subsection{$\Psi_{\rm H}$}
\label{section:HertzH}
\subsubsection{{Contribution of angular momentum perturbation}}
Keidl, Friedman, and Wiseman (2007) \cite{kfw} illuminated
that some of parameters are physical parameters and others are pure gauge.
They found that $\Re(b_1)$ and $\Re(b_2)$ contribute to {the} mass perturbation of
the space{-}time and $\Im(a_2)$ contributes to {the} angular momentum perturbation of
the space{-}time. Specifically, it is found that
\begin{equation}
\begin{split}
& \delta M = -A(3M\Re(b_1)+\Re(b_2))~
\\
& \delta J = -A\Im(a_2)~
\label{deltaMdeltaJ}
\end{split}
\end{equation}
The latter relation is obtained as below \cite{kfw}.
The metric perturbation due to small
angular momentum to the Schwarzschild space-time is given in the Boyer{--}Lindquist coordinates as
\begin{equation}
h_{ab}^{\rm Kerr} = -\frac{4\delta J}{r}\sin^2\theta (\dev t)_{(a}(\dev\phi)_{b)}
~{.}
\end{equation}
The corresponding tetrad components are
\begin{equation}
h_{23}^{\rm Kerr} = -\im \frac{ \delta J}{\sqrt{2} r^2}\sin\theta~,~~~~~
h_{13}^{\rm Kerr} = -\im \frac{2\delta J}{\sqrt{2}\Delta}\sin\theta~.
\end{equation}
We can transform these into ingoing radiation gauge, with the gauge vector
\begin{equation*}
\xi^a = \xi^3 m^a +\xi^4\overline{m}^a;
\end{equation*}
\begin{equation}
\xi^3 = -\xi^4 = -\frac{\im \delta J}{\sqrt{2}M}
\left( 1+\frac{r}{2M}\ln\left(1-\frac{2M}{r}\right) \right)~.
\end{equation}
The resultant nonzero component of $h_{ab}=h_{ab}^{\rm Kerr}+\mathcal{L}_\xi g_{ab}$ is
\begin{equation}
h_{23} = -\im \frac{\sqrt{2}\delta J}{r^2}\sin\theta~.
\end{equation}
{The metric associated with the imaginary part of $a_2$ can be obtained
by inserting \eref{PsiHabcd} and \eref{abcd} into \eref{irg}, and becomes
$h_{23}^{\rm H}=\im (\sqrt{2}A\Im(a_2)/r^2)\sin\theta$.
We thus obtain $\delta J = -A\Im(a_2)$. }
In our case, $\delta M$ and $\delta J$ are the energy and angular momentum
of the rotating ring, respectively. They are
\begin{equation}
M_{\rm ring} \equiv -2\pi\mass u_a (\partial_t)^a~,~~~~~J_{\rm ring} \equiv 2\pi\mass u_a (\partial_\phi)^a~,
\label{MJring}
\end{equation}
where $u^a$ is the four-velocity of the ring,
\begin{equation*}
u^a = \sqrt{\frac{r_0}{r_0-3M}}\left( (\partial_t)^a +\sqrt{\frac{M}{r_0{}^3}}(\partial_\phi)^a \right)~.
\end{equation*}
Interestingly, the jumps of $\Im(\psi_1)$, $\Im(\psi_2)$, and $\Im(\psi_3)$ disappeared
when we choose $\Im(a_2)=0$ for $r<r_0$ and $\Im(a_2)=-\delta J/A$ for $r>r_0$.
Namely, the imaginary parts of $\psi_1$, $\psi_2$, and $\psi_3$ {are continuous at $r=r_0$}
if we choose
\begin{eqnarray}
\Psi = \left\{ \begin{array}{ll}
\Psi_{\rm P}, & (2M < r < r_0)~ \\
\Psi_{\rm P} + \frac{2\im\delta J}{\sin^2\theta}\left( \frac{1}{6}\cos^3\theta -\frac{1}{2}\cos\theta \right). &(r_0 < r)~
\end{array} \right.
\nonumber\\
\label{eq:Psiangmom}
\end{eqnarray}
Further, they also look smooth at $r=r_0$ (Fig. \ref{fig:Impsi123}).
{Although we want to determine other parameters in a similar way,
we can not do it.
One reason is that since the mass perturbation in \eref{deltaMdeltaJ} contains two parameters,
$\Re(b_1)$ and $\Re(b_2)$, it is not possible to determine them from only one equation.
Further, we don't have similar equations for other parameters
which are not related to the mass and angular momentum perturbation. }
\subsubsection{Determination of all parameters in $\Psi_{\rm H}$}
We now determine all other parameters so that the discontinuity of all the fields at $r=r_0$
{disappears}.
Details are in the appendix.
First, we obtain four conditions by demanding that the metric perturbation and the Weyl scalars
should not diverge at $\theta=0$ and ${\theta=}\pi$. This can be satisfied when the Hertz potential $\Psi$
does not diverge at $\theta=0$ and $\theta=\pi$.
From the condition at $\theta=0$, we obtain
\begin{equation}
\begin{split}
& 3 d_1 = a_1~,~~~c_1=Ma_1-b_1~,\\
& c_2=2M^2a_1-b_2~,~~~6d_2=2a_2-3Mb_2
~{.}
\label{fromNpole}
\end{split}
\end{equation}
From the condition at $\theta=\pi$, we obtain
\begin{equation}
\begin{split}
& 3 d_1 = -a_1~,~~~c_1=Ma_1+b_1~,\\
& c_2=2M^2a_1+b_2~,~~~6d_2=-2a_2-3Mb_2
~{.}
\label{fromSpole}
\end{split}
\end{equation}
These sets of conditions are simultaneously satisfied if and only if
$a_1=a_2=b_1=b_2=c_1=c_2=d_1=d_2=0$, i.e. $\Psi_{\rm H}=0$.
This means that we can not have the contribution from the mass and the angular momentum
perturbation.
This implies that we can not obtain the regular solution globally.
However, we find that if we divide the space{-}time into several region,
we can obtain regular solution in each region.
{Namely, we divide the region into three regions:
$(2M<r<r_0)$, $(r>r_0, ~0\leq \theta<\pi/2)$, and $(r>r_0, ~\pi/2<\theta\leq \pi)$.
We denote each region by $I$, $N$, and $S$, respectively (Fig. \ref{fig:regions}).}
\begin{figure}[ht]
\begin{center}
\includegraphics[width=8cm]{./region.eps}
\caption{$r$-$\theta$ plane. The three regions are divided by dashed lines.
The filled black circle at the center is {the region within the} event horizon of the black hole.
The two black dots represent the position of the ring. }
\label{fig:regions}
\end{center}
\end{figure}
We look for the set of parameters that satisfy \eref{fromNpole} {in $N$} and \eref{fromSpole} {in $S$.}
Since these are four equations among eight unknown parameters,
the remaining parameters we have to determine are four.
As in the case of the contribution of the angular momentum perturbation,
\eref{eq:Psiangmom}, we add $\Psi_{\rm H}$ only at $r>r_0$.
Here, we note the symmetry of $\Psi_{\rm P}$.
From \eref{PsiP}, we find that,
just like $\psi_0$ and $\psi_4$, the real and imaginary part of $\Psi_{\rm P}$
are symmetric and antisymmetric about the equatorial plane respectively.
In order to kill the jump of $\Psi_{\rm P}$ at $r=r_0$, $\Psi_{\rm H}$ at $r>r_0$
must have the same symmetry about the equatorial plane.
Therefore we get
\begin{equation}
\begin{split}
&
a_{N}(r) = -\overline{a_{S}}(r)~,~~~~~b_{N}(r) = \overline{b_{S}}(r)~,\\
&
c_{N}(r) = -\overline{c_{S}}(r)~,~~~~~d_{N}(r) = \overline{d_{S}}(r)~.
\label{NSrelation}
\end{split}
\end{equation}
Here, $a_{N}(r)$ means $a(r)$ in $N$, and $a_{S}(r)$ means $a(r)$ in $S$, etc.
It is sufficient if we determine four complex parameters
only in the region $N$ or $S$.
From \eref{fromNpole}, we adopt $a_1$, $a_2$, $b_1$ and $b_2$ of $\Psi_{\rm H}$ in region $N$
as independent parameters.
When the parameters satisfy \eref{fromNpole},
the fields corresponding to $\Psi_{\rm H}$ and $\Psi_{\rm H}$ in $N$ can be written
as they include only $a_1$, $a_2$, $b_1$ and $b_2$ (equations \eref{psiH}-\eref{HertzH}).
We numerically determine values of these parameters that satisfy the {continuity} conditions
\begin{equation*}
\begin{split}
\left[ F_{\rm P}(r,\theta) \right]_{r_0} + F_{\rm H}(r_0, \theta) = 0~
\end{split}
\end{equation*}
for $F=\psi_1,~\psi_2,~\psi_3,~h_{22},~h_{23},~h_{33},~\Psi$, where
\begin{equation}
\left[ F_{\rm P}(r,\theta)\right]_{r_0}
\equiv \lim_{r\rightarrow r_0{}^+}F_{\rm P}(r,\theta) - \lim_{r\rightarrow r_0{}^-} F_{\rm P}(r,\theta)~.
\end{equation}
By using the relations between these four parameters, $a_1$, $a_2$, $b_1$ and $b_2$,
with $F_{\rm H}$ above given in \eref{psiH}-\eref{HertzH}, we obtain
\begin{equation*}
\begin{split}
& (a_1)_N = -0.0000025233 - 4.2486\im~,\\
& (a_2)_N = -134.33 - 2123.8\im~,\\
& (b_1)_N = 67.169 + 34.993\im~,\\
& (b_2)_N = -738.86 - 0.079440\im~.\\
\end{split}
\end{equation*}
when $M=1~,m=M/100~,r_0=10M$.
\begin{figure*}[htbp]
\begin{center}
\includegraphics[width=5.6cm]{./Repsi1.eps}
\includegraphics[width=5.6cm]{./Repsi2.eps}
\includegraphics[width=5.6cm]{./Repsi3.eps}
\end{center}
\caption{Radial dependence of the real part of $\psi_1$ (left), $\psi_2$ (center), and $\psi_3$ (right)
derived from $\Psi_{\rm P}+\Psi_{\rm H}$, with $\theta=\pi/4$ fixed.
The radius of the ring is $r_0=10M$. {It is clear that they are continuous at $r=r_0$.}}
\label{fig:Repsi123}
\end{figure*}
\begin{figure*}[htbp]
\begin{center}
\includegraphics[width=5.6cm]{./h22.eps}
\includegraphics[width=5.6cm]{./Reh23.eps}
\includegraphics[width=5.6cm]{./Reh33.eps}
\\
\includegraphics[width=5.6cm]{./Imh23.eps}
\includegraphics[width=5.6cm]{./Imh33.eps}
\end{center}
\caption{
Radial dependence of the each component of $h_{ab}$ derived from $\Psi$ at $\theta=\pi/4$.
The radius of the ring is $r_0=10M$. They are continuous at $r=r_0$. }
\label{fig:MP}
\end{figure*}
The plots of $\Re(\psi_1)$, $\Re(\psi_2)$ and $\Re(\psi_3)$
derived from $\Psi_{\rm P}+\Psi_{\rm H}$ are shown in Fig. \ref{fig:Repsi123}.
We find that all of the discontinuity disappeared.
Note that because of the relations \eref{NSrelation},
each of parameters $\Re(b_1)$, $\Re(b_2)$, and $\Im(a_2)$ is the same value
in $N$ and $S$.
Thus, $\delta M$ and $\delta J$ in \eref{deltaMdeltaJ} is the same in $N$ and $S$.
Interestingly, we numerically obtain the very good agreement between ($\delta M$, $\delta J$)
and the mass and angular momentum of the ring, \eref{MJring}.
We obtain from \eref{deltaMdeltaJ},
\begin{equation}
\begin{split}
\delta M &= -A(3M\Re(b_1)+\Re(b_2)) = {0.0600781}
~{,}\\
\delta J &= -A\Im(a_2) = {0.237451}
~.\\
\end{split}
\end{equation}
On the other hand, from \eref{MJring}
\begin{equation}
\begin{split}
M_{\rm ring} &= {0.06007874270}
~,\\
J_{\rm ring} &= {0.2374820823}
~{.}\\
\end{split}
\end{equation}
Although the method to determine the $\Psi_{\rm H}$ here is rather heuristic,
this excellent agreement suggests the validity of the method and the results.
{Further discussion on the the accuracy of the numerical results is given
at the end of Appendix \ref{section:DeterminationPsiH}.}
The results in the case of $r_0/M=6, 10, 20, 50$ are shown in Table \ref{r0depM} and \ref{r0depJ}.
\begin{table}[htb]
\caption{$\delta M$}
\begin{tabular}{| c | | l | l | l |} \hline
$r_0/M$ &
$\delta M$ & $M_{\rm ring}$ & $|(M_{\rm ring}-\delta M)/M_{\rm ring}|$ \\ \hline
6 & 0.0592444 & 0.05923843916 & 1.008027909 $\times 10^{-4}$ \\
10 & 0.0600781 & 0.06007874270 & 1.005730101 $\times 10^{-5}$ \\
20 & 0.0613351 & 0.06133564195 & 8.821135362 $\times 10^{-6}$ \\
50 & 0.0622144 & 0.06221386387 & 7.995806223 $\times 10^{-6}$ \\
100 & 0.0625205 & 0.06252015946 & 5.948001469 $\times 10^{-6}$ \\
\hline
\end{tabular}
\label{r0depM}
\end{table}
\begin{table}[htb]
\caption{$\delta J$}
\begin{tabular}{| c | | l | l | l |} \hline
$r_0/M$
& $\delta J$ & $J_{\rm ring}$ & $|(J_{\rm ring}-\delta J)/J_{\rm ring}|$ \\ \hline
6 & 0.217649 & 0.2176559237 & 3.301954698 $\times 10^{-5}$ \\
10 & 0.237451 & 0.2374820823 & 1.308149216 $\times 10^{-4}$ \\
20 & 0.304774 & 0.3047792551 & 1.758912364 $\times 10^{-5}$ \\
50 & 0.458263 & 0.4582483860 & 3.190540426 $\times 10^{-5}$ \\
100 & 0.637962 & 0.6379608107 & 1.972221458 $\times 10^{-6}$ \\
\hline
\end{tabular}
\label{r0depJ}
\end{table}
Finally, we show the radial dependence of the metric perturbation,
$h_{22}, \Re(h_{23}), \Re(h_{33}), \Im(h_{23})$, and $\Im(h_{33})$,
computed from \eref{irg} in Fig. \ref{fig:MP}.
These are the cases for $\theta=\pi/4$.
We find that they are smooth at $r=r_0$.
\section{Summary and Discussion}
\label{section:summary}
We computed the metric perturbation produced by a rotating circular mass ring
around a Schwarzschild black hole by using {the} CCK formalism.
In {the} CCK formalism, {the} Weyl scalars and the metric perturbation are expressed
by the Hertz potential in a radiation gauge.
The Hertz potential can be obtained by integrating an equation
which relates the Hertz potential with the Weyl scalars $\psi_0$ or $\psi_4$.
We used $\psi_4$ to obtain the Hertz potential.
The Hertz potential contains two parts, $\Psi_{\rm P}$ and $\Psi_{\rm H}$.
$\Psi_{\rm P}$ is derived directly from $\psi_4$ and
$\Psi_{\rm H}$ is the part which contains the integration constants.
We first obtained $\Psi_{\rm P}$
which has discontinuity on the surface of the sphere
at the radius of the ring. $\Psi_{\rm H}$, on the other hand, has 8 complex parameters,
given in \eref{abcd}. Among them, $\Im(a_2)$ is related to the angular momentum
perturbation and $\Re(b_1)$ and $\Re(b_2)$ are related to the mass perturbation.
We found that if we determine $\Im(a_2)$ by setting the angular momentum perturbation equal to the
angular momentum of the ring, the imaginary parts of $\psi_1$, $\psi_2$ and $\psi_3$
become continuous at the radius of the ring.
We determined other parameters by requiring the continuity condition at the radius of the ring.
We found that if we require the {regularity} condition both at $\theta=0$ and $\theta=\pi$,
we only have a trivial solution and $\Psi_{\rm H}$ becomes zero.
This fact shows the impossibility to obtain a globally regular solution
which were discussed previously (\cite{ori03}, \cite{kfw}, \cite{pmb}).
We divided the space time into 3 regions, $N$, $S$ and $I$, as in
Fig. \ref{fig:regions}, and tried to obtain a solution which is regular in each region and continuous
on the surface of the sphere at the ring radius.
We set $\Psi_{\rm H}=0$ in the inner region $I$, and
determined all unknown parameters of $\Psi_{\rm H}$
in the region $N$ and $S$ numerically
by requiring the continuity at the ring radius.
As a result, the Weyl scalars, $\psi_1$, $\psi_2$ and $\psi_3$, and
the components of the metric perturbation $h_{\mu\nu}$ become continuous
at the ring radius.
We also found that the mass perturbation determined in this method
agreed with the mass of the ring.
This fact suggests the validity of the method and the results in this paper.
The metric perturbation we obtained has a discontinuity on the equatorial plane outside the ring.
This is similar to the metric perturbation of a Schwarzschild black hole by a particle at rest,
which was discussed by Keidl et al. \cite{kfw}
Their metric perturbation has radial string singularity inside or outside the particle.
One of the major difference between Ref. \cite{kfw} and this paper is the presence of the
angular momentum perturbation in this paper.
We found that the angular momentum perturbation was important to remove
the discontinuity of $\Im(\psi_1^{\rm P})$, $\Im(\psi_2^{\rm P})$,
and $\Im(\psi_3^{\rm P})$.
However, in order to remove the discontinuity
of the real part of {the} Weyl scalars and that of the metric perturbation,
the mass perturbation $M_{\rm ring}$ and the gauge freedom
must be added outside the ring.
A natural extension of this work is to apply to the Kerr black hole case.
{In the case of Schwarzschild black hole,
the radial functions $R_l^{(2)}$ and $R_l^{(-2)}$ were expressed in terms of the associated Legendre functions.
In the case of Kerr,
{the radial functions become more complicated.}
Further, the relations between the perturbed Weyl scalars and the Hertz potential
{become} more complicated.
{Besides these complication,
it would be useful to derive the relation between the parameters in $\Psi_{\rm H}$
and the mass and angular momentum perturbation in the Kerr case. }
}
{Will \cite{will74, will75} derived a solution of rotating mass ring around a slowly rotating black hole.
The method used in those papers are completely different from our method.
Further, the gauge condition used is different from ours.
We have to treat these issues to compare our results with \cite{will74, will75},
and this is also one of our future works. }
An another interesting and important problem is the case of a particle
orbiting around a black hole. (e.g., Ref. \cite{pmb})
{In that case, since the problem becomes non-stationary,
the Teukolsky equation and the spin-weighted spheroidal harmonics
must be solved numerically. Although the problem must be solved fully numerically,
it would be straightforward to obtain the gravitational field produced by a orbiting particle
by using the method in this paper. }
{Pound et al. \cite{pmb} discussed a method to compute the gravitational self-force
on a orbiting point mass in {a} radiation gauge by using a local gauge transformation.
Once we obtain the gravitational field in {a} radiation gauge,
it would be possible to compute the self-force with the prescription of \cite{pmb}.}
We will work on these problem in the future.
|
\section{Introduction}
Direct integration of the classical $N$-body problem is an important tool for studying astrophysical systems. Examples include planetary systems, open and globular clusters dynamics, large-scale dynamics of galaxies, and structure formation in the universe. In many cases the calculations involve systems where the intensity of gravitational interactions spans multiple orders of magnitude with corresponding timescale variations. For example, the initial stages of cluster formation are now thought to resemble multi-scale fractal structures \citep{GoodwinWhitworth2004}, and stellar systems are invariably formed with a high fraction of binaries and hierarchical multiples that affect the dynamical evolution in crucial ways \citep{PortegiesZwart2010}.
In practice, integrating multiscale systems requires specialised methods that vary the resolution at which we treat different parts of the simulation. The aim of this is to obtain a solution with an acceptable accuracy without unnecessarily spending computational resources on the slowly evolving parts of the simulation. In generic $N$-body integrators, this idea is most commonly implemented via {\em particle-based block time steps} --- every particle in the system maintains an individual time step limited to discrete values in a power of two hierarchy. These block time steps are then typically used to determine the frequency of calculating the total force acting on a particle \citep[e.g.][]{McMillan1986, Makino1991,Konstantinidis:2010hx}.
While considerably speeding up calculations, particle-based block time steps are nevertheless limited in their ability to treat the extreme scale differences often present in $N$-body systems. Hence, complementary strategies, such as binary regularisation and neighbourhood schemes, have been devised. These approaches complicate the implementation of $N$-body integrators, and often introduce new method-specific free parameters. It is also unclear whether these combinations of multiple strategies represent the best possible approach for integrating multiscale $N$-body systems. These issues provide a clear incentive to explore alternative methods.
In \cite{Pelupessy:2012if}, we derived generic $N$-body integrators that recursively and adaptively split the Hamiltonian of the system. These methods show improved conservation of the integrals of motion by always evaluating partial forces between particles in different time-step bins symmetrically, and by using an approximately time-symmetric time-step criterion.
In the present work, we introduce a new Hamiltonian-splitting integration method that is particularly adept at integrating initial conditions with significant hierarchical substructure. Our approach is based on assigning time steps to individual interactions, followed by partitioning the system Hamiltonian based on a graph formed by the set of interactions that are faster than a fixed threshold time step. The successive partitioning produces closed Hamiltonians such that we can easily use specialised solvers for situations where more efficient solvers are available. Numerical experiments show that our integrator compares favourably to existing methods even for an ordinary Plummer sphere where the prevalence of isolated subsystems is not immediately obvious. For astrophysically realistic systems explicitly chosen for their multi-scale substructure, the performance gains increase can be orders of magnitude. An implementation of the method is incorporated in the HUAYNO code, which is freely available as a part of the AMUSE framework\citep{PortegiesZwart2013b, Pelupessy2013Amuse} and which was used for the tests presented in this paper.
Our method is similar in spirit, and accelerates the calculation of the N-body problem for much the same reasons as the well known neighbour schemes. The main idea is to divide the total force acting on a particle into a fast and a slow component based the distance to the given particle. Different approaches have been used for treating fast and slow components. The Ahmad-Cohen neighbourhood scheme\citep{Ahmad:1973kn} treats fast components with a more strict time step criteria. Alternatively, the PPPT scheme\citep{Oshino:2011wv} integrates fast components with a fourth-order Hermite method while using a leapfrog-based tree code for the long range interactions. The criteria for determining neighbourhood memberships are heuristics known to work in numerical experiments, e.g. a sphere with a fixed radius centred on the acting particle. These methods need to continuously update neighbourhood memberships as the system state changes throughout the simulation. In addition, neighbourhood schemes only make a single distinction between treating small subsystems such as hard binaries or many-body close encounters, and the large scale dynamics. It is difficult to generalise a neighbourhood scheme beyond a binary differentiation of the particle distribution.
In Section \ref{sec:method} we describe a bottleneck in existing general $N$-body splitting methods, and derive our novel splitting scheme that overcomes this bottleneck. Section \ref{sec:tests} presents the results of numerical tests comparing of our method to existing approaches. Finally, in Section \ref{sec:discussion} we discuss possible improvements and extensions to our work, including the feasibility of integrating general $N$-body systems using purely interaction-specific time steps.
\section{Method}\label{sec:method}
\subsection{Deriving time stepping schemes via Hamiltonian splitting}\label{sec:splitting-schemes}
The Hamiltonian for a system of $N$ particles $i=1\ldots N$ under gravitational interaction can be represented as a sum of momentum terms $T_{i}$ and potential terms $V_{ij}$:
\begin{eqnarray}
\label{eq:Hamiltonian}
H(\mathbf{p}_{i},\mathbf{q}_{i}) & = & T+V=\underset{i=1}{\overset{N}{\sum}}T_{i}+\underset{\substack{i,j=1\\i<j}}{\overset{N}{\sum}}V_{ij}\\
T_{i} & = & \frac{\left|\mathbf{p}_{i}\right|^{2}}{2m_{i}}\\
V_{ij} & = & -G\frac{m_{i}m_{j}}{\sqrt{q_{ij}^{2}+\varepsilon^{2}}}
\end{eqnarray}
where $m_{i}$ is the mass, $\mathbf{q}_{i}$ is the position and $\mathbf{p}_{i}$ is the momentum of the $i$-th particle of the system, and $q_{ij} = \|\mathbf{q}_{i}-\mathbf{q}_{j} \|$ . The evolution of the state of the system for a time step $h$ is given formally by the flow operator $E_{H}(h) = \exp( h \mathbb{H})$ where $\mathbb{H}$ is the Hamiltonian vector field corresponding to $H$.
If the Hamiltonian $H$ of the system is representable as a sum of two sub-Hamiltonians, $H=A+B$, we can approximate the time evolution under $H$ with a sequence of time evolution steps under the sub-Hamiltonians $A$ and $B$. A straightforward successive application of the time evolution under $A$ followed by the time evolution under $B$ gives a first-order approximation of the full time evolution under $A+B$, while a second-order accurate approximation can be obtained with one additional operator evaluation (\cite{SSC94}, Sec 12.4, also \cite{Hairer2006}).
\begin{equation}\label{eq:second-order-split}
\operatorname{E}_{A+B}(h)=\operatorname{E}_{A}(h/2)\operatorname{E}_{B}(h)\operatorname{E}_{A}(h/2)+\operatorname{O}\left(h^{2}\right)
\end{equation}
The sub-Hamiltonian $A$ is evolved in two steps of $h/2$ and the sub-Hamiltonian $B$ is evolved in a single step $h$. We can take advantage of this property of the splitting formula by dividing terms associated with fast interactions into $A$ and terms associated with slow interactions into $B$. We can proceed by applying this splitting procedure to different sub-Hamiltonians multiple times, thereby constructing an integrator that evaluates parts of the Hamiltonian at $h$, $h/2$, $h/4$ etc, similarly to the power of two hierarchy used in block time step schemes. This approach was followed in \cite{Pelupessy:2012if}, below we will introduce some notation and give a rough derivation of the integrators there.
Hamiltonians consisting of a single momentum term and Hamiltonians consisting of a single potential term have analytic solutions. For a momentum term of the $i$-th particle
\begin{equation}
H_{i}(\mathbf{p}_{i},\mathbf{q}_{i})=T_{i}=\frac{\left\langle \mathbf{p}_{i},\mathbf{p}_{i}\right\rangle }{2m_{i}}
\end{equation}
the solution consists of updating the position of the $i$-th particle under the assumption of constant velocity for a time period of $h$ (all positions except the position of the $i$-th particle and the momenta of all particles remain unchanged).
\begin{equation}
\label{eq:drift}
\mathbf{q}_{i}(t+h)=\mathbf{q}_{i}(t)+h\mathbf{v}_{i}(t)
\end{equation}
We call the time evolution operator for the momentum term of the $i$-th particle the \emph{drift operator} and write $D_{h,T_{i}}$.
For a single potential term between particles $i$ and $j$
\begin{equation}
\label{eq:kick}
H_{ij}(\mathbf{p}_{i},\mathbf{q}_{i})=V_{ij}=-G\frac{m_{i}m_{j}}{\sqrt{q_{ij}^{2}+\varepsilon^{2}}}
\end{equation}
the solution consists of updating the momenta of the $i$-th and $j$-th particles under the assumption of constant force for a time period of $h$ (all momenta except the momenta of the $i$-th and $j$-th particles and the positions of all particles remain unchanged).
\begin{eqnarray}
\label{eq:kickij}
\mathbf{p}_{i}(t+h) & = & \mathbf{p}_{i}(t)+h\mathbf{F}_{ij}(t)\\
\label{eq:kickji}
\mathbf{p}_{j}(t+h) & = & \mathbf{p}_{j}(t)+h\mathbf{F}_{ji}(t)
\end{eqnarray}
We call the time evolution operator for the potential term between the $i$-th and $j$-th particles the kick operator and write $K_{h,V_{ij}}$.
In addition to the kick and drift operators, the two-body Hamiltonian
\begin{equation}
\label{eq:kepler}
H_{ij}(\mathbf{p}_{i},\mathbf{q}_{i})=T_{i}+T_{j}+V_{ij}=
{\frac{\langle \mathbf{p}_{i},\mathbf{p}_{i}\rangle }{2m_{i}}} +
{\frac{\langle \mathbf{p}_{j},\mathbf{p}_{j}\rangle }{2m_{j}}} -
G\frac{m_{i}m_{j}}{\sqrt{q_{ij}^{2}+\varepsilon^{2}}}
\end{equation}
is solved (semi-) analytically by the Kepler solution\footnote{even the case with $\varepsilon \ne 0$ can be solved in a universal variable formulation (Ferrari, priv. comm.)}.
In \cite{Pelupessy:2012if}, we derive multiple integrators that recursively and adaptively split the system Hamiltonian through the second-order splitting formula \eqref{eq:second-order-split}. At every step in the recursion, all particles under consideration are divided into a slow set $S$ and a fast set $F$ by comparing the particle-specific time step function $\tau(i)$ to a pivot time step $h$.
\begin{eqnarray}
S & = & \left\{ i\in1\ldots N:\,\tau(i)\geq h\right\} \\
F & = & \left\{ i\in1\ldots N:\,\tau(i)<h\right\}
\end{eqnarray}
Using the two sets $S$ and $F$, we can rewrite the system Hamiltonian as follows.
\begin{eqnarray}
H & = & H_{S}+H_{F}+V_{SF}\label{eq:SF-splitting-rule}
\end{eqnarray}
The sub-Hamiltonian $H_{S}$ can be thought of as a ``closed Hamiltonian'' of the particles in $S$. Specifically, it consists of all drifts of particles in $S$ and all kicks where both participating particles are in $S$. The same property holds for the sub-Hamiltonian $H_{F}$ and the particles in $F$. The mixed term $V_{SF}$ contains all kicks where one particle is in $S$ and the other is in $F$.
We proceed by applying the second-order splitting rule \eqref{eq:second-order-split}:
\begin{eqnarray}
\operatorname{E}_{h,H} & = & \operatorname{E}_{h,H_{S}+H_{F}+V_{SF}}\\
& \approx & \operatorname{E}_{h/2,H_{F}}\,\operatorname{E}_{h,H_{S}+V_{SF}}\operatorname{E}_{h/2,H_{F}}\label{eq:HOLD}
\end{eqnarray}
(this is not the only conceivable approximation).
The sub-Hamiltonian $H_{F}$ is closed, and consists of particles where $\tau(i)<h$. We integrate $H_{F}$ by recursively applying the entire ``slow/fast'' partitioning, but using a smaller pivot $h/2$. In contrast, both $H_{S}=T_{S}+V_{S}$ and $V_{SF}$ are explicitly decomposed into individual kicks and drifts which are applied using the current pivot time step $h$. We refer to this particular choice as the HOLD method (since it 'holds' $V_{SF}$ for evaluation at the slow timestep).
\begin{eqnarray}
\operatorname{E}_{h,H_{S}+V_{SF}} & = & \operatorname{E}_{h,T_{S}+V_{S}+V_{SF}}\\
& \approx & \operatorname{D}_{h/2,T_{S}}\operatorname{K}_{h,V_{S}+V_{SF}}\operatorname{D}_{h/2,T_{S}}
\end{eqnarray}
The pivot time step $h$ is halved with each consecutive partitioning, and the recursion terminates when all remaining particles are placed into the $S$ set.
As noted previously, recursively and adaptively splitting the system Hamiltonian using the second order splitting rule (Eq \ref{eq:second-order-split}) is similar to conventional block time steps. Both approaches evolve different parts of the system using time steps that belong to a power of two hierarchy. However, the Hamiltonian splitting method derived above evaluates pairwise particle forces symmetrically in the sense that a ``kick'' from particle $i$ to particle $j$ (Eq \ref{eq:kickij}) is always paired with an opposite kick from particle $j$ to particle $i$ (Eq \ref{eq:kickji}). Furthermore, the kicks acting upon a particle at any given timestep typically correspond to partial forces only. This is in contrast to conventional block time steps where we always calculate the {\em total} force acting on a particle at the frequency determined by the particle-specific time step criteria $\tau(i)$
\begin{eqnarray}
\label{eq:Hamiltonian}
\mathbf{p}_{i}(t+h) & = & \mathbf{p}_{i}(t)+h\underset{j=1, j\neq i}{\overset{N}{\sum}}\mathbf{F}^{\star}_{ij}(t)
\end{eqnarray}
where, $\mathbf{F}^{\star}_{ij}(t)$ is the force acting on particle $i$ due to particle $j$, derived from extrapolated positions if necessary. Specifically, in situations where the position of particle $j$ has not been calculated for time $t$, we calculate the force by extrapolating the position at $t$ from the last known position. This can happen when particle $i$ is assigned a smaller time step than particle $j$. We refer to this method as BLOCK, and include it as a reference in our numerical tests to determine whether more ``aggressive'' splitting methods (such as HOLD) reduce the number of kicks and drifts while maintaining the accuracy of the solution.
The HOLD method evolves all kicks between fast particles at the fast time step. This is inefficient in the presence of isolated fast subsystems, as interactions between particles that belong to different subsystems could be evolved at a slower time step. As an extreme example, consider a Plummer sphere with each star being replaced by a stable hard binary. Here, every star has a close binary interaction that needs to be evaluated at a fast time step. However, the HOLD integrator will in this case integrate all interactions, including long-range interactions between stars in different binaries at a time step determined the binary interactions. The behaviour of the method becomes equivalent to evolving the entire system with a shared global time step!
In addition to the dramatic example just discussed, the same inefficiency --- evaluating long-range interactions between isolated fast subsystems at time steps determined by fast interactions inside the subsystems --- can manifest itself in other situations, such as the following.
\begin{itemize}
\item In a system with multiple globular clusters, each individual globular cluster is a subsystem.
\item In a globular cluster with planets around some of the stars, each star with planets is a subsystem.
\item In a single globular cluster, each close encounter between two or more stars is a subsystem.
\end{itemize}
\subsection{Hamiltonian splitting with connected components}\label{sec:cc-split}
The partitioning used in the HOLD method is based on a \emph{particle-specific} time step criteria $\tau(i)$, which by definition cannot separate slow and fast interactions in situations where all particle-specific time steps have the same (fast) value. We therefore introduce the \emph{interaction-specific} time step criterion $\tau(i,j)$
\begin{equation}
\label{eq:tauij}
\tau(i,j) = \eta \min \left(
\frac{\tau_\textrm{freefall}(i,j)}{\left(1-\frac{1}{2}\frac{d\tau_\textrm{freefall}(i,j)}{dt}\right)},
\frac{\tau_\textrm{flyby}(i,j)}{\left(1-\frac{1}{2}\frac{d\tau_\textrm{flyby}(i,j)}{dt}\right)}
\right)
\end{equation}
where $\tau_\textrm{freefall}(i,j)$ and $\tau_\textrm{flyby}(i,j)$ are proportional to the interparticle free-fall and interparticle flyby times as defined by eqs (13) and (16) in \cite{Pelupessy:2012if}, and $\eta$ is an accuracy parameter.
We split the system Hamiltonian using the connected components\cite[Sec B.4]{Cormen:2001uw} of the undirected graph generated by the time step criteria $\tau(i,j)$. Specifically, the particles of the system correspond to the vertices of the graph, and there is a edge between particles $i$ and $j$ if their interaction cannot be evaluated at the threshold time step $h$.
\begin{equation}
\tau(i,j)<h
\end{equation}
\begin{figure*}
\includegraphics[width=0.32\textwidth]{figs/cc-1.6-005.eps}
\includegraphics[width=0.32\textwidth]{figs/cc-1.6-006.eps}
\includegraphics[width=0.32\textwidth]{figs/cc-1.6-007.eps}
\includegraphics[width=0.32\textwidth]{figs/cc-2.3-005.eps}
\includegraphics[width=0.32\textwidth]{figs/cc-2.3-006.eps}
\includegraphics[width=0.32\textwidth]{figs/cc-2.3-007.eps}
\includegraphics[width=0.32\textwidth]{figs/cc-3.0-005.eps}
\includegraphics[width=0.32\textwidth]{figs/cc-3.0-006.eps}
\includegraphics[width=0.32\textwidth]{figs/cc-3.0-007.eps}
\caption{\label{fig:time-step-graph} Time step graphs generated by $\tau(i,j)$ at different levels of the time step hierarchy (left to right) for three values (top to bottom rows) of the fractal dimension. We plot particles that have been passed on as a part of a connected component from the previous time step level as black, grey points indicate points that are inactive on a given level (because they formed a single or binary component at a lower level). Thin grey lines indicate interactions with $\tau(i,j)<h$. Indicated in each frame are the fractal dimension (top left), and (on the bottom right) the level in the hierarchy, the number of connected components (cc) at this level, as well as the number of components that are single (s) and binary (b), and the total number of particles. }
\end{figure*}
Figure \ref{fig:time-step-graph} we visualises the time step graphs at varying values of the pivot time step $h$ for three different fractal initial conditions with different fractal dimension (described further in Section \ref{sec:tests}). As the pivot time step $h$ decreases, the set of interactions (and associated particles) that \emph{cannot} be evaluated at the current pivot time step gradually decreases as well. Although for visualisation purposes, we plot the time step graph of the entire system for varying $h$, the CC (\emph{Connected Components}) splitting method we are about to introduce typically calculates connected components for the entire system only once, at the largest pivot time step. At smaller pivot time steps, the connected components search is only calculated for parts of the system. The intuition behind this partitioning comes from clustering by maximising the margin between individual clusters as described in \cite{Duan:2009kh}.
Given a fixed pivot time step $h$, let the sets $C_{i},\, i=1\ldots K$ contain vertices of $K$ non-trivial connected components, and the set $R$ (``remainder set'') contain all particles in trivial connected components.\footnote{A \emph{trivial connected component} is a connected component with exactly one vertex and a \emph{non-trivial connected component} is a connected component with at least two vertices.} Based on the particle sets $C_{i}$ and $R$, we rewrite the Hamiltonian of the system in the following form
\begin{equation}
H = H_{C}+H_{R}+V_{CC}+V_{CR}
\end{equation}
where the individual terms are defined as follows.
\begin{eqnarray}
H_{C} & = & \underset{i=1}{\overset{K}{\sum}}H_{C_{i}}=\underset{i=1}{\overset{K}{\sum}}\left(\underset{j\in C_{i}}{\overset{}{\sum}}T_{j}+\underset{\substack{j,k\in C_{i}\\
i<j
}
}{\overset{}{\sum}}V_{jk}\right)\\
H_{R} & = & T_{R}+V_{R}=\underset{i\in R}{\overset{}{\sum}}T_{i}+\underset{\substack{i,j\in R\\
i<j
}
}{\overset{}{\sum}}V_{ij}\\
V_{CC} & = & \underset{\substack{i,j=1\\
i<j
}
}{\overset{K}{\sum}}V_{C_{i}C_{j}}=\underset{\substack{i,j=1\\
i<j
}
}{\overset{K}{\sum}}\left(\underset{\substack{k\in C_{i}\\
l\in C_{j}
}
}{\overset{}{\sum}}V_{kl}\right)\\
V_{CR} & = & \underset{i=1}{\overset{K}{\sum}}V_{C_{i}R}=\underset{i=1}{\overset{K}{\sum}}\left(\underset{\substack{j\in C_{i}\\
k\in R
}
}{\overset{}{\sum}}V_{jk}\right)
\end{eqnarray}
The term $H_{C}$ is the sum of all closed Hamiltonians $H_{C_{i}}$, each corresponding to one of the $K$ connected components. In every $H_{C_{i}}$ all drifts and some kicks cannot be evolved at the time step $h$ without violating the time step criteria. The term $H_{R}$ consists of the closed Hamiltonian formed by all of the particles in the rest system. All drifts and kicks in $H_R$ can be evolved at the current time step $h$.
The term $V_{CC}$ contains all kicks between particles that are in \emph{different} connected components. These kicks can be evaluated at the time step $h$. We explicitly point out that $V_{CC}$ explicitly contains the terms that are evolved inefficiently in the HOLD method. Similarly, $V_{CR}$ contains all kicks where one of the particles is in a connected component $C_i$, and the other is in the rest set $R$.
We split the system Hamiltonian $H$ by applying the second-order splitting rule \eqref{eq:second-order-split}:
\begin{eqnarray}
\operatorname{E}_{h,H} & = & \operatorname{E}_{h,H_{C}+H_{R}+V_{CC}+V_{CR}}\\
& \approx & \operatorname{E}_{h/2,H_{C}}\,\operatorname{E}_{h,H_{R}+V_{CC}+V_{CR}}\operatorname{E}_{h/2,H_{C}}
\end{eqnarray}
such that individual connected components $C_i$ are independently evolved at a higher pivot time step $h/2$ via recursion.
\begin{equation}
\operatorname{E}_{h/2,H_{C}}=\overset{K}{\underset{i=1}{\prod}}\operatorname{E}_{h/2,H_{C_{i}}}
\end{equation}
All remaining terms (including $V_{CC}$) are decomposed into individual drifts and kicks using the second-order splitting rule \eqref{eq:second-order-split}.
\begin{eqnarray}
\operatorname{E}_{h,H_{R}+V_{CC}+V_{CR}} & = & \operatorname{E}_{h,T_{R}+V_{R}+V_{CC}+V_{CR}}\\
& \approx & \operatorname{D}_{h/2,T_{R}}\operatorname{K}_{h,V_{R}+V_{CC}+V_{CR}}\operatorname{D}_{h/2,T_{R}}\\
& = & \operatorname{D}_{h/2,T_{R}}\operatorname{K}_{h,V_{R}}\operatorname{K}_{h,V_{CC}}\operatorname{K}_{h,V_{CR}}\operatorname{D}_{h/2,T_{R}}
\end{eqnarray}
As with the HOLD method, the pivot time step $h$ is halved at each successive partitioning such that at some point, all remaining particles are in the remainder set $R$.
Finally, the partitioning can lead to situations where $H_{C_{i}}$ is a two-body Hamiltonian (Eq.~\ref{eq:kepler}). We can use this property by integrating these cases with a dedicated a Kepler solver (we discuss this further in Section \ref{sec:plum}).
\subsection{Implementation}
\begin{figure*}
\begin{verbatim}
evolve_cc(H, h):
// split_cc() decomposes particles in H (eq 25) into:
// 1) K non-trivial connected components C_1..C_K
// 2) Rest set R
(C_1..C_K, R) = split_cc(H, h);
// Independently integrate every C_i at reduced pivot time step h/2 (eq 27)
for C_i in C_1..C_K:
evolve_cc(C_i, h/2)
// Apply drifts and kicks at current pivot time step h (eq 30)
drift(R, h/2) // evolves T_R
kick(R, R, h) // evolves V_RR
kick(C_1..C_K, C_1..C_K, h) // evolves V_CC (eq 23)
kick(C_1..C_K, R, h) // evolves V_CR (eq 24)
drift(R, h/2) // evolves T_R
// Independently integrate every C_i at reduced pivot time step h/2 (eq 27)
for C_i in C_1..C_K:
evolve_cc(C_i, h/2)
\end{verbatim}
\caption{\label{fig:cc-pseudo-code}Pseudocode for a second-order CC splitting routine for integrating a set of particles {\tt H} for time step {\tt h}.
}
\end{figure*}
We implemented the CC split in the HUAYNO code, which is freely available as a part of the AMUSE framework. Figure \ref{fig:cc-pseudo-code} sketches the main routine of the CC integrator in pseudocode, including explicit references to the corresponding equations and variables used in the derivation of the method (Section \ref{sec:cc-split}).
Subroutines and data structures that store the system state, calculate time steps, apply kicks and drifts to groups of particles, and gather statistics, are shared with other integrators such as the HOLD method. All particle states are kept in a contiguous block of memory. The connected component algorithm is implemented as a breadth-first search. It reshuffles particle states such that particles in the same connected component or rest set are kept adjacent to each other. Connected components are represented by a start and an end pointer to the contiguous array of particle states.
The time complexity of the connected component decomposition for $N$ particles has an upper bound of $O(N^2)$. This matches the time complexity of the splitting step of the HOLD method --- while the actual shuffling of the particles into $S$ and $F$ sets is $O(N)$, this division is based on the preceding step of calculating particle-based time steps $\tau(i)$ for all particles, which is $O(N^2)$.
For the special case where all interactions between the $N$ particles are below the threshold $h$, the complexity of the connected components decomposition is $O(N)$. This can happen multiple times (at consecutive recursion levels) when the initial value of the pivot time step $h$ is sufficiently large. Figuratively, if particle $X$ has a known connected component while particle $Y$ is unassigned, we can assign particle $Y$ to the connected component of particle $X$ based on a single time step evaluation $\tau(X,Y)$. A key step of the connected components search is choosing a particle $X$ with a known connected component, followed by assigning the membership of $X$ to all unassigned particles $U_k$ where $\tau(X, U_k)<h$. For the special case under consideration, a single iteration of this step is sufficient to assign membership to all particles (irrespective of the choice of the initial particle $X$), leading to a time complexity of $O(N)$. Further, while the splitting step is bounded from above by $O(N^2)$ for both HOLD and CC, the HOLD split always calculates time steps for all interactions. This is not the case with the CC method, and numerical tests in Section \ref{sec:tests} indicate that the reduction in time step evaluations does translate into improved performance.
\section{Tests}\label{sec:tests}
\begin{table*}
\begin{tabular}{|p{0.2\linewidth}|p{0.8\linewidth}|}
\hline
Method & Description\tabularnewline
\hline
\hline
BLOCK & Conventional particle-based block time steps --- positions of particles in lower time step bins are extrapolated when calculating the movement of particles in faster time step bins. \tabularnewline
\hline
HOLD & Individual timestepping method based on Hamiltonian splitting. Particles in different timestep bins interact by exchanging symmetric kicks
\citep{Pelupessy:2012if}. \tabularnewline
\hline
CC & An implementation of the connected components splitting (\mbox{Section \ref{sec:cc-split}}). Iterative partitioning based on the connected
components of the graph generated by the pairwise timestep criterion. \tabularnewline
\hline
CC\_KEPLER & An extension of the CC method that uses a Kepler solver to evolve connected components with two particles (Section \ref{sec:plum}).\tabularnewline
\hline
\end{tabular}
\caption{\label{table:methods}
An overview of the integrators used in the numerical tests. The implementations share a significant amount of the code and data structures used for storing system state, calculating time steps, and evaluating (partial) forces.
}
\end{table*}
We present results of numerical experiments of the connected components (CC) splitting method described in the previous section.
We confirm that the CC method works as intended conceptually by comparing to alternative Hamiltonian splitting methods (see Table \ref{table:methods} for an overview). Specifically, we demonstrate that the connected components search does not use excessive computational resources, reduces the number of elementary operations (kick, drift and time step evaluations) while maintaining the accuracy of the solution, and performs particularly well on multi-scale problems. Finally, we compare the CC method to established $N$-body codes.
We use $N$-body units as described in \cite{1986LNP...267..233H}.
\subsection{Smoothed Plummer sphere test}\label{sec:smoothplum}
\begin{figure*}
\centering
\includegraphics[width=0.8\textwidth]{figs/2_nitadori2008b.eps}
\caption{\label{fig:plummer-unsoftened} A comparison of the BLOCK, HOLD and CC methods on integrating an softened $1024$-body Plummer sphere: conservation of the integrals of motion (top two rows), evolution of the mass distribution of the solution (middle row) and performance metrics (two bottom rows). Lagrangian radii are plotted for 90\%, 50\%, 10\% and 1\% of the system mass. The performance metrics are normalised to the CC method at the start of the simulation. The CC method performs $1.1 \times 10^9$ kick, $6.6 \times 10^8$ time step, and $5.0 \times 10^6$ drift evaluations, taking 62 seconds for the first global time step ($1/512$-th of the simulation)
on a laptop with a 1.3GHz Intel Core i5 processor).
}
\end{figure*}
We begin by integrating an equal-mass $1024$-body Plummer sphere with softening ($\varepsilon=1/256$) for $700$ $N$-body time units using a time step accuracy parameter of $\eta=0.01$. We choose initial velocities such that the Plummer sphere is in a dynamic equilibrium. This setup is chosen to match the long-term integration tests in \cite[their section 3.2]{Nitadori:2008gt}.
Figure \ref{fig:plummer-unsoftened} visualises the conservation of the integrals of motion, the time evolution of the mass distribution, and performance metrics. While all three methods show similar energy conservation properties, only HOLD and CC maintain centre of mass, linear momentum and angular momentum near machine precision. As noted previously in \cite{Pelupessy:2012if}, this is caused by unsynchronised kicks which are only present in the BLOCK scheme. The solutions obtained by all three methods reproduce known results in terms of Lagrangian radii, the core radius and the core density. The CC scheme is about twice as fast than the HOLD scheme at the beginning of the simulation, and remains the fastest scheme throughout the run. The overall runtime measurements correlate with the number of time step formula evaluations and, to a lesser extent, the number of kick and drift formula evaluations. This indicates that the improved runtime is attributable to a reduction of time step, kick and drift formula evaluations.
The left plot of Figure \ref{fig:second-order} visualises energy error of evolving the softened Plummer sphere as described previously, but for $1$ $N$-body units and under varying time step accuracy $\eta$. As predicted, all three methods show second order behaviour. On the corresponding wall-clock time vs energy error plot on the right CC consistently outperforms HOLD, followed by BLOCK. We emphasise that the Plummer sphere is a spherically symmetric configuration with a smoothly changing mass distribution, and a non-zero softening length $\varepsilon$ sets an upper limit on the hardness of the binaries that can form during the simulation. Hence, we would not expect the CC scheme to have a significant advantage over the HOLD method.
\begin{figure*}
\centering
\includegraphics[width=0.85\textwidth]{figs/0_second_orderb.eps}
\caption{\label{fig:second-order}
Left: time step accuracy parameter $\eta$ vs energy error for integrating a $1024$-body Plummer sphere for $1$ $N$-body units. Right: corresponding wall-clock time vs energy error from the same set of tests.
}
\end{figure*}
\begin{figure*}
\centering
\includegraphics[width=0.8\textwidth]{figs/3_kk2010b.eps}
\caption{\label{fig:plummer-core-collapse} A comparison of the HOLD, CC and CC\_KEPLER methods on integrating an unsoftened $1024$-body Plummer sphere through core collapse: conservation of the integrals of motion (top two rows), evolution of the mass distribution of the solution (third row) and wall-clock time (bottom left). Lagrangian radii are plotted for 90\%, 50\%, 10\% and 1\% of the system mass. Wall-clock times are normalised by CC\_KEPLER at the first global time step ($1/64$-th of the simulation). On a laptop with a 1.3GHz Intel Core i5 processor, CC\_KEPLER integrates the first global time step in roughly $5$ minutes while the entire simulation takes about $7$ hours.
}
\end{figure*}
\begin{figure}
\centering
\includegraphics[width=0.49\textwidth]{figs/fractal_dimension_dE_time.eps}
\caption[]{\label{fig:fractal-initial-conditions} Energy conservation (top) and runtime (bottom) for integrating a $1024$-particle system of fractal initial conditions for $0.25$ $N$-body units as a function of the fractal dimension.
}
\end{figure}
\subsection{Unsoftened Plummer sphere test}\label{sec:plum}
We proceed by evolving an equal-mass $1024$-body Plummer sphere without softening through core collapse. We choose initial velocities consistent with a dynamic equilibrium, as in the softened case considered previously. This setup is chosen to match a test used on a modern implementation of a fourth-order Hermite scheme with block time steps in \cite[their section 3.4.1]{Konstantinidis:2010hx}.
In addition to the HOLD and CC schemes that have been introduced previously, we also test a modification of the CC scheme with a dedicated Kepler solver (CC\_KEPLER). In this scheme a Kepler solver is used for evolving connected components consisting of two particles. This is a form of algorithmic regularization of binaries, but note that the regularization follows naturally from the structure of the integrator and no separate binary detection or additional free parameters are necessary. The implementation of the Kepler solver is based on a universal variable formulation \citep{bate1971fundamentals}.
Results of the core collapse simulation are visualised in Figure \ref{fig:plummer-core-collapse}. All three methods produce solutions that are realistic in terms of the evolution of the mass distribution. Energy conservation is comparable to what is observed in \cite{Konstantinidis:2010hx}. Other integrals of motion show conservation around machine precision with the exception of a jump in the HOLD method around core collapse (this is caused by a high speed particle escaping from the system, causing a loss of precision in the force evaluations).
Before core collapse, execution times are roughly equivalent to the softened case considered previously (section~\ref{sec:smoothplum}) --- CC shows a modest improvement over HOLD, and CC\_KEPLER is very close to CC. Around core collapse, execution times of the HOLD and CC methods gradually increase by an order of magnitude (the CC method still consistently outperforms the HOLD method). In contrast, execution time used by the CC\_KEPLER method remains relatively uniform throughout the simulation, including core collapse.
The Sakura integrator achieves a similarly efficient treatment of close binaries by decomposing the evolution of an $N$-body Hamiltonian into a sequence of Kepler problems \citep{GoncalvesFerrari:2014uk}. The main source of errors in Sakura comes from many-body close encounters, as these are difficult to decompose into two-body interactions. In contrast, CC\_KEPLER only uses the binary solver for an isolated binary system, and switches to the regular many-body integrator when necessary (this is further discussed in Section \ref{sec:cc-ext}).
\subsection{Fractal distributions}\label{sec:fractaltest}
Since our new methods are based on the partitioning of the particle distribution in connected subsystems, we expect the method to be especially well suited to situations where substructure with extreme density contrasts exist. We therefore proceed by integrating a set of initial conditions developed with the aim of describing a star cluster with fractal substructure \citep{GoodwinWhitworth2004}. These initial conditions mimic the observed distribution of young stellar associations. They are parametrized by a fractal dimension: a low fractal dimension leads to an inhomogeneous (``structured'') distribution of stars whereas a high fractal dimension leads to a more homogenous (``spherical'') distribution (Figure \ref{fig:time-step-graph}). For the highest possible fractal dimension value of $3$, the initial conditions approximate a constant density sphere. We use $\eta=0.03$, and integrate a $1024$-particle system under an unsoftened potential for $0.25$ $N$-body units for varying fractal dimensions.
In Figure \ref{fig:fractal-initial-conditions} we plot the energy error and runtime of the simulation, averaged over 10 runs, as a function of the fractal dimension $f$. While all integrators show similar energy conservation, CC and CC\_KEPLER consistently outperform BLOCK and HOLD irrespective of fractal dimensions in terms of runtime. Further, runtime increases for decreasing fractal dimension for the BLOCK and HOLD integrators, while runtime remains essentially flat (and even decreases slightly) for decreasing fractal dimension for the CC and CC\_KEPLER integrators.
\subsection{Plummer sphere with binaries}\label{sec:binaries}
We proceed by looking at how our methods perform on systems containing a large number of binaries. Specifically, we take a Plummer sphere and replace every particle with a binary system. The positions and velocities of the particles are chosen such that under the absence of external perturbations, they would form a stable binary with a randomly oriented orbital plane, and a semi-major axis drawn uniformly in log space between $log(a)$ and $-0.5$. We integrate a system of $512$ binaries (=$1024$ individual particles) for $0.25$ $N$-body units with $\eta=0.03$.
Figure \ref{fig:plummer-of-binaries-scaling} visualises energy conservation and runtime of the initial conditions as a function of minimum semi-major axis $a$. For large $a$, the introduced binaries are generally unbounded, and the results are equivalent to evolving an ordinary Plummer sphere. As the minimum $a$ decreases, the introduced binaries become bounded and their interactions start dominating in the integration time, leading to a significant advantage for CC and CC\_KEPLER methods.
\begin{figure}
\centering
\includegraphics[width=0.49\textwidth]{figs/plummer_w_binary512.eps}
\caption{\label{fig:plummer-of-binaries-scaling}Energy conservation (top) and runtime (bottom) for integrating $512$-binary (=$1024$-particle) Plummer sphere for $0.25$ $N$-body units as a function of the initial semi-major axis $a$.}
\end{figure}
\subsection{Cold collapse test}\label{sec:coldcollapse}
As a final test we evaluate the performance of our integrators in a cold collapse scenario. Specifically, we use the fractal initial conditions described in Section \ref{sec:fractaltest} with the initial velocities set to zero. We consider a ``structured'' case with the fractal dimension $f_d=1.6$, and a ``spherical'' case with $f_d=3.0$. We evolve initial conditions for 2 $N$-body time units. For the spherical case, this is past the moment of collapse that occurs around $1.5$ $N$-body time units. For the structured case the moment of collapse is less well-defined, as different substructures collapse at different times.
We compare CC and CC\_KEPLER to two recent $N$-body codes, Ph4 (McMillan, in preparation) and HiGPUs \citep{CapuzzoDolcetta2013}. Both codes use a Hermite scheme with conventional block time steps. Ph4 implements a fourth-order scheme with the option of using the GPU-accelerated SAPORRO library \citep[][and B\'{e}dorf et al. in prep.]{Gaburov2009}. HiGPUs implements a sixth-order scheme and requires a GPU to run. We conduct our tests on a workstation --- running on a single core of an Intel i7-2720QM CPU, and a GTX460M GPU. The FLOPS performance of the GPU is roughly 40 times larger than a single core of the CPU. The hardware setup is thus indicative only of the intrinsic algorithmic scaling, rather than representative of the performance in production simulations (which would use multiple and/or more powerful CPUs/GPUs). We use unsoftened potential for CC, CC\_KEPLER, and Ph4 without GPU acceleration. We use a very small softening parameter ($\varepsilon=10^{-4}$) for Ph4 with GPU acceleration, and HiGPUS, as these run into severe slowdowns and/or crashes with unsoftened gravity --- probably because of the limited precision of their GPU kernels. We set code-specific time step accuracy parameters to $\eta=0.01$ for CC/CC\_KEPLER, $\eta_4=0.1$ for Ph4, and $\eta_4=0.05$ and $\eta_6=0.6$ for HiGPUs.
Figure \ref{fig:cold-collapse-scaling} visualises energy conservation, momentum conservation and the wall-clock time of the initial conditions as a function of the system size for structured and homogenous initial conditions. In the spherical case, setups that take advantage of the GPU (PH4\_gpu and HiGPUs) outperform the alternatives, but note that both CC and CC\_KEPLER show very similar scaling to Ph4 without GPU acceleration. In contrast, for structured case, CC\_KEPLER and CC show a marked speed up in comparison with the conventional block time step schemes, being faster for this particular calculation than Ph4\_GPU and HiGPUs, despite the latter having the advantage of using the GPU acceleration and integrating with softened gravity. The differences between the structured and the spherical cases highlight the relative advantage that the connected component approach has with respect to conventional block time steps when applied to multi-scale initial conditions.
\begin{figure*}
\centering
\includegraphics[width=0.32\textwidth]{figs/cold_collapse_dE_3.0.eps}
\includegraphics[width=0.32\textwidth]{figs/cold_collapse_dp_3.0.eps}
\includegraphics[width=0.32\textwidth]{figs/cold_collapse_time_3.0.eps}
\includegraphics[width=0.32\textwidth]{figs/cold_collapse_dE_1.6.eps}
\includegraphics[width=0.32\textwidth]{figs/cold_collapse_dp_1.6.eps}
\includegraphics[width=0.32\textwidth]{figs/cold_collapse_time_1.6.eps}
\caption{\label{fig:cold-collapse-scaling}
Energy conservation (left column), momentum (middle column) and wall-clock time (right column) as a function of system size $N$ for the cold collapse test. The top row shows results for a homogenous sphere (fractal dimension $f_d=3$), while the bottom row shows a highly structured fractal ($f_d=1.6$). We evolve the initial conditions for 2 $N$-body time units, and plot the mean values across 5 runs. CC, CC\_KEPLER and Ph4 use unsoftened gravity ($\varepsilon=0$) , while PH4\_gpu and HIGPUS use very small softening ($\varepsilon=10^{-4}$).}
\end{figure*}
\section{Discussion}\label{sec:discussion}
\subsection{Using and extending the CC method}\label{sec:cc-ext}
We introduced a novel method for direct integration of $N$-body systems based on splitting the system Hamiltonian using connected components of the time step graph (the CC split). We were motivated by the need for a more efficient divide-and-conquer strategy for reducing the intractable Hamiltonian (Eq.~\ref{eq:Hamiltonian}) to the least possible number of analytically solvable Hamiltonians (Eqs.~\ref{eq:kick},\ref{eq:drift}). In comparison to existing splitting methods, notably the HOLD split introduced in \cite{Pelupessy:2012if}, the CC split is particularly effective at splitting multi-scale systems. We have not encountered a situation where the HOLD split would be preferable over the CC split. The practical advantages of Hamiltonian splitting are similar to what is usually achieved with block time-steps. However, as our splitting methods, including the CC method, do not extrapolate particle states for evaluating the total force acting on a particle, we conserve linear and angular momentum to machine precision.
We went on to show on the example of the CC\_KEPLER method that the connected components partitioning has additional uses beyond improved splitting efficiency. Specifically, we were able to incorporate regularization of two-body close encounters by simply checking for the condition where the successive partitioning leads to a connected component with two particles, and evolving the corresponding two-body Hamiltonian (Eq \ref{eq:kepler}) using a dedicated Kepler solver. This approach can be extended to many-body close encounters by using a suitable specialised solver \citep[or e.g. chain regularization methods,][]{Mikkola2008} to evolve isolated Hamiltonians corresponding to connected components with certain properties. Possible selection criteria include having a specific number of particles and/or a maximum time step below a threshold value or the structure of the timestep graph.
The numerical experiments of Section \ref{sec:tests} were chosen to mainly study the splitting aspect of $N$-body integration. We focused on normalised performance metrics, and the scaling of the wall-clock time as a function of the ``multi-scaleness'' in the initial conditions. Our current implementations would benefit from additional optimisations typically used in production-level $N$-body codes. Specifically, there is inherent parallelism in the CC method, as recursive calls for evolving successively smaller closed Hamiltonians only affect the state of the particles in the ``current'' closed component. It may be possible to parallelise the method based on this property. However, tests show that a naive approach does not scale well due to load-balancing issues, as subsystems can vary substantially in size. Alternatively, it could be feasible to implement the CC method on a GPU, as
the major components --- $N$-body force evaluation\citep{PortegiesZwart:2007gl, Belleman:2008fv, CapuzzoDolcetta2013} and graph processing algorithms \citep{Harish07} --- have individually been successfully implemented on GPUs.
It may be possible to speed up the evaluation of long-range interactions between different connected components ($V_{CC}$ in the CC decomposition formula) through a centre-of-mass (or multipole) approximation that form the basis of tree codes \cite{Barnes:1986ed}. As long-range interactions between two connected components are evaluated symmetrically, this approach could make it possible to obtain most of the speedup of a tree code while maintaining good linear and angular momentum conservation. A potential pitfall with this approach could arise from the fact that the time step criterion $\tau(i,j)$ used in finding the connected components is only partially determined by the coordinates of the particles. As such, particles in the same connected component may occupy a ``non-compact'' region in physical space, making multipole approximation difficult.
\subsection{Formally optimal Hamiltonian splitting}
The HOLD integrator determines the accuracy of a kick between particles $i$ and $j$ from the particle-based time steps $\tau(i)$ and $\tau(j)$. In the CC integrator the accuracy of a kick is determined by the time step graph generated directly from interaction-based time steps $\tau(i,j)$. Could we further improve the splitting by applying kicks directly based on the interaction-specific time step criteria $\tau(i,j)$?
We implemented this idea in an experimental integrator which we named the OK split (OK stands for \emph{Optimal Kick}). The method partitions a list of all {\em interactions} in the system (based on a pivot time step $h$) just like the HOLD split partitions a list of all {\em particles} in the system. The partitioning is formally optimal in the sense that every kick is evaluated at the time step closest to $\tau(i,j)$ in the power of two hierarchy based on the pivot time step $h$. While the possibility of direct $N$-body integration with interaction-based time steps has been previously considered in \cite{Nitadori:2008gt}, the OK split is the first workable implementation of this idea that we are aware of.
In numerical tests, the OK split is not competitive compared to other methods such as the CC split. For example, in the $1024$-body smoothed Plummer sphere test from Section \ref{sec:smoothplum}, the relative energy error at the end of the simulation is around $10^{-2}$ (several orders of magnitude worse than HOLD and CC, but possibly still enough for drawing statistically correct conclusions, \cite{PortegiesZwart2014}). The remaining integrals of motion are conserved at machine precision, as the OK split applies kicks in pairs. Finally, the evolution of the mass distribution is comparable to HOLD and CC with the OK split using fewer kick and time step evaluations.
Could we improve the OK split by changing the time step criteria? For example, consider $\tau_{\star}(i,j)=\min\left(\tau(i),\tau(j)\right)$ where $\tau$ is the particle-based time step criteria as defined in \cite{Pelupessy:2012if}. Formally, combining the OK split with $\tau_{\star}(i,j)$ would result in a splitting with the exact same kicks and drifts as the HOLD integrator. This somewhat contrived example only serves the point of illustrating that the time step criteria can qualitatively change the behaviour of the OK split. While it is unknown whether practical interaction-based time step criteria even exist, we do believe that a closer look at the various simplifications made during the derivation of the explicit and approximately time-symmetric time step criteria that we've used throughout this work (eq \ref{eq:tauij}) would serve as a good starting point.
\begin{acknowledgements}
We thank Guilherme Gon\c{c}alves Ferrari and the anonymous referee
for a critical reading of the manuscript. This work was supported by the Netherlands Research Council NWO (Grants \#643.200.503, \#639.073.803 and \#614.061.608) and by the Netherlands Research School for Astronomy (NOVA). J\"urgen J\"anes was supported by the Archimedes Foundation, Estonian Students' Fund USA, Estonian Information Technology Foundation and Skype.
\end{acknowledgements}
\bibliographystyle{aa}
|
\section{Introduction}
In 1927, Artin proposed the following conjecture: \emph{If $g$ is not a square and $g\neq -1$, then there are infinitely many primes $p$ for which $g$ is a primitive root modulo $p$.} Artin's conjecture remains unsolved, but investigations in this direction have led to many deep and beautiful results (see \cite{moree12}).
In 1967, Hooley \cite{hooley67} showed that Artin's conjecture is a consequence of the Generalized Riemann Hypothesis for Dedekind zeta functions (hereafter GRH). In \cite{pollack14}, it was shown how Hooley's proof could be merged with the method of Maynard--Tao for producing bounded gaps between primes: \emph{On GRH, for every nonsquare $g\neq -1$ and every $m$, there are infinitely many runs of $m$ consecutive primes all possessing $g$ as a primitive root and lying in an interval of length $O_m(1)$.}
There is not a single $g$ for which the conclusion of Artin's conjecture is known to hold unconditionally. However, in 1984 Gupta and Ram Murty \cite{GM84} described how to produce many finite sets of integers some member of which satisfies Artin's conjecture. Their method was refined by Ram Murty and Srinivasan \cite{MS87}, Gupta, Ram Murty, and Kumar Murty \cite{GMM87}, and by Heath-Brown \cite{HB86}. It follows from the results in this last paper that Artin's conjecture holds for at least one $g \in\{2, 3, 5\}$. We prove a result in this direction where the primes produced are consecutive and contained in an interval of bounded length.
Recall that nonzero $q_1, \dots, q_r\in \Z$ are said to be \emph{multiplicatively independent} if $q_1^{e_1} \cdots q_r^{e_r}=1$ in integers $e_1, \dots, e_r$ only when $e_1 = \dots = e_r = 0$.
\begin{thm}\label{thm:main} Let $\Qq$ be a set of $r$ multiplicatively independent integers. Assume that the elements $q_1, \dots, q_r$ of $\Qq$ satisfy the following technical condition:
\[\tag{*} \parbox{0.8\linewidth}{If $e_0, e_1, \dots, e_r $ are nonnegative integers for which $(-3)^{e_0} q_1^{e_1}\cdots q_r^{e_r}$ is a square, then $\sum_{i=0}^{r} e_i$ is even.}\]
Let $m$ be a natural number. If $r \ge \exp(C m)$, then there are infinitely many runs of $m$ consecutive primes $p_1 < \dots < p_m$ all of which possess some element of $\Qq$ as a primitive root, where also
\[ p_m - p_1 \leq \mathfrak{f}(\Q(\sqrt{q_1}, \dots, \sqrt{q_r})/\Q) \cdot \exp(C'm).\]
Here $C$ and $C'$ are (positive) absolute constants, and $\mathfrak{f}(K/\Q)$ denotes the conductor of the abelian extension $K/\Q$.
\end{thm}
\begin{remark} Of course, (*) holds whenever $q_1, \dots, q_r$ are distinct (positive) primes.\end{remark}
The techniques used to attack Artin's conjecture can also be used to answer statistical questions about reductions of elliptic curves. Here the general setup is as follows:
Let $E/\Q$ be an elliptic curve. For all but finitely many primes $p$, one can reduce $E$ mod $p$ to obtain an elliptic curve defined over $\F_p$. What can one say about the structure of the group $E(\F_p)$ as $p$ varies? It is known that $E(\F_p)$ is always generated by two elements, and so it is particularly natural to ask when one suffices. In other words, how often is $E(\F_p)$ a cyclic group?
If all of the $2$-torsion of $E$ is defined over $\Q$, then $(\Z/2\Z)^2$ sits inside $E(\Q)$, and so $E(\F_p)$ is cyclic for at most finitely many primes $p$. So assume $E$ has an irrational $2$-torsion point. Assuming GRH, Serre showed that there are infinitely many primes $p$ with $E(\F_p)$ cyclic, using Hooley's approach \cite{hooley67} to Artin's conjecture. In fact, Serre \cite{serre79} obtained an asymptotic formula for the number of such $p \leq x$, as $x \to \infty$. Ram Murty \cite{murty83} showed that when $E$ has CM, Serre's asymptotic formula can be proved unconditionally; a simpler argument for the same conclusion has been given by Cojocaru \cite{cojocaru03}. See \cite{CM04} and \cite{AM10} for investigations into the size of the error term in Serre's formula.
We prove the following bounded gaps result.
\begin{thm}\label{thm:EC} Let $E/\Q$ be an elliptic curve with an irrational $2$-torsion point. Let $m$ be a natural number. If GRH holds, then there are infinitely many runs of $m$ consecutive primes $p_1 < p_2 < \dots < p_m$ for which $E(\F_p)$ is cyclic, where
\[ p_m-p_1 \leq \mathrm{rad}(\Delta_{E}) \cdot \exp(C'' m). \]
Here $\mathrm{rad}(\Delta_{E})$ is the product of the primes of bad reduction, and $C''$ is an absolute constant. If $E$ has $CM$, then the GRH assumption can be removed.
\end{thm}
The CM case of Theorem \ref{thm:EC} is particularly easy because of the abundance of supersingular primes. According to a criterion of Deuring (see, e.g., \cite[Theorem 12, p. 182]{lang87}), a prime $p$ of good reduction is supersingular precisely when there is a unique prime in the CM field lying above $p$. As we explain below, this implies that $\E(\F_p)$ is cyclic for all primes from a certain arithmetic progression. This allows us to appeal to a recent theorem of Banks--Freiberg--Turnage-Butterbaugh \cite{BFTB14} about long runs of such primes in short intervals.
It is perhaps slightly unsettling that we produce only supersingular primes in the CM case. In general, this is unavoidable. For instance, consider the curve $E$ given by $y^2=x^3+x$, whose $2$-torsion points are defined over $\Q(i)$. Since $E$ has CM by $\Z[i]$, Deuring's criterion tells us that a prime $p$ of good ordinary reduction splits in $\Q(i)$, and so $E(\F_p)$ contains $(\Z/2\Z)^2$ for all such $p$. Our final theorem says that \emph{if} there are infinitely many $p$ of good ordinary reduction with $E(\F_p)$ cyclic, then the set of these $p$ has bounded gaps.
\begin{thm}\label{thm:CM} Let $E/\Q$ be a CM elliptic curve. Assume that there are infinitely many primes $p$ of good ordinary reduction for which $E(\F_p)$ is cyclic. Then there are infinitely many tuples of $m$ such primes $p_1 <\dots < p_m$ with $p_m-p_1 \ll \exp(O_E(m))$.
\end{thm}
\noindent Unfortunately, the method of proof of Theorem \ref{thm:CM} does not allow us to impose the condition that the primes produced here are consecutive.
It would be desirable to remove the GRH assumption altogether from Theorem \ref{thm:EC}. We note that in \cite{GM90}, Gupta and Ram Murty showed unconditionally that if $E$ has an irrational $2$-torsion point, then there are always infinitely many primes $p$ with $E(\F_p)$ cyclic (but they do not get the order of magnitude for the count predicted by Serre's asymptotic formula). Their proof relies on a sieve result seemingly unavailable in our context.
\subsection*{Notation} The letters $\ell$ and $p$ are reserved for primes. We write $p^{-}(n)$ for the smallest prime factor of $n$, with the convention that $p^{-}(1)=\infty$. We use $\mathrm{rad}(n)$ to denote the largest squarefree divisor of $n$. We use $C_1, C_2, \dots$ for absolute positive constants that are be thought of as large. If $F$ is a number field, $\Z_F$ denotes its ring of integers, and we write $\Delta_F$ for the absolute discriminant of $F$. If $E$ is an elliptic curve defined over $\Q$, we let $\Delta_E$ denote the minimal discriminant of $E/\Q$. We write $\Prob(\cdot)$ for the probability of an event and $\E[\cdot]$ for the expectation of a random variable.
\section{Preliminaries for the proof of Theorem \ref{thm:main}}
If $q_1, \dots, q_r \in \Z$ and $p\nmid q_1 \cdots q_r$, we write $\langle q_1, \dots, q_r\bmod{p}\rangle$ for the subgroup of $\F_p^{\times}$ generated by the mod $p$ reductions of the $q_i$. The next lemma is due to Ram Murty and Srinivasan \cite{MS87} (compare with \cite[Lemma 2]{GM84}).
\begin{lem}\label{lem:MS} Let $q_1, \dots, q_r$ be multiplicatively independent integers, and let $Y \ge 1$. The number of primes $p$ for which
\[ \#\langle q_1, \dots, q_r \bmod{p}\rangle \leq Y \]
is $O(Y^{1+ \frac{1}{r}})$. Here the implied constant may depend on the $q_i$.
\end{lem}
\begin{proof} We include the short proof. Suppose that $\#\langle q_1, \dots, q_r \bmod{p}\rangle \leq Y$. By the pigeonhole principle, as $e_1, \dots, e_r$ run independently from $0$ through $\lfloor Y^{1/r}\rfloor$, two expressions of the form $q_1^{e_1} \cdots q_r^{e_r}$ must coincide mod $p$. Consequently, for some choice of integers $e_i'$ with each $|e_i'| \le Y^{1/r}$ and not all $e_i'=0$,
$p$ divides the numerator of the nonzero rational number $q_1^{e_1'} \cdots q_r^{e_r'}-1$. This numerator is (crudely) bounded above by $2\max\{|q_1|, \dots, |q_r|\}^{rY^{1/r}}$ and so has $O(Y^{1/r})$ prime divisors. Summing over the $O(Y)$ possibilities for the $e_i'$ completes the proof.
\end{proof}
The following lemma is used to construct an admissible collection of linear functions to which Maynard's machinery can be applied.
\begin{lem}\label{lem:HB} Let $q_1, \dots, q_r$ be nonzero integers satisfying {\normalfont\rmfamily{(*)}}. Let $v = 16 \prod_{\ell \mid q_1\cdots q_r,~\ell>2} \ell$. One can select an integer $u$ coprime to $v$ so that both of the following hold:
\begin{enumerate}
\item[\normalfont\rmfamily{(1)}] For every $p\equiv u\pmod{v}$, the Legendre symbols $\leg{q_1}{p}=\dots=\leg{q_r}{p}=-1$.
\item[\normalfont\rmfamily{(2)}] If $T$ is the largest power of $2$ dividing $u-1$, then $T \in \{2, 4, 8\}$, and $\gcd(\frac{u-1}{T},v)=1$.
\end{enumerate}
\end{lem}
\begin{proof} For $r=3$, this lemma was proved by Heath-Brown. Since the argument for the general case is the same, we only outline the main steps, referring the reader to \cite[pp. 35--36]{HB86} for the details. By estimating $\sum_{p \le x}\left(1-\leg{-3}{p}\right)\prod_{i=1}^{r} \left(1-\leg{q_i}{p}\right)$ from below --- keeping (*) in mind --- one shows that there are infinitely many primes $p$ with $\leg{-3}{p} = \leg{q_1}{p} = \dots = \leg{q_r}{p}=-1$. Fix one and call it $p_0$. For each odd prime $\ell$ dividing $q_1 \cdots q_r$, put
\[ u_{\ell} = \begin{cases} p_0 &\text{if $\ell \nmid p_0-1$,}\\
4p_0 &\text{otherwise},
\end{cases}
\quad\text{and put}\quad
u_2 = \begin{cases} p_0 &\text{if $16 \nmid p_0-1$}, \\
p_0-8 &\text{otherwise}.
\end{cases}
\]
Then for all odd primes $\ell \mid q_1\cdots q_r$, we see that $\ell \nmid u_{\ell}-1$. (We have used here that $p_0 \equiv -1\pmod{6}$, since $\leg{-3}{p}=-1$.) One checks that it suffices to choose $u$ as a solution to the simultaneous congruences
\[ u \equiv u_{\ell} \pmod{\ell}~\forall \text{ odd $\ell \mid q_1\dots q_r$} \quad\text{and} \quad u \equiv u_2 \pmod{16}. \qedhere \]
\end{proof}
Let $\Ll$ be a set of $k$ distinct linear functions, say $L_1(n) = a_1 n + b_1, \dots, L_k(n)= a_k n + b_k$, where each $a_i, b_i \in \Z$ and every $a_i>0$. We say that $\Ll$ is \emph{admissible} if for each prime $p$, there is some integer $n_p$ for which $p \nmid \prod_{i=1}^{k} L_i(n_p)$. Note that if each $(a_i, b_i)=1$, to check admissibility it suffices to check primes $p \leq k$.
\begin{lem}\label{lem:admissible} Let $q_1, \dots, q_r$ be nonzero integers satisfying {\normalfont\rmfamily{(*)}}, and let $u$ and $v$ be chosen as in Lemma \ref{lem:HB}. Let $\kappa$ be a natural number. There are integers $a_1 < \dots < a_{\kappa}$, each congruent to $u\bmod{v}$, for which the $2\kappa$ linear functions
\begin{alignat*}{3} L_1(n) &= vn + a_1, &\quad \dots, \quad &L_{\kappa}(n) = vn + a_{\kappa}, \\
\tilde{L}_1(n) &= \frac{v}{T}n + \frac{a_1-1}{T}, &\quad \dots, \quad &\tilde{L}_{\kappa}(n) = \frac{v}{T} n + \frac{a_{\kappa}-1}{T} \end{alignat*}
make up an admissible family. Moreover, we can select the $a_{i}$ in such a way that
\[ a_{\kappa} - a_1 \leq v\cdot (2\kappa)^{C_1}. \]
\end{lem}
\begin{proof} By the fundamental lemma of the sieve, if $C_1$ is large enough, then the number of integers $A \in [0,(2\kappa)^{C_1}]$ for which $p^{-}((vA+u)(\frac{v}{T}A+\frac{u-1}{T})) > 2\kappa$ exceeds \[ \frac{1}{2} ((2\kappa)^{C_1}) \prod_{p \leq 2\kappa} (1-2/p).\]
Increasing $C_1$ if necessary, this lower bound exceeds $\kappa$. Pick $\kappa$ of these integers, say $A_1 < \dots< A_{\kappa}$. The theorem follows upon choosing $a_i = vA_i + u$. Indeed, for primes $p \leq 2\kappa$, we have arranged matters so that $p \nmid \prod_{i=1}^{\kappa} L_{i}(0) \tilde{L}_{i}(0)$.
\end{proof}
\begin{remark} In the next section, we will show that all of the $\tilde{L}_i$ are almost primes at the same time that several of the ${L}_i$ are prime. A similar strategy appears in work of Li and Pan \cite{LP14}, who seem to have been the first to notice that the Maynard--Tao method can be applied with auxiliary `almost prime' conditions added. In the context of the earlier GPY method, this observation was made by Pintz \cite{pintz10}.
\end{remark}
\section{Proof of Theorem \ref{thm:main}}
The following key proposition is contained in recent work of Maynard \cite{maynard14d}.
\begin{prop}\label{prop:maynard} Fix an admissible family $\Ll$ of $k$ distinct linear functions, where $k \ge 2$. Suppose that $x$ is sufficiently large, $x > x_0(\Ll)$. There is a probability measure on
\[ \A(x):= \{n \in \Z: x \leq n < 2x\} \]
with all of the following properties:
\begin{enumerate}
\item[\normalfont\rmfamily{(1)}] The probability mass at any single $n \in \A(x)$ is \[ \ll x^{-1} (\log{x})^k \left(\prod_{i=1}^{k} \prod_{p \mid L_i(n)} 4\right) \exp(O(k\log{k})). \]
\item[\normalfont\rmfamily{(2)}] For each $L \in \Ll$,
\[ \Prob(L(n)\text{ is prime})\gg \frac{\log{k}}{k}. \]
\item[\normalfont\rmfamily{(3)}] Suppose that $\rho \in [k\frac{(\log\log{x})^2}{\log{x}}, \frac{1}{25}]$. For each $L \in \Ll$,
\[ \E\bigg[\sum_{\substack{p \mid L(n) \\ p < x^{\rho}}} 1\bigg] \ll \rho^2 k^4 (\log{k})^2. \]
\item[\normalfont\rmfamily{(4)}] Suppose that $L(n)=a_0n+b_0$ is a linear function not belonging to $\Ll$. Suppose also that $|a_0|, |b_0| \le x$ and that $\Delta_L$, defined by
\[ \Delta_L:= a_0 \prod_{j=1}^{k} |a_0 b_j - b_0 a_j|, \]
is nonzero. Then
\[ \Prob(p^{-}(L(n)) > x^{1/25}) \ll \frac{\Delta_L/\phi(\Delta_L)}{\log{x}}. \]
\end{enumerate}
Although $x_0$ may depend on $\Ll$, all implied constants in this statement are absolute.
\end{prop}
\begin{proof}[Proof (sketch)] This follows from \cite[Proposition 6.1]{maynard14d}. In the setup of that proposition, $\A$ is the set of natural numbers, $\Ll$ is as above, $\curly{P}$ is the set of all primes, $B=1$, $\theta = 2/5$, and $\alpha=1$. The probability measure on $\A(x)$ assigns to each $n$ the probability mass $w(n)/\sum_{n \in \A(x)} w(n)$. Our (1) follows from Proposition 6.1(1) together with the immediately preceding estimate for $w_n$; we also use Maynard's lower bounds on $\mathfrak{S}(\Ll)$ and $I_k(F)$ given in (8.2) and Lemma 8.6, respectively. Our (2) is deduced from Proposition 6.1(1,2); here we use the estimate $J_k/I_k \gg \log{k}/k$ and the observation that for each $L(n) = a_L n + b_L \in \Ll$, we have (in Maynard's notation)
\[ \#\curly{P}_{L,\A}(x) = \sum_{\substack{a_L x+b_L \leq p < 2a_L x+b_L \\ p \equiv b_L\pmod{a_L}}} 1 \sim \frac{1}{\phi(a_L)} \frac{a_L x}{\log{x}} \sim \frac{a_L}{\phi(a_L)}\ \frac{\#\A(x)}{\log{x}}. \]
Our (3) comes from Proposition 6.1(1,4), and (4) comes from Proposition 6.1(1,3).
\end{proof}
We now prove Theorem \ref{thm:main}.
\begin{proof} Assume that $r \ge \exp(C_2 m)$, and let $\kappa = \lceil \exp(C_3 m)\rceil$. Let $c$ be a small positive absolute constant. The necessary constraints on the constants $C_2$, $C_3$, and $c$ will emerge in the proof.
Let $q_1', \dots, q_r'$ be the integers obtained from $q_1, \dots, q_r$ by replacing each $q_i$ with its squarefree part. That is, $q_i'$ is the unique squarefree integer for which $q_i/q_i'$ is a square. Since $q_1, \dots, q_r$ satisfy (*), so do $q_1', \dots, q_r'$. Let $k=2\kappa$, and let $L_i$ and $\tilde{L}_i$, for $1 \leq i \leq \kappa$, be the linear functions produced by Lemma \ref{lem:admissible} applied with $q_1', \dots, q_r'$. Every prime dividing $q_1' \cdots q_r'$ divides $f:=\mathfrak{f}(\Q(\sqrt{q_1}, \dots, \sqrt{q_r})/\Q)$, and thus $v = 16 \prod_{\ell \mid q_1'\dots q_r',~\ell >2}\ell$ divides $16f$.
We now invoke Proposition \ref{prop:maynard}. We will show that with positive probability, an $n \in \A(x)$ satisfies all of
\begin{enumerate}
\item[(i)] at least $m$ of $L_1(n)$, \dots, $L_\kappa(n)$ are prime,
\item[(ii)] $p^{-}(L_i(n)) \ge x^{\frac{c}{k^3\log{k}}}$ and $p^{-}(\tilde{L}_i(n)) \ge x^{\frac{c}{k^3\log{k}}}$ for all $i=1,\dots, \kappa$,
\item[(iii)] all integers in the interval $[L_1(n),L_\kappa(n)]$ that are not one of the $L_i(n)$ are composite,
\item[(iv)] whenever $p=L(n)$ is prime with $L \in \{L_1, \dots, L_{\kappa}\}$, $p$ possesses some element of $\Qq$ as a primitive root.
\end{enumerate}
If (i)--(iv) hold for $n$, then the set of primes in $[L_1(n), L_{\kappa}(n)]$ has at least $m$ elements, each one of which possesses some element of $\Qq$ as a primitive root. Moreover, the difference between the largest and smallest such primes is at most
\[ L_{\kappa}(n) - L_1(n) = a_{\kappa}-a_1 \leq v (2\kappa)^{C_1} \leq f \exp(C_4 m), \]
provided that $C_4$ is large enough in terms of $C_1$ and $C_3$. Thus, we obtain Theorem \ref{thm:main} with $C = C_2$ and $C' = C_4$.
To begin analyzing (i)--(iv), let $\Pp$ be the set of primes, and consider the random variable $X := \sum_{i=1}^{\kappa} \mathbf{1}_{\Pp}(L_i(n))$. Proposition \ref{prop:maynard}(2) and our choice of $\kappa$ yield $\E[X] \gg C_3 m$. We assume $C_3$ is large enough that $\E[X] > m$. Noting the inequality \[ \mathbf{1}_{X\geq m} \geq \kappa^{-1}(X-(m-1)), \] and taking expectations, we find that (i) holds with probability at least $\kappa^{-1}$.
Let $L$ be one of the linear functions in $\Ll$. Then \[\Prob(p^{-}(L(n)) < x^{c/(k^3\log{k})})\le \E\bigg[\sum_{p \mid L(n),~p < x^{c/(k^3\log{k})}}1\bigg].\] So from Proposition \ref{prop:maynard}(3), (ii) fails with probability
\[ \ll k \cdot \left(\frac{c}{k^3\log{k}}\right)^2 k^4 (\log{k})^2. \]
We may assume $c>0$ is small enough that the odds of failure are less than $\frac{1}{2}\kappa^{-1}$. Then (i) and (ii) hold simultaneously with probability at least $\frac{1}{2}\kappa^{-1}$.
We claim that (iii) fails with probability $o(1)$, as $x\to\infty$. It is enough to show that if $a$ is a fixed integer from $[a_1, a_\kappa]$, and $a \not\in\{a_1,\dots, a_\kappa\}$, then the probability that $L(n):=vn+a$ is prime is $o(1)$. This follows immediately from Proposition 6.1(4) if $L$ is not a rational multiple of any $L_i$ or $\tilde{L}_i$. Since $L$ has leading coefficient $v$ and $a\not\in \{a_1,\dots, a_\kappa\}$, $L$ is not a multiple of any $L_i$. Since each $\tilde{L}_i$ has leading coefficient $v/T$, if $L$ is a multiple of some $\tilde{L}_i$, then $L = T \tilde{L_i}$; but then $T \mid L(n)$ and so $L(n)$ is composite for all $n\in \A(x)$.
Now assume that (ii) holds but that (iv) fails. We will show that this occurs with probability $o(1)$, as $x\to\infty$. In view of our previous estimates, this will complete the proof of Theorem \ref{thm:main}.
Assume that $p=L(n)$ is prime, with $L \in \{L_1, \dots, L_{\kappa}\}$, but that $p$ fails to have any $q\in \Qq$ as a primitive root. From Lemma \ref{lem:HB}(1) and our choice of $\Ll$, each $q \in \Qq$ is is a nonsquare modulo $p$. Thus, for each $q \in \Qq$, there is a prime $s=s_q$ dividing $(p-1)/T$ for which $q$ is an $s$th power modulo $p$. Put $t= \lceil 2c^{-1} k^3\log{k}\rceil$. Since (ii) holds, $\Omega((p-1)/T) \leq t$. (We assume here, as elsewhere in the proof, that $x$ is sufficiently large.) Recalling that $k=2\lceil \exp(C_3m)\rceil$, we assume $C_2$ is large enough that
\[ \exp(C_2 m) > (t-1)t. \]
Since $\#\Qq = r \geq \exp(C_2 m)$, the pigeonhole principle guarantees that at least $t$ values of $q \in \Qq$ share the same value of $s_q$; call this common value $s$. Relabeling, we can assume these are $q_1,\dots, q_t$. Then
\[ \#\langle q_1, \dots, q_t \bmod{p}\rangle \leq \frac{p-1}{s}\le \frac{p-1}{p^{-}((p-1)/T)}\leq x^{1-\frac{c}{2 k^3\log{k}}} \le x^{1-\frac{1}{t}}. \]
By Lemma \ref{lem:MS}, $p$ is restricted to a set of size $\ll_{\Qq} (x^{1-1/t})^{1+1/t} = x^{1-\frac{1}{t^2}}$.
Given $L$, the prime $p=L(n)$ determines $n$, restricting $n$ also to a set of size $O_{\Qq}(x^{1-1/t^2})$. Since there are $O_m(1)$ possibilities for $L$, the number of $n\in \A(x)$ for which (ii) holds but (iv) fails is $O_{\Qq,m}(x^{1-1/t^2})$.
Since $n$ satisfies (ii), each of $L_1(n),\dots,L_{\kappa}(n)$ has at most $t$ prime factors. So from Proposition \ref{prop:maynard}(i), the probability mass at $n$ is at $O_m(x^{-1} (\log{x})^{k})$. Thus, the probability of selecting an $n$ detected in the previous paragraph is
$O_{m,\Qq}((\log{x})^k x^{-1/t^2})$, which is $o(1)$ as $x\to\infty$.
\end{proof}
\section{Preparation for the proof of Theorem \ref{thm:EC}}
We begin with some background on elliptic curves. For each prime $\ell$, let $K_\ell$ denote the $\ell$-torsion field $\Q(E[\ell])$. It is well-known and easy to check that $K_\ell$ is a Galois extension of $\Q$. Now let $p$ be a prime of good reduction for $E$. Clearly, $E(\F_p)$ is cyclic if and only if it does not contain $(\Z/\ell\Z)^2$ for any prime $\ell$.
The following lemma, due to Ram Murty \cite[p. 159]{murty83}, shows that whether or not $E(\F_p)$ is cyclic amounts to a series of conditions on the splitting of $p$ in the fields $K_\ell$.
\begin{lem}\label{lem:murty} Let $p$ be a prime of good reduction for $E$. If $\ell$ is a prime with $\ell \neq p$, then $E(\F_p)$ contains $(\Z/\ell\Z)^2$ if and only if $p$ splits completely in $K_{\ell}$. As a consequence, $\E(\F_p)$ is cyclic if and only if for all primes $\ell \neq p$,
\begin{equation}\label{eq:notsplit} \text{$p$ does not split completely in $K_{\ell}$}. \end{equation}
\end{lem}
\begin{remark} If $E(\F_p)$ contains $(\Z/\ell\Z)^2$, then $\ell^2 \mid \#E(\F_p) \leq (\sqrt{p}+1)^2$, and so it suffices to test \eqref{eq:notsplit} for
\begin{equation}\label{eq:sufficetotest} \ell \le \sqrt{p}+1. \end{equation}
\end{remark}
If we assume the GRH, then the following theorem of Lagarias and Odlyzko \cite{LO77} gives a satisfactory estimate for the frequency with which primes split completely in $K_\ell$. We state the result incorporating a small improvement by Serre \cite[\S2.4]{serre81}.
\begin{prop}[Effective Chebotarev theorem, on GRH]\label{thm:serre} Let $K$ be a finite Galois extension of $\Q$, and let $\C$ be a conjugacy class of $\Gal(K/\Q)$. The number of unramified primes $p \leq x$ with $\legq{K/\Q}{p} = \C$ is given by
\[ \frac{\#\C}{[K:\Q]}\Li(x)+ O\left(\#\C \cdot x^{1/2}\left(\frac{\log|\Delta_K|}{[K:\Q]} + \log{x}\right)\right), \]
for all $x\geq 2$. Here the $O$-constant is absolute.
\end{prop}
\begin{remark} To estimate the $O$-term, we will use the following estimate valid for any Galois extension $K/\Q$ (see \cite[Proposition 6]{serre81}):
\begin{equation}\label{eq:hensel} \frac{1}{[K:\Q]} \log|\Delta_K| \leq \log~[K:\Q]+ \sum_{p \mid \Delta_K} \log{p}. \end{equation}
\end{remark}
To apply \eqref{eq:hensel}, we need to understand which primes ramify in $K_\ell$. The following result can be derived from a criterion of N\'eron--Ogg--Shafarevich \cite[Theorem 7.1, p. 201]{silverman09}.
\begin{lem} Let $E/\Q$ be an elliptic curve. Every prime that ramifies in $K_{\ell}$ divides $\ell \cdot \Delta_{E}$.
\end{lem}
We will find bounded gaps among primes $p$ produced by certain linear functions, with coefficients chosen to give $p$ a ``leg up'' in terms of $E(\F_p)$ being cyclic. To build these functions, we need the following analogue of Lemma \ref{lem:HB}.
\begin{lem}\label{lem:HB2} Let $M$ be either a quadratic or abelian cubic extension of $\Q$. Let $f = \mathfrak{f}(M/\Q)$, and let $v= 2^4 3^3 \prod_{\ell \mid f,~\ell > 3} \ell$. One can select an integer $u$ coprime to $v$ so that both of the following hold:
\begin{enumerate}
\item[\normalfont\rmfamily{(1)}] For every prime $p\equiv u\pmod{v}$, $p$ is inert in $M$.
\item[\normalfont\rmfamily{(2)}] If $T$ is the largest power of $2$ dividing $u-1$, then $T \in \{2,4,8\}$, and $\gcd(\frac{u-1}{T},v)=1$.
\end{enumerate}
\end{lem}
\begin{proof} We make free use of the correspondence between abelian extensions of $\Q$ and groups of primitive Dirichlet characters, as reviewed in \cite[Chapter 3]{washington97}.
If $M/\Q$ is quadratic, then $f$ is the absolute value of a fundamental discriminant, whereas if $M/\Q$ is abelian cubic, then
\[ f = 9 q_1 \cdots q_k \quad\text{or}\quad f = q_1\cdots q_k,\quad\text{for distinct primes}\quad q_i \equiv 1\pmod{6}. \]
Thus, $2^4 \nmid f$, $3^3\nmid f$, and every prime $\ell > 3$ that divides $f$ appears to the first power only. Let $H$ be the subgroup of $\Gal(\Q(\zeta_f)/\Q)$ that fixes $M$. We identify $\Gal(\Q(\zeta_f)/\Q)$ with $(\Z/f\Z)^{\times}$. Note that $H$ has index $[M:\Q] > 1$. Since $M$ is cyclic of prime degree, an unramified prime $p$ either remains inert or splits completely, the latter holding exactly when $p\bmod{f} \in H$.
Choose an integer $u_0$ with
\begin{equation}\label{eq:u00} \gcd(u_0,f)=1, \quad u_0\bmod{f} \not \in H, \quad\text{and}\quad u_0 \equiv 2\pmod{3}.\end{equation} This is clearly possible if $3\nmid f$. If $3\mid f$, we argue by contradiction: If there is no such $u_0$, then
$\#H > \#\{1 \leq h \leq f: \gcd(h,f)=1, h\equiv 2\pmod{3}\} = \frac{1}{2}\phi(f)$, where the inequality is strict since $1\bmod{f} \in H$. But then $H = \Gal(\Q(\zeta_f)/\Q)$, a contradiction.
We can also assume that
\begin{equation}\label{eq:u01} u_0 \equiv 1\pmod{2}. \end{equation}
Indeed, if $f$ is even, this condition is automatic, whereas if $f$ is odd but $u_0$ is even, we can replace $u_0$ by $u_0 + 3f$. Finally, we can assume that
\begin{equation}\label{eq:u02} 16 \nmid u_0-1, \end{equation}
by replacing $u_0$ with $u_0 + \lcm[24,f]$ if necessary.
If $M/\Q$ is quadratic, then for each prime $\ell > 3$ dividing $f$, put
\[ u_{\ell} = \begin{cases} u_0 &\text{if $\ell \nmid u_0-1$}, \\
4u_0&\text{otherwise}.
\end{cases} \]
Then $\ell \nmid u_{\ell}-1$. If $M/\Q$ is abelian cubic, then for each prime $\ell > 3$ dividing $f$, put
\[ u_{\ell} = \begin{cases} u_0&\text{if $\ell \nmid u_0-1$}, \\
-8u_0&\text{otherwise}.
\end{cases} \]
In this case, we again have that $\ell \nmid u_{\ell}-1$. Finally, select $u$ so that
\[ u \equiv u_0 \pmod{2^4\cdot 3^3}\quad\text{and}\quad u\equiv u_{\ell}\pmod{\ell}~\text{$\forall \ell \mid f$ with $\ell > 3$}. \]
This puts $u$ in a well-defined coprime residue class modulo $v$.
We now check (1) and (2). In the case when $M/\Q$ is quadratic, $u\equiv u_0 g^2\pmod{f}$ for some integer $g$. Since $H$ has index $2$, $g^2\bmod{f} \in H$. Since $u_0 \bmod{f}\not\in H$, we find that $u\bmod{f}\not\in H$. So if $p\equiv u\pmod{v}$, then $p\bmod{f} \not\in H$ (notice $f \mid v$) and so $p$ is inert in $M$. An analogous argument works when $M/\Q$ is abelian cubic; in that case, $H$ has index $3$ and $u\equiv u_0 g^3\pmod{f}$ for some $g$. This completes the verification of (1). Since $u\equiv u_0\pmod{16}$, \eqref{eq:u01} and \eqref{eq:u02} yield $T \in \{2,4,8\}$. Since $u\equiv u_0\pmod{3}$, \eqref{eq:u00} shows that $3\nmid u-1$. For each prime $\ell > 3$ dividing $v$, our choices of $u_{\ell}$ ensure that $\ell \nmid u-1$. Hence, $\gcd(\frac{u-1}{T},v)=1$, which completes the proof of (2).
\end{proof}
By imitating the deduction of Lemma \ref{lem:admissible} from Lemma \ref{lem:HB}, we obtain the following consequence of Lemma \ref{lem:HB2}.
\begin{lem}\label{lem:admissible2} Let $M$ be either a quadratic or abelian cubic extension of $\Q$. Let $u$ and $v$ be chosen as in Lemma \ref{lem:HB2}. Let $\kappa$ be a natural number. There are integers $a_1 < \dots < a_{\kappa}$, each congruent to $u\bmod{v}$, for which the $2\kappa$ linear functions
\begin{alignat*}{3} L_1(n) &= vn + a_1, &\quad \dots, \quad &L_{\kappa}(n) = vn + a_{\kappa}, \\
\tilde{L}_1(n) &= \frac{v}{T}n + \frac{a_1-1}{T}, &\quad \dots, \quad &\tilde{L}_{\kappa}(n) = \frac{v}{T} n + \frac{a_{\kappa}-1}{T} \end{alignat*}
make up an admissible family. Moreover, we can select the $a_{i}$ in such a way that
\[ a_{\kappa} - a_1 \leq v\cdot (2\kappa)^{C_5}. \]
\end{lem}
\section{Proof of Theorem \ref{thm:EC}}
\subsection{The GRH case}
By assumption, $K_2 \neq \Q$. Since $K_2$ is the splitting field of a cubic polynomial, it has a subfield $M$ that is either quadratic or abelian cubic over $\Q$. Let $\kappa = \lceil\exp(C_6 m) \rceil$, where $C_6$ is a large absolute constant. Let $k=2\kappa$, and let $\Ll$ consist of the linear functions $L_1, \dots, L_{\kappa}$, $\tilde{L}_1, \dots, \tilde{L}_{\kappa}$ constructed in Lemma \ref{lem:admissible2}. Recall that each $L_i$ has leading coefficient $v=2^4 3^3 \prod_{\ell \mid f,~\ell > 3}\ell$, where $f$ is the conductor of $M$. If $\ell \mid f$, then $\ell \mid \Delta_{K_2}$, and so $\ell =2$ or $\ell$ is a prime of bad reduction. Consequently,
\[ v \mid 2^4 3^3 \cdot \mathrm{rad}(\Delta_{E}). \]
We warn the reader of the following innocuous abuse of notation: If $L=L_i$, we will write $\tilde{L}$ for $\tilde{L}_i$.
Assume $x$ is large. We will show that if $c>0$ is a sufficiently small absolute constant, then with positive probability, an $n \in \A(x)$ satisfies all of
\begin{enumerate}
\item[(i)] at least $m$ of $L_1(n)$, \dots, $L_\kappa(n)$ are prime,
\item[(ii)] $p^{-}(L_i(n)) \ge x^{\frac{c}{k^3\log{k}}}$ and $p^{-}(\tilde{L}_i(n)) \ge x^{\frac{c}{k^3\log{k}}}$ for all $i=1,\dots, \kappa$,
\item[(iii)] all integers in the interval $[L_1(n),L_\kappa(n)]$ that are not one of the $L_i(n)$ are composite,
\item[(iv)] if $p=L(n)$ is prime with $L \in \{L_1, \dots, L_{\kappa}\}$, then $p$ is inert in every $K_\ell$ with $\ell> x^{1/3}$ and $\ell \neq p$,
\item[(v)] if $p=L(n)$ is prime with $L \in \{L_1, \dots, L_{\kappa}\}$, then $E(\F_p)$ is cyclic.
\end{enumerate}
If all of (i)--(v) hold for $n$, then the set of primes $p \in [L_1(n), L_\kappa(n)]$ has at least $m$ elements, all of these have $E(\F_p)$ cyclic, and the gap between the largest and smallest is at most
\[ L_{\kappa}(n)-L_1(n) \leq v \cdot (2\kappa)^{C_5} \leq \mathrm{rad}(\Delta_E) \cdot \exp(C_7 m); \]
the GRH half of Theorem \ref{thm:EC} follows.
For the sake of readability, in the remainder of the proof we suppress the dependence of implied constants on $E$.
To handle (i)--(iii), we proceed as in the proof of Theorem \ref{thm:main}. Arguments given there show that if we fix $C_6$ sufficiently large and $c$ sufficiently small, then (i) and (ii) hold simultaneously with probability at least $\frac{1}{2}\kappa^{-1}$, while (iii) fails with probability $o(1)$, as $x\to\infty$.
Now suppose that (i)--(iii) hold for $n$ but that (iv) fails. We will show that this occurs with probability $o(1)$. Observe that for each $n \in [x,2x)$ and each $L \in \{L_1, \dots, L_\kappa\}$, the integer $L(n)$ is smaller than $3vx$.
We start by bounding the number of $p \leq 3vx$ which split completely in $K_\ell$ for some $\ell > x^{1/3}$ with $\ell \neq p$. In that case, $(\Z/\ell\Z)^2$ sits inside $E(\F_p)$, and so
\begin{equation}\label{eq:firstdiv} \ell^2 \mid p+1-a_p. \end{equation}
Since $\Q(\zeta_l)\subset K_\ell$ (by properties of the Weil pairing \cite[Corollary 8.1.1]{silverman09}),
\begin{equation}\label{eq:2nddiv} \ell \mid p-1. \end{equation}
Comparing \eqref{eq:firstdiv} and \eqref{eq:2nddiv} shows that $\ell \mid 2-a_p$. If $a_p \neq 2$, then
\[ 0 < |2-a_p| < 2+2\sqrt{3vx} < x^{2/3} < \ell^2; \]
hence $\ell$ is uniquely determined by $a=a_p$, as the largest prime dividing $|2-a|$. Fixing $a\ne 2$, \eqref{eq:firstdiv} shows that the number of corresponding $p\le 3vx$ is $\ll \frac{x}{\ell^2} + 1 \ll x^{1/3}$. By the Hasse bound, $|a| \ll \sqrt{x}$, and so summing on the possible values of $a$ shows that $O(x^{5/6})$ values of $p$ arise in this way. On the other hand, when $a=2$, \eqref{eq:sufficetotest} and \eqref{eq:firstdiv} imply that the number of corresponding $p$ is
\[ \ll \sum_{\ell \in (x^{1/3}, \sqrt{3vx}+1]} \left(\frac{x}{\ell^2}+1\right) \ll x^{2/3}. \]
So there are a total of $O(x^{5/6})$ of these primes $p$.
Since (i)--(iii) hold while (iv) fails, there is an $L \in \{L_1, \dots, L_\kappa\}$ such that $p=L(n)$ is among the primes counted in the previous paragraph. There are $O_m(1)$ possibilities for $L$, and so $O_m(x^{5/6})$ possibilities for $n\in\A(x)$. From (ii) and Proposition \ref{prop:maynard}(1), the probability mass at each such $n$ is $O_m(x^{-1} (\log{x})^k)$. So the probability that (i)--(iii) hold but (iv) fails is $O_m(x^{-1/6} (\log{x})^k)$, which is $o(1)$ as $x\to\infty$.
To complete the proof, we show the probability (i)--(iv) hold but (v) fails is also $o(1)$.
Suppose $p=L(n)$ is prime, with $L \in \{L_1, \dots, L_\kappa\}$, but that $E(\F_p)$ is not cyclic. From Lemma \ref{lem:HB2}(1) and our choice of $\Ll$, $p$ is inert in $M$, and a fortiori does not split completely in $K_2$. So $\E(\F_p)$ must split completely in $K_\ell$ for some $\ell > 2$. Since $\ell \mid p-1 = T \cdot \tilde{L}(n)$, (ii) and (iv) imply that
\begin{equation}\label{eq:lrange} x^{c/(k^3\log{k})} \le \ell \leq x^{1/3}. \end{equation}
We now count how many $p \leq 3vx$ split completely in $K_\ell$ for some $\ell$ in the range \eqref{eq:lrange} with $\ell\neq p$. Making the same appeal to Proposition \ref{prop:maynard}(1) we saw earlier in the proof, it is enough to prove that the number of these $p$ is
\begin{equation}\label{eq:suffices} \ll x^{1-\frac{c}{k^3\log{k}}} + x^{5/6} \log{x}. \end{equation}
We invoke GRH. By effective Chebotarev, the number of $p\leq 3vx$ splitting completely in $K_\ell$ is
\[ \ll \frac{x}{[K_\ell:\Q] \log{x}} + x^{1/2}\left(\log{x}+ \frac{1}{[K_\ell:\Q]} \log|\Delta_{K_\ell}|\right). \]
Since every prime dividing $\Delta_{K_\ell}$ divides $\ell \cdot \Delta_E$, \eqref{eq:hensel} shows that this upper bound is
\begin{equation}\label{eq:upper2} \ll \frac{x}{[K_\ell:\Q] \log{x}} + x^{1/2} \log([K_\ell:\Q]\cdot \ell x ). \end{equation}
If $E$ has CM, then for all large primes $\ell$, the degree of $K_\ell/\Q$ is either $2(\ell-1)^2$ or $2(\ell^2-1)$, according to whether or not $\ell$ splits in the CM field. In particular, $[K_\ell:\Q] \asymp \ell^2$. If the non-CM case, we have $[K_\ell:\Q]= \#\mathrm{GL}_2(\Z/\ell\Z) \asymp \ell^4$ for all large primes $\ell$. (These results are due to Serre \cite{serre72}; see \cite[Theorem 18]{CCS13} for a detailed discussion of the CM case.) Thus, the sum of \eqref{eq:upper2} over the range \eqref{eq:lrange} is
\[ \ll \frac{x}{\log{x}} \sum\frac{1}{\ell^2} + x^{1/2}\log{x}\sum_{\ell} 1 \ll x^{1-\frac{c}{k^3\log{k}}} + x^{5/6}\log{x}, \]
which agrees with \eqref{eq:suffices}. This completes the proof in the GRH case.
\subsection{Unconditional proof in the CM case} As already mentioned in the introduction, we will deal entirely with supersingular primes in this part of the proof.
Suppose that $p\ge 5$ is supersingular but that $E(\F_p)$ is not cyclic. Choose an $\ell \neq p$ for which $p$ splits completely in $K_\ell$. Then $\ell^2 \mid \#E(\F_p) = p+1$ and $\ell \mid p-1$, forcing $\ell=2$. Consequently, $p$ splits in the quadratic or abelian cubic subfield $M$ of $K_2$.
Let $F$ be the CM field. We look for primes $p$ of good reduction that are inert in $F$ --- guaranteeing that $p$ is supersingular --- and inert in $M$. If $F=M$, it is clear that there are infinitely many such primes; otherwise, this follows from the linear disjointness of $F$ and $M$ over $\Q$. Since $F/\Q$ and $M/\Q$ are abelian, the set of such $p$ contains all primes in a certain arithmetic progression modulo $q:=f_1 f_2$, where $f_1=\mathfrak{f}(F/\Q)$ and $f_2=\mathfrak{f}(M/\Q)$. Since $E$ is defined over $\Q$, its CM field $F$ must be one of the nine imaginary quadratic fields of class number $1$ (see, e.g., Serre's chapter in \cite{CF86}), and so $f_1 \ll 1$. On the other hand, since every odd prime dividing $f_2$ divides $\Delta_E$, and since $f_2$ is squarefree apart from bounded powers of $2$ and $3$, the modulus $q = f_1 f_2 \ll f_2 \ll \mathrm{rad}(\Delta_E)$.
Corollary 3 of \cite{BFTB14} asserts that for any fixed coprime progression mod $q$, there are infinitely many tuples of $m$ consecutive primes $p_1 < p_2 < \dots < p_m$ with $p_m - p_1 \ll_{q,m} 1$. In fact, it is straightforward to modify their argument to get an upper bound of $q \exp(O(m))$ (cf. \cite[Theorem 2(2)]{thorner14} when $m=2$). The theorem follows.
\begin{remark} In the non-CM case, we do not have an unconditional bounded gaps result for primes $p$ with $E(\F_p)$ cyclic. But if `cyclic' is replaced by `has an element of order $> p^{3/4-\epsilon}$', then such a result follows quickly from work of Duke \cite{duke03}.
Let $E/\Q$ be any elliptic curve. (No assumption on the rational torsion is needed here.) For each prime $p$ of good reduction, write $E(\F_p) \cong \Z/d_p\Z \oplus \Z/e_p\Z$ for natural numbers $d_p$ and $e_p$ where $d_p \mid e_p$. Clearly, $d_p^2 \le \#E(\F_p) \leq (\sqrt{p}+1)^2$, so that $d_p \leq 2\sqrt{p}$.
Duke shows (see \cite[eq. (8)]{duke03}) that for each $n \le 2\sqrt{x}$, the number of $p \leq x$ for which $n \mid d_p$ is $O(x^{3/2} n^{-3})$. A fortiori, the same bound holds for how often $d_p=n$. Consequently, the number of $p\le x$ with $d_p > x^{1/4+\epsilon/2}$ is $O(x^{1-\epsilon})$. Whenever $d_p \leq x^{1/4+\epsilon/2}$, the group $E(\F_p)$ has an element of order
\[ e_p \ge \frac{\#E(\F_p)}{x^{1/4+\epsilon/2}} \gg p x^{-\frac14-\frac{\epsilon}{2}}. \]
Summing dyadically, we conclude that $E(\F_p)$ has an element of order $> p^{\frac{3}{4}-\epsilon}$ for all but $O_\epsilon(x^{1-\epsilon})$ primes $p \le x$. This exceptional set is so sparse that it follows immediately from Maynard's lower bound results (see \cite[Theorem 3.1]{maynard14d}) that the set of nonexceptional $p$ has bounded gaps. More precisely, this set contains arbitrarily long runs of primes contained in bounded length intervals.\end{remark}
\section{Proof of Theorem \ref{thm:CM}}
We begin by stating a variant of Proposition \ref{prop:maynard} for sets of primes described by Chebotarev conditions.
\begin{prop}\label{prop:maynard2} Let $K/\Q$ be a Galois extension, and let $\C$ be a fixed conjugacy class of $\Gal(K/\Q)$. Let \[ \Pp(\C)= \{p: p \nmid \Delta_K, \legq{K/\Q}{p} = \C\}. \] Suppose $a_1 < a_2 < \dots < a_\kappa$ are odd integers for which the $k=2\kappa$ linear functions
\begin{multline}\label{eq:simpleadmissible} L_1(n) = 2n+a_1, \quad L_2(n) = 2n+a_2, \quad\dots,\quad L_\kappa(n) = 2n+a_\kappa, \\
\tilde{L}_1(n) = n+\frac{a_1-1}{2}, \quad \tilde{L}_2(n) = n+\frac{a_2-1}{2}, \quad\dots,\quad \tilde{L}_\kappa(n) = n+\frac{a_\kappa-1}{2}
\end{multline}
form an admissible collection; call this collection $\Ll$.
Suppose that $x$ is sufficiently large, $x > x_0(K,\Ll)$. There is a probability measure on $\A(x)= \{n \in \Z: x \leq n < 2x\}$
with all of the following properties:
\begin{enumerate}
\item[\normalfont\rmfamily{(1)}] The probability mass at any single $n \in \A(x)$ is \[ \ll_{K} x^{-1} (\log{x})^k \left(\prod_{i=1}^{k} \prod_{\substack{p \mid L_i(n) \\ p \nmid 2 \Delta_K}} 4\right) \exp(O(k\log{k})). \]
\item[\normalfont\rmfamily{(2)}] For each $L \in \Ll$,
\[ \Prob(L(n)\text{ belongs to $\Pp(\C)$})\gg_K \frac{\log{k}}{k}. \]
\item[\normalfont\rmfamily{(3)}] Let $\rho \in [k\frac{(\log\log{x})^2}{\log{x}}, \frac{1}{30 [K:\Q]}]$. For each $L \in \Ll$,
\[ \E\bigg[\sum_{\substack{p \mid L(n) \\ p \leq x^{\rho}, ~p\nmid 2\Delta_K}} 1\bigg] \ll \rho^2 k^4 (\log{k})^2. \]
\end{enumerate}
The implied constant in (3) is absolute.
\end{prop}
\begin{proof}[Proof (sketch)] The main technical input is supplied by a variant of the Bombieri--Vinogradov theorem due to Murty and Murty \cite{MM87}, which asserts that $\Pp(\C)$ has level of distribution $\theta$ for any fixed \[ \theta < \min\{\frac{1}{2}, \frac{2}{[K:\Q]}\};\] here the moduli of the arithmetic progressions are assumed coprime to $\Delta_K$. We now argue as in the proof of Proposition \ref{prop:maynard}. Specifically, the Murty--Murty theorem allows us to apply \cite[Proposition 6.1]{maynard14d} with $\A=\N$, $\Ll$ as given, $\curly{P}= \Pp(\C,K)$, $B=2 \Delta_K$, $\theta = \min\{\frac{1}{3}, \frac{1}{[K:\Q]}\}$, and $\alpha=1$. Defining the probability mass at $n$ as $w(n)/\sum_{n \in \A(x)} w(n)$, the result follows. (For similar applications of the Murty--Murty theorem, see \cite{thorner14} and \cite[Theorem 3.5]{maynard14d}.)
\end{proof}
The proof of Theorem \ref{thm:CM} also uses the following criterion, which is contained in work of Cojocaru \cite[Lemmas 2.2 and 2.3]{cojocaru03}.
\begin{lem}\label{lem:cojocaru} Suppose that $E/\Q$ has CM by an order in the imaginary quadratic field $F$. Let $p$ be a prime of good ordinary reduction, and let $\ell$ be a prime with $\ell \neq p$. If $p$ splits completely in $K_\ell$, then there is a $\pi \in \Z_F$ with $\pi \equiv 1\pmod{\ell}$ and $N(\pi)=p$.\end{lem}
\begin{proof}[Proof of Theorem \ref{thm:CM}] We begin by specifying the parameters needed for our application of Proposition \ref{prop:maynard2}.
Let $F$ be the CM field of $E$. Let $\Qq$ be the set of primes dividing $2\Delta_F \Delta_E$, and let $K$ be the compositum of $F$ and all of the fields $K_\ell:=\Q(E[\ell])$ for $\ell \in \Qq$. Then $K/\Q$ is Galois and every prime dividing $\Delta_K$ belongs to $\Qq$. Choose a conjugacy class $\C$ of $\Gal(K/\Q)$ where every prime $p \in \Pp(\C)$ is such that
\begin{itemize}
\item $p$ splits in $F$,
\item $p$ does not split completely in any of the fields $K_\ell$ with $\ell \in \Qq$.
\end{itemize}
Any large prime $p$ of ordinary reduction for which $E(\F_p)$ is cyclic satisfies both of these conditions, and so such a $\C$ must exist.
Let $\kappa = \lceil\exp(C_K m)\rceil$, where $C_K$ is a sufficiently large constant depending on $K$. Mimicking the proof of Lemma \ref{lem:admissible}, we can choose odd integers $a_1 < \dots < a_{\kappa}$ for which \eqref{eq:simpleadmissible} is admissible, with $a_{\kappa} - a_1 \leq (2\kappa)^{C_8}$. Then \begin{equation}\label{eq:finalgapbound} a_{\kappa}-a_1 \ll \exp(O_E(m)).\end{equation}
We are now in a position to apply Proposition \ref{prop:maynard2}. If $C_K$ is sufficiently large and $c$ is sufficiently small (both allowed to depend on $K$), then an $n\in \A(x)$ satisfies both of the following conditions with probability $\gg_m 1$:
\begin{enumerate}
\item[(i)] at least $m$ of $L_1(n), \dots, L_\kappa(n)$ belong to $\Pp(\C)$,
\item[(ii)] whenever a prime $\ell \leq x^{c/(k^3\log{k})}$ divides $\prod_{i=1}^{\kappa}L_i(n) \tilde{L}_i(n)$, $\ell$ also divides $2\Delta_K$.
\end{enumerate}
Indeed, this follows from arguments seen already in the proofs of Theorems \ref{thm:main} and \ref{thm:EC}, the only difference being that we appeal to Proposition \ref{prop:maynard2} instead of Proposition \ref{prop:maynard}. We now introduce the statement
\begin{enumerate}
\item[(iii)] Whenever $p = L(n) \in \Pp(\C)$, with $L \in \{L_1, \dots, L_\kappa\}$, the group $E(\F_p)$ is cyclic.
\end{enumerate}
We will show that the probability (i) and (ii) hold but (iii) fails is $o(1)$, as $x\to\infty$, so that (i)--(iii) hold with positive probability for all large $x$. This will complete the proof; indeed, if $n\in \A(x)$ satisfies (i)--(iii), and $p_1 < p_2 < \dots < p_m$ are primes from $\Pp(\C)$ drawn from $\{L_1(n), \dots, L_\kappa(n)\}$, then the claimed bound on $p_m-p_1$ follows from \eqref{eq:finalgapbound}, while the fact that each of the primes is of good ordinary reduction follows from the choice of $\C$.
Suppose (i) and (ii) hold and that $p = L_i(n) \in \Pp(\C)$, where $i\in \{1,2,\dots, \kappa\}$. As we have just remarked, $p$ is a prime of good ordinary reduction. If $E(\F_p)$ is not cyclic, then $p$ spits completely in $K_\ell$ for some $\ell \neq p$. Then $\ell \mid p-1 = 2\tilde{L}_i(n)$, so that either $\ell \mid 2\Delta_K$ or $\ell \ge x^{c/(k^3\log{k})}$. But if $\ell \mid 2\Delta_K$, then $\ell \in \Qq$, and so the choice of $\C$ guarantees that $p$ does not split completely in $K_\ell$. So it must be that $\ell \ge x^{c/(k^3\log{k})}$. Since $p \leq 5x$ for large $x$, we also have that $\ell \leq \sqrt{5x}+1$, after recalling \eqref{eq:sufficetotest}.
Let us count primes $p\le 5x$ of good ordinary reduction that split completely in $K_\ell$ for some $\ell\ne p$ with
\begin{equation}\label{eq:lrange2}
x^{c/(k^3\log{k})}\le \ell \le \sqrt{5x}+1.
\end{equation}
From Lemma \ref{lem:cojocaru}, there is a $\pi_p \in \Z_{F}$ with $\pi_p\equiv 1\pmod{\ell}$ and $N(\pi_p)=p$. The number of $\pi \in \Z_F$ with $\pi \equiv 1\pmod{\ell}$ and $N(\pi) \le 5x$ is $O(\frac{x}{\ell^2} + 1)$, by an elementary lattice point counting argument (e.g., see \cite[Lemma 5]{murty83} or \cite[Lemma 2.6]{cojocaru03}). Summing on $\ell$ in the range \eqref{eq:lrange2} shows that the number of $p$ in question is \[ \ll x^{1-\frac{c}{k^3\log{k}}} + x^{1/2}.\]
If $n$ satisfies (i) and (ii), Proposition
\ref{prop:maynard2}(1) shows that the probability mass at $n$ is $O_{K,m}(x^{-1} (\log{x})^k)$. Consequently, the probability that $L(n)$ is one of the primes counted in the preceding paragraph is $o(1)$, as $x\to\infty$.
\end{proof}
\subsection*{Acknowledgments} We are indebted to Pete L. Clark for helpful conversations on the theory of elliptic curves. This research began while the second author enjoyed a very pleasant visit to Brigham Young University. He thanks the BYU mathematics department for their hospitality and the NSF for their support under award DMS-1402268.
\providecommand{\bysame}{\leavevmode\hbox to3em{\hrulefill}\thinspace}
\providecommand{\MR}{\relax\ifhmode\unskip\space\fi MR }
\providecommand{\MRhref}[2]{%
\href{http://www.ams.org/mathscinet-getitem?mr=#1}{#2}
}
\providecommand{\href}[2]{#2}
|
\section{Introduction}
\IEEEPARstart{T}{he} automated understanding of textual information in natural scene images is receiving increasing attention from computer vision researchers over the last decade. Text localization, extraction and recognition methods have evolved significantly and their accuracy has increased drastically in recent years~\cite{karatzas2013icdar}. However, the problem is far from being considered solved: note that the winner methods in the last ICDAR competition achieve only 66\% and 74\% recall in the tasks of text localization and text segmentation respectively. The main difficulties of the problem stem from the extremely high variability of scene text in terms of scale, rotation, location, physical appearance, and typeface design. Moreover, although standard benchmark datasets have traditionally focussed on horizontally-aligned English text, new datasets have recently appeared covering much more unconstrained scenarios including multi-script and arbitrary oriented text~\cite{yao2012detecting,kumar2013multi}.
Hierarchical organisation is an essential feature of text. Induced by typography and layout the hierarchical arrangement of text strokes leads to the structural formation of text component groupings at different levels (e.g. words, text lines, paragraphs, etc.), see Figure~\ref{fig:text_hierarchy}. This hierarchical property applies independently of the script, language, or style of the glyphs, thus it allows us to pose the problem of text detection in natural scenes in a holistic framework rather than as the classification of individual patches or regions as text or non-text. In fact, as Figure ~\ref{fig:atomic_vs_group} shows, when text-parts are viewed independently out of context they lose their distinguishable text traits, although they become structurally relevant and easily identifiable when observed as a group.
\begin{figure}[t]
\centering
\includegraphics[width=0.99\linewidth]{hierarchy_scene2.jpg}
\caption{A natural scene image and a hierarchical representation of its text. Atomic objects (characters) extracted in the bottom layer are agglomerated into text groupings at different levels of the hierarchy.}
\label{fig:text_hierarchy}
\end{figure}
\begin{figure}[b]
\centering
\includegraphics[width=0.99\linewidth]{atomic_vs_group02.jpg}
\caption{Individual text-parts are less distinguishable when viewed separately, but become structurally relevant and easily identifiable when perceived as a group.}
\label{fig:atomic_vs_group}
\end{figure}
Most existing scene text extraction methods include a grouping step, but more often than not this is done as a heuristic post-processing of regions (e.g. connected components or superpixels) previously classified as text in order to create word or text-line bounding boxes, and it is not part of the core text extraction process~\cite{Epshtein2010,Chen2011}.
Contrary to existing methods, this paper addresses the problem of text segmentation in natural scenes from a hierarchical perspective tackling directly the problem of the detection of groups of text regions, instead of individual regions. The method is driven by an agglomerative clustering process exploiting the strong similarities expected between text components in such groups:
irrespective on the script or language, text is formed by aligned and equally separated glyphs with noticeable contrast to their background, with constant stroke width (thickness), similar color and sizes.
The main contributions of this paper are the following.
First, we learn an optimal feature space that encodes the similarities between text components thus allowing the Single Linkage Clustering algorithm to generate text group hypotheses with high recall, independently of their orientations and scripts. Second, we couple the hierarchical clustering algorithm with novel discriminative and probabilistic stopping rules, that allow the efficient detection of text groups in a single grouping step. Third, we propose a new set of features for text group classification, that can be efficiently calculated in an incremental way, able to capture the inherent structural properties of text related to the arrangement of text parts and the intra-group similarity.
Note that the proposed method is distinctly different from other grouping based state of the art approaches~\cite{Yin2013} in that the text group classification is not an a-posteriori step, but an inherent part of the hierarchical clustering process.
Our findings are positioned in line with recent advances in object recognition~\cite{uijlings2013selective} where bottom-up grouping of an initial segmentation is used to generate object location hypotheses, producing a substantially reduced search space in comparison to the traditional slidding window approaches. We make use of incrementally computable group descriptors in order to make possible the direct evaluation of group hypotheses generated by the clustering algorithm without affecting the overall time complexity of the method. Experiments demonstrate that our algorithm outperforms the state of the art in the MSRRC, MSRA-TD500 and KAIST datasets of multi-script and arbitrary oriented text~\cite{kumar2013multi,yao2012detecting,Lee2010}.
It is important to notice that our method produces state of the art results in four different datasets with a single (mixed) training set, i.e. it can be seen as a general purpose robust method applicable in many different scenarios. This is afforded by the relatively high-level modelling of text as a group of individual elements, a model which is valid for practically every writing system.
\section{Related Work}
Scene text detection methods can be categorized into texture-based and region-based approaches. Texture-based methods usually work by performing a sliding window search over the image and extracting certain texture features in order to classify each possible patch as text or non-text. Coates \textit{et al.}~\cite{Coates2011}, and in a different flavour Wang \textit{et al.}~\cite{Wang2012} and Netzer \textit{et al.}~\cite{Netzer2011}, propose the use of unsupervised feature learning to generate the features for text versus non-text classification. Wang \textit{et al.}~\cite{Wang2011}, extending their previous work~\cite{Wang2010}, have built an end-to-end scene text recognition system based on a sliding window character classifier using Random Ferns, with features originating from a HOG descriptor. Mishra \textit{et al.}~\cite{Mishra2012} propose a closely related end-to-end method based on HOG features and a SVM classifier. Texture based methods yield good text localisation results, although they do not directly address the issue of text segmentation (separation of text from background). Their main drawback compared to region based methods is their lower time performance, as sliding window approaches are confronted with a huge search space in such an unconstrained (i.e. variable scale, rotation, aspect-ratio) task. Moreover, these methods are usually limited to the detection of a single language and orientation for which they have been trained on, therefore they are not directly applicable to the multi-script and arbitrary oriented text scenario.
Region-based methods, on the other hand, are based on a typical bottom-up pipeline: first performing an image segmentation and subsequently classifying the resulting regions into text or non-text ones. Yao \textit{et al.}~\cite{yao2012detecting} extract regions in the Stroke Width Transform (SWT) domain, proposed earlier for text detection by Epshtein \textit{et al.}~\cite{Epshtein2010}. Yin \textit{et al.}~\cite{Yin2013} obtain state-of-the-art performance with a method that prunes the tree of Maximally Stable Extremal Regions (MSER) using the strategy of minimizing regularized variations. The effectiveness of MSER for character candidates detection is also exploited by Chen \textit{et al.}~\cite{Chen2011} and Novikova \textit{et al.}~\cite{Novikova2012}, while Neumann \textit{et al.}~\cite{Neumann2012} propose a region representation derived from MSER where character/non-character classification is done for each possible Extremal Region (ER).
Most of the region-based methods are complemented with a post-processing step where regions assessed to be characters are grouped together into words or text lines. The hierarchical structure of text has been traditionally exploited in a post-processing stage with heuristic rules~\cite{Epshtein2010,Chen2011} usually constrained to search for horizontally aligned text in order to avoid a combinatorial explosion of enumerating all possible text lines. Neumann and Matas~\cite{Neumann2012} introduce an efficient exhaustive search algorithm using heuristic verification functions at different grouping levels (i.e. region pairs, triplets, etc.), but still constrained to horizontal text. Yao \textit{et al.}~\cite{yao2012detecting} make use of a greedy agglomerative clustering where regions are grouped if their average alignment is under a certain threshold. Yin \textit{et al.}~\cite{Yin2013} use a self-training distance metric learning algorithm that can learn distance weights and clustering thresholds simultaneously and automatically for text groups detection in a similarity feature space.
\begin{figure*}[t]
\centering
\includegraphics[width=\linewidth]{hierarchy_dendrogram01_new.jpg}
\caption{A bottom-up agglomerative clustering of individual regions produces a dendrogram in which each node represents a text group hypothesis. Our work focuses on learning the optimal features allowing the generation of pure text groups (comprising only text regions) with high recall, and designing a stopping rule that allows the efficient detection of those groups in a single grouping step.}
\label{fig:hierarchy_dendrogram00}
\end{figure*}
In this paper we present a novel hierarchical approach in which region hierarchies are built efficiently using Single Linkage Clustering in a weighted similarity feature space. The hierarchies are built in different color channels in order to diversify the number of hypotheses and thus increase the maximum theoretical recall. Our method is less heuristic in nature and faster than the greedy algorithm of Yao \textit{et al.}~\cite{yao2012detecting}, because the number of atomic objects in our clustering analysis is not increased by taking into account all possible region pairs; besides our method uses similarity and not collinearity for grouping. Yin \textit{et al.}~\cite{Yin2013}, and also our previous work~\cite{Gomez2013}, make use of a two step architecture first doing an automatic clustering analysis in a similarity feature space and then classifying the groups obtained in the first step. The method presented here differs from such approaches in that our agglomerative clustering algorithm integrates a group classifier, acting as a stopping rule, that evaluates the conditional probability for each group in the hierarchy to correspond to a text group in an efficient manner through the use of incrementally computable descriptors. In this sense our work is related with Matas and Zimmerman~\cite{matas2005} region detection algorithm, while the incremental descriptors proposed here are designed to find relevant groups of regions in a similarity dendrogram instead of the detection of individual regions in the component tree of the image. There is also a relationship between our method with the work of Van de Sande \textit{et al.}~\cite{van2011} and Uijlings \textit{et al.}~\cite{uijlings2013selective} on using segmentation and grouping as selective search for object recognition. However, our approach is distinct in that their region grouping algorithm agglomerates regions in a class-independent way while our hierarchical clusterings are designed in order to maximize the chances of finding specifically text groups. Thus, our algorithm can be seen as a task-specific selective search.
\section{Hierarchy guided text extraction}
Our hierarchical approach to text extraction involves an initial region decomposition step where non-overlapping atomic text parts are identified. Extracted regions are then grouped together in a bottom-up manner with an agglomerative process guided by their similarity. The agglomerative clustering process produces a dendrogram where each node represents a text group hypothesis. We can then find the branches corresponding to text groups by simply traversing the dendrogram with an efficient stopping rule. Figure~\ref{fig:hierarchy_dendrogram00} shows an example of the main steps of the pipeline.
We make use of the Maximally Stable Extremal Regions (MSER)~\cite{Matas2004} algorithm to get the initial set of low-level regions. MSERs have been extensively used in recent state of the art methods for detecting text character candidates~\cite{Neumann2011,Chen2011,Novikova2012,Gomez2013,Yin2013}.
Recall in character detection is increased by extracting regions from different single channel projections of the image (i.e. gray, red, green and blue channel). This technique, denoted MSER++~\cite{Neumann2011}, is a way of diversifying the segmentation step in order to maximize the chances of detecting all text regions.
In the following we address the problem of designing a grouping algorithm exploiting the hierarchical structure of text, in order to detect text regions in a holistic manner. Our solution involves the learning of the optimal clustering feature space for text regions grouping and the design of novel discriminative and probabilistic stopping rules, that allows the efficient detection of text groups in a single clustering step.
\subsection{Optimal clustering feature space}
\label{sec:weighted}
It is usually expected that text parts belonging to the same word or text line share similar colors, stroke widths, and sizes.
Although the previous assumption does not always hold (see Figure~\ref{fig:dissimilarities}), in this work we consider that it is possible to weight those simple similarity features obtaining an optimal feature space projection that maximizes the probabilities of finding pure text groups (groups comprising only regions that correspond to text parts) in a Single Linkage Clustering (SLC) dendrogram.
\begin{figure}
\centering
\subfloat[]{\includegraphics[width=0.3\linewidth]{dissimilarity-color}\label{fig:dissimilarity-color}}
\hspace{0.1cm}
\subfloat[]{\includegraphics[width=0.3\linewidth]{dissimilarity-stroke}\label{fig:dissimilarity-stroke}}
\hspace{0.1cm}
\subfloat[]{\includegraphics[width=0.3\linewidth]{dissimilarity-size}\label{fig:dissimilarity-size}}
\caption{There is no single best feature for character clustering: Characters in the same word may appear with different color (a), stroke width (b) or sizes (c).}
\label{fig:dissimilarities}
\end{figure}
\begin{figure*}[t]
\centering
\subfloat[]{\includegraphics[width=0.23\linewidth]{max_recall_dendrograms_a.jpg}\label{fig:max_recall_a}}
\hspace{0.1cm}
\subfloat[]{\includegraphics[width=0.23\linewidth]{max_recall_dendrograms_b.jpg}\label{fig:max_recall_b}}
\hspace{0.1cm}
\subfloat[]{\includegraphics[width=0.23\linewidth]{max_recall_dendrograms_c.jpg}\label{fig:max_recall_c}}
\hspace{0.1cm}
\subfloat[]{\includegraphics[width=0.23\linewidth]{max_recall_dendrograms_d.jpg}\label{fig:max_recall_d}}
\caption{(a) Scene image, (b) its MSER decomposition, and (c,d) two possible hierarchies built from two different weight configurations, red nodes indicate pure text groupings. The first configuration (c) yields a 28\% text group recall ($\mathcal{TGR}$) while the second (d) achieves 100\% for this particular image. }
\label{fig:max_recall_dendrograms}
\end{figure*}
Let $\mathcal{R}_c$ be the initial set of individual regions extracted with the MSER algorithm from channel $c$.
We start an agglomerative clustering process, where initially each region $r\in\mathcal{R}_c$ starts in its own cluster and then the closest pair of clusters ($A,B$) is merged iteratively, using the single linkage criterion ($ \min \, \{\, \mathrm{d}(r_a,r_b) : r_a \in A,\, r_b \in B \,\} $), until all regions are clustered together ($C \equiv \mathcal{R}_c$). The distance between two regions $\mathrm{d}(r_a,r_b)$ is calculated as the squared Euclidean distance between their weighted feature vectors, adding a spatial constraint term (the squared Euclidean distance between their centers' coordinates) in order to induce neighbouring regions to merge first:
\begin{equation}
\mathrm{d}(\mathbf{a},\mathbf{b}) = {\sum\limits^D_{i=1}}{(w_i\cdot (a_i - b_i))^2} + \{(x_a-x_b)^2+(y_a-y_b)^2\}
\label{eq:distance}
\end{equation}
\noindent
where we consider the 5-dimensional feature space ($D=5$) comprising the following features: mean gray value of the region, mean gray value in the immediate outer boundary of the region, region's major axis, mean stroke width, and mean of the gradient magnitude at the region's border.
It is worth noting that using the squared Euclidean distance for the spatial term in equation~\ref{eq:distance}, our clustering analysis remains rotation invariant, thus the obtained hierarchy generates the same text group hypotheses independently of the image orientation. For example, rotating the image in Figure~\ref{fig:max_recall_a} by any degree would produce exactly the same dendrograms shown in Figures~\ref{fig:max_recall_c} and \ref{fig:max_recall_d}.This is intentional as we want our method to be capable of detecting text in arbitrary orientations. In this way, our algorithm deals naturally with arbitrary oriented text without using any heuristic assumption or threshold.
Given a possible set of weights $\mathbf{w}$, SLC produces a dendrogram $D_w$ where each node $H \in {D_w}$ is a subset of $\mathcal{R}_c$ and represents a text group hypothesis. The text group recall ($\mathcal{TGR}$) represents the ability of a particular weighting configuration to produce pure text groupings (comprising only text regions) corresponding to words or text lines in the ground-truth. Figure~\ref{fig:max_recall_dendrograms} shows an example of how different weight configurations lead to different text group recall.
Given a set of hypotheses $H \in D_w$, and a set of ground-truth text-group objects (i.e. words and text-lines) $G \in GT$, $\mathcal{TGR}$ is defined as:
\begin{equation}
\begin{split}
\mathcal{TGR}(D_w,GT) &= {1\over{|GT|}} \sum_{G \in GT} (\max_{H \in D_w}{{|H|}\over{|G|}} \\
& \mid \forall \, {r_h}\in{H} \, \exists \, {r_g}\in{G} \mid m_r(r_h,r_g) > 0.9 )
\end{split}
\end{equation}
\noindent
where $|\cdot|$ indicates cardinality of a set, and $m_r(\cdot,\cdot)$ is the overlap ratio between two regions. Thus, for each text-group in the ground-truth ($G \in GT$) we look for the largest group hypothesis in the dedrogram ($H \in D_w$) such that all regions in $H$ ($r_h \in H$) match with regions in the ground-truth group (${r_g}\in{G}$).
Our optimization problem for learning the optimal clustering feature space is defined as finding the set of weights $w_{opt}$ that maximise $\mathcal{TGR}$:
\begin{equation}
w_{opt} = \argmax_w \{ \mathcal{TGR}(D_w,GT) \}
\label{eq:argmax}
\end{equation}
As discussed before, there is no single best way to define similarity between text parts, hence there is no single best set of weights for our strategy, instead missing groups under a particular configuration may be potentially found under another.
An alternative to using a single feature space would be to diversify our clustering strategy, adding more hypotheses to the system by building different hierarchies obtained from different weight configurations (similarly to what we do with different color channels).
At training time, we use grid search strategy over the weights parameter space in order to solve equation~\ref{eq:argmax} for our training dataset. We assembled a mixed set of training examples using the MSRRC and ICDAR training sets. The MSRRC training set contains 167 images and the ICDAR training set 229. We have manually separated all text-lines and words in the ground truth data of these images, giving rise to 1611 examples of text groups. Figure~\ref{fig:train_sample} shows the group examples extracted from one of the training set images.
\begin{figure}
\centering
\includegraphics[width=\linewidth]{train_sample.png}
\caption{Our training set is assembled by manually separating the pixel level ground-truth of train images into all possible text groups (lines and words).}
\label{fig:train_sample}
\end{figure}
The optimized weights $w_{opt}$, obtained with grid search by maximizing the text group recall using equation~\ref{eq:argmax}, yield a cross-validated $\mathcal{TGR}$ of 0.87 in the training set, and are:
{$w_1 = 0.65$ (for intensity mean)}
, {$w_2 = 0.65$ (for outer boundary intensity mean)}
, {$w_3 = 0.49$ (for border gradient mean)}
, {$w_4 = 0.67$ (for diameter)}
, {$w_5 = 0.91$ (for stroke width mean)}.
As diversification strategy, after an optimized set of weights is obtained we subsequently remove from the training set the groups that have been detected, and then search again for new optimal weights $(w_{opt_{2}}, \dots, w_{opt_{n}})$ with the remaining groups. We evaluate this diversification strategy in section~\ref{sec:experiments}.
At test time, each of the optimal weight configurations is used to generate a dendrogram where each node represents a text group hypothesis. Selecting the branches corresponding to text groups is done by traversing the dendrogram with an efficient stopping rule.
\subsection{Discriminative and Probabilistic Stopping Rules}
Given a dendrogram representing a set of text groups hypotheses from the SLC algorithm, we need a strategy to determine the partition of the data that best fits our objective of finding pure text groups. A rule to decide the best partition in a Hierarchical Clustering is known as a \emph{stopping rule} because it can be seen as stopping the agglomerative process.
Differently from standard clustering stopping rules here we do not expect to obtain a full partition of regions in $\mathcal{R}_c$. In fact we do not even know if there are any text clusters at all in $\mathcal{R}_c$. Moreover, in our case we have a quite clear model for the kind of groups sought, corresponding to text. These particularities motivate the next contribution of this paper. We propose a stopping rule, to select a subset of meaningful clusters in a given dendrogram $D_w$, comprising the combination of the following two elements:
\begin{itemize}
\item{A text group discriminative classifier.}
\item{A probabilistic measure for hierarchical clustering validity assessment~\cite{Cao2004}.}
\end{itemize}
\subsubsection{Discriminative Text Group Classifier}
\label{sec:discriminative}
The first part of our stopping rule takes advantage of supervised learning, building a discriminative classification model $\mathcal{F}$ in a group-level feature space. Thus, given a group hypothesis $H$ and its feature vector $\mathbf{h}$, our stopping rule will accept $H$ only if $\mathcal{F}(\mathbf{h}) = 1$. We use a Real AdaBoost classifier~\cite{schapire1999improved} with decision stumps.
Our group-level features originate from three different types: 1) Group intra-similarity statistics, since we expect to see regions in the same word having low variation in color, stroke width, and size; 2) Shape similarity of participating regions, in order to discriminate repetitive patterns, such as bricks or windows in a building, which tend to be confused with text; 3) Collinearity and equidistance features, measure the text-like structure of text groups by using statistics of the 2-D Minimum Spanning Tree (MST) built with their regions centers. The list of used features is as follows:
\begin{itemize}
\item{FG intensities standard deviation.}
\item{BG intensities standard deviation.}
\item{Major axis coefficient of variation.}
\item{Stroke widths~\cite{Chen2011} coefficient of variation.}
\item{Mean gradient standard deviation.}
\item{Aspect ratios coefficient of variation.}
\item{Hu's invariant moments~\cite{hu1962} average Euclidean distance.}
\item{Convex hull compactness~\cite{Neumann2012} mean and standard deviation.}
\item{Convexity defects coefficient of variation.}
\item{MST angles mean and standard deviation.}
\item{MST edge widths coefficient of variation.}
\item{MST edge distances mean vs. diameters mean ratio.}
\end{itemize}
The AdaBoost classifier is trained using the same training set described in section~\ref{sec:weighted}. We have two sources of positive samples: 1) Using each GT group as if it were the output of the region decomposition step; 2) we run MSER and SLC ($w_{opt}$) against a train image and use as positive samples those pure-text groups in the SLC tree with more than 80\% match with a GT group. From the same tree we extract negative examples as nodes with 0 matchings. This gives us around 3k positive and 15k negative samples. We balance the positive and negative data and train a first classifier that is used to select 100 hard negatives that are used to re-train and improve accuracy.
\subsubsection{Incrementally computable descriptors}
Since at test time we have to calculate the group level features at each node of the similarity hierarchy, it is important that they are fast to compute. We take advantage of the inclusion relation of the dendrogram's nodes in order to make such features incrementally computable when possible. This allows us to compute the probability of each possible group of regions to be a text group without affecting the overall time complexity of our algorithm.
Group level features consisting of simple statistics over individual region features (e.g. diameters, strokes, intensity, etc.) can be incrementally computed straightforwardly with a few arithmetic operations and so have a constant complexity $\mathcal{O}(1)$.
Regarding the MST based features, an incremental algorithm (i.e. propagating the MST of children nodes to their parent) computing the MST on each node of the dendrogram takes $\mathcal{O}(n \times {\log}^2{n})$ in the worst case. Although this complexity is much lower than the $\mathcal{O}(n^2)$ complexity of the SLC step and thus does not affect the overall complexity of the algorithm, this has noticeable impact in run time. For this reason we add an heuristic rule on the maximum size of valid clusters: clusters with more than a certain number of regions are immediately discarded and there is no need to compute their features. By taking the length of the largest text line in the MSRRC training set (50) as the maximum cluster size, the run-time growth due to the features calculation in our algorithm is negligible and the obtained results are not affected at all.
\subsubsection{Probabilistic cluster meaningfulness estimation}
The way our classifier $\mathcal{F}$ is designed may eventually make the discriminative stopping rule to accept groups with outliers. For example, Figure~\ref{fig:nfa_rule} shows the situation where a node of the dendrogram consisting in a correctly detected word is merged with a (character like) region which is not part of the text group (outlier). In order to increase the discriminative power of our \emph{stopping rule} in such situations, we make use of a probabilistic measure of cluster meaningfulness~\cite{Desolneux2003,Cao2004}. This probabilistic measure, also used for text detection in our previous work~\cite{Gomez2013}, provides us with a way to compare clusters' qualities in order to decide if a given node in the dendrogram is a better text candidate than its children.
The Number of False Alarms ($\mathcal{NFA}$)~\cite{Desolneux2003,Cao2004}, based on the principle on non-accidentalness, measures the meaningfulness of a particular group of regions in $\mathcal{R}_c$ by quantifying
how the distribution of their features deviates from randomness.
Consider that there are $n$ regions in $\mathcal{R}_c$ and that a particular group hypothesis $H$ of $k$ of them have a feature in common. Assuming that the observed quality has been distributed randomly and uniformly across all regions in $\mathcal{R}_c$, the probability that the observed distribution for $H$ is a random realisation of this uniform process is given by the tail of the binomial distribution:
\begin{equation}
\mathcal{NFA}(G) = \mathcal{B}_G(k,n,p) = \sum_{i=k}^{n} \binom{n}{i} p^i (1-p)^{n-i}
\label{eq:nfa}
\end{equation}
\noindent
where $p$ is the probability of a single object having the aforementioned feature. The lower the $\mathcal{NFA}$ is, the more meaningful the group is.
We make use of this metric in each node of a dendrogram $D_w$ to assess the meaningfulness of all produced grouping hypotheses. We calculate (\ref{eq:nfa}) for each possible group hypothesis $H$ using as $p$ the ratio of the volume defined by the distribution of the feature vectors of the comprising regions ($h \in H$) with respect to the total volume of the $5-D$ feature space defined in Section~\ref{sec:weighted}.
\begin{figure}
\centering
\includegraphics[width=\linewidth]{nfa_rule.png}
\caption{A node in a similarity dendrogram consisting in a correctly detected word ($H_1$) is merged with a cluster consisting of a single region outlier ($H_2$). Our stopping rule will not consider valid the resulting cluster $H = \{H_1 \cup H_2\}$ although the classifier has labelled it as a text group ($\mathcal{F}(\mathbf{h}) = 1$) because $\mathcal{NFA}(H)$ is larger than $\mathcal{NFA}(H_1)$. The scatter plot simulates the arrangement of the feature vectors of the regions forming $H_1$, $H_2$, and $H$ in the similarity feature space.}
\label{fig:nfa_rule}
\end{figure}
Our stopping rule is defined recursively in order to accept a particular hypothesis $H$ as a valid group \emph{iif} its classifier predicted label is "text" ($\mathcal{F}(\mathbf{h}) = 1$) and its meaningfulness measure is higher than the respective meaningfulness measures of every successor $A$ and every ancestor $B$ labelled as text, i.e. the following inequalities hold:
\begin{equation}
\mathcal{NFA}(H) < \mathcal{NFA}(A), \forall A \in successors(H) \mid \mathcal{F}(A)=1
\end{equation}
\begin{equation}
\mathcal{NFA}(H) < \mathcal{NFA}(B), \forall B \in ancestors(H) \mid \mathcal{F}(B)=1
\end{equation}
Notice that by using this criteria no region is allowed to belong to more than one text group at the same time. The clustering analysis is done without specifying any parameter or cut-off value and without making any assumption on the number of meaningful clusters, but merely comparing the values of (\ref{eq:nfa}) at each node in the dendrogram for which the discriminative classifier label is "text" ($\mathcal{F}(H)=1$).
See Figure~\ref{fig:nfa_rule} for an example on how this stopping rule is able to detect outliers. As a side effect, the stopping rule is also able to correctly separate different words in a text line.
\setlength{\tabcolsep}{4pt}
\begin{table*}[t]
\centering
\caption{Segmentation results (a) and Precision-Recall curves (b) comparing different variants of our method.}
\subfloat[]{\baselinetab}
\hspace{0.6cm}
\subfloat[]{\includegraphics[width=0.34\linewidth]{pr_curve_tip2014.png}\label{fig:pr}}
\label{table:results_msrrc_diversification}
\end{table*}
\setlength{\tabcolsep}{1.4pt}
\begin{figure*}
\centering
\includegraphics[width=\linewidth]{msrrc_ok.jpg}
\caption{Qualitative segmentation results on the MSRRC 2013 dataset.}
\label{fig:msrrc_ok}
\end{figure*}
At this point, applying the method described so far our algorithm is able to produce results for the scene text segmentation task. The segmentation task is evaluated at pixel-level, this is the algorithm must provide a binary image where white pixels correspond to text and black pixels to background. All segmentation results given in section~\ref{sec:experiments} are obtained with this algorithm, trained with a single mixed dataset and without any further post-processing, by setting to white the pixels corresponding to the detected text groups. Figures~\ref{fig:msrrc_ok}, \ref{fig:kaist_ok}, and \ref{fig:icdar_ok} show segmentation results of our method in different datasets.
\subsection{From Pixel Level Segmentation to Bounding Box Localization}
In order to evaluate our method in the text localization task we extend our method with a simple post-processing operation in order to obtain word and text line bounding boxes depending on the semantic level ground truth information is defined (e.g. words in the case of ICDAR and MSRRC datasets, lines in the case of the MSRA-TD500 dataset). This is because the text groups detected by our stopping rule may correspond indistinctly to words, lines, or even paragraphs in some cases, depending on the particular typography and layout of the detected text.
First of all, region groups selected as text by our stopping rule in the different dendrograms are combined in a procedure that serves to de-duplicate repeated groups (e.g. the same group may potentially be found in several channels or weights configurations) and to merge collinear groups that may have been detected by chunks. Two given text groups are merged if they are collinear, and their relative distance and height ratio are under thresholds learned during training.
After that, if needed by the granularity of the ground-truth level, we split resulted text lines into words by considering as word boundaries all spaces between regions with a larger distance than a certain threshold, learned during training, proportional to the group's average inter-region distance.
\section{Experiments}
\label{sec:experiments}
The proposed method has been evaluated on three multi-script and arbitrary oriented text datasets and one English-only dataset for the tasks of text extraction and localization. All the segmentation evaluation is done at the pixel level, i.e. precision $p$ and recall $r$ are defined as $p = |E \cap T| / |E|$ and $r = |E \cap T| / |T|$, where $E$ is the set of pixels labelled as text and $T$ is the set of pixels corresponding to text in the ground truth. The localization results are evaluated with different frameworks depending on the dataset. The ICDAR~\cite{karatzas2013icdar,Lucas2003} and MSCCR~\cite{kumar2013multi} datasets have ground-truth defined at the word level and the proposed evaluation framework is the one of Wolf and Jolion~\cite{wolf2006}. The MSRA-TD500~\cite{yao2012detecting} has ground-truth defined at the line level and uses it's own evaluation framework. We use the standard evaluation frameworks for each dataset to be able to compare with the state of the art.
\subsection{Baseline analysis}
We have evaluated different variants of our method in order to assess the contribution of each of the proposed techniques. This baseline analysis is performed in the MSRRC test set. The baseline method is configured by setting all weights to $1$ ($w_I$) and accepting all group hypotheses which are labelled as text by the classifier ($\mathcal{F}(H)=1$). We compare this baseline with the variants making use of the learned optimal weights $w_{opt}$, and with including the meaningfulness criteria to our \emph{stopping rule}. Finally we have evaluated the impact of different diversification strategies to the initial segmentation, both in the number of image channels (MSER vs. MSER++), and in the number of weight configurations by adding a variable number of optimal weighted configurations $w_{opt_2}, \dots, w_{opt_n}$ into the system.
Table~\ref{tab:baselinetab} shows segmentation results of our method in the MSRRC 2013 test set comparing different variants of our method and different diversification strategies. The table includes also two variants of our previously published work~\cite{Gomez2013} for comparison. We chose to use the MSRRC dataset for this analysis as it is representative of the targeted scenario of multi-script and arbitrarily oriented text. Figure~\ref{fig:pr} plots the Precision-Recall curves, obtained by varying the acceptance threshold of the discriminative classifier in the stopping rule, for the five top scoring variations in Table~\ref{tab:baselinetab}.
From the obtained results we can see that the optimized weights $w_{opt}$ have a noticeable impact in the method recall, while the \emph{stopping rule} leads to a considerable increase in precision without any recall deterioration.
Regarding diversification, if one wants to maximize the harmonic mean between precision and recall, the use of MSER++ is well justified even though it produces a slight precision drop. However, examining the effect of further diversification using more optimal weighting configurations, we can see that the obtained gain in recall by adding more hypotheses does not help improving the f-score as it produces a significant precision deterioration. Such a diversification strategy should be considered only if one wants to maximize the system's recall.
\subsection{Multi-script and arbitrary oriented scene text extraction}
\label{sec:experimets_msrrc}
\begin{figure*}
\centering
\includegraphics[width=\linewidth]{kaist_ok.jpg}
\caption{Qualitative segmentation results on the KAIST dataset.}
\label{fig:kaist_ok}
\end{figure*}
\begin{figure*}
\centering
\includegraphics[width=\linewidth]{msra_localization_6.jpg}
\caption{Qualitative localization results on the MSRA-TD500 dataset.}
\label{fig:msra_localization}
\end{figure*}
We evaluate our method in three standard multi-script and arbitrary oriented text datasets. The MSRA-TD500 dataset~\cite{Yao2012} does not have pixel-level segmentation ground truth and thus we only evaluate on it for the text localization task, while in the MSRRC~\cite{kumar2013multi} and KAIST~\cite{Lee2010} datasets evaluation is done for both segmentation and localization tasks.
The MSRA-TD500 dataset~\cite{Yao2012} contains arbitrary oriented text in both English and Chinese languages. The dataset contains 500 images in total, with varying resolutions from $1296\times864$ to $1920\times1280$. The evaluation for text localization is done as proposed in~\cite{Yao2012} using minimum area rectangles. For an estimated minimum area rectangle $D$ to be considered a true positive, it is required to find a ground truth rectangle $G$ such that:
\begin{equation*}
{A(D' \cap G')}/{A(D' \cup G')} > 0.5 , abs(\alpha_D - \alpha_G) < {\pi}/{8}
\end{equation*}
where $D'$ and $G'$ are the axis oriented versions of $D$ and $G$, $A(D' \cap G')$ and $A(D' \cup G')$ are respectively the area of their intersection and union, and $\alpha_D$ and $\alpha_G$ their rotation angles. The definitions of precision $p$ and recall $r$ are: $p = |TP|/|E|$, $r = |TP|/|T |$ where $TP$ is the set of true positive detections while $E$ and $T$ are the sets of estimated rectangles and ground truth rectangles. Table~\ref{table:results_msra} compares our results with other state of the art methods on the MSRA-TD500 dataset and Figure~\ref{fig:msra_localization} show qualitative localization results.
\setlength{\tabcolsep}{4pt}
\begin{table}[h]
\begin{center}
\caption{Localization results in the MSRA-TD500 dataset.}
\label{table:results_msra}
\begin{tabular}{l c c c}
\hline\noalign{\smallskip}
Method & Precision & Recall & F-score\\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
\cellcolor{blue!25}\textbf{Ours} & \cellcolor{blue!25}\textbf{0.69} & \cellcolor{blue!25}0.54 & \cellcolor{blue!25}\textbf{0.61}\\
TD-Mixture~\cite{Yao2012} & 0.63 & \textbf{0.63} & 0.60\\
Gomez \& Karatzas \cite{Gomez2013}& 0.58 & 0.54 & 0.56\\
TD-ICDAR~\cite{Yao2012}& 0.53 & 0.52 & 0.54 \\
Epshtein \textit{et al.}~\cite{Epshtein2010}& 0.25 & 0.25 & 0.25\\
Chen \textit{et al.}~\cite{Chen2004}& 0.05 & 0.05 & 0.05\\
\hline
\end{tabular}
\end{center}
\end{table}
\setlength{\tabcolsep}{1.4pt}
The MSRRC dataset~\cite{kumar2013multi} comprises 334 camera-captured scene images, 167 in the training and 167 in the test set respectively, with sizes around $1.2MP$ for text localization and segmentation tasks. It covers Latin, Chinese, Kannada, and Devanagari scripts, and includes text with multiple orientations. Tables~\ref{table:results_msrrc} and ~\ref{table:results_msrrc2} compares our results with the participants in the segmentation and localization tasks of the 2013 Multi-script Robust Reading Competition, while Figure~\ref{fig:msrrc_ok} show examples of qualitative results. The average run-time of our algorithm in this dataset is 1.67 seconds per image on a standard PC.
\begin{figure}
\centering
\includegraphics[width=0.8\linewidth]{inverse_tip2014.png}
\caption{Inverse grade curves of different methods in the MSRRC dataset~\cite{kumar2013multi}.}
\label{fig:inverse}
\end{figure}
\setlength{\tabcolsep}{4pt}
\begin{table}[h]
\begin{center}
\caption{Segmentation results in the MSRRC 2013 dataset.}
\label{table:results_msrrc}
\begin{tabular}{l c c c}
\hline\noalign{\smallskip}
Method & Precision & Recall & F-score\\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
\cellcolor{blue!25}\textbf{Ours} & \cellcolor{blue!25}\textbf{0.75} & \cellcolor{blue!25}\textbf{0.71} & \cellcolor{blue!25}\textbf{0.73}\\
Yin \textit{et al.} \cite{Yin2013,kumar2013multi} & 0.71 & 0.67 & 0.69\\
Gomez \& Karatzas \cite{Gomez2013,kumar2013multi}& 0.64 & 0.58 & 0.61\\
Sethi \textit{et al.} \cite{kumar2013multi}& 0.33 & 0.72 & 0.45 \\
OTCYMIST \cite{kumar2012otcymist,kumar2013multi}& 0.50 & 0.29 & 0.37\\
\hline
\end{tabular}
\end{center}
\end{table}
\setlength{\tabcolsep}{1.4pt}
\setlength{\tabcolsep}{4pt}
\begin{table}[h]
\begin{center}
\caption{Localization results in the MSRRC 2013 dataset.}
\label{table:results_msrrc2}
\begin{tabular}{l c c c}
\hline\noalign{\smallskip}
Method & Precision & Recall & F-score\\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
\cellcolor{blue!25}\textbf{Ours} & \cellcolor{blue!25}0.63 & \cellcolor{blue!25}\textbf{0.54} & \cellcolor{blue!25}\textbf{0.58}\\
Yin \textit{et al.} \cite{Yin2013,kumar2013multi} & \textbf{0.64} & 0.42 & 0.51\\
\hline
\end{tabular}
\end{center}
\end{table}
\setlength{\tabcolsep}{1.4pt}
Figure~\ref{fig:inverse} show the inverse grade curves of different methods in the MSRRC dataset. The inverse grade curve plots the f-score divided by the ratio of text pixels for each image, and inversely sorts these values by the amount of text pixels, thus larger values in the x-axis correspond to images with less text. As can be seen our curve is the nearest to follow the ground-truth benchmark curve.
\begin{figure*}
\centering
\includegraphics[width=\linewidth]{icdar_localization_6.jpg}
\caption{Qualitative localization results on the ICDAR 2013 dataset.}
\label{fig:icdar_localization}
\end{figure*}
The KAIST dataset~\cite{Lee2010} comprises 3000 natural scene images, with a resolution of $640\times480 pixels$, categorized according to the language of the scene text captured: Korean, English, and Mixed (Korean + English). For our experiments we use the subset of 800 images corresponding to the Mixed subset in accordance to other reported results. Quantitative results are given in Table~\ref{table:results_kaist} while a set of example qualitative results are shown in Figure~\ref{fig:kaist_ok}. Our method archives a $0.77$ f-score in the localization task, with a $0.71$ precision and $0.83$ recall, not comparable to other methods as being the first to report localization results for this dataset. The average run-time of our algorithm in this dataset is 0.5 seconds per image on a standard PC.
\setlength{\tabcolsep}{4pt}
\begin{table}[h]
\begin{center}
\caption{Segmentation results in the KAIST dataset.}
\label{table:results_kaist}
\begin{tabular}{l c c c}
\hline\noalign{\smallskip}
Method & Precision & Recall & F-score\\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
\cellcolor{blue!25}\textbf{Ours} & \cellcolor{blue!25}\textbf{0.67} & \cellcolor{blue!25}\textbf{0.89} & \cellcolor{blue!25}\textbf{0.76}\\
Gomez \& Karatzas \cite{Gomez2013}& 0.66 & 0.78 & 0.71 \\
Lee \textit{et al.} \cite{Lee2010} & 0.69 & 0.60 & 0.64 \\
OTCYMIST \cite{kumar2012otcymist} & 0.52 & 0.61 & 0.56 \\
\hline
\end{tabular}
\end{center}
\end{table}
\setlength{\tabcolsep}{1.4pt}
The interpretation of the high increase in recall observed in the KAIST dataset compared to the obtained in MSRRC follows the fact that in KAIST dataset small text characters are not labelled in the groud-truth. These small text components are in general the ones more difficult to detect. On the other hand, in some cases precision suffers when such small text is correctly detected as it counts as false positive.
\subsection{English horizontal scene text extraction}
The proposed method has been evaluated on the ICDAR2013 Robust Reading Dataset~\cite{karatzas2013icdar}. The ICDAR2013 dataset contains 462 images, of which 229 comprise the training set and 233 images the test set. Table~\ref{table:results_icdar} compares the results of our method with the participants in the 2013 ICDAR Robust Reading Competition for the task of text segmentation. The average run-time of our algorithm in this dataset is 1.78 seconds per image on a standard PC.
\setlength{\tabcolsep}{4pt}
\begin{table}[h]
\begin{center}
\caption{Segmentation results in the ICDAR Robust Reading Competition 2013 dataset.}
\label{table:results_icdar}
\begin{tabular}{l c c c}
\hline\noalign{\smallskip}
Method & Precision & Recall & F-score\\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
I2R NUS FAR \cite{karatzas2013icdar} * & \textbf{0.82} & \textbf{0.75} & \textbf{0.78}\\
I2R NUS \cite{karatzas2013icdar} * & 0.79 & 0.73 & 0.76\\
\cellcolor{blue!25}\textbf{Ours} & \cellcolor{blue!25}0.74 & \cellcolor{blue!25}0.71 & \cellcolor{blue!25}0.73\\
USTB FuStar \cite{Yin2013,karatzas2013icdar}& 0.74 & 0.70 & 0.72\\
Text Detection \cite{karatzas2013icdar}& 0.76 & 0.65 & 0.70 \\
NSTextractor \cite{karatzas2013icdar}& 0.76 & 0.61 & 0.68 \\
NSTsegmentator \cite{karatzas2013icdar}& 0.64 & 0.68 & 0.66\\
Gomez \& Karatzas 2013 \cite{Gomez2013}& 0.63 & 0.59 & 0.61\\
OTCYMIST \cite{kumar2012otcymist,karatzas2013icdar}& 0.46 & 0.59 & 0.52\\
\hline
\end{tabular}
\end{center}
\end{table}
\setlength{\tabcolsep}{1.4pt}
\setlength{\tabcolsep}{4pt}
\begin{table}[h]
\begin{center}
\caption{Localization results in the ICDAR Robust Reading Competition 2013 dataset.}
\label{table:results_icdar2}
\begin{tabular}{l c c c}
\hline\noalign{\smallskip}
Method & Precision & Recall & F-score\\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
USTB TexStar \cite{karatzas2013icdar,Yin2013} & 0.88 & 0.66 & 0.76\\
TextSpotter \cite{karatzas2013icdar,Neumann2012}& 0.88 & 0.65 & 0.74\\
CASIA NLPR \cite{karatzas2013icdar} & 0.79 & 0.68 & 0.73\\
\cellcolor{blue!25}\textbf{Ours} & \cellcolor{blue!25}0.78 & \cellcolor{blue!25}0.67 & \cellcolor{blue!25}0.72\\
Text detector CASIA \cite{karatzas2013icdar}& 0.85 & 0.63 & 0.72\\
I2R NUS FAR \cite{karatzas2013icdar} * & 0.75 & 0.69 & 0.72\\
I2R NUS \cite{karatzas2013icdar} * &0.73 & 0.66 & 0.69\\
TH-TextLoc \cite{karatzas2013icdar}&0.70 & 0.65 & 0.67\\
Text Detection \cite{karatzas2013icdar}& 0.74 & 0.53 & 0.62\\
Baseline \cite{karatzas2013icdar}& 0.61 & 0.35 & 0.44\\
Inkam \cite{karatzas2013icdar}& 0.31 & 0.35 & 0.33\\
\hline
\end{tabular}
\end{center}
\end{table}
\setlength{\tabcolsep}{1.4pt}
\begin{figure*}
\centering
\includegraphics[width=\linewidth]{icdar_ok.jpg}
\caption{Qualitative segmentation results on the ICDAR dataset.}
\label{fig:icdar_ok}
\end{figure*}
\begin{figure*}
\centering
\includegraphics[width=\linewidth]{icdar_fails.jpg}
\caption{Common errors in the ICDAR dataset include false positives due to repetitive patterns and missing text in images with strong highlights, degraded text, or individual characters.}
\label{fig:icdar_fails}
\end{figure*}
As can be seen in Tables~\ref{table:results_icdar} and \ref{table:results_icdar2} our method produces competitive results although it does not perform better than the winner methods in the last ICDAR competition. This has a coherent interpretation as we aim for the highest generality of our method, addressing the unconstrained problem of detecting text irrespective of its language, script, and orientation. Contrary to our method, most methods listed in Tables~\ref{table:results_icdar} and \ref{table:results_icdar2} have been trained explicitly for horizontally aligned English text and address only this particular scenario. For example the TextSpotter \cite{Neumann2012} method is reported to be specifically designed to detect only English and horizontal text. The USTB TextStar~\cite{Yin2013} is a multi-script method but has two different variants for horizontal and arbitrary oriented text, while scoring first in the ICDAR dataset, Tables~\ref{table:results_msrrc} and ~\ref{table:results_msrrc2} show that our method is superior in other scenarios. Methods marked with an asterisk in Tables~\ref{table:results_icdar} and \ref{table:results_icdar2} have not been published.
Finally, we evaluate our method in the ICDAR2003 dataset~\cite{Lucas2003}. This is a slightly different version of the ICDAR dataset, with almost the same images but proposing a distinct evaluation framework. Table~\ref{table:results_icdar3} compare our results in the ICDAR2003 dataset with other state of the art methods. It is important to notice that some of the top scoring methods in this table have been evaluated in the MSRA-TD500 arbitrary oriented text dataset with a much worse performance compared to the method proposed here, as can be seen in Table~\ref{table:results_msra}. This is again because such methods are designed specifically for the solely detection of English horizontal text.
\setlength{\tabcolsep}{4pt}
\begin{table}[h]
\begin{center}
\caption{Localization results in the ICDAR 2003 dataset.}
\label{table:results_icdar3}
\begin{tabular}{l c c c}
\hline\noalign{\smallskip}
Method & Precision & Recall & F-score\\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
\cellcolor{blue!25}\textbf{Ours} & \cellcolor{blue!25}0.74 & \cellcolor{blue!25}0.65 & \cellcolor{blue!25}0.69\\
TD-Mixture \cite{yao2012detecting} & 0.69 & 0.66 & 0.67\\
TD-ICDAR \cite{yao2012detecting} & 0.68 & 0.66 & 0.66\\
Epshtein \textit{et al.}~\cite{Epshtein2010}& 0.73 & 0.60 & 0.66\\
Yi \textit{et al.} \cite{yi2011}&0.71 & 0.62 & 0.62\\
Becker \textit{et al.} \cite{Lucas2003}&0.62 & 0.67 & 0.62\\
Chen \textit{et al.}~\cite{Chen2004}& 0.60 & 0.60 & 0.58\\
Zhu \textit{et al.} \cite{Lucas2003}&0.33 & 0.40 & 0.33\\
Kim \textit{et al.} \cite{Lucas2003}&0.22 & 0.28 & 0.22\\
Ezaki \textit{et al.} \cite{Lucas2003}&0.18 & 0.36 & 0.22\\
\hline
\end{tabular}
\end{center}
\end{table}
\setlength{\tabcolsep}{1.4pt}
\section{Conclusions}
This paper presents a scene text extraction method in which the exploitation of the hierarchical structure of text plays an integral part. We have shown that the algorithm can efficiently detect text groups whith arbitrary orientation in a single clustering process that involves: a learned optimal clustering feature space for text region grouping, novel discriminative and probabilistic stopping rules, and a new set of features for text group classification that can be efficiently calculated in an incremental way.
Experimental results demonstrate that the presented algorithm outperforms other state of the art methods in three multi-script and arbitrary oriented scene text standard datasets while it stays competitive in the more restricted scenario of horizontally-aligned English text ICDAR dataset.
Moreover, the presented results in all datasets are obtained with a single (mixed) training set, demonstrating the general purpose character of the method which yields robust performance in a variety of distictly different scenarios.
Finally, the baseline analysis of the algorithm reveals that overall system recall can be substantially increased if needed by using feature space diversification.
\section*{Acknowledgment}
\label{sec:acknowledgement}
This project was supported by the Spanish project TIN2011-24631 the fellowship RYC-2009-05031, and the Catalan government scholarship 2013FI1126.
\ifCLASSOPTIONcaptionsoff
\newpage
\fi
\bibliographystyle{IEEEtran}
|
\section{Introduction}
Biological and soft matter systems are characterized by the existence
of processes with many different length and time scales. These
processes are usually coupled, making their theoretical, experimental,
and computer simulation description a daunting task. The functioning
of a protein, for example, involves chemical processes at active sites
as well as the overall dynamics of the protein and its environment
\cite{Neri2005}. Crack propagation is another example in which the
atomic processes occuring at the crack tip affect crucially the
overall elastic behaviour of the sample, and vice-versa
\cite{Csanyi2004c}. From a computational point of view, the brute
force approach of treating the system with full molecular detail is
not possible, and one needs to deal with simplified, or coarse-grained
versions of the system \cite{Voth2009}. By definition, in any coarse-grained model
some atomic/molecular detail is lost. In some fortunate cases, the
need for atomistic detail is confined in small regions of space as in
the examples above, and there is hope that a hybrid scheme coupling
all atom (AA) with coarse-grained (CG) descriptions may be a successful
approach. The coupling of different models describing the system at
different resolutions is an active field of research \cite{Park2009,Wang2013}
and in our opinion it will become useful in a broad range of calculations,
beyond the multi-scale community \cite{Wang2013}.
We have recently developed an Hamiltonian Adaptive Resolution Scheme (\hadr)
\cite{Potestio:PRL:110,Potestio:PRL:111}. Other proposals for
Hamiltonian hybrid (AA/CG) schemes have been presented \cite{Park2009}
which are technically challenging as compared with \hadr. As opposed
to previous versions of AdResS, where a \textit{force} interpolation
principle was the crucial element, in \hadr \textit{potentials} are
interpolated. The proposed Hamiltonian in \hadr includes a switching
field that allows for a swift interpolation between the truly
microscopic Hamiltonian and a CG version of it. When a molecule
crosses the interface between the AA and CG regions, its interaction
with other molecules changes accordingly. Usually the CG potential of interaction
used in the CG region is only an approximate version of the actual potential of mean force.
The discrepancies between the CG potential and the potential of mean
force are taken into account in the \hadr Hamiltonian through a {\em
free energy compensation term} \cite{Potestio:PRL:110,Potestio:PRL:111}.
The idea of interpolating AA an CG potentials through a hybrid region
is not new and was introduced in Refs.
\cite{Ensing2007},\cite{Nielsen2010} under the name of adaptive
Multiscale Molecular Dynamics (MMD)\cite{Nielsen2010}. However, the
detailed form of the interpolation is slightly different in \hadr and
leads to the existence of a well-defined Hamiltonian that allows for
the natural use of the principles of Statistical Mechanics. In the
original version of MMD, energy was not conserved
\cite{DelleSite2007},\cite{Praprotnik2011a} and thermostats were
required \cite{Ensing2007},\cite{Nielsen2011}. In the original
thermostatted \adr \cite{Matej_JCP07} and also in more recent versions \cite{Wang2013},
the mass in the atomistic domain fluctuates according to the
Grand-Canonical ensemble; at least up to the second moment of the
probability density function, as it has numerically
\cite{Delgado-Buscalioni2008}, and theoretically shown
\cite{Wang2013}.
Density fluctuations are determined by the fluid
compressibility, specified by the integral of the radial distribution
function, and by finely tunning the CG potential one can match the
compressilibities of the CG and AA domains. Having the same
compressibility does not however ensure the same pressure equation of state
and to ensure a constant density profile over the CG and AA domains, a
recent work \cite{FritschPRL108} proposes the imposition of a ``correction force field'', which
is iteratively evaluated according to the idea of imposing pressure balance
(and thus involving compressibilities). The existence of a Hamiltonian permits
us to derive a fundamental relation between the force density and the
density gradient, which turns out to be independent on the
compressibility. This relation explains the basis of the ``correction
force field'' used to control the density profile, not only
in Ref. \cite{FritschPRL108} but also in many other algorithms using domain
decomposition (see e.g. Refs. \cite{Issa2014,kotsalis07}).
The Hamiltonian
system of \hadr also allows to extend the working ensemble also to the
$E,N,V$ microcanonical ensemble, with no need of any extraneous
thermostats. However, the benefits of a Hamiltonian description will
prove to be substantially broader, as already shown in
Ref. \cite{Potestio:PRL:111}.
In the present paper, we derive the Statistical Mechanics basis for
the \hadr method. Several exact results concerning the local equations
of state for the pressure and temperature allow for the formulation of the
free energy compensation term in an iterative way. We also show that
under a local equilibrium approximation, valid when the hybrid region is wide,
the iterative procedure can be simplified leading to an approximate but very efficient way for the
calculation of the free energy compensation term in the Hamiltonian.
We have analyzed the effect of the width of the transition layer
where molecules gradually change their resolution. A relevant outcome
is that the \hadr {\em total} free energy compensation is independent
on the layer, even for widths of the same order of the molecular
diameter. Another very important observation is that the \hadr total
free energy correction is equal, within error bars, to the free energy
difference between both fluids (atomistic and coarse-grained)
evaluated from Kirkwood thermodynamic integration \cite{Kirkwood1935}.
Although more research is required in this direction, this would allow
\hadr to be used as a flexible tool for estimation of free energy
differences in different scenarios.
In what follows, we first present the \hadr Hamiltonian formulation in
Sec. \ref{sec:ham}. The free energy corresponding to the \hadr
Hamiltonian is introduced in Sec. \ref{sec:free}. In Sec. \ref{sec:PT}
we derive expressions for the temperature and the pressure tensor fields. In
Sec. \ref{sec:trans} we demonstrate that the condition of constant
pressure over the \hadr simulation stems from the condition of
translational invariance of the free energy. The force balance
equation derived in Sec. \ref{sec:itera} permits to rationalize the
different types of \hadr compensation terms, for either constant
pressure or density fields. Section \ref{sec:lea} shows that under local
equilibrium (LE) conditions, the FEC is just the Kirkwood free energy
difference \cite{Kirkwood1935}, thus justifying the non-iterative
route used in our previous works
\cite{Potestio:PRL:110,Potestio:PRL:111}. Finally, the theoretical
framework is validated through simulations in Sec.
\ref{sec:simulations} where we also provide relaxational schemes for
the iterative route to the FEC. We also study the effect of the
transition layer width and the deviation from the Kirkwood
approximation to the FEC. Conclusions and some future perspectives
are given in Sec. \ref{sec:conclu}.
\section{The AdResS Hamiltonian }
\label{sec:ham}
Consider a classic molecular system composed of $N$ constituent atoms. The
microscopic state of the system is described by the positions and
momenta of the atoms, denoted generically by $r,p$. The system is
coarse-grained by considering the centers of mass (CoM) of $M$ groups
of atoms that are bound together and that are termed {\em blobs}. A
blob may be, for example, a single molecule or a part of a bigger
molecule. The position of the $\mu$-th blob CoM is $\hat{\bf R}_\mu$
which is defined as the following phase function
\begin{align}
\hat{\bf R}_\mu(r)&=\sum_i^N\delta_\mu(i){\bf r}_i\frac{m_i}{M_\mu}=\sum^{N_\mu}_{i_\mu}{\bf r}_{i_\mu}\frac{m_{i_\mu}}{M_\mu}
\nonumber\\
M_\mu&=\sum_i^N\delta_\mu(i)m_i
\label{1}
\end{align}
where the indicator symbol $\delta_\mu(i)$ takes the value 1 if atom
$i$ is in blob $\mu$ and zero otherwise. The last definition makes
use of the notation $i_\mu$ that corresponds to the $i$-th atom of
blob $\mu$ and ${N_\mu}$ is the number of atoms of blob $\mu$. The
microscopic Hamiltonian governing the dynamics of the atoms is
\begin{align}
H^1(r,p) &=\sum_{i}^{N}\frac{{\bf p}_{i}^2}{2m_{i}}+\sum_\mu^MV^{\rm intra}_\mu(r)+V^1(r)
\label{H1}
\end{align}
where the total potential energy of interaction of the atoms is
decomposed into the potential of interaction between atoms within a
blob $V^{\rm intra}_\mu(r)$ and the potential of interaction between
atoms of different blobs $V^1$. This potential energy can be
decomposed as $V^1=\sum_\mu^MV^1_\mu$ where the terms $V^1_{\mu}(r)$
are the potential energy of interaction of the atoms of different
blobs where one of the atoms of the pair is in blob
$\mu$. Explicitly
\begin{align}
V^{1}_{\mu}(r)&=\frac{1}{2}\sum_{ij}^N \delta_\mu(i)\phi^{\rm inter}(r_{ij})
\end{align}
where $\phi^{\rm inter}$ is the pair potential between atoms $i,j$ of
different blobs. It is understood that $\phi^{\rm inter}(r_{ij})$ is zero
if atoms $i,j$ belong to the same blob. Note that any Hamiltonian
that differs from the one in Eq. (\ref{H1}) by a constant term will
produce exactly the same dynamics. The usual convention is to chose
the zero of potential energy in such a way that when the particles are
very far apart and, therefore, non-interacting, the potential energy
is zero. This fixes the origin of the energy scale. We will assume
that the above Hamiltonian has, through a mixing property of its
Hamiltonian flow, a well defined equilibrium ensemble. For this to be
true it is necessary that the {\em finite} system of $N$ atoms were
confined, either by a time-independent external field (not included in
(\ref{H1})) or through periodic boundary conditions.
The central idea of \hadr is to introduce a switching field
$\lambda({\bf r})$ that takes the value 1 in the region of space where
the system is described in full all atomic (AA) detail, and the value
0 in the region of space where the system is described in a
coarse-grained (CG) way. In the transition region between the two
zones the switching field changes monotonously from 0 to 1. The field
$\lambda({\bf r})$ gives the degree of detail of the description.
Instead of the microscopic Hamiltonian (\ref{H1}), the dynamics of the
atoms is modified with the following \hadr Hamiltonian,
\begin{align}
H_{[\lambda]}(r,p)&=\sum_i^N \frac{{\bf p}_i^2}{2m_i}+V_{[\lambda]}(r)
\nonumber\\
V_{[\lambda]}(r)&=\sum_\mu^M V^{\rm intra}_\mu(r)+
\sum_{\mu}^M\lambda(\hat{\bf R}_\mu)V^{1}_{\mu}(r)
\nonumber\\
&+\sum_{\mu}^M(1-\lambda(\hat{\bf R}_\mu))V^0_{\mu}(R)
+\sum_\mu^M\calf(\lambda(\hat{\bf R}_\mu))
\label{HA}
\end{align}
The potential $V^{0}_\mu(R)$ is assumed to depend on the atomic
coordinates $r$ {\em only} through the position of the centers of
mass, denoted collectively as $R=\{\hat{\bf
R}_\mu(r),\mu=1,\cdots,M\}$. Although the present formalism is general and allows
for multi-body CG potentials, in most practical cases a pair-wise form will be assumed, this is
\begin{align}
V^{0}_\mu(R)&=\frac{1}{2}\sum_{\nu}^MV^{0}(\hat{\bf R}_\mu-\hat{\bf R}_\nu)
\label{pw}
\end{align}
The potential $V^{0}_\mu(R)$ describes the interaction between blobs in
a coarse-grained way.
The term $\sum_\mu^M\calf(\lambda(\hat{\bf R}_\mu))$ in the
Hamiltonian is referred to as {\em the free energy compensation
term}. Its effect is very much like an external field acting on the
blobs. We require that $\calf(1)=0$.
The rationale for postulating the above Hamiltonian is the
following. When $\lambda({\bf r}) =1$ the above Hamiltonian coincides
with the microscopic Hamiltonian (\ref{H1}), this is $H_{[1]}=H^1$.
On the other hand, when $\lambda({\bf r})=0$ the Hamiltonian becomes
\begin{align}
H_{[0]}(r,p)&=\sum_i^N\frac{{\bf p}_i^2}{2m_i}+\sum_\mu^M V^{\rm intra}_\mu(r)
\nonumber\\
&+
\sum_{\mu}^MV^{0}_{\mu}(R)
+\sum_\mu^M\calf(0)
\label{HAl=0}
\end{align}
where, apart from the constant term $\sum_\mu^M\calf(0)$,
the potential of interaction between atoms of different blobs is given
by the CG interaction. Therefore, the idea is that with a spatially
varying $\lambda({\bf r})$ the blobs change its interaction from its
real microscopic interaction $V^{1}(r)$ to a CG interaction through
its centers of mass $V^{0}(R)$. In fact, the equations of motion
produced by the Hamiltonian (\ref{HA}) are (assume that particle $i$
belongs to blob $\mu$)
\begin{align}
\dot{\bf r}_i &=\frac{{\bf p}_i}{m_i}
\nonumber\\
\dot{\bf p}_i &= -\frac{\partial V_\mu^{\rm intra}}{\partial {\bf r}_i}
-
\sum^M_{\nu}\lambda(\hat{\bf R}_\nu)\frac{\partial V_\nu^{1}}{\partial {\bf r}_i}
-\sum^M_{\nu}(1-\lambda(\hat{\bf R}_\nu))\frac{\partial V^{0}_{\nu}}{\partial {\bf r}_i}
\nonumber\\
&-\nabla \lambda({\bf R}_\mu)\frac{m_i}{m_\mu}\left(V^{1}_{\mu}-V^{0}_{\mu}+
\calf'(\lambda(\hat{\bf R}_\mu))
\right),
\label{eqmot1}
\end{align}
where the prime ($\calf' = d\calf/d\lambda$) denotes derivative with
respect to $\lambda$. When $\lambda=1$ Eq. (\ref{eqmot1}) correspond to
the fully resolved microscopic dynamics, this is
\begin{align}
\dot{\bf r}_i &=\frac{{\bf p}_i}{m_i}
\nonumber\\
\dot{\bf p}_i &= -\frac{\partial V_\mu^{\rm intra}}{\partial {\bf r}_i}
-\sum^M_{\nu}\frac{\partial V_\nu^{1}}{\partial {\bf r}_i}
\end{align}
When $\lambda=0$ Eq. (\ref{eqmot1}) become
\begin{align}
\dot{\bf r}_i &=\frac{{\bf p}_i}{m_i}
\nonumber\\
\dot{\bf p}_i &= -\frac{\partial V_\mu^{\rm intra}}{\partial {\bf r}_i}
-\sum^M_{\nu}\frac{\partial V^{0}_{\nu}}{\partial {\bf r}_i}
\end{align}
that describes the motion of the atoms as given in terms of
microscopic forces due to the atoms of the same blob and CG
interactions between the centers of mass of the blobs. In this way,
{\em in the CG region, the Hamiltonian of \hadr moves the atoms with CG
interactions.}
In the transition region when $0<\lambda<1$ the atoms move with a
combination of the microscopic and CG potentials and, in addition,
feel the presence of an ``external field'', represented in the last
term of the momentum equation (\ref{eqmot1}), which is proportional to
the gradient of $\lambda$. The contribution $\calf'(\lambda)$ that
appears in Eq. (\ref{eqmot1}) has the mission to
make this ``external field'' effect as small as possible, in a
statistical sense. We will give in the next section a thermodynamic
interpretation to the $\calf(\lambda)$ contribution in the
Hamiltonian. A molecular dynamics simulation with the Hamiltonian
(\ref{HA}) can be coded in a way that the simulation proceeds much
faster than the one given by the full microscopic Hamiltonian
(\ref{H1}). Indeed, in the CG region the forces on the atoms need a
search only of the neighbouring {\em blobs} whose number is much
smaller than the number of atoms required in the microscopic
evaluation and indeed in the CG domain,
the number of force evaluations is drastically reduced.
Note that the way in which the AA and CG potentials are interpolated
in the Hamiltonian (\ref{HA}) is different from the interpolation in
the MMD method \cite{Ensing2007,Nielsen2011} where in the latter
method the switching function depends on the position of the centers
of mass of two blobs instead of just one blob in \hadr.
\section{ The free energy}
\label{sec:free}
The thermodynamic free energy corresponding to the AdResS Hamiltonian
(\ref{HA}) is given by the usual statistical mechanics formula
\begin{align}
F_{[\lambda]}&=-k_BT\ln\int{d^{3N}rd^{3N}p}\exp\left\{-\beta
H_{[\lambda]}(r,p)\right\}
\nonumber\\
&=-k_BT\ln\int \frac{d^{3N}r}{\Lambda^{3N}}\exp\left\{-\beta
V_{[\lambda]}(r)\right\}
\label{Flam}
\end{align}
and it is a functional of the switching field $\lambda({\bf r})$. In
this expression the momentum integrals of the kinetic energy in the
Hamiltonian have been performed giving rise to the factor
$\Lambda^{3N}$
\begin{align}
\Lambda^{3N}\equiv\prod_{\mu}^M\prod_{i_\mu}^{N_\mu}\Lambda_{i_\mu}^{3}
\end{align}
where the thermal wavelength of atom $i_\mu$ is defined as
\begin{eqnarray}
\Lambda_{i_\mu}=\left(\frac{h^2}{2\pi k_BT m_{i_\mu}}\right)^{1/2}
\end{eqnarray}
The macroscopic thermodynamic free energy can be expressed in
terms of a potential of mean force by introducing the identity in
the form
\begin{align}
1=\int d^{3M}R\prod_\mu^M\delta({\bf R}_\mu-\hat{\bf
R}_\mu(r) )
\label{1delta}
\end{align}
Recall that $\hat{\bf R}_\mu(r) $ is a phase function that depends on the positions of the
atoms of blob $\mu$, i.e. Eq. (\ref{1}).
\begin{widetext}
By inserting (\ref{1delta}) inside the free energy (\ref{Flam}) leads to
\begin{align}
F_{[\lambda]}
&=-k_BT\ln\int \frac{d^{3M}R}{\Lambda_0^{3M}}\exp\left\{-\beta\left[
\sum_{\mu}^M(1-\lambda({\bf R}_\mu))V^{0}_{\mu}(R)+V^{\rm mf}_{[\lambda]}(R)
+\sum_\mu^M\calf(\lambda({\bf R}_\mu))\right]\right\}
\label{Fm1}
\end{align}
where the potential of mean force is defined as
\begin{align}
V^{\rm mf}_{[\lambda]}(R)&\equiv-k_BT\ln \int\frac{d^{3N}r}{\Lambda^{3N}}\exp\left\{-\beta
\left[V^{\rm intra}(r)+
\sum_{\mu}^M\lambda({\bf R}_\mu)V^{1}_{\mu}(r)
\right]\right\} \Lambda_0^{3M} \prod_\mu^M\delta({\bf R}_\mu-\hat{\bf R}_\mu(r))
\label{Vmf}
\end{align}
$\Lambda_0$ is an {\em arbitrary} length scale that renders the argument of
the logarithms in Eqs. (\ref{Fm1}) and (\ref{Vmf}) dimensionless.
The potential of mean force (\ref{Vmf}) is a functional of the
switching field $\lambda$. When $\lambda({\bf r})=1$, the effective
potential $V^{\rm mf}_{[1]}(R)$ coincides with the potential of mean
force of the fully microscopic Hamiltonian $H_{[1]}(r,p)$, this is
\begin{align}
V^{\rm mf}_{[1]}(R)&\equiv-k_BT\ln \int\frac{d^{3N}r}{\Lambda^{3N}}\exp\left\{-\beta
\left[V^{\rm intra}(r)+
\sum_{\mu}^MV^{1}_{\mu}(r)
\right]\right\}\Lambda_0^{3M}\prod_\mu^M\delta({\bf R}_\mu-\hat{\bf R}_\mu)
\label{Vmf1}
\end{align}
On the other hand, when $\lambda({\bf r})=0$, we have
\begin{align}
V^{\rm mf}_{[0]}(R)
&\equiv-k_BT\ln \int\frac{d^{3N}r}{\Lambda^{3N}} \exp\left\{-\beta V^{\rm
intra}(r)\right\}\Lambda_0^{3M}\prod_\mu^M\delta({\bf
R}_\mu-\hat{\bf R}_\mu(r))
\nonumber\\
&=-k_BT\ln \prod_\mu^M\int\frac{d^{3N_\mu}r}{\Lambda^{3N}}\exp\left\{-\beta
V_\mu^{\rm intra}(r_\mu)\right\} \Lambda_0^3\delta({\bf
R}_\mu-\hat{\bf R}_\mu) = \sum_\mu^M F_\mu^{\rm intra}
\label{veff0}
\end{align}
where we have introduced the actual thermodynamic free
energy $F^{\rm intra}_\mu$ that a blob would have should it be
isolated from the rest of blobs, this is
\begin{align}
\exp\left\{-\beta F_\mu^{\rm intra}\right\}
&\equiv \int\frac{d^{3N_\mu}r}{\Lambda^{3N_\mu}}
\exp\left\{-\beta V_\mu^{\rm intra}(r_\mu)\right\}\Lambda_0^3
\delta({\bf R}_\mu-\hat{\bf R}_\mu)
\label{Fmu}\end{align}
\end{widetext}
Note that, in spite of the appearance of the Dirac delta function in Eq. (\ref{Fmu})
depending on ${\bf R}_\mu$, this internal blob free energy $F_\mu^{\rm
intra}$ is independent of ${\bf R}_\mu$ due to translational
invariance. Therefore, we may integrate both sides of (\ref{Fmu})
with respect to ${\bf R}_\mu$ leading to
\begin{align}
\exp\left\{-\beta F_\mu^{\rm intra}\right\}& =\frac{\Lambda_0^3}{V}\int\frac{d^{3N_\mu}r}{\Lambda^{3N_\mu}}\exp\left\{-\beta
V_\mu^{\rm intra}(r_\mu)\right\}
\end{align}
where $V$ is the total volume of the system.
Therefore, in the two limits $\lambda({\bf r})=1$, $\lambda({\bf
r})=0$, the free energy (\ref{Flam}) becomes
\begin{align}
F_{[1]}
&=-k_BT\ln \int\frac{ d^{3M}R}{\Lambda_0^{3M}}\exp\left\{-\beta
V_{[1]}^{\rm mf}(R)\right\}
\nonumber\\
F_{[0]}
&=-k_BT\ln\int \frac{ d^{3M}R}{\Lambda_0^{3M}}\exp\left\{-\beta
\sum_{\mu}^M\left[V^{0}_{\mu}(R)
+F^{\rm intra}_\mu\right]\right\}
\nonumber\\
&+M \calf(0)
\label{Fm4}
\end{align}
The requirement of thermodynamic consistency between both levels of
resolution enforces that the thermodynamic free energy should be
exactly the same in both limits, that is,
\begin{align}
F_{[0]}&= F_{[1]}
\label{F0F1}
\end{align}
This thermodynamic consistency requirement gives light to the meaning
of the free energy compensating term $\calf (\lambda)$. In the spirit
of changing the resolution, we expect that $V_0(R)$ in Eq. (\ref{HA})
is given by the potential of mean force of the microscopic Hamiltonian
(\ref{H1}). This potential of mean force can be measured in different
ways, from Boltzmann inversion \cite{Faller2004} to relative entropy
\cite{Shell2008a} methods. These methods allow one to obtain $V_0(R)$
{\em up to an arbitrary constant}. Indeed $V_0(R)$ is a mesoscopic
free energy for which only relative values may be computed. This
constant is usually fixed by requiring that $V_0(R)$ vanishes as the
centers of mass become apart, i.e. $|{\bf R}_\mu-{\bf
R}_\nu|\to\infty$. On the other hand, the potential of mean force
$V_{[1]}^{\rm mf}(R)$ of the microscopic Hamiltonian contains
information of not only the interactions between blobs but also about
the internal free energy of the molecules. One way in which this
clearly manifests is when the blobs in which we have grouped the atoms
correspond to full molecules. In that case it makes sense to look at
the low density regime in which the molecules are very far from each
other. In this limit, we obtain from Eq. (\ref{Vmf1}) that when the
centers of mass are separated beyond the range of interaction of the
potentials, then we may neglect the term $V_\mu^1(r)$ in Eq.
(\ref{Vmf1}), leading to $V^{\rm mf}_{[1]}(R)=\sum_\mu F_\mu^{\rm
intra}$. As a result, the potential of
mean force $V_{[1]}^{\rm mf}(R)$ does not vanish as the distance
between particles goes to infinity, as opposed to $V_0(R)$. If we
momentarily assume that the many-body potential of mean force $V^{\rm
mf}_{[1]}(R)$ could be very well approximated by a pair-wise form,
we would choose the pair-wise potential $V^{0}(R)$ as $V^{0}(R)=V^{\rm
mf}_{[1]}(R)-\sum_\mu F_\mu^{\rm intra}$ (vanishing as the CoM
separate). In that situation, the consistency (\ref{F0F1}) would
imply $\calf(0)=0$. It is clear, therefore, that the contribution
$\calf(0)$ has the effect of ``curing'', at the level of
thermodynamics, the errors due to the use of an approximate pair-wise
potential $V_0(R)$ for the actual many-body potential of mean force
$V_{[1]}^{\rm mf}(R)$.
The free energy (\ref{Flam}) is a functional of the switching field
$\lambda({\bf r})$. For future reference, we compute explicitly the
functional derivative of the free energy with respect to $\lambda({\bf
r})$, this is
\begin{align}
\frac{\delta F_{[\lambda]}}{\delta\lambda({\bf r})}
&=\left\langle \frac{\delta H_{[\lambda]}}
{\delta \lambda({\bf r})}\right\rangle^{[\lambda]}
\label{dhdla}
\end{align}
In this expression, $\langle\cdots\rangle^{[\lambda]}$ is a
canonical average with the AdResS Hamiltonian $H_{[\lambda]}$
in Eq. (\ref{HA}). By using
\begin{align}
\frac{\delta H_{[\lambda]}}{{\delta \lambda({\bf r})}}&
=\left( u^1_{{\bf r}}-u^0_{{\bf r}}
+\calf'(\lambda({\bf r}))n_{{\bf r}}\,\right)
\label{dhdl}
\end{align}
where have defined the potential energy
densities $ {u}^0_{\bf r}, {u}^1_{\bf r}$ and the center of mass
density ${n}_{\bf r}$ as
\begin{align}
{u}^1_{\bf r}&\equiv\sum^M_\mu V^1_\mu\delta(\hat{\bf R}_\mu-{\bf r})
\nonumber\\
{u}^0_{\bf r}&\equiv\sum^M_\mu V^{0}_\mu\delta(\hat{\bf R}_\mu-{\bf r})
\nonumber\\
{n}_{\bf r}&\equiv\sum^M_\mu\delta(\hat{\bf R}_\mu-{\bf r})
\label{defu1u0}
\end{align}
we finally obtain the explicit expression for the functional derivative of the free energy of \hadr
\begin{align}
\frac{\delta F_{[\lambda]}}{\delta\lambda({\bf r})}=\left\langle {u}^1_{\bf r}-{u}^0_{\bf r}\right\rangle^{[\lambda]}+
\calf'(\lambda({\bf r})) \left\langle {n}_{\bf r}\right\rangle^{[\lambda]}
\label{dfdl=01}
\end{align}
This expression will be used below.
\section{The temperature and pressure fields}
\label{sec:PT}
In the previous section, we have presented a consistency argument in
Eq. (\ref{F0F1}) based on the global thermodynamics of the \hadr
system. In this section we formulate the local thermodynamics of
\hadr in terms of the equations of state for the temperature and the
pressure. In order to achieve this, it is convenient to look at the
{\em molecular} momentum density field because its time derivative
will give information about mechanical equilibrium and, hence,
pressure. The molecular momentum density field is defined as
\begin{align}
\hat{\bf g}_{\bf r}(z) &\equiv\sum_\mu^M \hat{\bf P}_\mu\delta(\hat{\bf R}_\mu-{\bf r})
\label{hatg}
\end{align}
where the momentum $\hat{\bf P}_\mu$ of blob $\mu$ is given by
\begin{align}
\hat{\bf P}_\mu(r)&=\sum_i^N\delta_\mu(i){\bf p}_i
\end{align}
The time derivative of the phase function (\ref{hatg}) is
obtained by applying the Liouville operator onto this
function, providing
\begin{align}
iL\hat{\bf g}_{\bf r}&=\hat{\bf f}_{\bf r} -\nabla\hat{\bf K}_{\bf r}
\label{lrlg}
\end{align}
where the {\em kinetic part of the stress tensor} is defined as
\begin{align}
\hat{\bf K}_{\bf r}&\equiv\sum_\mu^M \hat{\bf P}_\mu\hat{\bf V}_\mu\delta(\hat{\bf R}_\mu-{\bf r}).
\label{kinst}
\end{align}
The velocity is $\hat{\bf V}_\mu=\hat{\bf P}_\mu/M_\mu$, and the {\em force density} is defined as
\begin{align}
\hat{\bf f}_{\bf r}&\equiv \sum_\mu^M \hat{\bf F}_\mu\delta(\hat{\bf R}_\mu-{\bf r})
\label{fr}
\end{align}
Here, $\hat{\bf F}_\mu$ is the force on molecule $\mu$ which is given by
\begin{align}
\hat{\bf F}_\mu&\equiv-\sum_i\delta_\mu(i) \frac{\partial H_{[\lambda]}}{\partial {\bf r}_i}.
\label{fmu}
\end{align}
In the Appendix \ref{Ap:force}, it is shown that the force ${\bf F}_\mu$ on molecule $\mu$
introduced in Eq. (\ref{fmu}) has the following form
\begin{align}
\hat{\bf F}_\mu
&=\sum_{\nu} \hat{\bf G}_{\mu\nu}
\nonumber\\
&-\nabla\lambda(\hat{\bf R}_\mu)(V^{1}_{\mu}(r)-V^{0}_{\mu}(R)+\calf'(\lambda_\mu(R)))
\end{align}
where we have introduced the pair force
\begin{align}
\hat{\bf G}_{\mu\nu}&\equiv
\left[\frac{\lambda(\hat{\bf R}_\mu)+\lambda(\hat{\bf R}_\nu)}{2}\right]
{\bf F}^{1}_{\mu\nu}(R_{\mu\nu})
\nonumber\\
&+\left[1-\frac{\lambda(\hat{\bf R}_\mu)+\lambda(\hat{\bf R}_\nu)}{2}\right]
{\bf F}^{0}_{\mu\nu}(R_{\mu\nu})
\end{align}
This force satisfies Newton's Third Law $ \hat{\bf G}_{\mu\nu}=-
\hat{\bf G}_{\nu\mu}$. The forces ${\bf F}^1_{\mu\nu},{\bf
F}^0_{\mu\nu}$ introduced in Appendix \ref{Ap:force} are the
original microscopic and CG forces between blobs, respectively. We may
compute now the force density $\hat{\bf f}_{\bf r}$ in Eq. (\ref{fr})
and obtain
\begin{align}
\sum_\mu^M \hat{\bf F}_\mu\delta(\hat{\bf R}_\mu-{\bf r})
&=\sum_{\mu \nu}
\delta(\hat{\bf R}_\mu-{\bf r}) \hat{\bf G}_{\mu\nu}
\nonumber\\
&-\nabla\lambda({\bf r})\left[
\hat{u}^{1}_{\bf r}-\hat{u}^{0}_{\bf r} +\calf'(\lambda({\bf r}))\hat{n}_{\bf r}\right]
\label{fr2}
\end{align}
Note that the last term may be written as the divergence of a tensor, because
\begin{align}
\nonumber\\
\sum_{\mu \nu}
\delta(\hat{\bf R}_\mu-{\bf r}) \hat{\bf G}_{\mu\nu}
&=\sum_{\mu\nu} \hat{\bf G}_{\mu\nu}
\frac{1}{2}\left[\delta(\hat{\bf R}_\mu-{\bf r})-\delta(\hat{\bf R}_\nu-{\bf r})\right]
\nonumber\\
&=-\nabla \hat{\boldsymbol{\Pi}}_{\bf r}
\label{gmunu}
\end{align}
where we have used the usual trick \cite{Grabert1982}
\begin{align}
\delta(\hat{\bf R}_\mu-{\bf r})-\delta(\hat{\bf R}_\nu-{\bf r})&=\int_0^1d\epsilon
\frac{d}{d\epsilon}\delta(\hat{\bf R}_\nu+\epsilon\hat{\bf R}_{\mu\nu}-{\bf r})
\nonumber\\
&=-\nabla{\hat\bf R}_{\mu\nu}\int_0^1d\epsilon
\delta(\hat{\bf R}_\nu+\epsilon\hat{\bf R}_{\mu\nu}-{\bf r})
\end{align}
where we have defined $\hat{\bf R}_{\mu\nu}=\hat{\bf R}_{\mu}-\hat{\bf R}_{\nu}$ and
$R_{\mu\nu}=|{\bf R}_{\mu\nu}|$
and introduced {\em the virial part of the stress tensor}
\begin{align}
\hat{\boldsymbol{\Pi}}_{\bf r}&\equiv\frac{1}{2}\sum_{\mu\nu} \hat{\bf G}_{\mu\nu} \hat{\bf R}_{\mu\nu}
\int_0^1d\epsilon
\delta(\hat{\bf R}_\nu+\epsilon\hat{\bf R}_{\mu\nu}-{\bf r})
\label{Pivir}
\end{align}
In summary, we may write the force density as
\begin{align}
\hat{\bf f}_{\bf r}
&=-\nabla \hat{\boldsymbol{\Pi}}_{\bf r}
-\nabla\lambda({\bf r})\frac{\delta H^{[\lambda]}}{\delta \lambda({\bf r})}
\label{frfin}
\end{align}
where we have used (\ref{dhdl}). As a consequence, the momentum equation (\ref{lrlg}) takes the form
\begin{align}
i L\hat{\bf g}_{\bf r}&=-\nabla\hat{\boldsymbol{\Sigma}}_{\bf r}
-\nabla\lambda({\bf r})
\frac{\delta H^{[\lambda]}}{\delta \lambda({\bf r})}
\label{lrlg2}
\end{align}
where the full stress tensor $\hat{\boldsymbol{\Sigma}}_{\bf r}=\hat{\bf
K}_{\bf r}+\hat{\boldsymbol{\Pi}}_{\bf r}$ is given by the
Irwing-Kirkwood (IK) form, generalized for \hadr,
\begin{align}
\hat{\boldsymbol{\Sigma}}_{\bf r}
&=\sum_\mu^M \hat{\bf P}_\mu\hat{\bf V}_\mu\delta(\hat{\bf R}_\mu-{\bf r})
\nonumber\\
&+\frac{1}{2}\sum_{\mu\nu} \hat{\bf G}_{\mu\nu} {\bf R}_{\mu\nu}
\int_0^1d\epsilon
\delta({\bf R}_\nu+\epsilon{\bf R}_{\mu\nu}-{\bf r})
\label{Sigma0}
\end{align}
\subsection{The temperature}
It is worth considering the equilibrium average computed with the
canonical ensemble of the kinetic part of the stress tensor in
Eq. (\ref{kinst}). It is computed easily because momentum is
distributed according to the Gaussian Maxwell distribution, with the
result
\begin{align}
\langle \hat{\bf K}_{\bf r}\rangle^{[\lambda]}&=
k_BT\langle n_{\bf r}\rangle^{[\lambda]}{\bf 1}
\label{eqkin}
\end{align}
Closely related to the kinetic part of the stress tensor is the
kinetic energy density field of the centers of mass which is defined
as
\begin{align}
k_{\bf r}&\equiv\sum_\mu^M\frac{m_\mu}{2}{\bf V}_\mu^2\delta({\bf r}-{\bf R}_\mu)
\end{align}
and whose average is
\begin{align}
\langle k_{\bf r}\rangle^{[\lambda]}
&= \frac{3k_BT}{2}\langle n_{\bf r}\rangle^{[\lambda]}
\label{prediction}\end{align}
We may introduce a CoM temperature field as the kinetic energy density
divided by the number density, providing an idea of the local kinetic
energy of the system, through the following definition
\begin{align}
k_BT({\bf r}) &\equiv\frac{2}{3}
\frac{\langle k_{\bf r}\rangle^{[\lambda]}}{\langle n_{\bf r}\rangle^{[\lambda]}} =k_BT
\label{tcte}
\end{align}
where the last identity is just Eq. (\ref{prediction}). This result
states that in all space including the transition region the
temperature field is constant, $T({\bf r})=T$.
\subsection{The stress and the pressure}
The equilibrium average of the time rate of change of the momentum
density field is zero at equilibrium, this is $\langle iL{\bf g}_{\bf
r}\rangle^{[\lambda]}=0$ (as can be shown by integrating by parts
the Liouville operator and use of $LH^{[\lambda]}=0$). By taking the
equilibrium average of Eq. (\ref{lrlg}) we obtain then
\begin{align}
0&=-\nabla \langle \hat{\bf K}_{\bf r}\rangle^{[\lambda]}+ \langle {\bf f}_{\bf r}\rangle^{[\lambda]}
\end{align}
which, on account of Eq. (\ref{eqkin}) gives an explicit form for the force density
field
\begin{align}
\label{fandn}
\langle {\bf f}_{\bf r}\rangle^{[\lambda]}&=k_BT\nabla\langle n_{\bf r}\rangle^{[\lambda]}
\end{align}
In passing, we note that Eq. (\ref{fandn}) is valid for {\em any} Hamiltonian system:
notably, this intimate relation between the force density field and the density gradients is {\em independent}
on the fluid compressibility. It explains the essence of many algorithms \cite{FritschPRL108,Issa2014,kotsalis07}
designed to impose a flat density profile by adding an external force ``correction'' to the system (which,
according to Eq. (\ref{fandn}) has to ensure vanishing total force density field ${\bf f}_{\bf r}=0$).
Figure \ref{fig:ener} (middle panel) offers a numerical check of the relation (\ref{fandn}) in one of our \hadr systems
(in that case with ${\bf f}_{\bf r} \ne 0$).
Now, let us consider the equilibrium average of Eq. (\ref{lrlg2}) by introducing
\begin{align}
\boldsymbol{\Sigma}({\bf r})&\equiv\langle \hat{\boldsymbol{\Sigma}}_{\bf r}\rangle^{[\lambda]}
=k_BT n({\bf r})+ \boldsymbol{\Pi}({\bf r})
\nonumber\\
\boldsymbol{\Pi}({\bf r})&=\langle \hat{\boldsymbol{\Pi}}_{\bf r}\rangle^{[\lambda]}
\label{Sigma}
\end{align}
Here $\boldsymbol{\Sigma}({\bf r})$ is the average of the
Irwing-Kirkwood (IK) stress tensor in Eq. (\ref{Sigma0}), which is decomposed into its ideal
and interaction (or excess over ideal) parts. With the IK stress
tensor, Eq. (\ref{lrlg2}) gives
\begin{align}
\nabla \boldsymbol{\Sigma}({\bf r})=
k_BT\nabla n({\bf r})+\nabla \boldsymbol{\Pi}({\bf r})
&=-\frac{\delta F^{[\lambda]}}{\delta \lambda({\bf r})}\nabla\lambda({\bf r})
\label{crux}
\end{align}
Under equilibrium conditions, Eq. (\ref{crux}) just represents the
hydrostatic balance \cite{LandauFL} i.e. the response of the system's
equilibrium stress field to an external force. When the switching
field is sufficiently smooth, we expect from symmetry reasons that the
average of the interaction part of the stress tensor is isotropic
\begin{align}
\boldsymbol{\Pi}({\bf r})&=p^{\rm ex}({\bf r}){\bf 1}
\label{Piiso}
\end{align}
where we have introduced the excess (over ideal) part of the pressure. The total pressure
is defined as
\begin{align}
p({\bf r}) &\equiv p^{\rm id}({\bf r})+ p^{\rm ex}({\bf r})
\nonumber\\
p^{\rm id}({\bf r})&\equiv k_BTn({\bf r})
\nonumber\\
p^{\rm ex}({\bf r})&\equiv \frac{1}{3}{\rm Tr}\left[\boldsymbol{\Pi}({\bf r})\right]
\end{align}
Therefore, Eq. (\ref{crux})
takes the form
\begin{align}
\boldsymbol{\nabla }p({\bf r})=k_BT \boldsymbol{\nabla }n({\bf r})+ \boldsymbol{\nabla }p^{\rm ex}({\bf r})&=-\frac{\delta F^{[\lambda]}}{\delta \lambda({\bf r})}\nabla\lambda({\bf r})
\label{Gradp}
\end{align}
The two exact results (\ref{tcte}) and (\ref{crux}) give the local
thermodynamics of the system in terms of its equations of state. They
are one of the main important results of the present work.
\section{Translation invariance}
\label{sec:trans}
\subsection{The free energy}
A nice theorem about the free energy involves its behaviour under
translations. Assume that there are no external potential fields and
that the system is either infinite or has periodic boundary
conditions. We may perform in the definition (\ref{Flam}) the change
of variables ${\bf r}_i={\bf r}_i'+{\bf a}$ where ${\bf a}$ is an
arbitrary translation vector. Because all the potentials are
translational invariant, we arrive at the identity
\begin{align}
F_{[\lambda]}&= F_{[T_{\bf a}\lambda]}
\label{FFa}
\end{align}
where $T_{\bf a}$ is a translational operator that when applied to a function
gives
\begin{align}
T_{\bf a}\lambda({\bf r}) &=\lambda({\bf r}+{\bf a})
\label{Ta}
\end{align}
We may now take the derivative of both sides of Eq. (\ref{FFa}) with
respect to ${\bf a}$ and obtain
\begin{align}
0&=\frac{\partial F_{[T_a\lambda]}}{\partial {\bf a}}=\int d{\bf r}\frac{\delta F_{[T_{\bf a}\lambda]}}{\delta \lambda({\bf r})}\frac{\partial}{\partial {\bf a}}T_{\bf a}\lambda({\bf r})
\label{dfdl=00}
\end{align}
where the chain rule has been used. By using (\ref{Ta}) and evaluating
the result at ${\bf a}=0$ we obtain
\begin{align}
\int d{\bf r}\frac{\delta F_{[\lambda]}}{\delta \lambda({\bf r})}\nabla\lambda({\bf r})=0
\label{dfdl=0}
\end{align}
One consequence of the translation invariance of the free energy
(\ref{FFa}) is that the average total force on the system is zero. The
average total force is
\begin{align}
\langle {\bf F}\rangle^{[\lambda]}&=\frac{1}{Z[\lambda]}\int \frac{d^{3N}r}{\Lambda^{3N}}
\exp\{-\beta H_{[\lambda]}\}\sum_i^N\left(-\frac{\partial H_{[\lambda]}}{\partial {\bf r}_i}\right)
\nonumber\\
&=k_BT \frac{1}{Z}\int \frac{d^{3N}r}{\Lambda^{3N}}
\sum_i^N\frac{\partial }{\partial {\bf r}_i}\exp\{-\beta H_{[\lambda]}\}
\end{align}
We may again perform a translation of the origin of coordinates and produce the change
of variables ${\bf r}_i={\bf r}_i'+{\bf a}$ that becomes
\begin{align}
\langle {\bf F}\rangle^{[\lambda]}&=
k_BT \frac{1}{Z[\lambda]}\int \frac{d^{3N}r'}{\Lambda^{3N}}
\sum_i^N\frac{\partial }{\partial {\bf r}'_i}\exp\{-\beta H_{[T_{\bf a}\lambda]}\}
\nonumber\\
&=
k_BT \frac{1}{Z[\lambda]}\frac{\partial}{\partial {\bf a}}\int \frac{d^{3N}r'}{\Lambda^{3N}}
\exp\{-\beta H_{[T_{\bf a}\lambda]}\}
\nonumber\\
&=
k_BT \frac{1}{Z[T_{\bf a}\lambda]}\frac{\partial}{\partial {\bf a}}\int \frac{d^{3N}r'}{\Lambda^{3N}}
\exp\{-\beta H_{[T_{\bf a}\lambda]}\}
\nonumber\\
&=-
\frac{\partial}{\partial {\bf a}}F_{[T_{\bf a}\lambda]} =0
\end{align}
where the last identity follows from Eq. (\ref{dfdl=00}).
More generally, we have derived an important relation
between the derivative of the free energy functional and the total
force on the system,
\begin{equation}
\langle {\bf F}\rangle^{[\lambda]} = - \int d{\bf r}\frac{\delta F_{[\lambda]}}{\delta \lambda({\bf r})} \nabla\lambda({\bf r})
\label{netforce}
\end{equation}
which indicates that $-\nabla\lambda({\bf r}) \delta F_{[\lambda]}/\delta \lambda({\bf r})$
is the force density field induced by the jump in potential energy densities (``the drift force'' in Ref. \cite{Potestio:PRL:110})
and the free energy correction (see Eq. \ref{dfdl=01}).
However to reach a well-defined equilibrium state,
any prescription for computing the free
energy compensating term entering the free energy $F_{[\lambda]}$ has
to comply with Eq. (\ref{dfdl=0}). Otherwise, a net force (\ref{netforce})
will appear in the system.
In this sense,
the requirement (\ref{dfdl=0}) provides {\em global thermodynamic consistency}.
By integrating Eq. (\ref{crux}) over the system volume and
using Gauss theorem, leads to
\begin{equation}
\oint {\boldsymbol{\Sigma}}_{\bf r} \cdot {\bf n} d r^2 = - \int \nabla\lambda({\bf r})\frac{\delta F^{[\lambda]}}{\delta \lambda({\bf r})} d {\bf r}.
\label{gauss}
\end{equation}
Therefore in periodic systems (where by construction $\oint
\boldsymbol{\Sigma}_{\bf r} \cdot {\bf n} =0$) translational
invariance (\ref{dfdl=0}) and global thermodynamic consistency (in
particular, mechanical equilibrium) are trivially satisfied for {\em
any choice} of the free energy correction.
\subsection{Averages of local functions}
Consider a {\em local} function based on the CoM of the form
\begin{align}
A_{{\bf r}}(r,p)&=\sum_\mu^M A_\mu(r,p)\delta(\hat{\bf R}_\mu-{\bf r})
\end{align}
where $A_\mu(r,p)$ is traslationally invariant, so the effect of
changing ${\bf r}_i$ with ${\bf r}_i+{\bf a}$ for any vector ${\bf a}$
leaves $A_\mu$ invariant.
Examples of local functions are those defined in Eqs. (\ref{defu1u0}). In this case, we have the following identity
\begin{align}
\left\langle A_{\bf r}\right\rangle^{[T_{\bf a}\lambda]}&= \left \langle A_{{\bf r}+{\bf a}}\right\rangle^{[\lambda]}\label{atrans}
\end{align}
\begin{widetext}
as we can check explicitly
\begin{align}
\left\langle A_{\bf r}\right\rangle^{[T_{\bf a}\lambda]}&=
\frac{1}{Z[T_{\bf a}\lambda]}\int d^{3N}rd^{3N}pA_{\bf r}(r,p)
\nonumber\\
&\times\exp\left\{
-\beta\left[K+V^{\rm intra}+\sum_\mu^M\lambda(\hat{\bf R}_\mu+{\bf a})V^{1}_\mu
+\sum_\mu^M(1-\lambda(\hat{\bf R}_\mu+{\bf a}))V^0_\mu+\calf(\lambda(\hat{\bf R}_\mu+{\bf a}))\right]
\right\}
\nonumber\\
&=\left\langle A_{{\bf r}+{\bf a}}\right\rangle^{[\lambda]}
\label{check}
\end{align}
where we have performed a change of variables ${\bf r}_i\to{\bf r}_i-{\bf a}$ in the last identity.
By taking the derivative of Eq. (\ref{atrans}) with respect to ${\bf a}$ and
setting afterwards ${\bf a}=0$ we have
\begin{align}
\nabla \left\langle A_{{\bf r}}\right\rangle^{[\lambda]}
&= \int d{\bf r}'\nabla'\lambda({\bf r}')\frac{\delta }{{\delta \lambda({\bf
r}')}}\left\langle A_{{\bf r}}\right\rangle^{[\lambda]}
\label{na}
\end{align}
The functional derivative of the average is given by
\begin{align}
\frac{\delta }{{\delta \lambda({\bf
r}')}}\left\langle A_{{\bf r}}\right\rangle^{[\lambda]}
&=\beta\langle A_{\bf r}\rangle^{[\lambda]}
\left\langle\frac{\delta H_{[\lambda]}}{{\delta \lambda({\bf
r}')}}
\right\rangle^{[\lambda]}
-\beta
\left\langle A_{\bf r}\frac{\delta H_{[\lambda]}}{{\delta \lambda({\bf
r}')}}
\right\rangle^{[\lambda]}
=-\beta
\left\langle \delta A_{\bf r}\frac{\delta H_{[\lambda]}}{{\delta \lambda({\bf
r}')}}
\right\rangle^{[\lambda]}
\end{align}
where $\delta A_{\bf r}=A_{\bf r}-\langle A_{\bf r}\rangle^{[\lambda]}$.
By using (\ref{dhdl}) we obtain the exact result for local functions
\begin{align}
\nabla \left\langle A_{{\bf r}}\right\rangle^{[\lambda]} &=-\beta
\int d{\bf r}'\nabla'\lambda({\bf r}')
\left\langle \delta A_{\bf
r}(u^1_{{\bf r}'}-u^0_{{\bf r}'}+{\cal F}'(\lambda({\bf
r}'))n_{{\bf r}'}) \right\rangle^{[\lambda]}
\label{nablaAr}
\end{align}
This expression clearly shows that the inhomogeneities of any local
function along space will show up basically in the transition region
$0<\lambda<1$ for which $\nabla \lambda\neq0$ and are exclusively due
to the correlations of this local function with the functional
derivative of the Hamiltonian. For example, take the center of mass
density field $n_{\bf r}$ as the local function $A_{\bf r}$. The above
expression gives
\begin{align}
\nabla \left\langle n_{{\bf r}}\right\rangle^{[\lambda]} &=-\beta
\int d{\bf r}'\nabla'\lambda({\bf r}')
\left[\left\langle \delta n_{\bf
r}(u^1_{{\bf r}'}-u^0_{{\bf r}'})\right\rangle^{[\lambda]}
+
{\cal F}'(\lambda({\bf
r}'))\left\langle \delta n_{\bf
r}n_{{\bf r}'} \right\rangle^{[\lambda]}
\right]
\label{nablan}
\end{align}
This expression connects (linearly) the gradients of the density field
with the gradients of the switching function. It explains why there
should be molecular density variations in the region where the switching
function changes its value.
\end{widetext}
\section{The free energy compensation term
$\calf(\lambda)$ through an iterative route}
\label{sec:itera}
Up to now we have presented a number of exact results in Eqs. (\ref{tcte}),
(\ref{Gradp}), and (\ref{nablaAr}), that are valid for a general
Hamiltonian of the form (\ref{HA}). The particular functional form of
the free energy compensation term $\calf(\lambda)$ has not
yet been specified. We will now use these exact results in order to
fix the functional form of the free energy compensating term.
\subsection{Constant stress field}
The basic requirement that the free energy in the AA region coincides
with the free energy of the CG region, $F_{[1]}=F_{[0]}$ (i.e. that
the free energy does not depend on the actual value of $\lambda$) can
be generalized to the case that the parameter $\lambda$ is space
dependent. We require that {\em the actual free energy is independent
of the switching field $\lambda({\bf r})$}. This requirement is
mathematically expressed as the vanishing of the functional derivative
\begin{align}
\frac{ \delta F_{[\lambda]}}{\delta \lambda({\bf r})}=0
\label{df0}
\end{align}
The condition (\ref{df0}) will be referred to as the {\em local
thermodynamic consistency requirement of} \hadr. Note that the
requirement (\ref{df0}) ensures automatically the translational
invariance of the system expressed in Eq. (\ref{dfdl=00}). It also
ensures, through Eq. (\ref{Gradp}), that the stress field and, therefore the pressure, is
constant through space. In general, however, the density field will
not be constant and the system may experience differences between the
value of the density in the AA region and the GG region. Of course,
the variations of the density are compensated with the variations of
the excess pressure $p^{\rm ex}({\bf r})$ in order to have a constant
pressure field.
By using Eq. (\ref{dfdl=01}), Eq. (\ref{df0}) becomes
\begin{align}
0=\left\langle {u}^1_{\bf r}-{u}^0_{\bf r}\right\rangle^{[\lambda]}+
\calf'(\lambda({\bf r})) \left\langle {n}_{\bf r}\right\rangle^{[\lambda]}
\label{dfdl=01b}
\end{align}
This equation can be understood as a non-linear functional equation to
be solved for $\calf(\lambda)$ (where $\calf(\lambda)$ appears explicitly as well as implicitly in the
definition of the averages $\langle\cdots\rangle^{[\lambda]}$). An iterative
method to solve Eq. (\ref{dfdl=01b}) is given in Sec. \ref{sec:simulations}.
\subsection{Constant density field }
\label{sec:nconts}
The Hamiltonian (\ref{HA}), with $\calf(\lambda)$ obtained
from the condition that its free energy does not depend on the field
$\lambda({\bf r})$ (i.e. conditions (\ref{df0}) and (\ref{dfdl=01b})),
ensures that the pressure field is constant through the simulation
box. However, it does not ensure that the molecular mass density or
the molecular energy density are the same in the AA and CG regions. We
expect that, to the extend that the CG model is a good model in that
it reproduces correctly the molecular radial distribution function,
the density mismatch between AA and CG regions cannot be very large.
However, the CG potential is approximate and there may be situations in which keeping the molecular
density field through the system may be more important than keeping
the pressure field constant. In these situations, an alternative
definition of the term $\calf(\lambda({\bf R}_\mu))$ in the
Hamiltonian (\ref{HA}) is required. Eq. (\ref{crux}) suggests a route
to an alternative definition of $\calf(\lambda)$ that ensures a
constant density field. By setting $\nabla \langle \hat{n}_{\bf
r}\rangle^{[\lambda]}=0$ in Eq. (\ref{crux}) we obtain
\begin{align}
\nabla\lambda({\bf r})\frac{\delta F^{[\lambda]}}{\delta \lambda({\bf r})}
+\nabla \left\langle\hat{\boldsymbol{\Pi}}_{\bf r}\right\rangle^{[\lambda]}&=0
\label{crux0}
\end{align}
this is
\begin{align}
\nabla\lambda({\bf r})\left[
\langle\hat{u}^1_{\bf r}\rangle^{[\lambda]}-\langle\hat{u}^0_{\bf r}\rangle^{[\lambda]}+\calf'(\lambda({\bf r}))\langle\hat{n}_{\bf r}\rangle^{[\lambda]}\right]
+\nabla\langle\hat{\boldsymbol{\Pi}}_{\bf r}\rangle^{[\lambda]}&=0
\label{gibbs}\end{align}
This equation is a non-linear implicit equation for
$\calf'(\lambda({\bf r}))$ that may be computed iteratively in a
simulation because all terms, except $\calf'$ are explicitly
computable. This $\calf(\lambda)$ will, by construction, ensure that
$\nabla \langle \hat{n}_{\bf r}\rangle^{[\lambda]}=0$, but will not
satisfy, in general, the thermodynamic consistency property
(\ref{dfdl=01b}). The pressure field $\boldsymbol{\Sigma}({\bf r})$
will not be constant across the system and its gradient will be given
by
\begin{align}
\nabla \boldsymbol{\Sigma}({\bf r}) &=
-\nabla\lambda({\bf r})\frac{\delta F^{[\lambda]}}{\delta \lambda({\bf r})}
\label{nsn0}
\end{align}
where we have used (\ref{Sigma}) and (\ref{crux0}). Note that in
general, (\ref{crux0}) does not comply with the {\em global
thermodynamic consistency} requirement (\ref{dfdl=0}) that the free
energy (\ref{Flam}) is translationally invariant. However, as stated
[see Eq. (\ref{gauss})], such requirement is automatically fulfilled
in periodic systems, where global mechanical equilibrium is always
guaranteed.
\section{The free energy compensating term $\calf(\lambda)$ through local equilibrium}
\label{sec:lea}
In this section, we explore the simplifications that result in the
calculation of the free energy compensation term when the switching
field is sufficiently smooth in the length scale of the molecular
correlations. As formally justified in Appendix \ref{localeq}, in
this case, we may resort to a {\em local equilibrium approximation}
(LEA). The LEA essentially consists on assuming that the
average of any local microscopic quantity $\langle \hat{A}_{\bf
r}\rangle^{[\lambda]}$ obtained from the \hadr Hamiltonian
$H_{[\lambda]}$, is close to the average where the field $\lambda$ is
constant $\overline{\lambda}$ and equal to the value
$\overline{\lambda}=\lambda({\bf r})$ at the space point ${\bf r}$.
Each value of this function determines a {\em hybrid molecular model}.
The (canonical) average of such ``hybrid'' fluid (using
$H_{\overline{\lambda}}$) is denoted as $\langle \hat{A}
\rangle^{\overline{\lambda},n,T}$, where the prescribed values of $n$
and $T$ are indicated. If $\lambda({\bf r})$ is smooth enough, the
\hadr local average at ${\bf r}$ is close to the standard canonical
average of a fluid model with a constant $\overline{\lambda}=\lambda({\bf r})$
(see Eq. Appendix \ref{localeq}),
\begin{align}
\langle \hat{A}_{\bf r}\rangle^{[\lambda]} \approx
\langle \hat{A} \rangle^{\overline{\lambda}=\lambda({\bf r}),
\langle n_{\bf r} \rangle^{[\lambda]}, \langle T_{\bf r} \rangle^{[\lambda]}} \equiv
\langle \hat{A}_{\bf r}\rangle^{\lambda_{\bf r}}
\label{AAaprox}
\end{align}
where the last definition is introduced to alleviate the fully explicit
heavy notation of the local average.
In what follows we use the LEA expressed in Eq. (\ref{AAaprox}) with
two purposes. First, we derive a non-iterative route to find the free
energy correction ${\cal F}(\lambda)$. This non-iterative procedure
connects the \hadr formalism to the process used in
thermodynamic integration \cite{Kirkwood1935,Frenkel.book}, from which
the \hadr idea actually stems. Second, we use the LEA to explore the
relations between the thermodynamic variables along the transition
region for the different forms of the free energy corrections proposed
hereby and in previous papers
\cite{Potestio:PRL:110,Potestio:PRL:111}.
\subsection{Kirkwood route to constant stress field}
\label{kirk_pres}
When the switching field varies very smoothly, we may use the
approximation (\ref{AAaprox}) in Eq. (\ref{dfdl=01}) in order to
obtain a method that does not require an iterative procedure. Indeed, to first order
in gradients of $\lambda({\bf r})$ we have
\begin{align}
0=\left\langle {u}^1_{\bf r}-{u}^0_{\bf r}\right\rangle^{\lambda}+
\calf'(\lambda) \left\langle {n}_{\bf r}\right\rangle^{\lambda}
\label{dfdl=02a}
\end{align}
where the actual value of $\lambda$ is $\lambda({\bf r})$. According
to the LEA, this identity can be also understood in terms of averages
of hybrid fluids with constant $\lambda$. By integrating over space
and using the definitions (\ref{defu1u0}) we obtain
\begin{align}
0&=
\left\langle {U}^1-{U}^0\right\rangle^{\lambda}+ \calf'(\lambda)M
\label{dfdl=02}
\end{align}
where we have defined the inter-blob potential energy of the
microscopic and CG systems as
\begin{align}
{U}^1&=\sum_\mu^M V^1_\mu
\nonumber\\
{U}^0&=\sum_\mu^M V^{0}_\mu
\end{align}
By integrating with respect to $\lambda$, we may write Eq. (\ref{dfdl=02}) as
\begin{align}
\calf^K(\lambda) = - \frac{1}{M} \int_0^\lambda d\lambda'\left\langle\frac{\partial U}{\partial \lambda'}\right\rangle^{\lambda'} + {\rm C}
\label{K}
\end{align}
where we have defined the potential energy $U\equiv\lambda U^1 +
(1-\lambda) U^0$. For consistency with Eq. (\ref{Fm4}), the arbitrary
constant C should be set to fix $\calf^K(1) =0$ (i.e. the free energy
correction is zero in the atomistic domain). On the right hand side
of Eq. (\ref{K}) one recognizes the Kirkwood formula for standard
thermodynamic integration \cite{Kirkwood1935} which indicates that
$\calf^K(0)$ is the change in free energy over an alchemic
transformation of the interblob interaction from $U^1$ to $U^0$. This
is consistent with the interpretation given after Eq. (\ref{F0F1}).
Evaluation of the RHS of Eq. (\ref{K}) from a series of simulations at
fixed $\lambda$ offers a non-iterative protocol to the free energy
correction $\calf$.
Kirkwood calibration of $\calf$ relies however on the local
thermodynamic equilibrium [see (\ref{AAaprox})] as Eq. (\ref{K}) does
not ensure the thermodynamic consistency (\ref{df0}), except if the
switching function is smooth enough. Simulations presented in Sec.
\ref{sec:simulations} show that in practice Kirkwood non-iterative
approximation works quite well, at least for the test cases considered
here. This was also observed in previous works with different fluid
models \cite{Potestio:PRL:110,Potestio:PRL:111}, although a study of
the validity of Kirkwood TI as a function of the transition layer
length and the coupled fluid models was not considered. We will
perform such study in Sec. \ref{sec:simulations}.
\subsection{Kirkwood route to constant density field.}
\label{kirk_dens}
We now consider the local equilibrium approximation (LEA)
to find a non-iterative way to compute the free energy compensation term when
the target is to keep the density field constant across the simulation
box. The exact result in Eq. (\ref{crux}) can be written as
\begin{eqnarray}
k_BT \boldsymbol{\nabla }\langle \hat{n}_{\bf r}\rangle^{[\lambda]}+ \boldsymbol{\nabla }
\left\langle\hat{p}^{\rm ex}_{\bf r}\right\rangle^{[\lambda]}
+\frac{\delta F^{[\lambda]}}{\delta \lambda({\bf r})}\nabla\lambda({\bf r})&=&0
\nonumber\\
k_BT \boldsymbol{\nabla }\langle \hat{n}_{\bf r}\rangle^{[\lambda]}+ \boldsymbol{\nabla }
\left\langle\hat{p}^{\rm ex}_{\bf r}\right\rangle^{[\lambda]} +\left[\langle\hat{u}^1_{\bf r}-\hat{u}^0_{\bf
r}\rangle^{[\lambda]}+\calf'(\lambda)\langle\hat{n}_{\bf
r}\rangle^{[\lambda]} \right]\nabla\lambda({\bf r})&=&0
\label{Gradp2}
\end{eqnarray}
where the microscopic excess pressure is defined by
\begin{align}
\hat{p}^{\rm ex}_{\bf r} = \frac{1}{3}{\rm Tr}\left[\boldsymbol{\Pi}_{\bf r} \right]
\end{align}
We assume that $\langle \hat{n}_{\bf r}\rangle^{[\lambda]}=n$ is
constant and, therefore, the first term in Eq. (\ref{Gradp2})
vanishes. The second term, with the local equilibrium approximation
(\ref{AAaprox}), becomes
\begin{align}
\boldsymbol{\nabla }
\left\langle\hat{p}^{\rm ex}_{\bf r}\right\rangle^{[\lambda]} &\approx
\frac{d}{d\lambda}
\left. \langle \hat{p}^{\rm ex}_{\bf r}\rangle^{\lambda}\right|_{\lambda=\lambda({\bf r})}
\nabla \lambda({\bf r})
\end{align}
The term involving the difference between potential energy densities is,
under the local equilibrium approximation (\ref{AAaprox})
\begin{align}
\left \langle\hat{u}^1_{\bf r}-\hat{u}^0_{\bf r}\right\rangle^{[\lambda]}\approx
\left. \left \langle\hat{u}^1_{\bf r}-\hat{u}^0_{\bf r}\right\rangle^{\lambda}\right|_{\lambda=\lambda({\bf r})}
\end{align}
This
may be written as
a total derivative with respect to $\lambda$ as
\begin{align}
\left \langle\hat{u}^1_{\bf r}-\hat{u}^0_{\bf r}\right\rangle^{\lambda}
&=\frac{d}{d\lambda}\int_0^\lambda d\lambda'
\left \langle\hat{u}^1_{\bf r}-\hat{u}^0_{\bf r}\right\rangle^{\lambda'}
\end{align}
By collecting these last results, Eq. (\ref{Gradp2}) becomes
\begin{align}
&\left. \frac{d}{d\lambda} \left[
\langle \hat{p}^{\rm ex}_{\bf r}\rangle^{\lambda}
+\int_0^\lambda d\lambda'
\left \langle\hat{u}^1_{\bf r}-\hat{u}^0_{\bf r}\right\rangle^{\lambda'}
+\calf(\lambda) n \right]\right|_{\lambda=\lambda({\bf r})}
\nonumber\\
&\times\nabla\lambda({\bf r})=0
\end{align}
One way to ensure this identity and, therefore, a constant density
field through the system is by requiring
\begin{align}
\langle \hat{p}^{\rm ex}_{\bf r}\rangle^{\lambda}
+\int_0^\lambda d\lambda'
\left \langle\hat{u}^1_{\bf r}-\hat{u}^0_{\bf r}\right\rangle^{\lambda'}
+\calf(\lambda) n &= {\rm C}
\label{req1}
\end{align}
where C is a constant.
Because the averages are performed with a constant switching field, we
have translation invariance and we can get rid off the position
dependence by simply averaging (\ref{req1}) over the whole volume. This gives
\begin{align}
\langle \hat{P}^{\rm ex}\rangle^{\lambda}
+\frac{1}{V}\int_0^\lambda d\lambda'
\left \langle\hat{U}^1-\hat{U}^0\right\rangle^{\lambda'}
+\calf(\lambda) n &= {\rm C}
\end{align}
where
\begin{align}
\hat{P}^{\rm ex}&\equiv\frac{1}{V}\int d{\bf r}\hat{p}_{\bf r}^{\rm ex}
=\frac{1}{V}\frac{1}{6}\sum_{\mu\nu} \hat{\bf G}_{\mu\nu}\esc \hat{\bf R}_{\mu\nu}
\end{align}
where we have used (\ref{Piiso}) and (\ref{Pivir}). Therefore, the
non-iterative prescription for the free energy compensating term,
valid for smooth switching fields, that produces a constant density
field is
\begin{align}
\calf^K(\lambda) &= -\frac{1}{M}\int_0^\lambda d\lambda'
\left \langle\hat{U}^1-\hat{U}^0\right\rangle^{\lambda'}
-\frac{ \langle \hat{P}^{\rm ex}\rangle^{\lambda}}{n} +
{\rm C}
\label{calFcdens}
\end{align}
to be compared with the prescription (\ref{K}) that produces a
constant pressure field. Again, the constant C should be set to fix
$\calf(1)=0$. The non-iterative calibration of $\calf$ based on Eq.
(\ref{calFcdens}) involves a series of simulations of
constant-$\lambda$ fluids in the canonical ensemble at the target
density $n=M/V$ and temperature $T$. The first term in the RHS of
Eq. (\ref{calFcdens}) is then the difference in the Helmholtz
excess free energy (per particle) $f^{ex}(0)-f^{ex}(\lambda)$ between
the CG fluid model ($\lambda=0$) and a fluid model with fixed
$\lambda$. The free energy correction $\calf$ acts like an external
potential field in the system so the system's chemical potential is
\cite{LandauSP} $\mu = g(\lambda) + \calf(\lambda)$ where,
$g(\lambda)=f(\lambda) + p/n$ is the Gibbs free energy per particle,
containing ideal and excess parts $g=g^{id}(n) +g^{ex}$. At constant
density, the ideal part contribution of any thermodynamic function is
constant and Eq. (\ref{calFcdens}) can be written as,
\begin{align}
g(\lambda)+\calf(\lambda) = g(1)=\mu,
\label{mu0}
\end{align}
showing that the constant density \hadr
consistently provides a constant chemical potential $\mu$
over the system.
\section{Simulations}
\label{sec:simulations}
This section presents molecular dynamics (MD) simulations to
illustrate and validate the \hadr theoretical framework. Simulations
of the microcanonical ensemble of the \hadr Hamiltonian in
Eq. (\ref{H1}) were done in periodic boxes with dimensions
$L_x,L_y=L_z$. We have used the tetrahedral fluid model
\cite{praprotnik:224106,Matej_JCP07,rdb08,Potestio:PRL:110} which has become one of the benchmark
models for Adaptive Resolution. Each tetrahedral molecule contains
four atoms bonded by FENE potentials. Non-bonded interactions are
described by a purely repulsive Lennard-Jones potential (cutoff at
$r_{cut}=2^{1/6}\sigma$ (where $\sigma$ is the atomic LJ-diameter).
The coarse-grained potential used for $\lambda=0$ (CG domain)
corresponds to the Morse potential proposed in
Ref. \cite{praprotnik:224106,rdb08}
\begin{equation}
U_{cg}(r)= \gamma \left(1.0-\exp\left[-\kappa (r-r_0)\right]\right)^2
\end{equation}
The parameters $\gamma=0.105$, $\kappa=2.4$ and
$r_0=2.31$, were originally fitted so as to correctly reproduce the
molecular radial distribution function of the polyatomic fluid and its
pressure. In order to study the flexibility of H-AdResS to compensate for
free energy differences between the coarse-grained and atomistic model we
have tweaked the CG potential to consider two cases,
\begin{itemize}
\item Fitted CG: $\gamma=0.105$, $\kappa=2.4$ and $r_0=2.31$,
\item Non-fitted CG: $\gamma=0.305$, $\kappa=2.4$ and $r_0=2.31$
\end{itemize}
The Inverse Boltzmann procedure was used to set the fitted CG
potential for a molecular density $\rho_m =0.1 \sigma^{-3}$ (atomic
density $n=4\rho_m$) and temperature $T\simeq 1.0 \epsilon/K_B$. The
CG potential also ensures $p^0(n,T) = p^1(n,T)$.
We consider a simple \hadr set-up where the switching function only
depends on the $x$-coordinate, $\lambda=\lambda(x)$
and its gradient is directed in $x$-direction, $\nabla \lambda({\bf r})=\lambda'(x){\bf e}_x$.
The resolution function $\lambda(x)$ is $\lambda=1$ at the AA domain and $\lambda=0$ at the CG domain
while in the transition layer it varies like,
\begin{equation}
\label{lambda}
\lambda(x)=\cos^2\left[\frac{\pi}{2}\frac{x-x_1}{\lh}\right]
\end{equation}
with $\lh=|x_1-x_0|$ the width of the transition region, where $\lambda'\ne 0$.
Here $x_1=x(\lambda=1)$ is the position of the AA-HYB border
and $x_0$ the location of the $\lambda=0$ border.
\subsection{Basic equilibrium thermodynamics of \hadr}
The MD algorithm was implemented in single precision arithmetic using
a standard second order velocity-Verlet integrator and a Verlet list
for neighbours search. As shown in Fig. \ref{fig:ener} (top panel) the total
energy is conserved (up to about $0.1\%$ deviation) and the energy
drift over long runs is practically zero. Figure \ref{fig:ener} (middle panel)
illustrates a numerical cross-check of the interesting relation
(\ref{fandn}), that relates the force density with the density
gradient (in the figure, for a system without free energy correction).
\begin{figure}[t]
\includegraphics[scale=0.8]{e.eps}
\caption{Top panel: Energy of a H-Adress simulation. There is practically no
drift in total energy over long simulation runs (here $5\times
10^5 \tau$, with $\tau=\sigma \sqrt(m/\epsilon)$ the standard
Lennard-Jones time unit of the atomic potential). Middle panel: A numerical
cross-check of the relation (\ref{fandn}). Bottom panel: The temperature profile
over the system. Simulations were done at density
$n=0.4\sigma^{-3}$ with fitted CG-potentials.}
\label{fig:ener}
\end{figure}
Also, in Fig. \ref{fig:ener} (bottom panel) the temperature profiles obtained in
several type of \hadr simulations (with or without correction) is
presented. In all cases, thermal equilibrium is attained and ensures
a constant temperature profile over the simulation box. In
microcanonical simulations the temperature is not an input simulation
parameter so one should expect small variations in temperature upon
inclusion of some form of the free energy correction (see for instance
Fig. (\ref{fig:ener}). In fact, a modification of the
FEC term changes the overall Hamiltonian of the system and in general its second derivatives
(e.g. the heat capacity) determining the caloric equation of state.
For this reason, here we use a standard (canonical)
thermostat {\em while adjusting} the free energy compensation in the
iterative way.
\subsection{Iterative evaluation of the free energy correction}
The iterative evaluation of the free energy correction (FEC) is based
on the force balance in Eq. (\ref{crux}), where the free energy
derivative is given by Eq. (\ref{dfdl=01}). The virial pressure
gradient in Eq. (\ref{crux}) stems from the inter-blob
forces. Instead of evaluating its gradient, it is more efficient to
use Eq. (\ref{gmunu}). We assume that the field $\lambda({\bf r})$
changes only along the $x$ axis, i.e. $\lambda({\bf r})=\lambda(x)$
and that there is translation invariance along the $y,z$ axis due to
the periodic boundary conditions. This allows to average (\ref{crux})
with respect $y,z$. We introduce the following $x$ dependent fields
\begin{align}
g(x)&\equiv \langle \nabla \hat{\boldsymbol{\Pi}}_{\bf r}\cdot {\bf e}_x\rangle^{[\lambda]} =
\left\langle\sum_{\mu \nu} \delta(\hat{X}_\mu-x) \hat{\bf G}_{\mu\nu} \cdot {\bf e}_x \right \rangle^{[\lambda]}\nonumber\\
{u}^1(x)-{u}^0(x) &\equiv\left\langle\sum^M_\mu \left(V^1_\mu -V^{0}_\mu\right) \delta(\hat{X}_\mu-x)\right\rangle^{[\lambda]}
\nonumber\\
n(x)&\equiv\left\langle\sum^M_\mu\delta(\hat{X}_\mu-x)\right\rangle^{[\lambda]}
\label{defu1u0x}
\end{align}
The density field $a(x)$ of any microscopic quantity $A_{\mu}$ is
numerically evaluated by a binned Dirac delta: $\delta_h(r) =
\Theta_h(r)/V_{h}$ where $V_{h}$ is the volume of the bin and in 1D the
characteristic function is $\Theta_h(x)=1$ if $|x| \le h/2$ and zero
otherwise. As customary we assume ergodicity and use temporal
averages instead of ensemble averages
\begin{equation}
a(x) = \frac{1}{T_{\rm sample}} \int_{T_{\rm sample}} dt \sum_{\mu} A_{\mu}(t) \delta_{\Delta x} \left(x-x_{\mu}\right).
\end{equation}
The sampling time is $T_{\rm sample}$ and the volume of the bin is
$V_{\Delta x} = \Delta x L_y\, L_z$ with $L_{\alpha}$ the
system's size in $\alpha$ direction.
With the definitions (\ref{defu1u0x}), the mechanical equilibrium
equation Eq. (\ref{crux}) becomes in the 1D setting
\begin{align}
\ffec(x) =& \frac{u^1(x)-u^0(x)}{n(x)} \lambda^{\prime}(x)
\nonumber\\
&+ \frac{g(x)}{n(x)} -k_BT \frac{d \ln n(x)}{dx}
\label{gibbs2}
\end{align}
where we have introduced the ``compensation'' force
\begin{align}
\ffec(x)\equiv- \calf'(\lambda(x))\lambda^{\prime}(x)
\end{align}
As it is clear from Eq. (\ref{eqmot1}), this is the $x$ component of the force due
to the FEC acting on the atoms of the system when they have the $x$ coordinate.
Eq. (\ref{gibbs2}) is valid for \textit{any} form of
the FEC $\calf(\lambda)$ as it reflects the condition of mechanical
equilibrium. The prescription to have a constant pressure field
in all the system, i.e. Eq. (\ref{dfdl=01b}), becomes in the 1D setting
\begin{align}
\ffec(x)&= \frac{u^1(x)-u^0(x)}{n(x)} \lambda^{\prime}(x)
\label{1Dpconst}
\end{align}
while the condition of constant density field, Eq. (\ref{gibbs}), becomes
\begin{align}
\ffec(x)&= \frac{u^1(x)-u^0(x)}{n(x)} \lambda^{\prime}(x)
+ \frac{g(x)}{n(x)}
\label{1Dnconst}
\end{align}
Note the the fields $n(x),u^0(x),u^1(x),g(x)$ depend implicitly on
$\calf(\lambda)$ because they are given in terms of equilibrium
averages computed with a Hamiltonian that contains $\calf(\lambda)$.
Therefore, we need to solve (\ref{1Dpconst}) and (\ref{1Dnconst})
iteratively. The general structure of Eqs. (\ref{1Dpconst}),(\ref{1Dnconst}) is
\begin{align}
\ffec=\Phi(\ffec)
\end{align}
One way to solve this equation iteratively is
\begin{align}
F_{\mathrm{c}}^{n+1}=\Phi(F_{\mathrm{c}}^n)
\label{a1}
\end{align}
with some initial good guess $\ffec^0$. In the present case, the Kirkwood
estimate for $\calf(\lambda)$ is a good guess that allows to use
(\ref{a1}). If we do not have such a good initial estimate, we need to
change the atomic forces $\ffec(x)$ slowly, otherwise the abrupt
change in the forces on the atoms may lead to undesirable
perturbations such as heat production (here we use thermostats
{\em only} during the FEC calibration), density
waves (that in a periodic system take a long time to be adsorbed), or
even the system explosion. For this reason it is better to consider
the iterative protocol
\begin{align}
\ffec^{n+1}=\ffec^n+\alpha(\Phi(\ffec^n)-\ffec^n)
\label{relax_eq}
\end{align}
where $\alpha$ is sufficiently small. When convergence is reached $\ffec^{n+1}\approx \ffec^n$ implying
$\ffec^n\approx\Phi(\ffec^n)$.
Note that
$\alpha$ can be seen as the inverse of a relaxation time (the solution
ideally converging exponentially fast to the converged solution,
$\ffec^{n+1} = \ffec^n$). We update Eq. (\ref{relax_eq}) each
sampling interval $T_{sample}=N_s \Delta t$ (with $N_s \sim 10^3$ time
steps) and in such case
$\alpha=\widehat{\alpha}\,\delta_{Kr}\left[\mathtt{mod}(n,N_f);0\right]$,
where $\delta_{Kr}$ is the Kronecker delta, $n$ is the time step,
$\mathtt{mod}(n;m)$ is the modulus function and $\hat{\alpha}<1$.
The iterative solution of the constant pressure FEC equation (\ref{1Dpconst}) becomes now
\begin{align}
F_{\mathrm{c}}^{n+1}(x)&=F_{\mathrm{c}}^{n}(x)
\nonumber\\
&+\alpha\left(\left[ \frac{u^1(x)-u^0(x)}{n(x)}\right]^n \lambda^{\prime}(x)-F_{\mathrm{c}}^n(x)\right)
\label{1DpconstIter}
\end{align}
where the notation $[\cdots]^n$ means that all averages are computed with the force $F_{\mathrm{c}}^n(x)$ known at the $n$-th iteration.
The iterative solution of the constant density FEC equation
(\ref{1Dnconst}) requires a further step in order to have a faster
convergence rate. The idea is to first perform an iteration of the
type (\ref{relax_eq})
\begin{align}
F^*_{\mathrm{c}}&=F^{n}_{\mathrm{c}}(x)
+\alpha\left(\left[ \frac{u^1(x)-u^0(x)}{n(x)}\right]^n \lambda^{\prime}(x)\right.
\nonumber\\
&+\left.
\left[\frac{g(x)}{n(x)}\right]^n-F^n_{\mathrm{c}}(x)\right)
\label{F*}
\end{align}
Then, we iterate the equivalent condition $k_BT\nabla\ln n(x)=0$
\begin{align}
F_{\mathrm{c}}^{n+1}(x)&= F_{\mathrm{c}}^{n}(x)+\alpha k_BT\frac{d}{dx}\ln n(x)
\end{align}
which we further integrate over the hybrid layer to have
\begin{align}
\label{intcal}
\calf^{n+1}(0)&= \calf^n(0)+\alpha k_T\ln \frac{n(x_1)}{n(x_0)}
\end{align}
where we have introduced
\begin{align}
\label{intcal2}
\calf^n(0)&\equiv\int_{x_0}^{x_1}dxF^n_{\mathrm{c}}(x)
\end{align}
and finally correct the result (\ref{F*}) as
\begin{align}
F_{\mathrm{c}}^{n+1}(x)&=F_{\mathrm{c}}^*(x)\frac{\calf^{n+1}(0)}{\calf^{n}(0)}
\label{Fcorrect}
\end{align}
The step in Eq. (\ref{intcal}) involves the integral (\ref{intcal2}) over the transition layer
so it permits to substantially reduce the fluctuations of the (total) free energy jump estimation
[$\calf(0)$]. This fastens up the iterative evaluation of the compensation force $F_{\mathrm{c}}(x)$.
An analysis of the convergence rates is however left for future work.
\subsection{Fitted CG potentials}
\label{fittedcg}
\subsubsection{Kirkwood TI versus iterative evaluation of $\calf$:
the effect of hybrid layer width $\lh$.}
This section analyzes the dependence of $\calf(\lambda)$ on the width
$\lh$ of the transition layer. Results will be compared with the
Kirkwood thermodynamic integration $\calf^K(\lambda)$ whose value
$\calf^K(0)$ at $\lambda=0$ is the free energy difference between both
fluid models (CG and AA). Recall that by construction $\calf(1)=0$,
and that for fitted CG potentials, by definition of fitted, we have
that $\calf(0)=0$. At some $0<\lambda(x)<1$, the agreement between the
Kirkwood free energy $\calf^K(\lambda)$ and the iterative evaluation
of $\calf(\lambda)$ will indicate the validity of the local
equilibrium approximation introduced in Sec. \ref{sec:lea}.
For large enough CG and AA domains the value of $\calf(0)$
has to be independent on the width of the transition layer.
The optimal result would be $\calf(0)=\calf(0)^{K}$ for any $\lambda$,
(i.e. for any width $\lh$). Such result would allow the \hadr scheme
to act as a flexible and efficient tool for free energy differences
evaluation. Although we will not focus here on this important
thermodynamic aspect of \hadr, we will analyze the effect of $\lh$ on
$\calf$ by considering systems with fitted CG potentials
($\calf(0)=\calf(1)=0$) in constant pressure \hadr simulations. These
issues will be also considered later when analyzing constant density
\hadr under non-fitted potentials, $\calf(0)\ne 0$.
The convergence of $\calf'$ is particularly fast in constant pressure
simulations because it only involves averages of extensive quantities
(energies). To get enough statistics for $\calf'$ in each iteration,
$T_{\rm sample}$ can be chosen to be few molecular collision times.
We usually started the iterative FEC evaluation using
$\calf(\lambda)=0$ as starting seed which is certainly a benefit, as
it avoids the pre-evaluation of the Kirkwood free energy $\calf^K$ as
starting point for the iterative route. It has to be said that
Molecular Dynamics \hadr only requires the derivative of the FEC
$\calf'$ for time stepping. In this context, MD-\hadr
\cite{Potestio:PRL:110} offers a benefit over
Monte Carlo \hadr \cite{Potestio:PRL:111} because it permits to use a
force balance like Eq. (\ref{1Dpconst}) to iteratively evaluate/update
the FEC on-the-fly.
Fig. \ref{kirk} compares
the Kirkwood approximation to $\calf$ with the iterative solution of
Eq. (\ref{gibbs2}) in a case with $\lh=5\sigma$. For large enough
transition layers, molecular correlations effects lessen and we expect
$\calf'$ to approach to Kirkwood's value. To analyze how molecular
correlations affect $\calf'$ we have reduced the width of the hybrid
layer $\lh$ up to quite small values. Fig. \ref{kirk} presents
results for $\lh=2$, $2.5$ and $5 \sigma$, which are similar to the
molecules' diameters (about $2.5\sigma$). Remarkably, ${\cal F'}$
becomes quite close the Kirkwood free energy as soon as $\lh$ is
larger than about twice the molecular cutoff radius. Maybe not
unexpectedly, deviations between the iterative $\calf$ and $\calf^K$
(Kirkwood) increase around $\lambda=0$ and $\lambda=1$. Despite
differences in $\calf(\lambda)$, it is important to stress that for
any choice of $\lh$ (see Fig. \ref{kirk}b) the iterative evaluation of
$\calf$ correctly predicts $\calf(0)=\calf(1)$. We shall come back to this
later in the case of non-fitted potentials.
\begin{figure}[t]
\includegraphics[scale=0.6]{FEC_pressure_FITTED.eps}
\caption{The derivative $\calf'$ (top) and FEC $\calf(\lambda)$
(bottom) between the atomistic tetrahedral fluid and the fitted CG
model as a function of $\lambda$ in constant pressure simulations.
Comparison is made between the Kirkwood TI (\ref{K}) and the
iterative solution of Eq. (\ref{gibbs2}) for several transition
layer widths $\lh$.}
\label{kirk}
\end{figure}
Fig. \ref{lhyb} illustrates the effect of reducing $\lh$ in the
density and pressure profiles in \hadr simulations with constant
pressure. An interesting observation is that the jump of
not-compensated quantities over the transition layer (here density)
does not significantly increase as $\lh$ is made shorter. It is
important to notice that in a closed system, any mass difference in
the transition layer (which is a lower density region in
Fig. \ref{lhyb}) induces finite size effects. The mass excluded from
the transition domain is transferred to the CG and AA domains
(according to their local chemical potential) so the density in both
domains will increase over the mean value $\bar{n}=M/V$ (which is
indicated with a dashed line in Fig. \ref{lhyb}a). Paradoxically,
for this reason the density profile using $\lh=2.5\sigma$ is closer to
$\bar{n}$ than the profile using from $\lh=5.0\sigma$ (see Fig.
\ref{lhyb}a). This mismatch in the bulk densities is reflected in the
total pressure, whose (constant) value slightly depends on $\lh$ (see
Fig. \ref{lhyb}b).
\begin{figure}[t]
\includegraphics[scale=0.35]{LHYB_profiles.eps}
\caption{
The effect of reducing $\lh$ in the density and pressure profiles.
Results corresponds to fitted CG potential at constant \hadr pressure.}
\label{lhyb}
\end{figure}
\subsubsection{Other finite box effects in closed systems}
Fig. \ref{fitcg} shows the density and pressure profiles for
$\lh=5\sigma$ in the case of fitted CG potentials. Comparison is made
between simulations with $\calf$ given by the pressure correction Eq.
(\ref{gibbs2}) and with $\calf=0$. Some conclusions can be extracted.
First the non-compensated version
presents a larger density jump over the transition regime, when
compared with the pressure compensated \hadr. The overall density
mismatch across the transition region is slightly larger in the
non-compensated \hadr, although it
is not a large difference neither. Second, in closed boxes (here
periodic) a rarefied transition region induces finite size effects on
the bulk densities which become larger than $\bar{n}=M/V$. The effect
is larger for $\calf=0$, although this effect is observed in both
simulations. This brings about consequences in the kinetic and virial
pressure profiles, shown in Fig. \ref{fitcg}b. Notably, the kinetic
pressure $p^{id}=\langle k_x\rangle$ is equal to $k_B T\langle n_x
\rangle$ (see Eq. \ref{tcte}) so any mismatch in density is reproduced
in $p^{id}$. The total pressure $p(x)=p^{id}(x)+p^{ex}(x)$ is robustly
fixed to a constant value $p(x)=P$ by the FEC. Consequently $p^{ex}$
compensates any variation in $p^{id}$ across the transition layer.
\begin{figure}[t]
\includegraphics[scale=0.4]{FITCG_dens_pres.eps}
\caption{(a) Density profile obtained from \hadr with fitted CG potentials
with constant pressure FEC and without FEC.
(b) The kinetic $p^{id}$ and virial $p^{ex}$ contributions to the total
pressure $p=p^{id}+p^{ex}$ in the pressure compensated \hadr with fitted CG potential. The mean pressure is $P=(1/L) \int_0^L p(x) dx$.
\label{fitcg}
}
\end{figure}
\subsection{Non-fitted CG potentials}
We now explore one of the main benefits of \hadr which is the
possibility of working with non-fitted CG potentials. This benefit is
not only to alleviate the time consuming and computational effort
related to pre-evaluation of CG potentials. In fact, fitting the CG
potential is a good practice as we have already seen that it minimizes
the mismatch in non-fitted thermodynamic variables. The benefits arise
from the possibility of performing simulations involving thermodynamic
{\em processes}, which involve changes in the global environmental
variables (temperature, pressure, chemical potential). In these cases
\hadr permits to work with a single CG model whose $\calf$ is
self-adapted over the whole process to keep the desired global
constraint (pressure, density, etc). In this sense \hadr offers an
alternative to the (probably more involved) problem of {\em potential
transferability}. Other benefits to be considered are the evaluation
of free energies differences in systems involving large solute
molecules. For these applications the estimation of the {\em total}
free energy difference between (CG and AA) models should be {\em
independent on the choice of the hybrid layer} and should coincide
with the Kirkwood thermodynamic value. On the other hand we expect
that the iterative evaluation of $\calf'$ will reduce or suppress the
oscillations in the density (or pressure) profiles around the
transition layer. As stated around Eq. (\ref{nablan}), these are due
to molecular correlations and have been reported in Kirkwood based
pre-evaluated FEC corrections (see
e.g. \cite{Potestio:PRL:110,Potestio:PRL:111}).
We start by presenting the free energy differences, pressure and
density profiles obtained for the three cases considered (constant
pressure and constant density FEC and no FEC) of a tetrahedral fluid
facing a non-fitted CG fluid. These results are shown in Fig.
\ref{FEC_NF} (FEC) and Fig. \ref{fig:nfp} (pressure and
densities). Note that in this case the Kirkwood free energy $\calf^K$
is practically equal to the constant pressure FEC correction,
reflecting again the strong connection of \hadr with standard
statistical mechanics. We will in fact hereafter focus on the
constant density FEC and on its iterative evaluation. Constant
density results of Fig. \ref{FEC_NF} and \ref{fig:nfp}, obtained with
the Kirkwood route $\calf^K$, reveal a relatively large free energy
difference between both fluids, of about $\calf(0) \simeq 2.7\,k_BT$
{\em per molecule}. Under no-FEC contribution, this leads to
substantial deviations in density and pressure across the simulation
box as reflected in Fig. \ref{fig:nfp}.
\begin{figure}[t]
\includegraphics[scale=0.35]{DFREE_nonFitted.eps}
\caption{(Top) The FEC $\calf$ evaluated from Kirkwood TI for constant
pressure and constant density. (Bottom) Derivatives of Kirkwood free
energies. In the constant pressure case, the \hadr FEC derivative
$\calf'$ is compared with Kirkwood's result. In this case,
the total Helmholtz free energy difference (Kirkwood) is
$\calf^K(0)=0.85(2)$ while the iterative \hadr provides
$\calf(0)=0.86(4)$.}
\label{FEC_NF}
\end{figure}
\begin{widetext}
\begin{figure}[t]
\includegraphics[scale=0.6]{dens_pres_tetralow_nf.eps}
\caption{The density (left) and pressure (right) obtained using
\hadr in simulations of tetrahedral molecules with the non-fitted
CG potential. The corresponding FEC $\calf$ is shown in
Fig. \ref{FEC_NF}}
\label{fig:nfp}
\end{figure}
\end{widetext}
\begin{figure}[ht!]
\includegraphics[scale=0.7]{Constant_density_NF.eps}
\caption{Free energies (top) and density profiles (bottom) for
constant density \hadr simulations of non-fitted CG
potentials. The Kirkwood free energy $\calf^K(\lambda)$ is
compared with the relaxation algorithm of Eqs.
(\ref{F*})-(\ref{Fcorrect}) for the FEC. Kirkwood total free energy jump is
$\calf^{K}(0)=2.67(0)$ and compares quite well with the \hadr
iterative result $\calf(0)=2.69(7)$. }
\label{fig:nfd}
\end{figure}
\begin{figure}[hb!]
\includegraphics[scale=0.3]{Balance_forces_w_Kirkwood_right.eps}
\caption{Details of the force balance of Eq. (\ref{1Dnconst}) at one
of the hybrid layers of a constant density \hadr simulation. The
iterative evaluation of $\calf'$ was performed with the algorithm
of Eqs. (\ref{F*})-(\ref{Fcorrect})
using $\Delta t=0.005$, $N_f=5000$ and $\hat{\alpha}=0.01$
and $\Phi=\lambda^{\prime}(x) \left(u^1(x)-u^0(x)\right)/n(x) + g(x)/n(x)$ [see Eq. (\ref{1Dnconst})].}
\label{fig:nfd_fb}
\end{figure}
\subsubsection{Iterative constant density FEC}
Fig. \ref{fig:nfd} compares the results for $\calf'$ and $\calf$
using the iterative evaluation in Eqs. (\ref{F*})-(\ref{Fcorrect})
and the Kirkwood TI in Eq. (\ref{calFcdens}) for constant density
field. The first thing to highlight from Fig. \ref{fig:nfd}
(top,half) is that although $\calf'(\lambda)$ (and its integral
$\calf(\lambda)$) differ substantially, the {\em overall} \hadr free
energy difference $\calf(0)$ results to be equal to the Kirkwood TI
value. For the reasons explained before, this is an important
result. Second, the density profile resulting from the iterative
protocol are not completely flat, although the oscillations
deviating from the mean density are softer and smaller than those
obtained from Kirkwood $\calf^K$ (maximum density deviations are
$2\%$ while about $5\%$ for Kirkwood). To understand the origin of
the density differences resulting from the iterative protocol
(\ref{F*})-(\ref{Fcorrect}) we plot in Fig. \ref{fig:nfd_fb} the
terms involved in the force balance over the $x$-direction. The
system's average force per molecule [RHS of Eq.
\ref{1Dnconst})] is compared with the imposed compensation force
$\ffec$. Density variations along $x$ arise with any difference
between both terms; from Eq. (\ref{gibbs2}), such difference is
precisely $k_BT d\ln n/dx$ and for clarity it has been amplified by
a factor 10 in Fig. \ref{fig:nfd_fb}. Indeed, $\ffec=0$ inside
the atomistic domain but due to the small width of the transition
layer and the sharp decay to zero of $\ffec(x)$ (particularly near
$\lambda=1$, indicated with an arrow in Fig. \ref{fig:nfd_fb}) the fluid is compressed and
creates density oscillations. It seems reasonable that the density
oscillations are larger where the difference in compressibility is
(i.e. near the atomistic border, $\lambda=1$). Fig.
\ref{fig:nfd_fb} shows that the transition of $\ffec$ to zero is
softer at $\lambda=0$, where the density profile is also softer.
These observations indicate two things: first, that density
variations should eventually decrease with increasing $\lh$ (by
allowing smaller values of $|d\ffec/dx|$ within the transition
layer) and second, that there might also be an optimal shape of
$\lambda(x)$. A study of these issues is however left for future
work.
\section{Conclusions}
\label{sec:conclu}
This work presented the statistical mechanics foundations of \hadr
\cite{Potestio:PRL:110,Potestio:PRL:111}.
Because the method is based on a Hamiltonian, the standard techniques
of Statistical Mechanics allow one to obtain a wealth of information
about the thermodynamics of AA and CG models. The Hamiltonian in
\hadr is an interpolation of the actual microscopic potential with a
CG representation of the system in terms of blobs. In this way, when a
blob moves from the AA region to the CG region its interactions change
accordingly. We have shown why and how \hadr can be adapted to
``connect'' two different fluid models (here the atomistic and the
coarse-grained models) by keeping both to coexist in the same fixed
ensemble (for instance, same density or same pressure) over the same
simulation box. The work required to do that is precisely the free
energy compensation $\calf$ which is the central ingredient of \hadr.
We have proved that $\calf(\lambda)$ is close to the free
energy difference obtained from Kirkwood thermodynamic integration
$\calf^K(\lambda)$ and that both energies are equal in the limit
of local thermodynamic equilibrium (in practice, wide enough
transition layers). We have developed schemes to iteratively evaluate
the free energy correction under either constant pressure or constant
density simulations. This iterative route has several benefits. The
first is a practical one, because it avoids the extra burden of
implementing Kirkwood thermodynamic integration each time a FEC needs
to be evaluated. Moreover, iterative evaluation of $\calf$ will permit
to self-adapt the FEC under a (slow enough) thermodynamic process.
It is important to stress that the {\em overall} free energy
jump in \hadr is a thermodynamic quantity which
does not depend on the shape or width of the
transition layer.
This is confirmed by simulation results
which agree withing error bars with the Kirkwood TI
free energy evaluation and indeed explains
the good performance of Kirkwood TI
approximations to $\calf$ used in Refs.
\cite{Potestio:PRL:110,Potestio:PRL:111}.
The limits and
potentiality of \hadr as a flexible, fast and self-adaptive free
energy estimator will surely deserve further studies
on denser and more disparate systems.
The emphasis in the present paper has been on \textit{equilibrium}
Statistical Mechanics. In order to look at problems in which
\textit{dynamics} is of importance, it is necessary to include the
possibility in the algorithm of interpolating the full CG dynamics.
In addition to the CG potential of interaction, the full CG dynamics
requires the presence of friction and stochastic forces in order to
fully account for eliminated degrees of freedom in the CG region
\cite{Hijon2009}. As it is well-known, the equilibrium properties
should not be affected by the presence of these additional forces that
are, however, crucial in non-equilibrium or dynamic situations. This
further development is left for future work.
\section{Acknowledgments}
We acknowledge the KAVLI Institute in Santa Barbara where this work
was initiated for its hospitality and support. MINECO provided
support through projects FIS2010-22047-C05-01 and
FIS2010-22047-C05-03. Comunidad Aut\'onoma de Madrid has financially
supported this work through the project MODELICO.
\begin{widetext}
\section{Appendix: Local equilibrium in the transition layer}
\label{localeq}
The exact result (\ref{nablaAr}) leads to an interesting result when
the typical length of variation of $\lambda({\bf r})$ is much larger
than the typical length of decay of the correlations. In this case,
because in the length scale in which the correlation decays, the field
$\lambda({\bf r}')$ hardly changes, we may approximate (\ref{nablaAr})
by taking $\nabla \lambda({\bf r}')\approx\nabla\lambda({\bf r})$
outside the integral as follows
\begin{align}
\nabla \left\langle A_{{\bf r}}\right\rangle^{[\lambda]} &\approx-\beta\nabla\lambda({\bf r})
\int d{\bf r}'\left\langle \delta A_{\bf
r}(u^1_{{\bf r}'}-u^0_{{\bf r}'}+{\cal F}'(\lambda({\bf
r}'))n_{{\bf r}'}) \right\rangle^{[\lambda]} \label{nablaArloc}
\end{align}
This approximation is equivalent to
set, in Eq. (\ref{na})
\begin{align}
\nabla \left\langle A_{{\bf r}}\right\rangle^{[\lambda]}
&\approx \nabla\lambda({\bf r}) \int d{\bf r}'\frac{\delta }{{\delta \lambda({\bf
r}')}}\left\langle A_{{\bf r}}\right\rangle^{[\lambda]}
\label{na2}
\end{align}
Now, let us consider the average of the local function $\left\langle
A_{{\bf r}}\right\rangle^{[\lambda]}$, when $\lambda({\bf r})$
changes smoothly. Consider the following rewriting of the Hamiltonian
\begin{align}
H_{[\lambda]}(r,p)&=H_{\lambda({\bf r})}+\delta H_{[\lambda]}
\end{align}
where we have added and subtracted a $\lambda({\bf r})$ term by defining
\begin{align}
H_{\lambda({\bf r})}&\equiv\sum_i\frac{{\bf p}_i^2}{2m_i}
+\sum_\mu^M V^{\rm intra}_\mu(r)+
\lambda({\bf r})\sum_{\mu}^MV^{1}_{\mu}(r)
+(1-\lambda({\bf r}))\sum_{\mu}^M V^{0}_{\mu}(R)
+\sum_\mu^M\calf (\lambda({\bf r}))
\nonumber\\
\delta H_{[\lambda]}&\equiv
\sum_{\mu}^M(\lambda(\hat{\bf R}_\mu)-\lambda({\bf r}))V^{1}_{\mu}(r)
-\sum_{\mu}^M(\lambda(\hat{\bf R}_\mu)-\lambda({\bf r}))V^{0}_{\mu}(R)
+\sum_\mu^M\calf(\lambda(\hat{\bf R}_\mu))-\sum_\mu^M\cal F(\lambda({\bf r}))
\nonumber\\
&=
\int d{\bf r}'(\lambda({\bf r}')-\lambda({\bf r}))u^{1}_{{\bf r}'}(r)
-\int d{\bf r}'(\lambda({\bf r}')-\lambda({\bf r}))u^{0}_{{\bf r}'}(R)
+\int d{\bf r}'(\cal F(\lambda({\bf r}'))-\cal F(\lambda({\bf r})))
\hat{n}_{{\bf r}'}(r)
\label{69}
\end{align}
Clearly, $H_{\lambda({\bf r})}$ is the Hamiltonian of a constant switching field where the value
of the constant is picked to be the local value $\lambda({\bf r})$.
We can now consider the average of a local function of the form
\begin{align}
\langle \hat{A}_{\bf r}\rangle^{[\lambda]}&=
\int dz\frac{1}{Z[\lambda]}\exp\{-\beta H_{[\lambda]}\}\sum_\mu^M A_\mu\delta({\bf r}-{\bf R}_\mu)
\end{align}
By expanding the exponential with respect to $\delta H_{[\lambda]}$ we have
\begin{align}
\langle \hat{A}_{\bf r}\rangle^{[\lambda]}&=\langle \hat{A}_{\bf r}\rangle^{\lambda=\lambda({\bf r})}
+\langle \delta H_{[\lambda]}\delta\hat{A}_{\bf r}\rangle^{\lambda=\lambda({\bf r})}+\cdots
\end{align}
By using the definition (\ref{69}), we have
\begin{align}
\langle \delta H_{[\lambda]}\delta\hat{A}_{\bf r}\rangle^{\lambda=\lambda({\bf r})}&=
\int d{\bf r}'(\lambda({\bf r}')-\lambda({\bf r})) \langle u^{1}_{{\bf r}'}(r)\delta\hat{A}_{\bf r}\rangle^{\lambda=\lambda({\bf r})}
-\int d{\bf r}'(\lambda({\bf r}')-\lambda({\bf r}))\langle u^{0}_{{\bf r}'}(R)\delta\hat{A}_{\bf r}\rangle^{\lambda=\lambda({\bf r})}
\nonumber\\
&+\int d{\bf r}'(\calf(\lambda({\bf r}'))-\calf(\lambda({\bf r})))
\langle\hat{n}_{{\bf r}'}(r)\delta\hat{A}_{\bf r}\rangle^{\lambda=\lambda({\bf r})}
\end{align}
\end{widetext}
It is apparent that if the switching field does not changes much on the
length scale of decay of the correlations, all the above contributions may be neglected
and we have
\begin{align}
\langle \hat{A}_{\bf r}\rangle^{[\lambda]} \approx \langle
\hat{A}_{\bf r}\rangle^{\lambda=\lambda({\bf r})}
\label{AAaproxApp}
\end{align}
This is a very natural result that tells that when the switching field
does not vary appreciably in the length scale of the molecular
correlations, the average of a local function in the spatially varying
switching field is very well approximated with the average at a
constant value of the switching field with the local value at the
point ${\bf r}$ that we are considering. By using this approximation
in Eq. (\ref{na2}), we obtain finally
\begin{align}
\nabla \langle \hat{A}_{\bf r}\rangle^{[\lambda]}
&\approx \nabla\lambda({\bf r}) \int d{\bf r}'\frac{\delta }{{\delta \lambda({\bf
r}')}}\left\langle A_{{\bf r}}\right\rangle^{\lambda=\lambda({\bf r})}
\nonumber\\
&= \nabla\lambda({\bf r})
\left.\frac{d}{d\lambda}\left\langle A_{{\bf r}}\right\rangle^{\lambda}\right|_{\lambda=\lambda({\bf r})}
\int d{\bf r}'\frac{\delta \lambda({\bf r})}{{\delta \lambda({\bf r}')}}
\nonumber\\
&=
\frac{d}{d\lambda}
\left. \langle \hat{A}_{\bf r}\rangle^{\lambda}\right|_{\lambda=\lambda({\bf r})}
\nabla \lambda({\bf r})
\label{le1}
\end{align}
This expression allows one to express gradients of local functions as
simply proportional to the gradients of the switching function
whenever the switching function changes smoothly on the length scale
of correlations of the CoM variables. Eq. (\ref{le1}) could be very
roughly interpreted as a sort of ``chain rule'' where space
derivatives are expressed in terms of derivatives with respect to the
switching field. The results (\ref{AAaprox}) and (\ref{le1}) will be
referred as the local equilibrium approximation for the averages and
its gradients.
\begin{widetext}
\section{Appendix: The force ${\bf F}_\mu$}
\label{Ap:force}
In this appendix we compute explicitly the force
\begin{align}
\hat{\bf F}_\mu&
=-\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}
\left[\sum_\nu V^{\rm intra}_\nu(r)+
\sum_{\nu}\lambda(\hat{\bf R}_\nu)V^{1}_{\nu}(r)
+\sum_{\nu}(1-\lambda(\hat{\bf R}_\nu))V^{0}_{\nu}(R)
+\sum_\nu\calf(\lambda(\hat{\bf R}_\nu))\right]
\end{align}
Consider the intra potential energy of molecule $\nu$ which is defined as
\begin{align}
V^{\rm intra}_\nu(r) &=\frac{1}{2}\sum_{i'j'}\delta_\nu(i')\delta_\nu(j')
\phi^{\rm intra}(r_{i'j'})
\end{align}
where $\phi^{\rm intra}(r_{i'j'})$ is the pair potential of particles
$i',j'$ due to intramolecular interactions. Then
\begin{align}
-\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}
\sum_\nu V^{\rm intra}_\nu(r)&=
-\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}\sum_\nu
\frac{1}{2}\sum_{i'j'}\delta_\nu(i')\delta_\nu(j')
\phi^{\rm intra}(r_{i'j'})
\nonumber\\
&=
-\sum_i\delta_\mu(i) \sum_\nu
\frac{1}{2}\sum_{i'j'}\delta_\nu(i')\delta_\nu(j')
\frac{\partial}{\partial {\bf r}_i}\phi^{\rm intra}(r_{i'j'})
\nonumber\\
&=
\sum_i\delta_\mu(i) \sum_\nu
\frac{1}{2}\sum_{i'j'}\delta_\nu(i')\delta_\nu(j')
f^{\rm intra}(r_{i'j'}){\bf e}_{i'j'}(\delta_{ii'}-\delta_{ij'})
\nonumber\\
&=
\sum_i\delta_\mu(i) \sum_\nu
\sum_{i'j'}\delta_\nu(i')\delta_\nu(j')
f^{\rm intra}(r_{i'j'}){\bf e}_{i'j'}\delta_{ii'}
\nonumber\\
&=
\sum_i\delta_\mu(i) \sum_\nu
\sum_{j'}\delta_\nu(i)\delta_\nu(j')
f^{\rm intra}(r_{ij'}){\bf e}_{ij'}
\nonumber\\
&=
\sum_i \sum_\nu
\sum_{j'}\delta_{\mu\nu}\delta_\nu(i)\delta_\nu(j')
f^{\rm intra}(r_{ij'}){\bf e}_{ij'}
\nonumber\\
&=
\sum_{ij'}\delta_\mu(i)\delta_\mu(j')
f^{\rm intra}(r_{ij'}){\bf e}_{ij'}=0
\end{align}
because ${\bf e}_{ij}=-{\bf e}_{ji}$ and the indices are dummy.
Indeed the total force on the molecule due to internal forces vanishes.
Consider now the term
\begin{align}
-\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}\lambda(\hat{\bf R}_\nu)
&=-\sum_i\delta_\mu(i)\nabla \lambda(\hat{\bf R}_\nu) \frac{\partial}{\partial {\bf r}_i}
\sum_{i'}\delta_\nu(i')\frac{m_{i'}}{m_\nu}{\bf r}_{i'}
\nonumber\\
&=-\sum_i\delta_\mu(i)\nabla \lambda(\hat{\bf R}_\nu)
\sum_{i'}\delta_\nu(i')\frac{m_{i'}}{m_\nu}\delta_{ii'}
=-\nabla \lambda(\hat{\bf R}_\nu)
\delta_{\mu\nu}
\end{align}
Next, the term
\begin{align}
-\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}
V^{1}_{\nu}(r)
&=-\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}
\frac{1}{2}\sum_{\nu'\neq\nu}\sum_{i'j'}\delta_{\nu}(i')\delta_{\nu'}(j')\phi^{\rm inter}(r_{i'j'})
\nonumber\\
&=
\frac{1}{2}\sum_{\nu'\neq\nu}\sum_{i i'j'}\delta_\mu(i)\delta_{\nu}(i')\delta_{\nu'}(j')
{\bf F}^{1}_{i'j'}(\delta_{ii'}-\delta_{ij'})
\end{align}
where we have introduced the force ${\bf F}^{1}_{i'j'}$ that atom $j'$ exerts on atom $i'$. Therefore
\begin{align}
\nonumber\\
-\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}
V^{\rm inter}_{\nu}(r)
&=
\frac{1}{2}\sum_{\nu'\neq\nu}\sum_{i j'}\delta_\mu(i)\delta_{\nu}(i)\delta_{\nu'}(j')
{\bf F}^{1}_{ij'}
-
\frac{1}{2}\sum_{\nu'\neq\nu}\sum_{i i'}\delta_\mu(i)\delta_{\nu}(i')\delta_{\nu'}(i)
{\bf F}^{1}_{i'i}
\nonumber\\
&=
\delta_{\mu\nu}\frac{1}{2}\sum_{\nu'\neq\nu}\sum_{i j}\delta_\mu(i)\delta_{\nu'}(j)
{\bf F}^{1}_{ij}
-
\frac{1}{2}\sum_{\nu'\neq\nu}\delta_{\mu\nu'}\sum_{i j}\delta_\mu(i)\delta_{\nu}(j)
{\bf F}^{1}_{ji}
\nonumber\\
&=
\delta_{\mu\nu}\frac{1}{2}\sum_{\nu'\neq\nu}{\bf F}^{1}_{\mu\nu'}
+
\frac{1}{2}\sum_{\nu'\neq\nu}\delta_{\mu\nu'}{\bf F}^{1}_{\mu\nu}
\nonumber\\
&=\sum_{\nu'\neq\nu}{\bf F}^{1}_{\nu\nu'}(R_{\nu\nu'})
\frac{1}{2}\left[\delta_{\mu\nu}-\delta_{\mu\nu'}\right]
\end{align}
where we have introduced the force that molecule $\mu $ exerts on molecule $\nu$ as
\begin{align}
{\bf F}^{1}_{\mu\nu}&\equiv
\sum_{i j}\delta_\mu(i)\delta_{\nu}(j)
{\bf F}^{1}_{ij}
\end{align}
Next, the term
\begin{align}
-\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}
V^{0}_{\nu}(R)&=
-\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}
\frac{1}{2}\sum_{\nu'}V^{0}_{\nu\nu'}(R)
\nonumber\\
&=
\frac{1}{2}\sum_{\nu'}F^{\rm cm}_{\nu\nu'}(R_{\nu\nu'})
\sum_i\delta_\mu(i) \frac{\partial{R}_{\nu\nu'}}{\partial {\bf r}_i}
\nonumber\\
&=
\frac{1}{2}\sum_{\nu'}F^{\rm cm}_{\nu\nu'}(R_{\nu\nu'})
\sum_i\delta_\mu(i) \frac{\partial{R}_{\nu\nu'}}{\partial {\bf r}_i}
\end{align}
where we assumed pair-wise interactions. Then
\begin{align}
\sum_i\delta_\mu(i) \frac{\partial{\bf R}_{\nu}}{\partial {\bf r}_i}&=
\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}
\sum_{i'}\delta_\nu(i')
\frac{m_{i'}}{m_\nu}{\bf r}_{i'}
\nonumber\\
&=
\sum_i\delta_\mu(i)
\sum_{i'}\delta_\nu(i')
\frac{m_{i'}}{m_\nu}{\bf 1}\delta_{ii'}
=
\sum_i\delta_\mu(i)
\delta_\nu(i)
\frac{m_{i}}{m_\nu}{\bf 1}=\delta_{\mu\nu}{\bf 1}
\end{align}
\begin{align}
\sum_i\delta_\mu(i) \frac{\partial{R}_{\nu\nu'}}{\partial {\bf r}_i}&=
{\bf e}_{\nu\nu'}\cdot \sum_i\delta_\mu(i)\frac{\partial{\bf R}_{\nu\nu'}}{\partial {\bf r}_i}
= {\bf e}_{\nu\nu'}\left[\delta_{\mu\nu}-\delta_{\mu\nu'}\right]
\end{align}
then
\begin{align}
-\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}
V^{0}_{\nu}(R)
&=
\frac{1}{2}\sum_{\nu'}F^{\rm cm}_{\nu\nu'}(R_{\nu\nu'})
\sum_i\delta_\mu(i) \frac{\partial{R}_{\nu\nu'}}{\partial {\bf r}_i}
\nonumber\\
&=
\frac{1}{2}\sum_{\nu'}F^{\rm cm}_{\nu\nu'}(R_{\nu\nu'})
{\bf e}_{\nu\nu'}\left[\delta_{\mu\nu}-\delta_{\mu\nu'}\right]
\nonumber\\
&=
\frac{1}{2}\sum_{\nu'}{\bf F}^{0}_{\nu\nu'}(R_{\nu\nu'})
\left[\delta_{\mu\nu}-\delta_{\mu\nu'}\right]
\end{align}
In summary, we have
\begin{align}
\hat{\bf F}_\mu&\equiv
-\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}
\left[\sum_\nu V^{\rm intra}_\nu(r)+
\sum_{\nu}\lambda(\hat{\bf R}_\nu)V^{\rm inter}_{\nu}(r)
+\sum_{\nu}(1-\lambda(\hat{\bf R}_\nu))V^{0}_{\nu}(R)
+\sum_\nu\calf(\lambda(\hat{\bf R}_\nu))\right]
\end{align}
and have to substitute in this expression the following results
\begin{align}
-\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}
\sum_\nu V^{\rm intra}_\nu(r)&=0
\nonumber\\
-\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}\lambda(\hat{\bf R}_\nu)
&=-\nabla \lambda(\hat{\bf R}_\nu)
\delta_{\mu\nu}
\nonumber\\
-\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}
V^{\rm inter}_{\nu}(r)
&=\frac{1}{2}\sum_{\nu'}{\bf F}^{\rm intra}_{\nu\nu'}(R_{\mu\nu'})
\left[\delta_{\mu\nu}-\delta_{\mu\nu'}\right]
\nonumber\\
-\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}
V^{0}_{\nu}(R)
&=\frac{1}{2}\sum_{\nu'}{\bf F}^{0}_{\nu\nu'}(R_{\nu\nu'})
\left[\delta_{\mu\nu}-\delta_{\mu\nu'}\right]
\end{align}
with the result
\begin{align}
\hat{\bf F}_\mu&=
-\left[\sum_{\nu}\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}
\lambda(\hat{\bf R}_\nu)(V^{\rm inter}_{\nu}(r)
-V^{0}_{\nu}(R)-\calf'(\lambda_\nu(R))
)\right]
\nonumber\\
&-\left[\sum_{\nu}\lambda(\hat{\bf R}_\nu)\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}V^{\rm inter}_{\nu}(r)
+\sum_{\nu}(1-\lambda(\hat{\bf R}_\nu))\sum_i\delta_\mu(i) \frac{\partial}{\partial {\bf r}_i}V^{0}_{\nu}(R)
\right]
\nonumber\\
&=-\nabla\lambda(\hat{\bf R}_\mu)(V^{\rm inter}_{\mu}(r)-V^{0}_{\mu}(R)-\calf'(\lambda_\mu(R)))
\nonumber\\
&+\sum_{\nu}\lambda(\hat{\bf R}_\nu)\frac{1}{2}\sum_{\nu'}{\bf F}^{1}_{\nu\nu'}(R_{\mu\nu'})
\left[\delta_{\mu\nu}-\delta_{\mu\nu'}\right]
+\sum_{\nu}(1-\lambda(\hat{\bf R}_\nu))
\frac{1}{2}\sum_{\nu'}{\bf F}^{0}_{\nu\nu'}(R_{\nu\nu'})
\left[\delta_{\mu\nu}-\delta_{\mu\nu'}\right]
\nonumber\\
&=-\nabla\lambda(\hat{\bf R}_\mu)(V^{\rm inter}_{\mu}(r)-V^{0}_{\mu}(R)-\calf'(\lambda_\mu(R)))
\nonumber\\
&+\sum_{\nu}\frac{\lambda(\hat{\bf R}_\mu)+\lambda(\hat{\bf R}_{\nu})}{2}
{\bf F}^{1}_{\mu\nu}(R_{\mu\nu})
+\sum_{\nu}\left(1-\frac{\lambda(\hat{\bf R}_\mu)+\lambda(\hat{\bf R}_{\nu})}{2}\right)
{\bf F}^{0}_{\nu\nu'}(R_{\mu\nu})
\end{align}
We may introduce the following pair force
\begin{align}
\hat{\bf G}_{\mu\nu}&=
\left[\frac{\lambda(\hat{\bf R}_\mu)+\lambda(\hat{\bf R}_\nu)}{2}\right]
{\bf F}^{1}_{\mu\nu}(R_{\mu\nu})
+\left[1-\frac{\lambda(\hat{\bf R}_\mu)+\lambda(\hat{\bf R}_\nu)}{2}\right]
{\bf F}^{0}_{\mu\nu}(R_{\mu\nu})
\end{align}
The pair force satisfies Newton's Third Law. With this definition we have
\begin{align}
\hat{\bf F}_\mu&=
-\nabla\lambda(\hat{\bf R}_\mu)(V^{\rm inter}_{\mu}(r)-V^{0}_{\mu}(R)-\calf'(\lambda_\mu(R)))
+\sum_{\nu} \hat{\bf G}_{\mu\nu}
\end{align}
\end{widetext}
|
\section{Introduction}
Approval Voting systems are widely considered~\cite{brams_av_book} as an alternative to traditional elections, where each voter may select and support at most some small number of candidates.
In Approval Voting each voter decides about every single candidate if he approves the candidate or does not approve him/her.
A result is obtained by applying a predefined election rule to the set of collected votes.
In this paper we study the problem of implementing an appropriate election rule and focus on the Minimax objective \cite{brams}: we minimize the biggest dissatisfaction over voters.
The resulting optimization problem is denoted $MAV$, and it is to select a committee composed of exactly $k$ candidates, and minimizing the maximal symmetric difference between the committee and the set of approved candidates by a single voter.
Using the string terminology, votes are encoded as strings, and the goal is to find a string encoding a committee minimizing the maximal Hamming distance to an input string.
Unlike in the related Closest String problem, in $MAV$ there is also a constraint: the selected committee must be of fixed size $k$, and hence in the string terminology there must be exactly $k$ ones in the string.
\subsection{Related work and our results}
Many different objective functions have been proposed and studied in the context of selecting the committee based on the set of votes collected in an Approval Voting system~\cite{computations_av,brams_av_book}.
Clearly, optimizing the sum of Hamming distances to all votes is an easy task and can be done by simply selecting the $k$ candidates approved by the largest number of voters.
By contrast, Minimax Approval Voting was shown by LeGrand~\cite{nphard} to be NP-hard.
LeGrand et al.~\cite{kcompletion} obtained $3$-approximation by a very simple $k$-completion algorithm. Next, Carragianis et al.~\cite{markakis} gave the currently best $2$-approximation algorithm.
The algorithm was obtained by rounding a fractional solution to the natural LP relaxation of the problem, and obtained approximation ratio essentially matches the integrality gap of the LP.
In this paper we give a PTAS for the Minimax Approval Voting problem. Our work is based on the PTAS for Closest String \cite{ptascs},
which is a similar problem to $MAV$ but there we do not have the restriction on the number of 1's in the result.
Technically, our contribution is the method of handling the number of 1's in the output. We also believe that our presentation is somewhat more intuitive.
Approval Voting systems are also analyzed in respect of manipulability, see e.g., \cite{computations_av} or \cite{markakis}. In particular, \cite{markakis} proved that each strategy-proof algorithm for $MAV$ must have approximation ratio at least $2-\frac{2}{k+1}$, which implies that our PTAS cannot be strategy-proof.
\subsection{Definitions}
We will use the following notation:\\
$n$ -- number of voters,\\
$m$ -- number of candidates,\\
$s_i \in \{0,1\}^m$ -- a vote of voter $i$,\\
$s_i[j]=1$ if voter $i$ approves candidate $j$,\\
$s_i[j]=0$ if voter $i$ does not approve candidate $j$,\\
$S=\{s_1,s_2,\dotsc,s_n\}$ -- the set of collected votes,\\
$s^{(1)}=\big|\{j:s[j]=1\}\big|$ -- the number of 1's in $s$.\\
For $x,y \in [\,0,1]^m$ we define a distance $d(x,y) = \sum_{j=1}^m \big|x[j]-y[j]\big| = \lVert x-y \rVert _{1}$.\\
For $x,y \in \{0,1\}^m$, $d(x,y)$ is called the Hamming distance.
\begin{definition}\label{def_mavk}
\[ OPT=\min_{\substack{x \in \{0,1\}^m\\x^{(1)}=k}} \: \max_{i \in \{1,2,\dots,n\}} d(x,s_i)\]
Let $s_{OPT}$ be an optimal solution, i.e., $\max_{i \in \{1,2,\dots,n\}} d(s_{OPT},s_i)=OPT$.
\end{definition}
WLOG we assume that $n>k$. If not, we copy the first string $k-n+1$ times.
\subsection{The main idea behind our algorithm}\label{name_consensus}
The general idea behind our PTAS is to find a small enough subset $X$ of votes that is a ``good representation'' of the whole set of votes $S$.
Then the candidates are partitioned into those for which voters in $X$ agree and the rest of candidates. For the ``consensus candidates'' we fix our decision
to the decision induced by votes in $X$ (additionally correcting the number of selected candidates in the ``consensus'' set).
Next, we consider the optimization problem of finding a proper subset of the remaining candidates to join the committee.
The key insight is that there exists a small enough subset $X$ such that the induced decision for the ``consensus candidates'' will not be a big mistake.
\subsection{Organization of the paper}
First, in Section~\ref{sec:info_subsets} we formalize the information we may extract from subset of votes, and introduce a measure of inaccuracy of such a subset.
Next, in Section~\ref{sec:existence_of_subset} we prove the existence of a small subset of votes with stable inaccuracy.
In Section~\ref{sec:aux_prob} we show that the optimization problem of deciding the part of the committee not induced by the subset of votes can be approximated
with only a small additional loss in the objective function.
Finally, in Section~\ref{sec:alg} we give an algorithm considering all subsets of a fixed size and show that, in the iteration when the algorithm
happens to consider a subset with stable inaccuracy, it will produce a $(1+\epsilon)$-approximate solution to $MAV$.
\section{Extracting information from subsets}
\label{sec:info_subsets}
We consider subsets of votes and analyze the information they carry. We measure the inaccuracy of this information with respect to the set of all votes. We show that there exists a small subset with stable inaccuracy, i.e., the drop of inaccuracy after including one more vote is small.
Let us define an inaccuracy function $ina:2^S \mapsto \mathbb{R}_{\geqslant 0}$ that measures the inaccuracy if we will consider subset $Y \subseteq S$ instead of $S$. The smaller the $ina(Y)$ is the better the common parts of strings in $Y$ represent $s_{OPT}$.
\begin{definition}\label{def_ina}
For all $Y \subseteq S, Y \neq \emptyset$ we define functions $t(Y) \in \{0,1\}^m$ and $ina(Y) \in \mathbb{R}_{\geqslant 0}$ as follows:
\[\big(t(Y)\big)[j] =
\begin{cases}
0 & \text{if } \forall_{y \in Y} \quad y[j]=0\\
1 & \text{if } \forall_{y \in Y} \quad y[j]=1\\
s_{OPT}[j] & \text{otherwise,}
\end{cases}\]
\[ ina(Y) = d(t(Y),s_{OPT}).\]
\end{definition}
Intuitively $t(Y)$ is the optimal solution $s_{OPT}$ changed at positions where all strings from $Y$ agree. Also we define the pattern of a subset of votes.
\begin{definition}\label{def_p}
For all $Y \subseteq S, Y \neq \emptyset$ we define pattern $p(Y) \in \{0,1,*\}^m$ as:
\[ \big(p(Y)\big)[j]=
\begin{cases}
0 & \text{if } \forall_{y \in Y} \quad y[j]=0\\
1 & \text{if } \forall_{y \in Y} \quad y[j]=1\\
* & \text{otherwise.}
\end{cases}
\]
\end{definition}
It represents positions that all strings in $Y$ agree. ``$*$'' encodes a mismatch.
Note that (from Definitions \ref{def_ina} and \ref{def_p}) $t(Y)$ is an optimal solution $s_{OPT}$ overwritten by a pattern $p_r$ on no-star positions:
\[\big(t(Y)\big)[j] =
\begin{cases}
s_{OPT}[j] & \text{if } \big(p(Y)\big)[j] = * \\
\big(p(Y)\big)[j] & \text{otherwise.}
\end{cases}
\]
The inaccuracy function has the following properties:
\begin{lemma}\label{decreasing_ina}
$\forall_{s_{i_1} \in S}$, for all sequences $\{s_{i_1}\}=Y_1 \subseteq Y_2 \subseteq \dots \subseteq Y_n = S$ we have
\[OPT \geqslant ina(Y_1) \geqslant ina(Y_2) \geqslant \dots \geqslant ina(Y_n) = 0\]
\end{lemma}
\begin{proof}
It is easy to see that
\[ina(Y_1) \stackrel{\text{def.}}{=} d\big(t(Y_1),s_{OPT}\big)=d\big(t(\{s_{i_1}\}),s_{OPT}\big)=d\big(s_{i_1},s_{OPT}\big) \leqslant OPT,\]
\[ina(Y_n) = ina(S) = d(s_{OPT},s_{OPT}) = 0.\]
Still we need to prove $ina(Y_i) \geqslant ina(Y_{i+1})$. Pattern $p(Y_{i+1})$ is built on strings from $Y_i \subseteq Y_{i+1}$ and strings from $Y_{i+1} \setminus Y_i$. So $p(Y_{i+1})$ has at least as many $*$ as $p(Y_i)$ has. Therefore $t(Y_{i+1})$ has at least as many positions as $t(Y_i)$ has that agree with optimal solution $s_{OPT}$, so $d\big(t(Y_i),s_{OPT}\big) \geqslant d\big(t(Y_{i+1}),s_{OPT}\big)$. Using definition of the inaccuracy function (Definition \ref{def_ina}) we prove the lemma. $\hfill\blacksquare$
\end{proof}
Intuitively $ina(Y)-ina(Y \cup \{y\})$ is the decrease of the inaccuracy from adding element $y$ to set $Y$. We will show that, when adding one more element $y$ to sets $Y,Z$ such that $Y \subseteq Z$, the inaccuracy decrease more in a case of adding $y$ to the smaller set $Y$ than adding $y$ to the bigger set $Z$.
\begin{lemma}\label{supermodular_ina}
If we artificially extend the $ina(\cdot)$ function for the empty set:\\ $ina(\emptyset)=2\cdot OPT$, then function $ina(\cdot)$ is supermodular\footnote{according to \cite{schrijver}, $f:2^S\mapsto \mathbb{R}$ is supermodular iff\\ $\forall_{Y,Z \subseteq S} \quad f(Y)+f(Z)\leqslant f(Y \cup Z)+f(Y\cap Z)$ which is equivalent with $\forall_{Y \subseteq Z \subseteq S} \quad\forall_{s\in S} \quad f(Z)-f(Z\cup\{s\})\leqslant f(Y)-f(Y \cup\{s\})$.}, i.e.,
\begin{equation}\label{supermodular_ina_ieq}
\forall_{Y \subseteq Z \subseteq S} \quad\forall_{s \in S} \quad ina(Z)-ina(Z \cup \{s\}) \leqslant ina(Y)-ina(Y \cup \{s\})
\end{equation}
\end{lemma}
\begin{proof}
Let fix $Y,Z$ and $s$ such that $Y \subseteq Z \subseteq S$ and $s \in S$.
\paragraph{\textbf{Case 1:}} $Z = \emptyset$:
Then also $Y = \emptyset$, and inequality (\ref{supermodular_ina_ieq}) holds obviously.
\paragraph{\textbf{Case 2:}} $Z \neq \emptyset, Y=\emptyset$:
We have:
\[ina(Z)-ina(Z \cup \{s\}) \leqslant OPT = \]
\begin{equation}\label{supermodular_ina_c2}
= 2\cdot OPT-OPT \leqslant ina(\emptyset)-ina(\{s\}) = ina(Y)-ina(Y \cup \{s\}),
\end{equation}
because we use respectively: Lemma \ref{decreasing_ina} and the fact that $Z$ has at least one element; definition of $ina(\cdot)$ for empty set and upperbound for $ina(\cdot)$ function; assumption that $Y = \emptyset$.
\paragraph{\textbf{Case 3:}} $Z \neq \emptyset, Y \neq \emptyset$:
From definition of $ina(\cdot)$ we have:
\[ ina(Z)-ina(Z \cup \{s\}) = d\big(t(Z),s_{OPT}\big)-d\big(t(Z \cup \{s\}),s_{OPT}\big) = \]
counting a difference by considering two cases for value of $s_{OPT}$ we obtain
\[ = \Big| \big\{ j: s_{OPT}[j]=1 \wedge t(Z \cup \{s\})[j]=1 \wedge t(Z)[j]=0 \big\}\Big| + \]
\[ +\;\, \Big| \big\{ j: s_{OPT}[j]=0 \wedge t(Z \cup \{s\})[j]=0 \wedge t(Z)[j]=1 \big\}\Big| = \]
using definition of function $t(\cdot)$:
\[ = \Big| \big\{ j: s_{OPT}[j]=1 \wedge s[j]=1 \wedge \;\forall_{z \in Z} \; z[j]=0 \big\}\Big| + \]
\[ +\;\, \Big| \big\{ j: s_{OPT}[j]=0 \wedge s[j]=0 \wedge \;\forall_{z \in Z} \; z[j]=1 \big\}\Big| \leqslant \]
taking an universal quantifier over a smaller subset we obtain:
\[ \leqslant \Big| \big\{ j: s_{OPT}[j]=1 \wedge s[j]=1 \wedge \;\forall_{y \in Y} \; y[j]=0 \big\}\Big| + \]
\[ + \;\,\Big| \big\{ j: s_{OPT}[j]=0 \wedge s[j]=0 \wedge \;\forall_{y \in Y} \; y[j]=1 \big\}\Big| = \]
reversing all previous transformations finally we obtain:
\[ = ina(Y)-ina(Y \cup \{s\}). \]$\hfill\blacksquare$
\end{proof}
\section{Existence of a stable subset}
\label{sec:existence_of_subset}
\begin{lemma}\label{lem_exists_x}
For any fixed $R\in\mathbb{N}_{\geqslant 1}$ there exists a subset $X \subseteq S, |X| = R$ such that
\begin{equation}\label{lem_x_exists_ina}
\forall_{s \in S \setminus X} \quad ina(X) - ina(X \cup \{s\}) \leqslant \frac{OPT}{R}.
\end{equation}
We say such $X$ is $\frac{OPT}{R}$-stable.
\end{lemma}
It means that there exists such a subset of votes $X$ that adding one more vote into $X$ the inaccuracy decreases by at most $\frac{OPT}{R}$.
\begin{proof}
First, we construct $S_{\underline{r}}$ satisfying (\ref{lem_x_exists_ina}) with at most $R$ elements.
Let us construct a sequence of subsets $S_1 \subset S_2 \subset \dotsc \subset S_n = S, |S_i|=i$. We take $S_1=\{s_{i_1}\}$, where $s_{i_1}$ is any element of $S$ and for $r \in \{2,3,\dots,n\}$ we take $S_{r}=S_{r-1}\cup\{s_{i_r}\}$ where $s_{i_r}$ is such a vote that after adding it the~inaccuracy function decreases the most, i.e.,
\begin{equation}\label{thm_sir}
s_{i_r} = \argmax_{s \in S\setminus S_{r-1}} \big(ina(S_{r-1})-ina(S_{r-1} \cup \{s\})\big).
\end{equation}
\begin{figure}\label{fig:ina}
\centering
\includegraphics[scale=1.15]{ina2}
\vspace{-10pt}
\caption{The $ina(\cdot)$ function for the sequence of subsets $S_1 \subset S_2 \subset \dotsc \subset S_n = S$.}
\end{figure}
We have
\[ \min_{r \in \{1,2,\dots,R\}} \quad ina(S_r)-ina(S_{r+1}) \leqslant \frac{1}{R} \left(\sum_{r=1}^R ina(S_r)-ina(S_{r+1})\right) = \]
\begin{equation}\label{thm_min_r}
= \frac{1}{R} \big( ina(S_1)-ina(S_{R+1}) \big) \leqslant \frac{OPT}{R},
\end{equation}
because (from Lemma \ref{decreasing_ina}) we know that $ina(S_1) \leqslant OPT$ and $ina(S_{R+1}) \geqslant 0$. Let $\underline{r}$ be a minimizer for the left-hand side of (\ref{thm_min_r}), then (by the choice of $s_{i_{\underline{r}}}$ in (\ref{thm_sir})) we have:
\begin{equation}\label{thm_max_diff_ina}
\max_{s \in S\setminus S_{\underline{r}}} \big(ina(S_{\underline{r}})-ina(S_{\underline{r}} \cup \{s\})\big)\leqslant \frac{OPT}{R},
\end{equation}
thus $S_{\underline{r}}$ satisfies (\ref{lem_x_exists_ina}), see Figure 1. If $S_{\underline{r}}$ has less elements than $R$ we can extend $S_{\underline{r}}$ to an $R$-elements subset $X$ by adding any elements of $S$. It follows from the supermodularity of $ina(\cdot)$. From Lemma \ref{supermodular_ina} we have:
\[\forall_{s \in S \setminus S_{\underline{r}}} \quad ina(X)-ina(X \cup \{s\}) \leqslant ina(S_{\underline{r}})-ina(S_{\underline{r}} \cup \{s\}), \]
and hence also:
\begin{equation}\label{thm_max_sx_x}
\max_{s \in S \setminus S_{\underline{r}}} \big(ina(X)-ina(X \cup \{s\})\big) \leqslant \max_{s \in S \setminus S_{\underline{r}}} \big(ina(S_{\underline{r}})-ina(S_{\underline{r}} \cup \{s\})\big).
\end{equation}
Finally, taking (\ref{thm_max_diff_ina}) and (\ref{thm_max_sx_x}) we obtain:
\[\max_{s \in S \setminus X} \big(ina(X)-ina(X \cup \{s\})\big) \leqslant \frac{OPT}{R}.\]$\hfill\blacksquare$
\end{proof}
Of course we cannot construct such a subset efficiently if we do not know $s_{OPT}$. How to find a proper subset $X$? For constructing our PTAS we will fix $R \in \mathbb{N}_{\geqslant 1}$ and consider all subsets $Y \subseteq S$ with cardinality $R$. There is less than $n^R \in Poly(n)$ such subsets. For clarity, we will use $Y \subseteq S$ in arguments valid for all subsets considered by the algorithm, and $X \subseteq S$ for a $\frac{OPT}{R}$-stable subset of votes.
For a fixed $Y \subseteq S, Y \neq \emptyset$, WLOG we reorder candidates in such a way that $p(Y)$ is a lexicographically smallest permutation:
\[p(Y)=**\dotsc *00\dotsc 011\dotsc 1.\]
The first part (from the left) is called ``star positions'' or ``star part''. The remaining part is called ``no-star part''. We define $p^{(*)}(Y)$ as the number of $*$ in $p(Y)$ and we denote it $\beta$:
\[ \beta = p^{(*)}(Y) = \Big|\left\{ j:\big(p(Y)\big)[j]=* \right\}\Big|.\]
In our PTAS we essentially fix the ``no-star part'' of the answer to the pattern $p(Y)$ and optimize over the choices for the ``star part'' of the outcome. If the~number of stars or number of 1's on star positions of $s_{OPT}$ is small enough, then there is only $Poly(m,n)$ possible solutions and we can consider all of them. Let us analyze the size of the ``star part''.
\begin{lemma}\label{size_p_star_y}
For all $Y \subseteq S$ we have
\[\beta = p^{(*)}(Y) \leqslant |Y| \cdot OPT\]
\end{lemma}
\begin{proof}
Consider an arbitrary $Y=\{y_1,y_2,\dots,y_{|Y|}\}$. We can construct $Y$ in the~following 3 phases:
\begin{enumerate}
\item[1.] $Y := \{s_{OPT}\}$
\item[2.] for $i=1$ to $|Y|$ do
\item[] \quad $Y := Y \cup \{y_i\}$
\item[3.] $Y := Y \setminus \{s_{OPT}\}$
\end{enumerate}
After that we obtain set Y. Let us calculate how many stars $p(Y)$ has. In Phase~1 there are no stars. In each step in Phase~2 we add at most $OPT$ stars, because $\forall_{i\in \{1,2,\dots,|Y|\}} \quad d(y_i,s_{OPT}) \leqslant OPT$. In Phase 3 we can at most decrease the~number of stars. So $\beta \leqslant |Y| \cdot OPT$.$\hfill\blacksquare$
\end{proof}
Note that for $X$ from Lemma \ref{lem_exists_x} we have
\begin{equation}\label{size_p_star}
p^{(*)}(X) \leqslant |X| \cdot OPT = R \cdot OPT.
\end{equation}
Let us now introduce some more notation. Assuming $Y \subseteq S$ and hence also $\beta = p^{(*)}(Y)$ are fixed,
we will use the following notation to denote the ``star part'' and the ``no-star part'' of a string $x \in \{0,1\}^m$:
\[x'=x[1]\cdot x[2]\cdot \dotsc\cdot x[\beta],\]
\[x''=x[\beta+1]\cdot x[\beta+2]\cdot\dotsc\cdot x[m],\]
where``$\cdot$'' is a concatenation of strings (letters).
So we divide $x$ into two parts: $x=x'\cdot x''$.
Let us now define a $k$-completion of $x \in \{0,1\}^m$ (definition from \cite{kcompletion}) to be a $y \in \{0,1\}^m$ such that $y^{(1)}=k$ and $d(y,x)$ is the minimum possible Hamming distance between $x$ and any vector with $k$ of 1's. To obtain a $k$-completion we only add or only delete a proper number of 1's. To be more specific in this paper we assume the $k$-completion is always obtained by changing bits at positions with the smallest possible index\footnote{Any other deterministic rule would work for us just as well.}.
In the following lemma we will show that for the pattern from a stable subset $X$ we can change the number of 1's in the ``no-star part'' to the properly guessed number of 1's loosing only twice the stability constant.
\begin{lemma}\label{lemma_kbiscompletion}
If $X \subseteq S$ is $(\epsilon_1 \cdot OPT)$-stable, $z''$ is a $k''$-completion of $\big(p(X)\big)''$, where $k''=(s_{OPT}'')^{(1)}$, then
\begin{equation}\label{lemma_kbis_completion}
\forall_{i \in \{1,2,\cdots,n\}} \quad d(s_{OPT}'\cdot z'',s_i) \leqslant (1+2\epsilon_1) \cdot OPT
\end{equation}
\end{lemma}
\begin{proof}
WLOG there is insufficient number of 1's in no-star part of pattern $p(X)$, i.e., $k'' \geqslant \big((p(X))''\big)^{(1)}$. The other case is symmetric.
Let us fix $s_i \in S$ and consider all combinations of values in strings $\big(p(X)\big)''$, $z''$, $s_i''$, $s_{OPT}''$ at the same position $j$. $\alpha_a \in \mathbb{N}$, for $a\in \{1,2,\cdots,12\}$, counts the number of positions $j$ with combination $a$, see Table 1.
\begin{table}[h]\label{table_cases}
\centering
\begin{tabular}{|c|c|c|c|c|c|c|c|c|c|c|c|c|}
\hline
& \multicolumn{12}{c|}{combinations} \\ \hline
\quad index of a combination \quad & 1 & 2 & 3 & 4 & 5 & 6 & 7 & 8 & 9 & 10& 11& 12\\ \hline
$(p(X))''[j]$ & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 0 & 1 & 1 & 1 & 1 \\
$z''[j]$ & 0 & 0 & 0 & 0 & 1 & 1 & 1 & 1 & 1 & 1 & 1 & 1 \\
$s_i''[j]$ & 0 & 1 & 0 & 1 & 0 & 1 & 0 & 1 & 0 & 1 & 0 & 1 \\
$s_{OPT}''[j]$ & 0 & 0 & 1 & 1 & 0 & 0 & 1 & 1 & 0 & 0 & 1 & 1 \\ \hline
number of occurrences & $\alpha_1$ & $\alpha_2$ & $\alpha_3$ & $\alpha_4$ & $\alpha_5$ & $\alpha_6$ & $\alpha_7$ & $\alpha_8$ & $\alpha_9$ & $\alpha_{10}$ & $\alpha_{11}$ & $\alpha_{12}$ \\ \hline
$d(z''[j],s_i''[j])$ & 0 & 1 & 0 & 1 & 1 & 0 & 1 & 0 & 1 & 0 & 1 & 0 \\ \hline
$d(s''_{OPT}[j],s_i''[j])$ & 0 & 1 & 1 & 0 & 0 & 1 & 1 & 0 & 0 & 1 & 1 & 0 \\ \hline
\end{tabular}
\vspace{5pt}
\caption{Combinations of values in strings $(p(X))'', z'', s_i'', s_{OPT}''$. There is only 12 combinations (no $2^4=16$), because by the assumption $k'' \geqslant ((p(X))'')^{(1)}$ we never change from 1 in $((p(X))'')^{(1)}$ to 0 in $z''$.}
\vspace{-15pt}
\end{table}
We have:
\[ d(z'',s_i'') = \big|\{j: z''[j] \neq s_i''[j] \}\big| = \]
we consider two cases for value of $s_{OPT}$ at position $j$:
\[ = \big|\{j: z''[j] \neq s_i''[j] \wedge (z''[j] = s_{OPT} \vee z''[j] \neq s_{OPT})\}\big| = \]
we divide it into two components:
\begin{alignat*}{4}
= &\big|\{j: s_{OPT} = && \; z''[j] \neq s_i''[j] &&\}\big| + \\
+ &\big|\{j: && \; z''[j] \neq s_i''[j] = s_{OPT}) &&\}\big| =
\end{alignat*}
we use case counts from Table 1 to count positions in both components:
\[ = ( \underbrace{\alpha_2 + \alpha_7 + \alpha_{11}}_\text{first component} + \underbrace{\alpha_3 + \alpha_6 + \alpha_{10} - \alpha_3 - \alpha_6 - \alpha_{10}}_{=0}) + \underbrace{(\alpha_4 + \alpha_5 + \alpha_9)}_\text{second component} = \]
and we use the definition of the Hamming distance:
\begin{equation}\label{d_z_bis_s_i_bis}
= \big(d(s_{OPT}'',s_i'') - \alpha_3 - \alpha_6 - \alpha_{10}\big) + (\alpha_4 + \alpha_5 + \alpha_9).
\end{equation}
Since $(z'')^{(1)} = k'' = \big(s_{OPT}''\big)^{(1)}$,
\[ \sum_{k=5}^{12} \alpha_k = \alpha_3 + \alpha_4 + \alpha_7 + \alpha_8 + \alpha_{11} + \alpha_{12} \]
\begin{equation}\label{alpha5}
\alpha_5 = \alpha_3 + \alpha_4 - \alpha_6 - \alpha_9 - \alpha_{10}.
\end{equation}
Also
\begin{equation}\label{epsilon1opt_stable}
\alpha_4 + \alpha_8 + \alpha_9 \leqslant \epsilon_1 \cdot OPT,
\end{equation}
because $X$ is $\epsilon_1 \cdot OPT$-stable.
Now we are ready to prove equation (\ref{lemma_kbis_completion}).
\[ d(s_{OPT}'\cdot z'',s_i) \stackrel{\rm def.}{=} d(s_{OPT}',s_i') + d(z'',s_i'') \stackrel{(\ref{d_z_bis_s_i_bis})}{=} \]
\[ \stackrel{(\ref{d_z_bis_s_i_bis})}{=} d(s_{OPT}',s_i') + d(s_{OPT}'',s_i'') - \alpha_3 - \alpha_6 - \alpha_{10} + \alpha_4 + \alpha_5 + \alpha_9 \stackrel{(\ref{alpha5})}{=} \]
\[ \stackrel{(\ref{alpha5})}{=} \underbrace{d(s_{OPT},s_i)}_{\leqslant OPT} + 2(\hspace{-10pt} \underbrace{\alpha_4}_{\stackrel{(\ref{epsilon1opt_stable})}{\leqslant} \epsilon_1 \cdot OPT} \hspace{-15pt} - \alpha_6 - \alpha_{10}) \stackrel{(\ref{epsilon1opt_stable})}{\leqslant} (1+2\epsilon_1) \cdot OPT. \]
$\hfill\blacksquare$
\end{proof}
\section{An auxiliary optimization problem}
\label{sec:aux_prob}
In this section we will consider the optimization problem obtained after guessing the number of 1's in the two parts and fixing the ``no-star part'' of the outcome.
It has variables for all the positions of the ``star part'' and constraints for all the original votes $s_i \in S$.
Let us define the optimization problem in terms of the integer program $IP_{(\ref{ip_q})-(\ref{ip_p01})} (Y,k')$ by (\ref{ip_q})-(\ref{ip_p01}):
\begin{equation}\label{ip_q}
\min q
\end{equation}
\begin{equation}\label{ip_no1}
(s')^{(1)} = k'
\end{equation}
\begin{equation}\label{ip_dist}
\forall_{i \in \{1,2,\dots,n\}} \quad d(s',s_i') \leqslant q-d(s_{ALG}'',s_i'')
\end{equation}
\begin{equation}\label{ip_qg0}
q \geqslant 0
\end{equation}
\begin{equation}\label{ip_p01}
\forall_{j \in \{1,2,\dots,\beta\}} \quad s'[j] \in \{0,1\}
\end{equation}
where $Y \subseteq S, k=k'+k''$, and $s_{ALG}''$ is the $k''$-completion of $(p(Y))''$. Recall that $\beta=p^{(*)}(Y)$ and $(p(Y))''$ is the ``no-star part'' of the~pattern $p(Y)$.
In the LP relaxation (\ref{ip_p01}) is replaced with:
\begin{equation}\label{lp_p01}
\forall_{j \in \{1,2,\dots,\beta\}} \quad s'[j] \in [0,1]
\end{equation}
Constraints (\ref{ip_q})-(\ref{ip_qg0}),(\ref{lp_p01}) are linear because
\[(s')^{(1)}=\sum_{j=1}^\beta s'[j],\]
\[d(s',s_i') = \sum_{j=1}^\beta \Big(\chi(s_i'[j]=0)\cdot s'[j] + \chi(s_i'[j]=1)\cdot(1-s'[j])\Big)\] are linear functions of $s'[j]$, where $j \in \{1,2,\dots,\beta\}$.
\begin{lemma}\label{lem_aprox_ip}
$\forall_{R \in \mathbb{N}_{\geqslant 1}, Y \subseteq S, |Y| \leqslant R, k' \in \mathbb{N}, \epsilon_2 > 0}$ we can find $(1+2\epsilon_2)$-approximation solution for $IP_{(\ref{ip_q})-(\ref{ip_p01})} (Y,k')$ by solving the $LP$ and considering at most
\[ (3n)^{\frac{3R \ln(2)}{(\epsilon_2)^2}} + m^{\frac{3R^2\ln(6)}{(\epsilon_2)^2}} \quad\text{cases.} \]
\end{lemma}
\begin{proof}
Let us fix constants $\epsilon_2 \in (0,\frac{1}{2})$ (for $\epsilon_2 \geqslant \frac{1}{2}$ we could use $2$-approximation from \cite{markakis}). Consider three cases:
\paragraph{\textbf{Case 1:}} $\beta \leqslant \frac{3 R\ln(3n)}{(\epsilon_2)^2}$\\
There is $2^\beta$ possibilities for $s'$.\\
\[2^\beta \leqslant 2^{\frac{3R\ln(3n)}{(\epsilon_2)^2}} = e^{\ln(3n) \frac{3R \ln(2)}{(\epsilon_2)^2}} = (3n)^{\frac{3R \ln(2)}{(\epsilon_2)^2}} \in Poly(n),\]
because $\epsilon_2$ and $R$ are fixed constants.
So we will check (in polynomial time) all possibilities for $s'$ and we will find optimal solution for the integer program.
\paragraph{\textbf{Case 2:}} $k'\leqslant \frac{3R^2\ln(6)}{(\epsilon_2)^2}$\\
There is $Poly(m)$ possibilities for $s'$ because we can upperbound the number of possibilities of setting 1's into $\beta$ positions by:
\[ {\beta \choose k'} \leqslant \beta^{k'} \leqslant \beta^{\frac{3R^2\ln(6)}{(\epsilon_2)^2}} \leqslant m^{\frac{3R^2\ln(6)}{(\epsilon_2)^2}} \in Poly(m), \]
because $\epsilon_2$ and $R$ are fixed constants.
\paragraph{\textbf{Case 3:}} $\;\beta > \frac{3R\ln(3n)}{(\epsilon_2)^2} \;\wedge\; k' > \frac{3R^2\ln(6)}{(\epsilon_2)^2}$\\
We denote an optimal solution of the~$IP_{(\ref{ip_q})-(\ref{ip_p01})} (Y,k')$ by $\big((s')^{IP},q^{IP}\big)$. Let us use LP relaxation and denote an optimal solution of the LP by $\big((s')^{LP},q^{LP}\big)$. Obviously we have $q^{LP} \leqslant q^{IP}$. We can solve the LP in polynomial time but we may obtain a fractional solution. We want to round it independently. We will use a randomized rounding defined by distributions on each position $j \in \{1,2,\dots,\beta\}$:
\begin{equation}\label{pse1}
P\big(s'[j]=1\big)= (s')^{LP}[j], \quad P\big(s'[j]=0\big)= 1-(s')^{LP}[j].
\end{equation}
We can estimate the expected value of a distance to such a random solution $s'$:
\[ \forall_{i \in \{1,2,\cdots,n\}} \quad \mathbb{E}\big[ d(s',s_i') \big] \stackrel{\text{def.}}{=} \mathbb{E}\left[ \sum_{j=1}^\beta \Big|s'[j]-s_i'[j]\Big| \right] = \]
\begin{alignat*}{6}
= \mathbb{E}\Bigg[ &\sum_{j=1}^\beta& \Big( \quad &\chi(s_i'[j]=0)\cdot s'[j] \quad &+ \quad &\chi(s_i'[j]=1)\cdot (1-s'[j]) &&\Big) \Bigg] \stackrel{\text{lin. of }\mathbb{E}}{=} \\
\stackrel{\text{lin. of }\mathbb{E}}{=} &\sum_{j=1}^\beta& \Big( \quad &\chi(s_i'[j]=0)\cdot \mathbb{E}\big[s'[j]\big]\quad &+ \quad &\chi(s_i'[j]=1)\cdot \mathbb{E}\big[1-s'[j]\big] &&\Big) \stackrel{(\ref{pse1})}{=} \\
\stackrel{(\ref{pse1})}{=} &\sum_{j=1}^\beta& \Big( \quad &\chi(s_i'[j]=0)\cdot (s')^{LP}[j]\quad &+ \quad &\chi(s_i'[j]=1)\cdot \big(1-(s')^{LP}[j]\big) &&\Big) \stackrel{\text{def.}}{=}
\end{alignat*}
\begin{equation}\label{exp_d_p_si}
\stackrel{\text{def.}}{=} d\big((s')^{LP},s_i'\big) \stackrel{(\ref{ip_dist})}{\leqslant} q^{LP}-d(s_{ALG}'',s_i'').
\end{equation}
$d(s',s_i')$ is a sum of $\beta$ independent 0-1 variables. For $\epsilon' \in (0,1)$ using Chernoff's bound \cite{motvani} we have:
\[ P\Big(d(s',s_i') \geqslant (1+\epsilon') \cdot\mathbb{E}\big[d(s',s_i')\big] \Big) \leqslant \exp\left( -\frac{1}{3}(\epsilon')^2 \cdot \mathbb{E}\big[d(s',s_i')\big] \right). \]
If we take $\epsilon' = \frac{\epsilon_2 \cdot q^{IP}}{\mathbb{E}[d(s',s_i')]}$ then we obtain:
\[ \exp\left( -\frac{1}{3} \cdot \frac{(\epsilon_2)^2 \cdot (q^{IP})^2}{\mathbb{E}\big[d(s',s_i')\big]} \right) \geqslant P\Big(d(s',s_i') \geqslant \mathbb{E}\big[d(s',s_i')\big] + \epsilon_2 \cdot q^{IP} \Big) \stackrel{(\ref{exp_d_p_si})}{\geqslant} \]
\begin{equation}\label{expgpbb}
\stackrel{(\ref{exp_d_p_si})}{\geqslant} P\Big(d(s',s_i') \geqslant q^{LP}-d(s_{ALG}'',s_i'') + \epsilon_2 \cdot q^{IP} \Big).
\end{equation}
We want to know an upperbound for the probability that we make an error greater than $\epsilon_2 \cdot q^{IP}$ for at least one vote:
\[ P\Big(\exists_{i \in \{1,2,\dots,n\}}: d(s',s_i') \geqslant q^{LP}-d(s_{ALG}'',s_i'') + \epsilon_2 \cdot q^{IP} \Big) \stackrel{(\ref{expgpbb})}{\leqslant} \]
\begin{equation}\label{pbb_error_all}
\stackrel{(\ref{expgpbb})}{\leqslant} n \cdot \exp\left( -\frac{1}{3} \cdot \frac{(\epsilon_2)^2 \cdot (q^{IP})^2}{\mathbb{E}\big[d(s',s_i')\big]} \right) \leqslant n \cdot \exp\left( -\frac{1}{3} (\epsilon_2)^2 \cdot q^{IP} \right),
\end{equation}
where the last inequality is because of:
\[ \mathbb{E}\big[ d(s',s_i') \big] \stackrel{(\ref{exp_d_p_si})}{\leqslant} q^{LP}-d(s_{ALG}'',s_i'') \leqslant q^{IP}. \]
We want to further upperbound the probability in (\ref{pbb_error_all}). From the assumption about $\beta$ and from Lemma~\ref{size_p_star_y} we have:
\[ \frac{3R\ln(3n)}{(\epsilon_2)^2} < \beta \stackrel{\rm{Lem. \ref{size_p_star_y}}}{\leqslant} |Y| \cdot OPT \leqslant R \cdot OPT \leqslant R \cdot q^{IP}, \quad \text{equivalently}\]
\begin{equation}\label{1deltalnexp}
\frac{1}{3} > n \cdot \exp\left( -\frac{1}{3} (\epsilon_2)^2 \cdot q^{IP} \right).
\end{equation}
So, finally we have:
\begin{equation}\label{pexistslqd}
P\Big(\exists_{i \in \{1,2,\dots,n\}}: d(s',s_i') \geqslant q^{LP}-d(s_{ALG}'',s_i'') + \epsilon_2 \cdot q^{IP} \Big) \stackrel{(\ref{pbb_error_all}),(\ref{1deltalnexp})}{<} \frac{1}{3}.
\end{equation}
So with probability at least $\frac{2}{3}$ we obtain:
\[ \forall_{i\in\{1,2,\dots,n\}} \quad d(s' \cdot s_{ALG}'', s_i) = d(s', s_i') + d(s_{ALG}'', s_i'') \stackrel{(\ref{pexistslqd})}{<} \]
\begin{equation}\label{eps2_approx_ip}
\stackrel{(\ref{pexistslqd})}{<} q^{LP}-d(s_{ALG}'', s_i'') +\epsilon_2\cdot q^{IP} + d(s_{ALG}'', s_i'') \leqslant (1+\epsilon_2)\cdot q^{IP}.
\end{equation}
We can also obtain a wrong number o 1's. The solution $s_{ALG}'$ for that is to take the $k'$-completion of $s'$. We will show that the additional error for such operation is not so big. Expected number of 1's in $s'$ is equal $k'$:
\[ \mathbb{E}\big[ (s')^{(1)} \big] \stackrel{\rm{def.}}{=} \mathbb{E}\left[ \sum_{j=1}^\beta s'[j] \right] \stackrel{\text{lin. of }\mathbb{E}}{=} \sum_{j=1}^\beta (s')^{LP}[j] \stackrel{\rm{def.}}{=} \big((s')^{LP}\big)^{(1)} \stackrel{(\ref{ip_no1})}{=} k'. \]
We want to know how much we lose taking the $k'$-completion. Similar as before, $(s')^{(1)} = \sum_{j=1}^\beta s'[j]$ is a sum of $\beta$ independent 0-1 variables. For $\epsilon'' \in (0,1)$ using Chernoff's bound \cite{motvani} we have:
\[ P\left( (s')^{(1)} \geqslant (1+\epsilon'')\cdot k' \right) \leqslant \exp\left( -\frac{1}{3}(\epsilon'')^2 \cdot k' \right), \]
\[ P\left( (s')^{(1)} \leqslant (1-\epsilon'')\cdot k' \right) \leqslant \exp\left( -\frac{1}{2}(\epsilon'')^2 \cdot k' \right). \]
Taking both inequalities together, $\epsilon'' = \frac{\epsilon_2}{R}$ and using assumption $k' > \frac{3R^2\ln(6)}{(\epsilon_2)^2}$ we have:
\[ P\left( \left|(s')^{(1)}-k'\right| \geqslant \epsilon''\cdot k' \right) \leqslant 2\cdot \exp\left( -\frac{1}{3}(\epsilon'')^2 \cdot k' \right) \leqslant \]
\[ \leqslant 2\cdot \exp\left( -\frac{1}{3}\frac{(\epsilon_2)^2}{R^2} \cdot k' \right) < \frac{1}{3}. \]
So with probability at least $\frac{2}{3}$ the error from taking the $k'$-completion is not greater than $ \epsilon'' \cdot k' = \frac{\epsilon_2}{R} \cdot k' \leqslant \frac{\epsilon_2}{R} \cdot \beta \stackrel{\rm{Lem. \ref{size_p_star_y}}}{\leqslant} \frac{\epsilon_2}{R} \cdot |Y| \cdot OPT \leqslant \epsilon_2 \cdot OPT \leqslant \epsilon_2 \cdot q^{IP}$.
Combining the above with (\ref{eps2_approx_ip}) we obtain a $(1+2\epsilon_2)$-approximate solution with probability at least $\frac{1}{3}$. We may derandomize the algorithm analogously to how it was done in the PTAS for the Closest String problem \cite{ptascs}. For more on derandomization techniques see~\cite{derandomization}. $\hfill\blacksquare$
\end{proof}
\section{Algorithm and its complexity analysis}
\label{sec:alg}
Now we are ready to combine the ideas into a single algorithm.
\begin{algorithm}
\caption{ALG(R)}
\label{alg_mav}
\begin{algorithmic}[1]
\REQUIRE $S=\{s_1,s_2,\dots,s_n\} \in (\{0,1\}^m)^n, 0 \leqslant k \leqslant m,R \in \mathbb{N}_{\geqslant 1}$
\ENSURE $s_{ALG} \in \{0,1\}^m$
\FOR{each $R$-element subset $Y=\{s_{i_1},s_{i_2},\dots,s_{i_R}\} \subseteq S$}\label{forRsubset}
\FOR{each division $k$ into two parts $k=k'+k''$}
\STATE{$s''_{ALG} \leftarrow k''$-completion of $(p(Y))''$ \\(if not possible, then skip this inner iteration)}
\STATE{$s'_{ALG} \leftarrow$ approximation solution of $IP_{(\ref{ip_q})-(\ref{ip_p01})} (Y,k')$ using Lemma \ref{lem_aprox_ip}\\ (if $LP_{(\ref{ip_q})-(\ref{ip_qg0}),(\ref{lp_p01})}(Y,k')$ infeasible, then skip this inner iteration)} \label{approx_opt_problem}
\STATE{evaluate $s'_{ALG}\cdot s''_{ALG}$ by computing $ \max_{i\in \{1,2,\dots,n\}} d(s_i,s'_{ALG}\cdot s''_{ALG})$}
\ENDFOR
\ENDFOR\label{forRsubsetend}
\STATE{$s_{ALG} \leftarrow$ the best solution from a loop in lines \ref{forRsubset}-\ref{forRsubsetend}}
\end{algorithmic}
\end{algorithm}
It remains to argue that for a large enough parameter $R$ the above algorithm will at some point consider a subset of votes $X$
that leads to an accurate enough approximation of the Minimax objective function of our problem.
\begin{theorem}
$\forall_{\epsilon \in (0,1)}$ we may compute a $(1+\epsilon)$-approximate solution to Minimax Approval Voting in $O\big(Poly(n,m)\big)$ time.
\end{theorem}
\begin{proof}
Let $\epsilon_0 = \frac{\epsilon}{3} < \frac{1}{3}$.
By Lemma \ref{lem_exists_x}, there exists an $\frac{\epsilon_0 \cdot OPT}{2}$-stable set of votes $X\subseteq S$ of cardinality $|X|=R=\lceil \frac{2}{\epsilon_0} \rceil$.
Consider algorithm ALG(R). In one iteration it will consider $X$ and $k',k''$ such that $(s_{OPT}')^{(1)}=k'$. Recall that $s_{ALG}''$ is the specific $k''$-completion of $\big(p(X)\big)''$. By Lemma \ref{lemma_kbiscompletion} we have:
\[ d(s'_{OPT} \cdot s_{ALG}'',s_i) \leqslant (1+\epsilon_0)\cdot OPT,\]
hence $\big(s'=s_{OPT}',q=(1+\epsilon_0)\cdot OPT \big)$ is a feasible solution to $IP_{(\ref{ip_q}-\ref{ip_p01})} (X,k')$ and the optimal value of $IP_{(\ref{ip_q}-\ref{ip_p01})} (X,k')$ is at most $(1+\epsilon_0)\cdot OPT$.
By Lemma \ref{lem_aprox_ip} with $\epsilon_2 = \frac{\epsilon_0}{2}$ we find a $(1+\epsilon_0)$-approximate solution $\big( s_{ALG}',q_{ALG} \big)$ to $IP_{(\ref{ip_q}-\ref{ip_p01})} (X,k')$. So we have:
\[q_{ALG} \leqslant (1+\epsilon_0) \cdot (1+\epsilon_0) \cdot OPT \stackrel{\epsilon_0<1}{\leqslant} (1+3\epsilon_0) \cdot OPT = (1+\epsilon) \cdot OPT.\]
It remains to observe, that $s_{ALG}=s_{ALG}' \cdot s_{ALG}''$ is a solution to $MAV$ of cost $q_{ALG} \leqslant (1+\epsilon) \cdot OPT$.
The algorithm examined $O(n^R)=O\big(n^{\lceil \frac{6}{\epsilon} \rceil} \big) \in O(Poly(n))$ subsets $Y$, $O(m)$ choices of $k'$ and each time considered
\[ O\left( (3n)^{\frac{108 \cdot \lceil \frac{6}{\epsilon} \rceil \cdot \ln(2) }{\epsilon^2}} + m^{\frac{108 \cdot {\lceil \frac{6}{\epsilon} \rceil}^2 \cdot \ln(6)}{\epsilon^2}} \right) \in O(Poly(n,m)) \;\, \text{cases.}\] $\hfill\blacksquare$
\end{proof}
\section{Concluding remarks}
We showed the existence of a PTAS for Minimax Approval Voting by considering all subsets of a fixed size $R$. If not the discovered supermodularity for the~inaccuracy function $ina(\cdot)$, we would simply consider all subsets of size at most $R$. Although the supermodularity was not essential for our result, it shows that larger subsets of votes are generally more stable (in the sense of definition in Lemma \ref{lem_exists_x}). It seems to suggest that an algorithm considering a smaller number of larger subsets of votes would potentially be more efficient in practice. Perhaps the most interesting open question is whether by randomly sampling a number of subsets of votes to examine, one could obtain a more practical FPRAS for the~problem.
Another interesting direction is the optimization of the Minimax objective function subject to a restriction that the voting system must be incentive compatible. According to~\cite{markakis} the best possible approximation ratio in this setting is between $2-\frac{2}{k+1}$ and $3-\frac{2}{k+1}$, and a natural challenge is to narrow this gap.
Finally, we know the complexity of the two extreme objectives, i.e., Minimax and Minisum. The latter is easily optimized by selecting the $k$ most often approved candidates. The optimization problem for intermediate objectives such as optimizing the sum of squares of the Hamming distances remains unexplored, and it would be interesting to learn which objective functions are more difficult to approximate than Minimax in the context of Approval Voting systems.
\section*{Acknowledgments}
We want to thank Katarzyna Staniewicz for many helpful proofreading comments. Also we want to thank reviewers for their valuable suggestions. Krzysztof Sornat was supported by local grant 2139/M/II/14.
|
\section{Introduction}
The majority of galaxies in the local universe reside in small,
gravitationally bound groups \citep{Ekeetal04}, whose low velocity
dispersions ($\lesssim$500\ensuremath{\km \ps}) and small galaxy separations are conducive
to tidal interactions and mergers between group members. X-ray observations
have shown that many groups host extended halos of hot gas, but that the
existence of a hot intra-group medium (IGM) appears to be linked to the
presence of early-type galaxies. X-ray luminous groups are sometimes
described as miniature galaxy clusters; their IGM is highly enriched,
particularly in the group core, they host significant early-type galaxy
populations, are often dominated by a single, centrally-located giant
elliptical, and follow morphology-density and morphology-radius relations
similar to (but offset from) those seen in clusters
\citep{HelsdonPonman03a}. However, they are typically poor in cold gas,
with neutral hydrogen restricted to spiral galaxies in the group outskirts
\citep{Kilbornetal06}.
Conversely, spiral-rich groups are typically poor in hot gas, and
consequently X-ray faint. \citet{Mulchaeyetal03} find that spiral-rich
groups tend to be less X-ray luminous than their elliptical-dominated
cohorts, and detect none of the twelve spiral-only groups in their \emph{ROSAT}\
atlas of 109 systems. \citet{OsmondPonman04} detect only galaxy-scale
emission in the ten spiral-only groups in the GEMS sample; the only
possible exception, the NGC~3783 group, appears to be biased by the
exceptionally luminous Seyfert nucleus of the dominant spiral. However,
spiral-rich groups usually contain significant quantities of cold gas in and
around their galaxies. Examination of X-ray faint, spiral-rich compact
groups has led to the suggestion of an evolutionary sequence, with galaxy
interactions stripping the \ensuremath{\mathrm{H}\textsc{i}}\ from spiral galaxies to form intergalactic
clouds and filaments or even a diffuse cold IGM
\citep{VerdesMontenegroetal01,Johnsonetal07,Konstantopoulosetal10}. The
redistribution of the \ensuremath{\mathrm{H}\textsc{i}}\ component is accompanied by the transformation
of some member galaxies from late to early-type, and in some cases by star
formation.
While the role of tidal interactions in driving the evolution of the galaxy
population and \ensuremath{\mathrm{H}\textsc{i}}\ component is clear, the origin of the hot IGM and its
link to the development of the group is not. Infall and gravitational shock
heating is believed to be the primary source of the hot gas which makes up
the dominant baryonic component of massive clusters, and the same mechanism
probably provides most of the IGM in the most massive groups. However, the
connection between galaxy evolution and the presence of a hot IGM in low
mass groups suggests a link. One possibility is that star-formation-driven
galactic winds could contribute to the formation of the IGM. Another is
that intergalactic \ensuremath{\mathrm{H}\textsc{i}}\ could be shock heated by collisions within the
group. The latter process is observed in one system, Stephan's Quintet
(HCG~92), in which an infalling spiral galaxy has collided with a tidal \ensuremath{\mathrm{H}\textsc{i}}\
filament, heating it to a temperature of $\sim$0.6~keV
\citep{VanderhulstRots81,Sulenticetal01,Trinchierietal03,OSullivanetal09}.
Understanding the development of the hot IGM is clearly central to any
study of galaxy evolution in groups, or of structure formation involving
groups. However, relatively few suitable systems have been examined in
detail, and their X-ray faintness makes such examination challenging. Only
two groups at an earlier evolutionary stage than Stephan's Quintet have
been shown to possess a hot IGM using modern high-spatial-resolution X-ray
instruments. \citet{Trinchierietal08} confirmed the presence of diffuse hot
gas in SCG0018-4854, a spiral-only southern compact group of four galaxies,
but the short \emph{XMM}\ observation provided only limited information on the
state and origin of the IGM and its relation to the galaxy population. The
other example is the well known spiral-dominated group HCG~16, which we
have chosen to study in this paper. HCG~16 appears to contain significant
quantities of diffuse hot and cold gas (\ensuremath{\mathrm{H}\textsc{i}}), and its galaxy members
include starburst systems with outflowing winds which may be enriching
their surroundings. The group therefore appears to be at the start of the
process of galaxy merger, hot gas build-up and enrichment which will
eventually produce an X-ray bright, metal-rich IGM and an
elliptical-dominated group.
In order to examine the process of IGM development in HCG~16, we have
obtained deep \emph{Chandra}\ X-ray observations of the group, as well as new
\textit{Giant Metrewave Radio Telescope} (GMRT) 610~MHz radio data. In this
paper we combining these observations with archival \emph{Chandra}\ and
\textit{Very Large Array} (VLA) 1.4~GHz data, with the goal of determining
the physical properties and origin of the diffuse gas component, and its
relationship to the galaxy population. We describe the observations, data
reduction and analysis techniques in detail in \citet[hereafter Paper
I]{OSullivanetal14b_special}. All five of the major galaxies in the group show
evidence of star formation and/or nuclear activity, with two of the
galaxies hosting galactic superwinds. A full discussion of the properties
of the galaxies and their point source populations can be found in Paper I,
and we summarise the results relevant to the current paper in
Section~\ref{sec:winds}.
Throughout this paper we adopt a redshift of $z$=0.0132 for the group
\citep{Hicksonetal92} and a Galactic hydrogen column density of
\NH=2.56$\times$10$^{20}$\ensuremath{\cm^{-2}}\ for the four original group member
galaxies and the surrounding diffuse emission \citep[taken from the
Leiden/Argentine/Bonn survey,][]{Kalberlaetal05}. For NGC~848 we adopt a
hydrogen column of \NH=2.75$\times$10$^{20}$\ensuremath{\cm^{-2}}. All fluxes and
luminosities are corrected for Galactic absorption. A redshift-independent
distance measurement is available for one of the five major galaxies, a
Tully-Fisher distance of 56.5~Mpc for NGC~848 \citep{Theureauetal07}. This
is consistent, within errors, with redshift-based estimates for all five of
the galaxies, correcting for infall toward the Virgo cluster, great
attractor and Shapley Supercluster, for a cosmology with $H_0$=70\ensuremath{\km \ps \pMpc\,}.
We therefore adopt this distance estimate for the group as a whole, which
gives an angular scale of 1\mbox{\arcm\arcm}=273~pc.
\section{HCG~16}
\label{sec:review}
HCG~16, also known as Arp~318, was originally identified \citep{Hickson82}
as a compact group of four spiral galaxies, NGC~833 (HCG~16B), NGC~835 (A),
NGC~838 (C) and NGC~839 (D), with later studies identifying a fifth large
spiral galaxy member (NGC~848) and a surrounding halo of dwarf galaxies
\citep[e.g.,][]{Ribeiroetal98}. Figure~\ref{fig:Xopt} includes a Digitized
Sky Survey image showing the relative positions of the group members. All
five major galaxies host AGN and/or starbursts
\citep[e.g.,][]{Martinezetal10,DeCarvalhoCoziol99,Continietal98}, and tidal
structures suggest an an ongoing or recent interaction between NGC~833 and
NGC~835 \citep{Konstantopoulosetal13}.
\begin{figure*}
\centerline{
\includegraphics[width=1.05\columnwidth,bb=36 175 576 616]{HCG16_5gals_DSS_HI.pdf
\includegraphics[width=1.05\columnwidth,bb=36 175 576 616]{allfive_0_5_2_smth_v4.pdf
}
\centerline{
\includegraphics[width=1.05\columnwidth,bb=36 175 576 616]{HCG16_VLA_on_DSS.pdf
\includegraphics[width=1.05\columnwidth,bb=36 175 576 616]{HCG16_HI_on_X.pdf
}
\caption{\label{fig:Xopt} \textit{Upper left:} Digitized Sky Survey 2 (DSS2) $R$-band image of the five largest galaxies in HCG~16, with the four galaxies originally identified as a compact group to the northwest. VLA \ensuremath{\mathrm{H}\textsc{i}}\ contours from Verdes-Montenegro et al. (2014, in prep.) are overlaid, with levels N(\ensuremath{\mathrm{H}\textsc{i}})$\simeq$10,20,40,65,85,110,140,160,200,250,350,450,570$\times$10$^{-19}$\ensuremath{\cm^{-2}}. \textit{Upper right:} Adaptively smoothed \emph{Chandra}\ 0.5-2~keV image using data from the S3 CCD in all five observations. Contours are overlaid in red to help elucidate the distribution of diffuse emission. Dashed ellipses indicate the \ensuremath{D_{25}}\ contours of the four main galaxies, white boxes the spectral extraction regions used to characterize the ridge of diffuse emission. \textit{Lower left:} VLA 1.4~GHz contours overlaid on a DSS2 image of the four main galaxies. Contours start at 3$\times$ the 0.2~mJy~beam$^{-1}$ rms noise level and increase in steps of factor 2. \textit{Lower right:} VLA \ensuremath{\mathrm{H}\textsc{i}}\ contours overlaid on the adaptively smoothed \emph{Chandra}\ 0.5-2~keV image. All four images have the same orientation.}
\end{figure*}
Neutral hydrogen mapping of the group revealed a $\sim$20\hbox{$^\prime$}\ long
complex filament of cold gas surrounding the four original members of the
group and linking them to NGC~848 \citep{VerdesMontenegroetal01}, almost
certainly as the result of tidal interactions between group members. The
total mass of \ensuremath{\mathrm{H}\textsc{i}}\ in the group is $>$2.63$\times$10$^{10}$\ensuremath{\mathrm{~M_{\odot}}}, and
Verdes-Montenegro et al. estimate that the group is $<$30\% \ensuremath{\mathrm{H}\textsc{i}}-deficient.
The four original member galaxies are $\sim$50-80\% deficient, while
NGC~848 is only $\sim$7\% deficient. This suggests that the majority of the
intergalactic \ensuremath{\mathrm{H}\textsc{i}}\ originated in the four main galaxies, perhaps being
transported out into the IGM by interactions among them, and then drawn
into its current morphology by a close passage of NGC~848.
\citet{Borthakuretal10} show that the \ensuremath{\mathrm{H}\textsc{i}}\ velocity distribution covers the
range $\sim$3650-4100\ensuremath{\km \ps}, confirming its association with the major
member galaxies.
\emph{ROSAT}\ studies of HCG~16 in the X-ray band were able to separate emission from the galaxies and diffuse inter-galactic gas, finding a gas temperature $\sim$0.3~keV and tracing emission out to $\sim$8\hbox{$^\prime$}\ \citep{Ponmanetal96,DosSantosMamon99}. \emph{ROSAT}\ imaging found the gas distribution to be irregular, with the brightest emission around the four original member galaxies and to the southeast of NGC~839, with fainter extension west or southwest of NGC~833 \citep{DosSantosMamon99}.
First light data from \emph{XMM-Newton}\ were used to investigate further the gas distribution, but this was hampered by a combination of uncertain calibration and scattered light from a bright background source close to the edge of the field of view. Nonetheless \citet{Belsoleetal03} reported a highly elliptical diffuse emission component surrounding the four main galaxies, with a temperature of $\sim$0.5~keV and abundance $\sim$0.07\ensuremath{\mathrm{~Z_{\odot}}}. A short (12.5~ks) \emph{Chandra}\ observation in cycle~1 provided no additional evidence of group-scale emission \citep{Jeltemaetal08}.
Optical spectroscopic studies of the starburst galaxies NGC~838 and NGC~839
have provided a detailed characterization of their stellar structures and
outflowing galactic winds. In NGC~839 the wind has formed a biconical polar
outflow \citep{Richetal10}, while in NGC~838 wind outflows have inflated
bubbles north and south of the galaxy, above and below the galactic disk
\citep{Vogtetal13}. These bubbles are likely confined by the surrounding
IGM and \ensuremath{\mathrm{H}\textsc{i}}, although there is some indication that material is leaking
from the southern bubble. Examination of optical spectra of the stellar
populations suggests that star formation peaked in NGC~838 and NGC~839 500
and 400~Myr ago respectively \citep{Vogtetal13,Richetal10} and still
continues in the core of each galaxy. Our own stellar population modelling
confirms that star formation is ongoing in NGC~838, but shows that star
formation in AGN-dominated NGC~833 has been minimal ($<$3\ensuremath{\mathrm{~M_{\odot}~yr^{-1}}}) over the
past few hundred Myr.
\section{Observations and Data Reduction}
\label{sec:obs}
Paper I describes our observations and reduction techniques in detail. We
used the five available \emph{Chandra}\ observations of HCG~16, all of which were
made with the ACIS-S3 CCD at the focal point. Three of the observations
(ObsIDs~\dataset[ADS/Sa.CXO#obs/15181]{15181}, \dataset[ADS/Sa.CXO#obs/15666]{15666} and \dataset[ADS/Sa.CXO#obs/15667]{15667}) totaling $\sim$75.7~ks, were made in 2013
July, using the full CCD and the same roll angle. The earliest
observation, ObsID~\dataset[ADS/Sa.CXO#obs/923]{923}, was made in 2000 Nov, for $\sim$12.5~ks. In 2008
Nov (ObsID~\dataset[ADS/Sa.CXO#obs/10394]{10394}) the group was observed in 1/2 subarray for $\sim$13.8~ks.
ObsID 923 and the 2013 observations all cover the four original member
galaxies, but ObsID 10394 only covers NGC~835, NGC~838 and part of NGC~833.
All five pointings were reduced using CIAO 4.6.1
\citep{Fruscioneetal06} and CALDB 4.5.9 following techniques similar to
those described in \citet{OSullivanetal07} and the \emph{Chandra}\ analysis
threads\footnote{http://asc.harvard.edu/ciao/threads/index.html}. Point
sources were identified using the \textsc{ciao} task \textsc{wavdetect},
and excluded. Spectra were extracted from each dataset using the
\textsc{specextract} task. Abundances were measured relative to the
abundance ratios of \citet{GrevesseSauval98}. 1$\sigma$ errors are reported
for all fitted values.
Reduction and analysis of the GMRT and VLA data was performed in the NRAO
Astronomical Image Processing System (\textsc{aips}) package following the
standard procedure (Fourier transform, clean and restore). Phase-only
self-calibration was applied to remove residual phase variations and
improve the quality of the images. The final VLA (GMRT) image has an
angular resolution of 25$\times$18.1\mbox{\arcm\arcm}\ (5.6$\times$5.4\mbox{\arcm\arcm}) and an rms
noise level (1$\sigma$) of 0.2 (0.06)~mJy beam$^{-1}$.
\section{Results}
We initially examined X-ray images of the group to determine the basic
structures associated with the galaxies and whether any large scale
emission was visible. Heavily smoothed or binned images showed evidence of
emission between the four galaxies on the ACIS-S3 CCD, located in a ridge
connecting the galaxies. Brighter diffuse emission was also visible in the
disks of several of the galaxies, between NGC~833 and NGC~835, and in
regions north and south of the disks of NGC~838 and NGC~839.
Figure~\ref{fig:Xopt} shows the core of the group, imaged in the optical,
\ensuremath{\mathrm{H}\textsc{i}}, 1.4~GHz radio continuum, and soft (0.5-2~keV) X-ray bands. The X-ray
image has been processed to remove point sources outside the galaxy cores
and refill the resulting holes using the \textsc{dmfilth} tool. The image
has then been adaptively smoothed using the \textsc{csmooth} task with
signal-to-noise limits of 3-5$\sigma$, and exposure corrected using a
0.91~keV exposure map smoothed to the same scales. The energy of the
exposure map was chosen to match the modal event energy in the
observations.
The ridge of faint diffuse emission linking the galaxies is clear, and
appears to extend past NGC~839 to the southeast, while surface brightness
declines to the southwest and northeast. The apparent brightness of
features close to the edges of the field may be affected by the exposure
correction and adaptive smoothing processes, and the surface brightness of
the diffuse emission is low. However, as mentioned above, heavily binned
images show the same basic ridge structure.
Comparison with the \ensuremath{\mathrm{H}\textsc{i}}\ map shows that the hot and cold gas structures are
similar, with the \ensuremath{\mathrm{H}\textsc{i}}\ filament overlapping the X-ray ridge over most of
its length. This is most obvious around NGC~838 and NGC839, the brightest
diffuse X-ray sources in the group, which appear to be embedded in some of
the highest density \ensuremath{\mathrm{H}\textsc{i}}\ gas. The extension of the X-ray ridge southeast of
NGC~839 also overlaps the section of the \ensuremath{\mathrm{H}\textsc{i}}\ filament extending toward
NGC~848. The \ensuremath{\mathrm{H}\textsc{i}}\ and X-ray emission agree less well between NGC838 and
NGC~835, with the X-ray ridge directly linking the two galaxies, while the
\ensuremath{\mathrm{H}\textsc{i}}\ filament curls to the north through a region of lower X-ray surface
brightness.
\subsection{Radio and X-ray imaging of the galactic superwinds}
\label{sec:winds}
We examined the galactic superwinds of NGC~838 and NGC~839 in detail in
Paper I, but as their outflows are possible contributors of hot gas to the
intra-group medium, we summarise the relevant results below, and then discuss
the wind morphology in more detail.
Our spectroscopic analysis of the X-ray and radio properties of the
galaxies confirms that NGC~833 and NGC~835 are dominated by emission from
absorbed AGN, while NGC~838 and NGC~839 are primarily starburst systems,
with only minimal AGN contribution to their X-ray luminosity. We estimate
the star-formation rate in NGC~838 (NGC~839) to be $\sim$7-17\ensuremath{\mathrm{~M_{\odot}~yr^{-1}}}\
($\sim$8-20\ensuremath{\mathrm{~M_{\odot}~yr^{-1}}}), with the lower, infra-red derived values probably
more representative of the current rate. The galactic winds have
temperatures $\sim$0.8~keV, with some evidence of a temperature decline in
the outer part of the southern bubble of NGC~838. We estimate the rate of
outflow in the winds to be 2.5\ensuremath{\mathrm{~M_{\odot}~yr^{-1}}}\ in NGC~839 and $\sim$17\ensuremath{\mathrm{~M_{\odot}~yr^{-1}}}\ in
NGC~838. However, since the NGC~838 wind appears to be largely confined, it is
unclear how much of this material escapes into the surrounding IGM.
Figures~\ref{fig:N838} and \ref{fig:N839} show X-ray images of the two galaxies. In both cases, a central area of X-ray emission, corresponding to the galaxy core, is visible. In the case of NGC~838 this is extended to the east and west in an ellipse. This includes emission from the northern wind bubble, and as the northern side of the galaxy disk is tilted toward the viewer, the galaxy centre is partly screened by this bubble and by gas and dust in the disk. More diffuse emission is clearly visible to the south, extending $\sim$25\mbox{\arcm\arcm}\ with a morphology suggestive of a bubble. Comparison with the H$\alpha$ imaging of the galaxy wind by \citet{Vogtetal13} confirms that the X-ray and optical morphology is similar, and that the southern bubble is viewed through the disk of the galaxy. Although the 610~MHz emission is dominated by the galaxy core, it is clearly extended to the south, coincident with the X-ray/optical bubble. Fainter diffuse X-ray and radio emission is visible north of the galaxy core, although interestingly both bands suggest that material outside the northern bubble is extended toward the northeast, rather than continuing the north-south axis of the bubbles.
\begin{figure}
\includegraphics[width=\columnwidth,bb=36 175 576 616]{NGC838_wind.pdf}
\caption{\label{fig:N838}\emph{Chandra}\ 0.5-2~keV exposure corrected image of
NGC~838, overlaid with GMRT 610~MHz contours. North is at the top of the
image, west to the right. The image has been smoothed with a
$\sim$2.5\mbox{\arcm\arcm}\ Gaussian. Contours begin at 180~$\mu$Jy~beam$^{-1}$
(3$\times$ the rms noise level) and increase in steps of factor 2. The
dashed green ellipse shows the approximate \ensuremath{D_{25}}\ contour of the galaxy
stellar light, and the scalebar indicates 30\mbox{\arcm\arcm}\ ($\sim$8~kpc).}
\end{figure}
\begin{figure}
\includegraphics[width=\columnwidth,bb=36 175 576 616]{NGC839_wind.pdf}
\caption{\label{fig:N839}\emph{Chandra}\ 0.5-2~keV exposure corrected image of
NGC~839. The image has been smoothed with a $\sim$2.5\mbox{\arcm\arcm}\ Gaussian. The
dashed green ellipse shows the approximate \ensuremath{D_{25}}\ contour of the galaxy
stellar light, and the scalebar indicates 30\mbox{\arcm\arcm}\ ($\sim$8~kpc).}
\end{figure}
NGC~839 is a point source at 610~MHz, dominated by the emission from its
dense star-forming core. This is also bright in the X-ray band, but it is
clear that gas emission extends north and south of the core, with a
morphology comparable to the biconical outflow observed in H$\alpha$
\citep{Richetal10}. However, there is also an indication of fainter X-ray
emission extending to the southeast. To test whether this apparent
extension is real, we measure the number of 0.5-2~keV counts in two annuli,
one with radius 4-15\mbox{\arcm\arcm}\ ($\sim$1-4~kpc) corresponding to the brighter conical wind
regions, one with radius 15-30\mbox{\arcm\arcm}\ ($\sim$4-8~kpc) corresponding to the
possible fainter emission. We break each annulus into 45\hbox{$^\circ$}\ sectors,
starting from north and proceeding anti-clockwise. Figure~\ref{fig:wedges}
shows the resulting azimuthal surface brightness measurements. It is clear
that in the inner annulus, emission is brightest in the 90\hbox{$^\circ$}\ sectors
centred on north and south, and faintest in the west-southwest. In the
outer annulus most of the sectors have comparable surface brightness
(though the western sectors are marginally fainter) but the south-southeast
sector is a factor $\sim$2 brighter than its neighbours (at 3$\sigma$
significance), confirming our identification of excess emission in this
direction in the image. Since the structure appears to be diffuse, it seems
likely that it is a more extended component of the galactic wind.
\begin{figure}
\includegraphics[width=\columnwidth,bb=30 210 570 740]{HCG16d_wedges_plot.pdf}
\caption{\label{fig:wedges}0.5-2~keV surface brightness around NGC~839 in the combined 2013 observations, measured in 45\hbox{$^\circ$}\ sectors of annuli with radii 4-15\mbox{\arcm\arcm}\ (black circles) and 15-30\mbox{\arcm\arcm}\ (grey triangles). Error bars indicate 1$\sigma$ uncertainties.}
\end{figure}
The 1.4~GHz radio continuum map (Figure~\ref{fig:Xopt}) shows unresolved
sources at the positions of NGC~838 and NGC~839, with low-surface brightness
diffuse emission to their south and east, extending $\sim$75\mbox{\arcm\arcm}\
($\sim$20~kpc) from NGC~838 and $\sim$110\mbox{\arcm\arcm}\ ($\sim$30~kpc) south of
NGC~839. There is no obvious source for this emission except relativistic
plasma ejected by the galactic superwinds of these two starburst galaxies,
or potentially by their AGN. Combined with the the X-ray and 610~MHz
imaging indicating that the winds bend to the east on scales of a few kpc,
this diffuse emission suggests that wind material transported out of the
galaxies is either driven eastward by the motion of the surrounding \ensuremath{\mathrm{H}\textsc{i}}\ or
IGM, or is left behind as the galaxies move westward. We will return to the
question of interaction between the winds and the surrounding environment
in Section~\ref{sec:IGM}.
\subsection{Group-scale diffuse X-ray emission}
To investigate the properties of the diffuse intergalactic emission, we
first exclude regions around the galaxies to avoid contamination. For
NGC~838 and NGC~839 we used $\sim$25\mbox{\arcm\arcm}\ radius circular regions, based
on a curve-of-growth analysis designed to ensure that 95\% of
the 0.5-2~keV flux from the galaxies is excluded. Since NGC~833 and NGC~835
are interacting, with much of their gas content between the two galaxies,
and some emission from a tidal arm extending east from NGC~835, we used a
polygonal region approximating the extent of the stellar component of the
galaxies. These regions are described in more detail in Paper I. Having
excluded the galaxies, we divided the remainder of the ACIS-S3 field of
view into several large regions, based on the adaptively smoothed contours.
The upper right panel of figure~\ref{fig:Xopt} shows the rectangular North
and South Ridge regions. We defined the remainder of the S3 field
of view as the Outer region.
The emission in all these regions is spectrally soft and extremely faint,
and probably fills the field of view. Since blank-sky fields scaled to
match the particle flux of our data may over- or under-subtract the soft
galactic foreground emission, we chose to model the background for these
regions following a method similar to that used by \citet{Snowdenetal04}
for \emph{XMM}\ data. We fit the regions simultaneously to provide the maximum
constraint on the background model. Since the observations prior to 2013
are relatively short and have different fields of view, we only use spectra
from ObsIDs 15181, 15666 and 15667, so as to simplify the model and avoid
any uncertainties associated with changes in effective sensitivity over the
life of the ACIS instrument.
\begin{deluxetable*}{lcccccc}
\tablewidth{0pt}
\tablecaption{\label{tab:diff}Best-fitting model parameters for the diffuse low-surface brightness emission}
\tablehead{
\colhead{Region} & \colhead{Background} & \colhead{kT} & \colhead{Abund.} & \colhead{$L_{0.5-7}$} & \colhead{Surface Brightness} & \colhead{red. $\chi^2$/d.o.f.} \\
\colhead{} & \colhead{} & \colhead{(keV)} & \colhead{(\ensuremath{\mathrm{~Z_{\odot}}})} & \colhead{(10$^{38}$\ensuremath{\erg \ps})} & \colhead{(10$^{38}$\ensuremath{\erg \ps}~arcmin$^{-2}$)} & \colhead{}\\
}
\startdata
North Ridge & Model & 0.27$\pm$0.03 & 0.05$^{+0.07}_{-0.03}$ & 94.72$^{+73.72}_{-48.13}$ & 8.63$^{+6.72}_{-4.39}$ & 1.31/2584$^\dagger$\\[+0.5mm]
& Outer & 0.24$\pm$0.03 & 0.3$^*$ & 58.06$\pm$1.15 & 5.27$\pm$0.12 & 1.07/253 \\[+0.5mm]
& Outer & 0.24$^{+0.03}_{-0.02}$ & $>$0.15 & 53.86$^{+1.53}_{-0.76}$ & 2.72$^{+0.15}_{-0.08}$ & 1.08/252 \\[+0.5mm]
South Ridge & Model & 0.34$^{+0.06}_{-0.03}$ & 0.05$^{+0.07}_{-0.03}$ & 190.21$^{+115.35}_{-95.49}$ & 13.18$^{+8.02}_{-6.61}$ & 1.31/2584$^\dagger$\\[+0.5mm]
& Outer & 0.45$^{+0.13}_{-0.09}$ & 0.3$^*$ & 77.54$\pm$1.15 & 5.35$\pm$0.12 & 1.11/345 \\[+0.5mm]
& Outer & 0.50$^{+0.15}_{-0.13}$ & 0.04$^{+0.05}_{-0.03}$ & 92.81$^\pm$1.15 & 6.45$\pm$0.12 & 1.10/344 \\[+0.5mm]
Outer & Model & 0.30$^{+0.07}_{-0.05}$ & 0.05$^{+0.07}_{-0.03}$ & 301.36$^{+290.66}_{-176.46}$ & 6.38$^{+6.15}_{-3.74}$ & 1.31/2584$^\dagger$
\enddata
\tablecomments{$^*$ Parameter fixed during fitting.\\
$^\dagger$ Fit statistic for simultaneous fit to all regions including background model.}
\end{deluxetable*}
Our spectral model consists of several components: 1) A broken power law
representing high-energy particles, which is convolved with the instrument
Response Matrix File (RMF) but not the Auxiliary Response File (ARF) since
the particles are to first order unaffected by the X-ray mirrors; 2)
Gaussians representing fluorescent emission lines within the detector, the Si K$\alpha$ and Au M$\alpha\beta$ in the case of the S3 CCD; 3) A powerlaw with $\Gamma$=1.46 and
initial normalization 8.88$\times$10$^{-7}$ per square arcminute,
representing the cosmic hard X--ray background; 4) Three thermal models
representing emission from the local hot bubble and the Galactic halo, two with
temperature 0.1~keV (one with Galactic absorption and one without) and the
third with an initial temperature of 0.25~keV; 5) A source component
consisting of an APEC thermal plasma model at the systemic redshift
($z$=0.0132) with fixed Galactic absorption. The normalisations of
components 1, 2 and 5 are allowed to fit independently for each region, but
normalisations for components 3 and 4 are linked across the regions,
scaling for area.
In addition to the \emph{Chandra}\ spectra we also fit components 3 and 4 (the
X-ray foreground and background emission) to a \emph{ROSAT}\ All-Sky Survey
spectrum extracted from an annulus between 0.5-0.75\hbox{$^\circ$}\ from the group
centroid (490-740~kpc at our adopted distance). This provides additional
constraints on the soft emission, particularly useful given the low
temperature of the group emission. Fits are carried out in the 0.5-10~keV
band, the inclusion of 7-10~keV emission helping to constrain the particle
background component in the \emph{Chandra}\ data.
Table~\ref{tab:diff} shows our results. Our fits suggest the presence of weak thermal emission throughout the S3 field of view. For the set of three regions (North and South Ridge plus Outer) we experimented with thawing the normalization of the cosmic hard background component and the temperature of the 0.25~keV Galactic soft foreground. The former falls 22\% below its initial value, which we consider acceptable as we have excluded a number of bright background point sources. The latter is poorly constrained when fitted, and we therefore fix it at 0.25~keV in our final fit, which has reduced $\chi^2$=1.31 for 2584 degrees of freedom.
As a test of the reliability of the background modelling approach, we
also fitted the spectra of the north and south ridge regions using the
spectrum of the outer region as the background. We fit each region with a
simple absorbed APEC model, and tried fits with abundance free to vary or
fixed at 0.3\ensuremath{\mathrm{~Z_{\odot}}}. The results are listed in Table~\ref{tab:diff}. In the
north ridge the temperatures agree within the 1$\sigma$ errors with our
background modelling fit. The agreement in the south ridge is poorer, but
still at the 2$\sigma$ level. As expected, the fluxes are systematically
lower than those found from the background modelling approach, suggesting
that the background is over-subtracted owing to the presence of source flux
in the outer region. For the north ridge the abundance is poorly constrained
when fitted, owing to the small number of net counts in the spectrum after
background subtraction. In the south ridge abundance is constrained to
0.01-0.09\ensuremath{\mathrm{~Z_{\odot}}}, but fixing it at 0.3\ensuremath{\mathrm{~Z_{\odot}}}\ only makes the fit slightly
poorer. In general, these fits confirm the accuracy of our background
modelling approach.
To search for evidence of any more extended emission, we extracted spectra
from the S2 CCD for all three 2013 observations. The S4 CCD was not active
during the 2013 observations. S2 covers the area immediately
south-southeast of the group, but it is a front illuminated CCD and
therefore less sensitive to spectrally soft emission than S3. After
examining a heavily smoothed image, we elected to use a
$\sim$5.8$\times$8\hbox{$^\prime$}\ region excluding strips of the CCD close to the
edge of the ACIS-S array. These regions include most of the point sources
most strongly blurred by the point spread function (PSF), and could be
affected by any imperfections in the calibration of absorption by the
contaminant which has built up on the ACIS optical filters.
We excluded point sources and used the same fitting approach described
above, with additional Gaussian components to fit the Ni K$\alpha$ and Au
L$\alpha$ fluorescent emission lines which are visible on this CCD. We
found that the background model did a good job of describing the spectrum,
with no residual features indicative of source emission. When we added a
thermal model (with abundance fixed at 0.3\ensuremath{\mathrm{~Z_{\odot}}}) to represent diffuse IGM
emission, we found that with temperature free to vary, the best fit had
kT$<$0.11 (1$\sigma$ limit) and an extremely poorly constrained
normalization. Fixing kT at 0.3~keV we found that the normalization was
consistent with zero, with a 1$\sigma$ upper limit on surface brightness of
1.36$\times$10$^{38}$\ensuremath{\erg \ps}~arcmin$^{-2}$. These are very weak constraints,
and would not rule out emission consistent with the extended IGM detected
in the ``Outer'' region on S3. They do however suggest that the bright
ridge seen around the galaxies does not extend any significant distance
into the region covered by S2.
\section{Discussion}
Smoothed images of the group show extended low surface-brightness emission linking the four main galaxies and probably extending southeast beyond NGC~839. Our spectral fitting suggests that this gas has a low abundance and temperature $\sim$0.3~keV, in agreement with previous \emph{ROSAT}\ observations \citep{Ponmanetal96,DosSantosMamon99}. This temperature is somewhat lower than that reported from \emph{XMM-Newton}\ \citep{Belsoleetal03} except in the south ridge, where the two are comparable within errors, but this is unsurprising given the uncertain calibration of that dataset. Our low abundance of Z=0.05$^{+0.07}_{-0.03}$\ensuremath{\mathrm{~Z_{\odot}}}\ agrees well with the \emph{ROSAT}\ and \emph{XMM-Newton}\ measurements \citep{DosSantosMamon99,Belsoleetal03}.
Our \emph{Chandra}\ observations seem to agree better with the \emph{ROSAT}\ than \emph{XMM-Newton}\
in terms of the morphology of the diffuse gas. \citet{DosSantosMamon99}
found the diffuse emission to be clumpy and filamentary, while
\citet{Belsoleetal03} were able to model the diffuse emission as a smooth,
if highly elliptical, $\beta$-model. Belsole et al. argued that the
superior collecting area and smaller PSF of \emph{XMM}\ allowed them to better
resolve and excise point sources, and that part of the clumpiness of the
\emph{ROSAT}\ image arose from point source contamination. The specific example
they raised, the C4 region identified by Dos Santos \& Mamon southwest of
NGC~833, is outside the \emph{Chandra}\ field of view, so a direct comparison is
not possible. However, we find that the distribution of diffuse emission in
the S3 field of view is not consistent with a $\beta$-model. Even if
extreme ellipticities are allowed, the model overestimates the flux in the
region between NGC~835 and NGC~838, while underestimating the flux at the
ends of the ridge, particularly in the southeast corner of the field. The
curve of the ridge and the lack of a clear decline in surface brightness
southeast of NGC~839 both suggest that the gas distribution is not a simple
ellipsoid. The disagreement with the Belsole et al. model is perhaps
unsurprising, since imaging analysis of the relatively shallow first light
\emph{XMM}\ data presented a number of difficulties, including high electronic
noise, the uncertain calibration of the EPIC-MOS cameras and the need to
exclude the then-uncalibrated EPIC-pn, scattered light from bright sources
in the field, and the difficulty of modelling the various components of the
background. We are fortunate in having deep observations from a mature,
well-calibrated instrument with superb spatial resolution.
\subsection{Physical properties of the hot intra-group medium}
\label{sec:IGM}
We can estimate the properties of the gas in the northern and southern
parts of the ridge from the results of our spectral fits, assuming a
cylindrical geometry. We approximate the north ridge as a cylinder of
length 4.3\hbox{$^\prime$}\ (70.4~kpc) and radius 1.6\hbox{$^\prime$}\ (25.9~kpc), and for the
south ridge use a length of 6.35\hbox{$^\prime$}\ (104~kpc) and radius 1.2\hbox{$^\prime$}\
(20.1~kpc). Based on the normalization of our background modelling fits to
the diffuse emission, we find that the electron number density of the
diffuse gas is 1.16$^{+1.03}_{-0.83}$$\times$10$^{-3}$\ensuremath{\cm^{-3}}\ and
1.16$^{+0.91}_{-0.83}$$\times$10$^{-3}$\ensuremath{\cm^{-3}}\ for the north and south
ridge regions respectively, neglecting any uncertainties in volume. The
bolometric luminosities for the two regions are $L_{X,{\rm
bolo}}$=1.68$^{+1.31}_{-0.86}\times$10$^{40}$\ensuremath{\erg \ps}\ (north ridge) and
$L_{X,{\rm bolo}}$=3.08$^{+1.87}_{-1.55}\times$10$^{40}$\ensuremath{\erg \ps}\ (south).
The gas pressure in these regions is
$\sim$5-6.5$\times$10$^{-13}$~dyne~cm$^{-2}$ (equivalent to \ensuremath{\erg \pcmcu}), entropy is $\sim$25-30 keV\ensuremath{\cm^{-2}}\
and the isobaric cooling times are $\sim$7-10~Gyr. While the low gas
entropy is comparable to that found in the cool cores of relaxed, X-ray
bright groups \citep{Sunetal09}, the cooling times are long, suggesting
that radiative cooling will probably play only a minor role in the future
development of the gas.
The mass of gas in the two regions, excluding the denser material in and
around the galaxies, is 3.5$^{+3.1}_{-2.5}$$\times$10$^9$\ensuremath{\mathrm{~M_{\odot}}}\ for the
north ridge and 3.1$^{+2.4}_{-2.2}$$\times$10$^9$\ensuremath{\mathrm{~M_{\odot}}}\ for the south
ridge. This is a significant mass of gas, comparable to the overall \ensuremath{\mathrm{H}\textsc{i}}\
deficiency estimated for the group, $\lesssim$10$^{10}$\ensuremath{\mathrm{~M_{\odot}}}\
\citep{VerdesMontenegroetal01}. However, the spectral fits show that IGM
emission extends outside these regions, and probably outside the field of
view. Estimating the total gas mass in the system requires a model of
surface brightness. Since a 2-dimensional $\beta$-model is a poor fit to
the diffuse emission, we modelled the 1-dimensional surface-brightness
profile across the ridge. We measured the 0.5-2~keV surface brightness
using 20$\times$600\mbox{\arcm\arcm}\ boxes aligned at an angle of 30\hbox{$^\circ$}\ (measured
anti-clockwise from west) with their long axes along the ridge. Again, the
regions used to extract spectra for the galaxies were excluded, as were
point sources outside the galaxies.The resulting profile extend from the
southwest (lower right) corner of the S3 CCD to the NE (upper right)
corner, and is shown in Figure~\ref{fig:SB}.
\begin{figure}
\includegraphics[width=\columnwidth,bb=20 205 565 745]{profile_SB.pdf}
\caption{\label{fig:SB}One dimensional exposure corrected 0.5-2~keV surface brightness profile running from southwest to northeast across the ridge of diffuse emission in HCG~16. The solid line shows the best-fitting $\beta$-model plus constant background.}
\end{figure}
We modelled the exposure corrected profile with a constant background component plus a $\beta$-model. The best-fitting model is extremely flat, with core radius $R_c$=12.03\mbox{\arcm\arcm}$^{+57.44}_{-13.04}$ (3.28$^{+15.68}_{-3.56}$~kpc) and $\beta$=0.181$^{+0.26}_{-0.01}$. This is probably unphysical; the 1$\sigma$ upper limit on $\beta$ is more comparable to other cool groups \citep[e.g.,][]{Mulchaeyetal03}. We therefore emphasize that, while we have used this model to estimate the gas mass and luminosity of the group at large radius, the results cannot be considered as reliable measurements.
To determine how far to extrapolate this model, we must make some
assumptions about the structure of the gas, based on its likely origin. If
the gas is a virialized IGM, bound within the gravitational well of the
group, we can estimate a typical scale size for a system of this
temperature from measured scaling relations. We use the mass-temperature
relation of \citet{Sunetal09}, which was derived from a collection of
groups and clusters observed with \emph{Chandra}.
For a group with kT=0.3~keV we estimate that
M$_{500}$=4.0$\times$10$^{12}$\ensuremath{\mathrm{~M_{\odot}}} and $R_{500}$=240~kpc
($\sim$14.6\hbox{$^\prime$}). Taking a mean radius of 80\mbox{\arcm\arcm}\ for the two cylinders
used to model the north and south ridges, we find that extrapolating to
14.6\hbox{$^\prime$}\ would increase the gas mass by a factor $\sim$33.9 for
$\beta$=0.181 or a factor 6.2 if $\beta$ is at the 1$\sigma$ upper limit
value of 0.45. These values assume a cylindrical geometry with length
$\sim$10\hbox{$^\prime$}, so an additional factor $\sim$2 should be added if the IGM
is ellipsoidal. This gives a hot gas mass of $M_{\rm
gas}$(R$<$$R_{500}$)$\sim$0.8-4.5$\times$10$^{11}$\ensuremath{\mathrm{~M_{\odot}}}.
If the hot gas is associated with, and has an extent similar to, the \ensuremath{\mathrm{H}\textsc{i}}\
filament that links the group galaxies, it is likely to be considerably
smaller, with a minor axis radius of only 4-5\hbox{$^\prime$}\ and a major axis
$\sim$20\hbox{$^\prime$}. In this case we expect an extrapolation factor of
$\sim$8-14, for a total hot gas mass of 5.2-9.4$\times$10$^{10}$\ensuremath{\mathrm{~M_{\odot}}},
2-3.5 times the total \ensuremath{\mathrm{H}\textsc{i}}\ mass in the system.
A similar process of extrapolation is required to estimate the total
luminosity of the IGM. Assuming emission to be proportional to density
squared (i.e., neglecting any temperature or abundance variation) we find
that the uncertainties in the surface brightness model lead to very large
uncertainties in the total luminosity for the case of an ellipsoidal
virialized IGM, $L_{X,{\rm
bolo}}$(R<$R_{500}$)$\simeq$1.1-8.6$\times$10$^{41}$\ensuremath{\erg \ps}. For a
filamentary distribution similar to that of the \ensuremath{\mathrm{H}\textsc{i}}, $L_{X,{\rm
bolo}}\simeq$2.5-6.1$\times$10$^{41}$. The lower end of these estimates
are comparable to the measured luminosity in the S3 field of view without
any extrapolation, $L_{X,{\rm
bolo}}$=1.87$^{+1.03}_{-0.66}$$\times$10$^{41}$\ensuremath{\erg \ps}.
\begin{figure}
\includegraphics[width=\columnwidth,bb=40 300 560 740]{LTplots.pdf}
\caption{\label{fig:LT}Luminosity and temperature for the \citet{OsmondPonman04} sample of galaxy groups, with estimates for HCG~16 overlaid. Black squares indicate the Osmond \& Ponman G sample of bona-fide groups, grey circles their H sample of groups with only galaxy-scale X-ray emission. Open circles indicate H sample systems containing only spirals. The diamonds show our own estimates using no extrapolation (red), extrapolation assuming a filamentary distribution (green) or a virialized halo (blue). Open, cyan points represent estimates by \citet[triangle]{Belsoleetal03}, \citet[star]{DosSantosMamon99} and Osmond \& Ponman (square). The dashed line shows the measured luminosity-temperature relation for groups and clusters from \citet{Eckmilleretal11}, a study based on \emph{Chandra}\ data for systems with kT$>$0.5~keV.}
\end{figure}
Figure~\ref{fig:LT} shows a comparison of our luminosity estimates with the luminosity-temperature distribution of a sample of X-ray bright groups from \citet{OsmondPonman04}. Their sample was divided into confirmed groups with diffuse X-ray emission extending $>$65~kpc (their G subsample, black squares) and less luminous systems which only host galaxy-scale extended emission (H sample, grey circle). Apart from HCG~16, none of their G groups were spiral-only systems. The H sample did include four spiral-only systems, which fall on the lower edge of the luminosity-temperature (L-T) distribution. All three of our estimates fall within the scatter of the distribution, which is large at low temperatures, though the upper bound of our luminosity estimate assuming a fully virialized halo is at the upper edge.
For comparison, we also show previous estimates of the luminosity and temperature of HCG~16, corrected to our adopted distance. However, we note that these values are not truly comparable, owing to the different apertures used and different approaches to exclusion of galaxy emission and point sources. The \emph{XMM}\ estimate of \citet{Belsoleetal03} differs from ours by the largest factor, with a marginally higher temperature and significantly lower luminosity. The \emph{ROSAT}\ estimate of \citet{DosSantosMamon99}, using their fit with abundance fixed at 0.1\ensuremath{\mathrm{~Z_{\odot}}}, is in agreement with our unextrapolated luminosity estimate, while the Osmond \& Ponman estimate is comparable to all three of our luminosity estimates. In summary, our \emph{Chandra}\ measurements are consistent with the measured luminosity-temperature relation for groups, within the large observed scatter.
\subsection{The origin of the hot IGM}
Three possible sources of the hot gas in the intra-group medium of HCG~16
can be suggested: 1) Primordial gas which has fallen into the group
potential and been shock heated to the virial temperature; 2) Shock-heated
\ensuremath{\mathrm{H}\textsc{i}}\ from the large-scale filament, heated by a high-speed collision as in
Stephan's Quintet; 3) Material ejected from the member galaxies by galactic
winds during phases of intense star formation. Perhaps the most notable
feature of the IGM is the partial correlation between the positions of the
hot and cold gas in the group, with the X-ray and \ensuremath{\mathrm{H}\textsc{i}}\ emission co-located
around the starburst galaxies NGC~838 and NGC~839. This, combined with the
ridge morphology of the hot gas, strongly suggests a connection between the
two different gas phases, at least in the group core.
In most relaxed, X-ray bright groups, the IGM produces an elliptical or
circular surface brightness distribution \citep[see
e.g.,][]{Mulchaeyetal03}. Ridge-like structures linking galaxies and
superimposed on an underlying ellipsoidal structure are not unknown
\citep[e.g., the NGC~5171 group,][]{Osmondetal04} but are uncommon. The
only other spiral-only group detected in the X-ray band, SCG0018-4854, is
too X-ray faint to allow a detailed morphological study
\citep{Trinchierietal08}, but Stephan's Quintet, whose galaxy population is
only slightly more evolved than HCG~16, has an apparently relaxed
large-scale IGM outside the complex group core
\citep{Trinchierietal05,OSullivanetal09}. The S3 field of view is too small
to allow us to determine whether the large-scale X-ray emission outside the
ridge is relaxed or not, but the failure to fit a $\beta$-model to the
ridge rules out a simple ellipsoidal model such as that used by
\citet{Belsoleetal03}. Certainly the ridge is the dominant component of
the IGM in the \emph{Chandra}\ field of view, and the distribution of emission
outside the ridge is very flat.
We can consider the stability and likely lifespan of the ridge structure as
a constraint on its likely origin. Taking a radius of 25~kpc and a typical
temperature of 0.3~keV, the sound crossing time of the ridge is $\sim$112~Myr.
The ridge has a higher density (and therefore pressure) than the
surrounding IGM, and should therefore expand and disperse over the course
of a few hundred Myr, unless it is somehow confined. An extended
filamentary structure like the ridge could not be formed by the relaxed
gravitational potential of a virialized group, but might be temporarily
formed by the close association of the group member galaxies, with their
individual dark matter halos helping to retain some of the gas around the
galaxies. The lifetime of the southern ridge might be extended by the ongoing
starburst winds from NGC~838 and NGC~839, which could inject new
higher-density gas as older material expands outward. The southern ridge
has a surface brightness a factor 1.5 greater than the northern ridge, and
starburst winds may explain this difference. The north ridge has only marginally higher surface brightness than its surroundings (8.63$^{+6.71}_{-4.39}$$\times$10$^{38}$\ensuremath{\erg \ps}~arcmin$^{-2}$ compared to 6.38$^{+6.15}_{-3.74}$$\times$10$^{38}$\ensuremath{\erg \ps}~arcmin$^{-2}$ for the IGM) and this may be related to the lack of recent star formation in NGC~833 and NGC~835.
Another possibility is that our spectral fits to the diffuse X-ray emission
are misleading, and that the ridge is actually a region of enhanced
abundance rather than enhanced density. Galaxy winds could have enriched
the IGM around the galaxies, producing higher surface brightness through
enhanced line emission. Quite small increases in abundance can
significantly increase surface brightness at the low temperatures observed
in the group, and doubling the observed 0.05\ensuremath{\mathrm{~Z_{\odot}}}\ abundance could explain
the factor $\sim$1.3 change in mean surface brightness across the ridge.
This scenario is again consistent with a higher surface brightness in the
southern ridge, since its enrichment is ongoing. However, it is difficult
to see how the ridge can be a product of enrichment alone; the same galaxy
winds which transport heavy elements out into the ridge also bring higher
density gas. It seems likely that all of these possibilities contribute,
and that the ridge is a temporary structure formed by the gravitational
interaction of the major group members, with starburst winds helping to
boost its surface brightness, extent and lifetime.
\subsubsection{A virialized halo}
The estimated gas mass in the IGM for a fully virialized 0.3~keV system
implies a (hot) gas mass fraction within $R_{500}$ of $f_{\rm gas}^{\rm
hot}$=0.02-0.11, a stellar fraction of $f_{*}$=0.08, and a baryon
fraction of $f_{\rm baryon}>$0.11-0.20, taking
$M_{500}$=4$\times$10$^{12}$\ensuremath{\mathrm{~M_{\odot}}}\ from the M:T relation of
\citet{Arnaudetal05} and neglecting the contribution of dwarf galaxies and
any intergalactic stellar component. The lower bound of this range is just
comparable with the upper limit of scatter seen in the lowest mass groups
and poor clusters for which accurate measurements have been made
\citep[e.g.,][]{Sandersonetal13,Gonzalezetal13}, while the upper bound
exceeds the universal baryon fraction. We also note that the Sanderson et
al. and Gonzalez et al. baryon fraction measurements include a significant
contribution from intra-cluster stars, which our estimate neglects.
Including this component in HCG~16 would increase the baryon fraction,
making the agreement with other groups even poorer. Our alternative
estimate, assuming the hot gas distribution is comparable to that of the
\ensuremath{\mathrm{H}\textsc{i}}, would give a somewhat more reasonable $f_{\rm baryon}>$0.12-0.13, again neglecting intra-cluster stars.
Similarly, the stellar fraction is $\sim$50\% of the universal baryon
fraction, considerably higher than observed in more massive groups and
clusters \citep{Sandersonetal13,Gonzalezetal13}. Stellar fraction is
expected to peak in systems with mass M$_{500}\sim$10$^{12}$\ensuremath{\mathrm{~M_{\odot}}}\
\citep{Leauthaudetal12}, but $f_{*}\sim$0.5 would place HCG~16 at the
extreme upper edge of the likely range.
It therefore seems unlikely that the IGM of HCG~16 (or at least that part
of it we can observe) is relaxed, or that it has formed entirely through
gravitational infall. The \emph{Chandra}\ results thus support the conclusions
drawn from the \emph{ROSAT}\ data by \citet{Ponmanetal96} and
\citet{DosSantosMamon99}. This raises the question of why the group falls
on or above the L-T relation. We might expect that the IGM in a virialised
group would be hotter than in one that has yet to virialize,
since the fully collapsed system is denser and better able to compress the
gas. However, if a large fraction of the gas is heated by non-gravitational
processes such as star formation or shocks, its temperature will depend on
the balance between those heating processes and losses from radiative
cooling and gas mixing.
\subsubsection{Shock heating}
Shock heating of the \ensuremath{\mathrm{H}\textsc{i}}, as observed in Stephan's Quintet, would require a
high velocity collision. The line-of-sight velocities of the five major
galaxies in HCG~16 cover a range of only 227\ensuremath{\km \ps}, with the largest
difference, between NGC~833 and NGC~835, probably arising from their
interaction. However, if we consider that the age of the starbursts in
NGC~838 and NGC~839 are likely to correspond to tidal encounters with
NGC~848, we can estimate the velocity of NGC~848 in the plane of the sky to
be 455-475\ensuremath{\km \ps}. A collision at this velocity would produce a strong shock
in a 100K \ensuremath{\mathrm{H}\textsc{i}}\ cloud; a head-on collision would raise the temperature of
the \ensuremath{\mathrm{H}\textsc{i}}\ to $\sim$0.42~keV ($\sim$5$\times$10$^6$K). This is quite similar
to the observed temperature of the IGM, suggesting that if shock heating is
the primary source of gas, cooling must have had little impact since the
shock occurred.
NGC~848 would also have caused a weak shock in any hot gas it encountered
while passing through the group core. A temperature of 0.3~keV implies a
sound speed of $\sim$220\ensuremath{\km \ps}\ in the IGM, so NGC~848 would have produced a
Mach$\sim$2.1 shock, a temperature increase of factor $\sim$2.2 and a
density increase of factor $\sim$2.4. A weak shock would therefore produce
a large increase, of a factor $\sim$10, in the bolometric X-ray luminosity
of any hot gas in the system. The current data are not capable of
differentiating between a ridge primarily formed from shocked \ensuremath{\mathrm{H}\textsc{i}}, and one
with a significant contribution from shocked hot gas, but the effects of
such shocks in compact groups would make an interesting subject for
investigation with future numerical simulations.
Only the S1 CCD of ObsID 923 covers NGC~848, where any shock front would
currently be located, and the short exposure of the observation and large
off-axis angle mean that it lacks the depth and spatial resolution to
detect such a feature. None of the other indicators of a shock observed in
Stephan's Quintet (optical and infra-red line emission, radio continuum
emission, star formation outside the major galaxies) are present in HCG~16,
though most of them are short lived and would likely have faded over the
$\sim$500~Myr since NGC~848 passed the other galaxies. The large mass of
\ensuremath{\mathrm{H}\textsc{i}}\ observed in NGC~848 and in the large-scale filament strongly suggests
that shock heating can only have affected a small fraction of the gas. We
can place a limit on the mass of \ensuremath{\mathrm{H}\textsc{i}}\ that could have been shock heated of
$\lesssim$1.1$\times$10$^{10}$\ensuremath{\mathrm{~M_{\odot}}}, based on the \ensuremath{\mathrm{H}\textsc{i}}\ deficiency of the
group \citep{VerdesMontenegroetal01}. At best, shock heated \ensuremath{\mathrm{H}\textsc{i}}\ may
therefore account for $\sim$20\% of the IGM.
\subsubsection{Galaxy winds}
The third possible source of gas is the star formation driven galactic
winds in the member galaxies. Our observations of NGC~838 and NGC~839 show
that the winds of these starburst galaxies have temperatures of
0.8-0.9~keV, with some indication of a temperature decline in their outer
parts (see Paper I). Although the radiative cooling times of the hot gas in
the winds are $>$1~Gyr, we might expect more rapid cooling to be caused by
mixing with cold gas entrained in the winds or encountered as wind material
interacts with the \ensuremath{\mathrm{H}\textsc{i}}\ filament. This might explain the reduction in
temperature by a factor of 3 from the galaxy winds to the IGM. It might
also explain the marginally higher temperature of the diffuse gas in the
south ridge (0.34~keV) where the two starburst galaxies reside, compared to the north ridge, which is occupied by AGN-dominated systems (0.27~keV).
Our estimate of the mass of gas which may have been ejected by NGC~838 and
NGC~839 (9.5$\times$10$^9$\ensuremath{\mathrm{~M_{\odot}}}) exceeds the mass of hot gas in the
southern ridge by a factor $\sim$3, suggesting that it could be largely
formed from ejected wind material. For the northern ridge to have formed in
a similar way, the star formation rates of NGC~833 and NGC~835 would have
to have been significantly greater in the past. Given the limits on recent
star formation from our stellar population modelling of NGC~833 (see paper
I) their starburst periods would need to be significantly older, perhaps
triggered by an initial tidal encounter between the two galaxies. From the
velocity estimated above, NGC~848 seems likely to have passed NGC~833/835
$\sim$600~Myr ago, and a starburst triggered then would be clearly detected
in the SDSS spectrum of NGC~833. If we assume that NGC~833 and NGC~835 did
go through a superwind phase, then a total of
$\sim$2$\times$10$^{10}$\ensuremath{\mathrm{~M_{\odot}}}\ of hot gas would have been produced by the
four original group members, $\sim$20-40\% of our estimated IGM mass. For a
Salpeter initial mass function, stellar populations are expected to lose
$\sim$30\% of their mass over a Hubble time \citep{White91,Davidetal91}.
For the four original members this puts a strong upper limit of
$\sim$9$\times$10$^{10}$\ensuremath{\mathrm{~M_{\odot}}}\ of hot gas which could be produced from the
galaxies, roughly five times the amount we expect to have been ejected by
stellar winds.
If wind material does make up a significant fraction of the IGM, we might
expect to see higher abundances. The abundance measured in NGC~838 is only
$\sim$0.16\ensuremath{\mathrm{~Z_{\odot}}}, but is probably biased low because even the hot phase of
the wind is multi-temperature and because of mass loading; as the enriched
hot gas produced by stellar winds and supernovae flows out of the galaxy it
entrains less enriched neutral hydrogen which effectively dilutes the
metallicity. This is likely to be the case in the IGM as well since we
observe \ensuremath{\mathrm{H}\textsc{i}}, X-ray and radio continuum emission from the same regions,
indicating that the IGM is multi-phase.
There is also the question of whether hot wind material can escape the
immediate neighbourhood of the starburst galaxies and diffuse out through
the surrounding \ensuremath{\mathrm{H}\textsc{i}}. The diffuse radio continuum emission demonstrates that
gas from the winds can reach distances of 20-30~kpc, and the X-ray and
radio morphology of the winds suggests that while they may be affected by
the motion of the \ensuremath{\mathrm{H}\textsc{i}}\ relative to the galaxies, they are not completely
confined by it. It seems likely that the large-scale \ensuremath{\mathrm{H}\textsc{i}}\ filament is
actually a complex of smaller structures, interspersed with hotter
material. While some mixing must take place, resulting in heating of the
\ensuremath{\mathrm{H}\textsc{i}}\ and cooling of the hot X-ray emitting gas, the presence of the X-ray
ridge demonstrates that this is not an efficient process. Hot gas clearly
coexists with the \ensuremath{\mathrm{H}\textsc{i}}\ over significant timescales.
\citep{Borthakuretal10} have estimated that \ensuremath{\mathrm{H}\textsc{i}}\ clouds of radius
$\geq$200~pc (0.73\mbox{\arcm\arcm}) can survive for a few hundred Myr in a hot IGM, so
this is not implausible.
We can question whether the energy available for star formation is
sufficient to heat this mass of gas to its current temperature. The total
energy of the gas in the north and south ridges is
$\sim$4$\times$10$^{57}$~erg. The number of core collapse supernovae
expected per unit star formation is 0.01-0.015 \citep{Botticellaetal12},
and we assume the standard value for the energy available from each
supernova, 10$^{51}$~erg. This suggests that over the 500~Myr (400~Myr)
timescale of star formation in NGC~838 (NGC~839), a star formation rate of
0.25-0.35\ensuremath{\mathrm{~M_{\odot}~yr^{-1}}}\ (0.35-0.5\ensuremath{\mathrm{~M_{\odot}~yr^{-1}}}) would be needed to heat the gas
through supernovae, assuming perfect efficiency and no losses. Accounting
for radiative losses from the hot gas, the much higher star formation rates
for the two galaxies measured in paper I suggest that efficiencies of 2-4\%
would be sufficient to heat the gas in the ridge to its observed
temperature.
However, the enrichment of the IGM by heavy elements in the galactic winds
must also be considered. If the abundance of the ridge is truly 0.05\ensuremath{\mathrm{~Z_{\odot}}},
this implies that the ridge contains $\sim$10400\ensuremath{\mathrm{~M_{\odot}}}\ of Fe. Production of
heavy elements is likely to be dominated by core collapse supernovae in the
superwind galaxies, so adopting an Fe yield of 0.07\ensuremath{\mathrm{~M_{\odot}}}\ per supernova
\citep{Finoguenovetal00}, we would require only $\sim$1.5$\times$10$^5$
supernovae to produce the observed enrichment. This is insufficient to heat
the ridge gas to the observed temperature. Even if the abundance were
0.3\ensuremath{\mathrm{~Z_{\odot}}}, we would still only require $\sim$9$\times$10$^5$ supernovae,
between 1/5 and 1/7 of the number required to heat the gas in the ridge.
If we use our estimate of the number of supernovae required to heat the
gas, we find that the expected Fe abundance is $\sim$1.3\ensuremath{\mathrm{~Z_{\odot}}}. However, we
only expect galactic winds to have contributed 20-40\% of the gas in the
ridge. If the remaining 60-80\% is low-abundance, unenriched gas heated by
some other process (gravitational collapse and/or shocks) mixing will
reduce the expected abundance to $\sim$0.25-0.5\ensuremath{\mathrm{~Z_{\odot}}}. The lower end of this
range could be consistent with our abundance measurements, once the effect
of biases arising from the multiphase nature of the IGM are taken in to
account. We therefore conclude that it is plausible both energetically and
in terms of mass, that galactic winds from the member galaxies have played
a significant part in forming the IGM we observe in HCG~16, though they
cannot be the only source of hot gas.
\section{Conclusions}
\label{sec:conc}
HCG~16 is one of only two spiral-only groups known to support an X-ray
luminous hot IGM. As such, it provides a unique view of the early stages of group evolution, in which hot and cold gas phases coexist and tidal interactions have begun to reshape the galaxy population. The possibility that dynamical interactions and galaxy winds may play a significant role in the build up of the hot IGM is of obvious importance to our understanding of the development of galaxy groups and their member galaxies. We have analysed new deep \emph{Chandra}\ observations of HCG~16 with the goal of determining the properties and likely origin of its hot gaseous halo. Our results can be summarised as follows:
\begin{enumerate}
\item We confirm the presence of an extended hot IGM in HCG~16 with a
temperature $\sim$0.3~keV and abundance 0.05$^{+0.07}_{-0.03}$\ensuremath{\mathrm{~Z_{\odot}}}. Hot
gas is found to extend through the ACIS-S3 field of view, with the
brightest emission forming a ridge linking the four original group
members and extending to the southeast, in rough agreement with previous
\emph{ROSAT}\ and \emph{XMM-Newton}\ results. The ridge partly overlaps the \ensuremath{\mathrm{H}\textsc{i}}\ tidal
filament that links the five major galaxies, particularly in the region
around NGC~838 and NGC~839, suggesting that the two gas phases are
intermingled. The ridge contains 6.6$^{+3.9}_{-3.3}$$\times$10$^9$\ensuremath{\mathrm{~M_{\odot}}}\
of gas with an entropy $\sim$25-30~keV~cm$^2$. This mass is similar to
the \ensuremath{\mathrm{H}\textsc{i}}\ deficiency of the group, while the entropy is comparable to the
limit below which cooling is thought to fuel star formation and nuclear
activity in the cooling flows of galaxy clusters \citep{Voitetal08}.
However, the cooling time of the gas is long, 7-10~Gyr, suggesting that
radiative cooling is unlikely to be an important factor for the hot IGM.
\item We consider three possible mechanisms which may have contributed to
the formation of the IGM in HCG~16. The correlated X-ray and \ensuremath{\mathrm{H}\textsc{i}}\
morphology suggests that in the group core neither the hot nor the cold
gas component is relaxed, while the radio and X-ray morphologies of the
galactic winds provide evidence that hot gas can escape the galaxies and
mix into the IGM despite the surrounding cold \ensuremath{\mathrm{H}\textsc{i}}. The assumption that the
group is virialized with an IGM formed through gravitational accretion
leads to unrealistically high gas and baryon fraction estimates. A more
limited gas distribution similar to that of the \ensuremath{\mathrm{H}\textsc{i}}\ implies a hot gas
mass of 5.2-9.4$\times$10$^{10}$\ensuremath{\mathrm{~M_{\odot}}}\ and luminosity $L_{X,{\rm
bolo}}$=2.5-6.1$\times$10$^{41}$\ensuremath{\erg \ps}. We consider the possibility
that part of the hot IGM may have been formed through the shock heating
of neutral hydrogen as NGC~848 passed through the group core, but while
we find it likely that NGC~848 was supersonic during this passage, there
is no direct evidence of such a shock. Taking our rates of outflow for
hot gas in the winds of NGC~838 and NGC~839 from Paper I, we find that
they may have ejected $\sim$2$\times$10$^{10}$\ensuremath{\mathrm{~M_{\odot}}}\ of gas since their
starbursts began. If NGC~833 and NGC~835 underwent similar starburst
phases at an earlier period, galactic winds could have contributed
20-40\% of the hot gas observed in the IGM.
\end{enumerate}
We therefore conclude that starburst winds have played a significant role in the development of the IGM in HCG~16, alongside gravitational infall and possibly collisional shock heating of \ensuremath{\mathrm{H}\textsc{i}}. Given the very small numbers of low-mass, spiral-rich groups at this early evolutionary stage in which a hot diffuse IGM is known to exist, it is difficult to judge how widely applicable this result is to the general population. However, the importance of understanding the process by which the IGM is formed and its relationship to other aspects of group evolution and structure formation makes it clear that further studies of similar systems are required.
\acknowledgements The authors thank L. Verdes-Montenegro for providing her
VLA \ensuremath{\mathrm{H}\textsc{i}}\ map of the group, and the anonymous referee for helpful comments on the paper. Support for this work was provided by the
National Aeronautics and Space Administration (NASA) through Chandra Award
Number G03-14143X issued by the Chandra X-ray Observatory Center (CXC),
which is operated by the Smithsonian Astrophysical Observatory (SAO) for
and on behalf of NASA under contract NAS8-03060. SG acknowledges the
support of NASA through the Einstein Postdoctoral Fellowship PF0-110071
awarded by the CXC, and this research has made use of data obtained from
the Chandra Data Archive and software provided by the CXC in the
application packages CIAO, ChIPS, and Sherpa, as well as SAOImage DS9,
developed by SAO. We thank the staff of the GMRT for their help during
observations. GMRT is run by the National Centre for Radio Astrophysics of
the Tata Institute for Fundamental Research. We acknowledge the usage of
the HyperLeda database (http://leda.univ-lyon1.fr).
\textit{Facilities:} \facility{CXO} \facility{VLA} \facility{GMRT}
\bibliographystyle{apj}
|
\section{Introduction}
Active Galactic Nuclei (AGN) are galaxies currently accreting significant amounts of material onto their central supermassive black holes (SMBHs). SMBHs are now believed to exist in the centres of all galaxies that contan a significant bulge component \citep{kr95}, and an observed relation between the mass of the SMBH and the bulge stellar mass \citep{fm00,geb00} suggests a coupled assembly history between the two. Galaxy clusters provide a unique environment in which to study the relationship between AGN and the galaxies in which they reside.
The dense cluster environment is known to affect the evolution of cluster galaxies, producing a significant population of early-type galaxies in contrast to the lower-density field, which is comprised mainly of late-type galaxies \citep[e.g.,][]{hh31,morgan61,abell65,oemler74}. For AGN, the cluster environment could impact the onset and continued fuelling of the accretion process. The two most important fuel sources for AGN are believed to be cold gas reservoirs near the central black hole and galaxy mergers, which cause inflows of gas into the nucleus \citep[e.g.,][]{bh92,sdh05}. Since few cluster galaxies have significant reservoirs of cold gas \citep[e.g.,][]{gh85} and major mergers are rare in clusters due to high relative speeds between cluster galaxies, AGN were expected to be rare in clusters. The existence of an AGN would be an indicator that such a reservoir of available gas exists in the central region of the host galaxy. AGN can thus provide information about the efficiency of cold gas stripping in galaxy clusters, as well as the extent to which the central supermassive black hole may grow in such an environment.
Cluster AGN can be used to probe galaxy/AGN evolution in dense environments \citep[e.g.,][]{galametz09,haines12}. Understanding the important factors in transforming star-forming galaxies into passive galaxies and the effects of environment on this process will provide a clearer picture of galaxy evolution. For example, it has been proposed that radiative feedback from the AGN may result in the quenching of star-formation and the transition to passively-evolving galaxies on the red sequence \citep[e.g.,][]{hopkins04, croton06, alexander05}. Feedback from AGN activity may also significantly contribute to the heating of the intracluster medium (ICM) and act as an important factor in galaxy evolution within the cluster \citep[e.g.,][]{mcnamara05}. Understanding the link between active galaxies and their environment is thus an important step in the process of understanding galaxy clusters and their evolution.
Early spectroscopic studies \citep{gisler78,dressler85} suggested that AGN were less common in rich clusters than in the field. However, more recent X-ray surveys have found an AGN fraction approximately five times higher than that of optical spectroscopic surveys \citep{martini02,martini06} and close to the X-ray AGN fraction in the field \citep{haggard10}. These results suggest that spectroscopic searches are missing many AGN, as emission from an AGN may be hidden due to weak line strengths, dust obscuration, or dilution of the source by the host galaxy light. A true understanding of the fraction of AGN in galaxy clusters thus relies upon the successful identification of AGN over a range of luminosities and obscuration levels.
To address the AGN fraction in clusters and understand the role environment plays in their onset and continued fuelling, we studied a sample of 12 massive clusters at 0.5 $<$ z $<$ 0.9, in which AGN were identified via optical variability, X-ray emission and mid-IR power-law SEDs fitting. The identification of AGN candidates is presented in \citet{ks12a} (hereafter referred to as KS12). We identified an average of $\sim$25 AGN per cluster and found no significant difference between the fraction of AGN among galaxies in clusters and similarly-detected AGN in field galaxy studies ($\sim$2.5\%). This result indicates that the dense cluster environment does not appear to significantly hinder accretion onto the central SMBH to fuel an AGN.
In this paper, we examine the radial distribution and colours of AGN host galaxies in our clusters to investigate the impact of local environment on AGN and their hosts. In Section~\ref{clusters}, we briefly describe the galaxy cluster and AGN sample presented in KS12. In Section~\ref{radist}, we compare the radial distributions of AGN and normal galaxies in our cluster sample. In Sections~\ref{morph} and~\ref{AGNprop}, we describe how different cluster morphologies and AGN properties relating to obscuration may impact the AGN distribution in clusters. In Sections~\ref{colors} and~\ref{radcolors}, we discuss the overall distribution of galaxy colours among AGN as compared to normal galaxies in the cluster and as a function of cluster radius. Finally, we present our conclusions in Section~\ref{conclusions}. Throughout the paper we assume a standard cosmology with H$_{0}$ = 70, $\Omega_{M}$ = 0.3, and $\Omega_{\Lambda}$ = 0.7.
\section{Galaxy Cluster Sample and AGN Selection}
\label{clusters}
To investigate the role of environment on the AGN phenomenon, we conducted a census of AGN in several galaxy cluster fields (KS12). We chose clusters for which archival data were available that would allow AGN to be detected via optical variability, X-ray emission and mid-IR properties. With this multi-wavelength, multi-technique approach, we produced a largely unbiased sample with which to investigate the relationship between AGN, their environment, and their host galaxies. These same AGN identification techniques were also used to select AGN within the GOODS fields \citep{sarajedini11}, providing an excellent way to compare the field and cluster environments. All clusters were part of an HST ACS survey for supernovae in massive clusters \citep{sharon2010}. Every cluster had also been observed with the Chandra X-ray Observatory, while seven had mid-IR observations from the Spitzer Infrared Array Camera (IRAC). Table~\ref{clusterinfo} lists these clusters and several cluster properties.
\begin{table*}
\centering
\begin{minipage}{140mm}
\caption[Galaxy Cluster Properties]{Galaxy Cluster Properties \label{clusterinfo}}
\begin{tabular}{lcccr}
\hline
\multicolumn{1}{c}{Cluster} &
\multicolumn{1}{c}{Redshift} &
\multicolumn{1}{c}{M$_{200} (10^{15} M_{\odot})$} &
\multicolumn{1}{c}{r$_{v}$ (Mpc)} &
\multicolumn{1}{c}{Morphology Class\footnotemark[1]} \\
\hline
$CL0152-1357$ & 0.831 & 0.45 & 1.14 & Disturbed$^{b}$ \\
$CLJ1226.9+3332$ & 0.888 & 1.4 & 1.66 & Relaxed$^{b}$ \\
$MACSJ0257-2325$ & 0.506 & 1.41\footnotemark[2] & 1.95 $\pm$ 0.42 & Semi-Relaxed$^{a}$ \\
$MACSJ0717+3745$ & 0.548 & 2.75\footnotemark[2] & 1.93 $\pm$ 0.28 & Disturbed$^{a}$ \\
$MACSJ0744+3297$ & 0.686 & 0.87\footnotemark[2] & 1.47 $\pm$ 0.18 & Semi-Relaxed$^{a}$ \\
$MACSJ0911+1746$ & 0.505 & 0.84\footnotemark[2] & 1.61 $\pm$ 0.50 & Disturbed$^{a}$ \\
$MACSJ1149+2223$ & 0.544 & 1.13\footnotemark[2] & 2.64 $\pm$ 0.14 & Disturbed$^{a}$ \\
$MACSJ1423+2404$ & 0.545 & 0.78\footnotemark[2] & 1.35 $\pm$ 0.19 & Relaxed$^{a}$ \\
$MACSJ2214-1359$ & 0.504 & 1.22\footnotemark[2] & 1.54 $\pm$ 0.16 & Semi-Relaxed$^{a}$ \\
$MS0451.6-0305$ & 0.550 & 1.4 & 2.6 & Relaxed$^{b}$ \\
$MS1054.4-0321$ & 0.830 & 1.1 & 1.8 & Semi-Disturbed$^{b}$ \\
$SDSS1004+41$ & 0.680 & 0.42 & 1.35 & Relaxed$^{b}$ \\
\hline
\end{tabular}
\footnotetext{$^{1}$ a: morphology classification from \citet{ebeling07}, b: morphology classification from \citet{ks12a}}
\footnotetext{$^{2}$ The value listed for M$_{200}$ corresponds to M$_{X,200}$ derived from X-ray gas as measured by \citet{barrett06}}
\end{minipage}
\end{table*}
We describe the AGN identification and selection process in detail in KS12, and thus provide only a brief summary here. Since our primary data set is the HST ACS images of each cluster, we limit our survey regions to the area slightly smaller than the ACS image centered at the cluster. This region is $\sim$3 $\arcmin$ across and corresponds to $\sim$1 Mpc at the cluster redshift. This is the region for which optical variability of the galaxies could be measured (i.e. where two or more ACS images of the cluster overlapped in multi-epoch imaging of the sources). Throughout the paper, we use the terms ``cluster regions'' or ``cluster field'' to refer to this optical variability survey field-of-view centered approximately at each cluster location.
Each cluster in our sample had two to three epochs of HST ACS observations in the I band, with epochs separated by $\sim$one year. These data were used to photometrically identify varying nuclei in galaxies, which are likely to be AGN. We found 178 varying nuclei in $\sim$12,500 galaxies surveyed in the 12 cluster regions. These nuclei reach apparent magnitudes of I$_{nuc}$ $\leq$ 27, corresponding to absolute magnitudes of M$_I$ = -15.3 at z = 0.5 and M$_I$ = -16.8 at z = 0.9. We also identified AGN that appeared as X-ray point sources in each of our cluster regions using data from the Chandra Data Archive. These data were sensitive to point sources down to 7 x 10$^{-16}$ erg/cm$^{2}$/s in the full band. We found 74 X-ray point sources, with an average of six X-ray point sources per cluster region. Our X-ray flux limit in the full band corresponds to an X-ray luminosity of $\sim$6.7$\times$10$^{41}$ ergs/s at z = 0.5 and $\sim$2.8$\times$10$^{42}$ ergs/s at z = 0.9. Finally, AGN were identified in the infrared using Spitzer IRAC observations, which were available for seven of the 12 clusters. Power-law SED fits through the IRAC channels (3.6 to 8 microns) identify dust-obscured AGN via their re-radiated X-ray, UV, and optical light. We found 64 sources in our seven cluster regions with mid-IR power-law SEDs, resulting in $\sim$9 IR power-law sources per region. The limiting magnitude of the infrared survey was $\sim$18 in the 8-micron band. This corresponds to L$_{IR}$ = 1.4$\times$10$^{42}$ ergs/s at z = 0.5 and L$_{IR}$ = 6.0$\times$10$^{42}$ ergs/s at z = 0.9. Many of the AGN candidates were identified in more than one wavelength, with 24\% of X-ray sources and 4\% of mid-IR power-law sources also detected as optical variables. Combining all three detection techniques, we find an average of $\sim$25 AGN candidates per cluster region with a range of 12 to 49 AGN candidates in each region.
\section{Radial Distribution of AGN}
\label{radist}
We compare the radial distribution of our AGN candidates to the radial distribution of normal galaxies for sources identified in our cluster fields. For this comparison, we define three samples based on the cluster membership probability of the galaxies and AGN. The first sample includes all sources found within the cluster survey region (i.e., everything within the overlapping ACS image field-of-view for the cluster). This sample will contain many foreground and background field sources, as well as cluster galaxies/AGN. The second sample contains those sources with $>$80\% probability of cluster membership as defined in KS12. These higher probability cluster members are generally those galaxies closer to the cluster centre, those with red colours, and spectroscopically-confirmed cluster members. The third sample contains only sources (AGN and galaxies) that are spectroscopically-confirmed cluster members and thus have 100\% probability of cluster membership.
\subsection{Individual Clusters}
We first compare the radial distribution of AGN to that of the normal galaxy population for our three samples in each individual cluster region. We perform a Kolmogorov-Smirnov test (KS test) to determine the probability that the AGN and galaxy radial distributions come from the same parent population. The results of these statistical tests are summarized in Table~\ref{rdistclusters}, along with other relevant cluster information. For each cluster, we give the redshift and fraction of the virial radius out to which we detect sources. We then report the KS probability that the radial distribution of AGN and normal galaxies are drawn from different parent populations for each of our three samples. We also give the total number of galaxies and AGN in the cluster field, as well as the number with greater than 80\% cluster membership probability and 100\% cluster membership probability (i.e., spectroscopically-confirmed cluster members).
\begin{table*}
\centering
\begin{minipage}{140mm}
\caption[Radial Distribution of Galaxies and AGN]{Radial Distribution of Galaxies and AGN \label{rdistclusters}}
\begin{tabular}{lcccc}
\hline
\multicolumn{1}{c}{Cluster} &
\multicolumn{1}{c}{Redshift} &
\multicolumn{1}{c}{KS\%$_{all}$ (\#Gal)(\#AGN)\footnotemark[1]} &
\multicolumn{1}{c}{KS\%$_{>80\%}$ (\#Gal)(\#AGN)\footnotemark[2]} &
\multicolumn{1}{c}{KS\%$_{100\%}$ (\#Gal)(\#AGN)\footnotemark[3]} \\
\hline
$CL0152-1357$ & 0.831 & 70.9 (727) (25) & 99.0 (185) (8) & 98.0 (39) (5) \\
$CLJ1226.9+3332$ & 0.888 & 83.6 (937) (34) & 2.7 (581) (16) & 79.1 (25) (2) \\
$MACSJ0257-2325$ & 0.506 & 87.6 (1188) (14) & 95.9 (1103) (14) & 78.3 (9) (1) \\
$MACSJ0717+3745$ & 0.548 & 85.0 (1455) (46) & 52.1 (1141) (28) & 85.2 (81) (3) \\
$MACSJ0744+3297$ & 0.686 & 69.3 (1183) (30) & 70.8 (1090) (28) & 89.7 (27) (2) \\
$MACSJ0911+1746$ & 0.505 & 11.4 (1020) (19) & 42.2 (945) (16) & --- (15) (0) \\
$MACSJ1149+2223$ & 0.544 & 15.5 (1444) (14) & 3.0 (1160) (11) & 76.7 (40) (1) \\
$MACSJ1423+2404$ & 0.545 & 87.6 (1157) (51) & 86.6 (953) (37) & 79.2 (32) (7) \\
$MACSJ2214-1359$ & 0.504 & 12.3 (1077) (10) & 54.7 (1020) (8) & 88.2 (41) (1) \\
$MS0451.6-0305$ & 0.550 & 99.9 (834) (25) & 98.4 (334) (15) & 77.5 (55) (4) \\
$MS1054.4-0321$ & 0.830 & 99.5 (859) (26) & 77.8 (533) (11) & 96.6 (107) (4) \\
$SDSS1004+41$ & 0.680 & 69.3 (694) (20) & 73.8 (620) (18) & --- (0) (0) \\
All & -- & 99.6 (12575) (314) & 56.5 (9665) (210) & 99.3 (471) (30) \\
All (Variable AGN only) & -- & 52.0 (12575) (178) & 15.6 (9665) (132) & 89.1 (471) (30) \\
All (X-ray AGN only) & -- & 79.3 (12575) (74) & 94.3 (9665) (47) & 87.8 (471) (11) \\
All (IR AGN only) & -- & 100.0 (12575) (64) & 62.5 (9665) (31) & 98.3 (471) (4) \\
Relaxed & -- & 94.9 (7070) (186) & 94.3 (5701) (136) & 95.2 (189) (17) \\
Disturbed & -- & 97.0 (5505) (131) & 43.9 (3964) (74) & 99.9 (282) (13) \\
\hline
\end{tabular}
\footnotetext{$^{1}$ The probability that AGN and galaxy radial distributions are drawn from different parent populations for all sources identified in the cluster field.}
\footnotetext{$^{2}$ The probability that AGN and galaxy radial distributions are drawn from different parent populations for sources with 80\% cluster membership probability.}
\footnotetext{$^{3}$ The probability that AGN and galaxy radial distributions are drawn from different parent populations for sources which are spectroscopically-confirmed cluster members.}
\end{minipage}
\end{table*}
Four of the twelve clusters in our survey show a significant difference between the AGN and normal galaxy distributions in at least one of our three samples. CL0152 shows significant differences among the samples with $>$80\% and 100\% probability of cluster membership, while MACSJ0257 shows a significant difference (96\% confidence) for the $>$80\% sample only. Examining the radial distributions of these two clusters reveals that the AGN sample is primarily less centrally concentrated than the normal galaxies except within the inner 20\% of the virial radius, where the AGN are either more centrally concentrated or closely follow the normal galaxy distribution. Two additional clusters, MS1054 and MS0145, reveal differences in the radial distribution of AGN and normal galaxies with even greater significance($<$97\% confidence for two of the three samples).
MS1054 and MS0145 are representative of the ways in which the radial distributions of galaxies and AGN differ from one another in our cluster sample. In Figure~\ref{cumhistcomp}, we show the cumulative distributions of galaxies in MS0451 (top panels) and MS1054 (bottom panels). The distribution for MS0145 extends to $\sim$32\% of the virial radius due to its low redshift (z = 0.55), whereas the distribution for MS1054 extends to more than 60\% of the virial radius at z = 0.83. What is most striking about these two clusters is the way in which their distributions differ from one another. In MS0145, the AGN are clearly more centrally concentrated than the normal galaxies, such that the differences are significant even in the sample containing many foreground and background sources (left, upper panel). In MS1054, we see the opposite trend. The AGN appear to be less centrally concentrated at all virial radii compared to the normal galaxies. To better understand these differences, we combine all cluster radial distributions in the next section and examine the distributions among different types of AGN. Based on our individual cluster analysis, we find that the AGN population is generally consistent with the distribution of cluster galaxies in two-thirds of our clusters, while one-third reveals AGN distributed differently from the normal galaxy population.
\subsection{Combined Cluster Distributions}
In an effort to improve the statistics of our sample and explore differences with AGN type, we combine the radial distributions of all 12 clusters scaled by the virial radius. In Figure~\ref{allch}, the left panel shows all sources detected in the cluster fields. The central panel contains only those sources with $>$80\% cluster membership probability, while the right panel shows spectroscopically-confirmed cluster members only. We show the distribution for normal galaxies (solid line) and AGN (thick solid line), as well as the distribution for AGN detected using different identification techniques: optical variability (green dotted line), X-ray (blue dot-dashed line) and mid-IR SED power-law sources (red dashed line).
The KS test reveals significant differences in the radial distributions of AGN and normal galaxies among the first (all sources in the cluster field) and third (spectroscopically-confirmed only) samples ($>$99\% significance; see Table~\ref{rdistclusters}). The AGN radial distribution appears to be more centrally concentrated than normal galaxies within 20\% of the virial radius, then becomes less centrally concentrated than normal galaxies beyond this radius. The division of AGN types reveals that the X-ray sources (blue dot-dashed line) are responsible for the central concentration of AGN in the inner part of the clusters, while the IR sources (red dashed line) appear to be the least centrally concentrated of the different AGN types. The KS test for X-ray sources does indicate a significantly different distribution (94\%) among the sources with $>$80\% cluster membership probability and a slightly significant distribution difference (88\%) for the spectroscopically-confirmed cluster members only. IR AGN also reveal significantly different radial distributions (100\% and 98\% significance) among two of our three samples. The radial distribution of optical variables appears to lie somewhere between the two other types of AGN samples, revealing a distribution that follows the normal galaxies closely in the inner regions of the clusters, and then becomes slightly less centrally concentrated farther out with just under 90\% significance among the spectroscopically-confirmed cluster members.
Our result is consistent with \citet{galametz09}, who detected a slight (1.2$\sigma$) overdensity of X-ray sources in the centres of their clusters (r $<$ 0.5--1 Mpc) at redshifts similar to our clusters. Similarly, \citet{re05} find an excess of X-ray point sources within the inner r $<$ 0.5 Mpc of their sample of 51 MACS clusters. A recent study by \citet{ehlert13} finds that while their sample of clusters at similar redshifts contains an overdensity of X-ray sources, the X-ray sources are actually less centrally concentrated within the cluster than the normal galaxy distribution. This differs from our result, but may be explained with the following reasons: first, our X-ray survey contains fainter X-ray sources than the survey of \citet{ehlert13} (7$\times$10$^{-16}$ erg/cm$^{2}$/s in our survey, compared with 5$\times$10$^{-15}$ erg/cm$^{2}$/s; this corresponds to X-ray luminosities of 6.7$\times$10$^{41}$ erg/s in our survey and 4.8$\times$10$^{42}$ erg/s in their survey at z = 0.5). Second, we compare our AGN source distribution directly to that of the cluster galaxy distribution, rather than a King or NFW model distribution. If the difference is due primarily to the first reason, it may indicate that more luminous X-ray sources tend to avoid the centres of clusters, while less luminous ones prefer the denser regions, which is consistent with the results of \citet{kauffmann04}. Sensitive X-ray surveys of more clusters will provide additional data to test this theory.
\citet{galametz09}, also identified an overdensity of IR sources within the central 0.3 Mpc of clusters, which differs from our result. However, their IR sources were detected using mid-IR colours, rather than the power-law fitting approach that we use, and thus may contain a broader range of AGN candidates. \citet{atlee11} examine the radial distribution of X-ray point sources and IR power-law sources in a sample of low-redshift (z $\sim$ 0.06--0.31) galaxy clusters and find that both samples are consistent with the distribution of cluster galaxies, though with better agreement among the IR sources than the X-ray sources. We find an overall trend that the IR sources are less centrally concentrated than the normal cluster galaxies.
\section{Cluster Morphology and Radial Distribution}
\label{morph}
Galaxy clusters can have a range of morphologies and density distributions. In KS12, we discuss the morphology classifications for our clusters, which are fairly evenly divided between ``relaxed'' and ``disturbed" morphologies according to the criterion of \citet{re05}. We list these classifications in Table~\ref{clusterinfo}. \citet{re05} found that the observed central excess of X-ray sources among their clusters was more pronounced in relaxed clusters than disturbed ones. Relaxed clusters have a central cooling core dominated by their massive cD galaxies; \citet{re05} speculate that interactions (e.g., mergers, tidal interactions) with the cD galaxies and other giant ellipticals in the cores of these clusters could result in the excess of X-ray AGN they detect.
As shown previously, we find AGN distributions significantly different from the normal galaxy distributions in four of our 12 cluster regions. Of these four clusters, two are ``relaxed'' or ``semi-relaxed'' and two are ``disturbed'' or ``semi-disturbed.'' To further investigate the effects of cluster morphology on the AGN population, we compare the radial distributions of AGN and normal galaxies for relaxed and disturbed clusters in Figure~\ref{agnchmorph}. We combine the ``relaxed'' and ``semi-relaxed'' clusters (seven total) and the ``disturbed'' and ``semi-disturbed'' clusters (5 total) in this figure. In relaxed clusters, we find that the AGN population is more centrally concentrated than the normal galaxies for all sources in the cluster field (left, upper panel) and for spectroscopically-confirmed cluster members only (right, upper panel) at a 95\% significance level. The AGN appear more centrally concentrated within the inner $<$20--30\% of the virial radius, with the distribution dominated by the X-ray selected AGN (blue dot-dashed line). In contrast, the disturbed clusters do not have a centrally concentrated AGN distribution (lower panels). While the AGN distribution is still significantly different from the galaxy distribution (97\% significance and 99.9\% significance for the samples with all sources and spectroscopically-confirmed only, respectively), the AGN appear less centrally concentrated for all AGN types. Disturbed clusters by nature have significant substructure and their centres are not as well-defined. \citet{re05} explain that the lack of an AGN excess in the central regions of these clusters may be due to this fact, since any change in the distribution of AGN with respect to the cluster galaxies will be spread out over larger radii.
\citet{re05} find an ``AGN depletion zone'' in relaxed clusters with a reduced density of AGN at radii $\sim$0.5 to 2 Mpc, followed by an increase of AGN at roughly the cluster virial radius where the cluster merges with the field. They explain this finding as a result of the shorter timescale for depleting a merger-induced accretion disk around a central supermassive black hole when compared with the cluster crossing time. \citet{re05} also suggest that the increase of AGN at the virial radius is the result of AGN triggered by mergers or galaxy harassment as galaxies first fall into the cluster. Our radial distribution results for relaxed clusters are not inconsistent with a depletion zone at intermediate cluster radii. The upper panels of Figure~\ref{agnchmorph} reveal an AGN distribution that rises much more slowly beyond the inner 30\% of the virial radius. This change in slope of the AGN distribution corresponds to a radius of $\sim$0.5 Mpc, since the average relaxed cluster virial radius is 1.7 Mpc. To quantify this, we calculate the ratio of AGN to normal galaxy number density as a function of virial radius. Within 0.1 virial radii for confirmed cluster members, we find the fraction of AGN to be 0.205$\pm$0.080 (8/39). This fraction drops to 0.049$\pm$0.019 (7/144) at virial radii between 0.1 and 0.5. From 0.5 to 0.6 virial radii, the fraction rises again to 0.333$\pm$0.272 (2/6). The drop between the inner and intermediate radius fractions is significant at just 1.6$\sigma$, and the rise beyond 0.5 virial radii is even less statistically significant due to the small number of sources. Therefore, these results appear largely consistent with the findings of \citet{re05}, though we cannot confirm their observed increase in AGN at the cluster virial radius.
\section{AGN Properties and Radial Distribution}
\label{AGNprop}
Next we explore correlations between the radial location of an AGN within a cluster and various AGN properties that relate to physical conditions within the AGN, such as obscuration. Obscured AGN generally have higher X-ray hardness ratios, lower variability significance \citep{ks07}, and steeper mid-IR power-law slopes \citep{ah06}. Figure~\ref{sigmavsdist} plots variability significance vs. distance from the cluster centre for all optical variables. We see no clear correlation between variability significance and distance from the cluster centre, though the most significant optical variables appear to lie at distances that are 20\% to 60\% of the cluster virial radius.
Figure~\ref{hrvsdist} shows the X-ray hardness ratio vs. cluster radius for all X-ray point sources observed in both the hard and soft bands. Hardness ratio is calculated as
\begin{equation}\label{} HR = \frac{F_{X}(2-8keV)}{F_{X}(0.5-2keV)}\end{equation}
\noindent
as done in KS12. We detect both hard and soft X-ray sources at all radii, though the hardest sources appear at radii less than $\sim$60\% of the cluster virial radius (we note that we do not detect many X-ray sources outside of this radius in general). There is no clear trend with radial distance from the cluster centre, though we find that all X-ray point sources with spectroscopically-confirmed cluster redshifts are among the softer sources and have a small range of X-ray hardness ratios, with only two cluster members having hardness ratios $>$5. This may be a selection effect, since softer X-ray sources may also be optically brighter and easier to detect in spectroscopic follow-up surveys.
Finally, Figure~\ref{alphavsdist} plots the slope of IR power-law vs. cluster radius for all IR power-law sources. \citet{ah06} find that the slope of the power law fit relates to AGN type, where steeper (i.e., more negative) values represent NLAGNs and shallower power-law SEDs are classified as BLAGNs. We see that sources with SEDs resembling BLAGNs ($\alpha$ $>$ -0.9) are found at all radii, but NLAGN-like SEDs ($\alpha$ $<$ -0.9) appear primarily at radii between $\sim$20--50\% of the cluster virial radius. There are only four IR power-law sources with spectroscopically-confirmed cluster redshifts and only one falls between 0.2 and 0.6 of the cluster virial radius, though it does have a steeper IR SED than the other three objects.
In summary, we find no significant correlations between AGN properties such as variability significance, X-ray hardness, or IR power-law slope and the radial location of an AGN within the cluster.
\section{AGN Host Galaxies in Clusters}
\label{colors}
The link between AGN activity and the evolution of galaxies may be reflected in the properties of galaxies which currently host actively accreting supermassive black holes. There is also a clear correlation between environment, galaxy type and star formation rate. We examine the colours of our AGN candidate host galaxies in an effort to explore the relationship between the presence of an AGN, the host galaxy, and the cluster environment.
\subsection{Variable, X-ray and IR AGN Host Galaxies}
As described in KS12, our cluster fields have been observed in both the HST V (F555W) and I (F814W) bands. To compare the observed V-I colours and minimize k-corrections, we limit our analysis to those clusters that fall in the small redshift range between 0.504 $<$ z $<$ 0.550, which includes seven of our 12 cluster fields. In Figure~\ref{groupcolormag}, we show the observed colour-magnitude diagram for all galaxies in the cluster field (cluster and field galaxies) for these z $\sim$ 0.5 clusters, with AGN candidates indicated by coloured symbols. The expected bimodal distribution of ``red sequence'' bulge-dominated, passively-evolving galaxies and ``blue cloud'' star-forming, disk-dominated galaxies \citep[e.g.,][]{strateva01,blanton03,kauffmann03} can be clearly seen in the normal galaxy population, indicated with black dots in the figure.
Many studies have found that the optical colours of AGN selected via X-ray, IR, and optical variability largely reflect the colours of the host galaxy. \citet{cardamone10} find that X-ray-selected AGN show less that 0.1 mag of contamination by the AGN on the host galaxy rest-frame U-V colours. \citet{hickox09} find that only 0.4--0.5 mag of correction in u-r is required to remove AGN contamination from the bluest X-ray and mid-IR sources in their field AGN survey. We expect the most luminous blue AGN candidates to be those with the most colour contamination from the nuclear emission on the host galaxy light. An AGN-dominated galaxy would lie in the blue cloud at the bright end of the distribution in Figure~\ref{groupcolormag}. We find only five AGN candidates among blue sources at I $<$ 20 (corresponding to an absolute magnitude of M$_{I}$ $\sim$ -22 at a redshift of 0.5 and M$_{I}$ $\sim$ -23.5 at a redshift of 0.9). Thus, we do not see evidence for a large amount of AGN contamination among the host galaxies in our AGN sample and assume that the colours represent the AGN host galaxy light as a reasonable approximation. We note, however, that some fraction of AGN sources may be shifted towards bluer colours. We estimate this effect to be less than half a magnitude in the most extreme cases.
Figure~\ref{groupcolors} shows the distribution of observed V-I colour for both normal galaxies and those hosting AGN for our z $\sim$ 0.5 clusters. The left panel includes all sources in the cluster field (cluster and field galaxies/AGN) and the right panel contains only spectroscopically-confirmed cluster members. The spectroscopically-confirmed members are mainly red galaxies with V-I $>$ 2. This reflects the selection effect for the spectroscopic sample, since most objects for which spectra were obtained were chosen based on their red colours \citep{barrett06}. For this reason, the spectroscopically-confirmed cluster member colour distribution is not a good measure of the true galaxy distribution in our clusters. We rely on the distribution of all sources in the cluster fields and compare to pure field survey results to determine if a difference can be detected.
The left panel containing all sources in the cluster field reveals optical variables extending over a broad colour range, peaking in the ``green valley'' between the red sequence and blue cloud at V-I $\sim$ 1.5 with an average colour of $\langle$V-I$\rangle$ $\sim$ 1.9. There are about 100 optically variable AGN in our z $\sim$ 0.5 cluster fields, compared to only $\sim$25 IR and X-ray-selected AGN. Even with this small number of AGN, we see that the IR-selected and X-ray-detected AGN follow a similar trend, with a range of colours and $\langle$V-I$\rangle$ = 2, though there is no obvious peak in either distribution.
\citet{hickox09} examined the host colours of X-ray- and IR colour-selected AGN at 0.25 $<$ z $<$ 0.8. They found that X-ray AGN tend to lie in the ``green valley'' with a small tail of blue host galaxies. The distribution of their IR AGN sample is similar to both ours and that of \citet{atlee11}, which show IR AGN residing in galaxies that have slightly bluer hosts than the X-ray AGN and a less pronounced green peak. We also compare our results to \citet{sarajedini11}, who examined the colours of GOODS field AGN identified via optical variability, X-ray emission, and IR power-law SED fits. They found that X-ray sources and optical variables inhabit host galaxies with a range of colours, with the X-ray point sources peaking in the green valley and the variables showing a flatter distribution through the green valley. Their IR power-law sources showed a relatively flat distribution across the range of U-V colours. These results agree with our findings as well, indicating that the host galaxy colours of AGN in these cluster regions do not differ significantly from that of pure field galaxy surveys. Since this conclusion is based on a sample that contains a mixture of cluster and field galaxies/AGN, the cluster environment cannot be ruled out as having no effect on AGN host galaxy colour because the impact may be too subtle to discern in our mixed population.
\citet{hickox09} present a general picture of galaxy evolution in accordance with the model presented by \citet{hopkins08}, where galaxies begin as disk-dominated gas-rich, star-forming systems characterised by blue optical colours. Once a galaxy undergoes some sort of triggering event, such as a major merger, it displays a relatively short ($\sim$10$^{8}$ yr) phase of AGN activity, which is subsequently quenched along with the star formation in the galaxy. As AGN and star formation activity decline and disappear, the host galaxy's spectrum evolves toward intermediate colours that result from a composite spectrum of emission from the older stellar population and a declining contribution from younger stars. Finally, after $\sim$1--2 Gyr, the galaxy evolves into a red, passively-evolving spheroid-dominated system in the red sequence portion of the colour-magnitude diagram. Several studies have shown that AGN hosts have colours that indicate this decline in star formation and transition from blue to red \citep{schawinski07,silverman08,bundy08}. These findings support the idea that nuclear activity suppresses star formation through feedback mechanisms such as radiation pressure and/or jets, which can inject energy (either radiative or mechanical) into the gas surrounding the nucleus and prevent it from cooling and fuelling any further star formation \citep[e.g.,][]{fabian03,mcnamara05,forman07,mn07}.
The dense cluster environment is expected to have some effect on the star formation rates and colours of cluster galaxies. Studies have found evidence of star formation being triggered in infalling galaxies, which later couples with physical mechanisms such as ram pressure stripping or galaxy harassment to quench star formation \citep{bai07,marcillac07}. \citet{chung10} found evidence for an elevated IR luminosity function in the Bullet Cluster, a galaxy cluster at z = 0.296 that is currently undergoing a major supersonic merger. The observed excess in star-forming IR galaxies can be associated with the infalling group of galaxies that have not yet been processed by the cluster environment into more quiescent galaxies. This suggests that the processes responsible for quenching star formation occur on a longer timescale than that of the infalling group's accretion into the Bullet Cluster ($\sim$250 Myr). One such slower process, known as ``strangulation,'' can occur in galaxies being assimilated into galaxy clusters. A galaxy's loosely-bound halo gas is gently pushed away through interactions with the intracluster medium, which can affect both the morphology and star formation rate of the galaxy \citep{larson80,balogh00}. This longer timescale for cluster environment effects on galaxy colours may be the reason we do not see an obvious difference between the AGN hosts in clusters when compared to field galaxy studies. The feedback mechanism that may affect the AGN hosts individually could also be operating on a larger scale in the ICM. Relativistic jets from radio AGN (such as the central cD galaxy in a cluster) could heat the surrounding intracluster medium and aid in quenching star formation in cluster galaxies \citep[e.g.,][]{birzan04,rafferty06,croston08}. To further investigate the cluster environmental effects and possible feedback effects within individual AGN, we compare the AGN host colours to the normal galaxies as a function of cluster radius in Section~\ref{radcolors}.
\subsection{Cluster Morphology and AGN Host Galaxies}
We investigate the galaxy colours in clusters divided by morphology classification to further explore the interplay of AGN and cluster environment on AGN host galaxies. Figure~\ref{morphcolors} shows a histogram of the optical V-I colours of the galaxies and AGN in relaxed (top panels) and disturbed clusters (bottom panels). The figures include data from our seven z $\sim$ 0.5 cluster fields, which are fairly evenly divided among disturbed (3) and relaxed (4) clusters.
The left panels of the figure include all galaxies in the cluster fields (i.e., field plus cluster sources). The disturbed clusters reveal a distribution of optical variables peaking in the green valley at V-I = 1.65 with $\langle$V-I$\rangle$ = 1.9. In relaxed clusters, the variables have roughly the same mean V-I colour, but extend through the green valley and peak in the red sequence at V-I = 2.3. While the colour distributions are not significantly different (we calculate a KS probability of 82\% that they are drawn from different parent populations), the absence of the red peak among the disturbed clusters is noticeable. The X-ray and IR AGN have host galaxy colours distributed throughout the full range of colours in both relaxed and disturbed clusters, with similar $\langle$V-I$\rangle$ $\sim$ 2. Among the spectroscopically-confirmed cluster members (right panels), the colour selection effect is apparent. The distribution here mainly reflects that of the left panels limited to sources redder than V-I = 2, where the spectroscopic follow-up observations were targeted.
If galaxy interactions and substructure in the hot X-ray gas are more common in disturbed clusters than relaxed clusters, this may be the reason that the AGN hosts appear to tend toward bluer colours in the disturbed clusters. These clusters are characterised by X-ray substructure and at least two of these clusters are either currently undergoing a merger or have recently experienced one in the past. Minor mergers, galaxy harassment, and interactions with substructure in the intracluster medium in disturbed clusters may be more common than in the X-ray-smooth, structurally-symmetric relaxed clusters. These processes may be responsible for triggering or maintaining star formation or AGN activity in cluster galaxies, pushing them toward bluer colours rather than passively-evolving red sequence galaxies. In contrast, galaxies in relaxed clusters may undergo fewer interactions as they fall into the smoother cluster density profile, and thus may be evolving more passively than those in disturbed clusters.
\section{Radial Distribution of AGN Host Galaxy Colours}
\label{radcolors}
In the dense cluster environment, star formation activity in galaxies is subdued compared to field galaxies and a higher fraction of early-type galaxies is observed \citep[e.g.,][]{hubble36,dressler80}. Furthermore, there is a well-known morphology-density relation observed in clusters, which is characterised by a higher fraction of early-type galaxies in dense cluster cores and an increasing fraction of late-type galaxies with increasing distance from the cluster centre as densities approach that of the field \citep[e.g.,][]{dressler80,oemler74,ms77}. A galaxy's distance from the centre of a cluster may also correlate with the time since infall into the cluster \citep{gao04}, as galaxies at large cluster radii have not yet encountered the densest parts of the cluster while galaxies in the cluster core were either formed in the dense environment or have already crossed from the outskirts to the cluster core at least once.
Given that the presence of an AGN may also affect star formation in the host galaxy and thus impact the galaxy colour, we compare the radial colour distribution of the host galaxies of our AGN candidates with that of the normal galaxy population to determine whether the AGN has any significant effect on the host galaxy as a function of radial distance from the cluster centre. In Figure~\ref{slopesbytype}, we show the radial colour distribution of the host galaxies for our three types of AGN candidates and the normal galaxy population for all galaxies within 60\% of the virial radius of the cluster. These figures include all sources within the cluster region and therefore contain both cluster and field galaxies/AGN. We have combined the data for the nine cluster regions at redshifts between z $\sim$ 0.5 and z $\sim$ 0.7 to provide the largest number of sources while avoiding significant redshift effects. We conducted this analysis using all sources in the cluster region, even though the sample will be contaminated by foreground and background galaxies. Since the spectroscopically-confirmed sources have a severe colour selection effect, any trend with radius has already been effectively removed from that sample. Likewise, our sources with greater than 80\% cluster membership probability have also had a colour and radial selection criterion imposed upon them, which would impact our analysis. Therefore, only the sample that includes all sources in the cluster fields is free from imposed selection effects.
To identify radial colour trends, the distributions were fitted with a straight line using $\chi^{2}$ minimisation to determine the best fit and the error on the fit. The slopes of the fits to the optical variables, IR power-law sources, and X-ray sources are -0.16 $\pm$ 0.6, 1.95 $\pm$ 1.8 and -0.25 $\pm$ 1.2, respectively. Combining the AGN samples yields a slope of -0.07 $\pm$ 0.54, consistent with a slope of zero. The normal galaxy distribution colour slope was found to be -1.2 $\pm$ 0.10. Thus, the normal galaxies are generally bluer with increasing radius and demonstrate a significantly steeper relation than the host galaxies of any AGN type or the combined AGN sample, which show no significant relation between cluster radius and colour. We found that cluster morphology (relaxed vs. disturbed) has no discernible effect on the slopes of the fits to the AGN or the normal galaxy distribution.
The galaxies and AGN analyzed with these fits represent the cluster plus field populations, where the field galaxies are expected to have no colour relation with cluster radius. Thus, any observed slope of this line is partly due to the true colour gradient expected among cluster galaxies and partly due to the decreasing cluster density and increasing dominance of the bluer field population as a function of increasing cluster radius. Since our fit to the AGN host galaxies also includes all sources in the cluster field, the AGN samples are subject to the same changes in cluster density and field contamination as a function of radius. The negative slope (with $>$3$\sigma$ confidence) of the normal galaxies reveals that the cluster population is dominant enough to reveal this expected trend. As discussed previously, several studies have found a relationship between normal galaxy colour and cluster radius. This relationship has been interpreted as due to the slow quenching of star formation over timescales of a few Gyr \citet{vdl10}, possibly through the process of strangulation (\citet{chung10,balogh00}). Our observations of the normal cluster galaxy population are consistent with these results.
The lack of a similar relationship among the AGN host galaxy colours is interesting, since the AGN population should contain a similar mix of cluster and field sources as function of cluster radius. While the colour of a normal galaxy within a cluster appears to depend to some extent on the radial distance and consequently the local environment, the AGN host colours do not reveal this dependency. This observation supports the possibility that processes related to the accreting supermassive black hole have a more significant impact on the star-forming properties of the host galaxy than does the intracluster medium and/or the local environmental density.
\section{Conclusions}
\label{conclusions}
We have explored several issues concerning the AGN population in dense cluster environments by analysing 12 galaxy clusters at redshifts 0.5 $<$ z $<$ 0.9 to determine the AGN fraction in clusters and address issues of AGN triggering and fuelling in massive galaxy clusters. In our first paper, \citep{ks12a}, we compiled a catalog of cluster AGN candidates using a combination of three detection techniques: optical variability, X-ray point source detection, and mid-IR power-law SEDs. We also discussed the overall AGN fraction in our cluster sample. In this paper, we focused on the analysis of the radial distributions of AGN and colours of AGN host galaxies in clusters.
We found that the radial distribution of the AGN population is consistent with the radial distribution of normal galaxies in two-thirds of our cluster fields, while one-third reveal an AGN population distributed differently than the normal galaxies. Among those with significantly different distributions, the AGN are generally less centrally concentrated than the normal galaxies except within the central 20\% of the virial radius, where the AGN are either more centrally concentrated or follow the normal galaxy distribution closely. When we combine the distributions of our cluster regions scaled by virial radius, we find that the AGN distribution rises slightly faster than the normal galaxies in the inner 20\% of the clusters, mainly due to the X-ray-selected AGN, and then becomes less centrally concentrated than the normal galaxies at greater radii. The slight over-concentration of AGN toward the cluster centres may be explained by interactions (e.g., mergers, tidal interactions) with the cD galaxies and other giant ellipticals in the cores of these clusters. At intermediate cluster radii, the cumulative distribution of AGN appears to rise more slowly than the normal galaxies and may be explained by the following scenario: as galaxies fall into the cluster for the first time, nuclear activity may be triggered by interactions at the cluster-field interface. The AGN triggered as a result will eventually deplete their accretion disks over a timescale less than the cluster crossing time, thus resulting in a less centrally concentrated AGN distribution at intermediate radii. Since our observations do not extend to the virial radius, we cannot confirm a higher concentration of AGN at the cluster virial radius.
We find that relaxed clusters, those with a clear centre and smooth radial distribution of galaxies, have more centrally-concentrated AGN than the normal galaxy distribution at $>$95\% confidence and these are primarily X-ray selected AGN. The disturbed clusters have AGN that are less centrally concentrated than the normal galaxies. In both types of clusters, the IR power-law sources appear to be the least centrally concentrated AGN type. Part of the reason for the difference between relaxed and disturbed clusters may be due to the fact that disturbed cluster radial distributions are spread over a larger range of radii since they do not have well-defined cores or spherical symmetry. This could make the detection of central concentrations difficult to identify in this type of cluster.
We explored correlations between the radial location of galaxies hosting AGN within the cluster field and various AGN properties that relate to physical conditions within the AGN. However, we found no significant correlations between the optical variability significance, X-ray hardness ratio, or IR power-law slope and the radial location of the AGN candidates within the cluster fields.
The host galaxies of the AGN candidates in our survey appear to have a different colour distribution than the normal galaxies in our cluster fields. The AGN tend to occupy the ``green valley" avoided by most normal galaxies. Comparing our results with the GOODS field, we found that the host galaxy colours of AGN in the cluster fields are not significantly different from pure field samples. While we cannot rule out the impact of cluster environment effects on these distributions, we note that the effect may be subtle and thus consistent with the quenching of star formation taking place over longer timescales, as found in recent cluster studies.
In disturbed clusters, the distribution of AGN candidates peaks in the green valley, whereas in relaxed clusters the distribution extends through the green valley but peaks among redder galaxies, especially among the optically variable AGN candidates. If galaxy interactions or substructure in the hot X-ray gas is more common in disturbed clusters than relaxed clusters, minor mergers, galaxy harassment, and interactions with substructure in the intracluster medium may be more common as a result of this cluster morphology, which is not present in the structurally-symmetric relaxed clusters. This may be responsible for triggering more star formation in the AGN hosts within disturbed clusters, pushing the galaxies toward bluer colours instead of allowing their star formation to be mostly quenched and the hosts to transition to passively-evolving red sequence galaxies.
Given that the presence of an AGN as well as the intracluster medium may affect star formation in the host galaxy and thus impact the galaxy colour, we have examined the radial colour distribution of the host galaxies of our AGN candidates compared with that of the normal galaxy population in our cluster fields. In this way, we attempted to determine if the presence of an AGN has any significant effect on the host galaxy as a function of radial distance from the cluster centre. We found that the AGN hosts display no apparent change in colour with cluster radius, while the normal galaxies become bluer at greater radii with $>$3$\sigma$ confidence. The fact that we see no colour-radius relation among AGN hosts supports the theory that processes related to the accreting supermassive black hole have a more significant impact on the star-forming properties of the host galaxy than does the intracluster medium and/or the local environmental density.
\section{Acknowledgements}
\label{acknowledgements}
The results in this paper are based on observations made with the NASA/ESA Hubble Space Telescope, obtained at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555. These observations are associated with program \# GO 10493, cycle 14 and GO 10793, cycle 15. This work is based in part on archival data obtained with the Spitzer Space Telescope, which is operated by the Jet Propulsion Laboratory, California Institute of Technology under a contract with NASA. This research has made use of data obtained from the Chandra Data Archive and the Chandra Source Catalog, and software provided by the Chandra X-ray Center (CXC) in the application packages CIAO, ChIPS, and Sherpa (ObsIDs 512, 520, 529, 902, 913, 1654, 1655, 1656, 1657, 3180, 3196, 3197, 3199, 3529, 3581, 3584, 3585, 3587, 3589, 3595, 4195, 4200, 5011, 5012, 5014, 5794). Funding was provided by NSF CAREER grant 0346691. We also thank the referee for a careful reading of this work and helpful comments and suggestions, which improved the quality of this paper.
|
\section{Introduction}
In 1962, F. Dyson \cite{Dy1, Dy2} observed that the eigenvalues of
the $N\times N$ Hermitian matrix valued Brownian motion is an
interacting $N$-particle system with the logarithmic Coulomb
interaction and derived their statistical properties. Since then,
the Dyson Brownian motion has been used in various areas in
mathematics and physics, including statistical physics and the
quantum chaotic systems. See e.g. \cite{Meh} and reference therein.
In \cite{RS93}, Rogers and Shi proved that the empirical measure of
the eigenvalues of the $N\times N$ Hermitian matrix valued
Ornstein-Uhlenbeck process weakly converges to the nonlinear
McKean-Vlasov equation with quadratic external potential as $N$
tends to infinity. This also gave a dynamic proof of Wigner's famous
semi-circle law for Gaussian Unitary Ensemble. See also \cite{AGZ,
Gui}.
The purpose of this paper is to study the generalized Dyson Brownian
motion and the associated McKean-Vlasov equation with the
logarithmic Coulomb interaction and with general external potential.
More precisely, let $\beta\geq 1$ be a parameter, $V:
\mathbb{R}\rightarrow \mathbb{R}^+$ be a continuous function, let
$(W^{1},\ldots, W^{N})$ be an $N$-dimensional Brownian motion
defined on a filtered probability space~$(\Omega, \mathscr{F},
(\mathscr{F}_t)_{t\geq 0}, \mathbb{P})$ satisfying the usual
conditions. Let
~$\lambda_N(0)=(\lambda^1_N(0),\ldots,\lambda^N_N(0))\in
\bigtriangleup_{N}=\{(x_{i})_{1\leq i\leq N}\in\mathbb{R}^{N}:
x_{1}<x_{2}<\ldots<x_{N}\}$. The generalized Dyson Brownian motion
$({\rm GDBM})_V$ is an interacting $N$-particle system
$\lambda_N(t)=(\lambda^1_N(t), \ldots, \lambda_N^N(t))$ with the
logarithmic Coulomb interaction and with external potential $V$, and
is defined as the solution to the following SDEs
\begin{eqnarray} d\lambda_N^i(t)=\sqrt{\frac{2}{\beta
N}}dW^i_{t}+\frac{1}{N}\sum\limits_{j:j\neq i}
\frac{1}{\lambda^i_N(t)-\lambda^j_N(t)}dt-{1\over 2}V'(\lambda_N^i(t))dt, \ \ \ i=1, \ldots, N, \label{SDE1}
\end{eqnarray}
with initial data ~$\lambda_N(0)$. It is a SDE for $N$-particles
with a singular drift of the form ${1\over x-y}$ due to the
logarithmic Coulomb interaction, and an additional nonlinear drift
due to non quadratic external potential. When $V=0$ and $\beta=1, 2,
4$, it is the standard Dyson Brownian motion \cite{Dy1, Dy2}. When
$V(x)={x^2\over 2}$ and $\beta>1$, it has been studied by Chan
\cite{Ch}, Rogers and Shi \cite{RS93}, C\'epa and L\'epingle
\cite{CL}, Fontbona \cite{JF1, JF2}, Guionnet \cite{Gui}, Anderson,
Guionnet and Zeitouni \cite{AGZ} and references therein. When $N=\infty$, see \cite{KT}.
By It\^o's calculus, $({\rm GDBM})_V$ is an interacting $N$-particle
system with the Hamiltonian
$$H(x_1, \ldots, x_N):=-\frac{1}{2N}\sum\limits_{1\leq i\neq
j\leq N}\log |x_i-x_j|+\frac{1}{2}\sum\limits_{i=1}^NV(x_i),$$
and the infinitesimal generator of $({\rm GDBM})_V$
is given by
\begin{eqnarray*}
\mathscr{L}^\beta_{N}f=\frac{1}{\beta
N}\sum\limits_{k=1}^N\frac{\partial^2f}{\partial
x_k^2}+\sum\limits_{k=1}^{N}\left({\rm P.V.}
\int_{\mathbb{R}}\frac{L_N(dy)}{x_k-y}-\frac{1}{2}V'(x_k)\right)\frac{\partial
f}{\partial x_k},
\end{eqnarray*}
where $f\in C^2(\mathbb{R}^N)$ and
$L_N=\frac{1}{N}\sum\limits_{i=1}^N\delta_{x_i}\in\mathscr{P}(\mathbb{R})$.
Under suitable condition on $V$, we prove that the SDEs
$(\ref{SDE1})$ for $({\rm GDBM})_V$ admit a unique strong solution
$\lambda_N(t)\in \bigtriangleup_{N}$ with infinite lifetime. See
Theorem \ref{Th1} below. Let
$$L_N(t)=\frac{1}{N}\sum\limits^N_{i=1}\delta_{\lambda^i_N(t)}\in\mathscr{P}(\mathbb{R}), \ \ \ t\in [0, \infty).$$Standard argument shows that the family
$\{L_N(t), t\in [0, T]\}$ is tight on $C([0, T],
\mathscr{P}(\mathbb{R}))$, and the limit of any weakly convergent
subsequence of $L_N(t)$, denoted by $\mu_t$, is a weak solution to
the following nonlinear McKean-Vlasov equation: for all $f\in
C^2_b(\mathbb{R})$,
\begin{eqnarray}\label{DBM7}
\frac{d}{dt}\int_{\mathbb{R}}
f(x)\mu_t(dx)=\frac{1}{2}\int\int_{\mathbb{R}^2}\frac{\partial_xf(x)-\partial_yf(y)}{x-y}\mu_t(dx)\mu_t(dy)-\frac{1}{2}\int_{\mathbb{R}}
V'(x)f'(x)\mu_t(dx).
\end{eqnarray}
In the case $\mu_t$ is absolutely continuous with respect to the
Lebesgue measure on $\mathbb{R}$, integrating by parts, one can
verify that the probability density $\rho_t={d\mu_t\over dx}$
satisfies the following nonlinear McKean-Vlasov equation (also called nonlinear Fokker-Planck equation in the literature)
\begin{eqnarray}
{\partial \rho_t\over \partial t}={\partial \over \partial
x}\left(\rho_t\left({1\over 2}V'-{\rm
H}\rho_t\right)\right),\label{NFK1}
\end{eqnarray}
where
$$
{\rm H}\rho(x)={\rm P.V.}\int_{\mathbb{R}}{\rho(y)\over x-y}dy$$ is
the Hilbert transform of $\rho$.
It seems that one can not find well-established result in the literature
on the uniqueness of weak solutions to the above McKean-Vlasov
equation with general external potential $V$. By lack of this, one
can not find established result in the literature on the Law of
Large Numbers for the GDBM with non quadratic potentials. One of the
main observations of this paper is to find (and prove) the fact
that the McKean-Vlasov equation is indeed the gradient flow of the
Voiculescu free entropy $\Sigma_V$ on the Wasserstein space
$\mathscr{P}_2(\mathbb{R})$ equipped with Otto's infinite
dimensional Riemannian structure, and to use the optimal
transportation theory to prove the uniqueness of weak solutions to
the McKean-Vlasov equation for general potentials $V$ with natural
condition. This allows us to further derive the Law of Large Numbers
for the empirical measures of
the generalized Dyson Brownian motion.
Following Voiculescu
\cite{VD1}and Biane \cite{Bian03}, for every $\mu\in
\mathscr{P}(\mathbb{R})$, we introduce the Voiculescu free entropy
as follows
\begin{eqnarray*}
\Sigma_V(\mu)=-\int_{\mathbb{R}}\int_{\mathbb{R}}
\log|x-y|d\mu(x)d\mu(y)+\int_{\mathbb{R}}V(x)d\mu(x).
\end{eqnarray*}
By \cite{Joh98}, it is well-known that if $V$ satisfies the growth condition
\begin{eqnarray}
V(x)\geq (1+\delta)\log(x^2+1),\ \ \ \ x\in \mathbb{R},\label{grow}
\end{eqnarray}
then there exists a unique minimizer (called the equilibrium
measure) of $\Sigma_V$, denoted by
\begin{eqnarray*}
\mu_V={\rm arg min}_{\mu\in \mathscr{P}(\mathbb{R})} \Sigma_V(\mu).
\end{eqnarray*}
Moreover, $\mu_V$ satisfies the Euler-Lagrange equation
\begin{eqnarray*}
{\rm H}\mu_V(x)={1\over 2}V'(x),\ \ \ \ \forall x\in \mathbb{R}.
\end{eqnarray*}
The relative free entropy is defined as follows
\begin{eqnarray*}
\Sigma_V(\mu|\mu_V)=\Sigma_V(\mu)-\Sigma_V(\mu_V).
\end{eqnarray*}
Following \cite{VD1, Bian03}, the relative free Fisher information
is defined as follows
\begin{eqnarray*}
{\rm I}_V(\mu)=\int_{\mathbb{R}}\left({\rm H}\mu(x)-{1\over
2}V'(x)\right)^2d\mu(x).
\end{eqnarray*}
Note that
\begin{eqnarray*}
{\rm I}_V(\mu_V)=0.
\end{eqnarray*}
We now state the main results of this paper. Our first result
establishes the existence and uniqueness of the strong solution to
SDEs $(\ref{SDE1})$ and the tightness of the associated empirical
measure for a class of $V$ with reasonable condition.
\begin{theorem}\label{Th1} \footnote{Under the condition $-xV'(x)\leq C$ for all $x\in \mathbb{R}$, Rogers and Shi \cite {RS93}
proved the non-collision of the strong solution to $(\ref{SDE1})$, but they did not precisely state the condition $(i)$ which is need for the existence of solution. In \cite{GM}, Graczyk and Malecki proved the existence and uniqueness of strong
solution to SDE $(\ref{SDE1})$ under the assumption that $V'$ is global Lipschitz. The conditions in Theorem \ref{Th1} require that $V'$ satisfies the local monotonicity condition, i.e., $(i)$, and one-side growth condition at infinity, i.e., $(ii)$.
We would like to point out that the local monotonicity condition $(i)$ in Theorem \ref{Th1} is weaker than the condition $V'$ is local Lipschitz, and the one-side growth condition $(ii)$ is
also weaker than the condition $V'$ is global Lipschitz . } Let $V$ be a $C^1$ function satisfying the growth condition $(\ref{grow})$ and the following conditions\\
(i) For all $R>0$, there is $K_R>0,$~such that for all
~$x,y\in\mathbb{R}$ with $|x|, |y|\leq R$, $$(x-y)(V'(x)-V'(y))\geq
-K_R|x-y|^2,$$ (ii) There exists a constant $\gamma>0$ such that
\begin{eqnarray}
-xV'(x)\leq \gamma(1+|x|^2), \ \ \forall \ x\in
\mathbb{R}.\label{Cond1}
\end{eqnarray}
Then, for all $\beta\geq 1$, and for any given $\lambda_N(0)\in
\bigtriangleup_{N}$, there exists a unique strong solution
$(\lambda_N(t))_{t\geq 0}$ taking values in $\bigtriangleup_{N}$
with infinite lifetime to SDEs $(\ref{SDE1})$ with initial value
$\lambda_N(0)$.\\
Moreover, suppose that
$L_N(0)\rightarrow\mu\in\mathscr{P}(\mathbb{R})$ as
$N\rightarrow\infty$, and
\begin{eqnarray*}
\sup\limits_{N\geq 0}\int_{\mathbb{R}}\log(x^2+1)dL_N(0)<\infty.
\end{eqnarray*}
Then, the family $\{L_N(t), t\in [0, T]]\}$ is tight in $\mathcal
{C}([0,T],\mathscr{P}(\mathbb{R}))$, and the limit of any weakly convergent
subsequence of $\{L_N(t), t\in [0, T]]\}$ is a weak solution of the
McKean-Vlasov equation $(\ref{DBM7})$.
\end{theorem}
Inspired by previous works due to Biane \cite{Bian03},
Biane-Speicher \cite{BS01}, Carrillo-McCann-Villani \cite{CMV} (see Theorem 3.1 below) and Sturm \cite{Stu}, we can prove the following result which might be already known by experts even though we cannot find the explicit statement in the literature.
\begin{theorem}\label{GFV}\label{MVGF} For all $V: \mathbb{R}\rightarrow [0,
\infty)$ being a $C^1$ function satisfies the condition
$(\ref{Cond1})$, the nonlinear McKean-Vlasov equation
$(\ref{NFK1})$, i.e.,
\begin{eqnarray*}
{\partial \rho_t\over \partial t}=-{\partial \over \partial
x}(\rho_t({\rm H}\rho_t-\frac{1}{2}V'))
\end{eqnarray*}
is indeed the gradient flow of \ $\Sigma_V$ on the Wasserstein space
$\mathscr{P}_2(\mathbb{R})$.
\end{theorem}
In the optimal transportation theory, it is well known that if the free energy $F$ on the Wasserstein space is $K$-convex, then the
$W_2$-Wasserstein distance between the solutions of the gradient flow $\partial_t \mu=-{\rm grad} F(\mu)$ with initial datas $\mu_1(0)$ and $\mu_2(0)$ satisfies $W_2(\mu_1(t), \mu_2(t))\leq e^{-Kt}W_2(\mu_1(0), \mu_2(0))$.
See \cite{Ot, OV, Stu, StR, Vi1, Vi2}.
In view of this and Theorem \ref{GFV}, and using the Hessian calculation for nonlinear diffusions with interaction on the Wasserstein space as developed in \cite{CMV, Stu}, we can prove the following result, which ensures the uniqueness
of weak solutions to the McKean-Vlasov equation with general
potential $V$ satisfying the condition $V''\geq K$.
\begin{theorem}\label{Th2}
Suppose that $V$ is a $C^2$ function satisfying the same condition as in Theorem \ref{Th1}, and there exists a constant $K\in \mathbb{R}$ such that
$$V''(x)\geq K, \ \ \ \ \forall\ x\in \mathbb{R}.$$
Then the Voiculescu free entropy $\Sigma_V$ on the Wasserstein space $\mathscr{P}_2(\mathbb{R})$ is $K$-convex, i.e., its Hessian on $\mathscr{P}_2(\mathbb{R})$ satisfies
\begin{eqnarray*}
{\rm Hess}_{\mathscr{P}_2(\mathbb{R})}\Sigma_V\geq K.
\end{eqnarray*}
Let $\mu_i(t)$ be two solutions of the McKean-Vlasov equation
$(\ref{NFK1})$ with initial data $\mu_i(0)$, $i=1, 2$. Then for all
$t>0$, we have
\begin{eqnarray*}
W_2(\mu_1(t), \mu_2(t))\leq e^{-Kt}W_2(\mu_1(0), \mu_2(0)).
\end{eqnarray*}
In particular, the Cauchy problem of the McKean-Vlasov equation
$(\ref{NFK1})$ has a unique weak solution.
\end{theorem}
We would like to point out that C\'spa and L\'epingle \cite{CL} proved the uniqueness of weak solution to the McKean-Vlasov equation $(\ref{NFK1})$ with quadratic potential function $V(x)=ax^2+bx$ with two constants, $a\geq 0$ and $b\in \mathbb{R}$, and
Fontbana \cite{JF2} proved the uniqueness of weak solution to the McKean-Vlasov equation $(\ref{NFK1})$ with external potential $V$ such that
$V'(x)=\theta x+b_1(x)$, where $\theta\in \mathbb{R}$ is a constant and $b_1\in C^1(\mathbb{R})$ is a bounded function with bounded derivative. See also \cite{JF1}. Theorem \ref{Th2} establishes the uniqueness of weak solution to the
McKean-Vlasov equation $(\ref{NFK1})$ with more general external potentials $V$ satisfying the condition $V''\geq K$ for some $K\in \mathbb{R}$.
As a consequence of Theorem \ref{Th1} and Theorem \ref{Th2}, we can
derive the Law of Large Numbers for the empirical measures of the
generalized Dyson Brownian motion.
\begin{theorem}\label{Th3}\footnote{For $V(x)=Kx^2$ with $K\in \mathbb{R}$, the result in Theorem \ref{Th3} also holds for $p=2$}
Suppose that $L_N(0)$ weakly converges to $\mu(0)\in \mathscr{P}(\mathbb{R})$. Let $V$ be a $C^2$ function satisfying the same condition as in Theorem \ref{Th1} and $V''\geq K$ for some constant $K\in \mathbb{R}$. Then the empirical measure $L_N(t)$ weakly converges to the unique solution $\mu_t$ of the McKean-Vlasov equation $(\ref{NFK1})$. Moreover, for all $p\in [1, 2)$, we have
\begin{eqnarray*}
W_p(\mathbb{E}(L_N(t), \mu_t)\rightarrow 0 \ \ \ {\rm as}\ \ N\rightarrow \infty,\label{speed1}
\end{eqnarray*}
where the convergence is uniformly with respect to $t\in [0, T]$ for
all fixed $T>0$.
\end{theorem}
The notion of propagation of chaos, which was introduced by M. Kac,
plays a critical role in the study of the large $N$ limit of
$N$-particle systems. According to Sznitman-Tanaka's theorem
\cite{ST}, for exchangeable systems, propagation of chaos is
equivalent to the law of large numbers for the empirical measures of
the system. In view of this and Theorem \ref{Th3}, we have the
following result, which is a dynamic version of a result due to Johansson (Theorem 2 in \cite{Joh98}).
\begin{theorem}\label{THPC}
Assume the conditions in Theorem \ref{Th3} holds. Let $M_{N;k}(t;dx_1,\cdots,dx_k)$ be the $k$-th moment measure for the random probability measure $L_N(t,\cdot)$, that is, for any Borel sets $A_1,\cdots,A_k$,
\begin{eqnarray*}
M_{N;k}(t; A_1,\cdots,A_k):=\mathbb{E}(L_N(t,A_1)\cdots L_N(t,A_k)).
\end{eqnarray*}
Then we have
\begin{eqnarray*}
\lim\limits_{N\rightarrow \infty}\int_{\mathbb{R}^k}\varphi(x_1,\cdots,x_k) M_{N;k}(t;dx_1,\cdots,dx_k)=\int_{\mathbb{R}^k}\varphi(x_1,\cdots,x_k)\mu_t(dx_1)\cdots\mu_t(dx_k)
\end{eqnarray*}
for any continuous, bounded $\varphi$ on $\mathbb{R}^k$.
\end{theorem}
By the ergodic theory of SDE, for a wide class of potentials $V$,
and for any fixed $N$, it is known that $L_N(t)$ converge to $L_N$,
as $t\rightarrow \infty$. On the other hand, the large $N$-limit of
$L_N(t)$, i.e., $\mu_t(dx)=\rho_t(x)dx$, satisfies the nonlinear
Fokker-Planck equation $(\ref{NFK1})$. It is natural to ask the
question whether $\mu_t$ converges to $\mu_V$ in the weak
convergence topology or with respect to the $W_2$-Wasserstein
distance for general potentials $V$. If this is true, then, with
respect to the weak convergence on $\mathscr{P}(\mathbb{R})$ or the
$W_2$-Wasserstein topology on $\mathscr{P}_2(\mathbb{R})$, the
following diagram is commutative
\begin{eqnarray*}
L_N(t)&\Longrightarrow& \mu_t\\
\Downarrow& & {\Downarrow }\\
L_N&\Longrightarrow &\mu_V
\end{eqnarray*}
In other words, we have
\begin{eqnarray*}
\lim\limits_{N\rightarrow \infty}\lim\limits_{t\rightarrow
\infty}L_N(t)= \lim\limits_{t\rightarrow
\infty}\lim\limits_{N\rightarrow \infty}L_N(t).
\end{eqnarray*} In the literature, Chan \cite{Ch} and Rogers-Shi \cite{RS93} proved that this is true
for $V(x)={x^2\over 2}$. See also \cite{AGZ, Gui}. In particular,
this gives a dynamic proof of Wigner's semi-circle law for the
Gaussian Unitary Ensemble. The following result provides some
positive answers to this problem for $C^2$-convex potentials.
\begin{theorem} \label{Th5}(i) Suppose that $V$ is $C^2$-convex, i.e., $V''\geq 0$. Then $\mu_t$ converges to $\mu_V$ with respect to the Wasserstein distance in $\mathscr{P}_2(\mathbb{R})$, i.e.,
\begin{eqnarray*}
W_2(\mu_t, \mu_V)\rightarrow 0 \ \ \ {\rm as}\ \ t\rightarrow
\infty.
\end{eqnarray*}
(ii) Suppose that $V$ is $C^2$ and there exists a constant $K\in
\mathbb{R}$ such that
$$V''(x)\geq K, \ \ \ \forall x\in \mathbb{R}.$$
Then for all $t>0$, we have
\begin{eqnarray*}
\Sigma_V(\mu_t|\mu_V)&\leq& e^{-2Kt}
\Sigma_V(\mu_0|\mu_V),\\
W_2(\mu_t, \mu_V)&\leq& e^{-Kt}W_2(\mu_0, \mu_V).
\end{eqnarray*}
In particular, if $V$ is $C^2$-uniform convex with $V''\geq K>0$, then $\mu_t$ converges to $\mu_V$ with the exponential rate $K$ in the $W_2$-Wasserstein topology on $\mathscr{P}_2(\mathbb{R})$. \\
(iii) Suppose that $V$ is a $C^2$, convex and there exists a
constant $r>0$ such that
$$V''(x)\geq K>0, \ \ \ \ |x|\geq r.$$
Then $\mu_t$
converges to $\mu_V$ with an
exponential rate in the $W_2$-Wasserstein topology on $\mathscr{P}_2(\mathbb{R})$. More precisely, there exist two constants $C_1>0$ and $C_2>0$ such that
\begin{eqnarray*}
W^2_2(\mu_t, \mu_V)\leq {e^{-C_1t}\over C_2}\Sigma_V(\mu_0|\mu_V), \ \ \ \ \ t>0.
\end{eqnarray*}
\end{theorem}
As a corollary of Theorem \ref{Th5}, for $C^2$-convex potentials, we
can give a dynamic proof of the well-known result due to Boutet de
Monvel-Pastur-Shcherbina \cite{BPS} and Johansson \cite{Joh98}.
Their result says that, for $V$ satisfying the growth condition
$(\ref{grow})$, the empirical measure $L_N={1\over
N}\sum\limits_{i=1}^N \delta_{x_i}$ weakly converges to the
equilibrium measure $\mu_V$, where $(x_i, i=1, \ldots, N)$,
satisfies the following probability distribution
\begin{eqnarray*}
P^N_\beta(dx_1,\ldots,dx_N)=\frac{1}{Z^\beta_N} \Pi_{i\neq
j}|x_i-x_j|^{\frac{\beta}{2}}\exp\left(-\frac{\beta N
}{2}\sum\limits_{i=1}^NV(x_i)\right)\prod_{i=1}^N dx_i,
\end{eqnarray*}
where $\beta>0$ is a parameter. We would like to mention that, for
non-convex potentials $V$, we do not know how to give a dynamic
proof of the above result. We would like to mention a recent paper by Bourgade, Erd\"os, and Yau \cite{BEY} in which the authors proved
the bulk universality of the $\beta$-ensembles with non-convex regular analytic potentials for any $\beta>0$. Whether or not their idea of introducing a ``convexified measure" can be used to
extend the results in this paper to non-convex case, will be an interesting problem for study in future.
Finally, let us mention that, for $\beta=2$ and for real analytic
function $V$, we can prove that the generalized Dyson Brownian
motion can be realized as the eigenvalues process of the $N\times N$
real Hermitian matrix valued diffusion process defined by
\begin{eqnarray*}
dX^N_t={1\over \sqrt{N}}dB^N_t-{1\over 2}V'(X^N_t)dt,\label{DBM1}
\end{eqnarray*}
where $B^N_t$ is the $N\times N$ Hermitian matrix valued Brownian
motion. Moreover, we can prove that $X_t^N$ converges in
distribution to the free diffusion process $X_t$, which was defined
by Biane and Speicher \cite{BS01}. This extends a famous result, due
to Voiculescu \cite{VD2, VD3} and Biane \cite{Bian95}, which states
that the renormalized Hermitian Brownian motion ${1\over
\sqrt{N}}B^N_t$ converges in distribution to the free Brownian
motion $S_t$. See \cite{LX2014}.
The rest of this paper is organized as follows. In Section $2$, we
prove Theorem \ref{Th1}. In Section $3$, we prove Theorem \ref{GFV},
Theorem \ref{Th2} and Theorem \ref{Th3}, Theorem \ref{THPC}. In
Section $4$, we prove Theorem \ref{Th5}. In Section $5$, we discuss
the case of double-well potential and raise some conjectures. Finally, let us mention that this paper is an update revised version of our previous paper entitled Generalized Dyson Brownian motion,
McKean-Vlasov equation and eigenvalues of random matrices (arxiv.org/abs/1303.1240v1).
\section{Proof of Theorem \ref{Th1}}
The proof of Theorem \ref{Th1} is adapted from classical argument coming back to McKean and exposed in \cite{RS93, CL, AGZ}. \\
\noindent{\bf Proof of existence and uniqueness of GDBM}. First, for
fixed $R>0$, let~$\phi_R(x)=
x^{-1}$~ if~$|x|\geq R^{-1}$,~
and~$\phi_R(x)=R^2x$ if $|x|< R^{-1}$. Since ~$\phi_R$~is uniformly Lipschitz and $V$ satisfies $(i)$ and $(ii)$, by Theorem 3.1.1 in \cite{Ro}, the following SDE for the truncated Dyson Brownian
motion
\begin{equation}
d\lambda^i_{N,R}(t)=\sqrt{\frac{2}{\beta
N}}dW^i_{t}+\frac{1}{N}\sum\limits_{j:j\neq i}
\phi_R(\lambda^i_{N,R}(t)-\lambda^j_{N,R}(t))dt-\frac{1}{2}V'(\lambda^i_{N,R}(t))dt,\label{E2.1}
\end{equation}
with $\lambda^i_{N,R}(0)=\lambda^i_N(0)$~for~$1\leq i\leq N$,
has a unique strong
solution. Let
\begin{eqnarray*}
\tau_R:=\inf\{t:\min_{i\neq
j}\mid\lambda^i_{N,R}(t)-\lambda^j_{N,R}(t)\mid<R^{-1}\}.
\end{eqnarray*}
Then $\tau_R$ is monotone increasing in $R$ and $\lambda_{N,R}(t)=\lambda_{N,R'}(t)$ for all $t\leq
\tau_R$ and $R<R'$.
Second, let $\lambda_N(t)=\lambda_{N, R}(t)$ on $t\in [0, \tau_R)$.
To prove that $\lambda_N(t)$ is a global solution
to SDE $(\ref{SDE1})$, we need
only to prove $\lambda_N(t)$ does not explode, and $\lambda_N^i(t)$ and $\lambda_N^j(t)$ never collide for all $t>0$, $i\neq j$.
To prove that $\lambda_N(t)$ does not explode, let
$R_t=\frac{1}{2N}\sum\limits_{j=1}^N\lambda_N^j(t)^2$. By It\^{o}'s
formula, and by Levy's characterization, we can introduce a new
Brownian motion~$B$,~such that ~$$dR_t={2\over N}\sqrt{R_t\over
\beta}dB_t+\left(\frac{1}{\beta N}+\frac{N-1}{2N}-\frac{1}{2}\langle
L_N(t), xV'(x)\rangle\right)dt.$$ Let $R'$ be the solution of
$$dR'_t={2\over N}\sqrt{R'_t\over \beta}dB_t+\left(\frac{1}{\beta
N}+\frac{N-1}{2N}+\frac{1}{2}\gamma+\gamma R'_t\right)dt,$$ with
$R'_0=R_0.$ Under the assumption $(\ref{Cond1})$, and using the
comparison theorem of one dimensional SDEs, cf. \cite{IW}, we can
derive that
$$R_t\leq
R'_t,\ \ \ \forall \ t\geq 0,~~{\rm a.s}.$$Moreover, by Ikeda and
Watanabe \cite{IW} (p. 235-237), the process $R'$ never explodes. So
the process $R$ (and hence $\lambda_N(t)$) ~does not explode in
finite time \footnote{ In \cite{RS93}, Rogers and Shi proved the
non-explosion of GDBM for $V$ satisfying $-xV'(x)\leq \gamma$,
$\forall x\in \mathbb{R}$. }.
To prove that $\lambda_N^i(t)$ and $\lambda_N^j(t)$ never collide for all $t>0$, $i\neq j$, let us introduce the Lyapunov function $f(x_1,\ldots,x_N)=\frac{1}{N}\sum\limits_{i=1}^N
V(x_i)-\frac{1}{N^2}\sum\limits_{i\neq j}\log|x_i-x_j|$. Similarly to \cite{Gui, AGZ}, we can prove
\begin{eqnarray*}
df(\lambda_N(t))&=&dM_N(t)+\frac{1}{N^3}\left(\frac{1}{\beta}-1\right)\sum\limits_{k\neq
i}\frac{1}{(\lambda_N^i(t)-\lambda_N^k(t))^2}dt-\frac{1}{2N}\sum\limits_{i=1}^N|V'(\lambda_N^i(t))|^2dt\\
& & +\frac{1}{N^2}\left({1\over \beta}\sum\limits_{i=1}^N
V''(\lambda_N^i(t))+\frac{3}{2}\sum\limits_{j\neq
i}\frac{V'(\lambda_N^i(t))-V'(\lambda_N^j(t))}{\lambda_N^i(t)-\lambda_N^j(t)}\right)dt,
\end{eqnarray*}
where~$M_N$~is the following local
martingale~$$dM_N(t)=\frac{2^{\frac{1}{2}}}{\beta^{\frac{1}{2}}
N^{\frac{3}{2}}}\sum\limits_{i=1}^N \left(
V'(\lambda_N^i(t))-\frac{1}{N}\sum\limits_{k:k\neq
i}\frac{1}{\lambda_N^i(t)-\lambda_N^k(t)} \right)dW_t^i.$$
Fix $K>0$ and $R>0$ such that $\lambda_N^i(0)\in
[-K, K]$ and $|\lambda_N^i(0)-\lambda_N^j(0)|\geq R^{-1}$ for all $i\neq j$, $i,
j=1, \ldots, N$. Let $C_1(K)\geq 0$ be such that $\sup\limits_{x\in [-K, K]}V''(x)\leq
C_1(K)$. Let $A_N(t)dt=df(\lambda_N(t))-dM_N(t)$, and
$\zeta_K=\inf\limits\{t\geq 0: \lambda_N^i(t)\notin [-K, K],\ {\rm
for\ some}\ i=1, \ldots, N\}$, then for any fixed $T>0$, $\sup\limits_{t\in [0, T]}A_N(t\wedge \zeta_K)\leq C_1(K)$ and
$\{f(\lambda_N(t\wedge \zeta_K)-C_1(K)(t\wedge \zeta_K),\ \ t\in [0,
T]\}$ is a supermartingale. Let $C_2(K):= \inf\limits\{V(x):|x|\leq K\}$, we can prove
\begin{eqnarray*}
\mathbb{P}(\tau_R\leq \zeta_K\wedge T)\leq
\frac{N^2(f(\lambda_N(0))+TC_1(K))+N(N-1)\log(2K)-C_2(K)}{\log(2K)+\log R}.
\end{eqnarray*}
Letting $R$, $T$ and $K$ tend to infinity, we can prove $\mathbb{P}(\tau_\infty<\zeta)=0$, where $\zeta:=\inf\{t:\lambda_N^i(t)=\lambda_N^j(t)~\exists~1\leq i\neq j\leq N\}$.
This proves that $\lambda_N^1(t), \ldots, \lambda_N^N(t)$ does not collide.
Finally, by the continuity of the trajectory of $\lambda_N(t)$, we
have $\lambda_N(t)\in\triangle_N$ for all $t\geq 0$. The same
argument as used in the proof of Theorem 12.1 in \cite{Gui} proves
the uniqueness of the weak solution to SDEs $(\ref{SDE1})$. The
proof of Theorem \ref{Th1} is completed.
\medskip
\noindent{\bf Proof of tightness and identification of McKean-Vlasov
limit}
We follow the argument used in \cite{RS93} to prove the tightness of
$\{L_N(t),t\in[0,T]\}$. Let us pick functions $f_j\in
C_b^\infty(\mathbb{R}, \mathbb{C}), j=1,2,\ldots,$ which is dense in
$C_b(\mathbb{R})$. Thus
$$\langle\mu, f_j\rangle=\langle\mu', f_j\rangle, \ \ \forall
j\Rightarrow \mu=\mu'.$$ We also pick a $C^\infty$ function
$f_0:\mathbb{R}\rightarrow[1, \infty)$ with the properties
$$f_0(x)=f_0(-x),\ \ \ f_0(x)\rightarrow\infty \ \ \ {\rm as} \ \ x\rightarrow\infty, \ x\in \mathbb{R}^{+}.$$ Taking test functions in the
Schwartz class of smooth functions whose derivatives (up to second order) are rapidly decreasing, we
may assume that
$$f_j, \ f_j'', \ V'f'_j\ \ \ {\rm are\ \ uniformly \ bounded\ \ for\ \ all}\ \
j\geq1.$$ By Ethier and Kurtz \cite{EK} (p.107), to prove the
tightness of $\{L_N(t),\ t\in [0, T],\ N\geq 1\}$, it is sufficient
to prove that for each $j$ the sequence of continuous real-valued
functions
$$\{\langle L_N(t), f_j\rangle,\ t\in [0, T],\ N\geq1\}$$ is relatively compact.
To this end, note that, by the first part of Theorem \ref{Th1},
there is non-collision and non-explosion for the particles
$\lambda_N^i(t)$ for all $t\in [0, \infty)$. By It\^{o}'s formula,
we have
\begin{eqnarray}\label{LN}
d\langle L_N(t), f\rangle&=&\frac{1}{N}\sqrt{\frac{2}{\beta
N}}\sum\limits_{i=1}^Nf'(\lambda_N^i(t))dW_t^i+ \left\langle L_N(t),
\left(\frac{2}{\beta}-1\right)\frac{1}{2N}f''-\frac{1}{2}V'f'\right\rangle
dt\nonumber\\
& &\hskip2cm
+\frac{1}{2}\int\int_{\mathbb{R}^2}\frac{f'(x)-f'(y)}{x-y}L_N(t,dx)L_N(t,dy)dt.
\end{eqnarray}This yields
\begin{eqnarray}
\langle L_N(t), f_j\rangle&=&\langle L_N(0),
f_j\rangle+\frac{1}{2}\int_0^t\int_{\mathbb{R}}\int_{\mathbb{R}}\frac{f_j'(x)-f_j'(y)}{x-y}L_N(s,dx)L_N(s,dy)ds\nonumber\\
& &-\frac{1}{2}\int_0^t\langle L_N(s), V'f_j'\rangle
ds+\int_0^t\left\langle L_N(s),
\left(\frac{2}{\beta}-1\right)\frac{1}{2N}f_j''\right\rangle ds+M_N^{f_j}(t)\nonumber\\
&=&I_1(N)+I_2(N)+I_3(N)+I_4(N)+M_N^{f_j}(t),\label{E3.A}
\end{eqnarray}
where $$M_N^{f_j}(t)=\frac{1}{N}\sqrt{\frac{2}{\beta
N}}\int_0^t\sum\limits_{i=1}^Nf_j'(\lambda_N^i(s))dW_s^i.$$ Note
that, as $L_N(0)$ is weakly convergent, $I_1(N)$ is convergent. By
the assumption that $f_j$ and $f_j''$ are uniformly bounded (hence
$f_j'$ are uniformly bounded) , we can easily show that
$\{M_N^{f_j}(t), t\in [0, T]\}$ and $I_4(N)$ converge to zero.
Moreover, by the assumption that $V'f_j'$ and $f_j''$ are uniformly
bounded, the Arzela-Ascoli theorem implies that $I_2(N)$ and
$I_3(N)$ are relatively compact in $C([0, T], \mathbb{R})$. Thus the sequence
$\{(L_N(t))_{t\geq0}:\ N\geq1\}$ is tight in $C([0, T],
\mathbb{R})$. Tightness also follows for $j=0$ if we have
$$\langle L_N(0),f_0\rangle\rightarrow\ \ {\rm finite\ \ limit\ \ as}\ \ N\rightarrow\infty.$$
So let us suppose that the initial distribution $L_N(0)$ have the
property $\langle L_N(0),f_0\rangle\leq K$ for some $K,$ for all
$N.$ For given $\mu_0,$ we could always find $L_N(0)$ and $f_0$ to
satisfy this and the other conditions, and this gives the tightness
for $j=0$ also.
Finally, we identify the limit process of any weakly convergent
subsequence of $\{L_N(t)\}$. Assuming that $\{L_{N_j}(t), t\in [0,
T]\}$ is a weakly convergent subsequence in $C([0, T],
\mathscr{P}(\mathbb{R}))$. Then, for all $f\in C_b^2(\mathbb{R})$,
the It\^o's formula $(\ref{E3.A})$ and the above argument show that
$\langle \mu_t,f\rangle=\lim\limits_{j\rightarrow \infty}\langle
L_{N_j}(t),f\rangle$ satisfies the following equation
\begin{eqnarray*}
\int_{\mathbb{R}} f(x)\mu_t(dx)&=&\int_{\mathbb{R}}
f(x)\mu_0(dx)+\frac{1}{2}\int_0^t\int\int_{\mathbb{R}^2}\frac{\partial_xf(x)-\partial_yf(y)}{x-y}\mu_s(dx)\mu_s(dy)ds\\
& &\ \ \ \ \ \ \ -\frac{1}{2}\int_0^t\int_{\mathbb{R}}
V'(x)f'(x)\mu_s(dx)ds.
\end{eqnarray*}
This proves that $\mu_t$ is a weak solution to the McKean-Vlasov
equation $(\ref{DBM7})$. The proof of Theorem \ref{Th1} is
completed.\hfill $\square$
\section{McKean-Vlasov equation: gradient flow and uniqueness}
To characterize the McKean-Vlasov limit $\mu_t$, we need only to use
the test function $f(x)=(z-x)^{-1}$, where
$z\in\mathbb{C}\backslash\mathbb{R}$, instead of using all test
functions $f\in C_b^2(\mathbb{R})$ in the McKean-Vlasov equation
$(\ref{DBM7})$. Let
\begin{eqnarray*}
G_t(z)=\int_{\mathbb{R}} {\mu_t(dx)\over z-x}
\end{eqnarray*}
be the Stieltjes transform of $\mu_t$. Then $G_t(z)$ satisfies the
following equation
\begin{eqnarray}
{\partial\over \partial t}G_t(z)=-G_t(z) {\partial\over \partial
z}G_t(z)-\frac{1}{2}\int_{\mathbb{R}} {V'(x)\over
(z-x)^2}\mu_t(dx).\label{CHSV-1}
\end{eqnarray}
In particular, in the case $V(x)=\theta x^2$, since
\begin{eqnarray*}
-\int_{\mathbb{R}} {x\over (z-x)^2}\mu_t(dx)=z{\partial\over
\partial z}G_t(z)+G_t(z),
\end{eqnarray*}
the Stieltjes transform of $\mu_t$ satisfies the complex Burgers
equation
\begin{eqnarray}
{\partial\over \partial t}G_t(z)=\left(-G_t(z)+\theta z\right)
{\partial\over
\partial z}G_t(z)+\theta G_t(z).\label{CHSV-2}
\end{eqnarray}
In \cite{Ch, RS93}, Chan and Rogers-Shi proved that the complex
Burgers equation $(\ref{CHSV-2})$ (equivalently, the McKean-Vlasov
equation with potential $V(x)=\theta x^2$) has a unique solution,
and $\lim\limits_{t\rightarrow \infty}G_t(z)$ exists and coincides
with the Stieltjes transform of the Wigner semi-circle law
$\mu_{SC}$. This yields a dynamic proof of the Wigner's theorem,
i.e., $L_N(\infty)$ weakly converges to $\mu_{SC}$.
However, for non quadratic potential $V$, $\int_{\mathbb{R}}
{V'(x)\over (z-x)^2}\mu_t(dx)$ in $(\ref{CHSV-1})$ cannot be
expressed in terms of $G_t(z)$ and its derivatives with respect to
$z$. Thus, one cannot derive an analogue of the complex Burgers
equation $(\ref{CHSV-2})$ for non quadratic potential $V$, and we
need to find a new approach to prove the uniqueness of the weak
solutions of the Mckean-Vlasov equation for general potential $V$.
In this section, we use the theory of gradient flow on the
Wasserstein space $\mathscr{P}_2(\mathbb{R})$ and the optimal
transportation theory to study this problem.
\subsection{Proof of Theorem \ref{GFV}}
By Theorem \ref{Th1}, we have proved the existence of weak solution
to the McKean-Vlasov equation $(\ref{DBM7})$. Assuming that the weak
solution $\mu_t$ of the McKean-Vlasov equation $(\ref{DBM7})$ is
absolutely continuous with respect to the Lebesgue measure $dx$, we
derive the existence of the weak solution of the nonlinear
Fokker-Planck equation $(\ref{NFK1})$. Thus, to prove the law of
large numbers for $L_N(t)$, we need only to show the uniqueness of
the nonlinear Fokker-Planck equation $(\ref{NFK1})$. Note that,
letting
\begin{eqnarray*}
W(x)=-2\log |x|, \ \ \ \ x\neq 0,
\end{eqnarray*}
then the nonlinear Fokker-Planck equation $(\ref{NFK1})$ can be
rewritten as follows
\begin{eqnarray}
\partial_t \rho=\nabla\cdot (\rho \nabla (V+W*\rho)).\label{MV-0}
\end{eqnarray}
To study the uniqueness and the longtime behavior of the nonlinear
Fokker-Planck equation $(\ref{NFK1})$ (i.e., $(\ref{MV-0})$), we
first recall Otto's infinite dimensional Riemannian structure on the
Wasserstein space $\mathscr{P}_2(\mathbb{R}^d)$. Fix $fdx\in
\mathscr{P}_2(\mathbb{R}^d)$, the tangent space of
$\mathscr{P}_2(\mathbb{R}^d)$ at $fdx$ is given by
$$
T_{fdx}\mathscr{P}_2(\mathbb{R}^d)=\{sdx: s\in W^{1,
2}(\mathbb{R}^d, \mathbb{R}), \ \ \int_{\mathbb{R}}sdx=0\}.$$ By
\cite{Ot}, for all $s_idx\in T_{fdx}\mathscr{P}_2(\mathbb{R}^d)$,
$i=1, 2$, there exist a unique $p_i\in W^{1,2}(\mathbb{R}^d,
\mathbb{R}^d)$, $i=1, 2$, such that
$$s_i=-\nabla.(f \nabla p_i)$$
In view of this, Otto's infinite dimensional Riemannian metric on
$T_{fdx}\mathscr{P}_2(\mathbb{R}^d)$ is defined by
\begin{eqnarray*}
g_{fdx}(s_1, s_2)=\int_{\mathbb{R}^d} \langle \nabla p_1, \nabla
p_2\rangle fdx.
\end{eqnarray*}
Next we recall some results and ideas that we borrow from
\cite{CMV}, in which Carrillo, McCann and Villani studied the
following type McKean-Vlasov evolution equation of the granular
media
\begin{eqnarray}
\partial_t \rho=\nabla\cdot (\rho \nabla (\log \rho+V+W*\rho)). \label{MV}
\end{eqnarray}
They proved that the McKean-Vlasov evolution equation can be
realized as a gradient flow of a free energy functional on the
infinite Wasserstein space. More precisely, they proved
\begin{theorem} (Carrillo-McCann-Villani\cite{CMV})\label{th0} Let $V, W$ be
nice functions on $\mathbb{R}^d$, and
\begin{eqnarray}
F(f)=\int_{\mathbb{R}^d} \rho\log \rho dv+\int_{\mathbb{R}^d} \rho V
dv+{1\over 2}\int_{\mathbb{R}^d}\int_{\mathbb{R}^d}
W(x-y)\rho(x)\rho(y)dxdy.\label{F}
\end{eqnarray}
Then the McKean-Vlasov equation $(\ref{MV})$ is the gradient flow of
$F$ with respect to Otto's infinite dimensional Riemannian metric
on $\mathscr{P}_2(\mathbb{R}^d)$.
\end{theorem}
Moreover, based on Otto's infinite dimensional geometric calculation
on the Wasserstein space, Carrillo, McCann and Villani \cite{CMV}
proved the following entropy dissipation formula
\begin{theorem}(Carrillo-McCann-Villani\cite{CMV}) \label{th1} Denote $\xi:=\nabla(\log \rho+V+W*\rho)$. The
\begin{eqnarray}
{d\over dt} F(\rho_t)&=&-\int_{\mathbb{R}^n} \rho |\xi|^2dv, \label{ED1}\\
{d^2\over dt^2} F(\rho_t)&=&2\int_{\mathbb{R}^n}\rho {\rm Tr}(D\xi)^T(D\xi)dx+2\int_{\mathbb{R}^n} \langle D^2V\cdot \xi, \xi\rangle\rho dx\nonumber\\
& &+\int_{\mathbb{R}^{2n}} \langle D^2W(x-y)\cdot [\xi(x)-\xi(y)],
[\xi(x)-\xi(y)]\rangle d\rho(x)d\rho(y). \label{ED2}
\end{eqnarray}
\end{theorem}
\medskip
Inspired by the earlier works due to Biane \cite{Bian03} and
Biane-Speicher \cite{BS01}, and Carrillo-McCann-Villani \cite{CMV},
we can prove the following results, which play a crucial r\^ole in
the proof of the main results of this paper.\footnote{In
\cite{BS01}, Biane and Speicher gave a heuristic proof of the fact
that the probability density of the large $N$-limit of $L_N(t)$
satisfies the McKean-Vlasov equation $(\ref{NFK1})$ (called the free Fokker-Planck equation in \cite{BS01}). Theorem
\ref{Th1} says that $\mu_t$ satisfies the McKean-Vlasov equation
$(\ref{DBM7})$ and integration by parts shows that $\rho_t$
satisfies $(\ref{NFK1})$. Combining this with
Carrillo-McCann-Villani's result in Theorem \ref{th0}, we obtained
Theorem \ref{MVGF} in August 2012.}
\begin{theorem}\label{MVGF} For all $V: \mathbb{R}\rightarrow [0,
\infty)$ being a $C^2$ function satisfies the condition
$(\ref{Cond1})$, the nonlinear Fokker-Planck equation
$(\ref{NFK1})$, i.e.,
\begin{eqnarray*}
{\partial \rho_t\over \partial t}=-{\partial \over \partial x}(\rho_t({\rm
H}\rho_t-\frac{1}{2}V'))
\end{eqnarray*}
is indeed the gradient flow of $\Sigma_V$ on the Wasserstein space
$\mathscr{P}_2(\mathbb{R})$.
\end{theorem}
{\it Proof}. Taking $W(x)=2\log|x|^{-1}$, and noting that $\nabla (W*\rho)=H\rho$, Theorem
\ref{th1} follows from Theorem \ref{th0}. \hfill
$\square$
\begin{theorem} \label{th2} Under the notation of Theorem
\ref{GFV}, we have
\begin{eqnarray}
\label{entro-diss-3}
{d\over dt}\Sigma_V(\mu_t|\mu_V)&=&-2\int_{\mathbb{R}}\left[V'(x)-2{\rm H}\rho_t(x)\right]^2\rho_t(x)dx,\\
{d^2\over dt^2}\Sigma_V(\mu_t|\mu_V)&=&2\int_{\mathbb{R}} V^{''}(x)|V'(x)-2{\rm H}\rho_t(x)|^2\rho_t(x) dx\nonumber\\
& &+\int\int_{\mathbb{R}^2} {\left[V'(x)-V'(y)-2({\rm
H}\rho_t(x)-{\rm H}\rho_t(y))\right]^2\over (x-y)^2}
\rho_t(x)\rho_t(y)dxdy.
\end{eqnarray}
\end{theorem}
{\it Proof}. By analogue of the proof of Theorem \ref{th1} in
\cite{CMV}, and observing that for $W(x)=-2\log |x|$, we have
$\xi:=\nabla(V+W*\rho) =V'-2{\rm H}\rho$, we can prove Theorem
\ref{th2}. \hfill $\square$
\medskip
\subsection{Proof of Theorem \ref{Th2}}
The proof follows the same argument as used in \cite{Ot, OV, CMV}.
We use the fact that the nonlinear Fokker-Planck equation
$(\ref{NFK1})$ is the gradient flow of the Voiculescu entropy
$\Sigma_V$ on the Wasserstein space $\mathscr{P}_2(\mathbb{R})$, and
that $\Sigma_V$ is $K$-convex along the geodesic displacement
between two probability measures in $\mathscr{P}_2(\mathbb{R})$.
\begin{theorem}\label{K}\footnote{After we proved Theorem \ref{K} in August 2012, we noticed later from Villani's book \cite{Vi2} that Blower \cite{Blo} has proved the $K$-convexity of the Voiculescu entropy.}\label{K-convex} Assuming that $V\in C^2(\mathbb{R},
\mathbb{R}^+)$ and there exists a constant $K\in \mathbb{R}$ such
that $V''\geq K$. Then
$$
{\rm Hess}_{\mathscr{P}_2(\mathbb{R})} \Sigma_V(\mu) \geq K.
$$
\end{theorem}
{\it Proof}. Let $(\rho_s, v_s)$ be the solution to the following continuity equation and the Hamilton-Jacobi equation
\begin{eqnarray*}
\partial_s\rho+\nabla\cdot(\rho v)&=&0,\\
\partial_s (\rho v)+\nabla\cdot(\rho v\otimes v)&=&0.
\end{eqnarray*}
Let $\mu_s=\rho_sdx$.
By analogue of the calculus of the Hessian of the free energy in \cite{CMV}, we can prove that
\begin{eqnarray*}
{d^2\over ds^2}\Sigma_V(\mu_s)=\int_{\mathbb{R}} V^{''}(x)|v_s(x)|^2\rho_s(x) dx+{1\over 2}\int\int_{\mathbb{R}^2} {|v_s(x)-v_s(y)|^2\over (x-y)^2} \rho_s(x)\rho_s(y)dxdy.
\end{eqnarray*}Thus, under the assumption $V''\geq K$, we have
\begin{eqnarray*}
{\rm Hess}_{\mathscr{P}_2(\mathbb{R})}
\Sigma_V(\mu)(v, v)=\left. {d^2\over ds^2}\Sigma_V(\mu_s)\right|_{s=0}\geq K\int_{\mathbb{R}} |v(x)|^2\rho_s(x) dx=K\|v\|^2.
\end{eqnarray*} \hfill $\square$
\begin{proposition}\label{MVI}
Let $\mu(0)(dx)=\rho(0)dx$ and $\mu(1)(dx)=\rho(1)dx$ be two probability
measures with compact support on $\mathbb{R}$, let
$\mu(s)(dx)=\rho(s)dx$ be the unique geodesic in the Wasserstein space
$\mathscr{P}_2(\mathbb{R})$ linking $\mu(0)$ and $\mu(1)$. Then
\begin{eqnarray}
\left.\left\langle \frac{d\rho(s)}{ds}, {\rm grad}_W \Sigma_V
(\rho(s))\right\rangle\right|_{s= 1} - \left.\left\langle
\frac{d\rho(s)}{ds}, {\rm grad}_W \Sigma_V(\rho(s))
\right\rangle\right|_{s= 0} \geq K W_{2}^{2}(\rho(0),
\rho(1)),\label{DDD1}
\end{eqnarray}
where ${\rm grad}_W\Sigma_V$ denotes the gradient of $\Sigma_V$ on
the Wasserstein space $\mathscr{P}_2(\mathbb{R})$ equipped with
Otto's infinite dimensional Riemannian metric.
\end{proposition}
{\it Proof}. By the assumptions, we have
\begin{eqnarray*}
{d^2\over ds^2}\Sigma_V(\mu(s)) ={\rm
Hess}_{\mathscr{P}_2(\mathbb{R})}
\Sigma_V(\rho(s))\left(\frac{\partial \rho(s)}{\partial s},
\frac{\partial \rho(s)}{\partial s}\right) \geq
K\left\|\frac{\partial \rho(s)}{\partial
s}\right\|^{2}_{\mathscr{P}_2(\mathbb{R})}.
\end{eqnarray*}
By the mean value theorem, for some $\sigma^*\in (0, 1)$,
\begin{eqnarray*}
\Sigma_V(\rho(1)) - \Sigma_V(\rho(0)) &=& \left.{d\over ds}\right|_{s=0}\Sigma_V(\rho(s)) + {1\over 2}\left.{d^2\over ds^2}\right|
_{s=\sigma^*}\Sigma_V(\rho(\sigma))\\
&\geq& \left.\left\langle \frac{d\rho(s)}{ds}, {\rm grad}_W
\Sigma_V(\rho(s)) \right\rangle\right|_{s= 0} +
{K \over 2}\int^{1}_{0}\left\|\frac{\partial \rho(s)}{\partial s}\right\|^{2}_{\mathscr{P}_2(\mathbb{R})}d\sigma \\
&=&\left.\left \langle \frac{d\rho(s)}{ds}, {\rm
grad}_W\Sigma_V(\rho(s)) \right\rangle\right|_{s= 0}
+ {K \over 2} W_{2}^{2}(\rho(0), \rho(1)).
\end{eqnarray*}
Similarly,
\begin{eqnarray*}
\Sigma_V(\rho(0)) - \Sigma_V(\rho(1)) &\geq& -\left.\left\langle
\frac{d\rho(s)}{ds}, {\rm grad}_W \Sigma_V(\rho(s))
\right\rangle\right|_ {s= 1} + {K\over 2} W_{2}^{2}(\rho(0),
\rho(1)).
\end{eqnarray*}
Summing the two inequalities together, we obtain $(\ref{DDD1})$.
\hfill $\square$
\medskip
We are ready to give the proof of Theorem \ref{Th2} as follows.\\
{\bf Proof of Theorem \ref{Th2}}. Let $\rho_{t}(s, x)dx:[0, 1]
\rightarrow \mathscr{P}_2(\mathbb{R})$ be the unique geodesic
between $\mu_1(t)$ and $\mu_2(t)$. By Otto \cite{Ot}, we have the
following derivative formula of the Wasserstein distance
\begin{eqnarray*}
{d \over dt}W^{2}_{2}(\mu_1(t), \mu_2(t)) &=&
-2\int_{\mathbb{R}}\left.\left\langle\frac{d\rho_t(s)}{ds}(x),
\xi_t\right\rangle\right|_{s=0}d\mu_1(t)
+ 2\int_{\mathbb{R}}\left.\left\langle \frac{d\rho_t(s)}{ds}(x), \xi_t\right\rangle\right|_{s=1} d\mu_2(t)\\
&=& 2 \left.\left\langle \frac{d\rho_t(s)}{ds}(x), {\rm grad}_W
\Sigma_V(\mu_2(t))\right\rangle\right|_{s=0}-2\left.\left\langle
\frac{d\rho_t(s)}{ds}(x), {\rm grad}_W
\Sigma_V(\mu_1(t))\right\rangle\right|_{s=1}.
\end{eqnarray*}
By Proposition \ref{MVI}, we have
\begin{eqnarray*}
{d \over dt}W^{2}_{2}(\mu_1(t), \mu_2(t))
\leq -2KW_{2}^{2}(\mu_1(t),
\mu_2(t)).
\end{eqnarray*}
The Gronwall inequality implies
\begin{eqnarray*}
W_{2}(\mu_1(t), \mu_2(t)) \leq e^{-Kt}W_{2}(\mu_1(0), \mu_2(0)).
\end{eqnarray*}
As a consequence, the McKean-Vlasov equation $(\ref{DBM7})$ has a
unique weak solution. This finishes the proof of Theorem \ref{Th2}.
\hfill $\square$
\subsection{Proof of Theorem \ref{Th3}}
{\bf Proof of Theorem \ref{Th3}}. By Theorem \ref{Th1}, the family
$\{L_N(t), t\in [0, T]\}$ is tight with respect to the weak
convergence topology on $\mathscr{P}(\mathbb{R})$, and the limit of
any weakly convergent subsequence of $\{L_N(t), t\in [0, T]\}$ is a
weak solution of $(\ref{DBM7})$. By the uniqueness of weak solutions
to $(\ref{NFK1})$, we conclude that $L_N(t)$ weakly converges to
$\mu_t$, and hence $\mathbb{E}[L_N(t)]$ weakly converges to $\mu_t$
as $N\rightarrow \infty$.
Taking $f(x)=x^2$ in $(\ref{DBM7})$ and $(\ref{LN})$ respectively,
we can derive that
\begin{eqnarray}
\label{mut}
\frac{d}{dt}\int_{\mathbb{R}}x^2\mu_t(dx)=1-\int_{\mathbb{R}}
xV'(x)\mu_t(dx),\label{DBM18}
\end{eqnarray}
and
\begin{eqnarray}\label{LN8}
d\langle L_N(t), x^2\rangle&=&\frac{2}{N}\sqrt{\frac{2}{\beta
N}}\sum\limits_{i=1}^N\lambda_N^i(t)dW_t^i+ \left\langle L_N(t),
\left(\frac{2}{\beta}-1\right)\frac{1}{N}-xV'\right\rangle dt+1.
\end{eqnarray}
Taking expectation, we have
\begin{eqnarray}\label{LN8a}
{d\over dt}\int_{\mathbb{R}} x^2 \mathbb{E}[L_N(t, dx)]=1+
\left(\frac{2}{\beta}-1\right)\frac{1}{N}-\int_{\mathbb{R}}xV'(x)L_N(t,
dx).
\end{eqnarray}
On the other hand, from the proof of Theorem \ref{Th1}, we have
\begin{eqnarray*}
\int_{\mathbb{R}} x^2 \mathbb{E}[L_N(t)]=\mathbb{E}\left[{1\over N}\sum\limits_{i=1}^N \lambda_N^i(t)^2\right]=\mathbb{E}[R_t]\leq\mathbb{E}[R'_t].
\end{eqnarray*}
Note that
$$dR'_t={2\over N}\sqrt{R'_t\over \beta}dB_t+\left(\frac{1}{\beta
N}+\frac{N-1}{2N}+\frac{1}{2}\gamma+\gamma R'_t\right)dt,$$
which yields
\begin{eqnarray*}
{d\over dt}\mathbb{E}[R'_t]={1\over \beta N}+\frac{N-1}{2N}+\frac{1}{2}\gamma+\gamma \mathbb{E}[R'_t]\leq {3+\gamma\over 2}+\gamma \mathbb{E}[R'_t].
\end{eqnarray*}
The Gronwall inequality implies
\begin{eqnarray*}
\sup\limits_{t\in [0, T]}\sup\limits_{N}\mathbb{E}[R_t]\leq C(\gamma, \mathbb{E}[R_0]) e^{\gamma T}<\infty.
\end{eqnarray*}
That is
\begin{eqnarray*}
\sup\limits_{t\in [0, T]}\sup\limits_{N}\int_{\mathbb{R}}x^2d\mathbb{E}[L_N](x)\leq C(\gamma, \mathbb{E}[R_0]) e^{\gamma T}<\infty.
\end{eqnarray*}
By H\"older inequality, for all $p\in [1, 2)$,
\begin{eqnarray*}
\int_{|x|\geq A}x^pd\mathbb{E}[L_N(t)](x)\leq \left(\int_{\mathbb{R}}x^2d\mathbb{E}[L_N(t)](x)\right)^{p/2}\left(\mathbb{E}[L_N(t)](|X|\geq A)\right)^{(2-p)/2}.
\end{eqnarray*}
By the tightness of $\mathbb{E}[L_N(t)]$, we have
\begin{eqnarray*}
\lim\limits_{A\rightarrow \infty} \sup\limits_{t\in [0, T]}\sup\limits_{N}\int_{|x|\geq A}x^pd\mathbb{E}[L_N(t)](x)=0.
\end{eqnarray*}
By the characterization of the $W_p$-convergence on
$\mathscr{P}(\mathbb{R})$, see \cite{Vi1, Vi2}, for all $p\in [1, 2)$, we have
\begin{eqnarray*}
\lim\limits_{N\rightarrow \infty}\sup\limits_{0\leq t\leq T}W_p(\mathbb{E}[L_N(t)], \mu_t)=0.
\end{eqnarray*}
When $V(x)={Kx^2\over 2}$, we have
\begin{eqnarray}
\label{mut}
\frac{d}{dt}\int_{\mathbb{R}}x^2\mu_t(dx)=1-K\int_{\mathbb{R}}
x^2\mu_t(dx),\label{DBM18a}
\end{eqnarray}
\begin{eqnarray}\label{LN8b}
d\langle L_N(t), x^2\rangle&=&\frac{2}{N}\sqrt{\frac{2}{\beta
N}}\sum\limits_{i=1}^N\lambda_N^i(t)dW_t^i-K \left\langle L_N(t),
x^2\right\rangle dt+\left(\frac{2}{\beta}-1\right)\frac{1}{N}+1,
\end{eqnarray}
and
\begin{eqnarray}\label{LN8a}
{d\over dt}\int_{\mathbb{R}} x^2\mathbb{E}[L_N(t, dx)]=1+
\left(\frac{2}{\beta}-1\right)\frac{1}{N}-K\int_{\mathbb{R}}x^2\mathbb{E}[L_N(t,
dx)].
\end{eqnarray}
Hence
\begin{eqnarray*}
\int_{\mathbb{R}} x^2\mathbb{E}[L_N(t, dx)]-\int_{\mathbb{R}} x^2\mu_t(dx)=e^{-Kt}\left[\int_{\mathbb{R}} x^2\mathbb{E}[L_N(0, dx)]-\int_{\mathbb{R}} x^2\mu_0(dx)\right]+{1\over N}\left(\frac{2}{\beta}-1\right){1-e^{-Kt}\over K}.
\end{eqnarray*}
The proof of Theorem \ref{Th3} is completed. \hfill $\square$
\subsection{\bf Proof of Theorem \ref{THPC}}
By the conditions in
Theorem~\ref{Th3} and the Theorem
of~Sznitman~and~Tanaka's~\cite{ST}, we know that,
$M_N(0)$~is~$\mu_0$-chaotic. Since $L_N(t)$~weakly converges to the
deterministic measure~$\mu_t$ for every $t\in [0,T]$, and the
systems~${\rm (GDBM)_V}$ are exchangeable systems, then we have this
propagation of chaos by ~Sznitman~and~Tanaka's Theorem~\cite{ST}.
\hfill $\square$
\medskip
\section{Proof of Theorem \ref{Th5}}
{\bf Proof of Theorem \ref{Th5} (i)}.
By Corollary 3.2 in Biane \cite{Bian03}, for any $C^2$-convex $V$,
there exists a unique equilibrium measure $\mu$ (indeed $\mu=\mu_V$)
with a density $\rho$ satisfying the Euler-Lagrange equation
${\rm H}\rho(x)={1\over 2}V'(x)$ for all $x\in {\rm supp}(\mu)$. Thus,
$\Sigma_V$ has a unique minimizer $\mu_V$. Moreover,
as $V$ is $C^2$-convex, Theorem \ref{th2} implies that
$\Sigma_V$ is a geodesically convex on $\mathscr{P}_2(\mathbb{R})$.
By the fact that $\Sigma_V$ is lower semi-continuous and with
respect to the weak convergence topology, see e.g. \cite{AGZ, Gui},
we see that it is also lower semi-continuous with respect to the
Wasserstein topology on $\mathscr{P}(\mathbb{R})$. Moreover, for all
$c\in \mathbb{R}$ the level set $\{\mu: \Sigma(\mu)\leq c\}$ of
$\Sigma_V$ is relatively compact in the weak convergence topology on
$\mathscr{P}(\mathbb{R})$. By the characterization of the
convergence in the Wasserstein space $\mathscr{P}_2(\mathbb{R})$, we
see that for all $c\in \mathbb{R}$ and $R>0$, $\{\mu:
\Sigma(\mu)\leq c\}\cap B(\mu_0, R)$ is relatively compact with
respect to the topology induced by the Wasserstein distance on
$\mathscr{P}_2(\mathbb{R})$, where $B(\mu_0, R)=\{\mu\in
\mathscr{P}_2(\mathbb{R}): W_2(\mu_0, \mu)\leq R\}$. Hence
$\Sigma_V$ is proper on any geodesic balls of
$\mathscr{P}_2(\mathbb{R})$.
By Proposition $4.1$ in Kloekner \cite{Kloe}, we know that
$\mathscr{P}_2(\mathbb{R})$ has vanishing sectional curvature in the
sense of Alexandrov. More precisely, for any $\mu_1, \mu_2, \mu_3\in
\mathscr{P}_2(\mathbb{R})$ and for any Wasserstein geodesic $\gamma:
[0, 1]\rightarrow \mathscr{P}_2(\mathbb{R})$ such that
$\gamma(0)=\mu_1$ and $\gamma(1)=\mu_2$, for all $t\in [0, 1]$, it
holds that
\begin{eqnarray*}
W_2^2(\mu_3, \gamma(t))=tW_2^2(\mu_3, \mu_1)+(1-t)W_2^2(\mu_3, \mu_2)-t(1-t)W_2^2(\mu_1, \mu_2).
\end{eqnarray*}
Therefore, $\mathscr{P}_2(\mathbb{R})$ is a nonpositively curved
(NPC) space in the sense of Alexandrov (even though
$\mathscr{P}_2(\mathbb{R}^n)$ is an Alexander space with nonnegative
curvature for $n\geq 2$, see e.g. \cite{AGS}).
By Mayer \cite{Mey} and \cite{Kuw}, we can conclude that $W_2(\mu_t,
\mu_V)\rightarrow 0$ holds if we only assume that $V$ is a
$C^2$-convex potential. The proof of Theorem \ref{Th5} (i) is
completed.\\
\medskip
\noindent{\bf Proof of Theorem \ref{Th5} (ii)}. Taking
$\mu_1(t)=\mu_t$ and $\mu_2(t)\equiv \mu_V$ in Theorem \ref{Th2}, we
have
\begin{eqnarray*}
W_{2}^{2}(\mu_t, \mu_V) \leq e^{-2Kt}W_{2}^{2}(\mu(0), \mu_V).
\end{eqnarray*}
By the fact that $\mu_t$ is the gradient flow of $\Sigma_V$ on
$\mathscr{P}_2(\mathbb{R})$ and using the uniform $K$-convexity of
$\Sigma_V$, we can use the same argument as in \cite{Ot} to prove
\begin{eqnarray*}
\Sigma_V(\mu_t|\mu_V)\leq e^{-2Kt}\Sigma_V(\mu_0|\mu_V).
\end{eqnarray*}
Indeed, by Otto's calculus, we have
\begin{eqnarray*}
\frac{d}{dt}\|{\rm
grad}_W\Sigma_V(\mu_t)\|^2_{\mathscr{P}_2(\mathbb{R})} &=&
2\left\langle {\rm grad}_W\|{\rm
grad}_W\Sigma_V(\mu_t)\|^2_{\mathscr{P}_2(\mathbb{R})}, \frac{d\mu_t}{dt}\right\rangle\\
&=& -2{\rm Hess}_{\mathscr{P}_2(\mathbb{R})} \Sigma_V(\mu_t)\left(\frac{d\mu_t}{dt}, \frac{d\mu_t}{dt}\right)\\
&\leq& -2K\left\|\frac{d\mu_t}{dt}\right\|^2_{\mathscr{P}_2(\mathbb{R})}\\
&=& -2K\|{\rm grad}_W
\Sigma_V(\mu_t)\|^2_{\mathscr{P}_2(\mathbb{R})}.
\end{eqnarray*}
Note that ${\rm grad}_W \Sigma_V(\mu_V)=0$. Thus
\begin{eqnarray*}
\frac{d}{dt}\Sigma_V(\mu_t|\mu_V) &=& \left\langle {\rm
grad}_W \Sigma_V(\mu_t), \frac{d\mu_t}{dt}\right\rangle\\
&=& -\|{\rm grad}_W \Sigma_V(\mu_t)\|^2_{\mathscr{P}_2(\mathbb{R})}\\
&=& \int^{\infty}_{t}\frac{d}{ds}\|{\rm grad}_W \Sigma_V(\mu_s)\|^2_{\mathscr{P}_2(\mathbb{R})}ds\\
&\leq& -2K\int^{\infty}_{t}\|{\rm grad}_W \Sigma_V(\mu_s)\|^2_{\mathscr{P}_2(\mathbb{R})}ds\\
&=& 2K\int^{\infty}_{t}\frac{d}{ds}\Sigma_V(\mu_s)ds\\
&=& -2K\Sigma_V(\mu_t|\mu_V),
\end{eqnarray*}
where in the last step we have used the fact
$\Sigma_V(\mu(\infty))=\Sigma_V(\mu_V)=0$. The Gronwall inequality
implies
\begin{eqnarray*}
\Sigma_V(\mu_t|\mu_V)\leq e^{-2Kt}\Sigma_V(\mu_0|\mu_V).
\end{eqnarray*}
The proof of Theorem \ref{Th5} (ii) is completed. \hfill $\square$
\medskip
To prove Theorem \ref{Th5} (iii), we need the following free
logarithmic Sobolev inequality and free Talagrand transportation
cost inequality due to Ledoux and Popescu \cite{Le-Po09}.
\begin{theorem}\label{LP09} (Ledoux-Popescu \cite{Le-Po09}) Suppose that $V$ is a $C^2$, convex and there exists a constant $r>0$ such that
$$V''(x)\geq K>0, \ \ \ \ |x|\geq r.$$
Then there exists a constant $c=C(K, r)>0$ such that the free
Log-Sobolev inequality holds: for all probability measure $\mu$ with
$I_V(\mu)<\infty$,
\begin{eqnarray*}
\Sigma_V(\mu|\mu_V)\leq {2\over c}{\rm I}_V(\mu).
\end{eqnarray*}
Moreover, the free Talagrand transportation inequality holds: there
exists a constant $C=C(K, r, V)>0$ such that
\begin{eqnarray*}
CW_2^2(\mu, \mu_V)\leq \Sigma_V(\mu|\mu_V).
\end{eqnarray*}
\end{theorem}
{\bf Proof of Theorem \ref{Th5} (iii)}. By Biane and Speicher
\cite{BS01}, we have the following entropy dissipation formula
\begin{eqnarray*}
{\partial \over \partial t}\Sigma_V(\mu_t|\mu_V)=-{1\over 2 }{\rm I}_V(\mu_t).\label{entro-diss-2}
\end{eqnarray*}
By Theorem \ref{LP09}, there exists a constant $C_1>0$ such that the {free LSI}
holds
\begin{eqnarray*}
\Sigma_V(\mu|\mu_V)\leq {2\over C_1}{\rm I}_V(\mu),
\end{eqnarray*}
which yields
\begin{eqnarray*}
{d\over dt} \Sigma_V(\mu_t|\mu_V)\leq -{C_1\over 4}
\Sigma_V(\mu_t|\mu_V).
\end{eqnarray*}
By the Gronwall inequality, we have
\begin{eqnarray*}
\Sigma_V(\mu_t|\mu_V)\leq e^{-C_1t/4}\Sigma_V(\mu_0|\mu_V).
\end{eqnarray*}
By Theorem \ref{LP09} again, there exists a constant $C_2>0$ such that the free
transportation cost inequality holds
\begin{eqnarray*}
W_2^2(\mu_t, \mu_V)\leq {1\over C_2}\Sigma_V(\mu_t|\mu_V).
\end{eqnarray*}
Therefore
\begin{eqnarray*}
W_2^2(\mu_t, \mu_V)\leq {e^{-C_1t/4}\over C_2}\Sigma_V(\mu_0|\mu_V).
\end{eqnarray*}
This finishes the proof of Theorem \ref{Th5} (iii). \hfill $\square$
\medskip
\begin{remark} By the same argument as used in Otto \cite{Ot} and Otto-Villani
\cite{OV}, we can prove the following HWI inequality: Suppose that
there exists a constant $K\in \mathbb{R}$ such that
$$V''(x)\geq K, \ \ \ \forall x\in \mathbb{R}.$$
Let $\mu_i\in \mathscr{P}_2(\mathbb{R})$, $i=1, 2$. Then for all
$t>0$, the HWI inequality holds
\begin{eqnarray}
\Sigma_V(\mu_1)-\Sigma_V(\mu_2)\leq W_2(\mu_1, \mu_2)\|{\rm
grad}_W\Sigma_V(\mu_1)\|_{\mathscr{P}_2(\mathbb{R})}-{K\over
2}W_2^2(\mu_1, \mu_2).\label{HWI-1}
\end{eqnarray}
In particular, for any solution to the McKean-Vlasov equation
$(\ref{DBM7})$, we have
\begin{eqnarray}
\Sigma_V(\mu_t)\leq W_2(\mu_t, \mu_V)\|{\rm
grad}_W\Sigma_V(\mu_t)\|_{\mathscr{P}_2(\mathbb{R})}-{K\over
2}W_2^2(\mu_t, \mu_V).\label{HWI-2}
\end{eqnarray}
where
\begin{eqnarray*}
\|{\rm
grad}_W\Sigma_V(\rho)\|^2_{\mathscr{P}_2(\mathbb{R})}=\int_{\mathbb{R}}\rho
|V'(x)-2{\rm H}\rho(x)|^2dx.
\end{eqnarray*}
\end{remark}
To save the length of the paper, we leave the proof to the reader.
\medskip
\section{Double-well potentials and some conjectures}
In this section we discuss again the problem of the longtime
convergence of the McKean-Vlasov equation towards to the equilibrium
measure. More precisely, we want to study the question under which
condition on the external potential $V$ the following double limits
are exchangeable. That is,
\begin{eqnarray*}
\lim\limits_{N\rightarrow \infty}\lim\limits_{t\rightarrow \infty}L_N(t)=
\lim\limits_{t\rightarrow \infty}\lim\limits_{N\rightarrow \infty}L_N(t).
\end{eqnarray*} By \cite{Ch, RS93}, see also \cite{AGZ, Gui}, this is the case when $V(x)={x^2\over 2}$.
Theorem \ref{Th5} ensures the longtime convergence of the weak
solution of the McKean-Vlasov equation to the equilibrium measure
$\mu_V$ for $C^2$-convex potentials $V$. In particular, Theorem
\ref{Th5} applies to {$V(x)=a|x|^{p}$} with $a>0$ and $p\geq 2$.
When $V(x)={x^2\over 2}$ and $\beta=1, 2, 4$, this corresponds to
the cases of GUE, GOE and GSE. Moreover, Theorem \ref{Th5} also
applies to the Kontsevich-Penner model on the Hermitian random
matrices ensemble with external potential (cf. \cite{CM})
$$
V(x)={ax^4\over 12}-{bx^2\over 2}-c\log|x|.
$$
provided that $a>0, c>0$ and $4ac\geq b^2$.
Can we establish the longtime convergence of the McKean-Vlasov
equation in the non-convex case of external potential? In
\cite{BS01, Bian03}, Biane and Speicher gave a non-convex potential
$V$ to which the longtime convergence of $\mu_t$ fails. Indeed, as
$\mu_t$ satisfies the gradient flow of the Voiculescu free entropy
$\Sigma_V$ on $\mathscr{P}(\mathbb{R})$, $\mu_t$ may converge to a
local minimizer of $\Sigma_V$ which is not necessary the global
minimizer $\mu_V$. In statistical physics, this indicates that there
might be a phase transition for the large $N$-GDBM model with
non-convex potentials.
Let us consider the double-well potential
\begin{eqnarray*}
V(x)={1\over 4}x^4+{c\over 2}x^2, \ \ \ \ x\in \mathbb{R},\label{DW}
\end{eqnarray*}
where $c\in \mathbb{R}$ is a constant. By
\cite{Joh98, BI}, it has been known that the density function of the
equilibrium measure $\mu_V$ can be explicitly given as follows: \\
$(i)$ When $c<-2$, $\rho(x)={1\over
2\pi}|x|\sqrt{(x^2-a^2)(b^2-x^2)}1_{[a, b]}$,
where
$a^2=-2-c$ and $b^2=2-c$. \\
$(ii)$ When $c=-2$, $\rho(x)={1\over 2\pi}x^2\sqrt{4-x^2}1_{[-2, 2]}$. \\
$(iii)$ When $c>-2$, $\rho(x)={1\over
\pi}(b_2x^2+b_0)\sqrt{a^2-x^2}1_{[-a, a]}$ ,
where $a^2={\sqrt{4c^2+48}-2c\over 3}$, $b_0={c+\sqrt{{c^2\over
4}+3}\over 3}$, and $b_2={1\over 2}$.\\
When $c\in [0, \infty)$, $V$ is $C^2$ convex and $V''(x)\geq 3$ for $|x|\geq 1$. In this case, Theorem \ref{Th5} (ii) implies that $W_2(\mu_t, \mu_V)\rightarrow 0$ with an exponential convergence rate.
When $c\in(-\infty, -2)$, $\mu_V$ has two supports $[-b, -a]$ and
$[a, b]$ which are disjoint. By Section $7.1$ in Biane-Speicher
\cite{BS01}, it is known that $\mu_t$ does not converge to $\mu_V$.
See also Biane \cite{Bian03}. This also indicates that one cannot
simultaneously prove a free version of the Holley-Stroock
logarithmic Sobolev inequality and a free version of the Talagrand
$T_2$-transportation cost inequality under bounded perturbations of
$p_N(dx)=Z_N^{-1}\prod_{i<j}|x_i-x_j|^2\prod_{i=1}^Ne^{-NV(x_i)}dx$.
Otherwise, by analogue of the proof of Theorem \ref{Th5} (ii), we
may prove that $\mu_t$ converges to $\mu_V$ with respect the
$W_2$-Wasserstein distance and hence in the weak convergence
topology on $\mathscr{P}(\mathbb{R})$. See also \cite{Le-Po09, MMS}
for a discussion on non-convex potentials.
In the case $c\in [-2, 0)$, as the global minimizer $\mu_V$ of
$\Sigma_V$ has a unique support, and all stationary point of $\mu_V$
must satisfy the Euler-Lagrange equation ${\rm H}\mu={1\over 2}V'$,
one can see that the Voiculescu free entropy $\Sigma_V$ has a unique
minimizer which is $\mu_V$. As $\mu_t$ is the gradient flow of
$\Sigma_V$ on $\mathscr{P}_2(\mathbb{R})$, and since ${d\over
dt}\Sigma_V(\mu_t)=-2\int_{\mathbb{R}}\left[V'(x)-2{\rm
H}\rho_t(x)\right]^2\rho_t(x)dx$, we see that $\Sigma_V(\mu_t)$ is
strictly decreasing in time $t$ unless $\mu_t$ achieves the
minimizer $\mu_V$. This yields that the limit of $\Sigma_V(\mu_t)$
exists as $t\rightarrow \infty$. If $\{\mu_t\}$ is tight, and
$\lim\limits_{t\rightarrow \infty}\Sigma_V(\mu_t)=\Sigma_V(\mu_V)$,
we can derive that $\mu_t$ weakly converges to $\mu_V$. By lack of
the tightness of $\{\mu_t\}$, the question whether $W_2(\mu_t,
\mu_V)\rightarrow 0$ (or even $\mu_t$ weakly converges to $\mu_V$)
as $t\rightarrow \infty$ for the above double-well potential $V$
remains open.
We would like to raise the following conjectures.
\begin{conjecture} \label{conj1} Consider the double-well potential $V(x)={1\over 4}x^4+{c\over 2}x^2$ with $c\in [-2, 0)$.
Then $\mu_t$ converges to $\mu_V$ with respect the $W_2$-Wasserstein distance and hence in the weak convergence topology
on $\mathscr{P}(\mathbb{R})$.
\end{conjecture}
\begin{conjecture}\label{conj2}
Suppose that the potential $V$ is a $C^2$ potential function with $V''(x)\geq K_1$ for all $|x|\geq r$
and $V''(x)\geq -K_2$ for all $|x|\leq r$, where $K_1, K_2, r>0$ are some positive constants.
Suppose further that $\Sigma_V$ has a unique minimizer which has a single compact support. Then $\mu_t$ converges to $\mu_V$ with respect the $W_2$-Wasserstein distance and in the weak convergence topology on $\mathscr{P}(\mathbb{R})$.
\end{conjecture}
Finally, let us mention the following conjecture due to Biane and
Speicher \cite{BS01}.
\begin{conjecture}\label{conj3} Consider the double-well potential given by $V(x)={1\over 2}x^2+{g\over 4}x^4$, where $g$ is a negative constant but very close to zero. Then $\mu_t$ weakly converges to $\mu_V$.
\end{conjecture}
\medskip
|
\section{Introduction}
We shall consider complex projective space $\bC\bP^n$ equipped with the standard Fubini-Study metric, and we shall denote by $\omega_n$ its associated $(1,1)$-form. In affine coordinates $z=(z_1,\ldots, z_n)$ in an affine chart $\cong \bC^n\subset \bC\bP^n$, we have
\begin{equation}\label{e:FB}
\omega_n=\frac{\sqrt{-1}}{2\pi}\partial \bar\partial \log \left(1+\sum_{i=1}^n|z_i|^2\right)=\frac{\sqrt{-1}}{2\pi}\partial \bar\partial \log \left(1+\norm z\norm^2\right).
\end{equation}
Let $\Omega\subset \bC\bP^n$ be an open subset and $F\colon \Omega\to \bC\bP^d$ a holomorphic mapping. For a nonnegative real number $\lambda$, the $\lambda$-modification of the Fubini-Study metric (in $\Omega$) induced by this mapping is the metric whose associated $(1,1)$-form is given by $\omega_{n,F,\lambda}=\omega_n+\lambda F^*\omega_p$. The considerations in this paper are local, so we shall assume that $\Omega$ and $F(\Omega)$ are contained in affine charts of $\bC\bP^n$ and $\bC\bP^d$, respectively; thus, if we express $F$ in affine coordinates, $F(z)=[1:f(z)]$ with $f(z)=(f_1(z),\ldots, f_n(z))$, then $\omega_{n,f,\lambda}=\omega_{n,F,\lambda}$ is given by
\begin{equation}\label{e:FBmod}
\omega_{n,f,\lambda}:=\omega_n+\lambda\frac{\sqrt{-1}}{2\pi}\partial \bar\partial \log \left(1+\sum_{i=1}^p|f_i(z)|^2\right)=\omega_n+\lambda\frac{\sqrt{-1}}{2\pi}\partial \bar\partial \log \left(1+\norm f\norm^2\right).
\end{equation}
We shall further assume that there is a positive integer $m$, a positive real number $\mu$, and a holomorphic mapping $h\colon \Omega\subset \bC^n\to \bC^m\subset \bC\bP^m$ such that $\omega_{n,f,\lambda}=\mu h^*\omega_m$; i.e.,
\begin{equation}\label{e:modeq}
\omega_n=\mu h^*\omega_m-\lambda f^*\omega_d.
\end{equation}
This situation, but in a more general setting where the source is a general simply connected K\"ahler manifold and the target is a product of projective spaces, was considered in the recent paper \cite{HuangY13} by X. Huang, and Y. Yuan. They show that strong rigidity properties hold under suitable number theoretic conditions on the conformal factors $\mu$ and $\lambda$. In the restrictive setting considered here, their result would state that if there are no positive rational numbers $s,t$ such that $s\lambda=t\mu$, then $f$ and $h$ extend as global holomorphic immersions $f\colon \bC\bP^n\to \bC\bP^d$, $h\colon \bC\bP^n\to \bC\bP^m$, and furthermore, $f$ and $h$ are both conformal isometries (i.e., $f^*\omega_d=a\omega_n$, $h^*\omega_m=b\omega_n$) with integral conformal factors $a,b$ such that $1=a\mu-b\lambda$. The reader is referred to \cite{HuangY13} for a discussion of the relevance and general context of this problem.
In this paper, we shall consider in some sense the opposite case, where the conformal factors $\lambda$ and $\mu$ are rational numbers, in which case the number theoretic condition in \cite{HuangY13} of course fails. In this case, the rigidity properties established in \cite{HuangY13} also fail, as is pointed out in that paper: The mappings $f$ and $h$ do not extend as global mappings and they are not conformal isometries, in general. However, there are range estimates that hold for the rank of $h$, in general depending on the dimension $d$ of the modification as well as the conformal factors $\mu$ and $\lambda$. In the special case where the conformal factor $\lambda$ equals one, there are "gaps" in the range of possible ranks of $h$ such that the integers in these gaps cannot occur as the rank of an $h$ satisfying \eqref{e:modeq} for any $f$ or integral $\mu$. (This phenomenon is akin to the codimensional gaps that are predicted by the Huang-Ji-Yin Gap Conjecture \cite{HuangJiYin09} for CR mappings between spheres. Indeed, the underlying reasons are similar, in both cases boiling down to rank properties of certain sums of squares; see Section \ref{s:SOS}.)
We shall say that a mapping $G\colon \Omega\to \bC\bP^N$ is {\it minimally embedded} if the image $G(\Omega)$ is not contained in a proper projective plane. Since projective space equipped with the Fubini-Study metric is a homogeneous space, there is no loss of generality in assuming that $0\in \Omega$ and that $f(0)=h(0)=0$. In this case, $g$ being minimally embedded is equivalent to the components of $g=(g_1,\ldots, g_N)$, in affine coordinates near $0$, being linearly independent. We first state our result in the special case where the conformal factors $\lambda$ and $\mu$ are both one:
\begin{theorem}\label{thm:main0} Let $\Omega\subset \bC^n\subset\bC\bP^n$ be a connected open set, $f\colon\Omega\to \bC^d\subset \bC\bP^d$ a minimally embedded holomorphic mapping, and $\omega_{n,f}=\omega_n+f^*\omega_d$ the $1$-modification of the Fubini-Study metric $\omega_n$ induced by $f$. Then, there is a minimally embedded holomorphic mapping $h\colon \Omega\to \bC^m\subset \bC\bP^m$, unique up to multiplication by a unitary $m\times m$ matrix, such that $\omega_{n,f}=h^*\omega_m$ or, equivalently,
\begin{equation}\label{e:hmod0}
\omega_{n}=h^*\omega_m-f^*\omega_d,
\end{equation}
and the dimension $m$ satisfies the following:
\smallskip
{\rm (i)} If $d\leq n$, then
\begin{equation}\label{e:main0bound}
n+\sum_{l=0}^{d-1} (n-l)=n(d+1)-\frac{d(d-1)}{2}\leq m\leq dn+n+d=n(d+1)+d.
\end{equation}
{\rm (ii)} If $d\geq n$, then $m\geq \max(n(n+3)/2,d)$.
\end{theorem}
\begin{remark}{\rm Observe that, for fixed $n$, the function $d\mapsto n(d+1)-\frac{d(d-1)}{2}$ is strictly increasing in $d$ for $1\leq d\leq n$.
}
\end{remark}
The existence of the mapping $h$ is trivial in this case (see the proof of Theorem \ref{thm:main0} below), and the uniqueness is a consequence of a well known lemma by D'Angelo \cite{D'Angelobook}. The main point of the theorem is the range estimates in (i) on the dimension $m$ for low dimensional ($d\leq n$) modifications. We note that there are "gaps" in the range of possible dimensions $m$ that can occur as target dimensions for $h$, {\it regardless} of the modification (i.e., regardless of the modifying mapping $f$ and dimension $d$). For instance, if $d=1$, then $2n\leq m\leq 2n+1$. If $d=2$, then $3n-1\leq m\leq 3n+2$, etc. As $d$ grows towards $n$, these possible ranges of $m$ will initially be disjoint, but the "gaps" between them will shrink until eventually (when $d\sim \sqrt{2n}$) they disappear. The gap intervals of dimensions $m$ for which {\it no} minimally embedded $h\colon \Omega\to \bC^m\subset \bC\bP^m$ is isometric to a $1$-modification induced by any $f$ go as follows (until they disappear):
\begin{equation}\label{e:1gaps}
(0,2n),\ (2n+1,3n-1),\ (3n+2,4n-3),\ \ldots,
\end{equation}
with the $d$th gap being given by $(n(d+1)+d,n(d+2)-d(d+1)/2)$, which as the reader can readily verify becomes empty when $d$ is sufficiently large, $d\sim \sqrt{2n}$ as mentioned above. The estimate provided in (ii) is of less interest. It will become very poor as the dimension $d$ of the modification grows; for generic choices of $f$ the dimension $m$ will grow on the order of the right hand side of \eqref{e:main0bound}. We formulate the gap result as a corollary:
\begin{corollary}\label{cor:maingap} Let $\Omega\subset \bC^n\subset\bC\bP^n$ be a connected open set, $h\colon\Omega\to \bC^m\subset \bC\bP^m$ a minimally embedded holomorphic mapping such that
\begin{equation}\label{e:gapd}
n(k+1)+k<m<n(k+2)-\frac{k(k+1)}{2},
\end{equation}
for some $k$. Then, $h^*\omega_m$ is not the $1$-modification $w_{n,f}=w_n+f^*\omega_d$ for any $f\colon\Omega\to \bC^d\subset \bC\bP^d$.
\end{corollary}
In order to formulate a result with general, rational conformal factors $\lambda$ and $\mu$, we need to introduce some terminology and notation. Let $\phi=(\phi_1,\ldots,\phi_a)$ and $\psi=(\psi_1,\ldots, \psi_b)$ be holomorphic mappings $\Omega\to \bC^a\subset \bC\bP^{a}$ and $\Omega\to \bC^b\subset \bC\bP^{b}$, respectively. The {\it tensor product} $\phi\otimes \psi$ is defined to be the mapping $\Omega\to \bC^{ab}\subset \bC\bP^{ab}$ whose components are $\phi_i\psi_j$ as $i$ and $j$ run over the sets $\{1,\ldots,a\}$ and $\{1,\ldots,b\}$, respectively, in some predetermined ordering of pairs $(i,j)$. The notation $\phi^{\otimes k}$ denotes the tensor product of $\phi$ with itself $k$ times. The {\it rank} of a holomorphic mapping $\phi\colon \Omega\to \bC^a\subset \bC\bP^{a}$ is the smallest integer $r$ such that the image $\phi(\Omega)$ is contained in an affine complex plane (or, equivalently, projective plane if considered as a mapping into $\bC\bP^{a}$) of dimension $r$; in particular, $\phi$ is minimally embedded in $\bC\bP^a$ if and only if the rank is $a$.
\begin{theorem}\label{thm:main1} Let $\Omega\subset \bC^n\subset\bC\bP^n$ be a connected open set, $f\colon\Omega\to \bC^d\subset \bC\bP^d$ a minimally embedded holomorphic mapping, and $a,b,c$ positive integers without a common prime factor. Let $\omega_{n,f, c/b}=\omega_n+(c/b) f^*\omega_d$ be the $c/b$-modification of the Fubini-Study metric $\omega_n$ induced by $f$. Assume that there is a minimally embedded holomorphic mapping $h\colon \Omega\to \bC^m\subset \bC\bP^m$ such that $\omega_{n,f,c/b}=(a/b)h^*\omega_m$ or, equivalently,
\begin{equation}\label{e:hmodrat}
\omega_{n}=\frac{a}{b}h^*\omega_m-\frac{c}{b}f^*\omega_d.
\end{equation}
If $(1,f)^{\otimes c}$ has rank $e+1$, then
\begin{equation}\label{e:ftensorrk}
cd\leq e\leq \sum_{k=1}^c\binom{d+k-1}{k}
\end{equation}
and the following hold:
\smallskip
{\rm (i)} If $e\leq n$ and $b=1$, then the following two inequalities hold:
\begin{equation}\label{e:mest1}
\left\{
\begin{aligned}
n(e+1)-\frac{e(e-1)}{2}\leq& \sum_{k=1}^a\binom{m+k-1}{k}\\
am\leq& n(e+1)+e.
\end{aligned}
\right.
\end{equation}
{\rm (ii)} If $e\geq n$ or $b\geq 2$, then
\begin{equation}\label{e:mest2}
\frac{n(n+3)}{2}\leq \sum_{k=1}^a\binom{m+k-1}{k}.
\end{equation}
\end{theorem}
We note that if $a=1$ in Theorem \ref{thm:main1}, then the estimates on $m$ are of the same type as those of Theorem \ref{thm:main0}, with the rank $e$ playing the role of the dimension $d$ in the latter theorem.
If $a\geq 2$, then the existence of the mapping $h$ has to be assumed, as it may not exist in general. We also note that in this case the sum on the right in the first inequality of \eqref{e:mest1} and on the right in \eqref{e:mest2} is a polynomial of degree $a$ in $m$ with positive coefficients. Consequently, the lower bound on $m$ provided by these two estimates will be roughly on the order of the $a$th root of the left hand sides. In the case (i) in Theorem \ref{thm:main1}, this means (as the reader can verify) that the possible intervals of $m$ provided by \eqref{e:mest1} will in general not be disjoint for different values of $e\geq cd$, as is the case in Theorem \ref{thm:main0}. Thus, the gaps in possible values of $m$, regardless of the modification, that exist when the conformal factors are both one, cannot be predicted (although they may still exist) for general rational conformal factors by Theorem \ref{thm:main1}, {\it except} for the first gap that exists for sufficiently small values of $a$ (compared to the dimension $n$): For fixed $c\geq 1$, we have $e\geq c$ and hence \eqref{e:mest1} and \eqref{e:mest2} imply that
$$
n(c+1)-\frac{c(c-1)}{2}\leq \sum_{k=1}^a\binom{m+k-1}{k}.
$$
(If $c\geq n$, we replace the left hand side by $n(n+3)/2$, but let us assume here that $c\leq n$.)
Observe that if $m=1$, then the right hand side equals $a$. It follows, regardless of the modification, that if $a<n(c+1)-c(c-1)/2$, then $m\geq 2$. Similarly, if $m=2$, then the right hand side equals $(a+1)(a+2)/2 -1$ and, hence, if $(a+1)(a+2)/2 -1<n(c+1)-c(c-1)/2$, then $m\geq 3$, etc. In general, Theorem \ref{thm:main1} should be regarded as estimates on $m$ for a given rank $e$, but estimates that do not depend on the modifying mapping $f$ itself. As above, the main point is the estimates for low ranks $e\leq n$. We should point out, however, that the gap phenomenon described in Corollary \ref{cor:maingap} holds for the possible {\it ranks} of $(1,h)^{\otimes a}$. This follows directly from the proofs of Theorems \ref{thm:main0} and \ref{thm:main1}.
\begin{remark}\label{rem:bad}{\rm If $a\geq 2$, then the lower bounds provided by Theorem \ref{thm:main1} are at best $m\geq n$, and in general worse than this. To see this, observe that in the case $a\geq 2$, if
\begin{equation}\label{e:bestbound}
\frac{n(n+3)}{2}\leq m+\frac{m(m+1)}{2},
\end{equation}
then the inequalities for lower bounds on $m$ in \eqref{e:mest1} and \eqref{e:mest2} hold. Thus, any lower bound $m\geq A$ that follows from Theorem \ref{thm:main1} is implied by the lower bound that follows from \eqref{e:bestbound}, and the reader can readily verify that this bound is precisely $m\geq n$.
}
\end{remark}
If $f$ in Theorem \ref{thm:main1} is assumed to be a rational mapping, then one can show (using Huang's Lemma; see below) that in fact $m\geq n$. Thus, in view of Remark \ref{rem:bad}, for rational mappings, the lower bounds in Theorem \ref{thm:main1} do not yield any new information beyond $m\geq n$. It is also possible, in this case, that estimates that are {\it linear} in $m$ could be proved, if further progress is made on the SOS problem (see Section \ref{s:SOS}) in the general situation. The case where $f$ is rational is discussed in Section \ref{s:polyf}.
Standard arguments will reduce the proofs of Theorems \ref{thm:main0} and \ref{thm:main1} to statements about ranks of certain sums of squares. (A reader unfamiliar with these standard arguments might want to read Section \ref{s:proofs} before reading Section \ref{s:SOS}.) The results concerning the latter that are needed for the proofs are stated and proved in Section \ref{s:SOS}. The proofs of Theorems \ref{thm:main0} and \ref{thm:main1} are then given in Section \ref{s:proofs}. A discussion of the case where the modifying map is rational is conducted in Section \ref{s:polyf}. Some examples are also given in the latter section.
The connection between results concerning sums of squares and isometric embedding problems has also been explored in, e.g., \cite{CatlinJPD97}, \cite{CatlinJPD99}.
\section{Ranks of Sums of Squares}\label{s:SOS}
In this section, we shall consider some Sums of Squares (SOS) problems that arise in the context of holomorphic mappings between projective spaces with (modified) Fubini-Study metrics. The (standard) connection will be made in Section \ref{s:proofs}.
\subsection{Setup and Basics} We shall first consider the following Sums of Squares (SOS) Equation in $z=(z_1,\ldots,z_n)\in \bC^n$:
\begin{equation}\label{e:SOS}
\left(\sum_{j=1}^n|z_j|^2\right)a(z,\bar z)=\sum_{k=1}^m|h_k(z)|^2,
\end{equation}
where $a(z,\bar z)$ is a real-analytic, Hermitian function; $h=(h_1,\ldots,h_m)$ is a local (germ at $0$ of a) holomorphic mapping $(\bC^n,0)\to (\bC^m,0)$ whose components are linearly independent over $\bC$ (or, equivalently, whose image is not contained in a proper subspace of $\bC^m$). For brevity, we shall also use the notation
$$
\norm z\norm^2:=\sum_{j=1}^n|z_j|^2,\quad \norm h\norm^2=\norm h(z)\norm^2:=\sum_{k=1}^m|h_k(z)|^2;
$$
the dimension of the complex space whose Euclidian norm is used will be clear from the context. Thus, equation \eqref{e:SOS} can then be written
\begin{equation}\label{e:SOS'}
\norm z\norm^2 a(z,\bar z)=\norm h\norm^2.
\end{equation}
If the Hermitian function $a(z,\bar z)$ is a polynomial, then it can then be written as a difference of finite squared norms:
\begin{equation}\label{e:DOS}
a(z,\bar z)=\sum_{i=1}^p|f_i(z)|^2-\sum_{j=1}^q|g_j(z)|^2=\norm f\norm^2-\norm g\norm^2,
\end{equation}
where $f=(f_1,\ldots, f_p)$, $g=(g_1,\ldots,g_q)$ are polynomial mappings. If $a$ is real-analytic but not polynomial, then a similar decomposition can be achieved with (in general, infinite dimensional) Hilbert space valued $f$ and $g$. In what follows, we shall assume that $a$ can be decomposed as a difference of finite squared norms (as in \eqref{e:DOS}), which is always the case if $a$ is polynomial. When the components of $f$ and $g$ are linearly independent (as can always be achieved), then the pair $(p,q)$ is called the {\it rank} of $a$. If $q=0$ (meaning that \eqref{e:DOS} can be achieved with $g\equiv 0$), then $a$ is said to be a (finite) SOS and we will simply refer to $p$ (rather than $(p,0)$) as its rank; thus, e.g., $\norm h\norm^2$ above is a finite SOS of rank $m$. A fundamental problem of general interest (and whose solution would have direct implications for the Huang-Ji-Yin Conjecture in CR geometry mentioned in the introduction) can be described as follows:
\medskip
\noindent
{\bf SOS Problem.} {\it Let $a(z,\bar z)$ be a Hermitian real-analytic function in neighborhood of $0$ in $\bC^n$
and assume that $\norm z\norm^2 a(z,\bar z)$ is a finite {\rm SOS}, i.e., there exists a holomorphic mapping $h=(h_1,\ldots, h_m)$
satisfying \eqref{e:SOS'}. Relate the possible values of the rank $m$ of the {\rm SOS} $\norm h\norm^2$ to the rank $(p,q)$ of $a$ and the dimension $n$.}
\medskip
The only general result known, to the best of the author's knowledge, about the SOS Problem is Huang's Lemma \cite{Huang99}, which states that if $a(z,\bar z)$ is not identically zero, then the rank $m\geq n$. In this paper, we shall only consider the SOS problem in the special case where $a(z,\bar z)$ itself is an SOS.
In what follows, we shall also use the following notation: Let $F=(F_1,\ldots,F_a)$ and $G=(G_1,\ldots, G_b)$ be local holomorphic mappings $(\bC^n, 0)\to \bC^a$ and $(\bC^n, 0)\to \bC^b$, respectively. Then, $F\oplus G$ denotes the mapping $(\bC^n,0)\to \bC^{a+b}$ given by $F\oplus G:=(F,G)$, and $F\otimes G$ the mapping $(\bC^n,0)\to \bC^{ab}$ whose components are $F_iG_j$ as $i$ and $j$ run over the sets $\{1,\ldots,a\}$ and $\{1,\ldots,b\}$, respectively, in some predetermined ordering of pairs $(i,j)$. We observe immediately that
\begin{equation}
\norm F\oplus G\norm^2=\norm F\norm^2+\norm G\norm^2;\quad \norm F\otimes G\norm^2=\norm F\norm^2\norm G\norm^2.
\end{equation}
We shall use the notation $V_F\subset \bC\{z\}\cong \mathcal O_n$ for the vector space over $\bC$ spanned by the components of $F$. Clearly, the rank of the SOS $\norm F\norm^2$ equals the dimension of $V_F$.
\subsection{The SOS problem when $a(z,\bar z)$ is an SOS} The following result concerning the case when $a$ is a bihomogeneous Hermitian polynomial SOS was proved in
\cite{GrHa13}
(Proposition 3) using an estimate by Macauley on the growth of the Hilbert function of a homogeneous polynomial ideal:
\begin{proposition}\label{prop:GrHa13}{\rm (\cite{GrHa13})}
Let $A(Z,\bar Z)$ be a bihomogeneous Hermitian polynomial in $Z=(Z_0, Z_1,\ldots,Z_n)$ and $\bar Z$, and assume that $A(Z,\bar Z)$ is an SOS of rank $p$, i.e.,
$$
A(Z,\bar Z)=\sum_{i=1}^p |F_i(Z)|^2,
$$
where $F_1(Z), \ldots, F_p(Z)$ are linearly independent homogeneous polynomials. If $p\leq n+1$, then the rank $R$ of the SOS $\norm Z\norm^2A(Z,\bar Z)$ satisfies
\begin{equation}\label{e:GrHa13}
\sum_{l=0}^{p-1} (n+1-l)=(n+1)p-\frac{p(p-1)}{2}\leq R\leq p(n+1),
\end{equation}
and if $p\geq n+1$, then $R\geq (n+1)(n+2)/2$.
\end{proposition}
\begin{remark}\label{rem:GrHasharp} {\rm We note that both the lower and upper bound in
can be achieved for each $p\leq n$. It is easy to see that the lower bound is achieved with, e.g., $F_i(Z)=Z_i$ for $i=1,\ldots, p$. The upper bound is achieved for "generic" choices of $F_i(Z)$.
}
\end{remark}
A straightforward argument using homogenization of polynomials yields the following:
\begin{theorem}\label{thm:ASOS} Let $a(z,\bar z)$ be a Hermitian real-analytic function near $0$ in $\bC^n$, and assume that $a(z,\bar z)$ is a finite SOS of rank $p$, i.e.,
\begin{equation}\label{e:aSOSdecomp}
a(z,\bar z)=\sum_{i=1}^p |f_i(z)|^2,
\end{equation}
where $f_1(z), \ldots, f_p(z)$ are linearly independent holomorphic functions near $0$. If $p\leq n$, then the rank $r$ of the SOS $\norm z\norm^2a(z,\bar z)$ satisfies
\begin{equation}\label{e:ASOSbound}
\sum_{l=0}^{p-1} (n-l)=np-\frac{p(p-1)}{2}\leq r\leq pn,
\end{equation}
and if $p\geq n$, then $r\geq \max(n(n+1)/2,p)$.
\end{theorem}
\begin{proof} Let us first assume that $a(z,\bar z)$ is a polynomial of bidegree $(d,d)$ and that \eqref{e:aSOSdecomp} can be achieved with linearly independent polynomials $f_i(z)$ of degree at most $d$. Let us introduce homogeneous coordinates $Z=(Z_0,Z_1,\ldots,Z_n)=(Z_0,\tilde Z)$ and define homogeneous polynomials of degree $d$ by
\begin{equation}\label{e:homoF}
F_i(Z):=Z_0^df_i(\tilde Z/Z_0),\quad i=1,\ldots, p.
\end{equation}
Clearly, the $F_i$ are linearly independent since the $f_i$ are.
It follows that the bihomogeneous Hermitian polynomial
$$
A(Z,\bar Z):=|Z_0|^{2d}a\left(\tilde Z/Z_0, \overline{\tilde Z/Z_0}\right)
$$
then satisfies
$$
A(Z,\bar Z)=\sum_{i=1}^p |F_i(Z)|^2,
$$
and has rank $p$. Let us first assume that $p\leq n$. By Proposition \ref{prop:GrHa13}, the rank $R$ of the SOS $\norm Z\norm^2 A(Z,\bar Z)$ satisfies $R\geq (n+1)p-p(p-1)/2$. Since
$$
\norm Z\norm^2 A(Z,\bar Z)=(|Z_0|^2+\norm\tilde Z\norm^2)A(Z,\bar Z)=
\sum_{i=1}^p|Z_0|^2|F_i(Z)|^2+\norm \tilde Z\norm^2 A(Z,\bar Z)
$$
and $\norm \tilde Z\norm^2 A(Z,\bar Z)$ is also an SOS, it is then clear that the rank $r$ of $\norm \tilde Z\norm^2 A(Z,\bar Z)$ must satisfy
\begin{equation}\label{e:rankbound}
r\geq (n+1)p-p(p-1)/2 - p=np-p(p-1)/2,
\end{equation}
which when $p=n$ reduces to
\begin{equation}\label{e:rankbound'}
r\geq (n+1)n-n(n-1)/2 - n= n(n+1)/2.
\end{equation}
In other words, there are linearly independent homogeneous polynomials $H_i(Z)$ of degree $d+1$ such that
\begin{equation}\label{e:prelim}
\norm \tilde Z\norm^2 A(Z,\bar Z)=\sum_{i=1}^r |H_i(Z)|^2,
\end{equation}
where $r$ satisfies the lower bound \eqref{e:rankbound}. Noting that $f_i(z)=F_i(1,z)$ and $a(z,\bar z)=A\big ((1,z),\overline{(1,z)}\big)$, we conclude by substituting $Z=(1,z)$ in \eqref{e:prelim} that
\begin{equation}\label{e:ASOSdecomp}
\norm z\norm^2 a(z,\bar z)=\sum_{i=1}^r |h_i(z)|^2,
\end{equation}
where $h_i(z)=H_i(1,z)$. The $h_i$ are linearly independent since the $H_i$ are, and therefore the rank of $\norm z\norm^2 a(z,\bar z)$ equals $r$, where $r$ satisfies \eqref{e:rankbound}. Clearly, by construction we have $r\leq pn$ (since $pn$ is the total number of terms obtained when the product $\norm z\norm^2 a(z,\bar z)$ is multiplied out), proving \eqref{e:ASOSbound} when $p\leq n$. If $p\geq n$, then the rank $r$ will be greater than or equal to the corresponding rank obtained when $a(z,\bar z)$ is replaced by the sum on the right in \eqref{e:aSOSdecomp} truncated after $n$ terms, i.e., $r\geq n(n+1)/2$. The fact that $r\geq p$ is trivial. This establishes the statement of Theorem \ref{thm:ASOS} in the polynomial case.
Next, let $a(z,\bar z)$ be a Hermitian real-analytic function near $0$ in $\bC^n$ satisfying \eqref{e:aSOSdecomp}, where the $f_i$ are linearly independent holomorphic functions near $0$, and let
\begin{equation}\label{e:ASOSdecomp2}
\norm z\norm^2 a(z,\bar z)=\sum_{i=1}^r |h_i(z)|^2,
\end{equation}
be a SOS decomposition of $\norm z\norm^2 a(z,\bar z)$ with the $h_i(z)$ being linearly independent holomorphic functions near $0$. If we truncate the Taylor series of the $f_i(z)$ at degree $d$ and those of the $h_i(z)$ at degree $d+1$, then we obtain a Hermitian polynomial $a^d(z,\bar z)$ of bidegree $(d,d)$ such that
\begin{equation}\label{e:adSOSdecomp}
a^d(z,\bar z)=\sum_{i=1}^p|f^d(z)|^2,\quad \norm z\norm^2 a^d(z,\bar z)=\sum_{i=1}^p|h^{d+1}(z)|^2
\end{equation}
where the $f^d_i(z)$ and $h^{d+1}(z)$ denote the truncated Taylor polynomials of $f_i(z)$ and $h_i(z)$ at degrees $d$ and $d+1$, respectively. Since the sets of holomorpic functions $f_1,\ldots, f_p$ and $h_1,\ldots, h_r$ both are linearly independent, it is clear that the sets of polynomials $f^d_1,\ldots, f^d_p$ and $h^d_1,\ldots, h^d_r$ both are linearly independent for $d$ sufficiently large. In other words, for $d$ sufficiently large, the Hermitian polynomial $a^d(z,\bar z)$ is an SOS of rank $p$ and $\norm z\norm^2 a^d(z,\bar z)$ is an SOS of rank $r$. The conclusion of Theorem \ref{thm:ASOS} now follows from the corresponding statement in the polynomial case, already established above. This concludes the proof of Theorem \ref{thm:ASOS}.
\end{proof}
We are now ready to state and prove a "nonhomogeneous" version of Theorem \ref{thm:ASOS} that will be used in the proofs of Theorems \ref{thm:main0} and \ref{thm:main1} below.
\begin{theorem}\label{thm:nonhomoSOS} Let $f_1(z),\ldots,f_p(z)$ be linearly independent local holomorphic functions vanishing at $0$ in $\bC^n$ such that
\begin{equation}\label{e:nonhomoSOSdecomp}
(1+\norm z\norm^2)(1+\sum_{i=1}^p |f_i(z)|^2)=1+\sum_{i=1}^r|h_i(z)|^2,
\end{equation}
where $h_1(z), \ldots, h_r(z)$ are linearly independent local holomorphic functions near $0$. If $p\leq n$, then the rank $r$ of the SOS $\norm h\norm^2=\sum_{i=1}^r |h_i(z)|^2$ satisfies
\begin{equation}\label{e:nonhomoASOSbound}
n+\sum_{l=0}^{p-1} (n-l)=n(p+1)-\frac{p(p-1)}{2}\leq r\leq pn+n+p=n(p+1)+p,
\end{equation}
and if $p\geq n$, then $r\geq n(n+3)/2$.
\end{theorem}
\begin{remark}{\rm Both the upper and lower bound in \eqref{e:nonhomoASOSbound} can be achieved by examples similar to those in Remark \ref{rem:GrHasharp}.
}
\end{remark}
\begin{proof} If we write $f=(f_1,\ldots, f_p)$ and $h=(h_1,\ldots, h_r)$, then \eqref{e:nonhomoSOSdecomp} can be written
$$
(1+\norm z\norm^2)(1+\norm f\norm^2)=1+\norm h\norm^2,
$$
which when multiplied out is equivalent to
\begin{equation}\label{e:nonhomo1}
\norm z\norm^2 +\norm f\norm^2+ \norm z\norm^2 \norm f\norm^2=\norm h\norm^2.
\end{equation}
It is clear that the vector spaces $V_z$ and $V_{f\otimes z}$ only intersect at $0$ (since the Taylor series of the $z_lf_i(z)$ have no constant or linear terms by the assumption that $f_i(0)=0$) and therefore the rank of the SOS $\norm z\norm^2 + \norm z\norm^2 \norm f\norm^2$ ($=\dim_\bC V_z\oplus V_{f\otimes Z}$) is $\geq$ the rank of $\norm z\norm^2 \norm f\norm^2$ plus $n$. It then follows immediately from Theorem \ref{thm:ASOS} that if $p\leq n$, then
$$
r\geq np -\frac{p(p-1)}{2} +n =n(p+1)-\frac{p(p-1)}{2},
$$
which is the lower bound in \eqref{e:nonhomoASOSbound}. The lower bound $r\geq n(n+3)/2$ when $p\geq n$ follows from Theorem \ref{thm:ASOS} in the same way. The upper bound in \eqref{e:nonhomoASOSbound} is obtained directly by counting the number of squares on the left in \eqref{e:nonhomo1}. This completes the proof of Theorem \ref{thm:nonhomoSOS}.
\end{proof}
\subsection{Another Sums of Squares Problem} For the proof of Theorem \ref{thm:main1}, we shall also need the following result concerning the rank of powers of $(1+\norm f\norm^2)$:
\begin{proposition}\label{prop:powerSOS} Let $f_1(z),\ldots,f_p(z)$ be linearly independent local holomorphic functions vanishing at $0$ in $\bC^n$. Let $t\in \bZ_+$ and express $(1+\sum_{i=1}^p |f_i(z)|^2)^t$ as follows:
\begin{equation}\label{e:powerSOSdecomp}
(1+\sum_{i=1}^p |f_i(z)|^2)^t=1+\sum_{i=1}^r|h_i(z)|^2,
\end{equation}
where $h_1(z), \ldots, h_r(z)$ are linearly independent local holomorphic functions near $0$. Then the rank $r$ of the SOS $\norm h\norm^2=\sum_{i=1}^r |h_i(z)|^2$ satisfies
\begin{equation}\label{e:powerASOSbound}
tp\leq r\leq \sum_{k=1}^t\binom{p+k-1}{k}.
\end{equation}
\end{proposition}
\begin{remark} {\rm The lower bound in \eqref{e:powerASOSbound} is realized by taking $f_i(z)=z_1^i$, and the upper bound can be realized by choosing the $f_i$ to be "suitably spaced" monomials. For instance, to realize the upper bound if $p\leq n$, we can simply take $f_i(z)=z_i$; if $p>n$, then we can take the first $n$ $f_i$'s of this form, and then take subsequent ones to be monomials with an increment in the degrees so that the degrees of monomials up to order $t$ in $f_1,\ldots, f_k$ is lower than the degree of the monomials $f_{k+1},\ldots,f_p$.
}
\end{remark}
\begin{proof} Observe that the rank $r$ is the dimension of the complex vector space $V_F\subset \bC\{z\}$ spanned by the collection $F$ of all monomials $f^\alpha:=f_1^{\alpha_1}\ldots f_p^{\alpha_p}$ with $1\leq |\alpha|\leq t$. The upper bound in \eqref{e:powerASOSbound} is easily seen to hold. The number on the right in \eqref{e:powerASOSbound} is the number of distinct monomials of degree $\leq t$ in $p$ variables, and the rank $r$ can clearly not exceed this. To prove the lower bound, we proceed as follows. First, a moments reflection will convince the reader that, by iteratively replacing the $k$th generator $f_k(z)$ with a suitable linear combination $f_k(z)-\sum_{i=1}^{k-1}c_i f_i(z)$ if necessary, we may assume that the generators are of the form $f_i(z)=q_i(z)+O(|z|^{s_i+1})$ where the $q_i(z)$ are homogeneous polynomials of degree $s_i$ such that $q_1,\ldots, q_p$ are linearly independent. We shall need the following lemma, in which the notation $\bC[w]_{\leq t}$ is used for the space of polynomials in $w=(w_1,\ldots,w_q)$ of degree $\leq t$.
\begin{lemma}\label{lem:polyinj} For any positive integers $t$ and $n$, there are positive integers $a_1,\ldots a_n$ such that the algebra homomorphism $\bC[z]\to \bC[\zeta]$ induced by the map
$$p(z)\mapsto p(\zeta^{a_1},\ldots, \zeta^{a_n})
$$
is injective when restricted as a linear map
$\bC[z]_{\leq t}\to \bC[\zeta]_{\leq t\max(a_1,\ldots a_n)}$.
\end{lemma}
\begin{proof}[Proof of Lemma $\ref{lem:polyinj}$] It suffices to prove the lemma for $n=2$, since this result can then be repeated iteratively to "collapse" two variables to one until the desired map $\bC[z]_{\leq t}\to \bC[\zeta]_{\leq t\max(a_1,\ldots a_n)}$ is obtained. Thus, assume $n=2$. Choose $a_1\neq a_2$ to be prime numbers $\geq t+1$. Then, the induced homomorphism maps the monomial $z^\alpha:=z_1^{\alpha_1}z_2^{\alpha_2}$ to $\zeta^{a_1\alpha_1+a_2\alpha_2}$. Suppose, in order to reach a contradiction, that the linear map $\bC[z]_{\leq t}\to \bC[\zeta]_{\leq t\max(a_1,a_2)}$ is not injective. Then, there must be $|\alpha|\leq t$ and $|\beta|\leq t$ such that $z^\alpha$ and $z^\beta$ get mapped to the same monomial $\zeta^k$; i.e.,
$a_1\alpha_1+a_2\alpha_2=a_1\beta_1+a_2\beta_2$ or, equivalently, $a_1(\alpha_1-\beta_1)=a_2(\beta_2-\alpha_2)$. This is clearly a contradiction since $a_1\neq a_2$ are prime numbers $\geq t+1$ and $|\alpha_1-\beta_1|$, $|\beta_2-\alpha_2|$ are both $\leq t$. This completes the proof.
\end{proof}
We now return to the proof of Proposition \ref{prop:powerSOS}. By Lemma \ref{lem:polyinj}, we can find positive integers $a_1,\ldots, a_n$ such that the one-variable polynomials $\tilde f_i(\zeta)=f_i(\zeta^{a_1},\ldots, \zeta^{a_n})$, for $1=1,2,\ldots, p$, are linearly independent. The rank $r$ in Proposition \ref{prop:powerSOS} will be $\geq$ the dimension $\tilde r$ of the complex vector space $V_{\tilde F}\subset \bC\{z\}$ spanned by the collection $\tilde F$ of all monomials $\tilde f^\alpha:=\tilde f_1^{\alpha_1}\ldots \tilde f_p^{\alpha_p}$ with $1\leq |\alpha|\leq t$. Again, by iteratively replacing the $k$th generator $\tilde f_k(\zeta)$ with a suitable linear combination $\tilde f_k(\zeta)-\sum_{i=1}^{k-1}c_i \tilde f_i(\zeta)$ and then renumbering if necessary, we may assume that $\tilde f_i(\zeta)=b_i\zeta^{s_i}+O(\zeta^{s_1+1})$ where $b_i\neq 0$ and $1\leq s_1<\ldots<s_p$. As a final reduction, we note that the dimension $\tilde r$ is $\geq$ the dimension $r'$ of the complex vector space $V'\subset \bC\{z\}$ spanned by the collection $M$ of all monomials
$$\zeta^{\alpha_1s_1+\ldots +\alpha_ps_p},\quad 1\leq |\alpha|\leq t.
$$
To finish the proof, we shall show that $r'$ satisfies the lower bound in \eqref{e:powerASOSbound}. We shall prove this by induction on $p$. Clearly, if $p=1$, then $r'=t=tp$. Next, assume that the lower bound $r'\geq tp$ has been proved for $p<p_0$. Then, with $p=p_0$, we obtain at least $t(p-1)$ distinct monomials $\zeta^q$ with $q=\alpha_1s_1+\ldots+\alpha_{p-1}s_{p-1}$ such that $|\alpha|=|(\alpha_1,\ldots,\alpha_{p-1},0)|\leq t$. We must prove that we obtain at least $t$ new distinct monomials $\zeta^q$ with $q=\alpha_1s_1+\ldots+\alpha_{p-1}s_{p-1}+\alpha_ps_p$ where $\alpha_p\geq 1$ and $|\alpha|\leq t$. We claim that every $q=ks_{p-1}+(m-k)s_p$, for $k=0,\ldots, t-1$, is such that it cannot be obtained as $q=\alpha_1s_1+\ldots+\alpha_{p-1}s_{p-1}$ with $|\alpha|=|(\alpha_1,\ldots,\alpha_{p-1},0)|\leq m$. To see this, note that since $1\leq s_1<\ldots<s_p$ we have
$\alpha_1s_1+\ldots+\alpha_{p-1}s_{p-1}\leq ts_{p-1}$. Moreover, we have
$$
ks_{p-1}+(t-k)s_p - ts_{p-1} =(t-k)(s_p-s_{p-1})>0,\quad k=0,\ldots, t-1,
$$
which proves the claim, and shows that $r'\geq (t-1)p+t=tp$ also for $p=p_0$. This completes the proof of the proposition.
\end{proof}
\section{Proof of Theorems \ref{thm:main0} and \ref{thm:main1}}\label{s:proofs}
\begin{proof}[Proof of Theorem $\ref{thm:main0}$] Let $f\colon\Omega\to \bC^d\subset \bC\bP^d$ and $h\colon \Omega\to \bC^m\subset \bC\bP^m$ be minimally embedded holomorphic mappings satisfying \eqref{e:hmod0}. Since projective space equipped with the Fubini-Study metric is a homogeneous space, we may assume that $0\in \Omega$ and $f(0)=0$, $h(0)=0$. By \eqref{e:FB} and \eqref{e:FBmod}, we conclude that $\log (1+\norm z\norm^2)+\log(1+\norm f\norm^2)=\log (1+\norm h\norm^2)+\Phi$, where $\Phi$ is a real-valued polyharmonic function, i.e., $\Phi(z,\bar z)=\phi(z)+\overline{\phi(z)}$ near $0$ for some local holomorphic function $\phi(z)$. By comparing the Taylor series of the log-terms (no constant terms and no pure terms in $z$ or $\bar z$) with that of $\Phi$, we conclude that $\Phi\equiv 0$, and hence
$$
\log (1+\norm z\norm^2)+\log(1+\norm f\norm^2)=\log (1+\norm h\norm^2).
$$
By exponentiating this, we obtain the SOS identity
\begin{equation}\label{e:FBSOS1}
(1+\norm z\norm^2)(1+\norm f\norm^2)=(1+\norm h\norm^2).
\end{equation}
Conversely, if $f$ and $h$ satisfy this SOS identity, then \eqref{e:hmod0} holds. It follows that if $f$ is a minimally embedded holomorphic mapping, then there exists a mapping $h$ satisfying \eqref{e:FBSOS1} (simply carry out the multiplication on the left), and elementary linear algebra shows that we may assume (after replacing the $h_i$ obtained by multiplying out the left hand side of \eqref{e:FBSOS1} by a suitable basis for the vector space spanned by these) that $h$ is also minimally embedded. If there are two minimally embedded $h$ and $\tilde h$ that both satisfy \eqref{e:FBSOS1}, then $\norm h\norm^2=\norm\tilde h\norm^2$ and hence, by a lemma of D'Angelo \cite{D'Angelobook}, it follows that $h=U\tilde h$ for some unitary $m\times m$ matrix $U$. To finish the proof of Theorem \ref{thm:main0}, we recall that $f$ and $h$, with $f(0)=0$ and $h(0)=0$, are minimally embedded precisely when their components are linearly independent. Thus, the estimates in (i) and (ii) follow immediately from Theorem \ref{thm:nonhomoSOS}.
\end{proof}
\begin{proof}[Proof of Theorem $\ref{thm:main1}$] As in the proof of Theorem \ref{thm:main0} above, we may assume that $0\in \Omega$ and $f(0)=0$, $h(0)=0$. The same argument as in that proof also shows that the mappings $f$, $h$ satisfy
\begin{equation}\label{e:FBSOS2}
(1+\norm z\norm^2)^b(1+\norm f\norm^2)^c=(1+\norm h\norm^2)^a.
\end{equation}
As noted in Section \ref{s:SOS}, we have $(1+\norm f\norm^2)^c=\norm (1,f)^{\otimes c}\norm^2$. Thus, if the rank of $(1,f)^{\otimes c}$ is $e+1$, then we can find a minimally embedded holomorphic mapping $g\colon \Omega\to \bC^{e}$, with $g(0)=0$, such that
\begin{equation}
(1+\norm f\norm^2)^c=\norm (1,f)^{\otimes c}\norm^2=(1+\norm g\norm^2).
\end{equation}
The estimate \eqref{e:ftensorrk} for $e$ follows from Proposition \ref{prop:powerSOS}.
We may now rewrite \eqref{e:FBSOS2} as follows:
\begin{equation}\label{e:FBSOS3}
(1+\norm z\norm^2)^b(1+\norm g\norm^2)=(1+\norm h\norm^2)^a.
\end{equation}
The estimates in (i) and (ii) in the case $b=1$ now follow immediately by combining the estimates in Theorem \ref{thm:nonhomoSOS} and Proposition \ref{prop:powerSOS}. If $b\geq 2$, we observe that
$$
(1+\norm z\norm^2)^b(1+\norm g\norm^2)=(1+\norm z\norm^2)^{b-1}\left((1+\norm z\norm^2)(1+\norm g\norm^2)\right)
$$
and $(1+\norm z\norm^2)(1+\norm g\norm^2)$ has rank at least $n$ by Theorem \ref{thm:nonhomoSOS}. The estimate in (ii) now follows by again combining the estimates in Theorem \ref{thm:nonhomoSOS} and Proposition \ref{prop:powerSOS}. This completes the proof of Theorem \ref{thm:main1}.
\end{proof}
\section{The case where the mapping $f$ is rational; Examples}\label{s:polyf}
Let us examine closer the case where $f\colon \Omega\to \bC\bP^d$ is a rational mapping in Theorem \ref{thm:main1}. Clearly, by the uniqueness property of the mapping $h$ (up to linear transformations as explained in the proof of Theorem \ref{thm:main0}), it follows that $h\colon \Omega\to \bC\bP^m$ is also rational. Let us choose homogeneous coordinates $Z=[Z_0:Z_1:\ldots :Z_n]$ in $\bC\bP^n$. The Fubini-Study metric then has the associated $(1,1)$-form
\begin{equation}\label{e:FBhomo}
\omega_n=\frac{\sqrt{-1}}{2\pi}\partial \bar\partial \log \left(\sum_{i=0}^n|Z_i|^2\right)=\frac{\sqrt{-1}}{2\pi}\partial \bar\partial \log \norm Z\norm^2.
\end{equation}
We shall denote the rational mappings $\bC\bP^n\to \bC\bP^d$ and $\bC\bP^n\to \bC\bP^m$ corresponding to $f$ and $h$ by $F$ and $H$, respectively. Thus, in homogeneous coordinates on the target projective spaces, we have $F(Z)=[F_0(Z):\ldots: F_d(Z)]$ and $H(Z)=[H_0(Z):\ldots:H_m(Z)]$, where $F_i(Z)$ and $H_j(Z)$ are homogeneous polynomials, and the assumptions in Theorem \ref{thm:main1} are equivalent to the identity
\begin{equation}\label{e:polyid}
(\norm Z\norm^2)^b(\norm F\norm^2)^c=(\norm H\norm^2)^a.
\end{equation}
Let us proceed under the assumption that $a\geq 2$.
By complexifying \eqref{e:polyid}, i.e.\ replacing $\bar Z$ by an independent complex variable $\chi$ and using the notation $\bar \phi(\chi):=\overline{\phi(\bar\chi)}$, we obtain
\begin{equation}\label{e:polyidcplx0}
\left(\sum_{j=0}^n Z_j\chi_j\right)^b\left(\sum_{i=0}^dF_i(Z)\bar F_i(\chi)\right)^c=
\left(\sum_{k=0}^mH_k(Z)\bar H_k(\chi)\right)^a.
\end{equation}
Since the polynomial $\sum_{j=0}^n Z_j\chi_j$ is irreducible, the identity \eqref{e:polyidcplx0} implies that $\sum_{j=0}^n Z_j\chi_j$ divides $\sum_{k=0}^mH_k(Z)\bar H_k(\chi)$, and we conclude that there exists a (homogeneous) polynomial $R(Z,\chi)$ such that
\begin{equation}\label{e:polyidcplx1}
\left(\sum_{j=0}^n Z_j\chi_j\right)R(Z,\chi)=
\sum_{k=0}^mH_k(Z)\bar H_k(\chi),
\end{equation}
or
\begin{equation}\label{e:polyid2}
\norm Z\norm^2 R(Z,\bar Z)=
\norm H\norm^2.
\end{equation}
Substituting this in \eqref{e:polyid}, we obtain
\begin{equation}\label{e:polyid3}
\left(\norm Z\norm^2\right)^b\left(\norm F\norm^2 \right)^c=
\left(\norm Z\norm^2 \right)^a R(Z,\bar Z)^a.
\end{equation}
From \eqref{e:polyid3}, we can conclude that
the Hermitian polynomial $R(z,\bar z)$ belongs to various "positivity classes" introduced by D'Angelo and Varolin (see \cite{JPDVar04}; see also \cite{JPD11}). For instance, if $b\geq a$, then it follows that $R(Z,\bar Z)^a$ is an SOS; {\it however}, $R(z,\bar z)$ may not be an SOS itself in general. Thus, in order to use \eqref{e:polyid3} instead of \eqref{e:polyid} to estimate the dimension $m$, we would need to solve the SOS Problem in Section \ref{s:SOS} in its general form. Nevertheless, Huang's Lemma (described in Section \ref{s:SOS} above) applied to \eqref{e:polyid2} implies that $m\geq n$, which is a stronger lower bound in the first gap than that provided by Theorem \ref{thm:main1} for general, not necessarily rational mappings (see the discussion following the statement of Theorem \ref{thm:main1} and Remark \ref{rem:bad}). When $R(z,\bar z)$ is not an SOS, then the modification of $\omega_n$ obtained by applying $(\sqrt{-1}/2\pi)\bar\partial\partial$ to $\log$ of the left side of \eqref{e:polyid2} arises from the pullback of a mapping into a projective space with an indefinite Fubini-Study type "metric" (as in \cite{Calabi53}).
We should remark that for local holomorphic mappings $f$ and $h$, reducing the identity $$
(1+\norm z\norm^2)^b(1+\norm f\norm^2)^c=(1+\norm h\norm^2)^a
$$
to an identity of the form
\begin{equation}\label{e:notuseful}
(1+\norm z\norm^2)(1+Q(z,\bar z))=1+\norm h\norm^2
\end{equation}
is not useful, unless $Q(z,\bar z)$ itself is an SOS. Any mapping $h$ has a local analytic identity of this form with
$$
1+Q(z,\bar z)=\frac{1+\norm h\norm^2}{1+\norm z\norm^2}.
$$
We conclude this section by giving an example (taken from \cite{JPDVar04} and \cite{JPD11}) illustrating that even in the rational (polynomial) case, the situation in Theorem \ref{thm:main1} cannot be reduced to a situation covered by Theorem \ref{thm:main0}.
\begin{example}\label{ex:1}
Let $n=1$ and
\begin{equation}
R_\lambda(z,\bar z):=(1+|z|^2)^4-\lambda|z|^4=1+4|z|^2+(6-\lambda)|z|^4+4|z|^6+|z|^8.
\end{equation}
We observe that $R_\lambda(z,\bar )=1+Q_\lambda(z,\bar z)$ where $Q_\lambda(z,\bar z)$ is an SOS if and only if $\lambda\leq 6$. A straightforward calculation shows that
\begin{equation}
(1+|z|^2)R_\lambda(z,\bar z)=1+5|z|^2+(10-\lambda)|z|^4+(10-\lambda)|z|^6+5|z|^8+|z|^{10},
\end{equation}
which is of the form $1+\norm h\norm^2$ if and only $\lambda \leq 10$. Another calculation shows that
\begin{multline}\label{e:Rsq}
R_\lambda(z,\bar z)^2=1+1+8|z|^2+2(14-\lambda)|z|^4+8(7-\lambda)|z|^6+\left((6-\lambda)^2+34\right)|z|^8
\\+8(7-\lambda)|z|^{10}
+2(14-\lambda)|z^{12}+9|z|^{14}+|z|^{16},
\end{multline}
which is of the form $1+\norm f\norm^2$ if and only $\lambda \leq 7$. (According to \cite{JPDVar04} and \cite{JPD11}, there exists $k\geq 2$ such that $R_\lambda(z,\bar z)^k= 1+$ an SOS if and only if $\lambda<8$.) Thus, if we choose $\lambda=7$ and denote $R_7$ by $R$, then we have
\begin{equation}\label{e:Rprod}
(1+|z|^2)R(z,\bar z)=1+\norm h\norm^2,
\end{equation}
with $h=(h_1,\ldots, h_m)$ and $m=5$, but $R(z,\bar z)$ is not an SOS. However, if we square \eqref{e:Rprod} and use \eqref{e:Rsq}, then we obtain
\begin{equation}
(1+|z|^2)^2(1+\norm f\norm^2)=(1+\norm h\norm^2)^2,
\end{equation}
with $h=(h_1,\ldots, h_m)$ and $m=5$, $f=(f_1,\ldots,f_d)$ and $d=6$. Thus, this is an example where the conditions in Theorem \ref{thm:main1} are satisfied (with $a=b=2$, $c=1$, $e=d=6$), but which cannot be reduced to a situation covered by Theorem \ref{thm:main0}. Of course, in this example the lower bound for $m$ (which is 5) provided by Theorem \ref{thm:main1} (ii) (or the one provided by Huang's Lemma as above) is a poor estimate, as it reduces to $m\geq 1$ (which is a trivial bound) when $n=1$. We do note, however, that $m<d=e$ in contrast to the situation in Theorem \ref{thm:main0}.
\end{example}
\def\cprime{$'$}
|
\section{INTRODUCTION}
\begin{figure*}
\begin{center}
\includegraphics[width=11cm]{Figure1.pdf}
\end{center}
\caption{
{\bf Ligand-Receptor Interactions can give rise to Turing Patterns.} (Color online) (A) Spatial patterns via a Turing mechanism can result from cooperative receptor-ligand interactions, where $m$ receptors ($R$) and $n$ ligand molecules ($L$) form an active complex that upregulates the receptor concentration by increasing its expression, limiting its turn-over, or similar. Importantly, the highest receptor and ligand concentrations are observed in different places. (B) In case of the standard network (panel A), Turing patterns emerge only for a small subset of the parameter range of the receptor and ligand production rates, $a$ and $b$. $a_{\textup{max}}$ denotes the maximal value of the receptor production rate, while $b_{\textup{min}}$ and $b_{\textup{max}}$ denote the minimal and maximal ligand production rates. (C) Additional feedbacks (red and dashed, blue arrows) can be mediated by the ligand-receptor complex, $R^2L$; $\leftrightarrow$ indicates receptor-ligand interactions, $\dashv$ inhibitory interactions, and $-\bullet$ up-regulating interactions. (D) The negative feedbacks in panel C (network U5 in Fig. S1) result in a larger Turing space when the response threshold $p$ is lowered from $p=1$ (blue, shaded area) to $p=0.1$ (yellow area). (E) The size of the Turing space for the network in panel C (network U5 in Fig. S1) increases as the response threshold $p$ is lowered. As a measure for the size of the Turing space, we record the maximum of the receptor production rate, $a_{\textup{max}}$, and the ratio of the maximal and minimal ligand production rates $b_{\textup{ratio}} = \frac{b_{\textup{max}}}{b_{\textup{min}}}$, for which Turing patterns can emerge. $a=0$ is part of the Turing space and negative values of $a$ have no physiological interpretation.}
\label{Fig0}
\end{figure*}
The development of complex organisms requires the repeated, reliable emergence of pattern in a cell or tissue from a homogenous, noisy distribution of components, also in the absence of any polarizing queues. It is a long-standing question how stereotyped patterns can emerge during development. Alan Turing proposed a simple reaction-diffusion-based mechanism \cite{Turing:1952} that has since been shown to have the potential to give rise to a large variety of patterns from noisy, homogenous starting conditions \cite{Kondo:Science:2010, Murray:MathematicalBiology3RdEditionIn2Volumes:2003, Vanag2001}.
Mathematical analysis reveals the types of interactions between the molecular components that can give rise to Turing patterns \cite{Gierer:Kybernetik:1972, Murray:MathematicalBiology3RdEditionIn2Volumes:2003, Prigogine:JChemPhys:1967, Prigogine:TheJournalOfChemicalPhysics:1968}. While many different Turing mechanisms have been proposed to explain pattern formation in biology, it has remained difficult to identify the molecular components \cite{Kondo:Science:2010}. The suggested Turing components are typically two diffusible, extracellular proteins \cite{Cho:Development:2011, Economou:NatGenet:2012, Sick:Science:2006}. However, one of the requirements for Turing patterns is a large difference in the diffusion coefficient between the two Turing components. While a number of chemical systems have been engineered where the diffusion speed of one of the components of the Turing system is strongly reduced, e.g. the Belousov--Zhabotinsky reaction in water-in-oil aerosol microelmulsion \cite{Vanag2001} or in a system with a low-mobility complexing agent \cite{Horvath2009}, these setups do not readily translate to biological systems. For biological systems, it has been suggested that differences in diffusion speed may arise from transient differences in the interactions with the extracellular matrix \cite{Muller:Science:2012}. A number of theoretical studies seek to overcome the requirement of a large difference in diffusivity of Turing components, and an emergence of Turing pattern has been shown to be possible also in the presence of a single diffusive specie coupled to a quenched oscillator \cite{Hsia:PlosComputBiol:2012}; cell migration rather than diffusion has been proposed to result in Turing instabilities \cite{Nakamasu:Pnas:2009, Payne:MolSystBiol:2013}. Finally, cross-diffusion and non-linear diffusion have been shown to support the formation of Turing-type patterns, such that Turing patterns can arise for any ratio of the main diffusivities \cite{Pang2004, Hamidi2012, Butler2011, Ridolfi2011, Fanelli:TheEuropeanPhysicalJournalB:2013, Madzvamuse:JMathBiol:2014}. Cross-diffusion has been shown to arise in crowded environments with finite carrying capacity, i.e. if diffusion is limited when local concentrations or densities reach the carrying capacity \cite{Fanelli:TheEuropeanPhysicalJournalB:2013, Bullara:PhysicalReviewE:2013}.
Another problem with the applicability of Turing mechanisms to biological pattern formation concerns the size of the parameter space that gives rise to Turing patterns, the Turing space. This parameter space is small for all known Turing mechanisms in the sense that kinetic parameters can be varied only a few fold as long as physiological constraints on the kinetic constants and relative diffusion constants are respected \cite{Murray:JournalOfTheoreticalBiology:1982}. It is therefore unclear how nature could have evolved such mechanism in the first place and how it could have been re-used in different settings during the evolution of new species. Moreover, biological systems are noisy, and time delays as may arise from the multi-step nature of protein expression as well as domain growth and the resulting changes in source and sink terms, may severely affect the existence and type of Turing patterns, though some of these effects as well as further regulatory interactions may somewhat increase the size of the Turing space \cite{Crampin1999, Gjorgjieva2007, Madzvamuse:JMathBiol:2010, Woolley2011, Maini:Interface:2012, Woolley2012, Klika2012, Gaffney2013}.
We recently noticed that ligand-receptor interactions of the form shown in Fig.\ \ref{Fig0}A can give rise to Turing patterns \cite{Menshykau:PhysicalBiology:2013, Tanaka:PhysicalBiology:2013, Menshykau:PlosComputBiol:2012, Badugu:SciRep:2012} as long as the following constraints are met by the the receptor-ligand interaction:
\begin{itemize}
\item Ligands must diffuse much faster than receptors ($d \gg 1$), as is generally the case \cite{Choquet:2003fk,Ries:2009p20248,Kumar2010, Hebert2005}.
\item Receptor-dependent ligand removal must dominate over receptor-independent ligand decay, as is generally the case because unspecific decay is typically much slower than active protein turn-over.
\item Ligands and receptors must bind cooperatively, as is the case for many ligand-receptor pairs \cite{Parkash2008, Jing1996, SpivakKroizman1994, Plotnikov1999, DiGabriele1998, Ibrahimi2005, Goetz2006, Scheufler1999, Nickel2001}.
\item Ligand-receptor complex formation must be fast compared to the other processes, such that we have a quasi-steady state for the ligand-receptor complex concentration. This is the case if the on-rate is very high, i.e.\ binding is diffusion limited, as is the case for many ligand-receptor pairs \cite{ReceptorBook}.
\item The receptor-ligand complex must upregulate the receptor concentration, as has been observed for several receptor systems \cite{Pepicelli1997, Lu2009, Estival1996, Ota2012, Chen1996, Weaver2003, Merino1998}. This positive feedback needs to operate far from saturation, i.e. if we describe the positive regulation by a Hill function of the form $\frac{R^2L}{R^2L + K}$, we require $R^2L \ll K$. Thus, this positive feedback must be rather inefficient.
\end{itemize}
If these conditions are met, the interactions between the receptor, $R$, and the ligand, $L$, result in Schnakenberg-type kinetics \cite{Schnakenberg:JTheorBiol:1979} of the form
\begin{eqnarray}
\hspace*{-0.4cm} \frac{\partial R}{\partial t} &=& \Delta R+ \gamma f(R,L) \ \textrm{with} \ f(R,L)=a-R+R^2 L \label{eq:1} \\
\hspace*{-0.4cm} \frac{\partial L}{\partial t} &=& d \Delta L+ \gamma g(R,L) \ \textrm{with} \ g(R,L)=b-R^2 L, \label{eq:2}
\end{eqnarray}
which correspond to the so-called activator-depleted substrate Turing kinetics, first described by Gierer and Meinhardt \cite{Gierer:Kybernetik:1972}, and which are very similar to the chemical Turing system first described by Prigogine and co-workers \cite{Prigogine:TheJournalOfChemicalPhysics:1968}. The detailed derivation of these equations for receptor-ligand interactions can be found in previous publications \cite{Menshykau:PhysicalBiology:2013, Tanaka:PhysicalBiology:2013, Menshykau:PlosComputBiol:2012, Badugu:SciRep:2012} and in the Appendix I. The $\Delta R$ and $d\Delta L$ terms represent the diffusion terms, where $d$ is the relative diffusion constant of ligand and receptor. Ligands typically diffuse faster than their receptors, $d \gg1$ \cite{Choquet:NatRevNeurosci:2003, Kumar:BiophysJ:2010, Ries:NatMethods:2009}, thus naturally meeting the Turing condition of different diffusivities. Receptor diffusion is restricted to single cells, and we have previously shown that patterns also emerge on such cellularized domains \cite{Menshykau:PlosComputBiol:2012}. The constants $a$ and $b$ are the receptor and ligand production rates. The $-R$ term describes the ligand-independent decay of the receptor at a rate proportional to the available receptor concentration, so-called linear decay. The term $R^2 L$ represents the quasi-steady state concentration of the receptor-ligand complex. Signaling complexes with a different stochiometry also result in Turing patterns \cite{Menshykau:PlosComputBiol:2012}. The 'minus' term in Eq.\ \eqref{eq:2} then reflects the receptor-dependent ligand removal rate, while the 'plus' term in Eq.\ \eqref{eq:1} reflects the combined effects of ligand-induced receptor removal and ligand-induced receptor accumulation on the cell membrane (by increased transcription, translation, recycling, less constitutive removal or similar). The $\gamma$ term arises in the non-dimensionalization of the model (Eq. \ref{eq:Schnakenberg}) and is useful as it is proportional to the domain area, and it gives the relative strength of the reaction and diffusion terms \cite{Murray:MathematicalBiology3RdEditionIn2Volumes:2003}.
A number of ligand-receptor systems meet the above conditions, including Hedgehog and its receptor PTCH \cite{Tanaka:PhysicalBiology:2013, Menshykau:PlosComputBiol:2012, Chen1996, Weaver2003, Goetz2006}, BMPs and their BMP receptors \cite{Badugu:SciRep:2012, Scheufler1999, Nickel2001, Merino1998}, GDNF and its receptor RET \cite{Menshykau:PhysicalBiology:2013, Parkash2008, Jing1996, Pepicelli1997, Lu2009}, as well as FGFs and their FGF receptors \cite{SpivakKroizman1994, Plotnikov1999, DiGabriele1998, Ibrahimi2005, Estival1996, Ota2012}. Thus, all of these proteins are multimers, and, by a range of mechanisms, the formation of the multimeric ligand-receptor complexes enhances the concentration of receptors on the membrane, as recently reviewed \cite{Iber:OpenBiol:2013}. We further showed that models based on these proteins could recapitulate the relevant wildtype and mutant expression patterns in the respective developmental systems \cite{Menshykau:PhysicalBiology:2013, Tanaka:PhysicalBiology:2013, Menshykau:PlosComputBiol:2012, Badugu:SciRep:2012, Celliere:BiologyOpen:2012}.
Here we show that ligand-receptor based Turing mechanisms can have significantly enlarged Turing spaces if we include negative feedbacks or couple several Turing modules, as generally found in biological systems. Similarly, the restriction of receptors to single cells and their clustering further increases the size of the Turing space. We conclude that a receptor-ligand based Turing mechanism offers a realistic mechanism to implement the Turing mechanism in a biological setting. The observation that the restriction of receptors to cells is sufficient to massively increase the Turing space offers an explanation of how Turing patterns may have first evolved in nature; additional feedbacks could then further enlarge the Turing space.
\section{RESULTS}
The Turing mechanism has been analysed extensively, and the parameter space that permits Turing patterns to emerge can easiest be determined with the help of a linear stability analysis \cite{Murray:MathematicalBiology3RdEditionIn2Volumes:2003}; see the Appendix II. To keep the analysis feasible, it is advisable to consider as simple models as possible, and to restrict the number of parameters to a minimum. The non-dimensional ligand-receptor based Turing model (Eq.\ \eqref{eq:1}-\eqref{eq:2}) has four parameters: the relative ligand/receptor diffusion constant $d$, the receptor production rate $a$, the ligand production rate $b$, and the scaling factor $\gamma$. The parameters $a$, $b$, $d$ determine whether Turing patterns can emerge, while the scaling factor $\gamma$ determines whether the domain is sufficiently large for Turing patterns to emerge. We therefore do not need to analyse $\gamma$ here. The relative diffusion constant of ligands and receptors, $d$, affects the size of the Turing space in that a larger $d$ results in a larger Turing space \cite{Murray:MathematicalBiology3RdEditionIn2Volumes:2003}. Since this effect is well documented, but limited by the physiological difference between the diffusion constants of ligands and receptors we fixed the relative diffusion constant in our analysis. For a simple receptor-ligand based Turing system, in which receptor and ligand bind cooperatively and upregulate the receptor concentration (Fig.\ \ref{Fig0}A), both parameter values $a$ and $b$ produce Turing patterns only within a small range (Fig.\ \ref{Fig0}B), i.e.\ the ligand production rate can at most be halved or doubled without leaving the Turing space. The Turing space is thus very small, even though the relative diffusion constant, $d=50$, between ligands and receptors was chosen to be rather large compared to what could be justified for two soluble ligands. We will now analyse the impact of feedbacks, receptor clustering, and the restriction of receptors to single cells on the size of the Turing Space.
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth]{Figure2.pdf}
\end{center}
\caption{
{\bf Negative Feedbacks by Receptor-Ligand Complexes result in Turing patterns with large Turing Spaces.} (Color online) (A) The simulated network architecture. Two receptors $R$ interact with one dimeric ligand $L$ to form a receptor-ligand complex $R^2L$ (black arrows, $\leftrightarrow$). The receptor-ligand complex upregulates the presence of receptor ($-\bullet$). In addition to these core interactions that can result in a Turing mechanism, we considered negative feedbacks ($\dashv$) on the ligand production (red arrow) and / or the receptor production (blue, dashed arrow). (B) A negative feedback on the receptor production rate (blue dashed arrow in panel A) increases the Turing parameter space for the receptor production rate, $a$ (blue squares in panel B) compared to the standard network (black part of the network in panel A and black star in panel B). A negative feedback on ligand production (red, solid arrow in panel A) enlarges the Turing parameter space for the ligand production rate, $b$, (red circles in B). In the presence of both feedbacks the Turing parameter space is enlarged along both axes (green triangles in panel B). The feedback effects are stronger, the lower the feedback threshold, $p$ ($p$= 0.01, 0.1, 1, 10, 100). The grey arrow indicates the direction, in which the feedback threshold, $p$ decreases.}
\label{Fig1u}
\end{figure}
\subsection{The impact of Feedbacks on the Turing Space of a single receptor-ligand based Turing Module}
Feedbacks are ubiquitous in biological signaling systems. In the framework of receptor-ligand-based Turing mechanisms, feedbacks result from regulatory interactions of the receptor-ligand complex, $R^2L$ (Fig.\ \ref{Fig0}C). To encode feedbacks mediated by receptor-ligand signaling, we modified the reaction terms $f(R,L)$ and $g(R,L)$ in the Turing model (Eq.\ \eqref{eq:1}-\eqref{eq:2}). See Supplemental Material \cite{Supp} for the list of all tested models with additional feedbacks. Thus a positive feedback on receptor or ligand expression would be obtained by adding a term $p R^2L$ to the respective equation and/or by multiplying the constitutive receptor and ligand production rates $a$ and $b$ with the factor $\frac{R^2 L}{R^2 L+p}$. A negative feedback would be obtained by multiplying the constitutive receptor and ligand expression rates $a$ and $b$ with the factor $\frac{1}{R^2 L / p+1}$. The new parameter $p$ represents the response threshold to the receptor-ligand complex. Figure \ref{Fig0}D illustrates the impact of feedbacks on the Turing space for the regulatory system with two additional negative feedbacks shown in Fig.\ \ref{Fig0}C. For a large response threshold ($p=1$) the Turing space is similar in size to the non-feedback case (compare the blue shaded area in Fig.\ \ref{Fig0}D to the Turing space in Fig.\ \ref{Fig0}B). As we lower the response threshold to $p=0.1$ and thus increase the strength of the negative feedbacks the Turing space increases in size, i.e.\ both the maximal receptor production rate, $a_{\textup{max}}$, as well as the range of ligand expression rates $[b_{\textup{min}}, b_{\textup{max}}]$ increases (yellow area in Fig.\ \ref{Fig0}D); the minimum of $a$ is negative and $a_{\textup{max}}$ thus defines the size of the physiological parameter range, $[0, a_{\textup{max}} ]$. As the response threshold $p$ is lowered further, the size of the Turing space further increases (Fig.\ \ref{Fig0}E).
We next systematically analysed eleven positive, negative, and mixed feedback architectures that were obtained by including feedbacks of the receptor-ligand complex ($R^2L$) on the receptor ($a$) and/or ligand production rates ($b$), as well as on the rate of receptor up-regulation upon receptor-ligand binding (for details see Appendix II, Fig.\ S1). Figure \ref{Fig1u}A,B shows the three cases with the largest Turing space out of the eleven cases analyzed. For better readability, we only record the maximal receptor production rate, $a_{\textup{max}}$ as well as the ratio, $b_{\textup{ratio}} = \frac{b_{\textup{max}}}{b_{\textup{min}}}$, of the maximal and minimal ligand production rates that permit Turing patterns to emerge. We note that the ratio $b_{\textup{ratio}} = \frac{b_{\textup{max}}}{b_{\textup{min}}}$ is biologically more relevant than the absolute size of the Turing space, $\Delta b= b_{\textup{max}} - b_{\textup{min}}$, because in biology relative changes in regulatory control and thus in production rates are particularly relevant; the absolute values are typically very difficult to measure. The largest Turing spaces are obtained with negative feedbacks. When the negative feedback is applied to the constitutive receptor expression, $a$ (blue), the maximal value of $a$ increases relative to the standard model (black spot) as the response threshold, $p$, is lowered; the minimum of $a$ is negative and $a_{\textup{max}}$ thus defines the size of the physiological parameter range, $[0, a_{\textup{max}}]$. If a feedback is applied to the ligand expression rate, $b$, then, as the response threshold, $p$, is lowered, the range of $b$ increases (red) compared to the standard model (black spot). The largest Turing spaces, expanded both along the $a$ and $b$ axes, are observed when negative feedbacks are applied to both, the receptor and ligand expression rates (green). The impact of the negative feedbacks can be observed for a wide range of the new parameters, $p$, and becomes stronger the smaller the value of the response threshold $p$ (Fig.\ \ref{Fig1u}B). As the response threshold $p$ is increased, the maximal values of $a$, and the range of $b$ all attain the value of the standard receptor-ligand model and thus all converge in the black spot in Fig.\ \ref{Fig1u}B. In summary, substantially enlarged Turing spaces are observed when signaling by the the receptor-ligand complex lowers the receptor production rate (Fig.\ \ref{Fig1u}B, blue), or the ligand production rate (Fig.\ \ref{Fig1u}B, red), or both (Fig.\ \ref{Fig1u}B, green).
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth]{Figure3.pdf}
\end{center}
\caption{
{\bf Coupling of several Receptor-Ligand based Turing modules further enlarges the Turing Space.} (Color online) (A) The simulated network architecture. Two receptor-ligand based Turing modules, as analysed in Fig. 2 (black arrows, $\leftrightarrow$, $-\bullet$), are coupled via additional negative feedbacks ($\dashv$) on the ligand production rates (red, solid arrows) and / or the receptor production rates (blue, dashed arrows). (B) A negative feedback on the receptor production rate (dashed, blue line in panel A) increases the Turing parameter space for the receptor production rate, $a$ (blue squares in panel B) compared to the standard network (black part of the network in Fig. 2A and black star in panel B). A negative feedback on ligand production (red, solid arrow in panel A) enlarges the Turing parameter space for the ligand production rate, $b$, (red circles in B). In the presence of both feedbacks the Turing parameter space is enlarged along both axes (green triangles in panel B). The feedback effects are stronger, the lower the feedback threshold, $p$ ($p$= 0.01, 0.1, 1, 10, 100). The grey arrow indicates the direction, in which the feedback threshold, $p$ decreases.}
\label{Fig1c}
\end{figure}
\subsection{Coupled Turing Modules}
In patterning processes, several receptor-ligand systems often interact, e.g.\ SHH, FGF10, and BMP together with their receptors regulate branching morphogenesis of the lung and several glands, while GDNF, FGF10, and WNT and their receptors regulate kidney branching morphogenesis, as recently reviewed \cite{Iber:OpenBiol:2013}. We were therefore interested how the interaction of several such Turing modules would affect the Turing space.
To that end, we carried out a systematic analysis of possible feedback interactions between two separate receptor-ligand-based Turing systems (for details see the Appendix II. E). The studied network architectures, systems of equations, and Turing spaces are shown in Fig.\ S1. Figure \ref{Fig1c}A summarizes the coupled Turing modules with the largest Turing spaces. Here, similar as for uncoupled modules (Fig.\ \ref{Fig1u}), the largest Turing space is observed when a negative feedback acts on the production rates (Fig.\ \ref{Fig1c}). We notice that coupling of the two Turing systems via a negative feedback on the constitutive receptor expression rates, $a$, results mainly in an increase in the parameter space of $a$ (Fig.\ \ref{Fig1c}B, blue), while coupling the two Turing systems via a negative feedback on the constitutive ligand production rate $b$ results mainly in an increase in the parameter space for $b$ (Fig.\ \ref{Fig1c}B, red). The asymmetrically coupled modules with one feedback on $a$ and one on $b$ have a very large (possibly infinitely large) parameter space (Fig.\ S1 panels C6, C8 and C10). However, the parameter range is very narrow and extends towards infinity only along the $b$-axis, while it is bounded above on the $a$-axis. A massive increase in the size of the Turing space is observed when the two Turing modules are coupled by four negative feedbacks, such that all constitutive receptor and ligand expression rates are regulated by negative feedbacks (Fig.\ \ref{Fig1c}B, green and Fig. S1 panel C11). In this case, the parameter space dramatically increases in both directions as $p$ is lowered, such that already for $p=0.1$, the parameter ranges of both $a$ and $b$ expand by more than four orders of magnitude compared to a single receptor-ligand based Turing model, and further increase as $p$ is lowered (Fig.\ \ref{Fig1c}B, green and Fig. S1 panel C11).
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth]{Figure4.pdf}
\end{center}
\caption{
{\bf Negative Feedbacks enlarge the Turing space by limiting the effective production rates.} The plot of the (A) effective receptor production rate $a_\mathrm{eff}=\frac{a}{\max (R^2 L) / p+1}$ versus the receptor production rate $a$, and (B) the plot of the effective ligand production rate $b_\mathrm{eff}=\frac{b}{\max(R^2 L) / p+1}$ versus $b$ show that, as a result of the negative feedbacks, the effective production rates remain in a narrow range, even as $a$ and $b$ are greatly changed. The calculation was carried out for the symmetrically coupled Turing system, shown in green in Figure \ref{Fig1c}B. }
\label{Fig1r}
\end{figure}
\subsection{Negative Feedbacks enlarge the Turing space by limiting the effective production rates}
We wondered why negative feedbacks would enlarge the Turing space. To this end, we plotted the effective production rates $a_\mathrm{eff}=\frac{a}{\max (R^2 L) / p+1}$ and $b_\mathrm{eff}=\frac{b}{\max(R^2 L) / p+1}$ for the coupled Turing systems with the largest Turing space (Fig.\ \ref{Fig1c}B, green) versus $a$ and $b$, respectively (Fig.\ \ref{Fig1r}). We find that the effective production rates are much smaller than what the parameter values $a$ and $b$ would suggest, and almost lie within the standard small Turing space. Thus, the negative feedback effectively corrects the receptor and ligand production rates, and thereby enables the Turing mechanism to tolerate a much wider range of production rates.
\subsection{The restriction of receptors to single cells enlarges the Turing space}
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth]{Figure5.pdf}
\end{center}
\caption{
{\bf The restriction of receptors to single cells enlarges the Turing space.} (Color online) \small{ (A) Cartoon of the computational domain: diffusion of receptors is restricted to single cells, while ligand can diffuse over the entire computational domain. (B) Solution of the receptor-ligand model on a 1D, 2D and 3D (left to right) cellularized computational domain. The ligand (upper row), receptors (middle row), and ligand-receptor complexes (bottom row) pattern the domain. We provide the concentration levels (in arbitrary units) on the vertical axis for the 1D domain (left column), and intensities as colour code (blue- low; red - high) on the 2D and 3D domains. To distinguish cell boundaries on the 1D domain we alternate black and grey lines. (C) The size of the Turing space increases as the domain of fixed size is split into more cells, $N$. Triangles show the results for $N = 10$ and $N = 100$ cells. The black star reports the Turing space for the standard model, $N=1$. (D) Patterns of receptor-ligand complexes that extend over several cells can be obtained with a diffusive component, $T$, that is produced in response to the formation of receptor-ligand complexes, and that enhances the abundance of receptors on neighbouring cells. The grey arrow indicates the direction, in which the feedback threshold, $p$, decreases.}
}
\label{Fig2}
\end{figure}
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth]{Figure6.pdf}
\end{center}
\caption{
{\bf Receptor clustering enlarges the Turing space.} (Color online) (A) The simulated network architecture. Clusters of $2 n$ receptors $R$ interact with $n$ dimeric ligands $L$ to form a receptor-ligand complex $(R^2L)^n$ (black arrows, $\leftrightarrow$). The receptor-ligand complex upregulates the presence of receptor (black interaction, $-\bullet$). In addition to these core interactions that can result in a Turing mechanism, we considered negative feedbacks on the ligand production (red, solid arrow, $\dashv$) and / or the receptor production (blue, dashed arrow, $\dashv$). (B) Higher cooperativity, $n>1$, as may result from larger receptor-ligand clusters further increases the size of the Turing space. The $n$-dependent increase was calculated for $p$ = 0.01, 0.1, 1, 10, 100 for case U5 in Fig. S1. The grey arrow indicates the direction, in which the feedback threshold, $p$, decreases. }
\label{Fig6}
\end{figure}
\begin{figure*}
\begin{center}
\includegraphics[width=0.8\textwidth]{Figure7.pdf}
\end{center}
\caption{
{\bf Substantially enlarged Turing spaces for physiological networks.} (Color online) (A) The SHH-FGF10 network in the control of lung branching morphogenesis. For details see text. (B) Schematic representation of the regulatory network for lung branching morphogenesis in panel A. (C) The Turing space of such a physiological model is huge, and further increases as the feedback threshold, $p$, is lowered. The red triangles represent the Turing spaces for $p_1 = q = 0.1$ (positive feedback on ligand and receptor, respectively) and $p_2$ = 0.01, 0.1, 1, 10, 100 (negative feedback); the black star represents the size of the Turing space of the standard network in Fig. 2A (black part). The grey arrow indicates the direction, in which the feedback threshold, $p$, decreases. $\leftrightarrow$ indicates binding interactions, $\dashv$ indicates inhibitory interactions, and $-\bullet$ indicates up-regulating interactions.}
\label{Fig:physiol}
\end{figure*}
So far, we have treated receptors in the same way as the ligand, just with a smaller diffusion coefficient. However, receptors are confined to single cells and thus cannot diffuse from one cell to the next. Moreover, they often cluster on the cell surface. We therefore next studied Turing patterns on cellular domains where receptors are confined to single cells, while ligands can diffuse within the tissue (Fig. \ref{Fig2}A). The computational details of the implementation have previously been described \cite{Menshykau:ProceedingsOfComsolConference2012:2012}, and details of the implementation are given in Appendix III. In brief, to restrict diffusion of receptors to a single cell in 1D and 2D models (Fig. \ref{Fig2} B, left and middle panel, respectively), we set no-flux boundary conditions for receptor at the pseudo-cell boundary, while ligand was free to diffuse in the entire domain. In the 3D model the cell surfaces were approximated as spheres (Fig. \ref{Fig2}B, right panel), and both ligands and receptors were produced on the spheres' surfaces. Diffusion of receptors was restricted to the surface of each sphere, while ligand was free to diffuse also in the intercellular space; the details of the implementation have been previously described \cite{Vollmer:CordConferenceProceedings:2013}.
We observe the emergence of patterns on 1D, 2D and 3D cellularized domains (Fig. \ref{Fig2}B), and as a tissue domain of a given size is divided into more (and thus smaller) cells, to which the receptors are restricted, the Turing space increases (Fig. \ref{Fig2}C). Interestingly, however, cells with a high level of receptor-ligand complexes occur only as isolated spots (Fig. \ref{Fig2}B, red spots), while clusters of such active cells are not observed. To obtain clusters of active cells we have to include a second diffusively component, $T$, that is secreted by the active cells and that activates neighboring cells (Fig. \ref{Fig2}D).
\subsection{Receptor clustering enlarges the Turing space}
Receptors often cluster on cell membranes, either as pre-clusters or induced by multimeric ligand. Clustered receptor-ligand complexes may cooperate \cite{Bray:Nature:1998}, such that regulation is not mediated by a single ligand-receptor complex, but by the cluster. We then have $(R^2L/p)^n$ with $n>1$ in Eq.\ \eqref{eq:1}-\eqref{eq:2} instead of $R^2L/p$. As we increase $n$, we observe a further increase in the size of the Turing space (Fig.\ref{Fig6}A,B). In summary, both receptor clustering and the cellular restriction of receptors greatly increase the Turing space.
\subsection{Physiological Turing Models}
Physiological networks harbour many feedbacks and we wondered by how much the size of the Turing space would be increased in physiological settings. Here, we considered the network that controls branching morphogenesis in the lung (Fig. \ref{Fig:physiol}A); similar networks also operate in the prostate, salivary gland, and in the pancreas \cite{Iber:OpenBiol:2013}. Core to the control of lung branching morphogenesis are FGF10 and SHH as no branching is observed in the null mutants \cite{Abler:DevDyn:2009, Chiang:Nature:1996, Pepicelli:CurrBiol:1998, Bellusci:Development:1997b}, and expression of dominant negative \textit{Fgfr2} blocks lung branching, but not outgrowth \cite{Peters:EmboJ:1994}.
FGF10 upregulates \textit{Shh} expression \cite{Abler:DevDyn:2009} and the expression of its own receptor, FGFR2b \cite{Cui:IntJMolSci:2013}, while SHH signaling downregulates \textit{Fgf10} expression\cite{Bellusci:Development:1997} and upregulates the expression of its own receptor \textit{Ptch1} \cite{Ingham:GenesDev:2001} (Fig.\ref{Fig:physiol}A). We have previously shown that the SHH-PTCH kinetics can be described by Eq.\ref{eq:1}-\ref{eq:2} \cite{Tanaka:PhysicalBiology:2013, Menshykau:PlosComputBiol:2012}; similar equations can also be derived for the FGF10-FGFR2b kinetics; see Appendix I for a general derivation of the ligand-receptor kinetics. The particular stoichiometry in Eq. \eqref{eq:1}-\eqref{eq:2} assumes the binding of one ligand dimer to two receptor monomers. In case of FGF10, monomeric binding of one FGF10 dimer to its trivalent FGFR2b receptor triggers dimerization of the FGF10-receptor complex \cite{Ibrahimi:MolecularAndCellularBiology:2005}; SHH is a multimer that may form higher order complexes with its receptor PTCH1 \cite{Zeng:Nature:2001}. We have previously shown that similar Turing patterns can be observed also with such very different stochiometries \cite{Menshykau:PlosComputBiol:2012}. For ease of comparison, we stick to the standard model (Eq. \ref{eq:1}-\ref{eq:2}) for the FGF10 and SHH modules, though we note that larger SHH/PTCH1 clusters would further increase the Turing space (Fig. \ref{Fig2}E,F). The two signalling factors interact in that FGF10 upregulates \textit{Shh} expression \cite{Abler:DevDyn:2009}, while SHH signaling downregulates \textit{Fgf10} expression\cite{Bellusci:Development:1997}. The equations for the coupled network (Fig.\ref{Fig:physiol}B) are thus given by
\begin{eqnarray}\label{Eq_lung_network}
\textup{PTCH1:} \qquad & & \dot{R_1} = \Delta R_1 + f(R_1,L_1,R_2,L_2) \nonumber \\
\textup{SHH:} \qquad & & \dot{L_1} = d \Delta L_1 +g(R_1,L_1,R_2,L_2) \nonumber \\
\textup{FGFR2b:} \qquad & & \dot{R_2} = \Delta R_2 + \widetilde f(R_1,L_1,R_2,L_2) \nonumber \\
\textup{FGF10:} \qquad & & \dot{L_2} = d \Delta L_2 + \widetilde g(R_1,L_1,R_2,L_2)
\end{eqnarray}
with
\begin{eqnarray}\label{Eq_lung_network_detailed}
f(R_1,L_1,R_2,L_2) & =& a_1 - R_1 + qR_1^2L_1 \nonumber \\
g(R_1,L_1,R_2,L_2) & =& b_1- R_1^2L_1 + p_1R_2^2L_2 \nonumber \\
\widetilde f(R_1,L_1,R_2,L_2) & = &\frac{a_2}{1+\frac{R_2^2L_2}{p_2}} - R_2 + qR_2^2L_2 \nonumber \\
\widetilde g(R_1,L_1,R_2,L_2) & = &\frac{b_2}{1+\frac{R_1^2L_1}{p_2}} - R_2^2L_2
\end{eqnarray}
Here $R_1$ represents the receptor PTCH1, $L_1$ the ligand SHH, $R_2$ the receptor FGFR2b, and $L_2$ the ligand FGF10. The SHH-PTCH1 complex, $R_1^2L_1$ upregulates the receptor PTCH1 \cite{Ingham:GenesDev:2001}, i.e.\ $+ qR_1^2L_1$ in the term $ f(R_1,L_1,R_2,L_2) $, and inhibits the production of FGF10 \cite{Bellusci:Development:1997}, i.e.\ $\frac{b_2}{1+\frac{R_1^2L_1}{p_2}}$ in the term $\widetilde g(R_1,L_1,R_2,L_2)$. The FGF-receptor complex, $R_2^2L_2$, upregulates the production of SHH, i.e. $p_1R_2^2L_2$ in the term $g(R_1,L_1,R_2,L_2) $ and both upregulates \cite{Cui:IntJMolSci:2013}, i.e.\ $qR_2^2L_2$ in term $\widetilde f(R_1,L_1,R_2,L_2) $, and downregulates \cite{Monzat:JournalOfBiologicalChemistry:1996}, i.e.\ $\frac{a}{1+\frac{R_2^2L_2}{p_2}}$ in term $\widetilde f(R_1,L_1,R_2,L_2) $, the FGF receptor FGFR2b. The $- R_1^2L_1$ and $- R_2^2L_2$ terms represent ligand removal by receptor binding; receptor removal by ligand binding is absorbed in the $ + qR_1^2L_1$ and $+ qR_2^2L_2 $ terms as signalling-dependent receptor upregulation dominates ligand-induced receptor removal.
We find that the combination of these two modules (Fig. \ref{Fig:physiol}B) increases the range of the receptor production rate, $a$, by about $10^9$-fold as the threshold $p$ is lowered to 0.01, while the relative range of the ligand production rate, $b_2$, increases about 100-fold compared to the single receptor-ligand based Turing model (Fig.\ \ref{Fig:physiol}C). \\
\section{DISCUSSION}
Turing mechanisms can reproduce a wide range of biological patterning phenomena. However, it has remained unclear how they may be implemented on the molecular scale and how they could evolve in spite of the small sizes of their Turing spaces. We propose that ligand-receptor interactions give rise to Turing patterns, and we show that negative feedbacks, the coupling of Turing modules, and the restriction of receptors to single cells can greatly increase the size of the Turing space (Fig. \ref{Fig1u}, \ref{Fig1c}, \ref{Fig2}) and thus increase the range of parameter values for which Turing patterns will emerge in biological systems.
The conditions for ligand-receptor based Turing mechanisms, as summarised in the Introduction, are met by many different ligand-receptor pairs, and we have previously shown that receptor-ligand based Turing mechanisms can indeed well describe the patterning processes for a range of developmental systems \cite{Menshykau:PhysicalBiology:2013, Tanaka:PhysicalBiology:2013, Menshykau:PlosComputBiol:2012, Badugu:SciRep:2012,Iber:OpenBiol:2013}. Equally, negative feedbacks are prevalent in biological regulation and have previously been shown to enable robustness to noise \cite{Becskei:Nature:2000} and transient responsiveness \cite{Ma:Cell:2009}. We now propose that negative feedbacks enable robust patterning also for receptor-ligand based Turing mechanisms. Interestingly, also the restriction of receptors to single cells can further increase the size of the Turing space (Fig. \ref{Fig2}). This suggests a way of how Turing mechanisms may have first evolved. Cooperative interactions in receptor clusters and the introduction of feedbacks as well as the coupling of several Turing modules may then have further increased the size of the Turing space.
It will be important to test our theoretical insights by synthetically constructing such ligand-receptor-based Turing mechanism, and by establishing the key parameter values (rates of production, decay, diffusion coefficients, endogenous concentrations etc.) in the living systems. The Turing space of ligand-receptor systems with additional negative feedbacks should be sufficiently large that synthetic biology approaches can now obtain Turing patterns in spite of the difficulties to accurately control kinetic rates in synthetic biology approaches. Given their robustness and flexibility, we propose that receptor-ligand based Turing mechanisms are the likely standard way how Turing mechanisms are implemented in Nature.\\
\section{Materials \& Methods}
The Turing space was defined based on a Linear Stability Analysis \cite{Murray:MathematicalBiology3RdEditionIn2Volumes:2003} as described in Appendix II, using MATLAB. The implementation of cell-based models is described in Appendix III. The partial differential equations were solved in COMSOL Multiphysics 4.3 as described previously \cite{Vollmer:CordConferenceProceedings:2013, Germann:ProceedingsOfComsolConference2011:2011, Menshykau:ProceedingsOfComsolConference2012:2012}. Tests of the numerical methods are provided in Appendix IV. \\
\section{Acknowledgments}
The authors thank Patrick Fried and Jannik Vollmer for discussions.
\section*{Appendix I: Derivation of the Equations for the Receptor-Ligand Signaling Model}
As previously derived \cite{Badugu:2012ho,Menshykau:2012kg,Tanaka:2013wt}, the dynamics of receptors, $R$, ligands, $L$, and the ligand-receptor complex, $C$ (Fig. 1C), can be described by the following set of equations:
\begin{widetext}
\begin{eqnarray} \label{eqL}
\dot{\![\mathrm{L}]} = \underbrace{\overline{D}_\mathrm{L} \overline{\Delta} [\mathrm{L}]}_{\text{\ diffusion}} + \underbrace{ \overline{\rho}_{\mathrm{S}}}_{\text{\ production}} \underbrace{-\overline{\delta}_{\mathrm{L}}[\mathrm{L}]}_{\text{\ degradation}} \underbrace{-n\, \overline{\mathrm{k}}_\mathrm{on}[\mathrm{R}]^m[\mathrm{L}]^n +n \overline{k}_\mathrm{off}[\mathrm{C}]}_{\text{\ complex formation}}
\end{eqnarray}
\begin{eqnarray} \label{eqR}
\dot{\![\mathrm{R}]} &=& \underbrace{\overline{D}_\mathrm{R} \overline{\Delta} [\mathrm{R}]}_{\text{\ diffusion}} +\underbrace{ \overline{\rho}_{\mathrm{R}}+\mu([\mathrm{C}])}_{\text{\ production}} \underbrace{-\overline{\delta}_{\mathrm{P}}[\mathrm{P}]}_{\text{\ degradation}} \underbrace{-m\, \mathrm{k}_\mathrm{on}[\mathrm{R}]^m[\mathrm{L}]^n +m \overline{k}_\mathrm{off}[\mathrm{C}]}_{\text{\ complex formation}}
\end{eqnarray}
\begin{eqnarray} \label{eqC}
\dot{\![\mathrm{C}]} &=& \underbrace{\overline{D}_\mathrm{C} \overline{\Delta} [\mathrm{C}]}_{\text{\ diffusion}}+\underbrace{\, \mathrm{k}_\mathrm{on}[\mathrm{R}]^m[\mathrm{L}]^n - \overline{k}_\mathrm{off}[\mathrm{C}]}_{\text{\ complex formation}} \underbrace{-\overline{\delta}_{\mathrm{C}}[\mathrm{C}]}_{\text{\ degradation}}.
\end{eqnarray}
\end{widetext}
Here $[X]$ denotes the concentration of component $X$, $\overline{D}_X$ denotes the diffusion coefficient, $\overline{\rho}_X$ the production rate constant, and $\overline{\delta}_X$ the first order degradation rate constant of component $X$. $\mu([C])$ specifies a function that describes the ligand-receptor dependent up regulation of receptor production. $k_\mathrm{on}$ denotes the rate constant for the formation, and $k_\mathrm{off}$ the rate constant for the dissociation of the ligand-receptor complex. $m$ and $n$ specify the number of receptors and ligands that bind in the ligand-receptor complex.\\
Assuming that the dynamics of the complex are fast compared to those of the other components, we can introduce a quasi-steady state approximation
\begin{eqnarray} \label{eqC}
0 &=& \underbrace{\, \mathrm{k}_\mathrm{on}[\mathrm{R}]^m[\mathrm{L}]^n - \overline{k}_\mathrm{off}[\mathrm{C}]}_{\text{\ complex formation}} \underbrace{-\overline{\delta}_{\mathrm{C}}[\mathrm{C}]}_{\text{\ degradation}},
\end{eqnarray}
and thus arrive at the quasi-steady state concentration of complex [C]$_\mathrm{SS}$
\begin{eqnarray} \label{eqC_qstst}
[\mathrm{C}] _\mathrm{SS}= \frac{\overline{k}_\mathrm{on}}{\overline{k}_\mathrm{off} + \overline{\delta}_{\mathrm{C}}} [\mathrm{R}]^m [\mathrm{L}]^n = \overline{\Gamma} [\mathrm{R}]^m [\mathrm{L}]^n,
\end{eqnarray}
where $\overline{\Gamma}= \frac{\overline{k}_\mathrm{on}}{\overline{k}_\mathrm{off} + \overline{\delta}_{\mathrm{C}}}$. The concentration of bound receptor, [C], is thus proportional to $[\mathrm{R}]^m\mathrm{[S]^n}$. Furthermore, assuming that the rate of receptor upregulation in response to receptor-ligand signaling $\mu([\mathrm{C}]) = \overline{v} [\mathrm{C}]= \overline{v} \overline{\Gamma}[\mathrm{R}]^m[\mathrm{L}]^n $ depends linearly on the ligand-receptor complex concentration, $[C]$, we obtain the following set of PDEs:
\begin{eqnarray} \label{eq:Sf}
\dot{\![\mathrm{L}]} &=& \overline{D}_\mathrm{L}\overline{\Delta} [\mathrm{L}] + \overline{\rho}_{\mathrm{L}} - n\overline{\delta}_{\mathrm{C}}\overline{\Gamma} [\mathrm{R}]^m[\mathrm{L}]^n -\overline{\delta}_{\mathrm{L}}[\mathrm{L}]
\end{eqnarray}
\begin{eqnarray}\label{eq:Pf}
\hspace*{-0.2cm} \dot{\![\mathrm{R}]} &=& \overline{D}_\mathrm{R}\overline{\Delta} [\mathrm{P}] + \overline{\rho}_\mathrm{R}+ (\overline{v} -m \overline{\delta}_\mathrm{C})\overline{\Gamma}[\mathrm{R}]^m[\mathrm{L}]^n - \overline{\delta}_\mathrm{R}[\mathrm{ R}]
\end{eqnarray}
We note that the linear response of the receptor production rate to receptor-ligand signaling helps to increase the size of the Turing space. Based on the results in Fig.\ S1, case U6, we expect that a saturation of the response for higher ligand-receptor concentrations, as could be described by a Hill function of the form $\mu([\mathrm{C}]) =H(\mu([\mathrm{C}], K)= H(\mathrm{\overline{\Gamma} [\mathrm{R}]^n[\mathrm{L}]^m},K)$, would cause a shrinking of the Turing space.
Equations \eqref{eq:Sf}-\eqref{eq:Pf} converge to the classical Schnakenberg equations for the following conditions:
\begin{itemize}
\item Receptor-independent degradation of ligand is much less efficient than receptor-dependent ligand degradation, as is generally the case, i.e.\ $\overline{\delta}_{\mathrm{L}}[\mathrm{L}] \ll n\overline{\delta}_{\mathrm{C}}\overline{\Gamma}\mathrm{R}]^m[\mathrm{L}]^n$.
\item The stochiometry of the ligand-receptor interaction yields $m$=2, $n$=1; we note that other stochiometries also yield Turing patters \cite{Menshykau:2012kg}.
\item $\overline{v} = (m+ n) \overline{\delta}_\mathrm{C}$.
\end{itemize}
\subsection{Derivation of the Non-Dimensional Set of Equations for the Receptor-Ligand based Turing Mechanism}
In the following, we will adopt the standard notation that is used to describe Turing models, and we write $U$ for the receptor concentration and $V$ for the ligand concentration; $U^mV^n$ represents the quasi-steady state concentration of the receptor-ligand complex. We have previously shown that a wide range of stochiometries can yield Turing patters \cite{Menshykau:2012kg}. Using $m=2$, $n=1$, i.e.\ one ligand dimer $V$ binds to two monomeric receptors $U$, equations \eqref{eq:Sf}-\eqref{eq:Pf} can be written as
\begin{eqnarray}
\frac{\partial U}{\partial\tau} & =& D_\mathrm{U}\Delta U +k_1-k_2 U +(k_5-2 k_3) U^2V \label{Schnakenberg_U_dim} \\[2mm]
\frac{\partial V}{\partial\tau} & = & D_\mathrm{V}\Delta V +k_4 -k_6 V - k_3 U^2V, \label{Schnakenberg_V_dim}
\end{eqnarray}
where $U=U(\tau,{\bf X})$ and $V=V(\tau,{\bf X})$ are the unknown functions depending on the time variable $\tau$ and space variable ${\bf X}$. The coefficient $k_1$ then represents the constitutive receptor production rate, while $k_4$ represents the constitutive ligand production rate. The term $-k_2 U$ reflects the ligand-independent receptor turn-over rate while $-k_6 V$ reflects the receptor-independent ligand turn-over rate. $-k_3 U^2V$ reflects the turn-over of the receptor-ligand complex, which leads to the removal of one ligand dimer, $V$, and two receptor monomers, $U$. Most ligand is typically removed by this receptor-dependent process, and we can therefore make the simplifying approximation $k_6=0$. Finally, $+k_5 U^2V$ reflects the signalling-dependent increase in receptor emergence (which can happen by a wide range of mechanisms); we will set $k_5=3 k_3$ in the following to recover the classical Schnakenberg equations. Equations \eqref{Schnakenberg_U_dim}-\eqref{Schnakenberg_V_dim} then read
\begin{eqnarray}
\frac{\partial U}{\partial\tau}& = &D_\mathrm{U}\Delta U +k_1-k_2 U + k_3 U^2V\\[2mm]
\frac{\partial V}{\partial\tau}& = & D_\mathrm{V}\Delta V +k_4 - k_3 U^2V.
\end{eqnarray}
These equations can be rewritten in dimensionless form as
\begin{equation}\label{eq:Schnakenberg}
\begin{split}
\frac{\partial u}{\partial t}&=\Delta u +\gamma(a-u+u^2v)\\[2mm]
\frac{\partial v}{\partial t}&=d\Delta v +\gamma(b -u^2v),
\end{split}
\end{equation}
where
\begin{gather*}
u = U\left(\frac{k_3}{k_2}\right)^{1/2},\ v = V\left(\frac{k_3}{k_2}\right)^{1/2},\ t = \frac{D_\mathrm{U}\,\tau}{L^2},\ {\bf x} = \frac{{\bf X}}{L},\\ d=\frac{D_\mathrm{V}}{D_\mathrm{U}},\, a = \frac{k_1}{k_2} \left(\frac{k_3}{k_2}\right)^{1/2},\, b = \frac{k_4}{k_2}\left(\frac{k_3}{k_2}\right)^{1/2},\, \gamma = \frac{L^2 k_2}{D_\mathrm{U}}.
\end{gather*}
The function $u$ then represents the receptor, $v$ represents the ligand, and $u^2v$ represents the quasi-steady state concentration of the receptor-ligand complex. As before, one ligand dimer $v$ binds to two monomeric receptors $u$. We have previously shown that also other combinations $u^m v^n$ result in Turing patterns \cite{Menshykau:2012kg}. The constant $\gamma a$ then represents the constitutive receptor production rate, while $\gamma b$ represents the constitutive ligand production rate. The term $-\gamma u$ reflects the ligand-independent receptor turn-over rate, while $-\gamma u^2v$ reflects the receptor-dependent ligand removal rate. Finally, $+\gamma u^2v$ represents the net result of ligand-dependent receptor turn-over and the signalling-dependent increase in receptor emergence, where the latter dominates, thus the positive term.
\section*{Appendix II: Determination of Turing Spaces}
\subsection{The Turing Mechanism}
\label{sec:Tpf}
In this section we summarize briefly the criteria for the emergence of Turing pattern for reaction-diffusion systems with two species. We consider systems of the form
\begin{equation}
\label{eq:generalReDiffEq}
\begin{split}
\frac{\partial U}{\partial \tau} & = F(U, V) + D_U\Delta U \\[2mm]
\frac{\partial V}{\partial \tau} & = G(U, V) + D_V\Delta V
\end{split}
\end{equation}
defined on $(0,\infty)\times \Omega$ (with a given spatial domain $\Omega\subset\mathbb{R}^n$) subject to boundary and initial conditions, where the space and time-dependent functions $U$ and $V$ represent concentrations and the reaction kinetic terms $F$ and $G$ are generally nonlinear functions. After suitable changes of variables and nondimensionalization Eq.\ \eqref{eq:generalReDiffEq} can be transformed into the dimensionless system
\begin{equation}
\label{eq:nondimReDiffEq}
\begin{split}
u_t & = \gamma f(u, v) + \Delta u \\
v_t & = \gamma g(u, v) + d\Delta v,\\
\end{split}
\end{equation}
where $t$ is the rescaled time variable, $d$ denotes (or is proportional to) the quotient of the diffusion coefficients $D_U$ and $D_V$ and $\gamma = \textit{const} \cdot L^2$, where $L$ is a typical length scale of the domain. To ensure the uniqueness of the solution we endow system \eqref{eq:nondimReDiffEq} with initial and boundary conditions. We will use homogeneous Neumann boundary condition of the form
\begin{equation*}
\begin{split}
({\bf n}\cdot\nabla)\begin{pmatrix}u\\ v\end{pmatrix} = 0 \quad \mbox{on } [0,\infty)\times\partial\Omega \\
\qquad u(0,{\bf x}) = u_0({\bf x}),\ v(0,{\bf x}) = v_0({\bf x}),
\end{split}
\end{equation*}
because they are easy to handle and have a biological interpretation (impermeable boundary). We note, however, that other boundary conditions would not greatly alter the following analysis. A Turing instability appears when a reaction-diffusion system has a stable steady state in the absence of diffusion, which loses its stability in the presence of diffusion such that spatial patterns emerge.
\subsection{Linear stability in the absence of diffusion}
Let $u_0$ and $v_0$ denote the steady state of the diffusion-free system of ordinary differential equations (ODEs)
\begin{equation}
\label{eq:generalODE}
u_t = \gamma f(u,v),\qquad v_t = \gamma g(u,v),
\end{equation}
and linearize the system about $(u_0,v_0)$ by introducing the translated function ${\bf w} = (w_1, w_2)^T$ with $w_1 = u-u_0,\ w_2 = v-v_0$. Then the linearized system becomes
\[
{\bf w}_t = \gamma J{\bf w},
\]
where
\[
J = \begin{pmatrix} f_u & f_v\\ g_u & g_v\end{pmatrix}\Bigg|_{(u_0,v_0)} = \begin{pmatrix} f_u(u_0,v_0) & f_v(u_0,v_0)\\ g_u(u_0,v_0) & g_v(u_0,v_0)\end{pmatrix}
\]
is the Jacobian evaluated at the point $(u_0,v_0)$. From now on, we write the partial derivatives evaluated at the steady state without their arguments for brevity. The steady state of the linearized system is stable, i.e.\ the steady state of system \eqref{eq:generalODE} is linearly stable if $\Re \lambda(J) < 0$ for all eigenvalues of $J$ (see any textbook on ODEs), which for a 2-component system is ensured by the conditions
\begin{equation}
\label{eq:linstab}
\textit{tr} J = f_u + g_v < 0,\qquad \textit{det}(J) = f_ug_v - f_vg_u > 0.
\end{equation}
\subsection{Diffusion-driven instability}
\label{subsec:diffdrivinstab}
Now let us add diffusion to our system of ODEs and consider the reaction-diffusion system linearized about the steady state ${\bf w} = (0,0)^T$, which has the form
\begin{equation}
\label{eq:linwdiff}
{\bf w}_t = \gamma J{\bf w} + D\Delta {\bf w},
\end{equation}
where $D = \textit{diag}(1,d)$ is a diagonal matrix containing the diffusion coefficients of the nondimensionalized system \eqref{eq:nondimReDiffEq}. We look for a solution of the form
\begin{equation}
\label{eq:w}
{\bf w}(t, {\bf x}) = \sum_k {\bf C}_k \textup{e}^{\lambda_k t}{\bf W}_k({\bf x}),
\end{equation}
where the exponents $\lambda_k$ determine the temporal growth of the solution and the time-in\-de\-pen\-dent functions ${\bf W}_k$ are the solutions of the elliptic eigenvalue problem
\[
\Delta {\bf W}_k + k^2{\bf W}_k = 0,\qquad ({\bf n}\cdot\nabla){\bf W}_k = 0.
\]
For instance, in one dimension on the interval $[0,L]$ the eigenvalues are $k = n\pi/L\ (n=0,1,2,\ldots)$, also called wavenumbers, and the eigenfunctions are $W(x) = \cos(n\pi x/L) = \cos(kx)$. The constants ${\bf C}_k = (C^{(1)}_k,C^{(2)}_k)^T$ are the Fourier-coefficients of the initial conditions.
Inserting equation \eqref{eq:w} into equation \eqref{eq:linwdiff} and using the fact that the set of eigenfunctions of the Laplace operator $\{{\bf W}_k\}$ forms a complete orthonormal system, we obtain as linearized system
\begin{equation}
\label{eq:linwdiff}
{\bf w}_t = \gamma J{\bf w} + D k^2 \bf{w}
\end{equation}
for each wavenumber $k$. Writing
\[
\det(\lambda I - \gamma J + k^2 D) = 0,
\]
where $I=I_2$ is the $2$-by-$2$ identity matrix, we obtain the eigenvalues $\lambda = \lambda_k$ of the matrix $M=\gamma J - k^2 D$. Expanding the above determinant, we obtain that $\lambda_k$ is the root of the second order polynomial equation
\begin{widetext}
\[
\lambda^2 + \lambda\big(k^2(1+d) - \gamma(f_u+g_v)\big) + dk^4 - \gamma(df_u + g_v)k^2 + \gamma^2(f_ug_v-f_vg_u)=0.
\]
\end{widetext}
Since we look for unstable solutions, we require that $\Re \lambda_k > 0$ for some $k\ne 0$. This means that either the coefficient of $\lambda$ and/or the constant term must be negative. Since the steady state is required to be linearly stable in the absence of diffusion (which corresponds to the case $k=0$), we must have $k^2(1+d) - \gamma(f_u+g_v) > 0$. Hence, to obtain a $\lambda$ with positive real part in the presence of diffusion we require
\[
h(k^2):= dk^4 - \gamma(df_u + g_v)k^2 + \gamma^2(f_ug_v-f_vg_u) < 0
\]
for some nonzero wavenumber $k$. Since we require $f_ug_v-f_vg_u>0$ for linear stability in the absence of diffusion ($k=0$) \eqref{eq:linstab}, it follows that $df_u + g_v > 0$ must hold. This condition is not sufficient to ensure the negativity of the function $h$; an elementary calculation shows that the minimum of $h$ is attained at the point
\[
k^2_{\textup{m}} = \gamma\frac{df_u+g_v}{2d},
\]
and the minimum value of $h$ is
\[
h_{\textup{min}} = h(k_{\textup{m}}^2) = \gamma^2\left[(f_ug_v-f_vg_u) - \frac{(df_u+g_v)^2}{4d}\right],
\]
which is negative if the expression in the bracket is negative.
\bigskip
In summary, the well-known conditions (see \cite[Sec.\ 2.3]{MurrayBook}) for which a reaction-diffusion system with two species exhibits a Turing instability are as follows:
\begin{equation}
\label{eq:Turingcond}
\begin{split}
f_u + g_v < 0,\quad f_ug_v-f_vg_u > 0,\\
df_u + g_v >0,\quad (df_u+g_v)^2 - 4d(f_u+g_v-f_vg_u) > 0,
\end{split}
\end{equation}
where all partial derivatives are evaluated at the steady state $(u_0,v_0)$. We note that it is possible that these conditions are satisfied, but that no pattern emerges. This is the case when $h$ is not negative for any $k$ within the discrete set of wavenumbers, and only takes a negative value in between two of these discrete wavenumbers. The distance between wavenumbers shrinks as $\gamma$ is increased, and in the limit of infinite $\gamma$ the spectrum of $k$ is continuous. Since $\gamma$ is related to the size of the spatial domain, it follows that on small domains pattern formation may not happen, while on a sufficiently increased domain patterns may be observed.
\subsection{Turing instability in interacting systems}
\label{sec:interacting}
We now consider two identical reaction-diffusion systems, which we couple with each other in several ways. When the couplings are of the same type (i.e. when the first 2-component Turing system based on $u$ and $v$ is coupled with the second Turing system that is based on $\widetilde u$ and $\widetilde v$ via the same functions $f$ and $g$), then we can derive exact conditions for the Turing instability, as an extension of the classical results that were presented in Section \ref{sec:Tpf} (see \cite[Sec.\ 2.3]{MurrayBook} for more details). For this let us consider systems of the form
\begin{equation}
\label{eq:coupledsys}
\begin{split}
u_t & = \gamma f(u,v,\widetilde u, \widetilde v) + \Delta u \\
v_t & = \gamma g(u,v,\widetilde u, \widetilde v) + d\Delta v \\
\widetilde u_t & = \gamma f(\widetilde u, \widetilde v, u,v) + \Delta \widetilde u \\
\widetilde v_t & = \gamma g(\widetilde u, \widetilde v, u, v) + d\Delta \widetilde v,
\end{split}
\end{equation}
where the functions $f$ and $g$ describe the chemical reactions, $\gamma>0$ is a constant depending on the size of the domain and $d>0$ is a diffusion parameter. Let $(u_0,v_0,\widetilde u_0,\widetilde v_0)$ denote the steady state (assuming that there is only one, or at least they are isolated) of this system in the absence of diffusion (note that due to the symmetry $u_0 = \widetilde u_0$ and $v_0 = \widetilde v_0$) and -- just as in the uncoupled case \eqref{eq:generalODE} -- linearize the system about the steady state. The linearized system has the form
\[
{\bf w}_t = \gamma J{\bf w},
\]
where
\[
J = \begin{pmatrix}f_u & f_v & f_{\widetilde u} & f_{\widetilde v} \\ g_u & g_v & g_{\widetilde u} & g_{\widetilde v} \\ f_{\widetilde u} & f_{\widetilde v} & f_u & f_v \\ g_{\widetilde u} & g_{\widetilde v} & g_u & g_v \end{pmatrix}
\]
is the Jacobian matrix. Note the symmetry in $J$ that arises for this particular coupling. In this linearized system the steady state is stable if $\Re \lambda(J) < 0$ for all eigenvalues of $J$. The eigenvalues are the roots of the characteristic polynomial $k_J(\lambda)$ of $J$, which is now a fourth order polynomial for the coupled system. Due to the very special form of the coupling and the resulting symmetries in $J$, the polynomial $k_J$ can be factorized as
\begin{widetext}
\begin{equation}
\label{eq:kJ}
\begin{split}
k_J(\lambda) = & \left[\lambda^2 + \lambda(-f_u-g_v-f_{\widetilde u}-g_{\widetilde v}) + f_ug_v-f_vg_u + f_ug_{\widetilde v} - f_{\widetilde v}g_u + f_{\widetilde u}g_v - f_vg_{\widetilde u} + f_{\widetilde u}g_{\widetilde v} - f_{\widetilde v}g_{\widetilde u}\right] \times \\
& \left[\lambda^2 + \lambda(-f_u-g_v+f_{\widetilde u}+g_{\widetilde v}) + f_ug_v-f_vg_u - f_ug_{\widetilde v} + f_{\widetilde v}g_u - f_{\widetilde u}g_v + f_vg_{\widetilde u} + f_{\widetilde u}g_{\widetilde v} - f_{\widetilde v}g_{\widetilde u} \right].
\end{split}
\end{equation}
\end{widetext}
Hence $\Re \lambda(J) <0$ holds for all the four eigenvalues of $J$, that is the steady state of \eqref{eq:coupledsys} is linearly stable if both of the factors in \eqref{eq:kJ} have only roots with negative real part, i.e.\
\begin{widetext}
\begin{equation}
\label{eq:coupledlinstab}
f_u + g_v < \pm (f_{\widetilde u} + g_{\widetilde v}),\qquad f_ug_v - f_vg_u + f_{\widetilde u}g_{\widetilde v} - f_{\widetilde v}g_{\widetilde u} > \pm(f_ug_{\widetilde v} - f_{\widetilde v}g_u + f_{\widetilde u}g_v - f_vg_{\widetilde u}).
\end{equation}
\end{widetext}
Following the course of the uncoupled case, by adding diffusion, we again arrive at Eq.\ \eqref{eq:linwdiff}, now with the diffusion matrix $D = \textit{diag}(1,d,1,d)$. As before, we look for a solution of the form of Eq.\ \eqref{eq:w}. To this end, we determine the eigenvalues $\lambda = \lambda_k$ for $M=\gamma J -k^2 D$. The characteristic polynomial of this matrix -- given the special forms of $J$ and $D$ -- can be factorized as the product of two second order polynomials:
\begin{widetext}
\begin{equation}
\label{eq:kM}
\begin{split}
k_M(\lambda) = & \left[\lambda^2 + \lambda\left(k^2(1+d) - \gamma(f_u+g_v+f_{\widetilde u}+g_{\widetilde v})\right) + dk^4 - \gamma k^2(df_u + g_v + df_{\widetilde u}+ g_{\widetilde v}) + \right. \\ & \qquad \quad\left. \gamma^2(f_ug_v-f_vg_u + f_ug_{\widetilde v} - f_{\widetilde v}g_u + f_{\widetilde u}g_v - f_vg_{\widetilde u} + f_{\widetilde u}g_{\widetilde v} - f_{\widetilde v}g_{\widetilde u})\right] \times \\
& \left[\lambda^2 + \lambda\left(k^2(1+d) - \gamma(f_u+g_v - f_{\widetilde u}-g_{\widetilde v})\right) + dk^4 - \gamma k^2(df_u + g_v - df_{\widetilde u} - g_{\widetilde v}) +\right. \\ & \qquad \quad\left. \gamma^2(f_ug_v-f_vg_u - f_ug_{\widetilde v} + f_{\widetilde v}g_u - f_{\widetilde u}g_v + f_vg_{\widetilde u} + f_{\widetilde u}g_{\widetilde v} - f_{\widetilde v}g_{\widetilde u})\right].
\end{split}
\end{equation}
\end{widetext}
To obtain a Turing instability, at least one of the roots of $k_M$ has to have positive real part for some $k\ne 0$, i.e.\ one of the factors of $k_M$ must have a root with $\Re \lambda(M)>0$. The first factor of \eqref{eq:kM} has a root with positive real part if the coefficient of $\lambda$ is negative, or the constant term is negative. But since the steady state is stable in the absence of diffusion (linear stability conditions \eqref{eq:coupledlinstab}) the coefficient of $\lambda$ is always positive, i.e.\ $k^2(1+d) - \gamma(f_u+g_v+f_{\widetilde u}+g_{\widetilde v}) > 0$. Hence we require that
\begin{widetext}
\[
h^{(1)}(k^2):= dk^4 - \gamma k^2(df_u + g_v + df_{\widetilde u}+ g_{\widetilde v}) + \gamma^2(f_ug_v-f_vg_u + f_ug_{\widetilde v} - f_{\widetilde v}g_u + f_{\widetilde u}g_v - f_vg_{\widetilde u} + f_{\widetilde u}g_{\widetilde v} - f_{\widetilde v}g_{\widetilde u}) < 0
\]
\end{widetext}
holds for some wavenumber $k\ne 0$. Since we know from the linear stability conditions \eqref{eq:coupledlinstab} that the constant term is positive, i.e.\ $f_ug_v-f_vg_u + f_ug_{\widetilde v} - f_{\widetilde v}g_u + f_{\widetilde u}g_v - f_vg_{\widetilde u} + f_{\widetilde u}g_{\widetilde v} - f_{\widetilde v}g_{\widetilde u} > 0$, it follows that $df_u + g_v + df_{\widetilde u}+ g_{\widetilde v}>0$ must hold. We further need to ensure that the function $h^{(1)}$ attains a negative value for some of the wave numbers. The minimum of $h^{(1)}$ is attained at
\[
k^2_{1,\textup{m}} = \gamma\frac{df_u + g_v + df_{\widetilde u}+ g_{\widetilde v}}{2d},
\]
and the minimum value of $h^{(1)}$ is
\begin{widetext}
\[
h^{(1)}_{\textup{min}} = h^{(1)}(k_{1,\textup{m}}^2) = \gamma^2\left[(f_ug_v-f_vg_u + f_ug_{\widetilde v} - f_{\widetilde v}g_u + f_{\widetilde u}g_v - f_vg_{\widetilde u} + f_{\widetilde u}g_{\widetilde v} - f_{\widetilde v}g_{\widetilde u}) - \frac{(df_u + g_v + df_{\widetilde u}+ g_{\widetilde v})^2}{4d}\right].
\]
\end{widetext}
The minimum value of $h^{(1)}$ is thus negative if the expression in the bracket is negative. If the first factor of \eqref{eq:kM} does not have roots with positive real part, the second factor has to have at least one root with positive real part to obtain a Turing instability. By similar reasoning as before we know from \eqref{eq:coupledlinstab} that the coefficient of $\lambda$ is again always positive: $k^2(1+d) - \gamma(f_u+g_v - f_{\widetilde u}-g_{\widetilde v}) > 0$. Hence, it is required that
\begin{widetext}
\[
h^{(2)}(k^2):= dk^4 - \gamma k^2(df_u + g_v - df_{\widetilde u} - g_{\widetilde v}) + \gamma^2(f_ug_v-f_vg_u - f_ug_{\widetilde v} + f_{\widetilde v}g_u - f_{\widetilde u}g_v + f_vg_{\widetilde u} + f_{\widetilde u}g_{\widetilde v} - f_{\widetilde v}g_{\widetilde u})<0
\]
\end{widetext}
holds for some $k\ne 0$. A necessary condition for this is $df_u + g_v - df_{\widetilde u} - g_{\widetilde v} >0$, since the constant term in $h^{(2)}$ is positive again by \eqref{eq:coupledlinstab}. To obtain a sufficient condition we have to calculate the minimum of $h^{(2)}$ as before, i.e.\
\[
k^2_{2,\textup{m}} = \gamma\frac{df_u + g_v - df_{\widetilde u} - g_{\widetilde v}}{2d}.
\]
The minimum value of $h^{(2)}$ is
\begin{widetext}
\[
h^{(2)}_{\textup{min}} = h^{(2)}(k_{2,\textup{m}}^2) = \gamma^2\left[(f_ug_v-f_vg_u - f_ug_{\widetilde v} + f_{\widetilde v}g_u - f_{\widetilde u}g_v + f_vg_{\widetilde u} + f_{\widetilde u}g_{\widetilde v} - f_{\widetilde v}g_{\widetilde u}) - \frac{(df_u + g_v - df_{\widetilde u} - g_{\widetilde v})^2}{4d}\right].
\]
\end{widetext}
In summary, the steady state has to be linearly stable if no diffusion is present, which means that all roots of \eqref{eq:kJ} have negative real part, but instability appears when diffusion is added, which means that the polynomial in \eqref{eq:kM} has to have at least one root with positive real part. Hence for Turing instability in the coupled system \eqref{eq:coupledsys} one of the following sets of conditions has to be satisfied (\eqref{eq:Turingcondsysa} \emph{or} \eqref{eq:Turingcondsysb}):
\begin{widetext}
\begin{subequations}
\label{eq:Turingcondsys}
\begin{equation}
\label{eq:Turingcondsysa}
\begin{split}
f_u + g_v < \pm (f_{\widetilde u} + g_{\widetilde v}),\qquad f_ug_v - f_vg_u + f_{\widetilde u}g_{\widetilde v} - f_{\widetilde v}g_{\widetilde u} > \pm(f_ug_{\widetilde v} - f_{\widetilde v}g_u + f_{\widetilde u}g_v - f_vg_{\widetilde u}),\\
df_u + g_v + df_{\widetilde u}+ g_{\widetilde v}>0,\\
(df_u + g_v + df_{\widetilde u}+ g_{\widetilde v})^2 - 4d(f_ug_v-f_vg_u + f_ug_{\widetilde v} - f_{\widetilde v}g_u + f_{\widetilde u}g_v - f_vg_{\widetilde u} + f_{\widetilde u}g_{\widetilde v} - f_{\widetilde v}g_{\widetilde u}) > 0;
\end{split}
\end{equation}
\vspace{5mm}
\begin{equation}
\label{eq:Turingcondsysb}
\begin{split}
f_u + g_v < \pm (f_{\widetilde u} + g_{\widetilde v}),\qquad f_ug_v - f_vg_u + f_{\widetilde u}g_{\widetilde v} - f_{\widetilde v}g_{\widetilde u} > \pm(f_ug_{\widetilde v} - f_{\widetilde v}g_u + f_{\widetilde u}g_v - f_vg_{\widetilde u}),\\
df_u + g_v - df_{\widetilde u} - g_{\widetilde v}>0,\\
(df_u + g_v - df_{\widetilde u} - g_{\widetilde v})^2 -4d(f_ug_v-f_vg_u - f_ug_{\widetilde v} + f_{\widetilde v}g_u - f_{\widetilde u}g_v + f_vg_{\widetilde u} + f_{\widetilde u}g_{\widetilde v} - f_{\widetilde v}g_{\widetilde u}) >0,
\end{split}
\end{equation}
\end{subequations}
\end{widetext}
where the first line comes from the linear stability condition (hence they are the same in both cases) and the other two lines are derived from the diffusion-driven instability conditions.
\iffalse
\section{Receptor-ligand based Turing Mechanisms}
We have recently shown that the biochemical interactions in several receptor-ligand systems can result in Turing patterns \cite{Badugu:2012ho,Menshykau:2012kg,Tanaka:2013wt}, if the following three conditions are fulfilled:
\begin{enumerate}
\item The diffusion of ligands is substantially faster than that of receptors.
\item The interaction of receptor and ligand is cooperative.
\item Receptor-ligand binding results in an increase of the membrane concentration/density of the receptor (as a result of enhanced transcription, lower receptor turn-over, enhanced recycling to the membrane or similar).
\end{enumerate}
These conditions are met by many receptor-ligand systems, and if met, the biochemical interactions can be described by a variant of the activator-depleted substrate / Schnakenberg kinetics \cite{Gierer:1972vq, Schnakenberg:1979td}, which were first described by Gierer and Meinhardt in 1972, and which can result in Turing pattern.
\fi
\section*{Appendix III: Cellular Models}
Here we present the details of the implementation of the cellular models presented in Figure \ref{Fig2}. We consider 1D, 2D and 3D cellular models. In all cases we solved Eqs. \eqref{eq:1}-\eqref{eq:2}, but with some terms restricted to certain subdomains as specified below. All equations were solved on the same mesh.
\subsection*{1D Cellular Models}
We use a 1D domain, comprising $N$ subdomains of equal length (Fig.\ \ref{Fig2}b). On every subdomain the set of Eqs. \eqref{eq:1}-\eqref{eq:2} is solved. Ligand $L$ can diffuse freely in the entire domain, while receptor $R$ is restricted to each subdomain by no-flux boundary conditions. Ligand exchange between subdomains is obtained by enforcing continuous ligand profiles across the borders of the subdomains, i.e. by requiring that the ligand value $L$ on the right hand side boundary of subdomain $i$ is the same as the ligand value $L$ on the left hand side boundary of subdomain $i+1$.
\subsection*{2D Cellular Models}
We use a 2D square domain, containing $N\times N$ equal sized subdomains of square shape. The subdomains neither intersect nor overlap (Fig. \ref{Fig2}b). The following set of PDEs is defined on this 2D domain:
\begin{eqnarray}
\hspace*{-0.4cm} \frac{\partial R}{\partial t} &=& \Delta R+ \gamma (a-R+R^2 L) ~\text{on} ~ C\\ \label{eq:domainsC}
\hspace*{-0.4cm} \frac{\partial L}{\partial t} &=& d \Delta L + \gamma \begin{cases} (b-R^2 L) & ~\text{on} ~ C\\
0 & ~ \text{on} ~ EC \\\end{cases} \label{eq:domainsEC}
\end{eqnarray}
where $C$ represents the $N\times N$ array of rectangular cellular subdomains, and EC refers to the rest of the 2D domain, representing the extracellular space.
\subsection*{3D Cellular Models}
We use a 3D domain (Fig. \ref{Fig2}b), containing $N\times N\times 1$ non-overlapping spheres that are embedded into a cuboid. The following set of PDEs describes the ligand and receptor dynamics on the surface of the spheres, referred to as $C$:
\begin{eqnarray}
\hspace*{-0.4cm} \frac{\partial R}{\partial t} &=& \Delta R+ \gamma (a-R+R^2 L) \mbox{ on } C\\ \label{eq:domainsC}
\hspace*{-0.4cm} \frac{\partial L}{\partial t} &=& d \Delta L+ \gamma(b-R^2 L) \mbox{ on } C
\end{eqnarray}
Additionally, ligand is free to diffuse in the bulk of cuboid, referred to as $EC$:
\begin{eqnarray}
\hspace*{-0.4cm} \frac{\partial L}{\partial t} &=& d \Delta L \mbox{ on } EC \label{eq:domainsEC}
\end{eqnarray}
The concentration of the ligand on the surface of the spheres and in the bulk of the cuboid is linked via
\begin{eqnarray}
\hspace*{-0.4cm} d \,\vec{n} \cdot \nabla L &=&\gamma(b-R^2 L)
\end{eqnarray}
where $\vec{n}$ is the outward normal vector. The volume inside the spheres (i.e. the cell interior) is not included in the simulations because we do not consider ligand or receptor internalization.
\section*{Appendix IV: Numerical Solution of PDEs with COMSOL}
The partial differential equations were solved in COMSOL Multiphysics 4.x as described previously \cite{Vollmer:CordConferenceProceedings:2013, Germann:ProceedingsOfComsolConference2011:2011, Menshykau:ProceedingsOfComsolConference2012:2012}. COMSOL Multiphysics has previously been used to accurately solve a variety of reaction-diffusion equations which originate from chemical, biological and engineering applications \cite{Menshykau2012, Menshykau:PhysicalBiology:2013, Kotha2014, Cutress2010, Sun2006, Seymen2014, Brian2012, Adivarahan2013}. In the following we present two tests for the numerical accuracy of the solution of Turing type models obtained with COMSOL Multiphysics.
\begin{figure}[t!]
\begin{center}
\includegraphics[width=\columnwidth]{Figure8.png}
\end{center}
\caption{
{\bf Comparison of the Turing spaces calculated numerically and those derived analytically.} (Color online) (A, B) The shaded regions of the parameter space indicates the area, where the linear stability analysis identifies a Turing instability (yellow, light shading) or other instabilities (navy, dark shading) for Eqs \ref{eq:1}, \ref{eq:2} with zero-flux boundary conditions. The symbols indicate the points in the parameter space where the numerical solution of Eqs \ref{eq:1}, \ref{eq:2} with zero-flux boundary conditions yielded either pattern formation (+) or not (0). $\gamma$ was chosen sufficiently large that Turing patterns could emerge on the 1D domain. Panel A and B differ in the relative diffusion coefficient $d$, with (A) $d=100$, and (B) $d=10$. }
\label{fig:NumTest1}
\end{figure}
\subsection{Accuracy of the Turing space}
We first test whether we obtain the same Turing space numerically and analytically. To this end, we use Eq. \ref{eq:Turingcond} as analytical condition for a Turing instability for the Turing model given by Eqs \ref{eq:1}, \ref{eq:2}. To estimate the size of the Turing space numerically, we solve Eqs \ref{eq:1}, \ref{eq:2} with COMSOL. Figure \ref{fig:NumTest1} shows that the numerical solution of Eqs \ref{eq:1}, \ref{eq:2} in COMSOL yields pattern (+ symbols) in the part of the parameter space where the analytical criterion specifies either the classical Turing space (yellow region) or an unstable steady state both in the presence and absence of diffusion (blue region).
\subsection{Convergence of Numerical Solution}
\begin{figure}[t!]
\begin{center}
\includegraphics[width=\columnwidth]{Figure9.pdf}
\end{center}
\caption{
{\bf Convergence of the Numerical Solution.} (Color online) (A) Typical pattern of receptor-ligand complexes ($R^2L$) on a domain comprising two subdomains. Ligand is produced in the upper domain, but free do diffuse on the entire domains. Receptor is produced in the lower domain and its diffusion is restricted to the lower domain. (B) The maximum deviation of the receptor-ligand complex ($R^2L$) as computed with an FEM mesh with element size equal to 0.01 from that computed at other mesh sizes.}
\label{fig:NumTest2}
\end{figure}
Here we show that the numerical solution of a ligand-receptor based Turing model on a domain comprising two layers
converges with respect to the mesh size. We consider the model
\begin{eqnarray}
\hspace*{-0.4cm} \frac{\partial R}{\partial t} &=& \Delta R+ \gamma (a-R+R^2 L) ~\text{on} ~T_1 \label{eq:domainsR} \\
\hspace*{-0.4cm} \frac{\partial L}{\partial t} &=& d \Delta L + \gamma \begin{cases} (-R^2 L) & \quad \text{on} ~T_1\\
b & \quad \text{on} ~T_2 \\\end{cases} \label{eq:domainsL}
\end{eqnarray}
where $T_1$ and $T_2$ indicate two different tissue layers. Figure \ref{fig:NumTest2}A shows the calculated distribution of the receptor-ligand complex ($R^2L$); similar patterns were obtained for a range of finite element meshes with the maximum size of the mesh size in the range from 0.01 to 0.1. Figure \ref{fig:NumTest2}B shows that the maximum deviation in the solution decreases quadratically with respect to the maximum mesh size or equivalently decreases linearly with respect to the maximum mesh edge, as expected for FEM with first order Lagrange elements. These tests support the previous observations by others that COMSOL Multyphysics can solve Turing-type equations accurately.
\bibliographystyle{plos2009.bst}
|
\section{Supplemental Material}
\begin{figure*}[t]
\begin{center}
\includegraphics[width=0.95 \textwidth]{SiVGP_supp_Fig1}
\label{fig:SiV_supp_Fig1}
\caption{(a) Reduced level diagram for the $\sigma_-\sigma_+$-excitation scheme. (b) Effective ground state evolution for the single-shot readout. Nine times as many plasmons are emitted in the spin-down state. (c) Hyperfine states for Si:Bi. There is an electron(nuclear) Zeeman energy $\omega_{0(n) }$ and a Hyperfine coupling between pairs of states in the submanifolds. For the protocols in Si:Bi we use the pseudo-spin-1/2 subspace of $\ket{g,\downarrow}=\ket{m_s=-1/2,m_I=-9/2}$ and $\ket{g,\uparrow}=\ket{+,-4}$. At large magnetic field these two states correspond to up/down electron spin and the fully polarized nuclear spin $m_I=-9/2$.}
\end{center}
\end{figure*}
\emph{Device Properties --}
The graphene could be encapsulated between two thin layers of hexagonal-BN (h-BN), while the doping of the nanoribbon is controlled by either direct chemical doping or a metal top gate. Room temperature mobilities approaching 10$^6$ cm$^2$/Vs have been demonstrated for bulk graphene surrounded by hBN \cite{Mayorov11,Elias11}, which would allow mid-infrared plasmons to coherently propagate for several microns \cite{Jablan09}. The $1s$ states of the donor have a six-fold valley degeneracy, which splits into a single orbital ground state $A_1$ and five excited states lying 10-30 meV higher in energy: a doublet $E$ and triplet $T_2$. Due to their large overlap with the donor nucleus they have a relatively large spin-orbit coupling [$\sim0.02(1)$ meV in Si:P(Bi)]. The excited state transitions from $1s(A_1)$ to $2p$ states lie in the mid-infrared between 30-70 meV. The $2p$ states primarily decay via phonons in a two step process through the $1s(T_2/E)$ valley states, but have long lifetimes: observed as long as 250 ps \cite{Vinh08} and predicted to be as long as 1.3 ns \cite{Tyuterev10}. The magnitude of the dipole moment for the $1s$ to $2p$ transition is $\abs{\bra{e_s}\mu_I\ket{g,s}} \approx 0.3~e\cdot$nm. In addition to the valley and spin fine structure, there is a hyperfine interaction with the donor nucleus, which has been used to achieve long term quantum memory in these systems \cite{Mohammady12,Wolfowicz13}; however, for all the donors except Bi (see below), the hyperfine interaction can be neglected at moderate magnetic fields $\gtrsim 10$ mT.
The spin-$3/2$ character of the $D^0X$ state has been measured and used for spin state preparation and detection in ensembles of Si:P and Si:Bi \cite{Yang06,Sekiguchi10}.
The $D^0X$ states lie $1.1$ eV above the $1s(A_1)$ ground state and have a lifetime of 272 ns, limited by Auger recombination. In principle, a lambda transition from the ground state to the to the 2$p$ states, through the $D^0X$ states, is allowed, but it is parity non-conserving, which is only weakly broken in this system. Instead we propose to hybridize the $2s$ states with the $2p$ states through application of an external stress, to make them degenerate, and an external electric field to mix them. This requires a pressure of $\sim 25$ MP along one of Si principal axes \cite{Ramdas81}. To apply suitable electric fields, one could either use an external gate or the nanoribbon itself, which will induce a 2s-2p splitting as large as $\approx 1$ meV for doping levels of $n_e\approx 10^{12}$ cm$^{-2}$.
\emph{Fidelity of Single-Shot Readout} --
To achieve single-shot readout we use the $\sigma_-\sigma_+$-excitation scheme discussed in the main text and shown in Fig.\ S1(a). The Hamiltonian for the Raman Rabi driving field takes the form
\begin{align}
H_{c}&=\Omega'(t)\bigg( \frac{1}{3} \ket{e_\uparrow}\bra{g,\uparrow} + \ket{e_\downarrow}\bra{g,\downarrow} + h.c. \bigg)
\end{align}
The factor of $1/3$ follows from the ratio of Clebsch-Gordon coefficients in the two excitation pathways through the spin-3/2 exciton state.
When the impurity is driven resonantly below saturation, the fluorescence distribution of the emitted plasmons follows a Poisson distribution \cite{Kimble76}, which is different for the two spin states
\begin{equation}
p_s(n,t)=\frac{(R_s t)^n}{n!} e^{-R t}
\end{equation}
where $R_\downarrow=\frac{2 \Omega'^2}{(\gamma+\gamma_p)^2} \gamma_p$ and $R_\uparrow = \frac{2 \Omega'^2}{9(\gamma+\gamma_p)^2} \gamma_p$. When the excited state emits a phonon instead of a photon it decays through the $1s(T_2/E)$ valley states where it can flip its spin due to the large spin orbit coupling in these states \cite{Ramdas81}. As a worst case scenario, we assume that each time a phonon decay occurs, there is an electron spin flip with probability 1/2. This leads to the evolution for the ground state spin populations $n_{\uparrow/\downarrow}$
\begin{equation}
\dot{n}_\uparrow = - \frac{\gamma}{2\gamma_p} R_\uparrow\, n_\uparrow + \frac{\gamma}{2 \gamma_p} R_\downarrow\, n_\downarrow
\end{equation}
and $n_\downarrow \approx 1- n_\uparrow$ since we are far below saturation. The dynamics are illustrated in Fig.\ S1(b) The emitted plasmon number distribution $P(n,t)$ is
\begin{align}
P(n,t)&=n_\uparrow(t) p_\uparrow(n,t)+n_\downarrow(t) p_\downarrow(n,t), \\
n_\uparrow(t)& = n_\uparrow(0) e^{- \gamma (R_\uparrow+R_\downarrow) t/\gamma_p}+ \frac{1}{10}\big[1-e^{- \gamma (R_\uparrow+R_\downarrow) t/\gamma_p} \big].
\end{align}
In Fig.\ 2(b) of the main text we show an example of the distribution for the two initial conditions $p_\uparrow=1$ and 0. The distribution for the two initial spin states will be disjoint when $\gamma_p/\gamma \gg R_\downarrow \, t \gg 1$ which guarantees that many plasmons are emitted before the spin is depolarized. The optimum occurs when $R_\downarrow \, t \sim \sqrt{2\gamma_p/\gamma}$.
\emph{Hyperfine Interaction in Si:Bi} --
The Hamiltonian for the ground state spin interacting with the donor nucleus is
\begin{equation}
H= \omega_0 s_z-\omega_n I_z + A\, \bm{s}\cdot \bm{I}
\end{equation}
where $I_z$ is the donor nuclear spin operator, $I=9/2$ in the case of Si:Bi, $\omega_{0(n)}$ is the Zeeman energy of the electron(nuclear) spin, and $A$ is the hyperfine coupling constant. For Si:Bi the hyperfine coupling is rather large $A/\hbar = 1.47$ GHz. To find the eigenstates, we first notice that $H$ has the two decoupled eigenstates $\ket{m_s=\pm1/2,m_I=\pm I}$, where $m=m_s+m_I$ is the total z-angular momentum. The other eigenstates are formed from superpositions of pairs of states $\ket{m_s= \pm1/2,m_I=m\mp 1/2}$ with $\abs{m}<I+1/2$ as shown in Fig.\ S1(c). We can define Pauli operators within each of these two-state submanifolds, then we can rewrite the Hamiltonian for these submanifolds as \cite{Mohammady12}
\begin{align}
H^m_{sub}& = \frac{\Delta_m}{2} \sigma_z+ \frac{\Omega_m}{2} \sigma_x- \epsilon_m \\
\Delta_m & = A\, m+ \omega_0+\omega_n \\
\Omega_m &=A \sqrt{I(I+1)-m^2 }\\
\epsilon_m &=\frac{A}{4}+\omega_n m
\end{align}
Defining $\theta_m = \tan^{-1}(\Omega_m/\Delta_m)$, the eigenstates are
\begin{equation}
\begin{split}
\ket{\pm, m} &= \cos(\theta_m/2) \ket{m_s=\pm\frac{1}{2},m_I=m\mp \frac{1}{2} } \\
&\pm \sin(\theta_m/2) \ket{m_s=\mp \frac{1}{2},m_I=m\pm \frac{1}{2} }
\end{split}
\end{equation}
To apply the results discussed in the main text to case of Si:Bi we work in the pseudo-spin 1/2 subspace with $\ket{g,\downarrow}=\ket{m_s=-1/2,m_I=-9/2}$ and $\ket{g,\uparrow}=\ket{+,-4}$. At large magnetic field these two states correspond to up/down electron spin and the fully polarized nuclear spin $m_z=-9/2$. The $\ket{g,\downarrow}$ state can be prepared by optically pumping with the $\sigma_-\pi$-excitation scheme. In addition, we note that both the two-qubit C-Phase gate and the entanglement preparation carry through unaltered in this pseudo-spin-1/2 subspace.
\emph{Master Equation --}
When the propagation time between impurities $r/v_g$ is much faster than the other time scales we can integrate out the plasmon modes in a Markov approximation and the super-radiance can be clearly seen in the master equation for the density matrix $\rho$ of the two impurities
\begin{align} \nonumber
\dot{\rho} &= -\frac{i \omega_{d}}{2} \comm{\sigma_1^z+\sigma_2^z}{\rho}- \frac{i}{2} v \comm{(\sigma_1^+\sigma_2^- +\sigma_1^-\sigma_2^+)}{\rho} \\ \label{eqn:me}
&- \gamma_c \mathcal{D}[\sigma_1^-+ \alpha \,\sigma_2^-]\rho - {\delta \gamma}( \mathcal{D}[\sigma_1^-] + \mathcal{D}[\sigma_2^-]) \rho \\
v &= \gamma_p\, \sin k r \, e^{-r/L_p}, ~~~~~~ \gamma_c = \gamma_p \,\abs{\cos kr} \, e^{-r/L_p} \\
\delta \gamma& = \gamma+\gamma_p - \gamma_c,~~~~ L_p^{-1} = \frac{e\, v_F^2}{2 \mu\,v_g \omega_F}
\end{align}
where $\mathcal{D}[c] \rho =1/2 \{c^\dagger c,\rho\}-c \rho c^\dagger$ and $\alpha= \textrm{sign} (\cos kr)$. We have included the losses in the plasmon propagation in terms of the plasmon propagation length $L_p$ defined in the main text.
\end{document}
|
\section{\label{sec:intro} Introduction}
In this paper we aim at determining $\Lambda_{\overline{\rm MS}}$ by comparing lattice and perturbative results for the quark-antiquark ($Q\bar Q$) static potential%
\footnote{In agreement with the prevalent notation, particularly in the field of lattice QCD, we use the terms {\it static potential} and {\it static energy} synonymously.
Note, however, that in the literature sometimes a distinction is made and these terms refer to different quantities.}
in momentum space.
More precisely, we restrict ourselves to quantum chromodynamics (QCD) with $n_f=2$ dynamical quark flavors, i.e.\ exclusively focus on $\Lambda_{\overline{\textrm{MS}}}^{(n_f=2)}$.
The $Q\bar Q$ static potential $V$ amounts to the interaction energy of the color-singlet state made up of a static quark $Q$ and its antiquark $\bar Q$ separated by a distance $r=|\vec{r}|$ (in position space), or characterized by a momentum transfer $p=|\vec{p}|$ (in momentum space), respectively.
Accordingly, we denote the potential in position space by $V(r)$ and in momentum space by $V(p)$.
While many studies aiming at the extraction of $\Lambda_{\overline{\rm MS}}$ from the $Q\bar Q$ static potential have been performed so far, e.g.\ \cite{Michael:1992nj,Gockeler:2005rv,Brambilla:2010pp,Leder:2010kz,Jansen:2011vv,Bazavov:2012ka,Fritzsch:2012wq,Bazavov:2014soa}, to the best of our knowledge we are the first to transform the lattice data to momentum space and compare and match perturbative and lattice results in momentum space.
As will be discussed in detail below, our present study is mainly motivated by the significantly worse convergence behavior of the perturbative potential in position space as compared to momentum space \cite{Aglietti:1995tg,Jezabek:1998pj,Jezabek:1998wk,Beneke:1998rk}.
Apart from that, many other approaches aim at determining $\Lambda_{\overline{\rm MS}}$ or alternatively the strong coupling $\alpha_s$ at a specific momentum scale, e.g.\ the $Z$-mass scale.\footnote{Exclusively focusing on QCD with just two dynamical quark flavors, in the framework of the present study we favor the specification of $\Lambda_{\overline{\rm MS}}$ rather than $\alpha_s(M_Z)$.
A reasonable extraction of $\alpha_s(M_Z)$ from an $n_f=2$ study would at least require the discussion of flavor thresholds and extrapolations from $n_f\to n_f+1$, and in a sense obscure the main aim of our study.}
For recent lattice studies and results, see e.g.\ the lattice computations \cite{DellaMorte:2004bc,Shintani:2008ga,Aoki:2009tf,Sternbeck:2010xu,McNeile:2010ji,Blossier:2010ky,Sternbeck:2012qs,Blossier:2013ioa,Cichy:2013eoa} using the Schr\"odinger functional, vacuum polarization functions, ghost and gluon propagators, heavy quark correlators and the Dirac operator spectrum. Other works \cite{Abbate:2010xh,Kneur:2011vi,Boito:2012cr,Gehrmann:2012sc,Abbas:2012fi,Kneur:2013coa,Alekhin:2013nda,Andreev:2014wwa} focus e.g.\ on $\tau$ decays and electron-positron as well as electron-proton collisions.
Higher-order perturbative calculations in QCD are most conveniently performed in momentum space. This is particularly true for the perturbative $Q\bar Q$ static potential.
Hence, the perturbative expression for the static potential at the highest current accuracy, which encompasses all contributions up to ${\cal O}(\alpha_s^4)$, is directly accessible in momentum space.
Due to the fact that QCD is asymptotically free, perturbative calculations in QCD are viable only at large momentum transfers $p\gg\Lambda_{\rm QCD}$,
with $\Lambda_{\rm QCD}$ denoting the {\it QCD (momentum) scale}, which can be seen as separating the regimes of perturbative and non-perturbative physics.
However, note that in standard perturbation theory loop diagrams come along with integrations of the loop four-momentum over the full momentum regime, implying that such loops naturally also receive contributions from momenta $\lesssim\Lambda_{\rm QCD}$ for which perturbation theory is no longer trustworthy.
The leading uncontrolled contribution $\delta V(p)$ contained in the perturbative potential in momentum space arising from this kind of diagrams is quadratic in $\Lambda_{\rm QCD}$ and scales as $\delta V(p)\sim-\frac{4\pi\alpha_s}{p^2}\bigl(\frac{\Lambda_{\rm QCD}}{p}\bigr)^2$ \cite{Beneke:1998rk}.
Taking into account that $V(p)\sim-\frac{4\pi\alpha_s}{p^2}\bigl(1+{\cal O}(\alpha_s)\bigr)$ and $p\gg\Lambda_{\rm QCD}$, i.e.\ $\frac{\Lambda_{\rm QCD}}{p}\ll1$, this only amounts to a tiny correction.
For completeness, note that the strong coupling $\alpha_s$ depends on an {\it a priory} arbitrarily chosen renormalization momentum scale $\mu$, i.e.\ $\alpha_s=\alpha_s(\mu)$.
The $\mu$ dependence is such that $\alpha_s(\mu)\ll1$ only for $\mu\gg\Lambda_{\rm QCD}$.
A particularly obvious choice of the renormalization scale for the couplings in $V(p)$ is $\mu=p$, corresponding to an identification of $\mu$ with the {\it typical momentum scale} of the quantity under consideration.
Conversely, within lattice QCD the $Q\bar Q$ static potential is naturally computed in position space.
It can be extracted straightforwardly by studying the exponential decay of the rectangular Wilson loop as a function of its temporal extension \cite{Brown:1979ya}.
Of course, lattice simulations at a given lattice spacing $a$ cannot resolve arbitrarily small separations.
Moreover, the minimum attainable lattice spacing is limited by the available computing power.
The behavior of the static potential at small $Q\bar Q$ separations is intimately related to its behavior at large momenta. Consequently, after a Fourier transform to momentum space, lattice results are expected to allow for reliable insights only below a certain threshold momentum,
which is $\lesssim$ the maximum momentum $\overline{p}=\pi/a$ that can be resolved on a lattice of spacing $a$ (cf. Sec.~\ref{SEC572} below).
Taking into account the above constraints, a comparison and matching of perturbative and lattice results in momentum space is limited to momenta $p$ fulfilling $\Lambda_{\rm QCD}\ll p\ll \overline{p}$.
Analogous considerations can be invoked to delimit the fitting interval in position space (cf.\ e.g.\ \cite{Jansen:2011vv}),
employing that the typical momentum scale that can be attributed to a relative distance $r$ scales as $p\sim1/r$. In position space the manifestly perturbative regime is thus characterized by $1/r\gg\Lambda_{\rm QCD}$.
However, an important difference is that the leading uncontrolled contribution $\delta V(r)$ to the static potential in position space as determined by a standard Fourier transform from momentum to position space is more pronounced than in momentum space: it is linear in $\Lambda_{\rm QCD}$, scales as $\delta V(r)\sim\frac{\alpha_s}{r}(r\Lambda_{\rm QCD})=\alpha_s\Lambda_{\rm QCD}$,
and arises exclusively from the low momentum part of the Fourier integral over momenta $\lesssim\Lambda_{\rm QCD}$, for which perturbation theory is no longer trustworthy \cite{Beneke:1998rk}%
\footnote{Recall that in the perturbative regime both dimensionless quantities, $\Lambda_{\rm QCD}/p$ and $r\Lambda_{\rm QCD}$, are small and of the same order of magnitude, i.e.\ $\Lambda_{\rm QCD}/p\sim r\Lambda_{\rm QCD}\sim\epsilon\ll1$.
Correspondingly, $r\Lambda_{\rm QCD}={\cal O}(\epsilon)$, while $\bigl(\frac{\Lambda_{\rm QCD}}{p}\bigr)^2={\cal O}(\epsilon^2)$.}.
For the position space potential a seemingly obvious choice of the renormalization scale is $\mu=1/r$, which amounts to the {\it typical momentum scale} to be associated with the static quarks separated by a distance $r$.
However, it has been recognized long ago that in particular for the identification $\mu=1/r$ the convergence of the perturbative potential in position space as defined by a standard Fourier transform from momentum space is spoiled \cite{Billoire:1979ih},
while it can be significantly improved by subtracting just the uncontrolled contribution $\delta V(r)$ linear in $\Lambda_{\rm QCD}$ \cite{Beneke:1998rk}.
Unfortunately this necessitates the specification of an additional {\it subtraction scale} and thereby increases the number of free parameters.
Without any subtraction procedure a meaningful fit of perturbative expressions for the $Q\bar Q$ static potential to lattice results in position space is not possible:
the perturbative expressions only show a controlled convergence behavior for separations, which are much smaller than the minimum accessible separations on state-of-the-art lattice simulations
(cf.\ e.g.\ \cite{Pineda:2001zq,Pineda:2002se,Laschka:2011zr,Karbstein:2013zxa}).
Correspondingly, aiming at an accurate determination of $\Lambda_{\overline{\rm MS}}$ by comparing the results from lattice QCD simulations with perturbative calculations of the $Q\bar Q$ static potential in position space,
considerable efforts are needed to cope with this issue. Various strategies to retain or restore the significantly better convergence of $V(p)$ also for $V(r)$ have been devised in the literature \cite{Beneke:1998rk,Pineda:2001zq,Pineda:2002se,Brambilla:2009bi,Brambilla:2010pp,Laschka:2011zr,Karbstein:2013zxa}.
An extraction of $\Lambda_{\overline{\rm MS}}$ directly in momentum space of course does not involve a Fourier transform of the perturbative potential, such that one may hope to circumvent most of these limitations.
On the other hand, one now has to Fourier transform the lattice potential, which might eventually lead to similar problems.
As we will argue and demonstrate in detail in this paper, most favorable for us the latter concerns are not substantiated.
Our paper is organized as follows. Section~\ref{SEC572} is devoted to the $Q\bar Q$ static potential on the lattice. After briefly reviewing its computation in position space, we point out how we transform it to momentum space.
In Sec.~\ref{sec:perturbation} we summarize the present knowledge of the perturbative $Q\bar Q$ static potential and detail on the role of $\Lambda_{\overline{\rm MS}}$.
Special emphasis is put on the convergence behavior of both the perturbative $Q\bar Q$ static potential for $n_f=2$ in momentum space and the QCD $\beta$-function.
Section~\ref{SEC600} constitutes the main section of our paper. Here we describe in detail our momentum space analysis to extract $\Lambda_{\overline{\textrm{MS}}}^{(n_f=2 )}$ by fitting the perturbative expressions for the $Q\bar Q$ static potential $V(p)$ to the corresponding lattice results. The various error sources are identified and delineated and our final result for $\Lambda_{\overline{\textrm{MS}}}^{(n_f=2 )}$ is specified. Moreover, comparisons with the result of \cite{Jansen:2011vv}, amounting to a position space extraction of $\Lambda_{\overline{\textrm{MS}}}^{(n_f=2 )}$ based on the same lattice data, are made. Finally, we end with conclusions in Sec.~\ref{sec:Concl}.
\newpage
\section{\label{SEC572}The $Q\bar Q$ static potential in momentum space from lattice QCD }
\subsection{\label{SEC852}Gauge link configurations}
We use the same $n_f = 2$ gauge link configurations as for a recent determination of $\Lambda_{\overline{\textrm{MS}}}$, where, in contrast to this work, the lattice results and perturbative expressions for the static potential were compared and matched in position space \cite{Jansen:2011vv}. These gauge link configurations were generated by the European Twisted Mass Collaboration (ETMC) \cite{Boucaud:2007uk,Boucaud:2008xu,Baron:2009wt} with the tree-level Symanzik improved gauge action \cite{Weisz:1982zw},
\begin{equation}
S_\mathrm{G}[U] = \frac{\beta}{6} \bigg(b_0 \sum_{x,\mu\neq\nu} \textrm{Tr}\Big(1 - P^{1 \times 1}(x;\mu,\nu)\Big) + b_1 \sum_{x,\mu\neq\nu} \textrm{Tr}\Big(1 - P^{1 \times 2}(x;\mu,\nu)\Big)\bigg)
\end{equation}
with $b_0 = 1 - 8 b_1$ and $b_1 = -1/12$ and the Wilson twisted mass quark action \cite{Frezzotti:2000nk,Frezzotti:2003ni,Frezzotti:2004wz,Shindler:2007vp},
\begin{equation}
\label{EQN963} S_\mathrm{F}[\chi,\bar{\chi},U] = a^4 \sum_x \bar{\chi}(x) \Big(D_{\rm W} + i \mu_\mathrm{q} \gamma_5 \tau_3\Big) \chi(x)
\end{equation}
with
\begin{equation}
D_\mathrm{W} = \frac{1}{2} \Big(\gamma_\mu \Big(\nabla_\mu + \nabla^\ast_\mu\Big) - a \nabla^\ast_\mu \nabla_\mu\Big) + m_0 .
\end{equation}
Here $a$ denotes the lattice spacing, $\nabla_\mu$ and $\nabla^\ast_\mu$ are the gauge covariant forward and backward derivatives, $m_0$ and $\mu_\mathrm{q}$ are the bare untwisted and twisted quark masses, $\tau_3$ is the third Pauli matrix acting in flavor space, and $\chi = (\chi^{(u)} , \chi^{(d)})$ represents the quark fields in the so-called twisted basis.
The twist angle $\omega$ is given by $\omega = \arctan(\mu_\mathrm{R} / m_\mathrm{R})$, where $\mu_\mathrm{R}$ and $m_\mathrm{R}$ denote the renormalized twisted and untwisted quark masses.
For the ensembles of gauge link configurations considered in the present study (cf.~Table~\ref{TAB077}) $\omega$ has been tuned to $\pi / 2$ by adjusting $m_0$ appropriately.
This ensures automatic $\mathcal{O}(a)$ improvement for many observables including the static potential (cf.\ \cite{Boucaud:2008xu} for details).
The considered gauge link configurations cover several different values of the lattice spacing, the pion mass $m_\textrm{PS}$ and the spacetime volume $L^3 \times T$; cf.~Table~\ref{TAB077}, which also provides the number of gauge link configurations, used for the computation of the static potential, for each ensemble. The lattice spacing in physical units has been set via the pion mass and the pion decay constant, using chiral perturbation theory. The resulting value for the
hadronic scale\footnote{The hadronic scale $r_0$ is defined via $r_0^2 F(r_0) = 1.65$, with $F(r) = {\rm d}V(r) / {\rm d}r$ \cite{Sommer:1993ce}.} $r_0$
is $r_0 = 0.420(14) \, \textrm{fm}$ (cf.\ Sec.~5 of \cite{Boucaud:2008xu} and Tab.~8 of \cite{Baron:2009wt}). For further details on the generation of these gauge field configurations as well as on the computation and the analysis of standard quantities (e.g.\ lattice spacing and pion mass) we refer the reader to \cite{Boucaud:2008xu,Baron:2009wt}.
\begin{table}[htb]
\begin{center}
\begin{tabular}{|c|c|c|c|c|c|}
\hline
& & & & & \vspace{-0.40cm} \\
$\beta$ & $a$ in $\textrm{fm}$ & $(L/a)^3 \times T/a$ & $m_\textrm{PS}$ in $\textrm{MeV}$ & $r_0 / a$ & \# gauges \\
& & & & & \vspace{-0.40cm} \\
\hline
& & & & & \vspace{-0.40cm} \\
\hline
& & & & & \vspace{-0.40cm} \\
$3.90$ & $0.079(3)\phantom{00}$ & $24^3 \times 48$ & $340(13)$ & $5.36(4)\phantom{0}$ & $168$ \\
& & & & & \vspace{-0.40cm} \\
\hline
& & & & & \vspace{-0.40cm} \\
$4.05$ & $0.063(2)\phantom{00}$ & $32^3 \times 64$ & $325(10)$ & $6.73(5)\phantom{0}$ & $\phantom{0}71$ \\
& & & & & \vspace{-0.40cm} \\
\hline
& & & & & \vspace{-0.40cm} \\
$4.20$ & $0.0514(8)\phantom{0}$ & $48^3 \times 96$ & $284(5)\phantom{0}$ & $8.36(6)\phantom{0}$ & $\phantom{0}46$ \\
& & & & & \vspace{-0.40cm} \\
\hline
& & & & & \vspace{-0.40cm} \\
$4.35$ & $0.0420(17)$ & $32^3 \times 64$ & $352(22)$ & $9.81(13)$ & $146$\vspace{-0.40cm} \\
& & & & & \\
\hline
\end{tabular}
\caption{\label{TAB077}Ensembles of gauge link configurations employed in the present study.}
\end{center}
\end{table}
\subsection{\label{EQN808}Computation of the $Q\bar Q$ static potential in position space}
First we determine the $Q\bar{Q}$ static potential $V(\vec{r})$ in position space. In a second step, $V(\vec{r})$ is transformed to momentum space by means of a discrete Fourier transform.
To be able to perform this Fourier transform numerically we need the static potential for all $\vec{r} = \vec{n} a$ inside a finite periodic spatial volume $L'^3$ of side length $L'$. This is achieved as follows: First, we compute $V(\vec{n} a)$ with $n_i\in\{0,1,\ldots,N'/2\}$, where $N'$ is even and defined as $N' \equiv L'/a$. Second, we realize the periodicity by defining $V(n_x a , n_y a , n_z a) \equiv V(|n_x| a , |n_y| a , |n_z| a)$, where now $n_i \in\{-\frac{N'}{2} + 1 , -\frac{N'}{2} + 2 , \ldots ,\frac{N'}{2}\}$.
To keep finite volume effects on a negligible level, the spatial volume $L'^3$ needs to be sufficiently large (cf. also point (2) below).
While a lattice computation of the static potential $V(\vec{n} a)$ for all $n_x , n_y , n_z = 0 , 1 , \ldots , N'/2$ is possible in principle, it is extremely computer time consuming in practice: One needs to generate gauge link configurations for such large volumes and has to compute both on- and off-axis Wilson loops for all possible quark-antiquark separations $\vec{r}=\vec{n} a$. On the other hand the shape of the static potential at large separations is known to be accurately described by $V(r) = A_0 + \sigma r + A_1/r$, where $A_0$ denotes a constant offset parameter, $\sigma$ is the string tension and $A_1 \gtrsim -\pi/12$ \cite{Luscher:1980fr,Luscher:1980ac,Donnellan:2010mx}.
Several clarifying comments are in order here.
Accounting for a nonvanishing number $n_f$ of light dynamical quark flavors,
there exists a certain threshold distance $r_c\gtrsim1\,{\rm fm}$ such that for separations $r>r_c$ of the static quark $Q$ and its antiquark $\bar Q$ the $Q\bar Q$ state is energetically disfavored in comparison to a pair of static-light mesons, $B\bar B$ \cite{Bali:2005fu}. This effect is known as string breaking.
The ground state of the system is made up of dynamical quarks, gluons and static quarks $Q$ and $\bar Q$ separated by a distance $r$.
It scales string-like $\sim \sigma r$ in the nonperturbative regime below $r_c$, but becomes completely independent of $r$ for $r>r_c$, and saturates at about twice the $B$ meson mass.
However, the $Q\bar Q$ state scaling $\sim \sigma r$ can still be traced for $r>r_c$ also, where it corresponds to an excited state:
Apart from resulting in a narrow mini-gap feature at $r\approx r_c$, i.e.\ where the energy of the $Q\bar Q$ state equals that of the $B\bar B$ state, mixing effects between the $Q$ and $B$ sectors are tiny.
Wilson loops as also employed here to extract the static potential on the lattice [cf. \Eqref{EQN600} below] are particularly insensitive to the $B$ sector.
On the other hand, standard perturbation theory for the $Q\bar Q$ static potential manifestly focuses on the $Q$ sector of the theory:
While it accounts for dynamical light quarks in loop diagrams, in this framework the virtual light quarks can never become real.
For these reasons, even though there occurs string breaking for $n_f\neq0$, when focusing on the $Q\bar Q$ potential $V(r)$
we manifestly limit ourselves to the $Q$ sector, i.e.\ focus on the $r$ dependent, linearly rising component of the potential also for $r>r_c$.
Moreover, the large distance behavior of $V(r)$ is expected to have a rather weak effect on the Fourier transformed potential for $p\gg\Lambda_{\rm QCD}$, i.e.\ the momentum
regime used for the matching to perturbation theory and the $\Lambda_{\overline{\textrm{MS}}}$ determination in this work.\footnote{To substantiate this rather vague statement given here, we have explicitly checked and confirmed that the large distance behavior of the lattice potential $V(r)$ has only a very mild influence of the value of $\Lambda_{\overline{\rm MS}}$ to be extracted from the $Q\bar Q$ static potential in momentum space, by modeling the long distance behavior of $V(r)$ with different functional forms (cf. point (2) below, and the numerical results in Sec.~\ref{input_var}, particularly Table~\ref{TAB015}).} Therefore, we stick to the following strategy:
\begin{itemize}
\item[(1)] \textbf{Perform a standard lattice computation of }$V(\vec{r})$\textbf{ for }$|\vec{r}|=|\vec{n}|a \leq r_\textrm{max}$\textbf{:} \\
For quark-antiquark separations $|\vec{n}| a \leq r_\textrm{max}$ we extract $V(\vec{r})$ from the exponential decay of Wilson loop averages $\langle W(\vec{r},t) \rangle$ with respect to their temporal extent $t$, while keeping their spatial extent $r$ fixed.
To this end we first compute
\begin{equation}
V^\textrm{(effective)}(\vec{r},t) = \frac{1}{a} \ln\bigg(\frac{\langle W(\vec{r},t) \rangle}{\langle W(\vec{r},t+a) \rangle}\bigg) .
\end{equation}
Somewhat arbitrarily, we choose $r_\textrm{max}\approx0.42\,{\rm fm}$ corresponding to $r_\textrm{max} = 10a$ for our smallest lattice spacing (cf. Table~\ref{TAB077}).
In a second step the $t$-independent quantity $V(\vec{r})$ is obtained by performing an uncorrelated $\chi^2$ minimizing fit to $V^\textrm{(effective)}(\vec{r},t)$ in a suitable $t$ range. This range is chosen such that excited states are strongly suppressed, while statistical errors are still small.
We use the ensembles listed in Table~\ref{TAB077} and consider on- and off-axis Wilson loops formed by APE smeared spatial links ($N_\textrm{APE} = 60$, $\alpha_\textrm{APE} = 0.5$ for all our gauge link ensembles) and ordinary, i.e .\ unsmeared temporal links. For a detailed explanation regarding the construction of off-axis Wilson loops cf.\ \cite{Jansen:2011vv}. For a definition of APE smearing we refer to \cite{Jansen:2008si}.
\item[(2)] \textbf{Model }$V(\vec{r})$\textbf{ for }$|\vec{r}|=|\vec{n}| a > r_\textrm{max}$\textbf{ and }$|r_i| = |n_i|a \leq L'/2$\textbf{:} \\
For quark-antiquark separations $|\vec{n}| a > r_\textrm{max}$ while $|n_i|a \leq L'/2$ we model the lattice potential by
\begin{equation}
\label{EQN600} V(\vec{r}) = V_M(r) \equiv A_0 + \sigma r + \sum_{m=1}^M \frac{A_m}{r^m} .
\end{equation}
For our finest lattice spacing $a \approx 0.0420 \, \textrm{fm}$ we use $L'/a = 256$, obviously fulfilling $L' \gg L$ (cf. Table~\ref{TAB077}).
To ensure that the extracted value of $\Lambda_{\overline{\rm MS}}$ is independent of the choice for $L'$, we also performed computations with $L'/a = 128$ and $L'/a = 512$ and found essentially identical results for $\Lambda_{\overline{\rm MS}}$
(the deviations are below $1\,{\rm MeV}$).
In Sec.~\ref{SEC600} different values of $M\in\{1,2,3,4\}$ are used to quantify systematic errors associated with this modeling of the long range part of the lattice potential. The string tension is fixed to $\sigma = 1550 \, \textrm{MeV}/\textrm{fm}$ (corresponding to $r_0 = 0.420 \, \textrm{fm}$ determined on the same gauge link configurations we are using throughout this work \cite{Baron:2009wt}). While $A_1 = -\pi/12$ in the bosonic string picture \cite{Luscher:1980fr,Luscher:1980ac}, lattice simulations with $n_f=2$ quark flavors yield a larger value $A_1 \approx -0.3 \ldots -0.5$ \cite{Donnellan:2010mx}. We determine $A_m$ with $m\in\{0,1,\ldots,M\}$ by a $\chi^2$ minimizing fit of \Eqref{EQN600} to the lattice results determined in step (1) in the region $r_\textrm{min} \leq r \leq r_\textrm{max}$. In order to have $\chi^2\lesssim1$, for our smallest lattice spacing $a = 0.0420 \, \textrm{fm}$ we choose $r_{\rm min}=8a$ for $M=1$, $r_{\rm min}=6a$ for $M=2$, and $r_\textrm{min} =4 a$ for $M =\{3,4\}$.
The resulting function $V_3(r)$ for $M=3$ is shown in Figure~\ref{FIG600} together with the lattice results for $V(\vec{r})$.
\end{itemize}
\begin{figure}[htb]
\begin{center}
\includegraphics[width=0.8\textwidth]{fit_results}
\caption{\label{FIG600}Position space potential $V_3(r)$ (green curve) obtained by a $\chi^2$ minimizing fit to the lattice results $V(\vec{r})$ (red dots) in the region $4 a \leq |\vec{r}| \leq 10 a$ with $a \approx 0.0420 \, \textrm{fm}$. Note that the statistical errors (actually also depicted here) are so tiny that the curve for $V_3(r)$ and its error band fall on top of each other and are indiscernible by eye.}
\end{center}
\end{figure}
\subsection{Computation of the $Q\bar Q$ static potential in momentum space}
We define the $Q\bar Q$ static potential in momentum space, $V(\vec{p})$ with $\vec{p} = 2 \pi \vec{k} / L'$, by the discrete Fourier transform of $V(\vec{r})$,
\begin{equation}
\label{EQN601} V(\vec{p}) = V(2 \pi \vec{k} / L') = \sum_{n_x , n_y , n_z} a^3 \exp\bigg(-\frac{2 \pi i \vec{k} \vec{n}}{N'}\bigg) V(\vec{n} a) ,
\end{equation}
where the sum is also over all possible values of $n_i\equiv-\frac{N'}{2}+\nu_i$ with $\nu_i\in\{1,2,\ldots,N'\}$, and $k_i\equiv -\frac{N'}{2} + \kappa_i$ with $\kappa_i\in\{1,2,\ldots,N'\}$.
In the continuum and in infinite volume $V(\vec{p})$ is rotationally symmetric, i.e.\ $V(\vec{p})=V(p)$.
Since both a cubic lattice discretization and a cubic periodic volume break rotational symmetry, this is no longer true for the lattice potential~\eqref{EQN601}.
The deviations from the rotationally invariant continuum and infinite volume potential are expected to be particularly small, when restricting the dimensionless lattice momenta $\vec k$ to values inside a cylinder of unit radius around the lattice diagonal, i.e.\ demanding
\begin{equation}
\vec{k}^2 - (\vec{k} \vec{d})^2 \leq 1\, , \quad \vec{d} = \frac{1}{\sqrt{3}} (1 , 1 , 1)\, .
\end{equation}
This so-called cylinder cut is frequently used in lattice momentum space computations of propagators. For details cf.\ \cite{Sternbeck:2006rd}.
The maximal momentum value along each of the three axes is $\overline{p} = \pi / a \approx 15 \, \textrm{GeV}$ for our smallest lattice spacing $a \approx 0.0420 \, \textrm{fm}$. Similar to position space, where the minimum on-axis separation is $a$ and the static potential is essentially free of discretization errors for $r \gtrsim 3 a$, one might expect rather small discretization errors for $p \lesssim \overline{p} / 3 \approx 5 \, \textrm{GeV}$. These expectations will be confirmed in Sec.~\ref{input_var} below, where in particular the results depicted in Figure~\ref{FIG003} (c) and (d) show that a variation of the maximum momentum employed in the extraction of $\Lambda_{\overline{\rm MS}}$ in a range $\lesssim3 \, \textrm{GeV}$
basically does not change the value of $\Lambda_{\overline{\rm MS}}$.
The final result $V(p)$ for our smallest lattice spacing $a \approx 0.0420 \, \textrm{fm}$, obtained with $V_3(r)$ [cf. \Eqref{EQN600} and Figure~\ref{FIG600}] is shown in Figure~\ref{FIG601}\footnote{Throughout this paper all computations are performed in units of the lattice spacing $a$, e.g.\ we work with quantities $aV(r)$, $V(p) / a^2$ and $a\Lambda_{\overline{\rm MS}}$, which are independent of any potential errors or uncertainties regarding scale setting; for a recent review cf.\ \cite{Sommer:2014mea}. Nevertheless, the axes in the presented plots as well as the numbers quoted in the main text often given in units of $\textrm{MeV}$ or $\textrm{fm}$. This is intended to make the physical scales more obvious. To this end we use $a = 0.0420 \, \textrm{fm}$ for our smallest lattice spacing corresponding to $\beta = 4.35$ (cf.\ also Table~\ref{TAB077}).}.
\begin{figure}[htb]
\begin{center}
\includegraphics[width=0.8\textwidth]{cylinder_data}
\caption{\label{FIG601}$V(p)$ for $a \approx 0.0420 \, \textrm{fm}$ obtained from $V_3(r)$ after applying the cylinder cut.}
\end{center}
\end{figure}
\newpage
\section{\label{sec:perturbation}Perturbative insights into the $Q\bar Q$ static potential}
\subsection{The $Q\bar Q$ static potential in perturbation theory}
To allow for easier reference and to keep this paper self-contained, we briefly summarize the present status of the $Q\bar Q$ static potential in perturbation theory.
Quantities which depend on the particular renormalization scheme used are given in the $\overline{\rm MS}$ scheme \cite{Bardeen:1978yd,Furmanski:1981cw}.
In the perturbative momentum regime, i.e.\ for momenta $p=|\vec{p}|\gg\Lambda_{\rm QCD}$, the $Q\bar Q$ static potential is conveniently represented as
\begin{equation}
V(p)=-C_F\frac{4\pi}{p^2}\alpha_V[\alpha_s(\mu),L(\mu,p)] \label{eq:V(p)}
\end{equation}
with $C_F$ the eigenvalue of the quadratic Casimir operator for the fundamental representation of the gauge group; $C_F=4/3$ for ${\rm SU}(3)$.
Pulling out the overall factor $\sim1/p^2$ the entire non-trivial structure of $V(p)$ can be encoded in the dimensionless quantity $\alpha_V$,
which in turn is a function of both the coupling $\alpha_s(\mu)$ and $L\equiv L(\mu,p)=\ln\frac{\mu^2}{p^2}$. $\alpha_s(\mu)$ is evaluated at an {\it a priori} arbitrarily chosen renormalization scale $\mu$ in the perturbative regime, i.e.\ $\mu\gg\Lambda_{\rm QCD}$, guaranteeing that $\alpha_s(\mu)\ll1$.
The running of the coupling $\alpha_s(\mu)$ as a function of the renormalization scale $\mu$ is governed by the QCD $\beta$-function defined as
\begin{equation}
\beta[\alpha_s(\mu)]\equiv\frac{\mu}{\alpha_s(\mu)}\frac{d}{d \mu}\alpha_s(\mu)\,, \label{eq:beta}
\end{equation}
whose series expansion in powers of $\alpha_s$ is presently known with the following accuracy,
\begin{equation}
\beta(\alpha_s)=-\frac{\alpha_s}{2\pi}\beta_0\biggl[1+\frac{\alpha_s}{4\pi}\frac{\beta_1}{\beta_0}+\biggl(\frac{\alpha_s}{4\pi}\biggr)^2\frac{\beta_2}{\beta_0}+\biggl(\frac{\alpha_s}{4\pi}\biggr)^3\frac{\beta_3}{\beta_0}
+ \ldots\,\biggr]\,. \label{eq:betaseries}
\end{equation}
While the expansion coefficients $\beta_0$ and $\beta_1$ are independent of the renormalization scheme,
$\beta_2$ and $\beta_3$ are scheme-dependent. They have been
determined for arbitrary compact semi-simple Lie groups in the $\overline{\rm
MS}$ scheme \cite{vanRitbergen:1997va}.
For ${\rm SU}(3)$ with $n_f=2$ massless dynamical quark flavors they read
\begin{equation}
\beta_0 =\frac{29}{3}\,, \quad \beta_1= \frac{230}{3}\,,\quad \beta_2 = \frac{48241}{54}\,,\quad \beta_3 = \frac{18799309}{1458}+\frac{275524}{81}\zeta(3) .\!
\end{equation}
The same quantities for arbitrary values of $n_f$ can be found e.g.\ in \cite{Jansen:2011vv}.
As the static potential is a physical observable, it should of course be independent of the explicit value of the renormalization scale $\mu$
and form a renormalization group (RG) invariant, i.e.\ fulfill
\begin{equation}
\mu\frac{d}{d\mu}V(p)=0\,. \label{eq:RG_V}
\end{equation}
One might wonder, how this can come about with $\alpha_V$ in \Eqref{eq:V(p)} being a function of the two $\mu$-dependent quantities $\alpha_s(\mu)$ and $L(\mu,p)$.
However, knowing $V(p)$ -- and thus $\alpha_V$ -- at a certain
accuracy in perturbation theory, e.g.\ up to ${\cal O}(\alpha_s^{\bar k})$,
\Eqref{eq:RG_V} only has to hold to this order, i.e.\
\begin{equation}
\mu\frac{d}{d\mu}\alpha_V[\alpha_s,L]={\cal O}(\alpha_s^{\bar k+1})
\quad\quad\leftrightarrow\quad\quad\left(\frac{\partial
} { \partial L}+\frac{\alpha_s}{2}\beta(\alpha_s)\frac{\partial}{\partial
\alpha_s}\right)\alpha_V[\alpha_s,L]={\cal O}(\alpha_s^{\bar k+1}) \,. \label{eq:RG_alphaV}
\end{equation}
Presently, all terms are known explicitly for $\bar k=4$, and $\alpha_V$ is of the following form,
\begin{multline}
\alpha_{V}[\alpha_s(\mu),L(\mu,p)] = \alpha_s(\mu)\left\{1+\frac{\alpha_s(\mu)}{4\pi}\,P_1(L)
+\left(\frac{\alpha_s(\mu)}{4\pi}\right)^2P_2(L)\right. \\
\left.+\left(\frac{\alpha_s(\mu)}{4\pi}
\right)^3\Bigl[P_3(L)+a_{3 {\rm
ln}}\ln{\alpha_s}(\mu)\Bigr]+\ldots
\right\} . \label{eq:pert1}
\end{multline}
While $\alpha_V$ has a strict power-series expansion in $\alpha_s$ up to ${\cal O}(\alpha_s^3)$, beyond this order one also encounters logarithmic contributions in $\alpha_s$ \cite{Appelquist:1977es}, the first such term being $\sim \alpha_s^4\ln{\alpha_s}$ \cite{Brambilla:1999qa}.
Equation~\eqref{eq:RG_alphaV} constrains the $P_k(L)$ accounting for the entire $L$ dependence of $\alpha_V$ in \Eqref{eq:pert1} to be polynomials in $L$ of degree $k$, i.e.
\begin{equation}
P_k(L)=\sum_{m=0}^{k}\rho_{km}L^m \label{eq:Pk}
\end{equation}
with dimensionless expansion coefficients $\rho_{km}$, and implies that, apart from the explicit values of $a_k\equiv\rho_{k0}=P_k(0)$ with $a_0=1$, the $\rho_{km}$ are fully determined by the coefficients of the $\beta$-function \cite{Chishtie:2001mf,Karbstein:2013zxa}. Their explicit expressions for $k\leq3$ are given in our notations in \cite{Karbstein:2013zxa}.
The coefficients $a_1$ \cite{Fischler:1977yf,Billoire:1979ih} and $a_2$ \cite{Peter:1996ig,Peter:1997me,Schroder:1998vy} are known analytically.
For gauge group ${\rm SU}(3)$, $n_f=2$ and in the $\overline{\rm MS}$-scheme they read
\begin{equation}
a_1=\frac{73}{9}\,, \quad a_2=\frac{25139}{162}+9\pi^2\Bigl(4-\frac{\pi^2}{4}\Bigr)+\frac{94}{3}\zeta(3) \,.
\end{equation}
Also the coefficients $a_3$ and $a_{3{\rm ln}}$ are known \cite{Smirnov:2008pn,Smirnov:2009fh,Anzai:2009tm,Smirnov:2010zc,Anzai:2010td,Brambilla:1999qa}.
Specializing to ${\rm SU}(3)$ and $n_f=2$, they are given by (cf.\ \cite{Jansen:2011vv})
\begin{multline}
a_3=27c_1 +\frac{15}{16}c_2+9c_3+\frac{5}{48}c_4-\frac{968981}{729}-8\pi^2\Bigl(15-\frac{8\pi^2}{45}\Bigr) \\
+144\pi^2\Bigl(\ln3+\gamma_E\Bigr) +\frac{38192}{27}\zeta(3)+\frac{320}{9}\zeta(5)
\end{multline}
and $a_{3{\rm ln}}=144\pi^2$.
The constants $c_i$ ($i=1\ldots4$) are only known numerically:
\begin{equation}
c_1 = 502.24(1)\,, \quad
c_2 = -136.39(12)\,, \quad
c_3 = -709.717\,, \quad
c_4 = -56.83(1)\,.
\end{equation}
$c_1$ and $c_2$ have been determined independently by both \cite{Smirnov:2009fh} and \cite{Anzai:2009tm}.
We use the numerical values from \cite{Smirnov:2009fh}, who provide smaller statistical errors.\footnote{The errors associated with $c_i$ $(i=1\ldots4)$ turn out to be negligible in the context of our $\Lambda_{\overline{\textrm{\rm MS}}}$ determination; therefore, we will not discuss them any further.} $c_3$ and $c_4$ have been determined by \cite{Smirnov:2008pn}.
The coefficients $a_i$ ($i=1\ldots4$) for arbitrary values of $n_f$ can be found, e.g., in \cite{Jansen:2011vv}.
Therewith, all coefficients in the perturbative expansion of the static potential in momentum space up to order $\alpha_s^4$ have been assembled.
We emphasize again that the resulting expression is independent of the explicit choice for $\mu$, in the sense that
different choices for $\mu$ only lead to deviations at ${\cal O}(\alpha_s^5)$ [cf.\ \Eqref{eq:RG_alphaV}], which is beyond the accuracy of the contributions taken into account in \Eqref{eq:pert1}.
In order to prevent the logarithms $L$ in \Eqref{eq:pert1} from becoming large and thereby spoil the perturbative expansion, it is desirable to ensure that $\mu$ does not deviate much from $p$.
Hence, a particularly convenient choice for the renormalization scale is $\mu\equiv p$, implying $L=0$. We will exclusively stick to this choice throughout the remainder of this paper.
Adopting this choice, $\alpha_V$ simplifies significantly. It becomes a function of $\alpha_s(p)$ only and reads
\begin{equation}
\alpha_{V}[\alpha_s(p)] = \alpha_s(p)\left\{1+\frac{\alpha_s(p)}{4\pi}\,a_1
+\left(\frac{\alpha_s(p)}{4\pi}\right)^2a_2
+\left(\frac{\alpha_s(p)}{4\pi}
\right)^3\Bigl[a_3+a_{3 {\rm
ln}}\ln{\alpha_s}(p)\Bigr]+\ldots
\right\} . \label{eq:pert2}
\end{equation}
For completeness, note that -- as briefly mentioned in the introduction -- an analogous choice in position space, i.e.\ setting $\mu\equiv1/r$ completely spoils the convergence of the perturbative potential $V(r)$ \cite{Billoire:1979ih}.
As will also become obvious below, the choice $\mu\equiv p$ does not lead to any problems, the reason being the much less pronounced IR sensitivity of the perturbative potential $V(p)$ in momentum space \cite{Beneke:1998rk}.
Let us here also mention the papers by \cite{Brodsky:1982gc,Lepage:1992xa} who argue that a particularly convenient choice of the RG scale $\mu$ is given by $\mu=p\,{\rm e}^{-5/6}$, rendering \Eqref{eq:pert1} independent of $n_f$ up to ${\cal O}(\alpha_s^2)$. While this choice is especially convenient when comparing the results for $\alpha_V$ as an expansion in powers of $\alpha_s(\mu)$ for different numbers of dynamical quarks as the $n_f$ dependency is relegated to higher order expansion coefficients, for our analysis this choice has no advantages and does not provide a handle to improve the results:
Even though the expansion coefficients of powers of $\alpha_s^{1+n}(\mu)$ show a slightly less pronounced increase with $n$ for this choice (cf.\ also our detailed analysis for $\mu=p$ presented below),
for a given value of $p$ the explicit numerical value of $\alpha_s(p\,{\rm e}^{-5/6})$ is substantially increased as compared to $\alpha_s(p)$.
In the momentum regime where both lattice simulations and perturbative calculations for the $Q\bar Q$ static potential are viable, the combined effect of these two opposite tendencies does not favor $\mu=p\,{\rm e}^{-5/6}$ in comparison to $\mu=p$.
An alternative representation of \Eqref{eq:pert2} is
\begin{multline}
\frac{\alpha_{V}[\alpha_s(p)]}{\alpha_s(p)} = 1+x\,a_1
+\left(x\,a_1\right)^2\frac{a_2}{a_1^2} \\
+\left(x\,a_1
\right)^3\frac{a_3+a_{3 {\rm
ln}}\ln(4\pi/a_1)}{a_1^3}\Bigl[1+\frac{a_{3 {\rm
ln}}}{a_3+a_{3 {\rm
ln}}\ln(4\pi/a_1)}\ln(xa_1)\Bigr]+\ldots , \label{eq:pert3}
\end{multline}
where we employed the shortcut notation $x\equiv\frac{\alpha_s(p)}{4\pi}$.
In order to allow for more insights into the structure of $\alpha_V[\alpha_s(p)]$ for $n_f=2$, we insert the explicit numerical values for the coefficients $a_1$, $a_2$, $a_3$ and $a_{3{\rm ln}}$ into \Eqref{eq:pert3}, and rewrite it as
\begin{equation}
\frac{\alpha_{V}[\alpha_s(p)]}{\alpha_s(p)} \approx 1+x\,a_1
+5.00\left(x\,a_1\right)^2
+17.63\left(x\,a_1
\right)^3\Bigl[1+0.15\ln(xa_1)\Bigr]+\ldots , \label{eq:pert4}
\end{equation}
with $a_1\approx8.11$.
It is easy to check that for $xa_1\lesssim\frac{1}{5}$ $\leftrightarrow$ $\alpha_s(p)\lesssim\frac{4\pi}{5a_1}\approx0.31$ the contributions $\sim(xa_1)^n$ in \Eqref{eq:pert4} are ordered in the sense that they become increasingly less important, when increasing $n$ from $0$ to $3$.
For larger values of $\alpha_s(p)\gtrsim0.31$ this ordering is spoiled.
Correspondingly, the perturbative expansion of $\alpha_V[\alpha_s(p)]$ with $n_f=2$ can in particular be considered as well-behaved and controllable for $\alpha_s(p)\ll 0.31$.
As the highest order contribution $\sim(xa_1)^3$ in \Eqref{eq:pert4} is still significantly smaller than the next-to-highest one $\sim(xa_1)^2$, in the present paper we will consider the perturbative expansion as trustworthy even up to $\alpha_s(p)\lesssim 0.31$.
Taking into account the value of $\Lambda_{\overline{\rm MS}}$ as determined in \cite{Jansen:2011vv}, $\Lambda_{\overline{\rm MS}}^{(n_f=2)}=315 \, \textrm{MeV}$, this directly translates into a restriction to momenta $p \gtrsim 1500 \, \textrm{MeV}$ (cf.\ Sec.~\ref{sec:LambdaMSbar}).
\subsection{\label{sec:LambdaMSbar}The scale $\Lambda_{\overline{\rm MS}}$}
So far we have not discussed, how an explicit numerical value can be attributed to the strong coupling $\alpha_s(\mu)$ evaluated at a given momentum $\mu$.
We emphasize that such an identification is renormalization scheme dependent, the most widely used scheme being the $\overline{\rm MS}$ scheme \cite{Bardeen:1978yd,Furmanski:1981cw}, which we also adopt here.
A straightforward integration of \Eqref{eq:beta} yields
\begin{equation}
\ln\frac{\mu}{\Lambda
=\biggl[\int\frac{{\rm d}\alpha_s}{\alpha_s}\frac{1}{\beta(\alpha_s)}\biggr]\bigg|_{\alpha_s=\alpha_s(\mu)} + C\,, \label{eq:beta2}
\end{equation}
where the conventional definition of $\Lambda$ in the $\overline{\rm MS}$ scheme, i.e.\ $\Lambda\to\Lambda_{\overline{\rm MS}}$,
corresponds to the choice of $C=\frac{\beta_1}{2\beta_0^2}\ln(\frac{\beta_0}{4\pi})$ \cite{Chetyrkin:1997sg}\footnote{For completeness, note that our conventions slightly differ from those of \cite{Chetyrkin:1997sg},
who write \Eqref{eq:beta} in terms of $a\equiv\alpha_s/\pi$. Moreover, $\beta_n|_{\textrm{\cite{Chetyrkin:1997sg}}}=\beta_n/4^{n+1}$ and $b_n|_{\textrm{\cite{Chetyrkin:1997sg}}}=\beta_n/(4^n\beta_0)$.}.
Employing some elementary manipulations and rearrangements, \Eqref{eq:beta2} adopted to the $\overline{\rm MS}$ scheme can be written as
\begin{equation}
\Lambda_{\overline{\rm MS}}\equiv\mu\left(\frac{\beta_0\alpha_s(\mu)}{4\pi}\right)^{-\frac{\beta_1}{2\beta_0^2}}\exp\left\{-\frac{2\pi}{\beta_0\alpha_s(\mu)}-\int_0^{\alpha_s(\mu)}\frac{{\rm d}\alpha_s}{\alpha_s}\left[\frac{1}{\beta(\alpha_s)}+\frac{2\pi}{\beta_0\alpha_s}-\frac{\beta_1}{2\beta_0^2}\right]\right\}, \label{eq:Lambda2-4}
\end{equation}
involving a definite integral over $\alpha_s$ (cf.\ e.g.\ \cite{Capitani:1998mq}). The additional terms apart from the factor of $1/\beta(\alpha_s)$ have been included in the integrand to make the finiteness of the integral over the interval from $0$ to $\alpha_s(\mu)$ manifest.
In general \Eqref{eq:Lambda2-4} cannot be solved explicitly to provide the coupling $\alpha_s(\mu)$ at a given momentum scale $\mu$ as a function of the ratio $\mu/\Lambda_{\overline{\rm MS}}$. An exact closed form solution is only possible at leading order -- i.e.\ taking into account only the leading contribution of the $\beta$-function~\eqref{eq:betaseries}, $\beta(\alpha_s)\approx-\frac{\alpha_s}{2\pi}\beta_0$ -- and reads $\alpha_s(\mu)=\bigl[\frac{\beta_0}{2\pi}\ln(\mu/\Lambda)\bigr]^{-1}$.
Note, however, that approximate results for $\alpha_s(\mu)$ as a function of the ratio $\mu/\Lambda_{\overline{\rm MS}}$ are available also for higher order contributions:
the derivation of such expressions involves expansions in terms of $\ln(\mu/\Lambda_{\overline{\rm MS}})$; see e.g.\ \cite{Chetyrkin:1997sg,Beringer:1900zz}.
Aiming at the determination of $\Lambda_{\overline{\rm MS}}$ by fitting the perturbative expression for the static potential~\eqref{eq:V(p)} to numerical data from lattice simulations
we do not see any reason to resort to these further approximations. Our strategy is rather to solve the implicit equation~\eqref{eq:Lambda2-4} for $\alpha_s(\mu)$ numerically.
Hence, we will always use the relation~\eqref{eq:Lambda2-4} between $\alpha_s(\mu)$ and $\mu/\Lambda_{\overline{\rm MS}}$ at the presently best available accuracy, irrespectively of the order of the expansion of the perturbative potential in $\alpha_s(p)$.
However, note that there is still some freedom left in deciding, how to proceed with the evaluation of \Eqref{eq:Lambda2-4}.
We can
\begin{itemize}
\item[(I)] either plug in the perturbative expression of the $\beta$-function~\eqref{eq:betaseries} at the presently best known accuracy and then do the integration over $\alpha_s$ numerically,
\item[(II)] or adopt a Taylor expansion of the integrand in \Eqref{eq:Lambda2-4} and do the integral analytically, keeping only those terms, whose coefficients are known explicitly,
\begin{multline}
\int_0^{\alpha_s(\mu)}\frac{{\rm d}\alpha_s}{\alpha_s}\left[\frac{2}{\beta(\alpha_s)}+\frac{4\pi}{\beta_0\alpha_s}-\frac{\beta_1}{\beta_0^2}\right] \\
= \frac{\beta_0\beta_2 - \beta_1^2}{\beta_0^3}\frac{\alpha_s(\mu)}{4\pi} + \frac{\beta_0^2\beta_3 - 2\beta_0\beta_1\beta_2+\beta_1^3}{2\beta_0^4}\left(\frac{\alpha_s(\mu)}{4\pi}\right)^2 +{\cal O}(\alpha_s^3). \label{eq:auch}
\end{multline}
\end{itemize}
Of course, limiting ourselves to the truly perturbative regime, i.e.\ the regime, where the perturbative expansion of the $\beta$-function~\eqref{eq:betaseries} is such that higher order corrections become increasingly less important,
both choices should be equally justified.
As pointed out in Sec.~\ref{sec:perturbation}, when fitting the perturbative static potential~\eqref{eq:V(p)} in momentum space to lattice data we will always stick to the identification $\mu\equiv p$.
Correspondingly, higher order corrections in \Eqref{eq:betaseries} should remain small throughout the integration interval from $0$ to $\alpha(p)$ in \Eqref{eq:Lambda2-4} and \Eqref{eq:auch}, respectively.
To get a feeling, how the perturbative expansion of the $\beta$-function behaves,
we employ the same strategy as adopted in the context of \Eqref{eq:pert4} and study its behavior at the upper integration limit in \Eqref{eq:Lambda2-4}, i.e.\ at $\alpha_s\equiv\alpha_s(p)$.
First we rewrite \Eqref{eq:betaseries} as
\begin{equation}
\beta(\alpha_s)=-\frac{\alpha_s}{2\pi}\beta_0\biggl[1+x\,\frac{\beta_1}{\beta_0}+\biggl(x\,\frac{\beta_1}{\beta_0}\biggr)^2\frac{\beta_0\beta_2}{\beta_1^2}+\biggl(x\,\frac{\beta_1}{\beta_0}\biggr)^3\frac{\beta_0^2\beta_3}{\beta_1^3}
+ \ldots\,\biggr] \label{eq:betaseries2}
\end{equation}
with $x\equiv\frac{\alpha_s(p)}{4\pi}$. Second we insert the explicit numerical coefficients of the $\beta$-function for $n_f=2$ into this equation, resulting in
\begin{equation}
\beta(\alpha_s)\approx-\frac{\alpha_s}{2\pi}\beta_0\biggl[1+x\,\frac{\beta_1}{\beta_0}+1.47\biggl(x\,\frac{\beta_1}{\beta_0}\biggr)^2+3.52\biggl(x\,\frac{\beta_1}{\beta_0}\biggr)^3
+ \ldots\,\biggr] \label{eq:betaseries3}
\end{equation}
with $\frac{\beta_1}{\beta_0}\approx7.93$.
Using the same reasoning as in the context of \Eqref{eq:pert4} above, we find that for the contributions $\sim(x\beta_1/\beta_0)^n$ to become increasingly less important with $n$,
we have to demand $x\beta_1/\beta_0\lesssim0.4$, which corresponds to $\alpha_s(p)\lesssim0.63$.
Hence, in particular for $\alpha_s(p)\lesssim0.31$ -- which was the value found to crudely delimit the range of validity of a perturbative expansion of the static potential in momentum space for $n_f=2$ [cf.\ in the context of \Eqref{eq:pert4} above] -- higher order terms in the perturbative expansion of the $\beta$-function~\eqref{eq:betaseries} for $n_f=2$ are expected to become much less important.
In turn both possibilities (I) and (II) discussed above to numerically solve \Eqref{eq:Lambda2-4} for $\alpha_s(p)$ as a function of $p/\Lambda_{\overline{\rm MS}}$ should yield very similar results.
\newpage
\section{\label{SEC600}Determination of $\Lambda_{\overline{\rm MS}}$}
In this section, we determine $\Lambda_{\overline{\textrm{MS}}}$ in units of the lattice spacing, i.e.\ $a\Lambda_{\overline{\textrm{MS}}}$, by fitting perturbative expressions for the static potential in momentum space (cf.\ Sec.~\ref{sec:perturbation}) to corresponding lattice results (cf.\ Sec.~\ref{SEC572}). Using the values of the lattice spacing listed in Table~\ref{TAB077}, these results can easily be converted to physical units, i.e.\ $\textrm{MeV}$. Unless explicitly stated otherwise, the errors provided for $\Lambda_{\overline{\textrm{MS}}}$ do not account for the uncertainties associated with the lattice spacing $a$. These errors will however be accounted for when quoting our final result for $\Lambda_{\overline{\textrm{MS}}}$ [cf.\ \Eqref{EQN844}, below].
For completeness, we will also quote $r_0\Lambda_{\overline{\textrm{MS}}}$, which is dimensionless and, hence, unaffected by any potential uncertainty in $a$ (uncertainties in $r_0 / a$, which are collected in Table~\ref{TAB077}, are, of course, included).
\subsection{Fitting procedures}
The perturbative $Q\bar Q$ potential to be fitted to lattice results is given by
\begin{equation}
\label{EQN678} V(p) = -\frac{4}{3}\frac{4\pi}{p^2}\alpha_s(p) \biggl\{
\underbrace{\underbrace{\underbrace{\underbrace{1}_{\textrm{LO}}
+ \frac{\alpha_s(p)}{4 \pi} a_1}_{\textrm{NLO}}
+ \bigg(\frac{\alpha_s(p)}{4 \pi}\bigg)^2 a_2}_{\textrm{NNLO}}
+ \bigg(\frac{\alpha_s(p)}{4 \pi}\bigg)^3 \Bigl[a_3 + a_{3 {\rm ln}} \ln{\alpha_s(p)}\Bigr]}_{\textrm{NNNLO}}
\biggr\}+V_0,
\end{equation}
where we included an overall constant offset $V_0$ of the potential.
We use different orders in the expansion of $V(p)$ in powers of $\alpha_s(p)$ to test and judge the convergence behavior of our results; the abbreviation ${\rm N}^n{\rm LO}$ stands for (next-to-)$^n$leading-order.
As only energy differences are measurable $V_0$ does not have any observable consequences. However, it is necessary to allow for a meaningful matching of the perturbative potential to lattice results.
For the relation between $\alpha_s(p)$ and $\Lambda_{\overline{\rm MS}}$ we employ (cf.\ Sec.~\ref{sec:perturbation})
\begin{itemize}
\item[(I)] either \Eqref{eq:Lambda2-4} with $\beta(\alpha)$ given by the terms written explicitly in \Eqref{eq:betaseries},
\item[(II)] or \Eqref{eq:Lambda2-4} with the integral expression substituted for the terms written explicitly on the left-hand side of \Eqref{eq:auch}.
\end{itemize}
These implicit equations are solved numerically to yield $\alpha_s(p)$ as a function of $p/\Lambda_{\overline{\rm MS}}$.
We employ an uncorrelated $\chi^2$ minimizing fit with two degrees of freedom, $V_0$ and $\Lambda_{\overline{\rm MS}}$, to fit the perturbative $Q\bar Q$ potential~\eqref{EQN678} to the lattice potential in momentum space.
Our fitting interval is delimited by $p_{\textrm{min}}$ and $p_{\textrm{max}}$.
Note that the lattice potential in momentum space originates from position space results whose large distance behavior is modeled by \Eqref{EQN600} with $M \in \{1,2,3,4\}$ -- and thus also depends on $M$ (cf. Sec.~\ref{SEC572}).
\subsection{\label{SEC329}Systematic and statistical errors of $\Lambda_{\overline{\rm MS}}$}
\subsubsection{\label{input_var}Individual variation of input parameters}
We investigate the stability of $\Lambda_{\overline{\rm MS}}$ with respect to variations of the input parameters $p_\textrm{min}$, $p_\textrm{max}$ and $M$, using our finest lattice spacing $a \approx 0.042 \, \textrm{fm}$ and $(L'/a)^3 = 256^3$ lattice sites. To this end the input parameters delimiting the fitting range are varied in the following intervals:
\begin{itemize}
\item $p_\textrm{min} = 1500 \, \textrm{MeV} \ldots 2250 \, \textrm{MeV}$: \\
As discussed in detail in Sec.~\ref{sec:perturbation}, for $p_\textrm{min} \lesssim 1500 \, \textrm{MeV}$ the validity of perturbative expressions for the static potential is rather questionable.
\item $p_\textrm{max} = 2250 \, \textrm{MeV} \ldots 3000 \, \textrm{MeV}$: \\
The maximum momentum on our finest lattice along an axis is $\overline{p} = \pi / a \approx 15 \, \textrm{GeV}$. For $p\lesssim \overline{p} / 3$, it seems reasonable to expect rather small discretization errors.
These expectations have also been confirmed by numerical investigations.
Fitting the perturbative potential to lattice results we obtain an essentially stable value of $\Lambda_{\overline{\rm MS}}$ up to $3000 \, \textrm{MeV} \ldots 3500 \, \textrm{MeV}$;
cf.\ also Figure~\ref{FIG003} (c) and (d) below.
\end{itemize}
The different choices for the parameter $M$ are
\begin{itemize}
\item $M\in\{1,2,3,4\}$: \\
The parameter $M$ is varied to estimate the systematic errors associated with our modeling of the long range part of the lattice potential in position space; cf.\ Sec.~\ref{EQN808}.
\end{itemize}
Below, we demonstrate that the fit results for $\Lambda_{\overline{\rm MS}}$ are rather stable with respect to these parameter variations, i.e.\ we confirm that a meaningful and rather precise matching of perturbative expressions for the $Q\bar Q$ static potential and lattice results is possible.
Exemplary fits of the perturbative static potential~\eqref{EQN678} at LO, NLO, NNLO and NNNLO to lattice results are depicted in Figure~\ref{FIG002}.
Noteworthily, all four orders are suited to describe the lattice potential within statistical errors, i.e.\ fulfill $\chi^2 / \textrm{dof} \lesssim 1.0$.
Similar $\chi^2 / \textrm{dof}$ are obtained when varying input parameters (cf. below).
\begin{figure}[htb]
\centering
\subfigure[LO - $\Lambda_{\overline{\rm MS}}=534(3)$MeV]{\includegraphics[width=0.4\textwidth]{LOgefittet_BStrich}}\qquad
\subfigure[NLO - $\Lambda_{\overline{\rm MS}}=414(3)$MeV ]{\includegraphics[width=0.4\textwidth]{NLOgefittet_BStrich}}\\
\subfigure[NNLO - $\Lambda_{\overline{\rm MS}}=339(2)$MeV ]{\includegraphics[width=0.4\textwidth]{NNLOgefittet_BStrich}}\qquad
\subfigure[NNNLO - $\Lambda_{\overline{\rm MS}}=315(2)$MeV ]{\includegraphics[width=0.4\textwidth]{NNNLOgefittet_BStrich}}
\caption{\label{FIG002}Exemplary fits of the perturbative static potential~\eqref{EQN678} to lattice results obtained with $\beta = 4.35$ and $M = 3$. We employ fitting procedure (I) with $p_\textrm{min} = 1875 \, \textrm{MeV}$ and $p_\textrm{max} = 2625 \, \textrm{MeV}$. The errors quoted below the plots refer to the fitting and are statistical errors only; systematic errors will be discussed below.}
\end{figure}
To understand, how the value of $\Lambda_{\overline{\rm MS}}$ depends on the input parameters $p_\textrm{min}$ and $p_\textrm{max}$, we vary them individually.
Our results are summarized graphically in Figure~\ref{FIG003}:\footnote{\label{footnote}Statistical errors associated with the $\chi^2$ minimizing fits, which are rather small ($\approx 2\,\textrm{MeV}\ldots4{\rm MeV}$), are not considered in this subsection. They are, however, included in the final results presented in Sec.~\ref{SEC790}.}
\begin{itemize}
\item The plots in the left column
are obtained with fitting procedure (I), while those in the right column
result from fitting procedure (II).
\item For the plots in the first line
we vary $p_\textrm{min} = 1500 \, \textrm{MeV} \ldots 2250 \, \textrm{MeV}$, while keeping $p_\textrm{max} = 2625 \, \textrm{MeV}$ fixed at the center of the interval defined above, and set $M=3$. We find:
\begin{itemize}
\item Variation of $\Lambda_{\overline{\rm MS}}$ (NNLO): \\
$337 \, \textrm{MeV} \ldots 344 \, \textrm{MeV}$ (fitting procedure (I)), \\
$335 \, \textrm{MeV} \ldots 342 \, \textrm{MeV}$ (fitting procedure (II)).
\item Variation of $\Lambda_{\overline{\rm MS}}$ (NNNLO): \\
$313 \, \textrm{MeV} \ldots 317 \, \textrm{MeV}$ (fitting procedure (I)), \\
$312 \, \textrm{MeV} \ldots 315 \, \textrm{MeV}$ (fitting procedure (II)).
\end{itemize}
\item For the plots in the second line
we vary $p_\textrm{max} = 2250 \, \textrm{MeV} \ldots 3000 \, \textrm{MeV}$, while keeping $p_\textrm{min} = 1875 \, \textrm{MeV}$ fixed at the center of the interval defined above, and set $M=3$. We obtain:
\begin{itemize}
\item Variation of $\Lambda_{\overline{\rm MS}}$ (NNLO): \\
$338 \, \textrm{MeV} \ldots 343 \, \textrm{MeV}$ (fitting procedure (I)), \\
$337 \, \textrm{MeV} \ldots 341 \, \textrm{MeV}$ (fitting procedure (II)).
\item Variation of $\Lambda_{\overline{\rm MS}}$ (NNNLO): \\
$314 \, \textrm{MeV} \ldots 317 \, \textrm{MeV}$ (fitting procedure (I)), \\
$313 \, \textrm{MeV} \ldots 315 \, \textrm{MeV}$ (fitting procedure (II)).
\end{itemize}
\item For the plots in the third line
we vary the center of the fitting range $(p_\textrm{min} + p_\textrm{max}) / 2 = 1875 \, \textrm{MeV} \ldots 2625 \, \textrm{MeV}$, while keeping the width of the fitting range $p_\textrm{max} - p_\textrm{min} = 750 \, \textrm{MeV}$ fixed, and set $M=3$. This results in:
\begin{itemize}
\item Variation of $\Lambda_{\overline{\rm MS}}$ (NNLO): \\
$335 \, \textrm{MeV} \ldots 340 \, \textrm{MeV}$ (fitting procedure (I)), \\
$334 \, \textrm{MeV} \ldots 338 \, \textrm{MeV}$ (fitting procedure (II)).
\item Variation of $\Lambda_{\overline{\rm MS}}$ (NNNLO): \\
$313 \, \textrm{MeV} \ldots 315 \, \textrm{MeV}$ (fitting procedure (I)), \\
$312 \, \textrm{MeV} \ldots 314 \, \textrm{MeV}$ (fitting procedure (II)).
\end{itemize}
\end{itemize}
To allow for a meaningful determination of $\Lambda_{\overline{\rm MS}}$, it is important to ensure that the fit results for $\Lambda_{\overline{\rm MS}}$ only exhibit a weak -- preferably negligible -- dependence on $p_\textrm{min}$ and $p_\textrm{max}$. In plots of $\Lambda_{\overline{\rm MS}}$ as a function of $p_\textrm{min}$ and $p_\textrm{max}$ this should manifest itself in the formation of plateaus. This behavior is clearly visible in Figure~\ref{FIG003} for all six plots and at all orders in the expansion~\eqref{EQN678}.
Moreover, the curves shown in Figure~\ref{FIG003} clearly reveal a convergence behavior when increasing the order in \Eqref{EQN678} from LO to NNNLO: First, $\Lambda_{\overline{\rm MS}}|_{\rm LO}>\Lambda_{\overline{\rm MS}}|_{\rm NLO}>\Lambda_{\overline{\rm MS}}|_{\rm NNLO}>\Lambda_{\overline{\rm MS}}|_{\rm NNNLO}$, second
the ratios of their relative differences scale as
\begin{equation}
\Lambda_{\overline{\rm MS}}|_{\rm LO}-\Lambda_{\overline{\rm MS}}|_{\rm NLO}\ :\ \Lambda_{\overline{\rm MS}}|_{\rm NLO}-\Lambda_{\overline{\rm MS}}|_{\rm NNLO}\ :\ \Lambda_{\overline{\rm MS}}|_{\rm NNLO} - \Lambda_{\overline{\rm MS}}|_{\rm NNNLO}\ \approx\ 5\ :\ 3\ :\ 1\,.
\end{equation}
Third, the values of $\Lambda_{\overline{\rm MS}}$ extracted from fits of \Eqref{EQN678} at NNLO and NNNLO to lattice results yield quite similar results. This can be considered as indication that the neglected orders beyond NNNLO in \Eqref{EQN678} will not alter the value of $\Lambda_{\overline{\rm MS}}$ significantly.
As an -- rather conservative -- estimate of the systematic error due to the truncation of the perturbative expansion~\eqref{EQN678} at ${\cal O}(\alpha_s^4)$ we take the difference between the NNLO and the NNNLO results for $\Lambda_{\overline{\rm MS}}$ (cf.\ Sec.~\ref{SEC444}).
\begin{figure}[p]
\centering
\subfigure[fitting proc.\ (I), $p_{\textrm{max}}=2625 \, \textrm{MeV}$]{\includegraphics[width=0.47\textwidth]{expression_I_p_min_runs}}\qquad
\subfigure[fitting proc.\ (II), $p_{\textrm{max}}=2625 \, \textrm{MeV}$]{\includegraphics[width=0.47\textwidth]{expression_II_p_min_runs}}\\
\subfigure[fitting proc.\ (I), $p_{\textrm{min}}=1875 \, \textrm{MeV}$]{\includegraphics[width=0.47\textwidth]{expression_I_p_max_runs}}\qquad
\subfigure[fitting proc.\ (II), $p_{\textrm{min}}=1875 \, \textrm{MeV}$]{\includegraphics[width=0.47\textwidth]{expression_II_p_max_runs}}\\
\subfigure[fitting proc.\ (I), $p_{\textrm{max}}-{p_{\textrm{min}}}=750 \, \textrm{MeV}$]{\includegraphics[width=0.47\textwidth]{expression_I_interval_fix}}\qquad
\subfigure[fitting proc.\ (II), $p_{\textrm{max}}-{p_{\textrm{min}}}=750 \, \textrm{MeV}$]{\includegraphics[width=0.47\textwidth]{expression_II_interval_fix}}
\caption{\label{FIG003}Results for $\Lambda_{\overline{\rm MS}}$ obtained by fitting \Eqref{EQN678} to lattice results using $M = 3$; \textbf{left column:} fitting procedure (I); \textbf{right column:} fitting procedure (II); \textbf{first line:} $\Lambda_{\overline{\rm MS}}$ as a function of $p_\textrm{max}$; \textbf{second line:} $\Lambda_{\overline{\rm MS}}$ as a function of $p_\textrm{min}$; \textbf{third line:} $\Lambda_{\overline{\rm MS}}$ as a function of $(p_\textrm{min} + p_\textrm{max})/2$.}
\end{figure}
Another source of uncertainty in our analysis is the modeling~\eqref{EQN600} used to extrapolate the lattice potential computed from Wilson loops to larger $r$. In order to scrutinize this uncertainty, we study the dependence of $\Lambda_{\overline{\rm MS}}$ on $M\in\{ 1, 2, 3, 4\}$, resorting to fitting procedure (I) and keeping $p_\textrm{min} = 1875 \, \textrm{MeV}$ and $p_\textrm{max} = 2625 \, \textrm{MeV}$ fixed. The corresponding results, which are collected in Table~\ref{TAB015}, indicate that $\Lambda_{\overline{\rm MS}}$ is also rather stable with respect to variations of $M$.
\begin{table}[htb]
\begin{center}
\begin{tabular}{|c|c|c|c|c|}
\hline
& $V_1(r)$ & $V_2(r)$ & $V_3(r)$ & $V_4(r)$ \\
\hline
NNLO, fitting procedure (I) & 349(2) & 343(2) & 340(3) & 346(4) \\
NNLO, fitting procedure (II) & 347(2) & 341(2) & 338(2) & 345(4) \\
\hline
NNNLO, fitting procedure (I) & 323(2) & 318(2) & 315(2) & 321(4) \\
NNNLO, fitting procedure (II) & 322(2) & 317(2) & 314(3) & 320(4) \\
\hline
\end{tabular}
\caption{\label{TAB015}Results for $\Lambda_{\overline{\rm MS}}^{(n_f=2)}$ in $\textrm{MeV}$ for different $M \in \{1, 2, 3, 4\}$ using fitting procedure (I) and $p_\textrm{min} = 1875 \, \textrm{MeV}$ and $p_\textrm{max} = 2625 \, \textrm{MeV}$.}
\end{center}
\end{table}
The variation of $\Lambda_{\overline{\rm MS}}$ as a function of the input parameters $p_\textrm{min}$, $p_\textrm{max}$ and $M$ (cf. also Figure~\ref{FIG003} and Table~\ref{TAB015}) when fitting~\eqref{EQN678}, both at NNLO (red) and at NNNLO (green), to lattice results is visualized in Figure~\ref{FIG003_error}. As systematic uncertainty one could e.g.\ quote the whole range of values covered by the NNNLO variations,
\begin{equation}
\label{EQN257} \Lambda_{\overline{\textrm{MS}}}^{(n_f=2)} = 314 \, \textrm{MeV} \ldots 323 \, \textrm{MeV} .
\end{equation}
In particular note the drastic improvement in stability as compared to the previous position space analysis~\cite{Jansen:2011vv}, where a systematic uncertainty larger by a factor of $\approx 4$ has been obtained, $\Lambda_{\overline{\rm MS}}^{(n_f=2)} = 304 \, \textrm{MeV} \ldots 344 \, \textrm{MeV}$ (cf.\ Eq.\ (46) and Figure~5 of \cite{Jansen:2011vv}).
\begin{figure}[htb]
\begin{center}
\includegraphics[width=0.8\textwidth]{summary}
\end{center}
\caption{\label{FIG003_error}Graphical summary of the variation of $\Lambda_{\overline{\rm MS}}$, when varying the input parameters $p_\textrm{min}$, $p_\textrm{max}$ and $M$ (cf.\ also Figure~\ref{FIG003}, Table~\ref{TAB015} and the corresponding paragraphs in the main text).
For each line, we indicate the fitting procedure used, i.e., either (I) and (II), and specify which parameter is varied. The variation is over the intervals specified at the begin of Sec.~\ref{input_var}.
The other input parameters are held fixed at the center of the respective intervals, and $M=3$ unless otherwise stated.}
\end{figure}
The sources of the systematic error discussed above might, however, be correlated. A method to determine the overall systematic error accounting for potential correlation is discussed in the following subsection.
\subsubsection{\label{SEC444}Consideration of correlations between different systematic error sources}
To account for possible correlations, we perform a large number of fits, with the input parameters chosen randomly and uniformly in the intervals specified in Sec.~\ref{input_var}. As systematic error we then take the variance of the fit results.
This procedure is analogous to that used for a $\Lambda_{\overline{\rm MS}}$ determination from the $Q\bar Q$ static potential in position space based upon the same lattice data \cite{Jansen:2011vv} to which we will compare our results in the following.
We have performed 40,000 fits (i.e., sufficiently many to render the statistical error of the variance negligible):
\begin{itemize}
\item 10,000 NNLO fits, fitting procedure (I);
\item 10,000 NNLO fits, fitting procedure (II);
\item 10,000 NNNLO fits, fitting procedure (I);
\item 10,000 NNNLO fits, fitting procedure (II).
\end{itemize}
In these fits we randomly vary $p_{\rm min}$ and $p_{\rm max}$ in the intervals $p_\textrm{min} = 1500 \, \textrm{MeV} \ldots 2250 \, \textrm{MeV}$ and $p_\textrm{max} = 2250 \, \textrm{MeV} \ldots 3000 \, \textrm{MeV}$ while imposing the constraint $p_\textrm{max} - p_\textrm{min} \geq 375 \, \textrm{MeV}$. Moreover, we cyclically vary $M \in\{ 1, 2, 3, 4\}$. From this analysis we obtain an average and a variance, i.e.\ a systematic error, of
\begin{equation}
\label{EQN791} \Lambda_{\overline{\rm MS}}^{(n_f=2)} = 331(13) \, \textrm{MeV} .
\end{equation}
Again it is most instructive to compare the error with that obtained for a position space analysis of the $Q\bar Q$ static potential based upon the same lattice data \cite{Jansen:2011vv}:
\begin{itemize}
\item The final result of Ref.~\cite{Jansen:2011vv} is $\Lambda_{\overline{\rm MS}}^{(n_f=2)} = 315(26) \, \textrm{MeV}$. Hence, the momentum space analysis pursued in the present paper yields a result which is more precise by a factor of $\approx 2$.
\item While fitting procedure (A) used in \cite{Jansen:2011vv} is rather similar to (I) and (II) of this paper, fitting procedure (B) used in \cite{Jansen:2011vv} is somewhat different.\footnote{In fact, fitting procedure (B) of \cite{Jansen:2011vv} is expected to increase the error because the one-loop result for $\alpha_s$, employed in this fitting procedure, deviates significantly from the higher-order expressions for $\alpha_s$ throughout the considered fitting interval.} Hence, in order to allow for a consistent and fair comparison of our present momentum space analysis and the previous position space analysis, only fitting procedure (A) should be used. We carried out such an analysis, resulting in $\Lambda_{\overline{\rm MS}}^{(n_f=2)} = 331(20) \, \textrm{MeV}$. While there is now perfect agreement with respect to the average, the error associated with a momentum space analysis is still smaller by a factor of $\approx 1.5$ than the error obtained in a position space analysis.
\item Finally, one could be less conservative with regard to the estimate of the uncertainty associated with the truncation of the perturbative expressions and only take into account the NNNLO results. A momentum space analysis would then result in $\Lambda_{\overline{\rm MS}}^{(n_f=2)} = 318(3) \, \textrm{MeV}$, while a position space analysis (fitting procedure (A) only) would lead to $\Lambda_{\overline{\rm MS}}^{(n_f=2)} = 326(13) \, \textrm{MeV}$. In this case the error of $\Lambda_{\overline{\rm MS}}$ as obtained from a momentum space analysis is even smaller by a factor of $\approx 4$ as compared to an analogous position space analysis. This rather drastic difference can be attributed to the almost perfect stability of $\Lambda_{\overline{\rm MS}}$ at NNNLO with respect to variations of the fitting range in momentum space (cf.\ Figure~\ref{FIG003}). In position space, a similar stability has not been observed \cite{Jansen:2011vv}.
\end{itemize}
Let us emphasize again that all these results for $\Lambda_{\overline{\rm MS}}^{(n_f=2)}$ agree within statistical errors.
As already mentioned in footnote~\ref{footnote}, the fitting procedure introduces an additional statistical uncertainty of $\approx 2 \, \textrm{MeV} \ldots 4 \, \textrm{MeV}$, which we add in quadrature when generating the final result (cf.\ Sec.~\ref{SEC790}).
Further systematic uncertainties, exclusively originating from the lattice static potential, are discussed in the next section.
\subsubsection{\label{SEC330}Systematic errors of $\Lambda_{\overline{\rm MS}}$ associated with the lattice computation}
\subsubsection*{Lattice discretization errors}
To estimate the order of magnitude of lattice discretization errors, we utilize the lattice ensembles listed in Table~\ref{TAB077}, featuring pions of roughly of the same mass $\approx 284 \, \textrm{MeV} \ldots 352 \, \textrm{MeV}$.
Keeping the input parameters $p_\textrm{min} = 1300 \, \textrm{MeV}$, $p_\textrm{max} = 2050 \, \textrm{MeV}$ and $M=3$ fixed, we extract values for both $r_0\Lambda_{\overline{\rm MS}}$ and $\Lambda_{\overline{\rm MS}}$ by performing the corresponding fits.
For these investigations we exclusively adopt fitting procedure (I) and use the perturbative $Q\bar Q$ static potential~\eqref{EQN678} at the best known accuracy, i.e.\ NNNLO.\footnote{To allow for stable fits with sufficiently many lattice points for $V(p)$ inside the interval $p_\textrm{min} \ldots p_\textrm{max}$, with $p_\textrm{max}$ not too close to the maximum lattice momentum along one of the three spatial axes, we had to choose $p_\textrm{min}$ slightly smaller than the minimum value of $p_{\rm min}=1500 \, \textrm{MeV}$ defined in Sec.~\ref{input_var}.} In Figure~\ref{FIG954} we depict our results for $r_0\Lambda_{\overline{\rm MS}}$ and $\Lambda_{\overline{\rm MS}}$ as a function of $a^2$. The results determined at the four available lattice spacings perfectly agree within errors, i.e.\ there is no indication of any sizable lattice discretization errors.
\begin{figure}[htb]
\centering
\includegraphics[width=0.48\textwidth]{continuum_Lambda_r0}\includegraphics[width=0.48\textwidth]{continuum}
\caption{\label{FIG954}Dependence of $r_0\Lambda_{\overline{\rm MS}}$ and $\Lambda_{\overline{\rm MS}}$ on the lattice spacing. The result of an continuum extrapolation is depicted in green.}
\end{figure}
For completeness note that the errors associated with the extracted values for $\Lambda_{\overline{\rm MS}}$ in Figure~\ref{FIG954} are rather large.
The reason for this is that we have here included the errors associated with the lattice spacings and also with $r_0 / a$ (cf.\ Table~\ref{TAB077}). This is absolutely essential when comparing results obtained at different values of the lattice spacing.
Figure~\ref{FIG954} also shows continuum extrapolations assuming a dependence $\propto a^2$ for both $r_0\Lambda_{\overline{\rm MS}}$ and $\Lambda_{\overline{\rm MS}}$. Within errors the results of these extrapolations agree with the results obtained for our smallest lattice spacing ($\beta = 4.35$, $a\approx 0.042\,{\rm fm}$). Even though no clear indication of any systematic upward or downward tendency as a function of $a^2$ is visible in Figure~\ref{FIG954}, we are conservative in our error estimate and infer additional systematic errors of $\pm 8 \, \textrm{MeV}$ for $\Lambda_{\overline{\rm MS}}$ and $\pm 0.010$ for $r_0\Lambda_{\overline{\rm MS}}$ to be accounted in our final result. These estimates are obtained by taking the difference between the central values of $r_0\Lambda_{\overline{\rm MS}}$ and $\Lambda_{\overline{\rm MS}}$ at our smallest lattice spacing and the corresponding continuum extrapolations.
\subsubsection*{Finite volume effects and non-vanishing light quark mass corrections}
Finite volume effects were investigated in detail in Ref.~\cite{Jansen:2011vv}. In this analysis such effects were found to be negligible compared to the other errors discussed above.
Similarly, potential corrections on $\Lambda_{\overline{\rm MS}}$ due to non-vanishing light quark masses were examined in detail in \cite{Jansen:2011vv} by studying the variation of $\Lambda_{\overline{\rm MS}}$ for different pion masses in the range $m_\textrm{PS} \approx
325 \, \textrm{MeV} \ldots 517 \, \textrm{MeV}$ at fixed lattice spacing and spacetime volume.
In the quark mass region investigated, $\Lambda_{\overline{\rm MS}}$ was found to be constant within tiny statistical errors of $\approx\pm 1 \,\textrm{MeV}$.
Therefore, we do not expect the non-vanishing light quark masses on the lattice to induce any significant deviations from the zero quark mass limit, for which the perturbative expressions in Sec.\ \ref{sec:perturbation} were derived.
In other words, we consider the systematic error introduced by comparing a lattice computation featuring massive light quarks with a perturbative calculation in the zero quark mass limit negligible compared to the other uncertainties discussed above.
For the above reasons we do not take any potential errors arising from finite volume effects and due to non-vanishing light quark mass corrections into account when quoting our final result for $\Lambda_{\overline{\rm MS}}$ in this paper.
\subsection{\label{SEC790}Final results for $\Lambda_{\overline{\rm MS}}$}
In the following, we present our final results for $\Lambda_{\overline{\rm MS}}^{(n_f=2)}$. We quote these results both in units of $\textrm{MeV}$ and in units of $r_0$. Since there seem to be up to $\mathcal{O}(10\%)$ unresolved differences regarding scale setting between different lattice QCD collaborations \cite{Sommer:2014mea}, we prefer the scale setting-independent quantity $r_0\Lambda_{\overline{\rm MS}}$.
The determination is based on the lattice results at our finest lattice spacing, i.e.\ at $\beta = 4.35$, as explained in Sec.~\ref{SEC444}. The individual error contributions, which we combine by adding them in quadrature, are
\begin{itemize}
\item[(1)] the correlated systematic errors associated with the unknown contributions beyond ${\cal O}(\alpha_s^4)$ of the perturbative $Q\bar Q$ static potential and the input parameters of the fitting procedure (cf.\ Sec.~\ref{SEC444}),
\item[(2)] the statistical errors associated with the $\chi^2$ minimizing fits,
\item[(3)] the estimated lattice discretization errors (cf.\ Sec.~\ref{SEC330}), and
\item[(4)] the errors associated with the lattice spacing $a = 0.0420(17) \, \textrm{fm}$ and with $r_0 / a = 9.81(13)$.
\end{itemize}
Taking all these contributions into account, we finally obtain
\begin{equation}
\label{EQN844}r_0\Lambda_{\overline{\rm MS}}^{(n_f=2)}\ = 0.692(31) \, , \quad \Lambda_{\overline{\rm MS}}^{(n_f=2)} = 331(21) \, \textrm{MeV}.
\end{equation}
For completeness, note that being less conservative with regard to the estimate of the uncertainty associated with the truncation of the perturbative expressions and just taking NNNLO fits into account (cf. also Sec.~\ref{SEC444}), would result in $r_0\Lambda_{\overline{\rm MS}}^{(n_f=2)}= 0.665(16)$ and $\Lambda_{\overline{\rm MS}}^{(n_f=2)} = 318(16) \, \textrm{MeV}$.
The final results~\eqref{EQN844} for $\Lambda_{\overline{\rm MS}}$ extracted from our momentum space analysis are more precise than those obtained in position space using the same lattice data: $r_0\Lambda_{\overline{\rm MS}}^{(n_f=2)}= 0.658(55)$ and $\Lambda_{\overline{\rm MS}}^{(n_f=2)} = 315(30) \, \textrm{MeV}$.
Let us emphasize again that the errors quoted here and in \Eqref{EQN844} also account for contributions which are not directly related to the extraction of $\Lambda_{\overline{\rm MS}}$, as e.g.\ the uncertainties of $a$ and of $r_0 / a$.
Hence, with regard to a comparison of the results for $\Lambda_{\overline{\rm MS}}$ obtained from an analysis of the $Q\bar Q$ static potential in momentum space as opposed to a similar analysis in position space, we consider the comparison in Sec.~\ref{SEC444} -- not accounting for these additional uncertainties -- as most relevant and meaningful.
For easy reference, we collect all error contributions for our momentum space analysis in Table~\ref{TAB490}, where we also confront them to the analogous error contributions for a position space analysis \cite{Jansen:2011vv}.
\begin{table}[htb]
\begin{center}
\begin{tabular}{|c|c|c|c|c|c|}
\hline
& & & \vspace{-0.40cm} \\
error source & momentum space & position space & comments and remarks \\
& & & \vspace{-0.40cm} \\
\hline
& & & \vspace{-0.40cm} \\
\hline
& & & \vspace{-0.40cm} \\
(1) correlated & $13 \, \textrm{MeV}$ & $20 \ldots 26 \, \textrm{MeV}$ & NNLO and NNNLO \\
\cline{2-4}
& & & \vspace{-0.40cm} \\
%
systematic errors & $\phantom{0}3 \, \textrm{MeV}$ & $13 \ldots 14 \, \textrm{MeV}$ & NNNLO only \\
\hline
& & & \vspace{-0.40cm} \\
(2) statistical & $\approx 2 \ldots 4 \, \textrm{MeV}$ & $\approx 2 \, \textrm{MeV}$ & The statistical error of the lattice \\
errors & & & static potential is propagated \\
& & & through to $\Lambda_{\overline{\rm MS}}$ via jackknife. \\
\hline
& & & \vspace{-0.40cm} \\
(3) lattice & $\ll 8 \, \textrm{MeV}$ & $\ll 6 \, \textrm{MeV}$ & These values amount to rather \\
discretization errors & & & conservative upper bounds.\\
\hline
& \multicolumn{2}{|c|}{} & \vspace{-0.40cm} \\
(4) errors associated & \multicolumn{2}{|c|}{$\approx 13 \, \textrm{MeV}$} & $\approx \Lambda_{\overline{\rm MS}} \times (\Delta a / a) \approx \Lambda_{\overline{\rm MS}} \times 0.04$ \\
with the lattice & \multicolumn{2}{|c|}{} & \\
spacing & \multicolumn{2}{|c|}{} & \vspace{-0.40cm} \\
& \multicolumn{2}{|c|}{} & \\
\hline
\end{tabular}
\caption{\label{TAB490}Error contributions for a momentum space analysis (this work) confronted to those for a position space analysis (Ref.~\cite{Jansen:2011vv}).}
\end{center}
\end{table}
\newpage
\section{\label{sec:Concl}Conclusions}
We have determined $\Lambda_{\overline{\rm MS}}^{(n_f=2)}$ by fitting perturbative expressions for the $Q\bar Q$ static potential at the presently best know accuracy, i.e. up to ${\cal O}(\alpha_s^4)$ in momentum space, to lattice results obtained at a rather fine lattice spacing $a \approx 0.042 \, \textrm{fm}$.
In contrast to previous works in this direction (cf.\ e.g.\ \cite{Michael:1992nj,Gockeler:2005rv,Brambilla:2010pp,Leder:2010kz,Jansen:2011vv,Bazavov:2012ka,Fritzsch:2012wq}) we have employed a discrete Fourier transform to transform the lattice results for the $Q\bar Q$ static potential to momentum space. The extraction of $\Lambda_{\overline{\rm MS}}$ by fitting perturbative expressions of the static potential to lattice results has exclusively been performed in momentum space.
Resorting to a previous position space extraction of $\Lambda_{\overline{\rm MS}}^{(n_f=2)}$ based on the same lattice data \cite{Jansen:2011vv}, we could show that the momentum space analysis allows for a more precise determination of $\Lambda_{\overline{\rm MS}}$.
We could reduce the associated errors by a factor of $\approx 1.5 \ldots 4$, depending on the details of the fitting procedures and the estimate of the error associated with the truncation of the perturbative expressions used (cf.\ in particular Sec.~\ref{SEC444}).
This improvement can mainly be attributed to the nearly perfect stability of the results for $\Lambda_{\overline{\rm MS}}$ with respect to variations of the momentum fitting range (cf.\ Figure~\ref{FIG003}).
Such behavior has not been observed in position space, where comparably rather strong variations have been observed (cf.\ also Figures~3 and 4 of \cite{Jansen:2011vv}).
In units of the hadronic scale $r_0$ our final result for $\Lambda_{\overline{\rm MS}}$ reads
\begin{equation}
\label{eq:final_} r_0\Lambda_{\overline{\rm MS}}^{(n_f=2)} = 0.692(31) ,
\end{equation}
while in physical units it is given by
\begin{equation}
\label{eq:final} \Lambda_{\overline{\rm MS}}^{(n_f=2)} = 331(21) \, \textrm{MeV} .
\end{equation}
The quoted uncertainties include all sources of systematic error: the neglect of higher orders in the perturbative expansion, the dependence of the fit results on the fitting range $p_\textrm{min} \ldots p_\textrm{max}$ and the parameter $M$, lattice discretization errors, finite volume effects and uncertainties due to nonvanishing quark masses on the lattice. The uncertainties are dominated by the variations of $\Lambda_{\overline{\rm MS}}$, when switching from NNLO to NNNLO, and in the case of \Eqref{eq:final} by the uncertainties associated with the lattice spacing.
We note that the results (\ref{eq:final_}) and (\ref{eq:final}) compare well with other determinations of $\Lambda_{\overline{\textrm{MS}}}^{(n_f=2)}$ from the literature, e.g.\ \cite{DellaMorte:2004bc,Leder:2010kz,Gockeler:2005rv,Sternbeck:2010xu,Blossier:2010ky,Kneur:2011vi,Kneur:2013coa,Cichy:2013eoa} (cf.\ also the more detailed discussion and graphical summary in \cite{Jansen:2011vv}).
In the future it would be interesting to adopt a similar momentum space analysis to lattice results for the $Q\bar Q$ static potential with $n_f = 0$, $n_f = 2+1$ and $n_f = 2+1+1$ dynamical quark flavors.
For all these cases we expect a momentum space analysis to benefit from the better convergence behavior of the perturbative static potential in momentum space as compared to position space (cf. also the detailed discussion in the introduction, Sec.~\ref{sec:intro}), which -- as demonstrated in the present work -- can reduce the error on the extracted values of $\Lambda_{\overline{\rm MS}}$ as compared to a position space analysis of the $Q\Bar Q$ static potential.
\newpage
\section*{Acknowledgments}
We would like to thank Karl Jansen for helpful discussions, and the European Twisted Mass Collaboration for the generation of gauge link configurations used in this paper.
M.W.\ acknowledges support by the Emmy Noether Programme of the DFG (German Research Foundation), grant WA 3000/1-1.
F.K. is grateful to M. \& M.~Tingu for inspiring comments, and thankful to Karl-Heinz ``Serv$\bar{\rm u}$s'' Dubansky for providing a clean environment and Gummib\"aren.
This work was supported in part by the Helmholtz International Center for FAIR within the framework of the LOEWE program launched by the State of Hesse.
|
\section*{Reductions to a point}
Let us recall that a useful way to think about matrix models is as reductions of field theories to a single point, namely to zero
dimensions \cite{rmm1,rmm2,rmm3}.
Consider for example the
bosonic sector of maximal supersymmetric YM theory in 10 (Euclidean) dimensions. Its action is simply
\be
\int\dd^{10}x ~\sfrac 14\text{Tr}~F\w\star F~,
\ee
where
\be
F=\sfrac 12(\partial_MA_N-\partial_NA_M+i[A_M,A_N])\dd x^M\w\dd x^N~,
\ee
and the index $M$ takes values from 0 to 9. In order to perform a trivial dimensional reduction from 10 to 0 dimensions, we
must assume that the gauge field in 10 dimensions does not depend on any of them, i.e. $\partial_MA_N=0$.
Then we directly find the reduced classical bosonic action,
\be \label{ikktaction}
S_{\text{B}}=-\sfrac 14\text{Tr}~[A_M,A_N][A_{M'},A_{N'}]g^{MM'}g^{NN'}~.
\ee
This is the starting point to define the partition function that yields the IIB matrix model \cite{ikkt},
\be \label{pf1}
\mc Z=\int \prod_{M=0}^{9} \dd A_M ~\text{Pf}(A_M)~ e^{-S_{\text{B}}}~,
\ee
where the Pfaffian appears by integrating out the matter fields after the model is supersymmetrized.
Note that the components of the 1-form
$A=A_M\dd x^M$ in 10D become (Hermitian) matrices in the 0D theory, having no dependence on any spacetime coordinates, which are
anyway absent in 0 dimensions.
Of course, $A_M$ are already Hermitian matrices in 10 dimensions, since the gauge field lives in the adjoint representation
of the gauge group. The integral in Eq. (\ref{pf1}) is over those matrices. It is remarkable that in certain cases this partition
function, as well as similarly defined correlation functions, are convergent for the Euclidean model \cite{Austing:2001pk,Krauth:1998yu}.
The EOMs for the action (\ref{ikktaction}) are
\be
g^{MM'}[A_M,[A_{M'},A_N]]=0~.
\ee
Classes of classical solutions to these equations were described in many works, such as
the basic ones in Ref. \cite{ikkt} and more in Refs. \cite{Chatzistavrakidis:2011su,Steinacker:2011wb,Arnlind:2012cx}
and \cite{Kim:2011cr,Kim:2011ts,Kim:2012mw} (in the Lorentzian model).
The usual interpretation is that the matrices $A_M$ are associated to coordinates and therefore the solutions correspond to
noncommutative spacetimes.
This is fine, although the origin of the matrices is in the cotangent bundle and
they naturally carry a lower index.
This remark implies that the matrices $A_M$ could also be associated to momenta and generate the momentum space
instead of spacetime. A relevant discussion on this may be found in Ref. \cite{Aoki:1999vr}. However, there is no clear way to
obtain the full structure of phase space from the IIB model.
On the other hand, the momenta in
matrix noncommutative geometry are typically related to the coordinates, since they correspond to inner derivations of the algebra
$\mc A$ of coordinate operators \cite{Madorebook}. Moreover, they involve two copies of $\mc A$, say $\mc A_L$ and $\mc A_R$, that correspond to the left and the right action
of the operators respectively \cite{Chatzistavrakidis:2014tda}.
The momenta are then related to the difference $\hat x_L-\hat x_R$ of coordinate operators in the
two representations. All these suggest that there should exist an extended model which is associated to the dynamics of phase space.
This is desirable for the reasons explained in the introduction, primarily for a better understanding of the gravitational
field in the framework of matrix models.
\section*{YM theories and Courant algebroids}
In order to construct the extended matrix model, we need some elementary concepts from generalized complex geometry
\cite{Hitchin:2004ut,gg} and the theory of
Courant algebroids \cite{wein}. The reader who is interested in the model itself may jump to the next section.
Consider the generalized tangent bundle of a manifold $\text{M}$ of dimension $d$ {\footnote{We often set $d$=10 in the
following, although the discussion is general and holds for any $d$.}}, which is given by
the sum of the tangent and cotangent ones,
$
\mc T\text{M}=\text{TM}\oplus\text{T}^{\star}\text{M}~.
$
The sections $\G(\mc T\text{M})$ of this bundle are
generalized vectors $\mathfrak{X}$, which can be written as the sum of an 1-vector
and an 1-form,
\be
\mathfrak{X}=X+\eta~,\quad X\in \G(\text{TM})~,\quad \eta\in\G(\text{T}^{\star}\text{M})~.\nn
\ee
The standard Courant algebroid is obtained by equipping the above bundle with the Courant bracket \cite{dirac},
\bea
\label{cour}[\mathfrak X,\mathfrak Y]_C=[X,Y]_{L}+{\cal L}_{X}\xi-{\cal L}_Y\eta-\sfrac 12\dd(X(\xi)-Y(\eta))~,\nn
\eea
a pairing,
\bea
\langle \mathfrak X,\mathfrak Y\rangle=\frac{1}{2}(X(\xi)+Y(\eta))~,
\eea
and a smooth map,
$
\rho:{\mc T}\text{M}\to\text{TM}
$,
the anchor.
A notion with particular interest for physics is that of Dirac structures \cite{dirac}. These are vector subbundles
$L\subset{\mc T}\text{M}$ of the generalized
tangent bundle such that
\bea
\langle\mathfrak X_L,\mathfrak Y_L\rangle=0~,\quad
[\mathfrak X_L,\mathfrak Y_L]\in \G(L)~, \nn
\eea
for any $\mathfrak X_L,\mathfrak Y_L\in \G(L)$. The rank of these bundles is exactly half of the rank of $\mc T\text{M}$.
Dirac structures are valuable for physical problems because arbitrary elements of
$\w^{\bullet}\mc T\text{M}$ do not generically transform as tensors, however elements of $\w^{\bullet}L$ do \cite{Gualtieri:2007bq}.
Moreover, the Courant bracket satisfies the Jacobi identity when restricted on a Dirac structure, although it does not satisfy it
on the generalized tangent bundle.
On a vector bundle, a generalized notion of a connection can be defined \cite{Gualtieri:2007bq}. Here we consider just
the simplest possibility,
\be \label{connection}
\mc D=\dd +A+V=\dd x^M\partial_M +A_M\dd x^M+V^M\partial_M~,
\ee
on the vector bundle $\mc T\text{M}$.
The curvature of a generalized connection is defined in a way that directly generalizes the usual definition,
\be
\label{gcurvgen}
\mc F(\mathfrak X,\mathfrak Y)=[\mc D_{\mathfrak X},\mc D_{\mathfrak Y}]-\mc D_{[\mathfrak X,\mathfrak Y]}~.
\ee
For the connection (\ref{connection}) this field
strength is
\bea
\mc F=&&\sfrac 12F_{MN}\dd x^M\w\dd x^N\nn\\ &&+(\partial_MV^N+i[A_M,V^N])\dd x^M\w\partial_N \nn\\
&&+\sfrac i2[V^M,V^N]\partial_M\w\partial_N~,
\eea
where the bracket is just the Lie algebra commutator associated to the gauge group.
Next we consider the volume form on the generalized tangent bundle. This is given as
\bea
\text{vol}_{\mc T\text{M}}&=&\pm\dd
x^0\w\dots\w\dd x^9\w\partial_{0}\w\dots\w\partial_{9}~,
\eea
where the choice of sign is a choice of orientation.
We choose the plus sign, which fixes the ordering of basis 1-forms
and 1-vectors.
Note that the metric does not enter, or rather the individual metric factors from
the tangent and the cotangent bundle cancel
each other. This becomes clear when the generalized metric
\be
{\cal H}=\begin{pmatrix}
g-bg^{-1}b & bg^{-1} \\
-g^{-1}b & g^{-1}
\end{pmatrix}
\ee
is considered, where $g$ is a Riemannian metric on M and $b$ is a 2-form. This generalized metric transforms covariantly under O($d,d$) transformations
$\cal O$,
\be \label{gmtrafo}
{\cal H} \quad \to \quad {\cal O}^{T}{\cal H}{\cal O}~.
\ee
Its inverse is
\be
{\cal H}^{-1}=\begin{pmatrix}
g^{-1} & -g^{-1}b \\
bg^{-1} & g-bg^{-1}b
\end{pmatrix}~,
\ee
and its determinant is
$
\text{det}~{\cal H}=1,
$
thus it drops out from any relevant formula.
In order to construct a YM theory, we need a Hodge star operator on the $\mc T\text{M}$. This acts as
\bea
\star_{\tb}:\w^p\text{TM}\w^q\text{T}^{\star}\text{M}\to \w^{d-p}\text{TM}\w^{d-q}\text{T}^{\star}\text{M}~,
\eea
and we define it such that
$\star_{\tb}\one=\text{vol}_{\tb}$.
Applying this operation to the generalized curvature $\mc F$, we are able to compute the product $\mc F\w\star_{\tb}\mc F$ and we obtain
\bea
\mc F\w\star_{\tb}\mc F&=&\big({\cal H}^{MM'}{\cal H}^{NN'}{\cal F}_{MN}{\cal F}_{M'N'}\big)\text{vol}_{\tb}~.\nn
\eea
The reader should be cautious with the exhibited index structure of the
generalized metric, which is purely conventional
since its components
have both upper and lower indices.
The expression in the parentheses can be identified with an inner product $(\mc F,\mc F)$, so that
\be
\mc F\w\star_{\tb}\mc F=(\mc F,\mc F) \text{vol}_{\tb}~.
\ee
The issue with this expression and the problem one faces in the corresponding generalized YM theory, is that the
generalized curvature ${\cal F}$ does not transform as a tensor at the level of the Courant algebroid \cite{Gualtieri:2007bq}. This can be overcome by defining the theory on Dirac structures, where ${\cal F}$ transforms tensorially. This was done and examined in Ref. \cite{ChaGau}. Here we adopt a different point of view. In particular, we overcome the above problem by projecting the theory to zero dimensions, thus defining
a matrix model, where harmful derivatives are dropped and the welcome transformation properties are restored.
\section*{The SO(10,10) matrix model and its symmetries}
Let us first examine how the matrix model with action (\ref{ikktaction}) is obtained in this formalism. This can be approached in two
ways. The first way is to trace the steps that led to the type IIB matrix model. Considering the YM theory on the Dirac structure $L=\text{T}\text{M}$ of
the full Courant algebroid
and setting $b=0$, the corresponding generalized YM theory is identical
to the standard YM in 10D and the model follows from its dimensional reduction, as previously.
Alternatively, one can consider instead the Dirac structure $L=\text{T}^{\star}\text{M}$ and the generalized YM
theory on it. In order to reach a 0D theory, we use the technique of Refs.
\cite{Ellwood:2006ya,Mylonas:2012pg}, also used in Ref. \cite{ChaGau}, where a map to momentum space was introduced.
Integrating out the volume of this
momentum space we obtain the action
\be \label{ikktalt}
S'_{\text{B}}=-\sfrac 14\text{Tr}~g_{MM'}g_{NN'}[V^M,V^N][V^{M'},V^{N'}]~.
\ee
This is equivalent to the action that appears in Eq. (\ref{ikktaction}) upon the identification $A_M=g_{MM'}V^{M'}$, and it
has the same classical solutions. It is a dual model that describes the same physics. However, the two actions were obtained from two very special but different Dirac structures. Here we show that
a more general model is obtained when we utilize the full structure of $\tb$, which has solutions that are not captured by the IIB
matrix model.
Consider the full generalized YM theory described in the previous section and its trivial reduction to a point. In the present case
the 2-form $b$ is not dropped.
The result is a reduced model with bosonic action
\bea \label{action}
S&=&-\sfrac 14\text{Tr}~\biggl(\tilde{g}_{MM'}\tilde{g}_{NN'}[V^M,V^N][V^{M'},V^{N'}]\nn\\
&&+g^{MM'}g^{NN'}[A_M,A_N][A_{M'},A_{N'}]\nn\\
&&+2~g^{MM'}\tilde{g}_{NN'}[A_M,V^N][A_{M'},V^{N'}]\nn\\
&&-2g^{MP}g^{M'Q}b_{QN}b_{PN'}[A_M,V^N][A_{M'},V^{N'}]\nn\\
&&+2g^{MP}g^{NQ}b_{PM'}b_{QN'}[A_{M},A_{N}][V^{M'},V^{N'}]\nn\\
&&+4g^{MM'}g^{NP}b_{N'P}[A_{M},A_N][A_{M'},V^{N'}]\nn\\
&&+4g^{MP}\tilde g_{NN'}b_{M'P}[A_M,V^{N}][V^{M'},V^{N'}]\biggl)
~,
\eea
where we defined
$
\tilde g=g-bg^{-1}b.
$
It should be clear that the dynamical degrees of freedom are the $A_M$ and $V^M$, while $g$ and $b$ are related to the geometry
of the embedding space and they are not dynamical.
Note that due to the terms that appear after the first two lines,
the model is more than a simple addition of the two dual actions for the IIB model.
Recalling the origin of the action (\ref{action}), its terms can be
collected accordingly. First, noting the symmetric role of $A_M$ and $V^M$, it is useful to define the extended matrix
\be \label{collectX}
X_M=\begin{pmatrix}
A_M \\ V^M
\end{pmatrix}~,
\ee
where once more the position of its index is conventional and has nothing to do with its transformation properties.
Then, the action can be cast into the following simple form:
\be \label{collectaction}
S=-\sfrac 14\text{Tr}~{\cal H}^{MM'}{\cal H}^{NN'}[X_M,X_N][X_{M'},X_{N'}]~.
\ee
A subtle point is that the bracket in
Eq. (\ref{collectaction}) is not precisely a commutator, since the $X_M$ are not square matrices, unlike $A_M$ and $V^M$.
Its actual definition is
\be
[X_M,X_N]:=\begin{pmatrix}
[A_M,A_N] & [A_M,V^N] \\
[V^M,A_N] & [V^M,V^N]
\end{pmatrix}~.
\ee
The action (\ref{action}), or equivalently (\ref{collectaction}), leads to two sets of EOMs.
Varying with respect to $A_M$ or $V^M$ independently, these are
\bea
\square A_M=0~,\quad
\square V^M=0~,
\eea
where we defined the box operator
\bea
\square\cdot&=&g^{MM'}[A_M,[A_{M'},\cdot]]+\tilde{g}_{MM'}[V^M,[V^{M'},\cdot]]\nn\\&&+g^{MP}b_{M'P}\big([A_M,[V^{M'},\cdot]]
+[V^{M'},[A_{M},\cdot]]\big)~.\nn
\eea
Note that these equations already appear coupled when one varies with respect to $A_M$ or $V^M$ alone.
We are going to discuss some benchmark classical solutions in the next section.
The bosonic model with action (\ref{action}) exhibits a number of symmetries. First of all, it has the obvious translational
symmetries $A_M\to A_M+c_M\one_{d}$ and $V^M\to V^M+c^M\one_{d}$, with $c_M, c^M\in \R$, which is an extension of the analogous property of the IIB model.
Moreover, it has the gauge symmetry
$X_M\to UX_MU^{-1}$, with $U\in U(N)$, ${N}$ being the size of the matrices ($N\to\infty$, as usual for large-$N$
models). This is again the same as in the IIB model and it reflects the fact that the extended set
of degrees of freedom originate from the same 10D generalized YM theory. Finally, there is a global rotational symmetry.
Recall that the Euclidean IIB model has such a symmetry too, but it is SO(10). Here we encounter the main difference, in that the model
(\ref{action}) exhibits a SO(10,10) global symmetry.
This can be directly verified by performing SO(10,10) transformations in the action (\ref{action}),
keeping in mind that aside $A_M$ and $V^M$, $g$ and $b$ transform too. Their transformation is
determined via the corresponding transformation of the generalized metric, given in Eq. (\ref{gmtrafo}).
The model also possesses a symmetry that is not present in the IIB model, which exchanges $A_M$ and $V^M$ as
\be
A^M\quad \to \quad V_M\quad \text{and}\quad V_M\quad \to \quad -A^M~.\label{xp}
\ee
We will comment on
this symmetry after we present some basic classical solutions.
\section*{Dynamical phase space}
One of the prime attractive features of the IIB matrix model is that it addresses the issue of the emergence of spacetime and its dynamics
(see e.g. Ref. \cite{Aoki:1998vn} and Refs. \cite{Steinacker:2010rh,Nishimura:2012xs} for reviews on some recent approaches).
The model that we defined in the previous section is similarly the appropriate arena to study the emergence and the dynamics of
phase space, which is valuable for the reasons explained in the introduction.
Let us search for solutions of the classical EOMs of the model. In order to simplify our analysis, we
consider $b=0$ {\footnote{This has the effect of the global symmetry of the model being just SO($d$)$\times$SO($d$).}}. The general case of $b\ne 0$ is very rich and interesting and we are going to report on this is the
future. The EOMs simply become
\bea
g^{MM'}[A_M,[A_{M'},A_N]]+g_{MM'}[V^M,[V^{M'},A_N]]&=&0~,\nn\\
g_{MM'}[V^M,[V^{M'},V^N]]+g^{MM'}[A_M,[A_{M'},V^N]]&=&0~.\nn
\eea
Consider the following vacuum ansatz:
\be \label{ansatz}
A_a=\hat p_a~,\quad V^a=\hat x^a~,\quad a=1,\dots,2m,~ 2m\le d~,
\ee
where $\hat x^a$ and $\hat p_a$ are to be identified with position and momentum operators, and
$A_{2m+1}=\dots=A_d=V^{2m+1}=\dots=V^d=0$. They satisfy the
canonical commutation relations (CCR)
\be
[\hat x^a,\hat p_b]=i\hbar\d^a_b~.
\ee
Then the EOMs are simplified to
\bea
[\hat p_a,[\hat p_{a},\hat p_b]]=0\quad \text{and} \quad
[\hat x^a,[\hat x^{a},\hat x^b]]=0~,\label{eomsimple}
\eea
which look very simple but actually include rather rich structures.
We split the rest of our analysis into two parts. The first part is rather degenerate, it refers to flat spacetimes and phase spaces, and
most of its features are essentially captured already by the IIB matrix model. It simply includes the algebra
\be
[\hat x^a,\hat x^b]=i\theta^{ab}~,\quad [\hat p_a,\hat p_b]=i\o_{ab}~,
\ee
with $\theta^{ab}$ and $\o_{ab}$ constant parameters, plus the CCR. This algebra is the one of noncommutative quantum mechanics
with a constant magnetic source \cite{Duval:2000xr,Morariu:2001dv}.
The second and more interesting class of solutions contains a subset of noncommutative phases spaces recently described in Ref. \cite{Chatzistavrakidis:2014tda}. These are phase spaces of noncommutative manifolds, whose underlying
commutative counterparts are general symplectic manifolds which are parallelizable,
i.e. they admit a global section of their tangent bundle, and they are not necessarily flat. It was shown in Ref. \cite{Chatzistavrakidis:2014tda} that in
such cases it is necessary to consider two copies of the noncommutative algebra ${\cal A}$ of position operators, one acting
from the left and denoted ${\cal A}_L$ with elements $\hat x_L^a$ and one acting from the right, denoted as ${\cal A}_R$ and
generated by $\hat x^a_R$. The two sets are commuting, namely $[\hat x^a_L,\hat x^b_R]=0$, and they are symplectic dual with respect to the symplectic 2-vector $\theta^{ab}$, i.e. $[\hat x^a_L,\hat x^b_L]=-[\hat x^a_R,\hat x^b_R]=i\theta^{ab}$.
In relation to the vacuum ansatz (\ref{ansatz}) fot the matrix model, $V^a$ are identified with $\hat x^a_L$, while $\hat x^a_R$
do not appear explicitly in the model but only indirectly as we immediately explain.
Recall that in the flat case, the momentum operators act as
\be
\hat p_a=\hbar\o_{ab}(\hat x^b_L-\hat x^b_R)~,
\ee
$\o_{ab}$ being the symplectic 2-form, and they are inner operators in the algebra $\mc A$.
However, when the manifold is not flat these operators do not correspond to the translations generated by invariant
vector fields.
In that case the correct momentum operators are
\be
\hat p_i=e_i^{\ a}(\hat x_R)\hat p_a~,
\ee
and this translates in the vacuum ansatz of Eq. (\ref{ansatz}) to $A_a=e_a^{\ i}\hat p_i$.
The important aspect in this formulation is that the momenta contain the non-constant frame $e_a^{\ i}$, which is
associated to the
gravitational field. In particular, the general form of the algebra of the operators $\hat x^a$ and $\hat p_i$ turns out to be
\bea
[\hat x^a_L,\hat x^b_L]&=&-[\hat x^a_R,\hat x^b_R]=i\theta^{ab}~,\nn\\
{[}\hat x^a_L,\hat p_i]&=&i\hbar e^a_{\ i}~,\nn\\
{[}\hat x^a_R,\hat p_i]&=&i\hbar e^a_{\ i}-e^k_{\ b}K^{ba}_i\hat p_k~,\nn\\
{[}\hat p_i,\hat p_j]&=&M_{ij}+N_{ij}^{\ k}\hat p_k+P_{ij}^{kl}\hat p_k\hat p_l~, \label{psa}
\eea
with exactly computable coefficients in terms of the frame and the symplectic structure, such that
all the Jacobi identities are satisfied \cite{Chatzistavrakidis:2014tda}.
We observe that the gravitational field is identified with the commutation relation among the position and momentum operators,
as in Refs. \cite{Madorebook,Buric:2006di,Buric:2011dd}.
When the geometric data are identified with that of symplectic nilmanifolds in dimensions 4 and 6, the set of relations
(\ref{psa}), along with the identifications $A_a=e_a^{\ i}\hat p_i$ and $V^a=\hat x^a_L$, provides many non-trivial solutions to the equations (\ref{eomsimple}) of the model, which are not captured by the IIB matrix model. A more direct way to see this, is to
consider the matrix model and its EOMs this time with a non-coordinate index structure.
This happens when the starting point is a generalized coonection of the form
\be
{\cal D}=(\theta_I+A_I) e^I+V^I\theta_I~,
\ee
where $e^I$ and $\theta_I$ are the 1-forms and 1-vectors of the non-coordinate basis respectively. The general form of the matrix model and its EOMs remains the same in this basis, but now they are written in terms of $A_I$ and $V^I$. The Ansatz for solutions now is
\be \label{ansatz2}
A_i=\hat p_i~,\quad V^i=\d^i_{\ a}\hat x^a_L~,\quad i=1,\dots,2m,~ 2m\le d~.
\ee
The EOMs in this basis become:
\bea
[\hat p_i,[\hat p_i,\hat p_j]]+[\hat x^a,[\hat x^a,\hat p_j]]&=&0~,\label{ceom1}\\
{[}\hat x^a,[\hat x^a,\hat x^b]]+[\hat p_i,[\hat p_i,\hat x^b]]&=&0~.\label{ceom2}
\eea
Assuming the phase space algebra (\ref{psa}) with constant parameters $\theta^{ab}$, we immediately obtain
\bea
[\hat x^a,[\hat x^a,\hat p_j]]&=&[\hat x^a,i\hbar e^a_{\ j}]=0~, \nn\\ {[}\hat x^a,[\hat x^a,\hat x^b]]&=&[\hat x^a,\theta^{ab}]=0~,\nn
\eea
where in the first equation we used the commutativity of ${\cal A}_L$ and ${\cal A}_R$. Then, a direct computation shows that the Eqs.
(\ref{ceom1}) and (\ref{ceom2}) result in the conditions:
\begin{align}
&N_{ij}^{\ l}M_{il}+(N_{ij}^{\ l}N_{il}^{\ m}+2P_{ij}^{\ lm}M_{il})\hat p_m+\nn\\
&~~+(N_{ij}^{\ l}P_{il}^{\ mn}+2P_{ij}^{\ lm}N_{il}^{n})\hat p_m\hat p_n+\nn\\
&~~+2P_{ij}^{\ lm}P_{il}^{\ nr}\hat p_n\hat p_r\hat p_m=0~,\label{cond1}\\
&{[}\hat p_i,e^a_{\ i}]=0~.\label{cond2}
\end{align}
Now it is time to specify a class of particular cases with their parameters. For step 2 nilmanifolds in 4 and 6 dimensions, it
was shown in Ref. \cite{Chatzistavrakidis:2014tda} that
\be
M_{ij}=0,~ N_{ij}^{\ k}\propto f^k_{\ ij},~ P_{ij}^{\ kl}\propto f^k_{\ [i\underline{c}}f^l_{\ j]d}\theta^{cd}~,
\ee
while
\be
e^a_{\ i}= \d^a_{\ i}-\sfrac 12f^a_{\ ib}\hat x_R^b~,
\ee
where $f^k_{\ ij}$ are the structure constants of the nilpotent Lie algebra that is associated to the nilmanifold.
Then, simply using the defining relation $f^k_{\ ij}f^i_{\ lm}=0$ (no summation) for step 2 nilmanifolds, the conditions
(\ref{cond1}) and (\ref{cond2}) are satisfied.
A full classification of solutions, including $b\ne 0$ too, is an open issue which should be addressed in detail.
We close this section by observing that the symmetry (\ref{xp}) of the matrix model translates into
\be
\hat x^a \quad \to \quad \hat p_a\quad \text{and} \quad \hat p_a\quad \to \quad -\hat x^a~,
\ee
which is familiar in quantum-mechanical phase space, and its role in matrix models was already emphasized in Ref. \cite{Chatzistavrakidis:2012qj}.
\section*{Remarks on quantization}
Quantization in matrix models is defined via matrix integrals. For the SO(10,10) matrix model the partition function is defined as
\be
{\cal Z}=\int \prod_{M=0}^{9} \dd A_M\prod_{N=0}^{9} \dd V^N ~ e^{-S}~,
\ee
where $S$ is given by Eq. (\ref{action}). Correlation functions may be defined similarly.
A primary question is whether these integrals are convergent
under certain conditions. This is a technical issue which presents an interesting challenge. However, given that when $V^N$ vanish
the corresponding integrals are convergent for certain number of dimensions (including 10) and
certain gauge groups \cite{Austing:2001pk,Krauth:1998yu}, it is reasonable to expect that a careful evaluation will
reveal such cases for the extended model too. This will be addressed in future work.
\section*{Conclusions}
In the present work we argued that a better understanding of the dynamics of full phase space, rather than just spacetime,
can be relevant for physics at the Planck scale and ultimately for quantum gravity. Similar ideas were already emphasized
before \cite{Freidel:2014qna,Madorebook}. Here we constructed a theory that captures the dynamics of phase space.
It is given by a matrix model which extends in a consistent way previous matrix models that proved to be
successful in the description of spacetime dynamics \cite{Banks:1996vh,ikkt}. The model is derived from the trivial dimensional
reduction of a generalized Yang-Mills theory on a Courant algebroid to zero dimensions. This allows us to overcome the problem of
the nontensorial transformation of generalized fields on the Courant algebroid. The symmetries of the model include and
extend the ones
of the IIB model. Notably there is a global SO($d,d$) symmetry, as well as a quantum-mechanical symmetry that is interpreted as
exchange of positions and momenta in phase space. Certain noncommutative phase spaces that correspond to curved manifolds
are classical solutions of the EOMs. The key feature is that the commutator of positions and momenta can be associated to the gravitational field, and therefore (semiclassical) gravity naturally emerges on solutions of the model. Furthermore, quantization is in principle
possible, with the partition function and correlation functions defined via matrix integrals. Whether these integrals are
convergent remains an open issue which should be carefully addressed.
\paragraph*{Acknowledgments.~} The author is indebted to F. F. Gautason and L. Jonke for reading the manuscript and making many interesting
comments, as well as to H. Steinacker for useful remarks. This work was completed in the Simons Center for Geometry and Physics during the 2014 Simons Summer Workshop. The author thanks the organisers for the
excellent working environment and the center for financial support.
|
\section{Introduction}
\subsection{Tidal coupling and binary inspiral}
Tidal coupling in binary inspiral has been a topic of much recent
interest. A great deal of attention has focused in particular on
systems which contain neutron stars, where tides and their
backreaction on the binary's evolution may allow a new probe of the
equation of state of neutron star matter {\cite{rmsucf2009, hllr2010,
dnv2012}}. A great deal of work has been done to rigorously
define the distortion of fluid stars {\cite{dn2009, bp2009}}, the
coupling of the tidal distortion to the binary's orbital energy and
angular momentum {\cite{bdgnr2010}}, and most recently the importance
of nonlinear fluid modes which can be sourced by tidal fields
{\cite{wab2013, vzh2014}}.
Tidal coupling also plays a role in the evolution of binary black
holes. Indeed, the influence of tidal coupling on binary black holes
has been studied in some detail over the past two decades, but using
rather different language: instead of ``tidal coupling,'' past
literature typically discusses gravitational radiation ``down the
horizon.'' This down-horizon radiation has a dual description in the
tidal deformation of the black hole's event horizon. A major purpose
of this paper is to explore this dual description, examining
quantitatively how a black hole is deformed by an orbiting companion.
Consider the down-horizon radiation picture first. The wave equation
governing radiation produced in a black hole spacetime admits two
solutions {\cite{teuk73,tp74}}, one describing outgoing radiation very
far from the hole, and another describing radiation ingoing on the
event horizon. Both solutions carry energy and angular momentum away
from the binary, and drive (on average) a secular inspiral of the
orbit. After suitable averaging, we require (for example) the orbital
energy $E_{\rm orb}$ to evolve according to
\begin{equation}
\frac{dE_{\rm orb}}{dt} = -\left(\frac{dE}{dt}\right)^\infty -
\left(\frac{dE}{dt}\right)^{\rm H}\;,
\end{equation}
where $(dE/dt)^\infty$ describes energy carried far away by the waves,
and $(dE/dt)^{\rm H}$ describes energy carried into the event horizon.
The down-horizon flux has an interesting property. When it is
computed for a small body that is in a circular, equatorial orbit of a
Kerr black hole with mass $M$ and spin parameter $a$, we find that
\begin{equation}
\left(\frac{dE}{dt}\right)^{\rm H} \propto \left(\Omega_{\rm orb} -
\Omega_{\rm H}\right)\;,
\label{eq:horizonflux_prop}
\end{equation}
where $\Omega_{\rm orb} = M^{1/2}/(r^{3/2} + aM^{1/2})$ is the orbital
frequency\footnote{Throughout this paper, we use units with $G = 1 =
c$.}, and $\Omega_{\rm H} = a/2Mr_+$ is the hole's spin frequency
(Ref.\ {\cite{membrane}}, Sec VIID; see also synopsis in
Sec.\ {\ref{sec:downhoriz}}). The radius $r_+ = M + \sqrt{M^2 - a^2}$
gives the location of the event horizon in Boyer-Lindquist
coordinates. We assume that the orbit is prograde, so that the
orbital angular momentum is parallel to the hole's spin angular
momentum.
When $\Omega_{\rm orb} > \Omega_{\rm H}$ (i.e., when the orbit rotates
faster than the black hole spins), we have $(dE/dt)^{\rm H} > 0$ ---
radiation carries energy into the horizon, taking it from the orbital
energy. This is intuitively sensible, given that an event horizon
generally acts as a sink for energy and matter. However, when
$\Omega_{\rm orb} < \Omega_{\rm H}$ (the hole spins faster than the
orbit's rotation), we have $(dE/dt)^{\rm H} < 0$. This means that the
down-horizon component of the radiation {\it augments} the orbital
energy --- energy is transferred from the hole to the orbit. This is
far more difficult to reconcile with the behavior of an event horizon.
One clue to understanding this behavior is that, when $\Omega_{\rm H}
> \Omega_{\rm orb}$, the modes which contribute to the radiation are
{\it superradiant} {\cite{pt73,chandra}}. Consider a plane wave which
propagates toward the black hole. A portion of the wave is absorbed
by the black hole (changing its mass and spin), and a portion is
scattered back out to large radius. A superradiant mode (see, for
example, Sec.\ 98 of Ref.\ {\cite{chandra}}) is one in which the
scattered wave has higher amplitude than the original ingoing wave.
Some of the black hole's spin angular momentum and rotational energy
has been transferred to the radiation.
\subsection{Tidally distorted strong gravity objects}
Although the condition for superradiance is the same as the condition
under which an orbit gains energy from the black hole, superradiance
does not explain how energy is transferred from the hole to the orbit.
A more satisfying picture of this can be built by invoking the dual
picture of a tidal distortion. As originally shown by Hartle
{\cite{hartle73,hartle74}}, an event horizon's intrinsic curvature is
distorted by a tidal perturbation. In analogy with tidal coupling in
fluid systems, the tidally distorted horizon can gravitationally
couple to the orbiting body, transfering energy and angular momentum
from the black hole to the orbit.
Let us examine the fluid analogy in more detail for a moment.
Consider in particular a moon that raises a tide on a fluid body,
distorting its shape from spherical to a prolate ellipsoid. The tidal
response will produce a bulge that tends to point at the moon. Due to
the fluid's viscosity, the bulging response will lag the driving tidal
force. As a consequence, if the moon's orbit is faster than the
body's spin, then the bulge will lag behind. The bulge will exert a
torque on the orbit that tends to slow down the orbit; the orbit
exerts a torque that tends to speed up the body's spin. Conversely,
if the spin is faster than the orbit, the bulge will lead the moon's
position, and the torque upon the orbit will tend to speed it up (and
torque from the orbit tends to slow down the spin). In both cases,
the bulge and moon exert torques on one another in such a way that the
spin and orbit frequencies tend to be equalized\footnote{This is why
our Moon keeps the same face to the Earth: Tidal coupling has spun
down the Moon's ``day'' to match its ``year.'' Tidal forces from
the Moon likewise slow down the Earth's spin, lengthening the day at
a rate of a few milliseconds per century {\cite{dickeyetal1994}}.
Given enough time, this effect would drive the Earth to keep the
same face to the Moon.}. The action of this torque is such that
energy is taken out of the moon's orbit if the orbit frequency is
larger than the spin frequency, and vice versa.
Since a black hole's shape is changed by tidal forces in a manner
similar to the change in shape of a fluid body, one can imagine that
the horizon's tidal bulge likewise exerts a torque on an orbit.
Examining Eq.\ (\ref{eq:horizonflux_prop}), we see that the sign of
the ``horizon flux'' energy loss is exactly in accord with the tidal
fluid analogy --- energy is lost from the orbit if the orbital
frequency exceeds the black hole's spin frequency, and vice versa.
Using the membrane paradigm {\cite{membrane}}, one can assign a
viscosity to the horizon, making the fluid analogy even more
compelling.
However, as was first noted by Hartle {\cite{hartle73}}, the geometry
of a black hole's tidal bulge behaves in a rather counterintuitive
manner. At least using a weak-field, slow spin analysis, the bulge
{\it leads} the orbit when $\Omega_{\rm orb} > \Omega_{\rm H}$, and
{\it lags} when $\Omega_{\rm orb} < \Omega_{\rm H}$. This is opposite
to the geometry which the fluid analogy would lead us to expect. This
is because an event horizon is a teleological object: Whether an event
in spacetime is inside or outside a horizon depends on that event's
null future. At some moment in a given time slicing, an event horizon
arranges itself in anticipation of the gravitational stresses it will
be feeling in the future. This is closely related to the manner in
which the event horizon of a spherical black hole expands outward when
a spherical shell falls into it. See Ref.\ {\cite{membrane}},
Sec.\ VI\,C\,6 for further discussion.
Much of this background has been extensively discussed in past
literature
{\cite{hartle73,hartle74,membrane,tp08,dl2009,bp2009,pv2010,vpm11}}.
Recent work on this problem has examined in detail how one can
quantify the tidal distortion of a black hole, demonstrating that the
``gravitational Love numbers'' which characterize the distortion of
fluid bodies vanish for non-rotating black holes {\cite{bp2009}}, but
that the geometry's distortion can nonetheless be quantified assuming
particularly useful coordinate systems {\cite{dl2009,pv2010}} and in a
fully covariant manner {\cite{vpm11}}. Indeed, one can define
``surficial Love numbers,'' which quantify the distortion of a body's
surface, for Schwarzschild black holes {\cite{lp14}}. These
techniques have been used to study horizon distortion in the
Schwarzschild and slow spin limits, and for slow orbital velocities
{\cite{fl05,dl2009,vpm11}}.
\subsection{Our analysis: Strong-field, rapid spin tidal distortions}
The primary goal of this paper is to develop tools to explore the
distorted geometry of a black hole in a binary which are good for fast
motion, strong field orbits. We use techniques originally developed
by Hartle {\cite{hartle74}} to compute the Ricci scalar curvature
$R_{\rm H}$ associated with the 2-surface of the distorted horizon;
this is closely related to the intrinsic horizon metric developed in
Ref.\ {\cite{vpm11}}. We will restrict our binaries to large mass
ratios in order to use the tools of black hole perturbation theory.
We also develop tools to embed the horizon in a 3-dimensional space in
order to visualize the tidal distortions. In this paper, we restrict
our embeddings to black hole spins $a/M \le \sqrt{3}/2$. This is the
largest spin at which the horizon can be embedded in a global
Euclidean space; black holes with spins in the range $\sqrt{3}/2 < a/M
\le 1$ must either be embedded in a space that is partially Euclidean,
partially Lorentzian {\cite{smarr}}, or be embedded in another space
altogether {\cite{frolov,gibbons}}. Although no issue of principle
prevents us for examining larger spins, it does not add very much to
the physics we wish to study here, so we defer embeddings for $a/M >
\sqrt{3}/2$ to a later paper.
A secondary goal of this paper is to investigate whether there is a
simple connection between the geometry of the tidal bulge and the
orbit's evolution. In particular, we wish to see if the sign of
$dE^{\rm H}/dt$, which is determined by $\Omega_{\rm orb} -
\Omega_{\rm H}$, is connected to the bulge's geometry relative to the
orbit. This turns out to be somewhat tricky to investigate. The
orbit and the horizon are at different locations, so we must map the
orbit's position onto the horizon. There is no unique way to do
this\footnote{Indeed, the behavior of the map depends on the gauge
used for the calculation, and the time slicing that is used, neither
of which we investigate in this paper.}, so the results depend at
least in part on how we make the map. We present two maps from orbit
to horizon. One, based on ingoing zero-angular momentum light rays,
is useful for comparing with past literature. The other, based on the
geometry of the horizon's embedding and the orbit at an instant of
constant ingoing time, is useful for describing our numerical data (at
least for small spin). Another way to characterize the bulge geometry
is to examine the relative phase of the bulge's curvature to the tidal
field which distorts the black hole. Both of these quantities are
defined at $r = r_+$, so no mapping is necessary.
We find that, at the extremes, the response of a black hole to a
perturbing tide follows Newtonian logic (modulo a swap of ``lag'' and
``lead,'' thanks to the horizon's teleological nature). In
particular, when $\Omega_{\rm orb} \gg \Omega_{\rm H}$ (so that
$dE^{\rm H}/dt > 0$), the bulge leads the orbit, no matter how we
compare the bulge to the orbit. When $\Omega_{\rm orb} \ll
\Omega_{\rm H}$ ($dE^{\rm H}/dt < 0$), the bulge lags the orbit.
However, relations between lag, lead, and $dE^{\rm H}/dt$ are not so
clear cut when $\Omega_{\rm orb} \sim \Omega_{\rm H}$. Consider, in
particular the case $\Omega_{\rm orb} = \Omega_{\rm H}$, for which
$dE^{\rm H}/dt = 0$. For Newtonian, fluid bodies, the tidal bulge
points directly at the orbiting body in this case, with no exchange of
torque between the body and the orbit. For black holes, we find no
particular relation between the horizon's bulge and the orbit's
position. The relation between tidal coupling and tidal distortion is
far more complicated in black hole systems than it is for fluid bodies
in Newtonian gravity --- which is not especially surprising.
Soon after we submitted this paper and posted a preprint to the arXiv,
Cabero and Krishnan posted an analysis of tidally deformed spinning
black holes {\cite{ck14}}. Although their techniques and analysis
differ quite a bit from ours (focusing on the Bowen-York {\cite{by80}}
initial data set, and using the framework of isolated horizons), their
results seem broadly consistent with ours. It may be useful in future
work to explore this apparent consistency more closely, and to borrow
some of the tools that they have developed for the systems that we
analyze here.
\subsection{Outline of this paper, units, and conventions}
The remainder of this paper is organized as follows. Our formalism
for computing the geometry of distorted Kerr black holes is given in
Sec.\ {\ref{sec:formalism}}. We show how to compute the curvature of
a tidally distorted black hole, and how to quantify the relation of
the geometry of this distortion to the geometry of the orbit which
produces the tidal field. We also discuss how to compute $dE^{\rm
H}/dt$, demonstrating that the information which determines this
down-horizon flux is identical to the information which determines the
geometry of the distorted event horizon.
Sections {\ref{sec:schw_results}} and {\ref{sec:kerr_results}} present
results for Schwarzschild and Kerr, respectively. In both sections,
we first look at the black hole's curvature in a slow motion, slow
spin expansion (slow motion only for Schwarzschild). This allows us
to develop analytic expressions for the curvature, which are useful
for comparing to the fast motion, rapid spin numerical results that we
then compute. We visualize tidally distorted black holes by embedding
their horizons in a 3-dimensional space. This provides a useful way
to see how tides change the shape of a black hole. In
Sec.\ {\ref{sec:leadlag}}, we examine in some detail whether there is
a simple connection between a black hole's tidally distorted geometry
and the coupling between the hole and the orbit. In short, the answer
we find is ``no'' --- Newtonian, fluid intuition breaks down for black
holes and strong-field orbits.
Concluding discussion is given in Sec.\ {\ref{sec:conclude}}, followed
by certain lengthy technical details which we relegate to appendices.
Appendix {\ref{app:ethdetails}} describes in detail how to compute
$\bar\eth$, a Newman-Penrose operator which lowers the spin-weight of
quantities needed for our analysis. Appendix {\ref{app:embed}}
describes how to embed a distorted black hole's event horizon in a
3-dimensional Euclidean space. As mentioned above, one cannot embed
black holes with $a/M > \sqrt{3}/2$ in Euclidean space, but must use a
either a mixed Euclidean/Lorentzian space {\cite{smarr}}, or something
altogether different {\cite{frolov,gibbons}}. We will examine the
range $a/M > \sqrt{3}/2$ in a later paper. Appendix
{\ref{app:spheroidal_lin}} computes, to leading order in spin, the
spheroidal harmonics which are used as basis functions in black hole
perturbation theory. This is needed for the slow-spin expansions we
present in Sec.\ {\ref{sec:kerr_results}}. Finally, Appendix
{\ref{app:glossary}} summarizes certain changes in notation that we
have introduced versus previous papers that use black hole
perturbation theory. These changes synchronize our notation with that
used in the literature from which we have recently adopted our core
numerical method {\cite{ft04,ft05}}.
All of our calculations are done in the background of a Kerr black
hole. Two coordinate systems, described in detail in
Ref.\ {\cite{poisson}}, are particularly useful for us. The
Boyer-Lindquist coordinates ($t, r, \theta, \phi$) yield the line
element
\begin{eqnarray}
ds^2 &=& -\left(1 - \frac{2Mr}{\Sigma}\right)dt^2 -
\frac{4Mar\sin^2\theta}{\Sigma}dt\,d\phi + \frac{\Sigma}{\Delta}dr^2
\nonumber\\
&+& \!\! \Sigma\,d\theta^2 + \frac{(r^2 + a^2)^2 -
a^2\Delta\sin^2\theta}{\Sigma}\sin^2\theta\,d\phi^2\;,
\label{eq:Kerr_BL}
\end{eqnarray}
where
\begin{equation}
\Delta = r^2 - 2Mr + a^2\;,\qquad
\Sigma = r^2 + a^2\cos^2\theta\;.
\label{eq:DeltaSigma}
\end{equation}
The function $\Delta$ has two roots, $r_\pm = M \pm \sqrt{M^2 - a^2}$;
$r_+$ is the location of the event horizon. We will also often find
it useful to use ingoing coordinates $(v, r', \theta, \psi)$, related
to the Boyer-Lindquist coordinates by {\cite{poisson}}
\begin{eqnarray}
dv &=& dt + \frac{(r^2 + a^2)}{\Delta}\,dr\;,
\label{eq:vdef}
\\
d\psi &=& d\phi + \frac{a}{\Delta}\,dr\;.
\label{eq:psidef}
\\
dr' &=& dr\;,
\label{eq:rpdef}
\end{eqnarray}
These coordinates are well-behaved on the event horizon, and so are
useful tools for describing fields that fall into the hole. Although
the relation between $r$ and $r'$ is trivial, it can be useful to
distinguish the two as a bookkeeping device when transforming between
the two coordinate systems. When there is no ambiguity, we will drop
the prime on the ingoing radial coordinate. The Kerr metric in
ingoing coordinates is given by
\begin{eqnarray}
ds^2 &=& -\left(1 - \frac{2Mr'}{\Sigma}\right)dv^2 + 2dv\,dr' -
2a\sin^2\theta\,dr'\,d\psi
\nonumber\\
&-& \!\! \frac{4Mar'\sin^2\theta}{\Sigma}dv\,d\psi
\nonumber\\
&+&\!\! \Sigma\,d\theta^2 + \frac{[(r')^2 + a^2]^2 -
a^2\Delta\sin^2\theta}{\Sigma}\sin^2\theta\,d\psi^2\;.
\nonumber\\
\label{eq:Kerr_IN}
\end{eqnarray}
The quantities $\Sigma$ and $\Delta$ here are exactly as in
Eq.\ (\ref{eq:DeltaSigma}), but with $r \to r'$.
It is not difficult to integrate up Eqs.\ (\ref{eq:vdef}) and
(\ref{eq:psidef}) to find
\begin{equation}
v = t + r^*\;,\qquad
\psi = \phi + \bar r\;,
\label{eq:v_and_psi}
\end{equation}
where {\cite{poisson}}
\begin{eqnarray}
r^* &=& r + \frac{Mr_+}{\sqrt{M^2 - a^2}}\ln\left(\frac{r}{r_+} -
1\right)
\nonumber\\
& & \qquad- \frac{Mr_-}{\sqrt{M^2 - a^2}}\ln\left(\frac{r}{r_-} -
1\right)\;,
\label{eq:rstar}\\
\bar r &=& \frac{a}{2\sqrt{M^2 - a^2}}\ln\left(\frac{r - r_+}{r -
r_-}\right)\;.
\label{eq:rbar}
\end{eqnarray}
Notice that $\psi = \phi$ when $a = 0$.
For $r = r_+ + \delta r$, $\delta r \ll M$,
\begin{equation}
\bar r - \Omega_{\rm H}r^* = K(a) + O(\delta r)\;,
\label{eq:nearhorizradialbehavior}
\end{equation}
where
\begin{eqnarray}
K(a) &=& \frac{a}{2M(Mr_+ - a^2)}\biggl\{a^2 - Mr_+
\nonumber\\
&+& 2M^2{\rm arctanh}\left(\sqrt{1 - a^2/M^2}\right)
\nonumber\\
&+& M\sqrt{M^2 - a^2} \ln\left[\frac{a^2}{4(M^2 - a^2)}\right]\biggr\}
\nonumber\\
&=& -\frac{a}{2M} + \left[1 - 2\ln\left(\frac{a}{2M}\right)\right]
\left(\frac{a}{2M}\right)^3 + O(a^5)\;.
\nonumber\\
\label{eq:K_of_a}
\end{eqnarray}
This means that, near the horizon, the combination $\bar r -
\Omega_{\rm H}r^*$ cancels out the logarithms in both $r^*$ and $\bar
r$, trending to a constant $K(a)$ that depends only on spin. The
quantity $K(a)$ plays an important role in setting the phase of tidal
fields on the event horizon.
\section{Formalism}
\label{sec:formalism}
In this section, we develop the formalism we use to study the geometry
of deformed event horizons. The details of this calculation are
presented in Sec.\ {\ref{sec:geometry}}. Two pieces of this
calculation are sufficiently involved that we present them separately.
First, in Sec.\ {\ref{sec:ZH}}, we give an overview of how one solves
the radial perturbation equation to find the amplitude that sets the
magnitude of the tidal distortion. This material has been discussed
at great length in many other papers, so we present just enough detail
to illustrate what is needed for our analysis. We include in our
discussion the static limit, mode frequency $\omega = 0$. Since
static modes do not carry energy or angular momentum, they have been
neglected in almost all previous analyses. However, these modes
affect the shape of a black hole, so they must be included here.
Second, in Sec.\ {\ref{sec:barethbareth}} we provide detailed
discussion of the angular operator $\bar\eth\bar\eth$ and its action
upon the spin-weighted spheroidal harmonic.
Section {\ref{sec:bulge}} describes how we characterize the bulge in
the event horizon which is raised by the orbiting body's tide. The
bulge is a simple consequence of the geometry, but this discussion
deserves separate treatment in order to properly discuss certain
choices and conventions we must make. We conclude this section by
briefly reviewing down-horizon fluxes in Sec.\ {\ref{sec:downhoriz}}.
Although this discussion is tangential to our main focus in this
paper, we do this to explicitly show that the deformed geometry and
the down-horizon flux are just different ways of presenting the same
information about the orbiting body's perturbation to the black hole.
\subsection{The geometry of an event horizon}
\label{sec:geometry}
We will characterize the geometry of distorted black holes using the
Ricci scalar curvature $R_{\rm H}$ associated with their event
horizon's 2-surface. The scalar curvature of an undistorted Kerr
black hole is given by\footnote{Reference {\cite{smarr}} actually
computes the horizon's Gaussian curvature ${\cal R}_{\rm H}$. The
Gaussian curvature ${\cal R}$ of any 2-surface is exactly half that
surface's scalar curvature $R$, so $R_{\rm H} = 2{\cal R}_{\rm H}$.}
{\cite{smarr}}
\begin{equation}
R_{\rm H} = R^{(0)}_{\rm H} = \frac{2}{r_+^2}\frac{(1 + a^2/r_+^2)(1 -
3a^2\cos^2\theta/r_+^2)}{(1 + a^2\cos^2\theta/r_+^2)^3}\;.
\label{eq:Kerr_curvature}
\end{equation}
For $a = 0$, $R^{(0)}_{\rm H} = 2/r_+^2$, the standard result for a
sphere of radius $r_+$. For $a/M \ge \sqrt{3}/2$, $R^{(0)}_{\rm H}$
changes sign near the poles. This introduces important and
interesting complications to how we represent the tidal distortions of
a rapidly rotating black hole's horizon.
To first order in the mass ratio, tidal distortions leave the horizon
at the coordinate $r = r_+$, but change the scalar curvature on that
surface (at least in all ``horizon-locking gauges'' {\cite{vpm11}},
which we implicitly use in our analysis). Using the Newman-Penrose
formalism {\cite{np62}}, Hartle {\cite{hartle74}} shows that the
perturbation $R^{(1)}_{\rm H}$ to the curvature is simply related to
the perturbing tidal field $\psi_0$:
\begin{eqnarray}
R^{(1)}_{\rm H} &=& -4\,{\rm Im}\sum_{lmkn}\frac{ \bar\eth\bar\eth
\psi^{\rm HH}_{0,lmkn}}{p_{mkn}(ip_{mkn} + 2\epsilon)}
\nonumber\\
&\equiv& \sum_{lmkn} R^{(1)}_{{\rm H},lmkn}\;,
\label{eq:horizcurv1}
\end{eqnarray}
with all quantities evaluated at $r = r_+$. The quantity $\psi^{\rm
HH}_{0,lmkn}$ is a term in a multipolar and harmonic expansion of
the Newman-Penrose curvature scalar $\psi_0$, computed using the
Hawking-Hartle tetrad {\cite{hh72}}:
\begin{eqnarray}
\psi^{\rm HH}_0 &\equiv& -C_{\alpha\beta\gamma\delta}(l^\alpha)^{\rm HH}
(m^\beta)^{\rm HH} (l^\gamma)^{\rm HH} (m^\delta)^{\rm HH}
\nonumber\\
&=& \sum_{lmkn}\psi_{0,lmkn}^{\rm HH}\;.
\label{eq:psi0_def}
\end{eqnarray}
The tensor $C_{\alpha\beta\gamma\delta}$ is the Weyl curvature, and
the vectors $(l^\alpha)^{\rm HH}$ and $(m^\alpha)^{\rm HH}$ are
Newman-Penrose tetrad legs in the Hawking-Hartle representation. See
Appendix {\ref{app:ethdetails}} for detailed discussion of this tetrad
and related quantities.
We assume that $\psi_0$ arises from an object in a bound orbit of the
Kerr black hole. This object's motion can be described using the
three fundamental frequencies associated with such orbits: an axial
frequency $\Omega_\phi$, a polar frequency $\Omega_\theta$, and a
radial frequency $\Omega_r$. The indices $m$, $k$, and $n$ label
harmonics of these frequencies:
\begin{equation}
\omega_{mkn} = m\Omega_\phi + k\Omega_\theta + n\Omega_r\;.
\end{equation}
The index $l$ labels a spheroidal harmonic mode, and is discussed in
more detail below. The remaining quantities appearing in
Eq.\ (\ref{eq:horizcurv1}) are the wavenumber for ingoing
radiation\footnote{This wavenumber is often written $k$ in the
literature; we use $p$ to avoid confusion with harmonics of the
$\theta$ frequency.}
\begin{equation}
p_{mkn} = \omega_{mkn} - m\Omega_{\rm H}\;,
\label{eq:pmkndef}
\end{equation}
and
\begin{equation}
\epsilon = \frac{\sqrt{M^2 - a^2}}{4Mr_+} \equiv \frac{\kappa}{2}\;.
\label{eq:epsilondef}
\end{equation}
The quantity $\kappa$ is the Kerr surface gravity. We will find this
interpretation of $\epsilon$ to be useful when discussing the geometry
of the horizon's tidal distortion. We discuss the operator
$\bar\eth\bar\eth$ in detail in Sec.\ {\ref{sec:barethbareth}}. For
now, note that it involves derivatives with respect to $\theta$.
The calculation of $R^{(1)}_{\rm H}$ involves several computations
that use the Newman-Penrose derivative operator $D \equiv
l^\alpha\partial_\alpha$. Using the Hawking-Hartle form of $l^\alpha$
and ingoing Kerr coordinates (see Appendix {\ref{app:ethdetails}}), we
find that
\begin{equation}
D \to \frac{\partial}{\partial v} + \Omega_{\rm
H}\frac{\partial}{\partial\psi}
\end{equation}
as $r \to r_+$. The fields to which we apply this operator have the
form $e^{i(m\psi - \omega_{mkn} v)}$ near the horizon, so
\begin{equation}
D{\cal F} = i(m\Omega_{\rm H} - \omega_{mkn}){\cal F} = -ip_{mkn}{\cal
F}
\end{equation}
for all relevant fields ${\cal F}$. Hartle choses a time coordinate
$t$ such that $D \equiv \partial/\partial t$ near the horizon,
effectively working in a frame that corotates with the black hole. As
a consequence, his Eq.\ (2.21) [equivalent to our
Eq.\ (\ref{eq:horizcurv1})] has $\omega$ in place of $p$. Hartle's
(2.21) also corresponds to a single Fourier mode, and so is not summed
over indices.
The Hawking-Hartle tetrad is used in Eq.\ (\ref{eq:psi0_def}) because
it is well behaved on the black hole's event horizon {\cite{hh72}}.
In many discussions of black hole perturbation theory based on the
Teukolsky equation, we instead use the Kinnersley tetrad, which is
well designed to describe distant radiation
{\cite{teuk73,kinnersley}}. The Kinnersley tetrad is described
explicitly in Appendix {\ref{app:ethdetails}}. The relation between
$\psi_0$ in these two tetrads is [cf.\ Ref.\ {\cite{tp74}},
Eq.\ (4.43)]
\begin{equation}
\psi_0^{\rm HH} = \frac{\Delta^2}{4(r^2 + a^2)^2} \psi_0^{\rm K}\;.
\label{eq:psi0_convert}
\end{equation}
Further, we know that $\psi_0^{\rm K}$ on the horizon can be written
{\cite{tp74}}
\begin{equation}
\psi_{0,lmkn}^{\rm K} = \frac{W^{\rm
H}_{lmkn}\,{_{+2}S}_{lm}(\theta;a\omega_{mkn})}
{\Delta^2}e^{i(m\phi - \omega_{mkn} t - p_{mkn} r^*)}\;.
\label{eq:psi0_K}
\end{equation}
We have introduced $W^{\rm H}_{lmkn}$, a complex
amplitude\footnote{This amplitude is written $Y$ rather than $W$ in
Ref.\ {\cite{tp74}}; we have changed notation to avoid confusion
with the spherical harmonic.} which we will discuss in more detail
below, as well as the spheroidal harmonic of spin-weight $+2$,
${_{+2}S}_{lm}(\theta;a\omega_{mkn})$. Spheroidal harmonics are often
used in black hole perturbation theory, since the equations governing
a field of spin-weight $s$ in a black hole spacetime separate when
these harmonics are used as a basis for the $\theta$ dependence. In
the limit $a\omega_{mkn} \to 0$, they reduce to the spin-weighted
spherical harmonics:
\begin{equation}
{_sS}_{lm}(\theta;a\omega_{mkn}) \to {_sY}_{lm}(\theta) \quad\mbox{as}\quad
a\omega_{mkn} \to 0\;.
\end{equation}
${_sY}_{lm}(\theta)$ denotes the spherical harmonic without the axial
dependence: ${_s}Y_{lm}(\theta,\phi) = {_s}Y_{lm}(\theta)e^{im\phi}$.
In what follows, we will abbreviate:
\begin{equation}
{_{+2}S}_{lm}(\theta;a\omega_{mkn}) \equiv S^+_{lmkn}(\theta)\;.
\end{equation}
We will likewise write the spin-weight $-2$ spheroidal harmonic as
$S^{-}_{lmkn}(\theta)$.
Combining Eqs.\ (\ref{eq:psi0_convert}) and (\ref{eq:psi0_K}), we find
\begin{equation}
\psi^{\rm HH}_{0,lmkn} = \frac{W^{\rm H}_{lmkn}
S^+_{lmkn}(\theta)}{4(r^2 + a^2)^2} \, e^{i(m\phi - \omega_{mkn} t -
p_{mkn}r^*)}\;.
\label{eq:psi0HHexpand}
\end{equation}
Using Eqs.\ (\ref{eq:v_and_psi}) and (\ref{eq:pmkndef}), we can
rewrite the phase factor using coordinates that are well-behaved on
the horizon:
\begin{eqnarray}
m\phi - \omega_{mkn}t - p_{mkn}r^*\! &=& m(\psi - \bar r)
- \omega_{mkn}(v - r^*)
\nonumber\\
& & - (\omega_{mkn} - m\Omega_{\rm H})r^*
\nonumber\\
&=& m\psi - \omega_{mkn} v - m(\bar r - \Omega_{\rm H}r^*)\;.
\nonumber\\
\end{eqnarray}
Taking the limit $r \to r_+$ and using
Eq.\ (\ref{eq:nearhorizradialbehavior}), we find
\begin{equation}
\psi^{\rm HH}_{0,lmkn} = \frac{W^{\rm
H}_{lmkn}S^+_{lmkn}(\theta)}{16M^2r_+^2}e^{i\Phi_{mkn}(v,\psi)}\;,
\label{eq:psi0HHexpand2}
\end{equation}
where
\begin{equation}
\Phi_{mkn}(v,\psi) = m\psi - \omega_{mkn}v - m K(a)\;,
\label{eq:Phi_mkn}
\end{equation}
with $K(a)$ defined in Eq.\ (\ref{eq:K_of_a}). We finally find
\begin{eqnarray}
R^{(1)}_{{\rm H},lmkn} = -{\rm Im}\left[\frac{W^{\rm H}_{lmkn}
e^{i\Phi_{mkn}(v,\psi)}\bar\eth\bar\eth
S^+_{lmkn}(\theta)}{4M^2r_+^2p_{mkn}(ip_{mkn} +
2\epsilon)}\right]\;. \nonumber\\
\label{eq:gauss2}
\end{eqnarray}
We will use a Teukolsky equation solver
{\cite{h2000,dh2006,thd_inprep}} which computes the curvature scalar
$\psi_4$ rather than $\psi_0$. Although $\psi_4$ is usually used to
study radiation far from the black hole, one can construct $\psi_0$
from it using the Starobinsky-Churilov identities {\cite{tp74,sc73}}.
In the limit $r \to r_+$,
\begin{eqnarray}
\psi_4 &=& \frac{\Delta^2}{(r - ia\cos\theta)^4} \sum_{lmkn}
Z^{\rm H}_{lmkn} S^-_{lmkn}(\theta)
\nonumber\\
& & \qquad\qquad\times\,e^{i(m\phi - \omega_{mkn} t - p_{mkn}r^*)}\;.
\end{eqnarray}
We briefly summarize how we compute $Z^{\rm H}_{lmkn}$ in
Sec.\ {\ref{sec:ZH}}. Using the Starobinsky-Churilov identities, we
find that $Z^{\rm H}_{lmkn}$ and $W^{\rm H}_{lmkn}$ are related by
\begin{equation}
W^{\rm H}_{lmkn} = \beta_{lmkn} Z^{\rm H}_{lmkn}\;,
\label{eq:WlmknZHlmkn}
\end{equation}
where
\begin{equation}
\beta_{lmkn} = \frac{64(2Mr_+)^4p_{mkn}(p_{mkn}^2 + 4\epsilon^2)(p_{mkn} +
4i\epsilon)}{c_{lmkn}}\;,
\label{eq:betalmkn}
\end{equation}
and where the complex number $c_{lmkn}$ is given by
\begin{eqnarray}
|c_{lmkn}|^2 &=& \left\{\left[(\lambda + 2)^2 + 4ma\omega_{mkn} -
4a^2\omega_{mkn}^2\right]\right.
\nonumber\\
& & \left. \qquad\times \left(\lambda^2 + 36ma\omega_{mkn} -
36a^2\omega_{mkn}^2\right)\right.
\nonumber\\
& &\left. \qquad +\,(2\lambda + 3)(96a^2\omega_{mkn}^2 -
48ma\omega_{mkn})\right\}
\nonumber\\
& & + 144\omega_{mkn}^2(M^2 - a^2)\;,
\label{eq:clmknsqr}
\\
{\rm Im}\,c_{lmkn} &=& 12 M \omega_{mkn}\;,
\\
{\rm Re}\,c_{lmkn} &=& +\sqrt{|c_{lmkn}|^2 - 144M^2\omega_{mkn}^2}\;.
\end{eqnarray}
The real number $\lambda$ appearing here is
\begin{equation}
\lambda = {\cal E}_{lmkn} - 2 a m \omega_{mkn} + a^2\omega_{mkn}^2 - 2\;,
\end{equation}
with ${\cal E}_{lmkn}$ the eigenvalue of $S^-_{lmkn}(\theta)$. In the
limit $a\omega_{mkn} \to 0$, ${\cal E}_{lmkn} \to l(l+1)$. For our
later weak-field expansion, it will be useful to have $\lambda$ as an
expansion in $a\omega_{mkn}$. See Appendix {\ref{app:spheroidal_lin}}
for discussion of this.
Using these results, we can write the tidal distortion of the
horizon's curvature as
\begin{eqnarray}
R^{(1)}_{{\rm H},lmkn} &=& -{\rm Im}\left[
\frac{\beta_{lmkn}Z^{\rm H}_{lmkn}
e^{i\Phi_{mkn}(v,\psi)}\bar\eth\bar\eth S^+_{lmkn}(\theta)}
{4M^2r_+^2p_{mkn}(ip_{mkn} + 2\epsilon)}\right]
\nonumber\\
&\equiv& {\rm Im}\left[{\cal C}_{lmkn}Z^{\rm H}_{lmkn}e^{i \Phi_{mkn}(v,\psi)}
\bar\eth\bar\eth S^+_{lmkn}(\theta)\right]\;, \nonumber\\
\label{eq:gauss_kerr}
\end{eqnarray}
where
\begin{equation}
{\cal C}_{lmkn} = 256M^2r_+^2 c_{lmkn}^{-1}(p_{mkn} + 4i\epsilon)(ip_{mkn} -
2\epsilon)\;.
\label{eq:Clmkn}
\end{equation}
Equation (\ref{eq:gauss_kerr}) is the workhorse of our analysis. We
use a slightly modified version of the code described in Refs.
{\cite{h2000,dh2006,thd_inprep}} to compute the complex numbers
$Z^{\rm H}_{lmkn}$ and the angular function $\bar\eth\bar\eth
S^{+}_{lmkn}$. We briefly describe these calculations in the next two
subsections.
\subsection{Computing $Z^{\rm H}_{lmkn}$}
\label{sec:ZH}
Techniques for computing the amplitude $Z^{\rm H}_{lmkn}$ have been
discussed in great detail in other papers, so our discussion here will
be very brief; our analysis follows that given in
Ref.\ {\cite{dh2006}}. The major change versus previous works is that
we need the solution for static modes ($\omega = 0$). Our goal here
is to present enough detail to see how earlier studies can be modified
fairly simply to include these modes. It is worth noting that we have
changed notation from that used in previous papers by our group in
order to more closely follow the notation of Fujita and Tagoshi
{\cite{ft04,ft05}}. Appendix {\ref{app:glossary}} summarizes these
changes.
The complex number $Z^{\rm H}_{lmkn}$ is the amplitude of solutions to
the Teukolsky equation for spin-weight $s = -2$, so we begin there:
\begin{equation}
\Delta^2\frac{d}{dr}\left(\frac{dR_{lm\omega}}{dr}\right) -
V_{lm}(r)R_{lm\omega} = {\cal T}_{lm\omega}(r)\;.
\label{eq:teuk}
\end{equation}
This is the frequency-domain version of this equation, following the
introduction of a modal and harmonic decomposition which separates the
original time-domain equation; see {\cite{teuk73}} for further
details. The potential $V_{lm}$ is discussed in Sec.\ IIIA of
Ref.\ {\cite{dh2006}}; the source term ${\cal T}_{lm\omega}$ is
discussed in Sec.\ IIIB of that paper.
Equation (\ref{eq:teuk}) has two homogeneous solutions relevant to our
analysis: The ``in'' solution is purely ingoing on the horizon, but is
a mixture of ingoing and outgoing at future null infinity; the ``up''
solution is purely outgoing at future null infinity, but is a mixture
of ingoing and outgoing on the horizon. We discuss these solutions in
more detail below. For now, it is enough that these solutions allow
us to build a Green's function {\cite{poisson93}},
\begin{eqnarray}
G(r|r') &=& \frac{1}{{\cal W}}R^{\rm up}_{lm\omega}(r)R^{\rm
in}_{lm\omega}(r')\;, \quad r' < r\;,
\nonumber\\
&=& \frac{1}{{\cal W}}R^{\rm in}_{lm\omega}(r)R^{\rm
up}_{lm\omega}(r')\;, \quad r' > r\;,
\label{eq:Green}
\end{eqnarray}
where
\begin{equation}
{\cal W} =
\frac{1}{\Delta}\left[R^{\rm in}_{lm\omega}\frac{dR^{\rm up}_{lm\omega}}{dr}
- R^{\rm up}_{lm\omega}\frac{dR^{\rm in}_{lm\omega}}{dr}\right]
\end{equation}
is the equation's Wronskian. This is then integrated against the
source to build the general inhomogeneous solution:
\begin{eqnarray}
R_{lm\omega}(r) &=& \int_{r_+}^\infty G(r|r'){\cal T}_{lm\omega}(r')dr'
\nonumber\\
&\equiv& Z^{\rm in}_{lm\omega}(r)R^{\rm up}_{lm\omega}(r) +
Z^{\rm up}_{lm\omega}(r)R^{\rm in}_{lm\omega}(r)\;.
\nonumber\\
\label{eq:inhomogeneous}
\end{eqnarray}
We have defined
\begin{eqnarray}
Z^{\rm in}_{lm\omega}(r) &=& \frac{1}{\cal
W}\int_{r_+}^r\frac{R^{\rm in}_{lm\omega}(r'){\cal
T}_{lm\omega}(r')}{\Delta(r')^2}dr'\;,
\label{eq:Zin1}
\\
Z^{\rm up}_{lm\omega}(r) &=& \frac{1}{\cal
W}\int_r^\infty\frac{R^{\rm up}_{lm\omega}(r'){\cal
T}_{lm\omega}(r')}{\Delta(r')^2}dr'\;.
\label{eq:Zup1}
\end{eqnarray}
A key property of ${\cal T}_{lm\omega}$ is that it is the sum of three
terms, one proportional to $\delta[r - r_{\rm orb}(t)]$, one
proportional to $\delta'[r - r_{\rm orb}(t)]$, and one proportional to
$\delta''[r - r_{\rm orb}(t)]$ (where $'$ denotes $d/dr$). Putting
this into Eqs.\ (\ref{eq:Zin1}) and (\ref{eq:Zup1}), we find that
\begin{eqnarray}
Z^\star_{lm\omega}(r) &=& \frac{1}{\cal W}\Biggl\{ {\cal
I}^0_{lm\omega}\left[R^\star_{lm\omega}(r)\right] + {\cal
I}^1_{lm\omega}\left[\frac{dR^\star_{lm\omega}}{dr}\biggr|_r\right]
\nonumber\\
& & + {\cal I}^2_{lm\omega}\left[\frac{d^2R^\star_{lm\omega}}{dr^2}\biggr|_r
\right]\Biggr\}\;,
\label{eq:Zstargeneric}
\end{eqnarray}
(where $\star$ can stand for ``up'' or ``in''). The factors ${\cal
I}^{0,1,2}_{lm\omega}$ are operators which act on $R^\star_{lm\omega}$
and its derivatives. These operators integrate over the $r$ and
$\theta$ motion of the orbiting body.
In this analysis, we are concerned with the solution of the
perturbation equation on the event horizon, so we want $R_{lm\omega}$
as $r \to r_+$. In this limit, $Z^{\rm in}_{lm\omega} = 0$. We
define
\begin{equation}
Z^{\rm H}_{lm\omega} \equiv Z^{\rm up}_{lm\omega}(r_+)\;.
\end{equation}
For a source term corresponding to a small body in a bound Kerr orbit,
we find that Eq.\ (\ref{eq:Zstargeneric}) has the form
\begin{equation}
Z^{\rm H}_{lm\omega} = \sum_{kn}Z^{\rm H}_{lmkn}\delta(\omega - \omega_{mkn})\;.
\end{equation}
It is then not difficult to read off $Z^{\rm H}_{lmkn}$. See
Ref.\ {\cite{dh2006}} for detailed discussion of how to evaluate
Eq.\ (\ref{eq:Zstargeneric}) and read off these amplitudes.
Key to computing $Z^{\rm H}_{lmkn}$ is computing the homogeneous
solutions $R^{\rm up}_{lm\omega}(r)$, $R^{\rm in}_{lm\omega}(r)$, and
their derivatives. Our methods for doing this depend on whether
$\omega_{mkn}$ is zero or not.
\subsubsection{$\omega_{mkn} \ne 0$}
The homogeneous solutions for $\omega_{mkn} \ne 0$ have been amply
discussed in the literature; our analysis is based on that of
Ref.\ {\cite{dh2006}}. In brief, the two homogeneous solutions of
Eq.\ (\ref{eq:teuk}) have the following asymptotic behavior:
\begin{eqnarray}
R^{\rm in}_{lm\omega}(r \to r_+) &=& B^{\rm
trans}_{lm\omega}\Delta^2e^{-ipr^*}\;,
\label{eq:Rinr+}
\\
R^{\rm in}_{lm\omega}(r \to \infty) &=& B^{\rm
ref}_{lm\omega}r^3e^{i\omega r^*} + \frac{B^{\rm
inc}_{lm\omega}}{r}e^{-i\omega r^*}\;;
\nonumber\\
\label{eq:Rininf}
\\
R^{\rm up}_{lm\omega}(r \to r_+) &=& C^{\rm up}_{lm\omega}e^{ipr^*}
+ C^{\rm ref}_{lm\omega}\Delta^2e^{-ip r^*}\;,
\nonumber\\
\label{eq:Rupr+}\\
R^{\rm up}_{lm\omega}(r \to \infty) &=& C^{\rm
trans}_{lm\omega}r^3e^{i\omega r^*}\;.
\label{eq:Rupinf}
\end{eqnarray}
These asymptotic solutions yield the Wronskian:
\begin{equation}
{\cal W} = 2i\omega B^{\rm inc}_{lm\omega}C^{\rm trans}_{lm\omega}\;.
\end{equation}
An effective algorithm for computing all of the quantities which we
need is described by Fujita and Tagoshi {\cite{st03,ft04,ft05}}. It
is based on expanding the solution in a basis of hypergeometric and
Coulomb wave functions, with the coefficients of the expansion
determined by solving a recurrence relation; see Secs.\ 4.2 -- 4.4 of
Ref.\ {\cite{st03}} for detailed discussion. We use a code based on
these methods {\cite{thd_inprep}} for all of our $\omega_{mkn} \ne 0$
calculations; the analytic limits we present in
Secs.\ {\ref{sec:schw_analytic}} and {\ref{sec:kerr_analytic}} are
also based on these methods.
\subsubsection{$\omega_{mkn} = 0$}
Static modes have been neglected in much past work. They do not carry
any energy or angular momentum, and so are not important for many
applications. These modes do play a role in setting the shape of the
distorted event horizon, however, and must be included here.
It turns out that homogeneous solutions for $\omega_{mkn} = 0$ are
available as surprisingly simple closed form expressions. Teukolsky's
Ph.D.\ thesis {\cite{teukphd}} presents two solutions that satisfy
appropriate boundary conditions. Defining
\begin{equation}
x = \frac{r - r_+}{r_+ - r_-}\;,\quad\gamma = \frac{iam}{r_+ - r_-}\;,
\end{equation}
the two solutions of the radial Teukolsky equation for $s = -2$ are
\begin{eqnarray}
R^{\rm in}_{lm0}(r) &=& (r_+ - r_-)^4 x^2(1+x)^2\left(\frac{x}{1 +
x}\right)^\gamma\,\times
\nonumber\\
& & _2F_1(2-l, l+3; 3 + 2\gamma, -x)\;,
\label{eq:Rin_omega0}\\
R^{\rm up}_{lm0}(r) &=& (r_+ - r_-)^{(1 - l)}x^{(1 - l)}(1 +
1/x)^{(2-\gamma)}\,\times
\nonumber\\
& & _2F_1(l+3,l+1-2\gamma;2l+2,-1/x)\;.
\nonumber\\
\label{eq:Rup_omega0}
\end{eqnarray}
In these equations, $_2F_1(a,b;c,x)$ is the hypergeometric function.
These solutions satisfy regularity conditions at infinity and on the
horizon: $R^{\rm in}_{lm0}(r\to r_+) \propto \Delta^2$, and $R^{\rm
up}_{lm0}(r\to\infty)\propto 1/r^{l-1}$ \cite{teukphd}. We have
introduced powers of $r_+ - r_-$ to insure that we have the correct
asymptotic behavior in $r$, rather than in the dimensionless variable
$x$. The Wronskian corresponding to these solutions is
\begin{equation}
{\cal W} = -\frac{(2l+1)!}{(l+2)!}\frac{\Gamma(3 + 2\gamma)}{\Gamma(l
+ 1 + 2\gamma)}(r_+ - r_-)^{(2 - l)}\;.
\label{eq:Wronskian_omega0}
\end{equation}
Using Eqs.\ (\ref{eq:Rin_omega0}), (\ref{eq:Rup_omega0}), and
(\ref{eq:Wronskian_omega0}), it is simple to adapt existing codes to
compute $Z^{\rm H}_{lmkn}$ for $\omega_{mkn} = 0$.
The results we present in Secs.\ {\ref{sec:schw_results}} and
{\ref{sec:kerr_results}} will focus on circular, equatorial orbits,
for which $k = n = 0$. The zero-frequency modes in this limit have $m
= 0$, for which $\gamma = 0$. The Wronskian simplifies further:
\begin{equation}
{\cal W}_{(m = 0)} = -\frac{2(2l+1)!}{l!(l+2)!}(r_+ - r_-)^{(2 -
l)}\;.
\end{equation}
For generic orbit geometries, there will exist cases that have
$\omega_{mkn} = 0$ with $m \ne 0$, akin to the ``resonant'' orbits
studied at length in Refs.\ {\cite{fh12,fhr14}}. We defer discussion
of this possibility to a later analysis which will go beyond circular
and equatorial orbits.
\subsection{The operator $\bar\eth\bar\eth$}
\label{sec:barethbareth}
The operator $\bar\eth$, when acting on a quantity $\eta$ of
spin-weight $s$, takes the following form:
\begin{equation}
\bar\eth\eta = \left[\bar\delta - (\alpha - \bar\beta)\right]\eta\;;
\end{equation}
$\bar\eth\eta$ is then a quantity of spin-weight $s-1$. The
quantities $\alpha$ and $\beta$ are both Newman-Penrose spin
coefficients, and $\bar\delta$ is a Newman-Penrose derivative
operator. These quantities are all related to the tetrad legs ${\bf
m}$, $\bar{\bf m}$:
\begin{eqnarray}
\bar\delta &=& \bar m^\mu\partial_\mu\;,
\\
\alpha - \bar\beta &=& \frac{1}{2}\bar m^\nu\left(\bar m^\mu\nabla_\nu
m_\mu - m^\mu\nabla_\nu \bar m_\mu\right)\;.
\end{eqnarray}
We do this calculation using the Hawking-Hartle tetrad; details are
given in Appendix {\ref{app:ethdetails}}. The result for general
black hole spin $a$ is
\begin{eqnarray}
\bar\eth\eta &=& \frac{1}{\sqrt{2}(r_+ - ia\cos\theta)} \Biggl(L^s_- -
am\Omega_{\rm H}\sin\theta
\nonumber\\
& &\qquad\qquad\qquad\quad - \frac{isa\sin\theta}{r_+ -
ia\cos\theta}\Biggr)\eta\;.
\label{eq:kerr_eth1}
\end{eqnarray}
The operator\footnote{This operator is denoted $\bar\eth_0$ in
Ref.\ {\cite{hartle74}}. We will use the symbol $\bar\eth_0$ to
instead denote the Schwarzschild limit of $\bar\eth$.} $L^s_-$
lowers the spin-weight of the spherical harmonics by 1:
\begin{eqnarray}
L^s_- {_sY}_{lm} &=& \left(\partial_\theta + s\cot\theta +
m\csc\theta\right) {_sY}_{lm}
\nonumber\\
&=& \sqrt{(l+s)(l-s+1)}
{_{s-1}Y}_{lm}\;.
\label{eq:sphericalharmoniclower}
\end{eqnarray}
In a few places, we will need to evaluate
$L^s_-\left[\cos\theta\eta\right]$ and
$L^s_-\left[\sin\theta\eta\right]$. This requires that we rewrite
$\cos\theta$ and $\sin\theta$ in a form that properly indicates their
spin weight. We treat $\cos\theta$ as spin-weight zero, writing
\begin{equation}
\cos\theta = \sqrt{\frac{4\pi}{3}}\, {_0}Y_{10}\;.
\end{equation}
Likewise, we treat $\sin\theta$ as spin-weight $-1$, writing
\begin{equation}
\sin\theta = -\sqrt{\frac{8\pi}{3}}\, {_{-1}}Y_{10}\;.
\end{equation}
This accounts for the fact that $\sin\theta$ always appears in our
calculation inside operators that lower spin-weight.
With this, we find the following identities:
\begin{eqnarray}
L^s_-\left[\cos\theta\eta\right] &=&
\sqrt{\frac{4\pi}{3}} L^s_-\left[{_0}Y_{10}\,\eta\right]
\nonumber\\
&=& \sqrt{\frac{4\pi}{3}}\left({_0}Y_{10}\,L^s_-\eta +
\eta\, L^s_-\,{_0}Y_{10}\right)
\nonumber\\
&=& \sqrt{\frac{4\pi}{3}}\left({_0}Y_{10}\,L^s_-\eta +
\eta\,\sqrt{2}\,{_{-1}}Y_{10}\right)
\nonumber\\
&=& \cos\theta\,L^s_-\eta - \sin\theta\eta\;;
\label{eq:Lminus_s_costheta}
\end{eqnarray}
\begin{eqnarray}
L^s_-\left[\sin\theta\eta\right] &=&
-\sqrt{\frac{8\pi}{3}} L^s_-\left[_{-1}Y_{10}\eta\right]
\nonumber\\
&=& -\sqrt{\frac{8\pi}{3}}\left(_{-1}Y_{10}\,L^s_-\eta +
\eta\,L^s_-\,{_{-1}}Y_{10}\right)
\nonumber\\
&=& -\sqrt{\frac{8\pi}{3}}\,{_{-1}Y}_{10}L^s_-\,\eta
\nonumber\\
&=& \sin\theta\,L^s_-\eta\;.
\label{eq:Lminus_s_sintheta}
\end{eqnarray}
We used the fact that $L^s_-$ applied to ${_{-1}Y}_{10}$ yields zero.
Using these results, it follows that
\begin{eqnarray}
L^s_-\left(1 - \frac{ia\cos\theta}{r_+}\right)^{-s}\eta &=& \left(1 -
\frac{ia\cos\theta}{r_+}\right)^{-s}
\nonumber\\
& &
\!\!\!\!\!\!\!\!
\!\!\!\!\!\!\!\!
\!\!\!\!\times
\left(L^s_- - \frac{ias\sin\theta}{r_+ -
ia\cos\theta}\right)\eta\;.
\nonumber\\
\label{eq:usefulidentity}
\end{eqnarray}
We can next rewrite Eq.\ (\ref{eq:kerr_eth1}) as
\begin{eqnarray}
\bar\eth\eta &=& \frac{1}{\sqrt{2}r_+}\left(1 -
\frac{ia\cos\theta}{r_+}\right)^{s-1}
\nonumber\\
& &
\times\left(L^s_- -
am\Omega_{\rm H}\sin\theta\right)\left(1 -
\frac{ia\cos\theta}{r_+}\right)^{-s}\eta\;.
\nonumber\\
\label{eq:kerr_eth}
\end{eqnarray}
When $a = 0$, this reduces to
\begin{equation}
\bar\eth\eta = \frac{1}{2\sqrt{2}M}L^s_-\eta \equiv \bar\eth_0\;.
\end{equation}
When $\eta$ is of spin-weight 2, Eq.\ (\ref{eq:kerr_eth}) tells us that
\begin{equation}
\bar\eth\bar\eth\eta = \frac{1}{2r_+^2} \left(L^s_- -
am\Omega_{\rm H}\sin\theta\right)^2\left(1 -
\frac{ia\cos\theta}{r_+}\right)^{-2}\eta\;.
\label{eq:barethsqr_s=2}
\end{equation}
For $a \ll M$, Eq.\ (\ref{eq:barethsqr_s=2}) reduces to
\begin{equation}
\bar\eth\bar\eth\eta = \frac{1}{8M^2}\,L^s_-L^s_-\left(1 +
\frac{ia\cos\theta}{M}\right)\eta\;,
\label{eq:barethbareth_small_a}
\end{equation}
which reproduces Eq.\ (4.19) of Ref.\ {\cite{hartle74}}.
We will apply $\bar\eth\bar\eth$ to the spheroidal harmonic
$S^{+}_{lm}(\theta)$. Following Ref.\ {\cite{h2000}}, we compute this
function by expanding it using a basis of spherical harmonics, writing
\begin{equation}
S^+_{lm}(\theta) = \sum_{q = q_{\rm min}}^\infty b^l_q(a\omega_{mkn})
_{+2}Y_{qm}(\theta)\;,
\label{eq:spheroidexpand}
\end{equation}
where $q_{\rm min} = {\rm min}(2,|m|)$. Efficient algorithms exist to
compute the expansion coefficients $b^l_q(a\omega_{mkn})$
(cf.\ Appendix A of Ref.\ {\cite{h2000}}). Expanding
Eq.\ (\ref{eq:barethsqr_s=2}) puts it into a form very useful for our
purposes:
\begin{equation}
\bar\eth\bar\eth\eta = \frac{1}{2(r_+ -
ia\cos\theta)^2}\left[L^s_-L^s_- + {\cal A}_1L^s_- + {\cal
A}_2\right]\eta\;,
\label{eq:barethsqr_s=2_expand}
\end{equation}
where
\begin{eqnarray}
{\cal A}_1 &=& -2a\sin\theta\left[m\Omega_{\rm H} + \frac{2i}{r_+ -
ia\cos\theta}\right]\;,
\label{eq:barethbareth_A1}\\
{\cal A}_2 &=& a^2\sin^2\theta\biggl[m^2\Omega^2_{\rm H}
+ \frac{4im\Omega_{\rm H}}{r_+ - ia\cos\theta}
\nonumber\\
& &\qquad\qquad\qquad
-\frac{6}{(r_+ - ia\cos\theta)^2}\biggr]\;.
\label{eq:barethbareth_A2}
\end{eqnarray}
Combining Eqs.\ (\ref{eq:spheroidexpand}) and
(\ref{eq:barethsqr_s=2_expand}), and making use of
Eq.\ (\ref{eq:sphericalharmoniclower}), we finally obtain
\begin{eqnarray}
\bar\eth\bar\eth S^+_{lm} &=& \frac{1}{2(r_+ - ia\cos\theta)^2}
\sum_{q=q_{\rm min}}^\infty
b^l_q(a\omega_{mkn})
\nonumber\\
& &\!\!\!\!\!\!\!\!\!\!\!\!\!\!\!
\times
\biggl[\sqrt{(q+2)(q+1)q(q-1)}\,{_0}Y_{qm}
\nonumber\\
& &\!\!\!\!\!\!\!\!\!\!\!
+ {\cal A}_1\sqrt{(q+2)(q-1)}\,{_1}Y_{qm} + {\cal
A}_2\,{_2}Y_{qm}\biggr]\;.
\nonumber\\
\label{eq:barethbarethS}
\end{eqnarray}
This equation is simple to evaluate using the techniques presented in
Appendix A of Ref.\ {\cite{h2000}}.
\subsection{The phase of the tidal bulge}
\label{sec:bulge}
As we will see when we examine the geometry of distorted event
horizons in detail in Secs.\ {\ref{sec:schw_results}} and
{\ref{sec:kerr_results}}, a major effect of tides on a black hole is
to cause the horizon to bulge. As has been described in detail in
past literature (e.g., {\cite{membrane}}), the result is not so
different from the response of a fluid body to a tidal driving force,
albeit with some counterintuitive aspects thanks to the teleological
nature of the event horizon.
In this section, we describe three ways to characterize the tidal
bulge of the distorted event horizon. Two of these methods are based
on comparing the position at which the horizon is most distorted to
the position of the orbit. Because the orbit and the horizon are at
different locations, comparing their positions requires us to map from
one to the other. The notion of bulge phase that follows then depends
on the choice of map we use. As such, any notion of bulge phase built
from comparing orbit position to horizon geometry must be somewhat
arbitrary, and can only be understood in the context of the mapping
that has been used.
We use two maps from orbit to horizon. The first is a ``null map.''
Following Hartle {\cite{hartle74}}, we connect the orbit to the
horizon using an inward-going, zero-angular-momentum null geodesic.
This choice is commonly used in the literature, and so is useful for
comparing our results with past work. The second is an
``instantaneous map.'' We compare the horizon geometry to the orbit
position on a slice of constant ingoing time coordinate $v$. This is
particularly convenient for showing figures of the distorted horizon.
The third method of computing bulge phase directly compares the
horizon's response to the applied tidal field. Since both quantities
are defined on the horizon, no mapping is necessary, and no arbitrary
choices are needed. We do not use this notion of bulge phase very
much in this analysis, but anticipate using it in future work which
will examine more complicated cases than the circular, equatorial
orbits that are our focus here.
\subsubsection{Relative position of orbit and bulge I: Null map}
\label{sec:nullmap}
In his original examination of black hole tidal distortion, Hartle
{\cite{hartle74}} connects the orbit to the horizon with a zero
angular momentum ingoing light ray. Choosing our origins
appropriately, the orbiting body is at angle
\begin{equation}
\phi_{\rm o} = \Omega_{\rm orb}t
\end{equation}
in Boyer-Lindquist coordinates. We convert to ingoing coordinates
using Eq.\ (\ref{eq:v_and_psi}):
\begin{eqnarray}
\psi_{\rm o} &=& \Omega_{\rm orb}(v - r^*_{\rm o}) + \bar r_{\rm o}
\nonumber\\
&\equiv& \Omega_{\rm orb}v + \Delta\psi(r_{\rm o})\;,
\end{eqnarray}
where $\bar r_{\rm o} \equiv \bar r(r_{\rm o})$ and $r^*_{\rm o}
\equiv r^*(r_{\rm o})$ are given by Eqs.\ (\ref{eq:rbar}) and
(\ref{eq:rstar}), and where
\begin{equation}
\Delta\psi(r_{\rm o}) \equiv \bar r_{\rm o} - \Omega_{\rm orb} r^*_{\rm o}
\label{eq:Deltapsi_orb}
\end{equation}
is, for each orbital radius $r_{\rm o}$, a fixed angular offset
associated with the transformation from Boyer-Lindquist to ingoing
coordinates.
The orbit's location mapped onto the horizon is then
\begin{equation}
\psi^{\rm NM}_{\rm o} = \Omega_{\rm orb}v + \Delta\psi(r_{\rm o}) +
\delta\psi^{\rm null}\;,
\label{eq:psi_o_nullmap}
\end{equation}
where $\delta\psi^{\rm null}$ is the axial shift accumulated by the
ingoing null ray as it propagates from the orbit to the horizon. This
shift must in general be computed numerically, but to leading order in
$a$ (which will be sufficient for our purposes) it is given by
\begin{equation}
\delta\psi^{\rm null} = -\frac{a}{2M} + \frac{a}{r_{\rm o}}
= 2M\Omega_{\rm H}\left(\frac{2M}{r_{\rm o}} - 1\right)\;.
\label{eq:deltapsinull_expand}
\end{equation}
The second form uses $\Omega_{\rm H} = a/4M^2$ for small $a$ to
rewrite this formula, which will be useful when we compare our results
to previous literature for small spin. (One should also correct the
ingoing time, $v \to v + \delta v$, to account for the time it takes
for the ingoing null ray to propagate from the orbit to the horizon.
However, at leading order $\delta v \propto a^2$, so we can neglect it
for the applications we will use in this paper. The time shift is
also neglected in all previous papers we are aware of which examine
the angular offset of the tidal bulge {\cite{hartle74,fl05}}, since
they only consider $a = 0$ or $a/M \ll 1$.)
Let $\psi^{\rm bulge}$ be the angle at which $R^{(1)}_{\rm H}$ is
maximized. This value varies from mode to mode, but is easy to read
off once $R^{(1)}_{\rm H}$ is computed. The offset of the orbit and
bulge using the null map is then
\begin{eqnarray}
\delta\psi^{\rm OB-NM} &\equiv& \psi^{\rm bulge} - \psi^{\rm NM}_{\rm
o}
\nonumber\\
&=& \psi^{\rm bulge} - \Omega_{\rm orb}v - \Delta\psi(r_{\rm o}) -
\delta\psi^{\rm null}\;.
\nonumber\\
\label{eq:bulge_vs_orbit_null}
\end{eqnarray}
A positive value for $\delta\psi^{\rm OB-NM}$ means that the bulge
leads the orbit.
\subsubsection{Relative position of orbit and bulge II: Instantaneous map}
\label{sec:instmap}
Consider next a mapping that is instantaneous in ingoing time
coordinate $v$. This choice is useful for making figures that show
both bulge and orbit, since we simply show their locations at a given
moment $v$. This mapping neglects the term $\delta\psi^{\rm null}$,
but is otherwise identical to the null map:
\begin{equation}
\psi^{\rm IM}_{\rm o} = \psi_{\rm o} = \Omega v + \Delta\psi(r_{\rm o})\;.
\label{eq:psi_o_inst}
\end{equation}
The offset of the orbit and bulge in this mapping is
\begin{eqnarray}
\delta\psi^{\rm OB-IM} &\equiv& \psi^{\rm bulge} - \psi^{\rm IM}_{\rm
o}
\nonumber\\
&=& \psi^{\rm bulge} - \Omega_{\rm orb}v - \Delta\psi(r_{\rm o})\;.
\label{eq:bulge_vs_orbit_inst}
\end{eqnarray}
Since $\delta\psi^{\rm null} = 0$ for $a = 0$, the null and
instantaneous maps are identical for Schwarzschild black holes.
Before concluding our discussion of the tidal bulge phase, we
emphasize again that the phase in both the null map and the
instantaneous map follow from arbitrary choices, and must be
interpreted in the context of those choices. Other choices could be
made. For example, one could make a map that is instantaneous in a
different time coordinate, or that is based on a different family of
ingoing light rays (e.g., the principle ingoing null congruence, along
which $v$, $\psi$, and $\theta$ are constant; such a map would be
identical to the instantaneous map). These two maps are good enough
for our purposes --- the null map allows us to compare with other
papers in the literature, and the instantaneous map is excellent for
characterizing the plots we will show in
Secs.\ {\ref{sec:schw_results}} and {\ref{sec:kerr_results}}.
\subsubsection{Relative phase of tidal field and response}
\label{sec:phase_tide_horiz}
Our third method of characterizing the tidal bulge is to use the
relative phase of the horizon distortion $R^{(1)}_{\rm H}$ and
distorting tidal field $\psi_0$. For our frequency-domain study, this
phase is best understood on a mode-by-mode basis. Begin by
re-examining Eq.\ (\ref{eq:horizcurv1}):
\begin{eqnarray}
R^{(1)}_{{\rm H},lmkn} &=& -4\,{\rm Im}\left[\frac{\bar\eth\bar\eth
\psi^{\rm HH}_{0,lmkn}}{p_{mkn}(ip_{mkn} + 2\epsilon)}\right]
\nonumber\\
&\equiv& {\rm Im}\left[R^{\rm c}_{lmkn}\right]\;.
\end{eqnarray}
Let us define the phase $\delta\psi^{\rm TB}_{lmkn}$ by
\begin{equation}
\frac{R^{\rm c}_{lmkn}}{\psi^{\rm HH}_{0,lmkn}} =
\frac{|R^{\rm c}_{lmkn}|}{|\psi^{\rm
HH}_{0,lmkn}|}e^{-i\delta\psi^{\rm TB}_{lmkn}}\;.
\label{eq:deltapsiTB_def}
\end{equation}
As with $\delta\psi^{\rm OB-NM}$ and $\delta\psi^{\rm OB-IM}$,
$\delta\psi^{\rm TB}_{lmkn} > 0$ means that the horizon's response
leads the tidal field.
Using Eq.\ (\ref{eq:psi0HHexpand2}), we see that
\begin{equation}
\frac{R^{\rm c}_{lmkn}}{\psi^{\rm HH}_{0,lmkn}} =
-\frac{4}{p_{mkn}(ip_{mkn} + 2\epsilon)}\frac{\bar\eth\bar\eth
S^+_{lmkn}}{S^+_{lmkn}}\;.
\label{eq:deltapsiTB1}
\end{equation}
With a few definitions, this form expedites our identification of
$\delta\psi^{\rm TB}_{lmkn}$. First, note that $p_{mkn}$ and
$S^+_{lmkn}$ are both real, so the phase arises solely from the factor
$1/(ip_{mkn} + 2\epsilon)$ and the operator $\bar\eth\bar\eth$. The
first factor is easily rewritten in a more useful form:
\begin{equation}
\frac{1}{ip_{mkn} + 2\epsilon} =
\frac{e^{-i\arctan(p_{mkn}/2\epsilon)}}{\sqrt{p_{mkn}^2 +
4\epsilon^2}}\;.
\label{eq:deltapsiTB2}
\end{equation}
To clean up the phase associated with $\bar\eth\bar\eth$, we make a
definition:
\begin{equation}
\frac{\bar\eth\bar\eth S^+_{lmkn}}{S^+_{lmkn}} \equiv
\Sigma_{lmkn}(\theta) e^{-i{\cal S}_{lmkn}(\theta)}\;.
\label{eq:spheroidratio}
\end{equation}
The amplitude ratio $\Sigma_{lmkn}(\theta)$ and phase ${\cal
S}_{lmkn}(\theta)$ must in general be determined numerically. We
will show expansions for small $a$ and slow motion in
Sec.\ {\ref{sec:kerr_results}}. We include $S^+_{lmkn}$ in this
definition because it may pass through zero at a different angle than
$\bar\eth\bar\eth S^+_{lmkn}$ passes through zero. This will appear
as a change by $\pi$ radians in the phase ${\cal S}_{lmkn}$.
Combining Eqs.\ (\ref{eq:deltapsiTB_def}) -- (\ref{eq:spheroidratio})
and using the fact that $\epsilon = \kappa/2$ (where $\kappa$ is the
black hole surface gravity), we at last read out
\begin{equation}
\delta\psi^{\rm TB}_{lmkn} = \arctan\left(p_{mkn}/\kappa\right) + {\cal
S}_{lmkn}(\theta)\;.
\label{eq:deltapsiTB}
\end{equation}
Recall that the wavenumber $p_{mkn} = \omega_{mkn} - m\Omega_{\rm H}$.
In geometrized units, $\kappa^{-1}$ is a timescale which characterizes
how quickly the horizon adjusts to an external disturbance
(cf.\ Sec.\ VI C 5 of Ref.\ {\cite{membrane}} for discussion). The
first term in Eq.\ (\ref{eq:deltapsiTB}) is thus determined by the
wavenumber times this characteristic horizon time. For a circular,
equatorial orbit which has $\Omega_{\rm orb} = \Omega_{\rm H}$, this
term is zero, in accord with the Newtonian intuition that the tide and
the response are exactly aligned when the spin and orbit frequencies
are identical. This intuition does not quite hold up thanks to the
correcting phase ${\cal S}_{lmkn}(\theta)$. We will examine the
impact of this correction in Sec.\ {\ref{sec:kerr_results}}.
The phase $\delta\psi^{\rm TB}_{lmkn}$ is particularly useful for
describing the horizon's response to complicated orbits where the
relative geometry of the horizon and the orbit is dynamical. For
example, Vega, Poisson, and Massey {\cite{vpm11}} use a measure
similar to $\delta\psi^{\rm TB}_{lmkn}$ to describe how a
Schwarzschild black hole responds to a body that comes near the
horizon on a parabolic encounter, demonstrating that the horizon's
response leads the applied tidal field (cf.\ Sec.\ 5.2 of
Ref.\ {\cite{vpm11}}). We will examine $\delta\psi^{\rm TB}_{lmkn}$
briefly for the circular, equatorial orbits we focus on in this paper,
but will use it in greater depth in a follow-up analysis that looks at
tides from generic orbits.
When $a = 0$, the operator $\bar\eth\bar\eth$ is real, and ${\cal
S}_{lmkn}(\theta) = 0$. We have $p_{mkn} = \omega_{mkn}$ and
$\kappa = 1/4M$ in this limit, so
\begin{equation}
\delta\psi^{\rm TB}_{lmkn}\Bigr|_{a = 0} \to\quad \delta\phi^{\rm TB}_{mkn}
= \arctan\left(4M\omega_{mkn}\right)\;.
\label{eq:schw_bulge_vs_tide}
\end{equation}
We will show in Sec.\ {\ref{sec:schw_results}} that this agrees with
the phase shift obtained by Fang and Lovelace {\cite{fl05}}. It also
agrees with the results of Vega, Poisson, and Massey {\cite{vpm11}},
though in somewhat different language. They work in the time domain,
showing that a Schwarzschild black hole's horizon response leads the
field by a time interval $\kappa_{\rm Schw}^{-1} = 4M$. For a field
that is periodic with frequency $\omega$, this means that we expect
the response to lead the field by a phase angle $4M\omega$, exactly as
Eq.\ (\ref{eq:schw_bulge_vs_tide}) says.
\subsection{The down-horizon flux}
\label{sec:downhoriz}
Although not needed for this paper, we now summarize how one computes
the down-horizon flux. Our purpose is to show that the coefficients
$Z^{\rm H}_{lmkn}$ which characterize the geometry of the deformed
event horizon also characterize the down-horizon gravitational-wave
flux, showing that the ``deformed horizon'' and ``down-horizon flux''
pictures are just different ways of interpreting how the horizon
interacts with the orbit.
Our discussion follows Teukolsky and Press {\cite{tp74}}, which in
turn follows Hawking and Hartle {\cite{hh72}}, modifying the
presentation slightly to follow our notation. The starting point is
to note that a tidal perturbation shears the generators of the event
horizon. This shear, $\sigma$, causes the area of the event horizon
to grow:
\begin{equation}
\frac{d^2A}{d\Omega dt} = \frac{2Mr_+}{\epsilon}|\sigma|^2\;.
\label{eq:areagrowth}
\end{equation}
We also know the area of a black hole's event horizon,
\begin{equation}
A = 8\pi\left(M^2 + \sqrt{M^4 - S^2}\right)\;,
\end{equation}
where $S = aM$ is the black hole's spin angular momentum. Using this,
we can write the area growth law as
\begin{equation}
\frac{d^2A}{d\Omega dt} = \frac{8\pi}{\sqrt{M^4 - S^2}}
\left(2M^2r_+\frac{d^2M}{d\Omega dt} -
S\frac{d^2S}{d\Omega dt}\right)\;.
\end{equation}
Consider now radiation going down the horizon. Radiation carrying
energy $dE^{\rm H}$ and angular momentum $dL_z^{\rm H}$ into the hole
changes its mass and spin by
\begin{equation}
dM = dE^{\rm H}\;,\qquad dS = dL_z^{\rm H}\;.
\end{equation}
Angular momentum and energy carried by the radiation are related
according to
\begin{equation}
dL_z = \frac{m}{\omega_{mkn}} dE\;.
\end{equation}
Putting all of this together and using Eq.\ (\ref{eq:pmkndef}), we
find
\begin{eqnarray}
\frac{d^2E^{\rm H}}{dt d\Omega} &=& \frac{\omega_{mkn} Mr_+}{2\pi
p_{mkn}}|\sigma|^2\;,
\label{eq:dEHdtdOmega}
\\
\frac{d^2L_z^{\rm H}}{dt d\Omega} &=& \frac{mMr_+}{2\pi
p_{mkn}}|\sigma|^2\;.
\label{eq:dLzHdtdOmega}
\end{eqnarray}
So to compute the down-horizon flux, we just need to know the shear
$\sigma$. It is simply computed from the tidal field $\psi_0^{\rm
HH}$. First, expand $\sigma$ as
\begin{equation}
\sigma = \sum_{lmkn}\sigma_{lmkn}S^+_{lmkn}(\theta)e^{i[m\psi -
\omega_{mkn}v - mK(a)]}\;.
\end{equation}
The shear mode amplitudes $\sigma_{lmkn}$ are related to the tidal
field mode $\psi^{\rm HH}_{0,lmkn}$ by {\cite{tp74}}:
\begin{equation}
\sigma_{lmkn} = \frac{i\psi_{0,lmkn}^{\rm HH}}{p_{mkn} - 2i\epsilon}\;.
\label{eq:sigmapsi0relation}
\end{equation}
Combine Eq.\ (\ref{eq:sigmapsi0relation}) with
Eqs.\ (\ref{eq:psi0HHexpand}), (\ref{eq:WlmknZHlmkn}), and
(\ref{eq:betalmkn}). Integrate over solid angle, using the
orthogonality of the spheroidal harmonics. Equations
(\ref{eq:dEHdtdOmega}) and (\ref{eq:dLzHdtdOmega}) become
\begin{eqnarray}
\left(\frac{dE}{dt}\right)^{\rm H} &=& \sum_{lmkn} \alpha_{lmkn}
\frac{|Z^{\rm H}_{lmkn}|^2}{4\pi\omega_{mkn}^2}\;,
\label{eq:dEHdt}\\
\left(\frac{dL_z}{dt}\right)^{\rm H} &=& \sum_{lmkn} \alpha_{lmkn}
\frac{m|Z^{\rm H}_{lmkn}|^2}{4\pi\omega_{mkn}^3}\;.
\label{eq:dLzHdt}
\end{eqnarray}
The coefficient
\begin{eqnarray}
\alpha_{lmkn} &=& \frac{256(2Mr_+)^5p_{mkn}\omega_{mkn}^3}{|c_{lmkn}|^2}
\nonumber\\
& &\times\,
(p_{mkn}^2 + 4\epsilon^2)(p_{mkn}^2 + 16\epsilon^2)\;,
\end{eqnarray}
with $|c_{lmkn}|^2$ given by Eq.\ (\ref{eq:clmknsqr}), comes from
combining the various prefactors in the relations that lead to
Eqs.\ (\ref{eq:dEHdt}) and (\ref{eq:dLzHdt}). Notice that
$\alpha_{lmkn} \propto p_{mkn}$. This means that $\alpha_{lmkn} = 0$
when $\omega_{mkn} = m\Omega_{\rm H}$. The down-horizon fluxes
(\ref{eq:dEHdt}) and (\ref{eq:dLzHdt}) are likewise zero for modes
which satisfy this condition.
It is interesting to note that the shear $\sigma_{lmkn}$ and the tidal
field $\psi^{\rm HH}_{0,lmkn}$ are both proportional to $p_{mkn}$, and
hence both vanish when $\omega_{mkn} = m\Omega_{\rm H}$. The
horizon's Ricci curvature $R^{(1)}_{{\rm H},lmkn}$ does not, however,
vanish in this limit. Mathematically, this is because $R^{(1)}_{{\rm
H},lmkn}$ includes a factor of $1/p_{mkn}$ which removes this
proportionality [cf.\ Eq.\ (\ref{eq:horizcurv1})]. Physically, this
is telling us that when $\Omega_{\rm H} = \Omega_{\rm orb}$, the
horizon is deformed, but the deformation is static in the horizon's
reference frame. This static deformation does not shear the
generators, and does not carry energy or angular momentum into the
hole.
Equations (\ref{eq:dEHdt}) and (\ref{eq:dLzHdt}) illustrate the point
of this section: The fluxes of $E$ and $L_z$ into the horizon are
determined by the same numbers $Z^{\rm H}_{lmkn}$ used to compute the
horizon's deformed geometry, Eq.\ (\ref{eq:gauss_kerr}).
\section{Results I: Schwarzschild}
\label{sec:schw_results}
Using the formalism we have assembled, we now examine the tidally
deformed geometry of black hole event horizons. In this paper, we
will only study the circular, equatorial limit: The orbiting body is
at $r = r_{\rm o}$, $\theta = \pi/2$, and $\phi = \Omega_{\rm orb}t$.
Harmonics of $\Omega_\theta$ and $\Omega_r$ can play no role in any
physics arising from these orbits, so the index set $\{lmkn\}$ reduces
to $\{lm\}$, and the mode frequency $\omega_{mkn}$ to $\omega_m$. We
will consider general orbits in a later analysis.
Before tackling general black hole spin, it is useful to examine
Eq.\ (\ref{eq:gauss_kerr}) for Schwarzschild black holes. Several
simplifications occur when $a = 0$:
\begin{itemize}
\item The radius $r_+ = 2M$; the frequency $\Omega_{\rm H} = 0$, so
the wavenumber $p_m = \omega_m$; the factor $\epsilon = 1/8M$; the
phase factor $K(a) = 0$ [cf.\ Eq.\ (\ref{eq:K_of_a})]; and the ingoing
axial coordinate $\psi = \phi$.
\item The spin-weighted spheroidal harmonic becomes a spin-weighted
spherical harmonic: ${_{+2}S}_{lm}(\theta)
\to\ {_{+2}Y}_{lm}(\theta)$. The eigenvalue of the angular function
therefore simplifies, as does the complex number $c_{lm}$: ${\cal E}
= l(l+1)$, and $c_{lm} = (l+2)(l+1)l(l-1) + 12iM\omega_m$.
\item The angular operator $\bar\eth \equiv \bar\eth_0 =
1/(2\sqrt{2}M)L^s_-$. Using Eq.\ (\ref{eq:sphericalharmoniclower}),
we have
\begin{equation}
L^s_-L^s_-\, {_{+2}}Y_{lm}(\theta) =
\sqrt{(l+2)(l+1)l(l-1)}\,{_0}Y_{lm}\;,
\end{equation}
which tells us that
\begin{equation}
\bar\eth\bar\eth S^+_{lm}(\theta) =
\frac{1}{8M^2} \sqrt{(l+2)(l+1)l(l-1)}\,{_0}Y_{lm}
\label{eq:barethbarethS_schw}
\end{equation}
for $a = 0$.
\end{itemize}
Putting all of this together, for $a = 0$ we have
\begin{eqnarray}
R^{(1)}_{{\rm H},lm} &=& {\rm Im}\left[ {\cal C}_{lm} Z^{\rm H}_{lm}
e^{i\Phi_m}\right] \times
\nonumber\\
& &\!\!\!\!\!\!\!\!\!\!\!
\frac{1}{8M^2}\sqrt{(l+2)(l+1)l(l-1)}\,{_0}Y_{lm}(\theta)\;,
\label{eq:gauss_schw}
\end{eqnarray}
where
\begin{eqnarray}
{\cal C}_{lm} &=& \frac{1024M^2(iM\omega_m - 1/4)(M\omega_m +
i/2)}{(l+2)(l+1)l(l-1) + 12iM\omega_m}\;,
\label{eq:Clm_schw}\\
\Phi_m &=& m\phi - \omega_m v\;.
\end{eqnarray}
\subsection{Slow motion: Analytic results}
\label{sec:schw_analytic}
We begin our analysis of the Schwarzschild tidal deformations by
expanding all quantities in orbital speed $u \equiv (M/r_{\rm
o})^{1/2}$. We take all relevant quantities to $O(u^5)$ beyond the
leading term; this is far enough to see how the curvature behaves for
multipole index $l \le 4$. These results should be accurate for
weak-field orbits, when $u \ll 1$. In the following subsection, we
will compare with numerical results that are good into the strong
field.
Begin with ${\cal C}_{lm}$. Expanding Eq.\ (\ref{eq:Clm_schw}), we
find
\begin{eqnarray}
{\cal C}_{2m} &=& -\frac{16i}{3}M^2\exp\left(-\frac{13}{2}imu^3\right)\;,
\label{eq:C2m_schw}\\
{\cal C}_{3m} &=& -\frac{16i}{15}M^2\exp\left(-\frac{61}{10}
imu^3\right)\;,
\label{eq:C3m_schw}\\
{\cal C}_{4m} &=& -\frac{16i}{45}M^2
\exp\left(-\frac{181}{30}imu^3\right)\;.
\label{eq:C4m_schw}
\end{eqnarray}
To perform this expansion, we used the fact that, for $a = 0$,
$M\Omega_{\rm orb} = u^3$, so $M\omega_m = mu^3$.
Next, we construct analytic expansions for the amplitudes $Z^{\rm
H}_{lm}$, following the algorithm described in Sec.\ {\ref{sec:ZH}}.
All the results which follow are understood to neglect contributions
of $O(u^6)$ and higher. We also introduce $\mu$, the mass of the
small body whose tides deform the black hole.
For $l = 2$, the amplitudes are
\begin{eqnarray}
Z^{\rm H}_{20} &=& \sqrt{\frac{3\pi}{10}}\frac{\mu}{r_{\rm o}^3} \left(1 +
\frac{7}{2}u^2 + \frac{561}{56}u^4\right)\;,
\label{eq:ZHschw20}\\
Z^{\rm H}_{21} &=& -3 i \sqrt{\frac{\pi }{5}}\frac{\mu}{r_{\rm o}^3}
\left(u + \frac{8}{3} u^3 + \frac{10i}{3} u^4
+ \frac{152}{21} u^5\right)
\nonumber\\
&=& -3i\sqrt{\frac{\pi}{5}}\frac{\mu}{r_{\rm o}^3}
\left(u + \frac{8}{3} u^3 + \frac{152}{21} u^5\right)
\nonumber\\
& &\qquad\qquad\quad\times\,
\exp\left(\frac{10}{3}iu^3\right)\;,
\label{eq:ZHschw21}
\\
Z^{\rm H}_{22} &=& -\frac{3}{2}\sqrt{\frac{\pi }{5}}\frac{\mu}{r_{\rm o}^3}
\biggl(1 + \frac{3}{2}u^2 + \frac{23i}{3}u^3 + \frac{1403}{168}u^4
\nonumber\\
& &\qquad\qquad\qquad\qquad
+ \frac{473i}{30}u^5\biggr)
\nonumber\\
&=& -\frac{3}{2}\sqrt{\frac{\pi }{5}}\frac{\mu}{r_{\rm o}^3}
\left(1 + \frac{3}{2}u^2 + \frac{1403}{168}u^4\right)
\nonumber\\
& &\qquad\qquad\quad \times\,
\exp\left[i\left(\frac{23}{3}u^3 + \frac{64}{15}u^5\right)\right]\;.
\nonumber\\
\label{eq:ZHschw22}
\end{eqnarray}
For $l = 3$,
\begin{eqnarray}
Z^{\rm H}_{30} &=& -i \sqrt{\frac{30 \pi }{7}}\frac{\mu}{r_{\rm o}^3}
\left(u^3 + 4 u^5\right)\;,
\label{eq:ZHschw30}\\
Z^{\rm H}_{31} &=& -\frac{3}{2}\sqrt{\frac{5\pi}{14}}\frac{\mu}{r_{\rm o}^3}
\left(u^2 + \frac{13}{3} u^4 + \frac{43i}{30}u^5\right)
\nonumber\\
&=& -\frac{3}{2}\sqrt{\frac{5\pi}{14}}\frac{\mu}{r_{\rm o}^3}
\left(u^2 + \frac{13}{3} u^4\right)
\exp\left(\frac{43}{30}iu^3\right)\;,
\nonumber\\
\label{eq:ZHschw31}\\
Z^{\rm H}_{32} &=& 5 i \sqrt{\frac{\pi }{7}}\frac{\mu}{r_{\rm o}^3}
\left(u^3 + 4u^5\right)\;,
\label{eq:ZHschw32}\\
Z^{\rm H}_{33} &=& \frac{5}{2}\sqrt{\frac{3 \pi }{14}}\frac{\mu}{r_{\rm o}^3}
\left(u^2 + 3u^4 + \frac{43i}{10}u^5\right)
\nonumber\\
&=& \frac{5}{2}\sqrt{\frac{3 \pi }{14}}\frac{\mu}{r_{\rm o}^3}
\left(u^2 + 3u^4\right)\exp\left(\frac{43}{10}iu^3\right)\;.
\nonumber\\
\label{eq:ZHschw33}
\end{eqnarray}
Finally, for $l = 4$,
\begin{eqnarray}
Z^{\rm H}_{40} &=& -\frac{9}{14}\sqrt{\frac{5\pi}{2}}\frac{\mu}{r_{\rm o}^3}
u^4\;,
\label{eq:ZHschw40}\\
Z^{\rm H}_{41} &=& \frac{45i}{14} \sqrt{\frac{\pi }{2}}\frac{\mu}{r_{\rm o}^3}
u^5\;,
\label{eq:ZHschw41}\\
Z^{\rm H}_{42} &=& \frac{15}{14}\sqrt{\pi}\frac{\mu}{r_{\rm o}^3}
u^4\;,
\label{eq:ZHschw42}\\
Z^{\rm H}_{43} &=& -\frac{15i}{2} \sqrt{\frac{\pi}{14}}\frac{\mu}{r_{\rm o}^3}
u^5\;,
\label{eq:ZHschw43}\\
Z^{\rm H}_{44} &=& -\frac{15}{4}\sqrt{\frac{\pi}{7}}\frac{\mu}{r_{\rm o}^3}
u^4\;.
\label{eq:ZHschw44}
\end{eqnarray}
Note that $Z^{\rm H}_{l-m} = (-1)^l \bar Z^{\rm H}_{lm}$, where
overbar denotes complex conjugation.
\begin{widetext}
It is particularly convenient to combine the modes in pairs, examining
$R^{(1)}_{{\rm H},l-m} + R^{(1)}_{{\rm H},lm}$. Doing so, we find for
$l = 2$,
\begin{eqnarray}
R^{(1)}_{{\rm H},20} &=& -\frac{\mu}{r_{\rm o}^3}
\left(3\cos^2\theta - 1\right)
\left(1 + \frac{7}{2}u^2 + \frac{561}{56}u^4\right)\;,
\label{eq:R20schw}\\
R^{(1)}_{{\rm H},1-1} + R^{(1)}_{{\rm H},11} &=& 0\;,
\label{eq:R21schw}\\
R^{(1)}_{{\rm H},2-2} + R^{(1)}_{{\rm H},22} &=&
\frac{3\mu}{r_{\rm o}^3}\sin^2\theta \left(1 + \frac{3}{2}u^2 +
\frac{1403}{168}u^4\right) \cos\left[2\left(\phi - \Omega_{\rm orb} v
- \frac{8}{3}u^3 + \frac{32}{5}u^5\right)\right]\;.
\label{eq:R22schw}
\end{eqnarray}
For $l = 3$, we have
\begin{eqnarray}
R^{(1)}_{{\rm H},30} &=& 0\;,
\label{eq:R30schw}\\
R^{(1)}_{{\rm H},3-1} + R^{(1)}_{{\rm H},31} &=&
\frac{3}{2}\frac{\mu}{r_{\rm o}^3} \sin\theta\left(1 -
5\cos^2\theta\right) u^2\left(1 + \frac{13}{3}u^2\right)
\cos\left[\phi - \Omega_{\rm orb} v - \frac{14}{3}u^3\right]\;,
\label{eq:R31schw}\\
R^{(1)}_{{\rm H},3-2} + R^{(1)}_{{\rm H},32} &=& 0\;,
\label{eq:R32schw}\\
R^{(1)}_{{\rm H},3-3} + R^{(1)}_{{\rm H},33} &=&
\frac{5}{2}\frac{\mu}{r_{\rm o}^3}\sin^3\theta u^2\left(1 +
3u^2\right) \cos\left[3\left(\phi - \Omega_{\rm orb} v -
\frac{14}{3}u^3\right)\right]\;. \nonumber\\
\label{eq:R33schw}
\end{eqnarray}
And, for $l = 4$,
\begin{eqnarray}
R^{(1)}_{{\rm H},40} &=& \frac{9}{56}\frac{\mu}{r_{\rm o}^3}
\left(3 - 30\cos^2\theta + 35\cos^4\theta\right)u^4\;,
\label{eq:R40schw}\\
R^{(1)}_{{\rm H},4-1} + R^{(1)}_{{\rm H},41} &=& 0\;,
\label{eq:R41schw}\\
R^{(1)}_{{\rm H},4-2} + R^{(1)}_{{\rm H},42} &=&
\frac{15}{14}\frac{\mu}{r_{\rm o}^3}\sin^2\theta \left(1 -
7\cos^2\theta\right)u^4 \cos\left[2\left(\phi - \Omega_{\rm orb} v -
\frac{181}{30}u^3\right)\right]\;,
\label{eq:R42schw}\\
R^{(1)}_{{\rm H},4-3} + R^{(1)}_{{\rm H},43} &=& 0\;,
\label{eq:R43schw}\\
R^{(1)}_{{\rm H},4-4} + R^{(1)}_{{\rm H},44} &=&
\frac{15}{8}\frac{\mu}{r_{\rm o}^3}\sin^4\theta u^4
\cos\left[4\left(\phi - \Omega_{\rm orb} v - \frac{181}{30}u^3\right)
\right]\;.
\label{eq:R44schw}
\end{eqnarray}
\end{widetext}
In the next section, we will compare Eqs.\ (\ref{eq:R20schw}) --
(\ref{eq:R44schw}) with strong-field numerical calculations. Before
doing so, we examine some consequences of these results and compare
with earlier literature.
\subsubsection{Nearly static limit}
In Ref.\ {\cite{hartle74}}, Hartle examines the deformation of a black
hole due to a nearly static orbiting moon. To reproduce his results,
consider the $u \to 0$ limit of Eqs.\ (\ref{eq:R20schw}) --
(\ref{eq:R44schw}). Only the $l = 2$, $m = 0$, $m = \pm2$
contributions remain when $u \to 0$. Adding these contributions, we
find
\begin{equation}
R^{(1)}_{\rm H} = -\frac{\mu}{r_{\rm o}^3} \left[3\cos^2\theta -
3\sin^2\theta\cos\left(2\phi'\right) - 1\right]\;,
\label{eq:R1static}
\end{equation}
where $\phi' = \phi - \Omega_{\rm orb} v$ is the azimuthal coordinate
of the orbiting moon. Hartle writes\footnote{Note that the result
Hartle presents in Ref.\ {\cite{hartle74}} contains a sign error.
This can be seen by computing the curvature associated with the
metric he uses on the horizon [Eqs.\ (5.10) and (5.13) of
Ref.\ {\cite{hartle73}}]. The embedding surface Hartle uses,
Eq.\ (4.33) of Ref.\ {\cite{hartle74}} [or (5.14) of
Ref.\ {\cite{hartle73}}] is correct given this metric.} his result
\begin{equation}
R^{(1)}_{\rm Hartle} = \frac{4\mu}{r_{\rm o}^3}P_2(\cos\chi) =
\frac{2\mu}{r_{\rm o}^3}\left(3\cos^2\chi - 1\right)\;,
\label{eq:R1Hartle}
\end{equation}
where ``$\chi$ is the angle between the point of interest and the
direction to the moon'' [Ref.\ {\cite{hartle74}}, text following
Eq.\ (4.32)]. The angle $\chi$ can be interpreted as $\theta$ if we
place Hartle's moon at $\theta_{\rm moon} = 0$. To compare the two
solutions, we must rotate. One way to do this rotation is to note
that the equatorial plane in our calculation ($\theta = \pi/2$) should
vary with $\phi'$ as Hartle's result varies with $\chi$. Put $\theta
= \pi/2$ and $\phi' = \chi$ in Eq.\ (\ref{eq:R1static}):
\begin{eqnarray}
R^{(1)}_{\rm H}\bigr|_{\theta = \pi/2, \phi' = \chi}
&=& \frac{\mu}{r_{\rm o}^3}\left(3\cos2\chi + 1\right)
\nonumber\\
&=& \frac{2\mu}{r_{\rm o}^3}\left(3\cos^2\chi - 1\right)\;.
\end{eqnarray}
Another way to compare is to note that the $\phi' = 0$ circle should
vary with angle in a way that duplicates Hartle's result, modulo a
shift in angle, $\theta = \chi + \pi/2$:
\begin{eqnarray}
R^{(1)}_{\rm H}\bigr|_{\theta = \chi + \pi/2, \phi' = 0} &=&
-\frac{\mu}{r_{\rm o}^3}\bigl[3\cos^2(\chi + \pi/2)
\nonumber\\
& &\quad - 3\sin^2(\chi + \pi/2) + 1\bigr]
\nonumber\\
&=& -\frac{\mu}{r_{\rm o}^3}\left(3\sin^2\chi - 3\cos^2\chi + 1\right)
\nonumber\\
&=& \frac{2\mu}{r_{\rm o}^3}\left(3\cos^2\chi - 1\right)\;.
\end{eqnarray}
Both forms reproduce Hartle's static limit.
\subsubsection{Embedding the quadrupolar distortion}
\label{sec:l=2embed}
At various places in this paper, we will examine the geometry of a
distorted horizon by embedding it in a 3-dimensional Euclidean space.
The details of this calculation are given in Appendix
{\ref{app:embed}}; equivalent discussion for Schwarzschild, where the
results are particularly clean, is also given in Ref.\ {\cite{vpm11}}.
Briefly, a Schwarzschild horizon that has been distorted by a tidal
field has the scalar curvature of a spheroid of radius
\begin{equation}
r_{\rm E} = 2M\left[1 + \sum_{lm}\varepsilon_{lm}(\theta,\phi)\right]\;,
\end{equation}
where, as shown in Appendix {\ref{app:embed_schw}} and
Ref.\ {\cite{vpm11}},
\begin{equation}
\varepsilon_{lm} = \frac{2M^2}{(l + 2)(l - 1)}R^{(1)}_{{\rm H},lm}\;.
\label{eq:schw_spheroid}
\end{equation}
By considering a Schwarzschild black hole embedded in a universe
endowed with post-Newtonian tidal fields, Taylor and Poisson
{\cite{tp08}} compute $\varepsilon_{lm}$ in a post-Newtonian
framework. Specializing to the tides appropriate to a binary system,
they find
\begin{eqnarray}
& &\sum_m \varepsilon_{2m}(\theta,\phi) = \frac{\mu}{b^3}\frac{M^2}{2}
\left(1 + \frac{1}{2}u^2\right)(1 - \cos^2\theta)
\nonumber\\
& & \quad\quad + \frac{3\mu}{b^3}\frac{M^2}{2}\left(1 -
\frac{3}{2}u^2\right)\sin^2\theta \cos\left[2(\phi - \Omega_{\rm
orb}v)\right]\;.
\nonumber\\
\label{eq:taylorpoisson1}
\end{eqnarray}
This is Eq.\ (8.8) of Ref.\ {\cite{tp08}}, with $M_2 \to \mu$, $M_1
\to M$, $v_{\rm rel}/c \to u$, and expanded to leading order in $\mu$.
Their parameter $b$ is the separation of the binary in harmonic
coordinates. Using the fact that $r_{\rm H} = r_{\rm S} - M$ (with
``H'' and ``S'' subscripts denoting harmonic and Schwarzschild,
respectively), it is easy to convert to $r_{\rm o}$, our separation in
Schwarzschild coordinates:
\begin{eqnarray}
\frac{1}{b^3} = \frac{1}{r_{\rm o}^3(1 - M/r_{\rm o})^3}
\simeq \frac{1}{r_{\rm o}^3}\left(1 + 3u^2\right)\;.
\end{eqnarray}
Replacing $b$ for $r_{\rm o}$ and truncating at $O(u^2)$,
Eq.\ (\ref{eq:taylorpoisson1}) becomes
\begin{eqnarray}
& &\sum_m \varepsilon_{2m}(\theta,\phi) = \frac{\mu}{r_{\rm o}^3}\frac{M^2}{2}
\left(1 + \frac{7}{2}u^2\right)(1 - \cos^2\theta)
\nonumber\\
& & \quad\quad + \frac{3\mu}{r_{\rm o}^3}\frac{M^2}{2}\left(1 +
\frac{3}{2}u^2\right)\sin^2\theta \cos\left[2(\phi - \Omega_{\rm
orb}v)\right]\;.
\nonumber\\
\label{eq:taylorpoisson2}
\end{eqnarray}
Comparing with Eqs.\ (\ref{eq:R20schw}) and (\ref{eq:R22schw}) and
correcting for the factor $M^2/2$ which converts curvature
$R^{(1)}_{{\rm H},2m}$ to $\varepsilon_{2m}$, we see agreement to
$O(u^2)$.
\subsubsection{Phase of the tidal bulge}
\label{sec:schw_bulge}
Using these analytic results, let us examine the notions of bulge
phase introduced in Sec.\ {\ref{sec:bulge}}. First consider the
position of the bulge versus the position of the orbit according to
the null and instantaneous maps (which are identical for
Schwarzschild), Eq.\ (\ref{eq:bulge_vs_orbit_null}). The various
modes which determine the shape of the horizon all peak at angle $\phi
= \Omega_{\rm orb}v + \delta\phi(u)$, where $\delta\phi(u)$ can be
read out of Eqs.\ (\ref{eq:R20schw})--(\ref{eq:R44schw}). For
Schwarzschild $\bar r = 0$, and the ingoing angle $\psi = \phi$. The
orbit's position mapped onto the horizon is $\phi_{\rm o}^{\rm NM} =
\phi_{\rm o} = \Omega_{\rm orb} v + \Delta\phi(r_{\rm o})$, where
\begin{equation}
\Delta\phi(r_{\rm o}) = -\Omega_{\rm orb} r^*_{\rm o}
\end{equation}
is Eq.\ (\ref{eq:Deltapsi_orb}) for $a = 0$. The result for the
bulge's offset from the orbit is
\begin{eqnarray}
\delta\phi^{\rm OB}_{22} &=& \frac{8}{3}u^3 - \frac{32}{5}u^5 -
\Delta\phi(r_{\rm o})\;,
\label{eq:phi_ob_2}\\
\delta\phi^{\rm OB}_{31} &=& \delta\phi^{\rm OB}_{33} =
\frac{14}{3}u^3 - \Delta\phi(r_{\rm o})\;,
\label{eq:phi_ob_3}\\
\delta\phi^{\rm OB}_{42} &=& \delta\phi^{\rm OB}_{44} =
\frac{181}{30}u^3 - \Delta\phi(r_{\rm o})\;.
\label{eq:phi_ob_4}
\end{eqnarray}
For the multipoles which we do not include here, no useful notion of
bulge position exists: for $m = 0$ the bulge is axisymmetric, and for
the others, the bulge's amplitude is zero to this order. Our results
for $l = |m| = 2$ agree with Fang and Lovelace; cf.\ Eq.\ (4) of
Ref.\ {\cite{fl05}}.
Consider next the relative phase of the tidal bulge and the perturbing
field, Eq.\ (\ref{eq:schw_bulge_vs_tide}). For small $u$, we have
\begin{equation}
\delta\phi^{\rm TD}_{m} = 4mu^3\;.
\end{equation}
This again agrees with Fang and Lovelace --- compare Eq.\ (6) of
Ref.\ {\cite{fl05}}, bearing in mind that $m$ is built into their
definition of the offset angle [their Eq.\ (50)], and that they fix $m
= 2$.
In both cases, note that the bulge's offset is a positive phase. This
indicates that the bulge leads both the orbiting body's instantaneous
position, as well as the tidal field that sources the tidal
deformation. As discussed in the Introduction, this is consistent
with past work, and is a consequence of the horizon's teleological
nature.
\subsection{Fast motion: Numerical results}
\label{sec:schw_numerical}
Our numerical results for Schwarzschild black holes are summarized by
Figs.\ {\ref{fig:schw_converge}}, {\ref{fig:schw_Rvsphi}}, and
{\ref{fig:a0.0_shapes}}. We compute $R^{(1)}_{\rm H}$ by solving for
$Z^{\rm H}_{lm}$ numerically as described in Sec.\ {\ref{sec:ZH}}, and
then applying Eq.\ (\ref{eq:gauss_schw}). All of our results
illustrate quantities computed in the black hole's equatorial plane,
$\theta = \pi/2$. We include all contributions up to $l = 15$ in the
sum. Figure {\ref{fig:schw_converge}} shows that contributions to the
horizon's scalar curvature converge quite rapidly. The contributions
from $l = 15$ are about $10^{-9}$ of the total for the most extreme
case we consider here, $r_{\rm o} = 6M$.
\begin{figure}[ht]
\includegraphics[width = 0.48\textwidth]{schw_converge.pdf}
\caption{Convergence of contributions to the horizon's tidal
distortion. We show $R^{(1)}_{{\rm H},lm}$ summed over $m$ for a
given $l$, scaled by a factor $(r_{\rm o}^3/\mu)$ to account for the
leading dependence on small body mass and orbital radius. The
largest amplitude oscillation is for $l = 2$ (red in color). The
next largest is $l = 3$ (green), followed by $l = 4$ (blue), $l = 5$
(magenta), with the smallest oscillations shown for $l = 6$ (cyan).
(Higher order contributions are omitted since their variations
cannot be seen on the scale of this plot.) These curves are for a
circular orbit at $r_{\rm o} = 6M$, which has $u = 0.41$, the
largest value for the Schwarzschild cases we consider. As such,
this case has the slowest convergence among Schwarzschild orbits.
The falloff with $l$ is more rapid for all other cases.}
\label{fig:schw_converge}
\end{figure}
Figure {\ref{fig:schw_Rvsphi}} compares the analytic predictions for
$R^{\rm H}$ [Eqs.\ (\ref{eq:R20schw})--(\ref{eq:R44schw})] with
numerical results for $l = 2$, $l = 3$, and $l = 4$, and for two
different orbital radii ($r_{\rm o} = 50M$ and $6M$). The agreement
is outstanding for the large radius orbit. Our numerical and analytic
predictions can barely be distinguished at $l = 2$ and $l = 3$, and
differ by about $10\%$ at maximum for $l = 4$ (where our analytic
formula includes only the leading contribution to the curvature). The
agreement is much poorer at small radius. At $r_{\rm o} = 6M$,
disagreement is several tens of percent for $l = 2$, rising to a
factor $\sim 5$ for $l = 4$. For both the large and small radius
cases we show, the sum over modes is dominated by the contribution
from $l = 2$. The phase agreement between analytic and numerical
formulas is quite good all the way into the strong field, even when
the amplitudes differ significantly.
\begin{figure*}[ht]
\includegraphics[width = 0.48\textwidth]{schw_comp50.pdf}
\includegraphics[width = 0.48\textwidth]{schw_comp6.pdf}
\caption{Comparison of numerically computed scalar curvature
perturbation $R^{(1)}_{\rm H}$ for Schwarzschild with the analytic
expansion given in Eqs.\ (\ref{eq:R20schw})--(\ref{eq:R44schw}).
The four panels on the left compare numerical (dark gray curves; red
in color) and analytic (light gray; green in color) results for an
orbit at $r_{\rm o} = 50M$. Panels on the right are for $r_{\rm o}
= 6M$. In both cases, we plot $(r_{\rm o}^3/\mu)R^{(1)}_{\rm H}$,
scaling out the leading dependence on orbital radius and the
orbiting body's mass. We show contributions for $l = 2$, $l = 3$,
and $l = 4$, plus the sum of these modes. For $r_{\rm o} = 50M$, we
have $u = 0.14$, and we see very good agreement between the
numerical and analytic formulas. In several cases, the numerical
data lie on top of the analytic curves. For $r_{\rm o} = 6M$, $u =
0.41$, and the agreement is not as good. Although the amplitudes
disagree in the strong field (especially for large $l$), the two
computations maintain good phase agreement well into the strong
field.}
\label{fig:schw_Rvsphi}
\end{figure*}
\begin{figure}[ht]
\includegraphics[width = 0.48\textwidth]{a00_shapes.pdf}
\caption{Equatorial section of the embedding of a distorted
Schwarzschild horizon. Each panel shows the distortion for a
different orbital radius, varying from $r_{\rm o} = 50M$ to $r_{\rm
o} = 6M$. The black circles are the undistorted black hole, and
the red curves are the distorted horizons, embedded with
Eq.\ (\ref{eq:schw_spheroid}). These plots are in a frame that
corotates with the orbit, and are for a slice of constant ingoing
time $v$. The green dashed line in each panel shows the angle at
which the tidal distortion is largest; the black dotted line shows
the orbit's position. Notice that the bulge leads the orbit in all
cases, with the lead angle growing as the orbit moves to smaller
orbital radius. We have rescaled the horizon's tidal distortion by
a factor $\propto r_{\rm o}^3/\mu$ so that, at leading order, the
magnitude of the distortion is the same in all plots.}
\label{fig:a0.0_shapes}
\end{figure}
Figure {\ref{fig:a0.0_shapes}} shows distorted black holes by
embedding the horizon in a 3-dimensional Euclidean space, as discussed
in Sec.\ {\ref{sec:l=2embed}}. Now, we do not truncate at $l = 2$,
but include all moments that we calculate. We show the equatorial
slices of our embeddings for several different circular orbits
($r_{\rm o} = 50M$, $20M$, $10M$, and $6M$). In all of our plots, we
scale the horizon distortion $\varepsilon_{lm}$ by a factor
proportional to $r_{\rm o}^3/\mu$ so that the tide's impact is of
roughly the same magnitude for all orbital separations.
The embeddings are shown in a frame that corotates with the orbit at
an instant $v = {\rm constant}$. The $x$-axis is at $\phi = 0$, so
the orbiting body sits at $\phi = \Delta\phi(r_{\rm o}) = -\Omega_{\rm
orb}r^*_{\rm o}$. In each panel, we have indicated where the radius
of the embedding is largest (green dashed line, showing the angle of
greatest tidal distortion) and the angular position of the orbiting
body (black dotted line). In all cases, the bulge leads the orbiting
body's position, just as predicted in Sec.\ {\ref{sec:schw_bulge}}.
The numerical value of the bulge's position relative to the orbit,
$\delta\phi^{\rm num}$, agrees quite well with $\delta\phi^{\rm
OB}_{22}$, Eq.\ (\ref{eq:phi_ob_2}) From
Fig.\ {\ref{fig:a0.0_shapes}}, we have
\begin{eqnarray}
\delta\phi^{\rm num} &=& 9.56^\circ\quad r_{\rm o} = 50M\;,
\nonumber\\
&=& 17.3^\circ\quad r_{\rm o} = 20M\;,
\nonumber\\
&=& 27.8^\circ\quad r_{\rm o} = 10M\;,
\nonumber\\
&=& 37.6^\circ\quad r_{\rm o} = 6M\;.
\end{eqnarray}
Equation (\ref{eq:phi_ob_2}) tells us
\begin{eqnarray}
\delta\phi^{\rm OB}_{22}
&=& 9.54^\circ\quad r_{\rm o} = 50M\;,
\nonumber\\
&=& 17.1^\circ\quad r_{\rm o} = 20M\;,
\nonumber\\
&=& 26.8^\circ\quad r_{\rm o} = 10M\;,
\nonumber\\
&=& 35.0^\circ\quad r_{\rm o} = 6M\;.
\end{eqnarray}
In all cases, the true position of the bulge is slightly larger than
$\delta\phi^{\rm OB}_{22}$. This appears to be due in large part to
the contribution of modes other than $l = |m| = 2$; the agreement
improves if we calculate $\delta\phi^{\rm num}$ using only the $l = 2$
contribution to the embedding.
\section{Results II: Kerr}
\label{sec:kerr_results}
Now consider non-zero black hole spin. We begin with slow motion and
small black hole spin, expanding Eq.\ (\ref{eq:gauss_kerr}) using $u
\equiv (M/r_{\rm o})^{1/2} \ll 1$ and $q \equiv a/M \ll 1$, and derive
analytic results which are useful points of comparison to the general
case. We then show numerical results which illustrate tidal
deformations for strong-field orbits.
\subsection{Slow motion: Analytic results}
\label{sec:kerr_analytic}
Here we present analytic results, expanding in powers of $u =
(M/r_{\rm o})^{1/2}$ and $q = a/M$. We take all relevant quantities
to order $u^5$ and $q$ beyond the leading term; this is far enough to
see how quantities behave for $l \le 4$. We compare with strong-field
numerical results in the following subsection.
\begin{widetext}
Begin again with ${\cal C}_{lm}$. Neglect the $k$ and $n$ indices
which are irrelevant for circular, equatorial orbits, and expand
$\lambda = \lambda_0 + (a\omega_m)\lambda_1$, with $\lambda_0$ and
$\lambda_1$ given by Eqs.\ (\ref{eq:lambda0}) and (\ref{eq:lambda1})
for $s = -2$ [recall that $\lambda$ comes from the spheroidal harmonic
$S^-_{lm}(\theta)$]. Finally, expand to $O(u^5)$ and $O(q)$. Doing
so, Eq.\ (\ref{eq:Clmkn}) yields
\begin{eqnarray}
{\cal C}_{2m} &=& -\frac{16i}{3}M^2\left(1 - \frac{13}{3}qm^2u^3\right)
\exp\left[-im\left(\frac{13}{2}u^3 - \frac{3}{2}q\right)\right]\;,
\label{eq:C2m_kerr}\\
{\cal C}_{3m} &=& -\frac{16i}{15}M^2\left(1 - \frac{14}{3}qm^2u^3\right)
\exp\left[-im\left(\frac{61}{10}u^3 - \frac{3}{2}q\right)\right]\;,
\label{eq:C3m_kerr}\\
{\cal C}_{4m} &=& -\frac{16i}{45}M^2\left(1 - \frac{24}{5}qm^2u^3\right)
\exp\left[-im\left(\frac{181}{30}u^3 - \frac{3}{2}q\right)\right]\;.
\label{eq:C4m_kerr}
\end{eqnarray}
These reduce to the Schwarzschild results, Eqs.\ (\ref{eq:C2m_schw})
-- (\ref{eq:C4m_schw}), when $q \to 0$.
Next, the amplitudes $Z^{\rm H}_{lm}$, again following the algorithm
described in Sec.\ {\ref{sec:ZH}}. These results should be understood
to neglect contributions of $O(u^6)$, $O(q^2)$ and higher. As
elsewhere, $\mu$ is the mass of the smaller body. For $l = 2$, we
have
\begin{eqnarray}
Z^{\rm H}_{20} &=& \sqrt{\frac{3 \pi }{10}}\frac{\mu}{r_{\rm
o}^3}\left(1 + \frac{7}{2}u^2 - 4 q u^3 + \frac{561}{56} u^4 - 18
q u^5\right)\;,
\label{eq:ZHkerr20}\\
Z^{\rm H}_{21} &=& -3 i \sqrt{\frac{\pi }{5}}\frac{\mu}{r_{\rm o}^3}
\biggl\{\left(1 - \frac{i}{2}q\right)u - \frac{2}{3}qu^2 +
\left(\frac{8}{3}-\frac{4i}{3}q\right)u^3 + \left[\frac{10i}{3} +
\left(\frac{1}{6} - \frac{\pi^2}{3}\right)q\right]u^4 +
\left(\frac{152}{21} - \frac{368}{63}q\right)u^5\biggr\}
\nonumber\\
&=& -3i\sqrt{\frac{\pi}{5}}\frac{\mu}{r_{\rm o}^3} \biggl[u -
\frac{2}{3}qu^2 + \frac{8}{3} u^3 - \left(\frac{3}{2} +
\frac{\pi^2}{3}\right)qu^4 +
\frac{152}{21}u^5\biggr]\exp\left[i\left(\frac{10}{3}u^3 -
\frac{q}{2}\right)\right]\;,
\label{eq:ZHkerr21}\\
Z^{\rm H}_{22} &=& -\frac{3}{2} \sqrt{\frac{\pi}{5}}\frac{\mu}{r_{\rm
o}^3} \biggl\{1 - i q + \left(\frac{3}{2} -
\frac{3i}{2}q\right)u^2 + \left[\frac{23i}{3} + \left(15 -
\frac{4\pi^2}{3}\right)q\right]u^3 + \left(\frac{1403}{168} -
\frac{1403i}{168}q\right)u^4
\nonumber\\
& & + \left[\frac{473i}{30} + \left(\frac{2449}{90} -
2\pi^{2}\right)q\right]u^5\biggr\}
\nonumber\\
&=& -\frac{3}{2} \sqrt{\frac{\pi}{5}}\frac{\mu}{r_{\rm o}^3} \biggl[1
+ \frac{3}{2}u^2 + \left(\frac{22}{3} -\frac{4\pi ^2}{3}\right)q u^3
+ \frac{1403}{168}u^4 + \left(\frac{103}{9} -
2\pi^2\right)qu^5\biggr]\exp\left[i\left(\frac{23}{3}u^3 +
\frac{64}{15}u^5 - q\right)\right]\;.
\nonumber\\
\label{eq:ZHkerr22}
\end{eqnarray}
For $l = 3$,
\begin{eqnarray}
Z^{\rm H}_{30} &=& -i\sqrt{\frac{30\pi}{7}}\frac{\mu}{r_{\rm o}^3}
\left(u^3 - \frac{3}{4}qu^4 + 4u^5\right)\;,
\label{eq:ZHkerr30}
\\
Z^{\rm H}_{31} &=&
-\frac{3}{2}\sqrt{\frac{5\pi}{14}}\frac{\mu}{r_{\rm o}^3}
\left\{\left(1 - \frac{i}{6}q\right)u^2 + \left(\frac{13}{3} -
\frac{13i}{18}q\right)u^4 + \left[\frac{43i}{30} -
\left(\frac{247}{180} + \frac{\pi^2}{3}\right)q\right]u^5\right\}
\nonumber\\
&=& -\frac{3}{2}\sqrt{\frac{5\pi}{14}}\frac{\mu}{r_{\rm o}^3}
\left[u^2 + \frac{13}{3}u^4 - \left(\frac{29}{18} +
\frac{\pi^2}{3}\right)qu^5\right]\exp\left[i\left(\frac{43}{30}u^3 -
\frac{q}{6}\right)\right]\;,
\label{eq:ZHkerr31}
\\
Z^{\rm H}_{32} &=& 5i\sqrt{\frac{\pi}{7}}\frac{\mu}{r_{\rm o}^3}
\left[\left(1 - \frac{i}{3}q\right)u^3 - \frac{3}{4}qu^4 + 4\left(1 -
\frac{i}{3}q\right)u^5\right]\nonumber\\
&=& 5i\sqrt{\frac{\pi}{7}}\frac{\mu}{r_{\rm o}^3}\left(u^3 -
\frac{3}{4}qu^4 + 4 u^5\right)\exp\left(-iq/3\right)\;,
\label{eq:ZHkerr32}
\\
Z^{\rm H}_{33} &=& \frac{5}{2}\sqrt{\frac{3\pi}{14}}\frac{\mu}{r_{\rm
o}^3} \left\{\left(1 - \frac{i}{2}q\right)u^2 + \left(3 -
\frac{3i}{2}q\right)u^4 + \left[\frac{43i}{10} +
\left(\frac{393}{20} - 3\pi^2\right)q\right]u^5\right\}
\nonumber\\
&=& \frac{5}{2}\sqrt{\frac{3\pi}{14}}\frac{\mu}{r_{\rm o}^3} \left[u^2
+ 3u^4 + \left(\frac{35}{2} - 3\pi^2\right)qu^5\right]
\exp\left[i\left(\frac{43}{10}u^3 - \frac{q}{2}\right)\right]\;.
\label{eq:ZHkerr33}
\end{eqnarray}
And for $l = 4$,
\begin{eqnarray}
& &Z^{\rm H}_{40} =
-\frac{9}{14}\sqrt{\frac{5\pi}{2}}\frac{\mu}{r_{\rm o}^3} u^4\;,
\label{eq:ZHkerr40}
\\
& &Z^{\rm H}_{41} = \frac{45i}{14}\sqrt{\frac{\pi}{2}}
\frac{\mu}{r_{\rm o}^3} \left[\left(1 + \frac{i}{12}q\right)u^5\right]
= \frac{45i}{14}\sqrt{\frac{\pi}{2}}\frac{\mu}{r_{\rm
o}^3}u^5\exp\left(iq/12\right)\;,
\label{eq:ZHkerr41}
\\
& &Z^{\rm H}_{42} = \frac{15}{14}\sqrt{\pi}\frac{\mu}{r_{\rm o}^3}
\left[\left(1 + \frac{i}{6}q\right)u^4\right]
= \frac{15}{14}\sqrt{\pi}\frac{\mu}{r_{\rm o}^3}u^4
\exp\left(iq/6\right)\;,
\label{eq:ZHkerr42}
\\
& &Z^{\rm H}_{43} = -\frac{15i}{2} \sqrt{\frac{\pi}{14}}
\frac{\mu}{r_{\rm o}^3}\left[\left(1 + \frac{i}{4}q\right)u^5\right]
= -\frac{15i}{2}\sqrt{\frac{\pi}{14}}\frac{\mu}{r_{\rm o}^3} u^5
\exp\left(iq/4\right)\;,
\label{eq:ZHkerr43}
\\
& &Z^{\rm H}_{44} = -\frac{15}{4}\sqrt{\frac{\pi}{7}}\frac{\mu}{r_{\rm
o}^3} \left[\left(1 + \frac{i}{3}q\right)u^4\right] =
-\frac{15}{4}\sqrt{\frac{\pi}{7}}\frac{\mu}{r_{\rm
o}^3}u^4\exp\left(iq/3\right)\;.
\label{eq:ZHkerr44}
\end{eqnarray}
Equations (\ref{eq:ZHkerr20}) -- (\ref{eq:ZHkerr44}) reduce to
Eqs.\ (\ref{eq:ZHschw20}) -- (\ref{eq:ZHschw44}) when $q \to 0$.
Modes for $m < 0$ can be obtained using the rule $Z^{\rm H}_{l-m} =
(-1)^l\bar Z^{\rm H}_{lm}$, with overbar denoting complex conjugate.
Lastly, we need the angular function $\bar\eth\bar\eth S^+_{lm}$ to
leading order in $q$. Using Eqs.\ (\ref{eq:Lminus_s_costheta}),
(\ref{eq:Lminus_s_sintheta}), (\ref{eq:barethbareth_small_a}), and the
condition $q \ll 1$, we have
\begin{equation}
\bar\eth\bar\eth S^+_{lm} = \frac{1}{8M^2}L_-^sL_-^s\left(1 +
iq\cos\theta\right)S^+_{lm} = \frac{1}{8M^2} \left[(1 +
iq\cos\theta)L_-^sL_-^sS^+_{lm} - 2iq\sin\theta L_-^s
S^+_{lm}\right]\;.
\label{eq:barethbarethexpand}
\end{equation}
Following the analysis in Appendix {\ref{app:spheroidal_lin}}, the
spheroidal harmonic to this order is
\begin{equation}
S^+_{lm} = {_2}Y_{lm} + qM\omega_m\left[c^{l+1}_{lm} {_2}Y_{(l+1)m}
+ c^{l-1}_{lm} {_2}Y_{(l-1)m}\right]\;,
\label{eq:spheroidalexpand}
\end{equation}
where
\begin{eqnarray}
c^{l+1}_{lm} &=& -\frac{2}{(l+1)^2}\sqrt{\frac{(l + 3)(l - 1)(l + m + 1)(l
- m + 1)}{(2l + 3)(2l + 1)}}\;,
\\
c^{l-1}_{lm} &=& \frac{2}{l^2}\sqrt{\frac{(l + 2)(l - 2)(l + m)(l -
m)}{(2l + 1)(2l - 1)}}\;.
\end{eqnarray}
Using Eq.\ (\ref{eq:sphericalharmoniclower}) with
Eqs.\ (\ref{eq:barethbarethexpand}) and (\ref{eq:spheroidalexpand})
and expanding to leading order in $q$, we find
\begin{eqnarray}
\bar\eth\bar\eth S^+_{lm} &=& \frac{1}{8M^2}\biggl[(1 +
iq\cos\theta)\sqrt{(l+2)(l+1)l(l-1)}\,{_0}Y_{lm} -
2iq\sin\theta\sqrt{(l + 2)(l - 1)}\,{_1}Y_{lm}
\nonumber\\
& &\qquad + qM\omega_m\left(c^{l+1}_{lm}\sqrt{(l + 3)(l + 2)(l +
1)l}\,{_0}Y_{(l+1)m} + c^{l-1}_{lm}\sqrt{(l + 1)l(l - 1)(l - 2)}\,
{_0}Y_{(l-1)m}\right)\biggr]\;.
\label{eq:barethbarethSexpand}
\end{eqnarray}
As in Sec.\ {\ref{sec:schw_analytic}}, it is convenient to combine
modes in pairs. For $l = 2$, we find
\begin{eqnarray}
R^{(1)}_{{\rm H},20} &=& -\frac{\mu}{r_{\rm o}^3} \left(3\cos^2\theta
- 1\right) \left(1 + \frac{7}{2}u^2 - 4qu^3 + \frac{561}{56}u^4 -
18qu^5\right)\;,
\label{eq:R20kerr}\\
R^{(1)}_{{\rm H},2-1} + R^{(1)}_{{\rm H},21} &=& \frac{4\mu}{r_{\rm
o}^3} \left(5\cos^2\theta - 1\right)\sin\theta\, q\left(u +
\frac{8}{3}u^3 + \frac{152}{21}u^5\right) \cos\left(\psi - \Omega_{\rm
orb}v - \frac{3}{2}u^3 + 6M\Omega_{\rm H}\right)\;,
\label{eq:R21kerr}\\
R^{(1)}_{{\rm H},2-2} + R^{(1)}_{{\rm H},22} &=&
\frac{3\mu}{r_{\rm o}^3}\sin^2\theta \left[1 + \frac{3}{2}u^2 -
\left(10 + \frac{4\pi^2}{3}\right)qu^3 + \frac{1403}{168}u^4 -
\left(\frac{131}{9} + 2\pi^2\right)qu^5\right]
\times
\nonumber\\
& &\quad
\cos\left[2\left(\psi - \Omega_{\rm orb} v - \frac{8}{3}u^3 +
\frac{32}{5}u^5 + \frac{14}{3}M\Omega_{\rm H}\right)\right]\;.
\label{eq:R22kerr}
\end{eqnarray}
For $l = 3$,
\begin{eqnarray}
R^{(1)}_{{\rm H},30} &=& -\frac{\mu}{r_{\rm o}^3}\left(1 -
12\cos^2\theta + 15\cos^4\theta\right)qu^3(1 + 4u^2)\;,
\label{eq:R30kerr}\\
R^{(1)}_{{\rm H},3-1} + R^{(1)}_{{\rm H},31} &=&
\frac{3}{2}\frac{\mu}{r_{\rm o}^3} \sin\theta\left(1 -
5\cos^2\theta\right) u^2\left[1 + \frac{13}{3}u^2
-\left(\frac{113}{12} + \frac{\pi^2}{2}\right)qu^3\right]
\cos\left[\psi - \Omega_{\rm orb} v - \frac{14}{3}u^3 +
\frac{20}{3}M\Omega_{\rm H}\right]\;,
\nonumber\\
\label{eq:R31kerr}\\
R^{(1)}_{{\rm H},3-2} + R^{(1)}_{{\rm H},32} &=&
\frac{5}{3}\frac{\mu}{r_{\rm o}^3}q(u^3 + 4u^5)\left\{9\cos^2\theta
\cos\left[2\left(\psi - \Omega_{\rm orb} v - \frac{56}{10}u^3 +
\frac{22}{3}M\Omega_{\rm H}\right)\right]
\right.\nonumber\\
& &\qquad\qquad\qquad\qquad\left.
- \cos\left[2\left(\psi - \Omega_{\rm orb} v - \frac{158}{30}u^3 +
\frac{22}{3}M\Omega_{\rm H}\right)\right]\right\}\;,
\label{eq:R32kerr}\\
R^{(1)}_{{\rm H},3-3} + R^{(1)}_{{\rm H},33} &=&
\frac{5}{2}\frac{\mu}{r_{\rm o}^3}\sin^3\theta u^2\left[1 + 3u^2 -
\left(\frac{49}{2} + 3\pi^2\right)qu^3\right]
\cos\left[3\left(\psi - \Omega_{\rm orb} v - \frac{14}{3}u^3 +
\frac{20}{3}M\Omega_{\rm H}\right)\right]\;.
\nonumber\\
\label{eq:R33kerr}
\end{eqnarray}
And for $l = 4$,
\begin{eqnarray}
R^{(1)}_{{\rm H},40} &=& \frac{9}{56}\frac{\mu}{r_{\rm o}^3}
\left(3 - 30\cos^2\theta + 35\cos^4\theta\right)u^4\;,
\label{eq:R40kerr}\\
R^{(1)}_{{\rm H},4-1} + R^{(1)}_{{\rm H},41} &=&
-\frac{9}{28}\frac{\mu}{r_{\rm
o}^3}\sin\theta\,qu^5\left[98\cos^4\theta\cos\left(\psi -
\Omega_{\rm orb}v - \frac{169}{30}u^3 + \frac{25}{3}M\Omega_{\rm
H}\right)
\right.
\nonumber\\
& &\quad\left.
- 57\cos^2\theta\cos\left(\psi - \Omega_{\rm orb}v -
\frac{1568}{285}u^3 + \frac{25}{3}M\Omega_{\rm H}\right)
+ 3\cos\left(\psi - \Omega_{\rm orb}v - \frac{77}{15}u^3 +
\frac{25}{3}M\Omega_{\rm H}\right)\right]\;,
\nonumber\\
\label{eq:R41kerr}\\
R^{(1)}_{{\rm H},4-2} + R^{(1)}_{{\rm H},42} &=&
\frac{15}{14}\frac{\mu}{r_{\rm o}^3}\sin^2\theta \left(1 -
7\cos^2\theta\right)u^4 \cos\left[2\left(\psi - \Omega_{\rm orb} v -
\frac{181}{30}u^3 + \frac{119}{15}M\Omega_{\rm H}\right)\right]\;,
\label{eq:R42kerr}\\
R^{(1)}_{{\rm H},4-3} + R^{(1)}_{{\rm H},43} &=&
\frac{3}{4}\frac{\mu}{r_{\rm o}^3}\sin^3\theta\,qu^5
\left\{14\cos^2\theta\cos\left[3\left(\psi - \Omega_{\rm orb}v -
\frac{169}{30}u^3 + \frac{25}{3}M\Omega_{\rm H}\right)\right]
\right.
\nonumber\\
& &\quad\qquad\qquad\qquad\left.
- \cos\left[3\left(\psi - \Omega_{\rm orb}v - \frac{77}{15}u^3 +
\frac{25}{3}M\Omega_{\rm H}\right)\right]\right\}\;,
\label{eq:R43kerr}\\
R^{(1)}_{{\rm H},4-4} + R^{(1)}_{{\rm H},44} &=&
\frac{15}{8}\frac{\mu}{r_{\rm o}^3}\sin^4\theta u^4
\cos\left[4\left(\psi - \Omega_{\rm orb} v - \frac{181}{30}u^3 +
\frac{119}{15}M\Omega_{\rm H}\right) \right]\;.
\label{eq:R44kerr}
\end{eqnarray}
\end{widetext}
In writing these formulas, we have used the fact that $\Omega_{\rm H}
= q/4M$ in the $q \ll 1$ limit to rewrite certain terms in the phases
using $\Omega_{\rm H}$ rather than $q$. For example, in
Eq.\ (\ref{eq:R22kerr}) our calculation yields a term $7q/6$ in the
argument of the cosine, which we rewrite $14M\Omega_{\rm H}/3$. We
have found that this improves the match of Eqs.\ (\ref{eq:R20kerr}) --
(\ref{eq:R44kerr}) with the numerical results we discuss in
Sec.\ {\ref{sec:kerr_numerical}}.
\subsubsection{Phase of the tidal bulge: Null map}
We begin by examining the bulge-orbit offset using the null map,
Eq.\ (\ref{eq:bulge_vs_orbit_null}). The horizon's geometry is
dominated by contributions for which $l + m$ is even; modes with $l +
m$ odd are suppressed by $qu$ relative to these dominant modes (thus
vanishing in the Schwarzschild limit). The dominant modes peak at
$\psi^{\rm bulge}_{lm} = \Omega_{\rm orb}v + \delta\psi_{lm}(u) +
\delta\psi_{lm}(q)$, where $\delta\psi_{lm}(u)$ and
$\delta\psi_{lm}(q)$ can be read out of Eqs.\ (\ref{eq:R20kerr}) --
(\ref{eq:R44kerr}). The orbit mapped onto the horizon in the null map
is given by Eq.\ (\ref{eq:psi_o_nullmap}). Following discussion in
Sec.\ {\ref{sec:nullmap}}, the offset phases in the null map for the
dominant modes, to $O(u^5)$ and $O(q)$, are
\begin{eqnarray}
\delta\psi^{\rm OB-NM}_{22} &=& \frac{8}{3}\left(u^3 - M\Omega_{\rm H}\right)
- \frac{32}{5}u^5 - \frac{4M^2\Omega_{\rm H}}{r_{\rm o}}
\nonumber\\
& & - \Delta\psi(r_{\rm o})\;,
\label{eq:psi_obnm_2}\\
\delta\psi^{\rm OB-NM}_{31} &=& \delta\psi^{\rm OB-NM}_{33}
\nonumber\\
&=& \frac{14}{3}\left(u^3 - M\Omega_{\rm H}\right) -
\frac{4M^2\Omega_{\rm H}}{r_{\rm o}} - \Delta\psi(r_{\rm o})\;,
\nonumber\\
\label{eq:psi_obnm_3}\\
\delta\psi^{\rm OB-NM}_{42} &=& \delta\psi^{\rm OB-NM}_{44}
\nonumber\\
&=& \frac{181}{30}u^3 - \frac{89}{15}M\Omega_{\rm H}
- \frac{4M^2\Omega_{\rm H}}{r_{\rm o}} - \Delta\psi(r_{\rm o})\;.
\nonumber\\
\label{eq:psi_obnm_4}
\end{eqnarray}
We again see agreement with Fang and Lovelace for $l = m = 2$, who
correct a sign error in Hartle's {\cite{hartle74}} treatment of the
bulge phase; compare Eq.\ (61) and footnote 6 of Ref.\ {\cite{fl05}}
and associated discussion. In contrast to the Schwarzschild case, the
Kerr offset phases can be positive or negative, depending on the
values of $r_{\rm o}$ and $q$. To highlight this further, let us
examine Eq.\ (\ref{eq:psi_obnm_2}) for very large $r_{\rm o}$: we drop
the term in $u^5$, and expand $\Delta\psi(r_{\rm o})$. The result is
\begin{equation}
\delta\psi^{\rm OB-NM}_{22} \simeq \frac{8}{3}\left(u^3 - M\Omega_{\rm
H}\right) + \sqrt{\frac{M}{r_{\rm o}}}\;.
\label{eq:psi_obnm_2_v2}
\end{equation}
As $r_{\rm o} \to \infty$, we see that this bulge lags the orbit by
$\delta^{\rm OB-NM}_{22} = -8M\Omega_{\rm H}/3$, which reproduces
Hartle's finding for a stationary moon orbiting a slowly rotating Kerr
black hole [Eq.\ (4.34) of Ref.\ {\cite{hartle74}}, correcting the
sign error discussed in footnote 6 of Ref.\ {\cite{fl05}}]. We
discuss this point further in Sec.\ {\ref{sec:leadlag}}.
\subsubsection{Phase of the tidal bulge: Instantaneous map}
Consider next the instantaneous-in-$v$ map discussed in
Sec.\ {\ref{sec:instmap}}. The position of the orbit on the horizon
in this mapping is given by Eq.\ (\ref{eq:bulge_vs_orbit_inst}). To
$O(u^5)$ and $O(q)$, the offset phase for the dominant modes in this
map is
\begin{eqnarray}
\delta\psi^{\rm OB-IM}_{22} &=& \frac{8}{3}u^3 -
\frac{14}{3}M\Omega_{\rm H} - \frac{32}{5}u^5 - \Delta\psi(r_{\rm
o})\;,
\nonumber\\
\label{eq:psi_obim_2}\\
\delta\psi^{\rm OB-IM}_{31} &=& \delta\psi^{\rm OB-IM}_{33}
\nonumber\\
&=& \frac{14}{3}u^3 - \frac{20}{3}M\Omega_{\rm H} - \Delta\psi(r_{\rm
o})\;,
\label{eq:psi_obim_3}\\
\delta\psi^{\rm OB-IM}_{42} &=& \delta\psi^{\rm OB-IM}_{44}
\nonumber\\
&=& \frac{181}{30}u^3 - \frac{119}{15}M\Omega_{\rm H} -
\Delta\psi(r_{\rm o})\;.
\label{eq:psi_obim_4}
\end{eqnarray}
As in the null map, these phases can be positive or negative,
depending on the values of $r_{\rm o}$ and $q$. As we'll see when we
examine numerical results for the horizon geometry,
Eq.\ (\ref{eq:psi_obim_2}) does a good job describing the angle of the
peak horizon bulge for small values of $q$.
\subsubsection{Phase of the tidal bulge: Tidal field versus tidal
response}
Finally, let us examine the relative phase of tidal field modes
$\psi^{\rm HH}_{0,lm}$ and the horizon's response $R^{(1)}_{{\rm
H},lm}$. For $q \ll 1$, we have $\kappa^{-1} = 4M + O(q^2)$.
Expanding in the weak-field limit, Eq.\ (\ref{eq:deltapsiTB}) becomes
\begin{equation}
\delta\psi^{\rm TB}_{lm} = 4m(u^3 - M\Omega_{\rm H}) + {\cal
S}_{lm}(\pi/2)\;.
\label{eq:deltaTB_weak}
\end{equation}
For the modes with $l + m$ even which dominate the horizon's response,
it is not difficult to compute ${\cal S}_{lm}(\pi/2)$ to leading order
in $q$. Equation (\ref{eq:barethbarethSexpand}) and the definition
(\ref{eq:spheroidratio}) yield
\begin{equation}
{\cal S}_{lm}(\pi/2) = \frac{2q}{\sqrt{l(l+1)}}
\frac{{_1}Y_{lm}(\pi/2)}{{_0}Y_{lm}(\pi/2)} + O(q^2)\;.
\label{eq:Sphase_smallq}
\end{equation}
We also know [cf.\ Eq.\ (A8) of Ref.\ {\cite{h2000}}] that
\begin{equation}
{_1}Y_{lm}(\theta) = -\frac{1}{\sqrt{l(l+1)}}\left(\partial_\theta -
m\csc\theta\right){_0}Y_{lm}(\theta)\;.
\end{equation}
For $l + m$ even, $\partial_\theta\,{_0}Y_{lm} = 0$ at $\theta =
\pi/2$. Plugging the resulting expression for ${_1}Y_{lm}(\pi/2)$
into Eq.\ (\ref{eq:Sphase_smallq}), we find
\begin{equation}
{\cal S}_{lm}(\pi/2) = \frac{2mq}{l(l+1)} = \frac{8mM\Omega_{\rm
H}}{l(l+1)}\;,
\end{equation}
where in the last step we again used $q = 4M^2\Omega_{\rm H}$,
accurate for $q \ll 1$. With this, Eq.\ (\ref{eq:deltaTB_weak})
becomes
\begin{eqnarray}
\delta\psi^{\rm TB}_{lm} &=& 4m\left[u^3 - M\Omega_{\rm H}\left(1 -
\frac{2}{l(l+1)}\right)\right]
\nonumber\\
&=& 4m\left[u^3 - M\Omega_{\rm H}\frac{(l+2)(l-1)}{l(l+1)}\right]\;.
\label{eq:deltaTB}
\end{eqnarray}
Just as with the offset phases of the bulge and the orbit for Kerr,
this tidal bulge phase can be either positive or negative depending on
$r_{\rm o}$ and $q$, and so the horizon's response can lead or lag the
applied tidal field.
\subsection{Fast motion: Numerical results}
\label{sec:kerr_numerical}
Figures {\ref{fig:kerr_Rvspsi}}, {\ref{fig:a0.1_a0.4_shapes}}, and
{\ref{fig:a0.7_a0.866_shapes}} present summary data for our numerical
calculations of tidally distorted Kerr black holes. Just as in
Sec.\ {\ref{sec:schw_numerical}}, we compute $R^{(1)}_{\rm H}$ by
solving for $Z^{\rm H}_{lm}$ as described in Sec.\ {\ref{sec:ZH}}, and
then apply Eq.\ (\ref{eq:gauss_kerr}). As in the Schwarzschild case,
we find rapid convergence with mode index $l$. All the data we show
are for the equatorial plane, $\theta = \pi/2$, and are rescaled by
$(r_{\rm o}^3/\mu)$. We typically include all modes up to $l = 15$
(increasing this to $20$ and $25$ in a few very strong field cases).
Contributions beyond this are typically at the level of $10^{-9}$ or
smaller, which is accurate enough for this exploratory analysis.
Figure {\ref{fig:kerr_Rvspsi}} is the Kerr analog of
Fig.\ {\ref{fig:schw_Rvsphi}}, comparing numerical results for $R^{\rm
H}_{lm}$ with analytic predictions for selected black hole spins,
mode numbers, and orbital radii. For all modes we show here, we see
outstanding agreement in both phase and amplitude for $q = 0.1$ and
$r_{\rm o} = 50M$; in some cases, the numerical data lies almost
directly on top of the analytic prediction. The amplitude agreement
is not quite as good as we increase the spin to $q = 0.2$ and move to
smaller radius ($r_{\rm o} = 10M$), though the phase agreement remains
quite good for all modes.
\begin{figure*}[ht]
\includegraphics[width = 0.48\textwidth]{a01_comp.pdf}
\includegraphics[width = 0.48\textwidth]{a02_comp.pdf}
\caption{Comparison of selected modes for the numerically computed
scalar curvature perturbation $R^{(1)}_{{\rm H},lm}$ with the
analytic expansion given in Eqs.\ (\ref{eq:R20kerr}) --
(\ref{eq:R44kerr}). The four panels on the left are for orbits of a
black hole with $a = 0.1M$ at $r_{\rm o} = 50M$; those on the right
are for orbits of a black hole with $a = 0.2M$ at $r_{\rm o} = 10M$.
The mode shown is indicated by $(l,m)$ in the upper right corner of
each panel [we actually show the contributions from $(l,m)$ and
$(l,-m)$]. In all cases, we plot $(r_{\rm o}^3/\mu)R^{(1)}_{\rm
H}$, scaling out the leading dependence on orbital radius and the
orbiting body's mass. Curves in light gray (green in color) are the
analytic results, those in dark gray (red in color) are our
numerical data. Agreement for the large radius, low spin cases is
extremely good, especially for small $l$ where the numerical data
lies practically on top of the analytic predictions. As we increase
$q$ and decrease $r_{\rm o}$, the amplitude agreement becomes less
good, though the analytic formulas still are within several to
several tens of percent of the numerical data. The phase agreement
is outstanding in all of these cases. }
\label{fig:kerr_Rvspsi}
\end{figure*}
Figures {\ref{fig:a0.1_a0.4_shapes}} and
{\ref{fig:a0.7_a0.866_shapes}} show equatorial slices of the embedding
of distorted Kerr black holes for a range of orbits and black hole
spins. These embeddings are similar to those we used for distorted
Schwarzschild black holes (as described in Sec.\ {\ref{sec:l=2embed}}),
with a few important adjustments. The embedding surface we use has the
form
\begin{equation}
r_{\rm E} = r_{\rm E}^{0}(\theta) + r_+\sum_{\ell m}\varepsilon_{\ell
m}(\theta,\psi)\;.
\end{equation}
Both the undistorted radius $r_{\rm E}^{(0)}(\theta)$ and the tidal
distortion $\varepsilon_{\ell m}(\theta,\psi)$ are described in
Appendix {\ref{app:embed}}; see also Ref.\ {\cite{smarr}}. The
background embedding reduces to a sphere of radius $2M$ when $a = 0$,
but is more complicated in general. The embedding's tidal distortion
is linearly related to the curvature $R^{(1)}_{{\rm H},lm}$, but in a
way that is more complicated than the Schwarzschild relation
(\ref{eq:schw_spheroid}). In particular, mode mixing becomes
important: Different angular basis functions are needed to describe
the curvature $R^{(1)}_{{\rm H},lm}$ and the embedding distortion
$\varepsilon_{\ell m}$ when $a \ne 0$. Hence, the $\ell = 2$
contribution to the horizon's shape has contributions from all $l$
curvature modes, not just $l = 2$. See Appendix {\ref{app:embed}} for
detailed discussion.
In this paper, we only generate embeddings for $a/M \le \sqrt{3}/2$.
For spins greater than this, the horizon cannot be embedded in a
global 3-dimensional Euclidean space. A ``belt'' from $\pi -
\theta_{\rm E} \le \theta \le \theta_{\rm E}$ can always be embedded
in 3-dimensional Euclidean space, but the ``polar cones'' $0 \le
\theta < \theta_{\rm E}$ and $\pi - \theta_{\rm E} < \theta \le \pi$
must be embedded in a Lorentzian geometry (where $\theta_{\rm E}$ is
related to the root of a function used in the embedding; see Appendix
{\ref{app:embed}} for details). Alternatively, one can embed the
entire horizon in a different space, as discussed in
Refs.\ {\cite{frolov,gibbons}}. We defer detailed discussion of
embeddings that can handle the case $a/M > \sqrt{3}/2$ to a later
paper.
As with the Schwarzschild embeddings shown in
Fig.\ {\ref{fig:a0.0_shapes}}, the Kerr embeddings we show are all
plotted in a frame that corotates with the orbit at a moment $v = {\rm
constant}$. The $x$-axis is at $\psi = 0$, and the orbiting body
sits at $\psi = \Delta\psi(r_{\rm o}) = \bar r_{\rm o} - \Omega_{\rm
orb}r^*_{\rm o}$. As in Fig.\ {\ref{fig:a0.0_shapes}}, the green
dashed line labels the horizon's peak bulge, and the black dotted line
shows the position of the orbiting body.
For small $q$, we find that the numerically computed bulge offset
agrees quite well with the $l = 2$ analytic expansion in the
instantaneous map, Eq.\ (\ref{eq:psi_obim_2}). For $q = 0.1$, our
numerical results are
\begin{eqnarray}
\delta\psi^{\rm num} &=& 3.01^\circ\quad r_{\rm o} = 50M\;,
\nonumber\\
&=& 10.8^\circ\quad r_{\rm o} = 20M\;,
\nonumber\\
&=& 21.6^\circ\quad r_{\rm o} = 10M\;,
\nonumber\\
&=& 33.7^\circ\quad r_{\rm o} = 5.669M\;.
\label{eq:q=0.1_num}
\end{eqnarray}
These are within a few percent of predictions based on the weak-field,
slow spin expansion:
\begin{eqnarray}
\delta\psi^{\rm OB-IM}_{22}
&=& 2.95^\circ\quad r_{\rm o} = 50M\;,
\nonumber\\
&=& 10.7^\circ\quad r_{\rm o} = 20M\;,
\nonumber\\
&=& 20.6^\circ\quad r_{\rm o} = 10M\;,
\nonumber\\
&=& 30.0^\circ\quad r_{\rm o} = 5.669M\;.
\end{eqnarray}
As we move to larger spin, the agreement rapidly becomes worse. Terms
which we neglect in our expansion become important, and the mode
mixing described above becomes very important. For $q = 0.4$, the
agreement degrades to a few tens of percent in most cases:
\begin{eqnarray}
\delta\psi^{\rm num} &=& -13.5^\circ\quad r_{\rm o} = 50M\;,
\nonumber\\
&=& -6.17^\circ\quad r_{\rm o} = 20M\;,
\nonumber\\
&=& 3.55^\circ\quad r_{\rm o} = 10M\;,
\nonumber\\
&=& 21.4^\circ\quad r_{\rm o} = 4.614M\;;
\label{eq:q=0.4_num}
\end{eqnarray}
and
\begin{eqnarray}
\delta\psi^{\rm OB-IM}_{22}
&=& -17.9^\circ\quad r_{\rm o} = 50M\;,
\nonumber\\
&=& -9.76^\circ\quad r_{\rm o} = 20M\;,
\nonumber\\
&=& 0.82^\circ\quad r_{\rm o} = 10M\;,
\nonumber\\
&=& 13.6^\circ\quad r_{\rm o} = 4.614M\;.
\end{eqnarray}
The agreement gets significantly worse as $q$ is increased further.
Presumably, $q \sim 0.3$ is about as far as the leading order
expansion in $q$ can reasonably be taken.
\begin{figure*}[ht]
\includegraphics[width = 0.48\textwidth]{a01_shapes.pdf}
\includegraphics[width = 0.48\textwidth]{a04_shapes.pdf}
\caption{Equatorial section of the embedding of distorted Kerr black
hole event horizons, $a = 0.1M$ and $a = 0.4M$. Each panel
represents the distortion for a different radius of the orbiting
body, varying from $r_{\rm o} = 50M$ to the innermost stable
circular orbit ($r_{\rm o} = 5.669M$ for $a = 0.1M$, $r_{\rm o} =
4.614M$ for $a = 0.4M$). As in Fig.\ {\ref{fig:a0.0_shapes}}, the
green dashed line shows the angle at which the tidal distortion is
largest, and the black dotted line shows the position of the orbit.
As in Fig.\ {\ref{fig:a0.0_shapes}}, we have rescaled by a factor
$\propto r_{\rm o}^3/\mu$ to account for the leading dependence of
the tide on mass and orbital separation. In contrast to the
Schwarzschild results, the bulge does not lead the orbit in all
cases here. The amount by which the bulge leads the orbit grows as
the orbit moves to small orbital radius (in some cases, changing
from a lag to a lead as part of this trend).}
\label{fig:a0.1_a0.4_shapes}
\end{figure*}
\begin{figure*}[ht]
\includegraphics[width = 0.48\textwidth]{a07_shapes.pdf}
\includegraphics[width = 0.48\textwidth]{a0866_shapes.pdf}
\caption{Equatorial section of the embedding of distorted Kerr black
hole event horizons, $a = 0.7M$ and $a = 0.866M$. Each panel
represents the distortion for a different radius of the orbiting
body, varying from $r_{\rm o} = 50M$ to the innermost stable
circular orbit ($r_{\rm o} = 3.393M$ for $a = 0.7M$, $r_{\rm o} =
2.537M$ for $a = 0.866M$), with the green dashed and black dotted
lines labeling the locations of maximal distortion and position of
the orbit, respectively, and with the distortion rescaled by a
factor $\propto r_{\rm o}^3/\mu$. The bulge lags the orbit in most
cases we show here, with the lag angle getting smaller and
converting to a small lead as the orbit moves to smaller and smaller
orbital radius.}
\label{fig:a0.7_a0.866_shapes}
\end{figure*}
To conclude this section, we show two examples of embeddings for the
entire horizon surface, rather than just the equatorial slice. The
left-hand panel of Fig.\ {\ref{fig:surfaces}} is an example of a
relatively mild tidal distortion. The black hole has spin $a = 0.3M$,
and the orbiting body is at $r_{\rm o} = 20M$. The distortion is
strongly dominated by the $\ell = 2$ contribution, and we see a fairly
simple prolate ellipsoid whose bulge lags the orbit. The right-hand
panel shows a much more extreme example. The black hole here has $a =
0.866M$, and the orbiting body is at $r_{\rm o} = 1.75M$. The
horizon's shape has strong contributions from many multipoles, and so
is bent in a rather more complicated way than in the mild case. The
connection between the orbit and the horizon geometry is quite unusual
here. Note that this extreme case corresponds to an {\it unstable}
circular orbit, and so one might question whether this figure is
physically relevant. We include it because we expect similar horizon
distortions for very strong field orbits of black holes with $a/M >
\sqrt{3}/2$, and that such a horizon geometry will be produced
transiently from the closest approach of eccentric orbits around black
holes with $a/M \lesssim \sqrt{3}/2$. Both of these cases will be
investigated more thoroughly in later papers.
\begin{figure*}[ht]
\includegraphics[width = 0.497\textwidth]{q03_r20.png}
\includegraphics[width = 0.463\textwidth]{q0866_r175_l12.png}
\caption{Two example embeddings of the tidally distorted horizon's
surface. Both panels show the 3-dimensional Euclidean embedding
surface, $r_{\rm E}(\theta,\psi)$; the shading (or color scale)
indicates the horizon's distortion relative to an isolated Kerr
black hole. The hole is stretched (i.e., $r_{\rm E}$ increased by
the tides relative to an isolated hole; red in color) at the end
near to and opposite from the orbiting body. It is squeezed
($r_{\rm E}$ decreased by tides; blue in color) in a band between
these two ends. As in other figures illustrating the embedded
distorted horizon, we have rescaled the distortion by a factor
$\propto r_{\rm o}^3/\mu$. On the left, we show a relatively gentle
deformation around a moderately spinning black hole: $a = 0.3M$,
$r_{\rm o} = 20M$. The distortion here is dominated by a
quadrupolar deformation of the horizon (lagging the orbiting body,
whose angular position is indicated by the small blue ball). On the
right, we show a rather extreme case: $a = 0.866M$, $r_{\rm o} =
1.75M$. The deformation here is much more complicated, as many
multipoles beyond $l = 2$ contribute to the shape of the horizon.}
\label{fig:surfaces}
\end{figure*}
\section{Lead or lag?}
\label{sec:leadlag}
We showed in Sec.\ {\ref{sec:downhoriz}} that the orbital energy
evolves due to horizon coupling according to $dE^{\rm H}/dt \propto
(\Omega_{\rm orb} - \Omega_{\rm H})$. As discussed in the
Introduction, it is simple to build an intuitive picture of this in
Newtonian physics. For a Newtonian tide acting on a fluid body, when
$\Omega_{\rm H} > \Omega_{\rm orb}$ tidal forces raise a bulge on the
body which leads the orbit's position. This bulge exerts a torque
which transfers energy from the body's spin to the orbit. When
$\Omega_{\rm H} < \Omega_{\rm orb}$, the bulge lags the orbit, and the
torque transfers energy from the orbit to the body's spin. When
$\Omega_{\rm H} = \Omega_{\rm orb}$, $dE^{\rm H}/dt = 0$. The
Newtonian fluid expectation is thus that there should be no offset
between the bulge and the orbit. The tidal bulge should point
directly at the orbiting body, locking the body's tide to the orbit.
Consider now a fully relativistic calculation of tides acting on a
black hole. When $\Omega_{\rm orb} \gg \Omega_{\rm H}$ (e.g., the
Schwarzschild limit) and $\Omega_{\rm H} \gg \Omega_{\rm orb}$ (large
radius orbits of Kerr black holes), the Newtonian fluid intuition is
consistent with our results, modulo the switch of ``lead'' and ``lag''
thanks to the teleological nature of the event horizon. However, it
is not so clear if this intuition holds up when $\Omega_{\rm orb}$ and
$\Omega_{\rm H}$ are comparable in magnitude.
Let us investigate this systematically. Begin with the weak-field $l
= m = 2$ offset angles in the null and instantaneous maps,
Eqs.\ (\ref{eq:psi_obnm_2}) and (\ref{eq:psi_obim_2}). Dropping terms
of $O(u^5)$ and noting that $u^3 = M\Omega_{\rm orb} + O(qu^6)$, we
solve for the conditions under which $\delta\psi^{\rm OB-NM}_{22}$ and
$\delta\psi^{\rm OB-IM}_{22}$ are zero. In the null map, we find
\begin{equation}
\Omega_{\rm orb} = \Omega_{\rm H} + \frac{3M\Omega_{\rm H}}{2r_{\rm o}} +
\frac{3\Delta\psi_{\rm o}}{8M}\;.
\label{eq:NMoffsetzero}
\end{equation}
The bulge leads the orbit when the equals in the above equation is
replaced by greater than, and lags when replaced by less than. In the
instantaneous map,
\begin{equation}
\Omega_{\rm orb} = \frac{7}{4}\Omega_{\rm H} + \frac{3\Delta\psi_{\rm
o}}{8M}\;,
\label{eq:IMoffsetzero}
\end{equation}
with the same replacements indicating lead or lag.
{\it Neither of these conditions are consistent with $\Omega_{\rm orb}
= \Omega_{\rm H}$ indicating zero bulge-orbit offset.} In both the
null and instantaneous maps, we find $\Omega_{\rm orb} \ll \Omega_{\rm
H}$ when the bulge angle is zero. For example, for $a = 0.3M$
(roughly the largest $a$ for which the small spin expansion is
trustworthy), Eq.\ (\ref{eq:NMoffsetzero}) has a root at $r_{\rm o} =
35.9M$, for which $M\Omega_{\rm orb} = 0.00464$, $M\Omega_{\rm H} =
0.0768$. (A second root exists at $r_{\rm o} = 2.15M$, but this is
inside the photon orbit.) Using the instantaneous map changes the
numbers, but not the punchline: for $a = 0.3M$, the root moves to
$r_{\rm o} = 16.7M$, with $M\Omega_{\rm orb} = 0.0146$. Changing the
spin changes the numbers, but leaves the message the same: zero offset
in these maps does not correspond to $\Omega_{\rm orb} = \Omega_{\rm
H}$.
Equations (\ref{eq:NMoffsetzero}) and (\ref{eq:IMoffsetzero}) were
derived using a small spin expansion. Before drawing too firm a
conclusion from this, let us examine the situation using numerical
data good for large spin. In Fig.\ {\ref{fig:corotshapes}}, we
examine a sequence of ``corotating'' orbits --- orbits for which
$\Omega_{\rm H} = \Omega_{\rm orb}$, so that $dE^{\rm H}/dt = 0$. For
very small spins, the orbit leads the bulge. As the black hole's spin
increases, the lead becomes a lag. This lead gets smaller as the spin
gets larger. Since the lag becomes a lead as the spin is changed from
$a = 0.1M$ to $a = 0.2M$, there must be a spin value between $a =
0.1M$ and $a = 0.2M$ for which the lead angle is zero for the
corotating orbit. Our data also suggest that the lead angle may
approach zero as the spin gets very large. But this suggests that the
horizon locks to the orbit for at most only two spin values, in this
map --- a set of measure zero. We do not find any systematic
connection between the geometry and the horizon for these orbits.
\begin{figure*}[ht]
\includegraphics[width = 0.48\textwidth]{corot_shapes1.pdf}
\includegraphics[width = 0.48\textwidth]{corot_shapes2.pdf}
\caption{Embedding of distorted Kerr black hole event horizons for a
corotating orbit --- i.e., an orbit for which $\Omega_{\rm orb} =
\Omega_{\rm H}$. As in Figs.\ {\ref{fig:a0.0_shapes}},
{\ref{fig:a0.1_a0.4_shapes}}, and
{\ref{fig:a0.7_a0.866_shapes}}, the green dashed line points
along the direction of greatest horizon distortion, and the
black dotted line points to the orbiting body; the distortions
are all scaled by a factor $\propto r_{\rm o}^3/\mu$. At very
small spins (for which the corotating orbital radius is very
large), the bulge lags the orbit slightly, but the bulge leads
for all other spins.}
\label{fig:corotshapes}
\end{figure*}
Before concluding, let us examine the relative phase of the tidal
field and the horizon's curvature, Eq.\ (\ref{eq:deltaTB}). Setting
$\delta\psi^{\rm TB}_{lm} = 0$ yields
\begin{equation}
\Omega_{\rm orb} = \Omega_{\rm H}\,\frac{(l+2)(l-1)}{l(l+1)}\;.
\label{eq:TBoffsetzero}
\end{equation}
We again see $\Omega_{\rm orb} \ne \Omega_{\rm H}$ when the field and
the response are aligned (although $\Omega_{\rm orb} \to \Omega_{\rm
H}$ as $l$ gets very large).
The analytical expansions and numerical data indicate that the
Newtonian fluid intuition for the geometry of tidal coupling simply
does not work well for strong-field black hole binaries, even
accounting for the teleological swap of ``lag'' and ``lead.'' Only in
the extremes can we make statements with confidence: when $\Omega_{\rm
H} \gg \Omega_{\rm orb}$, the tidal bulge will lag the orbit; when
$\Omega_{\rm orb} \gg \Omega_{\rm H}$, the bulge will lead the orbit.
But when $\Omega_{\rm orb}$ and $\Omega_{\rm H}$ are of similar
magnitude, we cannot make a clean prediction.
The tidal bulge is {\it not} locked to the orbit when $dE^{\rm H}/dt =
0$, at least using any scheme to define the lead/lag angle that we
have examined.
\section{Conclusions}
\label{sec:conclude}
In this paper, we have presented a formalism for computing tidal
distortions of Kerr black holes. Using black hole perturbation
theory, our approach is good for fast motion, strong field orbits, and
can be applied to a black hole of any spin parameter. We have also
developed tools for visualizing the distorted horizon by embedding its
2-dimensional surface in a 3-dimensional Euclidean space. For now,
our embeddings are only good for Kerr spin parameter $a/M \le
\sqrt{3}/2$, the highest value for which the entire horizon can be
embedded in a globally Euclidean space. Higher spins require either a
piecewise embedding of an equatorial ``belt'' in a Euclidean space,
and a region near the ``poles'' in a Lorentzian space, or else
embedding in a different space altogether.
Although our formalism is good for arbitrary bound orbits, we have
focused on circular and equatorial orbits for this first analysis.
This allowed us to validate this formalism against existing results in
the literature, and to explore whether there is a simple connection
between the tidal coupling of the hole to the orbit, and the relative
geometry of the orbit and the horizon's tidal bulge. We find that
there is no such simple connection in general. Perhaps not
surprisingly, strong-field black hole systems are more complicated
than Newtonian fluid bodies.
We plan two followup analyses to extend the work we have done here.
First, we plan to extend the work on embedding horizons to $a/M >
\sqrt{3}/2$, the domain for which we cannot use a globally Euclidean
embedding. Work in progress indicates that the simplest and perhaps
most useful approach is to use the globally hyperbolic 3-space $H^3$
{\cite{gibbons}}. This allows us to treat the entire range of
physical black hole spins, $0 \le a/M \le 1$, using a single global
embedding space. Second, we plan to examine tidal distortions from
generic --- inclined and eccentric --- Kerr orbits. The circular
equatorial orbits we have studied in this first paper are stationary,
as are the tidal fields and tidal responses that arise from them. If
one examines the system and the horizon's response in a frame that
corotates with the orbit, the tide and the horizon will appear static.
This will not be the case for generic orbits. Even when viewed in a
frame that rotates at the orbit's mean $\phi$ frequency, the orbit
will be dynamical, and so the horizon's response will likewise be
dynamical. Similar analyses for Schwarzschild have already been
presented by Vega, Poisson, and Massey {\cite{vpm11}}; it will be
interesting to compare with the more complicated and less symmetric
Kerr case.
An extension of our analysis may be useful for improving initial data
for numerical relativity simulations of merging binary black holes.
One source of error in such simulations is that the black holes
typically have the wrong initial geometry --- unless the binary is
extremely widely separated, we expect each hole to be distorted by
their companion's tides. Accounting for this in the initial data
requires matching the near-horizon geometry to the binary's spacetime
metric; see {\cite{chu14}} for an up-to-date discussion of work to
include tidal effects in a binary's initial data. Much work has been
done on binaries containing tidally deformed Schwarzschild black holes
{\cite{alvi,yunesetal,mcdanieletal}}, and efforts now focus on the
more realistic case of binaries containing spinning black holes
{\cite{gallouinetal,chu14}}. With some effort (in order to get the
geometry in a region near the horizon, not just on the horizon), we
believe it should be possible to use this work as an additional tool
for extending the matching procedure to the realistic orbital
geometries of rotating black holes.
\acknowledgments
We thank Eric Poisson for useful discussions and comments, as well as
feedback on an early draft-in-progress on this paper; Robert Penna for
helpful comments and discussion, particularly regarding non-Euclidean
horizon embeddings; Nicol\'as Yunes for suggesting that this technique
might usefully connect to initial data for binary black holes; Daniel
Kennefick for helpful discussions as this project was originally being
formulated; and this paper's referee for very helpful comments and
feedback. This work was supported by NSF grant PHY-1068720. SAH
gratefully acknowledges fellowship support by the John Simon
Guggenheim Memorial Foundation, and sabbatical support from the
Canadian Institute for Theoretical Astrophysics and Perimeter
Institute for Theoretical Physics.
|
\section{Introduction}
\label{sec:Intro}
It is known that the homeomorphism relation on Polish spaces
is $\Sigma^1_2$~\cite{Gao} and
${\Sigma_1^1}$-hard~\cite[Thm 22]{FerLouRos}. On the other hand, it
is known that restricted to compact spaces
this homeomorphism relation is reducible
to an orbit equivalence relation induced by continuous action of a Polish group
which is known to be strictly below ${\Sigma_1^1}$-complete. This result
is extended to locally compact spaces in Theorem~\ref{thm:locCom}.
The main result of this paper is that this ``nice'' property of
locally compact spaces breaks when just one point is added to them:
The homeomorphism relation on the $\sigma$-compact spaces of the form
$V\cup \{x\}$ where $x\in \R^3$ is fixed and $V\subset \R^3$ is open
falls somewhere in between: it is ${\Sigma_1^1}$ and the equivalence relation
known as $E_1$ is continuously reducible to it
(Theorem~\ref{thm:NonClass}). This implies that this homeomorphism
relation is not classifiable in a Borel way by any orbit equivalence
relation arising from a Borel action of a Polish group.
The proof relies on known results in knot theory and low dimensional
topology. We hope that these
methods can be helpful in approaching Question~\ref{open:Main}
and other questions listed in Section~\ref{sec:Further}.
Sections \ref{sec:KnotTheory} and \ref{sec:BkgrndDST}
are devoted to the required preliminaries. In Section~\ref{sec:NonClass}
we prove the main non-classification result.
In the final sections the research topic of classifying
homeomorphism relations is looked at in more
detail: In Section~\ref{sec:Other} it is reviewed what
positive results there are in classification of homeomorphism
relations and in
Section~\ref{sec:Further} a list of open questions is given
in the area.
\section{Preliminaries in Topology and Knot Theory}
\label{sec:KnotTheory}
In this section we go through those definitions and lemmas in
knot theory and topology that we need in the proofs later.
We assume that the reader is familiar with the notion of
the first homology group $H_1(X)$ of a topological
space $X$. The
standard definitions can be found
for example in~\cite{Hatcher}.
We denote by $\R^n$ the $n$-dimensional Euclidean space and by
$\S^n$ the one-point compactification of it, i.e.
$\S^n=\R^n\cup\{\infty\}$ and the neighborhoods of $\infty$
are the sets of the form $\{\infty\}\cup(\R^n\setminus C)$
where $C$ is compact. By ${\operatorname{int}} A$ we denote
the topological interior of $A$ and by $\bar A$ the closure.
\subsection{Hausdorff Metric and Path Connected Subspaces}
\begin{Def}\label{def:HausdorffMetric}
Let $X$ be a compact metric space. The space of all non-empty
compact subsets of $X$ is denoted by $K(X)$. We equip $K(X)$ with
the Hausdorff-metric: An \emph{$\varepsilon$-collar} of a set $C\subset X$ is
the set
$$C_\varepsilon=\{x\mid d(x,C)< \varepsilon\}$$
and the Hausdorff-distance between two sets in $K(X)$ is determined
by
$$d_{K(X)}(C,C')=\max\{\inf\{\varepsilon\mid C\subset C'_\varepsilon\},\inf\{\varepsilon\mid C'\subset C_\varepsilon\}\}.$$
\end{Def}
The following facts are standard to verify.
\begin{Fact}\label{fact:Hausdorffmetric}
Let $X$ be a compact metric space.
Then $K(X)$ is compact
and if $(C_i)_{i\in\N}$ is a converging
sequence in $K(X)$ and $C_*$ is its limit, then
\begin{enumerate}
\item for every $x_*$ we have
$x_*\in C_*$ if and only if there is a sequence
$x_i$ converging to $x_*$ with $x_i\in C_i$
for all~$i\in\N$. \label{fact:Haus1}
\item if every $C_i$ is connected, then $C_*$ is connected.
\label{fact:Haus2}
\label{fact:Haus3}
\end{enumerate}
\end{Fact}
\begin{Def}\label{def:DenselyPathConnected}
A subset $A\subset \R^n$ is \emph{path metric} if the
distance between two
points is given by
$$d_E(x,y)=\inf\{L(\gamma)\mid \gamma\subset A\text{ is a path joining }x\text{ and }y\}$$
where $d_E$ is the Euclidean distance and
$L(\gamma)$ is the length of the path.
Equivalently $A$ is path metric if and only if
for every two points $x,y\in A$ and $\varepsilon>0$ there is a path $\gamma\subset A$
connecting $x$ to $y$ and $L(\gamma)<(1+\varepsilon)d_E(x,y)$.
\end{Def}
\begin{Lemma}\label{lem:HDPathMetric}
If the Hausdorff dimension of a closed $A\subset\R^n$ is less than
$n-1$, then $\R^n\setminus A$ is path metric.
\end{Lemma}
\begin{proof}
Let $D_0$ be the $(n-1)$-dimensional unit disc
$$D_0=\{(x_1,\dots,x_{n-1})\mid x_1^2+\dots+x_{n-1}^2<1\}.$$
and let $C_0$ be the cylinder
$$D_0\times [0,1]\subset \R^n.$$
For $x,y\in\R^n$ denote by $[x,y]$ the straight
line segment connecting $x$ and $y$.
Suppose $A_0\subset C_0$ and assume that for every
$(x_1,\dots,x_{n-1})\in D_0$ the set
$$A_0\cap [(x_1,\dots,x_{n-1},0),(x_1,\dots,x_{n-1},1)]$$
is non-empty.
Then $A_0$ must have Hausdorff dimension at least $n-1$:
$A_0$ can be projected onto $D_0$ with the Lipschitz map
$$(x_1,\dots,x_{n-1},x_n)\mapsto (x_1,\dots,x_{n-1},0),$$
the latter has Hausdorff dimension $n-1$ and the Hausdorff dimension
cannot increase in a Lipschitz map. Therefore we have the following claim:
\begin{claim}
If $A_0\subset C_0$ has Hausdorff dimension less than $n-1$, then
there is $(x_1,\dots,x_{n-1})\in D_0$ such that
$[(x_1,\dots,x_{n-1},0),(x_1,\dots,x_{n-1},1)]\cap A_0=\varnothing.$ \qed
\end{claim}
Let $x,y\in \R^n\setminus A$ and let $\varepsilon>0$. Since $A$ is closed there is
$\d<\varepsilon/2$ such that $\bar B(x,\d)\cap A=\bar B(y,\d)\cap A=\varnothing$.
Let $P_x$ and $P_y$ be $(n-1)$-dimensional affine hyperplanes passing
through $x$ and $y$ respectively and which are orthogonal to $x-y$.
Then there is an affine map $f\colon \R^n\to\R^n$ such that
$f[P_x]=\R^{n-1}\times\{0\}$, $f[P_y]=\R^{n-1}\times \{1\}$
and $f[P_x\cap \bar B(x,\d)]=D_0\times\{0\}$. Since $\dim_H(A)<n-1$,
also $\dim_H(f[A])<n-1$ (because $f$ is Lipschitz) and so by the claim
above there is a line segment $s$ passing from $f[\bar B(x,\d)\cap P_x]$
to $f[\bar B(y,\d)\cap P_y]$ outside $f[A]$ which is orthogonal to
$\R^{n-1}\times\{0\}$. By applying $f^{-1}$ to $s$,
we obtain a straight line segment passing from $\bar B(x,\delta)\cap P_x$
to $\bar B(y,\delta)\cap P_y$ orthogonal to~$P_x$.
Now by connecting the endpoints of $f^{-1}[s]$
to $x$ and $y$ we obtain a path outside $A$ of length
at most $d(x,y)+2\d=d(x,y)+\varepsilon$ connecting these two points.
\end{proof}
\begin{Lemma}\label{lemma:CauchyComponent}
Suppose $X,X'\subset \R^n$ are such that
$X$ is a path
metric space and there is a homeomorphism
$h\colon X\to X'$.
If $(x_i)_{i\in\N}$ is a Cauchy sequence
in $X$ converging in $\R^n$ to some
point $x\in \R^n\setminus X$,
then all the accumulation points of $(h(x_i))_{i\in\N}$
lie in the same component of~$\R^n\setminus X'$.
In particular, if this component is a singleton,
then $(h(x_i))_{i\in\N}$ is also a Cauchy sequence.
\end{Lemma}
\begin{proof}
Let $(x_i)_{i\in\N}$ be as in the statement.
Suppose for a contradiction
that $y^1$ and $y^2$ are two points in two different components of
$\R^n\setminus X'$ that are accumulation points
of $(h(x_i))_{i\in\N}$ and let $(x^1_{i})_{i\in\N}$ and
$(x^2_{i})_{i\in\N}$ be subsequences of $(x_i)_{i\in\N}$
such that $(h(x^1_i))_{i\in\N}$ and $(h(x^2_i))_{i\in\N}$
converge to $y^1$ and $y^2$ respectively.
For $k\in\{1,2\}$ and $i\in\N$ denote
$y^k_i=h(x^k_i)$.
For each $i\in \N$ let $\gamma_i$ be a path in $X$ connecting $x^1_i$ to
$x^2_{i}$ such that $L(\gamma)<(1+2^{-i})d(x^1_i,x^2_{i})$. We think of
the paths as compact subsets of $\S^n$. The sequence $\{\gamma_i\mid
i\in\N\}$ converges in $K(\S^n)$ to $\{x\}$. Consider the sequence
$(h[\gamma_i])_{i\in\N}$. It is a sequence of compact subsets of
$\S^n$, so it is a sequence of elements of $K(\S^n)$. The latter is
compact, so there is a converging subsequence:
$(h[\gamma_{i(k)}])_{k\in\N}$. Denote by $\gamma$ the limit of that
sequence. By Fact~\ref{fact:Hausdorffmetric}.\ref{fact:Haus1},
we have $y^1,y^2\in \gamma$ since $y^1_i,y^2_i\in h[\gamma_i]$ for
all~$i\in\N$ and additionally, since every element in the sequence
is connected, $\gamma$ is also connected by
Fact~\ref{fact:Hausdorffmetric}.\ref{fact:Haus2}.
Since $y^1$ and $y^2$ lie
in different components of $\R^n\setminus X'$,
there must be a point $z$ in $\gamma$
which is in $X'$. Now, by
Fact~\ref{fact:Hausdorffmetric}.\ref{fact:Haus1}
we can find a sequence $z_k\in h[\gamma_{i(k)}]$, $k\in\N$,
such that $(z_k)_{k\in\N}$ converges to $z$. But
$h^{-1}(z_k)$ lies in $\gamma_{i(k)}$ and so $(h^{-1}(z_k))_{k\in\N}$
converges to $x$. This is a contradiction, because
$x\notin \operatorname{dom} h=X$, but $z\in \operatorname{ran} h= X'$.
\end{proof}
\subsection{Separation Theorems}
\label{sec:Sep}
Here we state, for the sake of completeness,
two known results from finite dimensional topology that we
will need.
\begin{Thm}[Jordan-Brouwer Separation Theorem]\label{thm:JordanBrouwer}
Let $h\colon \S^{n-1}\to \S^n$ be an embedding. Then
$\S^n\setminus h[\S^{n-1}]$ consists of two open connected
components. \qed
\end{Thm}
\begin{Thm}(A Generalization of the Schönflies Theorem
by M. Brown~\cite{Brown})\label{thm:Brown}
Let $h\colon \S^{n-1}\times [0,1]\to \S^n$ be an embedding.
Then the closures of the complementary domains of
$h[\S^{2}\times \{\frac{1}{2}\}]$ are topological
$n$-cells, i.e. homeomorphic to closed balls.
In particular there is a self-homeomorphism of $\S^n$ which
takes $h[\S^{n-1}\times \{\frac{1}{2}\}]$ to the standard $\S^{n-1}$.
\qed
\end{Thm}
\subsection{Knot Theory}
We present the basics of knot theory here as neatly as possible and
account only for the facts necessary for the present paper. Unless a
specific reference is given below, the reader is referred to the classical
textbooks on knot theory \cite{BurZie,Kauf,Mur} for the details and
omitted proofs.
A \emph{knot} is an embedding
$K\colon \S^1\to \S^3$. We often identify a knot with its
image, $\operatorname{ran} K$.
This is in particular justified by the following equivalence
relation on knots:
\begin{Def}
Two knots $K_0,K_1\colon \S^1\to \S^3$ are equivalent, if there is
a homeomorphism $h\colon \S^3\to \S^3$ with
$$K_0=h\circ K_1.$$
In literature this homeomorphism is often required to be orientation
preserving in which case this equivalence relation coincides with
the so-called \emph{ambient isotopy}, but we do not require $h$
to be orientation preserving.
A knot is \emph{trivial} if it is equivalent to the standard
embedding $\S^1\hookrightarrow \S^3$. A knot is \emph{tame} if it is
equivalent to a smooth or a piecewise linear knot. As usual in knot
theory, we consider only tame knots.
\end{Def}
The following is a basic fact of knot theory:
\begin{Fact}
There are infinitely many non-equivalent knots. \qed
\end{Fact}
A not so basic fact is the following theorem:
\begin{Thm}(\cite{GorLue})\label{thm:GordonLuecke}
If two knots have homeomorphic complements,
then they are equivalent. \qed
\end{Thm}
\begin{Def}
Let $K$ be a knot in $\R^3$. A \emph{Seifert surface} $S$ of $K$
is a compact orientable connected $2$-manifold with boundary
$M\subset \R^3$ whose
interior lies in $\R^3\setminus K$ and the boundary
is exactly~$K$.
\end{Def}
\begin{Fact}\label{fact:ExistsSeifertS}
For every open ball $B$ containing $K$ there exists
a Seifert surface $S\subset B$ of $K$. \qed
\end{Fact}
\begin{Fact}(\cite[5.D]{Rolfsen})\label{fact:LinkingSeifert}
Let $K$ be a knot and $w$ a closed curve in $\R^3\setminus K$.
The following are equivalent:
\begin{itemize}
\item $w\cap S\ne\varnothing$ for every Seifert surface $S$ of $K$,
\item $w$ represents a non-trivial element in $H_1(\R^3\setminus K)$. \qed
\end{itemize}
\end{Fact}
\begin{Fact}\label{fact:NonTrHomology}
For every knot $K$ we have $H_1(\S^3\setminus K)\cong \Z$. \qed
\end{Fact}
\subsection{Preserving Knot Types}
The goal of this section is to prove Lemma~\ref{lemma:FindPoints}
which says that if we carve out infinitely many knots from $\R^3$
in a certain way, then a self-homeomorphism of the left-over space
will, in an approximate way,
respect the knot-types of the carved knots.
\begin{Def}\label{def:Properties}
Let $(B_n)_{n\in\N}$ be a sequence of closed balls in $\R^3$,
$(K_n)_{n\in\N}$ a sequence of knots, $Q\subset \R^3$ and $P\subset \R^3$.
Here we list some properties for these sets which we will later
refer to.
\begin{itemize}
\item[B1] All the balls are disjoint from each other
and are contained in a bounded region, i.e. there is $r$
such that $\Cup_{n\in\N}B_n\subset B(0,r)$.
\item[B2] If $x$ is a limit of a sequence $(x_i)_{i\in\N}$ such that
for all $i\in\N$ the point $x_i$ is in the ball $B_{n_i}$ and
for all $i<j$, $n_i\ne n_j$,
then $x$ is not in any of the balls.
$Q$ is the set of such points $x$.
\item[B3] $P\supset Q$, every connected component of $P$ contains
a point in $Q$ and for all $n$ there is $\varepsilon>0$ such that
$P\cap (B_n)_{\varepsilon}=\varnothing$.
(Recall the definition of $\varepsilon$-collar,
Definition~\ref{def:HausdorffMetric}).
\item[B4] $K_n\subset {\operatorname{int}} B_n$.
\item[B5] $X=\R^3\setminus (P\cup\Cup_{n\in\N}K_n)$ is path metric
(Definition~\ref{def:DenselyPathConnected}).
\end{itemize}
\end{Def}
\begin{Lemma}\label{lemma:FindPoints}
Suppose $(B_n)_{n\in\N}$,
$(K_n)_{n\in\N}$, $Q$ and $P$ as well as
$(B_n')_{n\in\N}$,
$(K_n')_{n\in\N}$, $Q'$ and $P'$
satisfy the properties B1 -- B5. Let
$$X=\S^3\setminus (P\cup\Cup_{n\in\N}K_n)$$
and
$$X'=\S^3\setminus (P'\cup\Cup_{n\in\N}K'_n).$$
Suppose further that
$X$ and $X'$ are homeomorphic and $h$ is the homeomorphism.
Then there is a bijection $\rho\colon\N\to\N$
such that for all $n\in\N$ we have
that $K_n$ and $K_{\rho(n)}'$ have the same knot-type and
for some $z\in B_n\setminus K_n$ we have
$h(z)\in B_{\rho(n)}'\setminus K_{\rho(n)}'$.
\end{Lemma}
\begin{proof}
Fix $n\in \N$. By the Jordan-Brouwer separation theorem
(Theorem~\ref{thm:JordanBrouwer}) the
complement of $h[\partial B_n]$ in $\S^3$ consists of two open connected
components, say $Y_1$ and $Y_2$. In this case, however, we can prove
even more, namely that $Y_1$ and $Y_2$ are homeomorphic to open
balls and that there is a self-homeomorphism of $\S^3$ wich takes
$h[\partial B_n]$ to $\S^2$. Let $\varepsilon$ be small enough so that $(\partial B_n)_\varepsilon\cap
B_k=\varnothing$ for all $k\ne n$, $(\partial B_n)_\varepsilon\cap P=\varnothing$
and $(\partial B_n)_{\varepsilon}\cap K_n=\varnothing$. This is possible by B2, B3
and~B4.
Let
$$f\colon \S^2\times [0,1]\to (\partial B_n)_{\varepsilon}$$
be a homeomorphism such that
$f[\S^2\times \{\frac{1}{2}\}]=\partial B_n$.
We can think of $h\circ f$ as an embedding of $\S^2\times [0,1]$
into $\S^3$. Now apply the the generalized Schönflies theorem
(Theorem~\ref{thm:Brown}) to~$h\circ f$.
Since $\partial B_n$ divides $X$ into two disjoint components
as well as $h[\partial B_n]$ divides $X'$, $h$ takes
them to one another. Assume without loss of generality that
$h[{\operatorname{int}} B_n\setminus K_n]=Y_1\cap X'$.
\begin{claim}\label{claim:ConnectedThing}
The space $Y_1\setminus X'$ is connected.
\end{claim}
\begin{proof}
For this we need a slight modification of the argument
used to prove Lemma~\ref{lemma:CauchyComponent}.
(Note that $B_n\setminus K_n$ is path metric.)
Suppose there was two components $A$ and $B$
of $Y_1\setminus X'$
and let $(x_1,y_1,x_2,y_2,\dots)$ be a sequence such that
$(x_i)_{i\in\N}$ converges (in $\S^3$) to a point in $A$
and $(y_i)_{i\in\N}$ converges to a point in $B$.
Now $(h^{-1}(x_1),h^{-1}(y_1)\cdots)$ can only have accumulation
points in~$K_n$ (because the accumulation points cannot be in $X$).
Pick Cauchy subsequences from both
$(h^{-1}(x_i))_{i\in\N}$ and $(h^{-1}(y_i))_{i\in\N}$ and
denote $(z_i)_{i\in\N}$ and~$(w_i)_{i\in\N}$.
Since $z=\lim_{i\to\infty}z_i$ and $w=\lim_{i\to\infty}w_i$
lie both in the knot, using the fact that $B_n\setminus K_n$
is path metric, it is possible to connect
$z_i$ to $w_i$ by a curve $\gamma_i$ lying in $B_n\setminus K_n$
such that the sequence $(\gamma_i)_{i\in\N}$ converges in $K(\S^3)$
to a subset of~$K_n$. Now pick (in $K(\S^3)$)
a converging subsequence
$(\xi_{j})_{j\in\N}$ of $(h[\gamma_i])_{i\in\N}$. These are
connected sets containing $h(z_i)$ and $h(w_i)$. Therefore
the limit in $K(\S^3)$ must intersect both $A$ and $B$ and
since it is connected, it must contain a point $p$
in $Y_1\cap X'= h[{\operatorname{int}} B_n]$.
By Fact~\ref{fact:Hausdorffmetric} there is a Cauchy sequence
$(p_j)_{j\in\N}$ with $p_j\in \xi_{j}$ converging
to $p$ but $((h^{-1}(p_j))_{j\in\N}$ does not have accumulation
points in $B_n\setminus K_n$. This is a contradiction.
\end{proof}
Thus, $Y_1\setminus X'$ is a connected component of $\S^3\setminus X'$
Note that this component must be in the interior of
$\bar Y_1$, so it cannot be a subset of~$P$, by~B2, B3 and B4.
Thus, it is $K_m'$
for some $m$.
Since $h$ is a homeomorphism we have that
$${\operatorname{int}} B_n\setminus K_n\approx Y_1\setminus K_m'.$$
Since $Y_1\approx {\operatorname{int}} B_n\approx \R^n$, we can conclude from
Theorem~\ref{thm:GordonLuecke}
that $K_n$ and $K_{m'}$ have the same knot-type.
By symmetry arguments using the fact that $h$ is a homeomorphism,
this establishes a map $n\mapsto m$ which is
actually bijective, so denote this bijection by~$\rho$.
Let $\gamma\subset B_n\setminus K_n$ be a closed curve representing
a non-trivial cycle in $H_1(B_n\setminus K_n)$ (such exists
by Fact~\ref{fact:NonTrHomology}).
Then $h[\gamma]$ will be a non-trivial cycle in $h[B_n]$.
We would like to show that $h[\gamma]$ is also non-trivial in
$(\S^3\setminus Y_1)\cup h[B_n]$. But since we established a homeomorphism
of $\S^3$ to itself taking $h[\partial B_n]$ to $\S^2$, we know that
if $\gamma$ bounds a disk $D\subset (\S^3\setminus Y_1)\cup h[B_n]$,
this disk can be isotoped to a disk $D'\subset Y_1$ keeping $Y_1\cap D$ fixed.
Let $S$ be a Seifert surface of $K_m'$
contained in $B_m'$
(see Fact~\ref{fact:ExistsSeifertS}). Then by
Fact~\ref{fact:LinkingSeifert} there is a point
$z'\in h[\gamma]\cap S$. Let $z=h^{-1}(z')$. This completes
the proof, since $z'\in B_{m'}$.
\end{proof}
\section{Preliminaries in Descriptive Set Theory}
\label{sec:BkgrndDST}
\begin{Def}
A \emph{Polish space} is a separable topological space which is
homeomorphic to a complete metric space.
\end{Def}
The most common examples of
Polish spaces are $\R$, $\C$ and $\N^\N$ in the Tychonov product
topology. Less common examples include the space of all homeomorphisms
${\operatorname{Hom}}(X)$ of a compact Polish space $X$ in the $\sup$-metric
(see Fact~\ref{fact:HomPolish}) and the space of compact subsets of
a compact space $X$ in the Hausdorff metric denoted $K(X)$
(see Fact~\ref{fact:Hausdorffmetric}).
\begin{Fact}(\cite{Kechris})\label{fact:Gdelta}
A subset of a Polish space is Polish in the subspace topology
if and only if it is a $G_\delta$ subset. \qed
\end{Fact}
\begin{Fact}(\cite[Theorem 3.11 and Example 9B(8)]{Kechris})
\label{fact:HomPolish}
For a compact Polish space $X$ equipped with the metric
$d_X$, the space ${\operatorname{Hom}}(X)$ of homeomorphisms of
$X$ in the $\sup$-metric,
$\delta(h,g)=\sup\{d_X(h(x),g(x))\mid x\in X\}$
is a Polish space. \qed
\end{Fact}
\begin{Def}\label{def:Reduction}
Suppose $E$ and $E'$ are equivalence relations on
Borel subsets $A$ and $A'$ of Polish
spaces $X$ and $X'$ respectively. The equivalence relation
$E$ is \emph{Borel reducible to} $E'$, denoted $E\le_B E'$,
if there is a Borel map $f\colon A\to A'$ such that
$$\forall x,y\in A\big((x,y)\in E\iff (f(x),f(y))\in E'\big).$$
We say that an equivalence relation $E$ is \emph{universal}
among a set $X$ of equivalence relations, if $E\in X$ and for all
$E'\in X$ we have $E'\le_B E$.
\end{Def}
A lot is known about the partial order $\le_B$ on
analytic equivalence relations which are defined on standard Borel spaces. A
thorough treatment can be found in~\cite{Gao}.
Preface in~\cite{Hjorth} gives a good glimpse of available applications.
Here is an example of an equivalence relation which we will need:
\begin{Def}\label{def:E1}
Let $(2^\N)^\N$ be the space of sequences of elements of $2^\N$
(the Cantor space). The topology on both $2^\N$ and $(2^\N)^{\N}$
is given by the Tychonov product topology. Let $E_1$
be the equivalence relation given by:
$$((r_n)_{n\in \N},(s_n)_{n\in \N})\in E_1\iff
\exists m\forall k>m(r_k=s_k).$$
\end{Def}
Another wide class of equivalence relations is given by Polish
group actions:
\begin{Def}
Let $G$ be a Polish group acting
in a Borel way on
a Polish space $X$. Let $E^X_G$ be the equivalence relation
where $x,y\in X$ are equivalent if and only if there exists
$g\in G$ such that $y=gx$. This is called the
\emph{orbit equivalence relation} induced by this (Borel)
action of a Polish group.
\end{Def}
Many natural equivalence relations, in particular the isomorphism
on countable structures (see the end of this section),
can be viewed as orbit equivalence relations
induced by Polish group actions.
A proof of the following can be found in~\cite[Theorem 10.6.1]{Gao}.
\begin{Thm}\label{thm:KecLou}(Kechris-Louveau \cite{KecLou})
Let $E$ be any orbit equivalence relation induced
by a Borel action of a Polish group.
Then $E_1\not\le_B E$. \qed
\end{Thm}
\begin{Def}\label{def:StrangeSpaces}
Let $X$ be a compact Polish space.
For a fixed closed (and hence compact) subset $F\subset X$, let
$$K^{F}(X)=\{A\in K(X)\mid F\subset A\}.$$
(See Definition~\ref{def:HausdorffMetric} for the definition of~$K(X)$.)
Then $K^{F}(X)$ is a closed subspace of $K(X)$ and so Polish
itself by Fact~\ref{fact:Gdelta}. Let
$$K^{F}_*(X)=\{(X\setminus A)\cup F\mid A\in K^F(X)\}.$$
The Polish topology on $K^{F}_*(X)$ is induced by the bijection
$K^F(X)\to K^F_*(X)$ given by $A\mapsto (X\setminus A)\cup F$.
\end{Def}
Let $F\subset X$ be closed. Then elements of
$K^F_*(X)$ are of the form $U\cup F$ where $U$ is an open
set disjoint from $F$. Therefore elements of this space are
$\sigma$-compact $G_\delta$-subsets.
Using Fact~\ref{fact:Gdelta} we obtain:
\begin{Fact}\label{fact:TheyreAllPolish}
For a fixed closed $F\subset X$ and $X$ compact
$K^F_*(X)$ consists of $\sigma$-compact Polish spaces.
\end{Fact}
\begin{Def}\label{def:Relations}
For a fixed closed $F\subset \S^3$,
let $\approx^F$ be the homeomorphism
relation on the space $K^F_*(\S^3)$.
\end{Def}
The main result of this paper (Theorem~\ref{thm:NonClass})
can be now stated: for a fixed $x\in \S^3$,
$E_1\le_B\ \approx^{\{x\}}$.
A countable model in a fixed vocabulary
with universe $\N$ can be coded as an element of
$2^{\N}$ in such a way that each $\eta\in 2^\N$ in fact represents
some model. There are many nice ways to do this, see for
example~\cite{Gao}. Let $\cong$ be the equivalence relation of isomorphism.
It is well known that given a vocabulary and any collection of countable models
in this vocabulary
whose set of codes is Borel, $\cong$ is reducible
to $\cong_G$ where $\cong_G$
is the isomorphism of graphs, i.e. vocabulary consists of one binary
symbol and the models are infinite graphs with domain~$\N$.
This equivalence relation is induced by the action of the infinite
symmetric group $S_\infty$ (which is Polish in
the standard product topology).
A corollary to Theorem~\ref{thm:NonClass} which follows from
Theorem~\ref{thm:KecLou} is that $\approx^{\{x\}}$
is not reducible to~$\cong_G$, although
this has been proved for the homeomorphism relation
on compact spaces already by Hjorth~\cite[]{Hjorth}.
The original motivation of this research was the following,
stronger, question:
\begin{Open}\label{open:Main}
Is $\approx^{\varnothing}$ reducible to $\cong_G$?
\end{Open}
Note that $\approx^{\varnothing}$ is just the homeomorphism relation
on open subsets of $\S^3$. See Section~\ref{sec:Further} for a discussion
on this and other open questions.
\subsection{Parametrization}
\label{ssec:Par}
As was pointed out, the space $K^F_*(X)$ consists of
$\sigma$-compact Polish spaces (Fact~\ref{fact:TheyreAllPolish}).
However, not \emph{all} $\sigma$-compact Polish spaces are found in~$K^F_*(X)$.
There are different ways to parametrize different classes of Polish spaces such
as compact, locally compact, $\sigma$-compact, $n$-manifolds and so on. In this section we will
present these different ways and show that essentially it does not matter which
one we choose, all of them being essentially equivalent in some sense.
Additionally in this section we introduce many new notations for various homeomorphism
relations. A helpful list of notations can be found in Section~\ref{sec:Conclusion}.
In \cite{HjoKec1} Hjorth and Kechris give a simple parametrization of
all Polish spaces. Their parametrization, let us call it the
\emph{Hjorth-Kechris parametrization}, consists of two-fold sequences
$\eta\in \R^{\N\times\N}$ which satisfy the requirements for a metric
on $\N$. The set of such $\eta$ is easily seen to be
Borel. Then the space $X(\eta)$ is obtained as a
completion of this countable metric space.
Another way to parametrize all Polish spaces is to view them as closed
subsets of the Urysohn universal space $U$. Denote the space of all closed
subsets of $U$ by $F(U)$. It can be equipped with a standard Borel structure which is inherited
from $K(\bar U)$ where $\bar U$ is a compactification of $U$
(see~\cite[Thm~12.6]{Kechris}). The Borel sets of $F(U)$ are generated by the sets
of the form
$$\{F\in F(U)\mid F\cap O\ne \varnothing\}$$
for some open $O\subset U$. This Borel structure is also generated by the
\emph{Fell topology} generated by the sets of the form
\begin{equation}
\label{eq:StandBor}
\{F\in F(U)\mid F\cap K=\varnothing\land F\cap O_1\ne\varnothing\land\dots\land F\cap O_n\ne\varnothing\},
\end{equation}
where $K$ varies over $K(U)$ and $O_i$ are open sets in~$U$ \cite[Exercise 12.7]{Kechris}.
Let us show that
these parametrizations are essentially equivalent. The universality property of $U$ is that
given any finite metric space $H$ and $x\in H$, every isometric embedding
of $H\setminus \{x\}$ into $U$ extends to an isometric embedding of $H$ into $U$.
Thus, given a countable metric space as defined by $\eta$ as above, it can
be isometrically embedded into $U$. The closure of the image will then
be homeomorphic (and even isometric) to $X(\eta)$.
We want to show that there are Borel reductions reducing the homeomorphism of
Polish spaces in one parametrization to the other.
To show this, let us
define an ``intermediate'' parametrization. Let $U^\N$ be the set of all
countable sequences in $U$. Each such sequence $\xi$ corresponds to the Polish space
$Y(\xi)$ obtained as its closure taken in $U$. Let $f_1\colon U^\N\to \R^{\N\times\N}$
be defined by $f_1(\xi)=\eta$ where $\eta(n,m)=d_U(\xi(n),\xi(m))$. Obviously
$X(\eta)$ and $Y(\xi)$ are isometric and $f_1$ is continuous. Let $f_2\colon U^\N\to F(U)$
be the map which takes $\xi$ to the closure of $\{\xi(n)\mid n\in\N\}$ in $U$.
\begin{Lemma}
There are Borel functions $g_1\colon \operatorname{ran} (f_1)\to U^\N$ and $g_2\colon F(U)\to U^\N$
such that $f_1\circ g_1=\operatorname{id}$ and $f_2\circ g_2=\operatorname{id}$.
\end{Lemma}
\begin{proof}
For $g_2$ we will use \cite[Cor 5.4]{Sri} which says that if $f\colon X\to Y$ is a
Borel function between Polish spaces $X$ and $Y$ such that $f[V]$ is open for
all open $V\subset X$, $f^{-1}[V]$ is $F_\sigma$ for all open $V\subset Y$ and
$f^{-1}\{y\}$ is $G_\delta$ for all $y\in Y$, then there is $g\colon Y\to X$
such that $f\circ g=\operatorname{id}_Y$.
It is easy to see that the inverse image under $f_2$ of a set of the form \eqref{eq:StandBor}
is $F_\sigma$ in $U^\N$, so in particular $f_2$ is Borel. Additionally, given a closed
set $C\in F(U)$, the inverse image of the singleton $f_2^{-1}\{C\}$ is $G_\delta$.
To see this, let $Q=\{q_n\mid n\in\N\}$ be a dense
countable subset of~$C$. Then $f_2^{-1}\{C\}$
is the intersection of $\{\xi\mid \operatorname{ran}(\xi)\subset C\}$ and the sets
$$O(k,m)=\{\xi\mid \exists n\in\N(\xi(n)\in B(q_k,1/m))\}.$$
The former is closed and the latter are open, so the intersection is~$G_\d$.
Let $V\subset U^\N$ be a basic open set. It is of the form
\begin{equation}
O_0\times\cdots\times O_n\times U^{\N\setminus \{0,\dots,n\}}.\label{eq:openset}
\end{equation}
To see that $f_2[V]$ is open in $U^\N$
note that it can be represented
in the form of \eqref{eq:StandBor} with $K=\varnothing$ and $O_i$ as in \eqref{eq:openset};
thus $f_2$ is an open map. By
\cite[Cor 5.4]{Sri} there is a Borel $g_2\colon F(U)\to U^\N$
such that $f_2\circ g_2=\operatorname{id}$.
Now consider $f_1$. It is continuous and the inverse image of a singleton
is closed. By \cite[Thm 35.46]{Kechris} (or again by \cite[Cor 5.4]{Sri})
there is $g_1\colon \operatorname{ran} (f_1)\to U^\N$
such that $f_1\circ g_1=\operatorname{id}$. Note that $\operatorname{ran} (f_1)$ is merely the Borel subset of
$\R^{\N\times\N}$ on which
the operation $\eta\mapsto X(\eta)$ is well defined and produces a Polish space.
\end{proof}
Let
$\approx_P$ be the equivalence relation on
$\operatorname{ran} (f_1)\subset\R^{\N\times\N}$
where $\eta$ and $\eta'$ are equivalent if and only if $X(\eta)$ and
$X(\eta')$ are homeomorphic,
let $\approx_P'$ be the equivalence relation on $U^\N$ where
two sequences $\xi$ and $\xi'$ are equivalent if and only if $Y(\xi)$ and $Y(\xi')$
are homeomorphic,
and let $\approx''_P$ be the equivalence relation on $F(U)$ where two closed $C$ and $C'$
are equivalent if and only if they are homeomorphic.
Then $f_1$, $g_1$, $f_2$ and $g_2$ witness that these three equivalence relations,
$\approx_P$, $\approx'_P$ and $\approx''_P$ are all Borel reducible to each other.
Hjorth and Kechris showed that
the set of those $\eta$ for which $X(\eta)$ is compact and the set of those
for which it is locally
compact are both Borel subsets of $\R^{\N\times\N}$. Taking Borel inverse images
under $f_1$ and $g_2$ we obtain the same conclusion for the other parametrizations.
In \cite{HjoKec1} it is shown
that the set of complex $n$-manifolds is Borel.
One has to replace
``biholomorphic'' by ``homeomorphic'' in order to relax from
complex manifolds to (conventional) manifolds. But as also proved
in \cite{HjoKec1}, in the case of locally compact spaces,
a function is defined to be a homeomorphism in a Borel way.
Thus, the set of those $\eta$ for which $X(\eta)$ is an $n$-manifold
is Borel. Using the functions $f_1$, $g_1$, $f_2$ and $g_2$ we finally
obtain that the sets of $n$-manifolds in all the other parametrizations
are also Borel.
Denote by $\approx_P$ the homeomorhism relation on all Polish spaces
and by
$\approx_{loc}$, $\approx_c$ and $\approx_n$ the same relation restricted
to the sets of locally compact, compact Polish spaces and $n$-manifolds
respectively. From what is shown above it follows that the chosen parametrization is
irrelevant. Recall also the notation $\approx^{\{x\}}$ from Definition~\ref{def:Relations}.
All of these equivalence relations are defined on Borel subsets of Polish spaces.
The following easily follows from \cite[Exercise (27.9)]{Kechris}:
\begin{Fact}
The set of those $C\in F(U)$ which are $\sigma$-compact is ${\Pi_1^1}$-complete. \qed
\end{Fact}
Because, as custom is, we require in Definition~\ref{def:Reduction} that the domains
of equivalence relations are Borel subsets of Polish spaces, we do not talk directly
about the homeomorphism relation restricted to the $\sigma$-compact spaces.
However, by removing that requirement and relaxing from Borel sets to \emph{relatively Borel}
one could also talk about the Borel reducibility of $\approx_\sigma$, the homeomorphism relation
on $\sigma$-compact spaces, to other equivalence relations. From our results it would follow
in particular that $E_1\le_B\ \approx_\sigma$ and that $\approx_\sigma$ is not classifiably by any
equivalence relation induced by a Borel group action.
Yet another way to parametrize compact and locally compact spaces is
to view them as subsets of the Hilbert cube $I^\N$
where $I$ is
the unit interval. It is known that for every compact Polish space there is a
homeomorphic copy as a subset of $I^\N$.
For locally compact spaces we also obtain a parametrization:
By \cite[Theorem 5.3]{Kechris}, the one-point compactification of
every locally compact Polish space is a compact Polish space.
Now fix a point $x\in I^\N$ and for each $\xi\in (I^\N\setminus\{x\})^\N$ let
$Z(\xi)$ be the space
$\overline{\{\xi(n)\mid n\in\N\}}\setminus \{x\}$.
If $P$ is any locally compact
space, then let $\bar P=P\cup \{\infty\}$ be its one-point compactification.
There is an embedding of $\bar P$ into $I^\N$ and since $I^\N$ is homogenous
(see e.g.~\cite{Fort}), there is an embedding such that $\infty$ is mapped
to $x$. Thus, in the notation of \ref{def:StrangeSpaces}, $K^{\{x\}}(I^\N)$
is a space parametrizing all locally compact spaces. By using the fact
that $I^\N$ can be also isometrically embedded into the Urysohn space $U$,
one can use the methods from above to conclude that this parametrization
is in our sense equivalent to all the other parametrizations (i.e. the homeomorphism
relation is Borel bireducible with the corresponding relation in other parametrizations
and the relevant subsets such as $n$-manifolds are Borel subsets).
\paragraph{Summary.}
When proving a classification or a non-classification result for
any of $\approx_P$, $\approx^{\{x\}}$, $\approx_{loc}$, $\approx_{c}$, $\approx_n$
it is irrelevant
which of the parametrizations is used.
Additionally the sets of locally compact and compact spaces,
of $n$-manifolds and of the spaces in $K^{\{x\}}_*(\S^3)$ are Borel
no matter which parametrization is used.
\section{Non-classification of $\approx^{\{x\}}$}
\label{sec:NonClass}
This section is devoted to proving the main result:
\begin{Thm}\label{thm:NonClass}
The equivalence relation $E_1$ (Definition~\ref{def:E1})
is continuously reducible to the homeomorphism relation
on $K^{\{x\}}_*(\S^3)$ for any fixed $x\in\S^3$.
\end{Thm}
\begin{proof}
As before, we parametrize $\S^3$ as $\R^3\cup \{\infty\}$.
Obviously the choice of $x$ does not matter. In our case
$x=(1,1,\frac{1}{2})\in \R^3$ as will be seen below.
For every $n\in\N$, $k\in\N$ and $l\in \{0,1\}$, let
$B_{n,k,l}\subset \R^3$ be a closed ball with the center at
$(1-2^{-n},1-2^{-k},l)$ and radius $2^{-4(n+1)(k+1)}$.
Define $Q$, $P'$ and $P$ as follows:
\begin{eqnarray*}
Q&=&\{(1-2^{-n},1,l)\mid n\in\N,l\in\{0,1\}\}
\cup\{(1,1-2^{-k},l)\mid k\in\N,l\in\{0,1\}\},\\
P'&=&Q\cup \Cup_{n\in\N}\{(1-2^{-n},1,t)\mid t\in [0,1]\}\\
P&=&P'\cup \{(1,1,t)\mid t\in [0,1]\}\setminus \{(1,1,\frac{1}{2})\}.
\end{eqnarray*}
Thus, $(B_{n,k,l})$, $Q$ and $P$ satisfy the assumptions
B1, B2 and B3 from Definition~\ref{def:Properties}.
Let $\{P_{n,k,l}\mid n\in\N,k\in\N,l\in\{0,1\}\}$ be the
set of all (mutually different) knot types
indexed by the set $\N\times\N\times \{0,1\}$. Let
$\bar r=(r_n)_{n\in\N}\in (2^{\N})^{\N}$ be a sequence of elements
of $2^\N$. For each $(n,k,l)\in (2^{\N})^{\N}$, let
$K^{\bar r}_{n,k,l}$ be a (piecewise linear)
knot inside the interior of $B_{n,k,l}$. The
knot-type of $K^{\bar r}_{n,k,l}$ is determined as follows:
\begin{itemize}
\item If $n$ is odd, then it is $P_{n,k,l}$,
\item If $n$ is even and $r_{n/2}(k)=0$, then it is $P_{n,k,l}$,
\item If $n$ is even and $r_{n/2}(k)=1$, then it is $P_{n,k,1-l}$.
\end{itemize}
Let $R(\bar r)$ be
$\S^3\setminus (P\cup\Cup_{n,k,l}K^{\bar r}_{n,k,l}).$
Note that $R(\bar r)$ corresponds to $X$ in
Definition~\ref{def:Properties} and properties B4 and B5
are now also satisfied (B5 follows easily
from Lemma~\ref{lem:HDPathMetric} and the fact that
$\S^3\setminus X$ is a countable union of piecewise linear
curves and points). Notice also that
$R(\bar r)\setminus \{(1,1,\frac{1}{2})\}$
is an open set, so $R(\bar r)\in K^{\{(1,1,\frac{1}{2})\}}_*(\S^3)$.
In the following three
claims we will show that
$F$ is a continuous reduction: $\bar r$ and
$\bar r'$ are $E_1$-equivalent if and only if
$R(\bar r)$ and $R(\bar r')$
are homeomorphic.
\begin{claim}
Suppose $\bar r$ and
$\bar r'$ are $E_1$-equivalent. Then
$R(\bar r)$ and $R(\bar r')$
are homeomorphic.
\end{claim}
\begin{proof}
For every $(n,k)\in \N\times\N$ let $C_{n,k}$ be the convex hull
of $B_{n,k,0}\cup B_{n,k,1}$, a ``capsule''
containing $B_{n,k,0}$ and $B_{n,k,1}$ disjoint from
all other balls and from~$P$.
Denote for
simplicity $X=R(\bar r)$ and $X'=R(\bar r')$.
Now $C_{n,k}\cap X$ and $C_{n,k}\cap X'$ are homeomorphic
because both are complements of
two knots of types $P_{n,k,0}$
and $P_{n,k,1}$.
If $n$ is odd or $n$ is even and $r_{n/2}(k)=r'_{n/2}(k)$
then identity on $C_{n,k}$ witnesses this. Otherwise
there is a homeomorphism $g_{n,k}$ of $\S^3$ fixing
$\S^3\setminus C_{n,k}$
and taking $C_{n,k}\cap X$ to $C_{n,k}\cap X'$.
For each $(n,k)$, if $n$ is even and $r_{n/2}(k)\ne r'_{n/2}(k)$,
let $h_{n,k}=g_{n,k}$. Otherwise let $h_{n,k}$ be the identity on
$\S^3$. Let $\pi\colon \N\to \N\times\N$ be a bijection and define
a sequence of functions $(t_m)_{m\in\N}$ by induction as follows:
\begin{eqnarray*}
t_0&=&h_{\pi(0)}\\
t_{m+1}&=&h_{\pi(m+1)}\circ t_m.
\end{eqnarray*}
We claim that for every $x\in R(\bar r)$ the limit
$t(x)=\lim_{m\to\infty}t_m(x)$ exists and defines a homeomorphism
$t$ from $R(\bar r)$ to $R(\bar r')$. Let
us define a \emph{support} of a homeomorphism $h$ to be the set
$\sprt h=\{x\in\operatorname{dom} h\mid h(x)\ne x\}$. Now obviously for $m\ne
m'$, the supports of $h_{\pi(m)}$ and $h_{\pi(m')}$ are disjoint,
so the existence of the limit follows easily. In fact if $x\in
C_{n,k}$ for some $n,k\in\N$, then $t(x)=h_{(n,k)}(x)$ and $t(x)=x$
otherwise. Same argument leads that $t$ is bijective. Let
$(x,y,z)\in X$ and let us show that $t$ is continuous at
$(x,y,z)$. If $y\ne 1$ and $x\ne 1$, then $(x,y,z)$ has a
neighborhood intersecting only finitely many $C_{n,k}$, so $t$ is
determined by a finite composition of continuous functions in this
neighborhood. If $y=1$ and $x\notin \{1\}\cup\{1-2^{-n}\mid
n\in\N\}$, then the same holds again and also if vice versa: If
$x=1$ and $y\notin \{1\}\cup\{1-2^{-n}\mid n\in\N\}$. If $y=1$
and $x\in \{1-2^{-n}\mid n\in\N\}$, then $(x,y,z)\in X$ only if
$z\notin [0,1]$ (by the definition of $P$) and in this case
$(x,y,z)$ has again an open neighborhood intersecting only
finitely many $C_{n,k}$. If $x=1$ and $y\in
\{1\}\cup\{1-2^{-n}\mid n\in\N\}$, then every neighborhood
intersects infinitely many $C_{n,k}$. Let $n_*$ be such that for
all $n>n_*$ we have $r_{n}(k)=r'_{n}(k)$ which exists because
$\bar r$ and $\bar r'$ are $E_1$-equivalent and
let $U$ be a neighborhood of $(x,y,z)$ of radius
$2^{-2n_*}$. Then $U$ intersects only those $C_{n,k}$ for which
$n/2>n_*$ and so by the definition of $h_{n,k}$
it is identity on $C_{n,k}$ for all such $n$. Thus, $t_{m}$
is identity in $U$ for all $m$ and so $t$ is continuous.
Now we should check that the inverse is also continuous.
But with just a little care in the definition of $g_{n,k}$
we can assume that $g_{n,k}=g_{n,k}^{-1}$ and so
$t=t^{-1}$. Thus by symmetry, $t^{-1}$ is also continuous.
\end{proof}
\begin{claim}
Suppose $\bar r$ and
$\bar r'$ are not $E_1$-equivalent.
Then
$R(\bar r)$ and $R(\bar r')$
are not homeomorphic.
\end{claim}
\begin{proof}
Denote again $X=R(\bar r)$ and $X'=R(\bar r')$
and assume on contrary that there is a homeomorphism
$h\colon X\to X'$. Since $\bar r$ and
$\bar r'$ are not $E_1$-equivalent,
there is a sequence $(n_i,k_i)_{i\in\N}$ such that
$(n_i)_{i\in\N}$ is increasing and unbounded in $\N$
and for all $i$, $r_{n_i}(k_i)\ne r'_{n_i}(k_i)$.
Suppose first that $(k_i)_{i\in\N}$ is bounded in $\N$.
Then there exists a subsequence $(n_{i(j)},k_{i(j)})_{j\in\N}$
such that $k_{i(j)}=k_*$ for all $j$ for some fixed $k_*$.
By the construction each knot-type appears exactly once
in either of the sets
$$\{K^{\bar r}_{n,k,l}\mid (n,k,l)\in\N\times\N\times\{0,1\}\}$$
and
$$\{K^{\bar r'}_{n,k,l}\mid (n,k,l)\in\N\times\N\times\{0,1\}\}.$$
For each $m\in\N$ define the point $x_{m}$ as follows: Let $x_{m}$
be the point in $B_{m,k_*,0}\setminus K^{\bar r}_{m,k_*,0}$ given
by Lemma~\ref{lemma:FindPoints}. We know that if $m$ is odd, then
$K^{\bar r}_{m,k_*,0}$ has the same knot-type as
$K^{\bar r'}_{m,k_*,0}$ and if $m/2=n_{i(j)}$ for
some $j$, then $K^{\bar r}_{m,k_*,0}$ has the same
knot-type as $K^{\bar r'}_{m,k_*,1}$. Thus there are infinitely
many $m$ such that
$h(x_m)\in B_{m,k_*,0}$ and infinitely many $m$ such that
$h(x_m)\in B_{m,k_*,1}$. Thus, both points $(1,1-2^{-k_*},0)$ and
$(1,1-2^{-k_*},1)$ are accumulation points of
$(h(x_m))_{m\in\N}$. But only $(1,1-2^{-k_*},0)$ is an
accumulation point of $x_m$ which is a contradiction with
Lemma~\ref{lemma:CauchyComponent}, because both
$\{(1,1-2^{-k_*},0)\}$ and $\{(1,1-2^{-k_*},1)\}$
are connected components of both $\S^3\setminus X$
and $\S^3\setminus X'$.
Suppose now that $(k_i)_{i\in \N}$ is unbounded in $\N$.
Now pick a subsequence
$(n_{i(j)},k_{i(j)})_{j\in\N}$
such that not only $n_{i(j)}$ is strictly increasing,
but also $k_{i(j)}$ is.
For all $j$, let $x_{2j}$ be the point in
$B_{2n_{i(j)},k_{i(j)},0}$ given by
Lemma~\ref{lemma:FindPoints}. By similar argumentation as
above we know that $h(x_{2j})\in B_{2n_{i(j)},k_{i(j)},1}$.
Now again for all $j$,
define the point $x_{2j+1}$ to be a point in
$B_{2j+1,k_{i(j)},0}$ given again by Lemma~\ref{lemma:FindPoints}.
By the construction we know that $h(x_{2j+1})$ is in
$B_{2j+1,k_{i(j)},0}$ too. Thus $(x_m)_{m\in\N}$ is now a Cauchy
sequence converging to $(1,1,0)$ and $(h(x_m))_{m\in\N}$ is
sequence with two accumulation points $(1,1,0)$ and $(1,1,1)$.
The first of these points belongs to the connected component
$\{(1,1,t)\mid 0\le t<\frac{1}{2}\}$ of both $\S^3\setminus X$
and $\S^3\setminus X'$
(by the definition of $P$) and the second belongs
to the other connected component
$\{(1,1,t)\mid \frac{1}{2}<t\le 1\}$. Thus, we obtain
a contradiction with Lemma~\ref{lemma:CauchyComponent} again.
\end{proof}
\begin{claim}
$F$ is continuous.
\end{claim}
\begin{proof}
The inverse image of an $\varepsilon$-neighborhood of $R(\bar r)$
consists of all $\bar r$ which are mapped inside the
$\varepsilon$-collar of $R(\bar r)$ and in whose $\varepsilon$-collar
$R(\bar r)$ is contained. It is evident that only finitely
many of the knot-types are determined by the $\varepsilon$-collar,
since the $\varepsilon$-collar of $Q$ (or $P$) ``swallows'' all but finitely
many knots.
\end{proof}
\end{proof}
By Theorem~\ref{thm:KecLou} we have:
\begin{Cor}
The homeomorphism relation $\approx^{\{x\}}$,
is not Borel reducible to any orbit equivalence relation induced
by a Borel action of a Polish group. \qed
\end{Cor}
\begin{Cor}
There is a collection $D$ of Polish spaces homeomorphic to
subsets of $\S^3$ of the form
$V\cup \{x\}$ where $V$ is open and $x\in \R^3$ such that
all elements of $D$ have the same fundamental group, but
are not classifiable up to homeomorphism by any equivalence relation
arising from a Borel action of a Polish group.
\end{Cor}
\begin{proof}
The fundamental group of $R(\bar r)$ is the same
for all $\bar r$ -- the free product of the knot groups --
as can be witnessed by the Seifert-van Kampen theorem
by considering $R(\bar r)$ as the union of its open subsets
$A_{n}=\S^3\setminus (P\cup K_n\cup \Cup_{k\ne n}B_k)$
(here we fall back to the easier enumeration of the
balls by just one index used in
Definition~\ref{def:Properties}).
\end{proof}
\section{Positive Classification Results}
\label{sec:Other}
The results in this section are either known or follow easily from
what is known. We give some of the proofs for the sake of completeness.
Since the main result of the paper deals with $\sigma$-compact spaces, we begin
this section with the following relatively simple observation:
\begin{Thm}
The homeomorphism relation on any Borel collection of $\sigma$-compact spaces,
such as $K^{\{x\}}_{*}$ is~${\Sigma_1^1}$.
\end{Thm}
\begin{proof}
Two $\sigma$-compact spaces $X$ and $X'$ are homeomorphic if and only if there exist
sequences of compact sets $(C_n)_{n\in\N}$ and $(C_n')_{n\in\N}$
and homeomorphisms $h_n\colon C_n\to C_n'$ for each $n$ such that
\begin{enumerate}
\item $C_n\subset C_{n+1}$ and $C'_n\subset C'_{n+1}$ for all $n$,
\item $h_n\subset h_{n+1}$ for all $n$,
\item $X=\Cup_{n\in\N}C_n$ and $X'=\Cup_{n\in\N}C'_n$.
\end{enumerate}
To see this suppose $h\colon X\to X'$ is a homeomorphism. Since $X$ is $\sigma$-compact,
there is a sequence of compact sets $(C_n)_{n\in\N}$ which satisfies the first parts of
(1) and (3).
Let $C'_n=h[C_n]$. Then $(C_n')_{n\in\N}$ and $(h_n)_{n\in\N}$ where $h_n=h\!\restriction\! C_n$
satisfy all the rest.
On the other hand suppose that such sequences
$(C_n)_{n\in\N}$, $(C_n')_{n\in\N}$ and $(h_n)_{n\in\N}$ exist. Then obviously
$\Cup_{n\in\N}h_n$ is a homeomorphism $X\to X'$.
Consider the space $F(U)$ defined in Section~\ref{ssec:Par} parametrizing all Polish spaces.
Then, according to the above, the homeomorphism relation restricted to $\sigma$-compact spaces
can be defined by saying that $C$ and $C'$ are equivalent if there exist sequences
satisfying (1), (2) and (3). But these properties are all Borel properties. Also for $h_n$ to
be a homeomorphism is Borel, because the domains $C_n$ are compact.
This shows that the equivalence relation is~${\Sigma_1^1}$.
\end{proof}
Now we turn to compact and locally compact spaces.
\begin{Def}
Let $I^{\N}$ be the Hilbert cube. A set of \emph{infinite deficiency}
$A\subset I^{\N}$ is a closed set whose projection onto infinitely
many interval-coordinates is a singleton.
A closed set $A$ is a \emph{$Z$-set}, if for every open $U\subset I^\N$
with all homotopy groups vanishing, all the homotopy groups
of $U\setminus A$ vanish too.
\end{Def}
Anderson proved in \cite{Anderson} the following:
\begin{Thm}(\cite{Anderson})\label{Anderson}
\begin{enumerate}
\item If a set is of an infinite deficiency, then it is a $Z$-set.
\item Each homeomorphism between two closed $Z$-subsets of $I^\N$
can be extended to a homeomorphism of $I^\N$ onto itself. \qed
\end{enumerate}
\end{Thm}
From this it is not difficult to obtain the following theorem:
\begin{Thm}[Kechris-Solecki]\label{thm:KechSol}
The homeomorphism relation on compact Polish spaces is
continuously reducible to an orbit equivalence relation
induced by a Polish group action.
\end{Thm}
\begin{proof}
Let $h_1\colon I^\N\to I^\N$ be an embedding
defined as follows:
$$h_1((x_i)_{i\in\N})=(y_i)_{i\in\N}$$
where for all $n$, $y_{2n}=x_n$ and $y_{2n+1}=0$.
Let $X=\operatorname{ran}(h_1)$. Then $X$ is homeomorphic to $I^\N$.
Let $\approx^*_{c}$ be the equivalence relation on $K(I^\N)$
where two compact sets $C$ and $C'$
are equivalent if there exists
a homeomorphism $h\colon I^\N\to I^\N$ taking $C$ onto~$C'$.
By Fact~\ref{fact:HomPolish}
this equivalence relation is induced by a Polish group action,
and it is standard to verify that this action is continuous.
Let $\approx_{c}$ be the homeomorphism relation on $K(I^\N)$ where
$C$ and $C'$ are equivalent if they are homeomorphic.
Thus, it is sufficient to find a reduction of this into
$\approx^*_{c}$. For each $C\in K(I^\N)$ let $F(C)=h_1[C]$.
Now, if $C\approx_{c} C'$, then there is a homeomorphism
between $F(C)$ and $F(C')$. But these are of infinite deficiency,
so by Theorem~\ref{Anderson} there is a homeomorphism
$h\colon I^\N\to I^\N$ taking $F(C)$ onto $F(C')$ and
so $F(C)\approx^*_{c} F(C')$. If $C$ and $C'$ are not
homeomorphic, then so are not $F(C)$ and $F(C')$, so no such
homeomorphism can exist.
\end{proof}
Arguments along the same lines give us a stronger result:
\begin{Thm}\label{thm:locCom}
The homeomorphism relation on all locally compact Polish spaces
is Borel reducible to an orbit equivalence relation induced by
a continuous Polish group action.
\end{Thm}
\begin{proof}
Let ${\operatorname{Hom}}^{\{x\}}(I^\N)$
be the subgroup of ${\operatorname{Hom}}(I^\N)$ which consists of those
homeomorphisms $h$ such that $h(x)=x$. As a closed subgroup of
${\operatorname{Hom}}(I^\N)$ it is also Polish and acts continuously on $K^{\{x\}}(I^\N)$
(see Definition~\ref{def:StrangeSpaces}).
As shown in Section~\ref{ssec:Par}, the space of locally compact
spaces can be paramet\-rized as $K^{\{x\}}(I^\N)$ each locally compact space
being homeomorphic to $C\setminus \{x\}$ for some $C\in K^{\{x\}}(I^\N)$.
Applying the homeomorphism $h_1$ from the proof of
Theorem~\ref{thm:KechSol}, we may as well
assume that this $C$ is of infinite deficiency. Now, if the two spaces
$C\setminus \{x\}$ and $C'\setminus \{x\}$ are homeomorphic, the
homeomorphism extends to their one-point compactifications and thus to $x$.
Further, since $C$ and $C'$ are $Z$-sets, the
homeomorphism extends to an element of ${\operatorname{Hom}}^{\{x\}}(I^\N)$. On the
other hand, it is obvious that if $C\setminus \{x\}$ and
$C'\setminus \{x\}$ are not homeomorphic, then no such element of
${\operatorname{Hom}}^{\{x\}}(I^\N)$ can exist.
\end{proof}
By combining these results with Theorem~\ref{thm:NonClass}
we can conclude that ``not locally compact at one point''
is in a sense the
strongest requirement for Polish spaces
to be non-classifiable by such an orbit equivalence relation.
This is also reflected in the following Corollary:
\begin{Cor}\label{cor:locsig}
Then $\approx^{\{x\}}\ \not\le_B\ \approx_{loc}$. \qed
\end{Cor}
We would like to apply Theorem~\ref{thm:locCom} to the homeomorphism on
$n$-manifolds. It was discussed in Section~\ref{ssec:Par} that the set
$M_n$ of $n$-manifolds is Borel as a subset of the space of all Polish spaces
(in any of the parametrizations). As before, denote the homeomorphism
relation on $M_n$ by $\approx_n$.
Since manifolds are locally compact
the inclusion into the locally compact spaces is a reduction
$\approx_n\ \le_B\ \approx_{loc}$. By applying Theorem~\ref{thm:locCom}
we get the following:
\begin{Cor}
$\approx_n$ is reducible to an orbit equivalence relation induced
by a Polish group action. \qed
\end{Cor}
More is known in the case $n=2$. There is a classification
of $\approx_2$ by algebraic structures using cohomology groups
by Goldman~\cite{Goldman}.
It is probably routine to verify that this gives a Borel reduction into
the isomorphism on countable structures, but for now I leave it open
in the form of a conjecture:
\begin{Conj}\label{con:firstCon}
The classification in \cite{Goldman}
is a Borel reduction into $\cong_G$, thus
$\approx_2\ \le_B\ \cong_G$.
\end{Conj}
If the conjecture holds, we obtain a consequence which follows from Theorem~\ref{thm:graphstomanifolds}
below:
\begin{Conj}\label{conj:2to3}
$\approx_2\ \le_B\ \approx_3$.
\end{Conj}
The converse to Conjecture~\ref{con:firstCon}
is known to hold: $\cong_G\ \le_B\ \approx_2$. In fact
$\cong_G\ \le_B \ \approx_n$ for all $n\ge 2$.
We sketch two proofs of this fact -- one is based
on results by Camerlo and Gao and extension theorems from topology --
the other one, for $n=3$, is based on the methods used in this paper,
just to illustrate how these methods can be used.
\begin{Thm}\label{thm:graphstomanifolds}
For all $n\ge 2$ we have $\cong_G\ \le_B\ \approx_n$.
\end{Thm}
\begin{proof}[Sketch 1]
I would like to thank Clinton Conley who came up with this proof
at \verb|mathoverflow.net|.
It is sufficient to find a reduction into the homeomorphism relation
on open subsets of $\S^n$ or $\R^n$.
As shown in \cite{CamGao}, $\cong_G$
is Borel reducible to the homeomorphism relation on $K(2^\N)$.
On one hand it is known that every homeomorphism of a
totally disconnected compact
subset of the plane extends to the whole plane
(\cite[Ch. 13, Thm 7]{Moise}). On the other hand, by an application of
Lemma~\ref{lemma:CauchyComponent}, every homeomorphism of
$\R^2\setminus C$ where $C$ is compact and totally disconnected,
induces a homeomorphism of $C$. Thus, we can define a reduction from
the homeomorphism relation on $K(2^\N)$ to $\approx_2$: let
$f\colon 2^\N\to \R^2$ be the standard embedding (the Cantor set)
and with
$C\subset 2^\N$ associate the open set $\R^2\setminus f[C]$.
Of course these homeomorphisms extend
to $\R^n$ for every $n>2$ as well, so in fact
we have $\cong_G\ \le_B\ \approx_n$ for all~$n$.
\end{proof}
\begin{proof}[Sketch 2 (for $n=3$)]
Again, we consider only open subsets of $\R^3$.
It was proved by H. Friedman and L. Stanley in~\cite{FrSt}
that $\cong_G$ is reducible to
the isomorphism relation on countable linear orders.
Given a countable linear order $L$ with domain $\{x_n\mid n\in\N\}$,
construct first a set of disjoint open intervals $U_n\subset [0,1]$
such that $\Cup_{n\in\N}U_n$ is open and dense in $[0,1]$,
$\sup U_n\le \inf U_m$ if and only if $x_n<_Lx_m$ and if $x_m$ is an
immediate successor of $x_n$ then $\sup U_n=\inf U_m$. Then, by
considering $[0,1]$ as a subset of $\R^3$ in a canonical way,
replace each open interval with a copy of the chain depicted on
Figure~\ref{fig:VadimsSuperLink}. Let $C(L)$ be the closure of the
union of all these chains in~$\R^3$ By using methods similar to
those above, one can show that two linear orders $L$ and $L'$ are
isomorphic if and only if the complements of $C(L)$ and $C(L')$ are
homeomorphic.
\begin{figure}
\centering
\includegraphics[width=0.7\textwidth]{Singular_link_trefoil.mps}
\caption{The singular link.}
\label{fig:VadimsSuperLink}
\end{figure}
The idea is that the knot-types fix the orientation within the chain,
and the set $Q$ -- in this case,
the set $[0,1]\setminus \Cup_{n\in\N}U_n$ --
is totally disconnected and the homeomorphism of the complement
extends to it. Moreover it extends to it in an order preserving
way and also preserves end-points of the chains.
On the other hand
all these chains are similar to one another,
so any isomorphism of $L$
can be realized as a homeomorphism of the complement
of~$C(L)$.
\end{proof}
\section{Further Research}
\label{sec:Further}
Let $O_n(\S^n)$ be the space of all open subsets of $\S^n$ and
let $\approx^o_n$ be the homeomorphism relation on this space.
As before, $\approx_n$ is the homeomorphism relation on general
non-compact $n$-manifolds without boundary.
As before, let $\approx_{P}$ be the homeomorphism relation on all Polish,
spaces, $\approx_{loc}$ the one on locally compact,
$\approx_{c}$ the one on compact Polish spaces and $\approx^{\{x\}}$
as in Definition~\ref{def:Relations}.
An open-ended research direction is to establish the places of these
and other topological
equivalence relations in the hierarchy of analytic
equivalence relations. Positive and negative, new and old results
have been reviewed in this paper.
We already stated the main open question:
\begin{Open}
Is $\approx_3\ \le_B\ \cong_G$? If not, is it universal among
orbit equivalence relations induced by a Polish group action?
\end{Open}
And further one can ask:
\begin{Open}\label{q:whichnm}
For which $n$ and $m$ do we have $\approx_n\ \le_B\ \approx_m$?
The same for $\approx_n^o$ and $\approx_{m}^o$.
\end{Open}
\begin{Open}
For which $n\in\N$ and known equivalence relations $E$
do we have \mbox{$\approx_n\ \le_B E$} or $E\le_B\ \approx_n$
and same for $\approx^o_n$?
\end{Open}
\begin{Open}
What about open subsets of the separable Hilbert space~$\ell_2$?
By the results of Henderson~\cite{Hend} this covers all reasonable
concepts of infinite-dimensional manifolds.
\end{Open}
\begin{Open}[\cite{Gao}]\label{q:S12}
What is the exact complexity of $\approx_P$? It is known
that it is $\Sigma^1_2$ \cite{Gao} and ${\Sigma_1^1}$-hard~\cite{FerLouRos}.
\end{Open}
\begin{Open}
What are the precise locations of $\approx_{loc}$ and $\approx_{c}$
and in particular are they bireducible?
Hjorth has shown
using turbulence theory \cite{Hjorth} that $\cong_G\ <_B\ \approx_c$
(notice strict inequality) and by the results above, $\approx_c$ as
well as $\approx_{loc}$ are below the universal equivalence relation
induced by the Polish group action.
\end{Open}
The following question has been asked already in \cite{FaToTo2}:
\begin{Open}
Is $\approx_c$ universal among all equivalence relations that
are reducible to an orbit equivalence relation induced by a Polish
group action?
\end{Open}
And of course the conjectures from the end of the previous section:
\begin{Conj}\label{conj:2to3dubl}
The classification in \cite{Goldman}
is a Borel reduction into $\cong_G$, thus
$\approx_2\ \le_B\ \cong_G$.
In particular
$\approx_2\ \le_B\ \approx_3$.
\end{Conj}
Concerning Question~\ref{q:whichnm} and Conjectures~\ref{conj:2to3} and~\ref{conj:2to3dubl}:
at first it might seem that it holds that
$\approx_n\ \le_B\ \approx_{n+1}$.
However, the obvious candidate for a reduction
$M\mapsto M\times\R$ does not work: as shown in~\cite{McMillan} there
are open subsets $O$ of $\R^3$ which are not homeomorphic with $\R^3$,
yet $O\times \R\approx \R^4$. There are no such manifolds in
dimension $2$~\cite{Daverman}, but it is still unclear to the author
whether the general map $M\mapsto M\times\R$ from $2$- to
$3$-manifolds provides a reduction between the homeomorphism
relations.
\section{Conclusion}\label{sec:Conclusion}
As a conclusion we provide a diagram of all the relevant
equivalence relations and which relations are knownbetween them.
We omit some obvious arrows that follow e.g. from transitivity.
In the diagram we use the following notation:
\begin{eqnarray*}
\xymatrix{E\ar[r] & \ E'} & & E \le_B E',\\
\xymatrix{E\ar@{.>}[r]|? & E'} & & \text{Not known whether or not } E \le_B E',\\
\xymatrix{E\ar@{.>}[r]|C & E'} & & \text{Conjectured in this paper that } E \le_B E',\\
\xymatrix{E\ar@{-|}[r] &\ E'} & & E \not\le_B E',\\
E_1 && \text{Definition \ref{def:E1},}\\
E_0 && \text{Two sequences $\eta,\xi\in \N^\N$ are equivalent if $\exists n\forall (m>n)(\eta(m)=\xi(m))$,}\\
E_{Gr} && \text{The universal equivalence relation induced by a Borel Polish group action,}\\
E_{{\Sigma_1^1}} && \text{The universal ${\Sigma_1^1}$ equivalence relation,}\\
\approx^{\{x\}}&& \text{See Definition~\ref{def:Relations},}\\
\approx_{loc}&& \text{Homeomorphism on locally compact Polish spaces,}\\
\approx_c&& \text{Homeomorphism on compact Polish spaces,}\\
\approx_P&& \text{Homeomorphism on all Polish spaces,}\\
\approx_n && \text{Homeomorphism on $n$-manifolds.}\\
\cong_G && \text{Isomorphism on countable graphs.}
\end{eqnarray*}
\begin{figure}
\centering
$$\xymatrix{
& \approx_P \ar@<1ex>@{.>}[d]|?& \\
& E_{{\Sigma_1^1}}\ar[u] \ar@{.>}[dl]|? &\\
\approx^{\{x\}}\ar@<1ex>[ur] &\approx_{loc}\ar@{|->}[u]\ar@<1ex>@{.>}[d]|? \ar@{|->}[l]\ar@<1ex>[r]& E_{Gr}\ar@{|->}[ul]\ar@{.>}[l]|? \\
& \approx_c \ar[u] & \approx_n \ar@<1ex>@{.>}[dl]|?\ar[u]\ar[ul]\ar@<1ex>@{.>}[d]|?&\approx_m\ar@{<.>}[l]|?\\
E_1 \ar[uu]& \cong_G\ar@{|->}[u]\ar[ur]\ar@<-1ex>[r]& \approx_2\ar@{.>}[l]|C\ar@<.5ex>@{.>}[u]|C \ar@/^/[uul]\\
& E_0 \ar@{|->}[ul]\ar@{|->}[u]\ar@{|->}[ur]&
}
$$
\caption{Diagram of reducibility. Here $n>2$ and $m\ne n$.}
\label{fig:Diag}
\end{figure}
\newpage
|
\section{Overview}
Almost 50 years since its theoretical inception,
the Higgs boson has been discovered at the LHC{\cite{higgs}}. Nonetheless, the Higgs boson remains a mystery, and its discovery has unlocked many questions
about its nature that are related to its special role in the Universe.
Now that the Higgs boson mass is known, the Standard Model (SM) predicts its interactions and properties with no free parameters. Any deviation
from these predictions provides unambiguous evidence for new physics, making a rigorous study of the Higgs a focus of upcoming operations at the LHC, as well as at
future colliders. This quest to determine the properties of the Higgs goes hand in hand with direct searches for new physics at the LHC. In particular, it is
crucial to understand how the two modes of exploration are intertwined. In this paper, we examine this connection within the framework of the
p(henomenological) MSSM{\cite{Djouadi:1998di}}.
The pMSSM provides an excellent structure for a systematic and comprehensive survey of constraints on Supersymmetry (SUSY) and for the investigation of complementary
approaches to detecting its existence. Towards this end, we have previously embarked on a detailed study of signatures
for the pMSSM at the 7, 8 and 14 TeV LHC~{\cite {us1,us2,usnew}} and have compared the LHC search reach to that of searches for dark matter via
direct and indirect detection~\cite{Cahill-Rowley:2013dpa}. Our focus on Supersymmetry stems from its attractiveness as a candidate for new physics.
Its presence at the weak-scale would stabilize the Higgs sector under quantum corrections, provide a natural thermal dark matter candidate, and accommodate
unification of the gauge couplings.
The pMSSM is the most general version of the R-parity conserving MSSM when it is subjected to a minimal set
of experimentally-motivated guiding principles: ($i$) CP conservation, ($ii$) Minimal Flavor Violation at the electroweak scale so that flavor
physics is controlled by the CKM mixing matrix and the Yukawa couplings of the SM fermions, ($iii$) degenerate 1\textsuperscript{st} and
2\textsuperscript{nd} generation sfermion masses. In addition, it is assumed that ($iv$) the Yukawa couplings and A-terms for the first two
generations can be safely neglected. In particular, theoretical assumptions about physics at high scales, {\it e.g.}, the nature of SUSY
breaking, are absent in order to capture electroweak scale phenomenology for which a UV-complete theory may not yet exist. Imposing these principles
decreases the number of free parameters in the MSSM at the TeV-scale from 105 to 19 for the case of a neutralino LSP. If the gravitino
mass is included so that it plays the role of the LSP, an additional parameter is required.
With respect to the production of new physics at an accelerator, a key question is whether its signature can be detected
given our understanding of the backgrounds arising from SM processes (provided the new particles are kinematically accessible). In
particular, it is important to determine how experimental analyses can probe the full parameter space of
interest within any specific model. This is certainly true in the case of Supersymmetry. However, even in the simplest SUSY scenario, the MSSM, the number of free
parameters ($\sim$ 100) is much too large to study in complete generality. A traditional approach is to assume the existence of a UV-complete
theory with minimal set of parameters (such as mSUGRA{\cite {SUSYrefs}) from which the properties of the sparticles at the TeV scale can be determined and studied in
detail. While such an approach is often quite valuable~\cite{Cohen:2013kna}, these scenarios can be phenomenologically
limiting and many are under increasing tension with a wide range of experimental data, including the $\sim 126$ GeV mass of the Higgs. At the
opposite end of the spectrum, simplified model scenarios can be employed to estimate constraints from the LHC, thereby bounding the model parameter space in a
process-by-process fashion. However, a concern in this case is that the simplified models are not capturing the `big picture' of what is occurring in the full
underlying theory. The more general pMSSM circumvents the limitations of these other approaches.
The increased dimensionality of the parameter space not only allows for a somewhat less prejudiced study of SUSY,
but also yields valuable information on unusual scenarios, identifies weaknesses in the LHC analyses, and can be used to combine
results from many individual and independent SUSY related searches.
To study the pMSSM, we generate large sets of models by randomly scanning the parameter space.
The 19/20 parameters and the ranges of values that we employ in our scans are listed in Table~\ref{ScanRanges}. In order to sample the pMSSM space as
thoroughly as possible, we generate many millions of model points (using SOFTSUSY{\cite{Allanach:2001kg}} and checking for consistency with
SuSpect{\cite{Djouadi:2002ze}}), with each point corresponding to a specific set of values for the parameters. We then subject these individual `models'
to a global set of collider, flavor, precision, dark matter and theoretical constraints~\cite{us1}. In particular,
we do not assume that the LSP relic density necessarily saturates the WMAP/Planck value{\cite{omega}}, $\Omega h^2 \simeq
0.12$, in order to allow for the possibility of multi-component DM. (For example, the axions introduced to solve the strong CP problem may contribute significantly to the DM relic density.)
Roughly $\sim$225k models for each type of LSP survive this initial selection and can be used for further physics studies. Decay patterns of the
SUSY partners and the extended Higgs sector are calculated using a privately modified version of SUSY-HIT{\cite{Djouadi:2006bz}} as well as the
most recent version of HDECAY{\cite {HDECAY}}. Since our scan ranges include sparticle masses up to 4 TeV, an upper limit chosen to enable phenomenological studies at the 14 TeV LHC, the majority of neutralinos and charginos are nearly pure electroweak eigenstates. This is due to the off-diagonal elements of the corresponding mass matrices
being at most $M_W$. This has important
implications for the resulting collider and DM phenomenology\cite{us1,us2,usnew,Cahill-Rowley:2013dpa}. We note that both of these model sets were generated before the Higgs boson was discovered. For the neutralino (gravitino) model set
we find that roughly $\simeq 20 (10)\%$ of the models are found to satisfy $m_h=126\,\pm\,3$ GeV; clearly, we will focus on these subsets in the analyses that
follow.
In addition to these two large pMSSM model sets, we have also generated a third, somewhat smaller, specialized set of
`natural' models with the neutralino being identified as the LSP. These models predict $m_h=126\,\pm\,3$ GeV, have an LSP that {\it does} saturate the
WMAP/Planck relic density, and yield values
of fine-tuning (FT) better than $1\%$ employing the traditional Ellis-Barbieri-Giudice measure~\cite{ebg}. This low-FT model
set will also be included as part of the present study.
In order to produce this model
set, we modified the parameter scan ranges as indicated in Table~\ref{ScanRanges} to greatly increase the likelihood of achieving both low FT and
a thermal relic density in the desired range. In addition to these modified scan ranges, we also required $|M_1/\mu|<1.2$ and $|X_t|/m_{\tilde t} >1$, where
$X_t=A_t-\mu\cot\beta$ quantifies the mixing between the stop-squarks with $m_{\tilde t}$ being the geometric mean of the tree-level stop masses.
Amongst other things, this requires a
bino-like LSP, light Higgsinos and highly mixed stops. We generated $\sim 3.3 \times 10^8$ low-FT points in this 19-dimensional parameter
space and required consistency with current precision, flavor, DM and collider constraints as before.
Due to the difficulty of satisfying this set of requirements, only $\sim$ 10.2k low-FT models were found to be viable for further study.
Within each pMSSM model, the characteristics of the lightest CP-even Higgs, $h$, as well as the entire superpartner spectrum, are calculable (to several loops) from
the chosen values of the soft-breaking parameters in the underlying Lagrangian.
Given this correspondence, we can address the connection between the predicted SUSY Higgs properties and the direct searches for SUSY at the LHC. In particular,
we seek to address two questions: ($i$)
How will potentially null searches for SUSY at the LHC influence the predicted properties of the Higgs boson? ($ii$) What can be learned about the properties of the superpartners
from precision measurements of the Higgs Boson couplings?
In what follows, we briefly discuss the impact on our model sets of the 7, 8, direct SUSY searches at the LHC,
as well as the expectations for 14 TeV searches in Section 2. In Section 3 we examine the predictions of the properties of the lightest Higgs in the pMSSM.
We discuss the impact of measurements of the Higgs properties on the pMSSM from current data and from future measurements at the 14 TeV LHC with 300 fb$^{-1}$ and 3 ab$^{-1}$,
as well as the proposed International Linear Collider (ILC), in Section 4. Our conclusions are given in Section 5.
\begin{table}
\centering
\begin{tabular}{|c|c|c|} \hline\hline
Parameter & General Neutralio/Gravitino Set & Low Fine-Tuned Set \\
\hline\hline
$m_{\tilde L(e)_{1/2,3}}$ & $100 \,{\rm GeV} - 4 \,{\rm TeV}$ & 100\,{\rm GeV} - 4\,{\rm TeV} \\
$m_{\tilde Q(u,d)_{1/2}}$ & $400 \,{\rm GeV} - 4 \,{\rm TeV}$ & 100\,{\rm GeV} - 4\,{\rm TeV} \\
$m_{\tilde Q(u,d)_{3}}$ & $200 \,{\rm GeV} - 4 \,{\rm TeV}$ & 100\,{\rm GeV} - 4\,{\rm TeV}\\
$|M_1|$ & $50 \,{\rm GeV} - 4 \,{\rm TeV}$ & 25\,{\rm GeV} - 552\,{\rm GeV} \\
$|M_2|$ & $100 \,{\rm GeV} - 4 \,{\rm TeV}$ & 100\,{\rm GeV} - 2.1\,{\rm TeV} \\
$|\mu|$ & $100 \,{\rm GeV} - 4 \,{\rm TeV}$ & 100\,{\rm GeV} - 460\,{\rm GeV} \\
$M_3$ & $400 \,{\rm GeV} - 4 \,{\rm TeV}$ & 400\,{\rm GeV} - 4\,{\rm TeV} \\
$|A_{t,b,\tau}|$ & $0 \,{\rm GeV} - 4 \,{\rm TeV}$ & 0\,{\rm GeV} - 2.3\,{\rm TeV} ($A_t$ only) \\
$M_A$ & $100 \,{\rm GeV} - 4 \,{\rm TeV}$ & 100\,{\rm GeV} - 4\,{\rm TeV}\\
$\tan \beta$ & $1 - 60$ & 1 - 60 \\
$m_{3/2}$ & 1 eV$ - 1 \,{\rm TeV}$ ($\tilde{G}$ LSP) & - \\
\hline\hline
\end{tabular}
\caption{Scan ranges for the 19 (20) parameters of the pMSSM with a neutralino (gravitino) LSP. The gravitino mass is scanned with
a log prior. All other parameters are scanned with flat priors; we expect this choice to have little qualitative impact on
our results for observables~\cite{Djouadi:1998di}.}
\label{ScanRanges}
\end{table}
\section{LHC SUSY Searches}
To begin this study, we first ascertain which models in each of our three sets are excluded at the 7,8 TeV LHC, and which can be probed at 14 TeV.
Once these current constraints and future expectations for the pMSSM parameter space are characterized, we can determine how the properties of
the lightest SUSY Higgs boson are affected by the direct searches, and quantify how they may differ from SM predictions. Such correlations between
the direct search results and the properties of the Higgs can address the questions posed above.
We begin this step of the analysis with a brief overview of our procedure for computing the effects of the LHC direct SUSY searches on the pMSSM.
In general, we replicate the suite of ATLAS SUSY search analyses as closely as possible employing fast Monte Carlo. We also include
several searches performed by CMS. The specific analyses applied to our pMSSM model sets are briefly summarized in
Tables~\ref{SearchList7} and~\ref{SearchList8}. We augment the standard MET-based SUSY channels by including searches for heavy stable charged particles and
a heavy neutral SUSY Higgs decaying into $\tau^+\tau^-$ as performed by CMS~\cite{CMSextra}, as well as measurements of the rare decay mode $B_s\rightarrow \mu^+\mu^-$
as discovered by CMS and LHCb~\cite{BSMUMU}. All of these play distinct and important roles in covering the pMSSM parameter space. Details of our
analysis and results are discussed at length in our previous work~\cite{us1,us2,usnew}, with the most recent description of
our final results for 7 and 8 TeV and expectations for 14 TeV given in \cite{usnew}. Here, we provide a concise summary of the salient features of this work
in order to provide a basis for investigating the properties of the Higgs.
\begin{table}
\centering
\begin{tabular}{|l|l|c|c|c|} \hline\hline
Search & Reference & Neutralino & Gravitino & Low-FT \\
\hline
2-6 jets & ATLAS-CONF-2012-033 & 21.2\% & 17.4\% & 36.5\% \\
multijets & ATLAS-CONF-2012-037 & 1.6\% & 2.1\% & 10.6\% \\
1 lepton & ATLAS-CONF-2012-041 & 3.2\% & 5.3\% & 18.7\% \\
HSCP & 1205.0272 & 4.0\% & 17.4\% & $<$0.1\% \\
Disappearing Track & ATLAS-CONF-2012-111 & 2.6\% & 1.2\% & $<$0.1\% \\
Muon + Displaced Vertex & 1210.7451 & - & 0.5\% & - \\
Displaced Dilepton & 1211.2472 & - & 0.8\% & - \\
Gluino $\rightarrow$ Stop/Sbottom & 1207.4686 & 4.9\% & 3.5\% & 21.2\% \\
Very Light Stop & ATLAS-CONF-2012-059 & $<$0.1\% & $<$0.1\% & 0.1\% \\
Medium Stop & ATLAS-CONF-2012-071 & 0.3\% & 5.1\% & 2.1\% \\
Heavy Stop (0$\ell$) & 1208.1447 & 3.7\% & 3.0\% & 17.0\% \\
Heavy Stop (1$\ell$) & 1208.2590 & 2.0\% & 2.2\% & 12.6\% \\
GMSB Direct Stop & 1204.6736 & $<$0.1\% & $<$0.1\% & 0.7\% \\
Direct Sbottom & ATLAS-CONF-2012-106 & 2.5\% & 2.3\% & 5.1\% \\
3 leptons & ATLAS-CONF-2012-108 & 1.1\% & 6.1\% & 17.6\% \\
1-2 leptons & 1208.4688 & 4.1\% & 8.2\% & 21.0\% \\
Direct slepton/gaugino (2$\ell$) & 1208.2884 & 0.1\% & 1.2\% & 0.8\% \\
Direct gaugino (3$\ell$) & 1208.3144 & 0.4\% & 5.4\% & 7.5\% \\
4 leptons & 1210.4457 & 0.7\% & 6.3\% & 14.8\% \\
1 lepton + many jets & ATLAS-CONF-2012-140 & 1.3\% & 2.0\% & 11.7\% \\
1 lepton + $\gamma$ & ATLAS-CONF-2012-144 & $<$0.1\% & 1.6\% & $<$0.1\% \\
$\gamma$ + b & 1211.1167 & $<$0.1\% & 2.3\% & $<$0.1\% \\
$\gamma \gamma $ + MET & 1209.0753 & $<$0.1\% & 5.4\% & $<$0.1\% \\
$B_s \rightarrow \mu \mu$ & 1211.2674 & 0.8\% & 3.1\% & * \\
$A/H \rightarrow \tau \tau$ & CMS-PAS-HIG-12-050 & 1.6\% & $<$0.1\% & * \\
\hline\hline
\end{tabular}
\caption{7 TeV LHC searches included in the present analysis, and the corresponding fraction of the neutralino, gravitino and low-FT pMSSM
model sets excluded by each channel. Note that in the case of the last two rows the experimental constraints have already been included
in the model generation process for the low-FT model set.}
\label{SearchList7}
\end{table}
\begin{table}
\centering
\begin{tabular}{|l|l|c|c|c|} \hline\hline
Search & Reference & Neutralino & Gravitino & Low-FT \\
\hline
2-6 jets & ATLAS-CONF-2012-109 & 26.7\% & 22.5\% & 44.9\% \\
multijets & ATLAS-CONF-2012-103 & 3.3\% & 5.6\% & 20.9\% \\
1 lepton & ATLAS-CONF-2012-104 & 3.3\% & 6.0\% & 20.9\% \\
SS dileptons & ATLAS-CONF-2012-105 & 4.9\% & 12.5\% & 35.5\% \\
2-6 jets & ATLAS-CONF-2013-047 & 38.0\% & 31.1\% & 56.5\% \\
HSCP & 1305.0491 & - & 23.0\% & - \\
Medium Stop (2$\ell$) & ATLAS-CONF-2012-167 & 0.6\% & 8.1\% & 4.9\% \\
Medium/Heavy Stop (1$\ell$) & ATLAS-CONF-2012-166 & 3.8\% & 4.5\% & 21.0\% \\
Direct Sbottom (2b) & ATLAS-CONF-2012-165 & 6.2\% & 5.1\% & 12.1\% \\
3rd Generation Squarks (3b) & ATLAS-CONF-2012-145 & 10.8\% & 9.9\% & 40.8\% \\
3rd Generation Squarks (3$\ell$) & ATLAS-CONF-2012-151 & 1.9\% & 9.2\% & 26.5\% \\
3 leptons & ATLAS-CONF-2012-154 & 1.4\% & 8.8\% &32.3\% \\
4 leptons & ATLAS-CONF-2012-153 & 3.0\% & 13.2\% & 46.9\% \\
Z + jets + MET & ATLAS-CONF-2012-152 & 0.3\% & 1.4\% &6.8\% \\
\hline\hline
\end{tabular}
\caption{Same as in the previous table but now for the 8 TeV ATLAS MET-based SUSY searches. Note that when all the channels from this table and the previous table are
combined, we find that $\sim 45.5~(61.3,~74.0)\%$ of these models are excluded by the LHC for the neutralino
(gravitino, low-FT) model set.}
\label{SearchList8}
\end{table}
Briefly stated, our procedure is as follows: We generate SUSY events for each model for all relevant (up to 85) production channels with PYTHIA
6.4.26~\cite{Sjostrand:2006za}, and then pass the events through fast detector simulation using PGS 4~\cite{PGS}. Both programs have been modified to,
{\it e.g.}, correctly deal with gravitinos, multi-body decays, hadronization of stable colored sparticles, and ATLAS b-tagging. We then scale our event rates to NLO by computing the relevant K-factors using Prospino 2.1~\cite{Beenakker:1996ch}. The
individual searches are then implemented using our customized analysis code{\cite {us}}, which follows the published experimental cuts and selection criteria as
closely as possible. This analysis code is validated for each of the many search regions for every channel, employing the benchmark model points
provided by ATLAS (and CMS). Models are then excluded using the 95\% $CL_s$ limits as employed by ATLAS (and CMS). For the purpose of obtaining
the direct SUSY search results on the two large model sets, we
perform this analysis {\it without} requiring the Higgs mass constraint, $m_h=126\,\pm\,3$ GeV (combined experimental and theoretical errors) so that we
can understand its influence on the search results. Recall that roughly $\sim 20 (10)\%$ of models in the neutralino(gravitino) model set predict a Higgs mass in
this range. While we observe some variation amongst the individual searches, we find that once the channels are combined, the overall pMSSM model
coverage is to an excellent approximation {\it independent} of the value of the Higgs mass
\cite{usnew}. Conversely, the fraction of neutralino and gravitino LSP models predicting the observed Higgs mass is also found to be approximately
independent of whether or not the direct SUSY search results have been enforced. This result is very powerful and demonstrates the approximate decoupling of the direct
SUSY search results from the mass of the Higgs boson. Of course, for this study, in which we specifically examine the properties of the
Higgs boson itself, we restrict our investigation to the subset of the neutralino and gravitino LSP model samples that predict $m_h=126\,\pm\,3$ GeV. No additional
requirements on the Higgs mass are necessary for the low-FT set, since in this case the Higgs mass constraint is imposed during the model generation
process.
Tables~\ref{SearchList7} and~\ref{SearchList8} also show
the coverage of our pMSSM model sets from the 7 and 8 TeV search constraints. We find that $\sim 45.5 (61.3, 74.0)\%$
of the neutralino (gravitino, low-FT) model samples are excluded by the LHC. In particular, we find that numerous models with light squarks
and gluinos (500-1000 GeV) are currently viable. These results demonstrate that much phase-space is left to
accommodate natural Supersymmetry.
In addition to the searches performed at 7 and 8 TeV, future LHC operations at $\sim 14$ TeV will greatly extend the coverage of the pMSSM parameter space.
For our 14 TeV analysis, we considered the impact of two of the most powerful searches to be performed by ATLAS~\cite{atlas14}, namely the zero-lepton jets +MET
and the zero- and one-lepton stop channels. We have simulated these channels~\cite{usnew} in a manner identical to that described above
for the 7 and 8 TeV searches. We have extrapolated the results expected by ATLAS at 300 fb$^{-1}$ of integrated luminosity to 3 ab$^{-1}$ by scaling
the required signal rate. Due to the large CPU required to generate events at these luminosities, we restricted our study to the subset of models that
remain viable after the 7,8 TeV constraints and predict the observed Higgs mass.
We find that with 300 (3000) fb$^{-1}$ of data, the combination of these searches covers 90.83\% (97.15\%) of the
neutralino LSP model set, 83.22\% (93.29\%) for the gravitino LSP model set,
and 97.69\% (100\%) of the low-FT model sample.
Clearly, the 14 TeV LHC will provide a more definitive statement on the existence of natural Supersymmetry, even in complex forms such as the pMSSM,
and the discovery space of the upcoming run is significant.
These results of the direct LHC SUSY searches will be employed below in our study of the Higgs couplings.
\section{Determination of Higgs Properties}
In this section, we show how the pMSSM parameter space can be constrained by
the measured properties of the Higgs. For this analysis, we must first
determine the extent that the couplings of the light CP-even Higgs boson in the pMSSM differ from the expectations for the SM Higgs, and then we can compare these results
to the current and expected future experimental determinations of the couplings. We make several such comparisons corresponding to the anticipated evolution
of our knowledge about the allowed values of the Higgs couplings: ($i$) current data{\cite {now}}, ($ii$) measurements that are expected to be attainable
at the 14 TeV LHC with an integrated luminosity of 0.3(3) ab$^{-1}${\cite {14TeVLHC}}, and finally ($iii$) projected measurements at the ILC with two different run plans being
250 \unit{fb^{-1}}\ at 250 GeV plus 250 \unit{fb^{-1}}\ at 500 GeV, as well as an upgrade to 1150 \unit{fb^{-1}}\ at 250 GeV plus 1600 \unit{fb^{-1}}\ at 500 GeV plus 2500 \unit{fb^{-1}}\ at 1000 GeV center of mass
energy~{\cite {ILC}}.
To calculate the Higgs couplings in the pMSSM and in the SM, we employ HDECAY 5.11.
We note that since the full set of computed SUSY loop corrections for the $h \rightarrow WW$ and $h\rightarrow ZZ$ partial
widths are not yet incorporated in HDECAY, we unfortunately can not employ these very important modes to constrain our pMSSM model sample. We follow the standard approach, using the narrow width approximation (NWA) and defining the signal strength for a given production channel (e.g. $gg,VBF\rightarrow h$), with the subsequent decay into the
final state, $h \rightarrow X$, normalized to the corresponding SM value, as
\begin{equation}
\mu_{gg,VBF}(X) = {{\sigma(gg,VV\rightarrow h)~B(h \rightarrow X)}\over {SM}}\,.
\end{equation}
For final states that do not involve the top quark, we can also define the ratio of the squares of the couplings to
their corresponding SM values by simply forming the ratio of the relevant partial decay widths,
\begin{equation}
r_X = {{\Gamma (h \rightarrow X)}\over {SM}}\,,
\end{equation}
for the final states $X=ZZ,~W^+W^-,~\bar b b, ~\bar c c, ~\tau^+\tau^-,~gg, ~\gamma\gamma, ~\gamma Z$. The case of the $ht\bar t$ coupling must be handled
separately and can only be directly accessed via associated $t\bar th$ production. We are, of course, also interested in the branching fraction
for Higgs decays into the lightest neutralino{\footnote {The LEP limits of $\sim$100 GeV on the mass of charged sparticles, which we apply strictly, constrain the possible invisible decay modes of the Higgs. We note that neutral winos, Higgsinos and sneutrinos are required to have a charged partner with a similar mass, thus preventing them from being decay products of the Higgs.}},
producing a final state which is purely invisible or accessed by jets+MET, depending on the production channel.
Searches for invisible decays into the LSP are very interesting because of their potential to place significant constraints on
the SUSY parameter space, particularly when results from ILC500 are employed, as we shall see below.
\begin{figure}[htbp]
\centerline{\includegraphics[width=3.5in]{muggfgamgam_log.png}
\hspace{-0.50cm}
\includegraphics[width=3.5in]{muvbfgamgam_log.png}}
\vspace*{0.50cm}
\centerline{\includegraphics[width=3.5in]{rtotal_log.png}}
\vspace*{-0.10cm}
\caption{Histograms of signal strengths for $h \rightarrow \gamma \gamma$ in the $gg$-fusion (top left) and vector boson fusion (top right) production channels for the subset
of neutralino models that predict $m_h=\,126\pm\,3$ GeV. The blue (red) histogram represents models before any ATLAS searches (after the 7 and 8
TeV SUSY searches) are applied, while the green (purple) histograms show models that are expected to survive the zero-lepton jets plus MET plus the 0,1-$\ell$ stop searches
at 14 TeV, assuming a luminosity of 300 (3000) fb$^{-1}$. The ratio $r_{total}$ for the total width of the Higgs is analogously shown in the bottom panel.}
\label{figA}
\end{figure}
To get an initial understanding of the distribution of Higgs properties in the various pMSSM model sets, it is instructive to first study a few examples. Figure~\ref{figA},
shows the distribution of the $h\rightarrow \gamma\gamma$ signal strength for both the $gg$-fusion and vector boson fusion production channels in the neutralino
LSP model set (with $m_h=126\,\pm\,3$ GeV), along with the effect of the current 7/8 TeV and future 14 TeV ATLAS searches on this
distribution~\cite{usnew} as indicated. Other than the obvious fact that these distributions peak near unity but have long tails,
the most important observation is that the shape of these distributions (up to statistical fluctuations) is essentially unaffected by the
imposition of the ATLAS direct SUSY searches. Furthermore, the shape of the distribution for the ration of total widths, $r_{total}= \Gamma(h\rightarrow All)/SM$, for the neutralino models
demonstrates that this shape invariance is maintained for the other observables. We therefore see that SUSY searches and Higgs boson properties are to a very good
approximation `orthogonal'. As we will show below, the other final states exhibit a similar behavior, answering our first question above:
Future null direct SUSY searches at the LHC will, to a good approximation (as is seen here except for statistical limitations),
not significantly modify the range of values that we expect for the SUSY Higgs
couplings.
\begin{figure}[htbp]
\centerline{\includegraphics[width=5.5in]{rgamgam_log_lines.png}}
\vspace*{0.50cm}
\centerline{\includegraphics[width=3.5in]{rgamgam_g_log_lines.png}
\hspace{-0.50cm}
\includegraphics[width=3.5in]{rgamgam_lft20_log_lines.png}}
\vspace*{-0.10cm}
\caption{Histograms of the ratio of partial widths for $h \rightarrow \gamma \gamma$ for the subset of neutralino (top), gravitino (lower left) and low-FT models (lower
right) that predict $m_h=126\,\pm\,3$ GeV. The blue (red) histogram represents models before any ATLAS searches (after the 7 and 8 TeV SUSY searches) are applied
while the green (purple) histograms show models that are expected to survive both the zero-lepton jets plus MET and the 0,1-$\ell$ stop searches at 14 TeV, assuming
an integrated luminosity of 300 (3000) fb$^{-1}$. The vertical lines show the expected future limits on $r_{\gamma \gamma}$, and are discussed in the text.}
\label{figB}
\end{figure}
\begin{figure}[htbp]
\centerline{\includegraphics[width=5.5in]{rgg_log_lines.png}}
\vspace*{0.50cm}
\centerline{\includegraphics[width=3.5in]{rgg_g_log_lines.png}
\hspace{-0.50cm}
\includegraphics[width=3.5in]{rgg_lft20_log_lines.png}}
\vspace*{-0.10cm}
\caption{Same as the previous Figure but now for $h\rightarrow gg$.}
\label{figC}
\end{figure}
\begin{figure}[htbp]
\centerline{\includegraphics[width=5.5in]{rbb_log_lines.png}}
\vspace*{0.50cm}
\centerline{\includegraphics[width=3.5in]{rbb_g_log_lines.png}
\hspace{-0.50cm}
\includegraphics[width=3.5in]{rbb_lft20_log_lines.png}}
\vspace*{-0.10cm}
\caption{Same as in Figure~\ref{figB} but now for $h\rightarrow b \bar b$.}
\label{figD}
\end{figure}
\begin{figure}[htbp]
\centerline{\includegraphics[width=5.5in]{rtautau_log_lines.png}}
\vspace*{0.50cm}
\centerline{\includegraphics[width=3.5in]{rtautau_g_log_lines.png}
\hspace{-0.50cm}
\includegraphics[width=3.5in]{rtautau_lft20_log_lines.png}}
\vspace*{-0.10cm}
\caption{Same as in Figure~\ref{figB} but now for $h\rightarrow \tau^+\tau^-$.}
\label{figE}
\end{figure}
We now turn our attention to the predicted distributions for the values of the various partial width distributions, $r_X$, in each model set, and the
effect of the future LHC direct SUSY searches on these distributions. We will return to these distributions later in our subsequent analysis to understand the effects
of the future Higgs coupling measurements.
Figure~\ref{figB} shows a histogram for the ratio $r_{\gamma\gamma}$ in the three different model sets. The
vertical lines appearing in these plots are discussed in detail in the next section, and represent the anticipated Higgs coupling measurement sensitivities provided by future measurements
at the 14 TeV LHC and ILC500 as discussed above and indicated in the figure. Qualitatively, we see that the effect of the LHC direct
searches on the $r_{\gamma\gamma}$ distributions is to decrease the normalization
while preserving the overall shapes of the distributions for all three model sets. Deviations from this general behavior are mainly seen in the tails of the
distributions, where the statistics are low. The different responses of each model set to the direct LHC SUSY searches can be seen by observing the
differing impacts of the searches on the distribution areas. Interestingly, the $r_{\gamma\gamma}$ distribution in the neutralino model
set has a very different shape compared with the corresponding distribution of diphoton signal strengths shown in Figure~\ref{figA}, which are coupled to the production
channels; this difference results mainly from
large corrections to the $h \rightarrow b \bar{b}$ and $h\rightarrow gg$ partial widths (which will be discussed below), and therefore to the total width. These corrections alter the
diphoton branching fraction, and therefore the signal strength, for a given value of $r_{\gamma\gamma}$. Note also that the distributions of
$r_{\gamma\gamma}$ in the neutralino and gravitino model sets are rather similar yet somewhat distinct from the corresponding distribution in the low-FT model
set, which exhibits a broader range of values for $r_{\gamma\gamma}$ despite the lower statistics. This larger spread in the low-FT distribution
arises from the mandatory presence of light charginos, stops, and (in many cases) sbottoms, typically resulting in larger SUSY corrections to the effective
$h \gamma\gamma$ coupling than in the large neutralino and gravitino model sets, in which charged sparticles are not required to be relatively light. Finally,
note that in all three model sets the value of $r_{\gamma\gamma}$ peaks at the roughly same value, slightly above unity. We will see below that this shift
is reasonably anticorrelated with a corresponding shift in the peak of the $r_{gg}$ distribution (as well as with the $ht\bar{t}$ coupling in the low-FT set).
Both offsets generally result from the large stop mixing that is necessary to obtain the correct value of the Higgs mass.
Figure~\ref{figC} displays analogous histograms of the ratio $r_{gg}$, showing the distribution for each pMSSM model set. Once again, we see that
the neutralino and gravitino distributions are quite similar while the low-FT distribution differs as a result of distinct requirements on the sparticle
spectra. Note that all three distributions peak below unity.
As shown in, {\it e.g.}, \cite{Carena:2013iba}, the large Higgs mass generally requires large stop mixing in Supersymmetry,
which results in a small ($\sim 5\%$) but important reduction in the $h \rightarrow gg$ partial width and a simultaneous, but somewhat smaller, enhancement in the $h \rightarrow \gamma\gamma$ partial width. This is a consequence of the non-decoupling nature of SUSY corrections to the Higgs sector. If the stop
sector radiative corrections were totally responsible for this deviation (which is a reasonable approximation in many cases), then the shift in $r_{gg}$ at the amplitude level would be
$\sim 3$ times larger than the corresponding change in $r_{\gamma\gamma}$, with the two displacements having opposite signs. As a result of this effect, essentially all
of our models predict $r_{gg}$ to be below unity; this observation will figure prominently in our subsequent discussion of future experimental constraints
on the Higgs couplings. Interestingly, we also see that the tails of the $r_{gg}$ distribution are not very large for the neutralino and gravitino pMSSM model sets.
The tails are slightly smaller in
the low-FT $r_{gg}$ distribution, since the relevant corrections tend to be larger as a result of the bias towards light stops. Since
the stops are playing an important role, we would expect corresponding shifts in the magnitude of the $ht\bar{t}$ coupling; as we will see below,
this is indeed the case.
Figure~\ref{figD} displays the results for the ratio $r_{bb}$ for the three pMSSM model sets, with the neutralino and gravitino distributions again being similar, yet
somewhat different from the low-FT scenario. Small differences between the neutralino and gravitino distributions arise from several reasons, but namely from the fact that
lighter stops/sbottoms can appear in the gravitino set, since the requirement for the stop to be heavier than the LSP is trivially satisfied when $m_{LSP} \sim 0$
as in most of the gravitino LSP models. For each pMSSM model set we see the now-familiar pattern in which the LHC direct SUSY searches do not significantly alter the shapes
of the partial width distributions. Unlike the previous cases, however, we now see that $r_{bb}$ may deviate from unity by a significant $O(1)$ factor. These
deviations result from large sbottom mixing that can make corresponding $O(1)$ changes in the $hb\bar b$ couplings through non-decoupling (mostly gluino)
loop effects. These loop effects are driven by the size of the off-diagonal element of the sbottom mass matrix, {\it i.e.}, $m_b(A_b-\mu \tan \beta)$, which is
enhanced for large values of $\tan \beta$. While the tails of this distribution mostly extend to larger values of $r_{bb}$, we see that models also exist
with $r_{bb}$ being significantly below unity. Since the $b\bar b$ mode dominates the Higgs width, the large variations in $r_{bb}$ also explain the large spread in the
distribution of $r_{total}$, presented in Fig.~\ref{figA}. In our neutralino and gravitino parameter scans,
$|A_b|$ and $|\mu|$ are typically of a similar size while $\tan \beta$ has typical values that are $\mathcal{O}(10)$, so the $\mu \tan \beta$ term
in the off-diagonal element dominates. However, in the low-FT set this is no longer applicable since the allowed size of $|\mu|$ (and therefore the sbottom mixing) is significantly reduced by naturalness requirements. Thus in the low-FT scenario we expect a considerably smaller range of values for $r_{bb}$, which agrees with the distributions shown in Fig.~\ref{figD}.
Figure~\ref{figE} displays the analogous results for the ratio $r_{\tau\tau}$ for the three different pMSSM model sets. Here we again see that the shapes of the
$r_{\tau\tau}$ histograms are not significantly altered by the ATLAS direct SUSY searches at this level of statistics. We also see that the peak occurs at a ratio
value that is slightly greater than unity (by $\sim 2\%$) with a significant tail extending to larger values. This is not surprising since there are also
non-decoupling effects in the corrections to the $h\tau \tau$ vertex. However, these corrections occur via electroweakino loops and are proportional to the $\tau$ mass.
This implies that the effect of these non-decoupling terms should be relatively small when compared with their corresponding effect in the ratio $r_{bb}$,
and that is indeed what we observe. Again, since this non-decoupling occurs via the off-diagonal $m_\tau (A_\tau -\mu \tan \beta)$ term in the stau mass
matrix, these effects should be somewhat suppressed in the low-FT model set in comparison to the other pMSSM model sets, and this is demonstrated in
Figure~\ref{figE}.
\begin{figure}[htbp]
\centerline{\includegraphics[width=4.5in]{rbb_b1.png}}
\vspace{-0.10cm}
\centerline{\includegraphics[width=4.5in]{rbb_b1_14tev.png}}
\vspace*{-0.10cm}
\caption{Values of the ratio $r_{bb}$ as a function of the lightest sbottom mass for the neutralino model set incorporating the influence of the ATLAS direct SUSY searches.
The lower panel shows those models probed by the searches at 14 TeV. }
\label{figF}
\end{figure}
\begin{figure}[htbp]
\centerline{\includegraphics[width=4.5in]{rbb_b1_g.png}}
\vspace{-0.10cm}
\centerline{\includegraphics[width=4.5in]{rbb_b1_14tev_g.png}}
\vspace*{-0.10cm}
\caption{Same as the previous figure but now for the gravitino model set. }
\label{figF2}
\end{figure}
Figure~\ref{figF} shows the dependence of the ratio $r_{bb}$ on the lighter sbottom mass for the neutralino LSP model set with the effects of the direct LHC SUSY
searches being imposed. Interestingly, measuring a value of this ratio near unity will not impose a constraint on the sbottom mass, regardless of the precision
of the measurement. On the other hand, very large deviations of this ratio from unity are seen to require a relatively light sbottom mass, meaning that
null SUSY search results should be able to reduce the expected range for $r_{bb}$. However, the non-decoupling nature of the corrections means that values
of $r_{bb}$ above 2 are predicted, even after the 14 TeV direct SUSY searches are included. Excluding O(1) deviations from $r_{bb} = 1$ (which can occur
for sbottoms as heavy as 2.5 TeV) through direct SUSY searches is clearly not feasible. The large sbottom mass direct search reach necessary to constrain
$r_{bb}$ significantly explains our earlier observation that this distribution is roughly independent of results from the LHC direct searches.
Figure~\ref{figF2} shows that the corresponding results for the gravitino model sample are qualitatively similar, although they differ in detail due to the improved reach of direct sparticle searches
in the gravitino set. Figure~\ref{figFP} shows the analogous results for the low-FT model set. As discussed above, the decreased range of $r_{bb}$ values in the low-FT model set arises from the requirement that $|\mu|$ is relatively small, decreasing the size of the off-diagonal
element in the sbottom mass matrix and therefore the corrections to $r_{bb}$.
\begin{figure}[htbp]
\centerline{\includegraphics[width=4.5in]{rbb_b1_lft20.png}}
\vspace{-0.10cm}
\centerline{\includegraphics[width=4.5in]{rbb_b1_lft20_14tev.png}}
\vspace*{-0.10cm}
\caption{Same as the previous Figure but now for the low-FT model set.}
\label{figFP}
\end{figure}
If the lightest neutralino is sufficiently light, then the Higgs can decay to neutralino pairs, being observed as an invisible decay mode of the Higgs.
The top panel in Fig.~\ref{figG} displays the branching fraction,
$B(h\rightarrow \chi\chi)$, as a function of the LSP mass for the few
neutralino LSP models where this channel
is kinematically allowed, and also indicates the influence of the direct SUSY searches at the LHC. Note that all of these models have values
of $B(h\rightarrow \chi\chi) <0.5$, meaning that they remain allowed by the current LHC constraints on invisible Higgs decays. Since these models are mostly bino-Higgsino
admixtures (to satisfy the WMAP/Planck relic density upper bound) and the coupling to the Higgs is proportional to the product of the bino and Higgsino content of the
neutralino, the branching fractions are seen to fall rapidly as the neutralino mass increases. This is due not only to a reduction in phase space, but also to a
decline in the neutralino Higgsino content as its mass increases. We note that all of these models will eventually be excluded (or discovered)
by sparticle searches, as well as by searches for Higgs $\rightarrow$ invisible at the 14 TeV LHC and/or ILC500. The lower left panel shows the corresponding results
for the gravitino set with a neutralino NLSP; here we see that a much smaller branching fraction is obtained since the WMAP/Planck constraint
does not apply to the neutralino NLSP. Of course for these gravitino pMSSM models the lightest neutralinos will only produce an invisible final state if they escape the
detector before decaying. Neutralinos with $c \tau \lesssim 1$ m will have visible decays, generally producing a (possibly displaced) diphoton + MET signature,
where the diphotons would of course fail to reconstruct the Higgs mass. However, the stability of the neutralino tends to be unimportant, since (with the possible
exception of the models with very light neutralinos) the $h\rightarrow \chi \chi$ branching fraction is far too small to be accessible at the 14 TeV LHC.
The bottom right panel displays the same distribution for the low-FT model set. Here we see that the additional constraints imposed on the pMSSM spectrum
during the model generation yield numerous light LSPs that are mainly bino-Higgsino admixtures, a sizable fraction of which pair-annihilate via the Z/Higgs funnel. Note that
these fall into two distinct branches, depending on the sign of the parameter $\mu$. In all cases, however, the invisible branching fraction is found to be
below $\sim 30-50\%$, which will eventually be accessible at the 14 TeV LHC. While many of these models are now excluded by LHC direct SUSY searches, the remainder would
be probed by the corresponding 14 TeV direct searches.
\begin{figure}[htbp]
\centerline{\includegraphics[width=5.5in]{invbr_log.png}}
\vspace*{0.50cm}
\centerline{\includegraphics[width=3.5in]{invbr_g_log.png}
\hspace{-0.50cm}
\includegraphics[width=3.5in]{invbr_lft20_log.png}}
\vspace*{-0.10cm}
\caption{Branching fraction of the invisible Higgs decay $h \rightarrow \tilde{\chi}_1^0 \tilde{\chi_1}^0$ for neutralino LSP models (top) with the observed Higgs mass. The points
are color-coded according to their coverage by the LHC direct SUSY searches. The analogous results for the gravitino (bottom left) in the case of a neutralino
NLSP, and for the low-FT (bottom right) model sets are also shown.}
\label{figG}
\end{figure}
As a final observable, we briefly consider the ratio $r_{tt}$, defined as the squared value of the $ht\bar{t}$ coupling normalized to its SM value. We calculate this quantity using the expressions given in
Ref.~\cite{Draper:2013oza}. The predicted values of this ratio, computed for each of the various model sets before the application of constraints from the
SUSY direct searches, are of some interest for future measurements at both the 14 TeV LHC and at the ILC. They are displayed in Figure~\ref{figH}.
Here we see that the deviation from the SM expectation is always less than $\sim 10\%$, which is below the anticipated sensitivity of both LHC14
and ILC500. However, the 1 TeV upgrade of the ILC should eventually be able to determine this quantity at the level of a few percent{\cite {ILC}}.
The different behavior of the histograms for the three model sets is easily understood when we recall that the deviations from unity are driven mostly by the
non-decoupling effects of the stop masses and, particularly, by the mixing in the stop sector that are controlled by the values of the parameters $A_t, \mu$ and
$\tan \beta$ via the quantity $X_t \sim A_t-\mu /\tan \beta$. If there
is no strong preference for the sign of $A_t$ and/or $\mu$ arising from the model generation procedure this distribution will be approximately
symmetric around the SM value; this is observed for the neutralino set. We note that while $A_t$ is, in fact, sign symmetric, the corresponding
distribution of the values of $\mu$ is found to be somewhat asymmetric in sign when $|\mu|$ is large. However, most of the models in this set have
relatively small values of $|\mu|$ (as Higgsino LSPs are common) so that the resulting distribution remains essentially symmetric, as shown in the Figure.
For the gravitino set, these same conditions hold except that Higgsino LSPs are somewhat less common and the values of $|\mu|$ tend to be correspondingly larger,
thus $r_{tt}$ is now more sensitive to this sign asymmetry in the $\mu$ distribution. Hence, for the gravitino set, we see a somewhat asymmetric distribution for $r_{tt}$.
The low-FT model set displays a different behavior, as here $|\mu|$ must be small and we simultaneously require both the observed
value of the Higgs mass and also less than than $1\%$ values of fine-tuning. This selects a specific sign for the stop mixing as well as a hierarchical
stop spectrum. This pushes $r_{tt}$ to somewhat larger deviations from the SM, on average, than in the other two model sets, with a strong preference towards increasing the $ht\bar{t}$ coupling with respect to its Standard Model value.
\begin{figure}[htbp]
\centerline{\includegraphics[width=6.5in]{top_histos.png}}
\vspace*{-0.10cm}
\caption{Histograms of the predicted values of the ratio $r_{tt}$, defined in the text, for the various pMSSM model
sets: neutralino (red), gravitino (green) and low-FT (blue).}
\label{figH}
\end{figure}
\section{Analysis and Results}
\label{sec:step3}
Now that we have assembled the necessary ingredients, we can determine how the future measurements of the various Higgs couplings at the LHC and ILC will
restrict the pMSSM parameter space, and compare these constraints with those from the direct SUSY searches at the LHC. In this analysis, we use the
numerical results for the current and expected future Higgs coupling measurements at the LHC and ILC
presented in Refs.~\cite{now,14TeVLHC,ILC}. We note that important higher-order corrections to the Higgs couplings have yet to be computed, and these
may be quite relevant compared to the claimed level of precision for the future collider measurements. We are thus unfortunately forced to ignore these potentially
significant theoretical uncertainties in quoting allowed ranges for the ratios of Higgs couplings in the pMSSM to those in the SM.
We remind the reader to keep this important issue in mind when interpreting our results, and note they should be treated as indicative only. Clearly,
more theoretical work will be necessary before (sub-)percent-level measurements are truly meaningful.
We also caution the reader that in obtaining the results shown below, we have necessarily made an assumption about the central value of the
future Higgs coupling measurements at the LHC and ILC.
Namely, we have assumed that the central values will coincide {\it exactly} with those predicted by the SM, {\it i.e.}, we take the measured central value to be $r_{X}=1$
for all couplings. As we will see from the discussion that follows, the observation of Higgs couplings not centered around the SM prediction
(even within the expected ranges) would probe a different fraction of our
model sets. This is particularly true for the case of the couplings generated at loop-level, $r_{gg}$ and $r_{\gamma\gamma}$, where the pMSSM predictions deviate from the SM values
essentially all in one direction. Of course, our qualitative results, which indicate that precise Higgs coupling measurements (when properly
understood) have significant sensitivity to the pMSSM, do not depend on the actual central values that will be observed for these couplings.
Assuming that the future measured central value of each parameter is equal to its SM prediction, we return to Figs.~\ref{figB},~\ref{figC},~\ref{figD} and~\ref{figE}
(as well as~\ref{figG}) and now concentrate on the vertical lines, which display the expected sensitivity arising from future experiments. These show the regions of the
various $r_X$ that will be allowed or excluded at the $95\%$ CL by Higgs coupling measurements at the LHC and HL-LHC{\cite {14TeVLHC}},
the 500 GeV ILC (ILC500) and the ILC500 with a luminosity upgrade{\cite {ILC}}, here denoted as HL-ILC500. Of course, it is important to once again note that
these future expected allowed regions can always be shifted, allowing for the estimation of implications of other possible experimental outcomes.
We first notice that current LHC data on the Higgs couplings does not significantly constrain the pMSSM parameter space, since the precision of the Higgs measurements
is still rather low in comparison to the deviations expected in the pMSSM. Once 14 TeV LHC, as well as ILC, data is available, this will no longer
be the case and the measurements will begin to probe pMSSM effects as their accuracy improves.
However, the key result here is that, regardless
of what central values are actually observed, {\it indirect Higgs coupling measurements will likely result in the exclusion (or the discovery) of pMSSM models
that are not accessible to the direct SUSY searches at the LHC}. An important caveat to this, of course, is that we need to include the full suite of 14 TeV direct SUSY
searches before this result is truly robust. However, given our 7 and 8 TeV studies~\cite{usnew}, the 0-$\ell$ jets plus MET search when combined
with the 0,1-$\ell$ stop searches will result in powerful parameter space coverage at 14 TeV, and so this qualitative conclusion is unlikely to change. This result
is found to hold for all of the model sets.
Taking these results at face value, we can extract some relevant numbers directly from these Figures. We can now determine
what fraction of the presently allowed pMSSM models, {\it i.e.}, those passing the 7 and 8 TeV ATLAS direct search analyses (with $m_h=126\,\pm\,3$ GeV),
will be indirectly probed by future measurements of the Higgs couplings. Next, we can ascertain how
these results will be modified by the 14 TeV LHC direct SUSY searches. Our results are presented in the set of
Tables~\ref{T1},~\ref{T2} and~\ref{T3} for both the 14 TeV LHC and the ILC.
In these Tables we see a number of important results: ($i$) at the LHC, constraining the $hb\bar b$ coupling yields the
strongest bounds on the allowed pMSSM parameter space. This measurement can be greatly improved at the ILC, which has the potential to yield exquisite precision on this coupling.
($ii$) However, given our assumption that the measured central values exactly correspond to the SM predictions, we see that the ILC determination of the $hgg$ coupling probes much,
if not all, of the remaining pMSSM parameter space. The reason for this is clear: Since $r_{gg}$ is forced to be less than unity by the non-decoupling effects associated with
the large stop mixing required
to generate the observed value of the Higgs mass, a determination of $r_{gg}=1$ with a very small error will probe essentially all of the model sets! If, on the other
hand, the central value were measured to be, say, only $\sim 2-3\%$ below unity, a very much smaller fraction of models would then be probed. For example,
if the central value of $r_{gg}$ were measured to be 0.97 with the same expected errors, then we find that this measurement is only sensitive to $2.7\%$ of the
neutralino LSP model set at the ILC500, so that $hb\bar b$ would remain the dominant constraint in this case. This
specific example demonstrates the sensitivity of our results to the assumption that the measured central values will always agree with the SM predictions. In any case,
($iii$) we see that both the LHC and ILC will provide very powerful probes of the pMSSM model space and have the potential to observe the effects of at least some of
the models that would otherwise remain viable, being missed by the 14 TeV direct SUSY searches. In particular, the precision attainable in Higgs coupling measurements at the ILC
will deeply probe the pMSSM parameter space.
Tables~\ref{T1}-\ref{T3} also show that ($iv$) although the general shapes of the $r_X$ distributions are somewhat similar, they differ in detail so
that the three pMSSM model sets will respond distinctly to constraints from the various indirect Higgs coupling measurements. Of course, the ILC500
is extremely powerful in all three cases. The last thing we notice is that ($v$) the entries in the Tables will not vary greatly as we include more channels
from future direct SUSY searches at the 14 TeV LHC. This is not surprising; in the limit that the shapes of the $r_X$ distributions are completely unaffected by the
SUSY search results, the Table entries should be essentially independent of which LHC searches have been applied. The limited statistical size of our model samples, and
the small changes in the $r_X$ distribution shapes, account for the observed variations.
\begin{table}
\begin{center}
\begin{tabular}{|c||c|c|c|c|}
\hline
Channel & 300 fb$^{-1}$ LHC & 3 ab$^{-1}$ LHC & 500 GeV ILC & HL 500 GeV ILC \\
\hline
\hline
$b\bar{b}$ & 16.6 (27.7, 0.5) & 33.4 (48.5, 5.5) & 78.4 (88.8, 49.1) & 91.1 (95.8, 77.3) \\
$\tau \tau$ & 0.7 (0.8, 2.9) & 3.1 (2.7, 5.7) & 11.5 (9.9, 11.9) & 36.9 (34.2, 32.9) \\
$gg$ & 0.02 (0.04, 0.5) & 0.5 (0.6, 3.1) & 99.4 (99.7, 99.7)& 100.0 (100.0, 100.0) \\
$\gamma \gamma$ & 0.02 (0.07, 0) & 0.02 (0.09, 0.2) & 0.02 (0.07, 0) & 0.1 (0.2, 0.6) \\
Invisible & 0 (0, 0) & 0 (0, 0) & 0.01 (0.01, 6.2) & 0.02 (0.01, 7.5) \\
\hline
\hline
All & 17.1 (28.2, 3.8) & 34.9 (49.6, 11.1) & 99.8 (99.96, 99.92) & 100.0 (100.0, 100.0) \\
\hline
\end{tabular}
\caption{The fraction in percent of the neutralino (gravitino, low-FT) model sets (with the correct Higgs mass), which remain viable after the current 7 and 8 TeV LHC searches,
that can be probed by future Higgs coupling measurements, {\it assuming} that the SM values for these couplings are observed.}
\label{T1}
\end{center}
\end{table}
\begin{table}
\begin{center}
\begin{tabular}{|c||c|c|c|c|}
\hline
Channel & 300 fb$^{-1}$ LHC & 3 ab$^{-1}$ LHC & 500 GeV ILC & HL 500 GeV ILC \\
\hline
\hline
$b\bar{b}$ & 20.5 (31.7, 0) & 39.1 (53.0, 5.4) & 82.6 (92.6, 46.4) & 93.1 (97.5, 75.0) \\
$\tau \tau$ & 0.5 (0.7, 1.8) & 3.3 (2.3, 1.8) & 12.9 (9.9, 5.4) & 38.9 (32.6, 23.2) \\
$gg$ & 0 (0, 0) & 0.09 (0.1, 0) & 99.9 (99.93, 100.0) & 100.0 (100.0, 100.0) \\
$\gamma \gamma$ & 0 (0, 0) & 0 (0, 0) & 0 (0, 0) & 0 (0, 0) \\
Invisible & 0 (0, 0) & 0 (0, 0) & 0 (0, 10.7) & 0 (0, 16.1) \\
\hline
\hline
All & 20.8 (31.9, 1.8) & 40.6 (53.7, 5.4) & 99.91 (100.0, 100.0) & 100.0 (100.0, 100.0) \\
\hline
\end{tabular}
\caption{Same as Table~\ref{T1} above, but now for the subset of models expected to remain viable after the ATLAS 14 TeV 0l jets + MET and 0l and 1l stop searches with 300 fb$^{-1}$ of
integrated luminosity.}
\label{T2}
\end{center}
\end{table}
\begin{table}
\begin{center}
\begin{tabular}{|c||c|c|c|c|}
\hline
Channel & 300 fb$^{-1}$ LHC & 3 ab$^{-1}$ LHC & 500 GeV ILC & HL 500 GeV ILC \\
\hline
\hline
$b\bar{b}$ & 19.6 (32.6, ---) & 38.4 (54.5, ---) & 82.9 (94.9, ---) & 93.4 (98.4, ---) \\
$\tau \tau$ & 0.7 (0.7, ---) & 3.3 (2.5, ---) & 14.7 (10.7, ---) & 41.6 (35.3, ---) \\
$gg$ & 0 (0,---) & 0 (0, ---) & 100.0 (100.0, ---) & 100.0 (100.0, ---) \\
$\gamma \gamma$ & 0 (0, ---) & 0 (0, ---) & 0 (0, ---) & 0 (0, ---) \\
Invisible & 0 (0, ---) & 0 (0, ---) & 0 (0, ---) & 0 (0, ---) \\
\hline
\hline
All & 29.9 (32.8, ---) & 39.3 (55.4, ---) & 100.0 (100.0, ---) & 100.0 (100.0, ---) \\
\hline
\end{tabular}
\caption{Same as Table~\ref{T1} above but now for the subset of models expected to remain viable after the ATLAS 0l jets + MET and 0l and 1l stop searches with 3 ab$^{-1}$ of
integrated luminosity. The entries for the low-FT set in this table are blank because no models survive the 3 ab$^{-1}$ LHC direct searches.}
\label{T3}
\end{center}
\end{table}
Lastly, we summarize our results in the $M_A - \tan\beta$ plane by combining the effects of anticipated Higgs coupling measurements in the $\gamma\gamma\,, \tau\tau\,, b\bar b$ channels
for the neutralino LSP model set.
We exclude the $hgg$ coupling from this analysis due to the complications and resulting shift in the central value of this parameter arising from the large stop mixing as discussed above.
Figure ~\ref{newplot}
shows the fraction of models in the large neutralino LSP sample that are probed in a particular bin by the anticipated measurements of these three channels
at the LHC and ILC500 and their luminosity upgrades. The fraction is color-coded, indicating the pMSSM coverage within a bin, ranging from 100\% (black)
coverage to 0\% (dark blue). The white curves represent the results from current heavy Higgs searches with decays into $\tau$ pairs~\cite{CMSextra}. Here, we clearly see the
effects of increasing precision for the Higgs coupling measurements, and the value of the anticipated ultra-precise determinations to be available at the ILC500, in covering the pMSSM
parameter space. We note that the Higgs coupling measurements cover a region of parameter space that is somewhat orthogonal to that of the heavy Higgs searches. Namely,
the coupling determinations probe essentially vertical slices of this plane, and most importantly, catch the low $M_A$, $\tan\beta$ region that is missed by the direct searches. This
demonstrates the complementarity of the direct and indirect approaches.
\begin{figure}[htbp]
\centerline{\includegraphics[width=3.5in]{matb_ybtau_LHC14.png}
\hspace{-0.50cm}
\includegraphics[width=3.5in]{matb_ybtau_HL-LHC14.png}}
\vspace*{0.50cm}
\centerline{\includegraphics[width=3.5in]{matb_ybtau_ILC500.png}
\hspace{-0.50cm}
\includegraphics[width=3.5in]{matb_ybtau_HL-ILC500.png}}
\vspace*{-0.10cm}
\caption{Coverage of the pMSSM parameter space for the neutralino model set in the $M_A - \tan\beta$ plane,
showing the fraction of models probed in each bin by the anticipated
sensitivity to the combined $\gamma\gamma$, $\tau\tau$, and $b\bar b$ Higgs couplings at various colliders as indicated. The white curves correspond to the
present limits from the direct searches for $H/A\rightarrow\tau\tau$.}
\label{newplot}
\end{figure}
\section{Conclusion}
In this paper we have examined SUSY signals and Higgs boson properties within the context of the pMSSM for models with either neutralino or gravitino LSPs as
well as for neutralino models with low FT that saturate the WMAP/Planck relic density measurement. Within this general scenario we then addressed the following
questions: `What will potentially null searches for SUSY at the LHC tell us about the possible properties of the Higgs boson?' and, conversely, `What do the
precision measurements of the couplings of the Higgs tell us about the possible properties of the various superpartners?' We again note that in
obtaining the results presented here we have ignored any theoretical errors associated with the as-yet to be computed corrections for the Higgs coupling ratios,
and employed the current version of the corrections as implemented in HDECAY.
Our results can be further refined once a better understanding of this uncertainty is provided by future theoretical work.
We saw in the above discussion that the answer to the first question was rather straightforward: Given an initial distribution of signal strengths $\mu_X$ or
branching fraction ratios $r_X$ for a specific final state, the LHC direct SUSY searches reduce the size of the distribution but to a very good approximation
do not change its {\it shape}. This was shown to be true for all three of the model sets we consider. This implies that to first order the direct (null) SUSY
searches at the LHC will not impact the range of possible deviations of Higgs branching fractions from their SM values. This is a very powerful result.
However, we found the answer to the second question to be much more complex and of potentially of even greater importance: Precision measurements of Higgs
couplings and branching fractions can and do lead to the exclusion of pMSSM models which cannot be probed by the powerful 14 TeV LHC direct SUSY searches, even
with an integrated luminosity of 3 ab$^{-1}$. This is true for both gravitino and neutralino model sets and also true whether or not the precise values of
the measured quantities are consistent with the SM expectation. Of course, the more precisely the Higgs couplings are measured, the greater the fraction of
the pMSSM models that can be probed. Since the $hb\bar b$ coupling can deviate the furthest from its SM value within the pMSSM framework, measurements of
its value generally will have the greatest impact {\it if} we do not assume that the central values measured for the Higgs couplings are given exactly by
their SM values. If this is indeed the case, however, then the $hgg$ coupling at the ILC will provide the strongest constraint as this quantity is necessarily
shifted in the pMSSM by stop loops with a central value crudely determined by the requirement of obtaining the observed Higgs mass. In such a case (or if the
observed central values for $r_{gg}$ -- or to a lesser extent $r_{\gamma\gamma}$ -- differ from the SM in the opposite direction from the pMSSM prediction),
essentially all of the pMSSM parameter space considered here would then be excluded.
Lastly,
we compared the reach of the Higgs coupling determinations to the direct heavy Higgs searches in the $M_A - \tan\beta$ plane
and show that they cover orthogonal regions.
Our analysis
demonstrates the complementarity of the direct and indirect approaches in searching for Supersymmetry, and the importance of precision studies of the properties of the
Higgs Boson.
\section{Acknowledgments}
We wish to thank M.~Spira for answering our many questions and for assistance with the implementation of HDECAY. This work was
supported by the Department of Energy, Contracts DE-AC02-06CH11357, DE-AC02-76SF00515, and DE-FG02-12ER41811.
|
\section{Introduction}
\label{sec:Introduction}
The nonperturbative solution of quantum field theories is
considerably more difficult than perturbative calculations.
Various approaches have been developed, with lattice theory~\cite{lattice}
being the most popular. The Dyson--Schwinger approach~\cite{DSE}
has also had some success. An alternative is a Hamiltonian
approach based on light-front quantization~\cite{DLCQreviews,Vary,ArbGauge},
which has the advantage of providing boost-invariant wave functions;
such a formulation is much more intuitive, and, of the three, it is
the only one formulated in Minkowski space.
Light-front quantization is done in terms of Dirac's light-front
coordinates~\cite{Dirac}, where $x^+\equiv t+z$ is time and the
corresponding spatial coordinate is $x^-\equiv t-z$. The transverse
coordinates $x$ and $y$ remain as they were. The conjugate variables
of light-front energy and momentum are $p^-\equiv E-p_z$ and $p^+\equiv E+p_z$,
respectively. Again, the transverse components $\vec p_\perp=(p_x,p_y)$
are unchanged. The mass-shell condition $p^2=m^2$ then implies
$p^-=(m^2+p_\perp^2)/p^+$. However, here we will be concerned with a
two-dimensional theory, and the transverse components do not enter.
The two-dimensional light-front Hamiltonian eigenvalue problem is~\cite{DLCQreviews},
\begin{equation}
{\cal P}^-|\psi(P^+)\rangle=\frac{M^2}{P^+}|\psi(P^+)\rangle \;\;
\mbox{and} \;\;
{\cal P}^+|\psi(P^+)\rangle=P^+|\psi(P^+)\rangle,
\end{equation}
with ${\cal P}^-$ and ${\cal P}^+$ the light-front energy and momentum operators.
The second equation is automatically satisfied by expanding
the eigenstate in Fock states that are themselves eigenstates
of ${\cal P}^+$. For $n$ bosons with individual momenta $p_i^+$,
we have the total momentum $\sum_i p_i^+=P^+$
and the Fock states $|p_i^+;P^+,n\rangle$. The eigenstate
expansion is then
\begin{equation} \label{eq:fockexp}
|\psi(P^+)\rangle=\sum_n (P^+)^{(n-1)/2}
\int \left(\prod_{i=1}^{n-1} dx_i\right) \psi_n(x_1,...,x_n) |x_iP^+;P^+,n\rangle,
\end{equation}
with $\psi_n$ the $n$-boson wave function. The factor $(P^+)^{(n-1)/2}$
is explicit in order that $\psi_n$ be independent of $P^+$.
The Hamiltonian eigenvalue problem now becomes an infinite set of
integral equations for the Fock-state wave functions. The
standard approximation made is to truncate the expansion to a
finite number of terms, yielding a finite set of equations.
In many theories, such a truncation causes difficulties with
respect to regularization and renormalization, because the
truncation in particle number removes infinite contributions that
would ordinarily cancel against contributions that are retained.
This is the nonperturbative analog of separating a Feynman diagram
into time-ordered diagrams and throwing away diagrams that have
more than some specified number of intermediate particles;
the remaining approximation to the Feynman diagram will
generally be divergent, even if the original covariant diagram
was finite.
One proposed resolution of this difficulty is sector-dependent
renormalizaton~\cite{Wilson,hb,Karmanov}, where infinities
caused by truncation are absorbed by allowing the bare masses and
couplings to depend on the Fock sectors involved. However, this
can lead to ill-defined wave functions~\cite{SecDep}.
The light-front coupled-cluster (LFCC) method~\cite{LFCC} avoids this
by not truncating Fock space. Instead, the eigenstate
is built as $\sqrt{Z}e^T|\phi\rangle$, where $\sqrt{Z}$ is
a normalizing factor, $|\phi\rangle$ is a valence state with
the smallest possible number of constituents, and $T$ is an
operator that increases particle number. The truncation
made is a truncation of $T$; the exponentiation then
builds all higher Fock states, giving them wave functions
that depend on the valence wave functions and on the functions
in the truncated $T$.
The Hamiltonian eigenvalue problem becomes a valence equation
\begin{equation} \label{eq:valence}
P_v\ob{{\cal P}^-}|\phi\rangle=\frac{M^2+P_\perp^2}{P^+}|\phi\rangle,
\end{equation}
with $\ob{{\cal P}^-}=e^{-T} {\cal P}^- e^T$ the effective LFCC
Hamiltonian and $P_v$ the projection onto the valence Fock sector.
This is supplemented by an auxiliary equation for $T$
\begin{equation} \label{eq:aux}
(1-P_v)\ob{{\cal P}^-}|\phi\rangle=0.
\end{equation}
Without truncation, these equations provide an exact
representation of the original eigenvalue problem.
When $T$ is truncated, the projection $1-P_v$ must
also be truncated, to retain only enough equations
to solve for the functions retained in $T$. Of course,
the valence eigenvalue problem and the auxiliary equation
must be solved simultaneously. One immediate advantage
is that the physical mass of the eigenstate appears in
the kinetic-energy terms of the effective Hamiltonian,
without use of sector-dependent renormalization. The
price to be paid is that the auxiliary equation is
nonlinear in the functions that determine $T$.
When this approach is applied to a simple soluble model~\cite{GreenbergSchweber},
one finds that the simplest truncation of $T$ is sufficient
to generate the exact answer analytically~\cite{LFCC}.
As a first realistic application, where a truncation is an approximation
and where numerical methods are required, we consider here the
odd-parity sector of light-front $\phi^4$ theory in two
dimensions~\cite{VaryHari}. The valence state is
the one-particle state, and the $e^T$ operator generates all
the higher odd-particle-number Fock states. The $T$ operator
is truncated to a single term that creates two additional particles.
By itself, this would only couple the three-particle state, but
the exponentiation of $T$ still generates all the higher odd
states.
The truncated $T$ operator depends on a single function of two
relative momenta. The auxiliary equation (\ref{eq:aux}) is restricted to
a projection onto the three-particle state, to provide an
equation for this function. We solve this equation numerically
with use of an expansion in a basis of symmetric multivariate
polynomials~\cite{SymPolys}. The valence eigenvalue problem
then yields the mass $M$ of the eigenstate.
We limit the calculation to one without symmetry breaking.
The negative mass-squared case, with its wine-bottle potential,
requires careful consideration of the vacuum
state~\cite{RozowskyThorn,Varyetal}. An
extension of the LFCC method to handle this has been developed~\cite{LFCCzeromodes},
but we do not apply it here, in order to not complicate the discussion
of this first numerical LFCC application.
Instead, we focus on a comparison of the LFCC calculation
with a Fock-space truncation calculation. Specifically,
we consider a truncation that yields the same number of
equations and therefore the same computational effort (aside
from the nonlinearity in the LFCC equations). This is a
truncation to at most three constituents. One can then
see explicitly the sector dependence induced by Fock-space truncation
and avoided by the LFCC truncation.
The details of the LFCC analysis are given in Sec.~\ref{sec:analysis},
along with those of the direct Fock-space truncation, for comparison.
Results of numerical solutions are shown and discussed
in Sec.~\ref{sec:results}. A brief summary is presented in Sec.~\ref{sec:summary}.
The specifics of the simplification of the LFCC equations are left to
Appendix~\ref{sec:reduction}, and Appendix~\ref{sec:methods}
summarizes the numerical methods and illustrates their convergence.
\section{Analysis}
\label{sec:analysis}
The Lagrangian for two-dimensional $\phi^4$ theory is
\begin{equation}
{\cal L}=\frac12(\partial_\mu\phi)^2-\frac12\mu^2\phi^2-\frac{\lambda}{4!}\phi^4,
\end{equation}
where $\mu$ is the mass of the boson and $\lambda$ is the coupling constant.
The light-front Hamiltonian density is
\begin{equation}
{\cal H}=\frac12 \mu^2 \phi^2+\frac{\lambda}{4!}\phi^4.
\end{equation}
The mode expansion for the field at zero light-front time is
\begin{equation} \label{eq:mode}
\phi=\int \frac{dp^+}{\sqrt{4\pi p^+}}
\left\{ a(p^+)e^{-ip^+x^-/2} + a^\dagger(p^+)e^{ip^+x^-/2}\right\},
\end{equation}
with the modes quantized such that
\begin{equation}
[a(p^+),a^\dagger(p^{\prime +})]=\delta(p^+-p^{\prime +}).
\end{equation}
The operator $a^\dagger(p^+)$ creates a boson with momentum $p^+$
and builds the Fock states from the Fock vacuum $|0\rangle$ in the form
\begin{equation}
|x_iP^+;P^+,n\rangle=\frac{1}{\sqrt{n!}}\prod_{i=1}^n a^\dagger(x_iP^+)|0\rangle.
\end{equation}
The light-front Hamiltonian for $\phi^4$ theory is
${\cal P}^-={\cal P}^-_{11}+{\cal P}^-_{22}+{\cal P}^-_{13}+{\cal P}^-_{31}$,
with
\begin{eqnarray} \label{eq:Pminus11}
{\cal P}^-_{11}&=&\int dp^+ \frac{\mu^2}{p^+} a^\dagger(p^+)a(p^+), \\
\label{eq:Pminus22}
{\cal P}^-_{22}&=&\frac{\lambda}{4}\int\frac{dp_1^+ dp_2^+}{4\pi\sqrt{p_1^+p_2^+}}
\int\frac{dp_1^{\prime +}dp_2^{\prime +}}{\sqrt{p_1^{\prime +} p_2^{\prime +}}}
\delta(p_1^+ + p_2^+-p_1^{\prime +}-p_2^{\prime +}) \\
&& \rule{2in}{0mm} \times a^\dagger(p_1^+) a^\dagger(p_2^+) a(p_1^{\prime +}) a(p_2^{\prime +}),
\nonumber \\
\label{eq:Pminus13}
{\cal P}^-_{13}&=&\frac{\lambda}{6}\int \frac{dp_1^+dp_2^+dp_3^+}
{4\pi \sqrt{p_1^+p_2^+p_3^+(p_1^++p_2^++p_3^+)}}
a^\dagger(p_1^++p_2^++p_3^+)a(p_1^+)a(p_2^+)a(p_3^+), \\
\label{eq:Pminus31}
{\cal P}^-_{31}&=&\frac{\lambda}{6}\int \frac{dp_1^+dp_2^+dp_3^+}
{4\pi \sqrt{p_1^+p_2^+p_3^+(p_1^++p_2^++p_3^+)}}
a^\dagger(p_1^+)a^\dagger(p_2^+)a^\dagger(p_3^+)a(p_1^++p_2^++p_3^+).
\end{eqnarray}
A graphical representation is given in Fig.~\ref{fig:PT}(a).
The subscripts indicate the number of creation and annihilation operators
in each term; this notation will allow a simple representation of the
terms in the effective Hamiltonian $\ob{{\cal P}^-}$.
Because the terms of ${\cal P}^-$ change particle
number by zero or by two, the eigenstates can be separated
according to the oddness or evenness of the number of constituents.
\begin{figure}[ht]
\vspace{0.2in}
\begin{center}
\begin{tabular}{c}
\includegraphics[width=15cm]{Pminus.eps} \\
(a) \\ \\
\includegraphics[width=4.5cm]{T.eps} \\
(b)
\end{tabular}
\end{center}
\caption{\label{fig:PT} Graphical representations of (a) the light-front
Hamiltonian ${\cal P}^-$ and (b) the truncated LFCC operator $T_2$.
Lines on the right represent annihilation operators, and those on
the left, creation operators. The cross is the kinetic energy contribution.
}
\end{figure}
We consider the odd case, for which the valence state $|\phi\rangle$ is just the
one-particle state $a^\dagger(P^+)|0\rangle$. The simplest contribution to
the $T$ operator is
\begin{equation} \label{eq:T2}
T_2=\int dp_1^+ dp_2^+ dp_3^+ t_2(p_1^+,p_2^+,p_3^+)
a^\dagger(p_1^+)a^\dagger(p_2^+)a^\dagger(p_3^+)a(p_1^++p_2^++p_3^+),
\end{equation}
with $t_2$ symmetric in its arguments, and we truncate $T$ to this,
as represented in Fig.~\ref{fig:PT}(b). The
projection $1-P_v$ for the auxiliary equation (\ref{eq:aux}) is then just
a projection onto the three-particle state $a^\dagger(p_1^+)a^\dagger(p_2^+)a^\dagger(p_3^+)|0\rangle$.
The effective Hamiltonian $\ob{{\cal P}^-}$ can be constructed from the
Baker--Hausdorff expansion $e^{-T}{\cal P}^- e^T={\cal P}^-+[{\cal P}^-,T]+\frac12[[{\cal P}^-,T],T]+\cdots$.
The series can be terminated, without additional approximation, when the net increase
in particle number is greater than what is needed for the truncated auxiliary equation.
However, this generates many terms that do not actually contribute to the
valence equation or to the auxiliary equation. Therefore, a more efficient approach
for the construction of these equations is to compute explicitly the matrix elements
of $\ob{{\cal P}^-}$ that enter into the projections onto the valence one-particle
sector and the three-particle sector.
The valence equation (\ref{eq:valence}) has the following contributions:
\begin{equation} \label{eq:valenceprojected}
\langle 0|a(Q^+)\left({\cal P}^-_{11}+{\cal P}^-_{13}T_2\right)a^\dagger(P^+)|0\rangle
=\frac{M^2}{P^+}\delta(Q^+-P^+).
\end{equation}
The auxiliary equation (\ref{eq:aux}) becomes
\begin{equation} \label{eq:auxprojected}
\langle 0|a(q_1^+)a(q_2^+)a(q_3^+)\left({\cal P}^-_{31}+({\cal P}^-_{11}+{\cal P}^-_{22})T_2-T_2{\cal P}^-_{11}
-T_2{\cal P}^-_{13}T_2+\frac12{\cal P}^-_{13}T_2^2\right)a^\dagger(P^+)|0\rangle=0.
\end{equation}
Graphical representations are given in Fig.~\ref{fig:LFCCeqns}.
\begin{figure}[ht]
\vspace{0.2in}
\begin{center}
\begin{tabular}{c}
\includegraphics[width=7.5cm]{valence.eps} \\
(a) \\ \\
\includegraphics[width=16cm]{auxeqn.eps} \\
(b)
\end{tabular}
\end{center}
\caption{\label{fig:LFCCeqns} Graphical representations of the (a) valence
and (b) auxiliary equations, given in Eqs.~(\ref{eq:valenceprojected}) and
(\ref{eq:auxprojected}) of the text.
}
\end{figure}
The reduction of these equations, including the evaluation of the individual
matrix elements, is carried out in Appendix~\ref{sec:reduction}.
The valence equation is reduced to
\begin{equation} \label{eq:LFCCvalence}
1+g\int\frac{dx_1 dx_2}{\sqrt{x_1 x_2 x_3}}\tilde t_2(x_1,x_2,x_3)=M^2/\mu^2,
\end{equation}
where the $x_i=p_i^+/P^+$ are longitudinal momentum fractions, with $x_3=1-x_1-x_2$,
$g$ is a dimensionless coupling constant, defined by
\begin{equation} \label{eq:g}
g=\frac{\lambda}{4\pi\mu^2},
\end{equation}
and $\tilde t_2$ is a rescaled function of longitudinal momentum fractions,
\begin{equation} \label{eq:tildet2}
\tilde t_2(x_1,x_2,x_3)=P^+t_2(x_1P^+,x_2P^+,x_3P^+).
\end{equation}
This leads to the definition of a dimensionless mass shift $\Delta$
\begin{equation} \label{eq:Delta}
\Delta\equiv g\int\frac{dx_1 dx_2}{\sqrt{x_1 x_2 x_3}}\tilde t_2(x_1,x_2,x_3),
\end{equation}
such that $M^2=(1+\Delta)\mu^2$.
Thus the valence equation determines the mass $M$, given the reduced
$T$ function $\tilde t_2$.
The final form of the auxiliary equation is~\cite{Elliott}
\begin{eqnarray} \label{eq:LFCCaux}
\lefteqn{\frac16\frac{g}{\sqrt{y_1 y_2 y_3}}
+(1+\Delta)\left(\frac{1}{y_1}+\frac{1}{y_2}+\frac{1}{y_3}-1\right)
\tilde t_2(y_1,y_2,y_3)}&& \\
&& +\frac{g}{2}\left[\int_0^{1-y_1}dx_1
\frac{\tilde t_2(y_1,x_1,1-y_1-x_1)}{\sqrt{x_1 y_2 y_3 (1-y_1-x_1)}}
+ (y_1 \leftrightarrow y_2) + (y_1 \leftrightarrow y_3)\right] \nonumber \\
&& -\frac{\Delta}{2} \left(\frac{1}{y_1}+\frac{1}{y_2}+\frac{1}{y_3}\right) \tilde t_2(y_1,y_2,y_3)
\nonumber \\
&& +\frac{3g}{2}\left\{\int_{y_1/(1-y_2)}^1 d\alpha_1 \int_0^{1-\alpha_1} d\alpha_2
\frac{\tilde t_2(y_1/\alpha_1,y_2,1-y_1/\alpha_1-y_2) \tilde t_2(\alpha_1,\alpha_2,\alpha_3)}
{\sqrt{\alpha_1 \alpha_2 \alpha_3 y_3 (\alpha_1-y_1-\alpha_1 y_2)}}\right. \nonumber \\
&& \rule{2.2in}{0mm} \left. +(y_1\leftrightarrow y_2)+(y_1\leftrightarrow y_3)
\rule{0mm}{0.25in} \right\} \nonumber \\
&& +\frac{3g}{2}\left\{ \left[ \int_{y_1+y_2}^1 d\alpha_1 \int_0^{1-\alpha_1} d\alpha_2
\frac{\tilde t_2(y_1/\alpha_1,y_2/\alpha_1,1-(y_1+y_2)/\alpha_1)\tilde t_2(\alpha_1,\alpha_2,\alpha_3)}
{\alpha_1 \sqrt{\alpha_2 \alpha_3 y_3 (\alpha_1-y_1-y_2)}} \right. \right. \nonumber \\
&& \rule{2.5in}{0mm} \left. + (y_2 \leftrightarrow y_3)
\rule{0mm}{0.25in}\right] \nonumber \\
&& \rule{1in}{0mm} \left. +(y_1\leftrightarrow y_2)+(y_1\leftrightarrow y_3)
\rule{0mm}{0.25in}\right\}=0, \nonumber
\end{eqnarray}
with $y_i=q_i^+/P^+$.
It is this form of the auxiliary equation that we solve numerically, to obtain
$\tilde t_2$, as discussed in Appendix~\ref{sec:methods}.
For comparison, we consider the Fock-space truncation approach, with the number
of constituents limited to three, so that the resulting equations have the
same number of independent variables as the LFCC equations derived above.
The eigenstate is then approximated by
\begin{equation}
|\psi(P^+)\rangle=\psi_0 a^\dagger(P^+)|0\rangle
+P^+\int dx_1 dx_2 \psi_3(x_1,x_2,x_3)\frac{1}{\sqrt{3!}}
a^\dagger(x_1P^+)a^\dagger(x_2P^+)a^\dagger(x_3P^+)|0\rangle.
\end{equation}
The action of the light-front Hamiltonian ${\cal P}^-$
then yields the following coupled system of integral equations:
\begin{eqnarray}
\lefteqn{\mu^2\psi_1
+\frac{\lambda}{6}\int\frac{dx_1 dx_2}{4\pi\sqrt{x_1 x_2 x_3}}\psi_3(x_1,x_2,x_3)=M^2\psi_1,}&& \\
&& \mu^2\left(\frac{1}{y_1}+\frac{1}{y_2}+\frac{1}{y_3}\right)\psi_3(y_1,y_2,y_3)
+\lambda\frac{\psi_1}{4\pi\sqrt{y_1y_2y_3}} \\
&& +\frac12\frac{\lambda}{4\pi}\left[ \int_0^{1-y_1} dx_1
\frac{\psi_3(x_1,y_1,1-y_1-x_1)}{\sqrt{x_1(1-y_1-x_1)y_2 y_3}}
+ (y_1 \leftrightarrow y_2) + (y_1 \leftrightarrow y_3)\rule{0mm}{0.3in}\right]
=M^2\psi_3(y_1,y_2,y_3). \nonumber
\end{eqnarray}
If we define $\tilde\psi_3=\psi_3/(6\psi_1)$, these reduce to forms
directly comparable to the LFCC equations
\begin{eqnarray} \label{eq:tildepsi1}
\lefteqn{1+g\int\frac{dx_1 dx_2}{\sqrt{x_1 x_2 x_3}}\tilde\psi_3(x_1,x_2,x_3)=M^2/\mu^2,} && \\
\label{eq:tildepsi3}
&& \frac16\frac{g}{\sqrt{y_1y_2y_3}}
+\left(\frac{1}{y_1}+\frac{1}{y_2}+\frac{1}{y_3}-\frac{M^2}{\mu^2}\right)\tilde\psi_3(y_1,y_2,y_3) \\
&& +\frac{g}{2}\left[ \int_0^{1-y_1} dx_1
\frac{\tilde\psi_3(x_1,y_1,1-y_1-x_1)}{\sqrt{x_1(1-y_1-x_1)y_2 y_3}}
+ (y_1 \leftrightarrow y_2) + (y_1 \leftrightarrow y_3)\rule{0mm}{0.3in}\right]
=0. \nonumber
\end{eqnarray}
Graphical representations are given in Fig.~\ref{fig:truncated}.
\begin{figure}[ht]
\vspace{0.2in}
\begin{center}
\begin{tabular}{c}
\includegraphics[width=8cm]{onebody.eps} \\
(a) \\ \\
\includegraphics[width=12cm]{threebody.eps} \\
(b)
\end{tabular}
\end{center}
\caption{\label{fig:truncated} Graphical representations of the (a) one-body
and (b) three-body equations in the case of Fock-space truncation,
given in Eqs.~(\ref{eq:tildepsi1}) and (\ref{eq:tildepsi3}) of the text.
}
\end{figure}
The first Fock-state equation (\ref{eq:tildepsi1}) is the same as the LFCC valence
equation (\ref{eq:LFCCvalence}); it relates
the physical mass to the bare mass through the self-energy correction due to the
three-particle intermediate state. However, the second equation differs in
several respects. The inhomogeneous term, from the coupling to the one-particle
state, and the two-two scattering term, the last term in (\ref{eq:tildepsi3}),
remain the same. The eigenvalue term that appears in the equation for
$\tilde\psi_3$ is also seen to be present in the LFCC auxiliary equation,
with use of the valence relation, $1+\Delta=M^2/\mu^2$, but the kinetic energy
contributions for the three individual constituents are not the same. In
the $\tilde\psi_3$ equation, they enter only as $\mu^2/y_i$, whereas in
the LFCC auxiliary equation, they are $M^2/y_i$, again with $1+\Delta=M^2/\mu^2$.
The sector-dependent approach~\cite{Wilson,hb,Karmanov,SecDep} would rectify this
by making $\mu$ sector dependent and equal to $M$ in the highest Fock sector;
this compensates for the lack of a self-energy correction in a sector
for which there can be no additional particles to generate such a loop.
The LFCC approach automatically inserts the correct mass without using
a sector-dependent bare mass.
The LFCC auxiliary equation also contains several terms that do not appear
in the second Fock-state equation. The fourth term is the nonperturbative
analog of the wave-function renormalization counterterm, which is
a subtraction from the loop contributions represented by the fifth
and sixth terms. These last two terms are a partial resummation
of the high-order loops generated by Fock sectors beyond the
three-particle sector. The resummation is partial, because
the truncated $T$ operator is an approximation.
The price to be paid for these additions is the nonlinearity
of the LFCC auxiliary equation. However, this does not present
any particular difficulty for its solution. Our numerical
methods are described in Appendix~\ref{sec:methods}. They rely
on an expansion of $\tilde t_2$ in a basis of fully symmetric
polynomials~\cite{SymPolys}, which converts the auxiliary
equation to a system of nonlinear algebraic equations. The
results obtained in this way are presented and discussed in
the next section. We also solve the Fock-state wave-function
equations in the same way, for numerical comparison.
\section{Results}
\label{sec:results}
The converged results for the mass-squared eigenvalues are
shown in Fig.~\ref{fig:M2vsg}. This figure compares results
for the LFCC approximation with those obtained with a
Fock-space truncation. In the latter case, two different
approximations are considered, with and without sector-dependent
bare masses. When sector-dependent masses are used, the
leading $\mu^2$ in Eq.~(\ref{eq:tildepsi3}) is replaced by
$M^2$; this compensates for the exclusion of
any self-energy corrections in the top Fock sector
by the Fock-space truncation.
There are significant
differences in behavior among the three cases, with
the LFCC result decreasing most rapidly with increasing
coupling. The sector-dependent case proves to be
intermediate, as expected; one of the improvements that
the LFCC method offers is to place sector-dependent
masses automatically. The remainder of the difference,
which neither Fock-space truncation can include, is
the resummation of contributions from all higher
Fock states.
\begin{figure}[ht]
\vspace{0.2in}
\centerline{\includegraphics[width=12cm]{M2vsg.eps}}
\caption{\label{fig:M2vsg} Mass-squared ratios $M^2/\mu^2$ versus
dimensionless coupling strength $g$ for the LFCC approximation (squares),
the Fock-space truncation (circles), and the Fock-space truncation
with sector-dependent masses (triangles).
}
\end{figure}
Also, unlike the two Fock-space truncation
cases, which continue to zero and below with increased
coupling, the LFCC eigenmass becomes complex at approximately
$g=1.5$ instead of continuing toward zero. Since the
approach to zero is associated with spontaneous symmetry
breaking, we expect that zero modes will be required to
do a proper analysis~\cite{LFCCzeromodes}.
Another signal that the LFCC approximation provides more
information is in the probabilities for higher
Fock states. In Fig.~\ref{fig:relprob}, we plot the
probability for the three-body Fock state, relative to
the one-body state, as a function of the coupling.
In general, the relative probability for the $n$-body
state is zero for $n$ even and given as follows for
$n$ odd:
\begin{equation}
R_n=\frac{1}{Z}\int\left(\prod_{i=1}^{n-1} dx_i\right) |\psi_n(x_i)|^2,
\end{equation}
where $Z$ is the probability for the one-body state and $\psi_n$ is
the wave function for the $n$-body state, as in (\ref{eq:fockexp}). In the LFCC method,
the chosen truncation (\ref{eq:T2}) of $T$ yields these wave functions as
\begin{equation}
\psi_n(x_i)=\int dP^{\prime +} \frac{1}{\sqrt{n!}}
\langle 0|\prod_{i=1}^n a(x_iP^{\prime +})
\frac{(P^+)^{(n-1)/2}\sqrt{Z}}{((n-1)/2)!}T_2^{(n-1)/2}a^\dagger(P^+)|0\rangle.
\end{equation}
The integration over $P^{\prime +}$ eliminates the momentum conserving
delta function from the projection. The plots in Fig.~\ref{fig:relprob} show
that the relative probability $R_3$ increases rapidly in the LFCC approximation
as the coupling approaches the value at which the mass value becomes complex.
The Fock-space truncation results remain slowly varying, even as the
coupling approaches values at which the mass value becomes zero.
\begin{figure}[ht]
\vspace{0.2in}
\centerline{
\includegraphics[width=12cm]{relprob.eps}}
\caption{\label{fig:relprob} Same as Fig.~\ref{fig:M2vsg},
but for the relative probability $R_3$.
}
\end{figure}
Figure~\ref{fig:contour} provides a comparison of
three-body wave functions. For the LFCC
case, what is actually plotted is the $T$ function $\tilde t_2$,
which also determines the higher Fock-state wave functions.
The main qualitative difference between cases is associated
with the two-two scattering process; when it is left out,
as in Fig.~\ref{fig:contour}(c), the structure of the three-body
wave function is much simpler.
\begin{figure}[ht]
\vspace{0.2in}
\begin{center}
\begin{tabular}{ccc}
\includegraphics[width=5cm]{t2.eps} &
\includegraphics[width=5cm]{psi3.eps} &
\includegraphics[width=5cm]{psi3no22.eps} \\
(a) & (b) & (c)
\end{tabular}
\end{center}
\caption{\label{fig:contour} Contour plots of the three-body
wave function for (a) the LFCC approximation,
(b) the Fock-space truncation, and (c) the Fock-space truncation
without two-two scattering. The dimensionless coupling strength
was $g=1$ in all three cases.
}
\end{figure}
\section{Summary}
\label{sec:summary}
We have illustrated the use of the LFCC method~\cite{LFCC} in
an application to a model theory that is simple enough to
make the method plain, yet complex enough to require
numerical techniques. This goes beyond the previous
illustration~\cite{LFCC} that used a soluble model~\cite{GreenbergSchweber}.
In particular, a comparison with a Fock-space truncation
approach shows explicitly that the LFCC method introduces
the physical mass for kinetic energy terms without use
of a sector-dependent parameterization~\cite{SecDep},
as is seen in the comparison of Eqs.~(\ref{eq:auxprojected})
and (\ref{eq:tildepsi3}). In the former, all kinetic
terms contain $1+\Delta\equiv M^2/\mu^2$ and in the
latter, the three-body terms contain only $1=\mu^2/\mu^2$.
Thus, if the physical mass is to be restored in the highest
sector of a truncated Fock space, where no self-energy
loops can occur, the bare mass must be sector dependent
and set equal to the physical mass only in that highest
sector. The LFCC method arranges the mass automatically,
without use of sector-dependent bare masses.
A comparison of numerical results is given in Figs.~\ref{fig:M2vsg},
\ref{fig:relprob},
and \ref{fig:contour}. The calculations done with Fock-space
truncation clearly yield results which differ those of the
LFCC calculation. This is to be expected, because the LFCC
method includes effects of higher Fock states. However, the
three-body wave functions are qualitatively similar, with
the dominant effect being the inclusion of two-two scattering.
Future work along these lines could include various extensions
of the $T$ operator, which could be used to study convergence
of the LFCC results as more terms are added. Also, the process
of symmetry breaking is of particular interest, for both positive
and negative $\mu^2$.
A partial analysis can be done by comparing odd and even
eigenstates, by looking for degeneracy in the lowest
states~\cite{RozowskyThorn,Varyetal};
this will require consideration of valence
states with two particles and at least one additional term
in $T$. A full analysis is best done with inclusion of
modes of zero longitudinal momentum; for this some preliminary
work has already been done~\cite{LFCCzeromodes}.
\acknowledgments
This work was supported in part by the Department of Energy
through Contract No.\ DE-FG02-98ER41087
and by the Minnesota Supercomputing Institute through
grants of computing time.
We thank J. Vary for comments on preliminary results.
|
\section{Introduction}
In modern Commercial Off-The-Shelf (COTS) multicore systems, many
parallel memory requests can be
sent to the main memory system at any given time for the following two
reasons. First, each core employs a variety of techniques---such as
non-blocking cache, out-of-order issues, and speculative
execution---to hide memory access latency. These techniques allow the
core continue to execute new
instructions while it is still waiting memory requests for previous
instructions to be completed. Second, multiple cores can run multiple
threads, each of which generates memory requests.
These parallel memory requests from the processor put high pressure
on the main memory system. To deliver high performance, modern DRAM
consists of multiple resources called banks that can be accessed in
parallel. For example, a typical DRAM module has 16 banks, supporting
up to 16 parallel accesses~\cite{micronddr3}. To efficiently utilize
the available bank level parallelism,
modern COTS DRAM controllers employ sophisticated techniques such as
out-of-order request processing, overlapped request dispatches, and
interleaved bank mapping~\cite{rixner2000memory,natarajan2004study,chatterjee2012staged}.
While parallel processing of multiple memory requests generally
improves overall memory performance, it is very difficult to understand
precise memory performance especially when multiple applications run
concurrently, because each memory request is more likely to be
interfered by other requests. Therefore, reducing interference and
improving performance isolation in COTS multicore systems has been an
important research topic in the real-time systems community.
To this end, software based DRAM bank partitioning ~\cite{yun2014rtas,liu2012software,suzuki2013coordinated}
and last-level cache (LLC) partitioning ~\cite{zhang2009towards, mancuso2013rtas,
ding2011srm, ward2013rtas, lin2008gaining} have been studied by many
researchers. These approaches reduce interference by allocating
dedicated cache space and/or DRAM banks. While effective, it is also
shown that partitioning these resources alone does not provide ideal
isolation due to interference in other parts of the memory
hierarchy, most notably in the DRAM
controller and the shared memory bus (command and data) which connects
the controller and the DRAM module ~\cite{yun2014rtas,
kim2014rtas}. As a result, it is still difficult to understand
worst-case memory interference delay, even when the LLC and DRAM banks
are partitioned.
Recently, Kim et al. proposed an analysis method that takes the DRAM
controller and the shared memory bus into
account~\cite{kim2014rtas}. They faithfully model a modern COTS DRAM
system, and provide an analytic upper bound on the worst-case memory
interference delay of each memory request of the task under analysis.
However, their analysis made a significant assumption that makes it
difficult to apply it to modern COTS multicore systems. Specifically, the
analysis assumes that each core can only generate one outstanding
memory request at a time and stalls until the memory request is
served. Unfortunately, this assumption is far from reality in modern
COTS platforms. As outstanding requests can interfere with the
memory request under analysis, the actual worst-case depends on the
number of outstanding requests, which is typically substantially
higher than the core count. For example, a COTS multicore processor
used in our evaluation supports up to 32 outstanding reads and 16
outstanding writes while it has only 4 cores. As a result, the computed
memory interference bounds can be significantly optimistic than the
reality (i.e., underestimating the actual delay), as we experimentally demonstrated
in Section ~\ref{sec:result_syn}.
In this work, we present a parallelism-aware memory interference delay
analysis. We model a COTS DRAM controller that has a
separate read and a write request buffer. Multiple outstanding memory
requests can be queued in the buffers and processed in
out-of-order to maximize memory performance. Also, reads are
prioritized over writes in our model. These features are commonly
found in modern COTS multicore systems and crucially important in
understanding memory interference.
To minimize interference, we only consider a system in which the LLC and
DRAM banks are partitioned. This is easily achievable on COTS multicore
systems via software~\cite{liu2012software,suzuki2013coordinated}.
Based on the system model, our analysis provides a safe analytic upper
bound on the worst-case memory interference delay for each memory
request of the task under analysis.
We evaluate the proposed analysis on a real COTS multicore platform
with a set of synthetic benchmarks as well as
SPEC2006 benchmarks. The synthetic benchmarks are specially
designed to simulate worst possible memory interference delay. As for
synthetic benchmarks, our analysis provides a tight and safe upper
bound while the analysis in ~\cite{kim2014rtas} significantly
under-estimates the actual delay (almost double than the computed
delay). As for SPEC2006 benchmarks, we
found our analysis provides safe upper bounds for all but
two benchmarks, while the compared analysis under-estimates 11 (out of 19)
benchmarks. Investigating the two benchmarks that our analysis
under-estimated (so did ~\cite{kim2014rtas}),
we find that space competition in miss status holding registers
(MSHRs)~\cite{kroft1981lockup}, which track the status of outstanding
cache misses, can be a considerable source of additional interference
in modern COTS systems.
Based on our analysis and empirical evaluation, we propose two simple
architectural supports, which can be easily
incorporated in modern COTS, to effectively reduce worst-case
interference delay.
Our contributions are as follows:
\begin{itemize}
\item To the best of our knowledge, our work is the first that considers
memory level parallelism and read prioritized DRAM controllers to
analyze memory interference delay.
\item We experimentally validate and compare our analysis with a state
of art analysis on a real COTS multicore
platform with a set of carefully designed synthetic benchmarks as well
as SPEC2006 benchmarks.
\item We propose two simple architectural supports that can
significantly reduce worst-case memory interference delay on COTS multicore
processors.
\end{itemize}
The remaining sections are organized as follows:
Section~\ref{sec:background} provides background on COTS multicore
systems and LLC and DRAM bank partitioning techniques.
Section~\ref{sec:motivation} discusses the state-of-art memory
interference delay analysis.
We present our analysis in Section~\ref{sec:analysis} and provide
evaluation results in Section~\ref{sec:evaluation}.
Section~\ref{sec:recommendation} discusses architectural recommendations.
Section \ref{sec:related} discusses related work.
Finally, we conclude in Section~\ref{sec:conclusion}.
\section{Background: Modern COTS Multicore systems} \label{sec:background}
A modern COTS multicore system, shown in
Figure~\ref{fig:architecture}, supports a high degree of memory
level parallelism through a variety of architectural features.
In this section, we provide some background on important
architectural features of modern COTS multicore systems, and review
existing software based resource partitioning techniques.
\begin{figure} [t]
\centering
\centering
\includegraphics[width=0.40\textwidth]{architecture}
\caption{Modern COTS multicore architecture}
\label{fig:architecture}
\end{figure}
\subsection{Non-blocking Cache and MSHR}
At the cache level, non-blocking caches are used to handle multiple
simultaneous cache-misses. This is especially crucial for the shared last
level cache (LLC), as it is shared by all cores. The state of the
outstanding memory requests are maintained by a set of miss status
holding registers (MSHRs). On a cache-miss, the LLC allocates a
MSHR entry to track the status of the ongoing request and the entry is
cleared when the corresponding memory request is serviced from the
main memory. As such, the number of MSHRs effectively determines the
maximum number of outstanding memory requests directed to the DRAM
controller.
\subsection{DRAM Controller}
The DRAM controller receives requests from the LLC (or other DMA
devices) and generates DRAM specific commands to access data in the
DRAM. Modern DRAM controllers often include separate
read and write \emph{request buffers} and \emph{prioritize reads} over
writes because writes are not on the critical path for
program execution. Write requests are buffered on the write buffer of
the DRAM controller and serviced when there are no pending read
requests or the write queue is near
full~\cite{natarajan2004study,chatterjee2012staged}.
The DRAM controller and the DRAM module are connected
through a command/address bus and a data bus. Modern DRAM modules are organized
into ranks and each rank is divided into multiple
\textit{banks}, which can be accessed in \emph{parallel} provided that
no collisions occur on either buses.
Each bank comprises a row-buffer and an array of storage
cells organized as \textit{rows} and \textit{columns}. In order to
access the data stored in a DRAM row, an activate
command (\textit{ACT}) must be issued to load the data into the row
buffer first before it can be read or written. Once the data is in the
row buffer, any numbers of subsequent read or write
commands (\textit{RD, WR}) can be issued to access data in the row. If, however, a
request wishes to access a different row from the same bank, the row
buffer must be written back to the array with a pre-charge
command (\textit{PRE}) first before the second row can be
activated.
\subsection{Memory Scheduling Algorithm}
Due to hardware limitations, the memory device takes time to perform
different operations and therefore timing constraints between various
commands must be satisfied by the controller. The operation and timing
constraints of memory devices are defined by the JEDEC standard
\cite{jedec}. The key facts concerning timing constraints are:
1) the latency for accessing a closed row is much longer than
accessing a row that is already open; 2) different banks can be
operated in parallel since there are no long timing constraints
between banks.
To maximize memory performance, modern DRAM controllers typically use
a first-ready first-come-first-serve
(FR-FCFS)~\cite{rixner2000memory} scheduling algorithm that
prioritizes:
\begin{enumerate}
\item Ready commands over non-ready commands,
\item Column (CAS) commands over row (RAS) commands,
\item Older commands over younger commands.
\end{enumerate}
This means that the algorithm can process memory requests in out-of-order of their
arrival times. Note that a DRAM command is said to be~\emph{ready}
when it can be scheduled immediately as it satisfies all timing
constraints imposed by previously scheduled commands and the current
DRAM status.
\subsection{DRAM Bank and Cache Partitioning}
In order to maximize memory level parallelism, most COTS DRAM
controllers employ a version of \emph{interleaved bank} addressing
strategy. Under this scheme, consecutive memory blocks in physical
address space, typically of the size of a memory page, are allocated
to different banks. This makes pending memory requests in the DRAM
controller are likely to target different banks, thereby maximizing
memory level parallelism. In the worst case, however, it is possible
that all programs allocate memory on the same bank, resulting in
much increased memory latency compared to the average case. This
dependency on run-time decisions by the memory allocator can be a
significant potential source of unpredictability.
Furthermore, since banks are interleaved, any core in the system can
access any bank. If two applications running in parallel on different cores
access two different rows in the same bank, they can force the memory
controller to continuously pre-charge the row buffer and open a new
row every time an access is performed. This loss of row locality can
result in a much degraded row hit ratio and thus a corresponding
latency increases for both applications.
Software bank partitioning~\cite{yun2014rtas,liu2012software,suzuki2013coordinated}
can be used to avoid the problems of shared banks. The technique
leverages the page-based virtual memory system of modern operating
systems and allow us to allocate memory to specific DRAM
banks. Each core, then, can be assigned to use its
private DRAM banks, effectively eliminates bank sharing among cores
without requiring any hardware modification.
Similar techniques can also be applied to partitioning the shared
LLC as explored in ~\cite{zhang2009towards, mancuso2013rtas,
ding2011srm, ward2013rtas, lin2008gaining}. It is shown that
partitioning DRAM banks and LLC substantially reduce memory
interference among the cores~\cite{yun2014rtas}.
However, the LLC cache space and DRAM banks are not the only shared
resources contributing to memory interference. Most notably, at the
DRAM chip level, all DRAM banks fundamentally share the common command
and data bus. Hence, contention in the buses can become a bottleneck.
Furthermore, as many memory requests can be buffered inside the DRAM
controller's request buffers, its scheduling policy can greatly affect
memory interference delay. Finally, at the LLC level, the MSHRs for
the LLC are also shared by all cores even if the cache space is
partitioned. We will show its performance impact in Section
~\ref{sec:result_spec}.
\textbf{Goal:} The goal of this paper is to analyze the worst-case
memory interference delay in a cache and DRAM bank partitioned
system, focusing mainly on delay in the DRAM controller and command
and data bus between the controller and the DRAM module.
\section{The State of Art Delay Analysis and the Problem}\label{sec:motivation}
\begin{figure*} [t]
\centering
\centering
\subfigure[Initial bank queue status] {
\includegraphics[width=0.14\textwidth]{cq-multi}
}
\subfigure[Timing diagram under FR-FCFS schedule]{
\includegraphics[width=0.82\textwidth]{timing-multi}
}
\caption{Inter-bank delay caused by three previously arrived
outstanding requests. (DRAM commands are numbered according to
their arrival time to the DRAM controller.) }
\label{fig:motivation1}
\end{figure*}
\begin{figure*} [t]
\centering
\centering
\subfigure[Initial bank queue status] {
\includegraphics[width=0.14\textwidth]{cq-pess}
}
\subfigure[Timing diagram under FR-FCFS schedule]{
\includegraphics[width=0.82\textwidth]{timing-pess}
}
\caption{Inter-bank delay caused by a previously arrived row-miss request.}
\label{fig:motivation2}
\end{figure*}
In this section, we first review a state of art memory interference delay
analysis for COTS memory systems, proposed by Kim et
al.~\cite{kim2014rtas}. We then investigate some of its assumptions
that are not generally applicable in modern COTS multicore systems.
The analysis models a modern COTS memory system in great detail. While
there has been a similar effort in the past ~\cite{zheng2013worst}, this
is the first work that considers the DRAM bank level request reordering
effect (i.e., out-of-order execution of young row-hit column requests
over older row-miss requests). Here, we briefly summarize the assumed
system model and part of the memory interference delay analysis,
relevant for the purpose of this paper.
The system model assumes a single channel DRAM controller and a DDR3
memory module. The DRAM controller
includes request buffers and uses the FR-FCFS scheduling algorithm
At the high level, the analysis computes the worst-case memory
interference delay of the task under analysis $\tau_i$ either (1) as a
function of number of memory requests $H_i$ of the task (referred as
Request driven approach) or (2) as a function of the number of memory
requests generated by the other tasks on different cores (referred as
Job driven approach)---it takes the minimum of the two---similar to prior
work~\cite{yun2012ecrts}. The unique characteristics of the analysis
is that it considers both inter-bank and intra-bank (including request
reordering) memory interference delay. For the purpose of this paper,
however, we focus on their inter-bank delay analysis that assumes each
core is assigned \emph{dedicated DRAM bank partitions} using software bank
partitioning systems~\cite{yun2014rtas,liu2012software}.
The analysis assumes that each memory request of $\tau_i$ is composed
of PRE, ACT, and RD/WR DRAM commands (i.e., a row-miss) and each of
the command can be delayed by DRAM commands generated by other tasks
on different cores, due to inter-bank timing constraints imposed by
the JEDEC DDR3 specification~\cite{jedec}. These timing constraint
imposed inter-bank delay for PRE, ACT, and RD/WR commands are denoted
as $L_{inter}^{PRE}$, $L_{inter}^{ACT}$, and $L_{inter}^{RW}$,
respectively.
One major assumption of the analysis is that each core can generate
only \emph{one outstanding memory request} to the DRAM controller.
Based on this assumption, the worst-case per-request inter-bank memory
interference delay on a core $p$, $RD_p^{inter}$, is simply expressed by
$
RD_p^{inter} = \sum_{\forall q: q \neq p} \times ( L_{inter}^{PRE} + L_{inter}^{ACT} + L_{inter}^{RW}).
$
Finally, the total memory interference delay of a task is calculated
by multiplying $RD_p^{inter}$ to the number of total LLC misses $H_i$
of $\tau_i$
The analysis, however, has two main problems when it is applied to
modern COTS multicore systems. On the one hand, it is
overly \emph{optimistic} as it assumes each interfering core only can
generate one outstanding memory request at a time. Hence, it
essentially limits the maximum number of competing requests to the
number of cores in the system. However this is far from reality as
modern COTS multicore can generate many parallel memory requests at a
time. For example, a quad-core processor used in our evaluation can
generate up to 48 concurrent DRAM requests at a time (see
Section~\ref{sec:setup} for details). Because DRAM
performance is much slower than CPU performance, these requests are
queued inside the DRAM controller and can aggravate the overall delay.
Figure~\ref{fig:motivation1} illustrates
this problem. In the figure, three parallel requests RD1, RD2, and RD3
are already in the command queue for Bank2, when the request RD4 has
arrived at Bank1. Note that the DRAM commands are numbered in the
order of their arrival times in the DRAM controller. At memory clock
0, both RD1 and RD4 are ready, but RD1 is scheduled as FR-FCFS policy
prioritizes older requests over younger ones. Similarly, RD2 and RD3
are prioritized over RD4 at time 4 and 8, respectively. At other times
such as at clock 1, RD4 cannot be scheduled due a channel timing
constraint ($tCCD$), even though it
is ready w.r.t. the Bank1.
On the other hand, it is also overly \emph{pessimistic} as a memory
request---composed of PRE, ACT, and RD/WR DRAM sub-commands---is
assumed to suffer inter-bank interference for each sub-command, while
in reality the delays of executing sub-commands of a memory request
are not additive on efficient modern COTS memory
controllers.
Figure~\ref{fig:motivation2} shows such a case. In the
figure, each bank has one row miss DRAM request. Hence, each has to
open a new row with a ACT command followed by a RD command. Following
the FC-FRFS policy, ACT1 on Bank2 is executed first at clock
0. Even though ACT2 is targeting to a different bank, it is not
scheduled immediately due to the required minimum separation time $tRRD$
between two inter-bank ACT commands. At clock 4, however, ACT2 can be
issued even though ACT1 on the Bank2 is still in progress. In other
words, the two memory requests are \emph{overlapped}. Hence, when RD2
is finally issued at time 11, there is no extra inter-bank delay other
than the initial delay of $tRRD$.
From the point of view of WCET analysis, the former problem is more
serious as it undermines the safety of the computed WCET.
We experimentally validated the former problem on our test platform with
carefully engineered synthetic tasks, as we will detail in
Section~\ref{sec:result_syn}. To summarize the result, the calculated
worst-case response time using the stated analysis is up to 53\% smaller
than the measured worst-case response time. The result motivates our
analysis in the next section.
\section{Parallelism-Aware Memory Interference Delay Analysis} \label{sec:analysis}
In this section, we present our parallelism-aware memory interference
delay analysis that is aimed to support modern COTS multicore
systems. We begin by defining the system model on which our analysis
is based. We then present the main analysis with examples.
\subsection{System Model}
We consider a modern multicore architecture described in
Section~\ref{sec:background}. Specifically, there are $N_{proc}$ identical cores
in a single processor chip. A single LLC and MSHRs are shared among
the cores. When there is a miss in the LLC, an entry is registered on
the MSHR and it is removed when the associated DRAM transaction is
completed.
We assume a typical shared L3 cache that employs write-back
write-allocate policy. Hence
a write to DRAM only occurs when there is a L3 miss (either read or
write) that evicts a modified cache-line in the L3 cache, and
program execution can proceed without waiting the write request to be
processed in the DRAM. Therefore, for the analysis purpose, we only
consider memory interference delay imposed to each read request of the
task under analysis $\tau_i$. Note that the number of DRAM read requests
$H_i$ is equal to the number of LLC misses because, in a write-back
write-allocate cache, a write miss also generates a DRAM read request to
allocate the line in the L3 cache and then write to it.
On the DRAM controller side, we assume a modern DRAM controller that
supports the FR-FCFS scheduling
policy~\cite{rixner2000memory,wang2005umd} which is connected to a DDR3
DRAM module. At each memory clock tick, we assume a highly efficient
FR-FCFS scheduler that picks the highest priority ready command among all
requests and can overlap multiple requests simultaneously as long as
DRAM bank and channel level timing constraints and the
FR-FCFS priority rules are satisfied~\cite{natarajan2004study}.
We also assume that the DRAM controller has a read request buffer and a
write request buffer, and prioritizes reads over writes. The writes are
only serviced when there is no pending read request or the
write buffer is full.
The maximum number of prior read requests queued in the read request
buffer is denoted as $N_{rq}$ and we assume it is much bigger than
$N_{proc}$. In processing write requests, we assume it processes at
least $N_{wq}$ requests in a batch to amortize the cost of the bus
turnaround delay ~\cite{chatterjee2012staged}.
The values of $N_{rq}$ and $N_{wq}$ are platform specific and
determined by the number of MSHRs, the size of read request buffer and
the write-scheduling algorithm in the DRAM controller.
Table~\ref{tbl:sysparams} shows the parameters we used throughput this
paper which closely model our evaluation platform described in
Section~\ref{sec:setup}. ~\footnote{
Each core in our evaluation platform can have up to 10 outstanding
memory requests~\cite{david2012bandit}. Hence, $N_{rq} = 10 \times
(N_{proc} - 1) = 30$. As for $N_{wq}$, we use the value of Intel 870
memory controller ~\cite{natarajan2004study}.}
All previously mentioned assumptions closely follow common
behaviors of commercial COTS DRAM controllers~\cite{natarajan2004study}.
We assume open-page policy is used for bank management to
maximize data locality.
We assume a single rank DRAM module for simplicity but our analysis
can be extended to consider a multi-rank DRAM module.
We assume DRAM banks and the LLC space are partitioned on
a per-core basis. In other words, each core is assigned its own private
DRAM banks and LLC space. This can be easily achieved by using
software partitioning techniques on COTS
systems~\cite{yun2014rtas,liu2012software}.
Finally, we assume that any increase in memory latency is additive to
the task's execution time as in~\cite{kim2014rtas}. This is a
pessimistic assumption given that we consider out-of-order cores that
can hide much of memory access latency. Modeling reduced memory
latencies by OoO cores is, however, out of the scope of this paper.
In short, our system model is similar to ~\cite{kim2014rtas}, but
significantly differs in that (1) it models multiple parallel memory
requests buffered in the DRAM controller, and (2) it maintains
separate read and write request queues in the DRAM controller and
reads are prioritized over writes. Both are common characteristics
of modern COTS multicore memory systems ~\cite{natarajan2004study}.
Lastly, Table~\ref{tbl:dramparams} shows the DRAM parameters we used
throughout this paper.
\begin{table}[htbp]
\centering
\caption{System parameters for our evaluation platform}
\begin{tabular}{rrr}
\hline
Symbols & Description & Value \\
\hline
$N_{rq}$ & Maximum no. of prior read requests & 30 \\
$N_{wq}$ & Maximum no. of prior write requests & 4 \\
$N_{proc}$ & Number of cores & 4 \\
\hline
\end{tabular}%
\label{tbl:sysparams}%
\end{table}%
\begin{table}[htbp]
\centering
\caption{DRAM timing parameters~\cite{micronddr3}}
\begin{tabular}{rrrr}
\hline
Symbols & Description & DDR3-1066 & Units \\
\hline
$tCK$ & DRAM clock cycle time & 1.87 & nsec \\
$tRP$ & Row precharge time & 7 & cycles \\
$tRCD$ & Row activation time & 7 & cycles \\
$CL$ & Read latency & 7 & cycles \\
$WL$ & Write latency & 6 & cycles \\
$tBURST$& Data burst duration & 4 & cycles \\
$tCCD$ & Column-to-Column delay & 4 & cycles \\
$tWTR$ & Write to read delay & 4 & cycles \\
$tRRD$ & Activate to activate delay & 4 & cycles \\
$tRTP$ & Read to precharge delay & 4 & cycles \\
$tFAW$ & Four activate windows & 20 & cycles \\
$tRC$ & Row cycle time & 27 & cycles \\
\hline
\end{tabular}%
\label{tbl:dramparams}%
\end{table}%
\begin{figure*} [t]
\centering
\centering
\subfigure[Initial bank queue status] {
\includegraphics[width=0.14\textwidth]{cq-reorder}
}
\subfigure[Timing diagram under FR-FCFS schedule]{
\includegraphics[width=0.82\textwidth]{timing-reorder}
}
\caption{Out-of-order request processing due to bank timing constraints.}
\label{fig:timing-reorder}
\end{figure*}
\begin{figure*} [t]
\centering
\centering
\includegraphics[width=0.95\textwidth]{timing-worst}
\caption{Worst-case per-request inter-bank interference delay.}
\label{fig:timing-worst}
\end{figure*}
\subsection{Delay Analysis}
We now present our analysis that considers parallel memory requests
in modern COTS multicore systems.
As mentioned in the previous section, write memory requests are not in
the critical path of program execution in modern COTS systems. Hence,
our primary concern is memory interference delay to read
requests of the task under analysis.
As we consider a system in which both the LLC cache and DRAM banks
are partitioned on a per-core basis, conventional cache space and share bank
level contention do not exist. However, because the command and data
bus are shared in processing the queued memory requests, DRAM
controller's request scheduling greatly affects memory interference
delay. We now detail our delay analysis. Because the DRAM
controller prioritize reads and the bus turn-around cost is high, the
DRAM controller process requests in a \emph{batch} for either reads or
writes. In each processing mode, we analyze the worst-case memory
interference delay to a newly arrived read request.
\subsubsection{Data Bus Contention Delay}
When the DRAM controller is in the read processing mode, the
worst-case to a newly arrived read request occurs when the request
buffer is fully occupied by previously arrived $N_{rq}$ read requests
from the other competing cores. Furthermore, regardless whether the
read request under analysis is row-hit or row-miss, the worst-case
interference delay occurs when all the previous reads are pipelined
(i.e., overlapped scheduling~\cite{natarajan2004study}).
Reads can be pipelined in two cases: consecutive reads on the
same row or reads over different banks (see Figure~\ref{fig:motivation2}).
When reads are pipelined, the data bus is fully occupied and the
newly arrived read (the request under analysis) must wait until all
the previous reads are processed (because FR-FCFS prioritize older
requests). If, however, the previous requests are not
pipelined, the read request under analysis (younger request) can be
processed ahead of older requests on the other banks
(i.e. out-of-order processing). Figure~\ref{fig:timing-reorder} shows
such an example. At time 0, both RD3 and RD1 are ready to be scheduled
and FR-FCFS schedules RD1 as it is older than RD3. At time 4, both
PRE1 and RD3 are ready, but this time FR-FCFS chooses RD3 as it first
prioritizes row-hit column commands (i.e., RD) over other commands
(i.e., PRE, ACT).
When read requests are pipelined, the processing time of each read is
$tBURST$. Therefore, the delay caused by previously arrived read
commands $L_{rq}$ is
\begin{equation}
L_{rq} = N_{rq} \times tBURST.
\end{equation}
Note that if a read request under analysis needs to execute PRE or ACT
commands (closing the previous row and open a new row, respectively),
they can be executed in parallel by the time the data bus becomes
free, without adding to the total delay.
\subsubsection{Write Draining and Bus Turn-around Delay} \label{sec:wqdrain}
When there is no pending reads or the write request buffer is full,
the DRAM controller switches the mode to process pending writes. It is
called write draining and once the drain process begins, new incoming
read requests must wait until at least
$N_{wq}$ writes are drained to amortize the bus switching cost
$tWTR$~\cite{chatterjee2012staged}.
In draining writes, the worst-case happens when all writes access
different rows in the same bank, forcing the memory
controller to close and open a new row for each write. In this case, the
required time between two successive writes is determined by the row
cycle time $tRC$.
Therefore, the write queue draining delay $L_{wq}$ is given by
\begin{equation}
\begin{split}
L_{wq} = N_{wq} \times tRC + tWTR. \\
\end{split}
\label{eq:wqdrain}
\end{equation}
Then, the worst-case delay for a read request for the core under analysis
arises when the request arrived right after (1) the write queue drain process
began and (2) $N_{rq}$ read requests from other cores
arrived. A simplified illustrative example is shown in
Figure~\ref{fig:timing-worst}. In the figure, at time 0, three events
occurred in-order: (1) the write queue drain started to process two pending
write requests (WR1 and WR2); (2)two read requests (RD1 and RD2) from
competing cores arrived; (3) a read request (RD3) from the core under
analysis arrived. In this case, the RD3 must wait until all previous
activities finish.
Therefore, the worst-case inter-bank delay $D_p$ for a read request on
the core under analysis $p$ is expressed as follows:
\begin{equation}
D_{p} = L_{rq} + L_{wq}.
\end{equation}
Finally, the total inter-bank memory delay of $t_i$ can be computed by
\begin{equation}
H_i \times D_p,
\end{equation}
where $H_i$ is the number of LLC misses.
\section{Evaluation}\label{sec:evaluation}
In this section, we first present details on the hardware and software
platform used in our evaluation. We then present our evaluation
results obtained using a set of synthetic and SPEC2006 benchmarks.
\subsection{Evaluation Setup} \label{sec:setup}
Our hardware platform is a quad-core Intel Xeon W3530
(Nehalem)~\cite{molka2009memory} based computer. Each core has a
private L1 cache (32K-I/32K-D) and a private L2 cache (256~KiB),
and all cores share a 8MiB L3 cache. Shared MSHRs (called Global
Queue or GQ~\cite{intel2012optimization}) track the status of
up to 32 read requests and 16 write requests from all cores. According
to ~\cite{david2012bandit}, a single core can generate up to 10
concurrent read requests at a time, which we also experimentally
verified to be true in our test platform.
The memory controller (MC) is integrated in the processor and clocked
at 1066~MHz. The computer equips a single-channel
dual-rank 4~GiB PC10666 DDR3 DIMM module which includes 16 DRAM
banks. We disabled all hardware prefetchers, dynamic voltage
and frequency scaling, and the turbo-boost feature for better
predictability.
We use PALLOC~\cite{yun2014rtas} to partition DRAM banks and the L3
cache. For the purpose of our evaluation, we assign one private DRAM bank and
1/4 (2MiB) private L3 cache partition to each core. Therefore, there are
neither cache space evictions nor DRAM bank conflicts caused by
memory accesses from contending cores.
For measurement, we use Linux kernel's $perf$ infrastructure to
monitor LLC miss hardware performance counter.
\subsection{Results with Synthetic Benchmarks} \label{sec:result_syn}
\begin{figure} [t]
\centering
\centering
\includegraphics[width=0.45\textwidth]{result_syn}
\caption{Measured and analytic worst-case response times of
\emph{Latency} benchmark under high memory interference.}
\label{fig:result_syn}
\end{figure}
We now investigate the validity of our analysis compared to
experimental results obtained using a set of carefully engineered
synthetic benchmarks.
In this experiment, our goal is to simulate and measure the worst-case
memory interferences on a system in which DRAM banks and the LLC are
partitioned. We use \emph{Latency} benchmark~\cite{yun2013rtas} as
the task under analysis. The benchmark is a pointer-chasing application
over a randomly shuffled linked-list. Due to data dependency, it only
can generate one outstanding memory request at a time. Furthermore,
because the size of linked list is two times bigger than the size of
the LLC, each memory access is likely to result in a cache miss, hence
generating a DRAM request. As a result, its execution time highly
depends on DRAM performance and any delay in its memory access will
directly contribute to its execution time increase. In effect, this
benchmark defeats any potential benefits from out-of-order instruction
processing and other memory latency hiding techniques (i.e., an
equivalent of in-order processing).
We first run the Latency benchmark alone on Core0 to collect
its solo execution time and the number of LLC misses. We then
co-schedule three memory intensive tasks on the
other cores (Core1-3) to generate high memory traffic and measure the
response time increase of the Latency benchmark.
Note that the number of L3 misses of Latency do not change between
solo and co-scheduled execution as the L3 cache space is
partitioned. Furthermore, because each core also has a dedicated DRAM
bank, the number of DRAM row hit and misses also would not
change. Therefore, any response time increase mainly comes from
contention in the DRAM controller and its shared command and
data bus which we modeled in Section~\ref{sec:analysis}.
We repeat the experiment with three different memory intensive
benchmarks: \emph{Bandwidth(read)}, \emph{Bandwidth(write)}, and
\emph{Stream}\footnote{We use the code provided by the authors of
\cite{kim2014rtas}.}. All three benchmarks essentially access a big
array continuously but differ in their access
patterns---Bandwidth(read) only performs
consecutive reads; Bandwidth(write) do writes only; Stream performs both
reads and writes. Because memory accesses of these benchmarks do not
have data dependencies, modern Out-of-Order (OoO) cores can generate
as many outstanding requests as possible, hence simulating the
worst-case as their requests will occupy most of the read request
buffer (and the write request buffer) in the DRAM controller.
Figure~\ref{fig:result_syn} shows both measured and analytically
calculated response times of the Latency benchmark (normalized to the
solo execution time). First, measured response times in the
left side of the figure show that Bandwidth(write) causes the
highest memory interference. This is because the benchmark generates
two memory requests---a DRAM read (a cache-line allocation) and a DRAM
write (a write-back)---for each LLC miss, and processing writes can add
high delays due to reasons described in Section~\ref{sec:wqdrain}.
Second, note that the state-of-art
analysis~\cite{kim2014rtas} significantly under-estimates the memory
interference delay---the computed WCET is just 53\% of the measured
worst-case response time. This is mainly due to the fact that the
analysis assumes only one memory request from each competing core
while in this experiment competing cores generate many requests at a
time occupying the request buffers. On the other hand, our analysis,
denoted as \emph{Ours}, provides a safe upper bound for all
cases. Note that our analysis that ignores write-queue induced
worst-case latency, denoted as \emph{Ours(nowq)}, provides an upper
bound when the co-scheduled task
performs read only---i.e., Bandwidth(read)---but fails to do so when
the co-scheduled task performs many writes---i.e., Stream and
Bandwidth(write)---because it does not account additional
delay caused by occasional write buffer draining.
Lastly, the calculated WCET of our analysis is considerably
higher than the measured worst-case response
time by about 29\%. We believe this is because
our analysis assumes that all
writes are row-misses (see Eq.~\ref{eq:wqdrain}) in draining the
write-queue, while the actual writes from the benchmark are mostly
row-hits. In other words, the analysis over-estimated write-queue
draining delay $L_{wq}$.
\begin{table}[t]
\centering
\caption{SPEC2006 benchmark characteristics}
\begin{tabular}{rrrc}
\hline
\multirow{2}[0]{*}{Benchmark} & Average & LLC misses & Memory \\
& IPC & per msec & intensity \\
\hline
462.libquantum & 0.52 & 32497 & \\
482.sphinx3 & 0.70 & 22429 & \\
437.leslie3d & 0.39 & 21478 & high \\
450.soplex & 0.24 & 17970 & \\
471.omnetpp & 0.30 & 16629 & \\
\hline
403.gcc & 0.89 & 8465 & \\
483.xalancbmk & 0.11 & 7035 & \\
465.tonto & 1.21 & 5995 & \\
447.dealII & 1.41 & 4941 & medium \\
445.gobmk & 1.12 & 2531 & \\
456.hmmer & 1.94 & 2001 & \\
454.calculix & 2.38 & 1970 & \\
458.sjeng & 1.33 & 1672 & \\
435.gromacs & 1.12 & 1334 & \\
\hline
400.perlbench & 0.37 & 907 & \\
464.h264ref & 1.92 & 759 & \\
444.namd & 1.61 & 372 & low \\
416.gamess & 2.08 & 40 & \\
453.povray & 1.35 & 0 & \\
\hline
\end{tabular}%
\label{tbl:spec2006}%
\end{table}%
\begin{table}[htbp]
\centering
\caption{Normalized response times of SPEC2006 benchmarks
with three memory intensive tasks.}
\begin{tabular}{r|c|rr||rr}
\hline
\multirow{2}[0]{*}{Benchmark} & \multirow{2}[0]{*}{Measured} &
\multicolumn{2}{c||}{Calculated} &
\multicolumn{2}{c}{Pessimism} \\
& & Kim\cite{kim2014rtas} & Ours
& Kim\cite{kim2014rtas} & Ours \\
\hline
462.libquantum & 3.22 & 5.19 & 15.10 & 61\% & 369\% \\
482.sphinx3 & 3.31 & 3.89 & 10.73 & 18\% & 224\% \\
437.leslie3d & 2.45 & 3.77 & 10.32 & 54\% & 321\% \\
450.soplex & 2.45 & 3.32 & 8.80 & 35\% & 259\% \\
471.omnetpp & 3.01 & 3.15 & 8.21 & 4\% & 173\% \\
\hline
403.gcc & 2.53 & 2.09 & 4.66 & -17\% & 85\% \\
483.xalancbmk & 1.68 & 1.91 & 4.05 & 14\% & 141\% \\
465.tonto & 1.78 & 1.77 & 3.60 & 0\% & 103\% \\
447.dealII & 1.59 & 1.64 & 3.14 & 3\% & 98\% \\
445.gobmk & 1.34 & 1.33 & 2.10 & -1\% & 57\% \\
456.hmmer & 1.32 & 1.26 & 1.87 & -5\% & 42\% \\
454.calculix & 1.31 & 1.25 & 1.85 & -4\% & 42\% \\
458.sjeng & 1.35 & 1.22 & 1.73 & -10\% & 28\% \\
435.gromacs & 1.20 & 1.17 & 1.58 & -2\% & 32\% \\
\hline
400.perlbench & 1.23 & 1.12 & 1.39 & -9\% & 14\% \\
464.h264ref & 1.18 & 1.10 & 1.33 & -7\% & 12\% \\
444.namd & 1.08 & 1.05 & 1.16 & -3\% & 7\% \\
416.gamess & 1.07 & 1.01 & 1.02 & -6\% & -5\% \\
453.povray & 1.35 & 1.00 & 1.00 & -26\% & -26\% \\
\hline
\end{tabular}%
\label{tbl:result_spec}%
\end{table}%
\subsection{Results with SPEC2006 Benchmarks} \label{sec:result_spec}
In this subsection, we evaluate the response times of SPEC2006
benchmarks. The main characteristics of 19 benchmarks we
used are given in Table~\ref{tbl:spec2006}. We exclude 10 (out of 29)
benchmarks whose memory footprints are bigger than a DRAM bank
partition size (i.e., 256MB) for the purpose of our evaluation.
The basic experiment setup is the same as the previous subsection except
that we now use each of SPEC2006 benchmark as the task under analysis
instead of the Latency benchmark. As for interfering tasks, we use
Bandwidth(write) benchmark, described in the previous subsection, as
it gives worst-case memory interference.
Table~\ref{tbl:result_spec} shows the measured and analytic response
times. The two rightmost columns show pessimism in the analysis
compared to the measured response times. Note first that the baseline
analysis ~\cite{kim2014rtas} under-estimates worst-case response times
of 11 out of 19 benchmarks, although the degree of under-estimation is
much less than the engineered synthetic tasks we used in
Section~\ref{sec:result_syn}. As explained earlier, this is because
the analysis does not take multiple outstanding memory requests into
account, resulting much less queuing delay in its calculation than
reality.
Interestingly, both analyses
under-estimate the response times of 453.povray and
416.gamess. This is because both benchmarks generate very little
(close to zero) DRAM traffic, as can be see in
Table~\ref{tbl:spec2006}, the added interference is not caused by DRAM
related interference that both analyses try to estimate. This is
interesting as we already partition the L3 cache space among
cores. It means that the observed interference delay is caused neither by
cache space competition nor DRAM related interference.
To further investigate the source of the delay, we varied the number of
interfering tasks---i.e., Bandwidth(write)---from 1 to 3, and found
that performance suffers only when there are more than two Bandwidth
instances. We believe this is because the MSHRs are shared by both L2
and L3 caches in our platform so that cache misses of both caches
compete the limited MSHR space. Specifically, there are a total of 32
entries for outstanding reads where each core can use up to 10
entries. When three Bandwidth instances run (on Core1-3), they use up
to 30 entries (10 entries/core x 3 cores), it leaves only two
entries for the task under analysis (on Core0)---\emph{a 80\% reduction}
(2 out of 10). Given that both benchmarks (453.povray and 416.gamess)
show relatively high L2 miss rates, they likely suffer from the
reduction in the available MSHR entries.
We currently do not consider this MSHR space
contention in the analysis as we have no control over the allocations
of MSHRs. We will discuss how we can provide better isolation
concerning MSHR competition in Section~\ref{sec:recommendation}.
Other than these two benchmarks, our analysis provides safe upper
bounds for the rest of benchmarks we tested, albeit pessimistic. We
argue, however, this is expected behavior
given the fact that our analysis---as well as ~\cite{kim2014rtas}---does
not consider latency hiding techniques used in modern OoO cores, which
are highly effective in reducing perceived memory access latency to
the task~\cite{wang2002memory}, and we assume increased memory
access latency is addictive to the task execution time. Another major source
of pessimism in our analysis comes from the fact that we assume the
read request queue in the DRAM controller is always fully occupied by
prior requests from the interfering tasks. However, when the task under
analysis itself is highly memory intensive and has a high degree of memory level
parallelism, such as 462.libquantum, the read request queue likely
contains many memory requests from the analyzed task.
\section{Desired Architectural Support for Real-Time Systems} \label{sec:recommendation}
In this section, we discuss two simple and low-cost architectural supports
that can greatly reduce worst-case memory interference delay on COTS
multicore systems.
\subsection{Software Controlled MSHR Reservation}
MSHRs are important shared resources that determine the amount of parallelism in
the system. As experimentally shown in Section~\ref{sec:result_spec},
when a highly memory intensive task generate many parallel requests,
MSHRs become scarce, thereby significantly lower achievable
memory level parallelism of competing tasks. As a result, competing
tasks' performance would suffer. To achieve better performance
isolation, it is desirable for each core to reserve a fraction of
MSHRs, preferably by software. This can be easily implemented in
hardware, as shown in ~\cite{ebrahimi2010fairness}, and can eliminate
unintended memory interference due to contention in MSHRs.
\subsection{Software Controlled Bank Prioritization}
The biggest factor in high worst-case memory interference comes from
the fact that a large number of previously arrived memory requests are
prioritized under FR-FCFS scheduling policy, even if the newly arrived
request is from a higher priority task, effectively creating a
priority inversion problem~\cite{sha2004real}. Hence, from the
real-time perspective, it is highly desirable if software can
influence on prioritization logic of the DRAM controller. If, for
example, software can prioritize a specific bank over the other banks,
memory requests to the prioritized bank can always be processed
almost immediately without waiting the all the queued requests are
serviced. It will be especially effective for a DRAM bank partitioned
system as assumed in our analysis.
\section{Related Work} \label{sec:related}
As memory performance is becoming increasingly important in modern multicore
systems, there have been great interests in the real-time research
community to minimize and analyze memory related interference delay for
designing more predictable real-time systems.
Initially, many researchers model the cost to access the
main memory as a constant and view the main memory as a single resource
shared by the cores~\cite{yun2012ecrts,pellizzoni2010worst,yao2012memory,schranzhofer2010timing}.
However, modern DRAM systems are composed of many sophisticated
components and the memory access cost is far from being a constant as it
varies significant depending on the states of the variety of components
comprising the system.
Many researchers turn to hardware approaches and develop specially
designed DRAM controllers that are highly predictable and provide
certain performance guarantees~\cite{reineke2011pret,zheng2013worst,akesson2007predator,paolieri2009analyzable,goossen2013conservative}.
The work in \cite{reineke2011pret} and \cite{zheng2013worst} both
implement hardware based private banking scheme which eliminate
interferences caused by sharing the banks. They differ in that the controller
in \cite{reineke2011pret} uses close page policy with TDMA scheduling
while the work in \cite{zheng2013worst} uses open page policy with
FCFS arbitration. AMC~\cite{paolieri2009analyzable} and Predator
\cite{akesson2007predator} utilize interleaved bank and close page
policy. Both approaches treat multiple memory banks as a single unit of
access to simplify resource management. They differ in that AMC uses a
round-robin arbiter while Predator uses the credit-controlled
static-priority (CCSP) arbitration \cite{akesson2008rtcsa}, which
assigns priorities to requestors in order to guarantee minimum bandwidth
and bounded latency. While these proposals are valuable,
especially for hard real-time systems, they are not available in COTS
systems.
To improve performance isolation in COTS systems, several recent papers
proposed software based bank partitioning
techniques~\cite{yun2014rtas,liu2012software,suzuki2013coordinated}. They
exploit the virtual memory of modern operating systems to allocate
memory on specific DRAM banks without requiring any other special
hardware support. Similar techniques has long been applied in
partitioning shared caches
~\cite{liedtke97ospart,lin2008gaining,zhang2009towards,soares2008reducing,ding2011srm,ward2013rtas,mancuso2013rtas}. These
resource partitioning techniques eliminate space contention of the
partitioned resources, hence improve performance isolation. However,
as shown in ~\cite{yun2014rtas,kim2014rtas}, modern COTS systems have
many other still shared components that affect memory performance.
A recent attempt to analyze these effects~\cite{kim2014rtas}, which is
reviewed in Section~\ref{sec:motivation}, greatly increased our
understanding on the DRAM controller, but its system model is still
far from real COTS systems, particularly on its assumption of one
outstanding memory request per core. In contrast, our work models
a more realistic COTS DRAM controller that handles multiple
outstanding memory requests from each core and out-of-order memory
request processing (i.e., prioritizing reads over writes). We believe
our system model and the analysis capture commonly found architectural
features in modern COTS systems, hence better applicable in analyzing
memory interference on COTS multicore systems.
\section{Conclusion} \label{sec:conclusion}
We presented a new parallelism-aware worst-case memory interference
delay analysis for COTS multicore systems. We model a COTS DRAM
controller that has a separate read and a write request
buffer. The modeled DRAM controller buffers multiple outstanding memory
requests from the LLC and processes them in out-of-order fashion. It
prioritizes reads over writes and row-hit over misses. By modeling
these architectural features, which are commonly found in COTS
multicore systems, our analysis can compute more accurate worst-case
memory access delay of COTS multicore systems.
We validated our analysis on a real COTS multicore platform with a set
of carefully designed synthetic benchmarks as well as SPEC2006
benchmarks. For synthetic benchmarks, our analysis produces a tight
and safe upper bound while the compared recent work~\cite{kim2014rtas}
significantly under-estimates the interference delay. For SPEC2006
benchmarks, our analysis is more pessimistic but safer than the
compared work. These evaluation results show that our analysis is
better applicable for modern COTS multicore systems.
As future work, we will examine several architectural supports that can
provide better isolation and reduce pessimism in the analysis.
\section*{Acknowledgements} \label{acknowledge}
This research is supported in part by NSF CNS 1302563.
\bibliographystyle{plain}
|
\section{Introduction}
Observationally, our knowledge of the large-scale structure of the Galactic magnetic field has primarily been derived from studies of background starlight polarimetry \citep{1970MmRAS..74..139M, 1996ApJ...462..316H, 2000AJ....119..923H, 2012ApJ...749...71P}, Faraday rotation \citep{1997AA...322...98H, 2006ApJ...642..868H, 2007ApJ...663..258B, 2010ApJ...714.1170M, 2012ApJ...755...21M, 2011ApJ...738..192P, 2011ApJ...728...97V}, and synchrotron emission \citep{2012ApJ...757...14J, 2012ApJ...761L..11J}. Yet only Faraday rotation has successfully resolved large-scale magnetic field structures with distance. This work presents a new method for resolving the large-scale magnetic field geometry with distance by using red clump stars as standard candles in the Galaxy. Combined with near-infrared (NIR) starlight polarimetry, the magnetic field properties of a line of sight can be decomposed with distance to map variations in the plane-of-the-sky magnetic field.
Previous Faraday rotation studies have applied similar methods to studies of the large-scale structure of the Galactic magnetic field. By combining observations of Galactic pulsars with distance estimates (that sample discrete portions of the Galactic magnetic field) with observations of polarized extragalactic sources (that sample the entire Galactic line of sight towardh that object) detailed models of the Galactic magnetic field have been generated \citep[e.g., ][]{2007ApJ...663..258B, 2008AA...477..573S, 2011ApJ...728...97V}. However, these studies suffer from the limitations of Faraday rotation, specifically that they only sense the line of sight component of magnetic fields. Also, pulsar distances can be uncertain. Few parallax measurements are available and only for nearby pulsars, H{\scriptsize I} kinematic distances typically only give upper or lower distance limits, and those distances have typical uncertainties of 0.5 kpc or greater \citep{1990AJ....100..743F, 2012ApJ...755...39V}.
Starlight polarization arises as an integral quantity of the line of sight between the background star and the observer. Regions of magnetically-aligned dust grains impose weak polarizations parallel to the local magnetic field direction on background starlight. Via this mechanism the orientation of magnetic fields are directly measurable \citep{1985ApJ...290..211L, 2011ApJ...740...21P}, the plane-of-sky magnetic field strength can be indirectly inferred \citep{1953ApJ...118..113C, 2008ApJ...679..537F, 2012ApJ...755..130M}, though the field direction is not obtainable. Statistical studies of observed starlight polarization have shed light on the general symmetry and spiral-type pitch angle of the Galactic magnetic field \citep{1996ApJ...462..316H, 2012ApJ...749...71P}, but even kpc-scale details of the morphology of the magnetic field remain hidden. To measure changes in the sky-projected magnetic field with distance, as is done with Faraday rotation for the line of sight field, reliable stellar distance markers are needed.
The first steps in using starlight polarimetry to disentangle magnetic fields with distance have already been taken. \citet{2009ApJ...690.1648N} attempted to measure magnetic fields in the Galactic Center with NIR polarimetry by subtracting the polarization associated with foreground ``bluer'' field stars from the polarization of ``redder'' stars located near the Galactic Center and showed that both populations were polarimetrically indistinguishable. This work was followed by \citet{2010ApJ...722L..23N} who used the same method to observe a transition from toroidal to poloidal magnetic field above and below the Galactic center. \citet{2013AJ....145...74C} use a similar foreground subtraction method to remove a foreground Galactic contribution and resolve the magnetic field of a galaxy seen through the Taurus molecular cloud. This work extends magnetic field decomposition by moving away from statistical treatments of foreground stars, and instead uses red clump stars as standard candles to identify and characterize changes in magnetic field orientation with distance.
In color-magnitude diagrams (CMDs) of star clusters, the red clump is an overdensity of core helium burning (horizontal branch) stars seen in intermediate-age ($\sim 1-10$ Gyr) clusters and old ($\geq 10$ Gyr), metal-rich clusters \citep{2000ApJ...539..732A}. The relative invariance of red clump star K-band ($2.2\mu$m) absolute magnitude with age or metallicity \citep[$M_K=-1.57\pm0.05$;][]{2007AA...463..559V} allows it to serve as a K-band standard candle \citep{1970MNRAS.150..111C, 2000ApJ...539..732A}. In a NIR CMD of field stars, red clump stars located at different distances form an easily distinguishable feature. The photometric properties and methods for identifying this feature have been studied in the Galactic disk \citep{2002AA...394..883L, 2012MNRAS.419.1637L, 2013ApSS.346...89K}, the Large Magellanic Cloud \citep{2007AJ....134..680G}, and in star clusters \citep{2009MNRAS.394L..74G, 2011AA...535A..33M}. These studies have further refined our knowledge of the intrinsic properties of these objects, and limitations to their use.
This work implements the NIR red clump selection algorithm described in \citet{2002AA...394..883L} to identify red clump stars at different distances. For a population of stars with a fixed absolute magnitude, the apparent magnitude is an excellent proxy for distance order, since extinction monotonically increases with distance along a given line of sight. Naively assuming a constant dust density along a sightline, the physical distances to these stars are estimated below. These stars are then used as distance markers along a line of sight to probe magnetic field properties between them via NIR starlight polarimetry.
The details of a new method for measuring the Galactic magnetic field structure with distance is presented in Section 2. The method is applied to example lines-of-sight toward the inner Galaxy in Section 3 and the outer Galaxy in Section 4. Section 5 presents new results on the morphology of the Galactic magnetic field toward these two lines-of-sight and also discusses evidence for intrinsically polarized red clump stars. A summary and conclusions are presented in Section 7.
\section{Decomposing Sky-Projected Magnetic Field Orientation Along the line of sight}
Starlight linear polarization arises from dichroic extinction via aligned dust grains along the line of sight between a star and an observer. Spinning, non-spherical dust grains preferentially align with their long axes perpendicular to the local magnetic field orientation \citep{2003JQSRT..79..881L, 2007JQSRT.106..225L, 2011ASPC..449..116L}. These aligned dust grains preferentially extinct photons polarized perpendicular to the local magnetic field direction (parallel to a grain's long axis), giving rise to a weak linear polarization along the magnetic field direction. \citet{2011ApJ...740...21P} describes the equations of polarized radiative transfer and discusses the effects of magnetic field geometry on the observed polarization properties.
\begin{figure}
\epsscale{1.15}
\centering
\plotone{GP144_CMD_2.eps}
\caption{\label{cmd}Near-infrared color-magnitude diagram showing field stars toward a $100' \times 100'$ region centered on $(\ell , \it{b})$ = (19.49\degr, +0.56\degr), from 2MASS. White contours and the background greyscale image show the relative density of stars. The red clump sequence found using the method of \citet{2002AA...394..883L} is shown by the dashed white lines. The solid white triangle approximately traces the main sequence. The locations of red clump stars with $H$-band polarimetry in the $10' \times 10'$ centered on the same coordinates are identified in the figure.}
\end{figure}
Summarizing \citet{2011ApJ...740...21P}, linear polarization will trace the average, extinction-weighted, sky projected orientation of the magnetic field between an unpolarized star and observer. Magnetic field components parallel to the line of sight do not contribute to this polarization, though these are the fields traced by Faraday rotation. The Galactic magnetic field consists of regular and turbulent components in approximate equipartition. Along long (kpc) lines-of-sight in the diffuse interstellar medium, the turbulent magnetic field contributions will tend to cancel, allowing linear polarimetry to trace the large-scale structure of the Galactic magnetic field.
If the distances to stars with measured polarizations can be estimated, then the average sky-projected magnetic field orientation in the material between those stars can be measured via differences in their Stokes parameters. These differences arise in the dusty, magnetized medium between the stars. This method of decomposition requires that the different stars sample the same line of sight so that polarizing foregrounds can be reliably removed. If the angular separation on the sky between stars is large, then their light may not be sampling the same polarizing layers. The typical size scales for significant variation in the random component of the Galactic magnetic field are $10-100$ pc \citep{1993MNRAS.262..953O, 2008ApJ...680..362H}. Two stars 5 kpc distant and separated by 10 arcmin on the sky would only be physically separated by 14.5 pc, which is smaller than the typical random fluctuations in the Galactic magnetic field, and much smaller than the size scales for variations in the regular magnetic field. As the light from these two stars approach an observer and converge, they propagate through almost identical paths. If instead one star is located between the observer and the other star, the conditions are favorable for measurement of the magnetic properties between the stars via Stokes subtraction.
\subsection{Identifying Red Clump Stars}
\label{red_clump}
Red clump stars are identified using the CMD method described in \citet{2002AA...394..883L}, illustrated in Figure \ref{cmd} with Two Micron All Sky Survey (2MASS) data \citep{2006AJ....131.1163S} for a $100' \times 100'$ region centered on $(\ell , b)= (19.49\degr,+0.56\degr)$. In Figure \ref{cmd}, two significant features are seen. Main sequence stars of various reddenings are located in the region $0.5 < (J-K) < 1.0$ and $K > 8 + 4(J-K)$ (approximately traced by the solid white triangle in Figure \ref{cmd}), and the roughly diagonal feature running from the top-center toward the lower-right is dominated by red clump stars.
To objectively quantify the stars in the red clump CMD sequence, \citet{2002AA...394..883L} first grouped all stars by $K$ in 0.5 mag wide groups for $8<K<13$. For each $K$ group, a histogram of $(J-K)$ was calculated with 0.05 mag wide $(J-K)$ bins. This histogram was fit with a Gaussian to identify the $(J-K)$ location of the peak of the red clump feature at that $K$. The locus of points tracing the red clump stars was fit with a polynomial. Stars within 0.2 mag of the $(J-K)$ center of the feature are considered red clump stars. In Figure \ref{cmd}, a dashed line outlines this region.
\begin{figure}
\epsscale{1.1}
\centering
\plotone{gauss_fit_fig.eps}
\caption{\label{gauss}Normalized histograms drawn from the CMD in Figure \ref{cmd} to illustrate the red clump fitting method. Each panel is labeled with the center of its $K$-band brightness. A Gaussian fit to each histogram is shown, the center of each fit is marked with a solid vertical line. The vertical dashed lines enclose $\pm0.2$ mag in $(J-K)$, the selection criteria for red clump stars.}
\end{figure}
The details of this fitting are shown in Figure \ref{gauss}. The well-separated red clump sequence in each $K$-band brightness bin (labeled in each panel) is shown. A similar histogram for the bluer main sequence is not shown in each brightness bin. The fitted Gaussian to each histogram is also shown. These centers of these fits are marked with vertical solid lines. Stars within 0.2 mag of each $(J-K)$ center are classified as red clump stars, and these color limits are shown by the vertical dashed lines in each panel. For fainter stars, the red end of this feature may suffer incompleteness (e.g., the lower right panel in Figure \ref{gauss}), which shifts the fitted center up to 0.15 mag bluer in $(J-K)$ than would be identified by eye. However, this remains a conservative method for identifying red clump stars. Some actual red clump stars are excluded so as to minimize false positive identifications. Sample contamination will be discussed in more detail in the next subsection.
The polarimetric observations described below cover single $10' \times 10'$ regions of sky, which typically contain too few stars with 2MASS photometry to reliably identify the red clump feature. To assist with identification of red clump stars within this small region, a much larger sample of 2MASS photometry is needed. As described above, a $100' \times 100'$ region was used for the red clump star identification. Then, only those red clump stars within the much smaller $10' \times 10'$ region were flagged for use. As an example, the four red clump stars identified in the $10'\times 10'$ region toward $(\ell , b)= (19.49\degr,+0.56\degr)$ are marked with diamonds in Figure \ref{cmd}
\subsection{Sample Contamination}
A key concern about the red clump identification method described above is the contamination of other stars into this region of the CMD. In particular, reddened main sequence stars could make up a significant fraction of the selected stars. \citet{2002AA...394..883L} comment on dwarf contamination and quantify the fraction of dwarfs in the total star counts as a function of magnitude (see Figure 5 in that paper). However, that analysis is for all stars in the field and does not quantify the effectiveness of the NIR color selection method at removing reddened interlopers.
\begin{figure}
\epsscale{1.1}
\centering
\plotone{contamination2.eps}
\caption{\label{contam}Two different measures of contamination toward the example direction $(\ell , \it{b})$ = (20\degr, 0\degr) using data from the stellar population synthesis model of \citet{2003A&A...409..523R}. The dashed line shows the fraction of dwarf stars at that magnitude. The solid line shows the dwarf and supergiant contamination fraction after applying the color selection criteria from \citet{2002AA...394..883L}. The shaded region estimates the combined intrinsic uncertainty in the contamination fraction and the uncertainty associated with varying the extinction per unit length toward this direction.}
\end{figure}
To estimate the total contamination along a line of sight, we use the stellar population synthesis model of \citet{2003A&A...409..523R} to generate sample star fields toward an example direction in the inner Galaxy, $(\ell , b)= (20\degr,0\degr)$. In addition to observed photometric properties, the intrinsic properties of each star are known. The average extinction per unit length was varied from $A_V /L = 0.7-1.8$ mag kpc$^{-1}$ in the different Monte Carlo samples. Using $A_V /L=$1.8 mag kpc$^{-1}$ is a reasonable assumption for the Galactic disk, but we also included lower values as a weak proxy for sightlines at higher Galactic latitudes.
The first quantity calculated was the dwarf fraction of all stars as a function of $K$-band apparent magnitude, the same quantity calculated in \citet{2002AA...394..883L} for the outer Galaxy. This quantity is plotted by the dashed line in Figure \ref{contam}. As shown in that figure, the dwarf contamination fraction rises for fainter stars. This contamination should be dominated by nearby, cool main sequence stars \citep[e.g., ][]{2007AJ....133..439C, 2008AJ....136.1778C, 2010AJ....139.2679B}. The contamination fractions calculated here are lower than the Outer Galaxy fractions from \citet{2002AA...394..883L}. This difference can be attributed to two effects: (1) lower extinction in the outer Galaxy allows intrinsically fainter stars to remain above a given brightness limit (reducing the Malmquist bias over a fixed volume), and (2) the distant (several kiloparsec) background in the outer Galaxy has a lower stellar space density of very bright sources (e.g., red clump stars) compared to the inner Galaxy.
Next, the contamination of the fitted red clump sequence was calculated. The photometric method described in Section \ref{red_clump} was applied to each of the simulated fields to identify red clump candidates. In this case dwarfs and supergiants, though rare, were considered as interlopers. The average fraction of interlopers is shown by the solid line in Figure \ref{contam} and the shaded region represents the uncertainty of that fraction. This uncertainty contains two components: the intrinsic uncertainty in the contamination fraction and the variation caused by different assumptions about the average extinction per unit length. In general, lower extinction per unit length is associated with more red clump contamination (toward the upper envelope of the shaded region), and larger extinction per unit length is associated with less contamination (toward the lower envelope of the shaded region). Later, the measured contamination fraction in Figure \ref{contam} will be used to estimate the probability of false detections.
\subsection{Distance Estimation}
To estimate the distance of a red clump star, the K-band absolute magnitude was assumed to be $M_K=-1.57\pm0.05$ \citep{2007AA...463..559V}. This is slightly dimmer than the value \citep[$M_K=-1.65$;][]{1992ApJS...83..111W,2000MNRAS.317L..45H} used by \citet{2002AA...394..883L}. K-band extinction was estimated from the apparent NIR (H-K) colors, assuming (H-K)$_o = 0.09\pm0.03$ \citep{2009BaltA..18...19S}, $R_V = 3.1$, and the \citet{1989ApJ...345..245C} extinction law.
\subsection{Isolating Polarizing Layers in the ISM}
Together, the position angle (PA) and degree of polarization ($P$) fully characterize the linear polarization properties of observed starlight, and correspond to a unique pair of linear Stokes parameters, $U$ and $Q$. Consider a series of $n$ optically thin polarizing layers characterized by Stokes parameters ($U_1$, $Q_1$), ($U_2$, $Q_2$),...,($U_n$, $Q_n$). The linear polarization ($U_f$, $Q_f$) of an initially unpolarized beam of light propagated through these layers becomes:
\begin{equation}
U_f = \displaystyle\sum\limits_{i=0}^n U_i
\end{equation}
\begin{equation}
Q_f = \displaystyle\sum\limits_{i=0}^n Q_i,
\end{equation}
under the assumption that all layers are optically thin. With measurements of ($U_f$, $Q_f$) through different numbers of polarizing layers, the properties of those polarizing layers can be measured by differencing Stokes parameters for stars at different distances. Similar techniques for the isolation of the polarization properties of stars in the Galactic center \citep{2009ApJ...690.1648N, 2010ApJ...722L..23N} and a background galaxy seen through Taurus \citep{2013AJ....145...74C} have been used before. However, those cases only considered the measurement and removal of a single, foreground polarization component.
In the following, this method is applied to the entire line of sight to decompose the sky-projected magnetic field orientation with distance. The first example below (Section \ref{inner}) demonstrates a coarse implementation of this procedure for an example line of sight with few identified red clump stars. In Section \ref{outer}, a more sophisticated implementation is shown later for a line of sight with many red clump identifications.
\subsection{The Effect of Discrete Clouds}
An individual (dense) molecular cloud along the line of sight probed by this method will have a significant effect on the observed starlight polarization. Since starlight polarimetry uses the {\em dust-weighted} magnetic field orientation, the presence of these higher density regions could contaminate the polarization signal expected from the large-scale Galactic magnetic field.
Comparative studies of sub-millimeter thermal dust polarization and optical starlight polarimetry have shown that the magnetic orientation of these clouds are generally aligned with the local magnetic field \citep{2009ApJ...704..891L, 2014arXiv1404.2024L}. This has also been observed in giant molecular clouds in M33 by \citet{2011Natur.479..499L}. The polarization signal superimposed by these clouds also depends on the degree of turbulence within the cloud: a magnetically turbulent cloud will produce a lower polarization signal per unit extinction than a less turbulent cloud \citep{2008ApJ...679..537F}.
However, these high-density regions can be identified via star counts \citep[e.g.,][]{2009ApJ...703...52L}, long wavelength continuum emission \citep[e.g.,][]{2011ApJS..192....4A}, or line emission from high-density tracers \citep[e.g., $^{12}$CO, $^{13}$CO, or CS;][]{2013ApJ...764..102F}. Velocity-resolved line emission at radio wavelengths is perhaps the best method for identifying individual high-density regions along a single line of sight. Once identified, the polarimetric effect of these clouds can be individually assessed. In the best case scenario, red clump stars with high polarimetric signal-to-noise on both sides of the cloud are found and the polarization signal from the cloud can be independently measured. In the more likely scenario the polarization signal from the cloud will dominate any signals from more distant stars making analysis beyond the cloud impossible. In this case the clouds will create ``shadow zones'' where the magnetic field cannot be measured by this method.
\section{Discrete Magnetic Field Decomposition with Few Red Clump Stars}
\label{inner}
$H$-band ($1.6\mu$m) starlight polarimetry toward $(\ell , b)= (19.49\degr,+0.56\degr)$ was drawn from the first data release of the Galactic Plane Infrared Polarization Survey \citep[GPIPS;][]{GPIPS_III,GPIPS_I}. These observations were made with the Mimir instrument \citep{2007PASP..119.1385C} on the 1.8m Perkins telescope of Lowell Observatory outside Flagstaff, AZ. Each observation covered a $10'\times 10'$ field-of-view and data were reduced with custom IDL-based software \citep{GPIPS_II, GPIPS_I}. For each observation, a catalog of sky coordinates, starlight polarization properties, and measured $H$-band photometry were produced. These catalogs were matched to 2MASS and GLIMPSE \citep{2003PASP..115..953B} point source catalogs. GPIPS polarimetric observations are complete to $H\approx 12$, which provides a faintness (and ultimately distance) limit on red clump stars with measurable polarimetry. Data from the $^{13}$CO Galactic Ring Survey (Jackson et al. 2006) shows no significant emission at any velocity toward this direction indicating that it is free from cloud contamination.
\begin{figure}
\epsscale{1.1}
\centering
\plotone{GP144_GPAvsK.eps}
\caption{\label{gpa_vs_k}Top panel: near-infrared polarization orientation as Galactic position angle (GPA) as a function of 2MASS K-band apparent magnitude for red clump stars. The horizontal dashed line represents a polarization angle parallel to the Galactic disk. Bottom panel: GPA vs. estimated distances for the same stars.}
\end{figure}
Using the photometric method described in Section \ref{red_clump}, four candidate red clump stars were identified in this $10'\times 10'$ field. Using the estimated contamination fractions presented in Figure \ref{contam}, the K-band magnitude maps directly to a false-positive probability for that star. These $K$-band magnitudes are 8.82, 10.57, 10.92, and 11.67; which become false-positive probabilities of 0.141, 0.049, 0.047, and 0.045. The probability that all four stars are actually red clump stars (a reliability probability) is only 0.743 with the brightest star being the least likely. Nevertheless, the following analysis assumes that all four stars are actually red clump stars.
\begin{deluxetable*}{ccccccc}
\tablewidth{0pt}
\tablecolumns{7}
\tablecaption{Inner Galaxy $H$-band Polarimetry\label{pol_table}}
\tablehead{
\colhead{Star} &
\colhead{$H_{2MASS}$} &
\colhead{Distance} &
\colhead{P\tablenotemark{a}} &
\colhead{GPA\tablenotemark{b}} &
\colhead{$U$} &
\colhead{$Q$} \\
\colhead{} &
\colhead{(mag)} &
\colhead{(kpc)} &
\colhead{(\%)} &
\colhead{(deg)} &
\colhead{(\%)} &
\colhead{(\%)} \\
\colhead{(1)} &
\colhead{(2)} &
\colhead{(3)} &
\colhead{(4)} &
\colhead{(5)} &
\colhead{(6)} &
\colhead{(7)} \\
}
\startdata
1 & $9.04\pm0.02$ & $1.07\pm0.23$ & $0.31\pm0.15$ & $83.8\pm12.5$ & $0.23\pm0.15$ & $0.25\pm0.15$ \\
2 & $10.89\pm0.04$ & $2.19\pm0.61$ & $2.87\pm0.69$ & $91.4\pm 6.9$ & $2.52\pm0.68$ & $1.54\pm0.72$ \\
3 & $11.31\pm0.02$ & $2.45\pm0.56$ & $2.08\pm0.68$ & $75.3\pm 9.2$ & $0.97\pm0.71$ & $1.96\pm0.68$ \\
4 & $12.09\pm0.03$ & $3.41\pm0.87$ & $4.67\pm1.94$ & $59.0\pm11.0$ & $-0.55\pm1.94$ & $5.03\pm1.94$ \\
\enddata
\tablenotetext{a}{These P values and the derived position angle uncertainties have been corrected for positive statistical bias via the equation $P=\sqrt{U^2+Q^2-\sigma_P^2}$ \citep{1974ApJ...194..249W,2006PASP..118.1340V}.}
\tablenotetext{b}{GPA values were derived from the equatorial Stokes parameters, then rotated to the Galactic coordinate system using the equation in \citet{1968ApJ...151..907A}.}
\end{deluxetable*}
For the four red clump stars identified toward $(\ell , b)= (19.49\degr,+0.56\degr)$, the Galactic position angle (GPA) is plotted against the 2MASS $K$-band magnitude in the top panel of Figure \ref{gpa_vs_k} and all of the H-band polarization properties are listed in Table \ref{pol_table}. GPA is given in the Galactic coordinate system, where GPA=$90\degr$ is parallel to the Galactic disk, shown by the dashed line in Figure \ref{gpa_vs_k}. In the bottom panel of that Figure, the $x$-axis shows the estimated distance to each red clump star.
Qualitative estimates of the magnetic properties with distance can be deduced from Figure \ref{gpa_vs_k}. The first two stars have GPAs consistent with $90\degr$, suggesting that the magnetic field in this direction is parallel to the Galactic plane out to a distance of $\sim$2 kpc. The next two stars show a monotonic change in GPA with apparent magnitude (i.e., distance). Beyond 2 kpc, the magnetic field orientation changes. An alternative interpretation is that there is a constant change in GPA with $K$ or distance (representable by a straight line with non-zero slope) instead of a knee. A more quantitative analysis can be made using the procedure described below.
\begin{deluxetable}{ccccc}
\tablewidth{0pt}
\tablecolumns{5}
\tablecaption{Inner Galaxy Magnetic Field Properties, Binned by Distance\label{mag_table}}
\tablehead{
\colhead{Interval\tablenotemark{a}} &
\colhead{Distance} &
\colhead{GPA$_i$} &
\colhead{$\Delta U_i$} &
\colhead{$\Delta Q_i$} \\
\colhead{} &
\colhead{Range} &
\colhead{} &
\colhead{} &
\colhead{} \\
\colhead{} &
\colhead{[kpc]} &
\colhead{[deg]} &
\colhead{[\%]} &
\colhead{[\%]} \\
\colhead{(1)} &
\colhead{(2)} &
\colhead{(3)} &
\colhead{(4)} &
\colhead{(5)} \\
}
\startdata
0$\rightarrow$1 & 0$\rightarrow$1.07 & $83.8\pm12.5$ & $0.23\pm0.15$ & $0.25\pm0.15$ \\
1$\rightarrow$2 & 1.07$\rightarrow$2.19 & $92.4\pm7.9$ & $2.29\pm0.70$ & $1.29\pm0.74$ \\
2$\rightarrow$3 & 2.19$\rightarrow$2.45 & $24.7\pm17.6$ & $-1.55\pm0.98$ & $0.42\pm0.99$ \\
3$\rightarrow$4 & 2.45$\rightarrow$3.41 & $49.0\pm17.3$ & $-1.52\pm2.07$ & $3.07\pm2.06$ \\
\enddata
\tablenotetext{a}{These numbers refer to the stars listed in Table \ref{pol_table}, zero being the observer.}
\end{deluxetable}
The properties of the polarizing layers toward this direction were derived using Stokes subtraction. Table \ref{mag_table} lists the derived properties for the polarizing intervals between each adjacent pairs of stars. Each interval is designated by the star identifiers in Table \ref{pol_table}, zero referring to an observer on Earth. The distance span of each interval is bounded by the estimated distances of the stars, zero for an observer. The differences in Stokes parameters for the stars bounding each interval, $\Delta U_i$ and $\Delta Q_i$, and the corresponding polarization P.A. (rotated from equatorial to the Galactic coordinate system) are also listed. The corresponding degree of polarization is not listed because this quantity correlates more with the intervening dust column than with the magnetic field properties \citep{2011ApJ...740...21P}.
\begin{figure}
\epsscale{1.1}
\centering
\plotone{GP144_mag_vs_dist.eps}
\caption{\label{mag_vs_dist}Sky-projected magnetic field orientation as a function of distance toward $(\ell , \it{b})$ = (19.49\degr,+0.56\degr). The horizontal error bar represents the range over which this estimate spans, and the vertical error bar is the uncertainty in the sky-projected magnetic field orientation for that distance span. The horizontal dashed line shows GPA=$90\degr$, which represents alignment with the Galactic plane.}
\end{figure}
The derived dependence of the magnetic field orientation with distance along this direction is shown in Figure \ref{mag_vs_dist}. There, the horizontal error bars represent the distance intervals listed in Table \ref{mag_table} and the vertical error bars are the uncertainty in the average, dust-weighted magnetic field orientation in that interval. Out to a distance of 2 kpc, the magnetic field orientation is consistent with GPA$=90\degr$. Beyond 2 kpc, the magnetic field orientation is significantly different, being tilted approximately $45\degr$ to the direction of the Galactic plane.
From the $^{13}$CO data, it is unlikely that there is a cloud along this particular line of sight that could create a false signal. To further test this possibility the extinction per unit length (A$_v$/L) for the four stars toward this direction was calculated. If there was a dense cloud near 2 kpc there should be a large jump in A$_v$/L beyond that distance. A$_v$/L for the four stars are (in order of increasing distance): 1.88, 1.90, 1.67, and 1.46. The lack of a jump in A$_v$/L is further evidence against the presence of an intervening cloud. In fact, this result suggests that there is less extinction at further distances (probably because of increasing Galactic height).
This result is somewhat robust against the individual false-positive probabilities calculated above. The removal of any one star may increase the distance uncertainty of the change in magnetic orientation, but it would still remain detected. For example, the brightest star in Figure \ref{mag_vs_dist} has the lowest probability of being a red clump star. Excluding this star from the previous analysis does not affect the presence of the change in magnetic field orientation beyond 2 kpc, but actually increases the red clump reliability from 0.743 to 0.865.
\section{Continuous Magnetic Field Decomposition with Many Red Clump Stars}
\label{outer}
An inherent drawback of this method, well demonstrated toward the inner Galaxy line of sight in the previous section, is the small number of identifiable red clump stars with measured NIR polarization in each $10'\times 10'$ GPIPS field. With so few stars, measurement errors could easily mimic a false signal and impede interpretation of the results, any intrinsic stellar polarization could also mask the signature of the Galactic magnetic field, and false detections would introduce noise into the results. Furthermore, unidentified dense regions along the line of sight can significantly affect interpretation of the measured polarizations. Larger numbers of red clump stars were available in a sample of starlight polarimetry drawn from \citet{2012ApJ...749...71P} and is here used to illustrate this method toward an example direction in the outer Galaxy, $(\ell , \it{b})$ = (150\degr, -2.5\degr).
Red clump stars were identified in this direction using the same procedure described in Section \ref{red_clump}, and many more stars were found than in the previous inner Galaxy example. The larger number of red clump stars found in this outer Galaxy example is likely caused by lower extinction allowing larger distances to be probed, which can be seen in the bottom panels of Figures \ref{gpa_vs_k} and \ref{outer_gal}. Plots of the measured polarization GPA with apparent magnitude and estimated distance are shown in Figure \ref{outer_gal}.
\begin{figure}
\epsscale{1.1}
\centering
\plotone{Outer_gal_GPAvsK2.eps}
\caption{\label{outer_gal}Same as Figure \ref{gpa_vs_k}, but toward the $10' \times 10'$ field centered on $(\ell , \it{b})$ = (150\degr, -2.5\degr).}
\end{figure}
\begin{figure}
\epsscale{1.1}
\centering
\plotone{Outer_gal_UQ2.eps}
\caption{\label{outer_gal_uq}Stokes $Q$ (top panel) and $U$ (bottom panel) as a function of distance for red clump stars toward the $(\ell , \it{b})$ = (150\degr, -2.5\degr) field. The solid lines are linear fits to the data, the dashed lines show the standard error of that fit. Stars with possible intrinsic polarization (marked with diamonds) were excluded from the fits.}
\end{figure}
The larger number of stars available toward this outer Galaxy line of sight allows a more sophisticated analysis of changes in magnetic field orientation with distance. By fitting the change in Stokes parameters with estimated distance, the geometry of the sky-projected magnetic field can be measured without needing to invoke a specific model. Figure \ref{outer_gal_uq} shows the variation of Stokes $U$ and $Q$ parameters with estimated distance. Excluding the two stars showing evidence for intrinsic polarization (discussed below), a least-squares fit to the data was calculated and is also plotted as solid lines. A statistical F-test showed that a linear fit was appropriate and that higher order fitting terms were insignificant in both cases. In Section \ref{inner}, the entire line of sight was divided into four regions, here the sampled line of sight is divided into infinitesimally small regions and the instantaneous change in Stokes parameters is calculated from the fits in Figure \ref{outer_gal_uq}. The change in Stokes parameters with distance maps directly to the orientation of the magnetic field via the equation:
\begin{equation}
{GPA_i}=\frac{1}{2} \, atan \left[ \frac{dU_i/ds}{dQ_i/ds} \right]
\end{equation}
where $i$ refers to the $i$th polarizing region along the line of sight and $ds$ is an infinitesimal change in distance along this direction. For the fits in Figure \ref{outer_gal_uq}, $dU/ds=-0.404 \pm 0.312$ kpc$^{-1}$ and $dQ/ds=0.497 \pm 0.582$ kpc$^{-1}$. Since these values are constants, this is interpreted as a constant magnetic field orientation in this direction between 2 and 5 kpc with equatorial PA$=160.5\degr \pm 19.6\degr$ and GPA$=120.9\degr \pm 19.6\degr$.
This result can be seen in Figure \ref{outer_gal}. The line of sight between the observer and the first star at 2 kpc imposes a net polarization with GPA$=99.0\degr \pm 2.1\degr$. Stars at larger distances show a weak monotonic increase in P.A. toward GPA$=118.6 \pm 3.2\degr$. The polarization of a star at a distance of 5 kpc is dominated by the polarizing layer between 2 and 5 kpc with a constant mean magnetic field orientation measured by the slopes of $dU/ds$ and $dQ/ds$.
\section{Identification of Intrinsically Polarized Stars}
The application of this method assumes that the stars are intrinsically unpolarized. Any stars violating this assumption could be interpreted as changes in the magnetic field geometry in the polarizing medium between the stars. This could be particularly troublesome for the `sparse' example in Section \ref{inner} where the polarization of each star was used to measure the magnetic field properties between adjacent stars. The `rich' example in Section \ref{outer} uses a statistically more robust sample to fit the magnetic field properties with distance. The polarization properties of this more robust sample can be used to tentatively identify intrinsically polarized stars by identifying significant polarization deviants toward a given direction.
The 2MASS photometry and $H$-band degree of polarization of these Outer Galaxy red clump stars are shown in Figure \ref{intrinsic}. In the figure, the top panels show two CMDs: $K$ versus $J-K$ (upper left panel) and $K$ versus $H-K$ (upper right panel). If they are truly red clump stars then all should have the same absolute $K$-band magnitude and intrinsic colors. The overall trend in each CMD illustrates the red clump selection method in Section \ref{red_clump}: more distant standard candles are fainter and redder. The lower panels plot the degree of $H$-band polarization against these same colors. In these lower panels, a linear fit to the data is shown by the solid line, which recovers the expected polarization properties for the magnetic field derived in Section \ref{outer}, and the shaded regions represents the area enclosed by the standard error of the fit. In both bottom panels, the same two outliers are seen outside of the shaded region.
\begin{figure}
\epsscale{1.1}
\centering
\plotone{jk_pol_stats2.eps}
\caption{\label{intrinsic}Photometric method for identifying candidate red clump stars with intrinsic polarization. Lower panels show the degree of H-band polarization against $(J-K)$ (lower left panel) and $(H-K)$ (lower right panel) colors for outer Galaxy red clump stars. The solid lines are linear fits to each data set and the grey bands represents the standard errors of the fits. Stars outside this band tentatively show anomalous polarizations. Upper panels show near-infrared color-magnitude diagrams for these red clump stars, the intrinsically polarized stars are marked with diamonds. (A color version of this figure is available in the online journal.)}
\end{figure}
Both polarization and extinction (traced by NIR color) are integral quantities along the line of sight, meaning that the observed properties depend on the conditions between the object and the observer. While extinction monotonically increases with distance, the Stokes $U$ and $Q$ parameters are signed quantities, meaning that the degree of polarization is not necessarily a monotonic function of distance. In Figure \ref{intrinsic}, the photometric properties of the stars are `well-behaved' in that there is a correlation between apparent $K$ magnitude and color, because both fundamentally depend on distance. For a line of sight with a constant magnetic field orientation, the degree of starlight polarization would also depend monotonically on distance, and therefore would also correlate with apparent brightness and reddening.
The degree of polarization has a more complicated dependence on distance if the line of sight suffers depolarization \citep{1989ApJ...346..728J, 1992ApJ...389..602J}. Strong depolarization, for example occurring as a result of a change in the magnetic field direction at some distance, would cause the polarization trend to change and all stars beyond that distance would be affected. Similarly, the presence of a polarizing cloud along the line of sight would also affect all stars beyond its distance. This would not give rise to individual outliers, but to a discreet offset in polarization versus distance at that distance. A more likely explanation for individual outliers is that some intrinsic property of the star is contributing to its measured polarization.
The variation in Stokes $U$ and $Q$ parameters with $K$ are shown in Figure \ref{outer_gal_uq} and can distinguish between intrinsically polarized stars and the presence of a change in field orientation or a polarizing cloud. In that figure the candidate intrinsically polarized red clump stars are marked with diamonds. A linear fit to the data, excluding the intrinsically polarized candidates, is shown by the solid line. The dashed lines represent the standard error of that fit. If there were a systematic change in the magnetic field orientation or a polarizing cloud along the line of sight there would be bend or offset in the trend of $U$ and $Q$ with distance.
The candidate stars appear as outliers in that figure as well, particularly in Stokes $U$ (lower panel). The intrinsic polarization of each star (assuming they are not interlopers) was estimated by the difference between the fit and the observed Stokes parameters. For the brighter candidate star at $K=11.01$ mag (2MASS J03534501+5026427), the intrinsic degree of polarization is $1.58\%\pm0.45\%$ with GPA$=114.0\pm14.9\degr$. The fainter candidate star (2MASS J03543283+5029199) at $K=11.75$ mag has an intrinsic polarization of $1.94\%\pm0.73\%$ with GPA$=116.3\pm21.4\degr$.
These stars were assumed to be red clump stars. However, they could be true interlopers in which case their distance estimates would be incorrect. If they were distant reddened supergiants or dwarfs then the measured polarizations could be consistent with the rest of that line of sight. No additional information is currently available in the literature for these two objects, so their true identity or the cause of this anomalous polarization remains unknown. Spectroscopic follow-up could easily determine whether these stars belong to the red clump or not. Some possible explanations for anomalous polarization of red clump stars are presented below.
\section{Discussion}
Red clump stars were used as standard candles to probe the structure of the sky-projected magnetic field along two Galactic lines-of-sight. toward one field in the inner Galaxy, the magnetic field showed evidence for a large change in magnetic field orientation at a specific distance. toward the outer Galaxy, the magnetic field orientation was uniform in the region probed, however not parallel to the Galactic disk.
\subsection{Evidence for a Magnetic Reversal?}
In Figure \ref{mag_vs_dist}, the H-band polarimetry toward the inner Galaxy showed a change in the magnetic field orientation at a distance of $\sim$2 kpc. Apparently, the sky-projected magnetic field orientation has changed, but this could also be evidence for the presence of magnetic reversals in the Galactic magnetic field.
Considering the Galactic center located at the origin of a cylindrical coordinate system, the magnetic field at any point can be defined by the vector sum of radial, vertical, and azimuthal components. In the disk, theoretically \citep[e.g., ][and references therein]{1978mfge.book.....M, 2000AA...358..125F, 2008RPPh...71d6901K} and observationally \citep[e.g., ][and references therein]{1976ARAA..14....1H, 1996ARAA..34..155B}, the azimuthal component dominates the total field strength in the disk of spiral galaxies, including the Milky Way.
In the Milky Way, the dominant azimuthal component of the magnetic field does not necessarily have the same sign everywhere. Evidence for discreet azimuthal magnetic field reversals have been seen in the Galactic disk in Faraday rotation measures \citep[RMs;][]{2008AA...477..573S,2011ApJ...728...97V}. As described in \citet{2011ApJ...740...21P}, starlight polarimetry is insensitive to magnetic field direction, only its orientation, leading to a $180\degr$ ambiguity in the direction of the magnetic field aligning the dust. Formally, starlight polarimetry should be insensitive to full $180\degr$ magnetic reversals. However, if instead of simply reversing the azimuthal magnetic field direction, the azimuthal reversal is associated with a large-scale perturbation in the vertical and/or radial magnetic field components, then the projection of that field onto the sky would also change and the effects of magnetic reversals would be expected to be seen with starlight polarimetry.
Magnetic reversals are expected in two main situations: (1) there is a field discontinuity, separated by a current sheet \citep{1954ApJ...119....1F}, or (2) if the magnetic field is continuous, then the amplitude of the azimuthal magnetic field must decrease to zero before increasing in the opposite direction, though no limits are placed on the vertical and radial field components that may also change, possibly in such a way that the total field strength is constant. In the first situation the magnetic field direction is reversed on opposite sides of a current sheet. Magnetic field rotations of $180\degr$ have been observed by in situ satellites passing through the heliospheric current sheet in the solar system \citep{2004JGRA..109.3107C}. The presence of a current sheet would have little effect on the observed polarization orientation since the magnetic fields on either side of an interstellar current sheet would be indistinguishable with this NIR background starlight observational tool, illustrated in Figure 1 of \citet{2011ApJ...740...21P}. In the second situation, as the dominant azimuthal magnetic field vanishes the other field components (radial and vertical) will be expressed. This would manifest as a change in the linear polarizing properties of the medium hosting this magnetic field.
A change in the magnetic field geometry toward the inner Galaxy has been modeled from observations of RMs of Galactic pulsars and polarized extragalactic objects. toward the direction shown in Figure \ref{mag_vs_dist} ($\ell=20\degr$, $b=0\degr$), \citet{2008AA...477..573S} saw evidence for magnetic reversals at Galactocentric radii (R) of 7.5, 6, and 5 kpc (corresponding to line of sight distances of 1.1, 2.7, and 3.9 kpc, respectively) in their ASS+RING Galactic magnetic field model. In this same direction \citet{2011ApJ...728...97V} concluded that a magnetic reversal occurs at R=5.8 kpc (a line of sight distance of 3.0 kpc). It seems that the same magnetic feature seen in the RMs is being observed by NIR starlight polarimetry. If so, this would corroborate the presence of a large-scale magnetic field change toward that direction.
For now, the NIR starlight polarimetry results, while intriguing, are tentative. Spectroscopic follow-up of the stars along this line of sight is required to derive reliable spectrophotometric distances, but will require a substantial observational program to test the photometric results presented here.
\subsection{Magnetic Fields Following the Galactic Warp}
Figure \ref{outer_gal} shows a line of sight toward ($\ell$, $b$) = ($150\degr$, $-2.5\degr$). Toward this direction in the outer Galaxy, the magnetic field orientation was shown to be uniform with distance, which agrees with the observationally-driven models of \citet{2008AA...477..573S} and \citet{2011ApJ...728...97V}. Observations and simulations conclude that in this direction the disk magnetic field should be roughly parallel the Galactic plane \citep{1996ARAA..34..155B,2009ASTRA...5...43B}.
However, the deviation of the polarization P.A., as described in Section \ref{outer} (GPA$=121\degr$), is inconsistent with a plane-parallel magnetic field in the Galactic disk (GPA$=90\degr$). Geometric effects cannot account for the discrepancy. This particular outer Galaxy line of sight samples the Galactic magnetic field below the disk ($b=-2.5$), and the combination of inclined viewing angle \citep{2011ApJ...740...21P} and the pitch angle of the Galactic disk magnetic field \citep{1988AJ.....95..750V, 1994AA...288..759H, 1996ApJ...462..316H, 2007EAS....23...19B, 2011ApJ...738..192P, 2012ApJ...749...71P} only contribute a 1.1$\degr$ twist to the projected magnetic field orientation for stars within $\sim 7$ kpc.
A possible explanation for the observed magnetic field orientation in the Outer Galaxy is the presence of the Galactic warp . Using the model of \citet{2002AA...394..883L}, the Galactic warp can be described by $z(\phi)\sim sin(\phi+5\degr)$ where $\phi = 0\degr$ is toward the Sun and $\ell=150\degr$ spans $\phi=0-12.3\degr$ from the Sun to a distance of 4 kpc. For these values of $\phi$, the Galactic plane is tilted upward $44\degr$ toward Galactic North. This leads to a predicted magnetic field orientation of GPA=$134\degr$, consistent with the measured value (GPA$=121\pm20\degr$). A similar result was seen in M33 where vertical magnetic fields were suggested to actually be disk fields perturbed by a Galactic warp \citep{2008A&A...490.1005T}.
\subsection{Intrinsically Polarized Red Clump Stars}
One unexpected result of this study was the possible identification of intrinsically polarized red clump star candidates. These stars were photometrically identified as red clump stars, but their polarization properties (shown in Figures \ref{intrinsic} and \ref{outer_gal_uq}) were inconsistent with the trend of Stokes $U$ and $Q$ versus brightness (or distance) seen in the other red clump stars. The reason for intrinsic polarization is unclear from the data available, though there are several possibilities. Rapid rotation in old red clump stars is unlikely, but its presence could distort the star from spherical symmetry leading to a net polarization signature \citep[up to 1.7\%;][]{1968ApJ...151.1051H}. This rotation would be easily seen in moderate resolution spectroscopy and also allow identification of the star's rotation axis. The presence of a disk or envelope could also give rise to a net polarization \citep[up to 4\%;][]{1996ApJ...461..828W}. If true, the observed NIR colors would suggest that this is not a red clump star but possibly a reddened main sequence star that is observed in the same color-magnitude space as true red clump stars. Future variability studies or spectroscopy could shed light on the physical mechanism creating these apparently anomalous polarizations.
The method described above can be used to measure the mean interstellar polarization signal with distance for some direction, then easily distinguish polarimetric deviants from stars following the nominal polarization trend. This allows for the possibility of surveying for intrinsically polarized stars over large areas of the sky. Intrinsic polarization is associated with a range of physical phenomena that may not otherwise be seen in large-scale photometric surveys. By looking for polarization deviants, a catalog of interesting targets can be developed for follow-up investigation.
\section{Summary and Conclusions}
This work presents a new method for measuring the structure of the large-scale Galactic magnetic field. Photometrically identified red clump stars were used as standard candles along two example directions in the sky, one with few red clump stars and one with many red clump stars. The false-detection probability as a function of $K$-band brightness when using the method of \citet{2002AA...394..883L} was calculated and used to inform the reliability of the results. H-band starlight polarimetry of these stars from GPIPS provides information about the integrated magnetic field properties toward each star. By decomposing the Stokes $U$ and $Q$ parameters with distance, changes in the magnetic field orientation in the plane of the sky were measured with distance.
One line of sight, toward $(\ell , b)= (19.49\degr,+0.56\degr)$ in the inner Galaxy, shows evidence for a strong magnetic field perturbation at a distance of 2 kpc that cannot be attributed to dusty clouds along the line of sight. This is approximately the same distance as a magnetic field reversal seen in Faraday rotation surveys \citep{2008AA...477..573S, 2011ApJ...728...97V}, which are sensitive to the electron-weighted magnetic field. Formally, starlight polarimetry should not be sensitive to magnetic field reversals \citep{2011ApJ...740...21P}, but changes in large-scale magnetic field geometry could manifest as both NIR polarization P.A. changes and as magnetic field reversals in Faraday rotation studies. While this remains circumstantial evidence for the presence of a magnetic field reversal toward this direction, the overall agreement between Faraday rotation studies of the Galactic magnetic field and this method is intriguing.
A different line of sight, toward $(\ell , b)= (150\degr,-2.5\degr)$ in the outer Galaxy, shows a constant magnetic field orientation between distances of 2-5 kpc. The larger number of identified red clump stars allowed a more detailed analysis of the relatively simple magnetic field in this direction. The measured magnetic field orientation (GPA=$121\degr$) does not fit the plane-parallel model typically assumed for the Galactic plane. However, this value can be explained by the Galactic warp toward this direction in the Galaxy. From this analysis, it appears that the large-scale Galactic magnetic field toward this direction follows the Galactic warp.
An unexpected outcome of the analysis toward the outer Galaxy was the possible detection of two intrinsically polarized red clump stars. Photometrically, these stars fall along the red clump sequence, but their polarization properties do not match the growth of percentage polarization with distance exhibited by the remaining stars. The deviations of the NIR polarization observations of these stars to the fits of the other stars' Stokes parameters allowed determination of these intrinsic polarizations. Moderate resolution spectra (to test whether these are red clump stars and to measure $v$ sin $i$) would provide additional insight into the physical natures of these objects.
This work represents a first step in decomposing with distance the magnetic fields probed by background starlight polarimetry. Additional work is required to fully develop this method. Photometric methods can statistically identify red clump stars; however, spectroscopic follow-up is needed. Also, while the relative distances, or at least distance order, of stars derived from photometry are reliable, physical distances require spectrophotometric follow-up. The future application of photometrically identified red clump stars to large, uniformly sampled background starlight polarimetry will allow the large-scale Galactic magnetic field to be unambiguously disentangled with distance for the first time.
\acknowledgements
The author would like to thank Dan P. Clemens, Brian W. Walsh, Kamen Kozarev, Daniel Jaffe, and Neal Evans for useful discussions. The author would also like to thank the anonymous referee for comments that greatly improved the manuscript. The author gratefully acknowledges financial support from McDonald Observatory. This research used data from the Boston University (BU) Galactic Plane Infrared Polarization Survey (GPIPS), funded in part by NSF grants AST 06-07500 and 09-07790. GPIPS used the Mimir instrument, jointly developed at BU and Lowell Observatory and supported by NASA, NSF, and the W.M. Keck Foundation.
|
\section{Introduction}
The B-factory experiments Belle and \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}} established CP violation in the neutral and charged B meson system, and experimentally confirmed
the source of CP violation to be one single complex phase in the three-family CKM quark mixing matrix~\cite{Belle2001,BaBar2001,Belle2004c,BaBar2004c}. The Belle and \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}} experiments
provide an excellent environment to study heavy-flavor decays. Most importantly the clean environment and the Lorentz boost caused by the asymmetric-energy $e^+e^-$ colliders KEKB and PEP-II,
and the quantum entanglement of neutral B mesons produced by $\Upsilon(4S)$ decays enable measurements of observables sensitive to the fundamental symmetries CP, T and CPT.
At the conference Flavor Physics \& CP Violation 2014 the current most precise measurements sensitive to the breaking of these symmetries have been presented, and compared to previous related results.
In this proceedings article the presented measurements are summarized, and references for further reading are provided.
\section{CPT Measurement by Belle}
The invariance under CPT, the combined operation of charge-conjugation C, parity-transformation P and time-reversal T, is of fundamental importance in physics. All local Lorentz-invariant
quantum field theories are assumed to conserve CPT~\cite{Lueders,Pauli,Greenberg}. A breaking of CPT symmetry would have important consequences such as the possibility for violation of
Lorentz invariance, or differences in lifetimes and masses of particles and corresponding antiparticles. Searches for the violation of CPT symmetry have been carried out before
in the neutral kaon system by the CPLEAR, KLOE and KTeV collaborations~\cite{CPLEAR2001,KLOE2006,KTeV2011}, and in the neutral B meson system by the Belle and \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}} experiments~\cite{Belle2003,BaBar2004a,BaBar2006}.
An observable sensitive to CPT violation in neutral B meson mixing is the complex parameter $z$, which is related to the light ($B_{L}$) and heavy ($B_{H}$) mass eigenstates and the complex mixing parameters $p$ and $q$ by:
\begin{eqnarray}
| B_{L} \rangle \propto& p \sqrt{1 - z} | B^{0} \rangle | + q \sqrt{1 + z} | \bar{B}^{0} \rangle \nonumber \\
| B_{H} \rangle \propto& p \sqrt{1 + z} | B^{0} \rangle | - q \sqrt{1 - z} | \bar{B}^{0} \rangle \nonumber
\end{eqnarray}
In 2012, Belle performed a measurement of the CPT violating parameter $z$ based on a data sample containing $535 \times 10^6$ $B\bar{B}$ pairs collected on the $\Upsilon(4S)$~\cite{Belle2012}.
The time-dependent analysis utilizes the coherent mixing of two entangled neutral B mesons in an $\Upsilon(4S)$ event~\cite{BaBar2004a,BaBar2004b}. The measurement reconstructs one B meson either in CP eigenstates,
or in flavor-specific hadronic or semileptonic decay modes. The reconstructed CP modes $B^0 \ensuremath{\rightarrow}\xspace J/\psi K^{0}_{S}$ and $B^0 \ensuremath{\rightarrow}\xspace J/\psi K^{0}_{L}$ provide sensitivity mainly on $Re(z)$, while others are sensitive to $Im(z)$.
Besides the measurement of $Re(z)$ and $Im(z)$, the analysis extracts $\Delta \Gamma_{d} / \Gamma_{d}$ from an unbinned maximum likelihood fit of the reconstructed data containing about $560k$ events
to the full decay rate of the entangled B meson pairs in an $\Upsilon(4S)$ event. The obtained results are:
\begin{eqnarray}
Re ( z ) =& \left[ +1.9 \pm 3.7 \, (\rm{stat}) \pm 3.3 \, (\rm{syst}) \right] \times 10^{-2} \nonumber \\
Im ( z ) =& \left[ -5.7 \pm 3.3 \, (\rm{stat}) \pm 3.3 \, (\rm{syst}) \right] \times 10^{-3} \nonumber \\
\frac{\Delta \Gamma_{d}}{\Gamma_{d}} =& \left[ -1.7 \pm 1.8 \, (\rm{stat}) \pm 1.1 \, (\rm{syst}) \right] \times 10^{-2} \nonumber
\end{eqnarray}
The results are consistent with the assumption of no CPT violation, and provide the current most precise estimation of observables sensitive to CPT violation in the neutral B meson system.
\section{T Violation Measurement by \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}}
To compensate for CP violating effects the invariance under CPT of Standard Model physics processes predicts T violating phenomena.
To observe T violation by rate comparisons, initial and final states of physical processes need to be exchanged.
Performing this exchange by the reversal of reactions of unstable particles is difficult to realize in experiments. In 2012, \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}} performed a measurement utilizing the coherent mixing of
two neutral B mesons in an $\Upsilon(4S)$ event
to probe for time-reversal invariance by measuring the rate differences of processes with exchanged initial and final states~\cite{BaBar2012}.
The measurement follows an idea by Banuls and Bernabeu~\cite{Banuls1999, Banuls2000} and applies a procedure called CP tagging in analogy to the flavor-tagging in standard time-dependent CP violation measurements.
This enables the inference of the CP state of one B meson at the instant of decay of the second B meson.
The approach allows the construction of proper states and the measurement of the rates of reactions related by T transformation.
For example, the transition of $\bar{B}^{0} \ensuremath{\rightarrow}\xspace B_{-}$ where $B_{-}$ denotes a CP-odd eigenstate is identified by the time-ordered final states $(l^{+}, J/\psi K^{0}_{S})$.
The related T reversed process $B_{-} \ensuremath{\rightarrow}\xspace \bar{B}^{0}$ is obtained by the final states $(J/\psi K^{0}_{L}, l^{-})$.
The measurement provides three further independent rate comparisons between $B_{+} \ensuremath{\rightarrow}\xspace B^{0}$ $(J/\psi K^{0}_{S}, l^{+})$, $\bar{B}_{0} \ensuremath{\rightarrow}\xspace B_{+}$ $(l^{+}, J/\psi K^{0}_{L})$, and
$B_{-} \ensuremath{\rightarrow}\xspace B^{0}$ $(J/\psi K^{0}_{L}, l^{+})$ and their corresponding T-conjugated transitions. Any difference in the rates of these pairs manifests the breaking of T symmetry.
Furthermore the measurement needs no assumptions about CP violation or CPT conservation as input, but measures asymmetry parameters sensitive to CP and CPT in addition to that sensitive to T violation.
For the main T violating parameters, \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}} measures using a data sample containing $468 \times 10^6$ $B\bar{B}$ pairs collected on the $\Upsilon(4S)$:
\begin{eqnarray}
\Delta S^{+}_{T} =& -1.37 \pm 0.14 \, (\rm{stat}) \pm 0.06 \, (\rm{syst}) \nonumber \\
\Delta S^{-}_{T} =& 1.17 \pm 0.18 \, (\rm{stat}) \pm 0.11 \, (\rm{syst}) \nonumber
\end{eqnarray}
The result directly observes T violation with a significance of $14\sigma$, and is in agreement with the assumption of CPT conservation and CP violation.
\begin{figure}[htb]
\centering
\includegraphics[width=0.7\textwidth]{figure_HFAG_average.pdf}
\caption{Comparison and average of $A_{SL}^{s}$ and $A_{SL}^{d}$ measurements provided by the Heavy Flavor Averaging Group (HFAG)~\cite{HFAG}.}
\label{fig:HFAG_average}
\end{figure}
\section{CP Violation in Mixing Measurement by \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}}}
CP violation in neutral B meson mixing provides a sensitive probe for models beyond the Standard Model. New particles might emerge in the processes mediated by box diagrams and modify the mixing asymmetry defined by~\cite{Lenz2007}:
\[
A_{SL}^{q} = \frac{ P \left( \bar{B}^{0}_{q} \ensuremath{\rightarrow}\xspace B^{0}_{q} \left(t\right) \right) - P \left( B^{0}_{q} \ensuremath{\rightarrow}\xspace \bar{B}^{0}_{q} \left(t\right) \right) }{ P \left( \bar{B}^{0}_{q} \ensuremath{\rightarrow}\xspace B^{0}_{q} \left(t\right) \right) + P \left( B^{0}_{q} \ensuremath{\rightarrow}\xspace \bar{B}^{0}_{q} \left(t\right) \right) }
= \frac{1 - |q/p|^{4}}{1 + |q/p|^{4}} \approx \frac{|\Gamma_{12}^{q}|}{|M_{12}^{q}|} \sin \phi_{q}
\]
The Standard Model predictions for the neutral $B_d$ meson system are $A_{SL}^{d} = (1.8 \pm 0.3) \times 10^{-5}$ and $\phi_{d} = (0.24 \pm 0.06)^{\circ}$~\cite{Nierste2012}. The current achievable experimental sensitivity
on $A_{SL}^{q}$ is $\mathcal{O}(10^{-3})$. Any significant deviation from 0 might point to possible beyond the Standard Model processes. An important motivation for measurements sensitive to CP violation in B mixing comes from the tension driven
by the results of the D0 experiment on the inclusive like-sign dimuon charge asymmetry~\cite{Dzero2014}.
In 2013, \mbox{\sl B\hspace{-0.4em} {\small\sl A}\hspace{-0.37em} \sl B\hspace{-0.4em} {\small\sl A\hspace{-0.02em}R}} performed a search for CP violation in $B^{0}$-$\bar{B}^{0}$ mixing using a data sample containing $468 \times 10^6$ $B\bar{B}$ pairs collected on the $\Upsilon(4S)$~\cite{BaBar2013}.
The measurement reconstructs one B meson as $B^{0} \ensuremath{\rightarrow}\xspace D^{*-} X l^{+} \nu$ by applying a partial reconstruction technique for $D^{0}$ mesons from $D^{*-} \ensuremath{\rightarrow}\xspace D^{0} \pi^{-}_{\mathrm{soft}}$ decays.
The flavor of the reconstructed B meson is inferred from the lepton charge, and the flavor of the second B meson is inferred from its decay products by employing a kaon tag.
A difficulty of the measurement is the correction for effects induced by physical and detector charge asymmetries. The strategy of the measurement is to avoid relying on control samples or Monte Carlo simulations, and to
estimate these effects directly from data by exploiting all available information from different reconstructed sub-samples accounting for electron or muon leptons, and mixed or unmixed neutral B events.
The signal fraction is estimated from a fit to distributions of the missing neutrino mass, and a yield of $(5.945 \pm 0.007) \times 10^6$ signal events obtained. The $A_{SL}^{d}$ asymmetry is obtained
by a binned four-dimensional fit to the $\Delta t$, $\sigma(\Delta t)$, $\cos(\theta_{lk})$ and $p_k$ distributions. The result of the measurement is:
\begin{eqnarray}
| \frac{q}{p} - 1 | =& ( -0.29 \pm 0.84 \, (\rm{stat}) \, ^{+1.88}_{-1.61} \, (\rm{syst}) ) \times 10^{-3} \nonumber \\
A_{SL}^{d} =& ( 0.06 \pm 0.17 \, (\rm{stat}) \, ^{+0.38}_{-0.32} \, (\rm{syst}) ) \% \nonumber \nonumber
\end{eqnarray}
The measurement provides the most precise single result on $A_{SL}^{d}$, and is in agreement with the Standard Model predictions. A comparison of results from $e^+e^-$ and hadron colliders provided by the Heavy Flavor Averaging Group is shown in Figure~\ref{fig:HFAG_average}.
|
\section*{Acknowledgements}
\label{sec:Acknowledgements}
This research was funded by Sullivan Nicolaides Pathology, Australia and the Australian Research Council (ARC) Linkage Projects Grant LP130100230.
\section*{Appendix}
\label{app:appendix}
\label{proof:ARCAD}
\noindent
We show the proof of Proposition.~\ref{prop:1} by deriving Eqn.~\ref{eqn:PiCoDes} from Eqn.~\ref{eqn:ARCAD}.
We first rewrite Eqn.~\ref{eqn:ARCAD}:
\begin{small}
\begin{equation}
\label{app:ARCAD}
\min_{\Mat{w}_{1 \dots K}, \Vec{b}_{1 \dots K}, \Mat{A}_{1 \dots J}}
\sum_{k=1}^K \left\{ \frac{1}{2} \| \Mat{w}_k \|^2 + \frac{\lambda}{N} \sum_{i=1}^N \ell \left[ y_{i,k} (\Vec{b}_k +
\sum^{J}_{j=1} \left( \frac{\Vec{w}_{k,j}}{N_{i,j}} \sum^{N_{i,j}}_{c=1} \Mat{A}^{\top}_j \Vec{x}_{i,j,c} \right) \right] \right\}
\end{equation}
\end{small}
\noindent
This equation can be rewritten as:
\begin{small}
\begin{equation}
\min_{\Mat{w}_{1 \dots K}, \Vec{b}_{1 \dots K}, \Mat{A}_{1 \dots J}}
\sum_{k=1}^K \left\{ \frac{1}{2} \| \Mat{w}_k \|^2 + \frac{\lambda}{N} \sum_{i=1}^N \ell \left[ y_{i,k} (\Vec{b}_k +
\sum^{J}_{j=1} \left( \Vec{w}_{k,j} \Mat{A}^{\top}_j \left( \frac{1}{N_{i,j}} \sum^{N_{i,j}}_{c=1} \Vec{x}_{i,j,c} \right) \right) \right] \right\}
\end{equation}
\end{small}
\noindent
Let $\hat{\Vec{x}}_{i,j}$ be the average of cell-level descriptor extracted from $j$-th region
$\hat{\Vec{x}}_{i,j} = \frac{1}{N_{i,j}} \sum^{N_{i,j}}_{c=1} \Vec{x}_{i,j,c}$, the equation can be transformed into:
\begin{small}
\begin{equation}
\min_{\Mat{w}_{1 \dots K}, \Vec{b}_{1 \dots K}, \Mat{A}_{1 \dots J}}
\sum_{k=1}^K \left\{ \frac{1}{2} \| \Mat{w}_k \|^2 + \frac{\lambda}{N} \sum_{i=1}^N \ell \left[ y_{i,k} (\Vec{b}_k +
\sum^{J}_{j=1} \left( \Vec{w}_{k,j} \Mat{A}^{\top}_j \hat{\Vec{x}}_{i,j} \right) \right] \right\}
\end{equation}
\end{small}
\noindent
Finally, this can be rewritten into Eqn.~\ref{eqn:PiCoDes}, with $\Vec{u}_i = [\hat{\Vec{x}}_{i,j} \dots \hat{\Vec{x}}_{i,J}]$:
\begin{small}
\begin{equation}
\min_{\Mat{w}_{1 \dots K}, \Vec{b}_{1 \dots K}, \Mat{A}_{1 \dots J}}
\sum_{k=1}^K \left\{ \frac{1}{2} \| \Mat{w}_k \|^2 + \frac{\lambda}{N} \sum_{i=1}^N \ell \left[ y_{i,k} (\Vec{b}_k +
\sum^{P}_{p=1} \left( w_{k,p} \Mat{A}^{\top}_j \Vec{u_i} \right) \right] \right\}
\end{equation}
\end{small}
\begin{center}
\begin{figure*}[!tb]
\includegraphics[width=\linewidth]{results_03.pdf}
\caption {The generated text description for each pattern.
The correct description is depicted in green colour and without bracket, whilst the text in square brackets and parentheses indicate the omitted and incorrect description, respectively.}
\label{fig:result_03}
\end{figure*}
\end{center}
\section{Main Findings}
\label{sec:conclusions}
The ANA test via Indirect Immunofluoresence protocol has been the gold standard to identifying Connective Tissue Diseases.
Unfortunately the protocol is subjective, time as well as labour intensive.
Despite the growing interest in this domain, prior works have primarily focused on classifying cell images extracted from specimen images.
In this work, we took a further step by addressing the specimen image classification problem.
To this end, we designed a specimen-level image descriptor to be: highly discriminative; short in descriptor length as well as semantically meaningful at cell level.
We achieved this goal by proposing two learning schemes which are based on the max-margin framework.
We later showed that under a certain condition, discovering such an image descriptor is equivalent to discovering discriminative image-level attributes.
We contrasted the proposed approaches to numerous hashing techniques as well as discriminative attribute learning approaches on a new HEp-2 cell dataset.
We found that the descriptors trained by the proposed schemes outperform the existing approaches with a much shorter code length.
The experiments also show that the proposed approaches outperform the recent approaches presented in~\cite{Soda2009,Li2013,Wiliem2013}.
Furthermore, we found that the descriptors can be used to provide a textual description of ANA patterns based on the cell attributes.
Despite the fact that the proposed approaches are able to discover semantically meaningful cell-level attributes, the constraints used to achieve this are not explicitly imposed in the proposed max-margin based training schemes.
As such, we will explore ways to make the constraints more explicit and study the effect of the constraints on the learned descriptor.
\section{Experiment}
\label{sec:experiment}
We first describe the novel {specimen-level} HEp-2 cell image classification dataset and experiment settings.
The proposed approaches are compared to various notable hashing techniques~\cite{charikar2002similarity,kulis2009kernelized,gong2012,weiss2008nips} as well as recent discriminative attribute learning methods~\cite{RastegariECCV2012,Bergamo2011}.
The proposed approaches are also compared to the existing methods for {specimen-level} image classification in this domain such as the Multiple Expert System (MES)~\cite{Soda2009}.
Finally, we showcase the ability of the discovered {cell-level} attributes to help physicians in describing the ANA pattern classes.
\subsection{Dataset}
We propose a new dataset~\footnote{Available at \url{http://www.itee.uq.edu.au/sas/datasets}}
to serve our goal because the existing benchmarking datasets such as SNPHEp-2~\cite{Wiliem2013_wacv},
ICIP2013~\footnote{\url{http://nerone.diiie.unisa.it/contest-icip-2013/index.shtml}} and
ICPR2012Contest~\cite{Cordelli2011} are primarily focussed on the {cell-level} classification problem.
Although the ICPR2012Contest dataset provides specimen images,
the number of images for each train and test are insufficient for this study
(\textit{i.e.}~ only 14 images for each train and test classified into six classes).
The proposed dataset was obtained from 262 patient sera in 2013.
The patient sera were diluted to 1:80 dilution.
The prepared specimen was then photographed using a monochrome
high dynamic range cooled microscopy camera fitted on a microscope.
In total there were 262 specimen images classified into eight classes: homogeneous, speckled, nucleolar, centromere, nuclear membrane (NuMem), cell cycle dependent (CCD), mitotic spindle (MitSp) and golgi aparatus (golgi).
Each specimen image contains a collection of interphase and mitotic cells which are important for the specimen-level classification (refer to Fig.~\ref{fig:dataset}).
The mitotic cells are the HEp-2 cells undergoing the mitosis phase.
In this phase each cell divides itself into two separate individual cells.
The mitotic cells are of importance as they produce a set of antigens which are either less concentrated or undetectable in the interphase stage.
Due to this fact, experts consider the pattern of both interphase and mitotic cells in classifying an ANA specimen.
Five-fold validations of training and test were created by randomly selecting from the pool of images.
Specifically, we randomly selected approximately 130 images from the image pool for each fold.
Then, the selected images were further divided into two equally sized sets for training and testing (\textit{i.e.}~ approximately 75 images for each set).
To our knowledge this is the first dataset to offer a
reasonable number of data samples for benchmarking
{specimen-level} classification approaches. We plan to release the dataset by the end of 2014.
\subsection{Experiment settings}
\label{sec:experiment_settings}
Although one could use any {cell-level} descriptor,
in the present work, we use the bag of words descriptor with Cell Pyramid Matching (CPM) structure recently proposed in~\cite{Wiliem2013} due to its robustness to various laboratory settings.
Specifically, we divide each cell into an inner region,
which covers the cell content and an outer region, which contains information related to cell edges and shape.
In addition, a descriptor extracted from the whole cell region is also used.
Thus, there are three regional descriptors extracted from each cell.
As there are two cell types: interphase and
mitotic cells, we use six regions in total (three regions of each type).
Note that, we only consider the regional descriptors extracted from the same cell type when computing $\hat{\Vec{x}}_{i,j}$ as well as $\hat{\Vec{h}}_{i,j}$.
Once the regional descriptors are extracted,
we lift up the each descriptor into a higher-dimensional space approximating the histogram intersection kernel by using the explicit feature maps proposed by Vedaldi and Zisserman~\cite{vedaldi11efficient}.
Specifically we map the original regional descriptor into the space three times larger by setting the parameter $n$ in~\cite{vedaldi11efficient} to 1.
We opt to use the lifted space of the histogram intersection
kernel as bag of words histogram comparisons are generally best to be done under histogram intersection kernel as suggested in~\cite{Bergamo2011}.
The lifted-up descriptor will be the regional descriptor for each $\Vec{x}_{i,j,c}$.
Although it is possible to vary the number of attributes extracted from each region, we choose to use the same number of attributes for all regions in order to reduce the total number of combinations.
\subsection{Evaluation of different descriptors}
We first contrast the proposed approaches to the recent hashing methods such as: Locality-Sensitive Hashing (LSH)~\cite{charikar2002similarity}, Kernelised LSH (KLSH)~\cite{kulis2009kernelized}, Spectral Hashing (SPH)~\cite{weiss2008nips} and Iterative Quantization (ITQ)~\cite{gong2012}.
In addition we also compare the proposed approaches to the recent discriminative attributes learning: PiCoDeS~\cite{Bergamo2011} and Discriminative Binary Code (DBC)~\cite{RastegariECCV2012}.
All methods were trained and tested using the {specimen-level} descriptor extracted with the setup previously described.
We used the unlifted regional descriptor in conjunction with the histogram intersection kernel for KLSH.
The results are presented in Fig.~\ref{fig:result_01}.
Both ARCAD and CRAD outperform all other methods even when small numbers of attributes are used.
This suggests that the proposed approaches are highly discriminative even with small code length.
The learning schemes successfully discover the essential cell attributes which are highly discriminative to form specimen image descriptor.
We note that both LSH and KLSH require longer code length to achieve similar performance to the proposed approaches.
This is consistent with the finding reported in previous works suggesting that LSH requires longer code length to achieve good performance~\cite{kulis2009kernelized}.
In our case having a small code length is advantageous since the experts require less time to name the discovered attributes.
It is noteworthy to mention that the proposed approaches outperform significantly PiCoDeS as well as DBC which were specifically designed to discover the discriminative attributes.
Both PiCoDeS and DBC consider attribute value in binary space.
Furthermore, the DBC learning scheme operates in the binary space which is intrinsically more complex.
In our case, we consider real attribute values (Refer Eqn.~\ref{eqn:h_j}).
In the light of this fact, we argue that presenting attribute
value as a the real number is more expressive than a binary value (~\textit{i.e.}~ 0 or 1).
\begin{figure}[!tb]
\centering
\includegraphics[width=0.7\columnwidth]{results_01.pdf}
\caption {Performance comparison of the proposed approaches (ARCAD and CRAD) to the existing methods; LSH: Locality-Sensitive Hashing~\cite{charikar2002similarity}; KLSH: Kernelised LSH~\cite{kulis2009kernelized}; SPH: Spectral Hashing~\cite{weiss2008nips};ITQ: Iterative Quantization~\cite{gong2012}; DBC: Discriminative Binary Code~\cite{RastegariECCV2012}.}
\label{fig:result_01}
\end{figure}
\begin{figure}[!tb]
\centering
\includegraphics[width=0.7\columnwidth]{results_02.pdf}
\caption {Performance comparison of the proposed approaches (ARCAD and CRAD) to the state of the arts; MES: Multiple Expert Systems~\cite{Soda2009}; Object bank proposed in~\cite{Li2013}.}
\label{fig:result_02}
\end{figure}
In the second evaluation we contrasted the proposed approaches to existing methods for HEp-2 specimen image classification.
The most common approach for classifying specimen images is to use the dominant pattern of the interphase cells~\cite{foggia2013benchmarking,Wiliem2013,
Faraki2013,Yang2013}.
Here, we call this \textit{baseline}.
Another approach, here denoted Multiple Expert System (MES), is to train individual classifier for each class (\textit{i.e.}~ one interphase cell classifier for each class) and use the classification reliability score to do weighted voting~\cite{Soda2009}.
We implemented both \textit{baseline} and the MES approach.
We also implemented the approach in~\cite{Li2013} which proposes the concept of Object Bank.
Technically, we train k one-versus-all classifiers.
Given a cell image, we apply all the k classifiers and consider the k classification output scores as the object bank representation of the cell.
The object bank representation of a specimen image is obtained by averaging the {cell-level} object bank representation.
Here, we trained two sets of classifiers: (1) eight classifiers trained on interphase cells (\textit{i.e.}~ one classifier for each pattern class), denoted \textit{Object bank interphase}; (2) sixteen classifiers consisting of eight classifiers trained on the interphase cells and eight on the mitotic cells, denoted \textit{Object bank both}.
Fig.~\ref{fig:result_02} presents the evaluation results.
The proposed approaches significantly outperform all other methods.
We note that the MES has poor performance which contradicts what was reported in~\cite{Soda2009}.
Upon a closer look we found that some classes such as Mitotic Spindle do not have specific characteristics on their interphase cells.
Henceforth, it is difficult to train a reliable cell classifier rendering much lower reliability score.
In other hand, traditional voting (\textit{i.e.}~ the baseline) has much better performance as it only counts the vote and does not consider the classification reliability score.
Furthermore, we found that combining the information extracted from both interphase and mitotic cells is of importance.
This can be observed from the fact that there is a significant increase in Object bank performance when information from both cell types is used.
\vspace{-1ex}
\subsection{Describing ANA pattern class}
In this section we use the attributes discovered by the proposed approaches to generate a textual description of the eight ANA patterns.
To that end, we use attributes trained by CRAD as it gives the most consistent results in the previous evaluation.
Specifically, we opt to use 32 attributes extracted from inner and outer regions from both cell types.
We first selected the most frequently appearing {cell-level} attributes from each pattern.
From the selected set, we further excluded the attributes which appear in at least more than four classes.
Finally, to name the cell attributes, we presented each cell attribute to the domain experts who were trained to read ANA by showing them both images classified as positive and negative by the attribute classifier.
We note that we presented the cell images in green colour which is similar to the colour of an ANA specimen under a fluorescent microscope.
Since the attributes are extracted from each cell region, we could ask more specific questions to the experts in relation to each region (\eg \textit{Please describe the property appearing at the cell boundary}).
The experts could opt not to name an attribute if they were not able to find any consistent property in the positive cell images.
Fig.~\ref{fig:result_04} and~\ref{fig:examples} present some examples of cell attributes successfully identified by the experts.
Once the description for each class was generated, we let the experts indicate the correctness of each text description.
Fig.~\ref{fig:result_03} presents the generated description of each pattern.
Most patterns could be reasonably described with minor errors or omissions in the description.
The mitotic spindle pattern was perfectly described with no errors or omissions in the description.
On the other hand, despite this system being able to detect the important property of Golgi (\textit{i.e.}~ golgi organelle is stained), the system had more mistakes on Golgi than the other patterns.
This is probably due to the fact that the Golgi pattern has only one prominent property.
\begin{figure}[tb]
\centering
\includegraphics[width=0.7\columnwidth]{results_04.pdf}
\caption {Example of successfully described attributes. from top row to bottom: \textit{parts of chromosome are stained}, \textit{golgi organelle is stained}, \textit{mitotic spindle staining}.}
\label{fig:result_04}
\end{figure}
\vspace{-2ex}
\section{Introduction}
\begin{figure}[!tb]
\centering
\includegraphics[width=0.5\columnwidth]{fig1.pdf}
\caption {Examples of {cell-level} discriminative attributes found by our proposed approach for some ANA patterns.
Each pattern is described using two cell types: interphase cell (left); mitotic cell (right).}
\label{fig:examples}
\end{figure}
The application of image analysis for various routine clinical pathology tests has been growing in recent years~\cite{foggia2013benchmarking,
labati2011lymphoblastic}.
When incorporated into subjective analysis from scientists, these can potentially not only lower test turn around time but also increase test result reliability and consistency across laboratories~\cite{foggia2013benchmarking,Wiliem2011}.
The common way of identifying the existence of connective tissue diseases is via the Anti-Nuclear Antibody (ANA) test using Indirect Immunofluorescence (IIF) protocol on Human Epithelial type 2 (HEp-2) cells~\cite{meroni2010ana}.
This is due to its high sensitivity and the large range expression of antigens.
Unfortunately, the protocol is time consuming, labour intensive and subjective~\cite{Bizzaro1998,Pham2005} leading to low reproducibility and large inter/intra- personnel/laboratory variations~\cite{Soda2009}.
One possible solution is to apply Computer Aided Diagnostic (CAD) systems for automated classification of ANA IIF digitally captured images.
Despite the large interest shown recently in the literature, most of the existing works only focus on the early steps of the CAD system, that is {cell-level} classification~\cite{foggia2013benchmarking,Wiliem2011,Wiliem2013_wacv,Wiliem2013,Faraki2013,Yang2013}.
Whilst, some methods which go beyond this scope assume that the {specimen-level} pattern can be simply estimated from the most dominant {cell-level} pattern~\cite{Soda2009}.
As each {specimen-level} image consists of a set of cells, ideally a CAD system should be able to extract more useful information from the cells distribution to infer the specimen image pattern (refer to Fig.~\ref{fig:dataset} for examples of specimen images).
Furthermore, the existing systems do not provide meaningful information as to why an ANA pattern differs to the others.
Having meaningful information such as textual description is of interest in this area since often a pattern has various descriptions amongst physicians~\cite{wiik2010ana}.
One way to address these issues is to learn discriminative and semantically meaningful descriptors.
Each element of the descriptor defines the existence/absence of a specific inherent property/characteristics in an image.
For instance, an image containing a car may have some properties such as \textit{has a wheel}, \textit{is metallic}, \textit{has windows and doors}~\cite{Ferrari07}.
These properties are popularly known as image attributes~\cite{Ferrari07}.
There is a growing interest to develop attribute-based approach for image classification~\cite{Ferrari07,
LampertCVPR2009,
Bergamo2011,
RastegariECCV2012,
parikh2009}.
For instance, Ferrari and Zisserman proposed a probabilistic generative model of visual attributes~\cite{Ferrari07}.
Lampert~\etal reported excellent results of the Direct Attribute Prediction (DAP) approach in a \textit{zero-shot} learning problem~\cite{LampertCVPR2009}.
Parikh~\etal extended the notion of image attribute into relative attribute which is related to adjective of a noun such as \textit{larger} and \textit{more open space}~\cite{parikh2009}.
An image attribute detector is essentially a binary classifier which determines the presence or absence of an image property.
As such, each attribute needs a training set which may be expensive to acquire for our problem domain due to the limited number of domain experts.
It is also almost impossible to use the \textit{Amazon Mechanical Turk} service~\cite{parikh2009,LampertCVPR2009} to acquire the training labels.
Therefore, it is desirable to automatically discover the smallest set of discriminative image attributes wherein the domain experts can name them.
To that end, one could apply ideas proposed in~\cite{Bergamo2011,RastegariECCV2012} to discover discriminative image attributes.
Here, the attribute classifiers are jointly learned with the image classifier in the max-margin framework.
Nevertheless, these approaches are mainly focussed on image attributes, whereas in our work, the system needs to discover {cell-level} attributes which, when summarised into the {specimen-level} descriptor, will be highly discriminative.
Henceforth, the cell attributes must be indirectly discovered via learning a discriminative specimen image descriptor.
The concept of discovering discriminative image attributes is also related to hashing techniques~\cite{charikar2002similarity,kulis2009kernelized,gong2012,weiss2008nips}.
However, the descriptors resulting from these techniques are not necessarily semantically meaningful.
For instance the spectral hashing approach aims to generate a set of binary codes which are short, easy to compute and maps similar items to similar binary codewords~\cite{weiss2008nips}.
To that end, the approach constraints that each bit has a 50\% chance of being zero and one and different bits independent to each other.
These are much weaker constraints to discover semantically meaningful code.
\begin{figure}[!tb]
\includegraphics[width=\columnwidth]{fig_dataset.pdf}
\caption {Sample images from the proposed dataset}
\label{fig:dataset}
\end{figure}
\textbf{Contributions} The aim of the present work is to devise an algorithm which learns image descriptors for ANA IIF specimen image classification problem with three properties: (1)~highly discriminative; (2)~semantically meaningful at cell level and (3)~having short descriptor length.
We achieve this by proposing two learning schemes for discovering {cell-level} attributes through a discriminative learning framework.
In contrast to previous approaches~\cite{Bergamo2011,RastegariECCV2012} which learns discriminative {image-level} attributes, our approach learns {cell-level} attributes where their values can be used to construct discriminative {image-level} descriptors.
Our theoretical results show that under a certain condition, it is possible to devise solutions based on {image-level} discriminative attribute learning for solving the posed problem.
Finally, we further showcase (refer to Fig.~\ref{fig:examples} for some examples of discovered meaningful cell attributes) that a textual description can be generated from the learned {cell-level} attributes and shares similarities to the description from experts.
We evaluated all the approaches on the new HEp-2 cell dataset proposed for the specimen image classification problem.
To our knowledge this is the first comprehensive dataset constructed for this purpose.
We continue our discussion as follows.
Section~\ref{sec:problem_def} discusses the ANA IIF {specimen-level} classification problem.
The proposed learning schemes are described in Section~\ref{sec:proposed_approach}.
We present the experiment and results in Section~\ref{sec:experiment}.
Finally, the main findings and future direction are discussed in Section~\ref{sec:conclusions}.
\section{Problem definition}
\label{sec:problem_def}
An ANA IIF specimen image $\Mat{I}$ is represented by the three-tuple $\{\Mat{I}, \Mat{M}, \delta\}$ which consists of: \textbf{(i)} the Fluorescein Isothiocyanate (FITC) image channel which carries pattern information $\Mat{I}$; \textbf{(ii)} a binary cell mask image $\Mat{M}$ which are extracted from the \textit{4',6-diamidino-2-phenylindole} (DAPI) image channel; \textbf{(iii)} the fluorescence intensity $\delta = \{\operatorname{weak}, \operatorname{strong} \}$.
The goal is to construct a classifier which classifies a specimen image into one of the known classes.
Our problem differs from~\cite{foggia2013benchmarking,Wiliem2013_wacv,Wiliem2013,
Faraki2013,Yang2013} in the way that these works focus on classification of individual cell images extracted from specimen images; our main attention is on the specimen image classification problem.
\section{Discovering cell attributes}
\label{sec:proposed_approach}
The goal of the present work is to discover {cell-level} attributes which can be used to form a discriminative {specimen-level} descriptor.
Once all cells are extracted from a specimen image using its mask, each cell image is divided into J regions from which a regional descriptor is extracted.
For clarity, we defer the discussion of how the regions are divided until Section~\ref{sec:experiment_settings}.
After the cell has been divided into regions, we derive the {cell-level} attributes from the extracted regional descriptor.
The specimen image descriptor is then formed by concatenating the overall cell attributes from all regions.
Let $\Vec{z}_i \in \mathbb{R}^P$ be the $P$ dimensional {specimen-level} descriptor of the $i$-th specimen image, $\Vec{z}_i = [\hat{\Vec{h}}_{i,1} \dots \hat{\Vec{h}}_{i,J}]$.
Each $\hat{\Vec{h}}_{i,j} \in \mathbb{R}^b$ represents the overall {cell-level} attribute descriptor extracted from the $j$-th region:
\vspace{-1ex}
\begin{equation}
\label{eqn:average_h_j}
\hat{\Vec{h}}_{i,j} = \frac{1}{N_{i,j}} \sum^{N_{i,j}}_{c=1}{\Vec{h}_{i,j,c}}
\end{equation}
\vspace{-1ex}
\noindent
where $N_{i,j}$ is the number of cells extracted from the specimen image for $j$-th region; $\Vec{h}_{i,j,c} \in \mathbb{R}^b$ is the {cell-level} attribute descriptor extracted from the $j$-th region in the $c$-th cell; $b$ is the number of cell attributes extracted from $j$-th region.
The above equation suggests that the specimen image is represented by the average of the {cell-level} attributes.
This approach differs from~\cite{Soda2009,foggia2013benchmarking} wherein each image is represented as the dominant pattern of extracted cells.
We will later show in the experiment that our strategy is considerably more effective.
Inspired from~\cite{Bergamo2011}, we define the value of each element of the {cell-level} attribute descriptor $\Vec{h}_{i,j,c}$ as the output of a set of basis linear classifiers as follows:
\vspace{-2ex}
\begin{equation}
\label{eqn:h_j}
\Vec{h}_{i,j,c} = \Mat{A}_j^{\top}\Vec{x}_{i,j,c}
\end{equation}
\vspace{-1ex}
\noindent
where each column of $\Mat{A_j} \in \mathbb{R}^{d \times b}$ is the model parameter for a single basis classifier; $\Vec{x}_{i,j,c} \in \mathbb{R}^d$ is the regional descriptor extracted from $j$-th region in $c$-th cell.
The value of each element in $\Vec{h}_{i,j,c}$ indicates the presence ($+$) or absence ($-$) of a particular {cell-level} attribute.
Here, each attribute classifier will be trained by maximising the margin between the positive and negative samples,
thereby, imposing an indirect constraint that an attribute should provide a meaningful concept (\textit{i.e.}~ the concept discriminating positive and negative samples).
We learn both the {image-level} classifier and the {cell-level} attribute basis classifiers simultaneously over all variables.
To this end, we propose two learning schemes via the one-versus-all linear SVM framework.
The first scheme, denoted \textit{All Region Cell Attribute Descriptor} (ARCAD), is designed to discover discriminative {cell-level} attributes extracted from each region at the same time.
On the other hand, the second scheme, namely \textit{Cell Regional Attribute Descriptor} (CRAD), is proposed to discover the most discriminative {cell-level} attributes for each individual region.
Let $\mathcal{G} \in \{(\Mat{I}_i,\Vec{y}_i)\}^N_{i=1}$ be the training sets wherein each image $I_i$ has label vector $\Vec{y}_i \in \{-1,+1\} ^ K$ encoding the class label from K possible patterns.
For instance, if $I_i$ belongs to the first class, then $\Vec{y}_{i,1} = +1$ and $-1$ for the rest of its elements.
Let $\mathcal{X} \in \{ \{\Vec{x}_{i,j,1}\}_{j=1}^{J} \dots \{\Vec{x}_{i,j,N_{i,j}}\}_{j=1}^{J}\}$ be the set of extracted regional descriptors of the specimen image $I_i$.
\subsection{All Region Cell Attribute Descriptor (ARCAD)}
The training objective for ARCAD is defined by:
\begin{center}
\vspace{-2ex}
\begin{small}
\begin{equation}
\min_{\Mat{w}_{1 \dots K}, \Vec{b}_{1 \dots K}}
\sum_{k=1}^K \left\{ \frac{1}{2} \| \Mat{w}_k \|^2 + \frac{\lambda}{N} \sum_{i=1}^N \ell \left[ y_{i,k} (\Vec{b}_k + \Mat{w}^{\top}_k \Vec{z}_i \right] \right\}
\end{equation}
\end{small}
\end{center}
\vspace{-2ex}
\noindent
where $\Vec{w}_k$, $\Vec{b}_k$ and $\lambda$ are the hyperplane, bias term and regularisation parameters for each SVM, respectively; $\ell [ \cdot ]$ is the hinge loss function.
If we expand the $\Vec{z}_i$ in the above equation by substituting Eqn.~\ref{eqn:average_h_j} into~\ref{eqn:h_j}, it becomes:
\begin{center}
\vspace{-3ex}
\begin{small}
\begin{equation}
\label{eqn:ARCAD}
\min_{\Mat{w}_{1 \dots K}, \Vec{b}_{1 \dots K}, \Mat{A}_{1 \dots J}}
\sum_{k=1}^K \left\{ \frac{1}{2} \| \Mat{w}_k \|^2 + \frac{\lambda}{N} \sum_{i=1}^N \ell \left[ y_{i,k} (\Vec{b}_k +
\sum^{J}_{j=1} \left( \frac{\Vec{w}_{k,j}}{N_{i,j}} \sum^{N_{i,j}}_{c=1} \Mat{A}^{\top}_j \Vec{x}_{i,j,c} \right) \right] \right\}
\end{equation}
\end{small}
\end{center}
\vspace{-1ex}
\noindent
where $\Vec{w}_{k,j}$ is the parameters for $k$-th
linear SVM of the $j$-th
region~\footnote{Here $\Vec{w}_{k,j} \in \mathbb{R}^{1 \times b}$ is a sub-vector of $\Vec{w}_{k} = [\Vec{w}_{k,1} \dots \Vec{w}_{k,J}]$.}.
Minimising the above equation requires to learn all the parameters simultaneously.
This includes the specimen image one-versus-all SVMs as well as the {cell-level} attribute basis classifiers $\{\Mat{A}_j\}^J_{j=1}$.
We note that although the Eqn.~\ref{eqn:ARCAD} shares similarities to the objective function discussed in the PiCoDes
approach of~\cite{Bergamo2011}, applying the solution proposed in PiCoDes to solve the above equation is not straightforward.
This is because in contrast to PiCoDes, the above equation is not aimed at learning discriminative {image-level} attributes.
In the present work, the {image-level} descriptor is formed by concatenating the overall {cell-level} attribute descriptors extracted from cell regions.
For completeness, we present the objective function of PiCoDes which discovers discriminative {image-level} attributes:
\begin{center}
\begin{small}
\begin{equation}
\label{eqn:PiCoDes}
\min_{\Mat{w}_{1 \dots K}, \Vec{b}_{1 \dots K}, \Mat{A}_{1 \dots J}}
\sum_{k=1}^K \left\{ \frac{1}{2} \| \Mat{w}_k \|^2 + \frac{\lambda}{N} \sum_{i=1}^N \ell \left[ y_{i,k} (\Vec{b}_k +
\sum^{P}_{p=1} \left( w_{k,p} \Mat{A}^{\top}_p \Vec{u}_i \right) \right] \right\}
\end{equation}
\end{small}
\end{center}
\noindent
where $P$ is the number of {image-level} attributes; $w_{k,p}$ is the $p$-th parameter value of the $k$-th SVM model~\footnote{Note that $w_{k,p}$ is a scalar value. $w_{k,p}$ differs from $\Vec{w}_{k,j}$ which is a sub-vector}; $\Vec{u}_i$ is the $i$-th {image-level} descriptor. Note that $\Vec{u}_i$ is not the same as $\Vec{z}_i$ as the former is image descriptor extracted using various image features.
To address our problem, we present the following proposition.
\vspace{-1ex}
\begin{proposition}
\label{prop:1}
The problem presented in Eqn.~\ref{eqn:ARCAD} is equivalent to Eqn.~\ref{eqn:PiCoDes} if and only if $\Vec{z}_i$ is formed by $\hat{\Vec{h}}_{i,j}$ defined in Eqn.~\ref{eqn:average_h_j} and $\Vec{u}_i = [\hat{\Vec{x}}_{i,j} \dots \hat{\Vec{x}}_{i,J}]$ where $\hat{\Vec{x}}_{i,j} = \frac{1}{N_{i,j}} \sum^{N_{i,j}}_{c=1} \Vec{x}_{i,j,c}$
\end{proposition}
\vspace{-1ex}
The proof for Proposition.~\ref{prop:1} is presented in the Appendix.
The above proposition allows us to design a
tractable solution for Eqn.~\ref{eqn:ARCAD}
using the existing PiCoDeS solution for Eqn.~\ref{eqn:PiCoDes}.
\subsection{Cell Regional Attribute Descriptor (CRAD)}
The ARCAD objective function learns {cell-level} attributes for all regions at the same time.
Another alternative is to learn discriminative {cell-level} attribute from each region exclusively.
To this end, we propose an objective function consisting of
the summation of $\operatorname{E} ( \cdot )$, defined by:
\begin{small}
\begin{equation}
\operatorname{min} \sum_{j = 1}^J \operatorname{E} \left( \Vec{w}_{1 \dots K}^{[j]}; \Vec{b}^{[j]}_{1 \dots K}; \Mat{A}_j; \{ ( \hat{\Vec{h}}_{i,j}, \Vec{y}_i ) \}_{i = 1}^N \right)
\end{equation}
\end{small}
\noindent
where $\operatorname{E} ( \cdot )$ is the objective function for learning the basis classifier for each {cell-level} attribute extracted from each region:
\begin{small}
\begin{equation}
\operatorname{E} \left( \Mat{A}_j; \{ \hat{\Vec{h}}_{n,j} \}_{n = 1}^N \right) =
\sum_{k=1}^K \left\{ \frac{1}{2} \| \Mat{w}^{[j]}_k \|^2 + \frac{\lambda^{[j]}}{N} \sum_{i=1}^N \ell \left[ y_{i,k} (\Vec{b}^{[j]}_k + \Mat{w}^{[j]\top}_k \hat{\Vec{h}}_{i,j} \right] \right\}
\end{equation}
\end{small}
\noindent
where $\Vec{w}^{[j]}_k$ and $\Vec{b}^{[j]}_k$ are the $k$-th SVM parameters of the $j$-th region {cell-level} attributes.
Similar to Eqn.~\ref{eqn:ARCAD}, the above equation can
be expanded by substituting Eqn.~\ref{eqn:average_h_j} into~\ref{eqn:h_j}. We get:
\begin{small}
\begin{equation}
\label{eqn:CRAD}
\operatorname{E} \left( \Mat{A}_j; \{ \hat{\Vec{h}}_{n,j} \}_{n = 1}^N \right) = \sum_{k=1}^K \left\{ \frac{1}{2} \| \Mat{w}^{[j]}_k \|^2 + \frac{\lambda^{[j]}}{N} \sum_{i=1}^N \ell \left[ y_{i,k} (\Vec{b}^{[j]}_k + \frac{\Mat{w}^{[j]\top}_k}{N_{i,j}} \sum_{c=1}^{N_{i,j}} \Mat{A}^{\top}_j \Vec{x}_{i,j,c} \right] \right\}
\end{equation}
\end{small}
\noindent
This brings us to similar challenge as posed in Eqn.~\ref{eqn:ARCAD} which can be solved by using Proposition.~\ref{prop:1}.
We note that the purpose of Eqn.~\ref{eqn:CRAD} is only to train the cell attribute basis classifiers from each region individually.
Once these are trained, we train the one-versus-all SVM classifiers
for the specimen-level classification.
\subsection{Optimisation algorithm}
\label{sub:optimisation_algorithm}
We present the optimisation algorithm used to solve Eqn.~\ref{eqn:PiCoDes} for solving Eqn.~\ref{eqn:ARCAD} and~\ref{eqn:CRAD}.
The algorithm uses block coordinate descent alternating between optimising the
SVM parameters and the attribute basis classifiers.
\subsubsection{Learning specimen image SVM parameters}
When $\Mat{A}_j$ is fixed, $\Vec{z}_i$ can be determined.
Thus, the problem is reduced to linear a SVM learning problem for both ARCAD and CRAD.
\subsubsection{Learning attribute basis classifiers}
More elaborate steps are required to learn individual
attribute basis classifiers.
It has been shown in~\cite{Bergamo2011} that when $\Vec{w}_k$ and $\Vec{b}_k$ are fixed, updating each {cell-level} attribute base classifier from each region can be done via learning the upper
bound of the objective function which resembles to learning
a linear SVM classifier by minimising the sum of weighed
mis-classifications. Our objective function is:~\footnote{Readers are encouraged to read~\cite{Bergamo2011} for further details.}
\vspace{-1ex}
\begin{equation}
\operatorname{\hat{E}} (\Vec{a_{b}}) = \sum^N_{i=1} v_i \ell( q_i \Vec{a_b} \hat{\Vec{x}}_{i,j} )
\label{eqn:basis_learning}
\end{equation}
\vspace{-1ex}
\noindent
where $\hat{\Vec{x}}_{i,j}$ is the average of the low level features extracted from the $j$-th region of image $I_i$; $\Vec{a_b}$ is the base classifier model of the \mbox{$b$-th} attribute; $q_i \in \{-1,+1\}$ and $v_i \in \mathbb{R}^{+}$ are known values computed via:
\vspace{-2ex}
\begin{align}
q_i & = \left| \sum^{K}_{k=1} \ell (\alpha_{i,k,b} + \beta_{i,k,b} ) - \ell( \beta_{i,k,b} ) \right| \\
v_i & = \operatorname{sgn} \left( \sum^{K}_{k=1} \ell (\alpha_{i,k,b} + \beta_{i,k,b} ) - \ell (\beta_{i,k,b}) \right)
\end{align}
\vspace{-1ex}
\noindent
where $\alpha$ and $\beta$ are defined by: $\alpha_{i,k,b} \equiv y_{i,k} w_{k,b}$ and $\beta_{i,k,b} \equiv y_{i,k} b_k + \sum_{b' \neq b} y_{i,k} w_{k,b'}\Vec{a_b^{\top}} \Vec{u}_i$
The main difference between CRAD and ARCAD is that CRAD uses the SVM parameters of each region (\textit{i.e.}~ ~$\Vec{w}^{[j]}_k$ and $\Vec{b}^{[j]}_k$) to update the base classifiers, whereas ARCAD uses the overall SVM parameters.
In other words, in ARCAD, the information learned from
other regions are used to optimise Eqn.~\ref{eqn:basis_learning},
while this is not the case for CRAD.
\section{Related Work}
\cite{berg2010} in the oposite direction, from text to attributes.
\subsection{HEp-2 cell classification}
\subsection{Attributes}
Some works have been done like in~\cite{duan2012}.
Duan et al~\cite{duan2012} only localizes the attributes, whereas in our case we do in cell level and compute the statistic of them to describe an image.
|
\section{Introduction}
Compressive sensing is a recent field in signal processing that predicts that sparse vectors can be stably reconstructed from incomplete measurements via efficient algorithms \cite{fora13,elku12}.
Traditionally, the synthesis sparsity model is used in this context, where it is assumed that the signal can be written as a linear combination of a few elements of an appropriate basis.
Recently, the analysis sparsity model (or cosparsity model) has attracted significant interest as well \cite{caelnera11,daelgrna11,daelgigrna14,kara13}, where it is assumed that the signal is sparse after a transformation.
In the case of a basis transformation, the synthesis and analysis models coincide, but if the transform is redundant then the analysis sparsity model is different, and in fact the class of analysis sparsity
models is richer than the class of synthesis sparsity models, see \cite{EladMilanfarRubinstein,daelgrna11} for further details.
While the naive recovery approach of $\ell_0$-minimization is NP-hard, one may use convex relaxations, i.e., $\ell_1$-minimization in both the synthesis and analysis sparsity cases. Especially in the synthesis sparsity case, $\ell_1$-minimization is by now rather well understood, see e.g.\ \cite{fora13}. Despite recent progress in the theory of
analysis $\ell_1$-minimization \cite{caelnera11,daelgrna11,daelgigrna14,kara13},
there still remain a number of questions to be explored. In particular, in the important special case of total variation minimization which is ubiquitious in image processing
only a few contributions analyzing bounds for recovery from underdetermined measurements are available \cite{newa12,NeedellWard13,CaiXu}. The results of \cite{newa12,NeedellWard13} cover stable and robust recovery of signals in $\mathbb{C}^{d^n}$ of arbitrary dimension $n\geq 2$. They rely on the restricted isometry property and bounds on the decay of wavelet coefficients of high-dimensional signals by its total variation semi-norm. The authors of \cite{CaiXu} provide a bound on the number of Gaussian measurements that guarantee stable and robust recovery of signals in $\mathbb{R}^{d^n}$ with $n\geq 1$. The recovery is proved by establishing the null space property for the measurement matrix. For $n=1$ and fixed signal dimension $d$ bounds on the number of measurements were computed using Gordon's escape through a mesh theorem. In the asymptotic regime when $d\to\infty$ the performance guarantees explore the Grassmann angle framework. This article contributes to the topic of total variation minimization by providing a bound on the number of Gaussian random measurements in order to recover a gradient sparse signal via total variation minimization.
Moreover, we also provide an alternative bound to the one in \cite{kara13} on the number
of required Gaussian measurements for recovery of analysis-sparse signals with respect to a frame.
In contrast to \cite{newa12,NeedellWard13,CaiXu} which establish uniform recovery of all signals simultaneously with a single draw of a measurement matrix, our main results concern so-called nonuniform recovery using a Gaussian random measurement matrix, i.e., we fix a sparse (or rather cosparse)
vector (with respect to a given analysis operator) and provide bounds that guarantee that the given vector is recovered via analysis $\ell_1$-minimization
with high probability. Our bounds are particularly good when the sparsity is not very small compared to the ambient dimension. In fact, in the case of analysis-sparsity with respect to a very redundant frame,
there are natural lower bounds for the sparsity, so that in this context
the bounds in \cite{kara13} may turn out to be trivial in certain situations as they require more measurements than the ambient dimension. In contrast, the bounds
derived in this paper are always non-trivial in the sense that the number of required measurements is always lower than the ambient dimension.
Following \cite{chparewi12}, our analysis is based on estimating widths of tangent cones of a transformed $\ell_1$-norm at the (co-)sparse vector to be recovered, similarly
to \cite{kara13} or \cite[Chapter 9.2]{fora13}.
This is in contrast to uniform recovery bounds which are often based on the restricted isometry property or the null space property \cite{fora13}, see also \cite{kara13} for precise bounds
for an analysis sparsity version of the null space property for Gaussian random measurements.
In mathematical terms, we wish to recover a signal $x\in\mathbb{R}^d$ from measurements
\begin{equation}\label{eqMeasurements}
y=Mx+w,
\end{equation}
where $M\in\mathbb{R}^{m\times d}$ is a measurement matrix and $w \in \mathbb{R}^m$ with $\norm{w}_2\leq\eta$ corresponds to noise.
When $m\ll d$, there are infinitely many solutions to (\ref{eqMeasurements}). However, the prior sparsity knowledge about the underlying signal $x$ makes its recovery possible.
We assume that $x$ possesses a structure generated by a matrix $\Omega \in \mathbb{R}^{p \times d}$, called the analysis operator, in the sense that the application of $\Omega$ to $x$ produces a vector
with a small number of non-zero entries. If $\Omega x$ has $s$ non-zero entries, then $x$ is called $\ell$-\emph{cosparse}, where the number $\ell:=p-s$ is refered to as \emph{cosparsity} of $x$
(with respect to $\Omega$).
The index set of the zero entries of $\Omega x$ is called the \emph{cosupport} of $x$. Analysis $\ell_1$-minimization tries to recover the signal by computing the minimizer of
\begin{equation}\label{eqProblemP1Noise}
\underset{z\in\mathbb{R}^d}\min\;\norm{\Omega z}_1\;\;\mbox{subject to}\;\;\; \norm{Mz-y}_2\leq\eta.
\end{equation}
In this paper we consider two prominent examples of the analysis operator. The method of total variation corresponds to the program (\ref{eqProblemP1Noise}), when $\Omega$ is a difference operator. In the one-dimensional case it is defined by the matrix
\begin{equation}\label{eq:OneDimDifferenceOperator}
\Omega= \left(\begin{matrix} -1 & 1 & 0 & \cdots & 0\\
0 & -1 & 1 & \cdots & 0 \\
\vdots & & \ddots & \ddots & \vdots \\
0 & \cdots & 0 & -1 & 1 \end{matrix}\right).
\end{equation}
In this setting (\ref{eqProblemP1Noise}) promotes piecewise constant signals with sparse gradient.
Another important example of the analysis operator appears when the rows $\omega_j$ of $\Omega$ form a frame, i.e., if there exist constants $0 < A\leq B < \infty$ such that
\begin{equation}\label{eq:Frame}
A \|x\|_2^2 \leq \| \Omega x \|_2^2 = \sum_{j=1}^p |\langle \omega_j,x\rangle|^2 \leq B \|x\|_2^2.
\end{equation}
We are interested in the minimal number of measurements in terms of the sparsity (or cosparsity) required to recover a cosparse vector from its measurements $y$ in \eqref{eqMeasurements} when
the matrix $M$ is a Gaussian random matrix, i.e., its entries are independent standard normal distributed random variables.
We rely on a recent result of Foygel and Mackey in \cite{foma14} which is in spirit of the geometric approach of \cite{chparewi12} in order to provide such a bound on the number of Gaussian
measurements needed to recover $x$ via (\ref{eqProblemP1Noise}), when $\Omega$ satisfies either (\ref{eq:OneDimDifferenceOperator}) or (\ref{eq:Frame}).
\subsection{Main Results}
We first provide a general bound for the required number $m$ of Gaussian measurements in order to (stably) recover an analysis-sparse signal $x \in \mathbb{R}^d$
via analysis $\ell_1$-minimization, see Theorem~\ref{thNumberOfMeasurementsForDecomposableNorms}.
An important feature of the result is that $m$ is always (essentially) less than the ambient dimension $d$. Based on this, we show that
a signal $x \in \mathbb{R}^d$
which is $s$-sparse with respect to the difference operator $\Omega$ in \eqref{eq:OneDimDifferenceOperator} (that is, whose gradient is $s$-sparse) can be recovered with high probability if ``roughly'', i.e., ignoring terms
of lower order
\begin{equation}\label{m:TV:rough}
m > d\left( 1- \frac{1}{\pi}(1- \frac{s+1}{d})^2\right),
\end{equation}
see Theorem~\ref{thNumberOfMeasurementsForL1NormDifOperator}. In \eqref{m:TV:rough}, the number of measurements is clearly less than $d$. Note that the usual bound in compressive sensing require
\begin{equation}\label{m:standard}
m > c s \log(d/s)
\end{equation}
for recovery of $s$-sparse vectors via $\ell_1$-minimization from Gaussian random measurements \cite[Chapter 9]{fora13}, \cite{chparewi12}.
One realizes a structural difference of this bound to the new one stated above.
In fact, even for sparsity $s=1$, \eqref{m:TV:rough} roughly requires $m > (1-\pi^{-1}) d \approx 0.6817 \, d$, while \eqref{m:standard} requires $m \geq c \log(d)$, which is of much smaller order. It should be pointed out that a bound of the form (\ref{m:standard}) is so far not available for TV-minimization with $1$-dimensional difference operator. For this particular case a bound which resembles (\ref{m:standard}) is obtained in \cite{CaiXu} and it requires
\begin{equation}\label{m:CaiXuResult}
m\geq c(sd)^{1/2} \log(d).
\end{equation}
However, if $s$ is proportional to $d$, i.e., $s = \alpha d$ for some $\alpha \in (0,1)$ then \eqref{m:TV:rough} requires $m \geq c_\alpha d$ with $c_\alpha = 1-\frac{1}{\pi}(1-\alpha)^2$ which is always less than $1$, while
\eqref{m:standard} and \eqref{m:CaiXuResult} give $m \geq c'_\alpha d$ with $c'_\alpha = c \alpha \ln( \alpha^{-1})$ which may become larger than $1$ because the available estimates for $c$ in \eqref{m:standard} are larger than $2$ \cite{fora13}.
As the second case, we consider $\Omega \in \mathbb{R}^{p \times d}$ to be a frame with frame bounds $A,B > 0$ in \eqref{eq:Frame}. Theorem~\ref{thNumberOfMeasurementsForL1NormFrameOperator}
below shows that a signal $x \in \mathbb{R}^d$ which is $\ell$-cosparse
with cosupport $\Lambda$, i.e., $\supp \Omega x = \Lambda^c$, can be recovered from $m$ Gaussian measurements via analysis $\ell_1$-minimization with high probability if, roughly speaking,
\begin{equation}\label{m:frame:rough}
m > d - \frac{2}{\pi p B}\left(\sum_{\ell \in \Lambda} \|\omega_\ell\|_2\right)^2.
\end{equation}
According to \eqref{m:frame:rough},
the number of measurements is always less than $d$, and even though there is no direct dependence on lower frame bound $A$, it is still independent of the scaling of $\Omega$. The bound derived in \cite{kara13} for analysis sparse recovery with respect to a frame roughly requires
\begin{equation}\label{m:prev:frame}
m \geq \frac{2B}{A} s \ln(ep/s),
\end{equation}
see also \cite{kara14}.
In order to place the bound \eqref{m:frame:rough} of Theorem~\ref{thNumberOfMeasurementsForL1NormFrameOperator} for the frame case into context, we recall some facts on the analysis sparsity model.
Consider the generic case, that the rows of the analysis operator $\Omega$, that is, the frame elements $\omega_j$, $j = 1,\hdots,p$, are in general linear position so that a subcollection of $d$ frame elements
is always linearly independent. Then any subspace $W_\Lambda = \operatorname{span} \{ \omega_j, j \in \Lambda \}^\perp$ with $\#\Lambda = \ell$, of $\ell$-cosparse vectors,
where ${}^\perp$ denotes the orthogonal complement, has dimension $d-\ell$, which means that the smallest value that the sparsity parameter $s = p- \ell$ can take for a nontrivial vector $x$
is $p-d+1$, see also \cite{daelgrna11,kara14}. Therefore, if $p = \kappa d$ for some $\kappa > 1$, then the sparsity is always proportional to the dimension $d$, i.e., $s \sim \alpha d$
which means that the bound \eqref{m:frame:rough} reads $m\geq c_{\kappa,\alpha}d$ with $c_{\kappa, \alpha}=1-\frac{2(\kappa-\alpha)^2}{\pi\kappa B}$, assuming that the frame is normalized. At the same time the $\log$ factor in \eqref{m:standard} becomes constant for $s$ proportional to $d$ and if $\kappa > 2$, then the sparsity is always at least as large as the ambient dimension, so that
\eqref{m:standard} becomes a trivial bound. Similar considerations
apply to the previous bound in \cite{kara13} on analysis-sparse recovery with respect to frames.
In contrast, let us consider the new bound \eqref{m:frame:rough} for a tight unit norm frame, i.e., $\|\omega_j\|_2=1$ and $A=B = p/d$.
For cosparsity $\# \Omega = \ell = p-s$, \eqref{m:frame:rough} yields then
\[
m > d - \frac{2d}{\pi p^2}(p-s)^2 = \left(1- \frac{2(p-s)^2}{\pi p^2}\right) d.
\]
For sparsity proportional to $p$, the number of measurements clearly scales like $\beta d$ where $\beta$ is always strictly smaller than $1$,
which shows that our new bound is better in the natural regime $p = \kappa d$ with
$\kappa \geq 2$, say. In contrast, the previous bound \eqref{m:prev:frame} requires $m$ to be larger than $d$ for a frame in general position with $p = \kappa d$, see Figure \ref{fig:ComparisonOfBounds}.
\begin{figure}
\centering
\includegraphics[scale=0.5]{FrameSparsityProportionVsMeasurements}
\caption{Comparison of the standard recovery bounds and the new one for the setting of a normalized tight frame with frame bounds $A=B=\frac{p}{d}$. The red line corresponds to the new bound on the rescaled number of measurements $\frac{m}{d}$ as a function of sparsity fraction $\frac{s}{p}$. The black and blue lines represent the bound given by (\ref{m:prev:frame}) for $p=d$ and $p=1.3d$ respectively.}
\label{fig:ComparisonOfBounds}
\end{figure}
\subsection{Notation}
We denote the rows of $\Omega\in\mathbb{R}^{p\times d}$ by $\omega_j$, $j=1,\dots,p$. We use $\Omega_{\Lambda}$ to refer to a submatrix of $\Omega$ with the rows indexed by $\Lambda$; $\Omega^T$ is the transpose of $\Omega$; $\norm{\Omega\cdot}_1$ denotes a semi-norm generated by the transform $\Omega$; $\alpha_{\Lambda}$ stands for the vector whose entries indexed by $\Lambda$ coincide with the entries of $\alpha$ and the rest are filled by zeros.
The sign of a real number $r\neq 0$ is $\sgn(r)=\frac{r}{\abs{r}}$.
For a vector $\alpha\in\mathbb{R}^p$ we define its sign vector $\sgn(\alpha)\in\mathbb{R}^p$ by
\[
(\sgn(\alpha))_i=\left\{\begin{array}{cl}
\frac{\alpha_i}{\abs{\alpha_i}}, & \mbox{if }\,\alpha_i\neq 0,\\
0, & \mbox{otherwise}.
\end{array}\right.
\]
We write $S^{d-1}$ for the unit sphere in $\mathbb{R}^d$.
\section{Recovery from Gaussian measurements}
\subsection{Theoretical guarantees}
Our analysis is based on work of Chandrasekaran et al.\ \cite{chparewi12}, where sufficient conditions for robust recovery make use of the tangent cone (also called descent cone) of
a convex function (e.g., the $\ell_1$-norm) at the signal to be recovered, see also \cite{tr14}.
In the case of analysis $\ell_1$-minimization these are formulated explicitly in \cite{kara13}. For a fixed $x\in\mathbb{R}^d$ we define a tangent cone as
\[
T(x)=\cone\{z-x:z\in\mathbb{R}^d,\;\norm{\Omega z}_1\leq\norm{\Omega x}_{1}\},
\]
where the notation ``cone'' stands for the conic hull of the indicated set. The set $T(x)$ consists of the directions from $x$, which do not increase the value of $\norm{\Omega x}_{1}$.
\begin{theorem}[\cite{kara13}]\label{thm:rec}
Let $x\in\mathbb{R}^d$. If
\begin{equation}\label{eqTangentConeWithNoise}
\underset{\substack{v\in T(x)\\ \norm{v}_2\leq 1}}\inf \norm{Mv}_2\geq\tau,
\end{equation}
for some $\tau>0$, then a minimizer $\hat x$ of (\ref{eqProblemP1Noise}) satisfies
\[
\norm{x-\hat x}_2\leq \frac{2\eta}{\tau}.
\]
\end{theorem}
\subsection{Gordon's escape through a mesh theorem}
According to (\ref{eqTangentConeWithNoise}), successful recovery of a signal is achieved, when the minimal gain of the measurement matrix over the tangent cone is greater than some positive constant. For Gaussian matrices the probability of this event can be estimated by Gordon's escape through a mesh theorem \cite{go88-2}, see also \cite[Theorem 9.21]{fora13}. In order to present Gordon's result formally, we introduce some notation.
Let $g\in\mathbb{R}^m$ be a standard Gaussian random vector. Then
\[
E_m:=\mean\norm{g}_2=\sqrt{2}\;\frac{\Gamma\brac{(m+1)/2}}{\Gamma\brac{m/2}},
\]
where $\Gamma$ is the standard Gamma function, satisfies
\[
\frac{m}{\sqrt{m+1}}\leq E_m\leq\sqrt{m}.
\]
For a set $T\subset\mathbb{R}^d$ we define its Gaussian width by
\[
\ell(T):=\mean\underset{x\in T}\sup\;\abrac{x,g},
\]
where $g\in\mathbb{R}^d$ is a standard Gaussian random vector.
\begin{theorem}[Gordon's escape through a mesh]\label{thGordonsEscapeThroughTheMesh}
Let $M\in\mathbb{R}^{m\times d}$ be a Gaussian random matrix and $T$ be a subset of the unit sphere $S^{d-1}=\{x\in\mathbb{R}^d: \norm{x}_2=1\}$. Then, for $t>0$,
\begin{equation}\label{eqGordonsEscapeThroughTheMesh}
\mathbb{P}\brac{\underset{x\in T}\inf\norm{Mx}_2> E_m-\ell(T)-t}\geq 1-e^{-\frac{t^2}{2}}.
\end{equation}
\end{theorem}
Thus, to provide a bound on the number of Gaussian measurements, we estimate the Gaussian width of the tangent cone $T(x)$ intersected with the unit sphere.
\subsection{Estimates for the Gaussian width}
A basic technique to estimate the Gaussian width of the tangent cone was developed in the work of Stojnic \cite{st09-6} and was later refined in a series of papers \cite{amlomctr13,chparewi12}, see also \cite{tr14}.
The result states that the Gaussian width of the tangent cone can be bounded by the Euclidean distance of the standard normal vector to the scaled cone generated by the subdifferential.
Recall that the Euclidean distance to a set $T \subset \mathbb{R}^d$ is the function defined by
\[
\dist(x,T):=\inf\fbrac{\norm{x-u}_2:u\in T}.
\]
The subdifferential of a convex function $f: \mathbb{R}^d \to \mathbb{R}$ at a point $x \in \mathbb{R}^d$ is the set
\[
\partial f(x) = \{ v \in \mathbb{R}^d : f(z) \geq f(x) + \langle z-x,v \rangle \mbox{ for all } z \in \mathbb{R}^d \}.
\]
It is shown in \cite{st09-6,chparewi12} that
\begin{equation}\label{eqGaussianWidthEuclideanDistance}
\ell(T(x)\cap\sph^{d-1})^2\leq\underset{t\geq 0}\inf\mean\sbrac{\dist(g,t\cdot\partial\norm{\Omega\cdot}_{1}(x))}^2,
\end{equation}
where $g\in\mathbb{R}^d$ is the standard normal vector.
To provide a bound on the expected squared distance from (\ref{eqGaussianWidthEuclideanDistance}) we generalize Proposition 3 in \cite{foma14} valid for $\Omega$ being a basis to the following result.
\begin{lemma}
Let $\Omega\in\mathbb{R}^{p\times d}$ be an analysis operator, $g\in\mathbb{R}^d$ be a standard normal random vector and $x\in\mathbb{R}^d$. Then
\begin{equation}\label{eqEstimateGWFoygel}
\underset{t\geq 0}\inf\mean\sbrac{\dist(g,t\cdot\partial\norm{\Omega\cdot}_{1}(x))}^2\leq d-\frac{\sbrac{\ell(\partial\norm{\Omega\cdot}_{1}(x))}^2}{\underset{\norm{z}_2\leq 1}\max\norm{\Omega z}_{1}^2}.
\end{equation}
\end{lemma}
\begin{proof}
Since $\partial \norm{\Omega\cdot}_1(x)$ is a compact set, there exists $w_0\in \partial \norm{\Omega\cdot}_1(x)$ such that
\[
\abrac{g,w_0}=\underset{w\in\partial \norm{\Omega\cdot}_1(x)}\max\abrac{g,w}.
\]
Then
\[
\dist(g,t\cdot\partial\norm{\Omega\cdot}_{1}(x))^2\leq \norm{g-tw_0}_2^2=\norm{g}_2^2-2t\abrac{g,w_0}+t^2\norm{w_0}_2^2.
\]
Moreover,
\[
\partial\norm{\Omega\cdot}_{1}(x)=\Omega^T\brac{\partial\norm{\cdot}_{1}(\Omega x)}
\]
and
\[
\partial\norm{\cdot}_1(\Omega x)=\fbrac{\alpha\in\mathbb{R}^p:\alpha_{\Lambda^c}=\sgn(\Omega x),\;\;\norm{\alpha_{\Lambda}}_{\infty}\leq 1},
\]
which implies that $w_0=\Omega^T\alpha_0$ for some $\alpha_0\in\partial\norm{\cdot}_1(\Omega x)$. Thus
\[
\dist(g,t\cdot\partial\norm{\Omega\cdot}_{1}(x))^2\leq\norm{g}_2^2-2t\underset{w\in\partial \norm{\Omega\cdot}_1(x)}\max\abrac{g,w}+t^2\norm{\Omega^T\alpha_0}_2^2.
\]
Due to duality and the fact that $\norm{\alpha_0}_{\infty}\leq 1$ we obtain
\[
\norm{\Omega^T\alpha_0}_2=\underset{\norm{z}_2\leq 1}\max\abrac{z,\Omega^T\alpha_0}=\underset{\norm{z}_2\leq 1}\max\abrac{\Omega z,\alpha_0}\leq\underset{\norm{z}_2\leq 1}\max\norm{\Omega z}_1\norm{\alpha_0}_{\infty}\leq\underset{\norm{z}_2\leq 1}\max\norm{\Omega z}_1.
\]
So altogether
\[
\dist(g,t\cdot\partial\norm{\Omega\cdot}_{1}(x))^2\leq \norm{g}_2^2-2t\underset{w\in\partial \norm{\Omega\cdot}_1(x)}\max\abrac{g,w}+t^2\underset{\norm{z}_2\leq 1}\max\norm{\Omega z}_1^2
\]
and by taking the expectation of both sides we obtain
\[
\mean\dist(g,t\cdot\partial\norm{\Omega\cdot}_{1}(x))^2\leq d-2t\ell(\partial\norm{\Omega\cdot}_1(x))+t^2\underset{\norm{z}_2\leq 1}\max\norm{\Omega z}_1^2.
\]
Setting $t=\frac{\ell(\partial\norm{\Omega\cdot}_1(x))}{\underset{\norm{z_2}\leq 1}\max\norm{\Omega z}_1^2}$ yields
\[
\underset{t\geq 0}\inf\mean\sbrac{\dist(g,t\cdot\partial\norm{\Omega\cdot}_{1}(x))}^2\leq d-\frac{\sbrac{\ell(\partial\norm{\Omega\cdot}_{1}(x))}^2}{\underset{\norm{z}_2\leq 1}\max\norm{\Omega z}_{1}^2}.
\]
\end{proof}
\subsection{Number of Gaussian measurements}
Gordon's escape through a mesh, Theorem \ref{thGordonsEscapeThroughTheMesh}, together with the estimates (\ref{eqGaussianWidthEuclideanDistance}) and (\ref{eqEstimateGWFoygel}) leads to the next result.
\begin{theorem}\label{thNumberOfMeasurementsForDecomposableNorms}
Let $\Omega\in\mathbb{R}^{p\times d}$ be an analysis operator and $x$ be cosparse with cosupport $\Lambda$. For a random draw $M\in\mathbb{R}^{m\times d}$ of a Gaussian matrix, let noisy measurements $y=Mx+w$ be given with $\norm{w}_2\leq\eta$ and $0<\varepsilon<1$. If
\begin{equation}\label{eqNumberOfMeasurementsByGW}
\frac{m^2}{m+1}\geq \brac{\sqrt{d-\frac{\brac{\mean\norm{\Omega_{\Lambda}g}_{1}}^2}{\underset{\norm{z}_2\leq 1}\max\norm{\Omega z}_{1}^2}}+\sqrt{2\ln(\varepsilon^{-1})}+\tau}^2,
\end{equation}
then with probability at least $1-\varepsilon$, any minimizer $\hat x$ of (\ref{eqProblemP1Noise}) satisfies
\[
\norm{x-\hat x}_2\leq\frac{2\eta}{\tau}.
\]
\end{theorem}
\begin{proof}
Estimates (\ref{eqGaussianWidthEuclideanDistance}) and (\ref{eqEstimateGWFoygel}) imply that
\begin{equation}\label{eq:PreEstimateWithGW}
E_m-\ell\brac{T(x)\cap \sph^{d-1}}-t\geq E_m-\sqrt{d-\frac{\sbrac{\ell(\partial\norm{\Omega\cdot}_{1}(x))}^2}{\underset{\norm{z}_2\leq 1}\max\norm{\Omega z}_{1}^2}}-t.
\end{equation}
For any $z\in\partial\norm{\Omega\cdot}_{1}(x)$ there is $\alpha\in\mathbb{R}^p$ with $\norm{\alpha_{\Lambda}}_{\infty}\leq 1$ such that
\[
z=\Omega^T\alpha=\Omega^T\sgn(\Omega x)+\Omega^T\alpha_{\Lambda}.
\]
Therefore,
\[
\ell(\partial\norm{\Omega\cdot}_{1}(x))=\mean\underset{z\in\partial\norm{\Omega\cdot}_{1}(x)}\max\abrac{g,z}=\mean\abrac{g,\Omega^T\sgn(\Omega x)}+\mean\underset{\norm{\alpha_{\Lambda}}_{1}\leq 1}\max\abrac{g,\Omega^T\alpha_{\Lambda}}
\]
\[
=\mean\underset{\norm{\alpha_{\Lambda}}_{1}\leq 1}\max\abrac{(\Omega g)_{\Lambda},\alpha_{\Lambda}}=\mean\norm{\Omega_{\Lambda}g}_{1}.
\]
Plugging this into (\ref{eq:PreEstimateWithGW}) gives
\[
E_m-\ell\brac{T(x)\cap \sph^{d-1}}-t\geq E_m-\sqrt{d-\frac{\brac{\mean\norm{\Omega_{\Lambda}g}_{1}}^2}{\underset{\norm{z}_2\leq 1}\max\norm{\Omega z}_{1}^2}}-t.
\]
Setting $t=\sqrt{2\ln(\varepsilon^{-1})}$, the choice of $m$ in (\ref{eqNumberOfMeasurementsByGW}) guarantees that
\[
E_m-\ell\brac{T(x)\cap \sph^{d-1}}-t\geq\frac{m}{\sqrt{m+1}}-\sqrt{d-\frac{\brac{\mean\norm{\Omega_{\Lambda}g}_{1}}^2}{\underset{\norm{z}_2\leq 1}\max\norm{\Omega z}_{1}^2}}-\sqrt{2\ln(\varepsilon^{-1})}\geq\tau.
\]
The monotonicity of probability together with Theorems~\ref{thm:rec} and \ref{thGordonsEscapeThroughTheMesh} yields
\[
\begin{aligned}
\mathbb{P}\brac{\underset{{x}\in T(x)\cap\sph^{d-1}}\inf\norm{Mx}_2\geq\tau}&\geq\mathbb{P}\brac{\underset{x\in T(x)\cap\sph^{d-1}}\inf\norm{Mx}_2\geq E_m-\ell\brac{T(x)\cap\sph^{d-1}}-t}\\
&\geq 1-\varepsilon.
\end{aligned}
\]
This concludes the proof.
\end{proof}
\section{Explicit choice of the analysis operator}
Theorem \ref{thNumberOfMeasurementsForDecomposableNorms} can be further refined for special choices of the operator $\Omega$. We start with the one-dimensional difference operator.
\begin{theorem}\label{thNumberOfMeasurementsForL1NormDifOperator}
Let $\Omega:\mathbb{R}^d\to\mathbb{R}^{d-1}$ be a one-dimensional difference operator. Let $x\in\mathbb{R}^d$ be $\ell$-cosparse with respect to $\Omega$ and $s=d-1-\ell$.
For a random draw $M\in\mathbb{R}^{m\times d}$ of a Gaussian matrix, let noisy measurements $y=Mx+w$ be given with $\norm{w}_2\leq\eta$ and $0<\varepsilon<1$. If
\begin{equation}\label{eqNumberOfMeasurementsForL1NormDifOperator}
\frac{m^2}{m+1}\geq \brac{\sqrt{d\brac{1-\frac{1}{\pi}\brac{1-\frac{s+1}{d}}^2}}+\sqrt{2\ln(\varepsilon^{-1})}+\tau}^2,
\end{equation}
then with probability at least $1-\varepsilon$, any minimizer $\hat x$ of (\ref{eqProblemP1Noise}) satisfies
\[
\norm{x-\hat x}_2\leq\frac{2\eta}{\tau}.
\]
\end{theorem}
\begin{proof}
According to the definition of $\Omega$, for any $z\in\mathbb{R}^d$
\begin{equation}\label{eq:PropertiesOfDiffOperator}
\norm{\Omega z}_1=\sum\limits_{i=1}^{d-1}\abs{z_{i+1}-z_i}\leq 2\norm{z}_1\leq 2\sqrt d\norm{z}_2.
\end{equation}
Hence, $\underset{\norm{z}_2\leq 1}\max\norm{\Omega z}_1^2\leq 4d$.
The rows $\omega_j$ of $\Omega\in\mathbb{R}^{(d-1)\times d}$ satisfy $\norm{\omega_i}_2=\sqrt 2$ and the properties of the Gaussian distribution imply
\begin{equation}\label{eq:PropertiesGaussianDistribution}
\mean\norm{\Omega_{\Lambda}g}_1=\mean\sum_{i\in\Lambda}\abs{\abrac{g,\omega_i}}=\sqrt{\frac{2}{\pi}}\sum\limits_{i\in\Lambda}\norm{\omega_i}_2=\frac{2}{\sqrt\pi}(d-1-s).
\end{equation}
Plugging (\ref{eq:PropertiesOfDiffOperator}) and (\ref{eq:PropertiesGaussianDistribution}) into (\ref{eqNumberOfMeasurementsByGW}) implies the desired result.
\end{proof}
A comparison of the theoretical bound to the bound obtained from the numerical experiments is shown on Figure \ref{fig:DiffOperatorNumericalExperiments}. We considered signals in $\mathbb{R}^{200}$. We ran the algorithm and counted the number of times the signal was recovered correctly out of 200 trials. A reconstruction error of $10^{-5}$ was considered as a successful recovery. Each pixel intensity represents
the ratio of the signals recovered perfectly with black being 100\% success.
\begin{figure}
\centering
\includegraphics[scale=0.3]{TV100TrialsResultMatrix}
\caption{The red line corresponds to the theoretical bound omitting the terms of lower order.}
\label{fig:DiffOperatorNumericalExperiments}
\end{figure}
We would like to point out that, even though the theoretical guarantee requires around 90\% of the samples to be taken, when 80 out of 199 entries of the gradient are non-zero, minimizing the TV-norm would still provide better performance than simply applying Tikhonov regularization with 90\% samples. We present a comparison of the results of the two reconstruction algorithms in Figure \ref{fig:ComparisonTVandLS}.
\begin{figure}
\minipage{0.32\textwidth}
\includegraphics[width=\linewidth]{x_init}
\endminipage\hfill
\minipage{0.32\textwidth}
\includegraphics[width=\linewidth]{x_TV}
\endminipage\hfill
\minipage{0.32\textwidth}%
\includegraphics[width=\linewidth]{x_LS}
\endminipage
\caption{The first plot depicts an initial signal in $\mathbb{R}^{200}$, whose gradient has 80 non-zero entries. The second and third one correspond to the reconstruction from 180 measurements via TV-minimization (exact) and Tikhonov regularization respectively.}\label{fig:ComparisonTVandLS}
\end{figure}
Theorem \ref{thNumberOfMeasurementsForL1NormDifOperator} can be extended to two dimensions. Let $X\in\mathbb{R}^{d\times d}$. The two-dimensional difference operator collects all vertical and horizontal derivatives of $X$ into a single vector. If we concatenate the columns of $X$ into the vector $x\in\mathbb{R}^{d^2}$, then we can represent this operator by the matrix $\Omega\in\mathbb{R}^{2d(d-1)\times d^2}$, whose action is given by
\[
\Omega x=\brac{X_{21}-X_{11},\ldots,X_{dd}-X_{d-1d},X_{12}-X_{11},\ldots,X_{dd}-X_{dd-1}}^T.
\]
Each row $\omega_i$ of $\Omega$ has exactly two non-zero entries with values $-1$ and $1$ at the proper locations.
\begin{theorem}\label{th2DNumberOfMeasurementsForDifOperator}
Let $\Omega\in\mathbb{R}^{2d(d-1)\times d^2}$ define a two-dimensional difference operator. Let $x\in\mathbb{R}^{d^2}$ be $\ell$-cosparse with respect to $\Omega$ and $s=2d(d-1)-\ell$.
For a random draw $M\in\mathbb{R}^{m\times d^2}$ of a Gaussian matrix, let noisy measurements $y=Mx+w$ be given with $\norm{w}_2\leq\eta$ and $0<\varepsilon<1$. If
\begin{equation}\label{eq2DNumberOfMeasurementsForDifOperator}
\frac{m^2}{m+1}\geq \brac{\sqrt{d^2\brac{1-\frac{1}{\pi}\brac{1-\frac{1}{d}-\frac{s}{2d^2}}^2}}+\sqrt{2\ln(\varepsilon^{-1})}+\tau}^2,
\end{equation}
then with probability at least $1-\varepsilon$, any minimizer $\hat x$ of (\ref{eqProblemP1Noise}) satisfies
\[
\norm{x-\hat x}_2\leq\frac{2\eta}{\tau}.
\]
\end{theorem}
\begin{proof}
Let $z\in\mathbb{R}^{d^2}$. Then
\[
\norm{\Omega z}_1=\sum_{j=1}^d\sum_{i=1}^{d-1}\abs{z_{i+1j}-z_{ij}}+\sum_{i=1}^d\sum_{j=1}^{d-1}\abs{z_{ij+1}-z_{ij}}\leq 4\norm{z}_1\leq 4d\norm{z}_2
\]
and it follows that
\[
\underset{\norm{z}_2\leq 1}\max\norm{\Omega z}_1^2\leq 16 d^2.
\]
As in the one-dimensional case,
\[
\mean\norm{\Omega_{\Lambda}g}_1=\sqrt{\frac{2}{\pi}}\sum_{i\in\Lambda}\norm{\omega_i}_2=\frac{2}{\sqrt{\pi}}(2d(d-1)-s).
\]
As the final step we apply formula (\ref{eqNumberOfMeasurementsByGW}).
\end{proof}
For an $\Omega$ being a frame, we obtain the following bound on the required number of measurements.
\begin{theorem}\label{thNumberOfMeasurementsForL1NormFrameOperator}
Let $\Omega:\mathbb{R}^d\to\mathbb{R}^{p}$ be a frame with an upper frame bound $B$. Let $x\in\mathbb{R}^d$ be cosparse with cosupport $\Lambda$. For a random draw $M\in\mathbb{R}^{m\times d}$ of a Gaussian matrix, let noisy measurements $y=Mx+w$ be given with $\norm{w}_2\leq\eta$ and $0<\varepsilon<1$. If
\begin{equation}\label{eqNumberOfMeasurementsForL1NormFrameOperator}
\frac{m^2}{m+1}\geq \brac{\sqrt{d-\frac{2}{\pi}\frac{\brac{\sum_{i\in\Lambda}\norm{\omega_i}_2}^2}{pB}}+\sqrt{2\ln(\varepsilon^{-1})}+\tau}^2,
\end{equation}
then with probability at least $1-\varepsilon$, any minimizer $\hat x$ of (\ref{eqProblemP1Noise}) satisfies
\[
\norm{x-\hat x}_2\leq\frac{2\eta}{\tau}.
\]
\end{theorem}
\begin{proof}
The only difference to the proof of Theorem \ref{thNumberOfMeasurementsForL1NormDifOperator} is the following estimate, which is due to the Cauchy inequality and the fact that $\Omega$ is a frame:
\[
\underset{\norm{z}_2\leq 1}\max\norm{\Omega z}_1^2\leq p\underset{\norm{z}_2\leq 1}\max\norm{\Omega z}_2^2\leq p\underset{\norm{z}_2\leq 1}\max B\norm{z}_2^2\leq pB.
\]
\end{proof}
The bound on the number of measurements (\ref{eqNumberOfMeasurementsForL1NormFrameOperator}) does not have an explicit dependence on the ratio of the frame bounds.
So it can not be directly compared to the results provided in \cite{kara13}, see also \eqref{m:prev:frame}.
However, an important observation is that the right hand side in (\ref{eqNumberOfMeasurementsForL1NormFrameOperator}) is strictly less than $d$ for any $p$ and any number of elements in the cosupport of the signal
(provided $\varepsilon^{-1}$ and $\tau$ are not too large).
\section{Conclusions}
We have presented results on the nonuniform recovery from Gaussian random measurements of analysis-sparse signals with respect to the one- and two-dimensional difference operators or with respect to a frame. The derived bound on required measurements is always smaller than the ambient dimension of a signal and it is particularly suitable for the case when the sparsity of the analysis representation of the signal is not very small.
\section{Acknowledgements}
M.~Kabanava and H.~Rauhut acknowledge support by the European Research Council through the grant StG 258926. H. Zhang is supported by China NSF Grants No. 61201328.
|
\section{Introduction}
\label{sec:intro}
In the last decade, many papers have studied consensus in networks, see \cite{boyd2005gossip,boyd2006randomized,consensus,freeman2006stability,kempe2003gossip,kempe2004spatial} for some early work on consensus formation and gossip algorithms. These and subsequent papers concentrate on various algorithms to achieve consensus in a network. Typical consensus averaging algorithm is composed of a distributed averaging where each node in the network averages its data with its neighbors using a (possibly time varying) weighting. The simplest algorithms to analyze involve fixed averaging where the weighting matrix in the graph is doubly stochastic. A more sophisticated approach involves gossip based algorithms in which each node randomly picks a neighbor (with the possibility of not choosing any) and updates its data with the neighbor, and updates the weighting as well. It is shown that this scheme also converges to the population average.
A significant amount of research concentrated on the rate of convergence to the consensus using analysis of the second eigenvlue of the network weighting matrix, on the design of such matrices and on analysis of distributed convergence to good weighting which accelerates the convergence \cite{olshevsky2009convergence}. Similar results are also available for the random gossip model, although the analysis is more intricate \cite{boyd2006randomized} due to the stochastic nature of the gossip process.
Much work has also been devoted to the analysis of quantization schemes \cite{kashyap2007quantized}, and to imperfect channels which introduce noise \cite{rajagopal2011network,kar2009distributed}. The close analogy between diffusion processes and consensus averaging has also been a fruitful analogy which led to many interesting results, see e.g., \cite{sardellitti2010fast,chen2012diffusion,diffusion}.
More recent results can be found in the review paper by Dimakis et al. \cite{dimakis2010gossip}.
Most papers in consensus analysis
focus on the case of {\em initial value problems}, where a distributed (possibly time varying) operator is applied to the data.
In contrast, this paper considers {\em boundary value problems} in consensus networks.
Such boundary value problems were originally motivated by continuous diffusion problems, where boundary conditions play an important role, and in many cases influence the nature of the solution. We first analyze the effect of fixed boundary conditions on the network, and show that even when a single node is subject to a boundary condition, it is the value of the data in this node that actually determines the limiting value. For simplicity we focus on the constant averaging matrix. However in an extension of this work we analyze the effect of boundary value conditions on randomized gossip algorithms. Not surprisingly, our study of boundary conditions leads to the analysis of general harmonic functions on graphs. A good overview of such functions, which also play an important role in analyzing random walks on graphs is given by Benjamini and Lovasz \cite{benjamini2003harmonic}. The paper by Bendito et al. \cite{bendito2000solving} consider the solution of discrete boundary value problems using Green's functions. The solution is quite complicated and in this paper we take a different path, which leads to closed form solution and better intuitions regarding the solution.
The main results of the paper determine the asymptotic value of the nodes in a consensus network, subject to boundary conditions. We show that in this case the value is no longer constant at all nodes, but consists of a general harmonic function on the graph. We then prove an analog of the theorem that an harmonic function with constant value on the boundary is constant, and conclude that even fixing a single node can be used to drive the network to any desired consensus. This fact has several important consequences: Distributed detection in consensus networks can be significantly harmed even in the presence of a single malicious node. This also has impact on the sensitivity to consistent errors.
Wang and Krim \cite{wang2013control,wang2014analysis} have recently discussed the control of beliefs in social networks. They studied the maximal change in beliefs as a function of the degree distribution of the network. In this paper, we show that the beliefs can be steered towards any constant value by impacting a single node. Furthermore, any given change in the beliefs can be implemented by carefully selecting a small number of nodes to influence and using a time varying periodic strategy. The choice of nodes to impact has mainly an effect on the speed of convergence. to achieve this we analyze network with periodic boundary conditions.
The applications of this paper go beyond the obvious networking setups. Interestingly, the results of this paper also generalize the results of Cohen and Peleg \cite{cohen2005convergence}, \cite{cohen2008convergence}, \cite{cohen2008local} on distributed convergence of groups of robots to the center of gravity of the robots. These results are a special case of our main theorem. According to the model Cohen and Peleg each robot moves in the average direction of the robots it sees. This can be considered as the application of two consensus processes for the $x$ and $y$ coordinates of the robots. Furthermore, our results show that when some robots are fixed and serve as beacons for the others, the rest of the group will converge to the center of gravity of the fixed robots.
The structure of the paper is as follows: Section~\ref{problem formulation} poses the problem of consensus with boundary value conditions. Section~\ref{s:main_theorem} presents the main result of the paper, namely Theorem~\ref{main_theorem}.
We provide several consequences for specific network topologies. Section~\ref{gossip} describes two randomized gossip based versions of the algorithm and prove their convergence. Section~\ref{consequences} discusses consequences to advertising in social networks as well as to distributed decision making in sensor networks. We present an experimental study in section~\ref{experimental} to investigate the existence of boundary nodes in a community. Interestingly, our experiment verified the existence of such boundary nodes (who are not affected by their neighbors) in the population under study (a group of undergraduate students). We discuss the sensitivity of sensors networks to malicious nodes and show that even a single malicious node can practically steer the network towards any desired decision. We also point out that consistent bias in a single node can make the entire network useless. Section~\ref{periodic} extends the results to time varying periodic boundary conditions and provide closed form to the limit of the network. Finally, numerical examples are presented in Section~\ref{simulations}.
\section{Problem formulation}
\label{problem formulation}
Consider a network with $K$ boundary nodes $1,\ldots,K$ and $M$ internal nodes ${K+1},\ldots,{K+M}$.
The connectivity graph of the network is denoted by $G=\left(V,E \right)$, where $V =\{1,2, \ldots,{K+M}\}$, and $E \subset V\times V$. Let $\mathbf A$ denote the adjacency (connectivity) matrix of the graph $G$ with elements $a_{i,j}$ given by:
\begin{equation}
a_{i,j}=1 \iff \{i,j\} \in E.
\end{equation}
Define the set of neighbors of node $n$ as $N_n=\left\{m \in V: \left\{n,m\right\}\in E \right\}$.
The aim of this paper is to analyze the dynamics of consensus averaging when the network is subject to boundary value constraints, i.e., when the boundary nodes do not average the estimates of their neighbors and keep their estimates constant.
At time $0$, each node $1 \le n \le M+K$ has an initial value $x_n(0)$. Subsequently, each internal node $n>K$ averages the input from its neighbors.
\begin{equation}
x_n(t+1)=p_{n,n} x_n(t)+\sum_{\left\{m \in N_n \right\}} p_{n,m}(t)x_m(t),
\end{equation}
where $0 \leq p_{i,j} \leq 1$ (for $K<i\leq K+M$ and $1\leq j \leq K+M$) are the averaging weights. The system continues until convergence is achieved. In the existence of boundary nodes, the behavior of the network is very different than the behavior of a standard consensus network.
Before proceeding, let us define the following matrices which are used to analyze the behavior of the system:
Let $\hbox{{\bf P}}_i$ be the weights used in averaging internal nodes with elements \begin{equation} \hbox{{\bf P}}_i(k,l) = p_{k+K, l+K} \quad 1 \leq k,l \leq M, \end{equation} and $\hbox{{\bf P}}_e$ be the weights assigned to the external boundary nodes with elements \begin{equation} \hbox{{\bf P}}_e(k,l) = p_{k+K, l} \quad 1 \leq k \leq M, 1 \leq l \leq K. \end{equation}
The system dynamics is described by the following equations:
\begin{equation}
\begin{array}{lll}
\hbox{Initial conditions: } & & \\
& \mbox{${\bf x}$}_b(0)=[x_1(0),...,x_K(0)]^T, & \\
& \mbox{${\bf x}$}_i(0)=[x_{K+1}(0),...,x_{M+K}(0)]^T & \\
\hbox{System dynamics: } & & \\
\hbox{For all $t>0$: } & & \\
& \mbox{${\bf x}$}_i(t+1)=\hbox{{\bf L}} \mbox{${\bf x}$}(t), &
\end{array}
\label{boundary_TI}
\end{equation}
where
\begin{equation}
\mbox{${\bf x}$}(t)=\left[ \mbox{${\bf x}$}_b^T(t),\mbox{${\bf x}$}_i^T(t)\right]^T=\left[x_1(t),...,x_{M+K}(t)\right]^T,
\end{equation}
and
\begin{equation}\label{eq:L}
\hbox{{\bf L}} = \left[
\begin{array}{c|c}
\mathbf I_{K\times K} & {\bf 0}_{K\times M} \\
\hline \\
\hbox{{\bf P}}_e & \hbox{{\bf P}}_i
\end{array}
\right].
\end{equation}
Here, $A^T$ is used to denote the transpose of matrix $A$. Note, that in this case the boundary conditions are time invariant. Later in Section~\ref{boundary_periodic}, an analysis is presented for the general case. Also note that $\hbox{{\bf L}}$ is a stochastic matrix, since it averages each $x_n(t)$ with its neighbors. For simplicity, in the next section, we analyze the case of constant averaging matrix. This will provide good insight into the general case of time varying averaging. The generalization to randomized gossip algorithms is investigated in the subsequent section.
\section{Asymptotic Analysis of Boundary Value Problems}
\label{s:main_theorem}
This section presents the
main result of the paper, namely the analysis of system (\ref{boundary_TI}). It is shown that the stable point of (\ref{boundary_TI}) is independent of the values stored at the internal nodes of the network. This has important consequences for consensus based networks. Then, some special cases of boundary conditions are discussed and we show that, in general, the network does not converge to a consensus, unless the boundary conditions are constant functions.
Our first goal is to study the stable points of (\ref{boundary_TI}). The following theorem provides existence and uniqueness of the stable point.
\begin{theorem}
Consider the system given in (\ref{boundary_TI}) with $\hbox{{\bf L}}$ defined in (\ref{eq:L}). Assume that the graph $G$ is connected.
and the stochastic matrix $\hbox{{\bf L}}$ describes a Markov chain with absorbing states $1,..,K$ and transient states $K+1,...,N$ (This amount to the fact that all the internal nodes are averaging inputs from nodes $1,...,K$ at some stage). Then the following holds:
\begin{enumerate}
\item $\mathbf I-\hbox{{\bf P}}_i$ is an invertible matrix.
\item
System (\ref{boundary_TI}) always converges to a limit point which is given by:
\begin{equation} \label{main_eq}
\begin{array}{lcl}
\mbox{${\bf x}$}_\infty(\mbox{${\bf x}$}_b(0))&=&\lim_{k {\rightarrow} \infty} \mbox{${\bf x}$}(k) \\
&=& \left[\mbox{${\bf x}$}_b^T, \left(\left( \mathbf I-\hbox{{\bf P}}_i \right)^{-1} \hbox{{\bf P}}_e\mbox{${\bf x}$}_b\right)^T \right]^T.
\end{array}
\end{equation}
\end{enumerate}
\label{main_theorem}
\end{theorem}
\begin{proof}
To prove the first part, note that $\hbox{{\bf P}}_i$ is a weakly sub-stochastic matrix, in which the sum of all rows is less or equal 1 and there exists at least one row which sums to less than 1. To show this, consider the matrix $\hbox{{\bf L}}$ as the transition probability matrix of a Markov chain. This Markov chain has absorbing states $1,..,K$, and the stationary distribution is concentrating on the absorbing states. Examining the powers of the matrix $\hbox{{\bf L}}^k$, this implies that $\lim_{k {\rightarrow} \infty} \hbox{{\bf P}}_i^k={\bf 0}$. Therefore, the spectral radius of $\hbox{{\bf P}}_i$ satisfies $\rho(\hbox{{\bf P}}_i)<1$ and, thus, $\mathbf I-\hbox{{\bf P}}_i$ is invertible.
Consider now, computing the limiting value of (\ref{boundary_TI}) to prove the second part. It follows via induction that for each $k$
\begin{equation}
\hbox{{\bf L}}^k = \left[
\begin{array}{c|c}
\mathbf I & {\bf 0} \\
\hline \\
\left(\mathbf I+\hbox{{\bf P}}_i+...+\hbox{{\bf P}}_i^{k-1}\right)\hbox{{\bf P}}_e & \hbox{{\bf P}}_i^k
\end{array}
\right].
\end{equation}
To see this, note that at $k=1$ this trivially holds. By induction:
\begin{equation}
\hbox{{\bf L}}^{k+1}=\hbox{{\bf L}}^k \hbox{{\bf L}}
\end{equation}
Using block by block multiplication, it is seen that the structure is preserved and the left bottom block satisfies:
\begin{equation}
\sum_{m=1}^{k}\hbox{{\bf P}}_i^m \hbox{{\bf P}}_e=\left(\sum_{m=1}^{k-1}\hbox{{\bf P}}_i^m \hbox{{\bf P}}_e\right) \mathbf I + \hbox{{\bf P}}_i^{k} \hbox{{\bf P}}_e.
\end{equation}
Similarly, the right bottom block is $\hbox{{\bf P}}_i^{k+1}$.
Since the spectral radius of $\rho(\hbox{{\bf P}}_i)<1$, it follows that $\mathbf I-\hbox{{\bf P}}_i$ is invertible.
Let
\begin{equation}
\hbox{{\bf L}}_{\infty}=\lim_{k {\rightarrow} \infty}\hbox{{\bf L}}^k =
\lim_{k {\rightarrow} \infty}
\left[
\begin{array}{c|c}
\mathbf I & {\bf 0} \\
\hline \\
\left(\mathbf I+\hbox{{\bf P}}_i+...+\hbox{{\bf P}}_i^k\right)\hbox{{\bf P}}_e & \hbox{{\bf P}}_i^k
\end{array}
\right]
\end{equation}
Since the spectral radius of $\hbox{{\bf P}}_i$ is less than 1:
\begin{equation}
\hbox{{\bf L}}_{\infty}=
\left[
\begin{array}{c|c}
\mathbf I & {\bf 0} \\
\hline \\
\left(\mathbf I-\hbox{{\bf P}}_i\right)^{-1}\hbox{{\bf P}}_e & 0
\end{array}
\right].
\end{equation}
\begin{claim}
\label{cl:Linf_stochastic}
$\hbox{{\bf L}}_{\infty}$ is stochastic.
\end{claim}
For each $k$, $\hbox{{\bf L}}^k$ is a stochastic matrix. Hence, by a simple continuity argument $\hbox{{\bf L}}_{\infty}$ is also stochastic.
Therefore, for any vector $\mbox{${\bf x}$}(0)=\left[\mbox{${\bf x}$}_b^T(0),\mbox{${\bf x}$}_i^T(0) \right]^T$
we obtain that
\begin{equation}\label{eq:asymptotic}
\lim_{k {\rightarrow} \infty} \mbox{${\bf x}$}(k) = \left[\mbox{${\bf x}$}_b^T, \left(\left( \mathbf I-\hbox{{\bf P}}_i \right)^{-1} \hbox{{\bf P}}_e\mbox{${\bf x}$}_b\right)^T \right]^T.
\end{equation}
An alternative way to observe the equilibrium equation is by writing
\begin{equation} \label{temp}
\mbox{${\bf x}$}_\infty=\hbox{{\bf L}}\mbox{${\bf x}$}_\infty.
\end{equation}
Let $\mbox{${\bf x}$}_i^\infty = \lim_{k\rightarrow \infty} \mbox{${\bf x}$}_i(k)$. Then, from (\ref{temp})
\begin{equation}
\mbox{${\bf x}$}_i^\infty = \hbox{{\bf P}}_e \mbox{${\bf x}$}_b+\hbox{{\bf P}}_i \mbox{${\bf x}$}_i^\infty,
\end{equation}
which translates into
\begin{equation}
\mbox{${\bf x}$}_i^\infty=\left(\mathbf I-\hbox{{\bf P}}_i \right)^{-1} \hbox{{\bf P}}_i \mbox{${\bf x}$}_b
\end{equation}
\end{proof}
This result implies that in the existence of boundary nodes, the network's final state does not depend on the values of the internal nodes. This simple observation has significant implications on consensus based networks which will be analyzed in the subsequent sections.
Here, an interesting property of the limiting function, which has important implications to sensor and social networks, is presented. The following theorem is a
\begin{theorem}
Consider the system given in (\ref{boundary_TI}) with $\hbox{{\bf L}}$ defined in (\ref{eq:L}). Assume that the initial values of boundary nodes $\mbox{${\bf x}$}_b(0)=\mu \mbox{${\bf 1}$}_K$ where $\mbox{${\bf 1}$}_K$ is a $K$ dimensional vector of $1$'s. Then,
$\mbox{${\bf x}$}_\infty(\mbox{${\bf x}$}_b(0))=\mu \mbox{${\bf 1}$}_N$.
\label{thm:constant}
\end{theorem}
\begin{proof} Recall from Claim~\ref{cl:Linf_stochastic} that, $\hbox{{\bf L}}_{\infty}$ is stochastic. Therefore, the vector $\mbox{${\bf 1}$}_N$ is an eigenvector of $\hbox{{\bf L}}_\infty$ with eigenvalue $1$ and $\left(\mathbf I-\hbox{{\bf P}}_i\right)^{-1}\hbox{{\bf P}}_e \mu \mbox{${\bf 1}$}_K = \mu\mbox{${\bf 1}$}_{N-K}$. \end{proof}
The above theorem implies that if the value on the boundary is constant the network will converge to the same value.
This result has significant implications to both social network and decision making in sensor networks which are discussed, in details, in Section~\ref{consequences}.
The operator $\hbox{{\bf L}}$ averages each $x_n(t)$ with its neighbors in the graph. Hence, in steady state, $\mbox{${\bf x}$}_\infty(\mbox{${\bf x}$}_b(0))$ is a generalized harmonic function on the graph $G$. Furthermore, by the above analysis, the boundary conditions completely determine the final state. It follows that, if the values of the
boundary nodes are not constant, the network will not converge to a consensus at all, as the following example shows:
\begin{example}
Assume that $G$ is a line graph as described in figure \ref{line_graph} where $V=\left\{0,...,N-1 \right\}$ and $\{i,j\}\in E \iff j=i+1$.
and for $0<n<N-1$ and $0<\alpha<1$:
\begin{equation}
x_n(t+1)=\alpha x_{n}(t)+ \frac{1-\alpha}{2} \left(x_{n-1}(t)+x_{n+1}(t)\right).
\label{line_dynamics}
\end{equation}
Assume that the boundary nodes are $0,N-1$ and that the boundary values are $x_0(t)=a, x_{N-1}(t)=b$ and $a \neq b$. The network converges to a linear function where $x_n^{\infty}=a+\frac{n(b-a)}{{N-1}}$.
To see this note that by (\ref{line_dynamics}) for $0<n<N-1$ we have at steady state:
\begin{equation}
x_n^{\infty}=\frac{1}{2} \left(x_{n-1}^{\infty}+x_{n+1}^{\infty}\right).
\end{equation}
$x_n^{\infty}=a+\frac{n(b-a)}{{N-1}}$ satisfies the asymptotic equation and by (\ref{eq:asymptotic}) there is a unique stable point which is the linear function.
\begin{figure}[htb]
\centering{\includegraphics[width = .4\textwidth]{line_network1.pdf}}
\caption{Linear network with boundary conditions.}
\label{line_graph}
\end{figure}
\end{example}
\section{Randomized gossip algorithms}\label{gossip}
This section extends the results of Section~\ref{s:main_theorem} to the case of random gossip algorithms. Here, two types of network operation, sensor polling and pair-wise averaging, are considered.
{\em Sensor polling:} Each sensor decides to poll all its neighbors with probability $p$. Once it decided to poll the neighbours, it computes its new state by weighted averaging with fixed weights given by the matrix $\hbox{{\bf L}}$ (for example based on the degree or any other predetermined set of weights). This is done independently for each sensor. For simplification purposes, it is assumed that sensors update their measurement synchronously.
In the sensor polling, each sensor averages its neighbors (applies a row of $\hbox{{\bf L}}$ to the data) with probability $p$ and otherwise, does not change its value (applies a row of the identity matrix) with probability $1-p$. Therefore, we obtain
\begin{equation}
\mathbb{E}\left(\hbox{{\bf L}}(k)\right)= p \hbox{{\bf L}} +(1-p) \mathbf I,
\end{equation}
where $\mathbb{E}(\cdot)$ denotes the expectation. Since sensors act independently over time (the choices are independent between sensors and times), the following expression is obtained for value of each node at time $k$:
\begin{equation}
\mathbb{E}\left(\mbox{${\bf x}$}(k)\right)=\left(p \hbox{{\bf L}} +(1-p) \mathbf I\right)^k \mbox{${\bf x}$}(0)
\end{equation}
By the convexity of the stochastic matrices, $\mathbb{E}\left(L(k)\right)$ is stochastic and the sub-block
\begin{equation}
\mathbb{E}\left(\hbox{{\bf P}}_i(k)\right)=p \hbox{{\bf P}}_i +(1-p) \mathbf I,
\end{equation}
is also weakly sub-stochastic.
Following the same argument as in Section~\ref{main_theorem}, we obtain that with probability~1
\begin{equation}
\begin{array}{lcl}
\mathbb{E}\left(\mbox{${\bf x}$}_\infty(\mbox{${\bf x}$}_b(0))\right)&=&\lim_{k {\rightarrow} \infty} \mathbb{E}\left(\mbox{${\bf x}$}(k)\right)\\
&=& \left[\mbox{${\bf x}$}_b^T, \left( \left(\mathbf I-\hbox{{\bf P}}_i \right)^{-1} \hbox{{\bf P}}_e\mbox{${\bf x}$}_b\right)^T \right].
\end{array}
\label{eq:gossip_const}
\end{equation}
Similar to Theorems~\ref{main_theorem} and~\ref{thm:constant}, when the boundary conditions are constant, the network converges in the mean to the same constant, and generally with a given boundary conditions, the network converges to the harmonic function defined by (\ref{eq:gossip_const}).
{\em Pair-wise averaging:} The second model is the pairwise interaction model where the matrix $\hbox{{\bf L}}$ is derived from a doubly stochastic matrix. In this model, two internal nodes choose to update their values by averaging, while whenever an interaction occurs between external node and a boundary node, only the internal node updates its value; that is, the value of boundary nodes always remains fixed.
The process can be viewed as follows: We define a probability distribution over pairs $1 \le i,j \le N$ (where $N = M+K$) given by $\pi_{i,j}$ and an averaging constant $0<\alpha_{i,j} <1$. For each $i \neq j$, a weighting matrix $\hbox{{\bf W}}_{i,j}$ is defined as
\begin{equation}
\left[\hbox{{\bf W}}_{i,j}\right]_{m,n}=\left\{
\begin{array}{lcl}
\delta (m-n)& \hbox{if} & \{m,n\} \not \subseteq \{i,j\} \\
\alpha_{i,j}& \hbox{if} & \{m,n\}=\{i,j\}\\
1-\alpha_{i,j}& \hbox{if} & m=n= (i \hbox{\ or \ }j),
\end{array}
\right.
\end{equation}
where $\delta(\cdot)$ is Dirac delta function. At each step a pair is chosen randomly according to $\pi_{i,j}$ and the matrix $\hbox{{\bf W}}_{i,j}$ is applied to the data, i.e. nodes $i,j$ exchange their data and average it with weights given by $\alpha_{i,j}, 1-\alpha_{i,j}$, with the exception that boundary nodes do not perform the averaging but maintain their original data.
By the convexity of the doubly stochastic matrices
\begin{equation}
\mathbb{E}\left(\hbox{{\bf W}}(k)\right)=\sum_{i \neq j} \pi_{i,j} \hbox{{\bf W}}_{i,j},
\end{equation}
is also doubly stochastic.
Furthermore, the operation of the system is given by a random matrix of the form
\begin{equation}\label{eq:Lpairwise}
\hbox{{\bf L}}(k) = \left[
\begin{array}{c|c}
\mathbf I_{K\times K} & {\bf 0}_{K\times M} \\
\hline \\
\hbox{{\bf W}}_e(k) & \hbox{{\bf W}}_i(k)
\end{array}
\right].
\end{equation}
where the lower part of the matrix is a random matrix which consists of the lower $M = N-K$, (recall that $M$ denotes the number of internal nodes) rows of $\hbox{{\bf W}}(k)$.
Using the same argument as before and assuming that there exists at least one pair $(i,j)$ with $\pi_{i,j}>0$ where $i$ is a boundary node and $j$ is an internal node, we obtain that
\begin{equation}
\hbox{{\bf P}}_e = \mathbb{E}\left(\hbox{{\bf W}}_e(k)\right)\neq \mbox{${\bf 0}$}
\end{equation}
and
\begin{equation}
\hbox{{\bf P}}_i = \mathbb{E}\left(\hbox{{\bf W}}_i(k)\right)
\end{equation}
is a sub-stochastic matrix. This suffices to prove convergence in the mean:
\begin{equation}\label{eq:pair}
\lim_{k {\rightarrow} \infty} \mathbb{E}\left(\mbox{${\bf x}$}(k)\right) = \left[\mbox{${\bf x}$}_b^T, \left( \left(\mathbf I-\hbox{{\bf P}}_i \right)^{-1} \hbox{{\bf P}}_e\mbox{${\bf x}$}_b\right)^T \right]^T.
\end{equation}
\section{Consequences for networks - Experimental Studies on Humans}
\label{consequences}
In this section, the implications of the main results of this paper (which are presented in Section~\ref{s:main_theorem}) on consensus in social and sensor networks are presented. First, the control of beliefs in social networks, which has significant impact on advertisement technology as well as steering opinions in a Bayesian setup, is discussed. We, further, present the results of an experimental study which verifies the existence of such boundary nodes in social networks. Then, an example is provided where a single malicious node can completely destroy the detection capability of a sensor network with distributed averaging.
\subsection{Implications to social networks}\label{sec:SNs}
Here, an example is provided to discuss the importance of the boundary value problems in the asymptotic agreement of agents in social networks. The propagation of beliefs and learning in non-Bayesian social networks is investigated in \cite{SL}. One protocol to experiment learning in social networks is consensus formation over graphs \cite{consensus, diffusion}. In this subsection, we investigate the effect of the boundary value problems in such consensus networks. Consider a social network where individuals (agents) interact to update their belief about an underlying state of nature simply by evaluating the weighted average of the beliefs of their neighbors at previous time-instant. Assume that there exists two types of agents in the social network: (i) ordinary agents (internal nodes) whose beliefs change according to the beliefs of their neighbors, and (ii) ``advertiser"s (boundary nodes). The belief of advertisers are not affected by their neighbors, thus, remain fixed over time. These advertisers can be viewed as boundary nodes that inject malignant beliefs into the social network. Such a problem, the effect of advertisers (they can also be considered as malicious agents) on the consensus formation in social networks is not investigated in the literature. Here, we can simply show that the belief dynamics in such social networks can be modeled with (\ref{boundary_TI}). Consequently, the asymptotic agreement of agents in social network can be computed from (\ref{eq:asymptotic}). This means that the asymptotic agreement of the network, no matter of the initial beliefs of the internal nodes, depends only on the malignant beliefs that are being injected into the network by the advertisers. From this, the asymptotic agreement of social network can be controlled and shaped by choosing proper values for the (initial) belief of the advertisers $\mbox{${\bf x}$}_b(0)$ in the presence of adverting agents (boundary conditions).
\subsection{Existence of boundary nodes in human social networks: An experimental study}\label{experimental} Individuals in a society who are not affected by their neighbors (they stand firm on their own belief) can be considered as boundary nodes in that society.
In collaboration with the Department of Psychology, University of British Columbia (UBC), we conducted an experimental study of the existence of such boundary nodes on a group of undergraduate students at UBC during the period October-November 2013. Below we report on these experimental results\footnote{We acknowledge Prof. Alan Kingstone and Dr. Grayden Solman of the Department of Psychology, University of British Columbia, for conducting the psychology experiment.}.
\paragraph*{Experimental Setup} The experimental study involved 1658 individual experiments. Each individual experiment comprised two participants.
The two participants were asked to perform a perceptual task interactively. Two arrays of circles were given to each pair of participants, then, they were asked to judge which array had the larger average diameter. One member was chosen randomly and started the experiment and chose either left side or right side as his judgment. Thereafter, each member saw their partner's previous response and his own previous judgment prior to making their own judgment; thereby providing a means for measuring social influence. The participants continued choosing actions according to this procedure until the experiment terminated. The experiment terminated
when the response of each of the two participants did not change for three successive iterations (the two participants did not necessarily have to agree
for the experiment to terminate). In this experimental study, each participant chose an action $a \in \mathbf{A} = \{0,1\}$; $a = 0$ when he judged that the left array of circles had the larger diameter and $a = 1$ when his judgments was that the right array of circles had the larger diameter. In each experiment, judgments (actions) of participants are recorded along with the amount of time taken by each participant to make that judgment.
\begin{figure*}[!t]
\centerline{
\includegraphics[width=.7\textwidth]{experiment.pdf}}
\caption{Two arrays of circles were given to each pair of participants. Their task is to interactively determine which side (either left or right) had the larger average diameter. In this example, the average diameter of the left array of circles is 8.4 mm and the right array is 9.1 mm.}
\label{Fig:SocialSensor}
\end{figure*}
\paragraph*{Experiment Results} Surprisingly, among 3316 individuals who participated in this experiment, 1336 participants (around 40\%) did not change their judgements after observing the action of their partners (these participants can be viewed as boundary nodes), while the other 60\% changed their initial judgment and got influenced by the action of their partners (internal nodes). Fig.~\ref{Fig:Samplepath} shows the sample path of two participant in a same group, Participant~1 is an internal node while Participant~2 is a boundary node.
\begin{figure}[htb]
\centerline{
\includegraphics[width=.5\textwidth]{SamplePath.pdf}}
\caption{Actions of two participants in a group at different epochs. Participant~1 can be considered as an internal node and Participant~2 can be viewed as a boundary node.}
\label{Fig:Samplepath}
\end{figure}
We, further, investigate the time taken by each participant to make his judgment. Let $\mu_{\rm judg.}$ and $\sigma_{\rm judg.}$ denote the mean and the standard deviation of the time taken by participants to make their judgments in milliseconds. The results of our experimental study, which are presented in Table~I, show that the internal nodes, in average, require more time to make their judgments compared to the boundary nodes; this is quite intuitive from the fact that the boundary nodes stand firm on their decisions and ignore the judgment of their partners and thus require less time to make their judgments.
\begin{table}[h]
\begin{center}
\begin{tabular}{ | c || c | l | l | }
\hline
Type of nodes & relative frequency & $\mu_{\rm judg.}$ & $\sigma_{\rm judg.}$ \\ \hline
Internal & 40 \% & 1058 ms & 315 ms \\ \hline
Boundary & 60 \% & 861 ms & 403 ms\\ \hline
\end{tabular}
\end{center}
\label{table1}
\caption{The frequency of the internal and the boundary nodes in a community of 3316 undergraduate students of the University of British Columbia along with the statistics of the time required by participants (of both types) to make their judgments in milliseconds.}
\end{table}
This psychology experiment illustrates the importance of investigating the consensus of networks with boundary nodes in social networks.
\subsection{Implications to sensor networks}
In this subsection, an example is provided which shows the sensitivity of consensus networks to malicious attacks, and even for an unintended bias in the computation of the mean in one of the elements of the network. To observe the effect, consider a binary distributed detection problem. Two hypotheses $H_0,H_1$ are given and each sensor $0 \le n \le N-1$ measures a realization $x(n)$
\begin{equation}
x(n)\sim\left\{
\begin{array}{ll}
P\left(x(n)|H_0\right) & H_0 \\
P\left(x(n)|H_1\right) & H_1
\end{array} \right.
\end{equation}
Here, $P(\cdot)$ denote the probability of an event. Assume that $x(n)$ are conditionally independent given $H_0$ and $ H_1$. To obtain an optimal decision, each sensor needs to compute the average log-likelihood over the network.
The log likelihood ratio is given by:
\begin{equation}
L(\mbox{${\bf x}$})=\log \frac{P\left(x(0),...,x(N-1)|H_1\right)}{P\left(x(0),...,x(N-1)|H_0\right)}
\end{equation}
which can be computed by conditional independence as
\begin{equation}
L(\mbox{${\bf x}$})=\sum_{n=0}^{N-1} \log \frac{P\left(x(n)|H_1\right)}{P\left(x(n)|H_0\right)}
\end{equation}
This detector can be easily computed using a gossip algorithm. However given any threshold designed for a fixed probability of false alarm $\tau$,
a malicious node, that is aware of the desired detection threshold $\tau$, can choose a value $\mu<\tau$ and decrease the probability of detection to $0$ by steering the network to $\mu$.
The system matrix is now given by
\begin{equation}\label{eq:A2}
\hbox{{\bf L}} = \left[
\begin{array}{c|c}
1 & {0} \\
\hline \\
\mu\mbox{${\bf e}$}_1 & \hbox{{\bf P}}_i
\end{array}
\right],
\end{equation}
where index $1$ is used to denote the faulty node.
By Theorem~\ref{thm:constant}, the limiting value at the network nodes, given by
\begin{equation}
\mbox{${\bf x}$}_{\infty}=\mu \mbox{${\bf 1}$}_N<\tau,
\end{equation}
is constant. Once the network converged to consensus, all nodes will agree that there is no target.
The level of confidence can be made arbitrarily high by choosing smaller $\mu$.
Note that the existence of a consistent bias in a single node can result in similar erroneous conclusion, or if the bias is larger than the detection threshold, it will generate false alarms.
\section{Networks with time periodic boundary conditions}
\label{periodic}
In this section, the analysis of Section~\ref{s:main_theorem} is extended to the case where the boundary conditions are periodic with period $\tau$.
Similar to Theorem~\ref{main_theorem}, consider the following system:
Let $\tau>0$ be fixed.
\begin{equation}
\begin{array}{ll}
\hbox{Initial conditions: } & \\
&\hspace{-1cm} \mbox{${\bf x}$}_i(0)=[x_{K+1}(0),...,x_{M+K}(0)]^T \\
\hbox{Boundary conditions: } & \\
\forall t \le \tau-1,n: &\hspace{-1cm} \mbox{${\bf x}$}_b(t+n\tau)=[x_1(t),...,x_K(t)]^T, \\
\hbox{System dynamics: }\\
\forall t > 0: & \hspace{-1cm} \mbox{${\bf x}$}_i(t+1)=\hbox{{\bf P}}_e \mbox{${\bf x}$}_b(t) + \hbox{{\bf P}}_i\mbox{${\bf x}$}_i(t).
\end{array}
\label{boundary_periodic}
\end{equation}
The matrix $\left[\hbox{{\bf P}}_e, \hbox{{\bf P}}_i \right]$ is stochastic, which implies that $\hbox{{\bf P}}_i$ is weakly sub-stochastic. As a consequence, $\rho(\hbox{{\bf P}}_i)<1$.
Similar to the time invariant case, we can show inductively the following theorem:
\begin{theorem}
The state of the internal system $\mbox{${\bf x}$}_i(k)$ at time $k$ is given by
\begin{equation}
\label{finite_periodic}
\mbox{${\bf x}$}_i(k)=\sum_{m=0}^{k-1} \hbox{{\bf P}}_i^m \hbox{{\bf P}}_e \mbox{${\bf x}$}_b(k-m \mod \tau)+\hbox{{\bf P}}_i^k \mbox{${\bf x}$}_i(0)
\end{equation}
Furthermore,
\begin{equation}
\label{lim_periodic}
\lim_{k{\rightarrow} \infty} \mbox{${\bf x}$}_i(k) = \sum_{m=0}^{\tau-1} \left(\mathbf I-\hbox{{\bf P}}_i^\tau\right)^{-1} \hbox{{\bf P}}_i^m \hbox{{\bf P}}_e \mbox{${\bf x}$}_b(m).
\end{equation}
\end{theorem}
\begin{proof} The proof of (\ref{finite_periodic}) is an easy induction. To show the second part, one splits the sequence in (\ref{finite_periodic}) according to the value of $m \mod \tau$, and noting that we have a geometric series with factor $\hbox{{\bf P}}_i^\tau$ for each subsequence, beginning with $\hbox{{\bf P}}_i^m \hbox{{\bf P}}_e \mbox{${\bf x}$}_b(m)$. Since $\rho(\hbox{{\bf P}}_i)<1$ the part depending on initial conditions in the interior part of the network tends to $\bf 0$ exponentially fast. \end{proof}
\section{Numerical Examples}
\label{simulations}
In this section, numerical examples are provided to verify the results of Sec.\ref{gossip} and Sec.\ref{consequences}. In the first numerical study, we investigate the dynamics of consensus averaging in networks with boundary constraints. A network comprising of $N = 100$ nodes ($K = 50$ boundary nodes and $M = 50$ internal nodes) is considered. In this network, internal nodes (in contrast to boundary nodes whose values remain fixed) average over their neighbors to update their values. Neighbors of each node and weights used in the averaging (that is matrix $\hbox{{\bf L}}$) are chosen randomly. The initial values of nodes $\mbox{${\bf x}$}(0)$ are simulated from normal distribution; that is $\mbox{${\bf x}$}_i(0) \sim \mathbf{N}(0,5)$ and $\mbox{${\bf x}$}_b(0) \sim \mathbf{N}(0,1)$. Fig.\ref{fig1} compares $\mbox{${\bf x}$}_\infty(\mbox{${\bf x}$}_b(0))$ obtained via (\ref{main_eq}) in Theorem~\ref{main_theorem} with $\mbox{${\bf x}$}(t)$ from simulation. Three different entries of $\mbox{${\bf x}$}_\infty(\mbox{${\bf x}$}_b(0))$ and $\mbox{${\bf x}$}(t)$ are shown in Fig.\ref{fig1}.
\begin{figure}[htb]
\centering{\includegraphics[width = .5\textwidth]{Figure1.pdf}}
\caption{Dynamics of beliefs in a network with boundary constraints.}
\label{fig1}
\end{figure}
In the next example, we study the effect of advertisers (boundary nodes) in formation of consensus in a network. A network with $N = 50$ nodes ($K = 1$ boundary node) is considered. In this scenario, the advertising node (boundary node) injects a fixed belief into the network (that is, it is not affected by belief of its neighbors) and the internal nodes average over their neighbors to update their beliefs. Similar to the previous example, the neighbors and the averaging weights are chosen randomly. Fig.~\ref{fig2} depicts $\mbox{${\bf x}$}(t)$ at each iteration for one of the non-advertiser (internal) agents for different values of initial belief $\mbox{${\bf x}$}_b(0)$. As can be seen in Fig.~\ref{fig2}, the beliefs of internal nodes converge to the initial belief of the advertising node (boundary node). This is quite interesting result in social networks, it shows that by the means of advertising node in a social network, we can control the consensus of the whole network.
\begin{figure}[htb]
\centering{\includegraphics[width=.5\textwidth]{Figure2.pdf}}
\caption{ Effect of an advertising node (boundary node) in the asymptotic belief of internal nodes.}
\label{fig2}
\end{figure}
To study the rate of convergence of beliefs to the asymptotic value, a network similar to the previous example with $N = 50$ nodes and $K = 1$ boundary node is considered. Fig.\ref{fig4} shows the value of $\mbox{${\bf x}$}(T)$ for one of the internal nodes versus the initial belief of the advertising node $\mbox{${\bf x}$}_b(0)$ for different values of $T$. For sufficiently large values of $T$ (for example $T = 10000$), belief of the internal node reaches its asymptotic value (which is equal to $\mbox{${\bf x}$}_b(0)$), therefore, $\mbox{${\bf x}$}(T) \approx \mbox{${\bf x}$}_b(0)$; the slope of the solid line which corresponds to $T= 10000$ is very close to one. Fig.~\ref{fig4} reveals that for smaller values of $T$, beliefs of the internal nodes converge faster to their asymptotic values when the initial belief of the advertising node $\mbox{${\bf x}$}_b(0)$ is larger.
\begin{figure}[htb]
\centering{\includegraphics[width=.5\textwidth]{Figure4.pdf}}
\caption{Belief of an internal node at time $T$ versus the initial belief of the advertising node $\mbox{${\bf x}$}_b(0)$.}
\label{fig4}
\end{figure}
Fig.\ref{fig3} illustrates the average of the distance between $\mbox{${\bf x}$}_\infty(\mbox{${\bf x}$}_b(0))$ and $\mbox{${\bf x}$}(t)$ for $t = 1,2,\ldots, 2000$ versus the initial belief of the advertising nodes $\mbox{${\bf x}$}_b(0)$. Similar to the previous example, it can be inferred from Fig.\ref{fig3} that the distance between beliefs of internal nodes and the initial belief of the advertising node is lower when $\mbox{${\bf x}$}_b(0)$ is larger. In other words, the belief of internal nodes converge faster to the asymptotic value when the bias (the belief of the advertising node $\mbox{${\bf x}$}_b(0)$) is large enough.
\begin{figure}[htb]
\centering{\includegraphics[width=.5\textwidth]{Figure3.pdf}}
\caption{The average distance between beliefs of an internal node at different times and the consensus of the network. }
\label{fig3}
\end{figure}
Here, we study the effect of the boundary value constraints in random gossip algorithms. A network comprising of $N = 100$ sensors ($K = 50$ boundary nodes and $M = 50$ internal node) is considered. In this scenario, each sensor independently decides to poll its neighbors with probability $p$; it averages over the neighbors to update its belief. In other words, with the probability $1 - p$ its belief remains unchanged. Neighbors of each sensor and the weights used in averaging; that is matrix $\hbox{{\bf L}}$ ($\hbox{{\bf P}}_e$ and $\hbox{{\bf P}}_i$) is chosen randomly. The initial values for beliefs $\mbox{${\bf x}$}$ are simulated from normal distribution; that is $\mbox{${\bf x}$}_i(0) \sim \mathbf{N}(0,5)$ and $\mbox{${\bf x}$}_b(0) \sim \mathbf{N}(0,1)$. This scenario is simulated $1000$ times and the average of an internal sensor's beliefs $E(\mbox{${\bf x}$}_\infty(\mbox{${\bf x}$}_b(0)))$ for four different values of $p$, ($p \in \{0.1, 0.2, 0.5, 0.8\}$) are depicted in Fig.~\ref{Fig:gossip} along with the expected consensus of the network computed via (\ref{eq:gossip_const}). As can be seen in Fig.~\ref{Fig:gossip}, the network converges to the harmonic function defined by (\ref{eq:gossip_const}) in the existence of boundary constraints.
\begin{figure}[h]
\centerline{
\includegraphics[width=.5\textwidth]{Gossip.pdf}}
\caption{Beliefs of an internal node in a network with sensor polling in the existence of boundary constraints for different probabilities of polling.}
\label{Fig:gossip}
\end{figure}
In the next example, a pairwise interaction model for the network is considered. Similar to the previous example, a network comprising of $N = 100$ nodes ($K = 50$ boundary nodes and $M = 50$ internal node) is considered. At each time, two nodes are randomly selected according to a uniform probability distribution over pairs. Then, these nodes (given that none of them are boundary nodes) update their beliefs by averaging. In the interaction between an internal node and a boundary node, only the internal node updates its belief, and in the interaction between two boundary nodes, none of them update their beliefs. We assume that for all pairs of nodes, the averaging weight is fixed, i.e., $\alpha_{i,j} = \alpha$ for $ 1 \leq i,j \leq N$ and $0< \alpha < 1$. The initial values for beliefs $\mbox{${\bf x}$}$ are simulated from normal distribution; that is $\mbox{${\bf x}$}_i(0) \sim \mathbf{N}(0,5)$ and $\mbox{${\bf x}$}_b(0) \sim \mathbf{N}(0,1)$. This scenario is simulated $1000$ times and we average over beliefs of one the internal sensors. $E(\mbox{${\bf x}$}_\infty(\mbox{${\bf x}$}_b(0)))$ for four different values of $\alpha$, ($ \alpha \in \{0.1, 0.2, 0.5, 0.8\}$) along with the expected consensus of the network computed via (\ref{eq:pair}) are depicted in Fig.\ref{Fig:pair}. As can be seen in Fig.~\ref{Fig:pair}, the network converges in the mean to the asymptotic value defined by (\ref{eq:pair}) in the existence of boundary constraints.
\begin{figure}[h]
\centerline{
\includegraphics[width=.5\textwidth]{Pairwise.pdf}}
\caption{Beliefs of an internal node in a network with pairwise interaction model between nodes in the existence of boundary constraints for different averaging weights.}
\label{Fig:pair}
\end{figure}
\section{Conclusions}
This paper has examined the effect of boundary valued conditions on consensus. The main result
was Theorem~\ref{main_theorem} which gave sufficient conditions for the
consensus only to depend on the boundary nodes and not the internal nodes. Applications in sensor polling and human interactions (with experimental data) were described.
Numerical examples were provided to illustrate the main result.
\section*{Acknowledgment}
We would like to acknowledge our colleagues in the Department of Psychology, University of British Columbia, Prof. Alan Kingstone and Dr. Grayden Solman, for conducting the psychology experiment discussed in Sec.\ref{sec:SNs}.
\bibliographystyle{IEEEtran}
|
\section{Introduction}
Finding the global behavior of an analytic function in terms of its Taylor
coefficients is a notoriously difficult problem, in fact one which is impossible in full generality since undecidable statements can be formulated in these terms. However, a very interesting and quite general criterion for the disk of convergence of a Taylor series to coincide with its maximal domain of analyticity was recently discovered by Breuer and Simon \cite{Simon} (see also \cite{Sauzin1}). The present paper complements this result by finding criteria on the Taylor coefficients, say at zero, for the associated analytic function not to have natural boundaries and to belong to the
class $\mathcal{M}$ of functions analytic in the complex plane with finitely
many cuts and with algebraic behavior at infinity (see Definition \ref{Def21}
below). Our condition is that the coefficients $c_k$ admit generalized Borel summable (or Ecalle-Borel summable, EB) transseries in $k$. Many general classes of problems in analysis are known to have EB transseries solutions. For details on generalized Borel summability, transseries and resurgence see \cite{Book,Ecalle-book,Ecalle,Ecalle2,Ecalle3}.
In particular it is known \cite{Braaksma, Ecalle, Book} that, if the $c_k$ are solutions of generic linear or
nonlinear recurrence relations of arbitrary but finite order with analytic coefficients, then they are
are EB-summable. Recurrence relations exist for instance
when the coefficients are obtained by solving differential equations by power
series. In a forthcoming paper we show that the Taylor coefficients of the Borel transform of solutions of generic systems of linear or nonlinear ODEs (in the setting of \cite{Duke}) also admit EB summable transseries. The global analytic structure of the Borel transform is crucial in understanding the monodromy of solutions of such equations.
We also extend our procedure to analyze the global behavior of entire functions, and to formal series, giving criteria directly on the coefficients for the formal series to be Borel summable.
Globally reconstructing function from its Taylor coefficients, when these admit EB summable transseries is effective, constructive and explicit -in the sense of producing integral representations far easier to analyze than the sums; this provides a new summation procedure, generalizing in some ways the Poisson summation formula.
We recently used this approach to analyze a class of linear PDEs with variable coefficients,
\cite{CHT}. One can obtain explicit integral representations for solutions of ODEs not known to be solvable such as
\begin{equation}
\label{eq:thfilm}
A\eta^2f^{(4)} +2A\eta f'''+\frac12 \eta f'-(1+a) f=0
\end{equation}
arising as the scaling pinching profile $ h\sim (t_c-t)f(x(t_c-t)^{-1/2})$ of the thin film
equation, $$h_t+(h_{xxx} h)_x=0, \ h\sim (t_c-t)f(x(t_c-t)^{-1/2})$$
where $t_c$ is the singularity time, where the interest is in the solution {\em analytic at zero, $f_0$}. Let $A\neq 0$, and $a+\frac{1}{2} \notin \mathbb{N}$. While it is not clear how to obtain representations of $f_0$ itself, the Taylor coefficients of $f_0$ are explicit. From this, our technique introduced in \S\ref{entire} by first considering the Laplace transform $F(p) = \mathcal{L}f (p)$, which is expressible in terms of integrals of Whittaker functions: with
\begin{equation}\label{whit}
F(p)= C e^{-\frac{p^{-2}}{8A}}p^{-3/2} \left[M_{-a-\frac{3}{4},\frac{1}{4}}\left(\frac{p^{-2}}{4A}\right)+ \frac{a\Gamma(a)}{2\sqrt{\pi}} W_{-a-\frac{3}{4},\frac{1}{4}}\left(\frac{p^{-2}}{4 A}\right) \right]
\end{equation}
we have $f = \mathcal{L}^{-1} F$. The proof of \eqref{whit} is sketched in \S\ref{4thoode}.
In particular, if $c_k=\varphi(k)$ where the function $\varphi$, defined in the right
half plane, is inverse Laplace transformable and its inverse Laplace
transform $\mathcal{L}^{-1} \varphi$ can be calculated in closed form, the function
$f$ has integral representations in terms of $\varphi$. We will use some particularly simple examples for illustration. For the first one, the generalized Hurwitz zeta function, our procedure quickly yields one of the known integral representations. For the other three, our procedure yields integral representations while the global behavior of $\sum_k c_k z^k$ does not follow in any other obvious way:
\begin{equation}
\label{eq:ln2}
c_k^{[1]}=\frac{1}{(k+a)^b}; \ \ c_k^{[2]} =\frac{1}{k^b+\ln
k},\ \ c_k^{[3]}= \frac{1}{k^{k+1}}, \
c_k^{[4]}=e^{\sqrt{k}}, \ (a,b>0)
\end{equation}
We find that
\begin{equation}
\label{eq:sum0}
f_1(z):=\sum_{k=1}^\infty c_k^{[1]}z^k= \frac{ z}{\Gamma(b)}\int_0^{\infty}\frac{[\ln(1+t)]^{b-1} \mathrm{d}t}{(1+t)^{a+1}(t-(z-1))}
\end{equation}
On the first Riemann sheet $f_1$ has only one singularity, at $z=1$, of
logarithmic type, and $f_1=o(z)$ for large $z$. General Riemann surface
information and monodromy follow straightforwardly from \eqref{eq:sum0}. A
similar complex analytic structure is shared by $f_2=\sum_{k=1}^\infty
c_k^{[2]}z^k$, which has one singularity at $z=1$ where it is analytic in
$\ln(1-z)$ and $(1-z)$; more precisely,
\begin{equation}
\label{intlog}
f_2(z)=-\frac{1}{2\pi i} \frac{\ln \ln
z}{z}\oint_{0}^{\infty}\frac{e^{-u\ln(z)}}{(-u)^{b}+\ln (-u)}\mathrm{d}u +E(z)
\end{equation}
see Definition \ref{def2}, where $E$ is entire.
The function $f_3(z)=\sum_{k=1}^\infty
c_k^{[3]}z^k$ is entire; questions answered regard
say the behavior for large negative $z$ (certainly not obvious from the series) or the asymptotic location of
zeros. It will follow that $f_3$ can be written as
\begin{equation}
\label{eq:invl}
f_3(z)=\int_0^{\infty}(1+u)^{-1}G(\ln(1+u))\left[\exp\left({\frac{ze^{-1}}{1+u}}\right)-1\right]\mathrm{d}u
\end{equation}
where $G(p)=s'_2(1+p)-s'_1(1+p)$ and $s_{1,2}$ are two branches of the
functional inverse of $s-\ln s$, cf. \S\ref{S115}. Using the integral
representation of $f_{3}$, its behavior for large $z$ can be obtained from
(\ref{eq:invl}) by standard asymptotics methods; in particular, for large
negative $z$, $f_3$ behaves like a constant plus $z^{-1/2}e^{-z/e}$ times a
factorially divergent series (whose terms can be calculated).
For
$ c_k^{[4]}$ we find
\begin{equation}
\label{ff4}
f_4(z)=\sum_{k=1}^\infty c_k^{[4]}z^k =-\frac{z}{2\sqrt{\pi}}\int_{C_1}\frac{p^{-3/2}e^{-\frac{1}{4p}}\mathrm{d}p}{e^p-z}
\end{equation}
where $C_1$ is a spiral $S_1$ followed by $[1,\infty)$, where $S_1$ starts at $0$ and ends at $1$, and is given in polar coordinates by $r = \theta e^{2\pi i \theta}$, $\left(\theta \in [0,1]\right)$.
We also show that Borel summation of divergent series or transseries of
resurgent functions with finitely many Borel-plane singularities, as well as
the Abel-Plana version of the Euler-Maclaurin summation formula (see also
\cite{Stavros1}) can be derived by a natural extension of our analysis. Another illustration is obtaining the closed form Borel summed formula for $\ln \Gamma$, cf. \eqref{eq:lngamma} below.
A separate category is represented by lacunary series. Their coefficients do
not satisfy our assumption; however a slightly different approach allows for a
detailed study of the associated functions as the natural boundary is
approached, \cite{Advances}.
\section{Main results}
$ $\ \ \ {\em A first class of problems} is finding the location and type of
singularities in $\mathbb{C}$ and the behavior for large values of the variable of
functions given by series with finite radius of convergence (Theorem
\ref{T1}), such as the first, second and fourth in (\ref{eq:ln2}).
{\em The second class of problems} amenable to the techniques presented
concerns the behavior at infinity (growth, decay, asymptotic location of
zeros etc.) of entire functions presented as Taylor series (Theorem \ref{T2}).
{\em The third class of problems} is essentially the converse of the two above: given a
function that has analytic continuation on some Riemann surface, how is this
reflected on $c_k$? (Theorem \ref{T1}.)
{\em The fourth class of problems} is to determine Borel summability of
series with zero radius of convergence such as
\begin{equation}
\label{eq:entire}
\tilde{f}_5=\sum_{n=0}^{\infty}n^{n+1} z^n
\end{equation}
in which
the coefficients of the series are analyzable (Theorem \ref{T3}).
\begin{Definition}\label{Def21}
{\rm Let $\{a_j : 1\leq j \leq N\}$ be a set of nonzero complex numbers with distinct arguments. Let $\mathcal{M}$ consist of the functions algebraically bounded
at $\infty$ and analytic in $\mathbb{C}\setminus \bigcup_{j=1}^N
\{a_j t: t\geq 1\}$ . By dividing by a power of $z$ and subtracting out the principal part ({\em i.e.,}the negative powers of $z$) we can assume that
$f\in\mathcal{M}'=\{f\in \mathcal{M}:f(z)=o(z)\ \text{as}\
|z|\to\infty\}$. }\\
{\rm This is one of the simplest settings often occurring in applications. We can see later from the proof that the approach is more general.}
\end{Definition}
\begin{Definition}
\label{def2}
{\rm Assume $g(s)$ is analytic in $U_\delta \backslash [0,\infty) $ for some $\delta > 0$, where $U_\delta = \{z : |\mathrm{Im} (z)| \leq \delta, \mathrm{Re}(z) \geq -\delta \}$ and $g(s) \to 0$ uniformly in $U_\delta$, as $\mathrm{Re}(s) \to \infty$.
Assume $\epsilon \leq \delta$. We define $\Gamma_\epsilon$ to be the contour around $\mathbb{R}^+$ consisting of two rays $l_{1,\epsilon}, l_{2,\epsilon}$ and a semicircle $\gamma_{\epsilon}$, where $l_{1,\epsilon} = \{x - \epsilon i: a\in[0,\infty)\}$ oriented towards the left, $l_{2,\epsilon} = \{x + \epsilon i: a\in[0,\infty)\}$ oriented towards the right; $\gamma_{\epsilon}$ is the left semicircle centered at origin oriented clockwise.
Assume also that $g(s)$ is absolutely integrable over $\Gamma_\epsilon$ for some $\epsilon$.
We denote by
\begin{equation}
\label{eq:sg}
\oint_0^{\infty}g(s)\mathrm{d}s
\end{equation}
the integral of $g$ over $\Gamma_\epsilon$. Since $g(s)$ vanishes at $\infty$, the integral is independent of the choice of $\epsilon$ as long as it is small enough.}
\end{Definition}
The following observations will simplify our proofs.
\begin{Note}\label{residue}
{\rm Let $g(s)$, $U_\delta$ and $\Gamma_\epsilon$ be as in Definition \ref{def2}. $\Gamma_\epsilon$ separates $\mathbb{C} \backslash \Gamma_\epsilon$ into two regions. We denote the region containing $\mathbb{R}^+$ by $S_1$ and the other by $S_2$. Let}
\begin{align}
G_1(z) = \int_{\Gamma_\epsilon} \frac{g(s)\mathrm{d}s}{s-z} \quad\quad (z\in S_1)\\
G_2(z) = \int_{\Gamma_\epsilon} \frac{g(s)\mathrm{d}s}{s-z} \quad\quad (z\in S_2)
\end{align}
{\rm Then $G_1$ is analytic in $S_1$ and $G_2$ is analytic in $S_2$. By slightly deforming $\Gamma_\epsilon$ we are able to see that each $G_i$ can be analytically continued to $S_i \cup \Gamma_\epsilon$, $i = 1,2$. On $\Gamma_\epsilon$ their analytic continuations satisfy}
\begin{equation}
G_2(z)- G_1(z) = 2\pi i g(z) \quad\quad (z \in \Gamma _\epsilon)
\end{equation}
{\rm Hence $G_2(z)$ can be analytically continued to $\mathbb{C} \backslash [0,\infty)$ and $G_1(z)$ can be analytically continued to at least $U_\delta$ and in regions where $g$ is analytic. For each $z \in \mathbb{C} \backslash [0,\infty)$}
\begin{equation}
G_2(z) = \oint_0^{\infty} \frac{g(s) \rm{d}s}{s-z}
\end{equation}
\end{Note}
\begin{Note}
{\rm A representation of the form \eqref{eq:sg} exists for Hilbert-transform-like integrals such as $h(t)= \int_0^{\infty}(s-t)^{-1}H(s)ds$ with $H$ analytic at
zero, for instance $h(t)=- (2\pi i)^{-1}
\oint_0^{\infty}(s-t)^{-1}H(s)\ln s\,ds$.}
\end{Note}
\begin{Note}\label{COV}
{\rm Consider the composition of $g$ with $s \mapsto \ln (1+s)$, the branch cut of which is chosen to be $(-\infty, -1]$. If $g$ is analytic in $U_\delta \setminus [0,\infty)$, then $g(\ln(1+s))$ is analytic in the set $-1+\exp(U_\delta \setminus [0,\infty))$. If in addition we have the decay condition $g(\ln(1+s)) = o(|s|^{-\alpha})$ for some $\alpha > 0$ as $|s| \to \infty$, then there exists a $\tilde{\delta}$ small enough such that $U_{\tilde{\delta}} \subseteq -1+\exp(U_\delta \setminus [0,\infty))$. It is easy to see from the decay condition that
\begin{equation*}
\int_{\Gamma_{\delta}} g(p) \mathrm{d}p = \int_{\exp(\Gamma_\delta) -1} \frac{g(\ln(1+s))}{1+s} \mathrm{d}s = \int_{\Gamma_{\tilde{\delta}}} \frac{g(\ln(1+s))}{1+s} \mathrm{d}s
\end{equation*}
\rm{and thus we can make the change of variable}
\begin{equation}
\oint_{0}^{\infty} g(p) \mathrm{d}p = \oint_{0}^{\infty} \frac{g(\ln(1+s))}{1+s} \mathrm{d}s
\end{equation}
\end{Note}
While providing integral formulae in terms of functions with known
singularities which are often rather explicit, the following result can also
be interpreted as a {\em duality of resurgence}. \footnote{After developing these methods,
it has been brought to our attention that a duality between resurgent
functions and resurgent Taylor coefficients has been noted in an unpublished
manuscript by \'Ecalle. This will be further explored in a forthcoming paper.}.
\begin{Theorem}\label{T1}
(i) Assume that $f(z)=\sum_{k=0}^{\infty}c_k z^k$ is a series with positive, finite radius of convergence, with $c_k$ having Borel sum-like representations of the form
\begin{equation}
\label{eq:trs1}
c_k=\sum_{j=1}^N a_j^{-k}\oint_0^{\infty}e^{-kp}F_j(p)\mathrm{d}p \hspace{0.4in} (k \ge 1)
\end{equation}
(\ref{eq:trs1}) with $a_j$ as in Definition 2.1, $F_j$ analytic in $U_\delta \backslash [0,\infty) $ for some $\delta > 0$ and algebraically bounded at $\infty$. Then, $f$ is given by
\begin{equation}
\label{eq:recon3}
f(z)=f(0)+z\oint_0^{\infty}\sum_{j=1}^N \frac{F_j(\ln (1+s))\mathrm{d}s}{(1+s)((1+s) a_j - z)}
\end{equation}
\z (ii) Furthermore, $f \in \mathcal{M}'$.The behavior of $f$ at $a_j$ and is of the same type as the behavior of $F_j(\ln(1+s))$ at $0$. More precisely, for small $z \notin [0,\infty)$,
\begin{equation}
\label{eq:singtype}
f(a_j (z+1)) = 2\pi i F_j (\ln (1+z)) + G(z)
\end{equation}
where $G(z)$ is analytic at 0.
\z (iii) Conversely, assume $f \in \mathcal{M}'$, and has finitely many singularities located at $\{a_j t_{j,l} \}$, $(1 \leq j \leq N, 1\leq l)$, with $1 =t _{j,1}$ and $ t _{j,l} < t_{j,l+1}$ for all $j, l$. Let
$c_k=f^{(k)}(0)/k!$; then $c_k$ have Borel sum-like representations of the form
\begin{equation}
\label{eq:trs12}
c_k=\frac{1}{2\pi i}\sum_{j=1}^N (a_j )^{-k}\oint_0^{\infty}e^{-ks}\ f( a_{j} e^s) \, \rm{d}s,\quad k \geq 1
\end{equation}
\end{Theorem}
The behavior at $a_j$ and
at $\infty$ will follow from the proof.
As it will be clear from the proofs, the method and results would apply, with
minor adaptations to functions of several complex variables.
\subsection{Entire functions}\label{entire}
We
restrict the analysis to entire functions of exponential order one, with
complete information on the Taylor coefficients. Such functions include of
course the exponential itself, or expressions such as $f_3$. It is useful to
start with $f_3$ as an example. The analysis is brought to the case in
Theorem \ref{T1} by first taking a Laplace transform. Note that
\begin{equation}
\label{eq:eq41}
\int_0^{\infty}e^{-xz}f(z)\mathrm{d}z=\frac{1}{x}\sum_{n=1}^\infty\frac{n!}{n^{n+1}x^{n}}
\end{equation}
The study of entire functions of exponential order one likely involves the
factorial, and then a Borel summed representation of the Stirling formula is
needed; this is provided in the Appendix.
\begin{Theorem}\label{T2}
Assume that the entire function $f$ is given by
\begin{equation}
\label{eq:enti22}
f(z)= \sum_{k=1}^\infty \frac{c_k z^k }{k!}
\end{equation}
with $c_k$ as in Theorem~\ref{T1} (i). Then,
\begin{equation}
\label{eq:recon31}
f(z)=
\oint_0^{\infty}\sum_{j=1}^N \left[\left(e^{\frac{z}{a_j(1+s)}}-1\right)
\frac{F_j(\ln (1+s))}{(1+s)}\right]\mathrm{d}s
\end{equation}
\end{Theorem}
As in the simple example, the behavior at infinity follows from the integral
representation by classical means.
\subsection{Borel summation}
We obtain from Theorem~\ref{T1}, in the same way as above, the
following.
\begin{Theorem}\label{T3}
Consider the formal power series
\begin{equation}
\label{eq:enti23}
\tilde{f}(z)= \sum_{k=1}^\infty {c_k k! x^{-k-1}}
\end{equation}
with coefficients $c_k$ as in Theorem~\ref{T1} (i).
Then the series \eqref{eq:enti23} is (generalized) Borel summable to
\begin{multline}
\label{eq:bs1}
\int_0^{\infty}dp e^{-px}p\sum_{j=1}^N\oint_0^{\infty}\frac{F_j(\ln(1+s))}{(1+s)(a_js+a_j-p)}\mathrm{d}s\\=
\sum_{j=1}^N\oint_0^{\infty}\frac{F_j(\ln (1+s))}{1+s}\left(-\frac{1}{x}+a_j(s+1)e^{-a_j(s+1)x}\,\mathrm{Ei}\left(a_j(s+1){x}\right)\right)\mathrm{d}s
\end{multline}
\end{Theorem}
The proof proceeds as in the previous sections, taking now a Borel
transform followed by Laplace transform.
\subsection{Other applications; the examples in the introduction}\label{Examples}
\subsubsection{Other growth rates}\label{Growthrates}
Series with coefficients with growth rates precluding a straightforward inverse Laplace transform can be accommodated, for instance by analytic continuation.
We have for positive $\gamma$,
\begin{equation}
\label{eq:eqinvl}
e^{-\gamma\sqrt{
n}}=\frac{\gamma}{2\sqrt{\pi}}\int_0^{\infty}p^{-3/2}e^{-\frac{\gamma^2}{4p}}e^{-np}\mathrm{d}p
\end{equation}
which can be analytically continued in $\gamma$. We note first that the
contour cannot be, for this function, detached from zero. Instead, we keep $[1,\infty)$ as part of the original contour fixed and, deform the part $[0,1]$ by simultaneously rotating $\gamma$ and $p$ to maintain $\gamma^2/p$ real
and positive near the
origin. We get
\begin{equation}
\label{eq:eqerfc}
e^{\sqrt{n}}=-\frac{1}{2\sqrt{\pi}}\int_{C_1} p^{-3/2} e^{-\frac{1}{4p}}e^{-np}\mathrm{d}p
\end{equation}
and \eqref{ff4} follows, for the same reason Theorem \ref{T1} (i) holds. In particular,
\begin{equation}
\label{eq:limit1}
\lim_{z\to -1^+}\sum_{n=1}^{\infty}e^{\sqrt{n}}z^n=
\frac{1}{2\sqrt{\pi}}\int_{C_1}\frac{p^{-3/2}e^{-\frac{1}{4p}}}{e^p+1}\mathrm{d}p
\end{equation}
The sum \eqref{eq:limit1} is unwieldy numerically,
while the integral \eqref{eq:eqinvl} can be evaluated accurately by standard
means. In a similar way we get
\begin{equation}
\label{eq:eqsum}
\sum_{k=0}^{\infty}\frac{e^{i\sqrt{k}}}{k^{a}}=-\frac{\gamma
2^{a-1/2}}{\sqrt{\pi}}
\int_C \frac{e^{-\frac{1}{8p}}U(2a+1/2;\frac{1}{\sqrt{2p}})}{p^{a-1}(e^p+1)}\mathrm{d}p
\end{equation}
for $a>1/2$ for which the series converges.
Here $C$ is a contour consisting of $S_2$ followed by $[1,\infty)$, where $S_2$ starts at $0$ and ends at $1$, and is given by $r = 1 - \theta/\pi$, $\left(\theta \in [0,\pi]\right)$ in polar coordinates. $U$ is the parabolic cylinder function
\cite{Abramowitz}.
{\em The coefficients $c_k^{[1]}$ in \eqref{eq:ln2}}.
We have
\begin{equation}
\label{eq:invlp3}
\mathcal{L}^{-1}\left[\frac{1}{(n+a)^b}\right]=\Gamma(b)^{-1}p^{b-1}e^{-ap}
\end{equation}
The rest follows in the same way \eqref{ff4} was obtained, after changing
variables to $1+t=e^p$.
{\em The coefficients $c_k^{[2]}$}.
We let $x=n$ and take the inverse
Laplace transform in $x$:
\begin{equation}
\label{eq:dual1}
G(p) = \frac{1}{2\pi i}\int_{1-i\infty}^{1+i\infty}\frac{e^{xp}}{x^{b}+\ln x}\mathrm{d}x
\end{equation}
where the contour can be bent backwards for $p \in \mathbb{R}^+$, to hang around $\mathbb{R}^-$. Then, with
the change of variable $x=-u$ (\ref{eq:dual1}) becomes
\begin{equation}
\label{eq:dual2}
G(p)= \frac{1}{2 \pi i}\oint_{0}^{\infty}\frac{e^{-up}}{(-u)^{b}+\ln (-u)}\mathrm{d}u
\end{equation}
and thus
\begin{equation*}
c_k = (\mathcal{L} G) (k) = \int_0^{\infty} G (p) e^{-kp}{\mathrm{d}p } = \oint_0^{\infty} \left[ \frac{-G(p) \ln p}{ 2 \pi i} \right] e^{-kp}{\mathrm{d}p}
\end{equation*}
We see that $F_1(p) = (- G(p) \ln p)/{2 \pi i}$ and by Theorem 2.1
\begin{equation*}
f_2 (z) = z \oint_0^{\infty} \frac {\tilde{G}(s)}{s - (z-1)} {\mathrm{d}s}
\end{equation*}
where $\tilde {G}(s) = F_1 (\ln(1+s))/(1+s)$.
Hence the singularity of $f_2(z)$, at $z=1$ on the first Riemann sheet, according to Note \ref{residue} is
that of $ \phi(z) = 2 \pi i \tilde{G}(z - 1)$, as in
(\ref{intlog}).
The example of {\em the coefficients $c_k^{[3]}$} is studied in a similar way as Theorem
\ref{T2}; related calculations can be found in \S\ref{S115}.
{\em The coefficients $c_k^{[4]}$} were treated at the beginning of this
section.
{\em Another example} is provided by the log of the Gamma function,
$\ln\Gamma(n)=\sum_{k=1}^n \ln k$. It is convenient to first subtract out the leading
behavior of the sum to arrange that the summand is inverse Laplace
transformable. With
$$g_n=\ln\Gamma(n+1)-\left((n+1)\ln (n+1)-n-\frac{1}{2}\ln (n+1)\right)$$ we get
\begin{multline}
g_N=\sum_{1}^N
\Big[1-\Big(\frac{1}{2}+n\Big)\ln\Big(1+\frac{1}{n}\Big)\Big]=\sum_{1}^N\int_0^\infty
e^{-np} \frac{ 1-\frac{p}{2}-(\frac{p}{2}+1)e^{-p}}{p^2}\mathrm{d}p
\end{multline}
where $\mathcal{L}^{-1}$ of the summand in the middle term is most easily
obtained by noting that its second derivative is a rational function.
Summing as usual $e^{-np}$ we get
\begin{equation}
\label{eq:lngamma}
\ln\Gamma(n)=n(\ln n -1)-\frac{1}{2}\ln n +\frac{1}{2}\ln(2\pi)+\int_0^{\infty}\frac{\displaystyle
1-\frac{p}{2}-\Big(\frac{p}{2}+1\Big)e^{-p}}{p^2(e^{-p}-1)}e^{-np}\mathrm{d}p
\end{equation}
Obviously, if the behavior of the coefficients is of the form $A^kc_k$ where
$c_k$ satisfies
the conditions in the paper, one simply changes the independent variable to $z'=Az$.
\section{Proof of Theorem \ref{T1}}
\z If $f\in\mathcal{M}'$ we write the Taylor coefficients in the form
\begin{equation}
\label{eq:cf1}
c_k=\frac{1}{2\pi i}\oint \frac{ f(p)\mathrm{d}p}{p^{k+1}} \quad\quad(k\geq 1)
\end{equation}
where the contour of integration is a small circle of radius $r$ around the
origin. We attempt to increase $r$ without bound. In the process, the contour
will hang around the singularities of $f$ as shown in Figure 1. Each integral over a curve that wraps around a ray $\{a_j t:t\geq 1\}$ converges by the decay assumptions and the contribution of the arcs
at large $r$ vanishes, since $f(z) = o(z)$ as $z \to \infty$.
\begin{figure}
\includegraphics[width=12cm]{cuts.pdf}
\caption{Singularities of $f$, cuts, direction of integration and Cauchy
contour deformation. }
\end{figure}
To be more precise, let $C_{j,\epsilon}$ be the part of the image of $\Gamma_{\epsilon}$ under the mapping $s \to a_j e^s $, let $C_{j,\epsilon, R}$ be the part of $C_{j,\epsilon}$ inside the disk $|s| \leq R$, and $C_R$ be the part of the contour on $|s| = R$. Then for $\epsilon$ small enough and $R$ large enough we have
\begin{equation*}
c_k=\frac{1}{2\pi i}\oint \frac{ f(p)\mathrm{d}p}{p^{k+1}} = \frac{1}{2\pi i} \left( \sum_{j = 1}^{N} \int_{C_{j, \epsilon, R}} \frac{ f(p)\mathrm{d}p}{p^{k+1} }+ \int_{C_R} \frac{ f(p)\mathrm{d}p}{p^{k+1} }\right)
\end{equation*}
By the change of variable $p = a_j e^s$ and letting $R \to \infty$ we get
\begin{equation*}
\int_{a_j C_{\epsilon, R}} \frac{ f(p)\mathrm{d}p}{p^{k+1} } = \int_{\Gamma_{\epsilon}} \frac{a_j e^s f(a_j e^s) \mathrm{d}s}{(a_j e^s)^{k+1} } = \oint_{0}^{\infty} {a_j}^{-k} e^{-ks} f(a_j e^s) \mathrm{d}s
\end{equation*}
and the integral over $C_R$ vanishes as $R \to \infty$ by decay condition for $k \geq 1$.
\bigskip
\z In the opposite direction, first let $\epsilon$ be small enough so that for all $j$ and $k = 1$
\begin{equation}
\label{eq:path}
\oint_{0}^{\infty}{e^{-kp}F_j(p) \mathrm{d}p} = \int_{\Gamma_{\epsilon}}{e^{-kp}F_j(p) \mathrm{d}p}
\end{equation}
Then for all $1\leq j \leq N$, $k \geq 1$ (\ref{eq:path}) is true. Also let $z$ be small so that
\begin{equation}
\label{eq:est1}
|{a_j}^{-1}e^{-p} z| \leq \delta^2 < 1
\end{equation}
for all $j$ and $p \in \Gamma_\epsilon$.
Then, by the dominated convergence theorem (which applies in this case, see (\ref{eq:estimate})) we have
\begin{align}
\label{eq:converse}
f(z) - f(0)&
= \sum_{k=1}^{\infty}c_k z^k
=\sum_{k=1}^{\infty}\left({\sum_{j=1}^N a_j^{-k}\int_{\Gamma_\epsilon}e^{-kp}F_j(p)\mathrm{d}p} \right) z^k\\ \notag
&= \int_{\Gamma_\epsilon} \sum_{j=1}^{N} \left( \sum_{k=1}^{\infty} (a_j^{-1}e^{-p} z)^k \right) F_j(p) \mathrm{d}p = \int_{\Gamma_\epsilon} \sum_{j=1}^{N} \left( \frac{a_j^{-1}e^{-p} z}{1 - a_j^{-1}e^{-p} z} \right) F_j(p) \mathrm{d}p\\ \notag
&=\sum_{j=1}^{N} \oint_0^{\infty} \frac{a_j^{-1}e^{-p} z}{1 - a_j^{-1}e^{-p} z} F_j(p) \mathrm{d}p =z\sum_{j=1}^N
\oint_0^{\infty}\frac{F_j(\ln (1+s))\mathrm{d}s}{(1+s)(s a_j+a_j-z)}
\end{align}
as stated. The third equality holds because we have, in view of \eqref{eq:est1},
\begin{align}
\label{eq:estimate}
\left| \sum_{j=1}^N (a_j^{-1}e^{-p} z)^kF_j(p)\right| \leq \sum_{j=1}^N |a_j^{-1} e^{-p} z|^{k/2} \left( |a_j^{-1} e^{-p} z|^{k/2} |F_j(p)|\right) \nonumber\\
\leq \sum_{j=1}^N \delta^k \left( |a_j^{-1} e^{-p} z|^{k/2} |F_j(p)|\right)
\end{align}
For each $j$, $|a_j^{-1} e^{-p} z|^{k/2} F_j(p)$ is integrable over $\Gamma_\epsilon$ since $F_j$ is algebraically bounded at $\infty$, so we may interchange the order of integration and summation over $k$. The last equality holds because $F_j(\ln (1+s))/(s a_j+a_j-z) = o(|s|^{-\alpha})$ for some $\alpha > 1$ so we can make the change of variable $p=\ln (1+s)$, see Note \ref{COV}. Hence \eqref{eq:trs1} holds for $z$ small.\\
\indent Given $j\in \{1, ..., N\}$ we may write
\begin{equation}\label{intj}
I_j(z) := \oint_0^{\infty}\frac{F_j\left(\ln (1+s)\right) \rm{d}s}{(1+s)(s a_j+a_j-z)} = \oint_0^{\infty}\frac{F_j\left(\ln (1+s)\right)/\left(a_j (1+s)\right)\, \rm{d}s}{\left(s-\left(z/a_j - 1\right)\right)}
\end{equation}
Since $F_j$ is analytic in $\mathbb{C} \backslash [0,\infty)$ and is algebraically bounded at $\infty$, $F_j(\ln (1+s))/\left(a_j (1+s)\right)$ is analytic in $\mathbb{C} \backslash [0,\infty)$ and vanishes uniformly as $\mathrm{Re}(s) \to \infty$. Then it becomes obvious from Note \ref{residue} that the integral $I_j$ in \eqref{intj} is analytic in $\mathbb{C} \backslash \{a_j t: t \geq 1\}$. Thus $f(z)$ can be analytically continued to $\mathbb{C} \backslash \bigcup_{j=1}^{N} \{a_j t: t \geq 1\}$.\\
To see that $f(z) = o(z)$ as $|z| \to \infty$, it suffices to show that for each $j$, $I_j(z) = o(1)$. Assume $3\epsilon < \delta$. We use the contours $\exp(\Gamma_\epsilon)-1$ and $\exp(\Gamma_{3{\epsilon}})-1$. If $|\arg(z/a_j)| \geq 2 {\epsilon}$ and $|z/a_j|$ is large enough we write:
\begin{equation}
I_j(z) = \int_{\exp(\Gamma_\epsilon) -1}\frac{F_j\left(\ln (1+s)\right)/\left(a_j (1+s)\right)\, \rm{d}s}{\left(s-\left(z/a_j - 1\right)\right)}
\end{equation}
Then there exist some positive number $\rho_1$ such that $|s-\left(z/a_j - 1\right)| \geq \rho_1 |s|$ for each $s\in \exp(\Gamma_\epsilon) -1$, so we can use dominated convergence to obtain $I_j(z) \to 0$ as $|z| \to \infty$. If $|\arg(z/a_j)| \leq 2 {\epsilon}$ we use Note \ref{residue} to write:
\begin{equation}\label{insideIj}
I_j(z) = \int_{\Gamma_{3{\epsilon}}}\frac{F_j\left(\ln (1+s)\right)/\left(a_j (1+s)\right)\, \rm{d}s}{\left(s-\left(z/a_j - 1\right)\right)} + 2\pi i F_j(\ln(z/a_j))/z
\end{equation}
Then there exist some positive number $\rho_2$ such that $|s-\left(z/a_j - 1\right)| \geq \rho_2 |s|$ for each $s\in \exp(\Gamma_3\epsilon) -1$, so we can use dominated convergence to prove the integral in the right hand side of \eqref{insideIj} is $o(1)$. By assumption it is obvious that $F_j(\ln(z/a_j))/z = o(1)$, so in this case we also have $I_j(z) \to 0$ as $|z| \to \infty$.\\
The nature of the singularities of $f$ is derived from Note \ref{residue}. Let $z\notin [0,\infty)$ be small, then for each $j \in \{1,...,N\}$
\begin{align*}
&f((z+1)a_j)\\ &=f(0) + ((z+1)a_j)\sum_{l \neq j} I_l((z+1)a_j)+ ((z+1)a_j)I_j((z+1)a_j)\\
&=f(0) + ((z+1)a_j)\sum_{l \neq j} I_l((z+1)a_j)\\
&\quad\quad+((z+1)a_j)\oint_0^{\infty}\frac{F_j\left(\ln (1+s)\right)/\left(a_j (1+s)\right)\, \rm{d}s}{s-z}\\
&=f(0) + ((z+1)a_j)\sum_{l \neq j} I_l((z+1)a_j) + A(t) + 2 \pi i F_j(\ln(1+z))\end{align*}
where $A(t)$ is analytic at $z=0$. The last equality is obtained by Note \ref{residue}. It is obvious from Note \ref{residue} that each $I_l((z+1)a_j)$ $(l\neq j)$ is analytic on $[0, \infty)$. Thus
\begin{equation}
f((z+1)a_j) = \tilde{G}(z) + 2 \pi i F_j(\ln(1+z))
\end{equation}
where $\tilde{G}(z)$ is analytic at $z=0$. Hence \eqref{eq:singtype} follows.
\section{Appendix}
\label{EB}
\subsection{Simple integral representation of $1/n!$ \cite{Abramowitz}}
We have
\begin{equation}
\label{eq:eqg}
\frac{1}{\Gamma(z)}=-\frac{ie^{-\pi i
z}}{2\pi}\oint_0^{\infty}s^{-z}e^{-s}ds=-\frac{ie^{-\pi i
z}z^{-z}}{2\pi}\oint_0^{\infty}s^{-z}e^{-zs}ds
\end{equation}
with our convention of contour integration. From here, one can proceed
as in \S\ref{S115}.
\subsection{The Gamma function and Borel summed Stirling formula}\label{S115}
We have
\begin{multline}
n!=\int_0^\infty t^ne^{-t}\mathrm{d}t=n^{n+1}\int_0^\infty e^{-n(s-\ln s)}\mathrm{d}s\\=
n^{n+1}\int_0^1 e^{-n(s-\ln s)}\mathrm{d}s+ n^{n+1}\int_1^\infty e^{-n(s-\ln s)}\mathrm{d}s
\end{multline}
\z On $(0,1)$ and $(1,\infty)$ separately, the function $s-\ln(s)$ is
monotonic and we may write, after inverting $s-\ln(s)=t$ on the two
intervals to get $s_{1,2}=s_{1,2}(t)$ \footnote{The functions $s_{1,2}$ are
given by branches of $-W(-e^t)$, where $W$ is the Lambert function.},
\begin{equation}\label{e25}
n!= n^{n+1}\int_1^{\infty}e^{-nt}(s'_2(t)-s'_1(t))\mathrm{d}t=n^{n+1}e^{-n}
\int_0^{\infty}e^{-np}G(p)\mathrm{d}p
\end{equation}
\z where $G(p)=s_2'(1+p)-s_1'(1+p)$. From the definition it follows that $G$
is bounded at infinity and $p^{1/2}G$ is analytic in $p$ at $p=0$.
Using now (\ref{e25}) and Theorem~\ref{T1} in (\ref{eq:eq41}) we get
\begin{equation}
\label{eq:lapl41}
\int_0^{\infty}e^{-xz}f_3(z)\mathrm{d}z=\frac{1}{ x^2}\int_0^{\infty}\frac{G(\ln(1+t))}{(te+(e-x^{-1}))(t+1)}\mathrm{d}t
\end{equation}
Upon taking the inverse Laplace transform we obtain (\ref{eq:invl}).
\subsection{Solution of \eqref{eq:thfilm}}\label{4thoode}
Assume the solution to \eqref{eq:thfilm} which is analytic at $\eta=0$ has Taylor expansion
$f(\eta) = \sum_{k=0}^{\infty} c_k \eta^k$. Then
\begin{equation}
c_k = \frac{c_1 (-1)^{k/2-1/2}}{A\Gamma(-1/2-a)} \frac{\Gamma(k/2-1-a) k}{\left(\Gamma(k+1)\right)^2} \quad\quad (k\,\,\rm{is}\,\,odd)
\end{equation}
\begin{equation}
c_k = 0 \quad\quad (k\,\,\rm{is}\,\,even)
\end{equation}
It is obvious that $f(\eta)$ is entire. Consider the Laplace transform $F(p) = \mathcal{L}f(p)$. Then $F(p) = \sum_{k=0}^{\infty} \Gamma(k+1) c_k p^{-k-1}$. Let $G(p) = F(1/p)$; then $G(z)$ is a solution to the differential equation
\begin{equation}
G''(z)+\left(-\frac{4}{z}+\frac{z}{2A}\right) G'(z) + \left(\frac{6}{z^2}-\frac{3/2+a}{A}\right)G(z) = 0
\end{equation}
and $G(z)$ is analytic at $z=0$.
The normalization transformation $G(z)=z^{3/2}e^{-z^2/A}h(z^2/A)$ (cf. \cite{Duke} for a general description of normalization methods) yields
\begin{equation}
\label{eq:eqW}
h''-\frac{7}{4}h'+\frac{1}{16}\left(12-\frac{3+4a}{s}+\frac{3}{s^2}\right)h=0
\end{equation}
solvable in terms of Whittaker functions \cite{Abramowitz}. Substituting back, by straightforward algebra, this yields \eqref{whit}.
\subsection{Acknowledgments} M. Kruskal introduced OC to questions of the
type addressed in the paper; OC is grateful to R. D. Costin, J. McNeal, D. Sauzin, and S. Tanveer for very useful discussions. Work supported in part
by NSF grants DMS-0600369 and DMS-1108794.
|
\section{Introduction}
Infinitesimal Perturbation Analysis (IPA) is an established sample-path technique for sensitivity estimation of performance functions
defined on discrete event dynamic systems and stochastic hybrid systems. In a typical scenario $L(\theta)$ is a real-valued, random function of a parameter $\theta\in R^n$, defined over a common probability space
$(\Omega,{\cal F},P)$, and for a particular realization corresponding to $\omega\in\Omega$, IPA computes its sample
($\omega$-dependent) derivative (gradient) $\nabla L(\theta)$. This sample gradient can act, under certain
circumstances, as an estimator of the gradient of the expected-value function $J(\theta):=E\big[L(\theta)\big]$, with
$E[\cdot]$ denoting expectation in $(\Omega,{\cal F},P)$. This can be used, in conjunction with gradient-descent
algorithms such as stochastic approximation, to minimize the function
$J(\theta)$ to the extent of computing a local minimum. For extensive presentation of IPA and its scope
in optimization, , please see
\cite{Ho91,Glasserman91,Cassandras99}.
In recent years there has been a mounting interest in the application of IPA to stochastic hybrid systems, and
especially to stochastic flow networks, comprising generalizations of fluid queues \cite{Cassandras02,Cassandras06}. The reason is that, for an extensive
class of performance functions in this setting, IPA was shown to be computable via simple algorithms directly from quantities that
are observable from realizations of the state of the system. Moreover, the resulting gradient estimators are quite
robust to modeling variations. For these reasons, for a variety of
applications, the IPA gradients arguably can be computed in real time without having concerns about the accuracy of the
underlying models. References \cite{Cassandras10,Wardi10} provide detailed
discussions of these points as well as unified frameworks
for IPA in the setting of stochastic hybrid systems.
Until recently the main areas mentioned for potential applications of IPA were in manufacturing and telecommunications,
but lately
there has been a growing interest in transportation networks as well, and especially in traffic-light control (see \cite{Fleck14} for
a survey).
The main objective
of traffic-light control is to reduce, or minimize congestion at
traffic intersections. Early techniques developed for these control and optimization problems include
dynamic programming \cite{Porche96} and linear complementary algorithms \cite{DeSchutter99},
while
more recent approaches are based on Markov-decision processes \cite{Yu06}, game theory \cite{Alvarez10},
and
mixed-integer programming \cite{Dujardin11};
\cite{Papageorgiou03} contains an early survey, and a recent one can be found in \cite{Fleck14}. Regarding applications of IPA,
early results were presented in
\cite{Fu03,Panayiotou05}, and a recent systematic approach has been
developed in
\cite{Geng12a,Geng12b,Geng13,Fleck14}. This approach defines the traffic-light control problem in the aforementioned
setting of stochastic hybrid systems \cite{Cassandras10}, and develops for it effective IPA-based algorithms.
The development of IPA since its inception has been motivated primarily by applications to
performance optimization in discrete event and hybrid systems. This paper follows a different track in pursuing
an application to performance regulation. The term ``regulation'' here means real-time tracking of
a set (reference) performance index by tuning a control parameter, and it is a common engineering practice. In
particular, following a system's design or optimization with an imprecise system model, regulation
can be used to ensure that performance meets specifications under changing system characteristics and operating
environments.
In devising our regulation technique we aim at effective and efficient real-time implementation. Effectiveness means that the
set-point tracking algorithm is to have fast convergence under a wide set of system parameters, efficiency means simple implementation
requiring low computing efforts, and the real-time requirement means that all input parameters to the controller be measurable
by observing the system's state. High degree of efficiency means that we may have to tilt the balance between speed and
precision
of computation in favor of the former requirement. Consequently we design the controller for maximum speed
and simplicity of computations, possibly by using imprecise models,
while guaranteeing its robustness under large variations in the system's parameters. {\it The main
contribution of this paper is in a control system with all of these properties. It is based on an integral control
with a variable gain, computed by using the IPA derivative of the plant function with respect to the control variable.}
The integral controller is explained in Section II in an abstract setting. Its application to a relevant example
of traffic-light control, including the derivation of the IPA derivative, is presented in Section III. Section IV
contains simulation results, and Section V concludes the paper.
\section{Regulation Algorithm: Integral Control with Adaptive Gain}
Consider the single-input-single-output discrete-time
control system shown in Figure~1, where $r$ is the reference (set point) input, $y_{n}$ is the output, $e_{n}$ is the error signal, and $u_{n}$ is the input to the plant. Suppose that the plant is a time-varying, memoryless nonlinearity
of the form
\begin{equation}
y_{n}=g_{n}(u_{n}),
\end{equation}
where $g_{n}:R\rightarrow R$, $n=1,2,\ldots$, is called the {\it plant function}.
Given a reference input $r$, the purpose of the control system is to ensure that
$\lim_{n\rightarrow\infty}y_{n}=r$. To this end it is natural to choose the controller to be an integrator having,
for example, the transfer function $G_{c}(z)=Az^{-1}/(1-z^{-1})$ for a given gain $A>0$.
However, integral controllers may display oscillatory behavior and have narrow stability margins. Furthermore,
due to the time-varying nature of the system it may not be easy (or possible) to choose a gain that fits all
possible scenarios.
For this reason we use an integral control with adaptive gain, $A_{n}$, having the time-domain representation
\begin{equation}
u_{n}\ =\ u_{n-1}+A_{n}e_{n-1},
\end{equation}
where the error signal is defined as
\begin{equation}
e_{n}=r-y_{n}.
\end{equation}
We choose the gain $A_{n}$ to be defined via the equation
\begin{equation}
A_{n}\ =\ \frac{1}{g_{n}^{\prime}(u_{n-1})},
\end{equation}
with ``prime'' denoting derivative. Equations (1) -- (4), computed cyclically in the order
$(4)\rightarrow(2)\rightarrow(1)\rightarrow(3)$, define the dynamics of the closed-loop system.
In fact, we implicitly consider the following scenario in a real-time setting: the quantities $u_{n-1},\ y_{n-1}$, and
$e_{n-1}$ have been computed or derived by the start of the $n$th computing cycle, and at that time the following sequence of
operations takes place and are completed before the end of the cycle: (i) $A_{n}$ is computed by Equation (4);
(ii) $u_{n}$ is computed by the controller via Equation (2); and (iii) the system yields $y_{n}$ and $e_{n}$ (see (1), (3)).
The main computation is that of $A_{n}$ in Equation (4), since an exact evaluation of
$g_{n}^{\prime}(u_{n-1})$ may not be possible in real time. In this case it may be necessary to trade off precision with
computational expediency, with the result that Equation (4) be replaced by
\begin{equation}
A_{n}\ =\ \frac{1}{g_{n}^{\prime}(u_{n-1})+\xi_{n}},
\end{equation}
where $\xi_{n}$ is a quantity representing the computing error. We will say more about this point in the sequel.
\begin{figure}[!t]
\centering
\includegraphics[width=3in]{Figure_1}
\caption{Control System}
\label{Figure 1}
\end{figure}
The rationale behind Equation (4) can be seen by considering the time invariant plant, where $g$ is not a function
of $n$ and hence Equation (1) has the form $y_{n}=g(u_{n})$. In this case, it can be seen that
the control loop, defined via Equations (1) --(4), implements Newton's method for solving the equation
$g(u)=r$, known to converge, in the sense that
\begin{equation}
\lim_{n\rightarrow\infty}e_{n}\ =\ 0,
\end{equation}
under broad assumptions. In the event that Equation (5) is used instead of (4),
namely (still assuming a time-invariant plant) $A_{n}=1/(g^{\prime}(u_{n-1})+\xi_{n})$,
the regulation scheme still converges in the sense of Equation (6), under broad assumptions,
as long as the relative error is under 100\%. For example, if the function $g$ is concave or convex,
and if, for some $\alpha\in(0,1)$, the relative error of the plant derivative, defined as $\varepsilon_{n}:=|\xi_{n}|/|g_{n}^{\prime}(u_{n+1})|$,
satisfies the inequality
$\varepsilon_{n}<\alpha$ for all $n=1,2,\ldots$, then Equation (6) is satisfied (see \cite{Almoosa12}).
In the case of a time-varying plant defined by
Equation (1),
Equation (6) no longer can be expected. However, the error sequence $\{e_{n}\}$
asymptotically gets close to 0 by an amount that depends, monotonically, on a measure of the
system's variability. For example, \cite{Almoosa12} proved, under the convexity assumption, that for every
$\epsilon>0$ there exists $\delta>0$ such that, if $|g_{n}(u_{n-1})-g_{n-1}(u_{n-1})|<\delta$ for every $n=1,2,\ldots$, then
\begin{equation}
\limsup_{n\rightarrow\infty}|e_{n}|\ <\ \epsilon.
\end{equation}
We point out that the convexity assumption can be relaxed to local convexity or concavity as long as guards are put in
place to ensure an appropriate upper bound on the terms $|u_{n}-u_{n-1}|$. Moreover, bounds on the relative errors
$\varepsilon_{n}:=|\xi_{n}|/|g_{n}^{\prime}(u_{n-1})|$ practically need not be computed a priori but can be verified from system
simulation as done in \cite{Almoosa12a}.
This control law was applied to regulate the dynamic core-power in computer processors by the applied frequency.
The plant, comprising the frequency-to-power relationship, has an established model based on
physical principles \cite{Floyd07}, represented by
a convex, time-varying, memoryless
nonlinearity. Its time variability is due to the activity factor of the program load, a quantity representing the amount
of switching activity of the logic gates at the core. This quantity generally is unpredictable and cannot be measured
in real time, and hence the plant functions cannot be computed. Although the output $y_{n}$ can be measured, the derivative $g_{n}^{\prime}(u_{n-1})$ requires a formula.
It turns out that this derivative is computable from quantities that can be measured in real time and hence
the regulation scheme could be applied; for simulation results with an industry-grade simulator,
please see \cite{Almoosa12}.
Reference \cite{Almoosa12a} considers regulating the instruction throughput in a similar core, also as a function
of frequency. Each control cycle consists of a fixed number of clock cycles and takes about 10 miliseconds, during which
the applied frequency is fixed. The output $y_{n}$ is defined as the average throughput over a given cycle,
and the model for the plant is a queueing network representing instruction-processing at the core.
The queueing model has multiple precedence constraints and is complicated in various other ways,
and hence defies analysis for deriving closed-form formulas for the plant function. However, we estimated the derivative term
$g_{n}^{\prime}(u_{n-1})$ by using IPA, which yielded simple formulas that could be computed in real time. We point
out that these IPA derivatives were statistically biased and hence ``wrong'', but extensive simulations
on various benchmark systems yielded maximum relative error of 30\%. Encouraged by these findings we applied the regulation scheme despite the bias of IPA, and the results can be seen in \cite{Almoosa12a}.
This example highlights two salient points of our proposed regulation framework. First, the plant is modeled as a queueing
network which is a highly dynamic system, but the cost function, consisting of the average throughput over
a certain amount of time, allows us to consider it as a {\it memoryless, time-varying} nonlinearity. The time variability
is due to the uncertain, random element as well as to variations that are inherent in the system's dynamics.
This view serves us better than the dynamic view by dint of Equation (7), and suggests that, under conditions of stochastic stability, longer control cycles
should result in smaller values of $\lim_{n\rightarrow\infty}|e_{n}|$.
Second, the important feature of IPA and its use in Equation (4) is its simplicity and on-line computability, and therefore
the main objective of the analysis below is to derive simple terms for the sample derivatives
$g_{n}^{\prime}(u_{n-1})$.
\section{Traffic-Light Regulation Problem}
This section concerns an application of the aforementioned control technique to a traffic-light intersection model.
In order to highlight the salient features of the regulation scheme, we consider only a simple model and defer
a discussion of more detailed models to a future study. Thus, consider a traffic-light intersection of two unidirectional (one-way) roads. Suppose
that the light in each direction alternates between red and green signals,
and for the sake of simplicity we assume that there is no orange light,
and hence, epochs of red signal in one direction (road) correspond to green signal in the other direction.
Let us focus attention on one of the roads and define a {\it light cycle} as a red period
followed by a green period (for the other road, the same cycle is comprised of green followed by red).
Suppose that the light cycle time is fixed, and denoted by $C$, and let the length of the red period comprise
the control variable, $\theta\in[0,C]$.
One of the ways to characterize congestion is by the traffic buildup in front of a traffic light,
and as in \cite{Geng12a,Geng13,Fleck14}, we model its dynamics by a fluid queue. Such a queue is driven by two random processes: the arrival rate
and the service rate. The arrival-rate process does not depend on $\theta$ and hence it is denoted by
$\{\alpha(t)\}$, while the service-rate process depends on $\theta$ in a manner described below, and hence it
is denoted by $\{\beta(\theta;t)\}$. Given an integer $N>0$ we define the performance function that we seek to regulate
as the time-average of the buffer contents (amount of fluid at the buffer) during $N$ light cycles. Defining $T:=NC$, we call the period $[0,T]$ a {\it control cycle}, and we note that a control cycle consists of $N$ light cycles.
We assume that $\theta$ remains fixed during each
control cycle and it is changed, by the regulation process,
only at the boundary points between consecutive control cycles.
Consider a control cycle at a given $\theta\in[0,C]$. We define the buffer contents during the cycle,
denoted by $x(\theta,t)$, by the one-sided differential equation
\begin{equation}
\frac{dx}{dt^{+}}(\theta;t)=\left\{
\begin{array}{ll}
\alpha(t)-\beta(\theta;t), & {\rm if}\ x(\theta,t)>0\\
0, & {\rm if}\ x(\theta,t)=0,
\end{array}
\right.
\end{equation}
where, for the sake of simplicity, we assume the initial condition $x(\theta,t)=0$
(our simulations use a different initial condition as will be explained in the sequel).
We assume that, $\forall\ \theta\in[0,C]$, w.p.1, $\alpha(t)$ and $\beta(\theta,t)$ are piecewise continuously
differentiable (but not necessarily continuous) in the interval
$t\in[0,T]$, and this ensures that
$x(\theta,t)$ is well defined by Equation (8).
The performance function $L(\theta)$ is defined as
\begin{equation}
L(\theta)=\frac{1}{T}\int_{0}^{T}x(\theta;t)dt.
\end{equation}
Given a reference set-point $r>0$, the objective of the regulation scheme is to compute a sequence
of control variables $\theta_{1},\theta_{2},\ldots$, such that the sequence $L(\theta_{1}),L(\theta_{2}),\ldots$ tracks
$r$ as best as possible.
Regarding the service rate process, we define a model that includes the cases where, upon the light switching from
red to green, the service rate either jumps to, or ramps up towards a given maximum value.
In the latter case the ramp-up period depends on the
queue length at the time the green epoch starts, and lasts until either one of the following two
events occur: (i) The buffer becomes empty, or (ii) the light switches back to red.
In the first event the service rate jumps to the maximum value, and in the second event, it jumps down to 0.
We model the ramp-up rate process as a random function to account for fluctuations that are hard
to model, and assume mutually independent realizations of it in
successive cycles.
Formally, let $\beta_{m}>0$ be a given constant, and let $b(t):[0,C]\rightarrow[0,\beta_{m}]$
be a monotone-nondecreasing random function that does not depend on $\theta$.
On a given light cycle
$[kC,(k+1)C)$, we define the service rate of the queue as follows,
\begin{equation}
\beta(\theta,t)=\left\{
\begin{array}{ll}
0, & {\rm if}\ t\in[kC,kC+\theta)\\
b(t-(kC+\theta)), & {\rm if}\ t\in[kC+\theta,kC+C),\\
& {\rm and}\ x(\theta,t)>0\\
\beta_{m}, & {\rm if}\ t\in[kC+\theta,kC+C),\\
& {\rm and}\ x(\theta,t)=0.
\end{array}
\right.
\end{equation}
Note that this includes the case where $\beta(\theta,t)$ jumps directly from $0$ to $\beta_{m}$ when the light turns green.
In the rest of this section we derive the IPA formula for the derivative term $L^{\prime}(\theta,t)$. This will be
done under the following assumption.
\begin{assumption}
(i). The random functions $\alpha(t)$ and $b(t)$ are independent of each other.
(ii). W.p.1, $\alpha(t)$ is piecewise monotone (nondecreasing/non-increasing) and piecewise continuously differentiable in
$t$.
(iii). W.p.1, the function $b(t)$ is monotone nondecreasing and piecewise continuously differentiable in $t$.
(iv). For every $\theta\in[0,T]$, w.p.1 $\alpha(t)$ is continuous at the points $kC$ and $kC+\theta$, $k=0,1,\ldots$.
(v). For every $\theta\in[0,T]$, w.p.1 the function $\alpha(t)$ is continuous at any point where $\beta(\theta;t)$ is discontinuous.
(vi). For every $\theta\in[0,C)$, w.p.1, for every open interval $I\subset[0,T]$, it is impossible to have the relation
$\alpha(t)=\beta(\theta,t)$ $\forall\ t\in I$ except for the case where $\alpha(t)=\beta(\theta;t)=0$.
\end{assumption}
Similar assumptions are routinely made in the literature on IPA in the setting of stochastic hybrid systems;
e.g., \cite{Cassandras02,Wardi10}. We remark that the attribute ``piecewise continuously differentiable'' means that it is continuously differentiable at all but
a finite set of time-points $t$. At those points it may be discontinuous.
This set of points may depend on the sample $\omega$, and its cardinality need not be upper-bounded over
$\omega\in\Omega$.
In the forthcoming discussion we use the `prime' notation for derivatives with respect to
$\theta$, and the `dot' notation for derivatives with respect to $t$. Thus,
$x^{\prime}(\theta;t):=\frac{\partial x}{\partial\theta}(\theta;t)$, while $\dot{x}(\theta;t)=\frac{\partial x}{\partial t}(\theta;t)$.
Fix $\theta\in[0,T]$. $x(\theta,t)$ is continuous in $t$ by Equation (8), and by (9),
\begin{equation}
L^{\prime}(\theta)=\frac{1}{T}\int_{0}^{T}x^{\prime}(\theta;t)dt.
\end{equation}
We next derive formulas for the derivative term $x^{\prime}(\theta;t)$.
First, suppose
that $t$ lies in the interior of an empty period
(namely, the continuous-queue analogue of an idle period; a maximal period when
$x(\theta;\cdot)=0$; see \cite{Cassandras02}). Then,
obviously, $x^{\prime}(\theta,t)=0$.
Next, consider $t$ lying in a non-empty period
(the complementary of empty periods; a supremal interval where $x(\theta;\cdot)>0$). Let $u_{t}(\theta)$ be the
starting time of the nonempty period, namely, $u_{t}(\theta):=\max\{u\leq t:x(\theta;u)=0\}$;
if no such $u$ exists, $u_{t}(\theta):=0$. Let $\ell C,(\ell+1)C,\ldots,mC$ denote the time-points in the interval
$(u_{t}(\theta),t)$ when the light switches from green to red, for some integers $\ell\geq 1$ and $m\geq \ell$.
We have the following result.
\begin{proposition}
Fix $\theta\in(0,C)$, and consider $t\in[0,T]$ such that $\beta(\theta;\cdot)$
is continuous at $t$,
and $x(\theta;t)>0$. Then
the term $x^{\prime}(\theta;t)$ has the following form,
\begin{equation}
x^{\prime}(\theta;t)=\sum_{k=\ell}^{m}\beta(\theta;(kC)^-)+\beta(\theta;t)-\beta(\theta;u_{t}(\theta)^+).
\end{equation}
For the proof we provide the diagram in Figure~2 as a visual aid.
\begin{figure}
\centering
\includegraphics[width=3.3in]{Figure_2.eps}
\caption{Evolution of $\beta(\theta;\cdot)$ (red curve) and $x(\theta;\cdot)$ (blue curve):
$u_{t}=(\ell-1)C$;
$\xi_{1}=(\ell-1)C+\theta$;
$\eta_{1}=\ell C$;
$\xi_{2}=\ell C+\theta$;
$\eta_{2}=(\ell+1)C$; $p=2$.
}
\label{Figure 2}
\end{figure}
\begin{proof}
By definition of $u_{t}(\theta)$ and Equation (8), we have that
\begin{equation}
x(\theta;t)=\int_{u_{t}(\theta)}^{t}\big(\alpha(\tau)-\beta(\theta;\tau)\big)d\tau.
\end{equation}
If $\tau$ lies in the interior of a red-signal period then
$\beta(\theta;\tau)=0$ and hence $\frac{\partial}{\partial\theta}\big(\alpha(\tau)-\beta(\theta;\tau)\big)=0$;
therefore only green-signal periods need be considered in the computation of $x^{\prime}(\theta;t)$ in the following way.
Let us denote the green-signal periods in the interval $[u_{t}(\theta),t]$ by $\Gamma_{j}$, $j=1,\ldots,p$, in increasing
order, and let $\xi_{j}(\theta)$ and $\eta_{j}(\theta)$ be the boundary points of $\Gamma_{j}$, so that
$\Gamma_{j}=[\xi_{j}(\theta),\eta_{j}(\theta)]$.
Define the functions $f_{j}(\theta)$, $j=1,\ldots,p$, as follows:
\begin{equation}
f_{j}(\theta):=\int_{\xi_{j}(\theta)^{-}}^{\eta_{j}(\theta)^{+}}\big(\alpha(\tau)-\beta(\theta;\tau)\big)d\tau,
\end{equation}
with only the following two possible exceptions: (i) For $j=1$, if $\xi_{1}(\theta)=u_{t}(\theta)$, then
the left - limit point of the integral in (14) is $u_{t}(\theta)$ and not $u_{t}(\theta)^-$;
and (ii) for $j=p$, if $\eta_{p}(\theta)=t$ then the right-limit point of the integral is
$t$ and not $t^+$. The reason for these exceptions is that the integral in (13) is taken over $[u_{t}(\theta),t]$ and its integrant
may not be valid outside this interval. Furthermore, it is obvious that
$x(\theta;t)=\sum_{j=1}^{p}f_{j}(\theta)$, and hence,
\begin{equation}
x^{\prime}(\theta;t)=\sum_{j=1}^{p}f_{j}^{\prime}(\theta).
\end{equation}
We next derive formulas for $f_{j}^{\prime}(\theta)$.
Consider a typical green-signal period $\Gamma:=[\xi(\theta),\eta(\theta)]\subset[u_{t}(\theta),t]$, and define
\begin{equation}
f(\theta)=\int_{\max\{\xi(\theta)^-,u_{t}(\theta)\}}^{\min\{\eta(\theta)^+,t\}}\big(\alpha(\tau)-\beta(\theta;\tau)\big)d\tau;
\end{equation}
note that this is like one of the functions $f_{j}(\theta)$ defined in (14) with the noted exceptions.
Let $\tau_{i}(\theta)$, $i=1,\ldots,q$, be the jump-times of $\beta(\theta;\cdot)$ in increasing order, in
the interval $(\xi(\theta),\eta(\theta))$. Suppose first that $\xi(\theta)>u_{t}(\theta)$ and $\eta(\theta)<t$, the cases where
$\xi(\theta)=u_{t}(\theta)$ or $\eta(\theta)=t$ will be considered later.
Taking derivative with respect to $\theta$ in (16), we obtain,
\begin{eqnarray}
f^{\prime}(\theta)\ =\
-\int_{\xi(\theta)}^{\tau_{1}(\theta)}\beta^{\prime}(\theta;\tau)d\tau\nonumber\\
-\sum_{i=1}^{q-1}\int_{\tau_{i}(\theta)}^{\tau_{i+1}(\theta)}\beta^{\prime}(\theta;\tau)d\tau
-\int_{\tau_{q}(\theta)}^{\eta(\theta)}\beta^{\prime}(\theta;\tau)d\tau\nonumber\\
-\sum_{i=1}^{q}\big(\beta(\theta;\tau_{i}(\theta)^{-})-\beta(\theta;\tau_{i}(\theta)^{+})\big)\tau_{i}^{\prime}(\theta)\nonumber\\
+\Big(\big(\alpha(\xi(\theta)^-)-\beta(\theta;\xi(\theta)^-)\big)\nonumber\\
-\big(\alpha(\xi(\theta)^+)
-\beta(\theta;\xi(\theta)^+)\big)\Big)\xi^{\prime}(\theta)\nonumber\\
+\Big(\big(\alpha(\eta(\theta)^-)-\beta(\theta;\eta(\theta)^-)\big)\nonumber\\
-\big(\alpha(\eta(\theta)^+)
-\beta(\theta;\eta(\theta)^+)\big)\Big)\eta^{\prime}(\theta);
\end{eqnarray}
we note that discontinuities in $\alpha(\cdot)$ do not matter since the process $\{\alpha(\cdot)\}$ is independent of
$\theta$, and further notice that $\alpha(\cdot)$ is continuous at jump-points of $\beta(\theta;\cdot)$ by Assumption
1.v.
Consider the integral terms in the RHS of (17). Since $\Gamma$ is a green-signal period contained in a nonempty period of
the queue, Equation (10) implies that, for every $\tau\in\Gamma$, $\beta(\theta;\tau)=b(\tau-kC-\theta)$ for some $k=1,2,\ldots$.
This implies that $\beta^{\prime}(\theta;\tau)=-\dot{\beta}(\theta;\tau)$, which allows
us to compute the integrals in the following way:
\begin{eqnarray}
-\int_{\xi(\theta)}^{\tau_{1}(\theta)}\beta^{\prime}(\theta;\tau)d\tau
=\int_{\xi(\theta)}^{\tau_{1}
(\theta)}
\dot{\beta}(\theta;\tau)\nonumber\\
=\beta(\theta;\tau_{1}(\theta)^-)-\beta(\theta;\xi(\theta)^+),
\end{eqnarray}
and similarly, for the rest of the integrals,
\begin{equation}
-\int_{\tau_{i}(\theta)}^{\tau_{i+1}(\theta)}\beta^{\prime}(\theta;\tau)=\beta(\theta;\tau_{i+1}(\theta)^-)-\beta(\theta;\tau_{i}(\theta)^+),
\end{equation}
and
\begin{equation}
-\int_{\tau_{q}(\theta)}^{\eta(\theta)}\beta^{\prime}(\theta;\tau)=\beta(\theta;\eta(\theta)^-)-\beta(\theta;\tau_{q}(\theta)^+).
\end{equation}
Substituting from Equations (18)-(20) in (17) we obtain,
\begin{eqnarray}
f^{\prime}(\theta)=\sum_{i=1}^{q}\big(1-\tau_{i}^{\prime}(\theta)\big)\big(\beta(\theta;\tau_{i}(\theta)^{-})
-\beta(\theta;\tau_{i}(\theta)^{+})\big)\nonumber\\
-\beta(\theta;\xi(\theta)^+)+\beta(\theta;\eta(\theta)^-)\nonumber\\
+\Big(\big(\alpha(\xi(\theta)^-)-\beta(\theta;\xi(\theta)^-)\big)\nonumber\\
-\big(\alpha(\xi(\theta)^+)
-\beta(\theta;\xi(\theta)^+)\big)\Big)\xi^{\prime}(\theta)\nonumber\\
+\Big(\big(\alpha(\eta(\theta)^-)-\beta(\theta;\eta(\theta)^-)\big)\nonumber\\
-\big(\alpha(\eta(\theta)^+)
-\beta(\theta;\eta(\theta)^+)\big)\Big)\eta^{\prime}(\theta).
\end{eqnarray}
Next, we observe that a green-signal period contained in the interval $(u_{t}(\theta),t)$ ends at a time-point
$kC$, $k=1,\ldots$, meaning that $\eta(\theta)=kC$ which is independent of $\theta$, and hence
$\eta^{\prime}(\theta)=0$. This implies that the last additive term in Equation (21) is zero.
Furthermore, each time $\tau_{i}(\theta)$ lies in the interior of the green-signal period $\Gamma$, and hence there exists an open interval
containing it where, by Equation (10), $\beta(\theta,\tau)=b(\tau-kC-\theta)$; this implies that $\tau_{i}^{\prime}(\theta)=1$, which annuls
the first additive term in the RHS of (21). All of this reduces (21) to the following equation,
\begin{eqnarray}
f^{\prime}(\theta)=
-\beta(\theta;\xi(\theta)^+)+\beta(\theta;\eta(\theta)^-)\nonumber\\
+\Big(\big(\alpha(\xi(\theta)^-)-\beta(\theta;\xi(\theta)^-)\big)\nonumber\\
-\big(\alpha(\xi(\theta)^+)
-\beta(\theta;\xi(\theta)^+)\big)\Big)\xi^{\prime}(\theta).
\end{eqnarray}
Next, the starting time of a green period has the form $\xi(\theta)=kC+\theta$,
$k=1,\ldots$, and hence $\xi^{\prime}(\theta)=1$. Moreover, by Assumption 1.iv, $\alpha(\tau)$ is continuous
at $\tau=\xi(\theta)$ and hence $\alpha(\xi(\theta)^-)=\alpha(\xi(\theta)^+)$, and finally,
$\beta(\theta;\xi(\theta)^-)=0$ since $\xi(\theta)^-$ lies in a red-signal period.
All of this reduces (22) to
\begin{equation}
f^{\prime}(\theta)=\beta(\theta;\eta(\theta)^-).
\end{equation}
Consider now the case where $\xi(\theta)=u_{t}(\theta)$. Then, the lower boundary of the integral
in (16) is $u_{t}(\theta)$, and the corresponding boundary condition in (17) becomes
$-\big(\alpha(u_{t}(\theta)^+)-\beta(\theta;u_{t}(\theta)^+)\big)u_{t}^{\prime}(\theta)$ instead of
$\Big(\big(\alpha(\xi(\theta)^-)-\beta(\theta;\xi(\theta)^-)\big)
-\big(\alpha(\xi(\theta)^+)
-\beta(\theta;\xi(\theta)^+)\big)\Big)\xi^{\prime}(\theta)$.
As a result, (22) becomes
\begin{eqnarray}
f^{\prime}(\theta)=-\beta(\theta;u_{t}(\theta)^+)+\beta(\theta;\eta(\theta)^-)\nonumber\\
-\big(\alpha(u_{t}(\theta)^+)-\beta(\theta;u_{t}(\theta)^+)\big)u_{t}^{\prime}(\theta).
\end{eqnarray}
We now assert that
\begin{equation}
\big(\alpha(u_{t}(\theta)^+)-\beta(\theta;u_{t}(\theta)^+)\big)u_{t}^{\prime}(\theta)=0.
\end{equation}
Recall that $u_{t}(\theta)$ is the time a non-empty period starts at the queue.
There are three ways a non-empty period can start: while the queue is empty, (i) $\alpha(\cdot)$ jumps up;
(ii) $\beta(\theta;\cdot)$ jump down; and (iii) $\alpha(\cdot)-\beta(\theta;\cdot)$ rises in a continuous fashion from non-positive to positive. In the first case $u_{t}(\theta)$ is a jump time of $\alpha(\cdot)$, and since the latter process is independent
of $\theta$, $u_{t}^{\prime}(\theta)=0$. In the second case, the only way $\beta(\theta,\cdot)$
can jump down is at the start of
red-signal periods. In that case $u_{t}(\theta)=kC$ for some $k=1,\ldots$, and again $u_{t}^{\prime}(\theta)=0$. In the third case,
$\alpha(u_{t}(\theta))-\beta(\theta;u_{t}(\theta))=0$. In all three cases,
(25) is satisfied.
Applying (25) to (24), we obtain that
\begin{equation}
f^{\prime}(\theta)=-\beta(\theta;u_{t}(\theta)^+)+\beta(\theta;\eta(\theta)^-).
\end{equation}
Finally, consider the case where $t=\eta(\theta)$. Then, in Equation (21) we have that $\eta^{\prime}(\theta)=0$ as before,
and hence the derivations of Equations (23) and (26) remain unchanged.
Consider now Equation (15). For $j=1$, if $\xi_{1}(\theta)=u_{t}(\theta)$ then Equation (26) applies
to $f_{1}^{\prime}(\theta)$. On the other hand, if $\xi_{1}(\theta)>u_{t}(\theta)$ then Equation (23) applies
to $f_{1}^{\prime}(\theta)$, but in this case (by definition of $u_{t}(\theta)$) $u_{t}(\theta)^+$ lies in a red-signal period
and hence $\beta(\theta;u_{t}(\theta)^+)=0$; implying that (26) applies as well. Thus, in any event,
$f_{1}^{\prime}(\theta)=-\beta(\theta;u_{t}(\theta)^+)+\beta(\theta;\eta_{1}(\theta)^{-})$. For every $j=2,\ldots,p$, $\xi_{j}(\theta)>u_{t}(\theta)$ and hence Equation (23) applies, namely, $f_{j}^{\prime}(\theta)=\beta(\theta;\eta_{j}(\theta)^-)$.
Therefore, by summing up all the terms in (15), Equation (12) follows.
\end{proof}
\end{proposition}
Equation (12) requires the on-line monitoring of traffic-flow rates, and this can be done by measuring the speed of automobiles
crossing the intersection. In the special case where the service rate alternates between 0 and $\beta_{m}$,
$\beta(\theta;t)$ can be directly determined by the color of the traffic light.
Finally, we point out that the sample function $x(\theta;t)$ evidently is continuous in $\theta$ for every given $t$,
and its derivative $x^{\prime}(\theta,t)$ is monotone nondecreasing in $\theta$ (see (12)). This implies that $L(\theta)$ is
continuous as well, and the IPA derivative
$L^{\prime}(\theta)$ is unbiased.
\section{Simulation Examples}
This section presents simulation examples for testing the effectiveness of our regulation technique.
The traffic-light cycle is $C=1$, and the control cycle consists of 20 light cycles.
The arrival rate consists of a off/on model where, in the {\it off} stage $\alpha(t)=0$, while for each {\it on}
stage $\alpha(t)$ is uniformly distributed in an interval
$[(1-\zeta)\bar{\alpha},(1+\zeta)\bar{\alpha}]$; we chose its mean to be $\bar{\alpha}=4.1$, and consider
different values of $\zeta>0$.
The arrival rate $\alpha(t)$ varies from one
{\it on} period to the next
but retains a constant value throughout each {\it on} period. The durations of {\it off} periods and {\it on}
periods are drawn from the uniform distributions
on the intervals $[0,0.02]$ and $[0,0.063]$, respectively. The service rate $\beta(\theta;t)$ ramps up at the start of
each green-signal period at the rate of $0.62$, until either it reaches the saturation level of $\beta_{m}:=5.0$ or an empty period starts. In
the latter case the service rate jumps to $\beta_{m}$, and in both cases $\beta(\theta;t)$ remains at the level of $\beta_{m}$ to the end of the green-cycle period.
The set-point reference value is $r=0.3$, and the initial control variable was set to
$\theta_{1}=0.9$.
We chose $\zeta=0.3$ and $\zeta=0.1$,
respectively, and thus, the variance of the arrival process is larger in the first experiment than in the second one.
Figure~3 depicts the graphs of the obtained outputs $L_{n}(\theta_{n})$ as functions of the counter $n=1,\ldots,50$,
while Figure~4 provides the same information for $n=10,\ldots,50$ in order to highlight the effects of the variance on the asymptotic behavior of the outputs.
In both figures the blue and red graphs correspond to the respective cases of $\zeta=0.3$ and $\zeta=0.1$.
In Figure 3 the graphs are hardly distinguishable, and both exhibit convergence to about $r=0.3$ in about 10 iterations.
In Figure 4 the differences are more evident, and the two graphs exhibit variability about the target value of $0.3$. This
is expected in view of the variations in $L_{n}(\theta_{n})$ which are due to the random elements of the system
and the fact that it is nowhere near steady state after $20$ light cycles. However, the respective means over the last 41 iterations, namely
the quantities $\frac{1}{41}\sum_{n=10}^{50}L_{n}(\theta_{n})$, are $0.3011$ for the case where $\zeta=0.3$, and $0.305$ for the case
where $\zeta=0.1$.
\begin{figure}
\centering
\includegraphics[width=3.6in]{Figure_3.eps}
\caption{Evolution of $L_{n}:\ n=1\;\dots,50$}
\label{Figure 3}
\end{figure}
\begin{figure}[!t]
\centering
\includegraphics[width=3.6in]{Figure_4.eps}
\caption{Evolution of $L_{n}:\ n=10\;\dots,50$}
\label{Figure 1}
\end{figure}
Finally, Figure~5 shows plots of the control variable $\theta_{n}$ as functions of $n$ for the case where $\zeta=0.3$,
for two runs with the respective starting values of $\theta_{1}=0.9$ (the blue graph) and $\theta_{1}=0.1$ (the green graph). Not surprisingly, both
settle to roughly the same value ($\theta\sim 0.25$) after 10 iterations.
We point out that the flat part of the blue curve at iterations 3-5 is due to a lower-bound guard on $\theta$
at $0.1$,
designed
to prevent extreme values which could destabilize the system.
\begin{figure}[!t]
\centering
\includegraphics[width=3.6in]{Figure_5.eps}
\caption{Evolution of $\theta_{n}:\ n=1,\ldots,50$}
\label{g}
\end{figure}
\section{Conclusions and Future Work}
This paper proposes a regulation technique for congestion management in a traffic-light intersection. The technique aims at
tracking a given reference queue level at the light in the face of variable traffic patterns. It
is based on the simple idea of an integral controller with a variable gain, adjusted according to the IPA derivative of the plant function. The main theoretical result concerns a simple formula for the IPA derivative, which is computable from traffic
rates that can be measured on-line.
Simulation results exhibit fast convergence towards the set value, and suggest the potential viability of
our approach in eventual applications.
Future research concerns extensions of our control formulation to grids of traffic light with cross-correlated traffic patterns. On the theoretical
side, the main question is how to design regulators for traffic-light systems with multiple controllers.
On the practical side, the main issue is how to apply the proposed technique to achieve
effective regulation under more realistic traffic conditions.
|
\section{Introduction}
In this contribution, we make use of a numerical
technique which can be used to compute the effective actions of
external field configurations. The technique, called either worldline
numerics or the Loop-Cloud Method, was first used by Gies and
Langfeld~\cite{Gies:2001zp} and has since been applied to computation
of effective actions
\cite{Gies:2001tj,Langfeld:2002vy,Gies:2005sb,Gies:2005ym,Dunne:2009zz}
and Casimir energies
\cite{Moyaerts:2003ts,Gies:2003cv,PhysRevLett.96.220401}. More
recently, the technique has also been applied to pair
production~\cite{2005PhRvD..72f5001G} and the vacuum polarization
tensor~\cite{PhysRevD.84.065035}. Worldline numerics
is able to compute quantum effective actions in the one-loop
approximation to all orders in both
the coupling and in the external field, so it is well suited to
studying non-perturbative aspects of quantum field theory in strong
background fields. Moreover, the technique maintains gauge invariance
and Lorentz invariance. The key idea of the technique is that a path
integral is approximated as the average of a finite set of $N_l$
representative closed paths (loops) through spacetime. These loops
are not mapped to any spacetime lattice, so the theory maintains
Lorentz invariance and is distinct from Lattice-based techniques. We use a
standard Monte-Carlo procedure to generate loops which have large
contributions to the loop average.
We will focus on calculations of the \ac{QED}
effective action in cylindrically symmetric, extended tubes of magnetic
flux using the worldline numerics method. These
configurations may be called flux tubes, strings, or vortices,
depending on the context. Flux tubes are of interest in astrophysics
because they describe magnetic structures near stars and planets,
cosmic strings~\cite{vilenkin2000cosmic}, and vortices in the
superconducting core of neutron stars \cite{2006pfsb.book..135S,
schmitt2010dense}. Outside of astrophysics, magnetic vortex systems
are at the forefront of research in condensed matter physics for the
role they play in superconducting systems and in \ac{QCD} research for
their relation to center vortices, a gluonic configuration analogous
to magnetic vortices which is believed to be important to quark
confinement~\cite{tHooft19781, 2003PrPNP..51....1G}. Currently,
we are most interested in the roles played by magnetic flux tubes in
neutron star cores.
Our motivation for discussing flux tubes comes from the fact that
superconductivity is predicted in the nuclear matter of neutron stars
and that some superconducting materials produce a lattice of flux
tubes when placed in an external magnetic field. Superconductivity is
a macroscopic quantum state of a fluid of fermions that, most notably,
allows for the resistanceless conduction of charge. In 1933, Meissner
and Ochsenfeld observed that magnetic fields are repelled from
superconducting materials~\cite{meisner33}. In 1935, F. and H. London
described the Meissner effect in terms of a minimization of the free
energy of the superconducting current~\cite{london35}. Then, in 1957,
by studying the superconducting electromagnetic equations of motion in
cylindrical coordinates, Abrikosov predicted the possible existence of
line defects in superconductors which can carry quantized magnetic
flux through the superconducting material \cite{abrikosov57}.
A more complete microscopic description of superconducting materials
is given by BCS (Bardeen, Cooper, and Schrieffer)
theory~\cite{PhysRev.108.1175}.
Interested readers may pursue more thorough
reviews of superconductivity and superfluidity in neutron
stars~\cite{2006pfsb.book..135S, schmitt2010dense}.
Our calculations use the worldline numerics method which is reviewed
in detail in section~\ref{sec:qed-effect-act}. In section
\ref{sec:background} of this article, we will briefly review the
physics of magnetic flux tubes and of nuclear superconductivity in
neutron stars to provide context and motivation. We present our models
for solitary flux tubes and dense flux tube lattices in section
\ref{sec:calculations} as well as the details of the calculations.
The results of our calculations for scalar and spinor \ac{QED} are
presented in section \ref{sec:periodic_results}. Our results suggest a
small but possibly influential Casimir interaction between flux tubes
in a dense lattice that may cause the flux tubes to form bunches. This
result and other implications of our calculations are discussed in
section \ref{sec:conclusion}.
\section{QED Effective Action on the Worldline}
\label{sec:qed-effect-act}
Worldline numerics is built on the worldline formalism which was
initially invented by Feynman~\cite{PhysRev.80.440, PhysRev.84.108}.
Much of the recent interest in this formalism is based on the work of
Bern and Kosower, who derived it from the infinite string-tension
limit of string theory and demonstrated that it provided an efficient
means for computing amplitudes in QCD~\cite{PhysRevLett.66.1669}.
For this reason, the worldline formalism is often referred to as
`string inspired'. However, the formalism can also be obtained
straight-forwardly from first-quantized field
theory~\cite{1992NuPhB.385..145S}, which is the approach we will adopt
here. In this formalism the degrees of freedom of the field are
represented in terms of one-dimensional path integrals over an
ensemble of closed trajectories.
\begin{widetext}
We begin with the QED effective action expressed in the proper-time
formalism \cite{Schwinger:1951},
\begin{equation}
\label{eqn:trln}
\mathrm{Tr~ln}\left[\frac{\slashed{p}+e\slashed{A}_\mu^0
-m}{\slashed{p}-m}\right] = -\frac{1}{2}\int d^4x \int_0^\infty
\frac{dT}{T}e^{-iTm^2}
\times \mathrm{tr}\biggr( \bra{x}e^{iT(\slashed{p}
+e\slashed{A}^0_\mu)^2}\ket{x}
- \bra{x}e^{iTp^2}\ket{x}\biggr).
\end{equation}
To evaluate $\bra{x}e^{iT(\slashed{p}_\mu +
e\slashed{A}_\mu)^2}\ket{x}$, we recognize that it is simply the
propagation amplitude $\braket{x,T}{x,0}$ from ordinary quantum
mechanics with $(\slashed{p}_\mu + e\slashed{A}_\mu)^2$ playing the
role of the Hamiltonian. We therefore express this factor in its path
integral form:
\begin{equation}
\bra{x}e^{iT(\slashed{p}_\mu + e\slashed{A}_mu)^2}\ket{x} = \mathcal{N}
\int \mathcal{D}x_\rho(\tau) e^{-\int_0^T d\tau \left[\frac{\dot{x}^2(\tau)}{4}
+ i A_\rho x^\rho(\tau)\right]}
\times \frac{1}{4} {\rm tr}e^{\frac{1}{2}\int_0^T d\tau \sigma_{\mu \nu}F^{\mu \nu}(x_{\rm CM}+x(\tau))}.
\end{equation}
$\mathcal{N}$ is a normalization constant that we can fix by using
our renormalization condition that the fermion determinant should
vanish at zero external field:
\begin{equation}
\bra{x}e^{iTp^2}\ket{x} = \mathcal{N}\int \mathcal{D}
x_p(\tau)e^{-\int_0^T d\tau\frac{\dot{x}^2(\tau)}{4}}
= \int \frac{d^4p}{(2\pi T)^4}\bra{x}e^{iTp^2}\ket{p}\braket{p}{x}
= \frac{1}{(4\pi T)^2},
\end{equation}
We may now write
\begin{equation}
\mathcal{N}\int \mathcal{D}x_\rho(\tau) e^{-\int_0^T d\tau[\frac{\dot{x}^2(\tau)}{4}+iA_\rho x^\rho(\tau)]}
\frac{1}{4} {\rm tr}e^{\frac{1}{2}\int_0^T d\tau \sigma_{\mu \nu}F^{\mu \nu}(x_{\rm CM}+x(\tau))}
= \frac{\left\langle e^{-i\int_0^T d\tau A_\rho x^\rho(\tau)}\frac{1}{4} {\rm tr}e^{\frac{1}{2}\int_0^T d\tau \sigma_{\mu \nu}F^{\mu \nu}(x_{\rm CM}+x(\tau))}\right\rangle_x}{(4\pi T)^2} ,
\end{equation}
where
\begin{equation}
\label{eqn:meandef}
\mean{\hat{\mathcal{O}}}_x = \frac{\int \mathcal{D}x_\rho(\tau) \hat{\mathcal{O}}
e^{-\int_0^T d\tau\frac{\dot{x}^2(\tau)}{4}}}{\int \mathcal{D}x_\rho(\tau)
e^{-\int_0^T d\tau\frac{\dot{x}^2(\tau)}{4}}}
\end{equation}
is the weighted average of the operator $\hat{\mathcal{O}}$ over an ensemble of closed particle
loops with a Gaussian velocity distribution.
Finally, combining all of the equations from this section results in the
renormalized one-loop effective action for \ac{QED} on the worldline:
\begin{equation}
\label{eqn:QEDWL}
\Gamma^{(1)}[A_\mu] = \frac{2}{(4\pi)^2}\int_0^\infty
\frac{dT}{T^3}e^{-m^2T}\int d^4x_\mathrm{CM} \times
\left[\left\langle e^{i\int_0^Td\tau A_\rho(x_\mathrm{CM}+x(\tau))\dot{x}^\rho(\tau)}
\frac{1}{4} {\rm tr}e^{\frac{1}{2}\int_0^T d\tau \sigma_{\mu \nu}F^{\mu \nu}(x_{\rm CM}+x(\tau))}\right\rangle _x -1\right].
\end{equation}
\end{widetext}
\subsection{Worldline Numerics}
The averages, $\mean{\hat{\mathcal{O}}}$, defined by equation (\ref{eqn:meandef})
involve
functional integration over every possible closed path through spacetime. The velocities along the paths are drawn from a Gaussian velocity distribution.
The prescription of the worldline numerics technique is to compute
these averages approximately using a finite set of $N_l$ representative loops
on a computer. The worldline average is then approximated as the mean of
an operator evaluated along each of the worldlines in the ensemble:
\begin{equation}
\mean{\hat{\mathcal{O}}[x(\tau)]} \approx
\frac{1}{N_l} \sum_{i=1}^{N_l} \hat{\mathcal{O}}[x_i(\tau)].
\end{equation}
\subsubsection{Loop Generation}
\label{sec:loopgen}
The velocity distribution for the loops depends on
the proper time, $T$. However, generating a separate ensemble of loops for
each value of $T$ would be very computationally expensive. This problem is alleviated by generating
a single ensemble of loops, $\vec{y}(\tau)$, representing unit proper time,
and scaling those loops accordingly for different values of $T$:
\begin{equation}
\vec{x}(\tau) = \sqrt{T}\vec{y}(\tau/T) ,
\end{equation}
\begin{equation} \int_0^T d\tau \vec{\dot{x}}^2(\tau) \rightarrow \int_0^1 dt
\vec{\dot{y}}^2(t).
\end{equation}
There is no way to treat the integrals as continuous as we generate our loop
ensembles. Instead, we treat the integrals as sums over discrete points
along the proper-time interval $[0,T]$. This is fundamentally different
from space-time discretization, however. Any point on the worldline loop
may exist at any position in space, and $T$ may take on any value. It is
important to note this distinction because the worldline method retains
Lorentz invariance while lattice techniques, in general, do not.
The challenge of loop cloud generation is in generating a discrete set
of points on a unit loop which obeys the prescribed velocity distribution.
There are a number of different algorithms for achieving this goal that have
been discussed in the literature. Four possible algorithms are compared
and contrasted in \cite{Gies:2003cv}. In this work, we chose an algorithm
dubbed ``d-loops", which was first described
in \cite{Gies:2005sb}. To generate a ``d-loop", the number of points is iteratively
doubled, placing the new points in a Gaussian distribution between the existing neighbour points.
We quote the algorithm directly:
\begin{enumerate}
\item Begin with one arbitrary point
$N_0=1$, $y_{N}$.
\item Create an $N_1=2$ loop, i.e., add a point $y_{N/2}$ that is
distributed in the heat bath of $y_N$ with
\begin{equation}
e^{-\frac{N_1}{4} 2 (y_{N/2} -y_{N})^2}. \label{yn2}
\end{equation}
\item Iterate this procedure, creating an $N_k=2^k$ points
per line (ppl) loop by adding $2^{k-1}$ points
$y_{{qN}/{N_k}}$, $q=1,3,\dots, N_k-1$ with distribution
\begin{equation}
e^{-\frac{N_k}{4} 2 [y_{qN/N_k} -\frac{1}{2}(y_{(q+1)N/N_k}+
y_{(q-1)N/N_k})]^2}. \label{ynk}
\end{equation}
\item Terminate the procedure if $N_k$ has reached $N_k=N$ for
unit loops with $N$ ppl.
\item For an ensemble with common center of mass, shift each
whole loop accordingly.
\end{enumerate}
The above d-loop algorithm was selected since it is simple and about
10\% faster than previous algorithms, according to its developers,
because it requires fewer algebraic operations. The generation of the
loops is largely independent from the main program. Because of this,
it was simpler to generate the loops using a Matlab script.
This function was used to produce text files containing the worldline data for
ensembles of loops. These text files were read into memory at the
launch of each calculation. The results of this generation routine can
be seen in figure \ref{fig:worldlineplot}.
When the \ac{CUDA} kernel is called, every thread
in every block executes the kernel function with its own unique
identifier. Therefore, it is best to generate worldlines in integer
multiples of the number of threads per block.
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=2cm 1cm 2cm 1cm]{images/worldlineplot}
\caption[Discretization of the worldline loop]
{A single discrete worldline loop shown at several levels
of discretization. The loops form fractal patterns and have a strong
parallel with Brownian motion. The colour
represents the phase of a particle travelling along the loop, and
begins at dark blue, progresses in a random walk through yellow,
and ends at dark
red. The total flux through this particular worldline at $T=1$ and
$B=B_k$ is about $0.08 \pi/e$.}
\label{fig:worldlineplot}
\end{figure}
\subsection{Cylindrical Worldline Numerics}
We now consider cylindrically symmetric external magnetic fields.
In this case, we may simplify (\ref{eqn:QEDWL}),
\begin{eqnarray}
\label{eqn:cylEA}
\frac{\Gamma^{(1)}_\mathrm{ ferm}}{T
L_z} &=& \frac{1}{4\pi} \int_0^\infty \rho_\mathrm{ cm}
\biggr[ \int_0^\infty \frac{dT}{T^3}e^{-m^2T} \times \\
& & ~~~~~~ \left\{\langle
W\rangle_{\vec{r}_\mathrm{ cm}} - 1 -\frac{1}{3}(eB_\mathrm{
cm}T)^2\right\}\biggr]d\rho. \nonumber
\end{eqnarray}
\subsubsection{Cylindrical Magnetic Fields}
\label{sec:cylindrical}
We have $\vec{B} =
B(\rho)$\bhattext{z} with
\begin{equation} \label{eqn:BWLN} B(\rho) = \frac{A_\phi(\rho)}{\rho} +
\frac{dA_\phi(\rho)}{d\rho} \end{equation} if we make the gauge choice that $A_0 =
A_\rho = A_z = 0$.
We begin by considering $A_\phi(\rho)$ in the form
\begin{equation} A_\phi(\rho) = \frac{F}{2\pi \rho}f_\lambda(\rho) \end{equation} so that \begin{equation}
B_z(\rho)=\frac{F}{2\pi\rho}\frac{df_\lambda(\rho)}{d\rho} \end{equation} and the total
flux is \begin{equation} \Phi=F(f_\lambda(L_\rho)-f_\lambda(0)). \end{equation} It is convenient
to express the flux in units of $\frac{2 \pi}{e}$ and define a dimensionless
quantity
\begin{equation}
\mathcal{F}=\frac{e}{2 \pi} F.
\end{equation}
\subsubsection{Wilson Loop}
The quantity inside the angled brackets in equation (\ref{eqn:QEDWL}) is a
gauge invariant observable called a Wilson loop. We note that the proper time
integral provides a natural path ordering for this operator.
The Wilson loop expectation value is
\begin{equation}
\label{eqn:wilsonloop}
\langle W\rangle_{\vec{r}_\mathrm{cm}} \!\! = \biggl \langle e^{ie\int_0^T d\tau
\vec{A}(\vec{r}_{\mathrm{ cm}} + \vec{r}(\tau)) \cdot \dot{\vec{r}}}
\frac{1}{4} \mathrm{ tr} e^{\frac{e}{2}\int_0^T d\tau
\sigma_{\mu \nu}F_{\mu \nu}(\vec{r}_{\mathrm{ cm}} + \vec{r}(\tau))}\biggr
\rangle_{\vec{r}_\mathrm{ cm}} \!\!\!\!\!\!\!\!,
\end{equation}
which we look at as a product between a scalar part ($e^{ie\int_0^T d\tau
\vec{A}(\vec{r}_{\mathrm{ cm}} + \vec{r}(\tau)) \cdot \dot{\vec{r}}}$)
and a fermionic part ($\frac{1}{4} \mathrm{ tr} e^{\frac{e}{2}\int_0^T d\tau
\sigma_{\mu \nu}F_{\mu \nu}(\vec{r}_{\mathrm{ cm}} + \vec{r}(\tau))}$).
\subsubsection{Scalar Part}
In a magnetic field, the scalar part is related to the flux through
the loop, $\Phi_B$, by Stokes theorem:
\begin{eqnarray}
e^{ie\int_0^T d\tau
\vec{A} \cdot \dot{\vec{r}}} &=&
e^{ie\oint \vec{A}\cdot d\vec{r}} = e^{ie\int_{\vec{\Sigma}} \vec{\nabla}\times\vec{A} \cdot d\vec{\Sigma}}\\
& = & e^{ie\int_{\vec{\Sigma}} \vec{B} d\vec{\Sigma}} = e^{ie\Phi_B}.
\end{eqnarray}
Consequently, this factor accounts for the Aharonov-Bohm phase acquired by
particles in the loop.
The loop discretization results in the following approximation of the
scalar integral:
\begin{equation}
\oint \vec{A}(\vec{r})\cdot d\vec{r} = \sum_{i=1}^{N_\mathrm{ppl}}
\int_{\vec{r}^i}^{\vec{r}^{i+1}}\vec{A}(\vec{r})\cdot d\vec{r}.
\end{equation}
Using a linear parameterization of the positions, the line integrals are
\begin{equation}
\int_{\vec{r}^i}^{\vec{r}^{i+1}}\vec{A}(\vec{r})\cdot d\vec{r} =
\int_0^1dt \vec{A}(\vec{r}(t))\cdot(\vec{r}^{i+1} - \vec{r}^i).
\end{equation}
Using the same gauge choice outlined above ($\vec{A}=A_\phi \hat{\phi}$),
we may write
\begin{equation}
\vec{A}(\vec{r}(t)) = \frac{\mathcal{F}}{e\rho^2}
f_\lambda(\rho^2)(-y,x,0),
\end{equation}
where we have chosen $f_\lambda(\rho^2)$ to depend on $\rho^2$ instead
of $\rho$ to simplify some expressions and to
avoid taking many costly square roots in the worldline numerics.
We then have
\begin{equation} \int_{\vec{r}^i}^{\vec{r}^{i+1}}\vec{A}(\vec{r})\cdot
d\vec{r} = \mathcal{F} (x^iy^{i+1}-y^i x^{i+1})\int_0^1
dt\frac{f_\lambda(\rho_i^2(t))}{\rho_i^2(t)}. \end{equation} The linear interpolation
in Cartesian coordinates gives
\begin{equation}
\label{eqn:rhoi} \rho_i^2(t) = A_i + 2B_it + C_i t^2,
\end{equation}
where
\begin{eqnarray}
A_i &=& (x^i)^2 + (y^i)^2 \\
B_i&=&x^i(x^{i+1} - x^i) + y^i(y^{i+1}-y^i)\\
C_i &= &(x^{i+1}-x^i)^2 + (y^{i+1}-y^i)^2.
\label{eqn:Ci}
\end{eqnarray}
In performing the integrals along the straight lines connecting
each discretized loop point, we are in danger of violating gauge invariance.
If these integrals can be performed analytically, than gauge invariance
is preserved exactly. However, in general, we wish to compute these integrals
numerically. In this case, gauge invariance is no longer guaranteed, but
can be preserved to any numerical precision that's desired.
\subsubsection{Fermion Part}
For fermions, the Wilson loop is modified by a factor,
\begin{eqnarray}
W^\mathrm{ ferm.} &=& \frac{1}{4}\mathrm{ tr}\left(e^{\frac{1}{2}
e\int_0^T d\tau \sigma_{\mu \nu}F^{\mu \nu}}\right)\\
&=& \frac{1}{4}\mathrm{ tr}\left(e^{\sigma_{x y}
e\int_0^T d\tau B\left(x(\tau)\right)}\right) \\
&=& \cosh{\left(e \int_0^T d\tau B\left(x(\tau)\right)\right)}\\
&=& \cosh{\left(2\mathcal{F}\int_0^T
d\tau f'_\lambda(\rho^2(\tau))\right)},
\label{eqn:Wfermfpl}
\end{eqnarray}
where we have used the relation
\begin{equation}
eB = 2\mathcal{F}\frac{d f_\lambda(\rho^2)}{d \rho^2} =
2\mathcal{F}f'_\lambda(\rho^2).
\end{equation}
This factor represents an additional contribution to the
action because of the spin interaction with the magnetic field.
Classically, for a particle with a magnetic moment $\vec{\mu}$
travelling through a magnetic field in a time $T$, the
action is modified by a term given by
\begin{equation}
\Gamma^0_\mathrm{ spin} = \int_0^T \vec{\mu} \cdot \vec{B}(\vec{x}(\tau)) d\tau.
\end{equation}
The magnetic moment is related to the electron spin
$\vec{\mu} = g\left(\frac{e}{2m}\right)\vec{\sigma}$,
so we see that the integral in the above quantum fermion factor is
very closely related to the classical action
associated with transporting a magnetic moment through a magnetic field:
\begin{equation}
\Gamma^0_\mathrm{ spin} = g\left(\frac{e}{2m}\right) \sigma_{x y} \int_0^T B_z(x(\tau))d\tau.
\end{equation}
Qualitatively, we could write
\begin{equation}
W^\mathrm{ ferm} \sim \cosh{\left(\Gamma^0_\mathrm{ spin}\right)}.
\end{equation}
As a possibly useful aside, we may want to express
the integral in terms of $f_\lambda(\rho^2)$ instead of its derivative.
We can do this by integrating by parts:
\begin{eqnarray}
\int_0^T d\tau f'_\lambda(\rho^2(\tau))&=&
\frac{T}{N_\mathrm{ ppl}}\sum_{i=1}^{N_\mathrm{ ppl}}\int_0^1
dt f'_\lambda(\rho^2_i(\tau)) \\
& = &
\frac{T}{N_\mathrm{ ppl}}\sum_{i=1}^{N_\mathrm{ ppl}}\biggr[
\frac{f_\lambda(\rho^2_i(t))}{2(B_i+C_it)}\biggr|^{t=1}_{t=0} \nonumber \\
& & +\frac{C_i}{2} \!\!\int_0^1\!\!\!\!
\frac{f_\lambda(\rho^2_i(t))}{(B_i+C_i t)^2} dt\biggr] \\
& = & \frac{T}{N_\mathrm{ ppl}}\sum_{i=1}^{N_\mathrm{
ppl}}\frac{C_i}{2}\!\!\int_0^1\!\!\!\! \frac{f_\lambda(\rho^2_i(t))}{(B_i+C_i t)^2} dt,
\end{eqnarray}
with $\rho_i^2(t)$ given by equations (\ref{eqn:rhoi}) to (\ref{eqn:Ci}).
The second equality is obtained from integration-by-parts. In the third
equality, we use the loop sum to cancel the boundary terms in pairs:
\begin{equation}
\label{eqn:Wfermfl}
W^{\mathrm{ ferm.}} = \cosh{\left(\frac{\mathcal{F}T}{N_\mathrm{
ppl}}\sum_{i=1}^{N_{\mathrm{ ppl}}}
C_i \int_0^1 dt \frac{f_\lambda(\rho^2_i(t))}{(B_i+C_i t)^2}\right)}.
\end{equation}
In most cases, one would use equation (\ref{eqn:Wfermfpl}) to compute the fermion factor
of the Wilson loop. However,
equation (\ref{eqn:Wfermfl}) may be useful in cases where $f'_\lambda(\rho^2(\tau))$
is not known or is difficult to compute.
\subsubsection{Renormalization}
The field strength renormalization counterterms result from the small $T$
behaviour of the worldline integrand. In the limit where $T$ is very small,
the worldline loops are very localized around their center of mass. So,
we may approximate their contribution as being that of a constant field
with value $\vec{A}(\vec{r}_{\mathrm{ cm}})$. Specifically, we require that the field
change slowly on the length scale defined by $\sqrt{T}$. This condition on
$T$ can be written
\begin{equation}
T \ll \left|\frac{m^2}{e B'(\rho^2)}\right|
= \left| \frac{m^2}{2\mathcal{F}f''_\lambda(\rho^2_\mathrm{ cm})}\right|.
\end{equation}
When this limit is satisfied, we may use the exact expressions for the
constant field Wilson loops to determine the small $T$ behaviour of the
integrands and the corresponding counterterms.
The Wilson loop averages for constant magnetic fields in scalar and
fermionic \ac{QED} are
\begin{equation}
\mean{W}_\mathrm{ ferm} = eBT\coth{(eBT)}
\end{equation}
and
\begin{equation}
\mean{W}_\mathrm{ scal} = \frac{eBT}{\sinh{(eBT)}}.
\end{equation}
\begin{widetext}
Therefore, the integrand for fermionic \ac{QED} in the limit of small $T$ is
\begin{eqnarray} \label{eqn:fermI} I_\mathrm{ ferm}(T) &=&
\frac{e^{-m^2T}}{T^3}\left[eB(\vec{r}_{cm})T\coth{(eB(\vec{r}_{cm})T)}
- 1 -\frac{e^2}{3} B^2(\vec{r}_{cm})T^2\right] \nonumber \\ &\approx&-\frac{(eB)^4
T}{45}+\frac{1}{45} (eB)^4 m^2 T^2+\left( \frac{2 (eB)^6}{945}-\frac{(eB)^4
m^4}{90}\right)T^3
+\frac{(7 (eB)^4 m^6-4 (eB)^6 m^2)T^4 }{1890}+O(T^5).
\end{eqnarray}
For scalar QED we have
\begin{eqnarray}
\label{eqn:scalI} I_\mathrm{ scal}(T) &=&
\frac{e^{-m^2T}}{T^3}\left[\frac{eB(\vec{r}_{cm})T}{\sinh{(eB(\vec{r}_{cm})T)}}
- 1 +\frac{1}{6} (eB)^2(\vec{r}_{cm})T^2\right] \nonumber \\ &\approx&\frac{7
(eB)^4 T}{360}-\frac{7 (eB)^4 m^2 T^2}{360}+\frac{(147 (eB)^4 m^4-31
(eB)^6)T^3 }{15120} +\frac{ (31 (eB)^6 m^2-49 (eB)^4
m^6)T^4}{15120}+O(T^5).
\end{eqnarray}
\end{widetext}
Beyond providing the renormalization conditions, these expansions can
be used in the small $T$ regime to avoid a problem with the Wilson
loop uncertainties in this region. Consider the uncertainty in the
integrand arising from the uncertainty in the Wilson loop:
\begin{equation}
\delta I(T) = \frac{\partial I}{\partial W} \delta W = \frac{e^{-m^2 T}}{T^3} \delta W.
\end{equation}
In this case, even though we can compute the Wilson loops for small $T$
precisely, even a small uncertainty is magnified by a divergent factor when
computing the integrand for small values of $T$. So, in order to perform
the integral, we must replace the small $T$ behaviour of the integrand with
the above expansions (\ref{eqn:fermI}) and (\ref{eqn:scalI}). Our worldline
integral then proceeds by analytically computing the integral for the small
$T$ expansion up to some small value, $a$, and adding this to the remaining
part of the integral~\cite{MoyaertsLaurent:2004}:
\begin{equation}
\int_0^{\infty} I(T) dT = \underbrace{\int_0^a I(T) dT}_\mathrm{ small ~T}
+ \underbrace{\int_a^\infty I(T)dT}_\mathrm{ worldline ~numerics}.
\end{equation}
Because this normalization procedure uses the constant field expressions for small values of
$T$, this scheme introduces a small systematic uncertainty. To improve on the
method outlined here, the derivatives of the background field can
be accounted for by using the analytic forms of the heat kernel expansion to perform the
renormalization~\cite{Gies:2001tj}.
\subsection{Uncertainty analysis in worldline numerics}
\label{ch:WLError}
So far in the worldline numerics literature, the discussions of uncertainty
analysis have been unfortunately brief. It has been suggested that the
standard deviation of the worldlines provides a good measure of the
statistical error in the worldline method~\cite{Gies:2001zp,
Gies:2001tj}. However, the distributions produced by the worldline
ensemble are highly non-Gaussian (see figure \ref{fig:hists}),
and therefore the standard error in
the mean is not a good measure of the uncertainties
involved. Furthermore, the use of the same worldline ensemble to
compute the Wilson loop multiple times in an integral results in
strongly correlated uncertainties. Thus, propagating uncertainties
through integrals can be computationally expensive due to the
complexity of computing correlation coefficients.
The error bars on worldline calculations impact the conclusions that
can be drawn from calculations, and also have important implications
for the fermion problem, which limits the domain of applicability of
the technique (see section \ref{sec:fermionproblem}). It is therefore
important that the error analysis is done thoughtfully and
transparently. The purpose of this section is to contribute a more
thorough discussion of uncertainty analysis in the worldline numerics technique
to the literature in hopes of avoiding any confusion associated with
the above-mentioned subtleties.
There are two sources of uncertainty in the worldline technique: the
discretization error in treating each continuous worldline as a set of
discrete points, and the statistical error of sampling a finite number
of possible worldlines from a distribution. In this section, we
discuss each of these sources of uncertainty.
\subsubsection{Estimating the Discretization Uncertainties}
\label{sec:discunc}
The discretization error arising from the integral over $\tau$ in the
exponent of each Wilson loop (see equation (\ref{eqn:wilsonloop})) is
difficult to estimate since any number of loops could be represented
by each choice of discrete points. The general strategy is to make
this estimation by computing the Wilson loop using several different
numbers of points per worldline and observing the convergence
properties.
The specific procedure adopted for this work involves dividing each
discrete worldline into several worldlines with varying levels of
discretization. Since we are using the d-loop method for generating
the worldlines (section \ref{sec:loopgen}), a
$\frac{N_\mathrm{ppl}}{2}$ sub-loop consisting of every other point
will be guaranteed to contain the prescribed distribution of
velocities.
To look at the convergence for the loop discretization,
each worldline is divided into three groups. One group of $\frac{N_\mathrm{ppl}}{2}$ points, and two groups of
$\frac{N_\mathrm{ppl}}{4}$.
This permits us to compute the average holonomy factors at three levels of
discretization:
\begin{equation}
\mean{W}_{N_\mathrm{ppl}/4} = \mean{e^{\frac{i}{2}\triangle} e^{\frac{i}{2}\Box}},
\end{equation}
\begin{equation}
\mean{W}_{N_\mathrm{ppl}/2} = \mean{e^{i\circ}},
\end{equation}
and
\begin{equation}
\mean{W}_{N_\mathrm{ppl}} = \mean{e^{\frac{i}{2}\circ} e^{\frac{i}{4}\Box} e^{\frac{i}{4}\triangle}},
\end{equation}
where the symbols $\circ$, $\Box$, and $\triangle$ denote the worldline integral,
$\int_0^T d\tau A(x_{CM}+x(\tau))\cdot \dot x$, computed using the sub-worldlines
depicted in figure \ref{fig:Division}.
\begin{figure}
\centering
\includegraphics[width=\linewidth]{images/DiscretizationDivision}
\caption[Illustration of convergence testing scheme]
{Diagram illustrating the division of a worldline into three smaller interleaved worldlines}
\label{fig:Division}
\end{figure}
We may put these factors into the equation of a parabola to extrapolate the result to an infinite
number of points per line (see figure \ref{fig:DiscErr}):
\begin{equation}
\mean{W}_{\infty} \approx \frac{8}{3}\mean{W}_{N_\mathrm{ppl}} - 2 \mean{W}_{N_\mathrm{ppl}/2} +\frac{1}{3}\mean{W}_{N_\mathrm{ppl}/4}.
\end{equation}
So, we estimate the discretization uncertainty to be
\begin{equation}
\delta \mean{W}_{\infty} \approx |\mean{W}_{N_\mathrm{ppl}} - \mean{W}_{\infty}|.
\end{equation}
Generally, the statistical uncertainties are the limitation in the precision of the
worldline numerics technique. Therefore, $N_\mathrm{ppl}$ should be chosen to be
large enough that the discretization uncertainties are small relative to the
statistical uncertainties.
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=1.8cm 0.3cm 1.8cm 1cm]{images/DiscErrPlot}
\caption[Illustration of extrapolation to infinite points per loop]
{This plot illustrates the method used to extrapolate the Wilson loop to
infinite points per loop and the uncertainty estimate in the approximation.}
\label{fig:DiscErr}
\end{figure}
\subsubsection{Estimating the Statistical Uncertainties}
We can gain a great deal of insight into the nature of the statistical uncertainties
by examining the specific case of the uniform magnetic field since we know the
exact solution in this case. Sections~\ref{sec:notnormal},~\ref{sec:correlations},
and~\ref{sec:groupingworldlines} discuss the peculiarities of the
statisical uncertainties in the worldline numerics method for the uniform magnetic
field.
\subsubsection{The Worldline Ensemble Distribution is not Normal}
\label{sec:notnormal}
A reasonable first instinct for estimating the error bars is to use the standard
error in the mean of the collection of individual worldlines:
\begin{equation}
\mathrm{ SEM}( W ) = \sqrt{\sum_{i=1}^{N_l}\frac{(W_i - \mean{W})^2}{N_l(N_l-1)}}.
\end{equation}
This approach has been promoted in early papers on worldline numerics~\cite{Gies:2001zp, Gies:2001tj}.
In figure \ref{fig:resids}, we have plotted the residuals and the corresponding
error bars for several values of the proper time parameter, $T$ in black. From this plot,
it appears that the error bars are quite large in the sense that we appear to
produce residuals which are considerably smaller than would be implied by the
sizes of the error bars. This suggests that we have overestimated the size of
the uncertainty.
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.8cm 0.3cm 1.8cm 1cm]{images/residsboth}
\caption[Comparison of error bars between standard error in
the mean and jackknife analysis] {The residuals of the
Wilson loops for a constant magnetic field showing the
standard error in the mean (thin error bars) and the
uncertainty in determining the mean (thick blue error
bars). For reasons discussed in this section, the standard
error in the mean overestimates the uncertainties involved
by more than a factor of 3 at each value of $T$.}
\label{fig:resids2} \label{fig:resids}
\end{figure}
We can see why this is the case by looking more closely at the
distributions produced by the worldline technique. An exact
expression for these distributions can be derived in the case of the
constant magnetic field~\cite{MoyaertsLaurent:2004}:
\begin{eqnarray}
\label{eqn:exactdist}
w(y)&=&\frac{W_\mathrm{ exact}}{\sqrt{1-y^2}}\sum_{n=-\infty}^{\infty}\biggl[f(\arccos(y)+2n\pi)+\nonumber \\
& & ~~~~~~~~~~~~ f(-\arccos(y)+2n\pi)\biggr]
\end{eqnarray}
with
\begin{equation}
f(\phi)=\frac{\pi}{4BT\cosh^2(\frac{\pi \phi}{2BT})}.
\end{equation}
Figure \ref{fig:hists} shows histograms of the worldline results along with the
expected distributions. These distributions highlight a significant hurdle in
assigning error bars to the results of worldline numerics.
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.8cm 0.3cm 1.8cm 1cm]{images/hist}
\caption[Histograms showing the worldline distributions]
{Histograms showing the worldline distributions of the residuals
for three values of $T$ in the constant magnetic field case.
Here, we are neglecting the fermion factor. The dark
line represents the exact distribution computed using equation
\ref{eqn:exactdist}. The worldline means are indicated with an arrow,
while the exact mean in each case is 0. There are 5120 worldlines in
each histogram. The vertical axes are normalized to a total area of unity.}
\label{fig:hists}
\end{figure}
Due to their highly non-Gaussian nature, the standard error in the mean is not a good
characterization of the distributions that are produced. We should not interpret each individual
worldline as an independent measurement of the mean value of these distributions; for large values of $BT$,
almost all of our worldlines will produce answers which are far away from the mean of the
distribution. This means that the variance of the distribution will be very large, even
though our ability to determine the mean of the distribution is relatively precise
because of the increasing symmetry about the mean as $T$ becomes large.
\subsubsection{Correlations between Wilson Loops}
\label{sec:correlations}
Typically, numerical integration is performed by replacing the integral with a sum over a finite set
of points from the integrand. We will begin the present discussion by considering the uncertainty
in adding together two points (labelled $i$ and $j$) in our integral over $T$. Two terms of the
sum representing the numerical integral will involve a function of $T$ times the two
Wilson loop factors,
\begin{equation}
I = g(T_i)\mean{W(T_i)} + g(T_j)\mean{W(T_j)}
\end{equation}
with an uncertainty given by
\begin{eqnarray}
\delta I &=& \left | \pderiv{I}{\mean{W(T_i)}} \right|^2 (\delta \mean{W(T_i)})^2 \\ \nonumber
& &
~~ + \left | \pderiv{I}{\mean{W(T_j)}} \right|^2 (\delta
\mean{W(T_j)})^2 \\ \nonumber
& & ~~ + 2 \left | \pderiv{I}{\mean{W(T_i)}}
\pderiv{I}{\mean{W(T_j)}} \right | \times \\ \nonumber
& & ~~\rho_{ij}
(\delta \mean{W(T_i)}) (\delta \mean{W(T_j)}) \\
&=& g(T_i)^2 (\delta \mean{W(T_i)})^2 + g(T_j)^2 (\delta \mean{W(T_j)})^2+ \nonumber \\
& & ~ 2 \left | g(T_i)g(T_j) \right | \rho_{ij} (\delta \mean{W(T_i)}) (\delta \mean{W(T_j)})
\end{eqnarray}
and the correlation coefficient $\rho_{ij}$ given by
\begin{equation}
\label{eqn:corrcoef}
\rho_{ij} = \frac{\mean{(W(T_i) - \mean{W(T_i)})(W(T_j)-\mean{W(T_j)})}}{\sqrt{(W(T_i)
-\mean{W(T_i)})^2}\sqrt{(W(T_j)-\mean{W(T_j)})^2}}.
\end{equation}
The final term in the error propagation equation takes into account correlations between the
random variables $W(T_i)$ and $W(T_j)$. Often in a Monte Carlo computation, one can
treat each evaluation of the integrand as independent, and neglect the uncertainty
term involving the correlation coefficient. However, in worldline numerics,
the evaluations are related because the same worldline ensemble is reused
for each evaluation of the integrand.
The correlations are significant (see figure \ref{fig:corr}), and this term
can't be neglected. Computing each correlation coefficient takes
a time proportional to the square of the number of worldlines. Therefore, it may
be computationally expensive to formally propagate uncertainties through an
integral.
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.8cm 0.3cm 1.8cm 1cm]{images/corrrand}
\caption[Correlation coefficients between different evaluations of the integrand]
{Correlation coefficients, equation (\ref{eqn:corrcoef}), between $\mean{W(T)}$
and $\mean{W(T=3)}$ computed using individual worldlines, groups of worldlines, and
shuffled groups of worldlines.}
\label{fig:corr}
\end{figure}
The point-to-point correlations were originally pointed out by Gies
and Langfeld who addressed the problem by updating (but not replacing
or regenerating) the loop ensemble in between each evaluation of the
Wilson loop average~\cite{Gies:2001zp}. This may be a good way of
addressing the problem. However, in the following section, we promote a
method which can bypass the difficulties presented by the correlations
by treating the worldlines as a collection of worldline groups.
\subsubsection{Grouping Worldlines}
\label{sec:groupingworldlines}
Both of the problems explained in the previous two subsections can be overcome
by creating groups of worldline loops within the ensemble. Each group of worldlines
then makes a statistically independent measurement of the Wilson loop average
for that group. The statistics between the groups of measurements are normally
distributed, and so the uncertainty is the standard error in the mean of the
ensemble of groups (in contrast to the ensemble of worldlines).
For example, if we divide the $N_l$ worldlines into $N_G$ groups of $N_l/N_G$
worldlines each, we can compute a mean for each group:
\begin{equation}
\mean{W}_{G_j} = \frac{N_G}{N_l}\sum_{i=1}^{N_l/N_G}W_i.
\end{equation}
Provided each group contains the same number of worldlines,
the average of the Wilson loop is unaffected by this grouping:
\begin{eqnarray}
\mean{W} & = & \frac{1}{N_G} \sum_{j=1}^{N_G} \mean{W}_{G_j} \\
& = & \frac{1}{N_l} \sum_{i=1}^{N_l} W_i.
\end{eqnarray}
However, the uncertainty is the standard error in the mean of
the groups,
\begin{equation}
\delta \mean{W} = \sqrt{\sum_{i=1}^{N_G}
\frac{(\mean{W}_{G_i} - \mean{W})^2}{N_G(N_G-1)}}.
\end{equation}
Because the worldlines are unrelated to one another, the choice of how to
group them to compute a particular Wilson loop average is arbitrary. For example,
the simplest choice is to group the loops by the order they were generated, so that
a particular group number, $i$, contains worldlines $iN_l/N_G$ through $(i+1)N_l/N_G -1$.
Of course, if the same worldline groupings are used to compute different Wilson
loop averages, they will still be correlated. We will discuss this problem in a moment.
The basic claim of the worldline technique is that the mean of the
worldline distribution approximates the holonomy factor. However, from
the distributions in figure \ref{fig:hists}, we can see that the
individual worldlines themselves do not approximate the holonomy
factor. So, we should not think of an individual worldline as an
estimator of the mean of the distribution. Thus, a resampling
technique is required to determine the precision of our statistics. We
can think of each group of worldlines as making an independent
measurement of the mean of a distribution. As expected, the groups of
worldlines produce a more Gaussian-like distribution (see figure
\ref{fig:uncinmean}), and so the standard error of the groups is a
sensible measure of the uncertainty in the Wilson loop value.
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.8cm 0.3cm 1.8cm 1cm]{images/uncinmean}
\caption[Histogram for reproducing measurements with groups of worldlines]
{The histogram demonstrating the precision with which we can reproduce
measurements of the mean using different groups of 100 worldlines at $BT=6.0$.
In this case, the distribution is Gaussian-like and meaningful error bars can
be placed on our measurement of the mean.}
\label{fig:uncinmean}
\end{figure}
We find that the error bars are about one-third as large as those
determined from the standard error in the mean of the individual
worldlines, and the smaller error bars better characterize the size of
the residuals in the constant field case (see figure
\ref{fig:resids2}). The strategy of using subsets of the available
data to determine error bars is called jackknifing. Several previous
papers on worldline numerics have mentioned using jackknife analysis to
determine the uncertainties, but without an explanation of the
motivations or the procedure employed \cite{2005PhRvD..72f5001G,
PhysRevLett.96.220401, Dunne:2009zz, PhysRevD.84.065035}.
The grouping of worldlines alone does not address the problem of
correlations between different evaluations of the integrands. Figure
\ref{fig:corr} shows that the uncertainties for groups of worldlines
are also correlated between different points of the integrand.
However, the worldline grouping does provide a tool for bypassing the
problem. One possible strategy is to randomize how worldlines are
assigned to groups between each evaluation of the integrand. This
produces a considerable reduction in the correlations, as is shown in
figure \ref{fig:corr}. Then, errors can be propagated through the
integrals by neglecting the correlation terms. Another strategy is to
separately compute the integrals for each group of worldlines, and
then consider the statistics of the final product to determine the
error bars. This second strategy is the one adopted for the work
presented in this paper. Grouping in this way reduces the amount of
data which must be propagated through the integrals by a factor of the
group size compared to a delete-1 jackknife scheme, for example. In
general, the error bars quoted in the remainder of this paper are obtained by
computing the standard error in the mean of groups of worldlines.
\subsubsection{Uncertainties and the Fermion Problem}
\label{sec:fermionproblem}
The fermion problem of worldline numerics is a name given to an enhancement of
the uncertainties at large $T$~\cite{Gies:2001zp,
MoyaertsLaurent:2004}. It should not be confused with the
fermion-doubling problem associated with lattice methods. In a
constant magnetic field, the scalar portion of the calculation
produces a factor of $\frac{BT}{\sinh{(BT)}}$, while the fermion
portion of the calculation produces an additional factor
$\cosh{(BT)}$. Physically, this contribution arises as a result of the
energy required to transport the electron's magnetic moment around the
worldline loop. At large values of $T$, we require subtle
cancellation between huge values produced by the fermion portion with
tiny values produced by the scalar portion. However, for large $T$,
the scalar portion acquires large relative uncertainties which make
the computation of large $T$ contributions to the integral very
imprecise.
This can be easily understood by examining the worldline distributions shown in figure
\ref{fig:hists}. Recall that the scalar Wilson loop average for these histograms is given
by the flux in the loop, $\Phi_B$:
\begin{equation}
\mean{W} = \left<\exp{\left(ie\int_0^Td\tau \vec{A}(\vec{x}_\mathrm{ cm}
+ \vec{x}(\tau))\cdot d\vec{x}(\tau)\right)}\right> = \left< e^{ie\Phi_B}\right>.
\end{equation}
For constant fields, the flux through the worldline loops obeys the distribution function
\cite{MoyaertsLaurent:2004}
\begin{equation}
f(\Phi_B) = \frac{\pi}{4BT\cosh^2\left(\frac{\pi \Phi_B}{2BT}\right)}.
\end{equation}
For small values of $T$, the worldline loops are small and the
amount of flux through the loop is correspondingly small. Therefore, the
flux for small loops is narrowly distributed about $\Phi_B = 0$. Since zero
maximizes the Wilson loop ($e^{i0}=1$),
this explains the enhancement to the right of the distribution for small values of $T$.
As $T$ is increased, the flux through any given worldline becomes very large and the
distribution of the flux becomes very broad.
For very large $T$, the width of the distribution is many
factors of $2\pi/e$. Then, the phase ($e \Phi_B\mod{2\pi}$) is nearly
uniformly distributed, and the
Wilson loop distribution reproduces the Chebyshev distribution ({\em i.e.}\
the distribution obtained from projecting uniformly distributed points on
the unit circle onto the horizontal axis),
\begin{equation}
\lim_{T\to\infty}w(y) = \frac{1}{\pi\sqrt{1-y^2}}.
\end{equation}
The mean of the Chebyshev distribution is zero due to its symmetry.
However, this symmetry is not
realized precisely unless we use a very large number of loops. Since the
width of the distribution is already 100$\times$ the value of the mean at
$T=6$, any numerical asymmetries in the distribution result in very large
relative uncertainties of the scalar portion. Because of these uncertainties,
the large contribution from the fermion factor are not cancelled precisely.
This problem makes it very difficult to compute
the fermionic effective action unless the fields are well localized
\cite{MoyaertsLaurent:2004}. For example, the fermionic factor for
non-homogeneous magnetic fields oriented along the z-direction is
\begin{equation}
\cosh{\left(e\int_0^T d\tau B(x(\tau))\right)}.
\end{equation}
For a homogeneous field, this function grows exponentially with $T$ and
is cancelled by the exponentially vanishing scalar Wilson loop.
For a localized field,
the worldline loops are very large for large values of $T$, and they primarily
explore regions far from the field. Thus, the fermionic factor grows more slowly
in localized fields, and is more easily cancelled by the rapidly vanishing scalar part.
\begin{figure}[ht]
\begin{center}
\includegraphics[width=\linewidth,clip,trim=0.8cm 0.3cm 1.8cm 1cm]{images/Integrand}
\caption[small $T$ behaviour of worldline numerics]
{The small $T$
behaviour of worldline numerics. The data points represent the numerical
results, where the error bars are determined from the jackknife analysis described
in chapter \ref{ch:WLError}.
The solid line represents the exact solution while the dotted line represents
the small $T$ expansion of the exact solution. Note the amplification of
the uncertainties.}
\end{center}
\end{figure}
\subsection{Computing an Effective Action}
The ensemble average in the effective action is simply the sum over the
contributions from each worldline loop, divided by the number of loops in
the ensemble. Since the computation of each loop is independent of the
other loops, the ensemble average may be straightforwardly parallelized by
generating separate processes to compute the contribution from each loop.
For this parallelization, four Nvidia Tesla C1060 \acp{GPU} were used through
the \ac{CUDA} parallel programming framework. Because \acp{GPU} can spawn
thousands of parallel processing threads
with much less computational overhead
than an \ac{MPI} cluster, they excel at handling a very large number of parallel
threads, although the clock speed is slower and fewer memory resources are typically available.
The \ac{GPU} architecture has
recently been used by another group for computing Casimir forces using
worldline numerics~\cite{2011arXiv1110.5936A}. More detailed information about the
technical implementation of these calculations on the \ac{GPU} architecture,
including a listing of the source code, can be
found in~\cite{2012PhDT........21M}.
Once the ensemble average of the Wilson loop has been computed,
computing the effective action is a straightforward matter of
performing numerical integrals. The effective action density is
computed by performing the integration over proper time, $T$. Then,
the effective action is computed by performing a spacetime integral
over the loop ensemble center of mass. In all cases where a numerical
integral was performed, Simpson's method was
used~\cite{burden2001numerical}. Integrals from 0 to $\infty$ were
mapped to the interval $[0,1]$ using substitutions of the form $x =
\frac{1}{1+T/T_\mathrm{ max}}$, where $T_\mathrm{ max}$ sets the scale
for the peak of the integrand. In the constant field case, for the
integral over proper time, we expect $T_\mathrm{ max} \sim 3/(eB)$ for
large fields and $T_\mathrm{ max} \sim 1$ for fields of a few times
critical or smaller. In section~\ref{ch:WLError}, we presented a
detailed discussion of how the statistical and discretization
uncertainties can be computed in this technique.
\subsection{Verification and Validation}
\label{sec:verify}
The worldline numerics software can be validated and verified by making sure that it
produces the correct results where the derivative expansion is a good approximation,
and that the results are consistent with previous numerical calculations of flux tube
effective actions. For this reason, the validation was done primarily with flux tubes with
a profile defined by the function
\begin{equation}
f_\lambda(\rho^2) = \frac{\rho^2}{(\lambda^2 + \rho^2)}.
\end{equation}
For large values of $\lambda$, this function varies slowly on the Compton wavelength
scale, and so the derivative expansion is a good approximation. Also, flux tubes
with this profile were studied previously using worldline numerics
\cite{Moyaerts:2003ts, MoyaertsLaurent:2004}.
Among the results presented in~\cite{MoyaertsLaurent:2004} is a comparison of
the derivative expansion and worldline numerics for this magnetic field
configuration. The result is that the next-to-leading-order term
in the derivative expansion is only a small correction to the the
leading-order term for $\lambda \gg \lambda_e$, where the derivative
expansion is a good approximation. The derivative expansion breaks
down before it reaches its formal validity limits
at $\lambda \sim \lambda_e$. For this reason,
we will simply focus on the leading order derivative expansion, which we call
the locally-constant-field (\ac{LCF}) approximation.
The effective action of \ac{QED} in the \ac{LCF} approximation is given
in cylindrical symmetry by
\begin{eqnarray}
\label{eqn:LCFferm}
\Gamma^{(1)}_\mathrm{ ferm} &=& \frac{1}{4\pi}\int_0^\infty dT
\int_0^\infty \rho_\mathrm{ cm} d\rho_\mathrm{ cm}\frac{e^{-m^2T}}{T^3} \times \nonumber \\
& &\biggl\{eB(\rho_\mathrm{ cm})T\coth{(eB(\rho_\mathrm{ cm})T)} \\
& &
- 1 -\frac{1}{3}(eB(\rho_\mathrm{ cm})T)^2\biggr\}. \nonumber
\end{eqnarray}
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.8cm 0.3cm 1.2cm 1cm]{images/igrandvsT4_F10l21}
\caption[Comparison with derivative expansion for $T$ integrand]
{The integrand of the proper time, $T$, integral for three different values
of the radial coordinate, $\rho$ for a $\lambda = 1$ flux tube. The solid lines
represent the zeroth-order derivative expansion, which, as expected, is a good approximation
until $\rho$ becomes too small.}
\label{fig:igrandvsT}
\end{figure}
Figure \ref{fig:igrandvsT} shows a comparison between the proper time integrand,
\begin{equation}
\frac{e^{-m^2T}}{T^3}\left[\langle W\rangle_{\vec{r}_\mathrm{ cm}}
- 1 -\frac{1}{3}(eB_\mathrm{ cm}T)^2\right],
\end{equation}
and the \ac{LCF} approximation result for a flux tube with $\lambda = \lambda_e$ and
$\mathcal{F} = 10$. In this case, the \ac{LCF} approximation is only appropriate far from the
center of the flux tube, where the field is not changing very rapidly. In the figure, we can
begin to see the deviation from this approximation, which gets more pronounced closer to the
center of the flux tube (smaller values of $\rho$).
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.2cm 0.3cm 1.8cm 1cm]{images/vsxcm_ferm_smooth}
\caption[Comparison with \acs{LCF} approximation for action density]
{The fermion term of the effective action density as a function
of the radial coordinate for a flux tube with width $\lambda = 10 \lambda_e$. }
\label{fig:vsrhocm}
\end{figure}
The effective action density for a slowly varying flux tube is plotted in figure
\ref{fig:vsrhocm} along with the \ac{LCF} approximation. In this case,
the \ac{LCF} approximation agrees within the statistics of the worldline numerics.
\section{Astrophysical Background}
\label{sec:background}
\subsection{Nuclear Superconductivity in Neutron Stars}
\label{subsec:superconductivity}
In the dense nuclear matter of a neutron star, it may be possible to
have neutron superfluidity, proton superconductivity, and even quark
colour superconductivity~\cite{2006pfsb.book..135S, schmitt2010dense}.
The first prediction of neutron-star superfluidity dates back to
Migdal in 1959~\cite{1959NucPh..13..655M}. The arguments which make
superfluidity seem likely are based on the temperatures of neutron
stars. A short time after their creation, neutron stars are very cold
compared to nuclear energy scales. The temperature in the interior may
be a few hundred keV. Studies of nuclear matter show that the
transition temperature is $T_c \gtrsim 500$
keV~\cite{1989tns..conf..457S}. So, it is expected that the nuclear
matter in a neutron star forms condensates of Cooper pairs.
Because of this, some fraction of neutrons in the inner crust
of a neutron star are expected to be superfluid. These neutrons make up
about a percent of the moment of inertia of the star and are
weakly coupled to the nuclear crystal lattice which makes up the
remainder of the inner crust. If the neutron vortices in the
superfluid component move at nearly the same speed as the nuclear
lattice, the vortices can become pinned to the lattice so that the
superfluid shares an angular velocity with the crust.
This pinning between the fluid and solid crust has observable impacts
on the rotational dynamics of the neutron star, for the same reasons that
hard-boiling an egg (pinning the yolk to the shell) produces an
observable difference in the way it spins.
This picture of a neutron superfluid co-rotating with a solid crust
has been used to interpret several types of pulsar timing anomalies.
Pulsars are nearly perfect clocks, although they gradually spin-down
as they radiate energy. Occasionally, though, pulsars demonstrate
deviations from their expected regularity. A glitch is an abrupt
increase in the rotation and spin-down rate of a pulsar, followed
by a slow relaxation to pre-glitch values over weeks or years. This
behaviour is consistent with the neutron superfluid suddenly
becoming unpinned from the crust and then dynamically relaxing
due to its weak coupling to the crust until it is pinned once
again~\cite{1975Natur.256...25A}.
The neutron star in Cassiopeia A has been observed to be rapidly cooling
\cite{2010ApJ...719L.167H}. The surface temperature has decreased by about
4\% over 10 years. This observation is also strong evidence
of superfluidity and superconductivity in neutron stars
\cite{2011PhRvL.106h1101P, 2011MNRAS.412L.108S}. The observed cooling
is too fast to be explained by the observed x-ray emissions
and standard neutrino cooling. However,
the cooling is readily explained by the emission of neutrinos during
the formation of neutron Cooper pairs. Based on such a model,
the superfluid transition temperature of neutron star matter is
$\sim 10^{9}$ K or $\sim 90$ keV.
Further hints regarding superfluidity in neutron stars come from
long-term periodic variability in pulsar timing data. For example,
variabilities in PSR B1828-11 were initially interpreted as free
precession (or wobble) of the star~\cite{2000Natur.406..484S}. If
neutron stars can precess, observations could strongly constrain the
ratio of the moments of inertia of the crust and the superfluid
neutrons. Moreover, the existence of flux tubes ({\em i.e.}\ type-II
superconductivity) in the crust is generally incompatible with the
slow, large amplitude precession suggested by PSR
B1828-11~\cite{2000Natur.406..484S}. The neutron vortices would have
to pass through the flux tubes, which should cause a huge dissipation
of energy and a dampening of the precession which is not
observed~\cite{PhysRevLett.91.101101}. However, recent arguments
suggest that the timing variability data is not well explained by free
precession and that it more likely suggests that the star is switching
between two magnetospheric states~\cite{2010Sci...329..408L}.
Nevertheless, other authors suggest not being premature in throwing
out the precession hypothesis without further observations
\cite{2012MNRAS.420.2325J}.
\subsection{Magnetic Flux Tubes in Neutron Stars}
If a magnetic field is able to penetrate the proton superfluid on a
microscopic level, it must do so by forming a triangular Abrikosov
lattice with a single quanta of flux in each flux tube. So, the
density of flux tubes is given simply by the average field strength.
If the distance between flux tubes is $l_f$, the flux in a circular
region within $l_f/2$ of a flux tube is given by
\begin{equation}
F = \frac{2\pi \mathcal{F}}{e} = 2\pi \int_0^{l_f/2} B \rho d\rho
\end{equation}
where we have introduced a dimensionless measure of flux $\mathcal{F}
= \frac{e}{2\pi}F$. So, the distance between flux tubes is
\begin{equation}
l_f = \sqrt{\frac{8 \mathcal{F}}{eB}}.
\end{equation}
If the magnetic field is the quantum critical field strength,
$B_k = \frac{m^2}{e} = 4.4 \times 10^{13}$ Gauss, then
the flux tubes are separated by a few Compton electron wavelengths.
This is particularly interesting since this is the distance scale
associated with non-locality in \ac{QED}.
The size of a flux tube profile in laboratory superconductors is on
the nanometer or micron
scale~\cite{poole2007superconductivity}. Because the flux is fixed,
the size of the tube profile determines the strength of the magnetic
field within the tube. For laboratory superconductors the field
strengths are small compared to the quantum critical field, and the
field is slowly varying on the scale of the Compton wavelength. In
this case, the quantum corrections to the free energy are known to be
much smaller than the classical contribution (see section
\ref{sec:EAisoflux}). The size scale for the flux tubes in a
superconductor is determined by the London penetration depth. In a
neutron star, this quantity is estimated to be a small fraction of a
Compton wavelength, much smaller than in laboratory superconductors
\cite{lrr-2008-10, PhysRevLett.91.101101}. In this case, the magnetic
field strength at the centre of the tube exceeds the quantum critical
field strength and the field varies rapidly, rendering the derivative
expansion description of the effective action unreliable.
The Ginzburg-Landau parameter, equation is the
ratio of the proton coherence length, $\xi_p \sim 30$ fm, and the
London penetration depth of a proton superconductor, $\lambda_p \sim 80$~fm
\cite{PhysRevLett.91.101101}.
\begin{equation}
\kappa = \frac{\lambda_p}{\xi_p} \sim 2
\end{equation}
where $\kappa>1/\sqrt{2}$ signals Type-II behavior.
We therefore expect that the proton Cooper pairs most likely form a
type-II superconductor~\cite{2006pfsb.book..135S}. However, it is
possible that physics beyond what is taken into account in the
standard picture affects the free energy of a magnetic flux tube. In
that case, the interaction between two flux tubes may indeed be
attractive in which case the neutron star would be a type-I
superconductor.
\subsection{\texorpdfstring{\acs{QED}}{QED} Effective Actions of Flux Tubes}
\label{sec:EAisoflux}
Vortices of magnetic flux have very important impacts on the quantum
mechanics of electrons. In particular, the phase of the electron's
wavefunction is not unique in such a magnetic field. This is
demonstrated by the Aharonov-Bohm
effect~\cite{0370-1301-62-1-303,PhysRev.115.485}. The first
calculations of the fermion effective energies of these configurations
were for infinitely thin Aharonov-Bohm flux
strings~\cite{Gornicki1990271, 1998MPLA...13..379S}. Calculations for
thin strings were also performed for cosmic string
configurations~\cite{0264-9381-12-5-013}. For these infinitely-thin
string magnetic fields, the energy density is singular for small
radii. So, it is not possible to define a total energy per unit
length. Another approach was to compute the effective action for a
finite-radius flux tube where the magnetic flux was confined entirely
to the surface of the tube~\cite{PhysRevD.51.810}. This approach
results in infinite classical energy densities as well.
Physical flux tube configurations would have a finite radius. The
earliest paper to deal with finite radius flux tubes in QED considered
the effective action of a step-function profiled flux tube using the
Jost function of the related scattering
problem~\cite{1999PhRvD..60j5019B}. One of the conclusions from this
research was that the quantum correction to the classical energy was
relatively small for any value of the flux tube size, for the entire
range of applicability of \ac{QED}. The techniques from this study
were soon generalized to other field profiles including, a
delta-function cylindrical shell magnetic
field~\cite{PhysRevD.62.085024}, and more realistic flux tube
configurations such as the Gaussian~\cite{2001PhRvD..64j5011P} and the
Nielsen-Olesen vortex~\cite{2003PhRvD..68f5026B}. Flux tube vacuum
energies were also analyzed extensively using a spectral method
\cite{2005NuPhB.707..233G, 2006JPhA...39.6799W, Weigel:2010pf}.
The effective actions of flux tubes have been previously analyzed
using worldline numerics~\cite{Langfeld:2002vy}. This research
investigated isolated flux tubes, but also made use of the loop cloud
method's applicability to situations of low symmetry to investigate
pairs of interacting vortices. One conclusion from that investigation
was that the fermionic effects resulted in an attractive force between
vortices with parallel orientations, and a repulsive force between
vortices with anti-parallel orientations. Due to the similarity in
scope and technique, the latter mentioned research is the closest to
the research presented in this paper.
\section{Calculations}
\label{sec:calculations}
In this section, we will further explore the nature of this phenomenon
in \ac{QED} using a highly parallel implementation of the worldline numerics
technique implemented on a heterogeneous \ac{CPU} and \ac{GPU} architecture.
Specifically, we
explore cylindrically symmetric magnetic field profiles for isolated
flux tubes and periodic profiles designed to model properties of a
triangular lattice. For these calculations, the classical magnetic field
configurations are a chosen input to the algorithms and the physical processes that
may have created the field configurations do not factor in to the calculations.
The worldline numerics algorithm cannot be straight-forwardly applied to spinor
\ac{QED} calculations in our model lattice because of the well-known
fermion problem of worldline numerics (see section \ref{sec:fermionproblem})
~\cite{Gies:2001zp, MoyaertsLaurent:2004}.
However, the problem does not affect the scalar QED (\ac{ScQED})
calculations. Therefore, we explore the quantum-corrected
energies of isolated flux tubes for both scalar and spinor electrons
and use this comparison to speculate about the relationship of our
cylindrical lattice model and the spinor \ac{QED} energies of an
Abrikosov lattice of flux tubes that may be found in neutron stars.
\subsection{Isolated Flux Tubes}
As discussed in section~\ref{sec:cylindrical} we will focus on fields
with a cylindrical symmetry. In particular to explore isolated
magnetic flux tubes, we consider the following profile function (as we
used earlier in section~\ref{sec:verify}):
\begin{equation}
f_\lambda(\rho^2) =
\frac{\rho^2}{\rho^2 + \lambda^2}.
\end{equation}
This gives a magnetic field with a profile
\begin{equation}
\label{eqn:isoB}
B_z(\rho^2) = \frac{2\mathcal{F}}{e}\frac{\lambda^2}{(\rho^2+\lambda^2)^2}.
\end{equation}
This profile is a smooth flux tube representation that can be evaluated quickly.
Moreover, flux tubes with this profile were studied previously using worldline numerics
\cite{Moyaerts:2003ts, MoyaertsLaurent:2004}.
\subsection{Cylindrical Model of a Flux Tube Lattice}
\begin{figure}
\centering
\includegraphics[width=\linewidth]{images/latticesymmetry}
\caption[Cylindrical model of a hexagonal lattice]
{In a type-II superconductor, there are neighbouring flux
tubes arranged in a hexagonal Abrikosov lattice which have a non-local
impact on the effective action of the central flux tube (left). In our model,
we account for the contributions from these neighbouring flux tubes in
cylindrical symmetry by including surrounding rings of flux (right).}
\label{fig:latticesymmetry}
\end{figure}
In a neutron star, we do not have isolated flux tubes. The tubes are
likely arranged in a dense lattice with the spacing between tubes on
the order of the Compton wavelength, with the size of a flux tube a
few percent of the Compton wavelength. Specifically, the maximum size
of a flux tube is on the order of the coherence length of the
superconductor, which for neutron stars has been estimated to be $\xi
\approx 30$ fm~\cite{PhysRevLett.91.101101}. This situation can be
directly computed in the worldline numerics technique. However, this requires us
to integrate over two spatial dimensions instead of one. Moreover, it
requires the use of more loops to more precisely probe the spatial
configurations of the magnetic field. Despite these problems, it is
very interesting to consider a dense flux tube lattice. Unlike the
isolated flux tube, the wide-tube limit of the configuration doesn't
have zero field, but an average, uniform background field. If this
background field is the size of the critical field, there are
interesting quantum effects even in the wide-tube limit.
In this section, we build a cylindrically symmetric toy model of a
hexagonal flux tube lattice. We focus on one central flux tube and
treat the surrounding six flux tubes as a continuous ring with six
units of flux at a distance $a$ from the central tube. The next ring
will contain twelve units of flux at a distance of $2a$, etc (see
figure \ref{fig:latticesymmetry}). Because of this condition, the
average strength of the field is fixed, and the field becomes uniform
in the wide tube limit instead of going to zero. For small values of
$\lambda$, we will have non-local contributions from the surrounding
rings in addition to the local contributions from the central flux
tube. This strategy will result in a simple model relative to the
sophisticated flux tube lattice models used in the
context of superconducting physics. However, the simplifications
are appropriate for a first study of the non-local QED interactions
between the regions of flux.
It is difficult to construct a model of this scenario if the flux
tubes bleed into one another as they are placed close together. For
example, with Gaussian flux tubes or flux tubes with the profile used
in the previous section, it is difficult to increase the width of the
flux tubes while accounting for the magnetic flux that bleeds out of
their regions. Moreover, it is difficult to integrate these schemes to
find the profile function $f_\lambda(\rho)$ which is needed to compute
the scalar part of the Wilson loop. In order to keep each tube as a
distinct entity which stays within its assigned region, we assign a
smooth function with compact support to represent each tube. This is
most easily done with the bump function, $\Psi(x)$, defined as
\begin{equation}
\Psi(x) =
\begin{cases}
e^{-1/(1-x^2)} & \mbox{ for } |x| < 1\\
0 & \mbox{ otherwise}
\end{cases}.
\end{equation}
This function can be viewed as a rescaled Gaussian.
We start by defining the magnetic field outside of the central flux
tube. Here, the magnetic field is a constant background field, with
the flux ring contributing a bump of width $\lambda$. The height of
the bump must go to zero as the width of the flux tube approaches the
distance between flux tubes, and should become infinite as the flux
tube width goes to zero:
\begin{equation}
B_z(\rho>\frac{a}{2}) = B_{\rm bg} + A\left(\frac{a-\lambda}{\lambda}\right)\left[\Psi(2(\rho-n a)/\lambda)-B\right].
\end{equation}
with $n\equiv \lfloor\frac{\rho+a/2}{a}\rfloor$.
If we require 6 units of flux in the first outer ring, 12 in the second,
and so on (see figure \ref{fig:latticesymmetry}), the size of the background field
is fixed to $B_{\rm bg} = \frac{6 \mathcal{F}}{ea^2}$. The total flux contribution due to the $\lambda$-dependent
terms must be zero:
\begin{equation}
\int_{(n-1/2)a}^{(n+1/2)a} \rho A\left(\frac{a-\lambda}{\lambda}\right)
\left[\Psi(2(\rho-n a)/\lambda)-B\right] d\rho = 0
\end{equation}
\begin{equation}
\frac{\lambda}{2}\int_{-1}^1 \left(\frac{\lambda}{2}x+na\right)\Psi(x)dx-Ba^2n = 0.
\end{equation}
This fixes the value of the constant $B$ to
\begin{equation}
B=\frac{q_1}{2}\frac{\lambda}{a}.
\end{equation}
The numerical constant $q_1$ is defined by
\begin{equation}
q_1 = \int_{-1}^{1}\Psi(x)dx \approx 0.443991.
\end{equation}
For a given bump amplitude, $A$, the magnetic field will become
negative if $\lambda$ becomes small enough. Therefore, we replace $A$
with its maximum value for which the field is positive if $\lambda >
\lambda_{\rm min}$ for some choice of minimum flux tube size:
\begin{equation}
A=\frac{12 \mathcal{F}}{e a q_1 (a-\lambda_{\rm min})}.
\end{equation}
The choice of $\lambda_{\rm min}$ sets the tube width at which the
field between the flux tubes vanishes. If $\lambda < \lambda_{\rm
min}$, the magnetic field between the flux tubes will point in the
$-\hat{\vec{z}}$-direction.
Because we are trying to fit a hexagonal peg into a round hole, we
must treat the central flux tube differently. For example, the
average field inside the central region for a unit of flux, is
different than the average field in the exterior region. Therefore,
even when $\lambda \rightarrow a$, the field cannot be quite uniform.
We consider the field in the central region to be a constant field
with a bump centered at $\rho = 0$:
\begin{equation}
B_z(\rho < \frac{a}{2}) = A_0 \Psi(2\rho/\lambda) + B_0.
\end{equation}
The constant term, $B_0$ is fixed by requiring continuity with the
exterior field at $\rho = a/2$:
\begin{equation}
B_0 = \frac{6\mathcal{F}}{e a^2}\left(1-\frac{a-\lambda}{a-\lambda_{\rm min}}\right).
\end{equation}
The bump amplitude, $A_0$, is determined by fixing the flux in the central region,
\begin{equation}
\int_0^{a/2}\rho\left[A_0\Psi(2\rho/\lambda) + B_0\right]d\rho = \frac{\mathcal{F}}{e}:
\end{equation}
\begin{equation}
A_0 \left(\frac{\lambda}{2}\right)^2\int_0^1 x \Psi(x)dx +
\frac{B_0}{2}\left(\frac{a}{2}\right)^2 = \frac{\mathcal{F}}{e}
\end{equation}
\begin{equation}
A_0 = \frac{4 \mathcal{F}}{\lambda^2 e q_2}\left(1-\frac{3}{4}
\left(1-\frac{a-\lambda}{a-\lambda_{\rm min}}\right)\right),
\end{equation}
where the numerical constant, $q_2$, is defined by
\begin{equation}
\label{eqn:q2}
q_2 \equiv \int_0^1 x \Psi(x) dx \approx 0.0742478.
\end{equation}
Finally, collecting together the important expressions,
the cylindrically symmetric flux tube lattice model is
\begin{eqnarray}
\label{eqn:ffbless}
B_z(\rho\le\frac{a}{2}) &=& \frac{4 \mathcal{F}}{\lambda^2 e q_2}
\left(1-\frac{3}{4}\left(\frac{\lambda-\lambda_{\rm min}}{a-\lambda_{\rm min}}\right)\right)\Psi(2\rho/\lambda) \nonumber \\
& & + \frac{6\mathcal{F}}{ea^2}\left(\frac{\lambda-\lambda_{\rm min}}{a-\lambda_{\rm min}}\right)
\label{eqn:Bzext} \\
B_z(\rho>\frac{a}{2}) &=& \frac{6 \mathcal{F}}{e a^2}\left(\frac{\lambda-\lambda_{\rm min}}{a-\lambda_{\rm min}}\right) + \nonumber \\
& & \frac{12 \mathcal{F}}{q_1ea\lambda}\left(\frac{a-\lambda}{a-\lambda_{\min}}\right)
\Psi(2(\rho-na)/\lambda) .
\end{eqnarray}
The magnetic field profile defined by these equations is shown in figures
\ref{fig:periodicB} and \ref{fig:B3d}. The current density required to
created fields with this profile is shown in figure \ref{fig:current}.
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.45in 0.1in 0.55in 0.3in]{images/periodicB}
\caption[Magnetic field in a cylindrical lattice model]
{The cylindrical lattice flux tube model for several
values of the width parameter $\lambda$. Here we have taken $a =\sqrt{8}\lambda_e$
and $\lambda_{\rm min} = 0.1a$. Note that in the limit $\lambda\rightarrow a$
the field is nearly uniform with a mound in the central region. This is a
consequence of the flux conditions in cylindrical symmetry requiring different
fields in the internal and external regions. The height of the $\lambda=0.1a$
flux tube extends beyond the height of the graph to about $61.9B_k$.
A 3D surface plot of the $\lambda=0.6a$ field profile is shown in figure
\ref{fig:B3d}.}
\label{fig:periodicB}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.45in 0.1in 0.55in 0.3in]{images/B3d}
\caption[Surface plot of magnetic field model]
{A 3D surface plot of the $\lambda = 0.6a$ (red in figure \ref{fig:periodicB}) magnetic field profile from
figure \ref{fig:periodicB}.}
\label{fig:B3d}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.4in 0.1in 0.55in 0.3in]{images/current}
\caption[Classical current required to create magnetic field model]
{The current densities required to create the $\lambda=0.6a$ (red) magnetic field profile.
The current is given by the curl of the magnetic field, $J_\psi(\rho) = -\frac{d B_z(\rho)}{d\rho}$.
The conversion to SI units is $1B_k/\lambda_e \approx 6\times 10^{38}{\rm A/m}^3$.}
\label{fig:current}
\end{figure}
\subsection{The Classical Action}
The classical action, $\Gamma^0$, is infinite for this configuration. To obtain
a finite result, we must look at the action per unit length in the z-direction,
per unit time, and per flux tube region ({\em i.e.}\ for $\rho < a/2$).
The action for such a region in
cylindrical coordinates is given by
\begin{equation}
\frac{\Gamma^0}{T L_z} = -\pi \int_0^{a/2}\rho B_z(\rho)^2d\rho.
\end{equation}
After substituting the value of equation (\ref{eqn:ffbless}), the magnetic field in the interior
region:
\begin{eqnarray}
\frac{\Gamma^0}{T L_z} &=& -\pi \int_0^{a/2}\rho \biggr[
\frac{4\mathcal{F}}{\lambda^2e^2q_2}\left(1-\frac{3}{4}
\left(\frac{\lambda-\lambda_{\rm min}}{a-\lambda_{\rm min}}\right)\right)
\Psi(2\rho/\lambda)\nonumber \\
& & + \frac{6\mathcal{F}}{ea^2}\left(\frac{\lambda-\lambda_{\rm min}}{a-\lambda_{\rm min}}\right)
\biggr]^2 d\rho.
\end{eqnarray}
After some algebra, we are left with an expression for the classical action,
\begin{eqnarray}
\frac{\Gamma_0}{TL_z} &=& \pi \int_0^{a/2}\rho B_z(\rho)^2d\rho \\
& = & -\frac{\pi \mathcal{F}^2}{e^2a^2}\biggr\{4\frac{a^2q_3}{\lambda^2q_2^2}
+\left(\frac{\lambda-\lambda_{\rm min}}{a - \lambda_{\rm lmin}}\right)
\biggr[ \nonumber \\
& &\left(\frac{9}{4}\frac{a^2q_3}{\lambda^2q_2^2}-\frac{9}{2}\right)
\left(\frac{\lambda-\lambda_{\rm min}}{a - \lambda_{\rm lmin}}\right)
-6\frac{a^2q_3}{\lambda^2q_2^2}+12 \biggr]\biggr\},
\end{eqnarray}
where $q_3$ is another numerical constant related to integrating the bump function:
\begin{equation}
\label{eqn:q3}
q_3 \equiv \int_0^1 x\left(\Psi(x)\right)^2 dx
= \int_0^1 x e^{-\frac{2}{1-x^2}}dx \approx 0.0187671.
\end{equation}
\subsection{Integrating to Find the Potential Function}
To compute the Wilson loops, it is generally required to use the
vector potential which describes the magnetic field. For us, this
means that we must find $f_\lambda(\rho)$ for our magnetic field
model. This could always be done numerically, but can be
computationally costly since it is evaluated by every discrete point
of every worldline in the ensemble. For computations on the \ac{CUDA}
device, an increase in the complexity of the kernel often means that
less memory resources are available per processing thread, limiting
the number of threads that can be computed concurrently. It is
therefore preferable to find an analytic expression for this function.
From equation (\ref{eqn:isoB}), this function is related to the
integral of the magnetic field with respect to $\rho^2$. For the inner
region, we have
\begin{eqnarray}
f_\lambda(\rho < a/2) &=& \frac{e}{2\mathcal{F}}\biggr[\frac{4\mathcal{F}}{\lambda^2eq_2}
\left(1-\frac{3}{4}\left(\frac{\lambda-\lambda_{\rm min}}{a-\lambda_{\rm min}}\right)\right) \times \nonumber \\
& & \int_0^{\rho^2}\Psi\left(\frac{2\rho'}{\lambda}\right)d \rho'^2 + \\
& &\frac{6 \mathcal{F}}{ea^2}\left(\frac{\lambda-\lambda_{\rm min}}{a-\lambda_{\rm min}}\right)\rho^2\biggr]. \nonumber
\end{eqnarray}
The integral over the bump function can be computed in terms of the exponential integral
$E_i(x) = \int_{-\infty}^x \frac{e^t}{t} dt$:
\begin{eqnarray}
\int_0^{\rho^2}\Psi\left(\frac{2\rho'}{\lambda}\right)d\rho'^2 &=&
\left(\frac{\lambda}{2}\right)^2\Biggr[2q_2+\left(\frac{4\rho^2}{\lambda^2}-1\right)
e^{-\frac{1}{1-\frac{4\rho^2}{\lambda^2}}}- \nonumber \\
& & ~~~ E_i\left(-\frac{1}{1-\frac{4\rho^2}{\lambda^2}}\right)\Biggr]
\end{eqnarray}
for $\rho < \lambda/2$ and
\begin{equation}
\int_0^{\rho^2}\Psi\left(\frac{2\rho'}{\lambda}\right)d\rho'^2 = \frac{q_2\lambda}{2}
\end{equation}
for $\rho \ge \lambda/2$. Our expression for the profile function in the inner region is
\begin{equation}
f_\lambda(\rho\le a/2) = \left(1-\frac{3}{4}\left(\frac{\lambda-\lambda_{\rm min}}{a-\lambda_{\rm min}}\right)\right)
\Phi(2\rho/\lambda) + \frac{3\rho^2}{a^2}\left(\frac{\lambda-\lambda_{\rm min}}{a-\lambda_{\rm min}}\right),
\end{equation}
with
\begin{equation}
\Phi(x) \equiv
\begin{cases}
1 + \frac{1}{2q_2}\left(x^2-1\right)e^{-\frac{1}{1-x^2}}
-\frac{1}{2q_2}E_i\left(-\frac{1}{1-x^2}\right)& \mbox{ for } x < 1\\
1 & \mbox{ for } x \ge 1
\end{cases}.
\end{equation}
The exterior integral is a bit more challenging, but we can make
significant progress and obtain an approximate expression. The first
term is a constant given by the value of the profile function at $\rho
= a/2$. This value is given by the flux in the central flux tube,
which we have already chosen to be 1,
\begin{equation}
f_\lambda(\rho>a/2) = 1 + \frac{e}{\mathcal{F}}\int_{a/2}^{\rho}\rho'B(\rho'>a/2)d\rho'.
\end{equation}
We may put the magnetic field, equation (\ref{eqn:Bzext}), into this expression to get
\begin{eqnarray}
f_\lambda(\rho>a/2) &=& 1 + \frac{3}{4}\left(\frac{4\rho^2}{a^2}-1\right)
\left(\frac{\lambda-\lambda_{\rm min}}{a-\lambda_{\rm min}}\right) + \\
& & \frac{12}{q_1a\lambda}\left(\frac{a-\lambda}{a-\lambda_{\rm min}}\right)
\int_{a/2}^{\rho}\rho'\Psi\left(\frac{2(\rho-na)}{\lambda}\right)d\rho'. \nonumber
\end{eqnarray}
The remaining integral is over every bump between $\rho'=a/2$ and
$\rho'=\rho$. We express the result as a term which accounts for each
completely integrated bump, and an integral over the partial bump if
$\rho$ is within a bump:
\begin{eqnarray}
f_\lambda(\rho>a/2) &=& 1 + \frac{3}{4}\left(\frac{4\rho^2}{a^2}-1\right)
\left(\frac{\lambda-\lambda_{\rm min}}{a-\lambda_{\rm min}}\right)+ \nonumber \\
& & 3n(n-1)\left(\frac{a-\lambda}{a-\lambda_{\rm min}}\right) + \\
& & \frac{3\lambda}{q_1a}\left(\frac{a-\lambda}{a-\lambda_{\rm min}}\right)\chi(2(\rho-na)/\lambda), \nonumber
\end{eqnarray}
where
\begin{equation}
\chi(x_0) =
\begin{cases}
0 & \mbox{ for } x_0 \le -1 \\
\int_{-1}^{x_0}xe^{-\frac{1}{1-x^2}}dx
+ \frac{2na}{\lambda}\int_{-1}^{x_0}e^{-\frac{1}{1-x^2}}dx& \mbox{ for } |x_0| < 1\\
\frac{2naq_1}{\lambda} & \mbox{ for } x_0 \ge 1
\end{cases}.
\end{equation}
One of the integrals in $\chi(x_0)$ can be expressed in terms of the exponential integral:
\begin{equation}
\int_{-1}^{x_0}xe^{-\frac{1}{1-x^2}}dx =\frac{1}{2}\left[(x_0^2-1)e^{-\frac{1}{1-x_0^2}}
-E_i\left(-\frac{1}{1-x_0^2}\right)\right].
\end{equation}
The remaining integral cannot be simplified analytically. To use this
integral in our numerical model, it must be computed for each discrete
point on each loop for each $\rho_{\rm cm}$ and $T$ value. Therefore,
it is worthwhile to consider an approximate expression which models
the integral, and can be computed faster than performing a numerical
integral each time. To find this approximation, we computed the
numerical result at 300 values between $x_0=-1.2$ and $x_0=1.2$. The
data was then input into Eureqa Formulize, a symbolic regression
program which uses genetic algorithms to find analytic representations
of arbitrary data~\cite{formulize}. A similar technique has been used
to produce approximate analytic solutions of ODEs~\cite{geneticODEs}.
The result is a model of the numerical data points with a maximum
error of 0.0001 on the range $|x_0| < 1.0$:
\begin{equation}
\int_{-1}^{x_0} e^{-\frac{1}{1-x^2}}dx \approx
\frac{0.444}{1+e^{-3.31x_0 - g(x_0)}}
\end{equation}
where
\begin{equation}
g(x_0) =
\frac{5.25x_0^3 - 3.31x_0^2\sin{(x_0)}\cos{(-0.907x_0^2-1.29x_0^8)}}{\cos{(x_0)}}
\end{equation}
This function evaluates ten times faster than the numerical
integral evaluated at the same level of precision with the \ac{GSL}
Gaussian quadrature library functions and with fewer memory registers.
Using this approximation introduces a systematic uncertainty which is
small compared to that associated with the discretization of the loop
integrals, and considerably smaller than the statistical error bars.
Using these expressions for the integrals, we can express
$f_\lambda(\rho)$ in any region in terms of exponential integrals,
exponential, and trigonometric functions with suitable precision.
Furthermore, for computation we may express the exponential integral
as a continued fraction. The profile function, $f_\lambda(\rho)$, is
plotted in figure \ref{fig:periodicfl}.
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.4in 0.1in 0.55in 0.3in]{images/periodicfl}
\caption[The profile function in a cylindrical lattice model]
{The function $f_\lambda(\rho)$ for the above described magnetic
field model. The flux conditions require the function to pass through the
black dots. A quadratic function corresponds to a uniform field while
a staircase function corresponds to delta-function flux tubes. The parameter
$\lambda$ smoothly makes the transition between these two extremes.
Each of these functions corresponds to a magnetic field profile in
figure \ref{fig:periodicB}.}
\label{fig:periodicfl}
\end{figure}
\section{Results}
\label{sec:periodic_results}
\subsection{Comparing Scalar and Fermionic Effective Actions}
Because of the fermion problem of worldline numerics \cite{Gies:2001zp,
MoyaertsLaurent:2004}, the 1-loop effective action for the
cylindrical flux tube lattice model was not computed for the case
of spinor \ac{QED}. Performing this calculation in the spinor case
would require subtle numerical cancellations between large terms,
and represents a significant numerical challenge.
Fortunately, the fermion problem does not affect the scalar
case. So, we will analyze this model for \ac{ScQED}. However,
in this section we will compare the scalar and fermionic effective
actions for isolated flux tubes to demonstrate that the behaviour of
both theories is qualitatively and numerically similar.
For isolated flux tubes, the decay of the magnetic field for large distances
protects the calculations from the fermion problem. Therefore, the
effective action can be computed for both scalar and spinor
\ac{QED}. In figure \ref{fig:scalfermiratio}, we plot the
ratio of the spinor to scalar 1-loop correction term for
identical magnetic fields, along with the prediction of the
\ac{LCF} approximation for large values of $\lambda$. The
\ac{LCF} approximation in \ac{ScQED} is given by
\begin{eqnarray}
\label{eqn:LCFscal}
\Gamma^{(1)}_{\rm scal} &=& -\frac{1}{2\pi}\int_0^\infty dT
\int_0^\infty \rho_{\rm cm} d\rho_{\rm cm}\frac{e^{-m^2T}}{T^3} \nonumber \\
& &\left\{\frac{eB(\rho_{\rm cm})T}{\sinh{(eB(\rho_{\rm cm})T)} }
- 1 +\frac{1}{6}(eB(\rho_{\rm cm})T)^2\right\}.
\end{eqnarray}
This can be compared to the spinor \ac{QED} approximation,
\begin{eqnarray}
\label{eqn:LCFferm2}
\Gamma^{(1)}_\mathrm{ ferm} &=& \frac{1}{4\pi}\int_0^\infty dT
\int_0^\infty \rho_\mathrm{ cm} d\rho_\mathrm{ cm}\frac{e^{-m^2T}}{T^3} \times \nonumber \\
& &\biggl\{eB(\rho_\mathrm{ cm})T\coth{(eB(\rho_\mathrm{ cm})T)} \\
& &
- 1 -\frac{1}{3}(eB(\rho_\mathrm{ cm})T)^2\biggr\}. \nonumber
\end{eqnarray}
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.30in 0.1in 0.55in 0.3in]{images/scalfermiratio}
\caption[Comparison of 1-loop term in \acs{QED} and \acs{ScQED}]
{The ratio of the 1-loop term in \ac{QED} to the 1-loop term
in \ac{ScQED}. The solid line is the \ac{LCF} approximation,
while the points are the result of worldline numerics calculations. Note that the
\ac{LCF} approximation breaks down near $\lambda = \lambda_e$ and
that the statistics from point to point are strongly correlated. This plot shows that
the 1-loop correction in \ac{ScQED} differs from the \ac{QED} correction by a factor
close to unity for a wide range of flux tube widths.}
\label{fig:scalfermiratio}
\end{figure}
There are two important notes to make about figure \ref{fig:scalfermiratio}.
Firstly, the \ac{LCF} approximation is only a good approximation
for $\lambda \gg \lambda_e$, and isn't accurate when pushed near its formal
validity limits~\cite{MoyaertsLaurent:2004}. The second note is that the
statistics of the points computed with worldline numerics are strongly correlated. So, we conclude
that the \ac{ScQED} 1-loop correction is larger than the \ac{QED} correction
for large $\lambda$, and this appears to be reversed for small $\lambda$.
However, the large worldline numerics error bars and the invalidity of the
\ac{LCF} approximation near $\lambda = \lambda_e$ prevent us from
seeing how this transition happens. Nevertheless, the main
conclusion from this figure is that the scalar 1-loop correction
reflects the behaviour of the full \ac{QED} 1-loop correction
to within a factor of about 2 over a wide range of flux tube widths
for isolated flux tubes.
Besides using a finite field profile, the fermion problem can also be
circumvented by increasing the electron mass. The square of the electron mass
sets the scale for the exponential suppression of the large proper
time Wilson loops that contribute to the fermion problem. However, if
we increase the fermion mass, we are reducing the Compton wavelength
of our theory so that the flux tube lattice is no longer dense in
terms of the modified Compton wavelength. It is the Compton wavelength of
the theory that determines what is meant by `dense'. We therefore
cannot avoid the fermion problem for dense lattice models by changing
the electron mass.
Based on the results presented in figure \ref{fig:scalfermiratio}, we
conclude that the coupling between the electron's spin and the
magnetic field do not have a dramatic effect on the vacuum energy for
isolated flux tubes. Therefore, we expect that \ac{ScQED} provides a
good model of the underlying vacuum physics near these flux tubes, at
least at the level our toy model flux tube lattice.
\subsection{Flux Tube Lattice}
The worldline numerics technique computes an effective action density which is then integrated to
obtain the effective action. This quantity differs from the Lagrangian in that it is
not determined by local operators, but encodes information about the field everywhere
through the worldline loops.
Like the classical action, the 1-loop term of the effective action per unit length
is infinite for a flux tube lattice because the field extends infinitely far.
For this reason, we define the effective action to be the action density integrated
over the region of a central flux tube $(0<\rho<a/2)$:
\small
\begin{eqnarray}
\frac{\Gamma}{\mathcal{T}L_z} &=& -\pi \int_0^{a/2}\rho B_z(\rho)^2d\rho - \nonumber \\
& &\frac{1}{2\pi}\int_0^{a/2}\rho_{\rm cm} d\rho_{\rm cm}\int_0^\infty \frac{dT}{T^3}e^{-m^2T} \times
\nonumber \\
& & \left\{ \mean{W}_{\rho_{\rm cm}} - 1 +\frac{1}{6}(eB(\rho_{\rm cm})T)^2\right\}.
\end{eqnarray}
\normalsize
The 1-loop term of the effective action density is plotted in figure
\ref{fig:EAscffpeek} for the cylindrical flux tube lattice model. The
most pronounced feature of this density is that there is a negative
contribution from the regions where the field is strong. This
contribution has the same sign as the classical term. Therefore,
the quantum correction tends to reinforce the classical action. A less
pronounced feature is that there is a positive contribution arising
from the $\rho_{\rm cm} > \lambda/2$ region, in between the lumps of magnetic
field which represent the flux tubes. In this region, the local
magnetic field is positive, but small.
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.4in 0.1in 0.55in 0.3in]{images/EAscffpeek}
\caption[The 1-loop action density for a cylindrical lattice in \acs{ScQED}.]
{The \ac{ScQED} effective action density for the central
flux tube in our cylindrical lattice model for several tube widths,
$\lambda$. The average field strength is the critical field, $B_K$.
The effective action is positive in between flux tubes due to non-local
effects.}
\label{fig:EAscffpeek}
\end{figure}
To interpret this feature, we consider the relative contributions between the
Wilson loops and the counterterm. These terms are shown in figure
\ref{fig:CounterTermDomination} for the constant field case.
For all values of proper time, $T$, the counterterms dominate, giving
an overall negative sign. In order for the action density to be positive, there
must be a greater contribution from the Wilson loop average than from the
counterterm, since this term tends to give a positive contribution to the
action. In our flux tube model, this seems to occur in the regions between the
flux tubes. In these regions, the local contribution from the counterterm is
relatively small because the field is small. However, the contribution from the
Wilson loop average is large because the loop cloud is exploring the nearby
regions where the field is much larger. The effect is largest where the field
is small, but becomes large in a nearby region. We therefore interpret the positive
contributions to the 1-loop correction from these regions as a non-local
effect. A similar example of such an effect from fields which vary on
scales of the Compton wavelength has been observed previously using the
worldline numerics technique~\cite{PhysRevD.84.065035}.
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.5in 0.1in 0.55in 0.3in]{images/CounterTermDominationScalar}
\caption[Wilson loop and counterterm contributions for a constant field]
{The Wilson loop and counterterm contributions to the integrand
of the effective action for a constant field in \ac{ScQED}. For constant fields, the
effective action is always negative due to the domination of the counterterm
over the Wilson loop. For non-homogeneous fields,
a positive effective action density signifies
that non-local ({\em i.e.}\ $T > 0$) effects dominate the counterterm.}
\label{fig:CounterTermDomination}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.15in 0.1in 0.55in 0.3in]{images/EAscffvsl}
\caption[The 1-loop \acs{ScQED} term versus flux tube width]
{The 1-loop \ac{ScQED} term of the effective action as a function
of flux tube width, $\lambda/a$, for several values of the flux tube spacing, $a$.
The dotted lines are computed from the \ac{LCF} approximation.}
\label{fig:EAscffvsl}
\end{figure}
In figure \ref{fig:EAscffvsl}, we plot the magnitude of the 1-loop
\ac{ScQED} term of the effective action as a function of the flux tube
width. As the flux tubes become smaller, there is an amplification of
the 1-loop term, just as there is for the classical action.
Similarly, for more closely spaced flux tubes, $a$ is smaller, and the
1-loop term increases in magnitude. The ratio of the 1-loop term to
the classical term is plotted in figure \ref{fig:ActRatscff}. The
quantum contribution is greatest for closely spaced, narrow flux
tubes, but does not appear to become a significant fraction of the
total action.
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.25in 0.1in 0.55in 0.3in]{images/EAscffresid}
\caption[The deviation of the cylindrical flux tube lattice from the \ac{LCF}
approximation]
{The residuals between the worldline numerics results and the \ac{LCF} approximation
for the cylindrical flux tube lattice. The level of agreement observed is not expected because
the field varies rapidly on the Compton wavelength scale. This agreement is believed to
be due to an averaging effect of integrating over the electron degrees roughly reproducing
the result of a mean-field approximation.}
\label{fig:EAscffresid}
\end{figure}
We observe that the \ac{LCF} approximation is surprisingly good despite the
fact that the magnetic field is varying rapidly on the Compton wavelength scale of the electron.
We plot the residuals showing the deviations between the worldline numerics results and the \ac{LCF}
approximation in figure \ref{fig:EAscffresid}. To understand this, recall the discussion
surrounding figure \ref{fig:CounterTermDomination}. The Wilson loop term is sensitive to the
average magnetic field through the loop ensemble, $\mean{B}_{\rm e}$. In contrast, the counterterm
is sensitive to the magnetic field at the center of mass of the loop, $B_{\rm cm}$.
Since these terms carry opposite signs, we can understand the difference from the
constant field approximation in terms of a competition between these terms. When
$B_{\rm cm} < \mean{B}_{\rm e}$, such as when the center of mass is in a local minimum of the
field, there is a reduction of the energy relative to the locally constant field case, with
a possibility of the quantum term of the energy density becoming negative.
However, when $B_{\rm cm} > \mean{B}_{\rm e}$, such as in a local maximum of the field,
there is an amplification of the energy relative to the constant field case. We can put a
bound on the difference between the mean field through a loop and the field at the
center of mass for small loops ({\em i.e.}\ small $T$),
\begin{equation}
|\mean{B} - B_{\rm cm}| \lesssim |B''(\rho_0)| T
\end{equation}
where $|B''(\rho_0)| \ge B''(\rho)$ for all $\rho$ in the loop. This expression is proved the
same way as determining the error in numerical integration using the midpoint rectangle
rule.
If the field varies rapidly about some mean value
on the Compton wavelength scale, the various contributions
from local minima and local maxima are averaged out and the mean field approximation
provided by the \ac{LCF} method becomes appropriate. A similar argument applies in the
fermion case, where the important quantity is the mean magnetic field along the circumference
of the loop. This quantity is also well served by a mean-field approximation when integrating
over rapidly varying fields.
Another interesting feature of figure \ref{fig:EAscffresid} is that the \ac{LCF}
approximation appears to describe narrower flux tubes better than wider ones, even when the
spacing between the flux tubes is held constant. This effect is likely a result of the compact
support given to the flux tube profiles. For narrow tubes, we are guaranteed to have many more
center of mass points outside the flux tube than inside, giving a smaller energy contribution than
for isolated flux tubes without compact support where the distinction between inside and outside is
not as abrupt. This also explains why narrow, closely spaced tubes produce a lower energy than
is predicted by the \ac{LCF} approximation.
This argument does not apply to the smooth isolated flux tubes given
by equation (\ref{eqn:isoB}).
For these flux tubes, the only region where there is a large discrepancy between $\mean{B}_{\rm e}$ and
$B_{\rm cm}$ is near the center of the flux tube. This
is a global maximum of the field, and the only maximum of $|B''(\rho)|$.
There are no regions where the average field in the loop ensemble is
much stronger than the center of mass magnetic field. So, we expect an amplification of the
energy near the flux tube relative to the constant field case.
In the flux tubes with compact support, however,
there is such a region just outside the flux tube. We can understand the surprisingly close agreement
of these results to the \ac{LCF} approximation in our model
in terms of competition between these
regions of local minima and maxima of the field (see figure \ref{fig:EAscffvsconst}).
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.25in 0.1in 0.55in 0.3in]{images/EAscffvsconst}
\caption[A comparison between the action density in the \ac{LCF} approximation and in worldline numerics]
{The action density in worldline numerics and the \ac{LCF} approximation, scaled by $\rho$ so
an area on the figure is proportional to a volume. The approximation
is poor everywhere, however, when there are regions of local minima and local maxima of the field about
a mean value, the effective action approximately agrees between these methods. This is due to
a partial cancellation between regions where the estimate provided by the approximation is too
large (green) and other regions where the estimate is too small (red).}
\label{fig:EAscffvsconst}
\end{figure}
Finally, we find that the quantum term remains small compared to the
classical action for the range of parameters investigated. This is shown
in figure \ref{fig:ActRatscff} where we plot the ratio of the \acs{ScQED}
term of the action to the classical action. The relative smallness of this correction
is consistent with the predictions from homogeneous fields and the derivative
expansion, as well as with previous studies on flux tube
configurations~\cite{1999PhRvD..60j5019B}.
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.15in 0.1in 0.55in 0.3in]{images/ActRatscff}
\caption[The 1-loop \acs{ScQED} quantum to classical ratio versus flux tube width]
{The 1-loop \ac{ScQED} term divided by the classical term of the effective action as a function
of flux tube width, $\lambda/a$ for several values of the flux tube spacing, $a$.}
\label{fig:ActRatscff}
\end{figure}
\subsection{Interaction Energies}
Using this model, we can study the energy associated with
interactions between the flux tubes. Since the flux tubes in our model
exhibit compact support, the interaction energy is entirely due to a
non-local interaction between nearby flux tubes. Thus, it contrasts
with previous research which has investigated the interaction energies
between flux tubes which have overlapping
fields~\cite{Langfeld:2002vy}. In this case, there is a classical
interaction energy ($\propto B_1 B_2$) as well as a quantum correction
($\propto (B_1 + B_2)^4 - B_1^4 - B_2^4$ in the weak-field limit).
Even when these field overlap interactions are not present, there are
also non-local energies in the vicinity of a flux tube due to the
presence of other flux tubes. For example, the energy from nearby flux
tubes can interact with a flux tube through the quantum diffusion of
the magnetic field. Because of this phenomenon, we expect an
interaction energy in the region of the central flux tubes due to the
proximity of neighbouring flux tubes, even though no changes are made
to the field profile or its derivatives in the region of interest.
Since this interaction represents a force due to quantum fluctuations
under the influence of external conditions, it is an example of a
Casimir force. The Casimir force between two infinitely thin flux
tubes in \ac{ScQED} has previously been found to be
attractive~\cite{duru1993}. Our model can shed light specifically on
this interaction, which is not predicted by local approximations such
as the derivative expansion.
Consider a central flux tube with a width $\lambda = 0.5\lambda_e$.
When $\lambda_{\rm min} = \lambda$, the magnetic field outside of the
flux tube, $B_{\rm bg}$, is zero. Then, if the distance between flux
tubes, $a$, is set very large, the energy density will be localized to
the central flux tube and there will be no non-local interaction energy
due to neighbouring flux tubes. We define the interaction energy,
$E_{\rm int}$, as the difference in energy within a distance $a/2$ of
the central flux tube between a configuration with a given value of
$a$ and a configuration with $a = \infty$. In practice, we use $a =
10,000\lambda_e$ as a suitable stand-in for $a = \infty$:
\begin{equation}
\frac{E_{\rm int}}{L_z} = -\frac{\Gamma_{\rm scal}(a)}{L_z \mathcal{T}}
+\frac{\Gamma_{\rm scal}(a=1\times 10^4 \lambda_e)}{L_z \mathcal{T}}.
\end{equation}
With this definition, the interaction energy is the energy associated with
lowering the distance between flux rings from infinity. This the the analogue
in our model of reducing the lattice spacing of the flux tubes.
One complication of this definition is that there is no clear
distinction between energy density which `belongs' to the central flux
tube and energy density which `belongs' to the neighbouring flux
tubes. We continue to use our convention that the total energy for the
central flux tube is determined by the integral over the non-local
action density in a region within a radius of $a/2$ of the flux
tube. As $a$ is taken smaller and smaller, some energy from nearby
flux tubes is included within this region, but also, some energy
associated with the central flux tube is diffused out of the
region. This ambiguity is unavoidable within this model. We can't
numerically compute the energy over all of space and subtract off
different contributions, because these energies are infinite.
\begin{figure}
\centering
\includegraphics[width=\linewidth,clip,trim=0.25in 0.1in 0.55in 0.3in]{images/interactionb}
\caption[The interaction energy versus the flux tube spacing]
{The interaction energy per unit length of flux tube as a function
of the flux tube spacing, $a$. The energy density of a critical
strength magnetic field is 17 GeV/$\lambda_e^3$, so this
energy density is small in comparison to the classical magnetic field energies,
or the 1-loop corrections. However, the local interactions are constant in $a$.
At $a<\lambda_e/2$,
the bump functions from neighbouring tubes overlap. This approximately corresponds
to the critical magnetic field which destroys superconductivity.}
\label{fig:interaction}
\end{figure}
The interaction energy is plotted in figure \ref{fig:interaction}. In
this plot, the error bars are 1-sigma error bars that account for the
correlations in the means computed for each group of worldlines,
\begin{equation}
\sigma_{E_{\rm int}} = \sqrt{\sigma_{E_a}^2 + \sigma_{E_{a=10000}}^2
- 2\operatorname{Cov}(E_a, E_{a=10000})},
\end{equation}
where $\operatorname{Cov}(a,b)$ is the covariance between random
variables $a$ and $b$. Recall from figure \ref{fig:EAscffpeek} that
there is a positive contribution to the effective action, and
therefore a negative contribution to the energy from the region
between flux tubes. As we reduce $a$, bringing the flux tubes closer
together, two considerations become important. Firstly, we are
increasing the average field strength meaning there tends to be more
flux through the worldline loops which tends to give a negative
contribution to the interaction energy. Secondly, we are reducing the
spatial volume over which we integrate the energy since we only
integrate $\rho$ from $0$ to $a/2$. This effect makes a positive
contribution to the interaction energy since we include less and less
of the region of negative energy density in our integral.
In figure \ref{fig:interaction}, there appears to be a landscape with both
positive and negative interaction energies at different values of $a$.
These appear to be consistent with the interplay between positive and
negative contributions described in the previous paragraph. This is consistent with
the usual expectation of attractive Casimir forces~\cite{duru1993}. The dominant contribution
for the positive energy values is caused by less of the negative energy region contributing
as the domain assigned to the flux tube is reduced. However, the point at $a/\lambda_e = 2$ is
negative ($-0.3\pm0.1$MeV/$\lambda_e$) indicating that an attractive interaction from
nearby flux tubes is dominant.
At $a/\lambda_e = 0.5$, the flux tubes are positioned right next to
one another, and the negative contribution from the increase in the
mean field appears to be slightly larger than the positive
contribution from the loss of the region of negative energy density
from the integral. Beyond this, the flux tubes would overlap each
other, which approximately corresponds to the critical background
field which destroys superconductivity. Based on the above
explanation, it appears that the non-local interaction energy between
magnetic fields has a strong dependence on the specific profile of the
classical magnetic field that was used. This makes it difficult to
predict if it will result in attractive or repulsive forces in a more
realistic model of a flux tube lattice.
The energy density of a critical strength magnetic field is about
$17{\rm GeV}/\lambda_e^3$, so the energy associated with this
interaction is relatively small. However, there are no other
interactions which affect the energy of moving the flux tubes closer
together when they are separated by many coherence lengths. Here, the
characteristic distance associated with the interaction, $\lambda_e$,
is considerably larger than coherence length or London penetration
depth, so the interactions between flux tubes through the order field
are heavily suppressed.
\section{Conclusion}
\label{sec:conclusion}
In this paper, we have used the worldline numerics numerical technique
to compute the effective action of \ac{QED} in
non-homogeneous, cylindrically symmetric magnetic fields. The method
uses a Monte Carlo generated ensemble of worldline loops to
approximate a path integral in the worldline formalism. These
worldline loops are generated using a simple algorithm and encode the
information about the magnetic field by computing the flux through the
loop and the action acquired from transporting a magnetic moment
around the loop. This technique preserves Lorentz symmetry exactly
and can preserve gauge symmetry up to any required precision.
Computing the quantum effective action for magnetic flux tube
configurations is a problem that has generated considerable interest
and has been explored through a variety of approaches~\cite{Gornicki1990271,
1998MPLA...13..379S, 0264-9381-12-5-013, PhysRevD.51.810, 1999PhRvD..60j5019B,
PhysRevD.62.085024, 2001PhRvD..64j5011P, 2003PhRvD..68f5026B,
2005NuPhB.707..233G, 2006JPhA...39.6799W, Weigel:2010pf, Langfeld:2002vy}
(see section \ref{sec:EAisoflux}).
Partly, this
is because it is a relatively simple problem for analyzing non-homogeneous
generalizations of the Heisenberg-Euler action and for exploring
limitations of techniques such as the derivative expansion. But, this
is also a physically important problem because tubes of magnetic flux
are very important for the quantum mechanics of electrons due
to the Aharonov-Bohm effect, and they appear
in a variety of interesting physical scenarios such as stellar astrophysics,
cosmic strings, in superconductor vortices, and quark confinement
\cite{2003PrPNP..51....1G}.
In the present context, we are concerned with the role that magnetic
flux tubes play in the superconducting nuclear material of compact stars.
In this scenario, the \ac{QED} effects are particularly interesting because
the magnetic flux tubes, if they exist, are confined to tubes which may be
only a few percent of the Compton wavelength, $\lambda_{\rm C}$, in radius. Specifically,
the flux must be confined to within the London penetration depth of the superconducting
material, which for neutron stars has been estimated to be 80 fm $=$
0.032 $\lambda_e$~\cite{PhysRevLett.91.101101}. Moreover, the
flux tube density is expected to be proportional to the
average magnetic field. For a background field near the quantum critical
strength, $B_k$, such as in a neutron star, the distance between flux tubes
is comparable to a Compton wavelength. This Compton wavelength scale is
also the scale at which the non-locality of \ac{QED} becomes important and
at which powerful local techniques like the derivative expansion
are no longer appropriate for computing the effective action.
The free energy associated with these flux tubes is a factor in determining
whether the nuclear material of a neutron star is a type-I or type-II
superconductor. The free energy of a flux tube is determined by
looking at the energies associated with the magnetic field, with the
creation of a non-superconducting region in the superconductor, and
with interactions between the flux tubes. Flux tubes can only form
if it is energetically favourable to do so compared to expelling the
field due to the Meissner effect. For a lattice of
flux tubes, there is also an energy contribution from the
presence of neighbouring flux tubes because of the non-local
nature of quantum field theory.
The energy of two flux tubes has been previously
computed using worldline numerics methods and for flux tubes with aligned fields, the
energy is larger than twice the energy of a single flux tube when the flux tubes
are closely spaced~\cite{Langfeld:2002vy}. This result implies that there is a
repulsive interaction between the flux tubes due to \ac{QED} effects,
strengthening the likelihood of the type-II scenario in neutron stars. This interaction
energy increases as the flux tubes are placed closer together, and can have a similar
magnitude as the \ac{QED} correction to the energy when the flux tubes are closely spaced.
We have developed a cylindrically symmetric magnetic field
model which reproduces some of the
features of a flux tube lattice: for a given central flux tube, there are
nearby regions of large magnetic field that interact non-locally, and
the large flux tube size limit goes to a large uniform magnetic field
instead of to zero field. We have investigated the 1-loop effects from
\ac{ScQED} in this model using the worldline numerics technique for various
combinations of flux tube size, $\lambda$, and flux tube spacing, $a$.
In contrast to isolated flux tubes, we find that there are some regions where the
worldline numerics results are greater than the \ac{LCF} approximation and other
regions where they are less than the \ac{LCF} approximation. This can be understood
by thinking of the difference from the \ac{LCF} approximation as a competition
between the local counterterm and the Wilson loop averages.
For magnetic fields that vary on the Compton wavelength scale
about some mean field strength, the \ac{LCF} approximation
provides a poor approximation of the energy density, but may provide a
good approximation to the total energy density of the field due to it
being a good mean field theory approximation to the energy density.
The appropriateness of the \ac{LCF} approximation in this case
can be understood as an approximate balance between regions where the
field is a local maximum and the magnitude of the quantum correction to the
action density is larger than in the constant field case, and regions where
the field is a local minimum and the quantum corrections predict a smaller
action density than the constant field case. This
washing out of the field structure due to non-local effects has also
been observed in worldline numerics studies of the vacuum polarization
tensor~\cite{PhysRevD.84.065035}.
There is a force between nearby magnetic flux tubes due
to the quantum diffusion of the energy density. This interaction is
non-local and is not predicted by the local derivative expansion.
It is an example of a Casimir force ({\em i.e.}\ a force resulting from
quantum vacuum fluctuations) and it is computed in a very similar
way as the Casimir force between conducting bodies in the worldline numerics
technique~\cite{Gies:2003cv}.
The size of the energy densities involved in this force are small even
compared to the 1-loop corrections to the energy densities, which are
in turn small compared to the classical magnetic energy density.
Although this interaction energy is small,
the interactions between flux tubes in a neutron star due to the
order field of the superconductor are suppressed because the distance between the
tubes is considerably larger than the coherence length and
London penetration depth. Therefore, this force is possibly important
for the behaviour of flux tubes in neutron star crusts and interiors.
For example, in our lattice model, this force could contribute
to a bunching of the worldlines, producing regions where flux tubes
are separated by $\sim2\lambda_e$ and other regions which have
no flux tubes. Consequently, this force may have important implications
for neutron star physics. However, investigating these implications is
outside the scope of this paper.
The nature of this interaction energy is expected to depend on the
model of the magnetic field profile for the reasons discussed in
section \ref{sec:periodic_results}. It is therefore reasonable that
forces of either sign, attractive or repulsive, may be possible
depending on the particular landscape formed by the magnetic field and
the particular definition used of the interaction energy. In a
superconductor, the profiles of the magnetic flux tubes are determined
by the minimization of the free energy for the interacting system
formed by the magnetic and order parameter fields. Therefore,
investigating this phenomenon using more realistic models ({\em i.e.}\ two-dimensional
triangular lattices with field profiles
motivated by the physics of superconductors) is an
interesting direction for future research. In particular, it would be
interesting to determine if certain conditions allowed for a non-local
interaction between magnetic flux tubes to be experimentally
observable despite the small forces involved.
These conclusions are directly applicable to \ac{ScQED}. However, we
have also investigated the relationship between spinor and scalar
\ac{QED} for isolated flux tubes where the worldline numerics technique can be
applied to both cases. We find that both theories have the same
qualitative behaviour, and agree within a factor of order unity
quantitatively. The arguments and explanations given for the
\ac{ScQED} results have strong parallels in the spinor \ac{QED}
case. The spinor case can also be understood in terms of a competition
between the Wilson loop averages and the local counterterm. We
therefore speculate that the results from this paper will hold in
the spinor case, at least qualitatively. However, addressing the
fermion problem so that the spinor case can be studied explicitly for
flux tube lattices would be valuable progress in this area of
research.
This work was supported by the Natural Sciences and Engineering Research
Council of Canada, the Canadian Foundation for Innovation, the British
Columbia Knowledge Development Fund. It has made used of the NASA ADS
and arXiv.org.
\bibliographystyle{prsty}
|
\section{Introduction}
In this paper we investigate the convergence of full discretizations, explicit in time, of stochastic evolution equations
\begin{equation}
\label{eq:see}
du(t) = A u(t) dt + B u(t) dW(t)\, , \,\, t \in [0,T]
\end{equation}
with the drift term
governed by a super-linearly growing operator.
When the operator appearing in the drift term grows at most linearly
then the classical explicit Euler scheme applied to stochastic evolution equations is convergent
(when coupled appropriately with the spatial discretization),
see, for example, Gy\"ongy and Millet~\cite{gyongy:millet:on:discretization}.
If the operator appearing in the drift term grows faster than linearly then one would, in general, not expect the explicit Euler scheme to be convergent
(this is the case even in the setting of fully deterministic evolution equations).
Instead, one would typically consider the implicit Euler scheme which is convergent in this situation (see, for example, Gy\"ongy and Millet~\cite{gyongy:millet:on:discretization}).
The price one pays is the increased computational effort required at each time step of the numerical scheme.
Hutzenthaler, Jentzen and Kloeden~\cite{HJK-div} have observed that
the absolute moments of explicit Euler approximations for
stochastic differential equations with super-linearly growing coefficients
may diverge to infinity at finite time.
This led to development of ``tamed'' Euler schemes for stochastic
differential equations.
This was pioneered in Hutzenthaler, Jentzen and Kloeden~\cite{hutzenthaler:jentzen:kloeden:strong:convergence}
and, using different techniques, in Sabanis~\cite{sabanis:note}.
A taming-like technique in the form of truncation has been proposed
by Roberts and Tweedie~\cite{roberts:tweedie:exponential} in the
context of Markov chain Monte Carlo methods.
Further work on explicit numerical approximations of stochastic differential
equations with super-linearly growing coefficients can be found
in Tretyakov and Zhang~\cite{tretyakov:zhang:fundamental},
Hutzenthaler and Jentzen~\cite{hutzenthaler:jentzen:numerical},
Sabanis~\cite{sabanis:euler}
as well Dareiotis, Kumar and Sabanis~\cite{dareitos:kumar:sabanis}.
Moreover Hutzenthaler, Jentzen and Kloeden~\cite{hutzenthaler:jentzen:kloeden:strong:divergence:multilevel} have demonstrated
that to apply multilevel Monte Carlo methods (see Heinrich~\cite{heinrich:monte}, \cite{heinrich:multilevel} and Giles~\cite{giles:multilevel})
to stochastic differential equations with super-linearly growing coefficients
one must ``tame'' the explicit Euler scheme.
In this paper we use the idea of ``taming'' to devise a new type of a convergent explicit scheme for a class of stochastic evolution equations with super-linearly growing operators in the drift term.
The article is organised as follows. In Section~\ref{sec:main} we
introduce the numerical scheme, give the precise assumptions and state
the main result in Theorem~\ref{thm:main}.
Essential a priori estimates are proved in Section~\ref{sec:apriori_est}.
In Section~\ref{sec:convergence} we first use the a priori estimates
and a compactness argument to extract weakly convergent subsequences
and limits of the approximation.
The remaining step is to identify the weak limit of the approximation
of the nonlinear term with the nonlinear term in the equation.
This is done using a monotonicity argument in Section~\ref{sec:convergence}
where Theorem~\ref{thm:main} is finally proved.
In Section~\ref{sec:examples} we provide examples of stochastic partial
differential equations where the numerical scheme can be applied.
\section{Main results}
\label{sec:main}
Let $T>0$.
Let $(\Omega, \mathcal{F}, \P)$ be a probability space and let $(\mathcal{F}_t)_{t\in [0,T]}$
be a filtration such that $\mathcal{F}_0$ contains all
the $\P$-null sets of $\mathcal{F}$.
Let $K>0$ and $p \in [2,\infty)$ be given constants.
Let $p^* := p / (p-1)$.
For a reflexive, separable Banach space $(X,\|\cdot\|_X)$ let $X^*$ and $\|\cdot\|_{X^*}$ denote its dual space and the norm on the dual space respectively.
For $f\in X^*$ and $v\in X$ we use $\langle f,v \rangle$ to denote the duality pairing.
By $L^p(0,T;X)$ we denote the Lebesgue--Bochner space of
equivalence classes of measurable functions
with values in $X$ that satisfy
\begin{equation*}
\|x\|_{L^p(0,T;X)} := \left(\int_0^T\|x(t)\|_X^p dt)\right)^{1/p} < \infty.
\end{equation*}
By $L^p(\Omega; X)$ we denote the Lebesgue--Bochner space of random variables with values in $X$ and such that the norm
\begin{equation*}
\|x\|_{L^p(\Omega;X)} := \left(\mathbb{E}(\|x\|_X^p)\right)^{1/p}
\end{equation*}
is finite.
Finally by $\mathcal{L}^p(X)$ we denote the Lebesgue--Bochner space of
$dt\times \P$-equivalence classes of
$(\mathcal{F}_t)_{t\in[0,T]}$-adapted and $X$-valued stochastic process that satisfy
\begin{equation*}
\|x\|_{\mathcal{L}^p(X)} := \left( \mathbb{E}\int_0^T \|x(t)\|_X^p dt \right)^{1/p} < \infty.
\end{equation*}
We say that an operator $C:X\times \Omega \to X^*$ is measurable with respect to some $\mathcal{G} \subseteq \mathcal{F}$ if for any $v,w\in X$ the real-valued random variable $\langle C v,w\rangle$ is $\mathcal{G}$-measurable.
We assume that, with respect to $(\mathcal{F}_t)_{t\in [0,T]}$,
$(W_t)_{t\in [0,T]}$ is a cylindrical $Q$-Wiener process with $Q=I$ on a separable Hilbert space $(U,(\cdot,\cdot)_U, |\cdot|_U)$.
We assume that there are $(V_1,\|\cdot\|_{V_1})$ and $(V_2,\|\cdot\|_{V_2})$, separable and reflexive Banach spaces that are densely and continuously embedded in $H$, where $(H,(\cdot,\cdot),|\cdot|)$ is a Hilbert space identified with its dual.
We thus have two Gelfand triples
\begin{equation*}
V_i \hookrightarrow H \hookrightarrow V_i^*\,, \,\,\, i \in \{1,2\}.
\end{equation*}
Let $A_i$ with $i\in\{1,2\}$ be operators defined on $V_i\times \Omega$
with values in $V_i^*$.
Let $B_i$ with $i\in\{1,2\}$ be operators defined on $V_i\times \Omega$
with values in $L_2(U,H)$,
where $L_2(U,H)$ is the space of Hilbert--Schmidt operators from $U$ to $H$.
Let $V:=V_1 \cap V_2$ and let the norm in $V$ be given by
$\|\cdot\| := \|\cdot\|_{V_1} + \|\cdot\|_{V_2}$.
Assume that $V$ is separable and dense in both $V_1$ and $V_2$.
Using Gajewski, Gr\"oger and Zacharias~\cite[Kapitel I, Satz 5.13]{ggz}
one observes that the dual $V^*$ of $V$ can be identified with
\begin{equation*}
V_1^* + V_2^* := \{f = f_1 + f_2 : f_1 \in V_1^*,\, f_2\in V_2^*\}
\end{equation*}
and that for all $f \in V^*$
\begin{equation*}
\|f\|_{V*} = \inf \{\max(\|f_1\|_{V_1^*}, \|f_2\|_{V_2^*}): f = f_1 + f_2, f_1 \in V_1^*,\,\, f_2\in V_2^* \}.
\end{equation*}
We consider stochastic evolution equations of the form
\begin{equation}
\label{eq:see2}
du(t) = \big[A_1 u(t) + A_2 u(t)\big] dt + \big[B_1 u(t) + B_2 u(t)\big] dW(t)\, , \,\, t \in [0,T],
\end{equation}
where $u(0) = u_0$ with $u_0$ a given $H$-valued and $\mathcal{F}_0$-measurable random variable.
Let $A:=A_1 + A_2$ and $B:=B_1 + B_2$.
The operator $A$ is defined on $V\times\Omega$ with values in $V^*$
and the operator $B$ is defined on $V\times\Omega$ with values in $L_2(U,H)$.
Then we can write~\eqref{eq:see2} as~\eqref{eq:see}.
We impose the following assumptions on the operators.
\begin{assumption}
\label{ass:operators}
Let $A_i : V_i \times \Omega \to V_i^*$
be $\mathcal{F}_0$-measurable operators for $i\in \{1,2\}$.
Let $B_i: V_i \times \Omega \to L_2(U,H)$ be such that
for any $v\in V_i$, $u\in U$ and $h\in H$ the real-valued
random variable $((B_i v)u,h)$ is $\mathcal{F}_0$-measurable for $i\in \{1,2\}$.
Moreover assume that the following conditions hold.
{\em Monotonicity:}
\begin{equation*}
2\langle A v - A w, v-w\rangle + \|B v-B w\|_{L_2(U,H)}^2 \leq K|v-w|^2\,\, \textrm{ for all }\,\, v,w \in V.
\end{equation*}
{\em Coercivity:} there is $\mu > 0$ such that
\begin{equation*}
2\langle A_1 v, v \rangle + \|B_1 v\|_{L_2(U,H)}^2 \leq - \mu \|v\|_{V_1}^2 + K(1+|v|^2)\,\, \textrm{ for all }\,\, v\in V_1.
\end{equation*}
and
\begin{equation*}
2\langle A_2 v, v \rangle + \|B_2 v\|_{L_2(U,H)}^2 \leq K(1+|v|^2)\,\, \textrm{ for all }\,\, v\in V_2.
\end{equation*}
{\em Growth:}
\begin{equation*}
\|A_1 v\|_{V_1^*}^2 \leq K(1+\|v\|_{V_1}^2)\,\, \textrm{ for all }\,\, v\in V_1
\end{equation*}
and
\begin{equation*}
\|A_2 v\|_{V_2^*}^{p^*} \leq K(1+\|v\|_{V_2}^p)\,\, \textrm{ for all }\,\, v\in V_2.
\end{equation*}
as well as
\begin{equation*}
\|Bv\|_{L_2(U,H)}^2 \leq K(1+|v|^2) \,\, \textrm{ for all }\,\, v\in H.
\end{equation*}
{\em Hemicontinuity:} for any $v,w$ and $z$ in $V$
\begin{equation*}
\langle A(v+\epsilon w),z\rangle \to \langle Av,z \rangle \,\,\textrm{ as }\,\, \epsilon \to 0.
\end{equation*}
\end{assumption}
We now define what is meant by solution of~\eqref{eq:see}.
\begin{definition}[Solution]
\label{def:soln}
Let $u_0$ be an $\mathcal{F}_0$-measurable $H$-valued random variable.
We say that a continuous, $H$-valued and $(\mathcal{F}_t)_{t\in [0,T]}$-adapted process $u$ is a solution to~\eqref{eq:see}
if $u$ is $dt \times \P$ almost everywhere $V$-valued,
if $u\in \mathcal{L}^2(V_1)\cap \mathcal{L}^p(V_2)$
and if for every $t\in [0,T]$ and every $v\in V$, almost surely,
\begin{equation*}
( u(t),v ) = ( u_0, v ) + \int_0^t \langle A u(s), v \rangle ds + \int_0^t (v,Bu(s) dW(s)).
\end{equation*}
\end{definition}
To the best knowledge of the authors, existence and uniqueness has not been proved for this class of stochastic evolution equations.
Pardoux~\cite{pardoux:thesis} considers the situation where the stochastic evolution equation is governed by a sum of monotone, coercive and hemicontinuous operators satisfying certain growth condition.
However the operator $A_2$ in our case only satisfies a type of ``degenerate'' coercivity condition.
Hence the existence theorem from Pardoux~\cite{pardoux:thesis} does not apply.
We prove that a solution to~\eqref{eq:see} must be unique
in Theorem~\ref{thm:uniqueness} and
we prove existence of the solution
in Theorem~\ref{thm:main}.
\begin{theorem}[Uniqueness]
\label{thm:uniqueness}
The solution of~\eqref{eq:see}, specified by Definition~\ref{def:soln}, is unique, provided that the Growth and Monotonicity conditions in Assumption~\ref{ass:operators} are satisfied.
\end{theorem}
We prove Theorem~\ref{thm:uniqueness} in Section~\ref{sec:convergence}.
Let us now describe the discretization scheme for the stochastic evolution
equation~\eqref{eq:see2}.
For the space discretization let $(V_m)_{m \in \mathbb{N}}$ be a Galerkin scheme for $V$.
To be precise we assume that $V_m\subseteq V$
are finite dimensional spaces with the dimension of $V_m$ equal to $m$.
We further assume that $V_m \subseteq V_{m+1}$ for all $m\in \mathbb{N}$ and that
\begin{equation*}
\lim_{m\to \infty} \inf\{\|v-\varphi\|\, :\, \varphi \in V_m\}=0\,\,\,\, \forall v\in V.
\end{equation*}
(this is referred to as the limited completeness of the Galerkin scheme).
We will need the following projection operators.
\begin{assumption}
\label{ass:projection}
For any $m\in \mathbb{N}$ let $\Pi_m:V^* \to V_m$ satisfy the following:
\begin{enumerate}
\item For any $v \in V_m$, $\Pi_m v = v$.
\item If $f\in V^*$ and $v\in V$ then $\langle f, \Pi_m v \rangle = \langle v, \Pi_m f \rangle$.
\item If $g,h \in H$ then $(\Pi_m g,h) = (\Pi_m h, g)$
and $|\Pi_m h| \leq |h|$.
\item There is a constant, depending on $m$ and denoted by $\mathfrak{c}(m)$, such that
\begin{equation*}
|\Pi_m f|^2 \leq \mathfrak{c}(m)\|f\|_{V^*}^2\,\, \textrm{ for all }\,\, f\in V^*.
\end{equation*}
\end{enumerate}
\end{assumption}
In applications this assumption is easily satisfied.
In particular if $\{\varphi_j \in V\,\, : \,\, j = 1,2,\ldots \}$
is an orthonormal basis in $H$ then taking
$V_m := \textrm{span}\{\varphi_1,\ldots,\varphi_m\}$
is a Galerkin scheme for $V$.
Taking $\Pi_m f := \sum_{j=1}^m \langle f, \varphi_j \rangle \varphi_j$
satisfies the first three conditions in Assumption~\ref{ass:projection}.
Moreover, the following holds
\begin{equation*}
|\Pi_m f|^2 = \bigg| \sum_{j=1}^m \langle f, \varphi_j\rangle \varphi_j \bigg|^2 = \sum_{j=1}^m \langle f, \varphi_j\rangle^2 \leq \|f\|_{V^*}^2 \sum_{j=1}^m \|\varphi_j\|_V^2 = \mathfrak{c}(m) \|f\|_{V^*}^2,
\end{equation*}
where $\mathfrak{c}(m) := \sum_{j=1}^m \|\varphi_j\|_V^2$.
Thus the fourth condition in Assumption~\ref{ass:projection} is also satisfied.
Let $\{\chi_i\}_{i\in \mathbb{N}}$ be an orthonormal basis of $U$.
Fix $k \in \mathbb{N}$ and define
\begin{equation*}
W_k(t) := \sum_{j=1}^{k} (W(t),\chi_j)_U \chi_j.
\end{equation*}
For the time discretization take $n\in \mathbb{N}$, let $\tau_n := T/n$ and define the grid points on an equidistant grid as $t^n_i := \tau_n i$, $i=0,1,\ldots,n$.
Further consider some sequence $((n_\ell, m_\ell, k_\ell))_{\ell \in \mathbb{N}}$ such that $n_\ell, m_\ell$ and $k_\ell$ all go to infinity as $\ell \to \infty$.
Let $c$ denote a generic positive constant that may depend on $T$,
on the constants arising in the continuous embeddings $V_i \hookrightarrow H \hookrightarrow V_i^*$, $i=1,2$
and on the constants arising in Assumptions~\ref{ass:operators}
and~\ref{ass:interpolation} but that is always independent of the discretization parameters $m$, $k$ and $n$.
Define $\kappa_{n_\ell}(t) = t^{n_\ell}_i$ if $t\in [t^{n_\ell}_i, t^{n_\ell}_{i+1})$ for $i=0,\ldots,n_\ell-1$ and $\kappa_{n_\ell}(T) = T$.
Fix some $\ell \in \mathbb{N}$ (and hence $m_\ell$, $n_\ell$ and $k_\ell$).
Let $u_\ell(0)$ be a $V_{m_\ell}$ valued $\mathcal{F}_0$-measurable approximation of $u_0$.
For example we can take $u_\ell(0) := \Pi_{m_\ell} u_0$ but other approximations are possible.
For $t>0$ we define a process $u_\ell$ by
\begin{equation}
\label{eq:disc1}
\begin{split}
u_\ell(t) = u_\ell(0) & + \int_0^t \Pi_{m_\ell} \left[A_1 u_\ell(\kappa_{n_\ell}(s))
+ A_{2,\ell} u_\ell(\kappa_{n_\ell}(s)) \right] ds\\
& + \int_0^t \Pi_{m_\ell} B u_\ell(\kappa_{n_\ell}(s))dW_{k_\ell}(s),
\end{split}
\end{equation}
where we use the ``tamed'' operator $A_{2,\ell}$ defined by
\begin{equation}
\label{eq:def:taming}
A_{2,\ell} v := \frac{1}{1+{n_\ell}^{-1/2}|\Pi_{m_\ell} A_2 v|}A_2 v
\end{equation}
for any $v \in V_2$.
We will use the following notation:
$\bar{u}_\ell(t) := u_\ell(\kappa_{n_\ell}(t))$
and $a_\ell(v) := \Pi_{m_\ell}[A_1 v + A_{2,\ell} v]$.
Then~\eqref{eq:disc1} is equivalent to
\begin{equation}
\label{eq:disc2}
\begin{split}
u_\ell(t) = u_\ell(0) & + \int_0^t a_\ell(\bar{u}_\ell(s))
ds + \int_0^t \Pi_{m_\ell} B \bar{u}_\ell(s) dW_{k_\ell}(s).
\end{split}
\end{equation}
In particular at the time-grid points we have, for $i=0,1,\ldots,n_\ell-1$,
\begin{equation*}
\begin{split}
u_\ell(t^{n_\ell}_{i+1}) = u_\ell(t^{n_\ell}_i) & + a_\ell(u_\ell(t^{n_\ell}_i))\tau_{n_\ell} + \Pi_{m_\ell} B u_\ell(t^{n_\ell}_i) \Delta W_{k_\ell}(t_{i+1}),
\end{split}
\end{equation*}
where $\Delta W_{k_\ell}(t^{n_\ell}_{i+1}) := W_{k_\ell}(t^{n_\ell}_{i+1})-W_{k_\ell}(t^{n_\ell}_i)$.
This in turn is equivalent to
\begin{equation*}
\begin{split}
\frac{u_\ell(t^{n_\ell}_{i+1}) - u_\ell(t^{n_\ell}_i)}{\tau_{n_\ell}} = & a_\ell(u_\ell(t^{n_\ell}_i)) + \Pi_{m_\ell} Bu_\ell(t^{n_\ell}_i) \frac{\Delta W_{k_\ell}(t^{n_\ell}_{i+1})}{\tau_{n_\ell}}.
\end{split}
\end{equation*}
We list below the properies which are satisfied by the tamed operator $A_{2,\ell}$.
These are consequences of its structure and the assumed properties of $A_2$.
For brevity let, for any $v\in V_2$,
\begin{equation}
\label{eq:def_of_T_ell}
T_\ell(v) := \frac{1}{1+n_\ell^{-1/2}|\Pi_{m_\ell} A_2 v|}.
\end{equation}
Then for any $v\in V_2$,
\begin{equation}
\label{eq:taming_less_n}
|\Pi_{m_\ell} A_{2,\ell} v| = T_\ell(v)|\Pi_{m_\ell} A_2 v| \leq n_\ell^{1/2}
\end{equation}
and also, using the Growth assumption on $A_2$,
\begin{equation}
\label{eq:taming_less_original}
\|A_{2,\ell} v\|_{V_2^*}^{p^*} = T_\ell(v)^{p^*}\|A_2 v\|_{V_2^*}^{p^*} \leq K(1+\|v\|_{V_2}^p).
\end{equation}
Furthermore, using the Coercivity assumption on $A_2$, we note that for all $v \in V_{m_\ell}$ we have
\begin{equation}
\label{eq:taming_retains_weak_coerc}
2\langle A_{2,\ell} v, v \rangle = 2T_\ell(v)\langle A_2 v,v \rangle \leq K(1+|v|^2).
\end{equation}
Thus the weaker coercivity assumption that has been made about $A_2$ is retained.
Consider, for a moment, that $A_2$ satisfies the ``usual'' coercivity condition
\begin{equation*}
2\langle A_2 v, v \rangle + \|B_2 v\|_{L_2(U,H)}^2 \leq - \mu \|v\|_{V_2}^p + K(1+|v|^2)\,\, \textrm{ for all }\,\, v\in V_2.
\end{equation*}
We see that in this case the best coercivity we can get from this for $A_{2,\ell}$ is again only~\eqref{eq:taming_retains_weak_coerc}.
Hence to obtain the necessary a priori estimates
we will need an interpolation inequality between $V_2$ and $V_1$ with $H$.
\begin{assumption}
\label{ass:interpolation}
There are constants $\lambda \in [0,2/p)$ and
$\Lambda>0$ such that for any $v\in V$
\begin{equation*}
\|v\|_{V_2} \leq \Lambda \|v\|_{V_1}^\lambda |v|^{1-\lambda}.
\end{equation*}
\end{assumption}
Note that in order to overcome the difficulty with coercivity it would suffice to have Assumption~\ref{ass:interpolation} satisfied with $\lambda \in [0,2/p]$.
However monotonicity of $A_2$ is not preserved by taming.
To overcome this we will need to show that $A_{2,\ell} \bar{u}_\ell - A_2 \bar{u}_\ell \to 0$ in $\mathcal{L}^{p*}(V_2^*)$.
To achieve this we use the fact that $\lambda \in [0,2/p)$ in Lemma~\ref{lemma:strong_conv_of_tam_correction} and the following observation:
Assumption~\ref{ass:interpolation} implies that there is $\eta > 0$ such that
\begin{equation*}
\|v\|_{V_2}^{p(1+\eta)} \leq c \|v\|_{V_1}^2 |v|^\rho,
\end{equation*}
where $\rho := (1-\lambda)p(1+\eta)$.
From this it follows that if $v\in L^2(\Omega; L^2(0,T;V_1))$ and $v\in L^{2\rho}(\Omega; L^\infty(0,T;H))$ then
\begin{equation*}
\mathbb{E}\int_0^T \|v(s)\|_{V_2}^{p(1+\eta)} ds \leq c\Bigg[ \mathbb{E} \sup_{s\leq t}|v(s)|^{2\rho} + \mathbb{E}\bigg(\int_0^T \|v(s)\|_{V_1}^2 ds\bigg)^2 \Bigg].
\end{equation*}
Thus we see that Assumption~\ref{ass:interpolation} allows us to control the approximate solution in the $L^{p(1+\eta)}((0,T)\times \Omega; V_2)$ norm,
provided that we can control the approximate solution in the norms of $L^2(\Omega; L^2(0,T;V_1))$
and $L^{2\rho}(\Omega; L^\infty(0,T;H))$.
Let us take $q_0 := \max(4,2\rho)$.
Now we can state the main result of this paper.
\begin{theorem}
\label{thm:main}
Let Assumptions~\ref{ass:operators},~\ref{ass:projection} and~\ref{ass:interpolation} be satisfied.
Let $u_0 \in L^{q_0}(\Omega; H)$ and let $u_\ell(0) \to u_0$ in $L^{q_0}(\Omega; H)$.
Assume that $\tfrac{\mathfrak{c}(m_\ell)}{n_{\ell}} \to 0$ as $\ell \to \infty$.
Then there exists a unique solution $u$ to~\eqref{eq:see} and
$\bar{u}_\ell \rightharpoonup u$ in $\mathcal{L}^2(V_1)$ and in $\mathcal{L}^p(V_2)$ and $u_\ell(T) \to u(T)$ in $L^2(\Omega;H)$ as $\ell \to \infty$.
\end{theorem}
In Section~\ref{sec:examples} we provide examples of stochastic partial
differential equations where Theorem~\ref{thm:main} can be applied.
We also compute $\mathfrak{c}(m)$ in case of the spectral Galerkin method to
make the implications of the space-time coupling constraint more explicit.
The crucial point is that the requirement is no more onerous than in the case
of equations with operators growing at most linearly.
\section{A priori estimates}
\label{sec:apriori_est}
We start with an important observation that allows us to use standard results on bounds of stochastic integrals driven by finite dimensional Wiener processes.
\begin{remark}
\label{remark:n_bound}
Recall that $(\chi_j)_{j\in \mathbb{N}}$ is an orthonormal basis in $U$.
Moreover recall that $\bar{u}_\ell(t) := u_\ell(\kappa_{n_\ell}(t))$
and that $a_\ell(v) := \Pi_{m_\ell}[A_1 v + A_{2,\ell} v]$.
For each $j\in \mathbb{N}$ a Wiener processes $\mathcal{W}_j$
is obtained by taking $\mathcal{W}_j(t) := (W(t),\chi_j)_U$.
If $i\neq j$ then $\mathcal{W}_i$ and $\mathcal{W}_j$ are independent.
Furthermore~\eqref{eq:disc2} is equivalent to
\begin{equation*}
u_\ell(t) = u_\ell(0) + \int_0^t a_\ell(\bar{u}_\ell(s))ds
+ \sum_{j=1}^{k_\ell} \int_0^t \Pi_{m_\ell} B \bar{u}_\ell(s) \chi_j d\mathcal{W}_j(s).
\end{equation*}
Fix $\ell\in \mathbb{N}$ (and thus $k_\ell, m_\ell$ and $n_\ell$ are alse fixed).
Then using the Growth assumptions on $A_1$ and $B$, Assumption~\ref{ass:projection}
as well as~\eqref{eq:taming_less_n} one observes that
$|a_\ell(v)|^2 \leq 2\mathfrak{c}(m)K(1+|v|^2) + 2n_\ell$ and
$|\Pi_{m_\ell}B v\chi_j|^2 \leq K(1+|v|^2)$.
Hence one knows that, for $q\geq 1$,
\begin{equation*}
\mathbb{E}\sup_{0\leq t \leq T} |u_\ell(t)|^q < \infty,
\end{equation*}
provided that $\mathbb{E}|u_\ell(0)|^q < \infty$.
Clearly, at this point, one cannot claim that this bound is independent of $\ell$.
\end{remark}
One applies It\^o's formula to~\eqref{eq:disc2} to obtain
\begin{equation*}
\begin{split}
|u_\ell(t)|^2 = |u_\ell(0)|^2 & + \int_0^t \!\! \bigg[2 \langle a_\ell(\bar{u}_\ell(s)),u_\ell(s)\rangle + \sum_{j=1}^{k_\ell}|\Pi_{m_\ell} B \bar{u}_\ell(s)\chi_j|^2\bigg] ds\\
& + \int_0^t 2 (B \bar{u}_\ell(s), u_\ell(s) dW_{k_\ell}(s))\,,
\end{split}
\end{equation*}
which can be rewritten as
\begin{equation}
\label{eq:apriori_start}
\begin{split}
|u_\ell(t)|^2 = |u_\ell(0)|^2 & + \int_0^t \!\! \bigg[2 \langle a_\ell(\bar{u}_\ell(s)),\bar{u}_\ell(s)\rangle + \sum_{j=1}^{k_\ell}|\Pi_{m_\ell} B \bar{u}_\ell(s)\chi_j|^2 \bigg] ds\\
& + 2\int_0^t \!\! \langle a_\ell(\bar{u}_\ell(s)),u_\ell(s)-\bar{u}_\ell(s)\rangle ds\\
& + \int_0^t 2 (B \bar{u}_\ell(s), u_\ell(s) dW_{k_\ell}(s))\, ,
\end{split}
\end{equation}
in order to apply the coercivity assumption so as to obtain the a priori estimates for the discretized equation.
First we concentrate on the term that arises from the ``correction'' that one has to make to use the coercivity assumption due to the use of an explicit scheme.
\begin{lemma}
\label{lemma:corr_term_est}
Let the Growth condition in Assumption~\ref{ass:operators} be satisfied.
Let Assumption~\ref{ass:projection} hold.
Let $q\geq 1$ be given.
Then
\begin{equation}
\label{eq:lemma_corr_term_est1}
\begin{split}
\mathbb{E} \bigg( & \frac{1}{\tau_{n_\ell}} \int_0^t |u_\ell(s)-\bar{u}_\ell(s)|^2 ds \bigg)^q\\
& \leq c_{T,q} \Bigg(1+ (\mathfrak{c}(m)\tau_{n_\ell})^q \mathbb{E}\bigg(\int_0^t \|\bar{u}_\ell(s)\|_{V_1}^2 ds \bigg)^q + \mathbb{E}\int_0^t |\bar{u}_\ell(s)|^{2q}ds \Bigg)
\end{split}
\end{equation}
and
\begin{equation}
\label{eq:lemma_corr_term_est2}
\begin{split}
\mathbb{E} & \bigg( \int_0^t |a_\ell(\bar{u}_\ell(s))||u_\ell(s)-\bar{u}_\ell(s)|ds\bigg)^q \\
& \leq c_{T,q}\Bigg(1+ (\mathfrak{c}(m)\tau_{n_\ell})^q \mathbb{E}\bigg(\int_0^t \|\bar{u}_\ell(s)\|_{V_1}^2 ds \bigg)^q + \mathbb{E}\int_0^t |\bar{u}_\ell(s)|^{2q}ds \Bigg).
\end{split}
\end{equation}
\end{lemma}
\begin{proof}
From~\eqref{eq:disc1} it is clear that
\begin{equation*}
\begin{split}
& I_{1,\ell}(t) := \mathbb{E} \bigg( \frac{1}{\tau_{n_\ell}} \int_0^t |u_\ell(s)-\bar{u}_\ell(s)|^2 ds \bigg)^q \\
& = \mathbb{E}\bigg( \int_0^t \frac{1}{\tau_{n_\ell}}\bigg| \int_{\kappa_{n_\ell}(s)}^s a_\ell(\bar{u}_\ell(r)) dr + \int_{\kappa_{n_\ell}(s)}^s \Pi_{m_\ell} B \bar{u}_\ell(r) dW_{k_\ell}(r) \bigg|^2 ds \bigg)^q\\
& \leq 2^q \mathbb{E}\bigg( \frac{1}{\tau_{n_\ell}}\int_0^t \bigg| \int_{\kappa_{n_\ell}(s)}^s \!\!\! a_\ell(\bar{u}_\ell(r)) dr \bigg|^2
+ \bigg|\int_{\kappa_{n_\ell}(s)}^s \Pi_{m_\ell} B \bar{u}_\ell(r) dW_{k_\ell}(r) \bigg|^2 \!\!ds \bigg)^q\!\!.
\end{split}
\end{equation*}
Applying H\"older's inequality yields
\begin{equation*}
\begin{split}
I_{1,\ell}(t) & \leq 2^q \mathbb{E}\Bigg( \frac{1}{\tau_{n_\ell}}\int_0^t\Bigg[ (s-\kappa_{n_\ell}(s)) \int_{\kappa_{n_\ell}(s)}^s |a_\ell(\bar{u}_\ell(r))|^2 dr \\
& \quad\quad\quad \quad\quad\quad + \bigg|\int_{\kappa_{n_\ell}(s)}^s \Pi_{m_\ell} B \bar{u}_\ell(r) dW_{k_\ell}(r) \bigg|^2 \Bigg]ds \Bigg)^q\\
& \leq c_q \mathbb{E}\bigg( \frac{1}{\tau_{n_\ell}}\int_0^t \tau_{n_\ell}^2 |a_\ell(\bar{u}_\ell(s))|^2 ds \bigg)^q \\
& \quad\quad\quad \quad\quad\quad + c_q\mathbb{E}\bigg( \frac{1}{\tau_{n_\ell}}\int_0^t \bigg|\int_{\kappa_{n_\ell}(s)}^s \Pi_{m_\ell} B \bar{u}_\ell(r) dW_{k_\ell}(r) \bigg|^2 ds \bigg)^q.
\end{split}
\end{equation*}
Using Assumption~\ref{ass:projection} and~\eqref{eq:taming_less_n} one obtains
\begin{equation}
\label{eq:corr_term_est_2}
\begin{split}
& I_{1,\ell}(t) \leq c_q \mathbb{E}\bigg(\int_0^t \tau_{n_\ell} [ 2\mathfrak{c}(m)\|A_1 \bar{u}_\ell(s)\|_{V_1^*}^2 + 2n_\ell] ds \bigg)^q \\
& + c_q\mathbb{E}\bigg( \int_0^t \frac{1}{\tau_{n_\ell}} \bigg|\int_{\kappa_{n_\ell}(s)}^s \Pi_{m_\ell} B \bar{u}_\ell(r) dW_{k_\ell}(r) \bigg|^2 ds \bigg)^q
:= I_{2,\ell}(t) + I_{3,\ell}(t).
\end{split}
\end{equation}
The Growth assumption on $A_1$ implies that
\begin{equation*}
I_{2,\ell}(t) \leq c_{T,q}\bigg(1 + (\mathfrak{c}(m)\tau_{n_\ell})^q\mathbb{E}\bigg(\int_0^t \|\bar{u}_\ell(s)\|_{V_1}^2 ds \bigg)^q \bigg).
\end{equation*}
Using H\"older's inequality leads to
\begin{equation*}
I_{3,\ell}(t) \leq c_{T,q} \mathbb{E}\int_0^t \frac{1}{\tau_{n_\ell}^q} \bigg|\int_{\kappa_{n_\ell}(s)}^s \Pi_{m_\ell} B \bar{u}_\ell(r) dW_{k_\ell}(r) \bigg|^{2q} ds.
\end{equation*}
Due to Remark~\ref{remark:n_bound} and the Growth assumption on $B$ one observes that
\begin{equation*}
\begin{split}
I_{3,\ell}(t) & \leq c_{T,q} \int_0^t \frac{1}{\tau_{n_\ell}^q} \mathbb{E} \bigg|\int_{\kappa_{n_\ell}(s)}^s \|B \bar{u}_\ell(r)\|_{L_2(U,H)}^2 dr \bigg|^q ds \\
& \leq c_{T,q}\mathbb{E} \int_0^t \frac{1}{\tau_{n_\ell}^q} (s-\kappa_{n_\ell}(s))^q \|B \bar{u}_\ell(s)\|_{L_2(U,H)}^{2q} ds\\
& \leq c_{T,q}\bigg(1+\mathbb{E}\int_0^t |\bar{u}_\ell(s)|^{2q} ds \bigg).
\end{split}
\end{equation*}
This implies~\eqref{eq:lemma_corr_term_est1}.
Moreover Assumption~\ref{ass:projection} and~\eqref{eq:taming_less_n} imply that
\begin{equation*}
\begin{split}
I_\ell(t) & := \mathbb{E}\bigg(\int_0^t |a_\ell(\bar{u}_\ell(s))||u_\ell(s)-\bar{u}_\ell(s)|ds\bigg)^q \\
& \leq 2^q \mathbb{E}\bigg(\int_0^t \tau_{n_\ell}|a_\ell(\bar{u}_\ell(s))|^2 + \frac{1}{\tau_{n_\ell}}|u_\ell(s)-\bar{u}_\ell(s)|^2 ds\bigg)^q\\
& \leq c_q \mathbb{E}\bigg(\int_0^t [\mathfrak{c}(m)\tau_{n_\ell}\|A_1 \bar{u}_\ell(s)\|_{V_1^*}^2 + \tau_{n_\ell} n_\ell]ds \bigg)^q \\
& \quad\quad\quad \quad\quad\quad + c_q \mathbb{E}\bigg(\frac{1}{\tau_{n_\ell}} \int_0^t |u_\ell(s)-\bar{u}_\ell(s)|^2 ds\bigg)^q.
\end{split}
\end{equation*}
Applying the Growth assumption on $A_1$ yields
\begin{equation}
\label{eq:corr_term_est1}
\begin{split}
I_\ell(t) & \leq c_{T,q}\bigg(1 + (\mathfrak{c}(m)\tau_{n_\ell})^q \mathbb{E}\bigg(\int_0^t \|\bar{u}_\ell(s)\|_{V_1}^2 ds \bigg)^q \bigg) \\
& \quad\quad\quad \quad\quad\quad + c_q \mathbb{E}\bigg(\frac{1}{\tau_{n_\ell}} \int_0^t |u_\ell(s)-\bar{u}_\ell(s)|^2 ds\bigg)^q.
\end{split}
\end{equation}
Using~\eqref{eq:lemma_corr_term_est1} in~\eqref{eq:corr_term_est1} concludes the proof.
\end{proof}
\begin{theorem}[A priori estimate]
\label{thm:apriori2}
Let the Coercivity and Growth conditions in Assumption~\ref{ass:operators} hold.
Let Assumption~\ref{ass:projection} be satisfied.
Let $q \geq 1$ be given
and assume that $\mathbb{E}|u_\ell(0)|^{2q} < c$ and that $u_\ell(0)$ is $\mathcal{F}_0$-measurable.
There is $\epsilon \in (0,\infty)$ such that for all
$\ell \in \{\ell' \in \mathbb{N}: \mathfrak{c}(m_{\ell'})\tau_{n_{\ell'}} < \epsilon\}$ we have,
for any $t\in [0,T]$,
\begin{equation*}
\mathbb{E} \sup_{s\in [0,t]}|u_\ell(s)|^{2q} + \mu^{q} \mathbb{E} \bigg(\int_0^t \|\bar{u}_\ell(s)\|_{V_1}^2 ds \bigg)^{q}
\leq c\left(1+ \mathbb{E}|u_\ell(0)|^{2q}\right).
\end{equation*}
\end{theorem}
\begin{proof}
Applying the Coercivity assumption in~\eqref{eq:apriori_start},
raising to power $q \geq 1$,
taking the supremum over $s\leq t$ and taking the expectation yields
\begin{equation}
\label{eq:higher_moments_proof1}
\begin{split}
& \mathbb{E} \sup_{s\leq t} |u_\ell(s)|^{2q} + \mu^{q} \mathbb{E} \bigg( \int_0^t \|\bar{u}_\ell(s)\|_{V_1}^2 ds \bigg)^{q} \leq c_{T,q} \bigg[1 + \mathbb{E}|u_\ell(0)|^{2q} \\
& + \mathbb{E} \bigg( \int_0^t |\bar{u}_\ell(s)|^2 ds \bigg)^{q}
+ \mathbb{E}\bigg(\int_0^t |a_\ell(\bar{u}_\ell(s))||u_\ell(s)-\bar{u}_\ell(s)|ds\bigg)^{q}\\
& + \mathbb{E} \sup_{s \leq t}\bigg|\int_0^s (u_\ell(s), B\bar{u}_\ell(s) dW_{k_\ell}(s))\bigg|^{q} \bigg].
\end{split}
\end{equation}
Using Lemma~\ref{lemma:corr_term_est} in~\eqref{eq:higher_moments_proof1} results in
\begin{equation}
\label{eq:higher_moments_proof2}
\begin{split}
& \mathbb{E} \sup_{s\leq t} |u_\ell(s)|^{2q} + \frac{\mu^{q}}{2} \mathbb{E} \bigg( \int_0^t \|\bar{u}_\ell(s)\|_{V_1}^2 ds \bigg)^{q}\\
& \leq c_{T,q} \bigg[1 + \mathbb{E}|u_\ell(0)|^{2q} + \mathbb{E} \bigg( \int_0^t |\bar{u}_\ell(s)|^2 ds \bigg)^{q} \\
& \quad + \mathbb{E} \int_0^t |\bar{u}_\ell(s)|^{2q} ds
+ \mathbb{E} \sup_{s \leq t}\bigg(\int_0^s (u_\ell(r), B\bar{u}_\ell(r) dW_{k_\ell}(r))\bigg)^{q}\bigg].
\end{split}
\end{equation}
Using Burkholder--Davis--Gundy inequality one obtains
\begin{equation*}
\begin{split}
I_\ell & := c_{T,q}\mathbb{E} \sup_{s \leq t}\bigg|\int_0^s (u_\ell(r), B\bar{u}_\ell(r) dW_{k_\ell}(r))\bigg|^{q} \\
& \leq c_{T,q} \mathbb{E} \bigg|\int_0^t |u_\ell(s)|^2 \|B\bar{u}_\ell(s)\|_{L_2(U,H)}^2 ds\bigg|^{q/2}\\
& \leq c_{T,q} \mathbb{E} \Bigg[\sup_{s\leq t}|u_\ell(s)|^{q}\bigg(\int_0^t \|B\bar{u}_\ell(s)\|_{L_2(U,H)}^2ds\bigg)^{q/2} \Bigg].
\end{split}
\end{equation*}
Young's inequality and the Growth assumption on $B$ imply that
\begin{equation*}
\begin{split}
I_\ell & \leq \frac{1}{2}\mathbb{E} \sup_{s\leq t}|u_\ell(s)|^{2q} + c \mathbb{E} \bigg(\int_0^t \|B\bar{u}_\ell(s)\|_{L_2(U,H)}^2ds\bigg)^{q}\\
& \leq \frac{1}{2}\mathbb{E} \sup_{s\leq t}|u_\ell(s)|^{2q} + c\bigg(1+\int_0^t \mathbb{E} \sup_{r\leq s}|u_\ell(r)|^{2q} ds \bigg).
\end{split}
\end{equation*}
Applying this in~\eqref{eq:higher_moments_proof2} leads to
\begin{equation*}
\begin{split}
\frac{1}{2}\mathbb{E} \sup_{s\leq t} |u_\ell(s)|^{2q} + \frac{\mu^{q}}{2} \mathbb{E} \bigg( \int_0^t \|\bar{u}_\ell(s)\|_{V_1}^2 ds \bigg)^{q}
& \leq c \bigg[1 + \mathbb{E}|u_\ell(0)|^{2q} \\
& + \int_0^t \mathbb{E} \sup_{r\leq s}|u_\ell(r)|^{2q} ds\bigg].
\end{split}
\end{equation*}
Application of Gronwall's lemma yields
\begin{equation}
\label{eq:higher_moments_proof3}
\mathbb{E} \sup_{s\in [0,t]}|u_\ell(s)|^{2q} + \mu^{q} \mathbb{E} \bigg(\int_0^t \|\bar{u}_\ell(s)\|_{V_1}^2 ds \bigg)^{q}
\leq c \left(1+\mathbb{E}|u_\ell(0)|^{2q}\right).
\end{equation}
\end{proof}
Now we use Theorem~\ref{thm:apriori2} and Assumption~\ref{ass:interpolation} to obtain the remaining required estimates.
\begin{corollary}[Remaining a priori estimates]
\label{corollary:remaining_apriori}
Let the Growth and Coercivity conditions in Assumption~\ref{ass:operators} be satisfied.
Let Assumptions~\ref{ass:projection} and~\ref{ass:interpolation} hold.
Let $u_\ell(0)$ be bounded in $L^{q_0}(\Omega; H)$, uniformly with respect to $\ell$.
There is $\epsilon \in (0,\infty)$ such that for all
$\ell \in \{\ell' \in \mathbb{N}: \mathfrak{c}({m_\ell'}\tau_{n_{\ell'}} < \epsilon\}$ we have
\begin{equation}
\label{eq:remaining_apriori1}
\mathbb{E}\int_0^T \|A_1 \bar{u}_\ell\|_{V_1^*}^2 ds \leq c,\quad \mathbb{E}\int_0^T \|B \bar{u}_\ell\|_{L_2(U,H)}^2 ds \leq c.
\end{equation}
Furthermore
\begin{equation}
\label{eq:remaining_apriori4}
\mathbb{E}\int_0^T |u_\ell(s) - \bar{u}_\ell(s)|^2 ds \leq c \tau_{n_\ell}.
\end{equation}
Finally, for some $\eta > 0$,
\begin{equation}
\label{eq:remaining_apriori2}
\mathbb{E}\int_0^T \|\bar{u}_\ell\|_{V_2}^{p(1+\eta)} ds \leq c
\end{equation}
and
\begin{equation}
\label{eq:remaining_apriori3}
\mathbb{E}\int_0^T \|A_2 \bar{u}_\ell\|_{V_2^*}^{p^*(1+\eta)} ds \leq c,\quad \mathbb{E}\int_0^T \|A_{2,\ell} \bar{u}_\ell\|_{V_2^*}^{p^*(1+\eta)} ds \leq c.
\end{equation}
\end{corollary}
\begin{proof}
Inequality~\eqref{eq:remaining_apriori1} follows directly from the Growth assumptions on $A_1$ and $B$ and from Theorem~\ref{thm:apriori2} with $q=1$.
Using~\eqref{eq:lemma_corr_term_est1},
together with Theorem~\ref{thm:apriori2} with $q=1$,
yields~\eqref{eq:remaining_apriori4}.
Since $u_\ell(0)$ is assumed to be bounded in $L^{q_0}(\Omega; H)$,
uniformly in $\ell$,
one can conclude, using Theorem~\ref{thm:apriori2}, that
\begin{equation*}
\mathbb{E} \sup_{s\in [0,t]}|u_\ell(s)|^{2\rho} \leq c \,\,\textrm{ and }\,\, \mathbb{E} \bigg(\int_0^t \|\bar{u}_\ell(s)\|_{V_1}^2 ds \bigg)^2 \leq c.
\end{equation*}
This, together with Assumption~\ref{ass:interpolation}, yields~\eqref{eq:remaining_apriori2}.
Finally,~\eqref{eq:remaining_apriori2}, the assumption on the growth of $A_2$ and~\eqref{eq:taming_less_original} lead to~\eqref{eq:remaining_apriori3}.
\end{proof}
\section{Convergence}
\label{sec:convergence}
Having obtained the required a priori estimates we can use compactness arguments to extract weakly convergent subsequences of the approximation.
\begin{lemma}
\label{lemma:weak_limits_from_compactness}
Let the Growth and Coercivity conditions in Assumption~\ref{ass:operators} hold.
Let Assumptions~\ref{ass:projection} and~\ref{ass:interpolation} be satisfied.
Let $u_\ell(0) \to u_0$ in $L^{q_0}(\Omega; H)$.
Let $\tfrac{\mathfrak{c}(m_\ell)}{n_\ell} \to 0$ as $\ell \to \infty$.
Then there is a subsequence of the sequence $\ell$, which we denote $\ell'$, and $u \in \mathcal{L}^2(V_1) \cap \mathcal{L}^p(V_2)$ such that, as $\ell' \to \infty$,
\begin{equation*}
\bar{u}_{\ell'} \rightharpoonup u\,\, \textrm{ in }\,\, \mathcal{L}^2(V_1)\,\, \textrm{ and in }\,\, \mathcal{L}^p(V_2).
\end{equation*}
Furthermore there are $a_1^\infty \in \mathcal{L}^2(V_1^*)$, $a_2^\infty \in \mathcal{L}^{p^*}(V_2^*)$ and $b^\infty \in \mathcal{L}^2(L_2(U,H))$ such that, as $\ell' \to \infty$,
\begin{equation*}
A_1\bar{u}_{\ell'} \rightharpoonup a_1^\infty \,\, \textrm{ in }\,\, \mathcal{L}^2(V_1^*),\,\, A_{2,\ell'}\bar{u}_{\ell'} \rightharpoonup a_2^\infty \,\, \textrm{ in }\,\, \mathcal{L}^{p^*}(V_2^*)
\end{equation*}
and
\begin{equation*}
B\bar{u}_{\ell'} \rightharpoonup b^\infty\,\, \textrm{ in } \mathcal{L}^2({L_2(U,H)}).
\end{equation*}
Finally, there is $\xi \in L^{q_0}(\Omega; H)$ such that $\bar{u}_{\ell'}(T) = u_{\ell'}(T) \rightharpoonup \xi$ in $L^{q_0}(\Omega; H)$.
\end{lemma}
\begin{proof}
The sequence $(\bar{u}_\ell)$ is bounded in $\mathcal{L}^2(V_1)$ due to Theorem~\ref{thm:apriori2} and in $\mathcal{L}^p(V_2)$ due to Corollary~\ref{corollary:remaining_apriori}.
The sequences $(A_1 \bar{u}_\ell)$, $(A_{2,\ell} \bar{u}_\ell)$ and $(B\bar{u}_\ell)$ are bounded in $\mathcal{L}^2(V_1^*)$, $\mathcal{L}^{p^*}(V_2^*)$ and in
$\mathcal{L}^{2}(L_2(U,H))$ respectively, due to Corollary~\ref{corollary:remaining_apriori}.
Finally the sequence $(u_\ell(T))$ is bounded in $L^{q_0}(\Omega; H)$ due to Theorem~\ref{thm:apriori2}.
Since it is assumed that $V_1$, $V_2$ are reflexive, it follows that $\mathcal{L}^2(V_1)$ and $\mathcal{L}^p(V_2)$ are reflexive.
A bounded sequence in a reflexive Banach space must have a weakly convergent subsequence (see e.g. Br\'ezis~\cite[Theorem 3.18]{brezis:functional}).
Applying this to the sequences in question concludes the proof of the lemma.
\end{proof}
Let $a^\infty := a_1^\infty + a_2^\infty$.
Then $a^\infty \in \mathcal{L}^{p^*}(V^*)$.
Due to Lemma~\ref{lemma:weak_limits_from_compactness} $a_\ell(\bar{u}_\ell) \rightharpoonup a^\infty$ in $\mathcal{L}^{p^*}(V^*)$
as $\ell' \to \infty$, provided that $\tfrac{\mathfrak{c}(m_\ell)}{n_\ell} \to 0$.
The following lemma provides the equation satisfied by the weak limits of the approximations.
\begin{lemma}
\label{lemma:limit_equation}
Let the Growth and Coercivity conditions in Assumption~\ref{ass:operators} hold.
Let Assumptions~\ref{ass:projection} and~\ref{ass:interpolation} be satisfied.
Let $u_\ell(0) \to u_0$ in $L^{q_0}(\Omega; H)$.
Let $\tfrac{\mathfrak{c}(m_\ell)}{n_\ell} \to 0$ as $\ell \to \infty$.
Then there is an $H$-valued adapted continuous process $\tilde{u}$ on $[0,T]$
such that $u=\tilde{u}$ $dt\times \P$-almost everywhere on $(0,T)\times \Omega$.
Furthermore, for almost every $(t,\omega) \in (0,T) \times \Omega$,
\begin{equation}
\label{eq:limit_equation1}
\tilde{u}(t) = u_0 + \int_0^t a^\infty (s) ds + \int_0^t b^\infty(s) dW(s)
\end{equation}
and almost surely
\begin{equation}
\label{eq:limit_equation2}
\tilde{u}(T) = u_0 + \int_0^T a^\infty (s) ds + \int_0^T b^\infty(s) dW(s).
\end{equation}
\end{lemma}
In the rest of this paper we will write $u$ instead of $\tilde{u}$ for notational simplicity.
\begin{proof}
Fix $M\in \mathbb{N}$.
Let $\varphi$ be a $V_M$-valued adapted stochastic process such that $|\varphi(t)| < M$ for all $t\in [0,T]$ and $\omega \in \Omega$.
For $g\in U$ let $\tilde{\Pi}_m g:= \sum_{j=1}^m (\chi_j,g)_U \chi_j$ and note that for any $v\in V$ one has $Bv\tilde{\Pi}_{m} \in L_2(U,H)$.
From~\eqref{eq:disc1} one observes that
\begin{equation}
\label{eq:limit_eq_proof_1}
\begin{split}
(u_{\ell'}(t),\varphi(t)) = (u_{\ell'}(0), & \varphi(t))
+ \bigg\langle \int_0^t a_{\ell'}(\bar{u}_{\ell'}(s)) ds, \varphi(t) \bigg\rangle\\
& + \bigg(\int_0^t B \bar{u}_{\ell'}(s)\tilde{\Pi}_{m_{\ell'}}dW(s) ,\varphi(t)\bigg).
\end{split}
\end{equation}
Let $G:\mathcal{L}^{p^*}(V^*) \to \mathcal{L}^{p^*}(V^*)$ be given by $(Gv)(t) := \int_0^t v(s)ds$.
Moreover, let $H:\mathcal{L}^2(L_2(U,H)) \to \mathcal{L}^2(H)$ be given by $(Hv)(t) := \int_0^t v(s) dW(s)$.
Integrating~\eqref{eq:limit_eq_proof_1} from $0$ to $T$ and taking the expectation yields
\begin{equation*}
\begin{split}
& \mathbb{E}\int_0^T (u_{\ell'}(t), \varphi(t)) dt = \mathbb{E}\int_0^T (u_{\ell'}(0),\varphi(t))dt
\\ & + \mathbb{E}\int_0^T \big\langle (G a_{\ell'}(\bar{u}_{\ell'}))(t), \varphi(t) \big\rangle dt + \mathbb{E} \int_0^T \big( (H \, B \bar{u}_{\ell'}\tilde{\Pi}_{m_{\ell'}} )(t) ,\varphi(t)\big) dt.
\end{split}
\end{equation*}
The operator $G$ is linear and bounded and as such it is weakly-weakly continuous.
This operator $H$ is clearly linear.
Furthermore, due to It\^o's isometry,
\begin{equation*}
\begin{split}
\|Hv & \|_{\mathcal{L}^2(H)}^2 = \mathbb{E}\int_0^T |(Hv)(s)|^2 ds\\
& = \int_0^T \mathbb{E} \bigg|\int_0^t v(s) dW(s)\bigg|^2 dt = \mathbb{E}\int_0^T \int_0^t |v(s)|^2 ds dt \leq T \|v\|_{\mathcal{L}^2(L_2(U,H))}^2.
\end{split}
\end{equation*}
Thus the operator $H$ is also bounded.
It follows that $H$ is also weakly-weakly continuous.
Therefore, taking the limit as $\ell' \to \infty$ and using Lemma~\ref{lemma:weak_limits_from_compactness}, one obtains
\begin{equation*}
\begin{split}
& \mathbb{E}\int_0^T (u(t), \varphi(t)) dt = \mathbb{E}\int_0^T (u_0,\varphi(t))dt
\\ & + \mathbb{E}\int_0^T \big\langle (G a^\infty)(t), \varphi(t) \big\rangle dt + \mathbb{E} \int_0^T \big( (H b^\infty)(t) ,\varphi(t)\big) dt.
\end{split}
\end{equation*}
This holds for any $\varphi$ as specified at the beginning of the proof.
By letting $M\to \infty$ and using the limited completeness of the Galerkin scheme it follows that this also holds for any $\varphi \in \mathcal{L}^{p}(V)$.
Thus
\begin{equation}
\label{eq:limit_eq_proof1a}
u(t) = u_0 + \int_0^t a^\infty (s) ds + \int_0^t b^\infty(s) dW(s)
\end{equation}
holds for almost all $(t,\omega) \in (0,T) \times \Omega$.
Let $\varphi$ be a $V_M$-valued and $\mathcal{F}_T$-measurable random variable such that $\mathbb{E}\|\varphi\|_V^2 < \infty$.
Setting $t=T$ in~\eqref{eq:limit_eq_proof_1} and taking the expectation yields
\begin{equation*}
\begin{split}
\mathbb{E} (u_{\ell'}(T), \varphi) = & \mathbb{E} (u_{\ell'}(0),\varphi)
\\ & + \mathbb{E} \big\langle (G a_{\ell'}(\bar{u}_{\ell'}))(T), \varphi \big\rangle + \mathbb{E} \big( (H \,B\bar{u}_{\ell'}\tilde{\Pi}_{m'} )(T) ,\varphi\big).
\end{split}
\end{equation*}
Let $\ell' \to \infty$.
The weak-weak continuity of the operators $G$ and $H$,
together with Lemma~\ref{lemma:weak_limits_from_compactness},
implies that
\begin{equation}
\label{eq:limit_eq_proof_2}
\mathbb{E} (\xi, \varphi) = \mathbb{E} (u_0,\varphi)
+ \mathbb{E} \big\langle (G a^\infty)(T), \varphi \big\rangle + \mathbb{E} \big( (H b^\infty)(T) ,\varphi\big).
\end{equation}
Letting $M\to \infty$ and again using the limited completeness of the Galerkin scheme shows that the above equality holds for any $\mathcal{F}_T$-measurable $\varphi \in L^2(\Omega;V)$.
If one now applies It\^o's formula to~\eqref{eq:limit_eq_proof1a} then one obtains an adapted process $\tilde{u}$ with paths in $C([0,T];H)$ that is equal to $u$ almost surely.
Furthermore, for any $\varphi \in L^2(\Omega; V)$ and due to continuity of $\tilde{u}$,
\begin{equation*}
\begin{split}
\mathbb{E}(\xi - \tilde{u}(T),\varphi) & = \lim_{t\to T}\mathbb{E}(\xi - \tilde{u}(t),\varphi) \\
& = \lim_{t\to T} \mathbb{E} \bigg\langle \int_t^T a^\infty(s) ds + \int_t^T b^\infty(s)dW(s),\varphi \bigg\rangle = 0.
\end{split}
\end{equation*}
Thus $\xi = \tilde{u}(T)$.
This together with~\eqref{eq:limit_eq_proof_2} implies~\eqref{eq:limit_equation2}.
\end{proof}
All that remains to be done to prove Theorem~\ref{thm:main} is to identify $a^\infty$ with $Au$ and $b^\infty$ with $Bu$ and to show strong convergence of $u_\ell(T)$ to $u(T)$.
To that end we would like to use monotonicity of $A$.
In order to overcome the difficulty arising from the fact that the tamed operator $A_{2,\ell}$ does not preserve the monotonicity property of $A_2$ we need the following lemma.
\begin{lemma}
\label{lemma:strong_conv_of_tam_correction}
Let the Growth and Coercivity conditions in Assumption~\ref{ass:operators} hold.
Let Assumptions~\ref{ass:projection} and~\ref{ass:interpolation} be satisfied.
Let $u_\ell(0) \to u_0$ in $L^{q_0}(\Omega; H)$.
Let $\tfrac{\mathfrak{c}(m_\ell)}{n_\ell} \to 0$ as $\ell \to \infty$.
Then
\begin{equation*}
\mathbb{E}\int_0^T \|A_2\bar{u}_\ell(s)-A_{2,\ell} \bar{u}_\ell(s)\|_{V_2^*}^{p^*} ds \to 0 \,\, \textrm{ as }\,\, \ell \to \infty.
\end{equation*}
\end{lemma}
\begin{proof}
Consider some $M>0$.
Recall that $T_\ell$ is given by~\eqref{eq:def_of_T_ell}.
Then
\begin{equation*}
\begin{split}
& I_\ell := \mathbb{E}\int_0^T \|A_{2,\ell} \bar{u}_\ell(s) - A_2\bar{u}_\ell(s)\|_{V_2^*}^{p^*} ds \\
& = \mathbb{E}\int_0^T \big(1-T_\ell(\bar{u}_\ell(s))\big)^{p^*}\|A_2 \bar{u}_\ell(s)\|_{V_2^*}^{p^*}\one{1}_{\{\|A_2 \bar{u}_\ell(s)\|_{V_2^*} \leq M\}} ds\\
& \quad+ \mathbb{E}\int_0^T \big(1-T_\ell(\bar{u}_\ell(s))\big)^{p^*}\|A_2 \bar{u}_\ell(s)\|_{V_2^*}^{p^*}\one{1}_{\{\|A_2 \bar{u}_\ell(s)\|_{V_2^*} > M\}} ds\\
& =: I_{1,\ell,M} + I_{2,\ell,M}.
\end{split}
\end{equation*}
It is observed that
\begin{equation*}
\begin{split}
I_{1,\ell,M} & \leq \mathbb{E}\int_0^T \frac{n_\ell^{-1/2}|\Pi_{m_\ell} A_2 \bar{u}_\ell(s)|}{1+n_\ell^{-1/2}|\Pi_{m_\ell} A_2 \bar{u}_\ell(s)|} \|A_2 \bar{u}_\ell(s)\|_{V_2^*}^{p^*}\one{1}_{\{\|A_2 \bar{u}_\ell(s)\|_{V_2^*} \leq M\}} ds\\
& \leq \mathbb{E}\int_0^T \frac{\tau_{n_\ell}^{1/2}T^{-1/2} \mathfrak{c}(m)^{1/2}M}{1+n_\ell^{-1/2}|\Pi_{m_\ell} A_2 \bar{u}_\ell(s)|} M^{p^*} ds
\leq (\mathfrak{c}(m)\tau_{n_\ell} )^{1/2} T^{1/2} M^{1+p^*}.
\end{split}
\end{equation*}
Recall that due to Corollary~\ref{corollary:remaining_apriori} one knows that
\begin{equation*}
\mathbb{E}\int_0^T \|A_2 \bar{u}_\ell(s)\|_{V_2^*}^{p^*(1+\eta)} ds < c
\end{equation*}
with $c$ independent of $\ell$.
Thus the sequence $\big(\|A_2 \bar{u}_\ell\|_{V_2^*}^{p^*}\big)_{\ell \in \mathbb{N}}$ is uniformly integrable on $(0,T)\times \Omega$ with respect to $dt \times P$.
Hence for any $\epsilon > 0$ there exists $M$ such that $I_{2,\ell,M} < \epsilon/2$ for all $\ell$.
Finally, since $\tfrac{\mathfrak{c}(m_\ell)}{n_\ell} \to 0$ as $\ell \to \infty$, one can choose $\ell$ large such that $I_{1,\ell,M} < \epsilon/2$.
\end{proof}
We now prove Theorem~\ref{thm:uniqueness}.
This is needed to later show that the whole sequence of approximations
converges rather than just a subsequence.
\begin{proof}[Proof of Theorem~\ref{thm:uniqueness}]
Assume that $u_1$ and $u_2$ are two distinct solutions to~\eqref{eq:see}
such that $u_1(0) = u_2(0) = u_0$.
One would now like to apply It\^o's formula for the square of the norm from Pardoux~\cite[Chapitre 2, Theoreme 5.2]{pardoux:thesis}.
To that end one immediately observes that $u_1 - u_2 \in \mathcal{L}^2(V_1) \cap \mathcal{L}^p(V_2)$ and that $u_1(0) - u_2(0) = 0 \in L^2(\Omega;H)$.
Moreover
\begin{equation*}
\|Au_1 - Au_2\|_{\mathcal{L}^2(V_1^*) + \mathcal{L}^{p^*}(V_2^*)} = \|A_1 u_1 - A_1 u_2\|_{\mathcal{L}^2(V_1^*)} + \|A_2 u_1 - A_2 u_2\|_{\mathcal{L}^{p^*}(V_2^*)}.
\end{equation*}
Using the Growth assumption on $A_1$ one observes that
\begin{equation*}
\begin{split}
& \mathbb{E} \int_0^T \|A_1 u_1(s) - A_1 u_2(s)\|_{V_1^*}^2 ds \leq \mathbb{E} \int_0^T 2\big[\|A_1 u_1(s)\|_{V_1^*}^2 + \|A_1 u_2(s)\|_{V_1^*}^2 \big] ds\\
& \leq K \mathbb{E} \int_0^T 2\big[(1+\|u_1(s)\|_{V_1}^2) + (1+\|u_2(s)\|_{V_1}^2) \big] ds\\
& \leq c(1 + \|u_1\|_{\mathcal{L}^2(V_1)}^2 + \|u_2\|_{\mathcal{L}^2(V_1)}^2) < \infty.
\end{split}
\end{equation*}
Also, using the Growth assumption on $A_2$ one obtains
\begin{equation*}
\begin{split}
\mathbb{E} \int_0^T \|A_2 u_1(s) - A_2 u_2(s)\|_{V_2^*}^{p^*} ds
\leq c(1 + \|u_1\|_{\mathcal{L}^p(V_2)}^p + \|u_2\|_{\mathcal{L}^p(V_2)}^p) < \infty.
\end{split}
\end{equation*}
Thus $Au_1 - Au_2 \in \mathcal{L}^2(V_1^*)\cup \mathcal{L}^{p^*}(V_2^*)$.
Finally, using the Growth assumption on $B$ one deduces that $Bu_1 - Bu_2 \in \mathcal{L}^2(L_2(U,H))$.
Hence the afromentioned It\^o's formula for the square of the norm can be applied, yielding
\begin{equation*}
\begin{split}
e^{-Kt}|u_1(t) - u_2(t)|^2 =
& -K \int_0^t e^{-Ks} |u_1(s) - u_2(s)|^2 ds\\
& + \int_0^t e^{-Ks}\big[ 2\langle Au_1(s) - Au_2(s),u_1(s)-u_2(s)\rangle \\
& + \|Bu_1(s) - Bu_2(s)\|_{L_2(U,H)}^2 \big] ds + M(t),
\end{split}
\end{equation*}
where
\begin{equation*}
M(t) := \int_0^t e^{-Ks}\big(u_1(s)-u_2(s),(Bu_1(s)-Bu_2(s))dW(s)\big).
\end{equation*}
One then observes, due to the monotonicity of $A:V\times \Omega \to V^*$, that
\begin{equation*}
e^{-Kt}|u_1(t)-u_2(t)|^2 \leq M(t)
\end{equation*}
and hence $M$ is non-negative.
It is also a real-valued continuous local martingale, and thus a supermartingale.
Furthermore it starts from $0$ and thus, almost surely, $M(t) = 0 $ for all $t\in [0,T]$.
One thus concludes that
$u_1(t) = u_2(t)$ for all $t\in [0,T]$ almost surely.
\end{proof}
Finally we can prove Theorem~\ref{thm:main}.
\begin{proof}[Proof of Theorem~\ref{thm:main}]
Recall that $a_\ell(v) := \Pi_{m_\ell} [A_1 v + A_{2,\ell} v]$.
Applying It\^o's formula to the scheme~\eqref{eq:disc1} and taking expectations yields
\begin{equation*}
\begin{split}
e^{-KT} & \mathbb{E} |u_{\ell'}(T)|^2 = \mathbb{E}|u_{\ell'}(0)|^2 -K \mathbb{E} \int_0^T e^{-Ks} |u_{\ell'}(s)|^2 ds \\
& + \mathbb{E} \int_0^T e^{-Ks}\Big( 2 \langle a_{\ell'}(\bar{u}_{\ell'}(s)), \bar{u}_{\ell'}(s) \rangle + \|\Pi_{m_{\ell'}} B \bar{u}_{\ell'}(s) \|_{L_2(U,H)}^2 \Big) ds\\
& + \mathbb{E} \int_0^T 2e^{-Ks} \langle a_{\ell'}(\bar{u}_{\ell'}(s)), u_{\ell'}(s) - \bar{u}_{\ell'}(s) \rangle ds.
\end{split}
\end{equation*}
Let
\begin{equation*}
I_{1,\ell'} := \mathbb{E} \int_0^T \langle a_{\ell'}(\bar{u}_{\ell'}(s)), u_{\ell'}(s) - \bar{u}_{\ell'}(s) \rangle ds.
\end{equation*}
Using H\"older's inequality results in
\begin{equation*}
I_{1,\ell'} \leq \bigg(\mathbb{E}\int_0^T |a_\ell( \bar{u}_{\ell'}(s))|^2 ds \bigg)^{1/2}\bigg(\mathbb{E}\int_0^T |u_{\ell'}(s) - \bar{u}_{\ell'}(s)|^2 ds \bigg)^{1/2}.
\end{equation*}
Using Assumption~\ref{ass:projection} and Corollary~\ref{corollary:remaining_apriori} yields
\begin{equation*}
I_{1,\ell'} \leq c (\mathfrak{c}(m_{\ell'})\tau_{n_{\ell'}})^{1/2}.
\end{equation*}
Thus,
\begin{equation*}
\begin{split}
& e^{-Ks}\mathbb{E} |u_{\ell'}(T)|^2 \leq \mathbb{E}|u_{\ell'}(0)|^2 -K \mathbb{E} \int_0^T e^{-Ks}|u_{\ell'}(s)|^2 ds\\
& + \mathbb{E} \int_0^T e^{-Ks}\big( 2 \langle a_{\ell'}(\bar{u}_{\ell'}(s)), \bar{u}_{\ell'}(s) \rangle + \|B \bar{u}_{\ell'}(s)\|_{L_2(U,H)}^2 \big) ds
+ c(\mathfrak{c}(m_{\ell'})\tau_{n_{\ell'}})^{1/2}
\end{split}
\end{equation*}
and one may proceed with a monotonicity argument.
Let $w\in \mathcal{L}^p(V)$. Then
\begin{equation*}
\begin{split}
&e^{-KT}\mathbb{E}|u_{\ell'}(T)|^2 \leq \mathbb{E}|u_{\ell'}(0)|^2-K \mathbb{E}\int_0^T e^{-Ks}|u_{\ell'}(s)|^2 ds\\
&\quad + \mathbb{E} \int_0^T e^{-Ks}\big[ 2 \langle a_{\ell'}(\bar{u}_{\ell'}(s)) - a_{\ell'}(w(s)), \bar{u}_{\ell'}(s) - w(s)\rangle \\
& \quad \quad + 2 \langle a_{\ell'}(w(s)), \bar{u}_{\ell'}(s)) - w(s)\rangle + 2\langle a_{\ell'}(\bar{u}_{\ell'}(s)), w(s) \rangle \\
& \quad \quad + 2(Bw(s),B\bar{u}_{\ell'}(s))_{L_2(U,H)\times L_2(U,H)} - \|Bw(s)\|_{L_2(U,H)}^2\\
& \quad \quad + \|B\bar{u}_{\ell'}(s)) - Bw(s)\|_{L_2(U,H)}^2 \big] ds
+ c(\mathfrak{c}(m_{\ell'})\tau_{n_{\ell'}})^{1/2}.
\end{split}
\end{equation*}
Using the Monotonicity assumption on $A$ one obtains
\begin{equation}
\label{eq:main_thm_proof1}
\begin{split}
&e^{-KT}\mathbb{E}|u_{\ell'}(T)|^2 \leq \mathbb{E}|u_{\ell'}(0)|^2\\
& + K\mathbb{E}\int_0^T e^{-Ks} \big(|\bar{u}_{\ell'}(s)|^2 -|u_{\ell'}(s)|^2 -2(\bar{u}_{\ell'}(s),w(s)) + |w(s)|^2 \big) ds \\
&+ \mathbb{E} \int_0^T e^{-Ks} \langle A_{2,\ell'}\bar{u}_{\ell'}(s) - A_2 \bar{u}_{\ell'}(s), \bar{u}_{\ell'}(s) - w(s)\rangle ds\\
& + \mathbb{E} \int_0^T e^{-Ks} \langle A_2 w(s) - A_{2,\ell'}w(s), \bar{u}_{\ell'}(s) - w(s)\rangle ds\\
& + \mathbb{E} \int_0^T e^{-Ks} \big[ 2 \langle a_{\ell'}(w(s)), \bar{u}_{\ell'}(s)) - w(s)\rangle\\
& \quad\quad + 2\langle a_{\ell'}(\bar{u}_{\ell'}(s)), w(s) \rangle \\
& \quad\quad + 2(Bw(s),B\bar{u}_\ell(s))_{L_2(U,H)\times L_2(U,H)} - \|Bw(s)\|_{L_2(U,H)}^2 \big] ds\\
& + c(\mathfrak{c}(m_{\ell'})\tau_{n_{\ell'}})^{1/2}.
\end{split}
\end{equation}
Taking limit inferior as $\ell' \to \infty$,
using the weak lower-semi-continuity of the norm,
Lemma~\ref{lemma:weak_limits_from_compactness},
Lemma~\ref{lemma:strong_conv_of_tam_correction}
and Corollary~\ref{corollary:remaining_apriori},
one observes that
\begin{equation}
\label{eq:main_thm_proof2}
\begin{split}
& e^{-KT}\mathbb{E}|u(T)|^2 \leq \mathbb{E}|u_0|^2 + K\mathbb{E}\int_0^T e^{-Ks} \big[ -2(u(s),w(s)) + |w(s)|^2 \big] ds \\
& + \mathbb{E}\int_0^T e^{-Ks}\big[ 2 \langle A w(s), u(s) - w(s) \rangle + 2\langle a^\infty(s), w(s) \rangle \\
& \quad\quad\quad + 2(Bw(s), b^\infty(s))_{L_2(U,H)\times L_2(U,H)} - \|Bw(s)\|_{L_2(U,H)}^2 \big] ds.
\end{split}
\end{equation}
Applying It\^o's formula to~\eqref{eq:limit_equation1} and taking expectations yields
\begin{equation}
\label{eq:main_thm_proof3}
\begin{split}
e^{-KT}\mathbb{E}|u(T)|^2 = & \mathbb{E}|u(0)|^2 - K\mathbb{E} \int_0^T e^{-Ks} |u(s)|^2 ds\\
& + \mathbb{E}\int_0^T e^{-Ks} \big[2\langle a^\infty(s),u(s)\rangle + \|b^\infty(s)\|_{L_2(U,H)}^2 \big]ds
\end{split}
\end{equation}
Subtracting this from~\eqref{eq:main_thm_proof2} one arrives at
\begin{equation}
\label{eq:main_thm_proof4}
\begin{split}
0 \leq \mathbb{E}\int_0^T & e^{-Ks}\big[ K|u(s)-w(s)|^2 + 2 \langle A w(s), u(s) - w(s) \rangle \\
& + 2 \langle a^\infty(s), w(s) - u(s)\rangle - \|Bu(s) - b^\infty(s)\|_{L_2(U,H)}^2 \big]ds.
\end{split}
\end{equation}
Note that so far $w$ was arbitrary.
It will now be used to identify the nonlinear terms.
First, one takes $w=u$ and observes that,
\begin{equation*}
0 \leq - \mathbb{E}\int_0^T e^{-Ks} \|Bu(s) - b^\infty(s)\|_{L_2(U,H)}^2 ds \leq 0
\end{equation*}
which implies $b^\infty = Bu$.
Next, one sets $w=u+\epsilon z$ with $\epsilon > 0$
and $z\in \mathcal{L}^p(V)$ in~\eqref{eq:main_thm_proof4}.
Dividing by $\epsilon > 0$ leads to
\begin{equation*}
0 \leq \mathbb{E}\int_0^T e^{-Ks} \big[ K\epsilon|z|^2+ 2 \langle A (u(s)+\epsilon z(s)), -z(s) \rangle + 2 \langle a^\infty(s), z(s)\rangle \big]ds.
\end{equation*}
Using hemicontinuity of $A$ while letting $\epsilon \to 0$ results in
\begin{equation*}
\mathbb{E}\int_0^T \langle A u(s), z(s) \rangle ds \leq \mathbb{E}\int_0^T \langle a^\infty(s), z(s)\rangle ds.
\end{equation*}
This holds for an arbitrary $z\in \mathcal{L}^p(V)$ and hence,
in particular, for $-z$.
Thus one obtains that $a^\infty = Au$.
Due to Theorem~\ref{thm:uniqueness}, the solution $u$ to~\eqref{eq:see} is unique.
Thus the whole sequences of approximations converges rather than just the subsequence denoted by $\ell'$.
Finally, in order to show that $u_\ell(T) \to u(T)$ in $L^2(\Omega;H)$,
one uses~\eqref{eq:main_thm_proof1} and~\eqref{eq:main_thm_proof2}
with $w=u$ together with $a^\infty = Au$ and $b^\infty = Bu$.
Consequently, the weak-lower-semi-continuity of the norm and
Lemma~\ref{lemma:weak_limits_from_compactness} lead to
\begin{equation*}
\begin{split}
e^{-KT}\mathbb{E}|u(T)|^2 \leq & \liminf_{\ell\to \infty} e^{-KT} \mathbb{E} |u_\ell(T)|^2 \leq \mathbb{E}|u_0|^2 -K \mathbb{E} \int_0^T e^{-Ks} |u(s)|^2 ds \\
& + \mathbb{E} \int_0^T e^{-Ks}\big[2\langle Au(s),u(s) \rangle + \|Bu(s)\|_{L_2(U,H)}^2 \big]ds.
\end{split}
\end{equation*}
Thus, due to~\eqref{eq:main_thm_proof3},
\begin{equation*}
0 \leq \liminf_{\ell\to \infty} \mathbb{E} |u_\ell(T)|^2 - \mathbb{E}|u(T)|^2 \leq 0.
\end{equation*}
From Lemma~\ref{lemma:weak_limits_from_compactness} and Lemma~\ref{lemma:limit_equation}, one already knows that $u_\ell(T) \rightharpoonup u(T)$ in $L^2(\Omega;H)$.
This is a uniformly convex space (as it is a Hilbert space).
Thus one concludes that $u_\ell(T) \to u(T)$ in $L^2(\Omega;H)$.
For this see, e.g., Br\'ezis~\cite[Proposition 3.32]{brezis:functional}.
\end{proof}
\section{Examples}
\label{sec:examples}
In this section we give examples of three equations which fit into our
framework.
In all three examples the interpolation inequality is a consequence of the Gagliardo--Nirenberg inequality (see, for example,~\cite[Theorem 1.24]{roubicek:nonlinear}).
The first example is the equation:
\begin{equation*}
du = \left[\nabla a(\nabla u) - |u|^{p-2} u\right]dt + u dW\,\, \textrm{ on } \,\, \mathscr{D}\times (0,T)
\end{equation*}
with $u=0$ on the boundary of the domain $\mathscr{D}$
and $u(\cdot,0) = u_0$ given.
Here $a:\R^d \to \R^d$ can be nonlinear
but it is assumed to be continuous, monotone and growing at most linearly.
If we take $a_i(z) = z_i$ then $\nabla a(\nabla u) = \Delta u$ and this equation
is the stochastic Ginzburg--Landau equation.
An example of a nonlinear function is
$a_i(z) = \tfrac{2+\exp(-z_i)}{1+\exp(-z_i)}$.
Moreover $\mathscr{D}$ is a bounded Lipschitz domain in $\R^d$, $d=1,2,3$ and $p\in [2,6)$ if $d=1$, $p\in [2,4)$ if $d=2$ and $p\in [2,10/3)$ if $d=3$.
In our framework $H=L^2(\mathscr{D})$, $V_1 = H^1_0(\mathscr{D})$ and $V_2 = L^p(\mathscr{D})$ (using the standard notation for Lebesgue and Sobolev spaces).
The second is the stochastic Swift--Hohenberg equation:
\begin{equation*}
du = \big[\left(\gamma^2 - (1+\Delta)^2\right)u - |u|^{p-2}u\big]dt + dW\,\, \textrm{ on } \,\, \mathscr{D}\times (0,T)
\end{equation*}
with appropriate boundary and initial conditions.
The domain $\mathscr{D}$ is assumed to be a bounded Lipschitz domain in $\R^2$.
With Dirichlet boundary conditions we would take $V_1 = H^2_0(\mathscr{D})$
and $V_2 = L^p(\mathscr{D})$ with $p\in [2,6)$.
The third example is the spatially extended stochastic FitzHugh--Nagumo system for signal propagation in nerve cells (originally stated by FitzHugh~\cite{fitzhugh:impulses} as a system of ordinary differential equations, see Bonaccorsi and Mastrogiacomo~\cite{bonaccorsi:analysis} for mathematical analysis of the spatially extended stochastic version):
\begin{equation*}
\begin{array}{ll}
du & = (\Delta u + u - u^3 - v)dt + dW\\
dv & = c_1(u - c_2 v + c_3)dt
\end{array}
\,\, \textrm{ on } \,\, (0,1)\times (0,T),
\end{equation*}
together with appropriate initial data for $u$ and $v$ as well as homogeneous Neumann boundary conditions for $u$ only.
In this situation $V_1 = H^1((0,1))\times L^2((0,1))$ while $V_2 = L^4((0,1))\times L^2((0,1))$.
We now provide estimates on the constant $\mathfrak{c}(m)$ in the particular
case when $\mathscr{D} = (0,\pi)^2 \subset \R^2$ and
we use a spectral Galerkin method to construct the spaces $V_m$.
To that end define
\begin{equation*}
\varphi_{n_1 n_2}(x_1,x_2) := \frac{2}{\pi}\sin(n_1 x_1)\sin(n_2 x_2).
\end{equation*}
Let $V_m = \textrm{span}\{\varphi_{n_1 n_2} : n_1 = 1,\ldots, m,
n_2 = 1,\ldots,m\}$.
Then
\begin{equation*}
\Pi_m f :=
\sum_{n_1=1, n_2=1}^m \langle f, \varphi_{n_1 n_2} \rangle \varphi_{n_1 n_2}
\end{equation*}
satisfies Assumption~\ref{ass:projection}.
Moreover we can calculate
\begin{equation*}
\begin{split}
\mathfrak{c}(m) & = \sum_{n_1=1, n_2=1}^m
\left(
\|\varphi_{n_1 n_2}\|_{L^2(\mathscr{D})}^2
+ \|\nabla \varphi_{n_1 n_2}\|_{L^2(\mathscr{D};\R^2)}^2
+ \|\varphi_{n_1 n_2}\|_{L^p(\mathscr{D})}^{2/p}
\right) \\
& = m^2 \left(1 + 2 + c_p\right),
\end{split}
\end{equation*}
where $c_p$ depends only on $p$.
Hence, in order to apply Theorem~\ref{thm:main}, we need a sequence
$(m_\ell, n_\ell, k_\ell)$ such that $\tfrac{m_\ell^2}{n_\ell} \to 0$
as $\ell \to \infty$.
This means that we need to choose $n_\ell = \lfloor m_\ell^{2+\delta} \rfloor$
for some $\delta > 0$.
We also note that if $\mathscr{D} = (0,\pi)^d$ then an analogous construction
of $V_m$ would lead to the conclusion that we need
$n_\ell = \lfloor m_l^{d+\delta}\rfloor$ for some $\delta > 0$.
Crucially we see that the space-time coupling requirement is no more onerous
than in the case of equations with operators growing at most linearly.
\section*{Acknowledgement}
The authors would like to thank the referees for their comments on the paper.
|
\section*{Introduction}
The intensive study of invariants of generic immersions $S^1 \to \R^2$ was
started by V.~I.~Arnold in \cite{Arnold-20} and continued in
\cite{Arnold-MIAN}, \cite{Arnold-21}, \cite{Viro}, \cite{Ai}, \cite{Tabach},
\cite{Shum-1}, \cite{P}, \cite{Merx-4}, etc.
The most interesting invariant of such objects, the {\em strangeness}, is in
fact an invariant of triple points free immersions $S^1 \to \R^2$ (with allowed
self-tangencies).
Almost simultaneously, \cite{V-11}, \cite{V-94}, I considered the {\em
ornaments}, i.e. collections of plane curves (maybe with singularities) without
intersections of three different components, and developed regular techniques
for calculating their invariants. The present work is the (promised in
\cite{V-11}) substitution of these methods into the theory of triple points
free plane curves.
Below we describe a natural filtration of invariants by their {\em orders}, and
a regular method of calculating all invariants of all finite orders for triple
points free plane curves. Following the idea of \cite{A2}, we reduce the study
of invariants (and other cohomology classes of the space of generic objects) to
that of the homology groups of the complementary {\em discriminant set} of
objects with forbidden singularities (i.e., in our case, of curves with triple
points). The more technical tools of this method are the simplicial resolutions
of discriminants (see \cite{V-2}) and the (arising from them) calculus of {\em
triangular diagrams} and {\em connected hypergraphs}, which are analogs of
chord diagrams and connected graphs arising in the theory of finite-order knot
invariants.
The simplest such invariants are described in following two theorems.
First, as was proposed in \cite{V-11} (see Problem 2 of \S~9 there), we
consider the space of all plane curves $\phi: S^1 \to \R^2$ having no triple
points and no singularities obtained as degenerations of triple points (i.e.,
either the double points at which one of two local branches has a singular
point with $\phi' = 0$, or the points at which $\phi' = \phi'' = 0.)$ The
problem of classifying such objects (called the {\em doodles}) is in the same
relation with the classification of ornaments, in which the isotopy
classification of links is with the homotopy classification.
In this setting, the curves {\large ``$0$''} and {\large ``$8$''} become
equivalent, and the strangeness fails to be an invariant of such objects; this
is an analog of the fact that the trivial chord diagram $\ominus$ does not
define a knot invariant.
\thm \label{doodinv} There are no invariants of doodles of orders 1, 2 or 3,
and there is exactly one invariant of order 4. \etheorem
\begin{figure}
\begin{center}
\unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(21.00,21.67)
\multiput(1.00,11.00)(0.08,1.17){2}{\line(0,1){1.17}}
\multiput(1.16,13.34)(0.12,0.51){4}{\line(0,1){0.51}}
\multiput(1.63,15.38)(0.11,0.25){7}{\line(0,1){0.25}}
\multiput(2.41,17.09)(0.11,0.14){10}{\line(0,1){0.14}}
\multiput(3.50,18.50)(0.14,0.11){10}{\line(1,0){0.14}}
\multiput(4.91,19.59)(0.25,0.11){7}{\line(1,0){0.25}}
\multiput(6.63,20.38)(0.51,0.12){4}{\line(1,0){0.51}}
\multiput(8.66,20.84)(1.17,0.08){2}{\line(1,0){1.17}}
\multiput(11.00,21.00)(1.17,-0.08){2}{\line(1,0){1.17}}
\multiput(13.34,20.84)(0.51,-0.12){4}{\line(1,0){0.51}}
\multiput(15.38,20.38)(0.25,-0.11){7}{\line(1,0){0.25}}
\multiput(17.09,19.59)(0.14,-0.11){10}{\line(1,0){0.14}}
\multiput(18.50,18.50)(0.11,-0.14){10}{\line(0,-1){0.14}}
\multiput(19.59,17.09)(0.11,-0.25){7}{\line(0,-1){0.25}}
\multiput(20.38,15.38)(0.12,-0.51){4}{\line(0,-1){0.51}}
\multiput(20.84,13.34)(0.08,-1.17){2}{\line(0,-1){1.17}}
\multiput(21.00,11.00)(-0.08,-1.17){2}{\line(0,-1){1.17}}
\multiput(20.84,8.66)(-0.12,-0.51){4}{\line(0,-1){0.51}}
\multiput(20.38,6.63)(-0.11,-0.25){7}{\line(0,-1){0.25}}
\multiput(19.59,4.91)(-0.11,-0.14){10}{\line(0,-1){0.14}}
\multiput(18.50,3.50)(-0.14,-0.11){10}{\line(-1,0){0.14}}
\multiput(17.09,2.41)(-0.25,-0.11){7}{\line(-1,0){0.25}}
\multiput(15.38,1.63)(-0.51,-0.12){4}{\line(-1,0){0.51}}
\multiput(13.34,1.16)(-1.17,-0.08){2}{\line(-1,0){1.17}}
\multiput(11.00,1.00)(-1.17,0.08){2}{\line(-1,0){1.17}}
\multiput(8.66,1.16)(-0.51,0.12){4}{\line(-1,0){0.51}}
\multiput(6.63,1.63)(-0.25,0.11){7}{\line(-1,0){0.25}}
\multiput(4.91,2.41)(-0.14,0.11){10}{\line(-1,0){0.14}}
\multiput(3.50,3.50)(-0.11,0.14){10}{\line(0,1){0.14}}
\multiput(2.41,4.91)(-0.11,0.25){7}{\line(0,1){0.25}}
\multiput(1.63,6.63)(-0.12,0.51){4}{\line(0,1){0.51}}
\multiput(1.16,8.66)(-0.08,1.17){2}{\line(0,1){1.17}}
\put(2.33,6.00){\line(1,0){4.67}} \put(9.00,6.00){\line(1,0){10.33}}
\multiput(19.33,6.00)(-0.12,0.24){17}{\line(0,1){0.24}}
\multiput(16.00,12.00)(-0.12,0.21){42}{\line(0,1){0.21}}
\multiput(11.00,21.00)(-0.12,-0.22){20}{\line(0,-1){0.22}}
\multiput(7.67,15.33)(-0.12,-0.21){45}{\line(0,-1){0.21}}
\multiput(11.00,1.00)(0.12,0.20){20}{\line(0,1){0.20}}
\multiput(14.33,6.67)(0.12,0.21){45}{\line(0,1){0.21}}
\put(19.67,16.00){\line(-1,0){5.00}} \put(13.00,16.00){\line(-1,0){10.67}}
\multiput(2.33,16.00)(0.12,-0.22){20}{\line(0,-1){0.22}}
\multiput(5.67,10.00)(0.12,-0.20){45}{\line(0,-1){0.20}}
\put(2.00,6.00){\circle*{1.33}} \put(2.00,16.00){\circle*{1.33}}
\put(11.00,21.33){\circle*{1.33}} \put(20.00,16.00){\circle*{1.33}}
\put(20.00,6.00){\circle*{1.33}} \put(11.00,1.00){\circle*{1.33}}
\end{picture}
\end{center}
\caption{Unique invariant of order 4 for doodles} \label{md}
\end{figure}
This invariant is depicted by the triangular diagram shown in Fig.~\ref{md}.
(This diagram is an adequate analog of the chord diagram $\bigoplus$ defining
the first nontrivial knot invariant: they both are simplest diagrams of
corresponding kinds, not containing elements with neighboring vertices.)
\begin{figure}
\begin{center}
\unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(40.00,40.00)
\multiput(20.00,40.00)(1.17,-0.08){2}{\line(1,0){1.17}}
\multiput(22.34,39.84)(0.51,-0.12){4}{\line(1,0){0.51}}
\multiput(24.38,39.38)(0.25,-0.11){7}{\line(1,0){0.25}}
\multiput(26.09,38.59)(0.14,-0.11){10}{\line(1,0){0.14}}
\multiput(27.50,37.50)(0.11,-0.14){10}{\line(0,-1){0.14}}
\multiput(28.59,36.09)(0.11,-0.25){7}{\line(0,-1){0.25}}
\multiput(29.38,34.38)(0.12,-0.51){4}{\line(0,-1){0.51}}
\multiput(29.84,32.34)(0.08,-1.17){2}{\line(0,-1){1.17}}
\multiput(20.00,40.00)(-1.17,-0.08){2}{\line(-1,0){1.17}}
\multiput(17.66,39.84)(-0.51,-0.12){4}{\line(-1,0){0.51}}
\multiput(15.63,39.38)(-0.25,-0.11){7}{\line(-1,0){0.25}}
\multiput(13.91,38.59)(-0.14,-0.11){10}{\line(-1,0){0.14}}
\multiput(12.50,37.50)(-0.11,-0.14){10}{\line(0,-1){0.14}}
\multiput(11.41,36.09)(-0.11,-0.25){7}{\line(0,-1){0.25}}
\multiput(10.63,34.38)(-0.12,-0.51){4}{\line(0,-1){0.51}}
\multiput(10.16,32.34)(-0.08,-1.17){2}{\line(0,-1){1.17}}
\multiput(10.00,30.00)(-1.17,-0.08){2}{\line(-1,0){1.17}}
\multiput(7.66,29.84)(-0.51,-0.12){4}{\line(-1,0){0.51}}
\multiput(5.63,29.38)(-0.25,-0.11){7}{\line(-1,0){0.25}}
\multiput(3.91,28.59)(-0.14,-0.11){10}{\line(-1,0){0.14}}
\multiput(2.50,27.50)(-0.11,-0.14){10}{\line(0,-1){0.14}}
\multiput(1.41,26.09)(-0.11,-0.25){7}{\line(0,-1){0.25}}
\multiput(0.63,24.38)(-0.12,-0.51){4}{\line(0,-1){0.51}}
\multiput(0.16,22.34)(-0.08,-1.17){2}{\line(0,-1){1.17}}
\multiput(0.00,20.00)(0.08,-1.17){2}{\line(0,-1){1.17}}
\multiput(0.16,17.66)(0.12,-0.51){4}{\line(0,-1){0.51}}
\multiput(0.63,15.63)(0.11,-0.25){7}{\line(0,-1){0.25}}
\multiput(1.41,13.91)(0.11,-0.14){10}{\line(0,-1){0.14}}
\multiput(2.50,12.50)(0.14,-0.11){10}{\line(1,0){0.14}}
\multiput(3.91,11.41)(0.25,-0.11){7}{\line(1,0){0.25}}
\multiput(5.63,10.63)(0.51,-0.12){4}{\line(1,0){0.51}}
\multiput(7.66,10.16)(1.17,-0.08){2}{\line(1,0){1.17}}
\multiput(10.00,10.00)(0.08,-1.17){2}{\line(0,-1){1.17}}
\multiput(10.16,7.66)(0.12,-0.51){4}{\line(0,-1){0.51}}
\multiput(10.63,5.63)(0.11,-0.25){7}{\line(0,-1){0.25}}
\multiput(11.41,3.91)(0.11,-0.14){10}{\line(0,-1){0.14}}
\multiput(12.50,2.50)(0.14,-0.11){10}{\line(1,0){0.14}}
\multiput(13.91,1.41)(0.25,-0.11){7}{\line(1,0){0.25}}
\multiput(15.63,0.63)(0.51,-0.12){4}{\line(1,0){0.51}}
\multiput(17.66,0.16)(1.17,-0.08){2}{\line(1,0){1.17}}
\multiput(20.00,0.00)(1.17,0.08){2}{\line(1,0){1.17}}
\multiput(22.34,0.16)(0.51,0.12){4}{\line(1,0){0.51}}
\multiput(24.38,0.63)(0.25,0.11){7}{\line(1,0){0.25}}
\multiput(26.09,1.41)(0.14,0.11){10}{\line(1,0){0.14}}
\multiput(27.50,2.50)(0.11,0.14){10}{\line(0,1){0.14}}
\multiput(28.59,3.91)(0.11,0.25){7}{\line(0,1){0.25}}
\multiput(29.38,5.63)(0.12,0.51){4}{\line(0,1){0.51}}
\multiput(29.84,7.66)(0.08,1.17){2}{\line(0,1){1.17}}
\multiput(30.00,10.00)(1.17,0.08){2}{\line(1,0){1.17}}
\multiput(32.34,10.16)(0.51,0.12){4}{\line(1,0){0.51}}
\multiput(34.38,10.63)(0.25,0.11){7}{\line(1,0){0.25}}
\multiput(36.09,11.41)(0.14,0.11){10}{\line(1,0){0.14}}
\multiput(37.50,12.50)(0.11,0.14){10}{\line(0,1){0.14}}
\multiput(38.59,13.91)(0.11,0.25){7}{\line(0,1){0.25}}
\multiput(39.38,15.63)(0.12,0.51){4}{\line(0,1){0.51}}
\multiput(39.84,17.66)(0.08,1.17){2}{\line(0,1){1.17}}
\multiput(40.00,20.00)(-0.08,1.17){2}{\line(0,1){1.17}}
\multiput(39.84,22.34)(-0.12,0.51){4}{\line(0,1){0.51}}
\multiput(39.38,24.38)(-0.11,0.25){7}{\line(0,1){0.25}}
\multiput(38.59,26.09)(-0.11,0.14){10}{\line(0,1){0.14}}
\multiput(37.50,27.50)(-0.14,0.11){10}{\line(-1,0){0.14}}
\multiput(36.09,28.59)(-0.25,0.11){7}{\line(-1,0){0.25}}
\multiput(34.38,29.38)(-0.51,0.12){4}{\line(-1,0){0.51}}
\multiput(32.34,29.84)(-1.17,0.08){2}{\line(-1,0){1.17}}
\put(10.00,10.00){\line(1,0){2.42}}
\multiput(12.42,10.08)(1.13,0.12){2}{\line(1,0){1.13}}
\multiput(14.69,10.31)(0.53,0.10){4}{\line(1,0){0.53}}
\multiput(16.80,10.70)(0.39,0.11){5}{\line(1,0){0.39}}
\multiput(18.75,11.25)(0.30,0.12){6}{\line(1,0){0.30}}
\multiput(20.55,11.95)(0.21,0.11){8}{\line(1,0){0.21}}
\multiput(22.19,12.81)(0.16,0.11){9}{\line(1,0){0.16}}
\multiput(23.67,13.83)(0.13,0.12){10}{\line(1,0){0.13}}
\multiput(25.00,15.00)(0.12,0.13){10}{\line(0,1){0.13}}
\multiput(26.17,16.33)(0.11,0.16){9}{\line(0,1){0.16}}
\multiput(27.19,17.81)(0.11,0.21){8}{\line(0,1){0.21}}
\multiput(28.05,19.45)(0.12,0.30){6}{\line(0,1){0.30}}
\multiput(28.75,21.25)(0.11,0.39){5}{\line(0,1){0.39}}
\multiput(29.30,23.20)(0.10,0.53){4}{\line(0,1){0.53}}
\multiput(29.69,25.31)(0.12,1.13){2}{\line(0,1){1.13}}
\put(29.92,27.58){\line(0,1){2.42}} \put(30.00,10.00){\line(0,1){2.42}}
\multiput(29.92,12.42)(-0.12,1.13){2}{\line(0,1){1.13}}
\multiput(29.69,14.69)(-0.10,0.53){4}{\line(0,1){0.53}}
\multiput(29.30,16.80)(-0.11,0.39){5}{\line(0,1){0.39}}
\multiput(28.75,18.75)(-0.12,0.30){6}{\line(0,1){0.30}}
\multiput(28.05,20.55)(-0.11,0.21){8}{\line(0,1){0.21}}
\multiput(27.19,22.19)(-0.11,0.16){9}{\line(0,1){0.16}}
\multiput(26.17,23.67)(-0.12,0.13){10}{\line(0,1){0.13}}
\multiput(25.00,25.00)(-0.13,0.12){10}{\line(-1,0){0.13}}
\multiput(23.67,26.17)(-0.16,0.11){9}{\line(-1,0){0.16}}
\multiput(22.19,27.19)(-0.21,0.11){8}{\line(-1,0){0.21}}
\multiput(20.55,28.05)(-0.30,0.12){6}{\line(-1,0){0.30}}
\multiput(18.75,28.75)(-0.39,0.11){5}{\line(-1,0){0.39}}
\multiput(16.80,29.30)(-0.53,0.10){4}{\line(-1,0){0.53}}
\multiput(14.69,29.69)(-1.13,0.12){2}{\line(-1,0){1.13}}
\put(12.42,29.92){\line(-1,0){2.42}} \put(30.00,30.00){\line(-1,0){2.42}}
\multiput(27.58,29.92)(-1.13,-0.12){2}{\line(-1,0){1.13}}
\multiput(25.31,29.69)(-0.53,-0.10){4}{\line(-1,0){0.53}}
\multiput(23.20,29.30)(-0.39,-0.11){5}{\line(-1,0){0.39}}
\multiput(21.25,28.75)(-0.30,-0.12){6}{\line(-1,0){0.30}}
\multiput(19.45,28.05)(-0.21,-0.11){8}{\line(-1,0){0.21}}
\multiput(17.81,27.19)(-0.16,-0.11){9}{\line(-1,0){0.16}}
\multiput(16.33,26.17)(-0.13,-0.12){10}{\line(-1,0){0.13}}
\multiput(15.00,25.00)(-0.12,-0.13){10}{\line(0,-1){0.13}}
\multiput(13.83,23.67)(-0.11,-0.16){9}{\line(0,-1){0.16}}
\multiput(12.81,22.19)(-0.11,-0.21){8}{\line(0,-1){0.21}}
\multiput(11.95,20.55)(-0.12,-0.30){6}{\line(0,-1){0.30}}
\multiput(11.25,18.75)(-0.11,-0.39){5}{\line(0,-1){0.39}}
\multiput(10.70,16.80)(-0.10,-0.53){4}{\line(0,-1){0.53}}
\multiput(10.31,14.69)(-0.12,-1.13){2}{\line(0,-1){1.13}}
\put(10.08,12.42){\line(0,-1){2.42}} \put(10.00,30.00){\line(0,-1){2.42}}
\multiput(10.08,27.58)(0.12,-1.13){2}{\line(0,-1){1.13}}
\multiput(10.31,25.31)(0.10,-0.53){4}{\line(0,-1){0.53}}
\multiput(10.70,23.20)(0.11,-0.39){5}{\line(0,-1){0.39}}
\multiput(11.25,21.25)(0.12,-0.30){6}{\line(0,-1){0.30}}
\multiput(11.95,19.45)(0.11,-0.21){8}{\line(0,-1){0.21}}
\multiput(12.81,17.81)(0.11,-0.16){9}{\line(0,-1){0.16}}
\multiput(13.83,16.33)(0.12,-0.13){10}{\line(0,-1){0.13}}
\multiput(15.00,15.00)(0.13,-0.12){10}{\line(1,0){0.13}}
\multiput(16.33,13.83)(0.16,-0.11){9}{\line(1,0){0.16}}
\multiput(17.81,12.81)(0.21,-0.11){8}{\line(1,0){0.21}}
\multiput(19.45,11.95)(0.30,-0.12){6}{\line(1,0){0.30}}
\multiput(21.25,11.25)(0.39,-0.11){5}{\line(1,0){0.39}}
\multiput(23.20,10.70)(0.53,-0.10){4}{\line(1,0){0.53}}
\multiput(25.31,10.31)(1.13,-0.12){2}{\line(1,0){1.13}}
\put(27.58,10.08){\line(1,0){2.42}}
\end{picture}
\end{center}
\caption{A plane curve not equivalent to the circle in the space of singular
triple points free curves (after A.~B.~Merkov)} \label{merx}
\end{figure}
This invariant proves, in particular, that the curve of Fig.~\ref{merx}
(discovered previously by A.~B.~Merkov) is not equivalent to a circle.
\medskip
Further, following \cite{Arnold-20}, let us consider the immersed curves in
$\R^2.$
\thm \label{idoodinv} There are only the following invariants of orders $\le 4$
of triple points free plane {\bf immersed} curves $S^1 \to \R^2$:
1) no invariants of order 1;
2) one invariant of order 2 (the Arnold's {\em strangeness}; by some reasons we
denote this invariant by the simplest "triangular diagram" of Fig.~\ref{one}a;
3) one more invariant of order 3 (its natural notation see in Fig.~\ref{one}b;
\begin{figure}
\begin{center}
\special{em:linewidth 0.4pt} \unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(46.00,21.67)
\multiput(10.00,21.00)(0.79,-0.09){2}{\line(1,0){0.79}}
\multiput(11.59,20.82)(0.30,-0.11){5}{\line(1,0){0.30}}
\multiput(13.09,20.28)(0.17,-0.11){8}{\line(1,0){0.17}}
\multiput(14.43,19.42)(0.11,-0.11){10}{\line(0,-1){0.11}}
\multiput(15.54,18.27)(0.12,-0.20){7}{\line(0,-1){0.20}}
\multiput(16.37,16.91)(0.10,-0.30){5}{\line(0,-1){0.30}}
\multiput(16.86,15.39)(0.07,-0.80){2}{\line(0,-1){0.80}}
\multiput(17.00,13.80)(-0.11,-0.79){2}{\line(0,-1){0.79}}
\multiput(16.77,12.22)(-0.12,-0.30){5}{\line(0,-1){0.30}}
\multiput(16.19,10.73)(-0.11,-0.16){8}{\line(0,-1){0.16}}
\multiput(15.29,9.42)(-0.13,-0.12){9}{\line(-1,0){0.13}}
\multiput(14.11,8.34)(-0.20,-0.11){7}{\line(-1,0){0.20}}
\multiput(12.72,7.55)(-0.38,-0.11){4}{\line(-1,0){0.38}}
\put(11.19,7.10){\line(-1,0){1.59}}
\multiput(9.60,7.01)(-0.52,0.09){3}{\line(-1,0){0.52}}
\multiput(8.03,7.28)(-0.25,0.10){6}{\line(-1,0){0.25}}
\multiput(6.56,7.90)(-0.16,0.12){8}{\line(-1,0){0.16}}
\multiput(5.27,8.84)(-0.12,0.13){9}{\line(0,1){0.13}}
\multiput(4.22,10.05)(-0.11,0.20){7}{\line(0,1){0.20}}
\multiput(3.48,11.46)(-0.10,0.39){4}{\line(0,1){0.39}}
\put(3.07,13.00){\line(0,1){1.60}}
\multiput(3.03,14.60)(0.11,0.52){3}{\line(0,1){0.52}}
\multiput(3.34,16.16)(0.11,0.24){6}{\line(0,1){0.24}}
\multiput(4.01,17.61)(0.11,0.14){9}{\line(0,1){0.14}}
\multiput(4.98,18.88)(0.14,0.11){9}{\line(1,0){0.14}}
\multiput(6.22,19.89)(0.24,0.12){6}{\line(1,0){0.24}}
\multiput(7.65,20.59)(0.59,0.10){4}{\line(1,0){0.59}}
\put(4.00,11.00){\line(1,0){12.00}}
\multiput(16.00,11.00)(-0.12,0.20){50}{\line(0,1){0.20}}
\multiput(10.00,21.00)(-0.12,-0.20){50}{\line(0,-1){0.20}}
\multiput(40.00,21.00)(0.79,-0.09){2}{\line(1,0){0.79}}
\multiput(41.59,20.82)(0.30,-0.11){5}{\line(1,0){0.30}}
\multiput(43.09,20.28)(0.17,-0.11){8}{\line(1,0){0.17}}
\multiput(44.43,19.42)(0.11,-0.11){10}{\line(0,-1){0.11}}
\multiput(45.54,18.27)(0.12,-0.20){7}{\line(0,-1){0.20}}
\multiput(46.37,16.91)(0.10,-0.30){5}{\line(0,-1){0.30}}
\multiput(46.86,15.39)(0.07,-0.80){2}{\line(0,-1){0.80}}
\multiput(47.00,13.80)(-0.11,-0.79){2}{\line(0,-1){0.79}}
\multiput(46.77,12.22)(-0.12,-0.30){5}{\line(0,-1){0.30}}
\multiput(46.19,10.73)(-0.11,-0.16){8}{\line(0,-1){0.16}}
\multiput(45.29,9.42)(-0.13,-0.12){9}{\line(-1,0){0.13}}
\multiput(44.11,8.34)(-0.20,-0.11){7}{\line(-1,0){0.20}}
\multiput(42.72,7.55)(-0.38,-0.11){4}{\line(-1,0){0.38}}
\put(41.19,7.10){\line(-1,0){1.59}}
\multiput(39.60,7.01)(-0.52,0.09){3}{\line(-1,0){0.52}}
\multiput(38.03,7.28)(-0.25,0.10){6}{\line(-1,0){0.25}}
\multiput(36.56,7.90)(-0.16,0.12){8}{\line(-1,0){0.16}}
\multiput(35.27,8.84)(-0.12,0.13){9}{\line(0,1){0.13}}
\multiput(34.22,10.05)(-0.11,0.20){7}{\line(0,1){0.20}}
\multiput(33.48,11.46)(-0.10,0.39){4}{\line(0,1){0.39}}
\put(33.07,13.00){\line(0,1){1.60}}
\multiput(33.03,14.60)(0.11,0.52){3}{\line(0,1){0.52}}
\multiput(33.34,16.16)(0.11,0.24){6}{\line(0,1){0.24}}
\multiput(34.01,17.61)(0.11,0.14){9}{\line(0,1){0.14}}
\multiput(34.98,18.88)(0.14,0.11){9}{\line(1,0){0.14}}
\multiput(36.22,19.89)(0.24,0.12){6}{\line(1,0){0.24}}
\multiput(37.65,20.59)(0.59,0.10){4}{\line(1,0){0.59}}
\put(45.00,2.00){\makebox(0,0)[cc]{{\large b}}}
\put(15.00,2.00){\makebox(0,0)[cc]{{\large a}}}
\put(40.00,14.00){\vector(1,1){4.67}} \put(44.67,9.33){\vector(-1,1){4.67}}
\put(40.00,14.00){\vector(-1,-1){4.67}} \put(35.33,18.67){\vector(1,-1){4.67}}
\put(4.00,11.00){\circle*{1.00}} \put(16.00,11.00){\circle*{1.00}}
\put(10.00,21.00){\circle*{1.00}} \put(35.00,19.00){\circle*{1.00}}
\put(35.00,9.00){\circle*{1.00}} \put(45.00,9.00){\circle*{1.00}}
\put(45.00,19.00){\circle*{1.00}}
\end{picture}
\end{center}
\caption{Notation for invariants of order 2 and 3 of immersed curves}
\label{one}
\end{figure}
4) five more invariants of order 4 (they are described in Fig.~\ref{two}).
\etheorem
\begin{figure}
\begin{center}
\unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(125.00,25.67)
\multiput(8.00,22.00)(0.79,-0.09){2}{\line(1,0){0.79}}
\multiput(9.59,21.82)(0.30,-0.11){5}{\line(1,0){0.30}}
\multiput(11.09,21.28)(0.17,-0.11){8}{\line(1,0){0.17}}
\multiput(12.43,20.42)(0.11,-0.11){10}{\line(0,-1){0.11}}
\multiput(13.54,19.27)(0.12,-0.20){7}{\line(0,-1){0.20}}
\multiput(14.37,17.91)(0.10,-0.30){5}{\line(0,-1){0.30}}
\multiput(14.86,16.39)(0.07,-0.80){2}{\line(0,-1){0.80}}
\multiput(15.00,14.80)(-0.11,-0.79){2}{\line(0,-1){0.79}}
\multiput(14.77,13.22)(-0.12,-0.30){5}{\line(0,-1){0.30}}
\multiput(14.19,11.73)(-0.11,-0.16){8}{\line(0,-1){0.16}}
\multiput(13.29,10.42)(-0.13,-0.12){9}{\line(-1,0){0.13}}
\multiput(12.11,9.34)(-0.20,-0.11){7}{\line(-1,0){0.20}}
\multiput(10.72,8.55)(-0.38,-0.11){4}{\line(-1,0){0.38}}
\put(9.19,8.10){\line(-1,0){1.59}}
\multiput(7.60,8.01)(-0.52,0.09){3}{\line(-1,0){0.52}}
\multiput(6.03,8.28)(-0.25,0.10){6}{\line(-1,0){0.25}}
\multiput(4.56,8.90)(-0.16,0.12){8}{\line(-1,0){0.16}}
\multiput(3.27,9.84)(-0.12,0.13){9}{\line(0,1){0.13}}
\multiput(2.22,11.05)(-0.11,0.20){7}{\line(0,1){0.20}}
\multiput(1.48,12.46)(-0.10,0.39){4}{\line(0,1){0.39}}
\put(1.07,14.00){\line(0,1){1.60}}
\multiput(1.03,15.60)(0.11,0.52){3}{\line(0,1){0.52}}
\multiput(1.34,17.16)(0.11,0.24){6}{\line(0,1){0.24}}
\multiput(2.01,18.61)(0.11,0.14){9}{\line(0,1){0.14}}
\multiput(2.98,19.88)(0.14,0.11){9}{\line(1,0){0.14}}
\multiput(4.22,20.89)(0.24,0.12){6}{\line(1,0){0.24}}
\multiput(5.65,21.59)(0.59,0.10){4}{\line(1,0){0.59}}
\put(2.00,18.00){\line(1,0){12.00}}
\multiput(14.00,18.00)(-0.18,0.12){34}{\line(-1,0){0.18}}
\multiput(8.00,22.00)(-0.18,-0.12){34}{\line(-1,0){0.18}}
\put(2.00,12.00){\line(1,0){12.00}}
\multiput(14.00,12.00)(-0.18,-0.12){34}{\line(-1,0){0.18}}
\multiput(8.00,8.00)(-0.18,0.12){34}{\line(-1,0){0.18}}
\multiput(28.00,22.00)(0.79,-0.09){2}{\line(1,0){0.79}}
\multiput(29.59,21.82)(0.30,-0.11){5}{\line(1,0){0.30}}
\multiput(31.09,21.28)(0.17,-0.11){8}{\line(1,0){0.17}}
\multiput(32.43,20.42)(0.11,-0.11){10}{\line(0,-1){0.11}}
\multiput(33.54,19.27)(0.12,-0.20){7}{\line(0,-1){0.20}}
\multiput(34.37,17.91)(0.10,-0.30){5}{\line(0,-1){0.30}}
\multiput(34.86,16.39)(0.07,-0.80){2}{\line(0,-1){0.80}}
\multiput(35.00,14.80)(-0.11,-0.79){2}{\line(0,-1){0.79}}
\multiput(34.77,13.22)(-0.12,-0.30){5}{\line(0,-1){0.30}}
\multiput(34.19,11.73)(-0.11,-0.16){8}{\line(0,-1){0.16}}
\multiput(33.29,10.42)(-0.13,-0.12){9}{\line(-1,0){0.13}}
\multiput(32.11,9.34)(-0.20,-0.11){7}{\line(-1,0){0.20}}
\multiput(30.72,8.55)(-0.38,-0.11){4}{\line(-1,0){0.38}}
\put(29.19,8.10){\line(-1,0){1.59}}
\multiput(27.60,8.01)(-0.52,0.09){3}{\line(-1,0){0.52}}
\multiput(26.03,8.28)(-0.25,0.10){6}{\line(-1,0){0.25}}
\multiput(24.56,8.90)(-0.16,0.12){8}{\line(-1,0){0.16}}
\multiput(23.27,9.84)(-0.12,0.13){9}{\line(0,1){0.13}}
\multiput(22.22,11.05)(-0.11,0.20){7}{\line(0,1){0.20}}
\multiput(21.48,12.46)(-0.10,0.39){4}{\line(0,1){0.39}}
\put(21.07,14.00){\line(0,1){1.60}}
\multiput(21.03,15.60)(0.11,0.52){3}{\line(0,1){0.52}}
\multiput(21.34,17.16)(0.11,0.24){6}{\line(0,1){0.24}}
\multiput(22.01,18.61)(0.11,0.14){9}{\line(0,1){0.14}}
\multiput(22.98,19.88)(0.14,0.11){9}{\line(1,0){0.14}}
\multiput(24.22,20.89)(0.24,0.12){6}{\line(1,0){0.24}}
\multiput(25.65,21.59)(0.59,0.10){4}{\line(1,0){0.59}}
\multiput(28.00,22.00)(-0.18,-0.12){34}{\line(-1,0){0.18}}
\multiput(22.00,18.00)(0.24,-0.12){50}{\line(1,0){0.24}}
\multiput(34.00,12.00)(-0.12,0.20){50}{\line(0,1){0.20}}
\multiput(34.00,18.00)(-0.12,-0.20){50}{\line(0,-1){0.20}}
\multiput(28.00,8.00)(-0.18,0.12){34}{\line(-1,0){0.18}}
\multiput(22.00,12.00)(0.24,0.12){50}{\line(1,0){0.24}}
\multiput(48.00,22.00)(0.79,-0.09){2}{\line(1,0){0.79}}
\multiput(49.59,21.82)(0.30,-0.11){5}{\line(1,0){0.30}}
\multiput(51.09,21.28)(0.17,-0.11){8}{\line(1,0){0.17}}
\multiput(52.43,20.42)(0.11,-0.11){10}{\line(0,-1){0.11}}
\multiput(53.54,19.27)(0.12,-0.20){7}{\line(0,-1){0.20}}
\multiput(54.37,17.91)(0.10,-0.30){5}{\line(0,-1){0.30}}
\multiput(54.86,16.39)(0.07,-0.80){2}{\line(0,-1){0.80}}
\multiput(55.00,14.80)(-0.11,-0.79){2}{\line(0,-1){0.79}}
\multiput(54.77,13.22)(-0.12,-0.30){5}{\line(0,-1){0.30}}
\multiput(54.19,11.73)(-0.11,-0.16){8}{\line(0,-1){0.16}}
\multiput(53.29,10.42)(-0.13,-0.12){9}{\line(-1,0){0.13}}
\multiput(52.11,9.34)(-0.20,-0.11){7}{\line(-1,0){0.20}}
\multiput(50.72,8.55)(-0.38,-0.11){4}{\line(-1,0){0.38}}
\put(49.19,8.10){\line(-1,0){1.59}}
\multiput(47.60,8.01)(-0.52,0.09){3}{\line(-1,0){0.52}}
\multiput(46.03,8.28)(-0.25,0.10){6}{\line(-1,0){0.25}}
\multiput(44.56,8.90)(-0.16,0.12){8}{\line(-1,0){0.16}}
\multiput(43.27,9.84)(-0.12,0.13){9}{\line(0,1){0.13}}
\multiput(42.22,11.05)(-0.11,0.20){7}{\line(0,1){0.20}}
\multiput(41.48,12.46)(-0.10,0.39){4}{\line(0,1){0.39}}
\put(41.07,14.00){\line(0,1){1.60}}
\multiput(41.03,15.60)(0.11,0.52){3}{\line(0,1){0.52}}
\multiput(41.34,17.16)(0.11,0.24){6}{\line(0,1){0.24}}
\multiput(42.01,18.61)(0.11,0.14){9}{\line(0,1){0.14}}
\multiput(42.98,19.88)(0.14,0.11){9}{\line(1,0){0.14}}
\multiput(44.22,20.89)(0.24,0.12){6}{\line(1,0){0.24}}
\multiput(45.65,21.59)(0.59,0.10){4}{\line(1,0){0.59}}
\multiput(48.00,22.00)(-0.12,-0.20){50}{\line(0,-1){0.20}}
\put(42.00,12.00){\line(1,0){12.00}}
\multiput(54.00,12.00)(-0.12,0.20){50}{\line(0,1){0.20}}
\put(42.00,18.00){\line(1,0){12.00}}
\multiput(54.00,18.00)(-0.12,-0.20){50}{\line(0,-1){0.20}}
\multiput(48.00,8.00)(-0.12,0.20){50}{\line(0,1){0.20}}
\put(28.00,2.00){\makebox(0,0)[cc]{{\large a}}}
\multiput(77.00,15.00)(0.08,1.17){2}{\line(0,1){1.17}}
\multiput(77.16,17.34)(0.12,0.51){4}{\line(0,1){0.51}}
\multiput(77.63,19.38)(0.11,0.25){7}{\line(0,1){0.25}}
\multiput(78.41,21.09)(0.11,0.14){10}{\line(0,1){0.14}}
\multiput(79.50,22.50)(0.14,0.11){10}{\line(1,0){0.14}}
\multiput(80.91,23.59)(0.25,0.11){7}{\line(1,0){0.25}}
\multiput(82.63,24.38)(0.51,0.12){4}{\line(1,0){0.51}}
\multiput(84.66,24.84)(1.17,0.08){2}{\line(1,0){1.17}}
\multiput(87.00,25.00)(1.17,-0.08){2}{\line(1,0){1.17}}
\multiput(89.34,24.84)(0.51,-0.12){4}{\line(1,0){0.51}}
\multiput(91.38,24.38)(0.25,-0.11){7}{\line(1,0){0.25}}
\multiput(93.09,23.59)(0.14,-0.11){10}{\line(1,0){0.14}}
\multiput(94.50,22.50)(0.11,-0.14){10}{\line(0,-1){0.14}}
\multiput(95.59,21.09)(0.11,-0.25){7}{\line(0,-1){0.25}}
\multiput(96.38,19.38)(0.12,-0.51){4}{\line(0,-1){0.51}}
\multiput(96.84,17.34)(0.08,-1.17){2}{\line(0,-1){1.17}}
\multiput(97.00,15.00)(-0.08,-1.17){2}{\line(0,-1){1.17}}
\multiput(96.84,12.66)(-0.12,-0.51){4}{\line(0,-1){0.51}}
\multiput(96.38,10.63)(-0.11,-0.25){7}{\line(0,-1){0.25}}
\multiput(95.59,8.91)(-0.11,-0.14){10}{\line(0,-1){0.14}}
\multiput(94.50,7.50)(-0.14,-0.11){10}{\line(-1,0){0.14}}
\multiput(93.09,6.41)(-0.25,-0.11){7}{\line(-1,0){0.25}}
\multiput(91.38,5.63)(-0.51,-0.12){4}{\line(-1,0){0.51}}
\multiput(89.34,5.16)(-1.17,-0.08){2}{\line(-1,0){1.17}}
\multiput(87.00,5.00)(-1.17,0.08){2}{\line(-1,0){1.17}}
\multiput(84.66,5.16)(-0.51,0.12){4}{\line(-1,0){0.51}}
\multiput(82.63,5.63)(-0.25,0.11){7}{\line(-1,0){0.25}}
\multiput(80.91,6.41)(-0.14,0.11){10}{\line(-1,0){0.14}}
\multiput(79.50,7.50)(-0.11,0.14){10}{\line(0,1){0.14}}
\multiput(78.41,8.91)(-0.11,0.25){7}{\line(0,1){0.25}}
\multiput(77.63,10.63)(-0.12,0.51){4}{\line(0,1){0.51}}
\multiput(77.16,12.66)(-0.08,1.17){2}{\line(0,1){1.17}}
\put(77.67,18.35){\circle*{1.33}} \put(96.33,18.35){\circle*{1.33}}
\put(87.00,25.00){\circle*{1.33}} \put(80.80,7.05){\circle*{1.33}}
\put(93.20,7.05){\circle*{1.33}}
\multiput(105.00,15.00)(0.08,1.17){2}{\line(0,1){1.17}}
\multiput(105.16,17.34)(0.12,0.51){4}{\line(0,1){0.51}}
\multiput(105.63,19.38)(0.11,0.25){7}{\line(0,1){0.25}}
\multiput(106.41,21.09)(0.11,0.14){10}{\line(0,1){0.14}}
\multiput(107.50,22.50)(0.14,0.11){10}{\line(1,0){0.14}}
\multiput(108.91,23.59)(0.25,0.11){7}{\line(1,0){0.25}}
\multiput(110.63,24.38)(0.51,0.12){4}{\line(1,0){0.51}}
\multiput(112.66,24.84)(1.17,0.08){2}{\line(1,0){1.17}}
\multiput(115.00,25.00)(1.17,-0.08){2}{\line(1,0){1.17}}
\multiput(117.34,24.84)(0.51,-0.12){4}{\line(1,0){0.51}}
\multiput(119.38,24.38)(0.25,-0.11){7}{\line(1,0){0.25}}
\multiput(121.09,23.59)(0.14,-0.11){10}{\line(1,0){0.14}}
\multiput(122.50,22.50)(0.11,-0.14){10}{\line(0,-1){0.14}}
\multiput(123.59,21.09)(0.11,-0.25){7}{\line(0,-1){0.25}}
\multiput(124.38,19.38)(0.12,-0.51){4}{\line(0,-1){0.51}}
\multiput(124.84,17.34)(0.08,-1.17){2}{\line(0,-1){1.17}}
\multiput(125.00,15.00)(-0.08,-1.17){2}{\line(0,-1){1.17}}
\multiput(124.84,12.66)(-0.12,-0.51){4}{\line(0,-1){0.51}}
\multiput(124.38,10.63)(-0.11,-0.25){7}{\line(0,-1){0.25}}
\multiput(123.59,8.91)(-0.11,-0.14){10}{\line(0,-1){0.14}}
\multiput(122.50,7.50)(-0.14,-0.11){10}{\line(-1,0){0.14}}
\multiput(121.09,6.41)(-0.25,-0.11){7}{\line(-1,0){0.25}}
\multiput(119.38,5.63)(-0.51,-0.12){4}{\line(-1,0){0.51}}
\multiput(117.34,5.16)(-1.17,-0.08){2}{\line(-1,0){1.17}}
\multiput(115.00,5.00)(-1.17,0.08){2}{\line(-1,0){1.17}}
\multiput(112.66,5.16)(-0.51,0.12){4}{\line(-1,0){0.51}}
\multiput(110.63,5.63)(-0.25,0.11){7}{\line(-1,0){0.25}}
\multiput(108.91,6.41)(-0.14,0.11){10}{\line(-1,0){0.14}}
\multiput(107.50,7.50)(-0.11,0.14){10}{\line(0,1){0.14}}
\multiput(106.41,8.91)(-0.11,0.25){7}{\line(0,1){0.25}}
\multiput(105.63,10.63)(-0.12,0.51){4}{\line(0,1){0.51}}
\multiput(105.16,12.66)(-0.08,1.17){2}{\line(0,1){1.17}}
\put(106.00,18.67){\circle*{1.33}} \put(124.00,18.67){\circle*{1.33}}
\put(115.00,25.00){\circle*{1.33}} \put(108.33,7.33){\circle*{1.33}}
\put(122.00,7.33){\circle*{1.33}} \put(87.00,25.00){\vector(1,-3){6.00}}
\put(93.00,6.67){\vector(-4,3){15.33}} \put(77.67,18.33){\vector(1,0){18.33}}
\put(96.00,18.33){\vector(-4,-3){15.33}} \put(80.67,6.67){\vector(1,3){6.00}}
\put(115.00,25.00){\vector(3,-2){9.33}}
\put(124.33,18.67){\vector(-1,-4){3.00}}
\put(121.33,7.33){\vector(-1,0){13.00}} \put(108.33,7.33){\vector(-1,4){3.00}}
\put(105.33,18.67){\vector(3,2){9.67}} \put(87.00,25.00){\vector(1,-3){5.00}}
\put(93.00,6.67){\vector(-4,3){12.33}} \put(77.67,18.33){\vector(1,0){15.33}}
\put(96.00,18.33){\vector(-4,-3){12.33}} \put(80.67,6.67){\vector(1,3){5.00}}
\put(115.00,25.00){\vector(3,-2){7.00}}
\put(124.33,18.67){\vector(-1,-4){2.33}}
\put(121.33,7.33){\vector(-1,0){11.00}} \put(108.33,7.33){\vector(-1,4){2.33}}
\put(105.33,18.67){\vector(3,2){9.67}}
\put(101.00,2.00){\makebox(0,0)[cc]{{\large b}}}
\put(8.00,22.00){\circle*{1.00}} \put(14.00,18.00){\circle*{1.00}}
\put(2.00,18.00){\circle*{1.00}} \put(2.00,12.00){\circle*{1.00}}
\put(14.00,12.00){\circle*{1.00}} \put(22.00,12.00){\circle*{1.00}}
\put(22.00,18.00){\circle*{1.00}} \put(28.00,22.00){\circle*{1.00}}
\put(28.00,8.00){\circle*{1.00}} \put(34.00,12.00){\circle*{1.00}}
\put(34.00,18.00){\circle*{1.00}} \put(42.00,18.00){\circle*{1.00}}
\put(42.00,12.00){\circle*{1.00}} \put(54.00,12.00){\circle*{1.00}}
\put(54.00,18.00){\circle*{1.00}} \put(48.00,22.00){\circle*{1.00}}
\put(48.00,8.00){\circle*{1.00}} \put(8.00,8.00){\circle*{1.00}}
\put(106.00,19.00){\vector(3,2){6.00}}
\end{picture}
\end{center}
\caption{Notation for invariants of order 4} \label{two}
\end{figure}
Our methods allow us to calculate also some higher-dimensional cohomology
classes of spaces of $k$-points free plane curves (both immersed or just
$C^\infty$-pa\-ra\-met\-ri\-zed) for any $k \ge 3.$
E.g., let $k=4.$ The set $\Sigma 4$ of all curves with 4-fold selfintersections
has codimension 2 in the space of all plane curves, thus the first interesting
problem is the calculation of the 1-dimensional cohomology group of the
complementary space of immersed plane curves without such points\footnote{The
problem of calculating such homology groups, posed by V.~I.~Arnold (see
\cite{Arnold-probl}, problem 1996-2) forced me to write this paper}.
This cohomology group $H^1(Imm(S^1, \R^2) \sm \Sigma 4)$ also has a natural
filtration, so that the {\em orders} of (some) its elements are well-defined.
\thm \label{4ptfree} For any connected component of the space of immersions
$S^1 \to \R^2$, the first few groups ${\mathcal F}_d$ of order $d$
1-dimensional integer cohomology classes of the space of four-points free
immersions lying in this component are as follows: $\F_1=\F_2 = \F_3 =\F_4=0,$
$\F_5 \simeq \Z^2.$
However, if we calculate the $\Z_2$-cohomology, then we have $\F_3 \simeq \Z_2,
$ $\F_4/\F_3 =0, $ $\F_5/\F_4 \simeq \Z^2_2,$ and if we calculate the
$\Z_5$-cohomology, then $\F_3=0, \F_4 \simeq \Z_5, \F_5/\F_4 \simeq \Z_5^2.$
\etheorem
Two generators of the group $\F_5/\F_4$ (with integer coefficients) are
naturally depicted by two chains shown in Fig.~\ref{hexagon}.
\begin{figure}
\begin{center}
\unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(61.00,22.67)
\multiput(1.00,12.00)(0.08,1.17){2}{\line(0,1){1.17}}
\multiput(1.16,14.34)(0.12,0.51){4}{\line(0,1){0.51}}
\multiput(1.63,16.38)(0.11,0.25){7}{\line(0,1){0.25}}
\multiput(2.41,18.09)(0.11,0.14){10}{\line(0,1){0.14}}
\multiput(3.50,19.50)(0.14,0.11){10}{\line(1,0){0.14}}
\multiput(4.91,20.59)(0.25,0.11){7}{\line(1,0){0.25}}
\multiput(6.63,21.38)(0.51,0.12){4}{\line(1,0){0.51}}
\multiput(8.66,21.84)(1.17,0.08){2}{\line(1,0){1.17}}
\multiput(11.00,22.00)(1.17,-0.08){2}{\line(1,0){1.17}}
\multiput(13.34,21.84)(0.51,-0.12){4}{\line(1,0){0.51}}
\multiput(15.38,21.38)(0.25,-0.11){7}{\line(1,0){0.25}}
\multiput(17.09,20.59)(0.14,-0.11){10}{\line(1,0){0.14}}
\multiput(18.50,19.50)(0.11,-0.14){10}{\line(0,-1){0.14}}
\multiput(19.59,18.09)(0.11,-0.25){7}{\line(0,-1){0.25}}
\multiput(20.38,16.38)(0.12,-0.51){4}{\line(0,-1){0.51}}
\multiput(20.84,14.34)(0.08,-1.17){2}{\line(0,-1){1.17}}
\multiput(21.00,12.00)(-0.08,-1.17){2}{\line(0,-1){1.17}}
\multiput(20.84,9.66)(-0.12,-0.51){4}{\line(0,-1){0.51}}
\multiput(20.38,7.63)(-0.11,-0.25){7}{\line(0,-1){0.25}}
\multiput(19.59,5.91)(-0.11,-0.14){10}{\line(0,-1){0.14}}
\multiput(18.50,4.50)(-0.14,-0.11){10}{\line(-1,0){0.14}}
\multiput(17.09,3.41)(-0.25,-0.11){7}{\line(-1,0){0.25}}
\multiput(15.38,2.63)(-0.51,-0.12){4}{\line(-1,0){0.51}}
\multiput(13.34,2.16)(-1.17,-0.08){2}{\line(-1,0){1.17}}
\multiput(11.00,2.00)(-1.17,0.08){2}{\line(-1,0){1.17}}
\multiput(8.66,2.16)(-0.51,0.12){4}{\line(-1,0){0.51}}
\multiput(6.63,2.63)(-0.25,0.11){7}{\line(-1,0){0.25}}
\multiput(4.91,3.41)(-0.14,0.11){10}{\line(-1,0){0.14}}
\multiput(3.50,4.50)(-0.11,0.14){10}{\line(0,1){0.14}}
\multiput(2.41,5.91)(-0.11,0.25){7}{\line(0,1){0.25}}
\multiput(1.63,7.63)(-0.12,0.51){4}{\line(0,1){0.51}}
\multiput(1.16,9.66)(-0.08,1.17){2}{\line(0,1){1.17}}
\put(2.33,7.00){\vector(1,0){17.00}} \put(19.33,7.00){\vector(-1,2){7.67}}
\put(11.67,22.00){\vector(-2,-3){10.00}} \put(2.33,17.00){\vector(1,0){17.33}}
\put(19.67,17.00){\vector(-1,-2){7.67}} \put(12.00,2.00){\vector(-2,3){10.00}}
\put(2.33,7.00){\vector(1,0){15.00}} \put(19.33,7.00){\vector(-1,2){6.67}}
\put(11.67,22.00){\vector(-2,-3){8.67}} \put(2.33,17.00){\vector(1,0){15.33}}
\put(19.67,17.00){\vector(-1,-2){6.67}} \put(12.00,2.00){\vector(-2,3){8.67}}
\multiput(41.00,12.00)(0.08,1.17){2}{\line(0,1){1.17}}
\multiput(41.16,14.34)(0.12,0.51){4}{\line(0,1){0.51}}
\multiput(41.63,16.38)(0.11,0.25){7}{\line(0,1){0.25}}
\multiput(42.41,18.09)(0.11,0.14){10}{\line(0,1){0.14}}
\multiput(43.50,19.50)(0.14,0.11){10}{\line(1,0){0.14}}
\multiput(44.91,20.59)(0.25,0.11){7}{\line(1,0){0.25}}
\multiput(46.63,21.38)(0.51,0.12){4}{\line(1,0){0.51}}
\multiput(48.66,21.84)(1.17,0.08){2}{\line(1,0){1.17}}
\multiput(51.00,22.00)(1.17,-0.08){2}{\line(1,0){1.17}}
\multiput(53.34,21.84)(0.51,-0.12){4}{\line(1,0){0.51}}
\multiput(55.38,21.38)(0.25,-0.11){7}{\line(1,0){0.25}}
\multiput(57.09,20.59)(0.14,-0.11){10}{\line(1,0){0.14}}
\multiput(58.50,19.50)(0.11,-0.14){10}{\line(0,-1){0.14}}
\multiput(59.59,18.09)(0.11,-0.25){7}{\line(0,-1){0.25}}
\multiput(60.38,16.38)(0.12,-0.51){4}{\line(0,-1){0.51}}
\multiput(60.84,14.34)(0.08,-1.17){2}{\line(0,-1){1.17}}
\multiput(61.00,12.00)(-0.08,-1.17){2}{\line(0,-1){1.17}}
\multiput(60.84,9.66)(-0.12,-0.51){4}{\line(0,-1){0.51}}
\multiput(60.38,7.63)(-0.11,-0.25){7}{\line(0,-1){0.25}}
\multiput(59.59,5.91)(-0.11,-0.14){10}{\line(0,-1){0.14}}
\multiput(58.50,4.50)(-0.14,-0.11){10}{\line(-1,0){0.14}}
\multiput(57.09,3.41)(-0.25,-0.11){7}{\line(-1,0){0.25}}
\multiput(55.38,2.63)(-0.51,-0.12){4}{\line(-1,0){0.51}}
\multiput(53.34,2.16)(-1.17,-0.08){2}{\line(-1,0){1.17}}
\multiput(51.00,2.00)(-1.17,0.08){2}{\line(-1,0){1.17}}
\multiput(48.66,2.16)(-0.51,0.12){4}{\line(-1,0){0.51}}
\multiput(46.63,2.63)(-0.25,0.11){7}{\line(-1,0){0.25}}
\multiput(44.91,3.41)(-0.14,0.11){10}{\line(-1,0){0.14}}
\multiput(43.50,4.50)(-0.11,0.14){10}{\line(0,1){0.14}}
\multiput(42.41,5.91)(-0.11,0.25){7}{\line(0,1){0.25}}
\multiput(41.63,7.63)(-0.12,0.51){4}{\line(0,1){0.51}}
\multiput(41.16,9.66)(-0.08,1.17){2}{\line(0,1){1.17}}
\put(51.00,22.00){\vector(-2,-1){9.00}} \put(51.00,22.00){\vector(2,-1){9.00}}
\put(60.00,7.00){\vector(0,1){10.00}} \put(60.00,7.00){\vector(-2,-1){9.00}}
\put(42.00,7.00){\vector(2,-1){9.00}} \put(42.00,7.00){\vector(0,1){9.67}}
\put(51.00,22.00){\vector(-2,-1){8.00}} \put(51.00,22.00){\vector(2,-1){8.00}}
\put(60.00,7.00){\vector(0,1){9.00}} \put(60.00,7.00){\vector(-2,-1){8.00}}
\put(42.00,7.00){\vector(2,-1){8.00}} \put(42.00,7.00){\vector(0,1){8.67}}
\put(51.00,2.33){\vector(0,1){19.33}} \put(51.00,2.33){\vector(0,1){18.33}}
\put(51.00,2.33){\vector(0,1){17.33}} \put(51.00,2.33){\vector(0,1){16.33}}
\put(60.00,16.67){\vector(-2,-1){17.67}}
\put(60.00,16.67){\vector(-2,-1){16.67}}
\put(60.00,16.67){\vector(-2,-1){15.67}}
\put(60.00,16.67){\vector(-2,-1){14.67}}
\put(42.33,16.67){\vector(2,-1){17.67}} \put(42.33,16.67){\vector(2,-1){16.67}}
\put(42.33,16.67){\vector(2,-1){15.67}} \put(42.33,16.67){\vector(2,-1){14.67}}
\put(2.23,7.00){\circle*{1.33}} \put(2.33,17.00){\circle*{1.33}}
\put(12.00,22.33){\circle*{1.33}} \put(19.67,17.00){\circle*{1.33}}
\put(19.67,7.00){\circle*{1.33}} \put(12.10,2.15){\circle*{1.33}}
\put(51.00,2.33){\circle*{1.33}} \put(60.10,7.50){\circle*{1.33}}
\put(60.00,16.50){\circle*{1.33}} \put(51.00,22.00){\circle*{1.33}}
\put(42.00,17.00){\circle*{1.33}} \put(42.00,7.67){\circle*{1.33}}
\end{picture}
\end{center}
\caption{Two generators of order 5 of the 1-cohomology group of the space of
plane immersed curves without 4-fold points} \label{hexagon}
\end{figure}
\medskip
Many invariants of immersions from Theorem 1 have elementary description.
Namely, the sum of three generators from Fig.~\ref{two}a is equal to the square
of the strangeness. Moreover, strangeness itself, the unique invariant of order
3 and the sum of two invariants of order 4 shown in Fig.~\ref{two}b, are
"index-type invariants," see \S~\ref{itype} below, thus initiating an infinite
series of finite-order invariants (one in each order) of this sort.
\medskip
{\sc Important Note.} Our notion of the order of invariants differs from the
one used in \cite{Arnold-20}---\cite{Arnold-21}, \cite{Shum-1}, \cite{Tabach}
etc. Any invariant of finite order $k$ in the sense of our work is also of
order $\le [k/2]$ in the sense of these works; the converse is false very much.
There are (among others) three equivalent definitions of finite order
invariants of knots in $\R^3$: 1) the "geometrical", in terms of resolved
discriminants and their filtrations, 2) the "axiomatic", in terms of finite
differences of knot diagrams, and 3) the "combinatorial" (developed in
\cite{PV}) in terms of homomorphisms of chord diagrams. The equivalence of two
first definitions was clear from the very beginning, their equivalence to the
third one is a nontrivial fact, conjectured by M.~Polyak and O.~Viro and proved
by M.~Goussarov.
There is a wide class of objects (including knots, ornaments, and doodles),
whose invariants can be calculated by the methods, developed in \cite{V-7},
\cite{V-11}, i.e. in the terms of resolved discriminants, thus leading to the
"geometrical" definition of finite-order invariants. An "axiomatic" elementary
reformulation of the resulting notion in our present situation also exists, but
it is not a straightforward translation of that from \cite{V-7}, see
\S~\ref{elemdef} below. I believe that it will lead to the most interesting
algebraic structures, reflecting the rich geometric structures staying behind
it.
The "combinatorial" definition and related aspects of the same invariants of
ornaments and doodles are introduced and investigated by A.~B.~Merkov,
\cite{Merx}---\cite{Merx-4} as a far generalization of the index-type
invariants from \cite{V-11}.
In particular, he proved that these invariants distinguish any two
nonequivalent collections of (arbitrarily many) plane curves without triple
intersections or selfintersections. However I believe that the techniques of
the present work allow to calculate all such invariants in the most direct and
regular way.
\medskip
I thank very much A.~B.~Merkov for numerous consultations and other multiform
help.
\section{Elementary theory}
This and the next sections are almost exact analogues of \S\S~1, 2 from
\cite{V-11}.
\subsection{First definitions and Reidemeister moves.}
\begin{definition}
A {\em doodle} is a $C^\infty$-map $\phi: S^1 \to \R^2$ such that for none
three different points $x,y,z \in S^1$ one the following conditions holds
\footnote{In a more general theory, see \cite{Khovanov}, \cite{Merx-4}, this
object is called an 1-doodle. We consider here only such one-component doodles
and call them simply {\em doodles}.}:
\begin{equation}
\label{main} \phi(x)=\phi(y)=\phi(z)
\end{equation}
\begin{equation}
\label{second} \phi'(x)=0, \phi(x) = \phi(y)
\end{equation}
\begin{equation}
\label{tert} \phi'(x)=\phi''(x)=0.
\end{equation}
An {\em I-doodle} (i.e. immersed doodle) is a doodle which is an immersion
(i.e. a map $\phi$ without degenerations of two types (\ref{main}) and
\begin{equation}
\label{imm} \phi'(x)=0 \ ).
\end{equation}
Two doodles (respectively, I-doodles) are {\em equivalent} if there is a
continuous family of doodles (I-doodles) connecting them. An {\em invariant} of
doodles or I-doodles is any function on the space of these objects, taking
equal values at equivalent objects.
A doodle is {\em regular} if it is an immersion having only transverse double
points.
\end{definition}
\prop \label{reidprop} Any equivalence class of doodles or I-doodles contains
regular doodles. Two regular doodles define equivalent doodles (respectively,
I-doodles) if and only if they can be transformed one into the other by a
finite sequence of isotopies of ${\bf R}^2$ (which do not change the
topological picture of the image of the doodle), and of local moves shown in
Fig.~\ref{reidem}a, b (respectively, \ref{reidem}a only). \eprop
\begin{figure}
\begin{center}
\unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(124.00,47.00)
\put(2.00,30.00){\line(1,0){25.00}}
\multiput(27.00,22.00)(-0.11,0.21){11}{\line(0,1){0.21}}
\multiput(25.77,24.26)(-0.11,0.18){11}{\line(0,1){0.18}}
\multiput(24.54,26.30)(-0.11,0.16){11}{\line(0,1){0.16}}
\multiput(23.32,28.10)(-0.11,0.14){11}{\line(0,1){0.14}}
\multiput(22.10,29.69)(-0.11,0.12){11}{\line(0,1){0.12}}
\multiput(20.89,31.05)(-0.12,0.11){10}{\line(-1,0){0.12}}
\multiput(19.68,32.18)(-0.15,0.11){8}{\line(-1,0){0.15}}
\multiput(18.47,33.09)(-0.20,0.11){6}{\line(-1,0){0.20}}
\multiput(17.27,33.78)(-0.30,0.11){4}{\line(-1,0){0.30}}
\multiput(16.08,34.23)(-0.60,0.12){2}{\line(-1,0){0.60}}
\put(14.89,34.47){\line(-1,0){1.19}}
\multiput(13.70,34.48)(-0.59,-0.11){2}{\line(-1,0){0.59}}
\multiput(12.52,34.26)(-0.29,-0.11){4}{\line(-1,0){0.29}}
\multiput(11.34,33.82)(-0.20,-0.11){6}{\line(-1,0){0.20}}
\multiput(10.17,33.15)(-0.15,-0.11){8}{\line(-1,0){0.15}}
\multiput(9.00,32.26)(-0.12,-0.11){10}{\line(-1,0){0.12}}
\multiput(7.83,31.14)(-0.12,-0.13){10}{\line(0,-1){0.13}}
\multiput(6.68,29.80)(-0.12,-0.16){10}{\line(0,-1){0.16}}
\multiput(5.52,28.23)(-0.12,-0.18){10}{\line(0,-1){0.18}}
\multiput(4.37,26.43)(-0.11,-0.20){10}{\line(0,-1){0.20}}
\multiput(3.22,24.41)(-0.11,-0.22){11}{\line(0,-1){0.22}}
\put(2.00,14.00){\line(1,0){25.00}}
\multiput(27.00,6.00)(-0.13,0.11){14}{\line(-1,0){0.13}}
\multiput(25.15,7.59)(-0.15,0.11){12}{\line(-1,0){0.15}}
\multiput(23.31,8.94)(-0.18,0.11){10}{\line(-1,0){0.18}}
\multiput(21.47,10.04)(-0.23,0.11){8}{\line(-1,0){0.23}}
\multiput(19.65,10.90)(-0.30,0.10){6}{\line(-1,0){0.30}}
\multiput(17.84,11.51)(-0.45,0.09){4}{\line(-1,0){0.45}}
\put(16.04,11.88){\line(-1,0){3.57}}
\multiput(12.47,11.88)(-0.44,-0.09){4}{\line(-1,0){0.44}}
\multiput(10.70,11.51)(-0.29,-0.10){6}{\line(-1,0){0.29}}
\multiput(8.94,10.90)(-0.22,-0.11){8}{\line(-1,0){0.22}}
\multiput(7.19,10.04)(-0.17,-0.11){10}{\line(-1,0){0.17}}
\multiput(5.45,8.94)(-0.14,-0.11){12}{\line(-1,0){0.14}}
\multiput(3.72,7.59)(-0.12,-0.11){14}{\line(-1,0){0.12}}
\put(14.00,2.00){\makebox(0,0)[cc]{{\large a}}}
\put(14.00,21.00){\makebox(0,0)[cc]{$\Updownarrow$}}
\put(37.00,11.00){\line(1,0){28.00}}
\put(50.00,18.00){\makebox(0,0)[cc]{$\Updownarrow$}}
\multiput(82.00,34.00)(0.12,-0.12){117}{\line(0,-1){0.12}}
\multiput(88.00,20.00)(0.12,0.12){117}{\line(0,1){0.12}}
\put(103.00,27.00){\line(-1,0){22.00}}
\multiput(88.00,15.00)(0.12,-0.12){117}{\line(0,-1){0.12}}
\multiput(96.00,15.00)(-0.12,-0.12){117}{\line(0,-1){0.12}}
\put(81.00,8.00){\line(1,0){22.00}}
\put(92.00,2.00){\makebox(0,0)[cc]{{\large c}}}
\put(50.00,2.00){\makebox(0,0)[cc]{{\large b}}}
\put(92.00,18.00){\makebox(0,0)[cc]{$\Updownarrow$}}
\multiput(116.00,34.00)(0.12,-0.12){67}{\line(0,-1){0.12}}
\multiput(116.00,26.00)(0.12,0.12){67}{\line(0,1){0.12}}
\put(120.00,26.00){\line(0,1){8.00}}
\put(116.00,32.00){\line(1,0){8.00}}
\multiput(116.00,24.00)(0.12,-0.12){67}{\line(0,-1){0.12}}
\multiput(116.00,16.00)(0.12,0.12){67}{\line(0,1){0.12}}
\put(120.00,16.00){\line(0,1){8.00}}
\put(116.00,20.00){\line(1,0){8.00}}
\multiput(116.00,14.00)(0.12,-0.12){67}{\line(0,-1){0.12}}
\multiput(116.00,6.00)(0.12,0.12){67}{\line(0,1){0.12}}
\put(120.00,6.00){\line(0,1){8.00}}
\put(116.00,8.00){\line(1,0){8.00}}
\put(120.00,2.00){\makebox(0,0)[cc]{{\large d}}}
\put(37.00,22.00){\line(1,0){3.60}}
\multiput(40.60,22.00)(1.61,0.07){2}{\line(1,0){1.61}}
\multiput(43.81,22.13)(0.94,0.09){3}{\line(1,0){0.94}}
\multiput(46.63,22.40)(0.61,0.10){4}{\line(1,0){0.61}}
\multiput(49.05,22.80)(0.41,0.11){5}{\line(1,0){0.41}}
\multiput(51.08,23.33)(0.27,0.11){6}{\line(1,0){0.27}}
\multiput(52.72,24.00)(0.18,0.11){7}{\line(1,0){0.18}}
\multiput(53.96,24.80)(0.11,0.12){8}{\line(0,1){0.12}}
\multiput(54.81,25.73)(0.11,0.27){4}{\line(0,1){0.27}}
\put(55.27,26.80){\line(0,1){1.20}}
\put(55.33,28.00){\line(0,1){2.43}}
\multiput(55.23,30.43)(-0.10,0.31){6}{\line(0,1){0.31}}
\multiput(54.60,32.28)(-0.11,0.13){10}{\line(0,1){0.13}}
\multiput(53.46,33.55)(-0.35,0.11){7}{\line(-1,0){0.35}}
\put(51.00,34.33){\line(-1,0){1.84}}
\multiput(49.16,34.41)(-0.28,-0.11){5}{\line(-1,0){0.28}}
\multiput(47.74,33.87)(-0.11,-0.13){9}{\line(0,-1){0.13}}
\multiput(46.73,32.71)(-0.12,-0.35){5}{\line(0,-1){0.35}}
\multiput(46.15,30.94)(-0.07,-1.47){2}{\line(0,-1){1.47}}
\multiput(46.00,28.00)(0.09,-0.64){2}{\line(0,-1){0.64}}
\multiput(46.18,26.73)(0.11,-0.22){5}{\line(0,-1){0.22}}
\multiput(46.70,25.60)(0.11,-0.12){8}{\line(0,-1){0.12}}
\multiput(47.58,24.63)(0.18,-0.12){7}{\line(1,0){0.18}}
\multiput(48.81,23.80)(0.26,-0.11){6}{\line(1,0){0.26}}
\multiput(50.39,23.12)(0.39,-0.11){5}{\line(1,0){0.39}}
\multiput(52.32,22.59)(0.57,-0.10){4}{\line(1,0){0.57}}
\multiput(54.61,22.20)(1.32,-0.12){2}{\line(1,0){1.32}}
\put(57.24,21.96){\line(1,0){2.99}}
\multiput(60.23,21.88)(2.39,0.06){2}{\line(1,0){2.39}}
\end{picture}
\end{center}
\caption{Standard moves of quasidoodles} \label{reidem}
\end{figure}
In other words, the move of Fig.~\ref{reidem}c is prohibited in the
classification of doodles, and both \ref{reidem}b, \ref{reidem}c in the case of
I-doodles.
The proof of this proposition is trivial.
\begin{definition} A {\em quasidoodle} is any $C^\infty$-map
$S^1 \to \R^2.$ The space of all such maps is denoted by $\K$. The space
$Imm(S^1,\R^2)$ of all immersions $S^1 \to \R^2$ will be denoted by $I\K$. The
{\em discriminant} (respectively, I-discriminant) $\Sigma \subset \K$
(respectively, $I\Sigma \subset I\K$) is the set of all maps from this space
for which one of prohibited conditions (\ref{main})---(\ref{tert})
(respectively, (\ref{main})) is satisfied.
\end{definition}
\prop The set $\Sigma$ is a closed subvariety of codimension 1 in $\K$. The set
of all maps, for which only (\ref{main}) is satisfied, is dense in $\Sigma,$
and the maps satisfying (\ref{second}) or (\ref{tert}) lie in its closure. The
set of quasidoodles having no degenerations of types (\ref{main}) and
(\ref{second}) but with allowed degenerations of type (\ref{tert}) is
path-connected in $\K.$ \eprop
The proof of the last assertion essentially coincides with that of the fact
that {\em all} embeddings $S^1 \to \R^3$ (maybe not regular) form a
path-connected subset in $C^\infty(S^1, \R^3),$ see e.g. \cite{CF}. All other
statements of the proposition are elementary.
\subsection{On Arnold's invariants of immersed plane curves.}
In \cite{Arnold-20} V.~I.~Arnold introduced three invariants of generic
immersed plane curves. One of them is the {\em strangeness}, defined as the
linking number in $I\K$ with the suitably (co)oriented variety $I\Sigma \subset
I\K.$ The coorientation of this variety, participating in this construction,
will be specified in \S~\ref{coorient}.
\subsection{Index-type invariants of I-doodles.}
\label{itype} Let us fix an orientation of the plane ${\bf R}^2.$
Recall that any closed oriented immersed curve $c$ in ${\bf R}^2$ defines an
integer-valued function $ind_c$ on its complement: for any point $t$ of the
complement, $ind_c(t)$ equals the (counterclockwise) rotation number of the
vector $(t,x)$ when $x$ runs one time along $c$.
Consider a regular I-doodle $\phi: S^1 \to \R^2$. To any self-intersection
point $x$ of the curve $c=\phi(S^1)$ assign its index $i(x)$ equal to the
arithmetical mean of four values of $ind_c$ in four neighboring components of
$\R^2 \sm c,$ see Fig.~\ref{ium}. Let us fix a regular (not intersection) point
$\ast $ in $c$ and define its index $i(\ast)$ as the greatest value of $ind_c$
in two neighboring domains of the complement of $c.$ For any selfintersection
point $x$ consider the frame in it, formed by (oriented) tangent vectors to
$c$. These vectors are ordered in correspondence with the number of visits of
$c$ after leaving the point $\ast$. Define the {\em sign} $\sigma(x)$ of the
point $x$ as the sign of the orientation of this ordered frame.
\begin{figure}
\begin{center}
\unitlength=1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(16.00,16.00)
\put(15.00,9.00){\makebox(0,0)[cc]{$i$}}
\put(3.00,9.00){\makebox(0,0)[cc]{$i$}}
\put(9.00,16.00){\makebox(0,0)[cc]{$i+1$}}
\put(9.00,2.00){\makebox(0,0)[cc]{$i-1$}} \put(2.00,2.00){\vector(1,1){14.00}}
\put(2.00,16.00){\vector(1,-1){14.00}}
\put(5.00,13.00){\makebox(0,0)[cc]{{\large $\ast$}}}
\end{picture}
\end{center}
\caption{} \label{ium}
\end{figure}
For any integer $i$ and natural $\beta,$ denote by $\frac{i}{\beta}$ the number
$i(i-1) \cdots (i-\beta+1)/\beta!$, cf. \cite{Merx-2}, \cite{V-11}. It is easy
to see that this number is always integer.
For any natural $\beta,$ define the {\em $\beta$-th moment} $M(\beta)$ of the
regular doodle by the equality
\begin{equation}
\label{moment} M(\beta) = \sum_{x} {\sigma(x)} \binom{i(x)}{\beta} + 2
\binom{i(\ast)}{\beta+1} .
\end{equation}
\begin{proposition}[cf. \cite{Shum-3}, \cite{V-11}]
\label{mom} All numbers $M(\beta),$ $\beta=1,2, \ldots$, are invariants of
I-doodles, in particular do not depend on the choice of the distinguished point
$\ast.$ The first of them, $M(1),$ is the Arnold's strangeness.
\end{proposition}
\begin{remark} In the very similar case of ornaments such invariants were
introduced in \cite{V-11}, \S~1.4, as first nontrivial examples of finite-order
invariants, see also \cite{Merx-2}. In \cite{Tabach} similar expressions
appeared as invariants of one-component {\em long curves}, i.e. essentially of
curves with a fixed nonsingular point $*$. The formulae for all these
invariants contained only the terms similar to the first term of the right-hand
part of (\ref{moment}). Finally, A.~Shumakovich \cite{Shum-3} introduced a
correcting second term and obtained invariants independent on the choice of
this point: these his invariants coincide with (\ref{moment}) up to a linear
transformation with rational coefficients. (The simplest version of this
correcting term, corresponding to the case $\beta=1$ and providing a
combinatorial expression for the Arnold's strangeness, appeared previously in
\cite{Shum-1}.) Numerous more general combinatorial expressions for invariants
of doodles, ornaments, I-doodles, etc. were given in
\cite{Merx}--\cite{Merx-4}. It seems likely that the method described below
(see also \cite{V-11}, \cite{Merx-2}) is the most universal algorithm for
guessing such expressions: first one calculates several first elements of a
spectral sequence converging to the group of all invariants of finite degree,
and then finds an elementary interpretation for them; cf. also \cite{V-7},
\cite{V-94}.
\end{remark}
\subsection{Coorientation of the discriminant.}
\label{coorient}
The discriminant set $\Sigma$ has a natural (co)orientation in its regular
points: if we go along a generic path in the space $\K$ and traverse the
discriminant, doing the local surgery shown in Fig.~\ref{reidem}c, then there
is an invariant way to say, which one of these two resolved pictures lies on
the positive side of the discriminant, and which on the negative. There are
numerous equivalent definitions of this coorientation, see e.g.
\cite{Arnold-20}, \cite{Shum-1}, \cite{Merx-2}. One of them can be formulated
as follows: we consider the sum like (\ref{moment}) with $\beta=1,$ but with
summation only over 3 points participating in the surgery. The positive
(negative) side of discriminant is that with the greater (smaller) value of
this sum.
This coorientation is well defined even if the curve has forbidden multiple
(or, in the theory of I-doodles, forbidden singular) points far away from the
location of the surgery: in fact, this is a coorientation of the locally
irreducible branch of the discriminant set.
\section{Elementary definition of the order of invariants}
\label{elemdef}
\subsection{Degeneration modes and characteristic numbers.}
The orders of invariants of doodles and I-doodles will be defined in the same
way.
\begin{definition} Suppose that $j$ is a natural number, $j \ge 2.$
A {\it degree $j$ standard singularity\/} of doodles is a pair of the form \{a
quasidoodle $\phi:S^1 \to \R^2$; a point $x \in \R^2$\} such that
$\phi^{-1}(x)$ consists of exactly $j+1$ points $z_1, \ldots, z_{j+1}$, the map
$\phi$ close to all these points is an immersion, and the corresponding $j+1$
local branches of the curve $\phi(S^1)$ are pairwise nontangent at $x$. A
quasidoodle is called a {\it regular quasidoodle of complexity $i$,} if it is
an immersion, all its forbidden points (i.e. the points, at which at least
three different components meet) are standard singular points, and the sum of
degrees of these singularities is equal to $i$.
\end{definition}
Any regular quasidoodle can be obtained from regular doodles by a sequence of
elementary degenerations. Namely, first we move along a generic path in the
space $\K$, up to the first instant when some three points of $\phi(S^1)$ meet
at the same point, forming a regular singularity of degree 2 (we do not watch
the surgeries shown if Figs.~\ref{reidem}a and \ref{reidem}b). Then we consider
the vector subspace in $\K,$ consisting of maps gluing together these three
points of $S^1,$ and go along a generic path in it; at some instant either
another triple point occurs or a fourth branch joins these three. Again, we fix
the smaller subspace, consisting of maps gluing together all the same points,
and move inside it. On the third step a point of multiplicity 5 can occur, or
two points of multiplicities 4 and 3, of 3 points of multiplicity 3, etc. (In
this case we do not watch also the nonessential local moves like the one shown
in Fig.~\ref{noness}.) At the last step we get our quasidoodle.
\begin{figure}
\begin{center}
\unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(75.00,30.00)
\put(0.00,15.00){\vector(1,0){30.00}} \put(15.00,0.00){\vector(0,1){30.00}}
\put(27.00,27.00){\vector(-1,-1){24.00}} \put(3.00,27.00){\vector(1,-1){24.00}}
\put(38.00,15.00){\makebox(0,0)[cc]{$\Longleftrightarrow$}}
\put(45.00,15.00){\vector(1,0){30.00}} \put(60.00,0.00){\vector(0,1){30.00}}
\put(72.00,27.00){\vector(-1,-1){24.00}}
\multiput(49.00,18.00)(-0.12,-0.21){12}{\line(0,-1){0.21}}
\multiput(47.59,15.53)(-0.11,-0.37){5}{\line(0,-1){0.37}}
\multiput(47.05,13.68)(0.12,-0.21){8}{\line(0,-1){0.21}}
\multiput(48.00,12.00)(0.24,-0.10){4}{\line(1,0){0.24}}
\put(48.97,11.61){\line(1,0){1.48}}
\multiput(50.45,11.62)(0.50,0.10){4}{\line(1,0){0.50}}
\multiput(52.44,12.03)(0.36,0.12){7}{\line(1,0){0.36}}
\multiput(54.94,12.85)(0.28,0.12){18}{\line(1,0){0.28}}
\put(72.00,3.00){\vector(1,-1){0.20}}
\multiput(71.00,12.00)(0.12,0.21){12}{\line(0,1){0.21}}
\multiput(72.41,14.47)(0.11,0.37){5}{\line(0,1){0.37}}
\multiput(72.95,16.32)(-0.12,0.21){8}{\line(0,1){0.21}}
\multiput(72.00,18.00)(-0.24,0.10){4}{\line(-1,0){0.24}}
\put(71.03,18.39){\line(-1,0){1.48}}
\multiput(69.55,18.38)(-0.50,-0.10){4}{\line(-1,0){0.50}}
\multiput(67.56,17.97)(-0.36,-0.12){7}{\line(-1,0){0.36}}
\multiput(65.06,17.15)(-0.28,-0.12){18}{\line(-1,0){0.28}}
\multiput(71.00,12.00)(-0.12,-0.16){11}{\line(0,-1){0.16}}
\multiput(69.69,10.27)(-0.12,-0.35){5}{\line(0,-1){0.35}}
\multiput(69.10,8.54)(0.06,-0.87){2}{\line(0,-1){0.87}}
\multiput(69.22,6.81)(0.12,-0.25){7}{\line(0,-1){0.25}}
\multiput(70.05,5.08)(0.11,-0.12){17}{\line(0,-1){0.12}}
\multiput(49.00,18.00)(0.12,0.16){11}{\line(0,1){0.16}}
\multiput(50.31,19.73)(0.12,0.35){5}{\line(0,1){0.35}}
\multiput(50.90,21.46)(-0.06,0.87){2}{\line(0,1){0.87}}
\multiput(50.78,23.19)(-0.12,0.25){7}{\line(0,1){0.25}}
\multiput(49.95,24.92)(-0.11,0.12){17}{\line(0,1){0.12}}
\end{picture}
\end{center}
\caption{Nonessential move of a complicated multiple point} \label{noness}
\end{figure}
Any such sequence of paths is called the {\em degeneration process} of our
regular quasidoodle.
\medskip
\noindent {\bf Examples.} If our quasidoodle has only one triple point, then it
can be obtained by two essentially different processes, corresponding to two
its resolutions shown in Fig.~\ref{reidem}c.
A quasidoodle with one singular point of multiplicity 4 has $16= 4\times 2
\times 2$ essentially different degeneration processes: at the first step any 3
of 4 points can meet at the same point of $\R^2$ (and this can be done in two
different ways, see Fig.~\ref{reidem}c), and on the second the fourth point
joins them in one of two different ways shown in Fig.~\ref{reidem}d.
If our quasidoodle has one more point of multiplicity 3, then there are $96 = 4
\times 3\times 2^3$ different degeneration processes: the points of the second
group can meet before, after, or between of two steps of degeneration of the
first group.
Any degeneration process of a regular quasidoodle and
any invariant of doodles (or I-doodles)
defines a {\it characteristic number}, cf. \S~2 of \cite{V-11}. Namely, if we
have a quasidoodle of some complexity $j$ and a degeneration process $DP$ of
it, then there are exactly two quasidoodles of smaller complexity, whose
degeneration processes coincide with $DP$ without its last step. If at this
last step some new group of multiplicity 3 occurs, then these are two
resolutions of this group shown in Fig.~\ref{reidem}c; if at this step one
branch of $\phi(S^1)$ joined an existing group of multiplicity $\ge 3,$ then we
can move this branch to exactly two sides from this point, see
Fig.~\ref{reidem}d.
In all cases these two resolutions are ordered, i.e. one of them can be called
positive and the other negative. In the first case this order is described in
\S~\ref{coorient}.
In the second we define the index of a multiple point as the arithmetical mean
of indices of points from all neighboring components of the complement of
$\phi(S^1)$, and call {\em positive} the side for which the close point of
multiplicity $\ge 3$ has greater index with respect to the curve.
The characteristic number, which an invariant defines at the pair \{a regular
quasidoodle, some its degeneration process\} is equal to the difference of
similar numbers at two corresponding one-step resolutions (at the positive
minus at the negative one) supplied with the same degeneration process less its
last step. (For these easier singularities these characteristic numbers are
well defined by the inductive conjecture.)
\begin{definition}
An invariant is {\em of order} $j$ if for any quasidoodle, whose complexity is
greater than $j$, and any its degeneration process the corresponding
characteristic number is equal to 0.
\end{definition}
\begin{proposition}[cf. Theorem 4 in \cite{V-11}]
\label{ordofind} Any index-type invariant $M(\beta)$ from \S~\ref{itype} is of
order $\beta+1$. \quad $\Box$
\end{proposition}
A wide class of invariants, generalizing these from \S~\ref{itype}, was
constructed by A.~B.~Merkov, see \cite{Merx-2}. A further generalization of
these invariants, \cite{Merx-4}, classifies all doodles up to equivalence; in
particular it is as strong as entire space of all finite-order invariants.
\subsection{Coding and calculation of finite-order invariants.}
\label{codcal}
Of course, the characteristic numbers of a degeneration process (and hence also
the notion of the order) depend not of its geometrical realization, but only of
some discrete data related with it, such as the combinatorial type of the set
of points in $S^1$ pasted together at different its steps. Let us describe
these data.
Let $A$ be a finite series of integer numbers $A = (a_1 , a_2, \ldots , a_m)$,
all of which are $\ge 3.$ Denote by $|A|$ the number $a_1 + \ldots + a_m$, and
by $\#A$ the number of elements $a_l$ of the series $A$ (denoted in the
previous line by $m$).
\begin{definition}
An $A$-{\it configuration} is a collection of $|A|$ pairwise different points
in $S^1$ divided into groups of cardinalities $a_1, \ldots, a_{\#A}$. Two
$A$-configurations are {\it equivalent} if they can be transformed one into the
other by an orientation-preserving diffeomorphism of $S^1$. A quasidoodle
$\phi:S^1 \to {\bf R}^2$ {\it respects} an $A$-configuration if it sends any of
corresponding $\#A$ groups of points into one point in ${\bf R}^2$. $\phi$ {\it
strictly respects} the $A$-configuration if, moreover, all these $\#A$ points
in ${\bf R}^2$ are distinct, have no extra preimages than these $|A|$ points,
and $\phi$ has no extra points in ${\bf R}^2$ at which images of three of more
different points of $S^1$ meet.
\end{definition}
Obviously, the space of all quasidoodles respecting a given
$A$-con\-fi\-gu\-ra\-tion $J$ is a linear subspace of codimension $2(|A|-\#A)$
in the space $\K$ of all quasidoodles. The number $|A|-\#A$ is thus called the
{\em complexity} of the configuration. We shall denote this subspace by
$\chi(J)$. The set of all quasidoodles, which {\em strictly} respect this
configuration, is an open dense subset in this subspace.
\begin{definition}
A {\em degeneration mode} of an $A$-configuration is some arbitrary order of
marking all the points of the configuration, satisfying the following
conditions: on any step we mark either some three points of some of $\#A$
groups (if none point of the same group is already marked) or one point of a
group, some three or more points of which are already marked.
\end{definition}
Any degeneration process of a regular quasidoodle $\phi$ defines in the obvious
way a degeneration mode of the $A$-configuration strictly respected by $\phi.$
Let $M$ be an invariant of doodles, and $J$ an $A$-configuration.
\begin{proposition}[cf. Theorem 5 in \cite{V-11}]
\label{symbol} If $M$ is an invariant of order $i$, and $|A| - \#A = i,$ then
for any regular quasidoodle $\phi,$ which strictly respects the
$A$-con\-fi\-gu\-ra\-ti\-on $J$, any characteristic number defined by the
triple consisting of $M$, $\phi$ and a degeneration process of $\phi$, depends
only on the pair consisting of the configuration $J$ and its degeneration mode
defined by this degeneration process. \quad $\Box$
\end{proposition}
\noindent {\bf Corollary.} {\em For any $j$ the group of order $j$ invariants
of doodles or I-doodles is finitely generated.}
\medskip
Indeed, the number of its generators does not exceed the sum (over all
equivalence classes of $A$-configurations of complexity $\le i$) of numbers of
their degeneration modes. \quad $\Box$
\medskip
Any invariant of order $i$ can be encoded by its {\it characteristic table}
which we now describe.
This table has $i+1$ levels numbered by $0, 1, \ldots, i$. The $l$-th level
consists of several cells, which are in one-to-one correspondence with all
possible pairs consisting of
a) an equivalence class of $A$-configurations of complexity $l$ in $S^1$,
b) a degeneration mode of this $A$-configuration.
In each cell we indicate
a) a picture (or a code) representing a ``model'' regular quasidoodle, which
strictly respects some $A$-confi\-gu\-ra\-ti\-on from the corresponding
equivalence class (this picture is the same for all invariants),
b) a degeneration process of this quasidoodle, defining this degeneration mode
(also not depending on the invariant), and
c) the characteristic number, which our invariant and the degeneration process,
corresponding to the cell, assign to this quasidoodle.
By the Proposition~\ref{symbol}, we may not specify the pictures and
degeneration processes in the cells of the highest ($i$-th) level of the table:
indeed, the corresponding characteristic numbers depend only on the data
indexing the cell.
For instance, the $0$-th level consists of the trivial $\bigcirc$-like doodle,
and the corresponding characteristic number equals $0$ (we can normalize all
invariants so that they take zero value on the trivial doodle). The $1$-st
level is empty, because there are no configurations of complexity $1$.
\medskip
Having these data, we can calculate our invariant by the inductive process,
coinciding identically with the one described in \S~3 of \cite{V-11} or \S~4.2
of \cite{V-7} (where, however, the characteristic numbers sometimes are called
"actuality indices", and all (quasi)doodles should be replaced by
(quasi)ornaments or (singular) knots).
\begin{remark}
Of course, the characteristic numbers corresponding to different degeneration
modes of the same regular quasidoodle satisfy some natural relations. For
instance, if two degeneration modes differ only by a reordering of markings,
preserving their order inside any group of the $A$-configuration, then the
corresponding characteristic numbers coincide.
Less trivial identities, relating different degeneration modes inside the same
group, follow from the differentials in the chain complex of {\em connected
hypergraphs}, see \cite{V-10}, \cite{BW}.
\end{remark}
\section{Invariants of doodles in terms of the resolved discriminant}
We shall work with the space $\K\equiv C^\infty(S^1,\R^2)$ as with an Euclidean
space of a very large but finite dimension $\Delta$. The justification of this
assumption uses the finite-dimensional approximations of this space and is
similar to that given in \cite{V-11}, \cite{V-7}. A rigorous reader can
everywhere below consider $\K$ as a generic finitedimensional subspace in
$C^\infty(S^1, \R^2).$
\medskip
In particular, we shall use the Alexander duality formula
\begin{equation}
\label{alex} \tilde H^i(\K \setminus \Sigma) \simeq \bar
H_{\Delta-1-i}(\Sigma),
\end{equation}
where $\tilde H^*$ is the usual reduced cohomology group (we are especially
interested in the group $\tilde H^0(\K \sm \Sigma)$ of invariants taking zero
value on the trivial doodle), and $\bar H_*$ is the {\em Borel--Moore homology
group}, i.e. the homology group of the one-point compactification reduced
modulo the added point.
\subsection{Simplicial resolution of the discriminant variety.}
Denote by $\Psi$ the {\em configuration space} of all unordered collections of
three points in $S^1,$ so that $\Psi=(S^1)^3/S(3).$ It is a smooth
3-dimensional manifold with corners, homotopy equivalent to $S^1$. More
precisely, it is the space of an orientable fiber bundle over $S^1,$ whose
projection $p$ sends a triple of points to their sum in the Lie group $S^1,$
the fiber is a closed filled triangle, and the monodromy over the entire base
$S^1$ provides the cyclic permutation of sides and vertices of the triangles.
The "zero section", consisting of centers of these fibers, consists of
configurations, all whose points are at the distance $2\pi/3$ one from the
other.
Let us fix a space $\R^N$ of a huge dimension (much greater than that of the
space $\K$) and fix a generic embedding $\lambda: \Psi \to \R^N.$ For any point
$\phi \in \Sigma$ consider all such points $\{x,y,z\} \in \Psi$ that one of
three conditions holds:
a) $x \ne y \ne z \ne x$ and $\phi(x)=\phi(y)=\phi(z)$;
b) $x=z \ne y$ and $\phi'(x)=0, \phi(x)=\phi(y)$;
c) $x=y=z$ and $\phi'(x)=\phi''(x)=0.$
Then consider all points $\lambda(\{x,y,z\}) \in \R^N$ for all such triples
$\{x,y,z\}$. Since our embedding is generic (and $N$ is sufficiently large)
then for any $\phi \in \Sigma$ the convex hull of all such points is a simplex
with vertices at all these points. (Using the {\em generic} finite-dimensional
approximations of the space $\K$ we can ignore the situation when the number of
such triples is infinite, moreover, we can assume that the number of such
triples has a finite upper estimate, uniform over all $\phi \in \Sigma.$)
Denote this simplex by $\sigma(\phi).$
Finally, define the {\em resolution set} $\sigma \subset \K \times \R^N$ as the
union of all simplices of the form $\phi \times \sigma(\phi)$ over all $\phi
\in \Sigma$.
\medskip
The obvious projection $\K \times \R^N \to \K$ provides the map $\pi: \sigma
\to \Sigma.$ By definition, this map is surjective, and by the previous
"finiteness assumption" it is also proper.
\begin{proposition}[cf. \cite{V-7}, \cite{V-11}]
\label{homeq} The map $\pi$ provides the homotopy equivalence of one-point
compactifications of spaces $\sigma$ and $\Sigma$. \quad $\Box$
\end{proposition}
In particular, the Borel--Moore homology groups (see (\ref{alex})) of these
spaces are canonically isomorphic.
\subsection{$A$-cliques and the main filtration of the resolved discriminant.}
\label{mainfil}
The space $\sigma$ admits a natural filtration, which can be defined in two
equivalent ways. To do it, we need to extend the notion of an $A$-configuration
used in \S~\ref{codcal}. Again, let $A$ be a finite series of $\#A$ integer
numbers $A = (a_1 , a_2, \ldots , a_{\#A})$, all of which are $\ge 3.$
\begin{definition} An $A$-{\em clique} is an unordered collection of
$a_1 + \cdots + a_{\#A}$ points in $S^1,$ divided into groups of cardinalities
$a_1, \ldots, a_{\#A},$ such that a) points of different groups do not coincide
geometrically; b) points inside a group can coincide, but with multiplicity at
most 3. Again, the {\em complexity} of a clique $J$ is the number $|A|-\#A$;
another important characteristic, $\rho(J)$, is the number of geometrically
distinct points in it, i.e. the dimension of the space of cliques equivalent to
it.
The map $\phi:S^1 \to \R^2$ {\em respects} an $A$-clique if it glues together
all geometrically distinct points inside any its group, satisfies the condition
$\phi'=0$ at all points of multiplicity 2, and satisfies the condition $\phi'=
\phi''= 0$ at all points of multiplicity 3. $\phi$ {\em strictly respects} it,
if additionally it does not respect any cliques of larger complexity.
\end{definition}
For any $A$-clique $J$, consider all triples of points in $S^1$ belonging to
the same its group. All such triples are the points of the configuration space
$\Psi.$ Consider the images in $\R^N$ of all these points under the embedding
$\lambda$ and define the simplex $\sigma(J) \subset \K \times \R^N $ as the
convex hull of all such points.
\begin{example} Suppose that $\#A=1,$ $a_1=4,$ and the unique group
of our $A$-clique is a quadruple of points $(x,x,y,z),$ exactly two of which
coincide. Then the simplex $\sigma(J)$ is a triangle with 3 vertices $(x,x,y)$,
$(x,x,z)$, $(x,y,z)$.
\end{example}
Now, for any natural $i$ we take all quasidoodles, {\em strictly} respecting
all possible $A$-cliques of complexities $\le i,$ then consider the union of
their complete preimages in $\sigma$ and, finally, define the term $\sigma_i$
of our {\em main filtration} as the closure of this union. It contains also
some points of the form $\phi \times \theta \in \K \times \R^N,$ where $\phi$
is a quasidoodle of complexity $ >i$, and $\theta$ is some boundary point of
the corresponding simplex $\sigma(\phi).$
Equivalently, for any $A$-clique $J$ we can define the linear subspace $\chi(J)
\subset \K$ consisting of all quasidoodles respecting (strictly or not) this
clique. The term $\sigma_i \subset \sigma$ of the main filtration is then
defined as the union of all subsets $\chi(J) \times \sigma(J) \subset \K \times
\R^N$ over all cliques $J$ of complexity $\le i$. It is easy to see that these
two definitions of the main filtration are equivalent.
\begin{definition} An element of the group
$\bar H_*(\Sigma) \equiv \bar H_*(\sigma)$ is {\em of order} $i$ if it can be
realized by a locally finite cycle lying in the term $\sigma_i$ of this
filtration. In particular, an invariant of doodles is of order $i$ if its class
in the group (\ref{alex}) can be realized as a linking number with the direct
image of a cycle lying in $\sigma_i.$
\end{definition}
\begin{proposition}[cf. \cite{V-11}, Theorem 7]
\label{equival} The last definition of the order of invariants is equivalent to
that given in \S~\ref{elemdef}. \quad $\Box$
\end{proposition}
\section{Calculation of invariants of doodles}
Consider the spectral sequence $E^r_{p,q}$ calculating the Borel--Moore
homology group of the space $\sigma$ and generated by our filtration. Its term
$E^1_{p,q}$ is isomorphic to $\bar H_{p+q}(\sigma_p \sm \sigma_{p-1})$.
\begin{proposition}[cf. \cite{V-11}]
\label{dimres} All groups $E^1_{p,q}$ with $p+q \ge \Delta$ are equal to 0.
\end{proposition}
The proof of this proposition will be given in \S~\ref{revbl}.
\medskip
Hence, for the calculation of the group of invariants (by (\ref{alex})
coinciding with $\bar H_{\Delta-1}(\sigma)$) only the $(\Delta-1)-$ and
$(\Delta-2)$-dimensional Borel--Moore homology groups of these spaces $\sigma_i
\sm \sigma_{i-1}$ are interesting.
In this section we calculate these groups for $i \le 4.$
\subsection{Stratification of the resolved discriminant.}
\label{primstrat}
By construction, any space $\sigma_i \sm \sigma_{i-1}$ consists of several
$\J$-{\em blocks}, numbered by all equivalence classes $\J$ of $A$-cliques of
complexity exactly $i$. Given such an equivalence class $\J,$ the corresponding
block $B(\J)$ is the space of a fiber bundle, whose base is the space of all
$A$-cliques $J$ of this class, and the fiber over a clique $J$ is the direct
product of
a) a linear subspace of codimension $2i$ in $\K$, consisting of all
quasidoodles respecting $J$ (the vector bundle of such subspaces is always
orientable) and
b) a dense subset of the simplex $\sigma(J)$ (namely, this simplex minus some
its faces, which may belong to $\sigma_{i-1}$\footnote{If $J$ is an
$A$-configuration, i.e. has no multiple points, then these are exactly the
faces corresponding to {\em not connected} 3-hypergraphs, see \cite{V-10},
\cite{BW}}).
\medskip
By the Thom isomorphism, the Borel--Moore homology group $\bar H_*$ of any such
block is canonically isomorphic to the group $\bar H_{*-(\Delta-2i)}$ of the
space of only the second bundle of complexes b). Such spaces will be called the
{\em reduced} $\J$-blocks.
The bases of these fiber bundles are $\rho(\J)$-dimensional manifolds, where
$\rho(\J)$ is the number of geometrically distinct points in any clique $J$ of
the class $\J.$
\subsection{The auxiliary filtration.}
\begin{definition} We introduce the {\em auxiliary filtration}
in the space $\sigma_i \sm \sigma_{i-1}$, defining its term $F_\alpha$ as the
union of all above-described blocks over all classes $\J$ of $A$-cliques such
that $\rho(\J) \le \alpha.$
\end{definition}
\begin{example} The term $\sigma_2$ of the main filtration
consists of exactly 3 terms $F_1 \subset F_2 \subset F_3 \equiv \sigma_2$ of
the auxiliary filtration, because the $(3)$-cliques $\{x,y,z\}$ can consist of
1, 2 or 3 geometrically different points.
\end{example}
\begin{definition} The spectral sequence, calculating the group
$\bar H_*(\sigma_i \sm \sigma_{i-1})$ and generated by this auxiliary
filtration, is called the {\em auxiliary spectral sequence} in contrast to the
{\em main} one generated by the main filtration and calculating the homology
groups of entire $\sigma$.
\end{definition}
\subsection{Columns $p=1$ and $p=2$ of the main spectral sequence.}
The term $\sigma_1$ is empty, because there are no cliques of complexity 1.
\prop \label{si2} The colum $\{p=2\}$ of the term $E^1$ of the main spectral
sequence has only the following nontrivial elements: $E^1_{2,\Delta-5} \simeq
\Z$ and $E^1_{2,\Delta-6} \simeq \Z.$ \eprop
\begin{proof}
The term $\sigma_2$ is the space of a fiber bundle, whose base is the
configuration space $\Psi,$ and the fiber over its point $\{x,y,z\}$ is the
linear subspace of codimension 4 in $\K$ consisting of all quasidoodles
respecting the corresponding $(3)$-clique. This bundle is orientable, therefore
$\bar H_*(\sigma_2) = \bar H_{*-\Delta+4}(\Psi) \simeq H_{*-\Delta+4}(S^1).$
\end{proof}
\subsection{Revised resolution.}
\label{revbl}
Before continuing, we improve slightly the construction of our $\J$-blocks,
described in the subsection \ref{primstrat}. Namely, any of these blocks
contains a deformation retract having the same Borel--Moore homology group; it
will be convenient to consider these new blocks $VB(\J)$, which we (in
accordance with the terminology of \cite{V-11}) shall call the {\em visible
blocks.}
Again, any such block, corresponding to an equivalence class $\J$ of
$A$-cliques, is the space of a fibered product of two bundles, whose base and
the first factor of the fiber are the same as previously (i.e. respectively the
space of all cliques $J \in \J$ and an oriented vector space of dimension
$\Delta-2(|A|-\#A)$). However, the second factors of the fibers of this new
bundle are some subcomplexes of the barycentric subdivision of the
corresponding fibers of the former bundle of simplices.
Namely, consider any $A$-clique $J$ and the corresponding simplex $\sigma(J)$.
Any vertex of this simplex, i.e. a triple of points of our $A$-clique lying
inside one its group, defines a vector subspace of codimension $4$ in $\K$.
Consider all these subspaces corresponding to our clique. (E.g., if the clique
is an $A$-configuration, then there are exactly $\sum_{i=1}^{\#A}
\binom{a_i}{3}$ such different subspaces.) These subspaces together with all
their possible intersections form a partially ordered set $\Pi(J)$ with respect
to the relation of the (inverse) inclusion. This poset has unique maximal
element: the subspace $\chi(J)$, i.e. the intersection of all our subspaces.
Having such a partially ordered set, we can define its {\em order complex}
$\Diamond(J)$ (see e.g. \cite{B}): this is a formal simplicial complex, whose
simplices are all the strictly monotone sequences of elements of our poset.
E.g., if the $A$-clique $J$ is an $A$-configuration, then such simplices of the
maximal dimension in $\Diamond (J)$ are nothing but the degeneration modes of
$J$ described in \S~\ref{codcal}.
This order complex $\Diamond(J)$ can be naturally considered as a subcomplex of
the barycentric subdivision of the simplex $\sigma(J)$ (while its maximal
element $\{\chi(J)\}$ corresponds to the center of $\sigma(J)$).
If the complexity $|A|-\#A$ of $J$ is equal to $j$, then the set $\Diamond(J)
\cap \sigma_{j-1}$ of "marginal" faces of this complex consists of all its
faces not containing the main vertex $\{\chi(J)\}$.
Define the {\em visible block} $VB(\J)$ corresponding to our equivalence class
$\J$ of $A$-cliques as the subset of the previously defined block $B(\J)$,
described in \S~\ref{primstrat}, in which elements of the fibers $\sigma(J)$
should belong to the subcomplex $\Diamond(J) \subset \sigma(J).$
Define the revised resolved discriminant $\Diamond \subset \sigma$ as the union
of all such revised blocks $VB(\J)$. It has a natural {\em main} filtration
$\{\Diamond_i\}$ induced by the identical embedding from the main filtration in
$\sigma$. Similarly, in any space $\Diamond_i \sm \Diamond_{i-1} $ the
auxiliary filtration is induced from that in $\sigma_i \sm \sigma_{i-1}.$
\begin{proposition}[see \cite{V-11}, \cite{Phasis}]
\label{homeq2} The inclusion $\Diamond \hookrightarrow \sigma$ induces a
homotopy equivalence of one-point compactifications of these spaces. The same
is true for the inclusion $\Diamond_j \hookrightarrow \sigma_j$ of any terms of
the main filtrations and also for the inclusions $VB(\J) \hookrightarrow B(\J)$
of blocks in them corresponding to the same classes of equivalent $A$-cliques
of complexity $j$. In particular this inclusion induces an isomorphism of both
the main and auxiliary spectral sequences calculating the Borel--Moore homology
groups of all these objects. \quad $\Box$
\end{proposition}
The Proposition \ref{dimres} follows immediately from this one and the
following lemma.
\medskip
\begin{lemma} The dimension of any block $VB(\J)$
of $\Diamond$ does not exceed $\Delta-1.$
\end{lemma}
\begin{proof}
This dimension consists of three numbers:
a) the dimension $\rho(\J) $ of the space of cliques of our class (which does
not exceed the number $|A|$);
b) the dimension $\Delta-2(|A|-\#A)$ of the standard fiber $\chi(J)$ of the
first (vector) bundle;
c) the dimension of the order complex of subspaces associated with the clique
$J$.
The last dimension is equal to the length of the maximal monotone chain of
subspaces constituting our poset $\Pi(J)$ minus 1. This length is equal to
$|A|-2\#A$. Indeed, all our subspaces are of even codimension in $\K$, their
maximal element is the space $\chi(J)$ of codimension $2(|A|-\#A)$, and there
are exactly $\#A$ steps in the chain, when the dimension jumps by 4 (when we
start a new group of points; the very first step is one of them).
Finally, we obtain that the dimension of our block is not greater than $|A| +
\Delta-2(|A|-\#A)+ |A|-2\#A-1.$
\end{proof}
\noindent {\bf Corollary.} Calculating the invariants of doodles, we can
consider only the $\J$-blocks corresponding to such $A$-cliques $J$, that
$|A|-\rho(A) \le 1,$ i.e. either all their points are geometrically distinct or
there is at most one point of multiplicity 2. Moreover, the blocks of the
latter kind can provide only relations in the group of invariants $\bar
H_{\Delta-1}(\sigma)$, but not its generators.
\subsection{The third column of the spectral sequence.}
\thm \label{s3} The group $E^1_{3,q} = \bar H_{3+q}(\Diamond_3 \sm \Diamond_2)$
of the main spectral sequence is trivial for all numbers $q$ not equal to
$\Delta-6$ and $\Delta-7$, while $E^1_{3,\Delta-6} \simeq \Z \simeq
E^1_{3,\Delta-7}.$ \etheorem
The proof of this theorem occupies the rest of this subsection.
\medskip
The term $\Diamond_3 \sm \Diamond_2$ consists of exactly 4 $\J$-blocks
corresponding to different equivalence classes $\J$ of cliques of complexity 3:
the main block A (all 4 points in the cliques are distinct), the block B of
auxiliary filtration 3 (exactly two points coincide), and two blocks of
auxiliary filtration 2: C (one simple point and one point of multiplicity 3)
and D (two double points).
Consider the corresponding reduced blocks (see \S~\ref{primstrat}) $[A], [B],
[C]$ and $[D].$
\medskip
The main reduced block $[A]$ is the space of a fiber bundle, whose base $\J$ is
the configuration space $B(S^1,4)$ of subsets of cardinality 4 in $S^1,$ and
the fiber is a cross with its four endpoints removed. The vertices of such a
cross correspond to all possible choices of some 3 points of the 4-point
configuration; this notation is transparent in Fig.~\ref{str3dif}a.
The base $B(S^1,4),$ in its turn, is the space of the fiber bundle
\begin{equation}
\label{projconf} p: B(S^1,4) \to S^1
\end{equation}
where the projection sends a quadruple of points in $S^1 = \R/\Z$ to their sum
(mod $\Z$), and the fiber is diffeomorphic to an open 3-dimensional disc. The
monodromy over the basis circle of (\ref{projconf}) violates the orientation of
the bundle of 3-dimensional discs and acts on the bundle of crosses as a cyclic
permutation of their edges.
Thus the Wang exact sequence of this bundle gives us the following assertion.
\prop \label{mstrat3} The Borel--Moore homology group of the reduced main block
$[A]$
of $\Diamond_3\sm \Diamond_2$
is nontrivial only in dimensions $5$ and $4$ and is isomorphic to $\Z$ in these
dimensions. The generator of the $5$-dimensional group is swept out by the
bundle over $B(S^1,4)$ of 1-chains shown in Fig.~\ref{one}b $\equiv$
Fig.~\ref{str3dif}a (i.e. the unique chains antiinvariant under the monodromy
action). The generator of the $4$-dimensional group is swept out by the fiber
bundle, whose base is the fiber $p^{-1}(0)$ of the bundle (\ref{projconf}), and
fibers are 1-chains shown in these pictures by the sum of two right-hand
arrows. \quad $\Box$ \eprop
\begin{figure}
\begin{center}
\unitlength 0.87mm \linethickness{0.4pt}
\begin{picture}(143.00,57.00)
\put(3.00,5.00){\circle*{2.00}} \put(8.00,5.00){\circle*{2.00}}
\put(8.00,10.00){\circle*{2.00}} \put(3.00,10.00){\circle*{2.00}}
\put(27.00,29.00){\rule{2.00\unitlength}{2.00\unitlength}}
\put(43.00,15.00){\vector(-1,1){14.00}}
\put(27.00,29.00){\vector(-1,-1){14.00}}
\put(13.00,45.00){\vector(1,-1){14.00}} \put(29.00,31.00){\vector(1,1){14.00}}
\put(48.00,5.00){\circle*{2.00}} \put(53.00,5.00){\circle*{2.00}}
\put(53.00,10.00){\circle*{2.00}} \put(48.00,10.00){\circle*{2.00}}
\put(3.00,50.00){\circle*{2.00}} \put(8.00,50.00){\circle*{2.00}}
\put(8.00,55.00){\circle*{2.00}} \put(3.00,55.00){\circle*{2.00}}
\put(48.00,50.00){\circle*{2.00}} \put(53.00,50.00){\circle*{2.00}}
\put(53.00,55.00){\circle*{2.00}} \put(48.00,55.00){\circle*{2.00}}
\multiput(55.00,45.00)(-0.12,0.12){100}{\line(0,1){0.12}}
\put(43.00,57.00){\line(1,0){12.00}}
\put(55.00,57.00){\line(0,-1){12.00}}
\multiput(55.00,15.00)(-0.12,-0.12){100}{\line(0,-1){0.12}}
\put(43.00,3.00){\line(1,0){12.00}}
\put(55.00,3.00){\line(0,1){12.00}}
\multiput(1.00,15.00)(0.12,-0.12){100}{\line(0,-1){0.12}}
\put(13.00,3.00){\line(-1,0){12.00}}
\put(1.00,3.00){\line(0,1){12.00}}
\multiput(1.00,45.00)(0.12,0.12){100}{\line(0,1){0.12}}
\put(13.00,57.00){\line(-1,0){12.00}}
\put(1.00,57.00){\line(0,-1){12.00}}
\put(28.00,2.00){\makebox(0,0)[cc]{{\large a}}}
\put(80.00,27.00){\circle*{2.00}} \put(80.00,33.00){\circle*{2.00}}
\put(86.00,30.00){\circle*{2.00}} \put(88.00,30.00){\circle*{2.00}}
\put(87.00,28.00){\line(0,1){4.00}}
\multiput(87.00,32.00)(-0.26,0.12){34}{\line(-1,0){0.26}}
\put(78.00,36.00){\line(0,-1){12.00}}
\multiput(78.00,24.00)(0.26,0.12){34}{\line(1,0){0.26}}
\put(109.00,29.00){\rule{2.00\unitlength}{2.00\unitlength}}
\put(133.00,47.00){\circle*{2.00}} \put(133.00,53.00){\circle*{2.00}}
\put(139.00,50.00){\circle*{2.00}} \put(141.00,50.00){\circle*{2.00}}
\put(133.00,7.00){\circle*{2.00}} \put(133.00,13.00){\circle*{2.00}}
\put(139.00,10.00){\circle*{2.00}} \put(141.00,10.00){\circle*{2.00}}
\put(143.00,8.00){\line(-1,0){5.00}}
\multiput(138.00,8.00)(-0.21,0.12){34}{\line(-1,0){0.21}}
\put(131.00,12.00){\line(0,1){4.00}}
\multiput(131.00,16.00)(0.29,-0.12){42}{\line(1,0){0.29}}
\put(143.00,11.00){\line(0,-1){3.00}}
\put(143.00,52.00){\line(-1,0){5.00}}
\multiput(138.00,52.00)(-0.21,-0.12){34}{\line(-1,0){0.21}}
\put(131.00,48.00){\line(0,-1){4.00}}
\multiput(131.00,44.00)(0.29,0.12){42}{\line(1,0){0.29}}
\put(143.00,49.00){\line(0,1){3.00}}
\put(128.00,12.00){\vector(-1,1){17.00}}
\put(111.00,31.00){\vector(1,1){17.00}}
\put(109.00,30.00){\line(-1,0){17.00}}
\put(111.00,2.00){\makebox(0,0)[cc]{{\large b}}}
\end{picture}
\vspace{1.5cm} \unitlength 0.87mm \linethickness{0.4pt}
\begin{picture}(143.00,12.00)
\put(53.00,10.00){\circle*{2.00}} \put(61.00,10.00){\circle*{2.00}}
\put(63.00,10.00){\circle*{2.00}} \put(65.00,10.00){\circle*{2.00}}
\put(3.00,10.00){\circle*{2.00}} \put(11.00,10.00){\circle*{2.00}}
\put(13.00,10.00){\circle*{2.00}} \put(15.00,10.00){\circle*{2.00}}
\put(14.00,12.00){\line(0,-1){4.00}}
\put(1.00,8.00){\line(0,1){4.00}}
\put(59.00,8.00){\line(0,1){4.00}}
\put(59.00,12.00){\line(1,0){8.00}}
\put(67.00,12.00){\line(0,-1){4.00}}
\put(67.00,8.00){\line(-1,0){8.00}}
\put(33.00,9.00){\rule{2.00\unitlength}{2.00\unitlength}}
\put(34.00,2.00){\makebox(0,0)[cc]{{\large c}}}
\put(14.00,8.00){\line(-1,0){13.00}}
\put(1.00,12.00){\line(1,0){13.00}}
\put(83.00,10.00){\circle*{2.00}} \put(85.00,10.00){\circle*{2.00}}
\put(92.00,10.00){\circle*{2.00}} \put(94.00,10.00){\circle*{2.00}}
\put(130.00,10.00){\circle*{2.00}} \put(132.00,10.00){\circle*{2.00}}
\put(139.00,10.00){\circle*{2.00}} \put(141.00,10.00){\circle*{2.00}}
\put(131.00,8.00){\line(0,1){4.00}}
\put(131.00,12.00){\line(1,0){12.00}}
\put(143.00,12.00){\line(0,-1){4.00}}
\put(143.00,8.00){\line(-1,0){12.00}}
\put(93.00,8.00){\line(0,1){4.00}}
\put(93.00,12.00){\line(-1,0){12.00}}
\put(81.00,12.00){\line(0,-1){4.00}}
\put(81.00,8.00){\line(1,0){12.00}}
\put(111.00,9.00){\rule{2.00\unitlength}{2.00\unitlength}}
\put(113.00,10.00){\vector(1,0){13.00}} \put(98.00,10.00){\vector(1,0){13.00}}
\put(112.00,2.00){\makebox(0,0)[cc]{{\large d}}}
\put(19.00,10.00){\vector(1,0){14.00}} \put(35.00,10.00){\vector(1,0){14.00}}
\end{picture}
\end{center}
\caption{Order complexes for blocks in $\Diamond_3$} \label{str3dif}
\end{figure}
The block $[B]$ also is the space of a (trivial) fiber bundle, whose base is
the space of all configurations of 3 points in $S^1,$ one of which (the double
point) is distinguished, and the fiber is a star with 3 rays without endpoints.
These fibers and their endpoints are shown in Fig.~\ref{str3dif}b: two
right-hand endpoints correspond to the subspaces in $\K$ given by the
conditions of the form $\phi(x)=\phi(y), \phi'(x)=0$ and $\phi(x)=\phi(z),
\phi'(x)=0$, while the left endpoint corresponds to the equation
$\phi(x)=\phi(y)= \phi(z)$, cf. Example in \S~\ref{mainfil}.
The base of this bundle is diffeomorphic to the direct product $S^1 \times
B^2$, where $B^2$ is an open 2-dimensional disc. Monodromy along the basic
circle $S^1$ acts trivially on the bundle of 3-stars. Therefore we have the
following statement.
\prop \label{strB3} The Borel--Moore homology group of the reduced block $[B]$
is nontrivial only in dimensions $4$ and $3$ it is isomorphic to $\Z^2$ in both
these dimensions. \quad $\Box$ \eprop
In a similar way we get the following statements.
\prop \label{strC3} The Borel--Moore homology group of the reduced block $[C]$
is nontrivial only in dimensions $3$ and $2$ and is isomorphic to $\Z$ in both
these dimensions. \quad $\Box$ \eprop
\prop \label{strD3} The Borel--Moore homology group of the reduced block $[D]$
is nontrivial only in dimensions $3$ and $2$ and is isomorphic to $\Z$ in both
these dimensions. \quad $\Box$ \eprop
The corresponding order complexes are shown in Figs.~\ref{str3dif}c and
\ref{str3dif}d, respectively.
\medskip
\noindent {\bf Corollary.} {\em The term $E^1$ of the spectral sequence,
calculating the Borel--Moore homology group of the (not reduced) term
$\Diamond_3 \sm \Diamond_2$ of the main filtration of our resolved discriminant
and generated by the auxiliary filtration in this term, looks as is shown in
Fig.~\ref{ss3a} (i.e., all its cells $E^1_{p,q}$ other than six indicated there
are trivial).}
\begin{figure}
\begin{center}
\unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(52.00,24.00)
\put(12.00,6.00){\vector(0,1){18.00}} \put(12.00,6.00){\vector(1,0){40.00}}
\put(9.00,22.00){\makebox(0,0)[cc]{$q$}}
\put(5.00,9.00){\makebox(0,0)[cc]{$\Delta-6$}}
\put(5.00,15.00){\makebox(0,0)[cc]{$\Delta-5$}}
\put(17.00,2.00){\makebox(0,0)[cc]{2}} \put(17.00,9.00){\makebox(0,0)[cc]{${\bf
Z}^2$}} \put(17.00,15.00){\makebox(0,0)[cc]{${\bf Z}^2$}}
\put(26.00,9.00){\makebox(0,0)[cc]{${\bf Z}^2$}}
\put(26.00,15.00){\makebox(0,0)[cc]{${\bf Z}^2$}}
\put(35.00,15.00){\makebox(0,0)[cc]{${\bf Z}$}}
\put(35.00,9.00){\makebox(0,0)[cc]{${\bf Z}$}}
\put(35.00,2.00){\makebox(0,0)[cc]{4}} \put(26.00,2.00){\makebox(0,0)[cc]{3}}
\put(49.00,2.00){\makebox(0,0)[cc]{$a$}}
\put(12.00,12.00){\line(1,0){33.00}}
\put(45.00,18.00){\line(-1,0){33.00}}
\put(21.00,6.00){\line(0,1){15.00}}
\put(30.00,21.00){\line(0,-1){15.00}}
\put(39.00,6.00){\line(0,1){15.00}}
\end{picture}
\end{center}
\caption{Auxiliary spectral sequence for the column $p=3$.} \label{ss3a}
\end{figure}
\prop \label{acyclABC} The Borel--Moore homology group of the subspace in
$\Diamond_3 \sm \Diamond_2$ formed only by the blocks A, B and C, is acyclic in
all dimensions. \eprop
This follows immediately from the shape of all these blocks and their
generators, and from accounting the limit positions of subspaces in $\K$
corresponding to the cliques of types A and B when they degenerate and form
configurations of types B and C, respectively.
E.g. let us consider the $(4)$-clique :: drawn at any endpoint of
Fig.~\ref{str3dif}a and let the two right-hand points of it move one towards
the other, forming at the last instant a clique of type B. The subspaces of
codimension 2 in $\K,$ corresponding to all endpoints of the cross, tend to
similar subspaces for endpoints of the star of Fig.~\ref{str3dif}b. Namely, to
both left-hand endpoints of the cross $\times$ there corresponds the unique
left-hand endpoint of the star, and other two endpoints "remain unmoved".
Therefore the boundary under this degeneration of the basic cycle shown in
Fig.~\ref{str3dif}a is equal to the basic cycle in $\bar H_{\Delta-3}(the \
block \ B)$ depicted by two arrows in Fig.~\ref{str3dif}b (i.e., swept out by
the fiber bundle of such 1-chains over entire base $S^1 \times B^2$ of this
block).
Other boundary operators $\bar H_*(A) \to \bar H_*(B)$ and $\bar H_*(B) \to
\bar H_*(C)$ can be considered in a similar way and give the assertion of
Proposition \ref{acyclABC}. \quad $\Box$
\medskip
\noindent {\bf Corollary.} {\em The Borel--Moore homology group of the space
$\Diamond_3 \sm \Diamond_2$ coincides with that of unique its block $D$
(described in Proposition \ref{strD3}).}
\medskip
This terminates the proof of Theorem \ref{s3}. \quad $\Box$
\subsection{Invariants of order 4.}
In this subsection, we shall be interested only in the $(\Delta-1)$-dimensional
homology group of the space $\Diamond_4 \sm \Diamond_3,$ which can provide
invariants of doodles.
In accordance with the Corollary of Proposition \ref{homeq2}, we shall consider
only blocks of complexity $4$ having at most one double point.
\prop \label{classcomp4} There are exactly 4 $\J$-blocks of complexity 4
corresponding to classes $\J$ of $A$-configurations (i.e. of $A$-cliques, all
whose points are geometrically distinct). Three of them correspond to
$(3,3)$-cliques consisting of 6 points in $S^1$ separated into two triples in
one of ways shown in Fig.~\ref{two}a. The fourth corresponds to the unique
class of $(5)$-configurations.
Also, there are exactly 5 blocks of complexity 4 corresponding to cliques with
exactly one point of multiplicity 2. Four of them also are of type $A=(3,3)$
and can be obtained from first two pictures in Fig.~\ref{two}a by some
degenerations, see Figs.~\ref{degen}a and \ref{degen}b respectively. The fifth
$\J$-block corresponds to $(5)$-cliques with exactly one double points, see
Fig.~\ref{degen}c. \quad $\Box$ \eprop
\begin{figure}
\unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(124.67,17.67)
\multiput(10.00,17.00)(0.79,-0.09){2}{\line(1,0){0.79}}
\multiput(11.59,16.82)(0.30,-0.11){5}{\line(1,0){0.30}}
\multiput(13.09,16.28)(0.17,-0.11){8}{\line(1,0){0.17}}
\multiput(14.43,15.42)(0.11,-0.11){10}{\line(0,-1){0.11}}
\multiput(15.54,14.27)(0.12,-0.20){7}{\line(0,-1){0.20}}
\multiput(16.37,12.91)(0.10,-0.30){5}{\line(0,-1){0.30}}
\multiput(16.86,11.39)(0.07,-0.80){2}{\line(0,-1){0.80}}
\multiput(17.00,9.80)(-0.11,-0.79){2}{\line(0,-1){0.79}}
\multiput(16.77,8.22)(-0.12,-0.30){5}{\line(0,-1){0.30}}
\multiput(16.19,6.73)(-0.11,-0.16){8}{\line(0,-1){0.16}}
\multiput(15.29,5.42)(-0.13,-0.12){9}{\line(-1,0){0.13}}
\multiput(14.11,4.34)(-0.20,-0.11){7}{\line(-1,0){0.20}}
\multiput(12.72,3.55)(-0.38,-0.11){4}{\line(-1,0){0.38}}
\put(11.19,3.10){\line(-1,0){1.59}}
\multiput(9.60,3.01)(-0.52,0.09){3}{\line(-1,0){0.52}}
\multiput(8.03,3.28)(-0.25,0.10){6}{\line(-1,0){0.25}}
\multiput(6.56,3.90)(-0.16,0.12){8}{\line(-1,0){0.16}}
\multiput(5.27,4.84)(-0.12,0.13){9}{\line(0,1){0.13}}
\multiput(4.22,6.05)(-0.11,0.20){7}{\line(0,1){0.20}}
\multiput(3.48,7.46)(-0.10,0.39){4}{\line(0,1){0.39}}
\put(3.07,9.00){\line(0,1){1.60}}
\multiput(3.03,10.60)(0.11,0.52){3}{\line(0,1){0.52}}
\multiput(3.34,12.16)(0.11,0.24){6}{\line(0,1){0.24}}
\multiput(4.01,13.61)(0.11,0.14){9}{\line(0,1){0.14}}
\multiput(4.98,14.88)(0.14,0.11){9}{\line(1,0){0.14}}
\multiput(6.22,15.89)(0.24,0.12){6}{\line(1,0){0.24}}
\multiput(7.65,16.59)(0.59,0.10){4}{\line(1,0){0.59}}
\put(4.00,7.00){\line(1,0){12.00}}
\multiput(16.00,7.00)(-0.18,-0.12){34}{\line(-1,0){0.18}}
\multiput(10.00,3.00)(-0.18,0.12){34}{\line(-1,0){0.18}}
\multiput(4.00,13.00)(0.35,0.12){26}{\line(1,0){0.35}}
\put(10.00,3.00){\circle*{1.33}} \put(16.00,7.00){\circle*{1.33}}
\put(4.00,7.00){\circle*{1.33}} \put(4.00,13.00){\circle*{1.33}}
\multiput(30.00,17.00)(0.79,-0.09){2}{\line(1,0){0.79}}
\multiput(31.59,16.82)(0.30,-0.11){5}{\line(1,0){0.30}}
\multiput(33.09,16.28)(0.17,-0.11){8}{\line(1,0){0.17}}
\multiput(34.43,15.42)(0.11,-0.11){10}{\line(0,-1){0.11}}
\multiput(35.54,14.27)(0.12,-0.20){7}{\line(0,-1){0.20}}
\multiput(36.37,12.91)(0.10,-0.30){5}{\line(0,-1){0.30}}
\multiput(36.86,11.39)(0.07,-0.80){2}{\line(0,-1){0.80}}
\multiput(37.00,9.80)(-0.11,-0.79){2}{\line(0,-1){0.79}}
\multiput(36.77,8.22)(-0.12,-0.30){5}{\line(0,-1){0.30}}
\multiput(36.19,6.73)(-0.11,-0.16){8}{\line(0,-1){0.16}}
\multiput(35.29,5.42)(-0.13,-0.12){9}{\line(-1,0){0.13}}
\multiput(34.11,4.34)(-0.20,-0.11){7}{\line(-1,0){0.20}}
\multiput(32.72,3.55)(-0.38,-0.11){4}{\line(-1,0){0.38}}
\put(31.19,3.10){\line(-1,0){1.59}}
\multiput(29.60,3.01)(-0.52,0.09){3}{\line(-1,0){0.52}}
\multiput(28.03,3.28)(-0.25,0.10){6}{\line(-1,0){0.25}}
\multiput(26.56,3.90)(-0.16,0.12){8}{\line(-1,0){0.16}}
\multiput(25.27,4.84)(-0.12,0.13){9}{\line(0,1){0.13}}
\multiput(24.22,6.05)(-0.11,0.20){7}{\line(0,1){0.20}}
\multiput(23.48,7.46)(-0.10,0.39){4}{\line(0,1){0.39}}
\put(23.07,9.00){\line(0,1){1.60}}
\multiput(23.03,10.60)(0.11,0.52){3}{\line(0,1){0.52}}
\multiput(23.34,12.16)(0.11,0.24){6}{\line(0,1){0.24}}
\multiput(24.01,13.61)(0.11,0.14){9}{\line(0,1){0.14}}
\multiput(24.98,14.88)(0.14,0.11){9}{\line(1,0){0.14}}
\multiput(26.22,15.89)(0.24,0.12){6}{\line(1,0){0.24}}
\multiput(27.65,16.59)(0.59,0.10){4}{\line(1,0){0.59}}
\put(24.00,7.00){\line(1,0){12.00}}
\multiput(36.00,7.00)(-0.18,-0.12){34}{\line(-1,0){0.18}}
\multiput(30.00,3.00)(-0.18,0.12){34}{\line(-1,0){0.18}}
\put(30.00,3.00){\circle*{1.33}} \put(36.00,7.00){\circle*{1.33}}
\put(24.00,7.00){\circle*{1.33}}
\multiput(56.00,17.00)(0.79,-0.09){2}{\line(1,0){0.79}}
\multiput(57.59,16.82)(0.30,-0.11){5}{\line(1,0){0.30}}
\multiput(59.09,16.28)(0.17,-0.11){8}{\line(1,0){0.17}}
\multiput(60.43,15.42)(0.11,-0.11){10}{\line(0,-1){0.11}}
\multiput(61.54,14.27)(0.12,-0.20){7}{\line(0,-1){0.20}}
\multiput(62.37,12.91)(0.10,-0.30){5}{\line(0,-1){0.30}}
\multiput(62.86,11.39)(0.07,-0.80){2}{\line(0,-1){0.80}}
\multiput(63.00,9.80)(-0.11,-0.79){2}{\line(0,-1){0.79}}
\multiput(62.77,8.22)(-0.12,-0.30){5}{\line(0,-1){0.30}}
\multiput(62.19,6.73)(-0.11,-0.16){8}{\line(0,-1){0.16}}
\multiput(61.29,5.42)(-0.13,-0.12){9}{\line(-1,0){0.13}}
\multiput(60.11,4.34)(-0.20,-0.11){7}{\line(-1,0){0.20}}
\multiput(58.72,3.55)(-0.38,-0.11){4}{\line(-1,0){0.38}}
\put(57.19,3.10){\line(-1,0){1.59}}
\multiput(55.60,3.01)(-0.52,0.09){3}{\line(-1,0){0.52}}
\multiput(54.03,3.28)(-0.25,0.10){6}{\line(-1,0){0.25}}
\multiput(52.56,3.90)(-0.16,0.12){8}{\line(-1,0){0.16}}
\multiput(51.27,4.84)(-0.12,0.13){9}{\line(0,1){0.13}}
\multiput(50.22,6.05)(-0.11,0.20){7}{\line(0,1){0.20}}
\multiput(49.48,7.46)(-0.10,0.39){4}{\line(0,1){0.39}}
\put(49.07,9.00){\line(0,1){1.60}}
\multiput(49.03,10.60)(0.11,0.52){3}{\line(0,1){0.52}}
\multiput(49.34,12.16)(0.11,0.24){6}{\line(0,1){0.24}}
\multiput(50.01,13.61)(0.11,0.14){9}{\line(0,1){0.14}}
\multiput(50.98,14.88)(0.14,0.11){9}{\line(1,0){0.14}}
\multiput(52.22,15.89)(0.24,0.12){6}{\line(1,0){0.24}}
\multiput(53.65,16.59)(0.59,0.10){4}{\line(1,0){0.59}}
\multiput(56.00,3.00)(-0.18,0.12){34}{\line(-1,0){0.18}}
\put(56.00,3.00){\circle*{1.33}} \put(62.00,7.00){\circle*{1.33}}
\put(50.00,7.00){\circle*{1.33}}
\multiput(76.00,17.00)(0.79,-0.09){2}{\line(1,0){0.79}}
\multiput(77.59,16.82)(0.30,-0.11){5}{\line(1,0){0.30}}
\multiput(79.09,16.28)(0.17,-0.11){8}{\line(1,0){0.17}}
\multiput(80.43,15.42)(0.11,-0.11){10}{\line(0,-1){0.11}}
\multiput(81.54,14.27)(0.12,-0.20){7}{\line(0,-1){0.20}}
\multiput(82.37,12.91)(0.10,-0.30){5}{\line(0,-1){0.30}}
\multiput(82.86,11.39)(0.07,-0.80){2}{\line(0,-1){0.80}}
\multiput(83.00,9.80)(-0.11,-0.79){2}{\line(0,-1){0.79}}
\multiput(82.77,8.22)(-0.12,-0.30){5}{\line(0,-1){0.30}}
\multiput(82.19,6.73)(-0.11,-0.16){8}{\line(0,-1){0.16}}
\multiput(81.29,5.42)(-0.13,-0.12){9}{\line(-1,0){0.13}}
\multiput(80.11,4.34)(-0.20,-0.11){7}{\line(-1,0){0.20}}
\multiput(78.72,3.55)(-0.38,-0.11){4}{\line(-1,0){0.38}}
\put(77.19,3.10){\line(-1,0){1.59}}
\multiput(75.60,3.01)(-0.52,0.09){3}{\line(-1,0){0.52}}
\multiput(74.03,3.28)(-0.25,0.10){6}{\line(-1,0){0.25}}
\multiput(72.56,3.90)(-0.16,0.12){8}{\line(-1,0){0.16}}
\multiput(71.27,4.84)(-0.12,0.13){9}{\line(0,1){0.13}}
\multiput(70.22,6.05)(-0.11,0.20){7}{\line(0,1){0.20}}
\multiput(69.48,7.46)(-0.10,0.39){4}{\line(0,1){0.39}}
\put(69.07,9.00){\line(0,1){1.60}}
\multiput(69.03,10.60)(0.11,0.52){3}{\line(0,1){0.52}}
\multiput(69.34,12.16)(0.11,0.24){6}{\line(0,1){0.24}}
\multiput(70.01,13.61)(0.11,0.14){9}{\line(0,1){0.14}}
\multiput(70.98,14.88)(0.14,0.11){9}{\line(1,0){0.14}}
\multiput(72.22,15.89)(0.24,0.12){6}{\line(1,0){0.24}}
\multiput(73.65,16.59)(0.59,0.10){4}{\line(1,0){0.59}}
\put(82.00,7.00){\circle*{1.33}} \put(12.33,16.34){\circle*{1.33}}
\put(13.67,16.00){\circle*{1.33}} \put(36.00,13.00){\circle*{1.33}}
\multiput(36.00,13.00)(-0.42,0.12){23}{\line(-1,0){0.42}}
\put(27.00,16.00){\circle*{1.33}} \put(25.99,15.66){\circle*{1.33}}
\put(62.00,13.33){\circle*{1.33}}
\multiput(62.00,13.33)(-0.12,-0.21){50}{\line(0,-1){0.21}}
\multiput(50.00,7.00)(0.23,0.12){53}{\line(1,0){0.23}}
\multiput(62.00,7.00)(-0.16,0.12){67}{\line(-1,0){0.16}}
\put(52.00,15.67){\circle*{1.33}} \put(51.00,14.66){\circle*{1.33}}
\put(82.00,13.00){\circle*{1.33}}
\multiput(82.00,13.00)(-0.12,-0.12){78}{\line(-1,0){0.12}}
\put(72.00,4.00){\circle*{1.33}} \put(73.34,3.66){\circle*{1.33}}
\multiput(82.00,7.00)(-0.12,0.20){50}{\line(0,1){0.20}}
\multiput(76.00,17.00)(-0.19,-0.12){34}{\line(-1,0){0.19}}
\multiput(69.67,13.00)(0.25,-0.12){50}{\line(1,0){0.25}}
\put(76.00,16.67){\circle*{1.33}} \put(70.00,13.00){\circle*{1.33}}
\multiput(117.00,17.00)(0.79,-0.09){2}{\line(1,0){0.79}}
\multiput(118.59,16.82)(0.30,-0.11){5}{\line(1,0){0.30}}
\multiput(120.09,16.28)(0.17,-0.11){8}{\line(1,0){0.17}}
\multiput(121.43,15.42)(0.11,-0.11){10}{\line(0,-1){0.11}}
\multiput(122.54,14.27)(0.12,-0.20){7}{\line(0,-1){0.20}}
\multiput(123.37,12.91)(0.10,-0.30){5}{\line(0,-1){0.30}}
\multiput(123.86,11.39)(0.07,-0.80){2}{\line(0,-1){0.80}}
\multiput(124.00,9.80)(-0.11,-0.79){2}{\line(0,-1){0.79}}
\multiput(123.77,8.22)(-0.12,-0.30){5}{\line(0,-1){0.30}}
\multiput(123.19,6.73)(-0.11,-0.16){8}{\line(0,-1){0.16}}
\multiput(122.29,5.42)(-0.13,-0.12){9}{\line(-1,0){0.13}}
\multiput(121.11,4.34)(-0.20,-0.11){7}{\line(-1,0){0.20}}
\multiput(119.72,3.55)(-0.38,-0.11){4}{\line(-1,0){0.38}}
\put(118.19,3.10){\line(-1,0){1.59}}
\multiput(116.60,3.01)(-0.52,0.09){3}{\line(-1,0){0.52}}
\multiput(115.03,3.28)(-0.25,0.10){6}{\line(-1,0){0.25}}
\multiput(113.56,3.90)(-0.16,0.12){8}{\line(-1,0){0.16}}
\multiput(112.27,4.84)(-0.12,0.13){9}{\line(0,1){0.13}}
\multiput(111.22,6.05)(-0.11,0.20){7}{\line(0,1){0.20}}
\multiput(110.48,7.46)(-0.10,0.39){4}{\line(0,1){0.39}}
\put(110.07,9.00){\line(0,1){1.60}}
\multiput(110.03,10.60)(0.11,0.52){3}{\line(0,1){0.52}}
\multiput(110.34,12.16)(0.11,0.24){6}{\line(0,1){0.24}}
\multiput(111.01,13.61)(0.11,0.14){9}{\line(0,1){0.14}}
\multiput(111.98,14.88)(0.14,0.11){9}{\line(1,0){0.14}}
\multiput(113.22,15.89)(0.24,0.12){6}{\line(1,0){0.24}}
\multiput(114.65,16.59)(0.59,0.10){4}{\line(1,0){0.59}}
\put(124.00,10.00){\circle*{1.33}} \put(110.00,10.00){\circle*{1.33}}
\put(117.00,17.00){\circle*{1.33}} \put(116.33,3.00){\circle*{1.33}}
\put(117.67,3.00){\circle*{1.33}} \put(109.00,1.00){\makebox(0,0)[cc]{{\large
c}}} \put(66.00,1.00){\makebox(0,0)[cc]{{\large b}}}
\put(20.00,1.00){\makebox(0,0)[cc]{{\large a}}}
\end{picture}
\caption{Degenerated blocks of complexity 4} \label{degen}
\end{figure}
Let us study all these $\J$-blocks and their homology groups.
\prop \label{two_hom} All three $\J$-blocks corresponding to three pictures of
Fig.~\ref{two}a are smooth orientable manifolds diffeomorphic to $S^1 \times
\R^{\Delta-2},$ in particular any of them has only two nontrivial Borel--Moore
homology groups $\bar H_{\Delta-1} \simeq \Z \simeq \bar H_{\Delta-2}.$ \quad
$\Box$ \eprop
More precisely, in the first and the third (respectively, in the second) case
the corresponding space $\J$ of equivalent cliques is diffeomorphic to a
nonorientable (respectively, orientable) fiber bundle over $S^1$ with fiber
$B^5.$ The order complexes $\Diamond(J)$ in all cases are homeomorphic to open
intervals, whose endpoints correspond to subspaces of codimension 2 in $\K$
defined by the $(3)$-cliques forming the triangles, and the center corresponds
to the subspace of codimension 4, defined by their intersection. The bundle of
these intervals is (non)orientable exactly in the same cases when the
corresponding configuration space is.
\prop \label{three_hom} The $(\Delta-1)$-dimensional Borel--Moore homology
group of the $\J$-block in $\Diamond_4 \sm \Diamond_3,$ corresponding to the
unique class $\J$ of $(5)$-{\em configurations}, is equal to $\Z^2.$ \eprop
(I thank very much A.~B.~Merkov, who proved this proposition, and also
suggested the following notation, convenient for the homological study of such
blocks.)
\begin{proof} Let $J$ be a configuration of 5 different points in
$S^1$. The corresponding order complex $\Diamond(J)$ is two-dimensional. Its 20
simplices of dimension 2 are in a natural one-to-one correspondence with the
triples of the form \{some 3 points of $J$; some 4 points containing these
three; all five points\}. Denote such a simplex by the arrow, connecting two
points {\em not participating} in the first triple and directed towards the
point not participating in the quadruple. The one-dimensional simplices of the
same order complex $\Diamond(J)$ are the segments of the following three kinds.
A) connecting (the vertex corresponding to the subspace in $\K$ defined by) a
triple of our points and (the vertex corresponding to its subspace defined by)
a quadruple containing this triple. Such edges do not belong to $\Diamond_4$
and are not interesting for us.
B) connecting (the vertex corresponding to) a triple and (that corresponding
to) the maximal element $\chi(J) \in \Diamond(J)$ (defined by the entire
$(5)$-clique $J$). These 10 edges are denoted by non-oriented edges connecting
two points not participating in the triple.
C) connecting (the vertices corresponding to) a quadruple and $\chi(J)$. These
5 edges are denoted by marking the point not participating in the quadruple.
\medskip
In this notation, the boundary of an arrow (i.e., a 2-simplex of $\Diamond(J)$)
is equal to the edge, obtained from this arrow by forgetting the orientation,
minus the endpoint of the arrow.
\begin{lemma}
The Borel--Moore homology group of this complex is located in dimension $2$ and
is isomorphic to $\Z^6.$
\end{lemma}
\begin{proof}
Indeed, the generating it 2-cycles look as follows.
First of all, any arrow can appear in such a cycle only together with its
opposite, taken with opposite coefficient: otherwise the boundary of this cycle
will contain an edge of type B) with a non-zero coefficient. Thus it is
sufficient to consider the 10-dimensional group, generated by the linear
combinations of the form \{an arrow minus its opposite\}. Such an element will
be depicted by a double arrow directed as the first arrow in this combination,
see Fig.~\ref{two}b. A linear combination of such double arrows is a cycle of
the complex of closed chains of $\Diamond(J) \sm \Diamond_3$ if and only if the
correspondingly oriented segments form a cycle of the complete graph with 5
vertices. The group $H_1$ of this complete graph is isomorphic to $\Z^6,$ and
lemma is proved. \end{proof}
Further, our $\J$-block is the space of a fiber bundle, whose base is the
configuration space $B(S^1,5) \cong S^1 \times B^4,$ and the fiber over the
configuration $J$ is the direct product of the oriented
$(\Delta-8)$-dimensional subspace $\chi(J) \subset \K$ and the complex
$\Diamond(J) \sm \Diamond_3$. The monodromy over the generator $S^1$ of the
fundamental group of the base acts on this complex (and its homology) as a
cyclic permutation of 5 vertices. Thus by the Wang exact sequence the group
considered in Proposition \ref{three_hom} is generated exactly by all cycles of
a complete 5-graph which are invariant under this action. This group is
two-dimensional; its generators are shown in Fig.~\ref{two}b.
\end{proof}
We have found all possible generators of the group
\begin{equation}
\label{sig4} \bar H_{\Delta-1}(\Diamond_4 \sm \Diamond_3),
\end{equation}
namely the following statement holds.
\prop \label{gensig4} Any element of the group (\ref{sig4}) is a linear
combination of five chains shown in Fig.~\ref{two}. \quad $\Box$ \eprop
Now let us study the boundaries of these chains in other blocks.
\prop \label{trivtwo} Two chains corresponding to two left pictures in
Fig.~\ref{two}a cannot participate in an element of the group (\ref{sig4}) with
nonzero coefficients. \eprop
Indeed, the boundary of the first (respectively, the second) of them contains
the sum of generators of $(\Delta-2)$-dimensional homology groups of two blocks
shown in Fig.~\ref{degen}a (respectively, \ref{degen}b). These generators do
not appear in the boundaries of any other of our 5 chains. \quad $\Box$
\medskip
\prop \label{triv3} Two chains corresponding to two pictures in Fig.~\ref{two}b
cannot participate in an element of the group (\ref{sig4}) with nonzero
coefficients. \eprop
To prove this proposition, let us consider the homology group of the
$\J$-block, corresponding to $(5)$-cliques $J = (x,x,y,z,w)$, as shown in
Fig.~\ref{degen}c. The corresponding order complex $\Diamond(J)$ again is
two-dimensional.
\prop \label{rel4} For any $(5)$-clique $J$ consisting of exactly 4
geometrically different points in $S^1,$ the group $\bar H_*(\Diamond(J) \sm
\Diamond_3)$ is concentrated in dimension 2 and is isomorphic to $\Z^3$. The
Borel--Moore homology group of the corresponding block in $\Diamond_4 \sm
\Diamond_3$ is concentrated in dimension $\Delta-2$ and also is isomorphic to
$\Z^3$. \eprop
\begin{proof}
Exactly as in the proof of Proposition \ref{three_hom}, almost all
two-di\-men\-si\-onal simplices of the complex $\Diamond(J)$ are naturally
depicted by arrows connecting some of 4 geometrically distinct points of the
clique $J$. (The unique extra triangle is the triple of subspaces, whose first
element is defined by three points of multiplicity 1: it should be denoted by a
loop edge, connecting the double point with itself. This triangle cannot
participate with non-zero coefficient in any cycle of the complex $\Diamond(J)
\sm \Diamond_3$.)
Again, the cycles of this complex are depicted by double arrows, forming cycles
(in the usual sense) of the complete graph on our 4 vertices. This proves the
first assertion of Proposition \ref{rel4}.
The entire $J$-block is the space of a fiber bundle over the space of all
cliques of this type (which is diffeomorphic to $S^1 \times B^3$), namely, of a
fibered product of an orientable $(\Delta-8)$-dimensional vector bundle and the
(trivial) bundle of complexes $\Diamond(J)\sm \Diamond_3$. This proves the last
assertion of Proposition. Moreover, the $(\Delta-2)$-dimensional cycles of the
block are in a one-to-one (K\"unneth) correspondence with 2-dimensional cycles
of $\Diamond(J) \sm \Diamond_3,$ where $J$ is any clique of this class.
\end{proof}
\prop \label{bound3} The boundaries in this block of two basic
$(\Delta-1)$-dimensional cycles, shown in Fig.~\ref{two}b, are the two cycles
shown in Fig.~\ref{boun3}a. \quad $\Box$ \eprop
\begin{figure}
\begin{center}
\unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(107.00,17.67)
\multiput(10.00,17.00)(0.79,-0.09){2}{\line(1,0){0.79}}
\multiput(11.59,16.82)(0.30,-0.11){5}{\line(1,0){0.30}}
\multiput(13.09,16.28)(0.17,-0.11){8}{\line(1,0){0.17}}
\multiput(14.43,15.42)(0.11,-0.11){10}{\line(0,-1){0.11}}
\multiput(15.54,14.27)(0.12,-0.20){7}{\line(0,-1){0.20}}
\multiput(16.37,12.91)(0.10,-0.30){5}{\line(0,-1){0.30}}
\multiput(16.86,11.39)(0.07,-0.80){2}{\line(0,-1){0.80}}
\multiput(17.00,9.80)(-0.11,-0.79){2}{\line(0,-1){0.79}}
\multiput(16.77,8.22)(-0.12,-0.30){5}{\line(0,-1){0.30}}
\multiput(16.19,6.73)(-0.11,-0.16){8}{\line(0,-1){0.16}}
\multiput(15.29,5.42)(-0.13,-0.12){9}{\line(-1,0){0.13}}
\multiput(14.11,4.34)(-0.20,-0.11){7}{\line(-1,0){0.20}}
\multiput(12.72,3.55)(-0.38,-0.11){4}{\line(-1,0){0.38}}
\put(11.19,3.10){\line(-1,0){1.59}}
\multiput(9.60,3.01)(-0.52,0.09){3}{\line(-1,0){0.52}}
\multiput(8.03,3.28)(-0.25,0.10){6}{\line(-1,0){0.25}}
\multiput(6.56,3.90)(-0.16,0.12){8}{\line(-1,0){0.16}}
\multiput(5.27,4.84)(-0.12,0.13){9}{\line(0,1){0.13}}
\multiput(4.22,6.05)(-0.11,0.20){7}{\line(0,1){0.20}}
\multiput(3.48,7.46)(-0.10,0.39){4}{\line(0,1){0.39}}
\put(3.07,9.00){\line(0,1){1.60}}
\multiput(3.03,10.60)(0.11,0.52){3}{\line(0,1){0.52}}
\multiput(3.34,12.16)(0.11,0.24){6}{\line(0,1){0.24}}
\multiput(4.01,13.61)(0.11,0.14){9}{\line(0,1){0.14}}
\multiput(4.98,14.88)(0.14,0.11){9}{\line(1,0){0.14}}
\multiput(6.22,15.89)(0.24,0.12){6}{\line(1,0){0.24}}
\multiput(7.65,16.59)(0.59,0.10){4}{\line(1,0){0.59}}
\put(9.00,3.00){\circle*{1.33}} \put(11.00,3.00){\circle*{1.33}}
\put(17.00,10.00){\circle*{1.33}} \put(10.00,17.00){\circle*{1.33}}
\put(3.00,10.00){\circle*{1.33}}
\multiput(37.00,17.00)(0.79,-0.09){2}{\line(1,0){0.79}}
\multiput(38.59,16.82)(0.30,-0.11){5}{\line(1,0){0.30}}
\multiput(40.09,16.28)(0.17,-0.11){8}{\line(1,0){0.17}}
\multiput(41.43,15.42)(0.11,-0.11){10}{\line(0,-1){0.11}}
\multiput(42.54,14.27)(0.12,-0.20){7}{\line(0,-1){0.20}}
\multiput(43.37,12.91)(0.10,-0.30){5}{\line(0,-1){0.30}}
\multiput(43.86,11.39)(0.07,-0.80){2}{\line(0,-1){0.80}}
\multiput(44.00,9.80)(-0.11,-0.79){2}{\line(0,-1){0.79}}
\multiput(43.77,8.22)(-0.12,-0.30){5}{\line(0,-1){0.30}}
\multiput(43.19,6.73)(-0.11,-0.16){8}{\line(0,-1){0.16}}
\multiput(42.29,5.42)(-0.13,-0.12){9}{\line(-1,0){0.13}}
\multiput(41.11,4.34)(-0.20,-0.11){7}{\line(-1,0){0.20}}
\multiput(39.72,3.55)(-0.38,-0.11){4}{\line(-1,0){0.38}}
\put(38.19,3.10){\line(-1,0){1.59}}
\multiput(36.60,3.01)(-0.52,0.09){3}{\line(-1,0){0.52}}
\multiput(35.03,3.28)(-0.25,0.10){6}{\line(-1,0){0.25}}
\multiput(33.56,3.90)(-0.16,0.12){8}{\line(-1,0){0.16}}
\multiput(32.27,4.84)(-0.12,0.13){9}{\line(0,1){0.13}}
\multiput(31.22,6.05)(-0.11,0.20){7}{\line(0,1){0.20}}
\multiput(30.48,7.46)(-0.10,0.39){4}{\line(0,1){0.39}}
\put(30.07,9.00){\line(0,1){1.60}}
\multiput(30.03,10.60)(0.11,0.52){3}{\line(0,1){0.52}}
\multiput(30.34,12.16)(0.11,0.24){6}{\line(0,1){0.24}}
\multiput(31.01,13.61)(0.11,0.14){9}{\line(0,1){0.14}}
\multiput(31.98,14.88)(0.14,0.11){9}{\line(1,0){0.14}}
\multiput(33.22,15.89)(0.24,0.12){6}{\line(1,0){0.24}}
\multiput(34.65,16.59)(0.59,0.10){4}{\line(1,0){0.59}}
\put(36.00,3.00){\circle*{1.33}} \put(38.00,3.00){\circle*{1.33}}
\put(44.00,10.00){\circle*{1.33}} \put(37.00,17.00){\circle*{1.33}}
\put(30.00,10.00){\circle*{1.33}} \put(10.00,3.00){\vector(-1,1){6.00}}
\put(10.00,3.00){\vector(-1,1){5.00}} \put(3.00,10.00){\vector(1,1){6.00}}
\put(3.00,10.00){\vector(1,1){5.00}} \put(10.00,17.00){\vector(1,-1){6.00}}
\put(10.00,17.00){\vector(1,-1){5.00}} \put(17.00,10.00){\vector(-1,-1){6.00}}
\put(17.00,10.00){\vector(-1,-1){5.00}} \put(37.00,3.00){\vector(-1,1){6.00}}
\put(37.00,3.00){\vector(-1,1){5.00}} \put(30.00,10.00){\vector(1,0){13.00}}
\put(30.00,10.00){\vector(1,0){12.00}} \put(44.00,10.00){\vector(-1,-1){6.00}}
\put(44.00,10.00){\vector(-1,-1){5.00}}
\put(24.00,2.00){\makebox(0,0)[cc]{{\large a}}}
\multiput(70.00,17.00)(0.79,-0.09){2}{\line(1,0){0.79}}
\multiput(71.59,16.82)(0.30,-0.11){5}{\line(1,0){0.30}}
\multiput(73.09,16.28)(0.17,-0.11){8}{\line(1,0){0.17}}
\multiput(74.43,15.42)(0.11,-0.11){10}{\line(0,-1){0.11}}
\multiput(75.54,14.27)(0.12,-0.20){7}{\line(0,-1){0.20}}
\multiput(76.37,12.91)(0.10,-0.30){5}{\line(0,-1){0.30}}
\multiput(76.86,11.39)(0.07,-0.80){2}{\line(0,-1){0.80}}
\multiput(77.00,9.80)(-0.11,-0.79){2}{\line(0,-1){0.79}}
\multiput(76.77,8.22)(-0.12,-0.30){5}{\line(0,-1){0.30}}
\multiput(76.19,6.73)(-0.11,-0.16){8}{\line(0,-1){0.16}}
\multiput(75.29,5.42)(-0.13,-0.12){9}{\line(-1,0){0.13}}
\multiput(74.11,4.34)(-0.20,-0.11){7}{\line(-1,0){0.20}}
\multiput(72.72,3.55)(-0.38,-0.11){4}{\line(-1,0){0.38}}
\put(71.19,3.10){\line(-1,0){1.59}}
\multiput(69.60,3.01)(-0.52,0.09){3}{\line(-1,0){0.52}}
\multiput(68.03,3.28)(-0.25,0.10){6}{\line(-1,0){0.25}}
\multiput(66.56,3.90)(-0.16,0.12){8}{\line(-1,0){0.16}}
\multiput(65.27,4.84)(-0.12,0.13){9}{\line(0,1){0.13}}
\multiput(64.22,6.05)(-0.11,0.20){7}{\line(0,1){0.20}}
\multiput(63.48,7.46)(-0.10,0.39){4}{\line(0,1){0.39}}
\put(63.07,9.00){\line(0,1){1.60}}
\multiput(63.03,10.60)(0.11,0.52){3}{\line(0,1){0.52}}
\multiput(63.34,12.16)(0.11,0.24){6}{\line(0,1){0.24}}
\multiput(64.01,13.61)(0.11,0.14){9}{\line(0,1){0.14}}
\multiput(64.98,14.88)(0.14,0.11){9}{\line(1,0){0.14}}
\multiput(66.22,15.89)(0.24,0.12){6}{\line(1,0){0.24}}
\multiput(67.65,16.59)(0.59,0.10){4}{\line(1,0){0.59}}
\put(63.67,7.33){\circle*{1.33}} \put(76.33,7.33){\circle*{1.33}}
\put(63.67,12.67){\circle*{1.33}} \put(70.33,3.00){\circle*{1.33}}
\put(73.33,16.00){\circle*{1.33}} \put(75.00,14.67){\circle*{1.33}}
\multiput(75.00,14.67)(-0.67,-0.12){17}{\line(-1,0){0.67}}
\multiput(63.67,12.67)(0.12,-0.17){56}{\line(0,-1){0.17}}
\multiput(70.33,3.00)(0.12,0.30){39}{\line(0,1){0.30}}
\put(76.33,7.33){\line(-1,0){12.66}}
\multiput(63.67,7.33)(0.13,0.12){73}{\line(1,0){0.13}}
\multiput(73.33,16.00)(0.12,-0.33){26}{\line(0,-1){0.33}}
\multiput(100.00,17.00)(0.79,-0.09){2}{\line(1,0){0.79}}
\multiput(101.59,16.82)(0.30,-0.11){5}{\line(1,0){0.30}}
\multiput(103.09,16.28)(0.17,-0.11){8}{\line(1,0){0.17}}
\multiput(104.43,15.42)(0.11,-0.11){10}{\line(0,-1){0.11}}
\multiput(105.54,14.27)(0.12,-0.20){7}{\line(0,-1){0.20}}
\multiput(106.37,12.91)(0.10,-0.30){5}{\line(0,-1){0.30}}
\multiput(106.86,11.39)(0.07,-0.80){2}{\line(0,-1){0.80}}
\multiput(107.00,9.80)(-0.11,-0.79){2}{\line(0,-1){0.79}}
\multiput(106.77,8.22)(-0.12,-0.30){5}{\line(0,-1){0.30}}
\multiput(106.19,6.73)(-0.11,-0.16){8}{\line(0,-1){0.16}}
\multiput(105.29,5.42)(-0.13,-0.12){9}{\line(-1,0){0.13}}
\multiput(104.11,4.34)(-0.20,-0.11){7}{\line(-1,0){0.20}}
\multiput(102.72,3.55)(-0.38,-0.11){4}{\line(-1,0){0.38}}
\put(101.19,3.10){\line(-1,0){1.59}}
\multiput(99.60,3.01)(-0.52,0.09){3}{\line(-1,0){0.52}}
\multiput(98.03,3.28)(-0.25,0.10){6}{\line(-1,0){0.25}}
\multiput(96.56,3.90)(-0.16,0.12){8}{\line(-1,0){0.16}}
\multiput(95.27,4.84)(-0.12,0.13){9}{\line(0,1){0.13}}
\multiput(94.22,6.05)(-0.11,0.20){7}{\line(0,1){0.20}}
\multiput(93.48,7.46)(-0.10,0.39){4}{\line(0,1){0.39}}
\put(93.07,9.00){\line(0,1){1.60}}
\multiput(93.03,10.60)(0.11,0.52){3}{\line(0,1){0.52}}
\multiput(93.34,12.16)(0.11,0.24){6}{\line(0,1){0.24}}
\multiput(94.01,13.61)(0.11,0.14){9}{\line(0,1){0.14}}
\multiput(94.98,14.88)(0.14,0.11){9}{\line(1,0){0.14}}
\multiput(96.22,15.89)(0.24,0.12){6}{\line(1,0){0.24}}
\multiput(97.65,16.59)(0.59,0.10){4}{\line(1,0){0.59}}
\put(93.67,7.33){\circle*{1.33}} \put(106.33,7.33){\circle*{1.33}}
\put(93.67,12.67){\circle*{1.33}} \put(100.33,3.00){\circle*{1.33}}
\put(104.33,15.33){\circle*{1.33}}
\put(106.33,7.33){\line(-1,0){12.66}}
\multiput(93.67,12.67)(0.12,-0.17){56}{\line(0,-1){0.17}}
\put(84.67,2.00){\makebox(0,0)[cc]{{\large b}}}
\put(101.00,9.00){\makebox(0,0)[cc]{\small +}}
\put(97.00,10.00){\makebox(0,0)[cc]{{\small $-$}}}
\end{picture}
\end{center}
\caption{Boundaries of basic cycles of Fig.~\protect\ref{two}b and
Fig.~\protect\ref{md}} \label{boun3}
\end{figure}
\prop \label{mdcycle} The chain in $\Diamond_4,$ shown in Fig.~\ref{md}
($\equiv$ the right-hand picture in Fig.~\ref{two}a), defines a cycle in
$\Diamond_4 \sm \Diamond_3.$ \eprop
\begin{proof}
The $(3,3)$-cliques of this type {\Large $*$} can degenerate only in the
following way: some two neighboring points of {\em different} groups coincide,
thus forming a $(5)$-con\-fi\-gu\-ra\-tion, see the left picture of
Fig.~\ref{boun3}b. Given a $(5)$-con\-fi\-gu\-ra\-tion $J$, the $\J'$-block of
the type {\Large $*$} adjoins the corresponding complex $\Diamond(J)$ exactly 5
times, because such coincidence can happen at any of its 5 points. The boundary
position of complexes $\Diamond(J'),$ when $J' \in ${\Large $*$} tends to $J$
in the way shown in Fig.~\ref{boun3}b left, is (in the notation used in the
proof of Proposition \ref{three_hom}) equal to the difference of two
1-dimensional simplices in $\Diamond(J)$ denoted by two edges in
Fig.~\ref{boun3}b right. The sum of such differences over all 5 vertices of the
configuration $J$ is equal to 0, and Proposition \ref{mdcycle} is proved.
\end{proof}
\begin{remark}
A much more general fact was proved in \cite{Merx-2}: any "horizontal" boundary
operator $d^1$ of the auxiliary spectral sequence, corresponding to the
collision of two points of two different groups, always is trivial.
\end{remark}
Summarizing the Propositions \ref{classcomp4}--\ref{mdcycle}, we get the
following statement.
\thm The group $\bar H_{\Delta-1}(\Diamond_4 \sm \Diamond_3)$ is isomorphic to
$\Z$ and is generated by the fundamental cycle of the $\J$-block corresponding
to the $(3,3)$-cliques shown in Fig.~\ref{md}. \etheorem
As the $(\Delta-1)$- and $(\Delta-2)$-dimensional Borel--Moore homology groups
of both spaces $\Diamond_2 $ and $\Diamond_3 \sm \Diamond_2$ are trivial (see
Proposition \ref{si2} and Theorem \ref{s3}), this implies Theorem \ref{doodinv}
of the Introduction.
\subsection{A nontrivial doodle.}
\label{merxdood}
The theory of finite-order invariants provides a method of constructing a
priori nontrivial (and nonequivalent) objects. E.g., imagine that we do not
know any nontrivial knot in $\R^3$ and wish to construct it. To do it, we can
calculate the simplest finite-order knot invariant (given by the chord diagram
$\bigoplus$), then draw the simplest singular knot respecting this diagram
(i.e. having two transverse selfintersections), and then consider four knots
obtained from it by all possible local resolutions of both these points. At
least one of obtained knots surely will be nonequivalent to the others (and
indeed, if we do all this in the simplest possible way, we get three trivial
knots and one trefoil).
In exactly the same way, we can construct the simplest quasidoodle with two
generic triple points, respecting the triangular diagram of Fig.~\ref{md}.
Perturbing it in four different ways, we shall obtain three trivial (equivalent
to a circle) doodles, and one equivalent to Fig.~\ref{merx}a.
(However, A.~B.~Merkov, who discovered this doodle, came to it from very
different considerations.)
\section{Invariants of I-doodles}
In this and the next sections we consider only the immersed curves in $\R^2$.
To any immersion $\phi: S^1 \to \R^2$ there corresponds a map $S^1 \to S^1:$
any point $x \in S^1$ goes to the direction of the tangent vector $\phi'(x).$
Accordingly to S.~Smale \cite{Smale}, this correspondence is a homotopy
equivalence between spaces $I \K \equiv Imm(S^1, \R^2)$ and $C(S^1,S^1).$ These
spaces split into countably many components labeled by the "winding numbers"
(i.e. the indices of corresponding maps $S^1 \to S^1$). Any of these components
is homotopy equivalent to $S^1,$ the homotopy equivalence being provided by the
image of (the tangent direction of $\phi$ at) the distinguished point of $S^1$.
The discriminant $I\Sigma$ in the space $I \K$ is just the intersection of this
space $I \K$ with the discriminant set $\Sigma \subset \K$ considered in the
previous sections. Its resolution $I\Diamond$ is a subset in $\Diamond,$ namely
the complete preimage of $I\Sigma.$
In its decomposition into $\J$-blocks only the $J$-{\em configurations}, i.e.
the $\J$-cliques without multiple points, can take part.
\prop \label{smale} For any connected component $\CC$ of the space $I \K$ and
for any $A$-con\-fi\-gu\-ra\-tion $J$ in $S^1,$ the space of immersions $S^1
\to \R^2$ respecting this configuration and lying in this component is a {\em
path-connected} open submanifold of the space $\chi(J) \subset \K.$ \eprop
This follows easily from the Smale's theorem. \quad $\Box$
\medskip
For any component $\CC,$ denote by $\CC I \Diamond$ and $\CC I \Diamond_i$ the
intersection of the space $I \Diamond $ (respectively, $I \Diamond_i$) with the
preimage of $\CC$ under the projection $\Diamond \to \K$.
\begin{example} The stratum $I\Diamond_2$ is
an open subset in the space of a fiber bundle, almost coinciding with that
considered in Proposition \ref{si2}, with unique difference that its base is
not the entire space $\Psi =(S^1)^3/S(3),$ but its open part $B(S^1,3)$. The
{\em strangeness} is the linking number in $\CC$ with the direct image of the
fundamental cycle of this subset. The existence of the strangeness as an
integer-valued invariant is due to the fact that this configuration space
$B(S^1,3)$ is orientable.
\end{example}
All other $(\Delta-1)$-dimensional blocks in all spaces $I\Diamond_3 \sm
I\Diamond_2 $ and $I\Diamond_4 \sm I\Diamond_3 $ are the open subsets of
similar blocks considered in the previous section; the corresponding virtual
generators of the group of invariants of I-doodles are shown in
Figs.~\ref{one}b, \ref{two}a and \ref{two}b.
All these generators define elements of corresponding groups $\bar
H_{\Delta-1}(\Diamond_i \sm \Diamond_{i-1}) \equiv E^1_{i,\Delta-i-1}$. For two
generators shown in Fig.~\ref{two}b this follows from the fact that the entire
boundaries of corresponding blocks in $I\Diamond_4 \sm I\Diamond_3 $ are empty.
For two remaining blocks of Fig.~\ref{two}a we should additionally check that
their boundaries in the block corresponding to the $(5)$-configurations are
trivial; the proof of this fact essentially coincides with that of Proposition
\ref{mdcycle}.
So, for any fixed component of the space $I \K$ the domain in the table
$\{E^1_{p,q}\}$ responsible for the calculation of invariants of orders 2, 3
and 4 of I-doodles from this component looks as is shown in Fig.~\ref{ssimm}.
\begin{figure}
\begin{center}
\unitlength=1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(52.00,38.00)
\put(12.00,6.00){\vector(0,1){32.00}} \put(12.00,6.00){\vector(1,0){40.00}}
\put(49.00,3.00){\makebox(0,0)[cc]{$p$}}
\put(6.00,27.00){\makebox(0,0)[cc]{$\Delta-3$}}
\put(6.00,21.00){\makebox(0,0)[cc]{$\Delta-4$}}
\put(6.00,15.00){\makebox(0,0)[cc]{$\Delta-5$}}
\put(17.00,27.00){\makebox(0,0)[cc]{${\bf Z}$}}
\put(17.00,21.00){\makebox(0,0)[cc]{${\bf Z}$}}
\put(27.00,21.00){\makebox(0,0)[cc]{${\bf Z}$}}
\put(27.00,15.00){\makebox(0,0)[cc]{${\bf Z}$}}
\put(37.00,15.00){\makebox(0,0)[cc]{${\bf Z}^5$}}
\put(9.00,35.00){\makebox(0,0)[cc]{$q$}}
\put(17.00,15.00){\makebox(0,0)[cc]{0}} \put(27.00,9.00){\makebox(0,0)[cc]{0}}
\put(27.00,27.00){\makebox(0,0)[cc]{0}} \put(37.00,21.00){\makebox(0,0)[cc]{0}}
\put(17.00,33.00){\makebox(0,0)[cc]{0}} \put(17.00,3.00){\makebox(0,0)[cc]{2}}
\put(27.00,3.00){\makebox(0,0)[cc]{3}} \put(37.00,3.00){\makebox(0,0)[cc]{4}}
\put(12.00,12.00){\line(1,0){33.00}} \put(45.00,18.00){\line(-1,0){33.00}}
\put(12.00,24.00){\line(1,0){33.00}} \put(45.00,30.00){\line(-1,0){33.00}}
\put(22.00,31.00){\line(0,-1){25.00}} \put(32.00,6.00){\line(0,1){25.00}}
\put(42.00,31.00){\line(0,-1){25.00}}
\end{picture}
\end{center}
\caption{Spectral sequence for invariants of orders 2, 3 and 4} \label{ssimm}
\end{figure}
\prop \label{degenss} For any connected component $\CC$ of the space $I \K$,
the fragment of the spectral sequence shown in Fig.~\ref{ssimm} degenerates at
the term $E^1,$ i.e. all its elements extend to well defined Borel--Moore
homology classes of the space $\CC I\Diamond.$ \eprop
\begin{proof}
For the group $E_{2, \Delta-3}$ this is obvious.
The group $E^1_{2,\Delta-4}$ is generated by the fundamental cycle of the
submanifold in $\CC I\Diamond_2 $, consisting of such pairs of the form \{a
3-configuration $(x,y,z) \in B(S^1,3)$; a map $\phi:S^1 \to \R^2$\} that $x+y+z
\equiv 0 (mod \, 2\pi).$ It is obviously a cycle in entire $\CC I\Diamond $,
let us prove that it is not homologous to zero. As $H^2(\CC) \simeq 0,$ it is
sufficient to construct two 1-dimensional cycles in $\CC \sm I\Sigma,$ defining
the same element in $H^1(\CC)$ but such that some (and then any) 2-chain
realizing the homology between these cycles has nonzero intersection number
with this fundamental cycle.
Consider a map $\phi \in I\Sigma \cap \CC$ with unique generic triple point,
and let $\phi_1, \phi_2$ be two its small nondiscriminant perturbations
resolving this triple point in two different ways, see Fig.~\ref{reidem}c.
For any $i=1,2,$ denote by $[\phi_i]$ the 1-cycle in $\CC \sm \Sigma$ swept out
by all maps obtained from $\phi_i$ by all cyclic reparametrizations of the
issue circle $S^1.$ These two cycles are obviously homologous in $\CC,$ and the
intersection index of such a homology with the above manifold is equal to $\pm
3.$
This proves the assertion of Proposition for cells $E_{2,\Delta-4}$ and
$E_{3,\Delta-4}$.
In particular, there exist two elements of the group $H_0(\CC \sm \Sigma)$,
i.e. two linear combinations of doodles in $\CC$, which cannot be distinguished
by the "strangeness" (generating the group $E^\infty_{2,\Delta-3}$ of second
order 0-cohomology classes), but can be distinguished by the invariant
generating the group $E^\infty_{3,\Delta-4}$. (In accordance with
\S~\ref{merxdood}, we can find these combinations by resolving unique point of
multiplicity 4. Indeed, the linear combination of four doodles, locally
situated as in Fig.~\ref{diffour} and coinciding outside it, provides such a
chain.)
\begin{figure}
\unitlength 0.90mm \linethickness{0.4pt}
\begin{picture}(131.00,22.00)
\put(10.00,10.00){\vector(1,0){20.00}} \put(20.00,0.00){\vector(0,1){20.00}}
\put(7.00,3.00){\vector(1,1){19.00}} \put(42.00,10.00){\vector(1,0){20.00}}
\put(52.00,0.00){\vector(0,1){20.00}} \put(39.00,3.00){\vector(1,1){19.00}}
\put(29.00,3.00){\vector(-1,1){18.00}} \put(59.00,1.00){\vector(-1,1){19.00}}
\put(74.00,10.00){\vector(1,0){20.00}} \put(84.00,0.00){\vector(0,1){20.00}}
\put(106.00,10.00){\vector(1,0){20.00}} \put(116.00,0.00){\vector(0,1){20.00}}
\put(93.00,3.00){\vector(-1,1){18.00}} \put(123.00,1.00){\vector(-1,1){19.00}}
\put(80.00,0.00){\vector(1,1){19.00}} \put(112.00,0.00){\vector(1,1){19.00}}
\put(101.00,10.00){\makebox(0,0)[cc]{{\large $+$}}}
\put(68.00,10.00){\makebox(0,0)[cc]{{\large $-$}}}
\put(36.00,10.00){\makebox(0,0)[cc]{{\large $-$}}}
\put(4.00,10.00){\makebox(0,0)[cc]{{\large $+$}}}
\end{picture}
\caption{The chain of I-doodles nontrivial by the invariant of order 3}
\label{diffour}
\end{figure}
Exactly as above we produce from them an 1-cycle in $\CC \sm \Sigma$,
homologous to zero in $\CC$, and having a nonzero intersection index with the
chain generating the group $E^1_{3,\Delta-5}.$ This proves our Proposition.
\end{proof}
In particular, we have proved that for any component $\CC$ of the space $I \K$
all 7 generators mentioned in Theorem \ref{idoodinv} and shown in
Figs.~\ref{one}, \ref{two} define independent elements of the group $\bar
H_{\Delta-1}(\Sigma \cap \CC).$
Finally, it is obvious that the intersection indices with all these
$(\Delta-1)$-di\-men\-si\-o\-nal Borel--Moore homology classes of the
discriminant define zero elements in the 1-dimensional cohomology group of the
component $\CC$, and hence the linking numbers with them are well defined
invariants of I-doodles.
Theorem \ref{idoodinv} is thus completely proved.
\section{1-dimensional cohomology of the space of immersions
$S^1 \to \R^2$ without points of multiplicity 4}
Define the discriminant $I \Sigma 4 \subset I \K$ as the set of immersions
$\phi: S^1 \to \R^2$ such that images of some 4 different points coincide.
Its resolution $I \Diamond 4$
is constructed in essentially the same
way as it was done above for the set $I\Sigma.$ In this section we are
interested in the 1-dimensional cohomology classes of the space $I\K \setminus
I\Sigma 4$ or, which is the same, in the $(\Delta-2)$-dimensional Borel--Moore
homology classes of spaces $I\Sigma 4$ or $I\Diamond 4.$
The first nonempty term of this resolution is of filtration 3. This is the
space of an orientable $(\Delta-6)$-dimensional vector bundle over the
configuration space $B(S^1,4).$ This configuration space is non-orientable,
therefore the fundamental cycle of this term defines a class only in the group
$\bar H_{\Delta-2}(I \Sigma_4, \Z_2),$ but not in the integer homology group,
see in Theorem \ref{4ptfree} the statement "$\F_3 \sim \Z_2$ over $\Z_2$".
\medskip
The next term $I {\Diamond 4}_4 \sm I {\Diamond 4}_3$ of our filtration also is
the space of a fiber bundle, whose base is the configuration space $B(S^1,5),$
and the fibers are direct products of stars $\star$ with 5 rays (without
endpoints, which belong to the smaller term of the filtration) and some
(canonically oriented) $(\Delta-8)$-dimensional vector subspaces in $\K$.
The base $B(S^1,5)$ of this bundle is orientable (and diffeomorphic to $S^1
\times \R^4$), and the monodromy over the circle generating the group
$\pi_1(B(S^1,5)) \sim \Z$ acts on the fibration of 5-stars $\star$ by cyclic
permutations of their rays.
Therefore the $(\Delta-2)$-dimensional Borel--Moore homology group of this term
coincides with the subgroup of the group $\bar H_1(\star)$ consisting of
elements invariant under the rotations of these stars. For any coefficient
group $G$ this group is isomorphic to $G^4$. If in $G$ the condition $5a=0$
implies $a=0,$ then its invariant subgroup is trivial; in the case $G=\Z_5$
this group is isomorphic to $\Z_5,$ see statement "$\F_4 \simeq \Z_5$" over
$\Z_5$ of Theorem \ref{4ptfree}.
Finally, consider the term $I{\Diamond 4}_5 \sm I{\Diamond 4}_4$ of our
filtration. It is the space of a fiber bundle over $B(S^1,6),$ whose fiber is
the product of $\R^{\Delta-10}$ and some two-dimensional order complex. This
complex is similar to the one considered in the proof of Proposition
\ref{three_hom}, with unique difference: its vertices correspond to choices of
some 4, 5 or 6 points of our 6, and not of 3, 4 or 5 points of 5. Absolutely as
previously, the two-dimensional cycles of this complex are the linear
combinations of (double) arrows with starts and ends at these 6 points, forming
the cycles (in the usual sense) of the complete graph on these 6 vertices.
However, unlike the case of 5-configurations, the base space $B(S^1,6)$ is
non-orientable. Therefore the $(\Delta-2)$-dimensional Borel--Moore homology
classes of our block are in a one-to-one correspondence with such cycles of the
complete 6-graph, which are {\em anti-invariant} under the cyclic permutations
of its 6 vertices.
The group of such cycles can be easily calculated and is isomorphic to $\Z^2$;
the pictures of its generators are given in Fig.~\ref{hexagon}.
By dimensional reasons, these homology classes of the space $I{\Diamond 4}_5
\sm I{\Diamond 4}_4$ can be extended to these of the space $I{\Diamond 4}_5,$
and hence to the 1-dimensional cohomology classes (of order 5) of the entire
space of immersions $S^1 \to \R^2$ without 4-fold points.
This proves Theorem \ref{4ptfree} of the Introduction.
|
\section{Introduction}
One of the paramount problems in modern cosmology is to elucidate how the first
generation of luminous objects, stars, accreting black holes (BHs) and galaxies, shaped the early universe
at the end of the cosmic dark ages \citep{barkana2001, loeb2013, wiklind2013}.
A key driver of this grand cosmic transformation was the gradual enrichment of the pristine
universe with heavy chemical elements in the wake of the first supernova (SN) explosions
\citep[reviewed in][]{karlsson2013}. According to the modern theory of cosmological structure
formation \cite[e.g.][]{mo2010}, the hierarchical assembly of dark matter (DM) halos provided
the gravitational potential wells that allowed gas to form stars and galaxies inside them.
Extending this model to the highest redshifts, one can determine the sites where the first
stars, the so-called Population~III (Pop~III), formed out of the pure H/He gas created in
the Big Bang. Within this framework,
Pop~III stars are predicted to form inside
of DM minihalos with total masses (DM and gas) of about $10^{6}\,{\ensuremath{\mathrm{M}_{\odot}}}$ at redshifts
of $z\sim 20 - 30$ \citep{couchman1986, haiman1996, tegmark1997, bromm2004, oshea2008, bromm2009}.
The first stars affected the early universe in several different ways.
Massive Pop~III stars were strong emitters of hydrogen and helium ionizing photons that built
up extensive {H~II}, {He~II}, and {He~III} regions \citep{yoshida2007a}. The metals forged in Pop~III stars
later were dispersed into the intergalactic medium (IGM) when they died as SNe, thus quickly polluting the primordial gas such that the second generation of (Population~II) stars could emerge.
It is convenient to classify Pop~III feedback mechanisms into different classes (Ciardi \& Ferrara
2005), specifically radiative feedback
\citep{oet05, abel2007, su07, wet08b, hus09, wet10, yoshida2007a, greif2009a}, mechanical and chemical feedback due to SNe \citep{wise2008, greif2010, ritter2012},
and X-ray feedback from accreting BH remnants \citep{kuhlen2005,alvarez2009, jeon2012, jeon2014, xu2014}.
In the Pop~III star-forming regions, metal
cooling was absent because the primordial gas consisted almost exclusively of
hydrogen ($\sim76\,\%$ by mass) and helium ($\sim24\,\%$). Molecular hydrogen was
thus the dominant coolant, but owing to its quantum-mechanical structure,
it was unable to cool the gas to the low temperatures typically encountered in star forming
clouds today. The primordial gas, therefore, remained relatively warm, with
typical temperatures of several hundred Kelvin. Hence the Jeans mass was correspondingly larger, as
well, leading to the expectation that Pop~III stars
might have been more massive than stars formed today, with a predicted mass scale of
$50\,{\ensuremath{\mathrm{M}_{\odot}}} - 100\,{\ensuremath{\mathrm{M}_{\odot}}}$ \citep{bromm1999, bromm2002, naka2001, abel2002, omukai2003}.
Because of their high surface temperatures
\citep{bond1984, bkl2001, schaerer2002}, Pop~III stars could effectively
produce copious amounts of ionizing UV photons.
Extended H~II{} regions with size of several
kpc were created before the stars died. Given
the shallowness of the gravitational potential well of the host DM halos, the
surrounding gas was subject to strong photo-heating, thus being able to easily escape the
host minihalos \citep{whalen2004, alvarez2006}. This photo-evaporation suppressed
further star formation inside the minihalos, thus delaying further star formation until
more massive host halos emerged \citep{bromm2011}.
The character of Pop~III feedback sensitively depends on the fate encountered
by massive Pop~III stars when they die after their
short lifetime of a few million years to trigger a SN explosion,
or directly collapse into
black holes. Those Pop~III stars with masses of $140\,{\ensuremath{\mathrm{M}_{\odot}}}$ -
$260\,{\ensuremath{\mathrm{M}_{\odot}}}$ are thought to die as pair-instability supernovae (PSNe)
\citep{barkat1967, heger2002, chen2014}, although this mass range may have to be revised in the
case of rapidly rotating progenitors \citep{chatz2012}.
Unlike gravitationally-powered core-collapse supernovae
(CCSNe), PSNe are hyper-energetic thermonuclear explosions, not leaving any
compact remnant behind. Because vast amounts of metals
are ejected during a PSN explosion, even a single event could
enrich about $10^7\,{\ensuremath{\mathrm{M}_{\odot}}}$ of primordial gas up to
$10^{-4}\ensuremath{Z_\odot}-10^{-3}\ensuremath{Z_\odot}$ \citep{karlsson2008,wise2008,ss09,greif2010}.
Even such a trace amount of metals could change the subsequent star
formation process and might cause a transition of the stellar initial mass
function (IMF) from the top-heavy mode predicted for
Pop~III stars to the standard IMF for later (Pop~I and Pop~II)
generations with typical masses comparable to that of our Sun
\citep{bfcl2001, omukai2005, maio2010, wise2012}.
Results from stellar archaeology \citep{beers2005, frebel2005}, which
studies the most metal-poor stars in the Local Group that retained the imprints from
nucleosynthesis in the early universe, possibly including those from Pop~III stars, in general
do not support the chemical abundance pattern predicted for PSN enrichment \citep{tumlinson2006}. Theoretical PSN
yields exhibit a pronounced odd-even effect resulting from a low
neutron excess \citep{heger2002}. In addition,
the lack of any neutron capture process results in the absence of all elements heavier
than the Fe peak (no r- or s-process).
The Pop~III CCSN
models \citep{umeda2003, heger2010}, however, can produce abundance
patterns in good accord with the observation of metal poor stars.
Recent simulations of Pop~III star formation that take into account the radiation-hydrodynamical
feedback from the growing central protostar have shown that
accretion can be halted, thus preventing
the formation of stars more massive
than 50 ${\ensuremath{\mathrm{M}_{\odot}}}$ \citep{hoso2011, stacy2012, hirano2014}. That implies that Pop~III stars typically might
die in a CCSN, instead of a PSN, in agreement with the observations.
Furthermore,
recent cosmological simulations with extremely high resolution have shown
that the primordial gas cloud is able to fragment and produce stars
of relatively lower mass of tens of solar masses \citep{turk2009, stacy2010, greif2011},
organized in binaries or multiple stellar
systems. These simulations suggest that binary systems may be the typical
channel for primordial star formation in minihalos \citep{stacyB2013}.
Since these simulations only follow the protostellar assembly process for at most
$\sim 10^3$\,yr, it is not yet clear how the final mass spectrum will look like \citep{bromm2013}.
One key uncertainty is the degree of merging of neighboring protostars \citep{greif2012}.
However, it appears likely that the first stars in minihalos typically formed in binary
or small multiple systems.
Since the evolution of binary systems
and their final fate are very different from those of single stars, it is
worthwhile to investigate whether binary Pop~III stars lead to significantly altered feedback effects.
Since their evolution is quite
different from a single
star, it is worth investigating how the Pop~III binary
systems affected the IGM and their host halos
that later merged into the first galaxies \citep{rgs02,wise2008,wise2009,wise2014,johnson2009,johnson2013,xu2013}.
The evolving binaries might exert different feedback
mechanisms, through the emission of UV and x-ray ionizing photons, and SN
explosions, all of which may be quite different from the feedback
of the single stars which has been well documented in the literature. Therefore,
we first study the impact of the first massive stars of masses $60\,{\ensuremath{\mathrm{M}_{\odot}}}$,
$45\,{\ensuremath{\mathrm{M}_{\odot}}}$, $30\,{\ensuremath{\mathrm{M}_{\odot}}}$, and $15\,{\ensuremath{\mathrm{M}_{\odot}}}$ on their parent halos.
We study the possible impacts of the first binary systems on the IGM and
present the results of cosmological simulations by considering possible outcomes
of the Pop~III binary models with stars of $45\,{\ensuremath{\mathrm{M}_{\odot}}}$+$15\,{\ensuremath{\mathrm{M}_{\odot}}}$ (S45+S15)
and $30\,{\ensuremath{\mathrm{M}_{\odot}}}$+$30\,{\ensuremath{\mathrm{M}_{\odot}}}$ (S30+S30). Our binary models consider the
non-interacting binaries during their stellar evolution. However, more
realistic binary models might have a much wider range of outcomes
\citep{langer2012}.
The structure of this paper is as follows: In Section~2, we
describe our initial setup, as well as our numerical methods. A discussion of our
protostellar evolution models, both for single and binary stars, follows in Section~3.
The simulation results are presented in Section~\ref{results}, and we conclude
by discussing the broader implications
in Section~\ref{flms_discussion}. All of the results presented in this paper use physical
coordinates instead of comoving coordinates.
\section{Numerical Methodology}
\subsection{Problem Setup}
\label{flms_nm}
The primary code used for our simulations is the well-tested,
massively-parallel cosmological code \texttt{GADGET}{} \citep{springel2005},
which computes gravitational forces with a hierarchical tree algorithm
and represents fluids by means of smoothed particle hydrodynamics
(SPH). In order to simulate the feedback exerted by the first stars,
additional physical processes, such as cooling and chemistry of the primordial gas, radiative
transfer of ionizing photons, and SN explosions, are
required and have been implemented in \texttt{GADGET}{}.
Our simulations employ the same initial conditions as in \cite{greif2010},
starting at $z = 100$ in a periodic box of linear size of $1\,{\ensuremath{\mathrm{Mpc}}}$ (co-moving). We choose
$\Lambda$CDM cosmological parameters with matter density $\Omega_m=0.3$, baryon density $\Omega_b=0.04$,
Hubble constant $H_0 = 70\,\unit{km}$\,$\unit{s^{-1} Mpc^{-1}}$, spectral index $n_{\rm s}=1.0$, and normalization
$\sigma_8=0.9$, based on the WMAP cosmic microwave background (CMB) measurement \citep{komatsu2009}.
\cite{greif2010} used a standard hierarchical zoom-in technique, generating the highest mass
resolution covering the Lagrangian region where the first galaxy is destined to form.
That way, all the relevant fine-structure, specifically the minihalo progenitors
of the first galaxy, can be resolved. The resulting mass of the DM and gas particles
in the highest resolution region is $m_{\rm DM}\sim33\,{\ensuremath{\mathrm{M}_{\odot}}}$ and $m_{\rm sph}\sim5\,{\ensuremath{\mathrm{M}_{\odot}}}$,
respectively. Because the molecular hydrogen cooling in primordial gas imprints a characteristic
density of $n_{\rm H}=10^4\,\unit{cm^{-3}}$ and temperature of $200\,\ensuremath{\mathrm{K}}$ \citep{bromm2002}, our simulations marginally resolve the corresponding Jeans mass of $M_{\rm J}\simeq 500 {\ensuremath{\mathrm{M}_{\odot}}}$.
Our cooling module and chemistry network are based on \cite{greif2010},
and include all relevant cooling mechanisms of primordial gas, such as H and He collisional
ionization, excitation and recombination cooling, bremsstrahlung,
and inverse Compton cooling; in addition, the collisional excitation cooling via ${\ensuremath{^{} \mathrm{H}_2}}$ and HD is also taken into account. For ${\ensuremath{^{} \mathrm{H}_2}}$ cooling,
collisions with protons and electrons are explicitly included. The chemical network includes $\rm H, H^+, H^-, H_2, H_2^+, He, He^+, He^{++}$, and $\unit{e^-}$, D, $\rm D^+$, and HD.
State-of-the-art cosmological simulations can potentially use billions of particles
to model the formation of the universe. However, it is still extremely challenging
to resolve mass scales from galaxies (10$^{10}\,{\ensuremath{\mathrm{M}_{\odot}}}$) to individual
stars ($1\,{\ensuremath{\mathrm{M}_{\odot}}}$). For example, the resolution length in our
simulation is about $1\,{\ensuremath{\mathrm{pc}}}$, many orders of magnitude removed from truly stellar scales.
Hence, modeling the process of star formation on cosmological scales from first
principles is currently still out of reach. A viable strategy to treat star
formation and its feedback is to employ sub-grid models, where sink particles approximately
represent unresolved single stars, or clusters thereof. The sinks can then act as
sources of radiation, with luminosities and spectra chosen in accordance with
stellar structure and evolution models, and eventually as sites for SN explosions.
Another reason to use sink particles is to overcome the so-called `Courant myopia'.
When the gas density somewhere inside the computational box becomes increasingly high,
the SPH smoothing length decreases, in turn enforcing the adoption of smaller and smaller timesteps, according to the Courant condition. When encountering a runaway collapse, the simulation
would effectively grind to a halt or fail.
Creating sink particles allows us to bypass this numerical bottleneck, such that the
simulations can be followed beyond the initial collapse, where much of the interesting
physics, related to the stellar feedback, occurs.
We here apply the sink particle algorithm of \cite{JB2007}.
The key criterion for sink creation,
and subsequent accretion of further SPH particles, is that the gas density exceeds a
pre-specified density threshold,
$n_c\,\sim\,10^4\,{\ensuremath{\mathrm{cm}}}^{-3}$, but we also test for gravitational boundedness, and
whether the gas is part of a converging flow, where
$\nabla\cdot\vec{v}<0$.
The sink particles
provide markers for the position of a Pop~III star and its remnants, such as a black hole or SN,
to which detailed sub-grid physics can be supplied.
\subsection{Radiative Transfer}
\label{ray_sec}
When a Pop~III star has formed inside the minihalo, the sink particle representing it
immediately turns into a point source of ionizing photons to mimic the
birth of a star. The rate of ionizing photons emitted depends on the
physical size of the star and its surface temperature based on our stellar subgrid
models. Instead of simply assuming constant rates of emission,
we use the results of one-dimensional stellar evolution calculations from \cite{heger2010}
to construct the luminosity history of Pop~III stars.
Indeed, luminosities exhibit considerable time variability when taking the
evolution off the main sequence into account.
The photons streaming from the star then establish an ionization front and build
up H~II{} regions. To trace the propagation of photons and the ionization front,
we use the ray-tracing algorithm from \citep{greif2009a}, which solves the
ionization front equation in a spherical grid by tracking $10^5$ rays
with 500 logarithmically spaced radial bins around the photon source.
The radiation transport is coupled to the hydrodynamics of the gas through
its chemical and thermal evolution. The transfer of the $\rm H_2$-dissociating photons
in the Lyman--Werner (LW) band ($11.2-13.6\,\ensuremath{\mathrm{e\!V}}$) is also
included.
For completeness, we here briefly present the key features of the ray-tracing algorithm,
and refer the reader to \cite{greif2009a} for details and tests.
To begin with, particle positions are transformed from Cartesian to
spherical coordinates, i.e. radius ($r$), zenith angle ($\theta$), and azimuth angle ($\phi$).
The effective volume of each particle is $\sim h^3$, where $h$ is the SPH
smoothing length. The corresponding sizes in spherical coordinates are $\Delta r\,=\,h$,
$\Delta \theta\,=\,h/r$, and $\Delta \phi\,=\,h/r\sin\theta$. Using spherical coordinates
facilitates the convenient calculation of the
ionization front around the star,
\begin{equation}
n_n r^2_{\rm I}\frac{d r_{\rm I}}{dt}\,=\, \frac{\dot{N}_{\rm ion}}{4\pi}\,-\,\alpha_{\rm B}\int_0^{r_{\rm I}}n_en_+r^2dr,
\label{rad1}
\end{equation}
where $r_{\rm I}$ is the position of the ionization front, $\dot{N}_{\rm ion}$ represents
the number of ionizing photons emitted from the star per second, $\alpha_{\rm B}$ is
the case B recombination coefficient, and $n_n$, $n_e$, and $n_+$ are the number densities of neutral
particles, electrons, and positively charged ions, respectively. The recombination coefficient
is assumed to be constant at temperatures around $2\times10^4\,\ensuremath{\mathrm{K}}$. The ionizing photon rates
are
\begin{equation}
\dot{N}_{\rm ion}\,=\,\frac{\pi L_*}{\sigma_{\rm SB} T_{\rm eff}^4}\,\int_{\nu_{\rm min}}^{\infty}\frac{B_{\nu}}{h_{\rm P}\nu} {\rm d}\nu,
\label{rad2}
\end{equation}
where $h_{\rm P}$ is Planck's constant, $\sigma_{\rm SB}$ the Stefan-Boltzmann constant, $\sigma_{\nu}$ the relevant photo-ionization
cross section, and $\nu_{\rm min}$ the ionization threshold for
H~I, He~I, and He~II.
By assuming a blackbody spectrum with effective temperature, $T_{\rm eff}$,
the flux of a Pop~III star can be written
\begin{equation}
F_{\nu}\,=\,\frac{L_*}{4 \sigma T_{\rm eff}^4r^2}B_{\nu}.
\end{equation}
The size of the H~II{} region is determined by solving Equation~\ref{rad1}.
The particles within the H~II{} regions store information
about their distance from the star, which is used to calculate the
ionization and heating rates,
\begin{displaymath}
k_{\rm ion}\,=\,\int_{\nu_{\rm min}}^{\infty}\frac{F_{\nu}\sigma_{\nu}}{h_{\rm P}\nu} {\rm d}\nu,
\end{displaymath}
\begin{equation}
\Gamma\,=\,n_n\int_{\nu_{\rm min}}^{\infty}F_{\nu}\sigma_{\nu}\Big(1\,-\,\frac{\nu_{\rm min}}{\nu}\Big){\rm d}\nu. \
\end{equation}
{ {\ensuremath{^{} \mathrm{H}_2}}{} is the most important coolant for cooling the
primordial gas, which leads to formation of the first stars. However, its hydrogen bond is weak
and can be easily broken by photons in the LW
bands between 11.2 and 13.6 eV. The small ${\ensuremath{^{} \mathrm{H}_2}}$ fraction in the IGM
creates only a little optical depth for LW photons, allowing them to propagate a much
larger distance than ionizing photons. In our algorithm, self-shielding of H$_2$ is not
included because it is only important when H$_2$ column densities are high.
Here we treat the photodissociation of ${\ensuremath{^{} \mathrm{H}_2}}$ in the optically thin limit and the dissociation rate in a
volume constrained by causality within a radius, $r =ct $. The dissociation rate
is given by $k_{\rm H_2}=1.1\times10^8 F_{\rm LW}\,{\ensuremath{\mathrm{sec}}}^{-1}$, where $F_{\rm LW}$
is the flux within LW bands \citep{greif2009a}.}
\subsection{X-Ray Emission}
\label{xray_sec}
A compact binary may be able to produce
radiative feedback in the form of x-rays. In this section, we
describe the treatment of the radiation from such an x-ray binary source, if present.
Our methodology is based on \cite{jeon2012}. In the
local universe, the non-thermal emission spectrum from accreting back holes or neutron
stars can be expressed as $F_{\nu}\propto \nu^{\beta}$, where
$\beta = -1$ is the spectral index. { More precisely, the full spectrum is a combination
of a power-law component, representing non-thermal synchrotron radiation with the luminosity
formula $L \sim \dot{M}c^2$, where $\dot{M}$ is the Bondi-Holye accretion model \citep{bondi1944},
and a thermal multi-color disk component}. We here ignore the latter, as we are only interested
in the x-ray feedback on the general IGM, where only the optically-thin, non-thermal photons
contribute.
We conservatively assume that the BH emission physics in the early universe is identical to the local case,
and we therefore apply the same spectra for the black holes originating from the first stars.
The propagation of high-energy photons is assumed to result in an isotropic radiation field,
$\propto 1/r^2$, which only depends on the distance from the BH.
The corresponding photo-ionization and photo-heating rates can be
written as \citep{jeon2012}:
\begin{equation}
\label{ion1}
k_{\rm ion} = \frac{\dot{K}}{r^2_{\unit{pc}}} \left(\frac{\dot{M}_{\rm BH}}{10^{-6}
{\ensuremath{\mathrm{M}_{\odot}}}\,\ensuremath{\mathrm{yr}}^{-1}}\right),
\end{equation}
where
\begin{equation}
\dot{K} = [ 1.96, 2.48, 0.49 ] \times 10^{-11} \hspace{0.5cm} \quad\ensuremath{\mathrm{s}}^{-1} .
\end{equation}
and
\begin{equation}
\label{heat1}
\Gamma= \frac{n_{j}\dot{H}}{r^2_{\unit{pc}}}
\left(\frac{\dot{M}_{\rm BH}}{10^{-6}{\ensuremath{\mathrm{M}_{\odot}}}\,\rm yr^{-1}}\right) \left(1-f_{\rm H/He}\right) \hspace{0.1cm},
\end{equation}
Here
\begin{equation}
\dot{H} = [ 7.81, 9.43, 1.63 ] \times 10^{-21} \hspace{0.3cm} \quad{\ensuremath{\mathrm{erg}}}\,\ensuremath{\mathrm{cm}^{-3}}\,\ensuremath{\mathrm{s}}^{-1}
\end{equation}
for H~I, He~I, He~II, respectively. $n_{j}$ is the corresponding number density,
$r_{\ensuremath{\mathrm{pc}}}$ the distance from the star in units of ${\ensuremath{\mathrm{pc}}}$, and $\dot{M}_{\rm BH}$ the mass accretion rate. Finally, $f_{\rm H/He}$ are the fractions of the total photon energy expended in
secondary ionizations \citep{shull1985}.
\subsection{Supernova Explosion and Metal Diffusion}
\label{sn_sec}
After several million years, the massive Pop~III
stars eventually exhaust their fuel, and many of
them might have died as supernovae or black holes.
As we discussed above, the first SN explosions may be extremely powerful,
accompanied by huge outputs of energy and metals.
Here, we briefly discuss how we model SN explosions in
our cosmological simulations.
When the star reaches the end of its lifetime, we initialize
a SN blast wave by
distributing the explosion energy among the SPH particles surrounding
the sink that had marked the location of the Pop~III star. Because the resolution of the simulation
is about 1~pc, we cannot resolve individual
SNe in both mass and space. Instead, we here select the particles within a region
of 10~pc to share the supernova's thermal energy and metal yield. The gas within
this region has mean temperatures about several million Kelvin. On this scale, the blast
wave is still close to its energy-conserving phase. The explosion energy of hypernovae
and pair-instability SNe can be up to $10^{52}-10^{53}\,{\ensuremath{\mathrm{erg}}}$, whereas a conventional
core-collapse SN has about $10^{51}\,{\ensuremath{\mathrm{erg}}}$.
Mixing plays a crucial role in the transport of metals, which could be the most
important coolant for subsequent star formation. We here cannot resolve the fine-grained
mixing due to
fluid instabilities in the early SN ejecta, developing on a scale far below 1~pc.
However, we approximately model the coarse-grain mixing on cosmologically relevant
timescales by applying the
SPH diffusion scheme from \cite{greif2009b}.
A precise treatment of the mixing of metals in cosmological simulations
is not available so far because the turbulent motions responsible for
mixing can cascade down to very small scales, far beyond the resolutions
we can achieve now. We here, therefore, approximately model the effect of
unresolved, sub-grid turbulence as a diffusion process, linking the corresponding
transport coefficients to the local physical conditions at the grid scale.
For further algorithmic details, we refer the reader to \cite{greif2009b}.
After the SN explosion, metal cooling must be considered in the cooling network. We assume that C, O,
and Si are produced with solar relative abundances, which are the dominant coolants for the gas contaminated
by the first SNe.
There are two distinct temperature regimes for these species. In low temperature gas,
$T\,<\,2\,\times\,10^4\,\ensuremath{\mathrm{K}}$, we use a chemical network presented in \cite{glover2007}, which follows
the chemistry of C, C$^+$, O, O$^+$, Si, Si$^+$, and Si$^{++}$, supplemental to the primordial species discussed above.
This module
considers the fine structure cooling of C, C$^+$, O, Si, and Si$^+$, whereas molecular
cooling is not taken into account. At high temperatures,
$T\,\geq\,2\,\times\,10^4\,\ensuremath{\mathrm{K}}$, due to the increasing number of ionization states, a full non-equilibrium
treatment of metal chemistry becomes computationally prohibitive. Instead of directly
solving the cooling network, we use the cooling rate table based on \cite{suther1993}, which gives effective
cooling rates for hydrogen and helium line cooling, as well as bremsstrahlung, at different metallicities.
Dust cooling is not included because it would only become important at densities much higher than what is
reached in our simulations.
\section{Stellar Models}
\subsection{Single Star Models}
We use the Pop~III stellar models of $15\,{\ensuremath{\mathrm{M}_{\odot}}}$ (S15), $30\,{\ensuremath{\mathrm{M}_{\odot}}}$ (S30),
$45\,{\ensuremath{\mathrm{M}_{\odot}}}$ (S45), and $60\,{\ensuremath{\mathrm{M}_{\odot}}}$ (S60) stars from the library of
\citet{heger2010}. These models are non-rotating stars and we assume
no mass loss during the stellar evolution. We summarize key model characteristics
in Table~\ref{flms_1ife_models}, where we distinguish the life
span of main-sequence (MS; central hydrogen burning) and post-MS
(until supernova) evolution. The S15 model evolves in total for about $10.5\,\ensuremath{\mathrm{Myr}}$
before encountering an iron core-collapse supernova with an explosion
energy of $1.2\times10^{51}\,{\ensuremath{\mathrm{erg}}} =1.2\,\ensuremath{\mathrm{B}}$ and a metal yield of $1.4\,{\ensuremath{\mathrm{M}_{\odot}}}$.
Similarly, the S30, S45, and S60 models evolve for $5.7$, $4.4$, and $3.7\,\ensuremath{\mathrm{Myr}}$, respectively.
In assigning the final fate of our three most massive models, current understanding is
still quite uncertain, and we here focus on a few illustrative possibilities.
Specifically, in the cases of
stars with masses of $30\,{\ensuremath{\mathrm{M}_{\odot}}}$ and above, we assume that they do not die as
conventional iron core-collapse SNe. Instead, we assume that each of them either collapses
into a black hole (BH), triggering no explosion, or explodes as a hypernova (HN) with
$10\,\ensuremath{\mathrm{B}}$ explosion energy, based on the collapsar model \citep{woosley1993}.
When the star dies as a BH, all metals within the star fall back into the BH, such
that no enriched material is ejected.
In the HN case, the S30, S45, and S60 models synthesized about $6.8$, $13.2$,
and $20.6\,{\ensuremath{\mathrm{M}_{\odot}}}$ of heavy elements, and disperse them into the primordial IGM. We summarize
the possible stellar fates in Table~\ref{flms_fate}. For simplicity and to limit the
number of cases in this study, we focus only on these simplified, limiting cases, and
refer the reader to \citet{heger2010} for an extended discussion
of Pop~III SN models.
Massive Pop~III stars are strong sources of UV radiation. Because of the
predicted weak stellar
winds from Pop~III stars \citep{kudri2002}, there is no notable x-ray source contributing
to the ionizing photon budget resulting from such winds. The UV radiation thus exclusively
emerges from the hot stellar surface. Ionizing photon fluxes for all stellar models are
given in Figure~\ref{flms_photons}, where evolutionary effects are evident. Specifically,
fluxes exhibit a gradual increase in the lower energy bands
(LW and H~I), and a decrease in the higher energy band (He~II). The hydrogen ionizing flux
in the S60 model is about 10 times larger than for S15. The flux of more energetic
photons is highly sensitive to stellar mass, such that for photons capable of ionizing
He~II{}, their ratio is 100:10:1 for S60:S30:S15. Because the life span
of each star is different, we evaluate the overall ionizing power of an individual star
by calculating the total amount of ionizing photons emitted before the star dies.
As shown in Table~\ref{flms_flux}, the S60 model produces about 2, 3, and 6 times more
H~I{}, He~I{}, He~II{} ionizing photons than S30. The cumulative ionizing power of S60 is also much
stronger than the production from four S15 models combined. This implies that the overall radiative feedback
of Pop~III stars becomes weaker if their mass scale shrinks due to fragmentation.
Since our ray-tracing scheme cannot resolve a time scale less than a year, the
radiation flash from the SN itself is not included here because the SN transit only lasts for
about a few weeks to months. In principle, there could be a flash
of hard radiation from the shock breakout that may eventually be observable
\citep{scannapieco2005}, but the total energy in this flash is small, due largely to
the typically small radii of the Pop~III stars at the time of death.
Besides, the radiation from the subsequent main part of the supernova light
curve is largely at longer wavelengths and does not contribute much to the ionization.
Our calculation shows that the total ionizing photon production during the SN is about $10^{-5}$
that of the MS phase. Furthermore, most of the SN explosion energy goes
into the thermal and kinetic energy of the ejecta \citep{souza2013, whalen2014a}.
\subsection{Binary Star Models}
The ubiquitous fragmentation of primordial star forming clouds allows the widespread
formation of binary stars. We consider binary stars with a total mass
of $60\,{\ensuremath{\mathrm{M}_{\odot}}}$, specifically a system with two stars of equal mass (mass ratio 1:1)
and another one with a mass ratio of 3:1, in accordance with current theoretical understanding.
\citet{shu1987} and \citet{larson2003} suggest that binary stars with (close to)
equal mass are common in the local universe. Our binary models thus contain both asymmetric
cases of $45\,{\ensuremath{\mathrm{M}_{\odot}}} + 15\,{\ensuremath{\mathrm{M}_{\odot}}}$ Pop~III stars and symmetric ones of
$30\,{\ensuremath{\mathrm{M}_{\odot}}} + 30\,{\ensuremath{\mathrm{M}_{\odot}}}$ Pop~III stars.
To keep the binary models simple, we do not consider binary star mergers, and neglect
any mass ejection as an idealized approximation. We are aware of and advise the
reader of the shortcomings of these simplifications, compared to more realistic binary
models \citep{langer2012}.
Below, we discuss the select binary scenarios considered here (see Table~\ref{binaryfate}).
\begin{enumerate}
\item{\bf S30+S30 (HN)}:
This model contains two $30\,{\ensuremath{\mathrm{M}_{\odot}}}$ Pop~III stars, each represented by our S30 model.
We assume both stars form at the same time
and evolve together for about $5.7\,\ensuremath{\mathrm{Myr}}$, after which they both die as a HN.
\item {\bf S30+S30 (BH)}:
This scenario is very similar to S30+S30 (HN), but now both stars
directly collapse into BHs, without triggering a SN.
\item {\bf S45+S15 (BH)}:
The binary contains two stars, represented by S45 and S15 models. Both of them form at the
same time and evolve together for $4.4\,\ensuremath{\mathrm{Myr}}$. Subsequently, the S45 component dies as a BH,
and the system becomes an x-ray binary source due to the transfer of mass
from the S15 companion onto the BH. We approximately assume that the entire mass of the primary
collapses into the BH, and that the system remains bound.
We employ a mass transfer rate of $10^{-6}\,{\ensuremath{\mathrm{M}_{\odot}}}$/yr, active for about
$10\,\ensuremath{\mathrm{Myr}}$. Because the S15 secondary would lose much of its mass during this x-ray binary phase,
we assume a white dwarf (WD) death, without a SN explosion.
\item {\bf S45+S15 (HN)}: Both stars again form at the same time
and evolve together for $4.4\,\ensuremath{\mathrm{Myr}}$, then the S45 primary dies as a
HN. The S15 component evolves for another $6\,\ensuremath{\mathrm{Myr}}$, then
dies as a CCSN. For a wide binary, the kick
velocity is small, and the star will die essentially at the
location of the original binary. However, in a close binary situation,
the orbital velocity is high, and the S15 secondary would acquire a kick of about
$100\,\ensuremath{\mathrm{km}\,\mathrm{s}^{-1}}$, assuming a binary separation of about 1~AU.
Such a velocity might allow the star to travel close to
1~kpc before dying as a SN. In this case, the ejected S15 model could
become a moving radiation source and disperse its metal production from a site
quite remote from where it formed. \citet{conroy2012} suggested that the
runaway massive stars could also contribute to the reionization of the Universe.
For simplicity, we do not explore this intriguing scenario in this paper.
\end{enumerate}
Since we are considering only non-interacting cases, their resulting UV flux can easily be obtained by
summing over the individual contributions from the two component stars. Figure~\ref{fbn_photons} shows the ionizing
photon flux of the S45+S15, S30+S30 and S60 models, respectively, and Table~\ref{ion.binary} lists the total
number of ionizing photons emitted. The UV fluxes of the three models are comparable
within a factor of 2. However, a single star, S60, still produces stronger flux than binary
systems with the same overall mass. Differences are largest in the
amount of ionizing He~II{} photons, as this rate is extremely sensitive to the mass of the primary.
\section{Results}
\label{results}
The impact of the first binary stars on the early universe can be divided
into three different classes; UV-radiation, supernova, and x-ray.
The UV-radiative feedback here refers to the soft (LW) and hard (ionizing) photons produced
by the binaries during their stellar evolution. To facilitate
convenient comparison, we also briefly discuss the feedback from single Pop~III stars,
referring the reader to the extensive literature on this topic for further detail.
We note that our single star models improve on earlier treatments by including some
key features, such as the realistic modeling of the time dependence of the UV fluxes
in response to the underlying stellar evolution.
We present the results from our cosmological simulations following the
chronologically order of how the first single or binary stars evolve:
birth, evolution, and demise. { Because the results contain many cosmological simulations
of different feedback models, we summarize the stellar models, their feedbacks, and associated results in Table~\ref{crossref}.}
\subsection{Radiative Feedback}
\label{flms_results}
The first stellar system forms inside a minihalo, with a total mass of about
$10^6\,{\ensuremath{\mathrm{M}_{\odot}}}$ and a virial radius of $\sim 100$\,pc, located in the
region with the highest mass resolution at $z \sim 27$. This allows us to resolve
key small-scale features of the ensuing stellar feedback. Once the gas density
inside the star-forming cloud exceeds the threshold for sink creation, either
a single star or a binary system is assumed to promptly form.
By using sink particles, the realistic assembly history of a protostar via an
extended phase of accretion is not modelled, which still is computationally
prohibitive. Instead, we assume that the sink particles immediately represent
fully developed Pop~III stars, acting as sources of UV radiation. The
gas inside the halo is rapidly photo-heated up to temperatures of $T\approx 2\times
10^4\,\ensuremath{\mathrm{K}}$. This drives the sound speed up to $30\,\ensuremath{\mathrm{km}\,\mathrm{s}^{-1}}$, whereas the escape
velocity of the host halo is only about $3\,\ensuremath{\mathrm{km}\,\mathrm{s}^{-1}}$. The photo-heated gas is thus blown
out of the shallow potential well of the minihalo.
UV photons not only heat up the gas but also
ionize the neutral hydrogen and helium. The ionization-front (I-front) propagation begins with a
short supersonic phase (R-type), then switches to a subsonic phase (D-type), because the
I-front is trapped behind a hydrodynamical shock \cite[e.g.][]{gb01, whalen2004,wet10}. The I-front eventually breaks out, jumping ahead of the shock, and supersonically propagates
into the low-density IGM.
Since the lifetime of the individual stars is different, we compare their radiative feedback
when they die and their UV radiation is terminated. We first show the resulting gas temperatures
around the host minihalo in Figure~\ref{flms_2d_rad}. A giant bubble of hot, ionized, gas is created around the central star, with an inner
region that has reached temperatures in excess of $10^4\,\ensuremath{\mathrm{K}}$.
The shapes of these bubbles are very irregular due to the inhomogeneous and anisotropic
distribution of the gas in the surrounding IGM. The bubble sizes reflect the strength of the UV emission rates, which highly depend on stellar mass.
Specifically, our S60 model creates the largest bubble with a size of
about $5\,\ensuremath{\mathrm{kpc}}$. The S15 model, on the other hand, only gives rise to an ionized region
with a size of $\sim 2\,\ensuremath{\mathrm{kpc}}$, and a much cooler gas temperature.
For binary stars, we compare cases of equal total stellar mass in
Figure~\ref{fbn_2d_rad}. Overall bubble sizes for the equal-mass models are comparable.
However, there is a significant difference in the resulting He$^{++}$ regions. Ionizing He$^{+}$
requires photon energies of $h\nu > 54.4$\,eV, four times higher than the threshold
to ionize neutral hydrogen. Here, it greatly matters how the available stellar
mass is divided among the individual components, such that the S60 model exhibits
significantly larger He$^+$ ionizing rates than the binary models of equal mass,
S30+S30, and S45+S15. This difference may be reflected in the strength of the He$^{+}$ recombination line at $1640\,\textup{\AA}$, providing a
distinctive signature to distinguish between Pop~III single and binary systems, since the latter create
much smaller He$^{++}$ regions, for a fixed total stellar mass.
To better evaluate the impact of the radiative feedback, we map the 3D structure of the hot bubbles onto
1D radial profiles in Figure~\ref{flms_1d_h2}. We first discuss the gas density profile. Due to the
UV photoheating, the gas density in the center of the minihalo has dropped to $0.2\,\ensuremath{\mathrm{cm}^{-3}}$ at the end of star's lifetime. The baryonic outflow
extends to a radius of $150\sim200~{\ensuremath{\mathrm{pc}}}$, slightly larger than the virial radius of the host halo, such that any subsequent star formation is suppressed by expelling the gas through
this photo-evaporation.
Besides their hydrodynamic feedback, UV photons also affect the chemistry of the
primordial gas by changing its ionization state, and releasing free electrons which
can catalyze H$_2$ formation. The weaker UV emission of the S15 model results in a
relatively smaller H$^{+}$ region, whereas those of
S60, S45, and S30 have radii close to $2\,\ensuremath{\mathrm{kpc}}$.
The difference in the central gas density profile is mainly due to the duration over which photo-heating is active.
Binary models exhibit longer overall lifetimes, allowing the gas to escape farther into the IGM, such that the resulting gas densities within the halo are lower.
It is not clear whether the UV radiation
can penetrate into nearby minihalos and affect their star formation or not. When the stars die, and if there
are no additional heating sources, the ionized gas will cool and then recombine, eventually extinguishing the fossil H$^{+}$ regions. The relevant timescales can be estimated as follows \citep{bromm2002}. The cooling time of primordial gas is approximately
$t_{\rm cool}\approx nk_{\rm B}T/\Lambda
\approx 10^5\sim 10^6\,\ensuremath{\mathrm{yr}}$, where $\Lambda \propto n^2$ is the cooling rate, and $k_{\rm B}$
the Boltzmann constant. For the
recombination timescale, we estimate $t_{\rm rec}\approx (k_{rec} n)^{-1}\approx
10^6\sim 10^7\,\ensuremath{\mathrm{yr}}$, where the recombination coefficient is $k_{\rm rec} \approx 10^{-12}\,{\ensuremath{\mathrm{cm}}}^3 {\rm s}^{-1}$,
and the IGM densities
$n\approx 0.01\,{\ensuremath{\mathrm{cm}}}^{-3}$.
\subsection{Supernova Feedback}
The majority of Pop~III stars may finally die as supernovae or directly collapse
into black holes. In this section, we discuss the supernova feedback from single
and binary stars. When the stars die, we assume for simplicity that their
UV radiation is immediately shut off. We employ the SN explosion energies and
metal yields discussed in Section~3, depending on the properties of the progenitor
star. The SN explosion creates a strong
shock wave, traveling with a velocity of $v_{\rm sn} \approx 10^{4}\,
\ensuremath{\mathrm{km}\,\mathrm{s}^{-1}}$. The energy carried by the shock is able to reheat the relic H~II{} region
for an additional $t_{\rm sn} \approx r_{h}/V_{\rm sn}\approx 0.4\,\ensuremath{\mathrm{Myr}}$, assuming
H~II{} region radii of $r_{h}\approx 4\,\ensuremath{\mathrm{kpc}}$. The shock heating in the simulation is
roughly about $0.6\,\ensuremath{\mathrm{Myr}}$ because the shock velocity is slowed down due to radiative
cooling. After the shock dissipates, a hot and metal-rich bubble is left behind in the IGM.
This bubble continues to expand for about another 5 million years, with an increasingly
ill-defined boundary. Eventually,
the thermal energy of the initial ejecta is radiated away, and the expansion stalls. The
mixing of the metals with primordial gas predominantly occurs before stalling.
On the other hand, if a single star directly dies
as a black hole, no feedback is taken into account (but see Section~4.3).
We first discuss the combined feedback from UV radiation and supernovae/black holes,
evaluated $15\,\ensuremath{\mathrm{Myr}}$ after the birth of the stars, in Figure~\ref{flms_2d_all}.
At this moment, the SN blastwave has stalled. The heating from the hot SN ejecta
maintains elevated ionization in the central part of the H~II{} region. The chemical
feedback of the S15 model is initiated by a CCSN with explosion energy of about
$1.2\,\ensuremath{\mathrm{B}}$ and an ejected metal mass of
$1.4\,{\ensuremath{\mathrm{M}_{\odot}}}$. These metals are dispersed out to a radius of $\sim 0.5\,\ensuremath{\mathrm{kpc}}$, resulting
in an average metallicity of $\sim 10^{-5}-10^{-4}\,\ensuremath{Z_\odot}$. According to our stellar
model assumptions, the S30, S45, and S60 cases all die as
a hypernova, dispersing 6.9, 13.3, and 20.7 ${\ensuremath{\mathrm{M}_{\odot}}}$ of heavy elements, respectively.
For the hypernova cases, the metal-enriched bubbles
have reached radial sizes of $\sim 1\,\ensuremath{\mathrm{kpc}}$ with corresponding metallicities
of $10^{-4}-10^{-2}\,\ensuremath{Z_\odot}$. Although all hypernova cases exhibit a similar overall range
of chemical enrichment, there are noticeable differences in the detailed distribution
of metals. These differences in turn may reflect how the preceeding radiative feedback from the
different model stars has shaped the gas distribution inside the IGM, which could
affect the transport of metals later on.
To further facilitate the comparison between the single-star cases, we present
1D profiles of gas density and metallicity in Figure~\ref{flms_1d_metal}.
The strong SN blast waves in all cases substantially suppress the gas density in the
host halo, resulting in $n < 0.01\,\ensuremath{\mathrm{g}\,\mathrm{cm}^{-3}}$, similar to that of the background IGM at this
redshift of $z \sim 26$.
The extent of the central void is about $200\,{\ensuremath{\mathrm{pc}}}$ for the CCSN
case (S15), and $500\,{\ensuremath{\mathrm{pc}}}$ for the HN cases, respectively.
The metallicity profiles for the HN cases are quite similar, but the S30 model
enriches the IGM out to a slightly larger radius in response to its higher
blast wave velocity. Finally, instead of
SN explosions, the stars may directly collapse into a BH, a possibility
which becomes increasingly likely with increasing progenitor mass.
For such a BH fate, the relic H~II{} regions would quickly begin to cool to
about $10^3\,\ensuremath{\mathrm{K}}$, and recombination would suppress the abundance of free electrons.
We now discuss the SN feedback from binary stars.
To enable a meaningful comparison, we fix the total mass in Pop~III
stars, here $60\,{\ensuremath{\mathrm{M}_{\odot}}}$, either locked up in a single star or distributed among
binary partners (S30+S30 or S45+S15). The key question then is whether there are
significant differences or not.
Stellar evolution models strongly
suggest that $60-80\,{\ensuremath{\mathrm{M}_{\odot}}}$ Pop~III stars are likely to die as BHs, possibly
accompanied by very weak SN explosions.
Hence, we also consider the case of a $60\,{\ensuremath{\mathrm{M}_{\odot}}}$ star dying as such a weak
SN with an explosion energy of $\sim 0.1 $B and a metal yield of
$\sim 0.1 {\ensuremath{\mathrm{M}_{\odot}}}$, due to strong fallback during the BH-forming explosion.
The basic trends can be gleaned from Figure~\ref{fbn_metal} and Figure~\ref{fbn_1d_metal}.
As long as at least one component explodes as a HN, the resulting
metal enrichment is very similar. However, the enrichment from the binary systems is
more robust, in the sense that the single S60 star is likely to experience only
a weak explosion accompanied by much reduced heavy element production and dispersal.
Thus, the recent revision of the Pop~III star formation paradigm away from a single-star
outcome to ubiquitous binarity in effect enhances chemical feedback in the early
universe.
\subsection{X-Ray Feedback}
One particularly interesting variant of the S45+S15 model is the presence of
long-lived x-ray feedback.
Because of the uncertainty in the stellar evolution model, the $45\,{\ensuremath{\mathrm{M}_{\odot}}}$ star
can possibly collapse into a BH instead of blowing up as a SN. In the case of
a close binary, mass transfer from the $15\,{\ensuremath{\mathrm{M}_{\odot}}}$ companion, still alive at this time,
becomes possible. Mass accretion onto the compact object
can efficiently extract the gravitational energy of the infalling material, converting
it into the thermal energy within the accretion disk that leads to x-ray
emission. Unlike the ionizing photons from stars, the x-rays can more easily penetrate
the IGM because of the much reduced opacity at high energies \cite[e.g.][]{mba03,kuhlen2005, jeon2012}.
Here, we assume a constant accretion rate onto the central BH of about
$10^{-6}\,{\ensuremath{\mathrm{M}_{\odot}}}\,\ensuremath{\mathrm{yr}}^{-1}$, which lasts for about 10 million years.
For simplicity, we further assume that the
$15\,{\ensuremath{\mathrm{M}_{\odot}}}$ companion star eventually dies as a white dwarf without a SN explosion, and
neglect its UV radiation during the accretion phase. We
show the impact of such a Pop~III x-ray binary, comparing it with the non x-ray case, in
Figure~\ref{fbn_xray}. In the absence of x-ray emission, the temperature of the
the relic H~II{} region quickly declines and the ionized hydrogen recombines.
An active x-ray binary, on the other hand, provides a prolonged heating source,
maintaining the temperature and ionization inside the relic H~II{} region, and further
heating up the gas beyond its boundary, up to several hundred
Kelvin. More importantly, the x-ray emission can change the free electron fraction
in the IGM which is critical for
H$_2$ formation.
Because the IGM is optically thin to x-ray photons, the resulting heating is
quite homogeneous and isotropic. We also find that the x-ray photons may penetrate the
gas of nearby halos, thus affecting their star formation properties.
\section{Discussions and Conclusions}
\label{flms_discussion}
We have presented the results from cosmological simulations of the
impact of Pop~III stars, specifically focusing on the new effects arising from the
presence of binaries. We improve on earlier simulations by using
updated Pop~III stellar models. In ascertaining the cosmological impact of
the first binaries, we consider their radiative, SN, and x-ray feedback.
By comparing a single $60\,{\ensuremath{\mathrm{M}_{\odot}}}$ star and the corresponding
binary systems with equal total mass, we find that the resulting numbers of
hydrogen-ionizing photons are very similar.
However, for He$^+$ ionizing radiation, the binary stars are significantly
weaker than the single star, because the more energetic UV photon production
is strongly reduced
in the less massive stars. If binary stars thus were the typical outcome of
Pop~III star formation in minihalos, detection of
the distinct emission lines from He$^{++}$ recombination, most prominently the
1640 $\textup{\AA}$ line, would be very challenging.
{Because the x-ray feedback strongly depends on the spectrum of
x-ray binaries which is uncertain in the high redshifts, its observational signature
is more difficult to predict. Recently, \citet{xu2014} and \cite{ahn2014} suggested
that the x-ray feedback of the Pop III binaries can be examined by the 21 cm observations.}
We here explore cases where the stars die as core-collapse SNe or as
hypernovae. In all cases, the mechanical feedback from the explosions
expels the gas from the host systems, thus suppressing any subsequent
star formation in the same halo, at least for of order 10\,Myr. To trigger any
further star formation, the expelled gas needs to be driven back to high
density, possibly as a result of halo mergers within bottom-up structure
formation later on.
For a $15\,{\ensuremath{\mathrm{M}_{\odot}}}$ star, the expected final fate is a core-collapse SN, whereas
the fate of $30\,{\ensuremath{\mathrm{M}_{\odot}}}$,
$45\,{\ensuremath{\mathrm{M}_{\odot}}}$, and $60\,{\ensuremath{\mathrm{M}_{\odot}}}$ Pop~III stars is still poorly understood. To
bracket parameter space, we here assume
that they die as energetic hypernovae, weak SNe, or directly collapse into BHs.
It is evident that for such more massive progenitors, chemical feedback
can span a broad range of possibilities. The most effective feedback is
provided by the hypernova models, where the nearby IGM is typically enriched to
average metallicities of $\sim 10^{-4} - 10^{-3}\,\ensuremath{Z_\odot}$, out to
$\sim 2\,\ensuremath{\mathrm{kpc}}$.
Even single, energetic SNe can impact the early universe on cosmological scales.
We demonstrate this in Figure~\ref{impacts_den_temp}, showing the feedback from
a single $60\,{\ensuremath{\mathrm{M}_{\odot}}}$ star, undergoing a HN explosion.
The resulting feedback significantly enhances the IGM temperature, smoothes out
density structures in nearby halos, and enriches the primordial gas over regions
of $\sim 2\,\ensuremath{\mathrm{kpc}}$. However, the formation of a single massive star inside a
minihalo increases the probability of collapsing into a BH without any
chemical enrichment. On the other hand, binary star formation greatly buttresses
the likelihood that metal enrichment will occur. { In effect, binary formation
is much less likely to keep the early universe metal free, with consequences for the
prompt transition in star formation mode to the low-mass dominated Population~II.}
Realistic binary stellar models could introduce a very rich phenomenology of
evolutionary pathways. Among them are ejection scenarios, where one component is
flung out, such that it may explode in the outskirts of its host halo, as opposed
to the location of its birth, as implicitly assumed here. There is also
rich physics related to the early evolution of the SN remnant, when shock
break-out occurs, and hydrodynamical mixing takes place. All of these
intriguing aspects of binary progenitors will be explored in future simulations,
with greatly improved spatial and temporal resolution. The exploratory models
presented here, however, already clearly indicate the importance of studying
binary-related feedback in the early universe. The imprint from the
violent death of Pop~III stars, in all its variety, might soon be amenable to
empirical testing with the {\it James Webb Space Telescope (JWST)}, to be launched
around 2018. One key aspect of this search will be to distinguish the possible signature
of binarity in primordial star formation.
\section*{Acknowledgments}
{Authors thank the anonymous referee for many constructive suggestions}.
We also thank Jarrett Johnson, Dan Whalen and Ryan Cooke for many useful discussions.
KC was supported by a IAU-Gruber Fellowship, a Stanwood
Johnston Fellowship, and a KITP Graduate Fellowship. Work at UCSC has been
supported by DOE HEP Program under contract DE-SC0010676; the NSF
(AST 0909129) and the NASA Theory Program (NNX14AH34G).
VB acknowledges support from NSF grant AST-1009928 and NASA grant NNX09AJ33G.
AH acknowledges support by an ARC Future Fellowship (FT120100363) and a
Monash University Larkins Fellowship. This work has been supported by the DOE
grants; DE-GF02-87ER40328, DE-FC02-09ER41618 and by the NSF grants;
AST-1109394, and PHY02-16783. All numerical simulations were performed with allocations from the University
of Minnesota Supercomputing Institute and the National Energy Research Scientific Computing Center.
|
\section{Introduction}
\label{int}
The Hall effect is a fundamental transport property of metals and
semiconductors. It can provide information on the carrier
densities as well as on other interesting features of the electronic band
structure. Surprisingly and in spite of considerable work in the
past, the Hall coefficient ($R_{\rm H}$) of graphite, a highly
anisotropic material composed by a
stack of graphene layers with Bernal stacking order
(ABA...), in particular the temperature and magnetic field
dependence of $R_{\rm H}$ reported in literature is far from
clear. For example, early data on the low-field Hall coefficient
obtained in single-crystalline natural graphite (SCNG) samples
showed that it is {\em positive} at fields below and negative
above $\sim 0.5~$T at a temperature $T = 77~$K \cite{sou58},
suggesting that holes are the majority carriers. This result
appears to be at odd to several other studies on the graphite band
structure obtained in highly oriented pyrolytic graphite (HOPG)
samples
\cite{kelly,gru08,zhou061,zhou062,lee08,sug07,sch09,orl08JP,gon12}
that suggest that electrons are the majority carriers, unless one
argues in terms of different mobilities of the majority carriers,
an interpretation that was used indeed in the past.
The difference between the reported Hall coefficient obtained in
the SCNG and different HOPG samples was the subject of a short
paper in 1970 where the authors concluded that the positive
low-field Hall coefficient is observed for samples with long
enough mean free path, i.e. less lattice defects, whereas the
negative sign results from boundary scattering in HOPG samples due
to the relatively small single crystalline grains \cite{coo70}. At
high enough applied magnetic fields or high enough temperatures,
this coefficient, however, turned negative \cite{sou58,coo70}. A
positive low-field Hall coefficient was already reported in 1953
for graphite samples with small crystal size (less annealing
temperature); it decreased with temperature and changed sign at a
sample dependent temperature \cite{kin53}. When the crystalline
grain size in the samples was larger than $\sim 0.5~\mu$m, the
Hall coefficient was always negative, at least at $T \ge 77~$K
\cite{kin53}. Similar results were obtained in carbons and
polycrystalline graphite samples with different crystal size in
Ref.~\onlinecite{mro56}, where the authors recognized further that
the Hall coefficient was highly dependent on the {\em alignment}
of crystallites in the samples. Note that these two last reports
\cite{kin53,mro56} are in apparent contradiction to the
relationship between crystal size and positive sign of $R_{\rm H}$
given in Ref.~\onlinecite{coo70}. It is therefore suggestive that
one extra parameter related to the alignment of the crystalline
grains in the samples could provide a hint to solve this
contradiction.
In the studies of Ref~\onlinecite{bra74} a positive Hall
coefficient was reported at 4.2~K that became negative at $\mu_0H
\ge 0.05$~T for different graphite samples. In that work
\cite{bra74} the positive low-field Hall coefficient was explained
within the two-band model arguing that it is due to the higher
mobility of the majority holes in comparison with the mobility of
the majority electrons. However, to understand its behavior as a
function of field and temperature, three types of carriers had to
be introduced in the calculations \cite{bra74}. The low-field
coefficient of different Kish graphite samples was reported in
Ref.~\onlinecite{osh82}. For the ``best" Kish sample, defined as
the one with the largest resistivity ratio $\rho(300)/\rho(4.2)$,
the Hall coefficient was positive at low fields and turned to
negative at $\mu_0 H \simeq 0.6~$T at 4.2~K, similarly to the
results for some of the graphite samples reported in
Ref.~\onlinecite{bra74}. The temperature dependence of the
zero-field Hall coefficient for the ``best" Kish specimen was
interpreted \cite{osh82} taking into account the trigonally warped
Fermi surfaces in the standard Slonczewski-Weiss-McClure's model
\cite{sw58,mcc2}. Interestingly, the lesser the perfection of the
Kish graphite samples the larger was the field where the Hall
coefficient changed sign \cite{osh82}.
Recently published Hall measurements in micrometer small and thin
graphite flakes, peeled off from HOPG samples, showed a positive and nearly field
independent Hall coefficient at $T = 0.1~$K up to 8~T
applied fields \cite{bun05}. A positive Hall coefficient
was also observed in similar graphite flakes at $77 \le T \le
300~$K, which decreased with temperature, it was field independent to $\mu_0H \simeq 1$~T, decreasing at higher fields
\cite{van11}. Interestingly, both results \cite{bun05,van11} are
in rather good quantitative agreement with the result for the
low-field Hall coefficient reported 56 years ago for the bulk SCNG
\cite{sou58}, in clear contrast to reports in HOPG bulk
samples \cite{kelly,coo70,yakovprl03,kempa06,kopepl06,sch09}.
Further studies on bulk HOPG samples showed the existence of an
anomalous Hall effect and a negative Hall coefficient at low
fields, interpreted as the result of a magnetic field induced
magnetic excitonic state \cite{kopepl06}. Also the quantum Hall
effect (QHE) (with electrons as majority carriers) has been
reported in some bulk HOPG samples at high enough fields
\cite{yakovprl03,kempa06}. But the QHE in graphite as well as
other interesting features of the Hall effect behavior like the
hole-like contribution with zero mass \cite{yakadv07} in bulk HOPG
samples, appear to be strongly sample dependent
\cite{sch09}. A short resume of the literature results can be
seen in Table~\ref{Tab.1}. Evidently, all these, apparently contradictory
results indicate us that we need a re-evaluation of the sign,
temperature and field behavior of the Hall coefficient. The whole
reported studies show us that we do not know with certainty, which is
the {\em intrinsic} value of the Hall coefficient in ideal
graphite and which is the origin of all the observed
differences between samples of different origins and microstructure.
\begin{widetext}
\begin{table}
\squeezetable
\label{Tab.1}
\begin{ruledtabular}
\begin{tabular}{l|c|c|c|c|c|l}
$\bf{Ref.}$&$\bf{year}$&$\bf{sample}$&$\bf{T-range}$&$\bf{field-range}$&$\bf{R_{\rm H}}$&$\bf{comment}$\\
\colrule
\colrule
\onlinecite{kin53} & 1953 & PCG & $\ge 4.2~$K & $< 1~$T & positive & for small grains only$^3$\\
& & & & & &\\
\onlinecite{mro56} & 1956 & PCG & 77~K/300~K & 0.65~T & positive & for small grains only$^4$\\
& & & & & &\\
\onlinecite{sou58} & 1958 & SCNG & $\le 77~$K & $<$ 0.5~T & positive & negative at large fields and temperatures\\
& & & & & &\\
\onlinecite{coo70} & 1970 & HOPG/SCG &77~K & $< 0.1~$T & positive & negative upon mean free path$^2$\\
& & & & & &\\
\onlinecite{bra74} & 1974 & $(\dag)$ & 4.2 K & $<$ 50 mT & positive & negative at higher fields\\
& & & & & &\\
\onlinecite{osh82} & 1982 & Kish graphite & 4.2 K & $<$ 600 mT &
positive & ${R_{\rm H}(H)}$ changes sign$^5$\\
& & & & & &\\
\onlinecite{yakovprl03} & 2003 & HOPG &0.1~K$ \le T \le 20~$K &$ \le 9~$T & negative & quantum Hall effect (QHE)$^6$ \\
& & & & & &\\
\onlinecite{bun05} & 2005 & graphite flakes$^1$ & 0.1 K & $<$ 8 T & positive &\\
& & & & & &\\
\onlinecite{kempa06} & 2006 & HOPG &2~K, 5~K & $\le 9~$T & negative & quantum Hall effect (QHE)$^6$\\
& & & & & &\\
\onlinecite{kopepl06} & 2006 & HOPG & 0.1~K$\le T \le 100~$K & $\le 0.5~$T& negative & Anomalous Hall effect\\
& & & & & &\\
\onlinecite{sch09} & 2010 & NG/HOPG & 10~mK & $<$ 10~T & negative & ($\ast$) \\
& & & & & &\\
\onlinecite{van11} & 2011 & graphite flakes$^1$ & $77 \le T \le 300$~K & $<$ 1 T & positive &\\
& & & & & &\\
\end{tabular}
\end{ruledtabular}
\caption{Selected publications that report on the Hall
coefficient $R_{\rm H}$ of different graphite samples at different
fields and temperatures. NG: natural graphite, SCNG: single
crystal natural graphite, PCG: polycrystalline graphite.
$^1$Peeled off from HOPG. $^2{R_{\rm H}}>0$ for long mean free
path, ${R_{\rm H}} < 0 $ for short mean free path or at higher
fields. $^3$Negative for grain size $ > 0.5 \mu$m. Crossover from
positive to negative at a sample dependent temperature.
$^4$Positive $R^0_{\rm H}$ for small grains, negative for grain
size $> 0.5 \mu$m, and strong orientation dependence of $R^0_{\rm
H}$. $^5$ Turns negative at large field for samples with imperfect
structure.$^6$With electrons as majority carriers. ($\ast$): No
Hall data shown at low fields. ($\dag$): The samples were obtained
by crystallization from a solution of carbon in iron and then
purified at 2000$^{\circ}$C in a flux of chlorine. The platelets
had a thickness of $\sim 100~\mu$m~$ \lesssim d \lesssim 9~$mm. }
\end{table}
\end{widetext}
In this work, we argue that one main reason for the observed
differences of the Hall coefficient between samples is related to
the existence of two dimensional (2D) interfaces
\cite{ina00,bar08,gar12}. Moreover, in some of them Josephson
coupled superconducting regions exist, oriented parallel to the
graphene layers of the graphite matrix
\cite{bal13,schcar,esqarx14}. The interfaces in graphite, whose
contribution to the Hall effect we discuss in this work, are grain
boundaries between crystalline domains with slightly different
orientations. Those crystalline domains and the two-dimensional
borders between them, can be recognized by transmission electron
microscopy when the electron beam is applied parallel to the
graphene planes of graphite, see e.g. Fig.~1 in
Ref.~\onlinecite{esqarx14}, Fig.~1 in Ref.~\onlinecite{bar08} or
Figs.~2.2 and 2.9 in Ref.~\onlinecite{ina00}. The interfaces can
be located at the borders of slightly twisted crystalline Bernal
stacking order regions (ABA...) or between regions with Bernal and
rhombohedral staking order (ABCAB...) regions. They can be
recognized usually by a certain gray colour in the TEM pictures
\cite{esqarx14,bar08}. From TEM pictures we obtain that the
distance between those interfaces can be between $\sim 30~$nm and
several hundreds of nm upon sample \cite{bar08,schcar}. Therefore,
the thinner the graphite sample the lower the probability to have
interfaces and to measure their contribution to any transport
property.
The twist angle $\theta_{\rm twist}$, i.e., a rotation with
respect to the $c-$axis between single crystalline domains of
Bernal graphite, may vary from $\sim 1^\circ$ to $< 60^\circ$
\cite{war09} while the tilting angle of the grains with respect to
the $c$-axis $\theta_{\rm c} \lesssim 0.4^\circ$ for the best
oriented pyrolytic graphite samples. As emphasized in
Ref.~\onlinecite{esqarx14}, in case the twist angle is small enough,
the grain boundary can be represented by a system of screw
dislocations or a system of edge dislocations if the misfit is in
the c-direction with an angle $\theta_{\rm c}\neq 0$. A system of
dislocations at the two-dimensional interfaces or topological line
defects can have a large influence in the dispersion relation of
the carriers \cite{fen12,San-Jose2013} and trigger localized
high-temperature superconductivity \cite{vol14}.
There have been several theoretical studies predicting high
temperature superconductivity at the rhombohedral (ABC) graphite
surface \cite{kop11,kop13,vol13} or at interfaces between
rhombohedral and Bernal (ABA) graphite \cite{mun13,kop12}. We note
that rhombohedral graphite regions were also recognized embedded
in bulk HOPG samples \cite{lin12,hat13}. Theoretical studies
indicate an unusual dependence of the superconductivity at the
surface of
rhombohedral graphite or at the interfaces between rhombohedral
and Bernal (ABA) graphite in multilayer graphene on doping \cite{mun13}.
Furthermore, calculations indicate that high-temperature
surface superconductivity survives throughout the bulk due to the
proximity effect between ABC/ABA interfaces where the order
parameter is enhanced \cite{mun13}. Following experimental results
that indicate the existence of granular superconductivity at
certain interfaces in bulk HOPG samples of high grade
\cite{bal13,schcar,bal14I}, it is then appealing to take the
contribution of the interfaces in the behavior of the Hall effect
into account. \color{black} Doing this we are able to interpret
several results from literature as well as the Hall coefficient
obtained from three, micrometer small graphite flakes described
below. The characteristics of the embedded interfaces, as for
example their size or area \cite{bal14I} or the twist angle
between the two Bernal graphite blocks \cite{esqarx14} can have a
direct influence on the temperature and magnetic field dependence
of the Hall coefficient of a graphite sample with interfaces.
Furthermore, the alignment dependence reported earlier
\cite{mro56} can be understood arguing that the interfaces are
created or get larger in area the higher the alignment of the
grains. From the presented work we can conclude that the intrinsic
low magnetic field Hall coefficient of ideal graphite is positive, i.e.
hole like. It can change to negative in samples with embedded
interfaces at fields and temperatures high enough to influence
their contribution. On the other hand one expects that if the
interfaces are not superconducting in a given sample, they will
provide an electron-like contribution to the Hall resistance,
which may or may not overwhelm the intrinsic Hall signal of
graphite.
\section{Experimental details}
The graphite flakes we have measured were obtained by a rubbing
method on Si-SiN substrates using a bulk HOPG sample of ZYA grade
from Advanced Ceramics Co. These samples show, in general, well
defined quasi-two dimensional interfaces between Bernal graphite
structures with slightly different orientation around the
$c-$axis. Their distance in the $c-$axis direction is sample
dependent and in general $< 500$~nm and of several micrometer
length in the ($a,b$) plane \cite{ina00,bar08,schcar}. The Pt/Au
contacts for longitudinal and transverse Hall resistance
measurements were prepared using electron beam lithography. The
samples with their substrates were fixed on a chip carrier.
Further details on the preparation can be read in \cite{bar08}. We
have measured three samples labeled S1, S2 and S3 with similar
lateral dimensions but with thickness: $~9 \pm 1~$nm (S2), $~20
\pm 4~$nm (S1) and $~30 \pm 3~$nm (S3). An example of one of the
samples is shown in Fig.~\ref{foto}.
\begin{figure}[]
\begin{center}
\includegraphics[width=0.9\columnwidth]{photo-r.eps}
\caption[]{Optical microscope picture of sample S1 with the two
current and four voltage gold contacts after lithography
development and evaporation of Pt/Au contacts.}\label{foto}
\end{center}
\end{figure}
The transport measurements were carried out using the usual
four-contacts methods in a conventional He$^4$ cryostat and two
of the samples (S1,S2) were also measured in a dilution
refrigerator. Static magnetic fields were provided by
superconducting solenoids applied always parallel to the $c-$axis.
The longitudinal and transverse resistances were measured using a
low-frequency AC bridge LR700 (Linear Research). After checking
the ohmic response in both voltage electrodes of the samples, all
the measurements were done with a fixed current of $1~\mu$A, which
means a dissipation $ \mathrm{d}Q/\mathrm{d}t < 0.1~$nW to avoid
self heating effects. In this work we focus on the Hall
coefficient obtained at fields $\mu_0 H \lesssim 0.2~$T.
\section{Results and discussion}
\subsection{Temperature dependence of the longitudinal resistance}
\begin{figure}[]
\begin{center}
\includegraphics[width=1\columnwidth]{Fig1.eps}
\caption[]{Resistance vs. temperature for the three multigraphene
samples discussed in this work. The measurements were made without
any applied field. The continuous lines were calculated following
the parallel-resistor model from Ref.~\onlinecite{gar12}, see
text. The inset shows the normalized resistance vs. temperature at
different fields applied normal to the interfaces for sample S2.
}\label{fig1}
\end{center}
\end{figure}
Figure~\ref{fig1} shows the electrical longitudinal resistance of
the three samples vs. temperature at zero applied field. Following
the results and the discussion exposed in Ref.~\onlinecite{gar12},
we assume that ideal graphite is a narrow-gap semiconductor. The
observed semiconducting-like temperature dependence in the
longitudinal resistance has been also reported recently for thin
graphite flakes
\cite{van11}. We assume that any
deviation from a semiconducting-like dependence in the
longitudinal resistance is due to extrinsic contributions. The
observed behavior in our samples is similar to that already
published \cite{bar08} and it can be understood taking into
account the contributions of the semiconducting graphene layers
in parallel to that from the embedded interfaces and of the
sample surfaces (open and with the substrate) \cite{gar12}. The
interfaces' contribution is responsible for the maximum in the
resistance observed at $\sim 50~$K and $\sim 20~$K for samples S2
and S3, respectively. We speculate that the sample surfaces are
responsible for the saturation of the resistance at low
temperatures, as in S1 for example. \color{black} The embedded
interface contribution appears to be weaker in sample S1 than in
the other two samples, therefore we expect for this sample a Hall
coefficient with less extrinsic contributions than for the other
two.
With a simple parallel-resistor model one can understand
quantitatively the measured temperature dependence using different
weights between the parallel contributions following the
relation \cite{gar12}:
\begin{equation}
R(T) = (R_i^{-1}(T) +
R_b^{-1}(T))^{-1}\,,
\label{eqrespar}
\end{equation}
see Fig.~\ref{fig1}. The bulk, intrinsic contribution of graphite
is semiconducting-like
\begin{equation}
R^{-1}_b(T) \simeq (A
\exp(E_g/2k_BT))^{-1} + R_d^{-1}\,,
\end{equation}
and the one from the interfaces
and surfaces can be simulated following
\begin{equation}
R_i(T) = R_0 + R_1 T +
R_2\exp(-E_a/k_BT)\,,
\end{equation}
whereas the parameters $A, R_j (j = 0,1,d)$ are free. The constant
term $R_d$ prevents an infinite resistance from the bulk
contribution, which is physically related to defects or (a)
surface band(s). $E_g, E_a$ denote the semiconducting gap and an
activation energy; their values depend on sample within the range
$250~$K~$ \lesssim E_g \lesssim 500~$K and $10~$K~$ \lesssim E_a
\lesssim 50~$K \cite{gar12}. The linear in temperature term ($R_1
T$) could be negative or positive and it is taken as a guess for
the contribution of the surfaces and/or metallic-like interfaces.
The thermally activated function can be interpreted as the
contribution of non-percolative, granular superconducting regions
inside the internal interfaces embedded in the graphite matrix
\cite{gar12}, similar to that observed in granular Al in a Ge
matrix \cite{sha83}, for example. Transport \cite{bal13} as well
as magnetization \cite{schcar} measurements support the existence
of granular superconductivity and Josephson coupling between
superconducting regions at these interfaces.
At fields of the order of 0.1~T applied perpendicular to the
interfaces, the metallic-like behavior (at $T < 50~$K) starts to
vanish, see the results of sample S2 in the inset of
Fig.~\ref{fig1}, as example. This behavior is assigned to the
field-driven superconductor- (or metal-) insulator transition
\cite{kempa00,yakovprl03,tok04,heb05}. Because this behaviour is
absent in thin enough graphite flakes \cite{bar08,gar12} or in
thicker graphite samples without well defined interfaces, the
field-driven transition is not intrinsic of ideal graphite and
should not be interpreted in terms of band models for graphite
with Bernal stacking order \cite{tok04,heb05}. If this
field-driven transition is compatible with the existence of
Josephson coupled superconducting regions at the interfaces
\cite{bal13,gar12,schcar}, we therefore expect that the Hall
coefficient should be influenced at similar applied fields, as we
show below.
\subsection{Temperature dependence of the low-field Hall coefficient}
The low-field Hall coefficient is defined as the $R^0_{\rm H} = d
\lim_{H \rightarrow 0} r_{\rm H} /\mu_0 H$, where $r_{\rm H}$ is
the Hall resistance and $d$ the measured thickness of the sample.
Figure \ref{fig5} shows the temperature dependence of $R^0_{\rm
H}(T)$ for the three samples measured in this work. We note that
it changes sign at $\sim 45~$K and $\sim 70~$K for samples S3 and
S2. For sample S1, $R^0_{\rm H}(T)$ remains positive in the
whole measured temperature range. The observed behavior of
$R^0_{\rm H}$ for sample S1 as well as its absolute value are
similar to the recently reported ones for a mesoscopic graphite
flake of similar thickness \cite{van11}. Due to the large
dispersion of Hall coefficients and their variation with
temperature found in literature (see Table I in Sec.~\ref{int}),
this agreement is remarkable and support early results on a
positive $R^0_{\rm H}$ at low enough fields and temperatures for
graphite \cite{sou58,coo70}.
All the three samples show positive, saturating $R^0_H(T \lesssim
25$~K), see Fig.~\ref{fig5}. The origin of the differences between the low-field Hall
coefficients at low temperatures is not known with certainty.
A quantitative comparison
of the absolute values of the Hall coefficient between different
graphite samples is not straightforward because different
samples have different
interface contributions. The density of these interfaces as well
as the number of those that have superconducting properties depend
on the sample and it is not simply proportional to the thickness
of the sample. Also different absolute values could arise from
differences between the properties of holes and electron
carriers, see Eq.~(\ref{hallc}), whose densities and mobilities
can be influenced by defects and impurity atoms. Therefore,
differences in the absolute values of the Hall coefficient between
the samples should be taken with some care.
\subsubsection{The intrinsic Hall coefficient}
Taking into account the $T-$dependence of the longitudinal
resistance, the results of Ref.~\onlinecite{gar12} and the fact
that $R^0_{\rm H}(T)$ is positive, at least a two-band model
\cite{kelly} is necessary to interpret the data. According to this
model, the Hall coefficient is given by:
\begin{equation}
R^0_H = \frac{\mu_p^2 p - \mu_n^2 n}{e(\mu_p p + \mu_n n)^2}\,,
\label{hallc}
\end{equation}
where $\mu_{p,n}$ are the mobilities for holes with density $p$
and electrons with density $n$, respectively; $e$ is the positive defined
electronic charge. Since $R^0_H>0$, then $\mu_p^2 p > \mu_n^2 n$.
Taking into account the expected band structure of graphite and
for practical purposes, we can assume either: (1) both mobilities
are equal and that the carrier densities are related through $p =
n + \delta > n$ with $\delta \ll p,n$, or (2) $\mu_p =
\mu_n+\delta>\mu_n$, $p= n$ and $\delta \ll \mu_p,\mu_n$. In both
cases $p + n \simeq 2p$ and Eq.~(\ref{hallc}) can be approximated
by either:
\begin{eqnarray}
R^0_{\rm H} & \simeq & \frac{\delta}{4ep^2}~\textrm{with~} \delta = p - n \label{hallc1}\\
\textrm{or} &&\nonumber\\
R^0_{\rm H} & \simeq & \frac{\delta}{2ep\mu_p}~\textrm{with~}
\delta = \mu_p - \mu_n \,. \label{hallc2}
\end{eqnarray}
Obviously, the use of the one-band model equation, i.e. $R^0_H =
1/pe$, provides incorrect values for the carrier concentration.
For example, in Ref.~\onlinecite{van11} the authors obtained
$p(77~$K) $\simeq 7.7 \times 10^{19}~$cm$^{-3}$ or $p_{2D}(77~$K)$
\simeq 2.5 \times 10^{12}~$cm$^{-2}$ per graphene layer. Using
the same one-band model we would obtain for sample S1, $p_{2D}
\simeq 8 \times 10^{12}~$cm$^{-2}$, a value four orders of
magnitude larger than the one obtained for similar samples using a
model independent constriction method \cite{dus11}. Moreover, for
such large $p$ values the use of Eq.~(\ref{hallc1}) would give
$\delta \gtrsim p$ using the measured $R^0_H$. However, if we take
the carrier concentration $p_{2D}(75~$K$) \simeq 5 \times
10^{8}$cm$^{-2}$ (or $p_{3D} \simeq 1.6 \times 10^{16}$cm$^{-3}$)
from Ref.~\onlinecite{dus11}, using Eq.~(\ref{hallc1}) we obtain
$\delta/p = 2.5 \times 10^{-4}$, indicating a very small
difference between electron and hole carrier densities. Or using
Eq.~(\ref{hallc2}) we obtain $\delta/\mu_p \simeq 1 \times
10^{-4}$, indicating a very small difference between the electron
and Hall mobilities.
For sample S1, $R^0_{\rm H}(T)$ can be understood as the parallel
contributions \cite{pet58} from the graphene layers with $p(T),
n(T) \propto \exp(-E_g/2k_{\rm B}T) $ and a roughly temperature
independent term (from the surfaces or some internal interfaces)
that prevents the divergence of $R^0_{\rm H}(T \rightarrow 0)$.
Using the same energy gap that fits the longitudinal resistance,
we can fit $R^0_{\rm H}(T)$ in the whole $T-$range, see
Fig.~\ref{fig5}.
\subsubsection{Interface contribution to the Hall coefficient}
\label{T-dep}
The results for $R^0_H(T)$ of sample S1 and those from
ref.~\onlinecite{van11} indicate us that the sign change above
certain temperature in samples S3 and S2 should be related with
the contribution of interfaces. Real graphite samples with
interfaces are rather complex systems in the sense that the
distribution of input electrical currents inside the sample is not
homogeneous. Without knowing this distribution and the intrinsic
conductivities of the different contributions, quantitative models
are only under certain assumptions applicable. To estimate the
interface contribution we use the model proposed in
Ref.~\onlinecite{pet58} for a bilayer, where the Hall coefficient
of the surface (in our case the interfaces $R^i_H(T)$) and bulk
($R^b_H(T))$ contribute in parallel. The total measured Hall
coefficient is given by
\begin{equation}
R_{\rm H} = \frac{d(R^b_{\rm H} \sigma^2_bd_b + R^i_{\rm H}
\sigma^2_i d_i)}{(\sigma_bd_b + \sigma_id_i)^2}\,, \label{eqnhall}
\end{equation}
where $\sigma_{b,i}$ are the conductivities of the bulk and
interface contributions and $d_{b,i}$ the respective effective
thicknesses, i.e. the total thickness of the sample is $d = d_b +
d_i$. Taking into account the interface density in our HOPG
samples obtained from transmission electron microscopy
\cite{bar08}, we estimate $d_i/d_b \lesssim 10^{-2}$. In this
case Eq.~(\ref{eqnhall}) can be written as:
\begin{eqnarray}
R_{\rm H} &\sim &\frac{R^b_H + R^i_H r_2}{r_1}\,,\\
r_1 &=& (1 + r'_1)^2, r'_1 = \frac{\sigma_id_i}{\sigma_bd_b}\,,\\
r_2 &=& \frac{\sigma_i^2d_i}{\sigma_b^2d_b} = r'^2_1
\frac{d_b}{d_i}\,. \label{eqnhall2}
\end{eqnarray}
The effective parallel contribution of the interfaces can be estimated from
\begin{equation}
R^i_H \sim \frac{R_{\rm H} r_1 - R^b_H}{r_2}\,. \label{eqhallrs}
\end{equation}
\begin{figure}[]
\begin{center}
\includegraphics[width=1.00\columnwidth]{Fig5.eps}
\caption[]{Temperature dependence of the low-field Hall
coefficient for the three studied samples. For sample S3 (black
squares) we show also the Hall coefficient (green squares)
obtained at a field $\mu_0 H = 0.2~$T applied normal to the
interfaces. The data from sample S1 in the main panel of the
figure were multiplied by a constant factor of 6.7. The inset
shows the same data but in a logarithmic temperature scale. The
line through the S1 data points follows the function $R^0_{\rm H}
= (11^{-1} + (4\times 10^{-3}
\exp(300/(2T)))^{-1})^{-1}$. All other lines
are only a guide to the eye.}\label{fig5}
\end{center}
\end{figure}
In clear contrast to sample S1, the $R^0_{\rm H}(T)$ of sample S3
turns to negative at $T > 45~$K. We propose that the origin for
the sign change of $R^0_{\rm H}(T)$ increasing $T$ is due to the
extra contribution of the interfaces when the superconducting
properties of the interfaces vanish.
At low enough temperatures the interfaces with the
superconducting regions \cite{bal13} do not
contribute substantially to the total low-field Hall effect.
Because we are dealing here with the zero- or low-field Hall
coefficient, vortices or fluxons (and their movement) are not
expected to influence the Hall signal.
For sample S3 and using Eq.~(\ref{eqhallrs}) we assume that
$R^i_{\rm H}(T \ll 50~$K$) \simeq 0$ implying $R^0_{\rm H} r_1
\ggg R^b_{\rm H}$. If we assume further that the total low-field
Hall coefficient for this sample and at low temperatures is mainly
given by the bulk contribution, we can roughly estimate the
expected contribution from the interfaces. Using the functions
that fit the temperature dependence of the measured resistance,
see Fig.~\ref{fig1}, the related conductivities can be estimated
from $\sigma^{-1}_i \sim g_i (140 - 0.045 T + 10 \exp(-4/T))$ and
similarly for the bulk contribution
$\sigma^{-1}_b \sim g_b (80 \exp(350/2T))$,
where the constant prefactors $g_{b,i} \sim w_{b,i} \times d_{b,i}/ \ell_{b,i}$
(width $\times$ thickness $/$ length)
are due to the different effective geometry of the two contributions.
Note that neither the samples have perfectly rectangular shape nor the internal interfaces
are expected to follow perfectly the measured external sample geometry.
We estimate the Hall coefficient due to the bulk contribution as
$R^b_{\rm H}(T) \sim ( 20^{-1} + (0.1 \exp(350/2T))^{-1})^{-1}$
assuming that the saturation of $R^0_{\rm H}$ we measured for this
sample, see Fig.~\ref{ri}, can be included in $R^b_{\rm H}$.
Assuming $d_b/d_i = 50$, we estimate the interface contribution to
the Hall coefficient using Eq.~(\ref{eqhallrs}) for three values
of the factor $w_b\ell_i/w_i\ell_b$ that enters in $r'_1$. The
three curves can be seen in Fig.~\ref{ri}. The uncertainty
in the geometrical factors and in the bulk Hall contribution
do not allow a better quantitative estimate of the interface
contribution. Nevertheless, qualitatively the obtained results
for $R^i_{\rm H}(T)$ appear reasonable.
\begin{figure}[]
\begin{center}
\includegraphics[width=0.9\columnwidth]{ri.eps}
\caption[]{Temperature dependence of the low-field Hall
coefficient for the S3 sample. The red line follows the equation
$R^b_{\rm H} = 0.03(20^{-1} + (0.1 \exp(350/(2T)))^{-1})^{-1}$
used as the bulk contribution to the Hall coefficient for that
sample. The dashed lines represent the contribution from the
interfaces $R^i_{\rm H}(T)$ calculated from Eq.~(\ref{eqhallrs})
assuming different geometrical ratios $w_b\ell_i/w_i\ell_b = 200,
300, 500$, from bottom to top lines, respectively.}\label{ri}
\end{center}
\end{figure}
\subsection{Magnetic field dependence of the Hall coefficient}
\label{B-dep}
\begin{figure}[]
\begin{center}
\includegraphics[width=1\columnwidth]{Fig4.eps}
\caption[]{Hall coefficient as a function of the absolute value of
the applied field at several temperatures for sample S3. The inset
shows the field dependence of the Hall resistance at two
temperatures.}\label{fig4}
\end{center}
\end{figure}
We emphasize that for graphite samples with no measurable evidence
for the interface contribution, in the electrical resistivity for
example, the Hall coefficient does not depend on the field, at
least up to 1~T applied normal to the graphene planes
\cite{bun05,van11}. Therefore, a direct way to test our
assumption that the Hall coefficient and its negative value is not
intrinsic -- but due to the extra contribution from the embedded
interfaces with superconducting regions -- can be independently
done measuring it at finite magnetic fields applied normal to the
interfaces. In this case we expect that a magnetic field will have
the same influence on the interface contribution as the
temperature. In other words, a large enough magnetic field will
destroy the coupling between the superconducting regions, or the
superconductivity itself at the interfaces and an extra,
electron-like contribution should be measurable, in principle in
the whole temperature range. Note that mostly electron-like
carriers are expected to be at the interfaces with a density of
the order of $\lesssim 10^{12}~$cm$^{-2}$. This is inferred from
Shubnikov-de Haas (SdH) oscillations in the magnetoresistance
obtained in samples with and without (or with less number of)
interfaces, see, e.g., Ref.~\onlinecite{bar08} (compare there
samples L5 and L7) or Ref.~\onlinecite{oha01} where a clear
decrease in the amplitude of the SdH oscillations decreasing the
thickness of the samples has been reported earlier. The reason why
mostly electron-like carriers appear to be at the interfaces is
related to the nature of the interfaces themselves, a subject that
is being discussed nowadays, see Ref.~\onlinecite{esqarx14} and
Refs. therein. For example, in addition to the twist angle between
the graphite Bernal blocks forming an interface, one has extra
doping through hydrogen or carbon vacancies that influence the
carrier density and the superconducting regions at the interfaces.
For interfaces without superconducting regions we expect an
electron-like contribution to the Hall coefficient with a weaker
field and temperature dependence.
How large should be the magnetic field to affect the coupling
between the superconducting regions or the superconductivity of
the regions itself? An estimate of this field can be directly
obtained from the longitudinal resistance and the metal-insulator
transition (MIT) observed in several high grade graphite samples,
as, e.g., in sample S2, see inset in Fig.~\ref{fig1}, or several
other samples reported in literature
\cite{yakovadv03,yakovprl03,tok04,heb05}. From all the results for
the longitudinal resistance we expect that a field of the order of
0.1~T should be sufficient to influence substantially the Hall
coefficient contribution of the interfaces. Note that the MIT of
graphite is absent for samples without or with negligible amount
of interfaces \cite{bar08,gar12,van11,bun05}. Figure~\ref{fig5}
shows the Hall coefficient of sample S3 at a field of 0.2~T. It
remains nearly temperature independent below 100~K and matches the
results obtained at low-fields at high enough temperatures.
Figure~\ref{fig4} shows the field dependence of $R_{\rm H}$ at
various constant temperatures where we recognize a transition from
positive to negative values at fields similar to those necessary
to trigger the metal-insulator transition observed in the
longitudinal resistance. The results shown in this figure clearly
indicate that a magnetic field has the same influence on the Hall
coefficient as temperature. Note also that a field of the order of
0.1~T is enough to change the sign of the Hall coefficient in the
sample with clear contribution of the interfaces.
Apart from the interface effects, one may expect a decrease of the
(positive) Hall coefficient with field, at large enough fields
when the cyclotron energy $\hbar \omega_c$ is of the order of the
energy gap $E_g$ ($\omega_c = e\mu_0H/m^\star$ and $m^\star$ the
effective electron mass, according to Ref.~\onlinecite{coh61}.
Because in graphite $E_g \lesssim 50~$meV this effect may start be
observable at $\mu_0 H > 1~$T. On the other hand, data obtained
from very thin graphite samples clearly show that the Hall
coefficient is field independent up to 10~T at $T = 0.1~$K
\cite{bun05}, or up to 1~T at $T \geqslant 77~$K \cite{van11}.
Therefore, we can clearly argue that the observed field dependence
in our sample, see Fig.~\ref{fig4}, is not intrinsic of the
graphite structure but has an extrinsic origin, similar to that
reported earlier \cite{sou58,coo70}.\\
\begin{figure}[]
\begin{center}
\includegraphics[width=1\columnwidth]{hall-c.eps}
\caption[]{Low-field Hall coefficient as a function of temperature
for sample S1 from this work (red lozenge); the line through the
points follows the equation $(11 + (0.004
\exp(300/2T))^{-1})^{-1}$. The other points are taken from
different graphite samples reported in literature. $(\star)$:
Taken from Ref.~\onlinecite{van11} for a graphite thin flake; the
line through the points follows the equation $(10 + (0.03
\exp(350/2T))^{-1})^{-1}$. $(\square)$: Taken from
Ref.~\onlinecite{kin53}; the line through the points follows the
equation $(6 + (0.15 \exp(400/2T))^{-1})^{-1}$. The vertical
dashed region at low temperatures with $0.05 \le R^0_H \le
0.2$~cm$^3$/C is from Ref.~\onlinecite{bra74}. $(\ast)$: Hall
coefficient at 100~mK obtained for a 5~nm thick graphite sample
from Ref.~\onlinecite{bun05}. }\label{fig6}
\end{center}
\end{figure}
\section{Comparison with literature and conclusion}
In a recent theoretical work\cite{pal13}, the magnetoresistance
and Hall resistivity for graphite has been calculated using the
usual 3D band structure described by the Slonczewski-Weiss-McClure
model and taking into account only some of its six free parameters
\cite{brandt88}. The obtained results indicate that at relatively
weak applied magnetic fields $< 1~$T, the magnetoresistance
increases linearly with field due to the presence of extremely
light, Dirac-like carriers. Interestingly, in the same field range
the authors found that the Hall coefficient should be positive and
proportional to $\ln|B|$. We note that a linear field
magnetoresistance was indeed reported in this field range and at
low $T$ in a large amount of graphite samples, especially for
relatively thick graphite samples, see for example the
magnetoresistance curves for sample L7 (75~nm thick) in
Ref.~\onlinecite{bar08}. Due to the observed positive Hall effect
at low fields $ < 0.2~$T and at low temperatures, it is of
interest to check whether such a linear field dependence is
observed in the samples described in this work. Figure~\ref{MR}
shows the magnetoresistance vs. applied field below 0.2~T for the
three samples and at low temperatures. Interestingly, none of the
samples show a clear linear field dependence. The sample S1, which
has the smallest contribution from interfaces (see
Fig.~\ref{fig1}) follows approximately a $H^{1.5}$ dependence. The
other two samples, S2 and S3 tends to follow a quadratic
dependence at low enough fields changing to a $\sim H^{1.3}$
dependence at higher fields, in agreement with similar
measurements but in bulk HOPG samples \cite{kop06}. Note that the
absolute value of the magnetoresistance at a given field for our
mesoscopic samples is much smaller than in bulk samples. As an
example, we show in Fig.~\ref{MR} the results for a bulk HOPG
sample of grade A. In the depicted field region the
magnetoresistance follows $\sim H^{1.25}$ but it is 200 times
larger than in the other mesoscopic samples. This difference might
be related to the size dependence of the magnetoresistance when
the mean free path is of the order of the sample size
\cite{gon07,dus11}. The field dependence in this low field region,
however, does not seem to be affected. \color{black}
\begin{figure}[]
\begin{center}
\includegraphics[width=1\columnwidth]{MR.eps}
\caption[]{Low-field magnetoresistance defined as $R(B) - R(0) /
R(0)$ vs. applied field in a double logarithmic scale, for the
three samples studied in this work. The data of samples S1 and S2
were taken at 0.1~K and of sample S3 at 2~K. In the same panel we
show the data for a bulk HOPG sample of grade A ($\circ$) obtained
at 2~K. }\label{MR}
\end{center}
\end{figure}
We note that in the same field range we did not see an
increasing Hall coefficient with field in any of the samples
studied here. Also, for the graphite samples reported in
Ref.~\onlinecite{van11} at $T \ge 77~$K the Hall coefficient is
constant below 1~T and decreases above. In the case of the $\simeq
1 \times 1~\mu$m$^2$ sample reported in Ref.~\onlinecite{bun05} we
note that the magnetoresistance is negligible to 10~T applied
field and $T = 0.1~$K, a result related to the ballistic behavior
of the carriers due to their large mean free path \cite{dus11},
and the Hall coefficient does not depend on field.
A comparison of the Hall coefficient and in general of the Hall
data from literature is not straightforward because the sample
quality as well as the existence of interfaces (of any kind) was
not provided in any of the publications and remains, in general,
unknown. Nevertheless, we can speculate the following trends and
provide a possible answer to the works listed in Tab.~\ref{Tab.1}.
In Refs.~\onlinecite{kin53,mro56}, a positive Hall effect was
observed for small grains only. Upon preparation conditions,
smaller grains should have less
interfaces, see section~\ref{int}, and therefore less contribution from them.
Note that the reported dependence of $R_H$ on the {\em alignment} of the crystallites can be
directly related to the larger probability to have well-defined and larger
interfaces the larger the alignment of the crystallites is. Note that effective critical temperature
depends on the size or area of the interfaces according to recently published experimental results\cite{bal14I}.
The crossover to a negative Hall coefficient at large enough
fields and temperatures observed in
Refs.~\onlinecite{sou58,bra74,osh82,sch09} can be understood in a
similar way as shown in sections~\ref{T-dep} and \ref{B-dep}. In
Ref.~\onlinecite{coo70} the Hall coefficient was reported to be
positive for samples with long mean free path of the carriers and
negative for samples with smaller mean free path or high fields.
The crossover to a negative $R_H$ can be understood in the same
way if some of the interfaces get normal conducting with field.
Now, the reported mean free path dependence of $R_H$ cannot be
simply interpreted in terms of interfaces contribution without
knowing the internal structure of the samples and whether there is
or not a crossover to negative $R_H$ at higher fields and
temperatures. If we assume that the carriers in the graphene
layers of graphite have much larger mobility\cite{dus11} than
those carriers at the interfaces in the normal state, we may
speculate that samples with smaller mean free path have larger
density of interfaces and therefore a negative $R_H$ should be
measured.
The negative QHE measured in the HOPG samples in
Refs.~\onlinecite{yakovprl03,kempa06} at high fields should come
from normal conducting interfaces with a relatively high density
of carriers. Note that those HOPG samples are the ones that show a
high density and well-defined two-dimensional interfaces
\cite{bar08}. That the QHE is not observed in all HOPG samples,
even when they are from the same grade\cite{sch09} (or even batch)
is related to a non-homogeneous distribution of the interfaces.
The observation of the AHE in Ref.~\onlinecite{kopepl06} is
related to defects that trigger magnetic order \cite{dim13}. Note
that even in a mesoscopic graphite sample one can find different
contributions to the transport depending how homogeneous the
sample is, see e.g., Ref.~\onlinecite{barjsnm10}. Therefore, in
this kind of samples it is difficult to measure the intrinsic
contribution coming from the ideal graphene layers. Finally, the
positive $R_H$ measured in Refs.~\onlinecite{bun05,van11} can be
expected since those samples were very probably free from
interfaces due to their small thickness.
\color{black} In general we can state that at low enough fields
and in thin enough samples with low density of interfaces, the
Hall coefficient should be closer the one from the ideal bulk
graphite than at higher fields. Therefore we show in
Fig.~\ref{fig6} low-field coefficients obtained from old and
recent publications at different temperatures, in case these data
were available. From all these data, we conclude that the
intrinsic, low magnetic field Hall coefficient of graphite appear
to be positive with a low-temperature value around $0.1~$cm$^3$/C
and a temperature dependence that follows closely that of a
semiconductor with an energy gap of the order of 400~K, in
agreement with the fits of the longitudinal resistance of
different samples \cite{gar12}. Note that the results shown in
Fig.~\ref{fig6} were obtained from graphite samples with very
different shapes, i.e. bulk samples in
Refs.~\onlinecite{kin53,bra74} and mesoscopic samples with
different areas in Refs.~\onlinecite{bun05,van11}. From this
comparison we would conclude that the low temperature value of the
Hall coefficient does not seem to strongly depend on the defined
Hall geometry.
\color{black}
In conclusion, taking into account the contribution of
superconducting regions at certain interfaces found in real
graphite samples, we provide a possible explanation for the
anomalous temperature and low magnetic field behavior of the Hall
coefficient as well as for its differences between samples of
different origins reported in the last 60 years.
We thank Yakov Kopelevich for fruitful discussion. Part of this
work has been supported by EuroMagNET II under the EC contract
228043.
|
\section{Introduction}
Evolution of open quantum systems is a subject of deep importance, since the
early days of quantum mechanics~\cite{vN55a} and has gained practical importance
since the boom of quantum information~\cite{breuer2007theory}. One is often
not interested in the evolution of the environment, and under very reasonable
assumptions, one can greatly simplify the
discussion~\cite{Lindblad1976,Gorini1976}. Random matrix theory (RMT) is used
to describe statistical aspects of ergodic quantum systems, and more recently,
to provide a framework for the description of
decoherence~\cite{Gorin2003,2008NJPh...10k5016G}. Yet the random matrix theory
of the states themselves, i.e. of the density matrix is still in its infancy. A
first approach is given in \cite{PhysRevLett.104.110501,Nadal2011}, where an
``unbiased'' ensemble of density matrices is constructed imposing on random
covariance matrices the condition that their trace must be one. However
correlations are important. We propose to make the next step in this direction,
by giving a recipe how to impose a fixed purity on such an ensemble, analogous
to the introduction of microcanonical ensembles in statistical mechanics.
Purity is chosen here, exclusively because of its analytic simplicity, but the
development presented is by no means restricted to this quantity. Towards the
end of the manuscript we discuss entropy as an alternate quantity to be fixed.
We shall focus on quantities that are known to measure decoherence or
entanglement. We are in part guided by the direct construction of ensembles
of scattering matrices~\cite{Mello1985254} and both differences and analogies will be
discussed.
The use of purity allows some analytic considerations (not possible with more
complex quantities) after which a very detailed discussion of a four level
central system will follow. There we can illustrate the behavior of density
matrices of fixed purity in terms of simple plots that show prominent features.
We shall also compare the results extensively with those obtained by using
dynamical RMT models for unitary evolutions of the total system (central system
plus environment)~\cite{2008NJPh...10k5016G}, i.e. RMT models for the Hamiltonian in a more traditional
setting~\cite{haakebook, pinedalong}, which have delivered interesting insights
in the field of decoherence. The models here considered range from the
bluntest description of an open quantum system~\cite{Gorin2003,
PhysRevLett.107.080404}, to models which take into account the internal
structure of the Hilbert space~\cite{pinedaRMTshort}, and for which a strong
dependence on the initial condition is discussed.
In section \ref{sec:micro} we shall outline possible approaches to our
problem and compare them with the situation known for random scattering
matrices, where a similar development occurred nearly 30 years ago. In section
\ref{sec:static} we explicitly show how to construct ensembles of fixed purity
obtaining analytic results in many cases. We study in detail how the structure
of the Jacobi determinants enters and why we should take it into account. In
the same section we specialize to a four degree of freedom central system in
order to illustrate our result. In section \ref{sec:rmtdynamical}
numerical calculations are presented for the more
standard approach of using RMT for Hamiltonians and couplings and compare the
results with the ones from the random density matrix ensembles. A wide variety
of RMT dynamics are used and we thus gain insight into the effectiveness of the
new method proposed.
The generality of our arguments is insinuated by the use of von Neumann entropy
as a criterion to establish classes of density matrices in section
\ref{sec:entropy}. Finally we reach some conclusions and give an
outlook.
\section{Microcanonical ensembles for random density matrices}
\label{sec:micro}
We shall devote this section to a discussion about the ensembles to be worked,
the motivation for their introduction and also some alternatives. We start by
recalling the random matrix descriptions of scattering, to contextualize the
situation. A dynamic random matrix theory for the $S$-matrix was developed
(see e.g. ~\cite{Agassi1975145, VWZ}) and has met great success. It is based
on the introduction of random Hamiltonians and channel couplings into standard
formulae for the $S$-matrix. Later a static approach, namely the direct
construction of a random ensemble of $S$-matrices was developed~\cite{SCali,
Mello1985254}, though the dynamical approach remains dominant in the
field. As we propose to develop a ensemble of random density matrices as an
alternative to the existing dynamical description in terms or random
Hamiltonians this back flash will prove helpful and we shall emphasize
analogies and differences.
Both for the scattering matrices and for density matrices a simple ensemble for
what we may consider an appropriate a priori was/is previously known. For
scattering problems this would be the unitary or unitary symmetric matrices
depending on whether we consider time reversal breaking or conserving
situations. Both are so-called ``circular'' versions of Cartan's classical
ensembles \cite{cartanRMT, mehta}. For density matrices
Nadal and Majumdar have recently proposed an ensemble~\cite{Nadal2011}.
These
ensembles in both cases are very useful as a starting point, but they need
some refinement to describe a wider set of physical situations, because in the
scattering case they describe ``total absorption'', which is not very realistic
and in the density matrix case equipartition, which is not very interesting.
For scattering problems the simplest relevant case is the one where an
``optical'' $S$-matrix describes direct processes which vary slowly as a
function of energy (and are taken as energy independent) and a ``compound''
part in the spirit of Niels Bohr~\cite{bohr1939mechanism} which fluctuates fast implying
long times. In this case the former part is assumed known, and usually easy to
measure and/or to evaluate, while the latter is represented by a random matrix
in the dynamical ansatz. The construction of an appropriate ensemble produces
the expected averaged $S$-matrix describing this slowly varying part often
known as the optical $S$-matrix. On one hand the average $S$-matrix is not
really a unitary scattering matrix; it is actually sub-unitary. On the other
hand an average density matrix will always again be a density matrix with all
its properties. This matter is complicated by the fact, that the density
matrix itself defines an ensemble of quantum systems as well as providing a
description of a subsystem of a larger system This is the first important
difference, because in the case of the scattering matrix we explicitly seek an
ensemble of $S$-matrices fluctuating around a mean value which is not a unitary
scattering matrix.
As an average over density matrices is in turn a density matrix,
fixing this average leads to a trivial ensemble.
In this paper we will concentrate on fixing a single scalar, mainly
purity, and construct the ensemble for its value.
As for the case of the $S$-matrix previous efforts exist to work with random
matrices in a dynamical way, but the above noted difference leads to
different approaches. Early work used an ensemble of Hamiltonians to evolve the
total system and then calculate the evolution of purity (and also concurrence
for a central system constituted by two qubits) obtained from the evolution of
an initial product state \cite{pinedaRMTshort, pineda:012305} while more recent
work calculates the average density matrix directly
\cite{gorin2008random,MGS2015} and then obtains the average value say of
purity for that corresponding density matrix. Note that both are dynamical
approaches, but for quantities non-linear in the density matrix, there will
be a difference in the result. In practical examples this difference seems
small, and it seems reasonable to compare the static approach to the dynamical
calculation of density matrices as we will indeed do.
Constructing ensembles with certain restrictions is usually done in one of two
ways: The more notorious way is the use of Jaynes principle \cite{jaynes1,
jaynes2}, where we start form an {\it a priori} probability distribution and
find the distribution that minimizes the information, using Lagrange
multipliers, while fulfilling other conditions. These averages (or
expectation values in a quantum context) may result from experiments or
theoretical considerations~\cite{rlevine87:dynamics}. The other alternative is
to impose the restriction strictly on every element of the ensemble
thus creating a ``microcanonical'' ensemble.
In the case of scattering matrices
the latter approach is impossible, because the optical
$S$-matrix is not
a unitary scattering matrix, and we wish all members of the ensemble to be
such. Historically the first non-trivial ensembles of scattering matrices where
introduced using Jaynes principle and fixing the average $S$-matrix to lowest
order \cite{Mello1985254, SM2Nuclphys, SM31chann}. Later analytic properties of
the $S$-matrix demonstrated that the Poisson kernel multiplied by the measure
of Cartan's ensembles leads to a measure with any desired fixed average $S$-matrix.
More general information-theoretical arguments show that this is in
some sense the minimal information solution among a family of ensembles that
yield the same average $S$-matrix \cite{SCali, Mello1985254}. This
result is used extensively (see e.g. \cite{baranger1994mesoscopic,
brouwer1995generalized, kuhl2005direct}) and yields, where applicable, similar
results to the dynamical model.
For the density matrices with fixed average purity we have more freedom. We
can choose the micro-canonical ensemble, where every member has the same purity
or a canonical ensemble, where purity has a given average value. We have
chosen the former because the original definition of the unbiased ensemble
\cite{Nadal2011} directly includes unit trace as a delta function, and thus it
seemed technically easier to include fixed purity the same way. We do not
expect a significant difference between the canonical and the micro-canonical
approach, but will not explore the conditions under which this is the case in the
present paper.
Existing static models have a limitation as time (or energy) have not been
included at this point: correlations between different times or different
energies (frequencies) cannot be obtained. On the other hand static ensembles
provide a very direct insight as to which properties govern the behavior of the
system beyond obvious things like phase space volume or noise.
\section{Eigenvalue distribution at fixed decoherence}
\label{sec:static}
\subsection{General considerations}
Consider the reduced density
matrix $\rho$ acting in our central system, resulting from taking the partial
trace of a random pure state $|\psi\>$, according to the Haar measure of the
unitary group from a larger space composed of our central system, and
an auxiliary environment~\cite{wishartoriginal}. Namely, we take
\begin{equation}
\rho = \tr_{\mathcal{H}_{\textrm{env} }} |\psi\>\<\psi|,\, \quad
|\psi\>\in \mathcal{H}
\label{eq:statelarger}
\end{equation}
with
\begin{equation}
\mathcal{H}= \mathcal{H}_{\textrm{env} } \otimes \mathcal{H}_{\textrm{cen } }{},
\label{eq:hilbert:space:structure}
\end{equation}
and set $n=\dim \mathcal{H}_{\textrm{cen } }{}$, and $m= \dim \mathcal{H}_{\textrm{env} }$.
Notice that this {\it ensemble} of density matrices coincides with a Wishart
ensemble, normalize to unit trace.
The density matrix $\rho$ has
eigenvalues $\lambda_i$, $i=1,\dots,n$ that (i) are connected by the
normalization condition
\begin{equation}
\sum_i \lambda_i=1,
\label{eq:normalization}
\end{equation}
and (ii) are required to be non-negative:
\begin{equation}
\lambda_i \ge 0, \quad \forall i.
\label{eq:semipositive}
\end{equation}
Within this ensemble, we shall consider manifolds of codimension one, resulting
from fixing a quantity that depends only on the eigenvalues of the reduced
density matrix. This is inspired by thinking about a system with a {\it fixed
degree of decoherence}, as measured by purity.
Purity is defined as
\begin{equation}
P=\tr \rho^2
= \sum_{i} \lambda_i^2.
\label{eq:defpurity}
\end{equation}
At this point it must be remembered that purity is just one of
infinitely many convex functions that can be used to characterize
decoherence; its main advantage is its simple analytic
structure. We shall later also consider the von Neumann entropy due to its
information theoretical meaning and general popularity. It is good to
remember that additivity, which is the main advantage of entropy among convex
functions, seems meaningless in the context of entanglement. Handling of the
full set of convex function and a use of the partial order implied, seems
unrealistically complicated despite of the availability of results for all
R\'enyi entropies~\cite{Nadal2011}. Following~\cite{Nadal2011} one finds that
the distribution of the eigenvalues of the reduced state $\rho$ is given by
\begin{equation}
\mathcal{P}(\vec \lambda) \propto
\delta\left( \sum_i \lambda_i -1 \right)
\delta\left( \sum_i \lambda_i^2 -P \right)
\prod_i \lambda_i^{|m-n|}
\prod_{i<j} (\lambda_i-\lambda_j)^2
\label{eq:distr}
\end{equation}
where a term that accounts for the restriction to fixed purity is included.
Up to unitary transformations in the central system,
there are $n$ parameters characterizing our state
(the real eigenvalues of the $n$-dimensional
density matrix $\rho$), constrained with a physicality
condition (trace equal to 1) and the additional fixed purity constrain, both
conditions being scalar. Thus a $n-2$ dimensional space of free parameters is
left. We want to do a mapping from the eigenvalues of a density matrix to a
meaningful and minimal space which can be plotted and thus have a deeper
understanding of the ideas to be developed. A simple idea is to choose the
first $n-2$ eigenvalues, and let the other two be determined by restrictions
\eref{eq:normalization} and \eref{eq:defpurity}. Given that $s_1 =
\sum_{i=1}^{n-2}\lambda_i$ and $s_2 = \sum_{i=1}^{n-2}\lambda_i^2$,
the other two eigenvalues are
\begin{equation}
\lambda_{n-1},\lambda_n = \frac{1-s_1 \pm \sqrt{2(P-s_2)-(1-s_1)^2}}{2}.
\label{eq:other:two}
\end{equation}
A term that accounts for the Jacobian of the transformation is needed to correctly
transform the probability densities.
The transformation will take
$\lambda_1,\dots,\lambda_n$ to
$\lambda_1,\dots,\lambda_{n-2}$, $\sum_{i=1}^{n}\lambda_i^2$, and
$\sum_{i=1}^{n}\lambda_i$.
That said, the determinant of the Jacobian has only a nontrivial contribution, given
by a $2\times2$ block:
\begin{equation}
J=\begin{vmatrix} 1& 1 \\-2\lambda_{n-1} & - 2\lambda_{n}\end{vmatrix}=2 \sqrt{2(P-s_2)-(1-s_1)^2}.
\end{equation}
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth]{examples2_l1l2Final}
\end{center}
\caption{\label{fig:example:smalldims} (Color online) Probability distribution
of 2 eigenvalues for $n=4$, at fixed purity $P=0.4$ for $m=16$ (main figure) and $m=4$ (inset), in
arbitrary units. We have supplied two color scales. On green/yellow, the
maximum correspond to the maximum when the eigenvalues are ordered. On
yellow/red when we allow arbitrary order. This allows to observe better the details.
The white region leads to nonphysical eigenvalues. For $m=4$, the probability
tends to the axes, whereas for larger $m$s, it has zero probability density
there. }
\end{figure}
We can visualize the distribution for $n=4$ in
\fref{fig:example:smalldims}.
This distribution has a peak at $n-1$ degenerate states,
modulated by a repulsion of levels, which, altogether creates
a ``cleaved peak''. Notice that at $m=n$ there is a qualitative
change of behavior, and a repulsion from the planes $\lambda_i=0$ is not observed,
since the term $\prod_i \lambda_i^{|m-n|}$ yields a constant.
Level repulsion, however, is always present.
In order to gain some insight into the behavior of the eigenvalue distribution,
and to make an approximation for large environments,
let us first rewrite~\eref{eq:distr} as
\begin{equation*}
\mathcal{P} = C_{m,P} J g^{|m-n|}(\lambda) V(\lambda),
\end{equation*}
with $\lambda$ being the vector that groups all $\lambda_i$ and
$C_{m,P}$ a normalization constant. The function
\begin{equation}
g(\lambda) = g(\lambda_{1,\dots,n-2},P) = \prod_{i=1}^n \lambda_i
\label{eq:g}
\end{equation}
is the product of all eigenvalues, but can be regarded as a function of the
first $n-2$ eigenvalues and the purity.
Finally,
\begin{equation}
V(\lambda) = V(\lambda_{1,\dots,n-2},P) =
\prod_{i<j} (\lambda_i-\lambda_j)^2
\label{eq:vandermonde}
\end{equation}
is a Vandermonde determinant. These two terms, $g$ and $V$,
shall be analyzed separately.
We will see that $g(\lambda)$ is responsible for a peak around a
$n-1$ degenerate state, whereas the determinant is responsible for the nodes
that appear modulating such peak. \par
\paragraph{The peak}
We find explicitly that
\begin{equation}
g(\lambda) =
\frac{
s_2
+\left( s_1 \right)^2
-2 s_1
+1-P
}{2}
\prod_{i=1}^{n-2} \lambda_i .
\label{eq:defgtwo}
\end{equation}
To obtain the extrema of the function, the partial derivatives are
set to zero. Assuming $\lambda_i >0$ and rearranging, we obtain from
the subtraction of the equations obtained from the partial derivative
of $g(\lambda)$ with
respect to $\lambda_j$ and $\lambda_k$:
\begin{equation}
(\lambda_j - \lambda_k)
\left(\lambda_j + \lambda_k + s_1 -1\right)=0
\label{eq:condition:one}
\end{equation}
and from the sum of all partial derivatives,
\begin{equation}
s_2
+\left( s_1 \right)^
+2\frac{1-n}{n} s_1
+\frac{n-2}{n}(1-P)
=0.
\label{eq:condition:two}
\end{equation}
In general, these equations have an exponentially large number of solutions, not all of them
physical, but a detailed analysis
turns out to be cumbersome. We first focus on the particular case
in which all $\lambda_i=\Lambda$, $i=1,\dots,n-2$, which clearly
leads to a solution of all eqs.~(\ref{eq:condition:one}). From
\eref{eq:condition:two}, and solving a simple quadratic equation,
one obtains two solutions,
\begin{equation}
\Lambda_{\pm} = \frac{1}{n} \pm \frac{1}{n}\sqrt{\frac{nP-1}{n-1}}.
\label{eq:lambdapm}
\end{equation}
Each of this highly degenerate eigenvalues determines the other two eigenvalues,
according to \eref{eq:other:two}. These
are $\lambda_{n-1}=\Lambda_{\pm}$ and
$\lambda_{n}=1-(n-1)\Lambda_{\pm}$. This corresponds to a state that is a mixture
of a pure state and the maximally mixed state. Notice that for
$P >1/(n-1)$, $\Lambda_+$ results in a negative (and thus inadmissible)
$\lambda_{n}$.
We can simplify the expression for the probability density
near the maximum corresponding to the point
$\lambda_i = \Lambda_-$, $i=1,\dots,n-1$. We expand in Taylor
series around the maximum, and note that
\begin{align}
\pder{^2 g}{\lambda_j^2}{\biggr |}_{\lambda_i=\Lambda_-}
= 2 \pder{^2 g}{\lambda_j \lambda_{k\ne j}}{\biggr |}_{\lambda_i=\Lambda_-}
= 2 \Lambda_-^{n-2}\left( n-\frac{1}{\Lambda_-} \right),
\label{eq:partial:derivatives}
\end{align}
with $j=1,\dots,n-2$. This leads to
\begin{equation}
g(\lambda_{1,\dots,n-2}, P) \approx
g_\text{max} - \alpha
\tilde \lambda
\cdot A \cdot
\tilde \lambda^T
\label{eq:approxpoly}
\end{equation}
with
$$A=
\begin{pmatrix}
2 & 1 & \dots & 1 \\
1 & 2 & \dots & 1 \\
\vdots & \vdots & \ddots & \vdots \\
1 & 1 & \dots & 2
\end{pmatrix},$$
$\alpha=
\Lambda_-^{n-2}\left( n-1/\Lambda_- \right)$,
$\tilde\lambda = \begin{pmatrix} \lambda_1-\Lambda_- & \cdots &
\lambda_{n-2}-\Lambda_- \end{pmatrix}$
and
$g_\text{max}=g|_{\lambda_i=\Lambda_-}$.
Recall that
$\rme^{-x} \approx 1-x$ for small values of $x$, so that, near a maximum,
one can approximate a polynomial with a Gaussian:
$ (1-x^2)^m \approx \rme ^{-mx^2}$
for $m\gg 1$ and $-1 < x < 1$.
It is then found that
\begin{equation}
g(\lambda_{1,\dots,n-2}, P)^{|m-n|} \approx
g_\text{max}^{|m-n|}
\exp \left(
-\frac{\tilde\lambda^T \cdot A \cdot \tilde\lambda}{\sigma_{m,P}^2}\right)
\label{eq:approximation:g}
\end{equation}
with the average standard deviation given by
with the factor
$\sigma_{m,P}^2=g_\text{max}/\alpha(m-n)$
determining the standard deviation of each of the eigenvalues.\par
\paragraph{The Vandermonde determinant}
We will study the situations in which $V(\lambda)=0$, as this will give information
about the nodes of the probability density that will help understand its behavior.
This happens when at least one of the terms $\lambda_i - \lambda_j=0$ ($i\ne j$).
In the space of the first $n-2$ eigenvalues, the conditions
\begin{align}
\lambda_1=\lambda_2,\quad & \lambda_1=\lambda_3, & \dots, \quad &\lambda_1=\lambda_{n-2} \nonumber\\
& \lambda_2=\lambda_3, &\dots, \quad &\lambda_2=\lambda_{n-2} \nonumber\\
& & \ddots & \nonumber\\
& & & \lambda_{n-3}=\lambda_{n-2}
\end{align}
are simple hyperplanes. On the other hand, replacing
the conditions $\lambda_j = \lambda_{n-1}$
and $\lambda_j=\lambda_n$ on
\eref{eq:other:two} and squaring the discriminant
for $j=1,\dots,n-2$, define the $n-2$ different quadratic
forms
\begin{equation}
Q_j(\lambda_i) = \left( 2\lambda_j + s_1 -1\right)^2 - 2(P-s_2)+ (1-s_1)^2=0.
\label{eq:all:ellipsoids}
\end{equation}
All these curves are ellipsoids, as can be noted by the fact that all
$\lambda$'s are bounded in these degree 2 polynomials.
The condition $\lambda_{n-1}-\lambda_n=0$
can be calculated directly form \eref{eq:other:two}, and
leads to the ellipsoid
\begin{equation}
Q_n = 2(P-s_2)-(1-s_1)^2 = 0,
\label{eq:physicalellipsoid}
\end{equation}
which marks the limits for which the expression~\eref{eq:other:two} lead to
real, rather than complex and thus inadmissible, values for $\lambda_{n-1,n}$.
Finally, the condition $\lambda_{n-1}=0$ constitutes another constraint, which
can also be calculated from~\eref{eq:other:two}, and one obtains
\begin{equation}
Q_{n-1} = (P-s_2)-(1-s_1)^2 = 0.
\label{eq:physicalellipsoid:other}
\end{equation}
Notice that the condition $\lambda_{n}=0$ is automatically taken care of, as
from \eref{eq:other:two}, $\lambda_n \ge \lambda_{n-1}$.
The ellipsoids are really symmetry planes, that acquire this skew form since
the last two eigenvalues are given a special status. One could focus the study
to one region defined by any side of each of the ellipsoids \eref{eq:all:ellipsoids}.
At the peak, there is an $n-1$ fold degeneracy of the eigenvalues.
The right hand side of~\eref{eq:vandermonde} may be approximated, considering
only linear terms with respect to $\lambda_j -\Lambda_-$. We obtain
from the linear approximation $q_j(\lambda_i)$ of $Q_j(\lambda_i)$
\begin{align}
q_j(\lambda_i) &=
(\lambda_j-\Lambda_-) \pder{Q_j}{\lambda_j}{\biggr |}_{\lambda_i=\Lambda_-} \nonumber
+\sum_{k\ne j} (\lambda_k-\Lambda_-) \pder{Q_j}{\lambda_k}{\biggr |}_{\lambda_i=\Lambda_-} \\
&=4(n\Lambda_--1)[\lambda_j+s_1-(n-1)\Lambda_-],
\label{eq:nodes:as:planes}
\end{align}
a series of $n-2$ hyperplanes. This approximation will be useful when combined with \eref{eq:approximation:g}
to yield a simple expression near the degeneracy.
\begin{figure}
\begin{center}
\includegraphics[scale=1.3]{gl1l2p4Final}
\end{center}
\caption{\label{fig:manchatwoD} (Color online) Boundaries for the possible
physical states, and visualization of the Vandermonde determinant in the
$\lambda_1$, $\lambda_2$ plane, for $P=0.4$ and $n=4$. The regions where the physical
density matrices live is above the horizontal [red] line ($\lambda_2\ge 0$), to
the right of the vertical [red] line ($\lambda_1\ge 0$), inside the outer
[green] ellipse [real values for $\lambda_3$ and $\lambda_4$,
eqs.~\eref{eq:physicalellipsoid} and \eref{eq:qthreefour}] and outside the
smaller [red] ellipse [$\lambda_3\ge0$, eqs.~\eref{eq:physicalellipsoid:other}
and \eref{eq:qfour}]. The geometric place of the
$\lambda_1=\lambda_2$ degeneracy is the diagonal line [blue], of the
$\lambda_1=\lambda_3$ and $\lambda_1=\lambda_4$ degeneracy is the most vertical
[blue] ellipse and the $\lambda_2=\lambda_3$ and $\lambda_2=\lambda_4$
degeneracy is the most horizontal [blue] ellipse, see
eqs.~\eref{eq:all:ellipsoids} and \eref{eq:qotf}. } \label{fig:symmetries}
\end{figure}
\subsection{The special case of $n=4$}
\label{sec:four:fixed}
The case $n=4$ deserves special attention, as the numerics are carried on there, and
a complete plot of the full distribution is possible.
The explicit form of $g$ and $\Lambda_-$ can be read directly from
eqs.~(\ref{eq:defgtwo}) and (\ref{eq:lambdapm}).
The analysis of the conditions of degeneracy results more interesting.
In this case, the ellipsoid
corresponding to $\lambda_1=\lambda_3$ and $\lambda_1=\lambda_4$ is
\begin{equation}
Q_1(\lambda_1,\lambda_2)= 3\lambda_1^2+\lambda_2^2+2\lambda_1\lambda_2-2\lambda_1-\lambda_2+\frac{1-P}{2}.
\label{eq:qotf}
\end{equation}
This is an ellipse centered in $(1/4, 1/4)$ rotated an angle $\theta$ such that
$\cot \theta=1+\sqrt2$ and with semi-axes of length $\sqrt{4P-1}/4\sqrt{1\pm2^{-1/2}}$.
$Q_2$ will be the same up to an interchange of
$\lambda_{1}$ and $\lambda_2$.
The border curve corresponding to $\lambda_3=\lambda_4$
leads to the polynomial
\begin{equation}
Q_4(\lambda_1, \lambda_2)=3\lambda_1^2+ 3\lambda_2^2+ 2\lambda_1\lambda_2-2\lambda_1-2\lambda_2-2P+1.
\label{eq:qthreefour}
\end{equation}
This defines an ellipse with center in $(1/4, 1/4)$, rotated by $\pi/4$
and with semiaxis $(\sqrt{4P-1}/3, \sqrt{(4P-1)/3})$.
Semi-definite positivity of the density operator is guarantied
if one restricts to the area delimited by $\lambda_i \ge 0$, which amounts to
consider the quadrant $\lambda_{1, 2}\ge 0$, and $\lambda_3 \ge 0$
as, by construction, $\lambda_4\ge\lambda_3$. The condition
$\lambda_3 \ge 0$ is met in the exterior of the ellipsoid defined by
$Q_3(\lambda_1, \lambda_2)=0$ with
\begin{equation}
Q_3(x, y)=2x^2+ 2y^2 +2xy-2x-2y+1-P.
\label{eq:qfour}
\end{equation}
This is an ellipsoid centered in $(1/3, 1/3)$, rotated $\pi/4$ and with semiaxis
$(\sqrt{3P-1}/3, \sqrt{(3P-1)/3})$. For $P<1/3$ this ellipse does not exist; as long
as the previous conditions are met, $\lambda_4$ will be real.
Using approximation \eref{eq:nodes:as:planes} the following simplified distribution is
obtained
\begin{equation}
\mathcal{P}(\lambda_1, \lambda_2) = C_P
\exp\left[
-\frac{\tilde\lambda^T \cdot A \cdot \tilde\lambda}{\sigma_{m,P}^2}\right]
\left[
(\lambda_1-\lambda_2)(2\lambda_1+\lambda_2-3\Lambda_-)(\lambda_1+2\lambda_2-3\Lambda_-) \right]^2,
\label{eq:approx}
\end{equation}
valid for large dimensions of the environment and sufficiently close to
$\lambda_{1,2,3} \approx \Lambda_-$.
It is now simple to read the behavior of the function. It is indeed a
6-fold peaked function arising from a Gaussian of width diminishing
as $1/\sqrt{m}$, modulated by some parabolas touching zero. In this
approximation the center of the Gaussian is
located in the point $\vec \Lambda_-=(\Lambda_-, \Lambda_-)$ with
$\Lambda_-=(3-\sqrt{12P-3})/12$; a point corresponding to a triply
degenerate state. The other peaks are in $\vec \Lambda_- +
\sigma_{m,P} \vec v_i$ where $ \vec v_{1,\ldots,6} \in \{(\pm \sqrt3, 0),
(0,\pm \sqrt3, 0), \sqrt2(\pm1, \mp1) \}$. The last thing that should be
calculated are single variable distributions,
which are presented in appendix~\ref{sec:distribution:smallest:eigenvalues}.
A glimpse at the complexity of the surface can be obtain analyzing the
stationary points of the surface. In this case, \eref{eq:condition:one} and
\eref{eq:condition:two} lead to four solutions. The first one corresponds to
$\lambda_1 = \lambda_2 = \Lambda_-$. The other two eigenvalues are $
\Lambda_-$ and $1-3\Lambda_-$. A second solution, corresponding to triply
degenerate $\Lambda_+$ remains physical for $P\le 1/2$, after which the points
migrate to the unphysical region corresponding to the interior of the red
ellipse in \fref{fig:manchatwoD}. The other two solutions correspond to saddle
points of the function $g$, but for $P > 1/2$ also correspond to unphysical
eigenvalues.
There are additional maximums that are not detected by conditions
\eref{eq:condition:one} and \eref{eq:condition:two}, which live in the edge of
the domain, and thus would have to be calculated separately. These also
correspond to the same triply degenerate states with $ \Lambda_-$, but
reordered and are located at the points in which the blue and green ellipses
in \fref{fig:manchatwoD} touch each other.
All these boundaries are plotted in \fref{fig:manchatwoD}, where one can visualize
the roll of each of the conditions earlier discussed. The global effect can
also be nicely seen in \fref{fig:example:smalldims}.
\section{Random states and random dynamics} \label{sec:rmtdynamical}
We now wish to compare the results obtained for the static model discussed in
the previous section with
those of previously studied dynamical random matrix
models, in which decoherence of a central system can be
studied~\cite{1464-4266-4-4-325, Gorin2003, PhysRevLett.107.080404}. These
models are time independent,
but will be used to evolve an initial state until a
time in which the purity (or any other quantity of interest) reaches
the particular value of interest, thus obtaining an equivalent ensemble
of density matrices.
That said, consider again a Hilbert space with the structure
${\mathcal H}\ =\ \mathcal{H}_{\textrm{cen } } \otimes \mathcal{H}_{\textrm{env} } $ [recall
\eref{eq:hilbert:space:structure}]. As in sec.~\ref{sec:four:fixed}, we also
restrict to the case in which $\mathcal{H}_{\textrm{cen } }$ is a 4-level system.
Special consideration will be given to the case
in which $\mathcal{H}_{\textrm{cen } }$ is composed of two
qubits; that is, when
\begin{equation}
\mathcal{H}_{\textrm{cen } } = {\mathcal H}_{{\textrm{q} }_1} \otimes {\mathcal H}_{{\textrm{q} }_2}.
\label{eq:two:qubits}
\end{equation}
where ${\mathcal H}_{{\textrm{q} }_i} $ denote
qubit spaces.
In this space, unitary dynamics will be generated by random Hamiltonians,
to be detailed later. Non-unitary dynamics
in the two qubit central system is thus induced by the unitary dynamics
in the whole space, plus partial tracing over $\mathcal{H}_{\textrm{cen } }$.
As initial states, pure separable states in the whole Hilbert
space are used:
\begin{equation}
\label{eq:initialconditiongeneral}
|\psi(0)\>=|\psi_{\textrm{cen } }\>|\psi_{\textrm{env} }\>,\quad
|\psi_{\textrm{cen } }\>\in
\mathcal{H}_{\textrm{cen } },\,|\psi_{\textrm{env} }\>\in \mathcal{H}_{\textrm{env} }; .
\end{equation}
We take $|\psi_{\textrm{env} }\>$ to be a random state, according to the Haar measure of
the unitary group in the environment, but $|\psi_{\textrm{cen } }\>$ is to be chosen
according to the particular case under
study. \par
Kolmogorov distance will be used to quantify the difference between two
distributions. Given two functions, $f(x)$ and $g(x)$, defined on
a domain $X$, it is
\begin{equation}
K(f, g) = \frac12 \int_X |f(x)-g(x)| \rmd x.
\label{eq:kolmogorov}
\end{equation}
Notice that since we are dealing with distributions, the normalization condition
for a distribution $\mathcal{P}(x)$ is stated as $K(\mathcal{P}, 0)=\frac12$, and thus the distance
between two non overlapping (i.e. totally distinct) distributions $\mathcal{P}_1$ and
$\mathcal{P}_2$ is always 1. \par
\subsection{Global Hamiltonian}
\begin{figure}
\begin{center}
\includegraphics[scale=1.3]{globalFinal}
\end{center}
\caption{\label{fig:global:hamiltonian}
(Color online)
Marginal distribution of the smallest two eigenvalues ($\lambda_1 \le \lambda_2$).
The points are obtained with nonunitary evolution
under the global Hamiltonian~\eref{eq:global:hamiltonian} with close to $10^6$
realizations, and the curves are obtained with the marginal distributions
\eref{eq:lambdaone} and \eref{eq:lambdatwo}, obtained by proper integration of
\eref{eq:distr}. We show different target purities and different environment
dimensions. }
\end{figure}
The simplest candidate for a random matrix model that describes decoherence is
simply choosing a Hamiltonian from one of the classical
ensembles~\cite{cartanRMT} that acts on the whole Hilbert space, i.e. on both
the central system and the environment~\cite{Gorin2003}. We shall call this
family {\it global Hamiltonian}. These Hamiltonians are attractive because of
their analytic simplicity~\cite{PhysRevLett.107.080404}. They also model the
strongest interaction between central system and environment. We shall choose
to pick the Hamiltonian from the GUE (Gaussian unitary ensemble) as it is the
simplest ensemble, from an analytical point of view.
For this case, we can write simply
\begin{equation}
H = H^{(\text{GUE})}_{{\textrm{env} },{\textrm{cen } }}
\label{eq:global:hamiltonian}
\end{equation}
where the superindex indicates the ensemble from which the operator is chosen, and the subindices
indicate the spaces in which they act non-trivially. For this case, and due to
the invariance of the GUE, one can choose the initial state of the central
system arbitrarily, with no effect on the results regarding
the eigenvalue density of the evolved state of the central system.
The eigenvalue density of the density matrices produced by the non-unitary
dynamics induced by the Hamiltonian \eref{eq:global:hamiltonian} are similar
to those for the static situation, \eref{eq:distr}. However,
differences can be observed by inspection, see \fref{fig:global:hamiltonian}.
These differences are larger for smaller purities, and seem to be independent
of the size of the environment. However, even for an intermediate purity of
$P=0.8$, it is difficult to see any differences.
If one is not interested in very precise data, or only in high purity,
quantitative information can be extracted from the static model regarding
the evolved state.
\begin{figure}
\begin{center}
\includegraphics[scale=1.3]{couplingFinal}
\end{center}
\caption{\label{fig:coupling:hamiltonian}
Marginal distribution of the smallest eigenvalue $\lambda_1$.
The points are obtained with nonunitary evolution
under the tunable coupling Hamiltonian~\eref{eq:weak:four:hamiltonian} with close to $10^5$
realizations, and the curves are obtained with the marginal distributions
\eref{eq:lambdaone} and \eref{eq:lambdatwo}, obtained by proper integration of
\eref{eq:distr}. We show two different target purities, different environment
dimensions and different couplings. For very small couplings, the ensemble
is not well approximated by \eref{eq:distr}, but already for moderate couplings,
the ansatz is very good.}
\end{figure}
\subsection{Tunable coupling
Hamiltonian \eref{eq:global:hamiltonian} couples the central system and
its environment. However it does not take into account the structure of
Hilbert space~\eref{eq:hilbert:space:structure}. One way to do that is
to provide the Hamiltonian governing the system with a similar
tensor product structure. Moreover this will lead the idea of tunable
coupling which is very convenient when studying open quantum systems.
We thus use
\begin{equation}
H = H^{(\text{GUE})}_{{\textrm{env} }} + \epsilon V^{(\text{GUE})}_{{\textrm{env} },{\textrm{cen } }},
\label{eq:weak:four:hamiltonian}
\end{equation}
which will be called {\it tunable coupling Hamiltonian}, or
coupling Hamiltonian, for short.
It must be noticed that the case
$\epsilon\to\infty$ corresponds, up to a rescaling in time, to the previous case,
i.e. \eref{eq:global:hamiltonian}.
Regarding the eigenvalue density of the evolved state of the central
system with
Hamiltonian \eref{eq:weak:four:hamiltonian}, a significant
dependency on the coupling strength was observed.
Four regimes should be considered: weak coupling [for which
$\epsilon = 0.01$ is taken as a representantive], intermediate coupling
[$\epsilon = 0.1$], strong coupling [$\epsilon = 1$], and
very strong coupling [$\epsilon \to \infty$].
\par
For weak coupling there is a clear discrepancy, for all purities and dimensions
examined. The agreement does not seem to improve for larger dimensions,
see \fref{fig:coupling:hamiltonian}. It must be noticed that even though there
is a significant disagreement between the two ensembles, the nodes can still be
observed, and for a qualitative description, the static ensemble is still
useful.
\par
For intermediate coupling $\epsilon=0.1$, the agreement is very good, and only
for smaller dimensions and purities can the difference be observed by
inspection, see \fref{fig:coupling:hamiltonian}. Here, it seems that both,
increasing the purity at fixed environment dimension, and increasing the
dimension at fixed purity leads to the same ensemble as the static one. For
strong coupling and very strong coupling, the results are indistinguishable
from the global Hamiltonian case, see \fref{fig:kolmogorov}.
\par
\subsection{Spectator Hamiltonian}
An interesting variation of \eref{eq:weak:four:hamiltonian} has
been studied in~\cite{pinedaRMTshort}. There, one studies a central system
whose Hilbert space is composed of two qubits. That is
when the Hilbert space is of the form \eref{eq:two:qubits}.
There, the
{\it spectator Hamiltonian}
\begin{equation}
\label{eq:hamiltonianspectator}
H= H^{(\text{GUE})}_{\textrm{env} } +\epsilon V^{(\text{GUE})}_{{\textrm{env} },{\textrm{q} }_1}
\end{equation}
was proposed where, again, the subindices indicate the part of the Hilbert
space where they act non trivially.
The first term correspond to the free dynamics of the environment. The next
represents the coupling of the first qubit to the
environment. Thus, the second qubit
is simply a \textit{spectator}, as it has no coupling
to an environment.
This is the
\textit{simplest} Hamiltonian for which we can analyze the effect of an
environment on a Bell pair \cite{pinedaRMTshort}.
The environment Hamiltonian $H_{\textrm{env} }$ will be
chosen from a classical ensemble~\cite{cartanRMT} of $m \times m$ matrices and
the coupling $V_{{\textrm{env} },{\textrm{q} }_1}$ from one of $2m \times 2m$ matrices.
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth]{entangledFinal}
\end{center}
\caption{Distribution of the two smallest eigenvalues for the spectator and the common
environment Hamiltonians eqs.~(\ref{eq:hamiltonianspectator}),
(\ref{eq:hamiltonian:common}),
for various degrees of entanglement of the initial condition in
the central system, modulated by
$\theta$ in \eref{eq:general:two:qubit}. The target purity
is, in this case, $P=0.8$ and we fix $m=8$. For the spectator case, low
entanglement tends to push the distribution towards the origin, and for
the common environment, small entanglement tends to push the distribution
towards the $y$ axis. Both models are in good agreement with the
ansatz distribution for a maximally entangled initial state. The figure
shows a disymmetrized distribution, assuming non-decreasing eigenvalues, and
the scale is arbitrary.
}
\label{fig:varyent}
\end{figure}
This Hamiltonian is quite interesting, since it does not involve one of the
qubits. At least not from the dynamical point of view. This Hamiltonian adds a
structure to the problem: it creates the notion of a particle. One obvious new
ingredient when the Hilbert space is split is entanglement. In this case we
shall thus have a new parameter which is the entanglement of the initial state
in the two qubits.
\begin{figure*}[ht]
\begin{center}
\includegraphics[width=\textwidth]{kolmogorovFinal}
\end{center}
\caption{We present a summary of the comparison between the static ensemble,
and the ensemble generated by random dynamics (left).
In the ordinate $\log_2 m$ with only take integer values. The symbols representing the different
cases will appear for arbitrary values within the interval around the fixed integer value
of $\log_2 m$ to make then discernible. On the abscissa the Kolmogorov distance between
the dynamical model and the corresponding static one $K(\mathcal P, \cdot)$ is plotted.
Good agreement between
both is observed for the global Hamiltonian,
the coupling Hamiltonian for large enough coupling, and for the spectator and common environment Hamiltonian
when the initial state has maximal entanglement. The symbols must be understood using the
table on the right hand side. The error bars are obtained with the Kolmogorov distance with
respect to a Monte Carlo simulation with the same number of points as the one with
the random dynamics, that is, $10^6$ points. }
\label{fig:kolmogorov}
\end{figure*}
For a totally disentangled initial state, the dynamics will produce in the
two qubits a reduced state of the form
\begin{equation}
\rho(t) = \rho_1(t) \otimes |\psi\> \< \psi|
\label{eq:separable:two:qubit}
\end{equation}
at all times. Notice that the smaller eigenvalues $0$ are double degenerate.
The other two eigenvalues are totally determined by the normalization condition
and the fixed purity condition.
Taking into account the invariance of the ensemble defined by
\eref{eq:hamiltonianspectator} under local unitary operations,
we notice that the initial states can be written with
absolute generality as
\begin{equation}
|\psi(0) \> = \sin \theta |00\> + \cos\theta |11\>,
\quad \rho(0) = |\psi(0) \> \<\psi(0) |,
\label{eq:general:two:qubit}
\end{equation}
where $\theta$ is the only relevant parameter of the state, and
also modulates the initial entanglement. $\theta=0$ corresponds
to an initially separable state, whereas $\theta=\pi/4$ to a
maximally entangled one.
If $\theta=0.2$ the situation is indeed close to the totally separable case.
In this case the distribution is totally different than
expected from the static analysis, that is from \eref{eq:distr}, see
\fref{fig:varyent}. However for a maximally entangled state $\theta = \pi/4$,
the results are similar as to the tunable coupling case with four levels. That
is, acceptable agreement for very small coupling, and much better agreement
for intermediate coupling. Again, the agreement improves
with increasing purity, but does not seems to converge for increasing dimension of
the environment at a fixed purity, see \fref{fig:kolmogorov}.
\par
\subsection{Common environment Hamiltonian}
Even though Hamiltonian (\ref{eq:hamiltonianspectator}) is the simplest one
that allows us to study entanglement evolution~\cite{pinedaRMTshort}, in many
situations one would have coupling of all constituents of the
central system to the environment. To be closer to many experimental situations,
consider what we call the {\it common environment} Hamiltonian
\begin{equation}
\label{eq:hamiltonian:common}
H= H^{(\text{GUE})}_{\textrm{env} }
+\epsilon V^{(\text{GUE})}_{{\textrm{env} },{\textrm{q} }_1}
+\epsilon V^{(\text{GUE})}_{{\textrm{env} },{\textrm{q} }_2}.
\end{equation}
Again, the indices indicate the subspaces in which they operate non-trivially,
with respect to \eref{eq:hilbert:space:structure} and \eref{eq:two:qubits}.
This Hamiltonian, includes coupling of both qubits to an environment, but also
takes into account a Hilbert space with the structure~\eref{eq:two:qubits}.
For the common environment Hamiltonian we have again a particular form of
the eigenvalues for an initially separable state. In this case, one can
approximate the dynamics as two channels acting independently on the two
qubits. Thus, neglecting correlations in the environment, for later times one
should have that
\begin{equation}
\rho(t) \approx \rho_1(t) \otimes \rho_2(t).
\label{eq:commong:two:qubit}
\end{equation}
One then has that the eigenvalues of the reduced density matrix are $(\lambda,
1-\lambda)\otimes (\lambda', 1-\lambda')$, which
gives an additional global restriction; the manifold in which the eigenvalues
lives has dimension one, instead of two as in the previous cases.
This will of course heavily influence the distributions of
eigenvalues. For a small amount of initial entanglement the situation should vary
continuously, and one indeed observes very significant differences with the
proposed ansatz, see \fref{fig:kolmogorov}. For an initially maximally entangled state,
the results are again similar and are well reproduced by the static ansatz, see
figs.~(\ref{fig:varyent}) and (\ref{fig:kolmogorov}).
\subsection{A comparison using Kolmogorov distance}
The results of this section can be summarized in a figure containing the
Kolmogorov distance between the results for the dynamical system, and the
statical ensemble, \eref{eq:distr}. We
present those results in \fref{fig:kolmogorov}.
Error bars, obtained using the Kolmogorov distance between the exact
distribution, and an equivalent Monte Carlo simulation [see
Appendix~\ref{sec:montecarlo}] with the same number of data points as the
dynamical situation, is also included. This number should be interpreted
as the error arising both from finite sampling and finite binning.
We can
see that the results for the static model and the dynamical models agree well
for the global Hamiltonian, and for the coupling Hamiltonian, as long as the
coupling is not too small. If the coupling is very small, there are
quantitative deviations, but the shape of the distribution remains similar. In
this case, one can see that there is a clear difference between the static and
the dynamical ensemble, which leads to intermediate values of $K$, very
different both from 0 and 1. As coupling becomes very small, these deviations
increase. Often such cases are associated with situations where the dynamical
model will actually not reach equipartition~\cite{PhysRevA.90.022107,
pinedalong}.
For models in which structured coupling is present (spectator and
common environment models), one has similar results as for the tunable coupling
model as long as the initial entanglement is maximal. For other initial states
the disagreement is progressive as the dimension of the environment increases.
We conjecture that in the limit of large $m$ there will be a dynamical quantum
phase transition between good agreement and maximum disagreement, for
arbitrarily small deviations from a Bell state. \par
\section{Other functions: von Neumann entropy} \label{sec:entropy}
\begin{figure}
\begin{center}
\includegraphics[scale=1.4]{EntropiaVsPurezaFinal}
\end{center}
\caption{\label{fig:entropy:for:purity} (Color online) Probability
distribution of the entropy $S$ for several values of purity. Each plot shows a
particular purity and several dimensions of the environment. Notice the scale
in the horizontal axis to observe that the distributions are already quite
peaked for the dimensions shown, the effect being larger for increasing dimension. }
\end{figure}
We have also studied the von Neumann entropy (henceforth called entropy) of the reduced density
matrix
\begin{equation}
S(\rho) = -\tr \left( \rho \ln \rho \right).
\label{eq:vonneumann}
\end{equation}
The algebraic structure is considerably more complicated than for purity, and
thus such a detailed program as was presented in the previous sections is in general only
possible in terms of solutions of transcendental equations. However much can be
said using the fact that purity and entropy are closely related in typical
cases.
For the ensembles studied in \sref{sec:static}, consider the entropy.
Its distribution is plotted in \fref{fig:entropy:for:purity}, for several fixed
purities, and several dimensions of the environment.
One can see that even for small dimensions of the environment ($m=8$), the
width of the entropy distribution is already of order $10^{-2}$, and its value
decreases with increasing $m$. Thus, even though the ensembles for fixed
purity and fixed entropy are different (not even the support is the same), they
are similar.
As an example, the equivalent to figures \ref{fig:example:smalldims}
and \ref{fig:manchatwoD}, but for fixed entropy,
is shown in \fref{fig:example:entropy}.
We obtained numerically the corresponding limiting curves and nodes, and the
probability distribution corresponding to the static ensemble. The resemblance
to the case of purity is striking, and one may conclude that many of the
results obtained for a fixed value of purity must hold qualitatively for a
fixed value of the entropy.
Notice that we are considering just two functions of infinitely many that define
a partial order over the eigenvalue vector, which mathematically corresponds to
a probability distribution. This indicates that the entire theory developed by
Hardy, Littlewood, and
Polya~\cite{hardy1952inequalities} applies, as noted in
reference~\cite{ru1,ru2, Uhlmann}. In particular it indicates that the ordering becomes
unique both near the pure states and near the maximally mixed state, thus
formalizing the analogy between purity and, say, entropy when one approaches
any of these limits.
\begin{figure}
\begin{center}
\includegraphics[width=\columnwidth]{entropy2dFinal}
\end{center}
\caption{\label{fig:example:entropy} (Color online) Probability distribution
of 2 eigenvalues at fixed entropy $S=1.5$ for $m=8$, in arbitrary units. The
white region leads to nonphysical eigenvalues. Nodes and limiting curves are
also present here. The color coding is exactly the same as in figures \ref{fig:example:smalldims} and
\ref{fig:manchatwoD}. }
\end{figure}
\section{Conclusions}
In this paper we set out to give a random matrix model for density matrices of
open quantum system by imposing further restrictions on an ensemble
originally obtained by simply restricting an ensemble of random covariance
matrices to matrices with unit trace. In section two we give some general
remarks about the alternate possibilities of restricting random matrix
ensembles in analogy to techniques used in statistical mechanics, as well as on
the alternative of a proposing a random matrix description directly for the
object of interest, as compared to proposing a random set of Hamiltonians to
describe the dynamics and from that calculate the ensemble of that same object.
In the latter case the condition is imposed by following evolution until it is
fulfilled exactly (i.e. up to numerical exactitude).
The conditions can either be imposed in a ``strict way'' for each member of the
ensemble resulting in microcanonical ensembles or on average in the spirit of
Jayne's principle leading to canonical ensembles. In view of the original
construction \cite{Nadal2011} we opt for the microcanonical approach.
After the construction of the ensemble we concentrate on the distribution of
eigenvalues of the density matrices using fixed purity as the additional
restriction. The choice of purity allows us to obtain analytical results but we
also consider alternate convex functions and explicitly the von Neumann entropy
towards the end of the paper. Qualitatively very similar rests are obtained
To visualize our result we restrict our considerations to a four dimensional
central system. The two restrictions imposed allow to give density maps over
the two remaining dimensions and one clearly sees the signature of the Jacobi
determinants on an otherwise smooth distribution limited by the boundaries
imposed by the restrictions on trace and purity. Again, this picture remains
qualitatively unchanged if we replace purity by von Neumann entropy.
For this four dimensional central system we present the comparison with
the results obtained if we run the time evolution of an initially pure state of
the central system under the random matrix model for decoherence given in terms
of several families of random Hamiltonians for the full system (central system
+ environment). We find fairly good agreement if the fouling between
environment and the central
system is not too weak, i.e. if we are not in the ``perturbative regime''.
Surprisingly this holds true even if the four dimensional central system
consists of two quits but one of the two does not interact with either the
other or the environment, under the condition that the two qubits are maximally
entangled.
We may speculate as to why the static and dynamic models agree so well, except
for the case of partially entangled central systems with a structured coupling
to the environment. Consider POSETS (partially ordered sets or lattices) of
density matrices introduced by Uhlmann~\cite{Uhlmann} as well as by Ruch and Mead
\cite{ru1, ru2} independently.
This work is based on a partial order introduced by Muirhead~\cite{muirhead} and connected to
classical bi-stochastic evolution by Hardy, Littlewood and
Polya~\cite{hardy1952inequalities} (see also
\cite{ru1}). The classical evolution can be generalized to POSETS of pairs of
distribution~\cite{HLP1,HLP2} and such concepts seem to have a growing impact on recent
work using majorization in the context of open quantum systems (e.g
\cite{wilming2014weak, faist2015gibbs, lostaglio2015quantum, gour2015resource}).
In this context it is well known, that near distributions with a dominant
component (i.e. for almost pure states) all convex functions will yield
essentially the same order, or, with other words, near its extreme points
POSETS are almost completely ordered sets. Therefore a unitary dynamic acting
on a initially separable density matrix will have no choice but to follow
closely the time-independent partial order, which governs the static ensemble.
Indeed if we look at \fref{fig:global:hamiltonian} we see immediately that the
agreement is best near maximal purity.
A number of steps along these lines are possible for the future. One may
investigate if a
partial orders for pairs, as suggested above, can be extended to pairs of density matrices.
is feasible and weather it it brings any
advantages.
A note of caution must be added here, as the partial order for pairs of
distributions in \cite{HLP1, HLP2} has not been successfully generalized to density
matrices because the two members of the pair typically cannot be diagonalized
with the same transformation.
If and when this mathematical problem is overcome we may well find this to be
the key to a deeper understanding of our results.
Another more obvious next step is to calculate other quantities for
fixed decoherence as measured by purity or some other convex function. For the
two qubit case concurrence would be the obvious candidate, as it would allow to
calculate $CP$ diagrams (concurrence purity) without referring to time
evolution.
\ack
We thank Satya Majumdar for very stimulating discussions at the onset of this project,
and Francois Leyvraz for informative discussions near its culmination.
We thank Isaac Perez for help with the Monte Carlo simulations.
Support by the projects CONACyT 153190, 79613, 219993,
UNAM-PAPIIT IA101713, IN111015
and IG101113 are acknowledged.
\section*{References}
\bibliographystyle{unsrt}
|
\section{Introduction}
\label{sec:Intro}
Observed rotation curves along with a mass model are one of the principal tools to derive the gravitational potential and distribution of mass in galaxies. The Inferred dynamical mass and luminosity profile suggest the presence of a dominant dark matter component in galaxies, but a study of rotation curves can give us a more detailed picture of this dark matter component.
For long time ago there has been tension between predicted and observed distributions of dark matter \citep{F1994,M1994} and an increasing number of recent high quality observations suggest that, for the dark matter distribution, a core-like behaviour is preferred in the central regions of dwarf Spheroidals (dSphs) \citep{G2006,Battaglia2008,Walter2011} and Low surface brightness (LSB) galaxies \citep{O11,R13,dB01,K03,A12}.
A model that tries to explain the nature of dark matter should be able to reproduce this core-like behaviour in the central region of the dark matter haloes. Cored density profiles predicted by the model can be tested against rotation curves. LSB galaxies and the dSphs of the Milky Way are overwhelmingly dark matter dominated, and so provide natural testing grounds for assessing these predictions.
On the other hand, it is a unique prediction of the Cold Dark Matter (CDM) model that the centers of dark haloes are cusped.
The finding that the haloes of at least some of the dSphs are cored may be interpreted in one
of two ways. Either it provides evidence that baryonic processes such as star formation and supernovae feedback have modified the pristine dark matter cusp \citep{Read2005,Mashchenko2008,C11,Po12}. Or it provides evidence of a new kind of DM, different from the one proposed by the CDM model that holds out the possibility of resolving some of these issues.
About the first way, is still not conclusive that baryonic processes such as star formation and supernova feedback will be able to erase the cusp predicted by CDM-only simulations \citep{Pe12,K11b}, specially in LSB and dSphs galaxies. For this reason is still open the possibility of an alternative explanation for the DM nature that produces cores naturally.
A dark matter scenario that has received much attention is the scalar field dark matter (SFDM) model. The main idea is that the nature of the DM is completely determined by a fundamental scalar field \citep{Bohmer2007,G00,M01,M12}. In this dark matter and galaxy formation scenario the DM haloes are naturally cored.
\citet{R13} found good agreement with data of rotation curves for LSB galaxies using the minimum disk hypothesis (neglecting the baryonic component).
Because baryons trace the DM potential, the aim of this work is to go forward within the SFDM model by building a galaxy mass model for dark matter dominated galaxies, including a baryonic disk. And because the rotation velocity of the tracer (gas, stars) is not always exactly equal to the circular velocity of a test particle on a circular orbit we simulate the HI tracer evolving the hydrodynamic equations for a gas distribution embedded in our modeled potential in order to measure the rotation curve directly on the gas and contrast with data.
The paper is organized as follow. In section \ref{sec:LSB} we describe briefly the importance of LSB galaxies and HI as a tracer. In section \ref{sec:massmodel} we present the mass model. Section \ref{sec:Initialconditions} contains the techniques an initial conditions for the simulations. Finally we present the results of our numerical simulations and a discussion.
\section{LSB Galaxies and HI as a dark matter tracer}
\label{sec:LSB}
LSB galaxies share common characteristics: late-type, gas-rich, poor star formation rates (SFR), low metallicities, diffuse stellar disks and extended HI gas disks \citep{Bothun97,Bell2000}.
Another important feature of LSB galaxies is that the mass-to-luminosity ratio is usually higher than that of a normal spiral galaxy due to the low luminosity, and the dark matter fraction is much higher.
The extended HI gas disk in LSB galaxies, although optically difficult to detect, are easily detected in HI surveys and the HI rotation curves of LSB galaxies have received great attention, with shapes different from those of high surface brightness (HSB) galaxies. The rotation curves of HSB spiral galaxies rise fairly steeply to reach an approximately flat part, whereas for LSB galaxies the rotation curves rise more slowly and often still rising at the outermost measured point. In order to describe the data well this galaxies need a massive dark matter halo with a nearly constant density core \citep{K11a}.
With a dark matter component dominating the total mass and HI gas disks far more extended that the stellar disks, LSB galaxies provide a suitable scenario to test the prediction of our mass model trought the study of the rotation curves of a tracer.
\section{Mass model}
\label{sec:massmodel}
For modeling a LSB galaxy we use an analytic two-component galaxy model which is composed of an axisymmetric disk and a spherically symmetric dark matter halo.
The gas and stellar components are represented by a three dimensional Miyamoto-Nagai disk profile \citep{MN75} that is mathematically simple, with a closed expression for the potential
\begin{equation}
\phi_d(R,z) = \frac{-GM_d}{\sqrt{R^2+(a+\sqrt{z^2+b^2})^2}},
\end{equation}
where $M_d$ is the total mass of the component, and $a$ and $b$ are the radial and vertical scale-length, respectively. This potential has continuous derivatives that make it particularly suitable for numerical work.
For the dark matter halo we made our choice based in numerous studies of rotation curves of LSB galaxies that have found the data to be consistent with a halo having a nearly constant density core \citep{dB01,K08,dB10,K11a}, and in the scalar field dark matter model (SFDM) the halos of galaxies featured a constant density core that comes out naturally in the model instead of just been assumed to fit the data.
Aimed to resolved some discrepancies when just taking in to account the condensed system at temperature $T=0$, \citet{R13} consider an scenario in which the dark matter, an auto-interacting real scalar field, is embedded in a thermal bath at temperature $T$. At high temperatures in the early universe, the system interacts with the rest of the matter. Its temperature will decrease due to the expansion of the universe and eventually, when the temperature is sufficiently small, the scalar field decouples from the interaction with the rest of the matter and follows its own thermodynamic history.
The Newtonian limit provides a good description at galactic scales so there is an exact analytic solution in the Newtonian approximation of this system \citep{R13}, a density profile that accounts for SFDM halos in condensed state or halos in a combination of exited states, $\rho = \rho_0\sin^2(r/a_h)/(r/a_h)^2$, where $\rho_0$ and $a_h$ are fitting parameters.
So the potential of the DM distribution can be written as
\begin{equation}
\label{eq:DMpotential}
\phi_h(r)=\frac{GM_0}{f(r_0)}\left( \ln(r)+\frac{\sin(2r/a_h)}{2r/a_h}-Ci(2r/a_h)\right)
\end{equation}
where $M_0$ is the mass enclosed within a radius $r_0$, $f(r_0)$ is a constant that depends on $r_0$ and $Ci$ is the cosine integral function.
For comparison we also use a different dark matter gravitational potential, a NFW halo that results from CDM simulations and characterized by a density profile that rise steeply towards the center \citep{N10}
\begin{equation}
\label{eq:NFW}
\rho_{NFW}(r) = \frac{\rho_s}{(r/r_s)(1+r/rs)^2}
\end{equation}
with $\rho_s$ and $r_s$ as fitting parameters. The gravitational potential from this cuspy halo can be written
\begin{equation}
\label{eq:NFW}
\phi_h(r) = \frac{-GM_0}{g(r_0)r}ln(1+r/r_s)
\end{equation}
with $M_0$ the mass enclosed within a radius $r_0$ and $g(r_0)$ a constant that depends on $r_0$.
The total gravitational potential for our galaxy model is then
\begin{equation}
\label{eq:potential}
\phi(R,z) = \phi_d(R,z) + \phi_h(R,z) ,
\end{equation}
from which we can compute, as usual, the circular velocity of a test particle.
But because the rotation velocity of the gas, that acts as a potential tracer, is not always exactly equal to the circular velocity of a test particle, we ran some simulations evolving the hydrodynamic equations (we do not model star formation or feedback) for a gas distribution embedded in our mass model potential. With this we are able to measure the rotation curves directly on the gas an compare with the data.
For the data we chose a set of four bulgeless LSB galaxies \citep{McGaugh01,K11b,K08}, because not including a bulge in our mass model reduce the number of adjustable parameters and because for this galaxies the total mass, the stellar mass and the HI mass are known.
For this data catalogue spectra was obtained for a few galaxies in order to check for the presence of noncircular motions, which apparently are minimum in LSB galaxies, as suggested by the authors of the study. Inclinations are also taken into account and errors regarding this only affect the absolute scale of rotation velocities, and not the shape of the rotation curve.
\section{Initial conditions}
\label{sec:Initialconditions}
To measure and study the temporal evolution of the baryonic component under the gravitational potential of our mass model we set up the initial condition of each LSB galaxy as an isolated Miyamoto-Nagai gas disk with a density profile
\begin{equation}
\label{eq:densityprofile}
\rho_d = \frac{b^2M_{HI}}{4\pi}\frac{aR^2+(a+3\sqrt{z^2+b^2})(a+\sqrt{z^2+b^2})^2}{(R^2+(a+\sqrt{z^2+b^2})^2)^{5/2}(z^2+b^2)^{3/2}}.
\end{equation}
where $M_{HI}$ is the HI mass of the galaxy disk, $a$ and $b$ are the radial and vertical scale-length, respectively.
\subsection{Dynamical stability of a gas disk}
Setting up a full three-dimensional dynamically stable astrophysical disk configuration is necessary in order to ensure that the evolution seen in the simulation is not because of some relaxation or evolution towards a new equilibrium configuration driven by non-stable initial conditions. Rotational forces, gravity and velocity dispersions must be balanced in order for the disk does not immediately disperse, collapse or deviate away from the desired density profile. We attempt to construct such a system following the epicyclic approximation as follow.
First we set up the density profile (eq \ref{eq:densityprofile}) and compute the principal circular velocity. Then according to the local stability criterion \citep{Toomre64} we obtain the velocity dispersion in $R$ and $z$ necessary for the disk to be stable against axisymmetric perturbations and be supported by random motions and rotation.
The radial and azimuthal dispersions are given by
\begin{equation}
\label{eq:radialdispersion}
\sigma_R = \frac{3.358\Sigma Q}{K},
\end{equation}
\begin{equation}
\label{eq:radialdispersion}
\sigma_{\phi} = \frac{\sigma_R K}{2\Omega},
\end{equation}
with $\Sigma$ the mass surface density, $Q$ the stability parameter whose initial value is $Q=1.2$ over the entire disk, $\Omega = \sqrt{(\partial_R \phi)/R}$ the angular frequency, and $K$ the epicyclic frequency given by
\begin{equation}
\label{eq:epicyclicfrecquency}
K = \left( 4\Omega^2 + R \frac{d\Omega^2}{dR} \right)^{1/2}.
\end{equation}
And for the vertical velocity dispersion
\begin{equation}
\label{eq:verticaldispersion}
\sigma_z = \sqrt{c\pi G\Sigma b},
\end{equation}
with $c$ a constant and $b$ the vertical scale-length. Finally the mean azimuthal velocity , corrected by non-circular motions (asymmetric drift) is given by \citep{Binney87}
\begin{equation}
\label{eq:asymmetricdrift}
\langle v_{\phi} \rangle^2 = v_c^2 - \sigma_{\phi}^2 -\sigma_R^2( -1 - \frac{R}{\Sigma}\partial_R(\Sigma\sigma_R^2) ).
\end{equation}
The velocity dispersions and the mean azimuthal velocity are then used to distribute the gas component in the velocity space.
\subsection{The code}
The simulations were performed using the latest version of the ZEUS-MP code. It is a fixed-grid, time-explicit Eulerian code that uses an artificial viscosity to handle shocks.
The physics suite in this code includes gas hydrodynamics, ideal MHD, flux-limited radiation diffusion, self gravity, and multispecies advection \citep{H06}.
In this work we focus in solving only the standard hydrodynamics equations for the baryonic component, where the description of the physical state of a fluid element is specified by the following set of fluid equations relating the mass density ($\rho$), velocity ($\textbf{v}$) and gas internal energy density ($e$).
\begin{eqnarray}
\label{eq:2e}
\frac{D\rho}{Dt} + \rho\nabla\cdot\textbf{v} &=& 0,\\
\label{eq:2i}
\rho\frac{D\textbf{v}}{Dt} &=& -\nabla P - \rho\nabla\phi,\\
\label{eq:energia4}
\rho\frac{D}{Dt}\left(\frac{e}{\rho}\right) &=& -P\nabla\cdot\textbf{v},
\end{eqnarray}
where the Lagrangean (or comoving) derivative is given by the usual definition:
\begin{equation}
\label{eq:2l}
\frac{D}{Dt} \equiv \frac{\partial}{\partial t} + \textbf{v}\cdot\nabla.
\end{equation}
The two terms on the right-hand side of the gas momentum equation (\ref{eq:2i}) denote forces due to thermal pressure gradients, and the external gravitational potential, respectively. The self gravity of the gas is neglected in our simulations.
Because most of the material in an astrophysical system is in a highly compressible gaseous phase at very low densities, a perfect adiabatic gas is a good approximation for this system. Here we assume an adiabatic index of $\gamma = 5/3$.
We set the gravitational constant $G$ to unity and the units of mass and length to $M_u = 2.32$ $\times10^7 \rm{M_\odot}$ (galactic mass units) and $R_u = 1 \rm{kpc}$, respectively. With this choice, the resulting units of time and velocity become $t_u = 1 \times 10^8 \rm{yr}$ and $v_u = 10 \rm{kms^{-1}}$.
Cylindrical coordinates $(R; \phi; z)$ are used.
\section{Results}
\label{sec:results}
In this section we describe the temporal evolution of the rotation curves of the gas distribution embedded in the mass models showed above, formed by static dark matter and stellar disk potentials. Then to compare with observations we show the rotation curve measured from the simulation of the isolated gas distribution within the mass model that better fits the data for each one of the LSB galaxies.
Our SFDM + disk mass model is also compared with the NFW + disk mass model.
The velocity profiles showed next are identical to lower resolution studies for the same parameters, indicating convergence in the results.
But first we tests our initial conditions against dynamical stability, Figure \ref{fig:density} shows the radial density profile for the simulated gas distribution at different times in the simulation, the gas density profile is evolving but basically oscillating around the Miyamoto-Nagai profile, as desired.
\begin{figure}
\begin{center}
\includegraphics[width=8.7cm]{density.eps}
\end{center}
\caption{Radial density profile for the gas distribution embedded in our mass model, the profiles taken at times $0$, $1Gyr$ and $2Gyr$ }
\label{fig:density}
\end{figure}
Figure \ref{fig:density2} shows the gas density sliced in the midplane. As seen in this color map, the density do not increase or diminish after a $2$ $\rm{Gyr}$ evolution, indicating stability in the initial conditions.
\begin{figure*}
\begin{center}$
\begin{array}{ccc}
\vspace{-0.3cm}
\includegraphics[width=6cm]{density0.ps} & \hspace{-1cm} \includegraphics[width=6cm]{density100.ps}& \hspace{-1cm} \includegraphics[width=6cm]{density200.ps}
\end{array}$
\end{center}
\caption{Temporal evolution of the gas density, sliced in the midplane. From left to right, the snapshots show the simulation at: t = 0, 1 Gyr and 2 Gyr.
}
\label{fig:density2}
\end{figure*}
Meanwhile Figure \ref{fig:density3} plots the edge-on projection of the gas distribution and shows roughly that the density profile do not deviate from the Miyamoto-Nagai during a $2$ \rm{$Gyr$} evolution, there is not spreading or collapsing, just oscillations around the initial configuration.
\begin{figure}
\begin{center}$
\begin{array}{c}
\vspace{-3.cm}
\includegraphics[width=6cm]{2densityy0.ps}\\
\vspace{-3.cm}
\includegraphics[width=6cm]{2densityy100.ps}\\
\includegraphics[width=6cm]{2densityy200.ps}
\end{array}$
\end{center}
\caption{Temporal evolution of the gas density, sliced edge-on From top to bottom the snapshots show the simulation at: t = 0, 1 Gyr and 2 Gyr.
}
\label{fig:density3}
\end{figure}
With this test we are sure that the gas distribution is not contracting, spreading or evolving towards a different profile and that the evolution seen in rotation curves obtained from simulations is not because of a drastic relaxation, as often occurs for artificial initial conditions.
\subsubsection{Time evolution of the rotation curve}
The are many physical processes that affect a gas distribution in the galaxy. With evolving just the standard hydrodynamic equations we account for processes that may induce non-circular motions such as velocity gradients, pressure gradients and even a flared or warped gas disk, often seen in HI disks. Figure \ref{fig:timeevolution} shows the temporal evolution of the rotation curve for one of our modeled galaxies evolved for some $2$ \rm{$Gyr$ }.
\begin{figure}
\begin{center}
\includegraphics[width=8.7cm]{timeevolution.eps}
\end{center}
\caption{Temporal evolution of the rotation curve of the gas distribution embedded in one of our mass models.}
\label{fig:timeevolution}
\end{figure}
Processes related to the physics of the tracer, like those mentioned above, could account for wiggles or oscillations seen in the rotation curves. And because they evolve with time is important to consider when trying to confront a mass model with rotation curves data.
From figure \ref{fig:timeevolution} we can see that the rotation curve measured directly on the gas develops wiggles but only at early times in the simulation, this feature smooths quickly and at the end do not deviate that much from the analytical circular velocity.
\subsection{LSB galaxies data}
By performing a manual grid search we locate the parameters for our mass model that represent best four specific, bulgeless, gas-rich, lsb galaxies \citep{McGaugh01,K11b} and contrast the rotation curves data with that obtained from our simulations.
F579V1: The SFDM + disk mass model for this Lsb galaxy fits better the data with parameters $a=1.3$ \rm{$kpc$}, $b=0.2$ \rm{$kpc$} for the disk and $a_h=5.2$ \rm{$kpc$} for the dark matter halo. Figure \ref{fig:F579V1} shows the rotation curve resulting from the temporal evolution contrasted with the data. By adjusting only this three parameters we get a relatively good fit, $\chi_r^2=1.3$.
In order to do a full comparison we perform the same hydrodynamics simulations evolving the gas within the NFW + disk mass model that better fits the data. The constrictions to this model are the same as for that with the SFDM halo, the same mass for the disk and the same mass enclosed at a radius $r_0$, this parameters fixed by observations.
The best fit with the NFW + disk mass model has parameters $a=3$ \rm{$kpc$}, $b=0.3$ \rm{$kpc$} for the disk and $r_s=10$ \rm{$kpc$} for the dark matter halo. With a reduced chi-square $\chi_r^2=1.5$ the fit is good but not better than that with the cored halo.
In the figures we include also the analytic rotation curves computed from each model in order to illustrate the contribution of the live gas to the fit of the data.
\begin{figure}
\begin{center}
\includegraphics[width=8.7cm]{2F579V1.eps}
\end{center}
\caption{F579-V1: Compared Rotation curve from data (square dots), simulation with the SFDM halo + disk (open circles), simulation with the NFW halo + disk (triangles), and the initial rotation curve for both models (dashed lines).}
\label{fig:F579V1}
\end{figure}
F5683: For this galaxy we used the parameters $a=1.3$ \rm{$kpc$}, $b=0.2$ \rm{$kpc$} for the disk and $a_h=5.2$ \rm{$kpc$} for the SFDM halo. Figure \ref{fig:F5683} still shows a relatively good fit to the data, $\chi_r^2=1.8$
Here the NFW + disk mass model that fits better the data with parameters $a=5$ \rm{$kpc$}, $b=0.3$ \rm{$kpc$} for the disk and $r_s=200$ \rm{$kpc$} for the dark matter halo, and a reduced chi-square $\chi_r^2=5.03$ has not a very good fit.
The mass model with a cored halo is more favored by the data, as we can see in Figure \ref{fig:F5683} that the cuspy halo fails in recovering the inner points of the rotation curve.
\begin{figure}
\begin{center}
\includegraphics[width=8.7cm]{2F5683.eps}
\end{center}
\caption{Lsb galaxy F568-3: Compared Rotation curve from data (square dots), simulation with the SFDM halo + disk (open circles), simulation with the NFW halo + disk (triangles), and the initial rotation curve for both models (dashed lines).}
\label{fig:F5683}
\end{figure}
F5831: The cored mass model for this Lsb galaxy fits better the data with parameters $a=1.3$ \rm{$kpc$}, $b=0.2$ \rm{$kpc$} for the disk and $a_h=5.2$ \rm{$kpc$} for the dark matter halo. Figure \ref{fig:F5831} shows a relatively good fit, $\chi_r^2=4.8$, and the simulation rotation curve recovers well the innermost part of the data.
The best fit to the data for the cuspy mass model has parameters $a=2.5$ \rm{$kpc$}, $b=0.15$ \rm{$kpc$} for the disk and $r_s=45$ \rm{$kpc$} for the dark matter halo and reduced chi-square $\chi_r^2=9.5$. Although Figure \ref{fig:F5831} shows the best fit for the cuspy mass model, its rotation curve is not able to meet the inner and outer points of the data at the same time. The fit in the inner region of the rotation curve is almost as good as the cored model but misses the rest of the data.
\begin{figure}
\begin{center}
\includegraphics[width=8.7cm]{2F5831.eps}
\end{center}
\caption{Lsb galaxy F583-1: Compared Rotation curve from data (square dots), simulation with the SFDM halo + disk (open circles), simulation with the NFW halo + disk (triangles), and the initial rotation curve for both models (dashed lines).}
\label{fig:F5831}
\end{figure}
Finally, the data for the lsb galaxy F583-4 is lest smooth, with a bump in the rotation curve. Is hard to fit with a smooth halo, but our mass model still recovers the most of the data points with $\chi_r^2=21.4$.
In contrast we see in Figure \ref{fig:F5834} that because of the universality of the rotation curve profile for a NFW halo, is more difficult to fit data like these.
The best fit to the data for the cuspy mass model has a a reduced chi-square $\chi_r^2=26.1$ that do not improves that of the cored mass model.
\begin{figure}
\begin{center}
\includegraphics[width=8.7cm]{2F5834.eps}
\end{center}
\caption{Lsb galaxy F583-4: Compared Rotation curve from data (square dots), simulation with the SFDM halo + disk (open circles), simulation with the NFW halo + disk (triangles), and the initial rotation curve for both models (dashed lines).}
\label{fig:F5834}
\end{figure}
Table \ref{table:table1} summarise the parameters used in our simulations with the cored mass model. The total mass ($M_{total}$), disk mass ($M_{disk}$) and HI mass ($M_{HI}$) are fixed by observations \citep{K11b}. The scale-lengths of the disk ($a$, $b$) and the halo ($a_h$) span a range of values in order to obtain good fits.
\begin{table*}\centering
\ra{1.3}
\begin{tabular}{@{}lccccccc@{}}\toprule
& $M_{total}$ & $M_{disk}$ & $M_{HI}$ & \phantom{ab}& $a$ & $b$ & $a_h$\\
&$(10^{10}\rm{M_{\odot}})$&$(10^{10}\rm{M_{\odot}})$&$(10^{10}\rm{M_{\odot}})$& \phantom{ab}&\rm{$(kpc)$}&\rm{$(kpc)$}&\rm{$(kpc)$}\\
\midrule
F579-V1& 4.3 & 0.63 & 0.11 & & 1.3 & 0.2 & 5.2 \\
F568-3 & 5.6 & 0.8 & 0.39 & & 4.5 & 0.5 & 7 \\
F583-1 & 2.5 & 0.185 & 0.16 & & 2.5 & 0.15 & 5.5\\
F583-4 & 0.76 & 0.077 & 0.077 & & 1.2 & 0.1 & 1.2, 7\\
\bottomrule
\end{tabular}
\caption{Parameters used in the simulations and mass models. Total dynamical mass $M_{total}$, baryonic disk mass $M_{disk}$, HI mass $M_{HI}$, disk scale-lengths $a$, $b$ and DM halo scale-length $a_h$.}
\label{table:table1}
\end{table*}
\citet{R13} use a combination of states to fit rotations curves that spans at large radii and are not smooth. In this work, for the SFDM halo, we used only one state in three of the four galaxies, while F583-4 needed two states. So for this three galaxies the number of free parameters was the same for the models compared here. The added parameter in F583-4 is taken into account when computing the reduced chi-square by penalizing for this extra parameter.
It is important to note that the rotation velocity for the gas deviates just a little from the analytical circular velocity for the two models compared here. The bumps and wiggles seen in the simulations with the SFDM halo are driven almost entirely for the mass model with a small effect due to the gas.
\section{Summary and Conclusions}\label{conclusions}
The Scalar field has been proposed as a galactic dark matter model and has been tested by confronting with rotation curve data. It has been shown that when using the minimum disk hypothesis (neglecting the baryonic component) in halos with one or more nodes in their density profiles, fits to rotation curves for LSB galaxies improve significantly.
Adding the baryonic component to the mass model results in a better fit, but including live gas is a step forward from the previous works using SFDM, as for example, the rotation velocity of the gas is not always exactly equal to the circular velocity of a test particle.
LSB galaxies are dark matter dominated and when building a mass model for this objects, a cored DM halo is favored by the data. When comparing the cored and the cuspy mass models, we see that a NFW halo does it relatively well in some cases, but a SFDM halo does it better for this kind of galaxies, specially in the inner regions.
There are differences with previous works with scalar field at zero temperature, considering a finite temperature for the scalar field configuration is translated in a non-universality for the rotation curves profile of the mass model, this allows us to describe well different lsb galaxies. As already showed in previous works with this density profile, the non-universality of the rotation curves allows also to account for wiggles and bumps in the velocities data. We found that this features are too wide to be addressed to the physics of the baryons, as modeled here, and can be explained within the SFDM model, being inherent to the DM halo.
The initial conditions presented here have been tested against dynamical stability ensuring that the obtained rotation curves profiles do not present artificial effects.
We have performed three-dimensional hydrodynamic simulations with simple physical processes for the gas (we do not model star formation or feedback) that evolves within a static dark matter and stellar disk potential. And although we argue that huge quantities of feedback events (star formation, supernovas) are not needed, those should be added to our simulations but in a way different to that used in CDM-based simulations because we are not trying to change the shape of the DM density profiles through feedback processes.
The results presented here show that the observed bumps and wiggles in the rotation curves can be explained by the dark matter halo, and that gas motions as modeled here are not sufficient to explain this features. But further work is required by considering a live dark matter halo that responds to the gaseous and stellar component, this becomes important for long term evolution of a bar or a spiral pattern that might also explain the observed bumps and wiggles.
\section*{Acknowledgments}
The authors wish to thank to Victor H. Robles for many helpful discussions and to the anonymous referee for
helping us to improve the manuscript.
The numerical computations were carried out in the "Laboratorio de Super-C\'omputo Astrof\'{\i}sico (LaSumA) del Cinvestav".
The authors acknowledge to the General Coordination of Information and
Communications Technologies (CGSTIC) at CINVESTAV for providing HPC
resources on the Hybrid Cluster Supercomputer "Xiuhcoatl".
This work was partially supported by CONACyT M\'exico under grants CB-2009-01, no. 132400, CB-2011, no. 166212, and I0101/131/07 C-234/07 of the Instituto Avanzado de Cosmologia (IAC) collaboration
(http://www.iac.edu.mx/). LAMM is supported by a CONACYT scholarship.
|
\section{Introduction} The need for high strength-to and high stiffness-to-weight ratio materials has led to the development of laminated composite materials. This class of material has seen increasing utilization as structural elements and as primary structures of large-scale aerospace structures~\cite{botelhosilva2006}. This is because of the possibility to tailor the properties to optimize the structural response. Conventionally, the fibre reinforced composites have straight and unidirectional fibres. In this type of construction, the stiffness of the laminate does not vary in the domain. However, recently, the composite materials with varying stiffness has received greater interest, as they may lead to better and efficient design~\cite{hyerlee1991,setoodehabdalla2008,lopesgurdal2010,khaniijsselmuiden2011}. The stiffness of the composite material can be varied by: (a) using curvilinear fibres~\cite{hyerlee1991}; (b) varying the volume fraction or varying the fibre spacing~\cite{muruganflores2012}; (c) dropping or adding plies to the laminate~\cite{mukherjeevarughese2001} and (d) attaching discrete stiffeners to the laminates~\cite{mejdiatalla2012}. Among these concepts related to variable stiffness composite laminates (VSCL), the approach of curvilinear fibre does not introduce major geometry variations, it however does impose the constraint on the curvature of the fibre. In the following, we restrict ourselves to work related to composite laminates with curvilinear fibres. Laminates with curvilinear fibres offers a wider degree of possibilities than variations of rectilinear fibre volume fraction and provides a solution to the problem of continuity when manufacturing a structure with different fibre angles in adjoining element. Furthermore, it exhibits a way to diminish the stress concentration in cutouts. Hyer and Lee~\cite{hyerlee1991} introduced the concept of variable stiffness panels to improve the structural response of panels with holes. G\"urdal and Olmedo~\cite{gurdalolmedo1993} studied the in-plane response of VSCL. Although the concept of tailored fibre orientation was developed in the nineties, the design based on VSCL has recently spurred interest among researchers~\cite{abdallagurdal2009,akhavanribeiro2011,hondanarita2011,houmat2013,grohweaver2014,rajuwu2012,akhavanribeiro2013} due to the improvement in the manufacturing capability~\cite{kimpotter2012,kimweaver2014}. Kim \textit{et al.,}~\cite{kimpotter2012,kimweaver2014} have demonstrated a process to manufacture variable angle composites. When composite laminates are modelled as plate structures, with VSCL, the plate stiffness coefficients vary with spatial coordinates. Such laminates not only have variable in-plane stiffness, in general may possess variable bending and coupling stiffness.
Abdallla \textit{et al.,}~\cite{abdallagurdal2009} by employing the classical lamination theory and the generalised reciprocal approximation, maximized the fundamental frequency of composite laminates. Their study showed that by employing VSCL significant increase in the fundamental frequency is achieved when compared to constant stiffness panels. This was later extended to study nonlinear dynamic response of VSCL. By using the $p-$version finite element, Akhavan and Ribeiro~\cite{akhavanribeiro2011} studied the fundamental frequencies and mode shapes of laminated composites with curvilinear fibres. It was inferred that VSCL introduces greater degree of flexibility in adjusting the frequencies and mode shapes. Honda and Narita~\cite{hondanarita2011,hondanarita2012}, studied the fundamental frequency of VSCL by employing classical plate theory. It was shown that the mode shapes of VSCL are significantly different from the constant stiffness laminated composites. Houmat~\cite{houmat2013} investigated the nonlinear free vibration of laminated rectangular plates with curvilinear fibres. Groh and Weaver~\cite{grohweaver2014} and Raju \textit{et al.,}~\cite{rajuwu2012,rajuwu2013} studied buckling of variable thickness curvilinear fibre panel using differential quadrature method. It was inferred that for VSCL shell-like description more accurately characterizes the buckling phenomenon than a plate-like description. A 2D analytical model based on equivalent single-layer formulation introduced recently in~\cite{grohweaver2013} in predicting the static response of curvilinear fibre laminates and the governing equations were solved by employing the differential quadrature method. Akhavan \textit{et al.,}~\cite{akhavanribeiro2013,akhavanribeiro2013a} employed the $p-$version finite element and studied the response of laminated composites with curvilinear fibres under static and dynamic loads. The plate kinematics was represented by third-order shear deformation theory.
It is evident from the literature that the dynamic response of VCSL has received greater attention, whilst, the studies on the static characteristics of VSCL structural elements are limited. The existing studies employed either classical lamination theory, first order or the third order shear deformation theory. However, these theories have limitation when employed to study the response of curvilinear fibre laminated composites. More accurate analytical/numerical models based on the three-dimensional models can be computationally involved and expensive. A layer wise theory is a possible candidature for this purpose, but it may be computationally expensive as the number of unknowns to be solved increases with increasing number of mathematical or physical layers. Hence, among the researchers, there is a growing appreciation of the importance of applying two-dimensional theories with new kinematics for accurate analysis. It is observed from the literature~\cite{murukami1986,alibhaskar1999,makhechaganapathi2001} that, for realistic structural analysis of laminated composites, higher-order theory with the inclusion of zig-zag function is necessary. The zig-zag function incorporated in the in-plane kinematics has been employed in~\cite{murukami1986,alibhaskar1999,makhechaganapathi2001,carrera2004} to study laminated composites. Makhecha \textit{et al.,}~\cite{makhechaganapathi2001} used higher-order accurate theory based on global approach for multilayered laminated composites by incorporating the realistic through the thickness approximations of the in-plane and transverse displacements by adding a zig-zag function and higher-order terms, respectively. A 8-noded quadrilateral element with 13 degrees of freedom per node was employed for the study. Recently, some researchers have attempted to combine the single layer and discrete layer theories to overcome the limitations of each one. Carrera~\cite{carrera2003} derived a series of axiomatic approaches for the general description of two-dimensional formulations for multilayered plates and shells.
The formulation is a valuable tool for gaining a deep insight into the complex mechanics of laminated structures. In the same spirit, the main objective of this paper is to study the static response of laminated composites with curvilinear fibres. From the higher-order accurate theory, four alternate structural models are deduced by deleting the appropriate degree of freedom. A $\mathcal{C}^o$ shear flexbile quadrilateral 8-noded element is used for the present study.
The paper commences with a discussion on the variable stiffness composite laminates. Section \ref{highertheory} presents an overview of the higher-order accurate theory to describe the plate kinematics. The various structural models deduced from the higher-order accurate theory and the element employed to discretize the plate is discussed in Section \ref{eledes}. The influence of various structural models on the deflection and through thickness stress distribution is numerically studied for three layered symmetric and four layer anti-symmetric composite laminate with curvilinear fibers in Section \ref{numexamples}, followed by concluding remarks in the last section.
\section{Curvilinear fibre composite laminate}
\label{cfiber}
In the conventional laminated composites, the fibres are straight and the stiffness of the laminate is constant in the in-plane direction. However in the current study we employ curvilinear fibres. In this type of construction, the fibre path is not straight, instead it is a function of the spatial coordinates, i.e., the fibre path cannot be described by a single orientation. The main advantage is that the resulting stiffness is a function of the spatial coordinate as well. It is assumed that the fibre path varies linearly with $x$ from a orientation $T_o$ at the center to $T_1$ at a distance $a/2$ from the center, where $a$ is the total length of the laminate. The orientation of a single fibre is denoted by $\phi \pm \langle T_o/T_1 \rangle$, where $\phi$ denotes the rotation of the fibre path with respect to the $x-$ axis. The reference fibre path is given by~\cite{houmat2013}:
\begin{equation}
\renewcommand{\arraystretch}{2}
y = \left\{ \begin{array}{ccc} \frac{a}{2(T_1-T_o)} \left\{ - {\rm ln} [\cos T_o] + {\rm ln} \left[ \cos \left(T_o - \frac{2(T_1-T_o)x}{a} \right) \right] \right\} & {\rm for} & -a/2 \le x \le 0 \\ \frac{a}{2(T_1-T_o)} \left\{ {\rm ln} [\cos T_o] - {\rm ln} \left[ \cos \left(T_o + \frac{2(T_1-T_o)x}{a} \right) \right] \right\} & {\rm for} & 0 \le x \le a/2 \end{array} \right.
\end{equation}
\fref{fig:reffibpath} shows a reference fibre path based on the above equation and the shifted fibres. In this study, the fibre orientation is assumed to be just a function of $x$ and is given by~\cite{houmat2013}:
\begin{equation}
\renewcommand{\arraystretch}{2}
\theta(x) = \left\{ \begin{array}{ccc} \frac{-2}{a}(T_1 - T_o)x + T_o & ~~\textup{for} & ~~ -a/2 \le x \le 0 \\
\frac{2}{a}(T_1 - T_o)x + T_o & ~~\textup{for} & ~~ 0 \le x \le a/2
\end{array} \right.
\label{eqn:anglevari}
\end{equation}
where $T_0$ and $T_1$ are the fibre angles at the ply centre, $x=0$ and at the ends $x=\pm a/2$, respectively and is denoted by $<T_0,T_1>$.
\begin{figure}[htpb]
\centering
\includegraphics[scale=0.6]{./Figures/fiberpath}
\caption{Reference fibre path ('dashed line') and the shifted fibres ('solid line') for a laminae.}
\label{fig:reffibpath}
\end{figure}
The curvature $\kappa$ is given by:
\begin{equation}
\kappa = \frac{ \frac{d^2y}{dx^2}}{ \left[ 1 + \left( \frac{dy}{dx} \right)^2 \right]^{3/2}}
\label{eqn:cur}
\end{equation}
The curvature of the fibre path in the positive $x$ axis is found by substituting the first and the second derivatives of the fibre path in~\Eref{eqn:cur}. At each point along the fibre path, the curvature must be less than the maximum prescribed value of 3.28 m$^{-1}$~\cite{waldhar1996,houmat2013}. \fref{fig:curvConst} shows the design space for curvilinear fibre composite laminates. If at any spatial coordinate, the fibre curvature constraint is not satisfied, the corresponding fiber orientation is discarded.
\begin{figure}[htpb]
\centering
\includegraphics[scale=0.6]{./Figures/FibreConst}
\caption{Curvilinear fibre composite laminate: design space. The `dark' shaded region depicts the fibre path that satisfies the curvature constraint. The values of fibre orientation, $T_o$ and $T_1$ are in degrees.}
\label{fig:curvConst}
\end{figure}
\section{Higher order accurate theory} \label{highertheory}
A curvilinear fibre composite plate is considered with the coordinates $x,y$ along the in-plane directions and $z$ along the thickness direction. The in-plane displacements $u^sk$ and $v^k$, and the transverse displacement $w^k$ for the $k^{th}$ layer, are assumed as~\cite{alibhaskar1999,makhechaganapathi2001}:
\begin{align}
u^k(x,y,z,t) &= u_o(x,y,t) + z \theta_x(x,y,t) + z^2 \beta_x(x,y,t) + z^3 \phi_x(x,y,t) + S^k \psi_x(x,y,t) \nonumber \\
v^k(x,y,z,t) &= v_o(x,y,t) + z \theta_y(x,y,t) + z^2 \beta_y(x,y,t) + z^3 \phi_y(x,y,t) + S^k \psi_y(x,y,t) \nonumber \\
w^k(x,y,z,t) &= w_o(x,y,t) + z w_1(x,y,t) + z^2 \Gamma(x,y,t)
\label{eqn:dispField}
\end{align}
The terms with even powers of $z$ in the in-plane displacements and odd powers of $z$ occurring in the expansion for $w^k$ correspond to the stretching problem. However, the terms with odd powers of $z$ in the in-plane displacements and the even ones in the expression for $w^k$ represent the flexure problem. $u_o, v_o$ and $w_o$ are the displacements of a generic point on the reference surface; $\theta_x$ and $\theta_y$ are the rotations of the normal to the reference surface about the $y$ and $x$ axes, respectively; $w_1,\beta_x,\beta_y,\Gamma,\phi_x$ and $\phi_y$ are the higher-order terms in the Taylor`s series expansions, defined at the reference surface. $\psi_x$ and $\psi_y$ are generalized variables associated with the zigzag function, $S^k$. The zigzag function, $S^k$, as given in~\cite{murukami1986,makhechaganapathi2001,roderiguesroque2011}, is defined by
\begin{equation}
S^k = 2(-1)^k \frac{z_k}{h_k}
\end{equation}
where $z_k$ is the local transverse coordinate with the origin at the center of the $k^{th}$ layer and $h_k$ is the corresponding layer thickness. Thus, the zigzag function is piecewise linear with values of -1 and 1 alternatively at different interfaces. The `zig-zag' function, as defined above, takes care of the inclusion of the slope discontinuities of $u$ and $v$ at the interfaces of the laminated plate as observed in the exact three-dimensional elasticity solutions of thick laminated composite plates. The main advantage of using such formulation is that such a function is more economical than a discrete layer approach~\cite{nosierkapania1993,ferreira2005}. The strains in terms of mid-plane deformation, rotations of normal and higher-order terms associated with displacements are:
\begin{equation}
\boldsymbol{\varepsilon} = \left \{ \begin{array}{c} \boldsymbol{\varepsilon}_{\rm bm} \\ \boldsymbol{\varepsilon}_s \end{array} \right\} - \left \{ \begin{array}{c} \overline{\boldsymbol{\varepsilon}}_t \\ \mathbf{0} \end{array} \right\}
\label{eqn:strain}
\end{equation}
The vector $\boldsymbol{\varepsilon}_{\rm bm}$ includes the bending and membrane terms of the strain components, vector $\boldsymbol{\varepsilon}_s$ contains the transverse shear strain terms and $\overline{\boldsymbol{\varepsilon}}_t$ contains the vector of thermal strains. These strain vectors are defined as:
\begin{eqnarray}
\boldsymbol{\varepsilon}_{\rm bm} &=& \left\{ \begin{array}{c} \varepsilon_{xx} \\ \varepsilon_{yy} \\ \varepsilon_{zz} \\ \gamma_{xy} \end{array} \right\} + \left\{ \begin{array}{c} u_{,x} \\ v_{,y} \\ w_{,z} \\ u_{,y} + v_{,x} \end{array} \right\} \nonumber \\
&=& \boldsymbol{\varepsilon}_0 + z \boldsymbol{\varepsilon}_1 + z^2 \boldsymbol{\varepsilon}_2 + z^3 \boldsymbol{\varepsilon}_3 + S^k \boldsymbol{\varepsilon}_4
\end{eqnarray}
\begin{eqnarray}
\boldsymbol{\varepsilon}_s &=& \left\{ \begin{array}{c} \gamma_{xz} \\ \gamma_{yz} \end{array} \right\} = \left\{ \begin{array}{c} u_{,z} + w_{,x} \\ v_{,z} + w_{,y} \end{array} \right\} \nonumber \\
&=& \gamma_o + z \gamma_1 + z^2 \gamma_2 + S^k_{,z} \gamma_3
\end{eqnarray}
where
\begin{align}
\boldsymbol{\varepsilon}_o &= \left\{ \begin{array}{c} u_{o,x} \\ v_{o,y} \\ w_1 \\ u_{o,y} + v_{o,x} \end{array} \right\},\hspace{1cm} \boldsymbol{\varepsilon}_1 = \left\{ \begin{array}{c} \theta_{x,x} \\ \theta_{y,y} \\ 2\Gamma \\ \theta_{x,y} + \theta_{y,x} \end{array} \right\}, \nonumber \\
\boldsymbol{\varepsilon}_2 &= \left\{ \begin{array}{c} \beta_{x,x} \\ \beta_{y,y} \\ 0 \\ \beta_{x,y} + \beta_{y,x} \end{array} \right\}, \hspace{1cm} \boldsymbol{\varepsilon}_3 = \left\{ \begin{array}{c} \phi_{x,x} \\ \phi_{y,y} \\ 0 \\ \phi_{x,y} + \phi_{y,x} \end{array} \right\}, \nonumber \\
\boldsymbol{\varepsilon}_4 &= \left\{ \begin{array}{c} \psi_{x,x} \\ \psi_{y,y} \\ 0 \\ \psi_{x,y} + \psi_{y,x} \end{array} \right\}.
\end{align}
and,
\begin{align}
\gamma_o &= \left\{ \begin{array}{c} \theta_x + w_{o,x} \\ \theta_y + w_{o,y} \end{array} \right\}, \hspace{1cm} \gamma_1 = \left\{ \begin{array}{c} 2\beta_x + w_{1,x} \\ 2\beta_y + w_{1,y} \end{array} \right\}, \nonumber \\
\gamma_2 &= \left\{ \begin{array}{c} 3\phi_x + \Gamma_{,x} \\ 3\phi_y + \Gamma_{,y} \end{array} \right\}, \hspace{1cm} \gamma_3 = \left\{ \begin{array}{c} \psi_x S_{,z}^k \\ \psi_y S_{,z}^k \end{array} \right\} .
\end{align}
The subscript comma denotes partial derivatives with respect to the spatial coordinate succeeding it. The thermal strain vector, $\overline{\boldsymbol{\varepsilon}}_t$ is given by:
\begin{align}
\overline{\boldsymbol{\varepsilon}}_t &= \left\{ \begin{array}{cccccc} \varepsilon_{xx} & \varepsilon_{yy} & \varepsilon_{zz} & \varepsilon_{xy} & \varepsilon_{yz} & \varepsilon_{xz} \end{array} \right\}^{\rm T} \nonumber \\
&= \Delta T \left\{ \begin{array}{cccccc} \alpha_{xx} & \alpha_{yy} & \alpha_{zz} & \alpha_{xy} & 0 & 0 \end{array} \right\}^{\rm T}
\end{align}
where $\Delta T$ is the rise in the temperature and is generally represented as a function of the spatial coordinates, $\alpha_{xx}, \alpha_{yy}, \alpha_{zz}$ and $\alpha_{xy}$ are thermal expansion coefficients in the plate coordinates and can be related to the thermal expansion coefficients in the material principal directions. The constitutive relations for an arbitrary layer $k$ can be expressed as:
\begin{eqnarray}
\boldsymbol{\sigma} &=& \left\{ \begin{array}{cccccc} \sigma_{xx} & \sigma_{yy} & \sigma_{zz} & \tau_{xz} & \tau_{yz} & \tau_{xy} \end{array} \right\}^{\rm T} \nonumber \\
&=& \bf{\overline{Q}}^k \left\{ \begin{array}{cc} \boldsymbol{\varepsilon}_{\rm bm} & \boldsymbol{\varepsilon}_s \end{array} \right\}^{\rm T} - \bf{\overline{Q}}^k \overline{\boldsymbol{\varepsilon}}_t
\end{eqnarray}
where the terms of $\bf{\overline{Q}}^k $ of $k^{th}$ ply are referred to the laminate axes and can be obtained from the $\bf{Q}^k$ matrix corresponding to the fibre directions with the appropriate transformations. The transformed stiffnesses are constant in a straight fibre laminate whilst for a curvilinear fibre laminate, it is a function of spatial coordinates $(x,y)$. The strain energy function $U$ is given by
\begin{equation}
U( \boldsymbol{\delta}) = \frac{1}{2} \iint \left[ \sum_{k=1}^n \int\limits_{h_k}^{h_{k+1}} \boldsymbol{\sigma}^{\rm T} \boldsymbol{\varepsilon} ~dz \right] dx dy - \iint {\bf q} w~dxdy
\label{eqn:potential}
\end{equation}
where $\boldsymbol{\delta}$ is the vector of degrees of freedom and ${\bf q}$ is the distributed force acting on the top surface of the plate. By following the Galerkin procedure outlined in~\cite{hughes2000}, one obtains the following governing equations for static deflection
\begin{equation}
\mathbf{K} \boldsymbol{\delta} = \mathbf{f}_m + \mathbf{f}_t
\label{eqn:staticdefl}
\end{equation}
where $\mathbf{K}$ is the global stiffness matrix and $\mathbf{f}_m$ and $\mathbf{f}_t$ are the global mechanical and thermal load vectors, respectively.
\section{Element description}
\label{eledes}
In this paper, $\mathcal{C}^o$ continuous, eight-noded serendipity quadrilateral shear flexible plate element is used. The finite element represented as per the kinematics based on \Eref{eqn:dispField} is referred to as HSDT13 with cubic variation. The 13 dofs are: $(u_o,v_o,w_o,\theta_x,\theta_y,w_1,\beta_x,\beta_y,\Gamma,\phi_x,\phi_y,\psi_x,\psi_y)$. Four more alternate discrete models are proposed to study the influence of higher-order terms in the displacement functions, whose displacement fields are deduced from the original element by deleting the appropriate degrees of freedom. These structural models, and the corresponding degrees of freedom are listed in Table \ref{table:alternatemodels}
\begin{table} [htpb]
\renewcommand\arraystretch{1.5}
\caption{Structural models employed in this study for describing the plate kinematics.}
\centering
\begin{tabular}{ll}
\hline
Finite element model & Degrees of freedom per node \\
\hline
HSDT13 & $u_o,v_o,w_o,\theta_x,\theta_y,w_1,\beta_x,\beta_y,\Gamma,\phi_x,\phi_y,\psi_x,\psi_y$ \\
HSDT11A & $u_o,v_o,w_o,\theta_x,\theta_y,\beta_x,\beta_y,\phi_x,\phi_y,\psi_x,\psi_y$ \\
HSDT11B & $u_o,v_o,w_o,\theta_x,\theta_y,w_1,\beta_x,\beta_y,\Gamma,\phi_x,\phi_y$ \\
TSDT7 & $u_o,v_o,w_o,\theta_x,\theta_y,\beta_x,\beta_y$ \\
FSDT5 & $u_o,v_o,w_o,\theta_x,\theta_y$ \\
\hline
\end{tabular}
\label{table:alternatemodels}
\end{table}
\section{Numerical results and discussion}
\label{numexamples}
In this section, the static bending response of curvilinear fibre composite laminated plates using the eight-noded shear flexible quadrilateral element is presented. The effect of various parameters, such as the plate thickness, the ply-angle and the lay-up sequence on the global response is numerically studied considering mechanical and thermal loads. Here, the laminate is assumed to be simply supported with movable boundary conditions. For the higher-order structural theory (HSDT13), the boundary conditions are given by:
\begin{eqnarray}
u_o = w_o = \theta_x = w_1 = \Gamma = \beta_x = \phi_x = \psi_x = 0, \hspace{0.2cm} ~\textup{on} ~ y = 0,b \nonumber \\
v_o = w_o = \theta_y = w_1 = \Gamma = \beta_y = \phi_y = \psi_y = 0, \hspace{0.2cm} ~\textup{on} ~ x= 0,a
\end{eqnarray}
where $a$ and $b$ refer to the length and width of the plate, respectively. The transverse shear stresses are evaluated by integrating the three-dimensional equilibrium equations for all types of elements. The material properties, unless otherwise specified, used in the present analysis are:
\begin{eqnarray}
E_L/E_T = 25, \hspace{10pt} G_{LT}/E_T = 0.5, \hspace{10pt} G_{TT}/E_T = 0.2 \nonumber \\
\nu_{LT}=0.25, \hspace{10pt} \alpha_T/\alpha_L = 1125, \hspace{10pt} \alpha_L = 10^{-6} \nonumber \\
E_T = 10^9 {\rm GPa}
\end{eqnarray}
where $E,G$ and $\nu$ are Young's modulus, shear modulus and Poisson's ratio, respectively. $L$ and $T$ are the longitudinal and transverse directions, respectively, with respect to fibres. All the layers are of equal thickness and the ply angle is measures with respect to the $x$-axis. For the present study, the following two types of loadings are considered:
\begin{itemize}
\item Mechanical loading: $q(x,y) = q_o \sin \frac{\pi x}{a} \sin \frac{\pi y}{b}$,
\item Thermal loading: $\Delta T(x,y) = \overline{T}_o \left( \frac{2z}{h} \right) \sin \frac{\pi x}{a} \sin \frac{\pi y}{b}$
\end{itemize}
where $q_o$ and $\overline{T}_o$ are the amplitude of the mechanical load and the thermal load, respectively. The physical quantities are non-dimensionalized by relations, unless stated otherwise:
\begin{align}
\renewcommand{\arraystretch}{2}
(\overline{u},\overline{v}) &= \frac{100E_{_T}}{q_o h S^3}(u,v) \nonumber \\
\overline{w} &= \frac{100E_T}{q_o h S^4}w \nonumber \\
\overline{\sigma}_{xx} &= \frac{\sigma_{xx}}{q_oS^2} \nonumber \\
\overline{\tau}_{xz} &= \frac{\tau_{xz}}{q_oS}
\end{align}
for the applied mechanical load and by
\begin{align}
\renewcommand{\arraystretch}{2}
(\overline{u},\overline{v}) &= \frac{E_T}{ h \alpha_L \overline{T}_o S}(u,v) \nonumber \\
\overline{w} &= \frac{E_T}{h \alpha_L \overline{T}_o h S^2}w\nonumber \\
\overline{\sigma}_{xx} &= \frac{\sigma_{xx}}{ E_T \alpha_L \overline{T}_o} \nonumber \\
\overline{\tau}_{xz} &= \frac{\tau_{xz}}{ E_T \alpha_L \overline{T}_o}
\end{align}
for the applied thermal load, where the $E_{T}, \alpha_{L}$ are the Young's modulus and the co-efficient of thermal expansion of the laminated composite in the transverse and longitudinal directions, respectively, and $S=a/h$. Considering different types of four-layered curvilinear fibre laminates, the convergence study is carried out by decreasing the mesh size and the results are presented in Table \ref{table:curvfibMeshconv} for both thick and thin plates ($a/h=$ 5 and $a/h=$ 50) by employing both the first- and higher-order (FSDT5, HSDT13) structural models. A very good convergence of the results is observed with decreasing element size. For the problem considered here, a structured quadrilateral mesh with 6$\times$6 mesh is found to be adequate to model the full plate.
\begin{table} [htpb]
\renewcommand\arraystretch{1.2}
\caption{Convergence study with mesh size for a four-layered curvilinear fibre composite plate under sinusoidally distributed pressure load.}
\centering
\begin{tabular}{ccrrrrrrrr}
\hline
$a/h$ & $T_o/T_1$ & Mesh size & \multicolumn{6}{c}{Structural model}\\
\cline{4-10}
& $(^\circ)$& & \multicolumn{3}{c}{HSDT13} && \multicolumn{3}{c}{FSDT5} \\
\cline{4-6}\cline{8-10}
& & & $\overline{w}$ & $\overline{\tau}_{xz}$ & $\overline{\sigma}_{xx}$ && $\overline{w}$ & $\overline{\tau}_{xz}$ & $\overline{\sigma}_{xx}$ \\
\hline
\multirow{15}{*}{5} & \multirow{5}{*}{$[0 \pm \langle 45/0 \rangle]_s$} & 2$\times$2 & 1.039 & 0.524 & 0.155 && 0.960 & 0.493 & 0.176 \\
&& 4$\times$4 & 1.277 & 0.296 & 0.360 && 1.189 & 0.333 & 0.252 \\
&& 6$\times$6 & 1.300 & 0.234 & 0.372 && 1.202 & 0.259 & 0.263 \\
&& 8$\times$8 & 1.305 & 0.198 & 0.378 && 1.204 & 0.210 & 0.268 \\
&& 16 $\times$16 & 1.305 & 0.196 & 0.379 && 1.204 & 0.210 & 0.270 \\
\cline{3-10}
& \multirow{5}{*}{$[0 \pm \langle 0/45 \rangle]_s$} & 2$\times$2 & 0.851 & -0.161 & 0.539 && 0.850 & -0.134 & 0.393 \\
&& 4$\times$4 & 1.179 & 0.153 & 0.372 && 1.125 & 0.254 & 0.321 \\
&& 6$\times$6 & 1.204 & 0.182 & 0.408 && 1.138 & 0.248 & 0.314 \\
&& 8$\times$8 & 1.210 & 0.228 & 0.414 && 1.141 & 0.298 & 0.313 \\
&& 16 $\times$16 & 1.211 & 0.232 & 0.416 && 1.142 & 0.302 & 0.311\\
\cline{3-10}
& \multirow{5}{*}{$[0 \pm \langle 90/0 \rangle]_s$} & 2$\times$2 &0.918 & 0.750 & 0.065 && 0.861 & 0.692 & 0.053 \\
&& 4$\times$4 & 1.267 & -0.037 & 0.096 && 1.204 & 0.022 & 0.093 \\
&& 6$\times$6 & 1.307 & -0.070 & 0.102 && 1.239 & 0.033 & 0.097 \\
&& 8$\times$8 & 1.318 & -0.080 & 0.103 && 1.248 & 0.032 & 0.095 \\
&& 16 $\times$16 & 1.319 & -0.085 & 0.104 && 1.248 & 0.031 & 0.092 \\
\cline{2-10}
\multirow{15}{*}{50} & \multirow{5}{*}{$[0 \pm \langle 45/0 \rangle]_s$} & 2$\times$2 & 0.466 & 0.417 & 0.149 && 0.434 & 0.374 & 0.244 \\
&& 4$\times$4 & 0.444 & 0.289 & 0.301 && 0.443 & 0.299 & 0.275 \\
&& 6$\times$6 & 0.445 & 0.285 & 0.277 && 0.443 & 0.283 & 0.277 \\
&& 8$\times$8 & 0.445 & 0.273 & 0.287 && 0.444 & 0.275 & 0.281 \\
&& 16 $\times$16 & 0.445 & 0.270 & 0.291 && 0.444 & 0.272 & 0.280 \\
\cline{3-10}
& \multirow{5}{*}{$[0 \pm \langle 0/45\rangle]_s$} & 2$\times$2 & 0.296 & 0.150 & 0.426 && 0.287 & 0.154 & 0.407 \\
&& 4$\times$4 & 0.322 & 0.207 & 0.384 && 0.319 & 0.260 & 0.390 \\
&& 6$\times$6 & 0.324 & 0.235 & 0.391 && 0.321 & 0.241 & 0.381 \\
&& 8$\times$8 & 0.325 & 0.249 & 0.376 && 0.322 & 0.299 & 0.378 \\
&& 16 $\times$16 & 0.325 & 0.252 & 0.372 && 0.323 & 0.303 & 0.375 \\
\cline{3-10}
& \multirow{5}{*}{$[0 \pm \langle 90/0 \rangle]_s$} & 2$\times$2 & 0.413 & 0.475 & 0.036 && 0.388 & 0.428 & 0.033 \\
&& 4$\times$4 & 0.445 & 0.068 & 0.047 && 0.439 & 0.073 & 0.049 \\
&& 6$\times$6 & 0.454 & 0.097 & 0.062 && 0.451 & 0.087 & 0.061 \\
&& 8$\times$8 & 0.454 & 0.057 & 0.064 && 0.452 & 0.064 & 0.063 \\
&& 16 $\times$16 & 0.455 & 0.055 & 0.065 && 0.452 & 0.060 & 0.064 \\
\hline
\end{tabular}
\label{table:curvfibMeshconv}
\end{table}
\begin{table} [htpb]
\renewcommand\arraystretch{1.2}
\caption{Comparison of deflection and stresses for a three-layered cross ply simply supported square plate subjected to sinusoidally distributed mechanical load against elasticity solution~\cite{bhaskarvaradan1996}.}
\centering
\begin{tabular}{crrrrrr}
\hline
$a/h$ & & $\overline{u}(-a/2,0,h/2)$ & $\overline{v}(0,-a/2,h/2)$ & $\overline{w}(0,0,h/2)$ & $\overline{\sigma}_{xx}(0,0,h/2)$ & $\overline{\tau}_{xz}(-a/2,0,0)$ \\
\hline
\multirow{4}{*}{4} & HSDT13 & -0.977 & -2.275 & 2.094 & 0.825 & 0.245 \\
& HSDT11A & -0.977 & -2.332 & 2.027 & 0.805 & 0.247 \\
& HSDT11B & -0.967 & -2.178 & 1.993 & 0.816 & 0.270 \\
& Elasticity~\cite{bhaskarvaradan1996}& -0.969 & -2.281 & 2.006 & 0.755 & 0.256 \\
\cline{2-7}
\multirow{4}{*}{10} & HSDT13 & -0.745 & -1.103 & 0.752 & 0.611 & 0.344 \\
& HSDT11A & -0.750 & -1.112 & 0.755 & 0.613 & 0.345 \\
& HSDT11B & -0.736 & -1.049 & 0.715 & 0.604 & 0.353 \\
& Elasticity~\cite{bhaskarvaradan1996}& -0.735 & -1.099 & 0.753 & 0.590 & 0.357 \\
\cline{2-7}
\multirow{4}{*}{20} & HSDT13 & -0.699 & -0.796 & 0.516 & 0.569 & 0.373 \\
& HSDT11A & -0.700 & -0.798 & 0.517 & 0.570 & 0.373 \\
& HSDT11B & -0.696 & -0.779 & 0.505 & 0.567 & 0.376 \\
& Elasticity~\cite{bhaskarvaradan1996}& -0.693 & -0.794 & 0.516 & 0.553 & 0.383 \\
\cline{2-7}
\multirow{4}{*}{50} & HSDT13 & -0.682 & -0.697 & 0.445 & 0.555 & 0.384 \\
& HSDT11A & -0.682 & -0.697 & 0.445 & 0.555 & 0.384 \\
& HSDT11B & -0.681 & -0.694 & 0.443 & 0.554 & 0.385 \\
& Elasticity~\cite{bhaskarvaradan1996}& -0.680 & -0.697 & 0.445 & 0.541 & 0.393 \\
\hline
\end{tabular}
\label{table:3L_mech_validation}
\end{table}
\begin{table} [htpb]
\renewcommand\arraystretch{1.2}
\caption{Comparison of deflection and stresses for a three-layered cross ply simply supported square plate subjected to sinusoidally distributed thermal load against elasticity solution~\cite{bhaskarvaradan1996}.}
\centering
\begin{tabular}{crrrrrr}
\hline
$a/h$ & & $\overline{u}(-a/2,0,h/2)$ & $\overline{v}(0,-a/2,h/2)$ & $\overline{w}(0,0,h/2)$ & $\overline{\sigma}_{xx}(0,0,h/2)$ & $\overline{\tau}_{xz}(-a/2,0,-h/6)$ \\
\hline
\multirow{4}{*}{4} & HSDT13 & -18.069 & -81.010 & 42.331 & 1216.860 & 82.926 \\
& HSDT11A & -14.457 & -77.187 & 26.222 & 919.206 & 85.649 \\
& HSDT11B & -18.396 & -75.714 & 42.040 & 1239.820 & 87.841 \\
& Elasticity~\cite{bhaskarvaradan1996}& -18.110 & -81.830 & 42.330 & 1183.000 & 84.810 \\
\cline{2-7}
\multirow{4}{*}{10} & HSDT13 & -16.738 & -31.687 & 17.374 & 1067.370 & 58.920 \\
& HSDT11A & -16.046 & -30.991 & 14.656 & 1009.980 & 58.621 \\
& HSDT11B & -16.689 & -29.978 & 16.901 & 1062.050 & 60.113 \\
& Elasticity~\cite{bhaskarvaradan1996}& -16.610 & -31.950 & 17.390 & 1026.000 & 60.540 \\
\cline{2-7}
\multirow{4}{*}{20} & HSDT13 & -16.255 & -20.183 & 12.116 & 1018.480 & 33.164 \\
& HSDT11A & -16.076 & -20.003 & 11.428 & 1003.610 & 32.855 \\
& HSDT11B & -16.229 & -19.694 & 11.961 & 1016.050 & 33.365 \\
& Elasticity~\cite{bhaskarvaradan1996}& -16.170 & -20.340 & 12.120 & 982.00 & 33.980 \\
\cline{2-7}
\multirow{4}{*}{50} & HSDT13 & -16.039 & -16.651 & 10.493 & 998.252 & 13.807 \\
& HSDT11A & -16.010 & -16.622 & 10.383 & 995.838 & 13.663 \\
& HSDT11B & -16.032 & -16.569 & 10.466 & 997.641 & 13.823 \\
& Elasticity~\cite{bhaskarvaradan1996}& -16.020 & -16.710 & 10.500 & 967.500 & 14.070 \\
\hline
\end{tabular}
\label{table:3L_therm_validation}
\end{table}
\begin{table}[htpb]
\centering
\caption{Comparison of different structural models in predicting the deflection of a four-layered curvilinear fibres composite laminate $([0 \pm \langle 45/0 \rangle]_s)$ under uniform pressure load against available solutions~\cite{grohweaver2013}.}
\begin{tabular}{rrrrr}
\hline
& \multicolumn{4}{c}{$a/h$} \\
\cline{2-5}
& 5 & 10 & 20 & 50 \\
\hline
HSDT13 & -3.823 & -2.393 & -1.995 & -1.882 \\
HSDT11A & -3.730 & -2.317 & -1.930 & -1.814 \\
HSDT11B & -3.822 & -2.393 & -1.995 & -1.882 \\
TSDT7 & -3.368 & -2.214 & -1.899 & -1.807 \\
FSDT5 & -3.671 & -2.298 & -1.922 & -1.811 \\
2D equivalent~\cite{grohweaver2013} & -4.002 & -2.396 & -1.943 & -1.805 \\
FEM S8R~\cite{grohweaver2013} & -3.761 & -2.311 & -1.913 & -1.796 \\
FEM C3D20R~\cite{grohweaver2013} & -4.630 & -2.520 & -2.000 & -1.820 \\
\hline
\end{tabular}
\label{table:curvFibLamValida}
\end{table}
Before proceeding with the detailed investigation, the efficacy of the present formulation is validated considering the static response of cross-ply straight fibre laminates subjected to mechanical and thermal loads. The results obtained with the present formulation are shown in Tables \ref{table:3L_mech_validation} - \ref{table:3L_therm_validation} for different plate side to thickness ratios. It is observed that the results from the present formulation agree very well with the results available in the literature~\cite{bhaskarvaradan1996}. Next, the various structural models proposed here are tested considering curvilinear fibre composite laminates subjected to mechanical load for which results are available in the literature~\cite{grohweaver2013}. The maximum transverse displacements calculated are tabulated in Table \ref{table:curvFibLamValida} for various plate side to thickness ratios. It can be observed from this Table that the present results are found to be in excellent agreement with those reported in the literature~\cite{grohweaver2013} and it may be further inferred that the performance of the present HSDT13 structural model is better than the analytical model and 2D element employed in~\cite{grohweaver2013}.
A systematic parametric study is further conducted to examine the suitability of an appropriate structural theory using different structural models deduced from the present formulation as given in Table \ref{table:alternatemodels}. The computed results pertaining to maximum displacements and stresses for different plate side to thickness ratios and curvilinear fibre angles $( \langle T_o/T_1 \rangle )$ are highlighted in Tables \ref{table:Table6symmMech} - \ref{table:Table7symmTherm}. For the present study, a three-layered symmetric laminate configuration is considered. It is observed from the Tables \ref{table:Table6symmMech} - \ref{table:Table7symmTherm} that, for the mechanical load case, the higher-order model, HSDT11A, is in close agreement with HDST13 in predicting the displacements, whereas, for the thermal loading case, the higher-order model HSDT11B is in close agreement with the HSDT13. However, depending on the curvilinear ply-angle and the plate aspect ratio, the stress values predicted either by HSDT11A or HSDT11B are close to the values of full model, HSDT13. Also, it may be opined that the influence of lower-order theories, HSDT7 and FSDT5 underestimates the deflection and are somewhat different from those of HSDT13 and HSDT11 (HSDT11A and HSDT11B). The displacements and the stress distribution through the thickness are shown in \frefs{fig:3LayerMechLoad} - \ref{fig:3LayerThermLoad} for thick laminate with curvilinear ply-angle $( \langle$ 45$^\circ$/-45$^\circ \rangle)$. The in-plane displacement is continuous through the thickness whereas the transverse displacement varies quadratically through the thickness. Also, it can be seen that the slope of the in-plane displacements shows discontinuity at the interface as expected. From Tables \ref{table:3L_mech_validation}-\ref{table:3L_therm_validation} and Tables \ref{table:Table6symmMech} - \ref{table:Table7symmTherm}, one can opine that, for the straight fibre laminates, either HSDT11A or HSDT11B may in general be employed depending on the loading types (mechanical or thermal), however, for curvilinear case, HSDT13 may have to be used for accurate results. Furthermore, it is observed that the differences in the performance among various theories are more for curvilinear case compared to the straight fibre case. It can be further noted that the lower order models (HSDT7 and FSDT5) cannot predict the through thickness variation of displacements.
\input{Curved-static-tables6_7_sym}
\begin{figure}
\centering
\newlength\figureheight
\newlength\figurewidth
\setlength\figureheight{10cm}
\setlength\figurewidth{12cm}
\subfigure[$\overline{u} = \overline{u}(-a/2,0,z)$]{\scalebox{0.5}{\input{./Figures/3L_mech_udisp.tikz}}}\hspace{60pt}
\subfigure[$\overline{w} = \overline{w}(0,0,z)$]{\scalebox{0.5}{\input{./Figures/3L_mech_wdisp.tikz}}}
\subfigure[$\overline{\sigma}_{xx} = \overline{\sigma}_{xx}(0,0,z)$]{\scalebox{0.5}{\input{./Figures/3L_mech_sxx.tikz}}}\hspace{60pt}
\subfigure[$\overline{\tau}_{yz} = \overline{\tau}_{yz}(-a/2,0,z)$]{\scalebox{0.5}{\input{./Figures/3L_mech_syz.tikz}}}
\subfigure[$\overline{\tau}_{xz}=\overline{\tau}_{xz}(-a/2,0,z)$]{\scalebox{0.5}{\input{./Figures/3L_mech_sxz.tikz}}}\hspace{60pt}
\subfigure[$\overline{\sigma}_{zz} = \overline{\sigma}_{zz}(0,0,z)$]{\scalebox{0.5}{\input{./Figures/3L_mech_szz.tikz}}}
\caption{The variation of non-dimensional deflections and stresses through the thickness of a three-layered symmetric curved fibre composite laminate under sinusoidally distributed mechanical load with $a/h=$ 5, $T_o=$ -45$^\circ$, $T_1=$ 45$^\circ$.}
\label{fig:3LayerMechLoad}
\end{figure}
\begin{figure}
\centering
\setlength\figureheight{10cm}
\setlength\figurewidth{12cm}
\subfigure[$\overline{u} = \overline{u}(-a/2,0,z)$]{\scalebox{0.5}{\input{./Figures/3L_therm_udisp.tikz}}}\hspace{60pt}
\subfigure[$\overline{w} = \overline{w}(0,0,z)$]{\scalebox{0.5}{\input{./Figures/3L_therm_wdisp.tikz}}}
\subfigure[$\overline{\sigma}_{xx} = \overline{\sigma}_{xx}(0,0,z)$]{\scalebox{0.5}{\input{./Figures/3L_therm_sxx.tikz}}}\hspace{60pt}
\subfigure[$\overline{\tau}_{yz} = \overline{\tau}_{yz}(-a/2,0,z)$]{\scalebox{0.5}{\input{./Figures/3L_therm_syz.tikz}}}
\subfigure[$\overline{\tau}_{xz}=\overline{\tau}_{xz}(-a/2,0,z)$]{\scalebox{0.5}{\input{./Figures/3L_therm_sxz.tikz}}}\hspace{60pt}
\subfigure[$\overline{\sigma}_{zz} = \overline{\sigma}_{zz}(0,0,z)$]{\scalebox{0.5}{\input{./Figures/3L_therm_szz.tikz}}}
\caption{The variation of non-dimensional deflections and stresses through the thickness of a three-layered symmetric curved fibre composite laminate under sinusoidally distributed thermal load with $a/h=$ 5, $T_o=$ -45$^\circ$, $T_1=$ 45$^\circ$.}
\label{fig:3LayerThermLoad}
\end{figure}
\begin{figure}
\centering
\setlength\figureheight{10cm}
\setlength\figurewidth{12cm}
\subfigure[$\overline{u} = \overline{u}(-a/2,0,z)$]{\scalebox{0.5}{\input{./Figures/4L_mech_udisp.tikz}}}\hspace{60pt}
\subfigure[$\overline{w} = \overline{w}(0,0,z)$]{\scalebox{0.5}{\input{./Figures/4L_mech_wdisp.tikz}}}
\subfigure[$\overline{\sigma}_{xx} = \overline{\sigma}_{xx}(0,0,z)$]{\scalebox{0.5}{\input{./Figures/4L_mech_sxx.tikz}}}\hspace{60pt}
\subfigure[$\overline{\tau}_{yz} = \overline{\tau}_{yz}(-a/2,0,z)$]{\scalebox{0.5}{\input{./Figures/4L_mech_syz.tikz}}}
\subfigure[$\overline{\tau}_{xz}=\overline{\tau}_{xz}(-a/2,0,z)$]{\scalebox{0.5}{\input{./Figures/4L_mech_sxz.tikz}}}\hspace{60pt}
\subfigure[$\overline{\sigma}_{zz} = \overline{\sigma}_{zz}(0,0,z)$]{\scalebox{0.5}{\input{./Figures/4L_mech_szz.tikz}}}
\caption{The variation of non-dimensional deflections and stresses through the thickness of a four-layered anti-symmetric curved fibre composite laminate under sinusoidally distributed mechanical load with $a/h=$ 5, $T_o=$ -45$^\circ$, $T_1=$ 45$^\circ$.}
\label{fig:4LayerMechLoad}
\end{figure}
\begin{figure}
\centering
\setlength\figureheight{10cm}
\setlength\figurewidth{12cm}
\subfigure[$\overline{u} = \overline{u}(-a/2,0,z)$]{\scalebox{0.5}{\input{./Figures/4L_therm_udisp.tikz}}}\hspace{60pt}
\subfigure[$\overline{w} = \overline{w}(0,0,z)$]{\scalebox{0.5}{\input{./Figures/4L_therm_wdisp.tikz}}}
\subfigure[$\overline{\sigma}_{xx} = \overline{\sigma}_{xx}(0,0,z)$]{\scalebox{0.5}{\input{./Figures/4L_therm_sxx.tikz}}}\hspace{60pt}
\subfigure[$\overline{\tau}_{yz} = \overline{\tau}_{yz}(-a/2,0,z)$]{\scalebox{0.5}{\input{./Figures/4L_therm_syz.tikz}}}
\subfigure[$\overline{\tau}_{xz}=\overline{\tau}_{xz}(-a/2,0,z)$]{\scalebox{0.5}{\input{./Figures/4L_therm_sxz.tikz}}}\hspace{60pt}
\subfigure[$\overline{\sigma}_{zz} = \overline{\sigma}_{zz}(0,0,z)$]{\scalebox{0.5}{\input{./Figures/4L_therm_szz.tikz}}}
\caption{The variation of non-dimensional deflections and stresses through the thickness of a four-layered anti-symmetric curved fibre composite laminate under sinusoidally distributed thermal load with $a/h=$ 5, $T_o=$ -45$^\circ$, $T_1=$ 45$^\circ$.}
\label{fig:4LayerThermLoad}
\end{figure}
Similar study is made considering four-layered anti-symmetric laminates and the maximum displacements and stresses evaluated considering both mechanical and thermal loads are presented in Tables \ref{table:4LmechAntiSymm} - \ref{table:4LthermAntiSymm}. The displacement and stress plots through the thickness of the laminate are depicted in \frefs{fig:4LayerMechLoad} - \ref{fig:4LayerThermLoad} for mechanical and thermal loads, respectively. The behaviour of different structural models are qualitatively similar to those of symmetric case. However, from \frefs{fig:3LayerMechLoad} - \ref{fig:4LayerThermLoad}, it is seen that the variation of stresses through the thickness predicted by lower-order theories are very much quantitatively different from those of symmetric case.
\begin{table}[htpb]
\centering
\caption{Maximum non-dimensionalized deflections and stresses for a four-layered anti-symmetric curvilinear fibre composite panels $( \langle -45^\circ,45^\circ \rangle, \langle 45^\circ,-45^\circ \rangle, \langle -45^\circ,45^\circ \rangle, \langle 45^\circ,-45^\circ \rangle)$ under sinusoidally distributed mechanical load.}
\begin{tabular}{crrrrr}
\hline
$a/h$ & Structural & $\overline{u}(-a/2,0,h/2)$ & $\overline{w}(0,0,h/2)$ & $\overline{\sigma}_{xx}(0,0,h/2)$ & $\overline{\tau}_{xz}(-a/2,0,0)$ \\
\hline
\multirow{5}{*}{5} & HSDT13 & -1.006 & 1.338 & 0.316 & -0.068 \\
& HSDT11A & -0.998 & 1.317 & 0.307 & -0.075 \\
& HSDT11B & -0.983 & 1.299 & 0.305 & -0.051 \\
& TSDT7 & -0.737 & 1.104 & 0.174 & 0.027 \\
& FSDT5& -0.565 & 1.104 & 0.279 & 0.122 \\
\cline{2-6}
\multirow{5}{*}{10} & HSDT13 & -0.727 & 0.667 & 0.286 & 0.013 \\
& HSDT11A & -0.728 & 0.671 & 0.285 & 0.010 \\
& HSDT11B & -0.720 & 0.647 & 0.277 & 0.032 \\
& TSDT7 & -0.653 & 0.600 & 0.237 & 0.068 \\
& FSDT5& -0.578 & 0.574 & 0.276 & 0.171 \\
\cline{2-6}
\multirow{5}{*}{50} & HSDT13 & -0.561 & 0.398 & 0.250 & 0.164 \\
& HSDT11A & -0.561 & 0.399 & 0.250 & 0.162 \\
& HSDT11B & -0.560 & 0.397 & 0.250 & 0.168 \\
& TSDT7 & -0.557 & 0.394 & 0.249 & 0.165 \\
& FSDT5& -0.531 & 0.381 & 0.257 & 0.311 \\
\hline
\end{tabular}
\label{table:4LmechAntiSymm}
\end{table}
\begin{table}[htpb]
\centering
\caption{Maximum non-dimensionalized deflections and stresses for a four-layered anti-symmetric curvilinear fibre composite panels $( \langle -45^\circ,45^\circ \rangle, \langle 45^\circ,-45^\circ \rangle, \langle -45^\circ,45^\circ \rangle, \langle 45^\circ,-45^\circ \rangle)$ under sinusoidally distributed thermal load.}
\begin{tabular}{crrrrr}
\hline
$a/h$ & Structural & $\overline{u}(-a/2,0,h/2)$ & $\overline{w}(0,0,h/2)$ & $\overline{\sigma}_{xx}(0,0,h/2)$ & $\overline{\tau}_{xz}(-a/2,0,-h/6)$ \\
\hline
\multirow{5}{*}{5} & HSDT13 & -18.956 & 28.981 & -166.942 & -136.980 \\
& HSDT11A & -16.059 & 18.141 & -279.092 & -124.888 \\
& HSDT11B & -19.015 & 28.861 & -169.577 & -128.414 \\
& TSDT7 & -13.978 & 15.957 & -509.318 & -104.787 \\
& FSDT5& -7.079 & 11.576 & -25.160 & -52.125 \\
\cline{2-6}
\multirow{5}{*}{10} & HSDT13 & -16.698 & 16.368 & -61.095 & -58.213 \\
& HSDT11A & -15.916 & 13.610 & -90.983 & -58.089 \\
& HSDT11B & -16.747 & 16.178 & -76.937 & -60.139 \\
& TSDT7 & -15.300 & 12.717 & -193.895 & -54.290 \\
& FSDT5& -12.498 & 10.791 & -0.795 & 28.503 \\
\cline{2-6}
\multirow{5}{*}{50} & HSDT13 & -14.283 & 10.432 & -75.392 & -6.851 \\
& HSDT11A & -14.255 & 10.325 & -76.578 & -6.977 \\
& HSDT11B & -14.266 & 10.398 & -75.355 & -6.955 \\
& TSDT7 & -14.165 & 10.232 & -73.767 & -7.103 \\
& FSDT5& -13.499 & 9.851 & -41.047 & -0.448 \\
\hline
\end{tabular}
\label{table:4LthermAntiSymm}
\end{table}
\section{Conclusion}
The performance of various structural models in predicting the static response of curvilinear fibre composite laminates is examined considering various parameters such as ply-angle, lay-up sequence, plate side to thickness ratio and loadings. From a detailed investigation, the following observations can be made:
\begin{itemize}
\item HSDT11A predicts accurately the in-plane and transverse deflection of the curvilinear laminates subjected to mechanical loading, whereas, for thermal loading, HSDT11B model yields accurate solutions in comparison with those of the full structural model considered here, viz., HSDT13.
\item The choice of HSDT11A or HSDT11B in evaluating the stress values and the variation of the stresses through the thickness highly depend on the plate side to thickness ratio and the curvilinear ply angle.
\item The performance of higher-order models for anti-symmetric lay-up case, in general, significantly noticeable over the lower-order models.
\item In-plane relative displacement has slope discontinuity at the layer interface whereas the transverse displacement varies quadratically through the thickness.
\item Higher-order model HSDT13 may be required for accurate prediction of displacements and stress variation though the thickness, irrespective of loading situation, ply-angle and lay-up sequence, in particular, for curvilinear case.
\end{itemize}
\section*{Acknowledgement}
S Natarajan would like to acknowledge the financial support of the School of Civil and Environmental Engineering, The University of New South Wales, for his research fellowship for the period September 2012 onwards.
\section*{References}
\bibliographystyle{elsarticle-num}
\subsubsection{Mechanical loading}
The static analysis is conducted for carbon nanotube reinforced functionally graded sandwich plate with homogeneous core. The following two types of mechanical loading are considered:
\begin{itemize}
\item Mechanical loading
\begin{itemize}
\item Sinusoidal loading: $q(x,y) = q_o \sin \frac{\pi x}{a} \sin \frac{\pi y}{b}$,
\item Uniformly distributed loading: $q(x,y) = q_o$.
\end{itemize}
\item Thermal loading: $T(x,y) = T_o \left( \frac{2z}{h} \right) \sin \frac{\pi x}{a} \sin \frac{\pi y}{b}$
\end{itemize}
where $q_o$ and $T_o$ are the amplitude of the mechanical load and the thermal load, respectively. The physical quantities are nondimensionalized by relations, unless stated otherwise:
\begin{align}
\renewcommand{\arraystretch}{2}
(\overline{u},\overline{v}) &= \frac{100E_{_H}}{q_o h S^3}(u,v) \nonumber \\
\overline{w} &= \frac{100E_H}{q_o h S^4}w \nonumber \\
\overline{\sigma}_{xx} &= \frac{\sigma_{xx}}{q_oS^2} \nonumber \\
\overline{\sigma}_{xz} &= \frac{\sigma_{xz}}{q_oS}
\end{align}
for the applied mechanical load and by
\begin{align}
\renewcommand{\arraystretch}{2}
(\overline{u},\overline{v}) &= \frac{1}{10 h \alpha_H T_o S}(u,v) \nonumber \\
\overline{w} &= \frac{1}{h \alpha_H T_o h S^4}w\nonumber \\
\overline{\sigma}_{xx} &= \frac{\sigma_{xx}}{100 E_H \alpha_H T_o} \nonumber \\
\overline{\sigma}_{xz} &= \frac{\sigma_{xz}}{10 E_H \alpha_H T_o}
\end{align}
for the applied thermal load, where the $E_{_H}, \alpha_{_H}$ are the Young's modulus and the co-efficient of thermal expansion corresponding to the homogeneous core, evaluated at $T_o=$ 300K and $S=a/h$.
Before proceeding to the detailed analysis of the static response of the sandwich plate for the applied mechanical load, the present formulation is validated against available solutions in the literature. In this case, only one layer is considered. The CNTs are either uniformly distributed or functionally graded along the thickness direction, given by:
\begin{equation}
V_{\rm CN}(z) = \left\{ \begin{array}{cl} V_{\rm CN}^\ast &\textup{UD} \\ \left(1 + \frac{2z}{h} \right) V_{\rm CN}^\ast & \textup{FG-V Type} \\ 2\left(\frac{2|z|}{h} \right) V_{\rm CN}^\ast & \textup{FG-X Type} \end{array} \right.
\end{equation}
The effective material properties, viz., Young's modulus, Poisson's ratio and the mass density are estimated from \Eref{eqn:effecprop}. The effect of type of CNT volume fraction distribution is also considered. PmPV~\cite{zhulei2012} is considered as the matrix with material properties: $E_m=$ 2.1 GPa at room temperature (300K), $\nu_m=$ 0.34. The SWCNTs are chosen as the reinforcements and the material properties for the SWCNT are taken from~\cite{zhulei2012} and given in Table \ref{table:cntproperty}. The CNT efficiency parameters $\eta_j$ are determined according to the effective properties of CNTRCs available by matrching the Young's moduli $E_1$ and $E_2$ with the counterparts computed by rule of mixtures. In this study, the efficiency parameters are: $\eta_1=$ 0.149 and $\eta_2=$ 0.934 for the case of $V_{\rm CN}^\ast=$ 0.11; $\eta_1=$ 0.150 and $\eta_2=$ 0.941 for the case of $V_{\rm CN}^\ast=$ 0.14 and $\eta_1=$ 0.149 and $\eta_2=$ 1.381 for the case of $V_{\rm CN}^\ast=$ 0.17. It is assumed that $\eta_2=\eta_3$ for this study. |
\section{Introduction}
If all mesons were $\bar qq$ states then there would
be no natural reason for poles in scattering amplitudes to occur very close to
thresholds. At large values of $N_{c}$, the number of colors in QCD, all
$\bar qq$ mesons become narrow with (nearly) unchanged mass \cite{'tHooft:1973jz}. Thus, their masses have
no relation to the masses of the mesons to which they couple\footnote{The same is true for a straightforward extension of
tetraquarks to large $N_c$ \cite{Weinberg:2013cfa} --- in case they existed at large $N_c$
\cite{Cohen:2014tga} --- although other possible extensions of tetraquarks to $N_c\neq3$ lead
to masses that grow when $N_c$ is increased \cite{Cohen:2014vta}.}.
Accordingly, the $\rho$ mass is not related to $2m_{\pi}$, nor is the $K^{*}$ mass
related to $m_{K}+m_{\pi}$. On the contrary, there is good reason for ``extraordinary'' hadrons,
often called hadronic molecules, to have masses close to
thresholds~\cite{Jaffe:2007id}. So the mere fact that the $f_{0}(980)$ and
$a_{0}(980)$ appear very near $K\bar K$ threshold is already a reason to be
suspicious that they may not be simple $\bar qq$ states. The same applies to
unusual charmonium states that have been found near charm-anticharm meson
thresholds like the famous $X(3872)$ located very close to the $D^0\bar D^{0 *}$ threshold --- for a recent review see~\cite{hqwg}.
In this paper we look carefully at the way
that the manifestations of poles in scattering amplitudes change as the poles
approach thresholds as some strength parameter is varied --- here one may
think of varying the quark masses in lattice QCD calculations. This has acquired
a renewed interest after the trajectory
of the $\sigma$ or $f_0(500)$ resonance pole as a function of the quark mass
was predicted by us
within unitarized Chiral Perturbation Theory \cite{Hanhart:2008mx}.
A similar trajectory as that of the $\sigma$ was soon shown to be followed by the controversial
$\kappa$ or $K(800)$ resonance in the isospin 1/2 scalar $\pi K$ scattering partial wave,
including the appearance of a virtual states at sufficiently large pion masses \cite{Nebreda:2010wv}.
Recently the existence of such a virtual bound state in $\pi K$ scattering at high pion masses has been confirmed by lattice
calculations \cite{Dudek:2014qha}. The subtleties of the extraction of resonance parameters from lattice QCD simulations performed at
a finite volume are outlined in detail in Refs.~\cite{Bernard:2010fp,Doring:2011vk} and
will not be discussed further here.
While finishing this work, we became aware of
a theoretical study \cite{Hyodo:2014bda} of the scaling of hadron masses
near an $s$-wave threshold, showing that the bound state energy is not continuously connected
to the real part of the resonance energy. In this paper we have another look at this issue
which allows us to provide various additional, non-trivial insights.
In particular, we
demonstrate that there is a qualitative difference between the pole
trajectories of resonances that couple to the relevant continuum channel is an
$s$--wave or in a higher partial wave: As a consequence of analyticity a
resonance is characterized by two poles on the second sheet, one located at
$s=s_R$ and one located at $s=s_R^*$. For narrow resonances only one of them
is close to the physical region. As some strength parameter is increased, the
two poles start to approach each other. We will show on general grounds
that while for higher partial waves the poles meet at the corresponding
two meson threshold, for $s$--waves the poles can still be located inside the
complex plane even for the real part of the pole position at or below
threshold.
As a consequence, $s$--wave--trajectories are controlled by an
additional dimensionful parameter, namely the value of $s$ where the two poles
meet below threshold which may be related to the structure of the state.
In other words, generic trajectories of $s$-wave resonances
do lead to poles
whose real part of the position is below threshold, but whose imaginary part of the position does not vanish,
before giving rise to virtual bound states, and then bound states, as some strength parameter is varied.
While this observation is in line with the findings of Refs.~\cite{Hanhart:2008mx,Guo:2009ct}, it
is in vast conflict with
``common wisdom'' that the imaginary part of a pole has to be identified with one half of its decaying width,
for this implies that, if the ``resonance mass'' --- identified with the real part of the pole position ---
lies below threshold, the pole necessarily has to
lie on the real axis. Of course, such identification is a fair approximation
for narrow resonances far from thresholds or other singularities, but very inappropriate
in the cases we will show below.
The paper is organized as follows: in the next Section we discuss general properties of the
poles that appear in the $S$-matrix, paying especial attention to poles that occur
in partial waves with angular momenta higher than 0, and in particular to the role of the centrifugal barrier which is absent in
the scalar partial waves. Next we consider the
trajectories of resonance poles in the complex plane as a function of some strength parameter,
and how they can become bound states. In the next section we briefly review Weinberg's compositeness criterion
and reformulate it in terms of the parameters introduced in the previous section.
The possible behaviors are then illustrated with two models of scattering in separable potentials
within non-relativistic scattering theory, one with a single channel and another one in a two-channel system.
In Section 4 we analyze
the realistic examples of the pole trajectories of the $\sigma$
or $f_0(500)$ scalar meson and the $\rho(770)$ as functions of the quark masses, obtained
from the combination of Chiral Perturbation Theory and a single channel dispersion relation
obtained in \cite{Hanhart:2008mx}.
We show how
the generic features discussed in this paper show up in these two cases.
In particular, we can conclude that the $f_0(500)$ or sigma meson would have a predominantly
molecular nature, if the pion mass were of the order of $450$ MeV or higher.
In the final Section we summarize our results.
\section{General properties of $S$--matrix poles}
In this work we only consider one continuum channel. This implies
that the $S$--matrix has one right hand cut, starting at
$s=(2m)^2$ --- the so
called unitarity cut\footnote{For simplicity we only consider the case of
scattering of two particles with equal mass, however, the
generalization to unequal masses is straightforward.}. As a consequence there are two sheets
and, as usual, we call first or physical sheet the one
corresponding to a momentum with a positive imaginary part. The $S$
matrix evaluated on sheet $I$ ($II$) is written as $S_I(s)$
($S_{II}(s)$). If no subscript is given, the expression holds for both
sheets. It follows directly from unitarity and analyticity
that~\cite{book}
\begin{equation}
S_I(s)=1/S_{II}(s) \quad \mbox{and} \quad \left[S(s)\right]^*=S(s^*) \ .
\label{sprops}
\end{equation}
As a consequence a pole on the second sheet immediately implies a zero
on the first and vice-versa. In addition, if there is a pole at $s=s_0$, there must
also be a pole at $s=s_0^*$, i.e., poles outside the real axis occur in conjugate pairs.
Furthermore, it can be shown that the only poles
allowed on the physical sheet are bound state poles, namely, those located
on the real axis below threshold.
A different, but equivalent, way to discuss the pole structure of the
$S$--matrix is to use the $k$--plane: instead of the Mandelstam
variable $s$, the center of mass momentum $k$ is used to characterize
the energy of the system. The two quantities are related via
\begin{equation}
k = \sqrt{s/4-m^2} \ .
\end{equation}
The obvious advantage is that there is no right hand cut with respect
to $k$ and correspondingly there is only one sheet. It follows
directly from the definition that the upper (lower) half plane of the
complex $k$--plane, defined by positive (negative) values of the
imaginary part of $k$, maps onto the first (second) sheet of the
$s$--plane. The conditions derived above from Eq.~(\ref{sprops}) translate
into the $k$--plane as follows: the only poles allowed in the upper half plane are
on the imaginary axis and in the lower half plane appear as mirror
images with respect to the imaginary axis.
The relation between the different planes is illustrated in Fig.~\ref{kvss}.
On the one hand, it becomes clear from the figure that the resonance pole located at $r$
is the one closest to the physical axis and therefore physically more
relevant than the one at $r'$ in the vicinity of the pole. On the other hand,
near threshold both poles are equally relevant regardless where they are located
in the second sheet.
Finally, in the $k$ plane virtual states appear as poles on the negative
imaginary axis (labeled as $v$ in the figure) and bound states
as poles on the positive imaginary axis (labeled as $b$ in the figure).
\begin{figure}[t!]
\begin{center}
\begin{center}
\epsfig{file=sheets.eps,width=0.6\textwidth}
\end{center}
\caption{Relation between $k$--plane and $s$--plane:
on the left the $k$--plane is shown. The (red) $x$s denote the
physical axis. On the right the two $s$--plane sheets are shown.
Here the broad band indicates the position of the unitarity cut.
The upper (lower) half plane of the $k$--plane
maps onto the first (second) sheet in the $s$--plane such that
the points $A$-$D$ get transferred as indicated in the figure.
In addition, the allowed pole positions in the complex plane
are also shown as $x$. They are labeled as $b$ for the
bound state, $v$ for the virtual state, and $r$ and $r'$ for
the two conjugate poles of the resonance state.}
\label{kvss}
\end{center}
\end{figure}
Now, assuming that there is at least one resonance pole, and that it is not too far away from threshold,
we are now in the position of writing down the most general expression
for the $S$--matrix in the vicinity of that pole or its conjugate partner.
For the derivation it is easier to use
the $k$ plane, and thus we assume that there is a resonance pole at
$k=k_{\rm p}-i\gamma$ with $\gamma>0$.
For a resonance, $k_{\rm p}$ is a real number and we choose $k_{\rm p}>0$, for, as commented above,
it corresponds to the pole closest to the physical axis.
Then, from the above considerations it follows that there is in addition
a pole at $k=-k_{\rm p}-i\gamma$ and zeros at $k=\pm k_{\rm p}+i\gamma$. We may
therefore, dropping terms of higher order in $k$,
and for a particular partial wave ${\ell}$,
write the following general expression for the
$S$--matrix element in the vicinity of the pole~\cite{book}:
\begin{equation}
S_{\ell}(k) =e^{i\phi(k)}\frac{(k-k_{\rm p}-i\gamma)(k+k_{\rm p}-i\gamma)}{(k-k_{\rm p}+i\gamma)(k+k_{\rm p}+i\gamma)} \ ,
\label{Spara}
\end{equation}
where $\phi(k)$ is a smooth function, real valued for real, positive values of $k$. For simplicity
this phase factor will be dropped in what follows.
Using the definition of the $T$ matrix, $S=1+2ikT$, we may write
\begin{equation}
T_{\ell}(k) =-\frac{2\,\gamma}{k^2-(\gamma^2+k_{\rm p}^2)+2i\gamma k} \ .
\label{Tpara}
\end{equation}
For elastic scattering unitarity provides a stringent link between the real and the
imaginary part of $T$, that actually allows the
$T$--matrix elements to be described in terms of its phase $\delta$ as
\begin{equation}
\arctan (\delta) = -\frac{2\,k\,\gamma}{k^2-(\gamma^2+k_{\rm p}^2)} \ .
\label{phase}
\end{equation}
Of course, it is straightforward to recast the above expressions in terms of $s$ instead of $k$.
One finds for example
\begin{equation}
S_{\ell}(k) =\frac{s-s_0-4i(s-4m^2)^{1/2}\gamma}{s-s_0+4i(s-4m^2)^{1/2}\gamma} \ ,
\label{Spara_s}
\end{equation}
with $s_0=4\left(k_{\rm p}^2+\gamma^2+m^2\right)$.
\subsection{$\ell>0$ partial wave threshold behavior and poles}
In general the centrifugal barrier demands, for momenta
much
smaller than some typical scale $\mu$,
that the scattering amplitude behaves as $T_{\ell}\propto k^{2\ell}$.
If we are only interested in this low energy region,
the constraint translates into the replacement
\begin{equation}
\gamma = \gamma(k)=\bar \gamma k^{2\ell} \ .
\label{gol}
\end{equation}
Of course, this amplitude should only be used for $k$ much smaller than
the typical scale $\mu$, not beyond.
Note that $\ell=0$ waves are unaffected by this change, but, for example, $\ell=1$ waves
now have poles whenever $i\bar \gamma k^2+k\pm k_{\rm p}=0$, namely at:
\begin{equation}
k_{\rm pole}=\frac{i}{2\bar\gamma}
\Big[ 1\pm\sqrt{1\mp4i\bar\gamma k_{\rm p}}\Big],\qquad (\ell=1\;\hbox{case}).
\end{equation}
These are four poles in conjugated pairs, but of course,
they are only meaningful if
they lie within the low momentum region of validity of our amplitude.
We can ensure that we have only one conjugated pair within this region
if we require $\bar\gamma k_{\rm p}\ll 1$, (which is nothing but assuming that
both parameters are natural, i.e., $\bar \gamma\ll 1/\mu$ and $k_{\rm p}\ll \mu$).
In such case we can expand
\begin{equation}
\label{squareroot}
\sqrt{1\mp 4i\bar\gamma k_{\rm p}}\simeq
1 \mp 2i\bar\gamma k_{\rm p} + 2\bar\gamma^2k_{\rm p}^2 + ...,
\end{equation}
so that the four poles lie at:
\begin{equation}
\label{kp}
k_{\rm pole}\simeq \frac{i}{2\bar\gamma}\Big[1 \pm
\left(1 \mp 2i\bar\gamma k_{\rm p} + 2\bar\gamma^2k_{\rm p}^2 \right)
\Big]=\left\{\mat{
&
\mp k_{\rm p} - i\bar\gamma k_{\rm p}^2. \qquad\qquad\hbox{ (Physical pair)}\\
&
\pm k_{\rm p}+\frac{i}{\bar\gamma}\Big(1+(\bar\gamma k_{\rm p})^2\Big)
\hbox{ (Unphysical pair)}
}\right.
\end{equation}
One should not worry about the unphysical conjugated pair of poles
lying on the first sheet, since our amplitude
has been constructed for $1/ \bar \gamma\gg k_{\rm p}$
and thus
these spurious poles are deep in the complex plane,
beyond the range of applicability of our approach, which is however valid for
the two poles not too far from threshold.
A similar pattern emerges for even higher partial waves, with a physical pair
for small $k$ and additional unphysical pairs of poles beyond the applicability
region of our amplitude.
\subsection{Pole trajectories as a function of a strength parameter}
In the construction presented in the previous section it was assumed that
$k_{\rm p}$ is a real number --- then the equations describe a resonance. We will
now generalize this investigation by considering the movement of the poles as
some strength parameter is varied. Therefore we will study how the resonance
properties change when varying $k_{\rm p}$. In particular, it is interesting to
observe the trajectories of the poles for $k_{\rm p}\rightarrow0$ especially very
close to threshold. Of course, as long as two
conjugate poles exist, their trajectories have to be symmetric with respect to
the imaginary $k$ axis.
\begin{figure}
\begin{center}
\includegraphics[width=7cm]{simplep-wave.eps}
\vspace*{-0.1in}
\end{center}
\caption[x]{\footnotesize Motion of the p-wave poles in the complex $s$-plane
with $m=1$ and $\bar \gamma=0.2$.}
\label{Fig:ppoles}
\end{figure}
Let us first follow the trajectories followed by conjugate poles
for $\ell>0$ partial waves. As a consequence of Eq.~(\ref{gol}), they will
come infinitesimally close to $k_{\rm pole}=0$.
However, as commented above
there can be no poles of the $S$--matrix on the physical cut.
This property is automatically implemented in Eq.~(\ref{Spara})
for when $k_{\rm p}=0$ and $\gamma=0$ simultaneously, the zeros
in the numerator and denominator
cancel to yield $S(k=0)=1$. The resulting pole trajectories in the $s$ plane are illustrated
for $\ell=1$ in Fig.~\ref{Fig:ppoles}.
In contrast, for $s$--waves
the point where the two conjugate poles meet each other on the imaginary $k$ axis
is not fixed except for the condition that no poles in the physical axis
should exist in the first Riemann sheet, and in particular not at $k=0$.
But that leaves the whole negative axis
for s-wave poles to meet when $k_{\rm p}$ decreases
and the point where the two trajectories meet, $-i\gamma$,
is a non--trivial parameter of the underlying dynamics.
This is one of the central messages of this paper.
In order to extend our discussion to poles
below threshold, we need to continue analytically
$k_{\rm p}$ to complex values $k_{\rm p}=i\kappa_{\rm p}$, with $\kappa_{\rm p}$ real and positive
so that $k_{\rm p}^2$ is real and negative. Of course, for our
amplitude to make sense we still keep the condition
$\bar\gamma\kappa_{\rm p}\ll1$.
Now, for the $\ell >0$ case, we find two physical poles at $k_{pole}=i\kappa_{\rm p}(\pm1+\bar\gamma \kappa_{\rm p})$.
In the s-plane, these are two poles
below threshold but one in the first and another one in
the second Riemann sheet.
The resonance has become a bound state.
Since $\bar\gamma\kappa_{\rm p}\ll1$ they lie almost symmetrically
with respect to the threshold.
As seen from the $s$-plane, this is the
typical structure of subthreshold poles
in the first and second Riemann sheets.
\begin{figure}
\begin{center}
\includegraphics[width=7cm]{simples-wave.eps}
\vspace*{-0.1in}
\end{center}
\caption[x]{\footnotesize Motion of the s-wave poles in the complex $s$-plane with $m=1$ and $\gamma=0.2$. }
\label{Fig:spoles}
\end{figure}
In contrast, for scalar waves, we
find two poles in the imaginary axis at $k_{pole}=i(\pm\kappa_{\rm p}-\gamma)$.
Note that, as $\kappa_{\rm p}$ grows,
the two poles start separating from each other and move apart from
the ``meeting point'', $-i\gamma$. In the $k$--plane
both move along the imaginary $k$--axis, the physically more relevant one
is located at $-i(\gamma-\kappa_{\rm p})$ while the other one is located at $-i(\gamma+\kappa_{\rm p})$.
Correspondingly in the $s$--plane the two poles move along the
real axis below threshold, but both of them lying on the second sheet
until $\kappa_{\rm p}=\gamma$. Eventually, when $\kappa_{\rm p}>\gamma$, the first pole moves
to the physical sheet --- the virtual state turns into a bound state.
The corresponding motion of the poles is illustrated in Fig.~\ref{Fig:spoles}.
One may define the mass $M$ of a particle as the real part of
the corresponding pole position in the complex plane. It is
therefore interesting to follow this point as $\gamma$ and $k_{\rm p}$
vary. For $s$--waves, in general one finds a striking non--analytic behavior
in $M$ at the point where $k_{\rm p}=0$. On the other hand, for partial
waves higher than $s$--waves the behavior much smoother
\footnote{This non-analytical behavior in hadron masses may also propagate to other
observables, like form factors in the $t$--channel~\cite{Guo:2011gc}.}.
This non-analyticity of the hadron mass when the conjugate poles reach the real axis
as $k_{\rm p}\to 0$
has been recently studied in detail within the general formalism of Jost functions in \cite{Hyodo:2014bda}.
The conclusion is a similar warning to the one we raised in Ref.~\cite{Hanhart:2008mx} about
the naive mass extrapolation formulas for states which appear near thresholds
on the lattice, although within a more general framework.
In this work we will illustrate this non-analyticity
in passing, when explaining the different possible pole trajectories
on the basis of various examples below, whereas for the analytic formalism of those
mass singularities we simply refer the reader to \cite{Hyodo:2014bda}.
\section{Summary of Weinberg's Criterion}
\label{nature}
In Refs.~\cite{weinberg} Weinberg developed a criterion for compositeness for
bound states that occur in the $s$-wave of scattering amplitudes (under which
circumstances this formalism
can be generalized to resonances is described in Ref.~\cite{evidence}).
The starting point is the scattering amplitude near threshold that may be expressed
in terms of the scattering length, $a$, and the effective range, $r$ as\footnote{Note
that sometimes a different sign convention is used for the scattering length.}
\begin{equation}
T_{0}(k)= \frac{1}{k\cot\delta_{0}(k)-ik} = \frac{1}{-1/a+rk^{2}/2 -ik} \ .
\label{e.9}
\end{equation}
Weinberg derived relations between the scattering length, $a$,
the effective range, $r$, and the wave function renormalization constant for
the particle described by the $S$-matrix pole, $Z$,
\begin{align}
a=2\left(\frac{1-Z}{2-Z}\right)R + {\cal O}(1/\beta) \ , \quad
r= -\left(\frac{Z}{1-Z}\right)R + {\cal O}(1/\beta) \ .
\label{e.7}
\end{align}
Here $\beta$ is the typical momentum scale of the binding interactions ---
in our case either $\beta\sim m_{\pi}$ or larger, depending on whether single pion exchange is important in the process ---
and $R$ is the inverse of the imaginary momentum corresponding to the $S$-matrix pole
\begin{equation}
R = \frac{1}{\kappa} = \sqrt{\frac{1}{2\mu B }}
\label{e.8}
\end{equation}
when the pole in $S$ occurs at $s=4(m^{2}-\kappa^{2})=4m^{2}-4mB$.
For a
bound state $\kappa>0$ and for a virtual state $\kappa<0$ (in both cases
$B>0$).
On general grounds one can show that in leading order in an expansion in $1/(R\beta)$ $Z$
can be interpreted as the probability to find the ordinary, compact component in the
wave function of the physical state; especially $0\le Z\le 1$.
Assuming that the effective range approximation is valid for all
momenta of interest, from Eq.~(\ref{e.9}) evaluated at the pole follows a single
kinematic relation among $a$, $r$, and $R$
\begin{equation}
\frac{1}{R}=\frac{1}{a} +\frac{r}{2R^{2}}
\label{e.10}
\end{equation}
Weinberg found that a predominantly composite state (or hadronic molecule
or extraordinary hadron) has $Z\approx 0$. In the case of a
weakly bound particle, where $R\gg 1/\beta$, this criterion reduces to $a\sim
R$ and $r\sim 1/\beta$, with the range term in Eq.~(\ref{e.10}) just
providing
a small correction.
Note that within potential scattering one can show that the terms of order
$1/\beta$ are typically positive.
On the other hand, a predominantly elementary state
has $Z\approx 1$ and therefore $a\sim 0$, or more accurately $a\sim
1/\beta$, and
and $|r|\gg R$. As a result, in order to get a bound state near threshold
for a predominantly elementary state a fine tuning between the range term and the scattering length term is
necessary in Eq.~(\ref{e.10})\footnote{This situation may be accompanied by quite unusual line
shapes as demonstrated in Refs.~\cite{interplay1,interplay2}.}.
This clearly demonstrates, that it is way more natural to find composite
states near thresholds than elementary states.
The argument can be easily expressed in terms of the parameters introduced in
the previous section. From Eq.~(\ref{Tpara}) one finds
\begin{equation}
\label{arofgammak}
a=-\frac{2\gamma}{\gamma^2+k_{\rm p}^2} \quad \mbox{and} \quad r=\frac1{\gamma} \ .
\end{equation}
Using Eq.~(\ref{e.7}) this can be translated to
\begin{equation}
\label{Zofgammak}
Z=1-\frac{\gamma}{\kappa_{\rm p}} \ .
\end{equation}
As explained in the previous section, for
a shallow bound state one has $\kappa_{\rm p}-\gamma=\kappa\ll \beta$. Thus,
we again recover that the most natural situation for a near threshold
pole is $Z\simeq 0$, since only if both $\kappa_{\rm p}$ and $\gamma$
are individually much smaller than $\beta$ and at the same time $\gamma\ll
\kappa_{\rm p}$, then $Z\simeq 1$, referring
to an elementary state. Clearly, for this to be realized
fine tuning is necessary.
In the following we will illustrate the patterns described above
on two simple models. (Some pole trajectories within a
specific coupled channel model were already shown in \cite{Lesniak}).
Both of these models are based on non-relativistic scattering in a separable
potential. Model A is a single channel separable potential. If the potential
is attractive and strong enough, it can generate an $S$-matrix pole. In
\cite{Jaffe:2007id} it was argued that this is a ``toy model'' for an extraordinary
hadron which would vanish as $N_{c}\to\infty$. Model B is a two channel model
where there is no diagonal interaction in the open channel, but there is a
bound state in the closed channel (a Feshbach resonance). In
\cite{Jaffe:2007id} it was argued that this is a model for an ``ordinary hadron'',
whose width would go to zero as $N_{c}\to\infty$.
\subsection{Model A }
Model A has a separable potential that only couples to a single partial wave
with angular momentum $l$. The scattering amplitude in all other partial
waves is zero. For the partial wave with angular momentum $l$, the
Schr\"odinger equation is
\begin{equation}
\la{schr}
-u''_{\ell} (r)+\frac{\ell(\ell+1)}{r^{2}}u_{\ell}(r)- \lambda \int_{0}^{\infty}dr'v(r)v(r')u_{\ell}(r')=Eu_{\ell}(r'),
\end{equation}
To make things simple
$v(r)$ is chosen such that the integrals can be done analytically:
\begin{equation}
\label{eq:V*}
v(r)=\sqrt{2} \mu^{3/2} e^{-\mu r},
\end{equation}
Although for $r\sim 1/\mu$ the behavior of the system depends
on the form chosen for $v(r)$, the behavior for $r\ll 1/\mu$ is genuine.
Then one finds for the scattering amplitude, $f_{l}(k)$,
\begin{equation}
f_{l}(k)=\frac{k\lambda \xi_{l}^{2}(k)}{1-\frac{2\lambda}{\pi}\int_{0}^{\infty}dq q^{2}\frac{\xi_{l}^{2}(q)}{q^{2}-k^{2}-i\epsilon}} \equiv\frac{N_l(k)}{D_l(k)}
\la{sepscatt}
\end{equation}
where $\xi_{l}(k)=\int_{0}^{\infty}rj_{l}(kr)v(r)$, and in particular
\begin{equation}
\xi_0(k)=\frac{\sqrt{2}}{k^2+1}.
\label{xinew}
\end{equation}
Here units are chosen such that $\mu=1$ --- accordingly
the model should reproduce the genuine behavior discussed above for $k\ll 1$.
One can compute $N(k)$ and $D(k)$ explicitly; for the $s$-wave ($\ell=0$),
\begin{align*}
N_{0}(k)= \frac{2k\lambda}{(k^2+1)^2},\quad
D_{0}(k)= 1+\frac{\lambda}{(k+i)^2}.
\end{align*}
For $s$-wave poles located near a threshold one may use Weinberg's criterion
to pin down the degree of compositeness of the corresponding physical state.
Here closeness to the threshold translates into $k\ll 1$. Then we may
read off from the expressions given above
\begin{equation}
\label{armodelAnew}
a_A=\frac{2\lambda}{\lambda-1} \quad \mbox{and} \quad
r_A=\frac{\lambda+2}{\lambda}.
\end{equation}
A bound state is present only if the interactions are attractive
and $a>0$, which translates into $\lambda>1$ ($\lambda<0$ refers to a repulsive interaction). In
addition, $r$ is always positive --- which means that in Eq.~(\ref{e.7}) for
the range the $1/\beta$ term dominates. Thus, the pole is located very near
threshold only for $r_A/(2a_A) \ll 1$,
as follows straightforwardly from
Eq.~(\ref{e.10}) --- within the model this ratio does not exceed 0.3 showing
that Model A produces extraordinary hadrons in the whole parameter range
where bound states are produced.
Equivalently one may also directly calculate the wave function renormalization
constant for the $s$-wave --- one finds it consistent with 0 within the
uncertainties. Thus, the single partial wave separable potential
generates an $S$-matrix pole dynamically that mimics an extraordinary hadron: a hadronic molecule.
\subsection{Model B }
This model is designed to show the scattering effects of a confined state when
it can be treated non-relativistically using the Schroedinger equation.
It is described in detail in \cite{Jaffe:2007id},
Section II.C, and will not be repeated here. Basically it maps onto a
separable potential model but with $\lambda\to \lambda/(E-E_{0})$, where
$E_{0}$ is the energy of the confined channel bound state. (Feshbach showed
that near such a state the two channel Schr\"odinger Equation collapses to a single channel
equation with a separable potential.) Here are the numerator and denominator
of the $s$-wave scattering amplitude for this model:
\begin{align}
N_{0}(k)&= \frac{2k\lambda}{(k^2+1)^2},\\
D_{0}(k)&= k^2-k_0^2+\frac{\lambda}{(k+i)^2}.
\label{feshbachswavenew}
\end{align}
The only difference between the two models is that the $1$ in $D_{0}(k)$ in
model A is replaced by $k^{2}-k_{0}^{2}$ in model B --- said differently:
model A is recovered from model B in the limit $k_0^2\to \infty$ while
$\bar\lambda=-\lambda/k_0^2$ is kept finite. In particular, this implies that in some areas of
parameter space model B describes, as model A, composite states.
However, because of the new extra parameter, $k_{0}$, the location of a near
threshold pole in the scattering amplitude is no longer directly linked to the
scattering length and as a consequence Weinberg's criterion for compositeness
can be evaded. This is intuitively clear: If the coupling to the confined
state is very weak or repulsive, it should appear like an elementary particle
in the scattering channel. This is the case for very small $\lambda$ and
negative $k_0^2$. If one takes, for instance, $\bar\lambda=0.1$ and
$k_0=0.3i$ one obtains a bound state with $Z\simeq 1$, which is
an elementary state.
In summary, for certain parameters, that need to be fine--tuned considerably, the ``Feshbach'' resonance model can describe a
bound state in a confined channel that couples to scattering in an open
channel that does not satisfy Weinberg' criterion for compositeness and thus
should be interpreted as a genuine (``ordinary'') state.
\section{A realistic example: Pole trajectories of the $\sigma$ and $\rho$ mesons
as a function of quark masses
}
\begin{figure
\begin{center}
\epsfig{file=poles.eps,width=0.5\textwidth}
\caption{ Movement of the $\sigma$ (dashed lines) and $\rho$ (dotted
lines) poles for increasing pion masses (direction indicated by the
arrows) on the second sheet as extracted from the IAM. The filled (open) boxes denote the
pole positions for the $\sigma$ ($\rho$) at pion masses $m_\pi=1,\
2,$ and $3 \times m_\pi^{\rm phys}$, respectively. Note, for
$m_\pi=3m_\pi^{\rm phys}$ three poles accumulate in the plot very near the
$\pi\pi$ threshold.}
\label{polosNormUpDown}
\end{center}
\end{figure}
\begin{figure}[t!]
\centering
\hbox{
\centering
\hspace{-1.9cm}
\includegraphics[scale=.9]{k0andGammaSigmaFit.delgado.eps}
\hspace{-4.5cm}
\includegraphics[scale=.9]{sigmakplane.delgado.eps}
}
\caption{Behavior of the $\sigma$ pole in the
$k$--plane. Left panel: $m_\pi$ dependence of $k_{\rm p}$ and $\gamma$.
The filled circles (boxes) show the results of
the numerical determination for $|k_p|$ ($\gamma$) from
the full calculation, while the lines are produced from
the fitting functions given in the text.
Right panel: the resulting pole movement for the $\sigma$
in the $k$--plane}
\label{sigink}
\end{figure}
\begin{figure}[t!]
\centering
\hbox{
\centering
\hspace{-1.9cm}
\includegraphics[scale=.9]{k0andGammaRhoFit.delgado.eps}
\hspace{-4.5cm}
\includegraphics[scale=.9]{rhokplane.delgado.eps}
}
\caption{Behavior of the $\rho$ pole in the
$k$--plane. Left panel: $m_\pi$ dependence of $k_{\rm p}$ and $\bar\gamma$.
The filled circles (boxes) show the results of
the numerical determination for $|k_p|$ ($\bar gamma k_p^2$) from
the full calculation, while the lines are produced from
the fitting functions given in the text.
Right panel: the resulting pole movement for the $\rho$
in the $k$--plane}
\label{rhoink}
\end{figure}
In order to illustrate with a realistic example
what was described in the previous
sections, we now show the results
for the pole trajectories of the $\rho$--meson and the $\sigma$--meson
calculated within the inverse amplitude method (IAM)~\cite{IAM}. The approach
uses Chiral Perturbation Theory (ChPT) predictions to a given order to fix the subtraction constants
of an elastic partial wave dispersion relation. This leads to
an amplitude consistent with elastic unitarity that by construction matches the ChPT
expansion when re-expanded at low energies and at the same time generates the poles associated to the $\sigma$ and $\rho$
resonances in pion-pion scattering~\cite{Dobado:1996ps}. Note that the numerical values of the
low energy constants obtained when fitting the IAM to scattering data
might slightly differ from those of ChPT since
they absorb higher order effects.
Since the whole QCD quark mass dependence is included up to the desired
order in terms of the ChPT expansion of the pion mass and decay constant,
one can study the quark mass dependence of both the $\sigma$ and $\rho$
resonances \cite{Hanhart:2008mx}. We will now discuss the resulting pole
trajectories is some more detail.
In Fig.~\ref{polosNormUpDown} we show the pole movement in the second sheet
for both $\sigma$ and $\rho$.
The pole movement of the $\sigma$ in the $k$--plane is shown in
the right panel of Fig.~\ref{sigink}.
Not only provides us the $k$--plane with a different look at the
positions and movements of $S$--matrix poles, it also allows us to
give a simple parameterization for the $m_\pi$--dependence of the
resonance poles shown in Figs.~\ref{polosNormUpDown} and ~\ref{sigink}.
Especially we get for the $\sigma$
\begin{equation}
\left(k_p^\sigma\right)^2=a_\sigma^2(b_\sigma^2-m_\pi^2) \quad \mbox{and} \quad \gamma^\sigma = \gamma_0^\sigma + c_\sigma(m_\pi/m_\pi^{\rm phys.})^2 \ ,
\end{equation}
\noindent
with $a_\sigma= 0.64$ MeV, $b_\sigma=320.8$ MeV, $c_\sigma=7.5$ MeV and $\gamma_0^\sigma=123$ MeV and analogously for the $\rho$
\begin{equation}
\left(k_p^\rho\right)^2=a_\rho^2(b_\rho^2-m_\pi^2) \quad \mbox{and} \quad \bar \gamma (k_p^\rho)^2 = \gamma_0^\rho + c_\rho(m_\pi/m_\pi^{\rm phys.})^2 \ ,
\end{equation}
\noindent
with $a_\rho= 0.75$ MeV, $b_\rho=480$ MeV, $c_\rho=-4.1$ MeV and $\gamma_0^\rho=44.1$ MeV.
A comparison of the fit functions and the full numerical results for the pole movements are
shown for the $\sigma$ and $\rho$ in the left panel of
Figs.~\ref{sigink} and \ref{rhoink}, respectively. We see
that for both $k_{\rm p}$ and $\gamma$ very simple two parameter fitting
functions provide a reasonable representation of the full results.
We start with the physical, non--vanishing values for both $\gamma$ and $k_{\rm p}$ for the
$\sigma$ as well as the $\rho$. As the pion mass gets increased
$k_{\rm p}$ decreases significantly and eventually vanishes while $\gamma$
changes relatively little.
At the point where $k_{\rm p}=0$ the two poles meet at the real axis below threshold
for the $\sigma$ and at exactly at threshold for the $\rho$, as explained
above. When the quark masses are increased further, one $\sigma$ pole moves
towards the $\pi\pi$ threshold, while the other one moves away from the
threshold
along the real $s$ (imaginary $k$) axis.
We can now come back to the discussion of Sec.~\ref{nature}
and apply the formalism to the $\sigma$ as derived from
the IAM.
In case of the $\sigma$ the range of forces is set by $m_\rho$. The $\sigma$
becomes a bound state at $m_\pi=450$ MeV. At this point we have
$$
\gamma \simeq \kappa \simeq 200 \ \mbox{MeV} \ \longrightarrow \ Z\simeq 0 \ .
$$
Thus we conclude from this analysis that at least for $m_\pi > 450$ MeV the
$\sigma$ is predominantly of molecular nature. Note that, both for simplicity and in order to
be conservative, we have
shown calculations for the IAM to one-loop from \cite{Hanhart:2008mx},
although the two-loop calculation has also been performed \cite{Pelaez:2010fj}.
In that case a similar behavior is found, including the appearance of a virtual pole, although
for pion masses $m_\pi>300$ MeV.
Given the large similarity of the pole trajectory of the $\sigma$ meson and that found for the controversial $K(800)$ scalar resonance
(or $\kappa$) with the IAM using SU(3) ChPT \cite{Nebreda:2010wv}, a similar conclusion seems unavoidable for the $K(800)$,
especially since the virtual pole predicted as the pion mass increases
was recently confirmed in a lattice-QCD calculation~\cite{Dudek:2014qha}.
\section{Summary}
In this paper we discussed on general grounds the properties of pole
trajectories as some strength parameter is varied for resonances coupling to
the continuum in different partial waves. There is a qualitatively different
behavior for states that couple in an $s$--wave compared to all higher
partial waves: only for $s$--wave states the two, complex conjugate resonance
poles on the second sheet meet at some value of the strength parameter below
the threshold --- for all other partial waves this meeting point is located
exactly at threshold. Using Weinberg's compositeness criterion we were able to
show that there is a connection between the value of the mentioned
subthreshold meeting point and the composition of the wave function of the
physical state. To illustrate the mentioned properties we investigated two
models: Model A gives hadronic
molecules, which one might also call extraordinary hadrons, for all values of the coupling that lead to a
pole, while the more general Model B
allows for a near threshold state with a prominent elementary
component, however, this requires a significant amount of fine tuning.
In lattice QCD resonance poles move as quark masses are varied. Since most
simulations at present are still performed at such values of the quark quark masses/lattice spacings
that the resonances can not decay to the continuum, so called chiral
extrapolations are necessary to relate the lattice results to the
real world parameters. For extraordinary $s$--waves those need to contain
striking non--analyticities. This is illustrated in this paper by
employing the quark mass dependence of the $\sigma$ pole as predicted
by the inverse amplitude method in combination with one loop chiral
perturbation theory. On the basis of this study we were also able to provide
a simple parameterization for the pole trajectories that contains the
mentioned non--analyticity and should proof useful in future lattice studies.
\section*{Acknowledgements}
We are particularly thankful towards R. L. Jaffe for his participation at early stages of this work.
The research was in part supported by the Spanish project FPA2011-27853-C02-02,
DFG funds to the Sino-German CRC 110 ``Symmetries
and the Emergence of Structure in QCD'' and CRC 16 "Subnuclear Structure of Matter", as well as the EU
I3HP ``Study of Strongly Interacting Matter'' under the Seventh Framework
Program of the EU.
|
\section{ Introduction}
The AdS/CFT correspondence \cite{Maldacena:1997re,Witten:1998qj,Gubser:1998bc,Aharony:1999ti} provides a window into the dynamics of strongly coupled systems by identifying the underlying field theory with a weakly coupled gravity dual. In recent years the methods and scope of the AdS/CFT correspondence have shifted from traditionally QCD-motivated problems to problems in the area of condensed matter systems (see reviews \cite{Hartnoll:2009sz,Herzog:2009xv,McGreevy:2009xe,Sachdev:2010ch}, and references therein). In particular, various models of holographic s-wave \cite{Hartnoll:2008vx,Hartnoll:2008kx} and p-wave \cite{Gubser:2008zu,Gubser:2008wv} superconductors have been constructed.
Among the various paradigms in condensed matter physics, disorder is a fundamental one as it provides a crucial step away from clean systems toward realistic ones. One striking manifestation of disorder in non-interacting quantum systems is the phenomenon of Anderson localization \cite{Anderson:1958vr}, where the conductivity can be completely suppressed by quantum effects. The study of the interplay between disorder and interactions in quantum systems has seen little progress on the theoretical side. Recently, however, in the context of disordered conductors, Basko, Aleiner and Altshuler presented compelling evidence in favor of a many-body localized phase, based on an analysis of the perturbation theory in electron-electron interaction to all orders \cite{Basko20061126}. Subsequent works (see \cite{PhysRevB.75.155111,PhysRevB.82.174411,PhysRevB.84.094203,0295-5075-101-3-37003} and references therein) have confirmed and sharpened the existence
of a phase transition separating the weakly and strongly interacting limits of electrons in disordered potentials.
Disorder is also particularly relevant in the context of superconductors; it has a rich history dating back to the pioneering work of Anderson in 1959 \cite{Anderson195926}. For many years Anderson's theorem, stating that superconductivity is insensitive to perturbations that do not destroy time-reversal invariance (pair breaking), provided the central intuition. Critiques to Anderson's argument were raised, for example, in \cite{doi:10.1143/JPSJ.51.1380,PhysRevLett.54.473,PhysRevB.33.3146,PhysRevB.32.5658} where the effects of strong localization were considered. More generally, the interplay between interactions and disorder in superconductors cannot be considered settled. In view of this situation, it makes sense to consider alternative models where the problem can be analyzed in full detail.
Indeed, in a previous work \cite{Arean:2013mta}, we initiated a program of directly studying the role of disorder in
holographic superconductors which arguably apply to strongly interacting superconductors. There have been other approaches
to disorder in holography
\cite{Hartnoll:2007ih, Hartnoll:2008hs,Fujita:2008rs,Ryu:2011vq,Adams:2011rj,Adams:2012yi,Saremi:2012ji}.
As in \cite{Arean:2013mta}, we follow a very direct approach to the realization of disorder by coupling an operator to a randomly distributed space-dependent source. Essentially, we directly translate a typical condensed matter protocol into the AdS/CFT framework. Namely, we choose a random space-dependent chemical potential by setting the boundary value of a $U(1)$ electric potential. The main rationale for this choice of disorder relies on the fact that the chemical potential defines the local energy of a charge carrier placed at a given position $x$, as it couples to the particle number $n(x)$ locally. Therefore, our choice of disorder
replicates a local disorder in the on-site energy. This is the simplest protocol one would implement.
Moreover, once disorder is introduced in such an strongly interacting system, all observables will become disordered and,
therefore, the physics is not expected to depend on the way disorder is originally introduced.
The direct approach outlined above has now been applied by other authors in the context of holography.
For example, it was used to argued for Anderson localization in \cite{Zeng:2013yoa}. Other interesting
applications include \cite{Hartnoll:2014cua} and \cite{Lucas:2014zea}.
It is worth mentioning that another very important motivation for our work is related to the more general and far-reaching problem of translational invariance in holography. Most holographic models respect translational invariance in the field theory directions. This underlying translational
invariance has adverse effects in applications involving transport properties in condensed matter. Since translational invariance implies momentum conservation, it means that the charge carriers have nowhere to dissipate their momentum, resulting
in a zero frequency delta function in the optical conductivity which obscures interesting questions
such as the temperature dependence of the DC resistivity. A lot of effort has recently been devoted to addressing this
shortcoming. Some progress has been reported in \cite{Hartnoll:2012rj,Sonner:2013aua,Herzog:2014tpa}.
Another approach to momentum
dissipation include models of massive gravity \cite{Vegh:2013sk}, \cite{Blake:2013owa},
\cite{Amoretti:2014zha,Amoretti:2014mma}.
However, this latter approach struggles with issues of UV completeness of the gravity models used.
The paper is organized as follows. In section \ref{Sec:ReviewPW}, we review the construction of the holographic p-wave
superconductor. In section \ref{Sec:Disorder} we introduce our implementation of disorder and present some typical results.
In section \ref{Sec:Thermo} we describe the different branches that emerge in our setup and determine which one wins the
thermodynamic competition by comparing the free energy. In section \ref{Sec:Phase} we present the disorder phase diagram.
Next, in section \ref{Sec:Correlated} we repeat the analysis for the case of noise with a non-flat power spectrum (which is correlated along the length of the system).
Section \ref{Sec:Spectrum} is devoted to the power spectra of the response. Namely we establish a fairly universal spectral
response for the charge density and the condensate as a function of the spectral description of the disordered chemical
potential. We conclude in section \ref{Sec:Conclusions} where we also point out some interesting directions.
\section{Review of the holographic p-wave superconductor}\label{Sec:ReviewPW}
To build a holographic $p$-wave superconductor in 2+1
dimensions we start with the action introduced originally in \cite{Gubser:2008zu} and further studied in
\cite{Gubser:2008wv}. Namely, we consider the dynamics of a $SU(2)$
Yang-Mills field in a gravitational background:
\bar{\varepsilon} S=\int d^4
x\,\sqrt{-g}\left(\frac{1}{16\pi G_N}\left(R-\Lambda\right)-{1\over4 q^2}Tr\,F_{\mu\nu}\,F^{\mu\nu} \right)\,.
\label{action}
\ee
In the limit where $G_N/q^2$ is very small, gravity can be considered decoupled and then, the Yang-Mills system is studied on the Schwarzschild-AdS
metric:
\begin{eqnarray}} \newcommand{\eea}{\end{eqnarray}
ds^2&=&{1\over z^2}\left(-f(z)dt^2+{dz^2\over
f(z)}+dx^2+dy^2
\right),\nonumber \\
f(z)&=&1-z^3\,,
\eea
where we have set the radius of AdS, $R=1$, and
the position of the horizon to $z_h=1$.
In \cite{Gubser:2008wv} an Ansatz was chosen such that the spatial
rotational symmetry is spontaneously broken when the condensate
breaking the gauge $U(1)$ symmetry subgroup arises at low temperatures.
The field strength in the action Eq. (\ref{action}) is given by
\bar{\varepsilon}
F^a_{\mu\nu}=\partial_\mu A^a_\nu -\partial_\nu A^a_\mu +f^a_{\,\,bc}A^b_\mu A^c_\nu\,,
\ee
and the corresponding Yang-Mills equations of motion:
\bar{\varepsilon}
\nabla_\mu F^{a \mu\nu}+f^a_{\,\,bc}A^b_\mu F^{c\mu\nu}=0\,.
\ee
In what follows we specialize to $SU(2)$ with the following generators
(further conventions are as in \cite{Amado:2013xya}):
\bar{\varepsilon}
\label{generators}
T_i=\frac{1}{2}\sigma_i\,,\hspace{1cm} \lbrace T_i,T_j \rbrace
=\frac{1}{2}\delta_{ij}\mathbb{I}\,.
\ee
It is possible to consider more general groups, for example $U(2)$, see \cite{Amado:2013aea,Amado:2013lia}.
Motivated by condensed matter applications and, in particular superconductivity,
we are interested in a system at finite chemical potential which develops an instability at low temperatures.
One simple Ansatz that achieves this goal is
\bar{\varepsilon}
A=\phi(z)dt\,T_3+w_x(z) dx \, T_1\,.
\ee
Following the AdS/CFT dictionary, one reads field theory information from the boundary data of the gravity fields.
Namely, the boundary ($z\to 0$) values of the fields are:
\begin{eqnarray}} \newcommand{\eea}{\end{eqnarray}
\label{homogeneous}
&&\phi(z)=\mu -\rho\,z+o(z^2)\,, \nonumber \\
&&w_x(z)=w_x^{(0)}+w_x^{(1)}\,z+o(z^2)\,.
\eea
The field theory interpretation in terms of these boundary values is as follows: $\mu$ is a chemical potential, $\rho$ is the charge
density, $w^{(0)}_x$ is the source and $w^{(1)}_x$ is the vacuum expectation value of the vector order parameter.
Since we are interested
in spontaneous symmetry breaking we will require the source to vanish $w^{(0)}_x=0$.
Notice that due to the rescaling that allowed us to set the horizon radius $z_h=1$, the chemical potential $\mu$ is actually
dimensionless and proportional to the ratio of the physical chemical potential to the temperature.
If one works in the grand canonical ensemble, where the chemical potential is held fixed, the temperature of the system is
thus given by $T\propto 1/\mu$. Hence in the rest of the paper we will only talk about $\mu$, with the understanding that it is
equivalent to the inverse of the temperature of the boundary field theory at fixed chemical potential.
An intuitive way of understanding the mechanism of condensation is as follows.
The gravity mode $w_x(z)$ has an effective mass of the form:
\bar{\varepsilon}
m_{eff}^2= q^2 \,\,g^{tt}\,\,\phi^2 \,.
\ee
Since $g^{tt}<0$, as we increase the value of $\mu$ the effective mass decreases and goes below the BF bound in a sufficiently
large region of space and, consequently, a zero mode of $w_x(z)$ develops at some $\mu_c$. Increasing $\mu$ above the
critical value, $\mu_c$, leads to the field condensing, and a new branch of solutions with nonzero condensate emerges.
This instability mechanism is fairly universal, appearing both in the $s$-wave \cite{Hartnoll:2008vx,Hartnoll:2008kx} and $p$-wave
\cite{Gubser:2008zu,Gubser:2008wv} holographic superconductors.
The asymptotic value of $\mu$ plays the role of chemical potential in the dual field theory. It is
worth mentioning that since the order parameter is determined by the asymptotic value of the field $w$ which is a vector,
we have a vectorial order parameter.
To complete the analogy with the superconducting phase transition
the conductivities were computed for this system \cite{Gubser:2008zu,Gubser:2008wv}, and qualitative agreement
was established.
More recent studies of this system, some taking into account the gravitational back-reaction,
include: \cite{Herzog:2014tpa,Basu:2009vv,Ammon:2009xh,Gangopadhyay:2012gx,Roychowdhury:2013aua,Arias:2012py}.
\section{ Holographic p-wave superconductor with disorder}\label{Sec:Disorder}
The main goal in this manuscript is the introduction of disorder in the $x$-direction of the field theory dual.
To be consistent with the
equations of motion, we are not allowed to choose the direction of the condensate freely as done in the previous section.
We, therefore, consider the following consistent Ansatz for the matter fields:
\begin{eqnarray}} \newcommand{\eea}{\end{eqnarray}
\label{Ansatz}
A=\phi(x,z)\,
dt\,T_3+w_x(x,z)\,T_1\, dx+ w_y(x,z)\,T_1\,dy + \theta(x,z)\,T_2\,dt
\,,
\eea
where $T_i$ are the $SU(2)$ generators presented before in Eq. (\ref{generators}).
The Yang-Mills equations of motion following from the Lagrangian (\ref{action}) and the Ansatz (\ref{Ansatz}) are:
\begin{eqnarray}} \newcommand{\eea}{\end{eqnarray}
&&-\partial_z^2\phi-\frac{1}{f}\,\partial_x^2\phi+\frac1 f (w_x^2+w_y^2)\phi+
\frac1 f \theta \partial_x w_x+
\frac2 f w_x\partial_x\theta =0\,,\label{eomphi}\\
&&\partial_z^2w_x+\frac{f'}{f}\partial_z w_x+\frac1{f^2}\left(\phi^2+\theta^2\right)w_x
+\frac1{f^2} \phi \partial_x\theta -\frac1{f^2} \theta\partial_x\phi =0\,,\label{eomwx} \\
&&\partial_z^2\theta+\frac1f\partial_x^2\theta-\frac1f (w_x^2+w_y^2) \theta
+\frac1{f}\phi\partial_x w_x+\frac2{f}w_x\partial_x\phi =0\,,\label{eomtheta} \\
&&\partial_z^2 w_y+\frac1f\partial_x^2 w_y+\frac{f'}{f}\partial_z w_y+
\frac1{f^2}\left(\phi^2+\theta^2\right)w_y=0\,.\label{eomwy} \eea
These equations of motion satisfy the constraint
\bar{\varepsilon}
\phi \partial_z
\theta-\theta\partial_z\phi-f\,\partial_z\partial_xw_x=0\,,
\ee
which is a consequence of gauge fixing: $A_z=0$.
{
As in the previous case, discussed around Eqs. (\ref{homogeneous}), to uncover the physics of the dual field theory
we need to examine the boundary values of the supergravity fields. The near boundary asymptotics of the solutions to equations (\ref{eomphi}-\ref{eomwy}) is given, at small values of $z$, by:
\begin{eqnarray}} \newcommand{\eea}{\end{eqnarray}
&&\phi(x,z)=\mu(x)-\rho(x)\,z+o(z^2)\,,\\
&&w_x(x,z)=w_x^{(0)}(x)+w_x^{(1)}(x)\,z+o(z^2)\,,\\
&&\theta(x,z)=\mu_2(x)-\rho_2(x)\,z+o(z^2)\,,\\
&&w_y(x,z)=w_y^{(0)}(x)+w_y^{(1)}(x)\,z+o(z^2)\,.
\eea
The values $\mu(x)$ and $\rho(x)$ correspond to space-dependent
chemical potential and charge density, respectively. Turning on a
chemical potential in the direction $T_3$ means breaking
$SU(2)\rightarrow U(1)_3$. The functions $w_i^{(0)}(x)$ and
$w_i^{(1)}(x)$ are identified, under the holographic duality, with the source
and VEV of vectorial operators in the $i$ direction. Finally, $\mu_2(x)$ and
$\rho_2(x)$ are, respectively, a new chemical potential and charge density that
are sourced by the space-dependent condensate.
The near horizon conditions on the gravity fields are completely determined by regularity of the solution. Regularity,
consequently, implies that $A_t$ vanishes at the horizon. Hence, we consider an asymptotic expansion
about $z\sim1$ of the form
\begin{eqnarray}} \newcommand{\eea}{\end{eqnarray}
&&\phi(x,z)=(1-z)\,\phi_h^{(1)}(x)+(1-z)^2\,\phi_h^{(2)}(x)+\ldots \,,\nonumber \label{phiIRexp} \\
&& \phi(x,z)=(1-z)\,\theta_h^{(1)}(x)+(1-z)^2\,\theta_h^{(2)}(x)+\ldots \,,\nonumber \label{thetaIRexp} \\
&&w_i(x,z)=w_{ih}^{(0)}(x)+(1-z)\,w_{ih}^{(1)}(x)+(1-z)^2\,w_{ih}^{(2)}(x)+\ldots\,, \label{wIRexp}
\eea
where the ellipses stand for higher order terms.
For numerical reasons, we find it convenient to redefine some of the fields involved in the equations of motion. Namely:
\bar{\varepsilon}
\chi_i(x,z) = (1-z)\,w_i(x,z)\,.
\label{chidef}
\ee
In terms of the redefined fields (\ref{chidef}) the equations (\ref{eomphi}-\ref{eomwy}) take the form
\begin{subequations}
\begin{eqnarray}} \newcommand{\eea}{\end{eqnarray}
&&\partial_z^2\phi+\frac{1}{f}\,\partial_x^2\,\phi
-{\chi^2\over(1-z)^2 f}\phi
-{1\over(1-z)\,f}\left(2\chi_x\,\partial_x\theta+(\partial_x\chi_x)\,\theta \right)=0\,,\label{eomrdphi}\\
%
&&\partial_z^2\theta+\frac1f\partial_x^2\theta
-{\chi^2\over f(1-z)^2} \theta
+{2\chi_x\,\partial_x\phi+(\partial_x\chi_x)\,\phi\over f(1-z)}=0\,,\label{eomrdtheta} \\
&&\partial_z^2\chi_x+\left(\frac{f'}{f}+\frac2{1-z}\right)\partial_z\chi_x
+{2f^2+(1-z)\,f\,f'+(1-z)^2(\theta^2+\phi^2))\over(1-z)^2\,f^2}\chi_x+\nonumber \\
&&+{1-z\over f^2}\left(\phi\,\partial_x\theta-(\partial_x\phi)\,\theta\right)
=0\,,\label{eomrdwx} \\
&&\partial_z^2 \chi_y+\frac1f\partial_x^2 \chi_y+\left(\frac{f'}{f}+\frac2{1-z}\right)\partial_z
\chi_y+
\left[
{2\over (1-z)^2}+{1\over f^2}(\theta^2+\phi^2)+{f'\over f\,(1-z)}
\right]\chi_y=0\,,\label{eomrdwy}\nonumber\\ \eea
\label{pderw}
\end{subequations}
where
$\chi^2=\chi_x^2+\chi_y^2$.
This redefinition leads to a simpler set of boundary conditions:
\begin{eqnarray}} \newcommand{\eea}{\end{eqnarray}
\chi_i(x,0)&=&0,\qquad \theta(x,0)=0\,, \qquad \phi(x,0)=\mu(x)\,,\; {\rm UV}\,\,\, z\to 0\,, \nonumber \\
\chi_i(x,1)&=&0\;,\qquad \theta(x,1)=0\,,\qquad
\phi(x,1)=0\,,\;\;\;\;\; {\rm IR}\,\,\, z\to 1\,.\label{bcirpsi}
\eea
This choice of boundary conditions corresponds to a spontaneous
breaking of the $U(1)$ symmetry with order parameter $\langle {\cal
O}\rangle\propto w^{(1)}_i(x)$. From now on we use the angle brackets
associated with ${\cal O}$ exclusively to refer to the average over $x$.
Moreover, we choose the condition $\mu_2=0$ (vanishing source for the charge
density in the $T_2$ direction) since we want a disordered
version of the p-wave superconductor. Hence, the charge density $\rho_2$
will be spontaneously induced. In this case, since the symmetry of our
action (\ref{action}) is the whole $SU(2)$, we are effectively realizing a two-component superfluid as recently discussed in the holographic framework in \cite{Amado:2013xya} following original ideas of \cite{PhysRevB.11.178},
an important
phenomenological paradigm in various condensed matter situations. We will however not pursue these questions in the present manuscript.
It is worth noticing that there are two simplified situations
that might be taken into account, given that they might include all the
interesting physics but require less computing power. It is easy to
see that the equations (\ref{eomphi}-\ref{eomwy}) allow to consistently set
$w_x=\theta=0$ or $w_y=0$. This is equivalent to going to the two
limits in which the system condenses in the direction of the noise
or in the direction perpendicular to the noise. We will devote ample attention to these branches in section \ref{Sec:Thermo}.
\subsection{Introducing disorder}
\label{Subsec:disorder}
We are interested in solving the system given by equations (\ref{pderw})
in the presence of disorder. Let us take the following form for the noisy chemical potential:
\begin{eqnarray}} \newcommand{\eea}{\end{eqnarray}
\nonumber
\mu(x)&=&\mu_0+\epsilon\sum_{k=k_0}^{k_*}{\sqrt{S_k}}\,\cos(k\,x+\delta_k)\\
&=&\mu_0+\epsilon\sum_{k=k_0}^{k_*}{1\over
k^\alpha}\,\cos(k\,x+\delta_k)\,, \label{noisefunc}
\eea
where $\delta_k$ is a random phase for each $k$ and $S_k$ is the power
spectrum.
For the case $\alpha=0$, corresponding to a flat spectrum, in the limit of infinitely many modes ($k_*\to\infty$)
the function in Eq. (\ref{noisefunc}) tends to a Gaussian distributed random function.
Instead, as we will see below, for $\alpha>0$, the typical length scale of the noise is of the order of the system size.
In this latter case, the power of $1/k$ determines the differentiability properties of $\mu(x)$.
Let us now discuss, in detail, the properties of the correlation function of the disorder we introduce above.
It takes the form
\bar{\varepsilon}
\left< \mu(x)\mu(0) \right>-\mu_0^2=\sum_{k=k_0}^{k_*}\frac1{k^{2\alpha}}\cos(kx)\,.
\label{corrfunc}
\ee
For $\alpha>0$ this correlation function is periodic and does not diverge for large $x$.
Moreover, one could try and go to the continuum limit, replacing the sum in Eq. (\ref{corrfunc}) by an integral over $k$. The corresponding expression for the correlation function is then given by
\bar{\varepsilon}
\left< \mu(x)\mu(0) \right>-\mu_0^2\approx |x|^{2
\alpha-1}\Gamma(1-2\alpha)\sin(\alpha\, \pi)+
\frac{k_0^{1-2\alpha}}{2\alpha-1} \,
_1F_2\left(\frac12-\alpha;\,\frac12,\, \frac32-\alpha ;\, -\frac14
k_0^2 x^2\right)\,.
\ee
It is clear from this expression that thanks to the IR cutoff $k_0$, no divergence appears in the correlation function.
Note, however, that the first term grows with distance, the second term kills that divergence and the result is a damped oscillatory correlation function.
More explicitly, the large $x$ limit of the expression above reads
\bar{\varepsilon}
\left< \mu(x)\mu(0) \right>-\mu_0^2\rightarrow \frac{k_0^{-2
\alpha}}{x}\cos(k_0 x)+\dots\,,
\label{xilargex}
\ee
The power law decay of the correlation function, reminiscent of a critical system, does not allow us to define a correlation length by the usual prescription. Lacking this, the oscillations define a length scale $\propto 1/k_0$, which could be as large as the system size.
Therefore, in a broad sense we can say that for $\alpha>0$ our noise is correlated along the whole system; we will denote this case as correlated noise.
Low correlation in the noise can be achieved by taking $\alpha=0$ in (\ref{noisefunc}), which results in a correlation function of the form
\bar{\varepsilon}
\left< \mu(x)\mu(0) \right>-\mu_0^2=\sum_{k=k_0}^{k_*}\cos(k x)
={\rm Re}\left(
e^{ik_0\,x}\,{e^{i(k_*-k_0+1)\,x}-1\over e^{ix}-1}
\right)\,.
\label{correlength}
\ee
Notice that this function, which we plot in Fig. \ref{corrfig},
is the closest to a delta function one can get with a finite number of modes.
Hence, in order to study a more realistic realization of noise, in the bulk of the work presented in this article we consider the case where $\alpha=0$, which we will denote as uncorrelated noise.
However, the correlated noise ($\alpha>0$) presents interesting features, such as a well defined continuum limit.
Moreover, as in \cite{Arean:2013mta}, it will allow us to study the power spectra of the response functions in our setup. Therefore, in sections \ref{Sec:Correlated} and \ref{Sec:Spectrum} we will analyze the case of correlated noise
($\alpha>0$).
To implement the noise given by Eq. (\ref{noisefunc}) and solve the coupled PDEs (\ref{pderw}),
we discretize the space and impose periodic boundary conditions in the $x$ direction, leading to a discretized
$k$ with values:
\bar{\varepsilon}\label{Eq:k-range}
k_n={2\pi\,n\over L}\quad {\rm with}\quad 1\leq n\leq{L\over 2a_x}\,,
\ee where $L$ is the length in the $x$ direction of our cylindrical
space, and $a_x$ is the lattice spacing in $x$. Note that there is an
IR scale given by $k_0$ and a UV scale defined by $k_*$ \footnote{Notice that both these two scales, as well
as the chemical potential $\mu$ are measured in terms of the temperature, since we have made use of the scaling symmetries of the problem to fix the horizon of the black hole at $z_h=1$.}.
Notice that the UV scale $k_*$ is given by $L/(2a_x)$, and
was chosen here to saturate the Nyquist limit\footnote{Nyquist frequency is the highest frequency that can be reconstructed from a signal given a sample rate. In order to recover all Fourier components of a periodic waveform, it is necessary to use a sampling rate at least twice the highest waveform frequency.
This can be understood from the fact that there are two Fourier coefficients to fit for each frequency.}.
However, as we explain below, when performing our numerical simulations we will take a $k_*$ sensibly smaller than
$L/(2a_x)$, in order to allow for our lattice to be sensible to higher harmonics sourced by our noise.
In order to parametrize the strength of the noise, which in Eq. (\ref{noisefunc}) is characterized by $\epsilon$, let
us introduce the variable $w$ defined through the expression
\bar{\varepsilon}
w={25\epsilon\over\mu_0}\,,
\label{wdef}
\ee
so that $w$ corresponds to a strength relative to the chemical potential $\mu_0$\footnote{The factor ${1\over25}$
is included to keep the same normalization as in \cite{Arean:2013mta}}.
Naturally, $w=0$ corresponds to the homogeneous case, while
the largest $w$ will be chosen by demanding that $\mu(x)$ remains positive all along the system.
Notice that this maximum value of $w$ will depend on the scales $k_0$ and $k_*$, and for the case of correlated noise also on the power $\alpha$ characterizing the power spectrum.
Our definition of $w$ corresponds, in the standard solid
state notation, to $1/k_F\, l$, where $k_F$ is the Fermi momentum and $l$ is the mean free path \cite{phillips2012advanced}.
\subsubsection*{Numerical Methods}
In order to solve the system of PDEs we have discretized it on a rectangular lattice of size $N_z\times N_x$, where $N_z$
and $N_x$ correspond, respectively, to the number of points in the $z$ and $x$ directions.
We used planar lattices for the $x$ direction, and Chebyshev grids along $z$.
Consequently, the
discretization of the derivatives was performed using pseudo spectral methods (with periodic boundary conditions
in the $x$ direction). To find solutions we employed a Newton-Raphson algorithm on lattices with a typical
size of $25\times 90$.
\begin{figure}[htp]
\begin{center}
\includegraphics[width=3.5in]{corr.pdf}
\caption{\label{corrfig} Plot of the correlation function of our chemical potential, Eq. (\ref{correlength}), for the parameters used in our numerical simulations: $L=2\pi$, $k_*=21$.}
\end{center}
\end{figure}
In the simulations performed to determine the phase diagram of our setup we set the system length to $L=2\pi$ and take $k_*=21$ and $\alpha=0$. We are therefore truncating the sum in Eq. (\ref{noisefunc}) at 21 modes. Notice that for a lattice with 90 points along the $x$ directions, saturating the Nyquist limit would correspond to a $k_*=45$, and hence 45 modes in the sum (\ref{noisefunc}).
For this parameters, from Eq. (\ref{correlength}) we can establish a correlation length
about $2.3\%$ of $L$ (the system size) or 2 lattice spacings.
Finally, in Fig. \ref{corrfig} we plot the correlation function (\ref{correlength}) for this choice of parameters.
\vspace{.5cm}
In Figs. \ref{ModelSimulationa0}, \ref{ModelSimulationa02} we plot the results of a single typical simulation. In Fig. \ref{ModelSimulationa0} we present the random chemical potential $\mu(x)$ (upper left panel) together with the solutions
for the fields $\phi$, $\theta$, $\chi_x$ resulting from solving the system (\ref{pderw}) with that chemical potential as boundary condition.
Instead, in Fig. \ref{ModelSimulationa02} we present the boundary data read from this same solution.
First, notice that the plots presented in these figures correspond to a solution for which the condensate lies purely in the $x$ component of the vector parameter. We will comment on
the competition between different branches of solutions (with condensate parallel or perpendicular to the direction of the noise) in section \ref{Sec:Thermo}.
As expected, the introduction of disorder leads to
a space-dependent charge density and condensate in some cases which we plot in those figures.
In a sense one could view the gravity equations of motion as a tool that provides precise answers to the question: Given a random chemical potential in a strongly coupled system with a superconducting transition, what is the value of the condensate and the charge density that the system uses to respond to the random chemical potential?
\begin{figure}[tb]
\begin{center}
\begin{subfigure}[b]{0.40\textwidth}
\includegraphics[width=\textwidth]{sim2da0_mu.pdf}
\end{subfigure}
~
\begin{subfigure}[b]{0.40\textwidth}
\includegraphics[width=\textwidth]{sim3da0_fi.jpg}
\end{subfigure}\\[4mm]
\begin{subfigure}[b]{0.40\textwidth}
\includegraphics[width=\textwidth]{sim3da0_th.jpg}
\end{subfigure}
~
\begin{subfigure}[b]{0.40\textwidth}
\includegraphics[width=\textwidth]{sim3da0_wx.jpg}
\end{subfigure}
\caption{Example of a simulation corresponding to a noisy chemical potential with
$w=2.95$, $\alpha = 0$, and $\mu_0=3.55$ below the critical value of the homogeneous chemical potential
($\mu_c=3.66$). We plot the original spatial-dependent chemical potential on the upper-left panel and the
corresponding solutions for the fields $\phi$, $\theta$ and $\chi_x$ ($\chi_y=0$ for this solution)
on the other three panels.} \label{ModelSimulationa0}
\end{center}
\end{figure}
\begin{figure}[tb]
\begin{center}
\begin{subfigure}[b]{0.40\textwidth}
\includegraphics[width=\textwidth]{sim2da0_mu.pdf}
\end{subfigure}
~
\begin{subfigure}[b]{0.40\textwidth}
\includegraphics[width=\textwidth]{sim2da0_rho.pdf}
\end{subfigure}\\[4mm]
\begin{subfigure}[b]{0.40\textwidth}
\includegraphics[width=\textwidth]{sim2da0_rho2.pdf}
\end{subfigure}
~
\begin{subfigure}[b]{0.40\textwidth}
\includegraphics[width=\textwidth]{sim2da0_wx.pdf}
\end{subfigure}
\caption{Boundary data corresponding to the same simulation as in Fig.
\ref{ModelSimulationa0}. From the upper left panel and in clockwise sense:
$\mu(x)$, $\rho(x)$, $w_x(x)$, and $\rho_2(x)$.} \label{ModelSimulationa02}
\end{center}
\end{figure}
Notice that the response to the source $\mu(x)$ is noisier for the VEV corresponding to the same field, that is, $\rho(x)$; while the VEVs realized by
other fields, namely $w_x(x)$ and $\rho_2(x)$, are visibly smoother. We will investigate this behavior in more detail in section \ref{Sec:Spectrum}.
It is worth pointing out, and we will use this result in the upcoming sections, that we find that for a chemical potential below the critical (in the homogeneous case), there are values of the strength of the disorder that render the expectation value everywhere non-vanishing; this is our definition of arriving at the superconducting phase via disorder.
\subsection{Thermodynamic limit and self-averaging condensate}
In the context of condensed matter physics it is quite important to consider the thermodynamic limit.
Namely, to study the properties of the system in the limit where its size goes to infinity.
The thermodynamic limit plays a particularly important role in systems dominated by quenched randomness.
In this subsection we address some issues related to the thermodynamic limit as they apply to the the problem at hand.
Let us review, once more, all the scales involved in the problem, both, physical and numerical.
The three physical dimensionless scales of the problem are $T/\mu$, $w$, and $L\,mu$ corresponding respectively to
the temperature, the strength of the noise, and the length of the system in the $x$ direction, all measured relative to
the chemical potential.
Since we are solving the problem numerically using a lattice we have two extra scales: $a_x$ and $a_z$ which are the sizes
of cells along the $x$ and the $z$ directions.
Discussing the thermodynamic limit is more than a mere academic question.
Given that the realization of disorder is intrinsically related to our way of solving the system we need to show that there
is a limit to which we are truly approximating. In general, we expect that in the thermodynamic limit certain quantities
will be self-averaging. A property $X$ is self-averaging if most realizations of the randomness have the same value of $X$.
More precisely, in the numerical context we use that: The system is said to be self-averaging with respect to property $X$ if
\bar{\varepsilon}
\frac{<X_n^2>-<X_n>^2}{<X_n>^2} \to 0,
\ee
as the size, $n$, of the system goes to infinity.
Here the angular brackets denote averages over the realizations of the quenched randomness of the system and $X_n$ is the value
of property $X$ when the system has size $n$ \cite{0305-4470-35-19-303,Binder:1986zz}.
\begin{figure}[tp]
\begin{center}
\includegraphics[width=6.5in]{thermo_lim.pdf}
\caption{\label{Fig:LogLog} On the left panel we plot the average of the condensate versus the number of lattice
sites $N_x$, notice that it stabilizes in the thermodynamic limit.
The right panel shows the variance of the condensate versus $N_x$, and the blue dashed line shows the fit
$\log({\rm var}(w_x)/\langle w_x\rangle^2) =-0.90-3.03\,\log(N_x)$. These figures show that the condensate $w_x$ is self-averaging. The data results from averaging over 50 realizations of a noisy chemical potential with $\alpha=0$,
$\mu_0=4.20$, and $w$ such that $w^2=160/N_x$ (so that the variance of the noisy chemical potential is kept
constant as $N_x$ is increased). }
\end{center}
\end{figure}
We will define our thermodynamic limit as the limit in which the correlation length of the disorder is negligible with respect to the length of the system.
In order to do so we will work with flat spectrum noise, which corresponds to setting $\alpha=0$ in Eq.
(\ref{noisefunc}).
Then
the scale $a_x$ (wich sets $k_{*}=\pi/a_x$) will determine the correlation length
of the disorder in our lattice.
Increasing the number of points $N_x$ of the lattice in the $x$ direction while keeping
the length of the system fixed will now imply to decrease the correlation length of the noise with respect
to the size of the system.
Using this prescription, we now study the self-averaging property of the condensate in the $x$-direction, $w_x$.
We claim that the average of the condensate stabilizes, while its variance goes to zero. Moreover,
we provide numerical evidence that this vanishing goes as a power law $\sim N_x^{-3}$ (see Log-Log plot in
Fig. \ref{Fig:LogLog}).
\begin{figure}[tp]
\begin{center}
\includegraphics[width=6.5in]{thermo_lim2.pdf}
\caption{\label{Fig:LogLogrho} On the left we show the average of the charge density $\rho$ versus the number of lattice sites $N_x$, while on the right we plot the variance of $\rho$ versus $N_x$. The blue dashed line results from the fit $\log({\rm var}(\rho)/\langle\rho\rangle^2) =-2.92-3.13\,\log(N_x)$. These plots follow from the same set of
data as those in Fig. \ref{Fig:LogLog}.}
\end{center}
\end{figure}
The same procedure as described above can be applied to study the self-averaging property of the charge density, $\rho$.
The result is presented in figure \ref{Fig:LogLogrho}, which clearly shows that the charge density is indeed self-averaging
in the thermodynamic limit.
\section{Free energy and competing solutions}\label{Sec:Thermo}
As already anticipated when we wrote the system of equations in section \ref{Sec:Disorder},
there are different branches or consistent truncations of the system of equations (\ref{pderw}).
One could expect three main types of solutions:
\begin{itemize}
\item Solutions with $\chi_x\equiv 0$ and also $\theta\equiv 0$, we will denote these solutions by $Y$, since the
vector order parameter lies in the direction transverse to the noise.
Note that in this approximation the system (\ref{pderw}) reduces significantly.
In particular, equations (\ref{eomrdtheta}) and (\ref{eomrdwx}) are identically satisfied.
\item Solutions where $\chi_y\equiv 0$, we denote them by $X$, as they correspond to a vector condensate along the $x$
direction. In this limit the system of equations (\ref{pderw}) leads to a system with equation (\ref{eomrdwy})
trivially satisfied.
\item A third possibility would be that of solutions where all functions in the system (\ref{pderw})
are nonzero. These would correspond to a vector condensate pointing along an intermediate direction in the $x\,y$ plane.
However, our numerics indicate that these solutions do not exist.
\end{itemize}
Let us elaborate a bit more about the absence of solutions where the condensate lies along an intermediate direction in the
$x\,y$ plane. As is clear from the equations, in the absence of noise the system in the normal phase is rotational invariant.
Therefore, symmetry-breaking solutions with the condensate along any arbitrary direction on the plane are equivalent.
However, as soon as some noise is turned on, our numerics converge to solutions with the condensate being either parallel or
orthogonal to the noise. We checked this fact by starting from a broad family of seeds.
We will now study the free energy of the $X$ and $Y$ solutions to decide which of them is energetically favorable.
The free energy of the system is given in terms of the on-shell action (\ref{action}) as
\begin{eqnarray}} \newcommand{\eea}{\end{eqnarray}
\Omega&=&- \frac{T S_{\rm on-shell}}{ L_y\,L}= \\
&=&-\frac1{4L}\int_0^L dx \,\mu\, \rho+\frac1{4L}\int_0^L dx\int_0^1
dz\,\frac1f\left[(\theta^2+\phi^2)\,(w_x^2+w_y^2)+w_x (\phi\, \partial_x
\theta-\theta\,\partial_x\phi)\right]\,,\nonumber
\eea
where $L_y$ is the length of the system in the $y$ direction; this is a regulator
we need in order to get a finite result and will simply cancel out when integrating along the $y$ direction since
the solutions are $y$ independent.
In Fig. \ref{SvW} we plot the free energy for the two kinds of superconducting solutions, subtracted from
that of the normal phase\footnote{In the normal phase $\theta=\chi_x=\chi_y=0$, and the system (\ref{eomrdphi}-\ref{eomrdwy}) reduces to the equation (\ref{eomrdphi}). The normal phase solution exists for all values of $\mu$.}.
We observe that, when it exists, the $X$ solution has always lower free energy.
Therefore, in the rest of the paper when we refer to the superconducting phase we will restrict ourselves to the X
solutions, namely those with condensate pointing in the direction parallel to the noise.
\begin{figure}[htp]
\begin{center}
\includegraphics[width=3.0in]{FE38a0.pdf}
\includegraphics[width=3.0in]{FE355a0.pdf}
\caption{\label{SvW} Subtracted free energy of competing solutions as a function of the strength of
the disorder, $w$. The blue (black) solid (dashed) line corresponds to the free energy of the X
(Y) solution subtracted from that of the normal phase. On the left panel we present the results for an average
chemical potential $\mu_0=3.8$ above the critical one. The plot on the right corresponds to $\mu_0=3.55$ below
the critical $\mu_c=3.66$. These plots result from averaging over 5 realizations on lattices of
size $25\times90$.}
\end{center}
\end{figure}
We shall now provide a heuristic explanation for the fact that the $X$ solution
is energetically favorable (see \cite{Basu:2008st,Erdmenger:2013zaa} for similar arguments).
The key observation follows from comparing the equations for the fields $w_x$ and $w_y$, namely
Eqs. (\ref{eomwx}-\ref{eomwy}), which we reproduce for convenience
\begin{eqnarray}} \newcommand{\eea}{\end{eqnarray}
&&\partial_z^2w_x+\frac{f'}{f}\partial_z w_x+\frac1{f^2}\left(\phi^2+\theta^2\right)w_x
+\frac1{f^2} \phi \partial_x\theta -\frac1{f^2} \theta\partial_x\phi =0\,, \\
&&\partial_z^2 w_y+\frac{f'}{f}\partial_z w_y+
\frac1{f^2}\left(\phi^2+\theta^2\right)w_y+\frac1f\partial_x^2 w_y=0\,,
\eea
We can, for example, consider that the mode $w_y$ is governed by an effective mass of the form:
\bar{\varepsilon}
m^2_{w_y}\propto -\frac1{f^2}\left(\phi^2+\theta^2\right)-\frac1{w_y f}\partial_x^2 w_y.
\ee
As noticed already in \cite{Erdmenger:2013zaa} the last term contributes a positive amount to the effective mass and
impedes condensation\footnote{Notice that at linear level one could Fourier transform the equation and consider the
effect of a single wave by replacing $\partial_x$ by $i\,k$.}.
Note, however, that the situation is different for the equation describing the effective mass for
the mode $w_x$.
\bar{\varepsilon}
m^2_{w_x}\propto -\frac1{f^2}\left(\phi^2+\theta^2\right)-\frac1{w_x f^2} \left(\phi \partial_x\theta -
\theta\partial_x\phi\right)\,.\label{meffwx}
\ee
There is no term $\sim \partial_x^2 w_x$ impeding condensation, and it can be argued that the last term in
the above effective mass is small, since the derivatives cancel at linear level.
Thus, condensation of the mode $w_y$ seems to be disfavored while for the mode $w_x$ it is not.
Then, one may expect the free energy of the $X$ solution to be lower than that of the $Y$ solution,
as our numerics show.
Finally, let us speculate on the consequences the outcome of this free energy computation may have
for more realistic systems with bidimensional inhomogeneities.
Since it turns out that the solutions with condensate parallel to the noise are always thermodynamically preferred,
one would expect that in the presence of disorder in two spatial directions the condensate would point in the stiffest
direction, thus following the gradient of the bidimensional noise.
\section{Toward the disordered phase diagram}\label{Sec:Phase}
In this section we present the phase diagram of the disordered holographic p-wave superconductor.
The key strategy is to repeat the simulations outlined in section \ref{Sec:Disorder} with a random chemical potential $\mu(x)$ given by Eq. (\ref{noisefunc}) in the regime illustrated by figure \ref{corrfig}, and to do that enough times so that we develop meaningful statistics.
\begin{figure}[htb]
\begin{center}
\includegraphics[width=0.95\textwidth]{phasediaga0.pdf}
\caption{\label{Cx-v-N} Spatial average of the condensate as a function of the strength of disorder.
Each line corresponds to an average over 25 realizations of noise (with $\alpha = 0$) on a lattice of
size $25\times90$.
The value of the condensate grows with increasing disorder strength, $w$.
Each line corresponds to a value of $\mu_0$ as indicated on the legend, but for the black dashed line which,
as explained in the text, marks the cut off used to define the critical temperature.}
\end{center}
\end{figure}
One of our main results is the dependence of the condensate on the strength of the disorder, $w$, presented in Fig.
\ref{Cx-v-N}.
The results are qualitatively similar to the s-wave results \cite{Arean:2013mta}.
As there, we observe that the average of the condensate grows as the strength of the disorder is increased. Moreover, for chemical potentials below the critical one, strong enough disorder drives the system into a phase where the average of the condensate is non-vanishing.
To make direct contact with the condensed matter literature we propose a disordered phase diagram in Fig. \ref{Tc}, where we track the value of the critical temperature of the normal to superconductor phase transition as a function of the strength
of the disorder. Let us explain, for the benefit of clarity how we have proceeded.
Since the value of the condensate increases with the the strength of the disorder, we have determined a value of the condensate above which we
consider the system in the superconducting phase (black dashed line in Fig. \ref{Cx-v-N}). We then read the average chemical potential\footnote{We refer to the average over realizations, not to be confused with the spatial average.} and
use the fact that the only relevant scale is $\mu/T$ to determine the critical temperature. Let us advance a potential
criticism to out method. Clearly, it would have been more relevant to compute the conductivities and determine the phase
diagram based on a conductivity criterion \cite{futureUS};
we expect, as in all previous cases, that there is a direct relation between the existence of a condensate and the transport
properties of the holographic solutions. One important aspect of Fig. \ref{Cx-v-N} is its robustness. Namely, the precise
form of the phase diagram varies quantitatively depending on where precisely we draw the cut off line defining the
``appearance'' of a nonzero condensate. However, qualitatively it is clear from the plot that the conclusions are stable
with respect to parallel shifts of the position of this cut off line.
\begin{figure}[htp]
\begin{center}
\includegraphics[width=3.5in]{tcplota0.pdf}
\caption{\label{Tc} Enhancement of $T_c$ with the noise strength $w$ ($T_c^{w=0}$ stands for the critical
temperature in the absence of disorder). For values of $w$ to the right of the red dashed line the chemical potential becomes negative at its minimum.}
\end{center}
\end{figure}
Finally, let us try and explain the mechanism behind this enhancement of the critical temperature.
Looking at Eq. (\ref{meffwx}) it is not evident that the noise would enhance condensation by lowering the effective mass.
Actually the effect of the noise on the average (along $x$) of that effective mass is almost negligible.
However, the noise does have the effect of producing regions (in the $x$ direction) where the effective mass is below the
critical value for condensation. When these regions are large enough they trigger the condensation, resulting in solutions
where the condensate is nonzero along the whole sample (see Fig. \ref{ModelSimulationa02}); even in the regions where the chemical potential is below its critical value (for the homogeneous case) the condensate
is nonzero.
\section{Correlated noise}\label{Sec:Correlated}
In this section we shall analyze the case of correlated noise, namely that when in Eq. (\ref{noisefunc}) we consider
$\alpha\neq 0$.
Although, as we have seen in Eq.(\ref{xilargex}), this noise is correlated along the whole system, it is still worth looking at its effect on the condensate,
and check if the main features of the response of the system are similar to those of the s-wave case studied in \cite{Arean:2013mta}.
Let us first specify the choice of parameters for our simulations. We will be using the function (\ref{noisefunc}) to implement a noisy chemical potential,
setting the power spectrum $\alpha=1.5$, the system length $L=2\pi$, and the UV cut off $k_* = 1/a_x$ saturating
the Nyquist limit. We again parametrize the strength of the noise in terms of $w={25\epsilon\over\mu_0}$, and
restrict $w$ to values for which the chemical potential $\mu(x)$ stays positive along the whole system.
We will run our simulations in lattices of size $N_z\times N_x=25\times75$, and use the numerical methods described in section \ref{Subsec:disorder}.
\begin{figure}[hbt]
\begin{center}
\begin{subfigure}[b]{0.40\textwidth}
\includegraphics[width=\textwidth]{sim2da15_mu.pdf}
\end{subfigure}
~
\begin{subfigure}[b]{0.40\textwidth}
\includegraphics[width=\textwidth]{sim2da15_rho.pdf}
\end{subfigure}\\[4mm]
\begin{subfigure}[b]{0.40\textwidth}
\includegraphics[width=\textwidth]{sim2da15_rho2.pdf}
\end{subfigure}
~
\begin{subfigure}[b]{0.40\textwidth}
\includegraphics[width=\textwidth]{sim2da15_wx.pdf}
\end{subfigure}
\caption{\label{ModelSimulationa15}
Example of a simulation corresponding to a noisy chemical potential with
$w=3.50$, $\alpha = 1.50$, and $\mu_0=3.50$ below the critical value of the homogeneous chemical potential
($\mu_c=3.66$). From the upper left panel and in clockwise sense: $\mu(x)$, $\rho(x)$, $w_x(x)$,
and $\rho_2(x)$.}
\end{center}
\end{figure}
In figure \ref{ModelSimulationa15} we present a typical example of a simulation for a solution with correlated noise and chemical
potential $\mu_0=3.50$ below the critical value.
As one can see, the main features of this solution are similar to those of that in Fig. \ref{ModelSimulationa02} for the uncorrelated noise,
although some of the effects of the disorder are more pronounced in this case.
\begin{figure}[htb]
\begin{center}
\includegraphics[width=3.0in]{FE_34.pdf}
\includegraphics[width=3.0in]{FE_38.pdf}
\caption{\label{SvWcorr} Free energy of competing solutions as a function of the strength of the disorder, $w$.
We consider values of $\mu_0$ below and above the critical one.
The left panel corresponds to $\mu_0=3.4<\mu_c$, and the right one to $\mu_0=3.8>\mu_c$ (both with $\alpha=1.50$). The black dashed line corresponds to the normal phase solution, the blue dot-dashed line to the $Y$ solution, and the red solid line to the $X$ solution.
These plots result from averaging over 5 realizations on lattices of size $22\times40$.}
\end{center}
\end{figure}
For example, the difference in free energy among the different branches is
visible directly in the graphs in Fig. \ref{SvWcorr}, where we plot the free energy of the X and Y condensed solutions plus that of the normal phase. Another interesting observation following from that figure is that
for the panel corresponding to
$\mu_0=3.4<\mu_c$, all solutions coincide up to a noise strength $w\sim 6$. This reflects the fact that
for $w < 6 $ only the solution corresponding to the normal phase exists.
As in section \ref{Sec:Phase}, by repeating our simulations for different values of the average chemical potential we can study the dependence of the condensate on the strength of the disorder.
The results are plotted in Fig. \ref{Cx-v-Ncorr}, and are qualitatively similar to those for the s-wave superconductor with correlated noise, presented in \cite{Arean:2013mta}. There are, however, some differences.
\begin{figure}[hbt]
\begin{center}
\begin{minipage}{.6\linewidth}
\includegraphics[width=\linewidth]{phasediaga15.pdf}
\end{minipage}
\begin{minipage}{.39\linewidth}
\vspace{.4cm}
\includegraphics[width=\linewidth]{tcplota15.pdf}
\end{minipage}
\caption{\label{Cx-v-Ncorr} On the left we plot the spatial average of the condensate as a function of the
strength of disorder. Each line corresponds to an average over 10 realizations of noise (with $\alpha = 1.50$)
on a lattice of size $22\times40$.
Each line corresponds to a value of $\mu_0$ as indicated on the legend, but for the black dashed
line which marks the cut off used to define the critical temperature. On the right panel is shown the resulting
phase diagram.}
\end{center}
\end{figure}
First, the saturation for large values of $\mu_0$ is different as it seems that the condensate takes relatively larger values
of the average chemical potential with respect to the critical one to stabilize. This observation is somewhat marred by the
fact that, in general, going to very high values of the chemical potential leads to a region where back-reaction
has to be included and, therefore, renders the result in the probe limit unreliable. Second, there seems to be the case that larger
values of the chemical potential for the p-wave superconductor allow for an enhancement of the condensate with a higher
slope that larger values for the s-wave. There are, certainly, some similarities. For example, as in \cite{Arean:2013mta},
the curves are noisier at large disorder
strength and even noisier for lower values of the chemical potential. It could be suspected that such behavior is the result
of numerical limitations but we have run extensive simulations to that effect and verified that the effect is real and has to
do more with the nature of the system of equations in this regime of parameters.
Finally, as for the case of uncorrelated noise, a tentative phase diagram is presented on the right panel of figure \ref{Cx-v-Ncorr}. It is constructed from the data on the left panel of that figure, following the same procedure as that described in section \ref{Sec:Phase} for the uncorrelated noise. This phase diagram clearly shows that
the introduction of correlated noise results in an enhancement of the superconductivity very similar to that observed in \cite{Arean:2013mta} for the case of the s-wave superconductor.
\section{Spectral properties and disorder}\label{Sec:Spectrum}
In previous sections we have mostly focused on the average properties of the condensate and the charge density.
For example, in Figs. \ref{Cx-v-N} and \ref{Cx-v-Ncorr} we followed the average value of the condensate as a function of the strength of
the noise.
Although this averaging is a good proxy at first, it is instrumental to the nature of disorder that we look into
properties depending on the spatial coordinate $x$. In this section we will go beyond that first order averaging study.
Following \cite{Arean:2013mta} we continue the study of the spectral properties of some of the quantities characterizing
our system. Our goal is to gain a quantitative understanding.
As in \cite{Arean:2013mta}, we establish certain universality of the power spectra of the condensate and charge density
as functions of the power spectrum of the signal defining the noise. Namely, for a given random signal with power spectrum
of the form $k^{-2\alpha}$ we study the power spectrum of the condensate $k^{-2\Delta(\alpha)}$, of the charge density
$k^{-2\Gamma(\alpha)}$, and of the $\rho_2$ charge density $k^{-2\Gamma_2(\alpha)}$; and report some interesting
universal behavior. We interpret this behavior as a particular form of renormalization of small wave-lengths.
We argue that this kind of smoothing/roughening points to a renormalization of sorts, where higher harmonics in ${\cal O}$
are suppressed or enhanced with respect to their spectral weight in $\mu$.
\begin{figure}[b]
\begin{center}
\includegraphics[width=.99\textwidth]{sp_renorm.pdf}
\caption{\label{RGToy} Renormalization of the disorder: Charge density $\rho$: $\Gamma= -1.00+ 1.00\,\alpha $ (left panel), charge density $\rho_2$: $\Gamma_2= 1.07+ 1.01\,\alpha $ (middle panel), and condensate:
$\Delta=2.02+1.03\,\alpha$ (right panel).
This plot was made considering $L=2 \pi$, $\mu_0=4$, $w=1$, and averaging over 5 realizations on a lattice of size
$22\times75$.
}
\end{center}
\end{figure}
Let us now carefully describe our setup.
To characterize this renormalization quantitatively we take a boundary chemical potential of the form presented in
equation (\ref{noisefunc}), but now considering different values for $\alpha$ (the choice of $\alpha$ determines
the degree of differentiability -smoothness- of the initial profile).
To make the concept of renormalization more precise we will study the power spectrum of the modes characterizing the response
of our system (the condensate, charge density $\rho$, and charge density $\rho_2$), as a function of the input power spectrum
determining our noise.
This input power spectrum of $\mu(x)$ is essentially proportional to $k^{-2\alpha}$. Remarkably, the power spectrum of the
condensate ${\cal O}(x)$ is numerically well approximated by $k^{-2\Delta}$ (see the rightmost panel of Fig.
\ref{RGToy}). We find $\Delta\simeq 2.02+1.03\alpha$, which is clearly larger than $\alpha$, meaning that
the weight of the high-$k$ harmonics is smaller in ${\cal O}$ than in $\mu$.
The power spectrum of the charge density $\rho$ is very well approximated by $k^{-2\Gamma(\alpha)}$ (Fig.
\ref{RGToy} left panel) with
$\Gamma \simeq -1.00+1.00 \alpha$, which implies that for the charge density the weight of the high-$k$ harmonics is
larger than in the spectrum of $\mu$. As for the charge density $\rho_2$, again the spectrum approximates very well to
a power law $k^{-2\Gamma_2(\alpha)}$ (Fig. \ref{RGToy} middle panel), with $\Gamma_2\simeq 2.02+1.03 \alpha$. As for
the condensate the weight of the high-$k$ harmonics is smaller than in $\mu$.
Let us stress that the spectra of all the response functions are given by power laws, and moreover, the
exponent of these response power laws is always of the form $\sim \alpha + {\rm integer}$.
This universality of RG is one of the main observations of our work and its origin seems to be in the strongly coupled
nature of the problem. The weak field theory intuition would dictate that $\Delta$ should be well approximated by
the conformal dimension associated with the order parameter and here we verify that it is not.
It is also interesting to point out that this behavior does not depend on any of the parameters of our theory, i.e.
$L$, $T/\mu$ or $w$. This means that we can redo Fig. \ref{RGToy} for the charge density in the normal phase.
This particular case is interesting, since the theory
becomes linear and we can therefore separate variables.
Being that the case, we can recompute the power spectrum solving the equations of motion using a
simple $Mathematica$'s NDSolve command. In this case we get $\Gamma= -1.00+ 1.00\,\alpha $,
which agrees with the result presented in Fig. \ref{RGToy}.
It is worth mentioning that the same scalings were found in \cite{Arean:2013mta} for the $s$-wave holographic superconductor,
and similar ones have been observed in a related holographic model
by other authors in \cite{Hartnoll:2014cua}.
\section{Conclusions}\label{Sec:Conclusions}
In this paper we have studied the influence of disorder in the holographic p-wave superconductor. We have found that
moderate disorder enhances the value of the order parameter and accordingly of the critical temperature.
We have also discussed various branches of solutions that appear in this particular setup since different solutions are
characterized by the dominant direction of the condensate.
We have established that the dominant solution, according to its free energy, is the one with the condensate along the
$x$-direction. We have also established that the condensate, $\langle {\cal O}_x\rangle/\mu^2$, is enhanced with the
disorder. Moreover, we have demonstrated the self-averaging property of $\langle {\cal O}_x\rangle$ under Gaussian
and uncorrelated randomness. We identify this enhancement with the ulterior enhancement of superconductivity.
The phase diagram is similar to the s-wave superconductor reported in \cite{Arean:2013mta} and we presented its
quantitative form in section \ref{Sec:Phase}. The key property is that the curve delimiting the normal and
superconducting phase shows an enhancement of the superconductivity with mild disorder. We have also studied some
universal properties of the power spectrum of the corresponding condensate and charge density.
We have found that the response is largely governed by a simple linear relation depending on the power spectrum of
the random chemical potential. These results expand those presented first in \cite{Arean:2013mta} to the case of a
disordered holographic p-wave superconductor. Similar behavior was also reported in \cite{Hartnoll:2014cua} in the
perturbative regime for a neutral scalar, and by some of the authors in gravity duals of brane intersections
\cite{futureUS}.
One of our main results, the enhancement of superfluidity with mild disorder,
is aligned with experimental and numerical claims of p-wave superfluidity enhancement by disorder
\cite{PhysRevB.79.214529,0295-5075-86-2-26004}.
We would like to finish by highlighting a few problems that are particularly interesting to us and some of which we hope
to pursue in future works. Having constructed the disordered solutions in \cite{Arean:2013mta} and the present manuscript,
it is natural to study transport properties and, in particular, the conductivities. It would also be interesting to
understand the effects of disorder in more general types of holographic p-wave superconductors. Recall that this type of
superconductors present a particularly interesting challenge to Anderson's theorem given its directional order parameter.
Some interesting models include \cite{Gubser:2008wv} and its extension to $p_x+ip_y$ along the lines of
\cite{Zayas:2011dw}. A recent study of conductivity in p-wave superconductors was presented in \cite{Herzog:2014tpa},
where some phenomenological similarities with high temperature cuprate superconductors were found even in a translational
invariant holographic model. It would be interesting to study the persistence or modification of such properties under the
effect of introducing disorder as a way of breaking translational invariance in these and similar systems.
Finally, as in \cite{Arean:2013mta}, we have established the existence of fairly universal response of the condensate and the charge density to the
power spectrum of the random disorder. We view this as evidence of some universality in cases of strongly coupled
systems under a sort of disorder renormalization. On the other hand, for very small values of $k_0$, we found evidence of a
new scaling for the expectation values of one point functions. We expect to study this potential universality in more detail
in the future \cite{futureUS}.
\section*{Acknowledgments}
We would like to thank M. Ara\'ujo, J. Sonner and T. Takayanagi for useful discussions.
D.A, L.A.P.Z. and I.S.L. thank the Abdus Salam ICTP, Italy for hospitality at various stages of this project.
I.S.L. thanks Max-Planck-Institut f\"ur Physik for hospitality.
D.A. thanks the FRont Of pro-Galician Scientists for unconditional support.
We also thank the Bivio; for being the epic place we needed to finish this project.
Some simulations
were performed in the University of Michigan Flux high-
performance computing cluster. This work is partially
supported by Department of Energy under grant DE-
FG02-95ER40899. The work of D.~Are\'an is supported by GIF, grant 1156
\bibliographystyle{JHEP}
|
\section{Introduction}
The notion of a free map arises naturally in free probability, the study of noncommutative rational functions \cite{AlpDu,BGM,HMV}, and systems theory \cite{HBJP, KV0}.
The study of these maps is in the realm of free analysis \cite{AM1,AM2,AY, AK,BV,KV,HKM12,MS,Pas,PT,Po1,Po2,Tay,Voc04,Voc10}.
The main contribution of this paper is to introduce powerful invariant-theoretic methods \cite{Pro} to free analysis.
We present an alternative, algebraic approach to free function theory.
While most of the current efforts in free analysis are focused on (involution-free) free maps -- free analogs of analytic functions in several complex variables -- where strong rigidity is observed, our main attention is to {\em free maps with involution}, e.g.~ noncommutative polynomials, rational function or power series in $x,x^t$.
Our methods are uniform in that they work in both cases with only minimal adaptations needed.
Thus we recover some of the existing results on (involution-free) free maps (cf. \cite{AM2,KV,Pas}).
We next give a list of the main results, that at the same time serves as a guide to the paper; we refer to Section \ref{sec2} for definitions and unexplained terminology.
\begin{enumerate}[\rm (1)]
\item
A free map with involution $f$ is a polynomial in $x,x^t$ if and only if there is $d\in \N$ such that each of the level functions $f[n]$ is a polynomial of degree $\leq d$ (Proposition \ref{hompol});
\item
Analytic free maps with involution admit convergent power series expansions
about scalar points (Theorem \ref{analit});
\item
Analytic free maps with involution admit convergent power series expansions about non-scalar points (Theorem \ref{rumanalit}, Theorem \ref{rumanalitO}), whose homogeneous parts are generalized polynomials. We present an invariant theoretic characterization of the latter in Subsection \ref{ssec4};
\item
Free inverse and implicit function theorems for differentiable free maps with involution are the theme of Section \ref{sec5}, see Theorem \ref{IFFT}, Corollary \ref{IFFT2}, and Theorem \ref{IFFTn};
\item
Section \ref{ex} presents several illustrating examples demonstrating non-rigidity properties
of free maps with involution. For instance, we give an example of a bounded smooth free map with involution that is not analytic (Example \ref{sin}).
\end{enumerate}
\def\cM{\mathcal M}
\section{Preliminaries}\label{sec2}
In this section we present preliminaries from free analysis, polynomial identities \cite{Dre,Row} and invariant theory \cite{Pro}
needed in the sequel.
\subsection{Notation}
Let $\F\in \{\RR,\CC\}$ and let $\cM(\F)$ stand for $\bigcup_n M_n(\F)$. We denote the monoid generated by $x_1,\dots,x_g$ by $\X$, and the free associative algebra in the variables $x=(x_1,\dots,x_g)$ by $\F\X$.
The free algebra with involution in the variables $x_1,x_1^t,\dots,x_g,x_g^t$ is denoted by $\F\Xt$.
The elements of degree $d$ in $\F\X$ (resp. $\F\Xt$) are denoted by $\F\X_d$ (resp. $\F\Xt_d$).
We write
\[C=\F\big[x_{ij}^{(k)}\mid 1\leq i,j\leq n,1\leq k\leq g\big]
\]
for the commutative polynomial ring in $gn^2$ variables.
We equip $M_n(C)$ with the transpose involution fixing $C$ pointwise.
The matrices $X_k=(x_{ij}^{(k)})\in M_n(C)$, $1\leq k\leq g$, are called {\bf generic matrices}.
By $\GM_n$
we denote the {unital} subalgebra of $M_n(C)$ generated by generic matrices,
and by $\GMt$ the subalgebra of $M_n(C)$ generated by generic matrices and their transposes.
We let $\Rn$ stand for the subalgebra of $M_n(C)$ generated by the generic matrices and traces $\tr(X_{i_1}\cdots X_{i_k})$ {of their products},
and $\Rnt$ for the subalgebra of $M_n(C)$ generated by generic matrices, their transposes, and traces $\tr(U_{i_1}\cdots U_{i_k})$, $U_j\in \{X_j,X_j^t\}$.
The center of $\Rn$ (resp. $\Rnt$) is generated by the traces, we denote it by $Z(\Rn)$ (resp. $Z(\Rnt)$).
\subsection{Free Sets and Free Maps}
Let $G=(G_n)_n$ be a sequence of groups {with $G_n\subseteq\GL_n(\F)$},
satisfying \beq\label{grgp}
G_n \oplus G_m = \begin{pmatrix} G_n & 0 \\ 0 &G_m \end{pmatrix} \subseteq G_{n+m}.
\eeq
We will be primarily concerned with the case $G_n=\GL_n(\F)$ for all $n$, or
$G_n$ is the orthogonal group $\OO_n(\RR)$ for all $n$. The modifications needed for the case
of the unitary groups $G_n=\U_n(\CC)$ will be discussed in Appendix \ref{apU}.
For simplicity of notation we write $\GL_n,\OO_n,\U_n$ instead of $\GL_n(\F),\OO_n(\RR),\U_n(\CC)$, respectively.
Let us denote $\GL=(\GL_n)_{n\in \N}$, $\OO=(\OO_n)_{n\in\N}$, $\U=(\U_n)_{n\in \N}$.
A subset $\cU\subseteq \cM(\F)^g$ is a sequence $\cU=(\cU[n])_{n\in \N}$, where each $\cU[n]\subseteq M_n(\F)^g$. The set $\cU$ is a {\bf $G$-free set}
if it is closed with respect to simultaneous $G$-similarity and with respect to direct sums; i.e., for every $m,n\in \N$:
\begin{equation}\label{eq:conj}
\s X\s^{-1}=(\s X_1\s^{-1},\dots,\s X_g\s^{-1})\in \cU[n]
\end{equation}
for all $X\in \cU[n]$, $\s\in G_n$, and
\begin{equation}\label{eq:oplus}
X\oplus Y=
\begin{pmatrix}
X&0\\
0&Y
\end{pmatrix}\in \cU[m+n]
\end{equation}
for all $X\in \cU[m], Y\in \cU[n]$.
Let $\cU$ be a $G$-free set.
We call a sequence of functions $f=(f[n])_{n\in\N}:(\cU[n])_{n\in\N}\to \MF$ a {\bf $G$-free map}, if it respects $G$-similarity and
direct sums; i.e, for every $m,n\in \N$:
\begin{equation}\label{eq:conj2}
f[n](\s X\s^{-1})=\s\, f[n](X)\,\s^{-1}
\end{equation}
for all $X\in \cU[n]$, $\s\in G_n$, and
\begin{equation}\label{eq:oplus2}
f[m+n](X\oplus Y)=f[m](X)\oplus f[n](Y)
\end{equation}
for all $X\in \cU[m], Y\in \cU[n]$.
{In the language of invariant theory} \cite{Pro,KP} the condition \eqref{eq:conj2} says that $f[n]$ is a $G_n$-concomitant. If $f$ satisfies only \eqref{eq:conj2} for all $n$ (and not necessarily \eqref{eq:oplus2}) we call it a {\bf free $G$-concomitant}.
Sometimes a $\GL$-free map is called simply a {\bf free map} and an $\OO$-free map is a {\bf free map with involution}.
With a slight abuse of notation we sometimes also refer to a map $f:\cU\to \cM$ as a $G$-free map if its domain $\cU$ is only closed under direct sums, $f$ respects direct sums and $f$ respects $G$-similarity on $\cU$; i.e, for every $n\in \N$:
$$f[n](\sigma X\sigma^{-1})=\sigma\, f[n](X)\,\sigma^{-1}$$
for all $X\in \cU[n]$, $\sigma\in G_n$ such that $\sigma X\sigma^{-1}\in \cU[n]$.
In this case we can canonically extend $f$ to the similarity invariant envelope of $\cU$
(cf.~\cite[Appendix A]{KV}), and remain in the framework of the given definition:
\begin{proposition}\label{simenv}
Let $\cU\subseteq \cM(\F)^g$ be closed under direct sums, and let $f:\cU\to \cM(\F)$ respect direct sums and $G$-similarity on $\cU$. Then
$$\td\cU=\{\sigma A\sigma^{-1}\mid A\in \cU[n], \sigma\in G_n,n\in \N\}$$
is a $G$-free set, and there exists a unique $G$-free map $\td f:\td\cU\to\cM(\F)$ such that $\td f|_\cU=f$, defined by $\td f(\s X\s^{-1})=\s f(X)\s^{-1}$ for $X\in \cU[n]$, $\s\in G_n$.
\end{proposition}
\begin{remark}
In \cite[Appendix A]{KV} the proof is given in the case $G=\GL$. The same proof with obvious modifications works also for any sequence of groups $G=(G_n)_n$ satisfying \eqref{grgp},
in particular for $G\in\{\OO,\U\}$.
\end{remark}
A $G$-free map $f$ is $\F$-analytic around $0$ if there exists a neighborhood
\beq\label{eq:nbhd}
\cB(0,\delta)=\bigcup_n\{X\in M_n(\F)^{g}\mid \|X\|<\delta_n\}
\eeq
of $0$ in $\cM(\F)^g$ such that $f[n]_{ij}$ is $\F$-analytic on $\cB(0,\delta)[n]$, $\delta=(\delta_n)_{n}$, and $\delta_n>0$ for every $n\in \N$.
It is a polynomial map of degree $m$ if $f[n]_{ij}$ are polynomials in $x_{ij}^{(k)}$ of degree $\leq m$ and at least one of the polynomials $f[n]_{ij}$ is of degree $m$; it is homogeneous of degree $m$ if $f[n]_{ij}$ are homogeneous polynomials of degree $m$ or zero polynomials, and $f[n]_{ij}$ is of degree $m$ for at least one triple $(n,i,j)$.
\subsection{Trace Polynomials}
The free algebra with trace $\Ftr$ is the algebra of free noncommutative polynomials in the variables $x_k$ over the polynomial algebra $T$ in the infinitely many variables $\tr(w)$, where $w$ runs over all representatives of the cyclic equivalence classes of words in the variables $x_k$; i.e., $w\in \X/_{\cyc}$.
Here two words $u,v\in\X$ are cyclically equivalent, $u\cyc v$, iff $u$ is a cyclic permutation of $v$.
The free $*$-algebra with trace $T^\dagger\Xt$ is the algebra of free noncommutative polynomials in the variables $x_k, x_k^t$
over the polynomial algebra $T^\dagger$ in the infinitely many variables $\tr(w)$, where $w$ runs over all representatives of the {$*$-cyclic} equivalence classes of words in the variables $x_k,x_k^t$; i.e., words $u$ and $v$ are equivalent if $u\cyc v$ {or $u^t\cyc v$}.
The elements of $\Ftr$ (resp. $T^\dagger\Xt$) are {\bf trace polynomials} (resp. {\bf trace polynomials with involution}) and elements of $T$ (resp. $T^\dagger$) are {\bf pure trace polynomials} (resp. {\bf pure trace polynomials with involution}).
The degree of a trace monomial $\tr(w_1)\cdots\tr(w_m)v$, $w_i,v\in \X$, equals $|v|+\sum_i |w_i|$, where $|u|$ denotes the length of a word $u$.
The degree of a trace polynomial is the maximum of the degrees of its trace monomials.
{\em Trace identities} of the matrix algebra $M_n(\F)$ (with involution) are the elements in the kernel of the evaluation map from the free algebra (with involution) with trace to $M_n(\F)$; i.e., trace identities of $M_n(\F)$ are trace polynomials that vanish on $n\times n$-matrices. {\em Pure trace identities} are trace identities that belong to $T$ (resp. $T^\dagger$).
The free ($*$-)algebra with trace $\Ftr$ (resp. $T^\dagger \Xt$) and the trace identities have its interpretation in terms of invariants of matrices. Let $G=\GL_n$ (resp. $G=\OO_n$) act by conjugation on $M_n(\F)$
and diagonally (i.e., componentwise) on $M_n(\F)^g$.
The first fundamental theorem for matrices (with involution) yields that a $\GL_n$- (resp. $\OO_n$-) concomitant is a trace polynomial (resp. with involution), see \cite[Theorem 2.1, Theorem 7.2]{Pro} or \cite[Chapter 11]{Pro3} for a broader perspective on the subject.
(For another take on the theory of polynomial identities we refer the reader to \cite{BCM}.)
Viewing a polynomial map $f:M_n(\F)^g\to M_n(\F)$ as an element $\td f\in M_n(C)$ we can see that the algebra of $\GL_n$- (resp. $\OO_n$-) concomitants is isomorphic to $\Rn$ (resp. $\Rnt$), and $\Rn$ (resp. $\Rnt$) is isomorphic to the quotient of $\Ftr$ (resp. $T^\dagger \Xt$) by the ideal of trace identities (resp. trace identities with involution).
\section{Analytic $G$-Free Maps and Power Series Expansions about Scalar Points}\label{0}
{In this section we investigate two distinguished classes of free maps,
namely polynomials and analytic free maps. We characterize free maps which
are polynomials in Subsection \ref{subsec:poly}, and use this to show that analytic free maps
admit power series expansions about scalar points
in Subsection \ref{subsec:analytic}.}
{These results are classical for $G=\GL$ (cf.~\cite{KV,Tay,Voc10}) and are -- to the best of our knowledge -- new for $G=\OO$.}
Throughout this section $G\in\{\GL,\OO\}$.
\subsection{Polynomial Free Maps}\label{subsec:poly}
{We start by characterizing free polynomial maps $f$ via their ``slices'' $f[n]$.
For $G=\GL$ this result is due to
Kaliuzhnyi-Verbovetskyi and Vinnikov \cite[Theorem 6.1]{KV} who deduce it from their power series expansion theorem for analytic free maps.
In contrast to this we shall first characterize free polynomial maps and employ this in Subsection \ref{subsec:analytic} to establish power series expansions for analytic $G$-free maps.
Our proofs are uniform in that they work for both $G=\GL$ and $G=\OO$, and are purely algebraic,
depending only on the invariant theory of matrices \cite{Pro}.}
\begin{comment}
\begin{proposition}\label{prop:freetr}
Let $f:\MF^g\to \MF $ be a free $G$-concomitant.
If $f$ is a polynomial map
and $\max_n \deg f[n]= d$, then $f$ is a trace polynomial of degree $d$.
\end{proposition}
\end{comment}
\begin{proposition}\label{hompol}
Let $f:\MF^g\to \MF $ be a $G$-free map.
If $f$ is a polynomial map
and $\max_n \deg f[n]= d$, then $f$ is a free
polynomial of degree $d$.
{That is, $f\in\F\X_d$ if $G=\GL$ and
$f\in\F\Xt_d$ if $G=\OO$.}
\end{proposition}
\begin{proof}
Since $f[n]:M_n(\F)^g\to M_n(\F)$ is a concomitant, it follows by \cite[Theorem 2.1, Theorem 7.2]{Pro} that $f[n]$ is a trace polynomial of degree $\leq d$ in the variables $x_k$ (resp. $x_k, x_k^t$).
Since there do not exist nontrivial trace identities for $M_n(\F)$ of degree less than $n$ by \cite[Theorem 4.5, Proposition 8.3]{Pro} (see also \cite{BK,Raz}), we can write $f[n]$ in the case $n\geq d+1$ uniquely as
$$f[n]=\sum_M \tr(h_M^n)M,$$
where $M$ runs over all monomials of degree $\leq d$ and $\deg \tr(h_M^n)+\deg M\leq d$. Choose $n\geq d+1$.
As $f$ is a free map, we have
\begin{multline*}
\sum_M \tr\big(h_M^{2n}(X\oplus Y)\big)M(X)\oplus \sum_M \tr\big(h_M^{2n}(X\oplus Y)\big)M(Y)=f[2n](X\oplus Y) \\
=f[n](X)\oplus f[n](Y)=\sum_M \tr\big(h_M^{n}(X)\big)M(X)\oplus \sum_M \tr\big(h_M^{n}(Y)\big)M(Y)
.
\end{multline*}
Comparing both sides of the above expression we obtain
$$\tr\big(h_M^{2n}(X\oplus Y)\big)=\tr\big(h_M^{n}(X)\big)=\tr\big(h_M^{n}(Y)\big)$$
since $M_n(\F)$ does not satisfy a nontrivial trace identity of degree $d$.
Thus,
$$\tr\big(h_M^{n}(X)\big)=\alpha=\tr\big(h_M^{n}(Y)\big)$$
for some $\alpha\in \F$.
Hence, for every $n>N$, $f[n]\in \GM_n$ (resp. $f[n]\in \GMt$)
is represented by an element $\tilde f\in \F\langle X\rangle$ (resp. $\tilde f\in \F\Xt$) of degree $d$.
Since $f$ is a free map, we can identify it with a free polynomial in the variables $x_k$ (resp. $x_k,x_k^t$).
\end{proof}
\begin{remark}
We note that Proposition \ref{hompol} holds also if $f$ is only defined on $\cB(0,\delta)$ (cf.~Proposition \ref{simenv}), since polynomial functions that agree on an open subset of $M_n(\F)^g$ represent the same function on $M_n(\F)^g$.
\end{remark}
\subsection{Analytic Free Maps}\label{subsec:analytic}
{We next turn our attention to analytic $G$-free maps. We show they admit unique convergent power series expansions about scalar points $a\in\F^g$, extending classical results for $G=\GL$, cf.~\cite{Tay,Voc04,Voc10,KV,HKM12}.} {By a translation we may assume without loss of generality that $a=0$.}
\begin{theorem}\label{analit}
Let $\cU$ be a $G$-free set and $f:\cU\to \MF$ an $\F$-analytic $G$-free map, and
let $\cB(0,\delta)\subseteq \cU$, where $\delta=(\delta_n)_{n\in\N}$, $\delta_n>0$ for every $n\in \N$.
{Then} there exists a unique formal power series
\beq\label{eq:pw}
F=\sum_{m=0}^\infty \sum _{|w|=m}F_w w,
\eeq
where $w\in \X$ $($resp. $w\in \Xt)$, which converges in norm on $\cB(0,\delta)$, {with} $f(X)=F(X)$ for $X\in \cB(0,\delta)$.
\end{theorem}
\begin{remark}
If $f$ is uniformly bounded, and $G=\GL$ then the convergence of the power series $F$ in \eqref{eq:pw} is uniform, cf.~\cite[Proposition 2.24]{HKM12}, while this conclusion does not hold when $G=\OO$. We present examples in Section \ref{ex}.
\end{remark}
We first prove the existence, the uniqueness will follow from Proposition \ref{unique} below.
\begin{proof}[Proof of the existence]
Since $f$ is analytic, there exists for every $X\in M_n(\F)^g$ a neighbourhood of $0$ such that the function $t\mapsto f[n](tX)$ is defined and analytic in that neighbourhood.
Hence, $f[n](tX)$ can be expressed in that neighbourhood as a convergent power series of the form $\sum_{m=0}^\infty t^m f[n]_{m}(X)$,
where $f[n]_{m}(X)$ is a function of $X$.
Note that for $X\in \cB(0,\delta)$, this power series converges for $t=1$.
{The} function $f[n]_{m}$ is a homogeneous polynomial function of degree $m$.
Indeed,
let $s\in \F$, $X\in M_n(\F)^g$ and choose $\delta'$ such that $tsX\in \cB(0,\delta)$ for $|t|\leq \delta'$.
Then
$$\sum_{m=0}^\infty t^m f[n]_{m}(sX)=f(tsX)=\sum_{m=0}^\infty (ts)^mf[n]_{m}(X),$$
and thus $f[n]_{m}(sX)=s^m f[n]_{m}(X)$.
Let us show that $f_{m}$ defined by $f_{m}[n]:=f[n]_{m}$ is an analytic free map.
Choose $\delta'$ such that $tX,tY,\sigma tX\s^{-1}\in \cB(0,\delta)$ for $|t|<\delta'$.
As $f$ is a free map we have
\begin{multline*}
\sum_{m=0}^\infty t^m f[n+n']_{m}(X\oplus Y)=f[n+n'](tX\oplus tY)
\\
=f[n](tX)\oplus f[n'](tY)=\sum_{m=0}^\infty t^m \big(f[n]_{m}(X)\oplus f[n']_{m}(Y)\big),
\end{multline*}
and
$$\sum_{m=0}^\infty t^m \s\, f[n]_{m}(X)\, \s^{-1}= \s\, f[n](tX)\, \s^{-1}=f[n](t\s X\s^{-1})=\sum_{m=0}^\infty t^m f[n]_{m}(\s X\s^{-1})$$
for all $|t|<\delta'$,
which implies that $f_{m}$ is a $G$-free map.
By construction, $f_{m}$ is a homogeneous polynomial function of degree $m$ (or $0$) for every $m$.
By Proposition \ref{hompol}, $f_m$ can be represented by a
free polynomial in the variables $x_k$ (resp. $x_k,x_k^t$) of degree $m$. Thus, $f$ can be expressed as a power series in noncommuting variables, $F=\sum f_{m}$.
By construction, this power series converges on $\cB(0,\delta)$.
\end{proof}
While the theories of $\GL$- and $\OO$-free maps enjoy certain similarities,
there are also major differences.
For instance, for $\GL$-free maps continuity implies analyticity and there is a very useful formula \cite[Proposition 2.5]{HKM11}, \cite[Theorem 7.2]{KV}
connecting function values with the derivative:
\begin{equation}\label{der}
f
\begin{pmatrix}
X&H\\
0& X
\end{pmatrix}=
\begin{pmatrix}
f(X)&{\delta}f(X)(H)\\
0& f(X)
\end{pmatrix},
\end{equation}
where ${\rm \delta}f(X)(H)$ denotes the G\^{a}teaux
{(directional)}
derivative of $f$ at $X$ in the direction $H$; i.e.,
\[{\rm \delta}f(X)(H)=\lim_{t\to 0}\frac{f(X+tH)-f(X)}{t}.\]
For $\OO$-free maps continuity does not imply differentiability; see Section \ref{ex} for examples.
However, for differentiable
$\OO$-free maps we do have an analog of formula \eqref{der}, which can be deduced from \cite[Lemma 2.3, Proposition 2.5]{PT}, but we prove it {here} for the sake of completeness. We write $\D f$ for a derivative of $f$, it can be either the G\^ateaux or the Fr\'echet derivative. The Lie bracket $[a,B]$ stands for $([a_,B_1],\dots,[a,B_g])$, where $a\in M_n(\F)$, $B=(B_1,\dots,B_g)\in M_n(\F)^g$.
\begin{lemma}
Let $f:\cU\to \cM(\F)$ be a real differentiable $G$-free map.
Then the identity
\begin{equation}
\D f(X)([a,X])=[a,f(X)]
\end{equation}
holds for all $X\in \cU[n]$, $a^t=-a\in M_n(\RR)$.
In particular, \begin{equation}\label{did}
\D f
\begin{pmatrix}
X_1&0\\
0& X_2
\end{pmatrix}
\begin{pmatrix}
0&X_1-X_2\\
X_1-X_2& 0
\end{pmatrix}=
\begin{pmatrix}
0&f(X_1)-f(X_2)\\
f(X_1)-f(X_2)& 0
\end{pmatrix}.
\end{equation}
\end{lemma}
\begin{proof}
Note that $e^{sa}$ is orthogonal for $a^t=-a\in M_n(\RR)$ and $s\in \RR$.
Thus we have
$$
f(e^{-sa}Xe^{sa})=e^{-sa}f(X)e^{sa}
$$
for every $X\in \cU[n]$.
Differentating with respect to $s$ at $0$ yields
$$
\D f(X)([a,X])=[a,f(X)].
$$
Take
$$a=\begin{pmatrix}
0&I_n\\
-I_n& 0
\end{pmatrix}\in M_{2n}(\RR),$$
where $I_n$ denotes the identity in $M_n(\RR)$.
Setting
$X=\begin{pmatrix}
X_1&0\\
0& X_2
\end{pmatrix}$
we get the identity \eqref{did}.
\end{proof}
We now show that the power series expansion is unique for a $G$-free function and give a way to recover its coefficients.
\begin{lemma}\label{matenote}
If $f(X)=\sum _{|w|\leq m}F_w w$, where {the sum is over words} in the variables $x_k$ (resp. $x_k,x_k^t$), then we can obtain {the coefficients} $F_w$ by evaluations of $f$ on $M_{m+1}(\F)$.
\end{lemma}
\begin{proof}
We proceed inductively.
Assume that we can obtain coefficients of $f(X)=\sum _{|w|\leq k}F_w w$ for $k<m$ by evaluations of $f$ on $M_{k+1}(\F)$.
The case $k=1$ is trivial. Suppose that $k=m$.
Let us determine the coefficient at
$w=u_{i_1}^{j_1}\cdots u_{i_s}^{j_s}$, where $\sum_{k=1}^s j_k=m$ and $u_{i_k}\in \{x_{i_k},x_{i_k}^t\}$.
We denote $s_k=\sum_{i=1}^{k} j_i$.
Setting $a_i=0$ at the beginning, we define a $g$-tuple $(a_i)\in M_{m+1}(\F)^g$ as follows.
We let $k$ {run} from $1$ to $s$, and at step $k$ we replace $a_{i_k}$ by
\[
a_{i_k}=
\begin{cases}
a_{i_k}+\sum_{u=s_{k-1}+1}^ {s_k}e_{u,u+1}&
\textrm{if $u_{i_k}=x_{i_k}$},\\[.1cm]
a_{i_k}+\sum_{u=s_{k-1}+1}^ {s_k}e_{u+1,u}&
\textrm{if $u_{i_k}=x_{i_k}^t$}.
\end{cases}
\]
We shall show that
$\tr(f(a_1,\dots,a_g)e_{m+1,1})=F_w$.
We need to find the coefficient of $f(a_1,\dots,a_g)$ expressed in {the} standard basis $e_{ij}$, $1\leq i,j\leq m+1$, of $M_{m+1}(\F)$ at $e_{1,m+1}$.
According to the definition of {the} $a_{i}$'s it suffices to show that $e_{1,m+1}$ can be obtained in only one way as a product of $\leq m$ matrix units from the set $S=\{e_{i,i+1},e_{i+1,i}\mid 1\leq i\leq m\}$. Note that the multiplication on the right of any matrix unit $e_{ij}$ by any element of $S$ either
{increases or decreases} $j$ by $1$. In order to obtain $e_{1,m+1}$ as a product of $\leq m$ elements from $S$, we can thus only choose matrix units which {increase} the second subscript of the preceding matrix unit in the product. Hence, $e_{1,m+1}=e_{12}\cdots e_{m,m+1}$, and any other product of $\leq m$ elements from $S$ {will be different from} $e_{1,m+1}$. As each $e_{i,i+1}$ appears only in one of the $a_i,a_i^t$, $1\leq i\leq g$, the order $e_{12},\dots,e_{m,m+1}$ corresponds to exactly one order of the $a_{i}'s$. By the definition of $a_i$ this order corresponds to $w$.
Now we can find the coefficients of $f-\sum _{|w|=m}F_w w=\sum _{|w|< m}F_w w$ by the induction hypothesis on $M_{m}(\F)\subset M_{m+1}(\F)$.
\end{proof}
\begin{proposition}\label{unique}
Suppose that a $G$-free map $f$
has a power series expansion in a neighbourhood $\cB(0,\delta)$ of $0$, $\delta=(\delta_n)_{n\in\N}$;
i.e.,
$$f(X)=\sum_{m=0}^\infty\sum _{|w|=m}F_w w(X),$$
for $X\in\cB(0,\delta).$
Then $F_w$ for $|w|=m$ is determined by the $m$-th derivative of the function $t\mapsto f[m+1](tX)$ at $0$ and hence by its evaluation on $M_{m+1}(\F)$.
\end{proposition}
\begin{proof}
Let $|t|<1$, then $tX\in \cB(0,\delta)[n]$ for every $X\in \cB(0,\delta)[n]$, and
$$f[n](tX)=\sum_{m=0}^\infty t^m f_m[n](X)$$
is a convergent power series in $t$, where $f_{m}$ are homogeneous free polynomials of degree $m$.
We can thus determine $f_m[n](X)$ as
$$\frac{1}{m!}\frac{\rm d}{{\rm d}t^m}f[n](tX)\bigg|_{t=0}.$$
Since $M_n(\F)$ does not admit a nontrivial polynomial identity (with involution) of degree $<n$
(see e.g.~\cite[Lemma 1.4.3, Remark 2.5.14]{Row}), $f_{m}$ is uniquely determined on $M_{m+1}(\F)$. Hence we can recover $f_{m}$ by the $m$-th derivative of the function $t\mapsto f[m+1](tX)$.
The coefficients of the polynomial $f_{m}$ can be constructively determined by evaluations on $M_{m+1}(\F)$ by Lemma
\ref{matenote}.
\end{proof}
\section{\RPS and Power Series Expansions about Non-scalar Points}\label{sec4}
{Theorem \ref{analit} gives a convergent power series expansion of a free analytic
map about a \emph{scalar point} $a\in\F^g$. In this section we present power series expansions
about non-scalar points $A\in M_n(\F)^g$, whose homogeneous components are generalized
polynomials.}
These are the topic of Subsection \ref{ssec4} and their obtained properties will be used in Subsection \ref{s2sec4} to deduce the desired power series expansion.
Our methods are algebraic, and work for $G=\GL$ and $G=\OO$.
For $G=\GL$ a similar result has been obtained earlier in \cite{KV} with a different proof.
Throughout this section $G\in\{\GL,\OO\}$.
\subsection{\RPS}\label{ssec4}
We call the elements of the free product $M_n(\F)\ast \F\X$ {\bf \Rps} (cf. \cite{Ami}, \cite[Section 4.4]{BMM}).
They can be written in the form
$$\sum a_{i_0}x_{k_1}a_{i_1}x_{k_2}\cdots a_{i_{\ell-1}}x_{k_\ell}a_{i_{\ell}},$$
where $a_{i_j}\in M_n(\F)$.
Let $e_{ij}$ denote the standard matrix units of $M_n(\F)$. Then a basis of $M_n(\F)\ast \F\X$ consists of monomials
$$ e_{i_0,j_0}x_{k_1}e_{i_1,j_1}x_{k_2}\cdots e_{i_{\ell-1},j_{\ell-1}}x_{k_\ell}e_{i_{\ell},j_{\ell}}$$
for $\ell\in\N_0$, $I,J\in \{1,\dots,n\}^{\ell+1}$, $K\in \{1,\dots,g\}^\ell$, where $I=(i_0,\dots,i_\ell),J=(j_0,\dots,j_\ell),K=(k_1,\dots,k_\ell)$.
The algebra $M_n(\F)\ast \F\X$ can be evaluated (as an algebra with unity) in $M_{ns}(\F)$ for $s\in \N$ and we have an isomorphism
\beq\label{redfun}
{\rm Hom}_{M_n}(M_n(\F)\ast \F\X,M_{ns}(\F))\cong {\rm Hom}(\w_n(\F\X),M_{s}(\F)),
\eeq
where $\w_n$ denotes the matrix reduction functor (see \cite[Section 1.7]{Coh}).
The isomorphism is a consequence of the identity
\beq\label{redfunid}
M_n(\F)\ast \F\X \cong M_n(\w_n(\F\X)).
\eeq
For the free algebra $\F\X=\F\langle x_1,\dots,x_g\rangle$ we have
$$\w_n(\F\X)=\F\langle y^{(k)}_{ij}\mid 1\leq i,j\leq n,1\leq k\leq g\rangle,$$
where $y_{ij}^{(k)}$, as the brackets suggest, denote free noncommutative variables.
For example, the evaluation of the element
$$e_{11}x_1e_{12}x_2e_{22}\in M_2(\F)\ast\F\X $$
in $M_4(\F)$, defined by mapping $x_1,x_2$ to $A,B\in M_4(\F)$, is
$$
\begin{pmatrix}
\multicolumn{2}{c}{\multirow{2}{*}{$\;I_2\;$}}&\multicolumn{2}{c}{\multirow{2}{*}{\quad}}\\
&&&\\
\multicolumn{2}{c}{\multirow{2}{*}{\quad}}&\multicolumn{2}{c}{\multirow{2}{*}{$\; \;$}}\\
&&&\\
\end{pmatrix}
\begin{pmatrix}
\multicolumn{2}{c}{\multirow{2}{*}{$A_{11}$}}&\multicolumn{2}{c}{\multirow{2}{*}{$A_{12}$}}\\
&&&\\
\multicolumn{2}{c}{\multirow{2}{*}{$A_{21}$}}&\multicolumn{2}{c}{\multirow{2}{*}{$A_{22}$}}\\
&&&\\
\end{pmatrix}
\begin{pmatrix}
\multicolumn{2}{c}{\multirow{2}{*}{$\;\;$}}&\multicolumn{2}{c}{\multirow{2}{*}{$\;I_2\;$}}\\
&&&\\
\multicolumn{2}{c}{\multirow{2}{*}{\quad}}&\multicolumn{2}{c}{\multirow{2}{*}{$\;\;$}}\\
&&&\\
\end{pmatrix}
\begin{pmatrix}
\multicolumn{2}{c}{\multirow{2}*{$B_{11}$}}&\multicolumn{2}{c}{\multirow{2}{*}{$B_{12}$}}\\
&&&\\
\multicolumn{2}{c}{\multirow{2}{*}{$B_{21}$}}&\multicolumn{2}{c}{\multirow{2}{*}{$B_{22}$}}\\
&&&\\
\end{pmatrix}
\begin{pmatrix}
\multicolumn{2}{c}{\multirow{2}{*}{$\; \;$}}&\multicolumn{2}{c}{\multirow{2}{*}{\quad}}\\
&&&\\
\multicolumn{2}{c}{\multirow{2}{*}{\quad}}&\multicolumn{2}{c}{\multirow{2}{*}{$\;I_2\;$}}\\
&&&\\
\end{pmatrix}
=
\begin{pmatrix}
&&\multicolumn{2}{c}{\multirow{2}{*}{$A_{11}B_{22}$}}\\
&&&\\
&&&\\
&&&\\
\end{pmatrix} ,
$$
where $I_2$ denotes the identity of $M_2(\F)$, and $A_{ij}$ (resp. $B_{ij}$) denotes the $(i,j)$-block entry of $A$ (resp. $B$), or
$$(e_{11}\tnz I_2)A(e_{12}\tnz I_2)B(e_{22}\tnz I_2)=e_{12}\tnz A_{11}B_{22},$$
viewed as en element in $M_2(\F)\tnz M_2(\F)\cong M_4(\F)$.
Note that \eqref{redfun} and \eqref{redfunid} imply that no \Rp vanishes on $M_{ns}(\F)$ for all $s$. In fact, two \Rps of degree $2d$ which agree on $M_{ns}(\F)$ for {some} $s>d$ are equal.
We denote by ${\rm g}\cT_{ns}$ the ideal of the elements in $M_n(\F)\ast\F\X$ that vanish when evaluated on $M_{ns}(\F)$ and let
\[
C_{ns}=\F\big[x_{ij}^{(k)}\mid 1\leq i,j\leq ns,1\leq k\leq g\big].
\]
The quotient algebra $\GMn=\big(\nA\big)/{\rm g}\cT_{ns}$ is isomorphic to the image of
$$\phi:\nA\to M_{ns}(C_{ns}),$$
defined by mapping $x_k$ to the corresponding generic matrix $(x_{ij}^{(k)}).$
We write $\Rns$ for the subalgebra of $M_{ns}(C_{ns})$ generated by $\GMn$ and traces of the elements in $\GMn$.
Note that every polynomial map $p:M_{ns}(\F)^g\to M_{ns}(\F)$ can be considered as an element $\td p\in M_{ns}(C_{ns})$.
Let $\GL_{ns}$ act on $M_{ns}(\F)$ by conjugation.
We will be interested in the action of its subgroup $I_n\tnz \GL_s$. In the next proposition we describe the invariants and concomitants of this action.
\begin{proposition}\label{partinv}
If $p: M_{ns}(\F)^g\to M_{ns}(\F)$
is an $I_n\otimes \GL_s$-concomitant, then $\td p\in \Rns$.
\end{proposition}
\begin{proof}
We can assume that $p$ is multilinear of degree $d$.
Then $p$ corresponds to an element in
$(M_{ns}(\F)^{\tnz d})^*\tnz M_{ns}(\F)$, which is canonically isomorphic to
$M_n(\F)^{\tnz {d+1}}\tnz (M_s(\F)^{\tnz d})^*\tnz M_s(\F)$ as $I_n\otimes \GL_s$-module.
The action of the group $I_n\tnz\GL_s$ reduces to the
action of $\GL_s$ on $(M_s(\F)^{\tnz d})^*\tnz M_s(\F)$.
The invariants of this action correspond to multilinear trace polynomials of degree $d$ in $M_s(C_s)$ by \cite[Theorem 2.1]{Pro}.
Moreover, the elements of the form
$$\sum_{I,J}e_{i_1j_1}\tnz\cdots\tnz e_{i_dj_d}\tnz \tau_{IJ},$$
where $\tau_{IJ}\in (M_s(\F)^{\tnz d})^*\tnz M_s(\F)$ is a $\GL_s$-concomitant map, can be identified with multilinear elements of degree $d$ in $\Rns$. \end{proof}
\subsubsection{Generalized Polynomials with Involution}
To consider the case of algebras with involution we need to introduce some additional notation.
We call the elements of the algebra $M_n(\F)\ast\F\Xt$ {\em \Rps with involution}.
By ${\rm g}\cT^{\dagger}_{ns}$ we denote the ideal of elements in $M_n(\F)\ast\F\Xt$ that vanish on $M_{ns}(\F)$.
The quotient algebra is isomorphic to the subalgebra $\GMn^\dagger$ of $M_{ns}(C_{ns})$ generated by $\GMn$ and transposes of elements in $\GMn$.
We write $\Rns^\dagger$ for the subalgebra of $M_{ns}(C_{ns})$ generated by $\GMn^\dagger$ and traces of elements in $\GMn^\dagger$.
We have the (usual) action of $\OO_{ns}$ on $M_{ns}(C_{ns})$.
The following proposition is the analog of Proposition \ref{partinv} for the action of $I_n\otimes \OO_s$ on $M_{ns}(C_{ns})$.
\begin{proposition}\label{partinvO}
If $p\in M_{ns}(\F)^g\to M_{ns}(\F)$ is an $I_n\otimes \OO_s$-concomitant, then $\td p\in \Rns^\dagger$.
\end{proposition}
\begin{proof}
The proof goes along the same lines as that of Proposition \ref{partinv}, we only need to invoke \cite[Theorem 7.2]{Pro} instead of \cite[Theorem 2.1]{Pro}.
\end{proof}
\subsubsection{Block ad centralizing $G$-concomitants}
Let us denote $\cM_n(\F)^k=\bigcup_s M_{ns}(\F)^k$, $k\in \N$.
We say that a map $f:\cM_{n}(\F)^g\to \cM_{n}(\F)$ is $I_n\tnz G$-concomitant if
$$f[ns]:\big(M_{n}(\F)\tnz M_s(\F)\big)^g\to M_{n}(\F)\tnz M_s (\F)$$
is a $I_n\tnz G_s$-concomitant for every $s\in \N$.
\begin{proposition}\label{freeRum}
If $f:\cM_{n}(\F)^g\to \cM_{n}(\F)$
is a homogeneous polynomial map of degree $d$ and $I_n\tnz \GL$-concomitant $($resp. $I_n\tnz \OO$-concomitant$)$
that preserves direct sums, then $f\in \nA$ $($resp. $f\in \nAt)$.
\end{proposition}
\begin{proof}
We prove the lemma only in the case $G=\GL$, the modifications needed to treat the case $G=\OO$ are minor.
We can assume that $f$ is multilinear.
Since $f[ns]$ is a $I_n\tnz \GL_s$-concomitant, $f[ns]\in \Rns$ {by Proposition \ref{partinv}}.
We can view $f[ns]$ as an element in $M_n(\F)^{\tnz d+1}\tnz (M_s(\F)^{\tnz d})^*\tnz M_s(\F)$ and write it in the form
$$f[ns]=\sum_{I,J}e_{i_1j_1}\tnz\cdots\tnz e_{i_dj_d}\tnz e_{i_{d+1}j_{d+1}}\tnz \tau^{(s)}_{IJ},$$
where $\tau^{(s)}_{IJ}$ is a $\GL_s$-concomitant.
Let $s>d$.
Since $f$ preserves direct sums we have
$$f[ns](X)\oplus f[ns](Y)=f[2ns](X\oplus Y).$$
We obtain for all $I,J$ an identity
\beq\label{n2n}
\tau^{(s)}_{IJ}(X)\oplus \tau^{(s)}_{IJ}(Y)=\tau^{(2s)}_{IJ}(X\oplus Y).
\eeq
Let us fix $I,J$. To simplify the notation we write $\tau^{(s)}$ instead of $\tau^{(s)}_{IJ}$.
We have
$$\tau^{(s)}=\sum_M h_M^{(s)} M,$$
where $h_M$ is a pure trace polynomial, $M$ is a monomial in the variables $x_k$, and $\deg M+\deg h_M=d$.
Then the identity \eqref{n2n} together with the fact that there are no trace identities of $M_s(\F)$ of degree $<s$ yields
$$h_M^{(s)}(X)=h_M^{(2s)}(X\oplus Y)=h_M^{(s)}(Y)$$
for all monomials $M$, which implies that \[\tau^{(s)}=\sum_M\alpha_M M\] for some $\alpha_M\in \F$.
Thus, $f[ns]\in \GMn$ for every $s>d$ is represented by the same \Rp $\td f$.
Since $f$ respects direct sums, we can identify it with $\td f$.
\end{proof}
For a subset $B$ of $M_n(\F)$ we denote by $C(B)$ its {\em centralizer} in $M_n(\F)$; i.e.,
\[
C(B)=\{c\in M_n(\F)\mid cb=bc \text{ for all $b\in B$}\},
\]
while $C_{G_n}(B)$ stands for $ C(B)\cap G_n$.
We say that a map $f: \cM_n(\F)^g\to \cM_n(\F)$ is a {\em \cg} if
$f[ns]$ is a $(C_{G_n}(B)\tnz M_s(\F))\cap G_{ns}$-concomitant for every $s\in \N$.
\begin{lemma}\label{centGL}
Let $B$ be a subalgebra of $M_n(\F)$.
If $f: \cM_n(\F)^g\to \cM_n(\F)$ is a homogeneous polynomial map of degree $d$ that is a \cgl, then $f\in C(C(B))\ast \F\X$.
\end{lemma}
\begin{proof}
By Lemma \ref{freeRum}, $f\in M_n(\F)\ast \F\X$.
Since $\GL_n$ is dense in $M_n(\F)$, the vector space spanned by $C_{\GL_n}(B)$ coincides with $C(B)$.
Thus we can choose a basis $\{c_1,\dots,c_t\}$ of $C(B)$ with $c_\ell\in \GL_n$.
Let $\{b_1,\dots,b_u\}$ be a basis of $C(C(B))$ and complete it to a basis $\{b_\ell\mid1\leq \ell\leq n^2\}$ of $M_n(\F)$.
We can write $f$ uniquely as
$$f=\sum_{I,K} \alpha_{IK} b_{i_1}x_{k_1} b_{i_2}\cdots x_{k_d} b_{i_{d+1}},$$
where $I$ runs over all $d+1$-tuples of elements in $\{1,\dots,n^2\}$, and $K$ over all $d$-tuples of elements in $\{1,\dots,g\}$.
Take $s>d$ and evaluate $f$ on $M_{2nts}(\F)\cong M_n(\F)\otimes M_{2t}(\F)\otimes M_s(\F)$.
Note that $f$ on $M_{2nts}(\F)$ can be identified with the evaluation of the generalized polynomial
\[
\sum_{I,K} \alpha_{IK} \Big(\sum_{i=1}^{2t}b_{i_1}\tnz e_{ii}\Big)x_{k_1} \Big(\sum_{i=1}^{2t}b_{i_2}\tnz e_{ii}\Big)\cdots x_{k_d} \Big(\sum_{i=1}^{2t}b_{i_{d+1}}\tnz e_{ii}\Big).
\]
in $M_{2nt}(\F)\ast F\X=(M_n(\F)\tnz M_{2t}(\F))\ast \F\X$,
and every element in $M_{2nt}(\F)\ast \F\X$ has a unique expression with the matrix coefficients $b_\ell\tnz e_{ij}$, $1\leq i,j\leq 2t$, $1\leq \ell\leq n^2$, on $M_{2nts}(\F)$ as $s>d$.
Let
\[
\s=\Big(\alpha1\tnz 1+\beta\sum_{\ell=1}^t(c_\ell\tnz e_{\ell,t+\ell}-c_\ell^{-1}\tnz e_{t+\ell,\ell})\Big)\tnz 1\in (C_{\GL_n}(B)\tnz M_{2t}(\F)\tnz M_s(\F))\cap \GL_{2nts}
\]
for $\alpha^2+\beta^2=1$, $\alpha,\beta\in \RR$.
Note that
\beq\label{centelem}
\s^{-1}=\Big(\alpha1\tnz 1-\beta\sum_{\ell=1}^t(c_\ell\tnz e_{\ell,t+\ell}-c_\ell^{-1}\tnz e_{t+\ell,\ell})\Big)\tnz 1.
\eeq
Since $f$ is a \cgl we have
\[
\sum_{I,K} \alpha_{IK} b_{i_1}^\s x_{k_1} b_{i_2}^\s\cdots x_{k_d} b_{i_{d+1}}^\s
=\sum_{I,K} \alpha_{IK} b_{i_1}x_{k_1} b_{i_2}\cdots x_{k_d} b_{i_{d+1}},
\]
where by a slight abuse of notation $b_{i}$ denotes $b_i\tnz 1\tnz 1$, and
\begin{multline}\label{eq:no1}
b_i^\s=\s^{-1}b_i\s= \alpha^2b_{i}\tnz 1\tnz 1+\sum_{\ell=1}^t\beta^2c_\ell b_{i}c_\ell^{-1}\tnz e_{\ell\ell}\tnz 1+\beta^2c_\ell^{-1}b_{i}c_\ell\tnz e_{t+\ell,t+\ell}\tnz 1\\
+\alpha\beta( b_{i}c_\ell-c_\ell b_{i})\tnz e_{\ell,t+\ell}\tnz 1-\alpha\beta(b_i c_\ell^{-1}-c_\ell^{-1}b_i)\tnz e_{t+\ell,\ell}\tnz 1.
\end{multline}
Since $s>d$ both sides of equation \eqref{eq:no1} have a unique expression as generalized polynomials in $M_{2tn}\ast \F\X$
with the generalized coefficients $b_\ell\tnz e_{ij}$, $1\leq i,j\leq 2t$, $1\leq \ell\leq n^2$.
We thus derive
\beq\label{eq:no2}
\sum_k \alpha_{I^j_kK}(b_kc_\ell-c_\ell b_k)=0
\eeq
for every $1\leq j\leq d+1$, $1\leq \ell\leq t$, where $I^j_k$ denotes a tuple of $d+1$-elements in $\{1,\dots,n^2\}$ with $k$ at the $j$-th position.
Equation \eqref{eq:no2} implies that
\[
\sum_k \alpha_{I^j_kK}b_k\in C(C(B)),
\]
which is by the choice of $b_\ell$, $1\leq \ell\leq n^2$, only possible if $\alpha_{I^j_kK}=0$ for $b_k\not\in C(C(B))$. Therefore we have $f\in C(C(B))\ast \F\X$.
\end{proof}
\begin{lemma}\label{preCentO}
If $B$ is a $*$-subalgebra of $M_n(\RR)$,
then the subalgebra generated by $C_{\OO_n}(B)$ is equal to $C(B)$, and $C(C_{\OO_n}(B))=C(C(B))=B$.
\end{lemma}
\begin{proof}
Since $B$ is a $*$-subalgebra of $M_n(\RR)$, $C(B)$ is also a $*$-subalgebra of $M_n(\RR)$, thus semisimple.
Notice that in order to show that $\RR\langle C_{\OO_n}(B)\rangle$, the subalgebra of $C(B)$ generated by $C_{\OO_n}(B)$, coincides with $C(B)$,
we can assume that $C(B)$ is simple.
We have $c^t-c\in {\rm span}\, C_{\OO_n}(B)$, the vector subspace of $M_n(\RR)$ spanned by $C_{\OO_n}(B)$, for every $c\in C(B)$.
Indeed, $e^{\lambda (c^t-c)}\in C_{\OO_n}(B)$ for every $\lambda\in \RR$, $c\in C(B)$ yields $c^t-c\in {\rm span}\,C_{\OO_n}(B)$.
If $C(B)$
is isomorphic to $\RR$, $M_2(\RR)$, $\CC$, or $M_2(\CC)$, where the involution on $\CC$ is the complex conjugation,
then one can easily verify that ${\rm span}\,C_{O_n}(B)=C(B)$.
Recall that a finite dimensional
simple $\RR$-algebra with involution
which is not
isomorphic to $\RR$, $M_2(\RR)$, $\CC$, or $M_2(\CC)$
coincides with
its subalgebra generated by the skew-symmetric elements (see e.g.~\cite[Lemma 2.26]{KMRT}).
Therefore $\RR\langle C_{\OO_n}(B)\rangle=C(B)$, which further implies $C(C_{\OO_n}(B))=C(C(B))$, and
the identity $C(C(B))=B$ follows from the double centralizer theorem (see e.g.~\cite[Theorem 1.5]{KMRT}).
\end{proof}
\begin{lemma}\label{centO}
Let $B$ be a $*$-subalgebra of $M_n(\RR)$.
If $f: \cM_n(\RR)^g\to \cM_n(\RR)$ is a homogeneous polynomial map of degree $d$ that is a \co, then $f\in B\ast \RR\Xt$.
\end{lemma}
\begin{proof}
Since the proof is similar to that of Lemma \ref{centO} we omit some of the details.
By Proposition \ref{partinvO} we have $f\in M_n(\RR)\ast \RR\Xt$.
Let $c_1,\dots,c_t$ be a basis of ${\rm span} \,C_{\OO_n}(B)$, the vector space spanned by $C_{\OO_n}(B)$, with $c_\ell\in \OO_n$.
Let us write
$$f=\sum_{I,K} \alpha_{IK} b_{i_1}u_{k_1} b_{i_2}\cdots u_{k_d} b_{i_{d+1}},$$
where $u_k\in \{x_k,x_k^t\}$.
Take $s>d$ and evaluate $f$ on $M_{2nts}(\F)$.
Let
\[
\s=\Big(\alpha1\tnz 1+\beta\sum_{\ell=1}^t(c_\ell\tnz e_{\ell,t+\ell}-c_\ell^t\tnz e_{t+\ell,\ell})\Big)\tnz 1\in (C_{\OO_n}(B)\tnz M_{2t}(\F)\tnz M_s(\F))\cap \OO_{2nts}
\]
for $\alpha^2+\beta^2=1$, $\alpha,\beta\in \RR$.
Note that $\s\in \OO_{2nts}$ and
\beq\label{centelem'}
\s^t=\Big(\alpha1\tnz 1-\beta\sum_{\ell=1}^t(c_\ell\tnz e_{\ell,t+\ell}-c_\ell^t\tnz e_{t+\ell,\ell})\Big)\tnz 1.
\eeq
Since $f$ is a \co we have
\[
\sum_{I,K} \alpha_{IK} b_{i_1}^\s u_{k_1} b_{i_2}^\s\cdots u_{k_d} b_{i_{d+1}}^\s
=\sum_{I,K} \alpha_{IK} b_{i_1}u_{k_1} b_{i_2}\cdots u_{k_d} b_{i_{d+1}},
\]
where $b_{i}$ denotes $b_i\tnz 1\tnz 1$, and
\begin{multline*}
b_i^\s=\s^{t}b_i\s= \alpha^2b_{i}\tnz 1\tnz 1+\sum_{\ell=1}^t\beta^2c_\ell b_{i}c_\ell^{t}\tnz e_{\ell\ell}\tnz 1+\beta^2c_\ell^{t}b_{i}c_\ell\tnz e_{t+\ell,t+\ell}\tnz 1+\\
+\alpha\beta( b_{i}c_\ell-c_\ell b_{i})\tnz e_{\ell,t+\ell}\tnz 1-\alpha\beta(b_i c_\ell^{t}-c_\ell^{t}b_i)\tnz e_{t+\ell,\ell}\tnz 1.
\end{multline*}
As $s>d$ both sides of the last identity have a unique expression as generalized polynomials in $M_{2tn}\ast \RR\Xt$
with the generalized coefficients $b_\ell\tnz e_{ij}$, $1\leq i,j\leq 2t$, $1\leq \ell\leq n^2$.
Thus, $\alpha_{I^j_kK}=0$ for $b_k\not\in C(C_{\OO_n}(B))$, where $I^j_k$ denotes a tuple of $d+1$-elements in $\{1,\dots,n^2\}$ with $k$ at the $j$-th position.
Since $C(C_{\OO_n}(B))=B$ by Lemma \ref{preCentO}, $f$ belongs to $B\ast \RR\Xt$.
\end{proof}
\subsection{Power Series Expansions about Non-Scalar Points}\label{s2sec4}
{We next turn to analytic free maps and exhibit
their power series expansions about a non-scalar point $A$. Homogeneous components of such an expansion will be \Rps.} For $G=\GL$ their matrix coefficients belong to the double centralizer $C(C(A))$, while for $G=\OO$ they lie in the $*$-subalgebra $\F\langle A,A^t\rangle$ generated by $A$.
Let us first introduce neighbourhoods of non-scalar points.
Given $A\in M_n(\F)^g$, set
$$\cB(A,\delta)=\bigcup_{s=1}^\infty\big\{X\in M_{ns}(\F)^g\mid \big\|X-\bigoplus_{i=1}^s A\big\|<\delta_s\big\},$$
where $\delta=(\delta_s)_{s\in \N}$, $\delta_s>0$ for every $s\in \N$.
\subsubsection{$GL$-free maps}\label{s1s2sec4}
The next theorem gives a power series expansion of a $\GL$-free map $f$ about $A=(A_1,\dots,A_g)\in M_n(\F)^g$, whose
matrix coefficients are elements of the double centralizer
algebra $C(C(\F \langle A\rangle))\subseteq M_n(\F)$ of the subalgebra $\F\langle A\rangle$ generated by $A_1,\dots,A_g$.
\begin{theorem}\label{rumanalit}
Let $\cU$ be a $\GL$-free set, $f:\cU\to\cM(\F)$ be an $\F$-analytic $\GL$-free map, and let $\cB(A,\delta)\subseteq \cU$, where $A\in M_n(F)^g$, and $\delta=(\delta_s)_{s\in\N}$, $\delta_s>0$ for every $s\in \N$.
Then there exist unique generalized polynomials $f_m\in C(C(\F\langle A))\rangle\ast \F\X$ of degree $m$
so that the formal power series
\beq\label{eq:ps}
F(X)=\sum_{m=0}^\infty f_m(X-A),
\eeq
converges in norm on the neighbourhood $\cB(A,\delta)$ of $A$ to $f$.
\end{theorem}
\begin{proof}
As $A\in \cU[n]$ and $\cU$ is a $\GL$-free set we have
\[A^{\oplus s}=\bigoplus_{i=1}^s A\in \cU[ns]\] for every $s\in \N$.
Since $f[ns]$ is analytic in a neighbourhood of $A^{\oplus s}$, the function
$$t\mapsto f[ns]\Big(A^{\oplus s}+t\big(X-A^{\oplus s}\big)\Big)$$
is defined and analytic for all $|t|<\delta_X$, where $\delta_X$ depends on $X\in M_{ns}(\F)$.
Thus, we can expand it in a power series
\beq\label{homt}
f[ns]\Big(A^{\oplus s}+t\big(X-A^{\oplus s}\big)\Big)=\sum_{m=0}^\infty t^mf[ns]_{m}\big(X-A^{\oplus s}\big)
\eeq
that converges for $|t|<\delta_X$.
If $X\in \cB(A,\delta)$, then we have $\delta_X\geq1$.
We claim that
$f[ns]_{m}$ is a homogeneous polynomial function of degree $m$.
Indeed, as
$$\sum_{m=0}^\infty t_1^mf[ns]_{m}\Big (t_2\big(X-A^{\oplus s}\big)\Big) =f[ns]\Big(A^{\oplus s}+t_1t_2\big(X-A^{\oplus s}\big)\Big)=\sum_{m=0}^\infty t_1^mt_2^mf[ns]_{m}\big(X-A^{\oplus s}\big)$$
for all $t_1$ that satisfy $|t_1|, |t_1t_2|<\delta_X$, we obtain \[f[ns]_{m}(tY)=t^mf[ns]_m(Y)\] for all $t\in \F$, $Y\in M_{ns}(\F)^g$.
Let us show that
$$f_{m}:\cM_{n}(\F)^g\to \cM_{n}(\F)$$
defined by
$f_{m}[ns]:=f[ns]_{m}$
is a \cgl
that preserves direct sums.
Take $s\in \N$, $\sigma\in (C_{\GL_n}(F\langle A\rangle)\tnz M_s(\F))\cap \GL_{ns}$
and note that
\[\s A^{\oplus s}\s^{-1}=A^{\oplus s}.\]
Then the identity
\begin{eqnarray*}
\sum t^m \s f[ns]_{m}\big(X-A^{\oplus s}\big) \s^{-1}&=& \s f[ns]\Big(A^{\oplus s}+t\big(X-A^{\oplus s}\big)\Big) \s^{-1}\\
&=&f[ns]\Big(A^{\oplus s}+t\big(\s X\s^{-1}-A^{\oplus s}\big)\Big)\\
&=&\sum t^m f[ns]_{m}\Big(\s\big( X-A^{\oplus s}\big)\s^{-1}\Big),
\end{eqnarray*}
for all small enough $t$
yields the desired conclusion.
To conclude the proof of the existence we proceed as at the end of the proof of existence in
Theorem \ref{analit}.
Thus, $f_{m}\in C(C(\F\langle A))\rangle\ast \F\X$
by Lemma \ref{centGL}.
Note that setting $t=1$ in \eqref{homt} establishes the existence of the desired power series.
For the uniqueness, we can also follow the proof of uniqueness in Theorem \ref{analit} carried out in Lemma \ref{matenote} and Proposition \ref{unique}, after recalling the identity \eqref{redfun}.
Hence we can recover $f_m$ by the $m$-th derivative of the function $t\mapsto f[n(m+1)](t(X-A))$ at $0$, and
the matrix coefficients of the generalized polynomial $f_m$ can be determined by evaluations on $M_{n(m+1)}(\F)$.
\end{proof}
\begin{remark}
If $f$ is a uniformly bounded $\GL$-free map then the convergence of $F$ in
\eqref{eq:ps} is uniform, which can be proved in the same way as the analogous statement for $\F=\CC$ and power series expansion about scalar points in the last part of the proof of \cite[Proposition 2.24]{HKM12}.
The only modification needed is to replace $\exp({\mathbbm{i}t})I_{ns},\exp(-\mathbbm{i}mt)I_{ns}\in M_{ns}(\CC)$ in the equation
\[
C\geq \Big\|\frac{1}{2\pi}\int f(\exp(\mathbbm{i}t)X)\exp(-\mathbbm{i}mt)dt\Big\|=\|f^{(m)}(X)\|
\]
with the corresponding matrices in $M_{2ns}(\RR)$.
\end{remark}
In general one cannot expect the matrix coefficients of the power series expansion of a $\GL$-free map $f$ about a non-scalar point $A$ to lie in $\F\langle A\rangle\ast\F\X$.
In this case one would have $f(A)\in \F\langle A\rangle$, which is not always the case by \cite[Theorem 7.7]{AM2}.
However, this does hold true in the case that $A$ is a generic point.
That is, if $g=1$, then $A$ is similar to a diagonal matrix with $n$ distinct eigenvalues, and
if $g>1$ then $\F\langle A\rangle=M_n(\F)$.
\begin{corollary}
Let $\cU$ be a $\GL$-free set, $f:\cU\to\cM(\F)$ be an $\F$-analytic $\GL$-free map, and let $\cB(A,\delta)\subseteq \cU$, where $A\in M_n(F)^g$ is a generic point, and $\delta=(\delta_s)_{s\in\N}$, $\delta_s>0$ for every $s\in \N$.
Then there exist generalized polynomials $f_m\in \nA$ of degree $m$
so that the formal power series
\[
F(X)=\sum_{m=0}^\infty f_m(X-A),
\]
converges in norm on the neighbourhood $\cB(A,\delta)$ of $A$ to $f$.
\end{corollary}
\subsubsection{$\OO$-free maps}
In the case of free maps with involution the matrix coefficients in the power series expansion of an $\OO$-free map about $A=(A_1,\dots,A_g)\in M_n(\F)^g$ lie in the $*$-subalgebra $\F\langle A,A^t\rangle$ of $M_n(\F)$ generated by $A_1,\dots,A_g$.
This contrasts the analogous result for $\GL$-free maps (Theorem \ref{rumanalit}) where the double centralizer of $\F\langle A\rangle$ is required.
\begin{theorem}\label{rumanalitO}
Let $\cU$ be an $\OO$-free set, $f:\cU\to\cM(\F)$ be an $\F$-analytic $\OO$-free map, and let $\cB(A,\delta)\subseteq \cU$, where $A\in M_n(F)^g$, and $\delta=(\delta_s)_{s\in\N}$, $\delta_s>0$ for every $s\in \N$.
Then there exist unique generalized polynomials $f_m\in\F\langle A,A^t\rangle\ast\F\Xt$ of degree $m$
so that the formal power series
\[
F(X)=\sum_{m=0}^\infty f_m(X-A),
\]
converges in norm on the neighbourhood $\cB(A,\delta)$ of $A$ to $f$.
\end{theorem}
\begin{proof}
The proof resembles that of Theorem \ref{rumanalit} with obvious modifications. One only needs to apply Lemma \ref{centO} instead of Lemma \ref{centGL}.
\end{proof}
\section{Inverse Function Theorem for Free Maps}\label{sec5}
As an application of the tools and techniques developed we present an inverse and implicit function theorem for free maps.
For $G=\GL$ these results have been obtained by Pascoe \cite{Pas}, Agler and McCarthy \cite{AM2},
Kaliuzhnyi-Verbovetskyi and Vinnikov (private communication).
Following \cite{KV} we recall two {topologies on $\cM(\F)^g$.}
The first is the {\bf finitely open topology}. Its basis are open sets $U$ such that the intersection of $U$ with $M_n(\F)^g$ is open for every $n\in \N$. The second topology is the {\bf uniformly open topology} and its basis consists of sets of the form
$${\mathcal B}(A,r)=\bigcup_{s=1}^\infty\big\{X\in M_{ns}(\F)^g\mid \big\|X-\bigoplus_{i=1}^s A\big\|<r\big\},$$
for $A\in M_n(\F)^g$, $n\in \N$, $r\geq 0$.
Further topologies in this free context are considered in \cite{AM1,AM2}.
Let us recall a version of the classical inverse function theorem, giving information on the injectivity domain (see e.g.~\cite[Theorem \rom{14}.1.2]{Lan}, \cite[Theorem 2.5.1]{KrPa}, \cite[Theorem 0.8.3]{KK}). We state it only in the case when $f:\cU\to V$ for $\cU\subset V$, $0$ is in the domain of $f$, $f(0)=0$, ${\rm D}f(0)=\id_V$,
to which the general case can be reduced by replacing the function $f:\cU\to V$ with the function $\ol{f}(x)={\rm D}f(x_0)^{-1}(f(x+x_0)-f(x_0))$, if $x_0$ is the point in the domain of $f$.
Here $\rm D$ denotes the Fr\'echet derivative. We say that $f\in \cC^r$ if all ${\rm D}^kf$, $1\leq k\leq r$,
exist and are continuous.
\begin{theorem}
\label{IFT}
Let $V$ be a Banach space, $\cU \subset V$ an open set containing $0$, $f:\cU\to V$, and let $f\in \cC^r$ for some $r\in \N$ $($resp. $f$ is analytic$)$.
Let ${\rm D}f(0):V\to V$ be a continuous bijective linear map.
If ${\rm Ball}(0,2\delta)\subseteq\cU$ and $\|{\rm D}(x-f(x))\|<\frac{1}{2}$ for $\|x\|<2\delta$, then $f$ is injective on ${\rm Ball}(0,\delta)$, and there exists $h:{\rm Ball}(0,\frac{\delta}{2})\to \cV$, where $\cV$ is an open subset of ${\rm Ball}(0,\delta)$, such that $hf=\id_\cV$, $fh=\id_{{\rm Ball}(0,\frac{\delta}{2})}$, and $h\in \cC^r$ $($resp. $h$ is analytic$)$.
\end{theorem}
With a slight abuse of notation, we call a $g'$-tuple of $G$-free maps $f=(f_1,\dots,f_{g'})$, $f_i:\cU \to\cM(\F)$, also a $G$-free map.
Throughout this section we let $G\in \{\GL,\OO\}$.
\subsection{Uniformly Open Topology}
In this subsection we work with the uniformly open topology. The Fr\'echet derivative ${\rm D}f$ is continuous in the uniformly open topology at $A\in M_n(\F)^g$ if for every $\varepsilon>0$ there exists $\delta >0$ such that $\|{\rm D}f(X)-{\rm D}f(A^{\oplus s})\|<\varepsilon$ if $s\in \N$ and $X\in \cB(A,\delta)[ns]$.
\begin{theorem}[Inverse free function theorem]\label{IFFT}
Let $\cU\subset \cM(\F)^g$ be an open $G$-free set containing $0$, $f:\cU\to \cM(\F)^{g'}$ a $G$-free map, and let $f\in \cC^r$
for $r\in \N$ $($resp. $f$ analytic$)$, with ${\rm D}f(0)$ invertible as a continuous linear map.
Then there exist open $G$-free sets $\cW\subset \cM(\F)^{g}$, $\cW'\subset \cM(\F)^{g'}$ containing $0$, $f(0)$ respectively, and a $G$-free map $h:\cW'\to \cW$ so that
$fh=\id_{\cW'}$, $hf=\id_{\cW}$, and $h\in \cC^{r}$ $($resp. $h$ analytic$)$.
Moreover, $h$ is analytic for every $r\in \N$ in the case $G=\GL$.
\end{theorem}
\begin{proof}
Since $\cM(\F)^g$ is not a Banach space we cannot directly apply Theorem \ref{IFT}.
However, we can use it levelwise.
Without loss of generality we can assume that $g=g'$, $f(0)=0$ and ${\rm D}f(0)=\id_{\cM(\F)^g}$ by replacing $f$ with the function
$$\ol f:\cM(\F)^g\to \cM(\F)^g,\quad \ol f={\rm D}f(0)^{-1}(f-f(0)).$$
As ${\rm D}f$ is continuous on $\cU$ and invertible at $0$ with a continuous inverse in the uniformly open topology,
there exists (by the definition of the topology) $\delta>0$ such that $\cB(0,2\delta)\subseteq \cU$ and $\|{\rm D}(x-f(x))\|<\frac{1}{2}$ for $\|x\|<2\delta$.
Theorem \ref{IFT} therefore implies that $f$ is injective on $\cB(0,\delta)$, and
provides
a $\cC^r$-map $h:\cB(0,\frac{\delta}{2})\to \cV$, where $\cV$ is an open subset of $\cB(0,\delta)$, that satisfies the desired identities.
Let us first show that $\cV$ is an $\OO$-free set and $h$ is an $\OO$-free map.
Let $u\in \OO_n$, $Y\in \cB(0,\frac{\delta}{2})[n]$.
As $uYu^t\in \cB(0,\frac{\delta}{2})[n]$ and $f$ is a $G$-free map we have
\beq\label{Ukon}
f\big(h(uYu^t)\big)=uYu^t=uf\big(h(Y)\big)u^t=f\big(uh(Y)u^t\big).
\eeq
Since $uh(Y)u^t\subset u\cV u^t\subset \cB(0,\delta)$ and $f$ is injective on $\cB(0,\delta)$, $h$ respects $\OO$-similarity.
In the same way one can show that $h$ respects direct sums, so it is indeed an $\OO$-free map.
In consequence, $\cV=h(\cB(0,\frac{\delta}{2}))$ is an $\OO$-free set.
Thus, in the case $G=\OO$, the proposition follows.
It remains to consider the case $G=\GL$.
We claim that $h$ is analytic in this case. In the case $\F=\CC$, $f$ is analytic (see \cite[Proposition 2.5]{HKM11} or \cite[Theorem 7.2]{KV}).
Our assumptions imply that $f$ is (uniformly) bounded in
$\cB(0,\delta)$, therefore we can apply \cite[Theorem 7.23, Remark 7.35]{KV} to deduce that $f$ is analytic also in the case $\F=\RR$. Thus, $h$ is analytic by Theorem \ref{IFT}.
Since $h$ is an $\OO$-free map according to the previous paragraph, it can be expanded in a power series \eqref{eq:pw} in $x,x^t$ about $0$ by Theorem \ref{analit}, which converges in $\cB(0,\frac{\delta}{2})$.
Note that \eqref{Ukon} holds also if we replace $u,u^t$ by $\s,\s^{-1}$ respectively, for $\s\in \GL_n$ such that
$\s Y\s^{-1}\in \cB(0,\frac{\delta}{2})$, $\s h(Y)\s^{-1}\in \cB(0,{\delta})$.
Note that for every $Y\in \cB(0,\frac{\delta}{2})$ there exists $\delta_\s>0$,
such that $t\s Y\s^{-1}\in \cB(0,\frac{\delta}{2}),\s h(tY)\s^{-1}\in \cB(0,\delta)$ for every $|t|<\delta_\s$. Thus,
\[
h(\s tY\s^{-1})=\s h(tY)\s^{-1}
\]
for every $|t|<\delta_\s$. Writing this identity as a power series in $t$, we can deduce that each homogeneous part $h_m$ of the power series $H$ of $h$ is a $\GL$-concomitant. Thus, $H$ is a power series in $x$, and $h$ is a $\GL$-free map on $\cB(0,\frac{\delta}{2})$.
Now notice that the $\GL$-similarity invariant envelopes
\[
\cW=\wtd\cV,\quad\cW'=\wtd{\cB(0,\frac{\delta}{2})}
\]
are open sets since the function $X\mapsto \s X\s^{-1}$ is an (analytic) isomorphism.
As $\cU$ is a $G$-free set, $\cW$ is contained in $\cU$.
Furthermore, $\td h$ (cf. Proposition \ref{simenv}) maps $\cW'$ to $\cW$.
Thus, we only need to check that $ f$ and $\td h$ satisfy the desired identities.
Let $\td X=\s X\s^{-1}\in \cW$, where $X\in \cV[n], \s\in \GL_n$. Then
$$\td h\Big(f\big(\s X\s^{-1}\big)\Big)=\td h\big(\s f(X)\s^{-1}\big)=\s h(f(X))\s^{-1}=\s X\s^{-1}$$
implies that $\td h f=\id_{\cW}$. The identity $f\td h=\id_{\cW'}$ can be checked similarly.
\end{proof}
The proof used in the classical setting to derive the implicit function theorem from the inverse function theorem can be also
utilized in the free setting. Thus, we obtain an implicit free function theorem.
We denote by ${\rm D}_2f(a,b)$, where $f:\cU\times \cV\to \cW$, and $(a,b)\in \cU\times \cV$, the Fr\'echet derivative of the function $y\mapsto f(a,y)$ evaluated at $b$.
\begin{corollary}[Implicit free function theorem]\label{IFFT2}
Let $\cU_1\times \cU_2\subseteq \cM(\F)^g\times \cM(\F)^{g'}$ be an open $G$-free set, $f:\cU_1\times \cU_2\to \cM(\F)^{g'}$ a $G$-free map, and let $f\in \cC^r$
for some $r\in \N$, with ${\rm D_2}f(0,0)$ invertible.
There exist an open $G$-free set $\cV_1\times\cV_2$ containing $(0,0)$, and a $G$-free map $h:\cV_1\to \cV_2 $, $h\in\cC^r$, such that
$f(x,y)=0$ for $(x,y)\in \cV_1\times \cV_2$ if and only if $y=h(x)$.
\end{corollary}
We now turn our attention to the inverse function theorem about neighbourhoods of non-scalar points.
Let us denote
$$C_{G}(A)=\{\s\in G_n\mid \s A_i=A_i\s, \,1\leq i\leq g\}$$
for $A=(A_1,\dots,A_g)\in M_n(\F)^g$.
We say that $\cU\subset \cM_n(\F)$ is a $C_G(A)\otimes G$-free set if it is closed under direct sums and simultaneous $C_G(A)\otimes G$-similarity.
By
$$\td\D f(A):\cM_n(\F)^g\to \cM_n(\F)^{g'}$$
for $f:\cU\to \cM_n(\F)^{g'}$, $A\in\cU\subseteq \cM_n(\F)^g$, we denote the linear map defined levelwise for every $s\in \N$ as
\[\td\D f(A)[ns](H):=\D f(A^{\oplus s})(H).\]
The next theorem generalizes Theorem \ref{IFFT} to the case of non-scalar center points.
\begin{theorem}\label{IFFTn}
Let $\cU\subset \cM(\F)^g$ be an open $G$-free set, $A\in \cU[n]$, $f:\cU\to \cM(\F)^{g'}$ a $G$-free map, and let $f\in \cC^r$
for $r\in \N$, with ${\td\D}f(A)$ invertible as a continuous linear map.
There exist open $C_G(A)\otimes G$-free sets $\cW\subset \cM_n(\F)^{g}$, $\cW'\subset \cM_n(\F)^{g'}$ containing $A$, $f(A)$ respectively,
and a $C_G(A)\otimes G$-free map $h:\cW'\to \cW$ so that
$fh=\id_{\cW'}$, $hf=\id_{\cW}$, and $h\in \cC^{r}$.
\end{theorem}
\begin{proof}
Note that
\[\D f(\s X\s^{-1})(\s H\s^{-1})=\s \D f(X)(H)\s^{-1}\]
for every $X,H\in M_n(\F)^g,\s\in G_n,n\in \N$. Since $A\in \cU$, which is an open $G$-free set, there exists $\delta>0$ such that $\cB(A,\delta)\subset \cU$. Then the function $\ol f:\cB(0,\delta)\cap\cM_n(\F)^g\to \cM_n(\F)^g$ defined by
\[{\ol f}[ns]:\cB(0,\delta)\cap M_{ns}(\F)^g\to M_{ns}(\F)^g,\quad {\ol f}[ns](X):=\D f\big (A^{\oplus s}\big)^{-1}\Big(f\big(X+A^{\oplus s}\big)-f\big(A^{\oplus s}\big)\Big )\]
is $C_G(A)\otimes G$-free with ${\ol f}(0)=0$, $\D {\ol f}(0)=\id_{\cM_n(\F)}$.
A similar reasoning to that in the proof of Theorem \ref{IFFT} with obvious modifications and using Theorem \ref{rumanalit} in the place of Theorem \ref{analit} now
yields the desired conclusions.
\end{proof}
\subsection{Finitely Open Topology}
Now we state a weak form of the inverse function theorem for the finitely open topology. The Fr\'echet derivative ${\rm D}f$ is continuous in the finitely open topology if ${\rm D}f[n]$ is continuous for every $n\in \N$.
\begin{proposition}\label{foi}
Let $\cU\subseteq \cM(\F)^g$ be an open $G$-free set, $f:\cU\to \cM(\F)^{g'}$ a $G$-free map, and let $f\in \cC^r$
for some $r>0$ with ${\rm D}f(0)$ be invertible.
There exist finitely open sets $\cW,\cV$, containing $0$, $f(0)$ respectively, and a free $\OO$-concomitant map $h:\cV\to \cW$ such that
$fh=\id_{\cV}$, $hf=\id_{\cW}$, and $h\in \cC^{r}$. In the case $\F=\CC$, $h$ is a a free $G$-concomitant map.
\end{proposition}
\begin{proof}
By the classical inverse function theorem we can find for every $n\in \N$
neighbourhoods $\cV_n$, $\cB(0,\delta_n)$ of $0$, $f[n](0)$ respectively, such that
$f[n]:\cV_n\to{\cB(0,\delta_n)}$ is a diffeomorphism with the inverse $h[n]\in \cC^r$.
Since $\cB(0,\delta_n)$ is $\OO_n$-invariant so is $\cV_n$ for every $n\in \N$.
As in the proof of Theorem \ref{IFFT} it is easy to show that $h(uYu^t)=uh(Y)u^t$ for every $u\in \OO_n$, $Y\in \cV_n$.
By the definition of the finitely open topology, the sets $\cV=\bigcup_n \cV_n$, $\cW=\bigcup_n\cB(0,\delta_n)$
are finitely open. This establishes the proposition in the case $G=\OO$. In the case $G=\GL_n$ and $\F=\CC$ we proceed as in the proof of Theorem \ref{IFFT}, and replace $\cV$, $\cW$ by $\wtd \cV$, $\wtd \cW$ respectively.
To show that $f,\td h$ satisfy the required identities one also only needs to follow the steps in the proof of Theorem \ref{IFFT}.
\end{proof}
We do not know whether $\cW$ and $\cV$ in Proposition \ref{foi} can be taken to be $G$-free sets, and consequently $h$ would be a $G$-free map; cf. \cite[Section 8]{AM2}.
\subsection{Global Free Inverse Function Theorem}
In \cite[Theorem 1.1]{Pas} it is proved that if $f$ is a $\GL$-free map and ${\rm D}f(X)$ is nonsingular for every $X\in \cM(\CC)$ then $f$ is injective, cf. \cite{AM2}. This also holds for $\OO$-free maps.
\begin{proposition}\label{GIFFT}
If $f:\cM(\F)^g\to \cM(\F)^{g'}$ is a differentiable $G$-free map such that ${\rm D}f(X)$ is nonsingular for every $X\in \cM(F)$ then $f$ is injective. If $f\in \cC^r$ for some $r\in \N$ then there exists a $G$-free map $h:f(\cM(\F)^g)\to \cM(\F)^{g'}$,
$h\in \cC^r$, such that $hf=\id|_{\cM(\F)^g}$, $fh=\id|_{f(\cM(\F)^g)}$.
\end{proposition}
\begin{proof}
Suppose that $f(X_1)=f(X_2)$ for some $X_1,X_2\in M_n(\F)^g$.
Then \eqref{did} yields
$$
\D f
\begin{pmatrix}
X_1&0\\
0& X_2
\end{pmatrix}
\begin{pmatrix}
0&X_1-X_2\\
X_1-X_2& 0
\end{pmatrix}
=
\begin{pmatrix}
0&0\\
0& 0
\end{pmatrix}.$$
Since $\D f
\begin{pmatrix}
X_1&0\\
0& X_2
\end{pmatrix}$
is nonsingular we have $X_1=X_2$, which implies the injectivity of $f$.
The proof of the existence of the free map $h$ satisfying the required properties is the same as that of Theorem \ref{IFFT}.
\end{proof}
\begin{remark}
We remark that a free real Jacobian conjecture can be deduced from Proposition \ref{GIFFT}.
(See e.g.~\cite[Theorem 1.3]{Pas}.)
\end{remark}
\section{Examples of $\OO$-Free Maps}\label{ex}
{The theory of $\GL$-free maps is very rigid to the point that many properties are stronger
than for complex analytic functions \cite{KV,HKM11,HKM12,Voc10}. In contrast to this is the theory of $\OO$-free maps as we shall now demonstrate. We start by presenting the following examples:
\begin{enumerate}[$\bullet$]
\item
a continuous $\OO$-free map which is not differentiable (Example \ref{cont}); more generally,
\item $C^k$-maps which are not $C^{k+1}$ (Example \ref{k});
\item
a smooth $\OO$-free map which is not analytic (Example \ref{sin}).
\end{enumerate}
}
\begin{example}\label{cont}
Consider the $\OO$-free map {$f_m:\cM(\RR)\to \cM(\RR)$ defined by}
\[f_m(x)=(xx^t)^{\frac{1}{m}} \quad\text{ for some $m\geq 2$}.\]
It is continuous by \cite[Theorem 1.1]{ZZ}.
Note that $f_m$ is not differentiable at $0$.
\end{example}
\begin{example}\label{k}
Let $k\in \N$ and
$$
f:\cM(\RR)\to \cM(\RR)\quad f(x)= (xx^t)^{k+\frac{1}{2}}.
$$
Then $f$ is an $\OO$-free $C^k$-map \cite[Theorem 1.1]{ZZ}, but is not $C^{k+1}$.
\end{example}
\begin{example}\label{sin}
For an example of a smooth nonanalytic $\OO$-free map consider the map
$$
{f:\cM(\RR)\to \cM(\RR), \quad}
f(x)=\nsum_{j=0}^{\infty} e^{\displaystyle -\sqrt{2^j}}\cos\big(2^j (x+x^t)\big).$$
Since $\|\cos(2^j(A+A^t))\|\leq 1$ for every $A\in \cM(\RR)$, the power series is convergent.
We show that there exist {derivatives of all orders} in all directions at all points of $\Mr$, but $f$ is not analytic.
Let us show {first} that $f$ is not analytic at $0$. This holds already for the function $f[1]:\RR\to \RR$.
Indeed, since
$$\limsup_{n\to \infty} \frac{|f[1]^{(n)}(0)|}{n!}\leq\limsup_{n\to \infty}\frac{e^{-\sqrt{n}}n^n}{n!}=\infty,$$
the radius of convergence of the Taylor series of $f[1]$ at $0$ is $0$.
Consider now the $\ell$-th order derivative of the function $x\mapsto \cos(kx)$ at a point $A\in M_n(\RR)$ in the direction $H\in M_n(\RR)$.
We define matrices
$$
A_{H}^\ell=
\begin{pmatrix}
A&H&&\\
&\ddots&\ddots&\\
&&A&H\\
&&&A\\
\end{pmatrix}\in M_{(\ell+1)n}(\RR).
$$
Let $F$ be an analytic function around $0$ with the radius of convergence $\infty$.
The $\ell!$-multiple of the $(1,\ell+1)$-entry of the matrix $F(A_{H}^\ell)$ equals the $\ell$-th order derivative of $F$ at the point $A$ in the direction $H$.
By \cite[Theorem 4.25]{Hig} we have
\[
||\cos(kA_H^\ell)||\leq (\ell+1)n\alpha k^{\ell n},
\]
where $\alpha$ depends only on $A$,
for $A=A^t, H=H^t\in M_n(\RR)$.
This implies that
$$\sum_{j=0}^\infty e^{-\sqrt{2^j}}\Big\|{\delta}^\ell\cos\big(2^j (A+A^t)\big)(H+H^t)\Big\|\leq
(\ell+1)!n\alpha\sum_{j=0}^\infty e^{-\sqrt{2^j}}2^{j\ell n}< \infty.$$
Hence the $\ell$-th order derivative of $f$ at $A$ in the direction $H$ exists and equals
$$\sum_{j=0}^\infty e^{-\sqrt{2^j}}{\delta}^\ell\cos\big(2^j (A+A^t)\big)(H+H^t).$$
\end{example}
Let $f:\cU\to \cM(\CC)$ be an analytic $\GL$-free map.
If $f$ is uniformly bounded on $\cU$ then the $m$-th homogeneous part of the corresponding power series is also uniformly bounded (see e.g.~the last part of the proof of \cite[Proposition 2.24]{HKM12}).
In the case of $\OO$-free maps this is no longer the case.
\begin{example}
The analytic $\OO$-free map
\[
x\mapsto \sin(xx^t)\]
is uniformly bounded on $\cM(\F)$, however its $(4m+2)$-th homogeneous part
\[(-1)^m\frac{1}{(2m+1)!} (xx^t)^{2m+1}\]
is not uniformly bounded.
\end{example}
If an analytic $\GL$-free map $f:\cU\to \cM(\CC)$ is uniformly bounded then it converges uniformly on $\cU$ by \cite[Proposition 2.24]{HKM12}.
The proof of the uniform convergence is easily established after noticing that the homogeneous parts of $f$ are also uniformly bounded by the same constant.
{As the previous example shows
this does not necessarily hold for $\OO$-free maps. Here is an explicit example of a uniformly bounded
analytic $\OO$-free map, which does not converge uniformly in a neighborhood of $0$. }
\begin{example}
We provide an example of a bounded analytic $\OO$-free map, such that the corresponding power series converges uniformly on $M_n(\RR)$ for all $n$ but does not converge uniformly on $\cM(\RR)$.
Define the homogeneous polynomials $z_{ij}=x_3^2x_2^{i-1}x_1^{j-1}-x_2^{i}x_1^{j}$ and let
$$h_k(x_1,x_2,x_3)=S_{2k}(z_{11},z_{22},z_{12},z_{33},\dots,z_{kk},z_{k-1,k},z_{k+1,k+1}),$$
where $S_{2k}$
denotes the standard polynomial of degree $2k$; i.e.,
\[
S_{2k}(x_1,\dots,x_{2n})=\sum_{\s\in {\rm Sym}(2n)}(-1)^\s x_{\s(1)}\cdots x_{\s(2n)}.
\]
We take
\beq\label{eq:greh}
f(x_1,x_2,x_3)=\sin\Big(\sum_{k=1}^\infty k!\big(h_k(x_1,x_2,x_3)+h_k(x_1,x_2,x_3)^t\big)\Big).
\eeq
Since $S_{2k}$ is a polynomial identity of $M_n(\RR)$ for $k\geq n$ by the Amitsur-Levitzki theorem (see e.g.~\cite[Theorem 1.4.1]{Row}), $f[n]$ can be defined by taking only a finite sum in the argument of $\sin$ in \eqref{eq:greh}.
Since $x\mapsto \sin(x)$ is analytic on $M_n(\RR)$, $f[n]$ is real analytic on $M_n(\RR)$. Moreover, $f$ is uniformly bounded by $1$, since the argument of $\sin$ in $f$ is symmetric.
Note that the corresponding power series $F=\sum f_m$, where $f_m$ is homogeneous of degree $m$, converges uniformly on $M_n(\RR)^3$ for every $n$, since the sum in the argument of $\sin$ in the definition of $f$ is finite on $M_n(\RR)^3$ and the power series corresponding to $\sin$ restricted to symmetric matrices converges uniformly.
We will now show that $F$ does not converge uniformly on $\cM(\RR)$.
{Assume for the sake of contradiction} that for every $\varepsilon>0$ there exist $N$ and $r>0$ such that
$$\Big\|f(X)-\sum_{m=0}^n f_m(X)\Big\|< \varepsilon \;\text{ for every $\|X\|<r$, $n\geq N$}.$$
Fix $\varepsilon<1$ and the corresponding $N$ and $r$.
Take $n>N$ such that
\beq\label{eq:weirdo}
n!\left(\frac{r}{2}\right)^{2n^2+3n+1}>\frac{\pi}{2}.
\eeq
Let
$$x_1=\sum_{i=1}^n e_{i,i+1},\quad x_2=\sum_{i=1}^{n}e_{i+1,i},\quad x_3=\sum_{i=1}^{n+1} e_{ii}+e_{n,n+1}$$
be elements in $M_{n+1}(\RR)$.
Note that $z_{ij}=e_{ij}$ for $1\leq i,j\leq n+1$, $i<j$. and $z_{ii}=e_{ii}+e_{n,n+1}$.
For $n>2$ we thus have
\[
h_k(x_1,x_2,x_3)=0 \quad \text{ for $k\neq n$, } \quad h_n(x_1,x_2,x_3)=(-1)^{n-1}(n+1)e_{1,n+1},
\]
where the last identity follows by the identities
\begin{multline*}
S_{2n}(e_{11},e_{22},e_{12},e_{33},\dots,e_{k-2,k-1},e_{n,n+1},e_{k-1,k},\dots,e_{n-1,n},e_{n+1,n+1})\\
=S_{2n}(e_{n,n+1},e_{22},e_{12},\dots,e_{n-1,n},e_{n-1,n},e_{n+1,n+1})=(-1)^{n-1}e_{1,n+1}
\end{multline*}
for $2\leq k\leq n+1$, and setting $e_{01}=e_{11}$.
{By \eqref{eq:weirdo} there is}
$r'<r$ such that
\[(n+1)!\left(\frac{r'}{2}\right)^{2n^2+3n+1}=\frac{\pi}{2}.
\]
Letting
\[y_i=\frac{r'}{2}x_i,\quad 1\leq i\leq 3,
\]
we have $||y||<r$ and
$$h_{n}(y_1,y_2,y_3)= (-1)^{n-1}\Big(\frac{r'}{2}\Big)^{2n^2+3n+1}(n+1)e_{1,n+1},$$
whence
\[f(y_1,y_2,y_3)=(-1)^{n-1}(e_{1,n+1}+e_{n+1,1}).\]
Note that $f_m(A_1,A_2,A_3)=0$ for $m<\ell$ if $h_{k}(A_1,A_2,A_3)=0$ for $k<\ell$.
Thus, \[\sum_{m=0}^N f_m(y_1,y_2,y_3)=0\] and
$$\Big\|f(y_1,y_2,y_3)-\sum_{m=0}^n f_m(y_1,y_2,y_3)\Big\|=1>\varepsilon,$$
a contradiction.
\end{example}
|
\section{Introduction}
The structure of incidences between points and various geometric objects is of central importance in discrete geometry, and theorems that elucidate this structure have had applications to, for example, problems from discrete and computational geometry \cite{guth2010erdos, pach2004geometric}, additive combinatorics \cite{chang2007sum, elekes1997number}, harmonic analysis \cite{guth2010algebraic, laba2008harmonic}, and computer science \cite{dvir2010incidence}.
The study of incidence theorems for finite geometry is an active area of research - e.g. \cite{vinh2011szemeredi, bourgain2004sum, covert2010generalized, iosevich2012areas, jones2012further, tao2012expanding, ellenberg2013incidence}.
The classical Szemr\'edi-Trotter Theorem~\cite{szemeredi1983extremal} bounds the maximum number of incidences between points and lines in 2 dimensional Euclidean space. Let $P$ be a set of points in $\mathbb{R}^2$ and $L$ be a set of lines in $\mathbb{R}^2$. Let $I(P,L)$ denote the number of incidences between points in $P$ and lines in $L$. The Szemer\'edi-Trotter Theorem shows that $I(P,L)$ is at most $O(|P|^{2/3}|L|^{2/3} + |P| + |L|)$. Ever since the original result, variations and generalizations of such incidence bounds have been intensively studied.
Incidence theorems for points and flats\footnote{We refer to $m$-dimensional affine subspaces of a vector space as $m$-flats.} in finite geometries is one instance of such incidence theorems that have received much attention, but in general we still do not fully understand the behavior of the bounds in this setting.
These bounds have different characteristics depending on the number of points and flats.
For example, consider the following question of proving an analog to the Szemer\'edi-Trotter theorem for points and lines in a plane over a finite field.
Let $q=p^n$ for prime $p$. Let $P$ be a set of points and $L$ a set of lines in $\mathbb{F}_q^2$, with $|P|=|L|=N$.
What is the maximum possible value of $I(P,L)$ over all point sets of size $|P|$ and sets of lines of size $|L|$?
If $N\leq O(\log\log\log(p))$, then a result of Grosu \cite{grosu2013f_p} implies that we can embed $P$ and $L$ in $\mathbb{C}^2$ without changing the underlying incidence structure.
Then we apply the result from the complex plane, proved by T\'oth \cite{toth2003szemeredi} and Zahl \cite{zahl2012szemeredi}, that $I(P,L) \leq O(N^{4/3})$.
This matches the bound of Szemer\'edi and Trotter, and a well-known construction based on a grid of points in $\mathbb{R}^2$ shows that the exponent of $4/3$ is tight.
The intermediate case of $N < p$ is rather poorly understood.
A result of Bourgain, Katz, and Tao \cite{bourgain2004sum}, later improved by Jones \cite{jones2012further}, shows that $I(P,L) \leq O(N^{3/2-\epsilon})$ for $\epsilon = 1/662 - o_p(1)$.
This result relies on methods from additive combinatorics, and is far from tight; in fact, we are not currently aware of any construction with $N<p^{3/2}$ that achieves $I(P,L) > \omega(N^{4/3})$.
For $N \geq q$, we know tight bounds on $I(P,L)$.
Using an argument based on spectral graph theory, Vinh \cite{vinh2011szemeredi} proved that $I(P,L) \leq N^2/q + q^{1/2}N$.
This bound meets the Szemer\'edi-Trotter bound of $O(N^{4/3})$ when $N=q^{3/2}$.
On the other hand, when $N = q$, this bound gives $O(N^{3/2})$, which may be proved from the Cauchy-Schwartz inequality using only the fact that each pair of points is contained in a single line.
A key difference between Vinh's bound and that of Bourgain, Katz, and Tao is that the argument that Vinh used is oblivious to the presence of subfields in $\mathbb{F}_q$.
Hence, it is hopeless to try to beat the Cauchy-Schwartz bound of $I(P,L) \leq O(N^{3/2})$ for small sets of points and lines using this method, since the Cauchy-Schwartz bound is tight when the points and lines are defined over a subfield of size $N^{1/2}$.
In this paper we generalize Vinh's argument to the purely combinatorial setting of balanced incomplete block designs (BIBDs), and show that this argument depends only on the combinatorial structure induced by flats in the finite vector space, and doesn't rely on the algebraic properties of $\mathbb{F}_q$ in any way.
We apply methods from spectral graph theory. We both generalize known incidence bounds for points and flats in finite geometries, and prove results for BIBDs that are new even in the special case of points and flats in finite geometries.
Finally, we apply one of these incidence bounds to improve a result of Iosevich, Rudnev, and Zhai \cite{iosevich2012areas} on the number of triangles with distinct areas determined by a set of points in $\mathbb{F}_q^2$.
\subsection{Outline}
In Section \ref{sec:definitions}, we state definitions of and basic facts about BIBDs and finite geometries.
In Section \ref{sec:incidenceTheorems}, we discuss our results on the incidence structure of designs.
In Section \ref{sec:triangleAreas}, we discuss our results on distinct triangle areas in $\mathbb{F}_q^2$.
In Section \ref{sec:graphTheory}, we introduce the tools from spectral graph theory that are used to prove our incidence results.
In Section \ref{sec:incidenceProofs}, we prove the results stated in Section \ref{sec:incidenceTheorems}.
In Section \ref{sec:triangleProofs}, we prove the result stated in Section \ref{sec:triangleAreas}.
\section{Results}\label{sec:results}
\subsection{ Definitions and Background}\label{sec:definitions}
Let $X$ be a finite set (which we call the points), and let $B$ be a set of subsets of $X$ (which we call the blocks).
We say that $(X,B)$ is an $(r, k, \lambda)$-BIBD if
\begin{itemize}
\item each point is in $r$ blocks,
\item each block contains $k$ points,
\item each pair of points is contained in $\lambda$ blocks, and
\item no single block contains all of the points.
\end{itemize}
It is easy to see that the following relations among the parameters $|X|,|B|,r,k,\lambda$ of a BIBD hold:
\begin{eqnarray}
r|X| & = & k|B|, \\
\lambda(|X| - 1) &=& r(k-1).
\end{eqnarray}
The first of these follows from double counting the pairs $(x,b) \in X \times B$ such that $x \in b$.
The second follows from fixing an element $x\in X$, and double counting the pairs $(x',b) \in (X \setminus \{x\}) \times B$ such that $x',x \in b$.
In the case where $X$ is the set of all points in $\mathbb{F}_q^n$, and $B$ is the set of $m$-flats in $\mathbb{F}_q^n$, we obtain a design with the following parameters \cite{ball2005introduction}:
\begin{itemize}
\item $|X| = q^n$,
\item $|B| = \binom{n+1}{m+1}_q - \binom{n}{m+1}_q$,
\item $r = \binom{n}{m}_q$,
\item $k = q^m$, and
\item $\lambda = \binom{n-1}{m-1}_q$.
\end{itemize}
The notation $\binom{n}{m}_q$ refers to the $q$-binomial coefficient, defined for integers $m \leq n$ by
\[ \binom{n}{m}_q = \frac{(q^n-1)(q^n-q) \ldots (q^n-q^{m-1})}{(q^m-1)(q^m-q) \ldots (q^m-q^{m-1})}.\]
We will only use the fact that $\binom{n}{m}_q = (1+o_q(1))q^{m(n-m)}$.
Given a design $(X,B)$, we say that a point $x \in X$ is incident to a block $b \in B$ if $x \in b$.
For subsets $P \subseteq X$ and $L\subseteq B$, we define $I(P,L)$ to be the number of incidences beween $P$ and $L$; in other words,
\[I(P,L) = |\{(x,b) \in P \times L: x \in b\}|.\]
Given a subset $L \subseteq B$, we say that a point $x$ is $t$-rich if it is contained in at least $t$ blocks of $B$, and we define $\Gamma_t(L)$ to be the number of $t$-rich points in $X$; in other words,
\[\Gamma_t(L) = |\{x \in X:|\{b \in L: x \in b\}| \geq t\}|.\]
Given a subset $P \subseteq X$, we say that a block $b$ is $t$-rich if it contains at least $t$ points of $P$, and we define $\Gamma_t(P)$ to be the number of $t$-rich blocks in $B$; in other words,
\[\Gamma_t(P) = |\{b \in B: |\{x \in P : x \in b \}| \geq t\}|.\]
\subsection{Incidence Theorems}\label{sec:incidenceTheorems}
The first result on the incidence structure of designs is a generalization of the finite field analog to the Szemer\'edi-Trotter theorem proved by Vinh \cite{vinh2011szemeredi}.
\begin{theorem}\label{th:designIncidenceBound}
Let $(X,B)$ be an $(r,k,\lambda)$-BIBD.
The number of incidences between $P \subseteq X$ and $L \subseteq B$ satisfies
\[\big | \thinspace I(P,L) - |P||L|r/|B| \thinspace \big | \leq \sqrt{(r-\lambda)|P||L|}.\]
\end{theorem}
Theorem \ref{th:designIncidenceBound} gives both upper and lower bounds on the number of incidences between arbitrary sets of points and blocks.
The term $|P||L|r/|B|$ corresponds to the number of incidences that we would expect to see between $P$ and $L$ if they were chosen uniformly at random.
When $|P||L|$ is much larger than $|B|^2(r-\lambda)/r^2 > |B|^2/r$, then $|P||L|r/|B|$ is much larger than $ \sqrt{(r-\lambda)|P||L|}$. Thus the theorem says that every set of points and blocks determines approximately the ``expected" number of incidences.
When $|P||L|<|B|^2/r$, the term on the right is larger, and Theorem \ref{th:designIncidenceBound} gives only an upper bound on the number of incidences.
The Cauchy-Schwartz inequality combined with the fact that each pair of points is in at most $\lambda$ blocks easily implies that $I(P,L) \leq \lambda^{1/2} |P| |L|^{1/2} + |L|$.
Hence, the upper bound in Theorem \ref{th:designIncidenceBound} is only interesting when $|P|>(r-\lambda)/\lambda$.
In the case of incidences between points and $m$-flats in $\mathbb{F}_q^n$, we get the following result as a special case of Theorem \ref{th:designIncidenceBound}.
\begin{corollary}\label{th:ffIncidenceBound}
Let $P$ be a set of points $L$ and let be a set of $m$-flats in $\mathbb{F}_q^n$. Then
\[\left | I(P,L) - |P| \thinspace |L|q^{m-n} \right | \leq (1 +o_q(1))\sqrt{q^{m(n-m)}|P|\thinspace |L|}.\]
\end{corollary}
Vinh \cite{vinh2011szemeredi} proved Corollary \ref{th:ffIncidenceBound} in the case $m = n-1$, and Bennett, Iosevich, and Pakianathan \cite{bennett2012three} derived the bounds for the remaining values of $m$ from Vinh's bound using elementary combinatorial arguments.
Vinh's proof is based on spectral graph theory, analogous to the proof of Theorem \ref{th:designIncidenceBound} that we present in Section \ref{sec:incidenceProofs}.
Cilleruelo proved a result similar to Vinh's using Sidon sets \cite{cilleruelo2012combinatorial}.
We also show lower bounds on the number of $t$-rich blocks determined by a set of points, and on the number of $t$-rich points determined by a set of points.
While reading the statements of these theorems, it is helpful to recall from equation $(1)$ that $|X|/k = |B|/r$.
\begin{theorem}\label{th:designRichBlocks}
Let $(X,B)$ be an $(r,k,\lambda)$-BIBD.
Let $\epsilon \in \mathbb{R}_{>0}$ and $t \in \mathbb{Z}_{\geq 2}$.
Let $P \subseteq X$ with
\[|P| \geq (1+\epsilon)(t-1)|X|/k.\]
Then, the number of $t$-rich blocks is at least
\[\Gamma_t(P) \geq a_{\epsilon,t, \mathcal{D}}|B|,\]
where
\[a_{\epsilon,t,\mathcal{D}} = \frac{\epsilon^2(t-1)}{\epsilon^2(t-1) + (1- \frac{\lambda}{r})(1+\epsilon)}.\]
\end{theorem}
\begin{theorem}\label{th:designRichPoints}
Let $(X,B)$ be an $(r,k,\lambda)$-BIBD.
Let $\epsilon \in \mathbb{R}_{>0}$ and $t \in \mathbb{Z}_{\geq 2}$.
Let $\L \subseteq B$, with
\[|L| \geq (1+\epsilon)(t-1)|X|/k.\]
Then, the number of points $t$-rich points is at least
\[\Gamma_t(L) \geq b_{\epsilon,t,\mathcal{D}}|X|,\]
where
\[b_{\epsilon,t,\mathcal{D}} = \frac{\epsilon^2(t-1)}{\epsilon^2(t-1) + \frac{r-\lambda}{k}(1+\epsilon)}.\]
\end{theorem}
Theorem \ref{th:designRichBlocks} is analogous to Beck's theorem \cite{beck1983lattice}, which states that, if $P$ is a set of points in $\mathbb{R}^2$, then either $c_1 |P|$ points lie on a single line, or $c_2 |P|^2$ lines each contain at least $2$ points of $P$, where $c_1$ and $c_2$ are fixed positive constants.
Both the $t=2$ case of Theorem \ref{th:designRichBlocks} and Beck's theorem are closely related to the de Bruijn-Erd\H{o}s theorem \cite{debruijn1948combinatorial, ryser1968extension}, which states that, if $P$ is a set, and $L$ is a set of subsets of $P$ such that each pair of elements in $P$ is contained in exactly $\lambda$ members of $L$, then either a single member of $L$ contains all elements of $P$, or $|L| \geq |P|$.
Each of Theorem \ref{th:designRichBlocks} and Beck's theorem has an additional hypothesis on the de Bruijn-Erd\H{o}s theorem, and a stronger conclusion.
Beck's theorem improves the de Bruijn-Erd\H{o}s theorem when $P$ is a set of points in $\mathbb{R}^2$ and $L$ is the set of lines that each contain $2$ points of $P$.
Theorem \ref{th:designRichBlocks} improves the de Bruijn-Erd\H{o}s theorem when $P$ is a sufficiently large subset of the points of a BIBD, and $L$ is the set of blocks that each contain at least $2$ points of $P$.
As special cases of Theorems \ref{th:designRichBlocks} and \ref{th:designRichPoints}, we get the following results on the number of $t$-rich points determined by a set of $m$-flats in $\mathbb{F}_q^n$, and on the number of $t$-rich $m$-flats determined by a set of points in $\mathbb{F}_q^2$.
\begin{corollary}\label{th:ffRichFlats}
Let $\epsilon \in \mathbb{R}_{\geq 0}$ and $t \in \mathbb{Z}_{\geq 2}$.
Let $P \subseteq \mathbb{F}_q^n$ with
\[|P| \geq (1+\epsilon)(t-1)q^{n-m}.\]
Then the number of $t$-rich $m$-flats is at least
\[\Gamma_t(L) \geq a_{\epsilon,t}q^{(m+1)(n-m)},\]
where
\[a_{\epsilon,t} = \frac{\epsilon^2(t-1)}{\epsilon^2(t-1) + (1 +\epsilon)}.\]
\end{corollary}
\begin{corollary}\label{th:ffRichPoints}
Let $\epsilon \in \mathbb{R}_{\geq 0}$ and $t \in \mathbb{Z}_{\geq 2}$.
Let $L$ be a subset of the $m$-flats in $\mathbb{F}_q^n$ with
\[|L| \geq (1+\epsilon)(t-1)q^{n-m}.\]
Then, the number of $t$-rich points is at least
\[\Gamma_t(L) \geq b_{\epsilon,t,q}q^n,\]
where
\[b_{\epsilon,t,q} = \frac{\epsilon^2(t-1)}{\epsilon^2(t-1) + q^{m(n-m-1)}(1+\epsilon)}.\]
\end{corollary}
The case $n=2,m=1, t=2$ of Corollary \ref{th:ffRichFlats} was proved (for a slightly smaller value of $a_{\epsilon,t}$) by Alon \cite{alon1986eigenvalues}.
Alon's proof is also based on spectral graph theory.
When $m=n-1$, Corollaries \ref{th:ffRichFlats} and \ref{th:ffRichPoints} are dual (in the projective sense) to each other; slight differences in the parameters arise since we're working in affine (as opposed to projective) geometry.
For the case $m < n-1$,the value of $b_{\epsilon,t,q}$ in Corollary \ref{th:ffRichPoints} depends strongly on $q$ when $\epsilon (t-1) < q^{m(n-m-1)}$.
This dependence is necessary.
For example, consider the case $n=3, m=1$, i.e. lines in $\mathbb{F}_q^3$.
Corollary \ref{th:ffRichPoints} implies that a set of $2q^2$ lines in $\mathbb{F}_q^3$ determines $\Omega(q^2)$ $2$-rich points, which is asymptotically fewer (with regard to $q$) than the total number of points in $\mathbb{F}_q^3$.
This is tight, since the lines may lie in the union of two planes.
By contrast, Corollary \ref{th:ffRichFlats} implies that a set of $2q^2$ points in $\mathbb{F}_q^3$ determines $\Omega(q^4)$ $2$-rich lines, which is a constant proportion of all lines in the space.
\subsection{Distinct Triangle Areas}\label{sec:triangleAreas}
Iosevich, Rudnev, and Zhai \cite{iosevich2012areas} studied a problem on distinct triangle areas in $\mathbb{F}_q^2$.
This is a finite field analog to a question that is well-studied in discrete geometry over the reals.
Erd\H{o}s, Purdy, and Strauss \cite{erdos1982problem} conjectured that a set of $n$ points in the real plane determines at least $\lfloor \frac{n-1}{2} \rfloor$ distinct triangle areas.
Pinchasi \cite{pinchasi2008minimum} proved that this is the case.
In $\mathbb{F}_q^2$, we define the area of a triangle in terms of the determinant of a matrix.
Suppose a triangle has vertices $a, b$, and $c$, and let $z_x$ and $z_y$ denote the $x$ and $y$ coordinates of a point $z$.
Then, we define the area associated to the ordered triple $(a,b,c)$ to be the determinant of the following matrix:
\[
\left[ {\begin{array}{ccc}
1 & 1 & 1 \\
a_x & b_x & c_x \\
a_y & b_y & c_y
\end{array} } \right]
\]
Iosevich, Rudnev, and Zhai \cite{iosevich2012areas} showed that a set of at least $64 q \log_2q $ points includes a point that is a common vertex of triangles having at least $q/2$ distinct areas.
They first prove a finite field analog of Beck's theorem, and then obtain their result on distinct triangle areas using this analog to Beck's theorem along with some Fourier analytic and combinatorial techniques.
Corollary \ref{th:ffRichFlats} (in the case $n=2,m=1,t=2$) strengthens their analog to Beck's theorem, and thus we are able to obtain the following strengthening of their result on distinct triangle areas.
\begin{theorem}\label{th:triangles_on_z_alternate}
Let $\epsilon \in \mathbb{R}_{>0}$.
Let $P$ be a set of at least $(1 + \epsilon)q$ points in $\mathbb{F}_q^2$.
Let $T$ be the set of triangles determined by $P$.
Then there is a point $z \in P$ such that $z$ is a common vertex of triangles in $T$ with at least $c_\epsilon q$ distinct areas,
where $c_\epsilon$ is a positive constant depending only on $\epsilon$, such that $c_\epsilon \to 1$ as $\epsilon \to \infty$.
\end{theorem}
Notice that Theorem \ref{th:triangles_on_z_alternate} is tight in the sense that fewer than $q$ points might determine only triangles with area zero (if all points are collinear). It is a very interesting open question to determine the minimum number of points $K_q$, such that any set of points of size $K_q$ determines all triangle areas.
In fact, we are not currently aware of any set of more than $q+1$ points that does not determine all triangle areas.
\section{Tools from Spectral Graph Theory}\label{sec:graphTheory}
\subsection{Context and Notation}
Let $G=(L,R,E)$ be a $(\Delta_L, \Delta_R)$-biregular bipartite graph; in other words, $G$ is a bipartite graph with left vertices $L$, right vertices $R$, and edge set $E$, such that each left vertex has degree $\Delta_L$, and each right vertex has degree $\Delta_R$.
Let $A$ be the $|L \cup R| \times |L \cup R|$ adjacency matrix of $G$, and let $\mu_1 \geq \mu_2 \geq \ldots \geq \mu_{|L|+|R|}$ be the eigenvalues of $A$.
Let $\mu = \mu_2/\mu_1$ be the normalized second eigenvalue of $G$.
Let $e(G)=\Delta_L|L| = \Delta_R|R|$ be the number of edges in $G$.
For any two subsets of vertices $A$ and $B$, denote by $e(A,B)$ the number of edges between $A$ and $B$.
For a subset of vertices $A \subseteq L \cup R$, denote by $\Gamma_t(A)$ the set of vertices in $G$ that have at least $t$ neighbors in $A$.
\subsection{Lemmas}
We will use two lemmas relating the normalized second eigenvalue of $G$ to its combinatorial properties.
The first of these is the expander mixing lemma \cite{alon1988explicit}.
\begin{lemma}[Expander Mixing Lemma]\label{th:expander_mixing}
Let $S \subseteq L$ with $|S|=\alpha |L|$ and let $T \subseteq R$ with $|T| = \beta |R|$.
Then,
\[\left | \frac{e(S,T)}{e(G)} - \alpha \beta \right | \leq \mu \sqrt{\alpha \beta (1- \alpha)(1- \beta)}.\]
\end{lemma}
Several variants of this result appear in the literature, most frequently without the $\sqrt{(1-\alpha)(1-\beta)}$ terms.
For a proof that includes these terms, see \cite{vadhan2012pseudorandomness}, Lemma 4.15.
Although the statement in \cite{vadhan2012pseudorandomness} is not specialized for bipartite graphs, it is easy to modify it to obtain Lemma \ref{th:expander_mixing}.
For completeness, we include a proof in the appendix.
Theorems \ref{th:designRichPoints} and \ref{th:designRichBlocks} follow from the following corollary to the expander mixing lemma, which may be of independent interest.
\begin{lemma}\label{th:becksForGraphs}
Let $\epsilon \in \mathbb{R}_{>0}$, and let $t \in \mathbb{Z}_{\geq 1}$.
If $S \subseteq L$ such that
\[|S| \geq (1+\epsilon)(t-1)|L|/\Delta_R,\]
then
\[|\Gamma_t(S)| \geq c_{\epsilon, t, G} |R|,\]
where
\[c_{\epsilon, t, G} = \frac{\epsilon^2(t-1)}{\epsilon^2(t-1) + \mu^2 \Delta_R (1+\epsilon)}.\]
\end{lemma}
\begin{proof}
Let $T = R \setminus \Gamma_{k}(S)$.
Let $\alpha = |S|/|L|$, and let $\beta = |T|/|R|$.
We will calculate a lower bound on $1-\beta = |\Gamma_k(S)|/|R|$, from which we will immediately obtain a lower bound on $|\Gamma_k(S)|$.
Since each vertex in $T$ has at most $k-1$ edges to vertices in $S$, we have $e(S,T) \leq (k-1)|T|$.
Along with the fact that $|T| = \beta |R|$, this gives
\[\begin{array}{rcl}
\alpha \beta - \frac{e(S,T)}{|R| \Delta_R} &\geq& \alpha \beta - (k-1) \beta / \Delta_R \\
&=& \beta(\alpha - (k-1)/\Delta_R). \end{array}\]
Lemma \ref{th:expander_mixing} implies that $\alpha \beta - e(S,T)/|R| \Delta_R \leq \lambda \sqrt{\alpha \beta (1-\alpha)(1-\beta)}$.
Since we expect $\alpha$ to be small, we will drop the $(1-\alpha)$ term, and we have
\[\begin{array}{rcl}
\lambda \sqrt{\alpha \beta (1-\beta)} & \geq & \beta(\alpha - (k-1)/\Delta_R) \\
\lambda^2 \alpha \beta (1-\beta) & \geq & \beta^2 (\alpha - (k-1)/\Delta_R)^2 \\
\lambda^2 (1-\beta)/\beta & \geq & \alpha - 2(k-1)/\Delta_R + ((k-1)/\Delta_R)^2 / \alpha.
\end{array}\]
By hypothesis, $\alpha \geq (1+\epsilon)(k-1)/\Delta_R$.
Let $c \geq 1$ such that $\alpha = c(1+\epsilon)(k-1)/\Delta_R$.
Then,
\[\begin{array}{rcl}
\lambda^2(1-\beta)/\beta & \geq & c(1+\epsilon)(k-1)/\Delta_R + (k-1)/\Delta_R(1+\epsilon)c - 2(k-1)/\Delta_R \\
& = & \left (c(1+\epsilon) + 1/c(1+\epsilon) - 2 \right ) \frac{k-1}{\Delta_R}.
\end{array}\]
Define $f(x) =x(1+\epsilon) + 1/x(1+\epsilon) - 2$ for $x\geq 1$.
Then
\[\frac{d (f(x))}{d x} = 1+\epsilon - 1/(1+\epsilon)x^2.\]
Since $x\geq 1$ and $1+\epsilon > 1$, we have $\frac{d (f(x))}{d x} > 0$.
Therefore, $f(c) \geq f(1)$, and
\[\begin{array}{rcl}
\lambda^2(1-\beta)/\beta & \geq & (1+ \epsilon + 1/(1+\epsilon) - 2) \frac{k-1}{\Delta_R} \\
& = &\frac{((\epsilon - 1)(1+\epsilon)+1)(k-1)}{(1+\epsilon)\Delta_R} \\
& = & \frac{\epsilon^2(k-1)}{(1+\epsilon)\Delta_R} \\
1/\beta - 1 & \geq & \frac{\epsilon^2(k-1)}{(1+\epsilon)\lambda^2\Delta_R} \\
\beta & \leq & \frac{(1+\epsilon)\lambda^2\Delta_R}{\epsilon^2(k-1) + (1+\epsilon)\lambda^2\Delta_R} \\
1 - \beta & \geq & \frac{\epsilon^2(k-1)}{\epsilon^2(k-1) + (1+\epsilon)\lambda^2\Delta_R}.
\end{array}\]
Recall that $1-\beta = |\Gamma_{k}(S)|/N_R$, so this completes the proof.
\end{proof}
\section{Proof of Incidence Bounds}\label{sec:incidenceProofs}
In this section, we prove Theorems \ref{th:designIncidenceBound}, \ref{th:designRichBlocks}, and \ref{th:designRichPoints}.
We first prove results on the spectrum of the bipartite graph associated to a BIBD.
We then use Lemmas \ref{th:expander_mixing} and \ref{th:becksForGraphs} to complete the proofs.
\begin{lemma}\label{th:eigenvaluesOfA}
Let $(X,B)$ be an $(r, k, \lambda)$-BIBD.
Let $G=(L,R,E)$ be a bipartite graph with left vertices $X$, right vertices $B$, and $(x,b) \in E$ if $x \in b$.
Let $A$ be the $(|X| + |B|) \times (|X| + |B|)$ adjacency matrix of $G$.
Then, the normalized second eigenvalue of $A$ is $\sqrt{(r - \lambda)/rk}$.
\end{lemma}
\begin{proof}
Let $N$ be the $|X| \times |B|$ incidence matrix of $\mathcal{D}$; that is, $N$ is a $(0,1)$-valued matrix such that $N_{i,j} = 1$ iff point $i$ is in block $j$.
We can write
\[A=
\left[ {\begin{array}{cc}
0 & N\\
N^T & 0
\end{array}} \right]. \]
Instead of analyzing the eigenvalues of $A$ directly, we'll first consider the eigenvalues of $A^2$.
Since
\[A^2=
\left[ {\begin{array}{cc}
NN^T & 0\\
0 & N^TN
\end{array}} \right] \]
is a block diagonal matrix, the eigenvalues of $A^2$ (counted with multiplicity) are the union of the eigenvalues of $NN^T$ and the eigenvalues of $N^TN$.
We will start by calculating the eigenvalues of $NN^T$.
The following observation about $NN^T$ was noted by Bose \cite{bose1949note}.
\begin{proposition}
\[NN^T = (r-\lambda)I + \lambda J,\]
where $I$ is the $|X| \times |X|$ identity matrix and $J$ is the $|X| \times |X|$ all-$1$s matrix.
\end{proposition}
\begin{proof}
The entry $(NN^T)_{i,j}$ corresponds to the number of blocks that contain both point $i$ and point $j$.
From the definition of an $(r,k,\lambda)$-BIBD, it follows that $(NN^T)_{i,i} = r$ and $(NN^T)_{i,j}=\lambda$ if $i \neq j$, and the conclusion of the proposition follows.
\end{proof}
We use the above decomposition to calculate the eigenvalues of $NN^T$.
\begin{proposition}
The eigenvalues of $NN^T$ are $rk$ with multiplicity $1$ and $r-\lambda$ with multiplicity $|X|-1$.
\end{proposition}
\begin{proof}
The eigenvalues of $I$ are all $1$.
The eigenvalues of $J$ are $|X|$ with multiplicity $1$ and $0$ with multiplicity $|X|-1$.
The eigenvector of $J$ corresponding to eigenvalue $|X|$ is the all-ones vector, and the orthogonal eigenspace has eigenvalue $0$.
Since $I$ and $J$ share a basis of eigenvectors, the eigenvalues of $NN^T$ are simply the sums of the corresponding eigenvalues of $(r-\lambda)I$ and $\lambda J$.
Hence, the largest eigenvalue of $NN^T$ is $r-\lambda + |X|\lambda$, corresponding to the all-ones vector, and the remaining eigenvalues are $r-\lambda$, corresponding to vectors whose entries sum to $0$.
From equation $(2)$, we have $\lambda(|X|-1) = r(k-1)$, and so we can write the largest eigenvalue as $rk$.
\end{proof}
Next, we use the existence of a singular value decomposition of $N$ to show that the nonzero eigenvalues of $N^TN$ have the same values and occur with the same multiplicity as the eigenvalues of $NN^T$.
The following is a standard theorem from linear algebra; see e.g. \cite[p. 429]{roman1992advanced}.
\begin{theorem}[Singular value decomposition.]
Let $M$ be a $m \times n$ real-valued matrix with rank $r$.
Then,
\[M=P \Sigma Q^T,\]
where $P$ is an $m \times m$ orthogonal matrix, $Q$ is an $n \times n$ orthogonal matrix, and $\Sigma$ is a diagonal matrix.
In addition, if the diagonal entries of $\Sigma$ are $s_1, s_2, \ldots, s_r, 0, \ldots, 0$, then the nonzero eigenvalues of $NN^T$ and $N^TN$ are $s_1^2, s_2^2, \ldots, s_r^2$.
\end{theorem}
It is immediate from this theorem that the nonzero eigenvalues of $N^TN$, counted with multiplicity, are identical with those of $NN^T$.
Hence, the nonzero eigenvalues of $A^2$ are $rk$ with multiplicity $2$ and $r-\lambda$ with multiplicity $2(|X|-1)$.
Clearly, the eigenvalues of $A^2$ are the squares of the eigenvalues of $A$; indeed, if $x$ is an eigenvector of $A$ with eigenvalue $\mu$, then $A^2x = \mu Ax = \mu^2x$.
Hence, the conclusion of the lemma will follow from the following proposition that the eigenvalues of $A$ are symmetric about $0$.
Although it is a well-known fact in spectral graph theory that the eigenvalues of the adjacency matrix of a bipartite graph are symmetric about $0$, we include a simple proof here for completeness.
\begin{proposition}
If $\mu$ is an eigenvalue of $A$ with multiplicity $w$, then $-\mu$ is an eigenvalue of $A$ with multiplicity $w$.
\end{proposition}
\begin{proof}
Let $x_1 \in \mathbb{R}^{|X|}$ and $x_2 \in \mathbb{R}^{|B|}$ so that $(x_1,x_2)^T$ is an eigenvector of $A$ with corresponding nonzero eigenvalue $\mu$.
Then,
\[A \left[ {\begin{array}{c} x_1 \\ x_2 \end{array}} \right] = \left[ {\begin{array}{c} Nx_2 \\ N^Tx_1 \end{array}} \right] = \mu \left[ {\begin{array}{c} x_1 \\ x_2 \end{array}} \right]. \]
Note that, since $Nx_2 = \mu x_1$ and $N^T x_1 = \mu x_2$ and $\mu \neq 0$, we have that that $x_1 \neq 0$ and $x_2 \neq 0$.
In addition,
\[A \left[ {\begin{array}{c} -x_1 \\ x_2 \end{array}} \right] = \left[ {\begin{array}{c} Nx_2 \\ -N^Tx_1 \end{array}} \right] = -\mu \left[ {\begin{array}{c} -x_1 \\ x_2 \end{array}} \right] \]
Hence, if $\mu$ is an eigenvalue of $A$ with eigenvector $(x_1,x_2)^T$, then $-\mu$ is an eigenvalue of $A$ with eigenvector $(-x_1, x_2)^T$.
Since $A$ is a real symmetric matrix, it follows from the spectral theorem \cite[p. 227]{roman1992advanced} that $A$ has an orthogonal eigenvector basis; hence, we can match the eigenvectors of $A$ with eigenvalue $\mu$ with those having eigenvalue $-\mu$ to show that the multiplicity of $\mu$ is equal to the multiplicity of $-\mu$.
\end{proof}
Now we can calculate that the nonzero eigenvalues of $A$ are $\sqrt{rk}$ and $-\sqrt{rk}$, each with multiplicity $1$, and $\sqrt{r-\lambda}$ and $-\sqrt{r-\lambda}$, each with multiplicity $|X|-1$.
Hence, the normalized second eigenvalue of $A$ is $\sqrt{(r-\lambda)/rk}$, and the proof of Lemma \ref{th:eigenvaluesOfA} is complete.
\end{proof}
\begin{proof}[Proof of Theorem \ref{th:designIncidenceBound}]
Lemma \ref{th:expander_mixing} implies that given a sets $P \subseteq X$ and $L \subseteq B$, the number of edges in $G$ between $P$ and $L$ is bounded by
\[ \left | \frac{e(P,L)}{r|X|} - \frac{|P| \thinspace |L|}{|X| \thinspace |B|} \right| \leq \sqrt{(r-\lambda)|P| \thinspace |L|/rk|X| \thinspace |B|}. \]
From equation (1), we know that $r|X| = k |B|$, so multiplying through by $r|X|$ gives
\[ \big | e(P,L) - |P| \thinspace |L| r/|B| \big | \leq \sqrt{(r-\lambda)|P| \thinspace |L|}.\]
From the construction of $G$, we see that $e(P,L)$ is exactly the term $I(P,L)$ bounded in Theorem \ref{th:designIncidenceBound}, so this completes the proof of Theorem \ref{th:designIncidenceBound}.
\end{proof}
\begin{proof}[Proof of Theorem \ref{th:designRichBlocks}]
Lemma \ref{th:becksForGraphs} implies that, for any $\epsilon \in \mathbb{R}_{>0}$ and $t \in \mathbb{Z}_{\geq 1}$, given a set $P \subseteq X$ such that
\[|P| \geq (1+\epsilon)(t-1)|X|/k,\]
there are at least
\[\Gamma_k(P) \geq \frac{\epsilon^2(t-1)|B|}{\epsilon^2 (t-1) + (r-\lambda)k (1+\epsilon)/rk}\]
vertices in $B$ that each have at least $t$ edges to vertices in $P$.
Rearranging slightly and again using the fact that edges in $G$ correspond to incidences in $\mathcal{D}$ gives Theorem \ref{th:designRichBlocks}.
\end{proof}
\begin{proof}[Proof of Theorem \ref{th:designRichPoints}]
Lemma \ref{th:becksForGraphs} also implies that, for any $\epsilon \in \mathbb{R}_{>0}$ and $t \in \mathbb{Z}_{\geq 1}$, given a set $L \subseteq B$ such that
\[|L| \geq (1+\epsilon)(t-1)|B|/r = (1+\epsilon)(t-1)|X|/k,\]
there are at least
\[ \Gamma_k(L) \geq \frac{\epsilon^2(t-1)|X|}{\epsilon^2 (t-1) + (r-\lambda)r (1+\epsilon)/rk}.\]
Simplifying this expression gives Theorem \ref{th:designRichPoints}.
\end{proof}
\section{Application to Distinct Triangle Areas}\label{sec:triangleProofs}
In this section, we will prove Theorem \ref{th:triangles_on_z_alternate}.
We will need the following theorem, which was proved by Iosevich, Rudnev, and Zhai \cite{iosevich2012areas} as a key ingredient of their lower bound on distinct triangle areas.
\begin{theorem}[\cite{iosevich2012areas}]\label{th:dotProductIncidenceBound}
Let $F,G \subset \mathbb{F}_q^2$. Suppose $0 \notin F$. Let, for $d \in \mathbb{F}_q$,
\[\nu(d) = |\{(a,b) \in F \times G : a \cdot b = d\}|,\]
where $a \cdot b = a_xb_x + a_yb_y$.
Then
\[\sum_d \nu^2(d) \leq |F|^2|G|^2q^{-1} + q |F| |G| \max_{x \in \mathbb{F}_q^2 \setminus \{0\}}|F \cap l_x|,\]
where
\[l_x = \{sx:s \in \mathbb{F}_q \}.\]
\end{theorem}
We will also need the following consequence of Corollary \ref{th:ffRichFlats}.
\begin{lemma}\label{th:centerPointBeck}
Let $\epsilon \in \mathbb{R}_{>0}$, and let $t \in \mathbb{Z}_{\geq 2}$.
There exists a constant $c'_\epsilon > 0$, depending only on $\epsilon$, such that the following holds.
Let $P$ be a set of $(1 + \epsilon)(t-1) q$ points in $\mathbb{F}_q^2$.
Then there is a point $z \in P$ such that at least $c'_\epsilon q$ or more $t$-rich lines are incident to $z$.
Moreover, if $\epsilon \geq 1$, then $c'_\epsilon \geq 1/3$.
\end{lemma}
\begin{proof}
By Corollary \ref{th:ffRichFlats},
\[ |\Gamma_t(P)| \geq a_{\epsilon,t} q^2,\]
where
\[a_{\epsilon,t} = \frac{\epsilon^2 (t-1)}{\epsilon^2 (t-1) + 1 + \epsilon}.\]
Denote by $I(P,\Gamma_t(P))$ the number of incidences between points of $P$ and lines of $\Gamma_t(P)$.
Since each line of $\Gamma_t(P)$ is incident to at least $t$ points of $P$, the average number of incidences with lines of $\Gamma_t(P)$ that each point of $P$ participates in is at least
\[\begin{array}{rcl}
I(P,\Gamma_t(P))/|P| & \geq & t|\Gamma_t(P)|/|P| \\
& \geq & t a_{\epsilon,t}q^2/|P| \\
&=& (1+\epsilon)(t-1)ta_{\epsilon,t} q \\
&=&c'_{\epsilon,t}q, \\
\end{array}\]
where $c'_{\epsilon,t} = t(\epsilon^2(t-1))/(1+\epsilon)(\epsilon^2(t-1)^2 + 1 + \epsilon)$.
We will now eliminate the dependence of $c'_{\epsilon,t}$ on $t$.
The maximum number of $t$-rich lines on any point increases monotonically in $\epsilon$.
Hence, it suffices to consider the cases that $\epsilon \in (0,1)$ and that $\epsilon = 1$.
In the first case, since $t$ is an integer that can take a finite number of values (since $t \geq 2$ and $(1+\epsilon)(t-1)q = |P| \leq q^2)$, there is some constant $\delta$ such that $c'_{\epsilon,t}\geq \delta$ for all $t$, and we set $c'_{\epsilon} = \min_{t}(c'_{\epsilon,t}) \geq \delta$.
In the second case, $c'_{1,t} = ((t-1)^2 + t)/(2(t-1)^2+4)$.
The minimum of $c'_{1,t}$ for $t \geq 2$ is $1/3$ for $t=2$, and so we set $c'_\epsilon = 1/3$.
Since the average point in $P$ is incident to at least $I(P,\Gamma_t(P))/|P|$ $t$-rich lines, there must be a point in $P$ incident to at least so many $t$-rich lines.
\end{proof}
\begin{proof}[Proof of Theorem \ref{th:triangles_on_z_alternate}]
Let $\delta = (1 + \epsilon)/(t-1) - 1$, so that $|P| = (1+\delta)(t-1)q$.
Let $c'_\delta$ be as in Lemma \ref{th:centerPointBeck}.
By Lemma \ref{th:centerPointBeck}, there is a point $z$ in $P$ incident to $c'_\delta q$ or more $t$-rich lines.
Let $P' \subseteq P - \{z\}$ be a set of points such that there are exactly $t-1$ points of $P'$ on exactly $\lceil c'_\delta q \rceil$ lines incident to $z$.
Clearly,
\[|P'| = (t-1) \lceil c'_\delta q \rceil \geq (t-1) c'_\delta q.\]
Let $P'_z$ be $P'$ translated so that $z$ is at the origin.
Each ordered pair $(a,b) \in P'_z \times P'_z$ corresponds to a triangle having $z$ as a vertex.
By the definition of area, given in Section \ref{sec:triangleAreas}, the area of the triangle corresponding to $(a,b)$ is $a_xb_y - b_xa_y$.
For any point $x \in \mathbb{F}_q^2$, let $x^\bot = (-x_y,x_x)$; let ${P'_z}^\bot = \{x^\bot:x \in P'_z\}$.
The area corresponding to $(a,b)$ is $a_xb_y - b_xa_y = a^\bot \cdot b$.
Hence, the number of distinct areas spanned by triangles with $z$ as a vertex is at least the number of distinct dot products $|\{a^\bot \cdot b: a^\bot \in {P'_z}^\bot, b \in P'_z\}|$.
To write this in another way, let $\nu(d)$ be as defined in Theorem \ref{th:dotProductIncidenceBound} with $F=P'_z$ and $G={P'_z}^\bot$.
Then, the number of distinct areas spanned by triangles containing $z$ is at least $|\{d:\nu(d) \neq 0\}|$.
Since no line through the origin contains more than $t-1$ points of $P'_z$, Theorem \ref{th:dotProductIncidenceBound} implies that
\[\sum_t \nu^2(d) \leq |P'|^4q^{-1} + q(t-1)|P'|^2.\]
By Cauchy-Schwarz, the number of distinct triangle areas is at least
\[ \begin{array}{rcl}
|\{d:\nu(d) \neq 0\}| & \geq& |\sum_d \nu(d)|^2 \left (\sum_d \nu^2(d) \right )^{-1} \\
& = & |\{(x,y) \in F \times G\}|^2 \left( \sum_d \nu^2(d) \right)^{-1} \\
& \geq & |P'|^4 \left( |P'|^4q^{-1} + q|P'|^2(t-1) \right)^{-1} \\
& = & q \left (1 + q^2(t-1)|P'|^{-2} \right)^{-1} \\
& = & q \left(1 + (t-1)^{-1}{c'}_\delta^{-2} \right)^{-1} \\
& = & \left((t-1){c'_\delta}^2 / ((t-1){c'_\delta}^2 + 1) \right ) q.
\end{array}\]
Hence, $P$ includes a point $z$ that is a vertex of triangles with at least
\[c_\epsilon q = \max_{t}\left (((t-1){c'_\delta}^2)/((t-1){c'}_\delta^2 + 1) \right )q\]
distinct areas.
To complete the proof, check that $c_\epsilon$ has the claimed properties that $c_\epsilon > 0$ for any $\epsilon$, and that $c_\epsilon \rightarrow 1$ as $\epsilon \rightarrow \infty$.
\end{proof}
\bibliographystyle{plain}
|
\section{Introduction}
Throughout this paper $G$ is a finite simple graph with vertex set $V(G)$ and
edge set $E(G)$. If $X\subseteq V\left( G\right) $, then $G[X]$ is the
subgraph of $G$ induced by $X$. By $G-W$ we mean either the subgraph
$G[V\left( G\right) -W]$, if $W\subseteq V(G)$, or the subgraph obtained by
deleting the edge set $W$, for $W\subseteq E(G)$. In either case, we use
$G-w$, whenever $W$ $=\{w\}$. If $A,B$ $\subseteq V\left( G\right) $, then
$(A,B)$ stands for the set $\{ab:a\in A,b\in B,ab\in E\left( G\right) \}$.
The \textit{neighborhood} $N(v)$ of a vertex $v\in V\left( G\right) $ is the
set $\{w:w\in V\left( G\right) $ \textit{and} $vw\in E\left( G\right) \}$,
while the \textit{closed neighborhood} $N[v]$\ of $v\in V\left( G\right) $
is the set $N(v)\cup\{v\}$; in order to avoid ambiguity, we use also
$N_{G}(v)$ instead of $N(v)$.
The \textit{neighborhood} $N(A)$ of $A\subseteq V\left( G\right) $ is
$\{v\in V\left( G\right) :N(v)\cap A\neq\emptyset\}$, and $N[A]=N(A)\cup A$.
We may also use $N_{G}(A)$ and $N_{G}\left[ A\right] $, when referring to
neighborhoods in a graph $G$.
A set $S\subseteq V(G)$ is \textit{independent} if no two vertices from $S$
are adjacent, and by $\mathrm{Ind}(G)$ we mean the family of all the
independent sets of $G$. An independent set of maximum size is a
\textit{maximum independent set} of $G$, and the \textit{independence number
}$\alpha(G)$ of $G$ is $\max\{\left\vert S\right\vert :S\in\mathrm{Ind}(G)\}$.
Let $\Omega(G)$ denote the family of all maximum independent sets, and le
\begin{align*}
\mathrm{core}(G) &
{\displaystyle\bigcap}
\{S:S\in\Omega(G)\}\text{ \cite{LevMan2002a}, and}\\
\mathrm{corona}(G) & =\cup\{S:S\in\Omega(G)\}\text{ \cite{BorosGolLev}.
\end{align*}
Clearly, $N\left( \mathrm{core}(G)\right) \subseteq$ $V\left( G\right)
-\mathrm{corona}(G)$, and there are graphs with $N\left( \mathrm{core
(G)\right) $ $\neq$ $V\left( G\right) -\mathrm{corona}(G)$ (for an example,
see Figure \ref{fig101}). The problem of whether $\mathrm{core}(G)\neq
\emptyset$ is \textbf{NP}-hard \cite{BorosGolLev}.
\begin{figure}[h]
\setlength{\unitlength}{1.0cm} \begin{picture}(5,1.75)\thicklines
\multiput(5,0.5)(1,0){5}{\circle*{0.29}}
\multiput(5,1.5)(1,0){5}{\circle*{0.29}}
\put(5,0.5){\line(1,0){4}}
\put(5,1.5){\line(1,-1){1}}
\put(6,1.5){\line(1,-1){1}}
\put(6,1.5){\line(1,0){1}}
\put(7,0.5){\line(0,1){1}}
\put(7,0.5){\line(1,1){1}}
\put(8,1.5){\line(1,0){1}}
\put(9,0.5){\line(0,1){1}}
\put(4.65,0.5){\makebox(0,0){$a$}}
\put(4.65,1.5){\makebox(0,0){$b$}}
\put(6,0.2){\makebox(0,0){$c$}}
\put(7,0.2){\makebox(0,0){$d$}}
\put(8,0.2){\makebox(0,0){$e$}}
\put(9.3,0.5){\makebox(0,0){$f$}}
\put(5.65,1.5){\makebox(0,0){$x$}}
\put(7.3,1.5){\makebox(0,0){$y$}}
\put(8,1.2){\makebox(0,0){$u$}}
\put(9.3,1.5){\makebox(0,0){$v$}}
\put(3.3,1){\makebox(0,0){$G$}}
\end{picture}\caption{\textrm{core}$(G)=\{a,b\}$ and $V(G)-$ \textrm{corona
$(G)=N\left( \mathrm{core}(G)\right) \cup\{d\}=\{c,d\}$.
\label{fig101
\end{figure}
A \textit{matching} is a set $M$ of pairwise non-incident edges of $G$. If
$A\subseteq V(G)$, then $M\left( A\right) $ is the set of all the vertices
matched by $M$ with vertices belonging to $A$. A matching of maximum
cardinality, denoted $\mu(G)$, is a \textit{maximum matching}.
For $X\subseteq V(G)$, the number $\left\vert X\right\vert -\left\vert
N(X)\right\vert $ is the \textit{difference} of $X$, denoted $d(X)$. The
\textit{critical difference} $d(G)$ is $\max\{d(X):X\subseteq V(G)\}$. The
number $\max\{d(I):I\in\mathrm{Ind}(G)\}$ is the \textit{critical independence
difference} of $G$, denoted $id(G)$. Clearly, $d(G)\geq id(G)$. It was shown
in \cite{Zhang1990} that $d(G)$ $=id(G)$ holds for every graph $G$. If $A$ is
an independent set in $G$ with $d\left( X\right) =id(G)$, then $A$ is a
\textit{critical independent set} \cite{Zhang1990}. All pendant vertices not
belonging to $K_{2}$ components are included in every inclusion maximal
critical independent set.
For example, let $X=\{v_{1},v_{2},v_{3},v_{4}\}$ and $I=\{v_{1},v_{2
,v_{3},v_{6},v_{7}\}$ in the graph $G$ of Figure \ref{fig511}. Note that $X$
is a critical set, since $N(X)=\{v_{3},v_{4},v_{5}\}$ and $d(X)=1=d(G)$, while
$I$ is a critical independent set, because $d(I)=1=id(G)$. Other critical sets
are $\{v_{1},v_{2}\}$, $\{v_{1},v_{2},v_{3}\}$, $\{v_{1},v_{2},v_{3
,v_{4},v_{6},v_{7}\}$. \begin{figure}[h]
\setlength{\unitlength}{1cm}\begin{picture}(5,1.9)\thicklines
\multiput(6,0.5)(1,0){6}{\circle*{0.29}}
\multiput(5,1.5)(1,0){4}{\circle*{0.29}}
\multiput(4,0.5)(0,1){2}{\circle*{0.29}}
\put(10,1.5){\circle*{0.29}}
\put(4,0.5){\line(1,0){7}}
\put(4,1.5){\line(2,-1){2}}
\put(5,1.5){\line(1,-1){1}}
\put(5,1.5){\line(1,0){1}}
\put(6,0.5){\line(0,1){1}}
\put(6,0.5){\line(1,1){1}}
\put(7,1.5){\line(1,0){1}}
\put(8,0.5){\line(0,1){1}}
\put(10,0.5){\line(0,1){1}}
\put(10,1.5){\line(1,-1){1}}
\put(4,0.1){\makebox(0,0){$v_{1}$}}
\put(3.65,1.5){\makebox(0,0){$v_{2}$}}
\put(4.65,1.5){\makebox(0,0){$v_{3}$}}
\put(6.35,1.5){\makebox(0,0){$v_{4}$}}
\put(6,0.1){\makebox(0,0){$v_{5}$}}
\put(7,0.1){\makebox(0,0){$v_{6}$}}
\put(7,1.15){\makebox(0,0){$v_{7}$}}
\put(8,0.1){\makebox(0,0){$v_{9}$}}
\put(8.35,1.5){\makebox(0,0){$v_{8}$}}
\put(9.65,1.5){\makebox(0,0){$v_{11}$}}
\put(9,0.1){\makebox(0,0){$v_{10}$}}
\put(10,0.1){\makebox(0,0){$v_{12}$}}
\put(11,0.1){\makebox(0,0){$v_{13}$}}
\put(2.5,1){\makebox(0,0){$G$}}
\end{picture}\caption{\textrm{core}$(G)=\{v_{1},v_{2},v_{6},v_{10}\}$ is a
critical set.
\label{fig511
\end{figure}
It is known that finding a maximum independent set is an \textbf{NP}-hard
problem \cite{GaryJohnson79}. Zhang proved that a critical independent set can
be find in polynomial time \cite{Zhang1990}. A simpler algorithm, reducing the
critical independent set problem to computing a maximum independent set in a
bipartite graph is given in \cite{Ageev}.
\begin{theorem}
\label{th3}\cite{ButTruk2007} Each critical independent set can be enlarged to
a maximum independent set.
\end{theorem}
Theorem \ref{th3} led to an efficient way of approximating $\alpha(G)$
\cite{Truchanov2008}. Moreover, it has been shown that a critical independent
set of maximum cardinality can be computed in polynomial time
\cite{Larson2007}. Recently, a parallel algorithm computing the critical
independence number was developed \cite{DeLaVinaLarson2013}.
Recall that if $\alpha(G)+\mu(G)=\left\vert V(G)\right\vert $, then $G$ is a
\textit{K\"{o}nig-Egerv\'{a}ry graph} \cite{Deming1979,Sterboul1979}. As a
well-known example, each bipartite graph is a K\"{o}nig-Egerv\'{a}ry graph as well.
\begin{theorem}
\label{Th5} \cite{LevMan2003} If $G$ is a K\"{o}nig-Egerv\'{a}ry graph, $M$ is
a maximum matching of $G$, and $S\in\Omega\left( G\right) $, then:
\emph{(i)} $M$ matches $V\left( G\right) -S$ into $S$, and $N(\mathrm{core
(G))$ into $\mathrm{core}(G)$;
\emph{(ii)} $N\left( \mathrm{core}(G)\right) =\cap\left\{ V(G)-S:S\in
(G)\right\} $, i.e., $N\left( \mathrm{core}(G\right) )=V\left( G\right)
-\mathrm{corona}(G)$.
\end{theorem}
The \textit{deficiency} $def(G)$ is the number of non-saturated vertices
relative to a maximum matching, i.e., $def(G)=\left\vert V\left( G\right)
\right\vert -2\mu(G)$ \cite{LovPlum1986}. A proof of a conjecture of
Graffiti.pc \cite{DeLaVina} yields a new characterization of
K\"{o}nig-Egerv\'{a}ry graphs: these are exactly the graphs, where there
exists a critical maximum independent set \cite{Larson2011}. In
\cite{LevMan2012b} it is proved the following.
\begin{theorem}
\label{th8}\cite{LevMan2012b} For a K\"{o}nig-Egerv\'{a}ry graph $G$ the
following equalities hold
\[
d(G)=\left\vert \mathrm{core}(G)\right\vert -\left\vert N(\mathrm{core
(G))\right\vert =\alpha(G)-\mu(G)=def(G).
\]
\end{theorem}
Using this finding, we have strengthened the characterization from
\cite{Larson2011}.
\begin{theorem}
\label{th5}\cite{LevMan2012b} $G$ is a K\"{o}nig-Egerv\'{a}ry graph if and
only if each of its maximum independent sets is critical.
\end{theorem}
For a graph $G$, let denote
\begin{align*}
\mathrm{\ker}(G) &
{\displaystyle\bigcap}
\left\{ S:S\text{ \textit{is a critical independent set}}\right\} \text{
\cite{LevMan2012a}, and}\\
\mathrm{diadem}(G) & =\bigcup\left\{ S:S\text{ \textit{is a critical
independent set}}\right\} \text{.
\end{align*}
In this paper we present several properties of $\mathrm{\ker}(G)$, in relation
with $\mathrm{core}(G)$, $\mathrm{corona}(G)$, and $\mathrm{diadem}(G)$.
\section{Preliminaries}
Let $G$ be the graph from Figure \ref{fig511}; the sets $X=\left\{
v_{1},v_{2},v_{3}\right\} $, $Y=\left\{ v_{1},v_{2},v_{4}\right\} $\ are
critical independent, and the sets $X\cap Y$, $X\cup Y$ are also critical, but
only $X\cap Y$ is also independent.\ In addition, one can easily see that
$\mathrm{\ker}(G)$ is a minimal critical independent set of $G$. These
properties of critical sets and $\mathrm{\ker}(G)$ are true even in general.
\begin{theorem}
\label{th4}\cite{LevMan2012a} For a graph $G$, the following assertions are true:
\emph{(i)} the function $d$ is supermodular, i.e., $d(A\cup B)+d(A\cap B)\geq
d(A)+d(B)$ for every $A,B\subseteq V(G)$;
\emph{(ii)} if $A$ and $B$ are critical in $G$, then $A\cup B$ and $A\cap B$
are critical as well;
\emph{(iii)} $G$ has a unique minimal independent critical set, namely,
$\mathrm{\ker}(G)$.
\end{theorem}
As a consequence, we have the following.
\begin{corollary}
For every graph $G$, $\mathrm{diadem}(G)$ is a critical set.
\end{corollary}
For instance, the graph $G$ from Figure \ref{fig511} has $\mathrm{diadem
(G)=\left\{ v_{1},v_{2},v_{3},v_{4},v_{6},v_{7},v_{10}\right\} $, which is
critical, but not independent.
\begin{figure}[h]
\setlength{\unitlength}{1cm}\begin{picture}(5,1.8)\thicklines
\multiput(1,0.5)(1,0){6}{\circle*{0.29}}
\multiput(1,1.5)(1,0){5}{\circle*{0.29}}
\put(1,0.5){\line(1,0){5}}
\put(1,1.5){\line(1,-1){1}}
\put(2,0.5){\line(0,1){1}}
\put(3,1.5){\line(1,-1){1}}
\put(3,1.5){\line(1,0){1}}
\put(4,0.5){\line(0,1){1}}
\put(5,1.5){\line(1,-1){1}}
\put(5,0.5){\line(0,1){1}}
\put(0.7,1.5){\makebox(0,0){$a$}}
\put(0.7,0.5){\makebox(0,0){$b$}}
\put(2.3,1.5){\makebox(0,0){$c$}}
\put(3,0.85){\makebox(0,0){$d$}}
\put(3,0){\makebox(0,0){$G_{1}$}}
\multiput(7,0.5)(1,0){7}{\circle*{0.29}}
\multiput(7,1.5)(1,0){6}{\circle*{0.29}}
\put(7,0.5){\line(1,0){6}}
\put(7,1.5){\line(1,-1){1}}
\put(8,0.5){\line(0,1){1}}
\put(9,0.5){\line(0,1){1}}
\put(9,0.5){\line(1,1){1}}
\put(9,1.5){\line(1,0){1}}
\put(11,0.5){\line(0,1){1}}
\put(12,1.5){\line(1,-1){1}}
\put(11,1.5){\line(1,0){1}}
\put(6.7,1.5){\makebox(0,0){$x$}}
\put(6.7,0.5){\makebox(0,0){$y$}}
\put(8.3,1.5){\makebox(0,0){$z$}}
\put(10,0.8){\makebox(0,0){$w$}}
\put(10,0){\makebox(0,0){$G_{2}$}}
\end{picture}\caption{Both $G_{1}$ and $G_{2}$\ are not K\"{o}nig-Egerv\'{a}ry
graphs.
\label{fig22
\end{figure}
The graph $G$ from Figure \ref{fig101} has $d\left( G\right) =1$ and
$d\left( \mathrm{corona}(G)\right) =0$, which means that $\mathrm{corona
(G)$ is not a critical set. Notice that $G$ is not a K\"{o}nig-Egerv\'{a}ry
graph. Combining Theorems \ref{th5} and \ref{th4}\emph{(ii)}, we deduce the following.
\begin{corollary}
\label{cor2}If $G$ is a K\"{o}nig-Egerv\'{a}ry graph, then both $\mathrm{core
(G)$ and $\mathrm{corona}(G)$ are critical sets.
\end{corollary}
Let consider the graphs $G_{1}$ and $G_{2}$ from Figure \ref{fig22}:
$\mathrm{core}(G_{1})=\left\{ a,b,c,d\right\} $ and it is a critical set,
while $\mathrm{core}(G_{2})=\left\{ x,y,z,w\right\} $ and it is not critical.
\begin{theorem}
If $\mathrm{core}(G)$ is a critical set, then
\[
\mathrm{core}(G)\subsete
{\displaystyle\bigcap}
\left\{ A:A\text{ \textit{is an inclusion maximal critical independent set
}\right\} .
\]
\end{theorem}
\begin{proof}
Let $A$ be an arbitrary inclusion maximal critical independent set. According
to Theorem \ref{th3}, there is some $S\in\Omega\left( G\right) $, such that
$A\subseteq S$. Since $\mathrm{core}(G)\subseteq S$, it follows that
$A\cup\mathrm{core}(G)\subseteq S$, and hence $A\cup\mathrm{core}(G)$ is
independent. By Theorem \ref{th4}, we get that $A\cup\mathrm{core}(G)$ is a
critical independent set. Since $A\subseteq A\cup\mathrm{core}(G)$ and $A$ is
an inclusion maximal critical independent set, it follows that $\mathrm{core
(G)\subseteq A$, for every such set $A$, and this completes the proof.
\end{proof}
\begin{remark}
By Theorem \ref{th3} the following inclusion holds for every graph $G$.
\[
\mathrm{corona}(G)\supsete
{\displaystyle\bigcup}
\left\{ A:A\text{ \textit{is an inclusion maximal critical independent set
}\right\} .
\]
\end{remark}
\section{Structural properties of $\mathrm{\ker}\left( G\right) $}
Deleting a vertex from a graph may change its critical difference. For
instance, $d\left( G-v_{1}\right) =d\left( G\right) -1$, $d\left(
G-v_{13}\right) =d\left( G\right) $, while $d\left( G-v_{3}\right)
=d\left( G\right) +1$, where $G$ is the graph of Figure \ref{fig511}.
\begin{proposition}
\cite{LevMan2013a} For a vertex $v$ in a graph $G$, the following assertions hold:
\emph{(i) }$d\left( G-v\right) =d\left( G\right) -1$ if and only if
$v\in\mathrm{\ker}(G)$;
\emph{(ii)} if $v\in\mathrm{\ker}(G)$, then $\mathrm{\ker}(G-v)\subseteq
\mathrm{\ker}(G)-\left\{ v\right\} $.
\end{proposition}
Note that $\mathrm{\ker}(G-v)$ may differ from $\mathrm{\ker}(G)-\left\{
v\right\} $. For example, $\mathrm{\ker}(K_{3,2})$ is equal to the partite
set of size $3$, but $\mathrm{\ker}(K_{3,2}-v)=\emptyset$ whenever $v$ is in
that set. Also, if $G=C_{4}$, then $\mathrm{\ker}(G)-\left\{ v\right\}
=\emptyset-\left\{ v\right\} =\emptyset$, while $\mathrm{\ker
(G-v)=N_{G}(v)$ for every $v\in V\left( G\right) $.
\begin{theorem}
\label{th2}\cite{Larson2007} There is a matching from $N(S)$ into $S$ for
every critical independent set $S$.
\end{theorem}
In the graph $G$ of Figure \ref{fig511}, let $S=\{v_{1},v_{2},v_{3}\}$. By
Theorem \ref{th2}, there is a matching from $N\left( S\right) $ into
$S=\left\{ v_{1},v_{2},v_{3}\right\} $, for instance, $M=\left\{ v_{2
v_{5},v_{3}v_{4}\right\} $, since $S$ is critical independent. On the other
hand, there is no matching from $N\left( S\right) $ into $S-v_{3}$.
\begin{theorem}
\cite{LevMan2013a}\label{th9} For a critical independent set $A$ in a graph
$G$, the following statements are equivalent:
\emph{(i) }$A=\mathrm{\ker}(G)$;
\emph{(ii)} there is no set $B\subseteq$ $N\left( A\right) ,B\neq\emptyset$
such that $\left\vert N\left( B\right) \cap A\right\vert =\left\vert
B\right\vert $;
\emph{(iii) }for each $v\in A$ there exists a matching from $N\left(
A\right) $ into $A-v$.
\end{theorem}
The graphs $G_{1}$ and $G_{2}$ in Figure \ref{fig333} satisfy $\mathrm{\ker
}(G_{1})=\mathrm{core}(G_{1})$, $\mathrm{\ker}(G_{2})=\left\{ x,y,z\right\}
\subset\mathrm{core}(G_{2})$, and both $\mathrm{core}(G_{1})$ and
$\mathrm{core}(G_{2})$\ are critical sets of maximum size. The graph $G_{3}$
in Figure \ref{fig333} has $\mathrm{\ker}(G_{3})=\{u,v\}$, the set
$\{t,u,v\}$\ as a critical independent set of maximum size, while
$\mathrm{core}(G_{3})=\left\{ t,u,v,w\right\} $\ is not a critical set.
\begin{figure}[h]
\setlength{\unitlength}{1.0cm} \begin{picture}(5,1.9)\thicklines
\multiput(1,0.5)(1,0){3}{\circle*{0.29}}
\multiput(2,1.5)(1,0){2}{\circle*{0.29}}
\put(1,0.5){\line(1,0){2}}
\put(2,0.5){\line(0,1){1}}
\put(2,0.5){\line(1,1){1}}
\put(3,0.5){\line(0,1){1}}
\put(1,0.84){\makebox(0,0){$a$}}
\put(1.7,1.5){\makebox(0,0){$b$}}
\put(2,0){\makebox(0,0){$G_{1}$}}
\multiput(4,0.5)(1,0){4}{\circle*{0.29}}
\put(4,0.5){\line(1,0){3}}
\multiput(4,1.5)(1,0){4}{\circle*{0.29}}
\put(4,1.5){\line(1,-1){1}}
\put(5,0.5){\line(0,1){1}}
\put(7,0.5){\line(0,1){1}}
\put(7,0.5){\line(-1,1){1}}
\put(6,1.5){\line(1,0){1}}
\put(6,0.84){\makebox(0,0){$q$}}
\put(4,0.84){\makebox(0,0){$x$}}
\put(4.27,1.5){\makebox(0,0){$y$}}
\put(5.27,1.5){\makebox(0,0){$z$}}
\put(5.5,0){\makebox(0,0){$G_{2}$}}
\multiput(8,0.5)(1,0){6}{\circle*{0.29}}
\multiput(8,1.5)(1,0){6}{\circle*{0.29}}
\put(8,0.5){\line(1,1){1}}
\put(8,1.5){\line(1,0){1}}
\put(9,0.5){\line(0,1){1}}
\put(9,0.5){\line(1,0){4}}
\put(10,0.5){\line(0,1){1}}
\put(10,0.5){\line(1,1){1}}
\put(10,1.5){\line(1,-1){1}}
\put(10,1.5){\line(1,0){1}}
\put(11,0.5){\line(0,1){1}}
\put(12,1.5){\line(1,-1){1}}
\put(12,1.5){\line(1,0){1}}
\put(13,0.5){\line(0,1){1}}
\put(8.3,0.5){\makebox(0,0){$v$}}
\put(8,1.15){\makebox(0,0){$u$}}
\put(9.3,0.84){\makebox(0,0){$t$}}
\put(12,0.84){\makebox(0,0){$w$}}
\put(10.5,0){\makebox(0,0){$G_{3}$}}
\end{picture}\caption{$\mathrm{core}(G_{1})=\left\{ a,b\right\} $,
$\mathrm{core}(G_{2})=\left\{ q,x,y,z\right\} $, $\mathrm{core
(G_{3})=\left\{ t,u,v,w\right\} $.
\label{fig333
\end{figure}
An independent set $S$ is \textit{inclusion minimal with} $d\left( S\right)
>0$ if no proper subset of $S$ has positive difference. For example, in Figure
\ref{fig333} one can see that $\mathrm{\ker}(G_{1})$ is an inclusion minimal
independent set with positive difference, while for the graph $G_{2}$ the sets
$\{x,y\},\{x,z\},\{y,z\}$ are inclusion minimal independent with positive
difference, and $\mathrm{\ker}(G_{2})=\{x,y\}\cup\{x,z\}\cup\{y,z\}$.
\begin{theorem}
\cite{LevMan2013a}\label{th1} If $\mathrm{\ker}(G)\neq\emptyset$, then
\begin{align*}
\mathrm{\ker}(G) &
{\displaystyle\bigcup}
\left\{ S_{0}:S_{0}\text{ is an inclusion minimal independent set with
}d\left( S_{0}\right) =1\right\} \\
&
{\displaystyle\bigcup}
\left\{ S_{0}:S_{0}\text{ is an inclusion minimal independent set with
}d\left( S_{0}\right) >0\right\} .
\end{align*}
\end{theorem}
In a graph $G$, the union of all minimum cardinality independent sets $S$ with
$d\left( S\right) >0$ may be a proper subset of $\mathrm{\ker}\left(
G\right) $. For example, consider the graph $G$ in Figure \ref{fig177}, where
$\left\{ x,y\right\} \subset\mathrm{\ker}\left( G\right) =\left\{
x,y,u,v,w\right\} $.
\begin{figure}[h]
\setlength{\unitlength}{1cm}\begin{picture}(5,1.3)\thicklines
\multiput(4,0)(1,0){3}{\circle*{0.29}}
\multiput(3,1)(1,0){5}{\circle*{0.29}}
\put(4,0){\line(1,0){2}}
\put(4,0){\line(0,1){1}}
\put(3,1){\line(1,-1){1}}
\put(5,0){\line(0,1){1}}
\put(5,0){\line(2,1){2}}
\put(6,0){\line(0,1){1}}
\put(6,0){\line(1,1){1}}
\put(2.7,1){\makebox(0,0){$x$}}
\put(3.7,1){\makebox(0,0){$y$}}
\put(4.7,1){\makebox(0,0){$u$}}
\put(5.7,1){\makebox(0,0){$v$}}
\put(7.3,1){\makebox(0,0){$w$}}
\put(2,0.5){\makebox(0,0){$G$}}
\multiput(9,0)(1,0){3}{\circle*{0.29}}
\multiput(9,1)(2,0){2}{\circle*{0.29}}
\put(9,0){\line(1,0){2}}
\put(9,1){\line(1,-1){1}}
\put(10,0){\line(1,1){1}}
\put(8.65,0){\makebox(0,0){$v_{1}$}}
\put(8.65,1){\makebox(0,0){$v_{2}$}}
\put(11.35,1){\makebox(0,0){$v_{3}$}}
\put(11.35,0){\makebox(0,0){$v_{4}$}}
\put(8,0.5){\makebox(0,0){$H$}}
\end{picture}\caption{Both $S_{1}=\{x,y\}$ and $S_{2}=\{u,v,w\}$ are inclusion
minimal independent sets satisfying $d\left( S\right) >0$.
\label{fig177
\end{figure}
Actually, all inclusion minimal independent sets $S$ with $d(S)>0$ are of the
same difference.
\begin{proposition}
\cite{LevMan2013a}\label{prop3} If $S_{0}$ is an inclusion minimal independent
set with $d\left( S_{0}\right) >0$, then $d\left( S_{0}\right) =1$. In
other words
\begin{gather*}
\left\{ S_{0}:S_{0}\text{ is an inclusion minimal independent set with
}d\left( S_{0}\right) >0\right\} =\\
=\left\{ S_{0}:S_{0}\text{ is an inclusion minimal independent set with
}d\left( S_{0}\right) =1\right\} .
\end{gather*}
\end{proposition}
The converse of Proposition \ref{prop3} is not true. For instance, $S=\left\{
x,y,u\right\} $ is independent in the graph $G$ of Figure \ref{fig177}\ and
$d\left( S\right) =1$, but $S$ is not minimal with this property.
\begin{proposition}
\cite{LevMan2013a} $\min\left\{ \left\vert S_{0}\right\vert :d\left(
S_{0}\right) >0,S_{0}\in\mathrm{Ind}(G)\right\} \leq\left\vert \mathrm{\ker
}\left( G\right) \right\vert -d\left( G\right) +1$ is true for every graph
$G$.
\end{proposition}
\section{Relationships between $\mathrm{\ker}\left( G\right) $ and
$\mathrm{core}(G)$}
Let us consider again the graph $G_{2}$ from Figure \ref{fig22}:
$\mathrm{core}(G_{2})=\left\{ x,y,z,w\right\} $ and it is not critical, but
$\mathrm{\ker}\left( G_{2}\right) =\left\{ x,y,z\right\} \subseteq
\mathrm{core}(G_{2})$. Clearly, the same inclusion holds for $G_{1}$, whose
$\mathrm{core}(G_{1})$ is a critical set.
\begin{theorem}
\label{th6}\cite{LevMan2012a} For every graph $G$, $\mathrm{\ker
(G)\subseteq\mathrm{core}(G)$.
\end{theorem}
Let $I_{c}$ be a maximum critical independent set of $G$, and $X=I_{c}\cup
N(I_{c})$. In \cite{Short} it is proved that $\mathrm{core}(G\left[ X\right]
)\subseteq\mathrm{core}(G)$. Moreover, in \cite{LevMan2012a}, we showed that
the chain of relationships $\mathrm{\ker}(G)=\mathrm{\ker}(G\left[ X\right]
)\subseteq\mathrm{core}(G\left[ X\right] )\subseteq\mathrm{core}(G)$ holds
for every graph $G$. Theorem \ref{th6} allows an alternative proof of the
following inequality due to Lorentzen.
\begin{corollary}
\label{cor1}\cite{Lorentzen1966,Schrijver2003,LevMan2012a} The inequality
$d\left( G\right) \geq\alpha\left( G\right) -\mu\left( G\right) $ holds
for every graph.
\end{corollary}
Following Ore \cite{Ore55}, \cite{Ore62}, the number $\delta(X)=d\left(
X\right) =\left\vert X\right\vert -\left\vert N\left( X\right) \right\vert
$ is the \textit{deficiency} of $X$, where $X\subseteq A$ or $X\subseteq B$
and $G=(A,B,E)$ is a bipartite graph. Let
\[
\delta_{0}(A)=\max\{\delta(X):X\subseteq A\},\quad\delta_{0}(B)=\max
\{\delta(Y):Y\subseteq B\}.
\]
A subset $X\subseteq A$ having $\delta(X)=\delta_{0}(A)$ is $A
-\textit{critical}, while $Y\subseteq B$ having $\delta(B)=\delta_{0}(B)$ is
$B$-\textit{critical}. For a bipartite graph $G=\left( A,B,E\right) $ let us
denote $\mathrm{\ker}_{A}(G)=\cap\left\{ S:S\text{ is
A\text{-\textit{critical}}\right\} $ and \textrm{diadem}$_{A}(G)=\cup\left\{
S:S\text{ is }A\text{-\textit{critical}}\right\} $. Similarly, $\mathrm{\ker
}_{B}(G)=\cap\left\{ S:S\text{ is }B\text{-\textit{critical}}\right\} $ and
\textrm{diadem}$_{B}(G)=\cup\left\{ S:S\text{ is }B\text{-\textit{critical
}\right\} $.
It is convenient to define $d\left( \emptyset\right) =\delta(\emptyset)=0$.
\begin{figure}[h]
\setlength{\unitlength}{1cm}\begin{picture}(5,1.9)\thicklines
\multiput(5,0.5)(1,0){7}{\circle*{0.29}}
\multiput(4,1.5)(1,0){6}{\circle*{0.29}}
\put(5,0.5){\line(-1,1){1}}
\put(5,0.5){\line(0,1){1}}
\put(5,0.5){\line(1,1){1}}
\put(6,0.5){\line(0,1){1}}
\put(6,0.5){\line(1,1){1}}
\put(7,0.5){\line(-1,1){1}}
\put(7,0.5){\line(0,1){1}}
\put(7,0.5){\line(1,1){1}}
\put(8,0.5){\line(0,1){1}}
\put(8,0.5){\line(1,1){1}}
\put(9,0.5){\line(0,1){1}}
\put(9,1.5){\line(1,-1){1}}
\put(9,1.5){\line(2,-1){2}}
\put(3.65,1.5){\makebox(0,0){$a_{1}$}}
\put(4.65,1.5){\makebox(0,0){$a_{2}$}}
\put(6.35,1.5){\makebox(0,0){$a_{3}$}}
\put(7.35,1.5){\makebox(0,0){$a_{4}$}}
\put(8.35,1.5){\makebox(0,0){$a_{5}$}}
\put(9.35,1.5){\makebox(0,0){$a_{6}$}}
\put(5,0.1){\makebox(0,0){$b_{1}$}}
\put(6,0.1){\makebox(0,0){$b_{2}$}}
\put(7,0.1){\makebox(0,0){$b_{3}$}}
\put(8,0.1){\makebox(0,0){$b_{4}$}}
\put(9,0.1){\makebox(0,0){$b_{5}$}}
\put(10,0.1){\makebox(0,0){$b_{6}$}}
\put(11,0.1){\makebox(0,0){$b_{7}$}}
\put(3,1){\makebox(0,0){$G$}}
\end{picture}\caption{$G$ is a bipartite graph without perfect matchings.
\label{fig233
\end{figure}For instance, the graph $G=(A,B,E)$ from Figure \ref{fig233} has:
$X=\left\{ a_{1},a_{2},a_{3},a_{4}\right\} $ as an $A$-critical set,
$\mathrm{\ker}_{A}(G)=\left\{ a_{1},a_{2}\right\} $, \textrm{diadem
$_{A}(G)=\left\{ a_{i}:i=1,...,5\right\} $ and $\delta_{0}(A)=1$, while
$Y=\left\{ b_{i}:i=4,5,6,7\right\} $ is a $B$-critical set, $\mathrm{\ker
}_{B}(G)=\left\{ b_{4},b_{5},b_{6}\right\} $, \textrm{diadem}$_{B
(G)=\left\{ b_{i}:i=2,...,7\right\} $ and $\delta_{0}(B)=2$.
As expected, there is a close relationship between critical independent sets
and $A$-critical\textit{ }or\textit{ }$B$-critical sets.
\begin{theorem}
\cite{LevMan2011b}\label{Th4} Let $G=\left( A,B,E\right) $ be a bipartite
graph. Then the following assertions are true:
\emph{(i) }$d(G)=\delta_{0}(A)+\delta_{0}(B)$;
\emph{(ii) }$\alpha\left( G\right) =\left\vert A\right\vert +\delta
_{0}(B)=\left\vert B\right\vert +\delta_{0}(A)=\mu\left( G\right)
+\delta_{0}(A)+\delta_{0}(B)=\mu\left( G\right) +d\left( G\right) $;
\emph{(iii)} if $X$ is an $A$-\textit{critical set and }$Y$\textit{ is a
$B$-\textit{critical set, then }$X\cup Y$ is a critical set;
\emph{(iv)} if $Z$ is a critical independent set, then $Z\cap A$ is an
$A$-\textit{critical set }and $Z\cap B$ is \textit{a }$B$-\textit{critical
set;}
\emph{(v)} if $X$ is either an $A$-\textit{critical set or a }$B
-\textit{critical set, then there is a matching from }$N\left( X\right) $
into $X$.
\end{theorem}
The following lemma will be used further to give an alternative proof for the
assertion that $\mathrm{\ker}(G)=\mathrm{core}(G)$ holds for every bipartite
graph $G$.
\begin{lemma}
\label{lem1}If $G=\left( A,B,E\right) $ is a bipartite graph with a perfect
matching, say $M$, $S\in\Omega\left( G\right) $, $X\in$ $\mathrm{Ind}(G)$,
$X\subseteq V\left( G\right) -S$, and $G\left[ X\cup M\left( X\right)
\right] $ is connected, then
\[
X^{1}=X\cup M\left( \left( N\left( X\right) \cap S\right) -M\left(
X\right) \right)
\]
is an independent set, and $G\left[ X^{1}\cup M\left( X^{1}\right) \right]
$ is connected.
\end{lemma}
\begin{proof}
Let us show that the set $M\left( \left( N\left( X\right) \cap S\right)
-M\left( X\right) \right) $ is independent. Suppose, to the contrary, that
there exist $v_{1},v_{2}\in M\left( \left( N\left( X\right) \cap S\right)
-M\left( X\right) \right) $ such that $v_{1}v_{2}\in E\left( G\right) $.
Hence $M\left( v_{1}\right) ,M\left( v_{2}\right) \in\left( N\left(
X\right) \cap S\right) -M\left( X\right) $.
If $M\left( v_{1}\right) $ and $M\left( v_{2}\right) $ have a common
neighbor $w\in X$, then $\left\{ v_{1},v_{2},M\left( v_{2}\right)
,w,M\left( v_{1}\right) \right\} $ spans $C_{5}$, which is forbidden for
bipartite graphs.
Otherwise, let $w_{1},w_{2}\in X$ be neighbors of $M\left( v_{1}\right) $
and $M\left( v_{2}\right) $, respectively. Since $G\left[ X\cup M\left(
X\right) \right] $ is connected, there is a path with even number of edges
connecting $w_{1}$ and $w_{2}$. Together with $\left\{ w_{1},M\left(
v_{1}\right) ,v_{1},v_{2},M\left( v_{2}\right) ,w_{2}\right\} $ this path
produces a cycle of odd length in contradiction with the hypothesis on $G$
being a bipartite graph.
To complete the proof of independence of the set
\[
X^{1}=X\cup M\left( \left( N\left( X\right) \cap S\right) -M\left(
X\right) \right)
\]
it is enough to demonstrate that there are no edges connecting vertices of $X$
and $M\left( \left( N\left( X\right) \cap S\right) -M\left( X\right)
\right) $. \begin{figure}[h]
\setlength{\unitlength}{1.0cm} \begin{picture}(5,4)\thicklines
\put(6.7,1){\oval(10,1.5)}
\put(5,1){\oval(4,1)}
\put(5,3){\oval(4,1)}
\put(6.7,3){\oval(10,1.5)}
\put(9,1){\oval(2.5,1)}
\put(9,3){\oval(2.5,1)}
\multiput(3.5,1)(0.7,0){2}{\circle*{0.29}}
\multiput(3.5,3)(0.7,0){2}{\circle*{0.29}}
\multiput(5.5,1)(0,2){2}{\circle*{0.29}}
\multiput(6.5,1)(0,2){2}{\circle*{0.29}}
\multiput(4.5,1)(0.35,0){3}{\circle*{0.1}}
\multiput(4.5,3)(0.35,0){3}{\circle*{0.1}}
\put(5.5,1){\line(0,1){2}}
\put(5.5,1){\line(3,2){3}}
\put(6.5,1){\line(0,1){2}}
\put(6.5,1){\line(3,2){3}}
\put(6.5,1){\line(1,1){2}}
\multiput(3.5,1)(0.7,0){2}{\line(0,1){2}}
\put(1,1){\makebox(0,0){$V-S$}}
\put(0.2,2.2){\makebox(0,0){$G$}}
\put(2.5,1){\makebox(0,0){$X$}}
\put(2.45,3){\makebox(0,0){$M(X)$}}
\put(1.2,3){\makebox(0,0){$S$}}
\multiput(8.5,1)(1,0){2}{\circle*{0.29}}
\multiput(8.5,3)(1,0){2}{\circle*{0.29}}
\put(8.5,1){\line(0,1){2}}
\put(9.5,1){\line(0,1){2}}
\put(10.85,1){\makebox(0,0){$M(Y)$}}
\put(10.75,3){\makebox(0,0){$Y$}}
\end{picture}\caption{$S\in\Omega(G)$, $Y=\left( N\left( X\right) \cap
S\right) -M\left( X\right) ${ and }$X^{1}=X\cup M\left( Y\right) $.
\end{figure}
Assume, to the contrary, that there is $vw\in E$, such that $v\in M\left(
\left( N\left( X\right) \cap S\right) -M\left( X\right) \right) $ and
$w\in X$. Since $M\left( v\right) \in\left( N\left( X\right) \cap
S\right) -M\left( X\right) $ and $G\left[ X\cup M\left( X\right)
\right] $ is connected, it follows that there exists a path with an odd
number of edges connecting $M\left( v\right) $ to $w$. This path together
with the edges $vw$ and $vM\left( v\right) $ produces cycle of odd length,
in contradiction with the bipartiteness of $G$.
Finally, since $G\left[ X\cup M\left( X\right) \right] $ is connected,
$G\left[ X^{1}\cup M\left( X^{1}\right) \right] $ is connected as well, by
definitions of set functions $N$ and $M$.
\end{proof}
Theorem \ref{th6} claims that $\mathrm{\ker}(G)\subseteq\mathrm{core}(G)$ for
every graph.
\begin{theorem}
\cite{LevMan2011b}\label{th10} If $G$ is a bipartite graph, then
$\mathrm{\ker}(G)=\mathrm{core}(G)$.
\end{theorem}
\begin{proof}
[Alternative Proof]The assertions are clearly true, whenever $\mathrm{core
(G)=\emptyset$, i.e., for $G$ having a perfect matching. Assume that
$\mathrm{core}(G)\neq\emptyset$.
Let $S\in\Omega\left( G\right) $ and $M$ be a maximum matching. By Theorem
\ref{Th5}\emph{(i)}, $M$ matches $V\left( G\right) -S$ into $S$, and
$N(\mathrm{core}(G))$ into $\mathrm{core}(G)$.
According to Theorem \ref{th9}\emph{(ii)}, it is sufficient to show that there
is no set $Z\subseteq N\left( \mathrm{core}(G)\right) $, $Z\neq\emptyset$,
such that $\left\vert N\left( Z\right) \cap\mathrm{core}(G)\right\vert
=\left\vert Z\right\vert $.
Suppose, to the contrary, that there exists a non-empty set $Z\subseteq$
$N\left( \mathrm{core}(G)\right) $ such that $\left\vert N\left( Z\right)
\cap\mathrm{core}(G)\right\vert =\left\vert Z\right\vert $. Let $Z_{0}$ be a
minimal non-empty subset of $N\left( \mathrm{core}(G)\right) $ enjoying this equality.
Clearly, $H=G\left[ Z_{0}\cup M\left( Z_{0}\right) \right] $ is bipartite,
because it is a subgraph of a bipartite graph. Moreover, the restriction of
$M$ on $H$ is a perfect matching.
\textit{Claim 1.} $Z_{0}$ is independent.
Since $H$ is a bipartite graph with a perfect matching it has two maximum
independent sets at least. Hence there exists $W\in\Omega\left( H\right) $
different from $M\left( Z_{0}\right) $. Thus $W\cap Z_{0}\neq\emptyset$.
Therefore, $N\left( W\cap Z_{0}\right) \cap\mathrm{core}(G)=M\left( W\cap
Z_{0}\right) $. Consequently,
\[
\left\vert N\left( W\cap Z_{0}\right) \cap\mathrm{core}(G)\right\vert
=\left\vert M\left( W\cap Z_{0}\right) \right\vert =\left\vert W\cap
Z_{0}\right\vert .
\]
Finally, $W\cap Z_{0}=Z_{0}$, because $Z_{0}$ has been chosen as a minimal
subset of $N\left( \mathrm{core}(G)\right) $ such that $\left\vert N\left(
Z_{0}\right) \cap\mathrm{core}(G)\right\vert =\left\vert Z_{0}\right\vert $.
Since $\left\vert Z_{0}\right\vert =\alpha\left( H\right) =\left\vert
W\right\vert $ we conclude with $W=Z_{0}$, which means, in particular, that
$Z_{0}$ is independent.
\textit{Claim 2.} $H$ is a connected graph.
Otherwise, for any connected component of $H$, say $\tilde{H}$, the set
$V\left( \tilde{H}\right) \cap Z_{0}$ contradicts the minimality property of
$Z_{0}$.
\textit{Claim 3.} $Z_{0}\cup$ $\left( \mathrm{core}(G)-M\left( Z_{0}\right)
\right) $ is independent.
By Claim 1 $Z_{0}$ is independent. The equality $\left\vert N\left(
Z_{0}\right) \cap\mathrm{core}(G)\right\vert =\left\vert Z_{0}\right\vert $
implies $N\left( Z_{0}\right) \cap\mathrm{core}(G)=M\left( Z_{0}\right) $,
which means that there are no edges connecting $Z_{0}$ and $\mathrm{core
(G)-M\left( Z_{0}\right) $. Consequently, $Z_{0}\cup$ $\left(
\mathrm{core}(G)-M\left( Z_{0}\right) \right) $ is independent.
\textit{Claim 4.} $Z_{0}\cup$ $\left( \mathrm{core}(G)-M\left( Z_{0}\right)
\right) $ is included in a maximum independent set.
Let $Z_{i}=M\left( \left( N\left( Z_{i-1}\right) \cap S\right) -M\left(
Z_{i-1}\right) \right) ,1\leq i<\infty$. By Lemma \ref{lem1} all the sets
$Z^{i}=\bigcup\limits_{0\leq j\leq i}Z_{j},1\leq i<\infty$ are independent.
Defin
\[
Z^{\infty}=\bigcup\limits_{0\leq i\leq\infty}Z_{i},
\]
which is, actually, the largest set in the sequence $\left\{ Z^{i},1\leq
i<\infty\right\} $.\begin{figure}[h]
\setlength{\unitlength}{1.0cm} \begin{picture}(5,4)\thicklines
\put(7.5,1){\oval(11,1.5)}
\put(6.7,3){\oval(12.5,1.7)}
\put(3.5,3){\oval(4.35,1.2)}
\put(2.5,3){\oval(1.5,0.8)}
\put(4.5,1){\oval(2,0.8)}
\put(4.5,3){\oval(2,0.8)}
\put(7.5,1){\oval(2,0.8)}
\put(7.5,3){\oval(2,0.8)}
\put(10.5,1){\oval(2,0.8)}
\put(10.5,3){\oval(2,0.8)}
\multiput(4,1)(1,0){2}{\circle*{0.29}}
\multiput(4,3)(1,0){2}{\circle*{0.29}}
\multiput(4,1)(1,0){2}{\line(0,1){2}}
\put(4,1){\line(3,2){3}}
\put(5,1){\line(3,2){3}}
\put(5,1){\line(1,1){2}}
\multiput(7,1)(1,0){2}{\circle*{0.29}}
\multiput(7,3)(1,0){2}{\circle*{0.29}}
\put(7,1){\line(0,1){2}}
\put(8,1){\line(0,1){2}}
\put(7,1){\line(3,2){3}}
\put(8,1){\line(3,2){3}}
\put(8,1){\line(1,1){2}}
\multiput(10,1)(1,0){2}{\circle*{0.29}}
\multiput(10,3)(1,0){2}{\circle*{0.29}}
\put(10,1){\line(0,1){2}}
\put(11,1){\line(0,1){2}}
\multiput(11.9,1)(0.35,0){3}{\circle*{0.1}}
\multiput(11.9,3)(0.35,0){3}{\circle*{0.1}}
\put(0.1,3){\makebox(0,0){$S$}}
\put(1,1){\makebox(0,0){$V-S$}}
\put(0.2,2){\makebox(0,0){$G$}}
\put(0.9,3){\makebox(0,0){$core$}}
\put(3,1){\makebox(0,0){$Z_{0}$}}
\put(4.5,3.1){\makebox(0,0){$Y_{0}$}}
\put(2.5,3.1){\makebox(0,0){$Q$}}
\put(6,1){\makebox(0,0){$Z_{1}$}}
\put(7.5,3.1){\makebox(0,0){$Y_{1}$}}
\put(9,1){\makebox(0,0){$Z_{2}$}}
\put(10.5,3.1){\makebox(0,0){$Y_{2}$}}
\end{picture}\caption{$S\in\Omega(G)$, $Q=\mathrm{core}\left( G\right)
-M\left( Z_{0}\right) $, $Y_{0}=$ $M\left( Z_{0}\right) $, $Y_{1}=\left(
N\left( Z_{0}\right) -M\left( Z_{0}\right) \right) \cap S$, $Y_{2}=...$,
and $Z_{i}=M\left( Y_{i}\right) ,i=1,2,...$ {.}
\end{figure}
The inclusion
\[
Z_{0}\cup\left( \mathrm{core}(G)-M\left( Z_{0}\right) \right)
\subseteq\left( S-M\left( Z^{\infty}\right) \right) \cup Z^{\infty
\]
is justified by the definition of $Z^{\infty}$.
Since $\left\vert M\left( Z^{\infty}\right) \right\vert =\left\vert
Z^{\infty}\right\vert $ we obtain $\left\vert \left( S-M\left( Z^{\infty
}\right) \right) \cup Z^{\infty}\right\vert =\left\vert S\right\vert $.
According to the definition of $Z^{\infty}$ the set
\[
\left( N\left( Z^{\infty}\right) \cap S\right) -M\left( Z^{\infty
}\right)
\]
is empty. In other words, the set $\left( S-M\left( Z^{\infty}\right)
\right) \cup Z^{\infty}$ is independent. Therefore, we arrive at
\[
\left( S-M\left( Z^{\infty}\right) \right) \cup Z^{\infty}\in\Omega\left(
G\right) .
\]
Consequently, $\left( S-M\left( Z^{\infty}\right) \right) \cup Z^{\infty}$
is a desired enlargement of $Z_{0}\cup$ $\left( \mathrm{core}(G)-M\left(
Z_{0}\right) \right) $.
\textit{Claim 5.} $\mathrm{core}(G)\cap\left( \left( S-M\left( Z^{\infty
}\right) \right) \cup Z^{\infty}\right) =\mathrm{core}(G)-M\left(
Z_{0}\right) $.
The only part of $\left( S-M\left( Z^{\infty}\right) \right) \cup
Z^{\infty}$ that interacts with $\mathrm{core}(G)$ is the subset
\[
Z_{0}\cup\left( \mathrm{core}(G)-M\left( Z_{0}\right) \right) .
\]
Hence we obtain
\begin{gather*}
\mathrm{core}(G)\cap\left( \left( S-M\left( Z^{\infty}\right) \right)
\cup Z^{\infty}\right) =\\
=\mathrm{core}(G)\cap\left( Z_{0}\cup\left( \mathrm{core}(G)-M\left(
Z_{0}\right) \right) \right) =\mathrm{core}(G)-M\left( Z_{0}\right) .
\end{gather*}
Since $Z_{0}$ is non-empty, by Claim 5 we arrive at the following
contradiction
\[
\mathrm{core}(G)\nsubseteq\left( S-M\left( Z^{\infty}\right) \right) \cup
Z^{\infty}\in\Omega\left( G\right) .
\]
Finally, we conclude with the fact there is no set $Z\subseteq$ $N\left(
\mathrm{core}(G)\right) ,Z\neq\emptyset$ such that $\left\vert N\left(
Z\right) \cap\mathrm{core}(G)\right\vert =\left\vert Z\right\vert $, which,
by Theorem \ref{th9}, means that $\mathrm{core}(G)$ and $\mathrm{\ker}(G)$ coincide.
\end{proof}
Notice that there are non-bipartite graphs enjoying the equality
$\mathrm{\ker}(G)=\mathrm{core}(G)$; e.g., the graphs from Figure \ref{fig14},
where only $G_{1}$ is a K\"{o}nig-Egerv\'{a}ry graph. \begin{figure}[h]
\setlength{\unitlength}{1cm}\begin{picture}(5,1.2)\thicklines
\multiput(2,0)(1,0){4}{\circle*{0.29}}
\multiput(3,1)(1,0){3}{\circle*{0.29}}
\put(2,0){\line(1,0){3}}
\put(3,0){\line(0,1){1}}
\put(5,0){\line(0,1){1}}
\put(4,1){\line(1,0){1}}
\put(3,0){\line(1,1){1}}
\put(1.7,0){\makebox(0,0){$x$}}
\put(2.7,1){\makebox(0,0){$y$}}
\put(1,0.5){\makebox(0,0){$G_{1}$}}
\multiput(8,0)(1,0){5}{\circle*{0.29}}
\multiput(9,1)(1,0){3}{\circle*{0.29}}
\put(8,0){\line(1,0){4}}
\put(9,0){\line(0,1){1}}
\put(10,0){\line(0,1){1}}
\put(10,1){\line(1,0){1}}
\put(11,1){\line(1,-1){1}}
\put(7.7,0){\makebox(0,0){$a$}}
\put(8.7,1){\makebox(0,0){$b$}}
\put(7,0.5){\makebox(0,0){$G_{2}$}}
\end{picture}\caption{$\mathrm{core}(G_{1})=\ker\left( G_{1}\right)
=\{x,y\}$ and $\mathrm{core}(G_{2})=\ker\left( G_{2}\right) =\{a,b\}$.
\label{fig14
\end{figure}
There is a non-bipartite K\"{o}nig-Egerv\'{a}ry graph $G$, such that
$\mathrm{\ker}(G)\neq\mathrm{core}(G)$. For instance, the graph $G_{1}$ from
Figure \ref{fig222} has $\mathrm{\ker}(G_{1})=\left\{ x,y\right\} $, while
$\mathrm{core}(G_{1})=\left\{ x,y,u,v\right\} $. The graph $G_{2}$ from
Figure \ref{fig222} has $\mathrm{\ker}(G_{2})=\emptyset$, while $\mathrm{core
(G_{2})=\left\{ w\right\} $.
\begin{figure}[h]
\setlength{\unitlength}{1cm}\begin{picture}(5,1.3)\thicklines
\multiput(4,0)(1,0){4}{\circle*{0.29}}
\multiput(3,1)(1,0){5}{\circle*{0.29}}
\put(4,0){\line(1,0){3}}
\put(4,0){\line(0,1){1}}
\put(3,1){\line(1,-1){1}}
\put(5,0){\line(0,1){1}}
\put(5,0){\line(1,1){1}}
\put(5,1){\line(1,-1){1}}
\put(6,0){\line(0,1){1}}
\put(7,0){\line(0,1){1}}
\put(2.7,1){\makebox(0,0){$x$}}
\put(3.7,1){\makebox(0,0){$y$}}
\put(4.7,1){\makebox(0,0){$u$}}
\put(6.3,1){\makebox(0,0){$v$}}
\put(2,0.5){\makebox(0,0){$G_{1}$}}
\multiput(10,0)(1,0){2}{\circle*{0.29}}
\multiput(12,0)(0,1){2}{\circle*{0.29}}
\put(10,0){\line(1,0){2}}
\put(11,0){\line(1,1){1}}
\put(12,0){\line(0,1){1}}
\put(10,0.3){\makebox(0,0){$w$}}
\put(9,0.5){\makebox(0,0){$G_{2}$}}
\end{picture}\caption{Both $G_{1}$ and $G_{2}$\ are K\"{o}nig-Egerv\'{a}ry
graphs. Only $G_{2}$\ has a perfect matching.
\label{fig222
\end{figure}
\section{$\mathrm{\ker}\left( G\right) $ and $\mathrm{diadem}(G)$ in
K\"{o}nig-Egerv\'{a}ry graphs}
There is a non-K\"{o}nig-Egerv\'{a}ry graph $G$ with $V\left( G\right)
=N\left( \mathrm{core}(G)\right) \cup\mathrm{corona}(G)$; e.g., the graph
$G$ from Figure \ref{fig1777}.
\begin{figure}[h]
\setlength{\unitlength}{1cm}\begin{picture}(5,1.2)\thicklines
\multiput(4,0)(1,0){8}{\circle*{0.29}}
\multiput(5,1)(2,0){2}{\circle*{0.29}}
\multiput(8,1)(2,0){2}{\circle*{0.29}}
\put(4,0){\line(1,0){7}}
\put(5,0){\line(0,1){1}}
\put(7,0){\line(0,1){1}}
\put(7,1){\line(1,0){1}}
\put(8,1){\line(1,-1){1}}
\put(10,0){\line(0,1){1}}
\put(10,1){\line(1,-1){1}}
\put(4,0.3){\makebox(0,0){$x$}}
\put(5.3,1){\makebox(0,0){$y$}}
\put(6,0.3){\makebox(0,0){$z$}}
\put(3.2,0.5){\makebox(0,0){$G$}}
\end{picture}
\caption{$G$ is not a K\"{o}nig-Egerv\'{a}ry graph, and $\mathrm{core
(G)=\left\{ x,y,z\right\} $.
\label{fig1777
\end{figure}
\begin{theorem}
\label{th11}If $G$ is a K\"{o}nig-Egerv\'{a}ry graph, then
\emph{(i)}$\ \left\vert \mathrm{corona}(G)\right\vert +\left\vert
\mathrm{core}(G)\right\vert =2\alpha\left( G\right) $;
\emph{(ii) }$\mathrm{diadem}(G)=\mathrm{corona}(G)$, while $\mathrm{diadem
(G)\subseteq\mathrm{corona}(G)$ is true for every graph;
\emph{(iii) }$\left\vert \ker\left( G\right) \right\vert +\left\vert
\mathrm{diadem}\left( G\right) \right\vert \leq2\alpha\left( G\right) $.
\end{theorem}
\begin{proof}
\emph{(i) }Using Theorems \ref{Th5}\emph{(ii) }and \ref{th8}, we infer that
\begin{gather*}
\left\vert \mathrm{corona}(G)\right\vert +\left\vert \mathrm{core
(G)\right\vert =\left\vert \mathrm{corona}(G)\right\vert +\left\vert N\left(
\mathrm{core}(G)\right) \right\vert +\left\vert \mathrm{core}(G)\right\vert
-\left\vert N\left( \mathrm{core}(G)\right) \right\vert =\\
=\left\vert V\left( G\right) \right\vert +d\left( G\right) =\alpha\left(
G\right) +\mu\left( G\right) +d\left( G\right) =2\alpha\left( G\right)
.
\end{gather*}
as claimed.
\emph{(ii)} Every $S\in\Omega\left( G\right) $ is a critical set, by Theorem
\ref{th5}. Hence we deduce that $\mathrm{corona}(G)\subseteq\mathrm{diadem
(G)$. On the other hand, for every graph each critical independent set is
included in a maximum independent set, according to Theorem \ref{th3}. Thus,
we infer that $\mathrm{diadem}(G)\subseteq\mathrm{corona}(G)$. Consequently,
the equality $\mathrm{diadem}(G)=\mathrm{corona}(G)$ holds.
\emph{(iii)} It follows by combining parts \emph{(i),(ii)} and Theorem
\ref{th6}.
\end{proof}
Notice that the graph from Figure \ref{fig1777} has $\left\vert
\mathrm{corona}(G)\right\vert +\left\vert \mathrm{core}(G)\right\vert
=13>12=2\alpha\left( G\right) $.
For a K\"{o}nig-Egerv\'{a}ry graph with $\left\vert \ker\left( G\right)
\right\vert +\left\vert \text{\textrm{diadem}}\left( G\right) \right\vert
<2\alpha\left( G\right) $ see Figure \ref{fig222}. Figure \ref{fig1777}
shows that it is possible for a graph to have $\mathrm{diadem
(G)\varsubsetneqq\mathrm{corona}(G)$ and $\mathrm{\ker}(G)\varsubsetneqq
\mathrm{core}(G)$.
\begin{figure}[h]
\setlength{\unitlength}{1cm}\begin{picture}(5,1.2)\thicklines
\multiput(2,1)(1,0){6}{\circle*{0.29}}
\multiput(4,0)(1,0){4}{\circle*{0.29}}
\put(2,0){\circle*{0.29}}
\put(2,0){\line(1,0){5}}
\put(2,1){\line(2,-1){2}}
\put(3,1){\line(1,-1){1}}
\put(3,1){\line(1,0){1}}
\put(4,0){\line(0,1){1}}
\put(4,0){\line(1,1){1}}
\put(5,1){\line(1,0){2}}
\put(6,0){\line(0,1){1}}
\put(1.2,0.5){\makebox(0,0){$G_{1}$}}
\multiput(9,0)(1,0){5}{\circle*{0.29}}
\multiput(10,1)(1,0){2}{\circle*{0.29}}
\put(13,1){\circle*{0.29}}
\put(9,0){\line(1,0){4}}
\put(10,0){\line(0,1){1}}
\put(11,0){\line(0,1){1}}
\put(11,1){\line(1,-1){1}}
\put(13,0){\line(0,1){1}}
\put(9,0.3){\makebox(0,0){$x$}}
\put(10.3,1){\makebox(0,0){$y$}}
\put(11.3,1){\makebox(0,0){$t$}}
\put(12.7,1){\makebox(0,0){$u$}}
\put(10.7,0.3){\makebox(0,0){$z$}}
\put(12,0.3){\makebox(0,0){$v$}}
\put(12.7,0.3){\makebox(0,0){$w$}}
\put(8.2,0.5){\makebox(0,0){$G_{2}$}}
\end{picture}\caption{$G_{1}$ is a non-bipartite K\"{o}nig-Egerv\'{a}ry graph,
such that $\mathrm{\ker}(G_{1})=\mathrm{core}(G_{1})$ and \textrm{diadem
$\left( G_{1}\right) =\mathrm{corona}(G_{1})$; $G_{2}$ is a
non-K\"{o}nig-Egerv\'{a}ry graph, such that $\mathrm{\ker}(G_{2
)=\mathrm{core}(G_{2})=\{x,y\}$; \textrm{diadem}$\left( G_{2}\right)
\cup\{z,t,v,w\}=\mathrm{corona}(G_{2})$.
\label{fig17888
\end{figure}The combination of $\mathrm{diadem}(G)\varsubsetneqq
\mathrm{corona}(G)$ and $\mathrm{\ker}(G)=\mathrm{core}(G)$ is realized in
Figure \ref{fig17888}.
\begin{proposition}
\label{Cor1} Let $G=\left( A,B,E\right) $ be a bipartite graph.
\emph{(i)} \cite{Ore62} If $X=\ker_{A}\left( G\right) $ \textit{and
$Y$\textit{ is a }$B$-\textit{critical set, then }$X\cap N\left( Y\right)
=N\left( X\right) \cap Y=\emptyset$;
\emph{(ii)} \cite{Ore55} $\ker_{A}\left( G\right) \cap N\left( \ker
_{B}\left( G\right) \right) =N\left( \ker_{A}\left( G\right) \right)
\cap\ker_{B}\left( G\right) =\emptyset$.
\end{proposition}
Now we are ready to describe both $\ker$ and \textrm{diadem} of a bipartite
graph in terms of its bipartition.
\begin{theorem}
Let $G=\left( A,B,E\right) $ be a bipartite graph. Then the following
assertions are true:
\emph{(i) }$\ker_{A}\left( G\right) \cup$ $\ker_{B}\left( G\right)
=\ker\left( G\right) $;
\emph{(ii) }$\left\vert \ker\left( G\right) \right\vert +\left\vert
\mathrm{diadem}\left( G\right) \right\vert =2\alpha\left( G\right) $;
\emph{(iii) }$\left\vert \ker_{A}\left( G\right) \right\vert +\left\vert
\mathrm{diadem}_{B}\left( G\right) \right\vert =\left\vert \ker_{B}\left(
G\right) \right\vert +\left\vert \mathrm{diadem}_{A}\left( G\right)
\right\vert =\alpha\left( G\right) $;
\emph{(iv) }$\mathrm{diadem}_{A}\left( G\right) \cup$ $\mathrm{diadem
_{B}\left( G\right) =$ $\mathrm{diadem}\left( G\right) $.
\end{theorem}
\begin{proof}
\emph{(i)} By Theorem \ref{Th4}\emph{(iii)}, $\ker_{A}\left( G\right) \cup$
$\ker_{B}\left( G\right) $ is critical in $G$. Moreover, the set $\ker
_{A}\left( G\right) \cup$ $\ker_{B}\left( G\right) $ is independent in
accordance with Proposition \ref{Cor1}\emph{(ii)}. Assume that $\ker
_{A}\left( G\right) \cup$ $\ker_{B}\left( G\right) $ is not minimal. Hence
the unique minimal $d$-critical set of $G$, say $Z$, is a proper subset of
$\ker_{A}\left( G\right) \cup$ $\ker_{B}\left( G\right) $, by Theorem
\ref{th4}\emph{(iii)}. According to Theorem \ref{Th4}\emph{(iv)}, $Z_{A}=Z\cap
A$ is an $A$-critical set, which implies $\ker_{A}\left( G\right) \subseteq
Z_{A}$, and similarly, $\ker_{B}\left( G\right) \subseteq Z_{B}$.
Consequently, we get that $\ker_{A}\left( G\right) \cup$ $\ker_{B}\left(
G\right) \subseteq Z$, in contradiction with the fact that $\ker_{A}\left(
G\right) \cup$ $\ker_{B}\left( G\right) \neq Z\subset\ker_{A}\left(
G\right) \cup$ $\ker_{B}\left( G\right) $.
\emph{(ii), (iii), (iv) }By Proposition \ref{Cor1}\emph{(i)}, we have
\[
\left\vert \ker_{A}\left( G\right) \right\vert -\delta_{0}(A)+\left\vert
\text{\textrm{diadem}}_{B}\left( G\right) \right\vert =\left\vert N\left(
\ker_{A}\left( G\right) \right) \right\vert +\left\vert
\text{\textrm{diadem}}_{B}\left( G\right) \right\vert \leq\left\vert
B\right\vert .
\]
Hence, according to Theorem \ref{Th4}\emph{(ii)}, it follows that
\[
\left\vert \ker_{A}\left( G\right) \right\vert +\left\vert
\text{\textrm{diadem}}_{B}\left( G\right) \right\vert \leq\left\vert
B\right\vert +\delta_{0}(A)=\alpha\left( G\right) .
\]
Changing the roles of $A$ and $B$, we obtain
\[
\left\vert \ker_{B}\left( G\right) \right\vert +\left\vert
\text{\textrm{diadem}}_{A}\left( G\right) \right\vert \leq\alpha\left(
G\right) .
\]
By Theorem \ref{Th4}\emph{(iv)}, \textrm{diadem}$\left( G\right) \cap A$ is
$A$-critical and \textrm{diadem}$\left( G\right) \cap B$ is $B$-critical.
Hence \textrm{diadem}$\left( G\right) \cap A\subseteq$ \textrm{diadem
$_{A}\left( G\right) $ and \textrm{diadem}$\left( G\right) \cap
B\subseteq$ \textrm{diadem}$_{B}\left( G\right) $. It implies both the
inclusion $\mathrm{diadem}\left( G\right) \subseteq\mathrm{diadem
_{A}\left( G\right) \cup\mathrm{diadem}_{B}\left( G\right) $, and the
inequality
\[
\left\vert \mathrm{diadem}\left( G\right) \right\vert \leq\left\vert
\mathrm{diadem}_{A}\left( G\right) \right\vert +\left\vert \mathrm{diadem
_{B}\left( G\right) \right\vert .
\]
Combining Theorem \ref{th10}, Theorem \ref{th11}\emph{(i),(ii)}, and part
\emph{(i)} with the above inequalities, we deduc
\begin{gather*}
2\alpha\left( G\right) \geq\left\vert \ker_{A}\left( G\right) \right\vert
+\left\vert \ker_{B}\left( G\right) \right\vert +\left\vert
\text{\textrm{diadem}}_{A}\left( G\right) \right\vert +\left\vert
\text{\textrm{diadem}}_{B}\left( G\right) \right\vert \geq\\
\geq\left\vert \ker\left( G\right) \right\vert +\left\vert
\text{\textrm{diadem}}\left( G\right) \right\vert =\left\vert
\text{\textrm{core}}\left( G\right) \right\vert +\left\vert
\text{\textrm{corona}}\left( G\right) \right\vert =2\alpha\left( G\right)
.
\end{gather*}
Consequently, we infer tha
\begin{gather*}
\left\vert \text{\textrm{diadem}}_{A}\left( G\right) \right\vert +\left\vert
\text{\textrm{diadem}}_{B}\left( G\right) \right\vert =\left\vert
\text{\textrm{diadem}}\left( G\right) \right\vert ,\\
\left\vert \ker\left( G\right) \right\vert +\left\vert \text{\textrm{diadem
}\left( G\right) \right\vert =2\alpha\left( G\right) ,\\
\left\vert \ker_{A}\left( G\right) \right\vert +\left\vert
\text{\textrm{diadem}}_{B}\left( G\right) \right\vert =\left\vert \ker
_{B}\left( G\right) \right\vert +\left\vert \text{\textrm{diadem}
_{A}\left( G\right) \right\vert =\alpha\left( G\right) .
\end{gather*}
Since $\mathrm{diadem}\left( G\right) \subseteq\mathrm{diadem}_{A}\left(
G\right) \cup\mathrm{diadem}_{B}\left( G\right) $ and $\mathrm{diadem
_{A}\left( G\right) \cap\mathrm{diadem}_{B}\left( G\right) =\emptyset$, we
finally obtain that
\[
\mathrm{diadem}_{A}\left( G\right) \cup\mathrm{diadem}_{B}\left( G\right)
=\mathrm{diadem}\left( G\right) ,
\]
as claimed.
\end{proof}
\section{Conclusions}
In this paper we focus on interconnections between $\ker$, \textrm{core,
diadem,} and \textrm{corona}. In \cite{LevManLemma2011} we showed that
$2\alpha\left( G\right) \leq\left\vert \text{\textrm{core}}\left( G\right)
\right\vert +\left\vert \text{\textrm{corona}}\left( G\right) \right\vert $
is true for every graph, while the equality holds whenever $G$ is a
K\"{o}nig-Egerv\'{a}ry graph, by Theorem \ref{th11}\emph{(i)}.
According to Theorem \ref{th6}, $\mathrm{\ker}(G)\subseteq\mathrm{core}(G)$
for every graph. On the other hand, Theorem \ref{th3}\emph{ }implies the
inclusion \textrm{diadem}$\left( G\right) \subseteq\mathrm{corona}(G)$. Henc
\[
\left\vert \ker\left( G\right) \right\vert +\left\vert \text{\textrm{diadem
}\left( G\right) \right\vert \leq\left\vert \text{\textrm{core}}\left(
G\right) \right\vert +\left\vert \text{\textrm{corona}}\left( G\right)
\right\vert
\]
for each graph $G$. These remarks together with Theorem \ref{th11}\emph{(iii)}
motivate the following.
\begin{conjecture}
$\left\vert \ker\left( G\right) \right\vert +\left\vert
\text{\textrm{diadem}}\left( G\right) \right\vert \leq2\alpha\left(
G\right) $ is true for every graph $G$.
\end{conjecture}
When it is proved one can conclude that the following inequalities:
\[
\left\vert \ker\left( G\right) \right\vert +\left\vert \text{\textrm{diadem
}\left( G\right) \right\vert \leq2\alpha\left( G\right) \leq\left\vert
\text{\textrm{core}}\left( G\right) \right\vert +\left\vert
\text{\textrm{corona}}\left( G\right) \right\vert
\]
hold for every graph $G$.
By Corollary \ref{cor2}, $\mathrm{core}(G)$ is critical for every
K\"{o}nig-Egerv\'{a}ry graph. It justifies the following.
\begin{problem}
Characterize graphs such that $\mathrm{core}(G)$ is a critical set.
\end{problem}
Theorem \ref{th10} claims that the sets $\mathrm{\ker}(G)$ and $\mathrm{core
(G)$ coincide for bipartite graphs. On the other hand, there are examples
showing that this equality holds even for some non-K\"{o}nig-Egerv\'{a}ry
graphs (see Figure \ref{fig14}). We propose the following.
\begin{problem}
Characterize graphs with $\ker\left( G\right) =\mathrm{core}(G)$.
\end{problem}
\section{Acknowledgments}
The authors would like to thank the organizers of the International Conference
in Discrete Mathematics (ICDM 2013) for an opportunity to give a special
invited talk including their recent findings.
|
\section{Introduction}
In this work we are interested in solving the following convex composite optimization problem
\begin{equation}
\label{Def_F}
\min_{x \in \mathbb{R}^N} F(x) := f(x) + \Psi(x),
\end{equation}
where $f(x)$ is a smooth convex function and $\Psi(x)$ is a (possibly) nonsmooth, block separable, extended real valued convex function (this will be defined precisely in Section \ref{S_Psi}).
Problems of the form of \eqref{Def_F} arise in many important scientific fields, and applications include machine learning \cite{yuanho}, regression \cite{IEEEhowto:Tibshirani} and compressed sensing \cite{IEEEhowto:CandesRombergTao,Candes06,Donoho06}. Often the term $f(x)$ is a data fidelity term, and the term $\Psi(x)$ represents some kind of regularization.
Frequently, problems of the form of \eqref{Def_F} are large-scale problems, i.e., the size of $N$ is of the order of a million or a billion. Large-scale problems impose restrictions on the types of methods that can be employed for the solution of \eqref{Def_F}. In particular, the methods should have low per iteration computational cost, otherwise completing even a single iteration of the method might require unreasonable time. The methods must also rely only on simple operations such as inner products or matrix vector products, and ideally, they should offer fast progress towards optimality.
First order methods, and in particular randomized coordinate descent methods, have found great success in this area because they can take advantage of the underlying problem structure (separability and block structure), and satisfy the requirements of low computational cost and low storage requirements. For example, in \cite{petermartin} the authors show that their randomized coordinate descent method was able to solve sparse problems with millions of variables in a reasonable amount of time.
Unfortunately, randomized coordinate descent methods have two significant drawbacks. First, due to its coordinate nature, it is efficient only on problems with high degree of separability, and performance suffers when there is a high dependency between variables. Second, as a first-order method, coordinate descent methods do not usually capture essential curvature information of the problem and have been shown to struggle on complicated sparse problems \cite{l1regSCfg}.
The purpose of this work is to overcome these drawbacks by equipping a randomized block coordinate descent method with approximate partial second-order information. In particular, at every iteration of RCD the direction is obtained by solving \textit{approximately} a block piece-wise quadratic model, where the model includes a matrix representing approximate second order information. Then, a line search is employed in order to guarantee a monotonic decrease of the objective function.
RCD randomly selects a block of coordinates at every iteration, which is inexpensive. Although the per iteration computational cost of the method may be higher than other randomized coordinate descent methods, we show that in practise the method is more robust and the total number of iterations decreases. In particular we show that RCD is able to solve difficult problems, on which other coordinate descent method may struggle. RCD uses an inexact search direction, (the termination condition for the block piecewise quadratic subproblem is inspired by \cite{sqa}), coupled with a line search to ensure a monotonic decrease in the function values, and we prove global convergence of RCD and study its local convergence properties.
\subsection{Literature review}
Coordinate descent methods are some of the oldest iterative methods, and they are often better known in the literature under various names such as Jacobi methods, Gauss-Seidel methods, among others. It has been observed that these methods suffer from poor practical performance, particularly on ill-conditioned problems. However, as we enter the era of big data, coordinate descent methods are coming back into favour, because of their ability to provide approximate solutions of some realistic very large/huge scale problems in a reasonable amount of time.
Currently, randomized coordinate descent methods include that of Richt\`{a}rik and Tak\`{a}\v{c} \cite{petermartin}, where the method can be applied to unconstrained convex composite optimization problems of the form \eqref{Def_F}. The algorithm is supported by theoretical convergence guarantees in the form of high probability iteration complexity results, and \cite{petermartin} also reports very impressive practical performance on highly separable large scale problems. The work has also been extended to the parallel case \cite{Richtarik12}, to include acceleration techniques \cite{Fercoq13}, and to include the use of inexact updates \cite{Tappenden13}.
Other important works on randomized coordinate descent methods include methods for huge-scale problems \cite{Nesterov12}, work in \cite{Lu13} that improves the complexity analysis of \cite{Richtarik12}, coordinate descent methods for group lasso \cite{Qin13,Simon12} and general regularizers \cite{ShalevSchwartz13,Wright12} and coordinate descent for constrained optimization problems \cite{Necoara14}.
Unfortunately, on ill-conditioned problems, or problems that are highly nonseparable, first order methods can display very poor practical performance, and this has prompted the study of methods that employ second order information. To this end, recently there has been a flurry of research on Newton-type methods for problems of the general form \eqref{Def_F}, or a special case where $\Psi(x)=\|x\|_1$. For example, Karimi and Vavasis \cite{Karimi14} have developed a proximal quasi-Newton method for $l_1$-regularized least squares problems, Lee, Sun and Saunders \cite{Lee12,Lee13} have proposed a family of Newton-type methods for solving problems of the form \eqref{Def_F} and Scheinberg and Tang \cite{Scheinberg13} present iteration complexity results for a proximal Newton-type method. Moreover, the authors in \cite{sqa} extended standard inexact Newton-type methods to the case of minimization of a composite objective involving a smooth convex term plus an $l_1$-regularizer term.
Finally, there exists parallel deterministic \cite{facchinei14} and sequential active set \cite{desantis14} block coordinate descent methods, where the authors incorporate some block
second-order information in the algorithmic process.
We believe that the works in \cite{sqa,facchinei14} can lead the way to allow general \textit{randomized} coordinate descent to practically incorporate second-order information and in this paper we propose a method which combines the ideas in \cite{sqa,facchinei14,petermartin}.
\subsection{Core ideas and Major Contributions}
In this section we list several of the core ideas and major contributions of this work on randomized block coordinate descent methods. The first two points briefly describe the idea of incorporating (approximate) partial second-order (curvature) information (which
have been also presented in a similar way in \cite{facchinei14}), whilst the last three are contributions of this paper.
\begin{enumerate}
\item \textbf{Incorporation of some second order information.} RCD uses a quadratic model to determine the search direction, which incorporates a user defined positive definite matrix $H^{(i)}(x_k)$. If $H^{(i)}(x_k)$ approximates the Hessian, then second order information is incorporated into the quadratic model, and the search direction obtained by minimizing the model is an approximate Newton-type direction. We stress that $H^{(i)}(x)$ can change at every iteration and this is an advantage over the method in \cite{petermartin} where the matrix is fixed at the start of the algorithm for each block of coordinates.
\item \textbf{Inexact updates.} To ensure that this method is computationally practical, it is imperative that the iterates are inexpensive, and RCD achieves this through the use of \emph{inexact} updates. Any algorithm can be used to approximately minimize the quadratic model. Moreover, the stopping conditions for inner solve are \emph{easy to verify}; they depend upon quantities that are easy/inexpensive to obtain, or may be available as a byproduct of the inner search direction solver.
\item \textbf{Blocks can vary throughout iterations.} If $\Psi(x)$ is completely separable into coordinates then we do not restrict ourselves to a fixed block structure; rather we allow the blocks of coordinates to \emph{change at any iteration}. This is important because every element of the Hessian can be accessed (this is discussed further in Section~\ref{S_HessianApprox}).
\item \textbf{Line search.} The algorithm includes a line search step to ensure a monotonic decrease of the objective function as iterates progress.
The line search is inexpensive to perform because, at each iteration, \textit{it depends on a single block of coordinates only}.
One of the major advantages of incorporating second-order information combined with line search is to allow in practice the selection of \emph{large step sizes} (close to one).
This is because unit step sizes can substantially improve the practical efficiency of a method. We prove that if $f$ is strongly convex, then close to the optimal solution unit step sizes are selected. In fact, for all experiments that we performed,
unit step sizes were accepted by line search for the majority of the iterations.
\item \textbf{Convergence theory.} We provide global convergence results to show that the RCD algorithm is guaranteed to converge in the limit. We also provide local convergence theory for strongly convex functions $f$. In particular, depending on the choice of stopping condition for the inner search direction solve and the matrix $H^{(i)}(x_k)$, we show that close to the optimal solution RCD has on expectation \textit{block} quadratic or superlinear rate of convergence.
\end{enumerate}
\subsection{Format of the paper}
The paper is organised as follows. In Section \ref{Section_Preliminaries} we introduce the notation and definitions that are used throughout this paper, as well as giving several technical results. We also define the quadratic model that is used in the algorithm, prove the equivalence of some stationarity conditions for problem \eqref{Def_F}, and define a continuous measure of the distance of the current point from the set of solutions of \eqref{Def_F}. A thorough description of the RCD algorithm is presented in Section \ref{S_Algorithm}, including how the blocks are selected/sampled at each iteration, a description of the search direction and line search, several suggestions for the matrices $H^{(i)}(x_k)$, and we also present several concrete examples.
The second half of the paper is devoted to providing convergence results and numerical experiments. In Sections \ref{S_GlobalConvergence}, global convergence results are presented, which do not require $f$ to be convex.
Local convergence theory for RCD is presented in Section \ref{S_LocalConvergence}. There we show that, close to optimality line search accepts unit step sizes. Moreover, if both the stopping conditions for the inner search direction solve and the matrix $H^{(i)}(x_k)$ are chosen appropriately, then RCD has on expectation block quadratic or superlinear rate of convergence. Finally, several numerical experiments are presented in Section \ref{S_Numerical}, which show that the algorithm performs very well in practice.
\section{Preliminaries}
\label{Section_Preliminaries}
In this section we introduce the notation and definitions that are used in this paper, and we also present some important technical results. Throughout the paper $\|\cdot\| \equiv \sqrt{\langle \cdot, \cdot \rangle}$
and $\|\cdot\|_A \equiv \sqrt{\langle \cdot, A\cdot \rangle}$, where $A$ is a positive definite matrix. Moreover, $\lambda_{\min}(\cdot)$ and $\lambda_{\max}(\cdot)$ denote the smallest and largest eigenvalue of $\cdot$, respectively.
\subsection{Subgradient and subdifferential}
For a function $\Phi: \mathbb{R}^N \to \mathbb{R} \cup \{+ \infty\}$ the elements $s\in\mathbb{R}^N$ that satisfy
$$
\Phi(y) \geq \Phi(x) + \langle s,y-x \rangle,
$$
are called the subgradients of $\Phi$ at point $x$. In words, all elements defining a linear function that supports the function $\Phi$ at point $x$ are subgradients. The set of all $s$ at a point $x$ is called the subdifferential of $\Phi$ and it is denoted by $\partial \Phi(x)$.
\subsection{Convexity}
\label{S_StrongConvex}
A function $\Phi: \mathbb{R}^N \to \mathbb{R} \cup \{+ \infty\}$ is strongly convex with convexity parameter $\mu_{\Phi} > 0$ if for all $x,y \in \mathbb{R}^N$, and where $s\in\partial \Phi(x)$,
\begin{equation*}
\label{strongly_convex_1}
\Phi(y) \geq \Phi(x) + \langle s,y-x \rangle + \tfrac{\mu_{\Phi}}{2}\|y-x\|^2.
\end{equation*}
If $\mu_\Phi = 0$. then function $\Phi$ is said to be convex.
\subsection{Convex conjugate and proximal mapping}
For a convex function $\Phi: \mathbb{R}^N \to \mathbb{R} \cup \{+ \infty\}$, its convex conjugate is defined as
$ \Phi^*(y) \equiv:= \sup_{u\in\mathbb{R}^N} \langle u,y \rangle - \Phi(u) . $
The proximal mapping of a convex function $\Psi$ at $x$ is
\begin{equation}
\label{Def_prox}
{\rm{prox}}_\Psi (x) := \arg \min_{y\in \mathbb{R}^N} \Psi(y) + \frac12 \|y-x\|^2,
\end{equation}
and the proximal mapping of its convex conjugate $\Psi^*$ is
\begin{equation}
\label{Def_dual_prox}
{\rm{prox}}_{\Psi^*} (x) := \arg \min_{y\in \mathbb{R}^N} \Psi^*(y) + \frac12 \|y-x\|^2.
\end{equation}
The following relation holds between the two proximal mappings.
\begin{lemma}[Chapter 1, ($1.4$) in \cite{Rockafellar06}]
\label{moreau}
Let $\Psi$ be a convex function and let $\Psi^*$ denote its convex conjugate. Then,
$
x = {\rm{prox}}_{\Psi}(x) + {\rm{prox}}_{\Psi^*}(x)
$
for all $x$.
\end{lemma}
From Chapter $1$ of \cite{Rockafellar06}, we also see that ${\rm{prox}}_{\Psi}(\cdot)$ and ${\rm{prox}}_{\Psi^*}(\cdot)$ are nonexpansive
\begin{equation}
\label{Eq_proxnonexpansive}
\|{\rm{prox}}_{\Psi}(y) - {\rm{prox}}_{\Psi}(x)\| \le \|y -x\|, \quad \mbox{and} \quad \|{\rm{prox}}_{\Psi^*}(y) - {\rm{prox}}_{\Psi^*}(x)\| \le \|y -x\|.
\end{equation}
Finally, from Chapter $1$ of \cite{Rockafellar06} we have
\begin{equation}
\label{eq:14}
{\rm{prox}}_{\Psi^*}(x) \in \partial \Psi ({\rm{prox}}_{\Psi}(x)).
\end{equation}
\subsection{Block decomposition of $\mathbb{R}^N$}
\label{S_Block_structure}
Let $U \in \mathbb{R}^{N \times N}$ be a column permutation of the $N \times N$ identity matrix and further let $U = [U_1,U_2,\dots,U_n]$ be a decomposition of $U$ into $n$ submatrices, where $U_i$ is $N \times N_i$ and $\sum_{i=1}^n N_i = N$. It is clear that any vector $x \in \mathbb{R}^N$ can be written uniquely as
$x = \sum_{i=1}^n U_ix^{(i)},$ where $x^{(i)} \in \mathbb{R}^{N_i}$ and block $i$ denotes a subset of $\{ 1,2,\dots,N\}$. Moreover, these vectors are given by
\begin{equation}\label{U_i}x^{(i)} := U_i^Tx.\end{equation}
\subsection{Block decomposition of $\Psi$}\label{S_Psi}
The function $\Psi:\mathbb{R}^N \to \mathbb{R}\cup \{+\infty\}$ is assumed to be block separable. That is, we assume that $\Psi(x)$ can be decomposed as:
\begin{equation}
\label{S2_separable_psi}
\Psi(x) = \sum_{i=1}^n \Psi_i (x^{(i)}),
\end{equation}
where the functions $\Psi_i: \mathbb{R}^{N_i} \to \mathbb{R}\cup \{+\infty\}$ are convex.
Notice that if $n = N$, $\Psi(x)$ is said to be separable (into coordinates), whereas if $n < N$, then $\Psi(x)$ is said to be \emph{block} separable (separable into blocks of coordinates).
The following relationship will be used repeatedly in this work:
\begin{eqnarray}\label{Eq_PsivsPsii}
\notag
\Psi(x+U_it^{(i)}) - \Psi(x) &=& \Big(\sum_{j\neq i}\Psi_j (x^{(j)})+ \Psi_i(x^{(i)}+ t^{(i)})\Big)-\Big(\sum_{j\neq i}\Psi_j (x^{(j)}) + \Psi_i(x^{(i)})\Big)\\
&=& \Psi_i(x^{(i)}+t^{(i)})-\Psi_i(x^{(i)}).
\end{eqnarray}
\subsection{Block Lipschitz continuity of $f$}
Throughout the paper we assume that the gradient of $f$ is block Lipschitz, uniformly in $x$. This means that, for all $x \in \mathbb{R}^N$, $i\subseteq \{ 1,2,\dots,n\}$ and $t^{(i)} \in \mathbb{R}^{N_i}$ we have
\begin{equation}
\label{S2_Lipschitz}
\| \nabla_{i} f(x + U_it^{(i)}) - \nabla_{i} f(x) \| \leq L_{i} \|t^{(i)}\|,
\end{equation}
where $ \nabla_{i} f(x) \overset{\eqref{U_i}}{=} U_i^T\nabla f(x)$. An important consequence of \eqref{S2_Lipschitz} is the following standard inequality \cite[p.57]{Nesterov04}:
\begin{equation}
\label{S2_upperbound}
f(x+ U_it^{(i)}) \leq f(x) + \langle \nabla_{i} f(x), t^{(i)} \rangle+ \tfrac{L_{i}}{2}\|t^{(i)}\|^2.
\end{equation}
\subsection{Piecewise Quadratic Model}
For fixed $x \in\mathbb{R}^N$, we define a piecewise quadratic approximation of $F$ around the point $(x + t)\,\in \mathbb{R}^N$ as follows:
\begin{equation}
\label{Def_Q}
F(x + t)\approx Q(x;t) := f(x) + \sum_{i=1}^n Q_i(x,t^{(i)}),
\end{equation}
where
\begin{equation}\label{Def_Qi}
Q_i(x,t^{(i)}) := \langle \nabla_i f(x), t^{(i)} \rangle + \frac12 \|t^{(i)}\|_{H^{(i)}(x)}^2 + \Psi_i(x^{(i)} + t^{(i)}),
\end{equation}
and $H^{(i)}(x) \in \mathbb{R}^{N_i \times N_i}$ is \emph{any} positive definite matrix, which possibly depends on $x$. Notice that $Q(x;0) = F(x)$ and that $Q_i(x,t^{(i)})$ is the quadratic model for block $i$.
\subsection{Stationarity conditions}
The following theorem gives the equivalence of some stationarity conditions of problem \eqref{Def_F}.
\begin{theorem}
\label{thm:optimality_conditions}
The following are equivalent first order optimality conditions of problem \eqref{Def_F}.
\begin{enumerate}
\item[(i)] $\nabla f(x) + s=0$ and $s\in \partial \Psi(x)$,
\item[(ii)] $-\nabla f(x)\in \partial \Psi(x)$,
\item[(iii)] $\nabla f(x) + \frac{1}{\beta}{\rm{prox}}_{(\beta\Psi)^*}\left(x - \beta \nabla f(x) \right) = 0$,
\item[(iv)] $x = {\rm{prox}}_{\beta\Psi}\left(x - \beta\nabla f(x) \right)$,
\end{enumerate}
where $\beta$ is any positive constant.
\end{theorem}
\begin{proof}
It is easy to see that $(i)$ are first-order optimality conditions of problem \eqref{Def_F}, which can be obtained by using the definition of subgradient.
It is trivial to show that $(i) \Longleftrightarrow (ii)$. By Lemma \ref{moreau}, we have that $(iii) \Longleftrightarrow (iv)$. We now show that $(iii) \Longleftrightarrow (ii)$.
We rewrite $(ii)$ as
\begin{equation*}
0\in \beta \nabla f(x) + y - x + \beta \partial \Psi(x) \quad \mbox{and} \quad y=x,
\end{equation*}
which is satisfied if and only if $(iv)$ holds, hence, if and only if $(iii)$ holds.
\end{proof}
Let us define the continuous function
\begin{equation}\label{defeq:1}
g(x;t) := \nabla f(x) + H(x) t + \frac{1}{\beta}{\rm{prox}}_{(\beta\Psi)^*}\big(x + t - \beta (\nabla f(x) + H(x) t) \big),
\end{equation}
where $\beta$ is a positive constant, which is used in the local convergence analysis (Section \ref{S_LocalConvergence}). By Theorem \ref{thm:optimality_conditions}, the points that satisfy $g(x;0) = \nabla f(x) + \frac1\beta {\rm{prox}}_{(\beta\Psi)^*}\left(x - \beta \nabla f(x)\right)=0$ are stationary points for problem \eqref{Def_F}. Hence, $g(x;0)$ is a continuous measure of the distance from the set of stationary points of problem \eqref{Def_F}.
Furthermore, let us define
\begin{eqnarray}
\label{defeq:2}
g_i(x;t^{(i)}) &:=& \nabla_i f(x) + H^{(i)} (x)t^{(i)} \\
& &+ \frac{1}{\beta}{\rm{prox}}_{(\beta\Psi)^*_i}\big(x^{(i)} + t^{(i)} - \beta(\nabla_i f(x) + H^{(i)}(x)t^{(i)}) \big),\notag
\end{eqnarray}
which will be used as a continuous measure for the distance from stationarity of the block piecewise quadratic function $Q_i(x_k;t^{(i)})$.
\section{The Algorithm}
\label{S_Algorithm}
In this section we present the Robust Coordinate Descent (RCD) algorithm for solving problems of the form \eqref{Def_F}. There are three key steps in the algorithm: (step $4$) the coordinates are sampled randomly; (step $5$) the quadratic model \eqref{Def_Qi} is solved approximately until the stopping conditions \eqref{Def_stoppingconditions} are satisfied to give a search direction; (step $6$) a line search is performed to find a step size that ensures a sufficient reduction in the objective value. Once these key steps have been performed, the current point $x_k$ is updated to give a new point $x_{k+1}$, and the process is repeated.
The following assumption is used in RCD. The reason this assumption is used will be made clear in Section \ref{S_blockstructure}.
\begin{assumption}
\label{Assump_sampling}
The block decomposition of $\mathbb{R}^N$ used within RCD, and the associated probability distribution, adhere to the block structure of $\Psi(x)$.
\end{assumption}
We now present pseudocode for the algorithm, while a thorough description of each of the key steps in the algorithm will follow in the rest of this section.
\begin{algorithm}[H]
\begin{algorithmic}[1]
\vspace{0.1cm}
\STATE \textbf{Input} Choose $x_0\in \mathbb{R}^N$, $\theta \in (0,1/2)$ and $\beta>0$.
\STATE{\textbf{Initialize} a decomposition of $\mathbb{R}^N$ and a probability distribution following Assumption~\ref{Assump_sampling}}\vspace{2mm}
\FOR{$k=1,2,\cdots$} \vspace{2mm}
\STATE Sample a block of coordinates $i$ with probability $p_i >0$.\\ \vspace{2mm}
\STATE If $g_i(x_k;0)=0$ then go to Step $3$; else approximately solve \begin{equation}\label{eq_subproblem}
t_k^{(i)} := \argmin_{t^{(i)}} Q_i(x_k;t^{(i)}),
\end{equation}
until the stopping conditions
\begin{equation}\label{Def_stoppingconditions}
Q(x_k;U_it^{(i)}_k) < Q(x_k;0) \quad \mbox{and} \quad \|g_i(x_k;t^{(i)}_k)\| \le \eta_k^i \|g_i(x_k;0)\|,
\end{equation}
are satisfied, (where $\eta_k^i \in [0,1)$).
\STATE Perform a backtracking line search along the direction $t_k^{(i)}$ starting from $\alpha=1$. That is, find $\alpha\in(0,1]$ such that
\begin{equation}\label{Def_linesearch}
F(x_k) - F(x_k+\alpha U_it_k^{(i)}) \ge \theta \left(\ell(x_k;0) - \ell(x_k;\alpha U_it_k^{(i)})\right),
\end{equation}
where
\begin{equation}\label{Def_lossfunctionli}
\ell(x_k; t) := f(x_k) + \langle \nabla f(x_k), t \rangle + \Psi(x_k + t).
\end{equation}
\STATE Update $x_{k+1} = x_k + \alpha U_i t_k^{(i)}$
\ENDFOR
\end{algorithmic}
\caption{Robust Coordinate Descent (RCD)}
\label{RCD}
\end{algorithm}
\subsection{Block structure and selection of coordinates (Steps $\boldsymbol 2$ \& $\boldsymbol 4$)}
\label{S_blockstructure}
One of the crucial ideas of this algorithm is that the block of coordinates to be updated at each iteration is chosen \emph{randomly}. This allows the coordinates to be selected very quickly. In this section we explain in detail, how the blocks are selected/sampled at each iteration. We also give examples of how coordinates can be randomly sampled such that Assumption \ref{Assump_sampling} is satisfied.
\subsubsection{$\Psi$ is block separable with $n < N$}
When $\Psi$ has a fixed block structure (i.e., $n < N$), the block decomposition of $\mathbb{R}^N$ (via the matrix $U = [U_1,\dots,U_n]$) described in Section \ref{S_Block_structure} is fixed at the start of the algorithm to coincide with the block structure of $\Psi$, and does not change as iterations progress. There are several ways to initialize a sampling scheme to use in RCD that follow Assumption \ref{Assump_sampling}.
\begin{enumerate}
\item Fix the $n$ blocks of coordinates according to the decomposition of $\mathbb{R}^N$ defined by $U$, and in the algorithm, select each block of coordinates with some probability $p_i$. (e.g., uniform probabilities $p_i = 1/n >0$ for all $i = 1,\dots,n$).
\item Perform (single) coordinate descent, where at each iteration of RCD, the coordinate $i$ is selected with some probability $p_i$ (e.g., uniform probabilities $p_i = 1/N$ for all $i$).
\item Perform block coordinate descent, where each block of coordinates has cardinality $N_{\min} := \min\{N_1,\dots,N_n\}$. The restriction is that, at any iteration $k$, the sampled coordinates forming block $i$, must all belong to the same block of $N_j$ coordinates defined by submatrix $U_j$. (i.e., Assumption \ref{Assump_sampling} is satisfied because the decomposition of $U$ is obeyed.) Recall that $\Psi$ is separable into $n$ blocks. Let the total number of subblocks be $l(>n)$, where we assume that each coordinate $1,\dots,N$ appears in at least one of the $l$ blocks. Then each subblock is selected with probability $p_i$.
\end{enumerate}
\subsubsection{$\Psi$ is separable with $n=N$}
When $\Psi$ is separable into coordinates, we have complete control over the indices that are updated at each iteration.
Let $\tau$ denote the block size (number of coordinates that are updated at any iteration $k$), where $1\leq \tau \leq N$. Note that there are $^NC_{\tau}$ subsets\footnote{Here $^NC_{\tau}$ denotes the usual `N choose $\tau$'. i.e., $^NC_{\tau} = N!/(\tau!(N-\tau))!$} of $\tau$ coordinates that can be made from the set $\{1,\dots,N\}$. At any iteration $k$ of RCD, a subset of coordinates $i_k$ with $|i_k| = \tau$ is sampled with some probability $p_i$ (e.g., uniform probabilities $p_{i} = 1/^NC_{\tau} >0$ for all $i$).
Note that in practice, one never explicitly forms the $^NC_{\tau}$ different blocks in order to randomly pick one with some probability $p_i$. Instead, $\tau$ coordinates are sampled randomly without replacement.
\subsection{The search direction and Hessian approximation (Step $\boldsymbol 5$)}
In this section we describe how RCD determines the search direction. In particular, RCD forms a quadratic model for block $i$, and minimizes the model approximately until the stopping conditions \eqref{Def_stoppingconditions} are satisfied, giving an `inexact' search direction.
We also describe the importance of the choice of matrix $H$, which is an approximate second order information term. From now on, we will often use the shorthand $H_k^{(i)} \equiv H^{(i)}(x_k)$.
\subsubsection{The search direction}
At each iteration the update/search direction is found as follows. The subproblem \eqref{eq_subproblem}, (where $Q_i(x_k;t^{(i)})$ is defined in \eqref{Def_Qi}) is approximately solved, and the search direction $t_k^{(i)}$ is accepted when the stopping conditions \eqref{Def_stoppingconditions} are satisfied, for some $\eta_k^i \in [0,1)$. Notice that
\begin{eqnarray}\label{eq:1}
Q(x;U_i t^{(i)}) - Q(x;0) &\overset{\eqref{Def_Q}}{=}& \langle \nabla_i f(x), t^{(i)} \rangle + \frac12 \|t^{(i)}\|_{H^{(i)}}^2 + \Psi(x + U_it^{(i)}) - \Psi(x) \nonumber \\
&\overset{\eqref{Eq_PsivsPsii}}{=}& \langle \nabla_i f(x), t^{(i)} \rangle + \frac12 \|t^{(i)}\|_{H^{(i)}}^2 + \Psi_i(x^{(i)} + t^{(i)}) - \Psi_i(x^{(i)}).
\end{eqnarray}
Hence, from \eqref{eq:1}, the stopping conditions \eqref{Def_stoppingconditions} depend on block $i$ only, and are therefore inexpensive to verify, meaning that they are \emph{implementable}.
\begin{remark}\label{Remark_tineq0}\quad
\begin{itemize}
\item[(i)] At some iteration $k$, it is possible that $g_i(x_k;0)=0$. In this case, it is easy to verify that the optimal solution of subproblem \eqref{eq_subproblem} is $t^{(i)}_k=0$. Therefore, before calculating $t^{(i)}_k$ we check a-priori if condition $g_i(x_k;0)=0$ is satisfied.
\item[(ii)] Notice that, unless at optimality (i.e., $g(x_k;0)=0$), there will always be blocks $i$ such that $g_i(x_k;0)\neq 0$, which implies that $t^{(i)}_k \neq 0$. Hence, RCD will not stagnate.
\item[(iii)] Following similar arguments as those made in \cite[p.4]{sqa}, we prove this in Lemma \ref{lem:Fdec} that both conditions are required to ensure that $t_k^{(i)}$ is a descent direction.
\end{itemize}
\end{remark}
\subsubsection{The Hessian approximation}
\label{S_HessianApprox}
Arguably, them most important feature of this method is that the quadratic model \eqref{Def_Qi} incorporates second order information in the form of a positive definite matrix $H_k^{(i)}$. This is key because, depending upon the choice of $H_k^{(i)}$, it makes the method robust. Moreover, at each iteration, the user has complete freedom over the choice of $H_k^{(i)}\succ0$.
We now provide a few suggestions for the choice of $H_k^{(i)}$. (This list is not intended to be exhaustive.) Notice that in each case there is a trade off between a matrix that is inexpensive to work with, and one that is a more accurate representation of the true block Hessian.
\begin{enumerate}
\item Clearly, the simplest option is to set $H_k^{(i)} = I$ for all $i$ and $k$. In this case \emph{no second order information is employed by the method.}
\item A second option is to let $H_k^{(i)} = \text{diag}(\nabla_i^2 f(x_k))$. In this case $H_k^{(i)}$ and it's inverse are inexpensive to work with. Moreover, if $f$ is quadratic, then $\nabla^2 f(x_k)$ is constant for all $k$, so $H = \text{diag}(\nabla^2 f(x))$ can be computed and stored at the start of the algorithm and elements can be accessed throughout the algorithm as necessary. This is very effective if $\text{diag}(\nabla^2 f(x))$ is a good approximation to $\nabla^2 f(x)$.
\item A third option is to let $H_k^{(i)} = \nabla_i^2 f(x_k)$ (i.e., $H_k^{(i)}$ is a principal minor of the Hessian). In this case, $H_k^{(i)}$ provides the most accurate second order information, but it is (potentially) more computationally expensive to work with.
In practice the matrix $\nabla_i^2 f(x_k)$ is used in a matrix-free way and is not explicitly stored. For example, there may be an analytic formula for performing matrix-vector products with $\nabla_i^2 f(x_k)$, or techniques from automatic differentiation could be employed, see \cite[Section $7$]{IEEEhowto:wrightbook2}.
\item Another option is to use a quasi-Newton type approach where $H_k^{(i)}$ is an approximation to $\nabla_i^2 f(x_k)$ based on the limited-memory BFGS update scheme, see \cite[Section $8$]{IEEEhowto:wrightbook2}.
This approach might be more suitable in cases that the problem is not very ill-conditioned and additionally performing matrix-vector products with $\nabla_i^2 f(x_k)$ is expensive.
\end{enumerate}
\begin{remark}\quad
If any of the matrices above are not positive definite, then they can be altered to make them so. For example, if $H_k^{(i)}$ is diagonal, any zero that appears on the diagonal can be replaced with a positive number. Moreover, if $\nabla_i^2 f(x_k)$ is not positive definite, a multiple of the identity can be added to it.
\end{remark}
An advantage of the RCD algorithm (if Option 3 is used for $H_k^{(i)}$) is that \emph{all elements of the Hessian can be accessed.} This is because the blocks of coordinates can change at every iteration, and so too can matrix $H_k^{(i)}$. This makes RCD extremely \emph{flexible} and is particularly advantageous when there are large off diagonal elements in the Hessian.
\subsection{The line search (Step $\boldsymbol 6$)}
\label{subsec:pract_line}
The stopping conditions \eqref{Def_stoppingconditions} ensure that $t_k^{(i)}$ is a descent direction, but if the full step $x_k + U_it_k^{(i)}$ is taken, a reduction in the function value \eqref{Def_F} is not guaranteed. To this end, we include a line search step in our algorithm in order to guarantee monotonic decrease of function $F$. Essentially, the line search guarantees the sufficient decrease of $F$ at every iteration, where sufficient decrease is measured by the loss function \eqref{Def_lossfunctionli}.
In particular, for fixed $\theta \in (0,1/2)$, we require that for some $\alpha \in (0,1]$, \eqref{Def_linesearch} is satisfied. (In Lemma \ref{lem:Fdec} we prove that there exists a subinterval $(0,\tilde{\alpha}]$ of $(0,1]$ in which \eqref{Def_linesearch} is satisfied.) Notice that
\begin{eqnarray}
\notag
\ell(x;U_it^{(i)})-\ell(x;0) &\overset{\eqref{Def_lossfunctionli}}{=}& \langle \nabla_i f(x),t^{(i)} \rangle + \Psi(x + U_it^{(i)}) - \Psi(x)\\
\label{Eq_differenceloss}
&\overset{\eqref{Eq_PsivsPsii}}{=}& \langle \nabla_i f(x),t^{(i)}\rangle + \Psi_i(x^{(i)} + t^{(i)}) - \Psi_i(x^{(i)}),
\end{eqnarray}
which shows that the calculation of the right hand side of \eqref{Def_linesearch} only depends upon block $i$, so it is inexpensive. Moreover, the line search condition (Step 5) involves the difference between function values $F(x_k) - F(x_k + \alpha U_i t_k^{(i)})$. Fortunately, while function values can be expensive to compute, the difference in the objective value between iterates need not be (this is discussed in more detail in Section \ref{S_Examples}).
\subsection{Examples}
\label{S_Examples}
In this section we provide several examples to demonstrate the practicality of the algorithm. These examples demonstrate that the difference of function values $F(x_k) - F(x_k + \alpha U_i t_k^{(i)})$
required by the line search conditions \eqref{Def_linesearch}, can be easy/inexpensive to implement and verify.
\subsubsection{Quadratic loss plus regularization example}
Suppose that
$f(x) = \tfrac12\|Ax-b\|^2$ and $\Psi(x) \neq 0,$
where $A\in\mathbb{R}^{m\times N}$, $b\in\mathbb{R}^m$ and $x \in \mathbb{R}^N$.
Then
\begin{eqnarray}
\label{Eq_Fvaldiff}
F(x_k) - F(x_k+\alpha U_i t_k^{(i)}) &\overset{\eqref{Eq_PsivsPsii}}{=}& f(x_k) + \Psi_i(x^{(i)}) \\ \notag
&& - f(x_k+\alpha U_i t_k^{(i)}) - \Psi_i(x_k^{(i)}+\alpha t_k^{(i)}) \\ \notag
&=& \Psi_i(x^{(i)})- \alpha \langle \nabla_i f(x), t_k^{(i)} \rangle - \frac{\alpha^2}2\|A_it_k^{(i)}\|_2^2 \\\notag
&& - \Psi_i(x_k^{(i)}+\alpha t_k^{(i)}).
\end{eqnarray}
Notice that calculation of $F(x_k) - F(x_k+\alpha U_i t_k^{(i)})$ as a function of $\alpha$ only depends on block $i$, hence, it is inexpensive.
Moreover, in some cases some of the quantities in \eqref{Eq_Fvaldiff} are already needed in the computation of the search direction $t$, so regarding the line search step, they essentially come ``for free''.
\subsubsection{Logistic regression example}
Suppose that
$$f(x) \equiv \sum_{j=1}^m \log(1 + e^{-b_j a_j^T x}) \quad \mbox{and} \quad \Psi(x) \neq 0,$$
where $a_j^T$ is the $j$th row of a matrix $A\in\mathbb{R}^{m\times n}$
and $b_j$ is the $j$th component of vector $b\in\mathbb{R}^m$. As before, we need to evaluate \eqref{Eq_Fvaldiff}. Let us split calculation of $F(x_k) - F(x_k+\alpha U_i t_k^{(i)}) $ in parts. The first part $\Psi_i(x^{(i)})- \Psi_i(x_k^{(i)}+\alpha t_k^{(i)}) $ is inexpensive, since it depends only on block $i$. The second part $f(x_k) - f(x_k+\alpha U_i t_k^{(i)})$ is more expensive because
is depends upon the logarithm.
In this case, one can calculate $f(x_0)$ \textit{once} at the beginning of the algorithm and then update $f(x_k+\alpha U_i t_k^{(i)})$ $\forall k\ge1$ less expensively.
In particular, let us assume that the following terms:
\begin{equation}
\label{inner_prod_log}
e^{-b_j a_j^Tx_0 } \quad \forall j \quad \mbox{and} \quad f(x_0)=\sum_{j=1}^m \log(1 + e^{-b_j a_j^Tx_0}),
\end{equation}
are calculated once and stored in memory.
Then, at iteration $1$, the calculation of
$f(x_0 + \alpha U_i t^{(i)}_0) = \sum_{j=1}^m \log(1 + e^{-b_j a_j^Tx_0}e^{-\alpha b_j a_j^T(U_i t^{(i)}_0)})$
is required for different values of $\alpha$ by the backtracking line search algorithm.
The most demanding task in calculating $f(x_0 + \alpha U_i t^{(i)}_0)$ is the calculation of the products $b_j a_j^T(U_i t^{(i)}_0)$ $\forall j$ \textit{once}, which is inexpensive since $\forall j$ this operation
depends only on block $i$. Having $b_j a_j^T(U_i t^{(i)}_0)$ $\forall j$ and \eqref{inner_prod_log} calculation of $f(x_0) - f(x_0 + \alpha U_i t^{(i)}_0)$ for different values of $\alpha$ is inexpensive.
At the end of the process, $f(x_1)$ and $e^{-b_j a_j^T x_1}$ $\forall j $ will be given for free, and the same process can be followed for the calculation of $f(x_1) - f(x_1 + \alpha U_i t^{(i)}_1)$ etc.
\section{Global convergence theory without convexity of $f$}
\label{S_GlobalConvergence}
In this section we provide global convergence theory for the RCD algorithm. Note that we \emph{do not assume that $f$ is convex}.
Throughout this section we denote $H^{(i)}_k \equiv H^{(i)}(x_k)$. The following assumptions are made about $H_k^{(i)}$ and $f$.
\begin{assumption}\label{Assump_HisPD}
There exist constants $0 < \lambda_i \leq \Lambda_i$, such that the sequence $\{H_k^{(i)}\}_{k\geq 0}$ satisfies
\begin{equation}\label{Assumption_lambdai}
0 < \lambda_i \leq \lambda_{\min}(H_k^{(i)}) \quad \text{and} \quad \lambda_{\max}(H_k^{(i)}) \leq \Lambda_i, \quad \text{for all } i \text{ and } k.
\end{equation}
\end{assumption}
\begin{assumption}\label{Assump_fisLipschitz}
The function $f$ is smooth, bounded below, and satisfies \eqref{S2_Lipschitz} for all $i$.
\end{assumption}
Assumption \ref{Assump_HisPD} explains that the Hessian approximation $H_k^{(i)}$ must be positive definite for all blocks $i$ at all iterations $k$. Assumption \ref{Assump_fisLipschitz} explains that $f$ must be block Lipschitz for all blocks $i$ and all iterations $k$.
Before proving global convergence of RCD, we present several technical results. The following lemma shows that if $t^{(i)}_k$ is nonzero, then $F$ is decreased.
\begin{lemma}
\label{lem:Fdec}
Let Assumptions \ref{Assump_HisPD} and \ref{Assump_fisLipschitz} hold. Let $\theta\in(0,1/2)$ and let $x_k$ and $i$ be generated by RCD. Then Step 6 of RCD will accept a step-size $\alpha$ that satisfies
\begin{equation}\label{Eq_alphamin}
\alpha \ge \tilde{\alpha} \qquad \text{where} \qquad \tilde{\alpha}:=(1-\theta)\frac{\lambda_i}{2L_i}.
\end{equation}
Furthermore,
\begin{equation}\label{Eq_F_ubont}
F(x_k) - F(x_k + \alpha U_it^{(i)}_k) > \theta(1-\theta)\frac{\lambda_i^2}{4L_i}\|t^{(i)}_k\|^2.
\end{equation}
\end{lemma}
\begin{proof}
The proof closely follows that of \cite[Theorem $3.1$]{sqa}.
From \eqref{Def_stoppingconditions},
\begin{equation}
\label{Eq_yo}
0 > Q(x_k ; U_it_k^{(i)}) - Q(x_k;0) = \ell(x_k;U_it_k^{(i)}) - \ell(x_k;0) + \frac12\|t_k^{(i)}\|_{H_k^{(i)}}^2.
\end{equation}
Rearranging gives
\begin{equation}\label{in:1}
\ell(x_k;0) - \ell(x_k; U_it_k^{(i)}) > \frac12\|t_k^{(i)}\|_{H_k^{(i)}}^2 \overset{\eqref{Assumption_lambdai}}{\geq} \frac12\lambda_i\|t_k^{(i)}\|^2.
\end{equation}
By Assumption \ref{Assump_fisLipschitz}, for $\alpha \in (0,1)$, we have
\begin{equation*}
F(x_k + \alpha U_it_k^{(i)}) \leq f(x_k) + \alpha \langle \nabla_i f(x), t_k^{(i)} \rangle + \frac{L_i}{2}\alpha^2 \|t_k^{(i)}\|^2 + \Psi(x_k+\alpha U_it_k^{(i)}).
\end{equation*}
Adding $\Psi(x_k)$ to both sides of the above and rearranging gives
\begin{eqnarray}
\label{in:3}
\notag
F(x_k) - F(x_k + \alpha U_it_k^{(i)}) &\ge& -\alpha \langle \nabla_i f(x_k), t_k^{(i)} \rangle - \frac{L_i}{2}\alpha^2 \|t_k^{(i)}\|^2 \\ \notag
& & - \Psi(x_k+\alpha U_it_k^{(i)}) + \Psi(x_k) \\
&\overset{\eqref{Eq_differenceloss}}{=}& \ell(x_k;0) - \ell(x_k ; \alpha U_it_k^{(i)}) - \frac{L_i}{2} \alpha^2 \|t_k^{(i)}\|^2.
\end{eqnarray}
By convexity of $\Psi(x)$ we have that
\begin{equation}\label{Eq_ell_convexity}
\ell(x_k;0) - \ell(x_k;\alpha U_it_k^{(i)})\ge \alpha(\ell(x_k;0) - \ell(x_k;U_it_k^{(i)})).
\end{equation}
Then
\begin{eqnarray*}
\label{in:10}
F(x_k) - F(x_k + \alpha U_it_k^{(i)}) &-& \theta (\ell(x_k;0) - \ell(x_k ; \alpha U_it_k^{(i)})) \nonumber \\
&\overset{\eqref{in:3}}{\geq}& (1-\theta) \big(\ell(x_k;0) - \ell(x_k ; \alpha U_it^{(i)})\big) - \frac{L_i}{2}\alpha^2 \|t_k^{(i)}\|^2 \nonumber \\
&\overset{\eqref{Eq_ell_convexity}}{\geq}& \alpha(1-\theta) \big(\ell(x_k;0) - \ell_i(x_k ; U_it_k^{(i)})\big) - \frac{L_i}{2}\alpha^2 \|t_k^{(i)}\|^2 \nonumber \\
& \stackrel{\eqref{in:1}}{>}& \frac12\big(\alpha(1-\theta)\lambda_i \|t_k^{(i)}\|^2 - L_i\alpha^2 \|t_k^{(i)}\|^2\big) \nonumber \\
& =& \frac{\alpha}{2}\big((1-\theta)\lambda_i - L_i\alpha\big) \|t_k^{(i)}\|^2.
\end{eqnarray*}
From the previous observe that if $\alpha$ satisfies $0 \le \alpha \le (1-\theta)\frac{\lambda_i}{L_i}$,
then $\alpha$ also satisfies the backtracking line search step of RCD. Suppose that any $\alpha$ that is rejected by the line search is halved for the next line search trial. Then, it is guaranteed that the $\alpha$ that is accepted satisfies \eqref{Eq_alphamin}.
Since the line search condition \eqref{Def_linesearch} is guaranteed to be satisfied for some step size $\alpha$, from \eqref{Def_linesearch}
and convexity of $\ell$ we obtain
$F(x_k) - F(x_k + \alpha U_it_k^{(i)}) \ge \theta \alpha (\ell(x_k;0) - \ell(x_k ; U_it_k^{(i)}))$. Using \eqref{in:1} and \eqref{Eq_alphamin} in the last inequality we obtain \eqref{Eq_F_ubont}.
\end{proof}
The following lemma bounds the norm of the direction $t_k^{(i)}$ in terms of $g_i(x_k,0)$.
\begin{lemma}\label{lem:Fdec2}
Let Assumptions \ref{Assump_HisPD} and \ref{Assump_fisLipschitz} hold. For $\beta>0$, and $x_k$ and $i$ generated by RCD, we have
\begin{equation}\label{in:8}
\|t_k^{(i)}\| \ge \gamma_i \|g_i(x_k;0)\|, \quad \mbox{where} \quad \gamma_i := \frac{1-\eta_k^i}{\frac{1}{\beta} + 2\Lambda_i},
\end{equation}
where $\eta_k^i$ is defined in \eqref{Def_stoppingconditions}. Moreover,
\begin{equation}\label{Eq_Fubong}
F(x_k) - F(x_k + {\alpha}U_it_k^{(i)}) > \theta(1-\theta)\frac{\lambda_i^2}{4L_i} \gamma_i^2 \|g_i(x_k;0)\|^2.
\end{equation}
\end{lemma}
\begin{proof}
This proof closely follows that of \cite[Theorem $3.1$]{sqa}.
Using the reverse triangular inequality and the fact that ${\rm{prox}}_{\Psi^*_i}(\cdot)$ is nonexpansive, we have that
\begin{eqnarray*}
(1-\eta_k^i)\|g_i(x_k;0)\| & \stackrel{\eqref{Def_stoppingconditions}}{\le}& \|g_i(x_k;0)\| - \|g_i(x_k;t^{(i)}_k)\| \\
& \le& \|g_i(x_k;t_k^{(i)}) - g_i(x_k;0) \| \\
& \stackrel{\eqref{defeq:2}}{=} & \|H_k^{(i)} t_k^{(i)} + \tfrac{1}{\beta}{\rm{prox}}_{(\beta\Psi)^*_i}\big(x_k^{(i)} + t_k^{(i)} - \beta(\nabla_i f(x_k) + H_k^{(i)} t_k^{(i)}) \big) \\
& &- \tfrac{1}{\beta}{\rm{prox}}_{(\beta\Psi)^*_i}\big(x_k^{(i)} - \beta\nabla_i f(x_k) \big)\| \\
& \overset{\eqref{Eq_proxnonexpansive}}{\le} & \|H_k^{(i)} t_k^{(i)}\| + \tfrac{1}{\beta}\| (I- \beta H_k^{(i)})t_k^{(i)} \| \\
& \le & (\tfrac{1}{\beta} + 2\|H_k^{(i)}\| )\|t_k^{(i)}\| \\
& \le & (\tfrac{1}{\beta} + 2\Lambda_i)\|t_k^{(i)}\|.
\end{eqnarray*}
Rearranging gives \eqref{in:8}, and combining \eqref{Eq_F_ubont} and \eqref{in:8} gives \eqref{Eq_Fubong}.
\end{proof}
We now have all the tools to prove global convergence of RCD.
\begin{theorem}
\label{thm:1}
Let Assumptions \ref{Assump_HisPD} and \ref{Assump_fisLipschitz} hold. Then the following hold for RCD:
\begin{equation*}
\lim_{k \to \infty} t^{(i)}_k = 0 \quad \forall i \quad \mbox{and} \quad \lim_{k \to \infty} g(x_k;0) = 0.
\end{equation*}
with probability one.
\end{theorem}
\begin{proof}
By Lemma \ref{lem:Fdec}, for $t^{(i)}_k \neq 0$ we have that $F(x_k) > F(x_{k} + {\alpha}U_it^{(i)}_k) $. Since $F$ is bounded from below and every block $(i)$ has positive probability to be selected, then
for $k\to \infty$ the following holds with probability one:
\begin{equation}
\label{eq:10}
\lim_{k\to \infty} F(x_k) - F(x_k + {\alpha}U_it^{(i)}_k) = 0.
\end{equation}
Using \eqref{Eq_F_ubont} in combination with \eqref{eq:10} we get that for $k\to\infty$ with probability one $ \|t_k^{(i)}\|\to 0$, which proves the first part.
Using \eqref{in:8} and $\|t_k^{(i)}\|\to 0$ for $k\to\infty$ with probability one we also get that $\|g_i(x_k;0)\|\to0$ with probability one for $k\to\infty$
Since $g_i(x_k;0)$ is a block of $g(x_k;0)$, i.e., $ g_i(x_k;0) \overset{\eqref{U_i}}{=} U_i^Tg(x_k;0)$ and since every block $g_i(x_k;0)$ tends to zero as $k\to\infty$ with probability one, then we have that
$g(x_k;0)\to0$ with probability one.
\end{proof}
\section{Local convergence theory}
\label{S_LocalConvergence}
In this section we present local convergence theory for RCD. First we discuss some common assumptions that are needed.
Throughout the section we set $H_k^{(i)} = \nabla_{i}^2 f(x_k)$ $\forall i$, where $\nabla_{i}^2 f(x)$ denotes the principal minor of the Hessian $\nabla^2 f(x)$ with row (equivalently column) indices in the subset $i$.
\begin{assumption}\label{Assump_stronglyconvex}
Function $f$ is strongly convex with strong convexity parameter $\mu_f>0$.
\end{assumption}
By continuity of $f$ we have that $\nabla^2 f(x)$ is symmetric, and by strong convexity of $f$ we have that $\mu_f I \preceq \nabla^2 f(x)$
The next theorem explains that, if the Hessian $H\equiv \nabla^2 f(x)$ is positive definite, then every principal submatrix of it is also positive definite.
\begin{theorem}[Theorem 4.3.15 in \cite{Horn85}]\label{Thm_eigs}
Let $A \in \mathbb{R}^{N \times N}$ be a Hermitian matrix, let $r$ be an integer with $1 \le r \le N$, and let $A_r$ denote any $r \times r$ principal submatrix of $A$ (obtained by deleting $N-r$ rows and the corresponding columns from $A$). For each integer $k$ such that $1\le k\le r$ we have
$\lambda_k(A) \le \lambda_k(A_r) \le \lambda_{k+N-r}(A),$
where $\lambda_k(\cdot)$ denotes the $k$th eigenvalue of matrix $\cdot$, and the eigenvalues are ordered $\lambda_1 \le \dots, \lambda_N$.
\end{theorem}
\begin{corollary}
If $f$ is strongly convex with strong convexity parameter $\mu_f>0$, then
\begin{equation}
\label{eq:11}
\mu_f I \preceq \nabla_i^2 f(x) \qquad \text{for all } i \subseteq \{1,\dots,N\}.
\end{equation}
\end{corollary}
\begin{assumption}\label{Assump_HkiSC}
We assume that the blocks of the Hessian of $f$ are Lipschitz continuous. This means that for all $x \in \mathbb{R}^N$, $i \subseteq \{1,2,\dots,n\}$ and $t^{(i)} \in \mathbb{R}^{N_i}$ we have
\begin{equation}\label{Def_LipschitzHessian}
\| \nabla_{i}^2 f(x + U_it^{(i)}) - \nabla_{i}^2 f(x) \| \leq M_i \|t^{(i)}\|.
\end{equation}
\end{assumption}
The following theorem is used to show that the backtracking line search accepts units step sizes close to optimality for any block $i$.
Similarly to \cite{sqa}, in order to prove the previous statement we have to impose sufficient decrease of the quadratic model \eqref{eq_subproblem}
at every iteration. This means that the inexactness conditions in \eqref{Def_stoppingconditions} are replaced by
\begin{eqnarray}
\label{Def_stoppingconditions_suff}
\xi(\ell(x_k;0)-\ell(x_k;U_it^{(i)}_k)) &\leq& Q(x_k;0) - Q(x_k;U_it^{(i)}_k),\\
\notag \text{and} \;\; \|g_i(x_k;t^{(i)}_k)\| &\leq& \eta_k^i \|g_i(x_k;0)\|,
\end{eqnarray}
where $\xi\in(\theta,1/2)$ and $\eta_k^i \in [0,1)$.
Notice that, substituting the equality in \eqref{Eq_yo} into \eqref{Def_stoppingconditions_suff}, gives
\begin{equation}
\label{eq:103}
\frac{1}{2}\|t^{(i)}_k\|_{H^{(i)}_k}^2 \le (1-\xi)\big(\ell(x_k;0)-\ell(x_k;U_it^{(i)}_k)\big).
\end{equation}
\begin{theorem}
\label{thm:unitstepsize}
Let Assumptions \ref{Assump_stronglyconvex} and \ref{Assump_HkiSC} hold. Let $x_k$ and $i$ be generated by RCD.
Moreover, let subproblem \eqref{eq_subproblem} of RCD be solved inexactly until the inexactness conditions \eqref{Def_stoppingconditions_suff}
are satisfied.
If $\|t^{(i)}_k\| \le 3{\lambda_i}(\xi-{\theta})/{M_i}$, where $\theta \in (0,1/2)$, $\xi\in(\theta,1/2)$
and $\lambda_i \equiv \lambda_{\min}(H_k^{(i)}) \geq \mu_f>0$ $\forall i$ (by Assumption \ref{Assump_stronglyconvex}), then the backtracking line search step in RCD accepts step sizes
$\alpha = 1$.
\end{theorem}
\begin{proof}The proof closely follows that of \cite[Lemma $3.3$]{Lee13}.
Using Lipschitz continuity of $H_k^{(i)} $ we have that
$$
f(x_k + U_it^{(i)}_k) \le f(x_k) + \langle \nabla_i f(x_k), t^{(i)}_k \rangle + \tfrac12 \|t_k^{(i)}\|_{H_k^{(i)}}^2 + \tfrac{M_i}{6}\|t^{(i)}_k\|^3.
$$
Adding $\Psi(x_k+U_i t^{(i)}_k)$ to both sides gives
\begin{eqnarray}
\notag
F(x_k + U_it^{(i)}_k) &\le& f(x_k) + \langle \nabla_i f(x_k), t^{(i)}_k \rangle + \tfrac12 \|t_k^{(i)}\|_{H_k^{(i)}}^2 + \tfrac{M_i}{6}\|t^{(i)}_k\|^3 + \Psi(x_k + U_i t^{(i)}_k)\\
\notag
&=& F(x_k) + \langle \nabla_i f(x_k), t^{(i)}_k \rangle + \tfrac12 \|t_k^{(i)}\|_{H_k^{(i)}}^2 + \tfrac{M_i}{6}\|t^{(i)}_k\|^3 \\
\notag
& & + \Psi(x_k + U_i t^{(i)}_k) -\Psi(x_k)\\
\notag
&=& F(x_k) - (\ell(x_k;0)-\ell(x_k;U_it^{(i)}_k)) + \tfrac12 \|t_k^{(i)}\|_{H_k^{(i)}}^2 + \tfrac{M_i}{6}\|t^{(i)}_k\|^3\\
\label{eq:104}
&\overset{\eqref{eq:103}}{\le}& F(x_k) - \xi (\ell(x_k;0)-\ell(x_k;U_it^{(i)}_k)) + \tfrac{M_i}{6}\|t^{(i)}_k\|^3.
\end{eqnarray}
Clearly, $\lambda_i \equiv \lambda_{\min}(H_k^{(i)}) \leq \|t_k^{(i)}\|_{H_k^{(i)}}^2/\|t_k^{(i)}\|^2$. Using this with \eqref{eq:103} in \eqref{eq:104} we get
\begin{eqnarray*}
F(x_k + U_i t^{(i)}_k) &\le& F(x_k) - \xi (\ell(x_k;0)-\ell(x_k;U_it^{(i)}_k)) \\
& &+ \tfrac{M_i}{3 \lambda_i}\|t^{(i)}_k\|(\ell(x_k;0)-\ell(x_k;U_it^{(i)}_k)) \\
&=&F(x_k) - (\xi - \tfrac{M_i}{3 \lambda_i}\|t^{(i)}_k\|)(\ell(x_k;0)-\ell(x_k;U_it^{(i)}_k)).
\end{eqnarray*}
If $\|t^{(i)}_k\| \le 3{\lambda_i}(\xi-{\theta})/{M_i}$, $\theta \in (0,\frac12)$ and $\xi\in(\theta,1/2)$ then
$
F(x_k) - F(x_k + U_i t^{(i)}_k) \ge \theta (\ell(x_k;0)-\ell(x_k;U_it^{(i)}_k)),
$
which implies that RCD accepts a step $\alpha = 1$.
\end{proof}
\begin{corollary}
\label{corollary:unit_step}
By Theorem \ref{thm:1}, $\|t^{(i)}_k\| \to 0$ $\forall i$ as $k\to \infty$.
Thus, there will be a region close to the optimal solution $x_*$ such that $\|t^{(i)}_{k}\| \le 3 {\lambda_i}(\xi-\theta)/{M_i}$ for all $i$ and for all $k$. Hence, by Theorem \ref{thm:unitstepsize}, in this region, the backtracking line search step in RCD accepts unit step sizes for any $i$.
\end{corollary}
The following assumption is mild since it is guaranteed to be satisfied by Corollary \ref{corollary:unit_step}.
\begin{assumption}\label{Assump_unitstep}
Iteration $x_k$ is close to the optimal solution $x_*$ of \eqref{Def_F}
such that unit step sizes are accepted by the backtracking line search algorithm of RCD.
\end{assumption}
The next lemma is a technical result that will be used in Theorem \ref{thm:quadratic}.
\begin{lemma}\label{strong_monotonicity}
Let Assumptions \ref{Assump_fisLipschitz} and \ref{Assump_stronglyconvex} hold. Let $L_{\max} = \max_i\{L_1,\dots,L_n\}$. Let $\beta < 1/L_{\max}$ in \eqref{defeq:2}. Then $g_i(x;t^{(i)})$ inherits strong monotonicity of $\nabla_i f(x)$ $\forall i$:
$$
(u - v)^T (g_i(x;u) - g_i(x;v)) \ge \frac{\mu_f}{2}\|u-v\|^2 \quad \forall u,v \in \mathbb{R}^{N_i}.
$$
\end{lemma}
\begin{proof}
The proof is the same as \cite[Lemma $3.9$]{Lee13}, but restricted to the $i$th block, so is omitted.
\end{proof}
In the following theorem we demonstrate that RCD has on expectation \textit{block} quadratic or superlinear local rate of convergence.
\begin{theorem}
\label{thm:quadratic}
Let Assumptions \ref{Assump_fisLipschitz}, \ref{Assump_stronglyconvex}, \ref{Assump_HkiSC} and \ref{Assump_unitstep} hold.
Let $x_{k+1} = x_k + t_k^{(i)}$, $\eta_k^i = \min\{1/2, \|g_i(x_k;0)\|\},$ and $\beta < 1/L_{\max}$ in \eqref{defeq:2}. Then, $\|g_{i}(x_k;0)\|$ has a quadratic rate of convergence in expectation:
$$
\lim_{k\to \infty} \mathbf{E}\left[\frac{\|g_i(x_{k+1};0)\|}{\|g_i(x_{k};0)\|^2}\;|\;x_k\right] = \mbox{c},
$$
where $c$ is a positive constant.
If $\eta_k^i \to 0$ for $k\to \infty$, then $\|g_{i}(x_k;0)\|$ has a superlinear rate of convergence in expectation:
$$
\lim_{k\to \infty} \mathbf{E}\left[\frac{\|g_i(x_{k+1};0)\|}{\|g_i(x_{k};0)\|}\;|\;x_k\right] = 0.
$$
\end{theorem}
\begin{proof}
For a given $x_k$ we define
\begin{equation}
x(\delta) := x_k + \delta U_i t^{(i)}_k, \quad \text{and} \quad x(\sigma) := x_k + \sigma U_i t^{(i)}_k.
\end{equation}
Using the Fundamental Theorem of Calculus (F.T.o.C.), we have
\begin{eqnarray}\label{Eq_FToC}
\notag
g_i(x(\delta);0) &=& \nabla_i f(x_k) + \delta \nabla_i^2 f(x_k) t^{(i)}_k + \int_0^\delta \int_0^u \nabla_i^3 f(x(\sigma))[t^{(i)}_k,t^{(i)}_k] d \sigma d u \\
&& + \frac{1}{\beta}{\rm{prox}}_{(\beta\Psi)^*_i}(x^{(i)}(\delta) - \beta\nabla_i f(x(\delta)) ).
\end{eqnarray}
Also, from the definition of a derivative we have
\begin{eqnarray}
\label{eqn_intupperbound}
\notag
\int_0^\delta \int_0^u \|\nabla_i^3 f(&x(\sigma)&)[t^{(i)}_k,t^{(i)}_k] \| d \sigma d u\\
\notag
&=& \int_0^\delta \int_0^u \lim_{\sigma \to 0}\|\frac{1}{\sigma}(t^{(i)}_k)^T(\nabla_i^2 f(x(\sigma))-\nabla_i^2 f(x_k)) \| d \sigma d u\\
&\le& \notag \|t^{(i)}_k\|\int_0^\delta \int_0^u \lim_{\sigma \to 0}\|\frac{1}{\sigma}(\nabla_i^2 f(x(\sigma))-\nabla_i^2 f(x_k)) \| d \sigma d u \\ &\overset{\eqref{Def_LipschitzHessian}}{\leq}&
M_i \|t^{(i)}_k\|^2 \int_0^\delta \int_0^u 1\; d \sigma d u = \frac{\delta^2}{2} M_i \|t^{(i)}_k\|^2.
\end{eqnarray}
Now, adding and subtracting $\tfrac1\beta {\rm{prox}}_{(\beta\Psi)^*_i}(x_k^{(i)} + t^{(i)}_k - \beta(\nabla_i f(x_k) + H_k^{(i)} t^{(i)}_k) )$ from \eqref{Eq_FToC}, followed by taking norms, applying the triangle inequality, and using \eqref{eqn_intupperbound}, gives
\begin{eqnarray}\label{Eq_yo2}
\notag
\|g_i(x(\delta);0)\|&\le & \|\nabla_i f(x_k) + \delta \nabla_i^2 f(x_k) t^{(i)}_k \\
\notag
&& +\frac{1}{\beta} {\rm{prox}}_{(\beta\Psi)^*_i}(x_k^{(i)} + t^{(i)}_k - \beta(\nabla_i f(x_k) + H_k^{(i)} t^{(i)}_k) )\|\\
\notag
&& + \frac{1}{\beta}\|{\rm{prox}}_{(\beta\Psi)^*_i}(x^{(i)}(\delta) - \beta\nabla_i f(x(\delta)) ) \\
&& -{\rm{prox}}_{(\beta\Psi)^*_i}(x_k^{(i)} + t^{(i)}_k - \beta(\nabla_i f(x_k) + H_k^{(i)} t^{(i)}_k) )\| +\frac{\delta^2}{2} M_i \|t^{(i)}_k\|^2.
\end{eqnarray}
By Assumption \ref{Assump_unitstep}, RCD accepts unit step sizes. Hence, setting $\delta = 1$ in \eqref{Eq_yo2} gives
\begin{eqnarray*}
\|g_i(x_{k+1};0)\| &\le& \|g_i(x_k;t^{(i)}_k)\| + \frac{1}{2}M_i \|t^{(i)}_k\|^2 \\
&& +\frac{1}{\beta}\|{\rm{prox}}_{(\beta\Psi)^*_i}(x_{k+1}^{(i)} - \beta\nabla_i f(x_{k+1})) \\
&& -{\rm{prox}}_{(\beta\Psi)^*_i}(x_{k+1}^{(i)} - \beta(\nabla_i f(x_k) + H_k^{(i)} t^{(i)}_k) )\|\\
&\le& \|g_i(x_k;t^{(i)}_k)\| + \|\nabla_i f(x_k) + H_k^{(i)} t^{(i)}_k - \nabla_i f(x_{k+1})\| + \frac12 M_i \|t^{(i)}_k\|^2.
\end{eqnarray*}
Using the same trick as before with the F.T.o.C. we have the bound\\
$\|\nabla_i f(x_k) + H_k^{(i)} t^{(i)}_k - \nabla_i f(x_{k+1})\| \le \frac12 M_i \|t^{(i)}_k\|^2,$
so that
\begin{eqnarray}
\notag
\|g_i(x_{k+1};0)\| &\le& \|g_i(x_k;t^{(i)}_k)\| + M_i \|t^{(i)}_k\|^2\\
\label{eq:3}
&\overset{\eqref{Def_stoppingconditions_suff}}{\le}& \eta_k^i \|g_i(x_k;0)\| + M_i \|t^{(i)}_k\|^2.
%
\end{eqnarray}
Setting $u=t^{(i)}_k$ and $v=0$ in Lemma \ref{strong_monotonicity} gives
$ (g_i(x_k;t^{(i)}_k) - g_i(x_k;0))^Tt^{(i)}_k \ge \frac{\mu_f}{2}\|t^{(i)}_k\|^2$. Then, using Cauchy-Schwarz we have
$$
\|g_i(x_k;t^{(i)}_k) - g_i(x_k;0)\| \ge \frac{\mu_f}{2}\|t^{(i)}_k\|.
$$
By the triangular inequality and stopping conditions \eqref{Def_stoppingconditions} we have
\begin{equation}
\label{eq:102}
(1+\eta^i_k) \|g_i(x_k;0)\| \ge \frac{\mu_f}{2}\|t^{(i)}_k\|.
\end{equation}
Replacing \eqref{eq:102} in \eqref{eq:3} we have
\begin{equation}
\label{eq:105}
\|g_i(x_{k+1};0)\| \le \eta_k^i \|g_i(x_k;0)\| + \frac{4 M_i (1 + \eta^i_k)^2}{\mu_f^2} \|g_i(x_k;0)\|^2.
\end{equation}
Moreover, by setting $\eta_k^i = \min\{1/2, \|g_i(x_k;0)\|\}$ we obtain
\begin{equation}
\label{eq:4}
\|g_i(x_{k+1};0)\| \le \left(1 + \frac{9 M_i}{\mu_f^2} \right)\|g_i(x_k;0)\|^2.
\end{equation}
Rearranging and taking expectation gives
\begin{equation}
\label{in:15}
\lim_{k\to \infty} \mathbf{E}\left[\frac{\|g_i(x_{k+1};0)\|}{\|g_i(x_{k};0)\|^2}\;|\;x_k\right] \le \lim_{k\to \infty} \mathbf{E}\left[1 + \frac{9 M_i}{\mu_F^2}\;|\;x_k\right].
\end{equation}
The right hand side of \eqref{in:15} is constant for all $i$, which implies quadratic convergence of $\|g_i(x_k;0)\|$ in expectation.
Moreover, if $\eta_k^i \to 0$ as $k\to \infty $, then from \eqref{eq:105}, Theorem \ref{thm:1} and Lemma \ref{lem:Fdec2},
$$
\lim_{k\to \infty} \mathbf{E}\left[\frac{\|g_i(x_{k+1};0)\|}{\|g_i(x_{k};0)\|}\;|\;x_k\right] = 0,
$$
which implies superlinear convergence of $\|g_i(x_k;0)\|$ in expectation.
\end{proof}
\section{Numerical Experiments}
\label{S_Numerical}
In this section we examine the performance of two versions of RCD and two versions of a Uniform Coordinate Descent method (UCDC) \cite{petermartin}
on two common optimization problems. The first problem is an $\ell_1$-regularized least squares problem of the form \eqref{Def_F} with
\begin{equation}
\label{least_squares}
f(x) =\tfrac12\|Ax-b\|^2 \quad \mbox{and} \quad \Psi(x) = c \|x\|_1,
\end{equation}
where $c>0$, $x\in\mathbb{R}^N$, $A\in\mathbb{R}^{m\times N }$ and $b\in\mathbb{R}^m$. The second problem is an $\ell_1$-regularized logistic regression problems of the form \eqref{Def_F} with
\begin{equation}
\label{logistic_regression}
f(x) = \sum_{j=1}^m\log(1+e^{-b_jx^T a_j}) \quad \mbox{and} \quad \Psi(x) =c \|x\|_1,
\end{equation}
where $c>0$, $a_j\in\mathbb{R}^N$ $\forall j=1,2,\ldots,m$ are the training samples and $b_j\in\{-1,+1\}$ are the corresponding labels.
For \eqref{least_squares} a synthetic sparse large scale experiment is performed and for \eqref{logistic_regression} we compare the methods on two real world large scale problems from machine learning. Notice that for both \eqref{least_squares} and \eqref{logistic_regression}, $\Psi(x) = c\|x\|_1$, which is fully separable into coordinates. This means that, for RCD, we have complete control over the block decomposition, and the indices making up each block can change at every iteration.
All algorithms are coded in MATLAB, and for fairness, MATLAB is limited to a single computational thread for each test run.
All experiments are performed on a Dell PowerEdge R920 running Redhat Enterprise Linux with four Intel Xeon E7-4830 v2 2.2GHz, 20M Cache, 7.2 GT/s QPI, Turbo (4x10Cores).
\subsection{Implementations of RCD and UCDC}
In this section we discuss some details of the implementations of methods RCD and UCDC.
\subsubsection{RCD}
For the RCD method, we fix the size of blocks $\tau>1$ (to be given in the numerical experiments subsections), and at every iteration of RCD, $\tau$ coordinates are sampled uniformly at random without replacement.
We implement two versions of RCD, which we denote by RCD v.1 and RCD v.2. The two versions only differ in how matrix $H_k^{(i)}$ is chosen. In particular, for RCD v.1 we set $H_k^{(i)}:= \text{diag}(\nabla_i^2 f(x_k))$ for all $i$ and $k$. In this case subproblem \eqref{eq_subproblem} is separable
and it has a closed form solution
$$t^{(i)}_k = \mathcal{S}(x^{(i)}_k - (H_k^{(i)})^{-1}\nabla_i f(x^{(i)}_k), c\, \text{diag}((H_k^{(i)})^{-1})),$$
where
\begin{equation}
\label{soft_thresholding}
\mathcal{S}(u,v) = \mbox{sign}(u) \max(|u| - v,0)
\end{equation}
is the well-known soft-thresholding operator which is applied component wise when $u$ and $v$ are vectors.
Notice that since the subproblem is solved exactly there is no need to verify the stopping conditions \eqref{Def_stoppingconditions}.
For RCD v.2, we set
\begin{equation}
\label{reg_H_k}
H_k^{(i)} := \nabla_i^2 f(x_k) + \rho I_{N_i}, \text{ for all } i \text{ and } k,
\end{equation}
where $\rho>0$ guarantees that $H_k^{(i)}$ is positive definite for all $i,k$. Hence, the subproblem \eqref{eq_subproblem} is well defined. The larger $\rho$ is the smaller the condition number of matrix $H_k$ becomes, hence, the faster subproblem \eqref{eq_subproblem} will be solved by an iterative solver. However, we do not want $\rho$ to dominate matrix $H_k$ because the essential second order information from $\nabla^2 f(x_k)$ will be lost.
In this setting of matrix $H_k^{(i)}$ we solve subproblems \eqref{eq_subproblem} iteratively using an Orthant Wise Limited-memory Quasi-Newton (OWL) method,
which can be downloaded from \url{http://www.di.ens.fr/~mschmidt/Software/L1General.html}.
We chose OWL because it has been shown in \cite{sqa} to result in a robust and efficient deterministic version of RCD, i.e. $\tau=N$ (one block of size $N$). Note that we never explicitly form matrix $H_k^{(i)}$, we only perform matrix-vector products with it in a matrix-free manner.
\subsubsection{UCDC}
We also implement two versions of a uniform coordinate descent method as it is described in Algorithm $2$ in \cite{petermartin}. For both versions the size $\tau$ of the blocks and the decomposition of $\mathbb{R}^N$ into $\ceil{N/\tau}$ blocks are \emph{fixed a-priori} and all blocks are selected by UCDC with uniform probability. We compare two versions of UCDC, denoted by UCDC v.1 and UCDC v.2 respectively. For UCDC v.1 we set $\tau = 1$ and for UCDC v.2 we set $\tau>1$ (the exact $\tau$ is given later).
One of the key ingredients of UCDC are the block Lipschitz constants, which are explicitly required in the algorithm. For single coordinate blocks, the Lipschitz constants can be computed with relative ease. However, for blocks of size greater than 1, the block Lipschitz constants can be far more expensive to compute. (For example, for problem \eqref{least_squares}, the block Lipschitz constants correspond to the maximum eigenvalue of $A_i^TA_i$, where $A_i := AU_i$.) For this reason, we do not compute the actual block Lipschitz constants, rather, we use an overapproximation.
To this end, let $L_j>0$ $\forall j=1,2,\cdots,N$ denote the coordinate Lipschitz constants of function $f$. Then the direction $t^{(i)}$ at every iteration is obtained by solving exactly subproblem \eqref{eq_subproblem} with
\begin{equation}
\label{H_k_ucdc}
H_k^{(i)}:= \Big(\sum_{j\in i}L_j\Big) I_\tau,
\end{equation}
using operator \eqref{soft_thresholding}. Notice that for problem \eqref{least_squares}, \eqref{H_k_ucdc} is equivalent to $H_k^{(i)} = \text{trace}(A_i^TA_i) I_{\tau}$, where $\text{trace}(A_i^TA_i)$ is an overapproximation of the maximum eigenvalue of $A_i^TA_i$.
Moreover, notice that Algorithm $2$ in \cite{petermartin} is a special case of RCD where the subproblem \eqref{eq_subproblem} is solved exactly
and line search is unnecessary. This is because, by setting $H_k^{(i)}$ as in \eqref{H_k_ucdc}, then subproblem \eqref{eq_subproblem} is an over estimator of function $F$ along block coordinate direction $t^{(i)}$ (for details we refer the reader to \cite{petermartin}).
\subsection{Termination Criteria and Parameter Tuning}
The only termination criteria that RCD and UCDC should have are maximum number of iterations or maximum running time. This is because using subgradients as a measure of distance from optimality or any other operation of similar cost are considered as expensive tasks for large scale problems. In our experiments RCD and UCDC are terminated when their running time exceeds the maximum allowed running time.
Furthermore, for RCD we set parameter $\eta^{(i)}_k$ in \eqref{Def_stoppingconditions} equal to $0.9$ $\forall i,k$ and
$\rho = 10^{-6}$ in \eqref{reg_H_k}.
The maximum number of backtracking line search iterations is set to 10 and $\theta=10^{-3}$.
For UCDC the coordinate Lipschitz constants $L_j$ $\forall j$ are calculated once at the beginning of the algorithm and this task is included in the overall running time.
Finally, all methods are initialized with the zero solution.
\subsection{$\ell_1$-Regularized Least Squares}
In this subsection we present the performance of RCD and UCDC on the $\ell_1$-regularized least squares problem \eqref{least_squares}.
For this problem the data $A$ and $b$ were synthetically constructed using a generator proposed in \cite[Section $6$]{nesterovgen}, and we set $c=1$.
The advantage of this generator is that it produces data $A$ and $b$
with a known minimizer $x_*$. We slightly modified the generator so that we could control the density of $A$.
The dimensions of the problem are $N=2^{21}$ and $m=N/4$ and the generated matrix $A$ is full rank (with at least one non zero component per column) and a density of $\approx10^{-4}mN$. The optimal solution is set to have $\ceil{0.01N}$ non zero components with values uniformly at random in the interval $[-1,1]$. For UCDC, the coordinate Lipschitz constants are
$
L_j := \|A_j\|_2^2 \quad j=1,2,\cdots,N,
$
and for RCD v.1, RCD v.2 and UCDC v.2, we set $\tau = \ceil{0.01 N}$.
The result of this experiment is shown in Figure \ref{fig1}.
\begin{figure}[h!]%
\centering
\subfloat[Objective function $F(x)$ against iterations]{%
\label{fig1a}%
\includegraphics[scale=0.32]{fig1a.pdf}}
\subfloat[Objective function $F(x)$ against time]{%
\label{fig1b}%
\includegraphics[scale=0.32]{fig1b.pdf}}\\
\caption{Performance of all four methods RCD v.1 and v.2 and UCDC v.1 and v.2 on a sparse large scale $\ell_1$-regularized least squares problem.
For practical purposes,
for UCDC v.1 results are printed every ten thousand iterations. \textit{Calculation of $F(x)$ is not included in running time of the methods.}
\textbf{Fig.1a} shows how the objective function $F(x)$ decreases as a function of the number of iterations. \textbf{Fig.1b} shows
how the objective function $F(x)$ decreases as a function of wall-clock time measured in seconds.}
\label{fig1}%
\end{figure}
In this figure notice that all methods were terminated after $10^4$ seconds. For practical purposes, for UCDC v.1 results are shown
every $10^4$ iterations. For all other methods results are shown after the first iteration takes place and then at every iteration. Observe in sub Figure \ref{fig1a} that block
methods RCD v.1, RCD v.2 and UCDC v.2 performed fewer iterations compared to the single coordinate UCDC v.1.
This is due to much larger per iteration computational complexity of the former methods compared to the latter.
RCD v.2 despite its larger per iteration computational complexity among all methods it was the only one that solved
the problem to higher accuracy within the required maximum time. Moreover, observe in sub Figure \ref{fig1b} that for purely practical purposes it might be better to have a combination of methods
RCD v.1 and v.2. The former could be used at the beginning of the process while the latter could be used at later stages in order to guarantee robustness and speed
closer to the optimal solution. \emph{Finally, it is important to mention that on this problem for both RCD versions unit step sizes $\alpha$ were accepted by the backtracking line search for a major part of the process. Hence, backtracking line search was inexpensive.}
\subsection{$\ell_1$-Regularized Logistic Regression}
In this section we present the performance of RCD and UCDC on the $\ell_1$-regularized logistic regression problems \eqref{logistic_regression}.
Such problems are important in machine learning and are used for training a linear classifier $x\in\mathbb{R}^N$ that separates input data into two distinct clusters, for example, see \cite{yuanho} for further details.
We present the performance of the methods on two sparse large scale data sets.
Problem details are given in Table \ref{LRprobs}, where $A\in\mathbb{R}^{m\times N}$ is a matrix whose rows are training samples.
\begin{table}[h!]
\centering
\caption{Properties of two $\ell_1$-regularized logistic regression problems. The second and third columns show the number of training samples and features, respectively.
The fourth column shows the sparsity of matrix $A$.}
\begin{tabular}{lccc}
\textbf{Problem}& \textbf{$\boldsymbol m$} & \textbf{$\boldsymbol N$} & \textbf{$\mathbf{ nnz(A)/(mN)}$} \\
webspam & $350,000$ & $16,609,143$ & $2.24e$-$4$ \\
kdd2010 (algebra) & $8,407,752$ & $20,216,830$ & $1.79e$-$6$ \\
\end{tabular}
\label{LRprobs}
\end{table}
The data sets can be downloaded from the collection of LSVM problems in \url{http://www.csie.ntu.edu.tw/~cjlin/libsvmtools/datasets/}.
For both experiments we set $c=10$, which resulted in more than $99\%$ classification accuracy of the used data sets.
By \cite[Table~10]{petermartin},
the coordinate Lipschitz constants for UCDC are
$$
L_j := \frac{1}{4}\sum_{q=1}^m (A_{qj}y_q)^2 \quad \forall j=1,2,\cdots,N,
$$
where $A_{qj}$ is the component of matrix $A$ at $qth$ row and $jth$ column.
For block versions RCD v.1, RCD v.2 and UCDC v.2, we set $\tau = \ceil{0.001 N}$.
The result of this experiment is shown in Figure \ref{fig2}.
\begin{figure}[h!]%
\centering
\subfloat[webspam, $F(x)$ against iterations]{%
\label{fig2a}%
\includegraphics[scale=0.32]{fig2a.pdf}}
\subfloat[webspam, $F(x)$ against time]{%
\label{fig2b}%
\includegraphics[scale=0.32]{fig2b.pdf}}\\
\subfloat[kdda, $F(x)$ against iterations]{%
\label{fig2c}%
\includegraphics[scale=0.32]{fig2c.pdf}}
\subfloat[kdda, $F(x)$ against time]{%
\label{fig2d}%
\includegraphics[scale=0.32]{fig2d.pdf}}\\
\caption{Performance of all four methods RCD v.1 and v.2 and UCDC v.1 and v.2 on two large scale $\ell_1$-regularized logistic regression problems.
The first and second rows of figures show the results for problems \textit{webspam} and \textit{kdda}, respectively.
\textit{Calculation of $F(x)$ is not included in running time of the methods.}}
\label{fig2}%
\end{figure}
In this experiment all methods were terminated after one hour of running time. Notice that RCD versions were more efficient than both UCDC versions,
with RCD v.1 being the fastest among all. An interesting observation in Figures \ref{fig2a} and \ref{fig2c} is that RCD versions had similar per iteration computational complexity
since they performed similar number of iterations within the maximum allowed running time. However, for RCD v.1, it seems that diagonal information from the second order derivatives of $f$ was enough
to decrease faster the objective function for all iterations compared to RCD v.2. Finally, in this experiment we observed that both RCD versions accepted unit step sizes for a major part
of the process.
\section{Conclusion}
We presented a robust randomized block coordinate descent method for composite function problems \eqref{Def_F}, which we name \textit{Robust Coordinate Descent} (RCD), that can properly handle second-order (curvature) information.
The proposed method can vary from first- to second-order; depending on how large the block updates are set, how accurate second-order information are used and how
inexactly the arising subproblems are solved.
Although the per iteration computational complexity might be higher for RCD, we present synthetic and real world large scale examples where the number of iterations substantially decreases, as well as the overall time.
From the theoretical point of view, we prove global convergence of RCD and under standard assumptions we show that RCD exhibits on expectation \textit{block} quadratic or superlinear rate of convergence.
\bibliographystyle{plain}
|
\section{\setcounter{equation}{0}\oldsection}
\renewcommand\thesection{\arabic{section}}
\renewcommand\theequation{\thesection.\arabic{equation}}
\newtheorem{claim}{\noindent Claim}[section]
\newtheorem{theorem}{\noindent Theorem}[section]
\newtheorem{lemma}{\noindent Lemma}[section]
\newtheorem{proposition}{\noindent Proposition}[section]
\newtheorem{definition}{\noindent Definition}[section]
\newtheorem{remark}{\noindent Remark}[section]
\newtheorem{corollary}{\noindent Corollary}[section]
\newtheorem{example}{\noindent Example}[section]
\title{The global smooth symmetric solution to 2-D
full compressible Euler system of Chaplygin gases}
\author{Ding,
Bingbing$^{1,*}$; \quad Witt, Ingo$^{2,**}$;\quad Yin,
Huicheng$^{1,}$\footnote{Ding Bingbing and Yin Huicheng are
supported by the NSFC (No.~10931007, No.~11025105), and by the
Priority Academic Program Development of Jiangsu Higher Education
Institutions.}\vspace{0.5cm}\\
\small 1. Department of Mathematics and
IMS, Nanjing University, Nanjing 210093, China.\\
\vspace{0.5cm}
\small 2.
Mathematical Institute, University of G\"{o}ttingen,
Bunsenstr.~3-5, D-37073 G\"{o}ttingen, Germany. }
\date{}
\maketitle
\centerline {\bf Abstract} \vskip 0.3 true cm
For one dimensional or multidimensional compressible Euler system of polytropic gases,
it is well known that the smooth solution will generally develop singularities in finite time.
However, for three dimensional Chaplygin gases, due to the crucial role of ``null condition'' in
the potential equation which is derived by the irrotational and isentropic flow,
P.Godin in [9] has proved the global existence of a smooth 3-D spherically symmetric flow with variable
entropy when the initial data are of small smooth perturbations with compact
supports to a constant state. It is noted that there are some essential
differences on the global solution or blowup problems between 2-D and 3-D hyperbolic systems.
In this paper, we will focus on the global symmetric solution problem of 2-D
full compressible Euler system of Chaplygin gases.
Through carrying out involved analysis and finding an appropriate weight
we can derive some uniform weighted energy estimates on the small
symmetric solution to 2-D compressible Euler system of Chaplygin gases and further establish the
global existence of smooth solution by continuous induction method.
\vskip 0.3 true cm
{\bf Keywords:} Full compressible Euler system, Chaplygin gases, global existence,
null condition, ghost weight, weighted energy estimate\vskip 0.3 true cm
{\bf Mathematical Subject Classification 2000:} 35L05, 35L72
\vskip 0.4 true cm
\centerline{\bf $\S 1$. Introduction and main results}
\vskip 0.3 true cm
In this paper, we are concerned with the global existence of a smooth symmetric solution
to 2-D full compressible Euler
system of Chaplygin gases. The 2-D full Euler system is
\begin{equation}
\left\{
\begin{aligned}
&\p_t\rho+div (\rho u)=0\qquad \qquad \qquad \qquad \qquad \qquad \qquad \qquad\text{(Conservation of mass)},\\
&\p_t(\rho u)+div (\rho u \otimes u) + \nabla P=0\qquad \qquad \qquad \qquad\qquad \text{(Conservation of momentum)},\\
&\p_t\bigl(\rho ( e + \ds\frac{|u|^2}{2})\bigr)+div \bigl( (\rho ( e + \ds\frac{|u|^2}{2}) + P
\bigr) u \bigr)=0\qquad\qquad \text{(Conservation of energy)},\\
&P=P(\rho, S),\qquad e=e(\rho, S)\qquad \qquad \qquad \qquad \qquad\qquad \text{(Equations of
state)},
\end{aligned}
\right.\tag{1.1}
\end{equation}
where $t\ge 0$, $x=(x_1, x_2)\in\Bbb R^2$, $\na=(\p_1, \p_2)$, and
$u=(u_1,u_2), \rho, P, e, S$ stand for the
velocity, density, pressure, internal energy, specific entropy
respectively. Moreover, the pressure function $P=P(\rho,S)$ and
the internal energy function $e=e(\rho,S)$ are smooth in their
arguments. In particular, $\p_{\rho}P(\rho,S)>0$ and
$\p_{S}e(\rho,S)>0$ for $\rho>0$. When $P(\rho, S)=A\rho^{\g}e^{\f{S}{c_v}}$
and $e(\rho,S)=\ds\f{A}{\g-1}\rho^{\g-1}e^{\f{S}{c_v}}$ hold for some positive
constants $A, c_v$ and $\g$ ($1<\g<3$), such flows
are then called the polytropic gases.
For the Chaplygin gases, the equation of pressure state (one can see [7] and so on)
is given by
$$P=P_0-\ds\f{A(S)}{\rho},\eqno{(1.2)}$$
where $P_0>0$ is a positive constant, $A(S)$ is a positive smooth function of $S$,
and $P>0$ for $\rho>0$.
If $(\rho, u, S)\in C^1$ is a solution of (1.1) with $\rho>0$, then (1.1) admits
the following
equivalent form
\begin{equation}
\left\{
\begin{aligned}
&\p_t\rho+div(\rho u)=0,\\
&\p_tu+u\cdot\na u+\ds\f{\na P}{\rho}=0,\\
&\p_tS+u\cdot\na S=0.
\end{aligned}
\right.\tag{1.3}
\end{equation}
We pose the symmetric initial data of (1.3) as follows:
\begin{equation}
\left\{
\begin{aligned}
\rho(0,x)&=\bar\rho+\varepsilon\rho_0(r), \\
u(0,x)&=\varepsilon U_0(r)\ds\f{x}{r},\\
S(0,x)&=\bar S+\varepsilon S_0(r),
\end{aligned}
\right.\tag{1.4}
\end{equation}
where $\bar\rho>0$ and $\bar S\in\Bbb R$ are constants, $\ve>0$
is a small parameter, $r=\sqrt{x_1^2+x_2^2}$ and $(\rho_0(r), U_0(r),$ $S_0(r))
\in C_0^{\infty}(B(0, M))$
(here and below $B(0, M)$ stands for a ball centered at the origin with a radius $M>0$).
Moreover, $\rho(0,x)>0$, $P_0-\ds\f{A(\bar S)}{\bar\rho}>0$ and $A'(\bar S)\not=0$ hold.
Our main result in this paper is
{\bf Theorem 1.1.} {\it Under the assumptions above, if $\ve>0$ is small enough, then (1.2)-(1.4) has a global
$C^{\infty}([0, \infty)\times\Bbb R^2)$ solution $(\rho(t,x), u(t,x), S(t,x))$ which admits such a symmetric
structure: $(\rho(t,x), u(t,x),$ $S(t,x))=(\rho(t, r), U(t,r)\ds\f{x}{r}, S(t,r))$.}
{\bf Remark 1.1.} {\it For the polytropic gases, it is well-known that the
bounded smooth solution of 1-D or multidimensional full compressible Euler system
will generally develop singularities in finite time whether the vacuum states appear or not,
one can see [1-2], [6-8], [14], [17]
[19-23] and the references therein. However, for the Chaplygin gases, by the results of 1-D case in [17],
3-D symmetric case in [9] and our Theorem 1.1 for 2-D symmetric case, the small perturbed symmetric
flows will exist globally.
Here we point out that the main difference for 2-D (or 3-D)
compressible Euler systems of polytropic gases and of Chaplygin gases is: if one neglects
the influences of rotations and entropy, the resulting potential equation of polytropic gases,
which is a 2-D (or 3-D) quasilinear wave equation,
does not fulfill the ``null conditions'' put forward in [4], [5] and [12]
but the related potential equation of Chaplygin gases does. In fact, when the 2-D or 3-D
quasilinear wave equations satisfy the null conditions, it is well-known that the small data
smooth solutions will exist globally (see [4-5], [11], [12] and [16]), otherwise the smooth
solution will blow up in finite time (one can see [3], [10-11] and the references therein).}
{\bf Remark 1.2.} {\it As in Remark 2.2 of [9], Theorem 1.1 still holds when
the initial data (1.4) is replaced by $(\rho(0,x), u(0,x), S(0,x))=(\bar\rho+\ve\rho_0(r),
\bar u+\ve U_0(r)\ds\f{x}{r}, \bar S+\varepsilon S_0(r))$ where $\bar u\in\Bbb R^2$ is a constant vector
since the compressible Euler system is invariable under the translation transformation.}
Let us give some comments on the proof of Theorem 1.1. As the usual first step to prove the global existence or blowup
of a smooth small data solution for the quasilinear hyperbolic equation, one should construct a suitable approximate solution
$(\rho_a, u_a, S_a)$ to (1.2)-(1.4). To this end, as in [8-9] and [21], a solution to the
related potential equation with suitable initial data will be taken as the approximate solution.
Through considering the difference of the real solution $(\rho, u, S)$ and approximate solution
$(\rho_a, u_a, S_a)$, and by applying a transformation of unknowns introduced in [21],
one can obtain a new symmetric hyperbolic system. From this system, we can get a potential
equation satisfying both null conditions in two space dimensions outside a compact support of
perturbed entropy. The so-called both null conditions in two space dimensions are defined in [4] as follows:
For the 2-D quasilinear wave equation $\square w+\ds\sum_{0\le i,j,k\le 2}g_{ij}^k\p_kw\p_{ij}^2w+
\ds\sum_{0\le i,j,k,l\le 2}g_{ij}^{kl}\p_kw\p_lw\p_{ij}^2w=0$, where $x_0=t$, $g_{ij}^k$ and
$g_{ij}^{kl}$ are certain constants, then $\ds\sum_{0\le i,j,k\le 2}g_{ij}^k\xi_k\xi_i\xi_j\equiv 0$
and $\ds\sum_{0\le i,j,k,l\le 2}g_{ij}^{kl}\xi_k\xi_l\xi_i\xi_j\equiv 0$ hold for $(\xi_0, \xi_1, \xi_2)
=(-1, cos\th, sin\th)$ with $0\le\th\le 2\pi$. Especially, the former
$\ds\sum_{0\le i,j,k\le 2}g_{ij}^k\xi_k\xi_i\xi_j\equiv 0$ and the
latter $\ds\sum_{0\le i,j,k,l\le 2}g_{ij}^{kl}\xi_k\xi_l\xi_i\xi_j\equiv 0$ are called the first and the second null condition respectively
by the terminology in [4]. Under both null conditions, S. Alinhac in [4] established the global existence of small data solution
to the 2-D quasilinear wave equation by looking for a crucial ``ghost weight" to derive an energy estimate.
In this paper, we will focus on the global existence of solution to (1.2)-(1.4). To this end,
we require to derive the uniform weighted energy estimate of solution $(\rho, u, S)$ so that the uniform bound
of $(\rho, u, S)$ for $(t,x)\in [0, \infty)\times\Bbb R^2$ can be obtained. Motivated by the methods in [4],
we will look for a suitable ``ghost weight'' to deal with the related compressible Euler system
(1.3) so that both null conditions and the variable entropy can be simultaneously considered.
Here we point out that the ghost weight introduced in [4] will not be applied for our case directly
since (1.3) can not be changed into a scalar quasilinear wave equation due to the influence of variable entropy,
and thus the ghost weight in [4] should be suitably adjusted for our uses (one can see more detailed explanations
in Remark 4.1 of $\S 4$ below).
On the other hand, although some procedures in this paper are somewhat analogous to those in [8-9] for considering the 3-D
symmetric Euler system, our analysis
is more involved since the decay rate of solution to 2-D free wave equation is lower
than that in 3-D case as well as the treatments on the both null conditions in 2-D quasilinear wave equation
are more complicated than the treatments on one null condition in 3-D quasilinear wave equation (one can compare the reference
[4] with [5] and [12]).
The rest of the paper is organized as follows. In $\S 2$, we will
construct an approximate solution of (1.2)-(1.4) and give some useful preliminary
knowledge. In $\S 3$ and $\S 4$, we will establish the uniform higher order
energy estimates on the solution $(\rho, u, S)$ near the light cone and in the whole time-space respectively.
Based on these uniform estimates, Theorem 1.1 will be proved by continuous induction method.
In what follows, we will use the following convention as in [8-9] and [21]:
$x_0=t$ $(\geq 0)$ denotes the time variable, $\partial=(\partial_0,\partial_1,\partial_2)=(\partial_t,\nabla)$;
$
\Omega=x^\perp\cdot\nabla$ with $x^\perp=(-x_2,x_1)$, $
X=t\partial_t+r\partial_r
$, $
L_j=t\partial_j+x_j\partial_t$ for $j=1,2$;
$\Lambda=(\Lambda_1, \cdots, \Lambda_7)=(\partial, \Omega, X, L_1, L_2),
$ $\Gamma=(\Gamma_1,\cdots,\Gamma_4)=(\partial, X)$;
$Z=(Z_1,Z_2)=(\pa_1+\omega_1\pa_t,\pa_2+\omega_2\pa_t)=\ds\frac{t-r}{t}(\pa_1,\pa_2)
+\frac{(x_1,x_2)X+(-x_2,x_1)\Omega}{rt}$ with $\o_i=\ds\f{x_i}{r}$ ($i=1,2$) and $\o=(\o_1, \o_2)$;
For $\Bbb{R}^l$-valued smooth functions $f(t,x)$ and $\tilde f(t,x)$, we put
$$
|f(t)|=\sup_{x\in\Bbb{R}^2}|f(t,x)|,\quad |f(t)|_\pm=\sup_{D_\pm}|f(t,x)|,
$$
$$
\langle f,\tilde f\rangle(t)=\int_{\Bbb{R}^2}f(t,x)\cdot\tilde f(t,x)\,{d x},
\quad \langle f,\tilde f\rangle_\pm(t)=\int_{D_\pm(t)}f(t,x)\cdot\tilde f(t,x)\,{d x},
$$
and
$$
\|f(t)\|=\langle f,f\rangle^{1/2}(t),\quad \|f(t)\|_\pm=\langle f,f\rangle_\pm^{1/2}(t),
$$
where $D_-(t)=\{x\in\Bbb{R}^2:
r\leq\ds\frac{t}{2}+M+1\}$ and $D_+(t)=\{x\in\Bbb{R}^2:
r\geq\ds\frac{t}{2}+M+1\}$;
If $g$ is only a function of $x$, then $\|g\|$ represents the usual $L^2$ norm;
Define $\langle\xi\rangle=1+|\xi|$ for $\xi\in\Bbb R$ and $\sigma(x)=\langle x\rangle$,
$\sigma_\pm(t,x)=\langle t\pm |x|\rangle$.
\vskip 0.4 true cm \centerline{\bf $\S 2$. The construction of approximate solution to
(1.2)-(1.4) and some preliminaries} \vskip 0.4 true cm
Without loss of generality, we will assume the sound speed
$\bar c=\big(\pa_{\rho} P(\bar\rho,\bar S)\big)^{1/2}\equiv 1$
in the whole paper. As in [8-9], denote $(\rho_a, u_a, \bar S)$ by a solution of
(1.3)-(1.4) satisfying the initial data condition $u_a(0,x)=u_0(0,x)$ and $P_a(0,x)=P(0,x)$, which means
the initial density $\rho_a(0,x)=\rho_a(0,r)=\ds\frac{A(\bar S)\rho(0,x)}{A(S(0,x))}$. At this time, since the initial
data of $(\rho_a, u_a, \bar S)$ are symmetric and isentropic, one can introduce a potential $\phi_a(t,r)$
such that $u_a=\na\phi_a$ and $\phi_a$ satisfies the following potential equation
\begin{equation}
\left\{
\begin{aligned}
&\pa_t^2\phi_a+2\ds\sum_{j=1}^2\pa_j\phi_a\pa_t\pa_j\phi_a+\ds\sum_{j,k=1}^2\pa_j\phi_a\pa_k\phi_a\pa_{jk}^2\phi_a
-(1+2\pa_t\phi_a+|\nabla \phi_a|^2)\triangle \phi_a=0,\\
&\phi_a(0,r)=-\displaystyle\ve\int_{r}^MU_0(s)ds,\\
&\pa_t\phi_a(0,r)=-\displaystyle\frac{1}{2}\ve^2U_0^2(r)-h\big(\rho_a(0,r)\big),
\end{aligned}
\right.\tag{2.1}
\end{equation}
where $h(\rho)=\ds\f12-\ds\f{A(\bar S)}{2\rho^2}$ is the enthalpy. Meanwhile, the density
$\rho_a$ is determined by the Bernoulli's law $h(\rho_a)=-\p_t\phi_a-\f12|\na\phi_a|^2$.
It is easy to verify that (2.1) satisfies both null conditions in two space dimensions
(i.e., the first null condition and second null condition, which have been illustrated in $\S 1$) posed in [4]. Then
(2.1) has a global smooth solution $\phi_a$ in terms of [4] and we have
{\bf Lemma 2.1.}
{\it If $\ve>0$ is small enough, then
$(1)\quad(|\Lambda^{\al}(\rho_a-\bar\rho)|+|\Lambda^{\al}u_a|)(t,x)
\leq C_{\al}\ve\si_-(t,x)^{-1}\si_+(t,x)^{-1/2}$;
$(2)\quad|\pa\Lambda^{\al}(\rho_a-\bar\rho)(t,x)|+|\pa\Lambda^{\al}u_a(t,x)|
\leq C_{\al}\ve\langle t\rangle^{-5/2}$ if $|x|\leq lt+M$ with $0\leq l<1$;
$(3)\quad|X^ku_a(t,x)|\leq C_{k\delta}\ve\langle t\rangle^{-5/2+\delta}$ if $0\leq\delta<1$
and $|x|\leq C\langle t\rangle^{\delta}$,
where $C_{\al}$, $C_{k\delta}$ and $C$ are some generic positive constants independent of $\ve$ and $(t,x)$.}
{\bf Proof.} Since (2.1) satisfies both null conditions in 2-D spaces, then one
has by the result of [4] or [13]
$$
|\Lambda^{\al}\pa^\beta \phi_a|\leq C_{\al}\ve\si_-(t,x)^{-|\beta|}\si_+(t,x)^{-1/2}.\eqno{(2.2)}
$$
From this, (1) and (2) can be obtained directly.
In addition, by $\ds X^ku_a(t,x)
=(X^k\p_r\phi_a)(t,r)\frac{x}{r}
=\biggl(\int_0^r\pa_\lambda (X^k\p_r\phi_a)(t,\lambda)\,{d \lambda}\biggr)\frac{x}{r}$ and (2.2), then (3) holds
obviously.
\qquad\qquad\qquad\qquad \qquad\qquad \qquad\qquad \qquad\qquad \qquad\qquad $\square$\bigskip
As in [8] and [21], we set $\theta(t,x)=1-\ds\frac{A(S(t,x))\bar\rho}{A(\bar S)\rho(t,x)}$,
$w(t,x)=u(t,x)$, $z(t,x)=\ds\frac{A(\bar S)}{A(S(t,x))}-1$, then it follows from (1.3)-(1.4) that
\begin{equation}
\left\{
\begin{aligned}
&\partial_t\theta+w\cdot\nabla\theta+(1-\theta)\nabla\cdot w=0,\\
&\partial_tw+w\cdot\nabla w+(1-\theta)(1+z)\nabla \theta=0,\\
&\partial_tz+w\cdot\nabla z=0,\\
&\theta(0,x)=1-\ds\frac{A(\bar S+\ve S_0(r))\bar\rho}{A(\bar S)(\bar\rho+\ve\rho_0(r))},\\
&w(0,x)=\varepsilon U_0(r)\ds\f{x}{r},\\
&z(0,x)=\ds\frac{A(\bar S)}{A(\bar S+\ve S_0(r))}-1.
\end{aligned}
\right.\tag{2.3}
\end{equation}
Corresponding to the approximate solution $(\rho_a,u_a,\bar S)$, define $\theta_a(t,x)=1-\ds\frac{\bar\rho}{\rho_a(t,x)}$
and $w_a(t,x)=u_a(t,x)$. Then a direct computation yields by Lemma 2.1 that
{\bf Lemma 2.2.}
{\it For small $\ve>0$, we have
$(1)\quad(|\Lambda^{\al}\theta_a|+|\Lambda^{\al}w_a|)(t,x)\leq C_{\al}\ve\si_-(t,x)^{-1}\si_+(t,x)^{-1/2}$;
$(2)\quad|\pa\Lambda^{\al}\theta_a(t,x)|+|\pa\Lambda^{\al}w_a(t,x)|\leq C_{\al}\ve\langle t\rangle^{-5/2}$
if $|x|\le lt+M$ with $0\le l< 1$;
$(3)\quad|X^kw_a(t,x)|\leq C_{k\delta}\ve\langle t\rangle^{-5/2+\delta}$ if $0\leq\delta<1$ and $|x|\leq C\langle t\rangle^{\delta}$.\hfil\break}
Let $\dot\theta=\theta-\theta_a$, $\dot w=w-w_a$ and $\dot z=z-z(0,x)$, then we have by (2.3)
\begin{equation}
\left\{
\begin{aligned}
&\partial_t\dot\theta+\nabla\cdot \dot w=-(w\cdot\nabla\theta-w_a\cdot\nabla\theta_a)+(\theta\nabla\cdot w
-\theta_a\nabla\cdot w_a),\\
&\partial_t\dot w+\nabla\dot \theta=-(w\cdot\nabla w-w_a\cdot\nabla w_a)+(\theta\nabla\theta-\theta_a\nabla\theta_a)-(1-\theta)z\nabla\theta,\\
&\partial_t\dot z=-w\cdot\nabla(z(0,x)+\dot z), \\
&\dot \theta(0,x)=0,\quad\dot w(0,x)=0,\quad \dot z(0,x)=0.
\end{aligned}
\right.\tag{2.4}
\end{equation}
To establish the global existence of solution to (2.4) in subsequent sections, we require to
give some preliminary analysis
on the related energies.
As usual, we define the energy for $n\in\Bbb N\cup\{0\}$
$$
E_n(t)=\sum_{|\al|\leq n}(\|\Gamma^{\al}\dot \theta(t)\|^2
+\|\Gamma^{\al} \dot w(t)\|^2+\|\Gamma^{\al} \dot z(t)\|^2).
$$
We also set $Q_n(t)=\ds\sum_{|\al|\leq n-1}\bigg(\|\si_-(t)\nabla\Gamma^{\al}\dot \theta(t)\|
+\|\si_-(t)\pa_t\Gamma^{\al}\dot w(t)\|
+\|\si_-(t)\nabla\cdot\Gamma^{\al}\dot w(t)\|\bigg)$ for $n\geq 1$, and define
$\widetilde Q_n(t)=Q_n(t)+E_{n-1}^{1/2}(t)$
for $n\geq 1$ and $\widehat Q_n(t)=Q_n(t)+E_{n-2}^{1/2}(t)$ for $n\geq 2$ as in [8-9]
and [21].
In addition, for our requirements to treat the 2-D full Euler system (1.3)-(1.4), it is necessary
to introduce some kinds of ``interior energies'' as follows:
Choose a smooth function
\begin{equation*}
\hat\chi(s) = \left\{
\begin{aligned}
& 1, \quad \text{if $s\leq1/2$},\\
& 0, \quad \text{if $s\geq3/4$},
\end{aligned}
\right.
\end{equation*}
and set $\chi(t,x)=\hat\chi(\ds\frac{|x|}{t+2M+2})$, then we define for $n\geq 1$
\begin{align*}
&Q_n^-(t)=\ds\sum_{|\al|\leq n-1}(\|\si_-(t)\nabla\Gamma^{\al}(\chi\dot\theta)(t)\|
+\|\si_-(t)\pa_t\Gamma^{\al}(\chi \dot w)(t)\|
+\|\si_-(t)\nabla\cdot\Gamma^{\al}(\chi \dot w)(t)\|),\\
&\widetilde Q_n^-(t)=Q_n^-(t)+E_{n-1}^{1/2}(t),\quad\text{for $n\geq 1$},\\
&\widehat Q_n^-(t)=Q_n^-(t)+E_{n-2}^{1/2}(t),\quad \text{for $n\geq 2$}.
\end{align*}
The following relations and properties on the above defined energies will be repeatedly utilized
later on.
{\bf Lemma 2.3.} {\it
\begin{align*}
(1)\quad&\|\si_-(t)\nabla v\|\leq C(\|\si_-(t)\nabla\cdot v\|+\|v\|)\quad\text{if $v\in C_0^\infty(\Bbb R^2,\Bbb R^2)$ and $\nabla\cdot v^\bot=0$,}\\
(2)\quad&|\si^{1/2}\Gamma^{\al}\dot w(t)|+|\si^{1/2}\Gamma^{\al}\dot\theta(t)|
+|\si^{1/2}\Gamma^{\al}\dot z(t)|\leq C_{\al}E_{|\al|+2}^{1/2}(t),\\
(3)\quad&|\si^{1/2}\si_-(t)\nabla\Gamma^{\al}\dot\theta(t)|\leq C_{\al}Q_{|\al|+3}(t),\\
(4)\quad&|\si^{1/2}\si_-(t)\nabla\Gamma^{\al}(\chi\dot\theta)(t)|\leq C_{\al}Q^-_{|\al|+3}(t),\\
(5)\quad&|\si^{1/2}\si_-(t)\nabla\Gamma^{\al}\dot w(t)|\leq C_{\al}\widetilde Q_{|\al|+3}(t),\\
(6)\quad&|\si^{1/2}\si_-(t)\nabla\Gamma^{\al}(\chi \dot w)(t)|\leq C_{\al}\widetilde Q^-_{|\al|+3}(t),\\
(7)\quad&|\si^{1/2}\si_-^{1/2}(t)\Gamma^{\al}\dot\theta(t)|\leq C_{\al}\widehat Q_{|\al|+2}(t),\\
(8)\quad&|\si^{1/2}\si_-^{1/2}(t)\Gamma^{\al}(\chi\dot\theta)(t)|\leq C_{\al}\widehat Q^-_{|\al|+2}(t),\\
(9)\quad&|\si^{1/2}\si_-^{1/2}(t)\Gamma^{\al}\dot w(t)|\leq C_{\al}\widetilde Q_{|\al|+2}(t),\\
(10)\quad&|\si^{1/2}\si_-^{1/2}(t)\Gamma^{\al}(\chi \dot w)(t)|\leq C_{\al}\widetilde Q_{|\al|+2}^-(t).
\end{align*}
}
{\bf Proof.} (1) comes from the formula (6.7) in [21] directly.
In addition, according to Lemma 1 of [21], one has for any smooth function $f(x)$
\begin{equation}
\left\{
\begin{aligned}
&\si(x)^{1/2}|f(x)|\leq C\ds\sum_{j=0}^1\sum_{|\a|=0}^{2-j}\|\nabla^\a\Omega^jf\|,\\
&\si(x)^{1/2}\si_-(t,x)|f(x)|\leq C\ds\sum_{j=0}^1\sum_{|\a|=0}^{2-j}\|\si_-(t)\nabla^\a\Omega^jf\|,\\
&\si(x)^{1/2}\si_-(t,x)^{1/2}|f(x)|\leq C\ds\sum_{j=0}^1\big(\|\Omega^jf\|+\sum_{|\a|=1}^{2-j}\|\si_-(t)\nabla^\a\Omega^jf\|\big).
\end{aligned}
\right.\tag{2.5}
\end{equation}
This, together with the facts of $\nabla\cdot(\Gamma^a(\chi \dot w))^\bot=0$
and $[\Omega,\Gamma^{\al}]=\ds\sum_{|\beta|\leq|\al|}C_{\al\beta}\Gamma^{\beta}$, yields
$(2)-(10)$ directly.\qquad \qquad \qquad \qquad \qquad \qquad \qquad $\square$
\bigskip
Applying $\pa^\a(X+1)^k$ to (2.4) and setting $\Gamma^{\mu}=\pa^\a X^k$, then we have
\begin{equation}
\left\{
\begin{aligned}
\pa_t\Gamma^{\mu}\dot\theta+\nabla\cdot\Gamma^{\mu}\dot w&=h_0^{\mu}\equiv \ds\sum_{1\leq j\leq6}f_j^{\mu},\\
\pa_t\Gamma^{\mu}\dot w+\nabla\Gamma^{\mu}\dot\theta&=h^{\mu}\equiv \ds\sum_{7\leq j\leq 13}f_j^{\mu},
\end{aligned}
\right.\tag{2.6}
\end{equation}
\break
$f_1^{\mu}=-\ds\sum_{{\nu}\leq \mu}\binom{\mu}{\nu}\Gamma^{\nu}w_a\cdot\nabla\Gamma^{\mu-\nu}\dot\theta,
\quad f_2^{\mu}=-\ds\sum_{\nu\leq \mu}\binom{\mu}{\nu}\Gamma^{\nu}\dot w\cdot\nabla\Gamma^{\mu-\nu}\theta_a,$\hfil\break
$f_3^{\mu}=-\ds\sum_{\nu\leq \mu}\binom{\mu}{\nu}\Gamma^{\nu}\dot w\cdot\nabla\Gamma^{\mu-\nu}\dot\theta,\quad
f_4^{\mu}=\ds\sum_{\nu\leq \mu}\binom{\mu}{\nu}\Gamma^{\nu}\theta_a\nabla\cdot\Gamma^{\mu-\nu}\dot w,$\hfil\break
$f_5^{\mu}=\sum_{\nu\leq \mu}\binom{\mu}{\nu}\Gamma^{\nu}\dot\theta\nabla\cdot\Gamma^{\mu-\nu}w_a,\quad
f_6^{\mu}=\sum_{\nu\leq \mu}\binom{\mu}{\nu}\Gamma^{\nu}\dot\theta\nabla\cdot\Gamma^{\mu-\nu}\dot w,$\hfil\break
$f_7^{\mu}=-\ds\sum_{\nu\leq\mu}\binom{\mu}{\nu}\Gamma^{\nu}w_a\cdot\nabla\Gamma^{\mu-\nu}\dot w,\quad f_8^{\mu}=-\ds\sum_{\nu\leq\mu}\binom{\mu}{\nu}\Gamma^{\nu}\dot w\cdot\nabla\Gamma^{\mu-\nu}w_a,
$\hfil\break
$f_9^{\mu}=-\ds\sum_{\nu\leq\mu}\binom{\mu}{\nu}\Gamma^{\nu}\dot w\cdot\nabla\Gamma^{\mu-\nu}\dot w,
\quad f_{10}^{\mu}=\ds\sum_{\nu\leq \mu}\binom{\mu}{\nu}\Gamma^{\nu}\theta_a\nabla\Gamma^{\mu-\nu}\dot\theta,$\hfil\break
$f_{11}^{\mu}=\sum_{\nu\leq \mu}\binom{\mu}{\nu}\Gamma^{\nu}\dot\theta\nabla\Gamma^{\mu-\nu}\theta_a,\quad
f_{12}^{\mu}=\sum_{\nu\leq\mu}\binom{\mu}{\nu}\Gamma^{\nu}\dot\theta\nabla\Gamma^{\mu-\nu}\dot\theta,$\hfil\break
$f_{13}^{\mu}=-\pa^\a(X+1)^k\big((1-\theta)z\nabla\theta\big)$,
\quad here we point out that the concrete expressions of
$f_i^{\mu}$ ($1\le i\le 12$) are important in order to derive the basic energy estimate in Lemma 3.5 below
(one can see the details in dealing with (3.21)) since
we require to utilize the first null condition by observing the main parts of $f_1^{\mu}$ and $f_4^{\mu}$ (or $f_7^{\mu}$ and $f_{10}^{\mu}$), and the second null condition by observing the left parts of $f_1^{\mu}$ and $f_4^{\mu}$
(or $f_7^{\mu}$ and $f_{10}^{\mu}$) except the terms in the first null condition together with the main parts of $f_i^{\mu}$ for $i= 2, 3, 5, 6$ (or $f_i^{\mu}$ for $i=8, 9, 11, 12$).
It follows from an analogous computation in [9] that
\begin{align*}
&(t-r)\nabla\Gamma^{\mu}\dot\theta=-\frac{t}{t+r}X\Gamma^{\mu}\dot w+\frac{t}{t+r}(\Omega\Gamma^{\mu}\dot w)^\bot-(X\Gamma^{\mu}\dot\theta)\frac{x}{t+r}\\
&\qquad \qquad \qquad -\frac{x^\bot}{t+r}\Omega\Gamma^{\mu}\dot\theta
+\frac{t}{t+r}h_0^{\mu}x+\frac{t^2}{t+r}h^{\mu},\tag{2.7}\\
&(t-r)\nabla\cdot\Gamma^{\mu}\dot w=-\frac{t}{t+r}X\Gamma^{\mu}\dot\theta-\frac{X\Gamma^{\mu}\dot w}{t+r}\cdot x-\frac{(\Omega\Gamma^{\mu}\dot w)^\bot}{t+r}\cdot x+\frac{t^2}{t+r}h_0^{\mu}
+\frac{t}{t+r}h^{\mu}\cdot x,\tag{2.8}\\
&(t-r)\pa_t\Gamma^{\mu}\dot w=\frac{t}{t+r}X\Gamma^{\mu}\dot w-\frac{t}{t+r}(\Omega\Gamma^{\mu}\dot w)^\bot+(X\Gamma^{\mu}\dot\theta)\frac{x}{t+r}+\frac{x^\bot}{t+r}\Omega\Gamma^{\mu}\dot\theta\\
&\qquad \qquad \qquad -\frac{t}{t+r}h_0^{\mu}x-\frac{r^2}{t+r}h^{\mu}.\tag{2.9}
\end{align*}
Based on (2.7)-(2.9), we can establish a useful energy inequality inside the light cone as follows:
{\bf Lemma 2.4.} {\it
For fixed $n\in\Bbb N$ with $n\geq 4$. There exist positive constants $\eta$ and $C$ such that
if $(\dot\theta, \dot w, \dot z)$ is a smooth solution of (2.4)
for $(t,x)\in [0, T]\times\Bbb R^2$ with $E_{[\frac{n}{2}]+2}^{1/2}(t)\leq\eta$, then for $0\leq t\leq T$ and small $\ve>0$
$$\ds Q_n^-(t)\leq C\biggl(E_n^{1/2}(t)+\frac{\ve^2}{\langle t\rangle^{3/2}}\biggr).\eqno{(2.10)}$$
}
\bigskip
{\bf Remark 2.1.} {\it For the free wave equation with compactly supported initial data,
it is well-known that the decay rate of smooth solution on the time $t$
in 2-D case is slower than that in 3-D case, therefore the author in [9] can obtain
an energy estimate of $Q_n(t)$
similar to (2.10) in the whole space $\Bbb R^3$ (one can see Proposition 4.1 of [9]) but
at present we only get (2.10) for
$Q_n^-(t)$ which is a kind of interior energy.}
{\bf Proof.} By (2.6), we have for $|\mu|\le n-1$
\begin{equation}
\left\{
\begin{aligned}
&\pa_t\Gamma^{\mu}(\chi\dot\theta)+\nabla\cdot\Gamma^{\mu}(\chi \dot w)=\tilde h_0^{\mu}\equiv
\ds\sum_{0\leq j\leq 6}\tilde f_j^{\mu},\\
&\pa_t\Gamma^{\mu}(\chi\dot w)+\nabla\Gamma^{\mu}(\chi\dot\theta)=\tilde h^{\mu}\equiv \ds\sum_{7\leq j\leq 14}\tilde f_j^{\mu},
\end{aligned}
\right.\tag{2.11}
\end{equation}
where\hfil\break
$\tilde f_0^{\mu}=\pa^\a(X+1)^k\bigg\{\ds\frac{\chi'}{t+2M+2}\biggl(-\frac{r}{t+2M+2}\dot\theta
+\dot w\cdot\omega+w\cdot\omega\dot\theta-\theta\omega\cdot \dot w\biggr)\bigg\},$\hfil\break
$\tilde f_3^\mu=-\ds\sum_{\nu\leq\mu}\binom{\mu}{\nu}\Gamma^\nu\dot w\cdot\nabla\Gamma^{\mu-\nu}(\chi\dot\theta),\quad\tilde f_6^\mu=\ds\sum_{\nu\leq\mu}\binom{\mu}{\nu}\Gamma^\nu\dot\theta\nabla\cdot\Gamma^{\mu-\nu}(\chi\dot w),
$\hfil\break
$\tilde f_9^\mu=-\ds\sum_{\nu\leq\mu}\binom{\mu}{\nu}\Gamma^\nu\dot w\cdot\nabla\Gamma^{\mu-\nu}(\chi \dot w),\quad
\tilde f_{12}^\mu=\ds\sum_{\nu\leq\mu}\binom{\mu}{\nu}\Gamma^\nu\dot\theta\nabla\Gamma^{\mu-\nu}(\chi\dot\theta),
$\hfil\break
$\ds\tilde f_{13}^\mu=-\pa^\a(X+1)^k\big(\chi(1-\theta)z\nabla\theta\big),$\hfil\break
$\ds\tilde f_{14}^\mu=\pa^\a(X+1)^k\bigg\{\frac{\chi'}{t+2M+2}\bigg(-\frac{r}{t+2M+2}
\dot w+\dot\theta\omega+w\cdot\omega \dot w-\theta\dot\theta\omega\bigg)\bigg\},$\hfil\break
and for $i=1,2,4,5,7,8,10,11$, the expressions of $\t f_i^{\mu}$ are the same as $f_i^{\mu}$ in (2.6)
when $\dot\th$ or $\dot w$
in $f_i^{\mu}$ is replaced by $\chi\dot\th$ or $\chi\dot w$ respectively.
By (2.11) together with the similar expressions of (2.7)-(2.9), an easy computation yields
$$
Q_n^-(t)\leq C\big(E_n^{1/2}(t)+\sum_{|\mu|\leq n-1}t\|\tilde h_0^\mu(t)\|+\sum_{|\mu|\leq n-1}\langle t\rangle\|\tilde h^\mu(t)\|\big).\eqno{(2.12)}
$$
We now focus on the estimates of $\|\tilde h_0^\mu(t)\|$ and $\|\tilde h^\mu(t)\|$ in the right hand side of (2.12).
According to Lemma 2.2 and by direct observations, we can obtain
$$
\|\tilde f_j^\mu\|\leq C_n\ve\langle t\rangle^{-3/2} E_n^{1/2}(t),\quad j\in\{1,2,4,5,7,8,10,11\},\eqno{(2.13)}
$$
and
$$
\|\tilde f_0^\mu(t)\|+\|\tilde f_{14}^\mu(t)\|\leq C_n\langle t\rangle^{-1}E_{n-1}^{1/2}(t).\eqno{(2.14)}
$$
In addition, applying Lemma 2.3 (2) and (4) respectively yield that
if $|\nu|\leq|\mu-\nu|$, then
$$
\|\Gamma^\nu\dot w\cdot\nabla\Gamma^{\mu-\nu}(\chi\dot\theta)(t)\|\leq C\langle t\rangle^{-1}|\Gamma^\nu \dot w(t)|\cdot\|\si_-(t)\nabla\Gamma^{\mu-\nu}(\chi\dot\theta)(t)\|\leq C_n\langle t\rangle^{-1}E_{[\frac{n-1}{2}]+2}^{1/2}(t)Q_n^-(t);
$$
if $|\nu|>|\mu-\nu|$, then
$$
\|\Gamma^\nu\dot w\cdot\nabla\Gamma^{\mu-\nu}(\chi\dot\theta)(t)\|\leq |\nabla\Gamma^{\mu-\nu}(\chi\dot\theta)(t)|\cdot\|\Gamma^{\nu}\dot w(t)\|\leq C_n\langle t\rangle^{-1}Q^-_{[\frac{n}{2}]+2}(t)E^{1/2}_{n-1}(t).
$$
Therefore, one has
$$\|\tilde f_3^\mu\|\leq C_n\langle t\rangle^{-1}\big(E_{[\frac{n-1}{2}]+2}^{1/2}(t)Q_n^-(t)+Q^-_{[\frac{n}{2}]+2}(t)E^{1/2}_{n-1}(t)\big).\eqno{(2.15)}$$
Similarly,
$$
\|\tilde f_j^\mu\|\leq C_n\langle t\rangle^{-1} \big(E_{[\frac{n-1}{2}]+2}^{1/2}(t) \widetilde Q_n^-(t)+\widetilde Q^-_{[\frac{n}{2}]+2}(t)E^{1/2}_{n-1}(t)\big),\quad j\in\{6,9,12\}.\eqno{(2.16)}
$$
Finally, we deal with $\tilde f_{13}^\mu$.
Set
\begin{align*}
&{L}_{1\nu}(t)=\|\Gamma^\nu z(0,x)\nabla\Gamma^{\mu-\nu}(\chi\theta_a)(t)\|,\\
&{L}_{2\nu}(t)=\|\Gamma^\nu z(0,x)\nabla\Gamma^{\mu-\nu}(\chi\dot\theta)(t)\|,\\
&{L}_{3\nu}(t)=\|\Gamma^\nu\dot z\nabla\Gamma^{\mu-\nu}(\chi\theta_a)(t)\|+\|\pa^\a(X+1)^k(\frac{\chi'}{t+2M+2}\dot z\theta_a\omega)(t)\|,\\
&{L}_{4\nu}(t)=\|\Gamma^\nu\dot z\nabla\Gamma^{\mu-\nu}(\chi\dot\theta)(t)\|+\|\pa^\a(X+1)^k(\ds\frac{\chi'}{t+2M+2}\dot z\dot\theta\omega)(t)\|.
\end{align*}
It is easy to obtain
\begin{align*}
&{L}_{1\nu}(t)\le C_n\ve^2\langle t\rangle^{-5/2},\\
&{L}_{2\nu}(t)\le C_n\ve\langle t\rangle^{-1}Q_n^-(t),\\
&{L}_{3\nu}(t)\leq C_n\ve\langle t\rangle^{-5/2}E_{n-1}^{1/2}(t),\\
&{L}_{4\nu}(t)\leq C_n\langle t\rangle^{-1}\bigg(E_{[\frac{n-1}{2}]+2}^{1/2}(t)Q_n^-(t)
+Q^-_{[\frac{n}{2}]+2}(t)E^{1/2}_{n-1}(t)\bigg)
+C_n\langle t\rangle^{-3/2}E_{[\frac{n-1}{2}]+2}^{1/2}(t)E_{n-1}^{1/2}(t).
\end{align*}
This, together with the expression of $\tilde f_{13}^\mu$, yields
$$\|\tilde f_{13}^\mu\|\leq C_n\big(1+E_{[\frac{n-1}{2}]+2}^{1/2}(t)\big)F_{n-1}(t)+C_nE_{n-1}^{1/2}(t)F_{[\frac{n}{2}]+1}(t),
\eqno{(2.17)}$$
where $F_{j-1}(t)$ is defined as $F_{j-1}(t)=\ve^2\langle t\rangle^{-5/2}+E_{j-1}^{1/2}(t)\big(\ve\langle t\rangle^{-5/2}+\langle t\rangle^{-1}Q^-_{[\frac{j}{2}]+2}(t)\big)+\langle t\rangle^{-1}Q_j^-(t)\big(\ve+E_{[\frac{j-1}{2}]+2}^{1/2}(t)\big)+\langle t\rangle^{-3/2}E_{[\frac{j-1}{2}]+2}^{1/2}(t)E_{j-1}^{1/2}(t)$ for $j\ge 1$.
Due to $E^{\f12}_{[\frac{n}{2}]+2}(t)\leq\eta$, (2.12) together with (2.13)-(2.17) derives
$$
Q_{[\frac{n}{2}]+2}^-(t)\leq C_n\eta\big(1+Q_{[\frac{n}{2}]+2}^-(t)\big)+C_n\ve^2\langle t\rangle^{-3/2},
$$
which means for small $\eta$
$$
Q_{[\frac{n}{2}]+2}^-(t)\leq C_n\big(\eta+\frac{\ve^2}{\langle t\rangle^{3/2}}\big).\eqno{(2.18)}
$$
And hence, $F_{n-1}(t)\leq C_n\langle t\rangle^{-1}\bigg(Q_n^-(t)(\ve+\eta)+\eta E_{n-1}^{1/2}(t)+\ds\frac{\ve^2}{\langle t\rangle^{3/2}}\bigg)$ and $F_{[\frac{n}{2}]+1}(t)\leq C_n\langle t\rangle^{-1}\bigg(\eta(\ve+\eta)+\ds\frac{\ve^2}{\langle t\rangle^{3/2}}\bigg)$. Substituting this into (2.17) and further combining with (2.12)-(2.16) yield
$$
Q_n^-(t)\leq C_n\bigg(E_n^{1/2}(t)+\frac{\ve^2}{\langle t\rangle^{3/2}}\bigg).\qquad\qquad\qquad\qquad\text{$\square$}
$$
Next, we derive the decay estimate of solution $(\dot\th,\dot w)$ inside the light cone, which is analogous to Proposition 4.2 in [9]. This is relatively easier
to be obtained than the one near the cone in subsequent $\S 3$.
{\bf Lemma 2.5.} {\it
For fixed $\la\in\Bbb N$ with $\la\geq 4$, if $(\dot\theta, \dot w, \dot z)$ is a smooth solution of (2.4) for $(t,x)\in [0, T]\times\Bbb R^2$ with $E_\la(t)\leq\ve^2$, and $\widetilde Q_{|\mu|+2}(t)\leq\eta\langle t\rangle^{1/2}$ holds for $|\mu|\geq\la-1$ and small constant $\eta>0$, then\hfil\break
$(1)\quad|\nabla\Gamma^\mu(\chi\dot\theta)(t)|+|\nabla\cdot\Gamma^\mu(\chi \dot w)(t)|\leq C\chi_{|\mu|}(t)$ for $|\mu|\leq 2\la-4$, where and below $\chi_k(t)\equiv \ds\frac{\widetilde Q_{k+3}^-(t)}{\langle t\rangle^{3/2}}+\frac{\ve^2}{\langle t\rangle^{5/2}}$.\hfil\break
$(2)\quad|\pa_t\Gamma^\mu(\chi\dot\theta)(t)|+|\pa_t\Gamma^\mu(\chi\dot w)(t)|\leq C\chi_{|\mu|}(t)$ for $|\mu|\leq 2\la-4$.
\hfil\break
$(3)\quad|\Gamma^\mu(\chi\dot w)(t)|\leq C\chi_{|\mu|-1}(t)$ for $|\mu|\leq 2\la-3$ and $\mu_2+\mu_3\neq0$.\hfil\break
$(4)\quad|\Gamma^\mu\dot w(t,x)|\leq Cr\chi_{|\mu|}(t)$ for $|\mu|\leq 2\la-4$, $\mu_2+\mu_3=0$, and $r\leq\ds\frac{t}{2}+M+1$.
}
{\bf Proof.}
(1) If $t\leq 8M+7$, (1) can be got directly by Lemma 2.3 (4) and (6). Otherwise, we know that $t^2-r^2$ is equivalent to $t^2$ when $r\leq\ds\frac{3}{4}t+\frac{3}{2}(M+1)$. Set $m_\mu(t)=|\nabla\Gamma^\mu(\chi\dot\theta)(t)|+|\nabla\cdot\Gamma^\mu(\chi\dot w)(t)|$ and $M_\mu(t)=\ds\sum_{|\nu|\leq |\mu|}m_\nu(t)$. According to the expressions (2.7)-(2.9) and Lemma 2.3, we have
$$
m_\mu(t)\leq C\Bigg(\frac{\widetilde Q^-_{|\mu|+3}(t)}{\langle t\rangle^{3/2}}+\sum_{j=1}^{13}|\tilde f_j^\mu(t)|\bigg),
\eqno{(2.19)}
$$
here we have applied $|\tilde f_0^\mu(t)|+|\tilde f_{14}^\mu(t)|\leq C\langle t \rangle^{-3/2}E^{1/2}_{|\mu|+2}(t)\leq C\langle t \rangle^{-3/2}\widetilde Q^-_{|\mu|+3}(t)$.
We now treat $\ds\sum_{j=1}^{13}|\tilde f_j^\mu(t)|$.
If $\Gamma^\nu=\pa_t^lX^k$, then $\Gamma^\nu(\chi\dot w)(t,x)=
\ds\f{1}{r}\biggl(\int_0^r s(\nabla\cdot\Gamma^\nu(\chi\dot w))(s)ds\biggr)\frac{x}{r}$. At this time,
it follows the inequality
in Lemma 2.1 (c) of [1] that
$$
|\pa_x^\a\Gamma^\nu(\chi\dot w)(t)|\leq C_\a\sum_{j\leq |\a|-1}|\nabla\cdot\pa_x^j\Gamma^\nu(\chi\dot w)(t)|.\eqno{(2.20)}
$$
This, together with Lemma 2.3 and the expressions of $\tilde f_j^\mu$ ($1\le j\le 12$), yields
$$
\sum_{j=1}^{12}|\tilde f_j^\mu(t)|\leq C\Bigg(\ds\frac{\ve\widetilde Q^-_{|\mu|+2}(t)}{\langle t\rangle^3}+E_{|\mu|+2}^{1/2}(t)M_\mu(t)+\frac{\ve}{\langle t\rangle^{3/2}}M_\mu(t)\Bigg),\eqno{(2.21)}$$
or
$$
\sum_{j=1}^{12}|\tilde f_j^\mu(t)|\leq C\Bigg(\ds\frac{\ve\widetilde Q^-_{|\mu|+2}(t)}{\langle t\rangle^3}+\frac{\widetilde Q_{|\mu|+2}(t)}{\langle t\rangle^{1/2}}M_\mu(t)+\frac{\ve}{\langle t\rangle^{3/2}}M_\mu(t)\Bigg).\eqno{(2.22)}
$$
Noticing $\tilde f_{13}^\mu=-\pa^\mu (X+1)^k(\chi z\nabla\theta)+\pa^\a(X+1)^k(\chi\theta z\nabla\theta)$
and ${\widetilde Q}_{|\mu|+2}^-(t)\le C\ve$ for $|\mu|\le\la-2$ in terms of Lemma 2.4,
then by the assumption $\widetilde Q_{|\mu|+2}(t)\leq\eta\langle t\rangle^{1/2}$
for $|\mu|\geq\la-1$, Lemma 2.3 and Lemma 2.2, we can obtain that $|\p^{\al'}X^{k'}\th(t)|$
is bounded for $|\al'|\le|\al|, k'\le k$ and further arrive at
\begin{align*}
|\tilde f_{13}^\mu(t)|\leq
&C\Big(\ve^2\langle t\rangle^{-5/2}+\ve M_\mu(t)+\ve\langle t\rangle^{-5/2}E^{1/2}_{|\mu|+2}(t)+\sum_{\nu\leq \mu}E^{1/2}_{|\nu|+2}(t)m_{\mu-\nu}(t)\\
&+\langle t\rangle^{-2}E_{[\frac{|\mu|}{2}]+2}^{1/2}\widetilde Q_{|\mu|+3}^-\Big).\tag{2.23}
\end{align*}
For $|\mu|\leq\la-2$, collecting (2.21), (2.23) yields
$$
\ds\sum_{j=1}^{13}|\tilde f_j^\mu(t)|\leq C\Big(\ds\f{\ve}{\langle t\rangle^{2}}{\widetilde Q}_{|\mu|+3}^-(t)
+\f{\ve^2}{\langle t\rangle^{5/2}}+\ve M_\mu(t)\Big),\eqno{(2.24)}
$$
which together with (2.19) means (1) holds.
For $\la-1\leq |\mu|\leq 2\la-4$, due to $E^{1/2}_{|\nu|+2}(t)m_{\mu-\nu}(t)\leq\ve M_\mu(t)$ if $|\nu|\leq\la-2$ and $E^{1/2}_{|\nu|+2}(t)m_{\mu-\nu}(t)\leq C\ve\langle t\rangle^{-3/2}E_{|\nu|+2}^{1/2}(t)$ otherwise, then by (2.22), (2.23)
$$
\sum_{j=1}^{13}|\tilde f_j^\mu(t)|\leq C\Big(\frac{\ve}{\langle t\rangle^2}\widetilde Q_{|\mu|+3}^-(t)+(\ve+\eta)M_\mu(t)+\frac{\ve^2}{\langle t\rangle^{5/2}}+\frac{\ve}{\langle t\rangle^{3/2}}E_{|\mu|+2}^{1/2}(t)\Big).
$$
This, together with (2.19), also yields (1).
(2) If $t$ is small, the estimate can be obtained by Lemma 2.3 easily.
Otherwise, (2) follows from (2.11) and (1).
(3) If $\mu_2+\mu_3\neq0$, as shown in (2.20), then we have for $|\mu|\leq 2\la-3$
$$
|\Gamma^\mu(\chi\dot w)(t)|\leq C\sum_{|\al|\leq \mu_2+\mu_3-1}|\nabla\cdot\pa_x^\al\pa_t^{\mu_1}X^{\mu_4}(\chi\dot w)(t)|\leq C\chi_{|\mu|-1}(t).
$$
(iv) If $\mu_2+\mu_3=0$, then $\Gamma^\mu\dot w(t,x)=\bigg(\ds\frac{1}{r}\int_0^rs(\nabla\cdot\pa_t^{\mu_1}X^{\mu_4}\dot w)(t,s){d s}\bigg)\ds\frac{x}{r}$ and (4) follows (1) immediately. $\square$
\bigskip
When $\ve>0$ is small enough, as in [8], we next show that (1.3) is isentropic outside the fixed ball $B(0, M+1)$
for any time $t$.
{\bf Lemma 2.6.} {\it For small $\ve>0$, if $(\dot\theta, \dot w, \dot z)$ is a smooth solution to (2.4)
for $(t,x)\in [0, T]\times\Bbb R^2$ and $E_\la(t)\leq\ve^2$ with $\la\ge 4$, then $\dot z(t,x)=0$ for $|x|\geq M+1$.
}
{\it Proof.}
From the third equation in (2.3), we know that $\pa_tz+W(t,r)\pa_rz=0$, where $W(t,r)=w(t,r)\cdot\ds\frac{x}{r}$.
Define the characteristics $r=r(t)$ by $r'(t)=W(t, r(t))$ with $r(0)=M$, then it is easy to see that $z(t,x)=0$
for $|x|\geq r(t)$ by the compact support property of $z(0,x)$.
By Lemma 2.2 (1) and Lemma 2.3 (9) together with Lemma 2.4, we have $|w(t)|_-\leq |w_a(t)|_-+|\dot w(t)|_-
\le C\ve\langle t\rangle^{-3/2}+C\langle t\rangle^{-1/2}\widetilde Q_2^-(t)\leq C\ve\langle t\rangle^{-1/2}$.
In addition, it follows from Lemma 2.3 (2) that $|w(t, x)|\le |w_a(t,x)|+|\dot w(t,x)|\leq C\langle t\rangle^{-1/2}(\ve+E_2^{1/2}(t))\leq C\ve\langle t\rangle^{-1/2}$ holds for $|x|\geq \ds\frac{t}{2}+M+1$. Therefore,
$|r(t)|\leq (M+1)\langle t\rangle^{1/2}$. From this, we can take use of Lemma 2.2 (3) and Lemma 2.5 (4)
to get $|W(t, r(t))|\leq C\ve\langle t\rangle^{-1}$, and hence $|r(t)|\leq (M+1)\langle t\rangle^{\delta}$ for any $0<\delta<\frac{1}{2}$. Applying Lemma 2.2 (3) and Lemma 2.5 (4) again, we have $|r'(t)|\leq C\ve\langle t\rangle^{-3/2+\delta}$
and further $|r(t)-M|\leq C\ve$. Consequently, $z(t, r)\equiv0$ as $|x|\geq M+1$, so does $\dot z$.
\qquad \qquad $\square$\bigskip
\vskip 0.4 true cm \centerline{\bf $\S 3$. Analysis on $(\dot\th, \dot w, \dot z)$ near light cone
and establishment of a crucial energy estimate} \vskip 0.4 true cm
From Lemma 2.6, we know that $z(t,x)\equiv0$ and then $S\equiv\bar S$ holds when $|x|\geq M+1$.
Assume that $(\dot\theta, \dot w, \dot z)$ is a smooth solution to (2.4) for $0\leq t\leq T$, and let
$\phi\in C^\infty(\{(t,x)\in [0, T]\times\Bbb R^2: |x|\geq M+1\})$ be the corresponding potential
vanishing in $|x|\geq M+t$ and satisfying $\nabla\phi=u$, $\pa_t\phi=-\frac{1}{2}|u|^2-h(\rho)$.
For $(\lambda,y)$,
$(\tilde\lambda,\tilde y)\in\Bbb R\times\Bbb R^2$, we define
\begin{align*}
&F_1(\lambda,y)=\frac{1}{2}(\lambda^2-|y|^2),\\
&F_2((\lambda,y),(\tilde\lambda,\tilde y))=\frac{1}{2}(\lambda\tilde\lambda-y\cdot\tilde y).
\end{align*}
Let $\dot\phi=\phi-\phi_a$, $\xi=(\theta, w)$, $\xi_a=(\theta_a, w_a)$ and $\dot\xi=\xi-\xi_a=(\dot\theta, \dot w)$,
where $\phi_a$ and $(\theta_a, w_a)$ are given in (2.1) and Lemma 2.2 respectively. Then we have
\begin{align*}
&\theta_a(t,x)=-\pa_t\phi_a(t,x)+F_1(\xi_a)(t,x),\\
&\theta(t,x)=-\pa_t\phi(t,x)+F_1(\xi)(t,x)\qquad\text{for}\quad|x|\geq M+1,\\
&\dot\theta(t,x)=-\pa_t\dot\phi(t,x)+F_2(\dot\xi, 2\xi_a+\dot\xi)(t,x)\qquad\text{for}\quad|x|\geq M+1.
\end{align*}
In order to make use of both null conditions introduced in [4] for the second order quasilinear
wave equations, we will pay more attention to the forms of $F_1(\xi_a)$ and $F_2(\dot\xi, 2\xi_a+\dot\xi)$, that is, if $|x|\geq M+1$, then
\begin{align*}
&F_1(\xi_a)=\frac{1}{2}(\pa_t\phi_a)^2-\frac{1}{2}|\nabla\phi_a|^2+F_1(\xi_a)\big(\theta_a-\frac{1}{2}F_1(\xi_a)\big),\tag{3.1}\\
&F_2(\dot\xi, 2\xi_a+\dot\xi)=\frac{1}{2}\pa_t\dot\phi(2\pa_t\phi_a+\pa_t\dot\phi)-\frac{1}{2}\nabla\dot\phi
\cdot(2\nabla\phi_a+\nabla\dot\phi)+\frac{1}{2}F_2(\dot\xi,2\xi_a+\dot\xi)(2\theta_a+\dot\theta)\\
&\qquad \qquad \qquad\quad -\frac{1}{2}\big(F_2(\dot\xi,2\xi_a+\dot\xi)
-\dot\theta\big)\big(2F_1(\xi_a)+F_2(\dot\xi,2\xi_a+\dot\xi)\big).\tag{3.2}
\end{align*}
Next we cite two fundamental estimates on the first and second null conditions which are shown in [4], Lemma 6.64-Lemma 6.65
of [11] or [13]
(one can see Lemma 3.1-Lemma 3.5 of [13] for some details).
{\bf Lemma 3.1.} {\it
Assume that $g_{i}^{jk}\in\Bbb R$ with $g_i^{jk}=g_i^{kj}$ $(0\le i, j, k\le 2)$ and $\ds\sum_{i,j,k=0}^2g_i^{jk}p_ip_jp_k=0$ for any $p=(p_0,p_1,p_2)\in\Bbb R^3$ satisfying $p_0^2=p_1^2+p_2^2$. Then, there exists a positive constant $C$ such that for any $f, g\in C^\infty(
[0,T]\times\Bbb R^2)$,
$$
|\ds\sum_{i,j,k=0}^2g_i^{jk}(\pa_i f\pa_{jk}^2g)(t, x)|\leq C\big(|Zf(t,x)||\pa^2g(t,x)|
+|\pa f(t,x)||Z\pa g(t,x)|\big)\eqno{(3.3)}
$$
and
\begin{align*}
\sum_{|\al|\le n}|\Gamma^{\al}(\ds\sum_{i,j,k=0}^2g_i^{jk}\p_i f\p_{jk}^2f)(t,x)&
-\sum_{i,j,k=0}^2g_i^{jk}(\p_i f\p_{jk}^2\Gamma^{\al} f)(t,x)|\\
&\le C_n\ds\sum_{|\beta+\nu|\le n+1,\quad |\beta|,|\nu|\le n}|Z\Gamma^\beta f(t,x)||\Gamma^{\nu}\p f(t,x)|.\tag{3.4}\\
\end{align*}
}
{\bf Lemma 3.2.} {\it
Assume that $g_{ij}^{kl}\in\Bbb R$ with $g_{ij}^{kl}=g_{ij}^{lk}$ and $g_{ij}^{kl}=g_{ji}^{kl}$ ($0\le i, j, k, l\le 2$),
and $\ds\sum_{i,j,k,l=0}^2g_{ij}^{kl}p_ip_jp_kp_l$ $=0$ for any $p=(p_0,p_1,p_2)\in\Bbb R^3$ satisfying $p_0^2=p_1^2+p_2^2$.
Then, there exists a positive constant $C$ such that for any $f,g,h\in C^\infty([0, T]\times\Bbb R^2)$,
\begin{align*}
&|\ds\sum_{i,j,k,l=0}^2g_{ij}^{kl}(\pa_if\pa_jg\pa_{kl}^2h)(t,x)|\leq C\big(|\pa f(t,x)||\pa g(t,x)||Z\pa h(t,x)|\\
&\qquad +|\pa f(t,x)||Z g(t,x)||\pa^2h(t,x)|+|Zf(t,x)||\pa g(t,x)||\pa^2h(t,x)|\big).\tag{3.5}
\end{align*}
}
\bigskip
Based on Lemma 3.1 and Lemma 3.2, and noting $
|Z\varphi(t,x)|\leq C\ds\Big(\frac{||x|-t|}{t}|\pa\varphi(t,x)|+\frac{1}{t}(|X\varphi(t,x)|$ $+|\Omega\varphi(t,x)|)\Big)
$, then one can obtain the following estimates under the assumptions of Lemma 3.1 and Lemma 3.2:
\begin{align*}
&|\ds\sum_{i,j,k=0}^2g_i^{jk}(\pa_i\Gamma^{\al}\phi_a\pa_{jk}^2\Gamma^{\beta}\dot\phi)(t,x)|\le\frac{C\ve}{t\langle t\rangle^{1/2}}\sum_{|\nu|\le |\beta|+1}|\Gamma^{\nu}\pa\dot\phi(t,x)|,\tag{3.6}\\
&|\ds\sum_{i,j,k=0}^2g_i^{jk}(\pa_i\Gamma^\al\dot\phi\pa_{jk}^2\Gamma^\beta\phi_a)(t,x)|\le\frac{C\ve}{t\langle t\rangle^{1/2}}\bigg(\sum_{|\nu|\le |\al|+1}|(\si_-^{-1}\Gamma^\nu\dot\phi)(t,x)|
+\sum_{|\nu|\le |\al|}|\Gamma^\nu\pa\dot\phi(t,x)|\bigg),\tag{3.7}\\
&|\ds\sum_{i,j,k,l=0}^2g_{ij}^{kl}(\pa_i\Gamma^\a\phi_a\pa_j\Gamma^\beta\phi_a\pa_{kl}^2\Gamma^\mu \dot\phi)(t,x)|\le\frac{C\ve^2}{t\langle t\rangle}\sum_{|\nu|\le |\mu|+1}|\Gamma^\nu\pa\dot\phi(t,x)|,\tag{3.8}\\
&|\ds\sum_{i,j,k,l=0}^2g_{ij}^{kl}(\pa_i\Gamma^\a\phi_a\pa_j\Gamma^\beta \dot\phi\pa_{kl}^2\Gamma^\mu \phi_a)(t,x)|\le\frac{C\ve^2}{t\langle t\rangle}
\bigg(\ds\sum_{|\nu|\le |\beta|+1}|(\si_-^{-1}\Gamma^\nu\dot\phi)(t,x)|\\
&\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad +\ds\sum_{|\nu|\le |\beta|}|\Gamma^\nu\pa\dot\phi(t,x)|\bigg)\tag{3.9}\\
\end{align*}
and
\begin{align*}
|\ds\sum_{i,j,k=0}^2&g_i^{jk}(\pa_i\Gamma^\al\dot\phi\pa_{jk}^2\Gamma^\beta\dot\phi)(t,x)|\le\frac{C}{t}\big(\sum_{|p|\le |\al|,\quad |q|\le|\beta|}|(\si_-\Gamma^{p}\pa\dot\phi\pa\Gamma^q\pa\dot\phi)(t,x)|\\
&+\sum_{|p|\le|\al|+1,\quad |q|\le |\beta|}|(\Gamma^p\dot\phi\pa\Gamma^q\pa \dot\phi)(t,x)|+\sum_{|p|\leq|\al|,\quad
|q|\le |\beta|+1}|(\Gamma^p \pa\dot\phi\Gamma^q\pa\dot\phi)(t,x)|\big),\tag{3.10}\\
|\ds\sum_{i,j,k,l=0}^2&g_{ij}^{kl}(\pa_i\Gamma^\a\dot\phi\pa_j\Gamma^\beta \phi_a\pa_{kl}^2\Gamma^\mu\dot\phi)(t,x)|\\
&\leq\frac{C\ve}{t\langle t\rangle^{1/2}}\big(\sum_{|p|\le |\a|,\quad |q|\le |\mu|+1}|(\Gamma^p\pa\dot\phi\Gamma^q\pa \dot\phi)(t,x)|
+\ds\sum_{ |p|\le |\a|+1\quad |q|\le |\mu|}|(\si_-^{-1}\Gamma^p\dot\phi\pa\Gamma^q\pa\dot\phi)(t,x)|\big),\tag{3.11}\\
\end{align*}
\begin{align*}
|\ds\sum_{i,j,k,l=0}^2&g_{ij}^{kl}(\pa_i\Gamma^\a\dot\phi\pa_j\Gamma^\beta\dot\phi\pa_{kl}^2\Gamma^\mu\phi_a)(t,x)|\\
&\leq\frac{C\ve}{t\langle t\rangle^{1/2}}\Big(\sum_{|p|\leq|\beta|+1,\quad |q|\leq|\a|}|(\si_-^{-1}\Gamma^p\dot\phi\Gamma^q\pa \dot\phi)(t,x)|+\ds\sum_{|p|\leq|\beta|,\quad |q|\leq|\a |+1}|(\si_-^{-1}\Gamma^q\dot\phi\Gamma^p\pa\dot\phi)(t,x)|\Big),\tag{3.12}\\
|\sum_{i,j,k,l=0}^2&g_{ij}^{kl}(\pa_i\Gamma^\a\dot\phi\pa_j\Gamma^\beta\dot\phi\pa_{kl}^2\Gamma^\mu\dot\phi)(t,x)|\\
&\leq\frac{C}{t}\Big(\sum_{|p|\leq|\a|,\quad |q|\leq|\beta|,|s|\leq|\mu|}|(\si_-\Gamma^p\pa\dot\phi\Gamma^q\pa \dot\phi\pa\Gamma^s\pa\dot\phi)(t,x)|\\
&+\sum_{|s|\leq|\mu|+1,\quad |q|\leq|\beta|,|p|\leq|\a|}|(\Gamma^p\pa\dot\phi\Gamma^q\pa\dot\phi\Gamma^s\pa\dot\phi)(t,x)|\\
\end{align*}
\begin{align*}
&+\sum_{|q|\leq|\beta|+1,\quad |s|\leq|\mu|,|p|\leq|\a|}|(\Gamma^p\pa \dot\phi\Gamma^q\dot\phi\pa\Gamma^s\pa\dot\phi)(t,x)|\\
&+\sum_{|p|\leq|\a|+1,\quad |q|\leq|\beta|,|s|\leq|\mu|}|(\Gamma^p\dot\phi\Gamma^q\pa\dot\phi\pa\Gamma^s\pa \dot\phi)(t,x)|\Big).\tag{3.13}
\end{align*}
To treat each term better in the right hand side of (3.6)-(3.13), we require the following two results
(Lemma 3.3 and Lemma 3.4):
{\bf Lemma 3.3.} {\it
Suppose that $f(t,x)\in C^1(\Bbb R_+\times\Bbb R^2)$ and $\text{supp} f\subset\{(t,x):|x|\leq M+t\}$.
Then there exists a positive constant $C$ independent of $t$ such that
$$
\|(\si_-^{-1}f)(t,\cdot)\|_{L^2(|x|\geq M+1)}\leq C\|\nabla f(t,\cdot)\|_{L^2(|x|\geq M+1)}.\eqno{(3.14)}
$$
}
{\bf Remark 3.1.} {\it Here we point out that the inequality $\|(\si_-^{-1}f)(t,\cdot)\|_{L^2}\leq C\|\nabla f(t,\cdot)\|_{L^2}$
in the whole space has been established in [15].}
\bigskip
{\bf Proof.}
Due to $\text{supp} f\subset\{|x|\leq M+t\}$, one then has
$$
f(t,x)=-\int_{|x|}^{M+t}\nabla f(t, s\omega)\cdot\omega ds.
$$
This derives
$$
|f(t,x)|^2\leq\bigg(\int_{|x|}^{M+t}|\nabla f(t, s\omega)|^2(1+|t-s|)^{1/2} ds\bigg)
\int_{|x|}^{M+t}(1+|t-s|)^{-1/2} ds.
$$
With easy computations, we have $\int_{|x|}^{M+t}(1+|t-s|)^{-1/2} ds\leq C(1+t-|x|)^{1/2}$ for $t>|x|$, and $\int_{|x|}^{M+t}(1+|t-s|)^{-1/2} ds\leq C$ for $t\leq|x|$. Hence, if $t\geq M+1$, then
\begin{align*}
&\int_{M+1}^{M+t}|f(t, r\omega)|^2(1+|t-r|)^{-2}r dr\\
&\leq C\int_{M+1}^t\Big(\int_r^t|\nabla f(t, s\omega)|^2(1+t-s)^{1/2} ds\Big)(1+t-r)^{-3/2}r dr\\
&\quad +C\int_{M+1}^t\Big(\int_t^{M+t}|\nabla f(t, s\omega)|^2 ds\Big)(1+t-r)^{-3/2}r dr\\
&\quad +C\int_t^{M+t}\Big(\int_r^{M+t}|\nabla f(t, s\omega)|^2 ds\Big)r(1+r-t)^{-2} dr\\
&\leq C\int_{M+1}^{M+t}s|\nabla f(t, s\omega)|^2 ds.\tag{3.15}
\end{align*}
If $1\leq t\leq M+1$, then
\begin{align*}
&\int_{M+1}^{M+t}|f(t,r\omega)|^2(1+|t-r|)^{-2}r dr\leq C\int_{M+1}^{M+t}\int_r^{M+t}|\nabla f(t,s\omega)|^2r ds dr\\
&\leq C\int_{M+1}^{M+t}s|\nabla f(t, s\omega)|^2 ds.\tag{3.16}
\end{align*}
Combining (3.15) with (3.16) yields (3.14). \qquad\qquad
\qquad\qquad\qquad\qquad\qquad\qquad\qquad$\square$\bigskip
{\bf Lemma 3.4.} {\it
Suppose that $f(t,x)\in C^2(\Bbb R_+\times\Bbb R^2)$ and $\text{supp} f\subset\{(t, x):|x|\leq M+t\}$. Then there exists a positive constant $C$ independent of $t$ such that for $|x|\geq M+1$
$$
\si(x)^{1/2}\si_-(t,x)^{-1}|f(t,x)|\leq C\sum_{j=0}^1\sum_{|\al|=0}^{2-j}\|\pa\nabla^\al\Omega^jf(t,\cdot)\|_{L^2(|x|\geq M+1)}.\eqno{(3.17)}
$$
}
\bigskip
{\bf Proof.}
Similar to the proof of Lemma 1 (3.1) in [21], we have for $|x|\geq M+1$ and $g(t,x)\in C^2(\Bbb R_+\times\Bbb R^2)$
with $\text{supp} g\subset\{(t, x):|x|\leq M+t\}$
$$
\si(x)^{1/2}|g(t,x)|\leq C\sum_{j=0}^1\sum_{|\al|=0}^{2-j}\|\nabla^\al\Omega^jg(t,\cdot)\|_{L^2(|x|\geq M+1)}.\eqno{(3.18)}
$$
Choosing $g(t,x)=\si_-(t,x)^{-1}f(t,x)$ in (3.18) and applying Lemma 3.3 yield (3.17) directly.
$\square$\bigskip
Next, we extend Lemma 2.4 so that an energy estimate in the whole space is established as in Proposition 4.1 of [9].
{\bf Lemma 3.5.} {\it
For fixed $n\in\Bbb N$ with $n\geq 5$, if $(\dot\theta, \dot w, \dot z)$ is a smooth solution of (2.4)
for $(t, x)\in [0, T]\times \Bbb R^2$, and $E_{[\frac{n+5}{2}]}^{1/2}(t)\leq\eta$ holds for small positive constant $\eta$,
then $\ds Q_n(t)\leq C(E_n^{1/2}(t)+\frac{\ve^2}{\langle t\rangle^{3/2}})$ for $0\leq t\leq T$.
}
\bigskip
{\bf Proof.} If $t\leq 1$, it is easy to see $Q_n(t)\leq C E_n^{1/2}(t)$. We now focus on the case of $t\geq 1$.
It is noted that for $|\nu|\leq n$
$$
\|\Gamma^\nu\pa \dot\phi(t)\|_+\leq C(\|\Gamma^\nu\dot\theta(t)\|_++\|\Gamma^\nu\dot w(t)\|_+)\leq C E^{1/2}_n(t),\eqno{(3.19)}
$$
here Lemma 2.3 (2) is applied in the first inequality.
Set $Q_n^+(t)=\ds\sum_{|\mu|\leq n-1}(\|\si_-(t)\nabla\Gamma^\mu\dot\theta(t)\|_++\|\si_-(t)\pa_t\Gamma^\mu\dot w(t)\|_++\|\si_-(t)\nabla\cdot\Gamma^\mu\dot w(t)\|_+)$. Similar to (2.12), the relationship between $E_j(t)$ and $Q_j^+(t)$
($1\le j\le n$) is
$$
Q_j^+(t)\leq C\big(E_j^{1/2}(t)+\sum_{|\mu|\leq j-1}t\|h_0^\mu(t)\|_++\sum_{|\mu|\leq j-1}\langle t\rangle\|h^\mu(t)\|_+\big).\eqno{(3.20)}
$$
When $|\mu|\leq j-1$, we have
$$
\|h_0^\mu(t)\|_+\leq\sum_{\nu\leq\mu}C_{\mu\nu}\big(\|I_1^{\mu\nu}(t)\|_++\|I_2^{\mu\nu}(t)\|_++\|I_3^{\mu\nu}(t)\|_+\big)
+\sum_{\nu\leq\mu}C_{\mu\nu}\|J_{\mu\nu}(t)\|_+,\eqno{(3.21)}
$$
where
\begin{align*}
I_1^{\mu\nu}(t,x)&=\sum_{i=1}^2\big((\Gamma^\nu\pa_i\phi_a)(\pa_i\Gamma^{\mu-\nu}\pa_t\dot\phi)
-(\Gamma^\nu\pa_t\phi_a)(\pa_i\Gamma^{\mu-\nu}\pa_i\dot\phi)\big)(t,x),\\
I_2^{\mu\nu}(t,x)&=\sum_{i=1}^2\big((\Gamma^\nu\pa_i\dot\phi)(\pa_i\Gamma^{\mu-\nu}\pa_t\phi_a)
-(\Gamma^\nu\pa_t\dot\phi)(\pa_i\Gamma^{\mu-\nu}\pa_i\phi_a)\big)(t,x),\\
I_3^{\mu\nu}(t,x)&=\sum_{i=1}^2\big((\Gamma^\nu\pa_i\dot\phi)(\pa_i\Gamma^{\mu-\nu}\pa_t\dot\phi)
-(\Gamma^\nu\pa_t\dot\phi)(\pa_i\Gamma^{\mu-\nu}\pa_i\dot\phi)\big)(t,x),
\end{align*}
and using (3.1)-(3.2) to get $J_{\mu\nu}(t,x)=\ds\sum_{j=1}^5\sum_{|\al|+|\beta|+|\gamma|\le\mu}C_{\al\beta\gamma}^{\nu}J^{\a\beta\gamma}_j(t,x)
+\tilde J_{\mu\nu}(t,x)$ with
\begin{align*}
J^{\a\beta\gamma}_1=\frac{1}{2}\sum_{k=1}^2&(\pa_t\Gamma^\a \phi_a)(\pa_t\Gamma^\beta \phi_a)(\pa_k^2\Gamma^\gamma \dot\phi)-\frac{1}{2}\sum_{k,l=1}^2(\pa_l\Gamma^\a \phi_a)(\pa_l\Gamma^\beta \phi_a)(\pa_k^2\Gamma^\gamma \dot\phi)\\
&-\sum_{k=1}^2(\pa_t\Gamma^\a \phi_a)(\pa_k\Gamma^\beta \phi_a)(\pa_k\pa_t\Gamma^\gamma \dot\phi)+\sum_{k,l=1}^2(\pa_l\Gamma^\a \phi_a)(\pa_k\Gamma^\beta \phi_a)(\pa_{kl}^2\Gamma^\gamma \dot\phi),
\end{align*}
\begin{align*}
J^{\a\beta\gamma}_2=-\sum_{k=1}^2&(\pa_t\Gamma^\beta \phi_a)(\pa_k\Gamma^\gamma\dot\phi)(\pa_k\pa_t\Gamma^\a \phi_a)+\sum_{k,l=1}^2(\pa_l\Gamma^\beta\phi_a)(\pa_k\Gamma^\gamma\dot\phi)(\pa_{kl}^2\Gamma^\a\phi_a)\\
&-\sum_{k=1}^2(\pa_k\Gamma^\beta\phi_a)(\pa_t\Gamma^\gamma\dot\phi)(\pa_k\pa_t\Gamma^\a\phi_a)+\sum_{k,l=1}^2(\pa_k\Gamma^\beta \phi_a)(\pa_l\Gamma^\gamma\dot\phi)(\pa_{kl}^2\Gamma^\a \phi_a)\\
&+\sum_{k=1}^2(\pa_t\Gamma^\beta\phi_a)(\pa_t\Gamma^\gamma\dot\phi)(\pa_k^2\Gamma^\a\phi_a)-\sum_{k,l=1}^2(\pa_k\Gamma^\beta \phi_a)(\pa_k\Gamma^\gamma\dot\phi)(\pa_l^2\Gamma^\a \phi_a),
\end{align*}
\begin{align*}
J^{\a\beta\gamma}_3=-\sum_{k=1}^2&(\pa_k\Gamma^\gamma\dot\phi)(\pa_t\Gamma^\beta\phi_a)(\pa_k\pa_t\Gamma^\a \dot\phi)+\sum_{k,l=1}^2(\pa_k\Gamma^\gamma\dot\phi)(\pa_l\Gamma^\beta\phi_a)(\pa_{kl}^2\Gamma^\a\dot\phi)\\
&-\sum_{k=1}^2(\pa_t\Gamma^\gamma\dot\phi)(\pa_k\Gamma^\beta\phi_a)(\pa_k\pa_t\Gamma^\a\dot\phi)+\sum_{k,l=1}^2(\pa_l \Gamma^\gamma \dot\phi)(\pa_k\Gamma^\beta\phi_a)(\pa_{kl}^2\Gamma^\a\dot\phi)\\
&+\sum_{k=1}^2(\pa_t\Gamma^\gamma\dot\phi)(\pa_t\Gamma^\beta\phi_a)(\pa_k^2\Gamma^\a \dot\phi)-\sum_{k,l=1}^2(\pa_k\Gamma^\gamma \dot\phi)(\pa_k\Gamma^\beta\phi_a)(\pa_l^2\Gamma^\a\dot\phi),
\end{align*}
\begin{align*}
J^{\a\beta\gamma}_4=-\sum_{k=1}^2&(\pa_k\Gamma^\gamma\dot\phi)(\pa_t\Gamma^\a\dot\phi)(\pa_k\pa_t\Gamma^\beta \phi_a)+\sum_{k,l=1}^2(\pa_k\Gamma^\gamma\dot\phi)(\pa_l\Gamma^\a\dot\phi)(\pa_{kl}^2\Gamma^\beta\phi_a)\\
&+\frac{1}{2}\sum_{k=1}^2(\pa_t\Gamma^\gamma\dot\phi)(\pa_t\Gamma^\a \dot\phi)(\pa_k^2\Gamma^\beta \phi_a)-\frac{1}{2}\sum_{k,l=1}^2(\pa_k\Gamma^\gamma\dot\phi)(\pa_k\Gamma^\a\dot\phi)(\pa_l^2\Gamma^\beta\phi_a),
\end{align*}
\begin{align*}
J^{\a\beta\gamma}_5=-\sum_{k=1}^2&(\pa_k\Gamma^\gamma\dot\phi)(\pa_t\Gamma^\beta\dot\phi)(\pa_k\pa_t\Gamma^\a \dot\phi)+\sum_{k,l=1}^2(\pa_k\Gamma^\gamma\dot\phi)(\pa_l\Gamma^\beta\dot\phi)(\pa_{kl}^2\Gamma^\a\dot\phi)\\
&+\frac{1}{2}\sum_{k=1}^2(\pa_t\Gamma^\gamma\dot\phi)(\pa_t\Gamma^\beta\dot\phi)(\pa_k^2\Gamma^\a \dot\phi)-\frac{1}{2}\sum_{k,l=1}^2(\pa_k\Gamma^\gamma\dot\phi)(\pa_k\Gamma^\beta\dot\phi)(\pa_l^2\Gamma^\a\dot\phi),
\end{align*}
and $\tilde J_{\mu\nu}$ is a higher order error term whose explicit expression is not needed.
In terms of Lemma 3.3 and (3.19), it is easy to find that
\begin{align*}
\|I_1^{\mu\nu}(t)\|_++\|I_2^{\mu\nu}(t)\|_+&\leq\frac{C\ve}{\langle t\rangle^{3/2}}E_j^{1/2}(t),\tag{3.22}\\
\|J_1^{\a\beta\gamma}(t)\|_++\|J_2^{\a\beta\gamma}(t)\|_+&\leq\frac{C\ve^2}{\langle t\rangle^2}E_j^{1/2}(t).\tag{3.23}
\end{align*}
To deal with $I_3^{\mu\nu}$, we should pay attention to (3.10). If $|\al|\leq |\beta|$ and $|\al+\beta|\leq j-1$,
then we can obtain with the help of Lemma 3.4
\begin{align*}
&\sum_{|p|\leq|\al|+1,\quad |q|\leq|\beta|}\|(\Gamma^p\dot\phi)(\pa\Gamma^q\pa\dot\phi)(t)\|_+\leq\sum_{ |p|\leq|\al|+1,\quad |q|\leq|\beta|}|\si_-^{-1}\Gamma^p\dot\phi(t)|_+\|\si_-\pa\Gamma^q\pa\dot\phi(t)\|_+\\
&\leq\frac{C}{\langle t\rangle^{1/2}}E_{[\frac{j+5}{2}]}^{1/2}(t)\widetilde Q_j(t).
\end{align*}
If $|\al|>|\beta|$, by using Lemma 2.3 and Lemma 3.3 we have
\begin{align*}
&\sum_{|p|\leq|\al|+1,\quad |q|\leq|\beta|}\|(\Gamma^p \dot\phi)(\pa\Gamma^q\pa\dot\phi)(t)\|_+\leq\sum_{ |p|\leq|\al|+1,\quad |q|\leq|\beta|}|\si_-\pa\Gamma^q\pa \dot\phi(t)|_+\|\si_-^{-1}\Gamma^p \dot\phi(t)\|_+\\
&\leq\frac{C}{\langle t\rangle^{1/2}}\widetilde Q_{[\frac{j}{2}]+2}(t)E_{j}^{1/2}(t).
\end{align*}
Similarly, we can arrive at for $|\al+\beta|\leq j-1$
\begin{align*}
\sum_{|p|\leq|\al|,\quad |q|\leq|\beta|}\|\si_-(\Gamma^{p}\pa \dot\phi)(\pa\Gamma^q\pa \dot\phi)(t)\|_+&+\sum_{ |p|\leq|\al|,\quad |q|\leq|\beta|+1}\|(\Gamma^p \pa\dot\phi)(\Gamma^q\pa \dot\phi)(t)\|_+\\
\leq\frac{C}{\langle t\rangle^{1/2}}&\Big(E^{1/2}_{[\frac{j+3}{2}]}\widetilde Q_j(t)+\widetilde Q_{[\frac{j}{2}]+2}(t)E_{j-1}^{1/2}(t)\Big),
\end{align*}
$$
\sum_{|p|\leq |\beta|+1,\quad |q|\leq |\al|}\|\si_-^{-1}\Gamma^p\dot\phi\Gamma^q\pa \dot\phi(t)\|_+\leq\frac{C}{\langle t\rangle^{1/2}}\Big(E^{1/2}_{[\frac{j}{2}]+2}(t)E^{1/2}_{j-1}(t)+E^{1/2}_{[\frac{j+3}{2}]}(t)E_j^{1/2}(t)\Big).
$$
Therefore,
\begin{align*}
\|I_3^{\mu\nu}(t)\|_+&\leq\frac{C}{\langle t\rangle^{3/2}}\Big(E^{1/2}_{[\frac{j+5}{2}]}(t)\widetilde Q_j(t)+\widetilde Q_{[\frac{j}{2}]+2}(t)E_j^{1/2}(t)\Big),\tag{3.24}\\
\|J_3^{\a\beta\gamma}(t)\|_+&\leq\frac{C\ve}{\langle t\rangle^2}\Big(E^{1/2}_{[\frac{j+3}{2}]}(t)\widetilde Q_j(t)+\widetilde Q_{[\frac{j}{2}]+2}(t)E_{j-1}^{1/2}(t)\Big),\tag{3.25}\\
\|J_4^{\a\beta\gamma}(t)\|_+&\leq\frac{C\ve}{\langle t\rangle^2}\Big(E^{1/2}_{[\frac{j}{2}]+2}(t)E^{1/2}_{j-1}(t)+ E^{1/2}_{[\frac{j+3}{2}]}(t)E_j^{1/2}(t)\Big).\tag{3.26}
\end{align*}
Analogously, we can also get the estimate of $J_5^{\a\beta\gamma}$ by means of
comparing the size of $|p|$, $|q|$ and $|s|$ in (3.13), that is,
\begin{align*}
&\|J_5^{\a\beta\gamma}(t)\|_+\\
&\leq\frac{C}{\langle t\rangle^2}\Big(\widetilde Q_{[\frac{j+5}{2}]}(t)E^{1/2}_{[\frac{j+3}{2}]}(t)E_{j-1}^{1/2}(t)+E_{[\frac{j}{2}]+1}(t)\widetilde Q_j(t)+E_{[\frac{j+5}{2}]}E^{1/2}_{j-1}(t)+E_{[\frac{j}{2}]+1}(t)E_j^{1/2}(t)\\
&\qquad +E_{[\frac{j+5}{2}]}^{1/2}(t)\widetilde Q_{[\frac{j+5}{2}]}(t)E^{1/2}_{j-1}(t)+E_{[\frac{j+3}{2}]}^{1/2}(t)\widetilde Q_{[\frac{j+5}{2}]}(t)E^{1/2}_{j}(t)+E^{1/2}_{[\frac{j}{2}]+1}(t)E^{1/2}_{[\frac{j}{2}]+2}(t)\widetilde Q_j(t)\Big).\tag{3.27}
\end{align*}
In addition, it is noted that $\tilde J_{\mu\nu}$ only contains the high-order error terms, then by a direct verification
one can derive that the $L^2$ norm of $\tilde J_{\mu\nu}$ near the light cone can be controlled by those terms in the right hand sides of (3.22)-(3.27).
Due to $Q_n^-(t)\leq C\big(E_n^{1/2}(t)+\ds\frac{\ve^2}{\langle t\rangle^{3/2}}\big)$ in terms of Lemma 2.4,
we apply the estimates (3.21)-(3.27) to obtain $\|h_0^\mu(t)\|_+\leq \ds\frac{C}{\langle t\rangle^{3/2}}\eta(\eta+\frac{\ve^2}{\langle t\rangle^{3/2}})+\frac{C\eta}{\langle t\rangle^{3/2}}Q^+_{[\frac{n+5}{2}]}(t)$ when $|\mu|\leq [\ds\frac{n+5}{2}]-1$ and $\eta$ is small enough. In addition, because of $z(t,x)\equiv0$ for $|x|\geq M+1$, then we can get the same result for $h^\mu$ by the analogous analysis to $h_0^{\mu}$, that is, $\|h^\mu(t)\|_+\leq \ds\frac{C}{\langle t\rangle^{3/2}}\eta(\eta+\frac{\ve^2}{\langle t\rangle^{3/2}})+\frac{C\eta}{\langle t\rangle^{3/2}}Q^+_{[\frac{n+5}{2}]}(t)$ for $|\mu|\leq [\ds\frac{n+5}{2}]-1$. Hence,
$Q^+_{[\frac{n+5}{2}]}(t)\leq C\eta$ can be derived by utilizing (3.20).
Taking $j=n$ in the estimates in (3.22)-(3.27) and applying (3.20) again, we obtain
$$
Q^+_{n}(t)\leq C\big(E_n^{1/2}(t)+\frac{\ve^2}{\langle t\rangle^{5/2}}\big).
$$
This, together with Lemma 2.4, yields Lemma 3.5.
\qquad \qquad $\square$
\bigskip
\vskip 0.4 true cm \centerline{\bf $\S 4$. Global energy estimates and proof of Theorem 1.1} \vskip 0.4 true cm
From (2.4), as in [9], we have
\begin{align*}
&(\pa_t+w\cdot\nabla)\Gamma^\mu\dot\theta+(1-\theta)\nabla\cdot\Gamma^\mu\dot w=\hat h_0^\mu\equiv
\ds\sum_{j\in\{2,5\}}f_j^\mu+\sum_{j\in\{1,3,4,6\}}\hat f_j^\mu,\tag{4.1}\\
&\ds(\frac{1}{1+z}\pa_t+\frac{w}{1+z}\cdot\nabla)\Gamma^\mu\dot w+(1-\theta)\nabla\Gamma^\mu\dot\theta=
\frac{\hat h^\mu}{1+z}\equiv\ds\f{1}{1+z}\bigg(\sum_{j\in\{8,11\}}f_j^\mu+\sum_{j\in\{7,9,10,12,13\}}\hat f_j^\mu\bigg),\tag{4.2}\\
&(\pa_t+w\cdot\nabla)\Gamma^\mu\dot z=\hat g^\mu,\tag{4.3}
\end{align*}
where $f_j^{\mu} (j=2, 5, 8, 11)$ have been defined in (2.6); if $j\not=13$, then $\hat f_j^0=0$, and $\hat f_j^\mu$ with $\mu\not=0$ are defined as $f_j^\mu$ but with the supplementary restriction condition $\nu\not=0$ in the sum; if $j=13$, then
$\hat f_j^\mu=f_j^\mu+(1-\theta)z\nabla\Gamma^\mu\dot\theta$. In addition, $\hat g^\mu=\ds\sum_{j\in\{1,2\}}\hat g_{aj}^\mu+\ds\sum_{j\in\{1,2\}}\hat g_{j}^\mu$ with $\hat g_{a1}^\mu=-\ds\sum_{\nu\leq \mu}\binom{\mu}{\nu}\Gamma^\nu w_a\cdot\nabla\Gamma^{\mu-\nu}z(0,x)$,
$ \hat g_{a2}^\mu =-\ds\sum_{0<\nu\leq \mu}\binom{\mu}{\nu}\Gamma^\nu w_a\cdot\nabla\Gamma^{\mu-\nu}\dot z$
and $\hat g_{1}^\mu=-\ds\sum_{\nu\leq \mu}\binom{\mu}{\nu}\Gamma^\nu \dot w\cdot\nabla\Gamma^{\mu-\nu}z(0,x)$,
$\hat g_{2}^\mu =-\ds\sum_{0<\nu\leq \mu}\binom{\mu}{\nu}\Gamma^\nu \dot w\cdot\nabla\Gamma^{\mu-\nu}\dot z$
if $\mu\neq0$, otherwise, $\hat g_{a2}^\mu= \hat g_{2}^\mu=0$ for $\mu=0$.
Set
$$\zeta^\mu=
\left(
\begin{array}{ccc}
\Gamma^\mu\dot\theta\\
\Gamma^\mu\dot w\\
\Gamma^\mu\dot z
\end{array}
\right), \quad
F^\mu= \left(
\begin{array}{ccc}
\hat h_0^\mu\\
\ds\frac{\hat h^\mu}{1+z}\\
\hat g^\mu
\end{array}
\right), \quad
A_0= \left(
\begin{array}{cccc} 1 & 0 & 0 & 0 \\ 0
& \ds\frac{1}{1+z} & 0 & 0 \\
0 & 0 & \ds\frac{1}{1+z} & 0 \\
0 & 0 & 0 & 1
\end{array}
\right),
$$
$$
A_1= \left(
\begin{array}{cccc}
w_1 & 1-\theta & 0 & 0 \\
1-\theta & \ds\frac{w_1}{1+z} & 0 & 0 \\
0 & 0 & \ds\frac{w_1}{1+z} & 0 \\
0 & 0 & 0 & w_1
\end{array}
\right),\quad
A_2= \left(
\begin{array}{cccc}
w_2 & 0 & 1-\theta & 0 \\
0 & \ds\frac{w_2}{1+z} & 0 & 0 \\
1-\theta & 0 &\ds \frac{w_2}{1+z} & 0 \\
0 & 0 & 0 & w_2
\end{array}
\right),
$$
where $w=(w_1, w_2)$.
In this case, (4.1)-(4.3) can be written as $A_0\pa_t\zeta^\mu+A_1\pa_1\zeta^\mu+A_2\pa_2\zeta^\mu=F^\mu$.
Taking inner product with $\zeta^\mu$ in the space $\Bbb R^2$ and applying integration by parts, we arrive at
$$
\frac{ d}{d t}\langle A_0\zeta^\mu,\zeta^\mu\rangle(t)=2\langle F^\mu,\zeta^\mu\rangle(t)+\sum_{0\leq j\leq2}\langle(\pa_jA_j)\zeta^\mu,\zeta^\mu\rangle(t).\eqno{(4.4)}
$$
As in [9], we set $H_j^\mu=\langle f_j^\mu,\Gamma^\mu\dot\theta\rangle$ if $j=2,5$, $H_j^\mu=\langle\hat f_j^\mu,\Gamma^\mu\dot\theta\rangle$ if $j=1,3,4,6$,
$H_j^\mu=\langle\ds\frac{f_j^\mu}{1+z},\Gamma^\mu\dot w\rangle$ if $j=8,11$, $H_j^\mu=\langle\ds\frac{\hat f_j^\mu}{1+z},\Gamma^\mu\dot w\rangle$ if $j=7,9,10,12,13$,
$\hat H_{aj}^\mu=\langle\hat g_{aj}^\mu,\Gamma^\mu\dot z\rangle$ and $\hat H_{j}^\mu
=\langle\hat g_{j}^\mu,\Gamma^\mu\dot z\rangle$ for $j=1,2$. Then
$$
\langle F^\mu,\zeta^\mu\rangle=\sum_{1\leq j\leq 13}H_j^\mu+\sum_{j=1,2}\hat H_{aj}^\mu
+\sum_{j=1,2}\hat H_{j}^\mu.\eqno{(4.5)}
$$
Although the local existence of solution to (2.4) has been early established (for example, one can see [18]),
we still give some detailed illustrations for later uses.
{\bf Lemma 4.1.} {\it
If $\ve>0$ is small, then (2.4) has a unique smooth solution $(\dot\theta, \dot w, \dot z)$ for $t\leq2/\ve$ and $E_m^{1/2}(t)\leq C\ve^2$.
}
\bigskip
{\bf Proof.} Define
$$
T_1=\sup\{T>0: E_m^{1/2}(t)\leq \ve,0<t\leq 2/\ve\}\quad\text{for}\quad m\geq5.
$$
It is noted that $|z(t,x)|\le C\ve$ holds for $t\in [0, T_1]$ by Lemma 2.3 (2). This, together with
Lemma 2.3 and Lemma 3.5, yields
$$
|\sum_{j=0}^2\pa_jA_j(t,x)|\leq C\big(|\nabla\theta(t,x)|+|\nabla\cdot w(t,x)|
+|w\cdot\nabla z(t,x)|\big)\leq\frac{C}{\langle t\rangle^{1/2}}(\ve+\widetilde Q_3(t))\leq\frac{ C\ve}{\langle t\rangle^{1/2}}.
$$
In addition, a direct computation yields
$$
|H_j^\mu(t)|\leq\frac{C\ve}{\langle t\rangle^{1/2}}E_m(t)\quad\text{for}\quad|\mu|\leq m\quad\text{and}\quad j\in\{1,2,4,5,7,8,10,11\}.\eqno{(4.6)}
$$
With respect to $H_3^\mu(t)$ for $|\mu|\leq m$, we will treat it under two kinds of cases:
If $|\nu|\leq |\mu-\nu|\leq m-1$, then by Lemma 2.3 and Lemma 3.5
$$
|\langle\Gamma^\nu\dot w\cdot\nabla\Gamma^{\mu-\nu}\dot\theta,\Gamma^\mu\dot\theta\rangle(t)|
\leq\frac{C}{\langle t\rangle^{1/2}}|\langle\si^{1/2}\Gamma^\nu\dot w\cdot\si_-\nabla\Gamma^{\mu-\nu}\dot\theta,\Gamma^\mu\dot\theta\rangle(t)|\le \ds\frac{C\ve}{\langle t\rangle^{1/2}}E_m(t)+\frac{C\ve^3}{\langle t\rangle^2}E_m^{1/2}(t).
$$
Similarly, for $|\nu|> |\mu-\nu|$,
$$
|\langle\Gamma^\nu\dot w\cdot\nabla\Gamma^{\mu-\nu}\dot\theta,\Gamma^\mu\dot\theta\rangle(t)|
\leq\frac{C}{\langle t\rangle^{1/2}}|\langle\si^{1/2}\si_-\nabla\Gamma^{\mu-\nu}\dot\theta\cdot\Gamma^\nu\dot w,\Gamma^\mu\dot\theta\rangle(t)|\leq\frac{C\ve}{\langle t\rangle^{1/2}}E_m(t).
$$
Therefore,
$$
|H_3^\mu(t)|\leq\frac{C\ve}{\langle t\rangle^{1/2}}E_m(t)+\frac{C\ve^3}{\langle t\rangle^2}E_m^{1/2}(t).\eqno{(4.7)}
$$
By the same way, we have
$$
|H_j^\mu(t)|\leq\frac{C\ve}{\langle t\rangle^{1/2}}E_m(t)+\frac{C\ve^3}{\langle t\rangle^2}E_m^{1/2}(t)
\quad\text{for}\quad|\mu|\leq m\quad\text{and}\quad j\in\{6,9,12\}.\eqno{(4.8)}
$$
On the other hand,
$$
\|\pa^\a(X+1)^k\big((1-\theta)z(0,x)\nabla\theta_a\big)\|
+\|\pa^\a(X+1)^k\big((1-\theta)\dot z\nabla\theta_a\big)\|\leq\frac{C\ve^2}{\langle t\rangle^{5/2}}+\frac{C\ve}{\langle t\rangle^{5/2}}E^{1/2}_m(t)\eqno{(4.9)}
$$
and
\begin{align*}
\|\pa^\a(X+1)^k\big((1-\theta)z\nabla\dot\theta\big)-(1-\theta)z\nabla\Gamma^\mu\dot\theta\|&\leq \frac{C}{\langle t\rangle}\sum_{0<\nu\leq \mu}\|\Gamma^\nu\big((1-\theta)z\big)\cdot\si_-\nabla\Gamma^{\mu-\nu}\dot\theta\|\\
&\le \ds\frac{C\ve}{\langle t\rangle}E^{1/2}_m(t)
+\frac{C\ve^3}{\langle t\rangle^{5/2}},\tag{4.10}
\end{align*}
which mean
$$
|H_{13}^\mu(t)|\leq\frac{C\ve^2}{\langle t\rangle^{5/2}}E^{1/2}_m(t)+\frac{C\ve}{\langle t\rangle}E_m(t).\eqno{(4.11)}
$$
Moreover, one can easily obtain
$$\|\hat g_{a1}^\mu(t)\|+\|\hat g_{1}^\mu(t)\|\leq\frac{C\ve^2}{\langle t\rangle^{5/2}}+ C\ve E_m^{1/2}(t),
$$
$$
\|\hat g_{a2}^\mu(t)\|+\|\hat g_{2}^\mu(t)\|\leq\frac{C\ve}{\langle t\rangle^{5/2}}E_m^{1/2}(t)+C\ve E_m^{1/2}(t)
$$
and then
$$
\sum_{j=1,2}|\hat H_{aj}^\mu|+\sum_{j=1,2}|\hat H_{j}^\mu|\leq\frac{C\ve^2}{\langle t\rangle^{5/2}}E_m^{1/2}(t)
+C\ve E_m(t).\eqno{(4.12)}
$$
By (4.6)-(4.9) and (4.11)-(4.12), we have from (4.4)
$$
\frac{d}{dt}E_m(t)\leq\frac{C\ve^2}{\langle t\rangle^{2}}E_m^{1/2}(t)+C\ve E_m(t),
$$
and then for sufficiently small $\ve$,
$$
E_m^{1/2}(t)\leq C\ve^2e^{C\ve t}\leq C\ve^2.\qquad\qquad\qquad\qquad \qquad\qquad
$$
Therefore, Lemma 4.1 is proved by the local existence of solution and continuous induction method.$\square$
\bigskip
As in Lemma 4.1, in order to prove Theorem 1.1 by the continuous induction method, we require to
establish a uniform estimate on the solution $(\rho, w, z)$. To this end, we will derive the a
priori estimate on the related potential $\phi$ in the domain
$\{x: |x|\ge M+1\}$.
{\bf Lemma 4.2.} {\it
Assume that $k$ and $\la$ are integers with $[\ds\frac{k+1}{2}]+3\leq\la\leq k$ and $k\geq7$,
$(\dot\theta, \dot w, \dot z)$ is a smooth solution of (2.4) for $(t,x)\in [\ds\f{1}{\ve}, T]\times\Bbb R^2$.
If $E_\la(t)\leq\ve^2$, then one can find a positive number $C$ independent of $\ve$ and $T$
such that the potential $\phi(t,x)$ of velocity $w(t,x)$ in the domain $\{x: |x|\ge M+1\}$ satisfies
\begin{align*}
&\sum_{|\mu|\leq k}\int_{1/\ve}^t\int_{D_+(\tau)}\si_-(t,x)^{-2}|Z\Gamma^\mu \phi(\tau, x)|^2d\tau dx \\
&\qquad \leq C\Big(\ve^4+\sum_{|\mu|\leq k}\int_{1/\ve}^t\langle\tau\rangle^{-1}\int_{\{|x|\geq M+1\}}|\Gamma^\mu\pa\phi(\tau, x)|^2d\tau dx \Big)\tag{4.13}
\end{align*}
}\bigskip
{\bf Remark 4.1.} {\it In order to apply energy integral method to derive (4.13), we require to choose a
different ``ghost weight'' from the one in [4] due to the following reason: notice that the both null conditions
of 2-D quasilinear wave equation are fulfilled in the whole space $\Bbb R^2$ in [4], but for our 2-D compressible
Euler system (1.3),
the null conditions hold only in the exterior domain $\{|x|\ge M+1\}$, and thus it is natural for us to multiply
a smooth cut-off function on the potential function $\phi$ so that a suitable weighted energy estimate can be obtained.
Due to this way of doing, the resulted ``ghost weight'' should be
reconsidered by comparison with that in [4]. More concretely speaking, the author in [4]
can obtain the a priori estimate $|\p\G^\mu v(t,x)|\le C\ve\si_-^{-1}(t,x)\si^{-\f12}(t,x)$ for the solution $v$ of
quasilinear wave equation $\p_t^2v-\Delta v+\ds\sum_{0\le i, j\le 2}g_{ij}(\p v)\p_{ij}^2v=0$ when the both null conditions hold, however, here we only get $|\p\G^\mu \phi(t,x)|\le C\ve\si_-^{-\f12}(t,x)\si^{-\f12}(t,x)$ for the potential $\phi$
(see (4.17) below) which will lead to a different choice of the ghost weight from that in [4].}
\bigskip
{\bf Proof.}
As in (2.1), the function $\phi(t,x)$ satisfies for $|x|\geq M+1$
$$
\displaystyle\pa_t^2\phi-\triangle \phi+2\sum_{j=1}^2\pa_j\phi\pa_t\pa_j\phi+\sum_{j,k=1}^2\pa_j\phi\pa_k\phi\pa_{jk}^2\phi
-(2\pa_t\phi+|\nabla\phi|^2)\triangle\phi=0.\eqno{(4.14)}
$$
Define
$$
\varphi(t,x)=\vp(|x|-t)=\int_{-\infty}^{|x|-t}\frac{1}{(1+|\rho|)^{3/2}} d\rho,
$$
then $0\leq\varphi\leq4$ and $\vp'(|x|-t)=\si_-(t,x)^{-3/2}$ hold. In addition, we set $\tilde\si(t,x)=\big(1+(|x|-t-M)^2\big)^{1/2}$ and $\chi(t,x)=\tilde\chi(\ds\frac{2r}{t+2M+2})$ with the smooth function
\begin{equation*}
\tilde\chi(s) =
\left\{
\begin{aligned}
0, & \quad \text{if $s\leq\frac{1}{2}$},\\
1, & \quad \text{if $s\geq1$}.
\end{aligned}
\right.
\end{equation*}
Let the corresponding energy be denoted by
$$
\widetilde E_n(t)=\ds\sum_{|\mu|\le n}\int_{\Bbb R^2}\tilde\si(t,x)^{-1/2}e^{\varphi(t,x)}\chi(t,x)\big(|\pa_t\G^\mu \phi(t,x)|^2+|\nabla\G^\mu\phi(t,x)|^2\big) dx
$$
for $n\in\Bbb N\cup\{0\}$. Motivated by the terminology in [4], the weight function $\t\si(t,x)^{-1/2}e^{\varphi(t,x)}\chi(t,x)$
is also called the ghost weight by us, which will display the null conditions and decay rate simultaneously.
Notice that
\begin{align*}
\ds\pa_t^2\Gamma^\mu \phi-&\triangle\Gamma^\mu \phi+2\sum_{j=1}^2\pa_j\phi\pa_t\pa_j\Gamma^\mu \phi+\sum_{j,k=1}^2\pa_j\phi\pa_k\phi\pa_{jk}^2\Gamma^\mu \phi-(2\pa_t\phi+|\nabla\phi|^2)\triangle \Gamma^\mu \phi\\
&=\sum_{\nu\leq \mu}C_\nu\Gamma^\nu\big(-2\sum_{j=1}^2\pa_j\phi\pa_t\pa_j\phi-\sum_{j,k=1}^2\pa_j\phi\pa_k\phi\pa_{jk}^2\phi
+(2\pa_t\phi+|\nabla\phi|^2)\triangle\phi\big)\\
&+2\sum_{j=1}^2\pa_j\phi\pa_t\pa_j\Gamma^\mu \phi+\sum_{j,k=1}^2\pa_j\phi\pa_k\phi\pa_{jk}^2\Gamma^\mu
\phi-(2\pa_t\phi+|\nabla\phi|^2)\triangle \Gamma^\mu \phi.\tag{4.15}
\end{align*}
Multiplying (4.15) by $\tilde\si^{-1/2}(t,x)e^{\varphi(t,x)}\chi(t,x)\pa_t\Gamma^\mu \phi$, integrating in the space $\Bbb R^2$ and using integration by parts, we can get
\begin{align*}
&\frac{1}{2}\frac{d}{dt}\int_{\Bbb R^2}\tilde\si^{-1/2}e^{\varphi}\chi\big(|\pa_t\Gamma^\mu \phi|^2+|\nabla\Gamma^\mu \phi|^2\big) dx+\frac{1}{2}\int_{\Bbb R^2}\tilde\si^{-1/2}e^{\varphi}\chi\varphi'|Z\Gamma^\mu \phi|^2 dx\\
&+\frac{ d}{dt}\int_{\Bbb R^2}\tilde\si^{-1/2}e^{\varphi}\chi\Bigg(\pa_t\phi|\nabla\Gamma^\mu\phi|^2-\frac{1}{2}\sum_{i,j=1}^2\pa_i \phi\pa_j\phi(\pa_j\Gamma^\mu\phi)(\pa_i\Gamma^\mu\phi)+\frac{1}{2}|\nabla\phi|^2|\nabla\Gamma^\mu\phi|^2\Bigg)dx\\
&-\frac{1}{4}\int_{\Bbb R^2}(|x|-t-M)\tilde\si^{-5/2}e^{\varphi}\chi|Z\Gamma^\mu\phi|^2 dx+\sum_{i=2}^5L^\mu_i(t)
=L^\mu_1(t),\tag{4.16}
\end{align*}
where
\begin{align*}
L^\mu_1(t)=&\int_{\Bbb R^2}\tilde\si^{-1/2}e^{\varphi}\chi\pa_t\Gamma^\mu\phi\Big\{\sum_{\nu\leq \mu}C_{\nu}\Gamma^\nu\bigg(-2\sum_{j=1}^2\pa_j\phi\pa_t\pa_j\phi-\sum_{j,k=1}^2\pa_j\phi\pa_k\phi\pa_{jk}^2\phi+(2\pa_t\phi\\
&+|\nabla\phi|^2)\triangle \phi\bigg)+2\sum_{j=1}^2\pa_j\phi\pa_t\pa_j\Gamma^\mu\phi+\sum_{j,k=1}^2\pa_j\phi\pa_k\phi\pa_{jk}^2\Gamma^\mu\phi
-(2\pa_t\phi+|\nabla\phi|^2)\triangle \Gamma^\mu\phi\Big\} dx,\\
L^\mu_2(t)=&-\int_{\Bbb R^2}\tilde\si^{-1/2}e^{\varphi}\varphi'\chi\Big\{\sum_{i=1}^2\omega_i\pa_i \phi(\pa_t\Gamma^\mu\phi)^2-2\sum_{i=1}^2\omega_i\pa_t\phi(\pa_i\Gamma^\mu\phi)(\pa_t\Gamma^\mu\phi)\\
&-\sum_{i=1}^2\pa_t\phi(\pa_i\Gamma^\mu\phi)^2+\sum_{i,j=1}^2\omega_i\pa_i\phi\pa_j\phi(\pa_j\Gamma^\mu\phi)
(\pa_t\Gamma^\mu\phi)+\frac{1}{2}\sum_{i,j=1}^2\pa_i\phi\pa_j\phi(\pa_i\Gamma^\mu\phi)(\pa_j\Gamma^\mu\phi)\\
&-\sum_{i=1}^2\omega_i|\nabla\phi|^2(\pa_i\Gamma^\mu\phi)(\pa_t\Gamma^\mu\phi)-\frac{1}{2}|\nabla\phi|^2|\nabla\Gamma^\mu \phi|^2\Big\}dx,\\
\end{align*}
\begin{align*}
L^\mu_3(t)=&-\int_{\Bbb R^2}\tilde\si^{-1/2}e^{\varphi}\chi\Big\{\sum_{i=1}^2\pa_i^2\phi(\pa_t\Gamma^\mu\phi)^2
-2\sum_{i=1}^2\pa_i\pa_t\phi(\pa_i\Gamma^\mu \phi)(\pa_t\Gamma^\mu\phi)+\pa_t^2\phi|\nabla\Gamma^\mu\phi|^2\\
&+\sum_{i,j=1}^2\pa_i(\pa_i\phi\pa_j\phi)(\pa_j\Gamma^\mu\phi)(\pa_t\Gamma^\mu\phi)
-\sum_{i,j=1}^2\pa_t(\pa_i\phi\pa_j\phi)(\pa_j\Gamma^\mu\phi)(\pa_i\Gamma^\mu\phi)\\
&-\sum_{i=1}^2\pa_i(|\nabla\phi|^2)(\pa_i\Gamma^\mu\phi)(\pa_t\Gamma^\mu\phi)+\frac{1}{2}\pa_t(|\nabla\phi|^2)|\nabla \Gamma^\mu\phi|^2\Big\} dx,\\
L^\mu_4(t)=&(t+2M+2)^{-1}\int_{\Bbb R^2}\tilde\si^{-1/2}e^{\varphi}
\tilde\chi'\Big\{\frac{r}{t+2M+2}\big((\pa_t\Gamma^\mu\phi)^2
+|\nabla\Gamma^\mu\phi|^2\big)-2\sum_{i=1}^2\omega_i\pa_i\phi(\pa_t\Gamma^\mu\phi)^2\\
&+2\sum_{i=1}^2\omega_i(\pa_t\Gamma^\mu\phi)(\pa_i\Gamma^\mu\phi)
+4\sum_{i=1}^2\omega_i\pa_t\phi(\pa_t\Gamma^\mu\phi)
(\pa_i\Gamma^\mu\phi)+2r(t+2M+2)^{-1}\pa_t\phi|\nabla\Gamma^\mu\phi|^2\\
&-2\sum_{i,j=1}^2\omega_i\pa_i\phi\pa_j\phi(\pa_j\Gamma^\mu\phi)(\pa_t\Gamma^\mu\phi)
-r(t+2M+2)^{-1}\pa_i\phi\pa_j\phi(\pa_j\Gamma^\mu\phi)(\pa_i\Gamma^\mu\phi)\\
&+2\sum_{i=1}^2\omega_i|\nabla\phi|^2(\pa_t\Gamma^\mu\phi)(\pa_i\Gamma^\mu\phi)
+r(t+2M+2)^{-1}|\nabla\phi|^2|\nabla\Gamma^\mu\phi|^2\Big\} dx,\\
L^\mu_5(t)=&\frac{1}{2}\int_{\Bbb R^2}(|x|-t-M)\tilde\si^{-5/2}e^{\varphi}
\chi\Big\{\sum_{i=1}^2\omega_i\pa_i \phi(\pa_t\Gamma^\mu\phi)^2-2\sum_{i=1}^2\omega_i\pa_t\phi(\pa_i\Gamma^\mu\phi)(\pa_t\Gamma^\mu\phi)\\
-&\sum_{i=1}^2\pa_t\phi(\pa_i\Gamma^\mu\phi)^2+\sum_{i,j=1}^2\omega_i\pa_i\phi\pa_j\phi(\pa_j\Gamma^\mu\phi)
(\pa_t\Gamma^\mu\phi)+\frac{1}{2}\sum_{i,j=1}^2\pa_i\phi\pa_j\phi(\pa_i\Gamma^\mu\phi)(\pa_j\Gamma^\mu\phi)\\
-&\sum_{i=1}^2\omega_i|\nabla\phi|^2(\pa_i\Gamma^\mu\phi)(\pa_t\Gamma^\mu\phi)
-\frac{1}{2}|\nabla\phi|^2|\nabla\Gamma^\mu\phi|^2\Big\} dx.
\end{align*}
By $\pa_t\dot\phi=-\dot\theta+F_2(\dot\xi,2\xi_a+\dot\xi)$ and $\nabla\dot\phi=\dot w$ for $|x|\geq M+1$, then
by Lemma 2.3 and Lemma 3.5 we have
$$
|\pa\Gamma^\nu\phi(t,x)|\leq C\ve\si_-(t,x)^{-1/2}\si(x)^{-1/2}\quad\text{for}\quad |\nu|\leq \la-2\quad\text{and}\quad |x|\geq M+1.
\eqno{(4.17)}
$$
Similarly,
$$
|\pa^2\Gamma^\nu \phi(t,x)|\leq C\ve\si_-(t,x)^{-1}\si(x)^{-1/2}\quad\text{for}\quad |\nu|\leq \la-3\quad\text{and}\quad |x|\geq M+1.
\eqno{(4.18)}
$$
In addition, it follows Lemma 3.4 that
$$
|\Gamma^\nu\phi(t,x)|\leq C\ve\si(x)^{-1/2}\si_-(t,x)\quad\text{for}\quad |\nu|\leq \la-2\quad\text{and}\quad |x|\geq M+1.
\eqno{(4.19)}
$$
Therefore, we derive from (4.16)-(4.19) that
\begin{align*}
&\widetilde E_k(t)+\sum_{|\mu|\leq k}\int_{1/\ve}^t\int_{\Bbb R^2}\tilde\si^{-1/2}e^\varphi\chi\varphi'|Z\Gamma^\mu\phi|^2 dx d\tau\\
&\qquad -\sum_{|\mu|\leq k}\int_{1/\ve}^t\int_{\Bbb R^2}(|x|-t-M)\tilde\si^{-5/2}e^{\varphi}\chi|Z\Gamma^\mu\phi|^2dxd\tau\\
&\leq C\Big(\widetilde E_k(\frac{1}{\ve})+\sum_{|\mu|\leq k}\sum_{i=1}^5\int_{1/\ve}^t|L^\mu_i(\tau)|d\tau\Big).\tag{4.20}
\end{align*}
Next we deal with each term $L^\mu_i (1\le i\le 5)$ in the right hand side of (4.20).
In this process, we will often use the fact that $\tilde\si(t,x)$ is equivalent to $\si_-(t,x)$.
First, we take $g_i^{0i}=g_i^{i0}=-1$, $g_0^{ii}=2$, $g_{ij}^{ij}=g_{ij}^{ji}=-\ds\frac{1}{2}$, $g_{ii}^{jj}=1$ for $i,j=1,2$, and the others are 0, then we have by Lemma 3.1 and Lemma 3.2 together with (4.17)-(4.19)
\begin{align*}
&|L_1^\mu (t)|\leq \int_{D_+(t)}\tilde\si^{-1/2}e^\varphi\chi|\Gamma^\mu\big(\sum_{i,j,k=0}^2g_i^{jk}\pa_i\phi\pa_{jk}^2\phi\big)
-\sum_{i,j,k=0}^2g_i^{jk}\pa_i\phi\pa_{jk}^2\Gamma^\mu\phi||\pa_t\Gamma^\mu\phi|{d}x\\
&\quad +\sum_{\nu\leq \mu}\int_{D_-(t)}\tilde\si^{-1/2}e^\varphi\chi|C_{\nu}\Gamma^\nu\big(\sum_{i,j,k=0}^2g_i^{jk}\pa_i\phi\pa_{jk}^2\phi\big)
-\sum_{i,j,k=0}^2g_i^{jk}\pa_i\phi\pa_{jk}^2\Gamma^\mu\phi||\pa_t\Gamma^\mu\phi|{d}x\\
&\quad +\sum_{\nu\leq\mu}\int_{\Bbb{R}^2}\tilde\si^{-1/2}e^\varphi\chi|C_{\nu}\Gamma^\nu\big(\sum_{i,j,k,l=0}^2g_{ij}^{kl}\pa_i\phi\pa_j\phi
\pa_{kl}^2\phi\big)-\sum_{i,j,k,l=0}^2g_{ij}^{kl}\pa_i\phi\pa_j\phi\pa_{kl}^2\Gamma^\mu\phi||\pa_t\Gamma^\mu\phi|{d}x\\
&\quad +\sum_{\nu<\mu}\int_{D_+(t)}\tilde\si^{-1/2}e^\varphi\chi|C_{\nu}\Gamma^\nu
\big(\sum_{i,j,k=0}^2g_i^{jk}\pa_i\phi\pa_{jk}^2\phi\big)||\pa_t\Gamma^\mu\phi|dx\\
&\leq C\Big\{\langle t\rangle^{-1}\int_{D_+(t)}\tilde\si^{-1/2}e^\varphi\chi\big(\si_-\sum_{|\nu|\leq [\frac{k+1}{2}]}|\pa \Gamma^\nu \phi|+\sum_{|\nu|\leq [\frac{k+1}{2}]+1}|\Gamma^\nu \phi|\big)\sum_{|\nu|\leq k}|\pa \Gamma^\nu \phi|^2{d}x\\
&\quad +\int_{D_-(t)}\tilde\si^{-1/2}e^\varphi\chi\sum_{|\nu_1|\leq[\frac{k+1}{2}]}|\pa \Gamma^{\nu_1}\phi|\sum_{|\nu_2|\leq k}|\pa \Gamma^{\nu_2} \phi|^2{d}x\\
&\quad +\int_{\Bbb{R}^2}\tilde\si^{-1/2}e^\varphi\chi\sum_{|\nu_1|\leq [\frac{k+1}{2}]}|\pa \Gamma^{\nu_1}\phi|^2\sum_{|\nu_2|\leq k}|\pa \Gamma^{\nu_2}\phi|^2{d}x\\
&\quad +\int_{D_+(t)}\tilde\si^{-1/2}e^\varphi\chi\sum_{|\nu_1|\leq[\frac{k}{2}]}|\pa \Gamma^{\nu_1}\phi|\sum_{|\nu_2|\leq k}|Z\Gamma^{\nu_2}\phi|\sum_{|\nu_3|\leq k}|\pa \Gamma^{\nu_3}\phi|{d}x\Big\}\tag{4.21}\\
&\leq C\ve \sum_{|\nu|\leq k}\int_{\Bbb{R}^2}\tilde\si^{-1/2}e^\varphi\chi{\varphi'}|Z\Gamma^\nu\phi|^2{d}x+C\ve \langle t\rangle^{-1}\sum_{|\nu|\leq k}\int_{\Bbb{R}^2}e^\varphi\chi|\pa\Gamma^\nu\phi|^2{d}x,\tag{4.22}
\end{align*}
here we give some explanations for the derivation process from (4.21) to (4.22), for example,
in order to treat the last term in (4.21), one can make use of (4.17) to get
\begin{align*}
&\int_{D_+(t)}\tilde\si^{-1/2}e^\varphi\chi\sum_{|\nu_1|\leq[\frac{k}{2}]}|\pa \Gamma^{\nu_1}\phi|\sum_{|\nu_2|\leq k}|Z\Gamma^{\nu_2}\phi|\sum_{|\nu_3|\leq k}|\pa \Gamma^{\nu_3}\phi|{d}x\\
&\le C\ve\int_{D_+(t)}{\t\si}^{-\f12}e^{\vp}\chi\si_-^{-\f12}\si^{-\f12}\sum_{|\nu_2|\leq k}|Z\Gamma^{\nu_2}\phi|\sum_{|\nu_3|\leq k}|\pa \Gamma^{\nu_3}\phi|{d}x\\
&\leq C\ve \sum_{|\nu|\leq k}\int_{\Bbb{R}^2}\tilde\si^{-1/2}e^\varphi\chi{\varphi'}|Z\Gamma^\nu\phi|^2{d}x+C\ve \langle t\rangle^{-1}\sum_{|\nu|\leq k}\int_{\Bbb{R}^2}e^\varphi\chi|\pa\Gamma^\nu\phi|^2{d}x\\
&\qquad \quad\qquad \quad\qquad \quad\qquad \quad\text{($\si_-$
is equivalent to $\t\si$ in $D_+(t)$)}.
\end{align*}
Analogously,
\begin{align*}
|L_2^\mu(t)|&\leq C\Big\{\int_{D_+(t)}\tilde\si^{-1/2}e^\varphi\chi{\varphi'}|\sum_{i=1}^2\omega_i Z_i\phi(\pa_t\Gamma^\mu \phi)^2-\pa_t \phi\sum_{i=1}^2(Z_i\Gamma^\mu\phi)^2|{d}x\\
&\quad +\int_{D_-(t)}\tilde\si^{-1/2}e^\varphi\chi{\varphi'}|\pa\phi||\pa\Gamma^\mu \phi|^2{d}x+\int_{\Bbb{R}^2}\tilde\si^{-1/2}e^\varphi\chi{\varphi'}|\pa\phi|^2|\pa\Gamma^\mu\phi|^2{d}x\Big\}\\
&\leq C\ve\langle t\rangle^{-1/2}\int_{\Bbb{R}^2}\tilde\si^{-1/2}e^\varphi\chi{\varphi'}|Z\Gamma^\mu\phi|^2{d}x+C\ve\langle t\rangle^{-1}\int_{\Bbb{R}^2}e^\varphi\chi|\pa\Gamma^\mu\phi|^2{d}x,\tag{4.23}\\
|L_3^\mu(t)|&\leq C\Big\{\int_{D_+(t)}\tilde\si^{-1/2}e^\varphi\chi\big(|\pa\Gamma^\mu\phi|^2|Z\pa\phi|
+|(\pa\Gamma^\mu\phi)(Z\Gamma^\mu\phi)(\pa^2\phi)|\big) dx\\
&\quad +\int_{D_-(t)}\tilde\si^{-1/2}e^\varphi\chi|(\pa^2\phi)(\pa\Gamma^\mu\phi)^2| dx
+\int_{\Bbb R^2}\tilde\si^{-1/2}e^\varphi\chi|\pa\phi(\pa^2\phi)(\pa\Gamma^\mu\phi)^2| dx\Big\}\\
&\leq C\ve\int_{\Bbb{R}^2}\tilde\si^{-1/2}e^\varphi\chi{\varphi}'|Z\Gamma^\mu\phi|^2{d}x+C\ve\langle t\rangle^{-1}\int_{\Bbb{R}^2}e^\varphi\chi|\pa\Gamma^\mu\phi|^2{d}x,\tag{4.24}\\
|L_4^\mu(t)|&\leq C\langle t\rangle^{-1}\int_{\{|x|\geq M+1\}}e^\varphi|\pa\Gamma^\mu\phi|^2{d}x,\tag{4.25}\\
|L_5^\mu(t)|&\leq C\Big\{-\int_{D_+(t)}(|x|-t-M)\tilde\si^{-5/2}e^\varphi\chi|\sum_{i=1}^2\omega_i Z_i\phi(\pa_t\Gamma^\mu\phi)^2-\pa_t\phi\sum_{i=1}^2(Z_i\Gamma^\mu\phi)^2|{d}x\\
&\quad +\int_{D_-(t)}\tilde\si^{-3/2}e^\varphi\chi|\pa\phi||\pa\Gamma^\mu\phi|^2{d}x+\int_{\Bbb{R}^2}\tilde\si^{-3/2}e^\varphi\chi|\pa \phi|^2|\pa\Gamma^\mu\phi|^2{d}x\Big\}\\
&\leq -C\ve\langle t\rangle^{-1/2}\int_{\Bbb{R}^2}(|x|-t-M)\tilde\si^{-5/2}e^\varphi\chi|Z\Gamma^\mu\phi|^2{d}x+C\ve\langle t\rangle^{-1}\int_{{R}^2}e^\varphi\chi|\pa\Gamma^\mu\phi|^2{d}x.\tag{4.26}
\end{align*}
Substituting (4.22)-(4.26) into (4.20) yields
\begin{align*}
&\widetilde E_k(t)+\sum_{|\mu|\leq k}\int_{1/\ve}^t\int_{\Bbb R^2}
\tilde\si^{-1/2}e^\varphi\chi\varphi'|Z\Gamma^\mu\phi|^2dx d\tau\\
&\qquad -\sum_{|\mu|\leq k}\int_{1/\ve}^t\int_{\Bbb R^2}(|x|-t-M)\tilde\si^{-5/2}e^{\varphi}\chi|Z\Gamma^\mu\phi|^2 dx d\tau\\
&\leq C\widetilde E_k(\frac{1}{\ve})+C\int_{1/\ve}^t\langle \tau\rangle^{-1}\sum_{|\mu|\leq k}\int_{\{|x|\geq M+1\}}|\pa\Gamma^\mu \phi|^2{d}x.\tag{4.27}
\end{align*}
This, together with Lemma 4.1, yields (4.13).\qquad\qquad\qquad\qquad $\square$ \bigskip
From (4.13) and (2.2), we have for $\dot\phi=\phi-\phi_a$
\begin{align*}
&\sum_{|\mu|\leq k}\int_{1/\ve}^t\int_{D_+(\tau)}\si_-(t,x)^{-2}|Z\Gamma^\mu\dot\phi(\tau,x)|^2 dx d\tau\\
&\leq C\Big(\ve^2\ln t+\sum_{|\mu|\leq k}\int_{1/\ve}^t\langle\tau\rangle^{-1}\int_{\{|x|\geq M+1\}}|\Gamma^\mu\pa\dot\phi(\tau,x)|^2dx d\tau\Big).\tag{4.28}
\end{align*}
{\bf Lemma 4.3.} {\it
Assume that $k$ and $\la$ are integers with $[\frac{k+7}{2}]\leq\la\leq k$ and $k\geq7$,
$(\dot\theta, \dot w, \dot z)$ is a smooth solution of
(2.4) for $(t,x)\in [1/\ve, T]\times \Bbb R^2$.
If $E_\la(t)\leq\ve^2$, then one can find a positive number $C$ independent of $\ve$ and $T$ such that for
$t\in[1/\ve,T]$ and sufficient small $\ve>0$,
$E_k(t)\leq C\ve^2\langle t\rangle^{C\ve}$ holds.
}
{\bf Proof.}
First, we come to estimate $H_j^\mu$ in (4.5) when $|\mu|\leq k$ and $j=1,\cdots,13$. It is easy to conclude that
$$
\sum_{|\mu|\leq k}\sum_{i=1}^{13}|H_i^\mu(t)|_-\leq C\ve\langle t\rangle^{-1}E_k(t)
+C\ve^2\langle t\rangle^{-5/2}E_k(t)^{1/2},\eqno{(4.29)}
$$
here the estimate on $H_{13}^\mu$ can be obtained as in (4.11).
We now continue to use the analogous notations as in (3.21), namely,
$$
\hat{h}_0^\mu=\sum_{0<\nu\leq\mu}C_{\mu\nu}\big(I_1^{\mu\nu}+I_3^{\mu\nu}\big)
+\sum_{0\leq \nu\leq\mu}C_{\mu\nu}I_2^{\mu\nu}+K^\mu,
$$
where
\begin{align*}
K^\mu=&\sum_{0<\nu\leq\mu}C_{\mu\nu}\big(-\Gamma^\nu w\cdot\nabla\Gamma^{\mu-\nu}F_2(\dot\xi,2\xi_a+\dot\xi)+\Gamma^\nu F_1(\xi)\nabla\cdot\Gamma^{\mu-\nu}\dot w\big)\\
&+\sum_{0\leq \nu\leq\mu}C_{\mu\nu}\big(-\Gamma^\nu \dot w\cdot\nabla\Gamma^{\mu-\nu}F_1(\xi_a)+\Gamma^\nu F_2(\dot\xi,2\xi_a+\dot\xi)\nabla\cdot\Gamma^{\mu-\nu}w_a\big).
\end{align*}
It follows from the similar analysis of (3.22) and (3.24) together with (4.28) that
\begin{align*}
&\sum_{0<\nu\leq\mu}C_{\mu\nu}\|I_1^{\mu\nu}(t)\|_+\leq C\ve \langle t\rangle^{-3/2}E^{1/2}_k(t),\qquad\qquad\qquad\qquad\qquad\qquad\qquad\tag{4.30}\\
&\ve^{-1}\sum_{0\leq \nu\leq \mu}C_{\mu\nu}\int_{1/\ve}^t\langle\tau\rangle\|I_2^{\mu\nu}(\tau)\|_+^2d\tau\\
&\quad \leq C\ve^{-1}\int_{1/\ve}^t\langle\tau\rangle\int_{D_+(\tau)}\sum_{|\nu_1|+|\nu_2|\leq k}\Big(|Z\Gamma^{\nu_1} \dot\phi|^2|\pa^2\Gamma^{\nu_2} \phi_a|^2+|\pa\Gamma^{\nu_1} \dot\phi|^2|Z\pa\Gamma^{\nu_2} \phi_a|^2\Big) dx d\tau\\
&\quad \leq C\ve\sum_{|\nu|\leq k}\int_{1/\ve}^t\int_{D_+(\tau)}\si_-(t,x)^{-2}|Z\Gamma^\nu\dot\phi(t,x)|^2dx d\tau+C\ve\int_{1/\ve}^t\langle\tau\rangle^{-1}E_k(\tau) d\tau\\
&\quad \leq C\ve^3\ln t+C\ve\int_{1/\ve}^t\langle\tau\rangle^{-1}E_k(\tau) d\tau,\tag{4.31}\\
&\ve^{-1}\sum_{0< \nu\leq\mu}C_{\mu\nu}\int_{1/\ve}^t\langle\tau\rangle\|I_3^{\mu\nu}(\tau)\|_+^2d\tau\\
&\quad \leq C\ve^{-1}\int_{1/\ve}^t\langle\tau\rangle\int_{D_+(\tau)}\sum_{|\nu_1|+|\nu_2|\leq k,|\nu_2|\leq k-1}\Big(|Z\Gamma^{\nu_1}\dot\phi|^2|\pa^2\Gamma^{\nu_2} \dot\phi|^2+|\pa\Gamma^{\nu_1}\dot\phi|^2|Z\pa\Gamma^{\nu_2}\dot\phi|^2\Big) dx d\tau\\
&\quad \leq C\ve\sum_{|\nu|\leq k-1} \int_{1/\ve}^t\langle\tau\rangle^{-2}\int_{D_+(\tau)}|\si_-\pa\Gamma^\nu\pa\dot\phi|^2{d}x d\tau\\
&\qquad +C\ve\sum_{|\nu|\leq k}\int_{1/\ve}^t\int_{D_+(\tau)}\si_-(x,t)^{-2}|Z\Gamma^\nu\dot\phi(t,x)|^2dxd\tau
+C\ve\int_{1/\ve}^t\langle\tau\rangle^{-1}E_k(\tau)d\tau\\
&\quad \leq C\ve^3\ln t+C\ve\int_{1/\ve}^t\langle\tau\rangle^{-1}E_k(\tau)d\tau.\tag{4.32}
\end{align*}
By using Lemma 2.2 and Lemma 2.3, we can get
$$
|K^\mu(t)|_+\leq C\ve^2\langle t\rangle^{-1}\sum_{|\nu|\leq k}\big(|\Gamma^\nu \dot\theta(t)|_+
+|\Gamma^\nu\dot w(t)|_+\big).\eqno{(4.33)}
$$
Hence, we have
$$
\int_{1/\ve}^t\sum_{|\mu|\leq k}\sum_{i=1}^{6}|H_i^\mu(\tau)|_+d\tau\leq C\ve^3\ln t+C\ve\int_{1/\ve}^t\langle\tau\rangle^{-1}E_k(\tau) d\tau,
$$
and similarly,
$$
\int_{1/\ve}^t\sum_{|\mu|\leq k}\sum_{i=7}^{12}|H_i^\mu(t)|_+ d\tau\leq C\ve^3\ln t+C\ve\int_{1/\ve}^t\langle\tau\rangle^{-1}E_k(\tau) d\tau.
$$
Second, we deal with the terms $\hat H^\mu_{aj}$ and $\hat H^\mu_{j}$ in (4.5) for $j=1,2$ and $|\mu|\leq k$.
Due to $\text{supp}\dot z\subset\{|x|\leq M+1\}$,
it is obvious that for $j=1,2$
$$|\hat H_{aj}^\mu|\leq C\ve\langle t\rangle^{-5/2}E_k^{1/2}(t)\big(E_k^{1/2}(t)+\ve\big).\eqno{(4.34)}$$
To treat the terms $\hat H_{j}^\mu$ ($j=1,2$), we will use the analogous method in $\S 5$ of [9] (see pages 101 of [9]).
For this end, we set $\la_a^{\mu\nu}=\Gamma^\nu \dot w\cdot\nabla\Gamma^{\mu-\nu}z(0,x)$ and $\la^{\mu\nu}=\Gamma^\nu \dot w\cdot\nabla\Gamma^{\mu-\nu}\dot z$.
If $\Gamma^\nu=\pa\Gamma^d$ with $|d|=|\nu|-1$, then
$$
\|\la_a^{\mu\nu}(t)\|\leq C\ve\langle t\rangle^{-1}\|(\si_-\pa\Gamma^d\dot w)(t)\|\leq C\ve\langle t\rangle^{-1}\big(E_k^{1/2}(t)+\ve^2\langle t\rangle^{-3/2}\big)\eqno{(4.35)}
$$
and
$$
\|\la^{\mu\nu}(t)\|\leq\big(\sup_{|x|\leq M+1}|\pa\Gamma^d\dot w(t,x)|\big)\|\nabla\Gamma^{\mu-\nu}\dot z(t)\|\leq C\ve\langle t\rangle^{-1}E_k^{1/2}(t)\quad\text{for}\quad |\nu|\leq|\mu-\nu|,\eqno{(4.36)}
$$
\begin{align*}
&\|\la^{\mu\nu}(t)\|\leq C\big(\sup\nabla\Gamma^{\mu-\nu} \dot z(t,x)\big)\langle t\rangle^{-1}\|\si_-\pa\Gamma^d\dot w(t)\|\leq C\ve\langle t\rangle^{-1}\big(E_k^{1/2}(t)+\ve^2\langle t\rangle^{-3/2}\big)\\
&\qquad \quad\qquad \quad\qquad \quad\text{for}\quad |\nu|>|\mu-\nu|.\tag{4.37}
\end{align*}
If $\Gamma^\nu=X^{|\nu|}$ with $|\nu|\leq k-1$, due to $\Gamma^\nu \dot w(t,x)=\big(\Gamma^\nu W(t,r)\big)\ds\frac{x}{r}$
holds for some smooth function $W(t,r)$, then we have
$$
\sup_{|x|\leq M+1}|\Gamma^\nu\dot w(t,x)|\leq C\langle t\rangle^{-1}\|\si_-\nabla\cdot\Gamma^\nu\dot w(t,x)\|\leq C\langle t\rangle^{-1}Q_{|\nu|+1}(t).\eqno{(4.38)}
$$
Hence, we derive from (4.38) that for $|\nu|\leq k-1$
$$
\|\la_a^{\mu\nu}(t)\|\leq C\ve\langle t\rangle^{-1}Q_k(t)\leq C\ve\langle t\rangle^{-1}\big(E_k^{1/2}(t)+\ve^2\langle t\rangle^{-3/2}\big),\eqno{(4.39)}
$$
$$
\|\la^{\mu\nu}(t)\|\leq C\langle t\rangle^{-1}Q_{|\nu|+1}(t)E_k^{1/2}(t)\leq C\ve\langle t\rangle^{-1}E_k^{1/2}(t)\quad\text{for}\quad |\nu|\leq|\mu-\nu|,\eqno{(4.40)}
$$
and
$$
\|\la^{\mu\nu}(t)\|\leq C\ve\langle t\rangle^{-1}\big(E_k^{1/2}(t)
+\ve^2\langle t\rangle^{-3/2}\big)\quad\text{for}\quad |\nu|>|\mu-\nu|.\eqno{(4.41)}
$$
If $\Gamma^\nu=X^k$, then exactly as in the proof of (5.10) in [9]
(here we will apply Lemma 2.5) together with the related estimates (4.35)-(4.41),
we can obtain
$$
|\hat H_{j}^\mu(t)-\frac{d}{dt}G_j(t)|\leq C\ve\langle t\rangle^{-1}\big(E_k(t)
+\ve^2\langle t\rangle^{-3/2}E^{1/2}_k(t)+\ve^5\langle t\rangle^{-3}\big)\qquad \text{for $j=1,2$},\eqno{(4.42)}
$$
where $|G_j(t)|\leq C\big(\ve E_k(t)+\ve^3\langle t\rangle^{-3/2}E^{1/2}_k(t)+\ve^6\langle t\rangle^{-3}\big)$.
Combining (4.34) with (4.42) yields
$$
\sum_{|\mu|\leq k}\sum_{j=1,2}(\hat H^\mu_{aj}(t)+\hat H^\mu_{j}(t))=\frac{ d}{ dt}\t G_1(t)+\t G_2(t),\eqno{(4.43)}
$$
where $|\t G_1(t)|\leq C\big(\ve E_k(t)+\ve^3\langle t\rangle^{-3/2}E^{1/2}_k(t)+\ve^6\langle t\rangle^{-3}\big)$, and $|\t G_2(t)|\leq C\ve\langle t\rangle^{-1}\big(E_k(t)+\ve^5\langle t\rangle^{-3}$ $+\ve\langle t\rangle^{-3/2}E^{1/2}_k(t)\big)$.
Third, according to the expressions of $A_j$ for $j=0, 1, 2$, it is easily known that
$$
\sum_{j=0}^2|\pa_j A_j(t)|_-\leq C\big(|\nabla\theta(t)|_-+|\nabla\cdot w(t)|_-+|w\cdot\nabla z(t)|_-\big)\leq C\ve\langle t\rangle^{-3/2},
$$
and then
$$|\ds\sum_{j=0}^2\langle(\pa_j A_j)\zeta^\mu,\zeta^\mu\rangle_-(t)|\leq C\ve\langle t\rangle^{-3/2} E_k(t).\eqno{(4.44)}$$
For the case of $|x|\geq \ds\frac{t}{2}
+M+1$, we know $z(t,x)=0$ by Lemma 2.6 and
\begin{align*}
&\sum_{j=0}^2\langle(\pa_j A_j)\zeta^\mu,\zeta^\mu\rangle_+(t)\\
=&\langle(\nabla\cdot w)\Gamma^\mu\dot\theta-\nabla\theta\cdot\Gamma^\mu\dot w,\Gamma^\mu
\dot\theta\rangle_+(t)+\langle(\nabla\cdot w)\Gamma^\mu \dot w-(\Gamma^\mu\dot\theta)\nabla\theta,\Gamma^\mu\dot w\rangle_+(t)\\
=&\langle \pa_t\nabla\phi\cdot\Gamma^\mu\nabla\dot\phi-\triangle\phi\Gamma^\mu\pa_t\dot\phi,\Gamma^\mu
\dot\theta\rangle_+(t)+\langle(\triangle\phi)\Gamma^\mu\nabla\dot\phi-(\Gamma^\mu\pa_t\dot\phi)\nabla\pa_t\phi,\Gamma^\mu \dot w\rangle_+(t)\\
&+\langle (\nabla\cdot w)\Gamma^\mu F_2(\dot\xi,2\xi_a+\dot\xi)
-\nabla F_1(\xi)\cdot\Gamma^\mu \dot w,\Gamma^\mu\dot\theta\rangle_+(t)\\
&+\langle(\Gamma^\mu F_2(\dot\xi,2\xi_a
+\dot\xi)-\Gamma^\mu\dot\theta)\nabla F_1(\xi)-\Gamma^\mu F_2(\dot\xi,2\xi_a+\dot\xi)\nabla\theta,
\Gamma^\mu\dot w\rangle_+(t).\tag{4.45}
\end{align*}
Notice that
\begin{align*}
&\int_{1/\ve}^t\langle \pa_t\nabla\phi\cdot\Gamma^\mu\nabla\dot\phi-\triangle\phi\Gamma^\mu\pa_t\dot\phi,\Gamma^\mu
\dot\theta\rangle_+(\tau)d\tau\\
&\quad \leq C\sum_{l=1}^2\sum_{\nu\leq\mu}\int_{1/\ve}^t|\langle\pa_t\pa_l\phi\pa_l\Gamma^\nu \dot\phi-\pa_l^2\phi\pa_t\Gamma^\nu\dot\phi,\Gamma^\mu\dot\theta\rangle_+(\tau)|
d\tau\\
&\quad \leq C\ve\sum_{|\nu|\leq k}\int_{1/\ve}^t\int_{D_+(\tau)}\si_-^{-2}|Z\Gamma^\nu\dot\phi|^2dxd\tau+C\ve\int_{1/\ve}^t
\langle\tau\rangle^{-1}E_k(\tau)d\tau\\
&\quad \leq C\ve^3\ln t+C\ve\int_{1/\ve}^t\langle\tau\rangle^{-1}E_k(\tau)d\tau,\tag{4.46}
\end{align*}
and in the same way, one has
$$
\int_{1/\ve}^t\langle(\triangle\phi)\Gamma^\mu\nabla\dot\phi-(\Gamma^\mu\pa_t\dot\phi)\nabla\pa_t\phi,
\Gamma^\mu\dot w\rangle_+(\tau)d\tau
\leq C\ve^3\ln t+C\ve\int_{1/\ve}^t\langle\tau\rangle^{-1}E_k(\tau)d\tau\eqno{(4.47)}
$$
and
\begin{align*}
&\langle (\nabla\cdot w)\Gamma^\mu F_2(\dot\xi,2\xi_a+\dot\xi)-\nabla F_1(\xi)\cdot\Gamma^\mu\dot w,
\Gamma^\mu\dot\theta\rangle_+(t)\\
&\quad+\langle(\Gamma^\mu F_2(\dot\xi,2\xi_a+\dot\xi)-\Gamma^\mu\dot\theta)\nabla F_1(\xi)-\Gamma^\mu F_2(\dot\xi,2\xi_a+\dot\xi)\nabla\theta,\Gamma^\mu \dot w\rangle_+(t)\\
&\leq C\ve^2\langle t\rangle^{-1}E_k(t).\tag{4.48}
\end{align*}
Substituting (4.46)-(4.48) into (4.45) and further combining with (4.44) yield
$$
\int_{1/\ve}^t\sum_{j=0}^2\langle(\pa_j A_j)\zeta^\mu,\zeta^\mu\rangle(\tau) d\tau
\leq C\ve^3\ln t+C\ve\int_{1/\ve}^t\langle\tau\rangle^{-1}E_k(\tau)d\tau.\eqno{(4.49)}
$$
Based on the estimates of the above three steps, if we set $N(t)=\ds\sum_{|\mu|\leq k}\langle A_0\zeta^\mu,
\zeta^\mu\rangle(t)-2\t G_1(t)$, then
$$
N(t)\leq C\ve^3\ln t+C\ve\int_{1/\ve}^t\langle\tau\rangle^{-1}E_k(\tau)d\tau
+C\ve^2\int_{1/\ve}^t\langle\tau\rangle^{-5/2}E_k^{1/2}(\tau)d\tau.\eqno{(4.50)}
$$
Note that $\tilde N(t)=N(t)+\ve^4\langle t\rangle^{-3}$ is equivalent to $E_k(t)+\ve^4\langle t\rangle^{-3}$,
then one can derive from (4.50)
$$
\tilde N(t)\leq C\ve^3\ln t+C\ve\int_{1/\ve}^t\langle\tau\rangle^{-1}\tilde N(\tau)d\tau
+C\ve^2\int_{1/\ve}^t\langle\tau\rangle^{-5/2}\tilde N^{1/2}(\tau) d\tau.\eqno{(4.51)}
$$
Set $g(t)=C\ve^3\ln t+C\ve\int_{1/\ve}^t\langle\tau\rangle^{-1}\tilde N(\tau) d\tau+C\ve^2\int_{1/\ve}^t\langle\tau\rangle^{-5/2}\tilde N^{1/2}(\tau) d\tau$, it follows from (4.51) that
$$
g'(t)\leq C\ve\langle t\rangle^{-1}\big(g(t)+\ve^2\big)+C\ve^2\langle t\rangle^{-5/2}g(t)^{1/2},
$$
furthermore, if we set $\tilde g(t)=g(t)+\ve^2$, then
$$
\tilde g'(t)\leq C\ve\langle t\rangle^{-1}\tilde g(t)+C\ve^2\langle t\rangle^{-5/2}\tilde g(t)^{1/2},
$$
thus, $g(t)\leq \tilde g(t)\leq C\ve^{2}\langle t\rangle^{C\ve}$ and further $E_k(t)\leq C\ve^2\langle t\rangle^{C\ve}$.
\qquad $\square$\bigskip
Based on Lemma 4.3, we next derive the uniform energy estimate on the solution $(\dot\th, \dot w, \dot z)$ of (2.4).
{\bf Lemma 4.4.} {\it
For a fixed integer $\la$ with $\la\geq 9$, it is assumed that $(\dot\theta, \dot w, \dot z)$ is a smooth solution of
(2.4) for $(t,x)\in [1/\ve, T]\times\Bbb R^2$.
If $E_\la(t)\leq\ve^2$ for $t\in[1/\ve,T]$, then $E_\la(t)\leq \frac{1}{2}\ve^2$ holds for small $\ve>0$.
}
\bigskip
{\bf Proof.}
Similar to the proof of Lemma 4.3, we will divide the whole proof procedure into three steps.
In the following, we always assume $|\mu|\leq\la$ and apply the same notations in (4.5).
First, by Lemma 2.2 and the assumption of $E_\la(t)\leq\ve^2$ for $t\in[1/\ve,T]$, we can get
$$
|H_j^\mu(t)|_-\leq C\ve^3\langle t\rangle^{-5/2},\quad j\in\{1,2,4,5,7,8,10,11\}.\eqno{(4.52)}
$$
Since $|\Gamma^\nu\dot w(t)|_-+|\Gamma^\nu\dot\theta(t)|_-\leq C\langle t\rangle^{-1/2}\widetilde Q_{\mu+2}(t)\leq C\ve\langle t\rangle^{-1/2+\delta}$ holds by Lemma 4.3 for sufficient small positive number $\delta$, we have
$$
|H_j^\mu(t)|_-\leq C\ve^3\langle t\rangle^{-3/2+\delta},\quad j\in\{3,6,9,12\}.\eqno{(4.53)}
$$
In addition, from (4.9) and (4.10) we can obtain $\|\hat f^\mu_{13}\|\leq C\ve^2\langle t\rangle^{-1}$ and further
$$
|H_{13}^\mu(t)|\leq C\ve^2\langle t\rangle^{-1}\sup_{|x|\leq M+1}|\Gamma^b\dot w|\leq C\ve^3\langle t\rangle^{-3/2+\delta}.\eqno{(4.54)}
$$
On the other hand, it follows (3.21) and (3.23)-(3.27) that
$$
\sum_{j=1}^{12} |H_j^\mu(t)|_+\leq C\ve^3\langle t\rangle^{-3/2+\delta}.\eqno{(4.55)}
$$
Second, it is easy to get
$$
|\hat H^\mu_{aj}(t)|\leq C\ve^3\langle t\rangle^{-5/2}\quad\text{for}\quad j=1,2.\eqno{(4.56)}
$$
Lemma 2.5 gives $|\Gamma^\nu\dot w(t,x)|\leq C\ve\langle t\rangle^{-3/2+\delta}$
for $|x|\leq M+1$ and $|\nu|\leq\mu$, and hence
$$
|\hat H^\mu_{j}(t)|\leq C\ve^3\langle t\rangle^{-3/2+\delta}\quad\text{for}\quad j=1,2.\eqno{(4.57)}
$$
Third, as in the proof of Lemma 4.3, one has
$$
|\sum_{j=0}^2\langle(\pa_j A_j)\zeta^\mu,\zeta^\mu\rangle_-(t)|\leq C\ve^3\langle t\rangle^{-3/2}.\eqno{(4.58)}
$$
Similar to the estimate (3.24) on $I_3^{\mu\nu}$, we have
$$
\|(\pa_t\nabla\phi\cdot\Gamma^\mu\nabla\dot\phi-\triangle\phi\Gamma^\mu\pa_t\dot\phi)(t)\|_+\leq C\ve\langle t\rangle^{-3/2}E_{\la+1}^{1/2}\leq C\ve^2\langle t\rangle^{-3/2+\delta}\eqno{(4.59)}
$$
and
$$
\|(\triangle\phi\Gamma^\mu\nabla\dot\phi-\Gamma^\mu\pa_t\dot\phi\nabla\pa_t\phi)(t)\|_+\leq C\ve\langle t\rangle^{-3/2}E_{\la+1}^{1/2}\leq C\ve^2\langle t\rangle^{-3/2+\delta}.\eqno{(4.60)}
$$
In addition, we can decompose
\begin{align*}
&\langle (\nabla\cdot w)\Gamma^\mu F_2(\dot\xi,2\xi_a+\dot\xi)-\nabla F_1(\xi)\cdot\Gamma^\mu\dot w,\Gamma^\mu\dot\theta\rangle_+(t)\\
&\quad +\langle(\Gamma^\mu F_2(\dot\xi,2\xi_a+\dot\xi)-\Gamma^\mu\dot\theta)\nabla F_1(\xi)-\Gamma^\mu F_2(\dot\xi,2\xi_a+\dot\xi)\nabla\theta,\Gamma^\mu\dot w\rangle_+(t)\\
&\equiv A^\mu_1(t)+A^\mu_2(t)+A^\mu_3(t),\tag{4.61}
\end{align*}
where
\begin{align*}
|A^\mu_1(t)|&=|\langle\triangle\phi\Gamma^\mu F_2(\dot\xi,2\xi_a+\dot\xi)-2\nabla F_1(\xi)\cdot
\Gamma^\mu\nabla \dot\phi,\Gamma^\mu F_2(\dot\xi,2\xi_a+\dot\xi)\rangle_+|\\
&\leq C\ve^3\langle t\rangle^{-3/2}E_\la(t)\leq C\ve^5\langle t\rangle^{-3/2},\tag{4.62}\\
|A^\mu_2(t)|&=|\langle\nabla\pa_t\phi\cdot\Gamma^\mu\nabla\dot\phi-\triangle\phi\Gamma^\mu\pa_t
\dot\phi,\Gamma^\mu F_2(\dot\xi,2\xi_a+\dot\xi)\rangle_+|\leq C\ve^4\langle t\rangle^{-2+\delta},\tag{4.63}\\
|A^\mu_3(t)|&=|2\langle\nabla F_1(\xi)\cdot\Gamma^\mu\nabla\dot\phi,\Gamma^\mu\pa_t\dot\phi\rangle_+|\leq C\ve^2\langle t\rangle^{-3/2}E_\mu(t)\leq C\ve^4\langle t\rangle^{-3/2},\tag{4.64}
\end{align*}
here we point out that the estimate of $\nabla F_1(\xi)$ in $A^\mu_3(t)$ follows from
$$
|\pa_l F_1(\xi)(t)|_+=|\big(\pa_t\phi\pa_l\pa_t\phi-\nabla \phi\cdot\nabla\pa_l \phi\big)(t)+\pa_l\big(F_1(\xi)(\theta-\frac{1}{2}F_1(\xi))\big)(t)|_+
\leq C\ve^2\langle t\rangle^{-3/2}.
$$
Therefore, by substituting (4.62)-(4.64) into (4.61) and further combining with (4.59)-(4.60),
it follows from (4.45) that
$$
|\sum_{j=0}^2\langle(\pa_j A_j)\zeta^\mu,\zeta^\mu\rangle_+(t)|
\leq C\ve^3\langle t\rangle^{-3/2+\delta}.\eqno{(4.65)}
$$
Finally, inserting (4.52)-(4.58) and (4.65) into (4.4) and combining with basic
energy inequalities
(similar to (2.12) and (3.22)) yield
$$\ds\frac{ d}{ dt}E_\mu(t)\leq C\ve^3\langle t\rangle^{-3/2+\delta}.\eqno{(4.66)}$$
In addition, we have $E_\mu(\ds\frac{1}{\ve})\leq C\ve^4$ by Lemma 4.1. This,
together with (4.66),
derives $E_\mu(t)\leq C\ve^3$ when we choose $\delta<\ds\frac{1}{2}$. Therefore,
$E_\mu(t)\leq \ds\f12\ve^2$ holds
for small $\ve>0$.\qquad $\square$
Finally, we start to show Theorem 1.1.
\bigskip
{\bf Proof of Theorem 1.1.} By Lemma 4.1 and Lemma 4.4, we know that (2.4) admits a
global smooth solution $(\dot\th, \dot w,
\dot z)$ in terms of the continuity induction method. Thus, (2.3) has a global smooth
solution $(\th, w, z)$
since the smooth solution of (2.1) exists globally. Therefore, Theorem 1.1 is proved.
\qquad \qquad $\square$
\vskip 0.4 true cm
|
\section{Introduction}
\label{sec:overview}
Morphoid type theory (MorTT) is a type-theoretic foundation for mathematics supporting isomorphism and distinguishing natural
maps from general functions. Morphoid type theory has been developed independently of Voevodsky's univalent foundations program
as realized in homotopy type theory (HoTT) \cite{HOTT}. HoTT is also a type-theoretic
foundation for mathematics supporting isomorphism.
This introduction is organized into two parts. The first discusses the general nature of isomorphism and the second discusses
the relationship between MorTT and HoTT.
\bigskip
\centerline{\large \bf Isomorphism}
\medskip
The notion of isomorphism in mathematics seems related to the notion of an application programming interface (API) in computer software.
An API specifies what information and behavior an object provides.
Two different implementations can produce identical behavior when interaction is restricted to that allowed by the API.
For example, textbooks on real analysis typically start from axioms involving multiplication, addition, and ordering.
Addition, multiplication and ordering define an abstract interface --- the well-formed statements
about real numbers are limited to those that can be defined in terms of the operations of the interface.
We can implement real numbers in different ways ---
as Dedekind cuts or Cauchy sequences. However, these different implementations provide identical behavior as viewed through the
interface --- the different implementations are isomorphic as ordered fields.
The axioms of real analysis specify the reals up to isomorphism for ordered fields. The second order Peano axioms
similarly specify the structure of the natural numbers up to isomorphism.
{\bf Isomorphism and Dependent Pair Types.} The general notion of isomorphism is best illustrated by considering dependent pair types.
Here we will write a dependent pair type as $\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}$
where the instances of this type are the pairs $\mathbf{Pair}(x,y)$ where $x$ is an instance of the type $\sigma$ and $y$ is an instance of $\tau[x]$.\footnote{We avoid the
more standard notation $\Sigma_{\intype{x\;}{\;\sigma}}\;\tau[x]$ so as to make the treatment more accessible to those readers not already familiar with type theory.}
The type of directed graphs can be written as $\pairtype{{\cal N}}{\mathrm{\bf type}}{\intype{P}{({\cal N} \times {\cal N}) \rightarrow \mathrm{\bf Bool}}}$
where ${\cal N}$ is a type representing the set of nodes of the graph and $P$ is a binary predicate on the nodes giving the edge relation.
Two directed graphs $\mathbf{Pair}({\cal N},P)$ and $\mathbf{Pair}({\cal M},Q)$ are isomorphic if there exists a bijection
from ${\cal N}$ to ${\cal M}$ that carries $P$ to $Q$. Some bijections will carry $P$ to $Q$ while others will not.
Two pairs $\mathbf{Pair}(a,b)$ and $\mathbf{Pair}(a',b')$ of a general dependent pair type
$\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}$ are isomorphic if there is a $\sigma$-isomorphism from $a$ to $a'$ that carries $b$ to $b'$.
Again consider two instances $\mathbf{Pair}(a,b)$ and $\mathbf{Pair}(a',b')$ of the type
$\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}$. When $\mathbf{Pair}(a,b)$ and $\mathbf{Pair}(a',b')$ are isomorphic in this type
we have that some $\sigma$-isomorphisms from $a$ to $a'$ will carry $b$ to $b'$ while others will not.
This implies that to define isomorphism at general dependent pair types
we need that for any type $\sigma$, and for any two isomorphic values $a$ and $a'$ of type $\sigma$, we can define the full set of $\sigma$-isomorphisms
from $a$ to $a'$.
As discussed in more detail below, MorTT takes type-isomorphisms to be bijections. This interpretation of type-isomorphism
gives the standard notion of isomorphism for dependent pair types of the form
$\mathbf{PairOf}(\intype{\alpha}{\mathrm{\bf type}},\;\intype{x}{\tau[\alpha]})$. The types ``graph'', ``group'' and ``topological space'' can all be written in this form
where $\tau[\alpha]$ is written in the language defined in section~\ref{sec:rules}.
{\bf Isomorphism as Observational Equivalence.}
Dedekind cuts and Cauchy sequences are ``observationally equivalent'' implementations of
the real numbers --- these different implementations cannot be distinguished by well-typed properties of ordered fields.
We can approach the concept of isomorphism by first defining the set of
well-typed properties and then seeking a notion of isomorphism which is as coarse as possible (which equates as many things as possible) while ensuring that
equated objects (isomorphic objects) are indistinguishable by well-typed properties.
When isomorphism is approached in this way, the concept of isomorphism (observational equivalence)
arises from, or is intuitively defined by,
the set of well-typed properties.
MorTT starts by defining a core system of inference rules which intuitively define the well-typed expressions,
including the well-typed properties of instances of a type $\tau$. The well-typed properties of instances of type $\tau$
correspond to the well-typed Boolean expressions $\Phi[x]$ for a variable $x$ of type $\tau$.
By specifying the set of observable properties for each type, and hence a notion of observational equivalence, the core inference rules (given in figures~\ref{fig:expressions},
\ref{fig:rules} and \ref{fig:pairs}) essentially dictate an appropriate notion of isomorphism.
{\bf Natural Maps and Voldemort's Theorem.}
There are many situations in mathematics where a type can be shown to be inhabited (have a member)
even though there is no natural or canonical member of that type. A classical example is that
for every finite dimensional vector space $V$ there is an isomorphism (a linear bijection) between $V$ and its dual $V^*$.
However, there is no natural or canonical isomorphism.
This is closely related to the fact that a vector space has no natural or canonical basis.
A simpler example is that there is no natural or canonical point on the topological circle $S^1$.
Formally this phenomenon can be captured by distinguishing general functions, functions whose definitions may require symmetry breaking,
from natural maps --- functions free of choices and definable without symmetry breaking.
The statement that there is no natural point on the topological circle $S^1$
corresponds to the fact (in MorTT) that for a variable $X$ ranging over topological spaces there there is no well-typed expression $e[X]$
such that $e[S^1]$ is a point on $S^1$ --- no point on the circle can be named.
An expression $e[X]$ that is well-typed
for a variable $X$ ranging over topological spaces is called a topological invariant --- a property of the topology such as the fundamental homotopy group.
We can prove that two topologies are different (not homeomorphic) by finding a topological invariant that distinguishes them.
Section~\ref{sec:natural} gives inference rules defining natural maps and section~\ref{sec:Voldemort} states Voldemort's theorem
which implies, for example, that there is no natural point on $S^1$ and no natural linear bijection between a finite dimensional vector space and its dual.
{\bf Cryptomorphism.}
Two types $\sigma$ and $\tau$ are cryptomorphic in the sense of Birkoff and Rota \cite{Rota} if they ``present the same data''.
For example a group can be defined as a four-tuple of a set, a group operation, an identity element and an inverse operation satisfying certain equations.
Alternatively, a group can be defined as a pair of a set and a group operation such that an identity element and an inverse elements exist.
These are different types with different elements (four-tuples vs.$\!$ pairs). However, these two types present the same data. Rota was fond of pointing out
the large number of different ways one can formulate the concept of a matroid. Any type theoretic foundation
for mathematics should account formally for this phenomenon. Here we suggest that two types are crytomorphic if there exist natural maps
$\intype{f}{\sigma \hookrightarrow \tau}$
and $\intype{g}{\tau \hookrightarrow \sigma}$ such that $f \circ g$ and $g \circ f$ are the identity natural maps on $\tau$ and $\sigma$ respectively.
This is discussed in section~\ref{sec:cryptomorphism}.
{\bf The Autonomy of Mathematics.}
MorTT has been developed from the perspective that mathematics exists independent of foundations.
The axioms of ZFC have a distinguished status in mathematics because they reflect pre-formal human intuition. This should also be true of a
type-theoretic foundation. MorTT has been developed to reflect mathematics, and possibly automate it,
but not to change it.
\bigskip
\centerline{\large \bf MorTT vs.$\!$ HoTT}
\medskip
Even a high level discussion of homotopy type theory (HoTT) is technical and difficult.
Readers not already familiar with HoTT should feel free to skip to section~\ref{sec:rules}.
\medskip
The fundamental difference between MorTT and HoTT is that MorTT extends classical predicate calculus while HoTT extends constructive type theory.
More specifically, HoTT extends Martin-L\"of type theory \cite{MLTT1971,COC} with Voevodsky's univalence axiom \cite{SimpSets} thus providing a treatment of isomorphism.
Martin-L\"of type theory is a formulation of constructive logic following Brauwer's constructivist program.
To accommodate classical (nonconstructive) inference, HoTT can be extended with a version of the law of the excluded middle and a version of the (nonconstructive) axiom of choice.
Here we take the law of the
excluded middle and the non-constructive axiom of choice to be self evident and consider only the classical version of HoTT.
HoTT inherits legacy features of constructive logic which have consequences for
the notion of isomorphism. Most significantly, HoTT inherits the representation of propositions as types.
Equality propositions are particularly significant in this respect. In MorTT one writes $G=_{\mathbf{Group}} H$
for the proposition that the group $G$ is group-isomorphic to the group $H$. In MorTT this is a Boolean proposition --- it is true or false.
In HoTT the expression $G =_{\mathbf{Group}} H$ is a type called an identity type --- it is the type whose elements are the group-isomorphisms
from $G$ to $H$.
This identity type is inhabited (has a member) if and only if $G$ is group-isomorphic to $H$.
The properties observable in MorTT
(the properties observable with Boolean-equality) are more limited than the properties observable in HoTT (the properties observable with identity types).
This leads to a notion of isomorphism in MorTT that is more abstract (is coarser) than that in HoTT. This is discussed in more detail below.
The classical version of HoTT also inherits the feature of constructive logic that each value is a member of only a single type.
In MorTT a single object can be a member of various types. For example, in MorTT a single group can be a member of the distinct types
``Abelian group'' and ``group''.
The classical version of HoTT also inherits Martin L\"off's axiom for equality. In the HoTT community this has come to be called path induction.
It has traditionally been called axiom J. This is a complex and subtle axiom.
MorTT, in contrast, inherits the classical axioms of reflexivity, symmetry, transitivity and substitution with direct compositional semantics.
In MorTT there is no path induction.
To accommodate the classical notion of isomorphism, HoTT includes ``squashing''.
An identity type can be squashed to a ``mere-proposition''. A mere-proposition is a type $\Phi$ (interpreted as a proposition)
such that for $\intype{x,y}{\Phi}$ we have $x =_{\Phi} y$.
A mere-proposition is Boolean in the sense of either being empty (false) or having only a single element (up to isomorphism).
In MorTT all propositions are Boolean and the MorTT inference rules do not involve squashing.
In HoTT, squashing also allows general types to be squashed to a special class of types called sets. It is important to note, however,
that the set-type distinction of HoTT is different from the traditional set-class distinction of set theory. This is in contrast to
MorTT where the set-class distinction corresponds to the classical set-theoretic distinction.
In MorTT all sets $\sigma$ (in the sense of the
traditional set-class distinction) are discrete in the sense that all the equivalence classes of the equivalence relation $=_\sigma$ are semantically singleton sets.
In MorTT only classes (and higher order types), such as the class of all groups, have non-trivial notions of isomorphism.
Using the terminology of HoTT, MorTT has the property that all types in $U_0$ are extensional in that propositional equality implies judgmental equality.
MorTT also includes an axiom of infinity stating that infinite sets (infinite types in $U_0$) exist.
{\bf Type-Isomorphism vs.$\!$ Cryptomorphism.}
In both MorTT and HoTT, closed types
(type expressions without free variables) have groupoid structure. The groupoid denoted by a closed type expression $\sigma$ consists of the set of elements of type
$\sigma$ together with the $\sigma$-isomorphisms between them. A groupoid is a category in which every morphism is an isomorphism. In MorTT the groupoid structure of a type $\alpha$ is not
observable by well-typed Boolean formulas $\Phi[\alpha]$. In MorTT two types are type-isomorphic if
they have the same cardinality (the same number of equivalence classes). The types
``finite graph'' and ``finite total order'' have different groupoid structure but the same cardinality --- they both have a countably infinite number of isomorphism classes.
These types are type-isomorphic in MorTT. In MorTT the difference in groupoid structure, although not observable by propositions on type variables,
still blocks the existence of certain natural maps. These types are type-isomorphic but are not cryptomorphic. MorTT makes a fundamental distinction between
type-isomorphism (same cardinality) and cryptomorphism (defined by a pair of natural maps).
In contrast, HoTT allows the groupoid structure of a type
to be observed by propositions on type variables.
In HoTT two types are type-isomorphic only when they have the same higher-order groupoid structure.
This has the consequence that two directed graphs of type $\mathbf{PairOf}(\intype{{\cal N}}{\mathrm{\bf type}},\;({\cal N} \times {\cal N}) \rightarrow
\mathrm{\bf Bool})$ fail to be isomorphic unless the node types have the same higher order groupoid structure. Directed graphs of type
$\mathbf{PairOf}(\intype{{\cal N}}{\mathbf{set}},\;({\cal N} \times {\cal N}) \rightarrow \mathrm{\bf Bool})$ have the standard notion of isomorphism.
As noted above, however, the set-type distinction of HoTT is different from the familiar set-class distinction of set theory.
In HoTT the cryptomorphism is not differentiated from type-isomorphism --- in HoTT type-isomorphism
itself is defined through the existence of a pair
of natural maps.
{\bf Dependent Functors vs.$\!$ Morphoids.}
It is useful to compare the groupoid model of Martin-L\"of type theory \cite{GRPD} with the morphoid model presented here.
The groupoid model of Martin L\"of type theory is simpler than, but related to, the simplicial set model of HoTT \cite{SimpSets}.
A central issue in any type-theoretic account of isomorphism is defining the semantics of type expressions with free variables.
Consider a type expression $\tau[x]$ containing the single free variable $x$ of type $\sigma$
where $\sigma$ is closed. In the groupoid model the closed type expression $\sigma$ denotes a groupoid.
The open type expression $\tau[x]$, however, is interpreted as a functor from the groupoid $\sigma$ into the groupoid GRPD --- the category of groupoids and their isomorphisms.
For an object $a$ in the groupoid $\sigma$ we have that $\tau[a]$ is a groupoid (a type). But the functor $\tau[x]$ maps morphisms as well as objects.
For any morphism (isomorphism) $\rho$ of $\sigma$ from object $a$ to object $b$ we have that $\tau[\rho]$ is a groupoid-isomorphism from the groupoid
$\tau[a]$ to the groupoid $\tau[b]$. Now consider a term $e[x]$ of type $\tau[x]$. In the groupoid model $e[x]$ is interpreted as a ``dependent functor''
where for an object $a$ of $\sigma$ we have that $e[a]$ is an object in the groupoid $\tau[a]$. But again, the functor $e[x]$ maps morphisms as well as objects.
For a morphism $\rho$ of the groupoid $\sigma$
from object $a$ to object $b$ we have that $e[\rho]$ is a morphism of $\tau[b]$ from the object $\tau[\rho](e[a])$ to the object $e[b]$.
The dependent functor $\intype{e[x]}{\tau[x]}$ satisfies the functorial equations
$$e[id_{a}] = id_{e[a]}$$
$$e[\gamma \circ \rho] = e[\gamma] \circ \tau[\gamma](e[\rho]).$$
The groupoid model can be contrasted with morphoid semantics. In morphoid semantics all semantic values are ``morphoids''.
Morphoids are defined recursively such that a morphoid is either a point (a morphoid ur-element), a pair of morphoids, a
set of morphoids satisfying certain conditions (a type), or a function from morphoids to morphoids satisfying certain conditions.
Every morphoid $x$ has a left interpretation $\mathbf{Left}(x)$, a right interpretation $\mathbf{Right}(x)$, and an inverse $x^{-1}$, all of which are also morphoids.
For any two morphoids $x$ and $y$ with $\mathbf{Right}(x) = \mathbf{Left}(y)$ we have the composition $x \circ y$ which is a morphoid.
The class of all morphoids satisfies the algebraic properties of a groupoid under these operations.
Since all values are morphoids, there is no need to define separate object and morphism value functions.
Instead we have a classical Tarskian
semantic value function where we write $\convalue{e}\rho$ for the semantic value of the expression $e$ where $\rho$ specifies a morphoid value for each free variable of $e$.
In MorTT the semantic value of a pair type is defined by
\begin{eqnarray}
\label{eq:meaning1}
\convalue{\mathbf{Pairof}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}\rho & = & \{\mathbf{Pair}(a,b):\;a \in \convalue{\sigma}\rho,\;b \in \convalue{\tau[x]}\rho[x \leftarrow a]\}.
\end{eqnarray}
For a morpoid $a\in\sigma$ the type $\tau[a]$ is a set of morphoids. However, every morphoid $a$ is an isomorphism from
$\mathbf{Left}(a)$ to $\mathbf{Right}(a)$. Each element of $\tau[a]$ is an isomorphism from an element of $\tau[\mathbf{Left}(a)]$ to an element of $\tau[\mathbf{Right}(a)]$.
We can then have $b \in \tau[a]$ but $b^{-1} \not \in \tau[a]$.
So $\tau[a]$ is a set of morphoids that is generally not closed under inverse and hence is not a groupoid. Instead, type expressions denote sets $u$
that satisfy the morphoid closure condition that for $b,c,d \in u$, with $b \circ c^{-1} \circ d$ defined,
we have $b \circ c^{-1} \circ d \in u$. It is possible to define $\mathbf{Left}$, $\mathbf{Right}$, inverse and composition on morphoid-closed sets
such that the groupoid equations are satisfied by these sets. The sets themselves are morphoids.
{\bf Abstract Homotopy Theory.}
Abstract homotopy theory arises in HoTT from identity types. In HoTT the elements of the identity type $a = _\sigma b$
are objects and for two elements $\intype{x,y}{(a =_\sigma b)}$ we can form the identity type
$x =_{(a =_\sigma b)} y$. This second order identity type has members and for $\intype{u,v}{(x =_{a =_\sigma b} y)}$
we have a third order identity type $u =_{(x =_{(a =_\sigma b)} y)} v$. We can carry this to arbitrarily high order
leading to a mathematical structure equivalent to abstract homotopy theory.
In MorTT there are types of the form $\mathrm{\bf iso}(\sigma,a,b)$ whose elements are the $\sigma$-isomorphisms from $a$ to $b$. However, MorTT distinguishes the Boolean
equation $a =_\sigma b$ from the type $\mathrm{\bf iso}(\sigma,a,b)$. Also, in MorTT we have that $\mathrm{\bf iso}(\sigma,a,b)$ is a subtype (literally a subset)
of $\sigma$ and for
$\intype{x,y}{\mathrm{\bf iso}(\sigma,a,b)}$ we have $\intype{x,y}{\sigma}$ with $x =_\sigma y$ and even $x =_{\mathrm{\bf iso}(\sigma,a,b)} y$.
Morphoid semantics does not involve abstract homotopy theory. Of course, as with any branch of mathematics, homotopy theory
can be formulated in MorTT.
{\bf Consistency vs.$\!$ Meaning.} When formal notation is introduced in mathematics
the meaning of the notation is generally defined. Equation (\ref{eq:meaning1})
defines the meaning of the pair type notation assuming that the meanings of certain parts of the notation are already defined.
The HoTT book \cite{HOTT} attempts to give intuitions for the type formalism in terms of homotopy theory but does not attempt to rigorously
define the meaning of the formal notation. While a rigorous semantics has been given for HoTT \cite{SimpSets}, the existence of this semantics
is only mentioned in passing in the HoTT book and only in noting that proofs of consistency exist.
Practitioners are not expected to think about rigorously defined meaning. Consistency is important, but Platonic thought requires meaning. Platonic thought
seems essential to the practice of mathematics.
\eject
\section{Rules and Semantics}
\label{sec:rules}
The core rules of morphoid type theory are described in section~\ref{sec:core} and
given in figures~\ref{fig:expressions} through~\ref{fig:additional}. Rules for deriving isomorphisms are given in section~\ref{sec:isorules}.
Section~\ref{subsec:semantics} gives a top level specification of the semantics
of MorTT. More specifically, section~\ref{subsec:semantics} defines a value function $\semvalue{e}\rho$ and a semantic entailment relation $\models$
but where a few key constructs used in these definitions are defined subsequently in section~\ref{sec:Morphoids}.
Sections~\ref{sec:natural} through~\ref{sec:cryptomorphism} present natural maps, Voldemort's theorem and Crypotomorphism
in terms of the inference rules.
\subsection{The Core Rules}
\label{sec:core}
Morphoid type theory starts from the syntax and semantics of classical predicate calculus.
In sorted first order logic every term has a sort and each function symbol $f$ specifies the sorts of its arguments
and the sort of its value.
We write $\intype{f}{\sigma_1 \times \cdots \times \sigma_n \rightarrow \tau}$ to indicate that $f$ is a function that takes
$n$ arguments of sort $\sigma_1$, $\ldots$, $\sigma_n$ respectively and which produces a value of sort $\tau$.
The syntax of sorted first order logic can be defined by the following grammar where function and predicate applications
must satisfy the sort constraints associated with the function and predicate symbols.
\begin{eqnarray*}
t & ::= & x \;|\; c \;|\; f(t_1,\ldots, t_n) \\
\\
\Phi & ::= & P(t_1,\ldots,t_n) \;|\; t_1 =_\sigma t_2 \\
& & |\; \Phi_1 \vee \Phi_2 \;|\; \neg \Phi \;|\; \forall \intype{x}{\sigma}\;\Phi[x]
\end{eqnarray*}
Note that in the above grammar the equality symbol $=_\sigma$ is subscripted with a sort $\sigma$ to which it applies.
The labeling of equality with sorts is important for the treatment of isomorphism.
Given this basic grammar it is standard to introduce the following abbreviations.
\begin{eqnarray*}
\Phi \wedge \Psi & \equiv & \neg(\neg \Phi \vee \neg \Psi) \\
\Phi \Rightarrow \Psi & \equiv & \neg \Phi \vee \Psi \\
\Phi \Leftrightarrow \Psi & \equiv & (\Phi \Rightarrow \Psi) \wedge (\Psi \Rightarrow \Phi) \\
\exists \intype{x}{\sigma}\;\Phi[x] & \equiv & \neg \forall \intype{x}{\sigma}\;\neg \Phi[x] \\
(\exists \intype{x}{\sigma}) & \equiv & \exists \intype{x}{\sigma}\;x=_\sigma x \\
\;\exists!\intype{x}{\sigma}\;\Phi[x] & \equiv & \left\{\begin{array}{l} \exists \intype{x}{\sigma}\;\Phi[x] \\
\wedge \forall \intype{x,y}{\sigma} \\
\;\;\;\;\; \Phi[x] \wedge \Phi[y] \Rightarrow x=_\sigma y \end{array}\right.
\end{eqnarray*}
\begin{figure}[t]
\begin{framed}
{\footnotesize
~ \hfill
$\epsilon \vdash \mathrm{\bf True}$
\hfill
$\begin{array}{l} \epsilon \vdash \intype{\mathrm{\bf type}_j}{\mathrm{\bf type}_i} \\
\mbox{for $j < i$} \\
\\
\mathbf{Set} \equiv \mathrm{\bf type}_0 \\
\mathbf{Class} \equiv \mathrm{\bf type}_1
\end{array}$
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{\tau}{\mathrm{\bf type}_i}}
\ant{\mbox{$x$ not declared in $\Sigma$}}}
{\ant{\Sigma;\;\intype{x}{\tau} \vdash \mathrm{\bf True}}}
\hfill
$\epsilon \vdash \intype{\mathrm{\bf Bool}}{\mathbf{Set}}$
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{\Phi}{\mathrm{\bf Bool}}}}
{\ant{\Sigma;\Phi \vdash \mathrm{\bf True}}}
\hfill ~
\vspace{-1em}
~ \hfill
\unnamed{
\ant{\Sigma;\Theta \vdash \mathrm{\bf True}}}
{\ant{\Sigma;\Theta \vdash \Theta}}
\hfill
\unnamed
{\ant{\Sigma;\Theta \vdash \mathrm{\bf True}}
\ant{\Sigma \vdash \Psi}}
{\ant{\Sigma;\Theta \vdash \Psi}}
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{\tau}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \intype{\sigma}{\mathrm{\bf type}_i}}}
{\ant{\Sigma \vdash \intype{(\tau \rightarrow \sigma)}{\mathrm{\bf type}_i}}}
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{f}{\sigma \rightarrow \tau}}
\ant{\Sigma \vdash \intype{e}{\sigma}}}
{\ant{\Sigma \vdash \intype{f(e)}{\tau}}}
\hfill ~
\vspace{-1em}
~ \hfill
\unnamed{\ant{\Sigma \vdash \intype{\Phi}{\mathrm{\bf Bool}}}
\ant{\Sigma \vdash \intype{\Psi}{\mathrm{\bf Bool}}}}
{\ant{\Sigma \vdash \intype{(\Phi \vee \Psi)}{\mathrm{\bf Bool}}}
\ant{\Sigma \vdash \intype{\neg \Phi}{\mathrm{\bf Bool}}}}
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{\tau}{\mathrm{\bf type}_i}}
\ant{\Sigma;\;\intype{x}{\tau} \vdash \intype{\Phi[x]}{\mathrm{\bf Bool}}}}
{\ant{\Sigma \vdash \intype{(\forall\intype{x}{\tau}\;\Phi[x])}{\mathrm{\bf Bool}}}}
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{\tau}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \intype{w}{\tau}}
\ant{\Sigma \vdash \intype{u}{\tau}}}
{\ant{\Sigma \vdash \intype{(w =_\tau u)}{\mathrm{\bf Bool}}}}
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{\sigma}{\mathrm{\bf type}_i}}}
{\ant{\Sigma \vdash \intype{\sigma}{\mathrm{\bf type}_j}\;\mbox{for}\;j> i}}
\hfill ~
}
\caption{{\bf Predicate Calculus Expressions}.
Here $\mathrm{\bf type}_0$, $\mathrm{\bf type}_1$, $\mathrm{\bf type}_2$, $\ldots$ are distinct constants and $\epsilon$ is a constant denoting the empty context.
The sequent $\Sigma \vdash \mathrm{\bf True}$ states that $\Sigma$ is a well-formed context.
The sequent $\epsilon \vdash \mathrm{\bf True}$ states that the empty context is well formed.
The requirement of $j < i$ in the second rule is needed to avoid Russel's paradox.
We will write $\mathbf{Set}$ as an alternate notation for $\mathrm{\bf type}_0$ and $\mathbf{Class}$ as an alternate notion for $\mathrm{\bf type}_1$.
Note that we have $\epsilon \vdash \intype{\mathbf{Set}}{\mathbf{Class}}$.
The first three rules
of the first row allow us to derive $\epsilon;\intype{\alpha}{\mathbf{Set}} \vdash \mathrm{\bf True}$ thereby declaring a primitive set.
We can then declare additional symbols such as $\intype{c}{\alpha}$ or $\intype{P}{\alpha \rightarrow \mathrm{\bf Bool}}$ which together
give $\intype{P(c)}{\mathrm{\bf Bool}}$.
A rule with multiple conclusions abbreviates multiple rules each with the same antecedents.
Dependent pair types, introduced in figure~\ref{fig:pairs}, allows us to define the type $\mathbf{Group}$ such that we have
$\epsilon \vdash \intype{\mathbf{Group}}{\mathbf{Class}}$.
The equality $G =_{\mathbf{Group}} H$ states that $G$ and $H$ are group-isomorphic. For sets $\alpha$ and $\beta$ the
equality $\alpha =_{\mathbf{Set}} \beta$ states that $\alpha$ and $\beta$ have the same cardinality.
}
\label{fig:expressions}
\end{framed}
\end{figure}
\begin{figure}[t]
\begin{framed}
{\footnotesize
\unnamed
{\ant{\Sigma;\Phi \vdash \Psi}
\ant{\Sigma;\neg \Phi \vdash \Psi}}
{\ant{\Sigma \vdash \Psi}}
\hfill
\unnamed
{\ant{\Sigma \vdash \Phi}
\ant{\Sigma \vdash \intype{\Psi}{\mathrm{\bf Bool}}}}
{\ant{\Sigma \vdash \Phi \vee \Psi}
\ant{\Sigma \vdash \Psi \vee \Phi}
\ant{\Sigma \vdash \neg \neg \Phi}}
\hfill
\unnamed
{\ant{\Sigma \vdash \neg \Psi}
\ant{\Sigma \vdash \neg \Phi}}
{\ant{\Sigma \vdash \neg(\Phi \vee \Psi)}}
\hfill
\unnamed
{\ant{\Sigma \vdash \forall \intype{x}{\sigma} \;\Phi[x]}
\ant{\Sigma \vdash \intype{e}{\sigma}}}
{\ant{\Sigma \vdash \Phi[e]}}
\hfill
\unnamed
{\ant{\Sigma;\intype{x}{\sigma} \vdash \intype{\Phi[x]}{\mathrm{\bf Bool}}}
\ant{\Sigma; \intype{x}{\sigma} \vdash \Phi[x]}}
{\ant{\Sigma \vdash \forall \intype{x}{\sigma}\;\Phi[x]}}
\vspace{-2em}
~ \hfill
\unnamed
{\ant{\Sigma \vdash \intype{e}{\tau}}}
{\ant{\Sigma \vdash e =_\tau e}}
\hfill
\unnamed
{\ant{\Sigma \vdash u=_\tau w}}
{\ant{\Sigma \vdash w=_\tau u}}
\hfill
\unnamed
{\ant{\Sigma \vdash u=_\tau w}
\ant{\Sigma \vdash w=_\tau s}}
{\ant{\Sigma \vdash u=_\tau s}}
\hfill
\unnamed
{\ant{\Sigma;\;\intype{x}{\sigma} \vdash \intype{e[x]}{\tau}}
\ant{\mbox{$x$ is not free in $\tau$}}
\ant{\Sigma \vdash w =_\sigma u}}
{\ant{\Sigma \vdash e[w] =_\tau e[u]}}
\hfill ~
\vspace{-2em}
\unnamed{
\ant{\Sigma \vdash \intype{f,g}{\sigma \rightarrow \tau}}
\ant{\Sigma \vdash \forall\intype{x}{\sigma}\;f(x) =_\tau g(x)}}
{\ant{\Sigma \vdash f =_{\sigma \rightarrow \tau}\;g}}
\hfill
\unnamed
{\ant{\Sigma;\;\intype{x}{\sigma};\;\intype{y}{\tau} \vdash \intype{\Phi[x,y]}{\mathrm{\bf Bool}}}
\ant{\mbox{$x$ is not free in $\tau$}}
\ant{\Sigma \vdash \forall \intype{x}{\sigma}\;\exists \intype{y}{\tau}\; \Phi[x,\;y]}}
{\ant{\Sigma \vdash \exists \intype{f}{\sigma \rightarrow \tau}\;\forall \intype{x}{\sigma}\;\Phi[x,f(x)]}}
\hfill
$\epsilon \vdash \left\{\begin{array}{l} \exists \intype{\alpha}{\mathbf{Set}}\;\exists\intype{f}{\alpha \rightarrow \mathbf{Set}} \\
\;\; \forall \intype{x}{\alpha}\; \exists \intype{y}{\alpha} \\
\;\;\;\;\;f(y) =_{\mathbf{Set}} (f(x) \rightarrow \mathrm{\bf Bool})
\end{array}\right.$
}
\caption{{\bf Predicate Calculus Inference Rules.} The first three rules of the first row give a complete set of rules for Boolean logic.
The substitution rule expresses the observational equivalence of isomorphics.
The last row gives the axioms of
extensionality, choice, and infinity.}
\label{fig:rules}
\end{framed}
\end{figure}
\begin{figure}[p]
\begin{framed}
{\scriptsize
\unnamed
{\ant{\Sigma \vdash \intype{\sigma}{\mathrm{\bf type}_i}}
\ant{\Sigma;\;\intype{x}{\sigma} \vdash \intype{\tau[x]}{\mathrm{\bf type}_i}}}
{\ant{\Sigma \vdash \intype{\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}}{\mathrm{\bf type}_i}}}
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \intype{u}{\sigma}}
\ant{\Sigma \vdash \intype{w}{\tau[u]}}}
{\ant{\Sigma \vdash \intype{\mathbf{Pair}(u,w)}{\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}}}
\ant{\Sigma \vdash \pi_1(\mathbf{Pair}(u,w)) \doteq u}
\ant{\Sigma \vdash \pi_2(\mathbf{Pair}(u,w)) \doteq w}}
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{p}{\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}}}}
{\ant{\Sigma \vdash \intype{\pi_1(p)}{\sigma}}
\ant{\Sigma \vdash \intype{\pi_2(p)}{\tau[\pi_1(p)]}}
\ant{\Sigma \vdash p \doteq \mathbf{Pair}(\pi_1(p),\;\pi_2(p))}}
~ \hfill
\unnamed
{\ant{\Sigma \vdash \intype{a}{\tau}}}
{\ant{\Sigma \vdash a \doteq a}}
\hfill
\unnamed
{\ant{\Sigma \vdash u \doteq w}}
{\ant{\Sigma \vdash w \doteq u}}
\hfill \unnamed
{\ant{\Sigma \vdash u \doteq w}
\ant{\Sigma \vdash w \doteq s}}
{\ant{\Sigma \vdash u \doteq s}}
\hfill
\unnamed
{\ant{\Sigma \vdash u\doteq w}
\ant{\Sigma \vdash \Theta[u]}}
{\ant{\Sigma \vdash \Theta[w]}}
\hfill ~
~ \hfill
\unnamed
{\ant{\Sigma \vdash \intype{\tau}{\mathrm{\bf type}_i}}
\ant{\Sigma;\;\intype{x}{\tau} \vdash \intype{\Phi[x]}{\mathrm{\bf Bool}}}}
{\ant{\Sigma \vdash \intype{\subtype{\intype{x}{\tau},\;\Phi[x]}}{\mathrm{\bf type}_i}}}
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{\subtype{\intype{x}{\tau},\;\Phi[x]}}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \intype{e}{\tau}}
\ant{\Sigma \vdash \Phi[e]}}
{\ant{\Sigma \vdash \intype{e}{\subtype{\intype{x}{\tau},\;\Phi[x]}}}}
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{e}{\subtype{\intype{x}{\tau},\;\Phi[x]}}}}
{\ant{\Sigma \vdash \intype{e}{\tau}}
\ant{\Sigma \vdash \Phi[e]}}
\hfill ~
}
\caption{{\bf Dependent Pair Types, Absolute Equality and Subtypes.}
Absolute equality (judgmental equality) accounts for the fact that we can have distinct but isomorphic groups ---
we can have $G =_{\mathbf{Group}} \;H$ with $G \not\doteq H$. It is important that we do {\em not} have $\intype{x}{\mathbf{Group}} \vdash \intype{(G \doteq x)}{\mathrm{\bf Bool}}$
as one could then derive that $G =_{\mathbf{Group}} H$ implies $G \doteq H$
by substitution into $G \doteq x$.}
\label{fig:pairs}
\end{framed}
\end{figure}
\begin{figure}
\begin{framed}
{\footnotesize
~ \hfill
\unnamed
{\ant{\Sigma;\intype{x}{\sigma} \vdash \intype{\Phi[x]}{\mathrm{\bf Bool}}}
\ant{\Sigma \vdash \exists!\intype{x}{\sigma}\;\Phi[x]}
\ant{\Sigma \vdash \intype{\sigma}{\mathbf{Set}}}}
{\ant{\Sigma \vdash \intype{\mathbf{The}(\intype{x}{\sigma},\;\Phi[x])}{\sigma}}
\ant{\Sigma \vdash \Phi[\mathbf{The}(\intype{x}{\sigma},\;\Phi[x])]}}
\hfill
\unnamed
{\ant{\Sigma;\intype{x}{\sigma} \vdash \intype{\Phi[x]}{\mathrm{\bf Bool}}}
\ant{\Sigma \vdash \exists!\intype{x}{\sigma}\;\Phi[x]}
\ant{\mathbf{The}(\intype{x}{\sigma}\;\Phi[x])\;\mbox{is closed}}}
{\ant{\Sigma \vdash \intype{\mathbf{The}(\intype{x}{\sigma},\;\Phi[x])}{\sigma}}
\ant{\Sigma \vdash \Phi[\mathbf{The}(\intype{x}{\sigma},\;\Phi[x])]}}
\hfill ~
~ \hfill
\unnamed
{\ant{\Sigma \vdash \intype{\sigma}{\mathbf{Set}}}
\ant{\Sigma \vdash a =_\sigma b}}
{\ant{\Sigma \vdash a \doteq b}}
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{f}{\sigma \rightarrow \tau}}
\ant{\Sigma \vdash a =_\sigma b}}
{\ant{\Sigma \vdash f(a) \doteq f(b)}}
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{f,g}\sigma \rightarrow \tau}
\ant{\Sigma;\intype{x}{\sigma} \vdash f(x) \doteq g(x)}}
{\ant{\Sigma \vdash f \doteq g}}
\hfill
\unnamed
{\ant{\Sigma; \intype{x}{\sigma} \vdash \intype{x}{\tau}}
\ant{\Sigma; \intype{x}{\tau} \vdash \intype{x}{\sigma}}}
{\ant{\Sigma \vdash \sigma \doteq \tau}}
\hfill ~
}
\caption{{\bf Definite Descriptions and Additional Absolute Equality Rules.} The natural numbers and the reals
can be written as closed definite descriptions involving axioms specifying their structure up to isomorphism.
The second rule of the second row expresses that a function on classes
(or higher level types) returns the same arbitrarily chosen member of the output isomorphism class for all the members of a given input isomorphism class.
Functions differ in this respect from natural maps introduced in section~\ref{sec:natural}.
}
\label{fig:additional}
\end{framed}
\end{figure}
We now replace the word ``sort'' with the word ``type''.
To define the set of well-formed terms and formulas we need to specify primitive types and a set of typed constant and function symbols.
In formal type systems this is done with symbol declarations.
We write $\Sigma \vdash \intype{t}{\sigma}$
to indicate that the symbol declarations in $\Sigma$ imply that $t$ is a well-formed expression
of type $\sigma$. For example we have the following.
\begin{eqnarray*}
\left.\begin{array}{l}\intype{\alpha}{\mathrm{\bf type}}; \\
\intype{\beta}{\mathrm{\bf type}}; \\
\intype{c}{\alpha}; \\
\intype{f}{\alpha \rightarrow \beta} \end{array}\right\} & \vdash & \intype{f(c)}{\beta}
\end{eqnarray*}
\begin{eqnarray*}
\left.\begin{array}{l}\intype{\alpha}{\mathrm{\bf type}}; \\
\intype{c}{\alpha}; \\
\intype{f}{\alpha \rightarrow \alpha}; \\
\intype{P}{\alpha \rightarrow \mathrm{\bf Bool}}\end{array}\right\} & \vdash & \intype{P(f(f(c)))}{\mathrm{\bf Bool}}
\end{eqnarray*}
An expression of the form $\Sigma \vdash \Theta$ is called a {\em sequent} where $\Sigma$
is called the context and $\Theta$ is called the judgement. The sequent $\Sigma \vdash \Theta$ says
that judgement $\Theta$ holds in context $\Sigma$. We allow a context to contain both symbol declarations
and Boolean assumptions. For example we have
$$\left.\begin{array}{l}
\intype{\alpha}{\mathrm{\bf type}};\;\intype{a}{\alpha};\;\intype{b}{\alpha}; \\
\intype{f}{\alpha \times \alpha \rightarrow \alpha}; \\
\forall\intype{x}{\alpha}\;\forall \intype{y}{\alpha} \\
\;\;f(x,y) =_\alpha f(y,x)
\end{array}\right\} \vdash f(a,b) =_\alpha f(b,a)$$
In higher order predicate calculus the type system
is extended to include not only primitive types but also function types
and we can write, for example, $P(f)$ where we have $\intype{f}{\sigma\rightarrow \tau}$
and $\intype{P}{(\sigma \rightarrow \tau) \rightarrow \mathrm{\bf Bool}}$. In the higher order case
we can use the following standard abbreviations due to Curry.
\begin{eqnarray*}
\sigma_1 \times \sigma_2 \rightarrow \tau & \equiv & \sigma_1 \rightarrow (\sigma_2 \rightarrow \tau) \\
f(a,b) & \equiv & f(a)(b)
\end{eqnarray*}
This extends in the obvious way to abbreviations of the form $\sigma_1 \times \cdots \times \sigma_n \rightarrow \tau$.
Without loss of generality we then need consider only single argument functions.
Figure~\ref{fig:expressions} gives a set of inference rules for forming the expressions of higher order predicate calculus.
Each rule in figure~\ref{fig:expressions} allows for the derivation of the sequent below the line provided that the sequents above the line are derivable.
A rule with no antecedents is written as a single derivable sequent.
The rules introduce the constant symbols $\mathrm{\bf type}_0$, $\mathrm{\bf type}_1$, $\mathrm{\bf type}_2$ $\ldots$ where we have $\intype{\mathrm{\bf type}_j}{\mathrm{\bf type}_i}$ for $j < i$.
The subscripts and the restriction that $j < i$ are needed to avoid Russell's paradox. We will use $\mathbf{Set}$ as an alternate notation for
$\mathrm{\bf type}_0$ and $\mathbf{Class}$ as an alternate notation for $\mathrm{\bf type}_1$. Note that we have $\epsilon \vdash \intype{\mathbf{Set}}{\mathbf{Class}}$.
Figure~\ref{fig:expressions} does not include rules introducing lambda expressions. In MorTT functions are introduced with the axiom of choice
given in the last row of figure~\ref{fig:rules}. Functions can also be written with the definite descriptions presented in figure~\ref{fig:additional}
where we have the following abbreviation.
$$(\lambda \intype{x}{\sigma}\; \intype{e[x]}{\tau}) \;\;\equiv\;\; \mathbf{The}(\intype{f}{\sigma \rightarrow \tau},\;\forall \intype{x}{\sigma}\; f(x) =_\tau e[x])$$
Figure~\ref{fig:rules} gives inference rules for predicate calculus and rules expressing the axioms
of extensionality, choice and infinity.
Figure~\ref{fig:pairs} gives inference rules for dependent pair types and subtypes.
A dependent pair type has the form $\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}$ and is the type whose instances
are the pairs $\mathbf{Pair}(x,y)$ where $x$ is an instance of $\sigma$ and $y$ is an instance of $\tau[x]$.
A subtype expression has the form $\mathbf{SubType}(\intype{x}{\sigma},\;\Phi[x])$ where $\Phi[x]$ is a Boolean expression.
This expression denotes the type whose elements are those elements $x$ in $\sigma$ such that $\Phi[x]$ holds.
We let $\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]}\; \mbox{s.t.}\; \Phi[x,y])$
abbreviate $\subtype{\intype{z}{\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}},\;\Phi[\pi_1(z),\pi_2(z)]}$.
The type of groups, abbreviated $\mathbf{Group}$, can then be written as
$$\mathbf{Group} \equiv \pairtype{\alpha}{\mathbf{Set}}{\intype{f}{(\alpha \times \alpha) \rightarrow \alpha}\;\mbox{s.t.}\;\Phi[\alpha,f]}$$
where $\Phi[\alpha,f]$ states the group axioms. For example,
the group axiom that an identity element exists can be written as
$$\exists \intype{x}{\alpha} \;\forall \intype{y}{\alpha} \;\;\;f(x,y) =_\alpha y\;\;\; \wedge \;\;\; f(y,x) =_\alpha y.$$
The type of topological spaces, denoted $\mathbf{TOP}$, can be written as
$$\mathbf{TOP} \equiv \mathbf{PairOf}\left(\begin{array}{l}
\intype{\alpha}{\mathbf{Set}}, \\
\intype{\mathbf{Open}}{(\alpha \rightarrow \mathrm{\bf Bool}) \rightarrow \mathrm{\bf Bool}}, \\
\;\mbox{s. t.}\;\Psi[\alpha,\mathbf{Open}]
\end{array}\right)$$
where $\Psi[\alpha,\mathbf{Open}]$ states the topology axioms.
Here the open sets of the topological space are represented by predicates.
Note that the types $\mathbf{Group}$ and $\mathbf{TOP}$ are closed type expressions --- these type expressions do not contain free variables.
We should note that subtypes are literally subsets and, for example, we can derive the sequent
$\intype{G}{\mathbf{AbelianGroup}} \vdash \intype{G}{\mathbf{Group}}.$
Figure~\ref{fig:additional} gives rules for definite descriptions. Care must be taken to ensure that these rules are sound under morphoid semantics.
Soundness proofs are given in section~\ref{sec:soundness}. The second definite description rule can be used to define
the natural numbers by
$${\cal N} \equiv \mathbf{The}(\intype{p}{\mathbf{PairOf}(\intype{\alpha}{\mathbf{Set}},\;\intype{s}{\alpha \rightarrow \alpha}),\;\Phi[p]})$$
where $\Phi[p]$ states Peano's axioms (the second order version).
The ordered field of real numbers can be defined similarly.
Figure~\ref{fig:additional} also gives some additional rules for absolute equality. These rules provide some insight
into morphoid semantics as defined in sections~\ref{subsec:semantics} and~\ref{sec:Morphoids}.
\subsection{Isomorphism Rules}
\label{sec:isorules}
The core rules of MorTT in figures~\ref{fig:expressions} through \ref{fig:additional} intuitively specify a notion of isomorphism
by specifying the observable properties of objects. We want a notion of
isomorphism that is as coarse as possible while not equating objects distinguishable by Boolean propositions of the core rules. Note that equating
$a$ and $b$ when $\Phi[a]$ can be proved true and $\Phi[b]$ can be proved false leads to an inconsistency. So we can alternatively say that
the rules for deriving isomorphism should equates as many things as possible while still being consistent with the core rules.
The rules for deriving isomorphism relations are given in figure~\ref{fig:simple}. The isomorphism rules
make use of the judgmental rules defined in figure~\ref{fig:jrules}. Judgmental rules manipulate judgements that are not propositions.
For example, as we have noted ealier, absolute equalities $a \doteq b$
are not propositional --- we cannot derive $\intype{(a \doteq b)}{\mathrm{\bf Bool}}$.
The rules in figure~\ref{fig:jrules} distinguish three kinds of expressions that can appear to the right of $\vdash$.
\begin{quotation}
\noindent {\bf Judgments.} A judgement is any expression that can appear on the right hand side of $\vdash$. Semantically we can think of
judgements as predicates on sets of variable interpretations. The sequent $\Sigma \vdash \Theta$ is valid if the property required by $\Theta$
holds of the set of semantic interpretations of $\Sigma$.
\medskip
\noindent {\bf Formulas.} A formula is a judgement $\Theta$ where the validity of $\Sigma \vdash \Theta$
has the form ``$\Theta$ is true in every interpretation of $\Sigma$''. The validity of $\Sigma \vdash \intype{e}{\tau}$ is not of this form
--- see requirements (V1) and (V2) in figure~\ref{fig:models}. We have that $\intype{e}{\tau}$ is a judgement but is not a formula.
A sequent of the form $\Sigma \vdash \inntype{e}{\tau}$ is valid
if in every semantic interpretation of $\Sigma$ we have that the semantic value of $e$ is a member of the value of $\tau$.
In contrast to $\intype{e}{\tau}$, the expression $\inntype{e}{\tau}$ is a formula. We have $\inntype{(\inntype{e}{\tau})}{\mathrm{\bf Bool}}$ but not
$\inntype{(\intype{e}{\tau})}{\mathrm{\bf Bool}}$.
\medskip
\noindent {\bf Propositions.} A formula $\Phi$ is said to be a proposition in context $\Sigma$ if we have
we have $\Sigma \vdash \intype{\Phi}{\mathrm{\bf Bool}}$. An absolute equation $x \doteq y$
is a formula but not a proposition --- it does not satisfy conditions (V1) and (V2) in figure~\ref{fig:models}.
\end{quotation}
\begin{figure}[t]
\begin{framed}
{\footnotesize
\unnamed
{\ant{\Sigma \vdash \intype{e}{\sigma}}}
{\ant{\Sigma \vdash \inntype{e}{\sigma}}}
\hfill
\unnamed
{\ant{\Sigma \vdash \inntype{\sigma}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \inntype{a}{\tau}}
\ant{\Sigma \vdash \inntype{b}{\gamma}}}
{\ant{\Sigma \vdash \inntype{(a \doteq b)}{\bool}}
\ant{\Sigma \vdash \inntype{(\inntype{a}{\sigma})}{\mathrm{\bf Bool}}}}
\hfill
\unnamed{\ant{\Sigma \vdash \isjudge{\Phi}}
\ant{\Sigma \vdash \isjudge{\Psi}}}
{\ant{\Sigma \vdash \isjudge{(\Phi \vee \Psi)}}
\ant{\Sigma \vdash \inntype{\neg \Phi}{\mathrm{\bf Bool}}}}
\unnamed
{\ant{\Sigma;\intype{x}{\sigma} \vdash \inntype{\Phi[x]}{\mathrm{\bf Bool}}}}
{\ant{\Sigma \vdash \inntype{(\forall \intype{x}{\sigma}\;\Phi[x])}{\mathrm{\bf Bool}}}}
\vspace{-2em}
~ \hfill
\unnamed{
\ant{\Sigma \vdash \inntype{\sigma}{\mathrm{\bf type}_i}}
\ant{\mbox{$x$ is not declared in $\Sigma$}}}
{\ant{\Sigma;\intype{x}{\sigma}\vdash \mathrm{\bf True}}}
\hfill
\unnamed{
\ant{\Sigma \vdash \isjudge{\Theta}}}
{\ant{\Sigma ;\Theta \vdash \mathrm{\bf True}}}
\hfill ~
\vspace{-2em}
~ \hfill
\unnamed
{\ant{\Sigma;\Phi \vdash \Psi}
\ant{\Sigma;\neg \Phi \vdash \Psi}}
{\ant{\Sigma \vdash \Psi}}
\hfill
\unnamed
{\ant{\Sigma \vdash \Phi}
\ant{\Sigma \vdash \inntype{\Psi}{\mathrm{\bf Bool}}}}
{\ant{\Sigma \vdash \Phi \vee \Psi}
\ant{\Sigma \vdash \Psi \vee \Phi}
\ant{\Sigma \vdash \neg \neg \Phi}}
\hfill
\unnamed
{\ant{\Sigma \vdash \neg \Psi}
\ant{\Sigma \vdash \neg \Phi}}
{\ant{\Sigma \vdash \neg(\Phi \vee \Psi)}}
\hfill
\unnamed
{\ant{\Sigma \vdash \forall \intype{x}{\tau} \;\Phi[x]}
\ant{\Sigma \vdash \intype{e}{\tau}}}
{\ant{\Sigma \vdash \Phi[e]}}
\hfill
\unnamed
{\ant{\Sigma;\intype{x}{\sigma} \vdash \inntype{\Phi[x]}{\mathrm{\bf Bool}}}
\ant{\Sigma; \intype{x}{\sigma} \vdash \Phi[x]}}
{\ant{\Sigma \vdash \forall \intype{x}{\sigma}\;\Phi[x]}}
\hfill ~
}
\caption{{\bf Judgmental Inference.} We write $\inntype{e}{\tau}$ to indicate $e \in \tau$ without requiring
that $e$ satisfy conditions (V1) and (V2) of figure~\ref{fig:models}.
These judgmental rules support the isomorphism rules in figure~\ref{fig:simple}.}
\label{fig:jrules}
\end{framed}
\end{figure}
\begin{figure}[p]
\begin{framed}
{\footnotesize
$$\mathbf{Bijection}[\sigma,\tau] \equiv \mathbf{SubType}(\intype{f}{\sigma \rightarrow \tau},\;\;\forall \intype{y}{\tau} \;\exists! \intype{x}{\sigma}\;f(x)=_\tau y)$$
$$\mathbf{PairOf}(\sigma,\tau) \;\;\equiv \;\; \mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau}) \;\mbox{with $x$ not free in $\tau$} $$
$$a \leftrightarrowtriangle_\sigma b\;\;\equiv \;\;\exists \intype{x}{\mathrm{\bf iso}(\sigma,a,b)}$$
\vspace{-1em}
\hfill ~
\unnamed{
\ant{\Sigma \vdash \intype{\sigma,\tau}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \intype{f}{\mathbf{Bijection}[\sigma,\tau]}}}
{\ant{\Sigma \vdash \inntype{\updownarrow\!(\sigma,\tau,f)}{\mathrm{\bf iso}(\mathrm{\bf type}_i,\sigma,\tau)}}
\ant{}
\ant{\Sigma \vdash \left\{\begin{array}{l}\forall \intype{x}{\sigma}\;\forall\intype{y}{\tau} \\
\;\;(x \leftrightarrowtriangle_{\updownarrow\!(\sigma,\tau,f)} y) \\
\;\;\Leftrightarrow f(x) =_{\tau} y
\end{array}\right.}}
\hfill
\unnamed{
\ant{\Sigma \vdash \inntype{a_3}{\mathrm{\bf iso}(\sigma,a_1,a_2)}}
\ant{\Sigma \vdash \inntype{b_3}{\mathrm{\bf iso}(\tau,b_1,b_2)}}}
{\ant{\Sigma \vdash \inntype{\mathbf{Pair}(a_3,b_3)}{\mathrm{\bf iso}\left(\begin{array}{l}\mathbf{PairOf}(\sigma,\tau), \\\mathbf{Pair}(a_1,b_1), \\\mathbf{Pair}(a_2,b_2) \end{array}\right)}}}
\hfill ~
~ \hfill
\unnamed
{\ant{\Sigma \vdash \intype{a,b}{\sigma}}}
{\ant{\Sigma \vdash \left\{\begin{array}{l} a \leftrightarrowtriangle_\sigma b \\ \;\Leftrightarrow \\ a =_\sigma b \end{array}\right.}}
\hfill
\unnamed{
\ant{\Sigma \vdash \inntype{a_3}{\mathrm{\bf iso}(\sigma,a_1,a_2)}}}
{\ant{\Sigma \vdash \inntype{a_3}{\sigma}}}
\hfill ~
\vspace{-1em}
~ \hfill
\unnamed{
\ant{\Sigma \vdash \intype{\mathbf{Pair}(a_1,b_1),\mathbf{Pair}(a_2,b_2)}{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}}
\ant{\Sigma \vdash \inntype{a_3}{\mathrm{\bf iso}(\sigma,a_1,a_2)}}
\ant{\Sigma \vdash b_1 \leftrightarrowtriangle_{\tau[a_3]} b_2}
}{
\ant{\Sigma \vdash \mathbf{Pair}(a_1,b_1) =_{\mathbf{PairOf}(\intype{x\;}{\;\sigma},\;\intype{y\;}{\;\tau[x]})} \mathbf{Pair}(a_2,b_2)}
}
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{a_1}{\sigma},\;\intype{a_2}{\sigma},\;\inntype{a_3}{\mathrm{\bf iso}(\sigma,a_1,a_2)}}
\ant{\Sigma;\;\intype{x}{\sigma} \vdash \intype{\mathbf{PairOf}(\tau_1[x],\;\tau_2[x])}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \intype{b_1}{\mathbf{PairOf}(\tau_1[a_1],\;\tau_2[a_1])}}
\ant{\Sigma \vdash \intype{b_2}{\mathbf{PairOf}(\tau_1[a_2],\;\tau_2[a_2])}}}
{\ant{\Sigma \vdash \left\{\begin{array}{l} (b_1 \leftrightarrowtriangle_{\mathbf{PairOf}(\tau_1[a_3],\;\tau_2[a_3])} b_2) \\
\;\;\;\Leftrightarrow \\
\pi_1(b_1) \leftrightarrowtriangle_{\tau_1[a_3]} \pi_1(b_2) \;\wedge \\
\pi_2(b_1) \leftrightarrowtriangle_{\tau_2[a_3]} \pi_2(b_2)
\end{array}\right.}}
\hfill ~
\vspace{-1em}
~ \hfill
\unnamed{
\ant{\Sigma \vdash \intype{a_1}{\sigma},\;\intype{a_2}{\sigma},\;\inntype{a_3}{\mathrm{\bf iso}(\sigma,a_1,a_2)}}
\ant{\Sigma; \intype{x}{\sigma} \vdash \intype{(\tau_1[x] \rightarrow \tau_2[x])}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \intype{b_1}{(\tau_1[a_1] \rightarrow \tau_2[a_1])}}
\ant{\Sigma \vdash \intype{b_2}{(\tau_1[a_2] \rightarrow \tau_2[a_2])}}}
{\ant{\Sigma \vdash \left\{\begin{array}{l} (b_1 \leftrightarrowtriangle_{\tau_1[a_3] \rightarrow \tau_2[a_3]} b_2) \\
\;\;\;\Leftrightarrow \\
\forall \intype{x_1}{\tau_1[a_1]}\; \;\forall \intype{x_2}{\tau_1[a_2]} \\
\;\;(x_1 \leftrightarrowtriangle_{\tau_1[a_3]} x_2) \\
\;\;\;\;\;\;\;\Rightarrow \\
\;\;b_1(x_1) \leftrightarrowtriangle_{\tau_2[a_3]} b_2(x_2)
\end{array}\right.}}
\hfill
\unnamed{
\ant{\Sigma \vdash \intype{a_1}{\sigma},\;\intype{a_2}{\sigma},\;\inntype{a_3}{\mathrm{\bf iso}(\sigma,a_1,a_2)}}
\ant{\Sigma; \intype{x}{\sigma} \vdash \intype{\mathbf{SubType}(\intype{y}{\tau[x]},\;\Phi[x,y])}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \intype{b_1}{\mathbf{SubType}(\intype{y}{\tau[a_1]},\;\Phi[a_1,y])}}
\ant{\Sigma \vdash \intype{b_2}{\mathbf{SubType}(\intype{y}{\tau[a_2]},\;\Phi[a_2,y])}}}
{\ant{\Sigma \vdash \left\{\begin{array}{l} (b_1 \leftrightarrowtriangle_{\mathbf{SubType}(\intype{y\;}{\;\tau[a_3]},\;\Phi[a_3,y])} b_2) \\ \;\;\;\Leftrightarrow \\ (b_1 \leftrightarrowtriangle_{\tau[a_3]} b_2)\end{array}\right.}}
\hfill ~
}
\caption{{\bf Isomorphism Rules.}
We have that $\mathrm{\bf iso}(\sigma,x,y)$ is the type whose members are the isomorphisms from $x$ to $y$.
The first two rules introduce isomorphisms between lists of types.
The first rule of third row derives equality (isomorphism) at dependent pair types
$\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})$ where $\tau[x]$ is simple in $x$.
It states that the pair $\mathbf{Pair}(a_1,b_1)$ is isomorphic to the pair $\mathbf{Pair}(a_2,b_2)$
if there exists an isomorphism from $a_1$ to $a_2$ that carries $b_1$ to $b_2$. The notation $b_1 \leftrightarrowtriangle_{\tau[a_3]} b_2$ indicates
that the isomorphism $a_3$ carries $b_1$ to $b_2$. The carrying relaiton $\leftrightarrowtriangle_{\tau[a_3]}$ is defined in the conclusions of the last three rules
plus bases cases defined by the conclusions of the first rule and the first rule of the second row.
The carrying relation $\leftrightarrowtriangle_{\tau[a_3]}$ always defines a bijection between $\tau[a_1]$ and $\tau[a_2]$.}
\label{fig:simple}
\end{framed}
\end{figure}
Figure~\ref{fig:simple} gives rules for deriving isomorphism for a subclass of the dependent pair types.
Isomorphism at other dependent pair types are derives from cryptomorphic equivalences
between the more general types and the types handled in figure~\ref{fig:simple}.
To understand the rules in figure~\ref{fig:simple} it is best to start with simple structure types.
To define simple structure types we first introduce the notation
$\mathbf{PairOf}(\sigma,\tau)$ as an abbreviation for $\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau})$ where $x$ does not occur free in $\tau$.
This is a simple pair type as opposed to a true dependent pair type.
We define a simple structure type to be a type expression of the form $\mathbf{PairOf}(\intype{\alpha}{\mathbf{Set}},\;\intype{y}{\delta[\alpha]})$ where
$\delta[\alpha]$ is simple in $\alpha$ as defined by the following grammar.
\begin{eqnarray*}
\delta[\alpha] & ::= & \alpha \;|\;\; \mbox{a type $\sigma$ with $\alpha$ not free in $\sigma$} \;\;|\; \delta_1[\alpha] \rightarrow \delta_2[\alpha] \\
& & \;|\; \mathbf{PairOf}(\delta_1[\alpha],\;\delta_2[\alpha]) \;|\; \mathbf{SubType}(\intype{x}{\delta[\alpha]},\;\Phi[\alpha,x])
\end{eqnarray*}
We note that simple structure types include most of the familiar concepts of mathematics such as groups, topological spaces and vector spaces. For example,
the type group can be written in the form
$$\mathbf{PairOf}(\intype{\alpha}{\mathbf{Set}},\;\mathbf{SubType}(\intype{f}{\alpha \times \alpha \rightarrow \alpha},\;\Phi[\alpha,f])).$$
Here we have placed the group axioms inside the dependent pair type. This can also be done for topological spaces and other
familiar concepts. The type of vector spaces over a given field $F$ can be written as
$$\mathbf{PairOf}(\intype{\alpha}{\mathbf{Set}},
\;\mathbf{SubType}(\mathbf{PairOf}(\intype{+}{\alpha \times \alpha \rightarrow \alpha},\;\intype{\ast}{\pi_1(F) \times \alpha \rightarrow \alpha}),
\;\Phi[F,\alpha,+,\ast])).$$
This is a simple structure type in which the field $F$ occurs as a free variable of the type expression.
The rules in figure~\ref{fig:simple} handle simple structure types as a special case. Consider the first rule of the third row which we will call the equality generation rule.
In this rule we can take $\sigma$ to be $\mathbf{Set}$. In that case $a_1$, $a_2$ and $a_3$ are sets where $a_3$ represents a bijection from $a_1$ to $a_2$
as derived by the first rule of the first row. Such an instance of the equality generation rule states that for a type of the form
$\mathbf{PairOf}(\intype{\alpha}{\mathbf{Set}},\;\intype{y}{\tau[\alpha]})$
we have that $\mathbf{Pair}(\sigma,b_1)$ is isomorphic to $\mathbf{Pair}(\delta,b_2)$ if there exists a bijection $f$ from $\sigma$ to $\delta$ that carries
$b_1$ to $b_2$. The statement that bijecition $f$ carries $b_1$ to $b_2$ is written as $b_1 \leftrightarrowtriangle_{\tau[\updownarrow(\sigma,\delta,f)]} b_2$.
The semantics is discussed in section~\ref{subsec:semantics}.
The rules in figure~\ref{fig:simple} recursively define the carrying relation $\leftrightarrowtriangle_{\tau[\updownarrow\!(\sigma,\delta,f)]}$ for all type expressions $\tau[\alpha]$ that are simple
in $\alpha$. In all cases $\leftrightarrowtriangle_{\tau[\updownarrow\!(\sigma,\delta,f)]}$ is a bijective relation between $\tau[\sigma]$ and $\tau[\delta]$.
The rules in figure~\ref{fig:simple} handle a slightly more general case. In the equality generation rule
the type $\sigma$ can be taken to be a pair of types, or a pair of pairs of types, or a list of types, or in general any tree over types.
For example we have
$$\Sigma; \intype{x}{\mathbf{PairOf}(\mathbf{Set},\mathbf{Set})} \vdash \intype{(\pi_1(x) \rightarrow \pi_2(x))}{\mathbf{Set}}.$$
As an example consider the type of colored graphs defined to be a pair of a graph and a coloring of that graph.
$$\mathbf{Graph} \equiv \mathbf{PairOf}(\intype{\alpha}{\mathbf{Set}},\;\intype{\mathbf{Edge}}{(\alpha \times \alpha) \rightarrow \mathrm{\bf Bool}})$$
$$\mathbf{CGraph} \equiv \mathbf{PairOf}(\intype{G}{\mathbf{Graph}},\;\mathbf{PairOf}(\intype{\beta}{\mathbf{Set}},\;\intype{c}{(\pi_1(G) \rightarrow \beta)}))$$
We can alternatively (cryptomorphically) define a colored graph as follows.
$$\mathbf{CGraph'} \equiv \mathbf{PairOf}\left(\intype{P}{\mathbf{PairOf}(\mathbf{Set},\mathbf{Set}),
\;\intype{y}{\mathbf{PairOf}\left(\begin{array}{l}\intype{\mathbf{Edge}}{(\pi_1(P) \times \pi_1(P)) \rightarrow \mathrm{\bf Bool}}, \\ \;\intype{c}{(\pi_1(P) \rightarrow \pi_2(P))}\end{array}\right)}}\right)$$
We can use the rules in figure~\ref{fig:simple} to derive equations at the type $\mathbf{CGraph'}$.
But we also have
$$\intype{G}{\mathbf{CGraph'}} \vdash \intype{\mathbf{Pair}(\mathbf{Pair}(\pi_1(\pi_1(G)),\pi_1(\pi_2(G))),\;\mathbf{Pair}(\pi_2(\pi_1(G)),\pi_2(\pi_2(G))))}{\mathbf{CGraph}}$$
By applying the substitution rule to the above sequent and an equation of the form $G' =_{\mathbf{CGraph'}} H'$
we can generate the desired isomorphisms at the type $\mathbf{CGraph}$.
\subsection{Semantics}
\label{subsec:semantics}
\begin{figure}[t]
\begin{framed}
\begin{itemize}
\item[(G1)] For any morphoid $x$ we have that $\mathbf{Left}(x)$, $\mathbf{Right}(x)$ and $x^{-1}$ are also morphoids.
\item[(G2)] $x \circ y$ is defined if and only if $\mathbf{Right}(x) = \mathbf{Left}(y)$
and when $x \circ y$ is defined we have that $x \circ y$ is a morphoid.
\item[(G3)] $\mathbf{Left}(x^{-1}) = \mathbf{Right}(x)$ and $\mathbf{Right}(x^{-1}) = \mathbf{Left}(x)$
\item[(G4)] $\mathbf{Left}(x \circ y) = \mathbf{Left}(x)$ and $\mathbf{Right}(x \circ y) = \mathbf{Right}(y)$.
\item[(G5)] $(x \circ y) \circ z$ = $x \circ (y \circ z)$.
\item[(G6)] $x^{-1} \circ x \circ y = y$ and $x \circ y \circ y^{-1} = x$.
\item[(G7)] $\mathbf{Right}(x) = x^{-1} \circ x$ and $\mathbf{Left}(x) = x \circ x^{-1}$
\item[(G8)] $(x^{-1})^{-1} = x$.
\item[(G9)] $(x \circ y)^{-1} = y^{-1} \circ x^{-1}$.
\end{itemize}
\caption{{\bf The Groupoid Properties of Morphoids.} In MorTT all semantic values are morphoids.}
\label{fig:GroupProps}
\end{framed}
\end{figure}
\begin{figure}[p]
\begin{framed}
{\footnotesize
~ \hfill
$$\begin{array}{ccc} \sigma & \mathbf{Left}(\sigma) & \mathbf{Right}(\sigma) \\
\\
\left\{\begin{array}{c}
\mathbf{Point}(a,A),\;\mathbf{Point}(a,\tilde{A}), \\
\mathbf{Point}(\tilde{a},A),\;\mathbf{Point}(\tilde{a},\tilde{A}), \\
\\
\mathbf{Point}(b,B),\;\mathbf{Point}(b,\tilde{B}), \\
\mathbf{Point}(\tilde{b},B),\;\mathbf{Point}(\tilde{b},\tilde{B}), \\
\\
\mathbf{Point}(c,C),\;\mathbf{Point}(c,\tilde{C}), \\
\mathbf{Point}(\tilde{c},C),\;\mathbf{Point}(\tilde{c},\tilde{C}) \\
\end{array}\right\}
&
\left\{\begin{array}{c}
\mathbf{Point}(a,a),\;\mathbf{Point}(a,\tilde{a}), \\
\mathbf{Point}(\tilde{a},a),\;\mathbf{Point}(\tilde{a},\tilde{a}), \\
\\
\mathbf{Point}(b,b),\;\mathbf{Point}(b,\tilde{b}), \\
\mathbf{Point}(\tilde{b},b),\;\mathbf{Point}(\tilde{b},\tilde{b}), \\
\\
\mathbf{Point}(c,c),\;\mathbf{Point}(c,\tilde{c}), \\
\mathbf{Point}(\tilde{c},c),\;\mathbf{Point}(\tilde{c},\tilde{c}) \\
\end{array}\right\}
&
\left\{\begin{array}{c}
\mathbf{Point}(A,A),\;\mathbf{Point}(A,\tilde{A}), \\
\mathbf{Point}(\tilde{A},A),\;\mathbf{Point}(\tilde{A},\tilde{A}), \\
\\
\mathbf{Point}(B,B),\;\mathbf{Point}(B,\tilde{B}), \\
\mathbf{Point}(\tilde{B},B),\;\mathbf{Point}(\tilde{B},\tilde{B}), \\
\\
\mathbf{Point}(C,C),\;\mathbf{Point}(C,\tilde{C}), \\
\mathbf{Point}(\tilde{C},C),\;\mathbf{Point}(\tilde{C},\tilde{C}) \\
\end{array}\right\}
\end{array}$$
\bigskip
\medskip
$$\begin{array}{ccc} \sigma^{-1} & \tau & \sigma \circ \tau \\
\\
\left\{\begin{array}{c}
\mathbf{Point}(A,a),\;\mathbf{Point}(\tilde{A},a), \\
\mathbf{Point}(\tilde{A},a),\;\mathbf{Point}(\tilde{A},\tilde{a}), \\
\\
\mathbf{Point}(B,b),\;\mathbf{Point}(\tilde{B},b), \\
\mathbf{Point}(\tilde{B},b),\;\mathbf{Point}(\tilde{B},\tilde{b}), \\
\\
\mathbf{Point}(C,c),\;\mathbf{Point}(\tilde{C},c), \\
\mathbf{Point}(\tilde{C},c),\;\mathbf{Point}(\tilde{C},\tilde{c})
\end{array}\right\}
&
\left\{\begin{array}{c}
\mathbf{Point}(A,X),\;\mathbf{Point}(A,\widetilde{X}), \\
\mathbf{Point}(\tilde{A},X),\;\mathbf{Point}(\tilde{A},\widetilde{X}), \\
\\
\mathbf{Point}(B,Y),\;\mathbf{Point}(B,\widetilde{Y}), \\
\mathbf{Point}(\tilde{B},Y),\;\mathbf{Point}(\tilde{B},\widetilde{Y}), \\
\\
\mathbf{Point}(C,Z),\;\mathbf{Point}(C,\widetilde{Z}), \\
\mathbf{Point}(\tilde{C},Z),\;\mathbf{Point}(\tilde{C},\widetilde{Z}) \\
\end{array}\right\}
&
\left\{\begin{array}{c}
\mathbf{Point}(a,X),\;\mathbf{Point}(a,\widetilde{X}), \\
\mathbf{Point}(\tilde{a},X),\;\mathbf{Point}(\tilde{a},\widetilde{X}), \\
\\
\mathbf{Point}(b,Y),\;\mathbf{Point}(b,\widetilde{Y}), \\
\mathbf{Point}(\tilde{b},Y),\;\mathbf{Point}(\tilde{b},\widetilde{Y}), \\
\\
\mathbf{Point}(c,Z),\;\mathbf{Point}(c,\widetilde{Z}), \\
\mathbf{Point}(\tilde{c},Z),\;\mathbf{Point}(\tilde{c},\widetilde{Z}) \\
\end{array}\right\}
\end{array}$$
}
\bigskip
\bigskip
~
{\footnotesize
\hspace{11em} $G$ \hspace{8em} $\mathbf{Left}(G)$ \hspace{6em} $\mathbf{Right}(G)$
\vspace{3ex}
\begin{verbatim}
Point(a,A) Point(a,a) Point(A,A)
\ \ \
Point(b,B) Point(b,b) Point(B,B)
/ / /
Point(c,C) Point(c,c) Point(C,C)
\end{verbatim}
\vspace{8ex}
\hspace{11em} $G^{-1}$ \hspace{8em} $H$ \hspace{8em} $G \circ H$
\vspace{3ex}
\begin{verbatim}
Point(A,a) Point(A,X) Point(a,X)
\ \ \
Point(B,b) Point(B,Y) Point(b,Y)
/ / /
Point(C,c) Point(C,Z) Point(c,Z)
\end{verbatim}
}
\medskip
\caption{{\bf Examples of $\mathbf{Left}$, $\mathbf{Right}$, inverse and composition.} Morphoids are built from points just as sets can be built from ur-elements.
A point has the form $\mathrm{\bf Point}(i,j)$ where we think of $i$ is a left index and $j$ as a right index.
We have $\mathrm{\bf Point}(i,j)^{-1} = \mathrm{\bf Point}(j,i)$ and $\mathrm{\bf Point}(i,j) \circ \mathrm{\bf Point}(j,k) = \mathrm{\bf Point}(i,k)$.
For any point type $\sigma$ we have that $\mathbf{Left}(\sigma)$ and $\mathbf{Right}(\sigma)$ are equivalence relations on indeces and $\sigma$
defines a bijection between the equivalence classes of $\mathbf{Left}(\sigma)$ and $\mathbf{Right}(\sigma)$. The last two rows show morphoid graphs where the nodes are
points.}
\label{fig:Examples}
\end{framed}
\end{figure}
\begin{figure}[t]
\begin{framed}
{\small
{\bf Structure.} A structure is defined to be a mapping from a finite set of variables to morphoids. The groupoid operations are defined
on structures $\rho$ and $\gamma$ by $\mathbf{Left}(\rho)(x) = \mathbf{Left}(\rho(x))$, $\mathbf{Right}(\rho)(x) = \mathbf{Right}(\rho(x))$,
$\rho^{-1}(x) = \rho(x)^{-1}$ and for $\mathbf{Right}(\rho) = \mathbf{Left}(\gamma)$ we define composition by $(\rho \circ \gamma)(x) = \rho(x) \circ \gamma(x)$.
For structures $\rho$ and $\gamma$ we define $\rho \preceq \gamma$ to mean that $\rho$ and $\gamma$ are defined on the same set of variables
and $\rho(x) \preceq \gamma(x)$ for all variables $x$ on which they are defined.
For a structure $\rho$, variable $x$, and morphoid $v$, where $x$ is not assigned a value in $\rho$, we define
$\rho[x \leftarrow v]$ to be the structure identical to $\rho$ but with the added assignment of value $v$ to variable $x$.
\medskip
${\bf \rho \in \convalue{\Sigma}}$. For a context $\Sigma$ the following clauses specify whether $\convalue{\Sigma}$ is defined and, if it is defined,
define it to be a set of structures.
\medskip
\begin{itemize}
\item We define $\convalue{\epsilon}$ to be the set containing the empty structure (the empty structure does not
assign any value to any variables).
\item For $\Sigma \models \inntype{\tau}{\mathrm{\bf type}_i}$ and $x$ not declared in $\Sigma$,
we have that $\convalue{\Sigma;\intype{x}{\tau}}$ is defined to be the set of structures of the form $\rho[x \leftarrow v]$ for $\rho \in \convalue{\Sigma}$
and $v \in \semvalue{\tau}\rho$.
\item For $\Sigma \models \inntype{\Phi}{\mathrm{\bf Bool}}$ we have that $\convalue{\Sigma;\Phi}$ is defined to be the set of all $\rho \in \convalue{\Sigma}$ such that $\semvalue{\Phi}\rho = \mathrm{\bf True}$.
\end{itemize}
\medskip
${\bf \Sigma \models \Phi}$. The relation $\Sigma \models \Phi$ is defined by the following clauses.
\medskip
\begin{itemize}
\item If $\convalue{\Sigma}$ and $\semvalue{\Phi}$ are defined, and for all $\rho \in \convalue{\Sigma}$ we have that $\semvalue{\Phi}\rho$
is a Boolean value, then we define $\Sigma \models \Phi$ to hold if and only if for all $\rho\in \convalue{\Sigma}$ we have that
$\semvalue{\Phi}\rho = \mathrm{\bf True}$.
\item The entailment $\Sigma \models \intype{e}{\tau}$ holds if $\Sigma \models \inntype{e}{\tau}$ and we have the following value (V) properties for $e$.
\begin{itemize}
\item[(V1)] For $\rho_1,\rho_2,\rho_3, \in \convalue{\Sigma}$ with $\rho_1 \circ \rho_2^{-1} \circ \rho_3$ defined and
$(\rho_1 \circ \rho_2^{-1} \circ \rho_3) \in \convalue{\Sigma}$ we have
$$\semvalue{e}(\rho_1 \circ \rho_2^{-1} \circ \rho_3) \;\;=\;\; (\semvalue{e}\rho_1) \circ (\semvalue{e}\rho_2)^{-1} \circ (\semvalue{e}\rho_3).$$
\item[(V2)] For $\rho_1$, $\rho_2 \in \convalue{\Sigma}$ with $\rho_1 \preceq \rho_2$ we have $\semvalue{e}\rho_1 \preceq \semvalue{e}\rho_2$.
\end{itemize}
\end{itemize}
}
\caption{{\bf Structures, the set $\convalue{\Sigma}$, and the relation $\models$.} Figure~\ref{fig:value} defines $\semvalue{(\inntype{e}{\tau})}\rho$
to be true if $\semvalue{e}\rho \in \semvalue{\tau}\rho$. We do not define $\semvalue{(\intype{e}{\tau})}$.
The relation $\preceq$ on morphoids is
defined in figure~\ref{fig:Abstraction}.}
\label{fig:models}
\end{framed}
\end{figure}
\begin{figure}[p]
\begin{framed}
{\small
\noindent $\bullet\;x.$ For $x$ declared in $\Sigma$ and for $\rho \in \convalue{\Sigma}$ we have that $\semvalue{x}$ is defined
with $\semvalue{x}\rho = \rho(x)$.
\medskip
\noindent $\bullet\;\mathrm{\bf Bool}.$ We have that $\semvalue{\mathrm{\bf Bool}}\rho$ is the type containing the two Boolean values $\mathrm{\bf True}$ and $\mathrm{\bf False}$.
\medskip
\noindent $\bullet\;\mathbf{Set}.$ We have $\semvalue{\mathbf{Set}}\rho$ is the type whose members are all discrete morphoid types in the Grothendiek
universe $V_{\kappa_0}$
where $\kappa_0$ is the smallest uncountable inaccessible cardinal.
\medskip
\noindent $\bullet\;\mathrm{\bf type}_i,\;i > 0.$ For $i > 0$ have $\semvalue{\mathrm{\bf type}_i}\rho$ is the type whose members are all morphoid types in the Grothendiek universe $V_{\kappa_i}$
where $\kappa_{i+1}$ is the smallest inaccessible cardinal larger than $\kappa_i$.
\medskip
\noindent $\bullet\;f(e).$ If $\semvalue{f}$ and $\semvalue{e}$ are defined, and for all $\rho \in \convalue{\Sigma}$
we have that $(\semvalue{f}\rho)(\semvalue{e}\rho)$ is defined, then $\semvalue{f(e)}$ is defined
with $\semvalue{f(e)}\rho = (\semvalue{f}\rho)(\semvalue{e}\rho).$
\medskip
\noindent $\bullet\;\sigma \rightarrow \tau.$ If $\Sigma \models \intype{\sigma}{\mathrm{\bf type}_i}$,
and $\Sigma\models \intype{\tau}{\mathrm{\bf type}_i}$, then $\semvalue{\sigma \rightarrow \tau}$ is defined
with $\semvalue{\sigma \rightarrow \tau}\rho = (\semvalue{\sigma}\rho) \rightarrow (\semvalue{\tau}\rho).$
\medskip
\noindent $\bullet\;\forall\; \intype{x}{\tau}\;\Phi[x].$ If
$\Sigma;\;\intype{y}{\tau} \models \inntype{\Phi[y]}{\mathrm{\bf Bool}}$ then $\semvalue{\forall\; \intype{x}{\tau}\;\Phi[x]}$ is defined
with $\semvalue{\forall\; \intype{x}{\tau}\;\Phi[x]}\rho$ being $\mathrm{\bf True}$ if
for all $v \in \semvalue{\tau}\rho$
we have $\subvalue{\Sigma;\intype{y\;}{\;\tau}}{\Phi[y]}\rho[y \leftarrow v] = \mathrm{\bf True}$.
\medskip
\noindent $\bullet\;\Phi \vee \Psi.$
If $\Sigma \models \inntype{\Phi}{\mathrm{\bf Bool}}$ and $\Sigma \models \inntype{\Psi}{\mathrm{\bf Bool}}$ then $\semvalue{\Phi \vee \Psi}$
is defined with $\semvalue{\Phi \vee \Psi}{\rho} = \semvalue{\Phi}{\rho} \vee \semvalue{\rho}{\rho}.$
\medskip
\noindent $\bullet\;\neg \Phi.$
If $\Sigma \models \inntype{\Phi}{\mathrm{\bf Bool}}$ then $\semvalue{\neg\Phi}$ is defined
with
$\semvalue{\neg \Phi}{\rho} = \neg\semvalue{\Phi}{\rho}.$
\medskip
\noindent $\bullet\;s =_\sigma w.$ If $\Sigma \models \inntype{s}{\sigma}$ and $\Sigma \models \inntype{w}{\sigma}$ then
$\semvalue{s=_\sigma w}$ is defined with $\semvalue{s =_\sigma w}\rho$ being $\mathrm{\bf True}$ if
$\semvalue{s}\rho \; =_{\semvalue{\sigma}\rho}\;\;\semvalue{w}\rho$.
\medskip
\noindent $\bullet\;s \doteq w.$ If $\semvalue{s}$ and $\semvalue{w}$ are defined then
$\semvalue{s \doteq w}$ is defined with $\semvalue{s \doteq w}\rho$ being $\mathrm{\bf True}$ if
$\semvalue{s}\rho \;\; =\;\;\semvalue{w}\rho$.
\medskip
\noindent $\bullet\;\inntype{e}{\sigma}.$ If $\semvalue{e}$ and $\semvalue{\sigma}$ are defined and for all $\rho \in \convalue{\Sigma}$
we have that $\semvalue{\sigma}\rho$ is a morphoid type, then
$\semvalue{\inntype{e}{\sigma}}$ is defined with $\semvalue{\inntype{e}{\sigma}}\rho$ being $\mathrm{\bf True}$ if
$\semvalue{e}\rho \in \semvalue{\sigma}\rho$.
\medskip
\noindent $\bullet\;\;\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}.$ If $\Sigma \models \intype{\sigma}{\mathrm{\bf type}_i}$
and $\Sigma;\intype{z}{\sigma} \models \intype{\tau[z]}{\mathrm{\bf type}_i}$ then $\semvalue{\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}}$ is defined
with $\semvalue{\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}}\rho$ being the type containing
the pairs $\mathbf{Pair}(v,w)$ for $v \in \semvalue{\sigma}\rho$ and $w \in \subvalue{\Sigma;\;\intype{z\;}{\;\sigma}}{\tau[z]}\rho[z \leftarrow v]$.
\medskip
\noindent $\bullet\;\mathbf{Pair}(u,w).$ If $\semvalue{u}$ and $\semvalue{w}$ are defined then $\semvalue{\mathbf{Pair}(u,w)}$ is defined with
$\semvalue{\mathbf{Pair}(u,w)}\rho = \mathbf{Pair}(\semvalue{u}\rho,\;\semvalue{w}\rho).$
\medskip
\noindent $\bullet\;\pi_i(e).$ If $\semvalue{e}$ is defined and for all $\rho \in \convalue{\Sigma}$ we have that
$\semvalue{e}{\rho}$ is a pair then $\semvalue{\pi_i(e)}$ is defined with $\semvalue{\pi_i(e)}\rho = \pi_i(\semvalue{e}\rho).$
\medskip
\noindent $\bullet\;\subtype{\intype{x}{\sigma},\;\Phi[x]}.$ If $\Sigma \models \intype{\sigma}{\mathrm{\bf type}_i}$ and
$\Sigma;\;\intype{y}{\sigma} \models \intype{\Phi[y]}{\mathrm{\bf Bool}}$ then $\semvalue{\mathbf{SubType}(\intype{x}{\sigma},\;\Phi[x])}$ is defined
with $\semvalue{\subtype{\intype{x}{\sigma},\;\Phi[x]}}\rho$ being the type whose members are those values $v \in \semvalue{\sigma}\rho$
with $\subvalue{\Sigma;\;\intype{y\;}{\;\sigma}}{\Phi[y]}\rho[y \leftarrow v] = \mathrm{\bf True}$.
\medskip
\noindent $\bullet\;\mathrm{\bf iso}(\sigma,a,b).$ If $\Sigma \models \inntype{\sigma}{\mathrm{\bf type}_i}$, and
$\semvalue{a}$ and $\semvalue{b}$ are defined, and for all $\rho \in \convalue{\Sigma}$
we have that $\mathrm{\bf iso}(\semvalue{\sigma}\rho,\;\semvalue{a}\rho,\;\semvalue{b}\rho)$ is defined,
then $\semvalue{\mathrm{\bf iso}(\sigma,a,b)}$ is defined with
$\semvalue{\mathrm{\bf iso}(\sigma,a,b)}\rho \;\;= \;\;\mathrm{\bf iso}(\semvalue{\sigma}\rho,\;\semvalue{a}\rho,\;\semvalue{b}\rho).$
\medskip
\noindent $\bullet\;\updownarrow\!(\sigma,\tau,f).$ If $\Sigma \models \inntype{\sigma}{\mathrm{\bf type}_i}$ and $\Sigma \models \inntype{\tau}{\mathrm{\bf type}_i}$
and $\Sigma \models \intype{f}{\mathbf{Bijection}[\sigma,\tau]}$ then $\semvalue{\updownarrow\!(\sigma,\tau,f)}$
is defined with
$\semvalue{\updownarrow\!(\sigma,\tau,f)}\rho = \updownarrow\!(\semvalue{\sigma}\rho,\;\semvalue{\tau}\rho,\;\semvalue{f}\rho).$
\ignore{
\medskip
\noindent $\bullet\;a \leftrightarrowtriangle_\sigma b.$ If $\semvalue{\mathrm{\bf iso}(\sigma,a,b)}$ is defined then $a \leftrightarrowtriangle_\sigma b$ is defined
with
$\semvalue{a \leftrightarrowtriangle_\sigma b}\rho$ being $\mathrm{\bf True}$ if $\semvalue{\mathrm{\bf iso}(\sigma,a,b)}\rho$ is non-empty.
}
\medskip
\noindent $\bullet\;\mathbf{The}(\intype{x}{\sigma},\;\Phi[x]).$ For $\Sigma;\;\intype{x}{\sigma} \models \intype{\Phi[x]}{\mathrm{\bf Bool}}$ and
$\Sigma \vdash \exists!\intype{x}{\sigma}\;\;\Phi[x]$ we have that $\mathbf{The}(\intype{x}{\sigma},\;\Phi[x])$ is defined with
$\semvalue{\mathbf{The}(\intype{x}{\sigma},\;\Phi[x])}\rho = \mathbf{The}(\semvalue{\mathbf{Subtype}(\intype{x}{\sigma},\;\Phi[x])}\rho)$.
}
\caption{{\bf The semantic value $\semvalue{e}\rho$.} The clauses specify whether $\semvalue{e}$ is defined and, if it is defined, also
specify the value of $\semvalue{e}\rho$ for all $\rho \in \convalue{\Sigma}$. For a closed expression $e$ we will write $\convalue{e}$ for
$\subvalue{\epsilon}{e}\epsilon$. For example, we have $\convalue{\mathbf{Set}}$, $\convalue{\mathrm{\bf Bool}}$ and $\convalue{\mathbf{Group}}$.
The definitions of the semantic constructs $x =_\sigma y$, $\updownarrow\!(\sigma,\tau,f)$ and $\mathrm{\bf iso}(\sigma,x,y)$, and their use in figure~\ref{fig:simple},
are discussed in the text of section~\ref{subsec:semantics} and defined formally in figures~\ref{fig:Morphoids} and~\ref{fig:MoreDefs}.}
\label{fig:value}
\end{framed}
\end{figure}
In morphoid semantics all values are morphoids.
Morphoids are built from ``points'' in much the same way that sets can be
built from ur-elements. A morphoid is either a point, a Boolean value, a set of morphoids (a type) satisfying certain conditions, a function from
morphoids to morphoids satisfying certain conditions, or a pair of morphoids. We define the operations $\mathbf{Left}$, $\mathbf{Right}$, $\circ$ and
$(\cdot)^{-1}$ on morphoids and show that the class of all morphoids forms a algebraic groupoid under these operations. More specifically,
these operations satisfy properties (G1) through (G9) in figure~\ref{fig:GroupProps}.
Figure~\ref{fig:Examples} gives examples of the groupoid operations acting on types and graphs.
Morphoids are defined rigorously in section~\ref{sec:Morphoids}.
The semantics of morphoid type theory is an extension of the
semantics of predicate calculus.
The semantics involves three concepts --- variable interpretations (structures), semantic entailment, and a semantic value function.
These are defined in figures~\ref{fig:models} and \ref{fig:value}. The definitions are mutually recursive but are well-founded by reduction
of the combined syntactic complexity of $\Sigma$ and $e$. The definitions use notation defined in later sections. This top down style of
definition, common in computer code, allows for insight into the high level structure of the system without requiring full mastery of details.
Figure~\ref{fig:models} defines a structure to be a variable interpretation. This is consistent with terminology from
first order logic where a structure is a thing that assigns meaning to predicate symbols, function symbols and constant symbols. In MorTT these symbols are simply
the variables declared by a context $\Sigma$. Here we identify the notion of signature from first order logic with the notion of context.
It is possible to include structures as first class morphoid values --- this was done in earlier versions of MorTT.
For simplicity we avoid that here.
Figure~\ref{fig:models} specifies when $\convalue{\Sigma}$ is defined, i.e., when $\Sigma$ is a well formed context, and for $\convalue{\Sigma}$ defined
figure~\ref{fig:models} defines $\convalue{\Sigma}$ to be a set of structures assigning values to the variables declared in $\Sigma$.
Intuitively this is the class of structures with signature $\Sigma$.
Figure~\ref{fig:models} also defines groupoid operations and the ordering $\preceq$ on structures assuming that these are already defined on morphoids.
Figure~\ref{fig:value} defines the value of an expression under a given interpretation of the free variables of that expression. For $\convalue{\Sigma}$ defined
figure~\ref{fig:value} defines when $\semvalue{e}$ is defined, i.e., when $e$ is a well formed expression in context $\Sigma$, and if $\semvalue{e}$ is defined
then for $\rho \in \convalue{\Sigma}$ the figure defines $\semvalue{e}\rho$.
Figure~\ref{fig:value} specifies that the value function is fully compositional. For example we have
$\semvalue{s=_\sigma w}\rho$ is true if $\semvalue{s}\rho =_{\semvalue{\sigma}\rho}\;\semvalue{w}\rho$. This definition
is incomplete without a definition of the meaning of the {\em semantic} notation $x =_\sigma y$. For a morphopid type $\sigma$
and for $x,y \in \sigma$ figure~\ref{fig:Morphoids} defines $x =_\sigma y$ to mean that there exists $z \in \sigma$
with $(x@\sigma) \circ z^{-1} \circ (y@\sigma)$ defined. Here the notation $x@\sigma$ denotes an abstraction of $x$ to an abstract member of $\sigma$.
For example consider a permutation group $G$ --- a group whose members are permutations of an underlying set and where the group operation is composition of
permutations. We have $G \in \mathbf{Group}$. However, ``abstract'' groups are groups whose group elements are points. A permutation group is a group representation
rather than an abstract group. Figure~\ref{fig:Abstraction} defines the abstraction operation $G@\mathbf{Group}$ such that $G@\mathbf{Group}$
is the result of abstracting the group elements of $G$ to points. We then have $G =_{\mathbf{Group}} H$ if there exists an (abstract) group $F$
such that $(G@\mathbf{Group}) \circ F^{-1} \circ (H@\mathbf{Group})$ is defined. We use $z^{-1}$ in the definition of $x =_\sigma y$
to handle the case where $\sigma$ not closed under composition. Types not closed under composition are needed to represent the type $\updownarrow\!(\sigma,\tau,f)$ as
discussed below.
In addition to the semantic notion of equality, figure~\ref{fig:value} relies on semantic meanings for $\mathrm{\bf iso}(\sigma,x,y)$ and $\updownarrow\!(\sigma,\tau,f)$.
For a morphoid type $\sigma$ and for morphoids $x$ and $y$ with $x@\sigma$
and $y@\sigma$ defined (but without requiring $x \in\sigma$ or $y \in \sigma$) figure~\ref{fig:MoreDefs} defines $\mathrm{\bf iso}(\sigma,x,y)$ to be the type whose members are those
morphoids $z \in \sigma$ such that $(x@\sigma) \circ z^{-1} \circ (y@\sigma)$ is defined. For $x,y \in \sigma$
we then have $x =_\sigma y$ if and only if $\mathrm{\bf iso}(\sigma,x,y)$ is non-empty. This yields soundness for the inference rules in the second row of figure~\ref{fig:simple}.
However, consider the carrying relation $b_1 \leftrightarrowtriangle_{\tau[a_3]} b_2$ used in the last two rows of figure~\ref{fig:simple}. The notation $b_1 \leftrightarrowtriangle_{\tau[a_3]} b_2$
is an abbreviation for $\exists\intype{z}{\mathrm{\bf iso}(\tau[a_3],b_1,b_2)}$. Here we do not have $b_1,b_2 \in \tau[a_3]$.
For morphoid types $\sigma$ and $\tau$ in $\convalue{\mathrm{\bf type}_i}$, and a bijection $f$ in $\sigma \rightarrow \tau$,
figure~\ref{fig:MoreDefs} defines $\updownarrow\!(\sigma,\tau,f)$ to be a type whose members
are points (a point type) such that $\updownarrow\!(\sigma,\tau,f) \in \mathrm{\bf iso}(\convalue{\mathrm{\bf type}_i},\sigma,\tau)$ and such that the first
rule of the first row in figure~\ref{fig:simple} is sound.
The judgmental rules in figure~\ref{fig:jrules} introduce ``impure'' contexts --- contexts which contain judgements as assumptions.
We can define a pure context to be either the empty context, a context of the form $\Sigma;\intype{x}{\tau}$ where $\Sigma$ is pure and $\Sigma \models \intype{\tau}{\mathrm{\bf type}_i}$
or a context of the form $\Sigma;\Phi$ where $\Sigma$ is pure and $\Sigma \models \intype{\Phi}{\mathrm{\bf Bool}}$.
If $\Sigma $ is pure then $\convalue{\Sigma}$ is closed under the groupoid operations and itself forms a groupoid.
Impure contexts such as $\intype{G}{\mathbf{Group}};\;\intype{H}{\mathbf{Group}};\;G\doteq H$
are not in general closed under composition.
Note that condition (V1) in figure~\ref{fig:models} is stated so as to accommodate impure contexts.
The soundness proofs of section~\ref{sec:soundness} accommodate impure contexts for all rules.
\subsection{Natural Maps}
\label{sec:natural}
\begin{figure}[t]
\begin{framed}
{\footnotesize
~ \hfill
\unnamed{
\ant{\Sigma;\;\intype{x}{\sigma} \vdash \intype{e[x]}{\tau[x]}}}
{\ant{\Sigma \vdash \inntype{(\Lambda \intype{x}{\sigma}\;e[x])\;}{\;\Pi_{\intype{x\;}{\;\sigma}}\;\tau[x]}}}
\hfill
\unnamed
{\ant{\Sigma \vdash \inntype{f}{\Pi_{\intype{x\;}{\;\sigma}}\;\tau[x]}}
\ant{\Sigma \vdash \intype{a}{\sigma}}}
{\ant{\Sigma \vdash \intype{f\langle a\rangle}{\tau[a]}}}
\hfill
\unnamed
{\ant{\Sigma;\; \intype{x}{\sigma} \vdash \intype{e[x]}{\tau[x]}}
\ant{\Sigma \vdash \intype{a}{\sigma}}}
{\ant{\Sigma \vdash (\Lambda\intype{x}{\sigma}\;e[x]) \langle a \rangle \doteq e[a]}}
\hfill ~
$$\sigma \hookrightarrow \tau \;\;\equiv\;\; \Pi_{\intype{x\;}{\;\sigma}}\;\tau \;\;\;\;\; \;\;\mbox{$x$ not free in $\tau$}$$
}
\caption{{\bf Natural Maps.} Note that we do {\em not} have $\intype{(\Pi_{\intype{x\;}{\;\sigma}}\;\tau[x])}{\mathrm{\bf type}_i}$ --- in MorTT dependent function types are not first class types.}
\label{fig:natural}
\end{framed}
\end{figure}
Figure~\ref{fig:natural} gives three inference rules for natural maps. These rules introduce lambda expressions of the form $\inntype{(\Lambda \intype{x}{\sigma} \; e[x])}{\Pi_{\intype{x\;}{\;\sigma}}\;\tau[x]}$
and a different application notation $f\langle a \rangle$ for the application of a natural map.
The simplest case of a natural map is a map of the form
$$\inntype{(\Lambda \intype{\alpha}{\mathbf{Set}}\;\lambda\intype{x}{\sigma[\alpha]}\;\intype{e[x]}{\tau[\alpha]})}{\Pi_{\intype{\alpha\;}{\;\mathbf{Set}}}\;\sigma[\alpha] \rightarrow \tau[\alpha]}.$$
Natural maps of this form are polymorphic functions in the sense of system F \cite{Girard71,Reynolds74}.
For example the operation of composition on functions of type $\alpha \rightarrow \alpha$
can be written as
$$\inntype{(\Lambda\intype{\alpha}{\mathbf{set}}\; \lambda\intype{f}{\alpha \rightarrow \alpha} \;\lambda\intype{g}{\alpha \rightarrow \alpha}\; \lambda \intype{x}{\alpha}\;f(g(x)))}
{\Pi_{\intype{\alpha\;}{\;\mathbf{Set}}}\;(\alpha \rightarrow \alpha) \times (\alpha \rightarrow \alpha) \rightarrow (\alpha \rightarrow \alpha)}$$
We can define a more general polymorphic composition operation for composing $\intype{f}{\alpha \rightarrow \beta}$ and $\intype{g}{\gamma \rightarrow \alpha}$
as a natural map whose first argument is a triple of sets.
Natural maps can also be defined on richer classes such as the class of groups or topological spaces.
There is a natural map taking an arbitrary group $G$ to the canonical permutation group on $G$ achieved by taking the natural action of group elements
on group elements. We can also define a natural map taking a pointed topological space to its fundamental group of loops at the selected point. More explicitly
we have
$$\intype{X}{\mathbf{TOP}};\intype{x}{\pi_1(X)} \vdash \intype{\mathbf{Pair}(P[X,x],C[X,x])}{\mathbf{Group}}$$
where $\pi_1(X)$ is the type of the points of $X$,
where $P[X,x]$ denotes the type whose elements are equivalence classes of loops from $x$ to $x$ and where $C[X,x]$ denotes the operation of composition
on these classes of loops. We can then define a natural map from a pointed topological space to its fundamental group as
$$\Lambda\intype{p}{\mathbf{PairOf}(\intype{X}{\mathbf{TOP}},\;\intype{x}{\pi_1(X)})}\;\mathbf{Pair}(P[\pi_1(p),\pi_2(p)],\;C[\pi_1(p),\pi_2(p)]).$$
We can show that for a connected space the group structure is independent of the choice of the base point. Using the axiom of choice in figure~\ref{fig:rules}
we can derive that there exists a function
$$\intype{\mathbf{FUND}\;}{\;\mathbf{CTOP} \rightarrow \mathbf{Group}}$$
where $\mathbf{CTOP}$ is the type of connected topological spaces and such that
$$\forall \intype{X}{\mathbf{CTOP}} \;\;\;\forall\intype{x}{\pi_1(X)}\;\;\;\;\mathbf{FUND}(X) =_{\mathbf{Group}} \mathbf{Pair}(P[X,x],C[X,x]).$$
We do not have that $\mathbf{FUND}(X)$ is a group of equivalence classes of loops --- we have that $\mathbf{FUND}(X)$
is an arbitrary group isomorphic to a group on loop classes.
There is no natural map that takes a connected topological space to a group of loop classes because,
by Voldemort's theorem (section~\ref{sec:Voldemort}), there is no natural base point on a sphere or a torus.
\subsection{Voldemort's Theorem}
\label{sec:Voldemort}
Voldemort's theorem implies that certain objects exist but cannot be named. For example, it is not possible to name any particular point on the
topological (or geometric) circle $S^1$. The statement of Voldemort's theorem appears at first to be unrelated to naming, but its implications for naming are
discussed below.
\begin{theorem}[Voldemort's Theorem]
The following rule can be derived from the rules in figures~\ref{fig:expressions} through~\ref{fig:pairs}.
~ \hfill
\unnamed
{\ant{\Sigma;\;\intype{x}{\sigma} \vdash \intype{e[x]}{\tau[x]}}
\ant{\Sigma \vdash \mathbf{Pair}(a,b) =_{\mathbf{PairOf}(\intype{x\;}{\;\sigma},\;\intype{y\;}{\;\tau[x]})}\;\;\mathbf{Pair}(a,e[a])}}
{\ant{\Sigma\;\vdash b =_{\tau[a]} e[a]}}
\hfill ~
\end{theorem}
\begin{proof}
We consider the following instance of the substitution rule.
~ \hfill
\unnamed
{\ant{\Sigma;\; \intype{p}{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})} \vdash \intype{\Phi[p]}{\mathrm{\bf Bool}}}
\ant{\Sigma \vdash q =_{\mathbf{PairOf}(\intype{x\;}{\;\sigma},\;\intype{y\;}{\;\tau[x]})}\;\;r}}
{\ant{\Sigma \vdash \Phi[q] \Leftrightarrow \Phi[r]}}
\hfill ~
Given the first antecedent of the lemma we can derive
$$\Sigma; \intype{p}{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})} \vdash \intype{(\pi_2(p) =_{\tau[\pi_1(p)]} e[\pi_1(p)])}{\mathrm{\bf Bool}}.$$
Applying the above instance of the substitution rule to the pairs $\mathbf{Pair}(a,b)$ and $\mathbf{Pair}(a,e[a])$ of the second premise of the lemma
gives $(b =_{\tau[a]} e[a]) \Leftrightarrow (e[a] =_{\tau[a]} e[a])$ which proves the theorem.
\end{proof}
Voldemort's theorem implies that it is not possible to name a point on the topological circle or the topological torus. Consider the topological sphere $S^2$
$$S^2 \equiv \mathbf{The}(\intype{X}{\mathbf{TOP}}\;X =_{\mathbf{TOP}} \mathbf{Pair}(\sigma,\;O))$$
where $\sigma$ is the set of points in $\mathbb{R}^3$ with unit length and $O$ is subspace topology of the standard topology on $\mathbb{R}^3$.
We can also define the class of topological spaces isomorphic (homeomorphic) to $S^2$ by
$$\mathbf{Sphere} \equiv \mathbf{SubType}(\intype{X}{\mathbf{TOP}},\;X =_{\mathbf{TOP}} S^2).$$
We now have $\intype{S^2}{\mathbf{Sphere}}$ and $\intype{\mathbf{Sphere}}{\mathbf{Class}}$ --- $S^2$ is a single (arbitrary) sphere
while $\mathbf{Sphere}$ is the class of all spheres.
\medskip
\noindent Now suppose that we could name a point on the sphere in the sense that we have
$$\intype{X}{\mathbf{Sphere}} \vdash \intype{e[X]}{\pi_1(X)}.$$
There exists a point $b$ on $S^2$ different from the point $e[S^2]$, i.e., such that $b \not =_{\pi_1(S^2)} e[S^2]$.
But we also have
$$\mathbf{Pair}(S^2,b) =_{\mathbf{PairOf}(\intype{X\;}{\;\mathbf{Sphere}},\;\intype{z\;}{\;\pi_1(X)})}\;\mathbf{Pair}(S^2,e[S^2]).$$
By Voldemort's theorem we then have $b =_{\pi_1(S^2)} e[S^2]$ giving a contradiction.
A similar argument yields that one cannot name any particular node of the complete graph or any particular isomorphism (linear bijection) of a finite dimensional
vector space with its dual.
\subsection{Cryptomorphism}
\label{sec:cryptomorphism}
Two types are cryptomorphic in the sense of Birkoff and Rota \cite{Rota}
if they ``provide the same data''. For example a group can be defined to be a four-tuple of a set,
a group operation, an identity element and an inverse operation satisfying certain equations. Alternatively, a group can be defined to be a
pair of a set and a group operation such that an identity element and an inverse operation exist. In the later case it can be shown that the identity
element and the group operation are unique. We might call these types $\mathbf{FourTupleGroup}$ and $\mathbf{PairGroup}$.
These are different types with different elements.
However, every four-tuple-group can be converted to a pair-group simply by dropping the second two components. Conversely, any pair-group can be converted to a four-tuple-group
by extending the pair to a four tuple whose last two elements are the unique identity element and inverse operation.
Here we suggest that two classes $\sigma$ and $\tau$ be considered cryptomorphic if there exists natural maps $\inntype{f}{\sigma \hookrightarrow \tau}$
and $\inntype{g}{\tau \hookrightarrow \sigma}$ such that $\forall \intype{x}{\sigma}\;g\langle f \langle x\rangle \rangle \doteq x$ and
$\forall \intype{y}{\tau}\;f\langle g \langle y \rangle \rangle \doteq y$.
As discussed in section~\ref{sec:isorules} we can use cryptomorphisms to strength the power of the rules in figure~\ref{fig:simple}.
More specifically, the following rule can be derived from the substitution rule.
~ \hfill
\unnamed
{\ant{\Sigma \vdash \inntype{f}{\sigma \hookrightarrow \tau}}
\ant{\Sigma \vdash \inntype{g}{\tau \hookrightarrow \sigma}}
\ant{\Sigma;\intype{x}{\sigma} \vdash x \doteq g(f(x))}
\ant{\Sigma;\intype{y}{\tau} \vdash y \doteq f(g(x))}
\ant{\Sigma \vdash \intype{G,H}{\sigma}}}
{\ant{\Sigma \vdash (G =_\sigma H) \Leftrightarrow (f\langle G \rangle =_\tau f\langle H \rangle)}}
\hfill ~
For this rule to strengthen the power of the rules in figure~\ref{fig:simple} it is important that we
use natural maps rather than morphoid functions. More specifically, to derive $f\langle G \rangle =_\tau f\langle H \rangle$ using the rules in figure~\ref{fig:simple} it is important
that the types within $G$ are carried over to types within $f\langle G \rangle$ up to absolute equality.
An example is the derivation of isomorphism equations for colored graphs as discussed in section~\ref{sec:isorules}.
\ignore{
In MorTT cryptomorphism is different from class-isomorphism.
In particular, in MorTT $A =_{\mathbf{Class}} B$ is treated the same as $a =_{\mathbf{Set}} b$ in that two classes $A$ and $B$
are class-isomorphic if there exists any bijection between them.
For example, the class of finite total orders is class-isomorphic to the class of finite graphs because
both of these types have a countably infinite number of isomorphism classes and hence there exists a bijection between them.
However, these two classes are not cryptomorphic ---
no absolute bijection between them can be defined by a pair of natural maps. There is no natural total ordering of the nodes in a complete graph.
Treating class-isomorphism as ``same cardinality'' seems consistent with mathematical practice.
For example, consider the following definition of the type of (large) categories where ${\cal O}$ is the class of objects of the category
and ${\cal M}$ is the class of morphisms of the category.
$$\mathbf{CAT} \equiv \mathbf{TupleOf}\left(\begin{array}{l}
\intype{{\cal O}}{\mathbf{Class}},\;\intype{{\cal M}}{\mathbf{Class}},\;\intype{\mathbf{Dom}}{{\cal M} \rightarrow {\cal O}},
\;\intype{\mathbf{Ran}}{{\cal M} \rightarrow {\cal O}}, \\
\intype{\mathbf{Comp}}{({\cal M} \times {\cal M} \times {\cal M}) \rightarrow \mathrm{\bf Bool}} \\
s.t.\;\Phi[{\cal O},\;{\cal M},\;\mathbf{Dom},\;\mathbf{Ran},\;\mathbf{Comp}]
\end{array}\right)$$
In addition to the usual axioms of category theory, $\Phi[{\cal O},\;{\cal M},\;\mathbf{Dom},\;\mathbf{Ran},\;\mathbf{Comp}]$ must specify that $\mathbf{Comp}$ is a partial function
where $\mathbf{Comp}(m_1,m_2)$ has a value only when $\mathbf{Ran}(m_2) = \mathbf{Dom}(m_1)$. This admitedly exhibits a weakness of MorTT --- dependent function types
(the type of natural maps) are not first class types in MorTT. However, the point being made here is independent of the issue of dependent function types.
Note that we have $\epsilon \vdash \intype{\mathbf{CAT}}{\mathrm{\bf type}_2}$ --- the collection of all large categories is not itself a class but is instead in the next higher
level of types.
However, two categories $C_1$ and $C_2$ are category-isomorphic if there exists a bijection from the objects of $C_1$ to the objects of $C_2$
and a bijection from the morphisms of $C_1$ to the morphisms of $C_2$ such that these bijections carry the
functions $\mathbf{Dom}$, $\mathbf{Ran}$ and the partial function $\mathbf{Comp}$
of $C_1$ to the corresponding functions of $C_2$ as specified in figure~\ref{fig:simple}. Category-isomorphism is of
the same nature as group-isomorphism.
In MorTT Boolean propositions on class variables cannot distinguish between classes with the same
cardinality (number of equivalence classes). Boolean propositions on category variables cannot distinguish
between isomorphic categories where isomorphism of categories is defined by arbitrary bijections.
}
\eject
\section{Morphoids}
\label{sec:Morphoids}
We will define (Platonic) Morphoid theory in terms of (Platonic) set theory.
In order to provide a semantics
for the type constants $\mathrm{\bf type}_i$ we assume an infinite number of inaccessible cardinals.
We will write set-theoretic equality (equality in the universe of pure sets) simply as $=$ and use this as the interpretation of
absolute equality $\doteq$.
\medskip
We let 0 denote the empty set and let 1 denote $\{0\}$.
We represent the pair $\struct{x,y}$ as $\{x,\;\{x,y\}\}$.
We represent lists by implementing the empty list as the empty set and implementing a non-empty list as a pair $\struct{x,r}$
where $x$ is the first element of the list and $r$ is the rest of the list (which might or might not be empty).
We represent bit strings as lists of 0s and 1s.
We represent a byte as a list of eight bits. We represent byte strings as lists of bytes.
We define a symbol to be a list of bytes and write symbols using the standard ASCII conventions. For example, we have the symbols
$\tagg{FOO}$ and $\tagg{BAR}$. An expression is either a symbol or a list of expressions.
\medskip
\noindent {\bf Left-Right Duality.}
Morphoid type theory involves a left-right duality. A duality is a cryptomorphism between a class and itself.
For example, the reversal of a partial order is a natural map on the class of partial orders that is its own inverse. Similarly the reversal of all points
in a morphoid might be interpretable as a natural map on the class of morphoids. But rather than try to apply morphoid theory to itself
we will simply claim that some kind of intuitive left-right duality exists.
This duality makes certain statements obvious. For example, we have $\semvalue{e}(\rho^{-1}) = (\semvalue{e}\rho)^{-1}$ and $(x^{-1})@{\cal T} = (x@{\cal T})^{-1}$.
Also, if $\rho \in \convalue{\Sigma}$ then $\rho^{-1} \in \convalue{\Sigma}$. We will generally use facts that follow from left-right duality without comment.
\subsection{Weak Morphoids}
\label{subsec:weak}
We first define weak morphoids. The morphoids will be a subclass of the weak morphoids.
Figure~\ref{fig:WeakMorphs} defines the weak morphoids and figure~\ref{fig:operations} defines the groupoid operations on
weak morphoids. The definitions are mutually recursive --- the definition of weak morphoids involves conditions stated in terms of the groupoid
operations and the groupoid operations are defined on the weak morphoids. However, as discussed in slightly more detail below,
these definitions are well founded
and do define the class of weak morphoids and the groupoid operations on them.
Condition (T1) in figure~\ref{fig:WeakMorphs} is central to morphoid type theory. Bijections are central to the concept of isomorphism and condition (T1)
allows types to represent bijections.
For any point type $\sigma$ we have that $\mathbf{Left}(\sigma)$ is an equivalence relation on the left indeces of $\sigma$ and $\mathbf{Right}(\sigma)$ is an equivalence relation
on the right indices of $\sigma$. Furthermore, $\sigma$ itself defines a bijection between the equivalence classes of $\mathbf{Left}(\sigma)$ and the equivalence classes of
$\mathbf{Right}(\sigma)$. This can be seen in the point types in figure~\ref{fig:Examples}. Any point type $\sigma$ satisfying condition (T1) has the structure
shown in the figure where each equivalence classes of $\sigma$ is a cross product of a left index set and a right index set. This is related to property ($\simeq$.B)
in figure~\ref{fig:HardProps}. Morphoid types are directed
from left to right. This is needed in order for types to represent bijections. For $x \in \sigma$ we do not in general have $x^{-1} \in \sigma$.
Hence a morphoid type $\sigma$ is not in general a groupoid.
\begin{figure}[p]
\begin{framed}
{\small
{\bf Weak Morphoid.} A weak morphoid is one of the following.
\begin{itemize}
\item A morphoid point --- a pair $(\tagg{POINT},(i,j))$ where $i$ and $j$ are arbitrary values (arbitrary elements of the set-theoretic universe).
We abbreviate $(\tagg{POINT},(i,j))$ as $\mathrm{\bf Point}(i,j)$.
\item A Boolean value --- one of two pairs $(\tagg{BOOL},0)$ or $(\tagg{BOOL},1)$. We will abbreviate $(\tagg{BOOL},0)$ by $\mathrm{\bf False}$
and $(\tagg{BOOL},1)$ by $\mathrm{\bf True}$.
\item A weak type --- a pair $\sigma = (\tagg{TYPE},s)$ where $s$ is a set of weak morphoids satisfying the following type (T) property.
\begin{itemize}
\item[(T1)] for $x,y,x \in s$ with $x \circ y^{-1} \circ z$ defined we have $(x \circ y^{-1} \circ z) \in s$.
\end{itemize}
We will write $x \in \sigma$ for $x \in s$.
\item A weak function --- a pair $f = (\tagg{FUN},s)$ where $s$ is a functional set of pairs of weak morphoids satisfying the function (F) condition below.
To state the conditions we write $\mathbf{Dom}(f)$ for $(\tagg{TYPE},w)$
where $w$ is the set of morphoids occurring as the first component of some pair in $s$ and for $x \in \mathbf{Dom}(f)$ we write $f[x]$ for the unique $y$
such that $(x,y) \in s$. A weak function must satisfy the condition that $\mathbf{Dom}(f)$ is a weak type containing only points (a point type),
that for $x \in \mathbf{Dom}(f)$ we have that $f[x]$ is a weak morphoid, and
\begin{itemize}
\item[(F1)] for $x,y \in \mathbf{Dom}(f)$ with $x \simeq_{\mathbf{Dom}(f)} y$ we have $f[x] = f[y]$ (absolute equality).
\end{itemize}
\item A weak pair --- a pair $(\tagg{PAIR},(x,y))$ where $x$ and $y$ are weak morphoids. We abbreviate $(\tagg{PAIR},(x,y))$
as $\mathbf{Pair}(x,y)$.
\end{itemize}
\medskip ${\bf x \simeq_\sigma y}$. For $x,y \in \sigma$ we defined $x \simeq_\sigma y$ to mean that there exists $z \in \sigma$
with $x \circ z^{-1} \circ y$ defined.
\medskip
{\bf Morphoid Rank $R(x)$}. For a point or Boolean value $x$, $R(x) = 0$.
For a weak type $\sigma$, $R(\sigma)$ is the least ordinal greater than $R(x)$ for all $x \in \sigma$.
For a weak function $f$, $R(f)$ is the least ordinal greater than $R(\mathbf{Dom}(f))$ and greater than $R(f[x])$ for all
$x \in \mathbf{Dom}(f)$. $R(\mathbf{Pair}(x,y))$ is the least ordinal greater than both $R(x)$ and $R(y)$.
\medskip
{$\bf (\sim)$} For weak morphoids $x$ and $y$ we write $x \sim y$ if there exists a weak morphoid $z$ with $x \circ z \circ y$ defined.
}
\caption{{\bf Weak Morphoids.} Weak morphoids (and morphoids) are built from points in much the same way that sets can be built from ur-elements.
Figure~\ref{fig:Morphoids} defines morphoids to be weak morphoids that satisfy additional type and function properties.
Weak morphoids satisfy the groupoid properties
but do not in general satisfy the ordering properties of morphoids. Note that the domain of a weak function is always a set of points.
We will define $f(x)$ to be $f[x@\mathrm{\bf Point}]$.}
\label{fig:WeakMorphs}
\end{framed}
\end{figure}
\begin{figure}[t]
\begin{framed}
{\small
\begin{eqnarray*}
\mathbf{Left}((\tagg{BOOL},v)) & = & \mathbf{Right}((\tagg{BOOL},v)) = (\tagg{BOOL},v)^{-1} = (\tagg{BOOL},v) \\
(\tagg{BOOL},v) \circ (\tagg{BOOL},v) & = & (\tagg{BOOL},v) \\
\\
\mathbf{Left}(\mathrm{\bf Point}(i,j)) & = & \mathrm{\bf Point}(i,i) \\
\mathbf{Right}(\mathrm{\bf Point}(i,j)) & = & \mathrm{\bf Point}(j,j) \\
\mathrm{\bf Point}(i,j)^{-1} & = & \mathrm{\bf Point}(j,i) \\
\mathrm{\bf Point}(i,j) \circ \mathrm{\bf Point}(j,k) & = & \mathrm{\bf Point}(i,k) \\
\\
\mathbf{Left}((\tagg{TYPE},s)) & = & (\tagg{TYPE},\{p\circ q^{-1}:\;p,q \in s\}) \\
\mathbf{Right}((\tagg{TYPE},s)) & = & (\tagg{TYPE},\{p^{-1}\circ q:\;p,q \in s\}) \\
(\tagg{TYPE},s)^{-1} & = & (\tagg{TYPE},\{p^{-1}:\;p \in s\}) \\
(\tagg{TYPE},s) \circ (\tagg{TYPE},w) & = & (\tagg{TYPE},\{p \circ q:\;p \in s,\;q \in w\}) \\
\\
\mathbf{Left}(\mathbf{Pair}(x,y)) & = & \mathbf{Pair}(\mathbf{Left}(x),\mathbf{Left}(y)) \\
\mathbf{Right}(\mathbf{Pair}(x,y)) & = & \mathbf{Pair}(\mathbf{Right}(x),\mathbf{Right}(y)) \\
\mathbf{Pair}(x,y)^{-1} & = & \mathbf{Pair}(x^{-1},y^{-1}) \\
\mathbf{Pair}(x,y) \circ \mathbf{Pair}(z,w) & = & \mathbf{Pair}(x\circ z,\;y\circ w) \\
\\
\mathbf{Left}((\tagg{FUN},s)) & = & (\tagg{FUN},\;\{(x_1 \circ x_2^{-1},\;y_1 \circ y_2^{-1}):\;(x_1,y_1),(x_2,y_2) \in s\}) \\
\mathbf{Right}((\tagg{FUN},s)) & = & (\tagg{FUN},\;\{(x_1^{-1} \circ x_2,\;y_1^{-1} \circ y_2):\;(x_1,y_1),(x_2,y_2) \in s\}) \\
(\tagg{FUN},s)^{-1} & = & (\tagg{FUN},\;\{(x^{-1},y^{-1}):\;(x,y) \in s\}) \\
(\tagg{FUN},s)\circ(\tagg{FUN},w) & = & (\tagg{FUN},\;\{(x_1\circ x_2,\;y_1\circ y_2):\; (x_1,y_1) \in s, \;(x_2,y_2) \in w\})
\end{eqnarray*}
}
\vspace{-1em}
\caption{{\bf The Groupoid Operations.} We have that $x \circ y$
is defined if and only if $\mathbf{Right}(x) = \mathbf{Left}(y)$.}
\label{fig:operations}
\end{framed}
\end{figure}
\begin{figure}[t]
\begin{framed}
{\small
\noindent
{\small \bf (Groupoid Properties)} The groupoid properties (G1) through (G9) in figure~\ref{fig:GroupProps} where (G1) and (G2) are modified
to state that the operations applied to weak morphoids yield weak morphoids.
\medskip {\small \bf (Fun-Left-Right)} For a weak function $f$ we have that $\mathbf{Left}(f)$ is the unique function with $\mathbf{Dom}(\mathbf{Left}(f)) = \mathbf{Left}(\mathbf{Dom}(f))$ and
such that $\mathbf{Left}(f)(\mathbf{Left}(x)) = \mathbf{Left}(f[x])$ and similarly for $\mathbf{Right}$.
\medskip {\small \bf (Funs-Composable)} For weak functions $f$ and $g$ we have that $f \circ g$ is defined if and only if $\mathbf{Dom}(f) \circ \mathbf{Dom}(g)$ is defined
and for $x \in \mathbf{Dom}(f)$ and $y \in \mathbf{Dom}(g)$ with $x \circ y$ defined we have $f[x] \circ g[y]$ defined.
\medskip {\small \bf (Fun-Composition)} For weak functions $f$ and $g$ with $f \circ g$ defined
we have that $f \circ g$ is the unique morphoid function such that $\mathbf{Dom}(f \circ g) = \mathbf{Dom}(f) \circ \mathbf{Dom}(g)$
and for $x \in \mathbf{Dom}(f)$ and $y \in \mathbf{Dom}(g)$ with $x \circ y$ defined we have $(f \circ g)[x \circ y] = f[x] \circ g[y]$.
\medskip {\small \bf (Composables-Equivalent)} For $x,y \in \sigma$, if either $x \circ y^{-1}$ or $x^{-1} \circ y$ are defined then
$x\simeq_\sigma y$.
\medskip {\small \bf ($\simeq$.A)} For any weak type $\sigma$ we have that $\simeq_\sigma$ is an equivalence relation on the members of $\sigma$.
\bigskip {\small \bf ($\simeq$.B)} For morphoid types $\sigma$ and $\tau$ with $\sigma \circ \tau$ defined and for $\incat{x_1,x_2}{\sigma}$ and $\incat{y_1,y_2}{\tau}$
with $x_1 \circ y_1$ and $x_2 \circ y_2$ defined, we have $(x_1 \circ y_1) \simeq_{\sigma\circ \tau} (x_2 \circ y_2)$ if and only if
$x_1 \simeq_\sigma x_2$ if and only if $y_1 \simeq_\tau y_2$.
\medskip {\small \bf (Partner)} For morphoid types $\sigma$ and $\tau$ with $\sigma \circ \tau$ defined we have that for all $x \in \sigma$ there exists $y \in \tau$
with $x \circ y$ defined and, similarly, for all $y \in \tau$ there exists $x \in \sigma$ with $x \circ y$ defined.
\medskip {\small \bf (Rank Preservation)} We have $R(x^{-1}) = R(\mathbf{Left}(x)) = R(\mathbf{Right}(x)) = R(x)$
and for $x \circ y$ defined we have $R(x) = R(y) = R(x \circ y)$.
\medskip {\small \bf ($\sim$.A)} The relation $\sim$ is an equivalence relation on weak morphoids.
\medskip {\small \bf ($\sim$.B)} We have $x \sim x^{-1} \sim \mathbf{Left}(x) \sim \mathbf{Right}(x)$
and for $x \circ y$ defined we have $x \sim y \sim (x \circ y)$.
}
\caption{{\bf Weak Morphoid Properties.}}
\label{fig:HardProps}
\end{framed}
\end{figure}
Figure~\ref{fig:WeakMorphs} also defines the notion of morphoid rank. Morphoid rank is analogous to set-theoretic rank except that the
rank of Boolean values and points is zero. Morphoid rank is well defined on an even larger class of tagged values
not involving any conditions on types and functions. The recursive definition
of morphoid rank is itself well-founded by reduction of set-theoretic rank.
The recursive definitions in figures~\ref{fig:WeakMorphs} and~\ref{fig:operations} are mutually well founded by reduction of morphoid rank.
Figure~\ref{fig:HardProps} states various properties of weak morphoids involving the groupoid operations.
We prove the properties in figure~\ref{fig:HardProps} except for ($\sim$.A) and ($\sim$.B) by a single simultaneous induction on morphoid rank.
Each instance of each property is associated with a rank. The rank of an instance of a property is the maximum rank of the weak morphoid variables in the statement of the property.
For example, property (G2) states that for two weak morphoids $x$ and $y$ with $\mathbf{Right}(x)$ = $\mathbf{Left}(y)$
we have that $x \circ y$ is a weak morphoid. The rank of an instance of this property is the maximum of the ranks of $x$ and $y$.
The property (Rank-Preservation) has two kinds of instances --- instances of the form $R(x^{-1}) = R(\mathbf{Left}(x)) = R(\mathbf{Right}(x)) = R(x)$
and instances of the form that $x \circ y$ defined implies that $R(x) = R(y) = R(x \circ y)$. The rank of an instance of the first type
is the rank of $x$ and the rank of an instance of the second type has rank equal to the maximum rank of $x$ and $y$.
The rank of an instance of a property is generally clear but it is also specified explicitly in the first line of the proof of each property.
Under the induction hypothesis that
all property instances of rank less than $\beta$ hold, we show that all property instances of rank $\beta$ hold.
This large simultaneous induction proof is spread over most of the remainder of this subsection.
At the end of section, after completion of the simultaneous induction, properties ($\sim$.A) and ($\sim$.B) are proved from the earlier properties.
In the proof of a property instance we can assume any property instances at any rank smaller than the rank of the instance we are proving
as well as previously proven lemmas at rank equal to the rank of the instance under consideration. The order of lemmas remains important.
The induction hypotheses ensure that expressions built from lower-rank morphoids denote lower-rank morphoids ---
the lower-rank rank morphoids are closed under the groupoid operations. This means that in each proof we have access to the full
algebra of the groupoid operations defined by properties (G1) through (G9) for the morphoids of lower rank.
We first prove the groupoid properties (G1) through (G9).
The groupoid properties are immediate for points and Boolean values. For pairs all properties follow straightforwardly from the induction
hypotheses. For example, to show that $\mathbf{Left}(\mathbf{Pair}(x,y))$ is a weak morphoid (property (G1)) we note that
by definition $\mathbf{Left}(\mathbf{Pair}(x,y)) = \mathbf{Pair}(\mathbf{Left}(x),\mathbf{Left}(y))$ and by the induction hypothesis $\mathbf{Left}(x)$ and $\mathbf{Left}(y)$
are weak morphoids. We explicitly prove the groupoid properties only for types and functions. We first consider types.
\begin{lemma}[G1 for Types]
For a weak type $\sigma$ we have that $\sigma^{-1}$, $\mathbf{Left}(\sigma)$ and $\mathbf{Right}(\sigma)$ are also weak types.
\end{lemma}
\begin{proof}
The morphoid rank of an instance of this lemma is the rank of $\sigma$.
\medskip
\noindent The duality of left and right implies the result for $\sigma^{-1}$.
We will show that $\mathbf{Left}(\sigma)$ is a weak type --- the case for $\mathbf{Right}(\sigma)$ is similar.
We let $x$ range over members of $\sigma$.
The elements of $\mathbf{Left}(\sigma)$ are the values of the form $x_1 \circ x_2^{-1}$. By the induction hypothesis for the groupoid properties and rank preservation
we have that every such value is a weak morphoid with rank less than the rank of $\sigma$.
We must show that $\mathbf{Left}(\sigma)$ satisfies (T1).
Suppose that
$(x_1 \circ x_2^{-1}) \circ (x_3 \circ x_4^{-1})^{-1} \circ (x_5 \circ x_6^{-1})$ is defined. By the induction hypothesis for the
groupoid properties and rank preservation we have
$$(x_1 \circ x_2^{-1}) \circ (x_3 \circ x_4^{-1})^{-1} \circ (x_5 \circ x_6^{-1}) = (x_1 \circ x_2^{-1} \circ x_4) \circ (x_6 \circ x_5^{-1} \circ x_3)^{-1}$$
which proves (T1) for $\mathbf{Left}(\sigma)$.
\end{proof}
\begin{lemma}[G2 for Types]
For two morphoid types $\sigma$ and $\tau$ with $\sigma \circ \tau$ defined, if $\sigma$ and $\tau$ are weak types
then $\sigma \circ \tau$ is a weak type.
\end{lemma}
\begin{proof}
The rank of an instance of this lemma is the maximum rank of $\sigma$ and $\tau$.
\medskip
\noindent We let $x$ range over elements of $\sigma$ and $y$ range over elements of $\tau$.
The elements of $\sigma \circ \tau$ are the values of the form $x \circ y$. By the induction hypotheses,
all such values are weak morphoids. We must show that $\sigma \circ \tau$ satisfies (T1).
We must show that for $(x_1\circ y_1)\circ (x_2 \circ y_2)^{-1} \circ (x_3 \circ y_3)$ defined
we have that this composition is in $\sigma \circ \tau$. Since $\mathbf{Right}(\sigma) = \mathbf{Left}(\tau)$,
every value of the form $y_1 \circ y_2^{-1}$ can be written as $x_1^{-1} \circ x_2$. By the induction hypotheses we have the following.
\begin{eqnarray*}
(x_1\circ y_1)\circ (x_2 \circ y_2)^{-1} \circ (x_3 \circ y_3) & = & x_1\circ (y_1 \circ y_2^{-1}) \circ x_2^{-1} \circ x_3 \circ y_3 \\
& = & x_1\circ (x_4^{-1} \circ x_5) \circ x_2^{-1} \circ x_3 \circ y_3 \\
& = & ((x_1\circ x_4^{-1} \circ x_5) \circ x_2^{-1} \circ x_3) \circ y_3 \\
& = & x_7 \circ y_3
\end{eqnarray*}
\end{proof}
\begin{lemma}[G3 for Types] $\mathbf{Left}(\sigma^{-1}) = \mathbf{Right}(\sigma)$ and $\mathbf{Right}(\sigma^{-1}) = \mathbf{Left}(\sigma)$.
\end{lemma}
\begin{proof}
The values in $\mathbf{Left}(\sigma^{-1})$ are the values of the form $x_1^{-1} \circ (x_2^{-1})^{-1}$.
But by the groupoid induction hypotheses these are the same as the values of the form $x_1^{-1} \circ x_2$. But these are exactly the values in $\mathbf{Right}(\sigma)$.
\end{proof}
\begin{lemma}[G4 for Types] $\mathbf{Left}(\sigma \circ \tau) = \mathbf{Left}(\sigma)$ and $\mathbf{Right}(\sigma \circ \tau) = \mathbf{Right}(\tau)$.
\end{lemma}
\begin{proof}
We will show $\mathbf{Left}(\sigma \circ \tau) = \mathbf{Left}(\sigma)$.
We will use $x$ to range over elements of $\sigma$
and $y$ range over elements of $\tau$.
We first show that every member of $\mathbf{Left}(\sigma \circ \tau)$ is an member of $\mathbf{Left}(\sigma)$.
A member of $\mathbf{Left}(\sigma \circ \tau)$ has the form
$(x_1 \circ y_1) \circ (x_2 \circ y_2)^{-1}$.
Since $\mathbf{Right}(\sigma) = \mathbf{Left}(\tau)$ we have that every value of the form $y_1^{-1} \circ y_2$ can be written as $x_1 \circ x_2^{-1}$.
By the groupoid induction hypotheses we then have the following.
\begin{eqnarray*}
(x_1 \circ y_1) \circ (x_2 \circ y_2)^{-1} & = & x_1 \circ (y_1 \circ y_2^{-1}) \circ x_2^{-1} \\
& = & x_1 \circ (x_3^{-1} \circ x_4) \circ x_2^{-1} \\
& = & (x_1 \circ x_3^{-1} \circ x_4) \circ x_2^{-1} \in \mathbf{Left}(\sigma)
\end{eqnarray*}
For the converse we consider a value $x_1 \circ x_2^{-1}$ in $\mathbf{Left}(\sigma)$. For this we have the following.
\begin{eqnarray*}
x_1 \circ x_2^{-1} & = & x_1 \circ x_2^{-1} \circ x_2 \circ x_2^{-1} \\
& =& x_1 \circ (x_2^{-1} \circ x_2) \circ x_2^{-1} \\
& =& x_1 \circ (y_1 \circ y_2^{-1}) \circ x_2^{-1} \\
& =& (x_1 \circ y_1) \circ (y_2^{-1} \circ x_2^{-1}) \\
& =& (x_1 \circ y_1) \circ (x_2 \circ y_2)^{-1} \in \mathbf{Left}(\sigma \circ \tau)
\end{eqnarray*}
\end{proof}
\begin{lemma}[G5 for Types] $(\sigma \circ \tau) \circ \gamma$ = $\sigma \circ (\tau \circ \gamma)$.
\end{lemma}
\begin{proof} Properties (G2) and (G4) proved above imply that $(\sigma \circ \tau)\circ \gamma$ is defined if and only if $\sigma \circ (\tau \circ \gamma)$ is defined.
The values in $(\sigma \circ \tau) \circ \gamma$ are the values of the form
$(x \circ y) \circ z$ for $\incat{x}{\sigma}$, $\incat{y}{\tau}$ and $\incat{z}{\gamma}$. But by the groupoid induction hypotheses
these are the same as the members of $\sigma \circ (\tau \circ \gamma)$.
\end{proof}
\begin{lemma}[G6 for Types] $\sigma^{-1} \circ \sigma \circ \tau = \tau$ and $\sigma \circ \tau \circ \tau^{-1} = \sigma$.
\end{lemma}
\begin{proof} We will show that if $\sigma \circ \tau$ is defined then $\sigma^{-1} \circ \sigma \circ \tau = \tau$.
We will let $x$ range over elements of $\sigma$ and $y$ range over elements of $\tau$.
We first show that every value $y$ in $\tau$ is in $\sigma^{-1} \circ \sigma \circ \tau$. For this we note
$$ y \; = \; (y \circ y^{-1}) \circ y \; = \; (x_1^{-1} \circ x_2) \circ y \;\in \; \sigma^{-1} \circ \sigma \circ \tau$$
Conversely, consider $x_1^{-1} \circ x_2 \circ y \in \sigma^{-1} \circ \sigma \circ \tau$.
For this case we have $(x_1^{-1} \circ x_2) \circ y_1 = (y_2 \circ y_3^{-1}) \circ y_1 \in \tau$.
\end{proof}
\begin{lemma}[G7 for Types] $\mathbf{Right}(\sigma) = \sigma^{-1} \circ \sigma$ and $\mathbf{Left}(\sigma) = \sigma \circ \sigma^{-1}$
\end{lemma}
\begin{proof}
We will show that $\mathbf{Left}(\sigma) = \sigma \circ \sigma^{-1}$. Property (G3) above implies that
$\sigma \circ \sigma^{-1}$ is defined. The result is then immediate from the definitions of $\mathbf{Left}(\sigma)$ and $\sigma \circ \sigma^{-1}$.
\end{proof}
It is a fact of groupoids that (G8) and (G9) follow from (G1) through (G7).
\begin{lemma}[(G1) for Functions and (Fun-Left-Right)]
\label{lem:fun1}
For a morphoid function $f$ we have that $\mathbf{Left}(f)$ is the unique weak function such that $\mathbf{Dom}(\mathbf{Left}(f)) = \mathbf{Left}(\mathbf{Dom}(f))$ and for $x \in \mathbf{Dom}(f)$
we have $\mathbf{Left}(f)(\mathbf{Left}(x))$ = $\mathbf{Left}(f[x])$. The dual statement holds for $\mathbf{Right}$.
\end{lemma}
\begin{proof}
The rank of an instance of this lemma is the rank of $f$.
\medskip
We note that $\mathbf{Dom}(f)$ and every element of $\mathbf{Dom}(f)$ has rank less than the rank of $f$.
We will let $x$ range over points in $\mathbf{Dom}(f)$.
For $x_1 \circ x_2^{-1}$ defined, the induction hypothesis for (Composables-Equivalent) implies $x_1 \simeq_{\mathbf{Dom}(f)} x_2$ and hence $f[x_1] = f[x_2]$.
This implies that the pair
\begin{eqnarray*}
((x_1 \circ x_2^{-1}),\;f[x_1]\circ f[x_2]^{-1}) & = & ((x_1 \circ x_2^{-1}),\;f[x_1]\circ f[x_1]^{-1}) \\
& = & ((x_1 \circ x_2^{-1}),\;\mathbf{Left}(f[x_1]))
\end{eqnarray*}
is a pair of $\mathbf{Left}(f)$. This implies that for each element $x_1 \circ x_2^{-1}$ of $\mathbf{Left}(\mathbf{Dom}(f))$ we have that this element is a first component
of some pair in $\mathbf{Left}(f)$. This implies that $\mathbf{Dom}(\mathbf{Left}(f)) = \mathbf{Left}(\mathbf{Dom}(f))$. By the induction hypothesis for (G1) we have that $\mathbf{Dom}(\mathbf{Left}(f))$ is a point type.
We now have that the pairs of $\mathbf{Left}(f)$ are all pairs of the form $((x_1 \circ x_2^{-1}),\;\mathbf{Left}(f[x_1]))$.
We must show that this set of pairs satisfies condition condition (F1).
Consider two elements $x_1 \circ x_2^{-1}$ and $x_3 \circ x_4^{-1}$ of $\mathbf{Left}(\mathbf{Dom}(f))$ with $(x_1 \circ x_2^{-1}) =_{\mathbf{Left}(\mathbf{Dom}(f))} (x_3 \circ x_4^{-1})$.
To show (F1) we must show that $f[x_1] = f[x_3]$.
By the definition of $\simeq$ we have that there exists $x_5,x_6 \in \mathbf{Dom}(f)$ with
$$(x_1 \circ x_2^{-1})\circ (x_5 \circ x_6^{-1})^{-1} \circ (x_3 \circ x_4^{-1})$$
defined. By the induction hypothesis for (Composables-Equivlent) we then have $x_1 \simeq x_3$ which by condition (F1) for $f$ implies that $f[x_1]= f[x_3]$.
Finally, we must show uniqueness. Consider a morphoid function $g$ with $\mathbf{Dom}(g) = \mathbf{Left}(\mathbf{Dom}(f))$
and with $g[\mathbf{Left}(x)] = \mathbf{Left}(f[x]) = \mathbf{Left}(f)[\mathbf{Left}(x)]$. To show that $f=g$ we must show
that $g[x_1 \circ x_2^{-1}] = f[x_1\circ x_2^{-1}]$. But (Composables-Equivalent) implies $x_1 \circ x_2^{-1} =_{\mathbf{Left}(\mathbf{Dom}(f))} x_1 \circ x_1^{-1} = \mathbf{Left}(x_1)$
which gives $g[x_1 \circ x_2^{-1}] = g[\mathbf{Left}(x_1)] = f[\mathbf{Left}(x_1)] = f[x_1\circ x_2^{-1}]$.
\end{proof}
\begin{corollary}[Funs-Composable]
\label{lem:fun2}
For two morphoid functions $f$ and $g$ we have that $f \circ g$ is defined if and only if $\mathbf{Dom}(f) \circ \mathbf{Dom}(g)$ is defined
and for $x \in \mathbf{Dom}(f)$ and $y \in \mathbf{Dom}(g)$ with $x \circ y$ defined we have that $f[x] \circ g[y]$ is defined.
\end{corollary}
\begin{proof} The rank of an instance of this lemma is the maximum rank of $f$ and $g$.
\medskip
We let $x$ range over elements of $\mathbf{Dom}(f)$ and $y$ range over elements of $\mathbf{Dom}(g)$.
First suppose that $f \circ g$ is defined. In this case we have $\mathbf{Right}(f) = \mathbf{Left}(g)$
and by the preceding lemma this implies that $\mathbf{Right}(\mathbf{Dom}(f)) = \mathbf{Left}(\mathbf{Dom}(g))$ which implies that $\mathbf{Dom}(f) \circ \mathbf{Dom}(g)$ is defined.
Furthermore, for $x \circ y$ defined we have $\mathbf{Right}(x) = \mathbf{Left}(y)$ which implies
$\mathbf{Right}(f)(\mathbf{Right}(x)) = \mathbf{Left}(g)(\mathbf{Left}(y))$ which by the preceding lemma implies that $\mathbf{Right}(f[x]) = \mathbf{Left}(g[y])$ and hence $f[x] \circ g[y]$ is defined.
Conversely suppose that $\mathbf{Dom}(f) \circ \mathbf{Dom}(g)$ is defined and for $x \circ y$ defined we have $f[x] \circ g[y]$ defined.
We must show that in this case $\mathbf{Right}(f) = \mathbf{Left}(g)$. We have that $\mathbf{Right}(\mathbf{Dom}(f)) = \mathbf{Left}(\mathbf{Dom}(g))$
which by the preceding lemma gives $\mathbf{Dom}(\mathbf{Right}(f)) = \mathbf{Dom}(\mathbf{Left}(g))$. Now consider $z \in \mathbf{Dom}(\mathbf{Right}(f)) = \mathbf{Dom}(\mathbf{Left}(g))$.
We must show that $\mathbf{Right}(f)(z) = \mathbf{Left}(g)[z]$. We have $z = x_1^{-1} \circ x_2 = y_1 \circ y_2^{-1}$. This gives $\mathbf{Right}(f)[z] = \mathbf{Right}(f)[\mathbf{Right}(x_2)]= \mathbf{Right}(f[x_2])$.
Similarly $\mathbf{Left}(g)[z] = \mathbf{Left}(g)[\mathbf{Left}(y_1)] = \mathbf{Left}(g[y_1])$. But we also have $\mathbf{Right}(x_2) = \mathbf{Left}(y_1)$ and hence $x_2 \circ y_1$ is defined
and hence $f[x_1] \circ f[y_1]$ is defined and hence $\mathbf{Right}(f[x_2]) = \mathbf{Left}(g[y_1])$. We now have $\mathbf{Right}(f)[z] = \mathbf{Left}(g)[z]$ which implies
that $\mathbf{Right}(f) = \mathbf{Left}(g)$.
\end{proof}
\begin{lemma}[(G2) for functions and (Fun-Composition)]
\label{lem:fun3}
For a morphoid functions $f$ and $g$ with $f \circ g$ defined
we have that $f \circ g$ is the unique morphoid function such that $\mathbf{Dom}(f \circ g) = \mathbf{Dom}(f) \circ \mathbf{Dom}(g)$
and for $x \in \mathbf{Dom}(f)$ and $y \in \mathbf{Dom}(g)$ with $x \circ y$ defined we have $(f \circ g)[x \circ y] = f[x] \circ g[y]$.
\end{lemma}
\begin{proof} The rank of an instance of this lemma is the maximum rank of $f$ and $g$.
\medskip
Let $x$ range over elements of $\mathbf{Dom}(f)$ and let $y$ range over elements of $\mathbf{Dom}(g)$. By definition we have that $f \circ g$
is the function consisting of the pairs of the form $((x \circ y),\;f[x] \circ g[y])$. By the preceding lemma we have that $f[x] \circ g[y]$
is defined for every $x$ and $y$ such that $x \circ y$ is defined and hence $\mathbf{Dom}(f \circ g) = \mathbf{Dom}(f) \circ \mathbf{Dom}(g)$. By (G2) we have that
$\mathbf{Dom}(f \circ g)$ is a point type. We must show that this set of pairs satisfies condition (F1). Consider $x_1 \circ y_1$ and $x_2 \circ y_2$
with $(x_1 \circ y_1) \simeq_{\mathbf{Dom}(f) \circ \mathbf{Dom}(g)} (x_2 \circ y_2)$.
By the induction hypothesis for ($\simeq$.B) we have that $x_1 \simeq_{\mathbf{Dom}(f)} x_2$ and $y_1 \simeq_{\mathbf{Dom}(g)} y_2$.
By condition (F1) on $f$ and $g$ we then have $f[x_1]=f[x_2]$ and $g[y_1]=g[y_2]$.
This implies that $f[x_1] \circ g[y_1] = f[x_2] \circ g[y_2]$ which establishes (F1).
Finally we must show uniqueness. It suffices to note that for any two point types $\sigma$ and $\tau$ with $\sigma \circ \tau$ defined,
and two functions $h_1$ and $h_2$ with $\mathbf{Dom}(h_1) = \mathbf{Dom}(h_2) = \sigma \circ \tau$, if for any $x \in \sigma$ and $y \in \tau$ with $x \circ y$ defined
we have $h_1[x \circ y] = h_2[x \circ y]$ then we have $h_1 = h_2$.
\end{proof}
\begin{lemma}[(G4) for functions] For two weak function $f$ and $g$ with $f \circ g$ defined we have $\mathbf{Left}(f \circ g) = \mathbf{Left}(f)$ and $\mathbf{Right}(f \circ g) = \mathbf{Right}(g)$.
\end{lemma}
\begin{proof}
By the preceding lemmas we have that $\mathbf{Left}(f)$ is the unique weak function such that $\mathbf{Dom}(\mathbf{Left}(f)) = \mathbf{Left}(\mathbf{Dom}(f))$ and
$\mathbf{Left}(f)[\mathbf{Left}(x)] = \mathbf{Left}(f[x])$.
But using the induction hypotheses and the preceding lemmas we have
$$\mathbf{Dom}(\mathbf{Left}(f \circ g)) = \mathbf{Left}(\mathbf{Dom}(f \circ g)) \mathbf{Left}(\mathbf{Dom}(f) \circ \mathbf{Dom}(g)) = \mathbf{Left}(\mathbf{Dom}(f))$$
and
\begin{eqnarray*}
\mathbf{Left}(f \circ g))[\mathbf{Left}(x)] & = & (\mathbf{Left}(f \circ g))[\mathbf{Left}(x \circ y)] \\
& = & \mathbf{Left}\left((f \circ g)[x \circ y]\right) \\
& = & \mathbf{Left}(f[x] \circ f[y]) \\
& = & \mathbf{Left}(f[x])
\end{eqnarray*}
which proves the result.
\end{proof}
The proofs of (G3), (G5), (G6) and (G7) for functions are similarly straightforward applications of the above lemmas and the induction hypotheses. Intuitively,
given lemmas~\ref{lem:fun1},~\ref{lem:fun2} and~\ref{lem:fun3} we have that functions act like pairs or tuples. For pairs the groupoid properties always follow directly
from the induction hypotheses.
\begin{lemma}[Composables-Equivalent]
\label{lem:definednesseq}
For any morphoid type $\sigma$ and for $\incat{x,y}{\sigma}$ with $x \circ y^{-1}$ defined or with $x^{-1} \circ y$ defined
we have $x \simeq_\sigma y$.
\end{lemma}
\begin{proof}
An instance of this lemma has rank equal to the rank of $\sigma$.
We consider the case of $x \circ y^{-1}$ defined. In this case $x \circ y^{-1} \circ y$ is also defined which gives $x \simeq_\sigma y$.
\end{proof}
\begin{lemma}[$\simeq$.A]
For any weak type $\sigma$ we have that $\simeq_\sigma$ is an equivalence relation on the elements of $\sigma$.
\end{lemma}
\begin{proof}
For any $\incat{x}{\sigma}$ we have that $x \circ x^{-1} \circ x$ is defined and hence $x \simeq_\sigma x$.
To show symmetry suppose $x \simeq_\sigma y$ with $x \circ z^{-1} \circ y$ defined.
In this case we have that $y \circ (x \circ z^{-1} \circ y)^{-1} \circ x$ is defined.
By condition (T1) we have $(x \circ z^{-1} \circ y) \in \sigma$ and hence $y \simeq_\sigma x$.
For transitivity suppose $x \simeq_\sigma y \simeq_\sigma z$. In this case there exist $s$ and $t$ in $\sigma$
that $x \circ s^{-1} \circ y \circ t^{-1} \circ z$ is defined. But in this case we have $x \circ (t \circ y^{-1} \circ s)^{-1} \circ z$
is defined. Condition (T1) implies $(s \circ y^{-1} \circ t) \in \sigma$ and the result follows.
\end{proof}
\begin{lemma}[$\simeq$.B Helper]
\label{lem:bijection}
For a weak morphoid type $\sigma$ and $x, y \in \sigma$ we have $x \simeq_\sigma y$ if and only if $\mathbf{Right}(x) \simeq_{\mathbf{Right}(\sigma)} \mathbf{Right}(y)$ and similarly for $\mathbf{Left}$.
\end{lemma}
\begin{proof}
An instance of this lemma has rank equal to the rank of $\sigma$.
\medskip
We show the case for $\mathbf{Right}$. First suppose $x \simeq_\sigma y$. In that case there exists $z \in \sigma$ with $x \circ z^{-1} \circ y$ defined.
But in this case we have that $\mathbf{Right}(x) \circ \mathbf{Right}(z)^{-1} \circ \mathbf{Right}(y) = x^{-1} \circ x \circ z^{-1} \circ z \circ y^{-1} \circ y$ is defined
and hence $\mathbf{Right}(x) \simeq_{\mathbf{Right}(\sigma)} \mathbf{Right}(y)$. Now suppose that $\mathbf{Right}(x) \simeq_{\mathbf{Right}(\sigma)} \mathbf{Right}(y)$.
In that case we have that there exist $z_1,z_2 \in \sigma$ with $x^{-1} \circ x \circ z_1^{-1} \circ z_2 \circ y^{-1} \circ y$ defined
and the result follows from ($\simeq$.A) and (Composables-Equivalent).
\end{proof}
\begin{lemma}[$\simeq$.B]
\label{lem:bijection2}
For morphoid types $\sigma$ and $\tau$ with $\sigma \circ \tau$ defined and for $\incat{x_1,x_2}{\sigma}$ and $\incat{y_1,y_2}{\tau}$
with $x_1 \circ y_1$ and $x_2 \circ y_2$ defined, we have $(x_1 \circ y_1) \simeq_{\sigma \circ \tau} (x_2 \circ y_2)$ if and only if $x_1 \simeq_\sigma x_2$ if and only if $y_1 \simeq_\tau y_2$.
\end{lemma}
\begin{proof} A instance of this lemma has rank equal to the maximum rank of $\sigma$ and $\tau$.
\medskip
We will let $x$ range over instances of $\sigma$ and $y$ range over instances of $\tau$.
We first show that $(x_1 \circ y_1) \simeq_{\sigma \circ \tau} (x_2 \circ y_2)$ implies $x_1 \simeq_\sigma x_2$.
If $(x_1 \circ y_1) \simeq_{\sigma \circ \tau} (x_2 \circ y_2)$ then by definition there exists
$x_3$ and $y_3$ with $(x_1 \circ y_1) \circ (x_3 \circ y_3)^{-1} \circ (x_2\circ y_2)$ defined.
We also have
\begin{eqnarray*}
x_1 \circ y_1 \circ (x_3 \circ y_3)^{-1} \circ x_2 & = & x_1 \circ (y_1 \circ y_3^{-1}) \circ x_3^{-1} \circ x_2
\end{eqnarray*}
Since we have $\mathbf{Right}(\sigma) = \mathbf{Left}(\tau)$ we have that $y_1 \circ y_3^{-1}$ can be written as $x_4^{-1} \circ x_5$
and we get
\begin{eqnarray*}
x_1 \circ y_1 \circ (x_3 \circ y_3)^{-1} \circ x_2 & = & x_1 \circ (x_4^{-1} \circ x_5) \circ x_3^{-1} \circ x_2 \\
& = & x_1 \circ (x_4 \circ x_5^{-1} \circ x_3)^{-1} \circ x_2 \\
& = & x_1 \circ x_6^{-1} \circ x_2 \\
\end{eqnarray*}
This gives $x_1 \simeq_\sigma x_2$.
We now show that $x_1 \simeq_\sigma x_2$ implies $y_1 \simeq_\tau y_2$. By ($\simeq$.B Helper) we have that $x_1 \simeq_\sigma x_2$
if and only if $\mathbf{Right}(x_1) \simeq_{\mathbf{Right}(\sigma)} \mathbf{Right}(x_2)$. But $\mathbf{Right}(x_1) = \mathbf{Left}(y_1)$, $\mathbf{Right}(x_2) = \mathbf{Left}(y_2)$ and $\mathbf{Right}(\sigma)$
equals $\mathbf{Left}(\tau)$. So we have that $x_1 \simeq_\sigma x_2$ if and only if $\mathbf{Left}(y_1) \simeq_{\mathbf{Left}(\tau)} \mathbf{Left}(y_2)$ if and only if $y_1 \simeq_\tau y_2$.
Finally we show that if $x_1 \simeq_\sigma x_2$ and $y_1 \simeq_\tau y_2$ then $x_1 \circ y_1 \simeq_{\sigma \circ \tau} x_2 \circ y_2$.
By definition there exists $x_3$ and $y_3$ with $x_1\circ x_3^{-1} x_2$ and $y_2 \circ y_3^{-1} \circ y_1$ defined.
This implies that $\mathbf{Right}(x_3) = \mathbf{Right}(x_1) = \mathbf{Left}(y_1) = \mathbf{Left}(y_3)$ and we have that $x_3 \circ y_3$ is defined.
We then have that $(x_1 \circ y_1) \circ (x_3 \circ y_3)^{-1} \circ (x_2 \circ y_2)$ is defined and hence $x_1 \circ y_1 \simeq_{\sigma \circ \tau} x_2 \circ y_2$.
\end{proof}
\begin{lemma}[Partner] For two pre-types $\sigma$ and $\tau$ such that $\sigma \circ \tau$ is defined we have that for all $x \in \sigma$ there exists
$y \in \tau$ with $x \circ y$ defined and for every $y \in \tau$ there exists $x \in \sigma$ with $x \circ y$ defined.
\end{lemma}
\begin{proof}
The rank of an instance of this lemma is the maximum rank of $\sigma$ and $\tau$.
\medskip
Consider $x \in \sigma$. By the induction hypotheses for the rank preservation and the groupoid properties we have $x^{-1} \circ x \in \mathbf{Right}(\sigma)$.
Since $\mathbf{Right}(\sigma) = \mathbf{Left}(\tau)$ we have $x^{-1} \circ x = y_1 \circ y_2^{-1}$ for some $y_1,y_2 \in \tau$. But by the induction hypothesis for the groupoid properties
and rank preservation this implies that $x \circ (y_1 \circ y_2^{-1})^{-1}$ is defined and hence $x \circ y_2$ is defined. The proof that every $y \in \tau$ has a partner in $\sigma$ is similar.
\end{proof}
\begin{lemma}[Rank Preservation]
For any morphoid $x$, we have $R(x^{-1}) = R(\mathbf{Left}(x)) = R(\mathbf{Right}(x)) = R(x)$ and if $x \circ y$ is defined then $R(x) = R(y) = R(x \circ y)$.
\end{lemma}
\begin{proof}
The rank of an instance of the form $R(x^{-1}) = R(\mathbf{Left}(x)) = R(\mathbf{Right}(x)) = R(x)$ is the rank of $x$. The rank of an instance of the implication that
if $x \circ y$ is defined then $R(x) = R(y) = R(x \circ y)$ is the maximum rank of $x$ and $y$.
By left-right duality we have $R(x^{-1}) = R(x)$.
We will show that $R(\mathbf{Left}(x)) = R(x)$, $R(\mathbf{Right}(x)) = R(x)$ then follows by left-right duality.
The result is immediate for Boolean values and points and follows straightforwardly from the induction hypothesis for pairs.
Now consider a weak type $\sigma$ and let $x$ range over the elements of $\sigma$.
We have that $\mathbf{Left}(\sigma)$ is the type containing the values of the form $x_1 \circ x_2^{-1}$.
By the induction hypothesis for the groupoid properties all values of the form $\mathbf{Left}(x) = x \circ x^{-1}$ are in $\mathbf{Left}(\sigma)$.
The induction hypothesis that $R(\mathbf{Left}(x)) = R(x)$ then implies that the rank of $\mathbf{Left}(\sigma)$ is at least as large as the rank of $\sigma$. Also, by the induction hypothesis
we have $R(x_1 \circ x_2^{-1}) = R(x)$ and hence the rank of $\mathbf{Left}(\sigma)$ is at most the rank of $\sigma$ which proves the result.
For a function $f$ the result follows from the induction hypothesis and the previously proved result that
$\mathbf{Left}(f)$ is the function whose domain is $\mathbf{Left}(\mathbf{Dom}(f))$ and such that $\mathbf{Left}(f)[\mathbf{Left}(x)] = \mathbf{Left}(f[x])$.
We now consider the instance that $x \circ y$ defined implies $R(x) = R(y) = R(x \circ y)$. The case of points and Booleans is immediate
and the case of pairs follows directly from the induction hypothesis. Now consider two types $\sigma$ and $\tau$ with $\sigma \circ \tau$
defined. The elements of this type are the values of the form $x \circ y$ for $x \in \sigma$ and $y \in \tau$.
By the induction hypothesis we have that any such value has the property that $R(x) = R(y) = R(x \circ y)$. By the previously proved property (Partner)
we have that for all $x \in \sigma$ there exists $y \in \tau$ with $x \circ y$ defined and every $y \in \tau$ similarly has a partner in $\sigma$.
This implies that the rank of $\sigma \circ \tau$ equals the rank of $\sigma$ and also the rank of $\tau$.
Now consider functions $f$ and $g$ with $f \circ g$
defined. For this case the result is implied by the previously proved fact that $\mathbf{Dom}(f \circ g) = \mathbf{Dom}(f) \circ \mathbf{Dom}(g)$ and
$(f\circ g)[x\circ y] = f[x] \circ g[y]$.
\end{proof}
This ends the simultaneous induction proof establishing the properties other than ($\sim$.A) and ($\sim$.B).
We now prove ($\sim$.A) and ($\sim$.B) from the earlier properties.
\begin{lemma}[$\sim$.A]
The relation $\sim$ is an equivalence relation on weak morphoids.
\end{lemma}
\begin{proof}
For any $\incat{x}{\sigma}$ we have that $x \circ x^{-1} \circ x$ is defined and hence $x \sim x$.
To show symmetry suxpose $x \sim y$ with $x \circ z \circ y$ defined.
In this case we have that $y \circ (y^{-1} \circ z^{-1} \circ x^{-1}) \circ x$ is defined which yields $y \sim x$
For transitivity suppose $x \simeq_\sigma y \simeq_\sigma z$. In this case there exist $s$ and $t$
with $x \circ s \circ y \circ t \circ z$ is defined which gives $x \sim z$.
\end{proof}
\begin{lemma}[$\sim$.B]
We have $x \sim x^{-1} \sim \mathbf{Left}(x) \sim \mathbf{Right}(x)$
and for $x \circ y$ defined we have $x \sim y \sim (x \circ y)$.
\end{lemma}
\begin{proof}
We not that $x \circ (x^{-1} \circ x)\circ x^{-1}$ is defined which yields $x \sim x^{-1}$.
Also we have $x \circ x^{-1} \circ (x \circ x^{-1})$ which yields $x \sim \mathbf{Left}(x)$.
Similarly we have $x \sim \mathbf{Right}(x)$. Finally consider $x$ and $y$ with $x \circ y$ defined.
We have that $x \circ x^{-1} \circ (x \circ y)$ is defined which yields $x \sim (x \circ y)$.
Similarly we have $(x \circ y) \sim y$.
\end{proof}
\subsection{Abstraction}
\begin{figure}[p]
\begin{framed}
{\small
{\bf Template.} A template is an expression generated by the nonterminal ${\cal T}$ of the following grammar.
\begin{eqnarray*}
{\cal T} & ::= & {\cal A} \;|\; \mathbf{TypeOf}({\cal A}) \;|\; \;|\; \mathbf{Pair}({\cal T}_1,\;{\cal T}_2) \\
{\cal A} & ::= & \mathrm{\bf Point} \;|\; \mathbf{TypeOf}(\mathrm{\bf Point}) \;|\; \mathrm{\bf Bool} \;|\; \mathrm{\bf Point} \rightarrow {\cal A} \;|\; \;\mathbf{Pair}({\cal A}_1,\;{\cal A}_2)
\end{eqnarray*}
{\bf Abstract Template.} A template is called abstract if it is generated by the nonterminal ${\cal A}$ in the above grammar.
\medskip
${\bf x@{\cal T}}$. For a morphoid $x$ and a template ${\cal T}$ we have that $x@{\cal T}$ is specified by the following rules
where the abstraction is undefined if no rule applies or if
the right hand side of the rule is itself undefined. For $\sigma@\mathbf{TypeOf}({\cal A})$ to be defined we need that $x@{\cal A}$ is defined for all $x \in \sigma$
and for $f@(\mathrm{\bf Point} \rightarrow {\cal A})$ to be defined we must have
$f[x]@{\cal A}$ defined for every $x \in \mathbf{Dom}(f)$.
$$x@\mathrm{\bf Point} = \left\{\begin{array}{ll}x & \mbox{for $x$ a point} \\ (\mathrm{\bf Point},\;(\mathbf{Left}(\mathbf{SubPoint}(x)),\;\mathbf{Right}(\mathbf{SubPoint}(x))))
& \mbox{otherwise} \end{array} \right.$$
\begin{eqnarray*}
\mathbf{SubPoint}((\tagg{BOOL},\;v)) & = & (\tagg{BOOL},\;v) \\
\mathbf{SubPoint}((\tagg{PAIR},\;(x,y))) & = & (\tagg{PAIR},\;(x@\mathrm{\bf Point},\;y@\mathrm{\bf Point})) \\
\mathbf{SubPoint}((\tagg{TYPE},\;s)) & = & (\tagg{TYPE},\;\{x@\mathrm{\bf Point},;x\in s\}) \\
\mathbf{SubPoint}((\tagg{FUN},\;s)) & = & (\tagg{FUN},\;\{(x,\;y@\mathrm{\bf Point}),\;(x,y) \in s) \\
\\
(\tagg{BOOL},\;v)@\mathrm{\bf Bool} & = & (\tagg{BOOL},\;v) \\
(\tagg{PAIR},\;(x,\;y))@\mathbf{Pair}({\cal A}_1,\;{\cal A}_2) & = & (\tagg{PAIR},\;(x@{\cal A}_1,\;y@{\cal A}_2)) \\
(\tagg{TYPE},\;s)@\mathbf{TypeOf}({\cal A}) & = & (\tagg{TYPE},\;\{x@{\cal A},;x\in s\}) \\
(\tagg{FUN},\;s)@(\mathrm{\bf Point} \rightarrow {\cal A}) & = & (\tagg{FUN},\;\{(x,\;y@{\cal A}),\;(x,y) \in s\})
\end{eqnarray*}
}
${\bf x \preceq y}$. For morphoids $x$ and $y$ we define $x \preceq y$ to mean that for any template ${\cal T}$ such that $y@{\cal T}$
is defined we have that $x@{\cal T}$ is also defined and $x@{\cal T} = y@{\cal T}$.
\medskip {\bf Minimal Template.} We say that ${\cal T}$ is a minimal template for morphoid $x$ if $x@{\cal T} \preceq x$, or equivalently,
if every abstraction of $x$ factors through ${\cal T}$.
\caption{{\bf Abstraction.} Each template ${\cal T}$ defines an abstraction operation mapping $x$ to $x@{\cal T}$.
A justification for this particular grammar of abstractions is discussed in the text.
The operation $\mathbf{SubPoint}$ is needed for property (Abs-Compression) in figure~\ref{fig:EasyProps}.
In particular we need that $(x@{\cal T})@\mathrm{\bf Point} = x@\mathrm{\bf Point}$.
Minimal templates are typically unique but are not unique in general. A simple example is that any template of the form $\typeof({\cal A})$ is a minimal template of the empty type.}
\label{fig:Abstraction}
\end{framed}
\end{figure}
\begin{figure}
\begin{framed}
{
\noindent {\small \bf (At-Point-Defined)} For any morphoid $x$ we have that $x@\mathrm{\bf Point}$ is defined.
\medskip
\noindent {\small \bf (Abs-Expansion)} If $x@{\cal T}$ is defined then $x@{\cal T}@{\cal T}$ is also defined.
\medskip
\noindent {\small \bf (Abs-Compression)} For $(x@{\cal T}_1)@{\cal T}_2$ defined we have $(x@{\cal T}_1)@{\cal T}_2 = x@{\cal T}_2$.
\medskip
\noindent {\small \bf (Abs-Alternation)} For $(x@{\cal T}_1)@{\cal T}_2$ defined and $(x@{\cal T}_2)@{\cal T}_1$ defined
we have $x@{\cal T}_1 = x@{\cal T}_2$.
\medskip {\small \bf (Abs-Rank-Preservation)} $R(x@{\cal T}) \leq R(x)$.
\medskip {\small \bf ($\preceq$.A)} The relation $\preceq$ is a preorder (reflexive and transitive).
\medskip {\small \bf ($\preceq$.B)} For $x@{\cal T}$ defined we have $x \preceq x@{\cal T}$.
\medskip
{\small \bf ($\sim$.C)} For morphoids $x$ and $y$ with $x \sim y$ we have that $x@{\cal T}$ is defined if and only if $y@{\cal T}$ is defined.
\medskip
{\small \bf ($\sim$.D)} For morphoids $x$ and $y$ with $x \sim y$ we have that $x@{\cal T} = x$ if and only if $y@{\cal T} = y$.
\medskip {\small \bf ($\sim$.E)} If $x \sim y$ then $x$ and $y$ have the same minimal templates.
}
\caption{{\bf Weak Morphoid Abstraction Properties.}}
\label{fig:EasyProps}
\end{framed}
\end{figure}
Figure~\ref{fig:Abstraction} defines templates and the abstraction operation.
We will have that for any (strong) morphoid type, as defined in figure~\ref{fig:Morphoids}, there exists an abstract template ${\cal A}_\sigma$, where the concept of an abstract template is
defined in figure~\ref{fig:Abstraction}, such that for all $x \in \sigma$ we have $x =_\sigma x@{\cal A}_\sigma$. The template ${\cal A}_\sigma$ will be called an interface template for $\sigma$.
For example any group $G$, such as a permutation group, or a group of
linear transformations on a vector space, can be abstracted to a group $G@{\cal A}_\mathbf{Group}$ where the
group elements of $G@{\cal A}_\mathbf{Group}$ are points. We can abbreviate $G@{\cal A}_{\mathbf{Group}}$ as $G@\mathbf{Group}$ and more generally for any (strong) morphoid type $\sigma$
and $x \in \sigma$ we have $x@\sigma \in \sigma$ and $x =_\sigma x@\sigma$. In general the abstraction from $x$ to $x@\sigma$ replaces all types occuring inside $x$ with point types.
This conversion of all types to point types is embodied in the definition of an abstract template in figure~\ref{fig:Abstraction}.
The definition of an abstract template in figure~\ref{fig:Abstraction} also embodies the fact that all morphoid functions have the property that their
domain type is a point type. The fundamental requirement is that for $x \in \sigma$ we have that the abstraction from $x$ to $x@\sigma$ replaces
all types occuring in $x$ with point types. This must include the domain types of functions.
This conversion of types to point types is ``forgetful'' and this forgetting of function domain types
is required for property (Abs-Distributes-Out) in figure~\ref{fig:MorphAbsProps}.
The property (Abs-Distributes-Out) is fundamental to the soundness theorems of morphoid type theory. It is not necessary to require that the domain type of every morphoid function
is a point type. However, the domain type of every morphoid function occurring in an object of the form $x@\sigma$ must be a point type and it is more convenient to simply require this of all
functions. Functions are introduced with the axiom of choice and the axiom of choice remains sound under this restriction. For a function $f$ on groups and a group $G$ we define
$f(G)$ to be $f[G@\mathrm{\bf Point}]$. We can get away with this because of property ($=$.C) in figure~\ref{fig:MorphAbsProps} which gives that
$G=_{\mathbf{Group}} H$ if and only if $G@\mathrm{\bf Point} =_{\mathbf{Group}@\typeof(\pointt)}\;H@\mathrm{\bf Point}$.
The above discussion motivates abstract templates but does not motivate more general templates of the form $\typeof({\cal A})$. These more general templates are needed
for abstract interpretation as defined in figure~\ref{fig:AbsEval}. Abstract interpretation plays an important role in the proof that all values are morphoids ---
that if $\semvalue{e}\rho$ is defined then $\semvalue{e}\rho$ is a morphoid.
Like the morphoid composition operation, the morphoid abstraction operation is partial --- for a weak morphoid $x$ and template ${\cal T}$ the abstraction
$x@{\cal T}$ may or may not be defined. For example, for a weak function $f$ the abstraction $f@\typeof(\pointt)$ is undefined --- functions cannot be abstracted to types.
Functions can only be abstracted to points or to functions. A similar observation applies to Booleans, pairs and types. For abstraction of a type to a type
or a function to a function to be defined, the abstraction must carry forward all elements of the type or all input-ouput pairs of the function. See figure~\ref{fig:Abstraction}.
Figure~\ref{fig:EasyProps} gives properties of abstraction that hold for all weak morphoids. Stronger properties hold over (strong) morphoids.
The properties that hold over weak morphoids are fairly straightforward. The property (At-Point-Defined) states that $x@\mathrm{\bf Point}$ is defined
for all weak morphoids $x$. This can be proved by a very straightforward induction on the morphoid rank of $x$. The property (Abs-Expansion)
states that if $x@{\cal T}$ is defined then $x@{\cal T}@{\cal T}$ is also defined. This can be proved by a straightforward structural induction on the
template ${\cal T}$. The proof of (Abs-Compression) is somewhat more involved and is given explicitly below.
Property (Abs-Alternation) states that if $x@{\cal T}_1@{\cal T}_2$ is defined and $x@{\cal T}_2@{\cal T}_1$ is defined
then $x@{\cal T}_1 = x@{\cal T}_2$. This can be proved by a straightforward structural induction on ${\cal T}_1$.
The property (Abs-Rank-Preservation) states that $R(x@{\cal T}) \leq R(x)$. This can be proved by a straightforward structural induction on ${\cal T}$.
Property ($\sim$.C) states that if $x \sim y$ then $x@{\cal T}$ is defined if and only if $y@{\cal T}$ is defined.
For this it suffices to prove that if $x \circ y$ is defined then $x@{\cal T}$ is defined if and only if $y@{\cal T}$ is defined.
This can be proved by a straightfoward structural induction on ${\cal T}$ where the case of types and functions uses property (Partner) in figure~\ref{fig:HardProps}.
The proofs of ($\sim$.D) and ($\sim$.E) are similar.
We now turn to the explicit proof of (Abs-Compression). We start with the following helper lemma.
\begin{lemma}[Abs-Compression Helper] For a weak morphoid $x$ and template ${\cal T}$ with $x@{\cal T}$ defined we have $x@{\cal T}@\mathrm{\bf Point} = x@\mathrm{\bf Point}$.
\label{lem:A2star}
\end{lemma}
\begin{proof} The proof is by structural induction on ${\cal T}$. The result is immediate for ${\cal T} = \mathrm{\bf Bool}$.
For ${\cal T}= \mathrm{\bf Point}$
we have $x@\mathrm{\bf Point}$ is a point and we have $x@\mathrm{\bf Point}@\mathrm{\bf Point} = x@\mathrm{\bf Point}$.
For ${\cal T} \not = \mathrm{\bf Point}$ we have
\begin{eqnarray*}
x@{\cal T}@\mathrm{\bf Point} & = &(\mathrm{\bf Point},\;(\mathbf{Left}(\mathbf{SubPoint}(x@{\cal T})),\;\mathbf{Right}(\mathbf{SubPoint}(x@{\cal T})))) \\
x@\mathrm{\bf Point} & = &(\mathrm{\bf Point},\;(\mathbf{Left}(\mathbf{SubPoint}(x)),\;\mathbf{Right}(\mathbf{SubPoint}(x)))) \\
\end{eqnarray*}
So for ${\cal T} \not = \mathrm{\bf Point}$ it suffices to show that $\mathbf{SubPoint}(x@{\cal T}) = \mathbf{SubPoint}(x)$.
For pairs we have the following.
\begin{eqnarray*}
& & \mathbf{SubPoint}(\mathbf{Pair}(x,y)@\mathbf{Pair}({\cal T}_1,{\cal T}_2)) \\
& = & \mathbf{Pair}(x@{\cal T}_1@\mathrm{\bf Point},\;y@{\cal T}_2@\mathrm{\bf Point}) \\
& = & \mathbf{Pair}(x@\mathrm{\bf Point},\;y@\mathrm{\bf Point}) \\
& = & \mathbf{SubPoint}(\mathbf{Pair}(x,y))
\end{eqnarray*}
For $f@(\mathrm{\bf Point} \rightarrow {\cal A})$ we have that $\mathbf{SubPoint}(f)$ consists of pairs of the form $(x,y@\mathrm{\bf Point})$ for $(x,y)$ a pair of $f$
and $\mathbf{SubPoint}(f@(\mathrm{\bf Point} \rightarrow {\cal A}))$ consists of the pairs $(x,y@{\cal A}@\mathrm{\bf Point})$ for $(x,y)$ a pair of $f$.
By the induction hypothesis these are the same set of pairs. The case of types is similar.
\end{proof}
The property (Abs-Compression) states that for $x@{\cal T}_1@{\cal T}_2$ defined we have $x@{\cal T}_1@{\cal T}_2 = x@{\cal T}_2$.
Given lemma~\ref{lem:A2star} to handle the base case, we can now prove (Abs-Compression) by a straightforward structural induction on ${\cal T}_2$.
\subsection{Morphoids}
\label{subsec:morphoids}
The class of (strong) morphoids is defined in figure~\ref{fig:Morphoids}. Morphoids are weak morphoids that hereditarily satisfy
the conditions that types have interface templates and functions have range templates. An interface template for a weak type $\sigma$
is an abstract template ${\cal A}$ such that for $x \in \sigma$ we have that $x@{\cal A}$ is defined and $x@{\cal A} \in \sigma$.
We have previously discussed the idea that the type $\mathbf{Group}$ has an interface template ${\cal A}_{\mathbf{Group}}$
such that for any group $G$ we have that $G@{\cal A}_{\mathbf{Group}}$ is a group whose group elements are points.
For the type $\mathbf{Group}$ the interface template is unique
and is $\mathbf{Pair}(\typeof(\pointt),\;\mathrm{\bf Point} \rightarrow (\mathrm{\bf Point} \rightarrow \mathrm{\bf Point})).$ In general, however, interface templates are not unique.
Any abstract template is vacuously an interface template of the empty type. This ambiguity at the empty type propagates to create
other examples of types with multiple interface templates. For example, consider a type of pairs where the second component of every pair is the empty type.
Typically, however, the interface template for a type is unique.
\begin{figure}[t]
\begin{framed}
{\small
{\bf Interface Template.} An interface template for a weak type $\sigma$ is an abstract template ${\cal A}$ such that for all $x \in \sigma$ we have
$x@{\cal A}$ is defined and $x@{\cal A} \in \sigma$.
\medskip {\bf Range Template.} A range template for weak function $f$ is an abstract template ${\cal A}$ such that for all $x \in \mathbf{Dom}(f)$
we have $f[x]@{\cal A} = f[x]$.
\medskip
{\bf Morphoid.} A morphoid is one of the following.
\begin{itemize}
\item A morphoid point or Boolean value.
\item A morphoid type --- a weak type $\sigma$ such that every member of $\sigma$ is a morphoid and
\begin{itemize}
\item[(T2)] there exists an interface template for $\sigma$.
\end{itemize}
\item A morphoid function --- a weak function $f$ such that for $x \in \mathbf{Dom}(f)$ we have that $f[x]$ is a morphoid and
\begin{itemize}
\item[(F2)] there exists a range template for $f$.
\end{itemize}
\item A pair $\mathbf{Pair}(x,y)$ where $x$ and $y$ are morphoids.
\end{itemize}
\medskip ${\bf x@\sigma}$. For a morphoid type $\sigma$ and for $x \in \sigma$ we define $x@\sigma$ to be $x@{\cal A}$ for any interface ${\cal A}$ for $\sigma$.
(Abs-Alternation) implies that this is independent of the choice of ${\cal A}$.
\medskip ${\bf x =_\sigma y}$. For a morphoid type $\sigma$ and for $x,y \in \sigma$ we define $x =_\sigma y$ to mean $x@\sigma \simeq_\sigma y@\sigma$.
}
\caption{{\bf Morphoids.} The morphoids are the weak morphoids which hereditarily satisfy the conditions (T2) for types and (F2) for functions.}
\label{fig:Morphoids}
\end{framed}
\end{figure}
\begin{figure}[t]
\begin{framed}
{\small
{\small \bf (Min-Template.A)} $\typeof({\cal A})$ is a minimal template for a morphoid type $\sigma$ if and only if ${\cal A}$ is an interface template for $\sigma$.
\medskip {\small \bf (Min-Template.B)} $\mathrm{\bf Point} \rightarrow {\cal A}$ is a minimal template for morphoid function $f$ if and only if ${\cal A}$
is a range template for $f$.
\medskip {\small \bf (Min-Template.C)} Every morphoid has a minimal template.
\medskip {\small \bf (Morphoid-Closure)} Morphoids are closed under the groupoid operations and abstraction.
\medskip
\noindent {\small \bf (Abs-Distributes-In)} For $(x \circ y)@{\cal T}$ defined we have $(x \circ y)@{\cal T} = (x@{\cal T}) \circ (y@{\cal T})$.
\medskip
\noindent {\small \bf (Abs-Distributes-Out)} For ${\cal A}$ and ${\cal B}$ abstract with $(x@{\cal A}@{\cal B}) \circ (y@{\cal A}@{\cal B})$ defined we have
$$(x@{\cal A}@{\cal B}) \circ (y@{\cal A}@{\cal B}) = ((x@{\cal A}) \circ (y@{\cal A}))@{\cal B}.$$
\medskip {\small \bf ($=$.A)} The relation $=_\sigma$ is an equivalence relation on the elements of $\sigma$.
\medskip {\small \bf ($=$.B)} We have that $x \simeq_\sigma y$ implies $x =_\sigma y$.
\medskip {\small \bf ($=$.C)} We have $x =_\sigma y$ if and only if $(x@\mathrm{\bf Point}) =_{\sigma@\mathbf{TypeOf}(\mathrm{\bf Point})} (y@\mathrm{\bf Point})$.
\medskip {\small \bf ($=$.D)} For $x \in \sigma$ we have $x =_\sigma x@\sigma$.
\medskip {\small \bf ($=$.V1)} For morphoid types $\sigma$ and $\tau$ and for $\incat{x_1,x_2}{\sigma}$ and $\incat{y_1,y_2}{\tau}$
with $\sigma \circ \tau$, $x_1 \circ y_1$ and $x_2 \circ y_2$ defined, we have $(x_1 \circ y_1) =_{\sigma\circ \tau} (x_2 \circ y_2)$ if and only if
$x_1 =_\sigma x_2$ if and only if $y_1 =_\tau y_2$.
\medskip
{\small \bf ($=$.V2)} For morphoid types $\sigma$ and $\tilde{\sigma}$ with $\sigma \preceq \tilde{\sigma}$ and $x_1,x_2 \in \sigma$ and $\tilde{x}_1,\tilde{x}_2 \in \tilde{\sigma}$ with
$x_1 \preceq \tilde{x}_1$ and $x_2 \preceq \tilde{x}_2$ we have $x_1 =_\sigma x_2$ if and only if $\tilde{x}_1 =_{\tilde{\sigma}} \tilde{x}_2$.
\medskip {\small ($\bf \mathrm{\bf type}_i$)} We have that $\convalue{\mathrm{\bf type}_i}$ is a morphoid type with interface template $\typeof(\pointt)$.
}
\caption{{\bf Morphoid Properties.} The restriction on (Abs-Distributes-Out) is needed. A counter example to unrestricted outward distribution is discussed in the text.}
\label{fig:MorphAbsProps}
\end{framed}
\end{figure}
Even when the interface template for $\sigma$ is not unique, for $x\in \sigma$ we can define $x@\sigma$ to be $x@{\cal A}$ where ${\cal A}$ is any interface template for $\sigma$.
This is well defined because for any two interface templates ${\cal A}$ and ${\cal B}$ and for $x \in \sigma$ we must have that $x@{\cal A}@{\cal B}$ and $x@{\cal B}@{\cal A}$
are both defined and by (Abs-Alternation) we then have that $x@{\cal A} = x@{\cal B}$.
Interface templates are central to defining isomorphism. More specifically, we have that $x =_\sigma y$ is defined to be $x@\sigma \simeq_\sigma y@\sigma$.
Property ($\simeq$.A) states that the relation $\simeq_\sigma$ is an equivalence relation on the elements of $\sigma$. This immediately implies property (=.A)
in figure~\ref{fig:MorphAbsProps} which states that $=_\sigma$ is also an equivalience relation on the elements of $\sigma$.
It is useful to consider groups. We have that $G =_{\mathbf{Group}} H$ is defined to mean that there exists an (abstract) group $F$ such that $(G@\mathbf{Group}) \circ F^{-1} \circ (H@\mathbf{Group})$
is defined. The domain of $F$ can be taken to be the point type $\updownarrow\!(\pi_1(G),\;\pi_1(H),\;f)$ for some appropriate bijection $f$ from $\pi_1(G)$ to $\pi_1(H)$.
A range template for a function $f$ is an abstract template ${\cal A}$ such that for all $x \in \mathbf{Dom}(f)$ we have that $f[x]@{\cal A} = f[x]$.
While requiring the existence of a range template for every function may seem like a severe restriction, it remains consistent the with axiom of choice where the
range template of a function $f$ in $\sigma \rightarrow \tau$ can be taken to the (or any) interface template for $\tau$. Range templates for functions
are important for the abstract interpretation defined in figure~\ref{fig:AbsEval}. As mentioned above, this abstract interpretation is needed for
proving that all values are morphoids.
\begin{lemma}[Min-Template.A]
$\typeof({\cal A})$ is a minimal template for a morphoid type $\sigma$ if and only if ${\cal A}$ is an interface template for $\sigma$.
\end{lemma}
\begin{proof}
First consider an interface template ${\cal A}$ for $\sigma$ and consider a template $\mathbf{TypeOf}({\cal B})$ with $\sigma@\mathbf{TypeOf}({\cal B})$
defined. We have $\sigma@\typeof({\cal A}) \subseteq \sigma$ which implies that $$\sigma@\typeof({\cal A})@\mathbf{TypeOf}({\cal B})$$ is defined and hence $\typeof({\cal A})$ is a minimal template for $\sigma$.
Now suppose that $\typeof({\cal A})$ is a minimal template for $\sigma$ and let ${\cal B}$ be an interface template for $\sigma$.
We then have that $\sigma@\mathbf{TypeOf}({\cal B})$ is defined. By the definition of a minimal template we then have that $$\sigma@\typeof({\cal A})@\mathbf{TypeOf}({\cal B})$$
is defined. This implies that for $x \in \sigma$ we have that $x@{\cal A}@{\cal B}$ is defined. But we also have that $\sigma@\mathbf{TypeOf}({\cal B}) \subseteq \sigma$
which implies that $\sigma@\mathbf{TypeOf}({\cal B})@\typeof({\cal A})$ is defined. This implies that for $x \in \sigma$ we have that $x@{\cal B}@{\cal A}$ is defined.
But by (Abs-Alternation) we then have that $x@{\cal A} = x@{\cal B}$ and therefore ${\cal A}$ is an interface template for $\sigma$.
\end{proof}
\begin{lemma}[Min-Template.B]
$\mathrm{\bf Point} \rightarrow {\cal A}$ is a minimal template for morphoid function $f$ if and only if ${\cal A}$
is a range template for $f$.
\end{lemma}
\begin{proof}
It is easy to check that for a range template ${\cal A}$ for $f$ we have that $f@(\mathrm{\bf Point} \rightarrow {\cal A}) = f$ which implies that
$(\mathrm{\bf Point} \rightarrow {\cal A}$ is a minimal template for $f$. Conversely suppose that $\mathrm{\bf Point} \rightarrow {\cal A}$ is a minimal template for $f$
and let ${\cal B}$ be a range template for $f$. We then get that for $x \in \mathbf{Dom}(f)$ we have that $f[x]@{\cal B} = f[x]@{\cal A}@{\cal B}$ is defined
and also $f[x]@(\mathrm{\bf Point} \rightarrow {\cal B})@(\mathrm{\bf Point} \rightarrow {\cal A})$ is defined and hence $f[x]@{\cal B}@{\cal A}$ is defined. (Abs-Alternation)
then gives that $f[x]@{\cal A} = f[x]@{\cal B}$ which gives that ${\cal A}$ is a range template for $f$.
\end{proof}
\begin{lemma}[Min-Template.C]
Every morphoid has a minimal template.
\end{lemma}
\begin{proof}
The proof is by induction on morphoid rank. The result is immediate for Booleans and points and follows straightforwardly form the induction hypothesis for pairs.
The cases for types and functions are implied by (Min-Template.A) and (Min-Template.B) respectively.
\end{proof}
We now prove (Abs-closure), (Abs-Distributes-In) and (Abs-Distributes-Out)
by a simultaneous induction on morphoid rank of the same style as the simultaneous induction in section~\ref{subsec:weak}. Each instance of these properties is
assigned a rank and we prove that all instances of rank $\beta$ hold assuming all instances of rank less than $\beta$ hold.
We start by proving closure under the groupoid operations --- that groupoid properties (G1) and (G2) hold over (strong) morphoids.
As in section~\ref{subsec:weak}, (G1) and (G2) are immediate for points and Booleans and follow immediately from the induction hypothesis for
pairs. We explicitly prove (G1) and (G2) only for types and functions.
\begin{lemma}[G1 for Types]
For a mophoid type $\sigma$ we have that $\sigma^{-1}$, $\mathbf{Left}(\sigma)$ and $\mathbf{Right}(\sigma)$ are also morphoid types.
\end{lemma}
\begin{proof}
The morphoid rank of an instance of this lemma is the rank of $\sigma$.
\medskip
\noindent The duality of left and right implies the result for $\sigma^{-1}$.
We will show that $\mathbf{Left}(\sigma)$ is a morphoid type --- the case for $\mathbf{Right}(\sigma)$ is similar.
By definition we have that every member of $\sigma$ is a morphoid and $\sigma$ is a weak morphoid. We have already
proved that $\mathbf{Left}(\sigma)$ is a weak morphoid. Every element
of $\mathbf{Left}(\sigma)$ is of the form $x_1 \circ x_2^{-1}$ for $x_1,x_2 \in \sigma$ and by
the induction hypothesis for (G2) we also have that every such element is a morphoid.
It remains only to prove that $\mathbf{Left}(\sigma)$ has an interface template.
We will show that any interface template ${\cal A}$ for $\sigma$ is also an interface template for $\mathbf{Left}(\sigma)$.
The values of $\mathbf{Left}(\sigma)$ are the values of the form $x_1 \circ x_2^{-1}$ for $x_1,x_2 \in \sigma$.
By property ($\sim$.C) in figure~\ref{fig:EasyProps} we have that $(x_1 \circ x_2^{-1})@{\cal A}$ is defined and by the
induction hypothesis for (Abs-Distributes-In) we then have $(x_1 \circ x_2^{-1})@{\cal A} = (x_1@{\cal A}) \circ (x_2@{\cal A})^{-1} \in \mathbf{Left}(\sigma)$.
\end{proof}
\begin{lemma}[G2 for Types]
For two morphoid types $\sigma$ and $\tau$ with $\sigma \circ \tau$ defined we have that $\sigma \circ \tau$ is a morphoid type.
\end{lemma}
\begin{proof}
The rank of an instance of this lemma is the rank of $\sigma$ which equals the rank of $\tau$.
\medskip
\noindent As in the previous lemma it suffices show that every element of $\sigma \circ \tau$ is a morphoid and that
$\sigma \circ \tau$ has an interface template. The elements of $\sigma \circ \tau$ are the values of the form
$x \circ y$ for $x \in \sigma$ and $y \in \tau$. By the induction hypothesis for (G2) we have that every such value is a morphoid.
Let ${\cal A}$ be an interface template for $\sigma$. We will show that ${\cal A}$ is also an interface template for $\sigma \circ \tau$.
For $x \circ y$ in $\sigma \circ \tau$ property ($\sim$.C) implies that $(x \circ y)@{\cal A}$
is defined and by the induction hypothesis for (Abs-Distributes-In) we have $(x \circ y)@{\cal A} = (x@{\cal A}) \circ (y@{\cal A}) \in \sigma \circ \tau$.
\end{proof}
\begin{lemma}[(G1) for functions]
For any morphoid function $f$ we have that $f^{-1}$, $\mathbf{Left}(f)$ and $\mathbf{Right}(f)$ are morphoid functions.
\end{lemma}
\begin{proof}
An instance of this lemma has rank equal to the rank of $f$.
\medskip
We consider $\mathbf{Left}(f)$. The range values of this function have the form $\mathbf{Left}(f[x])$ for $x \in \mathbf{Dom}(f)$.
We must show that every range value is a morphoid and that there exists a range template. The induction hypothesis for (G1)
implies that $\mathbf{Left}(f[x])$ is a morphoid. Let ${\cal A}$ be a range template for $f$.
By property ($\sim$.C) implies that $\mathbf{Left}(f[x])@{\cal A}$ is defined and the induction hypothesis for (Abs-Distributes-In)
implies that $\mathbf{Left}(f[x])@{\cal A} = \mathbf{Left}(f[x]@{\cal A}) = \mathbf{Left}(f[x])$.
\end{proof}
\begin{lemma}[(G2) for functions]
For morphoid functions $f$ and $g$ with $f \circ g$ defined we have that $f \circ g$ is a morphoid.
\end{lemma}
\begin{proof}
An instance of this lemma has rank equal to the rank of $f$ which equals the rank of $g$.
\medskip
The range values of $f \circ g$ are all of the form $f[x] \circ g[y]$. By the induction hypothesis for (G2) these are all morphoids.
To show that $f \circ g$ has a range template let ${\cal A}$ be a range template for $f$. By ($\sim$.C) we have that $f[x] \circ g[y]$
is defined and by the induction hypothesis for this function have the form $\mathbf{Left}(f[x])$ for $x \in \mathbf{Dom}(f)$.
We must show that every range value is a morphoid and that there exists a range template. The induction hypothesis for (G1)
implies that $\mathbf{Left}(f[x])$ is a morphoid. Let ${\cal A}$ be a range template for $f$.
By property ($\sim$.C) implies that $\mathbf{Left}(f[x])@{\cal A}$ is defined and the induction hypothesis for (Abs-Distributes-In)
and property ($\sim$.D) we have $(f[x] \circ g[y])@{\cal A} = (f[x]@{\cal A}) \circ (g[y]@{\cal A}) = f[x] \circ g[y]$.
\end{proof}
\begin{lemma}[Abstraction-Closure] For a morphoid value $x$ and template ${\cal T}$ with $x@{\cal T}$ defined we have the $x@{\cal T}$ is a morphoid.
\end{lemma}
\begin{proof} The rank of an instance of this lemma is the rank of $x$.
\medskip
The result is immediate for $x@\mathrm{\bf Point}$ or $x@\mathrm{\bf Bool}$ and follows from the induction hypothesis for $x@\mathbf{Pair}({\cal T}_1,{\cal T}_2)$.
For functions we have that the range elements of $f@(\mathrm{\bf Point} \rightarrow {\cal A})$ are all values of the form $f[x]@{\cal A}$.
By the induction hypothesis we have that $f[x]@{\cal A}$ is a morphoid. (Abs-Compression) implies that $f[x]@{\cal A}@{\cal A} = f[x]@{\cal A}$ and hence ${\cal A}$
is a range template for $f@(\mathrm{\bf Point} \rightarrow {\cal A})$.
For a morphoid type $\sigma$ we have that $\sigma@\typeof({\cal A})$ is the set of values of the form $x@{\cal A}$. By the induction hypothesis all such values are morphoids
and by (Abs-Compression) we have $x@{\cal A}@{\cal A} = x@{\cal A}$ and we have that ${\cal A}$ is an interface template for $\sigma@\typeof({\cal A})$. For types we must also show that
$\sigma@\typeof({\cal A})$ satisfies condition (T1). We must show that for $x,y,z \in \sigma$ with $(x@{\cal A}) \circ (y@{\cal A})^{-1} \circ (z@{\cal A})$ defined we have
$(x@{\cal A}) \circ (y@{\cal A})^{-1} \circ (z@{\cal A}) \in \sigma@\typeof({\cal A})$. We are given that $\sigma$ has an interface template ${\cal B}$.
Since $x@{\cal B} \in \sigma$ we must have $x@{\cal B}@{\cal A}$ is defined abd similarly for $y$ and $z$. The induction hypothesis for (Abs-Distributes-Out) then gives
\begin{eqnarray*}
(x@{\cal A}) \circ (y@{\cal A})^{-1} \circ (z@{\cal A}) & = & (x@{\cal B}@{\cal A}) \circ (y@{\cal B}@{\cal A})^{-1} \circ (z@{\cal B}@{\cal A}) \\
& = & ((x@{\cal B}) \circ (y@{\cal B})^{-1} \circ (z@{\cal B}))@{\cal A} \\
& \in & \sigma@\typeof({\cal A})
\end{eqnarray*}
\end{proof}
\begin{lemma}[Abs-Distributes-In Helper] If $x \circ y$ is defined for $x$ and $y$ other than points
then $\mathbf{SubPoint}(x\circ y) = \mathbf{SubPoint}(x) \circ \mathbf{SubPoint}(y)$.
\label{lem:A5star}
\end{lemma}
\begin{proof} The rank of an instance of this lemma is the rank of $x$ which equals the rank of $y$.
\medskip
If $x$ and $y$ are Boolean values or points then $\mathbf{SubPoint}(x) = x$ and $\mathbf{SubPoint}(y) = y$ and the result is immediate.
For pairs the induction hypothesis for (Abs-Distributes-In) gives
\begin{eqnarray*}
& & \mathbf{SubPoint}(\mathbf{Pair}(x,y) \circ \mathbf{Pair}(x',y')) \\
& = & \mathbf{SubPoint}(\mathbf{Pair}(x\circ x',y\circ y')) \\
& = & \mathbf{Pair}((x\circ x')@\mathrm{\bf Point},\;(y \circ y')@\mathrm{\bf Point}) \\
& = & \mathbf{Pair}((x@\mathrm{\bf Point}) \circ (x'@\mathrm{\bf Point}),\;(y@\mathrm{\bf Point}) \circ (y'@\mathrm{\bf Point})) \\
& = & \mathbf{Pair}(x@\mathrm{\bf Point},\;y@\mathrm{\bf Point}) \circ \mathbf{Pair}(x'@\mathrm{\bf Point},\;y'@\mathrm{\bf Point}) \\
& = & \mathbf{SubPoint}(\mathbf{Pair}(x,y)) \circ \mathbf{SubPoint}(\mathbf{Pair}(x',y'))
\end{eqnarray*}
Now consider two function $f$ and $g$ with $f \circ g$ defined. Let $x$ range over the points in $\mathbf{Dom}(f)$ and $y$ range over the points in $\mathbf{Dom}(g)$.
We have that $\mathbf{SubPoint}(f \circ g)$ is the function whose domain
is $\mathbf{Dom}(f) \circ \mathbf{Dom}(g)$ and such that $\mathbf{SubPoint}(f \circ g)[x \circ y] = (f[x] \circ g[y])@\mathrm{\bf Point}$. But by the induction hypothesis for (Abs-Distributes-in)
we have that this is the same as the function mapping $x \circ y$ to $(f[x]@\mathrm{\bf Point}) \circ (g[y]@\mathrm{\bf Point})$.
This gives that $\mathbf{SubPoint}(f \circ g)$ is the same function as $\mathbf{SubPoint}(f) \circ \mathbf{SubPoint}(g)$.
Finally, consider two morphoid types $\sigma$ and $\tau$ with $\sigma \circ \tau$ defined.
We will let $x$ range over elements of $\sigma$ and $y$ range over elements of $\tau$.
We first show that if $\mathbf{SubPoint}(\sigma) \circ \mathbf{SubPoint}(\tau)$ is also defined then
$\mathbf{SubPoint}(\sigma \circ \tau) = \mathbf{SubPoint}(\sigma) \circ \mathbf{SubPoint}(\tau)$. For this we must show that these two point types
contain the same points. The induction hypothesis for (Abs-Distributes-In) gives $(x \circ y)@\mathrm{\bf Point} =(x@\mathrm{\bf Point}) \circ (y@\mathrm{\bf Point}) \in \mathbf{SubPoint}(\sigma) \circ \mathbf{SubPoint}(\tau)$
which establishes $\mathbf{SubPoint}(\sigma \circ \tau) \subseteq \mathbf{SubPoint}(\sigma) \circ \mathbf{SubPoint}(\tau)$.
Conversely, consider $(x@\mathrm{\bf Point}) \circ (y@\mathrm{\bf Point}) \in \mathbf{SubPoint}(\sigma) \circ \mathbf{SubPoint}(\tau)$.
Let ${\cal B}$ be an interface template for $\sigma$. By (Min-Template.A) we have that $\mathbf{TypeOf}({\cal B})$ is a minimal template for $\sigma$.
By ($\sim$.E) we have that $\mathbf{TypeOf}({\cal B})$ is also a minimal template for $\tau$ and by (Min-Template.A) we have that ${\cal B}$ is also an interface template for $\tau$.
By (Abs-Compression) we have that $x@{\cal B}@\mathrm{\bf Point} = x@\mathrm{\bf Point}$ and $y@{\cal B}@\mathrm{\bf Point} = y@\mathrm{\bf Point}$.
The induction hypotheses for (Abs-Distributes-Out) then gives
\begin{eqnarray*}
(x@\mathrm{\bf Point}) \circ (y@\mathrm{\bf Point}) & = & (x@{\cal B}@\mathrm{\bf Point}) \circ (y@{\cal B}@\mathrm{\bf Point}) \\
& = & ((x@{\cal B}) \circ (y@{\cal B}))@\mathrm{\bf Point} \\
& \in & \mathbf{SubPoint}(\sigma \circ \tau)
\end{eqnarray*}
Next we show that for any type $\sigma$ we have $\mathbf{Left}(\mathbf{SubPoint}(\sigma)) = \mathbf{SubPoint}(\mathbf{Left}(\sigma))$ and similarly for right.
For this we note that $\mathbf{SubPoint}(\sigma) \circ \mathbf{SubPoint}(\sigma^{-1})$ is defined and hence
$\mathbf{SubPoint}(\sigma \circ \sigma^{-1}) = \mathbf{SubPoint}(\sigma) \circ \mathbf{SubPoint}(\sigma)^{-1}$. Finally, we must show that if $\sigma \circ \tau$
is defined then $\mathbf{SubPoint}(\sigma)\circ \mathbf{SubPoint}(\tau)$ is defined. For this we note that
$\mathbf{Right}(\mathbf{SubPoint}(\sigma)) = \mathbf{SubPoint}(\mathbf{Right}(\sigma)) = \mathbf{SubPoint}(\mathbf{Left}(\tau)) = \mathbf{Left}(\mathbf{SubPoint}(\tau))$.
\end{proof}
\begin{lemma}[Abs-Distributes-In] If $(x \circ y)@{\cal T}$ is defined then $(x@{\cal T}) \circ (y@{\cal T})$ is also defined and
$$(x \circ y)@{\cal T} = (x@{\cal T}) \circ (y@{\cal T}).$$
\end{lemma}
\begin{proof} The rank of an instance of this lemma is the rank of $x$ which equals the rank of $y$.
\medskip
We first consider $(x \circ y)@\mathrm{\bf Point}$.
If $x$ and $y$ are points we have $x@\mathrm{\bf Point} = x$ and $y@\mathrm{\bf Point} = y$
and the result is immediate. If $x$ and $y$ are not points then lemma~\ref{lem:A5star} gives
\begin{eqnarray*}
(x \circ y)@\mathrm{\bf Point} & = & (\mathrm{\bf Point},\;(\mathbf{Left}(\mathbf{SubPoint}(x \circ y)),\mathbf{Right}(\mathbf{SubPoint}(x \circ y)))) \\
& = & \parens{\mathrm{\bf Point},\;\parens{\begin{array}{l}\mathbf{Left}(\mathbf{SubPoint}(x) \circ \mathbf{SubPoint}(y)), \\\mathbf{Right}(\mathbf{SubPoint}(x) \circ \mathbf{SubPoint}(y))\end{array}}} \\
& = & (\mathrm{\bf Point},\;(\mathbf{Left}(\mathbf{SubPoint}(x)),\mathbf{Right}(\mathbf{SubPoint}(y)))) \\
& = & (\mathrm{\bf Point},\;(\mathbf{Left}(\mathbf{SubPoint}(x)),\mathbf{Right}(\mathbf{SubPoint}(x)))) \\
& & ~~ \circ (\mathrm{\bf Point},\;(\mathbf{Left}(\mathbf{SubPoint}(y)),\mathbf{Right}(\mathbf{SubPoint}(y)))) \\
& = & (x@\mathrm{\bf Point}) \circ (y@\mathrm{\bf Point}).
\end{eqnarray*}
We now consider $(x \circ y)@{\cal T}$ for ${\cal T} \not = \mathrm{\bf Point}$.
In this case if $x$ and $y$ are Boolean we must have $x@{\cal T} = x$ and $y@{\cal T} = y$ and the result is immediate.
For pairs the induction hypothesis for (Abs-Distributes-In) gives
\begin{eqnarray*}
& & (\mathbf{Pair}(s,\;t) \circ \mathbf{Pair}(u,\;w))@\mathbf{Pair}({\cal T}_1,\;{\cal T}_2) \\
& = & \mathbf{Pair}((s \circ u)@{\cal T}_1,\;(t\circ w)@{\cal T}_2) \\
& = & \mathbf{Pair}((s@{\cal T}_1 \circ u@{\cal T}_1),\;(t@{\cal T}_2 \circ w@{\cal T}_2)) \\
& = & (\mathbf{Pair}(s,\;t)@\mathbf{Pair}({\cal T}_1,\;{\cal T}_2)) \circ (\mathbf{Pair}(u,\;w)@\mathbf{Pair}({\cal T}_1,\;{\cal T}_2)).
\end{eqnarray*}
Now consider morphoid functions $f$ and $g$ with $(f \circ g)@(\mathrm{\bf Point} \rightarrow {\cal A})$ defined.
Let $x$ range over points in $\mathbf{Dom}(f)$ and let $y$ range over points in $\mathbf{Dom}(g)$.
We will first show that $(f@(\mathrm{\bf Point} \rightarrow {\cal A})) \circ (g@(\mathrm{\bf Point} \rightarrow {\cal A}))$ is defined.
For this we use the criterion for definedness of function composition given in property (Funs-Composable).
First we note that $\mathbf{Dom}(f@(\mathrm{\bf Point} \rightarrow {\cal A})) = \mathbf{Dom}(f)$ and $\mathbf{Dom}(g@(\mathrm{\bf Point} \rightarrow {\cal A})) = \mathbf{Dom}(g)$
and since $f \circ g$ is defined we have $\mathbf{Dom}(f) \circ \mathbf{Dom}(g)$ is defined.
For $x \in \mathbf{Dom}(f)$ and $y \in \mathbf{Dom}(g)$ with $x \circ y$ defined we must show that
$f@(\mathrm{\bf Point} \rightarrow {\cal A})[x] \circ g@(\mathrm{\bf Point} \rightarrow {\cal A})[y]$ is defined. But
$f[x] \circ f[y]$ is defined and by ($\sim$.C) we have that $(f[x] \circ f[y])@{\cal A}$ is defined.
By the induction hypothesis for (Abs-Distributes-In) we then have that $f[x]@{\cal A} \circ g[y]@{\cal A}$ is defined.
To prove that the equality holds we note that $(f@(\mathrm{\bf Point} \rightarrow {\cal A})) \circ (g @ (\mathrm{\bf Point} \rightarrow {\cal A}))$
and $(f \circ g)@(\mathrm{\bf Point} \rightarrow {\cal A})$ have the same domain and both map $x \circ y$ to $(f[x] \circ g[y])@{\cal A} = f[x]@{\cal A} \circ g[y]@{\cal A}$.
For types it suffices to consider $(\sigma \circ \tau)@\mathbf{TypeOf}({\cal A})$.
We let $x$ range over members of $\sigma$ and $y$ range over members of $\tau$.
The members of $(\sigma \circ \tau)@\mathbf{TypeOf}({\cal A})$ have the form
$(x \circ y)@{\cal A}$. But by the induction hypothesis for (Abs-Distributes-In)
every such member can be written as $(x@{\cal A}) \circ (y@{\cal A})$ and we have
$(\sigma \circ \tau)@\mathbf{TypeOf}({\cal A}) \subseteq (\sigma@\mathbf{TypeOf}({\cal A})) \circ (\tau@\mathbf{TypeOf}({\cal A}))$.
Conversely, consider $(x@{\cal A}) \circ (y@{\cal A}) \in (\sigma@\mathbf{TypeOf}({\cal A})) \circ (\tau@\mathbf{TypeOf}({\cal A}))$.
Let ${\cal B}$ be an interface template for $\sigma$. By ($\sim$.E) and (Min-Template.A) we have that ${\cal B}$ is also an interface template for
$\tau$. We then have $x@{\cal B} \in \sigma$ and $y@{\cal B} \in \tau$ which implies that
$x@{\cal B}@{\cal A}$ and $y@{\cal B}@{\cal A}$ are defined.
By (Abs-Compression) we have that $x@{\cal B}@{\cal A} = x@{\cal A}$ and $y@{\cal B}@{\cal A} = y@{\cal A}$.
By the induction hypothesis for (Abs-Distributes-Out) we then have
\begin{eqnarray*}
(x@{\cal A}) \circ (y@{\cal A}) & = & (x@{\cal B}@{\cal A}) \circ (y@{\cal B}@{\cal A}) \\
& = & ((x@{\cal B}) \circ (y@{\cal B}))@{\cal A} \\
& \in & (\sigma \circ \tau)@\mathbf{TypeOf}({\cal A}).
\end{eqnarray*}
\end{proof}
\begin{lemma}[Abs-Distributes-Out Helper] For an abstract template ${\cal A}$ and for $(x@{\cal A}@\mathrm{\bf Point}) \circ (y@{\cal A}@\mathrm{\bf Point})$ defined we have
that $(x@{\cal A} \circ y@{\cal A})@\mathrm{\bf Point}$ is also defined with
$$(x@{\cal A}@\mathrm{\bf Point}) \circ (y@{\cal A}@\mathrm{\bf Point}) = (x@{\cal A} \circ y@{\cal A})@\mathrm{\bf Point}.$$
\label{lem:A4star}
\end{lemma}
\begin{proof} The rank of an instance of this lemma is the maximum rank of $x$ and $y$.
\medskip
Given (Abs-Distributes-In) and rank-preservation at the current rank,
it suffices to show that if $(x@{\cal A}@\mathrm{\bf Point}) \circ (y@{\cal A}@\mathrm{\bf Point})$ is defined then $(x@{\cal A}) \circ (y@{\cal A})$ is defined.
If ${\cal A}$ is $\mathrm{\bf Point}$ then the result follows from (Abs-Compression).
For ${\cal A} \not = \mathrm{\bf Point}$ we have that $(x@{\cal A}@\mathrm{\bf Point}) \circ (y@{\cal A}@\mathrm{\bf Point})$ is defined if and only if $\mathbf{SubPoint}(x@{\cal A}) \circ \mathbf{SubPoint}(y@{\cal A})$ is defined.
So it suffices to show that for ${\cal A} \not = \mathrm{\bf Point}$, if $\mathbf{SubPoint}(x@{\cal A}) \circ \mathbf{SubPoint}(y@{\cal A})$ is defind then $(x@{\cal A}) \circ (y@{\cal A})$ is defined.
\medskip
Assume that $\mathbf{SubPoint}(x@{\cal A}) \circ \mathbf{SubPoint}(y@{\cal A})$ is defind.
We show must that $(x@{\cal A}) \circ (y@{\cal A})$ is defined.
The result is straightforward for ${\cal A} = \mathrm{\bf Bool}$.
If ${\cal A} = \typeof(\pointt)$ then $x@{\cal A}$ and $y@{\cal A}$ are point types and we have $\mathbf{SubPoint}(x@{\cal A}) = x@{\cal A}$ and $\mathbf{SubPoint}(y@{\cal A}) = y@{\cal A}$
and the result follows.
For ${\cal A} = \mathbf{Pair}({\cal B},{\cal C})$ we have that $\mathbf{SubPoint}(\mathbf{Pair}(x@{\cal B},y@{\cal C})) \circ \mathbf{SubPoint}(\mathbf{Pair}(x'@{\cal B},y'@{\cal C}))$ is defined
and must show that $\mathbf{Pair}(x@{\cal B},y@{\cal C}) \circ \mathbf{Pair}(x'@{\cal B},y'@{\cal C})$ is defined. Noting that $\mathbf{SubPoint}(\mathbf{Pair}(z,w)) = \mathbf{Pair}(z@\mathrm{\bf Point},y@\mathrm{\bf Point})$
we have that $(x@{\cal B}@\mathrm{\bf Point}) \circ (x'@{\cal B}@\mathrm{\bf Point})$ and $(y@{\cal C}@\mathrm{\bf Point}) \circ (y'@{\cal C}@\mathrm{\bf Point})$ must be defined.
By the induction hypothesis for (Abs-Distributes-Out) we then get that $(x@{\cal B})\circ (x'@{\cal B})$ and $(y@{\cal C}) \circ (y'@{\cal C})$ are both defined
which implies the lemma.
Now suppose ${\cal A} = \mathrm{\bf Point} \rightarrow {\cal B}$ and consider functions $f,g$
with $\mathbf{SubPoint}(f@(\mathrm{\bf Point} \rightarrow {\cal B})) \circ \mathbf{SubPoint}(g@(\mathrm{\bf Point} \rightarrow {\cal B}))$ defined.
Let $x$ range over $\mathbf{Dom}(f)$ and $y$ range over $\mathbf{Dom}(g)$.
We have $\mathbf{Dom}(\mathbf{SubPoint}(f@(\mathrm{\bf Point} \rightarrow {\cal N}))) = \mathbf{Dom}(f)$ and $\mathbf{Dom}(\mathbf{SubPoint}(g@(\mathrm{\bf Point} \rightarrow {\cal B}))) = \mathbf{Dom}(g)$.
By (Funs-Composable) we then have that the definedness of
$$\mathbf{SubPoint}(f@(\mathrm{\bf Point} \rightarrow {\cal B})) \circ \mathbf{SubPoint}(g@(\mathrm{\bf Point} \rightarrow {\cal B}))$$
implies the definedness of $\mathbf{Dom}(f@(\mathrm{\bf Point} \rightarrow {\cal B})) \circ \mathbf{Dom}(g@(\mathrm{\bf Point} \rightarrow {\cal B}))$.
Furthermore, for $x \circ y \in \mathbf{Dom}(f) \circ \mathbf{Dom}(g)$ the definedness of $\mathbf{SubPoint}(f@(\mathrm{\bf Point} \rightarrow {\cal B})) \circ \mathbf{SubPoint}(g@(\mathrm{\bf Point} \rightarrow {\cal B}))$ implies
that $(f[x]@{\cal B}@\mathrm{\bf Point}) \circ (g[y]@{\cal B}@\mathrm{\bf Point})$ is defined. By the induction hypothesis for (Abs-Distributes-Out) we then have
that $(f[x]@{\cal B}) \circ (g[y]@{\cal B})$ is defined which implies the lemma.
\end{proof}
\begin{lemma}[Abs-Distributes-Out] For ${\cal A}$ and ${\cal B}$ abstract with $(x@{\cal A}@{\cal B}) \circ (y@{\cal A}@{\cal B})$ defined we have
that $((x@{\cal A}) \circ (y@{\cal A}))@{\cal B}$ is also defined and
$$(x@{\cal A}@{\cal B}) \circ (y@{\cal A}@{\cal B}) = ((x@{\cal A}) \circ (y@{\cal A}))@{\cal B}.$$
\end{lemma}
\begin{proof} The rank of an instance of this lemma is the maximum rank of $x$ and $y$.
\medskip
The case of ${\cal B} = \mathrm{\bf Point}$ is handled by lemma~\ref{lem:A4star} and we can assume ${\cal B} \not = \mathrm{\bf Point}$.
Note that this implies that ${\cal A} \not = \mathrm{\bf Point}$.
Given (Abs-Distributes-In) and rank preservation at the current rank, it suffices to show that under the conditions of the lemma we have that $(x@{\cal A}) \circ (y@{\cal A})$ is defined.
We will now do a case analysis on ${\cal A}$.
If ${\cal A} = \mathrm{\bf Bool}$ the result is straightforward.
For pairs the result follows straightforwardly from the induction hypothesis.
Now consider two functions $f,g$ with $(f@(\mathrm{\bf Point} \rightarrow {\cal C})@(\mathrm{\bf Point} \rightarrow {\cal D})) \circ (g@(\mathrm{\bf Point} \rightarrow {\cal C})@(\mathrm{\bf Point} \rightarrow{\cal D}))$ defined.
Let $x$ range over elements of $\mathbf{Dom}(f)$ and let $y$ range over elements of $\mathbf{Dom}(g)$.
By an argument similar to that used in the proof of lemma~\ref{lem:A4star}
we get that for $x \circ y$ in $\mathbf{Dom}(f) \circ \mathbf{Dom}(g)$ we have $(f[x]@{\cal C}@{\cal D}) \circ (g[y]@{\cal C}@{\cal D})$ is defined.
By the induction hypothesis we then get that $f[x]@{\cal C} \circ g[y]@{\cal C}$ is defined which implies the lemma.
For types we note that the only abstract template for types is $\typeof(\pointt)$.
For ${\cal A} = \typeof(\pointt)$ then $x@{\cal A}$ and $y@{\cal A}$ must be point types and ${\cal B}$ must be either $\typeof(\pointt)$ or $\mathbf{TypeOf}(\mathrm{\bf Point})$. In either case
$x@{\cal A}@{\cal B} = x@{\cal A}$ and $y@{\cal A}@{\cal B} = y@{\cal A}$ and the result follows.
\end{proof}
\noindent This completes the simultaneous induction proof and properties (Abs-Closure), (Abs-Distributes-In) and (Abs-Distributes-Out) are now fully established.
\medskip
It is worth noting a counter example to unrestricted outward distribution. At a hight level unrestricted outward distribution fails
because the mapping from a type $\tau$ to $\tau@\typeof(\pointt)$ is forgetful ---
for two elements $x,y \in \tau$ with $\mathbf{Left}(x) = \mathbf{Left}(y)$ and $\mathbf{Right}(x) = \mathbf{Right}(y)$ we get $x@\mathrm{\bf Point} = y@\mathrm{\bf Point}$ and the distinction between $x$ and $y$ is forgotten.
For a concrete counter example let $\sigma$ be a type containing only points
of the form $\mathrm{\bf Point}(i,i)$. Let $\tau_1$ and $\tau_2$ be types representing
different permutation groups on $\sigma$. More explicitly, each element of $\tau_1$ is a type $\gamma_f$ representing a bijection $f$ from $\sigma$ to $\sigma$ ---
the elements of $\gamma_f$ are points of the form $\mathrm{\bf Point}(i,j)$ where $\mathrm{\bf Point}(i,j) \in \gamma_f$ if and only if $f(\mathrm{\bf Point}(j,j)) = \mathrm{\bf Point}(i,i)$. The type $\tau_2$ is similar
but represents a different permutation group on $\sigma$. Since $\tau_1$ and $\tau_2$ represent permutation groups we have that $\tau_1$ and $\tau_2$ are closed under both composition
and inverse. For any type $\tau$ closed under both composition and inverse we have that $\tau$ contains identity elements of the form $x \circ x^{-1}$ and $x^{-1} \circ x$
and we get that $\mathbf{Left}(\tau) = \tau$ and similarly for right. We then have that $\mathbf{Right}(\tau_1) = \tau_1 \not = \tau_2 = \mathbf{Left}(\tau_2)$ and hence $\tau_1 \circ \tau_2$
is not defined. But now consider $(\tau_1@\typeof(\pointt))\circ(\tau_2@\typeof(\pointt))$. The elements of $\tau_i@\typeof(\pointt)$
are the points of the form $\mathrm{\bf Point}(\mathbf{Left}(\mathbf{SubPoint}(\gamma_f)),\mathbf{Right}(\mathbf{SubPoint}(\gamma_f)))$. Since $\gamma_f$ is a point type
we have $\mathbf{SubPoint}(\gamma_f) = \gamma_f$. We also have that $\mathbf{Left}(\gamma_f) = \gamma_f \circ \gamma_f^{-1} = \gamma _{f \circ f^{-1}} = \gamma_I = \sigma$
and similarly for $\mathbf{Right}(\gamma_f)$.
So $\tau_1@\typeof(\pointt)$ contains only the single point $\mathrm{\bf Point}(\sigma,\sigma)$ and the same holds for $\tau_2$.
We then have that $(\tau_1@\typeof(\pointt)) \circ (\tau_2@\typeof(\pointt))$ is defined but distribution-out fails.
\medskip \noindent We now prove the remaining properties in figure~\ref{fig:MorphAbsProps}.
\begin{lemma}[$=$.A]
The relation $=_\sigma$ is an equivalence relation on the elements of $\sigma$.
\end{lemma}
\begin{proof}
By definition we have that $x =_\sigma y$ if and only if $x@\sigma \simeq_\sigma y@\sigma$. The result then follows from ($\simeq$.A) which states that
$\simeq_\sigma$ is an equivalence relation on the elements of $\sigma$.
\end{proof}
\begin{lemma}[$=$.B]
For $x,y \in \sigma$ we have $x \simeq_\sigma y$ implies $x=_\sigma y$.
\end{lemma}
\begin{proof}
Assume $x \simeq_\sigma y$. In this case there exists $z \in \sigma$ such that $x \circ z^{-1} \circ y$ is defined. By (Abs-Distributes-In)
we then have that $(x@\sigma) \circ (z@\sigma)^{-1} \circ (y@\sigma)$ is defined which implies $x =_\sigma y$.
\end{proof}
\begin{lemma}[$=$.C]
\label{lem:AbsEquivalence}
For any morphoid type $\sigma$, and
for $x,y \in \sigma$, we have $x =_\sigma y$ if and only if $x@\mathrm{\bf Point} =_{\sigma@\typeof(\pointt)} y@\mathrm{\bf Point}$.
\end{lemma}
\begin{proof}
Let ${\cal A}$ be an interface template for $\sigma$.
First assume $x =_\sigma y$. In this case there exists $z \in \sigma$ such that $(x@{\cal A}) \circ z^{-1} \circ (y@{\cal A})$ is defined. By (Abs-Distributes-In)
we then have that $(x@{\cal A}@\mathrm{\bf Point}) \circ (z@\mathrm{\bf Point})^{-1} \circ (y@{\cal A}@\mathrm{\bf Point})$ is defined. By (Abs-Compression) we then have that
$(x@\mathrm{\bf Point}) \circ (z@\mathrm{\bf Point})^{-1} \circ (y@@\mathrm{\bf Point})$ is defined which implies $$x@\mathrm{\bf Point} =_{\sigma@\typeof(\pointt)} y@\mathrm{\bf Point}.$$
Conversely, suppose that $$x@\mathrm{\bf Point} =_{\sigma@\typeof(\pointt)} y@\mathrm{\bf Point}.$$ This implies that there exists $z \in \sigma$ such that
$(x@\mathrm{\bf Point}) \circ (z@\mathrm{\bf Point})^{-1} \circ (y@\mathrm{\bf Point})$ is defined. We then have that $x@{\cal A}$ is defined and by (Abs-Compression) we have $x@\mathrm{\bf Point} = x@{\cal A}@\mathrm{\bf Point}$ and similarly for $z$ and $y$.
This gives that $(x@{\cal A}@\mathrm{\bf Point}) \circ (z@{\cal A}@\mathrm{\bf Point})^{-1} \circ (y@{\cal A}@\mathrm{\bf Point})$ is defined
and by (Abs-distributes-Out) we have that $(x@{\cal A}) \circ (z@{\cal A})^{-1} \circ (y@{\cal A})$ is defined
which implies $x =_\sigma y$.
\end{proof}
\begin{lemma}[$=$.D]
For $x \in \sigma$ we have $x =_\sigma x@\sigma$.
\end{lemma}
\begin{proof}
Let ${\cal A}$ be an interface template for $\sigma$. It suffices to note that $(x@{\cal A}) \circ (x@{\cal A})^{-1} \circ ((x@{\cal A})@{\cal A})$
is defined.
\end{proof}
\begin{lemma}[$=$.V1]
\label{lem:Equality.V1}
For morphoid types $\sigma$ and $\tau$ and for $x_1,x_2 \in \sigma$ and $y_1,y_2 \in \tau$
with $\sigma \circ \tau$, $x_1 \circ y_1$ and $x_2 \circ y_2$ defined, we have $(x_1 \circ y_1) =_{\sigma \circ \tau} (x_2 \circ y_2)$ if and only if
$x_1 =_{\sigma} x_2$ if and only if $y_1 =_{\tau} y_2$.
\end{lemma}
\begin{proof} Let ${\cal A}$ be an interface template for $\sigma$.
By (Min-Template.A) and ($\sim$.E) we then have that ${\cal A}$ is also an interface template for $\tau$ and $\sigma \circ \tau$.
By (Abs-distributes-in) and ($\simeq$.B) we have
\begin{eqnarray*}
& & (x_1 \circ y_1) =_{\sigma \circ \tau} (x_2 \circ y_2) \\
& \mbox{iff} & (x_1 \circ y_1)@{\cal A} \simeq_{\sigma \circ \tau} (x_2 \circ y_2)@{\cal A} \\
& \mbox{iff} & ((x_1@{{\cal A}}) \circ (y_1@{\cal A})) \simeq_{\sigma \circ \tau} ((x_2@{\cal A}) \circ (y_2@{\cal A})) \\
& \mbox{iff} & (x_1@{\cal A}) \simeq_\sigma (x_2@{\cal A}) \\
& \mbox{iff} & x_1 =_\sigma x_2
\end{eqnarray*}
Similarly we get equivalence to $y_1 =_\tau y_2$.
\end{proof}
\begin{lemma}[$=$.V2]
For morphoid types $\sigma$ and $\tilde{\sigma}$ with $\sigma \preceq \tilde{\sigma}$ and $x_1,x_2 \in \sigma$ and $\tilde{x}_1,\tilde{x}_2 \in \tilde{\sigma}$ with
$x_1 \preceq \tilde{x}_1$ and $x_2 \preceq \tilde{x}_2$ we have $x_1 =_\sigma x_2$ if and only if $\tilde{x}_1 =_{\tilde{\sigma}} \tilde{x}_2$.
\end{lemma}
\begin{proof}
The definition of $\preceq$ gives
$x_1@\mathrm{\bf Point} = \tilde{x}_1@\mathrm{\bf Point}$ and
$x_2@\mathrm{\bf Point} = \tilde{x}_2@\mathrm{\bf Point}$ and $\sigma@\typeof(\pointt) = \tilde{\sigma}@\typeof(\pointt)$. Property ($=$.C) then gives
\begin{eqnarray*}
x_1 =_\sigma x_2 & \mbox{iff} & x_1@\mathrm{\bf Point} =_{\sigma@\typeof(\pointt)}\;x_2@\mathrm{\bf Point} \\
& \mbox{iff} & \tilde{x}_1@\mathrm{\bf Point} =_{\tilde{\sigma}@\typeof(\pointt)}\;\tilde{x}_2@\mathrm{\bf Point} \\
& \mbox{iff} & \tilde{x}_1 =_{\tilde{\sigma}} \tilde{x}_2
\end{eqnarray*}
\end{proof}
\begin{lemma}[($\mathrm{\bf type}_0$) Helper]
If $\sigma$ and $\tau$ are discrete morphoid types with $\sigma \circ \tau$ defined then $\sigma \circ \tau$ is also discrete.
\label{lem:discrete}
\end{lemma}
\begin{proof}
This follows immediately from ($=$.V1).
\end{proof}
\begin{lemma}[$\mathrm{\bf type}_0$]
${\cal V}\double{\mathbf{Set}}$ is a mophoid type.
\end{lemma}
\begin{proof}
${\cal V}\double{\mathbf{Set}}$ is the type containing all discrete morphoid types in the universe $V_{\kappa_0}$ where $\kappa_0$ is the smallest uncountable inaccessible cardinal.
By definition every member of ${\cal V}\double{\mathbf{Set}}$ is a morphoid. We must show condition (T1) and that $\convalue{\mathbf{Set}}$
has an interface template.
For condition (T1) it suffices to show that $\convalue{\mathbf{set}}$ is closed
under inverse and composition. Left-right duality implies closure under inverse. (($\mathrm{\bf type}_0$) Helper) or ($=$.V1) implies that $\convalue{\mathbf{Set}}$
is closed under composition. We also have that $\typeof(\pointt)$ is an interface template for $\convalue{\mathbf{Set}}$.
To see this we first note that (Abs-Closure) implies that for $\sigma \in \convalue{\mathbf{Set}}$ we have that $\sigma@\typeof(\pointt)$ is a morphoid type.
The properties of Grothendiek universes impliy that $\sigma@\typeof(\pointt) \in V_{\kappa_0}$.
Property ($=$.C) implies that $\sigma@\typeof(\pointt)$ is discrete.
\end{proof}
\begin{lemma}[$\mathrm{\bf type}_i$]
In general we have that $\convalue{\mathrm{\bf type}_i}$ is a morphoid type.
\end{lemma}
\begin{proof}
The proof is the same as that for $\convalue{\mathbf{Set}}$ except that for $i > 0$ we do not have to check for discreteness.
\end{proof}
Note that $\convalue{\mathbf{Set}} \in \convalue{\mathbf{Class}}$ and that $\convalue{\mathbf{Set}}$ is not discrete --- for any two sets $\sigma$ and $\tau$
we have that $\sigma =_{\mathbf{Set}} \tau$ if and only if $\sigma$ and $\tau$ have the same cardinality.
\subsection{Additional Definitions and Properties}
\begin{figure}[p]
\begin{framed}
{\bf Discrete Type.} A type $\sigma$ is called discrete if for all $x,y \in \sigma$ we have that $x =_\sigma y$ implies $x = y$.
\medskip ${\bf f(x)}$. For a morphoid function $f$ and morphoid $x$ we have that $f(x)$ is defined if $x@\mathrm{\bf Point} \in \mathbf{Dom}(f)$.
If $f(x)$ is defined then we have that $f(x) = f[x@\mathrm{\bf Point}]$.
\medskip ${\bf \sigma \rightarrow \tau}$. For morphoid types $\sigma$ and $\tau$ we define $\sigma \rightarrow \tau$
to be the set of morphoid functions $f$ with $\mathbf{Dom}(f) = \sigma@\typeof(\pointt)$ and such that for all $x \in \sigma$ we have $f(x) \in \tau$ and
$f(x)@\tau = f(x)$.
\medskip $\mathbf{The}(\sigma)$. We assume a fixed global choice function $\mathbf{The}$ such that for any morphoid type $\sigma$ with exactly one equivalence class
we have $\mathbf{The}(\sigma) \in \sigma$ and $\mathbf{The}(\sigma)@\sigma = \mathbf{The}(\sigma)$.
\medskip $\updownarrow\!(\sigma,\tau,f)$. We define $\pi_i$ on points by $\pi_1(\mathrm{\bf Point}(i,j)) = i$ and $\pi_2(\mathrm{\bf Point}(i,j)) = j$.
For morphoid types $\sigma$ and $\tau$ and bijection $f \in \sigma \rightarrow \tau$ we define $\updownarrow\!(\sigma,\tau,f)$
to be the type containing the points $\mathrm{\bf Point}(\pi_1(y@\mathrm{\bf Point}),\pi_2(x@\mathrm{\bf Point}))$ for $x \in \sigma$ and $y \in \tau$ with $y =_\tau f(x)$.
\medskip ${\bf \mathrm{\bf iso}(\sigma,x,y)}$. For a morphoid type $\sigma$, and for morphoids $x$ and $y$ with $x@\sigma$ and $y@\sigma$ defined,
we define $\mathrm{\bf iso}(\sigma,x,y)$ to be the type whose members are the morphoids $z \in \sigma$
such that $(x@\sigma)\circ z^{-1} \circ(y@\sigma)$ is defined.
\caption{{\bf Additional Definitions.}}
\label{fig:MoreDefs}
\end{framed}
\end{figure}
\begin{figure}
\begin{framed}
{\small
\medskip {\small\bf ($\rightarrow$.A)} For morphoid types $\sigma$ and $\tau$ we have that $\sigma \rightarrow \tau$ is a morphoid type
and for any interface template ${\cal A}$ for $\tau$
we have that $\mathrm{\bf Point} \rightarrow {\cal A}$ is an interface template for $\sigma \rightarrow \tau$.
\medskip {\small \bf ($\rightarrow$.V1)}
For morphoid types $\sigma_1$, $\sigma_2$, $\tau_1$ and $\tau_2$ with $\sigma_1 \circ \sigma_2$ defined
and $\tau_1 \circ \tau_2$ defined we have that $(\sigma_1 \rightarrow \tau_1) \circ (\sigma_2 \rightarrow \tau_2)$
is also defined and $(\sigma_1 \circ \sigma_2) \rightarrow (\tau_1 \circ \tau_2)
= (\sigma_1 \rightarrow \tau_1) \circ (\sigma_2 \rightarrow \tau_2)$.
\medskip
\noindent {\small \bf ($\rightarrow$.V2)} For morphoid types $\sigma$, $\tilde{\sigma}$, $\tau$ and $\tilde{\tau}$ with $\sigma \preceq \tilde{\sigma}$ and $\tau \preceq \tilde{\tau}$
we have $(\sigma \rightarrow \tau) \preceq (\tilde{\sigma} \rightarrow \tilde{\tau})$.
\medskip {\small \bf (App.V1)} For functions $f$ and $g$ with $f \circ g$ defined and values $x$ and $y$ with $f(x)$ and $f(y)$ defined and $x \circ y$ defined
we have that $(f \circ g)(x \circ y)$ and $f(x) \circ f(y)$ are both defined and $(f \circ g)(x \circ y) = f(x) \circ g(y)$.
\medskip
\noindent {\small \bf (App.V2)}
For morphoid functions $f$ and $\tilde{f}$ with $f \preceq \tilde{f}$, and morphoids $x$ and $\tilde{x}$ with $x \preceq \tilde{x}$,
and such that $f(x)$ and $\tilde{f}(\tilde{x})$ are both defined, we have $f(x) \preceq \tilde{f}(\tilde{x})$.
\medskip {\small \bf ($\updownarrow$.A)} For morphoid types $\sigma,\tau$ and bijection $f \in \sigma \rightarrow \tau$
we have that $\updownarrow\!(\sigma,\tau,f)$ is a morphoid point type.
\medskip{\small \bf ($\updownarrow$.B)} For $\sigma,\tau \in \convalue{\mathrm{\bf type}_i}$ and for a bijection $f \in \sigma \rightarrow \tau$,
we have $\updownarrow\!(\sigma,\tau,f) \in \mathrm{\bf iso}(\convalue{\mathrm{\bf type}_i},\sigma,\tau)$.
\medskip {\small ($\mathbf{Iso}$)} For a morphoid type $\sigma$ and morphoids $x$ and $y$ we have that $\mathrm{\bf iso}(\sigma,x,y)$ is a morphoid type
and any interface template for $\sigma$ is also an interface template for $\mathrm{\bf iso}(\sigma,x,y)$.
}
\caption{{\bf Additional Properties.}}
\label{fig:MoreProps}
\end{framed}
\end{figure}
Figure~\ref{fig:MoreDefs} gives some additional definitions. These additional definitions cover all remaining constructs used the definition of the
value function given in figure~\ref{fig:value}. Some properties of these definitions are given in figure~\ref{fig:MoreProps} and proved here.
\begin{lemma}[$\mathbf{Iso}$]
For a morphoid type $\sigma$ and morphoids $x$ and $y$ we have that $\mathrm{\bf iso}(\sigma,x,y)$ is a morphoid type
and any interface template for $\sigma$ is also an interface template for $\mathrm{\bf iso}(\sigma,x,y)$.
\label{lem:isomorph}
\end{lemma}
\begin{proof}
We have that the elements of $\mathrm{\bf iso}(\sigma,x,y)$ are elements of $\sigma$ and hence are morphoids. We must show that
$\mathrm{\bf iso}(\sigma,x,y)$ satisfies condition (T1) and has an interface template. For condition (T1)
consider $z_1,z_2,z_3 \in \mathrm{\bf iso}(\sigma,x,y)$ with $z_1 \circ z_2^{-1} \circ z_3$ defined.
Since $z_i \in \sigma$ we have that there exists $\tilde{x}_i$ and $\tilde{y}_i$ with $x \preceq \tilde{x}_i$ and $y \preceq \tilde{y}_i$
and with $\tilde{x}_i \circ z_i^{-1} \circ \tilde{y}_i$ defined. We then have that
$$\tilde{x}_3 \circ (z_1 \circ z_2^{-1} \circ z_3)^{-1} \circ \tilde{y}_1$$
is defined and hence $(z_1 \circ z_2^{-1} \circ z_3) \in \mathrm{\bf iso}(\sigma,x,y)$.
For the second part of the lemma let ${\cal A}$ be an interface template for $\sigma$
and consider $z \in \mathrm{\bf iso}(\sigma,x,y)$. Let $\tilde{x}$ and $\tilde{y}$ be such that $x \preceq \tilde{x}$ and $y \preceq \tilde{y}$ and
$\tilde{x} \circ z^{-1} \circ \tilde{y}$ is defined. We have $z@{\cal A} \in \sigma$ and by (Abs-Distributes-In) we have
$(\tilde{x}@{\cal A}) \circ (z@{\cal A})^{-1} \circ (\tilde{y}@{\cal A})$ which establishes that $z@{\cal A} \in \mathrm{\bf iso}(\sigma,x,y)$.
\end{proof}
\begin{lemma}[$\updownarrow$.A] For morphoid types $\sigma$ and $\tau$ and a bijection $f \in \sigma \rightarrow \tau$ we have that $\updownarrow\!(\sigma,\tau,f)$ is a morphoid point type.
\label{lem:lastmorph}
\end{lemma}
\begin{proof}
Recall that we have defined $\pi_i$ on points by $\pi_1(\mathrm{\bf Point}(i,j)) = i$ and $\pi_2(\mathrm{\bf Point}(i,j)) = j$
and that $\updownarrow\!(\sigma,\tau,f)$ is defined to be the type whose members are the points of the form
$\mathrm{\bf Point}(\pi_1(y@\mathrm{\bf Point}),\pi_2(x@\mathrm{\bf Point}))$ for $x \in \sigma$ and $y \in \tau$ with $y =_\tau f(x)$.
Since $\updownarrow\!(\sigma,\tau,f)$ contains only points, we need only show condition (T1). We let $x$ range over elements of $\sigma$ and $y$
range over elements of $\tau$. To show (T1) we suppose that
$$\begin{array}{cl} & \mathrm{\bf Point}(\pi_1(y_1@\mathrm{\bf Point}),\pi_2(x_1@\mathrm{\bf Point})) \\
\circ & \mathrm{\bf Point}(\pi_1(y_2@\mathrm{\bf Point}),\pi_2(x_2@\mathrm{\bf Point}))^{-1} \\
\circ & \mathrm{\bf Point}(\pi_1(y_3@\mathrm{\bf Point}),\pi_2(x_3@\mathrm{\bf Point})) \end{array}$$
is defined. This composition is equal to $\mathrm{\bf Point}(\pi_1(y_1@\mathrm{\bf Point}),\pi_2(x_3@\mathrm{\bf Point}))$. We must show that this
point is in $\updownarrow\!(\sigma,\tau,f)$ and, in particular, that $f(x_3) =_\tau y_1$. We will show $f(x_3) =_\tau y_3 =_\tau y_2 =_\tau f[x_2@\mathrm{\bf Point}] =_\tau f[x_1@\mathrm{\bf Point}] = _\tau y_1$.
For this it now suffices to show $y_3 =_\tau y_2$ and $x_2@\mathrm{\bf Point} =_{\sigma@\typeof(\pointt)} x_1@\mathrm{\bf Point}$.
Since the above composition is defined we have $\pi_2(x_1@\mathrm{\bf Point}) = \pi_2(x_2@\mathrm{\bf Point})$ which implies $\mathbf{Right}(x_1@\mathrm{\bf Point}) = \mathbf{Right}(x_2@\mathrm{\bf Point})$ which by (Composables-Equivalent) implies
$x_1@\mathrm{\bf Point} =_{\sigma@\typeof(\pointt)} x_2@\mathrm{\bf Point}$. Similarly, the definedness of the above composition
implies $\mathbf{Left}(y_2@\mathrm{\bf Point}) = \mathbf{Left}(y_3@\mathrm{\bf Point})$ which by (Composables-Equivalent) implies that $(y_2@\mathrm{\bf Point}) =_{\tau@\typeof(\pointt)} (y_3@\mathrm{\bf Point})$.
By ($=$.C) we then have $y_2 =_\tau y_1$.
\end{proof}
\begin{lemma}[($\updownarrow\!$.B) First Helper] For morphoid type $\sigma$ we have that $\mathbf{Left}(\sigma@\typeof(\pointt))$ is the set of points
of the form $\mathrm{\bf Point}(\pi_1(x_1@\mathrm{\bf Point}),\;\pi_1(x_2@\mathrm{\bf Point}))$ for $x_1,x_2 \in \sigma$ with $x_1 =_\sigma x_2$.
The dual statement holds for $\mathbf{Right}$ and $\pi_2$.
\end{lemma}
\begin{proof}
We show containment in both directions.
Consider a point in $\mathbf{Left}(\sigma@\typeof(\pointt))$. Such a point has the form $(x_1@\mathrm{\bf Point})\circ (x_2@\mathrm{\bf Point})^{-1}$ for
$x_1,x_2 \in \sigma$. Such a point can be written as $\mathrm{\bf Point}(\pi_1(x_1@\mathrm{\bf Point}),\;\pi_1(x_2@\mathrm{\bf Point}))$ with
$\mathbf{Right}(x_1@\mathrm{\bf Point}) = \mathbf{Right}(x_2@\mathrm{\bf Point})$. (Composables-Equivalent) implies
$x_1@\mathrm{\bf Point} =_{\sigma@\typeof(\pointt)} x_2@\mathrm{\bf Point}$
and property ($=$.C) implies $x_1 =_\sigma x_2$.
\medskip
\noindent Conversly consider $x_1,x_2 \in \sigma$ with $x_1 =_\sigma x_2$. We must show
$$\mathrm{\bf Point}(\pi_1(x_1@\mathrm{\bf Point}),\;\pi_1(x_2@\mathrm{\bf Point})) \in \mathbf{Left}(\sigma@\typeof(\pointt)).$$
By ($=$.C) we have $x_1@\mathrm{\bf Point} =_{\sigma@\typeof(\pointt)} x_2@\mathrm{\bf Point}$. This implies that there exists $x_3 \in \sigma$
with $(x_1@\mathrm{\bf Point}) \circ (x_3@\mathrm{\bf Point})^{-1} \circ (x_2@\mathrm{\bf Point})$ defined. We then have $\mathbf{Left}(x_2@\mathrm{\bf Point}) = \mathbf{Left}(x_3@\mathrm{\bf Point})$ and we have
\begin{eqnarray*}
\mathrm{\bf Point}(\pi_1(x_1@\mathrm{\bf Point}),\;\pi_1(x_2@\mathrm{\bf Point})) & =& \mathrm{\bf Point}(\pi_1(x_1@\mathrm{\bf Point}),\;\pi_1(x_3@\mathrm{\bf Point})) \\
& = & x_1 \circ x_3^{-1} \\
& \in & \mathbf{Left}(\sigma@\typeof(\pointt))
\end{eqnarray*}
\end{proof}
\begin{lemma}[($\updownarrow$.B) Second Helper]
For any morphoid types $\sigma$ and $\tau$, and any bijection $f$ in $\sigma \rightarrow \tau$, we have
$\mathbf{Left}(\updownarrow\!(\sigma,\tau,f)) =\mathbf{Left}(\tau@\typeof(\pointt))$ and $\mathbf{Right}(\updownarrow\!(\sigma,\tau,f)) =\mathbf{Right}(\sigma@\typeof(\pointt))$.
\end{lemma}
\begin{proof}
We will show $\mathbf{Left}(\updownarrow\!(\sigma,\tau,f)) = \mathbf{Left}(\tau@\typeof(\pointt))$ by showing containment in both directions.
By definition
$\updownarrow\!(\sigma,\tau,f)$ is the set of points $\mathrm{\bf Point}(\pi_1(y@\mathrm{\bf Point}),\;\pi_2(x@\mathrm{\bf Point}))$
for $x \in \sigma$ and $y \in \tau$ with $f(x) =_\tau y$. This implies that $\mathbf{Left}(\updownarrow\!(\sigma,\tau,f))$
is the set of points of the form $\mathrm{\bf Point}(\pi_1(y_1@\mathrm{\bf Point}),\;\pi_1(y_2@\mathrm{\bf Point}))$ such that there exists $x_1,x_2 \in \sigma$
with $\mathbf{Right}(x_1@\mathrm{\bf Point}) = \mathbf{Right}(x_2@\mathrm{\bf Point})$ and with $y_1 =_\tau f(x_1)$ and $y_2 =_\tau f(x_2)$. (Composables-Equivalent) implies that
for any such $x_1$ and $x_2$ we have $x_1@\mathrm{\bf Point} =_{\sigma@\typeof(\pointt)} x_2@\mathrm{\bf Point}$ and hence $y_1 =_\tau f[x_1@\mathrm{\bf Point}] = f[x_2@\mathrm{\bf Point}] =_\tau y_2$.
By the preceding lemma we have that
any point of the form $\mathrm{\bf Point}(\pi_1(y_1@\mathrm{\bf Point}),\;\pi_1(y_2@\mathrm{\bf Point}))$ with $y_1 =_\tau y_2$ is a member
of $\mathbf{Left}(\tau@\typeof(\pointt))$ and we have now have $\mathbf{Left}(\updownarrow\!(\sigma,\tau,f)) \subseteq \mathbf{Left}(\tau@\typeof(\pointt))$.
Conversely the preceding lemma states that any member of $\mathbf{Left}(\tau@\typeof(\pointt))$ is
a point of the form $\mathrm{\bf Point}(\pi_1(y_1@\mathrm{\bf Point}),\;\pi_1(y_2@\mathrm{\bf Point}))$ with $y_1 =_\tau y_2$.
Let $x \in \sigma$ be such that $f(x) =_\tau y_1 =_\tau y_2$. Such an $x$ must exist because $f$ is a bijection.
We then have $\mathrm{\bf Point}(\pi_1(y_1@\mathrm{\bf Point}),\pi_2(x@\mathrm{\bf Point}))$
and $\mathrm{\bf Point}(\pi_1(y_2@\mathrm{\bf Point}),\;\pi_2(x@\mathrm{\bf Point}))$
in $\updownarrow\!(\sigma,\tau,f)$ which gives that
$$\mathrm{\bf Point}(\pi_1(y_1@\mathrm{\bf Point}),\;\pi_1(y_2@\mathrm{\bf Point})) \in \mathbf{Left}(\updownarrow\!(\sigma,\tau,f)).$$
\end{proof}
\begin{lemma}[$\updownarrow$.B]
For $\sigma,\tau \in \convalue{\mathrm{\bf type}_i}$ and for a bijection $f \in \sigma \rightarrow \tau$,
we have $\updownarrow\!(\sigma,\tau,f) \in \mathrm{\bf iso}(\convalue{\mathrm{\bf type}_i},\sigma,\tau)$.
\end{lemma}
\begin{proof}
($\updownarrow$.A) states that that $\updownarrow\!(\sigma,\tau,f)$ is a point type and the properties of Grothendiek universes
imply $\updownarrow\!(\sigma,\tau,f) \in \convalue{\mathrm{\bf type}_i}$. It then suffices to show
that $(\sigma@\typeof(\pointt)) \circ \updownarrow\!(\sigma,\tau,f)^{-1} \circ (\tau@\typeof(\pointt))$ is defined.
But this follows from the second helper lemma above.
\end{proof}
\begin{lemma}[$\rightarrow$.A] For morphoid types $\sigma$ and $\tau$ we have that $\sigma \rightarrow \tau$ is a morphoid type
and for any interfate template ${\cal A}$ for $\tau$
we have that $\mathrm{\bf Point} \rightarrow {\cal A}$ is an interface template for $\sigma \rightarrow \tau$.
\end{lemma}
\begin{proof}
We have that $\sigma \rightarrow \tau$ is the set of morphoid functions $f$ with $\mathbf{Dom}(f) = \sigma@\typeof(\pointt)$ and such that for $x \in \sigma$
we have $f(x) \in \tau$. By definition every element of
$\sigma \rightarrow \tau$ is a morphoid. It remains to show conditions (T1) and that for any interface template template ${\cal A}$ for $\tau$
we have that $\mathrm{\bf Point} \rightarrow {\cal A}$ is an interface template for $\sigma \rightarrow \tau$.
To show (T1) consider $f,g,h \in \sigma \rightarrow \tau$ with $f \circ g^{-1} \circ h$ defined.
By property (Fun-Composition) we have that $f \circ g^{-1} \circ h$ is the function with domain $\mathbf{Dom}(f) \circ \mathbf{Dom}(g)^{-1} \circ \mathbf{Dom}(h) = \sigma@\typeof(\pointt)$
and satisfying
$(f \circ g^{-1} \circ h)[x] = f[x] \circ g[x]^{-1} \circ h[x]$.
Now consider an interface template ${\cal A}$ for $\tau$. We must show that for $f \in \sigma \rightarrow \tau$ we have that $f@(\mathrm{\bf Point} \rightarrow \tau) \in \sigma \rightarrow \tau$.
This is straightforward.
\end{proof}
\begin{lemma}[$\rightarrow$.V1]
For morphoid types $\sigma_1$, $\sigma_2$, $\tau_1$ and $\tau_2$ with $\sigma_1 \circ \sigma_2$ defined
and $\tau_1 \circ \tau_2$ defined we have that $(\sigma_1 \rightarrow \tau_1) \circ (\sigma_2 \rightarrow \tau_2)$
is also defined and $(\sigma_1 \circ \sigma_2) \rightarrow (\tau_1 \circ \tau_2)
= (\sigma_1 \rightarrow \tau_1) \circ (\sigma_2 \rightarrow \tau_2)$.
\end{lemma}
\begin{proof}
Throughout the proof we will let $x$ range over elements of $\sigma_1$ and $y$ range over elements of $\sigma_2$.
We first show that if $(\sigma_1 \rightarrow \tau_1) \circ (\sigma_2 \rightarrow \tau_2)$ is defined then the equation holds.
We will show containment of instances in each direction. For $h \in (\sigma_1 \rightarrow \tau_1)\circ(\sigma_2 \rightarrow \tau_2)$
we must show
$h \in (\sigma_1\circ \sigma_2 \rightarrow \tau_1 \circ \tau_2)$.
By definition $h$ has the form $f \circ g$ with $f \in (\sigma_1 \rightarrow \tau_1)$ and $g \in (\sigma_2 \rightarrow \tau_2)$.
By property (Funs-Composable) we have $\mathbf{Dom}(f \circ g) = \mathbf{Dom}(f) \circ \mathbf{Dom}(g) = (\sigma_1@\typeof(\pointt)) \circ (\sigma_2@\typeof(\pointt))$.
Since $\sigma_1 \circ \sigma_2$ is defined, property (Abs-Commutes-In) gives $(\sigma_1@\typeof(\pointt)) \circ (\sigma_2@\typeof(\pointt)) = (\sigma_1 \circ \sigma_2)@\typeof(\pointt)$.
It remains only to show that $(f\circ g)(x \circ y) \in \tau_1 \circ \tau_2$.
But by (Abs-Commutes-In) and property (Fun-Composition) we have $(f \circ g)[(x \circ y)@\mathrm{\bf Point}] = (f \circ g)[(x@\mathrm{\bf Point}) \circ (y@\mathrm{\bf Point})] = f(x) \circ g(y) \in \tau_1 \circ \tau_2$.
\medskip
Conversely consider
$h \in (\sigma_1\circ \sigma_2 \rightarrow \tau_1 \circ \tau_2)$.
We must show
$h \in (\sigma_1 \rightarrow \tau_1)\circ(\sigma_2 \rightarrow \tau_2)$.
More specifically we must show that $h$ can be written as $f \circ g$ for $f \in (\sigma_1 \rightarrow \tau_1)$
and $g \in (\sigma_2 \rightarrow \tau_2)$. We have $h(x \circ y) \in \tau_1 \circ \tau_2$.
For each equivalence class $|x \circ y|$ of $\sigma_1 \circ \sigma_2$ we can select $f(|x \circ y|) \in \tau_1$ and $g(|x \circ y|) \in \tau_2$
such that $h(x \circ y) = f(|x \circ y|) \circ g(|x \circ y|)$. By lemma ($=$.V1) we have that $x_1 =_{\sigma_1} x_2$
if and only if $x_1 \circ y_1 =_{\sigma_1 \circ \sigma_2} x_2 \circ y_2$. This implies that we can define $f'(|x|) = f(|x \circ y|)$ and this definition
is independent of the choice of $y$. We can define $g'(|y|)$ similarly. We then have
$h(x \circ y) = f'(x) \circ g'(y)$. Property ($=$.C) implies that $f'$ and $g'$ can be defined with $\mathbf{Dom}(f') = \sigma_1@\typeof(\pointt)$ and $\mathbf{Dom}(g') = \sigma_2@\typeof(\pointt)$.
We can also select values so that $f'$ and $g'$ have range templates equal to an interface template of $\tau_1$ which by the properties ($\sim$) must also be an interface template of $\tau_2$.
This gives $\mathbf{Dom}(h) = (\sigma_1 \circ \sigma_2)@\typeof(\pointt) = \sigma_1@\typeof(\pointt) \circ \sigma_2@\typeof(\pointt) = \mathbf{Dom}(f') \circ \mathbf{Dom}(g')$.
We now have $h = f' \circ g' \in (\sigma_1 \rightarrow \tau_1) \circ (\sigma_2 \rightarrow \tau_2)$ as desired.
Finally, we must now show that if $\sigma_1 \circ \sigma_2$ is defined and $\tau_1 \circ \tau_2$ is defined then
$(\sigma_1 \rightarrow \tau_1) \circ (\sigma_2 \rightarrow \tau_2)$ is defined. For this we note that
$(\sigma \rightarrow \tau) \circ (\sigma \rightarrow \tau)^{-1}$ is defined and hence by above proof we have
$\mathbf{Left}(\sigma \rightarrow \tau) = \mathbf{Left}(\sigma) \rightarrow \mathbf{Left}(\tau)$. The corresponding statement holds for $\mathbf{Right}$
and we get
\begin{eqnarray*}
\mathbf{Right}(\sigma_1 \rightarrow \tau_1) & = & \mathbf{Right}(\sigma_1) \rightarrow \mathbf{Right}(\tau_1) \\
& = & \mathbf{Left}(\sigma_2) \rightarrow \mathbf{Left}(\tau_2) \\
& = & \mathbf{Left}(\sigma_1 \rightarrow \tau_2)
\end{eqnarray*}
\end{proof}
\begin{lemma}[$\rightarrow$.V2]
\label{lem:arrowabs}
For morphoid types $\sigma$, $\tilde{\sigma}$, $\tau$ and $\tilde{\tau}$ with $\sigma \preceq \tilde{\sigma}$ and $\tau \preceq \tilde{\tau}$
we have $(\sigma \rightarrow \tau) \preceq (\tilde{\sigma} \rightarrow \tilde{\tau})$.
\end{lemma}
\begin{proof} We must show that if $(\tilde{\sigma} \rightarrow \tilde{\tau})@\mathbf{TypeOf}(\mathrm{\bf Point} \rightarrow {\cal A})$
is defined then we have that $(\tilde{\sigma} \rightarrow \tilde{\tau})@\mathbf{TypeOf}(\mathrm{\bf Point} \rightarrow {\cal A})$ is also defined and
$$(1)\;\; (\sigma \rightarrow \tau)@\mathbf{TypeOf}(\mathrm{\bf Point} \rightarrow {\cal A}) \;\;=\;\; (\tilde{\sigma} \rightarrow \tilde{\tau})@\mathbf{TypeOf}(\mathrm{\bf Point} \rightarrow {\cal A}).$$
Since $\sigma \preceq \tilde{\sigma}$ we have $\sigma@\typeof(\pointt) = \tilde{\sigma}@\typeof(\pointt)$.
This implies that all functions in both function types have the same domain.
If $\sigma$ is empty then all the function types involved contain only the empty function and the result holds.
So we can assume that $\sigma$ is non-empty.
To show that $(\sigma \rightarrow \tau)@\mathbf{TypeOf}(\mathrm{\bf Point} \rightarrow {\cal A})$ is
defined consider $f \in \sigma \rightarrow \tau$. It suffices to show that $f@(\mathrm{\bf Point} \rightarrow {\cal A})$ is defined.
For this it suffices to show that for any $x \in \mathbf{Dom}(f)$ we have that
$f[x]@{\cal A}$ is defined. But since $(\tilde{\sigma} \rightarrow \tilde{\tau})@\mathbf{TypeOf}(\mathrm{\bf Point} \rightarrow {\cal A})$
is defined, and we have assumed there is some $x \in \tilde{\sigma}@\typeof(\pointt)$, we must have that $y@{\cal A}$ is defined
for all $y \in \tilde{\tau}$. Since $\tau \preceq \tilde{\tau}$ we must then have that $z@{\cal A}$ is defined for all $z \in \tau$
and hence $f[x]@{\cal A}$ is defined.
To show (1) let ${\cal B}$ be an interface template for $\tilde{\tau}$. We then have
$$(2)\;\;\tau@\mathbf{TypeOf}({\cal B}) = \tilde{\tau}@\mathbf{TypeOf}({\cal B}) \subseteq \tilde{\tau}.$$
To show (1) we show containment in both directions. First, let $f$ be a function in
$\sigma \rightarrow \tau$. We must show that $f@(\mathrm{\bf Point} \rightarrow {\cal A})$
is in $(\tilde{\sigma} \rightarrow \tilde{\tau})@\mathbf{TypeOf}(\mathrm{\bf Point} \rightarrow {\cal A})$.
Let $g$ be the function defined by $g[x] = f[x]@{\cal B}$. Equation (2) implies $g \in (\tilde{\sigma} \rightarrow \tilde{\tau})$.
Property (Abs-Compression) now implies
$$f@(\mathrm{\bf Point} \rightarrow {\cal A}) = g@(\mathrm{\bf Point} \rightarrow {\cal A}) \in (\tilde{\sigma} \rightarrow \tilde{\tau})@\mathbf{TypeOf}(\mathrm{\bf Point} \rightarrow {\cal A}).$$
For the converse consider a morphoid function $g \in \tilde{\sigma} \rightarrow \tilde{\tau}$. For each value of the form $g[x]@{\cal B}$
equation (2) implies that there exists $y \in \tau$ with $y@{\cal B} = g[x]@{\cal B}$. We can pick one such $y$ for each equivalence class of $\sigma@\typeof(\pointt)$
to get a function $f \in \sigma \rightarrow \tau$ with $f[x]@{\cal B} = g[x]@{\cal B}$. This gives
$$g@(\mathrm{\bf Point} \rightarrow {\cal A}) = f@(\mathrm{\bf Point} \rightarrow {\cal A}) \in (\sigma \rightarrow \tau)@\mathbf{TypeOf}(\mathrm{\bf Point} \rightarrow {\cal A}).$$
\end{proof}
\begin{lemma}[App.V1]
For functions $f$ and $g$ with $f \circ g$ defined and values $x$ and $y$ with $f(x)$ and $f(y)$ defined and $x \circ y$ defined
we have that $(f \circ g)(x \circ y)$ and $f(x) \circ f(y)$ are both defined and $(f \circ g)(x \circ y) = f(x) \circ g(y)$.
\end{lemma}
\begin{proof}
For $f \circ y$ defined we have $\mathbf{Dom}(f) \circ \mathbf{Dom}(g)$ defined. For $f(x)$ defined we have $x@\mathrm{\bf Point} \in \mathbf{Dom}(f)$
and for $g(y)$ defined we have $y@\mathrm{\bf Point} \in \mathbf{Dom}(g)$. By (Abs-Commutes-In) we have $(x \circ y)@\mathrm{\bf Point} = (x@\mathrm{\bf Point}) \circ (y@\mathrm{\bf Point}) \in \mathbf{Dom}(f \circ g)$.
By (Fun-Composition) we then have
\begin{eqnarray*}
(f \circ g)(x \circ y) & = & (f \circ g)[(x@\mathrm{\bf Point}) \circ (y@\mathrm{\bf Point})] \\
& = & f[x@\mathrm{\bf Point}] \circ g[y@\mathrm{\bf Point}] \\
& = & f(x) \circ g(y)
\end{eqnarray*}
\end{proof}
\begin{lemma}[App.V2]
For morphoid functions $f$ and $\tilde{f}$ with $f \preceq \tilde{f}$, and morphoids $x$ and $\tilde{x}$ with $x \preceq \tilde{x}$,
and such that $f(x)$ and $\tilde{f}(\tilde{x})$ are both defined, we have $f(x) \preceq \tilde{f}(\tilde{x})$.
\end{lemma}
\begin{proof}
Let ${\cal A}$ be a range template for $\tilde{f}$. By the definition of a range template we have that $\tilde{f}@(\mathrm{\bf Point} \rightarrow {\cal A}) = \tilde{f}$.
Since $f \preceq \tilde{f}$ we have $f@(\mathrm{\bf Point} \rightarrow {\cal A}) = \tilde{f}$.
This implies that $\mathbf{Dom}(f) = \mathbf{Dom}(\tilde{f})$. Since $x \preceq \tilde{x}$ we also have that $x@\mathrm{\bf Point} = \tilde{x}@\mathrm{\bf Point}$.
Now consider a template ${\cal B}$ such that $\tilde{f}(\tilde{x})@{\cal B}$ is defined. We now have
\begin{eqnarray*}
\tilde{f}(\tilde{x})@{\cal B} & = & \tilde{f}[x@\mathrm{\bf Point}]@{\cal A}@{\cal B} \\
& = & (f@(\mathrm{\bf Point} \rightarrow {\cal A}))[x@\mathrm{\bf Point}]@{\cal B} \\
& = & f[x@\mathrm{\bf Point}]@{\cal A}@{\cal B} \\
& = & f[x@\mathrm{\bf Point}]@{\cal B} \\
& = & f(x)@{\cal B}
\end{eqnarray*}
\end{proof}
\eject
\section{Soundness Proofs}
\label{sec:soundness}
We now prove the soundness of the inference rules under morphoid semantics. Section~\ref{subsec:AllValues} proves that every value is a morphoid --- for $\convalue{\Sigma}$
and $\semvalue{e}$ defined, and for $\rho \in \convalue{\Sigma}$, we have that $\semvalue{e}\rho$ is a morphoid. The main difficulty in this proof is establishing
that an interface template exists for pair type expressions --- that $\semvalue{\tau}\rho$ has an interface template in the case where $\tau$ is a dependent pair type.
This is done by defining an abstract interpretation function $\tempvalue{e}\eta$ with the property that for $\convalue{\Sigma}$ defined, and $\semvalue{e}$ defined,
and $\rho \in \convalue{\Sigma}$ and $\eta$ a minimal template for $\rho$, we have that $\tempvalue{e}\eta$ is a minimal template for $\semvalue{e}\rho$.
Section~\ref{subsec:PairTypeFormation} proves the soundness of the inference rule for forming pair types. Section~\ref{subsec:AllValues}
has already proved that pair type expressions denote morphoid types. However, to show $\Sigma \models \intype{\tau}{\mathrm{\bf type}_i}$
one must show that the expression $\tau$ satisfies conditions (V1) and (V2) in figure~\ref{fig:models}. The proof of soundness for this one rule
excercises most of the components of morphoid type theory.
Section~\ref{subsec:Substitution} proves the soundness of the substitution rule in figure~\ref{fig:rules} and of the isomorphism rules in figure~\ref{fig:simple}.
Section~\ref{subsec:remaining} proves the soundness of the remaining rules.
\medskip
\noindent Before considering the various soundness proofs we observe the following consequences of properties (V1) and (V2).
\begin{lemma}
\label{lem:convenient}
For $\Sigma;\;\intype{x}{\sigma} \models \intype{e[x]}{\tau[x]}$ and for $\rho \in \convalue{\Sigma}$ let $\sigma^*$ abbreviate $\semvalue{\sigma}\rho$,
and for $u \in \semvalue{\sigma}\rho$
let $e^*[u]$ abbreviate $\subvalue{\Sigma;\intype{x\;}{\;\sigma}}{e[x]}\rho[x \leftarrow u]$.
If $\rho(y)$ is a morphoid for all variables $y$ assigned a value by $\rho$ (which we prove below to be true) then we have the following corollaries of (V1) and (V2).
\begin{itemize}
\item[(1)] For $u_1,u_2,u_3 \in \sigma^*$ with $u_1 \circ u_2^{-1} \circ u_3$ defined we have
$$e^*[u_1 \circ u_2^{-1} \circ u_3] = e^*[u_1] \circ e^*[u_2^{-1}] \circ e^*[u_3].$$
\item[(2)] For $u \in \sigma^*$ we have $e^*[u] \preceq e^*[u@\sigma^*]$.
\item[(3)] For $u \in \sigma^*$ with $e^*[u@\sigma^*]@{\cal T}$ defined we have $e^*[u@\sigma^*]@{\cal T} = e^*[u]@{\cal T}$.
\end{itemize}
\end{lemma}
\begin{proof}
For (1) consider $u_1,u_2,u_3 \in \sigma^*$ with $u_1 \circ u_2^{-1} \circ u_3$ defined. By the definition of
$\Sigma\;\intype{x}{\sigma} \models \intype{e[x]}{\tau}$ we have that $\convalue{\Sigma;\intype{x}{\sigma}}$ is defined
and hence $\Sigma \models \inntype{\sigma}{\mathrm{\bf type}_i}$. This implies that $\sigma^*$ is a morphoid type and hence
$u_1 \circ u_2^{-1} \circ u_3 \in \sigma^*$. We then have $\rho[x \leftarrow (u_1 \circ u_2^{-1} \circ u_3)] \in \convalue{\Sigma;\intype{x}{\sigma}}$.
But since $\rho(y) = \rho(y) \circ \rho(y)^{-1} \circ \rho(y)$ we have $\rho = \rho \circ \rho^{-1} \circ \rho$ and
$$\rho[x \leftarrow (u_1 \circ u_2^{-1} \circ u_3)] \;\;= \rho[x \leftarrow u_1] \circ \rho[x \leftarrow u_2]^{-1} \circ \rho[x \leftarrow u_3].$$
By (V1) we then have
\begin{eqnarray*}
e^*[u_1 \circ u_2^{-1} \circ u_3] & = & \subvalue{\Sigma;\intype{x\;}{\;\sigma}}{e[x]}\rho[x \leftarrow (u_1 \circ u_2^{-1} \circ u_3)] \\
& = & \subvalue{\Sigma;\intype{x\;}{\;\sigma}}{e[x]}(\rho[x \leftarrow u_1] \circ \rho[x \leftarrow u_2]^{-1} \circ \rho[x \leftarrow u_3]) \\
& = & e^*[u_1] \circ e^*[u_2]^{-1} \circ e^*[u_3]
\end{eqnarray*}
For (2) we note that by property ($\preceq$.B) we have $u \preceq u@\sigma^*$ which implies $\rho[x \leftarrow u] \preceq \rho[x \leftarrow u@\sigma^*]$.
By (V2) we then have (2). Part (3) follows from part (2) and the definition of $\preceq$.
\end{proof}
\subsection{All Values are Morphoids}
\label{subsec:AllValues}
\begin{figure}[p]
\begin{framed}
{\bf Structure Template.} A structure template is a mapping from a finite set of variables to templates.
\medskip
{\bf Minimal Structure Template.}
For a structure $\rho$ and structure template $\eta$ defined on the same variables as $\rho$
we say that $\eta$ is a minimal template for $\rho$ if for each variable $x$ we have that $\eta(x)$ is a minimal template for
$\rho(x)$.
\medskip
${\bf \tempvalue{e}\eta}$. For an expression $e$ and structure
template $\eta$ the following rules define a template $\tempvalue{e}\eta$
where $\tempvalue{e}\eta$ is undefined if no rule applies or if
some expression on the right hand side of the matching rule is undefined.
\begin{eqnarray*}
\tempvalue{x}{\eta} & = & \eta(x) \\
\tempvalue{\mathrm{\bf Bool}}{\eta} & = & \mathbf{TypeOf}(\mathrm{\bf Bool}) \\
\tempvalue{\mathrm{\bf type}_i}{\eta} & = & \mathbf{TypeOf}(\typeof(\pointt)) \\
\tempvalue{\sigma \rightarrow \tau}{\eta} & = & \mathbf{TypeOf}(\mathrm{\bf Point} \rightarrow \mathbf{Mem}(\tempvalue{\tau}\eta)) \\
\tempvalue{\subtype{\intype{x}{\sigma},\;\Phi[x]}}{\eta} & = & \tempvalue{\sigma}{\eta} \\
\tempvalue{\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}}{\eta} & = & {\small
\mathbf{TypeOf}\left(\mathbf{Pair}\left(\begin{array}{l} \mathbf{Mem}(\tempvalue{\sigma}\eta), \\
\mathbf{Mem}(\tempvalue{\tau[x]}\eta[x \leftarrow \mathbf{Mem}(\tempvalue{\sigma}\eta)]) \end{array}\right)\right)} \\
\tempvalue{f(e)}{\eta} & = & \mathbf{Range}(\tempvalue{f}{\eta}) \\
\tempvalue{\mathbf{Pair}(u,\;w)}{\eta} & = & \mathbf{Pair}(\tempvalue{u}{\eta},\;\tempvalue{w}{\eta}) \\
\tempvalue{\pi_i[e]}{\eta} & = & \pi_i(\tempvalue{e}{\eta}) \\
\tempvalue{\updownarrow\!(\sigma,\tau,f)}\eta & = & \typeof(\pointt) \\
\tempvalue{\mathrm{\bf iso}(\sigma,a,b)}\eta & = & \tempvalue{\sigma}\eta \\
\tempvalue{\mathbf{The}(\intype{x}{\sigma},\;\Phi[x])}\eta & = & \mathbf{Mem}(\tempvalue{\sigma}\eta) \\
\tempvalue{\Phi}{\eta} & = & \mathrm{\bf Bool} \;\;\;\mbox{for $\Phi$ a subscripted equality, absolute equality,} \\
& & \mbox{expression of the form $(\inntype{e}{\sigma})$, disjunction, negation,} \\
& & \mbox{or quantified formula}
\end{eqnarray*}
\begin{eqnarray*}
\mathbf{Mem}(\mathbf{TypeOf}({\cal A})) & = & {\cal A} \\
\mathbf{Range}(\mathrm{\bf Point} \rightarrow {\cal A}) & = & {\cal A} \\
\pi_i(\mathbf{Pair}({\cal T}_1,\;{\cal T}_2)) & = & {\cal T}_i
\end{eqnarray*}
\caption{{\bf Template Evaluation.} For $\semvalue{e}$ defined, $\rho \in \convalue{\Sigma}$ and $\eta$ a minimal template of $\rho$ we have that $\tempvalue{e}\eta$
is a minimal template of $\semvalue{e}\rho$.}
\label{fig:AbsEval}
\end{framed}
\end{figure}
Figure~\ref{fig:AbsEval} defines a form of abstract interpretation \cite{Cousot77} which computes a minimal template for an expression $e$ using an assignment $\eta$ of a minimal
template for each free variable of $e$. We now prove the following theorem which simultaneously establishes that all values are morphoids and that the template evaluation
in figure~\ref{fig:AbsEval} produces minimal templates.
\begin{theorem}
\label{thm:AllValues}
~
\begin{itemize}
\item[(1)] For $\convalue{\Sigma}$ defined and $\rho \in \convalue{\Sigma}$ we have that $\rho$ is a morphoid structrure, i.e., $\rho(x)$ is a morphoid
for each variable $x$ declared in $\Sigma$.
\item[(2)]For $\semvalue{e}$ defined, for $\rho \in \convalue{\Sigma}$, and $\eta$ a minimal template of $\rho$,
we have that $\semvalue{e}\rho$ is a morphoid with minimal template $\tempvalue{e}\eta$.
\end{itemize}
\end{theorem}
\begin{proof}
The proof is by induction on the combined syntactic complexity of $\Sigma$ and $e$.
\medskip
Part (1) is immediate for the empty context and follows immediately from the induction hypothesis for
contexts $\Sigma;\Phi$ with $\convalue{\Sigma;\Phi} \subseteq \convalue{\Sigma}$. Now consider a context of the form $\Sigma;\intype{x}{\sigma}$
and consider $\rho[x \leftarrow v]$ with $v \in \semvalue{\sigma}\rho$.
By the induction hypothesis we have that $\rho$ is a morphoid structure and by the induction hypothesis for (2)
we have that $\semvalue{\sigma}\rho$ is a morphoid type and hence $v$ is a morphoid and $\rho[x \leftarrow v]$ is a morphoid structure.
\medskip
For the proof of (2) there is a case for each clause in the definition $\semvalue{e}\rho$.
Many of these cases are immediate. For example, the lemma is immediate for variables and Boolean expressions.
The case of function applications $f(w)$, pairs $\mathbf{Pair}(u,w)$ and projections $\pi_i(w)$ are also essentially immediate.
The case of $\mathrm{\bf type}_i$ is handled by property ($\mathrm{\bf type}_i$) in figure~\ref{fig:MorphAbsProps}. The cases of $\mathrm{\bf iso}(\sigma,x,y)$, $\updownarrow\!(\sigma,\tau,f)$, and $\sigma \rightarrow \tau$
are handled by properties ($\mathbf{Iso}$), ($\updownarrow$.A) and ($\rightarrow$.A) respectively in figure~\ref{fig:MoreProps}.
For the case of definite descriptions we note that the definition of $\mathbf{The}(\sigma)$ in figure~\ref{fig:MoreDefs} implies that
any interface template of $\sigma$ is a minimal template of $\mathbf{The}(\sigma)$. We now explicitly handle the cases of pair types and subtypes.
Each of these two cases is written as a lemma where the proof makes use of the induction hypothesis for theorem~\ref{thm:AllValues}.
\end{proof}
Before proving the case of pair types it is worth pointing out the delicate nature of this lemma and the need for template evaluation.
As a first example consider the following type of magmas --- structures with a binary operation not subject to any conditions.
$$\vdash \intype{\mathbf{Pair}(\intype{\alpha}{\mathbf{Set}},\;\intype{f}{\alpha \times \alpha \rightarrow \alpha})}{\mathbf{Class}}$$
The interface template for this type is $\mathbf{Pair}(\mathbf{TypeOf}(\mathrm{\bf Point}),\;\mathrm{\bf Point} \rightarrow (\mathrm{\bf Point} \rightarrow \mathrm{\bf Point}))$.
However, a particular magma $\mathbf{Pair}(\alpha,f)$ will in general have a minimal template different from the interface
template for magmas --- the domain type $\alpha$ is not required to be a point type.
The derivation of the magma type involves the sequent
$$\intype{\alpha}{\mathbf{Set}} \vdash \intype{\alpha \times \alpha \rightarrow \alpha}{\mathbf{Set}}.$$
When we consider an arbitrary set (type) $\alpha$ we have that interface template of $\alpha$ need not be $\mathbf{TypeOf}(\mathrm{\bf Point})$.
The interface template for $\alpha \times \alpha \rightarrow \alpha$
is $\mathrm{\bf Point} \rightarrow (\mathrm{\bf Point} \rightarrow {\cal A}_\alpha)$ where ${\cal A}_\alpha$ is the interface template of $\alpha$.
More generally, consider
$$\intype{x}{\sigma} \vdash \intype{\tau[x]}{\mathrm{\bf type}_i}.$$
As the magma example shows, the interface template for $\tau[x]$ in general depends on the choice of $x$.
However, we must show that the pair type $\mathbf{PairType}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})$ has a single interface template.
Another interesting example is the following where we take ${\cal N}$ to be the type of natural numbers.
$$\intype{f}{{\cal N} \rightarrow \mathbf{Set}} \vdash \intype{\mathbf{PairOf}(\intype{x}{{\cal N}},\;\intype{y}{f(x)})}{\mathbf{Set}}.$$
Here the requirement that every function has a range type is important. Since range types must be abstract, the range type of $f$ must be
$\typeof(\pointt)$.
\begin{lemma} Property (2) of theorem~\ref{thm:AllValues} holds for $\semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}\rho$.
\end{lemma}
\begin{proof}
Since $\semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}$ is defined we have
$\Sigma \models \intype{\sigma}{\mathrm{\bf type}_i}$ and $\Sigma;\;\intype{x}{\sigma} \models \intype{\tau[x]}{\mathrm{\bf type}_i}$ (for some $i$).
Let $\sigma^*$ abbreviate
$\semvalue{\sigma}\rho$ and for $u \in \sigma^*$ let $\tau^*[u]$ abbreviate $\subvalue{\Sigma;\;\intype{x\;}{\;\sigma}}{\tau[x]}\rho[x \leftarrow u]$.
We have
$$\semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}\rho = (\tagg{TYPE},\;\{\mathbf{Pair}(u,w),\; u \in \sigma^*,\;w \in \tau^*[u]\}).$$
We must show that this a morphoid type. By the induction hypothesis we have that $\sigma^*$ is a morphoid type
and for $u \in \sigma^*$ we have that $\tau^*[u]$ is a morphoid type. This implies that every member of the pair type is a morphoid pair.
To show condition (T1) Consider $u_1,u_2,u_3 \in \sigma^*$ with $u_1 \circ u_2^{-1} \circ u_3$ defined
and consider $w_1 \in \tau^*[u_1]$, $w_2 \in \tau^*[u_2]$ and $w_3 \in \tau^*[u_3]$ with $w_1 \circ w_2^{-1} \circ w_3$ defined.
We must show that
$$\mathbf{Pair}(u_1\circ u_2^{-1} \circ u_3,\;w_1 \circ w_2^{-1} \circ w_3)$$
is in the pair type.
By condition (T1) on $\sigma^*$ we have $u_1 \circ u_2^{-1} \circ u_3 \in \sigma^*$.
We must show that $w_1 \circ w_2^{-1} \circ w_3 \in \tau^*[u_1 \circ u_2^{-1} \circ u_3]$.
By lemma~\ref{lem:convenient} on the entailment $\Sigma;\;\intype{x}{\sigma} \models \intype{\tau[x]}{\mathrm{\bf type}_i}$
we have $\tau^*[u_1 \circ u_2^{-1} \circ u_3] = \tau^*[u_1] \circ \tau^*[u_2]^{-1} \circ \tau^*[u_3]$.
Since $w_1 \circ w_2^{-1} \circ w_3 \in \tau^*[u_1] \circ \tau^*[u_2]^{-1} \circ \tau^*[u_3]$
this proves the result.
We must also show that the pair type has an interface template and that template evaluation computes such a minimal template for the pair type.
By property (Min-Template.C) we have that there exists a minimal template $\eta$ for $\rho$.
We will show that $\tempvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}\eta$ is a minimal template for the pair type.
Let ${\cal A}$ abbreviate $\mathbf{Mem}(\tempvalue{\sigma}\eta)$ and let
${\cal B}$ abbreviate $\mathbf{Mem}(\tempvalue{\tau[x]}\eta[x \leftarrow {\cal A}])$.
We have that
$$\tempvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}\eta = \mathbf{TypeOf}(\mathbf{Pair}({\cal A},{\cal B})).$$
By (Min-Template.A) it now suffices to show that $\mathbf{Pair}({\cal A},{\cal B})$ is an interface template for the pair type.
By the induction hypothesis we have that ${\cal A}$ is an interface template for $\sigma^*$ and for $u \in \sigma^*$ we have that ${\cal B}$
is an interface template for $\tau^*[u@{\cal A}]$.
Consider $\mathbf{Pair}(u,w)$ with $u \in \sigma^*$ and $w \in \tau^*[u]$. We must show that $\mathbf{Pair}(u@{\cal A},w@{\cal B})$
is in the pair type.
By condition (V2) of $\Sigma;\intype{x}{\sigma} \models \intype{\tau[x]}{\mathrm{\bf type}_i}$ we have
that $\tau^*[u] \preceq \tau^*[u@{\cal A}]$.
We then get
$$\tau^*[u]@\mathbf{TypeOf}({\cal B}) = \tau^*[u@{\cal A}]@\mathbf{TypeOf}({\cal B}).$$
Since $w \in \tau^*[u]$ we then have that $w@{\cal B}$ is defined. Furthermore,
$w@{\cal B} = w'@{\cal B}\;\mbox{for some}\;w' \in \tau^*[v@{\cal A}]$.
We now have $\mathbf{Pair}(u@{\cal A},w@{\cal B}) = \mathbf{Pair}(u@{\cal A},w'@{\cal B})$ with $w' \in \tau^*[u@{\cal A}]$ which proves the result.
\end{proof}
\begin{lemma} Property (2) of theorem~\ref{thm:AllValues} holds for $\semvalue{\mathbf{SubType}(\intype{x}{\sigma},\;\Phi[x])}\rho$.
\end{lemma}
\begin{proof} Consider $\rho \in \convalue{\Sigma}$.
Let $\sigma^*$ abbreviate $\semvalue{\sigma}\rho$ and for $u \in \sigma^*$ let $\Phi^*[u]$ abbreviate
$\subvalue{\Sigma;\in\mathrm{\bf type}{x\;}{\;\sigma}}{\Phi[x]}\rho[x \leftarrow u]$. We have
$$\semvalue{\mathbf{SubType}(\intype{x}{\sigma},\;\Phi[x])}\rho = (\tagg{TYPE},\;\{u \in \sigma^* \;s.t. \;\Phi^*[u]\}).$$
We have that every element of this type is an element of $\sigma^*$ and hence is a morphoid.
Condition (T1) for the subtype type follows from condition (T1) for $\sigma^*$ and condition (V1) of $\Sigma;\intype{x}{\sigma} \models \intype{\Phi[x]}{\mathrm{\bf Bool}}$.
More explicitly, consider $u_1,u_2,u_3$ in the subtype. We must show that $u_1 \circ u_2^{-1} \circ u_3$ is in the subtype.
We have $u_1,u_2,u_3\in \sigma^*$ and $\Phi^*[u_1]$, $\Phi^*[u_2]$ and $\Phi^*[u_3]$.
By condition (T1) of $\sigma^*$ we have that $(u_1 \circ u_2^{-1} \circ u_3) \in \sigma^*$.
By condition (V1) of $\Sigma;\intype{x}{\sigma} \models \intype{\Phi[x]}{\mathrm{\bf Bool}}$
we have $\Phi^*[u_1 \circ u_2^{-1} \circ u_3] = \Phi^*[u_1] \circ \Phi^*[u_2] \circ \Phi^*[u_3] = \mathrm{\bf True}$
which implies that $u_1 \circ u_2^{-1} \circ u_3$ is in the subtype.
Now let ${\cal A}$ be an interface template for $\sigma^*$ and consider $u \in \sigma^*$ such that $\Phi^*[u]$.
By condition (V1) of $\Sigma;\intype{x}{\sigma} \models \intype{\Phi[x]}{\mathrm{\bf Bool}}$ we have
$\Phi^*[u] \preceq \Phi^*[u@{\cal A}]$ which implies $\Phi^*[u@{\cal A}]$ and hence $u@{\cal A}$ is in the subtype.
\end{proof}
\subsection{The Soundness of Pair Type Formation}
\label{subsec:PairTypeFormation}
The following lemma is relevant to the proof of the soundness of pair type formation.
\begin{lemma}
\label{lem:Corollary.V1}
If $\Sigma \models \intype{e}{\tau}$ then
\begin{itemize}
\item[(V3)] for $\rho,\gamma \in \convalue{\Sigma}$ with $\rho \circ \gamma$ defined we have
$(\semvalue{e}\rho) \circ (\semvalue{e}\gamma)$ is defined.
\end{itemize}
\end{lemma}
\begin{proof}
Consider $\rho_1,\rho_2 \in \convalue{\Sigma}$ with $\rho_1 \circ \rho_2$ defined. Let $e_i^*$ abbreviate $\semvalue{e}\rho_i$.
We have $\rho_1 \circ \rho_2 \circ \rho_2^{-1} = \rho_1 \in \convalue{\Sigma}$.
By property (V1) of $\Sigma \models \intype{e}{\tau}$ we then have $\semvalue{e}(\rho_1 \circ \rho_2 \circ \rho_2^{-1}) = e_1^* \circ e_2^* \circ {e_2^*}^{-1}$
which gives that $e_1^* \circ e_2^*$ is defined.
\end{proof}
We have that (V1) implies (V3). It will be convenient below to prove (V3) as a first step in proving (V1).
\medskip
\noindent The pair type formation rule is
~ \hfill
\unnamed
{\ant{\Sigma \vdash \intype{\sigma}{\mathrm{\bf type}_i}}
\ant{\Sigma;\;\intype{x}{\sigma} \vdash \intype{\tau[x]}{\mathrm{\bf type}_i}}}
{\ant{\Sigma \vdash \intype{\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}}{\mathrm{\bf type}_i}}}
\hfill ~
The previous section established that when the antecedents of this rule are valid we have $\Sigma \models \inntype{\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}}{\mathrm{\bf type}_i}$.
But we have not yet established that the pair type expression satisfies conditions (V1) and (V2) as required in figure~\ref{fig:models}.
Before giving this proof, it is insightful to point out the following corollary of condition (V1).
\begin{lemma}
\label{thm:core2}
The inference rule for pair type formation
is sound.
\end{lemma}
\begin{proof}
As discussed above, we must show that the conclusion of the rule satisfies conditions (V1) and (V2) in figure~\ref{fig:models}.
For condition (V1) consider $\rho_1,\rho_2,\rho_3 \in \convalue{\Sigma}$ with $(\rho_1 \circ \rho_2^{-1} \circ \rho_3) \in \convalue{\Sigma}$.
We must show that
$$\semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}(\rho_1 \circ \rho_2^{-1} \circ \rho_3)$$
$$=$$
$$\semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}\rho_1 \circ \semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}\rho_2^{-1}
\circ \semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}\rho_3.$$
\medskip
We will first show (V3) for the pair type. For this, consider $\rho_1,\rho_2 \in \convalue{\Sigma}$ with $\rho_1 \circ \rho_2$ defined.
Let $\sigma_i^*$ abbreviate $\semvalue{\sigma}\rho_i$ and for $u \in \sigma_i^*$ let
$\tau_i^*[u]$ abbreviate $\subvalue{\Sigma;\intype{x\;}{\;\sigma}}{\tau[x]}\rho[x \leftarrow u]$.
We must show that the set of pairs of the form $\mathbf{Pair}(\mathbf{Right}(u),\mathbf{Right}(w))$ for $u \in \sigma_1^*$ and $w \in \tau_1^*[u]$
is the same as the set of pairs of the form $\mathbf{Pair}(\mathbf{Left}(u),\mathbf{Left}(w))$ for $u \in \sigma_2^*$ and $w \in \tau_2^*[u]$.
We will show that every pair of the first form is also of the second form. The converse is similar.
Consider $u_1 \in \sigma^*_1$ and $w_1 \in \tau^*_1[u_1]$. It now suffices to show that there exists $u_2 \in\sigma^*_2$ and $w_2$ in $\tau^*_2[u_2]$
with $u_1 \circ u_2$ and $w_1 \circ w_2$ defined. By (V1) for the first antecedent and (V1-Corollary) we have (V3) for the first antecedent.
This gives that $\sigma^*_1 \circ \sigma^*_2$ is defined and by (Partner) there exists $u_2 \in \sigma^*_2$ with $u_1 \circ u_2$ defined. We then have
$\rho_1[x \leftarrow u_1]$ and $\rho_2[x \leftarrow u_2]$ are both in $\convalue{\Sigma;\;\intype{x}{\sigma}}$ with
$(\rho_1[x \leftarrow u_1]) \circ (\rho_2[x \leftarrow u_2])$ defined. By (V3) of the second antecedent we then get
that $\tau^*_1[u_1] \circ \tau^*_2[u_2]$ is defined. By (Partner) we then have that there exists $w_2 \in \tau^*_2[u_2]$
with $w_1 \circ w_2$ defined. We have now established (V3) for the pair type.
To show (V1) for the pair type consider $\rho_1,\rho_2,\rho_3 \in \convalue{\Sigma}$ with $\rho_1 \circ \rho_2^{-1}\circ \rho_3$ defined and
$(\rho_1 \circ \rho_2^{-1} \circ \rho_3) \in \convalue{\Sigma}$. (V3) implies that
$$\semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}\rho_1 \circ \semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}\rho_2^{-1}
\circ \semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}\rho_3.$$
is defined. We must show that this equals
$$\semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}(\rho_1 \circ \rho_2^{-1} \circ \rho_3).$$
We show containment in both directions. To show that every member of the former is a member of the latter consider $u_i \in \sigma^*_i$
and $w_i \in \tau_i^*(u_i)$ with $u_1 \circ u_2^{-1} \circ u_3$ defined and $w_1 \circ w_2^{-1} \circ w_3$ defined.
We must show
$$\mathbf{Pair}(u_1 \circ u_2^{-1} \circ u_3,\;w_1\circ w_2^{-1} \circ w_3) \in
\semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}(\rho_1 \circ \rho_2^{-1} \circ \rho_3).$$
By condition (V1) of the first antecedent we have
$$\semvalue{\sigma}(\rho_1 \circ \rho_2^{-1} \circ \rho_3) = \sigma_1^* \circ {\sigma_2^*}^{-1} \circ \sigma_3^*$$
which gives $$(u_1 \circ u_2^{-1} \circ u_3) \in \semvalue{\sigma}(\rho_1 \circ \rho_2^{-1} \circ \rho_3).$$
By condition (V1) of the second antecedent we have
\begin{eqnarray*}
& & \subvalue{\Sigma;\;\intype{x\;}{\;\sigma}}{\tau[x]}(\rho_1 \circ \rho_2^{-1} \circ \rho_3)[x \leftarrow (u_1 \circ u_2^{-1} \circ u_3)] \\
& = & \subvalue{\Sigma;\;\intype{x\;}{\;\sigma}}{\tau[x]}(\rho_1[x \leftarrow u_1] \circ \rho_2[x \leftarrow u_2]^{-1} \circ \rho_3[x \leftarrow u_3]) \\
& = & \tau_1^*[u_1] \circ \tau_2^*[u_2]^{-1} \circ \tau_3^*[u_3]
\end{eqnarray*}
which gives
$$(w_1 \circ w_2^{-1} \circ w_3) \in \subvalue{\Sigma;\;\intype{x\;}{\;\sigma}}{\tau[x]}(\rho_1 \circ \rho_2^{-1} \circ \rho_3)[x \leftarrow (u_1 \circ u_2^{-1} \circ u_3)].$$
For the converse consider
$$\mathbf{Pair}(u,w) \in \semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}(\rho_1 \circ \rho_2^{-1} \circ \rho_3).$$
We then have
$$u \in \semvalue{\sigma}(\rho_`1 \circ \rho_2^{-1} \circ \rho_3) = \sigma_1^* \circ {\sigma_2^*}^{-1} \circ \sigma_3^*.$$
Hence there exist $u_i \in \sigma_i^*$ such that $u = u_1 \circ u_2^{-1} \circ u_3$.
We also have
\begin{eqnarray*}
w & \in & \subvalue{\Sigma\;\intype{x\;}{\;\sigma}}{\tau[x]}(\rho_1 \circ \rho_2^{-1} \circ \rho_3)[x \leftarrow u] \\
& = & \subvalue{\Sigma\;\intype{x\;}{\;\sigma}}{\tau[x]}(\rho_1 \circ \rho_2^{-1} \circ \rho_3)[x \leftarrow (u_1 \circ u_2^{-1} \circ u_3)] \\
& = & \subvalue{\Sigma\;\intype{x\;}{\;\sigma}}{\tau[x]}(\rho_1[x \leftarrow u_1] \circ \rho_2[x \leftarrow u_2]^{-1} \circ \rho_3[x \leftarrow u_3]) \\
& = & \tau_1^*[u_1] \circ \tau_2^*[u_2]^{-1} \circ \tau_3^*[u_3]
\end{eqnarray*}
This now implies that there exist $w_i \in \tau_i[u_i]$ with $w = w_1 \circ w_2^{-1} \circ w_3$ which proves (V1) for the pair type.
To show (V2) consider $\rho,\tilde{\rho} \in \convalue{\Sigma}$ with $\rho \preceq \tilde{\rho}$.
We must show
$$\semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}\rho \preceq \semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}\tilde{\rho}.$$
Let $\sigma^*$ abbreviate $\semvalue{\sigma}\rho$ and let $\tilde{\sigma}^*$ abbreviate $\semvalue{\sigma}\tilde{\rho}$. For $u \in \sigma^*$
let $\tau^*[u]$ abbreviate $\subvalue{\Sigma;\intype{x\;}{\;\sigma}}{\tau[x]}\rho[x \leftarrow u]$ and for $\tilde{u} \in \tilde{\sigma}^*$
let $\tilde{\tau}[\tilde{u}]$ abbreviate $\subvalue{\Sigma;\intype{x\;}{\;\sigma}}{\tau[x]}\tilde{\rho}[x \leftarrow \tilde{u}]$.
We will let $u$ range over elements of $\sigma^*$ and $w$ range over elements of $\tau^*[u]$. Similarly we let $\tilde{u}$
range over elements of $\tilde{\sigma}^*$ and $\tilde{w}$ range over elements of $\tilde{\tau}[\tilde{u}]$.
Let $\eta$ be a minimal template for $\tilde{\rho}$
and let ${\cal A} = \mathbf{Mem}(\tempvalue{\sigma}\eta)$ and ${\cal B} = \mathbf{Mem}(\tempvalue{\tau[x]}\eta[x \leftarrow {\cal A}])$.
By theorem~\ref{thm:AllValues} we have that $\mathbf{Pair}({\cal A},{\cal B})$ is an interface template for the pair type.
Note that for any $\tilde{u} \in \tilde{\sigma}^*$ and $\tilde{w} \in \tau^*[\tilde{u}]$
we have
$$\mathbf{Pair}(\tilde{u}@{\cal A},\;\tilde{w}@{\cal B}) \in \semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}\tilde{\rho}.$$
We must show that for any abstract templates ${\cal C}$, ${\cal D}$ such that
$$\semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}\tilde{\rho}\;@\;\mathbf{TypeOf}(\mathbf{Pair}({\cal C},\;{\cal D}))$$
is defined, and any $u \in \sigma^*$ and $w \in \tau^*[u]$ we have that $u@{\cal C}$ and $w@{\cal D}$ are defined and,
furthermore, the set of pairs of the form $\mathbf{Pair}(u@{\cal C},\;w@{\cal D})$ with $u \in \sigma^*$ and $w \in \tau^*[u]$
is the same as the set of pairs of the form $\mathbf{Pair}(\tilde{u}@{\cal C},\;\tilde{w}@{\cal D})$ with $\tilde{u} \in \tilde{\sigma}^*$ and
$\tilde{w} \in \tilde{\tau}^*[\tilde{u}]$.
From property (V2) of the first antecedent we have
$\sigma^* \preceq \tilde{\sigma}^*$ which implies $$\sigma^*@\mathbf{TypeOf}({\cal A}) = \tilde{\sigma}^*@\mathbf{TypeOf}({\cal A}).$$
This implies that for any $u \in \sigma^*$ we have that $u@{\cal A}$ is defined and equal to $\tilde{u}@{\cal A}$ for some $\tilde{u} \in \tilde{\sigma}^*$.
This implies that $u@{\cal A} \in \tilde{\sigma}^*$. Furthermore, $u@{\cal A}@{\cal C} = u@{\cal C}$ is defined.
Also, we have $\rho[x \leftarrow u] \preceq \tilde{\rho}[x \leftarrow u@{\cal A}]$ and by property (V2) of the second antecedent
we have $\tau^*[u] \preceq \tilde{\tau}^*[u@{\cal A}]$ which implies
$$\tau^*[u]@\mathbf{TypeOf}({\cal B}) \;\;=\;\; \tilde{\tau}^*[u@{\cal A}]@\mathbf{TypeOf}({\cal B}).$$
This implies that $w@{\cal B}$ equals $\tilde{w}@{\cal B}$ for some $\tilde{w}$
in $\tilde{\tau}^*[u@{\cal A}]$. This implies that $\tilde{w}@{\cal B}@{\cal D} = w@{\cal B}@{\cal D} = w@{\cal D}$ is defined.
This also shows that every pair of the form $\mathbf{Pair}(u@{\cal C},\;w@{\cal D})$ is equal to a pair of the form $\mathbf{Pair}(\tilde{u}@{\cal C},\;\tilde{w}@{\cal D})$.
It remains only to show the converse. Consider a pair of the form $\mathbf{Pair}(\tilde{u}@{\cal C},\;\tilde{w}@{\cal D})$ with $\tilde{u} \in \tilde{\sigma}^*$
and $\tilde{w} \in \tilde{\tau}^*[\tilde{u}]$. Since $\sigma^*@\mathbf{TypeOf}({\cal A}) = \tilde{\sigma}^*@\mathbf{TypeOf}({\cal A})$
we have that $\tilde{u}@{\cal A}$ equals $u@{\cal A}$ for some $u \in \sigma^*$. As before, we then have $\tau^*[u] \preceq \tilde{\tau}^*[u@{\cal A}]$
which implies that $\tilde{w}@{\cal B}$ equals $w@{\cal B}$ for some $w \in \tau^*[u]$. We then have
\begin{eqnarray*}
\mathbf{Pair}(\tilde{u}@{\cal C},\;\tilde{w}@{\cal D}) & = & \mathbf{Pair}(\tilde{u}@{\cal A}@{\cal C},\;\tilde{w}@{\cal B}@{\cal D}) \\
& = & \mathbf{Pair}(u@{\cal A}@{\cal C},\;w@{\cal B}@{\cal D}) \\
& = & \mathbf{Pair}(u@{\cal C},\;w@{\cal D})
\end{eqnarray*}
\end{proof}
\subsection{The Soundness of Substitution and the Isomorphism Rules}
\label{subsec:Substitution}
\begin{lemma}
The substitution rule
\vspace{-1em}
~ \hfill
\unnamed
{\ant{\Sigma;\;\intype{x}{\sigma} \vdash \intype{e[x]}{\tau}}
\ant{\mbox{$x$ is not free in $\tau$}}
\ant{\Sigma \vdash w =_\sigma u}}
{\ant{\Sigma \vdash e[w] =_\tau e[u]}}
\hfill ~
\vspace{-1em}
is sound
\end{lemma}
\begin{proof}
Consider $\rho \in \convalue{\Sigma}$. Let $\sigma^*$ abbreviate $\semvalue{\sigma}\rho$ and similarly for $\tau^*$, $w^*$ and $u^*$.
For $u \in \sigma^*$ let $e^*[u]$ abbreviate $\subvalue{\Sigma;\intype{x\;}{\;\sigma}}{e[x]}\rho[x \leftarrow u]$.
Let ${\cal A}_\sigma$ be an interface template for $\sigma^*$ and let ${\cal A}_\tau$ be an interface template for $\tau^*$.
We must show that the validity of the antecedents of the rule implies $e^*[w^*] =_{\tau^*} e^*[u^*]$.
We have that $\Sigma \models w=_\sigma u$ implies $w^* \in \sigma^*$ and $u^*\in \sigma^*$ and that there exists $z \in \sigma^*$
such that $(w^*@{\cal A}_\sigma) \circ z^{-1} \circ (u^*@{\cal A}_\sigma)$ is defined.
Condition (V1) on the first antecedent then implies that $e^*[w^*@{\cal A}_\sigma] \circ e^*[z]^{-1} \circ e^*[u^*@{\cal A}_\sigma]$ is defined.
By (Abs-Distributes-In) we then have that
$$e^*[w^*@{\cal A}_\sigma]@{\cal A}_\tau \circ (e^*[z]@{\cal A}_\tau)^{-1} \circ e^*[u^*@{\cal A}_\sigma]@{\cal A}_\tau$$
is defined. By condition (V2) on the first antecedent we have $e^*[w^*] \preceq e^*[w^*@{\cal A}_\sigma]$ and hence $e^*[w^*@{\cal A}_\sigma]@{\cal A}_\tau = e^*[w^*]@{\cal A}_\tau$
and similarly for $u^*$. We now have that
$$e^*[w^*]@{\cal A}_\tau \circ (e^*[z]@{\cal A}_\tau)^{-1} \circ e^*[u^*]@{\cal A}_\tau$$
is defined which implies the result.
\end{proof}
We now consider the isomorphism inference rules in figure~\ref{fig:simple}. The first row of inference rules is the following.
\begin{lemma}
The inference ruls
\hfill ~
\unnamed{
\ant{\Sigma \vdash \intype{\sigma,\tau}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \intype{f}{\mathbf{Bijection}[\sigma,\tau]}}}
{\ant{\Sigma \vdash \inntype{\updownarrow\!(\sigma,\tau,f)}{\mathrm{\bf iso}(\mathrm{\bf type}_i,\sigma,\tau)}}
\ant{}
\ant{\Sigma \vdash \left\{\begin{array}{l}\forall \intype{x}{\sigma}\;\forall\intype{y}{\tau} \\
\;\;(x \leftrightarrowtriangle_{\updownarrow\!(\sigma,\tau,f)} y) \\
\;\;\Leftrightarrow f(x) =_{\tau} y
\end{array}\right.}}
\hfill ~
is sound.
\end{lemma}
\begin{proof}
Consider $\rho \in \convalue{\Sigma}$ and let $\sigma^*, \tau^*$ and $f^*$ be defined as usual.
The soundness of the first conclusion of the first rule follow from property ($\updownarrow$.B).
To show the soundness of the second conclusion consider $u \in \sigma^*$ and $w \in \tau^*$.
First suppose $u \leftrightarrowtriangle_{\updownarrow\!(\sigma^*,\tau^*,f^*)} w$. In this case there exists $p \in \; \updownarrow\!(\sigma^*,\tau^*,f^*)$ such that $(u@\mathrm{\bf Point}) \circ p^{-1} \circ (w@\mathrm{\bf Point})$ is defined.
By the definition of $\updownarrow(\sigma^*,\tau^*,f^*)$ the point $p$ has the form $\mathrm{\bf Point}(\pi_1(w'@\mathrm{\bf Point}),\pi_2(u'@\mathrm{\bf Point}))$
for $u' \in \sigma^*$ and $w' \in \tau^*$ with $f^*(u') =_{\tau^*} w'$. We now have that $(u@\mathrm{\bf Point}) \circ (u'@\mathrm{\bf Point})^{-1}$ is defined
which implies $(u@\mathrm{\bf Point}) \simeq_{\sigma^*@\typeof(\pointt)} (u'@\mathrm{\bf Point})$ which implies $u =_{\sigma^*} u'$.
Similarly we have $w =_{\tau^*} w'$. This gives $f^*(u) = f^*(u') =_{\tau^*} w' =_{\tau^*} w$.
Conversely suppose $w =_{\tau^*} f^*(u)$.
We must show that there exists $p \in \; \updownarrow\!(\sigma^*,\tau^*,f^*)$ such that $(u@\mathrm{\bf Point}) \circ p^{-1} \circ (w@\mathrm{\bf Point})$
is defined. But by the definition of $\updownarrow\!(\sigma^*,\tau^*,f^*)$ we have that $\mathrm{\bf Point}(\pi_1(w@\mathrm{\bf Point}),\pi_2(u@\mathrm{\bf Point})) \in \;\updownarrow\!(\sigma^*,\tau^*,f^*)$.
\end{proof}
\begin{lemma}
The inference rules
{\small
~ \hfill
\unnamed{
\ant{\Sigma \vdash \inntype{a_3}{\mathrm{\bf iso}(\sigma,a_1,a_2)}}
\ant{\Sigma \vdash \inntype{b_3}{\mathrm{\bf iso}(\tau,b_1,b_2)}}}
{\ant{\Sigma \vdash \inntype{\mathbf{Pair}(a_3,b_3)}{\mathrm{\bf iso}\left(\begin{array}{l}\mathbf{PairOf}(\sigma,\tau), \\\mathbf{Pair}(a_1,b_1), \\\mathbf{Pair}(a_2,b_2) \end{array}\right)}}}
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{a,b}{\sigma}}}
{\ant{\Sigma \vdash \left\{\begin{array}{l} a =_\sigma b \\ \;\Leftrightarrow \\ a \leftrightarrowtriangle_\sigma b \end{array}\right.}}
\hfill
\unnamed{
\ant{\Sigma \vdash \inntype{a_3}{\mathrm{\bf iso}(\sigma,a_1,a_2)}}}
{\ant{\Sigma \vdash \inntype{a_3}{\sigma}}}
\hfill ~}
are sound.
\end{lemma}
\begin{proof}
The soundness of these rules follows immediately from the definitions involved.
\end{proof}
\begin{lemma}
The inference rule
~ \hfill
\unnamed{
\ant{\Sigma \vdash \intype{\mathbf{Pair}(a_1,b_1),\mathbf{Pair}(a_2,b_2)}{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}}
\ant{\Sigma \vdash \inntype{a_3}{\mathrm{\bf iso}(\sigma,a_1,a_2)}}
\ant{\Sigma \vdash b_1 \leftrightarrowtriangle_{\tau[a_3]} b_2}
}{
\ant{\Sigma \vdash \mathbf{Pair}(a_1,b_1) =_{\mathbf{PairOf}(\intype{x\;}{\;\sigma},\;\intype{y\;}{\;\tau[x]})} \mathbf{Pair}(a_2,b_2)}
}
\hfill ~
is sound.
\end{lemma}
\begin{proof}
Consider $\rho \in \convalue{\Sigma}$ and let $a_i^*$, $b_i^*$, $\sigma^*$ and $\tau^*[u]$ for $u \in \sigma^*$ be defined as usual.
Let $\mathbf{PairOf}(\intype{x}{\sigma^*},\;\intype{y}{\tau^*[x]})$ abbreviate $\semvalue{\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})}\rho$.
Let ${\cal A}_\sigma$ be an interface template for $\sigma^*$, let ${\cal A}_\tau$ be an interface template for $\tau^*[a_3^*]$,
and let ${\cal A}_{\tilde{\tau}}$ be an interface template for $\tau^*[a_3^*@{\cal A}_\sigma]$. Theorem~\ref{thm:AllValues} implies that $\mathbf{Pair}({\cal A}_\sigma,{\cal A}_{\tilde{\tau}})$
is an interface for $\mathbf{PairOf}(\intype{x}{\sigma^*},\;\intype{y}{\tau^*[x]})$.
The validity of the third antecedent implies that there exists $b_3 \in \tau^*[a_3^*]$ such that $(b_1^*@{\cal A}_\tau) \circ b_3^{-1} \circ (b_2@{\cal A}_\tau)$
is defined. This immediately gives $\mathbf{Pair}(a_3^*,b_3) \in \mathbf{PairOf}(\intype{x}{\sigma^*},\;\intype{y}{\tau^*[x]})$
and that
$$(\mathbf{Pair}(a_1,b_1)@\mathbf{Pair}({\cal A}_\sigma,{\cal A}_\tau)) \circ \mathbf{Pair}(a_3^*,b_3^*)^{-1} \circ (\mathbf{Pair}(a_1,b_1)@\mathbf{Pair}({\cal A}_\sigma,{\cal A}_\tau))$$
is defined. By (Abs-Distributes-In) and (Abs-Compression) we then have that
$$(\mathbf{Pair}(a_1,b_1)@\mathbf{Pair}({\cal A}_\sigma,{\cal A}_{\tilde{\tau}})) \circ (\mathbf{Pair}(a_3^*,b_3^*)@\mathbf{Pair}({\cal A}_\sigma,{\cal A}_{\tilde{\tau}}))^{-1}
\circ (\mathbf{Pair}(a_1,b_1)@\mathbf{Pair}({\cal A}_\sigma,{\cal A}_{\tilde{\tau}}))$$
is defined which implies the lemma.
\end{proof}
\begin{lemma}
The rule
~\hfill
\unnamed
{\ant{\Sigma \vdash \intype{a_1}{\sigma},\;\intype{a_2}{\sigma},\;\inntype{a_3}{\mathrm{\bf iso}(\sigma,a_1,a_2)}}
\ant{\Sigma;\;\intype{x}{\sigma} \vdash \intype{\mathbf{PairOf}(\tau_1[x],\;\tau_2[x])}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \intype{b_1}{\mathbf{PairOf}(\tau_1[a_1],\;\tau_2[a_1])}}
\ant{\Sigma \vdash \intype{b_2}{\mathbf{PairOf}(\tau_1[a_2],\;\tau_2[a_2])}}}
{\ant{\Sigma \vdash \left\{\begin{array}{l} (b_1 \leftrightarrowtriangle_{\mathbf{PairOf}(\tau_1[a_3],\;\tau_2[a_3])} b_2) \\
\;\;\;\Leftrightarrow \\
\pi_1(b_1) \leftrightarrowtriangle_{\tau_1[a_3]} \pi_1(b_2) \;\wedge \\
\pi_2(b_1) \leftrightarrowtriangle_{\tau_2[a_3]} \pi_2(b_2)
\end{array}\right.}}
\hfill ~
is sound
\end{lemma}
\begin{proof}
Consider $\rho \in \convalue{\Sigma}$ and let $a_i^*$, $b_i^*$, $\sigma^*$ and $\tau_i^*[u]$ for $u \in \sigma^*$ be defined as usual.
Let ${\cal A}_1$ be an interface template for $\tau_1^*[a_3^*]$ and let ${\cal A}_2$ be an interface template for
$\tau_2^*[a_3^*]$. We have that $\mathbf{Pair}({\cal A}_1,{\cal A}_2)$ is an interface template for
$\mathbf{PairOf}(\tau_1^*[a_3^*],\;\tau_2^*[a_3^*])$.
First suppose that $b_1^* \leftrightarrowtriangle_{\mathbf{PairOf}(\tau_1^*[a_3^*],\;\tau_2^*[a_3^*])} b_2^*$. This implies that there exists $b_3 \in \mathbf{PairOf}(\tau_1^*[a_3^*],\;\tau_2^*[a_3^*])$ with
$(b_1^*@\mathbf{Pair}({\cal A}_1,{\cal A}_2)) \circ b_3^{-1} \circ (b_2^*@\mathbf{Pair}({\cal A}_1,{\cal A}_2))$ defined. This implies that $(\pi_1(b_1^*)@{\cal A}_1) \circ \pi_1(b_3)^{-1} \circ \pi_1(b_2)@{\cal A}_1)$
is defined which implies $\pi_1(b_1^*) \leftrightarrowtriangle{\tau_1^*[a_3]} \pi_1(b_2^*)$. The case of $\pi_2$ is similar.
Conversely suppose that $\pi_1(b_1^*) \leftrightarrowtriangle{\tau_1^*[a_3]} \pi_1(b_2^*)$ and $\pi_2(b_1^*) \leftrightarrowtriangle{\tau_2^*[a_3]} \pi_2(b_2^*)$.
In this case it is straightforward to show that there exist a pair $b_3 \in \mathbf{Pairf}(\tau_1^*[a_3^*],\tau_2^*[a_3^*])$
with $(b_1^*@\mathbf{Pair}({\cal A}_1,{\cal A}_2)) \circ b_3^{-1} \circ (b_2^*@\mathbf{Pair}({\cal A}_1,{\cal A}_2))$ defined.
\end{proof}
\begin{lemma}
The inference rule
~ \hfill
\unnamed{
\ant{\Sigma \vdash \intype{a_1}{\sigma},\;\intype{a_2}{\sigma},\; \inntype{a_3}{\mathrm{\bf iso}(\sigma,a_1,a_2)}}
\ant{\Sigma; \intype{x}{\sigma} \vdash \intype{(\tau_1[x] \rightarrow \tau_2[x])}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \intype{f_1}{(\tau_1[a_1] \rightarrow \tau_2[a_1])}}
\ant{\Sigma \vdash \intype{f_2}{(\tau_1[a_2] \rightarrow \tau_2[a_2])}}}
{\ant{\Sigma \vdash \left\{\begin{array}{l} (f_1 \leftrightarrowtriangle_{\tau_1[a_3] \rightarrow \tau_2[a_3]} f_2) \\
\;\;\;\Leftrightarrow \\
\forall \intype{x_1}{\tau_1[a_1]}\; \;\forall \intype{x_2}{\tau_1[a_2]} \\
\;\;(x_1 \leftrightarrowtriangle_{\tau_1[a_3]} x_2) \\
\;\;\;\;\;\;\;\Rightarrow \\
\;\;f_1(x_1) \leftrightarrowtriangle_{\tau_2[a_3]} f_2(x_2)
\end{array}\right.}}
\hfill ~
is sound.
\end{lemma}
\begin{proof}
Consider $\rho \in \convalue{\Sigma}$. Let $a_i^*$ abbreviate $\semvalue{a_i}\rho$ and similarly for $f_i^*$ and $\sigma^*$.
For $u \in \sigma^*$ let $\tau_i^*[u]$ abbreviate $\semvalue{\tau_i[x]}\rho[x \leftarrow u]$. Let ${\cal A}_\sigma$ be an interface template for
$\sigma^*$, let ${\cal A}_1$ be an interface template for $\tau_1^*[a_3^*]$ and let ${\cal A}_2$ be an interface template for $\tau_2^*[a_3^*]$.
First suppose that $f_1^* \leftrightarrowtriangle_{(\tau_1^*[a_3^*] \rightarrow \tau_2^*[a_3^*])} f_2^*$
and consider $x_1 \in \tau_1^*[a_1^*]$ and $x_2 \in \tau_1^*[a_2^*]$ with $x_1 \leftrightarrowtriangle_{\tau_1^*[a_3^*]} x_2$.
We must show $f_1^*(x_1) \leftrightarrowtriangle_{\tau_2^*[a_3^*]} f_2^*(x_2)$.
By the definition of $\leftrightarrowtriangle_{\tau_1^*[a_3^*]}$ there exists $x_3 \in \tau_1^*[a_3^*]$ with
$(x_1@{\cal A}_1) \circ x_3^{-1} \circ (x_2@{\cal A}_1)$ defined.
By (Abs-Distributes-In) and (Abs-Compression) we then have that
$(x_1@\mathrm{\bf Point}) \circ (x_3@\mathrm{\bf Point})^{-1} \circ (x_2@\mathrm{\bf Point})$ is defined.
By the definition of $\leftrightarrowtriangle_{(\tau_1^*[a_3^*] \rightarrow \tau_2^*[a_3^*])}$ there exists $f_3 \in \tau_1^*[a_3^*] \rightarrow \tau_2^*[a_3^*]$ with
$$g \equiv (f_1@(\mathrm{\bf Point} \rightarrow {\cal A})) \circ {f_3^*}^{-1} \circ (f_2@(\mathrm{\bf Point} \rightarrow {\cal A}))$$
defined where ${\cal A}$ is any interface template of $\tau_2^*[a_3^*]$. By (Funs-Composable) we have
$$g[(x_1@\mathrm{\bf Point}) \circ (x_3@\mathrm{\bf Point})^{-1} \circ (x_2@\mathrm{\bf Point})] = (f_1(x_1)@{\cal A}) \circ f_3^*(x_3)^{-1} \circ (f_2(x_2)@{\cal A})$$
The definedness of the right hand side gives $f_3(x_3) \in \mathrm{\bf iso}(\tau_2^*[x_3],f_1^*(x_1),f_2^*(x_2))$
which yields $f_1^*(x_1) \leftrightarrowtriangle_{\tau_2^*[a_3^*]} f_2^*(x_2)$ as desired.
Now suppose that for all
$x_1 \in \tau_1^*[a_1^*]$ and $x_2 \in \tau_1^*[a_2^*]$ with $x_1 \leftrightarrowtriangle_{\tau_1^*[a_3^*]} x_2$ we have $f_1^*(x_1) \leftrightarrowtriangle_{\tau_2^*[a_3^*]} f_2^*(x_2)$.
We must show $f_1^* \leftrightarrowtriangle_{(\tau_1^*[a_3^*] \rightarrow \tau_2^*[a_3^*])} f_2^*$. We must show that an appropriate witness $f_3 \in \tau_1^*[a_3^*] \rightarrow \tau_2^*[a_3^*]$
exists. For $x_3 \in \tau_1^*[a_3^*]$ we must define $f_3(x_3)$. By the first antecedent of the rule we have that
$(a_1^*@{\cal A}_\sigma) \circ {a_3^*}^{-1} \circ (a_2^*@{\cal A}_\sigma)$
is defined. The validity of the second antecedent implies $\Sigma;\intype{x}{\sigma} \models \intype{\tau_1[x]}{\mathrm{\bf type}_i}$.
Property (V1) of this entailment yields that
$$(1)\;\;\;\tau_1^*[a_1^*@{\cal A}_\sigma] \circ \tau_1^*[a_3^*]^{-1} \circ \tau_1^*[a_2^*@{\cal A}_\sigma]$$
is defined. We have that $\mathbf{TypeOf}({\cal A}_1)$
is a minimal template for $\tau_1^*[a_3^*]$ and by the properties of $\sim$ we can abstract the above composition to $\mathbf{TypeOf}({\cal A})$.
By (Abs-Distributes-In) and (Internal-Compression) we then have that
$$(\tau_1^*[a_1^*]@\mathbf{TypeOf}({\cal A})) \circ (\tau_1^*[a_3^*]@\mathbf{TypeOf}({\cal A}))^{-1} \circ (\tau_1^*[a_2^*]@\mathbf{TypeOf}({\cal A}))$$
is defined.
By (Partner) we then have that there exists $x_1 \in \tau_1^*[a_1^*]$ and $x_2 \in \tau_1^*[a_2^*]$
with $(x_1@{\cal A}) \circ (x_3@{\cal A})^{-1} \circ (x_2@{\cal A})$ defined.
We then have $x_1 \leftrightarrowtriangle_{\tau_1[a_3^*]} x_2$ which by assumption implies $f_1(x_1) \leftrightarrowtriangle_{\tau_2^*[a_3^*]} f_2(x_2)$.
For each equivalence class ${\cal C}$ of $\tau_1^*[a_3^*]$ we can select a value $y({\cal C}) \in \tau_2^*[a_3^*]$ such that there exists
$x_1 \in \tau_1^*[a_1^*]$, $x_2 \in \tau_1^*[a_2^*]$ and $x_3 \in {\cal C}$ with $(x_1@{\cal A}_\sigma)\circ (x_3@{\cal A}_\sigma)^{-1} \circ (x_2@{\cal A}_\sigma)$
defined and with $y({\cal C}) \in \mathrm{\bf iso}(\tau_2^*[a_3^*],x_1,x_2)$.
We can then define $f_3(x_3)$ to be $y(|x_3|)$ which gives $f_3 \in \tau_1^*[a_3^*] \rightarrow \tau_2^*[a_3^*]$. It remains to show
$f_3 \in \mathrm{\bf iso}(\tau_1^*[a_3^*] \rightarrow \tau_2^*[a_3^*],f_1,f_2)$. In particular we must show that
$$(f_1@(\mathrm{\bf Point} \rightarrow {\cal A})) \circ {f_3^*}^{-1} \circ (f_2@(\mathrm{\bf Point} \rightarrow {\cal A}))$$
is defined where ${\cal A}$ is the interace template for $\tau_2^*[a_3^*]$. By (1) above and (Abs-Distributes-In)
we get that $\mathbf{Dom}(f_1) \circ domop(f_3)^{-1} \circ \mathbf{Dom}(f_2)$ is defined. It remains only to show that for $x_i \in \mathbf{Dom}(f_i)$
with $x_1 \circ x_2^{-1} \circ x_3$ defined
we have that $(f_1[x_1]@{\cal A}_2) \circ f_3[x_3]^{-1} \circ (f_2[x_2]@{\cal A}_2)$ is defined. By the definition of $f_3$
there exists $x_i' \in \mathbf{Dom}(f_i)$ with $x_3' \simeq_{\mathbf{Dom}(f_3)} x_3$ and with $x_1' \circ x_3'^{-1} \circ x_2'$ defined
and $(f_1[x_1']@{\cal A}_2) \circ f_3[x_3']^{-1} \circ (f_2[x'_2]@{\cal A}_2)$ defined. By ($\simeq$.B) we have $x_i \simeq_{\mathbf{Dom}(f_1)} x_i'$
which implies that $f_i[x_i] = f_i[x_i']$ which now implies the result.
\end{proof}
\begin{lemma}
The inference rule
~ \hfill
\unnamed{
\ant{\Sigma \vdash \intype{a_1}{\sigma},\;\intype{a_2}{\sigma},\;\inntype{a_3}{\mathrm{\bf iso}(\sigma,a_1,a_2)}}
\ant{\Sigma; \intype{x}{\sigma} \vdash \intype{\mathbf{SubType}(\intype{y}{\tau[x]},\;\Phi[x,y])}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \intype{b_1}{\mathbf{SubType}(\intype{y}{\tau[a_1]},\;\Phi[a_1,y])}}
\ant{\Sigma \vdash \intype{b_2}{\mathbf{SubType}(\intype{y}{\tau[a_2]},\;\Phi[a_2,y])}}}
{\ant{\Sigma \vdash \left\{\begin{array}{l} (b_1 \leftrightarrowtriangle_{\mathbf{SubType}(\intype{y\;}{\;\tau[a_3]},\;\Phi[a_3,y])} b_2) \\ \;\;\;\Leftrightarrow \\ (b_1 \leftrightarrowtriangle_{\tau[a_3]} b_2)\end{array}\right.}}
\hfill ~
is sound.
\end{lemma}
\begin{proof}
Consider $\rho \in \convalue{\Sigma}$ and let $a_i^*$, $b_i^*$, $\sigma^*$, $\tau^*[u]$ for $u \in \sigma^*$, and $\Phi^*[u,w]$ for $u \in \sigma^*$ and $w \in \tau^*[u]$
be defined in the standard way. For $u \in \sigma^*$ let $\mathbf{Subtype}(\intype{y}{\tau^*[u]},\;\Phi^*[u,y])$ abbreviate
$\semvalue{\mathbf{Subtype}(\intype{y}{\tau[x]},\;\Phi[x,y])}\rho[x \leftarrow u]$. Let ${\cal A}_\tau$ be an interface template for $\tau^*[a_3^*]$.
First suppose that $b_1^* \leftrightarrowtriangle_{\mathbf{Subtype}(\intype{y\;}{\;\tau^*[a_3^*]},\;\Phi^*[u,y])} b_2^*$. This implies that there exists
$b_3 \in \tau^*[a_3^*]$ with $(b_1^*@{\cal A}_\tau) \circ b_3^{-1} \circ (b_2^*@{\cal A}_\tau)$ defined which yields
$b_1^* \leftrightarrowtriangle_{\tau^*[a_3^*]} b_2^*$. Conversely suppose that $b_1^* \leftrightarrowtriangle_{\tau^*[a_3^*]} b_2^*$. In this case there exists
$b_3 \in \tau^*[a_3^*]$ with $(b_1^*@{\cal A}_\tau) \circ b_3^{-1} \circ (b_2^*@{\cal A}_\tau)$ defined.
By the validity of the second antecedent we have $\Sigma;\;\intype{x}{\sigma};\;\intype{y}{\tau[x]} \models \intype{\Phi[x,y]}{\mathrm{\bf Bool}}$.
By property (V1) of this entailment we now have $\Phi^*[a_1^*,b_1^*] = \Phi^*[a_3^*,b_3] = \Phi^*[a_2^*,b_2^*]$.
This gives $b_3 \in \mathbf{Subtype}(\intype{y}{\tau^*[a_3^*]},\;\Phi^*[a_3^*,y])$ which proves the result.
\end{proof}
It is worth noting that we seem {\em unable} to prove the soundness of the following apparently natural rule.
~\hfill
\unnamed{
\ant{\Sigma \vdash \intype{a_1}{\sigma},\;\intype{a_1}{\sigma},\;\inntype{a_3}{\mathrm{\bf iso}(\sigma,a_1,a_2)}}
\ant{\Sigma;\intype{x}{\sigma} \vdash \intype{e[x]}{\tau[x]}}}
{\ant{\Sigma \vdash \inntype{e[a_3]}{\mathrm{\bf iso}(\tau[a_3],e[a_1],e[a_2])}}}
\hfill ~
\noindent We can consider two possible definitions of $\mathrm{\bf iso}(\sigma,x,y)$.
\begin{itemize}
\item[(1)] $z \in \mathrm{\bf iso}(\sigma,x,y)$ iff $(x@\sigma) \circ z^{-1} \circ (y@\sigma)$ is defined.
\item[(2)] $z \in \mathrm{\bf iso}(\sigma,x,y)$ iff $(x@\sigma) \circ (z@\sigma)^{-1} \circ (y@\sigma)$ is defined.
\end{itemize}
Under definition (1) we get
$e^*[a_3^*]@\tau^*[a_3^*] \in \mathrm{\bf iso}(\tau^*[a_3],e^*[a_1^*],e^*[a_2^*])$ for the conclusion and the conclusion fails to satisfy (1).
Under definition (2) we get
$e^*[a_3^*] \in \mathrm{\bf iso}(\tau^*[a_3^*@\sigma],e^*[a_1^*],e^*[a_2^*])$ for the conclusion. The approach taken here avoids this rule.
\subsection{Soundness of the Remaining Rules}
\label{subsec:remaining}
We now turn to proving the soundness of the inference rules in figures~\ref{fig:expressions} through~\ref{fig:jrules}
other than pair type formulation and substitution have been handled above. We consider each figure in turn.
\subsubsection{Proofs for figure~\ref{fig:expressions}}
Figure~\ref{fig:expressions} consists primarily of expression formation rules with conclusions of the form $\Sigma \vdash \intype{e}{\sigma}$.
For each such conclusion we must check that conditions (V1) and (V2) of figure~\ref{fig:models} hold. In addition to expression formation
rules, figure~\ref{fig:expressions} also includes various miscellaneous housekeeping rules. We prove soundness for the rules in the order in which they appear in the figure.
\begin{lemma}
The rules
{\footnotesize
~ \hfill
$\epsilon \vdash \mathrm{\bf True}$
\hfill
$\begin{array}{l} \epsilon \vdash \intype{\mathrm{\bf type}_j}{\mathrm{\bf type}_i} \\
\mbox{for $j < i$}
\end{array}$
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{\tau}{\mathrm{\bf type}_i}}
\ant{\mbox{$x$ not declared in $\Sigma$}}}
{\ant{\Sigma;\;\intype{x}{\tau} \vdash \mathrm{\bf True}}}
\hfill
$\epsilon \vdash \intype{\mathrm{\bf Bool}}{\mathbf{Set}}$
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{\Phi}{\mathrm{\bf Bool}}}}
{\ant{\Sigma;\Phi \vdash \mathrm{\bf True}}}
\hfill ~
}
are sound.
\end{lemma}
\begin{proof}
We will take $\mathbf{True}$ to be an abbreviation for
$\exists\intype{P}{\mathrm{\bf Bool}}$. We then have $\convalue{\mathrm{\bf True}} = \mathrm{\bf True}$ where the left occurrence of $\mathrm{\bf True}$ is an expression and the right occurrence of $\mathrm{\bf True}$
is a morphoid value (a Boolean value). We then have $\epsilon \models \mathrm{\bf True}$ which establishes the soundness of the first rule above.
For the soundness of the second rule we must show $\epsilon \models \intype{\mathrm{\bf type}_j}{\mathrm{\bf type}_i}$ for $j < i$. But this follows from the definition of $\convalue{\mathrm{\bf type}_i}$
in figure~\ref{fig:value} and theorem~\ref{thm:AllValues} which implies that $\convalue{\mathrm{\bf type}_j}$ is a morphoid type. This sequent satisfies (V1) and (V2) because
$\mathrm{\bf type}_i$ is closed and $\semvalue{\mathrm{\bf type}}\rho = \convalue{\mathrm{\bf type}_i}$.
For the third rule we must show that the definedness of $\convalue{\Sigma}$ implies the definedness of $\convalue{\Sigma;\;\intype{x}{\tau}}$ given that $x$ is not already in $\Sigma$
and $\Sigma \models \inntype{\tau}{\mathrm{\bf type}_i}$. But this follows directly from the definition of $\convalue{\Sigma}$. The soundness of the fourth and fifth rule are similarly straightforward.
\end{proof}
\begin{lemma}
The rules
~ \hfill
\unnamed
{\ant{\Sigma;\;\Theta \vdash \mathrm{\bf True}}}
{\ant{\Sigma;\;\Theta \vdash \Theta}}
\hfill
\unnamed
{\ant{\Sigma;\;\Theta \vdash \mathrm{\bf True}}
\ant{\Sigma \vdash \Psi}}
{\ant{\Sigma;\;\Theta \vdash \Psi}}
\hfill ~
are sound.
\end{lemma}
\begin{proof}
For the first rule it is possible that $\Theta$ has the form $\intype{x}{\sigma}$. In this case
we must show that conditions (V1) and (V2) are satisfied. But one can check that (V1) and (V2) are immediately satisfied for variables. For the second rule
it is possible that $\Psi$ has the form $\intype{e}{\sigma}$ where $e$ is not a variable. In this case (V1) and (V2) follow from the fact
that for judgments not involving a variable declared in $\Theta$ we can treat $\convalue{\Sigma,\Theta}$ as a subset of $\convalue{\Sigma}$.
One can readily show that (V1) and (V2) are monotone in the sense that if they are satisfied by a set $S$ of structures then they are satisfied by any subset of $S$.
\end{proof}
\begin{lemma}
The rule
\vspace{-1em}
~ \hfill
\unnamed
{\ant{\Sigma \vdash \intype{\sigma}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \intype{\tau}{\mathrm{\bf type}_i}}}
{\ant{\Sigma \vdash \intype{(\sigma \rightarrow \tau)}{\mathrm{\bf type}_i}}}
\hfill ~
\vspace{-1em}
is sound.
\end{lemma}
\begin{proof}
We prove conditions (V1) and (V2) for the conclusion. To prove (V1) consider $\rho_1,\rho_2,\rho_3 \in \convalue{\Sigma}$ with $\rho_1 \circ \rho_2^{-1} \circ \rho_3$
defined and $(\rho_1 \circ \rho_2^{-1} \circ \rho_3) \in \convalue{\Sigma}$. Let $\sigma_i^*$ abbrebiate $\semvalue{\sigma}\rho_i$ and let $\tau^*_i$ abbreviate
$\semvalue{\tau}\rho_i$.
By condition (V1) on the premises and property ($\rightarrow$.V1) we have the following calculation establishing (V1) for the conclusion.
\begin{eqnarray*}
\semvalue{\sigma\rightarrow \tau}(\rho_1 \circ \rho_2^{-1} \circ \rho_3) & = & \semvalue{\sigma}(\rho_1 \circ \rho_2^{-1} \circ \rho_3) \;\rightarrow \;
\semvalue{\tau}(\rho_1 \circ \rho_2^{-1} \circ \rho_3) \\
& = & (\sigma_1^* \circ (\sigma_2^*)^{-1} \circ \sigma_3^*) \rightarrow (\tau_1^* \circ (\tau_2^*)^{-1} \circ \tau_3^*) \\
& = & (\sigma_1^* \rightarrow \tau_1^*) \circ ((\sigma_2^*)^{-1} \rightarrow (\tau_2^*)^{-1}) \circ (\sigma_3^* \rightarrow \tau_3^*) \\
& = & (\sigma_1^* \rightarrow \tau_1^*) \circ (\sigma_2^* \rightarrow \tau_2^*)^{-1} \circ (\sigma_3^* \rightarrow \tau_3^*)
\end{eqnarray*}
To prove (V2) consider $\rho,\tilde{\rho} \in \convalue{\Sigma}$ with $\rho \preceq \tilde{\rho}$.
From (V2) of the antecedents we have $\semvalue{\sigma} \preceq \semvalue{\sigma}\tilde{\rho}$ and $\semvalue{\tau}\rho \preceq \semvalue{\tau}\tilde{\rho}$
and by ($\rightarrow$.V2) we then have
$\semvalue{\sigma \rightarrow \tau}\rho \preceq \semvalue{\sigma \rightarrow \tau}\tilde{\rho}$.
\end{proof}
\begin{lemma} The rule
\vspace{-1em}
~ \hfill
\unnamed
{\ant{\Sigma \vdash \intype{f}{\sigma \rightarrow \tau}}
\ant{\Sigma \vdash \intype{e}{\sigma}}}
{\ant{\Sigma \vdash \intype{f(e)}{\tau}}}
\hfill ~
\vspace{-1em}
is sound.
\end{lemma}
\begin{proof}
We show properties (V1) and (V2) of the conclusion.
To show property (V1) consider $\rho_1,\rho_2,\rho_3 \in \convalue{\Sigma}$ with $\rho_1 \circ \rho_2^{-1} \circ \rho_3$ defined
and $(\rho_1 \circ \rho_2^{-1} \circ \rho_3) \in \convalue{\Sigma}$.
Let $\sigma_i^*$ abbreviate $\semvalue{\sigma}\rho_i$, let
$f_i^*$ abbreviate $\semvalue{f}\rho_i$ and let $e_i^*$ abbreviate
$\semvalue{e}\rho_i$. By condition (V1) of the antecedents and property (Fun-Composition) we have the following calculation which proves (V1).
\begin{eqnarray*}
\semvalue{f(e)}(\rho_1 \circ \rho_2^{-1} \circ \rho_3) & = & (f_1^* \circ (f_2^*)^{-1} \circ f_3^*)(e_1^* \circ (e_2^*)^{-1} \circ e_3^*) \\
& = & f_1^*(e_1^*) \circ f_2^*(e_2^*)^{-1} \circ f_3^*(e_3^*)
\end{eqnarray*}
Property (V2) of the conclusion follows directly from property (V2) of the antecedents and (App.V2).
\end{proof}
\begin{lemma}[Bool.V1]
\label{lem:Bool.V1}
If $\sigma \models \inntype{\Phi}{\mathrm{\bf Bool}}$ then condition (V1) for $\Sigma \models \intype{\Phi}{\mathrm{\bf Bool}}$ is equivalent to the condition
that for $\rho_1,\rho_2 \in \convalue{\Sigma}$ with $\rho_1 \circ \rho_2$ defined
we have that $\semvalue{\Phi}\rho_1 = \semvalue{\Phi}\rho_2$.
\end{lemma}
\begin{proof}
First suppose that condition (V1) holds and consider $\rho_1,\rho_2 \in \convalue{\Sigma}$ with $\rho_1 \circ \rho_2$ defined.
We have that $\rho_1 \circ \rho_2 \circ \rho_2^{-1}$ is defined and equals $\rho_1$ and hence is in $\convalue{\Sigma}$.
By (V1) we then have
\begin{eqnarray*}
\semvalue{\Phi}(\rho_1 \circ (\rho_2^{-1})^{-1} \circ \rho_2^{-1}) & = & \semvalue{\Phi}\rho_1 \circ \semvalue{\Phi}\rho_2 \circ \semvalue{\Phi}\rho_2^{-1} \\
& = & \semvalue{\Phi}\rho_1 \\
& = & \semvalue{\Phi}\rho_2
\end{eqnarray*}
Conversely, suppose that for all $\rho_1,\rho_2 \in \convalue{\Sigma}$ with $\rho_1 \circ \rho_2$ defined we have
$\semvalue{\Phi}\rho_1 = \semvalue{\Phi}\rho_2$. Now consider $\rho_1,\rho_2,\rho_3 \in \convalue{\sigma}$ with
$\rho_1 \circ \rho_2^{-1} \circ \rho_3$ defined and with $(\rho_1 \circ \rho_2^{-1} \circ \rho_3) \in \convalue{\Sigma}$.
Let $\Phi_i^*$ abbreviate $\semvalue{\Phi}\rho_i$.
We immediately have $\Phi_1^* = \Phi_2^* = \Phi_3^*$ and we have that $\Phi_1^* \circ (\Phi_2^*)^{-1} \circ \Phi_3^*$ is defined and is equal to
$\Phi_1^*$. But we also have that $\rho_1^{-1} \circ (\rho_1 \circ \rho_2^{-1} \circ \rho_3)$ is defined and so
$\semvalue{\Phi}(\rho_1 \circ \rho_2^{-1} \circ \rho_3) = \semvalue{\Phi}\rho_1$.
We now have $\semvalue{\Phi}(\rho_1 \circ \rho_2^{-1} \circ \rho_3) = \Phi_1^* = \Phi_1^* \circ (\Phi_2^*)^{-1} \circ \Phi_3^*$ which proves
the result.
\end{proof}
\begin{lemma}[Bool.V2]
\label{Bool.V2}
If $\sigma \models \inntype{\Phi}{\mathrm{\bf Bool}}$ then condition (V2) for $\Sigma \models \intype{\Phi}{\mathrm{\bf Bool}}$ is equivalent to the condition
that for $\rho_1,\rho_2 \in \convalue{\Sigma}$ with $\rho_1 \preceq \rho_2$
we have that $\semvalue{\Phi}\rho_1 = \semvalue{\Phi}\rho_2$.
\end{lemma}
\begin{proof}
This follows from the observation that for Boolean value $P$ and $Q$ we have that $P \preceq Q$ is equivalent to $P = Q$.
\end{proof}
\begin{lemma} The rule
\vspace{-1em}
~ \hfill
\unnamed{\ant{\Sigma \vdash \intype{\Phi}{\mathrm{\bf Bool}}}
\ant{\Sigma \vdash \intype{\Psi}{\mathrm{\bf Bool}}}}
{\ant{\Sigma \vdash \intype{(\Phi \vee \Psi)}{\mathrm{\bf Bool}}}
\ant{\Sigma \vdash \intype{\neg \Phi}{\mathrm{\bf Bool}}}}
\hfill ~
\vspace{-1em}
is sound.
\end{lemma}
\begin{proof}
We consider conditions (V1) and (V2) for the conclusions. For (V1) we have that (Bool.V1)
implies that it is sufficient to consider $\rho_1,\rho_2 \in \convalue{\Sigma}$ with $\rho_1 \circ \rho_2$ defined.
Let $\Phi_i^*$ abbreviate $\semvalue{\Phi}\rho_i$ and let $\Psi_i^*$ abbreviate $\semvalue{\Psi}\rho_i$.
By property (Bool.V1) it suffices to show that $\Phi_1^* \vee \Psi_1^*$ equals $\Phi_2^* \vee \Psi_2^*$. But the
induction hypothesis for the antecedents and property (Bool.V1) give that $\Phi_1^* = \Phi_2^*$ and $\Psi_1^* = \Psi_2^*$.
The case of negation, and the argument for (V2), are similar.
\end{proof}
\begin{lemma}
The rule
\vspace{-1em}
~ \hfill
\unnamed
{\ant{\Sigma \vdash \intype{\sigma}{\mathrm{\bf type}_i}}
\ant{\Sigma;\;\intype{x}{\sigma} \vdash \intype{\Phi[x]}{\mathrm{\bf Bool}}}}
{\ant{\Sigma \vdash \intype{(\forall\intype{x}{\sigma}\;\Phi[x])}{\mathrm{\bf Bool}}}}
\hfill ~
\vspace{-1em}
is sound.
\end{lemma}
\begin{proof}
We show (V1) and (V2) for the conclusion. We note that for $\rho \in \convalue{\Sigma}$ we have that
$\semvalue{\forall \intype{x}{\sigma}\;\Phi[x]}\rho = \mathrm{\bf True}$ if and only if the type $\semvalue{\mathbf{SubType}(\intype{x}{\sigma},\;\neg\Phi[x])}\rho$ is empty.
Properties (V1) and (V2) for $\Sigma \models \intype{(\forall \intype{x}{\sigma}\;\Phi[x])}{\mathrm{\bf Bool}}$ now follow from properties (V1) and (V2)
of $\Sigma \models \intype{\mathbf{SubType}(\intype{x}{\sigma},\;\neg\Phi[x])}{\mathrm{\bf type}_i}$. We omit the details.
\end{proof}
\begin{lemma}
The rule
\vspace{-1em}
~ \hfill
\unnamed
{\ant{\Sigma \vdash \intype{\tau}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \intype{e}{\tau}}
\ant{\Sigma \vdash \intype{w}{\tau}}}
{\ant{\Sigma \vdash \intype{(e =_\tau w)}{\mathrm{\bf Bool}}}}
\hfill ~
\vspace{-1em}
is sound.
\end{lemma}
\begin{proof}
We show (V1) and (V2) for the conclusion. For (V1) we have that (Bool.V1) implies that it suffices
to consider $\rho_1,\rho_2 \in \convalue{\tau}$ with $\rho_1 \circ \rho_2$ defined and
show $\semvalue{e =_\tau w}\rho_1 = \semvalue{e =_\tau w}\rho_2$. Let $e_i^*$ abbreviate $\semvalue{e}\rho_i$
and similarly for $w_i^*$ and $\tau_i^*$. We must show that $e_1^* =_{\tau_1^*} w_1^*$ if and only if
$e_2^* =_{\tau_2^*} w_2^*$. By condition (V1) and property (Corollary.V1) (lemma~\ref{lem:Corollary.V1}) applied to
the antecedents we have that $\tau_1^* \circ \tau_2^*$, $e_1^* \circ e_2^*$ and $w_1^* \circ w_2^*$ are defined.
The result now follows from property ($=$.V1).
Now we show (V2). Consider $\rho_1,\rho_2 \in \convalue{\Sigma}$ with $\rho_1 \preceq \rho_2$.
We must show $\semvalue{e =_\tau w}\rho_1 = \semvalue{e =_\tau w}\rho_2$.
Let $e_i^*$, $w_i^*$ and $\tau_i^*$ be defined as before.
Property (V2) of the antecedents we have
$\tau_1^* \preceq \tau_2^*$, $e_1^* \preceq e_2^*$ and $w_1^* \preceq w_2^*$. The result is then implied by
($=$.V2).
\end{proof}
\begin{lemma}
The rule
~ \hfill
\unnamed
{\ant{\Sigma \vdash \intype{\sigma}{\mathrm{\bf type}_i}}}
{\ant{\Sigma \vdash \intype{\sigma}{\mathrm{\bf type}_j}\;\mbox{for}\;j> i}}
\hfill ~
is sound.
\end{lemma}
\begin{proof}
This follows straightforwardly form the definitions which imply that for $i < j$ we have $\convalue{\mathrm{\bf type}_i} \subseteq \convalue{\mathrm{\bf type}_j}$.
\end{proof}
\subsubsection{Proofs for Figure~\ref{fig:rules}}
The soundness of the first row of rules in figure~\ref{fig:rules} follows straightforwardly from the definitions in figures~\ref{fig:models} and~\ref{fig:value}.
The first three rules of the second row of figure~\ref{fig:rules} are implied by property (Equivalence-Relation) which states that $=_\sigma$ is an
equivalence relation. The soundness of the substitution rule is proved above.
We explicitly consider the rules of the last row.
\begin{lemma} The rule
\vspace{-1em}
~ \hfill
\unnamed{
\ant{\Sigma \vdash \intype{f,g}{\sigma \rightarrow \tau}}
\ant{\Sigma \vdash \forall\intype{x}{\sigma}\;f(x) =_\tau g(x)}}
{\ant{\Sigma \vdash f =_{\sigma \rightarrow \tau}\;g}}
\hfill ~
\vspace{-1em}
is sound.
\end{lemma}
\begin{proof}
Consider $\rho \in \convalue{\Sigma}$. Let $f^*$, $g^*$ and $\sigma^*$ be defined in the usual way. From the validity of the first antecedent
we have $f^*,g^* \in \sigma^* \rightarrow \tau^*$. We must show that the validity of the antecedents implies $f^* =_{\sigma^* \rightarrow \tau^*} g^*$.
From the second antecedent we have that for all $x \in \sigma^*$ there exists $z \in \tau^*$ such that
$f(x)@\tau^* \circ z^{-1} \circ g(x)@\tau^*$ is defined. Property ($=$.C) implies that for $x =_{\sigma^*} x'$ we have
$x@\mathrm{\bf Point} =_{\sigma^*@\typeof(\pointt)} x'@\mathrm{\bf Point}$ and hence
$f^*(x) = f^*[x@\mathrm{\bf Point}] = f^*[x'@\mathrm{\bf Point}] = f^*(x')$ and similarly for $g^*$. For each equivalence class $|x|_{\sigma^*}$
with $x \in \sigma^*$ we can select a value $h(|x_\sigma|) \in \tau^*$ such that $f(|x|_\sigma)@\tau^* \circ h(|x|_\sigma)^{-1} \circ g(|x|_\sigma)@\tau^*$
is defined. This gives a function $h \in \sigma^* \rightarrow \tau^*$
with $f@(\sigma^* \rightarrow \tau^*) \circ (h@(\sigma^* \rightarrow \tau^*))^{-1} \circ g@(\sigma^* \rightarrow \tau^*)$ defined and hence $f^* =_{(\sigma^* \rightarrow \tau^*)} g^*$.
\end{proof}
\begin{lemma}
The rule
\vspace{-1em}
~ \hfill
\unnamed
{\ant{\Sigma;\;\intype{x}{\sigma};\;\intype{y}{\tau} \vdash \intype{\Phi[x,y]}{\mathrm{\bf Bool}}}
\ant{\mbox{$x$ is not free in $\tau$}}
\ant{\Sigma \vdash \forall \intype{x}{\sigma}\;\exists \intype{y}{\tau}\; \Phi[x,\;y]}}
{\ant{\Sigma \vdash \exists \intype{f}{\sigma \rightarrow \tau}\;\forall \intype{x}{\sigma}\;\Phi[x,f(x)]}}
\hfill ~
\vspace{-1em}
is sound.
\end{lemma}
\begin{proof}
Consider $\rho \in \convalue{\Sigma}$. Let $\sigma^*$ abbreviate $\semvalue{\sigma}\rho$ and let $\tau^*$ abbreviate $\semvalue{\tau}\rho$.
We will let $u$ range over members of $\sigma^*$ and $v$ range over members of $\tau^*$.
Let $\Phi^*[u,v]$ abbreviate $\semvalue{\Phi[x,y]}\rho[x \leftarrow u][y \leftarrow v]$. We must show that there exists $f \in \sigma^* \rightarrow \tau^*$
such that for all $u \in \sigma^*$ we have $\Phi^*[u,f(u)]$.
We have that for each $u$ there exists a $v$ such that $\Phi^*[u,v]$. For each $u$ let $v(u)$ denote one such value.
We have
then have $\Phi^*[u,v(u)]$ for every $u$. But the mapping from $\sigma^*$ to $\tau^*$ defined by $u \mapsto v(u)$ need not satisfy condition (F1) in figure~\ref{fig:Morphoids}
which requires that equivalent inputs
yield absolutely equal outputs. Let $|u|$ denote the equivalence class of $u$ under the equivalence relation $=_{\sigma^*}$.
For each such class $|u|$ pick a representative member $w(|u|) \in |u|$. We then let $f$ be a morphoid function with $\mathbf{Dom}(f) = \sigma^*@\typeof(\pointt)$
and such that for $u \in \sigma^*$ we have $f(u) = f[u@\mathrm{\bf Point}] = v(w(|u|))$. By ($=$.C) we have
that if $u@\mathrm{\bf Point} =_{\sigma^*@\typeof(\pointt)} u'@\mathrm{\bf Point}$ for $u,u' \in \sigma^*$ then $u =_{\sigma^*} u'$.
This implies that this definition of $f[z]$ for $z \in \mathbf{Dom}(f)$ is well defined and that satisfies condition (F1).
Hence we have that $f \in {\cal M}(\mathrm{\bf Point} \rightarrow {\cal A})$ where ${\cal A}$ is any template such that $\tau^* \in {\cal M}(\mathbf{TypeOf}({\cal A}))$.
It remains to show that for $u \in \sigma^*$ we have $\Phi^*(u,\;f(u))$. But by the soundness of substitution we have
\begin{eqnarray*}
\Phi^*(u,\;f(u)) & = & \Phi^*(w[|u|],\;f(u)) \\
& = & \Phi^*(w[|u|],\;v(w(|u|))) \\
& = & \mathrm{\bf True}
\end{eqnarray*}
\end{proof}
The last rule in figure~\ref{fig:rules} is the axiom of infinity. The soundness of this rule follows from the fact that we have defined $\convalue{\mathbf{Set}}$
to be the type containing all discrete morphoid types in the universe $V_{\kappa_0}$ where $\kappa_0$ is an uncountable inaccessible cardinal.
\subsubsection{Soundness for Figure~\ref{fig:pairs}}
The proof of soundness of the first rule of figure~\ref{fig:pairs}, the rule for formation of dependent pair types, has been proved above.
We consider the remaining rules in the order in which they appear.
\begin{lemma}
The rule
~\hfill
\unnamed
{\ant{\Sigma \vdash \intype{\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \intype{u}{\sigma}}
\ant{\Sigma \vdash \intype{w}{\tau[u]}}}
{\ant{\Sigma \vdash \intype{\mathbf{Pair}(u,w)}{\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}}}
\ant{\Sigma \vdash \pi_1(\mathbf{Pair}(u,w)) \doteq u}
\ant{\Sigma \vdash \pi_2(\mathbf{Pair}(u,w)) \doteq w}}
\hfill ~
is sound.
\end{lemma}
\begin{proof}
We consider requirements (V1) and (V2) for the first conclusion. For the pair expression $\mathbf{Pair}(u,w)$ we have that conditions (V1) and (V2)
follow directly from conditions (V1) and (V2) for $\Sigma \models \intype{u}{\sigma}$ and $\Sigma \models \intype{w}{\tau[u]}$.
\end{proof}
\begin{lemma}
The rule
~\hfill
\unnamed
{\ant{\Sigma \vdash \intype{p}{\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}}}}
{\ant{\Sigma \vdash \intype{\pi_1(p)}{\sigma}}
\ant{\Sigma \vdash \intype{\pi_2(p)}{\tau[\pi_1(p)]}}
\ant{\Sigma \vdash p \doteq \mathbf{Pair}(\pi_1(p),\;\pi_2(p))}}
\hfill ~
is sound.
\end{lemma}
\begin{proof}
We consider requirements (V1) and (V2) for the first two conclusion. For the expression $\pi_1(p)$ we have that conditions (V1) and (V2)
follow directly from conditions (V1) and (V2) for $\Sigma \models \intype{p}{\pairtype{x}{\sigma}{\intype{y}{\tau[x]}}}$
and the fact that for pairs $p$ and $q$ with $p \circ q$ defined we have $\pi_1(p\circ q) = \pi_1(p) \circ \pi_1(q)$.
For pairs $p$ and $q$ with $p \preceq q$ we also have $\pi_1(p) \preceq \pi_1(p)$. Similar comments apply to $\pi_2$.
\end{proof}
The soundess of the rules for reflexivity, symmetry, transitivity and substitution for absolute equality follow immediately from the definition of absolute equality.
\begin{lemma}
The rule
~ \hfill
\unnamed
{\ant{\Sigma \vdash \intype{\tau}{\mathrm{\bf type}_i}}
\ant{\Sigma;\;\intype{x}{\tau} \vdash \intype{\Phi[x]}{\mathrm{\bf Bool}}}}
{\ant{\Sigma \vdash \intype{\subtype{\intype{x}{\tau},\;\Phi[x]}}{\mathrm{\bf type}_i}}}
\hfill ~
is sound.
\end{lemma}
\begin{proof}
We will show properties (V1) and (V2) of the conclusion. We first show soundness of the rule
\vspace{-1em}
~ \hfill \unnamed{
\ant{\Sigma \vdash \intype{\Phi}{\mathrm{\bf Bool}}}}
{\ant{\Sigma \vdash \intype{\mathbf{Subtype}(\intype{P}{\mathrm{\bf Bool}},\;\Phi)}{\mathrm{\bf type}_i}}}
\hfill ~
For $\rho \in \convalue{\Sigma}$ we have that $\semvalue{\mathbf{Subtype}(\intype{P}{\mathrm{\bf Bool}},\;\Phi)}\rho$ is either $\convalue{\mathrm{\bf Bool}}$ or the emtpy type depending on the value of
$\semvalue{\Phi}\rho$. We then get that properties (V1) and (V2) for the conclusion of this rule follow directly from the properties (V1) and (V2) of the antecedent.
Next we consider the rule
\vspace{-1em}
~ \hfill
\unnamed
{\ant{\Sigma \vdash \intype{\sigma}{\mathrm{\bf type}_i}}
\ant{\Sigma;\;\intype{x}{\sigma} \vdash \intype{\Phi[x]}{\mathrm{\bf Bool}}}}
{\ant{\Sigma \vdash \intype{\mathbf{PairType}(\intype{x}{\sigma},\;\intype{Q}{\mathbf{SubType}(\intype{P}{\mathrm{\bf Bool}},\;\Phi[x])})}{\mathrm{\bf type}_i}}}
\hfill ~
The validity of the antecedents and the soundness of the preceding rule give that the sequent $\Sigma;\;\intype{x}{\sigma} \vdash \intype{\mathbf{SubType}(\intype{P}{\mathrm{\bf Bool}},\;\Phi[x])}{\mathrm{\bf type}_i}$
is valid. The soundness of the pair type formation rule then gives that the conclusion of the above rule is valid.
For $\rho \in \convalue{\Sigma}$ we now have that $\semvalue{\mathbf{Subtype}(\intype{x}{\sigma},\;\Phi[x])}\rho$ is the set of values of the form $\pi_1[z]$
for $z \in \semvalue{\mathbf{PairType}(\intype{x}{\sigma},\;\mathbf{SubType}(\intype{P}{\mathrm{\bf Bool}},\;\Phi[x]))}\rho$. Properties (V1) and (V2) of the subtype formation rule
now follow from properties (V1) and (V2) of the conclusion of the above rule. We omit the details.
\ignore{
We will show properties (V1) and (V2) of the conclusion. For (V1) consider $\rho_1,\rho_2,\rho_3 \in \convalue{\Sigma}$ with $\rho_1 \circ \rho_2^{-1} \circ \rho_3$ defined
and with $\rho_1 \circ \rho_2^{-1} \circ \rho_3$ in $\convalue{\Sigma}$. Let $\tau_i^*$ abbreviate $\semvalue{\tau}\rho_i$
and for $u \in \tau_i^*$ the $\Phi_i^*[u]$ abbreviate $\subvalue{\Sigma;\intype{x\;}{\;\tau}}{\Phi[x]}\rho_i[x \leftarrow u]$.
Finally, let $\mathbf{SubType}_i^*(\intype{x}{\tau},\;\Phi[x])$ abbreviate $\semvalue{\mathbf{SubType}(\intype{x}{\tau},\;\Phi[x])}\rho_i$.
We must show that
$$\semvalue{\mathbf{SubType}(\intype{x}{\tau},\;\Phi[x])}(\rho_1 \circ \rho_2^{-1} \circ \rho_3)$$
$$=$$
$$\mathbf{SubType}_1^*(\intype{x}{\tau},\;\Phi[x]) \circ \mathbf{SubType}_2^*(\intype{x}{\tau},\;\Phi[x])^{-1} \circ \mathbf{SubType}_3^*(\intype{x}{\tau},\;\Phi[x])$$
We will show containment in both directions. First consider $u \in \semvalue{\mathbf{SubType}(\intype{x}{\tau},\;\Phi[x])}(\rho_1 \circ \rho_2^{-1} \circ \rho_3)$.
We have $u \in \semvalue{\tau}(\rho_1 \circ \rho_2^{-1} \circ \rho_3)$ and by condition (V1) of the first antecedent this gives
$u \in \tau_1^* \circ {\tau_2^*}^{-1} \circ \tau_3^*$. This implies that $u = u_1 \circ u_2^{-1} \circ u_3$ for $u_i \in \tau_i^*$.
Condition (V1) of the second antecedent implies that
\begin{eqnarray*}
\subvalue{\Sigma,\intype{x\;}{\;\tau}}{\Phi[x]}[x \leftarrow u] & = & \subvalue{\Sigma,\intype{x\;}{\;\tau}}{\Phi[x]}[x \leftarrow (u_1 \circ u_2^{-1} \circ u_3)] \\
& = & \Phi_1^*[u_1] \circ \Phi_2^*[u_2] \circ \Phi_3^*[u_3]
\end{eqnarray*}
This gives $u_i \in \mathbf{Subtype}^*_i(\intype{x}{\tau},\;\Phi[x])$ and we have
$$u \in \mathbf{SubType}_1^*(\intype{x}{\tau},\;\Phi[x]) \circ \mathbf{SubType}_2^*(\intype{x}{\tau},\;\Phi[x])^{-1} \circ \mathbf{SubType}_3^*(\intype{x}{\tau},\;\Phi[x])$$
as desired.
Conversely consider
$$u \in \mathbf{SubType}_1^*(\intype{x}{\tau},\;\Phi[x]) \circ \mathbf{SubType}_2^*(\intype{x}{\tau},\;\Phi[x])^{-1} \circ \mathbf{SubType}_3^*(\intype{x}{\tau},\;\Phi[x]).$$
This implies that that there exists $u_i \in \tau_i^*$ with $u = u_1 \circ u_2^{-1} \circ u_3$ and with $\Phi_i^*[u_i] = \mathrm{\bf True}$. By property (V1)
of the first antecedent we have $\semvalue{\tau}(\rho_1 \circ \rho_2^{-1} \rho_3) = \tau_1^* \circ {\tau_2^*}^{-1} \circ \tau_3^*$. This gives
$u \in \semvalue{\tau}(\rho_1 \circ \rho_2^{-1} \rho_3)$.
Condition (V1) of the second antecedent implies
\begin{eqnarray*}
& & \subvalue{\Sigma;\intype{x\;}{\;\tau}}{\Phi[x]}(\rho_1 \circ \rho_2^{-1} \circ \rho_3)[x \leftarrow u] \\
& = & \subvalue{\Sigma;\intype{x\;}{\;\tau}}{\Phi[x]}(\rho_1[x \leftarrow u_1] \circ \rho_2[x \leftarrow u_2]^{-1} \circ \rho_3[x \leftarrow u_3]) \\
& = & \Phi_1^*[u_1] \circ \Phi_2^*[u_2] \circ \Phi_3^*[u_3] \\
& = & \mathrm{\bf True}
\end{eqnarray*}
This gives $u \in \semvalue{\mathbf{SubType}(\intype{x}{\tau},\;\Phi[x])}(\rho_1 \circ \rho_2^{-1} \circ \rho_3)$ which proves (V1) for the conclusion of the rule.
To prove (V2) consider $\rho_1,\rho_2 \in \convalue{\Sigma}$ with $\rho_1 \preceq \rho_2$.
Let $\tau_i^*$ abbreviate $\semvalue{t}\rho_i$ and for $u \in \tau_i^*$ let $\Phi_i^*[u]$ be defined in the standard way.
Let $\mathbf{SubType}_i^*(\intype{x}{\tau},\;\Phi[x])$ abbreviate $\semvalue{\mathbf{SubType}(\intype{x}{\tau},\;\Phi[x])}\rho_i$.
We must show
$$\mathbf{SubType}_1^*(\intype{x}{\tau},\;\Phi[x]) \preceq \mathbf{SubType}_2^*(\intype{x}{\tau},\;\Phi[x]).$$
Let ${\cal A}$ be an interface template for $\mathbf{SubType}_2^*(\intype{x}{\tau},\;\Phi[x])$.
We then have that $\mathbf{TypeOf}({\cal A})$ is a minimal template for $\mathbf{SubType}_2^*(\intype{x}{\tau},\;\Phi[x])$
and it suffices to show that $\mathbf{SubType}_1^*(\intype{x}{\tau},\;\Phi[x])@\mathbf{TypeOf}({\cal A})$ is defined
and equals $\mathbf{SubType}_2^*(\intype{x}{\tau},\;\Phi[x])@\mathbf{TypeOf}({\cal A})$.
{\color{red} This is a mess.}}
\end{proof}
\begin{lemma}
The rules
~ \hfill
\unnamed
{\ant{\Sigma \vdash \intype{\subtype{\intype{x}{\tau},\;\Phi[x]}}{\mathrm{\bf type}_i}}
\ant{\Sigma \vdash \intype{e}{\tau}}
\ant{\Sigma \vdash \Phi[e]}}
{\ant{\Sigma \vdash \intype{e}{\subtype{\intype{x}{\tau},\;\Phi[x]}}}}
\hfill
\unnamed
{\ant{\Sigma \vdash \intype{e}{\subtype{\intype{x}{\tau},\;\Phi[x]}}}}
{\ant{\Sigma \vdash \intype{e}{\tau}}
\ant{\Sigma \vdash \Phi[e]}}
\hfill ~
are sound.
\end{lemma}
\begin{proof}
For the first rule we note that properties (V1) and (V2) of the conclusion follow immediately from properties (V1) and (V2) of the second antecedent.
For the second rule we note that properties (V1) and (V2) of the first conclusion follow immediately from (V1) and (V2) of the antecedent.
\end{proof}
\subsubsection{Soundness for Figure~\ref{fig:additional}}
\begin{lemma}
The rule
\vspace{-1em}
~ \hfill
\unnamed
{\ant{\Sigma;\intype{x}{\sigma} \vdash \intype{\Phi[x]}{\mathrm{\bf Bool}}}
\ant{\Sigma \vdash \exists!\intype{x}{\sigma}\;\Phi[x]}
\ant{\Sigma \vdash \intype{\sigma}{\mathbf{Set}}}}
{\ant{\Sigma \vdash \intype{\mathbf{The}(\intype{x}{\sigma},\;\Phi[x])}{\sigma}}
\ant{\Sigma \vdash \Phi[\mathbf{The}(\intype{x}{\sigma},\;\Phi[x])]}}
\hfill ~
\vspace{-1em}
is sound.
\end{lemma}
\begin{proof}
We show (V1) and (V2) for the first conclusion. We use the requirement that sets be discrete. For a discrete type the third antecedent implies
that there is one element of the type $\mathbf{SubType}(\intype{x}{\sigma;\;\Phi[x])}$. For singleton discrete types $\sigma_1$, $\sigma_2$ and $\sigma_3$
with $\sigma_1 \circ \sigma_2^{-1} \circ \sigma_3$ defined we have
we have $\mathbf{The}(\sigma_1 \circ \sigma_2^{-1} \circ \sigma_3)$ equals $\mathbf{The}(\sigma_1) \circ \mathbf{The}(\sigma_2)^{-1} \circ \mathbf{The}(\sigma_3)$
and $\sigma_1 \preceq \sigma_2$ if and only if $\mathbf{The}(\sigma_1) \preceq \mathbf{The}(\sigma_2)$.
\end{proof}
\begin{lemma}
The rule
\vspace{-1em}
~ \hfill
\unnamed
{\ant{\Sigma;\intype{x}{\sigma} \vdash \intype{\Phi[x]}{\mathrm{\bf Bool}}}
\ant{\Sigma \vdash \exists!\intype{x}{\sigma}\;\Phi[x]}
\ant{{\mathbf{The}(\intype{x}{\sigma},\;\Phi[x])}\;\mbox{is closed}}}
{\ant{\Sigma \vdash \intype{\mathbf{The}(\intype{x}{\sigma},\;\Phi[x])}{\sigma}}
\ant{\Sigma \vdash \Phi[\mathbf{The}(\intype{x}{\sigma},\;\Phi[x])]}}
\hfill ~
\vspace{-1em}
is sound.
\end{lemma}
\begin{proof}
Conditions (V1) and (V2) hold immediately for closed expressions.
\end{proof}
The remaining rules in figure~\ref{fig:additional} all follow directly from the definitions of the constructs involved.
\subsubsection{Soundness for Figure~\ref{fig:jrules}}
All of the rules in figure~\ref{fig:jrules} are manifestly sound.
\section{Final Comments on Platonism}
The foundations of mathematics seems more a branch of cognitive science than a branch of mathematics. The relationship between
logic and thought has been central to the development of logic from the ancient Greeks, through Leibniz, Boole, Frege and into mainstream Anglo-American analytic
philosophy. A type-theoretic foundation seems much closure to natural human thought than does untyped set theory. The excluded middle and the non-constructive
axiom of choice also seems to be inherent in human thought.
The existence of Platonic thought does not imply a causal connection between thought and the objects being considered.
We humans are presumably some form of machine and most human
mathematicians engage in Platonic thinking. We do not need to postulate any magical or mystical connection between thought and actual spheres and manifolds.
It seems more reasonable to model our thought process, even Platonic thought, as some form of symbolic computation.
But if Platonic thought is actually just symbolic computation what makes it ``Platonic''? A better question is what does one mean by a Platonic foundation
for mathematics? Formal semantics translates symbol strings into rigorous natural language --- the natural language
of the practice of mathematics. For example, the symbol string $\forall \intype{x}{\sigma}\;\Phi[x]$ is true if for all $x$ in $\sigma$ we have that $\Phi[x]$ is true.
Similarly, the symbol string $\mathbf{PairOf}(\intype{x}{\sigma},\;\intype{y}{\tau[x]})$ represents the type of pairs $(x,y)$ with $x$ in $\sigma$
and $y$ in $\tau[x]$. A formal system is Platonic to the extent that the formal semantics (the model) yields a translation from formal symbols to natural language that
is direct and trivial. A Platonic semantics establishes a tight correspondence between formal symbol strings and the naturally occurring language of thought.
The semantics of figure~\ref{fig:value} is designed to have this property.
As a final note we consider Wigner's famous comment on the unreasonable effectiveness of mathematics in physics.
By ``mathematics'' Wigner is referring to mathematical concepts such as topological spaces, manifolds and group
representations. A type-theoretic foundation of mathematics faithfully interprets these concepts as formal types.
Type theory is central to concept-based mathematics.
We can interpret Wigner's comment as stating the unreasonable effectiveness of type theory in physics. This is indeed
striking.
\bibliographystyle{apalike}
|
\section{Introduction}
\label{int}
There is a wide family of cohomology classes of spaces of knots $S^1
\hookrightarrow \R^n$ ($n \ge 3$), called {\em finite-type cohomology classes};
see \cite{V1}, \cite{fasis}, \cite{bjo}. For $n>3$ they cover all of the
cohomology group of the space of knots in $\R^n$, for $n=3$ their 0-dimensional
part are the finite-type knot invariants.
These classes are defined as linking numbers (in the space of all smooth maps
$S^1 \to \R^n$) with appropriate cycles (of infinite dimension but finite
codimension) in the {\em discriminant space} $\Sigma$ (cf. \cite{arnold}); in
our case this space consists of maps which are not smooth embeddings. The group
of all such classes is filtered by their {\em orders} induced by some
filtration of (some resolution of) the discriminant: roughly speaking, the
order of a cohomology class indicates how much complicated strata of $\Sigma$
participate in the definition of its dual variety.
In \cite{PV}, M.~Polyak and O.~Viro have proposed some combinatorial formulas
for the finite-type invariants of knots in $\R^3$. Later, M.~Goussarov has
proved that any finite-type invariant can be represented by a formula of this
type, see \cite{GPV}.
We describe some calculus for finding (and proving) combinatorial formulas for
arbitrary finite type cohomology classes, in particular show what the answers
can look like. {\em Any such combinatorial formula is nothing else than some
semialgebraic chain in the space of maps $S^1 \to \R^n,$ whose boundary lies in
$\Sigma$ and our cohomology class is equal to the linking number with this
boundary.} We introduce several natural families of semialgebraic subvarieties
of the space of such maps, of which the desired chains are built. These
varieties are defined by easy differential geometrical conditions; they arise
naturally in the direct calculation of the main spectral sequence converging to
the (finite type) cohomology group of the space of knots. It is not surprising
that some elements of this calculus repeat pictures from \cite{PV}, \cite{GPV},
and also from the A.~B.~Merkov's works on invariants of plane curves \cite{M},
\cite{MM}.
We accomplish these calculations explicitly for several cohomology classes of
low orders. Before describing them three remarks more.
{\bf 1. Long and compact knots.} We shall distinguish two kinds of knot spaces.
The {\em compact knots} in $\R^n$ are any smooth embeddings $S^1 \to \R^n,$
while the {\em long knots} are the smooth embeddings $\R^1 \to \R^n$ coinciding
with a standard linear embedding outside some compact subset in $\R^1$. The
invariants of knots of both types in $\R^3$ naturally coincide, but generally
the cohomology ring of the space of compact knots is more complicated: it is
built of the similar ring for long knots (playing the role of a "coefficient"
ring) and homology groups of the space $S^1$ and certain its configuration
spaces.
{\bf 2. Stabilization.} If numbers $n$ and $m$ are of the same parity, then the
theories of (finite type) cohomology groups of spaces of knots in $\R^n$ and
$\R^m$ are very similar. Namely, the first terms of spectral sequences
calculating both groups and generated by the natural filtration of resolved
discriminants coincide up to shifts of indices:
\begin{equation}
\label{stabil} E_1^{p,q-pn}(\R^n) \simeq E_1^{p,q-pm}(\R^m).
\end{equation}
Moreover, for spectral sequences calculating $\Z_2$-cohomology groups this
identity is true also if $n$ and $m$ are of different parities. M.~Kontsevich
has proved (but not published) that in the case of complex coefficients our
spectral sequence degenerates at the first term: $E_\infty^{p,q} \equiv
E_1^{p,q}$, therefore also the limit groups of finite type cohomology classes
are very similar. (I conjecture that in the case of long knots the similar
degeneration holds also for any coefficients.)
{\bf 3.} This paper is very much a work in the differential geometry of spatial
curves and their projections to different subspaces, although almost all
results of this kind are hidden in the formulas for boundary operators in our
homological calculations.
\subsection{Results for long knots.}
\label{reslong}
Accordingly to \cite{V1}, \cite{tetu}, \cite{bjo}, \cite{fasis}, all cohomology
classes of orders $\le 3$ of the space of long knots in $\R^n,$ $n \ge 3,$ are
as follows.
\begin{proposition}
\label{longres} There are no cohomology classes of order 1. The classes of
order 2 are only in dimension $2n-6$ and form a group isomorphic to $\Z$ $($for
$n=3$ it is generated by the simplest knot invariant$)$. In order 3 additional
classes can be in exactly two dimensions more: $3n-9$ and $3n-8$. In dimension
$3n-9$ they form a group isomorphic to $\Z$ $($for $n=3$ it is generated by the
next simple knot invariant$)$. In dimension $3n-8$ the same is true if $n >3,$
and for $n=3$ the similar $($1-dimensional$)$ cohomology group is cyclic
$($maybe of order 1 or $\infty )$.
\end{proposition}
It was conjectured in \cite{fasis}, \cite{bjo} that the latter group for $n=3$
also is isomorphic to $\Z$; we shall prove it in the present work.
For any $n$ we call the generator of this $(3n-8)$-dimensional cohomology group
the {\em Turchin--Teiblum cocycle}. In the case of odd $n$ its existence was
discovered by D.~M.~Teiblum and V.~E.~Turchin about 1995 (\cite{tetu}). Its
(quite different) superanalog for even $n$ was found in \cite{fasis},
\cite{bjo}. However, all these works contain only an implicit proof of the
existence of such a class: namely, the calculation of the third column of our
spectral sequence (which is responsible for the third order cohomology classes
and is isomorphic to $\Z$ for exactly two values of $q$), and the remark that
all further differentials acting from or to this column are trivial by some
dimensional reasons.
In \S \ref{proof1} we prove the following combinatorial expression for this
class reduced mod 2.
Let us choose a direction "up" in $\R^n$, and say that a point $x \in \R^n$ is
{\em above} the point $y$ if the vector $\overrightarrow{(yx)}$ has the chosen
direction. Let $\R^{n-1}$ be the quotient space of $\R^n$ by this direction,
and ${\bf p}: \R^n \to \R^{n-1}$ the corresponding projection. We choose a
direction "to the right" in $\R^{n-1}$ and say that the point $x \in \R^n$ {\em
is to the right} of the point $y$ if the vector $\overrightarrow{({\bf
p}(y),{\bf p}(x))} \in \R^{n-1}$ has this chosen direction.
\begin{theorem}
\label{main} For any $n \ge 3,$ the value of the reduced mod 2 Teiblum--Turchin
class on any generic $(3n-8)$-dimensional singular cycle in the space of long
knots in $\R^n$ is equal to the parity of the number of points of this cycle
corresponding to such knots $f:\R^1 \to \R^n$ that one of three holds:
a) there are five points $a<b<c<d<e$ in $\R^1$ such that $f(a)$ is above
$f(d)$, and $f(e)$ is above $f(c)$ and $f(b)$;
b) there are four points $a<b<c<d$ in $\R^1$ such that $f(a)$ is above $f(c)$,
$f(b)$ is below $f(d)$, and the projection of the derivative $f'(b)$ to
$\R^{n-1}$ is directed to the right;
c) there are three points $a<b<c$ in $\R^1 $such that $f(a)$ is above $f(b)$
but below $f(c)$, and the "exterior" angle in $\R^{n-1}$ formed by projections
of $f'(a)$ and $f'(b)$ contains the direction "to the right" $($i.e. this
direction is equal to a linear combination of these projections, and at least
one of coefficients in this combination is non-positive$)$.
\end{theorem}
We prove this theorem in \S \ref{proof1}. In the next works I am planning to
accomplish all the same calculations taking respect on the orientations, and
thus to obtain similar results with integer coefficients.
\begin{corollary}
\label{nontriv} The group of order 3 one-dimensional cohomology classes of the
space of long knots in $\R^3$ is free cyclic and generated by the
$($integral$)$ Teiblum--Turchin class.
\end{corollary}
More precisely, let us consider the connected sum of two equal (long) trefoil
knots in $\R^3$ and a path in the space of knots connecting this knot with
itself as in the proof of the commutativity of the knot semigroup: we shrink
the first summand, move it "through" the second, and then blow up again.
\begin{proposition}
\label{realiz} This closed path in the space of long knots has an odd number of
intersection points with the union of three varieties indicated in items a, b
and c of Theorem \ref{main}.
\end{proposition}
The proof will be given in \S~\ref{proreal}.
\medskip
On the other hand, for any $n$ the Teiblum--Turchin cocycle is a well-defined
integral cohomology class, and Corollary \ref{nontriv} is proved.
\subsection{Answers for compact knots}
\label{comp}
Nontrivial cohomology classes in the space of {\em compact knots} $S^1
\hookrightarrow \R^n$ appear already in filtrations 1 and 2. We assume that a
cyclic coordinate in $S^1$, i.e. an identification $S^1 \simeq \R^1/2\pi \Z$,
is fixed.
\begin{proposition}[see \cite{arman}, \cite{bjo}]
\label{ordone} For any $n \ge 3$ the group of $\Z_2$-cohomology classes of
order 1 of the space of compact knots in $\R^n$ is nontrivial only in
dimensions $n-2$ and $n-1$, and is isomorphic to $\Z_2$ in these dimensions.
Moreover, for $($only$)$ even $n$ similar integral cohomology groups in these
dimensions are isomorphic to $\Z.$ The generator of the $(n-2)$-dimensional
group is Alexander dual to the set ${\mathcal L}$ of discriminant maps $S^1 \to
\R^n$ gluing together some two {\em opposite} points of $S^1$, and the
generator of the $(n-1)$-dimensional group is dual to the set of maps gluing
some {\em chosen} opposite points, say $0$ and $\pi$.
\end{proposition}
\begin{proposition}[see \cite{fasis}, \cite{bjo}]
\label{ordtwo} Additional classes of order 2 exist in exactly two dimensions:
$2n-6$ and $2n-3$. In dimension $2n-6$ they for any $n$ form a group isomorphic
to $\Z$ $($for $n=3$ it is generated by the simplest knot invariant$)$. The
group in dimension $2n-3$ is isomorphic to $\Z$ for $n>3$ and cyclic for $n=3$;
its generator is Alexander dual to the cycle in the discriminant, whose {\em
principal part} $($see Definition \ref{fitype} in \S \ref{method} below$)$ in
the double self-intersection of $\Sigma$ is swept out by such maps $f:S^1 \to
\R^n$ that for some $\alpha \in S^1$ we have $f(\alpha) = f(\alpha+\pi),$
$f(\alpha+\pi/2) =f(\alpha+3\pi/2).$
\end{proposition}
Below we prove in particular that for $n=3$ the last group also is free cyclic,
see Corollary \ref{realiz3}. Now we give explicit combinatorial formulas for
all classes mentioned in Propositions \ref{ordone} and \ref{ordtwo}.
\begin{theorem}
\label{comain} For any $n \ge 3$, the values of any of these four basic
cohomology classes on any generic cycle of corresponding dimension in the space
${\mathcal K}_n \setminus \Sigma$ of compact knots $S^1 \hookrightarrow \R^n$
is equal to the number of points of this cycle, corresponding to knots
satisfying the following conditions $($and in the case of integer coefficients
taken with appropriate signs$)$.
A. For the $(n-1)$-dimensional class of order $1$: projections of $f(0)$ and
$f(\pi)$ into the plane $\R^{n-1}$ coincide, and $f(0)$ is above $f(\pi)$.
B. For the $(n-2)$-dimensional class of order 1, one of the following two
conditions:
a) there is a point $\alpha \in [0,\pi)$ such that the projections of
$f(\alpha)$ and $f(\alpha+\pi)$ to $\R^{n-1}$ coincide, and moreover
$f(\alpha)$ is above $f(\alpha+\pi)$;
b) the projection of the point $f(0)$ to $\R^{n-1}$ lies "to the right" from
the projection of $f(\pi)$.
C. For the $(2n-3)$-dimensional class of order $2$, one of following two
conditions:
a) there is a point $\alpha \in [0,\pi/2)$ such that projections of $f(\alpha)$
and $f(\alpha+\pi)$ to $\R^{n-1}$ coincide, projections of $f(\alpha+\pi/2)$
and $f(\alpha+3\pi/2)$ to $\R^{n-1}$ coincide, and additionally $f(\alpha+\pi)$
is above $f(\alpha)$ and $f(\alpha+\pi/2)$ is above $f(\alpha+3\pi/2)$;
b) projections of $f(0)$ and $f(\pi)$ to $\R^{n-1}$ coincide, $f(\pi)$ is above
$f(0)$, and the projection of $f(\pi/2)$ to $\R^{n-1}$ lies "to the right" from
the projection of $f(3\pi/2)$.
D. For the $(2n-6)$-dimensional class of order $2$, one of two conditions:
a) there are four distinct points $\alpha, \beta, \gamma, \delta \in S^1$
$($whose cyclic coordinates satisfy $0 \le \alpha < \beta < \gamma < \delta <
2\pi)$ such that projections of $f(\alpha)$ and $f(\gamma)$ to $\R^{n-1}$
coincide, projections of $f(\beta)$ and $f(\delta)$ to $\R^{n-1}$ coincide, and
additionally $f(\gamma)$ is above $f(\alpha)$ and $f(\beta)$ is above
$f(\delta)$.
b) If $n=3$ then second condition is void $($and we have only the first one
coinciding with the combinatorial formula from \mbox{\cite{PV}}$)$, but for
$n>3$ we have additional condition: there are three distinct points $ \beta,
\gamma, \delta$ $($whose cyclic coordinates satisfy $0 < \beta < \gamma <
\delta < 2\pi)$ such that projections of $f(\gamma)$ and $f(0)$ to $\R^{n-1}$
coincide, $f(\gamma)$ is above $f(0)$, and the projection of $f(\delta)$ to
$\R^{n-1}$ lies "to the right" of the projection of $f(\beta)$.
\end{theorem}
Proofs see in \S \ \ref{proof2}.
\medskip
\begin{corollary}
\label{realiz2} For any $n\ge 3$, the basic class of order $2$ and dimension
$2n-3$ takes value $\pm 1$ on the submanifold of the space of knots, consisting
of all naturally parametrized great circles of the unit sphere in $\R^n$.
\end{corollary}
Indeed, the variety a) of statement C does not intersect this submanifold, and
variety b) has with it exactly one intersection point. \quad $\square$
\medskip
In the case of even $n$ the fact that this variety in the space of knots is not
homologous to zero was proved in \cite{cotta} by very different methods.
\medskip
\begin{corollary}
\label{realiz3} The group of $(2n-3)$-dimensional cohomology classes of order 2
is free cyclic for $n=3$ as well. \quad $\square$
\end{corollary}
I am indebted to A.~B.~Merkov very much for many interesting conversations.
\section{Methodology and nature of combinatorial expressions.}
\label{method}
In fact, our main purpose is to develop a general method of finding
combinatorial formulas of this type.
Any such formula is just a relative cycle in the space of knots (modulo the
discriminant $\Sigma$) whose boundary in $\Sigma$ is Alexander dual to our
cohomology class. The problem is to construct such a variety explicitly and as
simply as possible.
Our method of doing it consists in the conscientious calculation of our
spectral sequence. In this subsection we outline the definition of this
sequence and this calculation. This spectral sequence for spaces of knots is
very analogous to that calculating the cohomology of complements of plane
arrangements (see \cite{congr}); let us demonstrate their main common features
on the latter more simple example.
\subsection{Simplicial resolutions for plane arrangements}
Let $L \subset \R^N$ be an {\em affine plane arrangement,} i.e. the union of
finitely many affine planes $L_i$ of any dimensions, $L = \bigcup_{i=1}^k L_i.$
The cohomology group of its complement is Alexander dual to the homology group
of $L$: $H^j(\R^N \setminus L) \simeq \bar H_{N-j-1}(L);$ here $\bar H_*$
denotes the {\em Borel--Moore homology group}, i.e. the homology group of the
one-point compactification reduced modulo the added point. To calculate the
latter group it is convenient to use the {\em simplicial resolution} of $L$
(which is a continuous version of the combinatorial formula of inclusions and
exclusions).
For some three line arrangements in $\R^2$ (shown in the lower part of
Fig.~\ref{arr}) the corresponding simplicial resolutions are given above them
in the same picture. These resolutions are constructed as follows.
\begin{figure}
\unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(128.00,90.00)
\put(0.00,10.00){\line(1,1){20.00}} \put(0.00,30.00){\line(1,-1){20.00}}
\put(20.00,40.00){\line(-1,1){20.00}} \put(0.00,70.00){\line(1,1){20.00}}
\put(10.10,80.10){\line(0,-1){30.20}} \put(9.90,79.90){\line(0,-1){29.80}}
\put(60.00,10.00){\line(-1,1){20.00}} \put(60.00,30.00){\line(-1,-1){20.00}}
\put(50.00,10.00){\line(0,1){20.00}} \put(60.00,40.00){\line(-1,1){20.00}}
\put(40.00,70.00){\line(1,1){20.00}} \put(60.00,55.00){\line(0,1){20.00}}
\put(50.00,50.00){\line(0,1){30.00}} \put(50.00,80.00){\line(2,-3){10.00}}
\put(60.00,65.00){\line(-2,-3){10.00}}
\put(60.00,77.00){\makebox(0,0)[cc]{$3$}}
\put(38.00,68.00){\makebox(0,0)[cc]{$2$}}
\put(38.00,62.00){\makebox(0,0)[cc]{$1$}}
\put(38.00,32.00){\makebox(0,0)[cc]{$1$}}
\put(38.00,8.00){\makebox(0,0)[cc]{$2$}}
\put(50.00,32.00){\makebox(0,0)[cc]{$3$}} \put(90.00,10.00){\line(1,1){20.00}}
\put(100.00,30.00){\line(1,-1){20.00}} \put(120.00,15.00){\line(-1,0){30.00}}
\put(82.00,48.00){\line(1,1){20.00}} \put(108.00,68.00){\line(1,-1){20.00}}
\put(120.00,40.00){\line(-1,0){30.00}} \put(114.90,40.00){\line(3,5){7.80}}
\put(115.10,40.00){\line(3,5){7.70}} \put(95.10,40.00){\line(-3,5){7.80}}
\put(94.90,40.00){\line(-3,5){7.70}} \put(97.10,63.10){\line(1,0){15.80}}
\put(96.90,62.90){\line(1,0){16.20}}
\put(14.75,19.75){\rule{0.50\unitlength}{0.50\unitlength}}
\put(10.00,2.00){\makebox(0,0)[cc]{$a$}}
\put(50.00,2.00){\makebox(0,0)[cc]{$b$}}
\put(105.00,2.00){\makebox(0,0)[cc]{$c$}} \put(52.00,53.00){\line(0,1){24.00}}
\put(54.00,74.00){\line(0,-1){18.00}} \put(56.00,59.00){\line(0,1){12.00}}
\put(58.00,68.00){\line(0,-1){6.00}}
\end{picture}
\caption{Examples of line arrangements} \label{arr}
\end{figure}
First, we embed the set of indices $\{1, \ldots, k\}$ into a space $\R^T$ of
dimension $T \ge k-1$ in such a way that their convex hull is a
$(k-1)$-dimensional simplex. The resolution will be constructed as a subset in
$\R^T \times \R^N.$ For any point $x \in L$ denote by $\tilde \Delta(x)$ the
convex hull in $\R^T$ of images of such indices $i$ that $L_i \ni x$, i.e. the
simplex with vertices at images of all these indices. Denote by $\Delta(x)$ the
simplex $\tilde \Delta(x) \times \{x\} \subset \R^T \times \R^N.$ Denote by
$L'$ the union of all simplices $\Delta(x)$, $x \in L.$ The obvious projection
$L' \to L$ (sending any $\Delta(x)$ to $x$) is a homotopy equivalence, as well
as its extension to the map of one-point compactifications $\bar L' \to \bar
L.$ In particular $\bar H_*(L') \equiv \bar H_*(L).$
On the other hand, $L'$ has a very useful filtration. For any set of indices $I
\subset \{1, \ldots, k\}$, denote by $L_I$ the plane $\cap_{i \in I} L_i. $
Let $L'_I \subset L'$ be the {\em proper preimage} of $L_I$, i.e. the closure
of the union of complete preimages of all {\em generic} points of $L_I$ (i.e.
of points not from even smaller strata $L_J \subset L_I,$ $L_J \ne L_I$). There
is obvious homeomorphism $L'_I \simeq \tilde \Delta(I) \times L_I,$ where
$\tilde \Delta(I) \subset \R^T$ is the simplex whose vertices correspond to all
indices $i$ such that $L_i \supset L_I.$ (It is equal to $\tilde \Delta(x)$
where $x$ is any generic point of $L_I.$)
By definition, $L' = \bigcup L'_I,$ where the union is taken over all {\em
geometrically different} planes $L_I$. We define the term $F_p$ of the desired
filtration of $L'$ as the similar union of prisms $L'_I$ over all planes $L_I$
{\em of codimension $\le p$}. Then we extend it to a filtration on the
one-point compactification $\overline{L'}$ of $L'$ setting $F_0=$ \{the added
point\}.
This filtration defines a spectral sequence calculating the group $\bar H_*(L')
\simeq \bar H_*(L)$: by definition its term $E_{p,q}^1$ is equal to $\bar
H_{p+q}(F_p \setminus F_{p-1}) \equiv
H_{p+q}(\overline{F_p}/\overline{F_{p-1}}).$ This space $F_p \setminus F_{p-1}$
splits into a disjoint union (over all spaces $L_I$ of codimension exactly $p$)
of spaces $\check L'_I \stackrel{def}{=} \check \Delta(I) \times L'_I$, where
$\check \Delta(I)$ is the simplex $\tilde \Delta(I)$ from which several faces
are removed: namely such faces $\tilde \Delta(J)$, $J \subset I,$ that the
plane $L_J$ is strictly greater than $L_I$. For instance, for the
configurations shown in pictures a), b), c) of Fig.~\ref{arr} the planes $L_I$
of codimension 2 are: the point $(1,2)$, the point $(1,2,3)$, and three points
$(1,2),$ $(1,3)$, $(2,3)$ respectively. The proper preimages of them are: one
segment, one triangle (shadowed vertically in the picture), and three segments.
In all these cases the corresponding spaces $\check L_I$ coincide with $\check
\Delta_I$, namely they are: an open interval, a triangle without vertices, and
three open intervals, respectively.
In general, any face of the simplex $\tilde \Delta(I)$ is characterized by its
vertices, i.e. some indices $i \in \{1, \ldots, k\}$. A face of $\tilde
\Delta(I)$ is called {\em marginal}\label{marginal} if the intersection of
planes $L_i$ labeled by its vertices is strictly greater than $L_I.$ $\check
\Delta(I)$ is equal to $\tilde \Delta(I)$ with all marginal faces removed. By
the K\"unneth formula, $E^1_{p,q}= \bigoplus \bar H_{p+q-(N-p)}(\check
\Delta(I)),$ summation over all planes $L_I$ of codimension $p$.
The geometrical sense of the corresponding filtration in the Alexander dual
group $H^*(\R^N \setminus L)$ is as follows: any element of this group has
filtration $p$ if and only if it is equal to a linear combination of finitely
many elements $\gamma_j$, any of which can be represented by the intersection
index with some semilinear\footnote{= semialgebraic distinguished by only
linear equations and inequalities} subvariety $V_j \subset \R^N,$ $\partial V_j
\subset L,$ invariant under the group $\R^{N-p_j}$ of translations in all
directions parallel to some $(N-p_j)$-dimensional plane $L_{I_j}$ with $p_j \le
p$.
\begin{proposition}[see \cite{congr}]
Our filtration of the space $\overline{L'}$ always homotopically splits, i.e.
we have the homotopy equivalence
\begin{equation}
\bar L' \sim \bar F_1 \vee (\bar F_2/\bar F_1) \vee \ldots \vee (\bar F_N/ \bar
F_{N-1}). \label{splitt}
\end{equation}
In particular, the spectral sequence degenerates in the first term: $E^\infty
\equiv E^1$, and we have $\bar H_{p+q}(\bar L') = \oplus_{p=1}^N E_{p,q}^1$.
\quad $\square$
\end{proposition}
An equivalent statement was proved in \cite{ZZ}.
\medskip
This theorem reduces the structure of cohomology groups of $\R^N \setminus L$
to dimensions of all spaces $L_I$. However, it does not allow us to calculate
the value of an arbitrary element of the group $E_{p,q}^1$ on any cycle in
$\R^N \setminus L$. For instance, in the case of the arrangement shown in
Fig.~\ref{arr}a, the entire group $E_{2,*}^1$ appears from the unique crossing
point $L_{\{1,2\}}$. This group is nontrivial only for $*=-1$, is isomorphic to
$\Z$ and is generated by the homology class of the segment $\Delta(1,2)$ modulo
its endpoints (lying in $F_1$). The splitting formula (\ref{splitt}) means that
we {\em can} extend this relative cycle of $\bar F_2 \ (\mbox{mod} \ \bar F_1)$
(or, equivalently, a closed locally finite cycle in $F_2 \setminus F_1$) to a
cycle in entire $\bar L'$ (respectively, in entire $L'$). However, to be able
to define the value of this point or of this segment on any $0$-dimensional
cycle (i.e. on a point) in $\R^2 \setminus L$ we need to {\em choose} such an
extension explicitly. Then we project it to $L$ and get a cycle there. Finally,
we need to choose a relative cycle in $\R^2 \ (\mbox{mod} \ L)$ whose boundary
coincides with this cycle. Then we call this relative cycle "a combinatorial
formula": its value on a point in $\R^2 \sm L$ is equal to the multiplicity of
this cycle in the neighborhood of this point.
If we have a more complicated plane arrangement, then the most convenient way
to extend an element of $E^1_{p,q}$ to a closed cycle in $L'$ is to do it step
by step over our filtration. Our starting element $\gamma \in E_{p,q}^1$ is
represented by a cycle with closed supports in $F_p \setminus F_{p-1}$. We take
its {\em first boundary} $d_1(\gamma)$, which is a cycle in $F_{p-1} \
(\mbox{mod} \ F_{p-2})$. Then we {\em span} it, i.e. construct a closed chain
$\tilde \gamma_1 $ in $ F_{p-1} \setminus F_{p-2}$ such that $\partial \tilde
\gamma_1 = d_1 \gamma $ there. Then we take the boundary of $\gamma+\tilde
\gamma_1$ in the space $F_{p-2} \setminus F_{p-3}$ and span it there by a chain
$\tilde \gamma_2,$ etc. The degeneration formula (\ref{splitt}) ensures that
all this sequence of choices can be accomplished. See \cite{ZZ}, \cite{M} for
some precise algorithms of doing it in the case of plane arrangements.
\subsection{All the same for knots}
The case of knots (say, of long knots) is very similar to that of plane
arrangements. A list of parallel notions is given in Table \ref{glos} (whose
left part was explained in the previous subsection, and the right-hand part
will be explained in the present one).
\begin{table}
\begin{tabular}{|l|l|}
\hline
Space $\R^N$ & Space ${\mathcal K}_n$ of smooth maps $\R^1 \to \R^n$ \\
& with a fixed behavior at $\infty$ \\
\hline Union of planes $L=\cup L_i \subset \R^N$ & Discriminant subset
$\Sigma \subset {\mathcal K}_n$ \\
\hline
Set of indices $\{1, \ldots, k\}$ & Chord space $\overline{B(\R^1,2)}$ \\
\hline
A plane $L_i$ & A subspace $L(a,b)$, $a,b \in \R^1$ \\
\hline Disjoint union of hyperplanes $L_i$ & Tautological resolution $F_1
\sigma$
of $\Sigma$ \\
\hline Simplicial resolution $L'$ of $L$ & Simplicial resolution
$\sigma$ of $\Sigma$ \\
\hline Subsets $I \subset \{1, \ldots, k\}$ & Combinatorial types of chord
configurations $J$ \\
with codim$L_I =p$ & with codim$L(J) = pn$ \\
\hline
A prism $L'_I$ & A $J$-block in $\sigma$ \\
\hline K\"unneth isomorphism for &
Thom isomorphism for the fibration of \\
homology of $\check L'_I = \check \Delta(I) \times L_I$ &
pure $J$-blocks by spaces $L(J')$ \\
\hline
Homotopy splitting (\ref{splitt}) & Kontsevich's degeneration theorem \\
\hline
\end{tabular}
\caption{} \label{glos}
\end{table}
So, instead of $\R^N$ we consider the affine space ${\mathcal K}_n$ of all
smooth maps $\R^1 \to \R^n$ coinciding with a fixed linear embedding "at
infinity", and instead of $L$ the discriminant space $\Sigma \subset {\mathcal
K}_n$ of all such maps which are not smooth embeddings.
Of course, the space ${\mathcal K}_n$ is infinite-dimensional, and formally we
cannot use the Alexander duality in it: the (finite-dimensional) cohomology
classes of the space of knots ${\mathcal K}_n \setminus \Sigma$ should
correspond to "infinite-dimensional cycles" in $\Sigma$, whose definition
requires some effort. The strict definition of such cycles corresponding to
finite-type cohomology classes was proposed in \cite{V1} and is as follows. We
consider a sequence of finite-dimensional approximating subspaces ${\mathcal
K}_n^j$ in ${\mathcal K}_n$, calculate (some) cohomology classes of ${\mathcal
K}^j_n \setminus \Sigma$ dual to certain cycles in $\Sigma$, and then prove a
stabilization theorem when $j \to \infty.$ It follows from the Weierstrass
approximation theorem that these stable cocycles are well-defined cohomology
classes in ${\mathcal K}_n \setminus \Sigma$. A rigorous reader can either read
\cite{V1} or \cite{fasis} for all justifications or to think of the spaces
${\mathcal K}_n$ as of such approximating spaces of very high but finite
dimension. Let us denote this virtual dimension of ${\mathcal K}_n$ by
$\omega.$
Again, $\Sigma$ is the union of a family of subspaces of very simple nature.
For any pair of points $(a,b)$ in $\R^1$, denote by $L(a,b)$ the space of all
maps $f \in {\mathcal K}_n$ such that
\begin{equation}
\label{cond} f(a)=f(b) \mbox{ (if }a\ne b\mbox{ ) or }f'(a)=0\mbox{ (if }a=b).
\end{equation}
Such spaces form a 2-parametric family parametrized by all points $(a,b)$ of
the space $\overline{B(\R^1,2)}$ of all unordered collections of two points in
$\R^1.$ Since \cite{V1} such pairs are depicted by arcs connecting the points
$a,b$ (called {\em chords} in \cite{BN}), so the space $\overline{B(\R^1,2)}$
will be called here the {\em chord space}. Its degenerated points
(corresponding to pairs $a=b$) are depicted by an asterisk at the point $a$.
The {\em tautological resolution} $F_1\sigma$ of $\Sigma$ is constructed as a
subspace of the direct product $\overline{B(\R^1,2)} \times {\mathcal K}_n $:
this is the space of pairs $((a,b),f)$ satisfying (\ref{cond}). It is the space
of an $(\omega-n)$-dimensional vector bundle over $\overline{B(\R^1,2)}.$
Therefore by the Thom isomorphism we have $\bar H_*(F_1\sigma)\simeq \bar
H_{*-(\omega-n)}(\overline{B(\R^1,2)}) \equiv 0:$ indeed,
$\overline{B(\R^1,2)}$ is homeomorphic to the closed half-plane. There is
obvious projection $F_1 \sigma \to \Sigma$; it is a map onto, and close to
generic points of $\Sigma$ is a homeomorphism.
Further, we insert simplices spanning preimages of non-generic points of
$\Sigma$. As previously, we embed the space $\overline{B(\R^1,2)}$ generically
and algebraically into a space $\R^T$ of a huge dimension ($T \approx \omega^3
$). Then for any point $f \in \Sigma$ we mark all the points $(a,b) \in
\overline{B(\R^1,2)}$ such that $L(a,b)\ni f,$ and denote by $\tilde \Delta(f)$
the convex hull of images of all these points in $\R^T.$
Of course, there exist points $f \in \Sigma$ having infinitely many preimages.
However they form a subset of infinite codimension in ${\mathcal K}_n$, and we
can ignore them by considering only finite-dimensional approximations
${\mathcal K}_n^j$ in general position with the stratification of $\Sigma$.
Then all the sets $\tilde \Delta(f),$ $f \in {\mathcal K}_n^j,$ still will be
the simplices with vertices at the images of all corresponding points $(a,b)$
of the chord space. The simplicial resolution $\sigma \subset \R^T \times
{\mathcal K}_n$ is defined as the union of all simplices $\Delta(f) \equiv
\tilde \Delta(f) \times \{f\}$.
Again, $\sigma$ has a useful increasing filtration. Let $I\subset
\overline{B(\R^1,2)}$ be a finite set of chords $(a,b)$. The intersection of
corresponding planes $L(a,b)$ is a subspace $L(I) \subset {\mathcal K}_n,$
whose codimension is a multiple of $n$. The proportionality coefficient
$\mbox{codim}L(I)/n$ is called the {\em complexity} of $I$. Consider all the
points $(a,b) \in \overline{B(\R^1,2)}$ such that $L(a,b) \supset L(I),$ and
denote by $\tilde \Delta(I) \subset \R^T$ the convex hull of images of all
these points. (It is equal to the space $\tilde \Delta(f)$ where $f$ is a {\em
generic} point of the space $L(I)$.) Set $L'(I)=\tilde \Delta(I) \times L(I)
\subset \R^T \times {\mathcal K}_n.$ Finally, define the term $F_p(\sigma)$ of
the filtration as the union of all simplices $\Delta(I)$ over all $I$ of
complexity $\le p.$
\begin{definition}
\label{fitype} {\rm A cohomology class of the space of knots ${\mathcal K}_n
\setminus \Sigma$ is a {\it finite type class of order} $p$ if it can be
defined as the linking number with the direct image in $\Sigma$ of a cycle
(with closed support) lying in the term $F_p$ of the standard filtration of
$\sigma$. For any such class of order $p$ and dimension $d$, its {\em principal
part} is the class of the corresponding cycle in the group $\bar
H_{\omega-d-1}(F_p \setminus F_{p-1})$.}
\end{definition}
The important property of this filtration is as follows: any its term $F_p
\setminus F_{p-1}$ is the space of an $(\omega-pn)$-dimensional affine bundle
over some finite-dimensional semialgebraic base: the projection of this bundle
is induced by the obvious projection $\R^T \times {\mathcal K}_n \to \R^T.$ In
particular, the Thom isomorphism reduces the Borel--Moore homology group of
this term to the homology group of locally finite chains of this base (in the
case of odd $n$ with coefficients in some system of groups locally isomorphic
to $\Z$, which is constant only for $p=1$).
These finite-dimensional bases, and hence also entire spaces $F_p \setminus
F_{p-1}$ of our filtration, admit an easy description, in particular their
one-point compactifications have a natural structure of $CW$-complexes. First
let us describe all the spaces $L(I)$ of complexity exactly $p$.
\begin{definition}[see \cite{V1}]{\rm
Let $A$ is a unordered finite collection of naturals $A=(a_1, \ldots, a_{\#
A}),$ $a_j \ge 2,$ and $b$ any nonnegative integer. Then an $(A,b)$-{\em
configuration} in $\R^1$ is any collection of distinct $a_1+ \cdots + a_{\# A}$
points in $\R^1$ separated into groups of cardinalities $a_1, \ldots, a_{\#
A}$, plus a collection of $b$ distinct points in $\R^1$ (some of which can
coincide with the points of the $A$-part). A map $f:\R^1 \to \R^n$ {\em
respects} an $(A,b)$-configuration $J$ if it maps all points of any of groups
of cardinality $a_j,$ $j=1, \ldots, \# A,$ into one point (these points for
different groups may coincide), and $f'=0$ at all points of the $b$-part of the
configuration. The space of all maps $f$ respecting a fixed
$(A,b)$-configuration $J$ is denoted by $L(J)$. Two $(A,b)$-configurations are
{\em equivalent} if they can be transformed one into the other by an
orientation-preserving homeomorphism of $\R^1$. The {\em complexity} of an
$(A,b)$-configuration is the number $\sum_{j=1}^{\# A}(a_j-1)+b.$}
\end{definition}
Obviously the codimension in ${\mathcal K}_n$ of any space $L(J)$ is equal to
$n$ times the complexity of $J$. The space of all $(A,b)$-configurations of
complexity $1$ is the chord space $\overline{B(\R^1,2)}$. Any intersection of
finitely many planes $L(a,b)$, $(a,b) \in\overline{B(\R^1,2)},$ is a plane of
form $L(J)$ for some $(A,b)$-configuration $J$. The corresponding simplex
$\tilde \Delta(J)$ in $\R^T$ has exactly $\sum_{i=1}^{\# A}\binom{a_i}{2} +b$
vertices. The $J$-{\em block} $\square(J)$ in $\sigma$ is the union of all
pairs $(x,f) \subset \R^T \times {\mathcal K}_n,$ such that $x$ belongs to the
simplex $\tilde \Delta(J')$ for some $(A,b)$-configuration $J'$ equivalent to
$J$, and $f$ respects this configuration $J'$. It belongs to the term $F_p$ of
our filtration, where $p$ is the complexity of $J$.
The {\em pure $J$-block} $\Check {\square}(J)$ is equal to $\square(J)
\setminus F_{p-1}$. It is fibered over the space of $(A,b)$-configurations $J'$
equivalent to $J$, with fiber equal to $\check \Delta(J) \times L(J)$, where
$\check \Delta(J)$ is the union of several (non-marginal in some sense) faces
of $\tilde \Delta(J)$. The base of this fiber bundle is an open cell, thus the
bundle is trivial, and we have a canonical decomposition of $\Check
{\square}(J)$ into open cells corresponding to all such non0marginal faces.
The canonical notation of any such cell is a {\em generalized chord diagram},
i.e. a finite collection of arcs connecting some points of $\R^1$ and of
asterisks marking some points, say as in the picture
\begin{equation}
\label{samcell} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(80.00,9.00)
\put(15.00,3.00){\oval(20.00,4.00)[t]} \put(30.00,3.00){\oval(10.00,4.00)[t]}
\put(20.00,3.00){\oval(30.00,8.00)[t]} \put(30.00,3.00){\oval(30.00,6.00)[b]}
\put(55.00,3.00){\oval(20.00,6.00)[b]} \put(55.00,3.00){\makebox(0,0)[cc]{$*$}}
\put(45.00,3.00){\makebox(0,0)[cc]{$*$}}
\put(50.00,3.00){\oval(50.00,12.00)[t]} \put(80.00,3.00){\line(-1,0){80.00}}
\end{picture}
\end{equation}
presenting one of cells of a certain equivalence class of
$((4,3),2)$-configurations. Namely, for any such cell related with a class of
equivalent $(A,b)$-configurations, we fix some configuration $J \subset \R^1$
of this class, mark by asterisks all points of its $b$-part ("singular points")
and draw a chord between any its two points $a,b$ such that the point $(a,b)
\in \overline{B(\R^1,2)}$ is a vertex of the face of $\tilde \Delta(J)$
corresponding to this cell.
The space $F_p \setminus F_{p-1}$ is the union of such pure blocks $\Check
{\square}(J)$ over (finitely many) equivalence classes of all
$(A,b)$-configurations of complexity exactly $p$. So we get also the
decomposition of this space into finitely many open cells. This decomposition
can be extended to the structure of a $CW$-complex on the one-point
compactification of $F_p \setminus F_{p-1}$. Its structure and incidence
coefficients are explicitly described in \cite{V1}, \cite{fasis}, which gives
also an algorithm for calculating the term $E^1$ of the spectral sequence
generated by this filtration and converging to the group of all finite-type
cohomology classes of the space of knots. In particular, if $n=3$ then all
$J$-blocks of complexity $p$ which (by dimensional reasons) can be valuable for
the calculation of knot invariants, are only the blocks with $(A,b)$ equal to
$((2,\ldots,2),0)$ (chord diagrams), $((2,\ldots,2),1)$ ({\em one-term
relations}, see \cite{BL}, \cite{BN}), and $((3,2,\ldots,2),0)$ (any such block
corresponds to the totality of {\em 4-term relations} arising from the
neighborhood of a triple point: there are three such relations, any two of
which are independent).
\begin{example}
\label{exf1} {\rm The term $F_1$ consists of exactly two cells, one of which is
the boundary of the other:
\begin{equation}
\label{dif01} \unitlength 1mm \linethickness{0.4pt}
\begin{picture}(72.00,4.00)
\put(3.00,0.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,0.00){\line(1,0){30.00}} \put(39.00,0.00){\makebox(0,0)[cc]{$=$}}
\put(42.00,0.00){\line(1,0){30.00}} \put(21.00,0.00){\oval(20.00,8.00)[t]}
\put(57.00,0.00){\makebox(0,0)[cc]{{\large $*$}}}
\end{picture} \ ,
\end{equation}
thus there are no cohomology classes of order 1 of the space of long knots.
(Here and in the next example we assume some natural orientations of such
cells, see \cite{V1}, \cite{fasis}.)}
\end{example}
\begin{example}
\label{exf2} {\rm The term $F_2 \setminus F_1$ consists of the following cells:
four cells of maximal dimension
\begin{equation}
\unitlength 1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(16.00,5.00)
\put(0.00,2.00){\line(1,0){16.00}} \put(6.50,2.00){\oval(9.00,4.00)[b]}
\put(9.50,2.00){\oval(9.00,4.00)[t]}
\end{picture}
\ , \
\unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(39.00,4.00)
\put(0.00,2.00){\line(1,0){16.00}} \put(3.50,2.00){\oval(5.00,4.00)[t]}
\put(12.50,2.00){\oval(5.00,4.00)[t]} \put(19.00,0.00){\makebox(0,0)[cc]{,}}
\put(23.00,2.00){\line(1,0){16.00}} \put(31.00,2.00){\oval(14.00,4.00)[t]}
\put(31.00,2.00){\oval(6.00,4.00)[b]}
\end{picture}
\ , \
\unitlength 1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(18.00,4.00)
\put(0.00,0.00){\line(1,0){16.00}} \put(8.00,0.00){\oval(14.00,8.00)[t]}
\put(4.50,0.00){\oval(7.00,4.00)[t]} \put(11.50,0.00){\oval(7.00,4.00)[t]}
\end{picture}
\label{maxcel}
\end{equation}
(only the first and the last of which will be interesting for us), three cells
forming the boundary of any of these two cells,
\begin{equation}
\label{ztripno} \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(20.00,4.00)
\put(0.00,0.00){\line(1,0){20.00}} \put(10.00,0.00){\oval(16.00,8.00)[t]}
\put(14.00,0.00){\oval(8.00,4.00)[t]}
\end{picture}
\ , \quad
\unitlength 1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(20.00,4.00)
\put(0.00,0.00){\line(1,0){20.00}} \put(14.00,0.00){\oval(8.00,4.00)[t]}
\put(6.00,0.00){\oval(8.00,4.00)[t]}
\end{picture}
\ , \quad
\mbox{and} \quad \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(20.00,4.00)
\put(0.00,0.00){\line(1,0){20.00}} \put(10.00,0.00){\oval(16.00,8.00)[t]}
\put(6.00,0.00){\oval(8.00,4.00)[t]}
\end{picture} \ ,
\end{equation}
and also six cells defined by maps with singular points (i.e. in whose notation
the asterisks $*$ participate):
\begin{equation}
\label{marg} \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(119.00,3.00)
\put(1.00,1.00){\line(1,0){15.00}} \put(21.00,1.00){\line(1,0){15.00}}
\put(41.00,1.00){\line(1,0){15.00}} \put(61.00,1.00){\line(1,0){15.00}}
\put(81.00,1.00){\line(1,0){15.00}} \put(101.00,1.00){\line(1,0){15.00}}
\put(10.50,1.00){\oval(7.00,4.00)[t]} \put(4.00,1.00){\makebox(0,0)[cc]{$*$}}
\put(28.50,1.00){\oval(11.00,4.00)[t]} \put(29.00,1.00){\makebox(0,0)[cc]{$*$}}
\put(46.50,1.00){\oval(7.00,4.00)[t]} \put(54.00,1.00){\makebox(0,0)[cc]{$*$}}
\put(68.50,1.00){\oval(9.00,4.00)[t]} \put(64.00,1.00){\makebox(0,0)[cc]{$*$}}
\put(88.50,1.00){\oval(9.00,4.00)[t]} \put(93.00,1.00){\makebox(0,0)[cc]{$*$}}
\put(104.00,1.00){\makebox(0,0)[cc]{$*$}}
\put(113.00,1.00){\makebox(0,0)[cc]{$*$}}
\put(18.00,1.00){\makebox(0,0)[cc]{,}} \put(38.00,1.00){\makebox(0,0)[cc]{,}}
\put(58.00,1.00){\makebox(0,0)[cc]{,}} \put(78.00,1.00){\makebox(0,0)[cc]{,}}
\put(98.00,1.00){\makebox(0,0)[cc]{,}} \put(119.00,1.00){\makebox(0,0)[cc]{.}}
\end{picture}
\end{equation}
The boundary operator in this term $F_2 \setminus F_1$ (i.e. the vertical
differential $d^0$ of the spectral sequence) acts as follows:
$$
\unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(125.00,53.00)
\put(3.00,34.00){\makebox(0,0)[cc]{$\partial$}}
\put(7.00,34.00){\line(1,0){22.00}} \put(33.00,34.00){\makebox(0,0)[cc]{$=$}}
\put(45.50,34.00){\line(1,0){21.00}} \put(125.00,34.00){\line(-1,0){22.00}}
\put(96.00,34.00){\line(-1,0){22.00}}
\put(3.00,49.00){\makebox(0,0)[cc]{$\partial$}}
\put(7.00,49.00){\line(1,0){22.00}} \put(33.00,49.00){\makebox(0,0)[cc]{$=$}}
\put(40.00,49.00){\line(1,0){22.00}} \put(125.00,49.00){\line(-1,0){22.00}}
\put(94.00,49.00){\line(-1,0){22.00}}
\put(3.00,4.00){\makebox(0,0)[cc]{$\partial$}}
\put(7.00,4.00){\line(1,0){22.00}} \put(33.00,4.00){\makebox(0,0)[cc]{$=$}}
\put(40.00,4.00){\line(1,0){22.00}} \put(125.00,4.00){\line(-1,0){22.00}}
\put(94.00,4.00){\line(-1,0){22.00}}
\put(3.00,19.00){\makebox(0,0)[cc]{$\partial$}}
\put(7.00,19.00){\line(1,0){22.00}} \put(33.00,19.00){\makebox(0,0)[cc]{$=$}}
\put(45.50,19.00){\line(1,0){21.00}} \put(125.00,19.00){\line(-1,0){22.00}}
\put(96.00,19.00){\line(-1,0){22.00}} \put(15.00,49.00){\oval(12.00,8.00)[t]}
\put(21.00,49.00){\oval(12.00,8.00)[b]}
\put(37.00,49.00){\makebox(0,0)[cc]{$-$}}
\put(50.50,49.00){\oval(15.00,8.00)[t]} \put(47.00,49.00){\oval(8.00,8.00)[b]}
\put(67.00,49.00){\makebox(0,0)[cc]{$+$}}
\put(78.50,49.00){\oval(9.00,8.00)[t]} \put(87.50,49.00){\oval(9.00,8.00)[t]}
\put(99.00,49.00){\makebox(0,0)[cc]{$-$}}
\put(114.00,49.00){\oval(18.00,8.00)[t]}
\put(118.50,49.00){\oval(9.00,8.00)[b]} \put(12.00,34.00){\oval(6.00,8.00)[t]}
\put(24.00,34.00){\oval(6.00,8.00)[t]}
\put(40.00,34.00){\makebox(0,0)[cc]{$(-1)^n$}}
\put(48.00,34.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(60.00,34.00){\oval(8.00,8.00)[t]}
\put(70.00,34.00){\makebox(0,0)[cc]{$+$}}
\put(80.50,34.00){\oval(9.00,8.00)[t]} \put(89.50,34.00){\oval(9.00,8.00)[t]}
\put(100.00,34.00){\makebox(0,0)[cc]{$-$}}
\put(109.50,34.00){\oval(9.00,8.00)[t]}
\put(123.00,34.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(18.00,19.00){\oval(18.00,8.00)[t]} \put(18.00,19.00){\oval(6.00,8.00)[b]}
\put(40.00,19.00){\makebox(0,0)[cc]{$(-1)^n$}}
\put(55.50,19.00){\oval(15.00,8.00)[t]} \put(52.00,19.00){\oval(8.00,8.00)[b]}
\put(70.00,19.00){\makebox(0,0)[cc]{$+$}}
\put(85.00,19.00){\oval(18.00,8.00)[t]}
\put(85.00,19.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(100.00,19.00){\makebox(0,0)[cc]{$-$}}
\put(114.00,19.00){\oval(18.00,8.00)[t]}
\put(118.50,19.00){\oval(9.00,8.00)[b]} \put(114.00,4.00){\oval(18.00,8.00)[t]}
\put(118.50,4.00){\oval(9.00,8.00)[b]} \put(99.00,4.00){\makebox(0,0)[cc]{$+$}}
\put(78.50,4.00){\oval(9.00,8.00)[t]} \put(87.50,4.00){\oval(9.00,8.00)[t]}
\put(67.00,4.00){\makebox(0,0)[cc]{$-$}} \put(50.50,4.00){\oval(15.00,8.00)[t]}
\put(47.00,4.00){\oval(8.00,8.00)[b]} \put(18.00,4.00){\oval(18.00,8.00)[t]}
\put(13.50,4.00){\oval(9.00,4.00)[t]} \put(22.50,4.00){\oval(9.00,4.00)[t]}
\end{picture}
$$
$$
\unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(105.00,38.00)
\put(3.00,34.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,34.00){\line(1,0){25.00}} \put(34.00,34.00){\makebox(0,0)[cc]{$=$}}
\put(47.00,34.00){\line(1,0){25.00}} \put(59.50,34.00){\oval(19.00,8.00)[t]}
\put(3.00,19.00){\makebox(0,0)[cc]{$\partial$}}
\put(3.00,4.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,19.00){\line(1,0){25.00}} \put(6.00,4.00){\line(1,0){25.00}}
\put(34.00,19.00){\makebox(0,0)[cc]{$=$}}
\put(34.00,4.00){\makebox(0,0)[cc]{$=$}} \put(47.00,19.00){\line(1,0){25.00}}
\put(40.00,4.00){\line(1,0){25.00}} \put(59.50,19.00){\oval(19.00,8.00)[t]}
\put(52.50,4.00){\oval(19.00,8.00)[t]} \put(18.50,34.00){\oval(19.00,8.00)[t]}
\put(14.00,34.00){\oval(10.00,8.00)[b]}
\put(42.00,34.00){\makebox(0,0)[cc]{$(-1)^n$}}
\put(50.00,34.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(92.50,19.00){\oval(19.00,8.00)[t]} \put(80.00,19.00){\line(1,0){25.00}}
\put(76.00,19.00){\makebox(0,0)[cc]{$+$}}
\put(42.00,19.00){\makebox(0,0)[cc]{$(-1)^n$}}
\put(50.00,19.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(102.00,19.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(62.00,4.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(18.50,4.00){\oval(19.00,8.00)[t]} \put(23.50,4.00){\oval(9.00,8.00)[b]}
\put(23.50,19.00){\oval(9.00,8.00)[t]} \put(14.00,19.00){\oval(10.00,8.00)[t]}
\end{picture}
$$
$$
\unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(112.00,38.00)
\put(3.00,34.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,34.00){\line(1,0){25.00}} \put(34.00,34.00){\makebox(0,0)[cc]{$=$}}
\put(42.00,34.00){\line(1,0){25.00}} \put(73.00,34.00){\line(1,0){25.00}}
\put(85.50,34.00){\oval(19.00,8.00)[t]}
\put(95.00,34.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(3.00,19.00){\makebox(0,0)[cc]{$\partial$}}
\put(3.00,4.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,19.00){\line(1,0){25.00}} \put(6.00,4.00){\line(1,0){25.00}}
\put(34.00,19.00){\makebox(0,0)[cc]{$=$}}
\put(34.00,4.00){\makebox(0,0)[cc]{$=$}} \put(42.00,19.00){\line(1,0){25.00}}
\put(42.00,4.00){\line(1,0){25.00}} \put(73.00,19.00){\line(1,0){25.00}}
\put(87.00,4.00){\line(1,0){25.00}} \put(85.50,19.00){\oval(19.00,8.00)[t]}
\put(95.00,19.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(90.00,4.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(109.00,4.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(45.00,34.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(64.00,34.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(14.00,34.00){\oval(10.00,8.00)[t]} \put(18.50,19.00){\oval(19.00,8.00)[t]}
\put(28.00,34.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(19.00,19.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(9.00,4.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(45.00,4.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(54.50,4.00){\oval(19.00,8.00)[t]}
\put(45.00,19.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(54.50,19.00){\oval(19.00,8.00)[t]}
\put(80.00,4.00){\makebox(0,0)[cc]{$(-1)^{n-1}$}}
\put(70.00,4.00){\makebox(0,0)[cc]{$+$}}
\put(70.00,19.00){\makebox(0,0)[cc]{$+$}}
\put(70.00,34.00){\makebox(0,0)[cc]{$+$}}
\put(39.00,34.00){\makebox(0,0)[cc]{$-$}}
\put(39.00,19.00){\makebox(0,0)[cc]{$-$}}
\put(39.00,4.00){\makebox(0,0)[cc]{$-$}} \put(23.50,4.00){\oval(9.00,8.00)[t]}
\end{picture} .
$$
In particular there is exactly one nontrivial group $\bar H_i(F_2 \setminus
F_1) \equiv E^1_{2,i-2}$, namely such group with $i=\omega-(2n-5)$ is
isomorphic to $\Z$ and is generated by the sum of the first and the last cells
in (\ref{maxcel}).}
\end{example}
Thus we obtain a proof of the first statement of Proposition \ref{longres}. The
group $\bar H_*(F_3 \setminus F_2)$ of (possible) principal parts of third
order cohomology classes was calculated in \cite{tetu} for odd $n$ (another
proof, not relying on the computer's honesty, see in \cite{bjo}) and in
\cite{bjo} for even $n$. In both cases, there are exactly two nontrivial groups
$\bar H_i(F_3 \setminus F_2) \simeq E^1_{3,i-3}$, namely with $i=\omega
-(3n-8)$ and $\omega-(3n-7)$; they both are isomorphic to $\Z$. By the
dimensional reasons both these groups for any $n \ge 3$ coincide with
corresponding groups $E^\infty_{3,i-3}$, and their generators extend to
well-defined cohomology classes of spaces of long knots in $\R^n$. By similar
considerations (see e.g. \cite{bjo}) for $n \ge 4$ these generators are
nontrivial and free. If $n=3$ then for the first of these classes the same
follows from the fact that it coincides with the second simple knot invariant
(calculated in \cite{V1}) whose nontriviality is well known. The other class is
exactly the Teiblum--Turchin class studied below; the fact that it also is
nontrivial and free for $n=3$ will follow from the proof of Corollary
\ref{nontriv}.
\begin{remark} {\rm
It is often convenient to replace formally our homological spectral sequence
calculating $\bar H_*(\sigma)$ by the "Alexander dual" cohomological spectral
sequence
$$
E_r^{p,q} \equiv E^r_{-p,\omega-q-1}.
$$
It lies in the second quadrant in the wedge $\{(p,q)| p \le 0, q +pn \ge 0\}$
and converges to some subgroup of the group $H^*(\K_n \setminus \Sigma)$ (if
$n>3$ then to entire this group).}
\end{remark}
\begin{remark} {\rm
There are beautiful algebraic structures on the above-described spectral
sequence, and hence on its limit filtered group $H^*(\K_n \setminus \Sigma)$
and the corresponding adjoint graded group, see \cite{Turchin}.}
\end{remark}
All of this theory can be extended almost literally to the cohomology of the
space of compact knots $S^1 \hookrightarrow \R^n$.\footnote{As well as of
compact links, i.e. of embeddings of a disjoint union of several circles into
$\R^n$; we shall not discuss here the latter theory} However, in this case the
chord space $\overline{B(S^1,2)}$ is not topologically trivial (it is a closed
M\"obius band); also the spaces of equivalent $(A,b)$-configurations are not
the cells. To get the cell decomposition of all spaces $F_p \setminus F_{p-1}$
we need to mark one point in $S^1$ ("the origin") and call two configurations
equivalent if they are transformed one into the other by a homeomorphism of
$S^1$ preserving the origin and the orientation.
\medskip
The direct calculation of the spectral sequence and obtaining the {\em
combinatorial formulas} for the finite-type cohomology classes of spaces of
knots is formally the same process as in the case of plane arrangements.
However, the exact choice of the spanning chains in all the consecutive terms
of the filtration and in entire ${\mathcal K}_n$ depends very much of the
features of the knot space.
\begin{remark}{\rm
There is another, sometimes more convenient construction of the resolution of
discriminant sets, namely the {\em conical resolutions} based on the notion of
a continuous order complex of a topologized partially ordered set, see e.g.
\cite{kotor}. In particular it allows us to resolve the points of $\Sigma$ with
infinitely many preimages in the tautological resolution space. However, for
the calculations in the present work it will be enough to use the "naive"
simplicial resolution described above.}
\end{remark}
\subsection{Finite type knot invariants and Polyak-Viro combinatorial
formulas}
Suppose that $n=3$ and we are interested in the knot invariants, i.e. the
0-dimensional cohomology classes of $\K_n \setminus \Sigma$. For any such class
of finite filtration $p$, its principal part in $F_p \setminus F_{p-1}$ is a
linear combination of cells depicted by $p$-chord diagrams (i.e. collections of
$p$ chords with distinct endpoints) and $\tilde p$-{\em configurations}, i.e.
collections of $p-2$ chords with different endpoints and one triple of points
joined by three chords. E.g., among all diagrams in
(\ref{samcell})--(\ref{marg}) only the left picture in (\ref{dif01}) and three
left pictures in (\ref{maxcel}) are chord diagrams, and only the last picture
in (\ref{maxcel}) is a $\tilde 2$-configuration. The coefficients with which
all these cells can enter the linear combination satisfy the homological
condition. In particular the coefficients at $\tilde p$-configurations are
determined by these at $p$-chord diagrams, and any admissible linear
combination is characterized uniquely only by the collection of latter
coefficients, which is called a {\em weight system}.
{\em The elementary characterization of these invariants} is as follows (see
e.g. \S 0.2 in \cite{V1}). Let us consider any immersion $\R^1 \to \R^3$ with
exactly $k$ transverse self-intersection points. We can resolve any of these
points in two locally distinct ways to get a knot without intersections. One of
these two local resolutions can be invariantly defined as a positive, and the
other as the negative one. The $k$-th index of a knot invariant at our singular
immersion is equal to the alternated sum of its values at all $2^k$ knots
obtained by all different possible resolutions of double points: the value at a
knot is counted with sign 1 or $-1$ depending on the parity of the number of
positive local resolutions. A knot invariant is of filtration $p$ if and only
if all its indices at all immersions with $k>p$ self-intersections are equal to
0. The same definition can be applied to define the filtration of invariants of
compact knots $S^1 \to \R^3$. On the other hand, it is easy to see that there
is a natural one-to-one correspondence between connected components of spaces
of long and compact knots, in particular the theories of their invariants
naturally coincide.
Some combinatorial formulas for the simplest finite-type knot invariants --- of
orders 2 and 3 --- were found in \cite{lannes}. Another, more convenient
formulas were introduced by M.~Polyak and O.~Viro in \cite{PV}. These formulas
for long knots look as the linear combinations of chord diagrams with oriented
chords. E.g. the formula \unitlength 1mm \linethickness{0.4pt}
\begin{picture}(22.00,6.00)
\put(0.00,3.00){\line(1,0){22.00}} \put(2.00,3.00){\vector(-4,-3){0.2}}
\bezier{68}(14.00,3.00)(8.00,9.00)(2.00,3.00)
\put(20.00,3.00){\vector(4,3){0.2}}
\bezier{68}(8.00,3.00)(14.00,-3.00)(20.00,3.00)
\end{picture} \ should be read as follows. Consider a generic
long knot $f : \R^1 \to \R^3$. A {\em representation} of the above picture in
this knot is any collection of points $a<b<c<d \subset \R^1$ such that $f(a)$
lies below $f(c)$ and $f(d)$ lies below $f(b)$. The value of this picture on
our knot is equal to the number of its representations (counted with
appropriate signs). An immediate calculation shows that this number is a knot
invariant of order 2.
In the case of compact knots, there are Polyak-Viro formulas of two types:
absolute and punctured ones. They also look as oriented chord diagrams, but
with endpoints in the oriented circle $S^1$ instead of $\R^1$; moreover, a
punctured Polyak-Viro diagram contains a point in $S^1$ not coinciding with the
endpoints of chords. E.g. a representation of the diagram \unitlength 1mm
\linethickness{0.4pt}
\begin{picture}(8.00,8.00)
\put(4.00,4.00){\circle{8.00}} \put(7.00,1.00){\vector(-1,1){6.00}}
\put(1.00,1.00){\vector(1,1){6.00}}
\end{picture}
in a compact knot is any collection of four points
in $S^1$ with cyclic order $a<b<c<d<a$, satisfying the same conditions as
previously. A representation of the punctured diagram \unitlength 1.00mm
\linethickness{0.4pt}
\begin{picture}(8.00,9.00)
\put(4.00,4.00){\circle{8.00}} \put(7.00,1.00){\vector(-1,1){6.00}}
\put(1.00,1.00){\vector(1,1){6.00}} \put(4.00,8.00){\circle*{1.00}}
\end{picture}
is such a collection of points in the parametrized
circle, whose cyclic coordinates satisfy a more strong condition
$0<a<b<c<d<2\pi;$ the origin $0=2\pi \in S^1$ corresponds to the marked point
in the diagram. It is easy to see that the number of representations of the
last punctured diagram (counted with natural signs) is a knot invariant, and
the similar number for the absolute diagram is not.
However, some of finite type knot invariants can be realized by absolute
diagrams: in particular the unique invariant of order 3 can be given by the
diagram \unitlength 1mm \linethickness{0.4pt}
\begin{picture}(29.00,8.00)
\put(9.00,4.00){\circle{8.00}} \put(25.00,4.00){\circle{8.00}}
\put(25.00,0.00){\vector(0,1){8.00}} \put(28.30,6.20){\vector(-3,-2){6.60}}
\put(21.70,6.20){\vector(3,-2){6.60}} \put(9.00,0.00){\vector(0,1){8.00}}
\put(5.60,6.00){\vector(1,0){6.80}} \put(12.40,2.00){\vector(-1,0){6.80}}
\put(2.00,4.00){\makebox(0,0)[cc]{$\frac{1}{2}$}}
\put(17.00,4.00){\makebox(0,0)[cc]{$+\frac{1}{3}$}}
\end{picture}
, see \cite{PV}.
M.~Goussarov has proved that any finite-type knot invariant of long knots can
be represented by a formula of Polyak--Viro type, see \cite{GPV}.
Similar (but more complicated) formulas appear naturally in the calculation of
higher-dimensional cohomology classes of spaces of knots, see the next
sections.
\begin{remark}{\rm
The homological calculations discussed below provide numerous possibilities to
make a mistake: to miss some component of the boundary, to calculate wrongly
some orientation, etc. Fortunately, we always can check our calculations. If we
have calculated some boundary operator and suspect that it is not correct, just
calculate the boundary of this boundary, and try to understand why it is not
equal to zero! My experience says that no mistakes survive this examination.}
\end{remark}
\section{Proof of Theorem 1}
\label{proof1}
\subsection{Principal part of the cocycle}
\label{f3}
In the original calculation \cite{tetu}, the principal part of the
Teiblum-Turchin cocycle in the term $F_3 \setminus F_2$ of the natural
filtration of the resolved discriminant was found as a linear combination of
some 8 cells of the canonical cell decomposition, see e.g. \cite{bjo},
\cite{fasis}.
This expression can be simplified, especially if $n$ is even.
\begin{proposition}
For any $n\ge 3$, the group of order 3 cohomology classes of dimension $3n-8$
of the space of long knots $\R^1 \to \R^n$ is cyclic; for $n\ge 3$ it is free
Abelian.
If $n$ is even, then this group is generated by the sum of only two cells:
\begin{equation}
\label{prinpart} \special{em:linewidth 0.4pt} \unitlength 1.00mm
\linethickness{0.4pt}
\begin{picture}(68.00,7.00)
\put(68.00,3.00){\line(-1,0){23.00}} \put(54.00,3.00){\oval(10.00,4.00)[t]}
\put(61.50,3.00){\oval(5.00,4.00)[t]} \put(56.50,3.00){\oval(15.00,8.00)[t]}
\put(59.00,3.00){\oval(10.00,6.00)[b]} \put(41.00,3.00){\makebox(0,0)[cc]{$+$}}
\put(19.50,3.00){\oval(15.00,6.00)[b]} \put(27.00,3.00){\oval(10.00,4.00)[t]}
\put(24.50,3.00){\oval(15.00,8.00)[t]} \put(36.00,3.00){\line(-1,0){28.00}}
\put(2.00,3.00){\makebox(0,0)[cc]{$TT =$}}
\end{picture} .
\end{equation}
For odd $n$ it is generated by the linear combination
\begin{equation}
\label{prinpar} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(125.00,7.00)
\put(18.00,3.00){\line(-1,0){18.00}} \put(6.00,3.00){\oval(10.00,6.00)[b]}
\put(8.50,3.00){\oval(5.00,4.00)[t]} \put(13.50,3.00){\oval(5.00,4.00)[t]}
\put(11.00,3.00){\oval(10.00,8.00)[t]} \put(21.00,3.00){\makebox(0,0)[cc]{$+$}}
\put(32.50,3.00){\oval(5.00,4.00)[t]} \put(37.50,3.00){\oval(5.00,4.00)[t]}
\put(35.00,3.00){\oval(10.00,8.00)[t]} \put(32.50,3.00){\oval(15.00,6.00)[b]}
\put(42.00,3.00){\line(-1,0){19.00}} \put(44.00,3.00){\makebox(0,0)[cc]{$+$}}
\put(70.00,3.00){\line(-1,0){24.00}} \put(55.50,3.00){\oval(15.00,6.00)[b]}
\put(63.00,3.00){\oval(10.00,4.00)[t]} \put(60.50,3.00){\oval(15.00,8.00)[t]}
\put(73.00,3.00){\makebox(0,0)[cc]{$+$}} \put(97.00,3.00){\line(-1,0){22.00}}
\put(86.00,3.00){\oval(10.00,6.00)[b]} \put(90.50,3.00){\oval(9.00,4.00)[t]}
\put(86.00,3.00){\oval(18.00,8.00)[t]}
\put(101.00,3.00){\makebox(0,0)[cc]{$-$}} \put(125.00,3.00){\line(-1,0){21.00}}
\put(112.50,3.00){\oval(15.00,8.00)[t]} \put(115.00,3.00){\oval(10.00,4.00)[t]}
\put(119.50,3.00){\oval(9.00,6.00)[b]}
\end{picture} .
\end{equation}
\end{proposition}
The first statement of this proposition for odd $n$ was essentially proved by
Teiblum and Turchin \cite{tetu}; the justification of entire statement see in
\S 6 of \cite{bjo} or \S V.8.8 of \cite{fasis}.
Since we consider our class mod 2, the stabilization formula (\ref{stabil})
allows us to use any of expressions (\ref{prinpart}), (\ref{prinpar}) in the
case of any $n$. We shall use the shorter "even" version (\ref{prinpart}).
All further calculations in this section are mod 2 only.
\subsection{On the pictures}
The system of notation in this work is an extension of that used in
\S~\ref{method} for the cells of the natural simplicial resolution of the
discriminant. Any of our pictures consists of a horizontal segment (the {\em
Wilson loop} symbolizing the line $\R^1$), several asterisks placed on it, and
several arcs ("chords") connecting some its points (these data determine such a
cell), plus some additional furniture consisting of broken lines ({\em
zigzags}) and subscripts, which distinguish certain subvarieties in these
cells.
For instance, the picture \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(21.00,5.00)
\put(8.50,2.00){\oval(13.00,4.00)[b]} \put(12.50,2.00){\oval(13.00,6.00)[t]}
\put(0.00,2.00){\line(1,0){21.00}} \put(19.00,2.00){\line(-2,1){4.00}}
\put(15.00,4.00){\line(-2,-1){4.00}}
\end{picture}
\
means, first of all, that we are in the cell \unitlength 1.00mm
\special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(16.00,5.00)
\put(0.00,2.00){\line(1,0){16.00}} \put(6.50,2.00){\oval(9.00,4.00)[b]}
\put(9.50,2.00){\oval(9.00,4.00)[t]}
\end{picture}
\ of the term $F_2 \setminus F_1$.
This cell can be considered as the space of all triples $(\alpha,t,f)$ where
$\alpha$ is some quadruple of points $a<b<c<d$ in $\R^1$, $f$ is a smooth map
$\R^1 \to \R^n$ such that $f(a)=f(c),$ $f(b)=f(d)$, and $t$ is a point of the
segment $\tilde \Delta(J),$ $J=((a,c);(b,d))$, participating in the
construction of the resolution: its endpoints correspond formally to the pairs
of points $(a,c)$ and $(b,d)$ glued together by $f$. The additional zigzag in
the picture \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(21.00,5.00)
\put(8.50,2.00){\oval(13.00,4.00)[b]} \put(12.50,2.00){\oval(13.00,6.00)[t]}
\put(0.00,2.00){\line(1,0){21.00}} \put(19.00,2.00){\line(-2,1){4.00}}
\put(15.00,4.00){\line(-2,-1){4.00}}
\end{picture}
\
distinguishes the subvariety in this cell, consisting of such triples
$(\alpha,t,f)$ that there exists one point $\lambda \in \R^1$ more, $b <
\lambda < c$, such that $f(\lambda)=f(d)$. By definition, this subvariety is
identical with the one encoded by the picture \unitlength 1.00mm
\special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(20.00,5.00)
\put(8.50,2.00){\oval(13.00,4.00)[b]} \put(11.50,2.00){\oval(13.00,6.00)[t]}
\put(0.00,2.00){\line(1,0){20.00}} \put(5.00,2.00){\line(2,1){4.00}}
\put(9.00,4.00){\line(2,-1){4.00}}
\end{picture} .
The picture \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(21.00,5.00)
\put(8.50,2.00){\oval(13.00,4.00)[b]} \put(12.50,2.00){\oval(13.00,6.00)[t]}
\put(0.00,2.00){\line(1,0){21.00}} \put(19.00,2.00){\line(-2,1){4.00}}
\put(15.00,4.00){\vector(-2,-1){4.00}}
\end{picture}
(respectively, \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(21.00,5.00)
\put(8.50,2.00){\oval(13.00,4.00)[b]} \put(12.50,2.00){\oval(13.00,6.00)[t]}
\put(0.00,2.00){\line(1,0){21.00}} \put(15.00,4.00){\vector(2,-1){4.00}}
\put(15.00,4.00){\line(-2,-1){4.00}}
\end{picture} )
will denote almost the same, but with the condition $f(\lambda)=f(d)$ replaced
by the condition that $f(\lambda)$ has the same projection to $\R^{n-1}$ as
$f(d)=f(b)$ and lies {\em below} (respectively, {\em above}) $f(d)$ in the line
of all points with the same projection.
The subscript of type \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(8.00,5.00)
\put(4.00,0.00){\vector(1,2){2.50}} \put(4.00,0.00){\vector(-1,2){2.50}}
\put(0.33,2.00){\makebox(0,0)[cc]{\small 1}}
\put(7.67,2.00){\makebox(0,0)[cc]{\small 2}}
\end{picture}
\
under a picture denotes the condition that the "vertical" direction in $\R^n$
lies in the angle between the tangent directions $f'(a_1)$ and $f'(a_2)$, where
$a_1$ and $a_2$ are the first and the second from the left points of $\R^1$
participating actively in the picture. Similarly, the subscript \unitlength
1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(5.00,6.00)
\put(0.00,3.00){\vector(3,1){5.00}} \put(0.00,3.00){\vector(3,-1){5.00}}
\put(2.50,0.00){\makebox(0,0)[cc]{\small 1}}
\put(2.50,6.00){\makebox(0,0)[cc]{\small 2}}
\end{picture}
(respectively, \special{em:linewidth 0.4pt} \unitlength 1.00mm
\linethickness{0.4pt}
\begin{picture}(8.00,5.00)
\put(4.00,5.00){\vector(1,-2){2.50}} \put(4.00,5.00){\vector(-1,-2){2.50}}
\put(7.67,2.00){\makebox(0,0)[cc]{\small 2}}
\put(0.33,2.00){\makebox(0,0)[cc]{\small 1}}
\end{picture}
)
says that the vertical direction lies in the angle between the vectors
$-f'(a_1)$ and $f'(a_2)$ (respectively, between the vectors $-f'(a_1)$ and
$-f'(a_2)$). The subscript \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(12.00,1.00)
\put(6.00,-1.00){\vector(1,0){6.00}} \put(6.00,-1.00){\vector(-1,0){6.00}}
\put(6.00,-1.00){\circle*{1.00}} \put(1.00,1.00){\makebox(0,0)[cc]{\small 1}}
\put(11.00,1.00){\makebox(0,0)[cc]{\small 2}}
\end{picture}
means that the tangent directions $f'(a_1)$ and $f'(a_2)$ are opposite in
$\R^n$. The notation of all these types appears only if the condition
$f(a_1)=f(a_2)$ is satisfied (and can be seen from the chords and zigzags on
the picture).
The subscript \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(6.00,5.00)
\put(0.00,1.00){\vector(2,-1){4.00}} \put(0.00,1.00){\vector(2,1){4.00}}
\put(0.00,1.00){\vector(1,0){6.00}} \put(2.00,-2.00){\makebox(0,0)[cc]{$2$}}
\put(2.00,4.00){\makebox(0,0)[cc]{$1$}} \put(0.00,-0.50){\line(0,1){3.00}}
\end{picture}
means that the distinguished direction "to the right" in ${\bf R}^{n-1}$ lies
between the projections of such tangents $f'(a_1),$ $f'(a_2)$ to $\R^{n-1}$
(i.e. this direction is the linear combination of these projections with
nonnegative coefficients). The subscript \unitlength 1.00mm
\special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(11.50,3.00)
\put(6.00,1.00){\makebox(0,0)[cc]{$1 \longleftrightarrow 2$}}
\put(6.00,-0.50){\line(0,1){3.00}}
\end{picture}
\/ means that the projections of tangents $f'(a_1), f'(a_2)$ to $\R^{n-1}$ have
opposite directions. The notation of last two types can appear only if the
projections of corresponding points $f(a_1), f(a_2)$ coincide in $\R^{n-1}$.
The sum of varieties distinguished by conditions of types \unitlength 1.00mm
\special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(8.00,5.00)
\put(4.00,0.00){\vector(1,2){2.50}} \put(4.00,0.00){\vector(-1,2){2.50}}
\put(0.33,2.00){\makebox(0,0)[cc]{\small 1}}
\put(7.67,2.00){\makebox(0,0)[cc]{\small 2}}
\end{picture}
\
and \special{em:linewidth 0.4pt} \unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(8.00,5.00)
\put(4.00,5.00){\vector(1,-2){2.50}} \put(4.00,5.00){\vector(-1,-2){2.50}}
\put(7.67,2.00){\makebox(0,0)[cc]{\small 2}}
\put(0.33,2.00){\makebox(0,0)[cc]{\small 1}}
\end{picture}
\
in one and the same cell is equal to the variety of type \unitlength 1.00mm
\special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(11.50,3.00)
\put(6.00,1.00){\makebox(0,0)[cc]{$1 \longleftrightarrow 2$}}
\put(6.00,-0.50){\line(0,1){3.00}}
\end{picture}
\ ;
"of type" here means that some other two numbers instead of 1 and 2 can stay in
all three pictures. This identity is not symmetric: indeed, the variety of type
\unitlength 1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(11.50,3.00)
\put(6.00,1.00){\makebox(0,0)[cc]{$1 \longleftrightarrow 2$}}
\put(6.00,-0.50){\line(0,1){3.00}}
\end{picture}
\
can be well defined even when the former two varieties have no sense.
The subscript $2 \updownarrow$ (respectively, $2 \uparrow$, respectively, $2
\downarrow$) means that the tangent vector $f'(a_2)$ is vertical (respectively,
vertical directed up, respectively, vertical directed down). The subscript of
type $2 \mapsto$ means that the projection of the tangent $f'(a_2)$ to
$\R^{n-1}$ is directed "to the right".
Abbreviation $f_1$ replaces the composition ${\bf p} \circ f: \R^1 \to
\R^{n-1}$. Finally, a collection of vectors in a framebox means that these
vectors are linearly dependent. For instance the subscript \unitlength 1.00mm
\special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(28.00,5.00)
\put(0.00,-1.00){\framebox(28.00,5.00)[cc]{$f_1'(1),f_1''(2),\mapsto$}}
\end{picture}
means that some three vectors in $\R^{n-1}$, namely the projection of
$f'(a_1)$, the projection of $f''(a_2)$, and the direction "to the right", span
a subspace of dimension $\le 2$. Several more specific abbreviations will be
explained later, close to their first use.
The boundary of the variety distinguished in any cell by the condition
\unitlength 1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(8.00,5.00)
\put(4.00,0.00){\vector(1,2){2.50}} \put(4.00,0.00){\vector(-1,2){2.50}}
\put(0.33,2.00){\makebox(0,0)[cc]{\small 1}}
\put(7.67,2.00){\makebox(0,0)[cc]{\small 2}}
\end{picture}
\ (respectively,
\special{em:linewidth 0.4pt} \unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(8.00,5.00)
\put(4.00,5.00){\vector(1,-2){2.50}} \put(4.00,5.00){\vector(-1,-2){2.50}}
\put(7.67,2.00){\makebox(0,0)[cc]{\small 2}}
\put(0.33,2.00){\makebox(0,0)[cc]{\small 1}}
\end{picture}
\ )
is equal to the sum of varieties distinguished by conditions \unitlength 1.00mm
\special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(12.00,1.00)
\put(6.00,-1.00){\vector(1,0){6.00}} \put(6.00,-1.00){\vector(-1,0){6.00}}
\put(6.00,-1.00){\circle*{1.00}} \put(1.00,1.00){\makebox(0,0)[cc]{\small 1}}
\put(11.00,1.00){\makebox(0,0)[cc]{\small 2}}
\end{picture}
\ ,
$1 \uparrow$ and $2 \uparrow$ (respectively, \unitlength 1.00mm
\special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(12.00,1.00)
\put(6.00,-1.00){\vector(1,0){6.00}} \put(6.00,-1.00){\vector(-1,0){6.00}}
\put(6.00,-1.00){\circle*{1.00}} \put(1.00,1.00){\makebox(0,0)[cc]{\small 1}}
\put(11.00,1.00){\makebox(0,0)[cc]{\small 2}}
\end{picture}
\ ,
$1 \downarrow$ and $2 \downarrow$) plus maybe something in the boundary of the
cell. Similarly, the boundary of the variety distinguished by the condition
\unitlength 1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(6.00,5.00)
\put(0.00,1.00){\vector(2,-1){4.00}} \put(0.00,1.00){\vector(2,1){4.00}}
\put(0.00,1.00){\vector(1,0){6.00}} \put(2.00,-2.00){\makebox(0,0)[cc]{$2$}}
\put(2.00,4.00){\makebox(0,0)[cc]{$1$}} \put(0.00,-0.50){\line(0,1){3.00}}
\end{picture}
is equal (modulo the boundary of the cell) to the sum of varieties
distinguished in the same cell by conditions \unitlength 1.00mm
\special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(11.50,3.00)
\put(6.00,1.00){\makebox(0,0)[cc]{$1 \longleftrightarrow 2$}}
\put(6.00,-0.50){\line(0,1){3.00}}
\end{picture}
\ ,
$1 \mapsto$, and $2 \mapsto$. The boundary of the condition $2 \mapsto$ is
equal to $2 \updownarrow$ plus something in smaller cells.
\subsection{The first differential}
\label{f2}
Formula (\ref{prinpart}) defines a relative cycle in the term $F_3$ of our
filtration modulo $F_2.$ In this subsection we calculate its boundary in the
term $F_2 \setminus F_1,$ and span it by some chain with closed supports in
this term (i.e. we represent it as the boundary of such a chain).
\begin{proposition}
The boundary of the cycle $($\ref{prinpart}$)$ in $F_2\setminus F_1$ is equal
to the chain
\begin{equation}
\label{difone} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(120.00,16.00)
\put(0.00,12.00){\line(1,0){20.00}} \put(25.00,12.00){\line(1,0){20.00}}
\put(50.00,12.00){\line(1,0){20.00}} \put(75.00,12.00){\line(1,0){20.00}}
\put(100.00,12.00){\line(1,0){20.00}} \put(110.00,12.00){\oval(16.00,8.00)[t]}
\put(106.00,12.00){\oval(8.00,4.00)[t]} \put(114.00,12.00){\oval(8.00,4.00)[t]}
\put(118.00,12.00){\line(-2,-1){6.00}} \put(112.00,9.00){\line(-2,1){6.00}}
\put(85.00,12.00){\oval(16.00,8.00)[t]} \put(81.00,12.00){\oval(8.00,4.00)[t]}
\put(89.00,12.00){\oval(8.00,4.00)[t]} \put(84.00,8.00){\vector(1,0){6.00}}
\put(84.00,8.00){\vector(-1,0){6.00}} \put(80.00,10.00){\makebox(0,0)[cc]{$1$}}
\put(88.00,10.00){\makebox(0,0)[cc]{$2$}} \put(84.00,8.00){\circle*{1.00}}
\put(58.50,12.00){\oval(13.00,8.00)[b]} \put(64.00,12.00){\oval(8.00,8.00)[t]}
\put(33.00,12.00){\oval(12.00,8.00)[b]} \put(36.50,12.00){\oval(13.00,8.00)[t]}
\put(43.00,12.00){\line(-2,1){6.00}} \put(37.00,15.00){\line(-4,-3){4.00}}
\put(11.50,12.00){\oval(13.00,8.00)[t]} \put(13.50,12.00){\oval(9.00,4.00)[t]}
\put(9.00,8.33){\line(-2,1){7.00}} \put(10.00,2.00){\makebox(0,0)[cc]{$A$}}
\put(35.00,2.00){\makebox(0,0)[cc]{$B$}}
\put(60.00,2.00){\makebox(0,0)[cc]{$C$}}
\put(85.00,2.00){\makebox(0,0)[cc]{$D$}}
\put(110.00,2.00){\makebox(0,0)[cc]{$E$}}
\put(22.50,12.00){\makebox(0,0)[cc]{$+$}}
\put(47.50,12.00){\makebox(0,0)[cc]{$+$}}
\put(72.50,12.00){\makebox(0,0)[cc]{$+$}}
\put(97.50,12.00){\makebox(0,0)[cc]{$+$}} \put(9.00,8.00){\line(3,2){6.00}}
\put(60.00,12.00){\line(-1,2){2.00}} \put(58.00,16.00){\line(-1,-2){2.00}}
\end{picture}
\ .
\end{equation}
$($Namely, the boundary of the first term of $($\ref{prinpart}$)$ consists of
four chains A, B, C and D in $($\ref{difone}$)$, and the boundary of the second
is equal to the fifth chain E.$)$
\end{proposition}
The unique nontrivial term of this formula is the 4th one: it appears when the
first point of the 5-configuration participating in the first term of
(\ref{prinpart}) tends to the second, and simultaneously the fourth point tends
to the third. \quad $\square$
\medskip
{\bf Exercise:} to check that the chain (\ref{difone}) actually is a cycle in
$F_2 \setminus F_1$.
\medskip
Now, let us span this cycle by a chain in the term $F_2 \sm F_1$. The cellular
structure of this term was described in Example \ref{exf2} of section
\ref{method}.
First we span the components D and E of (\ref{difone}) inside the cell
\unitlength 1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(18.00,4.00)
\put(0.00,0.00){\line(1,0){16.00}} \put(8.00,0.00){\oval(14.00,8.00)[t]}
\put(4.50,0.00){\oval(7.00,4.00)[t]} \put(11.50,0.00){\oval(7.00,4.00)[t]}
\end{picture}
, i.e. we construct the homology between their sum and some chain in the
boundary of this cell. It is natural to span a chain with condition of type
\unitlength 1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(12.00,1.00)
\put(6.00,-1.00){\vector(1,0){6.00}} \put(6.00,-1.00){\vector(-1,0){6.00}}
\put(6.00,-1.00){\circle*{1.00}} \put(1.00,1.00){\makebox(0,0)[cc]{\small 1}}
\put(11.00,1.00){\makebox(0,0)[cc]{\small 2}}
\end{picture}
\ by a similar chain with condition of type
\unitlength 1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(8.00,5.00)
\put(4.00,0.00){\vector(1,2){2.50}} \put(4.00,0.00){\vector(-1,2){2.50}}
\put(0.33,2.00){\makebox(0,0)[cc]{\small 1}}
\put(7.67,2.00){\makebox(0,0)[cc]{\small 2}}
\end{picture}
,
and a chain having zigzag without arrows by a similar chain with an arrow added
at one of endpoints of the zigzag. The chains obtained in this way from the
ones encoded by parts D and E of (\ref{difone}) are indicated in the left parts
of the next two equations (\ref{boundd}) and (\ref{bounde}) respectively.
In the right-hand parts of these formulas, as well as in all forthcoming
expressions for boundary operators in this work, we first count the components
of the boundary defined by the degenerations of the subvarieties in the
corresponding cells, distinguished by arrowed zigzags and subscripts. Then we
count the components defined by the limit positions of these varieties when the
cell itself degenerates because of the collision of some points forming its
underlying $J$-configuration in $\R^1$. The latter degenerations appear in the
lexicographic order: first by the number of colliding pairs of points in
$\R^1$, and then by their positions in $\R^1$.
\begin{equation}
\label{boundd} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(114.00,28.00)
\put(3.00,24.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,24.00){\line(1,0){20.00}} \put(16.00,24.00){\oval(16.00,8.00)[t]}
\put(12.00,24.00){\oval(8.00,4.00)[t]} \put(20.00,24.00){\oval(8.00,4.00)[t]}
\put(14.00,17.00){\vector(1,1){5.00}} \put(14.00,17.00){\vector(-1,1){5.00}}
\put(20.00,19.00){\makebox(0,0)[cc]{$1$}}
\put(8.00,19.00){\makebox(0,0)[cc]{$2$}}
\put(30.00,24.00){\makebox(0,0)[cc]{=}} \put(36.00,24.00){\line(1,0){20.00}}
\put(46.00,24.00){\oval(16.00,8.00)[t]} \put(42.00,24.00){\oval(8.00,4.00)[t]}
\put(50.00,24.00){\oval(8.00,4.00)[t]} \put(63.00,24.00){\line(1,0){20.00}}
\put(73.00,24.00){\oval(16.00,8.00)[t]} \put(69.00,24.00){\oval(8.00,4.00)[t]}
\put(77.00,24.00){\oval(8.00,4.00)[t]} \put(90.00,24.00){\line(1,0){20.00}}
\put(100.00,24.00){\oval(16.00,8.00)[t]} \put(96.00,24.00){\oval(8.00,4.00)[t]}
\put(104.00,24.00){\oval(8.00,4.00)[t]} \put(87.00,24.00){\makebox(0,0)[cc]{+}}
\put(60.00,24.00){\makebox(0,0)[cc]{+}} \put(41.00,20.00){\makebox(0,0)[cc]{$1
\uparrow$}} \put(70.00,20.00){\makebox(0,0)[cc]{$2 \uparrow$}}
\put(98.00,20.00){\vector(1,0){6.00}} \put(98.00,20.00){\vector(-1,0){6.00}}
\put(98.00,20.00){\circle*{1.00}} \put(102.00,22.00){\makebox(0,0)[cc]{$1$}}
\put(94.00,22.00){\makebox(0,0)[cc]{$2$}}
\put(114.00,24.00){\makebox(0,0)[cc]{$+$}} \put(36.00,7.00){\line(1,0){20.00}}
\put(46.00,7.00){\oval(16.00,8.00)[t]} \put(50.00,7.00){\oval(8.00,4.00)[t]}
\put(44.00,0.00){\vector(1,1){5.00}} \put(44.00,0.00){\vector(-1,1){5.00}}
\put(50.00,2.00){\makebox(0,0)[cc]{$1$}}
\put(38.00,2.00){\makebox(0,0)[cc]{$2$}} \put(63.00,7.00){\line(1,0){20.00}}
\put(69.00,7.00){\oval(8.00,4.00)[t]} \put(77.00,7.00){\oval(8.00,4.00)[t]}
\put(71.00,0.00){\vector(1,1){5.00}} \put(71.00,0.00){\vector(-1,1){5.00}}
\put(77.00,2.00){\makebox(0,0)[cc]{$1$}}
\put(65.00,2.00){\makebox(0,0)[cc]{$2$}} \put(90.00,7.00){\line(1,0){20.00}}
\put(100.00,7.00){\oval(16.00,8.00)[t]} \put(96.00,7.00){\oval(8.00,4.00)[t]}
\put(98.00,0.00){\vector(1,1){5.00}} \put(98.00,0.00){\vector(-1,1){5.00}}
\put(104.00,2.00){\makebox(0,0)[cc]{$1$}}
\put(92.00,2.00){\makebox(0,0)[cc]{$2$}}
\put(87.00,7.00){\makebox(0,0)[cc]{$+$}}
\put(60.00,7.00){\makebox(0,0)[cc]{$+$}}
\put(32.00,7.00){\makebox(0,0)[cc]{$+$}}
\end{picture}
,
\end{equation}
\begin{equation}
\label{bounde} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(108.00,23.00)
\put(3.00,19.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,19.00){\line(1,0){20.00}} \put(16.00,19.00){\oval(16.00,8.00)[t]}
\put(12.00,19.00){\oval(8.00,4.00)[t]} \put(20.00,19.00){\oval(8.00,4.00)[t]}
\put(24.00,19.00){\line(-2,-1){6.00}} \put(18.00,16.00){\vector(-2,1){6.00}}
\put(32.00,19.00){\line(1,0){20.00}} \put(42.00,19.00){\oval(16.00,8.00)[t]}
\put(38.00,19.00){\oval(8.00,4.00)[t]} \put(46.00,19.00){\oval(8.00,4.00)[t]}
\put(58.00,19.00){\line(1,0){20.00}} \put(68.00,19.00){\oval(16.00,8.00)[t]}
\put(64.00,19.00){\oval(8.00,4.00)[t]} \put(72.00,19.00){\oval(8.00,4.00)[t]}
\put(84.00,19.00){\line(1,0){20.00}} \put(94.00,19.00){\oval(16.00,8.00)[t]}
\put(90.00,19.00){\oval(8.00,4.00)[t]} \put(98.00,19.00){\oval(8.00,4.00)[t]}
\put(81.00,19.00){\makebox(0,0)[cc]{$+$}}
\put(55.00,19.00){\makebox(0,0)[cc]{$+$}}
\put(29.00,19.00){\makebox(0,0)[cc]{$=$}}
\put(36.00,14.00){\makebox(0,0)[cc]{$1 \downarrow$}}
\put(66.00,14.00){\makebox(0,0)[cc]{$2 \uparrow$}}
\put(102.00,19.00){\line(-2,-1){6.00}} \put(96.00,16.00){\line(-2,1){6.00}}
\put(108.00,19.00){\makebox(0,0)[cc]{$+$}} \put(32.00,3.00){\line(1,0){20.00}}
\put(42.00,3.00){\oval(16.00,8.00)[t]} \put(46.00,3.00){\oval(8.00,4.00)[t]}
\put(50.00,3.00){\line(-2,-1){6.00}} \put(44.00,0.00){\vector(-2,1){6.00}}
\put(58.00,3.00){\line(1,0){20.00}} \put(64.00,3.00){\oval(8.00,4.00)[t]}
\put(72.00,3.00){\oval(8.00,4.00)[t]} \put(76.00,3.00){\line(-2,-1){6.00}}
\put(70.00,0.00){\vector(-2,1){6.00}} \put(84.00,3.00){\line(1,0){20.00}}
\put(94.00,3.00){\oval(16.00,8.00)[t]} \put(90.00,3.00){\oval(8.00,4.00)[t]}
\put(102.00,3.00){\line(-2,-1){6.00}} \put(96.00,0.00){\vector(-2,1){6.00}}
\put(81.00,3.00){\makebox(0,0)[cc]{$+$}}
\put(55.00,3.00){\makebox(0,0)[cc]{$+$}}
\put(29.00,3.00){\makebox(0,0)[cc]{$+$}}
\end{picture}
.
\end{equation}
\begin{proposition}
The equalities $($\ref{boundd}$)$, $($\ref{bounde}$)$ are correct, i.e. the
algebraic boundaries $($mod 2$)$ in $F_2 \setminus F_1$ of the varieties
indicated in their left parts are equal to the sums of varieties indicated in
their right-hand parts. \quad $\square$
\end{proposition}
In (\ref{bounde}) first two summands are degenerations of the variety defined
by the zigzag when its arrowed endpoint tends to one of boundaries of the
corresponding segment, and the third summand belongs to its boundary as the
equality of type $\phi(x)=\phi(y)$ defines a component of the boundary of the
set defined by the inequality $\phi(x) \ge \phi(y)$.
The last three summands in both (\ref{boundd}) and (\ref{bounde}) belong to the
boundary (\ref{ztripno}) of the cell \unitlength 1.00mm \special{em:linewidth
0.4pt} \linethickness{0.4pt}
\begin{picture}(18.00,4.00)
\put(0.00,0.00){\line(1,0){16.00}} \put(8.00,0.00){\oval(14.00,8.00)[t]}
\put(4.50,0.00){\oval(7.00,4.00)[t]} \put(11.50,0.00){\oval(7.00,4.00)[t]}
\end{picture}
.
The sum of all varieties indicated in right-hand parts of (\ref{boundd}),
(\ref{bounde}) consists of part D + E of (\ref{difone}), some chain in the
boundary of the cell \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(18.00,4.00)
\put(0.00,0.00){\line(1,0){16.00}} \put(8.00,0.00){\oval(14.00,8.00)[t]}
\put(4.50,0.00){\oval(7.00,4.00)[t]} \put(11.50,0.00){\oval(7.00,4.00)[t]}
\end{picture}
,
and the first chain in the right-hand part of the equation
\begin{equation}
\label{nev} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(130.00,8.00)
\put(6.00,4.00){\line(1,0){20.00}} \put(16.00,4.00){\oval(16.00,8.00)[t]}
\put(12.00,4.00){\oval(8.00,4.00)[t]} \put(20.00,4.00){\oval(8.00,4.00)[t]}
\put(32.00,4.00){\line(1,0){20.00}} \put(42.00,4.00){\oval(16.00,8.00)[t]}
\put(38.00,4.00){\oval(8.00,4.00)[t]} \put(46.00,4.00){\oval(8.00,4.00)[t]}
\put(37.00,1.00){\makebox(0,0)[cc]{$1 \updownarrow$}}
\put(29.00,4.00){\makebox(0,0)[cc]{$=$}}
\put(3.00,4.00){\makebox(0,0)[cc]{$\partial$}}
\put(58.00,4.00){\line(1,0){20.00}} \put(68.00,4.00){\oval(16.00,8.00)[t]}
\put(72.00,4.00){\oval(8.00,4.00)[t]} \put(84.00,4.00){\line(1,0){20.00}}
\put(90.00,4.00){\oval(8.00,4.00)[t]} \put(98.00,4.00){\oval(8.00,4.00)[t]}
\put(110.00,4.00){\line(1,0){20.00}} \put(120.00,4.00){\oval(16.00,8.00)[t]}
\put(116.00,4.00){\oval(8.00,4.00)[t]} \put(81.00,4.00){\makebox(0,0)[cc]{$+$}}
\put(107.00,4.00){\makebox(0,0)[cc]{$+$}}
\put(55.00,4.00){\makebox(0,0)[cc]{$+$}} \put(116.00,1.00){\makebox(0,0)[cc]{$1
\mapsto$}} \put(90.00,1.00){\makebox(0,0)[cc]{$1 \mapsto$}}
\put(64.00,1.00){\makebox(0,0)[cc]{$1 \mapsto$}}
\put(12.00,1.00){\makebox(0,0)[cc]{$1 \mapsto$}}
\end{picture}
\ .
\end{equation}
In other words, the sum of three chains in the left parts of equations
(\ref{boundd}), (\ref{bounde}) and (\ref{nev}) realizes homology between the
sum of chains $D$ and $E$ and some chain in the boundary of the cell
\unitlength 1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(18.00,4.00)
\put(0.00,0.00){\line(1,0){16.00}} \put(8.00,0.00){\oval(14.00,8.00)[t]}
\put(4.50,0.00){\oval(7.00,4.00)[t]} \put(11.50,0.00){\oval(7.00,4.00)[t]}
\end{picture}
\/.
Now we span the summands B and C of (\ref{difone}) inside the open cell
\unitlength 1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(16.00,5.00)
\put(0.00,2.00){\line(1,0){16.00}} \put(6.50,2.00){\oval(9.00,4.00)[b]}
\put(9.50,2.00){\oval(9.00,4.00)[t]}
\end{picture}
. We need to find varieties in this cell,
whose boundaries include these summands. The obvious candidates for this are
the chains shown in the left parts of equations (\ref{boundb}) and
(\ref{boundc}) respectively.
\begin{equation}
\unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(121.00,29.00)
\put(3.00,25.00){\makebox(0,0)[cc]{$\partial$}}
\put(7.00,25.00){\line(1,0){23.00}} \put(17.00,25.00){\oval(16.00,8.00)[b]}
\put(20.00,25.00){\oval(16.00,8.00)[t]}
\put(33.00,25.00){\makebox(0,0)[cc]{$=$}} \put(36.00,25.00){\line(1,0){23.00}}
\put(46.00,25.00){\oval(16.00,8.00)[b]} \put(49.00,25.00){\oval(16.00,8.00)[t]}
\put(36.00,8.00){\line(1,0){23.00}} \put(46.00,8.00){\oval(16.00,8.00)[b]}
\put(49.00,8.00){\oval(16.00,8.00)[t]} \put(94.00,25.00){\line(1,0){23.00}}
\put(104.00,25.00){\oval(16.00,8.00)[b]}
\put(107.00,25.00){\oval(16.00,8.00)[t]}
\put(120.00,25.00){\makebox(0,0)[cc]{$+$}}
\put(91.00,25.00){\makebox(0,0)[cc]{$+$}}
\put(33.00,8.00){\makebox(0,0)[cc]{$+$}} \put(41.00,8.00){\line(3,-1){9.00}}
\put(50.00,5.00){\vector(4,3){4.00}} \put(114.00,19.00){\makebox(0,0)[cc]{$2
\downarrow$}} \put(62.00,25.00){\makebox(0,0)[cc]{$+$}}
\put(65.00,25.00){\line(1,0){23.00}} \put(76.50,25.00){\oval(19.00,8.00)[t]}
\put(74.00,25.00){\oval(14.00,8.00)[b]}
\put(92.00,8.00){\makebox(0,0)[cc]{$+$}} \put(95.00,8.00){\line(1,0){23.00}}
\put(106.50,8.00){\oval(19.00,8.00)[t]} \put(104.00,8.00){\oval(14.00,8.00)[b]}
\put(62.00,8.00){\makebox(0,0)[cc]{$+$}} \put(66.00,8.00){\line(1,0){23.00}}
\put(77.50,8.00){\oval(19.00,8.00)[b]} \put(80.00,8.00){\oval(14.00,8.00)[t]}
\put(116.00,0.00){\vector(-3,4){3.67}} \put(116.00,0.00){\vector(3,4){3.67}}
\put(121.00,3.00){\makebox(0,0)[cc]{$1$}}
\put(111.00,3.00){\makebox(0,0)[cc]{$2$}} \put(12.00,25.00){\line(4,-3){4.00}}
\put(16.00,22.00){\vector(4,3){4.00}} \put(41.00,25.00){\line(4,-3){4.00}}
\put(45.00,22.00){\line(4,3){4.00}} \put(81.00,25.00){\line(-4,3){4.00}}
\put(77.00,28.00){\vector(-4,-3){4.00}} \put(73.00,8.00){\line(4,-3){4.00}}
\put(77.00,5.00){\vector(4,3){4.00}}
\end{picture}
\label{boundb}
\end{equation}
\begin{equation}
\label{boundc} \unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(113.00,28.00)
\put(3.00,24.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,24.00){\line(1,0){21.00}} \put(14.00,24.00){\oval(14.00,8.00)[b]}
\put(20.50,24.00){\oval(11.00,8.00)[t]}
\put(30.00,24.00){\makebox(0,0)[cc]{$=$}} \put(33.00,24.00){\line(1,0){22.00}}
\put(41.50,24.00){\oval(15.00,8.00)[b]} \put(48.50,24.00){\oval(11.00,8.00)[t]}
\put(58.00,24.00){\makebox(0,0)[cc]{$+$}} \put(61.00,24.00){\line(1,0){21.00}}
\put(69.50,24.00){\oval(13.00,8.00)[b]} \put(73.00,24.00){\oval(14.00,8.00)[t]}
\put(66.00,24.00){\line(2,-1){6.00}} \put(72.00,21.00){\vector(4,3){4.00}}
\put(85.00,24.00){\makebox(0,0)[cc]{$+$}} \put(88.00,24.00){\line(1,0){21.00}}
\put(97.00,24.00){\oval(14.00,8.00)[b]}
\put(100.50,24.00){\oval(13.00,8.00)[t]}
\put(107.00,19.00){\makebox(0,0)[cc]{$2 \uparrow$}}
\put(113.00,24.00){\makebox(0,0)[cc]{$+$}} \put(7.00,9.00){\line(1,0){21.00}}
\put(13.50,9.00){\oval(9.00,8.00)[b]} \put(22.00,9.00){\oval(8.00,8.00)[t]}
\put(18.00,9.00){\line(-1,2){2.00}} \put(16.00,13.00){\vector(-1,-2){2.00}}
\put(58.00,9.00){\makebox(0,0)[cc]{$+$}} \put(61.00,9.00){\line(1,0){21.00}}
\put(67.50,9.00){\oval(9.00,8.00)[b]} \put(76.00,9.00){\oval(8.00,8.00)[t]}
\put(77.00,7.00){\vector(-2,-3){4.00}} \put(77.00,7.00){\vector(2,-3){4.00}}
\put(82.00,4.00){\makebox(0,0)[cc]{$1$}}
\put(72.00,4.00){\makebox(0,0)[cc]{$2$}}
\put(30.00,9.00){\makebox(0,0)[cc]{$+$}} \put(33.00,9.00){\line(1,0){21.00}}
\put(43.50,9.00){\oval(17.00,8.00)[b]} \put(47.50,9.00){\oval(9.00,8.00)[t]}
\put(43.00,9.00){\line(-1,2){2.00}} \put(41.00,13.00){\vector(-1,-2){2.00}}
\put(85.00,9.00){\makebox(0,0)[cc]{$+$}} \put(88.00,9.00){\line(1,0){21.00}}
\put(98.50,9.00){\oval(17.00,8.00)[b]} \put(102.50,9.00){\oval(9.00,8.00)[t]}
\put(15.00,24.00){\line(-3,4){3.00}} \put(12.00,28.00){\vector(-3,-4){3.00}}
\put(43.00,24.00){\line(-3,4){3.00}} \put(40.00,28.00){\line(-3,-4){3.00}}
\put(4.00,9.00){\makebox(0,0)[cc]{$+$}} \put(106.00,4.00){\vector(3,-1){6.00}}
\put(106.00,4.00){\vector(3,1){6.00}}
\put(109.00,7.00){\makebox(0,0)[cc]{{\small 3}}}
\put(109.00,1.00){\makebox(0,0)[cc]{{\small 1}}}
\end{picture}
\end{equation}
Again, all summands in lower rows of these equalities belong to the boundary of
the cell \unitlength 1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(16.00,5.00)
\put(0.00,2.00){\line(1,0){16.00}} \put(6.50,2.00){\oval(9.00,4.00)[b]}
\put(9.50,2.00){\oval(9.00,4.00)[t]}
\end{picture}
.
\begin{proposition}
\label{propbc} The equalities $($\ref{boundb}$)$, $($\ref{boundc}$)$ are
correct, i.e. the algebraic $($mod 2$)$ boundaries in $F_2 \setminus F_1$ of
the varieties indicated in their left parts are equal to the sums of varieties
indicated in their right-hand parts. \quad $\square$
\end{proposition}
In particular we get that the boundary of the sum of these two left-side
varieties is equal to the sum of varieties denoted in (\ref{difone}) by B and
C, plus some chain in the boundary of the cell \unitlength 1.00mm
\special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(16.00,5.00)
\put(0.00,2.00){\line(1,0){16.00}} \put(6.50,2.00){\oval(9.00,4.00)[b]}
\put(9.50,2.00){\oval(9.00,4.00)[t]}
\end{picture}
, plus the variety distinguished in this cell by the additional
condition $2 \updownarrow$. The last variety is a part of the boundary of the
similar set distinguished by the condition $2 \mapsto$. Entire boundary of this
set in $F_2 \setminus F_1$ is expressed by the formula
\begin{equation}
\label{nev2} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(127.00,11.00)
\put(16.00,0.00){\makebox(0,0)[cc]{$2 \mapsto$}}
\put(38.00,0.00){\makebox(0,0)[cc]{$2 \updownarrow$}}
\put(56.00,0.00){\makebox(0,0)[cc]{$1 \mapsto$}}
\put(74.00,0.00){\makebox(0,0)[cc]{$2 \mapsto$}}
\put(93.00,0.00){\makebox(0,0)[cc]{$2 \mapsto$}}
\put(1.00,7.00){\makebox(0,0)[cc]{$\partial$}}
\put(4.00,7.00){\line(1,0){14.00}} \put(9.50,7.00){\oval(9.00,8.00)[b]}
\put(12.50,7.00){\oval(9.00,8.00)[t]} \put(22.00,7.00){\makebox(0,0)[cc]{$=$}}
\put(25.00,7.00){\line(1,0){14.00}} \put(30.50,7.00){\oval(9.00,8.00)[b]}
\put(33.50,7.00){\oval(9.00,8.00)[t]} \put(42.00,7.00){\makebox(0,0)[cc]{$+$}}
\put(45.00,7.00){\line(1,0){15.00}} \put(50.50,7.00){\oval(9.00,8.00)[b]}
\put(52.50,7.00){\oval(13.00,8.00)[t]} \put(63.00,7.00){\makebox(0,0)[cc]{$+$}}
\put(66.00,7.00){\line(1,0){14.00}} \put(70.00,7.00){\oval(6.00,8.00)[b]}
\put(76.00,7.00){\oval(6.00,8.00)[t]} \put(83.00,7.00){\makebox(0,0)[cc]{$+$}}
\put(86.00,7.00){\line(1,0){14.00}} \put(93.00,7.00){\oval(12.00,8.00)[b]}
\put(95.00,7.00){\oval(8.00,8.00)[t]} \put(103.00,7.00){\makebox(0,0)[cc]{$+$}}
\put(107.00,7.00){\line(1,0){20.00}} \put(115.00,7.00){\oval(9.00,8.00)[t]}
\put(110.50,7.00){\makebox(0,0)[cc]{\large *}}
\put(118.00,4.00){\makebox(0,0)[cc]{\small $-f_{1}'''(1)/f_{1}''(1) \sim$}}
\put(118.00,0.00){\makebox(0,0)[cc]{\small $\sim \ \mapsto / f_{1}''(1)$}}
\end{picture}
\ .
\end{equation}
The subscript under the last term of (\ref{nev2}) means, that the projections
of second and third derivatives of $f$ at the point $a_1$ into $\R^{n-1}$ lie
in the same 2-plane as the direction "to the right", and two frames in this
2-plane obtained by adding to the projection of $f''(a_1)$ either the
projection of $f'''(a_1)$ or the direction "to the right" have {\em opposite}
orientations. This term occurs when both endpoints $a_1, a_3$ of the "lower"
arc in the left picture of (\ref{nev2}) tend from different sides to the first
endpoint $a_2$ of the "upper" arc.
Finally we get that the cycle $d^1(TT)$ shown in (\ref{difone}) is homologous
in $F_2 \setminus F_1$ to a chain lying in the union of cells of nonmaximal
dimensions listed in (\ref{ztripno}), (\ref{marg}); this homology is provided
by the sum of six varieties indicated in the left parts of equalities
(\ref{boundd}), (\ref{bounde}), (\ref{nev}), (\ref{boundb}), (\ref{boundc}),
and (\ref{nev2}).
Namely, this cycle homologous to $d^1(TT)$ is as follows. In the cell
\unitlength 1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(20.00,4.00)
\put(0.00,0.00){\line(1,0){20.00}} \put(10.00,0.00){\oval(16.00,8.00)[t]}
\put(6.00,0.00){\oval(8.00,4.00)[t]}
\end{picture}
it is zero, in the cell \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(20.00,4.00)
\put(0.00,0.00){\line(1,0){20.00}} \put(14.00,0.00){\oval(8.00,4.00)[t]}
\put(6.00,0.00){\oval(8.00,4.00)[t]}
\end{picture}
\ it is equal to the chain
\begin{equation}
\label{homztrip2} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(98.00,13.00)
\put(0.00,9.00){\line(1,0){20.00}} \put(6.00,9.00){\oval(8.00,8.00)[b]}
\put(14.00,9.00){\oval(8.00,8.00)[t]} \put(23.00,9.00){\makebox(0,0)[cc]{$+$}}
\put(26.00,9.00){\line(1,0){20.00}} \put(32.00,9.00){\oval(8.00,8.00)[b]}
\put(40.00,9.00){\oval(8.00,8.00)[t]} \put(49.00,9.00){\makebox(0,0)[cc]{$+$}}
\put(52.00,9.00){\line(1,0){20.00}} \put(58.00,9.00){\oval(8.00,8.00)[b]}
\put(66.00,9.00){\oval(8.00,8.00)[t]} \put(75.00,9.00){\makebox(0,0)[cc]{$+$}}
\put(78.00,9.00){\line(1,0){20.00}} \put(84.00,9.00){\oval(8.00,8.00)[b]}
\put(92.00,9.00){\oval(8.00,8.00)[t]} \put(91.00,1.00){\makebox(0,0)[cc]{$2
\mapsto$}} \put(65.00,6.00){\vector(-2,-3){4.00}}
\put(65.00,6.00){\vector(2,-3){4.00}} \put(70.00,3.00){\makebox(0,0)[cc]{1}}
\put(60.00,3.00){\makebox(0,0)[cc]{2}} \put(38.00,1.00){\makebox(0,0)[cc]{$1
\mapsto$}} \put(16.00,0.00){\vector(-2,3){4.00}}
\put(16.00,0.00){\vector(2,3){4.00}} \put(11.00,3.00){\makebox(0,0)[cc]{1}}
\put(21.00,3.00){\makebox(0,0)[cc]{2}}
\end{picture}
\ ,
\end{equation}
in the cell \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(20.00,4.00)
\put(0.00,0.00){\line(1,0){20.00}} \put(10.00,0.00){\oval(16.00,8.00)[t]}
\put(14.00,0.00){\oval(8.00,4.00)[t]}
\end{picture}
\ it is equal to the chain
\begin{equation}
\label{homztrip3} \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(101.00,28.00)
\put(0.00,24.00){\line(1,0){20.00}} \put(10.00,24.00){\oval(16.00,8.00)[t]}
\put(14.00,24.00){\oval(8.00,4.00)[t]}
\put(23.00,24.00){\makebox(0,0)[cc]{$+$}} \put(26.00,24.00){\line(1,0){20.00}}
\put(36.00,24.00){\oval(16.00,8.00)[t]} \put(40.00,24.00){\oval(8.00,4.00)[t]}
\put(49.00,24.00){\makebox(0,0)[cc]{$+$}} \put(52.00,24.00){\line(1,0){20.00}}
\put(62.00,24.00){\oval(16.00,8.00)[t]} \put(66.00,24.00){\oval(8.00,4.00)[t]}
\put(75.00,24.00){\makebox(0,0)[cc]{$+$}} \put(78.00,24.00){\line(1,0){20.00}}
\put(88.00,24.00){\oval(16.00,8.00)[t]} \put(92.00,24.00){\oval(8.00,4.00)[t]}
\put(101.00,24.00){\makebox(0,0)[cc]{$+$}} \put(52.00,4.00){\line(1,0){20.00}}
\put(62.00,4.00){\oval(16.00,8.00)[t]} \put(66.00,4.00){\oval(8.00,4.00)[t]}
\put(26.00,4.00){\line(1,0){20.00}} \put(37.00,4.00){\oval(14.00,8.00)[t]}
\put(40.00,4.00){\oval(8.00,4.00)[t]} \put(49.00,4.00){\makebox(0,0)[cc]{$+$}}
\put(10.00,15.00){\vector(-1,2){3.00}} \put(10.00,15.00){\vector(1,2){3.00}}
\put(15.00,18.00){\makebox(0,0)[cc]{2}} \put(5.00,18.00){\makebox(0,0)[cc]{1}}
\put(35.00,18.00){\makebox(0,0)[cc]{$1 \mapsto$}}
\put(88.00,18.00){\makebox(0,0)[cc]{$2 \mapsto$}}
\put(70.00,4.00){\line(-1,-2){2.00}} \put(68.00,0.00){\vector(-1,2){2.00}}
\put(40.00,4.00){\line(-5,-3){7.00}} \put(33.00,-0.33){\line(0,0){0.00}}
\put(33.00,0.00){\line(-3,2){6.00}} \put(23.00,4.00){\makebox(0,0)[cc]{$+$}}
\put(59.00,18.00){\vector(3,1){6.00}} \put(59.00,18.00){\vector(3,-1){6.00}}
\put(61.00,21.00){\makebox(0,0)[cc]{{\small 3}}}
\put(61.00,15.00){\makebox(0,0)[cc]{{\small 1}}}
\end{picture}
\ ,
\end{equation}
in the 4th cell of (\ref{marg}) it is equal to
\begin{equation}
\label{mar2} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(26.00,13.00)
\put(2.00,9.00){\line(1,0){24.00}} \put(13.00,9.00){\oval(13.00,8.00)[t]}
\put(6.50,9.00){\makebox(0,0)[cc]{\large *}}
\put(15.00,6.00){\makebox(0,0)[cc]{\small $-f_{1}'''(1)/f_{1}''(1) \sim$}}
\put(15.00,2.00){\makebox(0,0)[cc]{\small $\sim \ \mapsto / f_{1}''(1)$}}
\end{picture}
\ \qquad ,
\end{equation}
and its intersections with all other cells are empty.
The sum of the first and the third terms in (\ref{homztrip2}) is equal to the
variety denoted by the subscript \unitlength 1.00mm \special{em:linewidth
0.4pt} \linethickness{0.4pt}
\begin{picture}(11.50,3.00)
\put(6.00,1.00){\makebox(0,0)[cc]{$1 \longleftrightarrow 2$}}
\put(6.00,-0.50){\line(0,1){3.00}}
\end{picture}
\ .
To kill it (and something else) we consider the equality
\begin{equation}
\label{nev3} \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(125.00,17.00)
\put(5.00,14.00){\line(1,0){16.00}} \put(9.50,14.00){\oval(7.00,6.00)[t]}
\put(16.50,14.00){\oval(7.00,6.00)[t]} \put(9.00,5.00){\vector(2,1){6.00}}
\put(9.00,5.00){\vector(2,-1){6.00}} \put(9.00,5.00){\vector(1,0){10.00}}
\put(9.00,3.00){\line(0,1){4.00}} \put(12.00,9.00){\makebox(0,0)[cc]{\small
$1$}} \put(12.00,1.00){\makebox(0,0)[cc]{\small $2$}}
\put(2.00,14.00){\makebox(0,0)[cc]{$\partial$}}
\put(24.00,14.00){\makebox(0,0)[cc]{$=$}} \put(26.00,14.00){\line(1,0){16.00}}
\put(30.50,14.00){\oval(7.00,6.00)[t]} \put(37.50,14.00){\oval(7.00,6.00)[t]}
\put(47.00,14.00){\line(1,0){16.00}} \put(51.50,14.00){\oval(7.00,6.00)[t]}
\put(58.50,14.00){\oval(7.00,6.00)[t]} \put(68.00,14.00){\line(1,0){16.00}}
\put(72.50,14.00){\oval(7.00,6.00)[t]} \put(79.50,14.00){\oval(7.00,6.00)[t]}
\put(87.00,14.00){\makebox(0,0)[cc]{$+$}}
\put(66.00,14.00){\makebox(0,0)[cc]{$+$}}
\put(45.00,14.00){\makebox(0,0)[cc]{$+$}} \put(89.00,14.00){\line(1,0){16.00}}
\put(97.00,14.00){\oval(10.00,6.00)[t]}
\put(92.00,14.00){\makebox(0,0)[cc]{\large *}}
\put(108.00,14.00){\makebox(0,0)[cc]{$+$}}
\put(110.00,14.00){\line(1,0){15.00}} \put(117.50,14.00){\oval(9.00,6.00)[t]}
\put(97.00,8.00){\makebox(0,0)[cc]{\small $f_{1}'''(1)/f_{1}''(1) \sim$}}
\put(97.00,4.00){\makebox(0,0)[cc]{\small $\sim \ \mapsto /f_{1}''(1)$}}
\put(114.00,3.00){\line(0,1){4.00}} \put(114.00,5.00){\vector(1,0){10.00}}
\put(114.00,5.00){\vector(2,1){8.00}} \put(118.00,9.00){\makebox(0,0)[cc]{$1$}}
\put(122.00,1.00){\vector(-2,1){6.00}} \put(118.00,3.00){\vector(-2,1){4.00}}
\put(117.00,1.00){\makebox(0,0)[cc]{$2$}} \put(33.00,3.00){\line(0,1){4.00}}
\put(33.00,5.00){\vector(1,0){6.00}} \put(33.00,5.00){\vector(-1,0){6.00}}
\put(25.00,5.00){\makebox(0,0)[cc]{$1$}}
\put(41.00,5.00){\makebox(0,0)[cc]{$2$}} \put(53.00,5.00){\makebox(0,0)[cc]{$1
\mapsto$}} \put(74.00,5.00){\makebox(0,0)[cc]{$2 \mapsto$}}
\put(122.00,14.00){\makebox(0,0)[cc]{{\large *}}}
\end{picture}
\ .
\end{equation}
The variety in its left part consists of such points of the cell \unitlength
1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(18.00,4.00)
\put(0.00,0.00){\line(1,0){16.00}} \put(4.50,0.00){\oval(7.00,4.00)[t]}
\put(11.50,0.00){\oval(7.00,4.00)[t]}
\end{picture}
\
that the direction "to the right" in $\R^{n-1}$ lies between the projections of
$f'(a_1)$ and $f'(a_2)$ to $\R^{n-1}$. The sum of three first terms in the
right-hand part of (\ref{nev3}) is equal to entire (\ref{homztrip2}). The
subscript under the fourth term in (\ref{nev3}) means almost the same as in
(\ref{nev2}) or (\ref{mar2}), but now the two frames compared there should
define {\em equal} orientations.
Finally, the last term in (\ref{nev3}) belongs to the 5th cell in (\ref{marg}).
This cell can be considered as the space of triples $(\alpha,t,f)$ where
$\alpha$ is a pair of points $(a<b)$ in $\R^1$, $f$ a map $\R^1 \to \R^n$ such
that $f(a)=f(b), f'(b)=0$, and $t$ is a point of a segment participating in the
construction of the simplicial resolution (its endpoints formally correspond to
the above two linear conditions). The subscript under the picture of this cell
in (\ref{nev3}) denotes a subvariety in the space of such triples, defined by
the following additional condition: the direction "to the right" in $\R^{n-1}$
belongs to the angle between projections of vectors $f'(a)$ and $-f''(b)$. Here
the number of arrows labeled by $2$ shows us the order of the derivative at the
second point $b$ participating in this condition, and the reversed direction of
these arrows indicates that we need to take this derivative with the opposite
sign.
Now we span the chain (\ref{homztrip3}) inside the cell \unitlength 1.00mm
\special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(20.00,4.00)
\put(0.00,0.00){\line(1,0){20.00}} \put(10.00,0.00){\oval(16.00,8.00)[t]}
\put(14.00,0.00){\oval(8.00,4.00)[t]}
\end{picture}
. First of all we kill the 5th picture in
(\ref{homztrip3}) by the variety shown in the left part of the next equality:
\begin{equation}
\label{span3} \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(109.00,26.00)
\put(3.00,22.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,22.00){\line(1,0){21.00}} \put(18.50,22.00){\oval(13.00,8.00)[t]}
\put(21.50,22.00){\oval(7.00,4.00)[t]} \put(9.00,22.00){\line(2,-1){8.00}}
\put(17.00,18.00){\vector(1,1){4.00}} \put(33.00,22.00){\line(1,0){21.00}}
\put(45.50,22.00){\oval(13.00,8.00)[t]} \put(48.50,22.00){\oval(7.00,4.00)[t]}
\put(36.00,22.00){\line(2,-1){8.00}} \put(44.00,18.00){\line(1,1){4.00}}
\put(30.00,22.00){\makebox(0,0)[cc]{$=$}}
\put(57.00,22.00){\makebox(0,0)[cc]{$+$}} \put(60.00,22.00){\line(1,0){20.00}}
\put(71.00,22.00){\oval(14.00,8.00)[t]} \put(74.00,22.00){\oval(8.00,4.00)[t]}
\put(64.00,22.00){\line(3,-2){6.00}} \put(70.00,18.00){\vector(1,1){4.00}}
\put(83.00,22.00){\makebox(0,0)[cc]{$+$}} \put(86.00,22.00){\line(1,0){20.00}}
\put(97.00,22.00){\oval(14.00,8.00)[t]} \put(100.00,22.00){\oval(8.00,4.00)[t]}
\put(109.00,22.00){\makebox(0,0)[cc]{$+$}} \put(33.00,8.00){\line(1,0){20.00}}
\put(44.00,8.00){\oval(14.00,8.00)[t]} \put(47.00,8.00){\oval(8.00,4.00)[t]}
\put(56.00,8.00){\makebox(0,0)[cc]{$+$}} \put(60.00,8.00){\line(1,0){20.00}}
\put(71.00,8.00){\oval(14.00,8.00)[t]} \put(74.50,8.00){\oval(7.00,4.00)[t]}
\put(86.00,8.00){\line(1,0){20.00}} \put(97.00,8.00){\oval(14.00,8.00)[t]}
\put(100.50,8.00){\oval(7.00,4.00)[t]} \put(83.00,8.00){\makebox(0,0)[cc]{$+$}}
\put(92.00,18.00){\line(-5,4){5.00}} \put(92.00,18.00){\vector(1,1){4.00}}
\put(43.00,4.00){\line(-2,1){8.00}} \put(43.00,4.00){\vector(2,1){8.00}}
\put(71.00,6.00){\vector(-1,-2){3.00}} \put(71.00,6.00){\vector(1,-2){3.00}}
\put(76.00,3.00){\makebox(0,0)[cc]{1}} \put(66.00,3.00){\makebox(0,0)[cc]{2}}
\put(94.00,3.50){\vector(3,-1){6.00}} \put(94.00,3.50){\vector(3,1){6.00}}
\put(96.00,6.00){\makebox(0,0)[cc]{{\small 3}}}
\put(96.00,1.00){\makebox(0,0)[cc]{{\small 1}}}
\end{picture}
,
\end{equation}
thus reducing it to the sum of other five pictures in the right part of this
equality. The last picture in the upper row of (\ref{span3}) and the first
picture in the lower row denote one and the same set and annihilate. The first
term in (\ref{homztrip3}) together with the second from the end term in
(\ref{span3}) form a subvariety in the same cell defined by the condition of
the type \unitlength 1.00mm \special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(11.50,3.00)
\put(6.00,1.00){\makebox(0,0)[cc]{$1 \longleftrightarrow 2$}}
\put(6.00,-0.50){\line(0,1){3.00}}
\end{picture}
\ .
It is natural to kill it by the left part of the following equation:
\begin{equation}
\label{span4} \unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(125.00,15.00)
\put(3.00,11.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,11.00){\line(1,0){19.00}} \put(15.50,11.00){\oval(15.00,8.00)[t]}
\put(19.00,11.00){\oval(8.00,4.00)[t]} \put(31.00,11.00){\line(1,0){19.00}}
\put(40.50,11.00){\oval(15.00,8.00)[t]} \put(44.00,11.00){\oval(8.00,4.00)[t]}
\put(56.00,11.00){\line(1,0){19.00}} \put(65.50,11.00){\oval(15.00,8.00)[t]}
\put(69.00,11.00){\oval(8.00,4.00)[t]} \put(81.00,11.00){\line(1,0){19.00}}
\put(90.50,11.00){\oval(15.00,8.00)[t]} \put(94.00,11.00){\oval(8.00,4.00)[t]}
\put(106.00,11.00){\line(1,0){19.00}}
\put(103.00,11.00){\makebox(0,0)[cc]{$+$}}
\put(78.00,11.00){\makebox(0,0)[cc]{$+$}}
\put(53.00,11.00){\makebox(0,0)[cc]{$+$}}
\put(28.00,11.00){\makebox(0,0)[cc]{=}} \put(12.00,4.00){\vector(2,1){6.00}}
\put(12.00,4.00){\vector(2,-1){6.00}} \put(12.00,4.00){\vector(1,0){10.00}}
\put(12.00,2.00){\line(0,1){4.00}} \put(14.00,7.00){\makebox(0,0)[cc]{1}}
\put(14.00,1.00){\makebox(0,0)[cc]{2}} \put(62.00,6.00){\makebox(0,0)[cc]{$1
\mapsto$}} \put(90.00,6.00){\makebox(0,0)[cc]{$2 \mapsto$}}
\put(40.00,6.00){\line(0,-1){4.00}} \put(46.00,4.00){\vector(1,0){0.2}}
\put(40.00,4.00){\line(1,0){6.00}} \put(34.00,4.00){\vector(-1,0){0.2}}
\put(40.00,4.00){\line(-1,0){6.00}} \put(31.00,4.00){\makebox(0,0)[cc]{$1$}}
\put(49.00,4.00){\makebox(0,0)[cc]{$2$}}
\put(115.50,11.00){\oval(13.00,8.00)[t]}
\put(122.00,11.00){\makebox(0,0)[cc]{$*$}} \put(112.00,3.00){\line(0,1){4.00}}
\put(112.00,5.00){\vector(1,0){10.00}} \put(112.00,5.00){\vector(2,1){8.00}}
\put(116.00,9.00){\makebox(0,0)[cc]{$1$}}
\put(120.00,1.00){\vector(-2,1){4.00}} \put(116.00,3.00){\vector(-2,1){4.00}}
\put(115.00,1.00){\makebox(0,0)[cc]{$2$}}
\end{picture}
\ .
\end{equation}
Summing up all terms in right-hand parts of equations
(\ref{nev3})--(\ref{span4}) and subtracting the chains (\ref{homztrip2}),
(\ref{homztrip3}), (\ref{mar2}), we annihilate almost all of their summands
except for the term (\ref{mar2}) and the second from the right term of
(\ref{nev3}). The sum of these two terms is equal to the right-hand part of the
identity
\begin{equation}
\label{span5} \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(72.00,10.00)
\put(3.00,6.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,6.00){\line(1,0){24.00}} \put(18.00,6.00){\oval(18.00,8.00)[t]}
\put(12.50,6.00){\oval(7.00,4.00)[t]} \put(39.00,6.00){\makebox(0,0)[cc]{$=$}}
\put(48.00,6.00){\line(1,0){24.00}} \put(59.50,6.00){\oval(17.00,8.00)[t]}
\put(51.00,6.00){\makebox(0,0)[cc]{$*$}}
\put(6.00,0.00){\framebox(25.00,4.00)[cc]{\small
$f_{1}'(1),f_{1}'(2),\mapsto$}}
\put(46.00,0.00){\framebox(28.00,4.00)[cc]{\small
$f_{1}''(1),f_{1}'''(1),\mapsto$}}
\end{picture}
\ \ .
\end{equation}
The subscript under this right-hand part means that the projections of
$f''(a_1)$ and $f'''(a_1)$ to $\R^{n-1}$ and the direction "to the right"
should be linearly dependent; the subscript in the left part says the same
about projections of vectors $f'(a_1)$ and $f'(a_2)$. If $n=3$ then both these
subscripts mean nothing.
Summarizing, we get that for the desired chain spanning (\ref{difone}) in $F_2
\setminus F_1$ we can take the sum of varieties shown in left parts of
equalities (\ref{boundd})--(\ref{nev2}) and (\ref{nev3})--(\ref{span5}), i.e.
the chain
\begin{equation}
\label{spanlast} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(124.00,30.00)
\put(0.00,26.00){\line(1,0){19.00}} \put(9.50,26.00){\oval(15.00,8.00)[t]}
\put(6.00,26.00){\oval(8.00,4.00)[t]} \put(13.50,26.00){\oval(7.00,4.00)[t]}
\put(8.00,18.00){\vector(1,2){3.00}} \put(8.00,18.00){\vector(-1,2){3.00}}
\put(3.00,21.00){\makebox(0,0)[cc]{2}} \put(13.00,21.00){\makebox(0,0)[cc]{1}}
\put(22.00,26.00){\makebox(0,0)[cc]{$+$}} \put(25.00,26.00){\line(1,0){19.00}}
\put(34.50,26.00){\oval(15.00,8.00)[t]} \put(31.00,26.00){\oval(8.00,4.00)[t]}
\put(38.50,26.00){\oval(7.00,4.00)[t]}
\put(47.00,26.00){\makebox(0,0)[cc]{$+$}} \put(50.00,26.00){\line(1,0){19.00}}
\put(59.50,26.00){\oval(15.00,8.00)[t]} \put(56.00,26.00){\oval(8.00,4.00)[t]}
\put(63.50,26.00){\oval(7.00,4.00)[t]}
\put(72.00,26.00){\makebox(0,0)[cc]{$+$}} \put(35.00,26.00){\line(-1,-2){2.00}}
\put(33.00,22.00){\vector(-1,2){2.00}} \put(57.00,21.00){\makebox(0,0)[cc]{$1
\mapsto$}} \put(75.00,26.00){\line(1,0){19.00}}
\put(82.50,26.00){\oval(11.00,8.00)[b]} \put(86.50,26.00){\oval(11.00,8.00)[t]}
\put(81.00,26.00){\line(-1,4){1.00}} \put(80.00,30.00){\vector(-1,-4){1.00}}
\put(97.00,26.00){\makebox(0,0)[cc]{$+$}} \put(100.00,26.00){\line(1,0){19.00}}
\put(122.00,26.00){\makebox(0,0)[cc]{$+$}} \put(3.00,10.00){\line(1,0){19.00}}
\put(1.00,10.00){\makebox(0,0)[cc]{$+$}}
\put(10.50,10.00){\oval(11.00,8.00)[b]} \put(14.50,10.00){\oval(11.00,8.00)[t]}
\put(12.00,2.00){\makebox(0,0)[cc]{$2 \mapsto$}}
\put(25.00,10.00){\makebox(0,0)[cc]{$+$}} \put(28.00,10.00){\line(1,0){19.00}}
\put(34.00,10.00){\oval(8.00,6.00)[t]} \put(42.00,10.00){\oval(8.00,6.00)[t]}
\put(35.00,4.00){\vector(2,1){6.00}} \put(35.00,4.00){\vector(2,-1){6.00}}
\put(35.00,4.00){\vector(1,0){10.00}} \put(35.00,2.00){\line(0,1){4.00}}
\put(37.00,0.00){\makebox(0,0)[cc]{2}} \put(37.00,8.00){\makebox(0,0)[cc]{1}}
\put(50.00,10.00){\makebox(0,0)[cc]{$+$}} \put(53.00,10.00){\line(1,0){19.00}}
\put(63.50,10.00){\oval(13.00,8.00)[t]} \put(66.50,10.00){\oval(7.00,6.00)[t]}
\put(55.00,10.00){\line(3,-2){6.00}} \put(61.00,6.00){\vector(3,2){6.00}}
\put(75.00,10.00){\makebox(0,0)[cc]{$+$}} \put(78.00,10.00){\line(1,0){19.00}}
\put(88.00,10.00){\oval(14.00,8.00)[t]} \put(91.50,10.00){\oval(7.00,6.00)[t]}
\put(85.00,4.00){\vector(2,1){6.00}} \put(85.00,4.00){\vector(2,-1){6.00}}
\put(85.00,4.00){\vector(1,0){10.00}} \put(85.00,2.00){\line(0,1){4.00}}
\put(87.00,0.00){\makebox(0,0)[cc]{2}} \put(87.00,8.00){\makebox(0,0)[cc]{1}}
\put(108.50,26.00){\oval(15.00,8.00)[b]}
\put(111.00,26.00){\oval(14.00,8.00)[t]} \put(104.00,26.00){\line(5,3){5.00}}
\put(109.00,29.00){\vector(4,-3){4.00}} \put(103.00,10.00){\line(1,0){21.00}}
\put(113.50,10.00){\oval(15.00,8.00)[t]}
\put(109.50,10.00){\oval(7.00,4.00)[t]}
\put(100.00,10.00){\makebox(0,0)[cc]{$+$}}
\put(101.00,2.00){\framebox(25.00,5.00)[cc]{\small
$f_{1}'(1),f_{1}'(2),\mapsto$}}
\end{picture}
\ .
\end{equation}
\subsection{The second differential and its homology to zero}
Now let us consider the boundary of the chain (\ref{spanlast}) in the term
$F_1$ of the filtration. This term consists of two cells, one of which is
characterized by a single chord and the second by one asterisk; see
(\ref{dif01}). It is easy to see that the first three summands in
(\ref{spanlast}) do not have any homological boundary in these cells, and the
next seven have two components of the boundary each, and these pairs of
components are shown consecutively in the next formula (\ref{diftwo}):
\begin{equation}
\label{diftwo} \unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(120.00,77.00)
\put(0.00,72.00){\line(1,0){25.00}} \put(10.00,72.00){\oval(16.00,10.00)[b]}
\put(23.00,72.00){\line(-3,2){6.00}} \put(17.00,76.00){\line(-1,-1){4.00}}
\put(13.00,72.00){\line(-4,5){4.00}} \put(9.00,77.00){\vector(-3,-4){3.67}}
\put(28.00,72.00){\makebox(0,0)[cc]{$+$}} \put(31.00,72.00){\line(1,0){25.00}}
\put(47.50,72.00){\oval(13.00,10.00)[b]} \put(33.00,72.00){\line(2,1){10.00}}
\put(43.00,77.00){\line(6,-5){6.00}} \put(41.00,72.00){\line(-2,-5){2.00}}
\put(39.00,67.00){\vector(-2,3){3.33}}
\put(59.00,72.00){\makebox(0,0)[cc]{$+$}} \put(62.00,72.00){\line(1,0){25.00}}
\put(72.00,72.00){\oval(16.00,10.00)[b]} \put(85.00,72.00){\line(-2,1){10.00}}
\put(75.00,77.00){\line(-6,-5){6.00}} \put(69.00,72.00){\line(1,-2){2.00}}
\put(71.00,68.00){\vector(1,1){4.00}} \put(90.00,72.00){\makebox(0,0)[cc]{$+$}}
\put(93.00,72.00){\line(1,0){24.00}} \put(108.00,72.00){\oval(14.00,10.00)[b]}
\put(95.00,72.00){\line(2,1){10.00}} \put(105.00,77.00){\line(6,-5){6.00}}
\put(101.00,72.00){\line(1,-1){4.00}} \put(105.00,68.00){\vector(1,2){2.00}}
\put(120.00,72.00){\makebox(0,0)[cc]{$+$}} \put(6.00,54.00){\line(1,0){23.00}}
\put(14.50,54.00){\oval(15.00,10.00)[b]} \put(28.00,54.00){\line(-2,1){10.00}}
\put(18.00,59.00){\line(-6,-5){6.00}} \put(13.00,45.00){\makebox(0,0)[cc]{$2
\mapsto$}} \put(32.00,54.00){\makebox(0,0)[cc]{$+$}}
\put(35.00,54.00){\line(1,0){23.00}} \put(49.50,54.00){\oval(15.00,10.00)[b]}
\put(36.00,54.00){\line(2,1){10.00}} \put(46.00,59.00){\line(6,-5){6.00}}
\put(42.00,45.00){\makebox(0,0)[cc]{$2 \mapsto$}}
\put(61.00,54.00){\makebox(0,0)[cc]{$+$}} \put(65.00,54.00){\line(1,0){23.00}}
\put(71.50,54.00){\oval(11.00,10.00)[b]} \put(77.00,54.00){\line(1,-1){5.00}}
\put(82.00,49.00){\line(1,1){5.00}} \put(68.00,43.00){\vector(2,1){6.00}}
\put(68.00,43.00){\vector(2,-1){6.00}} \put(68.00,43.00){\vector(1,0){10.00}}
\put(68.00,41.00){\line(0,1){4.00}} \put(91.00,54.00){\makebox(0,0)[cc]{$+$}}
\put(94.00,54.00){\line(1,0){23.00}} \put(109.00,54.00){\oval(12.00,10.00)[b]}
\put(103.00,54.00){\line(-4,-5){4.00}} \put(99.00,49.00){\line(-4,5){4.00}}
\put(96.00,43.00){\vector(2,1){6.00}} \put(96.00,43.00){\vector(2,-1){6.00}}
\put(96.00,43.00){\vector(1,0){10.00}} \put(96.00,41.00){\line(0,1){4.00}}
\put(120.00,54.00){\makebox(0,0)[cc]{$+$}} \put(5.00,30.00){\line(1,0){25.00}}
\put(32.00,30.00){\makebox(0,0)[cc]{$+$}} \put(35.00,30.00){\line(1,0){23.00}}
\put(61.00,30.00){\makebox(0,0)[cc]{$+$}} \put(65.00,30.00){\line(1,0){23.00}}
\put(76.50,30.00){\oval(21.00,10.00)[b]} \put(87.00,30.00){\line(-6,5){6.00}}
\put(81.00,35.00){\line(-6,-5){6.00}} \put(91.00,30.00){\makebox(0,0)[cc]{$+$}}
\put(94.00,30.00){\line(1,0){23.00}} \put(110.00,30.00){\oval(10.00,10.00)[b]}
\put(115.00,30.00){\line(-2,1){12.00}} \put(103.00,36.00){\line(-4,-3){8.00}}
\put(77.00,19.00){\vector(2,1){6.00}} \put(77.00,19.00){\vector(2,-1){6.00}}
\put(77.00,19.00){\vector(1,0){10.00}} \put(77.00,17.00){\line(0,1){4.00}}
\put(96.00,19.00){\vector(2,1){6.00}} \put(96.00,19.00){\vector(2,-1){6.00}}
\put(96.00,19.00){\vector(1,0){10.00}} \put(96.00,17.00){\line(0,1){4.00}}
\put(70.00,46.00){\makebox(0,0)[cc]{1}} \put(70.00,39.00){\makebox(0,0)[cc]{2}}
\put(96.00,41.00){\line(0,1){4.00}} \put(98.00,46.00){\makebox(0,0)[cc]{1}}
\put(98.00,39.00){\makebox(0,0)[cc]{2}} \put(77.00,17.00){\line(0,1){4.00}}
\put(79.00,22.00){\makebox(0,0)[cc]{1}} \put(79.00,15.00){\makebox(0,0)[cc]{2}}
\put(96.00,17.00){\line(0,1){4.00}} \put(96.00,17.00){\line(0,1){4.00}}
\put(98.00,22.00){\makebox(0,0)[cc]{1}} \put(98.00,15.00){\makebox(0,0)[cc]{2}}
\put(6.00,30.00){\line(0,1){0.00}} \put(6.00,30.00){\line(1,-1){6.00}}
\put(12.00,24.00){\line(1,0){6.00}} \put(18.00,24.00){\vector(1,1){6.00}}
\put(20.00,30.00){\oval(18.00,10.00)[t]} \put(29.00,30.00){\line(-3,2){6.00}}
\put(23.00,34.00){\line(-3,-2){6.00}} \put(36.00,30.00){\line(1,-1){6.00}}
\put(42.00,24.00){\line(1,0){6.00}} \put(51.00,30.00){\oval(6.00,8.00)[t]}
\put(48.00,30.00){\line(-1,1){4.00}} \put(44.00,34.00){\line(-3,-4){3.00}}
\put(120.00,30.00){\makebox(0,0)[cc]{$+$}}
\put(31.00,10.00){\line(-1,0){25.00}} \put(18.50,10.00){\oval(19.00,10.00)[t]}
\put(36.00,10.00){\makebox(0,0)[cc]{$+$}} \put(41.00,10.00){\line(1,0){25.00}}
\put(48.00,10.00){\oval(8.00,6.00)[b]} \put(9.00,10.00){\line(2,-1){6.00}}
\put(15.00,7.00){\line(2,1){6.00}} \put(44.00,10.00){\line(2,1){10.00}}
\put(54.00,15.00){\line(2,-1){10.00}} \put(48.00,24.00){\vector(1,2){3.00}}
\put(6.00,0.00){\framebox(25.00,5.00)[cc]{\small $f_1'(1),f_1'(2),\mapsto$}}
\put(41.00,0.00){\framebox(25.00,5.00)[cc]{\small $f_1'(1),f_1'(2),\mapsto$}}
\put(3.00,10.00){\makebox(0,0)[cc]{$+$}}
\put(2.00,54.00){\makebox(0,0)[cc]{$+$}}
\put(2.00,30.00){\makebox(0,0)[cc]{$+$}}
\end{picture}
\ .
\end{equation}
The last pictures in the second and the third lines of this formula denote one
and the same variety and annihilate, so we get only the sum of remaining twelve
varieties.
Now let us span this sum by a chain in $F_1$. As usual, any time as we have a
variety characterized by a picture with a zigzag (without arrows) we represent
it as a component of the boundary of a variety, whose picture is obtained from
this one by adding an arrow at one of the ends of the zigzag. Performing this
systematically, we find some ten varieties of codimension 1 in the greater cell
of $F_1$. These varieties are encoded by the left parts of the following
equations (\ref{d2-1})--(\ref{d2-91}), whose right-hand parts express the
boundaries of these varieties.
\begin{equation}
\label{d2-1} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(124.00,29.00)
\put(3.00,24.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,24.00){\line(1,0){25.00}} \put(16.00,24.00){\oval(16.00,10.00)[b]}
\put(29.00,24.00){\line(-1,1){5.00}} \put(24.00,29.00){\vector(-1,-1){5.00}}
\put(19.00,24.00){\line(-3,5){3.00}} \put(16.00,29.00){\vector(-1,-1){5.00}}
\put(34.00,24.00){\makebox(0,0)[cc]{$=$}} \put(37.00,24.00){\line(1,0){25.00}}
\put(46.50,24.00){\oval(15.00,10.00)[b]} \put(60.00,24.00){\line(-1,1){5.00}}
\put(55.00,29.00){\line(-1,-1){5.00}} \put(50.00,24.00){\line(-3,5){3.00}}
\put(47.00,29.00){\vector(-1,-1){5.00}}
\put(65.00,24.00){\makebox(0,0)[cc]{$+$}} \put(68.00,24.00){\line(1,0){25.00}}
\put(77.50,24.00){\oval(15.00,10.00)[b]} \put(91.00,24.00){\line(-1,1){5.00}}
\put(86.00,29.00){\vector(-1,-1){5.00}} \put(81.00,24.00){\line(-4,5){4.00}}
\put(77.00,29.00){\line(-4,-5){4.00}} \put(96.00,24.00){\makebox(0,0)[cc]{$+$}}
\put(99.00,24.00){\line(1,0){22.00}} \put(107.50,24.00){\oval(13.00,10.00)[b]}
\put(119.00,24.00){\line(-1,1){5.00}} \put(114.00,29.00){\vector(-1,-1){5.00}}
\put(109.00,24.00){\line(-3,5){3.00}} \put(106.00,29.00){\vector(-1,-1){5.00}}
\put(124.00,24.00){\makebox(0,0)[cc]{$+$}}
\put(3.00,8.00){\makebox(0,0)[cc]{$+$}} \put(36.00,8.00){\line(1,0){25.00}}
\put(45.50,8.00){\oval(15.00,10.00)[b]} \put(60.00,8.00){\line(-3,4){3.00}}
\put(57.00,12.00){\vector(-1,-1){4.00}} \put(38.00,8.00){\line(1,1){5.00}}
\put(43.00,13.00){\vector(1,-1){5.00}} \put(64.00,8.00){\makebox(0,0)[cc]{$+$}}
\put(67.00,8.00){\line(1,0){24.00}} \put(79.00,8.00){\oval(20.00,10.00)[b]}
\put(89.00,8.00){\line(-3,5){3.00}} \put(86.00,13.00){\vector(-1,-1){5.00}}
\put(81.00,8.00){\line(-3,5){3.00}} \put(78.00,13.00){\vector(-1,-1){5.00}}
\put(94.00,8.00){\makebox(0,0)[cc]{$+$}} \put(7.00,8.00){\line(1,0){24.00}}
\put(16.00,8.00){\oval(14.00,10.00)[b]} \put(28.00,8.00){\line(-1,1){5.00}}
\put(23.00,13.00){\vector(-1,-1){5.00}} \put(27.00,2.00){\makebox(0,0)[cc]{$2
\uparrow$}} \put(34.00,8.00){\makebox(0,0)[cc]{$+$}}
\put(97.00,8.00){\line(1,0){24.00}} \put(106.00,8.00){\oval(14.00,10.00)[b]}
\put(119.00,8.00){\line(-3,4){3.00}} \put(116.00,12.00){\vector(-3,-4){3.00}}
\put(119.00,6.00){\vector(1,-2){3.00}} \put(119.00,6.00){\vector(-1,-2){3.00}}
\put(115.00,3.00){\makebox(0,0)[cc]{1}} \put(123.00,3.00){\makebox(0,0)[cc]{2}}
\end{picture}
\end{equation}
\begin{equation}
\label{d2-2} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(124.00,29.00)
\put(3.00,24.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,24.00){\line(1,0){25.00}} \put(15.50,24.00){\oval(15.00,10.00)[b]}
\put(29.00,24.00){\line(-1,1){5.00}} \put(24.00,29.00){\line(-1,0){11.00}}
\put(13.00,29.00){\vector(0,-1){5.00}} \put(13.00,24.00){\line(1,2){2.00}}
\put(15.00,28.00){\vector(3,-4){3.00}}
\put(34.00,24.00){\makebox(0,0)[cc]{$=$}} \put(37.00,24.00){\line(1,0){25.00}}
\put(46.50,24.00){\oval(15.00,10.00)[b]} \put(60.00,24.00){\line(-1,1){5.00}}
\put(55.00,29.00){\vector(-1,-1){5.00}} \put(50.00,24.00){\line(-3,5){3.00}}
\put(47.00,29.00){\line(-3,-5){3.00}} \put(65.00,24.00){\makebox(0,0)[cc]{$+$}}
\put(68.00,24.00){\line(1,0){25.00}} \put(77.50,24.00){\oval(15.00,10.00)[b]}
\put(91.00,24.00){\line(-3,5){3.00}} \put(88.00,29.00){\line(-1,0){14.00}}
\put(74.00,29.00){\line(0,-1){5.00}} \put(74.00,24.00){\line(3,4){3.00}}
\put(77.00,28.00){\vector(3,-4){3.00}}
\put(96.00,24.00){\makebox(0,0)[cc]{$+$}} \put(99.00,24.00){\line(1,0){22.00}}
\put(107.50,24.00){\oval(13.00,10.00)[b]} \put(119.00,24.00){\line(-1,1){5.00}}
\put(114.00,29.00){\vector(0,-1){5.00}} \put(101.00,24.00){\line(3,5){3.00}}
\put(104.00,29.00){\vector(1,-1){5.00}}
\put(124.00,24.00){\makebox(0,0)[cc]{$+$}}
\put(27.00,9.00){\makebox(0,0)[cc]{$+$}} \put(30.00,9.00){\line(1,0){19.00}}
\put(38.00,9.00){\oval(12.00,10.00)[b]} \put(48.00,9.00){\line(-1,1){5.00}}
\put(43.00,14.00){\vector(-1,-1){5.00}} \put(38.00,9.00){\line(-3,4){3.00}}
\put(35.00,13.00){\vector(-3,-4){3.00}} \put(2.00,9.00){\makebox(0,0)[cc]{$+$}}
\put(5.00,9.00){\line(1,0){18.00}} \put(12.00,9.00){\oval(12.00,10.00)[b]}
\put(22.00,9.00){\line(-1,1){5.00}} \put(17.00,14.00){\vector(-1,-1){5.00}}
\put(21.00,2.00){\makebox(0,0)[cc]{$2 \downarrow$}}
\put(51.00,9.00){\makebox(0,0)[cc]{$+$}} \put(54.00,9.00){\line(1,0){17.00}}
\put(62.50,9.00){\oval(15.00,10.00)[b]} \put(60.00,9.00){\line(1,1){5.00}}
\put(65.00,14.00){\vector(0,-1){5.00}} \put(55.00,9.00){\line(1,1){5.00}}
\put(60.00,14.00){\vector(0,-1){5.00}} \put(74.00,9.00){\makebox(0,0)[cc]{$+$}}
\put(77.00,9.00){\line(1,0){17.00}} \put(82.00,9.00){\oval(8.00,10.00)[b]}
\put(93.00,9.00){\line(-2,5){2.00}} \put(91.00,14.00){\vector(-1,-1){5.00}}
\put(91.00,-1.00){\vector(-1,2){3.00}} \put(91.00,-1.00){\vector(1,2){3.00}}
\put(95.00,2.00){\makebox(0,0)[cc]{2}} \put(87.00,2.00){\makebox(0,0)[cc]{1}}
\put(97.00,9.00){\makebox(0,0)[cc]{$+$}} \put(100.00,9.00){\line(1,0){16.00}}
\put(108.00,9.00){\oval(14.00,10.00)[b]} \put(115.00,9.00){\line(-3,5){3.00}}
\put(112.00,14.00){\vector(-1,-1){5.00}} \put(117.00,3.00){\vector(2,1){6.00}}
\put(117.00,3.00){\vector(2,-1){6.00}}
\put(119.00,-1.00){\makebox(0,0)[cc]{$1$}}
\put(119.00,7.00){\makebox(0,0)[cc]{$2$}}
\end{picture}
\end{equation}
\begin{equation}
\label{d2-3} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(125.00,29.00)
\put(2.00,24.00){\makebox(0,0)[cc]{$\partial$}}
\put(5.00,24.00){\line(1,0){25.00}} \put(14.50,24.00){\oval(15.00,8.00)[b]}
\put(28.00,24.00){\line(-6,5){6.00}} \put(22.00,29.00){\vector(-4,-3){6.67}}
\put(26.00,19.00){\makebox(0,0)[cc]{$2 \mapsto$}}
\put(33.00,24.00){\makebox(0,0)[cc]{$=$}} \put(36.00,24.00){\line(1,0){25.00}}
\put(45.00,24.00){\oval(14.00,8.00)[b]} \put(59.00,24.00){\line(-5,3){8.00}}
\put(51.00,28.67){\line(-5,-4){6.00}} \put(56.00,19.00){\makebox(0,0)[cc]{$2
\mapsto$}} \put(64.00,24.00){\makebox(0,0)[cc]{$+$}}
\put(67.00,24.00){\line(1,0){25.00}} \put(76.00,24.00){\oval(14.00,8.00)[b]}
\put(89.00,24.00){\line(-6,5){6.00}} \put(83.00,29.00){\vector(-4,-3){6.67}}
\put(87.00,19.00){\makebox(0,0)[cc]{$2 \updownarrow$}}
\put(95.00,24.00){\makebox(0,0)[cc]{$+$}} \put(98.00,24.00){\line(1,0){23.00}}
\put(106.50,24.00){\oval(13.00,8.00)[b]}
\put(117.00,19.00){\makebox(0,0)[cc]{$1 \mapsto$}}
\put(124.00,24.00){\makebox(0,0)[cc]{$+$}}
\put(1.00,8.00){\makebox(0,0)[cc]{$+$}} \put(4.00,8.00){\line(1,0){25.00}}
\put(12.00,8.00){\oval(12.00,8.00)[b]} \put(27.00,8.00){\line(-4,5){4.00}}
\put(23.00,13.00){\vector(-1,-1){5.00}} \put(18.00,0.00){\makebox(0,0)[cc]{$2
\mapsto$}} \put(32.00,8.00){\makebox(0,0)[cc]{$+$}}
\put(35.00,8.00){\line(1,0){24.00}} \put(47.00,8.00){\oval(20.00,8.00)[b]}
\put(57.00,8.00){\line(-1,1){5.00}} \put(52.00,13.00){\vector(-1,-1){5.00}}
\put(47.00,0.00){\makebox(0,0)[cc]{$2 \mapsto$}}
\put(62.00,8.00){\makebox(0,0)[cc]{$+$}} \put(65.00,8.00){\line(1,0){24.00}}
\put(76.00,8.00){\oval(16.00,8.00)[b]} \put(70.00,0.00){\makebox(0,0)[cc]{$1
\mapsto$}} \put(93.00,8.00){\makebox(0,0)[cc]{$+$}}
\put(96.00,8.00){\line(1,0){29.00}} \put(102.00,8.00){\makebox(0,0)[cc]{$*$}}
\put(122.00,8.00){\line(-2,1){10.00}} \put(112.00,13.00){\vector(-2,-1){10.00}}
\put(112.00,4.00){\makebox(0,0)[cc]{\small $f_1'''(1)/f_1''(1) \sim$}}
\put(112.00,0.00){\makebox(0,0)[cc]{\small $\sim \ \mapsto/ f_1''(1)$}}
\put(84.00,2.00){\vector(3,1){6.00}} \put(84.00,2.00){\vector(3,-1){6.00}}
\put(86.00,-1.00){\makebox(0,0)[cc]{$1$}}
\put(86.00,5.00){\makebox(0,0)[cc]{$2$}} \put(120.00,24.00){\line(-2,1){10.00}}
\put(110.00,29.00){\vector(-2,-1){10.00}}
\end{picture}
\end{equation}
\begin{equation}
\label{d2-4} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(125.00,38.00)
\put(3.00,33.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,33.00){\line(1,0){25.00}} \put(21.00,25.00){\vector(2,1){6.00}}
\put(21.00,25.00){\vector(2,-1){6.00}} \put(21.00,25.00){\vector(1,0){10.00}}
\put(21.00,23.00){\line(0,1){4.00}} \put(23.00,21.00){\makebox(0,0)[cc]{2}}
\put(23.00,29.00){\makebox(0,0)[cc]{1}}
\put(34.00,33.00){\makebox(0,0)[cc]{$=$}} \put(37.00,33.00){\line(1,0){25.00}}
\put(44.50,33.00){\oval(11.00,10.00)[b]} \put(60.00,33.00){\line(-1,1){5.00}}
\put(55.00,38.00){\line(-1,-1){5.00}} \put(13.50,33.00){\oval(11.00,10.00)[b]}
\put(29.00,33.00){\line(-1,1){5.00}} \put(24.00,38.00){\vector(-1,-1){5.00}}
\put(65.00,33.00){\makebox(0,0)[cc]{$+$}} \put(68.00,33.00){\line(1,0){25.00}}
\put(91.00,33.00){\line(-1,1){5.00}} \put(86.00,38.00){\vector(-1,-1){5.00}}
\put(75.50,33.00){\oval(11.00,10.00)[b]}
\put(96.00,33.00){\makebox(0,0)[cc]{$+$}} \put(52.00,25.00){\vector(2,1){6.00}}
\put(52.00,25.00){\vector(2,-1){6.00}} \put(52.00,25.00){\vector(1,0){10.00}}
\put(52.00,23.00){\line(0,1){4.00}} \put(54.00,21.00){\makebox(0,0)[cc]{2}}
\put(54.00,29.00){\makebox(0,0)[cc]{1}} \put(86.00,25.00){\makebox(0,0)[cc]{$1
\mapsto$}} \put(18.00,10.00){\makebox(0,0)[cc]{$+$}}
\put(98.00,33.00){\line(1,0){24.00}} \put(121.00,33.00){\line(-1,1){5.00}}
\put(116.00,38.00){\vector(-1,-1){5.00}}
\put(105.50,33.00){\oval(11.00,10.00)[b]}
\put(125.00,33.00){\makebox(0,0)[cc]{$+$}}
\put(116.00,25.00){\makebox(0,0)[cc]{$2 \mapsto$}}
\put(22.00,10.00){\line(1,0){25.00}} \put(29.50,10.00){\oval(11.00,10.00)[b]}
\put(45.00,10.00){\line(-1,1){5.00}} \put(40.00,15.00){\vector(-1,-1){5.00}}
\put(38.00,2.00){\makebox(0,0)[cc]{$1 \longleftrightarrow 2$}}
\put(38.00,0.50){\line(0,1){3.00}} \put(85.00,10.00){\makebox(0,0)[cc]{$+$}}
\put(88.00,10.00){\line(1,0){25.00}} \put(100.50,10.00){\oval(21.00,8.00)[t]}
\put(110.00,2.00){\makebox(0,0)[cc]{$2 \uparrow$}}
\put(51.00,10.00){\makebox(0,0)[cc]{$+$}} \put(54.00,10.00){\line(1,0){26.00}}
\put(77.00,10.00){\line(-2,1){10.00}} \put(67.00,15.00){\vector(-2,-1){10.00}}
\put(57.00,10.00){\makebox(0,0)[cc]{{\large $*$}}}
\put(67.00,5.00){\makebox(0,0)[cc]{\small $f_1'''(1)/f_1''(1) \sim$}}
\put(67.00,1.00){\makebox(0,0)[cc]{\small $\sim \ \mapsto/ f_1''(1)$}}
\put(90.00,5.00){\vector(1,0){12.00}} \put(90.00,5.00){\vector(3,1){9.00}}
\put(99.00,2.00){\vector(-3,1){6.00}} \put(96.00,3.00){\vector(-3,1){6.00}}
\put(90.00,2.00){\line(0,1){6.00}} \put(93.00,1.00){\makebox(0,0)[cc]{$2$}}
\put(93.00,8.00){\makebox(0,0)[cc]{$1$}}
\end{picture}
\end{equation}
\begin{equation}
\label{d2-5} \unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(126.00,42.00)
\put(3.00,37.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,37.00){\line(1,0){25.00}} \put(18.50,37.00){\oval(21.00,8.00)[b]}
\put(29.00,37.00){\line(-6,5){6.00}} \put(23.00,42.00){\vector(-1,-1){5.00}}
\put(15.00,27.00){\vector(2,1){6.00}} \put(15.00,27.00){\vector(2,-1){6.00}}
\put(15.00,27.00){\vector(1,0){10.00}} \put(15.00,25.00){\line(0,1){4.00}}
\put(16.50,30.20){\makebox(0,0)[cc]{1}} \put(16.50,23.50){\makebox(0,0)[cc]{2}}
\put(34.00,37.00){\makebox(0,0)[cc]{$=$}} \put(37.00,37.00){\line(1,0){24.00}}
\put(49.00,37.00){\oval(20.00,8.00)[b]} \put(59.00,37.00){\line(-6,5){6.00}}
\put(53.00,42.00){\vector(-1,-1){5.00}} \put(48.00,29.00){\makebox(0,0)[cc]{$1
\mapsto$}} \put(68.00,37.00){\line(1,0){24.00}}
\put(80.00,37.00){\oval(20.00,8.00)[b]} \put(90.00,37.00){\line(-6,5){6.00}}
\put(84.00,42.00){\vector(-1,-1){5.00}} \put(79.00,29.00){\makebox(0,0)[cc]{$2
\mapsto$}} \put(95.00,37.00){\makebox(0,0)[cc]{$+$}}
\put(95.00,37.00){\makebox(0,0)[cc]{$+$}} \put(98.00,37.00){\line(1,0){25.00}}
\put(110.50,37.00){\oval(21.00,8.00)[b]} \put(121.00,37.00){\line(-6,5){6.00}}
\put(115.00,42.00){\vector(-1,-1){5.00}}
\put(110.00,29.00){\makebox(0,0)[cc]{$1 \longleftrightarrow 2$}}
\put(110.00,27.50){\line(0,1){3.00}} \put(11.00,14.00){\makebox(0,0)[cc]{$+$}}
\put(14.00,14.00){\line(1,0){24.00}} \put(26.00,14.00){\oval(20.00,8.00)[b]}
\put(36.00,14.00){\line(-6,5){6.00}} \put(30.00,19.00){\line(-1,-1){5.00}}
\put(22.00,4.00){\vector(2,-1){6.00}} \put(22.00,4.00){\vector(2,1){6.00}}
\put(24.00,7.00){\makebox(0,0)[cc]{1}} \put(24.00,1.00){\makebox(0,0)[cc]{2}}
\put(22.00,4.00){\vector(1,0){10.00}} \put(22.00,2.00){\line(0,1){4.00}}
\put(48.00,14.00){\line(1,0){25.00}} \put(60.50,14.00){\oval(21.00,8.00)[t]}
\put(69.00,8.00){\makebox(0,0)[cc]{$2 \uparrow$}}
\put(43.00,14.00){\makebox(0,0)[cc]{$+$}}
\put(65.00,37.00){\makebox(0,0)[cc]{$+$}}
\put(126.00,37.00){\makebox(0,0)[cc]{$+$}}
\put(78.00,14.00){\makebox(0,0)[cc]{$+$}} \put(83.00,14.00){\line(1,0){26.00}}
\put(96.00,14.00){\oval(19.00,8.00)[t]} \put(84.00,6.00){\makebox(0,0)[cc]{$1
\downarrow$}} \put(103.00,8.00){\makebox(0,0)[cc]{\small $f_1'''(1)/f_1''(1)
\sim$}} \put(103.00,4.00){\makebox(0,0)[cc]{\small $\sim \/ \mapsto/
f_1''(1)$}} \put(49.00,6.00){\vector(1,0){12.00}}
\put(49.00,6.00){\vector(3,1){9.00}} \put(58.00,3.00){\vector(-3,1){6.00}}
\put(55.00,4.00){\vector(-3,1){6.00}} \put(49.00,3.00){\line(0,1){6.00}}
\put(52.00,2.00){\makebox(0,0)[cc]{$2$}}
\put(52.00,9.00){\makebox(0,0)[cc]{$1$}}
\end{picture}
\end{equation}
\begin{equation}
\label{d2-6} \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(125.00,32.00)
\put(3.00,27.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,27.00){\line(1,0){24.00}} \put(20.50,27.00){\oval(15.00,8.00)[b]}
\put(8.00,27.00){\line(3,5){3.00}} \put(20.00,32.00){\vector(1,-1){5.00}}
\put(11.00,32.00){\line(1,0){9.00}} \put(13.00,27.00){\line(3,4){3.00}}
\put(16.00,31.00){\vector(1,-2){2.00}}
\put(33.00,27.00){\makebox(0,0)[cc]{$=$}} \put(36.00,27.00){\line(1,0){24.00}}
\put(50.50,27.00){\oval(15.00,8.00)[b]} \put(38.00,27.00){\line(3,5){3.00}}
\put(41.00,32.00){\line(1,0){9.00}} \put(43.00,27.00){\line(3,4){3.00}}
\put(46.00,31.00){\vector(1,-2){2.00}} \put(50.00,32.00){\line(1,-1){5.00}}
\put(63.00,27.00){\makebox(0,0)[cc]{$+$}} \put(66.00,27.00){\line(1,0){24.00}}
\put(80.00,27.00){\oval(16.00,8.00)[b]} \put(68.00,27.00){\line(3,5){3.00}}
\put(71.00,32.00){\line(1,0){9.00}} \put(80.00,32.00){\vector(1,-1){5.00}}
\put(72.00,27.00){\line(3,4){3.00}} \put(75.00,31.00){\line(3,-4){3.00}}
\put(93.00,27.00){\makebox(0,0)[cc]{$+$}} \put(96.00,27.00){\line(1,0){24.00}}
\put(108.00,27.00){\oval(20.00,8.00)[b]} \put(98.00,27.00){\line(2,5){2.00}}
\put(100.00,32.00){\vector(1,-1){5.00}} \put(118.00,27.00){\line(-2,5){2.00}}
\put(116.00,32.00){\vector(-1,-1){5.00}}
\put(123.00,27.00){\makebox(0,0)[cc]{$+$}}
\put(2.00,11.00){\makebox(0,0)[cc]{$+$}} \put(5.00,11.00){\line(1,0){19.00}}
\put(18.50,11.00){\oval(9.00,8.00)[b]} \put(7.00,11.00){\line(6,5){6.00}}
\put(13.00,16.00){\vector(1,-1){5.00}} \put(11.00,5.00){\makebox(0,0)[cc]{$2
\downarrow$}} \put(27.00,11.00){\makebox(0,0)[cc]{$+$}}
\put(30.00,11.00){\line(1,0){19.00}} \put(44.00,11.00){\oval(8.00,8.00)[b]}
\put(32.00,11.00){\line(6,5){6.00}} \put(38.00,16.00){\vector(1,-1){5.00}}
\put(48.00,11.00){\line(0,1){5.00}} \put(48.00,16.00){\vector(-1,-1){5.00}}
\put(52.00,11.00){\makebox(0,0)[cc]{$+$}} \put(55.00,11.00){\line(1,0){19.00}}
\put(68.50,11.00){\oval(9.00,8.00)[b]} \put(57.00,11.00){\line(2,5){2.00}}
\put(59.00,16.00){\vector(1,-1){5.00}} \put(73.00,11.00){\line(-4,5){4.00}}
\put(69.00,16.00){\vector(0,-1){5.00}}
\put(77.00,11.00){\makebox(0,0)[cc]{$+$}} \put(105.00,11.00){\line(1,0){19.00}}
\put(114.50,11.00){\oval(15.00,8.00)[b]} \put(107.00,11.00){\line(3,5){3.00}}
\put(110.00,16.00){\vector(1,-1){5.00}}
\put(102.00,11.00){\makebox(0,0)[cc]{$+$}} \put(80.00,11.00){\line(1,0){20.00}}
\put(90.00,11.00){\oval(16.00,8.00)[b]} \put(98.00,11.00){\line(-3,5){3.00}}
\put(95.00,16.00){\vector(-1,-1){5.00}} \put(90.00,3.00){\makebox(0,0)[cc]{$1
\longleftrightarrow 2$}} \put(90.00,1.50){\line(0,1){3.00}}
\put(119.00,4.00){\vector(3,1){6.00}} \put(119.00,4.00){\vector(3,-1){6.00}}
\put(122.00,7.00){\makebox(0,0)[cc]{$3$}}
\put(122.00,1.00){\makebox(0,0)[cc]{$1$}}
\end{picture}
\end{equation}
\begin{equation}
\label{d2-7} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(102.00,30.00)
\put(2.00,25.00){\makebox(0,0)[cc]{$\partial$}}
\put(5.00,25.00){\line(1,0){19.00}} \put(16.50,25.00){\oval(11.00,8.00)[b]}
\put(7.00,25.00){\line(1,1){5.00}} \put(12.00,30.00){\vector(1,-1){5.00}}
\put(9.00,19.00){\makebox(0,0)[cc]{$2 \mapsto$}}
\put(27.00,25.00){\makebox(0,0)[cc]{$=$}} \put(30.00,25.00){\line(1,0){19.00}}
\put(41.50,25.00){\oval(11.00,8.00)[b]} \put(32.00,25.00){\line(1,1){5.00}}
\put(37.00,30.00){\line(1,-1){5.00}} \put(34.00,19.00){\makebox(0,0)[cc]{$2
\mapsto$}} \put(52.00,25.00){\makebox(0,0)[cc]{$+$}}
\put(55.00,25.00){\line(1,0){19.00}} \put(66.50,25.00){\oval(11.00,8.00)[b]}
\put(57.00,25.00){\line(1,1){5.00}} \put(62.00,30.00){\vector(1,-1){5.00}}
\put(58.00,19.00){\makebox(0,0)[cc]{$2 \updownarrow$}}
\put(77.00,25.00){\makebox(0,0)[cc]{$+$}} \put(80.00,25.00){\line(1,0){19.00}}
\put(89.50,25.00){\oval(15.00,8.00)[b]} \put(82.00,25.00){\line(3,5){3.00}}
\put(85.00,30.00){\vector(1,-1){5.00}} \put(81.00,19.00){\makebox(0,0)[cc]{$1
\mapsto$}} \put(102.00,25.00){\makebox(0,0)[cc]{$+$}}
\put(24.00,9.00){\line(1,0){19.00}} \put(33.00,9.00){\oval(14.00,8.00)[t]}
\put(27.00,4.00){\makebox(0,0)[cc]{$1 \mapsto$}}
\put(37.00,4.00){\vector(3,1){6.00}} \put(37.00,4.00){\vector(3,-1){6.00}}
\put(40.00,7.00){\makebox(0,0)[cc]{$2$}}
\put(40.00,1.00){\makebox(0,0)[cc]{$1$}}
\put(21.00,9.00){\makebox(0,0)[cc]{$+$}}
\put(53.00,9.00){\makebox(0,0)[cc]{$+$}} \put(63.00,9.00){\line(1,0){24.00}}
\put(75.00,9.00){\oval(18.00,8.00)[t]} \put(63.00,4.00){\makebox(0,0)[cc]{$1
\downarrow$}} \put(81.00,6.00){\makebox(0,0)[cc]{\small $f_1'''(1)/f_1''(1)
\sim$}} \put(80.00,2.00){\makebox(0,0)[cc]{\small $\sim \/ \mapsto/ f_1''(1)$}}
\end{picture}
\end{equation}
\begin{equation}
\label{d2-8} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(124.00,11.00)
\put(2.00,6.00){\makebox(0,0)[cc]{$\partial$}}
\put(5.00,6.00){\line(1,0){19.00}} \put(19.00,6.00){\oval(8.00,10.00)[b]}
\put(7.00,6.00){\line(6,5){6.00}} \put(13.00,11.00){\vector(1,-1){5.00}}
\put(15.00,6.00){\line(-4,-5){4.00}} \put(11.00,1.00){\vector(0,1){5.00}}
\put(27.00,6.00){\makebox(0,0)[cc]{$=$}} \put(30.00,6.00){\line(1,0){19.00}}
\put(44.00,6.00){\oval(8.00,10.00)[b]} \put(40.00,6.00){\line(-4,-5){4.00}}
\put(36.00,1.00){\vector(0,1){5.00}} \put(32.00,6.00){\line(6,5){6.00}}
\put(38.00,11.00){\line(6,-5){6.00}} \put(52.00,6.00){\makebox(0,0)[cc]{$+$}}
\put(55.00,6.00){\line(1,0){19.00}} \put(69.00,6.00){\oval(8.00,10.00)[b]}
\put(65.00,6.00){\line(-2,-5){2.00}} \put(63.00,1.00){\line(-2,5){2.00}}
\put(57.00,6.00){\line(6,5){6.00}} \put(63.00,11.00){\vector(1,-1){5.00}}
\put(77.00,6.00){\makebox(0,0)[cc]{$+$}} \put(80.00,6.00){\line(1,0){19.00}}
\put(93.50,6.00){\oval(9.00,10.00)[b]} \put(89.00,6.00){\line(-2,-5){2.00}}
\put(87.00,1.00){\vector(-1,1){5.00}} \put(82.00,6.00){\line(6,5){6.00}}
\put(88.00,11.00){\vector(1,-1){5.00}}
\put(102.00,6.00){\makebox(0,0)[cc]{$+$}} \put(105.00,6.00){\line(1,0){19.00}}
\put(118.50,6.00){\oval(9.00,10.00)[b]} \put(107.00,6.00){\line(6,5){6.00}}
\put(113.00,11.00){\vector(1,-1){5.00}} \put(109.00,1.00){\makebox(0,0)[cc]{$2
\uparrow$}}
\end{picture}
\end{equation}
\begin{equation}
\label{d2-9} \unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(125.00,16.00)
\put(2.00,11.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,11.00){\line(1,0){24.00}} \put(17.00,11.00){\oval(20.00,10.00)[t]}
\put(33.00,11.00){\makebox(0,0)[cc]{$=$}} \put(36.00,11.00){\line(1,0){24.00}}
\put(48.00,11.00){\oval(20.00,10.00)[t]} \put(38.00,11.00){\line(3,-2){5.00}}
\put(43.00,7.67){\line(3,2){5.00}} \put(64.00,11.00){\makebox(0,0)[cc]{$+$}}
\put(68.00,11.00){\line(1,0){24.00}} \put(80.00,11.00){\oval(16.00,10.00)[t]}
\put(69.00,7.50){\makebox(0,0)[cc]{$1 \downarrow$}}
\put(99.00,11.00){\line(1,0){24.00}} \put(112.00,11.00){\oval(16.00,10.00)[t]}
\put(122.00,7.50){\makebox(0,0)[cc]{$2 \uparrow$}}
\put(96.00,11.00){\makebox(0,0)[cc]{$+$}}
\put(67.00,0.00){\framebox(27.00,5.00)[cc]{\small
$f_1''(1),f_1'''(1),\mapsto$}}
\put(98.00,0.00){\framebox(27.00,5.00)[cc]{\small
$f_1''(1),f_1'''(1),\mapsto$}}
\put(35.50,0.00){\framebox(25.00,5.00)[cc]{\small $f_1'(1),f_1'(2),\mapsto$}}
\put(5.50,0.00){\framebox(25.00,5.00)[cc]{\small $f_1'(1),f_1'(2),\mapsto$}}
\put(7.00,11.00){\line(3,-2){6.00}} \put(13.00,7.00){\vector(3,2){6.00}}
\end{picture}
\end{equation}
\begin{equation}
\label{d2-91} \unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(126.00,16.00)
\put(3.00,11.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,11.00){\line(1,0){24.00}} \put(13.50,11.00){\oval(9.00,10.00)[b]}
\put(28.00,11.00){\line(-1,1){5.00}} \put(23.00,16.00){\vector(-1,-1){5.00}}
\put(34.00,11.00){\makebox(0,0)[cc]{$=$}} \put(69.00,11.00){\line(1,0){24.00}}
\put(91.00,11.00){\line(-2,1){10.00}} \put(81.00,16.00){\vector(-2,-1){10.00}}
\put(71.00,11.00){\makebox(0,0)[cc]{$*$}}
\put(66.00,11.00){\makebox(0,0)[cc]{$+$}} \put(37.00,11.00){\line(1,0){24.00}}
\put(44.50,11.00){\oval(9.00,10.00)[b]} \put(59.00,11.00){\line(-1,1){5.00}}
\put(98.00,11.00){\makebox(0,0)[cc]{$+$}} \put(101.00,11.00){\line(1,0){24.00}}
\put(113.00,11.00){\oval(18.00,10.00)[t]}
\put(119.00,7.50){\makebox(0,0)[cc]{$2 \uparrow$}}
\put(54.00,16.00){\line(-1,-1){5.00}}
\put(5.00,0.00){\framebox(26.00,5.00)[cc]{\small $f_1'(1),f_1'(2),\mapsto$}}
\put(67.00,0.00){\framebox(28.00,5.00)[cc]{\small
$f_1''(1),f_1'''(1),\mapsto$}}
\put(36.00,0.00){\framebox(26.00,5.00)[cc]{\small $f_1'(1),f_1'(2),\mapsto$}}
\put(100.00,0.00){\framebox(26.00,5.00)[cc]{\small $f_1'(1),f_1''(2),\mapsto$}}
\end{picture}
\end{equation}
Summing up the right-hand parts of these equations, we get the following
statement.
\begin{proposition}
The cycle $d^2(TT) \subset F_1$ presented by the linear combination
$($\ref{diftwo}$)$ is equal to the boundary $($mod 2$)$ of the sum of ten
varieties shown in the left parts of equations
$($\ref{d2-1}$)$--$($\ref{d2-91}$)$.
\end{proposition}
In this summation we use the following relations. Let us denote by (a;b) the
$b$-th summand in the right-hand part of the equation (a). Then (\ref{d2-1};6)
+ (\ref{d2-2};5) = (\ref{d2-3};2); \ (\ref{d2-1};7) + (\ref{d2-2};7) =
(\ref{d2-4};4); \ (\ref{d2-1};5) + (\ref{d2-2};6) = (\ref{d2-6};3); \
(\ref{d2-8};3) + (\ref{d2-6};6) = (\ref{d2-6};5); \ (\ref{d2-8};4) +
(\ref{d2-6};4) = (\ref{d2-7};2).
\subsection{The third differential}
Ten varieties described by left parts of equations (\ref{d2-1})--(\ref{d2-91})
form a chain in the term $F_1$ of the resolved discriminant, i.e. in the {\em
tautological resolution} of this discriminant, see \S \ref{method}.
Finally, we consider the image of this chain in the discriminant itself. The
image of any of ten components of this image is a subvariety in the space
${\mathcal K}_n$ of maps $\R^1 \to \R^n$, distinguished by conditions, whose
notation is obtained from the notation of the corresponding variety in $F_1$ by
replacing its unique chord by a zigzag with the same endpoints. It remains to
span the sum of these varieties by a chain in the space ${\mathcal K}_n$.
Proceeding as before, we find five varieties indicated in the left parts of the
following identities (\ref{d3-1})--(\ref{d3-5}).
\begin{equation}
\label{d3-1} \unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(123.00,46.00)
\put(3.00,40.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,40.00){\line(1,0){34.00}} \put(39.00,40.00){\line(-4,3){8.00}}
\put(31.00,46.00){\vector(-4,-3){8.00}} \put(23.00,40.00){\line(-2,3){4.00}}
\put(19.00,46.00){\vector(-2,-3){4.00}}
\put(43.00,40.00){\makebox(0,0)[cc]{$=$}} \put(46.00,40.00){\line(1,0){34.00}}
\put(79.00,40.00){\line(-4,3){8.00}} \put(71.00,46.00){\vector(-4,-3){8.00}}
\put(63.00,40.00){\line(-2,3){4.00}} \put(59.00,46.00){\vector(-2,-3){4.00}}
\put(86.00,40.00){\line(1,0){34.00}} \put(119.00,40.00){\line(-4,3){8.00}}
\put(111.00,46.00){\vector(-4,-3){8.00}} \put(87.00,40.00){\line(2,-1){12.00}}
\put(99.00,34.00){\vector(2,1){12.00}} \put(6.00,22.00){\line(1,0){34.00}}
\put(23.00,22.00){\line(-2,3){4.00}} \put(19.00,28.00){\vector(-2,-3){4.00}}
\put(83.00,40.00){\makebox(0,0)[cc]{$+$}}
\put(123.00,40.00){\makebox(0,0)[cc]{$+$}}
\put(3.00,22.00){\makebox(0,0)[cc]{$+$}}
\put(43.00,22.00){\makebox(0,0)[cc]{$+$}} \put(46.00,22.00){\line(1,0){34.00}}
\put(58.00,22.00){\line(-2,3){4.00}} \put(54.00,28.00){\vector(-1,-1){6.00}}
\put(83.00,22.00){\makebox(0,0)[cc]{$+$}} \put(86.00,22.00){\line(1,0){34.00}}
\put(110.00,16.00){\makebox(0,0)[cc]{$2 \uparrow$}}
\put(19.00,6.00){\makebox(0,0)[cc]{$+$}} \put(56.00,6.00){\line(-1,0){34.00}}
\put(44.00,6.00){\line(-2,3){4.00}} \put(40.00,12.00){\vector(-1,-1){6.00}}
\put(54.00,6.00){\line(-2,3){4.00}} \put(50.00,12.00){\vector(-1,-1){6.00}}
\put(59.00,6.00){\makebox(0,0)[cc]{$+$}} \put(62.00,6.00){\line(1,0){34.00}}
\put(64.00,6.00){\line(2,-1){12.00}} \put(76.00,0.00){\line(1,0){6.00}}
\put(82.00,0.00){\vector(2,1){12.00}} \put(94.00,6.00){\line(-2,3){4.00}}
\put(90.00,12.00){\vector(-1,-1){6.00}} \put(84.00,6.00){\line(-2,3){4.00}}
\put(80.00,12.00){\vector(-1,-1){6.00}}
\put(123.00,22.00){\makebox(0,0)[cc]{$+$}}
\put(119.00,22.00){\line(-1,1){6.00}} \put(113.00,28.00){\line(-1,0){9.00}}
\put(104.00,28.00){\vector(-1,-1){6.00}} \put(24.00,6.00){\line(1,-1){6.00}}
\put(30.00,0.00){\line(1,0){8.00}} \put(38.00,0.00){\vector(1,1){6.00}}
\put(78.00,22.00){\line(-1,1){6.00}} \put(72.00,28.00){\line(-1,0){8.00}}
\put(64.00,28.00){\vector(-1,-1){6.00}} \put(48.00,22.00){\line(1,-1){6.00}}
\put(54.00,16.00){\line(1,0){8.00}} \put(62.00,16.00){\vector(1,1){6.00}}
\put(7.00,40.00){\line(1,-1){6.00}} \put(13.00,34.00){\line(1,0){12.00}}
\put(25.00,34.00){\vector(1,1){6.00}} \put(39.00,22.00){\line(-4,3){8.00}}
\put(31.00,28.00){\line(-4,-3){8.00}} \put(103.00,40.00){\line(-2,3){4.00}}
\put(99.00,46.00){\line(-2,-3){4.00}} \put(47.00,40.00){\line(1,-1){6.00}}
\put(53.00,34.00){\line(1,0){12.00}} \put(65.00,34.00){\line(1,1){6.00}}
\put(7.00,22.00){\line(1,-1){6.00}} \put(13.00,16.00){\line(1,0){12.00}}
\put(25.00,16.00){\vector(1,1){6.00}} \put(88.00,22.00){\line(1,-1){6.00}}
\put(94.00,16.00){\line(1,0){8.00}} \put(102.00,16.00){\vector(1,1){6.00}}
\end{picture}
\end{equation}
\begin{equation}
\label{d3-2} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(123.00,51.00)
\put(3.00,45.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,45.00){\line(1,0){34.00}} \put(46.00,45.00){\line(1,0){34.00}}
\put(83.00,45.00){\makebox(0,0)[cc]{$+$}} \put(86.00,45.00){\line(1,0){34.00}}
\put(6.00,7.00){\line(1,0){34.00}} \put(46.00,7.00){\line(1,0){34.00}}
\put(83.00,7.00){\makebox(0,0)[cc]{$+$}} \put(86.00,7.00){\line(1,0){34.00}}
\put(6.00,26.00){\line(1,0){34.00}} \put(46.00,26.00){\line(1,0){34.00}}
\put(83.00,26.00){\makebox(0,0)[cc]{$+$}} \put(86.00,26.00){\line(1,0){34.00}}
\put(7.00,45.00){\line(2,-1){12.00}} \put(19.00,39.00){\vector(2,1){12.00}}
\put(39.00,45.00){\line(-4,3){8.00}} \put(31.00,51.00){\line(-1,0){8.00}}
\put(23.00,51.00){\vector(-4,-3){8.00}} \put(15.00,45.00){\line(1,-1){4.00}}
\put(19.00,41.00){\vector(1,1){4.00}} \put(43.00,45.00){\makebox(0,0)[cc]{$=$}}
\put(79.00,45.00){\line(-4,3){8.00}} \put(71.00,51.00){\line(-1,0){8.00}}
\put(63.00,51.00){\vector(-4,-3){8.00}} \put(55.00,45.00){\line(1,-1){4.00}}
\put(59.00,41.00){\vector(1,1){4.00}} \put(86.00,45.00){\line(1,0){34.00}}
\put(87.00,45.00){\line(2,-1){12.00}} \put(99.00,39.00){\vector(2,1){12.00}}
\put(119.00,45.00){\line(-4,3){8.00}} \put(111.00,51.00){\line(-1,0){8.00}}
\put(103.00,51.00){\vector(-4,-3){8.00}}
\put(123.00,45.00){\makebox(0,0)[cc]{$+$}}
\put(123.00,26.00){\makebox(0,0)[cc]{$+$}} \put(88.00,26.00){\line(4,-3){8.00}}
\put(96.00,20.00){\vector(2,1){12.00}} \put(118.00,26.00){\line(-4,3){8.00}}
\put(110.00,32.00){\vector(-2,-1){12.00}}
\put(109.00,19.00){\makebox(0,0)[cc]{$2 \downarrow$}}
\put(78.00,26.00){\line(-1,1){6.00}} \put(72.00,32.00){\line(-1,0){18.00}}
\put(54.00,32.00){\vector(-1,-1){6.00}} \put(48.00,26.00){\line(2,-3){4.00}}
\put(52.00,20.00){\vector(1,1){6.00}} \put(48.00,26.00){\line(3,1){12.00}}
\put(60.00,30.00){\vector(2,-1){8.00}}
\put(43.00,26.00){\makebox(0,0)[cc]{$+$}}
\put(3.00,26.00){\makebox(0,0)[cc]{$+$}} \put(7.00,26.00){\line(2,-1){12.00}}
\put(19.00,20.00){\vector(2,1){12.00}} \put(15.00,26.00){\line(1,-1){4.00}}
\put(19.00,22.00){\vector(1,1){4.00}} \put(3.00,7.00){\makebox(0,0)[cc]{$+$}}
\put(8.00,7.00){\line(1,-1){6.00}} \put(14.00,1.00){\line(1,0){8.00}}
\put(22.00,1.00){\vector(1,1){6.00}} \put(38.00,7.00){\line(-1,1){6.00}}
\put(32.00,13.00){\line(-1,0){8.00}} \put(24.00,13.00){\vector(-1,-1){6.00}}
\put(18.00,7.00){\line(2,1){8.00}} \put(26.00,11.00){\vector(1,-2){2.00}}
\put(43.00,7.00){\makebox(0,0)[cc]{$+$}} \put(48.00,7.00){\line(1,-1){6.00}}
\put(54.00,1.00){\line(1,0){18.00}} \put(72.00,1.00){\vector(1,1){6.00}}
\put(78.00,7.00){\line(-1,1){6.00}} \put(72.00,13.00){\line(-1,0){8.00}}
\put(64.00,13.00){\vector(-1,-1){6.00}} \put(58.00,7.00){\line(3,-2){6.00}}
\put(64.00,3.00){\vector(1,1){4.00}} \put(118.00,7.00){\line(-1,1){6.00}}
\put(112.00,13.00){\line(-1,0){18.00}} \put(94.00,13.00){\vector(-1,-1){6.00}}
\put(88.00,7.00){\line(3,-2){9.00}} \put(97.00,1.00){\vector(1,1){6.00}}
\put(109.00,2.00){\makebox(0,0)[cc]{$1 \longleftrightarrow 2$}}
\put(109.00,0.50){\line(0,1){3.00}} \put(15.00,26.00){\line(1,1){6.00}}
\put(21.00,32.00){\line(1,0){12.00}} \put(33.00,32.00){\line(1,-1){6.00}}
\put(47.00,45.00){\line(1,-1){6.00}} \put(53.00,39.00){\line(1,0){12.00}}
\put(65.00,39.00){\line(1,1){6.00}} \put(95.00,45.00){\line(1,-1){4.00}}
\put(99.00,41.00){\line(1,1){4.00}}
\end{picture}
\end{equation}
\begin{equation}
\label{d3-3} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(123.00,40.00)
\put(3.00,11.00){\makebox(0,0)[cc]{$+$}} \put(6.00,11.00){\line(1,0){22.00}}
\put(31.00,11.00){\makebox(0,0)[cc]{$+$}} \put(34.00,11.00){\line(1,0){22.00}}
\put(59.00,11.00){\makebox(0,0)[cc]{$+$}} \put(62.00,11.00){\line(1,0){22.00}}
\put(64.00,11.00){\line(3,2){9.00}} \put(73.00,17.00){\vector(3,-2){9.00}}
\put(82.00,11.00){\line(-3,-4){4.50}} \put(77.50,5.00){\vector(-3,4){4.50}}
\put(68.00,6.00){\makebox(0,0)[cc]{$2 \mapsto$}}
\put(36.00,11.00){\line(3,4){4.50}} \put(40.50,17.00){\vector(3,-4){4.50}}
\put(54.00,11.00){\line(-3,4){4.50}} \put(49.50,17.00){\vector(-3,-4){4.50}}
\put(42.00,6.00){\makebox(0,0)[cc]{$2 \mapsto$}}
\put(8.00,11.00){\line(3,-4){4.50}} \put(12.50,5.00){\vector(3,4){4.50}}
\put(26.00,11.00){\line(-3,2){9.00}} \put(17.00,17.00){\vector(-3,-2){9.00}}
\put(21.00,6.00){\makebox(0,0)[cc]{$1 \mapsto$}}
\put(3.00,35.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,35.00){\line(1,0){24.00}} \put(28.00,35.00){\line(-3,2){7.50}}
\put(20.50,40.00){\vector(-3,-2){7.50}} \put(8.00,35.00){\line(3,-2){7.50}}
\put(15.50,30.00){\vector(3,2){7.50}} \put(25.00,30.00){\makebox(0,0)[cc]{$2
\mapsto$}} \put(36.00,35.00){\line(1,0){24.00}}
\put(58.00,35.00){\line(-3,2){7.50}} \put(50.50,40.00){\vector(-3,-2){7.50}}
\put(38.00,35.00){\line(3,-2){7.50}} \put(45.50,30.00){\line(3,2){7.50}}
\put(55.00,30.00){\makebox(0,0)[cc]{$2 \mapsto$}}
\put(66.00,35.00){\line(1,0){24.00}} \put(88.00,35.00){\line(-3,2){7.50}}
\put(80.50,40.00){\line(-3,-2){7.50}} \put(68.00,35.00){\line(3,-2){7.50}}
\put(75.50,30.00){\vector(3,2){7.50}} \put(85.00,30.00){\makebox(0,0)[cc]{$2
\mapsto$}} \put(96.00,35.00){\line(1,0){24.00}}
\put(118.00,35.00){\line(-3,2){7.50}} \put(110.50,40.00){\vector(-3,-2){7.50}}
\put(98.00,35.00){\line(3,-2){7.50}} \put(105.50,30.00){\vector(3,2){7.50}}
\put(33.00,35.00){\makebox(0,0)[cc]{$=$}}
\put(63.00,35.00){\makebox(0,0)[cc]{$+$}}
\put(93.00,35.00){\makebox(0,0)[cc]{$+$}}
\put(123.00,35.00){\makebox(0,0)[cc]{$+$}}
\put(115.00,30.00){\makebox(0,0)[cc]{$2 \updownarrow$}}
\put(87.00,11.00){\makebox(0,0)[cc]{$+$}} \put(90.00,11.00){\line(1,0){29.00}}
\put(117.00,11.00){\line(-2,1){10.00}} \put(107.00,16.00){\line(-1,0){5.00}}
\put(102.00,16.00){\vector(-2,-1){10.00}} \put(92.00,5.00){\makebox(0,0)[cc]{$1
\downarrow$}} \put(111.00,7.00){\makebox(0,0)[cc]{\small $-f_1'''(1)/f_1''(1)
\sim$}} \put(111.00,3.00){\makebox(0,0)[cc]{\small $\sim \mapsto / f_1''(1)$}}
\end{picture}
\end{equation}
\begin{equation}
\label{d3-4} \unitlength=1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(122.00,35.00)
\put(2.00,30.00){\makebox(0,0)[cc]{$\partial$}}
\put(19.00,23.00){\vector(2,1){6.00}} \put(19.00,23.00){\vector(2,-1){6.00}}
\put(19.00,23.00){\vector(1,0){10.00}} \put(19.00,21.00){\line(0,1){4.00}}
\put(21.00,19.00){\makebox(0,0)[cc]{2}} \put(21.00,27.00){\makebox(0,0)[cc]{1}}
\put(5.00,30.00){\line(1,0){24.00}} \put(27.00,30.00){\line(-2,1){10.00}}
\put(17.00,35.00){\vector(-2,-1){10.00}} \put(7.00,30.00){\line(1,-1){5.00}}
\put(12.00,25.00){\vector(1,1){5.00}} \put(49.00,23.00){\vector(2,1){6.00}}
\put(49.00,23.00){\vector(2,-1){6.00}} \put(49.00,23.00){\vector(1,0){10.00}}
\put(49.00,21.00){\line(0,1){4.00}} \put(51.00,19.00){\makebox(0,0)[cc]{2}}
\put(51.00,27.00){\makebox(0,0)[cc]{1}} \put(35.00,30.00){\line(1,0){24.00}}
\put(57.00,30.00){\line(-2,1){10.00}} \put(37.00,30.00){\line(1,-1){5.00}}
\put(42.00,25.00){\vector(1,1){5.00}} \put(79.00,23.00){\vector(2,1){6.00}}
\put(79.00,23.00){\vector(2,-1){6.00}} \put(79.00,23.00){\vector(1,0){10.00}}
\put(79.00,21.00){\line(0,1){4.00}} \put(81.00,19.00){\makebox(0,0)[cc]{2}}
\put(81.00,27.00){\makebox(0,0)[cc]{1}} \put(65.00,30.00){\line(1,0){24.00}}
\put(87.00,30.00){\line(-2,1){10.00}} \put(77.00,35.00){\vector(-2,-1){10.00}}
\put(67.00,30.00){\line(1,-1){5.00}} \put(95.00,30.00){\line(1,0){24.00}}
\put(117.00,30.00){\line(-2,1){10.00}}
\put(107.00,35.00){\vector(-2,-1){10.00}} \put(97.00,30.00){\line(1,-1){5.00}}
\put(102.00,25.00){\vector(1,1){5.00}}
\put(32.00,30.00){\makebox(0,0)[cc]{$=$}}
\put(62.00,30.00){\makebox(0,0)[cc]{$+$}}
\put(92.00,30.00){\makebox(0,0)[cc]{$+$}}
\put(122.00,30.00){\makebox(0,0)[cc]{$+$}}
\put(111.00,24.00){\makebox(0,0)[cc]{$1 \mapsto$}}
\put(72.00,25.00){\line(1,1){5.00}} \put(47.00,35.00){\line(-2,-1){10.00}}
\put(9.00,9.00){\line(1,0){24.00}} \put(31.00,9.00){\line(-2,1){10.00}}
\put(21.00,14.00){\vector(-2,-1){10.00}} \put(11.00,9.00){\line(1,-1){5.00}}
\put(16.00,4.00){\vector(1,1){5.00}} \put(6.00,9.00){\makebox(0,0)[cc]{$+$}}
\put(36.00,9.00){\makebox(0,0)[cc]{$+$}} \put(25.00,3.00){\makebox(0,0)[cc]{$2
\mapsto$}} \put(39.00,9.00){\line(1,0){24.00}}
\put(61.00,9.00){\line(-2,1){10.00}} \put(51.00,14.00){\vector(-2,-1){10.00}}
\put(41.00,9.00){\line(1,-1){5.00}} \put(46.00,4.00){\vector(1,1){5.00}}
\put(69.00,9.00){\makebox(0,0)[cc]{$+$}} \put(58.00,5.00){\vector(-1,0){5.00}}
\put(58.00,5.00){\vector(1,0){5.00}} \put(58.00,3.00){\line(0,1){4.00}}
\put(65.00,5.00){\makebox(0,0)[cc]{$2$}}
\put(51.00,5.00){\makebox(0,0)[cc]{$1$}} \put(75.00,9.00){\line(1,0){29.00}}
\put(102.00,9.00){\line(-2,1){10.00}} \put(92.00,14.00){\line(-1,0){5.00}}
\put(87.00,14.00){\vector(-2,-1){10.00}} \put(77.00,3.00){\makebox(0,0)[cc]{$1
\downarrow$}} \put(95.00,5.00){\makebox(0,0)[cc]{\small $f_1'''(1)/f_1''(1)
\sim$}} \put(95.00,1.00){\makebox(0,0)[cc]{\small $\sim \mapsto / f_1''(1)$}}
\end{picture}
\end{equation}
\begin{equation}
\label{d3-5} \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(125.00,17.00)
\put(3.00,12.00){\makebox(0,0)[cc]{$\partial$}}
\put(6.00,12.00){\line(1,0){24.00}} \put(28.00,12.00){\line(-2,1){10.00}}
\put(18.00,17.00){\vector(-2,-1){10.00}} \put(8.00,12.00){\line(1,-1){5.00}}
\put(13.00,7.00){\vector(1,1){5.00}} \put(36.00,12.00){\line(1,0){24.00}}
\put(58.00,12.00){\line(-2,1){10.00}} \put(38.00,12.00){\line(1,-1){5.00}}
\put(43.00,7.00){\vector(1,1){5.00}} \put(66.00,12.00){\line(1,0){24.00}}
\put(88.00,12.00){\line(-2,1){10.00}} \put(78.00,17.00){\vector(-2,-1){10.00}}
\put(68.00,12.00){\line(1,-1){5.00}} \put(96.00,12.00){\line(1,0){29.00}}
\put(123.00,12.00){\line(-2,1){10.00}} \put(113.00,17.00){\line(-1,0){5.00}}
\put(108.00,17.00){\vector(-2,-1){10.00}}
\put(93.00,12.00){\makebox(0,0)[cc]{$+$}}
\put(63.00,12.00){\makebox(0,0)[cc]{$+$}}
\put(48.00,17.00){\line(-2,-1){10.00}} \put(73.00,7.00){\line(1,1){5.00}}
\put(99.00,8.00){\makebox(0,0)[cc]{$1 \downarrow$}}
\put(5.00,0.00){\framebox(26.00,5.00)[cc]{\small $f_1'(1),f_1'(2),\mapsto$}}
\put(35.00,0.00){\framebox(26.00,5.00)[cc]{\small $f_1'(1),f_1'(2),\mapsto$}}
\put(65.00,0.00){\framebox(26.00,5.00)[cc]{\small $f_1'(1),f_1'(2),\mapsto$}}
\put(96.50,0.00){\framebox(28.00,5.00)[cc]{\small
$f_1''(1),f_1'''(2),\mapsto$}} \put(34.00,12.00){\makebox(0,0)[cc]{$=$}}
\end{picture}
\end{equation}
It is easy to check that the homological sum of the right-hand parts of these
identities is equal to the sum of our ten discriminant varieties obtained from
left parts of equalities (\ref{d2-1})--(\ref{d2-91}). (We use the following
relations: (\ref{d3-1};5) + (\ref{d3-2};5) = (\ref{d3-3};3); \ (\ref{d3-3};6) +
(\ref{d3-4};4) = (\ref{d3-3};5); \ (\ref{d3-3};4) + (\ref{d3-4};6) =
(\ref{d3-5},3); \ (\ref{d3-1};4) + (\ref{d3-1};6) + (\ref{d3-1};7) +
(\ref{d3-2};4) + (\ref{d3-2};6) + (\ref{d3-2};7) = 0.)
Therefore the sum of the five varieties indicated in left parts of equalities
(\ref{d3-1})--(\ref{d3-5}) is the desired relative cycle in ${\mathcal K}_n \
(\mbox{mod } \Sigma)$.
The sum of the first and the second of these five varieties (respectively, the
third variety, respectively, the difference of the fifth and the fourth
varieties) is exactly the variety indicated in item a) (respectively, b),
respectively, c)) of Theorem 1, which is thus completely proved.
\subsection{Problems}
{\bf 1. Algorithmization.} How to do algorithmically all the same for any other
finite-type cohomology class of the space of knots?
Namely, let us consider any homology class $\gamma$ of the discriminant of the
space of knots, having some finite filtration ("order") $p$ and presented by
its "principal part", i.e. by the corresponding homology class in the term $F_p
\setminus F_{p-1}$ of the resolved discriminant. This class always is described
by some linear combination of pictures (generalized chord diagrams) as in
(\ref{prinpart}), (\ref{prinpar}), see \S \ \ref{method}. To get the
combinatorial description of a cohomology class with this principal part, we
need to calculate all the steps of the spectral sequence starting from this
part. To do it, on any step we need to find the chains spanning the consecutive
boundaries $d^r(\gamma) \subset F_{p-r} \setminus F_{p-r-1}$. Above we have
used some obvious rules: if a piece of our cycle $d^r(\gamma)$ is described by
a picture like in (\ref{difone})--(\ref{d3-4}), then it is natural to kill it
by a piece of the spanning chain, described by almost the same picture, only
replacing some one zigzag without arrows by the same zigzag with arrow at one
its end, or replacing some condition of type \unitlength 1.00mm
\special{em:linewidth 0.4pt} \linethickness{0.4pt}
\begin{picture}(12.00,1.00)
\put(6.00,-1.00){\vector(1,0){6.00}} \put(6.00,-1.00){\vector(-1,0){6.00}}
\put(6.00,-1.00){\circle*{1.00}} \put(1.00,1.00){\makebox(0,0)[cc]{\small 1}}
\put(11.00,1.00){\makebox(0,0)[cc]{\small 2}}
\end{picture}
\
by the condition of type \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(8.00,5.00)
\put(4.00,0.00){\vector(1,2){2.50}} \put(4.00,0.00){\vector(-1,2){2.50}}
\put(0.33,2.00){\makebox(0,0)[cc]{\small 1}}
\put(7.67,2.00){\makebox(0,0)[cc]{\small 2}}
\end{picture}
\ ,
or replacing some condition \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(11.50,3.00)
\put(6.00,1.00){\makebox(0,0)[cc]{$1 \longleftrightarrow 2$}}
\put(6.00,-0.50){\line(0,1){3.00}}
\end{picture}
\
by the condition \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(6.00,5.00)
\put(0.00,1.00){\vector(2,-1){4.00}} \put(0.00,1.00){\vector(2,1){4.00}}
\put(0.00,1.00){\vector(1,0){6.00}} \put(2.00,-2.00){\makebox(0,0)[cc]{$2$}}
\put(2.00,4.00){\makebox(0,0)[cc]{$1$}} \put(0.00,-0.50){\line(0,1){3.00}}
\end{picture}
\ , etc.
But how to decide, which of these fragments (and for which piece of the cycle)
to replace first? At which endpoint to put the arrow? Is it possible to do it
always in such a way that all the other components of the boundary of this
spanning variety would be in some sense "of lesser complexity" than the killed
one, so that our algorithm converges inductively? Which other subvarieties in
the cells of $F_p \setminus F_{p-1}$ can occur in the process of performing
this algorithm ? What are the formal rules for calculating their boundaries ?
I presume that the main filtering degree should be the number of points in
$\R^1$ participating in the definition of the subvariety, and the orientation
of arrows is not important: say, the algorithm will work if we orient all of
them from the right to the left (although, of course, other choice can provide
somewhat easier formulas).
\bigskip
{\bf 2. Orientable case.}
To do all the same for homology with integer coefficients, i.e. taking into
account orientations of our varieties. In this problem, the answers for odd and
even $n$ will be different: already the chain (\ref{prinpart}) is a $\Z$-cycle
in $F_3 \sm F_2$ only for even $n$. If $n$ is even, is it correct, that all the
calculations of \S~\ref{proof1} remain valid after imposing appropriate signs
before the pictures ?
\subsection{Proof of Proposition \protect{\ref{realiz}}}
\label{proreal}
First we specify a loop in the space of long knots as in this Proposition. We
can assume that the standard embedding $\R^1 \to \R^3$ (with which all long
knots should coincide close to the infinity) lies in the plane $\R^2$ and has
angle $\pi/4$ with the chosen direction "to the right". Let us consider the
standard long trefoil as shown in Fig.~\ref{cat}.
\begin{figure}
\begin{center}
\unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(69.00,77.00)
\bezier{132}(29.00,45.00)(28.00,28.00)(19.33,16.50)
\bezier{96}(29.00,45.00)(29.00,60.00)(38.00,59.00)
\bezier{32}(38.00,59.00)(42.00,58.00)(43.25,56.70)
\bezier{212}(46.67,54.33)(69.00,34.00)(56.00,16.00)
\bezier{212}(14.00,18.00)(39.00,1.00)(56.00,16.00)
\bezier{164}(27.00,39.00)(3.00,28.00)(14.00,18.00)
\bezier{156}(30.00,40.00)(42.00,46.00)(53.00,69.00)
\put(53.00,69.00){\vector(1,2){4.00}} \put(41.00,58.20){\circle{1.00}}
\put(10.70,25.00){\circle{1.00}} \put(46.94,63.04){\line(-2,-3){30.00}}
\put(47.06,62.96){\line(-2,-3){30.00}} \put(50.94,60.04){\line(-2,-3){30.00}}
\put(51.04,59.94){\line(-2,-3){30.00}} \put(18.94,12.04){\line(-2,-3){3.00}}
\put(19.06,11.94){\line(-2,-3){3.00}} \put(14.94,15.04){\line(-2,-3){3.00}}
\put(15.04,14.94){\line(-2,-3){3.00}}
\put(7.90,2.70){\framebox(1.00,1.00)[cc]{}}
\bezier{36}(47.06,62.96)(51.06,64.29)(51.06,59.96)
\bezier{36}(46.94,63.04)(50.94,64.37)(50.94,60.04)
\bezier{76}(17.00,14.00)(11.50,6.33)(5.50,0.00)
\bezier{20}(12.50,6.50)(14.00,6.00)(16.00,7.50)
\bezier{16}(16.00,7.50)(16.67,9.33)(15.30,10.90)
\bezier{20}(15.30,10.90)(13.50,11.50)(12.00,10.50)
\bezier{12}(12.00,10.50)(10.67,9.00)(11.00,8.00)
\end{picture}
\caption{} \label{cat}
\end{center}
\end{figure}
Namely, we assume that close to all crossing points the projections of tangent
directions to $\R^2$ are separated from the direction "to the right" or "to the
left": for certainty, let us make the angles between the direction "to the
right" and these projected tangent directions at consecutive 6 points to be
equal to $\pi/4$, $3\pi/10$, $-\pi/4$, $3\pi/4$, $2\pi/10$, and $\pi/4$
respectively.
We call this knot a "large" one, and tie a very small homothetic knot on its
initial segment indicated by a tiny square in Fig. \ref{cat}. Then we shrink
very much the large knot in the "vertical" direction (orthogonal to the plane
of our picture) so that it becomes almost flat and its derivative almost
horizontal, not changing the small knot. Then we move this small knot along the
large one in such a way as if it would be frozen in a small hard bead put on
this large knot. (On the same Fig.~\ref{cat} we show by the thick lines the
channel of the bead; in this case all the picture should be considered as that
of the small homothetic knot. In particular all the points of this knot where
the direction of its derivative is sufficiently far from the standard one, are
inside the bead.)
More precisely, we associate with this bead a orthonormal frame in $\R^3$ whose
first vector in the initial instant is vertical (i.e. orthogonal to the plane
of the picture) and the second vector is directed along the channel.
\begin{lemma}
Suppose that a) the ratio of the diameter of the channel to its length is equal
to a sufficiently small number $\varepsilon$, b) the coefficient of the
flattening of the large knot in the vertical direction is of order
$\varepsilon^2$, so that the absolute value of the "vertical" part of the
derivative of the large knot shown in Fig.~\ref{cat} is nowhere greater than
$\varepsilon^2$ times the length of its "horizontal" part, and c) the size of
the bead $($i.e. the homothety coefficient of two knots$)$ is equal to
$\varepsilon^3$.
Then we can move our bead along the entire large knot in such a way that
A) the first vector of its associated frame remains vertical all the time, and
B) there is a smooth one-parametric family of long knots in $\R^3$ such that at
any instant
i) they coincide with the large knot everywhere outside the convex hull of the
bead,
ii) their intersection with the bead itself remains fixed and is as shown in
Fig.~\ref{cat},
iii) in all the points of the knot inside the channel of the bead, the angle
between the derivative of the knot and the direction of the channel is less
than $\pi/4$. \quad $\square$
\end{lemma}
The first and last instants of this one-parameter family of knots obviously can
be joined by a homotopy not changing the topology of the knot diagram, and we
get a closed loop in the space of knots. Now let us calculate the intersection
number of this loop with the chain described in Theorem \ref{main}.
This loop can intersect the varieties indicated in statements a) and c) of this
Theorem only when triple intersections of the projection occur. This can happen
only if one of crossing points of the smaller knot moving along some branch of
the large knot passes above or below the other its branch: in total 18
suspicious instants. These instants should be counted with multiplicities. In
the case of variety described in statement a) the multiplicity is equal (mod 2)
to the number of other crossing points of the composite knot forming together
with this triple point a configuration satisfying all other conditions of this
statement; for variety described in c) the multiplicity is equal to 0 or 1
depending on the condition on the tangent frame.
It is easy to calculate that the desired configurations for the variety a)
exist only when our small knot passes the first time (i.e. along the lower
branch) the third crossing point of the large knot: moreover, all three
instants when one of crossing points of the small knot pass this point have
multiplicity 1. Therefore the total number of intersections of our path with
variety a) is equal to 3. Similarly, we meet the variety c) only once, when our
small knot (more precisely, its second crossing point) passes the first time
the first crossing point of the large knot. So, the intersection number with
variety c) is equal to 1.
The configurations of type b) can appear by two reasons. First, when the small
knot passes a crossing point of the large one (and namely an undercrossing)
then all its points go under the other branch of the large knot; at some
instant this happens with the point with the distinguished tangent direction.
Again, any such instant should be counted with multiplicities depending on the
order of other crossing points of the composite knot. It is easy to calculate
that only once this multiplicity can be not equal to zero. Namely, when our
small knot undercrosses the third crossing point, then at some instant this
situation appears with multiplicity 2. Further, when our small knot moves and
rotates together with the derivative of the large one, some of tangent lines at
its own crossing points can instantly become directed "to the right". (Namely,
only the tangent line at the undercrossing branch of the first or third
crossing point of the small knot is interesting for us.) There are exactly two
points of the large knot at which it happens: in Fig.~\ref{cat} they are
indicated by small circles. The multiplicity of the "lower" (in this picture)
point is equal to 1, and the multiplicity of the "upper" one is equal to 0.
Finally, the total number of intersection points of our path with the variety
indicated in Theorem \ref{main} is equal to $3+1+2+1=7$, and proposition
\ref{realiz} is completely proved.
\section{Comments on and proof of Theorem \protect{\ref{comain}}}
\label{proof2}
Four statements A, B, C, D of this theorem are discussed in corresponding parts
of this section.
\smallskip
{\bf A.} The variety in ${\mathcal K}_n$ given by the condition $f(0)=f(\pi)$
is a vector subspace of codimension $n.$ It is equal to the boundary of the
variety ${\mathcal A}$ described in statement A of Theorem \ref{comain}. The
map ${\mathcal K}_n \to \R^{n-1},$ sending any curve $f$ to the vector
$f_1(\pi)-f_1(0),$ defines an isomorphism between $\R^{n-1}$ and the normal
bundle of ${\mathcal A}$, in particular induces a coorientation of ${\mathcal
A}$ from any orientation of $\R^{n-1}$. Thus for any integral
$(n-1)$-dimensional cycle in ${\mathcal K}_n \setminus \Sigma$ its intersection
index with ${\mathcal A}$ is well defined and is equal to its linking number
with the subspace $\{f|f(0)=f(\pi)\}$. It follows from calculations in
\cite{arman}, \cite{fasis}, \cite{bjo} that such integral cycles exist only if
$n$ is even. For instance, let $S^{2k-1}$ be the unit sphere, and consider all
the fibers of the Hopf bundle $S^{2k-1} \to \CP^{k-1}$ supplied with natural
parametrizations respecting the natural orientations of these fibers. The set
of all these parametrized fibers is obviously homeomorphic to $S^{2k-1}$ (to
any parametrized fiber there corresponds the zero of the parameter) and has
exactly one intersection point with the variety ${\mathcal A}$.
\smallskip
{\bf B.} The algorithm of finding the spanning chain is as follows. The variety
${\mathcal L}$ described in Proposition \ref{ordone} is swept out by
1-parametric family of subspaces $L(\alpha) \subset {\mathcal K}_n$ of
codimension $n$: they are parametrized by points $\alpha$ of the half-circle
$S^1/\pm = \R^1/ \pi \Z$ and defined by conditions $f(\alpha)=f(\alpha+\pi)$.
Let us try to span all these spaces separately. Consider the trivial bundle
${\mathcal K}_n \times [0,\pi] \to [0,\pi]$ and subset in it consisting of
pairs $(\alpha,f)$ such that $f(\alpha)$ is above $f(\alpha+\pi)$ in $\R^n.$
This subset is a smooth submanifold with boundary, and its projection to
$[0,\pi]$ is a smooth fiber bundle. Forgetting the second coordinate $\alpha$
defines the projection of this manifold to ${\mathcal K}_n$. Its image is
exactly the variety ${\mathcal B}a$ described in statement Ba of Theorem
\ref{comain}. Its boundary consists of the variety ${\mathcal L}$ and images of
fibers of the above-described fiber bundle over the points $0$ and $\pi$. The
union of these two fibers is equal to the subspace distinguished by the
condition $f_1(0)=f_1(\pi)$, and is equal to the boundary of the half-space
${\mathcal B}b$ described in statement Bb of Theorem \ref{comain}.
Now we choose coorientations of these varieties. The variety ${\mathcal B}a$ is
singular. Any its regular point $f$ satisfies the condition
$f_1(\alpha)=f_1(\alpha+\pi)$ for exactly one $\alpha \in [0,\pi)$ and has
transverse self-intersection of the curve $f_1(S^1)$ at this point. Close to
such a point $f$ the coorientation of ${\mathcal B}a$ is defined as follows.
Fix our point $\alpha$ and define the map $({\mathcal K}_n,f) \to TS^{n-2}$
associating to any parametrized curve $g \approx f$ the point of $S^{n-2}$
equal to the direction of the vector $g'_1(\alpha+\pi)-g'_1(\alpha)$, and the
tangent vector at this point in $S^{n-2}$ equal to the projection of the vector
$g(\alpha+\pi)-g(\alpha)$ to the plane orthogonal to this vector
$g'_1(\alpha+\pi)-g'_1(\alpha)$. The preimage of the zero section of $TS^{n-2}$
under this map is tangent in ${\mathcal K}_n$ to the variety ${\mathcal B}a$,
in particular if we have a generic germ of a $(n-2)$-dimensional subvariety
(simplex) in ${\mathcal K}_n$ at the point $g$ then it is transversal to both
varieties and we can induce its desired orientation from (any fixed)
orientation of the bundle $TS^{n-2}$.
The coorientation of the variety ${\mathcal B}b$ is induced from a chosen
orientation of $\R^{n-2}$ by the map ${\mathcal K}_n \to \R^{n-2}$ by a map
sending any $f$ to the direction of the vector $f_2(\pi)-f_2(0) \in \R^{n-2}
\equiv \R^n/\{\uparrow, \mapsto\}$.
\smallskip
{\bf C}. Recall that the term $F_1$ of the simplicial resolution of $\Sigma$ is
the space of pairs
\begin{equation}
\label{f1} ((\alpha,\beta),f) \in \overline{B(S^1,2)} \times {\mathcal K}_n
\end{equation}
such that $f(\alpha)=f(\beta)$. In particular it is a vector bundle over
$\overline{B(S^1,2)}$. Let ${\mathcal M} \subset F_2 \setminus F_1$ be the
principal part of the $(2n-3)$-dimensional class of order 2 described in
Proposition \ref{ordtwo}. Its first differential $d^1(\mathcal M)$ is realized
by the subvariety in $F_1$ consisting of such pairs (\ref{f1}) that
$\beta=\alpha+\pi$ and $f$ satisfies not only the condition
$f(\alpha)=f(\alpha+\pi)$ but also the condition
$f(\alpha+\pi/2)=f(\alpha-\pi/2).$ The set of such pairs $(\alpha,\beta)$ is
the circle $\R^1 /\pi \Z,$ so our cycle $d^1({\mathcal M})$ is the space of a
vector bundle over the circle. To span it in $F_1$ consider the subvariety
${\mathcal M}'_1 \subset F_1$ consisting of such pairs (\ref{f1}) that again
$\beta=\alpha+\pi,$ $f(\alpha)=f(\beta),$ but the image of one of points
$f(\alpha \pm \pi/2)$ is {\em above} the other: namely, the image of those of
these two points which is separated from $0\in S^1$ by the points $\alpha,
\alpha+\pi$ is above the image of its antipode. This subvariety also forms a
fiber bundle over the circle $\R^1/\pi \Z$ of all such pairs
$(\alpha,\alpha+\pi)$. There is exactly one position of $\alpha$ over which
this fiber bundle fails to be locally trivial, namely $\alpha=0 (\mbox{mod }
\pi).$ The boundary of this subvariety is equal to the sum of the cycle
$d^1({\mathcal M})$ and the space of points (\ref{f1}) where $\alpha=
0(\mbox{mod } \pi)$, $f(0)=f(\pi)$ and $f_1(\pi/2)=f_1(-\pi/2)$. We span the
latter space by the similar {\em half}space ${\mathcal M}''_1$, defined by the
condition that $f(0)=f(\pi)$ and $f_1(\pi/2)$ lies to the right of
$f_1(-\pi/2)$. The sum ${\mathcal M}'_1 +{\mathcal M}''_1$ is the desired chain
in $F_1$ whose boundary is equal to $d^1({\mathcal M})$. Now we consider the
image $d_2({\mathcal M})$ of this chain in $\Sigma$ and try to represent it as
a boundary of some relative cycle in ${\mathcal K}_n (\mbox{mod }\Sigma).$ The
image of ${\mathcal M}''_2$ is obvious, and the image of ${\mathcal M}'_1$
consists of maps $f$ such that there exists $\alpha \in [0,\pi]$ such that
$f(\alpha)=f(\alpha+\pi)$, and the image of one of points $\alpha\pm \pi/2$
(namely, the one separated from 0 by $\alpha$ and $\alpha+\pi$) is above the
other.
It is natural to kill this variety by the space of all maps $f$ described in
statement Ca of Theorem \ref{comain}. Its boundary consists of this image of
${\mathcal M}'_1$
and the space of such maps $f$ that $f(\pi)$ is above $f(0)$
and $f_1(\pi/2)=f_1(-\pi/2).$ The boundary of the variety described in
statement Cb of Theorem \ref{comain} is equal to the sum of the latter space
and the image of ${\mathcal M}''_1$.
Statement C of Theorem \ref{comain} is thus proved for $\Z_2$-homology; the
proof of its integer version requires additionally only an accounting of
orientations.
\smallskip
{\bf D}. We shall use the pictures like in \S~\ref{proof1}, only the Wilson
loop will be shown not by a segment but by an oval with marked "zero" point on
its top. This point is referred to as $0$ in subscripts, and all the other
points participating in the definition of cells and their subvarieties are
numbered accordingly to the (counterclockwise) orientation of the Wilson loop.
All the calculus remains the same as in \S \ref{proof1}, only the boundary
operators will include the limit positions of our cells and their subvarieties
when some of defining them points tend to $0$.
As we are interested in integral homology classes, we shall take care of
orientations of all our varieties in the cells of the standard cell
decompositions of terms $F_i \setminus F_{i-1}.$ This orientation consists of
the orientation of the cell and the (co)orientation of the subvariety in it.
The choice of these orientations will follow the guidelines indicated in \S 3
of \cite{V1} or \S V.3.3 of \cite{fasis}. Namely, they consist of the following
orientations (taken in that order): a) the orientation of the simplex
participating in the construction of the simplicial resolution (i.e. the
simplex $\tilde \Delta(J)$ or some its non-marginal face); b) the coorientation
of the subspace $L(J)$ of the space $ {\mathcal K}_n$; c) the orientation of
the space of equivalent point configurations $J \subset S^1$; d) the
(co)orientation of the subvariety in the cell. The first three orientations are
specified exactly as in \S 3 of \cite{V1} (but now in c) we can move only {\em
nonzero} points). Often the subvariety in the cell is given by several
conditions of the form: "there are additional points in $\R^1$ whose images
$f(\cdot) \in \R^n$ (or their projections to some fixed subspace) coincide with
one another or with images of some points participating in the definition of
the cell", or at least our subvariety forms an open subset in a subvariety
defined in such a way. In this case the orientation d) also is defined by the
sequence consisting of $\alpha$) the (co)orientation of the vector subspace
defined by these conditions in the vector spaces counted in the step b) above,
and $\beta$) the orientation of the space of configurations of additional
points. These orientations also are specified as in \cite{V1}, \cite{fasis}; to
define the coorientations of subspaces we assume that the direction "up" in
$\R^n$ is the first vector of the canonical frame, and the direction "to the
right" is the second. All the forthcoming calculations refer to exactly this
choice of orientations.
\medskip
The principal part of the considered class in $F_2\setminus F_1$ is as follows:
\begin{equation}
\label{cprinpar} \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(80.00,14.50)
\put(25.00,9.00){\oval(30.00,10.00)[]}
\put(3.00,9.00){\makebox(0,0)[cc]{$V_2=$}}
\put(45.00,9.00){\makebox(0,0)[cc]{$+$}} \put(65.00,9.00){\oval(30.00,10.00)[]}
\put(65.00,14.00){\circle*{1.00}} \put(25.00,14.00){\circle*{1.00}}
\put(22.00,4.00){\oval(12.00,8.00)[t]} \put(28.00,4.00){\oval(12.00,8.00)[b]}
\put(65.00,4.00){\oval(16.00,8.00)[t]} \put(61.00,4.00){\oval(8.00,4.00)[t]}
\put(69.00,4.00){\oval(8.00,4.00)[t]}
\end{picture}
\ ,
\end{equation}
see \cite{V1}. The second summand has no boundary in $F_1$, the boundary of the
first is as follows:
\begin{equation}
\label{cdone} \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(79.00,15.50)
\put(24.00,10.00){\oval(30.00,10.00)[]} \put(64.00,10.00){\oval(30.00,10.00)[]}
\put(64.00,15.00){\circle*{1.00}} \put(24.00,15.00){\circle*{1.00}}
\put(24.00,5.00){\oval(12.00,8.00)[t]} \put(64.00,5.00){\oval(12.00,8.00)[t]}
\put(64.00,5.00){\line(1,-1){5.00}} \put(69.00,0.00){\line(1,1){5.00}}
\put(24.00,5.00){\line(-1,-1){5.00}} \put(19.00,0.00){\line(-1,1){5.00}}
\put(44.00,10.00){\makebox(0,0)[cc]{$-$}}
\put(3.00,10.00){\makebox(0,0)[cc]{$(-1)^n$}}
\end{picture}
\ .
\end{equation}
Arguing as previously, we try to span the two terms of this chain by the
varieties encoded in left parts of the next two equations, respectively:
\begin{equation}
\label{cspan1} \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(123.00,95.00)
\put(26.00,14.00){\oval(26.00,10.00)[]} \put(26.00,19.00){\circle*{1.00}}
\put(63.00,14.00){\oval(26.00,10.00)[]} \put(63.00,19.00){\circle*{1.00}}
\put(105.00,14.00){\oval(26.00,10.00)[]} \put(105.00,19.00){\circle*{1.00}}
\put(20.00,35.00){\oval(26.00,10.00)[]} \put(20.00,40.00){\circle*{1.00}}
\put(63.00,35.00){\oval(26.00,10.00)[]} \put(63.00,40.00){\circle*{1.00}}
\put(105.00,35.00){\oval(26.00,10.00)[]} \put(105.00,40.00){\circle*{1.00}}
\put(20.00,63.00){\oval(26.00,10.00)[]} \put(20.00,68.00){\circle*{1.00}}
\put(63.00,63.00){\oval(26.00,10.00)[]} \put(63.00,68.00){\circle*{1.00}}
\put(105.00,63.00){\oval(26.00,10.00)[]} \put(105.00,68.00){\circle*{1.00}}
\put(20.00,85.00){\oval(26.00,10.00)[]} \put(20.00,90.00){\circle*{1.00}}
\put(66.00,85.00){\oval(26.00,10.00)[]} \put(66.00,90.00){\circle*{1.00}}
\put(105.00,85.00){\oval(26.00,10.00)[]} \put(105.00,90.00){\circle*{1.00}}
\put(3.00,85.00){\makebox(0,0)[cc]{$\partial$}}
\put(44.00,85.00){\makebox(0,0)[cc]{$=(-1)^{n-1}$}}
\put(85.00,85.00){\makebox(0,0)[cc]{$+$}}
\put(123.00,85.00){\makebox(0,0)[cc]{$+$}}
\put(41.00,63.00){\makebox(0,0)[cc]{$-$}}
\put(84.00,63.00){\makebox(0,0)[cc]{$-$}}
\put(3.00,35.00){\makebox(0,0)[cc]{$\partial$}}
\put(41.00,35.00){\makebox(0,0)[cc]{$=(-1)^n$}}
\put(84.00,35.00){\makebox(0,0)[cc]{$-(-1)^n$}}
\put(123.00,35.00){\makebox(0,0)[cc]{$+$}}
\put(44.00,14.00){\makebox(0,0)[cc]{$-$}}
\put(84.00,14.00){\makebox(0,0)[cc]{$-(-1)^n$}}
\put(22.00,80.00){\oval(10.00,8.00)[t]} \put(22.00,80.00){\line(-1,-1){5.00}}
\put(17.00,75.00){\vector(-1,1){5.00}} \put(68.00,80.00){\oval(10.00,8.00)[t]}
\put(68.00,80.00){\line(-1,-1){5.00}} \put(63.00,75.00){\line(-1,1){5.00}}
\put(105.00,80.00){\line(0,-1){5.00}} \put(105.00,75.00){\line(-1,0){14.00}}
\put(91.00,75.00){\line(0,1){20.00}} \put(91.00,95.00){\line(1,0){14.00}}
\put(105.00,95.00){\vector(0,-1){4.50}} \put(105.00,58.00){\line(0,1){10.00}}
\put(110.00,58.00){\line(-1,-1){5.00}} \put(105.00,53.00){\vector(-1,1){5.00}}
\put(63.00,58.00){\oval(14.00,8.00)[t]} \put(66.00,58.00){\line(-1,-1){5.00}}
\put(61.00,53.00){\vector(-1,1){5.00}} \put(20.00,58.00){\oval(14.00,8.00)[t]}
\put(16.00,52.00){\vector(4,1){8.00}} \put(16.00,52.00){\vector(4,-1){8.00}}
\put(20.00,48.00){\makebox(0,0)[cc]{$2$}}
\put(20.00,56.00){\makebox(0,0)[cc]{$1$}}
\put(18.00,30.00){\oval(10.00,8.00)[t]} \put(18.00,30.00){\line(1,-1){5.00}}
\put(23.00,25.00){\vector(1,1){5.00}} \put(61.00,30.00){\oval(10.00,8.00)[t]}
\put(61.00,30.00){\line(1,-1){5.00}} \put(66.00,25.00){\line(1,1){5.00}}
\put(105.00,30.00){\line(0,-1){5.00}} \put(105.00,25.00){\line(1,0){14.00}}
\put(119.00,25.00){\line(0,1){20.00}} \put(119.00,45.00){\line(-1,0){14.00}}
\put(105.00,45.00){\vector(0,-1){4.50}}
\put(105.00,30.00){\oval(14.00,8.00)[t]}
\put(105.00,80.00){\oval(14.00,8.00)[t]} \put(105.00,9.00){\line(0,1){10.00}}
\put(100.00,9.00){\line(1,-1){5.00}} \put(105.00,4.00){\vector(1,1){5.00}}
\put(63.00,9.00){\oval(14.00,8.00)[t]} \put(60.00,9.00){\line(1,-1){5.00}}
\put(65.00,4.00){\vector(1,1){5.00}} \put(26.00,9.00){\oval(14.00,8.00)[t]}
\put(22.00,3.00){\vector(4,1){8.00}} \put(22.00,3.00){\vector(4,-1){8.00}}
\put(26.00,0.00){\makebox(0,0)[cc]{$2$}}
\put(26.00,7.00){\makebox(0,0)[cc]{$1$}}
\put(3.00,63.00){\makebox(0,0)[cc]{$+$}}
\put(6.00,14.00){\makebox(0,0)[cc]{$+(-1)^n$}}
\end{picture}
\end{equation}
The varieties shown by the second from the end pictures in both these equations
coincide geometrically, and their canonical orientations differ by the factor
$(-1)^{n-1}$. Therefore the linear combination of left parts of these equations
taken with coefficients $-1$ and $(-1)^{n-1}$ respectively is equal in $F_1$ to
the sum of the expression (\ref{cdone}) and two last varieties in these
equations. If $n=3$ then the sum of last two varieties is equal to zero.
Indeed, any of these varieties consists of pairs (\ref{f1}) with $\alpha=0,$
$f(\alpha)=f(\beta),$ taken with some multiplicities. These multiplicities
always are opposite, because they are equal (up to signs) to different
combinatorial expressions for the linking numbers of two ``smoothened'' loops
into which the point $f(0)=f(\beta)$ breaks the curve $f(S^1).$
However, if $n>3$ then the sum of these two varieties is only homologous to
zero, but not equal to it. We shall encode this sum by the picture in the
right-hand part of the following equation:
\begin{equation}
\label{cspan2} \unitlength 1.00mm \linethickness{0.4pt}
\begin{picture}(91.00,10.50)
\put(3.00,5.00){\makebox(0,0)[cc]{$\partial$}}
\put(24.00,5.00){\oval(32.00,10.00)[]} \put(24.00,0.00){\line(0,1){10.00}}
\put(24.00,10.00){\circle*{1.00}} \put(6.00,2.00){\line(0,1){6.00}}
\put(6.00,5.00){\vector(1,0){34.00}} \put(50.00,5.00){\makebox(0,0)[cc]{$=
(-1)^n$}} \put(75.00,5.00){\oval(32.00,10.00)[]}
\put(75.00,10.00){\circle*{1.00}} \put(75.00,10.00){\line(0,-1){10.00}}
\put(75.00,5.00){\vector(1,0){16.00}} \put(75.00,5.00){\vector(-1,0){16.00}}
\end{picture}
\ .
\end{equation}
The variety assumed in the left part of this equation consists of all pairs
(\ref{f1}) in $F_1$ such that $\alpha=0$ and additionally there are points
$\gamma \in (0,\beta)$ and $\delta \in (\beta,2\pi)$ such that the projection
of $f(\delta)$ to $\R^{n-1}$ lies "to the right" from that of $f(\beta)$.
So, the desired chain in $F_1$ spanning the cycle (\ref{cdone}) is equal to
\begin{equation}
\label{cspanlst} - \left( \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(108.00,8.50)
\put(12.00,3.00){\oval(24.00,10.00)[]}
\put(32.00,3.00){\makebox(0,0)[cc]{$+(-1)^n$}}
\put(52.00,3.00){\oval(24.00,10.00)[]}
\put(74.00,3.00){\makebox(0,0)[cc]{+$(-1)^n$}}
\put(96.00,3.00){\oval(24.00,10.00)[]} \put(82.00,0.00){\line(0,1){6.00}}
\put(82.00,3.00){\vector(1,0){26.00}} \put(50.00,-2.00){\oval(10.00,8.00)[t]}
\put(50.00,-2.00){\line(1,-1){5.00}} \put(55.00,-7.00){\vector(1,1){5.00}}
\put(52.00,8.00){\circle*{1.00}} \put(96.00,8.00){\circle*{1.00}}
\put(96.00,8.00){\line(0,-1){10.00}} \put(12.00,8.00){\circle*{1.00}}
\put(14.00,-2.00){\oval(10.00,8.00)[t]} \put(15.00,-2.00){\line(-1,-1){5.00}}
\put(10.00,-7.00){\vector(-1,1){5.00}}
\end{picture} \right) .
\end{equation}
Its image in $\Sigma$ is expressed by the formula
\begin{equation}
\label{cd-2} -\left( \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(101.00,8.50)
\put(12.00,3.00){\oval(24.00,10.00)[]} \put(34.00,3.00){\makebox(0,0)[cc]{$+
(-1)^n$}} \put(53.00,3.00){\oval(24.00,10.00)[]}
\put(70.00,3.00){\makebox(0,0)[cc]{$+$}} \put(89.00,3.00){\oval(24.00,10.00)[]}
\put(75.00,0.00){\line(0,1){6.00}} \put(75.00,3.00){\vector(1,0){26.00}}
\put(51.00,-2.00){\line(1,-1){5.00}} \put(56.00,-7.00){\vector(1,1){5.00}}
\put(53.00,8.00){\circle*{1.00}} \put(89.00,8.00){\circle*{1.00}}
\put(12.00,8.00){\circle*{1.00}} \put(15.00,-2.00){\line(-1,-1){5.00}}
\put(10.00,-7.00){\vector(-1,1){5.00}} \put(10.00,-2.00){\line(1,1){5.00}}
\put(15.00,3.00){\line(1,-1){5.00}} \put(56.00,-2.00){\line(-1,1){5.00}}
\put(51.00,3.00){\line(-1,-1){5.00}} \put(89.00,-2.00){\line(1,1){4.00}}
\put(93.00,2.00){\line(-2,3){4.00}}
\end{picture}
\right) .
\end{equation}
We need to span this chain by a relative cycle in ${\mathcal K}_n$ (mod
$\Sigma$). For this we have
\begin{equation}
\label{ppvv} \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(124.00,40.50)
\put(2.00,35.00){\makebox(0,0)[cc]{$\partial$}}
\put(15.00,35.00){\oval(22.00,10.00)[]} \put(15.00,40.00){\circle*{1.00}}
\put(57.00,35.00){\oval(22.00,10.00)[]} \put(57.00,40.00){\circle*{1.00}}
\put(85.00,35.00){\oval(22.00,10.00)[]} \put(85.00,40.00){\circle*{1.00}}
\put(111.00,35.00){\oval(20.00,10.00)[]} \put(111.00,40.00){\circle*{1.00}}
\put(23.00,10.00){\oval(24.00,10.00)[]} \put(23.00,15.00){\circle*{1.00}}
\put(53.00,10.00){\oval(24.00,10.00)[]} \put(53.00,15.00){\circle*{1.00}}
\put(83.00,10.00){\oval(24.00,10.00)[]} \put(83.00,15.00){\circle*{1.00}}
\put(112.00,10.00){\oval(20.00,10.00)[]} \put(112.00,15.00){\circle*{1.00}}
\put(18.00,30.00){\line(-1,1){5.00}} \put(13.00,35.00){\vector(-1,-1){5.00}}
\put(12.00,30.00){\line(1,-1){5.00}} \put(17.00,25.00){\vector(1,1){5.00}}
\put(36.00,35.00){\makebox(0,0)[cc]{$=(-1)^{n-1}$}}
\put(64.00,30.00){\line(-1,-1){5.00}} \put(59.00,25.00){\line(-1,1){5.00}}
\put(61.00,30.00){\line(-1,1){5.00}} \put(56.00,35.00){\vector(-1,-1){5.00}}
\put(78.00,30.00){\line(1,1){5.00}} \put(83.00,35.00){\line(1,-1){5.00}}
\put(81.00,30.00){\line(1,-1){5.00}} \put(86.00,25.00){\vector(1,1){5.00}}
\put(71.00,35.00){\makebox(0,0)[cc]{$-$}}
\put(99.00,35.00){\makebox(0,0)[cc]{$-$}}
\put(124.00,35.00){\makebox(0,0)[cc]{$+$}}
\put(106.00,30.00){\line(1,-1){5.00}} \put(111.00,25.00){\vector(1,1){5.00}}
\put(111.00,30.00){\vector(0,1){9.50}} \put(112.00,5.00){\vector(0,1){9.50}}
\put(117.00,5.00){\line(-1,-1){5.00}} \put(112.00,0.00){\vector(-1,1){5.00}}
\put(81.00,5.00){\line(1,-1){5.00}} \put(86.00,0.00){\vector(1,1){5.00}}
\put(90.50,5.00){\line(-3,2){7.50}} \put(83.00,10.00){\vector(-3,-2){7.50}}
\put(53.00,5.00){\line(3,-4){3.75}} \put(56.75,0.00){\vector(3,4){3.75}}
\put(53.00,5.00){\line(-3,4){3.75}} \put(49.25,10.00){\vector(-3,-4){3.75}}
\put(25.00,5.00){\line(-1,1){5.00}} \put(20.00,10.00){\vector(-1,-1){5.00}}
\put(15.00,5.00){\line(3,-2){7.50}} \put(22.50,0.00){\vector(3,2){7.50}}
\put(38.00,10.00){\makebox(0,0)[cc]{$+$}}
\put(8.00,10.00){\makebox(0,0)[cc]{$+$}}
\put(68.00,10.00){\makebox(0,0)[cc]{$+$}}
\put(98.00,10.00){\makebox(0,0)[cc]{$-$}}
\end{picture} ,
\end{equation}
\begin{equation}
\label{cnev1} \unitlength 1.00mm \special{em:linewidth 0.4pt}
\linethickness{0.4pt}
\begin{picture}(119.00,10.50)
\put(1.00,5.00){\makebox(0,0)[cc]{$\partial$}}
\put(19.00,5.00){\oval(26.00,10.00)[]} \put(4.00,8.00){\line(0,-1){6.00}}
\put(4.00,5.00){\vector(1,0){28.00}} \put(19.00,10.00){\circle*{1.00}}
\put(19.00,0.00){\vector(0,1){9.50}} \put(38.00,5.00){\makebox(0,0)[cc]{$=-$}}
\put(44.00,2.00){\line(0,1){6.00}} \put(44.00,5.00){\vector(1,0){28.00}}
\put(59.00,5.00){\oval(26.00,10.00)[]} \put(59.00,10.00){\circle*{1.00}}
\put(59.00,0.00){\line(1,1){4.00}} \put(63.00,4.00){\line(-2,3){4.00}}
\put(79.00,5.00){\makebox(0,0)[cc]{$+$}}
\put(87.50,5.00){\makebox(0,0)[cc]{$(-1)^n$}}
\put(106.00,5.00){\oval(26.00,10.00)[]} \put(106.00,10.00){\circle*{1.00}}
\put(106.00,5.00){\vector(1,0){13.00}} \put(106.00,5.00){\vector(-1,0){13.00}}
\put(106.00,0.00){\vector(0,1){9.50}}
\end{picture}
.
\end{equation}
The sum of the fourth, fifth and sixth terms in the right-hand part of
(\ref{ppvv}) is equal to zero. The sum of varieties encoded by the pictures in
the third and seventh terms is equal to the variety shown by the last picture
in (\ref{cnev1}). Therefore the desired relative cycle is equal to the linear
combination of left parts of (\ref{ppvv}) and (\ref{cnev1}) taken with
coefficients $1$ and $(-1)^n$ respectively. Theorem \ref{comain} is completely
proved.
|
\section{Introduction}
Despite not having a complete theory of quantum gravity, it is becoming more and more important to understand systems for which both quantum and general relativistic effects are important (see \cite{RuoBerchera13}-\cite{Lamine06}, for instance, and references therein). Indeed, it is the study of such systems that helps clarify the clash between quantum theory (QT) and general relativity (GR), in hope that we may find guidance towards a resolution of the many technical and conceptual problems one faces when attempting to unify these two pillars of physics.
The purpose of this work is to explore an ambiguity the results from taking both QT and GR seriously: quantum superpositions of different matter states are associated with different spacetime geometries, and hence different definitions of time evolution; how, then, can we use a single time-evolution operator to evolve superpositions of distinct spacetimes? Since the $1980$s, Roger Penrose has been arguing that this ambiguity results in an instability, and that in turn this instability leads to a type of ``decay'' that reduces the system to a state with a single well-defined geometry \cite{Penrose86}, \cite{Penrose}. It is not clear from Penrose's work, however, whether some sort of ``collapse'' occurs, or whether there is simply a form of ``intrinsic'' decoherence that removes phase correlations between states associated with sufficiently different geometries. In this paper, we consider the latter, and discuss whether or not a direct application of both QT and GR is enough to demonstrate the existence of this new type of ``intrinsic'' decoherence.
By ``intrinsic'' decoherence, we mean a decoherence effect that arises solely out of the internal behaviour of an isolated system, and not due to its interaction with the external world. For example, if we use a buckyball in a double-slit experiment, and prepare one of the slits to excite internal degrees of freedom of the buckyball, then the internal degrees of freedom carry ``which-way'' information and decohere the center-of-mass degree of freedom \cite{Banaszek13}. More generally, if a system carries an internal clock and is in a superposition of states corresponding to two paths that have different proper times associated with them, then again the internal clock read at the interference screen could provide which-way information, and decohere the center-of-mass \cite{PZCB13}-\cite{ZCPRB12}.
Whereas the decoherence produced by entangling internal degrees of freedom to a center-of-mass coordinate could be considered ``third-party'' decoherence \cite{Stamp06}, what we are concerned with here is whether or not there is something about gravity itself that could lead to such intrinsic decoherence. Penrose's intuition says yes: the path a mass takes alters the associated spacetime and especially the flow of time. Since the quantum phase is determined by the flow of time, the phase evolution is also altered by which path the mass takes. When one tries to interfere the two paths, these ``random'' phases (because it is impossible to uniquely map one spacetime onto another) cause decoherence. It is this \textit{gravitational} intrinsic decoherence that we explore here.
There are several proposals in the literature for a mechanism to describe gravitational intrinsic decoherence \cite{Diosi}-\cite{GambiniPullin06}, and other proposals for intrinsic decoherence mechanisms that are merely inspired by tension between QT and GR \cite{Milburn91}-\cite{Percival95}. Many of these approaches incorporate alterations to known physics, such as adding stochastic \cite{Percival95} or nonlinear \cite{Weinberg89I}-\cite{Penrose98} terms to the Schr\"{o}dinger equation, in order to achieve the desired decoherence effect. Such alterations are often ad hoc, and have historically faced difficulties maintaining consistency with experimental constraints; nonlinear additions to the Schr\"{o}dinger equation, for instance, have been shown under a wide range of conditions to lead to either superluminal signal propagation or to communication between different branches of the wavefunction \cite{Polchinsky91}. While it may be possible to obtain a sensible theory that allows communication between different wavefunction branches \cite{Stamp}, it remains to be seen whether a consistent interpretation results from this alteration. Instead, we take Penrose's initial arguments at face value, and entertain the possibility that a consistent combination of QT and GR can explain (gravitational) intrinsic decoherence without any assumptions about new physics.
Now, it is well-known that gravitational waves can carry away information from a system in a manner analogous to standard decoherence \cite{Wang06}, \cite{Hu14}. Penrose's suggestion is independent of such standard gravitational-wave-induced decoherence, so to distinguish between the two we will work in spherical symmetry. The restriction to spherical symmetry is not only a technical simplification, but avoids the occurrence of gravitational waves altogether.
Because this exploration requires both QT and GR, we will naturally be faced with some serious difficulties, which we will have to either overcome, sidestep, or ignore \cite{DeWitt67}. For instance, we will avoid issues with the factor-ordering ambiguity by working in the WKB regime (as in \cite{Visser}), and we will avoid issues with perturbative non-renormalizability by working in minisuperspace (i.e. enforcing spherical symmetry) and employing a reduced phase space approximation (as in \cite{Kuchar71}). It is still unclear what the exact connection is between the reduced phase space approximation, obtained by solving the GR constraints classically and then quantizing the reduced theory, and the standard Dirac quantization, obtained by quantizing the theory in the full kinematical Hilbert space and then enforcing the constraints at the quantum level. Following Hawking's path integral approach to quantum gravity \cite{Hawking78}, Halliwell has made some progress elucidating the connection between reduced phase space minisuperspace quantization and the standard Dirac approach in special cases \cite{Halliwell88}, but in general the connection is not well understood. Nonetheless, the limit we will work in has a rich structure, and in this paper we will explore whether or not it has a rich enough structure to contain evidence of intrinsic decoherence caused by gravity.
Since we aim to test whether or not gravity places a fundamental limit on the coherence of quantum systems, we develop a model of a self-gravitating interferometer. Interferometers are ideal for studying coherence, because interference is a key feature of coherent systems. We describe how the same interferometer would behave in the absence of gravity, and then we investigate the consequences of general relativistic corrections to this behaviour. In an interferometric setting, the intrinsic decoherence we seek to understand manifests itself as a phase-scrambling along different interferometer arms (for a general discussion see \cite{Aharonov90}), which in this case is attributed to gravity. According to Penrose, we should expect that the (interferometric) coherence should decay as the arm-length increases indefinitely, since this would correspond to a superposition of arbitrarily different spacetimes. We focus on the possibility that no collapse occurs, so we will simply analyze the interference pattern and search for departures from non-gravitational behaviour that indicate coherence loss. Conceptually, we are testing the idea that when one forms superpositions of geometries in the interferometer, the nature of time in GR leads at the quantum level to an imprint of which-way information, which is accompanied by a loss of fringe visibility \cite{Englert96}.
Still, an objection may be raised that if one describes the interferometer as a closed quantum system without tracing out over any physical degrees of freedom, then QT implies that coherence must prevail, regardless of whether the system is general relativistic. This objection was raised by Banks, Susskind, and Peskin \cite{BanksSusskind84} in the context of black hole evaporation, but it was later pointed out not only that the arguments in \cite{BanksSusskind84} were inconclusive, but that we have reason to support the possibility that pure states can effectively evolve into mixed states in black hole systems \cite{UnruhWald95}.
The more radical idea entertained here is that one might find pure states evolving to mixed states in gravitational systems without horizons. In general, this ``dissipationless'' type of decoherence has been explored to some degree \cite{Stamp88}-\cite{Gangopadhyay01}, but even the fact that it is possible has not been widely appreciated. Nonetheless, one can observe that the thermal character of acceleration radiation is approximately present even without the involvement of Rindler horizons (for recent analyses see \cite{Rovelli12}, \cite{Langlois}), and by the equivalence principle one might expect to find a gravitational analog of this thermal behaviour. This means, then, that one might expect that gravity generates an intrinsic form of entropy, even in systems without the horizon structure that one usually associates with entropy in black hole thermodynamics.
With this in mind, we will construct our interferometer, theoretically, out of a self-gravitating, spherically symmetric, infinitesimally thin shell of matter. The interferometer ``optics'' are encoded internally, by adding tangential pressure to the fluid that lives on the surface of the shell. The resulting model is reminiscent of an idea Einstein first proposed in $1939$ \cite{EinsteinCluster}, but in our case, the tangential pressure satisfies an equation of state that produces a beam-splitter and perfect reflectors. The fluid is ideal, in the sense that one obtains a perfect-fluid stress-energy tensor, if one projects the full four-dimensional spacetime stress-energy tensor onto the three-dimensional history of the shell. This approach ensures that the interferometric setup is manifestly invariant under coordinate transformations.
The configuration we construct resembles that of a Michelson interferometer in optics. Thus, we will send initial states at a beam-splitter, at which point the transmitted and reflected components travel in opposite directions until they encounter ``mirrors.'' The components will then reflect, travel back towards each other, and encounter the splitter once more. There will be two possible outputs, corresponding to final transmission and final reflection, which are comprised of different combinations of the initially split wave components.
What we mean by (interferometric) coherence, in this system, is the sustained phase relationships between different wave components that can allow us, for instance, to completely cancel either of the final outputs. In other words, if we are unable to obtain complete constructive or destructive interference in our interferometer (as predicted by Penrose), we can conclude that coherence is being limited in the system. The goal of our current investigation is to determine whether or not general relativistic effects could demonstrably produce such a limitation.
\section{Theory of Self-Gravitating Spherical Shells}
\subsection{Action Principle}
The perfect-fluid shell model we develop is a generalization of the dust shell model used by Kraus and Wilczek in their attempt to calculate self-interaction corrections to standard Hawking radiation \cite{KrausWilczek}-\cite{KrausCharged}. Generalizing the Kraus and Wilczek approach to include the required pressure effects is not without complications, even in the classical theory. In contrast to the approach to thin-shells pioneered by Israel that involves stitching two spacetimes together along the shell's history \cite{Israel}, the starting point for our theoretical considerations is an action that is composed of a gravitational part given by the Einstein-Hilbert action, plus some action for the shell that we can initially leave unspecified, written (in natural units) as
\begin{equation}
I = \frac{1}{16\pi}\int{d^4x\,\sqrt{-g^{(4)}} \hspace{3pt} \mathcal{R}^{(4)}} + I_{shell}.
\end{equation}
The superscripts on the metric determinant $g$ and the Ricci scalar $\mathcal{R}$ indicate that these quantities are constructed from components of the full spacetime metric $g_{\mu\nu}$, with $\mu,\nu \in \{0,1,2,3\}$.
We will express the metric in ADM form \cite{ADM}, which in spherical symmetry is given by
\begin{equation}
g_{\mu\nu}dx^{\mu}dx^{\nu}=-N^2 dt^2 + L^2\left(dr+N^r dt\right)^2+R^2d\Omega^2,
\label{eq:Metric}
\end{equation}
where $N$ is the lapse function, $N^r$ is the radial component of the shift vector, and $L^2$ and $R^2$ are the only nontrivial components of the spatial metric.
The angular variables are taken to be the polar angle $\theta$ and the azimuthal angle $\phi$, such that the angular metric takes the form $d\Omega^2=d\theta^2+\sin^2{\theta} d\phi^2$.
The shell action studied by Kraus and Wilczek takes the form
\begin{equation}
I_{dust}=-m \int{d\lambda\, \sqrt{-g_{\mu\nu}\frac{d x^\mu}{d\lambda}\frac{d x^\nu}{d\lambda}}},
\end{equation}
with $m$ being the rest-mass of the shell and all metric quantities evaluated on the shell history. The arbitrary parameter $\lambda$ can be chosen to coincide with the coordinate time $t$, which simplifies the integrand for the shell action.
To describe a more general fluid than dust, we need a more general action. There are well-established variational principles for regular perfect fluids in GR \cite{Schutz}, but the authors are unaware of any satisfactory variational principles for the perfect fluid shells we wish to describe. The stress-energy tensor for a perfect fluid with density $\sigma$ and pressure $p$ is given by
\begin{equation}
S^{ab}=\sigma u^a u^b + p ( \gamma^{ab}+u^a u^b ),
\label{eq:FluidTensor}
\end{equation}
where $u^a$ are the components of the fluid proper velocity in coordinates that cover the fluid history. For our purposes, the geometry along the fluid history of our shell is described by an induced metric $\gamma_{ab}dy^a dy^b = -d\tau^2+\hat{R}^2d\Omega^2$, with $\tau$ being the shell proper time. This induced metric obeys the relation
\begin{equation}
\gamma_{ab} = e_a^{\mu}e_b^{\nu}g_{\mu\nu},
\end{equation}
with the introduction of projectors onto the shell history given by
\begin{equation}
e_a^{\mu}=\frac{\partial x^{\mu} }{\partial y^a}=u^\mu \delta_a^\tau + \delta_\Omega^\mu \delta_a^\Omega.
\end{equation}
Here and elsewhere, the repeated $\Omega$ denotes a sum over angular coordinates. These projectors allow us to express the full spacetime stress-energy tensor of our perfect fluid shell as
\begin{equation}
T^{\mu\nu}=S^{ab}e_a^\mu e_b^\nu \delta (\chi),
\label{eq:FullTensor}
\end{equation}
where we have introduced a Gaussian normal coordinate $\chi$ in the direction of the outward-pointing space-like unit normal $\bm{\xi}$, with the shell location defined by $\chi=0$.
We want to obtain an action, expressed in terms of the full spacetime quantities, that yields the tensor (\ref{eq:FluidTensor}) in the intrinsic coordinates of the shell history. To convert derivative expressions from the intrinsic coordinates to the ADM coordinates given in equation (\ref{eq:Metric}), we can write infinitesimal changes in $r$ and $t$ as
\begin{equation}
dt = u^t d\tau + \xi^t d\chi, \hspace{5pt}
dr = u^r d\tau + \xi^r d\chi.
\label{eq:CoordDiffs}
\end{equation}
Taking advantage of the fact that $\bm{\xi}$ satisfies $u^\mu \xi_\mu=0$ and $\xi^\mu \xi_\mu = 1$, and suppressing the (vanishing) angular components for brevity, the outward normal can be written as
\begin{equation}
\xi_\alpha=\sqrt{g_{tr}^2-g_{tt}g_{rr}}\left(-u^r,u^t\right)=N^2 L^2\left(-u^r,u^t\right).
\end{equation}
For radial integration within an ADM slice, one has $dt=0$, and in this case we can solve for $\frac{dr}{d\chi}$ in equation (\ref{eq:CoordDiffs}) to obtain \cite{DeltaConversion}
\begin{equation}
\frac{dr}{d\chi}=\xi^r-\frac{u^r}{u^t}\xi^t.
\end{equation}
Also, since $u^\mu=(\partial t / \partial \tau)(1,\dot{X},0,0)$, the $4$-velocity normalization $u^\mu u_\mu=-1$ (evaluated on the shell) implies
\begin{equation}
\left(\frac{\partial t}{\partial \tau}\right)^2=(u^t)^2=\left(N^2-L^2(N^r+\dot{X})^2\right)^{-1}.
\label{eq:Normalization}
\end{equation}
This allows conversion of the delta function appearing in our expression (\ref{eq:FullTensor}) for the full spacetime stress-energy tensor:
\begin{equation}
\delta (\chi) = \frac{dr}{d\chi}\delta (r-X)=\frac{\sqrt{N^2-L^2(N^r+\dot{X})^2}}{N L}\delta(r-X).
\label{eq:Delta}
\end{equation}
Using equation (\ref{eq:FullTensor}), we find that our stress-energy tensor takes the form
\begin{equation}
T^{\mu\nu}=\left(\sigma u^\mu u^\nu + p \hspace{1pt} g^{\Omega\Omega}\delta_\Omega^\mu \delta_\Omega^\nu \right) \delta(\chi),
\label{eq:FullFluidTensor}
\end{equation}
where the repeated $\Omega$ indices denote a single sum over angular coordinates. In expression (\ref{eq:FullFluidTensor}), the ``tangential'' nature of the pressure is manifest, since the projection of this tensor onto the space-like normal $\bm{\xi}$ clearly vanishes.
The action we seek, then, yields (\ref{eq:FullFluidTensor}) upon taking variations with respect to the metric, in accordance with the definition
\begin{equation}
\delta I = \frac{1}{2}\int{d^4x\,\sqrt{-^4g}T^{\mu\nu}\delta g_{\mu\nu}}.
\label{eq:StressDefinition}
\end{equation}
We are especially interested in the contribution from the tangential pressure, which takes the form
\begin{equation}
\delta I_p = 8\pi\int{dt\,dr\,N L \delta\left(\chi\right)p R \delta R}.
\end{equation}
By inspection, we find that the action
\begin{equation}
I_{shell}=- \int{d\lambda \,\sqrt{-g_{\mu\nu}\frac{d x^\mu}{d\lambda}\frac{d x^\nu}{d\lambda}}} M(R),
\label{eq:ShellAction}
\end{equation}
with all quantities evaluated on the shell history, yields the appropriate stress-energy tensor: the relevant variational derivative of (\ref{eq:ShellAction}) with respect to the metric is
\begin{equation}
\delta I_{shell,p} = -\int{dt\,dr\,N L \delta\left(\chi\right)M'(R) \delta R},
\end{equation}
from which it follows that one has the pressure identification $p=-M'(R)/8\pi R$, along with the usual density identification $\sigma=M(R)/4\pi R^2$. We will use the freedom in choosing the function $M(R)$ to parametrize an equation of state that relates the density and pressure of our fluid. It should be noted that $R$ is not a coordinate, but a metric component that serves as a measure of the shell's internal energy.
The action (\ref{eq:ShellAction}) is reparametrization-invariant, as well as invariant under general (spherically symmetric) coordinate transformations, even with the inclusion of an $R$-dependent `mass'. As mentioned above, this is because $R$, when evaluated on the shell, is nothing more than the reduced area of the shell, and this area is independent of coordinate choices.
\subsection{Hamiltonianization}
Following the canonical formalism \cite{ADM}, one can perform a Legendre transformation $\mathcal{H}=P\dot{X}-\mathcal{L}$, for the shell variables. Here $\mathcal{L}$ is the Lagrangian defined by (\ref{eq:ShellAction}), subject to the condition that the shell history is parametrized by $t$. One then finds
\begin{equation}
\mathcal{L}=-\int{dr\,\sqrt{N^2-L^2(N^r+\dot{X})^2}M(R)\delta(r-X)},
\label{eq:ShellL}
\end{equation}
and it follows that the momentum conjugate to the shell position $X$ for the unreduced problem is given by
\begin{equation}
P=\frac{\partial \mathcal{L}}{\partial \dot{X}}=\int{dr\,\frac{L^2 (N^r+\dot{X}) M(R)}{\sqrt{N^2-L^2(N^r+\dot{X})^2}}\delta(r-X)}.
\label{eq:ShellMomentum}
\end{equation}
Explicitly, we can determine the Hamiltonian $\mathcal{H}$ to be
\begin{equation}
\mathcal{H}=P\dot{X}-\mathcal{L}=\int{dr \,\left(NH_0^s+ N^rH_r^s\right)},
\end{equation}
with the definitions
\begin{eqnarray}
H_0^s &=& \sqrt{L^{-2}P^2+M(R)^2}\delta(r-X), \nonumber \\
H_r^s &=& -P\delta(r-X).
\end{eqnarray}
Similarly, we can Hamiltonianize the gravitational action, and express the total action as
\begin{equation}
I = \int{dt\, P\dot{X}}+\int{dt\,dr\,\left(\pi_R \dot{R}+\pi_L \dot{L}-NH_0-N^r H_r\right)},
\end{equation}
for $H_0=H_0^s+H_0^G$ and $H_r=H_r^s+H_r^G$, such that
\begin{eqnarray}
H_0^G &=& \frac{L\pi_L^2}{2R^2}-\frac{\pi_L\pi_R}{R}+\left(\frac{R R'}{L}\right)'-\frac{(R')^2}{2L}-\frac{L}{2}, \nonumber\\
H_r^G &=& R'\pi_R-L\pi_L'.
\label{eq:GravConstraints}
\end{eqnarray}
\subsection{Equations of Motion}
Once in Hamiltonian form, the equations of motion for the system are obtained by varying the action with respect to the variables $N$, $N^r$, $\pi_L$, $\pi_R$, $L$, and $R$. Explicitly, these variations (respectively) lead to
\begin{eqnarray}
\label{eq:GravEOM}
H_0 &=& 0, \nonumber\\
H_r &=& 0, \nonumber\\
\dot{L} &=& \frac{N}{R}\left(\frac{L\pi_L}{R}-\pi_R\right)+\left(N^r L\right)', \nonumber\\
\dot{R} &=& -\frac{N\pi_L}{R}+N^r R', \\
\dot{\pi_L} &=& \frac{N}{2}\left(1-\frac{\pi_L^2}{R^2}-\frac{(R')^2}{L^2}\right)-\frac{N' R R'}{L^2} \nonumber\\
&&+ N^r \pi_L'+\frac{N P^2 \delta(r-X)}{L^2\sqrt{P^2+L^2 M^2}}, \nonumber\\
\dot{\pi_R} &=& \frac{N\pi_L}{R^2}\left(\frac{L\pi_L}{R}-\pi_R\right)-N\left(\frac{R'}{L}\right)'-\left(\frac{N'R}{L}\right)' \nonumber\\
&&+ \left(N^r \pi_R\right)' - \frac{NM\frac{dM}{dR}\delta(r-X)}{\sqrt{L^{-2} P^2+M^2}}. \nonumber
\end{eqnarray}
The first two equations are the Hamiltonian and momentum constraints, whereas the next four are the dynamical equations of motion for the gravitational variables.
For the shell variables, the equation of motion for $X$ can be easily obtained by varying the action with respect to $P$, or simply by solving equation (\ref{eq:ShellMomentum}) for $\dot{X}$. The result is
\begin{eqnarray}
\dot{X} &=& \int{dr\,\left(\frac{NP}{L\sqrt{P^2+L^2 M^2}}-N^r\right)\delta(r-X)} \nonumber\\
&=& \frac{\hat{N}P}{\hat{L}\sqrt{P^2+\hat{L}^2 \hat{M}^2}}-\hat{N^r},
\label{eq:XEOM}
\end{eqnarray}
with hats indicating that one evaluates the quantities at $r=X$.
The equation of motion for $P$ is more subtle, since a standard variation of the action with respect to $X$ is formally ambiguous, as noted in \cite{FriedmanLuoko}. The ambiguity arises because one must evaluate quantities on the shell ($L'$, $(N^r)'$, $N'$ and $R'$) that are (possibly) discontinuous at $r=X$:
\begin{equation}
\dot{P}=\left(N^r P - N\sqrt{L^{-2}P^2+M^2}\right)_{shell}'.
\label{ambiguous}
\end{equation}
However, it has been demonstrated in \cite{Menotti} that this ambiguity can be removed by requiring consistency with the constraints and the gravitational equations of motion (\ref{eq:GravEOM}), at least for the case of a dust shell. The argument described in \cite{Menotti} shows that one must average the discontinuous quantities when interpreting the equation of motion for the shell momentum, and similar reasoning leads to the same conclusion for the arbitrary perfect fluid shell described here. One then has the equation of motion
\begin{equation}
\dot{P}=\bar{(N^r)'}P-\frac{\bar{N'}}{\hat{L}}\sqrt{P^2+\hat{L}^2 \hat{M}^2}+\frac{\hat{N}\left(P^2\bar{L'}-\hat{L}^3 \hat{M} \bar{M'}\right)}{\hat{L}^2\sqrt{P^2+\hat{L}^2 \hat{M}^2}},
\label{eq:Pdot}
\end{equation}
with the average taken over $(N^r)'$ in the first term of the right-hand-side, and the last term containing the factor $\bar{M'}$ defined as $\bar{M'}=\hat{\frac{dM}{dR}}\bar{R'}$.
Let us briefly sketch the argument that leads to this result. To start, we take the time derivative of the (integrated and rearranged) momentum constraint:
\begin{equation}
\dot{P}=-\Delta\pi_L\frac{d}{dt}\left(\hat{L}\right)-\hat{L}\frac{d}{dt}\left(\Delta\pi_L\right).
\label{eq:constraintderivative}
\end{equation}
Then, by continuity of $\dot{L}$, we have
\begin{equation}
\frac{d}{dt}L(X)=L'(X\pm\epsilon)\dot{X}+\dot{L}(X\pm\epsilon)=\bar{L'}\dot{X}+\bar{\dot{L}}.
\end{equation}
Averaging the equation of motion for $L$, noting that $\frac{d}{dt}(\Delta\pi_L)=\Delta(\pi_L')\dot{X}+\Delta(\dot{\pi_L})$, and calculating $\Delta(\dot{\pi_L})$ from the equation of motion for $\pi_L$, we obtain
\begin{equation}
\dot{P}=\mathcal{\dot{P}}+\Phi,
\end{equation}
with $\mathcal{\dot{P}}$ representing the right side of equation (\ref{eq:Pdot}), and $\Phi$ defined such that
\begin{eqnarray}
\Phi=&-&P\frac{\hat{N}}{\hat{R}\hat{L}}\bar{\pi_R}+\frac{\hat{N}\Delta R'\bar{R'}}{\hat{L}}+\Delta N'\frac{\bar{R'}\hat{R}}{\hat{L}} \nonumber\\
&-&\hat{L}\Delta\pi_L'(\hat{N^r}+\dot{X})+\frac{\hat{N}\hat{M}\hat{\frac{dM}{dR}}\bar{R'}}{\sqrt{\hat{L}^{-2}P^2+\hat{M}^2}}.
\end{eqnarray}
To then demonstrate that $\Phi$ vanishes, one needs to take the jump of the momentum constraint across the shell to obtain $\hat{L}\Delta\pi_L'=\bar{R'}\Delta\pi_R+\bar{\pi_R}\Delta R'$, then integrate the equation of motion for $\pi_R$ across the shell, and use the result, combined with the fact that the delta contribution to $\dot{\pi_R}$ is given by $-\dot{X}(\Delta\pi_R)\delta(r-X)$ \cite{FriedmanLuoko}.
\subsection{Phase Space Reduction}
We now seek a description of the system in terms of only the shell coordinate $X$ and a conjugate momentum $P_c$. Note that it is not necessarily true that $P_c$ will coincide with the conjugate momentum $P$ for the unreduced problem, as will become clear in what follows.
To proceed with the Hamiltonian reduction, we will make use of the Liouville form $\mathcal{F}$ and the symplectic form $\Omega$, which on the full phase space (denoted by $\Gamma$) can be written as
\begin{equation}
\mathcal{F}=P_c \bm{\delta} X + \int{dr\left(\pi_L \bm{\delta} L +\pi_R \bm{\delta} R\right)}
\end{equation}
and
\begin{equation}
\Omega = \bm{\delta} P_c \wedge \bm{\delta} X + \int{dr \left(\bm{\delta} \pi_L \wedge \bm{\delta} L + \bm{\delta} \pi_R \wedge \bm{\delta} R \right)},
\end{equation}
respectively, with $\bm{\delta}$ denoting an exterior derivative in the associated functional space (see \cite{FriedmanLuoko} for more details). The reduced phase space $\bar{\Gamma}$ is defined as the set of equivalence classes in $\Gamma$ under changes of coordinates, and each (permissible) choice of coordinates defines a hypersurface $\bar{H}\subseteq \Gamma$ that is transversal to the orbits generated by coordinate transformations; this ensures that there exists an isomorphism between $\bar{\Gamma}$ and the representative hypersurface $\bar{H}$.
At this point we can determine the symplectic form $\bar{\Omega}$ induced on $\bar{H}$ as follows: first, consider the pullback of $\mathcal{F}$ to $\bar{H}$; this yields a quantity which we denote by $\mathcal{F}_{\bar{H}}$. Then, the symplectic form $\Omega_{\bar{H}}$ on the representative hypersurface $\bar{H}$ (corresponding to $\bar{\Omega}$) takes the form
\begin{equation}
\Omega_{\bar{H}}=\bm{\delta} \mathcal{F}_{\bar{H}}.
\end{equation}
This quantity defines the canonical structure of the reduced phase space.
To explicitly determine the (nonlocal) contribution of the gravitational variables to the dynamics on the reduced phase space, we can solve the GR constraints for the gravitational momenta, insert the solutions into the Liouville form on the full phase space, and perform the integration to express the gravitational contribution solely in terms of the (local) shell variables. Away from the shell, take the following linear combination of the constraints:
\begin{equation}
-\frac{R'}{L}H_0-\frac{\pi_L}{RL}H_r=\mathcal{M}',
\end{equation}
for
\begin{equation}
\mathcal{M}(r)=\frac{\pi_L^2}{2R}+\frac{R}{2}-\frac{R(R')^2}{2L^2}.
\end{equation}
The quantity $\mathcal{M}(r)$ corresponds to the ADM mass $H$ when evaluated outside of the shell, and vanishes inside the shell. This enables us to solve for the gravitational momenta $\pi_L$, $\pi_R$ away from the shell. The result is
\begin{equation}
\pi_L=\pm R\sqrt{\left(\frac{R'}{L}\right)^2-1+\frac{2\mathcal{M}}{R}},\hspace{8pt}\pi_R=\frac{L}{R'}\pi_L'.
\label{eq:GravMomenta}
\end{equation}
One then makes a coordinate choice, to pick out a representative hypersurface $\bar{H}$. The coordinates we will use resemble the flat-slice coordinates \{$L=1$, $R=r$\} described in \cite{KrausPainleveGullstrand} (also known as Painlev\'e-Gullstrand coordinates), though we will have a deformation region $X-\epsilon<r<X$ explicitly included, in order to both satisfy the constraints and yield a continuous spatial metric. The deformation region is related to a jump in $R'$ across the shell. This can be seen by first integrating the Hamiltonian and momentum constraints across the shell. Doing so yields, respectively,
\begin{equation}
\Delta R'=-\frac{\hat{V}}{\hat{R}}, \hspace{8pt} \Delta\pi_L=-\frac{P}{\hat{L}},
\label{eq:Jumps}
\end{equation}
where $V=\sqrt{P^2+M^2L^2}$ and $\Delta$ indicates the jump of a quantity across the shell. We therefore take
\begin{equation}
L=1, \hspace{8pt} R(r,t)=r-\frac{\epsilon}{X}\hat{V}f\left(\frac{X-r}{\epsilon}\right),
\label{eq:Coords}
\end{equation}
for a function $f$ having support in the interval $(0,1)$ with the property $f'(x)\rightarrow 1$ as $x\searrow0$. Outside of the deformation region, these coincide with flat-slice coordinates. For concreteness, let us suppose $f(x)$ takes the form
\begin{equation}
f(x)=x e^{-x^2/(1-x^2)}
\end{equation}
for all $x \in (0,1)$.
In what follows, it will be useful to note that $\hat{M}=M(\hat{R})=M(X)$, and that now $P$ is considered to be a function of $X$ and $H$, as a consequence of the gravitational constraints. We can implicitly determine this function by inserting the gravitational momentum solutions away from the shell given by equation (\ref{eq:GravMomenta}) into the jump equations (\ref{eq:Jumps}) and squaring. We are then left with
\begin{equation}
H=\hat{V}+\frac{\hat{M}^2}{2X}-P\sqrt{\frac{2H}{X}}.
\label{eq:DeterminesP}
\end{equation}
With this coordinate choice, the only gravitational contribution to the Liouville form comes from the $\pi_R$ term, and only from within the deformation region. Keeping in mind that we only care about terms that remain nonzero in the $\epsilon\rightarrow0$ limit, we have, in the deformation region,
\begin{equation}
\pi_R = \frac{X R''}{\sqrt{\left(R'\right)^2-1}}+\mathcal{O}(1),
\end{equation}
since $R=X+\mathcal{O}(\epsilon)$ and $R''=\mathcal{O}\left(\epsilon^{-1}\right)$. One can also note that $\bm{\delta} R = \left(1-R'\right)\bm{\delta} X+\mathcal{O}(\epsilon)$, and express the gravitational contribution to the Liouville form as
\begin{equation}
\int_{X-\epsilon}^{X}{dr\, \pi_R \bm{\delta} R}=X\bm{\delta} X\int_{X-\epsilon}^{X}{dr\,\frac{R'' \left(1-R'\right)}{\sqrt{\left(R'\right)^2-1}}}+\mathcal{O}(\epsilon).
\end{equation}
To evaluate this integral, one can change the integration variable from $r$ to $v=R'$:
\begin{equation}
\int_{X-\epsilon}^{X}{dr\, \pi_R \bm{\delta} R}=X\bm{\delta} X\int_{1}^{R_-'}{dv\,\frac{\left(1-v\right)}{\sqrt{v^2-1}}}+\mathcal{O}(\epsilon),
\end{equation}
with $R_-'$ being $R'$ evaluated just inside the shell. The integration is then straightforward, and after applying (\ref{eq:DeterminesP}) and making some rearrangements one arrives at
\begin{equation}
X\bm{\delta} X \left[-\frac{P}{X}-\sqrt{\frac{2 H}{X}}+\ln{\left(1+\sqrt{\frac{2H}{X}}+\frac{\hat{V}+P}{X}\right)}\right],
\end{equation}
plus terms that vanish as $\epsilon\rightarrow 0$. This completes the calculation of $\mathcal{F}_{\bar{H}}$, the pullback of the full Liouville form $\mathcal{F}$ to $\bar{H}$:
\begin{equation}
\mathcal{F}_{\bar{H}}=P_c \bm{\delta} X,
\end{equation}
with the reduced canonical momentum evidently given by
\begin{equation}
P_c = -\sqrt{2HX}+X\ln{\left(1+\sqrt{\frac{2H}{X}}+\frac{\hat{V}+P}{X}\right)}.
\end{equation}
This result agrees with \cite{FriedmanLuoko} in the limit of a dust shell ($\hat{M}'=0$).
To connect this with the expression derived by Kraus and Wilczek, we need only apply the expression (\ref{eq:DeterminesP}) to the argument of the logarithm, which leads to
\begin{equation}
P_c = -\sqrt{2HX}-X\ln{\left(\frac{X+\hat{V}-P-\sqrt{2HX}}{X}\right)}.
\label{eq:Pc}
\end{equation}
This form of the reduced momentum coincides with \cite{KrausWilczek} in the dust-shell limit.
\subsection{Boundary Terms}
To obtain a well-defined variational principle for the reduced problem, we must be careful with boundary terms, as first noted in \cite{RT1974} and \cite{Unruh1976}. In \cite{KrausWilczek}, it is observed that for asymptotically-flat spacetimes, we simply need to subtract the ADM mass (denoted suggestively by $H$) from our reduced Lagrangian. Specifically, as mentioned in \cite{FriedmanLuoko}, a nonzero boundary variation results from integrating by parts the term $\int{dtdrN^rL(\delta \pi_L)'}$, which is part of the momentum constraint. The only contribution comes from infinity, and in this case we have $N^r \rightarrow N\sqrt{2H/r}$, $N\rightarrow 1$, and
\begin{equation}
\delta (\pi_L)\rightarrow \delta(\sqrt{2Hr}) = \sqrt{\frac{r}{2H}}\delta H,
\end{equation}
so the variation of the boundary term is cancelled if we add to the action the term
\begin{equation}
I_{bdry}=-\int{dt\, H}.
\end{equation}
Including the boundary term to the action defined by $\mathcal{F}_{\bar{H}}$, one obtains the reduced action
\begin{equation}
I_{reduced}=\int{dt\, \left(P_c \dot{X} - H\right)},
\label{eq:ReducedAction}
\end{equation}
with the reduced momentum given by (\ref{eq:Pc}). From the form of the reduced action (\ref{eq:ReducedAction}), we can see that the ADM mass is the reduced Hamiltonian, defined implicitly by (\ref{eq:Pc}) and (\ref{eq:DeterminesP}).
Since (\ref{eq:DeterminesP}) has more than one solution $P=P(X,H)$, our conjugate momentum $P_c$ in turn becomes a multi-valued function of $X$ and $H$, as one expects from a theory that allows the degree of freedom to either increase or decrease. Explicitly, $P$ is given by
\begin{eqnarray}
&P = \frac{1}{1-\frac{2H}{X}}\left(\sqrt{\frac{2H}{X}}\left(H-\frac{\hat{M}^2}{2X}\right)\right)& \nonumber\\
&\pm \frac{1}{1-\frac{2H}{X}}\left(\sqrt{\left(H-\frac{\hat{M}^2}{2X}\right)^2-\hat{M}^2\left(1-\frac{2H}{X}\right)}\right),&
\end{eqnarray}
while the combination $\hat{V}-P$ that appears in the reduced momentum (\ref{eq:Pc}) is
\begin{equation}
\hat{V}-P=\frac{H-\frac{\hat{M}^2}{2X}\mp \sqrt{\left(H-\frac{\hat{M}^2}{2X}\right)^2-\hat{M}^2\left(1-\frac{2H}{X}\right)}}{1+\sqrt{\frac{2H}{X}}}.
\end{equation}
\subsection{Constructing Classical Spacetime}
Suppose one can find a solution $X(t)$ to the classical equations of motion for the reduced system (\ref{eq:ReducedAction}). Then, the gravitational constraints and equations of motion (\ref{eq:GravEOM}) can be solved to determine all the metric components $g_{\mu\nu}$. Therefore, from the reduced system solution $X(t)$ one can construct the classical spacetime structure, as we will now demonstrate.
By inserting the gravitational momenta solutions (\ref{eq:GravMomenta}) into the gravitational equations of motion (\ref{eq:GravEOM}), one can obtain the lapse function $N$ and the radial shift component $N^r$ that correspond to our coordinate choice (\ref{eq:Coords}).
Outside of the shell, one finds the familiar Schwarzschild structure, in flat-slice coordinates. The lapse function is constant, and unity if we want a time coordinate that increases towards the future, while the radial shift is given by
\begin{equation}
N^r(r\geq X)= \pm \sqrt{\frac{2H}{r}}.
\end{equation}
The $\pm$ here indicates two possible time-slicings, though we will often take the upper sign (this means that the gravitational momenta solution (\ref{eq:GravMomenta}) should take the upper sign as well, to ensure $N\rightarrow 1$ as $r\rightarrow \infty$).
Along with the expression (\ref{eq:DeterminesP}) for $P$ in terms of $X$ and $H$, we now have enough information to determine the classical path $X(t)$, since $H$ is constant along such paths. Specifically, the equation of motion for $X$ becomes
\begin{equation}
\dot{X}=\frac{P}{\sqrt{P^2+\hat{M}^2}}-\sqrt{\frac{2H}{X}},
\end{equation}
which leads to the expression
\begin{eqnarray}
\frac{dt}{dX} &=& \frac{\sqrt{2H X}}{X-2H} \\
&\pm& \frac{H-\frac{\hat{M}^2}{2X}}{\left(1-\frac{2H}{X}\right)\sqrt{\left(H-\frac{\hat{M}^2}{2X}\right)^2-\hat{M}^2\left(1-\frac{2H}{X}\right)}}. \nonumber
\end{eqnarray}
Therefore, finding the classical path $X(t)$ has been reduced to quadrature and inversion.
With the classical path known, one can also calculate the classical action, as done for the case of dust in \cite{KrausWilczek}:
\begin{equation}
S(t,X(t))=S(0,X(0))+\int_0^t{d\tilde{t}\left[P_c(\tilde{t})\dot{X}(\tilde{t})-H\right]},
\end{equation}
with
\begin{equation}
P_c(0)=\frac{\partial S}{\partial X}(0,X(0)).
\end{equation}
Unlike the (massless) dust case, however, our classical path $X(t)$ is not a null geodesic of the flat-slice metric
\begin{equation}
ds^2=-dt^2+\left(dr+\sqrt{\frac{2H}{r}}dt\right)^2,
\end{equation}
and so we cannot so easily determine explicit expressions for our shell trajectories.
\section{Interferometry}
\subsection{Equation of State Determination}
Up until this point, the function $M(X)$ has been left unspecified, though we have established the identifications $\sigma=M(X)/4\pi X^2$ for the density and $p=-M'(X)/8\pi X$ for the pressure. We would like to exploit this freedom for the purposes of interferometry. To maintain internal consistency, there should be a relationship $p=p(\sigma)$, which represents an equation of state for our fluid shell. The function $M(X)$ parametrizes this relationship, though not every choice of $M(X)$ yields a consistent (let alone physical) equation of state.
\begin{figure}[fig:Interferometer]
\includegraphics[width=3.5in]{Interferometer.jpg}
\caption{Schematic representation of the splitting, reflecting, and recombining that occur in our shell interferometer. $R$ and $T$ are effective reflection and transmission coefficients for the beam-splitter.}
\label{fig:Interferometer}
\end{figure}
The interferometric setup resembles that of Michelson, except we only have one spatial dimension to work with, since our system is spherically symmetric. Still, we would like the equation of state to produce two `reflectors' - one to reflect the shell outward if it gets too small, and one to reflect the shell inward if it gets too large. Also, we would like the equivalent of a `half-silvered mirror' to be in between the two reflectors, to act as a beam-splitter. This is depicted schematically in Figure~\ref{fig:Interferometer}, with $X_\pm$ being the shell radii that correspond to the reflectors, and $X_\delta$ the radius corresponding to the splitter. Accordingly, our equation of state $p=p(\sigma)$ must have a large positive peak for some large density, a large negative peak for some small density, and an intermediate peak for some intermediate density.
It would be convenient to use delta functions for these purposes, but due to the conversion between $\delta(\sigma-\sigma_0)$ and $\delta(X-X_0)$ and the resulting appearance of products of delta functions, this possibility seems problematic. Therefore, we have been considering the simplest alternative one could think of: rectangular barriers. These can be described with the use of step functions, which we will define such that $\Theta(x<0)=0$ and $\Theta(x>0)=1$.
The equation of state, then, takes the form
\begin{eqnarray}
p &=& p_1 \left( \Theta(\sigma-\sigma_1)-\Theta(\sigma-\sigma_2) \right) \nonumber \\
&& + p_2 \left( \Theta(\sigma-\sigma_3)-\Theta(\sigma-\sigma_4) \right) \nonumber \\
&& + p_3 \left( \Theta(\sigma-\sigma_5)-\Theta(\sigma-\sigma_6) \right),
\label{eq:EOS}
\end{eqnarray}
with $\sigma_{i+1}>\sigma_{i}$, $p_1<0$, and $p_2, p_3 > 0$. We may as well take $p_1=-p_3$, since both of these peaks serve the same purpose of reflecting, but we will not yet impose this condition.
\begin{figure}
\centerline{\epsfig{file=EOS.png, width=3.5in}}
\caption{
A sample equation of state represented by (\ref{eq:EOS}).
}
\label{fig:EOS}
\end{figure}
Figure~\ref{fig:EOS} illustrates the desired step function peaks, to enable our system to operate as an interferometer.
One would now like to find the function $M(X)$ that parametrizes the equation of state (\ref{eq:EOS}). If we could express (\ref{eq:EOS}) as $p = \sum_i{\tilde{p}_i \Theta (X-X_i)}$, then the identification $p=-M'(X)/8\pi X$ would imply
\begin{equation}
\label{eq:MassFunction}
M(X)=M_0+4\pi\sum_i{\tilde{p}_i\left(X_i^2-X^2\right)\Theta(X-X_i)},
\end{equation}
which would yield a density given by
\begin{equation}
\sigma = \frac{M_0}{4\pi X^2}+\sum_i{\tilde{p}_i\left(\frac{X_i^2}{X^2}-1\right)\Theta(X-X_i)}.
\end{equation}
The problem with this possibility is that, in general, it isn't necessarily true that $\Theta(X-X_i)$ produces the same (reversed) ordering as $\Theta(\sigma-\sigma_i)$, given that $\sigma_i=M(X_i)/4\pi X_i^2$. This problem can be avoided by making sure that the density $\sigma$ is a monotonically decreasing function of $X$. This leads to the condition
\begin{equation}
\frac{M_0}{4\pi} \geq -\sum_{i}\tilde{p}_i X_i^2 \Theta(X-X_i).
\end{equation}
To understand what this means in terms of the pressure peaks in our equation of state (\ref{eq:EOS}), we first note that if $\sigma$ monotonically decreases in $X$, then step functions can be converted by $\Theta(\sigma-\sigma_i) = 1-\Theta(X-X_i)$. This allows us to conclude that $\tilde{p_2}=-\tilde{p_1}=p_1$, $\tilde{p_4}=-\tilde{p_3}=p_2$, and $\tilde{p_6}=-\tilde{p_5}=p_3$. Then, one finds that monotonicity is maintained as long as
\begin{eqnarray}
\frac{M_0}{4\pi} > max\{ p_3 X_{5,6}^2, p_3 X_{5,6}^2+p_2 X_{3,4}^2, \nonumber \\
p_3 X_{5,6}^2+ p_2 X_{3,4}^2-p_1 X_2^2 \},
\end{eqnarray}
where the notation $X_{i,j}^2=X_i^2-X_j^2$ was introduced, for brevity.
Since an equation of state (\ref{eq:EOS}) is described by the pressure as a function of density, one should translate the conditions for monotonicity in terms of the step locations $\{\sigma_i\}$ and the step amplitudes $\{p_i\}$. To convert between the $\{X_i\}$ and the $\{\sigma_i\}$, one can use the relations
\begin{eqnarray}
X_6^2 &=&\frac{M_0}{4\pi \sigma_6}, \nonumber\\
X_5^2&=&\frac{M_0}{4\pi \sigma_6}\frac{(\sigma_6+p_3)}{(\sigma_5+p_3)}, \nonumber\\
X_4^2&=&\frac{M_0}{4\pi \sigma_6}\frac{(\sigma_6+p_3)}{(\sigma_5+p_3)}\frac{\sigma_5}{\sigma_4}, \nonumber\\
X_3^2&=&\frac{M_0}{4\pi \sigma_6}\frac{(\sigma_6+p_3)}{(\sigma_5+p_3)}\frac{\sigma_5}{\sigma_4}\frac{(\sigma_4+p_2)}{(\sigma_3+p_2)}, \nonumber\\
X_2^2&=&\frac{M_0}{4\pi \sigma_6}\frac{(\sigma_6+p_3)}{(\sigma_5+p_3)}\frac{\sigma_5}{\sigma_4}\frac{(\sigma_4+p_2)}{(\sigma_3+p_2)}\frac{\sigma_3}{\sigma_2}, \nonumber\\
X_1^2&=&\frac{M_0}{4\pi \sigma_6}\frac{(\sigma_6+p_3)}{(\sigma_5+p_3)}\frac{\sigma_5}{\sigma_4}\frac{(\sigma_4+p_2)}{(\sigma_3+p_2)}\frac{\sigma_3}{\sigma_2}\frac{(\sigma_2+p_1)}{(\sigma_1+p_1)}.
\label{eq:Convert}
\end{eqnarray}
With these expressions, one can write the monotonicity conditions in the much simpler form
\begin{equation}
\{\sigma_5 > 0,\sigma_3 > 0, \sigma_1+p_1 > 0 \}.
\end{equation}
Thus, as long as we keep the density $\sigma$ positive, it will be monotonic in $X$ provided $\sigma_1+p_1 > 0$.
\subsection{Flat Spacetime Limit}
To determine whether or not gravity produces some form of decoherence in our interferometer, let us first clarify the manner in which coherence manifests itself in the absence of gravity. In this case spacetime is flat, and along the arms of the interferometer defined by (\ref{eq:MassFunction}) the shell behaves as a free particle.
\begin{figure}
\centerline{\epsfig{file=MassFunction.png, width=3.5in}}
\caption{A sample mass function $\hat{M}$ is plotted with respect to the shell radius $X$. The approximate step function near $X=2$ serves as a beam-splitter, and the steep quadratic sides correspond to the inner and outer reflectors of the interferometer.}
\label{fig:MassFunction}
\end{figure}
As evident from Figure~\ref{fig:MassFunction}, the mass of the ``free" shell is different on each interferometer arm. Let us call the inner mass $M_-$, and the outer mass $M_+$, such that $M_->M_+$. For simplicity, suppose the reflectors are perfect, which for this system means that the quadratic walls of the mass function are large and steep. Similarly, let the quadratic beam-splitter interval be approximated by a step function, to ensure that only constant mass function basis states need to be used in the quantum analysis.
Further, let us treat each element of the interferometer separately, in a similar manner to that which is done in optical systems. The initial state will first encounter the splitter, at which point each incoming mode will transform into a reflected mode with a factor $R_\leftarrow$ and a transmitted mode with a factor $T_\leftarrow$ (subscripts are used here because the reflection/transmission coefficients depend on the direction the incoming state encounters the splitter from).
The split initial state components will then perfectly reflect off of the outer/inner reflectors, and travel back towards one another to the beam-splitter. Upon recombination there will be further splitting of the components coming from each direction of the splitter, which produces two outputs (one going in each direction from the splitter) that are themselves composed of two parts; it is the interference between these two parts of each output that we are interested in.
Let us now describe the process in detail. For the purposes of this paper, we will restrict our attention to a single-mode input, since the multi-mode analysis is more involved and will be presented elsewhere \cite{GU14}. We will approximate the single-mode input by an ingoing WKB state:
\begin{equation}
\Psi_0= \frac{e^{i\int{dX P_{-+}}}}{\sqrt{|\partial E/\partial P_{-+}}|}\equiv \psi_{-+},
\end{equation}
where the first set of plus/minuses of the reduced momentum $P$ indicating outgoing/ingoing, and the second set indicating evaluations of $P$ as $X$ approaches $X_\delta$ from above/below (we have dropped the subscript $c$ on the reduced momentum here and for the rest of the paper, for brevity). We will define the integration such that the lower bound in $X$ is $X_\delta$.
Treating the first splitting on its own, let us consider the wavefunction
\begin{displaymath}
\Psi = \left\{
\begin{array}{lr}
\psi_{-+}+R_{\leftarrow}\psi_{++} & : X > X_\delta \\
T_{\leftarrow}\psi_{--} & : X < X_\delta
\end{array}
\right.
\end{displaymath}
The (classical) flat spacetime Hamiltonian satisfies $H=\sqrt{P^2+\hat{M}^2}$, which in the nonrelativistic limit yields $H\approx \hat{M}+P^2/2\hat{M}$. Applying wavefunction continuity at $X_\delta$, and integrating the nonrelativistic Schr\"{o}dinger equation
\begin{equation}
i\frac{\partial \psi}{\partial t}= \hat{M}\psi- \frac{1}{2}\frac{\partial}{\partial X}\left( \frac{1}{\hat{M}}\frac{\partial \psi}{\partial X} \right)
\end{equation}
across $X_\delta$, one can obtain the reflection and transmission amplitudes $R_\leftarrow$ and $T_\leftarrow$. The equations take a simpler form after transforming to the variables $\bar{R}_\leftarrow$ and $\bar{T}_\leftarrow$, which are defined by
\begin{equation}
\bar{R}_{\leftarrow} \equiv \sqrt{\left|\frac{\frac{\partial E}{\partial P_{-+}}}{\frac{\partial E}{\partial P_{++}}}\right|} R_\leftarrow, \hspace{6pt}
\bar{T}_{\leftarrow} \equiv \sqrt{\left|\frac{\frac{\partial E}{\partial P_{-+}}}{\frac{\partial E}{\partial P_{--}}}\right|} T_\leftarrow.
\end{equation}
One can then easily solve for the new variables:
\begin{equation}
\bar{R}_{\leftarrow}=\frac{M_- P_{-+}-M_+ P_{--}}{M_+ P_{--}-M_- P_{++}}, \hspace{6pt} \bar{T}_\leftarrow=\frac{M_- (P_{++}-P_{-+})}{M_- P_{++}-M_+ P_{--}}.
\end{equation}
For convenience, we can also derive the reflection and transmission amplitudes from the left, which are found to be
\begin{equation}
\bar{R}_{\rightarrow}=\frac{M_- P_{++}-M_+ P_{+-}}{M_+ P_{--}-M_- P_{++}}, \hspace{6pt} \bar{T}_\rightarrow=\frac{M_+ (P_{+-}-P_{--})}{M_- P_{++}-M_+ P_{--}},
\end{equation}
using the similar definitions
\begin{equation}
\bar{R}_{\rightarrow} \equiv \sqrt{\left|\frac{\frac{\partial E}{\partial P_{+-}}}{\frac{\partial E}{\partial P_{--}}}\right|} R_\rightarrow, \hspace{6pt}
\bar{T}_{\rightarrow} \equiv \sqrt{\left|\frac{\frac{\partial E}{\partial P_{+-}}}{\frac{\partial E}{\partial P_{++}}}\right|} T_\rightarrow.
\end{equation}
Let us call the outgoing state after the split $\Psi^{(i)}_+$ and the ingoing state $\Psi^{(i)}_-$. We then can consider the first splitting a transformation of the wavefunction such that
\begin{equation}
\Psi_0=\psi_{-+} \rightarrow \left( \begin{array}{c} \Psi^{(i)}_+ \\ \Psi^{(i)}_- \end{array} \right)=\left( \begin{array}{c} R_{\leftarrow}\psi_{++} \\ T_{\leftarrow}\psi_{--} \end{array} \right).
\end{equation}
This splitting should preserve the probability current, for consistency. In the nonrelativistic, flat spacetime limit, the probability current $J$ satisfies the continuity equation
\begin{equation}
\frac{\partial}{\partial t}\left(|\psi|^2\right)+\frac{\partial}{\partial X} J=0
\label{eq:Cont}
\end{equation}
and is given by the usual quantum mechanics expression
\begin{equation}
\frac{1}{2im}\left(\psi^*\psi^\prime-\psi\psi^{*\prime}\right).
\label{eq:J0}
\end{equation}
Therefore, in this limit we have an input probability current of
\begin{equation}
J_0 = |\Psi_0|^2 \frac{P_{-+}}{M_+}.
\label{eq:JInit}
\end{equation}
After first encountering the beam-splitter, the probability current (\ref{eq:JInit}) splits into reflected and transmitted components
\begin{eqnarray}
|J_+^{(i)}| &=& \frac{1}{2iM_+}\left(\Psi_+^{(i)*}\left(\Psi_+^{(i)}\right)^\prime-\Psi_+^{(i)}\left(\Psi_+^{(i)}\right)^{*\prime}\right)\nonumber\\
&=& |J_0|\bar{R}_\leftarrow^2
\end{eqnarray}
and
\begin{eqnarray}
|J_-^{(i)}| &=& \frac{1}{2iM_-}\left(\Psi_-^{(i)*}\left(\Psi_-^{(i)}\right)^\prime-\Psi_-^{(i)}\left(\Psi_-{(i)}\right)^{*\prime}\right)\nonumber\\
&=& |J_0|\left(\frac{M_+ P_{--}}{M_- P_{-+}}\right)\bar{T}_\leftarrow^2 .
\end{eqnarray}
The splitting preserves probability current, as can be confirmed by observing that $\left| \frac{J_+^{(i)}}{J_0} \right|+\left| \frac{J_-^{(i)}}{J_0} \right|$ is unity. The terms $\left| \frac{J_+^{(i)}}{J_0} \right|$ and $\left| \frac{J_-^{(i)}}{J_0} \right|$ are usually called the reflection and transmission coefficients, respectively.
The second transformation propagates the modes along the interferometer arms, such that
\begin{equation}
\left( \begin{array}{c} \Psi^{(i)}_+ \\ \Psi^{(i)}_- \end{array} \right)
\rightarrow \left( \begin{array}{c} \Psi^{(ii)}_+ \\ \Psi^{(ii)}_- \end{array} \right)=
\left( \begin{array}{c} (E,_{++})^{-1/2}R_\leftarrow e^{i\Phi_{++}} \\ (E,_{--})^{-1/2}T_\leftarrow e^{i\Phi_{--}} \end{array} \right).
\end{equation}
For brevity, the notation $E,_{\pm\pm}$ was used to denote $\partial E / \partial P_{\pm\pm}$, and it is understood that we are evaluating these quantities at the outer walls of the interferometer. We have also introduced the quantities $\Phi_{\pm\pm}=\phi_{\pm\pm}-Et_{\pm\pm}$, for $\phi_{\pm\pm}=\int_{X_\delta}^{X_\pm}{dX P_{\pm\pm}}$, where $t_{++}$ and $t_{--}$ denote the travel times from the splitter to $X_+$ and $X_-$, respectively.
The modes then reflect off of the outer walls, as
\begin{equation}
\left( \begin{array}{c} \Psi^{(ii)}_+ \\ \Psi^{(ii)}_- \end{array} \right)
\rightarrow \left( \begin{array}{c} \Psi^{(iii)}_+ \\ \Psi^{(iii)}_- \end{array} \right)=
\left( \begin{array}{c} (E,_{-+})^{-1/2}R_\leftarrow e^{i\Phi_{++}}R^\rightarrow \\ (E,_{+-})^{-1/2}T_\leftarrow e^{i\Phi_{--}}R^\leftarrow \end{array} \right).
\end{equation}
The outer wall reflection amplitudes $\left(R^\rightarrow,R^\leftarrow\right)$ only depend on continuity of the wavefunction. To obtain the reflection amplitude from the left, for instance, consider the wavefunction
\begin{displaymath}
\Psi=\psi_{++} \rightarrow \left\{
\begin{array}{lr}
0 & : X > X_+ \\
\left(\psi_{++}+R^{\rightarrow}\psi_{--}\right) & : X < X_+
\end{array}
\right.
\end{displaymath}
By applying wavefunction continuity at $X_+$, one immediately obtains $R^\rightarrow$. $R^\leftarrow$ can be similarly determined, and the results are
\begin{equation}
\bar{R}^\rightarrow=-e^{i\left(\phi_{++}+\phi_{-+}\right)}, \hspace{6pt} \bar{R}^\leftarrow=-e^{i\left(\phi_{+-}+\phi_{--}\right)},
\end{equation}
with help of the simplifying definitions
\begin{equation}
\bar{R}^\rightarrow \equiv \sqrt{\left|\frac{\frac{\partial E}{\partial P_{++}}}{\frac{\partial E}{\partial P_{-+}}}\right|} R^\rightarrow, \hspace{4pt} \bar{R}^\leftarrow \equiv \sqrt{\left|\frac{\frac{\partial E}{\partial P_{--}}}{\frac{\partial E}{\partial P_{+-}}}\right|} R^\leftarrow.
\end{equation}
The phases are defined such that $\phi_{\pm\mp}=\int_{X_\mp}^{X_\delta}{dX P_{\pm\mp}}$ (signs chosen together). We will refer to the modes after reflection from the outer walls as $\Psi^{(iii)}_\pm$.
Propagation along the arms back to the splitter then proceeds as
\begin{equation}
\left(
\begin{array}{c} \Psi^{(iii)}_+ \\ \Psi^{(iii)}_- \end{array} \right)
\rightarrow
\left(\begin{array}{c} \Psi^{(iv)}_+ \\ \Psi^{(iv)}_- \end{array} \right),
\end{equation}
for
\begin{equation}
\left(\begin{array}{c} \Psi^{(iv)}_+ \\ \Psi^{(iv)}_- \end{array} \right)=\left( \begin{array}{c} (E,_{-+})^{-1/2}R_\leftarrow e^{i\Phi_{++}}R^\rightarrow e^{i\Phi_{-+}} \\ (E,_{+-})^{-1/2}T_\leftarrow e^{i\Phi_{--}}R^\leftarrow e^{i\Phi_{+-}} \end{array} \right).
\end{equation}
In this expression, the quantities $E,_{-+}$ and $E,_{+-}$ are evaluated at the splitter, and we have used the definitions $\Phi_{\pm\mp}=\phi_{\pm\mp}-Et_{\pm\mp}$ (signs again chosen together). Here $t_{-+}$ and $t_{+-}$ denote the travel times from $X_+$ to the splitter and from $X_-$ to the splitter, respectively.
The second encounter with the splitter occurs as it did before, as
\begin{equation}
\left(\begin{array}{c} \Psi^{(iv)}_+ \\ \Psi^{(iv)}_- \end{array} \right)\rightarrow \left(\begin{array}{c} \Psi^{(v)}_+ \\ \Psi^{(v)}_- \end{array} \right)=
\begin{pmatrix} \bar{R}_\leftarrow & \bar{T}_\rightarrow \\ \bar{T}_\leftarrow & \bar{R}_\rightarrow \end{pmatrix} \left( \begin{array}{c} \Psi^{(iv)}_+ \\ \Psi^{(iv)}_- \end{array} \right).
\label{eq:FinalStates}
\end{equation}
At the order we are working at in $\hbar$, the derivatives of our final outputs satisfy
\begin{equation}
\frac{d}{dX}\Psi^{(v)}_\pm = iP_{\pm\pm}\Psi^{(v)}_\pm,
\end{equation}
and so the currents for our final outputs are given by
\begin{eqnarray}
J^{(v)}_\pm &=& \frac{1}{2iM_\pm}\left(\Psi_\pm^{(v)*}\left(\Psi_\pm^{(v)}\right)^\prime-\Psi_\pm^{(v)}\left(\Psi_\pm^{(v)}\right)^{*\prime}\right)\nonumber\\
&=& \frac{P_{\pm\pm}}{M_\pm}\left|\Psi^{(v)}_\pm\right|^2.
\end{eqnarray}
We then have enough information to calculate the final reflected and transmitted probability currents, which can be written
\begin{equation}
|J_+^{(v)}| = |J_0|\left[1-4\bar{R}_\leftarrow^2\left(1-\bar{R}_\leftarrow^2\right)\sin^2{\varphi}\right]
\end{equation}
and
\begin{equation}
|J_-^{(v)}| = |J_0|4\bar{R}_\leftarrow^2\left(1-\bar{R}_\leftarrow^2\right)\sin^2{\varphi},
\end{equation}
where we have defined $\varphi=\phi_{++}+\phi_{-+}-\phi_{+-}-\phi_{--}$ and made use of the identity $\bar{R}_\leftarrow^2+\frac{M_+ P_{--}}{M_- P_{-+}}\bar{T}_\leftarrow^2=1$. The flat-spacetime interferometer thus manifestly conserves probability current in all regions of the parameter space.
One can now search for a nice region in the parameter space that cancels one of the outputs. First, we would like to avoid regions of the parameter space that don't describe splitting, i.e. complete initial reflection or transmission by the beam-splitter. We can accomplish this in a simple way by enforcing an equal splitting condition, $\bar{R}_\leftarrow^2=1/2$. This leads to compact expressions for the final reflection and transmission coefficients, given by
\begin{equation}
R_f \equiv \frac{|J_+^{(v)}|}{|J_0|} = \cos^2{\varphi}
\end{equation}
and
\begin{equation}
T_f \equiv \frac{|J_-^{(v)}|}{|J_0|}= \sin^2{\varphi},
\end{equation}
respectively.
We should also make sure that our shell velocity doesn't approach the speed of light, since we are working in the nonrelativistic limit. For small shell speeds, given an outer mass $M_+$ and an initial speed $v_+$, the initial splitting will be equal provided the inner mass satisfies
\begin{equation}
M_- \approx M_+ \left[1+\left(6\sqrt{2}-8\right)v_+^2-\left(99\sqrt{2}-140\right)v_+^4\right].
\label{eq:Mminus}
\end{equation}
In the quantum context, the ``speed'' $v_+$ is defined such that $E=M_+ + \frac{1}{2} M_+ v_+^2$, for a WKB state with energy $E$.
If we denote the interferometer arm lengths by $L_\pm \equiv \pm (X_\pm-X_\delta)$, we can see from the form of the reflection and transmission coefficients that one of the outputs will be completely cancelled if
\begin{eqnarray}
\varphi &=& 2 L_+\sqrt{2M_+\left(E-M_+\right)}-2L_-\sqrt{2M_-\left(E-M_-\right)}\nonumber\\
&=& 2 L_+ M_+ v_+-2L_- M_- v_-\nonumber\\
&=& \frac{n\pi}{2},
\label{eq:cancel}
\end{eqnarray}
for $n\in \mathbb{Z}$. Thus, as the outer arm length is increased or decreased, the outputs are alternately cancelled out for each value of $n$ (odd values cancel the transmitted output, and even values cancel the reflected output), with partial interference for intermediate arm lengths that don't correspond to solutions of (\ref{eq:cancel}). This behaviour is a direct reflection of coherence in the flat spacetime interferometer.
\subsection{General Relativistic Picture}
The current framework was designed to facilitate the inclusion of general relativistic corrections. Several expressions become messier once one includes gravity, and some expressions fundamentally change in structure. For instance, the probability current given by (\ref{eq:J0}) is no longer conserved in systems with more general Hamiltonians. In fact, a probability current for an arbitrary Hamiltonian system has never been constructed; only special cases are known.
For our purposes, since we are working in the WKB regime, we may sometimes wish to approximate a general Hamiltonian $H(X,P)$ by the first three terms in a Taylor expansion in $P$, given by
\begin{equation}
H_w=H(X,0)+\left(\frac{\partial H}{\partial P}\right) P + \frac{1}{2}\left(\frac{\partial^2 H}{\partial P^2}\right) P^2,
\label{eq:HAp}
\end{equation}
with the $P$-derivatives evaluated at $P=0$. In the quantum theory, we can symmetrize the term linear in $P$ to enforce Hermiticity, i.e.
\begin{equation}
\left(\frac{\partial H}{\partial P}\right)P\rightarrow \frac{1}{2}\left(\hat{\left(\frac{\partial H}{\partial P}\right)}\hat{P}+\hat{P}\hat{\left(\frac{\partial H}{\partial P}\right)}\right),
\end{equation}
as well as ordering the quadratic term as
\begin{equation}
\left(\frac{\partial^2 H}{\partial P^2}\right) P^2 \rightarrow \hat{P}\hat{\left(\frac{\partial^2 H}{\partial P^2}\right)}\hat{P}.
\end{equation}
If we take this operator ordering of the approximate form (\ref{eq:HAp}) as an exact Hamiltonian, then we can find a probability current $J$ that satisfies the continuity equation (\ref{eq:Cont}), which we can express as
\begin{equation}
J = \left(\frac{\partial H}{\partial P}\right) |\Psi|^2 +\frac{1}{2i}\left(\frac{\partial^2 H}{\partial P^2}\right) \left(\Psi^*\Psi^\prime-\Psi\Psi^{*\prime}\right).
\label{eq:JG}
\end{equation}
The $P$-derivatives in this expression are again evaluated at $P=0$, and for the special case of $\{\left(\frac{\partial H}{\partial P}\right)=0, \left(\frac{\partial^2 H}{\partial P^2}\right) = 1/m\}$, we are left with the nonrelativistic, flat spacetime limit described by (\ref{eq:J0}).
In the limit of large $X$, the WKB Hamiltonian for our shell system is given by
\begin{equation}
H_w\sim \left(\hat{M}-\frac{\hat{M}^2}{18 X}\right)-\frac{2}{3}\sqrt{\frac{2\hat{M}}{X}}P+\left(\frac{1}{2\hat{M}}+\frac{1}{3X}\right)P^2,
\label{eq:WKBHam}
\end{equation}
so the generalized probability current is given by
\begin{equation}
J\sim -\frac{2}{3}\sqrt{\frac{2\hat{M}}{X}}|\Psi|^2+\left(1+\frac{2\hat{M}}{3X}\right)J_{s},
\end{equation}
with $J_{s}$ being the standard (nonrelativistic) expression (\ref{eq:J0}) for the probability current. Note that although the functional form of $J_{s}$ with respect to $\Psi$ is the same as the nonrelativistic current (\ref{eq:J0}), in the above expression we are inserting the general relativistic WKB wavefunction $\Psi$.
Since our Schr\"{o}dinger equation now takes the asymptotic form
\begin{equation}
H_w \Psi = i \frac{\partial}{\partial t}\Psi,
\label{eq:Hw}
\end{equation}
taking the operator ordering mentioned above, we no longer have the simple reflection and transmission amplitudes obtained in the previous section. For instance, integrating (\ref{eq:Hw}) across $X_\delta$ yields
\begin{equation}
\left[\left(\frac{3}{2\hat{M}}+\frac{1}{X}\right)\Psi'\right]_\delta=i\sqrt{\frac{2}{X_\delta}}\left[\sqrt{\hat{M}}\right]_\delta \Psi(X_\delta).
\end{equation}
Here, $\left[\cdot\right]_\delta$ represents the jump of a quantity across $X_\delta$.
\begin{figure*}
\includegraphics[width=\textwidth,height=5.4cm]{Interference1B.jpg}
\caption{Sample interference pattern, for $M_+=15$, $v_+=0.003$, $X_-=5000$, $L_-=20$, and $M_-$ chosen to satisfy (\ref{eq:Mminus}), plotted against the outer mirror position, $X_+$.}
\label{fig:Interference1B}
\end{figure*}
\begin{figure*}
\includegraphics[width=\textwidth,height=5.4cm]{Interference2B.jpg}
\caption{Sample interference pattern, for $M_+=15$, $v_+=0.01$, $X_-=200$, $L_-=20$, and $M_-$ chosen to satisfy (\ref{eq:Mminus}), plotted against the outer mirror position, $X_+$.}
\label{fig:Interference2B}
\end{figure*}
To the order we are working at in $\hbar$, the new reflection and transmission amplitudes for scattering from the right are given by
\begin{widetext}
\begin{eqnarray}
\bar{R}_\leftarrow = \frac{\sqrt{\frac{2}{X_\delta}}\left[\sqrt{\hat{M}}\right]_\delta+\left(\frac{3}{2M_-}+\frac{1}{X_\delta}\right)P_{--}-\left(\frac{3}{2M_+}+\frac{1}{X_\delta}\right)P_{-+}}{-\sqrt{\frac{2}{X_\delta}}\left[\sqrt{\hat{M}}\right]_\delta-\left(\frac{3}{2M_-}+\frac{1}{X_\delta}\right)P_{--}+\left(\frac{3}{2M_+}+\frac{1}{X_\delta}\right)P_{++}}
\end{eqnarray}
and
\begin{eqnarray}
\bar{T}_\leftarrow =\frac{\left(\frac{3}{2M_+}-\frac{1}{X_\delta}\right)\left(P_{++}-P_{-+}\right)}{-\sqrt{\frac{2}{X_\delta}}\left[\sqrt{\hat{M}}\right]_\delta-\left(\frac{3}{2M_-}+\frac{1}{X_\delta}\right)P_{--}+\left(\frac{3}{2M_+}+\frac{1}{X_\delta}\right)P_{++}}.
\end{eqnarray}
Similarly, for scattering from the left we have
\begin{eqnarray}
\bar{R}_\rightarrow = \frac{\sqrt{\frac{2}{X_\delta}}\left[\sqrt{\hat{M}}\right]_\delta+\left(\frac{3}{2M_-}+\frac{1}{X_\delta}\right)P_{+-}-\left(\frac{3}{2M_+}+\frac{1}{X_\delta}\right)P_{++}}{-\sqrt{\frac{2}{X_\delta}}\left[\sqrt{\hat{M}}\right]_\delta-\left(\frac{3}{2M_-}+\frac{1}{X_\delta}\right)P_{--}+\left(\frac{3}{2M_+}+\frac{1}{X_\delta}\right)P_{++}}
\end{eqnarray}
and
\begin{eqnarray}
\bar{T}_\rightarrow = \frac{\left(\frac{3}{2M_-}-\frac{1}{X_\delta}\right)\left(P_{+-}-P_{--}\right)}{-\sqrt{\frac{2}{X_\delta}}\left[\sqrt{\hat{M}}\right]_\delta-\left(\frac{3}{2M_-}+\frac{1}{X_\delta}\right)P_{--}+\left(\frac{3}{2M_+}+\frac{1}{X_\delta}\right)P_{++}}.
\end{eqnarray}
\end{widetext}
Given the definition of probability current in this (more general) setting, we have
\begin{equation}
J^{(v)}_\pm = \left(\left(1+\frac{2M_\pm}{3 X_\delta}\right)\frac{P_{\pm\pm}}{M_\pm}-\frac{2}{3}\sqrt{\frac{2M_\pm}{X_\delta}}\right)\left|\Psi^{(v)}_\pm\right|^2.
\end{equation}
Just as in the flat spacetime limit, the final output states are given by (\ref{eq:FinalStates}), except that now the reflection/transmission amplitudes and the WKB phases are more complicated.
\begin{figure*}
\includegraphics[width=\textwidth,height=5.9cm]{Interference3B.jpg}
\caption{Sample interference pattern, for $M_+=0.05$, $v_+=0.0001$, $X_-\approx 1$, $L_- \approx 0.997$, and $M_-$ chosen to satisfy (\ref{eq:Mminus}), plotted against the outer mirror position, $X_+$.}
\label{fig:Interference3B}
\end{figure*}
The initial current can be expressed as
\begin{equation}
J_i = \left(1-\frac{2M_+}{3P_{-+}}\sqrt{\frac{2M_+}{X_\delta}}+\frac{2M_+}{3X_\delta}\right)J_0,
\end{equation}
with $J_0$ being the nonrelativistic initial current (\ref{eq:JInit}), and so the final reflection and transmission coefficients $R_f \equiv \frac{|J_+^{(v)}|}{|J_i|}$ and $T_f \equiv \frac{|J_-^{(v)}|}{|J_i|}$ (respectively) are fully determined.
Another similarity to the flat spacetime limit is that the oscillatory part of the final reflection and transmission coefficients is defined by $\varphi=\phi_{++}+\phi_{-+}-\phi_{+-}-\phi_{--}$, with the various $\phi$ terms involving integrals of the general relativistic momentum (\ref{eq:Pc}). In the weak-field limit, the initial ingoing momentum is given by
\begin{eqnarray}
P_{-+}\sim -\sqrt{H^2-M_+^2}+\frac{2}{3}\sqrt{\frac{2H}{X}}H\nonumber\\
-\frac{\left(H^2-M_+^2/2\right)H}{\sqrt{H^2-M_+^2}X},
\end{eqnarray}
to second order in $1/\sqrt{X}$. Care should be taken with these approximations, however, because our probability current (\ref{eq:JG}) is exactly conserved only in the quadratic momentum limit, which for the shell system is defined by (\ref{eq:WKBHam}). Also, the WKB solutions only approximately satisfy the Schr\"{o}dinger equation. Because of this, in order to control the errors involved in the approximations we find it useful to consider the ``WKB momentum,'' which we define by solving (\ref{eq:WKBHam}) for $P$, and expanding to second order in $1/\sqrt{X}$. For the initial ingoing momentum, the WKB momentum takes the form
\begin{eqnarray}
P_{w-+}\sim-\sqrt{2M_+\left(H-M_+\right)}+\frac{2}{3}\sqrt{\frac{2M_+}{X}}M_+\nonumber\\
-\sqrt{\frac{2M_+}{\left(H-M_+\right)}}\frac{\left(7M_+-4H\right)}{12 X}.
\end{eqnarray}
To understand the interference pattern described by $R_f$ and $T_f$, let us consider what $\varphi$ looks like in the weak-field limit, for slow speeds ($v_\pm \rightarrow 0$):
\begin{eqnarray}
\varphi\hspace{7pt} \sim &2L_+ M_+ v_+-2L_- M_- v_-\nonumber\\
&+\frac{M_+^2}{v_+}\ln{\left(\frac{X_+}{X_\delta}\right)}-\frac{M_-^2}{v_-}\ln{\left(\frac{X_\delta}{X_-}\right)}.
\label{eq:cancelGR}
\end{eqnarray}
Let us further imagine that we vary the outer arm length $L_+$, while keeping all other parameters constant. If the phase condition (\ref{eq:cancel}) from flat spacetime still approximately holds, then the corresponding expression (\ref{eq:cancelGR}) in the weak-field limit tells us that successive values of $n$ (say, $n$ to $n+1$) are associated with outer arm length values $L_{+n}$ and $L_{+(n+1)}$. Subtracting $\varphi_n=n\pi/2$ from $\varphi_{n+1}=(n+1)\pi/2$ yields
\begin{equation}
\frac{\pi}{2}=2M_+ v_+ \Delta L_n +\frac{M_+^2}{v_+}\ln{\left(\frac{X_{+n}+\Delta L_n}{X_{+n}}\right)},
\end{equation}
with the definitions $X_{+n}=X_\delta+L_{+n}$ and $\Delta L_n =L_{+(n+1)}-L_{+n}$. The distance between nodes of the interference pattern, denoted by $\Delta L_n$, is somewhat less than the outer mirror radius $X_+$, for the cases we are interested in; thus, we can expand the logarithm and solve for $\Delta L_n$, which gives us
\begin{equation}
\Delta L_n \approx \frac{\pi}{4M_+\left(v_++\frac{M_+}{2v_+X_{+n}}\right)}.
\label{eq:nodes}
\end{equation}
This result shows that gravity causes the node spacing in the interference pattern to increase with increasing outer arm length. In the flat space limit (i.e. as $X_\pm\rightarrow \infty$), we obtain the equal node spacing $\Delta L_n =\pi/4M_+v_+$, for all $n\in\mathbb{Z}$.
One can see from Figures~\ref{fig:Interference1B} and \ref{fig:Interference2B} that as we go from the essentially flat limit ($X_\pm \rightarrow \infty$) to less than $10$ Schwarzschild radii, we can still alternately cancel the reflection and transmission coefficients, even though the approximations lead to a probability current that is not fully conserved (note that the sum of the final probability currents is about $15\%$ less than the initial current). We take this as a direct indication that coherence is fully present in the single-mode system even with general relativistic corrections taken into account.
It is not clear from Figures~\ref{fig:Interference1B} and \ref{fig:Interference2B}, but the node spacing is indeed changing as (\ref{eq:nodes}) suggests. The reason it is not visible from these plots is that the node spacing changes noticeably only over a range of many wavelengths. Under more extreme circumstances, as depicted in Figure~\ref{fig:Interference3B}, there are visible changes in node spacing, though this represents a situation that is of less physical interest, since the de Broglie wavelength of the shell is larger than the interferometer arms.
\section{Discussion}
There are two problems with taking this result of no loss of coherence as the definitive answer to whether or not gravity, by itself, could decohere a system. The first is that these single-mode states correspond in some sense to energy eigenstates; one might expect it is only a superposition of energies that leads to decoherence, since from the above analysis one can see that the time-dependence cancels out of the final expressions for output probabilities in the interferometer. As mentioned above, a study of how wave-packets behave in this model will be presented in a forthcoming paper \cite{GU14}.
The second problem is that Penrose's intuition ties the loss of coherence to the inability to map one spacetime in any unique way onto a different spacetime. By our coordinate choice we have, in effect, chosen a unique way: two spacetime points are the same if they have the same coordinates. However, this is of course arbitrary and depends on the coordinate choice made. While the Painlev\'e-Gullstrand coordinates have many advantages, they are not the only possibile choice. A one-parameter family of generalized Painlev\'e-Gullstrand coordinates \cite{PainleveFamily} can also be used to perform analogous calculations to those above. Do all coordinate fixings produce the same maxima and minima in the interference pattern? The canonical momentum in the reduced system certainly depends on the coordinate choice, but one can show that a broad set of choices lead to the same classical action \cite{Menotti}. Still, it is unclear whether this is enough to ensure coordinate independence in the quantum setting. These issues will be examined in future work \cite{GQU14}.
While we do not report a result that demonstrates intrinsic decoherence due to gravity here, we have provided a model system for further analysis that could potentially lead to such a demonstration. This work can therefore be considered a first step towards a concrete derivation of Penrose's predictions within canonical quantum gravity. The main point argued here is that Penrose's initial arguments rely solely on the principles of QT and GR, and so we should thoroughly explore the possibility that his predicted decoherence could be shown to result solely from QT and GR before adding any assumptions about new physics.
It could also be argued that the reduced phase space approximation leads to an artificial form of time-evolution that is not entirely consistent with the ``timeless'' structure of canonical quantum gravity. For instance, the lack of a satisfactory interpretation of reduced phase space minisuperspace quantum cosmology was discussed in \cite{UnruhWald}. One might then be drawn to the conclusion that in the limited setting of our approximations, the evolution will necessarily be unitary (by construction), and we will escape Hawking's original arguments about pure states evolving into mixed states \cite{Hawking82} by virtue of our approximation scheme.
Certainly, our simple model does not have the features often associated with nonunitary modifications to standard Hamiltonian evolution (such as the inclusion of microscopic wormhole interactions \cite{Ellis89}), but there is still reason to believe the evolution defined by (\ref{eq:Pc}) could in principle exhibit decoherence. For one thing, we have in some places used an approximate Hamiltonian (\ref{eq:WKBHam}) that is quadratic in momenta and strictly Hermitian, but it may not be possible to define a Hermitian Hamiltonian operator that exactly corresponds to the solution of (\ref{eq:Pc}) (which is transcendental). For another, even if one could solve (\ref{eq:Pc}), the resulting Hamiltonian would be non-polynomial in both the momenta and the coordinate $X$. This means that the time evolution of the wavefunction at $X$ is not described by a finite number of derivatives at $X$, and is thus nonlocal, in the sense that the evolution equation is equivalent to an integro-differential equation with finitely-many derivatives \cite{Sucher63}-\cite{Garbaczewski14}. While some systems can be nonlocal in this way and yet maintain coherence (such as in the case of relativistic particles in flat spacetime \cite{Sucher63}, \cite{Lammerzahl92}), in other such systems there can be unexpected behaviour such as ``nonlocally-induced randomness'' \cite{Garbaczewski13}, \cite{Garbaczewski14}, which would in our case be attributable to gravity. These studies are still in their infancy, so it remains an open question whether or not this type of nonlocal behaviour can be connected with gravitational intrinsic decoherence.
Regardless of what the true theoretical mechanism is, the arguments for the existence of the decoherence effect studied here are compelling, and experimental investigations are already underway to test for signatures in micro-optomechanical systems \cite{PenroseBouwmeester}, \cite{PikovBouwm08}. There are many technical obstacles to overcome to minimize the effects of standard environmental decoherence (which obscures the desired behaviour), but there is hope that these types of experiments will bear fruit within the next decade \cite{Penrose14}.
\section{Acknowledgements}
The authors would like to thank the Natural Sciences and Engineering Research Council of Canada (NSERC) and the Templeton Foundation (Grant No. JTF $36838$) for financial support, as well as Friedemann Queisser for helpful discussions.
|
\section{Introduction}
Accurately determining the spiral structure of the Milky Way is a
difficult task. In principle, kinematic distances can be used to
construct the spiral structure of the Galaxy. However, recent work
on parallax and proper motion measurements has shown that kinematic
distances can be affected by large uncertainties. In some places of
the Galaxy, characterized by ``anomalous'' motions (as, for example,
the Perseus arm \citep{Xu:06a}), kinematic distances can be in error
by a factor as large as two. Relying solely on kinematic distances,
one cannot accurately determine the location of spiral arms. A more
secure method of distance measurement is required to re-construct
the Galactic spiral structure and its 3-D motion \citep{Reid:09a}.
Recently, \meth \ 12~GHz masers have been used as astrometric
targets to measure the trigonometric parallaxes and proper motions
of massive star-forming regions \citep[][hereafter called
Paper~I]{Reid:09b}. As part of that large project, here, we present
the results of our parallax measurement campaign toward the sources
\Ga,\ \Gb\ and \Gc.
\section{Observations and Calibration}
A series of observations of the 12~GHz \meth\ masers
in the \Ga, \Gb, and \Gc\ star-forming regions
were carried out with the NRAO\footnote{The National Radio Astronomy
Observatory is a facility of the National Science Foundation
operated under cooperative agreement by Associated Universities,
Inc.} Very Long Baseline Array (VLBA). Paper~I provides
a description of the general
observational setup and data calibration procedures, and
here we note only details specific to the observations of the three
sources presented in this paper.
\Ga\ and \Gb\ were observed at five epochs (VLBA program BR129A):
2007 October 18; April 17 and September 13 2008; March 23 and
October 23 2009. However, because of bad weather, the data quality
of the first epoch was not good enough for parallax measurement, and
the present results are based on the following four epochs only.
\Gc\ was observed at four epochs (VLBA program BR129C): 2007 October
27; April 24 and October 31 2008; 2009 April 16. The observing dates
have been chosen to sample the peaks of the parallax signature in
right ascension, as the amplitude of the parallax signature in
declination is considerably smaller.
For each maser source, we used three different background sources,
selected from the following calibrator surveys: the VLBI Exploration
of Radio Astrometry (VERA, \citet{Honma:01}) Galactic Plane Survey;
a VLA survey of compact NVSS sources \citep{Xu:06b}; the VCS2 and
VCS3 catalogs \citep{Fomalont:03, Petrov:05}. Only background
sources belonging to the VCS2 and VCS3 catalogs were detected. The
non-detected calibrators were J1815-1717 and J1837-0628 from the
VERA survey, and J1809-1804 and J1817-1614 from the VLA survey.
Table~\ref{table:positions} reports the peak positions and
intensities of the maser and background sources, listing also the
main observing parameters. We recorded four adjacent dual
circularly polarized 4~MHz bands with the maser signal in the second
band centered at the peak velocity of the 12~GHz maser emission,
corresponding to a LSR velocity, \Vlsr, of 40~\kms, 23~\kms\ and
100~\kms\ for source \Ga, \Gb\ and \Gc,\ respectively. The spectral
resolution was 0.38~\kms.
We used observations of the strong VLBA calibrators J1800$+$3848 and
3C345 (J1642+3948) to correct for instrumental delays and phase
offsets among different frequency bands. The spectral channel with
the strongest maser emission was used as the phase reference (see
Table 1). The maser reference features were detected at all epochs
and were relatively stable in flux, with the exception of the source
\Ga, whose correlated flux density on the shortest VLBA baselines
increased monotonically by a factor of $\approx$ 2, while
\citet{Goedhart:09} found a periodical variation with a period of
$\sim$30 days and flux variations of $\sim$50\%. After applying the
maser calibration, we integrated the data of the background
continuum sources over all four dual-polarized bands and imaged the
sources using the AIPS task IMAGR. For all maser targets, the
maser-referenced calibrator image has a dynamical range of about 10.
The naturally-weighted ``dirty'' beam was determined by the
availability of maser phase-reference data and, as expected for
low-declination sources, was strongly elongated close the N--S
direction. For each of the three maser targets,
Table~\ref{table:positions} reports the parameters of the ``dirty''
beam as derived by the AIPS task IMAGR by fitting an elliptical
Gaussian brightness distribution to the central portion of the
``dirty'' beam map. The images of the maser and corresponding
calibrators have been restored using a round "clean" beam, with a
FWHM size intermediate between the minor and major FWHM sizes of
the ``dirty'' beam. To make images from different epochs more
readily comparable, maps of a given source were restored using the
same beam.
After obtaining maps of both maser and background sources for all
epochs, we derived the parameters of the emission by fitting
elliptical Gaussian brightness distributions.
Table~\ref{table:positions} lists the position and peak intensity of the
maser and corresponding calibrator(s), as well as the angular
separation on the sky between the two sources. Absolute maser
positions are derived from the positions of the ICRF sources
J1825$-$1718 and J1834$-$0301, which are accurate within \
$\approx1$~mas.
\subsection{Methanol masers}
Emission in the 12.2 GHz \meth\ maser line generally is coextensive
with emission in the 6.7 GHz line -- maser spots at similar
velocities often arise from similar positions \citep{Menten:92}.
Since the 6.7 GHz emission is almost always (much) stronger than the
12.2 GHz emission, spectra in the former are much richer and more
maser spots are detected at 6.7 GHz \citep[see, e.g.
][]{Caswell1995}. For all our sources, the emission is dominated by
a single feature at the LSR velocity listed in
Table~\ref{table:positions}. The positions of any other weak
features (if present) were found to be close (withih a few mas) to
the strongest feature.
\section{Results}
Following the analysis described in Paper~I, the measured positions
of the masers were modeled as a linear combination of the elliptical
parallax and linear proper motion signatures. Because systematic
errors (owing to small uncompensated atmospheric delays and, in some
cases, varying maser and calibrator source structures) typically
dominate over thermal noise when measuring relative source
positions, we added ``error floors'' in quadrature to the formal
position uncertainties. We used different error floors for the Right
Ascension and Declination data and adjusted them to yield post-fit
residuals with $\chi^2$ per degree of freedom near unity for both
coordinates.
\subsection{\Ga} \label{PPM_g12}
Fig.~\ref{g12_masers} presents the images of the \Ga\ reference
maser channel at \Vlsr = 39.8~\kms\ and the background continuum
source J1825$-$1718 (phase referenced to the reference maser
channel) at the second epoch (2008 April 17). Both the maser and the
background source emission is dominated by a single compact
component, which serves as an astrometric target.
Fig.~\ref{g12_parallax} reports the positions of the emission in the
reference maser channel (relative to the background source
J1825$-$1718) as a function of time and the parallax fit. The large
declination errors of the parallax fit are probably caused by
residual phase errors due to the combination of the large
maser-calibrator separation (3\fdg3) and the low maser declination.
\begin{figure}
\includegraphics[angle=-90,scale=0.5]{g12+J1825_beam5}
\caption{The {\it left} and {\it right} panels present the map of
the \Ga\ reference maser channel (\Vlsr = 39.8~\kms)
and the background source J1825$-$1718, phase-referenced
to the reference maser channel, respectively.
Both maps are for the epoch 2008 April 17.
Contour levels are at integer multiples (with the zero contour suppressed)
of 10\% of the peak brightness of
2.2~Jy~beam$^{-1}$ for \Ga\ and 0.14~Jy~beam$^{-1}$ for
J1825$-$1718. The FWHM size of the restoring beam is given in the
lower left corner of each panel. \label{g12_masers}}
\end{figure}
\begin{figure}
\includegraphics[angle=-90,scale=0.67]{G12_3planes_new.eps}
\caption{Parallax and proper motion data and fits for \Ga. Plotted
are position offsets of the \Ga\ maser spot at \Vlsr = 39.8~\kms\
relative to the background source J1825$-$1718. {\it Left
Panel:} Positions on the sky with first and last epochs labeled. The
expected positions from the parallax and proper motion fit are
indicated {\it (empty circles)}.
{\it Middle Panel:} The position offsets of the maser along the East
and North direction versus time. The best-fit model of the variation of the
East and North offsets with time is shown as
{\it continuous} and {\it dashed} lines, respectively.
{\it Right Panel:}
Same as the {\it middle panel}, except the best fit proper motions
have been removed, allowing the effects
of only the parallax to be seen.
\label{g12_parallax}}
\end{figure}
Fitting for the parallax and proper motion simultaneously, we obtain
$\pi=0.428 \pm 0.022$~mas. The proper motions in the eastward and
northward directions are $0.16\pm0.03$ and $-1.90
\pm1.59$~mas~y$^{-1}$, respectively (see Table~\ref{table:allfits}).
This parallax corresponds to a distance of
$2.34^{+0.13}_{-0.11}$~kpc. Based on this distance, \Ga\ is most
likely in the Carina-Sagittarius arm of the Milky Way, and not in
the Crux-Scutum arm.
\subsubsection{Maser environment}
\Ga, better known as IRAS~18089$-$1732, has long be known as a
prominent maser source, showing strong emission from all the
widespread interstellar maser molecules (OH, H$_2$O, and \meth). It
was included in the sample of high mass protostellar objects (HMPOs)
defined by \citet{Sridharan:02} and further studied by \citet[][ and
references therein]{Beuther:02a}. Except for a single spot (the
component~F) detached by 1$\arcsec$.4, all of the 6.7 GHz \meth\
maser spots detected by \citet{Walsh:98} coincide with our VLBA
maser position within the errors ($\sim$1$\arcsec$) of the Australia
Telescope Compact Array (ATCA). Our VLBA maser emission is most
likely corresponding with the component~A of \citet{Walsh:98}, which
shows a similar \Vlsr = 39.2~\kms. H$_2$O maser emission has been
found with the VLA $\approx 1\as2$ to the NNE \citep{Beuther:02b}.
Interestingly, the H$_2$O maser position coincides with weak
cm-wavelength radio continuum emission and a (sub)mm dust emission
peak, while the \meth\ maser is offset from these peaks
\citep{Beuther2004}. The masers and continuum emission are
associated with a rich molecular hot core, which has been
interpreted as a rotating disk \citep{Beuther2005}.
\subsection{\Gb} \label{PPM_G15}
Fig.~\ref{g15_masers} presents the images of the \Gb\ reference
maser channel (\Vlsr = 23.4~\kms) and the background continuum
source J1825$-$1718 (phase referenced to the reference maser
channel), at the second epoch (2008 April 17).
Fig.~\ref{g15_parallax} reports the positions of emission in the
reference maser channel (relative to the background source
J1825$-$1718) as a function of time and the parallax fit.
\begin{figure}
\includegraphics[angle=-90,scale=0.5]{g15+J1825_beam5}
\caption{The {\it left} and {\it right} panels present maps of the
\Gb\ reference maser channel (\Vlsr = 23.4~\kms) and the background
source J1825$-$1718, phase-referenced to the reference maser
channel, respectively. Both maps are for 2008 April 17. Contour
levels are integer multiples (with zero contours suppressed) of 10\%
of the peak brightness of 2.9~Jy~beam$^{-1}$ for \Gb\ and
0.14~Jy~beam$^{-1}$ for J1825$-$1718. The FWHM size of the restoring
beams is given in the lower left corner of each panel.
\label{g15_masers}}
\end{figure}
\begin{figure}
\includegraphics[angle=-90,scale=0.67]{G15_3plans.eps}
\caption{Parallax and proper motion data and fits for \Gb. Plotted
are position offsets of the \Gb\ maser spot at \Vlsr = 23.4~\kms\
relative to the background source \ J1825$-$1718. {\it Left
Panel:} Positions on the sky with first and last epochs labeled. The
expected positions from the parallax and proper motion fit are
indicated {\it (empty circles)}.
{\it Middle Panel:} The position offsets of the maser along the East
and North direction versus time. The best-fit model of the variation of the
East and North offsets with time is shown as
{\it continuous} and {\it dashed} lines, respectively.
{\it Right Panel:}
Same as the {\it middle panel}, except the best fit proper motions
have been removed, allowing the effects
of only the parallax to be seen.
\label{g15_parallax}}
\end{figure}
Fitting for the parallax and proper motion simultaneously, we obtain
$\pi = 0.505\pm0.033$~mas, corresponding to a distance of \
$1.98^{+0.14}_{-0.12}$~kpc. At this distance, \Gb\ is likely in the
Carina-Sagittarius arm. The proper motions in the eastward and
northward directions are $0.68\pm0.05$ and \ $-1.42
\pm0.09$~mas~y$^{-1}$, respectively (see Table~\ref{table:allfits}).
\subsubsection{Maser environment}
The maser position coincides with the ``unusual radio point source
in M17'' found by \citet{Felli:80}, one of the first ultracompact
HII regions ever identified as such. The position of the 6.7 GHz
maser emission, as recently determined by \citet{Caswell:09} using
ATCA, is consistent within the ATCA measurement error ($\approx
1''$) with that we determine for the 12.2 GHz line.
\subsection{\Gc} \label{PPM_G27}
Fig.~\ref{g27_masers} presents the images of the \Gc\ reference maser
channel (\Vlsr = 99.6~\kms) and the two background continuum sources
J1834$-$0301 and J1846$-$0651 (phase referenced to the reference maser channel)
at the second epoch (2008 April 24). Fig.~\ref{g27_parallax} reports
the positions of the reference maser channel (relative to the
background sources \ J1834$-$0301) as a function of time and the
parallax fit.
For this source, the derived value of parallax has a large
fractional uncertainty. Using only data from the \ J1834$-$0301 \
calibrator, we obtain \ $\pi = 0.125\pm0.042$~mas, while using the
\ J1846$-$0651 \ data, we find \ $\pi = 0.198\pm0.133$~mas.
Since the result from the \ J1834$-$0301 \ calibrator appears
significantly more precise than that obtained from
J1846$-$0651, we take the former as the best parallax
measurement for \Gc, corresponding to a distance of
$8.0^{+4.0}_{-2.0}$~kpc. The proper motions in the eastward and
northward directions are $-1.81\pm0.08$ and
$-4.11\pm0.26$~mas~y$^{-1}$, respectively (see
Table~\ref{table:allfits}).
\begin{figure}
\includegraphics[angle=-90,scale=0.5]{G27+J1834+J1846_r}
\caption{The panels from {\it left} to {\it right} present the map
of the \Gc\ reference maser channel (\Vlsr = 99.6~\kms) and
the background sources J1834$-$0301 and \ J1846$-$0651,
phase-referenced to the reference maser channel.
The maps are for 2008 April 24.
Contour levels are integer multiples (with the zero contour suppressed)
of 10\% of the peak brightness of 2.6~Jy~beam$^{-1}$ for \Gc, and
0.07~and~0.04~Jy~beam$^{-1}$ for J1834$-$0301 and \ J1846$-$0651,
respectively. The FWHM size of the restoring beams is given in the
lower left corner of each panel. \label{g27_masers}}
\end{figure}
\begin{figure}
\includegraphics[angle=-90,scale=0.67]{G27_1834_3plane}
\caption{Parallax and proper motion data and fits for \Gc. Plotted
are position offsets of the \Gc\ maser spot at \ \Vlsr = 99.6~\kms
relative to the background source \ J1834$-$0301. {\it Left Panel:}
Positions on the sky with first and last epochs labeled. The
expected positions from the parallax and proper motion fit are
indicated {\it (empty circles)}.
{\it Middle Panel:} The position offsets of the maser along the East
and North direction versus time. The best-fit model of the variation of the
East and North offsets with time is shown as
{\it continuous} and {\it dashed} lines, respectively.
{\it Right Panel:}
Same as the {\it middle panel}, except for the fact that the best fit proper motions
have been removed, allowing the effects
of only the parallax to be seen. \label{g27_parallax}}
\end{figure}
\subsubsection{Maser environment}
The 12.2 GHz \meth\ maser in the \Gc\ region is offset by about
$2''$ from a compact $< 2''$ radio source found by
\citet{Becker:94}, which also has an associated IRAS source
(18391$-$0504). Much stronger emission than in this line has been
found in the 6.7 GHz \meth\ line \citep{Szymczak:02}, which has associated maser
emission from OH and H$_2$O \citep{Szymczak:04,Szymczak:05}. This
source appears to have been little studied in thermal molecular line
emission.
\section{Galactic Locations and 3-D Motions}
Combining the distances, LSR velocities and proper motions of the
masers yields their locations in the Galaxy and their full space
motions. Since internal motions of 12~GHz methanol masers are fairly
small, typically $\sim3$~\kms \citep{Mos02}, the maser motions
should be close to that of their associated young stars. Given a
model for the scale and rotation of the Milky Way, we can subtract
the effects of Galactic rotation and the peculiar motion of the Sun
from the space motions of the maser sources and estimate the
peculiar motions of the associated young stars. Following the
discussion by \citet{Reid:09a}, the motion of an individual massive
star (associated with the masers) can then be taken as
representative of that of the whole star-forming region to which the
massive star belongs, allowing for a velocity dispersion of
individual stars of $\approx 7$~\kms\ per each velocity coordinate.
Peculiar motions are given in a Galactocentric reference frame,
where $U, V \, {\rm and} \, W$ are the velocity components toward
the Galactic center, in the direction of Galactic rotation, and
toward the North Galactic Pole, respectively, at the location of a
given source in the Galaxy. We adopt the IAU values for the distance
to the Galactic center ($R_0=8.5$~kpc) and the rotation speed of the
Galaxy at this distance ($\Theta_0=220$~\kms), and assumes a flat
rotation curve. We use the Solar Motion value ($U = 11.1
^{+0.69}_{-0.75}$, $V = 12.24 ^{+0.47}_{-0.47}$, and $W = 7.25
^{+0.37}_{-0.36}$~\kms) by \citet{Schonrich:10}, which have recently
revised the Hipparcos satellite result. Table~\ref{table:3dmotion}
reports the peculiar motions derived for our maser targets. For
comparison, the results obtained with a different model of Galactic
rotation ($R_0 = 8.3 \pm 0.23$~kpc, $\Theta_0 = 239 \pm 7 $~\kms),
based on both the weighted average of recent four direct
measurements of the distance to the Galactic center and the proper
motion of Sgr A$^{\ast}$ \citep{Brunthale2011}, are also reported.
Looking at Table~\ref{table:3dmotion}, one can see that, except for
\Gc\, whose space motion may be strongly affected by the large
parallax uncertainty, the derived peculiar motions are fairly
independent on the adopted Galactic rotation model. For all sources,
the motion directed toward the north Galactic pole is only of a few
\kms. This is consistent with the expectation for the motion of
massive star-forming regions, which should be mainly in the Galactic
plane.
\section{Conclusions}
We have measured the parallax and proper motion of 12~GHz methanol
masers in three star-forming regions. For two sources, the derived
distances are accurate by better than 10\%:
$2.34^{+0.13}_{-0.11}$~kpc for \Ga\ and $1.98^{+0.14}_{-0.12}$~kpc
for \Gb. For the source \Gc, the derived distance is affected by a large
uncertainty: $8.0^{+4.0}_{-2.0}$.
Our precise absolute positions place the methanol masers near the center
of active regions of high mass star formation, as traced by molecular hot cores,
ultracompact HII regions and/or dust condensations or a combination of these.
\acknowledgments This work was supported by the Chinese NSF through
grants NSF 11073054, NSF 10733030, NSF 10703010 and NSF 10621303,
and NBRPC (973 Program) under grant 2007CB815403.
\vskip 0.5truecm
{\it Facilities:} \facility{VLBA}
|
\section{Flavor of flavor models}
Certain features of the leptonic mixing can be considered as an evidence of
discrete symmetry. Many different realizations
exist \cite{reviews}. However, in spite of a various interesting developments,
no convincing model based on discrete symmetries
has been proposed so far. The models require new extended structures,
many assumptions, {\it ad hoc} assignments of charges and selection of the group
representations for multiplets.
Additional auxiliary symmetries are needed which
sometime even more powerful the original one.
There are no natural and simple extensions of the leptonic symmetries
to the quark sector. In most of the models no connection
between mixing and masses exists and different physics (symmetry)
is responsible for the mass hierarchies.
So, what to do? Try further using the same
context~\footnote{In attempt to further pursue
this approach and to check whether something interesting
is overlooked, systematic scanning (including computerized one) of all
possible models within certain framework has been performed~\cite{scan}.}.
Be less ambitious and explain only some features (e.g., dominant
structures of the mass matrices) using the symmetry?
Or modify context (some important elements can be still missed).
Apply symmetry differently?
For illustration of these statements,
several recently proposed models will be
discussed. Generic features of the whole approach and its challenges
are formulated. For details of specific models see talks~\cite{others} at this conference.
The paper\footnote{Talk given at the Symposium on
``DISCRETE 2010'', 6 - 11 December 2010, La Sapienza, Rome, Italy}
is organized as follows.
In sect. 2 the experimental evidences in favor of discrete symmetries
will be presented. Sect. 3 is devoted to the Tri-Bimaximal (TBM)
mixing and ``symmetry building''.
In sect. 4 attempts to extend the leptonic symmetry to the quark sector are discussed.
Sect. 5 is devoted to alternatives to the TBM approach, in particular, to
the quark-lepton complementarity (QLC). Sect. 6 outlines perspectives in the field,
and conclusions are given in Sect. 7.
\section{Data and evidences}
The origin of excitement is the neutrino mass
and flavor spectrum shown in fig.~\ref{fig:spec}.
Regularities of the flavors distribution are obvious:
$\nu_\mu$ and $\nu_\tau$ flavors share $\nu_3$ equally
(bi-maximal mixing) and $\nu_e$ is absent in $\nu_3$,
all three flavors are presented
with the same weight in $\nu_2$ (trimaximal mixing). This can be formalized
as invariance of the picture with respect to the following
transformations:
(1) permutation of $\nu_\mu$ and $\nu_\tau$ flavors in all three mass states,
which also implies zero 1-3 mixing; (2) dilatation of the $\nu_e-$flavor part by factor 2 and
shrinking by 2 the rest of the state $\nu_2$,
then inverse operation in $\nu_1$: shrinking by factor 2 the $\nu_e-$ part, and
dilatation of the rest, and finally, permutation of
states $\nu_1$ and $\nu_2$. These transformations are the basis of possible discrete symmetry.
\begin{figure}[h]
\includegraphics[width=21pc]{spectrum.eps}\hspace{2pc}%
\begin{minipage}
[b]{13pc}\caption{\label{fig:spec}
Mass and flavor spectrum of neutrinos in the case of TBM mixing (left) and best fit
experimental values (right). The lengths of the colored parts are proportional to
the moduli squared of the mixing matrix elements, $|U_{\alpha
i}|$. }
\end{minipage}
\end{figure}
The regularities are described by the TBM mixing matrix~\cite{tbm}:
\begin{eqnarray}
U_{TBM} = U_{23}(\pi/4) U_{12}({\rm arcsin}(1/\sqrt{3})) = \left(
\begin{array}{ccc}
\sqrt\frac{2}{3} & \frac{1}{\sqrt3} & 0 \\
-\frac{1}{\sqrt6} & \frac{1}{\sqrt3} & -\frac{1}{\sqrt2} \\
-\frac{1}{\sqrt6} & \frac{1}{\sqrt3} & \frac{1}{\sqrt2}\\
\end{array}
\right).
\label{tbmmixing}
\end{eqnarray}
If the length of boxes is 1, the experimental data permit deviations
from the pattern shown in the left figure by amount 0.01 - 0.05.
The global fit of the oscillation data \cite{fit} agrees with the TBM mixing
within $(1 - 2) \sigma$. It shows non-zero best fit value for the 1-3 mixing,
some deviation of the 2-3 mixing
from maximal one with $\theta_{23} < \pi/4$, and
the 1-2 mixing slightly below $\sin^2 \theta_{12} = 1/3$:
\begin{equation}
\sin \theta_{13} = 0.118 ^{+0.038}_{-0.048}, ~~~
\sin^2 \theta_{23} = 0.463 ^{+ 0.071}_{-0.048}, ~~~
\sin^2 \theta_{12} = 0.321 ^{+ 0.016}_{-0.016}
\end{equation}
(with $1\sigma$ errors). The deviations from the TBM values
are small but robust appearing in the global analyses
of different groups. Important feature is that the 1-2 and 1-3 mixing angles
correlate in the data,
and therefore should be considered simultaneously.
The combined analyses of the solar and KamLaND results \cite{sno} \cite{kamland}
\cite{sk-sol} give non-zero 1-3 mixing at the level
$\sin \theta_{13} = 0.02$, $\sin^2 \theta_{12} < 0.33$,
and agreement with TBM is at $95 \%$ CL..
At the same time, the complete
$3\nu-$ analysis of the atmospheric neutrino data from SuperKamiokande
results in $\sin2 \theta_{13} = 0.0 ~( < 0.04 )~ 90\%$,
and $\sin2 \theta_{23} = 0.5~~(0.407 - 0.583)~ 90\%$ \cite{sk3}.
That is, the best fit value of the 2-3 mixing is maximal in the case
of normal mass hierarchy, in spite of the excess of $e-$like sub-GeV events
which gave an indication for $\theta_{23} < \pi/4$.
Several reservations are in order.
1. Large deviations from the symmetry case are still possible
and if correct (from the fundamental physics point of view)
measure of mixing is $\sin \theta$, then for
1-3 mixing we have $\sin \theta_{13} = (0.10 - 0.15)$ as the best fit
value which should be compared with $\sin \theta_{12} =
0.55$.
2. Mixing is not the RGE invariant, and
if theory is formulated at some high mass scales (e.g., GUT scale),
one may expect significant deviations from the symmetry case when running to
high energies.
3. There are several indications that new light ($m_s \sim 1$ eV$^2$)
and almost sterile neutrino state exists which mixes substantially
($\theta_{as} \sim 0.2$) with the active neutrinos. Since $m_s \sin^2 \theta_{as} \sim 0.04$ eV
is bigger than the largest element of the mass matrix
of active neutrinos (in the case of mass hierarchy),
the presence of this state (or states) destroys the present constructions
unless further complications and fine tunings are introduced.
4. Existence of the 4th generation of fermions can change
the present results completely.
All this may lead to different line of developments.
Concerning the mass hierarchy, there are several important features
which can be relevant for, but usually
not addressed in the models under consideration:
(1) The upper quarks have geometrical mass relation:
$m_u m_t = m_c^2$; this hints that masses of all three generation should be considered
on the same footing (and not in perturbative manner).
(2) The down quark masses satisfy the Gatto-Sartori-Tonin relation
$\sin \theta_C \approx \sqrt{m_d/m_s}$ \cite{gatto} -- explicit
relation between masses and mixing.
(3) The charged leptons satisfy the Koide relation \cite{koide} which
indicates certain symmetry, and again, an involvement of all three generations
of leptons. (4) Neutrinos have the weakest mass hierarchy among fermions which also
shows connection between masses and mixing.
\section{TBM and symmetry building}
To a large extend relevant symmetry can be systematically
derived from the data \cite{reviews}, \cite{lam} and this gives the main support of the
approach ``Mixing from discrete symmetries''.
\subsection{Deriving symmetry}
1. In assumption that neutrinos are the Majorana particles
the TBM mass matrix is given by $m_{TBM} = U_{TBM} m^{diag}U_{TBM}^T$, where
$ m^{diag} \equiv diag (m_1, m_2, m_3)$. In the flavor basis explicitly:
\begin{eqnarray}
m_{TBM} =
\left(
\begin{array}{ccc}
a & b & b \\
... & \frac{1}{2}(a+b+c) & \frac{1}{2}(a+b-c) \\
... & ... & \frac{1}{2}(a+b+c) \\
\end{array}
\right),
\label{mTBM}
\end{eqnarray}
\begin{equation}
a = \frac{1}{3} (2m_1 + m_2), ~~~
b = \frac{1}{3}(m_2 - m_1), ~~~c = m_3.
\end{equation}
Immediately one observes the $S_2$ symmetry of
the $\nu_\mu \leftrightarrow \nu_\tau$ permutations.
This symmetry plays a crucial role:
it is this symmetry of the dominant structure of the mass matrix
that ensures maximal
2-3 mixing and zero (small) 1-3 mixing which are robust features
of the lepton mixing. It fixes two out of three
angles, and probably, should be a starting
point of the symmetry building.
The TBM-symmetry can be expressed in the form of the
TBM-relations:
\begin{equation}
m_{e\mu} = m_{e \tau}, ~~~ m_{\mu \mu} = m_{\tau \tau}, ~~~
m_{e e} + m_{e \mu} = m_{\mu \mu} + m_{\mu \tau},
\label{third}
\end{equation}
or $\sum_\alpha m_{e\alpha} = \sum_\beta m_{\mu\beta}$ instead of
the last equality. Notice that in general, fixing three mixing angles
leads to three relations between the matrix elements.
What is non-trivial is that
the relations (\ref{third}) are simple and of particular type which
corresponds to certain symmetry.
(In general relations are complicated and no symmetry
can be found).
2. The TBM mass matrix (\ref{mTBM}) in the flavor basis
is invariant under transformations
\begin{equation}
V_i m_{TBM} V^T_i = m_{TBM},
\nonumber
\end{equation}
where
\begin{equation}
V_1 = S =
\frac{1}{3} \left(
\begin{array}{lll}
-1 & ~~ 2 & ~~ 2 \\
~ ... & - 1 & ~~ 2 \\
~ ... & ... & - 1
\end{array}
\right), ~~~
V_2 = U =
\left(
\begin{array}{lll}
1 & 0 & 0 \\
... & 0 & 1 \\
... & ... & 0
\end{array}
\right).
\label{v1v2}
\end{equation}
Two transformations (\ref{v1v2}) uniquely determine the form of the TBM mass matrix.
They do not depend on the mass spectrum, and in fact,
can be obtained immediately from the $U_{TBM}$.
At the same time, diagonality of the mass matrix squared of the charged leptons,
$m_e^{\dagger} m_e$, can be supported by
symmetry
\begin{equation}
V_3^\dagger (m_e^\dagger m_e) V_3 = m_e^\dagger m_e, ~~~~
V_3^T = diag(1, e^{i\alpha}, e^{i\beta}),
\end{equation}
where $\alpha \neq \beta \neq \pi k$.
The main idea is that $V_i$ are the generating elements
of certain discrete symmetry group $G_f$ which eventually determines mixing.
For instance, selecting $V_3^T = T \equiv diag(1, \omega, \omega^2)$,
where $\omega \equiv e^{i2\pi/3}$, one finds that
$S, T, U$ are the generating elements of the group $S_4$.
Furthermore, it was argued that $S_4$ group
is minimal structure which leads to TBM \cite{lam} (see
discussion in \cite{grimus}).
(Our consideration was in the flavor basis, however neither
mixing matrix nor symmetry depend on the basis once the change of basis is described
by transformation $V$ such that $V V^T = I$.)
3. The flavor symmetry $G_f$ (which contains $V_i$) should be broken.
(In fact, no exact flavor symmetry can be introduced in whole theory, see \cite{grimus2}).)
Neutrino mass matrix is not invariant under $T$, whereas
the charge lepton mass matrix is not invariant with respect to $S$ and $U$.
The idea is that symmetry $G_f$ is broken {\it differently and partially}
in the sectors which generate the neutrino masses and charged lepton masses.
Namely,
\begin{equation}
G_f \rightarrow {\rm breaking} \rightarrow \left\{
\begin{array}{lll}
G_\nu & (S, U) & {\rm neutrinos}\\
G_l & (T) & {\rm charged~~ leptons}
\end{array}
\right. .
\end{equation}
The {\it residual} symmetries $G_\nu$ and $G_l$
in the neutrino and charged lepton sectors are different.
Clearly $G_f$ is broken completely in whole theory.
Furthermore, the two sectors communicated in high orders,
and therefore $G_\nu$ and $G_l$ are broken
even in their own sectors being approximate.
As a result, the TBM symmetry is broken and the TBM mixing gets corrections.
Thus, the mixing appears as a result of different ways of the flavor
symmetry breaking in the neutrino and charge lepton sectors.
In turn, the difference of symmetry breaking can originate
(1) from different flavor assignments of
the right handed (RH) components of $N^c$ and $l^c$,
or/and (2) from Majorana nature of the neutrino mass terms
(absence of $N^c$): the neutrino and the charged lepton mass terms
(originate from $LL$ and $L l^c$ correspondingly)
can have different flavor properties.
If the symmetry $G_f$ is broken spontaneously, one should introduce different sets of
the Higgs (flavon) fields for neutrinos and charged leptons.
Then the form of the elements of residual symmetries, $S, U$ and $T$,
determines configurations of VEV's.
The items 1- 3 is a paradigm of the present day flavor model building.
4. It is possible to proceed even further in some particular way.
The TBM mass matrix can be presented as the sum of
three singular matrices. In the limit of small $m_1$
it becomes as
\begin{equation}
m_{TBM} \approx
A \left(
\begin{array}{l}
1 \\
1 \\
1
\end{array}
\right) \times (1, 1, 1) +
B \left(
\begin{array}{l}
0 \\
1 \\
- 1
\end{array}
\right) \times (0, 1, - 1).
\end{equation}
This form indicates the see-saw mechanism, for which the mass terms
may have the form
\begin{equation}
\sum_i \frac{1}{M_i} (L f_i) (L f_i)^T ,
\label{mass}
\end{equation}
where $M_i$ are the masses of RH neutrinos and $f_i$ are the triplets of flavon fields,
(see e.g. \cite{king}).
The VEV's of the triplets should be $\langle f_2\rangle (1, 1, 1)$ and
$\langle f_3 \rangle (0, 1, - 1)$, which can be obtained
in SUSY version of model.
So far so good. Problems appear when we
start to realize this program in consistent gauge models.
\subsection{Problems}
Generic problem originates from the fact that masses of fermions are given by
\begin{equation}
m = F(\{Y \}, \{ v \}),
\end{equation}
where $F$ is certain functional which depends on the mechanism
of neutrino mass generation. In general, it describes several
different contributions and includes various corrections.
$\{ Y \}$ are the Yukawa couplings or coupling constants
of different operators generating masses, and
$\{ v \}$ refer to a set of vev's of Higgs bosons.
$\{ Y \}$ and $\{ v \}$ follow from independent sectors:
from the Yukawa sector and the scalar potential. Yukawa couplings
are determined by symmetry (at least partially), whereas VEV's are fixed by pattern
of symmetry violation. In general, symmetry does not control how it is broken
and new ingredients (dynamics, symmetries) should be introduced to fix the VEV alignment.
To obtain TBM all these components should be correlated.
One step constructions do not work.
Essentially this means that TBM is not immediate consequence of symmetry but
a result of interplay of different factors, and in this sense -- accidental.
Another generic problem is related to the fact that
symmetry should be broken differently in the neutrino
and charged lepton Yukawa sectors. That is, different Higgs flavon multiplets
(typically - several for each sector) should be used.
To forbid unwanted couplings of these flavons one is
forced to introduce additional symmetries with {\it ad hoc} charge prescription.
\subsection{Flavons versus Flavored Higgses}
There are two types of models with
broken flavor symmetries:
1. Models with flavons $f$, singlets of the
gauge symmetry group which have non-zero ``flavor charges''.
In these models
usual Higgs doublet(s) $H$ are the flavor singlets, so that
the electroweak symmetry and flavor symmetry breakings
are separated.
The Yukawa couplings and mass terms are generated by the non-renormalizable
interactions:
\begin{equation}
\frac{1}{\Lambda^{n}_f} L e^c H f^{n} .
\label{eq:highdim}
\end{equation}
Here $\Lambda_f$ is the scale of flavor physics which can be above
the GUT scale. Clearly it is difficult to test such a scenario directly,
unless $\Lambda_f$ is not far above the electroweak scale.
Typical problem of this scenario is existence of high dimension operators
of the type (\ref{eq:highdim}) which contribute to masses.
Convergence of series is weak, $\langle f \rangle/ \Lambda_f \sim 0.2 - 0.5$,
especially if quarks (top quark) are included in consideration.
Further complications (e.g. additional symmetries) are needed to
control their effect.
2. Models with flavored Higgses:
The Higgs doublets carry flavor charges, and usually
large number of such doublets (which form flavor multiplets)
should be introduced. The flavor symmetry is broken
simultaneously with gauge symmetry at the EW scale.
These models are testable, and in fact, strongly restricted
by the FCNC, anomalous magnetic moment of muon, {\it etc.}.
One expects to see many scalar bosons at LHC \cite{multi}.
\subsection{Symmetry groups}
The Table 1 presents the list of small groups with
irreducible representation ${\bf 3}$, which are used
in the TBM model building.
Representation ${\bf 3}$ explains existence of three generations of fermions,
however all these groups have also singlets (and some - doublets).
There is no explanation why (non-trivial) singlet and doublet representations are missing.
An alternative is groups, like $S_3$, with non-trivial irreducible
representation ${\bf 2}$, so that the family structure appears as ${\bf 2} + {\bf 1}$ \cite{s333}.
In the Table 1 we give the order of group (number of elements),
irreducible representations and products of the representations
which contain invariants. The latter determines the flavor structure of a model.
\begin{table}[h]
\caption{\label{groups}The simplest groups with irreducible representations $\bf
3$.}
\begin{center}
\begin{tabular}{llll}
\br
group & order & representations & invariants \\
\mr
$A_4$ & 12 & $1,~ 1',~ 1'',~ 3$ & $3\times 3,~ 1'\times 1''$ \\
$T'$ & 24 & $1,~ 1',~ 1'',~ 2,~ 2',~ 2'',~ 3$ & $3\times 3,~ 1'\times 1'',~2\times$ \\
$S_4$ & 24 & $1,~ 1',~ 2,~ 3,~ 3'$ & $3\times 3,~ 3' \times 3',~ 2\times 2, ~ 1'\times 1'$ \\
$T_7$ & 21 & $1,~ 1',~ 1'',~ 3,~ 3^*$ & $3 \times 3^*$, $1'\times 1''$ \\
$\Delta (27)$ & 27 & $1_1 - 1_9$, ~ $3, ~ 3'$ & $3 \times 3'$, $1_2 \times 1_3$, $1_4 \times
1_7$, $1_5 \times 1_8$, $1_6 \times 1_9$ \\
\br
\end{tabular}
\end{center}
\end{table}
In what follows we present structures of different models based on
various discrete symmetries. The corresponding figures show explicitly {\it ad hoc}
character of selection of the flavon multiplets and prescription
of the flavor charges.
An open issue is ``missing'' representations: not all possible
low dimensional representations are used. This may create problem
if discrete symmetry originates from breaking of
some gauged continuous group \cite{luhn}.
For each model we indicate origins of mixing
and shortcomings with criteria based on
existence of auxiliary symmetries,
presence of new fields, possibility of extension
to the quarks sector and further embedding, {\it etc.}.
\subsection{$A_4$ symmetry and a simplest $A_4$ models}
The most popular group is $A_4$: the symmetry group of
even permutations of 4 elements, or symmetry of the tetrahedron~\cite{ma}.
It has order 12 and two generating elements $S$ and $T$ which are needed
to realize the TBM mixing.
The presentation of the group:
\begin{equation}
S^2 = 1, ~~~ T^3 = 1, ~~~ (ST)^3 = 1.
\label{repa4}
\end{equation}
The most important element, $U = A_{\mu \tau}$, is absent.
So, one needs either to introduce this permutation symmetry
in addition, thus extending the symmetry group,
or obtain it as an accidental symmetry: as a result of
particular selection of representation and the VEV alignment.
The flavor structure is determined
by properties of products of representations and invariants:
\begin{equation}
3 \times 3 = 3 + 3 + 1 + 1' + 1'', ~~~ 1' \times 1'' = 1
\end{equation}
(and of course, by the VEV alignment).
\begin{figure}[h]
\includegraphics[width=22pc]{rome10a4a.ps}\hspace{2pc}%
\begin{minipage}
[b]{12pc}\caption{\label{a4am}
Scheme of the lepton mass generation in the $A_4$ model \cite{alt-mel}.
Colors indicate representations of the flavor symmetry group ($A_4$), numbers
at the multiplets give transformation properties with respect to additional group ($Z_4$);
the lines show couplings. The closed loops correspond to self-coupling of a given multiplet,
e.g. $N^c N^c$.}
\end{minipage}
\end{figure}
The structure of a simplest $A4$ model for TBM mixing \cite{alt-mel}
with charge assignment is presented in fig.~\ref{a4am}.
The key features of the model include the following.
There are three RH neutrinos, $N^c$.
Four flavon multiplets participate in generation of masses.
The lepton mixing appears due to different
flavor assignments of the RH neutrinos and the charged leptons:
$N^c$ form triplet $N^c \sim {\bf 3}$, whereas $l^c$
are all singlets, ${\bf 1}$, of $A_4$.
The charged leptons get Dirac masses via non-renormalizable terms,
whereas neutrinos - via renormalizable ones.
The $U-$symmetry is accidental: due to particular selection of the
flavon representations and configuration of VEV's.
The auxiliary symmetry $Z_4$ with {\it ad hoc} prescribed charges is introduced;
in particular, all $N^c$ have the same
whereas $l^c$ have all different $Z_4$ prescriptions.
The vacuum alignment is achieved by using SUSY and additional driving fields.
The lepton mixing follows to a large extend from structure
of mass matrix of the RH neutrinos.
The model does not admit simple SO(10) embedding as well as the SU(5)
embedding (if quarks also form family symmetry):
$l^c$ are singlets of the flavor symmetry group and
the rest fermions form triplets.
To resolve this problem one can introduce additional
GUT matter multiplets locating in the same multiplets
known fermions together with new matter fields.
$A_4$ singlet representations of $L$ are missing;
the representation ${\bf 1''}$ is not used
neither for matter fields nor for flavons. Apparently introduction of these missing
multiplets will lead to new problems which could probably be
cured by further complications of model.
\subsection{Mixing and masses}
In majority of the models mixing and masses are unrelated or have indirect
relations. The latter may appear as a result of certain
choice of the Higgs representation within a given symmetry context.
Mixing follows from certain form invariance of the mass matrix.
Usually, additional $U(1)$ (Froggatt-Nielsen type) or/and discrete symmetries are used
for explanation of the mass hierarchies.
Mixing is a consequence of the relations between mass matrix elements
(like in eq. (\ref{third})), whereas masses depend on the absolute values of the elements.
For particular mass spectrum (set of values of masses), the mass matrix elements are
fixed and in some cases this may lead to more symmetric form of the mass matrix -
to additional symmetry. Then a covering symmetry group should fix both mixing and masses.
(E.g., one can consider the TBM mass matrix with equality of elements $a = b$,
which gives the spectrum with normal mass hierarchy and $m_1 \approx 0$, {\it etc}.)
\section{From leptons to quarks}
Do quarks need leptonic discrete symmetries? It is not accidental
that in the talks devoted to flavor physics in the quark sector
the leptonic symmetries proposed are not even mentioned.
Although originally the first the discrete symmetries have been applied
for flavor in the quark sector~\cite{pakvasa}.
Presently there is no clear attempts to go ``from quarks to leptons''
(approach which once has been failed). The $D_{14}$ symmetry has been proposed for
explanation of the Cabibbo angle value without extension to leptons
\cite{claudia}.
It is clear that the quark and lepton mixings are strongly different,
and probably this difference is directly related to smallness of neutrino mass.
\subsection{Extending symmetry to the quark sector}
There are two different ways to extend the leptonic symmetries
to the quark sector.
1) The first possibility is to use the same representations ${\bf 3}$ and ${\bf 1}$
for quarks as for leptons. In the lowest order
one can obtain
\begin{equation}
V_{CKM}^0 = I, ~~~~ U_{PMNS}^0 = U_{TBM}.
\end{equation}
This difference of mixings
can be attributed to the Majorana nature of neutrinos.
As a consequence of symmetry, the Dirac
mass matrices in the quark and lepton sectors
are the same, both leading to zero mixing.
(The Dirac matrices can be responcsible for the mass hierarchies
of the charged fermions.)
The TBM follows via seesaw from certain structure of the Majorana mass
matrix of the RH neutrinos. Then corrections from high order operators
generate the CKM mixing and the deviations of lepton mixing from the TBM form.
Generic problem is that corrections which would explain the Cabibbo
angle lead to too large deviations from TBM, so that additional tuning is required.
2) Another way is to use different representations
of the flavor symmetry group for quarks and leptons.
In particular, one can choose groups which contain not only
representations ${\bf 3}$ and ${\bf 1}$ but also ${\bf 2}$,
and assign three generations of quarks to the representations ${\bf 2}$ and
${\bf 1}$ instead of ${\bf 3}$ in lepton sector. This implies that family symmetry
does not exist and leaves another question: why quarks
and leptons have different symmetry properties, that is, fundamentally different.
As an example of realization of the second approach,
consider model based on the symmetry $T^\prime$.
The $T^\prime$ group has order 24 being the double covering of $A_4$ or
binary tetrahedron group. The generating elements of this group
are $S$, $T$ and $R$ and presentation of the group:
\begin{equation}
T^3 = I, ~~S^2 = R, ~~~R^2 = I, ~~~ (ST)^3 = 1.
\end{equation}
Here $R = 1$ for the odd-dimensional representations and
$R = -1$ for the even-dimensional representations. Again the element $U$ is missing.
Irreducible representations of the group include
${\bf 1, ~1', ~1'', ~ 2,~ 2'~, 2'', ~3}$.
The products of representations and invariants,
\begin{equation}
3 \times 3 = 1 + 1' + 1'' + 3 + 3, ~~~ 1' \times 1'' = 1,
\end{equation}
coincide with those in the $A_4$ case (see Table 1). New possibilities are related to
existence of the doublet representations
\begin{equation}
2^a \times 3 = 2 + 2' + 2'',
\end{equation}
where ${\bf 2}^a = {\bf 2}, ~{\bf 2'}, ~ {\bf 2''} $ and
\begin{equation}
2 \times 2 = 1 + 3, ~~~ 1 \times 2 = 2
\label{eq:news}
\end{equation}
with ``conservation'' of primes. The singlet which
appears in (\ref{eq:news}) allows one to produce new (in comparison with $A_4$)
flavor structures.
The mass generation scheme of the model~\cite{tpr-fer} based on
$T^{\prime}$ is shown in fig.~\ref{tprime}.
\begin{figure}[h]
\includegraphics[width=22pc]{rome10tpr.ps}\hspace{2pc}%
\begin{minipage}
[b]{12pc}\caption{\label{tprime}
Scheme of the mass generation in the $T^\prime$ model~\cite{tpr-fer}.
See the caption of fig.~\ref{a4am} for explanation.}
\end{minipage}
\end{figure}
Features of the model include the following.
The model has an auxiliary group $Z_3$.
There is no RH neutrinos, and neutrino masses are generated by D=6 operators
$L L H_u H_u f$, where $f = \phi_s,~ \xi$.
The origin of mixing is the Majorana nature of neutrinos.
In quark sector the two light generations form doublet, whereas
the third one is a singlet of $T'$.
The RH components of charged leptons are different singlets of $T^\prime$
but they have the same prescription of $Z_3$.
Four different flavon fields are introduced: two triplets and
one doublet ${\bf 2''}$ (no ${\bf 2, 2'}$).
Only doublets ${\bf 2''}$ are used for quarks.
$Z_3$ prescription looks random: doublets transform with $\omega$,
all RH components have $\omega^2$ transformation,
and there are various missing representations.
\subsection{TBM and GUT's}
Generic problem of many models is that the flavor prescriptions
required for explanation of difference of mass and mixing of quarks and leptons
prevents from their embedding into Grand Unified Theories.
The problem can be resolved by increasing number of matter fields and
locating the known fermions with new ones in the same multiplets.
One can start immediately from the GUT structure and known
matter fields:
\begin{equation}
GUT \times G_{flavor} + {\rm new~~ elements},
\nonumber
\end{equation}
where ``new elements'' may include singlet fermions and additional Higgses
or/and pairs of vector-like matter fields, their mixing with usual
matter, {\it etc.}.
As an example, structure of the model based on
$SU(5) \times A_4$ \cite{cooper} is shown in fig.~\ref{su5a4}.
The upper quarks get masses via interactions
$T_i T_j H_5 \{ f_{ij} \}$,
where $\{ f_{ij} \}$ is product of certain number (from zero to 5)
of flavon fields: e.g., $f_{33} = 1$,
$f_{32} = \phi_{123} \phi_3 \xi^2$.
\begin{figure}[h]
\includegraphics[width=22pc]{rome10gut1.ps}\hspace{2pc}%
\begin{minipage}
[b]{12pc}\caption{\label{su5a4}
Scheme of generation of masses of neutrinos, charged leptons and down quarks in
the $SU(5) \times A_4$ model \cite{cooper}.
See the caption of fig.~\ref{a4am} for explanation.}
\end{minipage}
\end{figure}
The following remarks are in order.
Extended auxiliary symmetry
$Z_2 \times Z_2 \times U(1) \times U(1)$ is imposed.
The singlet $N$ and
$\Sigma = 24$ - adjoint representation of fermions
are introduced apart from usual {\bf 10}-plets and $\bar{\bf 5}-$ plets;
neutrino masses are generated by a combination of type the I and type III seesaw.
Only $5-$plets form family structure, whereas $10-$plets and RH neutrinos are
singlets of $A_4$.
Matter fields are in ${\bf 3, ~ 1}$, but ${\bf 1'}$, ${\bf 1''}$ are missing.
Four different flavon multiplets
with ``random'' $Z_2 \times Z_2$ prescriptions generate masses.
Actually, $SU(5)$ has enough flexibility:
three different representations ${\bf 10}, {\bf 5} , {\bf 1}$
allow one to write independent terms for the upper quarks,
for down quarks and charged leptons and for neutrinos.
Models based on $SO(10)$
with all known fermions being in the same {\bf 16}-plets
are more constrained.
New elements should be added to the $SO(10) \times G_{flavor}$
structure. Two different ways are proposed:
1) add singlet fermions and ${\bf 16}_H$ Higgs multiplets
with couplings ${\bf 16} S {\bf 16}_H$ and flavons.
In \cite{hss} $G_{flavor} = T_7$ and screening of the Dirac
structures in the see-saw mechanism can be achieved which leads to
independent structures of the mass matrices of neutrinos and charged fermions.
Incomplete (partial) screening can be the origin of the TBM or bi-maximal mixings.
2). Another way of model building is to introduce ${\bf 126}$
and $\overline{\bf 126}$ plets,
and thus realize the seesaw type-II mechanism which opens up a possibility
to obtain (to a large extend) independent flavor mixing in the
quark and lepton sectors. Realistic model proposed
in \cite{dutta} and based on the symmetry $S_4 \times Z_n$
requires also introduction of vector-like pairs ${\bf 16}$,
$\overline{\bf 16}$ of matter fields, additional Higgs ${\bf 10}-$plet, and flavons.
\subsection{Is TBM accidental?}
Experiment still allows relatively large deviation of the mixing parameters
from the TBM values: $\Delta \sin^2 \theta_{23} \sim 0.05$,
$\Delta \sin^2 \theta_{12} \sim 0.02$, $\Delta \sin \theta_{13} \sim 0.15$.
The deviations can lead to strong (maximal)
violation of the TBM-conditions (\ref{third}),
and consequently, to significant deviation of $m_\nu$ from the TBM form.
For instance, instead of the first equality in (\ref{third}),
the equality with changed sign, $m_{e\mu} \sim - m_{e \tau}$,
is allowed without any modification of two others.
Leading structures of the mass matrix are relatively robust,
whereas the sub-leading structures
can change under these corrections completely.
It is therefore not excluded that
the approximate TBM is accidental being just
an interplay of several independent
factors (contributions) \cite{abbas}. Alternatively it can be
a manifestation of some other symmetry
which differs from TBM, or other structures.
This opens up new approaches to explain the data.
There are other possible applications of discrete symmetries.
In the universal approach to the quark and lepton masses based on certain
ansatz for the shape of the mass matrices discrete symmetries are used to get
texture zeros. In this context the corrected Fritzsch ansatz
has been explored in~\cite{simoes}.
\section{QLC and quark-lepton symmetries}
The Quark-lepton complementarity (QLC)~\cite{QLC} is an alternative
to description of the fermion mixings which is based
on observation that
\begin{equation}
\theta_{12}^l + \kappa_{12} \theta_{12}^q \approx \pi/4, ~~~
\theta_{23}^l + \kappa_{23} \theta_{23}^q \approx \pi/4 ,
\end{equation}
where $\kappa_{23}, \kappa_{12} \sim 1$, say (0.7 - 1).
Qualitatively, the QLC relations mean that
- the 2-3 leptonic mixing is close to maximal
because the 2-3 quark mixing is small;
- the 1-2 leptonic mixing deviates from maximal one substantially because
the 1-2 quark mixing is relatively large.
In other words,
\begin{equation}
{\rm ``Lepton ~mixing = bi-maximal~mixing - quark~mixing''}
\nonumber
\end{equation}
with possible implications being:
1. The quark-lepton symmetry,
which, in turn, implies the quark-lepton unification, or GUT, or common family
(horizontal) symmetry.
2. Existence of structure which produces the bi-maximal (BM) mixing.
The structure for BM could be related to see-saw with special properties of
the RH neutrino mass matrix.
The Bi-maximal mixing, $U_{bm} = U_{23}^m U_{12}^m$,
is characterized by maximal 1-2 and 2-3 rotations, and zero 1-3 rotation.
There is no CP-violation.
Possible scenario is that in the lowest order
\begin{equation}
V_{CKM}^0 = I, ~~ U_{PMNS}^0 = U_{bm},
\label{eq:zorder}
\end{equation}
and may be $m_1 = m_2 = 0$. If the BM structure in the lepton sector
is generated by the seesaw
mechanism, the corrections from the Dirac mass matrix produce (i) mass splitting,
(ii) CKM and (iii) deviation from the bi-maximal mixing.
The situation when the deviations and CKM mixing are related
by the quark - lepton symmetry (or the same flavor symmetry for quarks and leptons)
can be called ``strong complementarity''.
Another possibility is Cabibbo ``haze'' \cite{haze} \cite{QLC} or
the weak complementarity \cite{weakqlc}.
Deviations from BM are due to some corrections which can be of the
same order in the quark and lepton sectors but not necessarily related.
The corrections can be of the size of the Cabibbo angle.
Possible realization is that
the corrections are due to high order flavon interactions which generate
simultaneously the CKM mixing and deviation from BM.
In this case Grand Unification and family symmetries are not necessary.
\subsection{BM-symmetry}
A discrete symmetry can be behind the BM mixing as
the lowest order structure (\ref{eq:zorder}).
The bi-maximal mass relations
are
\begin{equation}
m_{e\mu} = m_{e \tau}, ~~~
m_{\mu \mu} = m_{\tau \tau}, ~~~
m_{e e} = m_{\mu \mu} + m_{\mu \tau} .
\end{equation}
The last equality distinguishes the bi-maximal case from TBM (see (\ref{third})).
The BM mass matrix, $m_{BM}$, is invariant under transformations
\begin{equation}
V^T_i m_{BM}V_i = m_{BM}, ~~~~~
V_i = S_{BM}, ~~ U,
\end{equation}
where
\begin{equation}
S_{BM} =
\frac{1}{2} \left(
\begin{array}{lll}
0 & ~~ \sqrt{2} & ~~ \sqrt{2} \\
~ ... & - 1 & ~~ 1 \\
~ ... & ... & - 1
\end{array}
\right) ~~~
\end{equation}
with property $S_{BM}^2 = I$.
One can select the matrix of transformation which keeps the charged leptons mass matrix
to be diagonal, in the form $T_{BM} = diag (- 1, - i, i)$. In this case $T_{BM}^4 = I$,
so that $T$ and $S_{BM}$ turn out to be the generating elements of
$S_4$ symmetry group \cite{s4}.
\subsection{$S_4$ symmetry and model}
$S_4$ has the order 24, it is the permutation
group of 4 elements. With generating elements $S_{BM}$ and $T_{BM}$
it has the following presentation
\begin{equation}
S_{BM}^2 = T_{BM}^4 = (T_{BM} S_{BM})^3 = I
\end{equation}
(compare with (\ref{repa4})).
It has irreducible representations
${\bf 1, ~1', ~ 2, ~3, ~ 3'}$.
The products of representations read
\begin{equation}
3 \times 3 = 3' \times 3' = 1 + 2 + 3 + 3', ~~~~
3 \times 3' = 1' + 2 + 3 + 3'
\nonumber
\end{equation}
\begin{equation}
1' \times 1' = 1, ~~~ 1' \times 2 = 2 ~~~~
2 \times 3 = 2 \times 3' = 3 + 3' ~~~
2 \times 2 = 1 + 1' + 2 ,
\end{equation}
and the latter contains singlet, thus leading to new flavor structures.
\begin{figure}[h]
\includegraphics[width=23pc]{rome10bmx1.ps}\hspace{2pc}%
\begin{minipage}
[b]{12pc}\caption{\label{bms4}
Scheme of the $S_4$ model for the weak quark-lepton complementarity \cite{afm}.
See the caption of fig.~\ref{a4am} for explanation.}
\end{minipage}
\end{figure}
Structure of the model~\cite{afm} based on $S_4 \times Z_4$ and $U(1)_{FN}$
is shown in fig.\ref{bms4}. It resembles the structure of $A_4$ model (see Fig.~\ref{a4am}).
The following are in order.
Only $S_4$ representations ${\bf 3}$, ${\bf 1}$, and ${\bf 1'}$ are used
for the matter fields, the representations ${\bf 3'}$, ${\bf 2}$ are missing.
Flavons are in ${\bf 3,~ 3',~ 1}$ representations,
and ${\bf 2,~ 1'}$ are absent.
The multiplets have {\it ad hoc} $Z_4$ prescription.
The Froggatt-Nielsen (FN) mechanism is introduced for the mass hierarchies and only RH
components of leptons have non-zero FN-charges.
Deviation from BM are due to high dimension operators with flavon fields.
\section{Perspectives and tests}
Minimal and simplest models
which lead to the lepton mixing of the TBM or BM type from discrete symmetries,
have been systematically explored.
The key problem is to check the models or at least some generic features of
the whole context. Unfortunately, majority of the models do not give specific
and precise predictions which can be tested.
Still one expects certain connections between the low energy observables and also
probably observables at high energy accelerators under certain additional conditions.
In this connection several phenomenological directions should be mentioned.
1. Precision measurements of the neutrino parameters.
Determination of the 1-3 mixing and the deviation of 2-3 mixing from
maximal one are of great importance. Some models predict $\theta_{13}$;
discrimination of models with large and very small $\theta_{13}$ will be possible.
Relations between $\theta_{13}$ and the deviation of $\theta_{23}$ from $\pi/2$
may reveal certain ways of realizations of the discrete symmetries.
Determination of the absolute scale of neutrino masses and mass hierarchies,
establishing possible relations between mass ratios and mixings
can give further insight.
2. Double beta decay. Some models lead to certain predictions for the
effective Majorana mass of the electron neutrino, $m_{ee}$,
as well as its connections to the effective electron neutrino mass, $m_e$, and sum of neutrino
masses $\sum m_i$~\cite{bbb}.
3. Rare decays with lepton flavor violation: $\mu \rightarrow e + \gamma$,
$\tau \rightarrow e + \gamma$, $\tau \rightarrow e + \gamma$
\cite{Feruglio} \cite{mueg}.
Equality of the rates of these decays may testify for certain class
of models with discrete symmetries~\cite{mueg}. Interesting predictions
for processes like $\tau^- \rightarrow e^+ \mu^- \mu^- $,
$\tau^- \rightarrow \mu^+ e^- e^- $ are given which depend on parameter of
violation of the discrete symmetry~\cite{Feruglio}.
4. LHC and high energy accelerators.
Models with flavored Higgses can be directly tested
in the collider experiments~\cite{multihi}.
Even for the SM Higgs one expects modifications
of the decay and production rates in the presence of
horizontal symmetries!\cite{lamhiggs}.
5. Leptogenesis. It is affected by discrete symmetries
and depends on the way the symmetries are broken \cite{leptogen}.
In some cases suppression of the leptonic asymmetry is expected.
6. Dark matter. Particles of the
dark matter (e.g. in the multi-Higgs models) can be stabilized
by some discrete symmetry which is related to the flavor symmetry~\cite{dmat}.
For instance, it may be a residual symmetry after breaking
$A_4 \rightarrow Z_2$.
7. As already mentioned, some future discoveries can simply reject the described
approach or require
its strong modification. That includes discoveries of
new fermions, like sterile neutrinos, the 4th generation of fermions,
the right handed neutrinos or $W_R$ of the left-right symmetric models, {\it etc.}.
\section{Conclusions}
In recent years,
it has been shown that the approximate
TBM mixing can be consistently obtained in the context of
gauge models with spontaneously broken flavor
symmetries and rather extended additional structures.
The considered examples of models show the price one should pay for realization of idea
``mixing from discrete symmetries''.
There are two opposite points of view on the obtained results:
I. The features of experimental data which testify for a symmetry behind
lepton mixing are actually accidental. The deviations from TBM can be significant.
Realizations are too complicated with
the number of assumptions being several times bigger than the number of mixing angles.
This indicates that alternative approaches to explanation of the data
should be pursued.
II. Some version of broken discrete symmetries
give correct explanation of the data. Physics behind neutrino mass and mixing
has rich extended structure and it leads to rich phenomenology.
(It may happen that still some important elements of the approach are missing.
Preferable scenario?
The difference of lepton and quark mixings is related
{\it directly} to smallness of neutrino mass and probably its Majorana character.
GUT's still look very appealing and
there is no point to sacrifice them in favor of
the present models with flavor symmetries.
The observed symmetry in the lepton mixing is related to a symmetry of
Hidden sector at some high mass scales.
It communicates with us via the neutrino portal -- mixing with neutrinos.
No analogy of this in the quark sector exists.
Another physics (but the same in $q-$ and $l-$ sectors)
is involved in generation of CKM and deviation of PMNS from the symmetric case.
Unfortunately, it is difficult, if possible, to check this possibility,
but this is not the problem of Nature...
\subsection{Acknowledgments}
Author is grateful to D. Hernandez for numerous discussions of the material presented
in this paper.
\section*{References}
|
\section{INTRODUCTION}
\label{intro}
Understanding the dynamics and evolution of the interstellar medium (ISM) is
critical to advancing our knowledge of many astrophysical phenomena spanning
a wide range of scales such as star
formation, cosmic ray physics, magnetic reconnection, galaxy evolution, and
magnetic dynamo theory (Elmegreen \& Scalo 2004). An essential component of the current paradigm of the
ISM is the ubiquitous existence of magnetohydrodynamic (MHD) turbulence (see
review by McKee \& Ostriker 2007). Turbulence and magnetic fields play a
crucial roll in each of these processes and are some of the main drivers of
ISM evolution. Evidence for the role of turbulence is seen in
the ``big power law'' of the electron density fluctuations (Armstrong, Rickett, \& Spangler 1995),
fractal structure in the molecular media (Elmegreen \& Falgarone 1996, Stutzki et al. 1998),
and intensity fluctuations contributed by both density and turbulent velocity
in channel maps (Crovisier \& Dickey 1983; Green 1993;
Deshpande, Dwarakanath, \& Goss 2000; Elmegreen, Kim, \& Staveley-Smith 2001).
Despite the obvious importance of MHD turbulence to astrophysics, few methods
exist to study it directly. Due to advances in computational power and the
general recognition of turbulence as an important ISM process,
major advances have been made in MHD
turbulence theory and observation in the last ten years. However, the issue
of turbulence and its effects on the ISM (and processes therein) remains one of the most
exciting and open problems in the field (Elmegreen \& Scalo 2004).
Astrophysical turbulence is a complex nonlinear phenomena
that can occur in a multiphase media with many energy injection
sources on scales ranging from kpc down to sub-AU.
Although limited in complexity, numerical simulations of turbulence provide one of the best avenues
for researchers of the ISM to understand the nature of magnetized
turbulence. The combined efforts of predictive theory and numerical tests have
greatly increased our knowledge of MHD turbulence, including its
anisotropy, intermittency, and imbalanced nature
(see Cho, Lazarian, \& Vishniac 2002, Kowal \& Lazarian 2010, Beresnyak \& Lazarian 2010).
Yet what about observationally driven studies of turbulence? The most common
observational techniques to study turbulence include scintillation studies,
which are limited to fluctuations in only ionized media (e.g.
Narayan \& Goodman 1989; Spangler \& Gwinn 1990), density fluctuations
via column density maps, and radio spectroscopic observations via centroids of
spectral lines (Falgarone et al. 1994; Miesch \& Bally 1994; Miesch \& Scalo
1995; Lis et al. 1998; Miesch, Scalo, \& Bally 1999). Column densities are the most abundant
and easily obtained observable data and have shown that density fluctuations can
be a useful and straightforward way of gauging turbulence parameters
(Monin \& Yaglom 1967; Lithwick \& Goldreich 2001; Cho \& Lazarian 2003).
Position-Position-Velocity (PPV) spectroscopic data has the
advantage over column density maps in that it contains information on the turbulent
velocity field. However, this type of data provides contributions of \textit{both}
density and velocity fluctuations entangled together, and the process of
separating the two has proven a challenging problem (see Lazarian 2006b).
In addition, one must use caution when dealing with PPV data, as structures seen
in velocity slices are not one-to-one with structures in three dimensional position space.
One of the main approaches for characterizing ISM turbulence is based on using
statistical techniques and descriptions. The most common ``go-to'' tool for both
observers and theorist alike is the spatial power spectrum. In fact, most of
the attempts to relate observations to models has been by obtaining the spectral
index (i.e. the log-log slope of the power spectrum) of column density and
velocity. While obtaining the spectral index of column density is
straightforward, more sophisticated techniques for obtaining the velocity
spectral index from PPV data have been recently developed. These include: Velocity
Channel Analysis (VCA) (Lazarian \& Pogosyan 2000; Esquivel et al. 2003;
Lazarian \& Pogosyan 2004; Lazarian et al. 2001; Padoan et al. 2003; Chepurnov
\& Lazarian 2009), the Spectral Correlation Function (SCF) (Rosolowsky et
al. 1999; Padoan, Rosolowsky, \& Goodman 2001), Velocity Coordinate Spectrum (VCS)(Lazarian \&
Pogosyan 2008, 2006; Chepurnov \& Lazarian 2006, 2009;
Padoan et al. 2009), and Modified Velocity Centroids (Lazarian \& Esquivel 2003;
Esquivel \& Lazarian 2005; Ossenkopf et al. 2006; Esquivel et al. 2007). For
cases where turbulence is supersonic, the VCA is most appropriate while
centroids can be used in subsonic cases.
Although the power spectrum is useful for obtaining information about energy
transfer over scales, it does not provide a full picture of turbulence,
partially because it only contains information on Fourier amplitudes. An
example of this is illuminated in a study by Chepurnov et al. (2008), who showed
that a substantially different distribution of density could have the same
power spectrum. In light of this, many other techniques have been developed
to study and parametrize observational magnetic turbulence. These include
higher order spectrum, such as the bispectrum, higher order statistical
moments, topological techniques (such as genus), clump and hierarchical
structure algorithms (such as dendrograms), principle component analysis, and structure/correlation
functions as tests of intermittency and anisotropy (for examples of such
studies see Heyer \& Schloerb 1997, Burkhart et al. 2009; Chepurnov \& Lazarian 2009;
Kowal, Lazarian, \& Beresnyak 2007;
Goodman et al. 2009; Burkhart et al. 2011). Wavelets methods, and variations on them such
as the $\Delta$-variance method, have also been shown to be very useful in characterizing
inhomogeneities in data (see Ossenkopf, Krips, \& Stutzki 2008a, 2008b).
In particular, many of the
studies mentioned above focus on obtaining the parameters of turbulence from
observations. These parameters include sonic and Alfv\'enic Mach numbers,
injection scale, gas temperature, and Reynolds number. In particular the sonic
and Alfv\'enic Mach numbers provide much coveted information on the gas
compressibility and magnetization. Many of these techniques, geared towards
obtaining the parameters of turbulence via density fluctuations studies, were
successfully applied to observational data (see Burkhart et al. 2010 and
Chepurnov et al. 2008, for examples). VCA and VCS were also applied to
Galactic HI data and successfully recovered the spectrum of velocity (see
Chepurnov et al. 2010).
In this vein, Esquivel \& Lazarian (2010), henceforth known as EL10, used the so-called
Tsallis statistic for studies of MHD turbulence. It is this
statistic that is the focus of this work, and here we will further illuminate
its uses. The Tsallis distribution is a function that can be fit to incremental
PDFs of turbulent density, magnetic field, and velocity.
In astrophysical settings, Tsallis statistics was originally used in
the context of solar wind observations (Burlaga \& Vi\~nas 2004a). EL10 applied this method to
3D MHD simulations with four varying values of sonic and Alfv\'enic Mach
number at $256^3$ resolution. They explored density, magnetic field, velocity,
and column densities, and found that the Tsallis distribution is a very good fit to PDFs of
increments of turbulence. They also found that the parameters of the fit that
describe the width and tails of the PDFs showed dependency on the
compressibility and magnetization of the simulation. The statistic is
particularly useful in that it is scale independent
and thus a comparison between the analysis of simulations and observations is not
burdened with complicated scaling relations. This opens up the possibility of
using the Tsallis method on observed ISM data in order to gain access to information
on these parameters. While EL10 was the first to implement this tool on simulations of ISM MHD turbulence,
they used low spatial resolution simulations,
a small parameter regime, and did not explore the dependencies on the amplitude fit parameter.
They also did not explore the use of Tsallis statistics
on spectroscopic data. In this paper, we will greatly extend their parameter
range and resolution from 4 at $256^3$ isothermal MHD simulations to 14 at
$512^3$. In addition, we will more explicitly explore the ability of the Tsallis function
to describe data of an observable nature, such as smoothed synthetic column density maps
and synthetic PPV data of varying velocity resolution. We also investigate the
quality of our fits and subsequent fit parameters.
The paper is organized as follows. In \S~\ref{tsallis} we describe the
Tsallis distribution which is fit to increments of MHD turbulence
PDFs. In \S~\ref{numericalsetup} we discuss our numerical scheme and
resulting simulations. In \S~\ref{3d} we apply Tsallis to
non-observational 3D quantities (density and directional components of
magnetic field and velocity) and test the accuracy of our fits. In \S~\ref{observ}
we apply this tool to observational quantities such as column density and PPV data. In
\S~\ref{disc} we discuss our results followed by the conclusions in
\S~\ref{conc}.
\section{TSALLIS STATISTICS}
\label{tsallis}
The Tsallis distribution (Tsallis 1988) was originally derived as
a means to extend traditional Boltzmann-Gibbs mechanics to fractal and
multifractal systems. The complex dynamics of multifractal systems apply to
many natural environments such as ISM turbulence
(Shivamoggi 1995). It is therefore fitting to
explore the extent to which Tsallis statistics can be used to describe these
systems and processes therein (for further discussion see section \ref{relate}).
Work of this nature was first carried out by Burlaga
and collaborators (Burlaga \& Vi\~{n}as 2004a, 2004b, 2005a, 2005b, 2006;
Burlaga, Ness, \& Acu{\~n}as 2006, 2007, 2009; Burlaga, Vi{\~n}as, \& Wang 2007)
to describe the temporal variation
in PDFs of magnetic field strength and velocity of solar wind measured by the
\emph{Voyager 1} \& \emph{2} spacecrafts. EL10 used Tsallis statistics to
describe the spatial variation in PDFs of density, velocity, and magnetic field of
MHD simulations similar to those used here. Both efforts
found that the Tsallis distribution provided adequate fits to their PDFs
and gave insight into statistics of turbulence.
The Tsallis function of an arbitrary incremental PDF $(\Delta f)$ has the form:
\begin{equation}
\label{(1)}
R_{q}= a \left[1+(q-1) \frac{\Delta f(r)^2}{w^2} \right]^{-1/(q-1)}
\end{equation}
The fit is described by the three dependent parameters $a$, $q$, and $w$. The
$a$ parameter describes the amplitude while $w$ is related to the width
or dispersion of the distribution. Parameter $q$, referred to as the ``non-extensivity
parameter'' or ``entropic index'', describes the sharpness and tail size of the
distribution.
The arbitrary function used to describe density and the directional
components of velocity and magnetic field in this application takes the form of
our incremental PDF. It has the form
$\Delta f(r)=(f(r)-\langle f(r)\rangle_{\bf{x}})/\sigma_{f}$, where
$\langle$...$ \rangle_{\bf{x}}$
refers to a spatial average. Depending on the quantity in question, we set
$f(x)=\rho(x + r) - \rho(x);
v_{x}(x + r) - v_{x}(x);
v_{y}(x + r) - v_{y}(x);
v_{z}(x + r) - v_{z}(x);
{\bf B}_{x}(x + r) - {\bf B}_{x}(x);
{\bf B}_{y}(x + r) - {\bf B}_{y}(x);
{\bf B}_{z}(x + r) - {\bf B}_{z}(x)$;
where ``$r$'' is the lag or spatial scale. For a given lag this calculation
is done for each pixel in the three cardinal directions. A normalized 100 bin
histogram of these values results in our incremental PDF which is then fit
with the Tsallis function (see Figure \ref{fig:3dhistsall} for an example PDF).
The Tsallis fit parameters are in many ways similar to statistical moments.
Moments, more specifically the third and fourth order moments, have been
used to describe the density distributions and have shown sensitivities to
simulation compressibility (Kowal, Lazarian, \& Beresnyak 2007; Burkhart et al. 2009). The
first and second order moments simply correspond to the mean and variance of a
distribution. Skewness, or third order moment, describes the asymmetry of a
distribution about its mode. Skewness can have positive or negative
values corresponding to right and left shifts of a distribution respectively. The
fourth order moment, kurtosis, is a measure of a distribution's peaked or flatness
compared to a Gaussian distribution. Like skewness, kurtosis can have positive or
negative values corresponding to increased sharpness or flatness.
In regards to the Tsallis fitting parameters, the $w$
parameter is similar to the second order moment variance while $q$ is
closely analogous to fourth order moment kurtosis. Unlike higher
order moments, however, the Tsallis fitting parameters are dependent least-squares
fit coefficients and are more sensitive to subtle changes in the PDF.
\section{MHD SIMULATIONS}
\label{numericalsetup}
We generate a database of 14 three dimensional numerical simulations (512$^3$
resolution) of isothermal compressible MHD turbulence by using the
Cho \& Lazarian (2002) code and varying the input values for the sonic
and Alfv\'enic Mach number (See Table \ref{tab:table1}). The sonic Mach number
is defined as ${\cal M}_s \equiv \langle |{\bf v}|/C_s \rangle$, where
$|{\bf v}|$ is the local velocity vector magnitude and $C_s$ is the sound speed. Averaging
is done over the entire simulation. Similarly, the Alfv\'enic Mach number is
${\cal M}_A\equiv \langle |{\bf v}|/v_A \rangle$, where $v_A = |{\bf
B}|/\sqrt{\rho}$ is the Alfv\'enic velocity, $|{\bf B}|$ is the local magnetic
field vector magnitude, and $\rho$ is density. Below, we briefly outline the major
points of the numerical setup (for more details see Cho \& Lazarian 2002).
The code is a second-order-accurate hybrid essentially
non-oscillatory (ENO) scheme which solves
the ideal MHD equations in a periodic box:
\newpage
\begin{eqnarray}
\frac{\partial \rho}{\partial t} + \nabla \cdot (\rho {\bf v}) = 0, \\
\frac{\partial \rho {\bf v}}{\partial t} + \nabla \cdot \left[ \rho {\bf v} {\bf v} + \left( p + \frac{B^2}{8 \pi} \right) {\bf I} - \frac{1}{4 \pi}{\bf B}{\bf B} \right] = {\bf f}, \\
\frac{\partial {\bf B}}{\partial t} - \nabla \times ({\bf v} \times{\bf B}) = 0,
\end{eqnarray}
with zero-divergence condition $\nabla \cdot {\bf B} = 0$,
and an isothermal equation of state $p = C_s^2 \rho$, where
$p$ is the gas pressure. On the right-hand side, the source term $\bf{f}$ is a
random large-scale driving force\footnote{${\bf f}= \rho d{\bf v}/dt$}. Boundary
conditions are periodic. We drive turbulence solenoidally in Fourier space at
wave scale k equal to about 2.5 (2.5 times smaller than L, the size
of the box). This defines the injection scale in our models
and the driving is done in Fourier space
to minimize the influence of the driving force on the generation of density structures.
The initial density and velocity fields are set to unity.
We do not set the viscosity and diffusion explicitly in our models.
The scale at which dissipation starts to act is defined by
the numerical diffusivity of the scheme. The ENO-type schemes
are considered to be relatively low diffusion (see Liu
\& Osher 1998; Levy, Puppo, \& Russo 1999). The numerical diffusion depends
not only on the adopted numerical scheme but also on the
smoothness of the solution, so it changes locally in the system.
In addition, it is also a time-varying quantity. All these problems
make its estimation very difficult and incomparable between
different applications. However, the dissipation scales can be estimated
approximately from the velocity spectra. In the case of
our models we estimate the dissipation scale at $k_{\nu}=30$ pixels.
\begin{deluxetable}{lllllc}
\tablewidth{0pt}
\tablecaption{Simulation Parameters}
\tablehead{
\colhead{Model} &
\colhead{B$_{ext}$} &
\colhead{$\mathcal{M}_{s}$} &
\colhead{$\mathcal{M}_{A}$} &
\colhead{Description}
}
\startdata
1 & 0.1 &10.0 & 2.0 & Supersonic \& super-Alfv\'{e}nic\\
2 & 1.0 &10.0 & 0.7 & Supersonic \& sub-Alfv\'{e}nic\\
3 & 0.1 & 7.0 & 2.0 & Supersonic \& super-Alfv\'{e}nic\\
4 & 1.0 & 7.0 & 0.7 & Supersonic \& sub-Alfv\'{e}nic\\
5 & 0.1 & 6.0 & 2.0 & Supersonic \& super-Alfv\'{e}nic\\
6 & 1.0 & 6.0 & 0.7 & Supersonic \& sub-Alfv\'{e}nic\\
7 & 0.1 & 4.0 & 2.0 & Supersonic \& super-Alfv\'{e}nic\\
8 & 1.0 & 4.0 & 0.7 & Supersonic \& sub-Alfv\'{e}nic\\
9 & 0.1 & 3.0 & 2.0 & Supersonic \& super-Alfv\'{e}nic\\
10& 1.0 & 3.0 & 0.7 & Supersonic \& sub-Alfv\'{e}nic\\
11& 0.1 & 0.7 & 2.0 & Subsonic \& super-Alfv\'{e}nic\\
12& 1.0 & 0.7 & 0.7 & Subsonic \& sub-Alfv\'{e}nic\\
13& 0.1 & 0.1 & 2.0 & Subsonic \& super-Alfv\'{e}nic\\
14& 1.0 & 0.1 & 0.7 & Subsonic \& sub-Alfv\'{e}nic\\
\enddata
\label{tab:table1}
\end{deluxetable}
As density fluctuations are generated by the interaction of MHD waves,
the time $t$ is in units of the large eddy turnover time
($\sim L/\delta V$) and the length in units of $L$, the energy injection scale.
The magnetic field consists of the uniform background
field and a fluctuating field: ${\bf B}= {\bf B}_\mathrm{ext} + {\bf
b}$. Initially ${\bf b}=0$. We divided our models into two groups
corresponding to sub-Alfv\'enic ($B_\mathrm{ext}=1.0$) and
super-Alfv\'enic ($B_\mathrm{ext}=0.1$) turbulence. For each group we
compute several models with different values of gas pressure (see Table
\ref{tab:table1}).
\begin{figure*}[bt]
\begin{center}
\includegraphics[keepaspectratio=true,scale=.70]{fig_01a.eps}\\
\includegraphics[keepaspectratio=true,scale=0.35]{fig_01b.eps}
\includegraphics[keepaspectratio=true,scale=0.35]{fig_01c.eps}
\end{center}
\caption{Top: 3D density rendering of a $512^3$ MHD simulation where
${\cal M}_s$=10 (highly turbulent) and ${\cal M}_A$=0.7 (high
magnetization), Model \#2 from Table \ref{tab:table1}. The mean
magnetic field is applied
along the x-direction for each simulation. Bottom: Visualizations of
compressed 2D column density (left) and cloud bounded column density (right).
Each is created along the line of sight parallel to the mean magnetic field.}
\label{fig:3dpic}
\end{figure*}
The top of Figure \ref{fig:3dpic} presents a 3D density rendering of simulation \#2 (Table
\ref{tab:table1}) where ${\cal M}_s$=10 and ${\cal M}_A$=0.7. The x and y axes
are labeled on the figure with the mean magnetic field ($<${\bf B}$>$)
parallel to the x direction. The bottom displays, for the same simulation, a column
density (left) and the same column density convolved with a radially decreasing Gaussian function
in order to create the effect of cloud-like boundaries. See section
\ref{observ} for descriptions of column density construction.
\section{Tsallis Fit of Density, Velocity, and Magnetic Field}
\label{3d}
For the first portion of our analysis we investigate Tsallis fits of PDFs
of 3D density and the three directional components of magnetic field and
velocity. We fit the Tsallis distribution to incremental PDFs (see section
\ref{tsallis}) using the Levenberg-Marquardt algorithm (Levenberg 1944;
Marquardt 1963), for spatial separations (lag) 1, 2, 4,
8, 16, 32, 64, and 128 pixels (up to 1.5 times smaller then the injection scale).
Fits and PDFs are shown in Figure \ref{fig:3dhistsall}
(symbols are the data from the simulations, lines represent the Tsallis fit).
Increasing lags are displayed vertically on the same logarithmic
vertical scale. We only present magnetic and velocity fields directed along the mean
magnetic field. EL10 saw no strong variation with LOS orientation
for magnetic and velocity fields, which we confirm.
PDFs of perpendicular components are therefore omitted from the figure.
\begin{figure*}[tbh]
\centering
\includegraphics[angle=90,keepaspectratio=true,scale=0.6]
{fig_02.eps}
\caption{PDFs of 3D density and components of velocity and magnetic
field parallel to the mean magnetic field
(red, green, and blue line respectively). Panels show four different
simulations ${\cal M}_s$=10 ${\cal M}_A$=0.7, ${\cal M}_s$=10 ${\cal M}_A$=2.0,
${\cal M}_s$=0.7 ${\cal M}_A$=0.7, and ${\cal M}_s$=0.7 ${\cal M}_A$=2.0
(left to right). PDFs of larger lags are displayed vertically on the
same scale for each simulation. Y-axis is logarithmic. Velocity and
magnetic field produced similar PDF perpendicular to the mean magnetic
field.}
\label{fig:3dhistsall}
\end{figure*}
\begin{figure*}[tb]
\begin{center}
\includegraphics[angle=90,keepaspectratio=true,scale=0.5]
{fig_03.eps}
\caption{From top to bottom, fit parameters $a$ (amplitude of fit),
$q$ (related to fitting tails of PDF), and $w$ (PDF width plotted
as $w^{-2}$) are displayed vs spatial lag for all 14 simulations showing 3D
density. The left and right columns correspond to sub-
and super-Alfv\'enic simulations respectively. Three solid lines are over-plotted
for $a$ and $w^{-2}$ averaging the highly supersonic (${\cal M}_s$=10, 7, 6),
mid supersonic (${\cal M}_s$=4, 3), and subsonic (${\cal M}_s$=0.7, 0.1)
simulations. Two solid lines are over-plotted for the $q$ parameter
averaging the supersonic (${\cal M}_s$=10, 7, 6, 4, 3) and subsonic
(${\cal M}_s$=0.7, 0.1) simulations. Parameters $a$ and $w^{-2}$ show
significant sensitivities to ${\cal M}_s$ and ${\cal M}_A$ (note
scale between left and right panels). Parameter $q$
is only slightly sensitive to the simulation's compressibility. Errors are
generally less than 25$\%$, 20$\%$, and 40$\%$ for $a$, $q$, and $w^{-2}$
respectively. Error bars are omitted for clarity.}
\label{fig:3dfinal}
\end{center}
\end{figure*}
Figure \ref{fig:3dhistsall} presents PDFs (and their corresponding Tsallis fits)
of 3D density and components of velocity and magnetic field parallel to the mean
magnetic field (red, green and blue lines respectively). Panels show four different
simulations ${\cal M}_s$=10 ${\cal M}_A$=0.7, ${\cal M}_s$=10 ${\cal M}_A$=2.0,
${\cal M}_s$=0.7 ${\cal M}_A$=0.7, and ${\cal M}_s$=0.7 ${\cal M}_A$=2.0 (left to
right). Visual analysis of these figures shows tightly correlated fits
with only slight deviation near the tails of the PDFs (y-axis is logarithmic).
Three outstanding trends can be
seen across the simulations. First, is the increase in
Gaussianity as ${\cal M}_s$ decreases (turbulence becomes subsonic).
For supersonic turbulence, density PDFs are highly kurtotic.
This corresponds to a higher probability that
$\rho(x + r)-\rho(x)=0$ due to shock filaments causing central spikes of in PDFs increasing their
kurtosis (agrees with trends seen in EL10 and Burlaga \& Vi\~{n}as 2004b).
The second trend is the increase in Gaussianity as the lag
increases. At low lags density, magnetic field, and velocity have
higher probabilities of being near the mean value (here normalized to zero).
This type of behavior is congruent with the results of
Falgarone et al.(1994) which analyzed the skewness and kurtosis of PDFs of
varying scales. In the case of subsonic turbulence, the PDFs of density look
very similar to PDFs of velocity and magnetic field for high lags.
Third, there is an increase in
PDF kurtosis for density, velocity, and magnetic field for simulations of a high magnetic
field (see the first and second panel of Figure \ref{fig:3dhistsall} for small
lags). Magnetization plays an intimate role in the development of turbulence
and density enhancements and this affect can be attributed to field freeze-in.
In the following subsections
we will further describe the 3D quantities and their fits individually.
\subsection{3D Density}
\label{3dd}
From top to bottom, Figure \ref{fig:3dfinal} displays fit parameters $a$, $q$, and $w$
(plotted as $w^{-2}$) versus the spatial lag for 3D density for all 14 simulations.
The figure is separated on the left and right corresponding to sub- and super-Alfv\'enic
Mach number simulations respectively. The top panels display parameter $a$
(corresponding the amplitude) and shows a strong sensitivity to the degree of
sonic number. To emphasize this fact we break our simulations up into three
categories; highly supersonic (${\cal M}_s$=10, 7, 6), mildly supersonic (${\cal M}_s$=4, 3),
and subsonic (${\cal M}_s$=0.7, 0.1). A solid line is plotted
through each subgroup's average value at each lag. Attention to the difference
in scale between the right and left panel shows a sensitivity
to ${\cal M}_A$ as well.
Errors in the fit parameters are calculated from the standard deviation about each
subgroup's mean (i.e. highly supersonic, mildly supersonic, subsonic) for each lag. \
Error bars are omitted from
Figure \ref{fig:3dfinal} for clarity but are displayed in Figure \ref{fig:3d4} for $w^{-2}$.
The $a$ parameter has a percent error generally $<$ 25$\%$ but reaches a maximum of 56$\%$ for
simulation ${\cal M}_s$=3.0 ${\cal M}_A$=2.0 (square symbol) at a lag of 1 pixel. The deviation
from the mean mildly
supersonic value is not large compared to the supersonic group but the lower numerical
values produce a higher percent error. This error drops off significantly with increasing lag.
This is to be expected for simulations since at low lags we are in the range of numerical
dissipation. Super-Alfv\'enic errors all fall below 20$\%$.
In the middle panels, $q$ (related to the PDF's
tails) shows a slight sensitivity to the simulations compressibility and no
magnetic sensitivity. For $q$ we break the simulations up into only
supersonic (${\cal M}_s$=10, 7, 6, 4, 3) and subsonic (${\cal M}_s$=0.7, 0.1)
and plot a solid line through their average values. The minimal variation in $q$
between simulations results in low errors that are below a 20$\%$ for each lag.
Parameter $w$, or width, displayed in the bottom panels of Figure \ref{fig:3dfinal}
is presented as the fit value $w^{-2}$ as it appears in Equation (1)
for convenience and clarity of representation. As with $a$, we plot solid lines
of the average of highly super-, mildly super-, and subsonic simulations showing the same sensitivity
to ${\cal M}_s$ although to a lesser degree. Both subsonic simulations lie along the lag (x)
axis. The most outstanding trend seen in $w$ is its sensitivity to ${\cal M}_A$.
Taking note of the large difference in scales between sub- and
super-Alfv\'enic panels shows elevated values for simulations with high magnetic fields. Figure
\ref{fig:3d4} summarizes this sensitivity displaying the sub- and super-Alfv\'enic versions of
two highly supersonic (${\cal M}_s$=10, 7), one mildly supersonic
(${\cal M}_s$=4), and one subsonic (${\cal M}_s$=0.7) simulation. For each ${\cal
M}_s$, the sub-Alfv\'enic ($B_\mathrm{ext}$=1.0) simulation produces a $w^{-2}$ value
$\geq$ 2 times its super-Alfv\'enic counter part at lag = 1 pixel.
Errors for $w^{-2}$ are the highest of the three parameters due to it large deviation between
simulations. Generally the errors are $<$ 40$\%$ but peak at 160$\%$ for the ${\cal M}_s$=3.0
${\cal M}_A$=2.0 simulation (filled circle) at 1 pixel lag. This is mainly due to its $<$ 1 value.
Even with these significant errors, the differences between sub- and super-Alf\'enic turbulence
is apparent.
Aside from the values of $a$, $q$, and $w$, the general shape of all three
shows trends toward ${\cal M}_s$. Both $a$ and $w^{-2}$ show relatively
consistent values over lag for subsonic simulations while supersonic
simulations show steep decreasing slopes.
\begin{figure*}[tbh]
\begin{center}
\includegraphics[angle=90,keepaspectratio=true,scale=0.5]
{fig_04.eps}
\caption{An enlarged version of Figure \ref{fig:3dfinal}'s bottom panel with ${\cal
M}_s$= 10, 7, 4, 0.7 (top left to bottom right respectively).
Sub-Alfv\'enic simulations are denoted with red diamonds while super-Alfv\'enic
simulations are denoted with blue triangles. Both super- and sub-sonic simulations show
larger $w^{-2}$ for high magnetization. Error bars are calculated from the standard
deviation of the mean of each ${\cal M}_s$ group (highly supersonic, mildly supersonic,
subsonic) at each lag.}
\label{fig:3d4}
\end{center}
\end{figure*}
\subsection{3D Velocity and Magnetic Field}
\label{3dVB}
The analysis of velocity is more directly related to investigations
of turbulence than density and is particularly important in regards to
Solar Wind measurements. Tsallis provides excellent fits to our velocity PDFs and
also shows dependencies on Mach numbers, although the Tsallis fit parameters show decreased
sensitivities to ${\cal M}_s$ and ${\cal M}_A$ compared to the density analysis.
The top panels of Figure \ref{fig:bvx} displays
$w^{-2}$ for sub- and super-Alfv\'enic simulations on the left and right respectively
for the component of velocity along the
mean magnetic field (x-direction). Solid lines represent the average values of highly super-,
mildly super-, and sub sonic simulations at each lag (for top and bottom panels).
A small sensitivity to ${\cal M}_s$ can be seen with increased
values for more turbulent simulations. The sensitivity to ${\cal M}_A$ is most
predominant parallel to $<${\bf B}$>$ (which is the LOS shown in this figure)
where there is a near factor of 2
increase in $w^{-2}$. Perpendicular to $<${\bf B}$>$, the increase is $\le50\%$.
For sub-Alfv\'enic simulations ($B_\mathrm{ext}$=1) a $\sim$20\% increase in $w^{-2}$ is
seen for velocity along $<${\bf B}$>$ while no preferred direction is seen for
super-Alfv\'enic simulations.
We do not show figures for $a$ and $q$ as they displayed less sensitivity
to Mach numbers then did $w^{-2}$.
Generally for velocity, the $a$ parameter displays a very small sensitivity to
compressibility and the values for all simulations span a limited range.
Compared to 3D density, velocity exhibits a two order of magnitude decrease in
standard deviation for small lags. A slight increase in $a$ can be seen for
velocity of highly magnetized simulations along the mean magnetic field (x
direction) but no similar relationship is seen for velocity perpendicular to
$<${\bf B}$>$. Fit parameter $q$ is even less descriptive, with no significant
Mach number dependencies. $w^{-2}$ is the most sensitive to ${\cal
M}_s$ and ${\cal M}_A$ maintaining the same trends seen in density but on
a fraction of the scale.
Analysis of directional magnetic field strength is shown in the bottom of
Figure \ref{fig:bvx} for $w^{-2}$. Tsallis fits the PDFs of magnetic field well.
Parameter $w^{-2}$ displays a sensitivity
to the ${\cal M}_s$ number of simulations of a given ${\cal M}_A$ as seen in the bottom
panels of Figure \ref{fig:bvx} for the magnetic field parallel to $<${\bf B}$>$.
In the presented figures $w^{-2}$ has a similar value for high and
low ${\cal M}_A$ but along other lines of sight the relationship between $w^{-2}$
values and increased magnetization generally does not hold. However, this may
be useful for determining mean field direction in the ISM.
Similar to velocity, we find that $a$ and $q$ are not as useful as $w$ in terms of describing
Mach numbers with component velocity and magnetic field, and hence we omit the Figures.
Parameter $a$ shows a very small sensitivity to compressibility while $q$ shows no significant
variation for any simulation.
Errors analysis for the fit parameters of velocity and magnetic field are carried out
in the same manner as density (calculating the standard deviation of each ${\cal M}_s$
subgroup at each lag). Velocity along the mean magnetic field
has fit errors consistently $<$ 18$\%$ for all lags and simulations. Magnetic field along
the same LOS has similarly low errors ($<$ 20$\%$) with the exclusion of the sub-Alfv\'enic
($B_\mathrm{ext}$=1) ${\cal M}_s$=4 \& 3 simulations (triangles and squares respectively) which
reach 43$\%$ and 109$\%$ error respectively.
One should keep in mind, however, that these measurements are done for one LOS
as there are significant variations in the trends of velocity and magnetic field depending on
the orientation.
The behavior of Tsallis fits to PDF increments of density, velocity, and magnetic field
is in agreement with results found in EL10 at higher resolution with a larger parameter range.
Our analysis of spatial variations provides insight into the underlying physics of MHD
turbulence, where as the analysis of temporal variations of velocity and magnetic field
in solar wind observations well described the multiphase structure
of this phenomenon (Burlaga \& Vi\~{n}as 2004a, 2004b, 2005a, 2005b, 2006;
Burlaga, Ness, \& Acu{\~n}as 2006, 2007, 2009; Burlaga, Vi{\~n}as, \& Wang 2007).
In the latter a time scale,
$\tau_m$, (opposed to spatial scale $r$) is used to describe incremental
fluctuations by $(B(t+\tau_m)-B(t))$.
\begin{figure*}[bh]
\begin{center}
\includegraphics[angle=90,keepaspectratio=true,scale=0.45]
{fig_05.eps}
\caption{Top: Fit parameter $w^{-2}$ for the velocity component parallel
to the mean magnetic field for 14 simulations. Sub- and super-Alfv\'enic
simulations are presented on the left and right respectively. Three
solid lines are over-plotted averaging the super, trans, and subsonic
simulations.
Bottom: Fit parameter $w^{-2}$ for the magnetic field strength parallel
to the mean magnetic field for 12 simulations. Three
solid lines are over-plotted averaging the super, trans, and subsonic
simulations. Errors for $\textbf{V}_x$ are all $<$ 18$\%$. Errors for
$\textbf{B}_x$ are all $<$ 20$\%$, excluding mildly subsonic
(${\cal M}_s$=4 \& 3) sub-Alfv\'enic ($B_\mathrm{ext}$=1) simulations.}
\label{fig:bvx}
\end{center}
\end{figure*}
\begin{figure*}[tb]
\begin{center}
\includegraphics[angle=90,keepaspectratio=true,scale=0.45]
{fig_06.eps}
\caption{$\chi^2$ and $R^2$ versus lag on the left and right respectively for
3D density simulations. The sub- and super-Alfv\'enic simulations are over plotted
(red diamonds and blue triangles respectively) for
${\cal M}_s$=10 (top) and ${\cal M}_A$=0.1 (bottom). Tightest fits are obtained
for medium lags for low magnetization and low turbulence. Largest deviations are
seen for sub-Alfv\'enic supersonic simulations at small lags (in the dissipation range).}
\label{fig:chiR}
\end{center}
\end{figure*}
\subsection{Quality of Fits}
\label{error}
In order to characterize the quality of our fits and reliability of results,
we calculate the Pearson's $\chi^2$ and coefficient of determination, $R^2$,
for our Tsallis PDF fits presented in Equations (5) and (6) below.
\begin{eqnarray}
\chi^2 = \sum\limits_{i=1}^n \frac{(y_i - f_i)^2}{f_i}, \\
R^2 = 1-\sum\limits_{i=1}^n \frac{(y_i-f_i)^2}{(y_i-\bar y)^2},
\end{eqnarray}
\begin{figure*}[tbh]
\begin{center}
\includegraphics[angle=90,keepaspectratio=true,scale=0.6]
{fig_07.eps}
\caption{PDFs of column density of four simulations with LOS averaged parallel to the
mean magnetic field. From left to right, panels
show ${\cal M}_s$=10 ${\cal M}_A$=0.7, ${\cal M}_s$=10 ${\cal M}_A$=2.0,
${\cal M}_s$=0.7 ${\cal M}_A$=0.7, and ${\cal M}_s$=0.7 ${\cal M}_A$=2.0.
Column density made
perpendicular to the mean magnetic field produced similar PDFs. Gaussianity
of PDFs begins to degrade at high lags for subsonic simulations (top right)}
\label{fig:cdxhists1}
\end{center}
\end{figure*}
In these equations, $y$ represents the observed PDF and $f$ is the Tsallis fit.
Each measurement describes the accuracy of the fit. $\chi^2$ approaches
zero for perfect fits while $R^2$ approaches one. In general, the
Tsallis function will fit PDFs of density, column density, magnetic field, and
velocity with a $\chi^2$$\le$0.1 and a $R^2$$\ge$0.85 resulting in tightly
correlated fits for the 100 degrees of freedom (number of bins in PDF histogram).
Figure \ref{fig:chiR} displays the $\chi^2$ and $R^2$ for the sub- and super-Alfv\'enic
counterparts of our most turbulent (top) and least turbulent 3D density
simulations (bottom). From these plots we find that the best fits are seen for
super-Alfv\'enic, subsonic simulations for lags greater than 16 pixels.
Due to the kurtotic nature of high turbulence, low lag PDFs are difficult
for Tsallis to fit at the central peak (see Figure \ref{fig:3dhistsall}, right
panel, bottom). In these extreme cases (${\cal M}_s$=10, ${\cal M}_A$=0.7) values
of $\chi^2$=0.67 and $R^2$=0.61 are obtained.
The low lag regimes are in the dissipation range of
the turbulence, and Tsallis seems to have a difficult time fitting turbulence on scales sampling either
the dissipation range or the injection scale. Despite these difficulties at low (or high) lag, Tsallis still can prove itself
useful for characterizing turbulence. For example, if one is in the dissipation range of turbulence (i.e. at scales similar to our lags
less then 32 pixels), then these ``goodness of fit tests'' can also be used to
determine what type of turbulence is present. Subsonic and super-Alfv\'enic turbulence both are better fit with the Tsallis distribution at low lags then
supersonic or sub-Alfv\'enic turbulence. At higher lags the fit quality converges as we enter the inertial range of turbulence.
These fit tests should not only be preformed to test fit quality but could also be used as an additional test of the Mach number range in a given data set.
An alternative source of potential error in our analysis is the quality of PDFs
at large lags. Not only are these lags on the upward scale of the inertial range,
but also are subjected to degraded resolution. These resolution effects can be seen particularly in the case
of the column density, which are 2D quantities discussed in the next section.
The inspection of Figure \ref{fig:cdxall} shows fits becoming less tight at large lags. This manifests as ``jumps''
in the fits parameters which can be seen in the subsonic case in Figure \ref{fig:cdx4}. This is
the most extreme example of this trend but it is seen to some degree for every
simulation at a large enough lag. In these cases the $\chi^2$ and $R^2$ remain constrained to the
values stated above. We conclude that, for this analysis, our
results are most stable for lags in the inertial range of turbulence and for lags at least 8 times smaller then the box size to provide
enough sampling. This
trend is consistent with the results seen using higher order moments to
analyze the incremental PDFs.
\begin{figure*}[tbh]
\begin{center}
\includegraphics[angle=90,keepaspectratio=true,scale=0.5]
{fig_08.eps}
\caption{From top to bottom fit parameters $a$ (amplitude of fit),
$q$ (related to fitting tails of PDF), and $w$ (PDF width plotted
as $w^{-2}$) are displayed vs spatial lag for all 14 simulations. Column
density are LOS averaged parallel to the mean magnetic field.
The left and right columns correspond to sub- and super-Alfv\'enic simulations
respectively. Three solid lines are over-plotted
for $a$ and $w^{-2}$ averaging the highly supersonic (${\cal M}_s$=10, 7, 6),
transonic (${\cal M}_s$=4, 3), and subsonic (${\cal M}_s$=0.7, 0.1)
simulations. Two solid lines are over-plotted for the $q$ parameter
averaging the supersonic (${\cal M}_s$=10, 7, 6, 4, 3) and subsonic
(${\cal M}_s$=0.7, 0.1) simulations. Parameters $a$ and $w^{-2}$ show
sensitivities to ${\cal M}_s$ and ${\cal M}_A$ (note
scale between left and right columns). Parameter $q$
is only slightly sensitive to the simulation's compressibility. Lags from 2 to
32 pixels provide the most consistent fit parameter errors where $a$, $q$,
and $w^{-2}$ maintain errors $<$ 22$\%$, 35$\%$, and 24$\%$ respectively.}
\label{fig:cdxall}
\end{center}
\end{figure*}
\newpage
\section{Tsallis Fit of 2D Column Density and PPV data}
\label{observ}
In section \ref{3d} we confirmed the results of EL10 using higher resolution
simulations and a substantially larger parameter range. In addition, we
demonstrate the sensitivity the $w$ (width) fit parameter has toward the
Alfv\'enic Mach number. However, observations of the ISM do not provide
direct 3D information of density. Combining density and velocity along the
line-of-sight (LOS) provides a 3D position-position-velocity (PPV) cube,
however these types of data cubes can almost never be reliably interpreted
as having a one-to-one correspondence with an actual 3D volume density.
Considering the difficulties of obtaining direct 3D ISM information, a
study of Tsallis statistics on observables such as column density and PPV
data is necessary.
\subsection{Column Density}
\label{coldn}
We create synthetic 2D column density maps perpendicular and parallel to
the mean magnetic field for all 14 of our simulations by averaging the 3D
density cubes along a given line of sight. We assume the emitting gas is
optically thin and that the emissivity is linearly proportional to density
(such as in the case of HI). An example is presented in Figure
\ref{fig:3dpic} (bottom left). Using the same method described in the section
\ref{tsallis}, PDFs are created using spatial lags (increments) in \emph{two}
directions which are then fit using the Tsallis function.
Figure
\ref{fig:cdxhists1} displays the distributions and fits
for 4 simulated column densities created parallel to the mean magnetic field.
The 4 simulations we show here are divided into panels:
${\cal M}_s$=10 ${\cal M}_A$=0.7, ${\cal M}_s$=10 ${\cal M}_A$=2.0,
${\cal M}_s$=0.7 ${\cal M}_A$=0.7, and ${\cal M}_s$=0.7 ${\cal M}_A$=2.0
(left to right). This is analogous to the arrangement in Figure
\ref{fig:3dhistsall}. While there is an
overall decrease in kurtosis of the column density PDFs compared to 3D density,
the same three trends are still apparent. PDFs become more Gaussian with
decreased ${\cal M}_s$, Gaussianity increases with lag for supersonic cases, and
Gaussianity increases with ${\cal M}_A$. A trend not seen in the 3D case is that
subsonic PDFs become less smooth at high lags. This is due to the decrease in
resolution going from 3D to 2D. EL10 also observed this trend
with their subsonic PDFs.
\begin{figure*}[tbh]
\begin{center}
\includegraphics[angle=90,keepaspectratio=true,scale=0.5]
{fig_09.eps}
\caption{An enlarged version Figure \ref{fig:cdxall}'s bottom panel with ${\cal M}_s$= 10,
7, 4, 0.7 (top left to bottom right respectively). High magnetization is
denoted with red diamonds and low magnetization is denoted with blue
triangles. Both super- and subsonic simulations show larger $w^{-2}$
for high magnetization. Error bars are calculated from the standard
deviation of the mean of each ${\cal M}_s$ group (highly supersonic, mildly supersonic,
subsonic) at each lag.}
\label{fig:cdx4}
\end{center}
\end{figure*}
Figure \ref{fig:cdxall} displays the fit parameters $a$,
$q$, and $w^{-2}$ from top to bottom for all 14 column density
simulations created along the x direction (parallel to $<${\bf B}$>$).
Each parameter is divided into two panels with
sub-Alfv\'enic and super-Alfv\'enic simulations on the left and right respectively. The
$a$ parameter (top panels) is presented with three over-plotted
lines, averaging the super-, mildly super-, and subsonic
simulations. Again, a strong sensitivity to ${\cal M}_s$ is seen both in the
value and the shape of the fit over spatial scale. The sensitivity to ${\cal
M}_A$ is less prevalent than in 3D density.
Errors for column density fit parameters are calculated in the same manner as 3D
density by measuring the standard deviation of each ${\cal M}_s$ subgroup at each
lag. The $a$ parameter achieves $<$ 23$\%$ errors for each lag and simulation
reaching the lowest errors ($<$ 5$\%$) at lags between 16 and 64 pixels. As in density,
error bars are excluded from Figure \ref{fig:cdxall} for clarity but displayed in
Figure \ref{fig:cdx4} for $w^{-2}$.
Parameter $q$ (middle panels)
is presented with averaging super- and subsonic lines showing
a slight sensitivity to ${\cal M}_s$ and no coherent sensitivity to ${\cal
M}_A$. Errors for $q$ are generally less than 30$\%$ but begin to reach
percent errors greater than 100$\%$ for subsonic simulations at lags greater
than 34 pixels due to random variations. Inspection of the right panel of Figure
\ref{fig:cdxhists1} shows that this is consistent with when the PDF start to become
nonuniform.
The $w$ parameter (bottom panels), here plotted as $w^{-2}$,
shows both a strong sensitivity to ${\cal M}_s$ and ${\cal M}_A$. Highly super-,
mildly super-, and subsonic averaging lines are over-plotted which emphasize that values of $w^{-2}$ are affected by ${\cal
M}_s$. Errors for $w^{-2}$ are $<$ 25$\%$ for lags of 2 to 32 pixels. Maximum
errors reach values of 70$\%$ at a 1 pixel lag and 60$\%$ error at lags greater than
64 pixels. Interestingly, as in EL10, subsonic column densities shows a more random behavior
at larger lags due to low resolution and low density contrasts.
Figure \ref{fig:cdx4} summarizes $w$'s Alfv\'enic sensitivity by
displaying the sub- and super-Alfv\'enic versions of two highly supersonic (${\cal M}_s$=10,
7), one mildly supersonic (${\cal M}_s$=4), and one subsonic (${\cal M}_s$=0.7)
simulation. In each case, the simulation having the higher magnetic field
results in a higher $w^{-2}$ value. Although the degree to which $w^{-2}$ is
elevated is less than for 3D density, a strong direct correlation remains.
Subsonic trends become less smooth at lag=32 pixels, which is $\approx$ 6.4 times smaller
then the injection scale. We may conclude that the column density has a stronger
dependency on lag then in the 3D cases, especially for subsonic turbulence.
We analyze fit parameters from column densities created along the y and z axes
(i.e. perpendicular to the mean magnetic field) and while
there were slight variations in the fits, sensitivities to Alfv\'en and
sonic Mach numbers were still observed. No discernible trend to detect
magnetic LOS orientation was seen.
\begin{figure*}[tbh]
\begin{center}
\includegraphics[angle=90,keepaspectratio=true,scale=0.5]
{fig_10.eps}
\caption{Fit parameter $w^{-2}$ for ${\cal M}_s$= 10 (top), and 0.7
(bottom) with sub- and super-Alfv\'enic simulations on the left and right
respectively. The original and smoothed data are over-plotted. Least squares
linear fits are over plotted in the same color. The enclosed legends displays
the slope of the respective linear fits. The central legend provide the
FWHM of smoothing (FWHM=0 is unsmoothed). Increased smoothing results in
decreased values and shallower slopes.}
\label{fig:lag1}
\end{center}
\end{figure*}
\begin{figure}[tbh]
\begin{center}
\includegraphics[angle=90,keepaspectratio=true,scale=0.3]
{fig_11.eps}
\caption{Slope of $w^{-2}$ (width) for the sub- and super-Alfv\'enic versions of the
${\cal M}_s$=7, 4, and 0.7 simulations plotted against the smoothing degree in FWHM
size. High and low magnetization are presented in red and blue
respectively. Supersonic simulations are more dramatically affected than
subsonic as turbulent density enhancements are smoothed out.}
\label{fig:mvsmooth}
\end{center}
\end{figure}
\subsubsection{Effects of Smoothing}
\label{smooth}
While the analysis simulated column density maps shows the
strength of Tsallis statistics in describing turbulent characteristics, analogous observational
column density data do not have pencil thin beam resolution. To address this issue,
we experiment with smoothing our column densities to different degrees to determine
the role degradation of resolution plays in distributions and fit
sensitivities. Using a Gaussian smoothing kernel, we degrade our column
density simulations with a full-width-half-maximum (FWHM) of 3, 5, and 10
pixels.
In an effort to characterize the effects of smoothing we increase our lag
resolution by creating the same number of distributions over a smaller
range of lags. Figure \ref{fig:lag1} presents the the original and smoothed
fit parameter $w^{-2}$ for sub- and super-Alfv\'enic simulations versions of simulations ${\cal
M}_s$=10 and 0.7 from top left to bottom right respectively. While we investigated the effects of smoothing
on all three fit parameters, we only present $w^{-2}$ as the effects on $q$ and $a$ are similar.
In order to characterize how the slope of the
Tsallis parameters vs. lag change with increasing smoothing a least squares
linear fit is applied to each scenario with the slopes displayed in the
respective legends. The red stars denote the non-smoothed case for comparison.
From this figure it is clear that increased smoothing
lowers the values of $w^{-2}$ for each lag and its overall slope.
These effects are present across all simulations.
While the supersonic (top row) simulations show smooth monotonic trends for
both smoothed and pencil beam data, subsonic simulations (bottom row) shows very bumpy
trends for the case where no smoothing is introduced (red stars and line)
especially as lag increases. However, as we introduce smoothing, the trends
become highly monotonic, even for the case of FWHM=3 pixels. Applying Tsallis
fits to incremental PDFs while varying the level of smoothing will act as an
additional tool to characterize turbulent parameters through the overall
change in slope.
Figure \ref{fig:mvsmooth} provides a summary of the effect smoothing has on
the slope of $w^{-2}$ by plotting it against the smoothing degree.
Simulations of the same sonic number are shown with the same symbol while
the color corresponds to the previously used color scheme (red being $B_\mathrm{ext}$=1,
sub-Alfv\'enic and blue being $B_\mathrm{ext}$=0.1, super-Alfv\'enic).
We see that increased smoothing flattens the $w^{-2}$-lag slope. A
sensitivity to the ${\cal M}_A$ is still present in both slope and value in
every smoothing cases (note scale between sub- and super-Alfv\'enic simulations in Figure
\ref{fig:lag1}). High sonic Mach number simulations show slopes that are the most altered
as smoothing is increased. This is due to shock density enhancements being smoothed out.
However, the supersonic cases show larger variation with differing ${\cal M}_A$ then
their subsonic counterparts. This leads us to believe that one can more easily tell the
magnetization strength of supersonic gas from Tsallis fits. Increasing the smoothing does
not affect the subsonic cases, as these simulations already have Gaussian distributions and have no peak density enhancements
from shocks to smooth out.
Parameter $a$ was affected by smoothing in the exact same way $w^{-2}$. Values and
slopes decreased with increased smoothing. Plots are excluded for brevity but both $a$ and
$w$ could be explored with increasing smoothing to put estimates on turbulence.
The $q$ parameter saw little deviations due to smoothing and a slightly weakened
${\cal M}_s$ sensitivity.
Fit parameter errors for our smoothing discussing can be considered to decrease with
smoothing from their unsmoothed values (quoted in the previous section) as differences
in ${\cal M}_s$ sub groups of simulations become washed out in parameter space.
\subsubsection{Artificial Noise}
\label{noise}
While analysis using Tsallis statistics is very adept at providing turbulence
parameters when smoothing
is applied to column density maps, more pressing maybe the issue of noise.
One may not expect noise to affect results at larger lags however, noise may make it
challenging to distinguish trends on smaller scales where the difference between
our simulations is prominent. It is our intent in this subsection to test the
effectiveness of this statistical tool with the addition of noise, determine our confidence
range between trends with noise and with no noise, and explore which
lags are most affected by the addition of noise. In order to achieve this end
we add random Gaussian noise to our column density maps. We do this by setting a
given average signal-to-noise (SNR) ratio and scaling the power of the noise and
signal to match this SNR. We look at column density integrated parallel to the mean
magnetic field with average SNR of 400, 200, 100, 50, 20 and unity.
\begin{figure*}[bh]
\begin{center}
\includegraphics[angle=90,keepaspectratio=true,scale=0.50]
{fig_12.eps}
\caption{Fit parameter $a$ for 4 simulations (sub and super-Alfv\'enic of simulations
${\cal M}_s$=7, 0.7) with varying levels of noise (see legend). Small amounts of
noise drastically
lower small lags values and has an overall smoothing affect. Increasing noise lowers $a$'s
sensitivity to ${\cal M}_s$ and ${\cal M}_A$.}
\label{fig:noisea}
\end{center}
\end{figure*}
\begin{figure*}[tb]
\begin{center}
\includegraphics[angle=90,keepaspectratio=true,scale=0.50]
{fig_13.eps}
\caption{Fit parameter $w^{-2}$ for 4 simulations (sub and super-Alfv\'enic of simulations
${\cal M}_s$=7, 0.7) with varying levels of noise (see legend). Small amounts of noise
drastically lower small lags values and has an overall smoothing affect. Increasing
noise lowers $w^{-2}$'s sensitivity to ${\cal M}_s$ and ${\cal M}_A$.}
\label{fig:noisew}
\end{center}
\end{figure*}
The addition of noise on the column density PDFs has a general smoothing effect
toward Gaussianity (PDFs not shown). This is to be expected as the noise added
to the column density is Gaussian white noise. Figures \ref{fig:noisea} and
\ref{fig:noisew} illustrate the effects of noise for the parameters $a$ and $w^{-2}$
respectively for four simulations with given sonic Mach numbers of ${\cal M}_s$= 7.0
and 0.7 and Alfv\'enic number ${\cal M}_A$= 2.0 and 0.7. From these figures it is
clear that the smaller spacial lags (lags = 1 - 32 pixels) are highly affected by the
addition of noise. The noise injected simulations (lines with colors and symbols) do
not converge to their no-noise counter part (shown as a solid black line) until lag
$\sim$20 pixels. At around lag 20, simulations with SNR greater than 50 show similar
shaped curves to the no-noise case. As the SNR decreases, the values of $w$ and $a$ both
decrease for the supersonic case. The subsonic case shows a more pronounced non-monotonic
behavior. In all cases above SNR of 20, the sub-Alfv\'enic supersonic turbulence shows
heightened values of $a$ and $w^{-2}$. The sonic number between these
two simulations can be determined within 1.5 $\sigma$ confidence at SNR 20 and above.
At SNR of unity it is impossible to distinguish between either sonic or Alfv\'enic numbers, which is to be expected.
A SNR greater than 100 is sufficient for showing differences in $w^{-2}$ and $a$ between
sub- and super-Alfv\'enic simulations, with sub-Alfv\'enic producing larger values.
In the case of supersonic turbulence a SNR of 20 is generally sufficient to discern between
sub- and super-Alfv\'enic gas using Tsallis fit parameters with 1.5$\sigma$ confidence
for all lags. While this is promising that all pixel scales show statistically significant
differences, the shape of the $w^{-2}$ or $a$ curve vs. lag converges to non-noise levels only for
lag $\sim$20 pixels and therefore, small lags may not be as reliable as larger lags in the presence of low SNR. Subsonic cases can
distinguish Alfv\'enic number with confident in the 1 $\sigma$ range at SNR 20 and greater
for lags larger then 32 pixels.
\subsubsection{Simulating Cloud Boundaries}
\label{cloud}
Studies of the ISM frequently deal with both clouds-like objects as well as diffuse gas.
While our simulations are most directly applicable to diffuse
ISM, we can also study the effects the Tsallis fits will have on ISM gas with
boundaries. An example of such a situation would be molecular clouds
which have characteristic radially decreasing density values. To mimic this,
we convolve our 3D density simulations with a spherical function which
maintains the value inside a given radius R from the center and creates a Gaussian decay
outside this radius. For our simulations we choose an R of 205,
90, and 10 pixels corresponding to 2.5, 6, and 51 times smaller
than box size respectively. The cloud convolved 3D simulations are then
averaged along a LOS creating a column density map as described in
section \ref{coldn}. For reference a column density map
with a radial decreasing Gaussian boundary starting at R=90 parallel to the mean magnetic field
is presented in Figure \ref{fig:3dpic} (bottom right) for simulation \#2 (${\cal M}_s$=10,
${\cal M}_A$=0.7).
\begin{figure*}[tbh]
\begin{center}
\includegraphics[angle=90,keepaspectratio=true,scale=0.5]
{fig_14.eps}
\caption{From top to bottom fit parameters $a$ (amplitude of fit),
$q$ (related to fitting tails of PDF), and $w$ (PDF width plotted
as $w^{-2}$) are displayed vs spatial lag for all 14 simulated cloud bounded column
density simulations created parallel to the mean magnetic field (R=205, 2.5 times
smaller than cube and the scale of the turbulent energy injection).
The left and right columns correspond to sub- and super-Alfv\'enic simulations
respectively. Three solid lines are over-plotted
for $a$ and $w^{-2}$ averaging the supersonic (${\cal M}_s$=10, 7, 6),
transonic (${\cal M}_s$=4, 3), and subsonic (${\cal M}_s$=0.7, 0.1)
simulations. Two solid lines are over-plotted for the $q$ parameter
averaging the supersonic (${\cal M}_s$=10, 7, 6, 4, 3) and subsonic
(${\cal M}_s$=0.7, 0.1) simulations. Parameters $a$ and $w^{-2}$ show
sensitivities to ${\cal M}_s$ and ${\cal M}_A$ (note
scale between left and right columns). Parameter $q$
is only slightly sensitive to the simulation's compressibility. Lowest error
are achieved for supersonic simulations at lags between 4 and 32 pixels
where percent errors are below 44$\%$, 40$\%$, and 84$\%$ for $a$, $q$, and $w^{-2}$
respectively. Outside this constraint, error $>$ 150$\%$ for each parameter.}
\label{fig:cloudall}
\end{center}
\end{figure*}
In general, a radially decreasing cloud boundary creates an increase in the
kurtosis of the PDFs as compared with unbounded column density (not shown). This change
comes from the increase in zero point values of $\rho(x+r)-\rho(x)=0$ as more
contrast is created. At a radius of 205 pixels the PDFs are still fairly uniform and
only see slight variations from Gaussianity for subsonic simulations and fits
remain similar to those seen in Figure
\ref{fig:cdxhists1}.
Figure \ref{fig:cloudall}
presents the $a$ (top), $q$ (middle) and $w^{-2}$ (bottom) fit parameters for the R=205
bounded simulations parallel to the mean magnetic field. As with 3D density and
unbounded column density, sub- and super-Alfv\'enic simulations are presented
on the left and right panels respectively. Three averaging lines are plotted
through the highly super-, mildly super-, and subsonic simulations for $a$ and $w$, while
$q$ has only super- and subsonic averaging lines. Sensitivities to ${\cal
M}_s$ are still apparent throughout the $a$ and $w^{-2}$ fits values and slopes but
emphasized more in sub-Alfv\'enic simulations. ${\cal M}_A$ remains strongly tied to $w$,
as presented in Figure \ref{fig:cloud4}. In each scenario, the simulation
containing a stronger magnetic field results in a $\geq$2 times higher value
of $w^{-2}$. The $q$ parameter shows little effect from a radially
decreasing boundary condition and still is only mildly affected by
simulation compressibility.
Fit parameter errors of cloud bounded column densities shows similar trends to those seen
in full resolution 2D density maps. Parameter $a$ maintains the lowest error from lag 2 to
32 pixels where errors remain $<$ 16$\%$. Outside this range, errors increase and reach a
maximum error of 46$\%$ at the 1 pixel lag. $q$ achieves errors below 36$\%$ for most
simulation lags excluding subsonic simulations and large lags. In this regime error reaches
90$\%$. Errors of the $w^{-2}$ parameter display similar trend as $q$ where errors below
50$\%$ are produced for mid range lags (4-32 pixels) and highly and mildly supersonic simulations. Errors
reach $>$ 150$\%$ at small lags and subsonic simulations at high lags. This further confirms
the trend that decreasing resolution has on the quality of fits and fit parameter values.
Even as errors increase, the $w^{-2}$ parameter is still able to distinguish ${\cal M}_A$ within the error bars.
Similar to the incremental column density PDFs, our largest clouds proved well
described by Tsallis statistics. However, the decrease in cloud size greatly affects
the PDFs and the ability of the Tsallis function to fit them. For example, in the case of our smallest
cloud (R=50 pixels), the PDF and fit for simulation \#2 (${\cal
M}_s$=10, ${\cal M}_A$=0.7) at even our largest lag (128 pixels, not shown), is extremely
kurtotic and skewed to negative values.
This is due to radially decreasing density, the value of $\rho(x+r)-\rho(x)$ is in general $>$ 0 for
three out of the four directional calculations (+x, -x, +y, -y). Tsallis is
a symmetric function about $\rho(x+r)-\rho(x)=0$ and this skewness of the
PDF prevents tight fits. While quality of the fit is no longer
intact, the relative amplitudes and widths change with lag, ${\cal M}_s$,
and ${\cal M}_A$.
Cloud convolved column densities with LOS perpendicular to the mean
magnetic field were also analyzed and showed the same trends and no discernible
correlation with the LOS orientation.
\begin{figure*}[tbh]
\begin{center}
\includegraphics[angle=90,keepaspectratio=true,scale=0.5]
{fig_15.eps}
\caption{An enlarged version of Figure \ref{fig:cloudall}'s bottom panels
for ${\cal M}_s$
= 10, 7, 4, and 0.7. Sub- and super-Alfv\'enic simulations are over-plotted in red
diamonds and blue triangles respectively. Both Super- and Sub-sonic
cases have elevated values for high magnetization. Error bars are calculated from the standard
deviation of the mean of each ${\cal M}_s$ group (highly supersonic, mildly supersonic,
subsonic) at each lag.}
\label{fig:cloud4}
\end{center}
\end{figure*}
\subsection{PPV Cubes}
\label{ppv}
The application of Tsallis statistics has proven itself to be a useful tool
for exploring a wide parameter range
of sonic and Alfv\'enic Mach numbers in density and column density data.
However, while density fluctuations are indeed useful for characterizing
turbulence, more appropriate is the use of velocity information.
We test the usefulness of Tsallis PDF fits on Position-Position-Velocity (PPV)
data in order to see if additional information is provided by including the
velocity axis. We create synthetic PPV cubes of our full simulation range
using density and velocity with LOS perpendicular to mean magnetic field. In
order to test a variety of velocity resolutions we create PPV cubes with
velocity widths of 0.15 ($\sim$30 channels), 0.07 ($\sim$60 channels), and
0.007 ($\sim$600 channels). Units remain scale free.
Figure \ref{fig:PPV} presents
$w^{-2}$ vs. lag for the sub- and super-Alfv\'enic versions of the ${\cal M}_s$=10, 4, and 0.7
simulations from top to bottom respectively. Each panel has the three velocity resolutions
over-plotted for comparison. Low, middle, and high resolutions are represented by the
circle, diamond, and triangle respectively. Within each resolution, red symbols correspond
to sub-Alfv\'enic ($B_\mathrm{ext}$=1.0) while blue corresponds to super-Alfv\'enic ($B_\mathrm{ext}$=0.1).
All three panels display a common trend of
decreasing $w^{-2}$ value as the velocity resolution increases
(width of the velocity channels decreases). Also, it is clear that the $w$
parameter is sensitive to the ${\cal M}_A$ number for PPV simulations and produces
elevated values for simulations with high magnetization.
The general shape of $w^{-2}$ over increasing lags is distinctly different for
PPV data than in density, velocity, or magnetic field. The PDFs, and therefore fit parameters
are more consistent with lag variations and produce smoother trends. This alternative trend
is only seen in supersonic PPV data while values of $w^{-2}$ for subsonic PPV data more closely
resemble the trends seen in density analysis.
Parameter $a$ shows minimal variations to ${\cal M}_A$ or ${\cal M}_s$ for PPV simulations.
Sensitivities to ${\cal M}_s$ are seen on a scale too small to be used observationally and
there is no coherent relationship to ${\cal M}_A$. Similar to previous discussions $q$ only
provides limited information on compressibility and none on ${\cal M}_A$.
It should be noted that during the distribution creation process, especially
for high velocity resolutions, velocity channels with little emission were
excluded. In addition, a majority of the small lag PDFs had a single multiple order of magnitude jump
at the center of the distribution (where $\rho(x+r)-\rho(x)=0$). This point was
excluded from the fitting process. This is simply due to the way the emission is spread over the channels in PPV with a Maxwellian velocity distribution.
Most gas is near the mean velocity and a Gaussian decrease in emission is seen as one goes to the extreme minimum and maximum channel
velocities.
Analysis of density distributions along the velocity axis for individual pixels were also
analyzed. An attempt to fit these distributions with the Tsallis function proved ineffective.
We find the distribution of density emission along velocity channels, especially for high
turbulence, are highly skewed and irregular which is consistent with the findings of Falgarone
et al. (1994). As Tsallis is a symmetric function it is not able to provide tight fits to these
highly irregular and varying distributions.
\begin{figure*}[tbh]
\begin{center}
\includegraphics[angle=90,keepaspectratio=true,scale=0.5]{fig_16.eps}
\caption{Parameter $w$ (PDF width plotted as $w^{-2}$) for ${\cal M}_s$= 10, 4, 0.7 from
top to bottom respectively. Each panel plots the sub- and super-Alfv\'enic simulations for the three
velocity channels (see legend). A sensitivity to ${\cal M}_A$ is seen in every case but
more pronounces for high turbulence simulations and few velocity channels. Subsonic
simulations shows trends that resemble density while supersonic simulations do not.}
\label{fig:PPV}
\end{center}
\end{figure*}
\section{DISCUSSION}
\label{disc}
\subsection{Summary and the Relation of Tsallis to Other ISM Statistics}
We studied Tsallis statistics as a way to characterize MHD ISM turbulence. While we do not address
the theoretical justification of the particular form of the fits predicted by the Equation (1),
our study shows that this fit provides a very good correspondence with experimentally obtained
distributions. We showed that the parameters that enter the Tsallis expression are dependent on
the Sonic Mach (${\cal M}_s$) and Alfv\'enic Mach (${\cal M}_A$) numbers, confirming and
extending the earlier claims of Esquivel \& Lazarian (2010). In addition, we explored the
differences in the dependencies of the fitting parameters on both ${\cal M}_A$,
${\cal M}_s$ and on the resolution. This allows one to find ${\cal M}_s$ and ${\cal M}_A$
independently. We also applied Tsallis statistics to PPV data, which, as far as we know,
is the first time such a study has been undertaken.
Another important point that we address here is the influence of telescope resolution in the form of smoothing.
This issue was largely omitted in our earlier publications, but our results show that
the influence of the telescope beam may be important for observational studies. Not only does smoothing affect
the statistics, but it could actually be a helpful addition to the statistical procedure to study Mach numbers.
Indeed, telescope measurements introduce their own smoothing,
which depends on the telescope resolution. As the injection scale of the turbulence is
frequently not known, the given measurements of the parameters of the Tsallis
statistics may not be straightforward to identify with the particular simulation. In
this situation, additional smoothing can help to see to what parameters of turbulence
the observed data corresponds to.
Burkhart et al. (2010) studied the Small Magellanic Cloud (SMC) to characterize the
properties of turbulence in its individual parts. The same approach is applicable with
Tsallis statistics. It is convenient to note, that one can gain information on regions
turbulence or magnetization without knowing the precise resolution of the telescope. Our
motivation for applying boundaries to our simulations was fueled by making Tsallis applicable to
more compact objects, such as those that may be found in molecular clouds.
Our results show clear tendencies to determine Mach number information from the Tsallis
fitting parameters.
In addition, our study shows that the use of the full PPV data can bring additional
advantages. For instance, the evolution of Tsallis parameters with is clearly different for
subsonic and supersonic cases (see Figure \ref{fig:PPV}) in PPV data.
We feel that the Tsallis fit provides useful way of studying turbulence, which should
be implemented along with other techniques used previously (see Kowal, Lazarian, \& Beresnyak 2007;
Burkhart et al. 2009, 2010). We advocate an approach which combines the use of different
techniques. For instance, the effect of the telescope beam smoothing depends on the ratio of
the angular size of the turbulence energy scale in the object under study and the telescope
resolution. The former can be obtained by studying the power spectra of turbulence.
While we are still searching for the basic approaches to study turbulence observationally,
the successes in theoretical and numerical studies provide us with a reliable guidance for
obtaining statistics of turbulence from observations. The development of the VCA
and VCS techniques (see Lazarian \& Pogosyan 2000, 2004, 2006, 2008, Chepurnov \& Lazarian
2008) for instance, quantified how velocity and density create structures observed in the PPV.
Results on the anisotropies of observational data statistics in the plane of
the sky in Lazarian, Pogosyan \& Esquivel (2002) and subsequent publications
(Esquivel \& Lazarian 2005) reveals strong dependencies on media magnetization.
Kurtosis and skewness measures of the density (Kowal, Lazarian, \& Beresnyak 2007; Burkhart et al. 2009)
provide good insight mostly into the intensity of turbulence, i.e. its ${\cal M}_s$. Genus
(see Lazarian 1999, Kim \& Park 2007, Chepurnov et al. 2008) provides a measure of topology,
while bispectrum (see Burkhart et al. 2009) provides insight into mode correlation and phase
information of the turbulence\footnote{The latter two measures were borrowed from
cosmological studies and their use for ISM studies was proposed in Lazarian (1999).
The process of appreciating of their value for the ISM is currently under way.}.
Additional information can be obtained by employing other tools, e.g. dendrograms
(Rosolowsky et al. 2008, Goodman et al. 2009, Burkhart et al. 2011). At the same time the
synergy of the simultaneous use of different techniques is still to be explored.
\subsection{Relation of This Study to Previous Tsallis Work}
\label{relate}
In this paper we fit the Tsallis distribution function to PDFs of turbulence with
little or no theoretical justification. Indeed, while
the history of turbulence studies can be traced back to Kolmogorov (1941), turbulence studies have
since taken many different directions. In this paper, we are particularly interested
in characterizing magnetized turbulence in the ISM. The Tsallis distribution
formalization comes from a different background, specifically the issue of quantities
which are invariant in the Navier-Stokes equations at high Reynolds numbers and the
assumption that the singularities due to the invariance distribute themselves multifractally
in physical space. Much work has been done to address both the theoretical frame work and
application of the Tsallis statistics to numerical and laboratory turbulence PDFs
(see Tsallis 1999 and Arimitsu \& Arimitsu (2000a, 2000b, 2001, 2002, 2003)). For example, in Arimitus \&
Arimitus 2003, the authors successfully fit the Tsallis distribution to PDFs of turbulence
in three different laboratory experiments. Burlaga and collaborators successfully fit
Tsallis to solar wind data (Burlaga, Vi{\~n}as, \& Wang 2009).
In this paper, we successfully fit Tsallis
to our numerical magnetized ISM turbulence, even in the case of observable quantities such
as column density and PPV emission cubes. In the case of the laboratory, space observations,
and numerical experiments; the geometry of the system, the Reynolds numbers, and the type of
turbulence in question, are all very different. Never-the-less, in all cases the Tsallis
distribution provided good fits to the data. Motivated by these past studies, we applied
the Tsallis function in an \emph{empirical} way to determine the parameters of turbulence, and
we direct the curious reader to the literature mentioned above for a more in depth treatment
of the theoretical foundations of the Tsallis function. Also, our paper is motivated towards
making Tsallis statistics useful for observational studies of turbulence, and in this case
the discussion of multifractal theory is beyond our current scope. We do not attempt to justify
or even test these ideas, but investigate to what degree the fit provided by the two parameter
Tsallis formula corresponds to our numerical data.
\subsection{The Issues of Numerical Resolution}
\label{numissue}
While there is no doubt that substantial progress has been made in last decade in numerical
simulations of turbulence, the community still has a long way to go in terms of achieving the
resolution required to fully capture ISM physics. One issue in particular is the discrepancy
between astrophysical values of Reynolds (Re) and magnetic Reynolds (Rm) and the numerical values.
As a rule, astrophysical environments are turbulent and astrophysical turbulence is
characterized by enormously large Re and Rm. For instance, the Rm number, which characterizes
the degree of frozenness of magnetic field within eddies, may differ in astrophysical environments
and numerical simulations by a factor larger than $10^{10}$. With these sorts of discrepancies
in mind (which exist for all numerical codes that simulate turbulence), how are we able to
justify our results?
There are several avenues to justify the independence of our results from Re and Rm values.
The particular numerical code used here provides enough resolution to allow us to see the
inertial range of turbulence. If turbulence is evolving along the inertial range it is
self-similar, and the change in Re and Rm does not change the structure of the large scale
motions which we study here. Indeed, according to the Lazarian \& Vishniac (1999) model of turbulent
reconnection (tested in Kowal et al. 2009), magnetic reconnection in
a turbulent media depends on the large scale field wondering determined by large scale motions.
Thus we do not expect the unresolved small scales to affect the physics of our simulations at
large scales that we study. Keeping all of this in mind, we also tested our Tsallis fits at
resolutions of $256^3$ and $512^3$ and found no deviation in how well Tsallis was able to both
fit the PDFs and describe the turbulence in question.
\section{CONCLUSIONS}
\label{conc}
In this paper we applied the Tsallis formalism to
simulated diffuse ISM isothermal ideal MHD turbulence using incremental PDFs of density,
velocity, and magnetic field. This method was also applied to simulated column
densities with measures taken to duplicate observed measurements such as
smoothing, noise, and radial decreasing cloud boundaries. The use of Tsallis
statistics on PPV data was also explored. A summary of our findings is as
follows:
\begin{itemize}
\item The Tsallis function is capable of well describing incremental PDFs of
a wide range of simulated MHD density, velocity, and magnetic field with
its three fit parameters $a$ (amplitude), $w$ (width), and $q$ (related to the
PDF's tails).
\item For 3D density, fit parameters $a$, $q$, and $w^{-2}$ are sensitive to
${\cal M}_s$. $a$ and $w^{-2}$ show sensitivities to ${\cal M}_A$ with
both displaying greater values for sub-Alfv\'enic simulations.
\item Magnetic field and velocity show sensitivities to ${\cal M}_s$
and ${\cal M}_A$ but these are not as strong as density.
\item PDFs of column density are equally well described by the Tsallis distribution.
Fit parameters $a$, $q$, and $w^{-2}$ remain sensitive to ${\cal M}_s$. $a$
looses some of its magnetic sensitivity but $w^{-2}$ consistently produces
elevated values for sub-Alfv\'enic simulations.
\item The affect of degrading resolution lowers both
$a$ and $w^{-2}$ vs. lag monotonically. Supersonic simulations are most
affected by smoothing. $a$ and $w^{-2}$ remain sensitive to ${\cal M}_s$ and
${\cal M}_A$ for mild smoothing. $q$ looses ${\cal M}_s$ sensitivity with
smoothing.
\item The addition of smoothing at varying degrees acts as an additional method to
constrain sonic Mach number through the analysis of the slope of
parameters $a$ or $w^{-2}$ over lag.
\item Noise has the largest affect on fit parameters. Small scale variations
strongly alter the PDFs of low lags and the shape of the fit
distributions. Sensitivities to ${\cal M}_A$ and ${\cal M}_s$ are still
seen on smaller scales as the spatial lag increases.
\item Radially deceasing cloud boundaries have little affect as long as a
there are enough points inside providing good signal. As clouds become
smaller, PDFs become more kurtotic and skewed preventing meaningful fits.
\item PPV data was fit with Tsallis distribution excluding the zero point and
parameter $w^{-2}$ showed sensitivities to
${\cal M}_s$ and ${\cal M}_A$.
\item Tsallis statistics of incremental PDFs is a successful tool in describing a
wide range of high resolution MHD simulations and their observational counter parts.
Along with its sensitivities to turbulence and magnetic fields, the Tsallis
distribution is highly complimentary to power spectrum and other ISM statistical tools.
\end{itemize}
\bibliographystyle{apj}
|
\section{Atomic model of lithium niobate substrates}
\begin{figure}
\centering
\includegraphics[angle=0, width=0.45\textwidth]{figS1.pdf}
\caption{Geometry of five-trilayer LiNbO$_3$ slab: (a) side view,
(b,c) top view of $c^+$, and (d,e) top view of $c^-$. Slabs are
represented using ball and stick (a) and space filling (b,d)
models, or the simplified model used in the main text (c,e). Blue
arrows in (a) denote ions (Li$'$ and O$'$) introduced for surface
charge passivation. Black lines in (b) and (d) indicate the
boundaries of the LiNbO$_3$ primitive surface unit cell. (b) and
(d) are redrawn from Ref.~\cite{Levchenko08p256101}.}
\label{f.slab}
\end{figure}
The ferroelectric phase of lithium niobate (LiNbO$_3$) has a bulk
structure consisting of layers in a Li-O$_3$-Nb
pattern~\cite{Saito04p2057}. Here we show LiNbO$_3$ with its
thermodynamically preferred surface terminations, as predicted by
Levchenko and Rappe (Fig.\ S1). Li atoms are added to the
positive ($c^+$) surface, and both Li and O atoms are added to the
negative surface ($c^-$) of LiNbO$_3$ for charge passivation.
Passivation atoms are denoted Li$'$ and O$'$. These surface
terminations make the overall stoichiometry of five trilayer slab
used in this study, from $c^+$ (left) to $c^-$ (right),
Li$'$-(Li-O$_3$-Nb)$_5$-Li$'$O$'$.
\section{Supplementary DFT methods}
In all density functional theory (DFT) calculations, we used
3~$\times$~3~$\times$~1 $k$-point grids~\cite{Monkhorst76p5188},
plane wave energy cutoffs of 50~Ry~\cite{Ihm79p4409}, and force
tolerances of 0.01~eV/\AA. When necessary, we included spin as a
degree of freedom in our calculations. We checked the convergence
of our monomer and dimer adsorption energies calculated in
$\sqrt{3}\times\sqrt{3}$ supercells on five layer thick slabs by
using either $\sqrt{7}\times\sqrt{7}$ supercells or seven trilayer
thick slabs. The resulting energies agreed within 0.1~eV/Pd.
\section{Adsorption Energies}
\begin{table}[t]
\begin{tabular}{c|cc|ccc|c}
\hline
\hline
\# of Pd & \multicolumn{2}{c|} {positive ($c^+$)} & \multicolumn{3}{c|}{ negative ($c^-$)} & free Pd$_n$ \\
\cline{2-6}
atoms & $E^{a}_{ads}$& $E^{c}_{ads}$ & $E^{a}_{ads}$ & $E^{c}_{ads}$ & $E^{\rm Pd-X}_{\rm bond}$ & $E^{\rm Pd-Pd}_{\rm bond}$ \\
\hline
1 & 0.95 & -- & 2.02 & -- & 1.01 & -- \\
2 & 1.20 & 0.62 & 2.01 & 1.44 & 1.01 & 1.15 \\
3 & 1.56 & 0.44 & 2.27 & 1.15 & 1.13 & 1.12 \\
4 & 1.84 & 0.26 & 2.34 & 0.77 & 1.02 & 1.05 \\
5 & -- & -- & 2.35 & 0.65 & 0.98 & 0.94 \\
\hline
\hline
\end{tabular}
\caption{Per-atom adsorption energies of Pd on LiNbO$_3$ (LNO),
relative to infinitely separated Pd atoms ($E^{a}_{ads}$) and free
Pd clusters ($E^{c}_{ads}$). We define $E^{a}_{ads} = (E_{\rm
Pd-LNO} - E_{\rm LNO} -nE_{\rm Pd})/n$ and $E^{c}_{ads} = (E_{\rm
Pd-LNO} - E_{\rm LNO} - E_{{\rm Pd}_{n}})/n$, where $n$ is the
number of Pd atoms. On the $c^-$ surface, we report average Pd-X
bond energies, $E^{\rm Pd-X}_{\rm bond}$= $n$$E^{a}_{ads}$/$N_{\rm
bond}$, where $N_{\rm bond}$ is the number of Pd-X bonds. We also
report average Pd-Pd bond energies, $E^{\rm Pd-Pd}_{\rm bond}$, in
free Pd clusters for comparison. All energies are given in eV.}
\label{t.Ead}
\end{table}
In our analysis of the energetics of Pd adsorption on LiNbO$_3$,
we consider two quantities for per-atom adsorption energy,
$E^{a}_{ads}$ and $E^{c}_{ads}$, defined as the difference between
the energy of Pd-LNO and the sum of the energies of the LiNbO$_3$
slab and either infinitely separated ($E^{a}_{ads}$) or clustered
($E^{c}_{ads}$) Pd atoms divided by the number of Pd atoms
(Table~\ref{t.Ead}). We note that on the $c^+$ surface,
$E^{c}_{ads}$ becomes very small as the number of adsorbed Pd
atoms increases, because of a mismatch between ideal Pd-Pd bond
lengths and distances between Pd binding sites, and because most
of the cluster not in contact with LNO. In contrast, $E^{c}_{ads}$
remains relatively high even for pentamers on $c^-$, largely due
to Pd-O$'$ bonding. In the main text, we explained many of our
conclusions about energetics on the $c^-$ surface in terms of the
total number of Pd-Pd and Pd-O$'$ bonds (hereafter denoted Pd-X
bonds). This analysis was justified by the fact that average Pd-X
bond energies ($E^{\rm Pd-X}_{\rm bond}$), defined as
$nE^{a}_{ads}$ divided by the number of Pd-X bonds ($N_{\rm
bond}$), were similar for all cluster sizes studied and were also
similar to average bond energies in free Pd clusters ($E^{c}_{\rm
bond}$). We show explicit values for $E^{\rm Pd-Pd}_{\rm bond}$
and $E^{\rm Pd-X}_{\rm bond}$ in Table~\ref{t.Ead}.
Since Pd-O$'$ bonding makes the key contribution to the difference
in adsorption behavior between the $c^+$ and $c^-$ surfaces, we
wanted to be sure that DFT-GGA does not misestimate the
favorability of Pd-O interactions. To do this, we considered the
case of bulk PdO. We compared our DFT-GGA energy of formation of
PdO from individual Pd and O atoms, defined as $E^{\rm atom}_{\rm
O}+E^{\rm atom}_{\rm Pd}-E_{\rm PdO}$, to the sum of the
experimental values for PdO formation energy~\cite{Nell96p2487},
Pd bulk cohesive energy~\cite{Kittel96ISSP}, and half of the O$_2$
bonding energy~\cite{Furche01p195120}. Our DFT-calculated value
was only 2.1\% greater than the experimental value, suggesting
that DFT predicts Pd-O interactions well and that the strong Pd-O
interactions seen on the $c^-$ surface are valid. As a note, we
did not use the DFT PdO formation energy as a metric for
comparison, because DFT-GGA is known to predict the energy of
O$_2$ quite poorly~\cite{Furche01p195120}.
\section{Kinetic Monte Carlo simulation details}
\begin{figure}
\centering
\includegraphics[angle=0, width=0.35\textwidth]{figS2.pdf}
\caption{ An FCC lattice is not sufficient to allow all possible
agglomeration events to occur. Of the two agglomeration processes
pictured, one is allowed by the FCC lattice, while the other,
denoted by a red 'X', would also require HCP lattice points on the
second layer of the simulation lattice. We thus used an
A--(ABC)--(ABC)$\cdots$ lattice as described in the text. Open
circles denote Pd atoms and `X's denote points on the second layer
of the simulation lattice.}
\label{f.lattice}
\end{figure}
Our kinetic Monte Carlo (KMC) simulations were performed in
periodic 10~nm~$\times$~10~nm cells. We included 15 layers of Pd
binding sites above the surface, though no Pd atoms reached the
top layer of the cell in any of our simulations. All Pd binding
sites on the layer in contact with the LiNbO$_3$ surface were
mapped onto their corresponding points on the triangular oxygen
lattice. Beginning with the second layer above the surface, we
used three sublattices to allow all possible Pd cluster
configurations such that the lattice is
substrate--A--(ABC)--(ABC)$\cdots$. The use of this lattice,
which is much denser than the face-centered cubic (FCC) lattice of
bulk Pd, was necessary to describe some diffusion events that can
occur for small numbers of Pd atoms, but is forbidden by the FCC
lattice. For example, half of all possible agglomeration
processes are forbidden using an FCC lattice, because lattice
points are present that correspond to only half of the surface's
hollow sites (Fig.~\ref{f.lattice}). Additionally, hopping of Pd
atoms between FCC and hexagonal close-packed (HCP) hollow sites on
(111) cluster facets is forbidden by a pure FCC lattice. To
compensate for the fact that some points in the lattice used in
our simulation are closer together than a reasonable Pd-Pd bond
distance, we imposed two rules. First, the same site cannot be
occupied on adjacent layers of the lattice. Additionally, within
the same layer, adjacent sites on different sublattices cannot be
occupied.
We built our simulations primarily based on our DFT calculations
of the activation barriers of diffusion and agglomeration of up to
four atoms. On the $c^+$ surface, it was unnecessary to consider
diffusion of larger clusters, because our deposition time interval
was sufficiently long that we observed monomers to aggregate into
clusters before other monomers are deposited in the same vicinity.
Diffusion of large clusters on the $c^-$ surface could also be
neglected, because they require the breaking of two or more Pd-X
bonds and thus have activation barriers sufficient to prevent
their occurrence on the time scale that we considered. As
described below, in cases where we had not calculated activation
barriers directly, such as monoatomic motion on cluster facets, we
interpolated reasonable barriers.
We permitted diffusion of Pd atoms that were directly in contact
with the LiNbO$_3$ surface or on cluster facets. On the LiNbO$_3$
surface, we permitted the following diffusion processes: 1)
nearest-neighbor monomer hopping, 2) agglomeration/deagglomeration
(movement of Pd atom from layer 1 (2) onto layer 2 (1)), 3) dimer
walking and sliding on $c^+$, 4) trimer walking and sliding on
$c^+$, 5) trimer flipping on $c^-$, 6) tetramer rolling and 7)
dissociation (some of these processes are pictured in Fig.~2 in
the main text). On cluster facets, we permitted monomer hopping
only. Though concerted movements of multiple atoms can occur on
metal cluster facets, these generally affect cluster morphology,
but not cluster size (number of atoms present). Since our primary
interest is only in seeing differences in cluster size, this
simplification of Pd behavior on cluster facets is acceptable.
\begin{figure*}[!htb]
\centering
\includegraphics[angle=0, width=0.8\textwidth]{figS3.pdf}
\caption{Pd adsorption geometries predicted by kinetic Monte
Carlo simulations on the $c^-$ surface of LiNbO$_3$ immediately
after deposition ends (a-c) and 2-15 minutes later (d-e). These
simulations were conducted at different coverages (0.1~ML and
1.0~ML), different temperatures (300 and 400~K), and with the two
different agglomeration barriers predicted by DFT (0.59~eV and
0.74~eV). All simulations were conducted at a deposition rate of
0.025~ML/s. Times waited after the completion of deposition were:
(d) 15 minutes, (e) 10 minutes, (f) 2 minutes, and (g) 3 minutes.
All simulations are conducted at 300~K, except for (f), which is
at 400~K.}
\label{f.ann}
\end{figure*}
It is well known that the activation barriers for monomeric
diffusion on the (111) surfaces of FCC metals are quite low
relative to the bulk metal-metal bond energy. Our nudged elastic
band (NEB) calculations show that the activation barrier for
monoatomic hopping from from FCC to HCP hollow sites on (111)
surface of palladium is 0.11~eV, despite the fact that the Pd-Pd
bulk bond energy is nearly six times larger (0.59~eV). This is
explained by the fact that during monomer hopping between
nearest-neighbor sites, new bonds start to form before others are
completely broken. In our KMC simulations, we interpolated
reasonable activation barriers for diffusion events on Pd cluster
facets using insights from Trunshin and coworker's studies of
nearest-neighbor monoatomic hopping on copper surfaces and
steps~\cite{Trushin05p115401}. By compensating for the difference
in cohesive energy between bulk Pd and Cu, we estimated that the
activation barrier for nearest neighbor monomer hopping is 0.2~eV
per Pd-Pd bond broken during transition state formation. We also
showed that the qualitative results of some of our simulations
were insensitive to the use of a 0.3~eV barrier for this process.
In next nearest neighbor monomer hopping, which is analogous to
the agglomeration processes pictured in Fig.~2 but generalized to
movements between any two levels in the simulation lattice
(starting with level 2), Pd-Pd bonds are broken completely during
transition state formation. Thus, we estimated the activation
barrier for this process based on the cost of Pd-Pd bond breaking.
The cost of breaking a bond in a Pd$_{20}$ cluster cut from bulk
Pd is 0.52~eV. Similarly the cost of dimer dissociation is
0.55~eV. Thus, for next-nearest neighbor hopping on Pd cluster
facets, we used a barrier of 0.5~eV per Pd-Pd bond broken.
\section{Aggregation and agglomeration on $c^-$ after deposition}
The relatively large diffusion barriers on the c$^-$ surface
($\approx$0.8~eV) were not sufficiently high to prevent
agglomeration completely, especially when our simulations were
extended until slightly after deposition was complete. We have
conducted KMC simulations during deposition and after annealing
for 2-15~minutes for both low (0.1~ML) and high (1.0~ML) coverages
and at different temperatures (300 or 400~K). We also conducted
our simulations with agglomeration barriers (0.59 and 0.74~eV) for
the two different possible cluster agglomeration processes
predicted by DFT on the $c^-$ surface (See Fig. 2(g,h) and Table I
in the main text). These data are reported in Fig.~\ref{f.ann}.
At low coverage (0.1~ML), Pd atoms are highly dispersed at the
end of deposition (Fig.~\ref{f.ann}(a)). However, within 15
minutes, all but two Pd clusters are immobile, suggesting that on
a slightly longer time scale, all Pd atoms will join immobile
clusters (Fig.~\ref{f.ann}(d)). However, these clusters are quite
small compared to the single cluster that forms on the $c^+$
surface under the same conditions. Two factors contribute to the
formation of smaller clusters on the $c^-$ surface than on $c^+$:
lower maximum mobile cluster size, and the higher number density
of clusters on the surface during deposition. On the $c^+$
surface, the mismatch between the distances between monomeric Pd
binding sites and Pd-Pd equilibrium bond distances makes the
activation barrier for motion of relatively large clusters rather
low. For example, our predicted barrier for tetramer rolling was
similar to that for monomer hopping, largely because this process
requires movement of only one Pd atom out of a monomeric potential
well. Other work suggests that this condition may be met by Pd
clusters as large as 13 atoms~\cite{Fan03p117}, and thus, that
clusters of this size may be mobile on the $c^+$ surface. In
contrast, the diffusion of clusters larger than a tetramer on the
$c^-$ surface requires breakage of at least two Pd-X bonds, making
these diffusion events extremely unlikely on the time scale of
deposition. A second factor contributing to the formation of
large clusters on the $c^+$ surface is the fact that, because
barriers for diffusion are so low, many diffusion events can occur
during the average waiting time between addition of new atoms to
the simulation cell on this surface. This means that only a few
atoms at a time will be separate from the large immobile cluster,
greatly reducing the probability that a second immobile cluster
will form. The opposite is true on the $c^-$ surface, where
larger diffusion barriers mean that on average multiple Pd atoms
are deposited inside the simulation cell between diffusion events
of a single atom or cluster. As a result, many individual
clusters are present in any given area of the simulation cell,
promoting the formation of many small immobile clusters.
At high coverage (1.0~ML), Pd atoms continue to rearrange after
deposition is complete until all atoms make at least three Pd-X
bonds (Fig.~\ref{f.ann}(b,e)). Once this occurs, agglomeration
can only occur on an extremely long time scale, because it
requires breakage of multiple Pd-X bonds. Thus, the Pd coverage
area of these clusters will remain relatively constant. This was
true even when our surface was annealed at a higher temperature
(400~K, Fig.~\ref{f.ann}(f)), consistent with the finding of Zhao
and colleagues that the Pd geometry on the $c^-$ surface is stable
up to 425~K~\cite{Zhao09p1337}.
Even when we used the lower of our two agglomeration barriers
predicted from DFT on the $c^-$ surface (Fig. 2(g) in the main
text), multiple distinct clusters formed at 1 ML coverage
(Fig.~\ref{f.ann}(c,g)). At low coverage we observed results
similar to those for the higher agglomeration barrier
(Fig.~\ref{f.ann}(d)), but with taller individual clusters.
Overall, our simulations with this lower agglomeration barrier on
the $c^-$ surface predicted Pd coverage areas that were
substantially higher than those on the $c^+$ surface, but lower
than the simulations on the $c^-$ surface with the higher
agglomeration barrier.
In sum, our DFT and KMC calculations predict that though Pd forms
slightly three dimensional clusters on the $c^-$ surface of
LiNbO$_3$, its overall adsorption pattern on this surface is much
more dispersed than that on the $c^+$ surface.
\bibliography {skim_PdLNO}
\end{document}
|
\section{Introduction}
The large-scale distribution of the arrival directions of Ultra-High Energy Cosmic Rays
(UHECRs) is, together with the spectrum and the mass composition, an important observable in
attempts to understand their nature and origin. The \textit{ankle}, a hardening of the
energy spectrum of UHECRs located at $E\simeq4\,$EeV~\cite{linsley-ankle,lawrence,nagano,
bird,auger-spectrum}, where 1~${\rm EeV}\equiv 10^{18}$~eV, is presumed to be either
the signature of the transition from galactic to extragalactic UHECRs~\cite{linsley-ankle},
or the distortion of a proton-dominated extragalactic spectrum due to $e^{\pm}$ pair
production of protons with the photons of the Cosmic Microwave Background (CMB)~\cite{hillas,berezinsky}.
If cosmic rays with energies below the ankle have a galactic origin, their escape from
the Galaxy might generate a dipolar large-scale pattern as seen from the Earth. The
amplitude of such a pattern is difficult to predict, as it depends on the assumed
galactic magnetic field and the charges of the particles as well as the distribution
of sources. Some estimates, in which the galactic cosmic rays are mostly heavy, show that
anisotropies at the level of a few percent are nevertheless expected in the EeV
range~\cite{ptuskin,roulet1}. Even for isotropic extragalactic cosmic rays, a dipole
anisotropy may exist due to our motion with respect to the frame of extragalactic
isotropy. This \emph{Compton-Getting effect}~\cite{compton} has been measured with
cosmic rays of much lower energy at the solar frequency~\cite{eastop, groom} as
a result of our motion relative to the frame in which they have no bulk motion.
Since January 2004, the surface detector (SD) array of the Pierre Auger Observatory has
collected a large amount of data. The statistics accumulated in the $1\,$EeV energy range
allows one to be sensitive to intrinsic anisotropies with amplitudes down to the 1\% level.
This requires determination of the exposure of the sky at a corresponding accuracy (see
Section~\ref{sec:exposure}) as well as control of the systematic uncertainty of the
variations in the counting rate of events induced by the changes of the atmospheric conditions
(see Section~\ref{sec:weather}). After carefully correcting these experimental effects, we
present in Section~\ref{sec:results} searches for first harmonic modulations in right-ascension
based on the classical Rayleigh analysis~\cite{linsley-fh} slightly modified to account for
the small variations of the exposure with right ascension.
Below $E\simeq1\,$EeV, the detection efficiency of the array depends on zenith angle and
composition, which amplifies detector-dependent variations in the counting rate.
Consequently, our results below 1~EeV are derived using simple event counting rate
differences between Eastward and Westward directions~\cite{ew}. That technique using
relative rates allows a search for anisotropy in right ascension without requiring any
evaluation of the detection efficiency.
From the results presented in this work, we derive in Section~\ref{sec:discussion}
upper limits on modulations in right-ascension of UHECRs and discuss some of their
implications.
\section{The Pierre Auger Observatory and the data set}
\label{sec:auger}
The southern site of the Pierre Auger Observatory~\cite{auger-nim} is located in Malarg\"{u}e,
Argentina, at latitude 35.2$^\circ\,$S, longitude 69.5$^\circ\,$W and mean altitude 1400
meters above sea level. Two complementary techniques are used to detect extensive air
showers initiated by UHECRs~: a \textit{surface detector array} and a \textit{fluorescence
detector}. The SD array consists of 1660 water-Cherenkov detectors covering an area of
about 3000~km$^2$ on a triangular grid with 1.5~km spacing, allowing electrons, photons
and muons in air showers to be sampled at ground level with a duty cycle of almost 100\%.
In addition, the atmosphere above the SD array is observed during clear, dark nights by
24 optical telescopes grouped in 4 buildings. These detectors observe the longitudinal
profile of air showers by detecting the fluorescence light emitted by nitrogen molecules
excited by the cascade.
The data set analysed here consists of events recorded by the surface detector from 1 January
2004 to 31 December 2009. During this time, the size of the Observatory increased from 154 to
1660 surface detector stations. We consider in the present analysis events\footnote{A
comprehensive description of the identification of shower candidates detected at the SD array
of the Pierre Auger Observatory is given in reference~\cite{auger-acc}.} with reconstructed
zenith angles smaller than 60$^\circ$ and satisfying a fiducial cut requiring that the six
neighbouring detectors in the hexagon surrounding the detector with the highest
signal were active when the event was recorded. Throughout this article, based on this fiducial
cut, any active detector with six active neighbours will be defined as
an \emph{unitary cell}~\cite{auger-acc}. It ensures both a good quality
of event reconstruction and a robust estimation of the exposure of the SD array, which is then
obtained in a purely geometrical way. The analysis reported here is restricted
to selected periods to eliminate unavoidable problems associated to the construction
phase, typically in the data acquisition and the communication system or due to hardware
instabilities~\cite{auger-acc}. These cuts restrict the duty cycle to $\simeq 85$\%. Above the
energy at which the detection efficiency saturates, 3~EeV~\cite{auger-acc}, the exposure
of the SD array is 16,323~km$^2$~sr~yr for six years used in this analysis.
The event direction is determined from a fit to the arrival times of the shower front at the
SD. The precision achieved in this reconstruction depends upon the accuracy on the GPS
clock resolution and on the fluctuations in the time of arrival of the first particle~\cite{ang-res}.
The angular resolution is defined as the angular aperture around the arrival directions of
cosmic rays within which 68\% of the showers are reconstructed. At the lowest observed energies,
events trigger as few as three surface detectors. The angular resolution of events having such a
low multiplicity is contained within $2.2^\circ$, which is quite sufficient to perform searches
for large-scale patterns in arrival directions, and reaches $\sim 1^\circ$ for events with
multiplicities larger than five~\cite{ang-res2}.
The energy of each event is determined in a two-step procedure. First, using the constant
intensity cut method, the shower size at a reference distance of $1000\,$m, $S(1000)$, is
converted to the value $S_{38^\circ}$ that would have been expected had the shower arrived
at a zenith angle 38$^\circ$. Then, $S_{38^\circ}$ is converted to energy
using a calibration curve based on the fluorescence telescope measurements~\cite{auger-prl}.
The uncertainty in $S_{38^\circ}$ resulting from the adjustment of the shower size, the
conversion to a reference angle, the fluctuations from shower-to-shower and the calibration
curve amounts to about 15\%. The absolute energy scale is given by the fluorescence
measurements and has a systematic uncertainty of 22\%~\cite{auger-prl}.
\section{The exposure of the surface detector}
\label{sec:exposure}
\begin{figure}[!t]
\centering
\includegraphics[width=7.5cm]{RelDetSize-solar.eps}
\includegraphics[width=7.5cm]{RelDetSize-sidereal.eps}
\caption{\small{Relative variation of the integrated number of unitary cells as a
function of the solar hour of the day in UTC (left panel), and as a function of the local sidereal
time (right panel).}}
\label{N_h}
\end{figure}
The instantaneous exposure $\omega(t,\theta,\phi,S_{38^\circ})$ of the SD array at the
time $t$ as a function of the incident zenith and azimuth\footnote{The angle $\phi$ is
the azimuth relative to the East direction, measured counterclockwise.} angles ($\theta,\phi$)
and shower size $S_{38^\circ}$ is given by~:
\begin{equation}
\label{eqn:omega}
\omega(t,\theta,\phi,S_{38^\circ})=n_{\mathrm{cell}}(t) \times a_{\mathrm{cell}} \cos{\theta}\times\epsilon(S_{38^\circ},\theta,\phi),
\end{equation}
where $a_{\mathrm{cell}} \cos{\theta}$ is the projected surface of a unitary cell under
the incidence zenith angle $\theta$, $n_{\mathrm{cell}}(t)$ is the number of unitary cells
at time $t$, and $\epsilon(S_{38^\circ},\theta,\phi)$ is the directional detection
efficiency at size parameter $S_{38^\circ}$ under incidence angles ($\theta,\phi$).
The conversion from $S_{38^\circ}$ to the energy $E$, which accounts for the changes of atmospheric
conditions, will be presented in the next section.
The number of unitary cells $n_{\mathrm{cell}}(t)$ is recorded every second using the trigger
system of the Observatory and reflects the array growth as well as the dead periods of each
detector. It ranges from $\simeq60$ (at the begining of the data taking in 2004) to
$\simeq1200$ (from the middle of 2008). From Eqn.~\ref{eqn:omega}, it is apparent that
$n_{\mathrm{cell}}(t)$ is the only time-dependent quantity entering in the definition of the
instantaneous exposure, modulating within any integrated solid angle the expected number of
events as a function of time. For any periodicity $T$, the total number of unitary cells
$N_{\mathrm{cell}}(t)$ as a function of time $t$ within a period and summed over all periods,
and its associated relative variations $\Delta N_{\mathrm{cell}}(t)$ are obtained from~:
\begin{equation}
N_{\mathrm{cell}}(t)=\sum_j n_{\mathrm{cell}}(t+jT), \hspace{1cm}
\Delta N_{\mathrm{cell}}(t)=\frac{N_{\mathrm{cell}}(t)}{\left<N_{\mathrm{cell}}(t)\right>},
\end{equation}
with $\left<N_{\mathrm{cell}}(t)\right>=1/T\,\int_0^T dt\,N_{\mathrm{cell}}(t)$.
A genuine dipolar anisotropy in the right ascension distribution of the events induces
a modulation in the distribution of the time of arrival of events with a period equal to
one sidereal day. A sidereal day indeed corresponds to the time it takes for the Earth to
complete one rotation relative to the vernal equinox. It is approximately $T_{sid} =$ 23 hours,
56 minutes, 4.091 seconds. Throughout this article, we denote by $\alpha^0$ the local sidereal
time and express it in hours or in radians, as appropriate. For practical reasons, $\alpha^0$
is chosen so that it is always equal to the right ascension of the zenith at the center of the
array.
On the other hand, a dipolar modulation \textit{of experimental origin} in the distribution of the
time of arrival of events with a period equal to one solar day may induce a spurious dipolar
anisotropy in the right ascension distribution of the events. Hence, it is essential to
control $\Delta N_{\mathrm{cell}}(t)$ to account for the variation of the
exposure in different directions. We show $\Delta N_{\mathrm{cell}}(t)$ in Fig.~\ref{N_h}
in 360 bins of 4 minutes at these two time scales of particular interest~:
the solar one $T=T_{sol}$ = 24 hs (left panel), and the sidereal one $T=T_{sid}$ (right
panel). A clear diurnal variation is apparent
on the solar time scale showing an almost dipolar modulation with an amplitude
of $\simeq 2.5$\%. This is due to both the working times of the construction phase of the
detector and to the outage of some batteries of the surface detector stations during nights.
When averaged over 6 full years, this modulation is almost totally smoothed out on the sidereal
time scale as seen in the right panel of Fig.~\ref{N_h}. This distribution will be used in
section \ref{sec:RayAna} to weight the events when estimating the Rayleigh parameters.
From the instantaneous exposure, it is straightforward to compute the integrated exposure
either in local coordinates $\omega(\theta,\phi,\alpha^0)$ by replacing
$n_{\mathrm{cell}}(t)$ by $\Delta N_{\mathrm{cell}}(\alpha^0)$ in Eqn.~\ref{eqn:omega},
or in celestial coordinates $\omega(\alpha,\delta)$ by expressing the zenith angle $\theta$
in terms of the equatorial right ascension $\alpha$ and declination $\delta$ through~:
\begin{equation}
\label{eqn:theta}
\cos{\theta}=\sin{\ell_{\mathrm{site}}}\sin{\delta}+\cos{\ell_{\mathrm{site}}}\cos{\delta}\cos{(\alpha-\alpha^0)},
\end{equation}
(where $\ell_{\mathrm{site}}$ is the Earth latitude of the site) and then by
integrating Eqn.~\ref{eqn:omega} over time. Besides, let us also mention
that to account for the spatial extension of the surface detector array making the latitude
of the site $\ell_{\mathrm{site}}$ varying by $\simeq 0.4^\circ$, the celestial
coordinates $(\alpha,\delta)$ of the events are calculated by transporting the showers
to the "center" of the Observatory site.
\section{Influence of the weather effects}
\label{sec:weather}
Changes in the atmospheric pressure $P$ and air density $\rho$ have been shown to affect the
development of extensive air showers detected by the surface detector array and these changes
are reflected in the temporal variations of shower size at a fixed energy~\cite{auger-weather}.
To eliminate these variations, the procedure used to convert the observed signal into energy
needs to account for these atmospheric effects. This is performed by relating the signal at 1~km
from the core, $S(1000)$, measured at the actual density $\rho$ and pressure $P$, to the one
$\tilde{S}(1000)$ that would have been measured at reference values $\rho_0$ and $P_0$,
chosen as the average values at Malarg\"ue, \textit{i.e.}
$\rho_0=1.06$ kg~m$^{-3}$ and $P_0=862$~hPa~\cite{auger-weather}~:
\begin{equation}
\tilde{S}(1000)=\left[1-\alpha_P(\theta)(P-P_0)-\alpha_\rho(\theta)(\rho_d-\rho_0)-\beta_\rho(\theta)(\rho-\rho_d)\right]S(1000),
\label{sweather}
\end{equation}
where $\rho_d$ is the average daily density at the time the event was recorded. The measured
coefficients $\alpha_\rho=(-0.80\pm0.02)\,$kg$^{-1}$m$^3$, $\beta_\rho=(-0.21\pm0.02)\,$kg$^{-1}$m$^3$
and $\alpha_P=(-1.1\pm0.1)\,10^{-3}\,$hPa$^{-1}$ reflect respectively the impact of the variation
of air density (and thus temperature) at long and short time scales, and of the variation of
pressure on the shower sizes~\cite{auger-weather}. It is worth pointing out that air density
coefficients are here predominant relative to the pressure one. The zenithal dependences of these
parameters, that we use in the following, were also studied in reference~\cite{auger-weather}.
It is this reference signal $\tilde{S}(1000)$ which has to be
converted, using the constant intensity cut method, to the signal size $\tilde{S}_{38^\circ}$
and finally to energy. For convenience, we denote hereafter the uncorrected (corrected)
shower size $S_{38^\circ}$ ($\tilde{S}_{38^\circ}$) simply by $S$ ($\tilde{S}$).
Carrying out such energy corrections is important for the study of large scale anisotro-
pies.
Above 3~EeV, the rate of events $R$ per unit time above a given \emph{uncorrected} size threshold
$S_{th}$, and in a given zenith angle bin, is modulated by changes of atmospheric conditions~:
\begin{equation}
R(>S_{th})\propto\int_{S_{th}}^{\infty}dS\,\frac{dJ}{d\tilde{S}}\,\frac{d\tilde{S}}{dS}\propto\bigg[1+(\gamma_S-1)\,\alpha_\xi\,\Delta\xi\bigg]\,\int_{S_{th}}^{\infty}dS\,S^{-\gamma_S},
\label{def_rate_sat}
\end{equation}
where hereafter $\xi$ generically denotes $P$, $\rho$ or $\rho_d$, and where we have
adopted for the differential flux $dJ/dS$ a power law with spectral index $\gamma_S$.
Hence, under changes of atmospheric parameters $\Delta\xi$, the following relative
change in the rate of events is expected~:
\begin{eqnarray}
\frac{1}{R}\frac{dR(>S_{th})}{d\xi}&\simeq&(\gamma_S-1)\,\alpha_\xi.
\label{rel_rate_sat}
\end{eqnarray}
Over a whole year, this spurious modulation is partially compensated in sidereal time,
though not in solar time. In addition, a seasonal variation of the modulation of the daily
counting rate induces sidebands at both the sidereal and anti-sidereal\footnote{The anti-sidereal
time is a fictitious time scale symmetrical to the sidereal one with respect to the solar one
and that reflects seasonal influences~\cite{far-sto}. It corresponds to a fictitious year of
$\simeq 364$ days.} frequencies. This may lead to
misleading measures of anisotropy if the amplitude of the sidebands significantly stands
out above the background noise~\cite{far-sto}. Correcting energies for weather effects, the
net correction in the first harmonic amplitude in \emph{sidereal time} turns out to be only
of $\simeq 0.2\%$ for energy thresholds greater than 3~EeV, thanks to large cancellations
taking place when considering the large time period used in this study.
\begin{figure}[!t]
\centering
\includegraphics[width=10cm]{RelMod.eps}
\caption{\small{Relative modulation of the rates above a given corrected signal size due
to the variations of the detection efficiency under changes of atmospheric conditions,
relatively to the factor modulating the rate of events above the corresponding uncorrected
signal size in units of $S_{1/2}$.}}
\label{rel_mod}
\end{figure}
In addition to the energy determination, weather effects can also affect the detection
efficiencies for showers with energies below 3~EeV, for the detection of which the surface
array is not fully efficient. Changes of the shower signal size due to changes of weather conditions
$\Delta\xi$ imply that showers are detected with the efficiency associated to the
\emph{observed} signal size $S$, which is related at first order to the one
associated to the \emph{corrected} signal size through~:
\begin{equation}
\epsilon(S)\simeq\epsilon(\tilde{S})+(S-\tilde{S})\frac{d\epsilon(S)}{dS}\bigg|_{S=\tilde{S}}\simeq\epsilon(\tilde{S})+\alpha_\xi\Delta\xi\,\tilde{S}\,\frac{d\epsilon(S)}{dS}\bigg|_{S=\tilde{S}},
\label{epsilon_variation}
\end{equation}
where we have made use of Eqn.~\ref{sweather}.
The second term modulates the observed rate of events, even after the correction of the
signal sizes. Indeed, the rate of events $R$ above a given \emph{corrected} signal size
threshold $\tilde{S}_{th}$ is now the integration of the cosmic ray spectrum weighted
by the corresponding detection efficiency expressed in terms of the \emph{observed} signal
size $S$~:
\begin{equation}
R(>\tilde{S}_{th})\propto\int_{\tilde{S}_{th}}^{\infty}d\tilde{S}\,\bigg[\epsilon(\tilde{S})+\alpha_\xi\Delta\xi\,\tilde{S}\,\frac{d\epsilon(S)}{dS}\bigg|_{S=\tilde{S}}\bigg]\,\frac{dJ}{d\tilde{S}}.
\label{def_rate}
\end{equation}
Hence, the relative change in the rate of events under changes in the atmosphere becomes~:
\begin{eqnarray}
\frac{1}{R}\frac{dR(>\tilde{S}_{th})}{d\xi}\simeq\frac{\alpha_\xi}{R}\int_{\tilde{S}_{th}}^{\infty}d\tilde{S}\,\tilde{S}\,\frac{dJ}{d\tilde{S}}\,\frac{d\epsilon(S)}{dS}\bigg|_{S=\tilde{S}},
\label{def_rel_rate}
\end{eqnarray}
which, after an integration by parts and at first order in $\alpha_\xi$, leads to~:
\begin{eqnarray}
\frac{1}{R}\frac{dR(>\tilde{S}_{th})}{d\xi}&\simeq&(\gamma_S-1)\,\alpha_\xi\,\bigg[1-\frac{\epsilon(\tilde{S}_{th})\int_{\tilde{S}_{th}}^{\infty}d\tilde{S}\,\tilde{S}^{-\gamma_S}}{\int_{\tilde{S}_{th}}^{\infty}d\tilde{S}\,\epsilon(\tilde{S})\,\tilde{S}^{-\gamma_S}}\bigg].
\label{rel_rate}
\end{eqnarray}
The expression in brackets gives the additional modulations (in units of the weather
effect modulation $(\gamma_s-1)\alpha_\xi$ when the detection efficiency is saturated)
due to the variation of the detection efficiency. Note that this expression is less than 1 for any
rising function $\epsilon$ satisfying $0\leq\epsilon(S)\leq 1$, and reduces to 0, as expected,
when $\epsilon(\tilde{S}_{th})=1$. As a typical example, we show in
Fig.~\ref{rel_mod} the expected modulation amplitude as a function of $S$ by adopting a
reasonable detection efficiency function of the form $\epsilon(S)=S^3/\big[S^3+S_{1/2}^3\big]$
where the value of $S_{1/2}$ is such that $\epsilon(S_{1/2})=0.5$. This relative amplitude is
about 0.3 for $S=S_{1/2}$, showing that for this signal size threshold the remaining modulation
of the rate of events after the signal size corrections is about $0.3\times(\gamma_s-1)\alpha_\xi$.
The value of $S_{1/2}$ being such that it corresponds to $\simeq 0.7$~EeV in terms of energy,
it turns out that within the current statistics the Rayleigh analysis of arrival directions can
be performed down to threshold energies of 1$\,$EeV by \emph{only} correcting the energy
assignments.
\section{Analysis methods and results}
\label{sec:results}
\subsection{Overview of the analyses}
The distribution in right ascension of the flux of CRs arriving at a detector can be
characterised by the amplitudes and phases of its Fourier expansion,
$I(\alpha) = I_0 (1 + r \cos(\alpha -\varphi) + r' \cos(2(\alpha - \varphi')) + \dots)$.
Our aim is to determine the first harmonic amplitude $r$ and its phase $\varphi$.
To account for the non-uniform exposure of the SD array, we perform two different
analyses.
\subsubsection{Rayleigh analysis weighted by exposure}
\label{sec:RayAna}
Above 1$\,$EeV, we search for the first harmonic modulation in right ascension
by applying the classical Rayleigh formalism~\cite{linsley-fh}
slightly modified to account for the non-uniform exposure to different parts of the sky.
This is achieved by weighting each event with a factor inversely proportional to the
integrated number of unitary cells at the local sidereal time of the event (given by
the right panel histogram of Fig.~\ref{N_h})~\cite{sommers,roulet2}:
\begin{equation}
\label{eqn:fh}
a=\frac{2}{\mathcal{N}}\sum_{i=1}^N w_i\cos{\alpha_i},
\hspace{1cm}b=\frac{2}{\mathcal{N}}\sum_{i=1}^N w_i\sin{\alpha_i},
\end{equation}
where the sum runs over the number of events $N$ in the considered energy range, the weights
are given by $w_i\equiv [\Delta N_{\mathrm{cell}}(\alpha^0_i)]^{-1}$ and the normalisation
factor is $\mathcal{N}=\sum_{i=1}^Nw_i$. The estimated amplitude $r$ and phase $\varphi$ are
then given by~:
\begin{equation}
\label{eqn:fh2}
r=\sqrt{a^2+b^2}, \hspace{1cm}\varphi=\arctan\frac{b}{a}.
\end{equation}
As the deviations from an uniform right ascension exposure are small,
the probability $P(>r)$ that an amplitude equal or larger than $r$ arises from an isotropic
distribution can be approximated by the cumulative distribution function of the Rayleigh
distribution $P(>r)=\exp{(-k_0)}$, where $k_0=\mathcal{N}r^2/4$.
\subsubsection{East-West method}
Below 1$\,$EeV, due to the variations of the event counting rate arising from Eqn.~\ref{rel_rate},
we adopt the differential \textit{East-West method}~\cite{ew}. Since the instantaneous exposure
of the detector for Eastward and Westward events is the same\footnote{The global tilt of the
array of $\simeq 0.2^\circ$, that makes it slightly asymmetric, is here negligible - see
sub-section~\ref{addcrosschecks}.}, with both sectors being equally affected by the instabilities
of the detector and the weather conditions, the difference between the event counting rate
measured from the East sector, $I_E(\alpha^0)$, and the West sector, $I_W(\alpha^0)$,
allows us to remove at first order the direction independent effects of experimental
origin without applying any correction, though at the cost of a reduced sensitivity.
Meanwhile, this counting difference is directly related to the right ascension modulation
$r$ by (see Appendix)~:
\begin{equation}
I_E(\alpha^0)-I_W(\alpha^0) = - \frac{N}{2\pi}\frac{2\left<\sin{\theta}\right>}
{\pi \langle \cos \delta \rangle} r \sin (\alpha^0 -\varphi).
\end{equation}
The amplitude $r$ and phase $\varphi$ can thus be calculated from the arrival times of each set
of $N$ events using the standard first harmonic analysis~\cite{linsley-fh} slightly modified to
account for the subtraction of the Western sector to the Eastern one. The Fourier
coefficients $a_{EW}$ and $b_{EW}$ are thus defined by~:
\begin{eqnarray}
\label{eqn:fhew}
a_{EW} = \frac{2}{N}\sum_{i=1}^{N} \cos{(\alpha^0_i+\zeta_i)},\hspace{1cm}
b_{EW} = \frac{2}{N}\sum_{i=1}^{N} \sin{(\alpha^0_i+\zeta_i)},
\end{eqnarray}
where $\zeta_i$ equals 0 if the event is coming from the East or $\pi$ if coming from the West
(so as to effectively subtract the events from the West direction).
This allows us to recover the amplitude $r$ and the phase $\varphi_{EW}$ from
\begin{equation}
\label{eqn:fh2ew}
r=\sqrt{a_{EW}^2+b_{EW}^2}, \hspace{1cm}\varphi_{EW}=
\arctan{\left(\frac{b_{EW}}{a_{EW}}\right)}.
\end{equation}
Note however that $\varphi_{EW}$, being the phase corresponding to the maximum
in the differential of the East and West fluxes, is related to $\varphi$ in
Eqn.~\ref{eqn:fh2} through $\varphi = \varphi_{EW}+\pi/2$. As in the previous analysis,
the probability $P(>r)$ that an amplitude equal or larger than $r$ arises from an isotropic
distribution is obtained by the cumulative distribution function of the Rayleigh distribution
$P(>r)=\exp{(-k_0^{EW})}$, where
$k_0^{EW}=(2\left<\sin{\theta}\right>/\pi\langle\cos\delta\rangle)^2\times Nr^2/4$. For the
values of $\left<\sin{\theta}\right>$ and $\langle\cos\delta\rangle$ of the events used in
this analysis, the first factor in the expression for $k_0^{EW}$ is 0.22. Then, comparing
it with the expression for $k_0$ in the standard Rayleigh analysis, it is seen that
approximately four times more events are needed in the East-West method to attain the same
sensitivity to a given amplitude $r$.
\subsection{Analysis of solar, anti-sidereal, and random frequencies}
\label{sec:solar-antisid}
\begin{figure}[!t]
\centering
\includegraphics[width=10cm]{PowerSpectrum.1eev.eps}
\caption{\small{Amplitude of the Fourier modes as a function of the frequency above 1$\,$EeV.
Thin dotted curve~: before correction of energies and exposure. Dashed curve~: after correction
of energies but before correction of exposure. Thick curve~: After correction of energies and
exposure. Dashed vertical lines from left to right~: anti-sidereal, solar and sidereal frequencies.}}
\label{spectralanalysis}
\end{figure}
The amplitude $r$ corresponds to the value of the Fourier transform of the arrival time distribution
of the events at the sidereal frequency. This can be generalised to other frequencies by performing
the Fourier transform of the modified time distribution~\cite{bil-let}~:
\begin{equation}
\label{eqn:modtimes}
\tilde{\alpha}^0_i=\frac{2\pi}{T_{sid}}t_i+\alpha_i-\alpha^0_i.
\end{equation}
Such a generalisation is helpful for examining an eventual residual spurious modulation
after applying the Rayleigh analysis after the corrections discussed in Sections
\ref{sec:exposure} and \ref{sec:weather}.
The amplitude of the Fourier modes when considering all events above 1$\,$EeV are shown
in Fig.~\ref{spectralanalysis} as a function of the frequency in a window centered on the
solar one (indicated by the dashed line at 365.25~cycles/year). The thin dotted curve is obtained
without accounting for the variations of the exposure and without accounting for the weather
effects. The large period of time analysed here, over 6 years, allows us to resolve the
frequencies at the level of $\simeq 1/6$~cycles/year. This induces a large decoupling of the
frequencies separated by more than this resolution~\cite{bil-let}. In particular, as the
resolution is less than the difference between the solar and the (anti-)sidereal frequencies
(which is of 1~cycle/year), this explains why the large spurious modulations standing out from
the background noise around the solar frequency are largely averaged out at both the sidereal
and anti-sidereal frequencies even without applying any correction.
The impact of the correction of the energies discussed in Section~\ref{sec:weather} is evidenced
by the dashed curve, which shows a reduction of $\simeq30$\% of the spurious modulations
within the resolved solar peak. In addition, when accounting also for the exposure variation at
each frequency, the solar peak is reduced at a level close to the statistical noise, as evidenced
by the thick curve. Results at the solar and the anti-sidereal frequencies are collected
in Tab.~\ref{tab}.
\begin{table*}[h]
\begin{center}
\begin{tabular}{c|c|c|c|c}
& $r_{\mathrm{solar}}$[\%] & $P(>r_{\mathrm{solar}})$[\%] & $r_{\mathrm{anti-sid}}$[\%] & $P(>r_{\mathrm{anti-sid}})$[\%] \\
\hline
\hline
no correction & 3.7 & $\simeq 2\,10^{-37}$ & 0.36 & 43 \\
energy corrections & 2.9 & $\simeq 4\,10^{-23}$ & 0.15 & 85 \\
+exposure correction & 0.96 & 0.2 & 0.49 & 19
\end{tabular}
\caption{\small{Amplitude and corresponding probability to get a larger amplitude from an
isotropic distribution at both the solar and the anti-sidereal frequencies for events with
energies $>1\,$EeV.}}
\label{tab}
\end{center}
\end{table*}%
\begin{figure}[!t]
\centering
\includegraphics[width=7.5cm]{RelDetSize-100.eps}
\includegraphics[width=7.5cm]{HistAmpRndmFreq.eps}
\caption{\small{Left~: Relative variation of the integrated number of unitary cells as a
function of the time $t_{100}$, where the time scale is such that corresponding frequency is
100~cycles/year. Right~: Rayleigh analysis above 1~EeV for 1600 random frequencies ranging
from 100 to 500~cycles/year. Thin histogram~: analysis without accounting for the exposure
variations. Thick histogram~: analysis accounting for the exposure variations. Smooth curve~:
expected Rayleigh distribution.}}
\label{fourierspectrum}
\end{figure}
To provide further evidence of the relevance of the corrections introduced to account
for the non-uniform exposure, it is worth analysing on a statistical basis the behaviour
of the reconstructed amplitudes at different frequencies (besides the anti-sidereal/solar/sidereal
ones). In particular, as the number of unitary cells $n_{\mathrm{cell}}$ has increased
from $\simeq$ 60 to $\simeq$ 1200 over the 6 years of data taking, an automatic
increase of the variations of $\Delta N_{\mathrm{cell}}(t)$ is expected at large time
periods. This expectation is illustrated in the left panel of Fig.~\ref{fourierspectrum},
which is similar to Fig.~\ref{N_h} but at a time periodicity $T\simeq87.5\,$h, corresponding
to a low frequency of 100~cycles/year. The size of the modulation is of the order
of the one observed in Fig.~\ref{N_h} at the solar frequency. In the right panel of
Fig.~\ref{fourierspectrum}, the results of the Rayleigh analysis applied above 1~EeV
to 1600 random frequencies ranging from 100 to 500~cycles/year are shown by histograming
the reconstructed amplitudes. The thin one is obtained \emph{without} accounting for the
variations of the exposure~: it clearly deviates from the expected Rayleigh distribution
displayed in the same graph. Once the exposure variations are accounted for through the
weighting procedure, the thick histogram is obtained, now in agreement with the expected
distribution. Note that in both cases the energies are corrected for weather effects, but
the impact of these effects is marginal when considering such random frequencies. This
provides additional support that the variations of the counting rate induced by the variations
of the exposure are under control through the monitoring of $\Delta N_{\mathrm{cell}}(t)$.
\begin{figure}[!t]
\centering
\includegraphics[width=10cm]{Amp.EBin-0.3.6T5.010110.eps}
\includegraphics[width=10cm]{Prob.EBin-0.3.Hex.Weath.6T5.010110.eps}
\caption{\small{Top~: Amplitude of the first harmonic as a function of energy. The dashed
line indicates the 99\% $C.L.$ upper bound on the amplitudes that could result from fluctuations
of an isotropic distribution. Bottom~: Corresponding probabilities to get at least the same
amplitude from an underlying isotropic distribution.}}
\label{amp_diff}
\end{figure}
\subsection{Results at the sidereal frequency in independent energy bins}
To perform first harmonic analyses as a function of energy, the choice of the size of the
energy bins, although arbitrary, is important to avoid the dilution of a genuine
signal with the background noise. In addition, the inclusion of intervals whose
width is below the energy resolution or with too few data is most likely to weaken the
sensitivity of the search for an energy-dependent anisotropy~\cite{alan}. To fulfill both
requirements, the size of the energy intervals is chosen to be $\Delta\log_{10}(E)=0.3$ below
8~EeV, so that it is larger than the energy resolution even at low energies. At higher
energies, to guarantee the determination of the amplitude measurement within an uncertainty
$\sigma\simeq2\%$, all events ($\simeq 5,000$) with energies above 8~EeV are gathered
in a single energy interval.
The amplitude $r$ at the sidereal frequency as a function of the energy is shown in
Fig.~\ref{amp_diff}, together with the corresponding probability $P(>r)$ to get a larger
amplitude in each energy interval for a statistical fluctuation of isotropy. The dashed
line indicates the 99\% $C.L.$ upper bound on the amplitudes that could result from
fluctuations of an isotropic distribution. It is apparent that there is no evidence of any
significant signal over the whole energy range. A global statement refering to the probability with
which the 6 observed amplitudes could have arisen from an underlying isotropic distribution can
be made by comparing the measured value $K=\sum_{i=1}^6k_{0_i}$ (where the sum is over all 6
independent energy intervals) with that expected from a random distribution for which
$\left<K\right>=6$~\cite{edge}. The statistics of $2K$ under the hypothesis of an isotropic sky
is a $\chi^2$ with $2\times6=12$ degrees of freedom. For our data, $2K=19.0$ and the associated
probability for an equal or larger value arising from an isotropic sky is $\simeq 9$\%.
\begin{figure}[!t]
\centering
\includegraphics[width=10cm]{PlotPhases.eps}
\caption{\small{Phase of the first harmonic as a function of energy. The dashed line,
resulting from an empirical fit, is used in the likelihood ratio test (see text).}}
\label{phase_diff}
\end{figure}
The phase $\varphi$ of the first harmonic is shown in Fig.~\ref{phase_diff} as a function
of the energy. While the measurements of the amplitudes do not provide any evidence for
anisotropy, we note that the measurements in adjacent energy intervals suggest a smooth
transition between a common phase of $\simeq 270^\circ$ in the first two bins below
$\simeq1$~EeV compatible with the right ascension of the Galactic Center
$\alpha_{GC}\simeq268.4^\circ$, and another phase ($\alpha\simeq100^\circ$) above
$\simeq5$~EeV. This is intriguing, as the phases are expected to be randomly distributed
in case of independent samples whose parent distribution is isotropic. Knowing the p.d.f.
of phase measurements drawn from an isotropic distribution, $p_0(\varphi)=(2\pi)^{-1}$,
and drawn from a population of directions having a non-zero amplitude $r_0$ with a phase
$\varphi_0$, $p_1(\varphi;r_0,\varphi_0)$~\cite{linsley-fh}, the likelihood functions of
any of the hypotheses may be built as~:
\begin{equation}
\label{eqn:likelihoods}
L_0=\prod_{i=1}^{N_{bins}}p_0(\varphi_i), \hspace{1cm} L_1=\prod_{i=1}^{N_{bins}}p_1(\varphi_i;r_0,\varphi_0).
\end{equation}
Without any knowledge of the expected amplitudes $r_0(E)$ in each bin, the values considered
in $L_1$ are the measurements performed in each energy interval. For the expected phases
$\varphi_0(E)$ as a function of energy, we use an arctangent function adjusted on the data
as illustrated by the dashed line in Fig.~\ref{phase_diff}. Since the smooth evolution
of the phase distribution is potentially interesting but observed \textit{a posteriori},
we aim at testing the fraction of random samples whose behaviour in adjacent energy bins would
show such a potential interest but with no reference to the specific values observed in the data.
To do so, we use the method of the likelihood ratio test, computing the $-2\ln(\lambda)$ statistic
where $\lambda=L_0/L_1$. Using only $N_{bins}=6$, the asymptotic behaviour of the $-2\ln(\lambda)$
statistic is not reached. Hence, the p.d.f. of $-2\ln(\lambda)$ under the hypothesis of isotropy
is built by repeating exactly the same procedure on a large number of isotropic samples~: in
each sample, the arctangent parameters are left to be optimised, and the corresponding value
of $-2\ln(\lambda)$ is calculated. In that way, any alignments, smooth evolutions or abrupt
transitions of phases in random samples are captured and contribute to high values of
the $-2\ln(\lambda)$ distribution. The probability that the hypothesis of isotropy better
reproduces our phase measurements compared to the alternative hypothesis is then calculated
by integrating the normalised distribution of $-2\ln(\lambda)$ above the value measured in the
data. It is found to be $\simeq 2\cdot 10^{-3}$.
It is important to stress that no confidence level can be built from this report as we did not
perform an \textit{a priori} search for a smooth transition in the phase measurements. To confirm
the detection of a real transition using only the measurements of the phases with an independent
data set, we need to collect $\simeq 1.8$ times the number of events analysed here to reach an
efficiency of $\simeq 90\%$ to detect the transition at 99\% $C.L.$ (in case the observed effect
is genuine). It is also worth noting that with a real underlying anisotropy, a consistency of the
phase measurements in ordered energy intervals is expected with lower statistics than the detection
of amplitudes significantly standing out of the background noise~\cite{edge,alan-linsley}. This
behaviour was pointed out by Linsley, quoted in~\cite{edge}~: ``if the number of events available
in an experiment is such that the RMS value of $r$ is equal to the true amplitude, then in a
sequence of experiments $r$ will be significant (say $P(>r)<1\%$) in one experiment out of ten
whereas the phase will be within 50$^\circ$ of the true phase in two experiments out of three.''
We have checked this result using Monte Carlo simulations.
An apparent constancy of phase, even when the significances of the amplitudes are relatively
small, has been noted previously in surveys of measurements made in the range
$10^{14}<E<10^{17}\,$eV~\cite{greisen,LW}. In~\cite{greisen} Greisen and his colleagues comment
that most experiments have been conducted at northern latitudes and therefore the reality of
the sidereal waves is not yet established. The present measurement is made with events coming
largely from the southern hemisphere.
\subsection{Additional cross-checks against systematic effects above 1$\,$EeV}
\label{addcrosschecks}
It is important to verify that the phase effect is not a manifestation of systematic effects,
the amplitudes of which are at the level of the background noise. We provide hereafter additional
studies above 1~EeV, where a few tests can cross-check results presented in Fig.~\ref{phase_diff}.
\begin{figure}[!t]
\centering
\includegraphics[width=7.5cm]{SystAmpCstExpo.eps}
\includegraphics[width=7.5cm]{SystPhasesCstExpo.eps}
\caption{\small{Rayleigh analysis above 1~EeV using a reduced data set built to get at a constant
exposure in right-ascension, using on-times of 98\% (triangles) and 99\% (squares) (see text).
Left~: analysis of the amplitude. The full (dashed) line indicates the 99\% $C.L.$ upper bound on
the amplitudes that could result from fluctuations of an isotropic distribution when using on-times
of 98\% (99\%). Right~: analysis of the phase, the dashed line being the same as the one plotted
in Fig.~\ref{phase_diff}. Results of Fig.~\ref{phase_diff} are also shown with circles.}}
\label{syst_cstexpo}
\end{figure}
The first cross-check is provided by applying the Rayleigh analysis on a reduced data set built
in such a way that its corresponding exposure in right ascension is uniform. This can be achieved
by selecting for each sidereal day only events triggering an unitary cell whose on-time was almost
100\% over the whole sidereal day. To keep a reasonably large data set, we present here the results
obtained for on-times of 98\% and 99\%. This allows us to use respectively $\simeq77$\%
and $\simeq63$\% of the cumulative data set without applying any correction to account for a
non-constant exposure. The results are shown in Fig.~\ref{syst_cstexpo} when considering on-time
of 98\% (triangles) and 99\% (squares). Even if more noisy due to the reduction of the statistics
with respect to the Rayleigh analysis applied on the cumulative data set, they are consistent
with the weighted Rayleigh analysis and support that results presented in Fig.~\ref{phase_diff}
are not dominated by any residual systematics induced by the non-uniform exposure.
\begin{figure}[!t]
\centering
\includegraphics[width=7.5cm]{Syst.AntiSidereal.Ray.Amp.200000.eps}
\includegraphics[width=7.5cm]{Syst.AntiSidereal.Ray.200000.eps}
\includegraphics[width=7.5cm]{AntiSid.Amp.Syst.eps}
\includegraphics[width=7.5cm]{AntiSid.Syst.eps}
\caption{\small{Top~:Anti-sidereal analysis of mock samples built such that a spurious
anti-sidereal amplitude stands out from the noise through the sideband mechanism between
1 and 2~EeV (see text). Thin histograms~: standard Rayleigh analysis of the amplitudes
(left panel) and phases (right panel) for isotropic samples. Thick histograms~: same for
samples biased by the energy variations induced by the atmospheric changes. Dashed histograms~:
same by \emph{reversing} the energy corrections. Filled histograms~: same by \emph{amplifying} by
10 the energy variations. Bottom~: Rayleigh analysis of the anti-sidereal amplitude (left) and
phase (right) of the real data sample without applying the energy corrections (squares), by
applying those corrections (triangles), and by reversing the energy corrections (circles).}}
\label{syst_antisid}
\end{figure}
From the Fourier analysis presented in section~\ref{sec:solar-antisid}, we have stressed
the decoupling between the solar frequency and both the sidereal and anti-sidereal ones
thanks to the frequency resolution reached after 6 years of data taking. However, as the
amplitude of an eventual sideband effect is \emph{proportional} to the solar amplitude
~\cite{far-sto}, it remains important to estimate the impact of an eventual sideband
effect persisting even after the energy corrections. To probe the magnitude of this sideband
effect, we use 10,000 mock data sets generated from the real data set (with energies
corrected for weather effects) by randomising the arrival times but meanwhile keeping
both the zenith and the azimuth angles of each original event. This procedure guarantees
the production of isotropic samples drawn from a uniform exposure with the same detection
efficiency conditions than the real data. The results of the Rayleigh analysis applied to
each mock sample between 1 and 2~EeV at the \emph{anti-sidereal} frequency are shown by the thin
histograms in top panels of Fig.~\ref{syst_antisid}, displaying Rayleigh distributions
for the amplitude measurements and uniform distributions for the phase measurements. Then,
after introducing in to each sample the temporal variations of the energies induced by the atmospheric
changes according to Eqn.~\ref{sweather}, it can be seen on the same graph (thick
histograms) that the amplitude measurements are almost undistinguishable with respect to
the reference ones, while the phase measurements start to show to a small extent a
preferential direction. The same conclusions hold when \emph{reversing} the energy corrections
(dashed histograms), but resulting in a phase shift of $\simeq 180^\circ$. Finally,
the filled histograms are obtained by \emph{amplifying} by 10 the energy variations induced by
the atmospheric changes. In this latter case, the large increase of the solar amplitude
induces a clear signal at the anti-sidereal frequency through the sideband mechanism,
as evidenced by the distributions of both the amplitudes and the phases. The sharp maximum
of the phase distribution points towards the spurious direction, while the amplitude
distribution follows a non-centered Rayleigh distribution with parameter $\simeq 1.4\times 10^{-2}$.
The spurious mean amplitude stands out from the noise ($\sim\sqrt{\pi/200,000}\simeq 4\times 10^{-3}$)
sufficiently to allow us to estimate empirically the original effect to be ten times
smaller, at the level of $\simeq 1.4\times 10^{-3}$. This will impact the analyses only in a marginal
way even if we had not performed the energy corrections, or if we had over-corrected the energies.
The energy corrections necessarily reduce even more the size of the sideband amplitude, well
below $\simeq 10^{-3}$. Hence, only small changes are expected in the anti-sidereal phase
measurements on the real data when applying (or not) or reversing the energy corrections. This is
found to be the case, as illustrated in the bottom panels of Fig.~\ref{syst_antisid}. In addition,
this phase distribution (bottom right panel of Fig.~\ref{syst_antisid}) does not show any
particular structure aligned with the spurious direction. This second cross-check support the
hypothesis that the phase measurements presented in Fig.~\ref{phase_diff} are not dominated by
any residual systematics induced by the sideband mechanism.
\begin{figure}[!t]
\centering
\includegraphics[width=7.5cm]{SystPhasesEW.eps}
\includegraphics[width=7.5cm]{Syst.10x.corrections.EW.200000.eps}
\caption{\small{Left~: East/West analysis at the sidereal frequency above 1~EeV. Right~:
East/West analyses at both the solar (dashed) and the anti-sidereal (thick) frequencies of mock
samples built by amplifying by 10 the energy variations induced by weather effects.}}
\label{ew_syst}
\end{figure}
In the left panel of Fig.~\ref{ew_syst} we show the results of the East/West analysis (circles).
They provide further support of the previous analyses (squares). As previously explained,
this method relies on the high symmetry between the Eastern and Western sectors.
However, the array being slightly tilted ($\simeq 0.2^\circ$), this symmetry is slightly broken,
resulting in a small shift between the Eastward and Westward counting rates. As this shift is
independent of time, it does not impact itself in the estimate of the first harmonic. However,
it is worth examining the effect of the \emph{combination} of the tilted array together with
the spurious modulations induced by weather effects, as this combination may mimic a real
East/West first harmonic modulation at the solar frequency. The size of such an effect can be probed
by analysing the mock data sets built by amplifying by 10 the energy variations induced by the
atmospheric changes. The results obtained between 1 and 2~EeV at both the solar and the
anti-sidereal frequencies are shown in the right panel of Fig.~\ref{ew_syst}~: while the phase
distribution starts to show a preferential direction at the solar frequency (dashed histogram),
the same distribution is still uniform at the anti-sidereal one (thick histogram). Hence, it is
safe to conclude that the results obtained at the sidereal frequency by means of the East/West
method are not affected by any systematics.
\subsection{Results at the sidereal frequency in cumulative energy bins}
\begin{figure}[!t]
\centering
\includegraphics[width=10cm]{Amp.ETh.6T5.010110.eps}
\includegraphics[width=10cm]{Prob.ETh.Hex.Weath.6T5.010110.eps}
\includegraphics[width=10cm]{Phase.ETh.Hex.Weath.6T5.010110.eps}
\caption{\small{Same as Fig.~\ref{amp_diff} and Fig.~\ref{phase_diff}, but as a function of
energy thresholds.}}
\label{amp_thres}
\end{figure}
Performing the same analysis in terms of energy thresholds may be convenient for optimizing the
detection of an eventual genuine signal spread over a large energy range, avoiding the arbitrary
choice of a bin size $\Delta\log_{10}(E)$. The bins are however strongly correlated, preventing
a straightforward interpretation of the evolution of the points with energy. The results on the
amplitudes are shown in Fig.~\ref{amp_thres}. They do not provide any further evidence in favor
of a significant amplitude.
\section{Upper limits and discussion}
\label{sec:discussion}
\begin{table*}[t]
\begin{center}
\begin{tabular}{c|c|c|c|c|c|c}
$\Delta E$ & $N$ & $r_{\mathrm{sidereal}}$[\%] & $P(>r_{\mathrm{sidereal}})[\%]$ & $\varphi$ [$^\circ$]& $\Delta \varphi$ [$^\circ$] & $d_\perp^{UL}$ [\%]\\
\hline
\hline
0.25 - 0.5 & 553639 & 0.4 & 67 & 262 & 64 & 1.3\\
0.5 - 1 & 488587 & 1.2 & 2 & 281 & 20 & 1.7\\
1 - 2 & 199926 & 0.5 & 22 & 15 & 33& 1.4\\
2 - 4 & 50605 & 0.8 & 47 & 39 & 46 & 2.3\\
4 - 8 & 12097 & 1.8 & 35 & 82 & 39 & 5.5\\
$>$8 & 5486 & 4.1 & 9 & 117 & 27 & 9.9
\end{tabular}
\caption{\small{Results of first harmonic analyses in different energy intervals, using the East/West
analysis below 1~EeV and the Rayleigh analysis above 1~EeV.}}
\label{tab2}
\end{center}
\end{table*}%
\begin{figure}[t]
\centering
\includegraphics[width=13cm]{UppLim.eps}
\caption{\small{Upper limits on the anisotropy amplitude of first harmonic as a function of
energy from this analysis. Results from EAS-TOP, AGASA, KASCADE and KASCADE-Grande experiments
are displayed too. An analysis of the KASCADE-Grande data with the East/West method delivers
an additional limit for $3\,10^{15}\,$eV.
Also shown are the predictions up to 1~EeV from two different galactic magnetic field models with
different symmetries ($A$ and $S$), the predictions for a purely galactic origin of UHECRs up to
a few tens of $10^{19}\,$eV ($Gal$), and the expectations from the Compton-Getting effect for an
extragalactic component isotropic in the CMB rest frame ($C$-$G\,Xgal$).}}
\label{UL}
\end{figure}
From the analyses reported in the previous Section, upper limits on amplitudes at 99\% $C.L.$ can
be derived according to the distribution drawn from a population characterised by an anisotropy of
unknown amplitude and phase as derived by Linsley~\cite{linsley-fh}~:
\begin{equation}
\label{eqn:ul}
\sqrt{\frac{2}{\pi}}\frac{1}{I_0(r^2/4\sigma^2)}\int_0^{r_{UL}} \frac{\mathrm{d}s}{\sigma}\,I_0\bigg(\frac{rs}{\sigma^2}\bigg)\,\exp{\bigg(-\frac{s^2+r^2/2}{2\sigma^2}\bigg)}=C.L.,
\end{equation}
where $I_0$ is the modified Bessel function of the first kind with order 0, and $\sigma=\sqrt{2/\mathcal{N}}$
in case of the Rayleigh analysis, and $\sigma_{EW}=(\pi\left<\cos{\delta}\right>/2\left<\sin{\theta}\right>)\times\sqrt{2/N}$
in case of the East/West analysis.
As discussed in the Appendix, the Rayleigh amplitude measured by an observatory
depends on its latitude and on the range of zenith angles considered.
The measured amplitude can be related to a real equatorial dipole component
$d_\perp$ by $d_\perp \simeq r/\langle \cos \delta\rangle$. This is the physical
quantity of interest to compare results from different experiments and from model
predictions. The upper limits on $d_\perp$ are given in Tab.~\ref{tab2} and shown
in Fig.\ref{UL}, together with previous results from EAS-TOP~\cite{eastop},
KASCADE~\cite{kascade}, KASCADE-Grande~\cite{grande} and AGASA~\cite{agasa}, and
with some predictions for the anisotropies arising from models of both galactic and
extragalactic UHECR origin. The results obtained in this study are not consistent
with the $\simeq4$\% anisotropy reported by AGASA in the energy range $1<E/\mathrm{EeV}<2$.
If the galactic/extragalactic transition occurs at the ankle energy~\cite{linsley-ankle},
UHECRs at 1$\,$EeV are predominantly of galactic origin and their escape from the galaxy
by diffusion and drift motions are expected to induce a modulation in this energy range.
These predictions depend on the assumed galactic magnetic field model as well as
on the source distribution and the composition of the UHECRs\footnote{The dependence of
the detection efficiency on the primary mass below 3$\,$EeV could affect the details of a direct
comparison with a model based on a mixed composition.}. Two alternative models are displayed
in Fig. \ref{UL}, corresponding to different geometries of the halo magnetic fields~\cite{roulet1}.
The bounds reported here already exclude the particular model with an antisymmetric halo
magnetic field ($A$) and are starting to become sensitive to the predictions of the model
with a symmetric field ($S$). We note that those models assume a predominantly heavy
composition galactic component at EeV energies, while scenarios in which galactic protons
dominate at those energies would typically predict anisotropies larger than the bounds
obtained in Fig.~\ref{UL}. Maintaining the amplitudes of such anisotropies within our
bounds necessarily translates into constraints upon the description of the halo magnetic
fields and/or the spatial source distribution. This is particularly interesting in the
view of our composition measurements at those energies compatible with a light
composition~\cite{auger-er}. Aternatively to a leaky galaxy model, there is still the
possibility that a large scale magnetic field retains all particles in the
galaxy~\cite{peters,jokipii}. If the structure of the magnetic fields in the halo is
such that the turbulent component predominates over the regular one, purely
diffusion motions may confine light elements of galactic origin up to $\simeq 1\,$EeV
and may induce an ankle feature due to the longer confinement of heavier elements at
higher energies~\cite{calvez}. Typical signatures of such a scenario in terms of large scale
anisotropies are also shown in Fig.~\ref{UL} (dotted line)~: the corresponding amplitudes are
challenged by our current sensitivity.
On the other hand, if the transition is taking place at lower energies around the
second knee at $\simeq5\times10^{17}\,$eV~\cite{berezinsky}, UHECRs above 1~EeV are
dominantly of extragalactic origin and their large scale distribution could be influenced
by the relative motion of the observer with respect to the frame of the sources. If the
frame in which the UHECR distribution is isotropic coincides with the CMB rest frame,
a small anisotropy is expected due to the Compton-Getting effect. Neglecting the effects
of the galactic magnetic field, this anisotropy would be a dipolar pattern pointing in
the direction $\alpha\simeq 168^\circ$ with an amplitude of about 0.6\%~\cite{kachelriess}.
On the contrary, when accounting for the galactic magnetic field, this dipolar anisotropy
is expected to also affect higher order multipoles~\cite{silvia}. These amplitudes are close
to the upper limits set in this analysis, and the statistics required to detect an amplitude
of 0.6\% at 99\% $C.L.$ is $\simeq 3$ times the present one.
Continued scrutiny of the large scale distribution of arrival directions of UHECRs as a
function of energy with the increased statistics provided by the Pierre Auger Observatory,
above a few times $10^{17}\,$eV, will help to discriminate between a predominantly galactic
or extragalactic origin of UHECRs as a function of the energy, and so benefit the search
for the galactic/extragalactic transition. Future work will profit from the lower energy
threshold that is now available at the Pierre Auger Observatory~\cite{infill}.
\section*{}
This article is dedicated to \emph{Gianni Navarra}, who has been deeply involved in this study for
many years and who has inspired several of the analyses described in this paper. His legacy lives
on.
\section*{Appendix}
The first harmonic amplitude of the distribution in right ascension of the detected cosmic rays can be
directly related to the amplitude $d$ of a dipolar distribution of the cosmic ray flux of the form
$J(\alpha,\delta) = J_0 (1+ d \ \hat d \cdot \hat u)$, where $\hat u$ and $\hat d$ denote respectively the
unit vector in the direction of an arrival direction and in the direction of the dipole. Eqn.~\ref{eqn:fh}
and Eqn.~\ref{eqn:fh2} can be used to express $a$, $b$ and $\mathcal{N}$ as~:
\begin{eqnarray}
a&=& \frac{2}{\mathcal{N}} \int_{\delta_{min}}^{\delta_{max}} d\delta \int_0^{2\pi}
d\alpha \cos \delta \ J(\alpha,\delta) \ \omega(\delta) \cos \alpha,\\
b&=& \frac{2}{\mathcal{N}} \int_{\delta_{min}}^{\delta_{max}} d\delta \int_0^{2\pi}
d\alpha \cos \delta \ J(\alpha,\delta) \ \omega(\delta) \sin \alpha, \nonumber \\
\mathcal{N}&=& \int_{\delta_{min}}^{\delta_{max}} d\delta \int_0^{2\pi}
d\alpha \cos \delta \ J(\alpha,\delta) \ \omega(\delta),\nonumber
\end{eqnarray}
where we have here neglected in the exposure $\omega$ the small dependence on right
ascension. Writing the angular dependence in
$J(\alpha,\delta)$ as $ \hat d \cdot \hat u = \cos \delta \cos \delta_d
\cos (\alpha-\alpha_d) + \sin \delta \sin \delta_d$, with $\delta_d$ the
dipole declination and $\alpha_d$ its right ascension, and performing the
integration in $\alpha$ in the previous equations, it can be seen that
\begin{equation}
\label{eqn:amplitudes}
r=\left| \frac{Ad_\perp}{1+Bd_z} \right|
\end{equation}
where
$$A =\frac{\int d\delta\,\omega(\delta) \cos^2 \delta}
{\int d\delta\,\omega(\delta) \cos \delta}, \hspace{1cm}
B =\frac{\int d\delta\,\omega(\delta) \cos \delta \sin \delta}
{\int d\delta\,\omega(\delta) \cos \delta}$$
and $d_z=d\sin{\delta_d}$ denotes the component of the dipole along the
Earth rotation axis while $d_\perp=d\cos{\delta_d}$ is the component in the
equatorial plane \cite{julien}. The coefficients $A$ and $B$ can be
estimated from the data as the mean values of the cosine
and the sine of the event declinations. In our case,
$A=\left<\cos{\delta}\right>\simeq 0.78$ and $B=\left<\sin{\delta}\right>\simeq -0.45$.
For a dipole amplitude $d$, the measured amplitude of the first harmonic
in right ascension $r$ thus depends on the
region of the sky observed, which is essentially a function of the latitude
of the observatory $\ell_{site}$, and the range of zenith angles considered. In the case of a
small $B d_z$ factor, the dipole component in the equatorial plane $d_\perp$
is obtained as $d_\perp\simeq r/\left<\cos{\delta}\right>$.
The phase $\varphi$ corresponds to the right ascension of the dipole direction
$\alpha_d$.
Turning now to the East-West method, the measured flux from the East sector
for a local sidereal time $\alpha^0$ can be similarly expressed as
\begin{equation}
I_E (\alpha^0)=\int_{-\pi/2}^{\pi/2} d\phi \int_0^{\theta_{max}} d\theta
\sin \theta\ \epsilon (\theta) \ J(\theta,\phi,\alpha^0),
\end{equation}
and analogously for the measured flux coming from the west sector
changing the azimuthal integration to the interval
$[\pi/2,3\pi/2]$. Expressing $\hat d \cdot \hat u$ in local coordinates ($\theta$, $\phi$
and $\alpha^0$), and performing the integration over $\phi$ we obtain for the leading order
\begin{equation}
\frac{I_E - I_W}{\langle I_E + I_W \rangle} (\alpha^0) =-\frac{2 d_\perp C
}{\pi (1+d_z D \sin \ell_{site})} \sin (\alpha^0 -\alpha_d),
\end{equation}
where
$$C =\frac{\int d\theta\,\epsilon(\theta) \sin^2 \theta
}{\int d\theta\,\epsilon(\theta) \sin \theta},\hspace{1cm}
D =\frac{\int d\theta\,\epsilon(\theta) \sin \theta \cos \theta}{
\int d\theta\,\epsilon(\theta) \sin \theta}.$$
In this calculation any dependence of the exposure on the local sidereal time $\alpha^0$ gives
at first order the same contribution to the East and West sectors flux, and thus gives a negligible
contribution to the flux difference\footnote{The exposure dependence on azimuth present in Auger
due to trigger effects at low energies has a $2\pi/6$ frequency that makes it cancel out in the
computation of the first harmonic, and thus does not affect the above result.}. The next leading
order, proportional to the equatorial dipole component times the sidereal modulation of the exposure,
is negligible. The coefficients $C$ and $D$ can be estimated from the observed zenith angles of the
events. In our case, $C =\left<\sin\theta\right>\simeq 0.58$ and
$D=\left<\cos\theta\right>\simeq 0.78$. The total detected flux averaged over the local sidereal
time can be estimated as $\langle I_E +I_W\rangle = N/2\pi$. In case $Dd_z\ll 1$, we get finally~:
\begin{equation}
(I_E - I_W) (\alpha^0) = - \frac{N}{2\pi}\frac{2 d_\perp \langle \sin \theta
\rangle}{\pi} \sin (\alpha^0 - \alpha_d).
\end{equation}
\section*{Acknowledgements}
The successful installation and commissioning of the Pierre Auger Observatory
would not have been possible without the strong commitment and effort
from the technical and administrative staff in Malarg\"ue.
We are very grateful to the following agencies and organizations for financial support:
Comisi\'on Nacional de Energ\'{\i}a At\'omica,
Fundaci\'on Antorchas,
Gobierno De La Provincia de Mendoza,
Municipalidad de Malarg\"ue,
NDM Holdings and Valle Las Le\~nas, in gratitude for their continuing
cooperation over land access, Argentina;
the Australian Research Council;
Conselho Nacional de Desenvolvimento Cient\'{\i}fico e Tecnol\'ogico (CNPq),
Financiadora de Estudos e Projetos (FINEP),
Funda\c{c}\~ao de Amparo \`a Pesquisa do Estado de Rio de Janeiro (FAPERJ),
Funda\c{c}\~ao de Amparo \`a Pesquisa do Estado de S\~ao Paulo (FAPESP),
Minist\'erio de Ci\^{e}ncia e Tecnologia (MCT), Brazil;
AVCR, AV0Z10100502 and AV0Z10100522,
GAAV KJB300100801 and KJB100100904,
MSMT-CR LA08016, LC527, 1M06002, and MSM0021620859, Czech Republic;
Centre de Calcul IN2P3/CNRS,
Centre National de la Recherche Scientifique (CNRS),
Conseil R\'egional Ile-de-France,
D\'epartement Physique Nucl\'eaire et Corpusculaire (PNC-IN2P3/CNRS),
D\'epartement Sciences de l'Univers (SDU-INSU/CNRS), France;
Bundesministerium f\"ur Bildung und Fors-
chung (BMBF),
Deutsche Forschungsgemeinschaft (DFG),
Finanzministerium Baden -
W\"urttemberg,
Helmholtz-Gemeinschaft Deutscher Forschungszentren (HGF),
Ministerium f\"ur Wissenschaft und Forschung, Nordrhein-Westfalen,
Ministerium f\"ur Wissenschaft, Fors-
chung und Kunst, Baden-W\"urttemberg, Germany;
Istituto Nazionale di Fisica Nucleare (INFN),
Istituto Nazionale di Astrofisica (INAF),
Ministero dell'Istruzione, dell'Universit\`a e della Ricerca (MIUR),
Gran Sasso Center for Astroparticle Physics (CFA), Italy;
Consejo Nacional de Ciencia y Tecnolog\'{\i}a (CONACYT), Mexico;
Ministerie van Onderwijs, Cultuur en Wetenschap,
Nederlandse Organisatie voor Wetenschappelijk Onderzoek (NWO),
Stichting voor Fundamenteel Onderzoek der Materie (FOM), Netherlands;
Ministry of Science and Higher Education,
Grant Nos. 1 P03 D 014 30 and N N202 207238, Poland;
Funda\c{c}\~ao para a Ci\^{e}ncia e a Tecnologia, Portugal;
Ministry for Higher Education, Science, and Technology,
Slovenian Research Agency, Slovenia;
Comunidad de Madrid,
Consejer\'{\i}a de Educaci\'on de la Comunidad de Castilla La Mancha,
FEDER funds,
Ministerio de Ciencia e Innovaci\'on and Consolider-Ingenio 2010 (CPAN),
Generalitat Valenciana,
Junta de Andaluc\'{\i}a,
Xunta de Galicia, Spain;
Science and Technology Facilities Council, United Kingdom;
Department of Energy, Contract Nos. DE-AC02-07CH11359, DE-FR02-04ER41300,
National Science Foundation, Grant No. 0450696,
The Grainger Foundation USA;
ALFA-EC / HELEN,
European Union 6th Framework Program,
Grant No. MEIF-CT-2005-025057,
European Union 7th Framework Program, Grant No. PIEF-GA-2008-220240,
and UNESCO.
\section*{References}
|
\section*{Universal low-rank matrix recovery from Pauli measurements:}
\section*{Supplementary material}
\section{Proof of the RIP for Pauli Measurements}
\label{app3}
\subsection{Overview}
We now prove Theorem \ref{thm-rip}. In this section we give an overview; proofs of the technical claims are deferred to later sections. The general approach involving Dudley's entropy bound is similar to \cite{RV06}, while the technical part of the proof (bounding certain covering numbers) uses ideas from \cite{Guedon08}.
Recall the definition of the restricted isometry property, with constant $0 \leq \delta < 1$. Let
\begin{equation}
U = \set{X \in \CC^{d\times d} \;|\; \norm{X}_* \leq \sqrt{r}\norm{X}_F}.
\end{equation}
Let us define
\begin{equation}
U_2 = \set{X\in\CC^{d\times d} \;|\; \norm{X}_F \leq 1, \; \norm{X}_* \leq \sqrt{r}\norm{X}_F},
\end{equation}
\begin{equation}
\vareps_r(\calA) = \sup_{X\in U_2} |(X, (\calA^*\calA-\calI) X)|.
\end{equation}
Also, define $\vareps = 2\delta - \delta^2$. We claim that, to show RIP, it suffices to show $\vareps_r(\calA) < \vareps$. To see this, note that the RIP condition is equivalent to the statement
\begin{equation}
\text{for all $X\in U$}, \quad
(1-\delta)^2 (X, X) \leq (X, \calA^*\calA X)
\leq (1+\delta)^2 (X, X),
\end{equation}
which is equivalent to
\begin{equation}
\text{for all $X\in U$}, \quad
(-2\delta+\delta^2) (X, X) \leq (X, (\calA^*\calA-\calI) X)
\leq (2\delta+\delta^2) (X, X),
\end{equation}
which is implied by
\begin{equation}
\text{for all $X\in U_2$}, \quad
|(X, (\calA^*\calA-\calI) X)| \leq \min\set{2\delta+\delta^2, 2\delta-\delta^2} = 2\delta-\delta^2.
\end{equation}
Thus our goal is to show $\vareps_r(\calA) < \vareps$. (Note that for $\delta$ in the range $[0,1]$, we have that $\vareps \geq \delta$.)
Let $\calB$ be the set of all self-adjoint linear operators from $\CC^{d\times d}$ to $\CC^{d\times d}$, and define the following norm on $\calB$:
\begin{equation}
\label{eqn-rnorm}
\norm{\calM}_{(r)} = \sup_{X\in U_2} |(X,\calM X)|.
\end{equation}
Suppose that $r \geq 2$ (this will suffice for our purposes, since RIP with $r=2$ implies RIP with $r=1$). We claim that $\norm{\cdot}_{(r)}$ is a norm, and that $\calB$ is a Banach space with respect to this norm.
To show these claims, we will consider the Frobenius norm $\norm{\cdot}_F$ on $\calB$, which is defined by viewing each element of $\calB$ as a ``matrix'' acting on ``vectors'' that are elements of $\CC^{d\times d}$. Then we will bound $\norm{\cdot}_{(r)}$ in terms of $\norm{\cdot}_F$. More precisely, let $\vec{e}_a$ ($a \in \set{0,1,\ldots,d-1}$) be the standard basis vectors in $\CC^d$, and let $E_{ab} = \vec{e}_a \vec{e}_b^*$ ($a,b \in \set{0,1,\ldots,d-1}$) be the standard basis vectors in $\CC^{d\times d}$. Then the Frobenius norm on $\calB$ can be written as
\begin{equation}
\norm{\calM}_F = \Bigl( \sum_{abcd} \Bigl| \Bra{E_{cd}}\calM\Ket{E_{ab}} \Bigr|^2 \Bigr)^{1/2}.
\end{equation}
We claim that, for all $\calM \in \calB$,
\begin{equation} \label{eqn-rnorm-geq-fnorm}
\norm{\calM}_{(r)} \geq \frac{1}{3\sqrt{2}d^2} \norm{\calM}_F.
\end{equation}
To see this, suppose that $\norm{\calM}_F \geq \mu$; then there must exist $a,b,c,d \in \set{0,1,\ldots,d-1}$ such that $\bigl| (E_{cd}| \calM |E_{ab}) \bigr| \geq \frac{1}{d^2} \mu$. If $E_{ab} = E_{cd}$, then we have $\norm{\calM}_{(r)} \geq \frac{1}{d^2}\mu$. Otherwise, we have $(E_{ab}|E_{cd}) = 0$. Now at least one of the following must be true:
\begin{equation}
\bigl| \text{Re} \Bra{E_{cd}}\calM\Ket{E_{ab}} \bigr| \geq \tfrac{1}{\sqrt{2} d^2} \mu \quad \text{(case 1)},
\end{equation}
\begin{equation}
\bigl| \text{Im} \Bra{E_{cd}}\calM\Ket{E_{ab}} \bigr| \geq \tfrac{1}{\sqrt{2} d^2} \mu \quad \text{(case 2)}.
\end{equation}
In case 1, let $X = \frac{1}{\sqrt{2}} (E_{ab} + E_{cd})$, and write
\begin{equation}
\text{Re} \Bra{E_{cd}}\calM\Ket{E_{ab}} = \Bra{X}\calM\Ket{X} - \tfrac{1}{2} \Bra{E_{ab}}\calM\Ket{E_{ab}} - \tfrac{1}{2} \Bra{E_{cd}}\calM\Ket{E_{cd}}.
\end{equation}
One of the three terms on the right hand side must have absolute value at least $\tfrac{1}{3\sqrt{2} d^2} \mu$. Since $X, E_{ab}, E_{cd}$ are in $U_2$, it follows that $\norm{\calM}_{(r)} \geq \tfrac{1}{3\sqrt{2} d^2} \mu$. In case 2, let $X = \frac{1}{\sqrt{2}} (E_{ab} + iE_{cd})$, and write
\begin{equation}
\text{Im} \Bra{E_{cd}}\calM\Ket{E_{ab}} = i\Bra{X}\calM\Ket{X} - \tfrac{1}{2} i\Bra{E_{ab}}\calM\Ket{E_{ab}} - \tfrac{1}{2} i\Bra{E_{cd}}\calM\Ket{E_{cd}}.
\end{equation}
By a similar argument, we get that $\norm{\calM}_{(r)} \geq \tfrac{1}{3\sqrt{2} d^2} \mu$. This shows (\ref{eqn-rnorm-geq-fnorm}).
In addition, it is straightforward to see that
\begin{equation} \label{eqn-rnorm-leq-fnorm}
\norm{\calM}_{(r)} \leq \sup_{X \::\: \norm{X}_F \leq 1} |(X, \calM X)|
\leq \norm{\calM}_{\text{op}} \leq \norm{\calM}_F.
\end{equation}
Finally, using (\ref{eqn-rnorm-geq-fnorm}) and (\ref{eqn-rnorm-leq-fnorm}), we see that $\norm{\cdot}_{(r)}$ is a norm, and $\calB$ is a Banach space with respect to $\norm{\cdot}_{(r)}$. (This follows since these same properties already hold for $\norm{\cdot}_F$.) In particular, $\norm{\cdot}_{(r)}$ is nondegenerate ($\norm{\calM}_{(r)} = 0$ implies $\calM = 0$), and $\calB$ is complete with respect to $\norm{\cdot}_{(r)}$.
Returning to our main proof, we can now write $\vareps_r(\calA) = \norm{\calA^*\calA-\calI}_{(r)}$. The strategy of the proof will be to first bound $\EE \vareps_r(\calA)$, then show that $\vareps_r(\calA)$ is concentrated around its mean.
We claim that
\begin{equation}
\EE \vareps_r(\calA) \leq 2\EE \Norm{\sum_{j=1}^m \vareps_j \Ket{S_j}\Bra{S_j} \tfrac{d^2}{m}}_{(r)},
\end{equation}
where the $\vareps_j$ are Rademacher (iid $\pm 1$) random variables. Here the round ket notation $\Ket{S_j}$ means we view the matrix $S_j$ as an element of the vector space $\CC^{d^2}$ with Hilbert-Schmidt inner product; the round bra $\Bra{S_j}$ denotes the adjoint element in the (dual) vector space. The above bound follows from a standard symmetrization argument: write $\calA^*\calA-\calI = \sum_{j=1}^m \calX_j$ where $\calX_j = \Ket{S_j}\Bra{S_j}\frac{d^2}{m} - \frac{\calI}{m}$, then let $\calX'_j$ be independent copies of the random variables $\calX_j$, and use equation (2.5) and Lemma 6.3 in \cite{LT91} to write:
\begin{equation}
\begin{split}
\EE\vareps_r(\calA) &= \EE\Norm{\sum_j \calX_j}_{(r)} \\
&\leq \EE\Norm{\sum_j (\calX_j-\calX'_j)}_{(r)} = \EE\norm{\sum_j \vareps_j(\calX_j-\calX'_j)}_{(r)} \\
&= \EE\Norm{\sum_j \vareps_j \Bigl( \Ket{S_j}\Bra{S_j} - \Ket{S'_j}\Bra{S'_j} \Bigr) \tfrac{d^2}{m}}_{(r)} \\
&\leq 2\EE\Norm{\sum_j \vareps_j \Ket{S_j}\Bra{S_j} \tfrac{d^2}{m}}_{(r)}.
\end{split}
\end{equation}
Now we use the following lemma, which we will prove later. This bounds the expected magnitude in $(r)$-norm of a Rademacher sum of a fixed collection of operators $V_1,\ldots,V_m$ that have small operator norm.
\begin{lem}
(restatement of Lemma \ref{lem-key})
Let $m\leq d^2$. Fix some $V_1,\ldots,V_m \in \CC^{d\times d}$ that have uniformly bounded operator norm, $\norm{V_i} \leq K$ (for all $i$). Let $\vareps_1,\ldots,\vareps_m$ be iid uniform $\pm 1$ random variables. Then
\begin{equation}
\EE_\vareps \Norm{\sum_{i=1}^m \vareps_i\Ket{V_i}\Bra{V_i}}_{(r)}
\leq C_5\cdot \Norm{\sum_{i=1}^m \Ket{V_i}\Bra{V_i}}_{(r)}^{1/2},
\end{equation}
where $C_5 = \sqrt{r}\cdot C_4 K \log^{5/2} d \log^{1/2} m$ and $C_4$ is some universal constant.
\end{lem}
We apply the lemma as follows. Let $\Omega = \set{S_1,\ldots,S_m}$ be the multiset of all the measurement operators that appear in the sampling operator $\calA$. Then we have
\begin{equation}
\EE\vareps_r(\calA) \leq 2\EE_\Omega \EE_\vareps \Norm{\sum_{J\in\Omega} \vareps_J \sqrt{d}\Ket{J}\Bra{J}\sqrt{d}}_{(r)} \cdot \tfrac{d}{m}.
\end{equation}
Using the lemma on the set of operators $\sqrt{d}J$ ($J\in\Omega$), we get
\begin{equation}
\begin{split}
\EE\vareps_r(\calA)
&\leq 2\EE_\Omega C_5\cdot \Norm{\sum_{J\in\Omega} \sqrt{d}\Ket{J}\Bra{J}\sqrt{d}}_{(r)}^{1/2} \cdot \tfrac{d}{m} \\
&\leq 2 \Bigl( \EE_\Omega \Norm{\sum_{J\in\Omega} \sqrt{d}\Ket{J}\Bra{J}\sqrt{d}}_{(r)} \Bigr)^{1/2} \cdot C_5 \cdot \tfrac{d}{m} \\
&= 2 \Bigl( \EE \norm{\calA^*\calA}_{(r)} \Bigr)^{1/2} \cdot C_5 \cdot \sqrt{\tfrac{d}{m}} \\
&\leq 2 (\EE\vareps_r(\calA) + 1)^{1/2} \cdot C_5 \cdot \sqrt{\tfrac{d}{m}},
\end{split}
\end{equation}
where $C_5 = \sqrt{r}\cdot C_4 K \log^3 d$.
To make the notation more concise, define $E_0 = \EE\vareps_r(\calA)$ and $C_0 = 2C_5\sqrt{\tfrac{d}{m}}$. Then, squaring both sides and rearranging, we have
\begin{equation}
E_0^2 - C_0^2E_0 - C_0^2 \leq 0.
\end{equation}
This quadratic equation has two roots, which are given by $\alpha_\pm = \tfrac{1}{2} (C_0^2 \pm C_0\sqrt{C_0^2+4})$, and we know that $E_0$ is bounded by
\begin{equation}
\alpha_- \leq 0 \leq E_0 \leq \alpha_+.
\end{equation}
Also, we can simplify the bound by writing $\alpha_+ \leq \tfrac{1}{2} (C_0^2 + C_0(C_0+2)) = C_0^2 + C_0$.
Now we use the fact that $m$ is large. Let $\lambda \geq 1$ (we will choose a precise value for $\lambda$ later). Assume that
\begin{equation}
m \geq \lambda d (2C_5)^2 = \lambda \cdot 4C_4^2 \cdot dr \cdot K^2 \log^6 d.
\end{equation}
Then $C_0 \leq 1/\sqrt{\lambda}$, and $\alpha_+ \leq \tfrac{1}{\lambda} + \tfrac{1}{\sqrt{\lambda}}$, and we have the desired result:
\begin{equation}
\EE\vareps_r(\calA) \leq \tfrac{1}{\lambda} + \tfrac{1}{\sqrt{\lambda}}.
\end{equation}
It remains to show that $\vareps_r(\calA)$ is concentrated around its expectation. We will use a concentration inequality from \cite{LT91} for sums of independent symmetric random variables that take values in some Banach space. Define $\calX = \sum_{j=1}^m \calX_j$ where $\calX_j = \tfrac{d^2}{m} \Ket{S_j}\Bra{S_j} - \tfrac{\calI}{m}$; then we have $\calA^*\calA - \calI = \calX$ and $\vareps_r(\calA) = \norm{\calX}_{(r)}$.
We showed above that $\EE \norm{\calX}_{(r)} \leq \tfrac{1}{\lambda} + \tfrac{1}{\sqrt{\lambda}}$. In addition, we can bound each $\calX_j$ as follows, using the fact that, for $X\in U_2$, $|(S_j,X)| \leq \norm{S_j} \norm{X}_* \leq (K/\sqrt{d}) \sqrt{r}\norm{X}_F \leq (K/\sqrt{d}) \sqrt{r}$.
\begin{equation}
\norm{\calX_j}_{(r)}
= \sup_{X\in U_2} \Bigl| \tfrac{d^2}{m} |(S_j,X)|^2 - \tfrac{1}{m} \norm{X}_F^2 \Bigr|
\leq \frac{drK^2 + 1}{m} \leq \frac{1}{\lambda\cdot 4C_4^2}.
\end{equation}
We use a standard symmetrization argument: let $\calX'_j$ denote an independent copy of $\calX_j$, and define $\calY_j = \calX_j - \calX'_j$, which is symmetric ($-\calY_j$ has the same distribution as $\calY_j$). Also define $\calY = \sum_{j=1}^m \calY_j = \calX-\calX'$. Using the triangle inequality, we have
\begin{equation}
\label{eqn-EY}
\EE \norm{\calY}_{(r)} \leq 2\EE \norm{\calX}_{(r)} \leq 2(\tfrac{1}{\lambda} + \tfrac{1}{\sqrt{\lambda}}),
\end{equation}
\begin{equation}
\label{eqn-Yj}
\norm{\calY_j}_{(r)} \leq 2\norm{\calX_j}_{(r)} \leq \frac{1}{\lambda\cdot 2C_4^2}.
\end{equation}
Using equation (6.1) in \cite{LT91}, we have, for any $u\geq 0$,
\begin{equation}
\label{eqn-Xtail}
\Pr\Bigl[ \norm{\calX}_{(r)} > 2(\tfrac{1}{\lambda} + \tfrac{1}{\sqrt{\lambda}}) + u \Bigr]
\leq \Pr\Bigl[ \norm{\calX}_{(r)} > 2\EE \norm{\calX}_{(r)} + u \Bigr]
\leq 2\Pr\Bigl[ \norm{\calY}_{(r)} > u \Bigr].
\end{equation}
We will use the following concentration inequality of Ledoux and Talagrand \cite{LT91}. This is a special case of Theorem 6.17 in \cite{LT91}, where we set $s=R\ell$ and use equation (6.19) in \cite{LT91}. This is the same bound used in \cite{RV06}.
\begin{thm}
Let $\calY_1,\ldots,\calY_m$ be independent symmetric random variables taking values in some Banach space. Assume that $\norm{\calY_j} \leq R$ for all $j$. Let $\calY = \sum_{j=1}^m \calY_j$. Then, for any integers $\ell \geq q$, and any $t>0$,
\begin{equation}
\Pr\Bigl[ \norm{\calY} \geq 8q\EE\norm{\calY} + 2R\ell + t\EE\norm{\calY} \Bigr]
\leq (C_7/q)^\ell + 2\exp(-t^2/256q),
\end{equation}
where $C_7$ is some universal constant.
\end{thm}
Now set $q = \lceil eC_7 \rceil$. Introduce a new parameter $s \geq \sqrt{q} + 1$, and set $\ell = \lfloor s^2 \rfloor$ and $t = s$. We get that the failure probability is exponentially small in $s$:
\begin{equation}
\Pr\Bigl[ \norm{\calY}_{(r)} \geq (8q + s) \EE\norm{\calY}_{(r)} + 2R s^2 \Bigr]
\leq e^{-s^2+1} + 2e^{-s^2/256q}.
\end{equation}
Then, using (\ref{eqn-EY}), (\ref{eqn-Yj}) and (\ref{eqn-Xtail}), we get
\begin{equation}
\Pr\Bigl[ \norm{\calX}_{(r)} \geq (1 + 8q + s) \cdot 2(\tfrac{1}{\lambda}+\tfrac{1}{\sqrt{\lambda}}) + \tfrac{1}{\lambda C_4^2} s^2 \Bigr]
\leq 2[e^{-s^2+1} + 2e^{-s^2/256q}].
\end{equation}
Now let $\lambda \geq (1+8q)^2 \cdot \frac{256}{\vareps^2}$ (note that $\lambda \geq 1$, as required). Then set $s = \frac{\vareps\sqrt{\lambda}}{16}$ (note that $s \geq 1+8q \geq \sqrt{q} + 1$, as required). Then we can write
\begin{equation}
\begin{split}
(1+8q+s) \cdot 2(\tfrac{1}{\lambda}+\tfrac{1}{\sqrt{\lambda}}) + \tfrac{1}{\lambda C_4^2} s^2
\leq \tfrac{8s}{\sqrt{\lambda}} + \tfrac{s^2}{C_4^2\lambda}
= \tfrac{\vareps}{2} + \tfrac{\vareps^2}{256C_4^2} \leq \vareps.
\end{split}
\end{equation}
Plugging into the previous inequality, we have
\begin{equation}
\Pr[ \norm{\calX}_{(r)} \geq \vareps ] \leq e^{-\Omega(s^2)} = e^{-\Omega(\vareps^2 \lambda)}.
\end{equation}
Therefore, we have $\vareps_r(\calA) \leq \vareps$, with a failure probability that decreases exponentially in $\lambda$. This completes the proof.
\subsection{Proof of Lemma \ref{lem-key} (bounding a Rademacher sum in $(r)$-norm)}
Let $L_0 = \EE_\vareps \norm{\sum_{i=1}^m \vareps_i\Ket{V_i}\Bra{V_i}}_{(r)}$; this is the quantity we want to bound. We can upper-bound it by replacing the $\pm 1$ random variables $\vareps_1,\ldots,\vareps_m$ with iid $N(0,1)$ Gaussian random variables $g_1,\ldots,g_m$ (see Lemma 4.5 and equation (4.8) in \cite{LT91}); then we get
\begin{equation}
L_0 \leq \EE_g \Norm{\sqrt{\tfrac{\pi}{2}} \sum_{i=1}^m g_i\Ket{V_i}\Bra{V_i}}_{(r)}.
\end{equation}
Using the definition of the norm $\norm{\cdot}_{(r)}$ (equation (\ref{eqn-rnorm})), we have
\begin{equation}
L_0 \leq \EE_g \sup_{X\in U_2} \sqrt{\tfrac{\pi}{2}} |G(X)|, \quad
G(X) = \sum_{i=1}^m g_i |(V_i,X)|^2.
\end{equation}
The random variables $G(X)$ (indexed by $X\in U_2$) form a Gaussian process, and $L_0$ is upper-bounded by the expected supremum of this process. In particular, using the fact that $G(0)=0$ and $G(\cdot)$ is symmetric (see \cite{LT91}, pp.298), we have
\begin{equation}
\begin{split}
L_0 &\leq \sqrt{\tfrac{\pi}{2}} \EE_g \sup_{X\in U_2} |G(X)-G(0)|
\leq \sqrt{\tfrac{\pi}{2}} \EE_g \sup_{X,Y\in U_2} |G(X)-G(Y)| \\
&= \sqrt{\tfrac{\pi}{2}} \EE_g \sup_{X,Y\in U_2} G(X)-G(Y)
= \sqrt{2\pi} \EE_g \sup_{X\in U_2} G(X).
\end{split}
\end{equation}
Using Dudley's inequality (Theorem 11.17 in \cite{LT91}), we have
\begin{equation}
\label{eqn-dudley}
L_0 \leq 24 \sqrt{2\pi} \int_0^\infty \log^{1/2} N(U_2,d_G,\vareps) d\vareps,
\end{equation}
where $N(U_2,d_G,\vareps)$ is a covering number (the number of balls in $\CC^{d\times d}$ of radius $\vareps$ in the metric $d_G$ that are needed to cover the set $U_2$), and the metric $d_G$ is given by
\begin{equation}
d_G(X,Y) = \Bigl( \EE[(G(X)-G(Y))^2] \Bigr)^{1/2}.
\end{equation}
We can simplify the metric $d_G$, using the fact that $\EE[g_ig_j] = 1$ when $i=j$ and 0 otherwise:
\begin{equation}
\begin{split}
d_G(X,Y) &= \Bigl( \EE\Bigl[ \Bigl( \sum_{i=1}^m g_i (|(V_i,X)|^2-|(V_i,Y)|^2) \Bigr)^2 \Bigr] \Bigr)^{1/2} \\
&= \Bigl( \sum_{i=1}^m \Bigl( |(V_i,X)|^2-|(V_i,Y)|^2 \Bigr)^2 \Bigr)^{1/2}
\end{split}
\end{equation}
Define a new norm (actually a semi-norm) $\norm{\cdot}_X$ on $\CC^{d\times d}$, as follows:
\begin{equation}
\norm{M}_X = \max_{i=1,\ldots,m} |(V_i,M)|.
\end{equation}
Note that
\footnote{Note that, for any complex numbers $a$ and $b$, $|a|^2 - |b|^2
= \tfrac{1}{2} (\bar{a}+\bar{b}) (a-b) + \tfrac{1}{2} (a+b) (\bar{a}-\bar{b})
\leq |a+b|\cdot |a-b|$.}
\begin{equation}
\begin{split}
\Bigl| |(V_i,X)|^2-|(V_i,Y)|^2 \Bigr|
&\leq \Bigl( |(V_i,X)|+|(V_i,Y)| \Bigr) \cdot |(V_i,X)-(V_i,Y)| \\
&\leq \Bigl( |(V_i,X)|+|(V_i,Y)| \Bigr) \cdot \norm{X-Y}_X.
\end{split}
\end{equation}
This lets us give a simpler upper bound on the metric $d_G$:
\begin{equation}
\begin{split}
d_G(X,Y)
&\leq \Bigl( \sum_{i=1}^m \Bigl( |(V_i,X)|+|(V_i,Y)| \Bigr)^2 \cdot \norm{X-Y}_X^2 \Bigr)^{1/2} \\
&\leq \Bigl[ \Bigl( \sum_{i=1}^m |(V_i,X)|^2 \Bigr)^{1/2} + \Bigl( \sum_{i=1}^m |(V_i,Y)|^2 \Bigr)^{1/2} \Bigr] \cdot \norm{X-Y}_X \\
&\leq 2\sup_{X\in U_2} \Bigl( \sum_{i=1}^m |(V_i,X)|^2 \Bigr)^{1/2} \cdot \norm{X-Y}_X \\
&= 2\Norm{\sum_{i=1}^m \Ket{V_i}\Bra{V_i}}_{(r)}^{1/2} \cdot \norm{X-Y}_X.
\end{split}
\end{equation}
Note that the last step holds for all $X,Y\in U_2$. To simplify the notation, let $R = \norm{\sum_{i=1}^m \Ket{V_i}\Bra{V_i}}_{(r)}^{1/2}$, then we have $d_G(X,Y) \leq 2R\norm{X-Y}_X$.
This lets us upper-bound the covering numbers in $d_G$ with covering numbers in $\norm{\cdot}_X$:
\begin{equation}
N(U_2, d_G, \vareps) \leq N(U_2, \norm{\cdot}_X, \tfrac{\vareps}{2R}) = N(\tfrac{1}{\sqrt{r}}U_2, \norm{\cdot}_X, \tfrac{\vareps}{2R\sqrt{r}}).
\end{equation}
Plugging into (\ref{eqn-dudley}) and changing variables, we get
\begin{equation}
\label{eqn-coverint}
L_0 \leq 48 \sqrt{2\pi} R\sqrt{r} \int_0^\infty \log^{1/2} N(\tfrac{1}{\sqrt{r}}U_2, \norm{\cdot}_X, \vareps) d\vareps.
\end{equation}
We will now bound these covering numbers. First, we introduce some notation: let $\norm{\cdot}_p$ denote the Schatten $p$-norm on $\CC^{d\times d}$, and let $B_p$ be the unit ball in this norm. Also, let $B_X$ be the unit ball in the $\norm{\cdot}_X$ norm.
Observe that
\begin{equation}
\tfrac{1}{\sqrt{r}}U_2 \subseteq B_1 \subseteq K\cdot B_X.
\end{equation}
(The second inclusion follows because $\norm{M}_X \leq \max_{i=1,\ldots,m} \norm{V_i} \norm{M}_* \leq K \norm{M}_*$.) This gives a simple bound on the covering numbers:
\begin{equation}
N(\tfrac{1}{\sqrt{r}}U_2, \norm{\cdot}_X, \vareps) \leq N(B_1, \norm{\cdot}_X, \vareps)
\leq N(K\cdot B_X, \norm{\cdot}_X, \vareps).
\end{equation}
This equals 1 when $\vareps\geq K$. So, in equation (\ref{eqn-coverint}), we can restrict the integral to the interval $[0,K]$.
When $\vareps$ is small, we will use the following simple bound (equation (5.7) in \cite{Pisier}): (this is equation (\ref{eqn-coversmall}))
\begin{equation}
N(K\cdot B_X, \norm{\cdot}_X, \vareps) \leq (1+\tfrac{2K}{\vareps})^{2d^2}.
\end{equation}
When $\vareps$ is large, we will use a more sophisticated bound based on Maurey's empirical method and entropy duality, which is due to \cite{Guedon08} (see also \cite{Aubrun09}): (this is equation (\ref{eqn-coverlarge}))
\begin{equation}
N(B_1, \norm{\cdot}_X, \vareps) \leq \exp(\tfrac{C_1^2 K^2}{\vareps^2} \log^3 d \log m), \quad \text{for some constant $C_1$}.
\end{equation}
We defer the proof of (\ref{eqn-coverlarge}) to the next section. Here, we proceed to bound the integral in (\ref{eqn-coverint}).
Let $A = K/d$. For the integral over $[0,A]$, we write
\begin{equation}
\begin{split}
L_1 &:= \int_0^A \log^{1/2} N(\tfrac{1}{\sqrt{r}}U_2, \norm{\cdot}_X, \vareps) d\vareps
\leq \int_0^A \sqrt{2}d \log^{1/2} (1+\tfrac{2K}{\vareps}) d\vareps \\
&\leq \sqrt{2}d \int_0^A \bigl( 1+\log(1+\tfrac{2K}{\vareps}) \bigr) d\vareps
= \sqrt{2}d\cdot A + \sqrt{2}d\cdot L'_1
\end{split}
\end{equation}
where
\begin{equation}
\begin{split}
L'_1 &:= \int_0^A \log(1+\tfrac{2K}{\vareps}) d\vareps
= \int_{1/A}^\infty \log(1+2Ky) \tfrac{dy}{y^2} \\
&\leq \int_{1/A}^\infty \log((A+2K)y) \tfrac{dy}{y^2}
= \int_{1/A}^\infty \log(A+2K) \tfrac{dy}{y^2} + \int_{1/A}^\infty \log y \tfrac{dy}{y^2}.
\end{split}
\end{equation}
Integrating by parts, we get
\begin{equation}
L'_1 \leq A\log(A+2K) + A\log\tfrac{1}{A} + A = A\log(1+\tfrac{2K}{A}) + A,
\end{equation}
and substituting back in,
\begin{equation}
L_1 \leq \sqrt{2}dA (2+\log(1+\tfrac{2K}{A})) = \sqrt{2}K (2+\log(1+2d)).
\end{equation}
For the integral over $[A,K]$, we write
\begin{equation}
\begin{split}
L_2 &:= \int_A^K \log^{1/2} N(\tfrac{1}{\sqrt{r}}U_2, \norm{\cdot}_X, \vareps) d\vareps
\leq \int_A^K \tfrac{C_1 K}{\vareps} \log^{3/2} d \log^{1/2} m\; d\vareps \\
&= C_1 K \log^{3/2} d \log^{1/2} m \log \tfrac{K}{A}
= C_1 K \log^{5/2} d \log^{1/2} m.
\end{split}
\end{equation}
Finally, substituting into (\ref{eqn-coverint}), we get
\begin{equation}
L_0 \leq 48 \sqrt{2\pi} R\sqrt{r} (L_1+L_2)
\leq C_4 R\sqrt{r} K \log^{5/2} d \log^{1/2} m,
\end{equation}
where $C_4$ is some universal constant. This proves the lemma.
\subsection{Proof of Equation (\ref{eqn-coverlarge}) (covering numbers of the nuclear-norm ball)}
Our result will follow easily from a bound on covering numbers introduced in \cite{Guedon08} (where it appears as Lemma 1):
\begin{lem}
Let $E$ be a Banach space, having modulus of convexity of power type 2 with constant $\lambda(E)$. Let $E^*$ be the dual space, and let $T_2(E^*)$ denote its type 2 constant. Let $B_E$ denote the unit ball in $E$.
Let $V_1,\ldots,V_m \in E^*$, such that $\norm{V_j}_{E^*} \leq K$ (for all $j$). Define the norm on $E$,
\begin{equation}
\norm{M}_X = \max_{j=1,\ldots,m} |(V_j,M)|.
\end{equation}
Then, for any $\vareps>0$,
\begin{equation}
\vareps \log^{1/2} N(B_E,\norm{\cdot}_X,\vareps) \leq C_2 \lambda(E)^2 T_2(E^*) K \log^{1/2} m,
\end{equation}
where $C_2$ is some universal constant.
\end{lem}
The proof uses entropy duality to reduce the problem to bounding the ``dual'' covering number. The basic idea is as follows. Let $\ell_p^m$ denote the complex vector space $\CC^m$ with the $\ell_p$ norm. Consider the map $S:\; \ell_1^m \rightarrow E^*$ that takes the $j$'th coordinate vector to $V_j$. Let $N(S)$ denote the number of balls in $E^*$ needed to cover the image (under the map $S$) of the unit ball in $\ell_1^m$. We can bound $N(S)$ using Maurey's empirical method. Also define the dual map $S^*:\; E \rightarrow \ell_\infty^m$, and the associated dual covering number $N(S^*)$. Then $N(B_E,\norm{\cdot}_X,\vareps)$ is related to $N(S^*)$. Finally, $N(S)$ and $N(S^*)$ are related via entropy duality inequalities. See \cite{Guedon08} for details.
We will apply this lemma as follows, using the same approach as \cite{Aubrun09}. Let $S_p$ denote the Banach space consisting of all matrices in $\CC^{d\times d}$ with the Schatten $p$-norm. Intuitively, we want to set $E = S_1$ and $E^* = S_\infty$, but this won't work because $\lambda(S_1)$ is infinite. Instead, we let $E = S_p$, $p = (\log d)/(\log d-1)$, and $E^* = S_q$, $q = \log d$. Note that $\norm{M}_p \leq \norm{M}_*$, hence $B_1 \subseteq B_p$ and
\begin{equation}
\vareps \log^{1/2} N(B_1,\norm{\cdot}_X,\vareps)
\leq \vareps \log^{1/2} N(B_p,\norm{\cdot}_X,\vareps).
\end{equation}
Also, we have $\lambda(E) \leq 1/\sqrt{p-1} = \sqrt{\log d-1}$ and $T_2(E^*) \leq \lambda(E) \leq \sqrt{\log d-1}$ (see the Appendix in \cite{Aubrun09}). Note that $\norm{M}_q \leq e\norm{M}$, thus we have $\norm{V_j}_q \leq eK$ (for all $j$). Then, using the lemma, we have
\begin{equation}
\vareps \log^{1/2} N(B_p,\norm{\cdot}_X,\vareps) \leq C_2 \log^{3/2} d\; (eK) \log^{1/2} m,
\end{equation}
which proves the claim.
\section{Proof of Proposition \ref{prop-errorbound2} (recovery of a full-rank matrix)}
\label{sec4}
In this section we will sketch the proof of Proposition \ref{prop-errorbound2}. We use the same argument as Theorem 2.8 in \cite{CP09}, adapted for Pauli (rather than Gaussian) measurements.
A crucial ingredient is the NNQ (``nuclear norm quotient'') property of a sampling operator $\calA$, which was introduced in \cite{CP09} and is analogous to the LQ (``$\ell_1$-quotient'') property in compressed sensing \cite{Woj09}. We say that a sampling operator $\calA:\: \CC^{d\times d} \rightarrow \CC^m$ satisfies the NNQ($\alpha$) property if
\begin{equation}
\calA(B_1) \supseteq \alpha B_2,
\end{equation}
where $B_1$ is the unit ball of the nuclear norm in $\CC^{d\times d}$, and $B_2$ is the unit ball of the $\ell_2$ (Euclidean) norm in $\CC^m$.
It is easy to see that the Pauli sampling operator $\calA$ defined in (\ref{eqn-A}) satisfies NNQ($\alpha$) with $\alpha = \sqrt{d/m}$. (Without loss of generality, suppose that the Pauli matrices $S_1,\ldots,S_m$ used to construct $\calA$ are all distinct. Let $\alpha = \sqrt{d/m}$ and choose any $y \in \alpha B_2$. Let $X = \frac{\sqrt{m}}{d} \sum_{i=1}^m y_i S_i$, so we have $\calA(X) = y$. Observe that $\norm{X}_* \leq \sqrt{d} \norm{X}_F = \sqrt{\frac{m}{d}} \norm{y}_2 \leq 1$, as desired.) We remark that this value of $\alpha$ is probably not optimal; if one could prove that $\calA$ satisfies NNQ($\alpha$) with larger $\alpha$, it would improve the bound in Proposition \ref{prop-errorbound2}.
We will need one more property of $\calA$. We want the following to hold: for any fixed matrix $M\in\CC^{d\times d}$ (which is not necessarily low-rank), almost all random choices of $\calA$ will satisfy
\begin{equation}
\norm{\calA(M)}_2^2 \leq 1.5 \norm{M}_F^2.
\end{equation}
(Note that this inequality is required to hold only for this one particular matrix $M$.) In our case (random Pauli measurements), it is easy to check that $\calA$ obeys this property as well.
The proof of Theorem 2.8 in \cite{CP09} actually implies the following more general statement, about low-rank matrix recovery when $\calA$ satisfies both RIP and NNQ:
\begin{thm}
\label{thm-errorbound3}
Let $M$ be any matrix in $\CC^{d\times d}$, and let $\sigma_1(M) \geq \sigma_2(M) \geq \cdots \geq \sigma_d(M) \geq 0$ be its singular values. Write $M = M_r+M_c$, where $M_r$ contains the $r$ largest singular values of $M$. Also write $M = M_0+M_e$, where $M_0$ contains only those singular values of $M$ that exceed $\lambda=16\sqrt{d}\sigma$.
Suppose the sampling operator $\calA:\: \CC^{d\times d} \rightarrow \CC^m$ satisfies RIP (for rank-$r$ matrices in $\CC^{d\times d}$), and NNQ($\alpha$) with $\alpha = \mu\sqrt{d/m}$. Furthermore, suppose that $\calA$ satisfies $\norm{\calA(M_c)}_2^2 \leq 1.5 \norm{M_c}_F^2$ and $\norm{\calA(M_e)}_2^2 \leq 1.5 \norm{M_e}_F^2$.
Say we observe $y = \calA(M)+z$, where $z \sim N(0,\sigma^2 I)$. Let $\Mhat$ be the Dantzig selector (\ref{eqn-tracemin}) with $\lambda = 16\sqrt{d}\sigma$, or the Lasso (\ref{eqn-lasso}) with $\mu = 32\sqrt{d}\sigma$. Then, with high probability over the choice of $\calA$ and the noise $z$,
\begin{equation}
\label{eqn-errorbound3}
\norm{\Mhat-M}_F^2 \leq C_0 \sum_{i=1}^r \min(\sigma_i^2(M), d\sigma^2) + \Bigl(C_1 + \frac{C_2 m}{\mu^2 rd}\Bigr) \sum_{i=r+1}^d \sigma_i^2(M),
\end{equation}
where $C_0$, $C_1$ and $C_2$ are absolute constants.
\end{thm}
To prove Theorem \ref{thm-errorbound3}, one follows the proof of Theorem 2.8 in \cite{CP09}. There is a slight modification to Lemma 3.10 in \cite{CP09}: one gets the more general bound,
\begin{equation}
\norm{\Mhat-M}_F \leq C_0 \lambda\sqrt{r} + \bigl(C_1 + \tfrac{C_2}{\mu} \sqrt{\tfrac{m}{rd}}\bigr) \norm{\calA(M_c)}_2 + \norm{M_c}_F.
\end{equation}
Combining Theorem \ref{thm-errorbound3} with the preceding facts gives us Proposition \ref{prop-errorbound2}.
\section{Introduction}
Low-rank matrix recovery is the following problem: let $M$ be some unknown matrix of dimension $d$ and rank $r\ll d$, and let $A_1,A_2,\ldots,A_m$ be a set of measurement matrices; then can one reconstruct $M$ from its inner products $\Tr(M^* A_1), \Tr(M^* A_2), \ldots, \Tr(M^* A_m)$? This problem has many applications in machine learning \cite{Fazel02, Srebro04}, e.g., collaborative filtering (the Netflix problem). Remarkably, it turns out that for many useful choices of measurement matrices, low-rank matrix recovery is possible, and can even be done efficiently. For example, when the $A_i$ are Gaussian random matrices, then it is known that $m=O(rd)$ measurements are sufficient to uniquely determine $M$, and furthermore, $M$ can be reconstructed by solving a convex program (minimizing the nuclear norm) \cite{RFP08,FCRP08,CP09}. Another example is the ``matrix completion'' problem, where the measurements return a random subset of matrix elements of $M$; in this case, $m=O(rd\poly\log d)$ measurements suffice, provided that $M$ satisfies some ``incoherence'' conditions \cite{CandesRecht08, CandesTao09, Gross09, Recht09, NW10}.
The focus of this paper is on a different class of measurements, known as Pauli measurements. Here, the $A_i$ are randomly chosen elements of the Pauli basis, a particular orthonormal basis of $\CC^{d\times d}$. The Pauli basis is a non-commutative analogue of the Fourier basis in $\CC^d$; thus, low-rank matrix recovery using Pauli measurements can be viewed as a generalization of the idea of compressed sensing of sparse vectors using their Fourier coefficients \cite{CandesTao04, RV06}. In addition, this problem has applications in quantum state tomography, the task of learning an unknown quantum state by performing measurements \cite{Grossetal09}. This is because most quantum states of physical interest are accurately described by density matrices that have low rank; and Pauli measurements are especially easy to carry out in an experiment (due to the tensor product structure of the Pauli basis).
In this paper we show stronger results on low-rank matrix recovery from Pauli measurements. Previously \cite{Grossetal09, Gross09}, it was known that, for every rank-$r$ matrix $M\in\CC^{d\times d}$, almost all choices of $m=O(rd \poly\log d)$ random Pauli measurements will lead to successful recovery of $M$. Here we show a stronger statement: there is a fixed (``universal'') set of $m=O(rd \poly\log d)$ Pauli measurements, such that for all rank-$r$ matrices $M\in\CC^{d\times d}$, we have successful recovery.\footnote{Note that in the universal result, $m$ is slightly larger, by a factor of $\poly\log d$.} We do this by showing that the random Pauli sampling operator obeys the ``restricted isometry property'' (RIP). Intuitively, RIP says that the sampling operator is an approximate isometry, acting on the set of all low-rank matrices. In geometric terms, it says that the sampling operator embeds the manifold of low-rank matrices into $O(rd \poly\log d)$ dimensions, with low distortion in the 2-norm.
RIP for low-rank matrices is a very strong property, and prior to this work, it was only known to hold for very unstructured types of random measurements, such as Gaussian measurements \cite{RFP08}, which are unsuitable for most applications. RIP was known to fail in the matrix completion case, and whether it held for Pauli measurements was an open question. Once we have established RIP for Pauli measurements, we can use known results \cite{RFP08,FCRP08,CP09} to show low-rank matrix recovery from a universal set of Pauli measurements. In particular, using \cite{CP09}, we can get nearly-optimal universal bounds on the error of the reconstructed density matrix, when the data are noisy; and we can even get bounds on the recovery of arbitrary (not necessarily low-rank) matrices. These RIP-based bounds are qualitatively stronger than those obtained using ``dual certificates'' \cite{CandesPlan-old} (though the latter technique is applicable in some situations where RIP fails).
In the context of quantum state tomography, this implies that, given a quantum state that consists of a low-rank component $M_r$ plus a residual full-rank component $M_c$, we can reconstruct $M_r$ up to an error that is not much larger than $M_c$. In particular, let $\norm{\cdot}_*$ denote the nuclear norm, and let $\norm{\cdot}_F$ denote the Frobenius norm. Then the error can be bounded in the nuclear norm by $O(\norm{M_c}_*)$ (assuming noiseless data), and it can be bounded in the Frobenius norm by $O(\norm{M_c}_F \poly\log d)$ (which holds even with noisy data\footnote{However, this bound is not universal.}). This shows that our reconstruction is nearly as good as the best rank-$r$ approximation to $M$ (which is given by the truncated SVD). In addition, a completely arbitrary quantum state can be reconstructed up to an error of $O(1/\sqrt{r})$ in Frobenius norm.
Lastly, the RIP gives some insight into the optimal design of tomography experiments, in particular, the tradeoff between the number of measurement settings (which is essentially $m$), and the number of repetitions of the experiment at each setting (which determines the statistical noise that enters the data) \cite{tradeoff11}.
These results can be generalized beyond the class of Pauli measurements. Essentially, one can replace the Pauli basis with any orthonormal basis of $\CC^{d\times d}$ that is \textit{incoherent}, i.e., whose elements have small operator norm (of order $O(1/\sqrt{d})$, say); a similar generalization was noted in the earlier results of \cite{Gross09}. Also, our proof shows that the RIP actually holds in a slightly stronger sense: it holds not just for all rank-$r$ matrices, but for all matrices $X$ that satisfy $\norm{X}_* \leq \sqrt{r}\norm{X}_F$.
To prove this result, we combine a number of techniques that have appeared elsewhere. RIP results were previously known for Gaussian measurements and some of their close relatives \cite{RFP08}. Also, restricted strong convexity (RSC), a similar but somewhat weaker property, was recently shown in the context of the matrix completion problem (with additional ``non-spikiness'' conditions) \cite{NW10}. These results follow from covering arguments (i.e., using a concentration inequality to upper-bound the failure probability on each individual low-rank matrix $X$, and then taking the union bound over all such $X$). Showing RIP for Pauli measurements seems to be more delicate, however. Pauli measurements have more structure and less randomness, so the concentration of measure phenomena are weaker, and the union bound no longer gives the desired result.
Instead, one must take into account the favorable correlations between the behavior of the sampling operator on different matrices --- intuitively, if two low-rank matrices $M$ and $M'$ have overlapping supports, then good behavior on $M$ is positively correlated with good behavior on $M'$. This can be done by transforming the problem into a Gaussian process, and using Dudley's entropy bound. This is the same approach used in classical compressed sensing, to show RIP for Fourier measurements \cite{RV06, CandesTao04}. The key difference is that in our case, the Gaussian process is indexed by low-rank matrices, rather than sparse vectors. To bound the correlations in this process, one then needs to bound the covering numbers of the nuclear norm ball (of matrices), rather than the $\ell_1$ ball (of vectors). This requires a different technique, using entropy duality, which is due to Gu\'edon et al \cite{Guedon08}. (See also the related work in \cite{Aubrun09}.)
As a side note, we remark that matrix recovery can sometimes fail because there exist large sets of up to $d$ Pauli matrices that all commute, i.e., they have a simultaneous eigenbasis $\phi_1,\ldots,\phi_d$. (These $\phi_i$ are of interest in quantum information --- they are called stabilizer states \cite{NC}.) If one were to measure such a set of Pauli's, one would gain complete knowledge about the diagonal elements of the unknown matrix $M$ in the $\phi_i$ basis, but one would learn nothing about the off-diagonal elements. This is reminiscent of the difficulties that arise in matrix completion. However, in our case, these pathological cases turn out to be rare, since it is unlikely that a random subset of Pauli matrices will all commute.
Finally, we note that there is a large body of related work on estimating a low-rank matrix by solving a regularized convex program; see, e.g., \cite{RT09
KLT10}.
This paper is organized as follows. In section 2, we state our results precisely, and discuss some specific applications to quantum state tomography. In section 3 we prove the RIP for Pauli matrices, and in section 4 we discuss some directions for future work.
Some technical details appear in sections \ref{app3}
and \ref{sec4}.
\textbf{Notation:} For vectors, $\norm{\cdot}_2$ denotes the $\ell_2$ norm. For matrices, $\norm{\cdot}_p$ denotes the Schatten $p$-norm, $\norm{X}_p = (\sum_i \sigma_i(X)^p)^{1/p}$, where $\sigma_i(X)$ are the singular values of $X$. In particular, $\norm{\cdot}_* = \norm{\cdot}_1$ is the trace or nuclear norm, $\norm{\cdot}_F = \norm{\cdot}_2$ is the Frobenius norm, and $\norm{\cdot} = \norm{\cdot}_\infty$ is the operator norm. Finally, for matrices, $A^*$ is the adjoint of $A$, and $(\cdot,\cdot)$ is the Hilbert-Schmidt inner product, $(A,B) = \Tr(A^* B)$. Calligraphic letters denote superoperators acting on matrices. Also, $\Ket{A}\Bra{A}$ is the superoperator that maps every matrix $X\in\CC^{d\times d}$ to the matrix $A\Tr(A^*X)$.
\section{Our Results}
We will consider the following approach to low-rank matrix recovery. Let $M\in\CC^{d\times d}$ be an unknown matrix of rank at most $r$. Let $W_1,\ldots,W_{d^2}$ be an orthonormal basis for $\CC^{d\times d}$, with respect to the inner product $(A,B) = \Tr(A^* B)$. We choose $m$ basis elements, $S_1,\ldots,S_m$, iid uniformly at random from $\set{W_1,\ldots,W_{d^2}}$ (``sampling with replacement''). We then observe the coefficients $(S_i,M)$. From this data, we want to reconstruct $M$.
For this to be possible, the measurement matrices $W_i$ must be ``incoherent'' with respect to $M$. Roughly speaking, this means that the inner products $(W_i,M)$ must be small. Formally, we say that the basis $W_1,\ldots,W_{d^2}$ is \textit{incoherent} if the $W_i$ all have small operator norm,
\begin{equation}
\label{eqn-smallopnorm}
\norm{W_i} \leq K/\sqrt{d},
\end{equation}
where $K$ is a constant.\footnote{Note that $\norm{W_i}$ is the maximum inner product between $W_i$ and any rank-1 matrix $M$ (normalized so that $\norm{M}_F = 1$).} (This assumption was also used in \cite{Gross09}.)
Before proceeding further, let us sketch the connection between this problem and quantum state tomography. Consider a system of $n$ qubits, with Hilbert space dimension $d = 2^n$. We want to learn the state of the system, which is described by a density matrix $\rho\in\CC^{d\times d}$; $\rho$ is positive semidefinite, has trace 1, and has rank $r\ll d$ when the state is nearly pure. There is a class of convenient (and experimentally feasible) measurements, which are described by Pauli matrices (also called Pauli observables). These are matrices of the form $P_1\tns\cdots\tns P_n$, where $\tns$ denotes the tensor product (Kronecker product), and each $P_i$ is a $2\times 2$ matrix chosen from the following four possibilities:
\begin{equation}
I = \begin{pmatrix} 1&0\\ 0&1 \end{pmatrix}, \quad
\sigma_x = \begin{pmatrix} 0&1\\ 1&0 \end{pmatrix}, \quad
\sigma_y = \begin{pmatrix} 0&-i\\ i&0 \end{pmatrix}, \quad
\sigma_z = \begin{pmatrix} 1&0\\ 0&-1 \end{pmatrix}.
\end{equation}
One can estimate expectation values of Pauli observables, which are given by $(\rho,(P_1\tns\cdots\tns P_n))$. This is a special case of the above measurement model, where the measurement matrices $W_i$ are the (scaled) Pauli observables $(P_1\tns\cdots\tns P_n)/\sqrt{d}$, and they are incoherent with $\norm{W_i} \leq K/\sqrt{d}$, $K=1$.
Now we return to our discussion of the general problem. We choose $S_1,\ldots,S_m$ iid uniformly at random from $\set{W_1,\ldots,W_{d^2}}$, and we define the \textit{sampling operator} $\calA:\: \CC^{d\times d} \rightarrow \CC^m$ as
\begin{equation}
\label{eqn-A}
(\calA(X))_i = \tfrac{d}{\sqrt{m}} \Tr(S_i^* X), \quad i=1,\ldots,m.
\end{equation}
The normalization is chosen so that $\EE \calA^*\calA = \calI$. (Note that $\calA^*\calA = \sum_{j=1}^m \Ket{S_j}\Bra{S_j} \cdot \tfrac{d^2}{m}$.)
We assume we are given the data
$y = \calA(M)+z$,
where $z\in\CC^m$ is some (unknown) noise contribution. We will construct an estimator $\Mhat$ by minimizing the nuclear norm, subject to the constraints specified by $y$. (Note that one can view the nuclear norm as a convex relaxation of the rank function --- thus these estimators can be computed efficiently.) One approach is the matrix Dantzig selector:
\begin{equation}
\label{eqn-tracemin}
\Mhat = \arg\min_X \: \norm{X}_* \text{ such that } \norm{\calA^*(y-\calA(X))} \leq \lambda.
\end{equation}
Alternatively, one can solve a regularized least-squares problem, also called the matrix Lasso:
\begin{equation}
\label{eqn-lasso}
\Mhat = \arg\min_X \tfrac{1}{2} \norm{\calA(X)-y}_2^2 + \mu\norm{X}_*.
\end{equation}
Here, the parameters $\lambda$ and $\mu$ are set according to the strength of the noise component $z$ (we will discuss this later). We will be interested in bounding the error of these estimators. To do this, we will show that the sampling operator $\calA$ satisfies the restricted isometry property (RIP).
\subsection{RIP for Pauli Measurements}
Fix some constant $0 \leq \delta < 1$. Fix $d$, and some set $U \subset \CC^{d\times d}$. We say that $\calA$ satisfies the \textit{restricted isometry property} (RIP) over $U$ if, for all $X \in U$, we have
\begin{equation}
\label{eqn-RIP}
(1-\delta) \norm{X}_F \leq \norm{\calA(X)}_2 \leq (1+\delta) \norm{X}_F.
\end{equation}
(Here, $\norm{\calA(X)}_2$ denotes the $\ell_2$ norm of a vector, while $\norm{X}_F$ denotes the Frobenius norm of a matrix.) When $U$ is the set of all $X \in \CC^{d\times d}$ with rank $r$, this is precisely the notion of RIP studied in \cite{RFP08, CP09}. We will show that Pauli measurements satisfy the RIP over a slightly larger set (the set of all $X\in\CC^{d\times d}$ such that $\norm{X}_* \leq \sqrt{r} \norm{X}_F$), provided the number of measurements $m$ is at least $\Omega(rd\poly\log d)$. This result generalizes to measurements in any basis with small operator norm.
\begin{thm}
\label{thm-rip}
Fix some constant $0 \leq \delta < 1$. Let $\set{W_1,\ldots,W_{d^2}}$ be an orthonormal basis for $\CC^{d\times d}$ that is incoherent in the sense of (\ref{eqn-smallopnorm}). Let $m = CK^2\cdot rd \log^6 d$, for some constant $C$ that depends only on $\delta$, $C = O(1/\delta^2)$. Let $\calA$ be defined as in (\ref{eqn-A}). Then, with high probability (over the choice of $S_1,\ldots,S_m$), $\calA$ satisfies the RIP over the set of all $X\in\CC^{d\times d}$ such that $\norm{X}_* \leq \sqrt{r} \norm{X}_F$. Furthermore, the failure probability is exponentially small in $\delta^2 C$.
\end{thm}
We will prove this theorem in section 3. In the remainder of this section, we discuss its applications to low-rank matrix recovery, and quantum state tomography in particular.
\subsection{Applications}
By combining Theorem \ref{thm-rip} with previous results \cite{RFP08, FCRP08, CP09}, we immediately obtain bounds on the accuracy of the matrix Dantzig selector (\ref{eqn-tracemin}) and the matrix Lasso (\ref{eqn-lasso}). In particular, for the first time we can show \textit{universal} recovery of low-rank matrices via Pauli measurements, and near-optimal bounds on the accuracy of the reconstruction when the data is noisy \cite{CP09}. (Similar results hold for measurements in any incoherent operator basis.) These RIP-based results improve on the earlier results based on dual certificates \cite{Grossetal09, Gross09, CandesPlan-old}. See \cite{RFP08, FCRP08, CP09} for details.
Here, we will sketch a couple of these results that are of particular interest for quantum state tomography. Here, $M$ is the density matrix describing the state of a quantum mechanical object, and $\calA(M)$ is a vector of Pauli expectation values for the state $M$. ($M$ has some additional properties: it is positive semidefinite, and has trace 1; thus $\calA(M)$ is a real vector.) There are two main issues that arise. First, $M$ is not precisely low-rank. In many situations, the \textit{ideal} state has low rank (for instance, a pure state has rank 1); however, for the \textit{actual} state observed in an experiment, the density matrix $M$ is full-rank with decaying eigenvalues. Typically, we will be interested in obtaining a good low-rank approximation to $M$, ignoring the tail of the spectrum.
Secondly, the measurements of $\calA(M)$ are inherently noisy. We do not observe $\calA(M)$ directly; rather, we estimate each entry $(\calA(M))_i$ by preparing many copies of the state $M$, measuring the Pauli observable $S_i$ on each copy, and averaging the results. Thus, we observe $y_i = (\calA(M))_i + z_i$, where $z_i$ is binomially distributed. When the number of experiments being averaged is large, $z_i$ can be approximated by Gaussian noise. We will be interested in getting an estimate of $M$ that is stable with respect to this noise. (We remark that one can also reduce the statistical noise by performing more repetitions of each experiment. This suggests the possibility of a tradeoff between the accuracy of estimating each parameter, and the number of parameters one chooses to measure overall. This will be discussed elsewhere \cite{tradeoff11}.)
We would like to reconstruct $M$ up to a small error in the nuclear or Frobenius norm. Let $\Mhat$ be our estimate. Bounding the error in nuclear norm implies that, for any measurement allowed by quantum mechanics, the probability of distinguishing the state $\Mhat$ from $M$ is small. Bounding the error in Frobenius norm implies that the difference $\Mhat-M$ is highly ``mixed'' (and thus does not contribute to the coherent or ``quantum'' behavior of the system).
We now sketch a few results from \cite{FCRP08,CP09} that apply to this situation. Write $M = M_r+M_c$, where $M_r$ is a rank-$r$ approximation to $M$, corresponding to the $r$ largest singular values of $M$, and $M_c$ is the residual part of $M$ (the ``tail'' of $M$). Ideally, our goal is to estimate $M$ up to an error that is not much larger than $M_c$. First, we can bound the error in nuclear norm (assuming the data has no noise):
\begin{prop}
\label{prop-errorbound-nuc}
(Theorem 5 from \cite{FCRP08})
Let $\calA:\: \CC^{d\times d} \rightarrow \CC^m$ be the random Pauli sampling operator, with $m = C rd \log^6 d$, for some absolute constant $C$. Then, with high probability over the choice of $\calA$, the following holds:
Let $M$ be any matrix in $\CC^{d\times d}$, and write $M = M_r + M_c$, as described above.
Say we observe $y = \calA(M)$, with no noise. Let $\Mhat$ be the Dantzig selector (\ref{eqn-tracemin}) with $\lambda = 0$. Then
\begin{equation}
\label{eqn-errorbound-nuc}
\norm{\Mhat-M}_* \leq C'_0 \norm{M_c}_*,
\end{equation}
where $C'_0$ is an absolute constant.
\end{prop}
We can also bound the error in Frobenius norm, allowing for noisy data:
\begin{prop}
\label{prop-errorbound}
(Lemma 3.2 from \cite{CP09})
Assume the same set-up as above, but say we observe $y = \calA(M)+z$, where $z \sim N(0,\sigma^2 I)$. Let $\Mhat$ be the Dantzig selector (\ref{eqn-tracemin}) with $\lambda = 8\sqrt{d}\sigma$, or the Lasso (\ref{eqn-lasso}) with $\mu = 16\sqrt{d}\sigma$. Then, with high probability over the noise $z$,
\begin{equation}
\label{eqn-errorbound}
\norm{\Mhat-M}_F \leq C_0 \sqrt{rd} \sigma + C_1 \norm{M_c}_* / \sqrt{r},
\end{equation}
where $C_0$ and $C_1$ are absolute constants.
\end{prop}
This bounds the error of $\Mhat$ in terms of the noise strength $\sigma$ and the size of the tail $M_c$. It is universal: one sampling operator $\calA$ works for all matrices $M$. While this bound may seem unnatural because it mixes different norms, it can be quite useful.
When $M$ actually is low-rank (with rank $r$), then $M_c = 0$, and the bound (\ref{eqn-errorbound}) becomes particularly simple. The dependence on the noise strength $\sigma$ is known to be nearly minimax-optimal \cite{CP09}. Furthermore, when some of the singular values of $M$ fall below the ``noise level'' $\sqrt{d}\sigma$, one can show a tighter bound, with a nearly-optimal bias-variance tradeoff; see Theorem 2.7 in \cite{CP09} for details.
On the other hand, when $M$ is full-rank, then the error of $\Mhat$ depends on the behavior of the tail $M_c$. We will consider a couple of cases. First, suppose we do not assume anything about $M$, besides the fact that it is a density matrix for a quantum state. Then $\norm{M}_* = 1$, hence $\norm{M_c}_* \leq 1 - \tfrac{r}{d}$, and we can use (\ref{eqn-errorbound}) to get
$\norm{\Mhat-M}_F \leq C_0 \sqrt{rd} \sigma + \frac{C_1}{\sqrt{r}}$.
Thus, even for \textit{arbitrary} (not necessarily low-rank) quantum states, the estimator $\Mhat$ gives nontrivial results.
The $O(1/\sqrt{r})$ term can be interpreted as the penalty for only measuring an incomplete subset of the Pauli observables.
Finally, consider the case where $M$ is full-rank, but we do know that the tail $M_c$ is small. If we know that $M_c$ is small in nuclear norm, then we can use equation (\ref{eqn-errorbound}). However, if we know that $M_c$ is small in Frobenius norm, one can give a different bound, using ideas from \cite{CP09}, as follows.
\begin{prop}
\label{prop-errorbound2}
Let $M$ be any matrix in $\CC^{d\times d}$, with singular values $\sigma_1(M) \geq \cdots \geq \sigma_d(M)$.
Choose a random Pauli sampling operator $\calA:\: \CC^{d\times d} \rightarrow \CC^m$, with $m = C rd \log^6 d$, for some absolute constant $C$. Say we observe $y = \calA(M)+z$, where $z \sim N(0,\sigma^2 I)$. Let $\Mhat$ be the Dantzig selector (\ref{eqn-tracemin}) with $\lambda = 16\sqrt{d}\sigma$, or the Lasso (\ref{eqn-lasso}) with $\mu = 32\sqrt{d}\sigma$. Then, with high probability over the choice of $\calA$ and the noise $z$,
\begin{equation}
\label{eqn-errorbound2}
\norm{\Mhat-M}_F^2 \leq C_0 \sum_{i=1}^r \min(\sigma_i^2(M), d\sigma^2) + C_2(\log^6 d) \sum_{i=r+1}^d \sigma_i^2(M),
\end{equation}
where $C_0$ and $C_2$ are absolute constants.
\end{prop}
This bound can be interpreted as follows.
The first term expresses the bias-variance tradeoff for estimating $M_r$, while the second term depends on the Frobenius norm of $M_c$. (Note that the $\log^6 d$ factor may not be tight.)
In particular, this implies:
$\norm{\Mhat-M}_F \leq \sqrt{C_0} \sqrt{rd}\sigma + \sqrt{C_2}(\log^3 d) \norm{M_c}_F$.
This can be compared with equation (\ref{eqn-errorbound}) (involving $\norm{M_c}_*$).
This bound will be better when $\norm{M_c}_F \ll \norm{M_c}_*$, i.e., when the tail $M_c$ has slowly-decaying eigenvalues (in physical terms, it is highly mixed).
Proposition \ref{prop-errorbound2} is an adaptation of Theorem 2.8 in \cite{CP09}. We sketch the proof in section \ref{sec4}.
Note that this bound is not universal: it shows that for all matrices $M$, a random choice of the sampling operator $\calA$ is likely to work.
\section{Proof of the RIP for Pauli Measurements}
We now prove Theorem \ref{thm-rip}. The general approach involving Dudley's entropy bound is similar to \cite{RV06}, while the technical part of the proof (bounding certain covering numbers) uses ideas from \cite{Guedon08}. We summarize the argument here; the details are given in section \ref{app3}.
\subsection{Overview}
Let
$U_2 = \set{X\in\CC^{d\times d} \;|\; \norm{X}_F \leq 1, \; \norm{X}_* \leq \sqrt{r}\norm{X}_F}$.
Let $\calB$ be the set of all self-adjoint linear operators from $\CC^{d\times d}$ to $\CC^{d\times d}$, and define the following norm on $\calB$:
\begin{equation}
\norm{\calM}_{(r)} = \sup_{X\in U_2} |(X,\calM X)|.
\end{equation}
(Suppose $r\geq 2$, which is sufficient for our purposes. It is straightforward to show that $\norm{\cdot}_{(r)}$ is a norm, and that $\calB$ is a Banach space with respect to this norm.) Then let us define
\begin{equation}
\vareps_r(\calA) = \norm{\calA^*\calA-\calI}_{(r)}.
\end{equation}
By an elementary argument, in order to prove RIP, it suffices to show that $\vareps_r(\calA) < 2\delta - \delta^2$. We will proceed as follows: we will first bound $\EE \vareps_r(\calA)$, then show that $\vareps_r(\calA)$ is concentrated around its mean.
Using a standard symmetrization argument, we have that
$\EE \vareps_r(\calA) \leq 2\EE \Norm{\sum_{j=1}^m \vareps_j \Ket{S_j}\Bra{S_j} \tfrac{d^2}{m}}_{(r)}$,
where the $\vareps_j$ are Rademacher (iid $\pm 1$) random variables. Here the round ket notation $\Ket{S_j}$ means we view the matrix $S_j$ as an element of the vector space $\CC^{d^2}$ with Hilbert-Schmidt inner product; the round bra $\Bra{S_j}$ denotes the adjoint element in the (dual) vector space.
Now we use the following lemma, which we will prove later. This bounds the expected magnitude in $(r)$-norm of a Rademacher sum of a fixed collection of operators $V_1,\ldots,V_m$ that have small operator norm.
\begin{lem}
\label{lem-key}
Let $m\leq d^2$. Fix some $V_1,\ldots,V_m \in \CC^{d\times d}$ that have uniformly bounded operator norm, $\norm{V_i} \leq K$ (for all $i$). Let $\vareps_1,\ldots,\vareps_m$ be iid uniform $\pm 1$ random variables. Then
\begin{equation}
\EE_\vareps \Norm{\sum_{i=1}^m \vareps_i\Ket{V_i}\Bra{V_i}}_{(r)}
\leq C_5\cdot \Norm{\sum_{i=1}^m \Ket{V_i}\Bra{V_i}}_{(r)}^{1/2},
\end{equation}
where $C_5 = \sqrt{r}\cdot C_4 K \log^{5/2} d \log^{1/2} m$ and $C_4$ is some universal constant.
\end{lem}
After some algebra, one gets that
$\EE\vareps_r(\calA)
\leq 2 (\EE\vareps_r(\calA) + 1)^{1/2} \cdot C_5 \cdot \sqrt{\tfrac{d}{m}}$,
where $C_5 = \sqrt{r}\cdot C_4 K \log^3 d$. By finding the roots of this quadratic equation, we get the following bound on $\EE\vareps_r(\calA)$.
Let $\lambda \geq 1$. Assume that
$m \geq \lambda d (2C_5)^2 = \lambda \cdot 4C_4^2 \cdot dr \cdot K^2 \log^6 d$.
Then we have the desired result:
\begin{equation}
\EE\vareps_r(\calA) \leq \tfrac{1}{\lambda} + \tfrac{1}{\sqrt{\lambda}}.
\end{equation}
It remains to show that $\vareps_r(\calA)$ is concentrated around its expectation. For this we use a concentration inequality from \cite{LT91} for sums of independent symmetric random variables that take values in some Banach space. See section \ref{app3}
for details.
\subsection{Proof of Lemma \ref{lem-key} (bounding a Rademacher sum in $(r)$-norm)}
Let $L_0 = \EE_\vareps \norm{\sum_{i=1}^m \vareps_i\Ket{V_i}\Bra{V_i}}_{(r)}$; this is the quantity we want to bound. Using a standard comparison principle, we can replace the $\pm 1$ random variables $\vareps_i$ with iid $N(0,1)$ Gaussian random variables $g_i$; then we get
\begin{equation}
L_0 \leq \EE_g \sup_{X\in U_2} \sqrt{\tfrac{\pi}{2}} |G(X)|, \quad
G(X) = \sum_{i=1}^m g_i |(V_i,X)|^2.
\end{equation}
The random variables $G(X)$ (indexed by $X\in U_2$) form a Gaussian process, and $L_0$ is upper-bounded by the expected supremum of this process. Using the fact that $G(0)=0$ and $G(\cdot)$ is symmetric, and Dudley's inequality (Theorem 11.17 in \cite{LT91}), we have
\begin{equation}
L_0 \leq \sqrt{2\pi} \EE_g \sup_{X\in U_2} G(X)
\leq 24 \sqrt{2\pi} \int_0^\infty \log^{1/2} N(U_2,d_G,\vareps) d\vareps,
\end{equation}
where $N(U_2,d_G,\vareps)$ is a covering number (the number of balls in $\CC^{d\times d}$ of radius $\vareps$ in the metric $d_G$ that are needed to cover the set $U_2$), and the metric $d_G$ is given by
\begin{equation}
d_G(X,Y) = \Bigl( \EE[(G(X)-G(Y))^2] \Bigr)^{1/2}.
\end{equation}
Define a new norm (actually a semi-norm) $\norm{\cdot}_X$ on $\CC^{d\times d}$, as follows:
\begin{equation}
\norm{M}_X = \max_{i=1,\ldots,m} |(V_i,M)|.
\end{equation}
We use this to upper-bound the metric $d_G$. An elementary calculation shows that $d_G(X,Y) \leq 2R\norm{X-Y}_X$, where $R = \norm{\sum_{i=1}^m \Ket{V_i}\Bra{V_i}}_{(r)}^{1/2}$. This lets us upper-bound the covering numbers in $d_G$ with covering numbers in $\norm{\cdot}_X$:
\begin{equation}
N(U_2, d_G, \vareps) \leq N(U_2, \norm{\cdot}_X, \tfrac{\vareps}{2R}) = N(\tfrac{1}{\sqrt{r}}U_2, \norm{\cdot}_X, \tfrac{\vareps}{2R\sqrt{r}}).
\end{equation}
We will now bound these covering numbers. First, we introduce some notation: let $\norm{\cdot}_p$ denote the Schatten $p$-norm on $\CC^{d\times d}$, and let $B_p$ be the unit ball in this norm. Also, let $B_X$ be the unit ball in the $\norm{\cdot}_X$ norm.
Observe that
$\tfrac{1}{\sqrt{r}}U_2 \subseteq B_1 \subseteq K\cdot B_X$.
(The second inclusion follows because $\norm{M}_X \leq \max_{i=1,\ldots,m} \norm{V_i} \norm{M}_* \leq K \norm{M}_*$.) This gives a simple bound on the covering numbers:
\begin{equation}
N(\tfrac{1}{\sqrt{r}}U_2, \norm{\cdot}_X, \vareps) \leq N(B_1, \norm{\cdot}_X, \vareps)
\leq N(K\cdot B_X, \norm{\cdot}_X, \vareps).
\end{equation}
This is 1 when $\vareps\geq K$. So, in Dudley's inequality,
we can restrict the integral to the interval $[0,K]$.
When $\vareps$ is small, we will use the following simple bound (equation (5.7) in \cite{Pisier}):
\begin{equation}
\label{eqn-coversmall}
N(K\cdot B_X, \norm{\cdot}_X, \vareps) \leq (1+\tfrac{2K}{\vareps})^{2d^2}.
\end{equation}
When $\vareps$ is large, we will use a more sophisticated bound based on Maurey's empirical method and entropy duality, which is due to \cite{Guedon08} (see also \cite{Aubrun09}):
\begin{equation}
\label{eqn-coverlarge}
N(B_1, \norm{\cdot}_X, \vareps) \leq \exp(\tfrac{C_1^2 K^2}{\vareps^2} \log^3 d \log m), \quad \text{for some constant $C_1$}.
\end{equation}
We defer the proof of (\ref{eqn-coverlarge}) to the next section.
Using (\ref{eqn-coversmall}) and (\ref{eqn-coverlarge}), we can bound the integral in Dudley's inequality. We get
\begin{equation}
L_0 \leq C_4 R\sqrt{r} K \log^{5/2} d \log^{1/2} m,
\end{equation}
where $C_4$ is some universal constant. This proves the lemma.
\subsection{Proof of Equation (\ref{eqn-coverlarge}) (covering numbers of the nuclear-norm ball)}
Our result will follow easily from a bound on covering numbers introduced in \cite{Guedon08} (where it appears as Lemma 1):
\begin{lem}
Let $E$ be a Banach space, having modulus of convexity of power type 2 with constant $\lambda(E)$. Let $E^*$ be the dual space, and let $T_2(E^*)$ denote its type 2 constant. Let $B_E$ denote the unit ball in $E$.
Let $V_1,\ldots,V_m \in E^*$, such that $\norm{V_j}_{E^*} \leq K$ (for all $j$). Define the norm on $E$,
\begin{equation}
\norm{M}_X = \max_{j=1,\ldots,m} |(V_j,M)|.
\end{equation}
Then, for any $\vareps>0$,
\begin{equation}
\vareps \log^{1/2} N(B_E,\norm{\cdot}_X,\vareps) \leq C_2 \lambda(E)^2 T_2(E^*) K \log^{1/2} m,
\end{equation}
where $C_2$ is some universal constant.
\end{lem}
The proof uses entropy duality to reduce the problem to bounding the ``dual'' covering number. The basic idea is as follows. Let $\ell_p^m$ denote the complex vector space $\CC^m$ with the $\ell_p$ norm. Consider the map $S:\; \ell_1^m \rightarrow E^*$ that takes the $j$'th coordinate vector to $V_j$. Let $N(S)$ denote the number of balls in $E^*$ needed to cover the image (under the map $S$) of the unit ball in $\ell_1^m$. We can bound $N(S)$ using Maurey's empirical method. Also define the dual map $S^*:\; E \rightarrow \ell_\infty^m$, and the associated dual covering number $N(S^*)$. Then $N(B_E,\norm{\cdot}_X,\vareps)$ is related to $N(S^*)$. Finally, $N(S)$ and $N(S^*)$ are related via entropy duality inequalities. See \cite{Guedon08} for details.
We will apply this lemma as follows, using the same approach as \cite{Aubrun09}. Let $S_p$ denote the Banach space consisting of all matrices in $\CC^{d\times d}$ with the Schatten $p$-norm. Intuitively, we want to set $E = S_1$ and $E^* = S_\infty$, but this won't work because $\lambda(S_1)$ is infinite. Instead, we let $E = S_p$, $p = (\log d)/(\log d-1)$, and $E^* = S_q$, $q = \log d$. Note that $\norm{M}_p \leq \norm{M}_*$, hence $B_1 \subseteq B_p$ and
\begin{equation}
\vareps \log^{1/2} N(B_1,\norm{\cdot}_X,\vareps)
\leq \vareps \log^{1/2} N(B_p,\norm{\cdot}_X,\vareps).
\end{equation}
Also, we have $\lambda(E) \leq 1/\sqrt{p-1} = \sqrt{\log d-1}$ and $T_2(E^*) \leq \lambda(E) \leq \sqrt{\log d-1}$ (see the Appendix in \cite{Aubrun09}). Note that $\norm{M}_q \leq e\norm{M}$, thus we have $\norm{V_j}_q \leq eK$ (for all $j$). Then, using the lemma, we have
\begin{equation}
\vareps \log^{1/2} N(B_p,\norm{\cdot}_X,\vareps) \leq C_2 \log^{3/2} d\; (eK) \log^{1/2} m,
\end{equation}
which proves the claim.
\section{Outlook}
We have showed that random Pauli measurements obey the restricted isometry property (RIP), which implies strong error bounds for low-rank matrix recovery. The key technical tool was a bound on covering numbers of the nuclear norm ball, due to Gu\'edon et al \cite{Guedon08}.
An interesting question is whether this method can be applied to other problems, such as matrix completion, or constructing embeddings of low-dimensional manifolds into linear spaces with slightly higher dimension. For matrix completion, one can compare with the work of Negahban and Wainwright \cite{NW10}, where the sampling operator satisfies restricted strong convexity (RSC) over a certain set of ``non-spiky'' low-rank matrices. For manifold embeddings, one could try to generalize the results of \cite{KW11}, which use the sparse-vector RIP to construct Johnson-Lindenstrauss metric embeddings.
There are also many questions pertaining to low-rank quantum state tomography. For example, how does the matrix Lasso compare to the traditional approach using maximum likelihood estimation? Also, there are several variations on the basic tomography problem, and alternative notions of sparsity (e.g., elementwise sparsity in a known basis) \cite{Shabani10}, which have not been fully explored.
\textbf{Acknowledgements:}
Thanks to David Gross, Yaniv Plan, Emmanuel Cand\`es, Stephen Jordan, and the anonymous reviewers, for helpful suggestions. Parts of this work were done at the University of California, Berkeley, and supported by NIST grant number 60NANB10D262.
This paper is a contribution of the National Institute of Standards and Technology, and is not subject to U.S. copyright.
|
\section{Introduction}
In type theory, a coercion is a function $k$ of type $S \rightarrow T$
that can be automatically inserted by an interactive theorem prover to
promote a value $v$ of type $S$ to a value $k~v$ of type $T$ whenever
$v$ is used in a context where it is expected to have type $T$. The
typical example is the promotion of natural numbers to integers.
In a theory with dependent types coercions are better generalized to
couples $(k,n)$ where $k$ is a function of type $\Pi
\overrightarrow{x_i:S_i(x_1,\ldots,x_{i-1})}.T(x_1,\ldots,x_m)$ and $n \leq m$ is
the index of one of the arguments of $k$. The coercion is
automatically inserted to promote a value $v$ of type
$S(t_1,\ldots,t_{n-1})$ to a value
$k~t_1~\dots~t_{n-1}~v~t_{n+1}~\dots~t_m$ of type $T(t_1,\ldots,t_m)$.
The arguments $t_1,\ldots,t_{n-1},t_{n+1},\ldots,t_m$ are partially
inferred from the actual type of $v$ or from the expected type; those
that are not inferred become proof obligations for the user. The
typical example for the latter situation is the coercion from lists to
non-empty lists, that opens a new proof obligation for the
non-emptiness of the argument. This kind of parametric coercions have
been heavily exploited by Sozeau in the Russell language~\cite{russell-shortbib}.
All the previous examples of coercions are \emph{uniform} in the sense
that, up to the inferred arguments, the very same function $k$ is used
to promote the value $v$. For instance, if $k$ is the promotion from
natural numbers to integers, $3$ and $5$ are promoted respectively to
$k~3$ and $k~5$. Nevertheless there are situations where we would
like to promote $v$ in a non uniform way, depending on the actual
value of $v$, but the language does not allow one to inspect (i.e. pattern
match over) $v$, only the meta-level allows it.
For example, consider the promotion of a type (the carrier of a
semi-group) to a semi-group (the carrier enriched with the operation).
Different types are
associated (even not uniquely) to
different semi-groups on them: the type $\ensuremath{\mathbb{N}}$ of natural numbers could
be promoted to the semi-group $(\ensuremath{\mathbb{N}},+)$ whereas the type $\L\ensuremath{\mathbb{N}}$ could
be promoted to $(\L\ensuremath{\mathbb{N}},@)$ (where `$@$' is the append operation). Since
no function in type theory can distinguish between $\ensuremath{\mathbb{N}}$ and
$\L\ensuremath{\mathbb{N}}$, there is no uniform coercion of type $\ensuremath{\mathbf{Type}}\to\ensuremath{\mathtt{SemiGroup}}$
that behaves in the expected way.
As far as we know, nonuniform coercions have not been considered so
far in the literature. Nevertheless, they arise quite naturally when
devices like canonical structures~\cite{canonical-structures}, type
classes~\cite{SozeauO08-shortbib} or unification hints~\cite{unification-hints}
are employed.
In Sect.~\ref{syntax} we introduce a syntax and semantics
for nonuniform coercion
declarations and in Sect.~\ref{mathcomp} we present the use of
nonuniform coercions as a complementary device to canonical structures
like mechanisms. In Sect.~\ref{uhintsnut} we recall the syntax and semantics
of \emph{unification hints}~\cite{unification-hints}.
In Sect.~\ref{uhints} we show how nonuniform
coercions can be efficiently implemented at the user level in an
interactive theorem prover equipped with a flexible coercion system
and unification hints.
Sections~\ref{currymorph} and~\ref{notation} are devoted to the solution of
two different problems that naturally arise when nonuniform coercions
are employed. Conclusions and future works follow in
Sect.~\ref{conclusions}.
\section{Syntax and semantics}
\label{syntax}
Let $\Gamma$ be a well-typed context, i.e. a sequence of variable declarations
of the form $x_i: R_i$ where each $R_i$ is a well typed type in the context
$x_1:R_1 \dots x_{i-1}:R_{i-1}$.
A nonuniform coercion declaration has the following syntax:
$$
\begin{array}{ccc}
\begin{array}{lcl}
\Gamma_1 & \vdash &
\begin{array}{lcl}
S_1 & \rightarrow & T_1 \\
s_1 & \mapsto & t_1
\end{array}
\end{array} &~~~~~ \ldots ~~~~~ &
\begin{array}{lcl}
\Gamma_n & \vdash &
\begin{array}{lcl}
S_n & \rightarrow & T_n \\
s_n & \mapsto & t_n
\end{array}
\end{array}
\end{array}
$$
where each $s_i$ has type $S_i$ in $\Gamma_i$; each $t_i$ has type $T_i$ in
$\Gamma_i$.
Promotion works as follows: let $\bar{s}$ be a term of type $\bar{S}$ which is
expected to have type $\bar{T}$ and let $\sigma$ be a substitution that
instantiates the variables declared in $\Gamma_i$ and such that
$s_i \sigma = \bar{s}$ and $S_i \sigma = \bar{S}$ and $T_i \sigma = \bar{T}$.
Then $\bar{s}$ can be promoted to $t_i \sigma$ of type $\bar{T}$.
Operationally, the substitution $\sigma$ is determined by looking for the
smallest index $i$ such that $\sigma$ is the most general unifier of
$s_i$ with $\bar{s}$, of $S_i$ with $\bar{S}$ and of $T_i$ with $\bar{T}$.
The substitution $\sigma$ can be made total by instantiating every
still uninstantiated variable $x_j:R_j\sigma$ with the result of a proof obligation for $R_j\sigma$.
Instead of stopping at the first index that yields a result, it could
also be possible to try all of them and let the user interactively choose the
desired promotion. Stopping at the first index is often desirable since it
allows one to add overlapping branches ordering them by number of open proof
obligations (see Example~\ref{ex3} below).
\begin{example}[Uniform coercions]
A uniform coercion $(k,n)$ where
$$
k:\Pi \overrightarrow{x_i:S_i(x_1,\ldots,x_{i-1})}.T(x_1,\ldots,x_m)
$$
is equivalent to a nonuniform coercion
$$
\begin{array}{lcl}
\ldots x_i:S_i(x_1,\ldots,x_{i-1}) \ldots & \vdash &
\begin{array}{lcl}
S_n(x_1,\ldots,x_{n-1}) & \rightarrow & T(x_1,\ldots,x_m) \\
x_n & \mapsto & k~x_1~\dots~x_m
\end{array}
\end{array}
$$
Conversely, every branch of a nonuniform coercion whose pattern $s_i$
is a variable $x_n$ is equivalent to a nonuniform coercion:
$$
\begin{array}{lcl}
\ldots x_i:S_i(x_1,\ldots,x_{i-1}) \ldots & \vdash &
\begin{array}{lcl}
S_n(x_1,\ldots,x_{n-1}) & \rightarrow & T(x_1,\ldots,x_m) \\
x_n & \mapsto & t
\end{array}
\end{array}
$$
is equivalent to the uniform coercion $(\lambda \overrightarrow{x_i:S_i(x_1,\ldots,x_{i-1})}.t,n)$:
$$
\lambda \overrightarrow{x_i:S_i(x_1,\ldots,x_{i-1})}.t:\Pi \overrightarrow{x_i:S_i(x_1,\ldots,x_{i-1})}.T(x_1,\ldots,x_m)
$$
\end{example}
\vspace{0.3em}
\begin{example}[Structure enrichment] Let $\pi$ be a proof
of the associativity of $+$ over $\ensuremath{\mathbb{Z}}$ and $\bar \pi$ a proof of
the associativity of the append operation over lists.
$$
\begin{array}{lcl}
& \vdash &
\begin{array}{lcl}
\ensuremath{\mathbf{Type}} & \rightarrow & \ensuremath{\mathtt{SemiGroup}} \\
\ensuremath{\mathbb{Z}} & \mapsto & (\ensuremath{\mathbb{Z}},~+,~\pi)
\end{array} \\
\end{array}
\quad\qquad
\begin{array}{lcl}
X: \ensuremath{\mathbf{Type}} & \vdash &
\begin{array}{lcl}
\ensuremath{\mathbf{Type}} & \rightarrow & \ensuremath{\mathtt{SemiGroup}} \\
\L X & \mapsto & (\L X,~@~X,~\bar\pi~X)
\end{array}
\end{array}
$$
Note that the second coercion is polymorphic in the type
of the list arguments. Seen as two
independent uniform coercions, the two coercions are
incoherent~\cite{coercivesubtyping}
since they both promote from and to the same couple of types and they
are not \mbox{$\alpha\beta\eta$-convertible}.
\end{example}
\vspace{0.3em}
\begin{example}[Property enrichment]\label{ex3}
Let $\pi$ be a proof that $+$ is associative, $\ensuremath{\mathtt{AssocFun}}$ the structure
of associative functions and $\ensuremath{\mathtt{assoc\_comp}}$ a proof that the composition of
objects in $\ensuremath{\mathtt{AssocFun}}$ is in $\ensuremath{\mathtt{AssocFun}}$.
$$
\begin{array}{lcl}
& \vdash &
\begin{array}{lcl}
\ensuremath{\mathbb{N}} \to \ensuremath{\mathbb{N}} \to \ensuremath{\mathbb{N}} & \rightarrow & \ensuremath{\mathtt{AssocFun}}~\ensuremath{\mathbb{N}} \\
+ & \mapsto & (+,\pi)
\end{array}\\
\\
\begin{array}{l}
X: \ensuremath{\mathbf{Type}}, *: X \to X \to X,\\
p: \forall a,b,c:X.a * (b * c)=(a * b) * c~
\end{array}
& \vdash &
\begin{array}{lcl}
X \to X \to X & \rightarrow & \ensuremath{\mathtt{AssocFun}}~X \\
~\!\!* & \mapsto & (*,p)
\end{array}\\
\\
X: \ensuremath{\mathbf{Type}}, f,g: \ensuremath{\mathtt{AssocFun}}~X & \vdash &
\begin{array}{lcl}
X \to X \to X & \rightarrow & \ensuremath{\mathtt{AssocFun}}~X \\
f \circ g & \mapsto & \ensuremath{\mathtt{assoc\_comp}}~f~g
\end{array}
\end{array}
$$
When used to promote $+$ to an associative operation, no proof
obligation is left open. On the other hand, when used to promote $*:\ensuremath{\mathbb{N}}
\to \ensuremath{\mathbb{N}} \to \ensuremath{\mathbb{N}}$ to an associative operation, the user is left with the
proof obligation $p : \forall a,b,c:\ensuremath{\mathbb{N}}.a * (b * c)=(a * b) * c$.
Finally, the composition of two associative operations is promoted to
an associative operation applying the theorem $\ensuremath{\mathtt{assoc\_comp}}$. The third case
can not be expressed as a uniform coercion since the pattern is not a
variable.
\end{example}
\section{Nonuniform coercions for mathematical structures}
\label{mathcomp}
Different ITPs provide the user different devices to fill the gap between
the high level language used to reason about abstract and complex mathematical
theories and the drastically lower level foundational language understandable
and checkable by a computer.
Among the most widespread machineries to aid the user we mention
decision procedures; general purpose proof searching facilities, usually
by means of external tools; model checkers and SMT solvers. All these tools
usually fall in the category called, with some abuse, automation.
Even if these tools shown to be quite effective in many areas,
some radically different techniques were recently successfully
employed~\cite{BigOps,canonical-structures,kacoq,type-classes-coq}
to aid the user in systems based on higher order languages.
Canonical structures, type classes (and their generalization into
unification hints) allow the user to instrument the ITP linking
objects and properties by means of structures.
Structures pack together axioms that give rise to a theory and that
are used to quantify theorems
(i.e. once defined the structure $\ensuremath{\mathtt{Group}}$ packing together the
group axioms, theorems about groups have usually the shape
$\forall G:\ensuremath{\mathtt{Group}}, \ldots$).
The user is then asked to link models of these structures
to their characterizing structure, so that the ITP is able to exploit
such link when needed.
For example, once proved that integers $\mathbb{Z}$ together with
addition $+$, $0$ and the inverse $-$ form a group $\ensuremath{\mathcal{Z}}$, every
theorem that holds on groups can be used over expressions laying in the
signature $(\mathbb{Z},+,-,0)$, even if the group structure $\ensuremath{\mathcal{Z}}$ is not
mentioned in any way in the context of the current conjecture.
In addition to that, limited forms of Prolog like proof search allow the
ITP to infer derived structures given the basic ones.
For example, the pair $(0,0)$ can be transparently considered as the unit
of the Cartesian product group $\ensuremath{\mathcal{Z}}\times\ensuremath{\mathcal{Z}}$, given
the general result that the Cartesian product of groups forms a group.
These forms of inference, although not as general
as other automatic devices, tend to be more predictable
and allow the user's formalization and proof strategy to be
closer to the widespread and natural mathematical argument:
\emph{since X together with Y form a Z, we have P(X,Y)}.
Moreover they shown to be very effective in building
reusable libraries of formalized
mathematics~\cite{canonical-structures}.
\subsection{Mathematical structures in dependently typed languages}
We briefly review here how the aforementioned devices are
implemented in an interactive theorem prover based on an higher order
and dependently typed language.
Suppose that the system library contains the definition of natural numbers
$\ensuremath{\mathbb{N}}$, of multiplication over them and the proof that multiplication
is associative, commutative and it has a neutral element. Suppose also that
the library contains the definition of unital semigroups (i.e. semigroups
that have a left neutral element), represented as
a dependently typed record~\cite{pollackFAC02,coercions-record-shortbib}:
$$
\begin{array}{l}
\mathtt{UnitalSemiGroup} := \\
\quad\{ S: \ensuremath{\mathbf{Type}};~1:S;~*:S \to S \to S;~
\pi: \mathtt{is\_unital\_semi\_group~S~1~*}\}
\end{array}
$$
The record projection $S$ is declared as a uniform coercion of type
$$
\mathtt{UnitalSemiGroup} \to \ensuremath{\mathbf{Type}}
$$
so that a quantification over a unital semigroup $G$
is automatically promoted to a quantification over the
group carrier $(S~G)$.
The library contains all the interesting theorems of the theory of unital
semigroups, stated by making the quantifiers range over dependently typed
records. One example is the unicity of the neutral element, if it exists:
$$\forall G: \mathtt{UnitalSemiGroup}.\forall x:G.
(\forall y:G. y*x=y) \Rightarrow x=1
$$
Suppose now that the user, during a proof over natural numbers, knows the
hypothesis $\forall y:\ensuremath{\mathbb{N}}. y+x=y$ (let's call it $H$) and she needs to
conclude that
$x=0$. In order to do that, she wants to apply the unicity property of the
neutral element of a unital semigroup, that is she wants to feed the
hypothesis $\forall y:\ensuremath{\mathbb{N}}. y+x=y$ to the theorem $L$ that proves the
unicity and that expects a unital semigroup $G$, an $x$ in the carrier of
$G$ and a proof that $\forall y:G. y+x=y$. The action of invoking the lemma $L$
generates the unification problem
$$\forall y:G. y*x=y \ensuremath{\stackrel{\scriptscriptstyle ?}{\equiv}} \forall y:\ensuremath{\mathbb{N}}. y+x=y$$
that can be solved by choosing for $G$ any unital semigroup over the natural
numbers whose operation is addition. The system automatically picks the
right group
definition if the user has already instructed the system
(by means of canonical structures or with the more flexible
mechanism of unification hints) to always enrich $\ensuremath{\mathbb{N}}$ to
$(\ensuremath{\mathbb{N}},0,+,\ldots)$ under the previous constraints.
More formally, the user has fed the system with the request to apply the proof
term $L~?_G~x~H$ where $?_G$ is a metavariable to be instantiated by unification
and the unifier instantiates $?_G$ with $(\ensuremath{\mathbb{N}},0,+,\ldots)$. Note that no
coercion has been applied in this case: $?_G$ is identified by the system
since it occurs in the type of $x$ and $H$, which are known.
Most of the time, the previous mechanism is sufficient to perform the
enrichment. Nevertheless, there are situations where the enriched structure
to be discovered does not occur dependently in the type of some other parameter
(or in the expected conclusion).
For instance, suppose that we already know that the exponential function is
injective and suppose that we want to inhabit the type
$(\mathtt{InjectiveFun}~\mathbb{R}~\mathbb{R})$ of injective functions over
real numbers with the function $\lambda x.e^{e^x}$. Since the previous
$\lambda$-abstraction is just a type theoretic function from $\mathbb{R}$ to
$\mathbb{R}$, what we need here is an automatic promotion of the function
from $(\mathbb{R} \to \mathbb{R})$ to
$(\mathtt{InjectiveFun}~\mathbb{R}~\mathbb{R})$, which can be achieved
by means of a nonuniform coercion similar to the one in Example~\ref{ex3}
that states that the composition of injective functions is injective.
Without nonuniform coercions, the user is obliged to manually feed the
system with the enriched structure itself, as in a traditional procedural
language.
This is, for instance, what happens currently in~\cite{canonical-structures},
where two distinct mechanisms (and notations) are used to enrich structure
when the canonical structure mechanism is not triggered.
The simplest one overloads every notation defining a scope for
every possible enrichment. Then, the explicit scope delimiter can be used
to interpret the notation as desired by the user. For example the
construction $A\cap B$,
even if $A$ and $B$ are groups, returns a set, while $(A\cap B)\%G$
returns a group.
This mechanism does not only force to redefine every notation for every
structure, but has also some technical limitations that make it fail if
one among $A$ and $B$ is not explicitly typed as a
group.\footnote{This
phenomenon happens, for example, when $B$ is a local definition for $C\cap D$.
Unfolding $B$ and then forcing the groups scope would make all instances of
$\cap$ to be interpreted at the group level, but if $B$ is not manually
unfolded, only one occurrence of $\cap$ is affected by the scope delimiter,
thus the full expression fails to be enriched.}
The second mechanism is way more verbose,
but does not suffer from the aforementioned limitations, and is thus
used as a fall back for the former. To
enrich G to the wanted $\mathtt{structure}$ the user types
``[the $\mathtt{structure}$ of G]'' and a rather complex machinery
generates behind the scenes an ad-hoc unification problem that triggers
canonical structures inference.\\
Note that in both approaches what the users types is what will then be
displayed, thus the more verbose an enrichment notation is, the more
cluttered the lemmas statements and the intermediate goals of their proofs
will be.
In the next section we see that,
quite surprisingly, nonuniform coercions can be implemented for free at the
user level in a system based on dependent types and unification hints.
\section{Unification hints in a nutshell}
\label{uhintsnut}
Unification hints~\cite{unification-hints} give solutions to (higher order)
unification problems that fall outside the domain of the regular unification
heuristic implemented by the ITP. They are presented as rules of the following
form:
\begin{prooftree}
\AxiomC{$\overrightarrow{?_x}\;:=\;\overrightarrow{H}$}
\LeftLabel{$\Gamma \vdash$}
\RightLabel{myhint}
\UnaryInfC{ $ P\;\equiv\;Q$ }
\end{prooftree}
where:
\begin{enumerate}
\item $\Gamma$ is a context that declares meta-variables $?_y$;
\item $\overrightarrow{?_x}\;:=\;\overrightarrow{H}$ is a telescope of definitions for
metavariables undeclared in $\Gamma$;
\item all meta-variables in $P$ and $Q$ are declared in $\Gamma$ or
defined in the telescope;
\item $P \equiv Q$ is a {\em linear} pattern in the meta-variables
defined in the telescope (i.e. every
meta-variable $?_x$ occurs just once either in $P$ or in $Q$);
\item the type checking rules for meta-variables mimick the corresponding
rules for variables; all the terms in $\Gamma$,
$\overrightarrow{H}$, $P$ and $Q$ must be well typed. We use meta-variables since
we expect them to be instantiated by unification\footnote{
Due to type dependencies, it is not always possible to give all the
meta-variable declarations first (in $\Gamma$) and then all the meta-variable
definitions. Interleaving declarations and definitions
poses no problem (and we implicitly assume that it is done). Nevertheless,
we prefer to present the hints in this way for clarity purposes.
}.
\end{enumerate}
A hint is {\em acceptable} if
$\Gamma \vdash P[\overrightarrow{H}/\overrightarrow{?_x}] \equiv Q[\overrightarrow{H}/\overrightarrow{?_x}]$,
i.e. if the two terms obtained by telescopic substitution,
are {\em convertible} in $\Gamma$.
Since convertibility
is (typically) a decidable relation, the system is able to discriminate
acceptable hints. As a consequence, for every substitution $\sigma$ that
instantiates only meta-variables in $\Gamma$, all unification problems of the
form $P \sigma \ensuremath{\stackrel{\scriptscriptstyle ?}{\equiv}} Q \sigma$ can be solved by the unifier $\rho$ that assigns
to each $?_x$ the term $H\sigma$. As a generalization,
if $\tau$ is a generic meta-variable substitution,
all unification problems of the form $P \tau \ensuremath{\stackrel{\scriptscriptstyle ?}{\equiv}} Q \tau$
can be solved by recursively solving the new problems
$\overrightarrow{?_x \tau \ensuremath{\stackrel{\scriptscriptstyle ?}{\equiv}} H \tau }$ (whose solution is $\rho$ in the
simple case where $\tau$ behaves as the identity function on the meta-variables
in the telescope).
Unification hints are triggered in case of a unification failure. If the
failing problem is $\bar P \ensuremath{\stackrel{\scriptscriptstyle ?}{\equiv}} \bar Q$, the system looks for an
hint such that there exists a unifier $\tau$ of $P$ with $\bar P$ and
$Q$ with $\bar Q$, and then try to solve the unification problem by triggering
the recursive problems $\overrightarrow{?_x \tau \ensuremath{\stackrel{\scriptscriptstyle ?}{\equiv}} H \tau }$.
When there are multiple hints matching the failing unification problem, the
system can either ask the user about the desired one, or it can
pick the first matching hint according to some user defined precedence
level (e.g. in order of declaration or by more precise matching).
We give a bird's-eye view on how they allow one to apply a simple property of
groups in a context where the group structure is implicit. Consider the
following statement and its corresponding form without (part of its) notational
sugar.
$$
a + 0 = a \qquad \ensuremath{\mathtt{zplus}}~a~0 = a
$$
\noindent
where $a$ is a point in $\ensuremath{\mathbb{Z}}$. The group theoretical property
we want to apply is the right-identity law for $+$ and $0$ that
is
$$
\mathtt{grid} : \forall G:\ensuremath{\mathtt{Group}},
\forall x: \ensuremath{\mathtt{carr}}~G, \ensuremath{\mathtt{op}}~G~x~(\ensuremath{\mathtt{unit}}~G) = x
$$
Instantiating this axiom to $a$ and leaving $G$ implicit (i.e. considering the
term $\mathtt{grid}~?_G~a$) triggers the unification problem
$\ensuremath{\mathtt{carr}}~?_G \ensuremath{\stackrel{\scriptscriptstyle ?}{\equiv}} \ensuremath{\mathbb{Z}}$, since $a:\ensuremath{\mathbb{Z}}$ is expected to have type
$\ensuremath{\mathtt{carr}}~?_G$. This problem can be solved only
by guessing a model of a group whose carrier is the set $\ensuremath{\mathbb{Z}}$.
Similarly, if we use $\mathtt{grid}$ to perform the rewriting without
instantiating it first (i.e. we use the term $\mathtt{grid}~?_G~?_x$ with
an expected type whose left hand side is $\ensuremath{\mathtt{zplus}}~a~0$), we trigger
the unification problem:
$$\ensuremath{\mathtt{op}}~?_G~?_x~(\ensuremath{\mathtt{unit}}~?_G) \ensuremath{\stackrel{\scriptscriptstyle ?}{\equiv}} \ensuremath{\mathtt{zplus}}~a~0$$
If the unification goes from left to right, the system must unify
$\ensuremath{\mathtt{op}}~?_G$ with $\ensuremath{\mathtt{zplus}}$; if it goes from right to left, it must
unify $\ensuremath{\mathtt{unit}}~?_G$ with $0$. In both cases we are facing again an
unification problem where a projection ($\ensuremath{\mathtt{op}},\ensuremath{\mathtt{unit}}$ or $\ensuremath{\mathtt{carr}}$) is applied to an implicit structure $?_G$ and a model of the structure must
be guessed by the system. Since any guess would be arbitrary and could involve
proof search, we expect the system to fail.
To force a solution to the unification problem, the user can
specify the following hints that link the carrier $\ensuremath{\mathbb{Z}}$, the operation
$\ensuremath{\mathtt{zplus}}$ and the constant $0$ to the group structure $\ensuremath{\mathcal{Z}}$.
\begin{center}
\AxiomC{$?_g := \ensuremath{\mathcal{Z}} $}
\LeftLabel{$\vdash$}
\UnaryInfC{ $ \ensuremath{\mathtt{carr}}~?_g \equiv \ensuremath{\mathbb{Z}}$ }
\DisplayProof
$\quad$
\AxiomC{$?_g := \ensuremath{\mathcal{Z}} $}
\LeftLabel{$\vdash$}
\UnaryInfC{ $ \ensuremath{\mathtt{op}}~?_g \equiv \ensuremath{\mathtt{zplus}}$ }
\DisplayProof
$\quad$
\AxiomC{$?_g := \ensuremath{\mathcal{Z}} $}
\LeftLabel{$\vdash$}
\UnaryInfC{ $ \ensuremath{\mathtt{unit}}~?_g \equiv 0$ }
\DisplayProof
\end{center}
Unification hints can also be used to drive proof search with clauses that
resemble the ones used in logic programming. For instance, to solve the
problem
$$
\ensuremath{\mathtt{carr}}~?_1 \ensuremath{\stackrel{\scriptscriptstyle ?}{\equiv}} \ensuremath{\mathbb{Z}} \times \ensuremath{\mathbb{Z}}
$$
the user can declare the following hint that recursively reduces the problem
to simpler ones:
\begin{prooftree}
\AxiomC{$?_A := \ensuremath{\mathtt{carr}}~?_h$}
\AxiomC{$?_B := \ensuremath{\mathtt{carr}}~?_q$}
\AxiomC{$?_g := ?_h \times ?_q$}
\LeftLabel{$?_h,?_q: \ensuremath{\mathtt{Group}} \vdash$}
\TrinaryInfC{
$ \ensuremath{\mathtt{carr}}~?_g \equiv ?_A \times ?_B$ }
\end{prooftree}
Intuitively, the hint says that, if
the carrier of a group $?_g$ is a product $?_A \times ?_B $,
where $?_A$ is the carrier of a group $?_h$ and $?_B$ is the carrier
of a group $?_q$ then a solution consists in choosing for $?_g$
the group product of $?_h$ and $?_q$.
\subsection{Hints and coercions indexing}
Another view at unification hints is that they define equivalence classes of
convertible terms that are considered indistinguishable by every functionality
of the system that works up to unification. Promotion of terms via (uniform)
coercions is the typical example: if the user declares a coercion from
$\ensuremath{\mathbb{Z}}$ to $\ensuremath{\mathbb{Q}}$, we expect the system to be able to promote also
$\ensuremath{\mathtt{carr}}~\ensuremath{\mathcal{Z}}$ to $\ensuremath{\mathbb{Q}}$.
This behaviour, however, does not come for free. Consider a failing unification
problem $\bar P \ensuremath{\stackrel{\scriptscriptstyle ?}{\equiv}} \bar Q$. In order to retrieve a coercion from $\bar P$
to $\bar Q$, it
would be inefficient to iterate over the whole set of coercions to find the
ones that goes from $M$ to $N$ such that $M$ unifies with $\bar P$ and $N$
unifies with $\bar Q$. The usual implementation strategy is to use a
discrimination tree~\cite{McCune,nieuwenhuis01evaluation-shortbib}
(or discrimination nets) to index the coercion and perform a quick
approximated search. These data structures, born in the field of
automatic theorem proving, are first order oriented and compare terms
according to their rigid structure. For example
$(\mbox{carr }\ensuremath{\mathcal{Z}})$ and $\ensuremath{\mathbb{Z}}$ would be considered different for at
least two reasons: both their head constant and head function symbol arity
differ.
Thus, in order to retrieve coercions up to hints, the
system must do some additional work, for instance to index every coercion
multiple times (on all $M'$ and $N'$ such that $M$ and $M'$ are in the same
hint-induced equivalence class, and the same for $N$ and $N'$), or to perform
the search multiple times (for every $P'$ and $Q'$ such that $\bar P$ and
$P'$ are in the same hint-induced equivalence class, and the same for
$\bar Q$ and $Q'$).
In the rest of the paper we assume our system to implement both unification
hints and retrieval of coercions up to unification hints. The
Matita{} interactive
theorem prover~\cite{ck-sadhana-shortbib,matita-jar-uitp-shortbib} satisfies these requirements.
\section{Non Uniform Coercions via Unification Hints}
\label{uhints}
We analyze at first a particular scenario where it is easier to explain the
ideas involved.
Suppose that we are interested in implementing the following nonuniform
coercion with two branches:
$$
\begin{array}{ccc}
\begin{array}{lcl}
& \vdash &
\begin{array}{lcl}
S & \rightarrow & T \\
s_1 & \mapsto & t_1
\end{array}
\end{array} & ~~~ &
\begin{array}{lcl}
& \vdash &
\begin{array}{lcl}
S & \rightarrow & T \\
s_2 & \mapsto & t_2
\end{array}
\end{array}
\end{array}
$$
The simplified scenario is characterized by the fact that all branches are
uniform in the types $S$ and $T$; the fact that we consider just two branches
and that we analyze only the case where all contexts $\Gamma_i$ are empty
is just to simplify the presentation and dropping these two limitations poses
no real problem.
Suppose also, to grant the coherence of the coercion graph,
that the system only allows one
to declare a single uniform coercion between two given
(equivalence classes of) types. The only way in which a single uniform coercion
can present the expected non uniform behaviour is that it takes $t_i$ in input:
$$
\begin{array}{l}
k : \forall S:\ensuremath{\mathbf{Type}}.S \rightarrow \forall T:\ensuremath{\mathbf{Type}}. T \to T\\
k~S~s~T~t := t
\end{array}
$$
so that $s_1$ can be promoted to $k~S~s_1~T~t_1$ and $s_2$ can be promoted to
$k~S~s_2~T~t_2$. However, when a term $\bar s$ of type $S$ is expected to have
type $T$, according to the standard semantics of uniform coercions, the user
will be presented with a proof obligation of type $T$ and the term $\bar s$ will
be promoted to $k~S~s~T~t$ (that reduces to $t$) where $t$ is the proof term
provided by the user in the proof obligation. This means that the coercion
maps any term of type $S$ to a term of type $T$ after asking the user what term
must be chosen. Instead, we would like the system to automatically pick $t_1$ for $t$
when $\bar s$ is $s_1$, to pick $t_2$ for $t$ when $\bar s$ is $s_2$ and to
fail in all the other cases, without bothering the user at all.
The idea to achieve the expected result is to use unification hints to
automatically suggest the term $t$. However, unification hints are only
triggered in case of a unification failure and apparently the only
failing unification problem is the initial one: $S \ensuremath{\stackrel{\scriptscriptstyle ?}{\equiv}} T$ where $s$ and $t$
do not occur at all.
\subsection{Lifting terms to types}
In a dependently typed language, terms can occur in types, or better
can be lifted to types when they occur as arguments of a dependently
typed function.
The trick we employ is to declare the nonuniform coercion not from $S$ to $T$,
but from $S$ to a type that is convertible to $T$ but that exposes $s$ and $t$.
We can make a first shot with the following definitions:
$$
\begin{array}{l}
\ensuremath{\mathtt{force}} :
\forall S:\ensuremath{\mathbf{Type}}. S \rightarrow \forall T:\ensuremath{\mathbf{Type}}. T \rightarrow \ensuremath{\mathbf{Type}}\\
\ensuremath{\mathtt{force}}~S~s~T~t := T
\end{array}
$$
\noindent
The $\ensuremath{\mathtt{force}}$ type has the property that
$\ensuremath{\mathtt{force}}~S~s~T~t$ is $\beta$-equivalent to $T$. The coercion $k$ is
now redefined as follows:
$$
\begin{array}{l}
k : \forall S:\ensuremath{\mathbf{Type}}.\forall s:S. \forall T:\ensuremath{\mathbf{Type}}.\forall t:T.
\ensuremath{\mathtt{force}}~S~s~T~t\\
k~S~s~T~t := t
\end{array}
$$
With these definitions, when a term $\bar s$ of type $S$ is used with type $T$,
the system tries to promote $\bar s$ to $k~S~\bar s~?_T~?_t$ (which reduces to $?_t$)
yielding a new unification problem
$$\ensuremath{\mathtt{force}}~S~\bar s~?_T~?_t \ensuremath{\stackrel{\scriptscriptstyle ?}{\equiv}} T$$
In this unification problem we have both $\bar s$ and $?_t$, and so we
have the data for choosing a term for $?_t$ in function of $\bar s$.
Moreover, since $?_t$ is also the argument the coercion $k$ is returning,
the solution for the unification problem can define the output of the
coercion.
However,
we also have that the left hand side reduces to $?_T$ and thus the system
can easily solve the unification problem by choosing $T$ for $?_T$, without
using any unification hint at all and, once again, opening a proof obligation
of type $T$ to determine the unconstrained term $?_t$.
\subsection{Locked reduction and lockpicking via hints}
In order to solve the problem, we need to change our definition of $k$
(and $\ensuremath{\mathtt{force}}$) again in order to produce in a similar way a
new unification problem whose solution is too difficult to be found
by the system without resorting to unification hints.
In other words, we need to change the definition of $\ensuremath{\mathtt{force}}$ in such
a way that $(\ensuremath{\mathtt{force}}~S~\bar s~?_T~?_t)$ is no longer always reducible to
$?_T$, but only in certain user-defined situations.
The latter behaviour can be achieved by adding an additional parameter to
$\ensuremath{\mathtt{force}}$ that must take a precise value to unlock the good reduction.
This can be achieved in many ways, one of them being to parameterize
$\ensuremath{\mathtt{force}}$ over an inhabitant of the unit type and using pattern matching
on it to block reduction until the canonical inhabitant is passed:
$$
\begin{array}{l}
\ensuremath{\mathtt{force}} :
\forall S:\ensuremath{\mathbf{Type}}. S \rightarrow \forall T:\ensuremath{\mathbf{Type}}. T \rightarrow \ensuremath{\mathbf{1}} \rightarrow \ensuremath{\mathbf{Type}}\\
\ensuremath{\mathtt{force}}~S~s~T~t~lock := \ensuremath{\mathtt{match~}} lock \ensuremath{\mathtt{~with~[~}} \star \Rightarrow T \ensuremath{\mathtt{~]}}
\end{array}
$$
\noindent
The $\ensuremath{\mathtt{force}}$ type has the property that
$(\ensuremath{\mathtt{force}}~S~s~T~t~\star)$ is $\beta$-equivalent to $T$ (where $\star$ is
the canonical inhabitant of $\ensuremath{\mathbf{1}}$), but
reduction of $(\ensuremath{\mathtt{force}}~S~s~T~t~?_l)$ is blocked.
The coercion $k$ is now redefined as follows:
$$
\begin{array}{l}
k : \forall S:\ensuremath{\mathbf{Type}}.\forall s:S. \forall T:\ensuremath{\mathbf{Type}}.\forall t:T.
\ensuremath{\mathbf{1}} \rightarrow \ensuremath{\mathtt{force}}~S~s~T~t\\
k~S~s~T~t~lock := \ensuremath{\mathtt{match~}} lock \ensuremath{\mathtt{~with~[~}} \star \Rightarrow t \ensuremath{\mathtt{~]}}
\end{array}
$$
The system now tries to promote $\bar s$ of type $S$ to
$(k~S~s~?_T~?_t~?_l)$ that generates the new failing unification problem
$$\ensuremath{\mathtt{force}}~S~\bar s~?_T~?_t~?_l \ensuremath{\stackrel{\scriptscriptstyle ?}{\equiv}} T$$
At last, this is where the unification hints come into play. Indeed, we can
define the following two acceptable unification hints:
\begin{center}
\AxiomC{ $ ?_T := T$}
\noLine
\UnaryInfC{ $ ?_t := t_1$}
\noLine
\UnaryInfC{ $ ?_l := \star$}
\LeftLabel{$\vdash$}
\UnaryInfC{$\ensuremath{\mathtt{force}}~S~s_1~?_T~?_t~?_l~\equiv~T$}
\DisplayProof
$\quad$
\AxiomC{$ ?_T := T$}
\noLine
\UnaryInfC{ $ ?_t := t_2$}
\noLine
\UnaryInfC{ $ ?_l := \star$}
\LeftLabel{$\vdash$}
\UnaryInfC{$\ensuremath{\mathtt{force}}~S~s_2~?_T~?_t~?_l~\equiv~T$}
\DisplayProof
\end{center}
\noindent
that suggests respectively the solution $t_1$ when $s_1$ is matched and
the solution $t_2$ when $s_2$ is, unblocking the reduction by fixing $?_l$ to
be $\star$ so that $(\ensuremath{\mathtt{force}}~S~s_i~T~t_i~\star)$ reduces to $T$ and
the promoted term $(k~S~s_i~T~t_i~\star)$ reduces to $t_i$ as expected.
\subsection{The general solution}
The simplest scenario of the previous section assumed every branch of the
nonuniform coercion to go from $S$ to $T$. The assumption was necessary to
type the very first failing attempts. However, it is useless for the final
working solution presented at the end of the section and thus it can be dropped.
Similarly, the number of branches in the nonuniform coercion also plays no
role since it just corresponds to the number of unification hints to be
declared. Also the order in which the branches are listed (and that determines
the branch that is triggered when multiple branches can) can be preserved by
defining the unification hints in the good order (or by giving an explicit
precedence to them). Finally, branches with a non empty context $\Gamma_i$ also
pose no problem since unification hints are also parameterized by a context.
Thus, to summarize, the final general solution is the following. First of all,
we declare at the beginning of the library the definition of $\ensuremath{\mathtt{force}}$
and $k$
$$
\begin{array}{l}
\ensuremath{\mathtt{force}} :
\forall S:\ensuremath{\mathbf{Type}}. S \rightarrow \forall T:\ensuremath{\mathbf{Type}}. T \rightarrow \ensuremath{\mathbf{1}} \rightarrow \ensuremath{\mathbf{Type}}\\
\ensuremath{\mathtt{force}}~S~s~T~t~lock := \ensuremath{\mathtt{match~}} lock \ensuremath{\mathtt{~with~[~}} \star \Rightarrow T \ensuremath{\mathtt{~]}}\\\\
k : \forall S:\ensuremath{\mathbf{Type}}.\forall s:S. \forall T:\ensuremath{\mathbf{Type}}.\forall t:T.
\ensuremath{\mathbf{1}} \rightarrow \ensuremath{\mathtt{force}}~S~s~T~t\\
k~S~s~T~t~lock := \ensuremath{\mathtt{match~}} lock \ensuremath{\mathtt{~with~[~}} \star \Rightarrow t \ensuremath{\mathtt{~]}}
\end{array}
$$
and we declare $k$ as the only (uniform) coercion.
Then, to declare a nonuniform coercion whose branches are all of the form
$$
\begin{array}{lcl}
\Gamma_i & \vdash &
\begin{array}{lcl}
S_i & \rightarrow & T_i \\
s_i & \mapsto & t_i
\end{array}
\end{array}
$$
the user declares for each branch the following acceptable unification hint:
\begin{prooftree}
\AxiomC{$?_T := T'_i$}
\AxiomC{$?_t := t'_i$}
\AxiomC{$?_l := \star$}
\LeftLabel{$\Gamma'_i \vdash$}
\TrinaryInfC{$\ensuremath{\mathtt{force}}~S'_i~s'_i~?_T~?_t~?_l~\equiv~T'_i$}
\end{prooftree}
\noindent
where $\Gamma'_i$ is obtained from $\Gamma_i$ by turning every variable
$x$ in a meta-variable $?_x$, and the same holds for
$T_i'$, $S'_i$ and $s'_i$ and $t'_i$.
Of course, the system could just implement the standard syntax for nonuniform
coercions as syntactic sugar for the unification hint itself.
In particular, since only one uniform coercion ($k$) is declared, the ITP
code that implements coercion can be simplified, for example dropping
all optimizations for fast indexing/retrieval. Actually, since the coercion
$k$ must be able to promote any type to $\ensuremath{\mathtt{force}}$, the ITP must allow the declaration
of coercions going from any type to something. This feature can be considered
quite unusual since it does not allow any form of efficient indexing. For
instance, the Coq system would not accept $k$ as a coercion. Matita{}, instead,
does not have this restriction. In any case, with a bit of work, we can avoid
the limitation when it is there by redeclaring $k$ multiple times as a coercion
from $T$ to $\ensuremath{\mathtt{force}}$ for every type $T$ whose elements can be promoted.
Remember that, in our implementation of nonuniform coercions via unification
hints, we have assumed the ITP to only allow a coherent set of uniform
coercions. With our implementation, uniform coercions can now be handled as
special cases of nonuniform coercions, with the side effect of having the
possibility to declare incoherent sets. Incoherence is then handled for free
by systematically choosing the first matching hint (or the one with the
greatest priority).
\section{Reasoning about Curryfied morphisms}
\label{currymorph}
During the last year the Matita ITP was redesigned, and one of the novelties
was the introduction of unification hints to support
the mathematical structures technique. Even if the old system (version 0.5.x)
was lacking any kind of structure inference support, we formalized with it a
hierarchy of algebraic and topological structures. It was
clear that porting this formalization to the new system embracing
the mathematical structures approach was a good test bench.
The following example, extracted from the formalization mentioned
above, convinced us to study the notion of nonuniform coercions
presented in this paper.
The setting of the formalization is deeply extensional: points belongs
to types equipped with an equivalence relation, usually called setoids.
Functions respecting the equivalence relation of their source and target types
are called morphisms and are denoted with $A \Rightarrow B$ where
$A$ and $B$ are setoids:
$$
\begin{array}{lcl}
\mathtt{setoid} & := & \{~
T : \ensuremath{\mathbf{Type}};~ \approx : T \to T \to \P;~
\pi_\approx : \mathtt{is\_equiv\_relation}~T~\approx ~\}
\\
A \Rightarrow B & := & \{~
f : A \to B;~
\pi_f : \forall x,y:A.x \approx_A y \to f~x \approx_B f~y ~\}
\end{array}
$$
One may expect that the concept of unary morphism
extends in a straightforward way via currying to n-ary
morphisms as the concept of unary function extends, in an higher order
languages, to n-ary functions. This is not the case
for morphisms, since currying comes at an extra price: in order
to form $A \Rightarrow (B \Rightarrow C)$, the target $B \Rightarrow C$ must
be a setoid, while $B \Rightarrow C$ is a type. We can obtain a setoid of
carrier $B \Rightarrow C$ equipping the carrier with the following extensional
equality relation over morphisms:
$$
A,B : \mathtt{setoid};~ f,g:A \Rightarrow B \vdash f \approx_\Rightarrow g :=
\forall x:A. f~x \approx_B g~x
$$
so that $(B \Rightarrow C, \approx_\Rightarrow B C, \ldots)$ is a setoid
(that we denote as $B \Rightarrow_\approx C$). It is now possible, but
cumbersome and distracting, to write a currified morphism as
$A \Rightarrow (B \Rightarrow_\approx C)$. A much better solution is to
exploit the following nonuniform coercion to allow the implicit promotion
of $B \Rightarrow C$ to $B \Rightarrow_\approx C$:
$$
\begin{array}{lcl}
A : \mathtt{setoid}, B : \mathtt{setoid} & \vdash &
\begin{array}{lcl}
\mathtt{Type} & \rightarrow & \mathtt{setoid} \\
A\Rightarrow B & \mapsto & (A\Rightarrow B, \approx_\Rightarrow, \pi_\Rightarrow)
\end{array}
\end{array}
$$
\noindent
where $\pi_\Rightarrow$ is a
proof that $\approx_\Rightarrow$ is an equivalence relation over morphisms.
Before developing the concept of non uniform coercions we tried other
solutions, but they turned out to be quite unsatisfactory.
First, note that a uniform coercion cannot distinguish the type $A \Rightarrow
B$ from another type, thus would apply in unexpected contexts too.
It would be also hard to achieve the same result relaying on the notational
support the interactive theorem prover offers to overload the $\Rightarrow$
and $\Rightarrow_\approx$ notations. Indeed,
in the very same expression $A \Rightarrow (B \Rightarrow C)$
the first occurrence of $\Rightarrow$ is intended to define a morphism, while
the other occurrence must define a setoid, but interpreting all the
occurrences as setoids, possibly inserting a projection in front,
makes sense too (even if it is not the intended meaning).
This approach would thus require the system to employ a non trivial mechanism
to disambiguate notational overloading.
Last, it is always possible to define different record types for unary, binary,
etc. morphisms, but it turns out to be rather inconvenient as soon as one has
to reason about partially instantiated morphisms.
For example a partially instantiated binary morphism of type
$A \Rightarrow B \Rightarrow C $ would have type $B \to C$, since
functions belonging to that morphism class have type
$A \to B \to C$ and not $A \to B \Rightarrow C$. The real shortcoming
is not that the system has to enrich partially instantiated morphism,
but that the user has to link every n-ary morphism to $n-1$ possible
enrichments (and all of them comprise a different proof for the
$\pi_f$ component of the morphism structure).
\section{Lifting types, not notations}
\label{notation}
Every ITP allows some degree of user configurable notational support.
Consider the following conjecture (where we used the respectively infix
notation $+$ for addition over integers and prefix notation $-$ for inverse)
$$x,y:\ensuremath{\mathbb{Z}} \vdash x + -(y + x) = -y$$
\noindent
and the following well known fact:
$$\mathtt{invmul}:\forall G:\ensuremath{\mathtt{Group}}.\forall a,b:G.(a * b)^{-1} = b^{-1} * a^{-1}$$
\noindent
where $*$ is a notation for $(\ensuremath{\mathtt{op}}~G)$ and $\cdot^{-1}$ for $(\ensuremath{\mathtt{inv}}~G)$.
Thanks to unification hints, we can rewrite the conjecture using
$\mathtt{invmul}$ since the system knows that
$\ensuremath{\mathbb{Z}}$ together with $+$ and $-$ forms the \ensuremath{\mathtt{Group}}{} $\ensuremath{\mathcal{Z}}$. The result, however,
is quite confusing:
$$x,y:\ensuremath{\mathbb{Z}} \vdash x + x^{-1} * y^{-1} = -y$$
To understand it, it is better to de-activate the notation for group,
obtaining
$$x,y:\ensuremath{\mathbb{Z}} \vdash x + \ensuremath{\mathtt{op}}~\ensuremath{\mathcal{Z}}~(\ensuremath{\mathtt{inv}}~\ensuremath{\mathcal{Z}}~x)~(\ensuremath{\mathtt{inv}}~\ensuremath{\mathcal{Z}}~y) = -y$$
The result is clearly correct, since $(\ensuremath{\mathtt{op}}~\ensuremath{\mathcal{Z}})$ reduces to $+$
and $(\ensuremath{\mathtt{inv}}~\ensuremath{\mathcal{Z}})$ to $-$, but, until we perform the reduction, the notation
we get is wrong.
This problem is already extremely frequent when we use the formalization
approach described in Section~\ref{mathcomp}, and it becomes worse
with nonuniform coercions, that may promote even an atomic term written by the
user (like $\ensuremath{\mathtt{zplus}}$, or $+$) to a richer operation, like multiplication in
a group (i.e. denoted with $*$).
A possible solution that avoids reducing the term is to overload the notation
for $+$ and $-$ over $(\ensuremath{\mathtt{op}}~\ensuremath{\mathcal{Z}})$ and $(\ensuremath{\mathtt{inv}}~\ensuremath{\mathcal{Z}})$. This is usually possible
since
the pattern $(\ensuremath{\mathtt{op}}~\ensuremath{\mathcal{Z}})$ is more precise than the pattern $(\ensuremath{\mathtt{op}}~\_)$ which is
matched in the generic group notation. Nevertheless, this is a bad solution
for two reasons: it requires to declare new notations every time we define
a new model of a given structure; and terms like $(\ensuremath{\mathtt{op}}~\ensuremath{\mathcal{Z}}~x~y)$ are left in the
proof term where we would expect to find the simpler $x + y$.
A superior solution is forcing the needed reductions, without requiring any
user intervention. We achieve this in two steps. First of all, we change
the way we declare our hints so that the term that is left by unification
is no longer a projection of the form $(\ensuremath{\mathtt{op}}~\ensuremath{\mathcal{Z}})$, but a redex of the form
$$\ensuremath{\mathtt{op}}~\langle \ensuremath{\mathbb{Z}}, +, -, \mathtt{assoc}~\ensuremath{\mathcal{Z}}, \mathtt{inv\_op\_cancel}~\ensuremath{\mathcal{Z}}, \mathtt{unit\_law}~\ensuremath{\mathcal{Z}}, \mathtt{closed\_op}~\ensuremath{\mathcal{Z}} \rangle$$
Then we change the implementation
of our ITP so that redexes of this kind are automatically reduced as soon
as they are formed. In particular, if the projection in the redex retrieves the
carrier or an operation of the structure (like $\ensuremath{\mathtt{op}}$ or $\ensuremath{\mathtt{inv}}$), the reduction
yields the plain operation of the structure ($+$ or $-$); if the projection
retrieves a property, the reduction yields a projection, like
$\mathtt{assoc}~\ensuremath{\mathcal{Z}}$ which is a compact proof term for the associativity
of addition.
The strategy of reducing this form of redexes
is consistent with the usually implemented one for
$\beta$-redexes. For example, in a system like Coq or Matita, every
rewriting step with an equation $a=b$ is performed by first changing
the conjecture $P$ to the $\beta$-redex $((\lambda x.P[x/a]) a)$ and then
applying the elimination principle for equality. The rewritten conjecture
would be $((\lambda x.P[x/a]) b)$, but, since $\beta$-redex are immediately
reduced, we obtain $P[b/a]$.
The following is the hint able to generate the redexes in the case of the
group of integers. Compare it with the similar hint given
in Section~\ref{uhintsnut}:
\begin{prooftree}
\AxiomC{$?_g := \langle \ensuremath{\mathbb{Z}}, +, -, 0,
\mathtt{assoc}~\ensuremath{\mathcal{Z}}, \mathtt{inv\_op\_cancel}~\ensuremath{\mathcal{Z}},
\mathtt{unit\_law}~\ensuremath{\mathcal{Z}}, \mathtt{closed\_op}~\ensuremath{\mathcal{Z}} \rangle $}
\LeftLabel{$\vdash$}
\UnaryInfC{ $ \ensuremath{\mathtt{op}}~?_g \equiv +$ }
\end{prooftree}
It is worth observing that the expanded form of the hinted solution can
obtained mechanically by separating fields that hold properties, that are kept
as projections of the $\ensuremath{\mathcal{Z}}$ structure, from fields holding types or operations,
that on the contrary are expanded.
It is natural to compare this approach with the alternative one consisting
in implementing an ad hoc simplification tactic that reduces only redexes
involving projections applied to concrete instances of a structure.
While this approach does not necessarily require any deep modification to
the interactive theorem prover,\footnote{Assuming the ITP features a language
to let the user define new tactics, like $\mathcal{L}$-tac.} it has to be
triggered manually and its effect is not recorded in the proof term. Reduction
is not a proof step in CIC, thus the effect of any reduction tactic is left
implicit in the resulting proof term. Hardwiring the greedy reduction strategy
we propose in the proof engine allows to obtain proof term in which the
aforementioned redexes are not present. This affects in a positive way not only
the type checking time, thanks to the smaller size of proof terms, but also the
output quality of any procedure manipulating proof terms.\\ For example
in~\cite{fguidi-proc} and~\cite{csc-decl} the authors reconstruct a proof
script, respectively based on a procedural and a declarative tactic language,
starting from a proof term. Another example is~\cite{natural,YANNTHESIS} where
an explanation of a proof in natural language is obtained solely processing a
proof term. In all these cases a proof term were operations are of the form
$(\ensuremath{\mathtt{op}}~\ensuremath{\mathcal{Z}})$ would be clearly explained with the nomenclature of abstract group
theory, while the original proof was carried on the concrete setting of
integers, and the user did resort to abstract group theory only to justify some
of his proof steps.
\section{Conclusions
\label{conclusions}
The most powerful tool of modern mathematics is abstraction in the following
sense: the mathematical corpus is no longer a flat collection of
facts, but a hierarchy of algebraic, topological and geometrical theories,
all characterized by its own set of axioms and often very rich in models.
Every concrete mathematical object
belongs, directly or via isomorphisms, to several models and the mathematician
effortlessly mixes lemmas from all the relative theories.
This paper aims at making the formalization technique pioneered by
Gonthier et al.~\cite{BigOps,canonical-structures} and the general
machinery behind it~\cite{unification-hints} run
smoothly, allowing the unification heuristic of ITPs to behave in a more
consistent way with respect to
the inference of the mathematical structure a model
belongs to. In particular, the inference is not triggered only when
projections of the structure are involved in a unification problem,
but more generally whenever a term (or type) has to be promoted to
a richer structure. The promotion is expressed in terms of nonuniform
coercions, a novel notion to the authors knowledge, that are shown to
be implementable in terms of uniform coercions and unification hints,
in an higher order logic equipped with dependent types and pattern matching.
\paragraph{Acknowledgments} We would like to thank Jeremy Avigad for
proofreading preliminary versions of this paper.
\bibliographystyle{eptcs}
|
\section{Introduction }
Environment-induced decoherence is a vexed issue in confronting the challenges to quantum information processes. Simple model calculations thus assume significance, to grasp the underlying parameter-regimes that can be manipulated, in order to minimize the effect of the environment. In quantum optics and solid state physics, the device that has gained popularity in recent times is a qubit (which can be approximately represented by a pair of two quantum dots)$^{2,3}$. This system is equivalent to a superconducting Josephson junction$^{4}$ or a spin-1/2 NMR nucleus$^{5}$. A qubit can be described by a two-level Hamiltonian, the quantum mechanics of which is rather straightforward. Environmental influences can also be easily incorporated in this Hamiltonian. There have been two distinctive attempts in modeling the environment, either in terms of ($i$) a classical stochastic process, such as a Gaussian or a telegraph one, or ($ii$) an explicit quantum collection of bosonic oscillators. One of the main objectives in this paper is to seek a unification of these two apparently disparate approaches.
We consider an electron hopping between two single level quantum dots, coupled via a tunneling coefficient $\Delta$ and an asymmetric bias $\epsilon$. Denoting by $|L\rangle$ and $|R\rangle$ the left and right dot states, the qubit Hamiltonian, indicated by the subscript $q$, can be written as,
\begin{equation}
\mathcal{H}_q=\Delta \left(|L\rangle \langle R|+|R\rangle \langle L|\right)+\epsilon\left(|L\rangle \langle L|-|R\rangle \langle R|\right)
\end{equation}
If we introduce the so-called bonding and antibonding states as the spin 'up' and spin 'down' states of the Pauli matrix $\sigma_z$, denoted by $|+\rangle$ and $|-\rangle$ respectively, as
\begin{eqnarray}
|+\rangle=\frac{1}{\sqrt{2}}\left(|L\rangle+|R\rangle\right), \nonumber \\
|-\rangle=\frac{1}{\sqrt{2}}\left(|L\rangle-|R\rangle\right).
\end{eqnarray}
$\mathcal{H}_q$ can be cast into more familiar notation:
\begin{equation}
\mathcal{H}_{q}=\Delta\sigma_{z}+\epsilon\sigma_{x},
\end{equation}
where $\sigma_x$ is the other Pauli matrix that is off-diagonal in the representation in which $\sigma_z$ is diagonal.
In order to incorporate the effect of the environment on the qubit our strategy is to expand the Hilbert space to re-express the system Hamiltonian as
\begin{equation}
\mathcal{H}_{s}=\mathcal{H}_{q}+\tau_{z}(\zeta_{\Delta}\sigma_{z}+\zeta_{\epsilon}\sigma_{x}),
\end{equation}
where $\tau_z$ is yet another Pauli matrix, which could represent a spin-1/2 impurity, for instance. The idea is to couple the $\tau_z$-system to a bath of bosonic oscillators, thereby causing fluctuations in $\tau_z$. The resulting full Hamiltonian can be written as
\begin{equation}
\mathcal{H}=\mathcal{H}_s+\tau_{x}\sum_{k}g_{k}(b_{k}+b_{k}^{\dagger})+\sum_{k}\omega_{k}b_{k}^{\dagger}b_{k}.\end{equation}
Here $b_k^{\dagger}$($b_k$) are boson creation (annihilation) operators, $g_k'^s$ are coupling constants and $\omega_k$ is the harmonic oscillator frequency of the $k^{th}$ mode.
It is interesting to physically assess the effect of the coupling term (proportional to $g_k$) on the system Hamiltonian $\mathcal{H}_s$. Because $\tau_x$ is purely off-diagonal in the $\tau_z$- representation, it would cause 'spin-flips' in $\tau_z$ between two allowed values +1 and -1, as in the celebrated Glauber model of Ising kinetics$^6$. These flips would occur with an amplitude field that would be proportional to $g_k$ and time-varying bosonic operators $b_k^{\dagger}(t)(b_k(t))$, in the interaction representation of the last term in Eq. (5). The net effect on $\mathcal{H}_s$ is a 'quantum noise', encapsulated by $\tau_z(t)$.
In a suitable limit, when the latter, i.e. $\tau_z(t)$, could be replaced by a classical noise $\eta(t)$ that jumps at random between $\pm 1$, the system Hamiltonian would be stochastic, given by,
\begin{equation}
\mathcal{H}_{s}(t)=\mathcal{H}_{q}+\eta(t)(\zeta_{\Delta}\sigma_{z}+\zeta_{\epsilon}\sigma_{x}),
\end{equation}
where $\eta(t)$ is a two-state jump process or a Telegraph Process. While we will present below a fully quantum mechanical treatment of Eq. (5), our aim will also be to set conditions under which a classical stochastic description via a telegraph process of $\eta(t)$, as expounded in detail by Tokura and Itakura$^7$, would ensue from the quantum formulation.
The form of the Hamiltonian in Eq. (5) falls under the general scheme of a system-plus-bath approach to nonequilibrium statistical mechanics, in which the full Hamiltonian is written as,
\begin{equation}
\mathcal{H}=\mathcal{H}_{s}+\mathcal{H}_{I}+\mathcal{H}_{B},
\end{equation}
where $\mathcal{H}_I$ is the interaction Hamiltonian and $\mathcal{H}_B$ is the bath Hamiltonian. In the present instance,
\begin{eqnarray}
\mathcal{H}_I&=&\tau_{x}\sum_{k}g_{k}(b_{k}+b_{k}^{\dagger}), \nonumber \\
\mathcal{H}_B&=&\sum_{k}\omega_{k}b_{k}^{\dagger}b_{k}.
\end{eqnarray}
As we will be employing a Liouvillean approach to the dynamics governed by $\mathcal{H}$, it is useful to explain the notation. The Liouville operator $\mathcal{L}$, associated with $\mathcal{H}$, is defined by
\begin{equation}
\mathcal{L}A=\frac{1}{\hbar}\left[\mathcal{H},A\right],
\end{equation}
where $A$ is an arbitrary but ordinary operator. Thus a Liouville operator yields an ordinary operator when it operates on an operator (as in the right hand side of Eq. (9)), just as an operator, operating on a wavefunction, yields another wavefunction. We will designate the 'states' of a Liouville operator by round brackets: $|)$, in analogy with the Dirac ket vectors: $|\rangle$ for the wavefunctions. The decomposition in Eq. (7) further allows $\mathcal{L}$ to be split as
\begin{equation}
\mathcal{L}=\mathcal{L}_s+\mathcal{L}_I+\mathcal{L}_B.
\end{equation}
The Liouville dynamics effected by $\mathcal{L}$ on the density operator $\rho$, leading to a non-Markovian master equation, in the context of the decoherence of a qubit (as well as a collection of qubits) have been extensively studied by Kurizki and collaborators$^9$. They have also considered the additional influence of external time-varying fields for controlling decoherence, dynamically. While we will be employing a very similar approach, our treatment will differ from Gordon etal$^9$. The latter utilize the Born approximation, valid for weak coupling constants $g_k$, whereas we will consider strong coupling in which $g_k$ will be treated to all orders.
Such strong coupling considerations are the hallmark of functional integral treatments of quantum dissipative systems eg. that of a spin-boson Hamiltonian$^{10}$. Furthermore our focus in the classical limit of the quantum noise will be a telegraph process as opposed to the much-studied Gaussian stochastic processes. Similar spin-boson Hamiltonians, akin to Eq. (5), have been considered in the past$^{11}$, though the comparison has not been dealt with, to the best of our knowledge.
With these preliminaries the paper is section wise organised as follows. In Sec. II we outline the mathematical steps for a resolvent expansion of the averaged time-development operator in the Laplace transform space, the average being carried out over the Hilbert spaces of $\tau_z$ and $\mathcal{H}_B$. The latter is expressed in terms of a 'self-energy' whose form is provided. In Sec. III explicit results are presented for the much explored Ohmic-dissipation model and comparisons are drawn with the classical telegraph process, results of which are given in the Appendix. For facilitating comparison we discuss fluctuation in bias and hopping separately. In a subsection of Sec. III, we indicate results for non-Ohmic situations as well. Finally in Sec. IV, we discuss the issue of partial decoherence, recently studied by us in the context of the telegraph process$^{12}$, to put it under the perspective of quantum noise. The section V contains a few concluding remarks.
\section{Theoretical Formulation}
In order to incorporate strong-coupling effects it is convenient to transform the Hamiltonian $\mathcal{H}$ in Eq. (5) with the aid of a unitary transformation defined by the operator:
\begin{equation}
S=\exp\left[-\sum_{q}\frac{g_{q}}{2\omega_{q}}\left(b_{q}-b_{q}^{\dagger}\right)\tau_{z}\right]\exp\left[-i\frac{\pi}{2}\tau_y\right].
\end{equation}
The transformed Hamiltonian is given by
\begin{equation}
\tilde{\mathcal{H}}=S\mathcal{H}S^{-1}.
\end{equation}
What the second term in $S$ does is to cause a rotation in the $\tau$-space by an angle of $\pi/2$ about the y-axis, in the anti-clockwise direction such that $\tau_x\rightarrow\tau_z$ and $\tau_z\rightarrow -\tau_x$. The first term in $S$
then eliminates the interaction term $\mathcal{H}_I$ but puts the onus of coupling on $\mathcal{H}_s$ itself (besides generating an innocuous 'counter term' that is constant, and can be dropped). The net result is
\begin{equation}
\tilde{\mathcal{H}}=\mathcal{H}_q-\frac{1}{2}\left(\tau^{-}B_{+}+\tau^{+}B_{-}\right)\left(\zeta_\Delta\sigma_z+\zeta_\epsilon\sigma_x\right)+\sum_{k}\omega_{k}b_{k}^{\dagger}b_{k},
\end{equation}
where
\begin{eqnarray}
\tau_\pm&=&\tau_x\pm i\tau_y, \nonumber \\
B_{\pm}&=&\exp\left[\pm\sum_{q}\frac{g_{q}}{2\omega_{q}}\left(b_{q}-b_{q}^{\dagger}\right)\right].
\end{eqnarray}
The transformed Hamiltonian $\mathcal{H}$ is now endowed with a new interaction term that may be written as
\begin{equation}
\tilde{\mathcal{H}_I}=-\frac{1}{2}\left(\tau^{-}B_{+}+\tau^{+}B_{-}\right)\left(\zeta_\Delta\sigma_z+\zeta_\epsilon\sigma_x\right).
\end{equation}
The important point however is that any perturbation treatment of $\tilde{\mathcal{H}_I}$ is tantamount to treating the coupling to all orders as $g_q$ now occurs in the exponent as is evident from the structure of $B_\pm$ (cf. Eq. (14)). The equation (13) is again of the generic form of Eq. (7) except that the 'system' is now the qubit itself as $\mathcal{H}_s$ is replaced by $\mathcal{H}_q$!
With reference to the Liovillean in Eq. (10) the equation of motion for the density operator $\rho$ reads
\begin{equation}
\frac{\partial}{\partial t}\rho(t)=-i\mathcal{L}\rho(t).
\end{equation}
Unlike other approaches$^{7,9}$ we find it convenient to work with the Laplace transforms, defined by
\begin{equation}
\rho(z)=\int_{0}^{\infty}dt e^{-zt}\rho(t).
\end{equation}
From Eq. (16) then
\begin{eqnarray}
\rho(z)&=&\left(U(z)\right)\rho(t=0), \nonumber \\
U(z)&=&\left(z+i\mathcal{L}\right)^{-1}.
\end{eqnarray}
We employ the usual factorization approximation that at $t=0$ the qubit and the bath are decoupled$^8$, so that
\begin{equation}
\rho(t=0)=\rho_q(0)\otimes\rho_B,
\end{equation}
where $\rho_q(0)$ is the qubit density operator at time $t=0$, $\rho_B=\exp(-\beta\mathcal{H}_B)/Z_B$, is the canonical density operator for the bath and $Z_B$ is the corresponding partition function. Underlying the prescription in Eq. (19) is the assumption that the bath always remains in quilibrium at a fixed temperature T while the qubit evolves from an arbitrary state. For most of our derived expressions we shall assume that the electron is on the left dot at $t=0$ i.e.
\begin{equation}
\rho_q(0)=|L\rangle\langle L|.
\end{equation}
Translated to the 'bonding' and 'anti-bonding' basis, this implies
\begin{equation}
\rho_q(0)=\frac{1}{2}\left(1+\sigma_x\right).
\end{equation}
(We shall return in Sec. IV to a more general initial condition.) The Laplace transform of the so-called reduced density operator can be written as
\begin{equation}
\rho_\mathcal{R}(z)={Tr_\tau\otimes Tr_B\left[\left(U(z)\right)\rho_B\right]}\rho_q(0),
\end{equation}
where $Tr_\tau(...)$ denotes trace over the eigenstates of $\tau_z$ whereas $Tr_B(...)$ specifies trace over the bath (i.e. eigenstates of $\mathcal{H}_B$). Because the trace is invariant under the unitary transformations (eg. in Eq. (11)), Eq. (22) can be equivalently expressed as
\begin{equation}
\rho_\mathcal{R}(z)={Tr_\tau\otimes Tr_B\left[\left(\tilde{U}(z)\right)\rho_B\right]}\rho_q(0),
\end{equation}
where the tilde on $U(z)$ implies that the corresponding time-evolution is now governed by $\tilde{U}$ of Eq. (13).
Denoting the Liouville operator associated with $\tilde{\mathcal{H}}_I$ as $\tilde{\mathcal{L}}_I$, we have upto second order in $\tilde{\mathcal{H}}_I^8$,
\begin{equation}
\rho_R(z)=\left[\left(z+i\mathcal{L}_q+\tilde{\sum}(z)\right)^{-1}\rho_q(0)\right],
\end{equation}
where the self-energy $\tilde{\sum}(z)$ is
\begin{equation}
\tilde{\sum}(z)=Tr_\tau\otimes {Tr_B\left[\tilde{\mathcal{L}}_I\left(z+i(\mathcal{L}_q+\mathcal{L}_B)\right)^{-1}\tilde{\mathcal{L}}_I\rho_B\right]}.
\end{equation}
The z-dependence (or frequency dependence) of $\tilde{\sum}(z)$ implies that the underlying dynamics is non-Markovian, in general. We emphasize once more that though the self-energy is computed to second order in $\tilde{\mathcal{L}}_I$, the theory is valid to arbitrary orders in the original coupling constants $g_k$. Hence, the resultant master equation in the time domain that emanates from Eq. (24) is of more general validity than the one obtained in the Born-approximation$^9$.
Our next step is to write the matrix elements of the self-energy $\tilde{\sum}(z)$. It is clear that $\tilde{\sum}(z)$ is a Liouvillean in the $\sigma$-subspace alone because the $\tau$ and the bath variables are averaged over, in Eq. (25). Hence, $\tilde{\sum}(z)$ has a $4\times 4$ matrix representation in the 'eigen basis' $|\mu\nu)$ (note the round brackets used for 'states' of the Liouvillean), wherein $|\mu\rangle$, $|\nu\rangle$,... are the eigenkets of $\sigma_z$. Clearly then, the rows and columns of the $4\times 4$ matrix would be labelled by $|++)$, $|--)$, $|+-)$ and $|-+)$, respectively.
Additionally, while performing the traces in Eq. (25), we would need to sum over the eigenstates of $\tau_z$ and $\mathcal{H}_B$, as in the following:
\begin{eqnarray}
\tau_z|s_1\rangle&=&s_1|s_1\rangle, \nonumber \\
\mathcal{H}_B|n_b\rangle&=&E_{n_b}|n_b\rangle,
\end{eqnarray}
etc, With these explanatory remarks about the notation, the matrix elements of $\tilde{\sum}(z)$ turn out to be generalized forms of those given in Ref. [13]:
\begin{widetext}
\begin{eqnarray}
(\mu\nu|\tilde{\sum}(z)|\mu'\nu')=\sum_{n_{b}n_{b}^{1}s_{1}s_{2}}&[&\delta_{\nu\nu'}\sum_{\eta}\frac{\left\langle \mu n_{b}s_{1}\right|H_{I}\left|\eta n_{b}^{1}s_{2}\right\rangle \left\langle \eta n_{b}^{1}s_{2}\right|H_{I}\left|\mu'n_{b}s_{1}\right\rangle }{z+i\left(E_{\eta}-E_{\nu}\right)+i\left(E_{n_{b}^{1}}-E_{n_{b}}\right)}\nonumber \\
&+&\delta_{\mu'\mu}\sum_{\eta}\frac{\left\langle \nu'n_{b}s_{1}\right|H_{I}\left|\eta n_{b}^{1}s_{2}\right\rangle \left\langle \eta n_{b}^{1}s_{2}\right|H_{I}\left|\nu n_{b}s_{1}\right\rangle }{z+i\left(E_{\mu}-E_{\eta}\right)+i\left(E_{n_{b}}-E_{n_{b}^{1}}\right)}\nonumber \\
&-&\frac{\left\langle \mu n_{b}s_{1}\right|H_{I}\left|\mu'n_{b}^{1}s_{2}\right\rangle \left\langle \nu'n_{b}^{1}s_{2}\right|H_{I}\left|\nu n_{b}s_{1}\right\rangle }{z+i\left(E_{\mu'}-E_{\nu}\right)+i\left(E_{n_{b}^{1}}-E_{n_{b}}\right)}\nonumber \\
&-&\frac{\left\langle \mu n_{b}s_{1}\right|H_{I}\left|\mu'n_{b}^{1}s_{2}\right\rangle \left\langle \nu'n_{b}^{1}s_{2}\right|H_{I}\left|\nu n_{b}s_{1}\right\rangle }{z+i\left(E_{\mu}-E_{\nu'}\right)+i\left(E_{n_{b}}-E_{n_{b}^{1}}\right)}]\langle n_b|\rho_B|n_b\rangle.
\end{eqnarray}
From the structure of $\tilde{\mathcal{H}}_I$ in Eq. (15) it is evident that the bath variables enter only through the operators $B_\pm$. Therefore, when we sum over the bath indices $n_b$, $n_b'$, etc., with the Boltzmann weight $\langle n_b|\rho_B|n_b\rangle$, we end up with correlation fucntions of $B_\pm$, as in the following:
\begin{equation}
\Phi_{ij}(t)=Tr_B\left(\rho_B B_i(0)B_j(t)\right) = \langle B_i(0)B_j(t)\rangle \quad(i,j=+ or -).
\end{equation}
As it happens, the only non-zero components of the correlation function are
\begin{eqnarray}
&&\Phi_{+-}(t)=\Phi_{-+}(t)= \langle B_\pm(0)B_\mp(t)\rangle \nonumber \\
&&=\exp\left\{-\sum_{q}\frac{4g_{q}^2}{\omega_{q}^2}\left[\coth{\left(\frac{\beta\omega_q}{2}\right)}\left(1-\cos\omega_qt\right)+i\sin\omega_qt\right]\right\}.
\end{eqnarray}
An essential attribute of quantum dissipative system is that the system $\mathcal{H}_B$ is taken to the limit of an infinitely large number of bosonic modes. This implies that the discrete sum over $q$ has to be replaced by an integral over continously varying frequencies $\omega$ with an appropriate weight called the 'spectral density' $J(\omega)$, thus
\begin{equation}
\phi(\tau)=\exp\left\{-2\int_{0}^{\infty}d\omega \frac{J(\omega)}{\omega^2}\left[\coth{\left(\frac{\beta\omega}{2}\right)}\left(1-\cos\omega\tau\right)+i\sin\omega\tau\right]\right\},
\end{equation}
where
\begin{equation}
J(\omega)=2\sum_{q}g_{q}^2\delta(\omega-\omega_q).
\end{equation}
\end{widetext}
\section{Explicit Results and Comparison with Telegraph Noise}
\subsection{Ohmic Dissipation}
In the literature on dissipative quantum systems one model for the spectral density $J(\omega)$ that has received the maximum amount of attention is what gives rise to Ohmic dissipation$^{14}$. In this, $J(\omega)$ is assumed to have a linear frequency dependence with an exponential cut-off $\omega_c$
\begin{equation}
J(\omega)=K\omega\exp\left(-\frac{\omega}{\omega_c}\right),
\end{equation}
where $K$ is a phenomenological damping parameter that measures the strength of the coupling with the heat bath. This form allows for an analytic expression for the Laplace transform as defined in Eq. (17), for $\beta\omega_c>>1$ as
\begin{equation}
\tilde{\phi}(z)=\frac{e^{i\pi K}}{\omega_c}\left(\frac{2\pi}{\beta\omega_c}\right)^{(2K-1)}\frac{\Gamma{(1-2K)}\Gamma{(K+z\beta/2\pi)}}{\Gamma{(1-K+z\beta/2\pi)}}.
\end{equation}
where $\Gamma(..)$ denotes gamma functions.
The significance of the condition $\beta\omega_c>>1$ can be ascertained from the fact that the quantity $\hbar\tau_q=\beta$ (Note: $\hbar$ has been set to unity, hitherto) sets the time scale over which quantum coherence is maintained. Thus, for all frequencies greater than the cut-off $\omega_c$, quantum fluctuations remain strictly coherent. Before we take up the comparison with the telegraph noise it is pertinent to point out that it is meaningful to think of a quantum bath as classical only in the incoherent regime i.e. over time scales larger than $\tau_q$. In other words, the classical limit of the bath obtains for frequencies $\omega<<1/\tau_q$, i.e. the temperature is significantly large, but not so large as to violate the condition $\omega_c\tau_q>>1$, if we want to restrict our discussion within the domain of validity of the analytical result of Eq. (33). If the restriction $\omega\tau_q<<1$ does apply, the frequency or z-dependence in the arguments of the Gamma functions in Eq. (33) can be dropped yielding
\begin{equation}
\lambda=\tilde{\phi}(z=0)=\frac{e^{i\pi K}}{\omega_c}\left(\frac{2\pi}{\beta\omega_c}\right)^{(2K-1)}\frac{\Gamma{(1-2K)}\Gamma{(K)}}{\Gamma{(1-K)}}.
\end{equation}
The parameter $\lambda$ will turn out later to be related to the jump rate of the telegraph process.
With this background we are ready to provide explicit results for the different elements of the density matrix in order to examine decoherence as well as to draw comparison with the results for the telegraph noise. Because the latter have been extensively studied recently by Itakura and Tokura$^7$, we will follow their lead in ignoring the bias term in the original qubit Hamiltonian $\mathcal{H}_q$, i.e. set $\epsilon=0$ though it should be evidently clear that the formalism can easily embrace asymmetric cases ($\epsilon\neq 0$) as well. With this, the relevant Hamiltonian for further considerations can then be written as (cf., Eqs. (3) and (13))
\begin{equation}
\tilde{\mathcal{H}}=\Delta\sigma_z-\frac{1}{2}\left(\tau^{-}B_{+}+\tau^{+}B_{-}\right)\left(\zeta_\Delta\sigma_z+\zeta_\epsilon\sigma_x\right)+\sum_{k}\omega_{k}b_{k}^{\dagger}b_{k}
\end{equation}
For ease of discussion and for remaining close to the treatment of Itakura and Tokura we take up the cases of fluctuation in bias ($\zeta_\Delta=0$) and fluctuation in hopping ($\zeta_\epsilon=0$) separately below.
\subsection{Fluctuation in Bias ($\zeta_\Delta=0$)}
The appropriate interaction Hamiltonian that has to be substituted in th expression for the self energy $\sum(z)$ in Eq. (25) is now given by
\begin{equation}
\tilde{H}_I=-\frac{1}{2}\zeta_\epsilon\sigma_x\left(\tau^{-}B_{+}+\tau^{+}B_{-}\right)
\end{equation}
Carrying out the summations implied in Eq. (25) we arrive at
\begin{equation}
\tilde{\sum}(z)=\left[\begin{array}{cccc}
a_1(z) & -a_1(z) & 0 & 0\\
-a_2(z) & a_{2}(z) & 0 & 0\\
0 & 0 & a_3(z) & -a_3(z)\\
0 & 0 & -a_3(z) & a_3(z)\end{array}\right],\end{equation}
where
\begin{eqnarray}
a_1(z)&=&2\zeta_{\epsilon}^{2}[\tilde{\phi}(z_{+})+\tilde{\phi}'(z_{-})], a_2(z)=2\zeta_{\epsilon}^{2}[\tilde{\phi}(z_{-})+\tilde{\phi}'(z_{+})] \nonumber \\
a_3(z)&=&2\zeta_{\epsilon}^{2}[\tilde{\phi}(z)+\tilde{\phi}'(z)], \quad z_\pm=z\pm2i\Delta.
\end{eqnarray}
Here $\tilde{\phi}(z)$ and $\tilde{\phi}'(z)$ are the Laplace transforms of the correlation functions $\phi(t)$ and $\phi(-t)$, respectively.
The matrix element of $\left(z+i\mathcal{L}_q+\tilde{\sum}(z)\right)$, required in Eq. (24), is obtained from a combination of Eq. (27) and the definition of the matrix elements of the Liouville operator $\mathcal{L}_q$ a la Eq. (A.7). The resultant matrix is of a block-diagonal form and can be easily inverted to yield
\begin{eqnarray}
\left(\rho_\mathcal{R}(z)\right)_{++}&=&\left(\rho_\mathcal{R}(z)\right)_{--}=\frac{1}{2z}, \nonumber \\
\left(\rho_\mathcal{R}(z)\right)_{+-}&=&\frac{z+2i\Delta+4\zeta_\epsilon^2\left(\tilde{\phi}(z)+\tilde{\phi}'(z)\right)}{2\left[z\left(z+4\zeta_\epsilon^2\left(\tilde{\phi}(z)+\tilde{\phi}'(z)\right)\right)+4\Delta^2\right]}.
\end{eqnarray}
\begin{figure}[h!]
\centering
\includegraphics[scale=0.50]{Fig1.jpg}
\caption{ $Im[\left(\rho_\mathcal{R}\right)_{LR}(t)]$ is plotted for $\beta=100$, $k=0.45$ and $\Delta=2$}
\end{figure}
\begin{figure}[h!]
\centering
\includegraphics[scale=0.50]{Fig2.jpg}
\caption{ $\left(\rho_\mathcal{R}\right)_{LL}(t)$ is plotted for $\beta=100$, $k=0.45$ and $\Delta=2$}
\end{figure}
\begin{figure}[h!]
\centering
\includegraphics[scale=0.50]{Fig3.jpg}
\caption{ $\left(\rho_\mathcal{R}\right)_{RR}(t)$ is plotted for $\beta=100$, $k=0.45$ and $\Delta=2$}
\end{figure}
It is of course possible to go back to the dot-basis with the aid of Eq. (2). Results are shown in Figs. 1, 2 and 3 in which $Im[\rho_{LR}(t)]$, $\rho_{LL}(t)$ and $\rho_{RR}(t)$ are plotted versus time t for a certain choice of parameters. While $Im[\rho_{LR}(t)]$ oscillates and decays to zero as $t\rightarrow\infty$, the diagonal elements, proportional to the populations of the two dot states, oscillate and settle to the common value of 0.5, as we expect for full decoherence. If we compare Figs 1-3 with Fig A.1 for the telegraph process below, it is evident that the quantum case exhibits persistence of oscillations over longer time scales.
\subsection{Fluctuation in Hopping ($\zeta_\epsilon=0$)}
The appropriate interaction Hamiltonian that has to be substituted in the expression for the self energy $\tilde{\sum}(z)$ in Eq. (25) is now given by
\begin{equation}
\tilde{H}_I=-\frac{1}{2}\zeta_\epsilon\sigma_z\left(\tau^{-}B_{+}+\tau^{+}B_{-}\right)
\end{equation}
Following the formalism developed in Sec. II and using Eq. (25), the self energy matrix $\tilde{\sum}(z)$ is given by
\begin{equation}
\tilde{\sum}(z)=\left[\begin{array}{cccc}
0 & 0 & 0 & 0\\
0 & 0 & 0 & 0\\
0 & 0 & b_3(z) & 0\\
0 & 0 & 0 & b_4(z)\end{array}\right],\end{equation}
where
\begin{eqnarray}
b_3(z)&=&4\zeta_{\Delta}^{2}[\phi'(z_{+})+\phi(z_{+})], \nonumber \\
b_4(z)&=&4\zeta_{\Delta}^{2}[\phi'(z_-)+\phi(z_-)], \quad z_\pm=z\pm2i\Delta,
\end{eqnarray}
where $\phi(z)$ and $\phi'(z)$ have been defined in the earlier section.
Using the form $\tilde{U}(z)=\left(z+i\mathcal{L}_s+\tilde{\sum}(z)\right)^{-1}$, it is evident from the diagonal form of the matrix that the population of the bonding and anti bonding state remains constant for all $t>0$. Using $\tilde{U}(z)$ the matrix elements are given by,
\begin{eqnarray}
\left(\rho_\mathcal{R}\right)_{++}(z)&=&\left(\rho_R\right)_{--}(z)=\frac{1}{2z}, \nonumber \\
\left(\rho_\mathcal{R}\right)_{+-}(z)&=&\frac{1}{2}\left[\frac{1}{z+2i\Delta+b_{3}}\right].
\end{eqnarray}
However in the dot basis we expect the usual saturation to the common value of 0.5 for both left and right dot populations. Fig. (4) shows the variation of left and right dot populations with time and also the loss of coherence over time from the decay of $Im(\rho_{LR}(t))$.
\begin{figure}
\vspace{0.5cm}
\begin{center}
\includegraphics[scale=0.50]{Fig4.jpg}
\end{center}
\caption{ $\left(\rho_\mathcal{R}\right)_{LL}(t)$ (solid), $\left(\rho_\mathcal{R}\right)_{RR}(t)$ (dashed) and $Im\left(\rho_\mathcal{R}\right)_{RR}(t)$ (dotted) are plotted for $\beta=100$, $k=0.45$ and $\Delta=2$}
\end{figure}
We are now in position to make a connection between the results of quantum and classical regimes considered separately in Sec.III and Appendix A. There are two critical distinctions between the two regimes. At the outset, the quantum behavior is non-Markovian, reflected in the z (or frequency)-dependence of the self-energy $\tilde{\sum}(z)$ in Eq. (25). On the other hand, in the classical stochastic case, the Markovian assumption is invoked, to begin with, as seen in Eq. (A.2) below. The second crucial ingredient in the quantum situation, even after the Markovian approximation is incorporated, is in the temperature-dependence of the relaxation rate $\lambda$. While the latter, in the classical domain, is usually endowed with an exponential dependence on temperature, with a barrier-activation in mind, it has a richer structure in the quantum regime, as discussed below.
We mention below in Appendix. A.2 that in the long time region the off-diagonal element of the system density operator follows an exponential decay with time constant $T_{2}=\frac{1}{\sqrt{ |\lambda^{2}-16\zeta_{\Delta}^{2}|}}$. From Eq. (43), as it can be clearly seen from the Laplace transform of $\rho_{+-}$, the decay is exponential only if we assume $b_{3}(z)$ to be independent of $z$, wherein the time constant is just the inverse of $Re[b_{3}]$. Thus the comparison with the classical telegraph process is meaningful only if one takes the Markovian limit of heat bath induced relaxation, at the outset.
At sufficiently high temperatures given by $k_{B}T>>\zeta_{\Delta}/K$ we can see from Eq. (33) that $\tilde{\phi}(z)$ can be approximated as $\tilde{\phi}(z)=\frac{e^{i\pi K}}{\omega_{c}}\left(\frac{2\pi}{\beta\omega_{c}}\right)^{2K-1}\frac{\Gamma{(1-2K)}\Gamma{(K)}}{\Gamma{(1-K)}}$ which leads to a rather simple expression for $b_{3}$ (for $z=0$) in Eq. (42) as $b_{3}=\frac{8\zeta_{\Delta}^2}{\omega_{c}}\cos{(\pi K)}(\frac{2\pi}{\beta\omega_{c}})^{2K-1}\frac{\Gamma{(1-2K)}\Gamma{(K)}}{\Gamma{(1-K)}}$, that is real. Hence $\lambda$, which appears as the jump rate in the case of telegraph noise, discussed in the Appendix in the sequel, can be compared with the relaxation rate $Re[b_{3}]$ in the case of quantum noise, is now a function of $K$ and $\beta$. Also, as it is obvious from the expression of the relaxation rate that the dependence on temperature is not an exponential, rather it shows a power law behavior. However, it is pertinent to emphasize that these conclusions hinge on our assumption that quantum dissipation is Ohmic. In the more general case, discussed below in Sec. III D, we will see that the temperature-dependence of the relaxation rate can in fact be exponential, under certain special situations.
At low temperatures and weak damping i.e. 'small' $K$, $\tilde{\phi}(z)$ is given by $\tilde{\phi}(z)=\frac{\mu^2}{z+2\pi/\beta}$ where $\mu$ is a renormalized form of tunneling frequency given by $\mu^2=4\zeta_{\Delta}^2(\frac{2\pi}{\beta\omega_{c}})^{2K-1}\Gamma{(1-2K)}$. Using this form of $\tilde{\phi}(z)$, and Eq. (43) we can deduce,
\begin{equation}
\left(\rho_\mathcal{R}\right)_{+-}(z)=\frac{1}{2}\left[\frac{1}{z+2i\Delta-\frac{4i\Delta\cos{(\pi K)}\mu^2}{(2\pi/\beta)^2+4\Delta^2}+\frac{4\mu^2(\pi/\beta)\cos{(\pi K)}}{(2\pi/\beta)^2+4\Delta^2}}\right].
\end{equation}
From Eq. (44), it is evident that the relaxation rate becomes $\frac{4\mu^2(\pi/\beta)\cos{(\pi K)}}{(2\pi/\beta)^2+4\Delta^2}$, which is far from an exponential in temperature. Thus it may be stressed that even in the Markovian limit the temperature dependence of both the relaxation rate as well as the Rabi frequency is much more complex than an exponential, as is usually assumed for classical activated processes.
These conclusions are similar to those obtained earlier in the context of spin relaxation of a muon, tunneling in a double well$^{15}$.
\subsection{Non-Ohmic case}
We had in the text presented results based on the assumption of Ohmic dissipation i.e. choosing $J(\omega)=K\omega\exp{(-\frac{\omega}{\omega_{c}})}$. The Ohmic model is a rather limited one, not applicable to phonons but is relevant to a bath consisting of electron-hole excitations near the Fermi surface, as in metals.$^{14,15}$ In this subsection we will allow for a general spectral density function, and will be analysing the relaxation rate in the case of quantum noise. We will try to make a connection with the classical Telegraph process and show that in the high-temperature limit, the relaxation rate indeed follows an exponential dependence on temperature with however quantum corrections, which was not the case in the Ohmic limit.
Our starting point is Eq.(30) which upon clubbing the oscillating terms together, can be rewritten as,
\begin{widetext}
\begin{equation}
\phi(\tau)=\exp\left[-2\int_{0}^{\infty}d\omega \frac{J(\omega)}{\omega^2}\left[\coth{\left(\frac{\beta\omega}{2}\right)}-\frac{\cos{(\omega(\tau-i\beta/2))}}{\sinh{(\frac{\beta\omega}{2})}}\right]\right].
\end{equation}
By making a change of variable from $\tau-i\beta/2$ to $t$ in Eq. (45), $\phi(z)$ can be written as,
\begin{equation}
\phi(z)= \exp{\left(-\frac{iz\beta}{2}\right)}\int_{-\frac{i\beta}{2}}^{\infty}dt e^{-zt}\exp\left[-2\int_{0}^{\infty}d\omega \frac{J(\omega)}{\omega^2}\left[\coth{\left(\frac{\beta\omega}{2}\right)}-\frac{\cos{(\omega t)}}{\sinh{(\frac{\beta\omega}{2})}}\right]\right],
\end{equation}
In the high temperature regime$^{19}$, the integrand in the second term in Eq. (46), can be replaced by its short time limit as,
\begin{equation}
\phi(z)= \exp{\left(-\frac{iz\beta}{2}\right)}\int_{-\frac{i\beta}{2}}^{\infty}dt e^{-zt}\exp\left[-2\int_{0}^{\infty}d\omega \frac{J(\omega)}{\omega^2}\left[\tanh{\left(\frac{\beta\omega}{4}\right)}+\frac{\omega^2t^2}{2\sinh{(\frac{\beta\omega}{2})}}\right]\right].
\end{equation}
Also in the high temperature regime i.e. low $\beta$, $\tanh(x)$ and $\sinh(x)$ can be replaced by the argument i.e. $x$, and hence the integrations in Eq. (47) can be carried out easily to yield,
\begin{equation}
\phi(z)=\exp{\left(-\beta(a+iz/2)\right)}\int_{-\frac{i\beta}{2}}^{\infty}dt e^{-zt}e^{\frac{4at^2}{\beta}},
\end{equation}
where $a=\frac{1}{2}\int_{0}^{\infty}d\omega \frac{J(\omega)}{\omega}$. Also $\phi'(z)$ is obtained by replacing $i$ by $-i$ in Eq. (48). Thus the relaxation rate which is given by Re($b_{3}$), for the quantum case in Sec. III.C, is obtained by imposing the Markovian limit i.e. putting $z=0$, leading to
\begin{equation}
b_{3}=4\zeta_{\Delta}^2\left(\exp{\left(-\beta(a-\Delta-\frac{\Delta^2}{4a})\right)}\int_{\frac{-i\beta}{2}+\frac{i\beta\Delta}{4a}}^{\infty}du \exp{\left(-4au^2/\beta \right)}+\exp{\left(-\beta(a+\Delta-\frac{\Delta^2}{4a})\right)}\int_{\frac{i\beta}{2}+\frac{i\beta\Delta}{4a}}^{\infty}du \exp{\left(-4au^2/\beta \right)}\right).
\end{equation}
\end{widetext}
Finally the lower limit in Eq. (49) can be set to zero and thus the integral is just a function of temperature which follows a power law. Thus the dominant behavior of the relaxation rate is an exponential dependence on temperature, which is what we expect at high temperatures. This analysis therefore provides a regime where we can more meaningfully compare the relaxation rate $\lambda$ for the case of classical telegraph process with the Laplace transform of the bath correlation function in the quantum case. Note that the present treatment does not depend on any specific assumption for the spectral density $J(\omega)$ but is more generally couched, irrespective of whether the phonons are acoustic or optic.
\section{The Issue of Partial Decoherence}
In the context of fluctuation in hopping, it is evident that the system Hamiltonian commutes with the interaction Hamiltonian, which means there is no energy exchange between the system and the bath. Such cases are important for studying partial decoherence$^8$. Thus we generalize our discussion to a more general initial state of a qubit (than of Eq. (20)) given by,
\begin{equation}
|\Psi\rangle=\cos{\left(\frac{\theta}{2}\right)}|L\rangle+e^{i\gamma}\sin{\left(\frac{\theta}{2}\right)}|R\rangle.
\end{equation}
In such a scenario it can be established that the density matrix does not
evolve to a completely mixed state, rather it approches a limiting
value, which does contain the off-diagonal components. In this
situtation coherence is not completely lost and the off-diagonal terms give information about the initial state of the system. We note that the initial qubit information can be retrieved, either from
experiments related to persistent current or by measuring the population
of states. The former path is not applicable in our case as we have not
allowed for the existence of an Aharonov-Bohm flux$^8$.
The density matrix (in $|L\rangle$ and $|R\rangle$ basis) at time $t=0$ is
given from Eq. (50) by,
\begin{equation}
\tilde{\rho_{\mathcal{R}}}(0)=\left[\begin{array}{cc}\cos^{2}(\theta/2) &
\cos(\theta/2)\sin(\theta/2)e^{-i\gamma}\\\cos(\theta/2)\sin(\theta/2)e^{i\gamma}
& \sin^{2}(\theta/2)\end{array}\right],
\end{equation}
which can be translated to the bonding-antibonding basis, via the transformation
given by Eq. (2). The off-diagonal term $\left(\rho_{\mathcal{R}}\right)_{LR}(t\rightarrow \infty)$, which measures coherence, does not go to zero unlike in the cases discussed in Sec.(III.C and A.2). These particular symmetric cases of $\epsilon=0$ and $\zeta_{\epsilon}=0$ show partial coherence which is important for decoherence-free quantum computation protocols.
\begin{figure}[h!]
\centering
\includegraphics[scale=0.50]{Fig5.jpg}
\caption{ $Re[\left(\rho_{\mathcal{R}}\right)_{+-}(t)]$ for quantum (solid) and classical (Dashed) cases are plotted versus time t.}
\end{figure}
\begin{figure}[h!]
\centering
\includegraphics[scale=0.50]{Fig6.jpg}
\caption{ $Im[\left(\rho_{\mathcal{R}}\right)_{+-}(t)]$ for quantum (solid) and classical (Dashed) cases are plotted versus time t.}
\end{figure}
The above formalism is general and is valid for any kind of environment, as long as the commutator of the system Hamiltonian and the interaction Hamiltonian is zero. As shown in the cases of classical and quantum noises (Figs. 5 and 6), the imaginary and real parts of $\left(\rho_{\mathcal{R}}\right)_{+-}(t)\rightarrow 0$ as $t \rightarrow \infty$, whereas the populations of bonding and anti-bonding states stay constant, as is evident from Eq. (43).
Thus, in general, the off-diagonal elements of the density matrix (in $|L\rangle$ and $|R\rangle$ basis) approach limiting values, emphasizing that the asymptotic state is not a fully mixed state:
\begin{equation}
\hat{\rho_{\mathcal{R}}}(t\rightarrow \infty) = \left[\begin{array}{cc}\frac{1}{2} &
Re(\left(\rho_\mathcal{R}\right)_{LR}(0))\\Re(\left(\rho_\mathcal{R}\right)_{RL}(0))
& \frac{1}{2}\end{array}\right].
\end{equation}
At this stage it is also worthwhile to underscore the importance of time scales involved in attaining partial decoherence. For this we define a function $C(t)$ as:
\begin{equation}
C(t)=\frac{Im(\left(\rho_\mathcal{R}\right)_{LR}(\infty))-Im(\left(\rho_\mathcal{R}\right)_{LR}(t))}{Im(\left(\rho_\mathcal{R}\right)_{LR}(\infty))-Im(\left(\rho_\mathcal{R}\right)_{LR}(0))}.
\end{equation}
This function is 1 at time $t=0$ and 0 at time $t=\infty$. When integrated over all times, we get a 'relaxation' time for substenance of decoherence as:
\begin{equation}
\tau=\int_{0}^{\infty} dt C(t).
\end{equation}
We carry out an explicit calculation for $\tau$ in both classical and quantum regimes. Using Eq. (A.14) and (43) and the initial density matrix given by Eq. (45), we can calculate $\left(\rho_\mathcal{R}\right)_{LR}(t)$ to obtain:
\begin{equation}
\tau=\frac{\frac{1}{2}I_1+I_2 Im\left(\left(\rho_\mathcal{R}\right)_{LR}(0)\right)}{Im\left(\left(\rho_\mathcal{R}\right)_{LR}(0)\right)},
\end{equation}
where, for the quantum case,
\begin{eqnarray}
I_1=\frac{2\Delta}{(2\Delta)^2+b_{3}^2}, \nonumber \\
I_2=\frac{b_{3}}{(2\Delta)^2+b_{3}^2},
\end{eqnarray}
On the other hand, for the classical telegraph process,
\begin{eqnarray}
I_1&=&\frac{\left(2\Delta\right)\left(-\lambda+k\right)}{(2\Delta)^2+\left(\frac{\lambda+k}{2}\right)^2}+\frac{\left(2\Delta\right)\left(\lambda+k\right)}{(2\Delta)^2+\left(\frac{\lambda-k}{2}\right)^2}, \nonumber \\
I_2&=&\frac{\left(\frac{\lambda+k}{2}\right)\left(-\lambda+k\right)}{(2\Delta)^2+\left(\frac{\lambda+k}{2}\right)^2}+\frac{\left(\frac{\lambda-k}{2}\right)\left(-\lambda+k\right)}{(2\Delta)^2+\left(\frac{\lambda-k}{2}\right)^2}, \nonumber \\
k&=&\sqrt{\left(\lambda^2-16\zeta_{\Delta}^2\right)}.
\end{eqnarray}
where $b_{3}$ is given by the $z=0$ limit of Eq. (42). It is evident from Fig. (6) that the partial decoherence is attained more rapidly in the case of classical noise.
\section{Concluding Remarks}
Our main emphasis in this paper has been a relative assessment of quantum and classical noise sources attached to a pair of quantum dots or a qubit. The analysis has been made with the aid of a unified formalism, comprising a resolvant expansion of an averaged time-development operator. The quantum formalism enabled us to provide a microscopic meaning and detailed temperature-dependence to the phenomenologically introduced parameter of the relaxation rate, that appears in the classical case of a telegraphic noise. Our treatment of the quantum noise has included the much studied ohmic model of dissipation that characterizes electron-hole excitations off the Fermi surface, in a metal, as well as non-ohmic dissipation which covers both acoustic and optic phonons$^{14}$. It is eventually in the phonon model of dissipation that the usually assumed exponential temperature dependence of the relaxation rate (of a Telegraph process) is realized, making the comparison between the classical and quantum cases more direct.
In the last subsection of the paper (see IV) we focussed on the important issue of partial decoherence that can be utilized for quantum computation, which has received recent attention$^8$. This case was analyzed when fluctuation is ascribed only to the hopping term, yielding a situation in which the system Hamiltonian commutes with the coupling to the heat bath. Here, the comparison between classical and quantum noises is quite striking--coherence persists over longer time scales for the quantum case. This attribute can be effectively exploited in the context of quantum computation, in which it is essential to be able to retrieve information about the initial quantum state, notwithstanding heat bath-induced effects.
\section{Acknowledgement}
EK thanks the INSPIRE support of the Department of Science and Technology for an MS thesis project that contains the present contribution. SD will like to record his gratitude to Amnon Aharony, Ora Entin-Wohlman and Shmuel Gurvitz for many helpful discussions.
|
\section{Introduction}
Since the early days of QCD \cite{Gros} it is known that the anomalous dimensions \(\gamma(S)\) of twist operators
\(\mathrm{Tr}(\Phi D^S\Phi)\) are related to parton splitting functions, which gives the probability of partons to split into another parton with certain momentum fractions. Nevertheless it is a recent observation that the
coefficient functions at three loop order in perturbation expansion of \(\gamma(S)\) are already determined by the two loop results
\cite{Moch, Moch2, Doks}.
In a large \(S\) expansion the terms organize as \cite{Bass, Becc}
\begin{equation}
\gamma(S)=f\ln(S)+f_c+f_{11}\frac{\ln(S)}{S}+f_{10}\frac{1}{S}+\mathcal{O}(S^{-2}),
\end{equation}
where
\begin{equation}
f_{11}=\frac{1}{2}f^2, \qquad f_{10}=\frac{1}{2}f\,f_c
\end{equation}
are the MVV relations.
In \cite{Bass, Becc} it was argued that similar relations hold also for coefficient functions of the higher order terms with odd
powers in \(S\)
provided the anomalous dimension has the parity preserving asymptotic expansion in the conformal spin
\(s=S+\frac{1}{2}\gamma(S)\):
\begin{equation}\label{eq:parity}
\gamma(S)=\mathrm{f}(S+\frac{1}{2}\gamma)\to g\ln(s)+g_c+\frac{g_1}{s^2}+\frac{g_2}{s^4}+\mathcal{O}(s^{-6}),
\end{equation}
where \(g_i\) are polynomials in \(\ln(s)\).
The parity preserving property implies also the Gribov-Lipatov reciprocity of the splitting functions \cite{Grib, Bass}.
The conjectured AdS/CFT correspondence \cite{Mald, Gubs, Gub2} relates
the anomalous dimension of twist operators in \(\mathcal{N}=4\) \(SU(N)\) SYM at strong coupling to the energy \(E(S)\) of semiclassical spinning string configurations with spin \(S\) in \(AdS_5\times S^5\). For recent reviews of the anomalous dimension
and Gribov-Lipatov reciprocity of twist operators in context of AdS/CFT see \cite{Bec3}.
In this letter we will use AdS/CFT to show that at strong coupling the MVV relations and the parity preserving and reciprocity properties can be derived from asymptotic series solutions of nonlinear ordinary differential equations. We obtain these equations by using
the close relation of the energy-spin relation of the dual spinning string to automorphic functions and Picard-Fuchs equations for complete elliptic integrals.
\section{The folded spinning string}
In the subspace \(AdS_3\) of \(AdS_5\times S^5\) with metric
\begin{equation}
\mathrm{d}s^2=-\cosh^2\rho\mathrm{d}t^2+\mathrm{d}\rho^2+\sinh^2\rho\mathrm{d}\phi^2,
\end{equation}
the folded spinning string can be described by the ansatz \cite{Gub2, Frol}
\begin{equation}\label{eq:RotAnsatz}
t=\kappa\tau,\;\;\phi=\omega\tau,\;\; \rho(\sigma)=\rho(\sigma+2\pi).
\end{equation}
The explicit solution of the equations of motions is given in terms of Jacobi's elliptic functions
\begin{equation}
\cosh^2\rho(\sigma)=\frac{1}{1-k^2}\mathrm{dn}^2(\omega\sigma+\mathbb{K}|k^2),
\end{equation}
and from the periodicity constraints one gets relations between the parameters in the ansatz:
\begin{equation}
\kappa=\frac{2}{\pi}k\mathbb{K},\qquad \omega=\frac{2}{\pi}\mathbb{K}.
\end{equation}
Also the energy and the spin can be parameterized in terms of complete elliptic integrals
\begin{eqnarray}
E&=&\sqrt{\lambda}\frac{2}{\pi}\frac{k}{1-k^2}\mathbb{E},\\
S&=&\sqrt{\lambda}\frac{2}{\pi}\left[\frac{1}{1-k^2}\mathbb{E}-\mathbb{K}\right],
\end{eqnarray}
where the modulus parameter \(k\) is related to the string length \(\rho_0\) as
\begin{equation}
\tanh\rho_0=k.
\end{equation}
In the semiclassical limit
\begin{equation}
\lambda,S,E\to \infty,\qquad \text{with }\, \mathcal{E}=\frac{E}{\sqrt{\lambda}},\qquad \mathcal{S}=\frac{S}{\sqrt{\lambda}}\; \text{ fix}
\end{equation}
the energy can be expanded as
\begin{equation}
E=\sqrt{\lambda}\mathcal{E}+E_1+\mathcal{O}\left(1/\sqrt{\lambda}\right),
\end{equation}
where the classical contribution for \(\mathcal{S}\to\infty\) is given as (with \(\mathcal{\bar S}=8\pi\mathcal{S}\)) \cite{Becc}:
\begin{eqnarray}\label{eq:classicalenergy}
& & \mathcal{E}(\mathcal{S})=\mathcal{S}+\frac{1}{\pi}\left(\ln\mathcal{\bar S}-1\right)+\frac{1}{2\pi^2\mathcal{S}}
\left(\ln\mathcal{\bar S}-1\right)-\nonumber\\
& & -\frac{1}{16\pi^3\mathcal S^2}\left(2\ln^2\mathcal{\bar S}-9\ln\mathcal{\bar S}+5\right)+\\
& &+\frac{1}{48\pi^4\mathcal{S}^3}\left(2\ln^3\mathcal{\bar S}-18\ln^2\mathcal{\bar S}+33\ln\mathcal{\bar S}-14\right)+
\mathcal{O}(\mathcal{S}^{-4}).\nonumber
\end{eqnarray}
In principle one has to insert the inverse function of \(\mathcal{S}\) into \(\mathcal{E}\) to get the expression for \(\mathcal{E}(\mathcal{S})\).
Since this was not possible, so far only this series expression was obtained.
\section{Picard-Fuchs differential equations of elliptic integrals}
In order to understand the modular derivatives of \(\mathcal{E}\) and \(\mathcal{S}\) we will mention some classic
results about Picard-Fuchs equations of complete elliptic integrals. Making the connection to the corresponding
Abelian periods it will then be straightforward to generalize this setting to the general hyperelliptic case \cite{Fuch, Koen}.
The classic canonical form of the complete elliptic integrals are defined by
\begin{eqnarray}
\mathbb{K}&=&\int_0^1\frac{\mathrm{d}x}{\sqrt{(1-x^2)(1-k^2x^2)}},\nonumber\\
\mathbb{E}&=&\int_0^1\frac{\mathrm{d}x\sqrt{1-k^2x^2}}{\sqrt{1-x^2}}.
\end{eqnarray}
In algebraic geometry one considers instead the Abelian periods
\begin{equation}\label{eq:abelperiod}
I_1=\int_0^1\frac{\mathrm{d}z}{\sqrt{P(z)}}\qquad I_2=\int_0^1\frac{\mathrm{d}z\,z}{\sqrt{P(z)}},
\end{equation}
with \( P(z)=z(1-z)(1-k^2z)\). Both expressions are related by (with \(z=x^2\))
\begin{equation}\label{eq:elliptic}
\mathbb{K}=\frac{1}{2}I_1,\qquad \mathbb{E}=\frac{1}{2}I_1-\frac{k^2}{2}I_2.
\end{equation}
Therefore the complete elliptic integrals are related to the Abelian periods by an additive combination.
Although traditionally the dependence on the elliptic modulus \(k\) is not mentioned explicitly one should keep in mind that
the Abelian periods and complete elliptic integrals are still functions of \(k\). It is a nice property of the Abelian periods that
their derivatives with respect to \(k\) can again be expressed as an additive combinations of the original Abelian periods
(with \(k'^2=1-k^2\)):
\begin{equation}
\frac{\mathrm{d}I_1}{\mathrm{d}k}=\frac{k}{k'^2}(I_1-I_2),\;\;\frac{\mathrm{d}I_2}{\mathrm{d}k}=\frac{1}{kk'^2}(I_1-(2-k^2)I_2).
\end{equation}
We avoid the term 'linear combination' since the coefficients in front of the Abelian periods are not independent constants but are again
expressed in terms of the elliptic modulus \(k\).
Using now (\ref{eq:elliptic}) one can immediately write down the modular derivatives for the complete elliptic integral:
\begin{equation}\label{eq:modulderiv}
\frac{\mathrm{d}\mathbb{K}}{\mathrm{d}k}=\frac{1}{kk'^2}(\mathbb{E}-k'^2\mathbb{K}),\qquad
\frac{\mathrm{d}\mathbb{E}}{\mathrm{d}k}=\frac{1}{k}(\mathbb{E}-\mathbb{K}).
\end{equation}
Repeating this procedure with the second derivative one finds the classic result by Legendre that the complete elliptic integral satisfies a second order linear differential equation
\begin{equation}\label{eq:moduldiff}
kk'^2\frac{\mathrm{d}^2\mathbb{K}}{\mathrm{d}k^2}+(1-3k^2)\frac{\mathrm{d}\mathbb{K}}{\mathrm{d}k}-k\mathbb{K}=0.
\end{equation}
This equation is today understood as a special case of Picard-Fuchs equations.
By the substitution \(x=k^2\) the differential equation (\ref{eq:moduldiff}) transforms to a hypergeometric equation,
which gives the well known expression for the complete elliptic integral in terms of hypergeometric functions
\begin{equation}
\mathbb{K}=\frac{\pi}{2}{}_2F_1\left(\frac{1}{2},\frac{1}{2},1;k^2\right).
\end{equation}
\section{The energy-spin function}
We can now take a look at the energy and spin of the folded string from the view point of Abelian periods and their
modular derivatives:
\begin{equation}\label{eq:Cartancharge}
\mathcal{E}=\frac{2}{\pi}\frac{k}{1-k^2}\mathbb{E},\qquad
\mathcal{S}=\frac{2}{\pi}\left[\frac{1}{1-k^2}\mathbb{E}-\mathbb{K}\right].
\end{equation}
Since (\ref{eq:Cartancharge}) are additive combinations of the complete elliptic integrals, they are also additive combinations of the Abelian periods.
And therefore they inherit the property that their modular derivatives can again be expressed not only as additive combination of complete
elliptic integrals or Abelian periods but also in terms of \(\mathcal{E}\) and \(\mathcal{S}\) again:
\begin{equation}\label{eq:Cartanderiv}
\frac{\mathrm{d}\mathcal{E}}{\mathrm{d}k}=\frac{1}{1-k^2}\left[\mathcal{S}+\frac{1}{k}\mathcal{E}\right],\;\;\;
\frac{\mathrm{d}\mathcal{S}}{\mathrm{d}k}=\frac{1}{1-k^2}\left[k\mathcal{S}+\mathcal{E}\right].
\end{equation}
This is not really a surprise but the essential point is: Although we have no analytic expression for \(\mathcal{E}(\mathcal{S})\) we can use now the results (\ref{eq:Cartanderiv})
to obtain a very nice compact expression for its derivative:
\begin{equation}\label{eq:firstderiv}
\frac{\mathrm{d}\mathcal{E}}{\mathrm{d}\mathcal{S}}=\frac{1}{k},
\end{equation}
where the elliptic modulus \(k\) should be understood as function of \(\mathcal{S}\).
On the other hand the elliptic modulus has a representation in terms of theta constants
\begin{equation}
k=\frac{\vartheta_{10}^2(0,\tau)}{\vartheta_{00}^2(0,\tau)},
\end{equation}
which has the property that it is invariant under modular transformations \(\Gamma(2)\) of \(\tau\)\cite{Mumf}, it is an
automorphic function. Further, transformations under the full modular group \(\Gamma\) may be
related to recently observed duality properties of the spinning string \cite{Geor}.
\(\mathcal{E}\) and \(\mathcal{S}\) are a very special combination of the complete elliptic integrals,
since from (\ref{eq:modulderiv}) one can see that e.g. \(\mathrm{d}\mathbb{K}/\mathrm{d}\mathbb{E}\) is not an algebraic expression
of an automorphic function.
We can proceed further and consider the second derivative, given by
\begin{eqnarray}
\frac{\mathrm{d}^2\mathcal{E}}{\mathrm{d}\mathcal{S}^2}&=&\frac{\mathrm{d}}{\mathrm{d}\mathcal{S}}
\left(\frac{\mathrm{d}\mathcal{E}}{\mathrm{d}\mathcal{S}}\right)=
\left(\frac{\mathrm{d}\mathcal{S}}{\mathrm{d}k}\right)^{-1}\frac{\mathrm{d}}{\mathrm{d}k}\left(\frac{1}{k}\right)=\nonumber\\
&=&-\frac{1-k^2}{k^2}\frac{1}{k\mathcal{S}+\mathcal{E}}.
\end{eqnarray}
Using now (\ref{eq:firstderiv}) for \(k\), we obtain a nonlinear second order ordinary differential equation of third degree:
\begin{equation}\label{eq:nonlin}
\left(\mathcal{S}+\mathcal{E}(\mathcal{S})\frac{\mathrm{d}\mathcal{E}}{\mathrm{d}\mathcal{S}}\right)
\frac{\mathrm{d}^2\mathcal{E}}{\mathrm{d}\mathcal{S}^2}+\left(\frac{\mathrm{d}\mathcal{E}}{\mathrm{d}\mathcal{S}}\right)^3-
\frac{\mathrm{d}\mathcal{E}}{\mathrm{d}\mathcal{S}}=0.
\end{equation}
This is the main result of this section. The energy-spin function \(\mathcal{E}(\mathcal{S})\) of the folded string is a solution of (\ref{eq:nonlin}).
Motivated by the previous result (\ref{eq:classicalenergy}) one can use the following asymptotic ansatz for
\(\mathcal{S}\to\infty\) as a solution of (\ref{eq:nonlin}):
\begin{eqnarray}
\mathcal{E}(\mathcal{S})&=&\mathcal{S}+f\ln\mathcal{S}+f_c+\frac{f_{11}\ln\mathcal{S}+f_{10}}{\mathcal{S}}+\nonumber\\
& &+\frac{f_{22}\ln^2\mathcal{S}+f_{21}\ln\mathcal{S}+f_{20}}{\mathcal{S}^2}+\nonumber\\
& &+\frac{f_{33}\ln^3\mathcal{S}+f_{32}\ln^2\mathcal{S}+f_{31}\ln\mathcal{S}+f_{30}}{\mathcal{S}^3}.
\end{eqnarray}
Then one obtains recurrence relations between the coefficients
\begin{eqnarray}\label{eq:recurrence1}
f_{11}&=&\frac{1}{2}f^2,\qquad f_{10}=\frac{1}{2}ff_c,\qquad f_{22}=-\frac{1}{8}f^3, \nonumber\\
f_{21}&=&\frac{1}{16}f^2(5f-4f_c), \nonumber\\
f_{20}&=&\frac{1}{16}f(2f^2+5ff_c-2f_c^2),\nonumber\\
f_{33}&=&\frac{1}{24}f^4,\qquad f_{32}=\frac{1}{8}f^3(-2f+f_c),\nonumber\\
f_{31}&=&\frac{1}{16}f^2(f^2-8ff_c+2f_c^2),\nonumber\\
f_{30}&=&\frac{1}{48}f(3f^3+3f^2f_c-12ff_c^2+2f_c^3).
\end{eqnarray}
Since we have a second order differential equation
all coefficients are determined when \(f\) and \(f_c\) are fixed by
some initial conditions. Apparently these are the MVV relations previously derived by assuming the parity preserving property
(\ref{eq:parity}). We see now that only two free coefficients are enough to determine the solution.
As a check one can convince oneself that the coefficients of the classical energy (\ref{eq:classicalenergy}) indeed satisfy the relations
(\ref{eq:recurrence1}).
\subsection{The weak involution property}
For the function \(\mathcal{S}(\mathcal{E})\), which is the inverse of \(\mathcal{E}(\mathcal{S})\), we have
\begin{equation}
\frac{\mathrm{d}\mathcal{S}}{\mathrm{d}\mathcal{E}}=k,
\end{equation}
from which we can derive the corresponding differential equation
\begin{equation}
\left(\mathcal{E}+\mathcal{S}(\mathcal{E})\frac{\mathrm{d}\mathcal{S}}{\mathrm{d}\mathcal{E}}\right)
\frac{\mathrm{d}^2\mathcal{S}}{\mathrm{d}\mathcal{E}^2}+
\left(\frac{\mathrm{d}\mathcal{S}}{\mathrm{d}\mathcal{E}}\right)^3-\frac{\mathrm{d}\mathcal{S}}{\mathrm{d}\mathcal{E}}=0
\end{equation}
Comparing with (\ref{eq:nonlin}) we see that the equations are invariant under \(\mathcal{S}\to\mathcal{E}\) and \(\mathcal{E}\to\mathcal{S}\).
This means that a function \(f_1(x)\) and also its inverse \(f_2(x)=f_1^{-1}(x)\) are solutions of the same differential equation:
\begin{equation}
f_1(\mathcal{E})=\mathcal{S},\qquad f_2(\mathcal{S})=\mathcal{E},
\end{equation}
\subsection{The reciprocity property}
If the anomalous dimension \(\gamma=\mathcal{E}-\mathcal{S}\) satisfies the parity preserving relation (\ref{eq:parity}) it should be
possible to write it as a function of the generalized spin variable \(s_0=\mathcal{E}+\mathcal{S}\) \cite{Becc}.
It is now easy to check that also the derivative of this function has the modular invariance property:
\begin{equation}
\frac{\mathrm{d}(\mathcal{E}-\mathcal{S})}{\mathrm{d}(\mathcal{E}+\mathcal{S})}=\frac{1-k}{1+k}.
\end{equation}
Along the same line as for \(\mathcal{E}(\mathcal{S})\) we find the differential equation for \(\gamma(s_0)\) as
\begin{equation}
\left(s_0+\gamma(s_0)\frac{\mathrm{d}\gamma}{\mathrm{d}s_0}\right)\frac{\mathrm{d}^2\gamma}{\mathrm{d}s_0^2}-
\left(\frac{\mathrm{d}\gamma}{\mathrm{d}s_0}\right)^3+\frac{\mathrm{d}\gamma}{\mathrm{d}s_0}=0.
\end{equation}
One can make a similar series ansatz for \(\gamma(s_0)\) as for \(\mathcal{E}(\mathcal{S})\):
\begin{eqnarray}
\gamma(s_0)=g_0s_0+g\ln(s_0)+g_c+\sum_{n=1}\frac{a_n(\ln(s_0))}{s_0^{n}},\nonumber\\
a_n(\ln(s_0))=\sum_{m=0}g_{nm}\ln^m(s_0).
\end{eqnarray}
Now we find the following recurrence relations:
\begin{eqnarray}
g_0&=&0,\qquad g_{10}=g_{11}=0,\qquad g_{22}=0,\nonumber\\
g_{21}&=&\frac{1}{4}g^3,\qquad g_{20}=\frac{1}{4}g^2(2g+g_c),\nonumber\\
g_{33}&=&g_{32}=g_{31}=g_{30}=0,\qquad g_{44}=g_{43}=0,\nonumber\\
g_{42}&=&-\frac{1}{8}g^5,\qquad
g_{41}=-\frac{1}{64}g^4(23g+16g_c)\nonumber\\
g_{40}&=&-\frac{1}{128}g^3(35g^2+46gg_c+16g_c^2),\nonumber\\
g_{55}&=&g_{54}=g_{53}=g_{52}=g_{51}=g_{50}=0.
\end{eqnarray}
The function \(\gamma(s_0)\) has all the desired properties \cite{Becc,Bec3}
\begin{itemize}
\item Reciprocity: Only even negative powers of \(s_0\) appear in the expansion,
\item The highest order of \(\ln^n(s_0)\) does not appear in the expansion.
\end{itemize}
\subsection{Short string expansion}
It is also easy to obtain from (\ref{eq:nonlin}) the short string expansion \(\mathcal{S}\to 0\) to arbitrary high order:
\begin{equation}\label{eq:shortexp}
\mathcal{E}=\sqrt{2\mathcal{S}}\sum_{n=0}a_n\mathcal{S}^n,
\end{equation}
with the following coefficients
\begin{eqnarray}
a_1&=&\frac{3}{8}\frac{1}{a_0},\qquad a_2=-\frac{21}{128}\frac{1}{a_0^3},\qquad a_3=\frac{187}{1024}\frac{1}{a_0^5},\nonumber\\
a_4&=&-\frac{9261}{32768}\frac{1}{a_0^7},\qquad a_5=\frac{136245}{262144}\frac{1}{a_0^9}.
\end{eqnarray}
This confirms and extends previous results for \(a_0=1\)\cite{Tirz}. A further numerical comparison of the asymptotic series (\ref{eq:shortexp})
with the exact result (\ref{eq:Cartancharge}) shows good agreement for \(\mathcal{S}<0.4\).
\section{Discussion}
We have shown that the energy-spin relation of the semiclassical folded spinning string has to satisfies a nonlinear second order
differential equation. Using the AdS/CFT correspondence the recurrence relations for the coefficients of asymptotic series ansatz at \(\mathcal{S}\to\infty\) are identified as the MVV relations for the anomalous dimensions of the twist operators in the dual gauge theory.
The analog asymptotic solution of the differential equations for the anomalous dimension expressed in terms of the conformal spin
has the reciprocity property, an expansion only in even negative powers of \(S\). So far this property was an
ad hoc assumption in order
to derive the MVV relations. Now we see it is inherit from the modular properties of the spinning string configuration.
Since the corresponding equations are of second order, there remain two undetermined coefficients, the scaling or cusp functions \(f\) and
the virtual scaling function \(f_c\) \cite{Frey}. Although the differential equation and the recurrence relations are derived from the classical string one can check by using the results of \cite{Bec2} that the relations are also satisfied up to \(\mathcal{O}(S^{-2})\) including the 1-loop
quantum corrections.
In order to generalize our results to other spin configurations we like to point out that it is not necessary to have a closed expression for
the constants of motion. All one needs are the integral expressions in terms of Abelian periods as (\ref{eq:abelperiod}) in the elliptic case. Integral expressions are accessible for a broad range of semiclassical string configuration \cite{Arut}. From these in principle the corresponding Picard-Fuchs equations can be derived in order to find some automorphic invariants.
It would also be interesting to understand the relation of the nonlinear differential equations to certain integral equations of the anomalous dimension, which follow from the asymptotic Bethe ansatz \cite{Beis, Bas2, Fior}.
I thank T. M\aa nsson and A. Tseytlin for useful comments on the manuscript. This work was supported by the G\"oran Gustafsson Foundation.
|
\section{Introduction}
The development of instruments with arrays of 100 to several 1000 bolometers to detect submillimeter and millimeter
radiation has become commonplace in the past decade. For ground-based instruments, the predominant signal in the
recorded data is emission from the atmosphere, particularly that from precipitable water vapor. Other undesirable features
in the recorded data include noise -- features in the data not traceable to incoming photons -- that are introduced via a
number of mechanisms. We can hope to remove sources of noise which are correlated from one detector to at least one
other, though many are common to the whole array or portions within. The removal of these undesirable, correlated features
from the data is a major hurdle in reducing the recorded data into cleaned data that are hopefully dominated by astrophysical
signal and unremovable random noise. In a typical AzTEC \citep{WilsonAztecInstrument} observation, $\sim$90 per cent of
the atmospheric emission is described by the average signal across the array. Thus the primary problem is to ``clean'' the
recorded data of the remaining atmospheric signal at higher moments and other sources of correlated noise without removing
signal to a degree that degrades signal-to-noise.
Techniques which target and remove specific modes from the recorded data are commonly used because they are
typically based upon models which incorporate physical phenomena and an understanding of the instrument.
In particular, linear techniques may be preferred because they have the distributive and scalar multiplicative
properties of linear operators. That is, if one knows how the signal of interest will manifest in the recorded data,
then one can estimate the filtering effect of the cleaning technique -- the ``transfer function'' -- on simulated data which
contain only such a signal, without being compelled to use real or simulated atmospheric signal and other sources of
noise. The final map can be appropriately normalized by this transfer function to produce a result in astrophysical units.
It is found, however, that purely linear techniques often provide unsatisfactory signal-to-noise performance in that they
remove insufficient noise or too much signal, particularly for signals that subtend a significant fraction of the array.
Thus non-linear, sometimes iterative, techniques have been developed to improve detection significance
\citep{Enoch2006, KovacsCRUSH2008, SayersMK2010}. The need for this development can be understood from some
simple properties of real-world instruments without resorting to measuring or modeling the properties of the atmospheric
emission (\textit{e.g.}, \cite{LayHalverson2000, SayersMK2010}) or any other undesirable feature in the recorded data.
An actual instrument employs detectors whose response to sky signal (both atmospheric and astrophysical in origin) may
be a varying function of time owing to, \textit{e.g.}, the nature of the detection mechanism, variation in the subsequent
electronic amplification or changes in the optical properties of the instrument. Though instruments are calibrated at regular
intervals, variations on time scales much shorter than the interval cannot be accounted for in the calibration. The relative
gain between detectors is important because modeling the largest undesirable feature, the atmosphere, requires converting
the recorded data to values proportional to physical units. A small fluctuation, or calibration imprecision, in the relative gain
between detectors can have significant impact because it is multiplied by the large correlated atmospheric signal. Allowing
the relative gains used in atmospheric removal to converge to a set of values independent of the calibration values, as in
\cite{SayersMK2010}, is an example of a non-linear technique because the data themselves are used to measure the
relative gain; \textit{i.e.}, the cleaned data are a function of the recorded data multiplied by the relative gain, which is no
longer independent of the data. Similarly, principal component analysis (PCA) allows the relative gain between detectors
to be determined by the covariance matrix calculated from the recorded data and is thereby non-linear.
To interpret a map produced by a non-linear analysis technique, we still require a transfer function for the signal of
interest. Though non-linear techniques will not, in general, have the distributive and multiplicative properties of
linear operators, the interpretation of the map depends only on the cleaned signal of interest being linearly
proportional to the input recorded data. Though it is non-linear, the technique described in \cite{KovacsCRUSH2008}
retains the transfer function estimation advantages of linear techniques because the data is modeled explicitly as the
summation of specified noise modes and astrophysical signal. The PCA technique, described further in Sec. \ref{sec:pca},
``adaptively'' uses the recorded data to identify the modes to be removed. Thus, calculation of a PCA transfer function must
ultimately rely on the recorded data themselves.
We describe herein applications of this approach to PCA on data from the AzTEC instrument and make comparisons to
approximations to the full non-linear problem. The resulting photometry is applied to the AzTEC data and revised versions of
previously released catalogues are presented. No changes to the correlated noise removal technique itself are made.
\section{Principal component analysis}
\label{sec:pca}
Principal component analysis is a popular technique for identifying the moments that describe the variance
in data without relying on having measured those moments in their natural coordinate frame. As applied to bolometric
arrays, the recorded data from $N$ bolometers with $N_\text{samples}$ each are decomposed into orthonormal eigenfunctions
by the standard eigen-decomposition technique (\textit{e.g.}, \cite[chap. 7]{linalg}). The eigenfunctions can be rank-ordered in
eigenvalue and thus also by their contribution to the variance in the recorded timestream. The largest eigenfunctions are
then supposed to have their origins in atmospheric signal as well as other strong correlations in the instrument. Since these
modes are determined by the data themselves, the process as a whole is non-linear, even though eigen-decomposition and
eigenfunction removal are individually linear.
The exact choice of the number of eigenfunctions to remove from the recorded data is somewhat arbitrary. It is empirically observed
that a logarithmic distribution of the eigenvalues will contain a large cluster of low eigenvalues\footnotemark~along with a
number of widely distributed larger eigenvalues. In the AzTEC pipeline, the width of the low-eigenvalue cluster is used to
calculate the number of eigenfunctions to remove. Though other cuts could be made, this choice allows a simple parameter, a
multiplier on the eigenvalue distribution width, to control the cleaning process. Eigenfunctions are removed from the data until no
further modes exist outside a region defined by the multiplier times the distribution width. It is observed that the
number of modes removed is unaffected upon addition of simulated sources of typical flux densities (a few to 10 mJy at 1.1 mm)
to the recorded data. The particular value of 2.5 for this multiplier has been empirically found to roughly maximize signal-to-noise for point sources. Typically 5-15 modes are removed from the data. The details of this technique are described in
\cite{ScottCOSMOS2008}.
\footnotetext{Simulated data that contain only random noise have such a feature, suggesting its origin.}
The advantage of the PCA technique is that the largest correlations are adaptively identified and removed. This removes
large correlated features that may be easily described by physically motivated models as well as features that do not lend
themselves to modeling. An example of the latter might be electromagnetic interference that couples to detectors
with a strength that varies with time or is found only in a subset of the data. By automatically removing these features, the observer's
time can be dedicated to interpretation of the interesting signals. However, the transfer function of PCA on signals is
dependent on what modes are adaptively identified and removed. The transfer function
estimation technique described in \cite{ScottCOSMOS2008} (and used in subsequent AzTEC publications) is
a linear approximation to the PCA cleaning operator because it assumes that the operator -- which identifies high power
modes that are correlated between detectors -- is unaffected by the presence of a simulated faint source. We might expect this
to be true because point sources subtend an angle that is small compared to the bolometer spacing and also because the
typical signal they contribute is small compared to that from the atmosphere, but it is not empirically observed to be true.
In fact, the eigenfunction spectrum at large eigenvalue is systematically affected by the addition of simulated sources of
typical flux densities to the recorded data. Comparing the eigenvectors\footnotemark~calculated from the recorded and source-added
data, we find that the cleaning operator components vary at the several percent level (with roughly equal fluctuations
upwards and downwards) for the largest eigenvalue eigenfunction. Because the largest eigenfunction is essentially the average
atmospheric signal across the array \citep{SayersThesis}, a several percent effect in the operator can be significant compared to
the source flux we intend to measure.
\footnotetext{The eigenvectors are an $N \times N$ matrix that transform the recorded data into the orthonormal basis of
eigenfunctions. Any changes in the eigenvector components corresponding to large eigenvalues are reflected in the PCA
cleaning operator that removes eigenfunctions from the data.}
This observation calls into question the accuracy of a linear approximation to the PCA cleaning operator. A full, non-linear
simulation of the cleaning operator is therefore necessary. As will be shown, the linear approximation results in a systematic
overestimation of the transfer function and an underestimate of the flux and noise present in the optimally filtered map.
\section{Simulation of the PCA Transfer Function}
If we clean and map data with simulated sources and difference them from unfiltered maps produced from the recorded
data, we can see how point sources are affected by PCA cleaning. For the purposes of this analysis, we have chosen
to apply this technique to several previously published AzTEC deep-field observations (described in greater detail in Sec.
\ref{sec:catalogues}) of size varying from $\sim$0.1 to 0.4 square degrees.
Prior to performing the simulation, we produce an initial filtered map using the linear prescription in \cite{ScottCOSMOS2008}. This
map can be used to estimate the final noise
level and to calculate a region of the map that will be used in subsequent analysis. Typically this region is defined by including
pixels whose noise-weighted time coverage is 50-70 per cent of the maximum coverage in the map. Simulated source locations
are chosen to be more than 60\arcsec~away from sources detected with significance greater than 3.5 in the selected region of the
initial map. Likewise, all simulated sources are chosen to have a flux density equal to 10 times the average noise level of the
selected region in the initial map. For each field, we insert 3 simulated sources per 0.05 square degrees with a maximum of 8.
These 3 choices ensure that the transfer function is measured on simulated sources that are comparable to typically observed
sources but are not affected by the true bright sources and do not themselves strongly affect the data.
The noise realisations in the recorded and source-added maps are similar but not precisely the same because the distributive
property does not hold for non-linear operators. Thus, simple differencing of the maps is insufficient to produce
a proper transfer function because it will include residual noise on pixel scales that is not a reflection of the actual effect of
cleaning on a point source signal. This residual is typically small compared to the noise level in the map; however, its use in
an optimal filter would wrongly couple noise into our estimate of source flux and detection significance. We mitigate this effect
through 3 additional steps: (1) stacking the difference map at the centre of the simulated sources and normalising by the
known inserted flux, (2) rotationally averaging the stacked signal, and (3) tapering the stacked signal at a distance 4 times the
FWHM from the beam centre. These steps must ultimately be justified \textit{a posteriori} -- do they produce a transfer function that
works? However, rotational averaging can be justified \textit{a priori} through an understanding of the AzTEC observing strategy.
AzTEC maps are produced by co-adding many individual maps (typically 60 or greater) taken at many elevations. Each individual
scan, whether raster or Lissajous, is performed in azimuth and elevation while tracking a fixed centre point. The software tracks
position angle of the beam and can detect when pixels are weakly cross-linked. Given this observing strategy, we expect that
the point source transfer function should exhibit significant cylindrical symmetry. Likewise, tapering the signal at the edges is
justified because any measured difference is unlikely to be physical in origin.
In Fig. \ref{fig:stacking}, we show a cut in elevation through the transfer function estimates for the previously selected field
resulting from the linear approximation, differencing/stacking, and differencing/stacking with the extra steps noted above. The
transfer function for each field will be slightly different owing to differing observing conditions and the non-linear nature of PCA
cleaning. This transfer function is representative of typical values seen for observations from the Atacama Submillimeter
Telescope Experiment (ASTE). It is seen that the linear approximation overestimates the peak signal and underestimates the
negative sidelobes that result from the effective high-pass filter of the cleaning operator. The lower peak value and larger
sidelobes can be understood as accounting for the effect of the source itself on the atmospheric model; the cleaning operator
mistakenly includes some source flux in its atmospheric removal thus reducing the peak signal and increasing the sidelobes
(which result in part from the source's contribution to the array average signal). Differencing and stacking simulated sources
results in a more accurate transfer function estimation, albeit with imperfect differencing of noise. This is effectively resolved by
rotationally averaging and tapering the map far from the source centre. The process was repeated many times using sources
at varying locations and produces a stable result, as indicated by the small scatter in values around the particular realisation
presented. Furthermore, this analysis was reproduced using 3 simulated sources of varying brightness at fixed locations
(Fig \ref{fig:intensity}). It is seen that the primary impact of non-linearity on the transfer function is to make the negative
sidelobes shallower as source brightness increases. The impact on measured flux is negligible for sources of typical detection
significance.
\begin{figure*}
\includegraphics[height=0.95\textwidth, angle=-90]{fig_kernel_comparison}
\caption{Comparison of the various techniques to estimate the PCA cleaning point source transfer function for a set of 43
observations of a single ASTE science field. See text for full discussion of interpretation. The grey shaded region indicates the
envelope of 20 calculations of the non-linear smoothed kernel using different locations for the simulated sources.}
\label{fig:stacking}
\end{figure*}
\begin{figure*}
\includegraphics[width=0.95\textwidth]{fig_kernel_intensity_dependence}
\caption{(a) Comparison of transfer functions calculated using 3 simulated sources of varying brightness at fixed locations. (b)
A zoomed portion of (a) showing that the primary impact of non-linearity is to make the negative sidelobes shallower as source
brightness increases. (c) The change in measured flux (relative to peak) for $20\sigma$ and $5\sigma$ sources when they are
optimally filtered using the the standard transfer function derived from $10\sigma$ sources. The peak flux and all other pixels
are shifted by $<1.5\%$ of the peak flux and therefore the transfer function does not introduce significant systematic error for
sources of typical detection significance.}
\label{fig:intensity}
\end{figure*}
The ultimate test of the revised transfer function is whether it succeeds in producing the correct flux when analyzing
data with simulated sources reduced blindly. 23 simulated maps were produced in each of which were inserted 4 simulated
sources at varying locations far from resolved sources with fluxes ranging from 3 to 20mJy. This spans a detection significance
range of $\sim$$4-30\sigma$. The sources were placed at the centre of 3\arcsec~pixels in the portion of the map with sufficient
and uniform coverage to be used for selecting true astrophysical sources. This simulation also tests for any impact that a
moderate increase in source density may have upon the transfer function as 7 simulated sources will ultimately be inserted into
the map (4 ``test'' sources whose flux we intend to measure and 3 sources whose sole purpose is to measure the transfer function).
The maps were then optimally filtered using the revised transfer function estimate and the detected flux and estimated noise at the
known source location was compared to the known input flux (Fig. \ref{fig:inout}). The input and observed fluxes are found to be
consistent; there is a small negative offset that is consistent with the mean value (-0.24 mJy) of the pixels at the chosen input
locations. Because the map has an overall mean of zero and we have chosen locations that are far from bright, positive sources
of flux, it is reasonable to find a small, negative offset. The absence of systematic effects from source location (Fig. \ref{fig:stacking})
or source flux density (Fig. \ref{fig:inout}) may be an indication that, although the transfer function must be varying as a function of
time (the number of eigenfunctions removed from each chunk of data is not constant), it varies more slowly than the time taken to
cover the useful coverage region of the maps. Thus the variations are captured equally well by any simulated source within this
region.
\begin{figure}
\includegraphics[width=0.95\columnwidth]{fig_inout_comparison}
\caption{A comparison of the observed flux in optimally filtered maps at the locations where simulated sources of known flux
have been inserted. The transfer function is consistent with unity and has a small, negative offset which can be explained by observing
that the mean value of the chosen pixel locations was -0.24 mJy in the recorded map. The best-fit line,
$y = -0.3 \pm 0.2 + (1.015\pm0.018)x$, is shown in red.}
\label{fig:inout}
\end{figure}
\begin{figure*}
\includegraphics[height=0.95\textwidth, angle=-90]{fig_kernel_comparison_all_fields}
\caption{Comparison of the linear approximation (dashed) and simulated source (solid) techniques to estimate the PCA cleaning
point source transfer function for previously published AzTEC catalogues. It is observed that the impact is greater for the 4 JCMT
fields than for the ASTE field shown as well as in internal analysis for ASTE fields not yet published. This may be due to the greater
atmospheric fluctuations at the JCMT site.}
\label{fig:allfields}
\end{figure*}
\section{Revised catalogues}
\label{sec:catalogues}
When this technique is applied to other fields (Fig. \ref{fig:allfields}), we observe, for AzTEC data taken while installed at the
ASTE, that the revised transfer function corrects both signal and noise by $\sim$+10 per cent, in each field. Similarly, we find a
correction of +15-25 per cent in the fields observed while AzTEC was installed at the James Clerk Maxwell Telescope (JCMT).
This is consistent with the notion that some of the point source signal is present in the largest eigenfunctions removed by PCA.
The larger impact for the JCMT data may be taken as a sign that the non-linear nature of PCA is more greatly affected by the
worse observing conditions at Mauna Kea as compared to the Atacama Desert in Chile. We wish to emphasize that,
for each observation, simulating the signal of interest and directly observing the impact of PCA or another non-linear technique
is a surer approach than building expectations based on prior results.
Several previous publications have released point source catalogues from the AzTEC instrument while it was installed at the
JCMT and the ASTE. These catalogues are reproduced below using the revised photometry, along with deboosted fluxes
calculated from a forthcoming number counts analysis to be presented in Scott, \textit{et al}., submitted\footnote{The best-fit parameters for the observed blank-field number counts were found to be $N_\text{3mJy} = 231 \unit{mJy}^{-1} \cdot \unit{deg}^{-2}$ and $S' = 1.84 \unit{mJy}$ while fixing $\alpha \equiv -2$ and following Eqns. 2 and 3 in \cite{AustermannSHADES2010}.}.
The AzTEC deboosting algorithm \citep{AustermannCOSMOS2009} accounts for the fact that, for a source population
that declines steeply with flux, any given source is more likely to be a relatively plentiful dim source ``boosted'' upward by
noise than a rare bright source on
top of a negative noise fluctuation. The algorithm makes the assumption that the flux in a given pixel is emitted from a single
source. This assumption may be invalid owing to source confusion from the finite AzTEC beam, however it has been shown
to yield results consistent with a parametric frequentist ``$P(d)$'' approach \citep{PereraGOODSN2008} that attempts to account
for confusion through comparison of the recorded map to simulations of noise and source populations convolved with the beam. The
number counts analysis incorporates the revised photometry and is based upon measurements in all blank fields, including
those below, surveyed by the AzTEC instrument. Because the effect of deboosting is a slowly changing function of the sky
model prior assumed, the deboosted values given below are reasonable even if the inclusion of other fields introduces a
potential bias to the sky model. The deboosted fluxes also include correction for a bias introduced by searching for a signal
peak in the presence of noise. The number counts analyses previously presented will be affected by the change in photometry
in 2 ways. First, the estimated counts at a given brightness value are now appropriate for a slightly increased source brightness.
Second, an integral step in computing number counts is dividing by the estimated completeness\footnotemark~for a given flux
in the map. Because the estimated error is higher under the revised transfer function, any given flux value will be less complete;
this effect will be more pronounced at moderate signal-to-noise ($\sim$3-5) where the completeness is rapidly dropping from unity
toward zero.
\footnotetext{The likelihood of detecting a source of a particular brightness given the sources of noise present in the map.}
The details of the analysis for each field do not differ appreciably from that previously performed, except for the change in
photometry. We provide references to these analyses for the interested reader. We present the catalogues using the
significance and spatial cuts applied by each catalogue's respective author. In each case the expected
false detection rate is estimated using the defined cuts and did not change greatly from that using the transfer function
derived from the linear approximation. Source names and numeric identifiers
for sources which were previously detected are retained; however a common format has been chosen: ``AzTEC/field\#'', where
`\#' indicates order of discovery rather than strictly being based upon detection significance in these revised catalogues. Thus newly
discovered sources with higher significance than a previously discovered source will appear at the end of these catalogs and a slight
shuffling (in significance) of previously discovered sources will occur for the reasons discussed above.
\subsection{COSMOS / JCMT}
The COSMOS survey \citep{ScottCOSMOS2008, AustermannCOSMOS2009} was undertaken by the AzTEC instrument in 2005
while it was installed at the JCMT. A revised catalogue is shown in Table \ref{tbl:cosmos}. A spatial cut (0.15 deg$^2$) was applied
by taking only pixels within the map whose weighting (a combination of noise in the data and amount of time spent observing that
pixel) are greater than 75 per cent of the map's characteristic value (roughly the maximum). A significance cut is applied by taking only
sources whose signal to noise ratio are greater than 3.5. We have conservatively over-estimated the false detection rate to be 25 per cent
($\sim$12 sources) using the number of sources ``detected'' in pure noise realisations of the field. This choice is conservative because
faint sources contribute to flux in nearly every pixel in the map and therefore the likelihood of finding a source at any pixel is greater than
it otherwise would be \citep{PereraGOODSN2008,ScottGOODSS2010}. The difference between this technique and one that attempts
to account for this effect through simulations of source populations can be a factor of 2 or greater.
\begin{table*}
\caption{The AzTEC point source catalogue for the COSMOS field as observed from the JCMT.}
\begin{tabular}{|l|l|c|c|c|c|c|c|c|}
\hline \hline
Source ID & Nickname & S/N & $S_{1.1\text{mm}}$ & $S^\text{corrected}_{1.1\text{mm}}$ & $P(<0)$ & Flux & Noise & $\theta$\\
& & & [mJy] & [mJy] & & increase & increase & \\
\hline
\input{cosmos_table1}
\hline
\multicolumn{9}{l}{The columns are as follows: (1) AzTEC source name, including RA and declination based on centroid position; (2) nickname;}\\
\multicolumn{9}{l}{(3) signal-to-noise of the detection; (4) measured $1100 \umu$m flux density and error; (5) flux density and 68 per cent confidence}\\
\multicolumn{9}{l}{interval of the deboosted flux density, including corrections for the bias to peak locations in the map; (6) probability that the}\\
\multicolumn{9}{l}{source will deboost to $S<0$ assuming the number counts prior based on all AzTEC measurements; (7,8) the relative increase}\\
\multicolumn{9}{l}{in flux and noise estimate for each source if it was detected in the previously release catalogue; (9) change in location of the}\\
\multicolumn{9}{l}{centroided source position if it was detected in both catalogs. ($^a$) indicates a source passed a significance test in the original}\\
\multicolumn{9}{l}{catalog, but not the same test in the new catalog. ($^b$) indicates a source passed a significance test in the new catalog, but not}\\
\multicolumn{9}{l}{the same test in the original catalog. In each case, an estimate for the missing quantity is made from the nearest pixel in the}\\
\multicolumn{9}{l}{map in which the test did not succeed.}
\end{tabular}
\label{tbl:cosmos}
\end{table*}
\subsection{GOODS North / JCMT}
The GOODS North field is commonly observed at many wavelengths. A revision of the catalogue presented in
\cite{PereraGOODSN2008} is shown in Table \ref{tbl:goodsn}. Similar cuts are taken at the 70 per cent coverage region (0.07 deg$^2$)
and detection significances above 3.5. The false detection rate was estimated using pure noise maps to be 13 per cent ($\sim$4-5 sources).
\begin{table*}
\caption{The AzTEC point source catalogue for the GOODS North field.}
\begin{tabular}{|l|l|c|c|c|c|c|c|c|}
\hline \hline
Source ID & Nickname & S/N & $S_{1.1\text{mm}}$ & $S^\text{corrected}_{1.1\text{mm}}$ & $P(<0)$ & Flux & Noise & $\theta$\\
& & & [mJy] & [mJy] & & increase & increase & \\
\hline
\input{goodsn_table1}
\hline
\multicolumn{9}{l}{Columns are as described in Table \ref{tbl:cosmos}.}
\end{tabular}
\label{tbl:goodsn}
\end{table*}
\subsection{Lockman Hole / JCMT}
The Lockman Hole survey \citep{AustermannSHADES2010} was undertaken by the AzTEC instrument in 2005 while it was
installed at the JCMT and formed part of the 1.1mm follow-up to the SCUBA/SHADES project. A revised catalogue is shown in
Table \ref{tbl:lh}. The 50 per cent coverage region (0.31 deg$^2$) was selected as a spatial cut, but a different significance cut was used.
Deboosting can also be used as a proxy for whether a source is likely to be real. If the likelihood of deboosting to 0 flux is significant,
then that source can be excluded from the catalogue. Only sources with less than 10 per cent likelihood of deboosting to 0 flux are taken.
The false detection rate using these cuts was estimated by the technique described in \cite{PereraGOODSN2008,ScottGOODSS2010}.
In this technique, false detection rates are estimated by fully simulating maps using noise estimates and a signal estimate using the
number counts used in deboosting measured fluxes. Using this technique, the false detection rate was estimated to be 20 per cent
($\sim$20 sources).
\begin{table*}
\caption{The AzTEC point source catalogue for the Lockman Hole field.}
\begin{tabular}{|l|l|c|c|c|c|c|c|c|}
\hline \hline
Source ID & Nickname & S/N & $S_{1.1\text{mm}}$ & $S^\text{corrected}_{1.1\text{mm}}$ & $P(<0)$ & Flux & Noise & $\theta$\\
& & & [mJy] & [mJy] & & increase & increase & \\
\hline
\input{lh05_table1}
\hline
\multicolumn{9}{l}{Columns are as described in Table \ref{tbl:cosmos}.}
\end{tabular}
\label{tbl:lh}
\end{table*}
\begin{table*}
\contcaption{}
\begin{tabular}{|l|l|c|c|c|c|c|c|c|}
\hline \hline
Source ID & Nickname & S/N & $S_{1.1\text{mm}}$ & $S^\text{corrected}_{1.1\text{mm}}$ & $P(<0)$ & Flux & Noise & $\theta$\\
& & & [mJy] & [mJy] & & increase & increase & \\
\hline
\input{lh05_table2}
\hline
\multicolumn{9}{l}{Columns are as described in Table \ref{tbl:cosmos}.}
\end{tabular}
\label{tbl:lh}
\end{table*}
\subsection{Subaru XMM-Newton Deep Field / JCMT}
The Subaru XMM-Newton Deep Field (SXDF) was also surveyed as part of the SHADES followup project. A revised
catalogue is shown in Table \ref{tbl:sxdf}. The same cuts as in the Lockman Hole field are applied with a resulting survey area
of 0.37 deg$^2$. Using the simulated map technique, the false detection rate was estimated to be 25 per cent ($\sim$16 sources).
The near doubling of the number of detected sources under the revised kernel is an artefact of using the deboosted flux values
and a different number counts model as a catalogue cut in a population of sources whose numbers grow rapidly with
declining flux.
\begin{table*}
\caption{The AzTEC point source catalogue for the Subaru XMM-Newton Deep Field.}
\begin{tabular}{|l|l|c|c|c|c|c|c|c|}
\hline \hline
Source ID & Nickname & S/N & $S_{1.1\text{mm}}$ & $S^\text{corrected}_{1.1\text{mm}}$ & $P(<0)$ & Flux & Noise & $\theta$\\
& & & [mJy] & [mJy] & & increase & increase & \\
\hline
\input{sxdf_table1}
\hline
\multicolumn{9}{l}{Columns are as described in Table \ref{tbl:cosmos}.}
\end{tabular}
\label{tbl:sxdf}
\end{table*}
\begin{table*}
\contcaption{}
\begin{tabular}{|l|l|c|c|c|c|c|c|c|}
\hline \hline
Source ID & Nickname & S/N & $S_{1.1\text{mm}}$ & $S^\text{corrected}_{1.1\text{mm}}$ & $P(<0)$ & Flux & Noise & $\theta$\\
& & & [mJy] & [mJy] & & increase & increase & \\
\hline
\input{sxdf_table2}
\hline
\multicolumn{9}{l}{Columns are as described in Table \ref{tbl:cosmos}.}
\end{tabular}
\label{tbl:lh}
\end{table*}
\subsection{GOODS South / ASTE}
The GOODS South field is another commonly observed field and among the first chosen for AzTEC when it was moved from
the JCMT to the ASTE in 2007. A revision of the confusion-limited catalogue presented in \cite{ScottGOODSS2010} is shown in
Table \ref{tbl:goodss}. Sources of significance greater than 3.5 in the 50 per cent coverage region (0.08 deg$^2$) are presented.
Using the same false detection estimate technique as was used for the COSMOS field, we estimate a false detection rate of
6 per cent ($\sim$3 sources). Many of the sources appear to be somewhat extended, a sign of the highly confused nature of the
map. Notably, we present the 2nd most significant detection under the assumptions that the flux is the result of a single source or,
alternatively, from two nearby sources. Other sources at modest signal-to-noise do not provide sufficient constraints to multiple
source models. 8 new sources are found in the revised catalogue, many from regions of the map which appear extended.
\begin{table*}
\caption{The AzTEC point source catalogue for the GOODS South field.}
\begin{tabular}{|l|l|c|c|c|c|c|c|c|}
\hline \hline
Source ID & Nickname & S/N & $S_{1.1\text{mm}}$ & $S^\text{corrected}_{1.1\text{mm}}$ & $P(<0)$ & Flux & Noise & $\theta$\\
& & & [mJy] & [mJy] & & increase & increase & \\
\hline
\input{goodss_table1}
\hline
\multicolumn{9}{l}{Columns are as described in Table \ref{tbl:cosmos}. ($^c$) indicates a source found in the extended, lower coverage regions of the map.}
\label{tbl:goodss}
\end{tabular}
\end{table*}
\subsection{Observations}
The detection significance of any given source may change owing to the greater accuracy of flux and noise estimation using the
revised photometry technique. Nonetheless, viewing the catalogues as a whole, detection significance is seen not to be greatly
impacted because the photometry affects both signal and noise. Likewise, the false detection rates in the fields were unchanged.
The vast majority of sources that passed a significance test in a prior catalogue pass it again in the revised catalogue. In fact, the
increased number of sources in the majority of the revised catalogues suggest a slight systematic upward shift in detection
significance; the exponentially increasing number of dim sources in these fields allow a slight shift to increase considerably the
number of detected sources.
\section{Future Work}
The reviewer suggested a potentially simpler technique for estimating the transfer function that would not require the noise
mitigating steps described. Rather than calculating the transfer function by differencing a source-added map from the
recorded map, one would calculate the eigenvectors from the source-added timestreams and remove them from source-only
timestreams. The eigenfunctions identified by PCA cleaning would have the proper impact of sources included, but the transfer
function estimated would not include the small difference in noise realisations between the two maps. This should produce a
result identical to the analysis shown in a much simpler fashion. This manuscript documents the technique used in
several AzTEC publications in 2011-2012 \citep{Humphrey2011, Yun2011, Aretxaga2011, Kim2012}. If the technique is revised further
to include this suggestion or others, it will be documented in subsequent publications.
\section{Conclusions}
A general approach to estimating the transfer function of non-linear techniques has been described and applied to the
specific case of principal component analysis (PCA). Simulations support the accuracy of the results and that
PCA has a transfer function which is effectively linear for point sources of typical detection significance. The resulting transfer
function has been used to correct the catalogue values for the flux, location and significance of point sources in existing AzTEC
maps. Mean source detection significance is not strongly impacted by the photometry correction and may be slightly enhanced.
\section{Acknowledgments}
This work has been made possible by generous support from the Kavli Foundation and the Gordon and Betty Moore Foundation.
Additional support for this analysis has been provided in part by National Science Foundation grants 0540852, 0838222 and
0907952. KSS is supported by the National Radio Astronomy Observatory, which is a facility of the National Science Foundation
operated under cooperative agreement by Associated Universities, Inc. We would like to acknowledge the assistance of Jack Sayers
and Stephan Meyer for useful comments on the analysis presented herein. We also thank the observatory staff of the JCMT and
ASTE who made these observations possible.
\bibliographystyle{mn2e}
|
\section{Effective Hamiltonian for a curved thin film}
|
\section{Introduction}
The Laplacian in tubular domains has been studied in various situations
\cite{Bou, Cr1, DE, Kre}.
A common tubular region~$\Omega$ is as follows:
let
$I \subseteq \mathbb R$ be an interval of $\mathbb R$,
$r:I \to \mathbb R^3$ a curve in $\mathbb R^3$, parametrized by its arc length~$s$,
and
$k(s)$ and $\tau(s)$ denote its curvature and torsion at the point $s
\in I$, respectively.
Let $S$ be an open, bounded, simply connected and nonempty subset of
$\mathbb R^2$.
Move
the region~$S$
along~$r(s)$ and at each point~$s$ allow the region to rotate by an
angle~$\alpha(s)$ (see details in Section~\ref{gddh}).
A problem of interest is the description of the spectral properties of the Laplacian in such
tubes and weakly effective operators (see the definition just after Theorem~\ref{mainTeor}) when the region~$\Omega$ is ``squeezed'' to the curve~$r(s)$, that is, one considers the
sequence of tubes~$\Omega_\varepsilon$
generated by the cross section~$\varepsilon S$ and analyze the limit~$\varepsilon \to 0$.
Let $-\Delta_\varepsilon$ be the Dirichlet Laplacian in
$\Omega_\varepsilon$.
For bounded tubes, i.e., when $I$ is a bounded interval of $\mathbb R$,
the spectrum of $-\Delta_\varepsilon$ is purely discrete because in
this case
its
resolvent is compact.
In \cite{Bou} it was analyzed the convergence of the eigenvalues
$\{\lambda_i^\varepsilon: i \in \mathbb N\}$
as $\varepsilon \to 0$ and
shown that
$$\lambda_i^\varepsilon = \frac{\lambda_0}{\varepsilon^2} +
\mu_i^\varepsilon, \qquad \mu_i^\varepsilon \to \mu_i,$$
where $\lambda_0$ is the first, i.e., the lowest, eigenvalue of the Laplacian in the Sobolev
space
${\mathcal H}_0^1(S)$, and $\mu_i$ are the eigenvalues of the
one-dimensional operator
\begin{equation}\label{bbb8}
w(s)\;\mapsto \; - w''(s) + \left[ C(S) (\tau(s) + \alpha'(s)) -
\frac{k(s)^2}{4} \right] w(s),
\end{equation} acting in~$\mathrm L^2(I)$.
Here $C(S)$ is a nonnegative number depending only on the transverse
region~$S$~\cite{Bou}.
Note that this effective operator explicitly depends on the geometric
shape of the reference curve~$r(s)$ (and so of the tube).
An interesting problem is to know if there exists a similar result about
convergence of
eigenvalues for unbounded tubes.
For such tubes, in~\cite{Kre} it is shown that
if $(\tau + \alpha')(s) = 0$ and $k(s) \neq 0$, then the discrete
spectrum is nonempty, whereas
if $(\tau + \alpha')(s) \neq 0$ and $k(s) = 0$, then the discrete spectrum is
empty.
In~\cite{Cr1}, by using
$\Gamma$-convergence in case of unbounded tubes, a strong resolvent convergence
was proven and the same action \eqref{bbb8} for the respective effective
operator (now acting in $\mathrm L^2(\mathbb R)$) was found as~$\varepsilon \to
0$.
The Dirichlet Laplacian in strips of $\mathbb R^2$ has been studied in
many works \cite{BF,Frie, Sol,Krej2}. For the case of the constraints of planar motion to curves there are results about the limit operator in
\cite{DellAntTen,ACF}, and the effective potential is written in terms of the curvature.
The main
novelty, when we pass from planar domain to tubes in~$\mathbb R^3$, as considered in \cite{Bou,Cr1,Kre, GJ}, is the additional presence of torsion and twisting (i.e., a nonzero $\tau(s) + \alpha'(s)$) in the
effective potential, since the case of untwisted tubes has also been previously studied (see, for instance, \cite{DE,CDF,ClaBra,Ekk,FreitasKrej,KS}).
In \cite{Frie,Sol} the authors consider a family of deformed strips
\[
\left\{(s, y) \in \mathbb R^2: s \in J, \,0 < y < \varepsilon h(s) \right\},
\]
where $J = [-a, b]$, $0< a, b \leq \infty$, and
$h(s) > 0$ is a continuous function satisfying:
\begin{itemize}
\item[(i)] $h(s)$ is a $\mathrm C^1$ function in $J\setminus\{0\}$ and $\|h'/h\|_\infty<\infty$;
\item[(ii)] near the origin $h$ behaves as
\begin{equation}\label{vvv1}
h(s) = M - s^2 + O( |s|^3), \qquad M > 0,
\end{equation}
and $s=0$ is a single point of global maximum for~$h$;
\item[(iii)] in case $I=\mathbb R$ it is assumed that $ \limsup_{|s| \to \infty} h(s) < M$.
\end{itemize}
In what follows we assume that $h$ satisfies the above conditions.
It was shown \cite{Frie,Sol} that, for $\varepsilon$ small enough, the discrete spectrum
of the Laplacian is
always nonempty and the eigenvalues
$\lambda_j(\varepsilon)$ have the following behavior
$$
\mu_j = \lim_{\varepsilon \to 0}\, \varepsilon \left(
\lambda_j(\varepsilon) -
\frac{\pi^2}{\varepsilon^2 M^2} \right),$$
where $\mu_j$ are the eigenvalues of the operator in
$\mathrm L^2(\mathbb R)$ (it acts on a subspace of~$\mathrm L^2(\mathbb R)$,
independently if the interval~$I$ is bounded or not)
given by
$$(T w)(s) = - w''(s) + 2 \frac{\pi^2 }{M^{3}} \; s^2 \; w(s),$$
so that we say that~$T$ is a {\em weakly effective operator} (WEO) in such situation.
In this work we show that these results hold in a more
general setting.
We consider a sequence of tubes~$\Omega_\varepsilon$ in the
space~$\mathbb R^3$, as presented at the beginning of this
introduction, but we deform them by multiplying their cross sections by the above function~$h(s)$.
Here the tubes may be bounded or not.
Then we analyze the asymptotic behavior of eigenvalues and the weakly effective operators in the limit~$\varepsilon\to0$.
The situation here differs from \cite{Frie,Sol}, since besides the
different dimensions (we consider regions in $3-$dimensional space), the reference curves defining our tubes are allowed to
have nontrivial curvatures and torsion.
These tubes, which we shall call {\em deformed tubes}, will generically be
denoted by~$\Lambda_\varepsilon$ (see details in Section~\ref{gddh}).
Our main goal is to study how curvature and torsion of
the reference curve, together with the deforming function $h$, influence the WEO and eigenvalues as~$\varepsilon \to 0$. To this end, we introduce some notation right now.
Recall that~$\lambda_0$ is the lowest eigenvalue of the negative
Laplacian with Dirichlet conditions in the region~$S$, and let $u_0$ be the corresponding (positive) normalized eigenfunction, that is,
\begin{equation}\label{eqU0lambda0}
-\Delta u_0 = \lambda_0 u_0, \quad u_0 \in {\mathcal H}_0^1(S),
\quad \int_S u_0(y)^2 dy = 1.
\end{equation} Furthermore, denote by~${\mathcal L}$ the subspace of~$\mathrm L^2(I\times S)$ generated by functions
$w(s) u_0(y)$ with $w \in \mathrm L^2(I)$.
We study three distinct cases. First, the tubes are bounded since the
interval~$I$ is of the form
$I=[-a,b]$ with $0< a, b < \infty$, and we consider the Dirichlet
condition at the
boundary
$\partial \Lambda_\varepsilon$.
In the second case, the tubes are bounded but the Dirichlet condition at
the vertical part of the $\partial \Lambda_\varepsilon$, that is,
$\{(-a) \times S \cup b \times S\}$,
is replaced by Neumann. In the third case we consider $I=\mathbb R$ with
Dirichlet
condition at~$\partial \Lambda_\varepsilon$.
If the tubes are not deformed, according to the results of \cite{Bou,Cr1}, the effective operator~\eqref{bbb8}
presents an additional potential
\[
C(S)\left(\tau+\alpha'\right)(s) - k^2(s)/4
\] derived from geometric features
of the tube. Hence, here there is a kind of competition between geometric
properties of the tube and the behavior at its single maximum of the
deformation function~$h$. Roughly speaking, it is expected that the behavior of $h$ at the single maximum will control the limit $\varepsilon\to0$, since the geometric effects gives a contribution of order zero, whereas the single maximum of the deformation function~$h$ gives a contribution of order $1/\varepsilon$. However, this requires a proof which turns out to be far from trivial, and so for the three cases mentioned in the
previous paragraph, we prove the following result:
\
\begin{teo}\label{mainTeor}
Let $I$ denote either $\mathbb R$ or a bounded interval $[-a,b]$ as above; in case $I=\mathbb R$ assume that $\lim_{|s| \to \infty} k(s) =0$. If $l_j(\varepsilon)$ denote the eigenvalues of the Dirichlet
$-\Delta_\varepsilon$ in the deformed tube~$\Lambda_\varepsilon$, then,
the limits
\begin{equation}\label{eee3}
\mu_j = \lim_{\varepsilon \to 0}\, \varepsilon \left( l_j(\varepsilon) -
\frac{\lambda_0}{\varepsilon^2 M^2} \right)
\end{equation}
exist, where $\mu_j= (2j+1)(2\lambda_0/M^3)^{1/2}$ are the eigenvalues of the self-adjoint
operator~$T$, acting in $\mathrm L^2(\mathbb R)$, given by
\begin{equation}\label{eqweo}
(T u)(s) = - u''(s) + 2\frac{ \lambda_0 }{ M^{3}} \; s^2 \; u(s).
\end{equation}
\end{teo}
\
Due to the conclusions of Theorem~\ref{mainTeor}, $T$ is a WEO for $-\Delta_\varepsilon$ as $\varepsilon\to0$. Note that $T$ has purely discrete spectrum since the potential
\[
V(s)=2\frac{\lambda_0}{M^3}\, s^2\to\infty,\qquad |s|\to\infty;
\] in this case it is the harmonic oscillator potential (but see~\eqref{aaa6} below). Therefore, for deformed tubes as above, the weakly effective operators~$T$ do not
depend on some geometric features of the tube, although the curvature of the reference curve must vanish at infinity. The additional potential $V(s)$ is related to the behavior of~$h(s)$
near it maximum (at the origin). Hence, the eigenvalues of the Laplacian in quite different deformed tubes are described by the same WEO as~$\varepsilon\to0$ !
In Section~\ref{gddh} we present a detailed construction of the
deformed tubes~$\Lambda_\varepsilon$. Our study and technique are focused on
analyzing the sequence of quadratic forms
\begin{equation}\label{seqfq}
F_\varepsilon (\psi) = \int_{\Lambda_\varepsilon} \left(|\nabla \psi|^2
- \frac{\lambda_0}{\varepsilon^2 M^2} |\psi|^2\right) dx,
\qquad \mathrm {dom}\, F_\varepsilon = {\mathcal H}_0^1(\Lambda_\varepsilon).
\end{equation}
In Section~\ref{ffqq} it will become clear why we subtract terms of the
form $\lambda_0 / (\varepsilon^2 M^2)|\psi|^2$ from the quadratic forms; we think
this is in fact a natural choice. In Section~\ref{ffqq} we also perform a
change of variables so that the integration region and the corresponding
domains in~\eqref{seqfq} remain fixed.
In Section~\ref{redim}, we show that our analysis can be restricted to a
specific subspace;
we will see that this subspace can be identified with the Sobolev space
${\mathcal H}_0^1(I)$, and we call this fact a
{\em reduction of dimension.}
Finally, in Sections~\ref{qqq3}, \ref{nnnn} and~\ref{iiii}, we discuss
details of the three cases previously
mentioned.
We remark that although we rely on \cite{Frie,Sol}, the generalization to our setting is not immediate and different techniques are added to those of the original works. Furthermore, as an alternative to~\eqref{vvv1}, all results can be easily adapted to cover more general
deformation functions $h(s)$, as considered in \cite{Frie,Sol}, so that
near the unique global maximum at the origin they behave as
\begin{equation}\label{aaa6}
h(s) = \left\{
\begin{array}{ccc}
M - c_+ s^m + O(s^{m+1}), & \hbox{if} & s>0 \\
M - c_- |s|^m + O(|s|^{m+1}), & \hbox{if} & s<0
\end{array}\right.,
\end{equation}
for some positive numbers $M,m,c_\pm$.
For the sake of simplicity, in Equation~\eqref{vvv1} we have particularized to $m=2$ and
$c_+=c_-=1$.
An interesting problem would be if the maximum of~$h$ would be reached at an interval of values of the parameter~$s$ instead of a single point (see~\cite{BF2} for results in this direction in case of bounded domains, as well as~\cite{BKRS,Grushin}); we are currently working on a related problem.
\section{Geometry of the tubes}\label{gddh}
Let $I=[-a, b]$, with either $0 < a,b< \infty$ or $a=b=\infty$, be an interval of $\mathbb R$,
$r: I \subseteq \mathbb R \rightarrow \mathbb R^3$ a simple $\mathrm C^2$ curve in
$\mathbb R^3$ parametrized by its arc length parameter~$s$ and, as in the previous section, $k(s)$ is its curvature.
The vectors
$$T(s) = r'(s), \qquad N(s) = \frac{1}{k(s)} T'(s), \qquad
B(s) = T(s) \times N(s),$$
denote, respectively, the tangent, normal and binormal vectors of the curve.
We assume that Frenet equations are satisfied, that is,
$$\left(\begin{array}{c}
T'\\
N'\\
B'
\end{array} \right)
=
\left(\begin{array}{ccc}
0 & k & 0 \\
-k & 0 & \tau \\
0 & -\tau & 0
\end{array}\right)
\left(\begin{array}{c}
T\\
N\\
B
\end{array} \right),
$$
where $\tau(s)$ is the torsion of the curve~$r(s)$.
Let $S$ be an open, bounded, simply connected and nonempty subset of~$\mathbb R^2$.
The set
$$\Omega = \left\{ x \in \mathbb R^3: x = r(s) + y_1 N(s) + y_2 B(s),\,
s \in I, y=(y_1, y_2) \in S \right\}$$
is obtained by translating the region $S$ along the curve~$r$. At each point
$r(s)$ we allow a rotation of the region~$S$ by an angle $\alpha(s)$ with
respect to $\alpha(0)=0$, so that the new region is given by
$$\Omega^\alpha = \left\{ x \in \mathbb R^3: x = r(s) + y_1 N_\alpha(s) + y_2
B_\alpha(s),
\,s \in I, (y_1, y_2) \in S \right\},$$
where
\begin{eqnarray*}
N_\alpha(s) & : = & \cos \alpha(s) N(s) + \sin \alpha(s) B(s), \\
B_\alpha(s) & : = & - \sin \alpha(s) N(s) + \cos \alpha(s) B(s).
\end{eqnarray*}
Next, for each $0<\varepsilon <1$, we ``squeeze'' the cross sections of the above region,
that is, we consider
$$\Omega_\varepsilon^\alpha = \left\{ x \in \mathbb R^3: x = r(s) + \varepsilon
y_1 N_\alpha(s) + \varepsilon y_2 B_\alpha(s),\,
s \in I, (y_1, y_2) \in S\right \}.$$
Note that $\Omega_\varepsilon^\alpha$ approaches the curve $r(s)$ as
$\varepsilon \to 0$.
Finally, we consider the function $h(s)$ defined in the Introduction, so that
each region $\Omega_\varepsilon^\alpha$ is properly deformed, and the
result is
$$\Lambda_\varepsilon^\alpha := \left\{ x \in \mathbb R^3: x = r(s) +
\varepsilon h(s) y_1 N_\alpha(s) + \varepsilon y_2 h(s) B_\alpha(s),\,
s \in I, (y_1, y_2) \in S \right\}.$$
From now on we will omit the symbol $\alpha$ in most notations and write $dx=dsdy_1dy_2$ and~$dy=dy_1dy_2$.
In this work we study the behavior of a free quantum particle that moves in
$\Lambda_\varepsilon$, and initially with Dirichlet boundary condition at the boundary
$\partial\Lambda_\varepsilon$.
Thus, we initially consider the family of quadratic forms
\begin{equation}\label{ddd6}
b_\varepsilon (\psi) := \int_{\Lambda_\varepsilon} |\nabla \psi|^2 dx,
\qquad
\mathrm {dom}\, b_\varepsilon = {\mathcal H}_0^1(\Lambda_\varepsilon),
\end{equation}
which is associated with the Dirichlet Laplacian operator $-\Delta_\varepsilon$ in
$\Lambda_\varepsilon$.
The symbol $\nabla = (\partial_s, \nabla_y)$, $\nabla_y=(\partial_{y_1},\partial_{y_2})$, denotes the gradient in the coordinates $(s,y_1,y_2)$ in~$\mathbb R^3$.
\section{Quadratic forms}\label{ffqq}
As usual in this kind of problems, in this section we perform a change of
variables so that the integration region
in \eqref{ddd6}, and consequently the domains, become independent
of~$\varepsilon > 0$. Then, for the singular limit $\varepsilon\to0$, customary ``regularizations'' will be employed.
Consider the mapping
$$\begin{array}{cccl}
f_\varepsilon: & I \times S & \to & \Lambda_\varepsilon \\
& (s, y_1, y_2) & \mapsto & r(s) + \varepsilon\, h(s) \left( y_1
N_\alpha(s) + y_2 B_\alpha(s)\right),
\end{array}$$
and suppose the boundedness $\| k\|_\infty, \|\tau\|_\infty,
\|\alpha'\|_\infty < \infty$.
These conditions are to guarantee that
$f_\varepsilon$ will be a diffeomorphism.
With this change of variables we work with a fixed region for all
$\varepsilon > 0$; more precisely, the domain of the quadratic form~\eqref{ddd6} turns out to be ${\mathcal H}_0^1(I \times S)$.
On the other hand, the price to be paid is a nontrivial Riemannian metric $G=
G_\varepsilon^\alpha$
which is induced by $f_\varepsilon$, i.e.,
$$G=(G_{ij}),\qquad G_{ij} = \langle e_i, e_j \rangle = G_{ji},
\qquad
1 \leq i, j \leq 3,$$
where
$$e_1 = \frac{\partial f_\varepsilon}{\partial s}, \qquad
e_2 = \frac{\partial f_\varepsilon}{\partial y_1}, \qquad
e_3 = \frac{\partial f_\varepsilon}{\partial y_2}.$$
Some calculations show that in the Frenet frame
\begin{eqnarray*}
J & = &
\left(\begin{array}{c}
e_1 \\
e_2 \\
e_3
\end{array}\right)
\\
& = &
\left(\begin{array}{ccc}
\beta_\varepsilon &
-\varepsilon h (\tau+\alpha') \langle z_\alpha^\perp, y \rangle +
\varepsilon h' \langle z_\alpha, y \rangle &
\varepsilon h (\tau+\alpha') \langle z_\alpha, y \rangle + \varepsilon h'
\langle z_\alpha^\perp, y \rangle \\
0 & \varepsilon h \cos \alpha & \varepsilon h \sin \alpha \\
0 & -\varepsilon h \sin \alpha & \varepsilon h \cos \alpha
\end{array}\right),
\end{eqnarray*}
where
$$\beta_\varepsilon(s, y) = 1 - \varepsilon h(s) k(s) \langle z_\alpha, y
\rangle, \quad
z_\alpha : = (\cos \alpha, - \sin \alpha), \quad
z_\alpha^\perp : = (\sin \alpha, \cos \alpha).$$
The inverse matrix of~$J$ is given by
$$J^{-1} =
\left(\begin{array}{ccc}
\frac{1}{\beta_\varepsilon} &
\frac{1}{\beta_\varepsilon} \left[(\tau + \alpha') y_2 - \frac{h'}{h} y_1
\right] &
\frac{1}{\beta_\varepsilon} \left[ - (\tau + \alpha') y_1 - \frac{h'}{h}
y_2 \right] \\
0 & \frac{\cos \alpha}{\varepsilon h} & \frac{-\sin \alpha}{\varepsilon
h} \\
0 & \frac{\sin \alpha}{\varepsilon h} & \frac{\cos \alpha}{\varepsilon h}
\end{array}\right).$$
Note that $J J^t = G$ and $\det J = | \det G|^{1/2} = \varepsilon^2
h^2(s) \beta_\varepsilon(s, y)$.
Since $k$ and $h$ are bounded functions, for $\varepsilon$ small enough
$\beta_\varepsilon$ does not vanish in $I \times S$. Thus,
$\beta_\varepsilon > 0$ and $f_\varepsilon$ is
a local diffeomorphism. By requiring that $f_\varepsilon$ is injective
(that is, the tube is not self-intersecting), a global
diffeomorphism is obtained.
Introducing the notation
$$\left| \left| \psi \right| \right|_G^2
:=\int_{I \times S} | \psi(s,y) |^2 \varepsilon^2 h^2(s)
\beta_\varepsilon(s, y) \,ds dy,$$
we obtain a sequence of quadratic forms
$$\tilde{b}_\varepsilon
(\psi) : = \left| \left| J^{-1}\nabla \psi \right| \right|_G,
\qquad
\mathrm {dom}\, \tilde{b}_\varepsilon = {\mathcal H}_0^1(I \times S, G).$$
More precisely, the above change of coordinates was obtained by a
unitary transformation
$$\begin{array}{cccl}
U_\varepsilon: & \mathrm L^2(\Lambda_\varepsilon) & \to & \mathrm L^2(I \times S, G) \\
& \phi & \mapsto & \phi \circ f_\varepsilon
\end{array}.$$
However, we still denote $U_\varepsilon \psi$ by $\psi$.
Recall that $\lambda_0$ is the lowest eigenvalue of the negative Laplacian with Dirichlet boundary conditions in the cross section region~$S$, and $u_0\ge0$ (see Equation~\eqref{eqU0lambda0}) the corresponding eigenfunction of this restricted problem. This eigenfunction~$u_0$ is directly related to transverse oscillations in~$\Lambda_\varepsilon$.
Due to this fact, in \cite{Bou,Cr1} the authors have remove the diverging energy $\lambda_0 / \varepsilon^2$ from their quadratic forms.
In our case, as the boundary of the tubes were multiplied by~$h(s)$, we
subtract the terms of the
form $\lambda_0 /( \varepsilon M)^2$, i.e., since
$0 < h(s) \le M$, for all $s \in I$,
we eliminate the possible ``least transverse energy.''
Therefore, we turn to the study of the sequence of quadratic forms
$$\tilde{g}_\varepsilon(\psi) : = \left(
\left| \left| J^{-1} \nabla \psi \right| \right|_G^2 -
\frac{\lambda_0}{\varepsilon^2 M^2} \| \psi\|_G^2
+ c \| \psi\|_G^2 \right),$$
where $c$ is a positive constant to be chosen later on. After the norms are written out, we obtain
\begin{eqnarray*}
\tilde{g}_\varepsilon (\psi) & = &
\varepsilon^2 \int_{I \times S}
\Big( \frac{1}{\beta_\varepsilon^2(s,y)}
\left|\psi' + \nabla_y \psi \cdot
R y (\tau+\alpha')(s) - \nabla_y \psi \cdot y \,\frac{h'(s)}{h(s)}
\right|^2 \\ &+& \frac{ |\nabla_y \psi|^2 }{\varepsilon^2 h(s)^2}
- \frac{\lambda_0}{\varepsilon^2 M^2} |\psi|^2 + c |\psi|^2 \Big)
h(s)^2 \beta_\varepsilon(s, y)\,ds dy\; .
\end{eqnarray*}
Note that $\mathrm {dom}\, \tilde{g}_\varepsilon = {\mathcal H}_0^1(I \times S)$ is a
subspace of
$\mathrm L^2(I \times S, h(s)^2 \beta_\varepsilon(s, y))$.
We observe that the factor
$|\nabla_y \psi|^2 / (\varepsilon h(s))^2$ is directly related to
transverse oscillations
of the particle. This term diverges as~$\varepsilon \to 0$, but we control
this fact by subtracting $\lambda_0 /( \varepsilon M)^2
|\psi|^2$ from the quadratic form (a renormalization).
It will be convenient to work in the space~$\mathrm L^2(I \times S, \beta_\varepsilon(s,
y))$; so we consider the isometry
$$\begin{array}{cccl}
& \mathrm L^2(I \times S, \beta_\varepsilon) & \to & \mathrm L^2(I \times S, h(s)^2
\beta_\varepsilon) \\
& v & \mapsto & v h^{-1}
\end{array}.$$
This change of variables and the division by the global factor~$\varepsilon^2$ (a common singular factor due to the ``change of dimension'' as $\varepsilon\to0$) leads to
\begin{eqnarray*}\hat{g}_\varepsilon (v) &: =&
\int_{I \times S} \Big( \frac{1}{\beta_\varepsilon(s,y)}
\left|v' - v \,\frac{h'(s)}{h(s)} + \nabla_y v \cdot
R y (\tau+\alpha')(s) - \nabla_y v \cdot y \,\frac{h'(s)}{h(s)} \right|^2
\\
&+& \frac{\beta_\varepsilon(s,y)}{\varepsilon^2 h(s)^2} |\nabla_y
v|^2
- \frac{\beta_\varepsilon(s, y)}{\varepsilon^2 M^2} |v|^2 + c
\beta_\varepsilon(s, y) |v|^2 \Big)\,ds dy,
\end{eqnarray*}
with $\mathrm {dom}\, \hat{g}_\varepsilon = {\mathcal H}_0^1(I \times S)$, again as a
subspace of
$\mathrm L^2(I \times S, \beta_\varepsilon(s, y))$.
However, this latter space can be identified with
$\mathrm L^2(I \times S)$, for all~$\varepsilon>0$, since $\beta_\varepsilon(s,
y)$ converges uniformly to $1$
as $\varepsilon \to 0$. Hence we introduce the form
\begin{eqnarray*}g_\varepsilon (v) &: =&
\int_{I \times S} \Big( \left|v' - v \,\frac{h'(s)}{h(s)} + \nabla_y v \cdot
R y (\tau+\alpha')(s) - \nabla_y v \cdot y \,\frac{h'(s)}{h(s)} \right|^2
\\ &+& \frac{\beta_\varepsilon(s,y)}{\varepsilon^2 h(s)^2} |\nabla_y
v|^2
- \frac{\beta_\varepsilon(s, y)}{\varepsilon^2 M^2} |v|^2 + c |v|^2
\Big)\,ds dy,
\end{eqnarray*}
with $\mathrm {dom}\, g_\varepsilon = {\mathcal H}_0^1(I \times S)$.
Let $\hat{G}_\varepsilon$ and $G_\varepsilon$ be the self-adjoint
operators associated with the quadratic forms~$\hat{g}_\varepsilon$ and~$g_\varepsilon$, respectively.
\
\begin{teo}\label{teoap}
For $\varepsilon$ small enough, there exists~$C > 0$ so that
$$\left\| \hat{G}_\varepsilon^{-1} - G_\varepsilon^{-1} \right\| \leq C \varepsilon.$$
\end{teo}
\
This theorem follows basically from the fact that $\beta_\varepsilon (s,
y) \to 1$
uniformly as $\varepsilon \to 0$. Its proof is presented in the Appendix.
Due to the above changes of variables and Theorem~\ref{teoap}, we may consider the
sequence of quadratic forms~$g_\varepsilon$ in what follows.
\section{Reduction of dimension}\label{redim}
Recall that
$u_0(y)$ is the positive and normalized eigenfunction corresponding to the
first eigenvalue $\lambda_0$ of the
Laplacian in ${\mathcal H}_0^1(S)$.
After the orthogonal decomposition $\mathrm L^2(\mathbb R \times S) = {\mathcal L}
\oplus {\mathcal L}^\perp,$
for $\psi \in \mathrm L^2(\mathbb R \times S)$, we can write
$$\psi(s, y) = w(s) u_0(y) + \eta(s, y),$$
with $w \in \mathrm L^2(I)$ and $\eta \in {\mathcal L}^\perp$.
We observe that $\eta\in {\mathcal L}^\perp$ implies
$$\int_S u_0(y) \eta(s, y) dy = 0, \qquad
\hbox{a.e.}[s].$$
Note that $wu_0 \in {\mathcal H}_0^1(I \times S)$ if $w \in {\mathcal
H}_0^1(I)$.
For $\psi \in {\mathcal H}_0^1(\mathbb R \times S)$, write
$\psi = w u_0 + \eta$ with $w \in {\mathcal H}_0^1(I)$ and $\eta \in {\mathcal
H}_0^1(\mathbb R \times S) \cap {\mathcal L}^\perp$.
First we study the quadratic form $g_\varepsilon$ restricted to
the subspace
\mbox{${\mathcal H}_0^1(I \times S) \cap {\mathcal L}$.}
For $w \in {\mathcal H}_0^1(I)$, some calculations show that
\begin{eqnarray*}
g_\varepsilon( w u_0) = \int_{I} \left[
|w'|^2 + \vartheta(s) |w|^2 +
\zeta_\varepsilon(s, y)
\left(\frac{\lambda_0}{\varepsilon^2h^2(s)}-\frac{\lambda_0}{\varepsilon^2M^2}\right)
|w|^2 + c |w|^2 \right] ds,
\end{eqnarray*}
where
$$\vartheta(s) = C_1(S) (\tau(s) + \alpha'(s))^2
+ (C_2(S) - 1) \left(\frac{h'(s)}{h(s)}\right)^2 - 2 C_3(S)
(\tau(s)+\alpha'(s)) \,\frac{h'(s)}{h(s)}$$
and
$$\zeta_\varepsilon(s, y) = 1 - \varepsilon\, k(s) h(s) \left\langle
z_\alpha(s), F(S) \right\rangle .$$
The constants $C_1(S)$, $C_2(S)$ and $C_3(S)$ that appear in the
definition of~$\vartheta$ depend only on the region $S$
and are explicitly given by
$$C_1(S) = \int_S |\langle \nabla_y u_0, R y \rangle|^2 dy, \qquad
C_2(S) = \int_S |\langle \nabla_y u_0, y \rangle|^2 dy,$$ and
$$C_3(S) = \int_S \langle \nabla_y u_0, R y \rangle \langle \nabla_y u_0,
y \rangle dy.$$
The vector $F(S) = (F_1(S), F_2(S))$ in the definition of
$\zeta_\varepsilon$ also depends only on the
region~$S$, and its components are given by
$$F_1(S) = \int_S y_1 |u_0|^2 dy\qquad \hbox{and}\qquad
F_2(S) = \int_S y_2 |u_0|^2 dy.$$
Under such restrictions, the quadratic form $b_\varepsilon$ in ${\mathcal H}_0^1(I)$ can be written in terms of the form $t_\varepsilon=t_{\varepsilon,c}$ given by
\begin{equation}\label{eqdefteps}
t_\varepsilon(w) : = g_\varepsilon(wu_0)=\int_I \left( |w'|^2 + W_\varepsilon(s) |w|^2
\right) ds,
\end{equation}
with
\begin{equation}\label{aaa1}
W_\varepsilon(s) : = \vartheta(s)+ c + \zeta_\varepsilon(s, y)
\left(\frac{\lambda_0}{\varepsilon^2 h^2(s)}-\frac{\lambda_0}{
\varepsilon^2 M^2}\right).
\end{equation}
We choose the constant~$c$ so that $c > \|v\|_\infty + (1 / M^2)\|
k(s)^2 / 4\|_\infty$.
Since $k(s)$ and $h(s)$ are bounded functions, there exist
$\varepsilon_1 > 0$ and $\delta > 0$ so that, for all $s\in I$,
$$
1 - \varepsilon\, k(s) h(s) \langle z_\alpha(s), F(S) \rangle > \delta \qquad \hbox{and} \qquad
1 - \varepsilon\, k(s) h(s) \langle z_\alpha(s), y \rangle > \delta,
$$
for all $\varepsilon < \varepsilon_1$.
In what follows, we tacitly assume that $\varepsilon < \varepsilon_1$.
The self-adjoint operator associated with $t_\varepsilon$ in $\mathrm L^2(I)$ is
$$(T_{\varepsilon,c} w)(s) := - w''(s) + W_\varepsilon(s) w(s), \qquad
\mathrm {dom}\,{T_{\varepsilon,c}} = {\mathcal H}^2(I)\cap {\mathcal H}_0^1(I).$$
From now on we denote by $-\Delta_{\varepsilon,c}$ the operator
$-\Delta_\varepsilon + c\mathbf 1$ and write $T_\varepsilon=T_{\varepsilon,c}-c\mathbf 1$.
Next we discuss how the resolvent operator
$\left(- \Delta_{\varepsilon,c} - \lambda_0 / \varepsilon^2 M^2 \mathbf 1 \right)^{-1}$
can be approximated by $T_{\varepsilon,c}^{-1}\oplus 0$, where $0$ is the
null operator on the
subspace~${\mathcal L}^\perp$. Such result gives a quantitative indication of how $-\Delta_\varepsilon$ is approximated by~$T_\varepsilon$.
\
\begin{lem}\label{ddd10}
{\rm Suppose that $I$ is a bounded interval. Then, there exists
$C_6 > 0$ so that
$$t_\varepsilon(w) \geq C_6^{-1}\, \varepsilon^{-1} \int_I |w|^2 ds,
\qquad \forall w\in {\mathcal H}_0^1(I),\;0 < \varepsilon < \varepsilon_1.$$
}
\end{lem}
\
By noting that
$$ \frac{\varepsilon^2 W_\varepsilon (s)}{s^2} \geq \frac{\lambda_0}{s^2}
\delta \left(
\frac{1}{h(s)^2} - \frac{1}{M^2} \right),$$
the proof of Lemma~\ref{ddd10} is similar to the proof of Lemma $2.1$
in~\cite{Frie}, and it will not be reproduced here.
By following \cite{Bou}, for each $\xi \in \mathbb R^2$, we consider the
following perturbed problem
$$-{\rm div}[(1 - (\xi \cdot y )) \nabla_y u] =
\lambda (1 - (\xi \cdot y)) u, \qquad
u \in {\mathcal H}_0^1(S).$$
By taking $\xi= \varepsilon h(s) k(s) z_\alpha$, for $\varepsilon$
small enough, the perturbed operator
is positive and with compact resolvent. Denote by $\lambda(\xi) > 0$
its first eigenvalue, i.e.,
$$\lambda(\xi) =
\inf_{\{u \in {\mathcal H}_0^1(S):\: u \neq 0\}}
\frac{\int_S (1 - (\xi \cdot y )) |\nabla_y u|^2 dy}{\int_S (1 - (\xi
\cdot y )) |u|^2 dy}.$$
Thus, for $v \in {\mathcal H}_0^1(\mathbb R \times S)$,
\begin{equation}\label{ddd8}
\frac{1}{\varepsilon^2} \int_S \beta_\varepsilon(s,y)
( | \nabla_y v|^2 - \lambda_0 |v|^2) dy \geq \gamma_\varepsilon(s) \int_S
\beta_\varepsilon(s,y) |v|^2 dy
\qquad
{\rm a.e.}[s],
\end{equation}
where
$$\gamma_\varepsilon(s) := \frac{\lambda(\varepsilon h(s) k(s)
z_{\alpha}(s)) - \lambda_0}{\varepsilon^2}.$$
Using the fact that $h(s)$ and $k(s)$ are bounded functions, it is possible to prove
that
$\gamma_\varepsilon(s)$ converges uniformly as $\varepsilon \to 0$ to a bounded function (see
Proposition $4.1$ in \cite{Bou}). This will be used in the proof of Lemma~\ref{eee4}.
\begin{lem}\label{eee4}
{\rm
Let $I$ denote either $\mathbb R$ or a bounded interval. Then, for $\eta \in
{\mathcal H}_0^1(I \times S) \cap {\mathcal L}^\perp$, there exists $C_7\in\mathbb R$ so
that,
for $\varepsilon$ small enough,
$$g_\varepsilon(\eta) \geq \frac{C_7}{\varepsilon^2 M^2} \| \eta \|^2.$$}
\end{lem}
\begin{dm}
Let $\lambda_1$ be the second eigenvalue of the Laplacian in ${\mathcal
H}_0^1(S)$, and pick
$\eta \in {\mathcal H}_0^1(\mathbb R \times S) \cap {\mathcal L}^\perp$.
Since
$h(s) \leq M$, for all $s \in I$, we have
$$\int_S
\beta_\varepsilon(s, y) \left(
\frac{|\nabla_y \eta|^2}{\varepsilon^2 h(s)^2}
- \lambda_1 \frac{|\eta|^2}{\varepsilon^2 M^2} \right)dy
\geq
\int_S \beta_\varepsilon(s, y) \left(
\frac{|\nabla_y \eta|^2}{\varepsilon^2 M^2}
- \lambda_1 \frac{|\eta|^2}{\varepsilon^2 M^2} \right)dy.$$
By \eqref{ddd8}, it follows that
$$\int_S \beta_\varepsilon(s, y) \left(
\frac{|\nabla_y \eta|^2}{\varepsilon^2 M^2}
- \lambda_1 \frac{|\eta|^2}{\varepsilon^2 M^2} \right)dy
\geq
\frac{\gamma_\varepsilon(s)}{M^2} \int_S \beta_\varepsilon(s, y) |\eta|^2 dy.$$
Since $\gamma_\varepsilon(s)$ converges uniformly as $\varepsilon \to 0$,
there exists $C_8 \in \mathbb R$ so that, for $\varepsilon$ small enough,
$$\frac{\gamma_\varepsilon(s)}{M^2} \geq C_8, \qquad \forall s \in I.$$
Thus,
$$\frac{\gamma_\varepsilon(s)}{M^2} \int_S \beta_\varepsilon(s, y) |\eta|^2 dy
\geq
C_8 \int_S \beta_\varepsilon(s, y) |\eta|^2 dy,$$
and so
$$\int_{I \times S}
\beta_\varepsilon(s, y) \left(
\frac{|\nabla_y \eta|^2}{\varepsilon^2 h(s)^2}
- \lambda_1 \frac{|\eta|^2}{\varepsilon^2 M^2} \right)dy ds
\geq
C_8 \int_{I \times S} \beta_\varepsilon(s, y) |\eta|^2 dy ds.$$
Adding and subtracting the term $ \frac{\lambda_0}{\varepsilon^2 M^2}
\int_{I \times S} \beta_\varepsilon (s, y) |\eta|^2 dyds$
on the left hand side of the above inequality, we obtain
\begin{eqnarray*}
\int_{I \times S}
&\beta_\varepsilon(s, y)& \left(
\frac{|\nabla_y \eta|^2}{\varepsilon^2 h(s)^2}
- \lambda_0 \frac{|\eta|^2}{\varepsilon^2 M^2} \right) dy ds\\
& \geq &
C_8 \int_{I \times S} \beta_\varepsilon(s, y) |\eta|^2 dy +
\frac{(\lambda_1 - \lambda_0)}{\varepsilon^2 M^2} \int_{I\times S}
\beta_\varepsilon(s, y) |\eta|^2 dy ds.
\end{eqnarray*}
Now, for $\varepsilon$ small enough, there exists~$C_9$ so that
\begin{eqnarray*}
g_\varepsilon(\eta) &\geq &
\int_{I \times S}
\beta_\varepsilon(s, y) \left(
\frac{|\nabla_y \eta|^2}{\varepsilon^2 h(s)^2}
- \lambda_0 \frac{|\eta|^2 }{\varepsilon^2 M^2} \right) dy ds + c \int_{I
\times S} |\eta|^2 dy ds\\
& \geq &
C_8 \delta \int_{I \times S} |\eta|^2 dy ds +
\frac{(\lambda_1 - \lambda_0)}{\varepsilon^2 M^2} \int_{I \times S}
\beta_\varepsilon(s, y) |\eta|^2 dy ds
+ c \int_{I \times S} |\eta|^2 dy ds\\
& \geq &
\frac{C_9}{\varepsilon^2 M^2} \int_{I \times S} \beta_\varepsilon(s, y)
|\eta|^2 dy ds\\
& \geq &
\frac{C_9}{\varepsilon^2 M^2} \delta \int_{I \times S} |\eta|^2 dy ds.
\end{eqnarray*}
Finally, it is enough to take $C_7= C_9\, \delta$ to complete the proof of the
lemma.
$\hfill \rule {2.5mm} {3mm}\vspace{0.5cm}$
\end{dm}
Now we are ready to state and prove the main result of this section; it will rest on results presented in Section~$3$ of~\cite{Sol}, combined with the previous lemmas.
\begin{teo}\label{aaa4}
{\rm Let $I$ denote either~$\mathbb R$ or a bounded interval. Then there exists~$C_{10} > 0$ so that, for $\varepsilon$ small enough,
$$\left|\left| \left(- \Delta_{\varepsilon,c} -
\frac{\lambda_0}{\varepsilon^2 M^2} \mathbf 1 \right)^{-1}
- \left(T_{\varepsilon,c}^{-1} \oplus 0\right) \right| \right| \leq C_{10}\,
\varepsilon^{3/2},$$
where $0$ denotes the null operator on the subspace ${\mathcal L}^\perp$.
}
\end{teo}
\begin{dm}{\rm
For $\psi \in {\mathcal H}_0^1(\mathbb R \times S)$ write
$$\psi(s,y) = w(s) u_0(y) + \eta(s, y),$$
with $w \in {\mathcal H}_0^1(I)$ and $\eta \in {\mathcal H}_0^1(\mathbb R
\times S) \cap {\mathcal L}^\perp$. Thus,
the quadratic form $g_\varepsilon(\psi)$ can be rewritten as
$$g_\varepsilon(\psi) = t_\varepsilon(w) + g_\varepsilon(\eta) + 2
m_\varepsilon(wu_0, \eta),$$
where $t_\varepsilon(w) = g_\varepsilon(wu_0)$ (see \eqref{eqdefteps}) and
\begin{eqnarray*}
m_\varepsilon (wu_0, \eta) & = &
\int_{I \times S} dy ds \Big[
\left( w' u_0 - w u_0 \,\frac{h'}{h} + w \nabla_y u_0 \cdot Ry
(\tau+\alpha') - w \nabla_y u_0 \cdot y \,\frac{h'}{h} \right) \\
& \times &
\left(\eta' - \eta \,\frac{h'}{h} + \nabla_y \eta \cdot Ry (\tau+\alpha') -
\nabla_y \eta \cdot y \,\frac{h'}{h} \right) \Big]\\
& - &
\int_{I \times S} dy ds\;
k(s) h(s) \langle z_\alpha, y \rangle w \left(\frac{\nabla_y u_0 \nabla_y
\eta}{\varepsilon h^2} -
\lambda_0 \frac{u_0 \eta}{\varepsilon M^2} \right).
\end{eqnarray*}
We are going to show that $t_\varepsilon(w)$, $g_\varepsilon(\eta)$ and
$m_\varepsilon(wu_0, \eta)$ satisfy the conditions
(3.2), (3.3), (3.4) and (3.5) in Section~$3$ of \cite{Sol},
and so the theorem will follow. Conditions (3.2), (3.3) and (3.4) are obtained by applying Lemmas~\ref{ddd10} and~\ref{eee4} above. We need only to verify condition~(3.5), i.e., that there exists a
function $q(\varepsilon)$ so that for each
$\psi \in {\mathcal H}_0^1(I \times S)$
\begin{equation}\label{eee2}
|m_\varepsilon(wu_0, \eta)|^2 \,\leq\, q(\varepsilon)^2\, t_\varepsilon(w)
\,g_\varepsilon(\eta), \qquad
q(\varepsilon) \to 0 \qquad (\varepsilon \to 0).
\end{equation}
We write
$$m_\varepsilon(wu_0, \eta) = m_\varepsilon^1(wu_0, \eta) -
m_\varepsilon^2(wu_0, \eta) + m_\varepsilon^3(wu_0, \eta) -
m_\varepsilon^4(wu_0, \eta)
- m_\varepsilon^5(wu_0, \eta),$$
where
\begin{eqnarray*}
m_\varepsilon^1(w u_0, \eta) & : = &
\int_{\mathbb R \times S}
w' u_0 \left( \eta' - \eta \,\frac{h'}{h} + \nabla_y \eta \cdot R y (\tau + \alpha')
- \nabla_y \eta \cdot y \,\frac{h'}{h} \right) \,ds dy\, , \\
m_\varepsilon^2(wu_0, \eta) & : = &
\int_{\mathbb R \times S}
w u_0 \,\frac{h'}{h} \left( \eta' - \eta \,\frac{h'}{h} + \nabla_y \eta \cdot R y
(\tau + \alpha')
- \nabla_y \eta \cdot y \,\frac{h'}{h} \right) \,ds dy\, , \\
m_\varepsilon^3(wu_0, \eta) & : = & \int_{\mathbb R \times S}
w \nabla_y u_0 \cdot Ry (\tau+\alpha')
\left( \eta' - \eta \,\frac{h'}{h} + \nabla_y \eta \cdot R y (\tau + \alpha')
- \nabla_y \eta \cdot y \,\frac{h'}{h} \right) dy ds\, , \\
m_\varepsilon^4(wu_0, \eta) & := &
\int_{\mathbb R \times S}
w \nabla_y u_0 \cdot y \,\frac{h'}{h}
\left( \eta' - \eta \,\frac{h'}{h} + \nabla_y \eta \cdot R y (\tau + \alpha')
- \nabla_y \eta \cdot y \,\frac{h'}{h} \right) \,ds dy\, , \\
m_\varepsilon^5(wu_0, \eta) & : = &
\int_{I \times S}
\frac{k(s) h(s) \langle z_{\alpha} , y \rangle}{\varepsilon}
w \left(\frac{\nabla_y u_0 \nabla_y \eta} {h^2} - \lambda_0
\frac{u_0 \eta}{M^2} \right) dy ds.
\end{eqnarray*}
Now we are going to estimate each of the above terms. Let
$$ H_1: = \left| \left| \frac{h'}{h} \right| \right|_\infty,
\qquad
H_2 : = \left| \left| \tau + \alpha ' \right| \right|_\infty,$$
and recall that
$$C_1(S) = \int_S |\langle \nabla_y u_0, R y \rangle|^2 dy \qquad
\hbox{and} \qquad
C_2(S) = \int_S |\langle \nabla_y u_0, y \rangle|^2 dy.$$
By Green identities and some calculations, we get
$$\int_S u_0 \langle \nabla_y \eta, Ry \rangle dy = - \int_S \langle \nabla_y
u_0, Ry \rangle \eta dy,$$
$$\int_S u_0 \langle \nabla_y \eta, y \rangle dy = - \int_S \langle \nabla_y
u_0, y \rangle \eta dy.$$
Hence:
\begin{eqnarray*}
\bullet \quad
|m_\varepsilon^1(wu_0, \eta)| & \leq &
H_2^{1/2} C_1(S) \left( \int_I |w'|^2 ds \right)^{1/2} \left( \int_{I
\times S} |\eta|^2 dy ds\right)^{1/2} \\
& + &
H_1^{1/2} C_2(S) \left( \int_I |w'|^2 ds \right)^{1/2} \left( \int_{I
\times S} |\eta|^2 dy ds \right)^{1/2} \\
& \leq &
C_{11}\, \varepsilon\, t_\varepsilon(w)^{1/2} g_\varepsilon(\eta)^{1/2}; \\
\end{eqnarray*}
\begin{eqnarray*}
\bullet \quad | m_\varepsilon^2(wu_0, \eta)| & \leq &
\left(\int_{I \times S} |w|^2 |u_0|^2 \left(\frac{h'}{h}\right)^2 dy
ds\right)^{1/2} g_\varepsilon(\eta)^{1/2}\\
& \leq &
H_1 \left(\int_I |w|^2 ds\right)^{1/2} g_\varepsilon(\eta)^{1/2}\\
& \leq &
\frac{H_1}{C_6^{1/2}}\, \varepsilon^{1/2}\, t_\varepsilon(w)^{1/2}
g_\varepsilon(\eta)^{1/2};\\
\end{eqnarray*}
\begin{eqnarray*}
\bullet \quad |m_\varepsilon^3(wu_0, \eta)| & \leq &
\left(\int_{I \times S} |w|^2 |\langle \nabla_y u_0, Ry \rangle|^2
(\tau+\alpha')^2 dy ds \right)^{1/2}
g_\varepsilon(\eta)^{1/2} \\
& = &
\left(\int_I |w|^2 C_1(S) (\tau+\alpha')^2 ds \right)^{1/2}
g_\varepsilon(\eta)^{1/2}\\
& \leq &
C_1(S)^{1/2} H_2 \left(\int_I |w|^2 ds\right)^{1/2} g_\varepsilon(\eta)^{1/2} \\
& \leq &
C_1(S)^{1/2} \frac{H_2}{C_6^{1/2}}\, \varepsilon^{1/2}\,
t_\varepsilon(w)^{1/2} g_\varepsilon(\eta)^{1/2};\\
\end{eqnarray*}
\begin{eqnarray*}
\bullet \quad
|m_\varepsilon^4(wu_0, \eta)| & \leq &
\left(\int_{I \times S} |w|^2 |\langle \nabla_y u_0, y \rangle |^2
\left(\frac{h'}{h}\right)^2 dy ds \right)^{1/2} g_\varepsilon(\eta)^{1/2} \\
& = &
\left(\int_I |w|^2 C_2(S) \left(\frac{h'}{h}\right)^2 ds \right)^{1/2}
g_\varepsilon(\eta)^{1/2}\\
& \leq &
C_2(S)^{1/2} H_1 \left(\int_I |w|^2 ds\right)^{1/2} g_\varepsilon(\eta)^{1/2} \\
& \leq &
C_2(S)^{1/2} \frac{H_1}{C_6^{1/2}}\, \varepsilon^{1/2}\,
t_\varepsilon(w)^{1/2} g_\varepsilon(\eta)^{1/2}.
\end{eqnarray*}
Additional calculations show that
$$\int_{S} y_1 \langle \nabla_y u_0, \nabla_y \eta \rangle =
- \int_{S} f_1(y) \eta dy ds,$$
$$\int_{S} y_2 \langle \nabla_y u_0 \nabla_y \eta \rangle =
- \int_{S} f_2(y) \eta dy ds,$$
where
$$f_1(y) : = \left(\frac{\partial u_0}{\partial y_1} + y_1 \frac{\partial^2
u_0}{\partial y_1^2}
+ y_1 \frac{\partial^2 u_0}{\partial y_2^2} \right),$$
$$f_2(y) : = \left(\frac{\partial u_0}{\partial y_2} + y_2 \frac{\partial^2
u_0}{\partial y_2^2}
+ y_2 \frac{\partial^2 u_0}{\partial y_1^2} \right).$$
Thus, there exist $C_{12}$ and $C_{13}$ so that
\begin{eqnarray*}
|m_\varepsilon^5(wu_0, \eta)| & \leq &
\left|
\int_{I \times S} k(s) \cos \alpha(s) y_1 w \frac{\nabla_y u_0 \nabla_y
\eta}{\varepsilon h} dy ds \right| \\
& + & \left| \int_{I \times S} k(s) h(s) \cos \alpha(s) y_1 \lambda_0 w
\frac{u_0 \eta}{\varepsilon M^2} dy ds \right| \\
& + &
\left|
\int_{I \times S} k(s) \sin \alpha(s) y_2 w \frac{\nabla_y u_0 \nabla_y
\eta}{\varepsilon h} dy ds \right| \\
& + & \left| \int_{I \times S} k(s) h(s) \sin \alpha(s) y_2 \lambda_0 w
\frac{u_0 \eta}{\varepsilon M^2} dy ds \right| \\
& \leq &
\frac{C_{12}}{\varepsilon} \left(\int_\mathbb R |w|^2 ds\right)^{1/2}
\left(\int_{\mathbb R \times S} |\eta|^2 dy ds\right)^{1/2}\\
& \leq &
C_{13}\, \varepsilon^{1/2}\, t_\varepsilon(w)^{1/2} g_\varepsilon(\eta)^{1/2}.
\end{eqnarray*}
By the above estimates it follows that there exists
$C_{14} > 0$ so that
$$|m_\varepsilon(wu_0, \eta)|^2 \leq C_{14}\, \varepsilon\, t_\varepsilon(w)\,
g_\varepsilon(\eta),$$
and so \eqref{eee2} is proven.
By applying Proposition~$3.1$ of \cite{Sol}, it is found that there exists
$C_{10}$ so that, for $\varepsilon$ small enough,
$$\left| \left| \left(- \Delta_{\varepsilon,c} - \frac{\lambda_0}{\varepsilon^2
M^2} \mathbf 1 \right)^{-1}
- \left(T_{\varepsilon,c}^{-1} \oplus 0\right) \right| \right| \leq C_{10}\,
\varepsilon^{3/2}.$$
The proof of the theorem is complete. $\hfill \rule {2.5mm} {3mm}\vspace{0.5cm}$
}
\end{dm}
\section{Bounded interval and Dirichlet condition}\label{qqq3}
In this section we suppose that $I=[-a,b]$ is a bounded interval and the condition at
the boundary $\partial \Lambda_\varepsilon$ is
Dirichlet.
Since $I$ is bounded, the spectrum of
$-\Delta_{\varepsilon,c}$ in $\Lambda_\varepsilon$ is purely discrete and we
denote its eigenvalues by
$l_j^c(\varepsilon)$.
The main result in this section, that is,
Theorem~\ref{aaa3}, is a version of Theorem~\ref{mainTeor} in this context.
\begin{teo}\label{aaa3}
The limits
\begin{equation}\label{eee3new}
\mu_j = \lim_{\varepsilon \to 0}\, \varepsilon\left( l_j^c(\varepsilon) -
\frac{\lambda_0}{\varepsilon^2 M^2} \right)
\end{equation}
exist, where $\mu_j$ are the eigenvalues of a self-adjoint operator~$T$ (see Equation~\eqref{eqweo}) acting in
$\mathrm L^2(\mathbb R)$.
\end{teo}
To prove this theorem we need some previous results; we will
follow~\cite{Frie}. Introduce the family of segments
$$I_\varepsilon = (-a \varepsilon^{-1/2}, b \varepsilon^{-1/2}),
\qquad\, \varepsilon > 0,$$ and the family of
unitary operators $J_\varepsilon: \mathrm L^2(I) \to \mathrm L^2 (I_\varepsilon)$
generated by the dilation
$s \mapsto s \varepsilon^{1/2}$, that is,
\[
(J_\varepsilon\psi)(s) = \varepsilon^{1/4}\, \psi(\varepsilon^{1/2}s),
\]and identify
$\mathrm L^2(I_\varepsilon)$ with the subspace
$$\left\{u \in \mathrm L^2(\mathbb R): u(s) = 0\, \hbox{ a.e. in }\, \mathbb R \backslash
I_\varepsilon\right \}.$$
Set
\begin{equation}\label{aaa7}
\hat{T}_{\varepsilon,c} : = \varepsilon J_\varepsilon T_{\varepsilon,c}
J_\varepsilon^{-1},
\end{equation}
which is a self-adjoint operator acting in $\mathrm L^2(I_\varepsilon)$.
\
\begin{teo}\label{aaa2}
{\rm In case $I=[-a,b]$ is a bounded interval, one has
$$\left| \left| \hat{T}_{\varepsilon,c}^{-1} \oplus 0 - T^{-1} \right| \right|
\to 0, \qquad \hbox{as}
\qquad
\varepsilon \to 0,$$
where $0$ is the null operator on the subspace $\mathrm L^2(\mathbb R
\backslash I_\varepsilon)$.}
\end{teo}
\
We have $\varepsilon J_\varepsilon W_\varepsilon(s) J_\varepsilon^{-1} = \varepsilon W_\varepsilon(\varepsilon^{1/2} s)$, and a direct calculation shows that
$$\varepsilon J_\varepsilon W_\varepsilon(s) J_\varepsilon^{-1} = \zeta_\varepsilon(\varepsilon^{1/2} s, y) \lambda_0 \left[
M^{-3} s^2
+ \rho (\varepsilon^{1/2} s) s^3
\varepsilon^{1/2} \right] + \varepsilon\, \vartheta(\varepsilon^{1/2} s) +
\varepsilon c,$$
with $\rho \in \mathrm L^\infty(I)$.
Since
$\zeta_\varepsilon(\varepsilon^{1/2} s, y) \to 1$ uniformly as
$\varepsilon \to 0$,
the proof of Theorem~\ref{aaa2} is similar to the proof of
Theorem~$1.3$ in \cite{Frie}, and so it will not be repeated here.
\
\noindent{\bf Proof of Theorem~\ref{aaa3}:}
Let $l_j(T_{\varepsilon,c})$, $l_j(\hat{T}_{\varepsilon,c})$ denote the
eigenvalues of
$T_{\varepsilon,c}$ and $\hat{T}_{\varepsilon,c}$ respectively.
Let $\psi_{j, \varepsilon}^c$ denote the eigenfunction associated with eigenvalue $l_j^c(\varepsilon)$
of $-\Delta_{\varepsilon, c}$.
Thus, there exist functions $w_{\varepsilon, c} \in \mathrm L^2(I)$ and $U \in \cal L $ so that
$\psi_{j, \varepsilon}^c = w_{j, \varepsilon}^c u_0 + U$. Since $\mathcal L$ is invariant under
$(-\Delta_{\varepsilon, c} - \lambda_0 / \varepsilon^2 M^2 \mathbf 1)$, it
follows that $w_{j, \varepsilon}^c u_0$ is the eigenfunction associated with the
eigenvalue $l_j(T_{\varepsilon, c})$.
Observe also that the nonzero
eigenvalues of
$T_{\varepsilon,c}^{-1} \oplus 0$ are exactly the eigenvalues of
$T_{\varepsilon,c}^{-1}$.
Hence, by Theorem~\ref{aaa4}, we have
\begin{eqnarray*}
\left| \left( l_j^c(\varepsilon) - \frac{\lambda_0}{\varepsilon^2 M^2}
\right)^{-1} - l_j^{-1}(T_{\varepsilon,c}) \right|
& \leq &
\left| \left| \left( -\Delta_{\varepsilon,c} -
\frac{\lambda_0}{\varepsilon^2 M^2}\mathbf 1 \right)^{-1} -
T_{\varepsilon,c}^{-1} \oplus 0 \right| \right| \\
& \leq &
C_{10}\, \varepsilon^{3/2}.
\end{eqnarray*}
Thus,
$$\left| \frac{1}{\varepsilon} \left( l_j^c(\varepsilon) -
\frac{\lambda_0}{\varepsilon^2 M^2} \right)^{-1}-
\frac{1}{\varepsilon l_j^{-1}(T_{\varepsilon,c})} \right| \leq C_{10}\,
\varepsilon^{1/2}.$$
Since $l_j(\hat{T}_{\varepsilon,c}) = \varepsilon l_j(T_{\varepsilon,c})$, by
Theorem~\ref{aaa2}, we find
$$\varepsilon l_j(T_{\varepsilon,c}) \to \mu_j, \quad\varepsilon \to 0,
$$
and~\eqref{eee3new} follows.
$\hfill \rule {2.5mm} {3mm}\vspace{0.5cm}$
\section{The Neumann case}\label{nnnn}
Here we again consider that $I=[-a,b]$ is a bounded interval, but the Dirichlet
condition at the vertical part of
the boundary $\partial(I \times S)$, that is, $\{(-a)\times S\cup b\times S\}$,
is replaced by Neumann condition.
Our point is that the conclusions of Theorem~\ref{aaa3} also hold true in
this case.
Although in our case the curvature and torsion can be nontrivial, the
proof in this case are similar to the proof of
Theorem~\ref{aaa3} above (and taking into account~\cite{Sol}); for this reason,
details will not be presented.
\section{The case $I = \mathbb R$ and Dirichlet condition}\label{iiii}
In this section we study the case $I = \mathbb R$. First we give sufficient conditions for a nonempty discrete spectrum of the Dirichlet Laplacian, and then discuss the WEO and eigenvalue approximations.
\subsection{The discrete spectrum}\label{udah}
Now the spectrum of the Laplacian $- \Delta_{\varepsilon,c}$ in
$\Lambda_\varepsilon$ is not
necessarily discrete, but in this section we will see that the essential spectrum $\sigma_{\mathrm{ess}}(- \Delta_{\varepsilon,c})$ depends on the behavior of the curvature at infinity; it will then follow that if $k(s)\to 0$ as $|s|\to\infty$, then the discrete spectrum of $- \Delta_{\varepsilon,c}$ is nonempty for $\varepsilon$ small enough.
Denote $\nu(\varepsilon) := \inf \sigma_{{\rm ess}}
(-\Delta_{\varepsilon,c})$ and let
$l_j^c(\varepsilon)$ be the eigenvalues of $-\Delta_{\varepsilon,c}$ (recall
the Dirichlet boundary condition).
\begin{teo}
If $I=\mathbb R$ and the curvature satisfies
\begin{equation}\label{eee5}
\lim_{|s| \to \infty} k(s) =0,
\end{equation}
then $\nu(\varepsilon) \to \infty$ as $\varepsilon \to 0.$
\end{teo}
\begin{dm}{\rm
Let $N : = \limsup_{|s| \to \infty} h(s) < M$ and $\hat{I} = [-a, a]$
and define
$$\Omega_{a, \varepsilon} =\left \{(s, y): s \in \hat{I} \right\} \qquad
\hbox{and} \qquad
\Omega_{a, \varepsilon}' = \left\{(s, y): s \notin \hat{I}\right \}.$$
Let $-\Delta_{a, \varepsilon, D}^c$, $-\Delta_{a, \varepsilon, D}^{'c}$ be
the Dirichlet Laplacian in
$\Omega_{a, \varepsilon}$ and $\Omega_{a, \varepsilon}'$ respectively.
Similarly, let
$-\Delta_{a, \varepsilon, DN}^c$, $-\Delta_{a, \varepsilon, DN}^{'c}$ be
the above Laplacian operators but with Neumann condition at the
vertical part of the boundaries of~$\Omega_{\alpha,\varepsilon}$ and~$\Omega_{\alpha,\varepsilon}^{'}$, respectively.
Note that
\begin{equation}\label{eee6}
- \Delta_{a, \varepsilon, DN}^c + \left(- \Delta_{a, \varepsilon,
DN}^{'c}\right)
< - \Delta_{\varepsilon, c} <
- \Delta_{a, \varepsilon, D}^c + \left(- \Delta_{a, \varepsilon,
D}^{'c}\right).
\end{equation}
Therefore
$\inf \sigma_{{\rm ess}} (-\Delta_{\varepsilon, c}) \geq \inf \sigma_{{\rm
ess}} (- \Delta_{a, \varepsilon, DN}^{'c})$.
Let $q_{a, \varepsilon, DN}'$ the quadratic form associated with the
operator $-\Delta_{a, \varepsilon, DN}^{'c}$.
Write $K_\varepsilon= \sup_{(s,y) \in \mathbb R \times S}
\beta_\varepsilon(s, y)$; we have
\begin{eqnarray*}
q_{a, \varepsilon, DN}' (\psi)
& \geq &
\left(
\inf_{(s,y) \in \mathbb R \times S}{\frac{\beta_\varepsilon(s,
y)}{\varepsilon^2 h(s)}} \right) \int_{(\mathbb R \backslash \hat{I})
\times S}
|\nabla_y \psi|^2 dy ds \\
& \geq &
\lambda_0 \left(
\inf_{(s,y) \in \mathbb R \times S}{\frac{\beta_\varepsilon(s,
y)}{\varepsilon^2 h(s)}} \right) \int_{(\mathbb R \backslash
\hat{I})\times S}
|\psi|^2 dy ds \\
& \geq &
\lambda_0
\left(\inf_{(s,y) \in \mathbb R \times S}{\frac{\beta_\varepsilon(s,
y)}{\varepsilon^2 h(s)}} \right)
\frac{1}{K_\varepsilon}
\int_{(\mathbb R \backslash \hat{I})\times S}
\beta_\varepsilon(s, y) |\psi|^2 dy ds,
\end{eqnarray*}
for all $\psi \in \mathrm {dom}\, q_{a, \varepsilon, DN}'$.
Since $k$ satisfies \eqref{eee5},
it follows that the essential spectrum of $-\Delta_{a, \varepsilon, DN}^{'c}$
is estimated from below by $\lambda_0$ times a function that converges to
$ \frac{1}{\varepsilon^2 N}$ as $a \to \infty$.
Since the essential spectrum is a closed subset, it follows that
$\nu(\varepsilon) \geq \frac{\lambda_0}{\varepsilon^2 N^2}$ and
consequently
$\nu(\varepsilon) \to \infty$ as $\varepsilon \to 0$.
$\hfill \rule {2.5mm} {3mm}\vspace{0.5cm}$
}
\end{dm}
We conclude that, under condition~\eqref{eee5}, for $\varepsilon$ small enough the discrete spectrum of
$-\Delta_{\varepsilon, c}$
is nonempty. We again stress that we have got another property that does
not depend on important geometric features of the tube.
Also the WEO~$T$ (see also Subsection~\ref{subsecEo}),
which weakly describes the asymptotic behaviors of the eigenvalues of
$-\Delta_{\varepsilon, c}$ in the sense of~\eqref{eee3new}, is not influenced by such geometric features.
\
\subsection{Weakly effective operator}\label{subsecEo}
The goal of this section is to show that Theorems~\ref{aaa4}, \ref{aaa3}
and~\ref{aaa2} have a similar counterpart in case~$I = \mathbb R$.
In \cite{Sol} these theorems are proven for two dimensional
strips, and here we argue that those proofs can be adapted to our three
dimensional setting. The proof of Lemma~\ref{bbb1} will be postponed to
the end of this subsection.
\begin{lem}\label{bbb1}
{\rm
There exists $C_6 > 0$ so that, for $\varepsilon$ small enough,
$$t_\varepsilon(w) \geq C_6^{-1}\, \varepsilon^{-1} \int_{\mathbb R} |w|^2
ds, \qquad \forall w\in {\mathcal H}_0^1(\mathbb R).$$
}
\end{lem}
The proof of the next theorem is similar to the proof of
Theorem~\ref{aaa4}; it is enough to take into account Lemma~\ref{eee4}, and then Lemma~\ref{bbb1}
instead of Lemma~\ref{ddd10}. Recall that ${\mathcal L}$ is the subspace generated by functions
$w(s) u_0(y)$ with $w \in \mathrm L^2(\mathbb R)$
\begin{teo}{\rm
Let $I=\mathbb R$. Then,
there exists $C_{10} > 0$ so that, for $\varepsilon$ small enough,
$$\left|\left| \left(- \Delta_{\varepsilon,c} -
\frac{\lambda_0}{\varepsilon^2 M^2} \mathbf 1 \right)^{-1}
- \left(T_{\varepsilon,c}^{-1} \oplus 0\right) \right| \right| \leq C_{10}\,
\varepsilon^{3/2},$$
where $0$ denotes the null operator on the subspace ${\mathcal L}^\perp$.
}
\end{teo}
As in the previous section, consider the self-adjoint operators
$$\hat{T}_{\varepsilon,c}:= \varepsilon J_\varepsilon T_{\varepsilon,c}
J_\varepsilon^{-1},$$
where $J_\varepsilon: \mathrm L^2(\mathbb R) \to \mathrm L^2(\mathbb R)$ is the previously discussed
unitary operator generated by the dilation
$s\mapsto s \varepsilon^{1/2}$.
\begin{teo}\label{ddd9}
{\rm
For $\varepsilon \to 0$ one has
$$\left| \left| \hat{T}_{\varepsilon,c}^{-1} - T^{-1} \right| \right| \to 0,
$$
where $T$ is the operator~\eqref{eqweo}.
}\end{teo}
As in the Section~\ref{qqq3},
we have that $\varepsilon J_\varepsilon W_\varepsilon(s) J_\varepsilon^{-1}$ equals
$$
\lambda_0 \,\zeta_\varepsilon(\varepsilon^{1/2} s, y) \left[ M^{-3} s^2+ \rho
(\varepsilon^{1/2} s) s^3
\varepsilon^{1/2} \right] + \varepsilon\, \vartheta(\varepsilon^{1/2} s) +
\varepsilon c.$$
Again, since
$\zeta_\varepsilon(\varepsilon^{1/2} s, y) \to 1$ uniformly as
$\varepsilon \to 0$,
the proof of Theorem~\ref{ddd9}
is similar to the proof of Theorem~$1.3$ in \cite{Frie}, and details will
be skipped.
\
\noindent{\bf Proof of Lemma~\ref{bbb1}:}
Theorem~\ref{ddd9} guarantees that
$$\varepsilon^{-1} \left| \left| T_{\varepsilon,c}^{-1} \right| \right| \to
\left| \left| T^{-1} \right| \right|
\qquad (\varepsilon \to 0),$$
and so there exists $C_6> 0$ so that
$$\left| \left| T_{\varepsilon,c}^{-1} \right| \right| \leq C_6
\,\varepsilon.$$ The proof is complete.
$\hfill \rule {2.5mm} {3mm}\vspace{0.5cm}$
|
\section{Introduction}
It is well-known that the long-time behaviour of solutions of partial
differential equations arising in mathematical physics can, in many cases,
be described in terms of global attractors of the associated semigroups (see
\cite{BV, CV, Ha, T} and references therein). For a large class of equations
of mathematical physics, including parabolic partial differential equations
modelling reaction, diffusion and drift, hyperbolic type equations, and so
on, the corresponding attractor has finite Hausdorff and fractal dimensions.
Thus, the dynamics on the attractor happens to be finite-dimensional, even
though the system is governed by a set of partial differential equations. As
the dimension of the attractor is indicative of the number of degrees of
freedom needed to simulate a given dynamical system, it is then crucial to
obtain more realistic estimates for its dimension in terms of observable
physical quantities.
Aside from some applied motivation, much of the mathematical interest
nowadays is centered on the dynamics of boundary value problems with \emph
static }boundary conditions of Dirichlet and Neumann-Robin type, or even
periodic boundary conditions. The influence of these dissipative boundary
conditions on a given model has only been recently investigated in
connection with a class of reaction-diffusion systems. In \cite{HR}, a first
contribution is made to the understanding of this problem with a Robin
boundary condition. In particular, it is shown, for a fixed nonlinearity,
how the flow defined by the reaction-diffusion system depends on the
interaction between diffusion $\nu $\ and another parameter $\theta $\
involved in the boundary condition (cf. also \cite{HR2}). A classification
of points in $\left( \nu ,\theta \right) $-space, as structurally stable, or
bifurcation points, for a one-dimensional scalar reaction-diffusion equation
with a cubic nonlinearity is discussed in detail in \cite{HR}. Other studies
on the influence of boundary conditions upon the solution structures of
partial differential equations have also been done by other scientists.
These studies have analyzed the detailed effect of boundary conditions on
the structure of global attractors (see, e.g., \cite{DMO, HS, MT, SG}). If
the equilibrium is nonhyperbolic and a bifurcation occurs, the structure of
attractors may vary with respect to boundary conditions. This has been
observed in the analysis of pattern formation in a 1D reaction-diffusion
system \cite{DMO}, in lattice systems \cite{S}, in the study of steady state
bifurcations \cite{HS, MT}, and finally in \cite{SG}, on mode-jumping of the
von Karman equations.
Although the global attractors of these systems will depend, for a given
nonlinearity, on the choice of the boundary conditions, their finite
dimension does generally \emph{not}. This result can be easily formulated
for a scalar reaction-diffusion equation, as follows. Consider the parabolic
partial differential equatio
\begin{equation}
\partial _{t}u=\nu \Delta u-f\left( u\right) +\lambda u+g,\text{ }\left(
x,t\right) \in \Omega \times \left( 0,+\infty \right) , \label{she}
\end{equation
where $u=u\left( x,t\right) \in \mathbb{R}$, $\Omega \subset \mathbb{R}^{n}
, $n\geq 1$, is a bounded domain with sufficiently smooth boundary $\Gamma ,$
$g=g\left( x\right) $, and $\nu $, $\lambda $ are positive constants. The
function $f:\mathbb{R\rightarrow R}$ is assumed to be $C^{1,1}$, that is,
continuous and with a Lipschitz continuous first derivative, which
satisfies, among other natural growth conditions (see, e.g., \cite[Chapter I
]{CV})
\begin{equation*}
f^{^{\prime }}\left( y\right) \geq -c_{f},\text{ for all }y\in \mathbb{R
\text{, for some }c_{f}>0.
\end{equation*
We may ask that $u$ satisfy either a Dirichlet ($K=D$) boundary condition or
a Neumann ($K=N$)\ boundary condition, and even a periodic ($K=P$) boundary
condition. It is well-known that equation (\ref{she}), supplemented with an
appropriate initial condition, generates a semigroup $\left\{ S_{t}\right\} $
acting on a suitable phase space $H$. This semigroup possesses the global
attractor $\mathcal{A}_{K}$, which may depend on the choice of the boundary
conditions, and $\mathcal{A}_{K}$ has finite fractal dimension for each
K\in \left\{ D,N,P\right\} $. In particular, the Haussdorf and fractal
dimensions of $\mathcal{A}_{K},$ for any $K\in \left\{ D,N,P\right\} $,
satisfy the following upper and lower bounds
\begin{equation}
c_{0}\left( \frac{\lambda }{\nu }\right) ^{n/2}\left\vert \Omega \right\vert
\leq \dim _{H}\mathcal{A}_{K}\leq \dim _{F}\mathcal{A}_{K}\leq c_{1}\left( 1
\frac{c_{f}+\lambda }{\nu }\left\vert \Omega \right\vert ^{2/n}\right)
^{n/2}, \label{lb1}
\end{equation
for some positive constants $c_{0},c_{1}$ that depend only on $n,$ $f$ and
the shape of $\Omega $ (see, e.g., \cite[Chapter III]{BV}; cf. also \cite{CV
, \cite[Chapter VI]{T}). Here, $\left\vert \Omega \right\vert $ stands for
the Lebesgue measure of $\Omega $. For a fixed domain $\Omega $, we observe
that these estimates are sharp with respect to $\nu \rightarrow 0^{+}$ (for
each fixed $\lambda >0$), or sufficiently large $\lambda $ (for each fixed
\nu >0$). Hence, these bounds for the dimension of $\mathcal{A}_{K}$\ are of
the same order for each $K\in \left\{ D,N,P\right\} $. These remarkable
estimates\ also depend linearly on the "volume" of the spatial domain
\Omega $, which is consistent with physical intuition. This property of the
dimension of the attractor has not been proved for all equations, such as,
the Kuramoto-Sivashinski equation.
Our main goal in this paper is to investigate the dependance of the
dimension of the global attractor for equation (\ref{she}) subject to a
completely new class of boundary conditions, which are sometimes dubbed as
\emph{Wentzell} boundary conditions, and which have some applications in
probability theory, specifically, Markov processes. But what are they
really? To put them into a context, let $L$ be an elliptic differential
operator of the second order (e.g., $L=\nu \Delta $) with coefficients that
are well-defined over $\overline{\Omega }$. It is known that there is a
one-to-one correspondence between $\left( C_{0}\right) $-semigroups and
Markov processes in $\overline{\Omega }$ which are homogeneous in time and
satisfy the condition of Feller \cite{Fe1} (that is, the range of the
resolvent operator coincides with a prescribed set). Thus, to each such
Markov process there is a corresponding semigroup of operator
\begin{equation*}
T_{t}v\left( x\right) =\int\limits_{\overline{\Omega }}v\left( y\right)
P\left( t,x,dy\right) ,
\end{equation*
where the Markov transition function $P\left( t,x,B\right) $ satisfies
P\left( t,x,B\right) \geq 0,$ for $t\geq 0,$ $x\in \overline{\Omega }$ and
any Borel set $B\subseteq \overline{\Omega }$. As a function of $B$,
P\left( t,x,\cdot \right) $ is a probability measure. What are the most
general boundary conditions which restrict the given operator $L$\ (more
correctly, its closure) to the infinitesimal operator of a semigroup of
positive contraction operators acting on $C\left( \overline{\Omega }\right)
? Wentzell \cite{W} gave a partial answer to this question in higher space
dimensions by finding a sufficiently large class of boundary conditions
which involve differential operators on the boundary that are of the same
order as the operator acting in $\Omega $. He discovered the following form
of boundary conditions
\begin{equation}
Lu+\nu b\partial _{\mathbf{n}}^{L}u=0\text{, on }\Gamma \times \left(
0,+\infty \right) , \label{wbc}
\end{equation
where $\mathbf{n}$ denotes the outward normal at $\Gamma $, $b$ is a
positive constant and $\partial _{\mathbf{n}}^{L}u$ is the outward co-normal
derivative of $u$ with respect to $L$. We refer also to the pioneering work
of \cite{Fe2}, for generation theorems for $L$ with Wentzell boundary
conditions in one space dimension. Until the work of \cite{FGGR}, the study
of the operator $L$\ with Wentzell boundary conditions was usually confined
to generation properties of this operator in the space $C\left( \overline
\Omega }\right) $. In 2002, the authors in \cite{FGGR} have found a way to
introduce the Wentzell boundary condition (\ref{wbc}) in an $L^{p}$-context,
which led to the discovery of the natural space for these type of problems
(see Section 2). The reader is referred to \cite{CFGGOR, Gi2} for an
extensive survey of these results and some history.
For the homogeneous linear heat equation (\ref{she}) (that is, $f=g=\lambda
=0$), the Wentzell boundary condition (\ref{wbc})\ is equivalent to a purely
differential equation of the for
\begin{equation}
\partial _{t}u+\nu b\partial _{\mathbf{n}}u=0,\quad \text{on}\;\Gamma \times
\left( 0,\infty \right) . \label{dyn}
\end{equation
Thus, the main attraction here is that there is a dynamic element introduced
into the boundary condition. The heat equation, supplemented by either (\re
{wbc}) or (\ref{dyn}), corresponds to the situation where there is a heat
source (if $b>0$) or sink (if $b<0$) acting on the boundary $\Gamma $.
Mathematically speaking, this kind of conditions (\ref{dyn}) arises due to
the presence of additional boundary terms in the free energy, which must
also account for the action of a source on $\Gamma $\ (see \cite{GW}). We
refer the reader to \cite{Gi} (cf. also \cite{GG0}), for an extensive
derivation and physical interpretation of (\ref{dyn}) for (\ref{she}). For
the nonlinear parabolic equation (\ref{she}), the boundary condition (\re
{wbc}) can be formally be transformed into a condition of the for
\begin{equation}
\partial _{t}u+\nu b\partial _{\mathbf{n}}u+f\left( u\right) -\lambda u=g
\text{ on }\Gamma \times \left( 0,\infty \right) . \label{dyn2}
\end{equation
Generally, one may replace $f-\lambda $ in (\ref{dyn2}) by another arbitrary
function $h$, satisfying suitable assumptions. With more sophisticated
arguments, using techniques from semigroup theory, and a variation of
parameter formula, it is possible to prove that the regularity of the
solution for (\ref{she}),(\ref{dyn2}) increases as $f$, $\Omega $ and $g$
become more regular (see Section 2). In particular, for $g=0$, if $\Omega $
is a bounded $\mathcal{C}^{\infty }$ domain and $f$ is a $\mathcal{C
^{\infty }$ function, regularity theory implies that the solution $u\left(
t\right) $ to (\ref{she}),(\ref{dyn2}) belongs $H^{k}\left( \Omega \right) ,$
for all $k\geq 0$ and all positive times. At least in this case, the
boundary condition (\ref{wbc}), for equation (\ref{she}), is equivalent to
the boundary condition (\ref{dyn2}). However, in general, this may not be so
if the solution, for the semilinear problem (\ref{she}) and the Wentzell
condition (\ref{dyn2}), is not smooth enough. Since we wish to treat the
most general case, by imposing the least regularity assumptions on $f,$ $g$
and $\Omega ,$ we will devote our attention only to the study of (\ref{she
), subject to linear boundary conditions of the form (\ref{dyn}). Our
results below can be immediately extended to other classes of nonlinear
Wentzell boundary conditions (see, e.g., \cite{GW} and references therein).
Boundary conditions of the form (\ref{dyn2}) arise for many known equations
of mathematical physics. They are motivated by heat control problems
formulated in the book of Duvaut and Lions \cite{DL}, problems in
phase-transition phenomena \cite{CGGM, G, GG1, GM, GM2, GMS, GMS2, MZ1, MZ2}
(and their references), special flows in hydrodynamics \cite{FL, GW, QWS, MW
, Stefan problems \cite{A, Ma, RSY}, models in climatology \cite{MW2}, and
many others. The reader is referred to \cite{GG_b} for a more complete list
of references involving the application of such boundary conditions to
real-world phenomena.
By keeping our treatment of the boundary condition simple, we wish to prove
that the problem (\ref{she}), (\ref{dyn}) generates a dynamical system on a
suitable phase-space, possessing a finite dimensional global attractor
\mathcal{A}_{W}.$ Then, we establish that the Haussdorf and fractal
dimensions of $\mathcal{A}_{W}$ satisfy the following \ upper and lower
bounds
\begin{equation}
c_{1}\left( \frac{\lambda }{C_{W}\left( \Omega ,\Gamma \right) \nu }\right)
^{n-1}\leq \dim _{H}\mathcal{A}_{W}\leq \dim _{F}\mathcal{A}_{W}\leq
c_{2}\left( 1+\frac{c_{f}+\lambda }{C_{W}\left( \Omega ,\Gamma \right) \nu
\right) ^{n-1}, \label{lb2}
\end{equation
for $n\geq 2,$ an
\begin{equation}
c_{3}\left( \frac{\lambda }{C_{D}\left( \Omega \right) \nu }\right)
^{1/2}\leq \dim _{H}\mathcal{A}_{W}\leq \dim _{F}\mathcal{A}_{W}\leq
c_{4}\left( 1+\frac{c_{f}+\lambda }{C_{D}\left( \Omega \right) \nu }\right)
^{1/2}, \label{lb3}
\end{equation
in one space dimension. The positive constants $c_{i},$ $i=1,...,4,$ depend
only on $n,$ $f$ and the shape of $\Omega ,$ while explicit estimates and
formulas for $C_{W}\left( \Omega ,\Gamma \right) $ and $C_{D}\left( \Omega
\right) ,$ respectively, are provided in the Appendix. We note again that,
for a fixed domain $\Omega $, these estimates are sharp with respect to $\nu
\rightarrow 0^{+}$ (for each fixed $\lambda >0$), and for sufficiently large
$\lambda $ (if $\nu >0$ is fixed). We remark that the bounds we obtain in
\ref{lb2})-(\ref{lb3}) are quite simple and their explicit dependance on the
physical parameters is transparent. Moreover, a careful analysis of the
constants involved in (\ref{lb2}) yields the following more explicit
two-sided estimate
\begin{equation}
c_{1}^{^{\prime }}\left( \frac{\lambda }{\nu b}\right) ^{n-1}\left\vert
\Gamma \right\vert \leq \dim _{H}\mathcal{A}_{W}\leq \dim _{F}\mathcal{A
_{W}\leq c_{2}^{^{\prime }}\left( 1+\frac{c_{f}+\lambda }{\nu b}\left\vert
\Gamma \right\vert ^{1/\left( n-1\right) }\right) ^{n-1}, \label{lb4}
\end{equation
in all space dimensions $n\geq 3$. It is worth pointing out that the bounds
in (\ref{lb4}) are proportional to the "surface area" $\left\vert \Gamma
\right\vert $ of $\Gamma ,$ and \emph{not} the "volume" $\left\vert \Omega
\right\vert $\ of $\Omega .$ This is remarkable; most nonlinear equations
arising in mathematical physics, involving the Laplacian on bounded domains,
have the dimension of the attractor of the order of $\left\vert \Omega
\right\vert ^{\alpha },$ for some $\alpha >0$ and for sufficiently large
domains. This property may have profound implications in the prediction of
weather and climate. The reader is referred to Section 4 where this
interesting physical observation is further discussed for the balance
equations governing the large-scale oceanic motion.
Our paper is organized as follows. In Section \ref{ub}, we obtain upper
bounds (cf. Theorem \ref{main1})\ for the fractal dimension of the global
attractor for equation (\ref{she}) with dynamic boundary conditions of the
form (\ref{dyn}). In Section \ref{lbb}, we employ the same technique of \cit
{BV} to derive a lower bound for the dimension of the unstable manifold of a
constant stationary solution $u^{\ast }$ of (\ref{she}), (\ref{dyn}). As a
consequence, we find a lower bound on the dimension of $\mathcal{A}_{W}$
(see Theorem \ref{lbbb}). Finally, in the Appendix, we recall some useful
results on the so-called Wentzell Laplacian, and prove an auxiliary
inequality, namely, we derive some kind of Sobolev-Lieb-Thirring inequality
that is required to prove the upper bound in (\ref{lb2}).
\section{Upper bounds on the dimension}
\label{ub}
We use the standard notation and facts from the dynamic theory of parabolic
equations (see, for instance, \cite{CFGGOR}, \cite{FGGR}, \cite{GG1}, \cit
{GW}). We denote by $\left\Vert \cdot \right\Vert _{p}$ and $\left\Vert
\cdot \right\Vert _{p,\Gamma },$ the norms on $L^{p}\left( \Omega \right) $
and $L^{p}\left( \Gamma \right) ,$ respectively. In the case $p=2$, $\langle
\cdot ,\cdot \rangle _{2}$ stands for the usual scalar product. The norms on
$H^{r}\left( \Omega \right) $ and $H^{r}\left( \Gamma \right) $ are
indicated by $\left\Vert \cdot \right\Vert _{H^{r}\left( \Omega \right) }$
and $\left\Vert \cdot \right\Vert _{H^{r}\left( \Gamma \right) }$,
respectively, for any $r>0$.
The natural phase-space for problem (\ref{she}), (\ref{dyn}) i
\begin{equation*}
\mathbb{X}^{p}:=L^{p}(\Omega )\oplus L^{p}(\Gamma )=\{F=\binom{f}{g}:f\in
L^{p}(\Omega ),\;g\in L^{p}(\Gamma )\},
\end{equation*
for all $p\in \left[ 1,\infty \right] $, endowed with the nor
\begin{equation}
\left\Vert F\right\Vert _{\mathbb{X}^{p}}^{p}=\int_{\Omega }\left\vert
f\left( x\right) \right\vert ^{p}dx+\int_{\Gamma }\left\vert g(x)\right\vert
^{p}\frac{dS}{b},\text{ }b>0, \label{xp}
\end{equation
if $p\in \lbrack 1,\infty ),$ and
\begin{equation*}
\Vert F\Vert _{\mathbb{X}^{\infty }}:=\Vert f\Vert _{L^{\infty }(\Omega
)}+\Vert g\Vert _{L^{\infty }(\Gamma )}.
\end{equation*
In the definition of $\mathbb{X}^{p}$, $dx$ denotes the Lebesgue measure on
\Omega ,$ and $dS$ denotes the natural surface measure on $\Gamma $.
Moreover, we have \cite{FGGR}
\begin{equation*}
\mathbb{X}^{p}=L^{p}\left( \overline{\Omega },d\mu \right) ,\text{ }p\in
\left[ 1,\infty \right] ,
\end{equation*
where the measure $d\mu =dx_{\mid \Omega }\oplus \frac{dS}{b}_{\mid \Gamma
}, $ on $\overline{\Omega },$ is defined for any measurable set $B\subset
\overline{\Omega }$ by $\mu (B)=|B\cap \Omega |+\left\vert B\cap \Gamma
\right\vert $. The Dirichlet trace map $\mathit{Tr}_{D}:C^{\infty }\left(
\overline{\Omega }\right) \rightarrow C^{\infty }\left( \Gamma \right) ,$
defined by $\mathit{Tr}_{D}\left( u\right) =u_{\mid \Gamma }$ extends to a
linear continuous operator $\mathit{Tr}_{D}:H^{r}\left( \Omega \right)
\rightarrow H^{r-1/2}\left( \Gamma \right) ,$ for all $r>1/2$, which is onto
for $1/2<r<3/2.$ This map also possesses a bounded right inverse $\mathit{Tr
_{D}^{-1}:H^{r-1/2}\left( \Gamma \right) \rightarrow H^{r}\left( \Omega
\right) $ such that $\mathit{Tr}_{D}\left( \mathit{Tr}_{D}^{-1}\psi \right)
=\psi ,$ for any $\psi \in H^{r-1/2}\left( \Gamma \right) $. Identifying
each function $v\in C\left( \overline{\Omega }\right) $ with the vector $V
\binom{v}{\mathit{Tr}_{D}\left( v\right) }\in C\left( \overline{\Omega
\right) \times C\left( \Gamma \right) $, it follows that $C\left( \overline
\Omega }\right) $ is a dense subspace of $\mathbb{X}^{p},$ for every $p\in
\lbrack 1,\infty ),$ and a closed subspace of $\mathbb{X}^{\infty }.$
Finally, we can also introduce the subspaces of $H^{r}\left( \Omega \right)
\times H^{r-1/2}\left( \Gamma \right) ,
\begin{equation*}
\mathbb{V}_{r}:=\left\{ \binom{u}{\psi }\in H^{r}\left( \Omega \right)
\times H^{r-1/2}\left( \Gamma \right) :\mathit{Tr}_{D}\left( u\right) =\psi
\right\} ,
\end{equation*
for every $r>1/2,$ and note that we have the following dense and compact
embeddings $\mathbb{V}_{r_{1}}\subset \mathbb{V}_{r_{2}},$ for any
r_{1}>r_{2}>1/2$. The linear subspace $\mathbb{V}_{r}$ is densely and
compactly embedded into $\mathbb{X}^{2},$ for any $r>1/2$. We emphasize that
$\mathbb{V}_{r}$ is not a product space and that, due to the boundedness of
the trace operator $\mathit{Tr}_{D},$ $\mathbb{V}_{r}$ is topologically
isomorphic to $H^{r}\left( \Omega \right) $ in the obvious way.
We begin by stating all the hypotheses on $f$ and $g$ that we need. We
assume that $g\in L^{2}\left( \Omega \right) $ and the following conditions
for $f\in C^{1}\left( \mathbb{R}\text{,}\mathbb{R}\right) $ hold
\begin{equation}
f^{^{\prime }}\left( y\right) >-c_{f},\text{ for all }y\in \mathbb{R}\text{,}
\label{n1}
\end{equation
\begin{equation}
\eta _{1}\left\vert y\right\vert ^{p}-C_{f}\leq f\left( y\right) y\leq \eta
_{2}\left\vert y\right\vert ^{p}+C_{f}, \label{n2}
\end{equation
for some $\eta _{1}$, $\eta _{2}>0,$ $C_{f}\geq 0$ and $p>2$.
We have the following rigorous notion of weak solution to (\ref{she}), (\re
{dyn}), with initial condition $u\left( 0\right) =u_{0},$ as in \cite{GW}.
\begin{definition}
\label{weak}The pair $U\left( t\right) =\binom{u\left( t\right) }{\psi
\left( t\right) }$ is said to be a weak solution if $\psi \left( t\right)
\mathit{Tr}_{D}\left( u\right) $ for almost all $t\in \left( 0,T\right) ,$
for any $T>0$, and $U$ fulfill
\begin{align}
U& \in C\left( \left[ 0,T\right] ;\mathbb{X}^{2}\right) \cap L^{\infty
}\left( 0,T;\mathbb{V}_{1}\right) \cap L^{p}\left( \Omega \times \left(
0,T\right) \right) , \label{reg} \\
u& \in H_{loc}^{1}(0,\infty ;L^{2}\left( \Omega \right) ),\text{ }\psi \in
H_{loc}^{1}(0,\infty ;L^{2}\left( \Gamma \right) ), \notag \\
\partial _{t}U& \in L^{2}\left( 0,T;\mathbb{V}_{1}^{\ast }\right) , \notag
\end{align
such that the identit
\begin{equation}
\left\langle \partial _{t}U,\Xi \right\rangle _{\mathbb{X}^{2}}+\nu
\left\langle \nabla u,\nabla \sigma \right\rangle _{2}+\left\langle f\left(
u\right) -\lambda u,\sigma \right\rangle _{2}=\left\langle g,\sigma
\right\rangle _{2}, \notag
\end{equation
holds almost everywhere in $\left( 0,T\right) $, for all $\Xi =\binom{\sigma
}{\varpi }\in \mathbb{V}_{1}$. Moreover, we have, in the space $\mathbb{X
^{2}$
\begin{equation}
U\left( 0\right) =\binom{u_{0}}{v_{0}}=:U_{0}, \label{ini}
\end{equation
where $u\left( 0\right) =u_{0}$ almost everywhere in $\Omega $, and $v\left(
0\right) =v_{0}$ almost everywhere in $\Gamma $. Note that in this setting,
v_{0}$ need not be the trace of $u_{0}$ at the boundary. Thus, in this
context equation (\ref{dyn}) is interpreted as an additional parabolic
equation, acting now on the boundary $\Gamma $.
\end{definition}
The following result is a direct consequence of results contained in \cite
Section 2]{GW}. The proof is based on the application of a Galerkin
approximation scheme which is not standard due to the nature of the boundary
conditions (see, also, \cite{CGGM}).
\begin{theorem}
\label{weaksol}Let the assumptions of (\ref{n1}), (\ref{n2}) be satisfied.
For any given initial data $U_{0}\in \mathbb{X}^{2},$ the problem (\ref{she
), (\ref{dyn}), (\ref{ini}) has a unique weak solution which depends
continuously on the initial data in a Lipschitz way. The following estimate
holds:
\begin{align}
& \left\Vert U\left( t\right) \right\Vert _{\mathbb{X}^{2}}^{2}+\in
\limits_{t}^{t+1}\left( \left\Vert U\left( s\right) \right\Vert _{\mathbb{V
_{1}}^{2}+\left\Vert u\left( s\right) \right\Vert _{L^{p}\left( \Omega
\right) }^{p}\right) ds \label{diss} \\
& \leq c\left( \left\Vert U\left( 0\right) \right\Vert _{\mathbb{X
^{2}}^{2}\right) e^{-\rho t}+c\left( 1+\left\Vert g\right\Vert _{L^{2}\left(
\Omega \right) }^{2}\right) , \notag
\end{align
for all $t\geq 0$, where the positive constants $c$, $\rho $ are independent
of time and initial data.
\end{theorem}
As a consequence, problem (\ref{she}), (\ref{dyn}), (\ref{ini}) defines a
(nonlinear) continuous semigroup $\mathcal{S}_{t}$ acting on the phase-space
$\mathbb{X}^{2}$
\begin{equation*}
\mathcal{S}_{t}:\mathbb{X}^{2}\rightarrow \mathbb{X}^{2},\text{ }t\geq 0,
\end{equation*
given b
\begin{equation*}
\mathcal{S}_{t}U_{0}=U\left( t\right) .
\end{equation*}
\begin{theorem}
\label{attr}Let $f$ satisfy assumptions (\ref{n1}), (\ref{n2}), let $g\in
L^{\infty }\left( \Omega \right) $ and $\Gamma \in \mathcal{C}^{2}$. Then,
\left\{ \mathcal{S}_{t}\right\} $ possesses the connected global attractor
\mathcal{A}_{W},$ which is a bounded subset of $\mathbb{V}_{2}\cap \mathbb{X
^{\infty }$. As a consequence, the global attractor contains only strong
solutions.
\end{theorem}
\begin{proof}
The existence of an absorbing set in $\mathbb{V}_{1}\cap L^{p}\left( \Omega
\right) $ and, hence, the existence of the global attractor $\mathcal{A
_{W}\subset \mathbb{V}_{1}$ follows from \cite[Theorem 2.8 and Corollary 3.1
]{GW}. We will now show that the attractor is bounded in $\mathbb{X}^{\infty
}$, and also in $\mathbb{V}_{2}$. All the calculations below are formal.
However, they can be rigorously justified by means of the approximation
procedures devised in \cite{GW} and \cite{GG0} (cf. \cite{CGGM} also). From
now on, $c$ will denote a positive constant that is independent of time and
initial data, which only depends on the other structural parameters of the
problem, that is, $\left\vert \Omega \right\vert ,$ $\left\vert \Gamma
\right\vert ,$ $\eta _{i},$ $\nu ,$ $b,$ $\left\Vert g\right\Vert _{\infty }$
and $n$. Such a constant may vary even from line to line.
\noindent \textbf{Step 1}. We will first establish the existence of a
bounded absorbing set in $\mathbb{X}^{\infty }$. First note that by (\re
{diss}), there is a constant $C_{0}>0,$ independent of time and initial
data, such that for any bounded subset $B$ of $\mathbb{X}^{2}$, $\exists $
\tau =\tau \left( \left\Vert B\right\Vert _{\mathbb{X}^{2}}\right) >0$ wit
\begin{equation}
\sup_{t\geq \tau }\left\Vert U\left( t\right) \right\Vert _{\mathbb{X
^{2}}\leq C_{0}. \label{diss2}
\end{equation
We shall now perform an Alikakos-Moser iteration argument. We multiply (\re
{she}) by $\left\vert u\right\vert ^{r_{k}-2}u,$ $r_{k}:=2^{k},$ $k\geq 1,$
and integrate over $\Omega $. We obtai
\begin{eqnarray}
&&\frac{1}{r_{k}}\frac{d}{dt}\left\Vert u\right\Vert
_{r_{k}}^{r_{k}}+\left\langle f\left( u\right) ,\left\vert u\right\vert
^{r_{k}-2}u\right\rangle _{2}+\nu \int_{\Omega }\nabla u\cdot \nabla \left(
\left\vert u\right\vert ^{r_{k}-2}u\right) dx \label{eqn2} \\
&=&\nu \int_{\Gamma }\partial _{\mathbf{n}}u\left\vert \psi \right\vert
^{r_{k}-2}\psi dS+\left\langle \lambda u+g,\left\vert u\right\vert
^{r_{k}-2}u\right\rangle _{2}. \notag
\end{eqnarray
Similarly, we multiply (\ref{dyn}) by $\left\vert \psi \right\vert
^{r_{k}-2}\psi /b$ and integrate over $\Gamma $. We hav
\begin{equation}
\frac{1}{br_{k}}\frac{d}{dt}\left\Vert \psi \right\Vert _{r_{k},\Gamma
}^{r_{k}}+\nu \int_{\Gamma }\partial _{\mathbf{n}}u\left\vert \psi
\right\vert ^{r_{k}-2}\psi dS=0. \label{eqn3}
\end{equation
Adding the equalities (\ref{eqn2}), (\ref{eqn3}), we deduc
\begin{align}
& \frac{1}{r_{k}}\frac{d}{dt}\left( \left\Vert U\right\Vert _{\mathbb{X
^{r_{k}}}^{r_{k}}\right) +\left\langle f\left( u\right) ,\left\vert
u\right\vert ^{r_{k}-2}u\right\rangle _{2}+\nu \int_{\Omega }\nabla u\cdot
\nabla \left( \left\vert u\right\vert ^{r_{k}-2}u\right) dx \label{eqn4} \\
& =\left\langle \lambda u+g,\left\vert u\right\vert ^{r_{k}-2}u\right\rangle
_{2}. \notag
\end{align
A simple manipulation of the third integral in (\ref{eqn4}), and employing
assumption (\ref{n2}) on $f$, we readily get the estimate
\begin{align}
& \frac{d}{dt}\left( \left\Vert U\right\Vert _{\mathbb{X}^{r_{k}}}^{r_{k}
\right) +\eta _{1}r_{k}\left\Vert u\right\Vert _{r_{k}+p-2}^{r_{k}+p-2}+\nu
\left( 2^{k}-1\right) 2^{2-k}\left\Vert \nabla \left\vert u\right\vert
^{2^{k-1}}\right\Vert _{2}^{2} \label{eqn5} \\
& \leq r_{k}\left\langle \lambda u+g+C_{f},\left\vert u\right\vert
^{r_{k}-2}u\right\rangle _{2}. \notag
\end{align
Next, using the fact that $\left\vert y\right\vert ^{r_{k}-2}\leq \left\vert
y\right\vert ^{r_{k}}+1,$ for all $k\geq 1$ and $y\in \mathbb{R}$, we
estimate the last term on the right-hand side of (\ref{eqn5})
\begin{equation}
\left\langle \lambda u+g+C_{f},\left\vert u\right\vert
^{r_{k}-2}u\right\rangle _{2}\leq c\left( \left\Vert u\right\Vert
_{r_{k}}^{r_{k}}+1\right) , \label{eqn6}
\end{equation
for some positive constant $c$ that depends on $\lambda $ and the $L^{\infty
}$-norm of $g$, but is independent of $k$. On the other hand, it follows
from Gagliardo-Nirenberg inequality, and Young's inequality for $\varepsilon
\in \left( 0,1\right) ,$ tha
\begin{equation}
\left\Vert v\right\Vert _{2}\leq c\left\Vert v\right\Vert _{H^{1}\left(
\Omega \right) }^{n/\left( n+2\right) }\left\Vert v\right\Vert
_{1}^{1-n/\left( n+2\right) }\leq \varepsilon \left\Vert v\right\Vert
_{H^{1}\left( \Omega \right) }+c\varepsilon ^{-n/2}\left\Vert v\right\Vert
_{1}, \label{estt}
\end{equation
which implie
\begin{equation*}
\left\Vert \nabla v\right\Vert _{2}^{2}\geq \frac{1-\varepsilon }
\varepsilon }\left\Vert v\right\Vert _{2}^{2}-c\varepsilon
^{-n/2-1}\left\Vert v\right\Vert _{1}^{2}.
\end{equation*
Note that the estimate (\ref{estt})\ remains valid if one replaces the
L^{2}\left( \Omega \right) $ and $L^{1}\left( \Omega \right) $-norms by the
L^{2}\left( \Gamma \right) $ and $L^{1}\left( \Gamma \right) $-norms,
respectively, and $n$ by $n-1$, respectively. Setting $v=\left\vert
u\right\vert ^{r_{k-1}}$ in the above inequality, noting that $\left(
2^{k}-1\right) 2^{2-k}\geq 2,$ for each $k,$ and the fact that $Tr_{D}$ maps
$H^{1}\left( \Omega \right) $ boundedly into $L^{2}\left( \Gamma \right) $,
we can estimate the gradient term in (\ref{eqn5}) in terms o
\begin{equation*}
c\frac{1-\varepsilon }{\varepsilon }\left( \left\Vert u\right\Vert
_{r_{k}}^{r_{k}}+\left\Vert \psi \right\Vert _{r_{k},\Gamma }^{r_{k}}\right)
-c\varepsilon ^{-n/2-1}\left( \left\Vert \left\vert u\right\vert
^{r_{k-1}}\right\Vert _{1}^{2}+\left\Vert \left\vert \psi \right\vert
^{r_{k-1}}\right\Vert _{1,\Gamma }^{2}\right) .
\end{equation*
(see, e.g., \cite[Chapter 5]{Ma}). This estimate together with (\ref{eqn5}),
(\ref{eqn6}) yiel
\begin{align}
& \frac{d}{dt}\left( \left\Vert U\right\Vert _{\mathbb{X}^{r_{k}}}^{r_{k}
\right) +c\left( \nu \frac{1-\varepsilon }{\varepsilon }-r_{k}\right) \left(
\left\Vert u\right\Vert _{r_{k}}^{r_{k}}+\left\Vert \psi \right\Vert
_{r_{k},\Gamma }^{r_{k}}\right) \label{eqn7} \\
& \leq c\varepsilon ^{-n/2-1}\left( \left\Vert \left\vert u\right\vert
^{r_{k-1}}\right\Vert _{1}^{2}+\left\Vert \left\vert \psi \right\vert
^{r_{k-1}}\right\Vert _{1,\Gamma }^{2}\right) +cr_{k}, \notag
\end{align
for all $k\geq 1,$ where $c>0$ is independent of $k$.
We shall now make use of an iterative argument to deduce the existence of a
bounded absorbing set in $\mathbb{X}^{r_{k}},$ for all $k\geq 1$. Thus,
noting that $r_{k}\leq \left( r_{k}\right) ^{n/2+1}$, then choosing
\varepsilon =\delta /r_{k}$ with small $\delta =\delta \left( \nu \right) >0$
such tha
\begin{equation*}
\left( \nu \frac{1-\varepsilon }{\varepsilon }-r_{k}\right) \geq r_{k},
\end{equation*
and settin
\begin{equation}
\mathcal{Y}_{k}\left( t\right) :=\int_{\Omega }\left\vert u\left( t,\cdot
\right) \right\vert ^{r_{k}}dx+\int_{\Gamma }\left\vert \psi \left( t,\cdot
\right) \right\vert ^{r_{k}}\frac{dS}{b}=\left\Vert U\right\Vert _{\mathbb{X
^{r_{k}}}^{r_{k}}, \label{def}
\end{equation
from (\ref{eqn7}) we derive the following estimate
\begin{equation}
\partial _{t}\mathcal{Y}_{k}\left( t\right) +cr_{k}\mathcal{Y}_{k}\left(
t\right) \leq c\left( r_{k}\right) ^{n/2+1}\left( \mathcal{Y}_{k-1}\left(
t\right) +1\right) ^{2}. \label{e11}
\end{equation
Let us now take two positive constants $t,$ $\mu $ such that $t-\mu
/r_{k}>0, $ for all $k\geq 1$. Their precise values will be chosen later. We
claim tha
\begin{equation}
\mathcal{Y}_{k}\left( t\right) \leq M_{k}\left( t,\mu \right) :=c\left(
r_{k}\right) ^{n/2+1}(\sup_{s\geq t-\mu /r_{k}}\mathcal{Y}_{k-1}\left(
s\right) +1)^{2} \label{claim}
\end{equation
holds for $\mathcal{Y}_{k},$ defined by (\ref{def}) and $k\geq 1$. To this
end, let $\zeta \left( s\right) $ be a positive function $\zeta :\mathbb{R
_{+}\rightarrow \left[ 0,1\right] $ such that $\zeta \left( s\right) =0$ for
$s\in \left[ 0,t-\mu /r_{k}\right] ,$ $\zeta \left( s\right) =1$ if $s\in
\left[ t,+\infty \right) $ and $\left\vert d\zeta /ds\right\vert \leq Cr_{k}
, if $s\in \left( t-\mu /r_{k},t\right) $. We define $Z_{k}\left( s\right)
=\zeta \left( s\right) \mathcal{Y}_{k}\left( s\right) $ and notice tha
\begin{equation*}
\frac{d}{ds}Z_{k}\left( s\right) \leq cr_{k}Z_{k}\left( s\right) +\zeta
\left( s\right) \frac{d}{ds}\mathcal{Y}_{k}\left( s\right) .
\end{equation*
Combining this estimate with (\ref{e11}) and noticing that $Z_{k}\leq
\mathcal{Y}_{k}$, we deduce the following estimate for $Z_{k}$
\begin{equation}
\frac{d}{ds}Z_{k}\left( s\right) +cr_{k}Z_{k}\left( s\right) \leq
M_{k}\left( t,\mu \right) ,\text{ for all }s\in \left[ t-\mu /r_{k},+\infty
\right) . \label{e12}
\end{equation
Integrating (\ref{e12}) with respect to $s$ from $t-\mu /r_{k}$ to $t$ and
taking into account the fact that $Z_{k}\left( t-\mu /r_{k}\right) =0,$ we
obtain that $\mathcal{Y}_{k}\left( t\right) =Z_{k}\left( t\right) \leq
M_{k}\left( t,\mu \right) \left( 1-e^{-C\mu }\right) $, which proves the
claim (\ref{claim}).
Let now $\tau ^{^{\prime }}>\tau >0$ be given with $\tau $ as in (\ref{diss2
), and define $\mu =2(\tau ^{^{\prime }}-\tau ),$ $t_{0}=\tau ^{^{\prime }}$
and $t_{k}=t_{k-1}-\mu /r_{k},$ $k\geq 1$. Using (\ref{claim}), we hav
\begin{equation}
\sup_{t\geq t_{k-1}}\mathcal{Y}_{k}\left( t\right) \leq c\left( r_{k}\right)
^{n/2+1}(\sup_{s\geq t_{k}}\mathcal{Y}_{k-1}\left( s\right) +1)^{2},\text{
k\geq 1. \label{e13}
\end{equation
Note that from (\ref{diss2}), we have $\left( \sup_{s\geq t_{1}=\tau
\mathcal{Y}_{1}\left( s\right) +1\right) \leq C_{0}+1=:\overline{C}$. Thus,
we can iterate in (\ref{e13}) with respect to $k\geq 1$ and obtain tha
\begin{align*}
\sup_{t\geq t_{k-1}}\mathcal{Y}_{k}\left( t\right) & \leq \left[ c\left(
r_{k}\right) ^{n/2+1}\right] \left[ c\left( r_{k-1}\right) ^{n/2+1}\right]
^{2}\cdot ...\cdot \left[ c\left( r_{1}\right) ^{n/2+1}\right] ^{2^{k}}
\overline{C})^{r_{k}} \\
& \leq c^{A_{k}}2^{B_{k}n/2+1}\left( \overline{C}\right) ^{r_{k}},
\end{align*
wher
\begin{equation}
A_{k}:=1+2+2^{2}+...+2^{k}\leq 2^{k}\sum_{i=1}^{\infty }\frac{1}{2^{i}}
\label{ak}
\end{equation
an
\begin{equation}
B_{k}:=k+2\left( k-1\right) +2^{2}\left( k-2\right) +...+2^{k}\leq
2^{k}\sum_{i=1}^{\infty }\frac{i}{2^{i}}. \label{bk}
\end{equation
Therefore
\begin{equation}
\sup_{t\geq t_{0}}\mathcal{Y}_{k}\left( t\right) \leq \sup_{t\geq t_{k-1}
\mathcal{Y}_{k}\left( t\right) \leq c^{A_{k}}2^{B_{k}\left( n/2+1\right)
\overline{C}^{r_{k}}. \label{e14}
\end{equation
Since the series in (\ref{ak}) and\ (\ref{bk})\ are convergent, we can take
the $r_{k}$-root on both sides of (\ref{e14}) and let $k\rightarrow +\infty
. We deduc
\begin{equation}
\sup_{t\geq t_{0}=\tau ^{^{\prime }}}\left\Vert U\left( t\right) \right\Vert
_{\mathbb{X}^{\infty }}\leq \lim_{k\rightarrow +\infty }\sup_{t\geq
t_{0}}\left( \mathcal{Y}_{k}\left( t\right) \right) ^{1/r_{k}}\leq C_{1},
\label{linf}
\end{equation
for some positive constant $C_{1}$ independent of $t,$ $k$, $U$ and initial
data.
\noindent \textbf{Step 2. }We claim that there is a positive constant
C_{2}, $ independent of time and initial data, and there exists $\tau
^{^{\prime \prime }}>0$ such that
\begin{equation}
\left\Vert U\left( t\right) \right\Vert _{\mathbb{V}_{2}}\leq C_{2},\qquad
\text{for all }\,t\geq \tau ^{^{\prime \prime }}. \label{h2est}
\end{equation
Before we prove (\ref{h2est}), let us recall the following estimate (see
\cite[Theorems 3.5, 3.10]{GW})
\begin{align}
& \sup_{t\geq \tau _{0}}\left( \left\Vert U\left( t\right) \right\Vert _
\mathbb{V}_{1}}^{2}+\left\Vert \partial _{t}u\left( t\right) \right\Vert
_{2}^{2}+\frac{1}{b}\left\Vert \partial _{t}\psi \left( t\right) \right\Vert
_{2,\Gamma }^{2}\right) \label{rec} \\
& +\sup_{t\geq \tau _{0}}\int_{t}^{t+1}\left\Vert \partial _{t}u\left(
s\right) \right\Vert _{H^{1}\left( \Omega \right) }^{2}ds \notag \\
& \leq C_{3}, \notag
\end{align
for some positive constant $C_{3}$ that is independent of time and the
initial data. In order to deduce (\ref{h2est}) from (\ref{rec}) and (\re
{linf}), we need to differentiate (\ref{she}) and (\ref{dyn}) with respect
to time. This yield
\begin{equation}
\partial _{t}^{2}u=\nu \Delta \partial _{t}u-f^{^{\prime }}\left( u\right)
\partial _{t}u+\lambda \partial _{t}u,\text{ }\left( \partial _{t}^{2}\psi
+\nu b\partial _{{\mathbf{n}}}\left( \partial _{t}u\right) \right) _{\mid
\Gamma }=0. \label{diff}
\end{equation
Then, we multiply the first equation of (\ref{diff}) by $\partial
_{t}^{2}u(t)$ and integrate over $\Omega ,$ using the boundary condition of
\ref{diff}). After standard transformations, we obtain
\begin{align*}
& \frac{1}{2}\frac{d}{dt}\left( \left\Vert \nabla \partial _{t}u\left(
t\right) \right\Vert _{2}^{2}\right) +\left\Vert \partial _{t}^{2}u\left(
t\right) \right\Vert _{2}^{2}+\frac{1}{b}\left\Vert \partial _{t}^{2}\psi
\left( t\right) \right\Vert _{2,\Gamma }^{2} \\
& =-\left\langle \left( f^{^{\prime }}\left( u\left( t\right) \right)
-\lambda \right) \partial _{t}u\left( t\right) ,\partial _{t}^{2}u\left(
t\right) \right\rangle _{2}.
\end{align*
Using H\"{o}lder and Young inequalities, we hav
\begin{align*}
& \frac{d}{dt}\left( \left\Vert \nabla \partial _{t}u\left( t\right)
\right\Vert _{2}^{2}\right) +\left\Vert \partial _{t}^{2}u\left( t\right)
\right\Vert _{2}^{2}+\frac{2}{b}\left\Vert \partial _{t}^{2}\psi \left(
t\right) \right\Vert _{2,\Gamma }^{2} \\
& \leq c\left( \left\Vert f^{^{\prime }}\left( u\left( t\right) \right)
\partial _{t}u\left( t\right) \right\Vert _{2}^{2}+\left\Vert \partial
_{t}u\left( t\right) \right\Vert _{2}^{2}\right) \\
& \leq Q\left( \left\Vert u\left( t\right) \right\Vert _{\infty }\right)
\left\Vert \partial _{t}u\left( t\right) \right\Vert _{2}^{2},
\end{align*
for some positive nondecreasing function $Q$ that depends only on $f$ and $c
. This estimate yields, owing to (\ref{linf}), (\ref{rec}),
\begin{equation*}
\frac{d}{dt}\left\Vert \nabla \partial _{t}u\left( t\right) \right\Vert
_{2}^{2}\leq c.
\end{equation*
Then, we can apply the so-called uniform Gronwall's lemma (see, e.g., \cite
Chapter III, Lemma 1.1]{T})\ to find a time $\tau _{1}\geq 1$, depending on
\tau _{0}$ and $\tau ,$ such that
\begin{equation}
\left\Vert \nabla \partial _{t}u\left( t\right) \right\Vert _{2}^{2}\leq
c,\qquad \text{for all }\,t\geq \tau _{1}. \label{4.29}
\end{equation
Therefore, (\ref{4.29}) and (\ref{rec}) allow us to deduce from (\ref{she})
and (\ref{dyn}), via standard elliptic regularity, the following estimate
\begin{equation}
\left\Vert u\left( t\right) \right\Vert _{H^{2}\left( \Omega \right)
}^{2}\leq c,\qquad \forall \,t\geq \tau _{1}. \label{4.30}
\end{equation
Summing up, we conclude by observing that (\ref{h2est}) follows from (\re
{4.30}) and the boundedness of the trace map $Tr_{D}:H^{2}\left( \Omega
\right) \rightarrow H^{3/2}\left( \Gamma \right) $. This completes the proof
of the theorem.
\end{proof}
\begin{remark}
The proof of Theorem \ref{attr} shows how to get an absorbing set in
\mathbb{V}_{2}$. Because of this, we can also prove the existence of a
global attractor for the dynamical system $\left( \left\{ \mathcal{S
_{t}\right\} _{t\geq 0},\mathbb{V}_{1}\right) .$
\end{remark}
\begin{theorem}
\label{reg2}If $\Omega $ is a bounded $\mathcal{C}^{\infty }$-domain, and
f,g$ are $\mathcal{C}^{\infty }$ functions, then the global attractor
\mathcal{A}_{W}$ is a bounded subset of $\mathbb{V}_{k},$ for every $k\geq 1
. In particular, if $U\in \mathcal{A}_{W}$ then $u\in \mathcal{C}^{\infty
}\left( \overline{\Omega }\right) .$
\end{theorem}
The proof of this result is standard and follows by successive time
differentiation of the equations in (\ref{diff}) and an induction argument.
We omit the details.
To prove the finite dimensionality of the global attractor $\mathcal{A}_{W}
, we can proceed in two different ways. One way is to establish the
existence of a more refined object called exponential attractor $\mathcal{E
_{W}$, whose existence proof is often based on proper forms of the so-called
squeezing/smoothing property for the differences of solutions. This can be
done by assuming smoother nonlinearities, i.e., $f\in C^{2}\left( \mathbb{R
\right) $ (see, e.g., \cite{GG0, GG1}). This has been carried out in \cit
{GG0}, and references therein, for a system of reaction-diffusion equations
with dynamic boundary conditions of the form (\ref{dyn}), without relating
the attractor dimension to the physical parameters of the problem. However,
since we wish to find explicit estimates of fractal or/and Hausdorff
dimension of $\mathcal{A}_{W}$, we shall employ the classical machinery for
proving the finite dimensionality of the global attractor $\mathcal{A}_{W}.$
This is based on the so-called volume contraction arguments and requires the
associated solution semigroup $\mathcal{S}_{t}$ to be (uniformly quasi-)
differentiable with respect to the initial data, at least on the attractor
(see, e.g., \cite{BV}).
We give without proof the following result, which follows as a consequence
of the boundedness of $\mathcal{A}_{W}$ into $\mathbb{V}_{2}\cap \mathbb{X
^{\infty }.$
\begin{proposition}
\label{propd}Provided that $f\in C^{2}\left( \mathbb{R}\right) $ satisfies
the conditions (\ref{n1}) and (\ref{n2}), the flow $\mathcal{S}_{t}$
generated by the reaction-diffusion equation (\ref{she}) and dynamic
boundary condition (\ref{dyn}) is uniformly differentiable on $\mathcal{A
_{W},$ with differentia
\begin{equation}
\mathbf{L}\left( t,U\left( t\right) \right) :\Theta =\binom{\xi _{1}}{\xi
_{2}}\in \mathbb{X}^{2}\mapsto V=\binom{v}{\varphi }\in \mathbb{X}^{2},
\label{4.9}
\end{equation
where $V$ is the unique solution t
\begin{align}
\partial _{t}v& =\nu \Delta v-f^{^{\prime }}\left( u\left( t\right) \right)
v+\lambda v,\text{ }\left( \partial _{t}\varphi +\nu b\partial _{\mathbf{n
}v\right) _{\mid \Gamma }=0, \label{var} \\
V\left( 0\right) & =\Theta . \notag
\end{align
Furthermore, $\mathbf{L}\left( t,U\left( t\right) \right) $ is compact for
all $t>0.$
\end{proposition}
The main result of this section is
\begin{theorem}
\label{main1}Let the assumptions of Proposition \ref{propd} be satisfied.
The fractal dimension of $\mathcal{A}_{W}$ admits the estimat
\begin{equation}
\dim _{F}\mathcal{A}_{W}\leq c_{0}\left( 1+\frac{c_{f}+\lambda }{C_{W}\left(
\Omega ,\Gamma \right) \nu }\right) ^{n-1},\text{ for }n\geq 2 \label{updim}
\end{equation
an
\begin{equation}
\dim _{F}\mathcal{A}_{W}\leq c_{0}\left( 1+\frac{c_{f}+\lambda }{C_{D}\left(
\Omega \right) \nu }\right) ^{1/2},\text{ for }n=1, \label{updim2}
\end{equation
where $c_{0}$ depends on the shape of $\Omega $ only. The positive constants
$C_{W},C_{D}$ depend only on $n,$ $\Omega ,$ $\Gamma $, $b$ and are given in
the Appendix.
\end{theorem}
\begin{proof}
In order to deduce (\ref{updim})-(\ref{updim2}), it is sufficient (see,
e.g., \cite[Chapter III, Definition 4.1]{CV})\ to estimate the $j$-trace of
the operato
\begin{equation*}
\mathbf{L}\left( t,U\left( t\right) \right) =\left(
\begin{array}{cc}
\nu \Delta -f^{^{\prime }}\left( u\left( t\right) \right) +\lambda I & 0 \\
-b\nu \partial _{\mathbf{n}} &
\end{array
\right) .
\end{equation*
We hav
\begin{align*}
Trace\left( \mathbf{L}\left( t,U\left( t\right) \right) Q_{m}\right) &
=\sum_{j=1}^{m}\left\langle \mathbf{L}\left( t,U\left( t\right) \right)
\varphi _{j},\varphi _{j}\right\rangle _{\mathbb{X}^{2}} \\
& =\sum_{i=1}^{m}\left\langle \nu \Delta \varphi _{j},\varphi
_{j}\right\rangle _{2}-\sum_{i=1}^{m}\left\langle \nu \partial _{\mathbf{n
}\varphi _{j},\varphi _{j}\right\rangle _{2,\Gamma } \\
& -\sum_{i=1}^{m}\left\langle f^{^{\prime }}\left( u\left( t\right) \right)
\varphi _{j},\varphi _{j}\right\rangle _{2}+\sum_{i=1}^{m}\lambda
\left\langle \varphi _{j},\varphi _{j}\right\rangle _{2},
\end{align*
where the set of vector-valued functions $\varphi _{j}\in \mathbb{X}^{2}\cap
\mathbb{V}_{1}$ is an orthonormal basis in $Q_{m}\mathbb{X}^{2}$. Since the
family $\varphi _{j}$ is orthonormal in $Q_{m}\mathbb{X}^{2},$ using
assumption (\ref{n1}) on $f$ (i.e., $f^{^{\prime }}\left( y\right) \geq
-c_{f},$ for all $y\in \mathbb{R}$), we fin
\begin{equation*}
Trace\left( \mathbf{L}\left( t,U\right) Q_{m}\right) \leq -\nu
\sum_{i=1}^{m}\left\Vert \nabla \varphi _{j}\right\Vert _{2}^{2}+\left(
c_{f}+\lambda \right) m.
\end{equation*
Let $n\geq 2$. From (\ref{LTineq}) (see Appendix, Proposition \ref{LT}), we
obtai
\begin{align*}
Trace\left( \mathbf{L}\left( t,U\right) Q_{m}\right) & \leq -\nu
c_{1}C_{W}\left( \Omega ,\Gamma \right) m^{\frac{1}{n-1}+1}+\left( c_{1}\nu
C_{W}\left( \Omega ,\Gamma \right) +c_{f}+\lambda \right) m \\
& =:\rho \left( m\right) .
\end{align*
The function $\rho \left( y\right) $ is concave. The root of the equation
\rho \left( d\right) =0$ i
\begin{equation*}
d^{\ast }=\left( 1+\frac{c_{f}+\lambda }{\nu c_{1}C_{W}\left( \Omega ,\Gamma
\right) }\right) ^{n-1}.
\end{equation*
Thus, we can apply \cite[Corollary 4.2 and Remark 4.1]{CV} to deduce that
\dim _{F}\mathcal{A}_{W}\leq d^{\ast },$ from which (\ref{updim}) follows.
The case $n=1$ is similar.
\end{proof}
\begin{remark}
Concerning the reaction-diffusion equation (\ref{she}), we can also handle
dynamic boundary conditions that involve surface diffusion:
\begin{equation}
\partial _{t}u-\alpha \Delta _{\Gamma }u+b\nu \partial _{{\mathbf{n}}}\phi
=0,\text{ on }\Gamma , \label{3.32}
\end{equation
where $\alpha >0$ and $\Delta _{\Gamma }$ is the Laplace-Beltrami operator
on $\Gamma $. Our method of establishing upper bounds, comparable to the
bounds (\ref{updim})-(\ref{updim2}), for the dimension of the global
attractor can be also extended to this case as well. The details will appear
elsewhere.
\end{remark}
\section{Lower bounds on the dimension}
\label{lbb}
Lower bounds on the dimension of the global attractor are usually based on
the observation that the unstable manifold of any equilibrium of the system
is always contained in the global attractor (see, e.g., \cite{BV}). Thus, a
lower bound on the dimension of the attractor $\mathcal{A}_{W}$ can be found
by analyzing the dimension of an unstable manifold associated with a
constant equilibrium $Z$ for (\ref{she}), (\ref{dyn}). We begin by assuming
that $g$ is constant, for the sake of simplicity. Steady-state solutions of
\ref{she}), (\ref{dyn}) satisf
\begin{equation*}
L_{0}\left( u\right) :=\nu \Delta u-f\left( u\right) +\lambda u-g=0,\text{
\left( \partial _{\mathbf{n}}u\right) _{\mid \Gamma }=0.
\end{equation*
We seek a solution of this system $U=\binom{u}{Tr_{D}\left( u\right) }\in
\mathbb{X}^{2}$\ which coincides with a constant vector $Z=\mathbf{c}=\binom
c}{c},$ $c$ is a constant. Such a stationary solution satisfies the equation
$\overline{L}_{0}\left( z\right) :=-f\left( z\right) +\lambda z-g=0.$ Sinc
\begin{equation*}
f\left( y\right) y\geq \eta _{1}\left\vert y\right\vert ^{p}-C_{f},\text{
for }p>2,
\end{equation*
we have $\overline{L}_{0}\left( z\right) z\leq -\widetilde{\eta
_{1}\left\vert z\right\vert ^{p}+\widetilde{C}_{f},$ for some positive
constants $\widetilde{\eta }_{1},\widetilde{C}_{f}$. Thus, $\overline{L
_{0}\left( z\right) z<0$ on the interval $I_{R}=\left( -R,R\right) ,$ if $R$
is large enough. It follows that $\overline{L}_{0}\left( z\right) =0$ has at
least one solution $Z=Z\left( \lambda \right) $ (see, e.g., \cite[Chapter II
]{CV}). By the implicit function theorem, this solution is of order
\thinspace $1/\lambda $ for sufficiently large $\lambda .$
Now fix this solution. In order to find a lower bound on the dimension of
the global attractor $\mathcal{A}_{W},$ it suffices to establish a lower
bound for $\dim E_{+}\left( Z\right) ,$ where $E_{+}\left( Z\right) $ is an
invariant subspace of $\mathbf{L}\left( Z\right) ,$ which corresponds t
\begin{equation*}
\mathbf{L}\left( Z\right) W=\binom{\nu \Delta w-f^{^{\prime }}\left(
z\right) w+\lambda w}{-b\nu \partial _{\mathbf{n}}w}
\end{equation*
with $\sigma \left( \mathbf{L}\left( Z\right) \right) \subset \left\{ \zeta
:\zeta >0\right\} $. We note that $\left( \mathbf{L}\left( Z\right) ,D\left(
\mathbf{L}\left( Z\right) \right) \right) $\ is self-adjoint on $X^{2}$\
with spectrum contained in $\left( -\infty ,c_{f}+\lambda \right] .$
The main result of this section is the following.
\begin{theorem}
\label{lbbb}Let $f\in C^{2}\left( \mathbb{R}\right) $ satisfy assumptions
\ref{n1})-(\ref{n2}). There exist a positive constant $c_{0}$, depending on
f,$ $g$ and the shape of $\Omega ,$ independent of $\lambda ,$ $\nu ,$ $b$,
\left\vert \Omega \right\vert ,$ $\left\vert \Gamma \right\vert ,$ such tha
\begin{equation*}
\dim _{F}\mathcal{A}_{W}\geq \dim _{H}\mathcal{A}_{W}\geq \dim E_{+}\left(
Z\right) \geq c_{0}\left( \frac{\lambda }{C_{W}\left( \Omega ,\Gamma \right)
\nu }\right) ^{n-1},
\end{equation*
for $n\geq 2$. In one space dimension, the same estimate is valid with
C_{W} $ replaced by $C_{D}$ and $n-1,$ replaced by $1/2$, respectively.
\end{theorem}
\begin{proof}
For a fixed constant solution $Z=\mathbf{c}$ of $\overline{L}_{0}\left(
z\right) =0$ and sufficiently large $\lambda \geq 1,$ we have $\chi \left(
\lambda \right) :=-f^{^{\prime }}\left( z\right) +\lambda >0$.
Let $\left\{ \varphi _{j}\left( x\right) \right\} _{ji\in \mathbb{N}_{0}}$
be an orthonormal basis in $\mathbb{X}^{2}$ consisting of eigenfunctions of
the Wentzell Laplacian $\Delta _{W}$ (see Appendix, Theorem \ref{sanalysis2
)
\begin{equation}
\Delta _{W}\varphi _{j}=\Lambda _{j}\varphi _{j},\text{ }j\in \mathbb{N}_{0}
\text{ }\varphi _{j}\in D\left( \Delta _{W}\right) \cap C\left( \overline
\Omega }\right) \label{evseq}
\end{equation
such tha
\begin{equation*}
0=\Lambda _{0}<\Lambda _{1}\leq \Lambda _{2}\leq ...\leq \Lambda _{,j}\leq
\Lambda _{j+1}\leq ....
\end{equation*
We shall seek for eigenvectors $W_{j}=\binom{w_{j}}{Tr_{D}\left(
w_{j}\right) }\in \mathbb{X}^{2}$, of the form $w_{j}\left( x\right)
=\varphi _{j}\left( x\right) p_{j},$ $p_{j}\in \mathbb{R}$, satisfying
equatio
\begin{equation}
\mathbf{L}\left( Z\right) W_{j}=\zeta _{j}W_{j},\text{ }W_{j}\in D\left(
\mathbf{L}\left( Z\right) \right) :=D\left( \Delta _{W}\right) . \label{eee}
\end{equation
Note that for $W_{j}\in D\left( \mathbf{L}\left( Z\right) \right) \subset
\mathbb{V}_{1},$ the trace of $w_{j}$ makes sense as an element of
H^{1/2}\left( \Gamma \right) $. Substituting such $w_{j}$ into (\ref{eee}),
taking into account (\ref{evseq}) and the fact tha
\begin{equation*}
\mathbf{L}\left( Z\right) W_{j}=-\nu \Delta _{W}W_{j}+\Pi _{\lambda }W_{j}
\text{ }\Pi _{\lambda }W_{j}:=\binom{\chi \left( \lambda \right) w_{j}}{0},
\end{equation*
we obtain the equatio
\begin{equation}
\left( -\nu \Lambda _{j}I+\Pi _{\lambda }\right) p_{j}=\zeta _{j}p_{j},\text{
}\Pi _{\lambda }=\left(
\begin{array}{cc}
\chi \left( \lambda \right) & 0 \\
0 &
\end{array
\right) . \label{pj}
\end{equation
A nonzero $p_{j}$ exists if $\zeta =\zeta _{j}$ is a root of the equatio
\begin{equation}
\det \left( -\nu \Lambda _{j}I+\Pi _{\lambda }-\zeta I\right) =0,\text{
\zeta >0. \label{dett}
\end{equation
When $\nu =0,$ this equation has at least one root $\zeta >0$ since $\chi
=\chi \left( \lambda \right) >0$ (in fact, $\zeta =\chi \left( \lambda
\right) $). Therefore, there exists $\delta >0$ such that when $\nu \Lambda
_{j}<\delta ,$ the equation (\ref{dett}) has a root $\zeta _{j}=\zeta
_{j}\left( \nu \right) $ with $\zeta _{j}>0$. Therefore, to any such root
\zeta _{j}$, we can assign a nontrivial $p_{j},$ which is a solution of (\re
{pj}), and thus an eigenvector $W_{j}=\binom{w_{j}}{Tr_{D}w_{j}},$
w_{j}=\varphi _{j}p_{j}$. Let us now compute how many $j$'s satisfy the
inequality $\nu \Lambda _{j}<\delta $. The asymptotic behavior of $\Lambda
_{j}$ is $\Lambda _{j}\sim C_{W}\left( \Omega ,\Gamma \right) j^{1/\left(
n-1\right) }$ as $j\rightarrow \infty $ (see, Appendix, Theorem \re
{asymptotic2}). The last inequality certainly holds whe
\begin{equation*}
1\leq j\leq c_{1}\delta ^{n-1}\left( C_{W}\nu \right) ^{1-n}=c_{2}\left(
\frac{1}{C_{W}\nu }\right) ^{n-1},\text{ for }n\geq 2
\end{equation*
an
\begin{equation*}
1\leq j\leq c_{1}\delta ^{1/2}\left( C_{D}\nu \right) ^{-1/2}=c_{2}\left(
\frac{1}{C_{D}\nu }\right) ^{1/2},\text{ for }n=1.
\end{equation*
The positive constants $c_{1},$ $c_{2}$ depend on $\lambda .$ In order to
get more explicit estimates for $c_{1},$ $c_{2}$, it is left to remark that
equation (\ref{dett}) may be rewritten in the for
\begin{equation*}
\det \left( -\nu \Lambda _{j}\lambda ^{-1}I+\lambda ^{-1}\Pi _{\lambda
}-\zeta _{1}I\right) =0
\end{equation*
with $\zeta _{1}=\lambda ^{-1}\zeta ,$ and to observe that a solution of
this equation clearly exists if $\nu \Lambda _{j}\lambda ^{-1}\leq \delta ,$
for sufficiently large $\lambda $ and small $\delta $. Employing the
asymptotic formula for $\Lambda _{j}$ once again, we find
\begin{equation*}
1\leq j\leq c_{1}^{^{\prime }}\delta ^{n-1}\lambda ^{n-1}\left( C_{W}\nu
\right) ^{1-n}=c_{2}^{^{\prime }}\left( \frac{\lambda }{C_{W}\nu }\right)
^{n-1},\text{ for }n\geq 2
\end{equation*
an
\begin{equation*}
1\leq j\leq c_{1}^{^{\prime }}\delta ^{1/2}\lambda ^{1/2}\left( C_{D}\nu
\right) ^{-1/2}=c_{2}^{^{\prime }}\left( \frac{\lambda }{C_{D}\nu }\right)
^{1/2},\text{ for }n=1.
\end{equation*
It follows that
\begin{equation*}
\dim E_{+}\left( Z\left( \lambda \right) \right) \geq c_{2}^{^{\prime
}}\left( \frac{\lambda }{C_{W}\nu }\right) ^{n-1},\text{ for }n\geq 2
\end{equation*
and
\begin{equation*}
\dim E_{+}\left( Z\left( \lambda \right) \right) \geq c_{2}^{^{\prime
}}\lambda ^{1/2}\left( C_{D}\nu \right) ^{-1/2},
\end{equation*
in one space dimension. The proof is complete.
\end{proof}
\section{Concluding remarks}
\label{cr}
In the textbook literature on theoretical geophysics, it was traditional to
use a Robin boundary condition with a nonlinear heat equation to describe
temperature variations at the upper surface of the ocean \cite{LMT, LMT2}.
But it was recognized that this was not always the physically correct
boundary condition \cite{MW2}. Among its applicability to a wide range of
phenomena, including phase-transitions in fluids, and so on \cite{GG0, QWS},
the reaction-diffusion system (\ref{she})-(\ref{dyn}) has important
applications in climatology and is essentially used to determine large and
rapid temperature changes in the ocean's surface as a response to changes
into deep water formations \cite{MW2}. In this paper, we provide explicit
bounds for the dimension of the attractor for this system and study the
effect of the dynamic term $b^{-1}\partial _{t}u$, representing change in
thermal energy\ in an infinitesimal layer near the surface. Unlike the
previous results, the dimension of the attractor is proportional to the
surface area $\left\vert \Gamma \right\vert ,$ for large domains $\Omega $
and fixed parameters $\nu ,$ $\lambda $ and $b$. Moreover, all the constants
involved in our estimates are given in an explicit form. We also observe
that in the case without $b^{-1}\partial _{t}u$ in (\ref{dyn}), \thinspace
i.e., $b=+\infty ,$ the dimension of the attractor is much larger (and
proportional to the volume $\left\vert \Omega \right\vert $ of $\Omega $)
than the dimension of the global attractor for the same system when $0<b\neq
+\infty $. Thus, we observe that the addition of the dynamic term
b^{-1}\partial _{t}u$, $b>0$ drastically changes the situation. This is a
remarkable fact that can have a profound effect onto the long-term dynamics
of other systems that are subject to dynamic boundary conditions of this
form. We will investigate these effects for other systems, such as the B\'{e
nard problem for nonlinear heat conduction, in forthcoming papers. Finally,
we note that it is also possible to extend the results of this paper to the
case when the boundary $\Gamma $\ consists of two disjoint open subsets
\Gamma _{1}$\ and $\Gamma _{2}$, each $\overline{\Gamma }_{i}\backprime
\Gamma _{i}$\ is a $S$-null subset of $\Gamma $\ and $\Gamma =\overline
\Gamma }_{1}\cup \overline{\Gamma }_{2}$\ with $\Gamma _{1}\subsetneqq
\Gamma $, such that $u$\ satisfies a Dirichlet boundary condition on $\Gamma
_{1}$\ and a dynamic boundary condition on $\Gamma _{2}$. We will come back
to this issue in a forthcoming article.
\section{Appendix}
\label{ap}
In this section, we shall recall several important results concerning a
certain realization of $L=\nu \Delta $ with the Wentzell boundary condition
\ref{wbc}). We have the following.
\begin{theorem}
\label{Wentzell2}Let $\Omega $ be a bounded open set of $\mathbb{R}^{n}$
with Lipschitz boundary $\Gamma $. Assume that $b>0$ and $0\leq q\in
L^{\infty }\left( \Omega \right) $. Define the operator $\Delta _{W}$ on
\mathbb{X}^{2},$ b
\begin{equation}
\Delta _{W}\binom{u_{1}}{u_{2}}:=\binom{-\Delta u_{1}+q\left( x\right) u_{1
}{b\partial _{\mathbf{n}}u_{1}}, \label{A_Wentzell1}
\end{equation
wit
\begin{equation}
D\left( \Delta _{W}\right) :=\left\{ U=\binom{u_{1}}{u_{2}}\in \mathbb{V
_{1}:-\Delta u_{1}\in L^{2}\left( \Omega \right) ,\text{ }\partial _{\mathbf
n}}u_{1}\in L^{2}\left( \Gamma ,\frac{dS}{b}\right) \right\} .
\label{A_Wentzell2}
\end{equation
Then, $\left( \Delta _{W},D\left( \Delta _{W}\right) \right) $ is
self-adjoint on $\mathbb{X}^{2}.$ Moreover, the resolvent operator $\left(
I+\Delta _{W}\right) ^{-1}\in \mathcal{L}\left( \mathbb{X}^{2}\right) $ is
compact.
\end{theorem}
We refer the reader to \cite{CFGGOR, GG_b, GGGRW} for an extensive survey of
recent results concerning the "Wentzell" Laplacian $\Delta _{W}$.
The eigenvalue problem associated with the operator $\Delta _{W}$ is given
by $\Delta _{W}\varphi =\Lambda \varphi ;$ this leads to the following
spectral problem for the perturbed Laplacia
\begin{equation}
-\Delta \varphi +q\left( x\right) \varphi =\Lambda \varphi \text{ in }\Omega
, \label{sp1}
\end{equation
with a boundary condition that depends on the eigenvalue $\Lambda $
explicitly
\begin{equation}
b\partial _{\mathbf{n}}\varphi =\Lambda \varphi \text{ on }\Gamma .
\label{sp2}
\end{equation
Such a function $\varphi $ will be called an eigenfunction associated with
\Lambda $ and the set of all eigenvalues $\Lambda $ of (\ref{sp1})-(\ref{sp2
) will be denoted by $\Lambda _{W}.$ Let $\varphi _{j}$ and $\Lambda _{W,j}
, $j\in J$, denote all the eigenfunctions and eigenvalues of (\ref{sp1})-
\ref{sp2}). We have the following (see, e.g., \cite{BBR, VVi}).
\begin{theorem}
\label{sanalysis1}Let $q\geq 0$ with $\int\limits_{\Omega }q\left( x\right)
dx>0$. Then, there exists a sequence of number
\begin{equation}
0<\Lambda _{W,1}\leq \Lambda _{W,2}\leq ...\leq \Lambda _{W,j}\leq \Lambda
_{W,j+1}\leq ..., \label{seq}
\end{equation
converging to $+\infty $, with the following properties:
(a) The spectrum of $\Delta _{W}$ is given b
\begin{equation*}
\sigma \left( \Delta _{W}\right) =\left\{ \Lambda _{W,j}\right\} _{j\in
\mathbb{N}},
\end{equation*
and each number $\Lambda _{W,j},$ $j\in \mathbb{N},$ is an eigenvalue for
\Delta _{W}$ of finite multiplicity.
(b) There exists a countable family of orthonormal eigenfunctions for
\Delta _{W}$ which spans $\mathbb{X}^{2}$. More precisely, there exists a
collection of functions $\left\{ \varphi _{j}\right\} _{j\in \mathbb{N}}$
with the following properties
\begin{align}
\varphi _{j}& \in D\left( \Delta _{W}\right) \text{ and }\Delta _{W}\varphi
_{j}=\Lambda _{W,j}\varphi _{j},\text{ }j\in \mathbb{N}, \label{prop} \\
\left\langle \varphi _{j},\varphi _{k}\right\rangle _{\mathbb{X}^{2}}&
=\delta _{jk}\text{, }j,k\in \mathbb{N}\text{,} \notag \\
\mathbb{X}^{2}& =\oplus \overline{lin.span\left\{ \varphi _{j}\right\}
_{j\in \mathbb{N}}}\text{ (orthogonal direct sum).} \notag
\end{align}
(c) If $\Gamma $ is Lipschitz, then every eigenfunction $\varphi _{j}\in
\mathbb{V}_{1}$, and in fact $\varphi _{j}\in C(\overline{\Omega })\cap
C^{\infty }(\Omega )$, for every $j$. If $\Gamma $ is of class $C^{2}$, then
every eigenfunction $\varphi _{j}\in \mathbb{V}_{1}\cap C^{2}\left(
\overline{\Omega }\right) ,$ for every $j.$
(d) The following min-max principle holds
\begin{equation}
\Lambda _{W,j}=\min_{\substack{ Y_{j}\subset \mathbb{V}_{1}, \\ \dim
Y_{j}=j }}\max_{0\neq \varphi \in Y_{j}}R_{W}\left( \varphi ,\varphi \right)
,\text{ }j\in \mathbb{N}\text{,} \label{minmax}
\end{equation
where the Rayleigh quotient $R_{W}$, for the perturbed Wentzell operators,
is given b
\begin{equation}
R_{W}\left( \varphi ,\varphi \right) :=\frac{\left\Vert \nabla \varphi
\right\Vert _{2}^{2}+\left\langle q\left( x\right) \varphi ,\varphi
\right\rangle _{2}}{\left\Vert \varphi \right\Vert _{\mathbb{X}^{2}}^{2}}
\text{ }0\neq \varphi \in \mathbb{V}_{1}. \label{rqw}
\end{equation}
\end{theorem}
Concerning the case $q\equiv 0$, we have the following.
\begin{theorem}
\label{sanalysis2}Let $q\equiv 0.$ Then, there exists a sequence of number
\begin{equation*}
0=\Lambda _{W,0}<\Lambda _{W,1}\leq \Lambda _{W,2}\leq ...\leq \Lambda
_{W,j}\leq \Lambda _{W,j+1}\leq ...,
\end{equation*
converging to $+\infty $, with the following properties:
(a) The spectrum of $\Delta _{W}$ is given b
\begin{equation*}
\sigma \left( \Delta _{W}\right) =\left\{ \Lambda _{W,j}\right\} _{j\in
\mathbb{N\cup }\left\{ 0\right\} },
\end{equation*
and each number $\Lambda _{W,j},$ $j\in \mathbb{N}_{0}=\mathbb{N\cup
\left\{ 0\right\} ,$ is an eigenvalue for $\Delta _{W}$ of finite
multiplicity. The eigenvalue $\Lambda _{W,0}$ is simple and its associated
eigenfunction is of constant sign.
(b) There exists a countable family of orthonormal eigenfunctions for
\Delta _{W}$ which spans $\mathbb{X}^{2}$. More precisely, the same
conclusion (b) of Theorem \ref{sanalysis1} holds in this case as well.
Finally, both conclusions (c) and (d) in Theorem \ref{sanalysis1} hold in
the case $q\equiv 0$ as well.
\end{theorem}
The asymptotic behavior of the eigenvalues $\Lambda _{W,j},$ as
j\rightarrow \infty ,$ was established in \cite{Fran, Fran2}. We refer the
reader to \cite{GG_b} for more details about the Wentzell Laplacian and
other generalizations. Let $J=\mathbb{N}_{0}$ or $\mathbb{N}$, according to
whether $q=0$ or $q>0$ respectively. Se
\begin{equation*}
C_{D}\left( \Omega \right) :=\frac{\left( 2\pi \right) ^{2}}{\left(
v_{n}\left\vert \Omega \right\vert \right) ^{2/n}}\text{ and }C_{S}\left(
\Gamma \right) =\frac{2\pi }{\left( v_{n-1}\left\vert \Gamma \right\vert
\right) ^{1/\left( n-1\right) }}.
\end{equation*
Here $v_{n}$ denotes the volume of the unit ball in $\mathbb{R}^{n}$, and we
recall that $\left\vert \Omega \right\vert $ stands for the $n$-dimensional
Euclidean volume of $\Omega $, while $\left\vert \Gamma \right\vert $ stands
for the usual $\left( n-1\right) $-dimensional Lebesgue surface measure on
\Gamma $.
We summarize these results in the following.
\begin{theorem}
\label{asymptotic2}The eigenvalue sequence $\left\{ \Lambda _{W,j}\right\}
_{j\in J}$ of the (un)perturbed Wentzell Laplacian $\Delta _{W}$ satisfies:
(i) For $n\geq 2$, we hav
\begin{equation}
\Lambda _{W,j}=C_{W}\left( \Omega ,\Gamma \right) j^{1/\left( n-1\right)
}+o\left( j^{1/\left( n-1\right) }\right) ,\text{ as }j\rightarrow +\infty ,
\label{ev}
\end{equation
for som
\begin{equation}
C_{W}\left( \Omega ,\Gamma \right) \in \left\{
\begin{array}{cc}
bC_{S}\left( \Gamma \right) \left[ 2^{-1/\left( n-1\right) },1\right] , &
\text{for }n\geq 3, \\
\left[ \frac{C_{D}\left( \Omega \right) C_{S}\left( \Gamma \right) }{2\left(
b^{-1}C_{D}\left( \Omega \right) +C_{S}\left( \Gamma \right) \right) },\min
\left\{ C_{D}\left( \Omega \right) ,bC_{S}\left( \Gamma \right) \right\}
\right] , & \text{for }n=2
\end{array
\right. \label{constant}
\end{equation}
(ii) For $n=1,$ we hav
\begin{equation}
\Lambda _{W,j}=C_{D}\left( \Omega \right) j^{2}+o\left( j^{2}\right) ,\text{
as }j\rightarrow +\infty . \label{ev2}
\end{equation}
\end{theorem}
The following version of the Lieb--Thirring inequality is essential.
\begin{proposition}
\label{LT}Let $\omega _{j},$ $1\leq j\leq m,$ be a finite family of $\mathbb
V}_{1},$ which is orthonormal in $\mathbb{X}^{2}$. We hav
\begin{equation}
\sum_{i=1}^{m}\left\Vert \nabla \omega _{j}\right\Vert _{2}^{2}\geq
c_{1}C_{W}\left( \Omega ,\Gamma \right) \left( m^{\frac{1}{n-1}+1}-m\right) .
\label{LTineq}
\end{equation
The constant $c_{1}>0$ depends only on $n$ and the shape of $\Omega ,$ and
is independent of the size of $\Omega ,$ $\Gamma ,$ of $m,$ and of the
\omega _{j}$'s.
\end{proposition}
\begin{proof}
Let $B_{W}:=\Delta _{W}+C_{W}\left( \Omega ,\Gamma \right) I$ and let
D\left( B_{W}\right) =D\left( \Delta _{W}\right) $. By Theorems \re
{Wentzell2}, \ref{sanalysis1}, $B_{W}$ is a linear positive unbounded
self-adjoint operator on $\mathbb{X}^{2},$ such that $B_{W}^{-1}$ is
compact. Thus, we can apply the abstract result of \cite[Chapter VI, Lemma
2.1]{T} to deduce tha
\begin{align}
\sum_{i=1}^{m}\left( \left\Vert \nabla \omega _{j}\right\Vert
_{2}^{2}+C_{W}\left\Vert \omega _{j}\right\Vert _{\mathbb{X}^{2}}^{2}\right)
& =\sum_{i=1}^{m}\left\langle B_{W}\omega _{j},\omega _{j}\right\rangle _
\mathbb{X}^{2}} \label{LTpp} \\
& \geq \Lambda _{W,1}\left( B_{W}\right) +\Lambda _{W,2}\left( B_{W}\right)
+...+\Lambda _{W,m}\left( B_{W}\right) \notag \\
& \geq C_{W}\left( 1^{1/\left( n-1\right) }+2^{1/\left( n-1\right)
}+...+m^{1/\left( n-1\right) }\right) \notag \\
& \geq c_{0}C_{W}m^{\frac{1}{n-1}+1}, \notag
\end{align
since, by (\ref{ev})-(\ref{ev2}), $\Lambda _{W,j}\left( B_{W}\right) \geq
C_{W}\left( \Omega ,\Gamma \right) j^{1/\left( n-1\right) },$ for all $j,$
and some positive constant $c_{0}$ (indeed, we have $\Lambda _{W,j}\left(
B_{W}\right) =\Lambda _{W,j}\left( \Delta _{W}\right) +C_{W}$)$.$ Thus, the
proof of (\ref{LTineq}) follows immediately from (\ref{LTpp}).
\end{proof}
|
\section{Introduction}
Serre \cite{Se} first developed the theory of $p$-adic Eisenstein series and there have subsequently been many results in the field of $p$-adic modular forms.
Several researchers have attempted to generalize the theory to modular forms with several variables. For example, we showed that a $p$-adic
limit of a Siegel Eisenstein series becomes a ``real" Siegel
modular form (cf. \cite{KN}). The same result has also been proved for Hermitian
modular forms (e.g., \cite{N}).
In the present paper, we study $p$-adic limits of quaternionic Eisenstein series. This study has two principal aims.
The first is to show that these $p$-adic limits become ``real" modular forms of level $p$ for higher
$p$-adical weights (Theorem \ref{ThmM}).
To prove this, we introduce a $U(p)$ type Hecke operator and study its properties; this is
a similar method to that used by B\"ocherer for Siegel modular forms \cite{Bo}. The second aim is
to show that a strange phenomenon occurs for low $p$-adical weights; namely, there exists a transcendental $p$-adic Eisenstein series in the quaternionic case (Theorem \ref{ThmM2}).
\section{Preliminaries}
\subsection{Notation and definitions}
Let $\mathbb{H}$ be Hamiltonian quaternions and $\mathcal{O}$ the Hurwitz order (cf. \cite{Kri2}). The half-space of quaternions of degree $n$ is defined as
$$
H(n;\mathbb{H}):=\{\; Z=X+iY\;|\; X,\,Y\in Her_n(\mathbb{H}),\;Y>0\;\}.
$$
Let $J_n:=\begin{pmatrix}O_n & 1_n \\ -1_n & O_n\end{pmatrix}$. Then, the group of symplectic similitudes
$$
\left\{\;M\in M(2n,\mathbb{H})\;|\; {}^t\overline{M}J_nM=qJ_n\ {\rm for\ some\ positive\ }q\in \mathbb{R}\;\right\}
$$
acts on $H(n;\mathbb{H})$ by
$$
Z\longmapsto M\langle Z\rangle =(AZ+B)(CZ+D)^{-1},\quad M=\begin{pmatrix}A & B \\ C & D \end{pmatrix}.
$$
Let $\varGamma_n$ denote the modular group of quaternions of degree $n$ defined by
\begin{align*}
&G_n:=\left\{\;M\in M(2n,\mathbb{H})\;|\; {}^t\overline{M}J_nM=J_n\;\right\}, \\
&\varGamma_n:=\varGamma_n(\mathcal{O})=M(2n,\mathcal{O})\cap G_n.
\end{align*}
For a given $q\in \mathbb{N}$, the congruence subgroup
$\varGamma _0 ^{(n)}(q)$ of $\varGamma_n$ is defined by
$$
\varGamma _0^{(n)}(q):=\left\{ \begin{pmatrix} A & B \\ C & D \end{pmatrix}\in
\varGamma_n \;|\; C\equiv O_n \bmod{qM(n,\mathcal{O})} \right\}.
$$
In this subsection, $\varGamma$ always denotes either $\varGamma _n$ or
$\varGamma _0^{(n)}(q)$.
Let $1=e_1$, $e_2$, $e_3$, $e_4$ denote the canonical basis of $\mathbb{H}$,
which is characterized by the identities
$$
e_4=e_2e_3=-e_3e_2,\quad e_2^2=e_3^2=-1.
$$
We consider the canonical isomorphism
\[
M(n,\mathbb{H}) \longrightarrow M(2n,\mathbb{C})
\]
given by $\overset{\vee}{A}=(\overset{\vee}{a}_{ij})$, where
$\overset{\vee}{a}=\begin{pmatrix}a_1+a_2i&a_3+a_4i\\-a_3+a_4i&a_1-a_2i\end{pmatrix}$,
if $a=a_1e_1+a_2e_2+a_3e_3+a_4e_4$
(cf. \cite{Kri2}).
We use the above isomorphism to define $\det(A)$ for $A\in M(n,\mathbb{H})$.
For a similitude $M=\begin{pmatrix} A & B \\ C & D \end{pmatrix}$ and
a function $f:\,H(n;\mathbb{H})\longrightarrow \mathbb{C}$,
we define the slash operator $\mid_k$ by
$$
(f|_kM)(Z)=\det(M)^{\frac{k}{2}}\det(CZ+D)^{-k}f((AZ+B)(CZ+D)^{-1}).
$$
A holomorphic function $f:\;H(n;\mathbb{H})\longrightarrow \mathbb{C}$ is called a
{\it quaternionic modular form of degree n and weight k for} $\varGamma $ if $f$ satisfies
$$
(f|_kM)(Z)=f(Z),
$$
for all $M\in \varGamma$. (The cusp condition is required if $n=1$.)
We denote by $M_k(\varGamma)$ the $\mathbb{C}$-vector space of all
quaternionic modular forms of degree $n$ and weight $k$ for $\varGamma$.
A modular form $f\in M_k(\varGamma)$ possesses a
Fourier expansion of the form
$$
f(Z)=\sum_{0\leq H\in Her_n^\tau (\mathcal{O})}a_f(H)e^{2\pi i\tau(H,Z)},\quad
Z\in H(n;\mathbb{H}),
$$
where $Her_n^\tau (\mathcal{O})$ denotes the dual lattice of $Her_n(\mathcal{O}):=
\{S\in M(n,\mathcal{O})\,|\,{}^t\overline{S}=S\}$ with respect to
the reduced trace form $\tau$ (cf. \cite{Kri2}). For simplicity,
we put $q^H:=e^{2\pi i\tau(H,Z)}$
for $H\in Her_n^\tau (\mathcal{O})$. Using this notation,
we write the above Fourier expansion simply as $f=\sum _Ha_f(H)q^H$.
For an even integer $k$, we consider the Eisenstein series
\begin{equation}
\label{Eisen}
E_k^{(n)}(Z):=\sum_{\binom{AB}{CD}\in\varGamma_{n0}\backslash\varGamma_n}
\det(CZ+D)^{-k},\quad Z\in H(n;\mathbb{H}),
\end{equation}
where $\varGamma_{n0}:=\left\{\begin{pmatrix} A & B \\ O_n & D \end{pmatrix}\in
\varGamma_n\right\}$. It is well known that this series belongs to $M_k(\varGamma_n)$ if $k>4n-2$.
We call this series the {\it quaternionic Eisenstein series of degree n and weight k}.
\subsection{Fourier coefficients of Eisenstein series}
In this section, we introduce an explicit formula for the Fourier coefficients
of the degree 2 quaternionic Eisenstein series obtained by
Krieg (cf. \cite{Kri3}).
Let $k>6$ be an even integer and let
$$
E_k^{(2)}(Z)=\sum_{0\leq H\in Her_2^\tau (\mathcal{O})}a_k(H)e^{2\pi i\tau (H,Z)}
$$
be the Fourier expansion of the degree $2$ quaternionic Eisenstein series $E_k^{(2)}$.
According to \cite{Kri3}, we introduce an explicit formula for $a_k(H)$. Given
$O_2\ne H\in Her_2^\tau (\mathcal{O})$, the ``greatest common divisor"
of $H$ is given by
$$
\varepsilon (H):=\text{max}\{d\in\mathbb{N}\;|\; d^{-1}H\in Her_2^\tau (\mathcal{O})\;\}.
$$
\begin{Thm}[Krieg \cite{Kri3}]
\label{Four}
\it{Let $k>6$ be even and $H\ne O_2$. Then, the Fourier
coefficient $a_k(H)$ is given by:
$$
a_k(H)=\sum_{0<d|\varepsilon (H)} d^{k-1}\alpha^*(2{\rm det}(H)/d^2)
$$
and
$$
\alpha^*(\ell)=
\begin{cases}\displaystyle
-\frac{2k}{B_k} & \text{if $\ell =0$},\\
\displaystyle
-\frac{4k(k-2)}{(2^{k-2}-1)\,B_k\,B_{k-2}}[\sigma_{k-3}(\ell)-2^{k-2}\sigma_{k-3}(\ell/4)]
& \text{if $\ell\in\mathbb{N}$},
\end{cases}
$$
where $B_m$ is the $m$-th Bernoulli number and
$$
\sigma_k(m):=
\begin{cases}
0 & \text{if $m\notin \mathbb{N}$},\\
\displaystyle
\sum_{0<d|m}d^k & \text{if $m\in\mathbb{N}$}.
\end{cases}
$$
}
\end{Thm}
\subsection{${\boldsymbol U(p)}$-operator}
In the remainder of this paper, we assume that $p$ is an odd prime. For a formal power series of the form
$F=\sum _Ha_F(H)q^H$, we define a $U(p)$ type operator as
$$
U(p):F=\sum _Ha_F(H)q^T\longmapsto F|U(p):=\sum _Ha_F(pH)q^H.
$$
In particular, for a modular form $F\in M_k(\varGamma _0^{(n)}(p))$, we may regard $U(p)$ as a Hecke operator
(cf. \cite{Bo}, \cite{Kri3}).
We prove this in this section. More precisely, we prove that
\begin{Prop}
\label{Prop1}
\it{If $F\in M_k(\varGamma _0^{(n)}(p))$ then $F|U(p)\in M_k(\varGamma _0^{(n)}(p))$.}
\end{Prop}
To prove this proposition, we introduce the following lemma.
\begin{Lem}
\label{Lem1}
\it{A complete set of representatives for the left cosets of
$$
\varGamma _0^{(n)}(p)\begin{pmatrix}O_n & -1_n \\ 1_n & O_n \end{pmatrix}\varGamma _0^{(n)}(p)
$$
is given by
$$
\left\{ \begin{pmatrix}O_n & -1_n \\ 1_n & T \end{pmatrix}|\: T\in Her_n({\mathcal O})/pHer_n({\mathcal O}) \right\}.
$$
}
\end{Lem}
\begin{proof}[Proof of Lemma \ref{Lem1}]
We set $\gamma _T:=\begin{pmatrix} O_n & -1_n \\ 1_n & T \end{pmatrix}$ and prove
\begin{align*}
\varGamma _0^{(n)}(p)\begin{pmatrix}O_n & -1_n \\ 1_n & O_n \end{pmatrix}\varGamma _0^{(n)}(p)
=\bigcup _{T\in Her_n({\mathcal O})/pHer_n({\mathcal O})}\varGamma _0^{(n)}(p) \gamma _T.
\end{align*}
By decomposition
\begin{equation}
\label{dec}
\begin{pmatrix}O_n & -1_n \\ 1_n & T \end{pmatrix}=\begin{pmatrix}O_n & -1_n \\ 1_n & O_n
\end{pmatrix}\begin{pmatrix}1_n & T \\ O_n & 1_n \end{pmatrix},
\end{equation}
we easily see the inclusion
\begin{equation}
\label{inc}
\varGamma _0^{(n)}(p)\begin{pmatrix}O_n & -1_n \\ 1_n & O_n \end{pmatrix}\varGamma _0^{(n)}(p) \supset
\bigcup _{T\in Her_n({\mathcal O})/pHer_n({\mathcal O})}\varGamma _0^{(n)}(p) \gamma _T.
\end{equation}
We shall prove the converse inclusion. Note that $T\equiv T'$ mod $pHer_n({\mathcal O})$ if and only if
$\varGamma _0^{(n)}(p)\gamma _T=\varGamma _0^{(n)}(p)\gamma _{T'}$. Hence, we have
$$
\bigcup _{T\in Her_n({\mathcal O})/pHer_n({\mathcal O})}\varGamma _0^{(n)}(p) \gamma _T=\bigcup _{T\in Her_n({\mathcal O})}\varGamma _0^{(n)}(p)\gamma _T
$$
as a set. Again, by the decomposition (\ref{dec}), it suffices to show that, for any
$\begin{pmatrix}A & B \\ C & D\end{pmatrix}\in \varGamma _0^{(n)}(p)$, there exists $S\in Her_n({\mathcal O})$ such that
$$
\begin{pmatrix}O_n & -1_n \\ 1_n & T \end{pmatrix}\begin{pmatrix}A & B \\ C & D \end{pmatrix}
\begin{pmatrix}O_n & -1_n \\ 1_n & S\end{pmatrix}^{-1}\in \varGamma _0^{(n)}(p).
$$
A direct calculation shows that
\begin{align*}
&\begin{pmatrix}O_n & -1_n \\ 1_n & T \end{pmatrix}\begin{pmatrix}A & B \\ C & D \end{pmatrix}\begin{pmatrix}O_n & -1_n \\ 1_n & S
\end{pmatrix}^{-1}=\begin{pmatrix}-CS+D & -C \\ (A+TC)S-(B+TD) & A+TC \end{pmatrix}.
\end{align*}
Hence, the proof is reduced to finding $S\in Her_n({\mathcal O})$ such that $AS\equiv B+TD$ mod
$p\,M(n,{\mathcal O})$.
Recall that $A{}^t\overline{D}-B{}^t\overline{C}=1_n$ and hence $A{}^t\overline{D}\equiv 1_n$ mod $p\,M(n,{\mathcal O})$.
If we choose $S$ as $S:={}^t\overline{D}(B+TD)$, then $AS\equiv B+TD$ mod $p\,M(n,{\mathcal O})$.
To complete the proof, we need to show that $S={}^t\overline{D}(B+TD)\in Her_n({\mathcal O})$. This assertion comes from the fact that
${}^t\overline{D}B$, ${}^t\overline{D}TD\in Her _n({\mathcal O})$.
\end{proof}
We now return to the proof of Proposition \ref{Prop1}.
\begin{proof}[Proof of Proposition \ref{Prop1}]
Let $F\in M_k(\varGamma _0^{(n)}(p))$. From Lemma \ref{Lem1}, we have
\begin{align*}
F|\varGamma _0^{(n)}(p)\begin{pmatrix}O_n & -1_n \\ 1_n & O_n \end{pmatrix}\varGamma _0^{(n)}(p)&=
\sum_TF|_k \begin{pmatrix}O_n & -1_n \\ 1_n & T \end{pmatrix}\\
&=\sum _TF|W_p|_k\begin{pmatrix}1_n & T \\ O_n & p1_n \end{pmatrix},
\end{align*}
where $W_p$ is the Fricke involution
$$
F\longmapsto F|W_p:=F|_k\begin{pmatrix}O_n & -1_n \\ p1_n & O_n\end{pmatrix}.
$$
We see by the usual way that $F|W_p\in M_k(\varGamma _0^{(n)}(p))$. If we write $G=F|W_p=\sum _Ha_G(H)q^H$, then
\begin{align*}
\sum_TF|W_p|_k\begin{pmatrix}1_n & T \\ O_n & p1_n \end{pmatrix}&=\sum _TG|_k\begin{pmatrix}1_n & T \\ O_n & p1_n \end{pmatrix}\\
&=\sum _H\left(\sum _Te^{\frac{2\pi i}{p}\tau(H,T)}\right)a_G(H)e^{\frac{2\pi i}{p}\tau(H,Z)}\\
&=c\cdot G|U(p),
\end{align*}
where $c:=\sharp Her _n({\mathcal O})/pHer _n({\mathcal O})$ and the last equality follows from the following lemma.
\begin{Lem}
\label{Lem2}
\it{
For fixed $H\in Her _n^\tau ({\mathcal O})$, we have
\begin{align}
\label{sum}
\sum _Te^{\frac{2\pi i}{p}\tau (H,T)}=\begin{cases} 0 \quad &{\rm if}\quad H\not \in pHer^\tau _n({\mathcal O}), \\ c \quad &{\rm if}
\quad H\in pHer^\tau _n({\mathcal O}). \\ \end{cases}
\end{align}
}
\end{Lem}
\begin{proof}[Proof of Lemma \ref{Lem2}]
For $H\in Her_n^\tau ({\mathcal O})$, we define
\begin{align*}
G(H):=\sum _{T\in Her_n({\mathcal O})/pHer_n({\mathcal O})}e^{\frac{2\pi i}{p}\tau (H,T)}.
\end{align*}
This definition is independent of the choice of the representation $T$. Replacing $T$ by $T+S$, we obtain
$$
G(H)=G(H)e^{\frac{2\pi i}{p}\tau (H,S)}.
$$
Hence, $G(H)=0$ unless $e^{\frac{2\pi i }{p}\tau (H,S)}=1$; i.e., $\tau (H,S)\in p\mathbb{Z}$. This implies
$\tau (\frac{1}{p}H,S)\in \mathbb{Z}$
for all $S\in Her_n ({\mathcal O})$. The definition of a dual lattice yields
$$
\frac{1}{p}H\in Her_n^\tau ({\mathcal O}).
$$
\end{proof}
From this lemma, we have
$$
F|\varGamma _0^{(n)}(p)\begin{pmatrix}O_n & -1_n \\ 1_n & O_n \end{pmatrix}\varGamma _0^{(n)}(p)=c\cdot F|W_p|U(p).
$$
Hence, the action of $U(p)$ is described by the action of the double coset
$$
\varGamma _0^{(n)}(p)\begin{pmatrix}1_n & O_n \\ O_n & p1_n\end{pmatrix}\varGamma _0^{(n)}(p).
$$
Therefore, we have $F|U(p)\in M_k(\varGamma _0^{(n)}(p))$, which completes the proof of
Proposition \ref{Prop1}.
\end{proof}
\begin{Rem}
The proof of Lemma \ref{Lem2} is due to Krieg.
\end{Rem}
\section{Main results}
\subsection{Modularity of $p$-adic Eisenstein series}
In this subsection, we deal with a suitable constant multiple of
the normalized quaternionic Eisenstein series
$$
G_k=G^{(2)}_k:=(2^{k-2}-1)\frac{B_kB_{k-2}}{4k(k-1)}E_k^{(2)}
$$
and show that certain $p$-adic limits of this Eisenstein series are ``real" modular forms for $\varGamma _0^{(2)}(p)$.
We write the Fourier expansion of $G_k$ as $G_k=\sum _{H}b_k(H)q^H$. We remark that
\[b_k(O_2)=(2^{k-2}-1)\frac{-B_kB_{k-2}}{4k(k-2)}.\]
For an odd prime $p$ we put
\begin{align*}
G^*_k&:=\frac{-1}{1+p^{k-3}}\left\{p^{2(k-3)}(G_k|U(p)-p^{k-1}G_k)-(G_k|U(p)-p^{k-1}G_k)|U(p)\right\}\\
&\in M_k(\varGamma _0^{(n)}(p)),
\end{align*}
where this modularity follows from Proposition \ref{Prop1}. The first main theorem is
\begin{Thm}
\label{ThmM}
\it{Let $p$ be an odd prime and $k$ an even integer with $k\ge 4$. Define a sequence
$\{k_m\}$ by
$$k_m:=k+(p-1)p^{m-1}. $$
Then, the corresponding sequence of Eisenstein series $\{G_{k_m}\}$ has a $p$-adic limit
$G^*_k$ and we have
\begin{align}
\label{Lim}
\lim _{m\to \infty }G_{k_m}=G^*_k\in M_k(\varGamma _0^{(2)}(p)).
\end{align}
}
\end{Thm}
\begin{proof}
The proof of (\ref{Lim}) is reduced to show that $G_k^*$ is obtained by
removing all $p$-factors of the Fourier coefficients of the quaternionic Eisenstein series.
To calculate the Fourier coefficients of $G^*_k$, we set
$$
F_k=G_k|U(p)-p^{k-1}G_k.
$$
We can then rewrite $G^*_k$ as
$$
G^*_k=\frac{-1}{1+p^{k-3}}(p^{2(k-3)}F_k-F_k|U(p)).
$$
We write the Fourier expansions as
$$
G^*_k=\sum _HA_k(H)q^H,\quad F_k=\sum _HB_k(H)q^H.
$$
First, we calculate the constant term of $G^*_k$.
Since
$$
b_k(O_2)=(2^{k-2}-1)\frac{-B_kB_{k-2}}{4k(k-2)},
$$
the constant term of $G^*_k$ becomes
\begin{align*}
A_k(O_2)&=\frac{-1}{1+p^{k-3}}\{p^{2(k-3)}(b_k(O_2)-p^{k-1}b_k(O_2))-(b_k(O_2)-p^{k-1}b_k(O_2))\}\\
&=(1-p^{k-1})(1-p^{k-3})(2^{k-2}-1)\frac{-B_kB_{k-2}}{4k(k-2)}.
\end{align*}
Second, we calculate the coefficient $A_k(H)$ for $H$ with ${\rm rank}(H)=1$.
\begin{align*}
B_k(H)&=b_k(pH)-p^{k-1}b_k(H)\\
&=(2^{k-2}-1)\frac{B_{k-2}}{2(k-2)}\left(\sum_{0<d|p\varepsilon (H)}d^{k-1}-p^{k-1}\sum_{0<d|\varepsilon (H) }d^{k-1}\right)\\
&=(2^{k-2}-1)\frac{B_{k-2}}{2(k-2)}\sigma _{k-1}^*(\varepsilon (H)),
\end{align*}
where $\sigma _{m}^*(N)$ is defined as
$$
\sigma _{m}^*(N):=\sum _{\substack{0<d|N \\ (p,d)=1}}d^{m}.
$$
Note that $B_k(pH)=B_k(H)$ when ${\rm rank}(H)=1$. Hence, we have
$$
A_k(H)=(1-p^{k-3})(2^{k-2}-1)\frac{B_{k-2}}{2(k-2)}\sigma _{k-1}^*(\varepsilon (H)).
$$
Finally, we consider the case ${\rm rank}(H)=2$.
\begin{align*}
B_k(H)&=b_k(pH)-p^{k-1}b_k(H)\\
&=\sum _{0<d| p\varepsilon (H)}d^{k-1}[\sigma _{k-3}\left(\tfrac{2p^2\det H}{d^2}\right)
-2^{k-2}\sigma _{k-3}\left(\tfrac{2p^2\det H}{4d^2}\right)]\\
&-p^{k-1}\sum _{0<d| \varepsilon (H)}d^{k-1}[\sigma _{k-3}
\left(\tfrac{2\det H}{d^2}\right)-2^{k-2}\sigma _{k-3}\left(\tfrac{2\det H}{4d^2}\right)]\\
&=\sum _{\substack{ 0<d|\varepsilon (H) \\ (p,d)=1}}d^{k-1}[\sigma _{k-3}
\left(\tfrac{2p^2\det H}{d^2}\right)-2^{k-2}\sigma _{k-3}\left(\tfrac{2p^2\det H}{4d^2}\right)].
\end{align*}
Here, the last equality was obtained from the elemental property:
\begin{Lem}
\label{Lem0}
\it{
Let $p$ be a prime and $N$ a positive integer. For a function $f:\mathbb{N}\rightarrow \mathbb{N}$, the following holds:
\begin{align*}
\sum _{0<d|pN}f(d)=\sum _{\substack{ 0<d|N \\ (p,d)=1}}f(d)+\sum _{0<d|N}f(pd).
\end{align*}
}
\end{Lem}
Therefore,
\begin{align*}
A_k(H)&=\frac{-1}{1+p^{k-3}}(p^{2(k-3)}B_k(pH)-B_k(H))\\
&=\frac{-1}{1+p^{k-3}}\Big( p^{2(k-3)}\sum _{\substack{ 0<d|\varepsilon (H) \\ (p,d)=1}}d^{k-1}
[\sigma _{k-3}\left(\tfrac{2p^2\det H}{d^2}\right)-2^{k-2}\sigma _{k-3}\left(\tfrac{2p^2\det H}{4d^2}\right)] \\
&-\sum _{\substack{ 0<d|\varepsilon (H) \\ (p,d)=1}}d^{k-1}[\sigma _{k-3}\left(\tfrac{2p^4\det H}{d^2}\right)-2^{k-2}
\sigma _{k-3}\left(\tfrac{2p^4\det H}{4d^2}\right)] \Big).
\end{align*}
By repeatedly applying Lemma \ref{Lem0}, we obtain
$$
p^{2m}\sigma _{m}(N)-\sigma _{m}(p^2N)=-(1+p^{m})\sum _{\substack{ 0<d|N \\ (p,d)=1}}d^m.
$$
From this, we have
$$
A_k(H)=\sum _{\substack{ 0<d|\varepsilon (H) \\ (p,d)=1}}d^{k-1}[\sigma ^*_{k-3}\left(\tfrac{2\det H}{d^2}\right)-2^{k-2}\sigma ^*_{k-3}
\left(\tfrac{2\det H}{4d^2}\right)].
$$
Summarizing these calculations, we obtain the following formula:
\begin{Prop}
\it{The following holds:}
\begin{align*}
A_k(H)=\begin{cases}
\displaystyle (1-p^{k-1})(1-p^{k-3})(2^{k-2}-1)\frac{-B_kB_{k-2}}{4k(k-2)},\ & \text{if}\; H=O_2,
\vspace{2mm}\\
\displaystyle (1-p^{k-3})(2^{k-2}-1)\frac{B_{k-2}}{2(k-2)}\sigma _{k-1}^*(\varepsilon (H)),\ &
\text{if\;rank}(H)=1,
\vspace{2mm}\\
\displaystyle \sum _{\substack{ 0<d|\varepsilon (H) \\ (p,d)=1}}d^{k-1}[\sigma ^*_{k-3}\left(\tfrac{2\det H}{d^2}\right)-2^{k-2}\sigma ^*_{k-3}
\left(\tfrac{2\det H}{4d^2}\right)],\ &
\text{if\;rank}(H)=2.
\end{cases}
\end{align*}
\end{Prop}
On the other hand,
\begin{align*}
b_{k_m}(H)=
\begin{cases}
\displaystyle (2^{k_m-2}-1)\frac{-B_{k_m}B_{k_m-2}}{4k_m(k_m-2)},\ &
\text{if}\; H=O_2,
\vspace{2mm}\\
\displaystyle (2^{k_m-2}-1)\frac{B_{k_m-2}}{2(k_m-2)}\sigma _{k_m-1}(\varepsilon (H)),\ &
\text{if\;rank}(H)=1,
\vspace{2mm}\\
\displaystyle \sum _{0<d|\varepsilon (H)}d^{k_m-1}[\sigma_{k_m-3}
\left(\tfrac{2\det H}{d^2}\right)-2^{k_m-2}\sigma_{k_m-3}\left(\tfrac{2\det H}{4d^2}\right)],\ &
\text{if\;rank}(H)=2.
\end{cases}
\end{align*}
Combining these formulas and the Kummer congruence, we can prove that
$$
\lim _{m\to \infty }b_{k_m}(H)=A_k(H)
$$
for all $H\in Her_2^\tau(\mathcal{O})$. This completes the proof of Theorem \ref{ThmM}.
\end{proof}
\begin{Rem}
\it{
Following Hida \cite{Hida}, our $G_k^*$ can be p-adic analytically interpolated with respect to the weight.
}
\end{Rem}
\subsection{Transcendental $p$-adic Eisenstein series}
As we have seen in the previous section, under certain conditions, a $p$-adic limit of a quaternionic
Eisenstein series becomes a ``real" modular form
with rational Fourier coefficients. This also holds for Siegel Eisenstein and Hermitian
Eisenstein series. More precisely, they coincide with the genus theta series
(cf. \cite{KN}, \cite{N}). In these cases (Siegel, Hermitian cases), the $p$-adic Eisenstein
series is algebraic. We shall show that there exists an example of a transcendental
$p$-adic Eisenstein series for quaternionic modular forms.
The second main theorem is
\begin{Thm}
\label{ThmM2}
\it{
Let $p$ be an odd prime and $\{k_m\}$ the sequence defined by
$$k_m:=2+(p-1)p^{m-1}.$$
Then, the $p$-adic Eisenstein series $\displaystyle\widetilde{E}=\lim_{m\to\infty}E_{k_m}^{(2)}$
is transcendental;
namely, $\widetilde{E}$ has transcendental coefficients
where $E_k^{(2)}$ is the normalized quaternionic Eisenstein series of degree 2 defined
in (\ref{Eisen}).
}
\end{Thm}
\begin{proof}
We calculate $\tilde{a}(H):=\displaystyle\lim_{m\to\infty}a_{k_m}(H)$ at
$H=\begin{pmatrix}1&\tfrac{e_1+e_2}{2}\\\tfrac{e_1-e_2}{2}&1\end{pmatrix}\in Her_2^\tau (\mathcal{O})$.
The convergence for general $H$ is proved similarly.
It follows from Theorem \ref{Four} that
$$
a_{k_m}(H)=-\frac{4k_m(k_m-2)}{(2^{k_m-2}-1)B_{k_m}B_{k_m-2}}.
$$
(We note that $\varepsilon (H)=1$ and $\det (H)=\frac{1}{2}$.)
We rewrite the right-hand side as
$$
-4\cdot \frac{2+(p-1)p^{m-1}}{B_{2+(p-1)p^{m-1}}}\cdot
\frac{1}{B_{(p-1)p^{m-1}}}\cdot
\frac{p^m}{2^{(p-1)p^{m-1}}-1}\cdot
\frac{p-1}{p}
$$
and calculate the $p$-adic limit separately:\\
(i)\; $\displaystyle\lim_{m\to\infty}\frac{2+(p-1)p^{m-1}}{B_{2+(p-1)p^{m-1}}}=\frac{B_2}{2}=\frac{1}{12}$.\\
This is a consequence of the Kummer congruence.\\
(ii)\; $\displaystyle\lim_{m\to\infty}B_{(p-1)p^{m-1}}=\frac{p-1}{p}$.\\
This identity comes from the fact that the residue of the $p$-adic $L$-function $L_p(s,\chi^0)$ at
$s=0$ is just $1-\frac{1}{p}$.\\
(iii)\; $\displaystyle\lim_{m\to\infty}\frac{2^{(p-1)p^{m-1}}-1}{p^m}=\frac{\log_p(2^{p-1})}{p}$.\\
where $\log_p$ is the $p$-adic logarithmic function defined by
$$
\log_p(x)=x-\frac{x^2}{2}+\frac{x^3}{3}-\cdots ,\qquad (|x|_p<1).
$$
Leopoldt's formula \cite{L} states that
$$
\lim_{m\to\infty}\frac{x^{p^m}-1}{p^m}=\log_p(x).
$$
if $|x-1|_p<1$ .
This implies that
$$
\lim_{m\to\infty}\frac{2^{(p-1)p^{m-1}}-1}{p^m}=\frac{1}{p}\cdot\log_p(2^{p-1}).
$$
Combining these formulas, we obtain
\begin{equation}
\tilde{a}(H)=\lim_{m\to\infty}a_{k_m}(H)=\frac{-48p}{\log_p(2^{p-1})}.
\end{equation}
We shall show that $\log_p(2^{p-1})$ is transcendental. Let $\text{exp}_p$ be the $p$-adic exponential function defined by
$$
\text{exp}_p(x)=1+x+\frac{x^2}{2!}+\frac{x^3}{3!}+\cdots,\qquad (|x|_p<p^{-\frac{1}{p-1}}).
$$
It is known that if $|x|_p<p^{-\frac{1}{p-1}}$, then
$$
\text{exp}_p(\log_p(1+x))=1+x,\quad (\text{e.g., \cite{G}}).
$$
To prove the transcendency of $\log_p(2^{p-1})$, we use the following theorem by Mahler:
\begin{Thm}[Mahler \cite{M}]\it{
Let $\mathbb{C}_p$ be the completion of the algebraic closure of $\mathbb{Q}_p$.
For any algebraic over $\mathbb{Q}$ $p$-adic number
$\alpha\in\mathbb{C}_p$ with
$0<|\alpha|_p<p^{-\frac{1}{p-1}}$, the quantity
$\text{exp}_p(\alpha)$ is transcendental.}
\end{Thm}
We note that $|x|_p<p^{-\frac{1}{p-1}}$ is equivalent to $|x|_p<1$ for odd prime $p$
(e.g., \cite{G}, p.114). We put $\alpha=2^{p-1}-1$. Since $|\alpha|_p<1$,
we have
$$
\text{exp}_p(\log_p(1+\alpha))=1+\alpha=2^{p-1}.
$$
The right-hand side is obviously algebraic. Hence, by Mahler's theorem,
$\log_p(1+\alpha)=\log_p(2^{p-1})$ must be transcendental. Thus, we can
prove the transcendency of $\tilde{a}(H)$ at
$H=\begin{pmatrix}1&\frac{e_1+e_2}{2}\\ \frac{e_1-e_2}{2}&1\end{pmatrix}$.
This completes the proof of Theorem \ref{ThmM2}.
\end{proof}
\begin{Rem}
By the above proof, we see that all coefficients $\tilde{a}(H)$ corresponding to
$H$ with rank $2$ are transcendental. However, $\tilde{a}(H)$ for $H$ with
$\text{rank}(H)\leq 1$ are rational.
\end{Rem}
\noindent
\textbf{Acknowledgments:}
We would like to thank Professor A. Krieg for helpful comments on the proof of
the modularity of $f|U(p)$.
We also thank Professor M.~Amou for pointing out the transcendency of
$\log_p(2^{p-1})$.
|
\section{Introduction}
The currently favored cosmological model, Lambda$+$ Cold Dark Matter
($\Lambda$CDM), is remarkably successful at reproducing the large-scale
structure of the Universe (Blumenthal et al.\ 1984; Springel et al.\ 2005).
However, small-scale observations have proven harder to explain.
High-resolution N-body simulations of $\Lambda$CDM structure formation predict
that the central density profiles of dark matter halos should rise steeply at
small radii, $\rho(r) \propto r^{-\gamma}$, with $\gamma\simeq 1 - 1.5$
(Navarro, Frenk \& White 1997, henceforth NFW; Navarro et al.\ 2004; Diemand
et al.\ 2005).
Observations of rotation curves of late-type disk galaxies and dwarf galaxies,
on the other hand, have shown that quite often, mass distributions with lower
than predicted densities or with constant density cores, where
$\gamma\simeq 0$ (i.e., a pseudo-isothermal profile), are preferred (Swaters et al.\ 2003; Gentile et al.\ 2004, 2005; Simon et al.\ 2005; Kuzio de Naray et
al.\ 2006, 2008; Shankar et al.\ 2006; Spano et al.\ 2008). This is known
as the cusp/core problem. One possibility is that these observations are
pointing to a real problem with $\Lambda$CDM cosmology, perhaps indicating
that the dark matter is not cold, but rather warm (Zentner \& Bullock 2002),
in which case it is easier to produce constant density cores at the centers
of dark matter halos. Another possibility is that these late-type galaxies
have constant density cores because of their late formation (Wechsler et al.\
2002) and that earlier-type bulge-dominated galaxies (which form at earlier
times) will tend to conform to the standard expectations of the theory. This
is because the central mass densities of galaxies tend to reflect the density
of the Universe at their formation time (Wechsler et al.\ 2002).
In this paper we have chosen to model the H{\tt I} rotation curve of M33 from
Crobelli \& Salucci (2000). Due to its proximity, M33 can be studied in
exquisite detail, and it therefore provides a crucial testing ground of our
ideas of galaxy formation. Its Hubble classification is SA(s)cd (de
Vaucouleurs et al.\ 1991), meaning that is of particularly late-type, with
little or no bulge. This is reflected in the central supermassive black hole
mass of $M_{\rm BH}<1500$ M$_{\odot}$ (Gebhardt et al.\ 2001), and black hole
masses tend to be related to the central bulge mass (Magorrian et al.\ 1998;
H\"aring \& Rix 2004). In this paper we model the rotation curve of M33 with
both a pseudo-isothermal profile dark matter halo density model and an NFW
dark matter halo density model. We then use parameters derived from these
fits to look at relations between the dark matter halo and other galaxy
properties, such as supermassive black hole mass and spiral arm pitch angle.
This paper is organized as follows. Section 2 describes the observed data and
data analysis. Section 3 describes how the rotation curve is modeled and how
we derive the baryonic and dark matter halo contributions to the rotation
curve. Section 4 discusses our results and Section 5 summarizes our findings.
Throughout this paper, we assume a flat $\Lambda$CDM cosmology with
$\Omega_m=0.27$ and a Hubble constant $H_0 = 75$ km s$^{-1}$ Mpc$^{-1}$.
\section{Observations and Data Reduction}
We have made use of the {\em Spitzer}/IRAC 3.6-$\mu$m image of M33. The
IRAC observations were taken as part of the Gehrz Guaranteed Time Observer
Program ID 5. The mapping sequence for each epoch consisted of $\simeq 148$
positions per channel. Each position was observed with three 12 s frames
dithered with the standard, small, cycling pattern. The FWHM of the
point-spread function (PSF) at 3.6-$\mu$m is 1.7$^{\prime\prime}$ or 6.9 pc
at the distance of M33. The final mosaic spans an area of $\sim1.^{\prime}0\times
1.^{\prime}2$. We adopt a distance to M33 of $d=840$ kpc (e.g., Magrini,
Corbelli \& Galli 2007), and
it has a redshift of $z=-0.000597$ (de Vaucouleurs et al.\ 1991).
For dynamical measurements, we make use of the H{\tt I} rotation curve of Corbelli \& Salucci (2000). We also make use of the inclination corrected H{\tt I} linewidth from HyperLeda\footnote{http://leda.univ-lyon.fr/} of $100.4\pm3.0$ km s$^{-1}$ (e.g., Paturel et al.\ 2003).
For the determination of the spiral arm morphology we have made use of an $R$
band image from the Digital Sky Survey (DSS).
\subsection{Measurement of spiral arm pitch angle}
\label{mpitch}
Spiral arm pitch angles are measured using a two-dimensional fast Fourier
decomposition technique, which employs a program described in Schr\"oder et
al. (1994). Logarithmic spirals are assumed in the
decomposition.
The amplitude of each Fourier component is given by
\begin{equation}
A(m,p)=\frac{\sum_{i=1}^{I}\sum_{j=1}^{J}I_{ij}(\ln{r},\theta)\exp{-[i(m\theta _p\ln{r})]}}{\sum_{i=1}^{I}\sum_{j=1}^{J}I_{ij}(\ln{r},\theta)}
\end{equation}
where $r$ and $\theta$ are polar coordinates, $I(\ln{r},\theta)$ is the
intensity at position $(\ln{r},\theta)$, $m$ represents the number of arms
or modes, and $p$ is the variable associated with the pitch angle $P$, defined
by $P=-(m/p_{max})$. Throughout this work we measure the pitch angle $P$ of the
$m=2$ component.
Pitch angles are
determined from peaks in the Fourier spectra, as this is the most powerful
method to find periodicity in a distribution (Consid\`ere \& Athanassoula 1988;
Garcia-Gomez \& Athanassoula 1993).
The image was first projected to face-on. Mean uncertainties of position
angle and inclination as a function of inclination were discussed by
Consid\`ere \& Athanassoula (1988). For a galaxy with low inclination, there
are clearly greater uncertainties in assigning both a position angle and an
accurate inclination. These uncertainties are discussed by Block et al.\ (1999)
and Seigar et al.\ (2005, 2006), who took a galaxy with low inclination
($<30^{\circ}$) and one with high inclination ($>60^{\circ}$) and varied the
inclination angle used in the correction to face-on. They found that for the
galaxy with low inclination, the measured pitch angle remained the same. M33
has a relatively low inclination of $\sim 30^{\circ}$, and so the uncertainty
in the inclination angle in this case, does not result in a large error in
the pitch angle we measure for M33.
Our deprojection method assumes that spiral galaxy
disks are intrinsically circular in nature.
\section{Mass modeling}
\subsection{The baryonic contribution}
Our goal is to determine a mass model for M33 from direct
fitting of mass models to its rotation curve. We perform a bulge-disk
decomposition in order to estimate the baryonic contribution. We then
determine several different models and try to recreate the nuclear spiral
by minimizing reduced-$\chi^2$.
We first extract the surface brightness of M33 using the {\em Spitzer}
3.6-$\mu$m image and the IRAF {\tt ELLIPSE} routine, which fits ellipses
to an image using an iterative method described by Jedrzejewski (1987). In
order to mask out foreground stars, SE{\tt XTRACTOR} (Bertin \& Arnouts 1996)
was used. An inclination correction was then applied to the surface brightness
profile (de Jong 1996; Seigar \& James 1998) as follows
\begin{equation}
\mu_i = \mu -2.5 C \log {\left(\frac{a}{b}\right)}
\end{equation}
where $\mu_i$ is the surface brightness when viewed at some inclination $i$,
$\mu$ is the corrected surface brightness, $a$ is the major axis, $b$ is the
minor axis and $C$ is a factor dependent on whether the galaxy is optically
thick or thin; if $C=1$ then the galaxy is optically thin; if $C=0$ then the
galaxy is optically thick (e.g., Seigar \& James 1998; de Jong 1996).
Graham (2001a)
showed that $C=0.91$ is a good value to use for the near-infrared $K_s$ band.
Adopting a simple reddening law, where extinction falls as the square of
wavelength, it can be shown that a value of $C=0.97$ is appropriate at
3.6-$\mu$m (Seigar, Barth \& Bullock 2008a) and we adopt this value here.
\begin{figure}
\special{psfile=m33_SB.eps
hscale=40 vscale=40 hoffset=70 voffset=-270 angle=0}
\vspace*{7cm}
\caption{The {\em Spitzer} 3.6-$\mu$m surface brightness profile with decomposition into bulge and disk components. The bulge has been fitted with a S\'ersic model (short-dashed line) and the disk has been fitted with an exponential model (long-dashed line).}
\label{SB}
\end{figure}
\begin{table}
\caption{M33 Observational data. The Hubble type is from de Vaucouleurs et al.\ (1991). The distance in kpc is taken from Magrini et al.\ (2007).}
\begin{footnotesize}
\begin{tabular}{ll}
\hline
Parameter & Measurement \\
\hline
Hubble Type & SA(s)cd \\
Distance (kpc) & 840 \\
Position angle of major axis ($^{\circ}$) & 23 \\
Bulge effective radius, $R_e$ (arcmin) & 1.60$\pm$0.11 \\
Bulge effective radius, $R_e$ (kpc) & 0.39$\pm$0.03 \\
Bulge surface brightness at the effective radius, $\mu_e$ (3.6 $\mu$m-mag arcsec$^{-2}$) & 19.57$\pm$0.98 \\
Bulge S\'ersic index, $n$ & 1.0 \\
Disk central surface brightness, $\mu_0$ (3.6 $\mu$m-mag arcsec$^{-2}$) & 18.08$\pm$1.02 \\
Disk scalelength, $h$ (arcmin) & 6.95$\pm$0.49 \\
Disk scalelength, $h$ (kpc) & 1.70$\pm$0.12 \\
Disk luminosity, $L_{disk}$ (L$_{\odot}$) & $(3.16\pm0.30)\times10^{9}$ \\
Bulge-to-disk ratio, $B/D$ & 0.03 \\
\hline
\end{tabular}
\end{footnotesize}
\end{table}
The resulting surface brighntess profile {Fig.\ \ref{SB} reaches a surface
brightness of $\mu_{3.6}\sim20.7$ mag arcsec$^{-2}$ at a radius of
$\sim$13.2 kpc
(equivalent to 54.0 arcmin). From this surface brightness profile, we
perform a one-dimensional bulge-disk decomposition, which employs the
S\'ersic model for the bulge component and an exponential law for the disk
component (e.g., Andredakis, Peletier \& Balcells 1995; Seigar \& James 1998;
Khosroshahi, Wadadekar \& Kembhavi 2000; D'Onofrio 2001; Graham 2001b;
M\"ollenhoff \& Heidt 2001; see Graham \& Driver 2005 for a review). The S\'ersic
(1963, 1968) $R^{1/n}$ model is most commonly expressed as a surface brightness
profile, such that
\begin{equation}
\mu(R)=\mu_e \exp{\left(-b_n\left[\left(\frac{R}{R_e}\right)^{1/n}-1\right]\right)},
\end{equation}
where $\mu_e$ is the surface brightness at the effective radius $R_e$ that
encloses half of the total light from the model (Ciotti 1991; Caon, Capaccioli
\& D'Onofrio 1993). The constant $b_n$ is defined in terms of the parameter
$n$, which described the overall shape of the light profile. When $n=4$,
the S\'ersic model is equivalent to a de Vaucouleurs (1948, 1959) $R^{1/4}$
model and when $n=1$ it is equivalent to an exponential model. The parameter
$b_n$ has been approximated by $b_n = 1.9992n - 0.3271$, for $0.5<n<10$
(Capaccioli 1989; Prugniel \& Simien 1997). The exponential model for the
disk surface brightness profile can be written as follows
\begin{equation}
\mu(R) = \mu_0 \exp{(-R/h)}
\end{equation}
where $\mu_0$ is the disk central surface brightness and $h$ is the disk
exponential scalelength. The results of our surface brightness fitting are
summarized in Table 1.
We now assign masses to the disk and bulge of M33. The stellar mass-to-light
ratio in the $K_s$ band is a well-calibrated quantity (Bell et al.\ 2003) which
depends on $B-R$ color. Seigar et al.\ (2008a) extended this to a 3.6-$\mu$m
image of M31 using the population synthesis codes of Bruzual \& Charlot (2003)
and Maraston (2005). Using their results, we find a central mass-to-light
ratio of $M/L_{3.6}\simeq 1.25\pm0.10$ with a gradient of -0.014 kpc$^{-1}$.
This results in a disk mass of $M_{\rm disk}=(3.81\pm0.47)\times10^{9}$
M$_{\odot}$ and a bulge mass of $M_{\rm bulge}=(1.14\pm0.14)\times10^{8}$
M$_{\odot}$ for M33.
A concern in using the 3.6-$\mu$m {\em Spitzer} waveband to determine
the underlying stellar mass, is the effect of emission from hot dust
in this waveband, although this is probably only important in or
near HII regions. In order to place some constraint on this, we have
chosen to explore the emission from dust in the near-infrared $K$
band at 2.2 $\mu$m. Using near-infrared spectroscopy at 2.2 $\mu$m, it has
been shown that hot dust can account for up to 30 per cent of the
continuum light observed at this wavelength in areas of active star
formation, i.e., spiral arms (James \& Seigar 1999). When averaged
over the entire disk of a galaxy, this reduces to a $2$ percent effect,
if one assumes that spiral arms can be up to 12$^{\circ}$ in width. At 3.6
$\mu$m,
this would therefore result in 3 percent of emitted light from dust.
Another concern for the 3.6-$\mu$m waveband would be the contribution
from the polycyclic aromatic hydrocarbon (PAH) emission
feature at 3.3 $\mu$m. However, an Infrared Space Observatory (ISO)
spectroscopic survey of actively star-forming galaxies by
Helou et al. (2000) found that the 3.3-$\mu$m feature was very
weak when they analysed the average 2.5--11.6-$\mu$m spectrum of 45
galaxies. The contribution of the PAH feature to the 3.6-$\mu$m Spitzer
waveband, is therefore not a major concern.
One other important contribution to the baryonic mass of M33 is the gas mass.
Corbelli \& Salucci (2000) have shown that beyond a radius of 10 kpc, the gas
contributes about the same to the rotation curve as the stars. Since, the
best current estimate of the gas distribution comes from Corbelli \& Salucci
(2000), we have chosen to adopt their model for the distribution of gas mass
in M33.
\subsection{The dark halo contribution}
\begin{figure}
\label{totrot}
\special{psfile=total_model.eps
hscale=40 vscale=40 hoffset=70 voffset=-270 angle=0}
\vspace*{7cm}
\caption{The H{\tt I} rotation curve from Corbelli \& Salucci (2000) modeled using a pseudo-isothermal model (core model; blue solid line) and a NFW model (red dotted line). The squares represent the total rotation velocities, and the circles represent the contribution of the dark matter to the rotation velocities (after subtraction of the stellar and gas mass).}
\end{figure}
A range of allowed dark matter halo masses and density profiles is
now explored, using two models for dark matter halo density profiles, the
pseudo-isothermal model (e.g., Simon et al.\ 2005; see Equation \ref{pseudoi})
and the Navarro, Frenk \& White (1997; hereafter NFW) profile. A pseudo-isothermal density profile is given by
\begin{equation}
\label{pseudoi}
\rho(R)=\rho_{0} \frac{R_{c}^{2}}{R_{c}^{2}+R^{2}},
\end{equation}
which in terms of rotational velocity becomes
\begin{equation}
V_{c}^{2}(R)=V_{c}^{2}(\infty)\left(1-\frac{R_c}{R}\tan^{-1}\frac{R}{R_c}\right),
\label{Vmod}
\end{equation}
where $R_c$ is the core radius, and
$\rho_0=V_{c}^{2}(\infty)/4\pi G R_{c}^{2}$.
The NFW profile is given by
\begin{equation}
\label{nfweqn}
\rho(R)=\frac{\delta_c \rho_c^0}{(R/R_s)(1+R/R_s)^2}
\end{equation}
where $R_s$ is a characteristic `inner' radius, and $\rho_c^0$ is the present
critical density and $\delta_c$ a characteristic overdensity.
This overdensity is defined as
\begin{equation}
\delta_c=\frac{100c_{\rm vir}^{3}}{3}
\end{equation}
where $c_{\rm vir}$=$R_{\rm vir}/R_s$ is the concentration parameter and
\begin{equation}
g(c_{\rm vir})=\frac{1}{\ln{(1+c_{\rm vir})}-c_{\rm vir}/(1+c_{\rm vir})}.
\end{equation}
The circular velocity associated with this density is given by Battaglia et
al.\ (2005) and is
\begin{equation}
V_{c}^{2}=\frac{V_{\rm vir}^{2}g(c_{\rm vir})}{s}\left[\ln{(1+c_{\rm vir}s)}-\frac{c_{\rm vir}s}{1+c_{\rm vir}s}\right]
\end{equation}
where $V_{\rm vir}$ is the circular velocity at the virial radius $R_{\rm vir}$ and $s=R/R_{\rm vir}$.
This NFW profile is a two parameter function and completely
specified by choosing two independent parameters, e.g., the virial
mass $M_{\rm vir}$ (or virial radius $R_{\rm vir}$) and concentration
$c_{\rm vir}=R_{\rm vir}/R_s$ (see
Bullock et al. 2001a for a discussion). Similarly, given a virial mass
$M_{\rm vir}$ and the dark matter circular velocity at any radius, the halo
concentration $c_{\rm vir}$ is completely determined.
We now proceed by finding the best-fitting NFW and pseudo-isothermal (or
constant density core) dark
matter halo density profiles that describe the complete H{\tt I}
rotation curve of
M33 as observed by Corbelli \& Salucci (2000). The result
of this is shown in Figure 2. The pseudo-isothermal fit is shown
as the solid blue line,
with best-fitting parameters of $V(\infty)=105.4\pm6.1$ km
s$^{-1}$ and $R_c=1.39\pm0.04$ kpc, and a reduced-$\chi^2$ value of
$\chi^2/\mu=3.19$, where $\mu$ is the degrees of freedom.
The NFW fit is shown as a dotted red line, with best-fitting parameters
$c_{\rm vir}=4.0\pm1.0$ and $M_{\rm vir}=(2.2\pm0.1)\times10^{11}$
M$_{\odot}$, with a
reduced-$\chi^2$ value of $\chi^2/\mu=1.18$. As can be seen from Figure
2, the pseudo-isothermal
model (or core model in the figure)
underestimates the rotation velocities beyond $\sim$7 kpc. However, the
NFW fit more closely recreates
the observed data. This is also clear from the values of reduced-$\chi^2$.
We therefore conclude that the NFW model best
represents these data, and this is consistent with the results of
Corbelli \& Salucci (2000). This is somewhat surprising for a late-type,
bulgeless galaxy like M33, since these late-type galaxies are often
shown to have constant density cores (e.g., Kuzio de Naray, 2006, 2008).
Table 2 lists the best-fit parameters of the
best-fit NSF and pseudo-isothermal models based upon direct fitting to the
H{\tt I} rotation curve data.
\begin{table*}
\label{compare}
\caption{M33 rotation curve modeling results, showing the best-fitting NFW and pseudo-isothermal models.}
\begin{center}
\begin{tabular}{ll}
\hline
Parameter & NFW model\\
\hline
$M_{\rm vir}$ & $(2.2\pm0.1)\times10^{11}$ M$_{\odot}$\\
$c_{\rm vir}$ & $4.0\pm1.0$\\
$\chi^2/\mu$ & 1.18\\
\hline
Parameter & Core model\\
\hline
$R_c$ & $1.39\pm0.04$ kpc\\
$V(\infty)$ & $105.4\pm6.1$ km s$^{-1}$\\
$\chi^2/\mu$ & 3.19 \\
\hline
\end{tabular}
\end{center}
\end{table*}
It is probably worthwhile noting that our best-fitting NFW model yields a
concentration parameter, $c_{\rm vir}=4.0\pm1.0$. This is somewhat lower than
the concentration parameter of $c_{\rm vir}=5.6$ reported by Corbelli \&
Salucci (2000). Furthermore, we derive a virial mass of
$M_{\rm vir}=(2.2\pm0.1)\times10^{11}$ M$_{\odot}$, which is significantly
lower than the virial mass of $M_{\rm vir}=7.4\times10^{11}$ M$_{\odot}$
found by Corbelli \& Salucci (2000). Here we discuss some reasons that could
account for these apparent differences. Since we use the same gas
distribution as Corbelli \& Salucci (2000), the only difference can come
from the stellar mass component. The main difference between our stellar
mass component, and that of Corbelli \& Salucci (2000), is that ours is
determined from a Spitzer 3.6-$\mu$m observed in 2007, and that of Corbelli \&
Salucci (2000) is determined from a $K$ band image reported by (Regan \&
Vogel 1994). The $K$ band image from 1994 was taken when near-infrared arrays
were really in their infancy, and so it is probably more important to rely
on the more modern datasets when possible. Furthermore, Corbelli \& Salucci
(2000) assume a distance to M33 of 0.7 Mpc, whereas we use the more accurate
measurement of 0.84 Mpc from Magrini et al.\ (2007). As a result of this
underestimate in the distance to M33, Corbelli \& Salucci (2000) have
underestimated the size of the visible galaxy by a factor of $\sim$17 percent,
and this in turn has probably affected the total mass of M33 that they
derive. Taking into account the different distances to M33, the disk
scalelength of $h=1.2$ kpc used by Corbelli \& Salucci (2000) would become
$h=1.4$ kpc if they had used the more accurate distance of 0.84 Mpc. This is
still lower than the scalelngth of $h=1.7$ kpc that we report here. In
converting this light distribution into stellar mass, we have then used a
combination of the stellar
mass-to-light ratios from Bell et al.\ (2003) and the
population synthesis codes from Maraston (2005). These papers provide the
best estimates currently available for determining the stellar mass-to-light
ratios, and they were not available to Corbelli \& Salucci when they performed
their analysis. One final difference between our results, and those of
Corbelli \& Salucci (2000), is that we include the bulge mass, although
considering the bulge-to-disk ratio of $B/D=0.03$ this is unlikely to have a
significant effect on the mass models. As a result, we conclude that the
differences between our results and those of Corbelli \& Salucci (2000), are
caused by the different treatment of the disk starlight, updated stellar
mass-to-light ratios, and more recent data.
Finally, it should be noted that Corbelli \& Walterbos (2007) revealed that
M33 has a weak central bar. This could potentially have the affect of
inducing non-circular motions in the central regions, i.e., within 1 kpc.
However, Kuzio de Naray \& Kaufmann (2010) have shown that, even in the case
of barred galaxies, it is difficult to confuse an NFW dark matter halo profile
with that of a pseudo-isothermal profile. In other words, our result that
M33 is best described by an NFW profile, still holds, and given that the
potential of the stellar bar is weak, the concentration is unlikely to change
significantly.
In the following discussion, we use
the NFW concentration parameter to reveal some interesting relationships.
\section{Discussion}
Seigar et al.\ (2004, 2005, 2006) have demonstrated that a relationship
exists between spiral arm pitch angle and rotation curve shear. Rotation curve
shear is defined as:
\begin{equation}
S=\frac{A}{\omega}=\frac{1}{2}\left(1-\frac{R}{V}\frac{dV}{dR}\right),
\end{equation}
where $A$ is the first Oort constant, $\omega$ is the angular velocity, and
$V$ is the velocity measured at radius $R$. Using this equation it is possible
to determine the shear from a rotation curve. We have performed such an analysis on the H{\tt I} rotation curve for M33 and found a value for its shear of
$S=0.46\pm0.01$. We have also measured the spiral arm pitch angle for M33,
which turns out to be $P=42.^{\circ}2\pm0.^{\circ}3$ (Seigar et al.\ 2008b).
This pitch angle is in good agreement with previous measurements
(Sandage \& Humphreys 1980; Block et al.\ 2004).
Figure 3 shows the relationship between
spiral arm pitch angle and rotation curve shear. One can easily see that the
pitch angle and shear values for M33 are consistent with the overall
relationship.
\begin{figure}
\label{shear_vs_pitch}
\special{psfile=pitch_vs_shear.eps
hscale=40 vscale=40 hoffset=70 voffset=-270 angle=0}
\vspace*{7cm}
\caption{Spiral arm pitch angle versus rotation curve shear, showing a strong
correlation. The solid squares represent galaxies with data measured by Block
et al.\ (1999), the red squares are galaxies from Seigar et al.\ (2005), the
blue squares are galaxies from Seigar et al.\ (2006), the cyan square is for
Malin 1 (Seigar 2008), the green square is for M31 (Seigar et al.\ 2008a) and
the magenta square represents the data for M33 (this paper).}
\end{figure}
Given the spiral arm pitch angles of a number of other galaxies, we can
also now compare this quantity with the NFW concentration parameters for
the galaxies listed in Table 3. Figure 4
shows a plot of NFW concentration as a function of spiral arm pitch angle
in degrees. This plot may only be for 5 galaxies, but a relatively strong
correlation appears to exist between these two quantities. Indeed Pearson's
linear correlation coefficient is 0.95 for this plot, although the significance
at which the null hypothesis of zero correlation is disproved in onlt 54 percent, probably due to low number statistics. Nevertheless, an interesting
correlation seems to exist between spiral arm morphology and dark matter concentration, and this could be further studied by targeting more galaxies in an observational campaign. Indeed, these data seem consistent with the suggestion
that pitch angle and mass concentration are related (Seigar et al.\ 2005, 2006).
\begin{table*}
\label{c_vs_pitch}
\caption{Spiral arm pitch angles, NFW concentration parameters, and central supermassive black hole mass for 5 galaxies. For the Malin 1 the spiral arm pitch angle is taken from Moore \& Parker (2006). For M31 the pitch angle is the average of values taken from Arp (1964) and Braun (1991). The NFW concentration value is taken from (1) Seigar (2008), (2) Klypin et al.\ (2002), (3) Seigar et al.\ (2008a), (4) Seigar et al.\ (2006). The black hole mass estimates are taken from (5) Ghez et al.\ (2005), (6) Bender et al.\ (2005), (7) Gebhardt et al.\ (2001).}
\begin{center}
\begin{footnotesize}
\begin{tabular}{llllll}
\hline
Galaxy name & Spiral arm pitch angle & $c_{\rm vir}$ & Source & $M_{\rm BH}$ & Source \\
& (degrees) & & & (M$_{\odot}$)\\
\hline
Malin 1 & $25.0\pm1.0$ & $8.0\pm1.0$ & (1) & -- & -- \\
Milky Way & -- & 12.0 & (2) & $(3.7\pm0.2)\times10^{6}$ & (5) \\
M31 & $7.1\pm0.4$ & $20.0\pm1.1$ & (3) & $(1.7\pm0.6)\times10^{8}$ & (6)\\
M33 & $42.2\pm3.0$ & $4.0$ & & $<1500$ & (7)\\
IC2522 & $38.8\pm1.6$ & $8.0\pm1.0$ & (4) & -- & -- \\
ESO582G12 & $22.6\pm0.6$ & $22.0\pm5.0$ & (4) & -- & -- \\
\hline
\end{tabular}
\end{footnotesize}
\end{center}
\end{table*}
\begin{figure}
\label{pitchplot}
\special{psfile=pitchplot.eps
hscale=40 vscale=40 hoffset=70 voffset=-270 angle=0}
\vspace*{7cm}
\caption{NFW concentration parameter versus spiral arm pitch angle, showing a correlation. The green point represents data for M31, the cyan point for Malin 1, the red point for IC2522, the blue point for ESO582G12 and the magenta point shows the data for M33.}
\end{figure}
Finally Figure 5 shows a plot of supermassive black hole mass as a function of
NFW concentration parameter. Unfortunately, here we only have data for three
galaxies. Nevertheless, a hint of a correlation is starting to show, and
seeing that such a correlation has been suggested by Seigar et al.\ (2008b),
as well as Satyapal et al.\ (2008) and Booth \& Schaye (2010), this plot is
somewhat intriguing. This hint of a correlation should, of course, be
expanded on by studying more galaxies along the Hubble sequence from type Sa
to Sd.
\begin{figure}
\label{bhplot}
\special{psfile=bhplot.eps
hscale=40 vscale=40 hoffset=70 voffset=-270 angle=0}
\vspace*{7cm}
\caption{Central supermassive black hole mass versus NFW concentration parameter, showing a correlation. The green point represents data for M31, the blue point for the Milky Way and the red point shows the data for M33.}
\end{figure}
\section{Summary}
We have shown that the H{\tt I} rotation curve of M33 can be best modeled with
a dark matter halo that follows a NFW profile, with low NFW concentration of
$c_{\rm vir}=4.0$. Using the NFW concentration parameter from this fit, we
find that interesting correlations between (1) spiral arm pitch angle and NFW
concentration and (2) central supermassive black hole mass and NFW
concentration, start to appear. Although the second correlation is only for
three galaxies, on the surface it appears to be in disagreement with the
argument made by Kormendy \& Bender (2011) that the dark matter halos of
galaxies have no affect on the masses of supermassive black holes found in
their centers. These correlations are very intriguing and our results warrant
further investigation, as we have been limited to data that was available for
just a few galaxies.
\section{Acknowledgements}
This research has made use of the NASA/ IPAC Infrared Science Archive, which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. The authors wish to thank the reviewers who helped to improve the content of this paper.
|
\section{List of figure captions}
\begin{enumerate}
\item Experimental image acquisition setup.
\item Algorithmic layout of holographic rendering.
\item Time-averaged holographic maps of the $\omega_{\rm M} / (2 \pi) = 10 \, \rm kHz$ vibrational mode of the paper sheet for the modulation sidebands of order $n=0$ (a), $n=1$ (b), $n=3$ (c), and $n=7$ (d). $1024 \times 1024$ pixels renderings.
\end{enumerate}
]
\end{document}
|
\section{Introduction}
The most important event in the formation and evolution of most close interacting binaries is the
so-called common envelope (CE) event, during which the components of a binary system
are engulfed by a common gaseous envelope, and the resulting interaction dramatically
shrinks their orbit \citep{BKS,Os76,Pa76}.
Depending on the envelope structure and companion masses, the envelope is ejected to leave behind a close binary,
or the two stars merge.
Dividing the parameter space into binaries that survive CE and those that do not,
and determining the final separation for the former, is critical for calculating formation rates of low mass X-ray binaries (LMXBs),
$\gamma$-ray bursts, as well as for LISA and LIGO events \citep{Bel08}.
In the standard treatment of CE outcomes,
the final separation of the binary is determined via ``energy formalism'' \citep{Webbink84,Livio88}, in which the binding energy of
the (shunned) envelope is equated to the decrease in the orbital energy $E_{\rm orb}$:
\begin{equation}
E_{\rm bind} = E_{\rm orb,i} - E_{\rm orb,f} = -\frac{ G m_1 m_2} {2 a_{\rm i}} + \frac{ G m_{1\rm,c} m_2} {2 a_{\rm f}}
\end{equation}
Here $a_{\rm i}$ and $ a_{\rm f}$ are the initial and final binary separations, $m_1$ and $m_2$ are the initial star masses
and $ m_{1\rm,c}$ is the final mass of the star that lost its envelope.
$E_{\rm bind}$ is assumed to be the energy expense needed to remove the envelope to infinity
and is commonly adopted to be the sum of the potential energy of the envelope and its internal energy.
To characterise the donor envelope central concentration and simplify calculations, specifically for population synthesis,
a parameter $\lambda$ was introduced:
\begin{equation}
E_{\lambda, \rm bind} = - \int_{\rm core}^{\rm surface} \left ( \Psi(m) + \epsilon (m) \right) dm = \frac {G m_1 m_{1,\rm e}} {\lambda R_1}
\label{lambda}
\end{equation}
Here $m_{1,\rm e}$ is the mass of the removed giant envelope and $R_1$ is the radius of the giant star at the onset of CE,
$\epsilon$ is the specific internal energy and $\Psi(m)=-Gm/r$ is the (gravitational) force potential (or specific potential energy).
$E_{\lambda, \rm bind}$ can be found directly from stellar structure for any accepted core mass.
Another parameter, $\alpha_{\rm CE}$, is introduced as a measure of the energy transfer efficiency from the orbital energy
into envelope expansion, and the balance of energy is written as
\begin{equation}
\alpha_{\rm CE} {\lambda} \left ( \frac{ G m_{1\rm,c} m_2} {2 a_{\rm f}} -\frac{ G m_1 m_2} {2 a_{\rm i}} \right ) =
\frac {G m_1 m_{1,\rm e}} {R_1}
\label{allam}
\end{equation}
Many authors choose to accept $\alpha_{\rm CE} \lambda \approx 1$, since for many (at least low-mass) stars $\lambda=1$,
as can be found from detailed stellar structure, and $\alpha_{\rm CE}$ is bounded above by 1.
The simplicity of the standard prescription resulted in its popular use
in the binary population synthesis calculations. However, once
accurate values of $\lambda$ from stellar structure calculations were determined,
this approach has shown inconsistencies with the observations, especially large for the formation of
black hole LMXBs: in massive giants $\lambda \ll 0.1$ \citep{podsi03}, and it has been shown that
with $\alpha_{\rm CE}\le 1$ only an intermediate-mass companion could avoid
a merger; this challenges the formation of a low-mass X-ray binary with a low-mass companion in general \citep{Justham06}.
Here, we revise the energy requirements necessary to force a common envelope to disperse
and discuss its possible application for LMXBs formation.
\section{Enthalpy considerations}\label{enthalpy}
\subsection{Total energy and instability}
While the definition of the gravitational binding energy is unequivocal, \citep[e.g.,][]{Chandra39},
there exists no authoritative source defining what expression should be used as the binding energy
of a gaseous sphere in the sense of \S~1. To clarify, in treating the common envelope problem we are interested
in the additional energy to be deposited in the envelope in order to disperse material to infinity.
The `industry standard' at the moment is to use the total of the gravitational $U$ and the gas thermal $E$
energies, which, as we shall argue in this section, is not correct.
As an illustrative limiting case, consider a star with the positive total energy\footnote{
with zero of energy defined for all material evacuated to infinity, and pressure set to zero
}
$W_0|_{t=0}>0$ to begin with, which is {\it ``kinetically stable''};
i.e. an energy barrier has to be overcome between the bound and unbound states \citep{BKZ67}.
It is clear that in such cases, the additional energy $\Delta$
required to unbind the star is the magnitude of the energy barrier, not the (unphysical) $-W_0<0$.
The secular stability of a star against small adiabatic perturbations
is not defined by the sign of the total energy $W$, but rather by the variational conditions:
an equilibrium state is an extremum of the total energy $(\delta W)_{\rm ad} =0$; and
secularly stable configurations are then found at local minima: $(\delta^2 W)_{\rm ad} > 0$.
\citep{Chiu68}.
If the first adiabatic exponent $\Gamma_1$ is approximately a constant throughout the star,
the secular stability criterion reduces to the condition \citep{Chiu68}
\begin{equation}{
\Gamma_1 = \left ( \frac {\partial \ln P}{\partial \ln \rho } \right )_{\rm ad} > 4/3
}\end{equation}
Here $P$ is the pressure and $\rho$ is the density.
For a one-zone model of a stellar envelope, a linearized version of the above condition,
the Baker's model \citep{Baker66} gives a similar criterion for the volume-averaged
adiabatic exponent $\Gamma_1 > 4/3$ in the envelope
On the other hand, the virial theorem, applied to the \emph{entire} star,
gives us the condition for the total energy $W<0$
in the case of constant third adiabatic exponent $\Gamma_3$,
as \citep[e.g.,][]{HKT04}:
\begin{equation}{
\Gamma_3:=\left ( \frac{\partial \ln T}{\partial \ln \rho }\right )_{\rm ad} +1 > 4/3 \ .
}\end{equation}
Here $T$ is is the temperature.
We note that this simple form depends crucially on Newton's third law applied to every pair of particles inside the star,
and thus needs to be substantially modified if we only consider the envelope.
We reiterate the importance of distinguishing these conditions:
instability does not have to occur in a state with positive total energy,
and $\Gamma_1$ does not have to coincide with $\Gamma_3$;
they are related as
\begin{equation}{
\Gamma_3=1 + \Gamma_1 \left ( \frac{\partial \ln T}{\partial \ln P }\right )_{\rm ad}
}\end{equation}
In particular, the ionization zones, where $
\nabla_{\rm ad} = \left ( \frac{\partial \ln T}{\partial \ln P }\right )_{\rm ad} < 0.4
$
and can become as low as 0.1, have local value of $\Gamma_3<4/3$ while $\Gamma_1>4/3$.
It is unfortunately non-trivial to find the volume-averaged $\Gamma_1$ in the giant envelope
after depositing some heat in it.
However, it is clear that the criterion of the total energy in the envelope to be $W=0$
is not applicable for the envelope stability, nor is it a sufficient condition
for the envelope to be dispersed to infinity. We will consider instead
the stability of a stellar envelope towards creating outflows.
\subsection{The condition for the envelope to outflow}
Since we are interested in the energy {\it requirements}, it is natural to consider
a lower bound on $\Delta$: all orbital energy is converted into heat, and the velocity
of ejecta at infinity is zero.
How precise these bounds are is not of great concern here and is a separate problem on
its own, although we expect them to be reasonably
close: the time scale for viscous friction between differentially rotating regions can exceed
the spiral-in time, but most angular momentum is lost in the outer envelope while most of the energy
is released at closer orbits, so the overall effect of angular momentum conservation can at
least be constructed to be small.
As any other thermodynamical system undergoing a steady process, the material in a star obeys
the first law of thermodynamics, which states that the change of the internal energy comes not
only from the heat transferred into system, but also from the work done by the system:
\begin{equation}
\delta E = \delta Q - P \delta \left ( \frac{1}{\rho} \right ) \ .
\end{equation}
Applying this equation for each mass shell in the envelope,
(see, e.g., \citealt{CW66, KS72}),
gives the energy conservation equation in Lagrangian coordinates as
\begin{equation}
\frac{\partial}{\partial t} \left(\frac{u^2}{2} + \Psi +\epsilon\right ) +
{P}\frac {\partial (1 / \rho) }{\partial t} = 0 \ .
\end{equation}
Here $u$ is the velocity.
The initial condition is that at the start the mass shells in the envelope are not moving,
but every shell in the envelope received some heat, in this case deposited from the companion's orbital contraction.
The heat has some arbitrary distribution ${\delta q(m)}>0$ in the envelope (${\delta q(m)}$ is per mass unit), such that
\begin{equation}
\int_{\rm core}^{\rm surface} \delta q(m) dm = Q.
\end{equation}
A given lagrangian shell, once it has been heated and started expansion,
will have reached the point of no return in its expansion when its $\Sigma>0$
\citep[e.g.,][]{BKZ67,SK72}
\begin{equation}
(\delta q(m) + \Psi + \epsilon + \frac{P}{\rho})_{\rm start}
= (\frac{1}{2} u^2 + \Psi + \epsilon + \frac{P}{\rho})_{\rm exp}
= \Sigma = const
\label{bern}
\end{equation}
In this form, the equation can be also recognized as version of the Bernoulli equation.
The quantity $\Sigma$, for stability analysis purposes, is more important than the total energy:
bipolitropic stellar model that has $\Sigma < 0$ in its envelope but positive total energy
is metastable as a whole, even though the run-away to infinity is energetically allowable;
but a star with $\Sigma>0$ in its envelope will be always quasi-steadily outflowing \citep{BKZ67,BK02}.
It also has been noted that it is a general feature of low-mass giants during double shell
burning to establish $\Sigma>0$
in larger and large parts of their envelopes as they evolve (with He shell providing large energy
inflow to the envelope with each He-shell flash),
and could possibly being responsible for envelope outflows and/or ejections with planetary nebula
formation \citep{SK72}.
In line with results of past stability analyses of when a star starts an outflow \citep[][and the references therein]{BK02}, we
suggest that once a \emph{part} of the common envelope has obtained positive $\Sigma$, it will start outflowing,
notwithstanding the sign of the envelope's total energy.
A stronger criterion would be to require the \emph{entire} envelope to have positive $\Sigma$.
Under assumption of the minimum energy requirement (velocity of gas at infinity is zero),
we can write the energy conservation for a whole envelope as
\begin{equation}
Q + \int_{\rm core}^{\rm surface}\left (\Psi(m) + \epsilon(m) + \frac{P(m)}{\rho(m)} \right ) dm = 0 \ .
\end{equation}
The timescale for start such quasi-steady surface outflows is one on which the envelope redistributes the dumped heat,
i.e. the thermal timescale of the envelope - recall our earlier note about most energy being deposited in lower orbits.
This time is about few hundred years \citep[e.g, it is about 1000 years for a $20 M_\odot$, see Fig.1 in ][]{Ivanova11}.
This makes the appearance of mass outflows {\it natural} during the self-regulating spiral-in stage;
such a stage could last for up to a thousand years \citep{podsi01};
but they likely will not take place if the common envelope event occurs on the dynamical time, as is the case
e.g., in the case of a physical collision, or if the swallowed companion is too small to expand the giant envelope
and establish a self-regulated slow spiral-in phase.
Note, that
\begin{itemize}
\item due to the presence of a non-negative term $P/\rho$, this condition occurs before
the envelope's total energy become positive;
\item the master equation (\ref{allam}) of the `$\lambda$-formalism` is a version of eq.(\ref{bern})
where all the velocities, as well as the work $P/\rho$ are simply neglected.
\end{itemize}
\noindent The term $P/\rho$ is of the order of magnitude of $\epsilon$;
for example, for an ideal gas, with no radiation pressure and ionization taken into account,
$P/\rho = {2}/{3}\, \epsilon$. The quantity $h=\epsilon + {P}/{\rho}$ is generally known as {\it enthalpy}.
In line with the classical `$\lambda$-formalism' and for ease of comparison, we introduce:
\begin{equation}
E_{h, \rm bind} = - \int_{\rm core}^{\rm surface} \left ( \Psi(m)+ h(m) \right )dm = \frac {G m_1 m_{1,\rm e}} {\lambda_{\rm h} R_1}
\end{equation}
\begin{figure}[t]
\includegraphics[height=.35\textheight]{fig1.eps}
\caption{$\Psi$ and $h$ inside $2\ M_\odot$ giant. The ``boiling pot'' zone has excess energy. Energy is the specific energy in erg per g.
\label{bpz}
}
\end{figure}
Following \cite{SK72} in considering the detailed stellar models of giants with respect to their $\Sigma$, we also
note that normal single giants have a "boiling pot'' zone (BPZ), where $\Sigma(m)=\Psi(m)+h(m) > 0$
due to high value of $P/\rho$ (Fig.~\ref{bpz}).
Once sufficient amount of the ``lid'' -- the star matter above this zone is removed,
all material with positive $\Sigma$ could freely stream away without the need to convert any additional
mechanical energy.
The outer boundary of the BPZ almost does not
change during the giant stage, and the inner boundary does
not change once the convective envelope is established (e.g.,
in case of low-mass red giants, RGs). The bottom of the BPZ
roughly (but not always exactly) coincides with the bottom of
the outer convective zone. Illustratively, the mass contained
in this zone is: $\sim 0.5 M_\odot$ for $2 M_\odot$ RG, up to $7.7 M_\odot$ for a $30 M_\sun$
giant evolved without wind mass loss, and about 3 times less for a $30 M_\sun$ giant evolved with mass loss.
We anticipate that the presence of the boiling pot zone could be a key to understand why RGs
are expanding when losing their mass.
\section{Comparison of $E_{\rm \lambda, bind}$ and $E_{\rm h, bind}$ in stellar models}
Just like $\lambda$, $\lambda_{\rm h}$ can be found from detailed stellar calculations.
For the purpose of this comparison,
we adopt the definition of the core mass for both cases as in \cite{Ivanova11}:
the post-CE core is the point with maximum compression $P/\rho$ inside the Hydrogen burning shell, $m_{1, \rm cp}$.
Such a core will not re-expand after the envelope loss, and it was shown that in most of the cases
it is energetically beneficial to remove the envelope down to exactly $m_{1, \rm cp}$.
In Table~1 we show the results for giants of different masses.
We used the set of stellar models described in \cite{Ivanova11}.
We find that in most giants the ratios of $\lambda_{\rm h}/\lambda$ are from $\sim2$ to $\sim5$, where
the largest ratios, for a giant of the same mass, are for an earlier giant.
The final separations allowed by the enthalpy-consideration are then larger by the same few times.
The minimal companion mass -- in the sense that no possible stripped core
value exists such that a binary would not merge -- changes by several times as well (Table~1).
Specifically, we want to point out the difference between the enthalpy consideration and standard prescription
for massive giants: while $\alpha_{\rm CE}\lambda$-formalism would predict the minimum surviving companion mass
of about several $M_\odot$, in line with the problem raised by \cite{Justham06}, the
enthalpy-formalism allows for an low-mass companion to survive and form in a future a black hole LMXB.
\begin{table}
\caption{CE outcomes comparison. }
\begin{tabular}{l r r r r r r r }
\hline
$m_{1} (m_{\rm zams})$ &$R_{1}$ & $m_{1,X}$ & $m_{1,\rm cp}$ & $\lambda$ & $\lambda_{\rm h}$ & $m_{2,\lambda}$ & $m_{2,h}$ \\
\hline
25.59(30) & 900 & 9.381 & 11.44 & 0.026 & 0.085 & 6.33 & 1.53 \\
25.53(30) & 1500 & 10.223 & 11.39 & 0.026 & 0.064 & 2.59 & 0.92 \\
18.5(20) & 600 & 5.59 & 6.48 & 0.065 & 0.299 & 2.84 & 0.46 \\
18.5(20) & 750 & 5.70 & 6.48 & 0.133 & 0.309 & 0.82 & 0.32 \\
16.8(20) & 850 & 6.75 & 6.92 & 0.067 & 0.142 & 0.72 & 0.31 \\
9.75(10) & 200 & 1.69 & 1.95 & 0.148 & 0.274 & 1.87 & 0.86 \\
9.75(10) & 300 & 1.73 & 2.04 & 0.136 & 0.244 & 1.28 & 0.62 \\
9.74(10) & 360 & 1.95 & 2.10 & 0.143 & 0.253 & 0.74 & 0.37 \\
5.09(10) & 380 & 2.87 & 2.94 & 0.061 & 0.109 & 0.16 & 0.09 \\
4.99(5) & 40 & 0.575 & 0.725 & 0.402 & 0.815 & 1.9 & 0.75 \\
4.99(5) & 80 & 0.702 & 0.784 & 0.425 & 0.822 & 0.56 & 0.25 \\
2 & 10 & 0.253 & 0.271 & 1.167 & 2.804 & 0.39 & 0.13 \\
2 & 40 & 0.526 & 0.529 & 0.730 & 1.652 & 0.04 & 0.02 \\
1 & 10 & 0.253 & 0.254 & 0.941 & 2.29 & 0.04 & 0.02 \\
\hline
\end{tabular}
\label{table}
\footnotesize{A post-CE mass is adopted to be the divergence point $m_{1,\rm cp}$, for a comparison is shown the mass of
the hydrogen-exhausted core $m_{1,\rm X}$ (where $X\le 10^{-10}$).
$\lambda$ and $\lambda_{h}$ connect the energies required to eject the envelope with their parametrizations in two formulations (eqs.~2 and 12).
$m_{2,\lambda}$ and $m_{2,h}$ are the {\it minimum} companions' masses
that could survive a CE event, where $m_{2,\lambda}$ is using standard prescription and
$m_{2,h}$ is with enthalpy-consideration.
$m_1$ are the current donor masses and $m_{\rm zams}$ are donor masses at the zero-age main sequence (if different from $m_1$).
$R_1$ are the current donor radii.
All masses are in $M_\odot$, $R_{1}$ is in $R_\odot$.
}
\end{table}
\section{Conclusions}
In this {\it Letter} we considered the termination of a common envelope event at the moment it establishes
a quasi-stationary mass outflow that would reach a point of no return.
Such outflow develops during the slow self-regulating spiral-in phase of the common envelope event.
We showed that if the common envelope is escaping the binary in this way,
neglecting the ${P}/{\rho}$ term in the standard $\alpha_{\rm CE}\lambda$ energy
conservation prescription is too gross.
If the enthalpy rather than internal energy
is calculated in the energy balance, it makes a crucial difference
in giants.
While we anticipate that the estimate of a proper energy requirement is paramount
for the formation rates of all kinds of post-CE binaries,
we especially emphasize the importance of this effect in the case of CE with a massive giant.
When the standard energy formalism predicts that only a several solar mass donor could survive CE,
and so the formation of a low-mass X-ray binary with a black hole accretor is forbidden unless
$\alpha_{\rm CE}$ exceeds 1,
the enthalpy-consideration naturally allows for a low-mass companion survival.
\section{Acknowledgment}
Natalia Ivanova acknowledges support from NSERC and Canada Research Chairs Program.
We are thankful to Jonathan Braithwaite for spotting a misprint in equation (8) of the version 1 of this paper.
|
\section{Introduction}
\label{sec31}
The purpose of this paper is to prove an inductive characterisation of simple $(2,1)$-tight graphs.
\begin{defn}[Lee and Streinu \cite{L&S}]
Let $k,\ell \in \bN$ and $\ell \leq 2k$. A graph $G=(V,E)$ is {\rm $(k,\ell)$-sparse} if for every subgraph $G'=(V',E')$,
$|E'|\leq k|V'|-\ell$ (where if $\ell=2k$ the inequality only applies if $|V'|\geq k$). $G$ is {\rm $(k,\ell)$-tight} if $G$ is
$(k,\ell)$-sparse and $|E|=k|V|-\ell$.
\end{defn}
In our notation a graph allows parallel edges and loops, whereas a simple graph allows neither.
The classes of $(2,\ell)$-tight simple graphs play an important role in the theory of $2$-dimensional bar-joint frameworks (see, for example,
\cite{A&R} and
\cite{GSS} for the general theory). When $l=3$ these graphs correspond to generic frameworks that are minimally rigid when joints corresponding to the
vertices are constrained to lie on a plane (since any framework on a plane has three independent rigid-body motions) \cite{Lam}. When $l=2$ these graphs correspond to generic frameworks
which are minimally rigid when the joints are constrained to lie on the surface of a cylinder (since this surface allows two independent rigid-body motions) \cite{NOP}. When $l=1$
we expect the graphs to correspond to frameworks that are rigid when the joints are constrained to a surface which admits one independent rigid-body motion. These surfaces include
linearly swept surfaces (such as an elliptical cylinder or any ruled surface with parallel rulings) and spun surfaces (such as a circular cone, torus or any
surface formed by rotating a smooth curve). These surfaces are important in engineering since they are easily manufactured using the processes of extrusion and turning.
The characterisation of generic framework rigidity typically involves two distinct steps - an inductive construction of the relevant class of graphs and then a proof
that the construction steps preserve the required rigidity properties.
The classical result of Henneberg \cite{Hen} characterises the class of $(2,3)$-tight graphs by recursive
operations. Combining this
with a result of Lovasz and Yemini \cite{L&Y}, extended by Recski \cite{Rec}, leads to:
\begin{thm}\label{thm1}[Henneberg \cite{Hen}, Lovasz and Yemini \cite{L&Y}, Recski \cite{Rec}]
For a graph $G=(V,E)$ the following are equivalent:
\begin{enumerate}
\item $G$ is $(2,3)$-tight,
\item $G$ is derivable from $K_{2}$ by the Henneberg 1 and Henneberg 2 moves,
\item for any edge $e\in E(K_{|V(G)|})$, $G\cup \{e\}$ is the edge-disjoint union of two spanning trees.
\end{enumerate}
\end{thm}
Laman \cite{Lam} then characterised generic minimal rigidity on the plane by showing that the Henneberg 1 and Henneberg 2 moves preserve this property on the plane.
Nixon, Owen and Power \cite{NOP} obtained a characterisation of simple $(2,2)$-tight graphs, Theorem \ref{thm2}.
\begin{thm}\label{thm2}[Nixon, Owen and Power \cite{NOP}, Nash-Williams \cite{N-W}]
For a simple graph $G=(V,E)$ the following are equivalent:
\begin{enumerate}
\item $G$ is $(2,2)$-tight,
\item $G$ is derivable from $K_{4}$ by the Henneberg 1, Henneberg 2 and graph extension moves,
\item $G$ is the edge-disjoint union of two spanning trees.
\end{enumerate}
\end{thm}
In this characterisation a graph extension move replaces a vertex in the graph by an arbitrary $(2,2)$-tight graph which thereby becomes a $(2,2)$-tight
subgraph in the extended graph. \cite{NOP} also characterised generic minimal rigidity on the cylinder by showing that the Henneberg 1, Henneberg 2 and graph extension moves preserve this property on the cylinder.
Our main result is the following inductive construction of $(2,1)$-tight simple graphs.
By $K_{5}\setminus e$ we mean the graph formed from the complete graph on $5$ vertices by removing an edge, and by
$K_{4}\sqcup K_{4}$ we mean the graph formed by taking two copies of $K_{4}$ that intersect in a copy of the complete graph $K_2$. The construction operations are defined at the start of Section \ref{tightchar}.
\begin{thm}\label{21theorem}
For a simple graph $G$ the following are equivalent:
\begin{enumerate}
\item $G$ is $(2,1)$-tight,
\item $G$ can be derived from $K_{5}\setminus e$ or $K_{4}\sqcup K_{4}$ by the Henneberg 1, Henneberg 2, vertex-to-$K_4$, edge joining and edge-to-$K_3$ moves,
\item $G$ is the edge-disjoint union of a spanning tree $T$ and a spanning subgraph $P$ in which every connected component contains exactly one cycle.
\end{enumerate}
\end{thm}
We expect that each of the construction moves in (2) of this theorem also preserves minimal generic rigidity on surfaces which admit one rigid body motion. We present this as a conjecture for subsequent investigation.
As a by-product of our arguments we also show the following result giving an alternative inductive construction of $(2,2)$-tight graphs. The construction should be easier to apply since we only insert prescribed small subgraphs rather than an arbitrary graph in the class.
\begin{thm}\label{22refine}
For a simple graph $G=(V,E)$ the following are equivalent:
\begin{enumerate}
\item $G$ is $(2,2)$-tight,
\item $G$ is derivable from $K_{4}$ by the Henneberg 1, Henneberg 2, vertex-to-$K_4$ and edge-to-$K_3$ moves.
\end{enumerate}
\end{thm}
The main difficulty in proving theorem \ref{21theorem} is the requirement that the inductive construction involves only simple graphs. This requirement arises because we are
interested in frameworks in which the distance between a pair of joints is the usual distance measured as a straight line in $3$-space. Minimal rigidity then clearly requires that
two vertices are joined by at most edge. Whitely \cite{Whi5} has considered frameworks embedded on surfaces in which the distance between a pair of joints is a geodesic distance over the surface. In this case a pair of vertices may be separated by more than one distinct geodesic distance and the class of graphs may be extended to include multiple edges between a pair of vertices. Similarly periodic frameworks \cite{B&S}, \cite{M&T}, \cite{Ros} on the plane may include edges connecting between different cells and result in graphs with multiple edges.
We note that for the case of $(k,\ell)$-tight graphs (permitting parallel edges and loops) there are elegant recursive constructions requiring Henneberg type operations only \cite{Fr&S}, \cite{F&S}.
A further motivation for our work is the hope that understanding the recursive constructions for $(2,\ell)$-tight graphs of the
various types will provide insight into $(3,6)$-tight graphs.
These are the graphs relevant to major open problems in $3$-dimensional rigidity theory \cite{GSS}, \cite{T&W}, \cite{Whi4}.
Note that these graphs are necessarily simple and are outside the matroidal range.
Indeed for $\ell<6$ adding \emph{any} $\ell-3$ edges to a $(3,6)$-tight graph results in a graph with a decomposition into three edge
disjoint spanning trees but for $l=6$ it does not, see \cite{Haa}.
From our main theorems one can quickly derive sparsity variants. That is, characterisations of $(2,\ell)$-sparsity in terms of
recursive operations. If Conjecture \ref{tjcon} is true then this has applications in computer aided design \cite{Owen} where the
emphasis is on establishing whether a system of constraint equations admits a matrix with linearly independent rows.
The paper is organised as follows. Section \ref{tightchar} defines the recursive moves we will consider. The key difficulty is the
construction theory of Section \ref{sec2*}, in which we discuss the sufficiency of the moves. The main step is Lemma \ref{jco4}.
Here we take a seemingly mild requirement that each edge in a copy of $K_3$ is in at least two copies of $K_3$ or is in a separate
$(2,1)$-tight subgraph. This leads to the strong conclusion that every copy of $K_3$ is in a copy of $K_4$. This convenient property
is used to derive the key implication in the proofs of Theorems \ref{21theorem} and \ref{22refine}. Finally Section \ref{21apps}
discusses rigidity theory and potential applications of our results therein.
We would like to thank Stephen Power for some helpful discussions and the anonymous referees for a number of helpful comments.
\section{Simple $(2,\ell)$-tight Graphs}
\label{tightchar}
It will be convenient for us to define $f(H):=2|V(H)|-|E(H)|$ for a graph $H$.
\begin{defn}
Let $\ell=1,2,3$. A simple graph $G$ is \emph{$(2,\ell)$-sparse} if $f(H) \geq \ell$ for all subgraphs $H$ of $G$ with at least one edge and is \emph{$(2,\ell)$-tight} if it is $(2,\ell)$-sparse and $f(G)=\ell$.
\end{defn}
We begin by recalling and formally defining the construction moves under consideration. Define the \emph{Henneberg $0$} move to be the addition of a vertex of degree $0$ or of degree $1$ to a graph. The inverse Henneberg $0$ move is the removal of a vertex of degree $0$ or degree $1$ from a graph.
The \emph{Henneberg $1$ move} \cite{Hen}, is the addition of a degree $2$ vertex to a graph. The inverse Henneberg $1$ move is the removal of a degree $2$ vertex from a graph.
The \emph{Henneberg $2$ move} \cite{Hen}, removes an edge $uv$ and adds a vertex $x$ and edges $xu,xv,xw$ for some vertex $w$. The inverse Henneberg $2$ move removes a degree $3$ vertex $x$ (and incident edges $xu,xv,xw$) and adds an edge $uv, uw$ or $vw$.
Let $G$ be $(2,1)$-sparse containing a copy of $K_4$. Write $G / K_4$ for the (possibly multi)graph formed by contracting this copy of $K_4$ to a vertex $v_*$. That is $G / K_4$ has vertex set $(V(G) \setminus V(K_4)) \cup \{v_{*}\}$ and edge set $(E(G) \setminus E(K_4)) \cup E_{*}$ where $E_{*}$ consists of the edges $vv_{*}$ associated with edges $vw$ where $v \in G / K_4$ and $w \in K_4$. We call this operation a \emph{$K_4$-to-vertex move}. The inverse move, he \emph{vertex-to-$K_4$ move} is illustrated in Figure \ref{fig:vtok4}.
The graph extension move mentioned in the introduction refers to the construction of $G$ from $G/H$ where $H$ is a proper induced $(2,2)$-tight subgraph of $G$. This move was used in \cite{NOP} and is similar to
vertex expansion moves used in graph theory, \cite{Die}.
\begin{center}
\begin{figure}[ht]
\centering
\includegraphics[width=5cm]{ext.eps}
\caption{With $H=K_4$, an example of the vertex-to-$K_4$ move and, with $H$ a proper induced $(2,2)$-tight subgraph of $G$, graph extension.}
\label{fig:vtok4}
\end{figure}
\end{center}
Let $G$ be a graph with an edge $uv$ such that the neighbours of $v$ are $a_1,\dots,a_n$. The \emph{edge-to-$K_3$ move}, see Figure \ref{Vertex splitting*}, (often referred to as vertex splitting in the literature, \cite{Whi6}) removes the edge $uv$ and the vertex $v$ and all the edges $va_i$, it replaces them with the vertices $v_1,v_2$ and edges $uv_1,uv_2,v_1v_2$, plus some bipartition of the remaining edges $v_1a_j$ and $v_2a_k$ (with one side possibly empty). The inverse move, called the \emph{$K_3$-to-edge move}, takes a copy of $K_3$ (with vertices $u,v_1,v_2$), removes the edges $uv_1,uv_2,v_1,v_2$, merges two vertices $v_1,v_2$ into a single vertex $v$ which is adjacent to all the vertices $v_1$ and $v_2$ were adjacent to and adds the edge $uv$.
\begin{center}
\begin{figure}[ht]
\centering
\includegraphics[width=6cm]{split.eps}
\caption{The edge-to-$K_3$ move.}
\label{Vertex splitting*}
\end{figure}
\end{center}
Let $K$ and $H$ be $(2,1)$-tight graphs with vertices $u \in K$ and $v \in H$. We will say that $G$ is formed from $K$ and $H$ by an \emph{edge joining move}, see Figure \ref{Join} if $V(G)=V(K)\cup V(H)$ and $E(G)=E(K)\cup E(H) \cup \{uv\}$. Further, if there is a single edge $uv$ joining two $(2,1)$-tight subgraphs $G$ and $H$, then we will call the inverse move an \emph{edge separation move}.
\begin{center}
\begin{figure}[ht]
\centering
\includegraphics[width=8cm]{join.eps}
\caption{The edge joining move.}
\label{Join}
\end{figure}
\end{center}
With respect to Theorem \ref{22refine} note that Figure \ref{edgetok3nec} illustrates the necessity of the $K_3$-to-edge move when we restrict graph contraction to the $K_4$-to-vertex move.
We note that $(1) \Leftrightarrow (3)$ in Theorem \ref{21theorem} can be proven in an elementary way by showing that the construction operations preserve the spanning subgraph decomposition. More efficiently, these implications follow from matroidal results; the $(1,1)$-tight graphs are the bases of the cycle matroid and the $(1,0)$-tight graphs are the bases of the bicycle matroid. The union (on the same ground set of vertices) of a cycle matroid and a bicycle matroid (with empty intersection) give the results, see \cite{Fr&S}, \cite{G&W}, \cite{N-W}, \cite{Whi5}.
\begin{center}
\begin{figure}[ht]
\centering
\includegraphics[width=5cm]{edgetok3nec.eps}
\caption{A $(2,2)$-tight graph that requires the $K_3$-to-edge move when we restrict the graph contraction move to the $K_4$-to-vertex move.}
\label{edgetok3nec}
\end{figure}
\end{center}
\section{Construction Theory}
\label{sec2*}
In this section we consider $(1) \Rightarrow (2)$ in Theorems \ref{21theorem} and \ref{22refine}. That is, we consider whether an arbitrary $(2,2)$ or $(2,1)$-tight graph can be reduced by applying one of our short list of moves (relevant to each case) to produce a smaller $(2,2)$-tight or $(2,1)$-tight graph.
We begin by showing that in a $(2,1)$-tight or $(2,2)$-tight graph, an inverse Henneberg 2 move is available unless all degree 3 vertices are in copies of $K_{4}$.
\begin{lem}\label{k4lem}
Let $G$ be $(2,\ell)$-tight for $\ell=1,2$ with a vertex $v\in V(G)$ of degree $3$ with neighbours $v_{1},v_{2},v_{3}$ in $G$. Then either $v$ is contained in a copy of
$K_{4}$ or $G'=(G\setminus v) \cup e$ (for $e=v_1v_2,v_2v_3$ or $v_3v_1$) is $(2,1)$-tight.
\end{lem}
\begin{proof}
With suitable labeling of vertices, we distinguish three cases corresponding to the possible edges among the neighbours of $v$. Either
\begin{enumerate}
\item $v_1v_2,v_1v_3,v_2v_3 \in E$,
\item $v_1v_2 \notin E,v_1v_3,v_2v_3 \in E$, or
\item $v_1v_2,v_2v_3 \notin E$.
\end{enumerate}
In case 1, $v,v_1,v_2,v_3$ induce a copy of $K_4$ in $G$.
\begin{center}
\begin{figure}[ht]
\centering
\begin{tikzpicture}
\filldraw[black]
(0,0) circle (3pt)node[anchor=north]{$v_3$}
(0,1) circle (3pt)node[anchor=south]{$v$}
(-1.5,2) circle (3pt)node[anchor=south]{$v_1$}
(1.5,2) circle (3pt)node[anchor=west]{$v_2$}
(6,0) circle (3pt)node[anchor=north]{$v_3$}
(6,1) circle (3pt)node[anchor=south]{$v$}
(4.5,2) circle (3pt)node[anchor=south]{$v_1$}
(7.5,2) circle (3pt)node[anchor=west]{$v_2$};
\draw[black,thick]
(0,0) -- (0,1) -- (-1.5,2) -- (0,0);
\draw[black,thick]
(0,1) -- (1.5,2) -- (0,0);
\draw[black,thick]
(6,0) -- (6,1) -- (4.5,2) -- (6,0);
\draw[black,thick]
(6,1) -- (7.5,2);
\draw[black] (0.2,2) ellipse (2.2 and 0.4)node{$Y_{12}$};
\draw[black] (6.2,2) ellipse (2.2 and 0.4)node{$Y_{12}$};
\draw[black, rotate around={54:(6.9,1)}] (6.9,1) ellipse (2 and 0.4)node{$Y_{23}$};
\end{tikzpicture}
\caption{The graph on the left illustrates case (2): if there was a subgraph $Y_{12}$ preventing the application on an inverse Henneberg 2 move on $v$ then the graph pictured would be over-counted. Similarly the graph on the right illustrates case (3): if there are subgraphs preventing the addition of $v_1v_2$ and $v_2v_3$ then the graph pictured would be over-counted. }
\label{fig:21h2proof}
\end{figure}
\end{center}
Figure \ref{fig:21h2proof} illustrates the proof in cases $2,3$. Define $Y_{12}$ to be a $(2,\ell)$-tight subgraph of $G$ containing $v_1,v_2$ but not $v_3,v$. Similarly define $Y_{13}$ and $Y_{23}$.
In case 2, $G'=(V\setminus v, (E\setminus \{vv_1,vv_2,vv_3\})\cup v_1v_2)$ is $(2,\ell)$-tight unless there exists a subgraph $Y_{12}$
of $G$. But then the addition of $v,v_3$ and their five incident edges to $Y_{12}$ gives a subgraph $Y$ of $G$ with $f(Y)=\ell-1$ which contradicts the fact that $G$ is $(2,\ell)$-tight.
In case 3, either $G'=(V\setminus v, (E\setminus \{vv_1,vv_2,vv_3\})\cup v_1v_2)$ or
$G'=(V\setminus v, (E\setminus \{vv_1,vv_2,vv_3\})\cup v_2v_3)$ is $(2,\ell)$-tight unless there exists subgraphs $Y_{12}$ and $Y_{23}$
of $G$. Then
\[f(Y_{12}\cup Y_{23})=f(Y_{12})+f(Y_{23})-f(Y_{12}\cap Y_{23}) \leq \ell+\ell-\ell=\ell \]
since $Y_{12}\cap Y_{23} \supseteq v_2$ and $Y_{12}\cap Y_{23}\subset G$. But then the addition of $v$ and its three incident edges to $ Y_{12}\cup Y_{23}$ gives a subgraph $Y$ of $G$ with $f(Y)=\ell-1$ which contradicts the fact that $G$ is $(2,\ell)$-tight.
\end{proof}
\begin{lem} \label{jco6}
Let $G=(V,E)$ be a $(2,\ell)$-tight graph for $\ell=1,2$. Then $G$ has either
an inverse Henneberg 1 move, an inverse Henneberg 2 move or at least $2l$ degree $3$ vertices, each of which is in a copy of $K_4$.
\end{lem}
\begin{proof}
$G$ has no degree $1$ vertices since this would imply that there is an edge $ab \in E(G)$ and
$G=Y \cup ab$ with $b \notin V(Y)$ and $f(Y) = l-1$.
Assume $G$ has no inverse Henneberg 1 move. Then every vertex has degree at least three.
Label the vertices $1,\dots, |V|$ and let $d(i)$ denote the degree of vertex $i$. The summation over the degree of all vertices in $G$ gives $2|E|$. Hence the
condition that $G$ is $(2,\ell)$-tight gives
\begin{equation}
\sum_{i=1}^{|V|}(4-d(i))=2l.
\label{degreecount}
\end{equation}
Since $d(i) \geq 3$ this implies $G$ has at least $2\ell$ degree $3$ vertices. By
Lemma \ref{k4lem} $G$ has an inverse Henneberg 2 move or each of these degree $3$ vertices is in a copy
of $K_4$.
\end{proof}
We will say that a $K_3$-to-edge or a $K_4$-to-vertex move is \emph{allowable} if it results in a graph
which is simple and $(2,\ell)$-tight.
The following lemma shows that a $K_4$-to-vertex move is allowable provided that the copy of $K_4$ does not have two vertices in a single copy of $K_3$.
We use the notation $K_n(v_1, \dots, v_n)$ for a subgraph of $G$ which is a copy of the complete
graph $K_n$ on the vertices $v_1, \dots, v_n$.
\begin{lem} \label{jco7}
Let $G$ be a $(2,\ell)$-tight graph with $|V(G)|>4$ and let $G \rightarrow G/K_4$ be a $K_4$-to-vertex move. Then
$G/K_4$ is
simple and $(2,\ell)$-tight unless there is a $K_3$ in $G$ with
$|V(K_3 \cap K_4)|=2$.
\end{lem}
\begin{proof}
$G/K_4$ is simple unless there is a vertex $v \in V(G) \setminus V(K_4)$ and edges $a,b \in E(G)$ with
$a,b \in V(K_4)$. In this case $|V(K_3(v,a,b) \cap K_4)|=2$.
$f(G/K_4)=f(G)$ so $G/K_4$ is $(2,\ell)$-tight unless there is a
$Y' \subset G/K_4$ with
$f(Y')<l$. There is a corresponding $Y \subset G$ such that $Y' = Y/K_4$. But then $f(Y)<l$ because $f(Y)=f(Y')$
which contradicts the $(2,\ell)$-sparsity of $G$.
\end{proof}
The following lemma describes when a $K_3$-to-edge move is allowable. Note that a $(2,\ell)$-tight graph containing no
copy of $K_3$ admits an inverse Henneberg move by Lemmas \ref{jco6} and \ref{k4lem}.
\begin{lem} \label{jco1}
Let $G$ be a $(2,\ell)$-tight graph and $G \rightarrow G'$ a $K_3$-to-edge move in which
the vertices $a, b \in K_3(a,b,c)$ are the vertices in $G$ which are merged. Then $G'$ is
simple unless there is a $K_3(a,b,d)$ in $G$ with $d \neq c$ and $G'$ is $(2,\ell)$-sparse unless there is a
$Y \subset G$ with $ab \in E(Y)$, $c \not\in V(Y)$ and $Y$ is $(2,\ell)$-tight.
\end{lem}
\begin{proof}
$G'$ is simple provided there is no vertex $d$ different from $c$ and two edges $da$,
$db$. This gives the first condition.
$G'$ is $(2,\ell)$-sparse provided it has no
subgraph $Y'$ with $f(Y') < l$. $Y'$ is also a subgraph of $G$ unless it
derives from a subgraph $Y \subset G$ with $ab \in E(Y)$ and $f(Y') < f(Y)$ only if
$c \not\in Y$.
\end{proof}
There are three possible $K_3$-to-edge moves which can be applied to a copy of $K_3$ in $G$. If none of these results in a simple graph then there are three
further copies of $K_3$ in $G$ and, if these are distinct, there are six further $K_3$-to-edge moves which might result in a simple graph. We will use this
growth in the number of copies of $K_3$ to show that if $G$ contains a copy of $K_3$ then either $G$ has an allowable $K_3$-to-edge move
or every copy of $K_3$ is in a copy of $K_4$ (Lemma \ref{jco4} below). This $K_4$ gives an allowable $K_4$-to-vertex move unless it is
adjacent to a copy of $K_3$ which, by this argument, must also be in another copy of $K_4$. This allows us to put a
strong constraint on the possible graphs which contain a copy of $K_3$ but no allowable $K_3$-to-edge or $K_4$-to-vertex move (Lemma \ref{jco5} below).
In order to keep track of the way in which copies of $K_3$ may share edges in a $(2,\ell)$-tight graph we first define a triangle sequence which is a set of
nested subgraphs of $G$
and derive some of its properties.
\begin{defn} \label{jco2}
Let $G$ be a simple graph. A triangle sequence in $G$ is a nested set of subgraphs
\[ M_3 \subset M_4 \subset \dots \subset M_i \dots \subset M_n \subseteq G\]
where $M_3$ is a copy of $K_3$, $E(M_i)$ and $V(M_i)$ are respectively the sets of edges and vertices of $M_i$,
$|V(M_i)|=|V(M_{i-1})|+1$ and if $V(M_i) \setminus V(M_{i-1})=v_i$ then $E(M_i) \setminus E(M_{i-1})=v_ia_i,v_ib_i$ where
$a_ib_i \in E(M_{i-1})$ and $a_ib_i$ is in exactly one copy of $K_3$ in $M_{i-1}$. We use $S(M_i)$ to denote the set of
edges in $E(M_i)$ which are in exactly one copy of $K_3$ in $M_i$ (so $a_ib_i \in S(M_{i-1}))$.
\end{defn}
We will often refer to a triangle sequence by the largest graph in the sequence. A maximal length triangle sequence is one which cannot
be extended by a single vertex in $G$. We
note that even for a maximal length triangle sequence with largest graph $M_n$ the graph $G$ may contain
copies $K_3(a,b,c)$ which are not subgraphs of $M_n$ even though $ab \in E(M_n)$. This may occur if $c \in V(M_n)$ or if
$c \notin V(M_n)$ and the edge $ab$ is in more than one copy of $K_3$ in $M_n$. Since $M_n$ is itself a graph we may form
different
triangle sequences within $M_n$ for example by starting with different copies of $K_3$ in $M_n$, see Figure \ref{johnfig}.
\begin{center}
\begin{figure}[ht]
\centering
\includegraphics[width=7cm]{johnfig.eps}
\caption{Two possible maximal length triangle sequences in $K_4 \sqcup K_4$. In each case the graph shown is the final graph $M_6$ in the
sequence. The dashed lines represent edges which are in $K_4 \sqcup K_4$ but not in $M_6$. The intermediate
graphs in the sequence are obtained by starting with any copy of $K_3$ and sequentially adding one vertex and two
edges from an adjacent copy of $K_3$}
\label{johnfig}
\end{figure}
\end{center}
\begin{lem} \label{jco3}
A triangle sequence in $G$ has the following properties.
\begin{enumerate}
\item $|V(M_i)|=i$ and $|E(M_i)| = 2i-3$.
\item The edges in $S(M_i)$ form a spanning cycle of $M_i$.
\item For every edge $ab \in E(M_i) \setminus S(M_i)$ the vertex pair $a,b$ separates $M_i$ (as a graph) with the property
that if
$aa_l,aa_r \in S(M_i)$ then $a_l,a_r$ are in different separation components.
\item If $K_3(a,b,c)$ is any copy of $K_3$ in $M_n$ then there is a triangle sequence
$M_3' \subset \dots \subset M_m'$ in $M_n$ such that $M_3'=K_3(a,b,c)$ and $M_m' = M_n$.
\end{enumerate}
\end{lem}
\begin{proof}
Property $(1)$ follows by induction since $|V(M_i)|=|V(M_{i-1})|+1$ and $|E(M_i)|=|E(M_{i-1})|+2$.
Property $(2)$ follows by induction. The edges of $S(M_3)$ form a spanning cycle of $M_3$. Assume
property $(2)$ is true for $M_{i-1}$. Let $V(M_i)=V(M_{i-1}),v_i$ and let $E(M_i)=E(M_{i-1}),v_ia_i,v_ib_i$. The edge $a_ib_i \in S(M_{i-1})$ is in $K_3(a_i,b_i,v_i) \subset M_i$
in addition to a copy of $K_3$ in $M_{i-1}$ so is not in $S(M_i)$. The two edges $a_iv_i$ and
$b_iv_i$ are both in $K_3(a_i,b_i,v_i)$ (and in no other copy of $K_3$ in $M_i$) so these are in
$S(M_i)$. If the edges in $S(M_{i-1})$ form a spanning cycle $C_{i-1}$ of $M_{i-1}$ then the cycle
\[ C_i = (C_{i-1} \setminus a_ib_i) \cup a_iv_i \cup b_iv_i\]
forms a spanning cycle of $M_i$.
Property $(3)$ is also proved by induction. It is trivially true for $M_3$. Assume it is true for any $M_{i-1}$.
Let $V(M_i)=V(M_{i-1}),v_i$ and let $E(M_i)=E(M_{i-1}),v_ia_i,v_ib_i$. Every vertex pair which separates $M_{i-1}$ also
separates $M_i$ with the same components because the vertices $a,b$ are adjacent and so are in the same component of any
vertex separation of $M_{i-1}$. Putting vertex $v_i$ in this component gives a corresponding vertex separation of $M_i$.
The edge $ab$ is the only edge which is in $E(M_i) \setminus S(M_i)$ and not in $E(M_{i-1}) \setminus S(M_{i-1})$. The
vertex pair $a,b$ separates the vertex $v_i$ from the vertices $V(M_{i-1}) \setminus a,b$. The neighbours of $a$ in
$S(M_i)$ are $v_i$ and a vertex $a_l \neq b \in V(M_{i-1}) \setminus a,b$ and these are separated by $a,b$.
For property $(4)$ we will show there is a triangle sequence in $M_n$ starting with $K_3(a,b,c)$ and
terminating with $M_m'$ for
which $m=n$. This implies $|E(M_m')|=|E(M_n)|$ and that $M_m'=M_n$. Let $M_3',M_4' \dots M_m'$ be a
maximal length triangle sequence in $M_n$ starting with $K_3(a,b,c)$. Suppose for a contradiction that $m < n$. The edges in
$S(M_m')$ form a spanning cycle
of $M_m'$ and there is a edge $a_mb_m$ in $S(M_m')$ which is not in $S(M_n)$
(since $m<n$ and a cycle contains no proper subcycles). Since $a_mb_m$ is in $E(M_n)$ but not in $S(M_n)$ there is a vertex $v_m$ in $V(M_n) \setminus V(M_m')$ such that
there is $K_3(a_m,b_m,v_m)$ which in is in $M_n$ and not in $M_m'$. The edge $a_mb_m$ is therefore in a subgraph
$K_3(a_m,b_m,v_m)$ of $M_n$ but is not in $M_m'$. This implies that
$v_m \in V(M_n) \setminus V(M_m')$ and $a_mv_m, b_mv_m \in E(M_n) \setminus E(M_m')$
with $a_mb_m \in S(M_m')$. This contradicts the maximality of the triangle sequence
in $M_n$.
\end{proof}
The following lemma uses a maximum length triangle sequence to show that if $G$ has a copy of $K_3$ but does not have a
$K_3$-to-edge move then every edge in a copy of $K_3$ in $G$ is in at least two copies of $K_3$ in $G$.
\begin{lem} \label{jco4a}
Let $G$ be a $(2,\ell)$-tight graph for $l=1,2$ containing a copy of $K_3$. Then either
\begin{enumerate}
\item[$(i)$] there is a $K_3$-to-edge move in $G$ which gives a $(2,l)$-tight graph or
\item[$(ii)$] every edge in a copy of $K_3$ in $G$ is in at least two copies of $K_3$ in $G$.
\end{enumerate}
\end{lem}
\begin{proof}
Suppose that an edge $e=ab \in E(G)$ is in exactly one copy $K_3(a,b,c) \subset G$. By Lemma \ref{jco1}, the $K_3$-to-edge
move which merges vertices $a$ and $b$ gives a simple graph $G^\prime$ and $G^\prime$ is $(2,\ell)$-tight unless $ab$ and
$c$
have the following property (P): there is a
$(2,\ell)$-tight subgraph $Y \subset G$ with $ab \in E(Y)$ and $c \notin V(Y)$.
Suppose for a contradiction to the lemma that every edge in $G$ which is in exactly one copy of $K_3$ satisfies this
property.
Let $M_3 \subset M_4 \dots \subset M_i \dots \subset M_n \subset G$
be a maximal length triangle sequence in $G$. Every edge in $E(M_n) \setminus S(M_n)$ is in two copies of $K_3$.
Suppose there is exactly one edge $ab$ in $S(M_n)$ which is in exactly one copy of $K_3$ in $G$ and therefore satisfies
property (P) with corresponding subgraph $Y$.
We will show by induction that
$V(M_n) \cap V(Y) = \{a,b\}$ and $f(Y \cup M_n) = \ell$. Since $ab \in S(M_n)$ there is a vertex $c$ in $V(M_n)$ such that
$K_3(a,b,c)$ is in $M_n$.
By property $(4)$ of Lemma \ref{jco3}
there is a triangle sequence $M_3' \subset \dots M_i' \dots \subset M_n$, starting with
$M_3' = K_3(a,b,c)$ and ending with $M_n$.
$V(M_3') \cap V(Y) = \{a,b\}$ and $f(Y \cup M_3') = \ell$.
Assume for the induction that $V(M_{i-1}) \cap V(Y) = \{a,b\}$ and that $f(Y \cup M_{i-1}) = \ell$.
Let $V(M_i) \setminus V(M_{i-1})=v_i$. If
$v_i \in V(Y)$ then
\[f(Y \cup v_ia_i \cup v_ib_i) = \ell-2\]
which contradicts the $(2,\ell)$-sparsity of $G$. Thus $V(M_i) \cap V(Y) = \{a,b\}$ and $f(Y \cup M_i) = \ell$.
Every edge $cd$ in $S(M_n) \setminus ab$ is in a subgraph $K_3(c,d,v)$ of $G$
where $K_3(c,d,v)$ is not a subgraph of $M_n$.
Since $M_n$ is the largest graph in a maximal length triangle sequence we must have $v \in V(M_n)$ else $M_n$ could be
extended to include $v$. But then
$f(Y \cup M_n \cup cv)=\ell-1$ and since $Y \cup M_n \cup cv$ is a subgraph of $G$ this contradicts the
$(2,\ell)$-sparsity of $G$.
Suppose there is more than one edge in $S(M_n)$ which is in exactly one copy of $K_3$ in $G$.
There are subgraphs $Y_1$ and $Y_2$
and edges $a_1b_1 \in Y_1 \cap M_n$ and $a_2b_2 \in Y_2 \cap M_n$. If the vertices
$a_1,b_1,a_2,b_2$ are distinct then
\[f(Y_1 \cup Y_2 \cup M_n) \leq 2\ell-3\]
because $f(Y_1 \cup Y_2) \leq 2l$ and there are $n-4$ vertices and $2n-5$ edges in $M_n$ which are not in
$Y_1 \cup Y_2$. If two of the vertices $a_1,b_1,a_2,b_2$ are the same then
\[f(Y_1 \cup Y_2 \cup M_n) = \ell-1\]
since $f(Y_1 \cup Y_2) = \ell$ and there are $n-3$ vertices and
$2n-5$ edges in $M_n$ which are not in $Y_1 \cup Y_2$. In either case this contradicts the $(2,\ell)$-sparsity of $G$ for
$l=1,2$.
\end{proof}
We say that an edge $ab \in E(G)$ is a $chord$ of $M_n$ if $a,b \in V(M_n)$ and $e \notin E(M_n)$. Let
$[M_n]$ denote the graph induced in $G$ by $V(M_n)$. Then
$E([M_n]) \setminus E(M_n)$ is the set of chords of $M_n$. The set $C$ defined in the next lemma is the set of edges in
$S(M_n)$ which are in two or more copies of $K_3$ in $[M_n]$. We will show that when $M_n$ is the largest subgraph in a
maximal length triangle sequence this is the same as a set of edges in $S(M_n)$ which are in two or more copies of $K_3$ in
$G$. This lemma can then be used to limit the length of a triangle sequence because the number of chords of $M_n$ is
limited to one for $\ell=2$ and to two for $\ell=1$ by the $(2,\ell)$-sparsity of $G$.
We use the notation
$\cup_{i=1}^m A_i$ to denote $A_1 \cup A_2,\dots,\cup A_m$ where $A_i$ are sets or graphs.
\begin{lem} \label{jco10}
Let $G$ be graph and let $M_n$ be a subgraph in a triangle sequence in $G$ with $n > 4$.
Let $e_1,\dots,e_m$ for $m > 0$ be chords of $M_n$,
let $C_i=\{f \in S(M_n) : \exists g \in E([M_n])$ such that
$K_3(e_i,f,g) \subset G\}$. Then $|C| \leq 3m$ where $C=\cup_{i=1}^m C_i$.
\end{lem}
\begin{proof}
Assume for induction that the lemma is true for all possible choices of $m-1$ chords of $M_n$ and suppose that
$e_1,\dots,e_m$ are a set of $m$ chords of $M_n$.
Suppose the chords $e_1,\dots,e_m$ determine a graph with $t$ distinct vertices and $c$ connected components. Then
$t \leq m+c$. Since the edges in $S(M_n)$ form a spanning cycle of $M_n$ each vertex of a
chord $e_i$ is incident to two edges in $S(M_n)$. This implies $|C| \leq 2(m+c)$ which implies $|C| \leq 3m$ unless $c > m/2$. We may assume therefore that
there is at least one component with exactly one
edge which we label as the edge $e_m$ where $e_m$ has no vertices in common with $e_i,i=1,\dots,m-1$.
Let $e_m=ab$ with $a,b \in V(M_n)$. Each of the vertices $a,b$ is
incident to exactly two edges in $S(M_n)$ which we label $aa_l,aa_r,bb_l,bb_r \in S(M_n)$. These
edges are all distinct because $ab \notin S(M_n)$ . Since the edges in $S(M_n)$ form a cycle we may label the vertices so that there is a
(possibly trivial) path $P(a_l,b_l) \in S(M_n)$ which connects $a_l,b_l$ and avoids $a,b,a_r,b_r$ and then
$a_r \neq b_l$ and $a_l \neq b_r$, see Figure \ref{lemma3.8}. We may also label so that $a_r \neq b_r$ since if $a_l=b_l$ and
$a_r=b_r$ the edges $aa_l,a_lb,ba_r,a_ra$ form a 4-cycle in $S(M_n)$ which contradicts $n > 4$.
\begin{center}
\begin{figure}[ht]
\centering
\begin{tikzpicture}
\filldraw[black]
(0,0) circle (3pt)node[anchor=south]{$b_l$}
(0,2) circle (3pt)node[anchor=north]{$a_l$}
(3,-.5) circle (3pt)node[anchor=north]{$b$}
(3,2.5) circle (3pt)node[anchor=south]{$a$}
(6,0) circle (3pt)node[anchor=south]{$b_r$}
(6,2) circle (3pt)node[anchor=north]{$a_r$};
\draw[black,thick]
(0,0) -- (3,-.5) -- (3,2.5) -- (0,2);
\draw[black,thick] (3,2.5) -- (6,2);
\draw[black,thick] (6,0) -- (3,-.5);
\draw[black] (0,1) ellipse (0.4 and 1.2);
\draw[black] (6,1) ellipse (0.4 and 1.2);
\draw[black,dashed] (3,2.5) -- (6,0);
\draw[black,dashed] (3,-.5) -- (6,2);
\end{tikzpicture}
\caption{A chord $ab$ of $M_n$ with $ab\in V(M_n)$ and adjacent to edges $aa_l, aa_r,bb_l, bb_r\in S(M_n)$. Edges $ab_r$ and $ba_r$ cannot both be in $E(M_n)$ because the vertex pair $b,a_r$ must then separate $M_n$.}
\label{lemma3.8}
\end{figure}
\end{center}
Any edge $f \in S(M_n)$ which is in a 3-cycle with $ab$ has a vertex in common with $ab$. Given edges $e,f$ there is at most one
3-cycle in $G$ which includes $e,f$. Thus we have shown $|C_m| \leq 4$. Furthermore, if $|C_m|=4$ the vertex triples $a,b,a_r$ and $a,b,b_r$ must
both induce 3-cycles in $[M_n]$. This implies $ba_r, ab_r \in E(M_n)$ because edges in $e_1,\dots,e_{m-1}$ have no vertices in common with $ab$. This contradicts
Lemma \ref{jco3} part (3) for the vertex pair $b,a_r$ because
the neighbours of $b$ in $S(M_n)$ are $b_l,b_r$ and there would be a path $b_ra,aa_l,P(a_l,b_l)$ which connects $b_r$ and $b_l$ and excludes $b,a_r$.
Thus $|C_m| \leq 3$ which combines with the induction hypothesis $\cup_{i=1}^{m-1} C_i \leq 3(m-1)$ to give $\cup_{i=1}^m C_i \leq 3m$ .
\end{proof}
\begin{lem} \label{jco4}
Let $G$ be a $(2,\ell)$-tight graph for $\ell=1,2$ with the property that
every edge $ab$ in a $K_3(a,b,c) \subset G$ is in at least two copies of $K_3$ in $G$.
Then every copy of $K_3$ in $G$ is in a copy of $K_4$.
\end{lem}
\begin{proof}
We will show first that every maximal length triangle sequence in $G$ with largest graph $M_n$ satisfies $n \leq 6$ for $\ell=1$ and $n \leq 4$
for $\ell=2$.
Since every edge $ab \in S(M_n)$ is in exactly one copy of $K_3$ in $M_n$ there is a vertex $c \in V(G)$ such that $K_3(a,b,c) \subset G$ and
$K_3(a,b,c) \not\subset M_n$. This implies that $c \in V(M_n)$ because otherwise the triangle sequence can be extended with vertex $c$. Since
$K_3(a,b,c) \not\subset M_n$ either $ac$ or $bc$ is a chord of $M_n$. Every edge in $S(M_n)$ must therefore be in the set $C$ defined in Lemma
\ref{jco10} and if $n>4$ by Lemma \ref{jco10} we have $n =|C| \leq 3m$ where $m$ is the number of chords of $M_n$ in $G$.
$f(M_n \cup_{i=1}^m e_i)=3-m$ because $f(M_n)=3$ so
$f(M_n \cup_{i=1}^m e_i) \ge \ell$ implies $m \leq 3-\ell$ and $n \leq 3(3-\ell)$. These imply $n \leq 4$ for $\ell=2$ and
$n \leq 6$ for $\ell=1$.
For $n=4$ there is a unique largest graph $M_4$ and a unique edge from $E([M_n]) \setminus E(M_n)$ which can be added to
the graph $M_4$ so that every edge of $S(M_4)$ is in two copies of $K_3$. This creates a copy of $K_4$.
An analysis of the subgraphs induced by the vertices of maximal length triangle sequences $M_n$ with $n \leq 6$ and with
the
property that every
edge in $E(M_n)$ is in two copies of $K_3$ in $G$ shows that for $l=1$, $[M_5]=K_5\setminus e$ or
$[M_6]=K_4 \sqcup K_4$. Since every $K_3$ is in a maximal length triangle sequence and every $K_3$ in
$K_4$, $K_5 \setminus e$ or $K_4 \cup K_4$ is in a copy of $K_4$ the proof is complete.
\end{proof}
\begin{lem} \label{jco5}
Let $G$ be a $(2,\ell)$-tight graph for $l=1,2$ which contains a copy of $K_3$. Then either $G=K_4$, $G$ has an
allowable $K_3$-to-edge move, an allowable $K_4$-to-vertex move or every copy of $K_3$ is in a copy
of $K_4 \sqcup K_4$ or $K_5 \setminus e$.
\end{lem}
\begin{proof}
Let the copy of $K_3$ be $K_3(a,b,c)$ and assume $G$ has no allowable $K_3$-to-edge move or
$K_4$-to-vertex move. By Lemmas \ref{jco4a} and \ref{jco4} $K_3(a,b,c)$ is in a $K_4(a,b,c,d)$. Since this does not give an allowable $K_4$-to-vertex move, by Lemma \ref{jco7} there is a $K_3(c,d,e)$ (say) with $a,b,c,d,e$ all distinct and again by
Lemma \ref{jco4} there is a $K_4(c,d,e,g)$. If $a,b,c,d,e,g$ are distinct then $K_3(a,b,c)$ is in a copy of $K_4 \sqcup K_4$ and if $g=a$ or $b$ then $K_3(a,b,c)$ is in a copy of $K_5 \setminus e$.
\end{proof}
We combine the lemmas in this section to show that all suitable $(2,\ell)$-tight graphs
have an allowable reduction move.
\begin{lem} \label{jco8}
Let $G$ be $(2,2)$-tight. Then $G=K_4$ or $G$ has an inverse Henneberg $1$ move,
an inverse Henneberg $2$ move, an allowable $K_3$-to-edge move or an allowable $K_4$-to-vertex move.
\end{lem}
\begin{proof}
Assume $G$ has no inverse Henneberg $1$ move and no inverse Henneberg $2$. By Lemma \ref{jco6} $G$ has a
copy of $K_4$ and thus a copy of $K_3$. The proof is completed by Lemma \ref{jco5} since neither
$K_4 \sqcup K_4$ nor $K_5 \setminus e$ is $(2,2)$-sparse.
\end{proof}
\begin{lem} \label{jco9}
Let $G$ be $(2,1)$-tight. Then $G=K_4 \sqcup K_4$ or $G=K_5 \setminus e$ or $G$ has an inverse
Henneberg $1$ move, an inverse Henneberg $2$ move, an allowable $K_3$-to-edge move, an allowable $K_4$-to-vertex move or an edge separation move.
\end{lem}
\begin{proof}
Assume $G$ has no inverse Henneberg 1 move, no inverse Henneberg 2 move, no allowable
$K_3$-to-edge move and no allowable $K_4$-to-vertex move. By Lemma \ref{jco6}
each of the degree-3 vertices in $G$ is in a copy of $K_4$ and thus in a copy of $K_3$. By Lemma
\ref{jco5} each of these $K_3$ is in a copy of $K_4 \sqcup K_4$ or $K_5 \setminus e$.
Let $Y=\{Y_1, \dots, Y_n\}$ be the set of subgraphs of $G$ which are each copies of
$K_4 \sqcup K_4$ or $K_5 \setminus e$.
The subgraphs $Y_i \in Y$ are vertex disjoint since
\[f(Y_i \cup Y_j)= f(Y_i)+f(Y_j)-f(Y_i \cap Y_j) =2-f(Y_i \cap Y_j)\]
and $(2,1)$-sparsity requires $f(Y_i \cap Y_j) \leq 1$. Every proper subgraph $X$ of $K_4 \sqcup K_4$
or $K_5 \setminus e$ has $f(X) \geq 2$ so this requires $Y_i$ and $Y_j$ to be vertex disjoint.
Let $V_0$ and $E_0$ be the sets of vertices and edges in $G$ which are in none of the $Y_i \in Y$.
Then
\[f(G)=\sum_{i=1}^nf(Y_i)+2|V_0|-|E_0|\]
so $|E_0|=2|V_0|+n-1$. Each of the vertices in $V_0$
is incident to at least 4 edges in $E_0$. If each $Y_i$ is incident to at least 2 edges in $E_0$
then $|E_0| \geq (4|V_0|+2n)/2$ for a contradiction.
At least one of the $Y_i$ is incident to at most one edge in $E_0$. If this $Y_i$ is incident to
no edges in $E_0$ then $G=K_4 \sqcup K_4$ or $G=K_5 \setminus e$ since $G$ is connected. Otherwise
$Y_i$ is incident to one edge $e \in E_0$ and $e$ provides an edge separation move.
\end{proof}
Using the above lemmas we reach the stated goal of this section.
\begin{proof}[Proof of $(1)\Rightarrow(2)$ in Theorem \ref{21theorem} or Theorem \ref{22refine}]
By induction using Lemma \ref{jco8} or Lemma \ref{jco9}.
\end{proof}
\section{Further Work}
\label{21apps}
We expect to be able to use Theorem \ref{21theorem} to prove the following conjecture discussed in the introduction.
\begin{con}\label{tjcon}
Let $\M$ be a cone, a torus, a union of concentric cones or a union of concentric tori and let $p$ be generic. Then $(G,p)$ is generically minimally rigid on $\M$ if and only if $G=K_{2},K_3, K_4$ or $G$ is $(2,1)$-tight.
\end{con}
It would also be interesting to consider surfaces that do not admit any rigid-body motions. For such surfaces there are immediate additional problems.
For example Equation \ref{degreecount} with $\ell=0$ shows that the minimum degree in a $(2,0)$-tight graph may be $4$ so additional Henneberg type
operations are required.
This actually provides additional motivation for studying these graphs since the obvious choices to take are $X$ and $V$-replacement
as studied by Tay and Whiteley \cite{T&W} in the $3$-dimensional setting. Indeed they conjecture that these operations (with additional
conditions for $V$-replacement)
preserve rigidity in $3$-dimensions.
It is also interesting to note that the $d$-dimensional version of the edge-to-$K_3$ move, known in the literature as vertex
splitting \cite{Whi6}, is one of a very short list of operations known to preserve rigidity in arbitrary dimension. Nevertheless there is no
conjectured inductive construction, even in $3$-dimensions, that makes use of this. We hope that our methods for dealing with the
edge-to-$K_3$ move for $(2,\ell)$-tight graphs may be useful in finding such a construction.
There are more exotic settings in which the class of $(2,1)$-tight graphs are the appropriate combinatorial tool needed to classify
generic minimal rigidity. For example we could take $\M$ to be two parallel (but not concentric) cylinders. Here there is only one
rigid-body motion of $\M$ in $\bR^3$, or we may take $\N$ to be a cylinder coaxial to a cone. Again there is only one rigid-body motion
(this time a rotation about the central axis). In such reducible settings there is a little more work to do to in considering which surface each framework point lies on. This extra requirement is particularly evident for
$\N$, but in either case a $(2,1)$-tight subgraph realised purely on one cylinder would be overconstrained.
A similar but deeper topic is the problem of when a framework realisation is unique (this is the topic of global rigidity, see for
example \cite{J&J} and \cite{Con2}). To characterise the global rigidity of frameworks supported on an algebraic surface one of the key steps is to analyse the circuits of the rigidity matroid $\R_{\M}$ (this is the linear matroid induced
by the linear independence of the rows of the surface rigidity matrix). Since the independent sets in $\R_{\M}$, for $\M$ a cylinder,
may be identified
with the $(2,2)$-tight graphs (\cite[Theorem $5.4$]{NOP}), the circuits may be identified with a sub-class of the
$(2,1)$-tight graphs. Such a recursive construction is given in \cite{Nix} and finding a similar construction for circuits in the $(2,1)$-tight
matroid is open.
|
\section{\bf Introduction}
Uniqueness for the Fourier transform acting on an integrable (or square integrable) function means that $\widehat{f}=0$ implies $f=0$. In other words, the Fourier transform is injective. For distributions, a related statement says that if the Fourier transform $\widehat{f}$ is supported at the origin then $f$ is a polynomial. In particular, if a bounded function has distributional Fourier transform supported at the origin, then it is a constant function.
This note establishes analogous uniqueness results for the continuous wavelet transform. Given a function $\psi$ (which we call a \emph{wavelet}) and a function $f$ (which we call a \emph{signal}), the \emph{continuous wavelet transform} of $f$ with respect to $\psi$ is the function
\[
(W_\psi f)(s,t) = \langle f , \psi_{s,t} \rangle = \int_{{\mathbb R}^d} f(x) \overline{\psi_{s,t}(x)} \, dx , \qquad s>0 , \quad t \in {{\mathbb R}^d} ,
\]
where
\[
\psi_{s,t}(x) = \frac{1}{s^{d/2}} \psi \Big( \frac{x-t}{s} \Big) .
\]
Notice $s$ denotes the scale, and $t$ the translation.
The scale and translation parameters vary continuously and so one calls $W_\psi$ the ``continuous'' wavelet transform, in distinction to the ``discrete'' wavelet transform which restricts to dyadic scales $s=2^j$ and translations $t=2^j k$. For more on wavelet transforms, readers may consult the texts of Daubechies \cite{D92},
Holschneider \cite{H95}, Mallat \cite{M09}, Meyer \cite{M92} and Pathak \cite{P09}. For precise relations between the continuous and discrete wavelet transforms, see Laugesen's work on translational averaging \cite{L01,L02}.
We will present four uniqueness (or injectivity) results for the continuous wavelet transform. The first result deals with signals in $L^2$, under slightly weaker assumptions than the Calder\'{o}n admissibility condition. The second result handles signals in $L^p$ for $1 \leq p<\infty$. The third treats $L^\infty$, and the case of polynomially bounded signals. The fourth result considers tempered distributions.
Our motivation comes from work of Sun and Sundararajan \cite{SS11} in theoretical economics. There the signal $f$ is a mixed partial derivative of the characteristic function of some \emph{attribution problem}. Such problems arise in cooperative game theory as cost-sharing problems, in investment finance as performance analyses of investment portfolios, and in operations research settings in the analysis of production process performance. The wavelet $\psi$ represents the difference between two path-generated attribution methods. With the help of our wavelet uniqueness results, these authors show, roughly speaking, that if two different path-generated attribution methods yield the same attributions, then the characteristic function must lie in some specific constrained class.
\section{\bf Wavelet uniqueness for square integrable signals}
We say that a function $\lambda$ on ${{\mathbb R}^d}$ is \emph{nontrivial in direction $\xi$} (where $\xi$ is a unit vector) if the set $\{ r>0 : \lambda(r\xi) \neq 0 \}$ has positive measure. For example, if $\lambda$ is continuous then nontriviality in a direction simply means $\lambda$ is nonzero at some point on the ray in that direction.
Our first result treats uniqueness for the continuous wavelet transform in $L^2$.
\begin{prop} \label{integrable}
Assume $f, \psi \in L^2({{\mathbb R}^d})$. If
\[
\langle f , \psi_{s,t} \rangle = 0 \qquad \text{for all $s>0 , \quad t \in {{\mathbb R}^d}$,}
\]
(in other words if $W_\psi f \equiv 0$), then
\begin{equation} \label{underlying}
\int_0^\infty |\widehat{f}(r\xi)|^2 r^{d-1} \, dr \, \int_0^\infty |\widehat{\psi}(s\xi)|^2 s^{d-1} \, ds = 0
\end{equation}
for almost every unit vector $\xi$ in ${{\mathbb R}^d}$.
In particular, if $W_\psi f \equiv 0$ then nontriviality of $\widehat{\psi}$ in almost every direction implies $f = 0$ a.e.
\end{prop}
In dimension $d=1$, the assumption that $\widehat{\psi}$ is nontrivial in almost every direction means that $\widehat{\psi}$ is nontrivial on each side of the origin: the two sets $ \{ r>0 : \widehat{\psi}(r) \neq 0 \}$ and $ \{ r>0 : \widehat{\psi}(-r) \neq 0 \}$ both have positive measure.
This nontriviality assumption on the Fourier transform is the standard \emph{Tauberian condition} in harmonic analysis.
The symmetry of equation \eqref{underlying} reflects the interchangeability of the signal and the wavelet, which one sees by a simple change of variable:
\[
\langle f , \psi_{s,t} \rangle = \langle f_{1/s,-t/s} , \psi \rangle
\]
\subsection*{Relation to the Calder\'{o}n condition.} The Calder\'{o}n admissibility condition for continuous wavelets says that
\[
\int_0^\infty |\widehat{\psi(s\xi)}|^2 \, \frac{ds}{s} = 1
\]
for almost every unit vector $\xi$ (see \cite[Section 2.4]{D92}, \cite{LWWW02}). This condition implies the hypothesis in Proposition~\ref{integrable} that $\widehat{\psi}$ is nontrivial in almost every direction, and indeed is stronger than that hypothesis because our $\widehat{\psi}$ need not vanish at the origin (or can vanish so slowly there that the Calder\'{o}n integral diverges).
On the other hand, the Calder\'{o}n condition guarantees more than just uniqueness for the wavelet transform: it guarantees a Plancherel formula
\[
\lVert f \rVert_2^2 = \int_0^\infty \int_{{\mathbb R}^d} |\langle f , \psi_{s,t} \rangle|^2 \, dt \, \frac{ds}{s^{d+1}} ,
\]
and hence (by polarization) a reproducing formula. Thus our Proposition assumes less and obtains less than the standard theory based on the Calder\'{o}n condition.
\begin{proof}[Proof of Proposition~\ref{integrable}]
Define the Fourier transform with $2\pi$ in the exponent:
\[
\widehat{\psi}(\omega) = \int_{{\mathbb R}^d} \psi(x) e^{-2\pi i \omega \cdot x} \, dx .
\]
Then by Parseval's identity and the vanishing of the wavelet transform we have
\[
0
= \langle f , \psi_{s,t} \rangle \\
= \langle \widehat{f} , \widehat{\psi_{s,t}} \rangle .
\]
Direct calculation of the Fourier transform shows that $\widehat{\psi_{s,t}}(\omega) = \widehat{\psi}(s\omega) e^{-2\pi i \omega \cdot t} s^{d/2}$, and so the previous formula says that
\[
0 = \big[ \widehat{f}(\cdot) \overline{\widehat{\psi}(s \cdot)} \, \big] \widehat{\ }(-t) \qquad \text{for each $t \in {{\mathbb R}^d}$.}
\]
Since the integrable function $\widehat{f}(\cdot) \overline{\widehat{\psi}(s \cdot)}$ has vanishing Fourier transform, it must equal zero a.e., which means
\[
0 = \widehat{f}(\omega) \overline{\widehat{\psi}(s\omega)} \qquad \text{for almost every $\omega \in {{\mathbb R}^d}$,}
\]
for each $s > 0$.
Next we square and multiply by $|\omega|^d s^{d-1}$, and then integrate to show that
\begin{align*}
0
& = \int_{{\mathbb R}^d} |\widehat{f}(\omega)|^2 \int_0^\infty |\widehat{\psi}(s\omega)|^2 |\omega|^d s^{d-1} \, ds d\omega \\
& = \int_{S^{d-1}} \int_0^\infty |\widehat{f}(r\xi)|^2 r^{d-1} \, dr \, \int_0^\infty |\widehat{\psi}(s\xi)|^2 s^{d-1} \, ds \, dS(\xi)
\end{align*}
by expressing $\omega=r\xi$ in spherical coordinates and then changing variable with $s \mapsto s/r$. The integrand therefore vanishes for almost every $\xi$, which proves equation \eqref{underlying}.
Finally, if $\widehat{\psi}$ is nontrivial in almost every direction then $ \int_0^\infty |\widehat{\psi}(s\xi)|^2 s^{d-1} \, ds$ is positive for almost every $\xi$, and so (by the preceding formula) the integral $\int_0^\infty |\widehat{f}(r\xi)|^2 r^{d-1} \, dr$ must vanish for almost every $\xi$. Thus $\widehat{f}=0$ a.e.\ and so $f=0$ a.e. The argument works also with the roles of $\psi$ and $f$ interchanged.
\end{proof}
\section{\bf Uniqueness for $p$-integrable signals}
Next we treat signals in $L^p$. The Fourier transform of the signal is a distribution, when $p>2$. We show that distribution has support at the origin, which implies the signal must vanish.
\begin{thm}\label{p-int}
Let $1 \leq p < \infty$ and $\frac{1}{p}+\frac{1}{p\prime}=1$. Assume $f \in L^p({{\mathbb R}^d}), \psi \in L^{p\prime}({{\mathbb R}^d}), \widehat{\psi} \in C^\infty({{\mathbb R}^d} \setminus \{ 0 \})$, and that the wavelet transform vanishes identically:
\[
\langle f , \psi_{s,t} \rangle = 0 \qquad \text{for all $s>0 , \quad t \in {{\mathbb R}^d}$.}
\]
If $\widehat{\psi}$ is nontrivial in every direction then $f = 0$ a.e.
\end{thm}
The smoothness hypothesis on $\widehat{\psi}$ away from the origin simply means that some function $\nu \in C^\infty({{\mathbb R}^d} \setminus \{ 0 \})$ represents the distribution $\widehat{\psi}$ when acting on test functions $\eta \in C^\infty_c({{\mathbb R}^d} \setminus \{ 0 \})$; in other words $\widehat{\psi}[\eta] = \int_{{\mathbb R}^d} \nu(\omega) \eta(\omega) \, d\omega$. We will write $\widehat{\psi}$ to mean both the distributional Fourier transform and the function $\nu$, as there will be no danger of confusion.
This smoothness hypothesis on $\widehat{\psi}$ is satisfied if $\psi$ has compact support, because then $\widehat{\psi}$ is smooth on all of ${{\mathbb R}^d}$, including at the origin.
For a non-compactly supported example, $\psi$ could be a Gaussian or one of its derivatives (such as the Mexican hat, the negative second derivative of the Gaussian), in which case the Fourier transform is smooth on all of ${{\mathbb R}^d}$.
For an example where $\widehat{\psi}$ is not smooth at the origin, suppose $\psi(x)=1/\pi(1+x^2)$, which is the Poisson kernel in $1$ dimension. Then $\widehat{\psi}(\xi)=e^{-2\pi|\xi|}$ is smooth away from the origin, but not at the origin. Similarly if $\psi$ is the first derivative of the Poisson kernel (in which case $\psi$ has integral equal to zero) then $\widehat{\psi}(\xi)=2\pi\xi e^{-2\pi|\xi|}$, which is again smooth except at the origin.
\begin{proof}[Proof of Theorem~\ref{p-int}]
We will show that the tempered distribution $\widehat{f}$ is supported at the origin. Then $f$ is a polynomial (see \cite[Corollary~2.4.2]{G08}), after suitable redefinition on a set of measure zero. Since $f$ belongs to $L^p$ by hypothesis, we conclude that the polynomial must be identically zero, as claimed in the theorem.
To show $\widehat{f}$ is supported at the origin, we start with a Schwartz function $\eta$ supported in ${{\mathbb R}^d} \setminus \{ 0 \}$. We must show $\widehat{f}[\eta]=0$, that is, $f[\widehat{\eta}]=0$.
\medskip
Write $\phi = \overline{\psi}$, so that (by hypothesis) $\widehat{\phi}$ is nontrivial in every direction. The proof proceeds in a number of steps.
\medskip
Step 1. [Cut-off function.] For each unit vector $\xi^\prime$, choose $r>0$ such that $\widehat{\phi}(r\xi^\prime) \neq 0$. By continuity of $\widehat{\phi}$, there exists a neighborhood $\Xi$ of $\xi^\prime$ on the unit sphere and a number $s>1$ such that $\widehat{\phi}(q\xi) \neq 0$ for all $\xi \in \Xi$ and all $q \in [r,sr]$. Cover the sphere with finitely many such neighborhoods $\Xi_1,\dots,\Xi_n$ having corresponding values $r_1,\dots,r_n$ and $s_1,\dots,s_n$.
Choose a nonnegative function $\lambda \in C^\infty_c({{\mathbb R}^d} \setminus \{ 0 \})$ such that $\lambda(q\xi)>0$ whenever $q \in [r_k,s_k r_k], \xi \in \Xi_k, k=1,\dots,n$; for example, one could take $\lambda$ to be a radially symmetric ``annular bump'' function that is zero near the origin and positive from radius $\min_k r_k$ out to radius $\max_k s_k r_k$.
\medskip
Step 2. [Satisfying the Calder\'{o}n condition.] Take $s = \min(s_1,\dots,s_n)$, and define a Schwartz function $\mu$ by letting its Fourier transform be
\[
\widehat{\mu}(\omega) = \frac{ \overline{\widehat{\phi}(\omega)} \lambda(\omega)}{ \sum_{j \in \mathbb{Z}} |\widehat{\phi}(s^j \omega)|^2 \lambda(s^j \omega)} , \qquad \omega \in {{\mathbb R}^d} \setminus \{ 0 \} .
\]
Note the term with $j=0$ in the denominator is positive at every point $\omega= q\xi$ with $q \in [r_k,s r_k], \xi \in \Xi_k, k=1,\dots,n$, and so by summing over $j$ we see the denominator is positive for every $\omega \neq 0$. Further, the series in the denominator converges because it involves only finitely many $j$ values (for each $\omega$), due to the compact support of $\lambda$ in ${{\mathbb R}^d} \setminus \{ 0 \}$.
Hence
\[
\sum_{j \in \mathbb{Z}} \widehat{\phi}(s^j \omega) \widehat{\mu}(s^j \omega) = 1 , \qquad \omega \in {{\mathbb R}^d} \setminus \{ 0 \} .
\]
\medskip
Step 3. [Calder\'{o}n reproducing formula.] Multiplying the result of Step 2 by $\eta(-\omega)$ shows that
\[
\sum_{j \in \mathbb{Z}} \widehat{\phi}(s^j \omega) \widehat{\mu}(s^j \omega) \eta(-\omega) = \eta(-\omega) .
\]
Only a finite range of $j$-values (independently of $\omega$) is needed in the sum, because both $\eta$ and $\widehat{\mu}$ have compact support in ${{\mathbb R}^d} \setminus \{ 0 \}$.
Applying the inverse Fourier transform gives a convolution formula:
\[
\sum_{j \in \mathbb{Z}} \phi_{s^j} * \mu_{s^j} * \widehat{\eta} = \widehat{\eta} ,
\]
where we use the notation $\phi_{s^j}(t)=\phi(t/s^j)/s^{jd}$ and so on. Convergence of the sum is guaranteed, because only finitely many $j$-values are summed.
\medskip
Step 4. [Applying the reproducing formula.] Hence the distribution $f$ acts on the Schwartz function $\widehat{\eta}$ according to
\begin{align*}
f[\widehat{\eta}]
& = \int_{{\mathbb R}^d} f(x) \widehat{\eta}(x) \, dx \\
& = \sum_{j \in \mathbb{Z}} \int_{{\mathbb R}^d} f(x) (\phi_{s^j} * \mu_{s^j} * \widehat{\eta})(x) \, dx \qquad \text{by Step 3} \\
& = \sum_{j \in \mathbb{Z}} \int_{{\mathbb R}^d} \int_{{\mathbb R}^d} f(x) \phi_{s^j}(x-t) \, (\mu_{s^j} * \widehat{\eta})(t) \, dt dx \\
& = \sum_{j \in \mathbb{Z}} \int_{{\mathbb R}^d} s^{-jd/2} \langle f , \psi_{s^j,t} \rangle \, (\mu_{s^j} * \widehat{\eta})(t) \, dt \\
& = 0 ,
\end{align*}
since $\langle f , \psi_{s^j,t} \rangle = 0$ by the hypothesis that the wavelet transform vanishes.
Thus the distribution $\widehat{f}$ is supported at the origin, as we needed to show.
\end{proof}
\subsection*{Comment on the literature.} The construction of $\mu$ in Steps 1 and 2 originated in Calder\'{o}n's work on his reproducing formula \cite{C64}. The discrete form used above (involving sums rather than integrals, over the scales) is a special case of a result by Str\"{o}mberg and Torchinsky \cite[Chapter V, Lemma 6]{ST89} (and see also \cite{H74}).
Both the continuous and discrete Calder\'{o}n reproducing formulas play an important role in the characterization of classical function spaces in mathematical analysis \cite{BPT96,BPT97,FJW91,W10}.
\section{\bf Uniqueness for bounded or polynomially bounded signals}
Next we treat signals in $L^\infty$. Uniqueness of bounded signals will hold only up to additive constants. More generally, we handle signals that grow polynomially.
\begin{thm}\label{bounded}
Assume $f$ is locally integrable with at most polynomial growth, meaning $f(x)(1+|x|)^{-k} \in L^\infty({{\mathbb R}^d})$ for some nonnegative integer $k$. Assume $\psi$ is integrable with $\psi(x)(1+|x|)^k \in L^1({{\mathbb R}^d})$ and $\widehat{\psi} \in C^\infty({{\mathbb R}^d} \setminus \{ 0 \})$, and suppose the wavelet transform vanishes identically:
\[
\langle f , \psi_{s,t} \rangle = 0 \qquad \text{for all $s>0 , \quad t \in {{\mathbb R}^d}$.}
\]
If $\widehat{\psi}$ is nontrivial in every direction, then $f$ is equal almost everywhere to a polynomial of degree $\leq k$. For example, if $f$ is bounded ($k=0$) then $f$ must be constant.
\end{thm}
\begin{proof}[Proof of Theorem~\ref{bounded}]
The proof proceeds exactly as for Theorem~\ref{p-int}, except that in the first paragraph of the proof, we do not know $f$ belongs to $L^p$ and so we cannot conclude the polynomial is identically zero. Instead, we simply conclude the polynomial has degree at most $k$.
\end{proof}
\subsection*{Vanishing moments} The conclusion of the theorem forces the wavelet $\psi$ to have vanishing moments. For example, if $f$ is bounded (the case $k=0$) then $f$ is constant by the theorem; and so either $f$ is identically zero or else $\psi$ has integral zero,
\[
\int_{{\mathbb R}^d} \psi(x) \, dx = 0,
\]
because of the hypothesis that $\langle f , \psi_{1,0} \rangle = 0$.
For higher moments, let us consider the $1$ dimensional case and write the polynomial $f$ as $f(x)=\sum_{\ell=0}^m c_\ell x^\ell$ for some coefficients $c_\ell$, where $m=\deg(f)$. Suppose $f$ is not identically zero, so that the leading coefficient is nonzero, $c_m \neq 0$. Then $\psi$ has vanishing moments up to order $m$, meaning
\[
\int_{\mathbb R} x^\ell \psi(x) \, dx = 0 , \qquad \ell = 0,1,\dots,m,
\]
we now show. From $\langle f , \psi_{1,t} \rangle = 0$ we deduce that $\sum_{\ell=0}^m c_\ell \int_{\mathbb R} (x+t)^\ell \, \overline{\psi(x)} \, dx = 0$. Differentiating $m$ times with respect to $t$ and then setting $t=0$ shows that $\int_{\mathbb R} \psi(x) \, dx = 0$, since $c_m \neq 0$. Differentiating $m-1$ times with respect to $t$ and setting $t=0$ then shows that $\int_{\mathbb R} x \psi(x) \, dx = 0$. Repeating this argument down to the $0$th derivative establishes the claimed vanishing moments.
\smallskip
We conclude by extending the uniqueness result to signals that are general tempered distributions.
\begin{thm}\label{distribution}
Assume $f$ is a tempered distribution, $\psi$ is a Schwartz function, and the wavelet transform vanishes identically:
\[
\langle f , \psi_{s,t} \rangle = 0 \qquad \text{for all $s>0 , \quad t \in {{\mathbb R}^d}$.}
\]
If $\widehat{\psi}$ is nontrivial in every direction, then $f$ is a polynomial.
\end{thm}
The proof requires only a rephrasing into distributional language of Step 4 in the proof of Theorem~\ref{p-int}. We leave this task to the reader.
\section*{Acknowledgments} We thank Y. Sun and M. Sundararajan for asking us about the continuous wavelet uniqueness problem, and explaining its implications in economic problems.
Laugesen thanks the Department of Mathematics and Statistics at the University of Canterbury, New Zealand, for hosting him during this research.
\vspace*{12pt}
|
\section*{Another Section}
Sections can only be used in Articles. Contributions should be
organized in the sequence: title, text, methods, references,
Supplementary Information line (if any), acknowledgements,
interest declaration, corresponding author line, tables, figure
legends.
Spelling must be British English (Oxford English Dictionary)
In addition, a cover letter needs to be written with the
following:
\begin{enumerate}
\item A 100 word or less summary indicating on scientific grounds
why the paper should be considered for a wide-ranging journal like
\textsl{Nature} instead of a more narrowly focussed journal.
\item A 100 word or less summary aimed at a non-scientific audience,
written at the level of a national newspaper. It may be used for
\textsl{Nature}'s press release or other general publicity.
\item The cover letter should state clearly what is included as the
submission, including number of figures, supporting manuscripts
and any Supplementary Information (specifying number of items and
format).
\item The cover letter should also state the number of
words of text in the paper; the number of figures and parts of
figures (for example, 4 figures, comprising 16 separate panels in
total); a rough estimate of the desired final size of figures in
terms of number of pages; and a full current postal address,
telephone and fax numbers, and current e-mail address.
\end{enumerate}
See \textsl{Nature}'s website
(\texttt{http://www.nature.com/nature/submit/gta/index.html}) for
complete submission guidelines.
\begin{methods}
Put methods in here. If you are going to subsection it, use
\verb|\subsection| commands. Methods section should be less than
800 words and if it is less than 200 words, it can be incorporated
into the main text.
\subsection{Method subsection.}
Here is a description of a specific method used. Note that the
subsection heading ends with a full stop (period) and that the
command is \verb|\subsection{}| not \verb|\subsection*{}|.
\end{methods}
\end{comment}
|
\section{Introduction}
Coronal rotation can be observed through various solar tracers at different frequencies, like coronal green line (Fe XIV emission line at 530.3 nm ), white light, He I line (at 1083 nm), soft X-rays, UV rays, radio waves. The coronal green line has been used to measure the rotation rate of the solar corona at higher latitudes by \citet{Waldemier50, Trellis57, Cooper62, Sykora71, Sime89, Rybak94, Badalyan06a, Badalyan06b} and others. The results of \citet{Waldemier50} and \citet{Cooper62} indicate a faster rate of rotation as compared to the rate of rotation of the sunspots, suggesting a much lower differential rotation rate in the corona. In his work on green corona \citet{Sykora71} found that the Sun shows little or no differential rotation for six latitudinal zones $\pm 7.5$, $\pm 27.5$ and $\pm 47.5$. For low latitudes the rotation period was near to that found by \citet{Trellis57}. The green (Fe XIV at 530.3 nm) emission line for the period 1973 - 2000 and red (Fe X at 637.4 nm) emission line for the period 1984 - 2000 were analyzed by \citet{Altrock97, Altrock03}. It was reported that the corona, at green and red emission lines, shows more rigid rotation than does the photosphere. \citet{Sime89} also concluded, after analyzing the Sacramento Peak Observatory (SPO) data observed between 1973 and 1985, that the Fe XIV corona rotates more rigidly than do features in the photosphere or chromosphere. The synodic period obtained by \citet{Rybak94} for the period 1964 - 89 again confirmed the differential rotation of the green corona. \citet{Badalyan06a, Badalyan06b} carried out a comprehensive analysis using a long database (1939 - 2001) on the brightness of the coronal green line. The results support previous conclusions that the differential rotation in the corona is less pronounced than in photospheric tracers.
\begin{figure*}
\centering{
\includegraphics[width=\textwidth]{Figure1.eps}}
\caption{The annual mean of solar radio flux at 2.8 GHz is compared with the annual mean of sunspot numbers for the period 1947-2009.}
\label{srf1}
\end{figure*}
\citet{Hansen69} used the K-coronometer for coronal rotation measurement at different latitudes, for heights ranging from 1.125 to 2 $R_{\sun}$. The rotation found at the equator is in good agreement with the sunspot's rotation results and shows less variation with latitude at higher latitudes in comparison to the rotation of the chromosphere. A detailed study of the white light corona, from 1.1 $R_{\sun}$ to 30 $R_{\sun}$, was done with the \emph{LASCO} coronagraphs on board the \emph{SoHO} spacecraft. It was concluded that the rotation of the corona displayed a radially rigid rotation of 27.5 days synodic period from 2.5 $R_{\sun}$ to $>$ 15 $R_{\sun}$ \citep{Lewis99}.
The He I 1083 nm maps, from the National Solar Observatory (NSO), have been used to determine the rotation. It is found, both from observations and from magnetic extrapolation methods, that the corona becomes more rigid with height. By considering coronal holes as tracers (from an atlas of coronal holes mapped in He I 1083 nm data) of the differential rotation \citet{Insley95} demonstrated that the mid-latitude corona rotates more rigidly than the photosphere, but still exhibits significant differential rotation, with an equatorial rate of 13.30 $\pm$ 0.04 deg/day, and at 45\degr latitude a rate of 12.57 $\pm$ 0.13 deg/day. An analysis of the rotation of coronal holes spanning 18 years (from 1973 to 1991) was done based on data from the Catalogue of Coronal Holes \citep{Navarro94}. Isolated coronal holes showed a typical differential rotation, but polar coronal holes extensions displayed two different types of behavior: a rotation rate below approximately $40\degr \pm 5\degr$ of heliographic latitude, increasing to the equator, and a rotation rate above the same heliographic latitude but increasing to the poles.
Coronal holes, as observed from the \emph{Skylab} and \emph{Yohkoh} spacecrafts, have also been used to determine the rotation rate of the outer corona. Soft X-ray (SXR) observations of an elongated coronal hole, shows the almost rigid rotation of the coronal hole \citep{Timothy75, Kozuka94}. The solar full disc (SFD) images, obtained by soft X-ray telescope (SXT) of \emph{Yohkoh} space observatory, were used by different scientific groups to study the rotation rate of the corona. \citet{Weber99, Weber02} and \citet{Chandra10} concluded, after analyzing SXT data by different methods, that the rotation profile of the corona across the latitude is shallower than the rotation profile of its lower atmospheric levels. \citet{Kariyappa08} tracked the X-ray bright points (XBPs) on SFD images observed through the SXT and XRT (X-ray telescope) on board the \emph{Yohkoh} and \emph{Hinode} spacecrafts, respectively. \citet {Kariyappa08} found, contrary to all expectations, that the corona rotates differentially with respect to latitude, as in the case of the photosphere and the chromosphere.
\citet{Karachik06} analyzed the coronal bright points (CBPs) on SFD filtergrams observed through \emph {SoHO}/EIT (Fe XII line at 19.5 nm ) and reported that the rotation of CBPs closely follows the latitudinal rotation profile of the photospheric magnetic field. It was also shown that coronal features at different heights in the corona exhibits different rotation rates. \citet{Brajsa02, Brajsa04, Mulec07} and \citet{Brajsa08} determined the solar differential rotation by tracing coronal bright points on SFD filtergrams observed through \emph {SoHO}/EIT (Fe XV line at 28.4 nm ). For the declining phase of solar cycle 23, \citet{Zaatri09} compared differential rotation of subphotospheric layers derived from GONG++ dopplergrams with the small bright coronal structures (SBCS) observed through \emph {SoHO}/EIT . It is found at the equator that the SBCS rotate faster than the upper subphotospheric layer (3Mm) by about 0.5 deg/day. The latitude gradient of the rotation rate of the SBCS and the subphotospheric layers were found in agreement.
\citet{Kane01, Vats01} used the disk integrated solar radio fluxes for the period 1997 to 1999 at different frequencies to study the coronal rotation at different heights. \citet{Vats01} shown that coronal rotation depends on the heights in the corona. The autocorrelation analysis of radio SFD images at 17 GHz, for the period 1999 - 2001, also ascertained that the solar corona rotates less differentially than does the photosphere and the chromosphere \citep{Chandra09}.
\citet{Rybak94} reported that the differential rotation of the green corona shows no clear cyclic variation of the rotation. \citet {Kariyappa08}, who tracks the XBPs on the SFD images of \emph{Yohkoh}/SXT and \emph{Hinode}/XRT, also found that the rate of coronal rotation is independent of the phases of the solar activity cycle. But, from the GONG (Global Oscillation Network Group) and MDI (Michelson Doppler Imager) data, \citet{Jain01}, have shown that the equatorial rotation rates show variation with the solar activity cycle. \citet{Badalyan06a, Badalyan06b} also reported that there is periodic variation in the rate of differential rotation of green corona which is related to the solar activity cycle, a correlation usually found in photospheric tracers \citep{Hathaway90}. With the SFD filtergrams obtained from \emph{SoHO}/EIT (Fe XV at 28.4 nm), \citet{Brajsa04} exhibited that the differential rotation profile corresponds roughly to the rotation of sunspot groups. \citet{Mulec07} and \citet{Brajsa08} found, with the same data of different epoch, that the equatorial rotation velocity and the gradient of differential rotation show similar values during periods of low and high activity.
\citet{Mouradian02} analyzed the 2.8 GHz radio emission flux with the maximum entropy method and showed that the 2.8 GHz radio emission rotation varies according to the activity level. But \citet{Mehta05}, after analyzing the radio emission data could not find any systematic relationship between the coronal rotation period and the phase of the solar cycle. \citet{Chandra09, Chandra10} calculated the equatorial rotation rate for each year separately (using radio images at 17 GHz for the years 1999-2001 and SXT data for the years 1992-2001, respectively) and compared it with the annual sunspot numbers. The comparison shows that the equatorial rotation rate seems largely to be a function of the phases of the solar cycle.
\citet{Javaraiah00} used the GPR (Greenwich Photoheliographic Results) data on sunspot groups compiled during 1879 - 1975 to study the variations of the differential rotation coefficients $A$ and $B$ during the odd numbered solar cycles (ONSCs) and during the even numbered solar cycles (ENSCs). The parameters $A$ and $B$ are measures of the equatorial rotation rate and latitude gradient of rotation rate, respectively. It is seen that the variation in $A$ is significant only in the ONSCs (Waldmeier cycle number 13, 15, 17 and 19). Whereas, the variation in $B$ is quite significant in both ONSCs and ENSCs (12, 14, 16, 18 and 20). There exists a good anticorrelation between the mean variation of $B$ during ONSCs and ENSCs, suggesting the existence of a 22-years periodicity in $B$.
Different methods and different tracers have been employed to determine the coronal rotation and its dependence on the phases of the solar cycle, but the estimates were not found to be in agreement with each other. Using radio measurements to estimate solar rotation is not very new, but since the last couple of decades many scientific groups have employed radio flux data at various frequencies to ascertain the rotation period \citep{Vats98a, Kane01, Mouradian02}, its differentiality as a function of altitude \citep{Vats01} and latitude \citep{Chandra09} and its relationship with the phases of the solar cycle \citep{Mehta05}. The present study is an attempt to estimate the coronal rotation rate and its relationship with the solar magnetic cycle by analyzing the time series of the radio emission data at 2.8 GHz frequency (wavelength 10.7 cm) for the period 1947 - 2009. The data covers almost six solar activity cycles and is currently one of the best and longest data of solar indices we have after the sunspot number. This radio emission at 2.8 GHz frequency originates deep in the solar corona and hence represents coronal rotation \citep{Vats01}.
The next section discusses the observation and importance of radio flux data at 2.8 GHz used in the present analysis. Its subsection briefly describes the autocorrelation and crosscorrelation analysis technique, primarily used to determine the rotation period and its correlation with the solar activity, respectively. The paper ends with a discussion of the results obtained.
\section{Data}
The data used in the present work comprises disc integrated flux measurements of radio waves at 2.8 GHz. It is expressed in solar flux units ($1$ SFU =$10^{-22}$ Watts/meter square-Hertz). It is almost thermal in origin because emission is mainly due to thermal free-free (bremsstrahlung) radiation \citep{Tapping90}. It is directly related to the total amount of magnetic flux. The data measurements have been continued through the Solar Radio Monitoring Programme, a service which is operated by the National Research Council of Canada. The solar flux density at 2.8 GHz is measured using two radio telescopes (called flux monitors). The two instruments record the strength of the solar radio emission at 2.8 GHz frequency each day for as long as the Sun is above the horizon. Between 1946 and 1990, the measurements were made in the Ottawa area at the Algonquin Radio Observatory. In 1990, following the closure of that observatory, the system was relocated to the Dominion Radio Astrophysical Observatory, near Penticton, British Columbia. It now forms a consistent, uninterrupted data base covering more than 63 years. The observations, which started in late 1946, are still continuing. The continuity and consistency of calibration of these measurements of the 2.8 GHz radio flux have helped to establish it as an internationally accepted index of solar activity.
These data are readily available through the National Geophysical Data Center (NGDC), NOAA, USA. The data
available at the site of NGDC are tabulated in two forms; the observed flux and the adjusted flux. The observed flux data are the actual measured values and are affected by the changing distance between the Earth and Sun, whereas the adjusted flux data are scaled to the standard distance of 1 AU. Therefore, the adjusted flux data are considered to be more appropriate for the study of the behavior of the solar corona. The annual mean of the daily measurements of the radio emission can be represented in the form of a time series (shown by a solid line in Figure \ref{srf1}). This figure shows the yearly variation of sunspots number as well as the radio emission at 2.8 GHz.
\citet{Covington69} made a comparison over more than a solar activity cycle and showed that there is indeed a linear correlation between the 2.8 GHz solar flux and the total photospheric magnetic flux in the active region. Various manifestations of solar activity are driven by the total amount of magnetic flux emerging through the photosphere into the chromosphere and the corona, and its spatial and temporal distribution. Besides this, the 2.8 GHz solar flux correlates well with some other indices of solar activity such as sunspot number and total sunspot area. The radio measurements have an advantage over other indices in that these measurements are completely objective, and can be made under almost any weather conditions. Since they are closely correlated with the magnetic activity which modulates the Sun's energy output with solar irradiance, they correlate closely with other solar activity indices \citep{Covington69}.
The fractal analysis of the solar radio emissions in the frequency range from 245 MHz to 15 GHz has established
that the 2.8 GHz is the most appropriate frequency for the study of the solar coronal rotation \citep{Vats98b, Mehta05}. It is so because the fractal dimension is found to be the least at a frequency near 3 GHz. Therefore, according to \citet{Vats98b} and \citet {Mehta05}, the time series of radio emission data available at 2.8 GHz frequency must have a very strong modulation due to the solar coronal rotation, in comparison to other frequencies. Thus, they used radio emission at 2.8 GHz frequency to determine the coronal rotation period. We, therefore, used this radio frequency to study long term variation in the coronal rotation period, so obtained.
From the knowledge of the yearly mean sunspot numbers (International Sunspot Number, credit: SIDC, 1947-2009) on the photosphere, the years 1954, 1964, 1976, 1986, 1996 and 2009 are the years of solar minima in the solar cycles numbered as 18, 19, 20, 21, 22 and 23, respectively (Figure \ref{srf1}). In fact the year 2006 would have been a solar minima and the number of sunspots would have increased after that, but surprisingly in cycle 23 the minima has become a much extended one. The sunspots number showed a decreasing trend up to the end of the year 2009. It is only now in 2010 that solar activity shows signs of increase. We can see that the radio flux at 2.8 GHz frequency clearly exhibits the 11-years cycle performed generally by the sunspots number on the photosphere. It is well established evidence of a good correlation between the sunspots and the radio emission.
\section{Methods}
There are many statistical methods to analyze such periodic time series. In the present work, the autocorrelation and crosscorrelation method are employed for the analysis of the time series of radio flux at 2.8 GHz.
\begin{figure*}
\centering{
\includegraphics[width=154mm]{Figure2.eps}}
\caption{A few typical time series of solar radio flux at 2.8 GHz (left panel) for the years of solar maxima (1947, 1969 and 1989) and minima (1986, 1996 and 2009) and their respective plots of autocorrelation function used in the present work for the estimation of the synodic rotation period (right panel).}
\label{srf2}
\end{figure*}
\subsection{Autocorrelation Analysis}
A time series is defined here as a sequential collection of data observations indexed over time. The series of autocorrelation coefficients represents the correlation between the successive observations of a single time series. The daily measurement of radio flux at 2.8 GHz frequency can be represented in the form of a time series, ($x_0, x_1, x_2, x_3, . . . , x_{n-1}$).
Now we can form $(n-l)$ pairs of observations from this series, where $n$ is the length of the data series (365/366 days in our analysis) and $l$ is referred to as the lag. They are $(x_0, x_l), (x_1, x_{1+l}), (x_2, x_{2+l}), . . . , (x_{n-l-1}, x_{n-1})$.
By regarding the first observation on each pair as one variable and the second observation on each pair as the second variable the autocorrelation coefficient $P_{x}(l)$ as a function of the lag $l$ is given by
\begin{equation}
P_{x}(l) = P_{x}(-l) = \frac{\sum^{n-l-1}_{k=0} (x_{k} -\overline{x})(x_{k+l} -\overline{x})}{\sum^{n-l}_{k=0} (x_{k} -\overline{x})^2}
\end{equation}
\begin{flushleft}
where $\overline{x}$ is the mean of the daily observations.
\end{flushleft}
The daily solar flux at 2.8 GHz is converted into time series of one year each (total 63 data sets) for the period 1947-2009, each time series beginning with the first day of each calendar year (except for the year 1947, because the data available for this year is from Feb 14 onwards). In this way, each time series of one year length can encompass several cycles of coronal rotation. The data set is almost continuous through out the period of study. We use linear interpolation to fill the few data gaps (less than 1\%) found in some data sets. A few typical examples of such time series are shown in the left panel of the Figure \ref{srf2}. The rotational modulation due to the solar rotation can be noticed in each of the time series. The time series (shown in the left panel of Figure \ref{srf2}) are particularly chosen from the years of solar maxima (1947, 1969 and 1989) and solar minima (1986, 1996 and 2009). The significant rotational modulation of the autocorrelation function in most of the years is clear display of few nice bright emissive tracers phased well in the longitudes. In some cases more pronounced autocorrelation function can appear even during minima of the solar cycle, if only one isolated (or only very few) active region is on the solar disc for several months.
The autocorrelation function for such time series of the solar radio flux at 2.8 GHz of each year is obtained. The right panels of Figure \ref{srf2} show six such plot of autocorrelation function out of the 63 obtained in the same manner as discussed above. The autocorrelation function shown here corresponds to the years of solar maxima (1947, 1969 and 1989) and solar minima (1986, 1996 and 2009). It can be observed from Figure \ref{srf2} that most of the autocorrelation function curves show a very smooth and cyclic rotational modulation with a fair amount of correlation.
The synodic coronal rotation period can be estimated from the position of the first secondary maximum. To determine the synodic rotation period with maximum possible accuracy, we fitted a sinusoidal wave whose amplitude is the same as first secondary maximum in the autocorrelation function curve and period accurate to a day. Then we considered 7 points (at interval of $\pm 1$ day) around the first secondary maximum of both the sine curve and autocorrelation curve and calculate the standard deviation. Next, we changed the period of sine wave in small steps (in the steps of 0.1 day) and tried to minimize the standard deviation. The value of wave period for which the standard deviation is found to be least would be the synodic rotation period. This fitting allowed us to determine the lag with the sub-day precision and to estimate the uncertainty of the lag determination, which is 0.1 day in present case.
The synodic rotation periods are then converted into sidereal rotation periods using the following expression.
\begin{equation}
T_{sidereal}=\frac{365.26 \times T_{synodic}}{365.26+T_{synodic}}
\end{equation}
This equation does an approximate conversion of synodic to sidereal rotation period. For more accurate conversion one may use the approach outlined by \citet{Rosa95}. Autocorrelation analysis can thus be used to advantage for investigating the periodical variations of time series, particularly the long ones. As we know that the indices of the Sun, derived from the observations or analysis of various solar activity phenomena, also form time series of such type. It is often used with the lag to determine the stationarity of a time series. This analysis is useful in predicting trends or cyclical functions present in a time series. This technique was first employed by \citet{Hansen69} to determine the synodic rotation period of the solar electron corona after analyzing the K-coronameter observations during 1964-1967. Later \citet{El-raey72} analyzed sunspot number, solar radio flux and interplanetary field for the period 1967 to 1970 to determine the solar atmospheric rotation. \citet{Parker82} showed that, within the limit of uncertainty, maximum entropy spectral analysis and autocorrelation analysis provide almost similar results. \citet{Fisher84, Sime89} applied the same analysis technique in their study of the rotational characteristics of white-light and Fe XIV corona, respectively. The differential rotation rates of soft X-ray features in the solar corona were determined by \citet{Weber99} using the Lomb-Scargle periodogram. These results were then compared with the results obtained through the autocorrelation method and were found to agree within $1\sigma$. \citet{Vats98a, Vats98b, Vats01} used the autocorrelation technique to determine the rotation period and the differential rotation as a function of height in the solar corona, respectively.
\begin{figure*}
\centering{
\includegraphics[width=\textwidth]{Figure3.eps}}
\caption{Temporal variation of the sidereal rotation period obtained using autocorrelation analysis of the solar radio emissions at 2.8 GHz during 1947-2009. The smoothed curve (eleven years running mean) of the sidereal rotation period is also displayed to compare it with its raw results, and also to see, if any long term component is present in it.}
\label{srf3}
\end{figure*}
\subsection{Crosscorrelation Analysis}
Crosscorrelation can be used to determine the degree of fit between two data sets $x$ and $y$. The day-to-day measurement of annual sunspot number and coronal rotational period, obtained from the analysis of the 2.8 GHz radio flux of the same time interval, can be represented in the form of two time series, $(x_0, x_1, x_2, x_3, . . . . ., x_{n-1})$ and $(y_0, y_1, y_2, y_3, . . . . ., y_{n-1})$.
Again, we can form $(n - l)$ pairs of observations from these two series. They are: $(x_0, x_l)$, $(x_1, x_{1+l})$, $(x_2, x_{2+l})$, . . . . ., $(x_{n-l-1}, x_{n-1})$ and $(y_0, y_l)$, $(y_1, y_{1+l})$, $(y_2, y_{2+l})$, . . . . ., $(y_{n-l-1}, y_{n-1})$.
The crosscorrelation coefficient $P_{xy}(l)$ of the two time series as a function of the lag $l$ is given by,
\begin{equation}
P_{xy}(l) = \frac{\sum^{n-\left|l\right|-1}_{k=0} (x_{k+\left|l\right|} -\overline{x})(y_{k} -\overline{y})}{\sqrt{{\left[\sum^{n-l}_{k=0} (x_{k} -\overline{x})^2\right]}{\left[\sum^{n-l}_{k=0} (y_{k} -\overline{y})^2\right]}}}
\end{equation}
\begin{equation}
P_{xy}(l) = \frac{\sum^{n-l-1}_{k=0} (x_{k} -\overline{x})(y_{k+l} -\overline{y})}{\sqrt{{\left[\sum^{n-l}_{k=0} (x_{k} -\overline{x})^2\right]}{\left[\sum^{n-l}_{k=0} (y_{k} -\overline{y})^2\right]}}}
\end{equation}
\begin{flushleft}
for $l < 0$ and $l > 0$, respectively. Here, $\overline{x}$ and $\overline{y}$ is the mean of the daily observations of two different time series.
\end{flushleft}
The cross-correlation analysis helps in determining the mutual correlation between solar rotation and other observed indices of the Sun as well as other stars. Spectroheliograms obtained in extreme ultraviolet (EUV) lines and the Lyman continuum are used to determine the rotation rate of the solar chromosphere, transition region, and corona \citep{Dupree72, Henze73}. A cross-correlation analysis of the observations indicates the presence of differential rotation through the chromosphere and the transition region. \citet{Bocchialini95} used this method for the diagnostics of the chromospheric dynamics. The correct algorithm to calculate the cross-correlation functions of the He I and the O V line intensities was proposed by \citet{Gomory04}.
\begin{table*}
\centering
\begin{minipage}{130mm}
\caption{Sidereal rotation period (in days) obtained through autocorrelation analysis of daily radio flux at 2.8 GHz for the years 1947-2009.}
\begin{tabular}{@{}llllllllll@{}}
\hline
Year &Sidereal&Year &Sidereal&Year &Sidereal&Year &Sidereal&Year &Sidereal\\
&Rotation& &Rotation& &Rotation& &Rotation& &Rotation\\
& Period& & Period& & Period& & Period& & Period\\
\hline
1947 & 24.5 & 1960 & 23.8 & 1973 & 24.4 & 1986 & 23.8 & 1999 & 25.1 \\
1948 & 22.6 & 1961 & 25.5 & 1974 & 23.9 & 1987 & 21.7 & 2000 & 22.6 \\
1949 & 25.3 & 1962 & 25.8 & 1975 & 23.4 & 1988 & 26.1 & 2001 & 22.3 \\
1950 & 22.6 & 1963 & 25.8 & 1976 & 26.4 & 1989 & 26.1 & 2002 & 22.7 \\
1951 & 25.3 & 1964 & 26.5 & 1977 & 20.0 & 1990 & 24.7 & 2003 & 25.7 \\
1952 & 23.2 & 1965 & 29.1 & 1978 & 23.4 & 1991 & 23.1 & 2004 & 25.6 \\
1953 & 24.7 & 1966 & 24.4 & 1979 & 25.4 & 1992 & 24.2 & 2005 & 24.5 \\
1954 & 23.4 & 1967 & 26.4 & 1980 & 24.1 & 1993 & 24.4 & 2006 & 23.4 \\
1955 & 21.7 & 1968 & 22.3 & 1981 & 26.8 & 1994 & 20.8 & 2007 & 29.5 \\
1956 & 24.4 & 1969 & 24.6 & 1982 & 24.9 & 1995 & 24.3 & 2008 & 23.5 \\
1957 & 27.0 & 1970 & 19.0 & 1983 & 24.1 & 1996 & 21.7 & 2009 & 24.4 \\
1958 & 28.1 & 1971 & 23.3 & 1984 & 25.1 & 1997 & 21.6 & \textbf{Mean} & 24.3 \\
1959 & 25.2 & 1972 & 25.4 & 1985 & 24.4 & 1998 & 20.9 & \textbf{std. dev.}& 02.0\\
\hline
\end{tabular}
\end{minipage}
\label{tab1}
\end{table*}
\section{Results}
The sidereal rotation period obtained from the time series of the radio flux of the years 1947 to 2009 are tabulated in Table \ref{tab1}. A plot is also given in the Figure \ref{srf3} (by scattered dots), which shows the sidereal rotation period as a function of the year. It is found that the mean sidereal rotation period during 1947-2009 is equal to 24.3 days. The maximum rotation period is found to be 29.5 days in the year 2007 and the minimum rotation period is 19.0 days in 1970. The precision is same in the determination of the sidereal rotation period throughout the period of study and is 0.1 day. The standard deviation of the sidereal rotation period during 1947 to 2009 is calculated and is equal to 2.0 days.
These values of the coronal rotation period show quite a large variability (from 19.0 to 29.5 days). Such large variability in the rotation period has also been reported by \citet{Howard84} in their study of the collection of Mount Wilson white-light plates from 1921 through 1982. From their Table 2, it can be seen that the sidereal rotation rate varies between 11.89 degree/day (in 1965) to 16.47 degree/day (in 1963) in the equatorial region. It means that the sidereal rotation period obtained with white-light plates also shows a large variability from 30.3 to 21.9 days, which is in good agreement with the coronal rotation period obtained in present work. Ironically, the year of maximum and minimum rotation rate (1963 and 1965, respectively) are quite nearby, which again shows analogy with the present work. In some cases, there is a large variation in rotation period in adjacent years (shown in Figure \ref{srf3}).
\citet{Chandra} obtained the coronal rotation period, as a function of latitude, using Nobeyama Radioheliogram (NoRH) images at 17 GHz for the year 1999 through 2005. The sidereal period obtained for the year 2003 shows quite a large variability from 29.2 to 22.1 days. This is again in agreement with our results obtained analizing radio flux data at 2.8 GHz.
To see the presence of any long term variability in it, the coronal rotation period data is smoothen by taking the 3, 5, 7, 9, . . , 17 years running mean of the sidereal rotation period. The smoothed rotation periods so obtained are then plotted against the year. It is found, after comparing all such plots, that the running mean of eleven years smoothed the raw data most. By taking the running mean of eleven years, the short term variation present in the rotation period data sets are smoothed and thus removed. The data obtained after taking the running mean will now show only the long term variation, \textit{i.e.} with the period of more than 11 years. The plot is shown in Figure \ref{srf3} (by solid line). An evidence of long term periodic variation can be seen in this plot. The overall variation has reduced from $\sim 11$ days to only $\sim 3$ days, as shown in Figure \ref{srf3}. This plot has several peaks and dips. However, there are two major maxima and two major minima which can be easily identified in Figure \ref{srf3}. The separation of consecutive maxima or minima is $\sim 22$ years. This is an evidence of $\sim 22$ years periodicity in the temporal variation of sidereal coronal rotation.
This periodicity may be better seen in the plot of autocorrelation function. Hence, we derived the autocorrelation coefficient of raw and smoothed (an eleven years running mean) of the sidereal rotation period. These are plotted against the lag in days in Figure \ref{srf4}. The dashed and solid curves are for raw and smoothed sidereal rotation period, respectively. The autocorrelation function exhibits the existence of a cycle close to a period of 22-years. As expected this periodicity is better seen in the autocorrelation of smoothed data set.
The crosscorrelation function between the smoothed rotation period and sunspots number is calculated and shown in Figure \ref{srf5} by solid line. It is natural that from a crosscorrelation of these time series, we expect to see evidence of any variation having a period of more than 11 years. The plot of crosscorrelation function in Figure \ref{srf5} (solid line) varies from $\sim -0.30$ to 0.53. One can easily identify two maxima and two minima in this curve and estimate the period. From the positions of these we get a periodicity of $\sim 22$ years. This periodicity in sunspot numbers exists due to the 22-years magnetic cycle. The temporal variation of the coronal rotation period also seems to have a component of the same. However, as the central maximum of the crosscorrelation function fall very close to 0 lag (at -1 year), the smoothed coronal rotation period and sunspot numbers can considered to be almost in phase.
\begin{figure}
\centering{
\includegraphics[width=80mm]{Figure4.eps}}
\caption{Autocorrelation function curves of raw and smoothed annual sidereal rotation period for the period 1947-2008.}
\label{srf4}
\end{figure}
\begin{figure}
\centering{
\includegraphics[width=80mm]{Figure5.eps}}
\caption{Crosscorrelation function of smoothed data of annual sidereal rotation period and sunspots numbers.}
\label{srf5}
\end{figure}
\section{Discussion}
Based on the model calculations, we know that this radio emission at 2.8 GHz originate from the lower solar corona near $\sim 60,000$ km above the solar surface \citep{Vats01}. Thus the present statistical analysis of solar flux at 2.8 GHz reveals some new long term rotational features of the solar corona at a height $\sim 60,000$ km above the photosphere. The results obtained through the autocorrelation and crosscorrelation analysis can be discussed in the following three parts.
\begin{enumerate}
\item
The temporal variation of the rotation period (Figure \ref{srf3}) does not show any systematic periodicity. On the other hand sunspots have a very clear 11 years periodicity. So from this, it seems that probably there is either no or very weak correlation between the coronal rotation period and the sunspot cycle.
\item
A reasonable presence of a long term variation in the coronal rotation period ( of $\sim$ 22-years) become visible simply by smoothing (11 point running mean) the values (Figure \ref{srf3}). This may correspond to the 22-years magnetic reversal cycle or Hale's cycle.
\item
The crosscorrelation analysis between the smoothed data of coronal rotation (Figure \ref{srf3}) and its correlation with the sunspots number also shows the clear presence of a 22-years Sun's magnetic field reversal cycle (Figures \ref{srf4} and \ref{srf5}).
\end{enumerate}
The position of minima in the Figure \ref{srf3} reveals that the years of minima of the 22-years cycle, obtained after smoothing the rotation period data, coincide well with the years of minimum activity of even numbered solar cycles (ENSCs), \textit{i.e.} at the years 1954, 1976 and 1996 (shown in Figure \ref{srf1}). Whereas, the years of maxima of the 22-years cycle matches well with the years of minimum activity (at the years 1964 and 1986) of odd numbered solar cycles (ONSCs). This result is in agreement with the results obtained by \citet{Javaraiah00}. The variation in equatorial rotation rate $A$ is found to be significant only in the years of ONSCs. Where $A$ is reciprocally related to the equatorial sidereal rotation period ($T$) by an expression ($T = 360/A$). It can be seen again in Figure \ref{srf3} (solid line) that the smoothed rotation period rises steeply in the ONSCs and falls back gradually in the ENSCs. A similar trend continues even in solar cycle 23. This shows a good agreement in both the results obtained through two entirely different methods and tracers.
The Magnetic butterfly diagram due to \citet{Hathaway98} shows clearly that the reversal of the Sun's magnetic polarity is triggered in the years close to the years of solar activity maxima, \textit{i.e.} years 1979, 1989 and 2000. But Figure \ref{srf3} indicates that the ascent or descent in the value of the smoothed sidereal rotation period is initiated in the years of solar minima. Therefore, it shows that there may be anticorrelation between the magnetic pole reversal and the rotation period.
\section*{Acknowledgments}
This work utilizes the radio flux data obtained through the Solar Radio Monitoring Programme, managed by the Dominion Radio Astrophysical Observatory, Penticton, British Columbia, Canada. The international sunspot number is produced by the Solar Influences Data Analysis Center (SIDC) at the Royal Observatory of Belgium. We are indebted to the observers who were involved in the acquisition of the useful solar indices used in the present work. The data (radio flux at 2.8 GHz and sunspot numbers) were acquired from the web page of the National Geophysical Data Center (NGDC). The authors are grateful to Dr. Kiran Biswas for his constructive comments regarding the manuscript. The authors wish to acknowledge the anonymous referee for his valuable suggestions for the improvement of the paper. The research at Physical Research Laboratory (PRL) is supported by the Department of Space, Government of India.
|
\section{INTRODUCTION}
An important result related to the improvement of the inequality
between arithmetic and geometric means (AM-GM) was obtained by D. I.
Cartwright and M. J. Field in \cite{CF}, which is given in the following
way: if $0<m=\min\{x_1,...,x_n\}$ and $M=\max\{x_1,...,x_n\}$, then
\begin{eqnarray}\label{ineq_1_1}
&\ds\frac{1}{2M}\sum_{i=1}^n\alpha_i\left(x_i-\sum_{k=1}^n \alpha_k
x_k\right)^2\leq \sum_{i=1}^n \alpha_i x_i-\prod_{i=1}^n
x_i^{\alpha_i}\leq&\notag\\
&\leq\ds \frac{1}{2m}\sum_{i=1}^n \alpha_i\left(x_i-\sum_{k=1}^n
\alpha_k x_k\right)^2,&
\end{eqnarray}
where $\alpha_i>0$ for all $i=1...n$ and $\alpha_1+...+\alpha_n=1$.
For $n=2$, this inequality may be written as follows:
\begin{equation}\label{ineq_1_2}
\frac{\lambda(1-\lambda)}{2M}(a-b)^2\leq \lambda
a+(1-\lambda)b-a^\lambda b^{1-\lambda}\leq
\frac{\lambda(1-\lambda)}{2m}(a-b)^2,
\end{equation}
where $a,b>0, \ m=\min\{a,b\}, \ M=\max\{a,b\}$ and
$\lambda\in[0,1]$.
Since $\ds\frac{\lambda(1-\lambda)}{2M}(a-b)^2\geq0$, we deduce
Young's inequality (see \cite{HLP,NP})
\begin{equation}\label{ineq_1_3}
a^\lambda b^{1-\lambda} \leq \lambda a+(1-\lambda)b
\end{equation}
Therefore, inequality (\ref{ineq_1_2}) is an improvement of Young's inequality
and at the same time gives a reverse inequality for the inequality
of Young.
In \cite{FM}, we presented two inequalities which give two different reverse inequalities for the Young's inequality, namely:
\begin{equation}\label{ineq_1_4}
0 \leq \lambda a+(1-\lambda) b -a^{\lambda}b^{1-\lambda} \leq
a^{\lambda}b^{1-\lambda}\exp\left\{\ds\frac{\lambda(1-\lambda)(a-b)^2}{m^2}\right\} -a^{\lambda}b^{1-\lambda}
\end{equation}
and
\begin{equation}\label{ineq_1_5}
0 \leq \lambda a+(1-\lambda) b -a^{\lambda}b^{1-\lambda} \leq
\lambda (1-\lambda)\left\{\log\left(\frac{a}{b}\right)\right\}^2M,
\end{equation}
where $a,b>0$, $m\equiv\min\{a,b\}$ and $M=\max\{a,b\}$ and
$\lambda\in[0,1]$.
\begin{rema}\label{rema_1_1}
The first inequality of (\ref{ineq_1_2}) clearly gives an improvement of the first inequality in (\ref{ineq_1_4}) and (\ref{ineq_1_5}).
For $0<a,b< 1$, we find the right hand side of the second inequality of (\ref{ineq_1_2}) gives tighter upper bound than that of (\ref{ineq_1_5}),
from the inequality $\frac{x-y}{\log x -\log y}<\frac{x+y}{2}$, for $x,y>0$.
For $a,b>1$, we find the right hand side of the second inequality of (\ref{ineq_1_5}) gives tighter upper bound than that of (\ref{ineq_1_2}),
from the inequality $\sqrt{xy}<\frac{x-y}{\log x -\log y}$, for $x,y>0$.
In addition, we find the right hand side of the second inequality of (\ref{ineq_1_2}) gives tighter upper bound than that of (\ref{ineq_1_4}) for $a,b>0$,
from $e^x > 1+ x$.
\end{rema}
Remark \ref{rema_1_1} supports the importance to study the inequality (\ref{ineq_1_2}) for several applications which will be given in the following sections.
\section{MAIN APPLICATIONS}
\begin{lema}
For $x>-1$ and $\lambda\in[0,1]$ there is the following inequality
\begin{equation}\label{ineq_2_1}
\frac{\lambda(1-\lambda)}{2M}x^2\leq \lambda x+1-(x+1)^\lambda \leq
\frac{\lambda(1-\lambda)}{2m}x^2,
\end{equation}
where $m=\min\{x+1,1\}$ and $M=\max\{x+1,1\}$.
\end{lema}
\noindent{\it Proof.} By replacing $\ds\frac{a}{b}$ of $t$ in
inequality (\ref{ineq_1_2}) we obtain the inequality
\begin{equation}\label{ineq_2_2}
\frac{\lambda(1-\lambda)}{2M}(t-1)^2\leq \lambda
t+1-\lambda-t^\lambda\leq \frac{\lambda(1-\lambda)}{2m}(t-1)^2,
\end{equation}
for all $t>0$ and $\lambda\in[0,1]$, where $m=\min\{t,1\}$ and
$M=\max\{t,1\}$.
Substituting $t=x+1$ in inequality (\ref{ineq_2_2}), we find the inequality
desired.
\hfill \qed
\begin{rema}
Inequality (\ref{ineq_2_1}) refines the inequality of Bernoulli, namely, for
$x>-1$ and $\lambda \in[0,1]$, we have
\begin{equation}
\lambda x+1\geq (x+1)^\lambda
\end{equation}
because $\ds\frac{\lambda(1-\lambda)}{2M}x^2\geq0$ in inequality
(\ref{ineq_2_1}).
\end{rema}
Next we will establish a refinement of the inequality between the
weighted power means, based on inequality (\ref{ineq_2_1}).
\begin{teo}
If $a_i>0, \ p_i>0, \ i=1...n, \ 0<r\leq s$,
$M_r(a,p)=\ds\left(\frac{\ds\sum_{i=1}^n p_i a_i^r}{\ds\sum_{i=1}^n
p_i}\right)^{1/r}$ and $M_s(a,p)=\left(\frac{\ds\sum_{i=1}^n p_i
a_i^s}{\ds\sum_{i=1}^n p_i}\right)^{1/s},$ then there is the
inequality
\begin{equation}\label{ineq_2_4}
\frac{A}{M}\leq[M_s(a,p)]^r-[M_r(a,p)]^r\leq\frac{A}{m},
\end{equation}
where
$$A=\frac{r(s-r)}{2s^2}[M_s(a,p)]^r\cdot\frac{\ds\sum_{i=1}^n p_i\left(\frac{a_i^s}{[M_s(a,p)]^s}-1\right)^2}{\ds\sum_{i=1}^n p_i},$$
$$m=\min_{i=\ov{1,n}}\left\{\frac{a_i^s}{[M_s(a,p)]^s},1\right\} \ {\rm and } \ M=\max_{i=\ov{1,n}}\left\{\frac{a_i^s}{[M_s(a,p)]^s},1\right\}.$$
\end{teo}
\noindent{\it Proof.} If $r=s$, then we have the equality in
relation
(\ref{ineq_2_4}). Let $r<s$.
In inequality (\ref{ineq_2_2}) we consider
$t=\ds\frac{a_i^s}{[M_s(a,p)]^s}$ and $\lambda=\frac{r}{s}<1$, thus,
we deduce the inequality
\begin{eqnarray}
\ds\frac{r(s-r)}{2s^2M}\left(\frac{a_i^s}{[M_s(a,p)]^s}-1\right)^2 &\leq& \frac{r}{s}\frac{a_i^s}{[M_s(a,p)]^s}+1-\frac{r}{s}-\frac{a_i^r}{[M_s(a,p)]^r} \nonumber \\
&\leq&\ds\frac{r(s-r)}{2s^2m}\left(\ds\frac{a_i^s}{[M_s(a,p)]^s}-1\right)^2.\label{ineq_2_5}
\end{eqnarray}
Multiplying by $p_i$ in inequality (\ref{ineq_2_5}) and taking the sum for
$i=1...n$, we obtain the following inequality
\begin{eqnarray*}
\frac{r(s-r)}{2s^2M}\frac{\ds\sum_{i=1}^n p_i\left(\frac{a_i^s}{[M_s(a,p)]^s}-1\right)^2}{\ds\sum_{i=1}^n p_i} &\leq & 1-\left[\frac{M_r(a,p)}{M_s(a,p)}\right]^r \\
&\leq& \frac{r(s-r)}{2s^2 m}\frac{\ds\sum_{i=1}^n p_i\left(\ds\frac{a_i^s}{[M_s(a,p)]^s}-1\right)^2}{\ds\sum_{i=1}^n p_i},
\end{eqnarray*}
which is equivalent to the inequality of the statement.
\hfill \qed
\begin{rema}
Since $\ds\frac{A}{M}\geq0$ in inequality (\ref{ineq_2_4}), we find the
inequality between the weighted power means \cite{HLP,NP},
\begin{equation}\label{ineq_2_6}
M_r(a,p)\leq M_s(a,p),
\end{equation}
for $0<r\leq s$.
The two means are equal if and only if $a_1=a_2=...=a_ n$.
\end{rema}
\begin{teo}
Let $p,q>1$ be real numbers satisfying
$\ds\frac{1}{p}+\frac{1}{q}=1$. If $a_i,b_i>0$ for all $i=1...n$,
then there is the following inequality
\begin{equation}\label{ineq_2_7}
\frac{A}{M}\leq\left(\sum_{i=1}^n
a_i^p\right)^{1/p}\left(\sum_{i=1}^n b_i^q\right)^{1/q}-\sum_{i=1}^n
a_ib_i\leq\frac{A}{m},
\end{equation}
where
$$A=\frac{1}{2pq}\left(\sum_{i=1}^n a_i^p\right)^{1/p}\left(\sum_{i=1}^n b_i^q\right)^{1/q}\sum_{i=1}^n \left(\frac{a_i^p}{\ds\sum_{i=1}^n a_i^p}-\frac{b_i^q}{\ds\sum_{i=1}^n b_i^q}\right)^2,$$
$$m=\min_{i=\ov{1,n}}\left\{\frac{a_i^p}{\ds\sum_{i=1}^n a_i^p}, \ \frac{b_i^q}{\ds\sum_{i=1}^n b_i^q}\right\} \ {\rm and } \ M=\max_{i=\ov{1,n}}\left\{\frac{a_i^p}{\ds\sum_{i=1}^n a_i^p}, \ \frac{b_i^q}{\ds\sum_{i=1}^n b_i^q}\right\}.$$
\end{teo}
\noindent{\it Proof.} By replacing $\lambda=\ds\frac{1}{p}, \
1-\lambda=\frac{1}{q}, \ a=\frac{a_i^p}{\ds\sum_{i=1}^n a_i^p}$;
$b=\ds\frac{b_i^q}{\ds\sum_{i=1}^n b_i^q}$ in inequality (\ref{ineq_1_2}) we
obtain the relation
$$\frac{1}{2pqM}\left(\frac{a_i^p}{\ds\sum_{i=1}^n a_i^p}-\frac{b_i^q}{\ds\sum_{i=1}^n b_i^q}\right)^2\leq \frac{a_i^p}{\ds p\sum_{i=1}^n a_i^p}+\frac{b_i^q}{q\ds\sum_{i=1}^m b_i^q}-\frac{a_i b_i}{\left(\ds\sum_{i=1}^na_i^p\right)^{1/p}\left(\ds\sum_{i=1}^n b_i^q\right)^{1/q}}$$
$$\leq \frac{1}{2pqm}\left(\frac{a_i^p}{\ds\sum_{i=1}^n a_i^p}-\frac{b_i^q}{\ds\sum_{i=1}^n b_i^q}\right)^2.$$
We observe that taking the sum for $i=1...n$ we deduce the
inequality of the statement.
\hfill \qed
\begin{rema}
\begin{itemize}
\item[(a)] H\"{o}lder's inequality is widely used in the theory of
inequalities and has the form \cite{HLP,NP}:
\begin{equation}\label{ineq_2_8}
\left(\sum_{i=1}^n a_i^p\right)^{1/p} \left(\ds\sum_{i=1}^n
b_i^q\right)^{1/q}\geq \sum_{i=1}^n a_i b_i.
\end{equation}
Because $\ds\frac{A}{M}\geq0$ in inequality (\ref{ineq_2_7}), we obtain a
proof of H\"{o}lder's inequality. It is easy to see that inequality
(\ref{ineq_2_7}) is a refinement of H\"{o}lder's inequality and contains a
reverse inequality for the inequality of H\"{o}lder.
\item[(b)] For $p=q=2$ in inequality (\ref{ineq_2_7}), we have an improvement of
Cauchy's inequality
\begin{equation}\label{ineq_2_9}
\ds\sum_{i=1}^n a_i^2\sum_{i=1}^n b_i^2\geq \left(\sum_{i=1}^n a_i
b_i\right)^2,
\end{equation}
given by the inequality
\begin{equation}\label{ineq_2_10}
\frac{A^2}{M^2}+\frac{2A}{M}\sum_{i=1}^n a_ib_i\leq
\left(\sum_{i=1}^n a_i^2\right)\left(\sum_{i=1}^n
b_i^2\right)-\left(\sum_{i=1}^n a_ib_i\right)^2\leq
\frac{A^2}{m^2}+\frac{2A}{m}\sum_{i=1}^n a_i b_i,
\end{equation}
where
$$A=\frac{1}{8}\sqrt{\ds\left(\sum_{i=1}^n a_i^2\right)\left(\sum_{i=1}^n b_i^2\right)}\cdot \sum_{i=1}^n \left(\frac{a_i^2}{\ds\sum_{i=1}^n a_i^2}-\frac{b_i^2}{\ds\sum_{i=1}^n b_i^2}\right)^2,$$
$$m=\min_{i=\ov{1,n}}\left\{\frac{a_i^2}{\ds\sum_{i=1}^n a_i^2}, \ \frac{b_i^2}{\ds\sum_{i=1}^n b_i^2}\right\} \ {\rm and} \ M=\max_{i=\ov{1,n}}\left\{\frac{a_i^2}{\ds\sum_{i=1}^n a_i^2}, \ \frac{b_i^2}{\ds\sum_{i=1}^n b_i^2}\right\}.$$
The equality holds for $\ds\frac{a_1}{b_1}=...=\frac{a_n}{b_n}$.
\item[(c)] In \cite{Pop}, O. T. Pop gave Bergstr\"{o}m's inequality,
\begin{equation}\label{ineq_2_11}
\frac{x_1^2}{a_1}+\frac{x_2^2}{q_2}+...+\frac{x_n^2}{a_n}\geq\frac{(x_1+x_2+...+x_n)^2}{a_1+a_2+...+a_n}
\end{equation}
for every $x_k\in\RR$ and $a_k>0, \ k\in\{1,2,...,n\}$.
If we make substitutions $a_i=\ds\frac{x_i}{\sqrt{a_i}}$ and
$b_i=\sqrt{a_i}$, for all $i=\{1,2,...,n\}$, in inequality (\ref{ineq_2_10}) we
find a new refinement of Bergstr\"{o}m's inequality, which is given as
follows
\begin{eqnarray*}
&& \left(\sum_{i=1}^n a_i\right)^{-1}\left(\frac{A^2}{M^2}+\frac{2A}{M}\sum_{i=1}^n |x_i|\right) \\
&& \leq \frac{x_1^2}{a_1}+\frac{x_2^2}{a_2}+...+\frac{x_n^2}{a_n}-\frac{(|x_1|+|x_2|+...+|x_n|)^2}{a_1+a_2+...+a_n}\\
&& \leq \left(\sum_{i=1}^n a_i\right)^{-1}\left(\frac{A^2}{m^2}+\frac{2A}{m}\sum_{i=1}^n |x_i|\right),
\end{eqnarray*}
where
$$A=\frac{1}{8}\sqrt{\ds\left(\sum_{i=1}^n \frac{x_i^2}{a_i}\right)\left(\sum_{i=1}^n a_i\right)}\cdot \sum_{i=1}^n \left(\frac{x_i^2}{a_i\ds\sum_{i=1}^n \frac{x_i^2}{a_i}}-\frac{a_i}{\ds\sum_{i=1}^n a_i}\right)^2,$$
$$m=\min_{i=\ov{1,n}}\left\{\frac{x_i^2}{a_i\ds\sum_{i=1}^n \frac{x_i^2}{a_i}},\frac{a_i}{\ds\sum_{i=1}^n a_i}\right\} \ and \ M=\max_{i=\ov{1,n}}\left\{\frac{x_i^2}{a_i\ds\sum_{i=1}^n \frac{x_i^2}{a_i}},\frac{a_i}{\ds\sum_{i=1}^n a_i}\right\}.$$
\end{itemize}
\end{rema}
\section{APPLICATION TO ARITHMETIC FUNCTIONS}
In the theory of the arithmetic functions \cite{Apo,Nat,SC}, for positive integer $n$,
several important functions have been studied. Among
these we found $\sigma_k(n), \ \tau(n), \ \sigma_k^*(n)$ and
$\tau^*(n)$, where $\sigma_k(n)$ is the sum of $k$th powers of the
divisors of $n$, $\tau(n)$ is the number of divisors of $n$,
$\sigma_k^*(n)$ is the sum of $k$th powers of the unitary divisors
of $n$ and $\tau^*(n)$ is the number of unitary divisors of $n$,
where $k\geq0$.
\begin{teo}
For $n\geq1$ and $k\geq0$, there are the following inequalities
\begin{eqnarray}
\ds\frac{1}{2n\tau(n)}\left[\sigma_{2k}(n)-\left(\frac{\sigma_k(n)}{\tau(n)}\right)^2\right]&\leq&
\frac{\sigma_k(n)}{\tau(n)}-\sqrt{n^k} \nonumber \\
&\leq& \ds\frac{1}{2\tau(n)}\left[\sigma_{2k}(n)-\left(\frac{\sigma_k(n)}{\tau(n)}\right)^2\right] \label{ineq_3_1}
\end{eqnarray}
and
\begin{eqnarray}
\ds\frac{1}{2n\tau^*(n)}\left[\sigma_{2k}^*
(n)-\left(\frac{\sigma^*_k(n)}{\tau^*(n)}\right)^2\right] &\leq&
\frac{\sigma_k^* (n)}{\tau^*(n)}-\sqrt{n^k} \nonumber \\
&\leq& \ds\frac{1}{2\tau^*(n)}\left[\sigma_{2k}^*(n)-\left(\frac{\sigma_k^*(n)}{\tau^*(n)}\right)^2\right]. \label{ineq_3_2}
\end{eqnarray}
\end{teo}
\noindent{\it Proof.} If $d_1,d_2,...,d_s$ are the divisors of $n$,
then we take $\alpha_i=\ds\frac{1}{s}$ and $x_i=d_i^k$ in inequality
(\ref{ineq_1_1}).\\
Therefore, we have $m=1, \ M=n$ and $s=\tau(n)$, so inequality (\ref{ineq_1_1})
becomes:
$$\frac{1}{2ns}\sum_{i=1}^s \left(d_i^k -\frac{\sigma_k(n)}{\tau(n)}\right)^2 \leq \frac{\sigma_k(n)}{\tau(n)}-\sqrt{n^k}
\leq\frac{1}{2s}\sum_{i=1}^s\left(d_i^k -\frac{\sigma_k(n)}{\tau(n)}\right)^2.$$
Making simple calculations and taking into account that
$\left(\ds\prod_{i=1}^s d_i^k\right)^{1/s}=\ds\left(\prod_{i=1}^s
d_i\right)^{\frac{k}{s}}=(n^{\frac{s}{2}})^{\frac{k}{s}}=n^{\frac{k}{2}},$
we observe that this inequality is equivalent to inequality (\ref{ineq_3_1}).
Similarly prove that inequality (\ref{ineq_3_2}) is true.
\hfill \qed
\begin{rema}
Inequality (\ref{ineq_3_2}) improves the inequality
\begin{equation}
\frac{\sigma_k(n)}{\tau(n)}\geq\sqrt{n^k},
\end{equation}
which is due to S. Sivaramakrishnan and C. S. Venkataraman \cite{SC} and
inequality (3.2) improves the inequality
$$\frac{\sigma_k^*(n)}{\tau^*(n)}\geq \sqrt{n^k},$$
which is due to J. S\'{a}ndor and L. T\'{o}th \cite{ST,SC}.
\end{rema}
\section{APPLICATIONS TO OPERATORS}
In this section, we consider bounded linear operators acting on a complex
Hilbert space $\mathcal{H} $. If a bounded linear operator $A$
satisfies $A=A^*$, then $A$ is called a self-adjoint operator. If a
self-adjoint operator $A$ satisfies $\langle x \vert A \vert x
\rangle \geq 0$ for any $\vert x \rangle \in \mathcal{H}$, then $A$
is called a positive operator and denoted by $A\geq 0$.
In addition, $A \geq B$ means $A -B \geq 0$.
We also define operator mean by $A\sharp_{\lambda}B\equiv
A^{1/2}\left(A^{-1/2}BA^{-1/2}\right)^{\lambda}A^{1/2}$ for $\lambda
\in [0,1]$, two invertible positive operator $A$ and $B$ \cite{KA}.
Note that we have the relation $B\sharp_{1-\lambda}A =
A\sharp_{\lambda}B$.
\begin{teo} \label{op_the1}
For $\lambda\in [0,1]$, two invertible positive operator $A$ and $B$, we have the following relations.
\begin{itemize}
\item[(i)] If $A \leq B$, then we have
\begin{eqnarray}
\frac{\lambda(1-\lambda)}{2} \left(AB^{-1}A-2A+B \right) &\leq& (1-\lambda) A +\lambda B -A \sharp_{\lambda} B \nonumber \\
&\leq & \frac{\lambda(1-\lambda)}{2} \left(BA^{-1}B-2B+A \right). \label{op_the_ineq01}
\end{eqnarray}
\item[(ii)] If $B \leq A$, then we have
\begin{eqnarray}
\frac{\lambda(1-\lambda)}{2} \left(BA^{-1}B-2B+A \right) &\leq& (1-\lambda) A +\lambda B -A \sharp_{\lambda} B \nonumber \\
&\leq & \frac{\lambda(1-\lambda)}{2}\left(AB^{-1}A-2A+B \right) .\label{op_the_ineq02}
\end{eqnarray}
\end{itemize}
\end{teo}
{\it Proof:}
We prove (i).
Exchanging $\lambda$ and $1-\lambda$ in the inequalities (1.2), we have
$$
\frac{\lambda(1-\lambda)}{2b}(a-b)^2\leq (1-\lambda)a+\lambda b-a^{1-\lambda}b^{\lambda} \leq \frac{\lambda(1-\lambda)}{2a}(a-b)^2
$$
in the case of $a \leq b$.
Thus we have the inequalities for $0<t \leq 1$:
$$
\frac{\lambda(1-\lambda)}{2}(t-1)^2\leq (1-\lambda)t+\lambda -t^{1-\lambda} \leq \frac{\lambda(1-\lambda)}{2}\left(\sqrt{t}-\frac{1}{\sqrt{t}}\right)^2
$$
putting $t \equiv \frac{a}{b}$.
Thus we have for $0<T\leq I$,
$$
\frac{\lambda(1-\lambda)}{2}(T-1)^2\leq (1-\lambda)T+\lambda -T^{1-\lambda} \leq \frac{\lambda(1-\lambda)}{2}(T^{1/2}-T^{-1/2})^2
$$
by standard operational calculus.
Putting $T=B^{-1/2}AB^{-1/2}$ and then multiplying $B^{1/2}$ from the both sides,
we obtain the desired results.
(ii) can be proven by the similar way to the proof of (i).
\hfill \qed
\begin{rema}
We have $AB^{-1}A-2A+B=A^{1/2}\left( A^{1/2}B^{-1}A^{1/2} +\left( A^{1/2}B^{-1}A^{1/2} \right)^{-1}-2I\right)A^{1/2} \geq 0$,
becasue we have $u+u^{-1}-2=\frac{(u-1)^2}{u}\geq 0$ for scalar $u\geq 0$.
By the similar way, we have $BA^{-1}B-2B+A \geq 0$.
Thus under the condition $A\leq B$ or $B\leq A$,
the inequalities in Theorem \ref{op_the1} improve the second inequality of the following inequalities (See \cite{FY,Furuichi} for example):
\begin{equation} \label{op_rem_ineq01}
\left\{ (1-\lambda)A^{-1}+\lambda B^{-1} \right\}^{-1} \leq A\sharp_{\lambda} B \leq (1-\lambda) A+\lambda B
\end{equation}
\end{rema}
\begin{cor}
For $\lambda\in (0,1)$, two invertible positive operator $A$ and $B$, we have the following relations.
\begin{itemize}
\item[(i)] If $A \leq B$, then we have
\begin{eqnarray}
&&A \sharp_{\lambda} B -\left(A \sharp_{\lambda} B \right) \left\{\frac{2}{\lambda(1-\lambda)}\left(B^{-1}AB^{-1}-2B^{-1}+A^{-1}\right)^{-1}+A \sharp_{\lambda} B \right\}\left( A \sharp_{\lambda} B \right) \nonumber\\
&&\leq \left\{ (1-\lambda) A^{-1} +\lambda B^{-1}\right\}^{-1} \nonumber \\
&&\leq A \sharp_{\lambda} B -\left(A \sharp_{\lambda} B \right) \left\{\frac{2}{\lambda(1-\lambda)}\left(A^{-1}BA^{-1}-2A^{-1}+B^{-1}\right)^{-1}+A \sharp_{\lambda} B \right\}\left( A \sharp_{\lambda} B \right). \label{op_the_ineq03}
\end{eqnarray}
\item[(ii)] If $B \leq A$, then we have
\begin{eqnarray}
&&A \sharp_{\lambda} B -\left(A \sharp_{\lambda} B \right) \left\{\frac{2}{\lambda(1-\lambda)}\left(A^{-1}BA^{-1}-2A^{-1}+B^{-1}\right)^{-1}+A \sharp_{\lambda} B \right\}\left( A \sharp_{\lambda} B \right) \nonumber\\
&&\leq \left\{ (1-\lambda) A^{-1} +\lambda B^{-1}\right\}^{-1} \nonumber \\
&&\leq A \sharp_{\lambda} B -\left(A \sharp_{\lambda} B \right) \left\{\frac{2}{\lambda(1-\lambda)}\left(B^{-1}AB^{-1}-2B^{-1}+A^{-1}\right)^{-1}+A \sharp_{\lambda} B \right\}\left( A \sharp_{\lambda} B \right). \label{op_the_ineq04}
\end{eqnarray}
\end{itemize}
\end{cor}
{\it Proof:}
Replacing $A$ and $B$ by $A^{-1}$ and $B^{-1}$ in the inequalities (\ref{op_the_ineq01}), respectively and taking the inverse of bothe sides, then
we have the inequalities (\ref{op_the_ineq03}), using $\left(A^{-1} \sharp_{\lambda}B^{-1}\right)^{-1}= A\sharp_{\lambda}B$ and
$$
\left( X^{-1}+Y^{-1} \right)^{-1} = X^{-1} -X^{-1}\left( X^{-1}+Y^{-1} \right)^{-1}X^{-1}
$$
for invertible positive operators $X$ and $Y$.
The inequalities (\ref{op_the_ineq04}) can be proven by the similar way to the inequalities (\ref{op_the_ineq03}).
\hfill \qed
\begin{rema}
Since $\left(A^{-1}BA^{-1}-2A^{-1}+B^{-1}\right)^{-1}\geq 0$, $\left(B^{-1}AB^{-1}-2B^{-1}+A^{-1}\right)^{-1}\geq 0$ and $A\sharp_{\lambda}B \geq 0 $,
then the right hand side of the inequalities (\ref{op_the_ineq03}) and (\ref{op_the_ineq04}) are further bounbed from the above by $A\sharp_{\lambda}B $.
Therefore two inequalities (\ref{op_the_ineq03}) and (\ref{op_the_ineq04}) improve the first inequality of the inequalities (\ref{op_rem_ineq01})
under the condition $A \leq B$ or $B \leq A$.
\end{rema}
\begin{cor}
If $0<A\leq B$, then we have
\begin{equation} \label{op_cor_eq01}
3(A-B) + BA^{-1}B-AB^{-1}A \geq 0.
\end{equation}
\end{cor}
The inequality (\ref{op_cor_eq01}) corresponds to the following relation:
$$
0< a \leq b \Rightarrow (b-a)^3 \geq 0
$$
in the commutative case. The inequality (\ref{op_cor_eq01}) can be directly proven by applying the standard operational calculus to
the scalar inequality $(t-1)^3\leq 0$ for $0<t\leq 1$.
\section*{Acknowledgement}
The first author (N.M.) was supported in part by the Romanian
Ministry of Education, Research and Innovation through the PNII Idei
project 842/2008.
The second author (S.F.) was supported in
part by the Japanese Ministry of Education, Science, Sports and
Culture, Grant-in-Aid for Encouragement of Young Scientists (B),
20740067.
|
\section{Covering numbers of the set of $1$-Lipschitz functions}
In this section, we provide bounds for the covering numbers of the set of $1$-Lipschitz functions over a precompact space.
Note that these results are likely not new.
However, we have been unable to find an original work, so we provide proofs for completeness.
Let $(K, d)$ be a precompact metric space, and let $\mathcal{N}(K, \delta)$ denote the minimal number of
balls of radius $\delta$ necessary to cover $K$. Let $x_0 \in K$ be a fixed point, and let
$\mathcal{F}$ denote the set of $1$-Lipschitz functions over $K$ vanishing at $x_0$.
This is also a precompact space when endowed with the metric of uniform convergence.
We denote by $\mathcal{N}(\mathcal{F}, \delta)$ the minimal number of balls of radius $\delta$ necessary to cover
$\mathcal{F}$. Finally, we set $R = \max_{x \in K} d(x, x_0)$.
Our first estimate is a fairly crude one.
\begin{prop}
We have
\begin{equation*}
\mathcal{N}(\mathcal{F}, \varepsilon) \leq \left( 2 + 2 \lfloor \frac{3 R}{\varepsilon} \rfloor \right)^{\mathcal{N}(K, \frac{\varepsilon}{3})}.
\end{equation*}
\end{prop}
\begin{proof}
For simplicity, write $n = \mathcal{N}(K, \varepsilon)$.
Let $x_1, \ldots, x_n$ be the centers of a set of balls covering $K$.
For any $f \in \mathcal{F}$ and $1 \leq i \leq n$, we have
\begin{equation*}
|f(x_i)| = |f(x_i) - f(x_0)| \leq R.
\end{equation*}
For any $n$-uple of integers $k = (k_1, \ldots, k_n)$ such that
$- \lfloor \frac{R}{\varepsilon} \rfloor - 1 \leq k_i \leq \lfloor \frac{R}{\varepsilon} \rfloor$,
$1 \leq i \leq n$,
choose a function $f_k \in \mathcal{F}$ such that $k_i \varepsilon \leq f_k(x_i) \leq (k_i + 1) \varepsilon$ if there exists one.
Consider $f \in \mathcal{F}$.
Let $l_i = \lfloor \frac{f(x_i)}{\varepsilon} \rfloor$ and $l = (l_1, \ldots, l_n)$.
Then the function $f_l$ defined above exists and $|f(x_i) - f_l(x_i)| \leq \varepsilon$ for $1 \leq i \leq n$. But then for any $x \in K$ there exists $i$, $1 \leq i \leq n$,
such that $x \in B(x_i, \varepsilon)$, and thus
\begin{equation*}
|f(x) - f_l(x) | \leq |f(x) - f(x_i)| + |f(x_i) - f_l(x_i)| + |f_l(x_i) - f_l(x)| \leq 3 \varepsilon.
\end{equation*}
This implies that $\mathcal{F}$ is covered by the balls of center $f_k$ and radius $3 \varepsilon$. As there are at most
$(2 + 2 \lfloor \frac{R}{\varepsilon} \rfloor)^n$ choices for $k$, this ends the proof.
\end{proof}
However, this bound is quite weak : as one can see by considering the case of a segment, for most choices of a $n$-uple, there
will not exist a function in $\mathcal{F}$ satisfying the requirements in the proof.
With the extra assumption that $K$ is connected, we can get a more refined result.
\begin{prop} \label{prop_covering_number_lip}
If $K$ is connected, then
\begin{equation} \label{covering_number_lip_2}
\mathcal{N}( \mathcal{F}, \varepsilon) \leq \left( 2 + 2 \lfloor \frac{4 R}{\varepsilon} \rfloor \right) \, 2^{ \displaystyle \mathcal{N}(K, \frac{\varepsilon}{16}) }.
\end{equation}
\end{prop}
\begin{remark}
The simple idea in this proposition is first to bring the problem to a discrete metric space
(graph), and then to bound the number of
Lipschitz functions on this graph by the number of Lipschitz functions on a spanning tree of the graph.
\end{remark}
\begin{proof}
In the following, we will denote $n = \mathcal{N}(K, \varepsilon)$ for simplicity.
Let $x_i$, $1 \leq i \leq n$ be the centers of a set of $n$ balls
$B_1, \ldots B_n$ covering $K$. Consider the graph $G$
built on the $n$ vertices $a_1, \ldots, a_n$,
where vertices $a_i$ and $a_j$ are connected if and only if
$i \neq j$ and the balls $B_i$ and $B_j$ have a non-empty intersection.
\begin{lemma}
The graph $G$ is connected. Moreover, there exists a subgraph $G'$
with the same set of vertices and whose edges are edges of $G$, which is a tree.
\end{lemma}
\begin{proof}
Suppose that $G$ were not connected . Upon exchanging the labels of the balls, there would
exist $k$, $1 \leq k < n$, such that for $i \leq k < j$ the
balls $B_i$ and $B_j$ have empty intersection.
But then $K$ would be equal to the disjoint reunion of the sets
$\bigcup_{i = 1}^k B_i$ and $\bigcup_{j = k+1}^n B_j$, which are both closed and
non-empty, contradicting the connectedness of $K$.
The second part of the claim is obtained by an easy induction on the size of the graph.
\end{proof}
Introduce the set $\mathcal{A}$
of functions $g : \{a_1, \ldots, a_n \} \rightarrow \mathbb{R}$ such that
$g(a_1) = 0$ and $|g(a_i) - g(a_j)| = 4 \varepsilon$ whenever $a_i$ and $a_j$ are connected in $G'$.
Using the fact that $G'$ is a tree, it is easy to see that $\mathcal{A}$ contains at most
$2^n$ elements.
Define a partition of $K$ by setting $C_1 = B_1$,
$C_2 = B_2 \backslash C_1$, $\ldots$, $ C_n = B_n \backslash C_{n-1}$ (remark that none of the $C_i$ is empty since
the $B_i$ are supposed to constitute a minimal covering).
Also fix for each $i$, $1 \leq i \leq n$, a point $y_i \in C_i$ (choosing $y_1 = x_1$). Notice that
$C_i$ is included in the ball of center $y_i$ and radius $2 \varepsilon$, and that
$d(y_i, y_j) \leq 4 \varepsilon$ whenever $a_i$ and $a_j$ are connected in $G$ (and therefore in $G'$).
To every $1$-Lipschitz function $f : K \rightarrow \mathbb{R}$ we associate $T(f) : \{a_1, \ldots, a_n \} \rightarrow \mathbb{R}$
defined by $T(f) (a_i) = f(y_i)$. For any $x \in K$, and $f_1$, $f_2 \in \mathcal{F}$, we have the following :
\begin{eqnarray*}
|f_1(x) - f_2(x) | & \leq & |f_1(x) - f_1(y_i) | + |f_1(y_i) - f_2(y_i) | + |f_2(y_i) - f_2(x)| \\
{} & \leq & 4 \varepsilon + \|T(f_1) - T(f_2)\|_{ \ell_\infty(G') }
\end{eqnarray*}
where $i$ is such that $x \in C_i$.
We now make the following claim :
\begin{lemma} \label{lip_approx}
For every $1$-Lipschitz function $f : K \rightarrow \mathbb{R}$ such that
$f(y_1) = 0$, there exists $g \in \mathcal{A}$ such that $\| T(f) - g \|_{\ell_\infty(G')} \leq 4 \varepsilon$.
\end{lemma}
Assume for the moment that this holds. As there are at most $2^n$ functions in $\mathcal{A}$, it is possible to choose
at most $2^n$ $1$-Lipschitz functions $f_1, \ldots, f_{2^n}$ vanishing at $x_1$ such that for any
$1$-Lipschitz function f vanishing at $x_1$ there exists $f_i$ such that $|T(f) - T(f_i)| \leq 8 \varepsilon$.
Using the inequality above, this implies that the balls of center $f_i$ and radius $12 \varepsilon$ for the uniform distance
cover the set of $1$-
Lipschitz functions vanishing at $x_1$.
Finally, consider $f \in \mathcal{F}$. We may write
\begin{equation*}
f = f - f(x_1) + f(x_1)
\end{equation*}
and observe that on the one hand, $f - f(x_1)$ is a $1$-Lipschitz function vanishing at $x_1$, and that on the other hand,
$|f(x_1)| \leq R$.
Thus the set $\mathcal{F}$ is covered by the balls of center $f_i + 4 k \varepsilon$ and radius $16 \varepsilon$,
where $- \lfloor \frac{R}{ 4 \varepsilon} \rfloor - 1 \leq k \leq \lfloor \frac{R}{4 \varepsilon} \rfloor$.
There are at most $(2 + 2 \lfloor \frac{R}{4 \varepsilon} \rfloor) 2^n$ such balls, which proves the desired result.
\end{proof}
We now prove Lemma \ref{lip_approx}.
\begin{proof}
Let us use induction again. If $K = B_1$ then $T(f) = 0$ and the property is straightforward.
Now if $K = C_1 \cup \ldots \cup C_n$, we may assume without loss of generality that $a_n$ is a leaf in $G'$, that
is a vertex with exactly one neighbor, and that it is connected to $a_{n-1}$.
By hypothesis there exists $\tilde{g}: \{ a_1, \ldots, a_{n-1} \} \rightarrow \mathbb{R}$
such that $|\tilde{g}(a_i) - \tilde{g}(a_j)| = 4 \varepsilon$
whenever $a_i$ and $a_j$ are connected in $G'$, and $|\tilde{g}(a_i) - f(a_i)| \leq 4 \varepsilon$ for $1 \leq i < n$.
Set $g = \tilde{g}$ on $ \{a_1, \ldots, a_{n-1} \}$, and
\begin{itemize}
\item $g(a_n) = g(a_{n-1}) + 4 \varepsilon$ if $f(y_n) - g(a_{n-1}) < 0$,
\item $g(a_n) = g(a_{n-1}) - 4 \varepsilon$ otherwise.
\end{itemize}
Since
\begin{equation*}
|f(y_n) - g(a_{n-1}) | \leq |f(y_n) - f(y_{n-1})| + |f(y_{n-1}) - g(a_{n-1}) | \leq 8 \varepsilon
\end{equation*}
it is easily checked that $|f(y_n) - g(a_n) | \leq 4 \varepsilon$. The function $g$ belongs to $\mathcal{A}$ and our claim is proved.
\end{proof}
\section{H\"{o}lder moments of stochastic processes}
We quote the following result from Revuz and Yor's book \cite{revuz1999continuous} (actually
the value of the constant is not given in their statement but is easily tracked in the proof).
\begin{theorem} \label{kolmogorov_theorem}
Let $X_t$, $t \in [0, 1]$ be a Banach-valued process such that there exist $\gamma, \varepsilon, c > 0$ with
\begin{equation*}
\mathbb{E} \left[|X_t-X_s| ^\gamma \right] \leq c |t-s|^{1+ \varepsilon},
\end{equation*}
then there exists a modification $\tilde{X}$ of $X$ such that
\begin{equation*}
\mathbb{E} \left[ \left( \sup_{s \neq t} \frac{|\tilde{X}_t - \tilde{X}_s|}{|t-s|^\alpha} \right)^\gamma \right]^{1/\gamma} \leq 2^{1 + \alpha} (2c)^{1/\gamma} \frac{1}{1 - 2^{\alpha - \varepsilon/\gamma}}
\end{equation*}
for all $0 \leq \alpha < \varepsilon/\gamma$.
\end{theorem}
\begin{corollary} \label{holder_moments_sbm}
Let $(B_t)_{0 \leq t \leq 1}$ denote the standard Brownian motion on $[0, 1]$.
Let $m_\alpha = \mathbb{E} \sup_t \| B_t \|_\alpha$ and $s^2_\alpha = \mathbb{E} (\sup_t \| B_t \|_\alpha)^2$, then $m_\alpha$ and $s_\alpha$ are bounded by
\begin{equation*}
C_\alpha = 2^{1+ \alpha}\frac{2^{(1-2 \alpha)/4}}{1 - 2^{(2 \alpha-1)/4}} \|Z\|_{4/(1-2\alpha)}
\end{equation*}
where $\|Z\|_p$ denotes the $L_p$ norm of a $\mathcal{N}(0, 1)$ variable $Z$.
\end{corollary}
\begin{proof}
Since the increments of the Brownian motion are Gaussian, we have for every $p > 0$
\begin{equation*}
\mathbb{E}[ |B_t - B_s|^{2p} ] = K_p |t-s|^p
\end{equation*}
with $K_p = \sqrt{2 \pi}^{-1} \int_{- \infty}^{+\infty} |x|^{2p} e^{-x^2/2} dx$.
Choose $p$ such that $\alpha < (p-1)/2p$, then
\begin{equation*}
\left( \mathbb{E} \|X\|_\alpha^{2p} \right)^{1/2p} \leq \frac{2^{1+ \alpha}}{1 - 2^{\alpha - 1/2 + 1/2p}} (2 K_p)^{1/2p}.
\end{equation*}
A suitable choice is $1/p = 1/2 - \alpha$, and the right-hand side becomes
\begin{equation*}
C_\alpha = \frac{2^{1+ \alpha}}{1 - 2^{(\alpha - 1/2)/2}} (2 G_\alpha)^{(1/2 - \alpha)/2}
\end{equation*}
with $G_\alpha = \sqrt{2 \pi}^{-1} \int_{- \infty}^{+\infty} |x|^{4/(1 - 2 \alpha)} e^{-x^2/2} dx$.
By H\"{o}lder's inequality, the result follows.
\end{proof}
\begin{corollary} \label{corollary_holder_sde}
Let $X_t$ be the solution on $[0, T]$ of
\begin{equation*}
d X_t = \sigma(X_t) dB_t + b(X_t) dt
\end{equation*}
with $\sigma, b : \mathbb{R} \rightarrow \mathbb{R}$ locally Lipschitz and satisfying the following hypotheses :
\begin{itemize}
\item $\sup_x | \sqrt{ \text{tr}\sigma(x)^t\sigma(x)} | \leq A$,
\item $\sup_x |b(x)| \leq B$.
\end{itemize}
Then for $\alpha < 1/2$, for $p$ such that $\alpha < (p-1)/2p$,
there exists $C < + \infty$ depending explicitely on $A$, $B$, $T$, $\alpha$ ,$p$
such that
\begin{equation*}
\mathbb{E}\|X\|_\alpha^p \leq C.
\end{equation*}
\end{corollary}
\begin{proof}
We first apply It\^{o}'s formula to the function $|X_t - X_s|^2$ : this yields
\begin{equation*}
\mathbb{E}|X_t - X_s|^2 \leq 2 B \int_s^t \mathbb{E}|X_u - X_s| du + A |t-s|.
\end{equation*}
Using the elementary inequality $x \leq 1/2(1 + x^2)$, we get
\begin{equation*}
\mathbb{E}|X_t - X_s|^2 \leq B \int_s^t \mathbb{E}|X_u - X_s|^2 du + (A + B)|t-s|.
\end{equation*}
Gronwall's lemma entails
\begin{equation*}
\mathbb{E}|X_t - X_s|^2 \leq (A + B) e^{B T} |t -s|
\end{equation*}
Likewise, applying It\^{o}'s formula to $|X_t - X_s|^4$, we get
\begin{align*}
\mathbb{E}|X_t - X_s|^4 & \leq 4 B \int_s^t \mathbb{E}|X_u - X_s|^3 ds + 6A \int_s^t \mathbb{E}|X_u-X_s|^2 du \\
{} & \leq (6A+2B) \int_s^t \mathbb{E}|X_u-X_s|^2 du + 2B \int_s^t \mathbb{E}|X_u-X_s|^4 du \\
{} &\leq \frac{1}{2}(6A+2B)(A+B)e^{BT}|t-s|^2 + 2B\int_s^t \mathbb{E}|X_u-X_s|^4 du
\end{align*}
and by Gronwall's lemma $\mathbb{E}|X_t - X_s|^4 \leq \frac{1}{2}(6A+2B)(A+B)e^{3BT}|t-s|^2$.
By an easy recurrence, following the above, one may show that
\begin{equation*}
\mathbb{E}|X_t - X_s|^{2p} \leq C(A, B, T, p) |t-s|^p.
\end{equation*}
To conclude it suffices to call on Theorem \ref{kolmogorov_theorem}.
\end{proof}
\section{Transportation inequalities for Gaussian measures on a Banach space} \label{appendix_gaussian}
\begin{lemma} \label{lemma_gaussian_tail}
Let $(E, \mu)$ be a Gaussian Banach space, and define $m = \int \|x\| \mu(dx)$. Also let $\sigma^2$ denote the weak variance of $\mu$.
The tail of $\mu$ is bounded as follows : for all $R \geq 0$,
\begin{equation*}
\mu \{x \in E, \; \|x\| \geq m + R \} \leq e^{- R ^2/2\sigma^2 }.
\end{equation*}
\end{lemma}
\bigskip
Finally we collect some (loose) results on the Wiener measure on the Banach space $(C([0, 1], \mathbb{R}), \|.\|_\infty)$.
\begin{lemma}
The Wiener measure satisfies a $\mathbf{T}_2(8)$ inequality (and therefore a $\mathbf{T}_1(8)$ inequality).
\end{lemma}
\begin{proof}
The Wiener measure satisfies the $\mathbf{T}_2(2 \sigma^2)$ inequality,
where
\begin{equation*}
\sigma^2 = \sup_{\mu} \mathbb{E} ( \int_0^1 B_s d \mu(s))^2
\end{equation*}
and the supremum runs over all Radon measures on $[0, 1]$ with total variation bounded by $1$.
Note that the weak variance $\sigma^2$ is bounded by the variance $s^2$ defined as
$s^2 = \mathbb{E} (\sup_t |B_t|) ^2$
(here
and hereafter $\sup_t |B_t|$ refers to the supremum on $[0, 1]$).
In turn we can give a (quite crude) bound on $s$ :
write $\sup_t |B_t| \leq \sup_t B_t - \inf_t B_t$, thus
$(\sup_t |B_t|)^2 \leq ( \sup_t B_t - \inf_t B_t)^2 \leq 2 (\sup_t B_t)^2 + 2 (- \inf_t B_t)^2$.
Remember the well-known fact that $\sup_t B_t$, $- \inf_t B_t$ and $|B_1|$ have the same law, so that
\begin{equation*}
\mathbb{E}( \sup_t |B_t|)^2 \leq 4 \mathbb{E} |B_1|^2 = 4.
\end{equation*}
\end{proof}
\begin{lemma} \label{lemma_small_exp_moment_wiener}
Let $\gamma$ denote the Wiener measure. For $a = \sqrt{2 \log 2}/3$, we have
\begin{equation*}
\int e^{a\|x\|_\infty} \gamma(dx) \leq 2
\end{equation*}.
\end{lemma}
\begin{proof}
We have
\begin{align*}
\int e^{a\|x\|_\infty} \gamma(dx) & = \int_0^{ + \infty} \mathbb{P} ( e^{a \|x\|_\infty} \geq t) d t \\
{} & = \int_0^{+ \infty} \mathbb{P}( \|x\|_\infty \geq u) a e^{a u} du \\
{} & = \int_0^{+ \infty} \mathbb{P}( \tau_u \leq 1) a e^{a u} du
\end{align*}
where $\tau_u$ is the stopping time $\inf \{t, |B_t| = u \}$. It is a simple exercise to compute
\begin{equation*}
\mathbb{E} e^{- \lambda^2 \tau_u / 2} = 1/ \cosh (\lambda u) \leq 2 e^{- \lambda u}.
\end{equation*}
This yields
\begin{equation*}
\int e^{a\|x\|_\infty} \gamma(dx) \leq 2 a e^{\lambda^2/2}\int_0^{+ \infty} e^{(a - \lambda) u} du = \frac{2 a e^{ \lambda^2/2}}{\lambda - a}.
\end{equation*}
We can choose $\lambda = 3 a$ to get $\int e^{a\|x\|_\infty} \gamma(dx) \leq e^{9a^2/2}$. In turn it implies the desired result.
\end{proof}
\section{Some facts on transportation inequalities} \label{appendix_transport_inequalities}
A crucial feature of transportation inequalities is that they imply the concentration of measure phenomenon, a fact first discovered by Marton
(\cite{marton1996bounding}). The following proposition is obtained by a straightforward adaptation of her famous argument :
\begin{prop}
If $\mu$ verifies a $\alpha(\mathcal{T}_d)$ inequality, then for all measurable sets $A \subset E$ with $\mu(A) \geq \frac{1}{2}$ and $r \geq r_0 = \alpha^{-1}(\log 2)$,
\begin{equation*}
\mu(A^r) \geq 1 - e^{-\alpha(r - r_0)}
\end{equation*}
where $A^r = \{ x \in E, \; d(x, A) \leq r \}$.
Moreover, let $X$ be a r.v. with law $\mu$. For all $1$-Lipschitz functions $f : E \rightarrow \mathbb{R}$ and all $r \geq r_0$, we have
\begin{equation*}
\mathbb{P} ( f(X) \geq m_f + r) \leq e^{- \alpha(r - r_0)}
\end{equation*}
where $m_f$ denotes a median of $f$.
\end{prop}
Bobkov and G\"{o}tze (\cite{bobkov1999exponential}) were the first to obtain an equivalent dual formulation of transportation inequalities.
We present it here in a more general form obtained by Gozlan and Leonard (see \cite{gozlan2010transport}), in the case when the transportation cost function is the distance.
\begin{definition}
Let $\alpha : [0, + \infty) \rightarrow \mathbb{R}$ be convex, increasing, left-continuous and vanishing at $0$.
The monotone conjugate of $\alpha$ is
\begin{equation*}
\alpha^\circledast (s) = \sup_{t \geq 0} st - \alpha(t).
\end{equation*}
\end{definition}
\begin{prop}[\cite{gozlan2010transport}] \label{bobkov_gotze_characterization}
Assume that $d$ is a metric defining the topology of $E$, and that there exist $a > 0$, $x_0 \in E$ such that
$\int \exp[ a d(x, x_0) ]\mu (d x) < + \infty$.
Then $\mu$ satisfies the $\alpha(\mathcal{T}_d)$ inequality
\begin{equation*}
\alpha( \mathcal{T}_d (\mu, \nu) ) \leq H( \nu | \mu)
\end{equation*}
for all $\nu \in \mathcal{P}(E)$ with finite first moment if and only if
for all $f : E \rightarrow \mathbb{R}$ $1$-Lipschitz and all $\lambda > 0$,
\begin{equation} \label{cond_laplace}
\int e^{\lambda (f(x) - \int f d \mu)} \mu(d x) \leq e^{ \alpha^\circledast (\lambda)}.
\end{equation}
\end{prop}
In the case $\mathbf{T}_1(C)$, Condition (\ref{cond_laplace}) becomes : for all $1$-Lipschitz $f : E \rightarrow \mathbb{R}$ and $\lambda > 0$,
\begin{equation}
\int e^{\lambda(f - \int f d \mu)} \mu (x) \leq e^{C\lambda^2/4}.
\end{equation}
\subsubsection{Integral criteria} \label{integral_criteria}
An interesting feature of transportation inequalities is that some of them are
characterized by simple moment conditions, making it tractable to obtain their existence.
In \cite{djellout_guillin_wu},
Djellout, Guillin and Wu showed that $\mu$ satisfies a $\mathbf{T}_1$ inequality if and only if
$\int \exp [ a_0 d^2(x_0, y)] \mu(dy) < + \infty$
for some $a_0$ and some $x_0$. They also connect the value of $a_0$ and of the Gaussian moment with the value of
the constant $C$ appearing in the transportation inequality. More generally,
Gozlan and Leonard provide in \cite{large_dev_gozlan_leonard} a nice criterion
to ensure that
a $\alpha( \mathbf{T}_d)$ inequality holds. We only quote here one side of what is actually an equivalence :
\begin{theorem}
Let $\mu \in \mathcal{P}(E)$.
Suppose there exists $a > 0$ with $\int e^{a d(x_0, x)} \mu(d x) \leq 2$ for some $x_0 \in E$, and
a convex, increasing function $\gamma$ on $[0, + \infty)$ vanishing at $0$ and $x_1 \in E$ such that
\begin{equation*}
\int \exp \gamma (d(x_1, x) ) \mu(d x) \leq B < + \infty.
\end{equation*}
Then $\mu$ satisfies the $\alpha(\mathcal{T}_d)$ inequality
\begin{equation*}
\alpha( W_1( \mu, \nu) ) \leq H( \nu | \mu)
\end{equation*}
for all $\nu \in \mathcal{P}(E)$ with finite first moment, with
\begin{equation*}
\alpha(x) = \max \left( (\sqrt{ax + 1} - 1)^2, \; 2 \gamma( \frac{x}{2} - 2 \log B )\right).
\end{equation*}
\end{theorem}
One particular instance of the result above was first obtained by Bolley and Villani, with sharper constants, in the case when $\mu$
only has a finite exponential moment (\cite{bolley2005weighted}), Corollary 2.6). Their technique involves the study of weighted Pinsker inequalities, and
encompasses more generally costs of the form $d^p$, $p \geq 1$ (we give only the case $p = 1$ here).
\begin{theorem}
Let $a > 0$ be such that $E_{a, 1} = \int e^{a d(x_0,x)} \mu(d x) < + \infty$. Then for $\nu \in \mathcal{P}_1(E)$, we have
\begin{equation*}
W_1(\mu, \nu) \leq C \displaystyle \left( H( \nu | \mu) + \sqrt{ H(\nu | \mu) } \right)
\end{equation*}
where $C = \frac{2}{a} \left( \frac{3}{2} + \log E_{a, 1} \right) < + \infty.$
\end{theorem}
And in the case when $\mu$ admits a finite Gaussian moment, the following holds (\cite{bolley2005weighted}, Corollary 2.4) :
\begin{theorem}
Let $a > 0$ be such that $E_{a, 2} = \int e^{a d^2(x_0,x)} \mu(d x) < + \infty$. Then $\mu$ satisfies a $\mathbf{T}_1(C)$ inequality
where $C = \frac{2}{a} \left( 1 + \log E_{a, 2} \right) < + \infty.$
\end{theorem}
\section{Proofs in the dependent case} \label{dependent_case}
Before proving Theorem \ref{theorem_markov}, we establish a more general result in the spirit of Lemma \ref{lemma_mean}.
As earlier, the first ingredient we need to apply our strategy of proof is
a tensorization property for the transport-entropy inequalities
in the case of non-independent sequences. To this end, we restate results from
\cite{djellout_guillin_wu}, where only $\mathbf{T}_1$
inequalities were investigated, in our framework.
For $x = (x_1, \ldots, x_n) \in E^n$, and $1 \leq i \leq n$, denote $x^i = (x_1, \ldots, x_i)$.
Endow $E^n$ with the distance $d_1(x, y) = \sum_{i _ 1}^n d(x_i, y_i)$.
Let $\nu \in \mathcal{P}(E^n)$, the notation $\nu^i(d x_1, \ldots, d x_i)$ stands for the marginal measure
on $E^i$, and $\nu^i(. | x^{i-1})$ stands for the regular
conditional law of $x_i$ knowing $x^{i-1}$, or in other words the conditional
disintegration of $\nu^i$ with respect to $\nu^{i-1}$ at $x^{I-1}$(its existence is assumed throughout).
The next theorem is a slight extension of Theorem 2.11 in \cite{djellout_guillin_wu}.
Its proof can be adapted without difficulty, and we omit it here.
\begin{theorem} \label{theorem_concentration_dependent_sequences}
Let $\nu \in \mathcal{P}(E^n)$ be a probability measure such that
\begin{enumerate}
\item For all $i \geq 1$ and all $x^{i-1} \in E^{i-1}$ ($E^0 = \{x_0\})$, $\nu^i(. |x^{i-1})$
satisfies a $\alpha(\mathcal{T}_d)$ inequality, and
\item
There exists $S > 0$ such that for every $1$-Lipschitz function
\begin{equation*}
f : (x_{k+1}, \ldots, x_n) \rightarrow f(x_{k+1}, \ldots, x_n),
\end{equation*}
for all $x^{k-1} \in E^{k-1}$ and $x_k, y_k \in E$, we have
\begin{equation} \label{contraction_property}
\begin{split}
| \mathbb{E}_\nu\left( f(X_{k+1}, \ldots, X_n) | X^k = (x^{k-1}, x_k) \right) - \\
\mathbb{E}_\nu \left( f(X_{k+1}, \ldots, X_n) | X^k = (x^{k-1}, y_k) \right) | \\
\leq S d(x_k, y_k)
\end{split}
\end{equation}
\end{enumerate}
Then $\nu$ verifies the transportation inequality $\tilde{\alpha}( \mathcal{T}_d) \leq H$
with
\begin{equation*}
\tilde{\alpha}(t) = n \alpha(\frac{1}{n (1+S)} t).
\end{equation*}
\end{theorem}
In the case of a homogeneous Markov chain $(X_n)_{n \in \mathbb{N}}$ with transition kernel $P(x, dy)$,
the next
proposition gives sufficient conditions
on the transition probabilities for
the laws of the variables $X_n$ and the path-level law
of $(X_1, \ldots, X_n)$ to satisfy some transportation inequalities.
Once again the statement and its proof are
adaptations of the corresponding Proposition 2.10 of \cite{djellout_guillin_wu}.
\begin{prop} \label{marginals_markov}
Let $P(x, dy)$ be a Markov kernel such that
\begin{enumerate}
\item \label{cond_a}the transition measures $P(x, .)$ satisfies $\alpha(\mathcal{T}_d) \leq H$ for all $x \in E$, and
\item \label{cond_b} $W_1(P(x, .), P(y, .) ) \leq r d(x, y)$ for all $x, y \in E$ and some $r < 1$.
\end{enumerate}
Then there exists a unique invariant probability measure $\pi$ for the Markov chain associated to $P$, and the measures
$P^n(x, .)$ and $\pi$ satisfy $\alpha'(\mathcal{T}_d) \leq H$, where $\alpha'(t) = \frac{1}{1 - r} \alpha( (1 - r) t)$.
Moreover, under these hypotheses, the conditions of Theorem \ref{theorem_concentration_dependent_sequences} are verified
with $S = \frac{r}{1-r}$ so that the law $P_n$ of the $n$-uple $(X_1, \ldots, X_n)$ under $X_0 = x_0 \in E$
verifies $\tilde{\alpha}(\mathcal{T}_d) \leq H$ where $\tilde{\alpha}(t) = n \alpha(\frac{1 - r}{n} t)$.
\end{prop}
\begin{proof}
The first claim is obtained exactly as in the proof of Proposition 2.10 in \cite{djellout_guillin_wu}, observing that
the contraction condition \ref{cond_b} is equivalent to
\begin{equation*}
W_1(\nu_1 P, \nu_2 P) \leq r W_1( \nu_1, \nu_2) \quad \text{for all } \nu_1, \nu_2 \in \mathcal{P}_1(E)
\end{equation*}
and also to
\begin{equation*}
\| Pf \|_\text{Lip} \leq r \|f\|_\text{Lip} \quad \text{for all } f.
\end{equation*}
This entails that whenever $f$ is $1$-Lipschitz, $P^n f$ is $r^n$-Lipschitz. Now, by condition \ref{cond_a}, we have
\begin{equation*}
\begin{aligned}
P^n(e^{s f}) & \leq P^{n-1}\left( \exp \left( sPf + \alpha^\circledast(s) \right) \right) \\
{} & \leq P^{n-2} \left( \left( s P^2 f + \alpha^\circledast(s) + \alpha^\circledast(rs) \right) \right) \\
{} & \leq \ldots \\
{} & \leq \exp \left( \left( s P^n f + \alpha^\circledast (s) + \ldots + \alpha^\circledast (r^n s) \right) \right).
\end{aligned}
\end{equation*}
As $\alpha^\circledast$ is convex and vanishes at $0$, we have $\alpha^\circledast (r t) \leq r \alpha^\circledast (t)$ for
all $t \geq 0$. Thus,
\begin{equation*}
P^n(e^{s f} \leq \exp \left( s P^n f + \sum_{k = 0}^{ + \infty} r^k \alpha^\circledast (s) \right) = \exp \left( s P^n f + \frac{1}{1 - r} \alpha^\circledast (s) \right)
\end{equation*}
It remains only to check that $\frac{1}{1 - r} \alpha^\circledast$ is the monotone conjugate of $\alpha'$ and to invoke Proposition \ref{bobkov_gotze_characterization}.
Moving on to the final claim, since the process is homogeneous, to ensure that (\ref{contraction_property}) is satisfied,
we need only show that for all $k \geq 1$, for
all $f : E^k \rightarrow \mathbb{R}$ $1$-Lipschitz w.r.t. $d_1$, the function
\begin{equation*}
x \mapsto \mathbb{E} \left[f(X_1, \ldots, X_k)|X_0 = x \right]
\end{equation*}
is $\frac{r}{1 - r}$-Lipschitz.
We show the following : if $g : E^2 \rightarrow \mathbb{R}$ is
such that for all $x_1, x_2 \in E$ the functions
$g(., x_2)$, resp. $g(x_1, .)$, are $1$-Lipschitz, resp. $\lambda$-Lipschitz,
then the function
\begin{equation*}
x_1 \mapsto \int g(x_1, x_2) P(x_1, d x_2)
\end{equation*}
is $(1 + \lambda r)$-Lipschitz.
Indeed,
\begin{align*}
| \int g(x_1, x_2) & P(x_1, d x_2) - \int g(y_1, x_2) P(y_1, d x_2) | \\
{} & \leq \int |g(x_1, x_2) - g(y_1, x_2) |P(x_1, d x_2) \\
{} & \quad + |\int g(y_1, x_2) (P(x_1, d x_2) - P(y_1, d x_2)) | \\
{} & \leq (1 + \lambda r) d(x_1, y_1).
\end{align*}
It follows easily by induction that the function
\begin{equation*}
f_k : x_1 \mapsto \int f(x_1, \ldots, x_k) P(x_{k-1}, d x_k) \ldots P(x_1, d x_2)
\end{equation*}
has Lipschitz norm bounded by $1 + r + \ldots r^k \leq \frac{1}{1 - r}$.
Hence the function $x \mapsto \int f_k(x_1) P(x, d x_1)$ has Lipschitz norm bounded by $\frac{r}{1 - r}$.
But this function is precisely
\begin{equation*}
x \mapsto \mathbb{E} \left[f(X_1, \ldots, X_k)|X_0 = x \right]
\end{equation*}
and the proof is complete.
\end{proof}
We are in position to prove the analogue of Lemma \ref{lemma_mean} in the Markov case.
\begin{lemma}
Consider the Markov chain associated to a transition kernel $P$ as in Proposition \ref{marginals_markov}.
Let $P_n$ denote the law of the Markov path $(X_1, \ldots, X_n)$ associated to $P$ under $X_0 = x_0$.
Introduce the averaged occupation measure $\pi_n = \mathbb{E}_{P_n} (L_n)$ and the invariant measure $\pi$.
Let $m_1 = \int d(x, x_0) \pi(dx)$.
Suppose that there exists $a > 0$ such that for all $i \geq 1$
$E_{a, i} = \int e^{a d(x, x_0)} P^i(d x) \leq 2$.
Let $\delta > 0$ and $K \in E$ be a compact subset containing $x_0$.
Let $\mathcal{N}_\delta$ denote the metric entropy of order
$\delta$ for the set $\mathcal{F}_K$ of $1$-Lipschitz functions on $K$
vanishing at $x_0$ (endowed with the uniform distance).
Also define $\sigma : [0, +\infty ) \rightarrow [1, + \infty)$ as the inverse function of $x \mapsto x \ln x - x + 1$ on $[1, + \infty)$.
The following holds :
\begin{align*}
\mathbb{E}_{P_n}[ W_1(L_n, \pi_n)] & \leq 2 \delta + \frac{8}{a} \frac{1}{n}\sum_{i = 1}^{n}\frac{1}{\sigma (\frac{1}{P^i(K^c)})} + \Gamma(\mathcal{N}_\delta, n) \\
W_1( \pi_n, \pi) & \leq \frac{m_1}{n(1 - r)}.
\end{align*}
where $\Gamma(\mathcal{N}_\delta, n) = \inf_{\lambda > 0} \frac{1}{\lambda} \left[ \log \mathcal{N}_\delta + n \alpha^\circledast (\frac{\lambda}{n(1 - r)}) \right] $
\end{lemma}
\begin{proof}
Convergence to the equilibrium measure is dealt with using the contraction hypothesis. Indeed, by convexity of the map
$\mu \mapsto W_1(\mu, \pi)$, we first have
\begin{equation*}
W_1(\pi_n, \pi) \leq \frac{1}{n} \sum_{i = 1}^n W_1( P^i(x_0, .), \pi).
\end{equation*}
Now, using that the contraction property (\ref{cond_b}) in Proposition \ref{marginals_markov} is equivalent to the inequality
$W_1( \mu_1 P, \mu_2 P) \leq r W_1( \mu_1, \mu_2)$ for all $\mu_1, \mu_2 \in \mathcal{P}_1(E)$, and using the fact that $\pi$
is $P$-invariant,
\begin{equation*}
W_1(\pi_n, \pi) \leq \frac{1}{n} \sum_{i = 1}^n r^i W_1( \delta_{x_0}, \pi) \leq \frac{W_1( \delta_{x_0}, \pi)}{n(1 - r)} = \frac{m_1}{n(1 - r)}.
\end{equation*}
In order to take care of the second term, we will use the same strategy (and notations) as in the independent case. Introduce once again a compact subset
$K \subset E$ and a covering of $\mathcal{F}_K$ by functions $f_1, \ldots, f_{\mathcal{N}_\delta}$ suitably extendend to $E$.
With the same arguments as before, we get
\begin{align*}
\mathbb{E}_{P_n}W_1( L_n, \pi_n) \leq & \mathbb{E}_{P_n} (\max_{j = 1, \ldots, \mathcal{N}_\delta} \Psi(f_j)) + 2 \delta + 2 \int d(x, x_0) 1_{K^c} d \pi_n \\
{} & + 2 \mathbb{E}_{P_n} (\int d(x, x_0) 1_{K^c} d L_n)
\end{align*}
Then,
\begin{equation*}
\int d(x_0, y) \pi_n(d y) = \frac{1}{n} \sum_{i = 1}^n \int d(x_0, y)1_{K^c} P^i(x_0, d y).
\end{equation*}
As before we can use Orlicz-H\"{o}lder's inequality to recover the bound
\begin{equation*}
\int d(x_0, y) d \pi_n \leq \frac{2}{a} \frac{1}{n} \sum_{i = 1}^{n} \frac{1}{\sigma (\frac{1}{P^i(K^c)})}.
\end{equation*}
And likewise,
\begin{align*}
\mathbb{E} (\int d(x, x_0) 1_{K^c} d L_n) & = \mathbb{E} \displaystyle \left[ \frac{1}{n} \sum_{i = 1}^n d(x_0, X_i) 1_{K^c} \right] \\
{} & = \frac{1}{n} \sum_{i = 1}^n \int d(x_0, y) 1_{K^c} P^i(x_0, dy)
\end{align*}
and we have the same bound as above.
As for the last term remaining : it will be possible to use the maximal inequality techniques just as in the proof of Theorem \ref{main_thm},
provided that we can find bounds on the terms $\mathbb{E} \displaystyle \left[ \exp \lambda \Psi (f_j) \right]$, where this time
\begin{equation*}
\Psi(f) = \int f d L_n - \int f d \pi_n.
\end{equation*}
Denote
\begin{equation*}
F_j(x_1, \ldots, x_n) = \frac{1}{n} \sum_{i = 1}^n f_j(x_i).
\end{equation*}
This is a $\frac{1}{n}$-Lipschitz function on $E^n$. Since $P_n$ satisfies a $\tilde{\alpha}(\mathcal{T}_d) \leq H$ inequality,
we have
\begin{equation*}
\int \exp \lambda F_j d P_n \leq \exp \left[ \lambda \int F_j d P_n + n \alpha^\circledast (\frac{\lambda}{n(1 - r)}) \right].
\end{equation*}
But this is exactly the bound
\begin{equation*}
\mathbb{E} \displaystyle \left[ \exp \lambda \Psi(f_j) \right] \leq e^{n \alpha^\circledast (\frac{\lambda}{n(1 - r)})}.
\end{equation*}
We may then proceed as in the independent case and obtain
\begin{equation*}
\mathbb{E} [ \max_{j = 1, \ldots, \mathcal{N}_\delta} \Psi(f_j) ] \leq \inf_{\lambda > 0} \frac{1}{\lambda} \left[ \log \mathcal{N}_\delta + n \alpha^\circledast (\frac{\lambda}{n(1 - r)}) \right].
\end{equation*}
\end{proof}
For any Lipschitz function $f : E^n \rightarrow \mathbb{R}$ (w.r.t. $d_1$), we have the concentration inequality
\begin{equation*}
P_n (x \in E^n, \quad f(x) \geq \int f d P_n + t) \leq \exp - n \alpha \left( \frac{ (1-r)t}{ n \|f\|_\text{Lip}} \right).
\end{equation*}
Remembering that $E^n \ni x \mapsto W_1( L_n^x, \pi_n)$ is $\frac{1}{n}$-Lipschitz, we get the bound
\begin{equation} \label{concentration_markov}
\mathbb{P}( W_1(L_n, \pi_n) \geq \mathbb{E}_{P_n}[ W_1(L_n, \pi_n) ] + t ) \leq \exp - n \alpha \left( (1 - r)t\right).
\end{equation}
Thanks to the triangular inequality $W_1(L_n, \pi_n) \geq W_1(L_n, \pi) - W_1(\pi_n, \pi)$, it holds that
\begin{equation} \label{concentration_ergodic}
\mathbb{P} ( W_1( L_n, \pi) \geq W_1(\pi_n, \pi) + \mathbb{E}_{P_n} [W_1(L_n, \pi_n)] + t) \leq \exp - n \alpha \left((1 - r)t \right).
\end{equation}
This in turn leads to an estimate on the deviations, under
the condition that we may exhibit a compact set with large measure for all the measures $P^i$.
We now move on to the proof of Theorem \ref{theorem_markov}.
\begin{proof}[Proof of Theorem \ref{theorem_markov}]
Fix $\delta = t/8$.
Set $m_1^i = \int |x| P^i(d x)$. We have
\begin{align*}
m_1^i & \leq m_1 + W_1( P^i, \pi) \\
{} & \leq m_1 + r^i W_1( \delta_0, \pi) \\
{} & \leq 2 m_1.
\end{align*}
Thus
\begin{equation*}
\int e^{a |x|} P^i(dx) \leq e^{a m_1^i + C a^2/4} \leq e^{2 m_1 a + C a^2 / 4}.
\end{equation*}
With $a$ as in the theorem, the above ensures that
$\int e^{a |x|} P^i(dx) \leq 2$.
Let $B_R$ denote the ball of center $0$ and radius $R$ :
we have $P^i(B_R^c) \leq 2 e^{- a R}$. Let
\begin{equation*}
R = \frac{1}{a} \log 2 \sigma^{-1}(\frac{32}{ a t}).
\end{equation*}
so that $ 2 \delta + \frac{8}{a} \frac{1}{n}\sum_{i = 1}^{n}\frac{1}{\sigma (\frac{1}{P^i(K^c)})} \leq t/2$.
As $\alpha(t) = \frac{1}{C} t^2$ we can compute
\begin{equation*}
\Gamma(\mathcal{N}_\delta, n) = \frac{1}{1-r}\sqrt{\frac{C}{n}} \sqrt{\log \mathcal{N}_\delta}.
\end{equation*}
We have chosen $K = B_R$ and $\delta = t/8$.
Working as in the proof of Proposition \ref{result_R}, when $t \leq 2/a$, we can bound $\log \mathcal{N}_\delta$ by
\begin{equation*}
\log \mathcal{N}_\delta \leq C_d (\frac{1}{at} \log \frac{1}{at})^d
\end{equation*}
where $C_d$ is a numerical constant depending on the dimension $d$.
Plugging the above estimates in (\ref{concentration_ergodic}) and using the inequality $(u-v)^2 \geq u^2/2 - v^2$
gives the desired result.
\end{proof}
\section{Introduction}
\subsection{Generalities}
In the whole paper, $(E, d)$ will denote a Polish space with metric $d$, equipped with its Borel $\sigma$-field and $\mathcal{P}(E)$ will denote
the set of probability measures over $E$.
Consider $\mu \in \mathcal{P}(E)$ and
a sequence of i.i.d. variables $X_i$, $1 \leq i \leq n$, with common law $\mu$. Let
\begin{equation}
L_n = \frac{1}{n} \sum_{i = 1}^{n} \delta_{X_i}
\end{equation}
denote the empirical measure
associated with the i.i.d. sample $(X_i)_{1 \leq i \leq n}$, then with probability 1, $L_n \rightharpoonup \mu$ as $n \rightarrow + \infty$ (here
the arrow denotes narrow convergence, or convergence against all
bounded continuous functions over $E$). This theorem is known as the empirical law of large number or Glivenko-Cantelli theorem and is due in this form to Varadarajan
\cite{varadarajan1958convergence}.
Quantifying the speed of convergence for an appropriate notion of distance between probability measures is an old problem, with notable importance in statistics.
For many examples, we refer to the book of Van der Vaart and Wellner \cite{van1996weak} and the Saint-Flour course of P.Massart \cite{massart1896concentration}.
Our aim here is to
study non-asymptotic deviations in $1$-Wasserstein distance. This is a problem of interest in the fields of
statistics and numerical probability.
More specifically, we provide bounds for the quantity
$\mathbb{P}( W_1( L_n, \mu) \geq t)$
for $t > 0$, i.e. we quantify the speed of convergence of the variable $W_1( L_n, \mu)$ to $0$ in probability.
This paper seeks to complement the work of F.Bolley, A.Guillin and C.Villani in \cite{quantit_conc_ineq}
where such estimates are obtained for measures supported in $\mathbb{R}^d$. We sum up (part of) their result here.
Suppose that $\mu$ is a probability measure on
$\mathbb{R}^d$ for $1 \leq p \leq 2$ that satisfies a $\mathbf{T}_p(C)$ transportation-entropy inequality, that is
\begin{equation*}
W_p(\nu, \mu) \leq \sqrt{C H(\nu | \mu)} \text{ for all } \nu \in \mathcal{P}_p(\mathbb{R}^d)
\end{equation*}
(see below for definitions).
They obtain a
non-asymptotic Gaussian deviation estimate for the $p-$Wasserstein distance between the empirical and true measures~:
\begin{equation*}
\mathbb{P} (W_p(L_n, \mu) \geq t) \leq C(t) \exp ( - K n t^2).
\end{equation*}
This is an effective result : the constants $K$ and $C(t)$ may be explicitely computed
from the value of some square-exponential moment of $\mu$ and the constant $C$ appearing in the transportation inequality.
The strategy used in \cite{quantit_conc_ineq} relies on a non-asymptotic version of (the upper bound in) Sanov's theorem.
Roughly speaking, Sanov's theorem states that the proper rate function for the deviations of empirical measures is the entropy functional,
or in other words that for 'good' subsets $A \in \mathcal{P}(E)$,
\begin{equation*}
\mathbb{P} (L_n \in A) \asymp e^{- n H(A | \mu)}
\end{equation*}
where $H(A | \mu) = \inf_{\nu \in A} H( \nu | \mu)$ (see \cite{dembo1993large} for a full statement of the theorem).
In a companion work \cite{boissard2011bounding}, we derive sharper bounds for this problem, using a construction originally due to R.M. Dudley \cite{dudley1969speed}.
The interested reader may refer to \cite{boissard2011bounding} for a summary of existing results.
Here, our purpose is to show that in the case $p = 1$, the results of \cite{quantit_conc_ineq} can be recovered with simple
arguments of measure concentration, and to give various extensions of interest.
\begin{itemize}
\item We would like to consider spaces more general than $\mathbb{R}^d$.
\item We would like to encompass a wide class of measures in a synthetic treatment. In order to do so we will consider more general transportation inequalities, see below.
\item Another interesting feature is to extend the result to dependent sequences such as the occupation measure of a Markov chain.
This is a particularly desirable feature in applications : one may wish to approximate a distribution that is unknown, or from which it is
practically impossible to sample uniformly, but that is known to be
the invariant measure of a simulable Markov chain.
\end{itemize}
\emph{Acknowledgements.} The author thanks his advisor Patrick Cattiaux for suggesting the problem and for his advice. Arnaud Guillin is also thanked for enriching conversations.
\bigskip
In the remainder of this section, we introduce the tools necessary in our framework : transportation distances and transportation-entropy inequalities.
In Section \ref{results}, we give our main results, as well as explicit estimates in several relevant cases. Section \ref{proof_of_main_thm} is devoted
to the proof of the main result. Section \ref{section_variant} is devoted to the proof of Theorem \ref{variant}.
In Section \ref{dependent_case} we show how our strategy of proof can extend to the dependent case.
\subsection{A short introduction to transportation inequalities}
\subsubsection{Transportation costs and Wasserstein distances}
We recall here basic definitions and propositions ; for proofs and a thorough account of this rich theory,
the reader may refer to \cite{optimal_transport_villani}.
Define $\mathcal{P}_p$, $ 1 \leq p < + \infty$, as the set of probability measures with a
finite $p$-th moment. The $p$-Wasserstein metric $W_p(\mu, \nu)$ between $\mu, \nu \in \mathcal{P}_p$ is defined by
\begin{equation*}
W_p^p(\mu, \nu) = \text{inf} \int d^p(x,y) \pi(dx, dy)
\end{equation*}
where the infimum is on probability measures $\pi \in \mathcal{P}(E \times E)$ with marginals $\mu$ and $\nu$.
The topology induced by this metric is slightly stronger than the weak topology : namely,
convergence of a sequence $(\mu_n)_{n \in \mathbb{N}}$ to a measure $\mu \in \mathcal{P}_p$ in the $p$-Wasserstein metric is equivalent to
the weak convergence of the sequence plus a uniform bound on the $p$-th moments of the measures $\mu_n$, $n \in \mathbb{N}$.
We also recall the well-known Kantorovich-Rubinstein dual characterization of $W_1$ : let $\mathcal{F}$ denote the set of $1$-Lipschitz functions
$f : E \rightarrow \mathbb{R}$ that vanish at some fixed point $x_0$. We have :
\begin{equation} \label{kantorovich_rubinstein}
W_1(\mu, \nu) = \inf_{f \in \mathcal{F}} \int f d \mu - \int f d \nu.
\end{equation}
\subsubsection{Transportation-entropy inequalities}
For a very complete overview of the subject, the reader is invited to consult the review \cite{gozlan2010transport}.
More facts and criteria are gathered in Appendix \ref{appendix_transport_inequalities}
For $\mu, \nu \in \mathcal{P}(E)$, define the relative entropy $H(\nu | \mu)$ as
\begin{equation*}
H( \nu | \mu) = \int_E \log \frac{d \nu}{d \mu} \nu( d x)
\end{equation*}
if $\nu$ is absolutely continuous relatively to $\mu$, and $H( \nu | \mu) = + \infty$ otherwise.
Let $\alpha : [0, \ + \infty) \rightarrow \mathbb{R}$ denote a convex, increasing, left-continous function such that $\alpha(0) = 0$.
\begin{definition}
We say that $\mu \in \mathcal{P}_p(E)$ satisfies a $\mathbf{T}_p(C)$ inequality
for some $C > 0$ if
for all $\nu \in \mathcal{P}_p(E)$,
\begin{equation}
W_p(\mu, \nu) \leq \sqrt{C H( \nu | \mu) }.
\end{equation}
We say that $\mu \in \mathcal{P}(E)$ satisfies a $\alpha(\mathcal{T}_d)$ inequality if for all $\nu \in \mathcal{P}(E)$,
\begin{equation} \label{t_c_eq}
\alpha( W_1(\mu, \nu)) \leq H(\nu | \mu).
\end{equation}
\end{definition}
Observe that $\mathbf{T}_1(C)$ inequalities are particular cases of $\alpha(\mathcal{T}_d)$ inequalities with
$\alpha(t) = \frac{1}{C} t^{2/p}$.
From here on, our focus will be on $\alpha(\mathcal{T}_d)$ inequalities.
\section{Results and applications} \label{results}
\subsection{General bounds in the independent case}
Let us first introduce some notation :
if $K \subset E$ is compact and $x_0 \in K$, we define the set $\mathcal{F}_K$ of $1$-Lipschitz functions over $K$ vanishing at $x_0$, which is
is also compact w.r.t. the uniform distance (as a consequence of the Ascoli-Arzela theorem). We will also need the following definition~:
\begin{definition}
Let $(A, d)$ be a totally bounded metric space. For every $\delta > 0$, define the \textit{covering number}
$\mathcal{N}(A, \delta)$ of order $\delta$ for $A$
as the minimal number of balls of radius $\delta$ needed to cover $A$.
\end{definition}
We state our first result in a fairly general fashion.
\begin{theorem} \label{main_thm}
Suppose that $\mu \in \mathcal{P}(E)$ satisfies a $\alpha(\mathcal{T}_d)$ inequality. Let $a > 0$ be such that
$E_{a, 1} = \int e^{a d(x_0, x)} \mu (dx) \leq 2$.
Choose a compact $K \subset E$ such that
\begin{equation*}
\mu(K^c) \leq \left[ \frac{32}{a t} \log \frac{32}{a t} - \frac{32}{a t} + 1 \right]^{-1}.
\end{equation*}
Denote
\begin{equation}
\mathcal{C}_t = \mathcal{N}(\mathcal{F}_K, t/8).
\end{equation}
We have
\begin{equation} \label{main_eq_bound}
\mathbb{P}( W_1(L_n, \mu) \geq t) \leq \exp -n \alpha \left[ t/2 -\Gamma(\mathcal{C}_t, n) \right]
\end{equation}
where $\Gamma(\mathcal{C}_t, n) = \inf_{\lambda > 0} 1/\lambda[ \log \mathcal{C}_t + n \alpha^\circledast (\lambda/n) ]$,
and with the convention that $\alpha(x) = 0$ for $x < 0$.
\end{theorem}
\begin{remark}
With a mild change in the proof, one may replace in (\ref{main_eq_bound}) the term $t/2$ by $ct$ for any $c < 1$,
with the trade-off of choosing a larger compact set, and thus a larger value of $\mathcal{C}_t$. For the sake of readability we do not
make further mention of this.
\end{remark}
The result in its general form is abtruse, but it yields interesting results as soon as one knows more about $\alpha$. Let us give a few examples.
\begin{corollary} \label{corollary_bt}
If $\mu$ satisfies $\mathbf{T}_1(C)$, we have
\begin{equation*}
\mathbb{P}( W_1(L_n, \mu) \geq t) \leq \mathcal{C}_t \exp - \frac{1}{8 C} n t^2.
\end{equation*}
\end{corollary}
\begin{corollary} \label{corollary_2}
Suppose that $\mu$ verifies the modified transport inequality
\begin{equation*}
W_1(\nu, \mu) \leq C \left( H(\nu | \mu) + \sqrt{H(\nu | \mu)} \right)
\end{equation*}
(as observed in paragraph \ref{integral_criteria}, this is equivalent to the finiteness of an exponential moment for $\mu$). Then, for $t \leq C/2$,
\begin{equation*}
\mathbb{P}( W_1(L_n, \mu) \geq t) \leq A(n, t) \exp - \frac{(\sqrt{2} - 1)^2}{2 C^2} n t^2
\end{equation*}
where
\begin{equation*}
A(n, t) = \exp \displaystyle \left[ 4 (\sqrt{2} - 1)^2 n (\sqrt{1 + \frac{n}{\log \mathcal{C}_t}} -1)^{-2} \right]
\end{equation*}
(observe that $A(n, t) \rightarrow \mathcal{C}_t$ when $n \rightarrow + \infty$).
\end{corollary}
\begin{proof}[Proof of Corollary \ref{corollary_bt}]
In this case, we have $\alpha(t) = \frac{1}{C}t^2$, and so
\begin{equation*}
\Gamma(\mathcal{C}_t, n) = \sqrt{\frac{C \log \mathcal{C}_t}{n}},
\end{equation*}
so that we get
\begin{equation*}
\mathbb{P}( W_1(L_n, \mu) \geq t) \leq \exp -\frac{n}{C} (\frac{t}{2} - \sqrt{\frac{C \log \mathcal{C}_t}{n}})^2
\end{equation*}
and conclude with the elementary inequality $(a - b)^2 \geq \frac{1}{2}a^2 - b^2$.
\end{proof}
\begin{proof}[Proof of Corollary \ref{corollary_2}]
Here, $\alpha(x) = \frac{1}{4}( \sqrt{1 + \frac{4 x}{C}} - 1)^2$, and one can get the bound
\begin{equation*}
\Gamma(\mathcal{C}_t, N) \leq \frac{C}{\sqrt{1 + \frac{N}{\log \mathcal{C}_t}} - 1}.
\end{equation*}
By concavity of the square root function, for $u \leq 1$, we have $\sqrt{1 + u} - 1 \geq (\sqrt{2}-1)u$. Thus, for $t \leq \frac{C}{2}$, we have
\begin{eqnarray*}
\alpha( \frac{t}{2} - \Gamma(\mathcal{C}_t, N) ) & \geq & \frac{(\sqrt{2} - 1)^2}{4} ( \frac{2}{C}t - \frac{4}{\sqrt{1 + \frac{N}{\log \mathcal{C}_t}} -1})^2 \\
{} & \geq & \frac{(\sqrt{2} - 1)^2}{2 C^2} t^2 - 4 (\sqrt{2} - 1)^2(\sqrt{1 + \frac{N}{\log \mathcal{C}_t}} -1)^{-2}
\end{eqnarray*}
(in the last line we have used again the inequality $(a-b)^2 \geq \frac{a^2}{2} - b^2$). This in turn gives the announced result.
\end{proof}
Our technique of proof, though related to the one in \cite{quantit_conc_ineq},
is based on different arguments : we make use of
the tensorization properties of transportation inequalities
as well as the estimates (\ref{cond_laplace}) in the spirit of Bobkov-G\"{o}tze,
instead of a Sanov-type bound.
The notion that is key here is the phenomenon of concentration of measure (see e.g. \cite{ledoux2001concentration}) : its relevance in statistics was put forth
very explicitely in \cite{massart1896concentration}.
We may sum up our approach as follows :
first, we rely on existing tensorization results to obtain concentration of $W_1(L_n, \mu)$
around its mean $\mathbb{E}[W_1(L_n, \mu)]$, and in a second time we estimate the decay of
the mean as $n \rightarrow + \infty$. Despite technical difficulties, the arguments are mostly elementary.
The next theorem is a variation on Corollary \ref{corollary_bt}. Its proof is based on different arguments, and it is postponed to Section \ref{section_variant}.
We will use this theorem to obtain bounds for Gaussian measures in Theorem \ref{variant}.
\begin{theorem} \label{variant}
Let $\mu \in \mathcal{P}(E)$ satisfy a $\mathbf{T}_1(C)$ inequality. Then :
\begin{equation*}
\mathbb{P}( W_1( \mu, L_n) \geq t) \leq K_t e^{-nt^2/8C}
\end{equation*}
where
\begin{equation*}
K_t = \exp \left[ \frac{1}{C} \inf_\nu \text{Card }(\text{Supp } \nu) (\text{Diam Supp } \nu)^2 \right]
\end{equation*}
and $\nu$ runs over all probability measures with finite support such that $W_1( \mu, \nu ) \leq t/4$.
\end{theorem}
\begin{remark}
As earlier, we could improve the factor $1/8C$ in the statement above to any constant $c < 1/C$, with the trade-off of a larger constant $K_t$.
\end{remark}
\subsection{Comments}
We give some comments on the pertinence of the results above. First of all, we argue that the asymptotic order of magnitude
of our estimates is the correct one. The term ``asymptotic'' here means
that we consider the regime $n \rightarrow + \infty$, and the relevant tool in this setting is Sanov's large deviation principle
for empirical measures.
A technical point needs to be stressed : there are several variations of Sanov's theorem, and the most common ones (see e.g. \cite{dembo1993large})
deal with the weak topology on probability measures. What we require is a version of the principle that holds for
the stronger topology induced by the $1$-Wasserstein metric, which leads to
slightly more stringent
assumptions on the measure than in Theorem \ref{main_thm}. With this in mind, we quote the following result from Wang \cite{wang2010sanov} :
\begin{prop}
Suppose that $\mu \in \mathcal{P}(E)$ satisfies
$\int e^{a d(x, x_0)} \mu(dx) < + \infty$
for all $a > 0$ and some $x_0 \in E$,
and a $\alpha(\mathcal{T}_d)$ inequality.
Then :
\begin{itemize}
\item for all $A \subset \mathcal{P}(E)$ closed for the $W_1$ topology,
\begin{equation*}
\limsup_{n \rightarrow + \infty} \frac{1}{n} \log \mu(A) \leq - \inf_{ \nu \in A} H(\nu | \mu)
\end{equation*}
\item for all $B \subset \mathcal{P}(E)$ open w.r.t. the $W_1$ topology,
\begin{equation*}
\liminf_{n \rightarrow + \infty} \frac{1}{n} \log \mu(B) \geq - \inf_{\nu \in B} H(\nu | \mu).
\end{equation*}
\end{itemize}
\end{prop}
Consider the closed set $A = \{ \nu \in \mathcal{P}(E), \, W_1(\mu, \nu) \geq t \}$, then we have according to the above
\begin{equation*}
\limsup_{n \rightarrow + \infty} \frac{1}{n} \log \mathbb{P}(W_1(L_n, \mu) \geq t) \leq - \alpha(t).
\end{equation*}
With Theorem \ref{main_thm} (and the remark following it), we obtain the bound
\begin{equation*}
\limsup_{n \rightarrow + \infty} \frac{1}{n} \log \mathbb{P}(W_1(L_n, \mu) \geq t) \leq - \alpha(ct)
\end{equation*}
for all $c < 1$, and since $\alpha$
is left-continuous, we indeed obtain the same asymptotic bound as from Sanov's theorem.
\bigskip
Let us come back to the non-asymptotic regime.
When we assume for example a $\mathbf{T}_1$ inequality, we get a bound
in the form $\mathbb{P} ( W_1( L_n, \mu) \geq t ) \leq C(t) e^{- C n t ^2}$ involving
the large constant $C(t)$. By the Kantorovich-Rubinstein dual formulation of $W_1$, this amounts to
simultaneous deviation
inequalities for all $1$-Lipschitz observables. We recall briefly the well-known fact
that it is fairly easy to obtain a deviation inequality for one
Lipschitz observable without a constant depending on the deviation scale $t$.
Indeed, consider a $1$-Lipschitz function $f$ and a sequence
$X_i$ of i.i.d. variables with law $\mu$. By Chebyshev's bound, for $\theta > 0$,
\begin{eqnarray*}
\mathbb{P} ( \frac{1}{n} \sum f(X_i) - \int f \mu \geq \varepsilon) & \leq & \exp - n [ \theta \varepsilon - \log (\int e^{\theta f(x)} \mu (dx) e^{- \theta \int f \mu}) ]
\end{eqnarray*}
According to Bobkov-G\"{o}tze's dual characterization of $\mathbf{T}_1$, the term inside the $\log$ is bounded above by $e^ {C \theta^2}$, for some positive $C$, whence
$\mathbb{P} ( \frac{1}{n} \sum f(X_i) - \int f \mu \geq \varepsilon) \leq \exp - n [ \theta \varepsilon - C \theta^2 ]$.
Finally, take $\theta = \frac{1}{2 C} \varepsilon$ to get
\begin{equation*}
\mathbb{P} ( \frac{1}{n} \sum f(X_i) - \int f \mu \geq \varepsilon) \leq e^{- C n t^2/2}.
\end{equation*}
Thus, we may see the multiplicative constant that we obtain as a trade-off
for the obtention of uniform deviation estimates on all Lipschitz observables.
\bigskip
\subsection{Examples of application}
For practical purposes, it is important to give the order of magnitude
of the multiplicative constant $\mathcal{C}_t$ depending on $t$.
We address this question on several important examples in this paragraph.
\subsubsection{The $\mathbb{R}^d$ case}
\begin{example} \label{result_R}
Denote $\theta(x) = 32 x \log \left[ 2 \left( 32 x \log 32 x - 32 x + 1 \right) \right]$.
In the case $E = \mathbb{R}^d$, the numerical constant $\mathcal{C}_t$ appearing in Theorem \ref{main_thm}
satisfies :
\begin{equation} \label{real_case}
\mathcal{C}_t \leq 2 \left( 1 + \theta ( \frac{1}{a t} ) \right) 2^{ \displaystyle C_d \theta( \frac{1}{a t} )^d }
\end{equation}
where $C_d$ only depends on $d$. In particular, for all $t \leq \frac{1}{2 a}$, there exist numerical constants $C_1$
and $C_2$ such that
\begin{equation*}
\mathcal{C}_t \leq C_1 (1 + \frac{1}{at} \log \frac{1}{at}) e^{ \displaystyle C_d C_2^d (\frac{1}{at} \log \frac{1}{a t})^d}.
\end{equation*}
\end{example}
\begin{remark}
The constants $C_d$, $C_1$, $C_2$ may be explicitely determined from the proof. We do not do so and only state that $C_d$ grows exponentially with $d$.
\end{remark}
\begin{proof}
For a measure $\mu \in \mathcal{P}(\mathbb{R}^d)$, a convenient natural choice for a compact set of large measure is a
Euclidean ball. Denote $B_R = \{x \in \mathbb{R}^d, |x| \leq R \}$. We will denote by $C_d$ a constant depending only on
the dimension $d$, that may change from line to line. Suppose that
$\mu$ satisfies the assumptions in Theorem \ref{main_thm}. By Chebyshev's bound,
$\mu(B_R^c) \leq 2 e^{-a R}$, so we may choose $K = B_{R_t}$ with
\begin{equation*}
R_t \geq \frac{1}{a} \log \left[ 2 \left(\frac{32}{a t} \log \frac{32}{ a t} - \frac{32}{a t} + 1 \right) \right] .
\end{equation*}
Next, the covering numbers for $B_R$ are bounded by :
\begin{equation*}
\mathcal{N}(B_R, \delta) \leq C_d \left( \frac{R}{\delta} \right)^d.
\end{equation*}
Using the bound (\ref{covering_number_lip_2}) of Proposition \ref{prop_covering_number_lip}, we have
\begin{equation*}
\mathcal{C}_t \leq \left( 2 + 2 \lfloor \frac{32 R_t}{t} \rfloor \right) 2^{ \displaystyle C_d \left(\frac{ 32 R_t}{t} \right)^d }.
\end{equation*}
This concludes the proof for the first part of the proposition.
The second claim derives from the fact that for $x > 2$, there exists
a numerical constant $k$ such that
$\theta(x) \leq k x \log x$.
\end{proof}
Example \ref{result_R} improves slightly upon the result for the $W_1$ metric in \cite{quantit_conc_ineq}.
One may wonder whether this order of magnitude is close to optimality. It is in fact not sharp, and we point out
where better results may be found.
In the case $d = 1$,
$W_1(L_n, \mu)$ is bounded above by the Kolmogorov-Smirnov divergence
$\sup_{x \in \mathbb{R}} | F_n(x) - F(x)|$ where $F_n$ and $F$ denote respectively the cumulative distribution functions (c.d.f.)
of $L_n$ and $\mu$. As a consequence of the celebrated Dvorestky-Kiefer-Wolfowitz theorem
(see \cite{massart1990tight}, \cite{van1996weak}), we have the following :
if $\mu \in \mathcal{P}(\mathbb{R})$ has a continuous c.d.f., then
\begin{equation*}
\mathbb{P}(W_1( L_n, \mu) > t ) \leq 2 e^{-2 n t^2}.
\end{equation*}
The behaviour of the Wasserstein distance between empirical and true distribution in one dimension has been very
thoroughly studied by del Barrio, Gin\'{e}, Matran, see \cite{del1999central}.
In dimensions greater than $1$, the result is also not sharp. Integrating (\ref{real_case}), one recovers a bound of the type
$\mathbb{E} (W_1(L_n, \mu)) \leq C n^{-1/(d + 2)} (\log n)^c$. Looking into the proof of our main result, one sees that any improvement of this bound will
automatically give a sharper result than (\ref{real_case}).
For the uniform measure over the unit cube, results have been known for a while. The pioneering work in this framework is
the celebrated article of Ajtai, Komlos and Tusn\'ady \cite{ajtai1984optimal}.
M.Talagrand \cite{talagrand1992matching} showed that when $\mu$ is the uniform distribution on the unit cube
(in which case it clearly satisfies a $\mathbf{T}_1$ inequality) and $d \geq 3$, there exists $c_d \leq C_d$ such that
\begin{equation*}
c_d n^{-1/d} \leq \mathbb{E} W_1(L_n, \mu) \leq C_d n^{-1/d}.
\end{equation*}
Sharp results for general measures are much more recent : as a consequence of the results of F. Barthe and C. Bordenave \cite{barthe2011combinatorial},
one may get an estimate of the type $\mathbb{E} W_1(L_n, \mu) \leq c n^{-1/d}$ under some polynomial moment condition on $\mu$.
\subsubsection{A first bound for Standard Brownian motion}
We wish now to illustrate our results on an infinite-dimensional case. A first natural candidate is the law of the standard Brownian motion,
with the sup-norm as reference metric. The natural idea that we put in place in this paragraph is to choose as large compact sets the $\alpha$-H\"{o}lder
balls, which are compact for the sup-norm. However the remainder of this paragraph serves mainly an illustrative purpose : we will obtain
sharper results, valid for general Gaussian measures on (separable) Banach spaces, in paragraph \ref{subsection_gaussian}.
We consider the canonical Wiener space
$\left( \mathcal{C}([0, 1], \mathbb{R}), \gamma, \| .\|_{\infty} \right)$,
where $\gamma$ denotes the
Wiener measure, under which the coordinate process $B_t : \omega \rightarrow \omega(t)$ is a standard Brownian motion.
\begin{example} \label{example_s_b_m}
Denote by $\gamma$ the Wiener measure on $\left( \mathcal{C}([0, 1], \mathbb{R}), \gamma, \| .\|_{\infty} \right)$, and
for $\alpha < 1/2$, define
\begin{equation*}
C_\alpha = 2^{1+ \alpha}\frac{2^{(1-2 \alpha)/4}}{1 - 2^{4/(1-2 \alpha)}} \|Z\|_{4/(1-2\alpha)}
\end{equation*}
where $\|Z\|_p$ denotes the $L_p$ norm of a $\mathcal{N}(0, 1)$ variable $Z$.
There exists $k > 0$ such that for every $t \leq 144/\sqrt{2 \log 2}$,
$\gamma$ satisfies
\begin{equation*}
\mathbb{P}( W_1( L_n, \gamma) \geq t) \leq \mathcal{C}_t e^{-n t^2/64}
\end{equation*}
with
\begin{equation*}
\mathcal{C}_t \leq \exp \exp (k C_\alpha \frac{\sqrt{\log 1/t}}{t})^{1/\alpha}.
\end{equation*}
\end{example}
\begin{proof}
For $0 < \alpha \leq 1$, define the $\alpha$-H\"{o}lder semi-norm as
\begin{equation*}
| x|_\alpha = \sup_{t,s \in [0, 1]} \frac{|x(t) - x(s)|}{|t-s|^\alpha}.
\end{equation*}
Let $0 < \alpha \leq 1$ and denote by $C_\alpha$ the Banach space of $\alpha$-H\"{o}lder continuous functions vanishing at $0$, endowed with the norm
$\|.\|_\alpha$.
It is a classical fact that the Wiener measure is
concentrated on $C_\alpha$ for all $\alpha \in ]0, 1/2[$.
By Ascoli-Arzela's theorem, $C_\alpha$ is compactly embedded in $\mathcal{C}([0, 1], \mathbb{R})$, or in other words
the $\alpha$-H\"{o}lder balls $B_{\alpha, R} = \{ x \in \mathcal{C}([0, 1], \mathbb{R}), \, \|x\|_\alpha \leq R \}$
are totally bounded for the uniform norm. This makes $B(\alpha, R)$ good candidates for compact spaces of large measure.
We need to evaluate how big $B(\alpha, R)$ is w.r.t. $\gamma$.
To this end we use the fact that the Wiener measure is also a Gaussian measure on $C_\alpha$ (see \cite{baldi1992large}).
Therefore Lemma \ref{lemma_gaussian_tail}
applies : denote
\begin{equation*}
m_\alpha = \mathbb{E} \sup_t \| B_t \|_\alpha, \quad s^2_\alpha = \mathbb{E} (\sup_t \| B_t \|_\alpha)^2,
\end{equation*}
we have
\begin{equation*}
\gamma ( B(\alpha, R)^c) \leq 2 e^{- (R - m_\alpha)^2/2 s^2_\alpha}
\end{equation*}
for $R \geq m_\alpha$. Choosing
\begin{equation} \label{ineq_requirement}
R_t \geq m_\alpha + \left[ 2 s^2_\alpha \log 2 ( \frac{32}{at} \log \frac{32}{at} - \frac{32}{at} + 1 )\right]^{1/2}
\end{equation}
guarantees that
\begin{equation*}
\gamma ( B(\alpha, R_t)^c) \leq \left( \frac{32}{at} \log \frac{32}{at} - \frac{32}{at} + 1 \right)^{-1}.
\end{equation*}
On the other hand, according to Corollary \ref{holder_moments_sbm}, $m_\alpha$ and $s_\alpha$ are bounded by $C_\alpha$.
And Lemma \ref{lemma_small_exp_moment_wiener} shows that choosing $a = \sqrt{2 \log 2}/3$ ensures
$\mathbb{E} e^{a \sup_t |B_t| } \leq 2$.
Elementary computations show that for $t \leq 144/\sqrt{2 \log 2}$, we can pick
\begin{equation*}
R_t = 3 C_\alpha\sqrt{ \log(96/(\sqrt{2 \log 2}t))}
\end{equation*}
to comply with the requirement in (\ref{ineq_requirement}).
Bounds for the covering numbers in $\alpha$-H\"{o}lder balls are computed in \cite{bolley2005quantitative} :
\begin{equation} \label{covering_number_holder}
\mathcal{N}(B(\alpha, R), \delta) \leq 10 \frac{R}{\delta} \exp \left[\log (3) 5^{\frac{1}{\alpha}} \left(\frac{R}{\delta}\right)^{\frac{1}{\alpha}} \right].
\end{equation}
We recover the (unpretty !) bound
\begin{align*}
\mathcal{C}_t & \leq 2(1 + 96 \frac{C_\alpha}{t} \sqrt{\log 96/(\sqrt{2 \log 2}t)}) \exp \left[ 240 \log 2 \frac{C_\alpha}{t} \sqrt{\log 96/(\sqrt{2 \log 2}t)} \right. \\
{} & \times \left. \exp \log 3 \left(120 \frac{C_\alpha}{t} \sqrt{\log 96/(\sqrt{2 \log 2}t)} \right)^{1/\alpha} \right].
\end{align*}
The final claim in the Proposition is obtained by elementary majorizations.
\end{proof}
\subsubsection{Paths of S.D.E.s}
H.Djellout, A.Guillin and L.Wu established a $\mathbf{T}_1$ inequality for paths of S.D.E.s that
allows us to work as in the case of Brownian motion. We quote their result from \cite{djellout_guillin_wu}.
Consider the S.D.E. on $\mathbb{R}^d$
\begin{equation} \label{s_d_e_example}
d X_t = b(X_t) dt + \sigma(X_t) d B_t, \quad X_0 = x_0 \in \mathbb{R}^d
\end{equation}
with $b : \mathbb{R}^d \rightarrow \mathbb{R}^d$, $\sigma : \mathbb{R}^d \rightarrow \mathcal{M}_{d \times m}$ and $(B_t)$ is a
standard $m$-dimensional Brownian motion. We assume that $b$ and $\sigma$ are locally Lipschitz and that
for all $x, y \in \mathbb{R}^d$,
\begin{equation*}
\sup_x | \sqrt{ \text{tr}\sigma(x)^t\sigma(x)} | \leq A, \quad \langle y - x, b(y) - b(x) \rangle \leq B(1 + |y-x|^2)
\end{equation*}
For each starting point $x$ it has a unique non-explosive solution denoted $(X_t(x)_{t \geq 0}$ and we denote its law on
$\mathcal{C}([0, 1], \mathbb{R}^d)$ by $\mathbb{P}_x$.
\begin{theorem} [\cite{djellout_guillin_wu}]
Assume the conditions above. There exists $C$ depending on $A$ and $B$ only such that
for every $x \in \mathbb{R}^d$, $\mathbb{P}_x$ satisfies a $\mathbf{T}_1(C)$ inequality on the space
$\mathcal{C}([0, 1], \mathbb{R}^d)$ endowed with the sup-norm.
\end{theorem}
We will now state our result. A word of caution : in order to balance readability, the following computations are neither
optimized nor made fully explicit. However it should be a simple, though dull, task for the reader to track the dependence of
the numerical constants on the parameters.
From now on we make the simplifying assumption that the drift coefficient is globally bounded by $B$
(this assumption is certainly not minimal).
\begin{example}
Let $\mu$ denote the law of the solution of the S.D.E. (\ref{s_d_e_example}) on
the Banach space $C([0, 1], \mathbb{R}^d)$ endowed with the sup-norm. Let $C$ be such that $\mu$ satisfies
$\mathbf{T}_1(C)$.
For all $0 < \alpha < 1/2$ there exist $C_\alpha$ and $c$ depending only on
$A$, $B$, $\alpha$ and $d$, and such that for $t \leq c$,
\begin{equation*}
\mathbb{P}(W_1(L_n, \mu) \geq t) \leq \mathcal{C}_t e^{- n t^2/ 8 C}
\end{equation*}
and
\begin{equation*}
\mathcal{C}_t \leq \exp \exp \left[ C_\alpha \left( \log \frac{1}{t} \right)^{-1 + 1/2\alpha} \left( \frac{1}{t} \right)^{-1 + 3/2\alpha} \right].
\end{equation*}
\end{example}
\begin{proof}
The proof goes along the same lines as the Brownian motion case, so we only outline the important steps.
First, there exists $a$ depending explicitely on $A$, $B$, $d$ such that $\mathbb{E}_{\mathcal{P}_x} e^{a\|X_.\|_\infty} \leq 2$ :
this can be seen by checking that the proof of Djellout-Guillin-Wu actually gives the value of a Gaussian moment for
$\mu$ as a function of $A$, $B$, $d$, and using standard bounds.
Corollary \ref{corollary_holder_sde} applies for $\alpha < 1/2$ and $p$ such that $1/p = 1/2 - \alpha$ :
there exists $C' < + \infty$ depending
explicitely on $A$, $B$, $\alpha$, $d$, such that $\mathbb{E} \|X_. \|_{\alpha}^p \leq C'$.
Consequently,
\begin{equation*}
\mu ( B(\alpha, R)^c) \leq C'/ R^p.
\end{equation*}
So choosing
\begin{equation*}
R = \left(C' (\frac{32}{at} \log \frac{32}{at} - \frac{32}{at} + 1) \right)^{1/p}
\end{equation*}
guarantees that
\begin{equation*}
\mu ( B(\alpha, R_t)^c) \leq \left( \frac{32}{at} \log \frac{32}{at} - \frac{32}{at} + 1 \right)^{-1}.
\end{equation*}
For $t \leq c$ small enough, $R \leq C'' \left( \frac{1}{t} \log \frac{1}{t} \right)^{1/p}$
with $c$, $C''$ depending on $A$, $B$, $\alpha$, $d$.
The conclusion is reached again by using estimate (\ref{covering_number_holder}) on the covering numbers of H\"{o}lder balls.
\end{proof}
\subsubsection{Gaussian r.v.s in Banach spaces} \label{subsection_gaussian}
In this paragraph we apply Theorem \ref{variant} to the case where $E$ is a separable Banach space with norm $\|.\|$,
and $\mu$ is a centered Gaussian random variable with values in $E$, meaning that the image of
$\mu$ by every continuous linear functional $f \in E^*$ is a centered Gaussian variable in $\mathbb{R}$.
The couple $(E, \mu)$ is said to be a Gaussian Banach space.
Let $X$ be a $E$-valued r.v. with law $\mu$, and define the weak variance of $\mu$ as
\begin{equation*}
\sigma = \sup_{f \in E^*, \, |f| \leq 1} \left( \mathbb{E} f^2(X) \right)^{1/2}.
\end{equation*}
The small ball function of a Gaussian Banach space $(E, \mu)$ is the function
\begin{equation*}
\psi(t) = - \log \mu(B(0, t)).
\end{equation*}
We can associate to the couple $(E, \mu)$ their Cameron-Martin Hilbert space $H \subset E$, see e.g. \cite{ledoux1996isoperimetry}
for a reference. It is known that the small ball function has deep links with the covering numbers of the unit ball of $H$,
see e.g. Kuelbs-Li \cite{kuelbs1993metric} and Li-Linde \cite{li1999approximation}, as well as with the
approximation of $\mu$ by measures with finite support in Wasserstein distance (the quantization or optimal quantization problem),
see Fehringer's Ph.D. thesis \cite{fehringer2001kodierung}, Dereich-Fehringer-Matoussi-Scheutzow \cite{dereich2003link}, Graf-Luschgy-Pag\`{e}s
\cite{graf2003functional}. It should thus come as no surprise that we can give a bound on the constant $K_t$ depending solely on $\psi$
and $\sigma$.
This is the content of the next example.
\begin{example} \label{theorem_gaussian_vector}
Let $(E, \mu)$ be a Gaussian Banach space. Denote by $\psi$ its small ball function and by $\sigma$ its weak variance.
Assume that $t$ is such that $\psi(t/16) \geq \log 2$ and $t/\sigma \leq 8 \sqrt{2 \log 2}$.
Then
\begin{equation*}
\mathbb{P} (W_1(L_n, \mu) \geq t) \leq K_t e^{- n t^2/16 \sigma^2}
\end{equation*}
with
\begin{equation*}
K_t = \exp \exp \left[ c (\psi(t/32) + \log(\sigma/t)) \right]
\end{equation*}
for some universal constant $c$.
\end{example}
A bound for $c$ may be tracked in the proof.
\begin{proof}
\emph{Step 1. Building an approximating measure of finite support.}
Denote by $K$ the unit ball of the Cameron-Martin space associated to $E$ and $\mu$, and by
$B$ the unit ball of $E$.
According to the Gaussian isoperimetric inequality (see \cite{ledoux1996isoperimetry}),
for all $\lambda > 0$ and $\varepsilon > 0$,
\begin{equation*}
\mu( \lambda K + \varepsilon B) \geq \Phi \left( \lambda + \Phi^{-1}( \mu( \varepsilon B)) \right)
\end{equation*}
where $\Phi(t) = \int_{- \infty}^t e^{- u^2/2} d u / \sqrt{2 \pi}$ is the Gaussian c.d.f.. Note
\begin{equation*}
\mu' = \frac{1}{ \mu( \lambda K + \varepsilon B)} \mathbf{1}_{ \lambda K + \varepsilon B} \mu
\end{equation*}
the restriction of $\mu$ to the enlarged ball. As proved in \cite{boissard2011bounding}, Appendix 1,
the Gaussian measure $\mu$ satisfies a $\mathbf{T}_2(2 \sigma^2)$ inequality,
hence a $\mathbf{T}_1$ inequality with the same constant. We have
\begin{align*}
W_1(\mu, \mu') & \leq \sqrt{ 2 \sigma^2 H( \mu' | \mu)} = \sqrt{-2 \sigma^2 \log \mu( \lambda K + \varepsilon B)} \\
{} & \leq \sqrt{-2 \sigma^2 \log \Phi( \lambda + \Phi^{-1}(\mu(\varepsilon B)))}.
\end{align*}
On the other hand,
denote $k = \mathcal{N}(\lambda K, \varepsilon)$ the covering number of $\lambda K$
(w.r.t. the norm of $E$).
Let $x_1, \ldots, x_k \in K$ be such that union of the balls
$B(x_i, \varepsilon)$ contains $\lambda K$.
From the triangle inequality we get the inclusion
\begin{equation*}
\lambda K + \varepsilon B \subset \bigcup_{i = 1}^k B(x_i, 2 \varepsilon).
\end{equation*}
Choose a measurable map $T : \lambda K + \varepsilon B \rightarrow \{ x_1, \ldots, x_k \}$
such that for all $x$, $|x - T(x)| \leq 2 \varepsilon$.
The push-forward measure $\mu^k = T_\# \mu'$ has support in the finite set
$\{ x_1, \ldots, x_k \}$, and clearly
\begin{equation*}
W_1( \mu', \mu^k) \leq 2 \varepsilon.
\end{equation*}
Choose $\varepsilon = t/16$, and
\begin{align}
\lambda & = \Phi^{-1}( e^{-t^2/(128 \sigma^2)}) - \Phi^{-1}( \mu( \varepsilon B)) \\
{} \label{bound_lambda} & = \Upsilon^{-1}( e^{ - \psi( t/16)} ) + \Phi^{-1} ( e^{-t^2/(128 \sigma^2)})
\end{align}
where $\Upsilon (t) = \int_t^{+ \infty} e^{- u^2/2} d u / \sqrt{2 \pi}$
is the tail of the Gaussian distribution (we have used the fact that $\Phi^{-1} + \Upsilon^{-1} = 0$, which comes from symmetry of the Gaussian distribution).
Altogether, this ensures that $W_1(\mu, \mu^k) \leq t/4$.
\emph{Step 2. Bounding $\lambda$.}
We can use the elementary bound
$\Upsilon(t) \leq e^{-t^2/2}$, $t \geq 0$ to get
\begin{equation*}
\Upsilon^{-1}(u) \leq \sqrt{-2 \log u}, \quad 0 < u \leq 1/2
\end{equation*}
which yields $\Upsilon^{-1} (e^{ - \psi( t/16)}) \leq \sqrt{ \psi(t/16)}$
as soon as $\psi(t/16) \geq \log 2$.
Likewise,
\begin{align*}
\Phi^{-1}(e^{-t^2/128 \sigma^2}) & = \Upsilon^{-1}(1 - e^{-t^2/128 \sigma^2}) \\
{} & \leq \sqrt{2 \log \frac{1}{1 - e^{- t^2/128 \sigma^2}}}
\end{align*}
as soon as $t^2/128 \sigma^2 \leq \log 2$. Moreover, for $u \leq \log 2$, we have
$1/(1 - e^{-u}) \leq 2 \log 2 / u$. Putting everything together, we get
\begin{equation} \label{bound_lambda_explicit}
\lambda \leq \sqrt{ \psi(t/16)} + c\sqrt{\log \sigma/t}
\end{equation}
for some universal constant $c > 0$.
Observe that the first term in (\ref{bound_lambda_explicit}) will usually be much larger than the second one.
\emph{Step 3.}
From Theorem \ref{variant} we know that
\begin{equation*}
\mathbb{P}( W_2( \mu, L_n) \geq t) \leq K_t e^{-nt^2/16\sigma^2}
\end{equation*}
with
\begin{equation*}
K_t = \exp \left[ \frac{1}{2 \sigma^2} \frac{k}{2} (\text{Diam } \{x_1, \ldots, x_k\})^2 \right].
\end{equation*}
The diameter is bounded by $\text{Diam }K = 2 \sigma \lambda \leq c \sigma ( \sqrt{ \psi(t/16)} + c\sqrt{\log \sigma/t})$.
We wish now to control $k = \mathcal{N}(\lambda K, t/16)$ in terms of the small ball function $\psi$.
The two quantities are known to be connected : for example, Lemma 1 in \cite{kuelbs1993metric} gives the bound
\begin{equation*}
\mathcal{N}( \lambda K, \varepsilon) \leq e^{\lambda^2/2 + \psi(\varepsilon/2)}.
\end{equation*}
Thus
\begin{equation*}
k \leq \exp \left[ \psi(t/16) + \psi(t/32) + c \log \sigma/t \right ].
\end{equation*}
With some elementary majorizations, this ends the proof.
\end{proof}
We can now sharpen the results of Proposition \ref{example_s_b_m}.
Let $\gamma$ denote the Wiener measure on $\mathcal{C}([0, 1], \mathbb{R}^d)$ endowed with the sup-norm, and denote by $\sigma^2$ its weak variance.
Let $\lambda_1$ be the first nonzero eigenvalue
of the Laplacian operator on the ball of $\mathbb{R}^d$ with homogeneous Dirichlet boundary conditions :
it is well-known that the small ball function for the Brownian motion on $\mathbb{R}^d$ is equivalent to $\lambda_1/t^2$
when $t \rightarrow + \infty$. for $t$ small enough.
As a consequence, there exists $C = C(d)$ such that for small enough $t > 0$ we have
\begin{equation}
W_1( L_n, \gamma) \leq \exp \exp \left[ C \lambda_1 / t^2 \right] e^{- n t^2/16 \sigma^2}.
\end{equation}
\subsection{Bounds in the dependent case : occupation measures of contractive Markov chains}
The results above can be extended to the convergence of the occupation measure for a
Markov process. As an example, we establish the following result.
\begin{theorem} \label{theorem_markov}
Let $P(x, dy)$ be a Markov kernel on $\mathbb{R}^d$ such that
\begin{enumerate}
\item the measures $P(x, .)$ satisfy a $\mathbf{T}_1(C)$ inequality
\item $W_1(P(x, .), P(y, .)) \leq r |x - y|$ for some $r < 1$.
\end{enumerate}
Let $\pi$ denote its invariant measure.
Let $(X_i)_{i \geq 0}$ denote the Markov chain associated with $P$ under $X_0 = 0$.
Set $a = \frac{2}{C} \left( \sqrt{4 m_1^2 + C \log 2} - 2 m_1 \right)$. There exists
$C_d > 0$ depending only on $d$ such that for $t \leq 2/a$,
\begin{equation*}
\mathbb{P}( W_1( L_n, \pi) \geq t) \leq K(n, t) \exp - n \frac{(1 - r)^2}{8C} t^2
\end{equation*}
where
\begin{equation*}
K(n, t) = \exp \left[ \frac {m_1}{\sqrt{n C}} + C_d (\frac{1}{at} \log \frac{1}{at})^\frac{d}{2} \right]^2.
\end{equation*}
\end{theorem}
\begin{remark}
The result is close to the one obtained in the independent case, and, as stressed in the introduction, it holds interest from the perspective of
numerical simulation, in cases where one cannot sample uniformly from a given probability distribution $\pi$ but may build a Markov chain
that admits $\pi$ as its invariant measure.
\end{remark}
\begin{remark}
We comment on the assumptions on the transition kernel. The first one ensures that the $\mathbf{T}_1$ inequality is propagated to the laws of $X_n$, $n \geq 1$.
As for the second one, it has appeared several times in the Markov chain literature
(see e.g. \cite{djellout_guillin_wu}, \cite{ollivier2009ricci}, \cite{joulin2010curvature})
as a particular variant of the Dobrushin uniqueness condition for Gibbs measures. It has a nice geometric interpretation as a positive lower bound on
the Ricci curvature of the Markov chain, put forward for example in \cite{ollivier2009ricci}. Heuristically,
this condition implies that the Markov chains started from two different points and suitably coupled tend to get closer.
\end{remark}
\section{Proof of Theorem \ref{main_thm}} \label{proof_of_main_thm}
The starting point is the following result, obtained by Gozlan and Leonard (\cite{large_dev_gozlan_leonard}, see Chapter 6)
by studying the tensorization properties of transportation inequalities.
\begin{lemma} \label{tensorization_lemma}
Suppose that $\mu \in \mathcal{P}(E)$ verifies a $\alpha(\mathcal{T}_d)$ inequality. Define on $E^n$ the metric
\begin{equation*}
d^{\oplus n}( (x_1, \ldots, x_n), (y_1, \ldots, y_n)) = \sum_{i = 1}^n d(x_i, y_i).
\end{equation*}
Then $\mu^{ \otimes n} \in \mathcal{P}(E^n)$ verifies a $\alpha'( \mathcal{T}_{d^{\oplus n}})$ inequality, where $\alpha'(t) = \frac{1}{n} \alpha( n t)$.
Hence, for all Lipschitz functionals
$Z : E^n \rightarrow \mathbb{R}$ (w.r.t. the distance $d^{\oplus n}$), we have the
concentration inequality
\begin{equation*}
\mu^{\otimes n} ( Z \geq \int Z d \mu^{\otimes n} + t) \leq \exp - n \alpha( \frac{t}{ n \|Z \|_\text{Lip}}) \quad \text{for all } t \geq 0.
\end{equation*}
\end{lemma}
Let $X_i$ be an i.i.d. sample of $\mu$.
Recalling that
\begin{equation*}
W_1(L_n, \mu) = \sup_{f 1-\text{Lip}} \frac{1}{n} \sum_{i = 1}^n f(X_i) - \int f d \mu
\end{equation*}
and that
\begin{equation*}
(x_1, \ldots, x_n) \mapsto \sup _{f 1-\text{Lip}} \frac{1}{n} \sum_{i = 1}^n f(x_i) - \int f d \mu
\end{equation*}
is $\frac{1}{n}$-Lipschitz w.r.t. the distance $d^{\oplus n}$ on $E^n$ (as a supremum of $\frac{1}{n}$-Lipschitz functions),
the following ensues :
\begin{equation} \label{deviation_estimate_with_mean}
\mathbb{P}( W_1(L_n, \mu) \geq \mathbb{E}[W_1(L_n, \mu)] + t) \leq \exp - n \alpha(t).
\end{equation}
Therefore, we are led to seek a control on $\mathbb{E}[W_1(L_n, \mu)]$. This is what we do in the next lemma.
\begin{lemma} \label{lemma_mean}
Let $a > 0$ be such that
$E_{a, 1} = \int e^{a d(x, x_0)} \mu(d x) \leq 2$.
Let $\delta > 0$ and $K \in E$ be a compact subset containing $x_0$.
Let $\mathcal{N}_\delta$ denote the covering number of order
$\delta$ for the set $\mathcal{F}_K$ of $1$-Lipschitz functions on $K$
vanishing at $x_0$ (endowed with the uniform distance).
Also define $\sigma : [0, +\infty ) \rightarrow [1, + \infty)$ as the inverse function of $x \mapsto x \ln x - x + 1$ on $[1, + \infty)$.
The following holds :
\begin{equation*}
\mathbb{E}[ W_1(L_n, \mu)] \leq 2 \delta + 8 \frac{1}{a} \frac{1}{\sigma (\frac{1}{\mu(K^c)})} + \Gamma(\mathcal{N}_\delta, n)
\end{equation*}
where
\begin{equation*}
\Gamma( \mathcal{N}_\delta, n) = \inf_{\lambda > 0} \frac{1}{\lambda} [ \log \mathcal{N}_\delta + n \alpha^*(\frac{\lambda}{n}) ].
\end{equation*}
\end{lemma}
\begin{proof}
We denote by $\mathcal{F}$
the set of $1$-Lipschitz functions $f$ over $E$ such that $f(x_0) = 0$.
Let us denote
\begin{equation*}
\Psi(f) = \int f d \mu - \int f d L_n,
\end{equation*}
we have for $f, g \in \mathcal{F}$ :
\begin{eqnarray*}
| \Psi(f) - \Psi(g)| & \leq & \int |f - g| 1_K d \mu + \int |f - g| 1_K d L_n \\
{} & {} & + \int (|f| + |g|)1_{K^c} d \mu + \int (|f| + |g|) 1_{K^c} d L_n \\
{} & \leq & 2 \| f - g \|_{L^\infty(K)} + 2 \int d(x, x_0) 1_{K^c} d \mu + 2 \int d(x, x_0) 1_{K^c} d L_n
\end{eqnarray*}
When $f : E \rightarrow \mathbb{R}$ is a measurable function,
denote by $f|_K$ its restriction to $K$. Notice that for every $g \in \mathcal{F}_K$, there
exists $f \in \mathcal{F}$ such that $f|_K = g$. Indeed, one may set
\begin{equation*}
f(x) =
\begin{cases}
g(x) \text{ if } x \in K \\
\inf_{y \in K} f(y) + d(x, y) \text{ otherwise}
\end{cases}
\end{equation*}
and check that $f$ is $1$-Lipschitz over $E$.
By definition of $\mathcal{N}_\delta$, there exist
functions $g_1, \ldots, g_{\mathcal{N}_\delta} \in \mathcal{F}_K$
such that the balls of center $g_i$ and radius $\delta$
(for the uniform distance) cover $\mathcal{F}_K$. We can extend
these functions to functions $f_i \in \mathcal{F}$ as noted above.
Consider $f \in \mathcal{F}$ and choose $f_i$ such that $|f - f_i| \leq \delta$ on $K$ :
\begin{eqnarray*}
\Psi(f) & \leq & | \Psi(f) - \Psi(f_i)| + \Psi(f_i) \\
{} & \leq & \Psi(f_i) + 2 \delta + 2 \int d(x, x_0) 1_{K^c} d \mu + 2 \int d(x, x_0) 1_{K^c} d L_n \\
{} & \leq & \max_{j = 1, \ldots, \mathcal{N}_\delta} \Psi(f_j) + 2 \delta + 2 \int d(x, x_0) 1_{K^c} d \mu + 2 \int d(x, x_0) 1_{K^c} d L_n
\end{eqnarray*}
The right-hand side in the last line does not depend on $f$, so it is also greater than $W_1( L_n, \mu) = \sup_\mathcal{F} \Psi(f)$.
We pass to expectations, and bound the terms on the right.
We use Orlicz-H\"{o}lder's inequality with the pair of conjugate Young functions
\begin{align*}
\tau(x) & = \begin{cases}
0 \text{ if } x \leq 1 \\
x \log x - x + 1 \text{ otherwise}
\end{cases} \\
\tau^*(x) & = e^x - 1
\end{align*}
(for definitions and a proof of Orlicz-H\"{o}lder's inequality, the reader may refer to \cite{rao1991theory}, Chapter 10).
We get
\begin{equation*}
\int d(x, x_0) 1_{K^c} d \mu \leq 2 \| 1_{K^c} \|_\tau \| d(x, x_0) \|_{\tau^*}
\end{equation*}
where
\begin{equation*}
\| 1_{K^c} \|_\tau = \inf \{ \theta > 0, \quad \int \tau \left( \frac{1_{K^c}}{\theta} \right) d \mu \leq 1 \}
\end{equation*}
and
\begin{equation*}
\| d(x, x_0) \|_{\tau^*} = \inf \{ \theta > 0, \quad \int \left[ e^{\frac{d(x, x_0)}{\theta}} - 1 \right] d \mu \leq 1 \}.
\end{equation*}
It is easily seen that $ \| 1_{K^c} \|_\tau = 1/\sigma (1/\mu(K^c))$.
And we assumed that $a$ is such that $E_{a, 1} = \int \exp a d(x, x_0) d \mu \leq 2$, so
$\| d(x, x_0) \|_{\tau^*} \leq 1/a$.
Altogether, this yields
\begin{equation*}
\int d(x, x_0) 1_{K^c} d \mu \leq 2 \frac{1}{a}\frac{1}{\sigma (\frac{1}{\mu(K^c)})}.
\end{equation*}
Also, if $X_1, \ldots, X_n$ are i.i.d. variables of law $\mu$,
\begin{equation*}
\mathbb{E} [ \int d(x, x_0) 1_{K^c} d L_n ] = \mathbb{E} [ d(X_1, x_0) 1_{K^c}(X_1) ] \leq \frac{2}{a} \frac{1}{\sigma (1/\mu(K^c))}
\end{equation*}
as seen above. Putting this together yields the inequality
\begin{equation*}
\mathbb{E} [W_1( L_n, \mu)] \leq 2 \delta + \frac{8}{a} \frac{1}{\sigma (1/\mu(K^c))} + \mathbb{E} [ \max_{j = 1, \ldots, \mathcal{N}_\delta} \Psi(f_j) ].
\end{equation*}
The remaining term can be bounded by a form of maximal inequality.
First fix some $i$ and $\lambda > 0$ : we have
\begin{eqnarray*}
\mathbb{E}[ \exp \lambda \Psi(f_i) ] & = & \mathbb{E} [ \exp \frac{\lambda}{n} \sum_{j = 1}^n (f(X_j) - \int f d \mu) ] \\
{} & = & ( \mathbb{E} [ \exp \frac{\lambda}{n} (f(X_1) - \int f d \mu) ] )^n \\
{} & \leq & e^ {n \alpha^\circledast(\lambda/n)}.
\end{eqnarray*}
In the last line, we have used estimate (\ref{cond_laplace}).
Using Jensen's inequality, we may then write
\begin{eqnarray*}
\mathbb{E} [ \max_{j = 1, \ldots, \mathcal{N}_\delta} \Psi(f_j) ] & \leq & \frac{1}{\lambda} \log \mathbb{E} [ \max_{j = 1, \ldots, \mathcal{N}_\delta} \exp \lambda \Psi(f_j) ] \\
{} & \leq & \frac{1}{\lambda} \log \sum_{j = 1}^{\mathcal{N}_\delta} \mathbb{E} [ \exp \lambda \Psi(f_j)] \\
{} & \leq & \frac{1}{\lambda} [ \log \mathcal{N}_\delta + n \alpha^*(\frac{\lambda}{n}) ]
\end{eqnarray*}
So minimizing in $\lambda$ we have
\begin{equation*}
\mathbb{E} [ \max_{j = 1, \ldots, \mathcal{N}_\delta} \Psi(f_j) ] \leq \Gamma( \mathcal{N}_\delta, n).
\end{equation*}
Bringing it all together finishes the proof of the lemma.
\end{proof}
We can now finish the proof of Theorem \ref{main_thm}.
\begin{proof}
Come back to the deviation bound (\ref{deviation_estimate_with_mean}).
Choose $\delta = t/8$, and choose $K$ such that
\begin{equation*}
\mu(K^c) \leq \left[ \frac{32}{a t} \log \frac{32}{a t} - \frac{32}{a t} + 1 \right]^{-1}.
\end{equation*}
We thus have
$2 \delta + 8 [a \sigma (1/\mu(K^c)) ]^{-1} \leq t/2$, which implies
\begin{equation} \label{general_estimate_mean}
\mathbb{E}( W_1(L_n,\mu) \leq t/2 + \Gamma(\mathcal{C}_t, n)
\end{equation}
and so
\begin{equation*}
\mathbb{P}( W_1(L_n, \mu) \geq t) \leq \exp -n \alpha(\frac{t}{2} -\Gamma(\mathcal{N}_\delta, n)),
\end{equation*}
with the convention $\alpha(y) = 0$ if $y < 0$.
\end{proof}
\section{Proof of Theorem \ref{variant}} \label{section_variant}
In this section, we provide a different approach to our result in the independent case.
As earlier we first aim to get a bound
on the speed of convergence on the average $W_1$ distance between empirical and true measure.
The lemma below provides another way to obtain such an estimate.
\begin{lemma} \label{lemma_alternative}
Let $\mu^k \in \mathcal{P}(E)$ be a finitely supported measure such that $|\text{Supp }\mu^k| \leq k$.
Let $D( \mu^k) = \text{Diam Supp } \mu^k$ be the diameter of $\text{Supp } \mu^k$.
The following holds :
\begin{equation*}
\mathbb{E} W_1(\mu, L_n) \leq 2 W_1 (\mu, \mu^k) + D(\mu^k) \sqrt{k / n}.
\end{equation*}
\end{lemma}
\begin{proof}
Let $\pi_\text{opt}$ be an optimal
coupling of $\mu$ and $\mu^k$ (it exists : see e.g. Theorem 4.1 in \cite{optimal_transport_villani}), and let
$(X_i, Y_i)$, $1 \leq i \leq n$, be i.i.d. variables on $E \times E$ with common law $\pi_\text{opt}$.
Let $L_n = 1/n \sum_{i = 1}^n \delta_{X_i}$ and $L_n^k = 1/n \sum_{i = 1}^n \delta_{Y_i}$.
By the triangle inequality, we have
\begin{equation*}
W_1(L_n, \mu) \leq W_1(L_n, L_n^k) + W_1(\mu, \mu^k) + W_1(\mu^k, L_n^k).
\end{equation*}
With our choice of coupling for $L_n$ and $L_n^k$ it is easily seen that
\begin{equation*}
\mathbb{E} W_1 (L_n, L_n^k) \leq W_1 (\mu, \mu^k)
\end{equation*}
Let us take care of the last term. We use Lemma \ref{lemma_measures_finite_support} below to obtain that
\begin{align*}
\mathbb{E} W_1 (L_n^k, \mu^k) & \leq D(\mu^k) \mathbb{E} \left( 1 - \sum_{i = 1}^k \mu^k(x_i) \wedge L_n^k(x_i) \right)\\
{} & = D(\mu^k) \sum_{i = 1}^k \mathbb{E} ( \mu^k(x_i) - \mu^k(x_i) \wedge L_n^k(x_i) ) \\
{} & \leq D(\mu^k) \sum_{i = 1}^k \mathbb{E} | \mu^k(x_i) - L_n^k(x_i)| \\
{} & \leq \frac{D(\mu^k)}{n} \sum_{i = 1}^k \sqrt{ \mathbb{E} | n\mu^k(x_i) - n L_n^k(x_i)|^2 }.
\end{align*}
Observe that the variables $n L_n^k(x_i)$ follow binomial laws with parameter $\mu^k(x_i)$ and $n$. We get :
\begin{equation*}
\mathbb{E} W_1(\mu^k, L_n^k) \leq \frac{D(\mu^k)}{n} \sum_{i = 1}^k \sqrt{n \mu^k(x_i) (1 - \mu^k(x_i))} \leq D(\mu^k) \sqrt{k/n}
\end{equation*}
(the last inequality being a consequence of the Cauchy-Schwarz inequality).
\end{proof}
\begin{lemma} \label{lemma_measures_finite_support}
Let $\mu, \nu$ be probability measures with support in a finite metric space $\{x_1, \ldots, x_k\}$
of diameter bounded by $D$. Then
\begin{equation*}
W_1(\mu, \nu) \leq D \left(1 - \sum_{i = 1}^k \left( \mu(x_i) \wedge \nu(x_i) \right)\right).
\end{equation*}
\end{lemma}
\begin{proof}
We build a coupling of $\mu$ and $\nu$ that leaves as much mass in place as possible,
in the following fashion :
set $f(x_i) = \mu(x_i) \wedge \nu(x_i)$ and $\lambda = \sum_{i _ 1}^k f_i$.
Set $q(x_i) = f_i/ \lambda$, and define the measures
\begin{align*}
\mu_1 & = \frac{1}{1- \lambda} (\mu - f) \\
\nu_1 & = \frac{1}{1- \lambda} (\nu - f).
\end{align*}
Finally, build independent random variables
$X_1 \sim \mu_1$, $Y_1 \sim \nu_1$, $Z \sim q$ and $B$ with Bernoulli law of parameter $\lambda$.
Define
\begin{equation*}
X = (1-B)X_1 + BZ, \quad Y = (1-B)Y_1 + BZ.
\end{equation*}
It is an easy check that $X \sim \mu$, $Y \sim \nu$.
Thus we have the bound
\begin{align*}
W_1 (\mu, \nu) & \leq \mathbb{E} |X - Y| \\
{} & = (1- \lambda) \mathbb{E}|X_1 - Y_1| \leq D (1 - \lambda)
\end{align*}
and this concludes the proof.
\end{proof}
\begin{proof}[Proof of Theorem \ref{variant}]
As stated earlier, we have the concentration bound
\begin{equation*}
\mathbb{P}( W_p(L_n, \mu) \geq t + \mathbb{E} W_p(L_n, \mu) ) \leq e^{-nt^2 / C}.
\end{equation*}
The proof is concluded by arguments similar to the ones used before, calling upon Lemma \ref{lemma_alternative} to bound the mean.
\end{proof}
|
\section{introduction}
Fluctuations witness the
interplay between quantum statistics and interactions and therefore their measurement
constitutes an important probe of
quantum many-body systems.
In particular, measurement of atom number fluctuations in ultracold
quantum gases has been a key tool
in the study of
the
Mott insulating phase in optical lattices \cite{Folling2005}, isothermal
compressibility of Bose and Fermi
gases \cite{Esteve06,Armijo10_2,Sanner10,Mueller10},
magnetic susceptibility of a strongly interacting
Fermi gas \cite{Sanner2011}, scale invariance of
a two-dimensional Bose gas \cite{Hung2011}, generation of atomic entanglement in
double-wells \cite{Esteve2008}, and relative number squeezing in pair-production via binary collisions
\cite{Jaskula2010,TwinAtomBeams}.
While
a simple account of quantum statistics can change the atom number
distribution, in a small volume of an ideal gas, from a classical-gas
Poissonian to super-Poissonian (for bosons) or sub-Poissonian (for
fermions) distributions,
many-body processes
can further modify the correlations and fluctuations. For
example, three-body losses may lead to sub-Poissonian fluctuations in
a Bose gas~\cite{Whitlock2010,Itah2010}. Even without dissipation, the
intrinsic interatomic interactions can also lead to sub-Poissonian
fluctuations, such as in a repulsive Bose gas in a periodic lattice
potential, where the energetically costly atom number fluctuations are
suppressed.
This effect has been
observed
for large ratios of
the on-site interaction energy to the inter-site tunnelling energy
\cite{Wei2007,Gross2010}, with the extreme limit
corresponding to the Mott insulator phase
~\cite{Bakr2010,Sherson2010}.
The same physics,
accounts for sub-Poissonian fluctuations
observed in double-well experiments~\cite{Sebby-Strabley2007,Esteve2008}.
Sub-Poissonian fluctuations of the total atom number have been also realised
via controlled loading of the atoms into very shallow traps
~\cite{Chuu2005}.
\begin{figure}
\includegraphics[width=7.9cm]{Fig_1.eps}
\caption{Phase diagram in the interaction-temperature
parameter space of a repulsive
uniform 1D Bose gas
\cite{Kheruntsyan05}.
The values of the local two-body correlation $g^{(2)}(0)$ are indicated
for the three main
regimes (white and grey areas).
The two horizontal lines show the parameters explored in this paper.
}
\label{fig.figphasediag}
\end{figure}
In this work, we observe for the first time sub-Poissonian atom
number fluctuations in
small slices of a \textit{single}
one-dimensional (1D) Bose gas with repulsive interactions,
where each slice approximates a uniform system.
Taking advantage of the long scale density variation due
to a weak longitudinal confinement, we
monitor -- at a given temperature -- the atom number fluctuations
in each slice
as a function of
the local density.
For a weakly interacting gas, the
measured fluctuations are super-Poissonian at low densities, and they
become sub-Poissonian
as the density is increased and the gas enters
the quantum quasi-condensate sub-regime that is
dominated by \textit{quantum} rather than thermal fluctuations
(see Fig.~\ref{fig.figphasediag}, with the interaction and temperature parameters, $\gamma$ and $t$, defined below).
When the strength of interactions is increased, the fluctuations are no longer
super-Poissonian at low densities and remain
sub-Poissonian at high densities.
The absence of super-Poissonian behavior implies
that the gas enters the strongly
interacting regime where the repulsive interactions
between bosonic atoms mimic fermionic Pauli blocking,
and the quantities involving only densities
are those of an ideal Fermi gas.
Our results in all regimes are in good agreement with the exact
Yang-Yang thermodynamic solution for the uniform 1D Bose gas
with contact interactions~\cite{YangYang69}.
We recall that the thermodynamics of a uniform 1D Bose gas can be
characterized by the dimensionless interaction and temperature parameters,
$\gamma\!=\!mg/\hbar^2 n$ and $t\!=\!2\hbar^2k_BT/mg^2$~\cite{Kheruntsyan05},
where $T$ is the temperature,
$n$ the 1D density, $g\simeq 2\hbar \omega_{\perp }a$ is the coupling constant, $a$ is the $s$-wave scattering length,
and $\omega_{\perp}$ is the frequency of the transverse harmonic
confining potential.
Figure~\ref{fig.figphasediag}
shows the different regimes of the gas,
characterized by the behavior of the
two-body correlation function $g^{(2)}$ and
separated by
smooth crossovers~\cite{Kheruntsyan05}.
Of particular relevance to the present work are the quantum quasi-condensate
sub-regime where $g^{(2)}(0)\lesssim 1$~\cite{Kheruntsyan05},
and the strongly interacting regime where
$g^{(2)}(0)\ll 1$~\cite{Kheruntsyan05,Weiss2005}.
The two different situations studied in this work
are shown in Fig.~\ref{fig.figphasediag}
by two horizontal lines at different values of $t$.
The experiments are performed using $^{87}$Rb atoms ($a\!=\!5.3$~nm)
confined in a magnetic trap
realised by current-carrying microwires on an atom chip.
For the data at $t=65$, as in Ref.~\cite{Armijo10_2}, we use
an H-shaped structure to
realise a very elongated harmonic trap at $\sim\!\!100~\mu$m
away from the wires, with
the longitudinal frequency $\omega_{\parallel}/2\pi\!=\!5.5$~Hz and
$\omega_{\perp}/2\pi\!=\!3.3$~kHz.
Using rf evaporation we produce clouds of $\sim\!\!3000$ atoms in
thermal equilibrium at $T\!=\!16.5~$nK, corresponding to $t\!=\!65$.
We extract the longitudinal density profile from \textit{in situ} absorption
images as detailed in~\cite{ArmijoSkew}. The local atom number
fluctuations in the image pixels, whose
length in the object plane is $\Delta=4.5~\mu$m, are
measured by repeating the same experiment
hundreds of times and performing statistical
analysis of the density profiles \cite{Armijo10_2}.
For each profile and pixel, we record the
atom number fluctuation $\delta N=N-\langle N\rangle$,
where $\langle N\rangle=n\Delta$ is the mean atom number.
The results are binned
according to $\langle N\rangle$ and for each
bin we compute the variance $\langle \delta N^2\rangle$.
The contribution of optical shot noise to $\langle \delta N^2\rangle$
is subtracted.
Figure~\ref{fig.varNchaud} shows the measured variance $\langle \delta N^2\rangle$ versus $\langle N\rangle$.
Since $l_c\ll\!\max\{\Delta,d\}\!\ll \!L$ in our experiment, where $L\!\simeq \!50$~$\mu$m is the cloud
rms length, $d$ is the imaging resolution,
and $l_c$ is the correlation length of density
fluctuations \cite{Kheruntsyan05,Deuar2009}, the local density
approximation is expected to correctly describe both the average density profile and the
fluctuations~\cite{Kheruntsyan05,Armijo10_2}.
Accordingly, $\langle \delta N^2\rangle$ is expected to follow the
thermodynamic prediction~\cite{Armijo10_2,ArmijoSkew}
\begin{equation}
\langle \delta N^2\rangle =\kappa k_B T\Delta (\partial n/\partial \mu)_{T},
\label{eq.thermo}
\end{equation}
where $n(\mu,T)$ is the linear density of a homogeneous gas,
and $\mu$ is the chemical potential.
The reduction factor $\kappa$ accounts for the finite resolution of the
imaging system; it is determined from the measured
correlation between adjacent pixels~\cite{ArmijoSkew}, from which we deduce the rms
width of the imaging impulse response function $\cal A$ (assumed
to be Gaussian), and find that $d=3.5$~$\mu$m and $\kappa=0.34$.
\begin{figure}
\includegraphics[width=6.5cm]{Fig_2.eps}
\caption{Variance of the atom number fluctuations in a weakly interacting gas, for $t=65$.
The measured data are shown as circles together with statistical uncertainties.
Predictions from Eq.~(\ref{eq.thermo}) and different thermodynamic models are shown
as solid
(Yang-Yang), short-dashed
(ideal Bose gas)
and dash-dotted (quasi-condensate) lines. The long-dashed line
is the Poissonian limit and the low lying dotted line is the
contribution from quantum fluctuations.
Here $\omega_\perp/2\pi\!=\!3.3$~kHz, $\omega_\parallel/2\pi\!=\!5.5$~Hz,
$T\!=\!16$~nK~($k_BT=\!0.1\hbar\omega_\perp$), and
$\kappa=0.34$.
}
\label{fig.varNchaud}
\end{figure}
The thermodynamic predictions for an ideal Bose gas and a quasi-condensate
are shown in Fig.~\ref{fig.varNchaud}. In the quasi-condensate regime
we use the equation of state $\mu=\hbar\omega_\perp(\sqrt{1+4na}-1)$~\cite{Fuchs03}.
The temperature is obtained by fitting
the quasi-condensate prediction to the measured fluctuations
at high densities.
Usual features of the quasi-condensation
transition are seen~\cite{Armijo10_2}: at low density, the gas lies within
the ideal gas regime where, for degenerate gases,
bosonic bunching raises the fluctuations well above the Poissonian limit;
at high density the gas lies in the quasi-condensate regime
where interactions level off the density fluctuations.
Within the quasi-condensate regime, the fluctuations go from super-Poissonian
to sub-Poissonian, with $\langle \delta N^2\rangle/\kappa\langle N\rangle$
going from $2$ to $0.44$.
Using the approximate 1D expression $\mu\!=\!gn$,~Eq.~(\ref{eq.thermo}) shows that
the transition
from super- to sub-Poissonian behavior
occurs at $k_BT\!\simeq\!gn$, which is the boundary between the thermal
and quantum quasi-condensate regimes~\cite{Kheruntsyan05,Deuar2009}.
The fluctuations in the whole explored density domain are in good agreement
with the exact 1D Yang-Yang predictions. The small
discrepancy at ~high densities~between the Yang-Yang and the quasi-condensate
models is due to the transverse swelling of the cloud
~\cite{ArmijoSkew,Armijo10_2}.
In the following, we neglect this 3D effect
and perform a purely 1D analysis.
Going beyond the thermodynamic relation~(\ref{eq.thermo}),
the variance $\langle \delta N^2\rangle$ in a pixel is given by
\begin{equation}
\langle \delta N^2\rangle =\!\!\int \!\! d^{4}{\cal {Z}}~\langle \delta n(z)\delta n(z')\rangle \;{\cal A}(z-Z){\cal A}(z'-Z'),
\label{eq.fluctu}
\end{equation}
where $\int \!d^4{\cal {Z}}\equiv\!\!\int_0^\Delta \!\!dZ\!\!\int_0^\Delta \!\!dZ'\!\!\int_{-\infty}^{\infty}\!\!dz\!\!\int_{-\infty}^{\infty}\!\!dz'$,
$\delta n(z)=n(z)-\langle n(z) \rangle$ is the density fluctuation, and $\int_{-\infty}^{+\infty}\!dZ{\cal A}(Z)\!=\!1$.
Isolating the one- and two-body terms, one has
\begin{equation}
\langle \delta n(z)\delta n(z')\rangle=n\delta(z-z')
+n^2 [g^{(2)}(z-z')-1].
\label{eq.deltan}
\end{equation}
The first term, when substituted into Eq.~(\ref{eq.fluctu}), accounts for Poissonian
level of fluctuations, $\kappa \langle N\rangle$. Therefore,
the measured \textit{sub}-Poissonian
fluctuations in Fig.~\ref{fig.varNchaud}
imply that
$g^{(2)}(z-z^{\prime})-1<0$.
Such
anti-bunching stems from quantum fluctuations. Indeed,
within the
Bogoliubov approximation, valid
for quasi-condensates,
one has~\cite{Deuar2009}
\begin{equation}
g^{(2)}(z\!-\!z')-1 \!=\int \frac{dk}{2\pi n} \left[2n_k f_k - (1-f_k)\right]e^{ik(z-z')},
\label{eq.g2Bogo}
\end{equation}
where $f_k=1/\sqrt{1+4/(l_\xi k)^2}$
and $n_k=1/\left(e^{\epsilon_k/k_BT}-1\right)$ is the thermal occupation
of the Bogoliubov collective mode of wavenumber $k$ and energy
$\epsilon_k=\hbar^2k^2/(2m\sqrt{1+4/(l_\xi k)^2})$, with
$l_\xi=\hbar/\sqrt{mgn}$ being
the healing length.
The first term in the rhs of Eq.~(\ref{eq.g2Bogo})
which accounts for thermal fluctuations is positive, whereas the second term
which is the contribution of quantum (i.e., zero temperature)
fluctuations is negative~\cite{Deuar2009}.
Therefore, the negativity of $g^{(2)}(z-z')-1$
implies that the quantum fluctuations give a larger contribution
to $g^{(2)}(z-z^{\prime})-1$ than the thermal ones.
It should be emphasised, however, that the quantity we measure is
$\langle \delta N^2\rangle $, and as we show below, for our large
values of $\Delta$ and $d$ it is still dominated
by thermal
(rather than quantum)
fluctuations.
This is because
the contribution to $\langle \delta N^2\rangle $ of the one-body
term
almost cancels out the contribution
of the zero-temperature two-body term.
Indeed, the contribution of quantum fluctuations to
$\langle \delta N^2\rangle$, calculated using Eqs.~(\ref{eq.fluctu}),~(\ref{eq.deltan}), and (\ref{eq.g2Bogo}),
is
\begin{equation}
\langle \delta N^2\rangle_{T=0}=
\frac{\langle N \rangle}{\Delta\pi}\int_{-\infty}^{\infty}
dk f_k\frac{1-\cos(k\Delta)}{k^2} e^{-k^2d^2}.
\label{eq.fluctuquant}
\end{equation}
Since $f_k \propto kl_\xi$ when $kl_\xi\ll 1$, we find that
for $\Delta\gg l_\xi,d$, $\langle \delta N^2\rangle_{T=0}$ scales
as $n l_\xi \ln(\Delta/l_\xi)$.
On the other hand, the thermal contribution
given by Eq.~(\ref{eq.thermo}), scales as
$\Delta T/g$. Therefore, the quantum
contribution becomes negligible
as $\Delta\rightarrow \infty$,
and the thermodynamic prediction of Eq.~(\ref{eq.thermo}) is
recovered~\cite{Klawunn2011}.
For our parameters, the contribution of Eq.~(\ref{eq.fluctuquant}) to
$\langle \delta N^2\rangle$ is
shown as a dotted
line in Fig.~\ref{fig.varNchaud}.
\begin{figure}
\includegraphics[width=8.5cm]{Fig_3.eps}
\caption{(a) Variance $\langle \delta N^2\rangle$ close to the strongly
interacting regime, for $t=5.4$. Different curves
are as in Fig.~\ref{fig.varNchaud}, but for
$\omega_{\perp}/2\pi\!=\!18.8$~kHz,
$\omega_{\parallel}\!=\!7.5$~Hz, $T\!=\!40$~nK~($k_BT\!=\!0.044\hbar\omega_\perp$),
and $\kappa\!=\!0.47$.
(b) Average density profile (solid line) together with the Yang-Yang prediction (dashes).
(c) The value of $t$ obtained from fits to the density profile (dotted line)
and atom number fluctuations (solid line) for different $\alpha$ (see text).
}
\label{fig.varNfroid}
\end{figure}
In weakly interacting gases, the atom number
fluctuations take super-Poissonian values
in the degenerate ideal gas and thermal quasi-condensate regimes,
$\langle \delta N^2\rangle/\langle N\rangle$ reaching its maximum at
the quasi-condensate transition where it scales as
$t^{1/3}$ \cite{Armijo10_2}.
When $t$ is decreased, the super-Poissonian zone is
expected to merge towards the Poissonian
limit and it vanishes when the gas enters the strongly interacting regime.
This trend is exactly what we observe in
Fig.~\ref{fig.varNfroid}(a), for $t\!=\!5.4$:
at large densities, we
see suppression of $\langle \delta N^2\rangle$ below the Poissonian level
but, most importantly, we no longer observe super-Poissonian fluctuations at lower densities
($\langle \delta N^2\rangle /\kappa \langle N\rangle\!<\!1.3$ within the
experimental resolution) \footnote{
Fluctuations are, however, still much larger than those
of a Fermi gas for our (not very small) value
of $t$.}.
Interestingly, no simple
analytic
theory is applicable to this
crossover region, and the only reliable
prediction here is the exact
Yang-Yang
thermodynamic
solution [solid line in Fig.~\ref{fig.varNfroid}(a)].
We now describe the experimental techniques that
allowed us to
increase significantly $\omega_\perp$
in order to reach $t=5.4$.
Keeping a reasonable heat dissipation in the wires,
increasing $\omega_\perp$ requires bringing the atomic cloud closer to the chip.
However, using dc micro-wire currents, one would observe
fragmentation of the cloud
due to wire imperfections and hence longitudinal roughness of the
potential~\cite{Esteve2004}.
To circumvent this problem, we use the modulation techniques
developed in~\cite{Trebbia2007,Bouchoule2008}. The atom chip schematic is
shown in Fig.~\ref{fig.modulatedguide}.
The transverse confinement is realized by three wires, carrying
the same ac current modulated at $200$~kHz,
and a longitudinal homogeneous
dc magnetic field of $\sim\!1.8$~G realized by external coils.
The modulation is fast enough so that the atoms experience the
time-averaged potential, transversely harmonic.
Monitoring dipole oscillations we measure
$\omega_\perp/2\pi$ varying from $2$
to $25$~kHz, for ac current amplitude varying from $40$ to $200$~mA.
The longitudinal confinement, with $\omega_z/2\pi$ varying
from $5$ to $12$ Hz, is realised by
wires perpendicular
to the $z$-direction, carrying
dc currents of a few tens of mA. After a
first rf evaporation stage in a dc trap
we load $6\times10^4$ atoms at a few $\mu$K in the ac trap where we
perform further rf evaporation at $\omega_{\perp}/2\pi\simeq 2$~kHz
and $\omega_{\parallel}/2\pi\simeq 12$~Hz. Next we lower the
longitudinal trapping frequency to about $7$~Hz and then ramp up the
transverse frequency to $18.8$~kHz
in $600$ ms keeping the rf evaporation on
during this compression. After ramping the rf power
down in $100$ ms and letting
the cloud to thermalize for $150$ ms, we switch off the wire currents and image
the atomic cloud after $50$~$\mu$s
with a $60$~$\mu$s long resonant probe pulse.
The probe
is circularly polarised and its intensity, chosen to
optimise the signal to noise ratio, is about $0.2\, I_{\mathrm{sat}}$,
where $I_{\mathrm{sat}} = 1.67$~mW is the saturation intensity of the
D$_2$ line.
Finally, we get
the longitudinal profile of the cloud by
summing over the transverse pixels.
We typically obtain clouds at $t\simeq 1-6$.
Taking a few hundreds
of images under the same conditions, we
measure $\langle \delta N^2\rangle$
the same way as for the results of Fig.~\ref{fig.varNchaud}.
\begin{figure}[t]
\includegraphics[width=8.4cm]{Fig_4.eps}
\caption{(color online) (a) Wire schematic of the atom chip: three gold wires
along $\mathbf{Z}$ carry an ac
current and produce a tight transverse confining potential.
The longitudinal
confinement is realized with dc currents $I_1$ and $I_2$.
(b) The wires are buried under a layer of resist,
which ensures electrical insulation and surface planarization.
The resist is covered with
200~nm thick gold mirror that reflects the probe beam.
The atoms are $15\!$~$\mu$m away
from the wires and see the interference pattern produced by the
probe and the reflected beam. (c)
Typical optical-density image of a gas of $10^3$ atoms.
}
\label{fig.modulatedguide}
\end{figure}
One crucial point to correctly determine the longitudinal profile
is the knowledge of the absorption cross section $\sigma$.
In our setup,
atoms sit in an interference pattern during the imaging pulse [see Fig.~\ref{fig.modulatedguide}(b)]
and are subjected to a magnetic field so that
the determination of $\sigma$ is not simple.
Following~\cite{Reinaudi}, we assume $\sigma = \alpha \sigma_\mathrm{o}/\left(1+\alpha I/I_{\mathrm{sat}}\right)$, where $I$ is the intensity of
the probe beam, $\sigma_\mathrm{o}=3\lambda^2/2\pi$ is the resonant
cross section
of the transition
$|F=2, m_F=2\rangle \rightarrow |F'=3, m_F'=3\rangle$,
and $\alpha$ is a numerical factor. Solving the
optical Bloch equations (OBE) for our probe intensity and duration,
we find that such a law is valid, and we obtain $\alpha = 0.75$.
In this calculation, we averaged $\alpha$ over the distance to the chip,
which is expected to be valid as atoms diffuse over a rms width of
about $1~\mu m$
during the imaging pulse, which is larger than the interference lattice period.
The factor $\alpha$ can be also deduced from the mean density
profile and/or the atom number fluctuations using the thermodynamic
Yang-Yang predictions.
Fitting both $\alpha$ and $t$ to either the mean profile or the
fluctuations leads to strongly correlated values of $\alpha$ and $t$ but with large
uncertainty in $\alpha$.
Combining both pieces of information, however, enables a precise
determination of $\alpha$.
More specifically, using the Yang-Yang theory, we extract $t_p$ and $t_f$ from fits
to the mean profile and the fluctuations, respectively,
for various values of $\alpha$ [see Fig.~\ref{fig.varNfroid}(b) and (c)].
The intersection $t_p=t_f$ gives the correct value of $\alpha$.
We find $\alpha=0.77$, in good
agreement with the OBE calculation. The corresponding value of $t$ is $5.4$ and hence $T=40$~nK.
In summary, we have realised for the first time a \textit{single} 1D Bose gas close to the strongly
interacting regime. In contrast to realisations of
arrays of multiple 1D gases in 2D optical lattices \cite{Phillips-2004,Weiss2005},
our experiments have allowed us to
perform atom number fluctuation measurements in small
slices of the gas, not possible with multiple 1D gases.
In the weakly interacting regime, we
reached the \textit{quantum} quasi-condensate regime (where $k_BT\!<\!\mu\!=\!gn$) in a strictly 1D situation with $k_BT\!\ll\!\hbar\omega_\perp$.
Although the two-body correlation function $g^{(2)}$ is
dominated by quantum fluctuations in this regime, we have
shown that the variance $\langle \delta N^2\rangle$ is
still dominated by thermal excitations.
To resolve quantum fluctuations one would need to access wavelengths
smaller than
the phonon thermal wavelength $\hbar^2/mk_BT l_\xi$
\cite{Klawunn2011}, which is in the submicron range for our parameters.
Our work opens up further opportunities in the study of 1D Bose gases,
such as better understanding of the mechanisms of thermalisation and the
role of three-body correlations.
\begin{acknowledgments}
The authors acknowledge support by the IFRAF Institute, the
ANR Grant No. ANR-08-BLAN-0165-03, the ARC Discovery Project Grant No. DP110101047, and the CoQuS Graduate school of the FWF and the Austro-French FWF-ANR Project I607.
\end{acknowledgments}
|
\section{Introduction}\label{sec1}
Dark energy (DE) and dark matter (DM) are the dominant components in
the current universe, according to the recent astronomical
observations \cite{Riess98,Tegmark04,Spergel03}. They together
account for about 96\% of the critical energy density of the
universe. However, ironically, we have known little about the
natures of DE and DM. Although DM is ``dark'' (nonluminous), its
gravitational property is normal, i.e., it gravitationally behaves
like the usual baryon matter and thus can form structures in the
universe. The property of DE is much more exotic in that it is
gravitationally repulsive and so responsible for the accelerated
expansion of the universe \cite{dereview}.
Since we are ignorant of the natures of both DE and DM, we cannot
ignore such an important possibility that there is some direct,
non-gravitational interaction between DE and DM. Intriguingly, such
a possible interaction between DE and DM plays a crucial role in
helping solve (or, at least alleviate) several important theoretical
problems of DE. For example, it can be used to understand the cosmic
coincidence problem \cite{intde1}, to avoid the cosmic doomsday
brought by phantom \cite{intde2}, and to solve the cosmic age
problem caused by old quasar \cite{intde3} as well. Therefore, it is
very meaningful to seriously study the interaction between dark
sectors. Owing to the lack of the knowledge of micro-origin of the
interaction, one has to first phenomenologically propose an
interacting DE model and then test its theoretical and observational
consequences. So far, lots of phenomenological interacting DE models
have been studied~\cite{intdeOdintsov,intde4}.
When considering the interaction between dark sectors, the
continuity equations for energy densities of DE and DM are of the
form: \be\label{eq1} \dot{\rho}_{de}+3H(1+w_{de})\rho_{de}=-Q, \ee
\be\label{eq2} \dot{\rho}_{dm}+3H\rho_{dm}=Q, \ee where $\rho_{de}$
and $\rho_{dm}$ are the energy densities of DE and DM, respectively,
$H=\dot{a}/a$ is the Hubble parameter, $a$ is the scale factor of
the Friedmann-Robertson-Walker (FRW) universe,
$w_{de}=p_{de}/\rho_{de}$ is the equation of state (EOS) parameter
of DE, a dot denotes the derivative with respect to cosmic time $t$,
and $Q$ denotes the phenomenological interaction term. Several forms
for $Q$ have been put forward and have been fitted with
observations~\cite{intde5}. Of course, all of these models are
phenomenological. Most of them are constructed specifically for
mathematical simplicity --- for example, models in which $Q\propto
H\rho$, where $\rho$ denotes the energy density of the dark sectors,
and usually it has three choices, namely, $\rho=\rho_{dm}$,
$\rho=\rho_{de}$, and $\rho=\rho_{de}+\rho_{dm}$. In addition, there
are also some models~\cite{11Valiviita:2009nu} in which the
assumption of that $Q$ is proportional to the Hubble parameter is
abandoned and thus $Q\propto\rho$. Such models are designed by
consulting the simple models of reheating, of dark matter decay into
radiation, and of curvaton decay --- i.e., where the interaction has
the form of a decay of one species into another, with constant decay
rate. However, it should be stressed that these models are severely
dependent on the man-made choice of the special interaction forms.
In other words, the predictions and observational consequences are
model-dependent. In particular, the abovementioned phenomenological
models exclude the important possibility that the interaction
changes sign during the cosmological evolution. It is of interest to
point out that a sign-changeable or oscillatory form of interaction
is possible, according to the current
observations
Recently, Cai and Su~\cite{12Cai:2009ht}, without choosing a special
phenomenological form of interaction, proposed a novel scheme in
which the whole redshift range is divided into a few bins and the
interaction term $\delta(z)$ (note that here $Q=3H\delta$) is set to
be a constant in each bin, and by fitting the observational data
they found that $\delta(z)$ is likely to cross the noninteracting
($\delta=0$) line. This study is fairly enlightening and inspires us
to open our mind to seriously consider the possibility of that the
interaction between dark sectors changes sign during the
cosmological evolution. If such an observation is conclusive, it is
suggested that more general phenomenological forms of interaction
term should be put forth. However, the work of Cai and
Su~\cite{12Cai:2009ht} seems not sufficient to prove that the
interaction changes its sign, since their conclusion is drawn based
upon the behavior of the best-fitted $\delta(z)$, and the errors of
the fitting results weaken the conclusion to a great degree due to
the fact that the observational data currently available cannot
determine more than two parameters in a piecewise parametrization
approach.
In this paper, our aim is to verify whether the interaction indeed
changes its sign (i.e., crosses the noninteracting line) during the
evolution, by using a different method. We propose a parametrization
form for the interaction term $Q$. In our work, we further abandon
the assumption that $Q$ is proportional to the Hubble expansion rate
$H$. So, the interaction term $Q$ is only characterized by the
coupling $b$: \be\label{eq3} Q(a)=3b(a)H_0\rho_0, \ee where the
dimensionless coupling $b(a)$ is variable with the cosmological
evolution. Note that here the occurrence of the present-day Hubble
parameter $H_0$ and the present-day density of dark sectors
$\rho_0=\rho_{de0}+\rho_{dm0}$ is only for a dimensional reason. By
the way, in the whole work the subscript ``0'' always indicates the
present-day value of the corresponding quantity. From the form of
Eq.~(\ref{eq3}), we see that the evolution of the interaction term
$Q$ is totally described by the running of the coupling constant,
$b(a)$, so our scenario can be called the ``running coupling''.
Furthermore, we assume that the coupling $b$ is described by a
constant $b_0$ at the late times, and determined by another constant
$b_e$ at the early times; and the whole evolution of $b(a)$ is
totally characterized by the two parameters, $b_0$ and $b_e$. For
continuously connecting the early-time and late-time behaviors, we
put forward the following two-parameter form for the coupling
$b(a)$: \be\label{eqb} b(a)=b_0a+b_e(1-a). \ee Though the
interaction term in our work depends on a particular parametrization
form, the parameters can be tightly constrained by the current
observational data, overcoming the disadvantage of the piecewise
fitting method. The reconstructed evolution of $b(a)$ will indicate
whether the coupling between the dark sectors crosses the
noninteracting line.
In this paper, we will investigate our coupling parametrization with
the latest observational data. For the EOS of DE, $w$, we consider
the following three cases: (1) the cosmological constant (vacuum
energy), $w=-1$; (2) the constant EOS, $w=w_0$; (3) the
time-variable EOS, namely, the Chevallier-Polarski-Linder (CPL)
parametrization, $w(a)=w_0+w_1(1-a)$~\cite{13CPL}. We will fit the
three interacting DE models with the data from the Union2 type Ia
supernovae (SNIa), the baryon acoustic oscillation (BAO), the cosmic
microwave background (CMB), the Hubble expansion rate, and the X-ray
gas mass fraction. We obtain the best-fitted parameters and
likelihoods by using the Monte Carlo Markov chain (MCMC) method. We
will show that the interaction $Q$ between DE and DM indeed changes
sign around $z=0.2-0.3$ during the cosmological evolution, at about
1$\sigma$ confidence level (CL).
\section{Methodology}\label{sec2}
We consider interacting DE models in a spatially flat FRW universe.
The Friedmann equation reads \be\label{eq4}
3M_{Pl}^2H^2=\rho_r+\rho_b+\rho_{de}+\rho_{dm}, \ee where $\rho_r$,
$\rho_b$, $\rho_{de}$ and $\rho_{dm}$ are the energy densities of
radiation, baryon, DE and DM, respectively, and $M_{Pl}$ is the
reduced Planck mass. It is convenient to introduce the fractional
energy densities $\Omega_i\equiv\rho_i/3M_{Pl}^2H^2$, with $i =r,\
b,\ de,$ and $dm$. Obviously, \be\label{eq5}
\Omega_r+\Omega_b+\Omega_{de}+\Omega_{dm}=1. \ee
Substituting Eqs.~(\ref{eq3}) and (\ref{eqb}) into Eqs.~(\ref{eq1})
and (\ref{eq2}), and defining the functions
$f_{de}=\rho_{de}/\rho_{de0}$ and $f_{dm}=\rho_{dm}/\rho_{dm0}$, we
obtain \be\label{eq8}
\frac{df_{de}(x)}{dx}+3(1+w)f_{de}(x)=-\frac{3}{E(x)}\left(1+\frac{1}{f_0}\right)\left[b_0e^x+b_e(1-e^x)\right],
\ee \be\label{eq9}
\frac{df_{dm}(x)}{dx}+3f_{dm}(x)=\frac{3}{E(x)}(1+f_0)\left[b_0e^x+b_e(1-e^x)\right],
\ee where $x\equiv\ln{a}$,
$f_0\equiv\rho_{de0}/\rho_{dm0}=\Omega_{de0}/\Omega_{dm0}=(1-\Omega_{r0}-\Omega_{b0}-\Omega_{dm0})/\Omega_{dm0}$,
and $E(x)\equiv
H(x)/H_0=[\Omega_{r0}e^{-4x}+\Omega_{b0}e^{-3x}+\Omega_{dm0}f_{dm}(x)+(1-\Omega_{r0}-\Omega_{b0}-\Omega_{dm0})f_{de}(x)]^{1/2}.$
Therefore, given the values of the parameters $\Omega_{r0}$,
$\Omega_{b0}$, $\Omega_{dm0}$ and $w$, Eqs.~(\ref{eq8}) and
(\ref{eq9}) can be numerically solved with the initial conditions
$f_{de}(0)=1$ and $f_{dm}(0)=1$. With the resulting functions
$f_{de}(x)$ and $f_{dm}(x)$, we finally obtain the function $E(x)$,
the dimensionless Hubble expansion rate. As aforementioned, in this
work we employ three DE models, namely, the cosmological constant
model ($\Lambda$CDM) with $w=-1$, the constant EOS model (XCDM) with
$w=w_0$, and the time-variable EOS model (CPL) with
$w(x)=w_0+w_1\left(1-e^x\right)$.
To fit the three interacting DE models with observations, we use the
data from the Union2 SNIa (557 data), the BAO from SDSS DR7, the CMB
from 7-year WMAP, the Hubble expansion rate (15 data), and the X-ray
gas mass fraction (42 data). The best-fitted parameters are obtained
by minimizing the sum \be\label{eq33}
\chi^2=\tilde{\chi}^2_{SN}+\chi^2_{BAO}+\chi^2_{CMB}+\chi^2_{H}+\chi^2_{X-ray}.
\ee We obtain the constraints by using a MCMC method.
{\it Supernovae.}--- We use the data points of the 557 Union2 SNIa
compiled in Ref.~\cite{14Amanullah:2010vv}. The theoretical distance
modulus is defined as \be\label{eq11} \mu_{th}(z_i)\equiv5\log_{10}
D_L(z_i)+\mu_0, \ee where $z={1/a}-1$ is the redshift,
$\mu_0\equiv42.38-5\log_{10} h$ with $h$ the Hubble constant $H_0$
in units of 100 km/s/Mpc, and the Hubble-free luminosity distance
\be\label{eq12} D_L(z)=(1+z)\int_0^z \frac{dz'}{E(z';{\bm \theta})},
\ee where ${\bm\theta}$ denotes the model parameters.
Correspondingly, the $\chi^2$ function for the 557 Union2 SNIa data
is given by \be\label{eq13}
\chi^2_{SN}({\bm\theta})=\sum\limits_{i=1}^{557}\frac{\left[\mu_{obs}(z_i)-\mu_{th}(z_i)\right]^2}{\sigma^2(z_i)},
\ee where $\sigma$ is the corresponding $1\sigma$ error of distance
modulus for each supernova. The parameter $\mu_0$ is a nuisance
parameter and one can expand Eq.~(\ref{eq13}) as \be\label{eq14}
\chi^2_{SN}({\bm\theta})=A({\bm\theta})-2\mu_0
B({\bm\theta})+\mu_0^2 C, \ee where $A({\bm\theta})$,
$B({\bm\theta})$ and $C$ are defined in Ref.~\cite{Nesseris:2005ur}.
Evidently, Eq.~(\ref{eq14}) has a minimum for $\mu_0=B/C$ at
\be\label{eq15}
\tilde{\chi}^2_{SN}({\bm\theta})=A({\bm\theta})-\frac{B({\bm\theta})^2}{C}.
\ee Since $\chi^2_{SN,\,min}=\tilde{\chi}^2_{SN,\,min}$, instead
minimizing $\chi_{SN}^2$ we will minimize $\tilde{\chi}^2_{SN}$
which is independent of the nuisance parameter $\mu_0$.
{\it Baryon acoustic oscillations.}--- We use the BAO data from SDSS
DR7~\cite{15Percival:2009xn}. The distance ratio ($d_z$) at $z=0.2$
and $z=0.35$ are \be\label{eq16}
d_{0.2}=\frac{r_s(z_d)}{D_V(0.2)},~~
d_{0.35}=\frac{r_s(z_d)}{D_V(0.35)}, \ee where $r_s(z_d)$ is the
comoving sound horizon at the baryon drag
epoch~\cite{16Eisenstein:1997ik}, and \be\label{eq17}
D_V(z)=\left[\left(\int_0^z\frac{dz'}{H(z')}\right)^2\frac{z}{H(z)}\right]^{1/3}
\ee encodes the visual distortion of a spherical object due to the
non Euclidianity of a FRW spacetime. The inverse covariance matrix
of BAO is \bq\label{eq18} (C^{-1}_{BAO}) & = &
\left(\begin{array}{ccc}
30124 & -17227 \\
-17227 & 86977\end{array}\right).\eq The $\chi^2$ function of the
BAO data is constructed as: \be\label{eq19}
\chi_{BAO}^2=(d_i^{th}-d_i^{obs})(C_{BAO}^{-1})_{ij}(d_j^{th}-d_j^{obs}),
\ee where $d_i=(d_{0.2}, d_{0.35})$ is a vector, and the BAO data we
use are $d_{0.2}=0.1905$ and $d_{0.35}=0.1097$.
{\it Cosmic microwave background.}--- We employ the ``WMAP distance
priors'' given by the 7-year WMAP observations~\cite{17WMAP7}. This
includes the ``acoustic scale'' $l_A$, the ``shift parameter'' $R$,
and the redshift of the decoupling epoch of photons $z_*$. The
acoustic scale $l_A$ describes the distance ratio
$D_A(z_*)/r_s(z_*)$, defined as \be\label{eq21} l_A\equiv
(1+z_*){\pi D_A(z_*)\over r_s(z_*)}, \ee where a factor of $(1+z_*)$
arises because $D_A(z_*)$ is the proper angular diameter distance,
whereas $r_s(z_*)$ is the comoving sound horizon at $z_*$. The
fitting formula of $r_s(z)$ is given by \be\label{eq22}
r_s(z)=\frac{1} {\sqrt{3}} \int_0^{1/(1+z)} \frac{da}
{a^2H(a)\sqrt{1+(3\Omega_{b0}/4\Omega_{\gamma0})a}}. \ee In this
paper, we fix $\Omega_{\gamma0}=2.469\times10^{-5}h^{-2}$,
$\Omega_{b0}=0.02246 h^{-2}$, given by the 7-year WMAP
observations~\cite{17WMAP7}, and $\Omega_{r0}=\Omega_{\gamma0}(1 +
0.2271N_{\rm eff})$ with $N_{\rm eff}$ the effective number of
neutrino species (in this paper we take its standard value, 3.04
\cite{17WMAP7}). We use the fitting function of $z_*$ proposed by Hu
and Sugiyama~\cite{18Hu:1995en} \be\label{eq23}
z_*=1048[1+0.00124(\Omega_{b0}
h^2)^{-0.738}][1+g_1(\Omega_{m0}h^2)^{g_2}], \ee where
$\Omega_{m0}=\Omega_{b0}+\Omega_{dm0}$ and \be\label{eq24}
g_1=\frac{0.0783(\Omega_{b0}h^2)^{-0.238}}{1+39.5(\Omega_{b0}h^2)^{0.763}},\quad
g_2=\frac{0.560}{1+21.1(\Omega_{b0}h^2)^{1.81}}. \ee The shift
parameter $R$ is responsible for the distance ratio
$D_A(z_*)/H^{-1}(z_*)$, given by~\cite{19Bond97} \be\label{eq25}
R(z_*)\equiv\sqrt{\Omega_{m0}H_0^2}(1+z_*)D_A(z_*). \ee
Following~\cite{17WMAP7}, we use the prescription for using the WMAP
distance priors. Thus, the $\chi^2$ function for the CMB data is
\be\label{eq26}
\chi_{CMB}^2=(x^{th}_i-x^{obs}_i)(C_{CMB}^{-1})_{ij}(x^{th}_j-x^{obs}_j),
\ee where $x_i=(l_A, R, z_*)$ is a vector, and $(C_{CMB}^{-1})_{ij}$
is the inverse covariance matrix. The 7-year WMAP
observations~\cite{17WMAP7} give the maximum likelihood values:
$l_A(z_*)=302.09$, $R(z_*)=1.725$, and $z_*=1091.3$. The inverse
covariance matrix is also given in~\cite{17WMAP7}: \bq\label{eq27}
(C_{CMB}^{-1})=\left(\begin{array}{ccc}
2.305 & 29.698 & -1.333 \\
29.698& 6825.27 & -113.180 \\
-1.333& -113.180 & 3.414 \\
\end{array}\right). \eq
{\it Hubble expansion rate.}--- For the Hubble parameter $H(z)$,
there are 15 observational data available, where twelve of them are
from Ref.~\cite{20Stern:2009ep}. In addition, in
Ref.~\cite{21Gaztanaga:2008xz}, the authors obtain the additional
three data: $H(z=0.24)=79.69\pm2.32$, $H(z=0.34)=83.8\pm2.96$, and
$H(z=0.43)=86.45\pm3.27$ (in units of km s$^{-1}$ Mpc$^{-1}$). The
$\chi^2$ function for the observational Hubble data is
\be\label{eq28} \chi_{H}^2({\bm\theta})=\sum_{i=1}^{15}
\frac{[H_{th}({\bm\theta};z_i)-H_{ obs}(z_i)]^2}{\sigma^2(z_i)}. \ee
{\it X-ray gas mass fraction.}--- For the $f_{gas}$ data, we use the
Chandra measurements in Ref.~\cite{22Allen:2007ue}. In the framework
of the $\Lambda$CDM reference cosmology, the X-ray gas mass fraction
is presented as~\cite{22Allen:2007ue,23Goncalves:2009kd}
\be\label{eq29} f_{gas}(z)=\frac{KA\gamma{b(z)}}{1+s(z)}
\left(\frac{\Omega_{b0}}{\Omega_{b0}+\Omega_{dm0}F(z)}\right)
\left[\frac{D_A^{\Lambda{CDM}}(z)}{D_A(z)}\right]^{1.5}, \ee where
the effect of the interaction between dark sectors has been
considered, resulting in an additional function
$F(z)=f_{dm}(-\ln(1+z))/(1+z)^3$. The parameters $K$, $\gamma$,
$b(z)$ and $s(z)$ model the abundance of gas in the clusters. We set
these parameters to their respective best-fit values of
Ref.~\cite{22Allen:2007ue}. $A$ is the angular correction factor,
which is caused by the change in angle for the current test model
$\theta_{2500}$ in comparison with that of the reference cosmology
$\theta_{2500}^{\Lambda{CDM}}$: \be\label{eq30}
A=\left(\frac{\theta_{2500}^{\Lambda{CDM}}}{\theta_{2500}}\right)^\eta\approx
\left(\frac{H(z)D_A(z)}{[H(z)D_A(z)]^{\Lambda{CDM}}}\right)^\eta,
\ee here, the index $\eta$ is the slope of the $f_{gas}(r/r_{2500})$
data within the radius $r_{2500}$, with the best-fit average value
$\eta=0.214\pm0.022$~\cite{22Allen:2007ue}. And the proper (not
comoving) angular diameter distance is given by \be\label{eq31}
D_A(z)=\frac{1}{(1+z)}\int_0^z\frac{dz'}{H(z')}. \ee The $\chi^{2}$
function for the $f_{gas}$ data from the 42 galaxy clusters reads
\be\label{eq32} \chi^2_{X-ray}({\bm\theta})=\sum_{i=1}^{42}
\frac{([f_{gas}({\bm\theta};z_{i})]_{th}-[f_{gas}(z_{i})]_{obs})^{2}}{\sigma^{2}(z_{i})}.
\ee It should be pointed out that the $f_{gas}$ data are rather
crucial for the fitting, since they can be used to break the
degeneracy between the parameters from the interaction and the EOS
of CPL model. So, the inclusion of the $f_{gas}$ data in our fitting
is indispensable.
\section{Results}\label{sec3}
Now we fit our interacting models with the observations. For the
interacting $\Lambda$CDM, XCDM and CPL models, the parameters are
${\bm\theta}=\{\Omega_{dm0},~b_0,~b_e,~h\}$,
$\{\Omega_{dm0},~w_0,~b_0,~b_e,~h\}$ and
$\{\Omega_{dm0},~w_0,~w_1,~b_0,~b_e,~h\}$, respectively. We use the
MCMC method and finally we obtain the best-fit parameters and the
corresponding $\chi^2_{min}$. The best-fit, $1\sigma$ and $2\sigma$
values of the parameters with $\chi^2_{min}$ of the three
interacting models are all presented in Table~\ref{table2}.
\begin{table*}\caption{The fitting results of the parameters with best-fit values as well as
$1\sigma$ and $2\sigma$ errors in the three interacting DE models.}
\begin{center}
\begin{tabular}{cc| cc cc cc}
\hline\hline model parameters & & $\Lambda$CDM & & XCDM & & CPL &
\\ \hline
$\Omega_{dm0}$ && $0.2262^{+0.0242 +0.0356}_{-0.0215 -0.0304}$ &
& $0.2267^{+0.0273 +0.0388}_{-0.0237 -0.0327}$ &
& $0.2271^{+0.0302 +0.0443}_{-0.0280 -0.0360}$ & \\
$w_0$ && N/A &
& $-0.9844^{+0.0915 +0.1282}_{-0.0932 -0.1357}$ &
& $-0.9768^{+0.3046 +0.4322}_{-0.2090 -0.2635}$ & \\
$w_1$ && N/A &
& N/A &
& $-0.0455^{+0.9426 +1.2742}_{-1.8125 -2.7810}$ & \\
$b_0$ && $0.0793^{+0.1220 +0.1797}_{-0.1092 -0.1580}$ &
& $0.0787^{+0.1353 +0.1962}_{-0.1159 -0.1654}$ &
& $0.0799^{+0.1512 +0.2039}_{-0.1342 -0.1894}$ & \\
$b_e$ && $-0.3274^{+0.4646 +0.6543}_{-0.5140 -0.7644}$ &
& $-0.3216^{+0.5007 +0.7082}_{-0.5771 -0.8317}$ &
& $-0.3290^{+0.5855 +0.8379}_{-0.6447 -0.8989}$ & \\
$h$ && $0.7120^{+0.0154 +0.0224}_{-0.0155 -0.0218}$ &
& $0.7094^{+0.0224 +0.0324}_{-0.0219 -0.0315}$ &
& $0.7097^{+0.0259 +0.0348}_{-0.0257 -0.0357}$ & \\
\hline $\chi^{2}_{min}$ && 595.968 & & 595.815 & & 595.808 & \\
\hline\hline
\end{tabular}
\label{table2}
\end{center}
\end{table*}
\begin{figure}[htbp]
\centering \noindent
\includegraphics[width=8cm]{LCDM.eps}
\caption{\label{fig1:LCDM}The probability contours at
$1\sigma$ and $2\sigma$ confidence levels in the parameter planes for the interacting $\Lambda$CDM model.}
\end{figure}
\begin{figure*}[htbp]
\centering \noindent
\includegraphics[width=14cm]{XCDM.eps}
\caption{\label{fig2:XCDM}The probability contours at $1\sigma$ and $2\sigma$ confidence levels in the parameter planes for the interacting XCDM model.}
\end{figure*}
\begin{figure*}[htbp]
\centering \noindent
\includegraphics[width=14cm]{CPL.eps}
\caption{\label{fig3:CPL}The probability contours at $1\sigma$ and $2\sigma$ confidence levels in the parameter planes for the interacting CPL model.}
\end{figure*}
\begin{figure}[htbp]
\centering \noindent
\includegraphics[width=5cm]{WCPL.eps}
\caption{\label{fig4:WCPL}The reconstructed evolutionary history for $w(z)$ of the interacting CPL model.}
\end{figure}
\begin{figure*}[htbp]
\centering\noindent
\includegraphics[width=5cm]{QLCDM.eps}\hskip.1cm
\includegraphics[width=5cm]{QXCDM.eps}\hskip.1cm
\includegraphics[width=5cm]{QCPL.eps}
\caption{\label{fig5:QALL} The reconstructed evolutionary histories for $b(z)$
in the three interacting models. The dashed line in each plot represents the noninteracting line.}
\end{figure*}
Figure~\ref{fig1:LCDM} shows the likelihood contours for the
interacting $\Lambda$CDM model. For this model, we have
$\Omega_{dm0}=0.2262$, $b_0=0.0793$, $b_e=-0.3274$ and $h=0.7120$,
with $\chi^2_{min}=595.968$. We plot the likelihood contours for the
interacting XCDM model in Fig.~\ref{fig2:XCDM}. For the interacting
XCDM model, the fitting results are $\Omega_{dm0}=0.2267$,
$w_0=-0.9844$, $b_0=0.0787$, $b_e=-0.3216$ and $h=0.7094$, with
$\chi^2_{min}=595.815$. For the interacting CPL model, we obtain the
fitting results: $\Omega_{dm0}=0.2271$, $w_0=-0.9768$, $w_1=-0.0455$
$b_0=0.0799$, $b_e=-0.3290$ and $h=0.7097$, with
$\chi^2_{min}=595.808$. The likelihood contours for this case are
shown in Fig.~\ref{fig3:CPL}.
From the fitting results, we first notice that the EOS of DE is near
$-1$ in both the constant $w$ and time-dependent $w$ scenarios. In
the constant $w$ case, we get the best-fitted value $w_0=-0.9844$,
rather close to $-1$, albeit slightly tends to a quintessence
($w>-1$). For the CPL case, the best-fitted values are $w_0=-0.9768$
and $w_1=-0.0455$; we also notice that $w_0$ is fairly near $-1$ and
$w_1$ is close to 0. So, in the interacting DE model with a running
coupling, even if the EOS of DE is allowed to vary, the data still
favor a slightly evolving EOS with the value approaching $-1$. To
see this clearly, we plot the reconstructed $w(z)$ for the
interacting CPL model in Fig.~\ref{fig4:WCPL} where the best fit and
the 1$\sigma$ uncertainties are shown. According the fitting
results, we realize that a time-varying vacuum scenario is favored
by the data. The time-varying vacuum model has been discussed
extensively \cite{runvac,Odintsov:2005ju}. Our work indicates that
the scenario of a time-varying vacuum with a running coupling
deserves more further investigations.
For the running coupling $b(z)$, we obtain similar results in all
the three scenarios. From Figs.~\ref{fig1:LCDM}--\ref{fig3:CPL} we
see that the parameters $b_0$ and $b_e$ are in strong
anti-correlation. The best-fitted values for $b_0$ and $b_e$ are:
$b_0\approx 0.08$ and $b_e\approx-0.3$, implying that the coupling
$b(z)$ crosses the noninteracting line $b=0$ during the cosmological
evolution, and the sign changes from negative to positive. Such a
feature for the interaction is favored by the data at about
1$\sigma$ level. To see the crossing phenomenon clearly, we
reconstruct the evolution of the coupling $b(z)$ by using a Fisher
Matrix technique, shown in Fig.~\ref{fig5:QALL}. From this figure we
read out that the crossing happens at $z=0.2-0.3$. Our results tell
such a story: at early times when DM dominates the universe, the
energy transfer direction is from DM to DE, and at late times when
DE becomes dominant, the decay direction reverses, from DE to DM.
The above phenomenon is favored at 1$\sigma$ CL. Note that once the
reconstruction of $b(z)$ is performed based on a Monte Carlo method,
the fluctuations will become larger, but the above distinctive
feature still stands by at about 1$\sigma$ CL.
In the work of Cai and Su~\cite{12Cai:2009ht}, a redshift-binned
parametrization method is used, but the limitation is that it is
hard to go beyond two parameters for tight constraints. Also, in
Ref.~\cite{12Cai:2009ht}, as the data give only weak constraint for
$z>1.8$, the additional assumption $Q=0$ for $1.8<z<1090$ is made.
In our method, the parametrization $Q(a)=3b(a)H_0\rho_0$ with
$b(a)=b_0a+b_e(1-a)$ has only two parameters, and the early-time
interaction can also be described. Our fitting results support the
conclusion of Cai and Su~\cite{12Cai:2009ht} that the interaction
$Q(z)$ crosses the noninteracting line, in a distinct way. The
limitations of our parametrization are: (1) whether or not there is
some oscillation feature in the interaction cannot be read out, and
(2) the future evolution cannot be described owing to the fact that
$b(z)$ will diverge as $z$ approaches $-1$, so the predictive power
of this parametrization is lost (see \cite{Ma:2011nc} for a similar
case for $w(z)$). We will go beyond these limitations in the future
work. In addition, now that the analysis of the current
observational data provides a hint that the interaction between dark
sectors might change sign during the cosmological evolution, more
general phenomenological forms for the interaction describing the
sign-changeable or oscillatory feature should be seriously
considered.
\section{Conclusion}\label{sec4}
Since the knowledge about the micro-origin of dark sector
interaction is absent, one has no way to know the form of the
interaction term $Q$ from a microscopic theory. A popular way to
investigate the interacting dark energy is to assume a specific
phenomenological form for $Q$; for instance, $Q\propto H\rho$ or
$Q\propto \rho$, where $\rho$ denotes the energy density of dark
sectors. However, the choice of the phenomenological forms is rather
arbitrary. In the face of this situation, let us recall the method
used to probe the dynamical evolution of the EOS of DE, $w(z)$, with
the observational data. Owing to our ignorance of DE, for probing
the dynamical evolution of $w$, one has to parameterize $w$
empirically, usually using two parameters, e.g., the most widely
used CPL form, $w(a)=w_0+w_1(1-a)$. Inspired by this method, we try
to parameterize the evolution of $Q$ in a similar way, and then use
the observational data to probe the dynamical evolution of $Q$. Such
a way may provide a guidance for finding a reasonable
phenomenological form for $Q$.
In this paper, we have put forward a running coupling scenario for
describing the interaction between DE and DM. In this scenario, the
dark sector interaction has the form $Q(a)=3b(a)H_0\rho_0$, where
$b(a)$ is the coupling which is variable during the cosmological
evolution. We have proposed a parametrization form for the running
coupling: $b(a)=b_0a+b_e(1-a)$. So, at the early times the coupling
is given by a constant $b_e$, while today the coupling is described
by another constant, $b_0$. We have constrained the parameters $b_0$
and $b_e$ with the observational data currently available, including
the Union2 SNIa, BAO (from SDSS DR7), CMB (from 7-year WMAP),
$H(z)$, and X-ray gas mass fraction data. It should be stressed that
the $f_{gas}$ data are very crucial in our analysis since they are
helpful in breaking the degeneracy between the parameters. And, in
our analysis, we employed three DE model, namely, the $\Lambda$CDM,
XCDM and CPL models.
Our fitting results show that the EOS of DE, $w$, is close to $-1$
both in the constant $w$ (XCDM) and time-dependent $w$ (CPL) cases.
Thus, a time-varying vacuum scenario is favored by the data,
according to this analysis. In addition, for the running coupling
$b(z)$, the results are also similar in all the three DE models. The
parameters $b_0$ and $b_e$ are in strong anti-correlation, and the
best-fitted values are: $b_0\approx 0.08$ and $b_e\approx-0.3$. This
implies that the coupling $b(z)$ crosses the noninteracting line
($b=0$) and the sign changes from $b<0$ to $b>0$. The reconstruction
of $b(z)$ indicates that the crossing of the noninteracting line
happens at around $z=0.2-0.3$, and the crossing behavior is favored
at about 1$\sigma$ CL. Therefore, our work tells a story about the
interacting DE model: DE is a time-varying vacuum; the coupling
between DE and DM runs with the expansion of the universe; at early
times when DM dominates the universe, DM decays to DE, while at late
times when DE becomes dominant, DE begins to decay to DM. If the
above scenario is true, then we should pay more attention to the
time-varying vacuum model, and seriously consider how to construct a
sign-changeable interaction between dark sectors phenomenologically.
For the time-varying vacuum model, previous work neglects the
perturbations of DE and only treats the decaying vacuum as a
background, however, if DE interacts with DM, the perturbations of
DE should also be taken into account even if $w=-1$. This deserves
further investigations.
Finally, we discuss the limitations of our parametrization for the
running coupling. Our parametrization cannot describe the possible
oscillation in the dark sector interaction. Moreover, such a
parametrization form cannot predict the future evolution of the
interaction, since the coupling $b(z)$ will encounter divergency in
the far future. We will go beyond these limitations in our future
work.
\begin{acknowledgments}
This work was supported by the Natural Science Foundation of China
under Grant Nos.~10705041 and 10975032, as well as the National
Innovation Experiment Program for University Students.
\end{acknowledgments}
|
\section{Introduction}
\noindent Half a century ago, visionary engineers and scientists
designed and built the world's first satellite communication station
choosing Goonhilly\ in Cornwall as its location (lat=50.0504$\,^{\circ}$
North, lon = 5.1835$\,^{\circ}$ West). Since then, Goonhilly\ has
pioneered many of the advances in global communication and inspired
budding scientists and engineers. Now in 2011, most needs for
satellite and tele-communications are met by using a combination of smaller antennae and
undersea cables, and there are plans afoot to use Goonhilly's valuable
assets for a variety of new purposes including radio astronomy.
\noindent There are three 30-m-class telescopes at Goonhilly, and two of these
could straightforwardly be adapted for radio astronomy. Goonhilly-1
(or "Arthur") can be most easily incorporated into e-MERLIN\ as its
design was based on the Jodrell Bank MKII Telescope that is already a
mainstay of the network. Goonhilly-3 (or "Guinevere") is perhaps a more suitable
Goonhilly antenna to use for higher-frequency (especially K-band)
observations.
Assuming that the chosen antennae can be outfitted with standard
e-MERLIN\ L-, C- or K-band receivers, and connected to the eighth "spare"
input of the correlator (information on e-MERLIN\ can be found via
http://www.e-merlin.ac.uk), the improvements that the addition of a
Goonhilly\ antenna would bring to the enhanced array performance, and
consequently the scientific return of the e-MERLIN\ Legacy observing
programmes, are discussed.
\section{Including Goonhilly and using the Lovell Telescope in
the e-MERLIN\ array}
\noindent The imaging capability of an interferometer is related to
the distribution of its antennae and their individual
sensitivities. Here the UV plane and imaging benefits are discussed if
a "Goonhilly" antenna\footnote{Note that in addition to the
already mentioned telescopes at the Goonhilly side there is another
30-m-class telescope (Goonhilly 6) available. If all three 30-m-class telescopes could be
phased up then the equivalent of a 50-m-class telescope could be delivered by Goonhilly resources.}
is included in the e-MERLIN\ array and if the Lovell Telescope\ (LT) is used.
\subsection{UV plane benefits}
\begin{figure}
\begin{center}
\includegraphics[width=.45\textwidth]{hrk_1a.png}
\includegraphics[width=.45\textwidth]{hrk_1b.png}\\
\includegraphics[width=.45\textwidth]{hrk_1c.png}
\includegraphics[width=.45\textwidth]{hrk_1d.png}\\
\end{center}
\caption{The left column shows the UV coverage of the e-MERLIN\ array consisting of the Lovell Telescope [76~m]
(the MKII [25~m] was not used), Cambridge [32~m], Defford [25~m], Knockin
[25~m], Darnhall [25~m] and Pickmere [25~m]) telescopes. The right column shows UV coverages generated by including
the Goonhilly telescope. The UV coverages in the top row are
based on a 12-hour, L-band simulation of a source at a declination
of 60$\,^{\circ}$, whereas the UV coverages in the lower row are for a
source at declination 0$\,^{\circ}$. Note that the UV coverage
exhibits a radial spread as a function of frequency, and the
individual tracks of the baseline should thus practically fill the
UV plane when e-MERLIN\ observes in full multi-frequency synthesis
mode. For clarity however, the simulation here presents a processed
bandwidth of 400~MHz which has been split into 8 channels of 50~MHz
each and therefore individual tracks are clearly visible,
particularly for the longest baselines. }
\label{fig1}
\end{figure}
\noindent Generally the angular resolution of an interferometer is determined by
the longest "projected" baseline (the value of which [in radians] is
approximately $\lambda / b$, for which $\lambda$ is the
observing wavelengh and b is the length of the longest baseline,
see e.g. Rohlfs \& Wilson 2004). With its current setup, e-MERLIN\ has a
maximum baseline of 217~km which provides an angular resolution
of 170~milli-arcsec (mas) at L-band. Introducing one of the
telescopes at Goonhilly would increase the maximum baseline to
440~km, thus doubling the angular resolution of e-MERLIN,
bringing it to values of 82~mas, 17~mas,
and 6~mas at L-, C- and K-band respectively.
Figure~\ref{fig1} affords a more detailed view on the
improved imaging capabilities, showing the L-Band UV coverages of
e-MERLIN\ and e-MERLIN\ including Goonhilly at two different declinations
(60$\,^{\circ}$ and 0$\,^{\circ}$). The UV coverage is based on a
12-hour synthesis\footnote{The UV coverages are derived from an array
simulation using the AIPS task {\tt UVCON}.}. For L-band
observations the e-MERLIN\ bandwidth will be of the order of 400~MHz
(1.3--1.7 GHz), which is simulated as 8 channels of 50~MHz bandwidth
each. At 60$\,^{\circ}$ declination the UV coverage is circularly
symmetric and would almost fully cover all spatial frequencies up to
the longest baselines. Due to the location of Goonhilly the array
would have almost equally long east-west and north-south baselines,
symmetrically extending the UV-coverage by a factor of $\sim$2. Such a
configuration would significantly improve the snap-shot imaging
capability with respect to the current e-MERLIN\ configuration. As the
declination of an observation moves towards zero the UV coverage
becomes increasingly "squashed" in the north-south direction, which results in an elongated
synthesized beam\footnote{The UV coverage and synthesised beams are
directly coupled in that they are Fourier transforms of one
another.}. This is a direct result of having "projected" baselines
which directly influences the imaging capability of the e-MERLIN\
interferometer. For example, a 12-hour observation on the celestial
equator and using uniform weighting would result in angular resolutions of 320~mas$\times$140~mas
for e-MERLIN, and 140~mas$\times$72~mas with the additional station at
Goonhilly. For further discussion on equatorial imaging using e-MERLIN\
and telescopes located in Chilbolton and in Goonhilly see Heywood et al. (2008,
2011).
In order to fully exploit such UV coverage new wide-band
imaging algorithms must be used. Properly accounting for sources which
are morphologically complex and exhibit components with different
spectral indices is a major challenge for calibration and
imaging. This is not the case for spectral line observations during which
the UV coverage advantage gained by multi-frequency synthesis is lost,
as each spectral channel will only occupy a narrow region of the UV
plane.
\subsection{The image sensitivity}
\noindent Adding a new antenna into an existing array provides
enhanced imaging sensitivity and depending on its location it will
potentially enhance the spatial sensitivity by adding baselines that
traces new sets of spatial scales. Here some basic radio astronomy
relationships are presented in order to quantify such benefits.
The baseline sensitivity [in units of Jy] at a single polarisation in the real or the imaginary component of the visibility
formed between antennas $i$ and $j$ is
given by:
\begin{eqnarray}
\Delta S^2_{ij} = \eta^2_{b} \, \frac{T_{i} T_{j}}{2 \Delta t \Delta \nu K_{i} K_{j}},
\end{eqnarray}
\noindent where $\eta^2_{b}$ is an efficiency term accounting for
losses due to digitisation of the signal for correlation, $T_{\,i}$ is
the system temperature of antenna $i$ or $j$ respectively [K], and for
example $K_{\,i}$ is the antenna sensitivity (where $K_{\,i} =
(\eta_{i} A_{i})/(2 k)$; $\eta_{i}$ is the antenna efficiency, $A_{i}$
is the geometrical area [m$^2$], and $k$ is the Boltzmann constant),
$\Delta t$ is the integration time per visibility [s], and $\Delta
\nu$ is the bandwidth [Hz]. In order to detect a single source, and
therefore to calibrate or self-calibrate the data, a sensitivity of
5~$\times$~$\Delta S_{\, ij}$ is required within the coherence time
(Walker 1989). A practical measure of the coherence time is the
averaging time at which the scalar- and the vector-averages of the
phases or the amplitudes are significantly different (e.g. the
amplitude vector average is calculated by $\sqrt{<re>^2 + <im>^2}$ and
the amplitude scalar average via $<\sqrt{re^2+im^2}>$, where $<>$
indicate the time averages and $re$ and $im$ are the real and imaginary
component of the visibilities).
The imaging sensitivity $\Delta I$ [Jy] is closely related to the
baseline sensitivity and for an inhomogeneous array can be determined
via:
\begin{eqnarray}
{\Delta I} = \frac{2 k \eta_{b}}{\sqrt{N_{\rm Stokes} \Delta \nu \Delta t}} \,\, \frac{1}{\sqrt{ (\sum \frac{\eta_i A_i}{T_i})^2 - \sum (\frac{\eta_i A_i}{T_i})^2 }},
\end{eqnarray}
\noindent where $N_{\rm Stokes}$ is the number of polarisation
products and $\Delta t$ is now the total on-source time. In case of a homogeneous array the image sensitivity is given by:
${\Delta I} = (\eta_{b} SEFD) / \sqrt{N\,(N - 1) \, N_{\rm Stokes} \Delta \nu \Delta t}$, where N is the number of antennae, SEFD is the system equivalent flux density, which is defined by SEFD = $T_i /
K_i$. For a full discussion on the derivation of the above equations see Walker (1989)
and Wrobel \& Walker (1999).
For a basic estimate of the relative improvement in imaging
sensitivity the following L-band SEFDs are considered: Lovell Telescope
36~Jy, MKII Telescope 350~Jy, and 220~Jy for the Telescope in
Cambridge (these SEFDs are provided in the EVN status table). For all other antennae in the array a SEFD of 350~Jy is
assumed. In this case the image sensitivity of e-MERLIN\ will improve by
a factor of $\sim$1.9 if the Lovell Telescope is used instead of the
MKII Telescope. Furthermore, comparing e-MERLIN\ including the Lovell
Telescope and Goonhilly with e-MERLIN\ and the MKII Telescope shows an even higher
imaging sensitivity of a factor of $\sim$2.1 or equivalently allows a reduction in the
observing time by a factor of $\sim$4.4. An equal sensitivity
improvement is expected for C-band observations, because the antennae
have similar SEFD values. For K-band observations this situation
changes significantly; the effect of adding sensitivity with the
Goonhilly Telescopes is of importance and of the order of $\sim$20\%,
as neither the Lovell nor the Defford Telescopes cannot operate at these
frequencies. Note that these sensitivity improvements are valid for
both continuum and spectral line observations.
In addition, to the imaging sensitivity is the spatial
sensitivity of the additional baselines formed by the antennae in
Goonhilly. The baselines in the triangle
Jodrell-Bank-Cambridge-Goonhilly would build the most sensitive part
of the e-MERLIN\ array and is particular important when the Lovell
Telescope is included to increase the long-baseline sensitivity.
\begin{figure}
\begin{center}
\includegraphics[width=.45\textwidth]{hrk_2a.png}
\includegraphics[width=.45\textwidth]{hrk_2b.png}\\
\includegraphics[width=.45\textwidth]{hrk_2c.png}
\includegraphics[width=.45\textwidth]{hrk_2d.png}
\end{center}
\caption{Simulated amplitude versus UV distance plots for e-MERLIN\
(upper left), e-MERLIN~+~Goonhilly (upper right) and the EVN (lower
left and lower right). An observation of two sources at
60$\,^{\circ}$ declination with equal flux densities of 0.5~Jy and
separated by 9~mas have been simulated with the AIPS tasks {\tt
UVCON} and {\tt UVSUB}. The simulation here presents a processed
bandwidth of 400~MHz which has been split into 8 channels of 50~MHz
each. The plot in the lower left displays the full EVN simulation,
whereas the lower right shows a sub-section of the EVN simulations
in order to highlight the shortest baselines (Effelsberg--Westerbork
baseline covers a range of 1--1.5 mega wavelength). The EVN simulations
are based on the following telescopes: Jodrell Bank (Lovell
Telescope), Effelsberg, Westerbork (phased-up), Onsala, Noto, Torun, Medicina.}
\label{fig2}
\end{figure}
\subsection{e-MERLIN\ and the European VLBI Network (EVN)}
\noindent The combination of e-MERLIN\ and EVN observations provide a
unique view of the nuclear composition of extragalactic
sources. The broad range of baselines within the combined datasets
correspondingly probe a broad range of important linear
size scales in extragalactic sources, from kilo- to (sub-)parsec.
With the anticipated imaging sensitivities of e-MERLIN\ normal
galaxies, starburst galaxies, radio quiet quasars, and hybrid
AGN-starburst galaxies will be observable at high redshifts and with mas
resolution. It is of vital importance to disentangle the
nuclear composition of these galaxies e.g. in order to determine the AGN
contribution to the total radio luminosity (Kl\"ockner et
al. 2009). However there is observational evidence that radio
emission can be missed between the baselines of the EVN and e-MERLIN . For
example, observations of the AGN-starburst system
J123642$+$621331 at redshift z=4.424 (Muxlow et al. 2005, Garrett et
al. 2001) show that the radio flux density does not change from
arcsecond to subarcsecond resolution between the baselines provided by the WSRT (489~$\mu$Jy),
the VLA-A (467~$\mu$Jy), and the MERLIN (472~$\mu$Jy) arrays, but
it clearly changes at the higher EVN resolution (248~$\mu$Jy). The missing-flux effect is
not only seen in continuum observations; it has also been noted for
spectral line emission, e.g. for the OH Megamaser emission in Mrk~273 (Yates et
al. 2000, Kl\"ockner \& Baan 2004).
Currently the only method available to compensate for such an effect is to perform EVN
observations which include the Lovell/MKII and the Cambridge
Telescopes from the e-MERLIN array. Obviously the time available to be allocated to such projects is rather
limited. If such an observation is not
possible, the only option one has is to combine both datasets by scaling the visibilities according to
the Effelsberg - Westerbork baseline (270~km) and the
MERLIN flux. In practice, the scaling factor is determined via the flux
density in the image plane and applied to both datasets before combination (e.g. via the AIPS
task {\tt DBCON}). Such a technique is only reliable if the source
structure is close to a point source. However if the target
source exhibits more complicated structure the reliability of such combined
MERLIN (or in future e-MERLIN) and EVN data sets is questionable.
Figure~\ref{fig2} illustrates the potential dangers of ill-defined
flux density measurement in both arrays. These amplitude versus
UV-distance plots are based on e-MERLIN\ and EVN simulations of two point
sources separated by 9~mas and with a flux density of about 0.5~Jy
each. Such a model would not be resolved by e-MERLIN\ as shown in the top
left-hand plot of Figure~\ref{fig2}, but when Goonhilly is added to
the array the source structure is detected by the additional
baselines. Furthermore, comparing the e-MERLIN\ simulation with those of
the EVN shows that there is hardly any overlap of the visibilities in
UV-distance. This spread in amplitude due to the source structure will
make it impossible to reliably model the datasets. However, including
a station at Goonhilly would allow reliable modeling and
cross-matching of the flux density at 1~M$\lambda$ to 2.5~M$\lambda$
scale sizes. Crucially, the baselines would fill the gap in the EVN
data and therefore provide new structural information\footnote{Note
that once the focal plane array program in Westerbork (APERTIF) is
installed (by $\sim$2012) the C-Band sensitivity of the
Effelsberg-Westerbork baseline drops and roughly equals the
sensitivity of the baseline between the Lovell Telescope and
Goonhilly.}. Once APERTIF is installed at the WSRT it would be of
particular importance if one of the Goonhilly antennae would be included to the
EVN observing session. Such setup would provide an almost uniform
structural sensitivity and would be ideal for the "the world's
leading real-time VLBI array" that is compiled by the EVN telescopes
(Szomoru et al. 2011).
The location of Goonhilly seems to be ideal in order to
produce better snap-shot imaging quality than is currently possible
with e-MERLIN . Furthermore, the baselines formed between Goonhilly and
Jodrell Bank, and Goonhilly and Cambridge (440~km) provide a
unique structural probe at 120~mas, which translates to kilo-parsec
scales at redshift 1. In addition, the new baselines formed by the
Goonhilly station will ensure that data from e-MERLIN\ can at all times
be reliably combined with EVN datasets due to the overlap in baseline
lengths. As discussed previously this will afford a totally unique
view of radio emission from galactic scales (10~km baseline Jodrell Bank to
Pickmere) to the central regions of galaxies.
\section{Impact on Legacy proposals}
\noindent The e-MERLIN\ Legacy programme consists of eleven projects
which spans a wide range of astrophysical themes and will address many
current challenges in astronomy and astrophysics. In particular, the
projects build a coherent set of surveys that addresses key questions
in star/planetary system formation, in radio outflow/jet physics, in
the interplay between accreting black holes and nuclear starbursts, and
galaxy/AGN evolution. The data products generated by the Legacy
programmes will be made available for exploitation by the whole
astronomical community.
Here a general overview of the basic observational
requirements of these programmes is given. The quoted hours are the allocated
hours for the proposals and not the proposed ones. The benefits in imaging
sensitivity of including the Lovell Telescope (LT) in the array is discussed. Furthermore, the enhanced angular
resolution brought by including a telescope at Goonhilly (GH) is
discussed. The e-MERLIN\ Legacy programmes and their
basic requirements are:
{\scriptsize
\begin{itemize}
\item {\bf A}strophysics of {\bf GA}laxy {\bf T}ransformation and {\bf E}volution (PI's: C.~Simpson, I.~Smail):\\
{\bf -} L-Band, targeted imaging (single pointings), 330~hours.\\
{\bf -} The LT is not considered due to FoV limitations, but the increased angular resolution due to GH
enables the study of the radio morphology at $\sim$500~pc scales (an average redshift of 0.5 is assumed).
\item The e-MERLIN\ {\bf C}yg {\bf OB}2 {\bf RA}dio {\bf S}urvey: Massive and Young stars in the Galaxy (PI: R.~Prinja):\\
{\bf -} L-Band, C-Band, several pointings to mosaic, multiple epochs at C-band, 294~hours.\\
{\bf -} Using the LT allows the detection of some 50 O stars, whereas without it only the most massive
stars can be observed, thus biasing the sample substantially. The enhanced angular resolution provided by GH
provides a better base to detect positional changes in order to study the kinematics of the OB association.
\item {\bf MER}LIN {\bf G}alaxy {\bf E}volution Survey (PI's: T.~Muxlow, I.~Smail, I.~McHardy):\\
{\bf -} L-band, C-band, single pointings and field mapping/mosaicing, (Tier 0 and Tier 1) 918~hours.\\
{\bf -} The science goals rely on reliably measuring the structural information of the radio sources and to
disentangle extended star-formation from compact AGN. The imaging sensitivity is crucial to this project therefore
the LT needs to be considered. In addition, the increased angular resolution provided by GH
will help to better explore the nuclear region at linear scales of 660~pc (L-band) and 140~pc (C-band) respectively (assuming redshift 1).
\item {\bf e}-MERLIN {\bf P}ulsar {\bf I}nterferometry Project (PI: W.~Vlemmings, B.~Stappers):\\
{\bf -} L-band, C-band, targeted imaging, visibility fitting, astrometry, multiple epoch, 160~hours.\\
{\bf -} The LT is considered for special cases, there are FoV restrictions in L-band due to in-beam
calibrators. The basic observable for astrometry is the position on the sky and with the enhanced angular resolution
provided by GH such measurements will have an enhanced accuracy of a factor of 2.
\item Feedback Process in Massive Star Formation (PI: M.~Hoare, W.~Vlemmings):\\
{\bf -} C-band; targeted imaging, spectral line imaging of methanol- and hydroxyl, full polarisation, 450~hours.\\
{\bf -} The LT is crucial for this project in order to test various star formation models. The enhanced angular
resolution due to GH provide better constrains on how to relate the radio outflow to the
spectral lines and therefore will provide important clues on the origin of the spectral line emission.
\item Gravitational Lensing and galaxy evolution with e-MERLIN\ (PI: N.~Jackson, S.~Serjeant):\\
{\bf -} L-band , C-band, targeted imaging, 228~hours.\\
{\bf -} Central images of lenses probe the central (10-50~pc) part
of the potential well of lens galaxies at significant cosmological
distance. GH baselines will add sensitivity on spatial scales
well-matched to the problem, allowing detection of fainter central
images; the LT is important in adding sensitivity. Inclusion of
GH/LT will also be important in high-fidelity mapping both of
existing radio lenses and new lenses discovered with Herschel ATLAS,
which is also part of the project. Many of these lenses are small
(<1 arcsecond), and at low declination, and in both cases GH
baselines are a major advantage.
\item {\bf L}egacy {\bf e}-{\bf M}ERLIN {\bf M}ulti-Band {\bf I}maging of {\bf N}earby {\bf G}alaxie{\bf s} (PI: R.~Beswick, I.~McHardy):\\
{\bf -} L-band, C-band, targeted imaging, line data mining considered, 810~hours.\\
{\bf -} The aim is to map out radio star-formation indicators in
nearby galaxies, such as HII regions and supernova remnants (SNR),
and to carry out a census of low luminosity AGN. Without the LT the
investigation is biased to the nearest and brightest sources making
the survey neither complete nor representative. The enhanced angular
resolution provided by GH will help to map out sub-structures at pc
scales to better constrain the nature of such indicators and will
help distinguish unresolved AGN from starburst activity.
\item {\bf L}uminous {\bf I}nfra-{\bf R}ed {\bf G}alaxy {\bf I}nventory (PI: J.~Conway, M. Perez-Torres):\\
{\bf -} L-band, C-band, targeted imaging, line imaging, multiple epoch, full polarisation, 353~hours.\\
{\bf -} The LT is considered necessary to match the sensitivity in polarisation and to detect a substantial rate of RSNe in
the most extreme starburst environments. Due to the compactness of the RSNe the enhanced angular resolution by GH will
provide a better constraint on the nature of the detected radio emission.
\item Morphology and Time Evolution of Thermal Jets Associated with Low Mass Young Stars (PI: L.~Rodriguez):\\
{\bf -} C-band, targeted imaging, multiple epoch, 180~hours.\\
{\bf -} The LT is needed to detect the hollow-core and the optically thin outer regions of the stellar jets and
hence lead to a fundamental understanding on the disk-jet relation. GH will provide a better estimate on the proper motion of the sources.
\item {\bf P}lanet {\bf E}arth {\bf B}uilding {\bf B}locks - a {\bf L}egacy {\bf e}-MERLIN {\bf S}urvey (PI: J. Greaves):\\
{\bf -} C-Band, targeted imaging, 72~hours.\\
{\bf -} Only with the LT is e-MERLIN able to detect the cm dust emission on a few AU scales for the first time.
The improved resolution with GH will better constrain the planet-forming zones and allow dust detections of a few Earth-masses.
\item Resolving Key Questions in Extragalactic Jet Physics (PI: R.~Laing):\\
{\bf -} L-band, C-band, targeted imaging, multi epoch, full polarisation, 375~hours.\\
{\bf -} For a sub-sample the LT is considered, without Lovell the observable sub-sample would
more than halve and hence would not be representative. The increase in angular resolution by GH (80~pc L-Band and 16~pc
for C-band at redshift 0.05) may be too high to investigate the radio sub-structure of the closest sources of the sample, but
for the other targets the additional resolution will help to better investigate the study of radio structure.
\end{itemize}}
\noindent Of the 11 Legacy projects, 9 require the additional sensitivity
provided by the Lovell Telescope. Apart from the great advantage in
imaging sensitivity there are some observing constraints due to the
field of view (FoV) limitation mentioned in the proposals, which precludes the use of
the Lovell Telescope. In particular, for in-beam calibration the smaller FoV of the
Lovell Telescope with respect to that of the MKII Telescope will reduce the
number of suitable calibrators. In addition, there is a trade off
between the mapping speed and the sensitivity, such that the mosaicing
efficiency might be reduced if Lovell Telescope is included in e-MERLIN .
At least 10 out of 11 proposals seem to accrue significant benefit
from the enhanced resolution provided by Goonhilly. Except for a
sub-program in one proposal the Legacy programmes will greatly benefit
from this factor 2 increase in spatial resolution, in particular
providing additional constraints on source structure, better
positional constraints of radio components for multi-epoch
observations, and improved astrometry. In addition to the above points
many breakthroughs in these and other areas of science will come from
the combination of e-MERLIN\ and EVN data. The much improved overlap of
baselines provided by Goonhilly (Figure 2) will be critical for
reliably achieving this combination, particularly for line
observations where multi-frequency synthesis cannot be used to improve
the UV coverage.
Finally, it should be noted that the existence of Goonhilly and the
opportunity to access southern fields (below Dec~30$\,^{\circ}$) in e-MERLIN\ may
mean that several of the Legacy programmes may wish to re-asses their
selection of targets or survey locations to benefit from overlap with
the ESO facilities such as ALMA or VISTA.
\section{Conclusions}
\noindent The addition of an antenna at Goonhilly to the e-MERLIN\ array
provides a near-doubling of the spatial
resolution and an improved UV-plane coverage. The increased N-S extent
also has strong positive imaging implications for equatorial imaging
with the e-MERLIN\ array. The very sensitive 440~km baseline that would be formed
between Goonhilly and the Lovell Telescope would result in
rich enhancements of the e-MERLIN\ Legacy surveys, roughly doubling the
overall imaging sensitivity and tripling the long-baseline
sensitivity. Finally, the Goonhilly station provides a unique overlap in baseline
lengths between e-MERLIN\ and the EVN, allowing robust combination of
data from the two instruments. Such combined data sets are unique and
open up the ability to study radio morphologies from kilo-parsec up to
parsec scales.
|
\section{Introduction}
The total energy $E_{12}$ of two static disjoint objects
may be decomposed as,
\begin{equation}
E_{12} = E_0 + \Delta E_1 + \Delta E_2 + \Delta E_{12},
\label{E12-2}
\end{equation}
where $E_0$ is the energy of the vacuum (medium) without objects,
$\Delta E_1$ and $\Delta E_2$ are (self)-energies required to create
the objects individually in isolation, and $\Delta E_{12}$ is the change
in energy due to their interaction.
The interaction energy $\Delta E_{12}$ is finite for
disjoint objects if it is mediated by an otherwise free quantum field
whose interaction with the objects is described by local potentials.
It is the only contribution to the total energy
that depends on the position and orientation of \emph{both} objects and
determines the forces between them. Casimir found that the
electromagnetic force between two parallel neutral metallic plates does
not vanish~\cite{Casimir:1948dh} and that the associated Casimir energy
$\Delta E_{12}$ may be interpreted as arising from changes in the
zero-point energy due to boundary conditions imposed on the
electromagnetic field by the metallic plates. Zero-point energy
contributions to the energies $\Delta E_i$ of
individual objects in general diverge but the change
$\Delta E_{12}$ due to the presence of two \emph{disjoint} objects is finite.
Reliable extraction of finite Casimir energies for a long time appeared
to be restricted to very special geometries,
like parallelepipeds~\cite{Lukosz:1971,Ambjorn:1981xw,Ambjorn:1981xv},
spheres~\cite{Boyer:1968uf,Milton:1978sf}, and
cylinders~\cite{DeRaad1981229}.
A multiple scattering formulation for computing Casimir energies of
smooth objects was developed by
Balian and Duplantier~\cite{Balian1977300,Balian1978165}
but it relied heavily on idealized boundary conditions. Kenneth and Klich
only recently observed~\cite{Kenneth:2006vr} that $\Delta E_{12}$ may
be computed independent of single-body contributions to the energy
and is always finite for disjoint objects. This two-body interaction
energy is compactly expressed~\cite{Kenneth:2006vr,Emig:2007me} in terms of
the free Green's function, $G_0$, and transition operators, $T_1$, $T_2$,
associated with the individual objects,
\begin{equation}
\Delta E_{12}
= \frac{1}{2} \int_{-\infty}^{\infty}\frac{d\zeta}{2\pi}
\,\text{Tr}\ln\big[ 1- G_0 T_1 G_0 T_2\big]
= \frac{1}{2} \int_{-\infty}^{\infty}\frac{d\zeta}{2\pi}
\,\text{Tr}\ln\big[ 1-\tilde T_1 \tilde T_2\big],
\label{trloggen}
\end{equation}
where we have defined partly amputed transition operators
$\tilde T_i=G_0T_i$, $i=1,2$.
For potential scattering the interaction energy may equivalently be
expressed~\cite{Milton:2007wz} in terms of the potentials
$V_i$ and the corresponding Green's functions $G_i$ satisfying
$G_i V_i=G_0T_i$.
In deriving Eq.~(\ref{trloggen}) one formally subtracts divergent self-energies
and avoids the question of whether these divergences have any physical
significance. One in particular circumvents the issue raised
in~\cite{Brown:1969na,Deutsch:1978sc} of how they should be treated
in the context of gravity, a problem that has so far only been
considered for parallel
plates~\cite{Fulling:2007xa,Milton:2007ar,Shajesh:2007sc}.
Although it does not address such conceptual points,
the irreducible contribution of Eq.~(\ref{trloggen}) suffices to explain
experimental measurements of Casimir forces between
two disjoint objects. Since the interaction energy
for disjoint objects is finite, errors due to numerical or other
approximations to Eq.~(\ref{trloggen}) can be controlled. This has
now been demonstrated by explicit calculations for a number of
geometries and physical situations
\cite{CaveroPelaez:2008tj,CaveroPelaez:2008tk,Maghrebi:2010wp,Maghrebi:2010jx}.
In this article we examine a recently proposed extension of these
ideas to more than two bodies. It was shown in \cite{Schaden:2010wv}
that the irreducible $N$-body part of the total energy remains finite
if the $N$ objects have no {\em common} intersection. We here explicitly
evaluate the irreducible three-body part, $\Delta E_{123}$, of the total energy,
\begin{equation}
E_{123} = E_0 + \Delta E_1 + \Delta E_2 + \Delta E_3
+ \Delta E_{12} + \Delta E_{23} + \Delta E_{31} + \Delta E_{123},
\label{E-123-ser}
\end{equation}
in several cases and verify that $\Delta E_{123}$ remains finite even as
irreducible two-body contributions diverge. The formal
expression for $\Delta E_{123}$ given in \cite{Schaden:2010wv} is
considerably more involved than~Eq.(\ref{trloggen}),
\begin{equation}
\Delta E_{123}
= \frac{1}{2} \int_{-\infty}^{\infty}\frac{d\zeta}{2\pi}
\,\text{Tr}\ln \Big[ 1 + X_{12}
\Big\{ \tilde T_1\tilde T_2\tilde T_3X_{23}
+\tilde T_1\tilde T_3\tilde T_2X_{32}
-\tilde T_1\tilde T_2\tilde T_1\tilde T_3
-\tilde T_1\tilde T_3\tilde T_2\tilde T_3X_{23}
-\tilde T_1\tilde T_2\tilde T_3\tilde T_2X_{32}
\Big\} X_{13} \Big],
\end{equation}
where the $X_{ij}$'s are solutions to the integral equations,
\begin{equation}
\big[ 1-\tilde T_i\tilde T_j\big]X_{ij}=1.
\label{def-Xij}
\end{equation}
We here obtain and evaluate alternate expressions for
irreducible Casimir energies that do not involve a logarithm.
The article is organized as follows. In Section~\ref{Nb-G-fun:s} we derive
Faddeev-like equations for the scattering matrix associated
with $N$-objects and extract their irreducible $N$-body parts.
The associated $N$-body Green's functions are expressed in terms of
one-body transition matrices describing scattering off each
object individually. The procedure is illustrated for $N=2$, and $N=3$,
for which explicit solutions are obtained.
The method is generalizable to higher $N$.
The general solutions of Section~\ref{Nb-G-fun:s}
are used to obtain the Green's functions for two and three
semitransparent plates in Section~\ref{G-fun-semi-Ps:s}.
The irreducible three-body contribution to the Green's function for
three semitransparent plates is found to exactly cancel the
two-body interaction of the outer plates when Dirichlet boundary
conditions are imposed on the central plate.
In Section~\ref{Nb-Cas-En:s} we express the irreducible
$N$-body contribution to the Casimir energy in terms of
the $N$-body part of the transition matrix.
This avoids the computation of the logarithm of an
integral operator, but requires one to solve a set of linear Faddeev-like
integral equations for the $N$-body transition matrix.
In Section~\ref{Cas-en-semiP:s} we use this formalism
to obtain irreducible two--and three-body contributions to the Casimir energy
of two and three semitransparent plates.
The three-body contribution again cancels the two-body
interaction from the outer plates when Dirichlet boundary conditions
are imposed on the central plate.
In Section~\ref{3body-sweak:s},
we specialize to the case when two of the three potentials are weak.
For point-potentials we prove that the irreducible two-body Casimir energy
is always negative whereas the irreducible three-body contribution
is positive. The proof immediately generalizes to any form of the weakly
interacting potentials. We also derive expressions for irreducible
two--and three-body contributions to the Casimir energy when
the weak potentials have translational symmetry and the third potential
represents a Dirichlet plate parallel to the symmetry axis.
Some of the expressions for the irreducible two-body
contributions appear to not have been noted in earlier studies.
We obtain the irreducible three-body contribution to the Casimir energy
in this semiweak approximation, and verify independently that
it is positive and finite.
In Section~\ref{tri-wedge:s} we use these results
to investigate the irreducible three-body
Casimir energy of a weakly interacting wedge placed atop
a Dirichlet plate forming a waveguide of triangular cross-section.
The potentials forming the triangular waveguide overlap
and the irreducible two-body contributions to the vacuum energy diverge.
However, the irreducible three-body Casimir energy
is well defined as long as the supports of the three potentials have
no {\em common} overlap.
The irreducible three-body Casimir energy is minimal (and vanishes)
when the shorter side of the wedge is perpendicular to the Dirichlet plate.
Inspired by the study in \cite{Abalo:2010ah}, we investigate the
dependence of the irreducible three-body Casimir energy
on the cross-sectional area and perimeter of the triangular waveguide.
To emphasize that the finiteness of the irreducible three-body Casimir
energies is not only due to the subtraction of corner divergences,
we, in Section~\ref{par-wedge:s}, consider a waveguide with weakly
interacting sides of parabolic cross-section that touch a Dirichlet plate.
The conclusions for weak triangular-wedges generalize to parabolic-wedges
with only minor changes in interpretation.
The explicit calculations in this article support the general results of
\cite{Schaden:2010wv}. The irreducible three-body contribution to the
Casimir energy in the examples considered here is always positive.
It furthermore is continuous (and in this sense is analytic in the
corresponding parameter) when two of the bodies approach each other
and intersect.
\section{Many-body Green's functions}
\label{Nb-G-fun:s}
The free Green's function of a massless scalar field in Euclidean space-time
satisfies the partial differential equation
\begin{equation}
[-{\bm\nabla}^2 +\zeta^2] G_0({\bf x},{\bf x}^\prime)
= \delta^{(3)}({\bf x}-{\bf x}^\prime),
\label{G0-def}
\end{equation}
where ${\bm\nabla}^2$ is the Laplacian of flat three-dimensional space.
It is related to the corresponding free Green's function
of Minkowski space-time by a Euclidean (Wick) rotation.
The ``one-body'' Green's function, $G_i$, associated with the
time-independent potential, $V_i({\bf x})$,
describing the interaction with the $i$-th object satisfies
\begin{equation}
\left[-{\bm\nabla}^2 +\zeta^2 +V_i({\bf x}) \right]
G_i({\bf x},{\bf x}^\prime) = \delta^{(3)}({\bf x}-{\bf x}^\prime).
\label{Gi-def}
\end{equation}
The two-body Green's function $G_{ij}$ solves a similar equation
with the potential $(V_i+V_j)$ associated with a pair of objects,
$G_{ijk}$ denotes the three-body Green's function to the potential
$(V_i+V_j+V_k)$, etc.
The potentials $V_i({\bf x})$ of this model~\cite{Bordag:1992cm}
are proportional to
$\delta$-functions that simulate the interaction of the scalar field
with classical objects. Infinitely strong $\delta$-function potentials
enforce Dirichlet boundary conditions at the surface of
the objects.
One obtains a formal solution to Eq.~(\ref{Gi-def}) by considering the
differential operator as an integral kernel, and manipulating the
kernels as if they were matrices. To emphasize the correspondence
between integral kernels and matrices, we replace:
$[-{\bm\nabla}_{\bf x}^2 +\zeta^2] \,\delta^{(3)}({\bf x}-{\bf x}^\prime)
\to G_0^{-1}$, $V_i({\bf x})\,\delta^{(3)}({\bf x}-{\bf x}^\prime) \to V_i$,
and $\delta^{(3)}({\bf x}-{\bf x}^\prime) \to 1$.
Using ordinary matrix algebra one obtains the formal solution to
Eq.~(\ref{Gi-def}) in the form
\begin{equation}
G_i=G_0 - G_0T_iG_0,
\label{Gi-gsol}
\end{equation}
where the transition matrix $T_i$ is given by
\begin{equation}
T_i = V_i (1+G_0V_i)^{-1} = (1+V_iG_0)^{-1} V_i
= V_i - V_iG_0V_i + V_iG_0V_iG_0V_i - \ldots.
\label{ti-def}
\end{equation}
The second term in Eq.~(\ref{Gi-gsol}) is interpreted as due to scattering
off the $i$-th object. It corresponds to the integral operator,
\begin{equation}
G_0T_iG_0\to{\textstyle\int} d^3x_1 {\textstyle\int} d^3x_1^\prime
\,G_0({\bf x}-{\bf x}_1) T_i({\bf x}_1,{\bf x}_1^\prime)
G_0({\bf x}_1^\prime-{\bf x}^\prime).
\label{G1=G0-RTR}
\end{equation}
In the following, symbolic equations often are more compactly
written\footnote{This is like setting $G_0=1$.}
in terms of partly amputated operators.
Equations for the physical operators are obtained by replacing
every partly amputated Greens-function $\tilde G_i$, potential $\tilde V_i$,
and scattering matrix $\tilde T_i$, by,
\begin{equation}
\tilde G_i\to G_iG_0^{-1}, \qquad
\tilde V_i\to G_0V_i, \qquad \text{and} \qquad \tilde T_i\to G_0T_i.
\end{equation}
\subsection{Many-body scattering theory}
The partly amputated $N$-body Green's function satisfies the equation
\begin{equation}
\Big[1 +\tilde V_1 +\tilde V_2 +\ldots+\tilde V_N\Big]\tilde G_{1\ldots N} =1.
\label{Ngeqn-def}
\end{equation}
The numbers in the subscript of $\tilde G_{1\ldots N}$ relate to the
respective potentials.
We may treat the sum of potentials in Eq.~(\ref{Ngeqn-def})
as a single potential and thus proceed as for a
single-body. The solution may again be written in the form
\begin{equation}
\tilde G_{1\ldots N}=1-\tilde T_{1\ldots N},
\label{G1-N=M1N}
\end{equation}
where the $N$-body transition matrix $\tilde T_{1\ldots N}$ satisfies the
equation
\begin{equation}
\Big[ 1 + (\tilde V_1+\tilde V_2+\dots +\tilde V_N)\Big] \tilde T_{1\ldots N}
= (\tilde V_1+\tilde V_2+\dots +\tilde V_N).
\label{T1N-deqn}
\end{equation}
The solution to Eq.~(\ref{T1N-deqn}) is an infinite series similar to
the one in Eq.~(\ref{ti-def}), whose terms can be regrouped into
components $\tilde T_{1\ldots N}^{ij}$, that begin with the $i$-th potential
and end with the $j$-th potential. For $N$ potentials we have $N^2$
such components representing transitions from the $i$-th to the $j$-th object.
This decomposition of the $N$-body transition matrix is of the form,
\begin{equation}
\tilde T_{1\ldots N} = \sum_{i=1}^N\sum_{j=1}^N \tilde T_{1\ldots N}^{ij}
=\text{Sum} \big[ \tilde{\bf T}_{1\ldots N} \big],
\label{T1N=T1Nij}
\end{equation}
where the symbol $\text{Sum}[{\bf A}]$ stands for the sum of all
elements of the matrix ${\bf A}$.
The matrix form of the $N$-body transition operator is,
\begin{equation}
\tilde{\bf T}_{1\dots N} = \left(
\begin{array}{cccc}
\tilde T_{1\dots N}^{11} & \tilde T_{1\dots N}^{12}
& \cdots & \tilde T_{1\dots N}^{1N} \\[2mm]
\tilde T_{1\dots N}^{21} & \tilde T_{1\dots N}^{22}
& \cdots & \tilde T_{1\dots N}^{2N} \\[2mm]
\vdots &\vdots &\ddots & \vdots \\[2mm]
\tilde T_{1\dots N}^{N1} & \tilde T_{1\dots N}^{N2}
& \cdots & \tilde T_{1\dots N}^{NN}
\end{array} \right),
\end{equation}
where each component is an integral operator.
Inserting Eq.~(\ref{T1N=T1Nij}), and introducing Kronecker-$\delta$ integral
operators, Eq.~(\ref{T1N-deqn}) is equivalent to the following set
of integral equations
\begin{equation}
\sum_k\Big[ \delta_{ik} +\tilde V_i\Big]\tilde T_{1\ldots N}^{kj}
=\tilde V_i\delta_{ij}.
\end{equation}
In matrix notation this set of equations is
\begin{equation}
\big[ {\bf 1}+ \tilde{\bf V}_\text{diag} + \tilde{\bf\Theta}_{1\ldots N}^V\big]
\cdot \tilde{\bf T}_{1\ldots N} = \tilde{\bf V}_\text{diag},
\label{1vtt=v}
\end{equation}
where we have introduced general matrix symbols
\begin{equation}
{\bm \Theta}_{1\ldots N}^A = \left(
\begin{array}{ccccc}
0 & A_1 & A_1 & \cdots & A_1 \\[1mm]
A_2 & 0 & A_2 & \cdots & A_2 \\[1mm]
A_3 & A_3 & 0 & \cdots & A_3 \\[1mm]
\vdots &\vdots &\vdots &\ddots & \vdots \\[1mm]
A_N & A_N & A_N & \cdots & 0
\end{array} \right),
\qquad \qquad
{\bm A}_\text{diag} = \left(
\begin{array}{ccccc}
A_1 & 0 & 0 & \cdots & 0 \\[1mm]
0 & A_2 & 0 & \cdots & 0 \\[1mm]
0 & 0 & A_3 & \cdots & 0 \\[1mm]
\vdots &\vdots &\vdots &\ddots & \vdots \\[1mm]
0 & 0 & 0 & \cdots & A_N
\end{array} \right).
\end{equation}
Using these definitions with Eq.~(\ref{ti-def}) we write,
\begin{equation}
\big[{\bf 1}+\tilde{\bf V}_\text{diag} \big]
\cdot \tilde{\bf T}_\text{diag} = \tilde{\bf V}_\text{diag}
\qquad \text{and} \qquad
\big[ {\bf 1} + \tilde{\bf V}_\text{diag} \big]
\cdot \tilde{\bm \Theta}_{1\ldots N}^T =\tilde{\bf\Theta}_{1\ldots N}^V,
\end{equation}
and use in Eq.~(\ref{1vtt=v}) to derive
\begin{equation}
\big[{\bf 1} +\tilde{\bm \Theta}_{1\ldots N}^T \big]
\cdot\tilde{\bf T}_{1\dots N}= \tilde{\bf T}_\text{diag}.
\label{fadeqn}
\end{equation}
The set of linear integral equations of Eq.~(\ref{fadeqn}) are often
referred to as Faddeev's equations~\cite{Faddeev:1965a, Merkuriev:1993a}
for nuclear many-body scattering, but apparently have been
known~\cite{francis:1953a, brueckner:1954a}
in the context of ``optical models'' for atomic nuclei since the 1950's
and may have been used in earlier optical studies.
The closely related approach in \cite{Martin:1959a, Puff1961317}
is known as Martin-Schwinger-Puff many-body theory.
Eq.~(\ref{G1-N=M1N}) relates the $N$-body Green's function
$\tilde G_{1,\dots N}$ to the $N$-body transition matrix
$\tilde T_{1,\dots N}$ satisfying Eq.~(\ref{T1N-deqn}).
Faddeev's equations of Eq.~(\ref{fadeqn}) reduce the problem
of solving Eq.~(\ref{T1N-deqn}) for the $N$-body transition matrix
to that of inverting
$\big[{\bf 1}+\tilde{\bf \Theta}_{1\ldots N}^T \big]$
by solving a set of $N$ linear integral equations.
Remarkably, ${\bf \Theta}_{1\ldots N}^T$ depends only on
single-body transition operators.
The norm of ${\bf \Theta}_{1\ldots N}^T$ is less than unity
(because the norm of single-body transition matrices is)
and Faddeev's equations can, at least in principle, be solved by
(numerical) iteration~\cite{Merkuriev:1993a}.
\subsection{$N=2$: Two-body interaction}
Using Eq.~(\ref{Ngeqn-def}) the Green's function equation for $N=2$
has solution given by Eq.~(\ref{G1-N=M1N}), where
the transition matrix is obtained by inverting the Faddeev's
equation in Eq.~(\ref{fadeqn}) to yield
\begin{equation}
\tilde{\bf T}_{12} = \big[ {\bf 1}+ \tilde{\bf \Theta}_{12}^T\big]^{-1}
\cdot \tilde{\bf T}_\text{diag}
= \left[ \begin{array}{cc} X_{12} & 0 \\ 0 & X_{21} \end{array} \right]
\left[ \begin{array}{cc} \tilde T_1 & -\tilde T_1\tilde T_2 \\
-\tilde T_2\tilde T_1 & \tilde T_2 \end{array}\right].
\label{12-T12}
\end{equation}
The integral operators $X_{ij}$ in Eq.~(\ref{12-T12})
satisfy Eq.~(\ref{def-Xij}).
Summing the components of $\tilde{\bf T}_{12}$
we obtain the total transition matrix as,
\begin{equation}
\tilde T_{12} = \text{Sum}\big[ \tilde{\bf T}_{12} \big]
= \big[ 1-X_{12}\tilde G_1\big] + \big[ 1-X_{21}\tilde G_2\big].
\end{equation}
The total two-body transition matrix $\tilde {\bf T}_{12}$ can be
decomposed into its irreducible one--and two-body parts,
\begin{equation}
\tilde {\bf T}_{12} = \tilde {\bf T}_1 + \tilde {\bf T}_2
+ \Delta \tilde {\bf T}_{12},
\label{T12=T1T2DT12}
\end{equation}
with
\begin{equation}
\tilde {\bf T}_{1}
= \left[ \begin{array}{cc} \tilde T_1 & 0 \\ 0 & 0 \end{array}\right],
\qquad
\tilde {\bf T}_{2}
= \left[ \begin{array}{cc} 0 & 0 \\ 0 & \tilde T_2 \end{array}\right],
\qquad
\Delta \tilde{\bf T}_{12} =
\left[ \begin{array}{cc} X_{12} & 0 \\ 0 & X_{21} \end{array} \right]
\left[ \begin{array}{cc}
\tilde T_1\tilde T_2\tilde T_1 & -\tilde T_1\tilde T_2 \\
-\tilde T_2\tilde T_1 & \tilde T_2\tilde T_1\tilde T_2
\end{array}\right].
\label{DT12=expl}
\end{equation}
The irreducible two-body transition matrix $\Delta \tilde{\bf T}_{12}$
includes all contribution with scattering off \emph{both} potentials.
Eq.~(\ref{T12=T1T2DT12}) and the definition in Eq.~(\ref{G1-N=M1N})
imply the following decomposition of the partly amputated
two-body Green's function
\begin{equation}
\tilde G_{12}=1-\tilde T_1-\tilde T_2-\Delta \tilde T_{12},
\label{G12-casc}
\end{equation}
where $\Delta \tilde T_{12}=\text{Sum}\big[\Delta \tilde{\bf T}_{12}\big]$.
Summing the four independent two-body transitions of Eq.~(\ref{DT12=expl}),
the irreducible two-body transition operator in terms of the
single-body transition matrices $\tilde T_1$ and $\tilde T_2$ is,
\begin{equation}
\Delta \tilde T_{12} = \text{Sum}\big[\Delta \tilde{\bf T}_{12}\big]
= (1-X_{12})\tilde G_1 + (1-X_{21})\tilde G_2.
\label{DT12full}
\end{equation}
\subsection{$N=3$: Three-body interaction}
One proceeds similarly for three bodies. In this case the formal solution
to the Faddeev's equation, of Eq.~(\ref{fadeqn}), is
\begin{equation}
\tilde{\bf T}_{123} =
\left[ \begin{array}{ccc}
X_{1[23]} & 0 & 0 \\[1mm] 0 & X_{2[31]} & 0 \\[1mm] 0 & 0 & X_{3[12]}
\end{array}\right]
\left[ \begin{array}{ccc}
\tilde T_1 & -\tilde T_1\tilde G_3\tilde T_2X_{32}
& -\tilde T_1\tilde G_2\tilde T_3X_{23} \\[1mm]
-\tilde T_2\tilde G_3\tilde T_1X_{31} & \tilde T_2
& -\tilde T_2\tilde G_1\tilde T_3X_{13} \\[1mm]
-\tilde T_3\tilde G_2\tilde T_1X_{21}
& -\tilde T_3\tilde G_1\tilde T_2X_{12} & \tilde T_3
\end{array}\right],
\label{123-rtr}
\end{equation}
where the $\tilde G_i$'s are related to $\tilde T_i$'s by Eq.~(\ref{Gi-gsol})
and the two-body effective Green's functions, $X_{ij}$, ($i\neq j$),
solve Eq.~(\ref{def-Xij}).
The three-body effective Green's functions, $X_{i[jk]}$,
($i\neq j\neq k$), satisfy the equation
\begin{equation}
X_{i[jk]} \big[ 1 -\tilde T_i\tilde T_{jk} \big]
=X_{i[jk]} \big[ 1 -\tilde T_i\tilde G_j\tilde T_k X_{jk}
-\tilde T_i\tilde G_k\tilde T_j X_{kj} \big] = 1.
\label{Xijk-def}
\end{equation}
The total transition matrix in this case is
\begin{equation}
\tilde T_{123} = \text{Sum}\big[ \tilde{\bf T}_{123} \big]
= \big[ 1 - X_{1[23]}\tilde G_1\big]
+ \big[ 1 - X_{2[31]}\tilde G_2\big] + \big[ 1 - X_{3[12]}\tilde G_3\big].
\end{equation}
The transition matrix may again be decomposed into irreducible parts
\begin{equation}
\tilde{\bf T}_{123} = \tilde{\bf T}_1 + \tilde{\bf T}_2 + \tilde{\bf T}_3
+ \Delta \tilde{\bf T}_{12}+\Delta\tilde{\bf T}_{23}+\Delta\tilde{\bf T}_{31}
+ \Delta \tilde{\bf T}_{123},
\label{T123=T1..DT123}
\end{equation}
where
\begin{equation}
\Delta {\bf\tilde T}_{12}+\Delta {\bf\tilde T}_{23}+\Delta {\bf\tilde T}_{31}
=\left[ \begin{array}{ccc}
X_{12}\tilde T_1\tilde T_2\tilde T_1
+X_{13}\tilde T_1\tilde T_3\tilde T_1&
-X_{12}\tilde T_1\tilde T_2 &-X_{13}\tilde T_1\tilde T_3 \\[1mm]
-X_{21}\tilde T_2\tilde T_1 & X_{21}\tilde T_2\tilde T_1\tilde T_2
+X_{23}\tilde T_2\tilde T_3\tilde T_2 &-X_{23}\tilde T_2\tilde T_3 \\[1mm]
-X_{31}\tilde T_3\tilde T_1
&-X_{32}\tilde T_3\tilde T_2 & X_{31}\tilde T_3\tilde T_1\tilde T_3
+X_{32}\tilde T_3\tilde T_2\tilde T_3
\end{array}\right]
\end{equation}
is obtained using the $N=2$ expressions of Eq.~(\ref{DT12=expl}).
The new irreducible three-body part is,
\begin{equation}
\Delta \tilde{\bf T}_{123} =
\left[ \begin{array}{ccc}
(1 -X_{12} -X_{13}+X_{1[23]})\tilde T_1
& [\tilde T_1X_{21} -X_{1[23]}\tilde T_1\tilde G_3X_{23}]\tilde T_2
& [\tilde T_1X_{31} -X_{1[23]}\tilde T_1\tilde G_2X_{32}]\tilde T_3 \\[1mm]
[\tilde T_2X_{12} -X_{2[31]}\tilde T_2\tilde G_3X_{13}]\tilde T_1
& (1-X_{23}-X_{21}+X_{2[31]})\tilde T_2
& [\tilde T_2X_{32} -X_{2[31]}\tilde T_2\tilde G_1X_{31}]\tilde T_3 \\[1mm]
[\tilde T_3X_{13} -X_{3[12]}\tilde T_3\tilde G_2X_{12}]\tilde T_1
& [\tilde T_3X_{23} -X_{3[12]}\tilde T_3\tilde G_1X_{21}]\tilde T_2
& (1-X_{31}-X_{32}+X_{3[12]})\tilde T_3
\end{array}\right].
\label{d123-rtr}
\end{equation}
The decomposition of Eq.~(\ref{T123=T1..DT123}) carries over to
the decomposition of the Green's function
\begin{equation}
\tilde G_{123}
=1 - \tilde T_1 - \tilde T_2 - \tilde T_3
- \Delta \tilde T_{12} - \Delta \tilde T_{23} - \Delta \tilde T_{31}
- \Delta \tilde T_{123}.
\label{G123-casc}
\end{equation}
Summing the nine independent three-body transitions in
Eq.~(\ref{d123-rtr}) we find that,
\begin{equation}
\Delta \tilde T_{123} = \text{Sum}\big[\Delta \tilde{\bf T}_{123}\big]
= -(1-X_{12}-X_{13}+X_{1[23]})\tilde G_1
-(1-X_{23}-X_{21}+X_{2[31]})\tilde G_2
-(1-X_{31}-X_{32}+X_{3[12]})\tilde G_3.
\end{equation}
Although not quite as obvious as for two-body scattering,
closer inspection reveals that each component of $\Delta \tilde{\bf T}_{123}$
indeed involves scattering off all three bodies. A similar procedure
can be used to obtain scattering matrices and their irreducible parts
for more than three bodies.
\section{Green's functions for parallel semitransparent $\delta$-plates}
\label{G-fun-semi-Ps:s}
We now apply this formalism to derive the Green's functions
for parallel semitransparent plates of infinite extent described by
$\delta$-function potentials
\begin{equation}
V_i({\bf x}) = \lambda_i \delta(z-a_i),
\label{Vi-def}
\end{equation}
where $a_i$ specifies the position of the $i$-th plate on the $z$-axis, and
$\lambda_i>0$ is the coupling parameter. In the limit
$\lambda_i\to\infty$ the potential of Eq.~(\ref{Vi-def})
simulates a plate with Dirichlet boundary conditions.
The translation symmetry in the $x$-$y$ plane can be exploited and
Eq.~(\ref{Gi-def}) written in terms of the dimensionally reduced
Green's function, $g_i(z,z^\prime)$, defined by
\begin{equation}
G_i({\bf x},{\bf x}^\prime;\zeta) = \int\frac{d^2k}{(2\pi)^2}
\,e^{i{\bf k}_\perp\cdot({\bf x}-{\bf x}^\prime)_\perp} g_i(z,z^\prime;\kappa),
\label{Gi=Igi}
\end{equation}
where ${\bf x}_\perp$ is the component of ${\bf x}$ in the
$x$-$y$ plane. ${\bf k}_\perp$ is the corresponding Fourier component,
and $\kappa^2=\zeta^2+{\bf k}^2_\perp$, ${\bf k}_\perp^2=k_x^2+k_y^2$.
Since the potentials of Eq.~(\ref{Vi-def}) do not depend on the
transverse dimensions, the Green's functions $G_{1\dots N}$
for $N$ parallel plates also correspond to dimensionally
reduced $g_{1\dots N}$.
\subsection{$N=1$: Green's function for a single semitransparent plate}
When substituted in Eq.~(\ref{Gi-def}), Eq.~(\ref{Gi=Igi})
implies that $g_i(z,z^\prime)$ solves a one-dimensional ordinary
inhomogeneous second order differential equation with a $\delta$-function
potential that can be solved explicitly, to obtain,
\begin{equation}
g_i(z,z^\prime) = \frac{1}{2\kappa}\,e^{-\kappa|z-z^\prime|}
-\frac{\bar{\lambda}_i}{1+\bar{\lambda}_i}
\frac{1}{(2\kappa)^2}
\,e^{-\kappa|z-a_i|}e^{-\kappa|z^\prime-a_i|},
\label{g1-exsol}
\end{equation}
where $\bar{\lambda}_i=\lambda_i/2\kappa$,
and $\kappa$ was defined after Eq.~(\ref{Gi=Igi}).
We also arrive at this solution by starting from Eq.~(\ref{Gi-gsol}),
which for the dimensionally reduced Green's function reads
\begin{equation}
g_i(z,z^\prime) = g_0(z-z^\prime) - r_i(z) t_i r_i(z^\prime),
\label{g1=g0t1}
\end{equation}
where $g_0(z,z^\prime)$ is the dimensionally reduced free Green's function,
and $r_i(z)=g_0(z-a_i)$.
Eq.~(\ref{G0-def}) implies that
\begin{equation}
g_0(z,z^\prime) = \frac{1}{2\kappa}\,e^{-\kappa|z-z^\prime|},
\quad r_i(z) = \frac{1}{2\kappa}\,e^{-\kappa|z-a_i|}.
\end{equation}
It will be convenient to define,
\begin{equation}
\bar{r}_i(z)=2\kappa\, r_i(z)=e^{-\kappa |z-a_i|},
\quad \text{and} \quad
\bar{r}_{ij}=e^{-\kappa |a_i-a_j|}=e^{-\kappa a_{ij}},
\end{equation}
where for notational convenience we have defined $a_{ij}=|a_i-a_j|$.
The dimensionally reduced transition matrix in Eq.~(\ref{g1=g0t1})
is found by summing the series in Eq.~(\ref{ti-def}).
Translational invariance in transverse directions and the
$\delta$-function potential render all integrals trivial and the
series can be re-summed to give
\begin{equation}
t_i(z,z^\prime) = 2\kappa\, \bar{t}_i
\,\delta(z-a_i)\delta(z^\prime-a_i),
\qquad \bar{t}_i= \frac{\bar{\lambda}_i}{1+\bar{\lambda}_i}.
\label{g1-exp}
\end{equation}
In the Dirichlet limit ($\lambda_i\to\infty$) the transition matrix
simplifies further to $\bar{t}_i^D=1$. Inserting the dimensionally
reduced transition matrix of Eq.~(\ref{g1-exp}) in Eq.~(\ref{g1=g0t1})
reproduces the explicit solution of Eq.~(\ref{g1-exsol}).
\subsection{$N=2$:
Green's function for parallel semitransparent plates}
The previous procedure is readily extended to compute the Green's
function of $N$ semitransparent plates located at
$z=a_i$, $i=1,2,\ldots,N$, and described by potentials of the form
given by Eq.~(\ref{Vi-def}) with associated strengths $\lambda_i$.
Generalization of Eq.~(\ref{g1=g0t1}) in particular gives the relation
\begin{equation}
g_{1\ldots N}(z,z^\prime) = g_0(z-z^\prime)
- {\bf r}(z)^T \cdot {\bf t}_{1\ldots N} \cdot {\bf r}(z^\prime),
\label{12-rtr}
\end{equation}
between the dimensionally reduced Green's function $g_{1\ldots N}(z,z^\prime)$
and the corresponding components of the dimensionally reduced
transition matrix. The vector ${\bf r}(z)$ constructed out of the
dimensionally reduced free Green's function originating or ending
at a plate is given by
\begin{equation}
{\bf r}(z)^T = \left[ r_1(z), r_2(z),\, \ldots\,, r_N(z) \right]
= \frac{1}{2\kappa} \left[
e^{-\kappa |z-a_1|}, e^{-\kappa |z-a_2|},\,\ldots\,,e^{-\kappa |z-a_N|}\right].
\label{vecrz}
\end{equation}
An advantage of this approach is that the Faddeev integral equations,
Eq.~(\ref{fadeqn}), collapse to algebraic equations for the dimensionally
reduced transition matrix ${\bf t}_{1\ldots N}$ due to the translational
symmetry and the $\delta$-function potentials.
The transition matrix decouples from the ${\bf r}$-vector, which leads
to considerable simplification in the evaluation of the Green's function. A
similar simplification occurs for concentric cylinders and concentric spheres.
The dimensionally reduced two-body transition matrix can be read out from
Eq.~(\ref{12-T12}) once the corresponding $X_{ij}$ has been evaluated.
With the single-body transition matrices of Eq.~(\ref{g1-exp}) all integrals
evaluate trivially and the solution of Eq.~(\ref{def-Xij}) for $X_{ij}$ is
\begin{equation}
X_{ij}=X_{ji}=\frac{1}{\Delta_{ij}},
\qquad \Delta_{ij} = 1-\bar t_i\bar r_{ij}\bar t_j\bar r_{ji}
= 1 - \bar{t}_i\bar{t}_j\,e^{-2\kappa a_{ij}}.
\label{Xij-sol}
\end{equation}
Using Eq.~(\ref{12-T12}), the dimensionally reduced transition matrix
for two plates is,
\begin{equation}
{\bf t}_{ij} = \frac{2\kappa}{\Delta_{ij}}
\left[ \begin{array}{cc}
\bar{t}_i & -\bar{t}_i\bar{r}_{ij}\bar{t}_j \\
-\bar{t}_j\bar{r}_{ji}\bar{t}_i & \bar{t}_j
\end{array}\right].
\label{t12i2p}
\end{equation}
Eq.~(\ref{t12i2p}) inserted in Eq.~(\ref{12-rtr})
gives the Green's function for two semitransparent parallel plates.
In the Dirichlet limit, $\bar{t}_i\to\bar{t}_i^D=1$,
we have $\Delta_{ij}\to\Delta_{ij}^D= (1-e^{-2\kappa a_{ij}})$,
and the transition matrix for two Dirichlet plates simplifies to,
\begin{equation}
{\bf t}_{ij}^D = \frac{\kappa}{\sinh\kappa a_{ij}}
\left[ \begin{array}{cc} e^{\kappa a_{ij}} & -1 \\ -1 & e^{\kappa a_{ij}}
\end{array}\right].
\end{equation}
From Eq.~(\ref{T12=T1T2DT12}) we similarly obtain the irreducible two-body
part of the dimensionally reduced transition matrix as
\begin{equation}
\Delta {\bf t}_{ij} = -2\kappa \left[ \begin{array}{cc}
\bar{t}_i \left\{1-\frac{1}{\Delta_{ij}}\right\} \hspace{5mm}
& \bar{t}_i\frac{\bar{r}_{ij}}{\Delta_{ij}}\bar{t}_j \\[2mm]
\bar{t}_j\frac{\bar{r}_{ji}}{\Delta_{ji}}\bar{t}_i
& \bar{t}_j \left\{1-\frac{1}{\Delta_{ij}}\right\}
\end{array}\right]
= \frac{2\kappa}{\Delta_{ij}} \left[ \begin{array}{cc}
\bar{t}_i \bar{r}_{ij} \bar{t}_j \bar{r}_{ji} \bar{t}_i
& -\bar{t}_i\bar{r}_{ij}\bar{t}_j \\[2mm]
-\bar{t}_j\bar{r}_{ji}\bar{t}_i
& \bar{t}_j \bar{r}_{ji} \bar{t}_i \bar{r}_{ij} \bar{t}_j
\end{array}\right],
\label{Dt12=sol}
\end{equation}
which in the Dirichlet limit simplifies to
\begin{equation}
\Delta {\bf t}_{ij}^D = \frac{\kappa}{\sinh\kappa a_{ij}}
\left[ \begin{array}{cc} e^{-\kappa a_{ij}} & -1 \\ -1 & e^{-\kappa a_{ij}}
\end{array}\right].
\label{t12D=exp}
\end{equation}
The two-plate Green's function has been obtained
previously~\cite{CaveroPelaez:2008tj} in a more direct manner.
We reproduced it using the multiple scattering method because
this approach readily generalizes to more than two plates.
\subsection{$N=3$:
Green's function for three parallel semitransparent plates}
\begin{center}
\begin{figure}
\includegraphics[scale=1.0]{figures/3Pplates.eps}
\caption{Three parallel plates. Plates $i$ and $k$ are separated by
distances $a_{ij}$ and $a_{jk}$ from the center plate $j$.}
\label{3paraPl}
\end{figure}
\end{center}
The three semitransparent plates $i$, $j$, and $k$, of infinite extent
and parallel to the $xy$-plane are described by potentials of the form
given in Eq.~(\ref{Vi-def}). Without loss of generality we assume that
$a_i<a_j<a_k$ (see FIG. \ref{3paraPl}). In the previously introduced
notation this implies that $a_{ij}+a_{jk}=a_{ik}$.
The vector ${\bf r}(z)$ in Eq.~(\ref{vecrz}) now has three components.
The dimensionally reduced three-body transition matrix is obtained by
solving Eq.~(\ref{Xijk-def}) for $X_{i[jk]}$ using the $X_{ij}$
of Eq.~(\ref{Xij-sol}). For three semitransparent parallel plates
one finds that,
\begin{equation}
X_{i[jk]}X_{jk} = \frac{1}{\Delta_{ijk}},
\qquad
\Delta_{ijk} = 1 -\bar{t}_i\bar{r}_{ij}\bar{t}_j\bar{r}_{ji}
-\bar{t}_j\bar{r}_{jk}\bar{t}_k\bar{r}_{kj}
-\bar{t}_k\bar{r}_{ki}\bar{t}_i\bar{r}_{ik}
+2\bar{t}_i\bar{r}_{ij}\bar{t}_j\bar{r}_{jk}\bar{t}_k\bar{r}_{ki}.
\label{Del123-def}
\end{equation}
Using Eq.~(\ref{Del123-def}) in Eq.~(\ref{123-rtr}) we obtain
\begin{equation}
{\bf t}_{ijk} = \frac{2\kappa}{\Delta_{ijk}}
\left[ \begin{array}{ccc}
\bar{t}_i\Delta_{jk}
&-\bar{t}_i\bar r_{ij[k]}\bar{t}_j &-\bar{t}_i\bar r_{ik[j]}\bar{t}_k \\[2mm]
-\bar{t}_j\bar r_{ji[k]}\bar{t}_i & \bar{t}_j\Delta_{ki}
&-\bar{t}_j\bar r_{jk[i]}\bar{t}_k \\[2mm]
-\bar{t}_k\bar r_{ki[j]}\bar{t}_i &-\bar{t}_k\bar r_{kj[i]}\bar{t}_j
&\bar{t}_k\Delta_{ij}
\end{array}\right],
\label{t123=expr}
\end{equation}
where
\begin{equation}
\bar r_{ij[k]} =\bar r_{ij}-\bar r_{ik} \bar{t}_k \bar r_{kj}.
\end{equation}
It is apparent from the solutions for the $N=2$ and $N=3$ case that
terms contributing to the transition matrix in the multiple scattering
expansion depend exponentially on the length of the path of propagation.
Expanding the inverse determinants $\Delta_{ij}$ and $\Delta_{ijk}$
gives all paths contributing to the scattering operator.
This is particularly transparent in our essentially 1-dimensional example.
From Eq.~(\ref{d123-rtr}) the dimensionally reduced, irreducible,
three-body transition matrix is similarly evaluated as,
\begin{equation}
\Delta {\bf t}_{ijk} = 2\kappa \left[ \begin{array}{ccc}
\bar{t}_i \left\{1 -\frac{1}{\Delta_{ij}} -\frac{1}{\Delta_{ik}}
+\frac{\Delta_{jk}}{\Delta_{ijk}} \right\}
& \bar{t}_i \left\{ \frac{\bar r_{ij}}{\Delta_{ij}}
-\frac{\bar r_{ij[k]}}{\Delta_{ijk}} \right\} \bar{t}_j
& \bar{t}_i \left\{ \frac{\bar r_{ik}}{\Delta_{ik}}
-\frac{\bar r_{ik[j]}}{\Delta_{ijk}} \right\} \bar{t}_k
\\[3mm]
\bar{t}_j \left\{ \frac{\bar r_{ji}}{\Delta_{ji}}
-\frac{\bar r_{ji[k]}}{\Delta_{ijk}} \right\} \bar{t}_i
&\bar{t}_j \left\{1 -\frac{1}{\Delta_{ji}} -\frac{1}{\Delta_{jk}}
+\frac{\Delta_{ik}}{\Delta_{ijk}} \right\}
& \bar{t}_j \left\{ \frac{\bar r_{jk}}{\Delta_{jk}}
-\frac{\bar r_{jk[i]}}{\Delta_{ijk}} \right\} \bar{t}_k
\\[3mm]
\bar{t}_k \left\{ \frac{\bar r_{ki}}{\Delta_{ki}}
-\frac{\bar r_{ki[j]}}{\Delta_{ijk}} \right\} \bar{t}_i
& \bar{t}_k \left\{ \frac{\bar r_{kj}}{\Delta_{kj}}
-\frac{\bar r_{kj[i]}}{\Delta_{ijk}} \right\} \bar{t}_j
& \bar{t}_k \left\{1 -\frac{1}{\Delta_{ki}} -\frac{1}{\Delta_{kj}}
+\frac{\Delta_{ij}}{\Delta_{ijk}} \right\}
\end{array}\right].
\label{Delt123-3psol}
\end{equation}
It is interesting to consider the situation when Dirichlet boundary
conditions hold on the $j$-th plate {\em between} the other two plates.
Taking the limit $\bar{t}_j\to \bar{t}_j^D=1$, the
determinant for three parallel plates is found to factorize
into a product of two-body determinants,
\begin{equation}
\Delta_{i\infty k}= \Delta_{i\infty} \Delta_{k\infty}
=(1-\bar{t}_i\,e^{-2\kappa a_{ij}}) (1-\bar{t}_k\,e^{-2\kappa a_{jk}}),
\label{det-fact}
\end{equation}
where replacing the subscript of a plate by $\infty$ denotes
Dirichlet boundary conditions on that plate. In this situation,
$\bar r_{ik[\infty]}=(1-\bar{t}_j^D)\bar r_{ik}=0$,
$\bar r_{i\infty [k]}=e^{\kappa a_{ij}}\Delta_{k\infty}$,
$\bar r_{k\infty [i]}=e^{\kappa a_{jk}}\Delta_{i\infty}$,
and Eq.~(\ref{t123=expr}) simplifies to
\begin{equation}
{\bf t}_{i\infty k} = 2\kappa
\left[ \begin{array}{ccc}
\bar{t}_i \frac{1}{\Delta_{i\infty}}
&-\bar{t}_i \frac{\bar r_{ij}}{\Delta_{i\infty}} &0\\[2mm]
-\frac{\bar r_{ji}}{\Delta_{\infty i}} \bar{t}_i \hspace{2mm}
& \frac{1}{\Delta_{i\infty}} + \frac{1}{\Delta_{k\infty}} -1 \hspace{2mm}
&-\frac{\bar r_{jk}}{\Delta_{k\infty}} \bar{t}_k \\[2mm]
0&-\bar{t}_k \frac{\bar r_{kj}}{\Delta_{k\infty}}
& \bar{t}_k \frac{1}{\Delta_{k\infty}}
\end{array}\right].
\end{equation}
This leads to the observation that
\begin{equation}
{\bf t}_{i\infty k}
={\bf t}_{i\infty} +{\bf t}_{k\infty}-{\bf t}_j^D
= {\bf t}_i + {\bf t}_j^D + {\bf t}_k
+ \Delta {\bf t}_{i\infty} + \Delta {\bf t}_{k\infty}.
\label{tiDk=allts}
\end{equation}
Comparing Eq.~(\ref{tiDk=allts})
with the decomposition of the three-body transition matrix
into irreducible one--and two-body parts in Eq.~(\ref{T123=T1..DT123})
this implies
\begin{equation}
\Delta {\bf t}_{i\infty k} +\Delta {\bf t}_{ik} = 0,
\label{etijk=2tik}
\end{equation}
which confirms the notion that modes in the two half-spaces on either side
of a Dirichlet plate are independent and that correlations between them
must vanish. The irreducible three-body correlations in this limit therefore
must cancel irreducible two-body correlations between objects on
opposite sides of the plate. Taking the Dirichlet limit on the central plate
in Eq.~(\ref{Delt123-3psol}) this is verified explicitly,
\begin{equation}
\Delta {\bf t}_{i\infty k} = 2\kappa \left[ \begin{array}{ccc}
\bar{t}_i \left\{1 -\frac{1}{\Delta_{ik}} \right\} & 0
& \bar{t}_i \frac{\bar r_{ik}}{\Delta_{ik}} \bar{t}_k \\[3mm]
0 & 0 & 0 \\[3mm]
\bar{t}_k \frac{\bar r_{ki}}{\Delta_{ki}} \bar{t}_i & 0
& \bar{t}_k \left\{1 -\frac{1}{\Delta_{ki}} \right\}
\end{array}\right] = -\Delta {\bf t}_{ik},
\label{tiDk=-tik}
\end{equation}
where we have used Eq.~(\ref{Dt12=sol}).
Let us finally consider the case when Dirichlet boundary conditions
are imposed on all three plates. The three-body determinant again
factorizes,
$\Delta_{ijk}^D = \Delta_{ij}^D \Delta_{jk}^D
=(1-e^{-2\kappa a_{ij}}) (1-e^{-2\kappa a_{jk}})$, and
\begin{equation}
{\bf t}_{ijk}^D =
2\kappa\left[ \begin{array}{ccc}
\dfrac{e^{\kappa a_{ij}}}{2\sinh\kappa a_{ij}}
& -\dfrac{1}{2\sinh\kappa a_{ij}} & 0 \\[3mm]
-\dfrac{1}{2\sinh\kappa a_{ij}} \hspace{4mm}
& \dfrac{e^{\kappa a_{ij}}}{2\sinh\kappa a_{ij}}
+ \dfrac{e^{\kappa a_{jk}}}{2\sinh\kappa a_{jk}} -1 \hspace{4mm}
& -\dfrac{1}{2\sinh\kappa a_{jk}} \\[3mm]
0 & -\dfrac{1}{2\sinh\kappa a_{jk}}
& \dfrac{e^{\kappa a_{jk}}}{2\sinh\kappa a_{jk}}
\end{array}\right]
= {\bf t}_i^D + {\bf t}_j^D + {\bf t}_k^D
+ \Delta {\bf t}_{ij}^D + \Delta {\bf t}_{jk}^D,
\end{equation}
in the limit of three Dirichlet plates.
This implies $\Delta {\bf t}_{ijk}^D +\Delta {\bf t}_{ik}^D = 0$,
and is explicitly verified by Eq.~(\ref{Delt123-3psol})
or Eq.~(\ref{tiDk=-tik}),
\begin{equation}
\Delta {\bf t}_{ijk}^D = -\frac{\kappa}{\sinh\kappa a_{ik}}
\left[ \begin{array}{ccc} e^{-\kappa a_{ik}} & 0 & -1 \\
0 & 0 & 0 \\
-1 & 0 & e^{-\kappa a_{ik}}
\end{array}\right] = -\Delta {\bf t}_{ik}^D,
\end{equation}
using Eq.~(\ref{t12D=exp}).
\section{Many-body Casimir energies}
\label{Nb-Cas-En:s}
Casimir energies are finite parts of the vacuum energy that describe
its dependence on configurations of macroscopic objects. The interaction
of classical objects with quantized fields at low energies can be
described by background potentials. It is
known~\cite{Deutsch:1978sc, kirsten2002spectral, Fulling:1989nb, Bordag:PRD70.045003, Schaden:2010wv}
that such a semiclassical description for the interaction with a
quantized field suffers of (local) ultraviolet divergences. A proper
treatment of the interaction at high energies requires modeling of
the quantum fluctuations associated with the objects. One fortunately
sometimes is able to isolate parts of the vacuum energy that depend only on
global changes of the system and can be reliably computed in
semiclassical approximation. In the following we systematically
determine irreducible parts of the vacuum energy for a
given number of classical objects. These irreducible $N$-body Casimir
energies diverge only if all $N$ potentials describing the classical
objects have a region of common support~\cite{Schaden:2010wv}.
Let $E_0$ be the (infinite) vacuum energy associated with zero-point
fluctuations of a massless scalar field in the absence of all
background potentials, $V_i({\bf x})$. The change in vacuum energy
in the presence of $N$ objects associated with the background potential,
$V=\sum_i V_i$, can be derived using field theory techniques,
for example in \cite{Bordag:1992cm, CaveroPelaez:2008tj}, to be
\begin{equation}
E_{1\ldots N}-E_0 = - \frac{1}{2} \int_{-\infty}^{\infty}\frac{d\zeta}{2\pi}
\,2\zeta^2\, \text{Tr}\, (G_{1\ldots N}-G_0)
= - \frac{1}{2} \int_{-\infty}^{\infty}\frac{d\zeta}{2\pi}
\,\text{Tr}\ln \tilde G_{1\ldots N}
= - \frac{1}{2} \int_{-\infty}^{\infty}\frac{d\zeta}{2\pi}
\,\text{Tr}\ln (1-\tilde T_{1\ldots N}).
\label{TrG-for}
\end{equation}
These expressions have recently been dubbed the Trace-G-formula and
Trace-Log-G-formula respectively. For frequency independent potentials,
the relation between them is established by differentiating
Eq.~(\ref{Gi-def}),
\begin{equation}
-\frac{d}{d\zeta^2}\, G = GG,
\label{GG-iden}
\end{equation}
and ignoring a boundary term.
To proceed further we generalize Eqs.~(\ref{G12-casc}) and (\ref{G123-casc})
and decompose a Green's function involving $N$ potentials
into irreducible $N$-body parts,
\begin{equation}
G_{1\ldots N} = G_0 + \sum_i\Delta G_i + \sum_{i<j}
\Delta G_{ij} + \dots.
\label{GN-casc}
\end{equation}
Using Eq.~(\ref{GG-iden}), the irreducible one--two--and three-body
parts of the Green's functions can be written in the form ($i\neq j\neq k$)
\begin{subequations}
\begin{eqnarray}
\Delta G_{i} &=& G_i-G_0 = -\frac{d}{d\zeta^2}\,\ln \frac{G_{i}}{G_0}, \\
\Delta G_{ij} &=& G_{ij}-\Delta G_i - \Delta G_j
=-\frac{d}{d\zeta^2}\,\ln \frac{G_{ij}G_0}{G_iG_j}, \\
\Delta G_{ijk} &=& G_{ijk} - \Delta G_{ij}- \Delta G_{jk}- \Delta G_{ki}
-\Delta G_i - \Delta G_j - \Delta G_k =-\frac{d}{d\zeta^2}\,\ln
\frac{G_{ijk}G_iG_jG_k}{G_{ij}G_{jk}G_{ki}G_0},
\end{eqnarray}
\end{subequations}
which is a (cascading) recursive definition
that can be extended to higher $N$.
Eqs.~(\ref{TrG-for}) and (\ref{GN-casc}) imply a corresponding decomposition
of the vacuum energy into irreducible $N$-body contributions,
\begin{equation}
E_{1\ldots N} = E_0 + \sum_i\Delta E_i + \sum_{i<j} \Delta E_{ij} + \dots,
\label{EN-casc}
\end{equation}
where
\begin{equation}
\Delta E_{1\ldots N} = - \frac{1}{2} \int_{-\infty}^{\infty}\frac{d\zeta}{2\pi}
\,2\zeta^2\, \text{Tr}\, \Delta G_{1\ldots N}.
\label{DE1N=DG1N}
\end{equation}
As shown in~\cite{Schaden:2010wv}, and as will be explicitly verified in
examples below, the irreducible $N$-body contribution to the vacuum energy
diverges only if all $N$ potentials have a common support.
One-body vacuum energies thus are generically divergent, whereas
two-body Casimir energies diverge only if the two bodies intersect
(and thus could be viewed as one). More interestingly though,
three-body Casimir energies diverge only when all three objects
have a \emph{common} intersection--the three bodies need not be
mutually disjoint and could, for instance, be arranged to form a triangle.
Eq.~(\ref{G1-N=M1N}) relates the irreducible $N$-body contribution
of the Green's functions to the irreducible $N$-body transition matrix,
\begin{equation}
\text{Tr}\, \Delta G_{1\ldots N} =-\text{Tr}\, \Delta T_{1\ldots N}G_0 G_0
=\text{Tr}\, \Delta T_{1\ldots N} \frac{d}{d\zeta^2} G_0.
\label{GN-TN}
\end{equation}
The support of delta-function potentials $V_i$ is restricted to
the surface $S_i$ of the $i$-th object and components of the transition-matrix
at most have support on the union of two such surfaces.
It is therefore convenient to formally define a vector ${\bf R}({\bf x})$,
and a matrix ${\bf R}$, with components
\begin{equation}
{\bf R}_{i}({\bf x}) :=G_0({\bf x}-{\bf y})\big|_{{\bf y}\in S_i},
\qquad {\bf R}_{ij}:=
G_0({\bf x}-{\bf y})\big|_{\genfrac{}{}{0pt}{}{{\bf x}\in S_i}{{\bf y}\in S_j}}.
\label{def-R}
\end{equation}
Using these definitions in Eq.~(\ref{GG-iden}) we have
\begin{equation}
-\frac{d}{d\zeta^2} {\bf R}
=\int d^3x\, {\bf R}({\bf x})\cdot {\bf R}({\bf x})^T.
\label{dR=RR}
\end{equation}
Writing the irreducible $N$-body transition
operator in Eq.~(\ref{GN-TN}) in matrix notation and using
Eq.~(\ref{dR=RR}), we express the irreducible
$N$-body contribution to the vacuum energy of Eq.~(\ref{DE1N=DG1N})
in the form
\begin{equation}
\Delta E_{1\ldots N} = -\frac{1}{2} \int_{-\infty}^{\infty}\frac{d\zeta}{2\pi}
\,2\zeta^2\, \text{Tr}
\left[ \Delta {\bf T}_{1\ldots N} \cdot \frac{d}{d\zeta^2} {\bf R} \right].
\label{Tr-TR-for}
\end{equation}
The trace in the last expression is over matrix indices and includes
integrals over the lower dimensional surfaces of the associated objects.
Note also that Eq.~(\ref{Tr-TR-for})
involves the irreducible $N$-body transition matrix, ${\bf T}_{1\ldots N}$,
not its partly amputated cousin $\tilde{\bf T}_{1\ldots N}$.
\section{Casimir energies for parallel semitransparent $\delta$-plates}
\label{Cas-en-semiP:s}
We illustrate this formalism by evaluating the irreducible (scalar)
Casimir energy for semitransparent parallel plates described by potentials
of the form given in Eq.~(\ref{Vi-def}). Exploiting translational
invariance parallel to the plates in Eq.~(\ref{Tr-TR-for}),
the irreducible $N$-body Casimir energy per unit area is described by
dimensionally reduced quantities
\begin{equation}
\frac{\Delta E_{1\ldots N}}{L_xL_y}
=-\frac{1}{6\pi^2} \int_0^\infty \kappa^4d\kappa \,\text{Tr}
\left[ \Delta {\bf t}_{1\ldots N} \cdot \frac{d}{d\zeta^2} {\bf r} \right],
\label{Tr-TR-forp}
\end{equation}
where $L_x$ and $L_y$ are the (infinite) lengths of the plates in
$x$ and $y$ direction, respectively.
The integrals on $\zeta$, $k_x$, and $k_y$, are performed using
polar varibales, which effectively amounts to replacing
$\zeta^2\to 2\kappa^2/3$, where $\kappa$ was defined after Eq.~(\ref{Gi=Igi}).
The dimensionally reduced transition matrices,
$\Delta {\bf t}_{1\ldots N}$, for $N=2$ and $N=3$ are, respectively,
given by Eqs.~(\ref{Dt12=sol}) and (\ref{Delt123-3psol}).
The derivative of the dimensionally reduced free Green's function in
this case is the matrix
\begin{equation}
-\frac{d}{d\zeta^2} {\bf r} =\int_{-\infty}^\infty dz\,
{\bf r}(z)\cdot {\bf r}(z)^T
= \frac{2}{(2\kappa)^3} \left[ \begin{array}{cccc}
1 & (1+\kappa a_{12})\,e^{-\kappa a_{12}} & \cdots
& (1+\kappa a_{1N})\,e^{-\kappa a_{1N}} \\[1mm]
(1+\kappa a_{21})\,e^{-\kappa a_{21}} &1 & \cdots
& (1+\kappa a_{2N})\,e^{-\kappa a_{2N}} \\[1mm]
\vdots & \vdots & \ddots & \vdots \\[1mm]
(1+\kappa a_{N1})\,e^{-\kappa a_{N1}} & (1+\kappa a_{N2})\,e^{-\kappa a_{N2}}
& \cdots & 1
\end{array} \right],
\label{pp-nmat}
\end{equation}
where $a_{ij}$ is the distance between the $i$-th and $j$-th parallel plate
defined previously.
\subsection{$N=1,2$: Irreducible one--and two-body Casimir energy for
semitransparent plates}
The irreducible one-body vacuum energy per unit area associated with
the $i$-th plate diverges. Eq.~(\ref{Tr-TR-forp}) gives it as the integral
\begin{equation}
\frac{\Delta E_i}{L_xL_y}
=\frac{1}{12\pi^2}\int_0^\infty\kappa^2 d\kappa\,\bar{t}_i,
\end{equation}
with $\bar{t}_i$ defined in Eq.~(\ref{g1-exp}).
The one-body vacuum energies are ultra-violet divergent at any
non-vanishing coupling, but do not depend on the relative position
of the plates and therefore do not contribute to forces between them.
The irreducible two-body Casimir energy per unit area associated with plates
$i$ and $j$ is obtained by inserting
Eqs.~(\ref{Dt12=sol}) and (\ref{pp-nmat}) (for $N=2$)
in Eq.~(\ref{Tr-TR-forp}),
\begin{equation}
\frac{\Delta E_{ij}}{L_xL_y}
= - \frac{1}{12\pi^2} \int_0^\infty \kappa^2d\kappa
\left[ \frac{1}{\Delta_{ij}} -1 \right]
\big[ 2\kappa a_{ij} + (1-\bar{t}_i) + (1-\bar{t}_j) \big],
\label{2-semi}
\end{equation}
where the two-body determinant is given by
Eq.~(\ref{Xij-sol}). Eq.~(\ref{2-semi}) for the Casimir interaction
energy of two semitransparent plates was obtained previously
in \cite{Bordag:1992cm}.
In the Dirichlet limit, $\bar{t}_i\to 1$, Eq.~(\ref{2-semi}) simplifies
to the well known Casimir energy for a massless scalar field satisfying
Dirichlet boundary conditions on a pair of parallel plates,
\begin{equation}
\frac{\Delta E_{ij}^D}{L_xL_y}
= - \frac{1}{12\pi^2} \int_0^\infty \kappa^2d\kappa
\frac{2\kappa a_{ij}}{e^{2\kappa a_{ij}}-1}
=-\frac{\pi^2}{1440}\frac{1}{a_{ij}^3}.
\label{DCasplates}
\end{equation}
Eqs.~(\ref{2-semi}) and (\ref{DCasplates}) are finite and negative for
two disjoint plates.
\subsection{$N=3$: Three-body Casimir energy for three parallel plates}
The irreducible three-body Casimir energy of three plates is
similarly obtained by inserting Eqs.~(\ref{Delt123-3psol}) and (\ref{pp-nmat})
(for $N=3$) in Eq.~(\ref{Tr-TR-forp}),
\begin{eqnarray}
\frac{\Delta E_{ijk}}{L_xL_y}
&=& \frac{1}{12\pi^2} \int_0^\infty \kappa^2d\kappa
\bigg[
\bar{t}_i \left\{ 1 -\frac{1}{\Delta_{ij}} -\frac{1}{\Delta_{ik}}
+\frac{\Delta_{jk}}{\Delta_{ijk}} \right\}
+ 2(1+\kappa a_{jk}) \left\{\frac{1}{\Delta_{jk}} -1\right\}
\left\{1-\bar r_{jk[i]}e^{\kappa a_{jk}}\frac{\Delta_{jk}}{\Delta_{ijk}}\right\}
\nonumber \\ && \hspace{25mm}
+ \bar{t}_j \left\{ 1 -\frac{1}{\Delta_{ji}} -\frac{1}{\Delta_{jk}}
+\frac{\Delta_{ik}}{\Delta_{ijk}} \right\}
+ 2(1+\kappa a_{ik}) \left\{\frac{1}{\Delta_{ik}} -1\right\}
\left\{1-\bar r_{ik[j]}e^{\kappa a_{ik}}\frac{\Delta_{ik}}{\Delta_{ijk}}\right\}
\nonumber \\ && \hspace{25mm}
+\bar{t}_k \left\{ 1 -\frac{1}{\Delta_{ki}} -\frac{1}{\Delta_{kj}}
+\frac{\Delta_{ij}}{\Delta_{ijk}} \right\}
+ 2(1+\kappa a_{ij}) \left\{\frac{1}{\Delta_{ij}} -1\right\}
\left\{1-\bar r_{ij[k]}e^{\kappa a_{ij}}\frac{\Delta_{ij}}{\Delta_{ijk}}\right\}
\bigg].
\label{DE123=gen}
\end{eqnarray}
When Dirichlet boundary conditions are imposed on the central $j$-th plate,
the relation between irreducible two--and three-body transition matrices
noted in Eq.~(\ref{etijk=2tik}) implies a corresponding relation
between two--and three-body Casimir energies,
\begin{equation}
\Delta E_{i\infty k} + \Delta E_{ik} = 0.
\end{equation}
This is explicitly verified by using the factorization of the
three-body determinant in Eq.~(\ref{det-fact}) and the Dirichlet limits for
$\bar r_{ij[k]}$ given after Eq.~(\ref{det-fact}) in Eq.~(\ref{DE123=gen}),
and identifying the irreducible two-body energy of Eq.~(\ref{2-semi})
in the result.
In the Dirichlet limit for all three plates the irreducible three-body
Casimir energy cancels the well-known two-body interaction between the
outer Dirichlet plates,
\begin{equation}
\frac{\Delta E_{ijk}^D}{L_xL_y} =\frac{\pi^2}{1440}\frac{1}{a_{ik}^3}
= -\frac{\Delta E_{ik}^D}{L_xL_y},
\end{equation}
where $a_{ik}$ is the distance between the outer plates.
This cancellation is to be expected on physical grounds and serves
to check the calculation. For semitransparent plates the cancellation
is not complete and the irreducible three-body contribution to
the total Casimir energy can be significant
for parallel plates. Note that the sign of the irreducible $N$-body
contribution to the scalar Casimir energy alternates.
Although not apparent from the expression of Eq.~(\ref{DE123=gen}),
this irreducible three-body
contribution to the Casimir energy is positive for \emph{any} positive
couplings $\lambda_1,\lambda_2,\lambda_3$ and \emph{any} relative
position of the three plates.
For parallel semitransparent plates we thus verify the more
general result obtained in~\cite{Schaden:2010wv}.
Also, as discussed in~\cite{Schaden:2010wv}
and noted previously, the three-plate Casimir energy diverges only if
\emph{all three} plates coincide.
In the following we will see that these generic results for the sign
and analyticity of the three-body scalar Casimir energy hold in the
limit where two of the three potentials are weak and need only be
accounted for to leading order.
\section{Three-body Scalar Casimir interaction for semiweak coupling}
\label{3body-sweak:s}
We now consider irreducible vacuum energies for three bodies when
two of the three potentials, $V_1$ and $V_2$,
are weak and need only be taken to leading
order. No restriction is imposed on the potential $V_3$ describing the
third body. To the leading order we thus approximate
$T_1\sim V_1$ and $T_2\sim V_2$ in Eq.~(\ref{ti-def}).
The three-body transition matrix of Eq.~(\ref{123-rtr})
in this semiweak approximation simplifies to
\begin{equation}
\tilde {\bf T}_{123}^W =
\left[ \begin{array}{ccc}
1 & 0&0 \\ 0&1&0 \\ 0&0& X_{3[12]}^W
\end{array}\right]
\left[ \begin{array}{ccc}
\tilde V_1 & -\tilde V_1(1-\tilde T_3)\tilde V_2
& -\tilde V_1(1-\tilde V_2+\tilde T_3\tilde V_2)\tilde T_3 \\
-\tilde V_2(1-\tilde T_3)\tilde V_1 & \tilde V_2
& -\tilde V_2(1-\tilde V_1+\tilde T_3\tilde V_1)\tilde T_3 \\
-\tilde T_3(1-\tilde V_2)\tilde V_1
& -\tilde T_3(1-\tilde V_1)\tilde V_2 & \tilde T_3
\end{array}\right].
\label{123-rtr-w}
\end{equation}
Here $X_{3[12]}^W$ satisfies Eq.~(\ref{Xijk-def}),
which to leading semiweak approximation is solved by
\begin{equation}
X_{3[12]}^W = 1 +\tilde T_3(1-\tilde V_1)\tilde V_2
+\tilde T_3(1-\tilde V_2)\tilde V_1
+\tilde T_3\tilde V_1\tilde T_3\tilde V_2
+\tilde T_3\tilde V_2\tilde T_3\tilde V_1.
\end{equation}
The transition matrix in semiweak approximation of Eq.~(\ref{123-rtr-w})
may again be decomposed into its irreducible
one--two--and three-body parts, leading to the semiweak version of
Eq.~(\ref{T123=T1..DT123}).
From Eq.~(\ref{DT12=expl}) the irreducible two-body transition matrices in
semiweak approximation are,
\begin{equation}
\Delta \tilde{\bf T}_{12}^W=
\left[ \begin{array}{cc} 0 & -\tilde V_1\tilde V_2 \\
-\tilde V_2\tilde V_1&0 \end{array}\right],
\qquad \Delta \tilde {\bf T}_{i3}^W=
\left[ \begin{array}{cc} 0 & -\tilde V_i\tilde T_3 \\
-\tilde T_3\tilde V_i& \tilde T_3\tilde V_i\tilde T_3 \end{array}\right],
\label{TW12}
\end{equation}
with $i=1,2$. Similarly approximating Eq.~(\ref{d123-rtr}),
the three-body transition matrix in semiweak approximation becomes,
\begin{equation}
\Delta \tilde{\bf T}_{123}^W
= \left[ \begin{array}{ccc}
0 & \tilde V_1\tilde T_3\tilde V_2
& \tilde V_1\tilde G_3\tilde V_2\tilde T_3 \\
\tilde V_2\tilde T_3\tilde V_1&0 & \tilde V_2\tilde G_3\tilde V_1\tilde T_3 \\
\tilde T_3\tilde V_2\tilde G_3\tilde V_1 \hspace{3mm}
& \tilde T_3\tilde V_1\tilde G_3\tilde V_2 \hspace{3mm} &
-\tilde T_3\tilde V_1\tilde G_3\tilde V_2\tilde T_3
-\tilde T_3\tilde V_2\tilde G_3\tilde V_1\tilde T_3
\end{array}\right],
\label{DT123W}
\end{equation}
where $\tilde G_3=1-\tilde T_3$.
Casimir energies in the semiweak approximation are obtained using
Eqs.~(\ref{DE1N=DG1N}) and (\ref{GN-TN}). Inserting Eq.~(\ref{TW12})
in Eq.~(\ref{GN-TN}) we have to this approximation,
\begin{subequations}
\begin{eqnarray}
-\text{Tr}\,\Delta G_{12}^W = \text{Tr}
\left[ \Delta {\bf T}_{12}^W \cdot \frac{d}{d\zeta^2} {\bf R} \right]
&=& \frac{d}{d\zeta^2} \text{Tr} \Big[ G_0V_1G_0V_2 \Big], \\
-\text{Tr}\,\Delta G_{i3}^W = \text{Tr}
\left[ \Delta {\bf T}_{i3}^W \cdot \frac{d}{d\zeta^2} {\bf R} \right]
&=& \frac{d}{d\zeta^2} \text{Tr} \Big[ G_0V_iG_0T_3 \Big], \qquad (i=1,2).
\end{eqnarray}
\label{DG12=dztr}
\end{subequations}
The corresponding irreducible three-body contribution
using Eq.~(\ref{DT123W}) in Eq.~(\ref{GN-TN}) is
\begin{equation}
-\text{Tr}\,\Delta G_{123}^W = \text{Tr}
\left[ \Delta {\bf T}_{123}^W \cdot \frac{d}{d\zeta^2} {\bf R} \right]
= -\frac{d}{d\zeta^2} \text{Tr} \Big[
G_0V_1G_0T_3G_0V_2 +G_0V_2G_0T_3G_0V_1 - G_0T_3G_0V_1G_0T_3G_0V_2 \Big].
\label{TrDT123R}
\end{equation}
Inserting Eq.~(\ref{DG12=dztr}) in Eq.~(\ref{DE1N=DG1N}),
and integrating by parts, the irreducible two-body Casimir energies
in semiweak approximation are
\begin{subequations}
\begin{eqnarray}
\Delta E_{12}^W &=& -\frac{1}{2} \int_{-\infty}^{\infty}\frac{d\zeta}{2\pi}
\,\text{Tr} \big[G_0V_1G_0V_2\big], \\
\Delta E_{i3}^W &=& -\frac{1}{2} \int_{-\infty}^{\infty}\frac{d\zeta}{2\pi}
\,\text{Tr} \big[G_0V_iG_0T_3\big], \qquad (i=1,2),
\label{DelE=2VVT-w}
\end{eqnarray}
\label{DelE=2VVT-wcom}%
\end{subequations}
verifying results reported in \cite{Milton:2007wz}.
The corresponding irreducible three-body contribution to the Casimir energy
in semiweak approximation
using Eq.~(\ref{TrDT123R}) in Eq.~(\ref{DE1N=DG1N}) is
\begin{equation}
\Delta E_{123}^W = \frac{1}{2} \int_{-\infty}^{\infty}\frac{d\zeta}{2\pi}
\,\text{Tr} \big[
G_0V_1G_0T_3G_0V_2 +G_0V_2G_0T_3G_0V_1 - G_0T_3G_0V_1G_0T_3G_0V_2 \big].
\label{DelE=VVT-w}
\end{equation}
In the following we evaluate Eqs.~(\ref{DelE=2VVT-wcom}) and
(\ref{DelE=VVT-w}) for some special cases.
\subsection{Weak point potentials}
Weak point potentials of the form,
\begin{equation}
V_i({\bf x})=\lambda_i\delta({\bf x}-{\bf x}_i),
\end{equation}
for $i=1,2$, allow one to explicitly perform the integrals
in Eqs.~(\ref{DelE=2VVT-wcom}) and (\ref{DelE=VVT-w}).
In this case we have
\begin{equation}
\Delta E_{12}^W
= -\frac{\lambda_1\lambda_2}{2} \int_{-\infty}^{\infty}\frac{d\zeta}{2\pi}
\,\big[ G_0({\bf x}_1-{\bf x}_2)\big]^2<0,
\end{equation}
and, using Eq.~(\ref{Gi-gsol}) in Eq.~(\ref{DelE=2VVT-w}),
\begin{equation}
\Delta E_{i3}^W
= -\frac{\lambda_i}{2} \int_{-\infty}^{\infty}\frac{d\zeta}{2\pi}
\,\Big\{ G_0(0)-G_3({\bf x}_i,{\bf x}_i)\Big\}<0, \qquad (i=1,2),
\label{2point-w}
\end{equation}
because the integrand in braces is positive for positive $V_3$.
The irreducible two-body contributions to the vacuum energy thus are negative
for weak point potentials. We similarly obtain that the irreducible three-body
correction to the vacuum energy,
\begin{equation}
\Delta E_{123}^W
= \frac{\lambda_1\lambda_2}{2} \int_{-\infty}^{\infty}\frac{d\zeta}{2\pi}
\,\Big\{ \big[G_0({\bf x}_1-{\bf x}_2)\big]^2
-\big[G_3({\bf x}_1,{\bf x}_2)\big]^2 \Big\}>0,
\label{3point-w}
\end{equation}
in this case is positive for any (positive) potential $V_3$.
Note that the irreducible three-body Casimir energy in semiweak
approximation diverges only if ${\bf x}_1={\bf x}_2$ is in the
support of $V_3$.
The pattern in the sign of the irreducible $N$-body contribution
is consistent with the findings of \cite{Schaden:2010wv}. Furthermore,
since any positive potential is a (positive) superposition of point
potentials, this pattern of the signs of irreducible $N$-body
contributions extend to any shape of the objects in semiweak approximation.
This is explicitly verified by the following examples.
\subsection{Weak potentials with translational symmetry
parallel to a Dirichlet plate}
We consider a Dirichlet plate and weak potentials that do not depend
on the Cartesian coordinate $x$,
\begin{equation}
V_i=V_i(y,z), \quad \text{for} \quad i=1,2;
\quad \text{and} \quad
V_3=\lambda_3\,\delta(z-a_3), \quad \text{with} \quad \lambda_3\to\infty.
\end{equation}
To evaluate Eqs.~(\ref{DelE=2VVT-wcom}) and (\ref{DelE=VVT-w})
for such potentials
we require the operator $G_0T_3^DG_0$ for a Dirichlet plate.
In order to exploit the translational symmetry in $x$-direction
we write the solution to Eq.~(\ref{G0-def}) for the free Green's function
in the form
\begin{equation}
G_0(|{\bf x}_1-{\bf x}_2|;\zeta)
= \int \frac{d^2k}{(2\pi)^2}\,e^{i{\bf k}\cdot({\bf x}_1-{\bf x}_2)_\perp}
\frac{e^{-\kappa |z_1-z_2|}}{2\kappa}
= \int_{-\infty}^{\infty}\frac{dk_x}{2\pi}\,e^{ik_x(x_1-x_2)}
\frac{K_0(\bar{\kappa}\, d_{12})}{2\pi}
=\frac{e^{-|\zeta| |{\bf x}_1-{\bf x}_2|}}{4\pi |{\bf x}_1-{\bf x}_2|},
\label{G0=K0}
\end{equation}
where $d_{12}=\sqrt{(y_1-y_2)^2+(z_1-z_2)^2}$ is the projected distance
in the $x_1=x_2$ plane, and $\bar{\kappa}^2= k_x^2+\zeta^2$.
$K_0(x)$ is the modified Bessel function of zero order.
Note that $\kappa$ defined after Eq.~(\ref{Gi=Igi}) satisfies
$\kappa^2 = \bar{\kappa}^2+k_y^2$.
Using the first equality of Eq.~(\ref{G0=K0})
and the dimensionally reduced transition matrix of a Dirichlet plate
given in Eq.~(\ref{g1-exp}) one can show that
\begin{equation}
-\Delta G_3^D({\bf x}_1,{\bf x}_2;\zeta)
=[G_0T_3^DG_0]({\bf x}_1,{\bf x}_2;\zeta)
= G_0(|{\bf x}_1-\bar{\bf x}_2|;\zeta)
= \int_{-\infty}^{\infty}\frac{dk_x}{2\pi}\,e^{ik_x(x_1-x_2)}
\frac{K_0(\bar{\kappa}\, \bar{d}_{12})}{2\pi},
\label{G0T3DG0}
\end{equation}
where $\bar{\bf x}_2=(x_2,y_2,-z_2+2a_3)$, and
$\bar{d}_{12}$ is the length of the shortest path between
${\bf x}_1$ and ${\bf x}_2$ in the $(x_1=x_2)$-plane
that reflects off the Dirichlet plate.
For a Dirichlet plate at $z=a_3$, this distance is given by
$\bar{d}^2_{12}=(y_1-y_2)^2+(|z_1-a_3|+|z_2-a_3|)^2$.
A geometrical interpretation of $d_{12}$ and $\bar{d}_{12}$
is shown in FIG.~\ref{r12-rb12}.
Substituting Eq.~(\ref{G0T3DG0}) in Eq.~(\ref{Gi-gsol}) leads to
\begin{equation}
G_3^D({\bf x}_1,{\bf x}_2;\zeta)
=G_0(|{\bf x}_1-{\bf x}_2|;\zeta)- G_0(|{\bf x}_1-\bar{\bf x}_2|;\zeta),
\label{G1D=G0G0}
\end{equation}
which is anti-symmetric under reflection about the Dirichlet plate.
Note that if ${\bf x}_1$ and ${\bf x}_2$ are on opposite
sides of the plate, $\bar{d}_{12}=d_{12}$, and $G_3^D$ vanishes.
\begin{center}
\begin{figure}
\includegraphics{figures/eff-dist.eps}
\caption{The distances $d_{12}$ and $\bar{d}_{12}$.
The effective distance
$\bar{d}_{12}$ is the shortest distance between the two points for a path
that reflects off the Dirichlet plate at $z=a_3$.
It also is the shortest distance between $(y_1,z_1)$ and a mirror
image of the point $(y_2,z_2)$ with respect to the $z=a_3$ line.}
\label{r12-rb12}
\end{figure}
\end{center}
Substituting Eq.~(\ref{G0=K0}) for the free Green's functions,
and Eq.~(\ref{G0T3DG0}) for the irreducible Green's function of a Dirichlet
plate in Eqs.~(\ref{DelE=2VVT-wcom}) and using the identity
\begin{equation}
\int_0^\infty \bar{\kappa}\,d\bar{\kappa} \,K_0(\bar{\kappa} x)
= \frac{1}{x^2},
\label{K0-iden}
\end{equation}
the irreducible two-body Casimir energies per unit length
in semiweak approximation for potentials with translational symmetry are
\begin{subequations}
\begin{eqnarray}
\frac{\Delta E_{12}^W}{L_x}
&=& -\frac{1}{32\pi^3} \int d^2r_1 \int d^2r_2
\frac{V_1({\bf r}_1) V_2({\bf r}_2)}{d_{12}^2},
\label{V1V2-w} \\[2mm]
\frac{\Delta E_{i3}^W}{L_x}
&=& -\frac{1}{32\pi^2} \int d^2r \frac{V_i({\bf r})}{|z|^2},
\qquad (i=1,2).
\label{E12W-ViDP}
\end{eqnarray}
\label{E12W-all}%
\end{subequations}
The Casimir energy in Eq.~(\ref{V1V2-w}) for two weakly interacting
objects with translational symmetry was previously obtained in
\cite{Wagner:2008qq}.
The Casimir energy for a Dirichlet plate weakly interacting with an
object possessing translational symmetry was obtained in \cite{Milton:2007wz},
but was given as a series involving modified Bessel functions.
The expression in \cite{Milton:2007wz} generally is much harder to
evaluate than Eq.~(\ref{E12W-ViDP}).
The simplification in Eq.~(\ref{E12W-ViDP}) was achieved by
using the effective Green's function for a Dirichlet plate in
Eq.~(\ref{G0T3DG0}). For many potentials, the evaluation of the
Casimir energy by Eq.~(\ref{E12W-ViDP}) is immediate.
We can for example
calculate the two-body Casimir energy for a cylinder of radius $a$,
described by the weak potential $V_\text{cyl}=\lambda\,\delta (r-a)$,
interacting with a Dirichlet plate positioned at $z=R>a$.
From Eq.~(\ref{E12W-ViDP}) one readily finds,
\begin{equation}
\frac{\Delta E_\text{Cyl-DP}^W}{L_x}
= -\frac{1}{32\pi^2} \int_0^\infty rdr \int_0^{2\pi} d\theta
\frac{\lambda\delta (r-a)}{|r\sin\theta -R|^2}
=-\frac{\lambda a}{16\pi}\frac{1}{R^2}
\left[1 - \frac{a^2}{R^2} \right]^{-\frac{3}{2}},
\end{equation}
which reproduces the expression in \cite{Milton:2007wz}.
A similarly simplified evaluation is expected for
an arbitrary surface with translational symmetry weakly interacting
with a Dirichlet plate parallel to the symmetry axis.
The irreducible three-body Casimir energies for translationally
invariant weak potentials
and a Dirichlet plate can be similarly evaluated using
Eq.~(\ref{DelE=VVT-w}).
The first two terms in Eq.~(\ref{DelE=VVT-w}) involve the product of
the free Green's function, $G_0$, with the irreducible Green's function
for a Dirichlet plate given in Eq.~(\ref{G0T3DG0}).
The last term requires the product of two irreducible one-body Green's
functions. A useful identity for the product of two modified Bessel functions
of zeroth order is
\begin{equation}
\int_0^\infty \bar{\kappa}\,d\bar{\kappa}
\,K_0(\bar{\kappa} x) K_0(\bar{\kappa} y)
= \frac{1}{x^2-y^2} \ln \left(\frac{x}{y}\right)
\xrightarrow{x\to y} \frac{1}{2x^2}.
\label{K0-iden2}
\end{equation}
Inserting Eqs.~(\ref{G0=K0}) and (\ref{G0T3DG0}) in Eq.~(\ref{DelE=VVT-w})
to write the Green's functions in terms of modified Bessel functions,
and then using Eq.~(\ref{K0-iden2}), we obtain
\begin{equation}
\frac{\Delta E_{123}^W}{L_x}
= \frac{1}{32\pi^3} \int d^2r_1 \int d^2r_2
\frac{V_1({\bf r}_1) V_2({\bf r}_2)}{\bar{d}_{12}^2}
\;Q\left(\frac{d_{12}^2}{\bar{d}_{12}^2} \right),
\label{DelE=pot-w}
\end{equation}
where the distances $d_{12}$ and $\bar{d}_{12}$ were introduced earlier
and are shown in FIG.~\ref{r12-rb12}.
The function
\begin{equation}
Q(x) = -\frac{2\ln x}{1-x}-1
\label{Q3-kernel}
\end{equation}
is bounded by $1\leq Q(x)\leq 1-2\ln x$
in the relevant domain $0<x=\frac{d_{12}^2}{\bar{d}_{12}^2}<1$.
This implies that the three-body Casimir energy of Eq.~(\ref{DelE=pot-w})
is always positive and bounded by
\begin{equation}
\frac{1}{32\pi^3} \int d^2r_1 \int d^2r_2
\frac{V_1({\bf r}_1) V_2({\bf r}_2)}{\bar{d}_{12}^2}
\leq \frac{\Delta E_{123}^W}{L_x}\leq
\frac{1}{32\pi^3} \int d^2r_1 \int d^2r_2
\frac{V_1({\bf r}_1) V_2({\bf r}_2)}{\bar{d}_{12}^2}
\left[ 1-2\ln\left(\frac{d^2_{12}}{\bar{d}^2_{12}}\right)\right].
\label{DE123W-bounds}
\end{equation}
$\bar{d}_{12}$ is the distance between a point on the first object
and another point on the reflected image of the second object
(see FIG.~\ref{r12-rb12}). It vanishes only at points where the
two weak objects \emph{and} the Dirichlet plate are concurrent.
The irreducible three-body Casimir energy in the semiweak approximation
of Eq.~(\ref{DelE=pot-w}) thus is finite if the three objects have
no point in common. This contribution in particular does not diverge
as the objects approach the plate or each other, corroborating the
findings in \cite{Schaden:2010wv}. Note that the lower bound in
Eq.~(\ref{DE123W-bounds}) is the two-body Casimir energy between
weak potentials of Eq.~(\ref{V1V2-w}), but with the reflected object
($d_{12}\to\bar{d}_{12}$) and of opposite sign. The irreducible
three-body Casimir energy approaches the lower bound for
$\frac{d_{12}^2}{\bar{d}_{12}^2}\sim 1$ and thus partially cancels
the irreducible two-body energy if one or both objects approach the
Dirichlet plate. In fact, if the two weakly interacting objects are
entirely on opposite sides of the Dirichlet plate, the lower bound
is achieved because $\bar{d}_{12}=d_{12}$ and the three-body
Casimir energy cancels the two-body interaction energy between them
precisely.
The following examples demonstrate the finiteness, sign, and analyticity,
of three-body contributions to Casimir energies for cases in which irreducible
one--\emph{and} two-body contributions to the vacuum energy diverge.
\section{Triangular-wedge on a Dirichlet plate}
\label{tri-wedge:s}
We first consider a triangular-wedge with two sides
described by weak potentials atop a Dirichlet plate at $z=0$,
forming a waveguide of triangular cross-section:
\begin{subequations}
\begin{eqnarray}
V_1(y,z) &=& \lambda_1 \delta (-z+m_\alpha(y-a)) \,\theta_1,
\quad \text{with} \quad
\theta_1\equiv\theta(y-\text{min}[0,a])\,\theta(\text{max}[0,a]-y), \\
V_2(y,z) &=& \lambda_2 \delta (-z+m_\beta(y-b)) \,\theta_2,
\quad \text{with} \quad
\theta_2\equiv\theta(y-\text{min}[0,b])\,\theta(\text{max}[0,b]-y),\\
V_3(z) &=& \lambda_3 \delta (z),
\quad \text{with} \quad \lambda_3\to\infty.
\end{eqnarray}
\label{wed-pot-w}%
\end{subequations}
The sides of the wedge have slopes
$m_\alpha=-\cot\alpha$ and $m_\beta=-\cot\beta$,
and lengths $\sqrt{h^2+a^2}$ and $\sqrt{h^2+b^2}$, respectively.
The constraint $m_\alpha a=m_\beta b=-h$ forces the sides to intersect
at $(y=0,z=h)$, where $h$ is the height of the triangle.
The base of the triangle formed then measures $|b-a|$.
Note that the Dirichlet plate at $z=0$ is of infinite extent.
This triangular-wedge on a Dirichlet plate is depicted in FIG.~\ref{wedoTri}.
Suitable parameters for describing the triangular waveguide are
$(h,\alpha,\beta)$, or $(h,\tilde{a}=a/h,\tilde{b}=b/h)$.
Without loss of generality we measure all lengths in multiples of
the height $h$. The triangle then has height $h=1$ and
the parameter space for the triangle is $-\infty<a,b<\infty$,
or, equivalently, $-\pi/2<\alpha,\beta<\pi/2$.
\begin{center}
\begin{figure}
\includegraphics{figures/tri-wedge.eps}
\caption{Weakly interacting triangular-wedge on a Dirichlet plate.
The objects are of infinite extent in the $x$-direction.
The weakly interacting sides of the wedge (in red) have finite length.}
\label{wedoTri}
\end{figure}
\end{center}
Observe that all irreducible two-body Casimir energies in Eq.~(\ref{E12W-all})
diverge due to ultra-violet contributions from the corners of the triangle
where pairs of potentials overlap.
More precisely, the integrand $\Delta E_{12}^W$ diverges
when $d_{12}\sim 0$ near the vertex of the wedge. The integrand of
$\Delta E_{i3}^W$ diverges when $z_i\sim 0$ near the corner with
the Dirichlet plate. The irreducible three-body Casimir energy,
$\Delta E_{123}^W$, in Eq.~(\ref{DelE=pot-w})
on the other hand is finite because $\bar{d}_{12}$
never vanishes in the integration region.
Substituting the potentials of Eq.~(\ref{wed-pot-w})
for the semiweak triangular waveguide in Eq.~(\ref{DelE=pot-w}) and
evaluating the $z$-integrals gives,
\begin{equation}
{\cal E}(\alpha,\beta)=
\frac{\Delta E_{123}^W}{L_x}
\left[ \frac{\lambda_1\lambda_2}{32\pi^3}\right]^{-1}
=|\tilde{a}\,\tilde{b}|
\int_0^1 \int_0^1 \frac{du_1 du_2}{\bar{u}_{12}^2}
\,Q\left( \frac{u_{12}^2}{\bar{u}_{12}^2} \right),
\label{E123=tri-wed}
\end{equation}
where we have rescaled the integration variables,
$y_1=|a|u_1$ and $y_2=|b|u_2$ by the respective lengths.
All distances have been expressed in units of $h$:
$d_{12}=hu_{12}$ and $\bar{d}_{12}=h\bar{u}_{12}$, with
\begin{subequations}
\begin{eqnarray}
\bar{u}_{12}^2 &=& (\tilde{a} u_1-\tilde{b} u_2 )^2 + [|1-u_1| +|1-u_2|]^2,\\
u_{12}^2 &=& (\tilde{a} u_1-\tilde{b} u_2)^2 + (u_1 -u_2)^2.
\end{eqnarray}
\end{subequations}
With the function $Q(x)$ defined in Eq.~(\ref{Q3-kernel})
the three-body interaction energy of Eq.~(\ref{E123=tri-wed})
is finite and can be evaluated numerically.
In FIG.~\ref{3dTri} we plot ${\cal E}(\alpha,\beta)$
as a function of the angles $\alpha$ and $\beta$.
The three-body interaction energy is always positive and vanishes
(and is minimized) only for $\alpha=0$, or $\beta=0$.
It is minimal when the shorter side of the wedge is perpendicular
to the Dirichlet plate. Wedges with angles
$\beta<0<\alpha$ or $\alpha<0<\beta$ are energetically
preferred over wedges with angles $\alpha,\beta>0$ or $\alpha,\beta<0$.
The three-body Casimir-energy diverges only when all three sides of
the triangle have a point in common, i.e.
when $\alpha =\beta$, or $\alpha=-\beta=\pm\pi/2$.
\begin{center}
\begin{figure}
\includegraphics[width=8cm]{figures/3Dplot-E123-versus-alpha-and-beta-for-triangular-wedge-on-Dplate.eps-from-jpg}%
\hspace{10mm}
\includegraphics{figures/tri-regions.eps}
\caption{ Casimir Landscape:
${\cal E}(\alpha,\beta)$ as a function of the opening angles
$\alpha$ and $\beta$ for a weakly interacting triangular-wedge
on a Dirichlet plate. The valley connecting the $\alpha>\beta$ region
with the $\alpha<\beta$ region is an artifact caused by limited numerical
accuracy. The valley should be replaced by a very thin
and infinitely high wall describing
the sharp change in energy when all surfaces overlap. On the right,
the shapes of the triangles are matched to the respective regions
of the $\alpha$-$\beta$ plane.}
\label{3dTri}
\end{figure}
\end{center}
Abalo, Milton, and Kaplan, recently~\cite{Abalo:2010ah} investigated
the dependence of the Casimir energy on the area and perimeter of
triangular waveguides on which Dirichlet boundary conditions were imposed.
Although only interior modes were taken into account and divergences
associated with corners and single-body vacuum energies were removed ad hoc,
they found that the dimensionless Casimir energy of their triangular
wave guides closely follow a universal function of the dimensionless
ratio ($P^2/A$) of the perimeter $P$ and area $A$ of the cross-section.
This would imply that the Casimir energy of triangular wave guides
depends on just one, rather than two, dimensionless parameters.
Although we cannot expect a similar dependence, the universal behavior
observed in \cite{Abalo:2010ah} prompted us to also investigate the
dependence of the semiweak three-body Casimir energy on the
dimensionless perimeter $\tilde p=(P/h)$ and dimensionless area
$\tilde s=(A/h^2)$ of the triangular waveguide.
It is also of interest to inquire for what configuration the energy of
a triangular waveguide is minimized if the cross-sectional area
is kept fixed.
The dimensionless area $\tilde s$ and perimeter $\tilde p$ of the
triangular wedge are given by,
\begin{subequations}
\begin{eqnarray}
\frac{A}{h^2}&=&\tilde{s}=\frac{1}{2}|\tilde{b}-\tilde{a}|,\label{s=ab} \\
\frac{P}{h} &=& \tilde{p}= |\tilde{b}-\tilde{a}|
+ \sqrt{1+\tilde{a}^2} +\sqrt{1+\tilde{b}^2}.%
\label{p=ab}
\end{eqnarray}%
\label{sp-eqns}%
\end{subequations}%
The parameter space of a triangular-wedge in this case is
$\tilde{s}\geq 0$, and
$\tilde{p}\geq 2\tilde s + 2\sqrt{1+\tilde s^2} \geq\text{Max}(2,4\tilde s)$.
See FIG.~\ref{3dspTri-wed}. The inequality, $\tilde p > 4\tilde s$,
is a consequence of the triangle inequality.
\begin{center}
\begin{figure}
\includegraphics{figures/e-versus-aoverh-for-constant-area.eps}
\caption{
${\cal E}(\tilde a,\tilde b)$ as a function of $\tilde a$ for fixed area,
$A=h^2$. The irreducible three-body Casimir energy is minimal when
the shorter side of the wedge is perpendicular to the Dirichlet plate
($\tilde a=0$ or $\tilde b=0$).
The maximum in the intermediate region corresponds to the unstable
equilibrium of an isosceles triangle. The dashed curves are the
approximation ${\cal E}(\tilde a,\tilde b)\sim |\tilde a\tilde b|$
obtained by replacing the integrals in Eq.~(\ref{E123=tri-wed}) with unity.
The dotted curves are reflections about the ${\cal E}=0$ line.}
\label{area-fix-fig}
\end{figure}
\end{center}
In FIG.~\ref{area-fix-fig} we plot the energy as a function of
$\tilde a$ for fixed
area: $A=h^2$, or $|\tilde b - \tilde a|=2$, or $|\tan\beta-\tan\alpha|=2$.
The three-body Casimir energy for a waveguide of given cross-sectional
area is minimal
when the shorter side of the wedge is perpendicular to the Dirichlet plate
($\alpha=0$ or $\beta=0~[\alpha=\tan^{-1}(-2)]$).
In the intermediate region the energy is extremal for an isosceles
triangle $(-\tilde a =\tilde b=1$) with ${\cal E}(-1,1)=0.893112\ldots$.
The dashed curve in FIG.~\ref{area-fix-fig} represents the
approximation ${\cal E}(\tilde a, \tilde b)\sim |\tilde a\tilde b|$
obtained by setting the dimensionless integral in Eq.~(\ref{E123=tri-wed})
to 1. Remarkably, this extremely simple expression for
the irreducible three-body energy is accurate to better than ten percent
everywhere. We also show reflections of the curves to illustrate that the
discontinuities in the slope are entirely due to the absolute value
in the pre-factor $|\tilde a \tilde b|$ and the integral itself is analytic.
We rewrite the irreducible three-body Casimir energy as a function
of the cross-sectional area and perimeter by inverting
Eqs.~(\ref{sp-eqns}) to obtain
\begin{subequations}
\begin{eqnarray}
\tilde{a} &=& \begin{cases}
\pm\tilde{\mu}-\tilde{s},\qquad\text{if}\quad \tilde{b}>\tilde{a}, \\
\pm\tilde{\mu}+\tilde{s},\qquad\text{if}\quad \tilde{b}<\tilde{a}, \\
\end{cases} \\
\tilde{b} &=& \begin{cases}
\pm\tilde{\mu}+\tilde{s},\qquad\text{if}\quad \tilde{b}>\tilde{a}, \\
\pm\tilde{\mu}-\tilde{s},\qquad\text{if}\quad \tilde{b}<\tilde{a}, \\
\end{cases}
\end{eqnarray}
\label{ab=spmu}
\end{subequations}
where
\begin{equation}
\tilde{\mu}= \frac{1}{2\tilde{p}}
\frac{(\tilde{p}-2\tilde{s})}{(\tilde{p}-4\tilde{s})}
\bigg[ \tilde{p} (\tilde{p}-4\tilde{s})
\Big\{ \tilde{p} (\tilde{p}-4\tilde{s}) -4\Big\}
\bigg]^\frac{1}{2}.
\end{equation}
Substituting Eqs.~(\ref{ab=spmu}) in Eq.~(\ref{E123=tri-wed}),
the three-body Casimir energy as of function of perimeter and area is
\begin{equation}
{\cal E}(\tilde s,\tilde p) =|\tilde{\mu}^2-\,\tilde{s}^2|
\int_0^1 \int_0^1 \frac{du_1 du_2}{\bar{u}_{12}^2}
\,Q\left( \frac{u_{12}^2}{\bar{u}_{12}^2} \right),
\label{E123sp=tri-wed}
\end{equation}
where the rescaled distances in terms of area and perimeter are given by
\begin{subequations}
\begin{eqnarray}
\bar{u}_{12}^2 &=& [\tilde{\mu} (u_1-u_2) +\tilde{s}(u_1+u_2) ]^2
+ [|1-u_1| +|1-u_2|]^2,\\
u_{12}^2 &=& [\tilde{\mu} (u_1-u_2) +\tilde{s}(u_1+u_2) ]^2 + (u_1 -u_2)^2.
\end{eqnarray}
\end{subequations}
In FIG.~\ref{3dspTri-wed} the irreducible three-body contribution
to the vacuum energy of a semiweak wedge is plotted
as a function of dimensionless area and perimeter of the cross-section.
The energy now is minimal along the curve
\begin{equation}
\tilde{p}=1+2\tilde{s}+\sqrt{1+4\tilde{s}^2}
=\begin{cases}
2+2\tilde{s}+{\cal O}(2\tilde{s})^2 \qquad & 2\tilde{s}<1,\\
1+4\tilde{s}+{\cal O}\left(\frac{1}{2\tilde{s}}\right) \qquad & 2\tilde{s}>1,
\end{cases}
\end{equation}
which corresponds to right-angled triangles.
The energy diverges along the line $\tilde{p}\geq2$, $\tilde s=0$,
which corresponds to the two sides of the wedge coinciding
($\alpha=\beta$). In FIG.~\ref{3dspTri-wed} the curve
$\tilde{p}=1+2\tilde{s}+\sqrt{1+4\tilde{s}^2}$ for $s\geq 0$
corresponds to right triangles of minimal energy, and the boundary of
the parameter space at $\tilde p =2\tilde{s}+2\sqrt{1+\tilde{s}^2}$
for $\tilde s \geq0$ is associated with isosceles triangles.
We do not observe that ${\cal E}(\tilde s,\tilde p)$ is a function of
$\tilde p^2/\tilde s$ only.
The rather good approximation obtained by ignoring the
dependence on the integral in Eq.~(\ref{E123sp=tri-wed}),
suggests that the energy approximately is given by
\begin{equation}
{\cal E}(\tilde s,\tilde p) \sim |\tilde{\mu}^2 -\tilde{s}^2|
=\frac{ \big| 4(\tilde p-2\tilde s)^2 -\tilde p^2 (\tilde p-4\tilde s)^2 \big|
}{4\tilde p (\tilde p -4\tilde s)},
\end{equation}
a somewhat involved function of the perimeter and area.
\begin{center}
\begin{figure}
\includegraphics[width=7.0cm]{figures/3Dplot-E123-versus-area-and-perimeter-for-triangular-wedge-on-Dplate.eps-from-jpeg}
\hspace{10mm}
\includegraphics[width=7.5cm]{figures/area-peri.eps}%
\caption{Irreducible three-body Casimir energy of a semiweak
triangular waveguide as a function of the cross-sectional area
and perimeter. This energy vanishes and is minimal along the line
$\tilde{p}=1+2\tilde{s}+\sqrt{1+4\tilde{s}^2}$.}
\label{3dspTri-wed}
\end{figure}
\end{center}
\section{Parabolic-wedge on a Dirichlet plate}
\label{par-wedge:s}
To explicitly verify that not just corner divergences have been subtracted
in the irreducible three-body contribution to the
vacuum energy~\cite{Schaden:2010wv},
we also consider a weakly interacting parabolic-wedge
on a Dirichlet plate. It is described by the potentials
\begin{subequations}
\begin{eqnarray}
V_1(y,z) &=& \lambda_1 \delta (-z+\alpha(y-a)^2) \,\theta_1,
\quad \text{with} \quad
\theta_1\equiv\theta(y-\text{min}[0,a])\,\theta(\text{max}[0,a]-y), \\
V_2(y,z) &=& \lambda_2 \delta (-z+\beta(y-b)^2) \,\theta_2,
\quad \text{with} \quad
\theta_2\equiv\theta(y-\text{min}[0,b])\,\theta(\text{max}[0,b]-y),\\
V_3(z) &=& \lambda_3 \delta (z), \quad \text{with} \quad \lambda_3\to\infty.
\end{eqnarray}
\label{parwed-pot-w}%
\end{subequations}%
The parameters $\alpha$ and $\beta$ here give the foci of the parabolas
and have dimensions of inverse length.
The constraint $\alpha a^2=\beta b^2=h$ implies that the two parabolas
intersect at $(y=0,z=h)$. See FIG.~\ref{parwedoTri}.
As in the case of the wave guide with triangular cross-section,
the base has length $|b-a|$ and the height of the wedge above the
Dirichlet plate is $h$. The parameter regions are:
$-\infty<a,b<\infty$, or, equivalently, $0\leq\alpha,\beta<\infty$.
We measure lengths in multiples of $h$ and
use parameters $(h,\tilde a=a/h,\tilde b=b/h)$ to describe it.
\begin{center}
\begin{figure}
\includegraphics{figures/par-wedge.eps}%
\caption{Weakly interacting parabolic-wedge on a Dirichlet plate.}
\label{parwedoTri}
\end{figure}
\end{center}
We proceed exactly as for the triangular-wedge and find that the
three-body Casimir energy of a parabolic-wedge is also given by
Eq.~(\ref{E123=tri-wed}), except that the distances now are given by
\begin{subequations}
\begin{eqnarray}
\bar{u}_{12}^2 &=& (\tilde{a} u_1-\tilde{b} u_2 )^2+[(1-u_1)^2 +(1-u_2)^2]^2,\\
u_{12}^2 &=& (\tilde{a} u_1-\tilde{b} u_2)^2+[(1-u_1)^2 -(1-u_2)^2]^2.
\end{eqnarray}
\end{subequations}
The three-body Casimir energy of a parabolic-wedge on a Dirichlet plate also
is minimized when either $\tilde{a}=0, \alpha=\infty$,
or $\tilde{b}=0, \beta=\infty$. Due to the constraint,
$\alpha a^2=\beta b^2=h$, the shorter side of the parabolic wedge in this
case degenerates to a straight line perpendicular to the Dirichlet plate.
Most of the analysis of the waveguide with two sides of parabolic
cross-section is the same as for a
triangular one with only minor changes in interpretation.
We note that the rescaled area and perimeter of the parabolic wedge are
\begin{subequations}
\begin{eqnarray}
\tilde{s}&=&\frac{1}{3}|\tilde{b}-\tilde{a}|,\label{par-s=ab} \\
\tilde{p} &=& |\tilde{b}-\tilde{a}| +\frac{\tilde{a}}{2}
\bigg[ \frac{\tilde{a}^2}{2}\sinh^{-1}\frac{2}{\tilde{a}^2}
+ \sqrt{1+\frac{4}{\tilde{a}^4}} \bigg]
+ \frac{\tilde{b}}{2}
\bigg[ \frac{\tilde{b}^2}{2}\sinh^{-1}\frac{2}{\tilde{b}^2}
+ \sqrt{1+\frac{4}{\tilde{b}^4}} \bigg].%
\label{par-p=ab}
\end{eqnarray}%
\label{par-sp-eqns}%
\end{subequations}%
The three-body Casimir energy of a semiweak parabolic-wedge is
shown in FIG.~\ref{par-area-fix-fig}.
The approximation of replacing the integral by unity
is not very accurate in this case, but the overall dependence of the
irreducible three-body energy of a parabolic waveguide on the parameters
$\tilde a$ and $\tilde b$ is qualitatively similar to that of a
triangular one.
\begin{center}
\begin{figure}
\includegraphics{figures/e-versus-aoverh-for-constant-area-par-wed.eps}
\caption{Irreducible three-body Casimir energy for a parabolic waveguide
of given cross-sectional area. See FIG.~\ref{area-fix-fig} for description.}
\label{par-area-fix-fig}
\end{figure}
\end{center}
\section{Discussion}
In Casimir studies one generally is interested in dependence of the
vacuum energy of massless quantum fields on the presence of objects
whose interaction with the quantum fields is treated semiclassically,
with quantum fluctuations of the fields describing the objects themselves
being disregarded. This leads to an effective action with ultraviolet
divergent contributions associated with geometrical properties of the
objects reflected by the coefficients~\cite{weyl,Minakshisundaram:1949xg}
in the asymptotic expansion of the heat
kernel~\cite{Kac:1966xd,Fulling:2003zx,kirsten2002spectral}.
The corresponding ultra-violet divergences in the vacuum energy are
proportional to the spatial volume, surface areas, curvatures,
as well as number and type of corners or intersections of the objects.
They depend only on \emph{local} geometric properties of the system.
Fortunately, there also are non-local contributions to the vacuum
energy that do not depend on the high energy description of the
model and can be reliably obtained in semiclassical
approximation. The best known of these is the force between disjoint
classical objects due to vacuum fluctuations, first obtained
for parallel metallic plates by Casimir~\cite{Casimir:1948dh}. It has
since been shown that this force is always finite~\cite{Kenneth:2006vr}
and that the associated finite contribution to the vacuum energy may be
computed for arbitrary objects in terms of the single-body scattering
matrices with Eq.~(\ref{trloggen}).
The investigation of generalized pistons
in~\cite{Schaden:PhysRevLett.102.060402,Schaden:PhysRevA.79.052105}
suggested that one may isolate finite parts of the vacuum energy
that describe the forces between objects even if these are not
mutually disjoint. These ideas were formalized in \cite{Schaden:2010wv}
where irreducible N-body contributions to the vacuum energy were
defined recursively and shown to be finite unless the N-bodies have
a common intersection. For a scalar field whose interaction with N-objects
are semiclassically described by positive local potentials,
the irreducible contribution to the vacuum energy furthermore
was found to be positive for an odd, and negative for an even,
number of objects.
We have here put these general considerations on a much more
practical and concrete footing and developed a formalism to extract
and compute irreducible $N$-body contributions from the single-body
transition matrices. Starting from Faddeev's equations in Eq.~(\ref{fadeqn}),
the irreducible parts of the $N$-body scattering matrix were extracted
recursively. We used this formalism to compute several examples
of irreducible two--and three-body Casimir energies. All our two-body
results have been obtained previously, but we were able to reproduce
some of them in a much simpler and direct manner.
Our three-body results for irreducible Casimir energies are new.
The irreducible three-body contributions
to the Casimir energy of parallel semitransparent plates was obtained
analytically and indeed remains finite when two of the three plates
overlap. We showed explicitly how the irreducible three-body contribution
precisely cancels the irreducible two-body Casimir energy of the outer
plates when Dirichlet boundary conditions are imposed on the central
plate--providing a raison-d'{\^e}tre for both, the existence, and sign,
of the three-body contribution to the force. For semitransparent plates
the cancellation is not complete but can be sizable.
In Section \ref{3body-sweak:s} we analyze
the irreducible three-body interaction in semiweak approximation.
In this approximation we are able to compute the irreducible
three-body Casimir energy for objects that are not mutually disjoint
and whose irreducible two-body contributions diverge. The irreducible
three-body contributions to the vacuum energy of a waveguide constructed by
placing a weakly interacting triangular--or parabolic-wedge on top of
a Dirichlet plate was found to be finite and computable without intermediate
regularization.
Our examples demonstrate that not only corner divergences, but also
divergences related to curvature are subtracted by this
procedure. We also explicitly verified that the irreducible
three-body contribution to the vacuum energy of a massless scalar
field is positive.
To develop a better understanding in a non-perturbative setting,
we are currently investigating the irreducible three-body vacuum energy
of a triangular waveguide formed by imposing Dirichlet boundary conditions
on three intersecting infinite planes (the geometry is similar
to that of FIG.~\ref{wedoTri}, but with sides of infinite extent).
In the limit of an extremely flat triangular cross-section, we intend
to compare the numerical results with analytic calculations. We
further wish to extend these methods to the physically relevant
electromagnetic case. Although irreducible three-body contributions
to the vacuum energy are expected to remain finite, we so far have
no rigorous statements about their sign for vector fields. Interestingly,
it is at least conceptually feasible to directly measure irreducible
electromagnetic three-body contributions to the vacuum energy by
balancing off irreducible two-body parts. We are investigating whether
this is experimentally feasible.
\begin{acknowledgments}
KVS would like to thank Prachi Parashar for useful comments and suggestions
at various stages of this project.
This work was supported by the National Science Foundation with
Grant no.~PHY0555580.
\end{acknowledgments}
|
\section{Introduction}
One of the missions of the Large Hadron Collider (LHC) is to piece together
some of nature's original symmetries. The search for these symmetries is a part of
the ultimate quest to unify all of the particles and forces within a grand unified theory
that exhibits overarching gauge symmetries. Theoretical clues to the original state
of symmetry may be present in conserved or nearly conserved quantities in nature
today. As remnants of symmetry breaking, extra gauge bosons exist in many models
of new physics (NP) that go beyond the standard model (SM). The discovery of new
neutral and charged gauge bosons and the establishment of their quantum numbers
would shed light on the gauge structure of NP~\cite{Robinett:1981yz,
Langacker:1984dc,Barger:1986hd,London:1986dk,Rosner:1986cv,Marciano:1990dp,
Rosner:1995ft,Rosner:1996eb, Carena:2004xs,
Rizzo:2007xs,Petriello:2008zr,Langacker:2008yv,Chiang:2009kb,delAguila:2009gz,
Diener:2010sy,Frank:2010cj,Grojean:2011vu}.
One salient property of new gauge bosons is the handedness of their couplings to
SM fermions, whether dominantly left-handed as the SM $W$ and $Z$ vector bosons
or possibly with large right-handed couplings. In this paper we focus on new
color-singlet $W^{\prime}$ and $Z^{\prime}$ production at the LHC and their decays
into the third generation SM fermions $t,b$. We explore quantitatively the measurement
of the chirality of the couplings of the new gauge bosons from the polarization of the top
quarks in their decays.%
\footnote{The top quark polarization can also be used to probe new gauge bosons
and scalars in exotic color representation such as sextet and anti-triplet;
see Ref.~\cite{Berger:2010fy,Zhang:2010kr} for details.}
The top quark is the only ``bare'' quark whose spin information
can be measured from its decay products since the decay proceeds promptly via the
weak interaction. Among the top quark decay products, the
charged lepton from $t \rightarrow b \ell \nu$ is the best analyzer of the top quark spin.
In the helicity basis, the polarization of the top quark can be determined from the distribution
in $\theta_\ell$, the angle of the lepton in the rest frame of top quark relative to
the top quark direction of motion in the overall center-of-mass (cm) frame. The
angular correlation of the lepton $\ell^+$ is ${1\over 2}(1\pm\cos\theta_\ell)$,
with the ($+$) choice for right-handed and ($-$) for left-handed top-quarks~\cite{Jezabek:1994zv,Mahlon:1995zn}.
In addition to the matter of handedness of couplings, there are other reasons to search for
the $Z^\prime/W^\prime$ in $t\bar{t}$ and $tb$
events. One is that searches in the leptonic decay modes would fail in the so-called
leptophobic models because the $Z^\prime$ and $W^\prime$ bosons in these models
do not couple to leptons. Searches in dijet invariant mass distributions are valuable but
cannot determine whether a dijet resonance is a $Z^\prime$ or $W^\prime$ boson because
the jet charge is not measurable. In such cases, the third generation quarks are necessary
for charge determinations of the heavy resonances; for example, the $Z^{\prime}$ bosons
decay into $t\bar{t}$ and $b\bar{b}$ pairs and the $W^{\prime}$ bosons into $t\bar{b}$
and $\bar{t}b$ pairs.
In Sec.~II, we describe models of new physics that contain extra gauge bosons and show how
patterns of symmetry breaking are manifest in the handedness of the couplings of the new
gauge bosons to SM fermions. We illustrate a few of the NP models from the current literature.
This section also includes a summary of the existing constraints on masses and couplings of new
gauge bosons. In Sec. III, we present $W^{\prime}$ and $Z^{\prime}$ production cross sections
at the LHC, both the inclusive rates and the rates of interest to us with all branching fractions
included. The collider signatures we study are $\ell^+ \slash{\!\!\!\!E}_T b\bar{b}$ for the $W^{\prime +}$ and the semileptonic decay of $t\bar{t}$, namely $\ell^\pm \slash{\!\!\!\!E}_T j j b \bar{b}$, for the $Z^\prime$.
The missing energy $\slash{\!\!\!\!E}_T$ is carried off by a neutrino in the top quark decay.
The dominant backgrounds are also computed and assessments are presented for
the $W^{\prime}$ and $Z^{\prime}$ discovery potential.
After we impose kinematic cuts and reconstruct the final states, we conclude that a $Z^\prime$ resonance with mass $1$ TeV could be seen above the SM $t\bar{t}$ background with a
statistical significance more than $5$ standard deviations ($5 \sigma$) for $10~{\text {fb}}^{-1}$ integrated luminosity
at $14$ TeV, provided its coupling $g_V \equiv \sqrt {g_L^2 + g_R^2}$ to the SM quarks is about
$0.4$, consistent with bounds from Tevatron searches in the dijet final state. The $W^\prime$
signal can be much larger than the SM background if a coupling strength $0.4$ to the
SM quarks is assumed. For purposes of comparison, the couplings of the SM $W$ boson to SM quarks are
$g_L^{Wu\bar{d}}=0.461$ and those of the SM $Z$ boson are
$g_L^{Zu\bar{u}}=-0.257$, $g_R^{Zu\bar{u}}=0.115$, $g_L^{Zd\bar{d}}=0.314$ and $g_R^{Zd\bar{d}}=-0.057$.
Section~IV is devoted to the measurement of the top quark polarization and the determination of
the handedness of the new gauge bosons. We apply our approach to three benchmark models,
the sequential SM-like $W^\prime/Z^\prime$ model (SSM), the top-flavor model, and the left-right symmetric model (LRM).
These models provide different predictions for the left-handed fraction of the coupling strengths of
the new gauge bosons. We show that the coupling of a $W^\prime$ to $tb$ can be determined precisely,
whereas the uncertainty is relatively large for a $Z^\prime$ to $t\bar{t}$, owing mainly
to better statistics and smaller SM backgrounds in the $W^\prime$ case.
For $m_{W^\prime} ~(\simeq m_{Z^\prime}) \sim 1$ TeV, our determinations of the handedness of the
$W^\prime$ and $Z^\prime$ couplings allow the three benchmark models be separated to varying
degrees with $100~{\text{fb}}^{-1}$ of accumulated data.
With this large data sample, one can distinguish the $Z^\prime$ models if the central values of
their top quark polarizations differ by $\sim 20\%$.
For a leptophobic $Z^\prime$, one can differentiate among different models if the
difference of the handedness of the coupling to SM quarks is $\gtrsim 10\%$ (for coupling
strength $\sim 0.4$)).
Our overall summary is found in Sec.~V.
\section{Models with extra gauge bosons}
\label{models}
\begin{figure}
\begin{center}
\includegraphics[scale=0.5,clip]{figure/221.eps}
\caption{Pictorial illustration of symmetry breaking patterns of $G(221)$ model.
\label{fig:221}}
\end{center}
\end{figure}
Extra gauge bosons may be classified according to their
electromagnetic charges: $W^\prime$ (charged bosons) and $Z^\prime$ (neutral bosons).
While a $Z^\prime$ can originate from an additional abelian $U(1)$ group,
a $W^\prime$ arises often in models with an extra non-abelian group.
In this section we consider the so-called $G(221)$ model~\cite{Hsieh:2010zr} which carries the
simplest non-abelian extension to the SM
\begin{equation}
G(221) = SU(2)_1 \otimes SU(2)_2 \otimes U(1)_X.
\end{equation}
The model represents a typical gauge structure of many interesting NP models
such as the non-universal model (NU)~\cite{Li:1981nk,Malkawi:1996fs,He:2002ha},
the ununified (UU) model~\cite{Georgi:1989ic, Georgi:1989xz},
the fermiophobic (FP) model~\cite{Barger:1980ix},
left-right (LR) models~\cite{Senjanovic:1975rk}, and so forth.
Both $W^\prime$ and $Z^\prime$ bosons appear after the $G(221)$ symmetry is
broken to the SM symmetry $G_{SM}=SU(2)_L\otimes U(1)_Y$.
As depicted in Fig.~\ref{fig:221}, these models can be categorized by two symmetry breaking patterns,
\begin{itemize}
\item[(a)] In the UU and NU models:\\
$U(1)_X$ is identified as the $U(1)_Y$ of the SM. The first stage of symmetry breaking
$SU(2)_{1}\times SU(2)_{2}\to SU(2)_{L}$ occurs at the TeV scale, while the second stage of
symmetry breaking $SU(2)_{L}\times U(1)_Y \to U(1)_{em}$ occurs at the electroweak scale;
\item[(b)] In the FP and LR models: \\
$SU(1)_1$ is identified as the $SU(2)_L$ of the SM. The first stage of symmetry breaking
$SU(2)_2 \times U(1)_{X}\to U(1)_{Y}$ occurs at the TeV scale, while the second stage of
symmetry breaking $SU(2)_{L}\times U(1)_Y \to U(1)_{em}$ occurs at the electroweak scale.
\end{itemize}
In the first pattern the couplings of new gauge bosons to the SM fermions
are predominately left-handed while in the second pattern the couplings are
right-handed. A measurement of the polarization of the fermions
from the new gauge bosons would decipher the handedness of the couplings.
Rather than focusing on a specific model, we explore the discovery potential
of $W^\prime$ and $Z^\prime$ bosons in a model independent method, and we comment
on a few new physics models later.
The most general interaction of the $Z^\prime$ and $W^\prime$ to the SM quarks is
\begin{eqnarray}
\mathcal{L} &=& \bar{q}\gamma^{\mu}(g^{Z^\prime}_L P_L + g^{Z^\prime}_R P_R) q~Z^{\prime}_{\mu} \nonumber \\
&+& \bar{q}\gamma^{\mu}(g^{W^\prime}_L P_L + g^{W^\prime}_R P_R) q^\prime~W^{\prime+}_{\mu} + h.c.
\label{eq:int}
\end{eqnarray}
where $P_{L/R}$ is the usual left- and right-handed projector and $q$ denotes the SM quarks.
The $Z^\prime$ and $W^\prime$ are understood here to be color singlet states,
but one can easily obtain the interaction of color octet bosons,
such as a $G^\prime$, from insertion of the $SU(3)_C$ color matrices $\lambda_A/2$
in Eq.~(\ref{eq:int}).
The couplings $g_{L/R}^{Z^\prime}$ and $g_{L/R}^{W^\prime}$ usually are not independent
when the $W^\prime$ and $Z^\prime$ originate from the same gauge group.
For example, in the left-right model, the SM right-handed quark singlets form a doublet $(u_R,d_R)$
which is gauged under the additional $SU(2)_R$ group. The $W^\prime$ and $Z^\prime$ transform as
a $SU(2)_R$ triplet and their couplings to the SM quarks are correlated.
In this work we first treat $g^{W^\prime}_{L/R}$ and $g^{Z^\prime}_{L/R}$ as independent in our
collider simulation to derive the experimental sensitivity on $Z^\prime$ and $W^\prime$ measurements.
We then consider the correlation between the two couplings in the context of some NP models.
We use $g_{L/R}$ to denote the left-handed and right-handed couplings of the $W^\prime$ and
$Z^\prime$ to the SM quarks.
For simplicity we assume the couplings of $Z^\prime$ to up- and down-type quarks are the same.
For illustration we study three benchmark NP models in this work:
\begin{itemize}
\item sequential SM-like $W^\prime/Z^\prime$ (SSM) model: the $W^\prime$ and $Z^\prime$
couplings to SM fermions
are exactly the same as the SM $W$ and $Z$ boson, and $m_{W^\prime}=m_{Z^\prime}$. Although it is difficult in a realistic model to have couplings which are the same as in the SM, we show such a case for comparison.
\item top-flavor model~\cite{Malkawi:1996fs}: the $W^\prime$ and $Z^\prime$ couplings are purely left-handed, and $m_{W^\prime}=m_{Z^\prime}$.
\item left-right symmetric model (LRM)~\cite{Senjanovic:1975rk}: Here, we consider a $SU(2)_L\times SU(2)_R\times U(1)_{B-L}$ model. The $W^\prime$ couplings to SM quarks are
purely right-handed while the $Z^\prime$ couplings are dominantly right-handed.
\end{itemize}
Table~\ref{table:coup} is a summary of the couplings of a $W^\prime$ and a $Z^\prime$ to SM third generation quarks in these models. The couplings to the quarks of first two generations are the same, except that one should replace $\sin\tilde{\phi}$ by $\cos\tilde{\phi}$ in the top-flavor model.
\begin{table}
\caption{Couplings of a $W^\prime$ to $tb$ and a $Z^\prime$ to $t\bar{t}$ for the sequential SM-like $W^\prime/Z^\prime$ (SSM) model, the left-right symmetric model (LRM), and the top-flavor model, where $s_w~(c_w, ~t_w) =\sin\theta_w~(\cos\theta_w,~\tan\theta_w)$, $\theta_w$ is the weak mixing angle, $g_2=e/{s_w}$ is the weak coupling,
$\alpha_{LR} \simeq 1.6$, and $\sin\tilde\phi$ is taken to be $1/\sqrt{2}$.\label{table:coup}}
\begin{tabular}{lll}
\hline
& $W^{\prime}tb$ & $Z^{\prime}t\bar{t}$\tabularnewline
\hline
SSM & $ \displaystyle{\frac{g_2}{\sqrt{2}}}\bar{b}\gamma_\mu P_L t W^{\prime\mu} $
& $ \displaystyle{\frac{g_2}{6c_w}}\bar{t}\gamma_\mu( (-3+4s_w^2) P_L + 4s_w^2 P_R ) t Z^{\prime\mu} $
\vspace*{3mm}
\tabularnewline
LRM & $ \displaystyle{\frac{g_2}{\sqrt{2}}} \bar{b}\gamma_\mu P_R t W^{\prime\mu}$
& $ \displaystyle{\frac{g_2t_w}{6}}\bar{t}\gamma_\mu( \frac{1}{\alpha_{LR}} P_L + (\frac{1}{\alpha_{LR}}-3\alpha_{LR}) P_R ) t Z^{\prime\mu}$
\vspace*{3mm}
\tabularnewline
Top-Flavor & $\displaystyle{\frac{g_2\sin\tilde{\phi}}{\sqrt{2}}} \bar{b}\gamma_\mu P_L t W^{\prime\mu} $
& $\displaystyle{\frac{g_2\sin\tilde{\phi}}{\sqrt{2}}}\bar{t}\gamma_\mu P_L t Z^{\prime\mu}$
\tabularnewline
\hline
\end{tabular}
\end{table}%
\subsection{Bounds on masses and couplings}
The masses and couplings of $Z^\prime$ and $W^\prime$ bosons are bounded
by various low energy measurements (mainly via the four-fermion
operators induced by exchanges of new heavy gauge bosons) such as the precision
measurements at the $Z$-pole at LEP-I~\cite{Nakamura:2010zzi}, the $W$-boson mass~\cite{Nakamura:2010zzi},
the forward-backward asymmetry in $b\bar{b}$ production at
LEP-II~\cite{Nakamura:2010zzi},
$\nu e$ scattering~\cite{Vilain:1996yf},
atomic parity violation~\cite{Vetter:1995vf,Wood:1997zq,Guena:2004sq},
Moller scattering~\cite{Anthony:2005pm}, and so forth.
The bounds are severe when new gauge bosons couple to leptons directly, but they can
be relaxed for a leptophobic model, as analyzed in Ref.~\cite{Hsieh:2010zr}.
Tevatron data place a lower bound about $1.1~\rm{TeV}$ on the mass of a $W^\prime$~\cite{Aaltonen:2010jj} and about $1.07~\rm{TeV}$ for a $Z^\prime$~\cite{Aaltonen:2011gp}, based on the charged lepton plus missing energy ($\ell^\pm {\not \! \!E}_T$) and $\mu^+\mu^-$ final states, respectively,
with the assumption that the couplings between the $W^\prime/Z^\prime$ and the SM fermions are
the same as those in the SM. Searches for the $W^\prime$ and $Z^\prime$ at the Tevatron in
dijet events yield lower bounds on $m_{W^\prime}$ and $m_{Z^\prime}$, assuming SM couplings, of $840$ GeV and $740$ GeV, respectively~\cite{Aaltonen:2008dn}.
Recent CMS and ATLAS collaboration searches for a $Z^\prime$ from dilepton final states and a $W^\prime$ from lepton plus missing energy events place lower bounds $m_{Z^\prime} > 1.14$ TeV
(CMS)~\cite{Collaboration:2011wq}, along with $m_{W^\prime} > 1.58$ TeV (CMS)~\cite{Collaboration:2011dx} and
$m_{W^\prime} > 1.49$ TeV (ATLAS)~\cite{Collaboration:2011fe}. The analyses assume the
$Z^\prime$ and $W^\prime$ have sequential standard model couplings.
CMS and ATLAS also present searches for a resonance in dijet events which also constrain the masses of a $W^\prime$ and a $Z^\prime$, but the lower bounds are looser than the Tevatron results~\cite{:2010bc,Khachatryan:2010jd}.
If the couplings between the $W^\prime/Z^\prime$ and the SM particles are not small, one can expect the discovery of the these heavy resonances sooner or later at the LHC.
Negative searches for a $W^\prime/Z^\prime$ through the $t\bar{b}$ and $t\bar{t}$ final states at the Tevatron impose upper bounds on the production cross section times decay branching ratio ($\sigma_{W^\prime} Br(W^\prime\to t\bar{b}),~\sigma_{Z^\prime} Br(Z^\prime\to t\bar{t})$) for masses up to
$950$ GeV and $900$ GeV for $W^\prime$ and $Z^\prime$, respectively~\cite{cdftt,cdftb}.
\section{LHC phenomenology}
We divide our discussion of $W^{\prime\pm}$ and $Z^\prime$ phenomenology into
two parts. In this section, we present our evaluation of the production cross sections and discuss
the pertinent backgrounds with a view toward understanding the discovery potential of the two states. Section~IV is then devoted to an examination of top quark polarization measurements as a means to learn more about the models that produce $W^{\prime}$ and $Z^\prime$ bosons.
The $Z^\prime$ production cross section is
\begin{equation}
\sigma_{Z^\prime}(\hat{s})=\frac{\beta}{192\pi}
\frac{\hat{s}}{(\hat{s}-m_{Z^\prime}^2)^2+m^2_{Z^\prime}\Gamma^2_{Z^\prime}}
\left[(g_L^2 + g_R^2)^2(3+\beta^2)+ 6g_Lg_R(g_L^2+g_R^2)(1-\beta^2)\right],
\end{equation}
where $\beta=\sqrt{1-4m_t^2/\hat{s}}$.
The $W^\prime$ production cross section is
\begin{equation}
\sigma_{W^\prime}(\hat{s})=\frac{(1-x_t^2)^2}{96\pi}
\frac{\hat{s}}{(\hat{s}-m_{W^\prime}^2)^2+m^2_{W^\prime}\Gamma^2_{W^\prime}}
(g_L^2 + g_R^2)^2\left(2+x_t^2\right),
\end{equation}
where $x_t=m_t/\sqrt{\hat{s}}$.
The partial decay width of $V^\prime \to q \bar{q}^\prime$ ($V^\prime=W^\prime,~Z^\prime$) is
\begin{equation}
\Gamma(V^\prime \to q\bar{q}^\prime) =
\frac{m_{V^\prime}}{8\pi} \beta_0
\left[(g_L^2 + g_R^2) \beta_1 + 6 g_L g_R \frac{m_q m_{q^\prime}}{m^2_{V^\prime}}\right],
\end{equation}
where
\begin{equation}
\beta_0 = \sqrt{1 - 2\frac{m_q^2+m^2_{q^\prime}}{m^2_{V^\prime}}+
\frac{(m_q^2-m_{q^\prime}^2)^2}{m^4_{V^\prime}}},
\qquad
\beta_1 = 1 - \frac{m^2_q + m^2_{q^\prime}}{2m^2_{V^\prime}}
-\frac{(m^2_q-m^2_{q^\prime})^2}{2m_{V^\prime}^4}.
\end{equation}
Evaluations of the cross sections are presented
in Fig.~\ref{fig:lhc} at the LHC with center of mass energy 14~TeV.
In the mass range of interest to us the $W^\prime$ and $Z^\prime$ bosons are
much heavier than the top quark, and $m_t$ can be ignored.
Hence, the decay branching ratio of $Z^\prime \to t\bar{t}$ is about 1/6 while
the ratio of $W^\prime \to t \bar{b}$ is 1/3.
The cross sections for other values of $g_L$ and $g_R$ can be obtained from
the curves in Fig.~\ref{fig:lhc} by a simple scaling, $\sigma_{V^\prime} \propto (g_L^2 + g_R^2)$.
\begin{figure}[t]
\begin{center}
\includegraphics[scale=0.7,clip]{figure/prod_lhc14.eps}
\caption{Production cross sections of the $W^\prime$ and $Z^\prime$ at 14~TeV
for the choice $g_L=0.4$ and $g_R=0$.
The solid curves provide results before decay branching fractions are included, whereas the dashed
curves include branching fractions.
\label{fig:lhc}}
\end{center}
\end{figure}
We use MadGraph/MadEvent~\cite{Alwall:2007st} to obtain the signal and background distributions.
The widths of $W^\prime$ and $Z^\prime$ for different $g_L$ and $g_R$ couplings are calculated
in BRIDGE~\cite{Meade:2007js}. These results are computed at leading order with
the renormalization scale ($\mu_R$) and factorization scale ($\mu_F$) chosen as
\begin{equation}
\mu_R = \mu_F = \sqrt{m_t^2 + 2 p_T^2(t)}.
\end{equation}
The CTEQ6.1L parton distribution functions (PDFs)~\cite{Pumplin:2002vw} are used.
The coupling strength is set at
$g_V \equiv \sqrt{g_L^2+g_R^2} = 0.4$, which respects the dijet constraints at the Tevatron~\cite{Aaltonen:2008dn}
in the leptophobic $W^\prime/Z^\prime$ models with a universal coupling.
We plot the production cross sections with the choice of $g_L=0.4,~g_R=0$ in the figure.
The curves shown in Fig.~\ref{fig:lhc} also represent the cross sections for
$g_R=0.4,~g_L=0$.
After including branching fractions, we also plot the cross
sections for $pp\to Z^\prime X \to t\bar{t} X$, $pp\to W^{\prime+} X\to t\bar{b} X$, and
$pp\to W^{\prime-} X\to \bar{t}bX$. Universal couplings of the $W^\prime/Z^\prime$ bosons to
three generation of quarks are understood. Because the $u$-quark parton density in the proton
is large, it is easy to understand that the $Z^\prime$ has the largest production cross section while
the $W^{\prime-}$ has the smallest one.
We explore the $W^\prime$ in the $t\bar{b}$ decay mode, and the $Z^\prime$ in the $t\bar{t}$ mode.
To be able to measure the top quark polarization, we focus on final states in which the top quark
decays leptonically. Therefore, the $W^\prime$ search is via the channel
$p p \to W^\prime X\to t b X\to \ell^\pm \nu b \bar{b} X$. The $Z^\prime$ search is done in the
channel $p p \to Z^\prime X\to t \bar{t} X\to \ell^\pm \nu j j b \bar{b} X$. We consider semileptonic decay
of only one of the top quarks in the $t\bar{t}$ mode of $Z^\prime$ decay because of its large branching ratio.
One can also use the dilepton channel to search for a $Z^\prime$ using the MT2-assisted method
discussed in Ref.~\cite{Berger:2010fy,Zhang:2010kr} to fully reconstruct the $Z^\prime$.
\subsection{$Z^\prime$ discovery potential}
\label{sec:lhc_z}
The collider signature of interest to us for the $Z^\prime$ boson is $\ell^\pm\nu jj b\bar{b}$, where one top quark decays semileptonically and the other decays hadronically. We consider in this subsection only $\mu^+$ final states, $t \to b \mu^+ \nu_\mu$, but the statistics will increase when the different flavors and charges of the leptons are combined. The major SM background is $t\bar{t}$ production
via the QCD interaction. The sum of other backgrounds, such as $W/Z(\to \ell\ell)+$jets, single top, $W+b\bar{b}$, etc., is a factor of $\gtrsim 20$ smaller after the usual semileptonic $t\bar{t}$ selection
cuts~\cite{Aad:2009wy}.
At the analysis level, all signal and background events are required to pass the acceptance cuts listed here:
\begin{eqnarray}
p_T(\ell,j) > 20~{\rm GeV}, &\quad |\eta(\ell, j)|<2.5,&\quad \Delta R_{jj,j\ell} > 0.4,\nonumber \\
\slash{\!\!\!\!E}_T >30~{\rm GeV}, &\quad H_T > 500~{\rm GeV}, &\quad M_T > 800~{\rm GeV},
\end{eqnarray}
where $p_T$($\eta$, $\slash{\!\!\!\!E}_T$) denotes the transverse momentum (rapidity, missing transverse momentum),
$\Delta R_{kl}\equiv\sqrt{\left(\eta_{k}-\eta_{l}\right)^{2}+\left(\phi_{k}-\phi_{l}\right)^{2}}$
is the separation in the azimuthal angle ($\phi$)-pseudorapidity
($\eta$) plane between the objects $k$ and $l$,
$H_T$ is the scalar sum of the transverse momenta of the final state visible particles plus
$\slash{\!\!\!\!E}_T$, and $M_T$ is the cluster transverse
mass defined as $M_T \equiv \sqrt{\sum_{i=j,\ell} p_i^2+\slash{\!\!\!\!E}_T^2} + \slash{\!\!\!\!E}_T $.
We model detector resolution effects by smearing the final state energy according to
$\delta E/E= \mathcal{A}/\sqrt{E/{\rm GeV}}\oplus \mathcal{B}$,
where we take $\mathcal{A}=10(50)\%$ and $\mathcal{B}=0.7(3)\%$ for leptons (jets). To
account for $b$-jet tagging efficiencies, we demand two $b$-tagged jets, each with a tagging efficiency of $60\%$.
We consider two masses $m_{Z^\prime} =1$TeV and $1.5$TeV and the coupling strength
$g_{Z^\prime}=0.4$ to satisfy the bounds on heavy resonance searches in the dijet channel
at the Tevatron. The cross sections for $t\bar{t}$ before and after the cuts are shown in
Table~\ref{table:xszprime}.
\begin{table}
\caption{Cross sections (in fb) for the signal process $pp\to Z^\prime \to t\bar{t} \to \mu^+ \slash{\!\!\!\!E}_T j j b \bar{b} $ and the SM backgrounds at 14~TeV.
Two $b$-jets are tagged. ``With cuts" means the cross sections after all of the kinematic cuts, b-tagging and reconstruction. The universal coupling of the $Z^\prime$ to the SM quarks is set to be $0.4$. The mass window cuts are $\Delta M = 150~(200)$ GeV for a $1~(1.5)$~TeV $Z^\prime$ resonance, respectively.}
\begin{tabular}{c|ccc|ccc|ccc}
\hline
& & $Z_R^\prime$& & &$Z_L^\prime$& & &Background & \tabularnewline
\hline
$M_{Z^\prime}$& No cut & With cuts & $\Delta M$ & No cut & With cuts & $\Delta M$ & No cut & With cuts & $\Delta M$ \tabularnewline
\hline
1~TeV & 275.6 & 29.5 & 28.3 & 275.6 & 27.2 & 26.2 & $3.75\times 10^4$ & 133.1 & 87.0 \tabularnewline
1.5~TeV & 51.4 & 3.0 & 2.6 & 52.5 & 3.9 & 3.5 & $3.75\times 10^4$ & 133.1 & 11.3 \tabularnewline
\hline
\end{tabular}
\label{table:xszprime}
\end{table}%
Since there is only one neutrino in the final state, the undetected $z$ component of the neutrino momentum can be reconstructed from the on-shell condition of the $W$-boson. This procedure
leads to a two-fold solution, but the ambiguity can be removed by the top quark on-shell condition
$m_{b\ell\nu}^2 = m_t^2$, where $m_t=173.3$ GeV~\cite{tevatron:1900yx} is used. If no real solution is obtained, we use the top quark on-shell condition first, and then the $W$-boson on-shell condition
to choose the better solution.
Since the $b$ and $\bar{b}$ are indistinguishable, one must pick which $b$-jet should be
combined with the charged lepton. For this, we use the on-shell condition of the top quark
in the hadronic decay to select one of the $b$-jets to pair with the two jets that are the
decay products of $W$-boson. The efficiency for choosing the correct $b$-jets to reconstruct
the semileptonic and hadronic decays of the top quarks can reach $\gtrsim99.7\%$.
In our simulations, after kinematic cuts and event reconstruction, the $Z^\prime$ resonance peak
can be seen above the SM continuum in the $ t\bar{t}$ invariant mass distribution, $m_{t\bar{t}}$,
especially for a 1 TeV $Z^\prime$ . The results are shown in Fig.~\ref{fig:mtt}. To enhance the discovery significance, we further focus on the events within a mass window around the resonance:
\begin{eqnarray}
|m_{t\bar{t}} - m_{Z^\prime}| \leq 150~(200)~{\rm GeV,~ for~1~(1.5)~TeV~Z^{\prime}}
\end{eqnarray}
\begin{figure}[t]
\begin{center}
\includegraphics[scale=0.45,clip]{figure/mtt_cut.eps}
\caption{Distribution in $m_{t\bar{t}}$ for a $Z^\prime$ and the SM $t\bar{t}$ background at the LHC. The coupling strength $g_{Z^\prime}$ is set to be $0.4$. }
\label{fig:mtt}
\end{center}
\end{figure}
\begin{figure}[t]
\begin{center}
\includegraphics[scale=0.4,clip]{figure/sig_lumi.eps}
\includegraphics[scale=0.4,clip]{figure/zpcontour.eps}
\caption{(a) Discovery potential for a $Z^\prime$ boson at 14~TeV
as a function of the integrated luminosity for a left-handed $Z^\prime$-$t$-$\bar{t}$ coupling
($g_L=0.4$, $g_R=0$) or a right-handed coupling ($g_L=0$, $g_R=0.4$).
(b) Luminosities required for 5 standard deviation ($5\sigma$) and 3 standard deviation ($3\sigma$)
discovery if the $Z^\prime$ mass is 1~TeV, as a function of the coupling strength $g_{L}(g_R)$.
\label{fig:lhc-potential}}
\end{center}
\end{figure}
Figure~\ref{fig:lhc-potential} shows the statistical significance for discovery as a function of
accumulated luminosity for $m_{Z^\prime} = 1~{\rm TeV}$ and $m_{Z^\prime} = 1.5~{\rm TeV}$.
Since the charged lepton from a right-handed top quark decay is boosted to a harder $p_T$,
more signal events survive after cuts in a right-handed $Z^\prime$ model (black-solid curve)
than in a left-handed $Z^\prime$ model (red-solid curve). Therefore, the statistical significance for a right-handed $Z^\prime$ is better when $m_{Z^\prime} = 1~{\rm TeV}$. However, the situation
reverses when the $Z^\prime$ becomes heavier because the selection cuts play a role. For a
heavy enough $Z^\prime$, the top quark is highly boosted. The leptons and jets from its decay are collimated and fail the $\Delta R$ separation cuts, as can be seen in Fig.~\ref{fig:deltaR}. The
peak position in the $\Delta R$ distribution for two light jets shifts down below $0.4$ when the mass of a right-handed $Z^\prime$ increases from $1$ TeV to $1.5$ TeV. As a result, the discovery potential for a very heavy, right-handed $Z^\prime$ is worse than for a left-handed $Z^\prime$, as illustrated in
the black-dashed and red-dashed curves in Fig.~\ref{fig:lhc-potential} for a 1.5~TeV $Z^\prime$.
\begin{figure}[t]
\begin{center}
\includegraphics[scale=0.55,clip]{figure/deltaR.eps}
\caption{(a) $\Delta R$ distributions for the $\mu^+$ and $b$ from $t$ semileptonic decay.
(b) $\Delta R$ distributions for two light jets from $\bar{t}$ hadronic decay. The vertical
dashed line indicates the $\Delta R$ cut applied in our event selection.}
\label{fig:deltaR}
\end{center}
\end{figure}
\subsection{$W^\prime$ discovery potential}
The search for a $W^\prime$ in the $tb$ mode involves the study of the $\ell^\pm b\bar{b}$ plus
missing energy final state, where the missing energy originates from the neutrino in top quark
decay. We examine only the $W^{\prime+}$ case since its production cross section is about a
factor of two larger than $W^{\prime-}$. We demand that two jets are $b$-tagged.
The SM background processes include production of
single top quarks, $Wb\bar{b}$, $Wjj$, and $t\bar{t}$. Only the
single-top background is considered in our study because it is about a factor of $10$ larger
than the others~\cite{Gopalakrishna:2010xm}.
There are three single-$t$ backgrounds: $q\bar{q}^\prime \to t \bar{b}$ (named $t\bar{b}$),
$q b \to q^\prime t$ (named $bq$) and $q g \to q^\prime t \bar{b}$ (called $Wg$-fusion).
The $bq$ channel has only one $b$-jet in the final state, but it is possible that the
other light-favor jet is misidentified as a $b$-jet as well.
The channel has a large cross section ($\sim 150~{\rm pb}$~\cite{Schwienhorst:2010je}).
We also apply a mistagging rate for charm-quarks $\epsilon_{c\to b}=10\%$ for
$p_{T}(c)>50\,{\rm GeV}$.
The mistag rate for a light jet is $\epsilon_{u,d,s,g\to b}= 0.67\%$ for $p_{T}(j)<100\,{\rm GeV}$
and $2\% $ for $p_{T}(j)>250\,{\rm GeV}$.
For $100\,{\rm GeV}<p_{T}\left(j\right)<250\,{\rm GeV}$, we linearly interpolate the fake rates given above.
\begin{table}
\caption{Cross sections (in fb) for the signal process $pp\to W^\prime \to t\bar{b} \to \mu^+ \slash{\!\!\!\!E}_T b \bar{b}$ and the SM backgrounds at 14~TeV. Two $b$-jets are tagged. ``With cuts" refers to cross sections after all of the kinematic cuts, b-tagging, and reconstruction. The value $0.4$ is used for the universal coupling of the $W^\prime$ to the SM quarks. The mass window cuts are
$\Delta M = 150~(200)$ GeV for a $1~(1.5)$~TeV $W^\prime$, respectively.}
\begin{tabular}{c|ccc|ccc|ccc}
\hline
& & $W_R^\prime$& & &$W_L^\prime$& & &Background & \tabularnewline
\hline
$M_{W^\prime}$& No cut & With cuts & $\Delta M$ & No cut & With cuts & $\Delta M$ & No cut & With cuts & $\Delta M$ \tabularnewline
\hline
1~TeV & 652.1 & 109.4 & 105.2 & 650.3 &112.9 & 108.5 & $3.04\times 10^4$ & 412.1 & 8.4 \tabularnewline
1.5~TeV & 131.7 & 23.9 & 22.7 & 129.2 & 26.1 & 24.8 & $3.04\times 10^4 $& 412.1 & 2.0 \tabularnewline
\hline
\end{tabular}
\label{table:xswprime}
\end{table}%
All signal and background events are required to pass the acceptance cuts:
\begin{eqnarray}
p_T(\ell) > 20~{\rm GeV},&\quad |\eta(\ell)|<2.5, &\quad \slash{\!\!\!\!E}_T > 25~{\rm GeV},\nonumber \\
p_T(j)>50~{\rm GeV}, &\quad |\eta(j)|<3.0, &\quad \Delta R(jj)>0.4,\quad \Delta R(j\ell)>0.3~.
\end{eqnarray}
The notation is the same as in the $Z^\prime$ search. The event reconstruction is done in a manner
similar to Ref.~\cite{Gopalakrishna:2010xm}.
A two-fold solution may be obtained for the undetected $z$ component of the momentum of the
missing neutrino if the $W$-boson on-shell condition is used. However, unlike $Z^\prime \to t \bar{t}$, there is not a second top quark in the final state to help in selecting the correct $b$-jet to pair with the charged lepton. Instead, we use the top quark on-shell condition, $m_{b\ell\nu}^2 = m_t^2$ and loop over the two $b$-jets to find the better paring of the $b$-jet and charged lepton, and the better solution
for the neutrino momentum. If no real solution is obtained, we discard the event. After the neutrino
momentum is obtained, we can reconstruct the momenta of the top quark and the $W^\prime$. To further suppress the SM backgrounds, we limit the event set to a mass window around the $W^\prime$
peak position. For a $1$ TeV ($1.5$ TeV) resonance, we adopt
\begin{equation}
\left|m_{t\bar{b}}-m_{W^\prime}\right|\leq 150~(200)~{\rm GeV}.
\end{equation}
After the cuts are imposed and the two b-jets are tagged, the SM backgrounds are at the fb
level while the signal rates are about 100~(20)~fb for $W^\prime$ masses of $1$ ($1.5$) TeV.
The signal and background cross sections are shown in Table~\ref{table:xswprime}.
After imposing the mass window cut and demanding two $b$-tagged jets, we find that
the $Wg$-fusion channel yields the largest background.
For a 1~TeV $W^\prime$, the single-$t$ processes give these background rates:
$Wg$-fusion, 6.3~fb; $bq$, 1.3~fb; $t\bar{b}$, 0.8~fb.
For a 1.5~TeV $W^\prime$, the $Wg$-fusion rate is 1.4~pb, $bq$ is 0.4~fb, and $t\bar{b}$ is 0.16~pb.
The acceptance
for a left-handed $W^\prime$ is slightly better than for a right-handed one. Because the top quark is boosted from $W^\prime$ decay, $\Delta R$ between the charged lepton and the $b$ quark from
the decay of a right-handed top quark is smaller than that from a left-handed top quark, similar to the situation described above for a $1.5$ TeV $Z^\prime$.
The coupling strength $g_L^{W^\prime}=0.4$ or $g_R^{W^\prime}=0.4$ is used in our analysis to satisfy constraints from the Tevatron dijet search.
Since the background is much smaller than the signal rate, $S/B\simeq {\cal O}(10)$, a $W^{\prime+}$ should be easy to discover. Moreover, good accuracy can be obtained for the top quark polarization measurement, as discussed in the next section. %
\section{Top quark polarization}
\label{sec:pol}
The symmetry breaking patterns mentioned in Sec.~II prefer either a purely left-handed
top quark ($SU(2)_1\times SU(2)_2 \to SU(2)_L$) or a purely right-handed top quark
($SU(2)_R\times U_Y^\prime \to U(1)_Y$). We can measure the top quark polarization
from the $\cos\theta$ distribution of the charged lepton in top quark decay after the top
quark kinematics are reconstructed in the $W^{\prime+}$ and $Z^{\prime+}$ final states,
as described in Sec.~III.
Figure~\ref{fig:cth_hel_z} displays the $\cos\theta_\ell$ distributions of the $\mu^+$ in the
rest frame of the top quark in $t\bar{t}$ events for the SM background and for a 1~TeV
$Z^\prime$ boson before and after cuts. A top quark produced at the LHC via QCD
interactions is unpolarized, as shown by the flat black curve in Fig.~\ref{fig:cth_hel_z} (a).
We see the $1\pm\cos\theta_\ell$ behaviors in Fig.~\ref{fig:cth_hel_z}(b) and (c) for purely
right- and left-handed polarized top quarks from $Z^\prime$ decay. After kinematic cuts are
imposed, the distributions are distorted and drop significantly in the region $\cos\theta_\ell \sim -1$, affected mainly by the $p_T$ and $\Delta R$ cuts. However, the main characteristic features remain, i.e. flatness and $1\pm\cos\theta_\ell$. While not shown here, the $\cos\theta_\ell$ distributions in right- and left-handed $W^\prime$ decay are similar to the $Z^\prime$ case.
\begin{figure}[t]
\begin{center}
\includegraphics[scale=0.7,clip]{figure/coscomp.eps}
\caption{Distributions in $\cos\theta_\ell$ of the lepton from the decay of top quarks
produced in $t\bar{t}$ events before and after cuts: (a) SM, (b) right-handed polarized top
quarks in $Z^\prime$ decay; (c) left-handed polarized top quarks in $Z^\prime$ decay.
The distributions for $W^\prime$ decay are similar to those for $Z^\prime$ decay.
\label{fig:cth_hel_z}}
\end{center}
\end{figure}
We denote the distributions from purely left-handed (right-handed) top quarks $F_L(y)$
($F_R(y)$) where $y=\cos\theta_\ell$. These are the distributions in
Fig.~\ref{fig:cth_hel_z} (b) and (c) with cuts. We use these as the basis functions to
fit the event distributions from our simulations of various models. We adopt a general
linear least squares fit in this study to estimate how well the degree of top quark polarization
can be determined.
An observed angular distribution $O(y)$ after the SM background is subtracted can be expressed as
\begin{equation}
O(y) = \epsilon_L~F_L(y) + \epsilon_R~F_R(y),
\end{equation}
where $\epsilon_L$ ($\epsilon_R$) is the fraction of left-handed (right-handed) top quarks.
The values of $\epsilon_L$ and $\epsilon_R$ are chosen as the best parameters that minimize $\chi^2$, defined as
\begin{equation}
\chi^2=\sum_{i=1}^N \left[\frac{O(y_i)-\epsilon_L F_L(y_i)-\epsilon_R F_R(y_i)}{\sigma_i}\right]^2,
\label{eq:chi2}
\end{equation}
where $N$ is the number of bins, and $\sigma_i = \sqrt{O(y_i)}$ is the statistical error
(standard deviation) of the $i$th data point.
The minimum of Eq.~\ref{eq:chi2} occurs where the derivative of $\chi^2$ with respect to
both $\epsilon_L$ and $\epsilon_R$ vanishes, yielding the normal equations of a least-squares problem:
\begin{equation}
0 = \sum_{i=1}^{N} \frac{1}{\sigma_i^2} \left[O(y_i)-\epsilon_L F_L(y_i)-\epsilon_R F_R(y_i)\right] F_l(y_i),
\quad {\rm where}~~l=L(R).
\end{equation}
Interchanging the order of summations, one can write the above equations as matrix equations,
\begin{equation}
\alpha_{LL} \epsilon_L + \alpha_{LR} \epsilon_R = \beta_L, \qquad \alpha_{RL} \epsilon_L + \alpha_{RR} \epsilon_R = \beta_R,
\label{normaleq}
\end{equation}
where
\begin{equation}
\alpha_{lm}=\sum_{i=1}^{N}\frac{F_l(y_i)F_m(y_i)}{\sigma_i^2},\qquad
\beta_l = \sum_{i=1}^N \frac{O(y_i)F_l(y_i)}{\sigma_i^2}.
\end{equation}
The coefficients $\epsilon_L$ and $\epsilon_R$ can be obtained from Eq.~\ref{normaleq} as
\begin{equation}
\epsilon_l=\sum_{m=L}^{R} [\alpha_{lm}]^{-1}\beta_{m}=\sum_{m=L}^{R} C_{lm}
\left[\sum_{i=1}^{N}\frac{O(y_i)F_m(y_i)}{\sigma_i^2}\right],\quad l=L,R~.
\end{equation}
The inverse matrix $C_{lm}\equiv [\alpha_{lm}]^{-1}$ is closely related to the standard uncertainties of the estimated coefficients $\epsilon_L$ and $\epsilon_R$. Assuming the data points are independent,
consideration of propagation of errors shows that the variance $\sigma_f^2$ in the value of $\epsilon_l$ is
\begin{equation}
\sigma^2(\epsilon_l)= \sum_{i=1}^{N} \sigma_i^2 \left(\frac{\partial \epsilon_l}{\partial O(y_i)}\right)^2=C_{ll},
\end{equation}
i.e. the diagonal elements of $[C]$ are the variances (squared uncertainties) of the fitted coefficients.
We illustrate this method with two toy models in which the $W^\prime tb$ and $Z^\prime t t$ couplings are $\epsilon_L=30\%$ and $\epsilon_L = 70\%$, where the fraction of left-handed top quarks is closely related to the $W^\prime$-$t$-$\bar{b}$ and
$Z^\prime$-$t$-$\bar{t}$ couplings as
\begin{equation}
\epsilon_L \equiv \frac{\sigma(t_L)}{\sigma(t_L)+\sigma(t_R)} \approx \frac{g_L^2}{g_L^2 + g_R^2}.
\end{equation}
For simplicity, we assume identical $\epsilon_L$ for $W^\prime$ and $Z^\prime$. (Note that
$\epsilon_L \approx 1 - \epsilon_R$.)
The coupling strength to the top quark ($g_V=\sqrt{g_L^2+g_R^2}$) and the masses of
the $Z^\prime$ and $W^\prime$ are taken to be the same: $g_V^{Z^\prime}=g_V^{W^\prime} = 0.4$;
$m_{Z^\prime} = m_{W^\prime} = 1~{\rm TeV}$.
We also adopt $5\%$ uncertainties in each bin to take into account the imperfect predictions of the template distributions $F_L(y)$ and $F_R(y)$.~\footnote{The $5\%$ variation may be too optimistic,
but the value to be adopted will not be obvious until a more precise next-to-leading order calculation
is done.}
After including SM backgrounds, we generate ten
bins of data in the $\cos\theta_\ell$ distribution. The first bin, $\cos\theta_\ell\sim -1$, is not used in
the fits because of the significant drop-off associated with cuts.
The results are shown in Fig.~\ref{fig:dop} with the legends ``True~(30,30)" and ``True~(70,70)". For an assumed $\epsilon_L = 30\%~(70\%)$ used in the event generation, the measured polarization ($\epsilon_L^{mea}$) is found to be $\epsilon_L^{mea} = 31.4\% \pm 15\%~(68.5\%\pm15\%)$ for a $Z^\prime$, and
$\epsilon_L^{mea} = 30.2\% \pm 3.4\%~(70.2\%\pm 3.4\%)$ for a $W^\prime$ for an integrated luminosity of $10~{\rm fb}^{-1}$. The uncertainties are reduced to $10.5\% $ and $2.1\%$ for a $Z^\prime$ and a $W^\prime$, respectively, if the integrated luminosity is raised to $100~{\rm fb}^{-1}$, while the central values remain the same. As shown in the figure, $\epsilon_L$ can be measured precisely in the
$W^\prime \to t\bar{b}$ mode, owing to the small SM backgrounds, while the measurement in the
$Z^\prime\to t\bar{t}$ channel is less accurate because of the large $t\bar{t}$ background.
The statistical uncertainty of each bin in the $\cos\theta_\ell$ distribution scales as
$\sqrt{N_i}=\sqrt{\sigma_i \mathcal{L}}$ where $\sigma_i$ is the
differential cross section of the $i$-th bin. An increase of the luminosity by a factor of
$k$ reduces the uncertainties by a factor of $\sqrt{k}$.
\begin{figure}[t]
\begin{center}
\includegraphics[scale=0.55,clip]{figure/top_polarization.eps}
\caption{Top quark polarizations determined from $\chi^2$ fits in
$pp\to W^{\prime+} \to t\bar{b}$ and $pp\to Z^\prime \to t\bar{t}$ for two assumed
$t$-polarizations $\epsilon_L=0.3$, 0.7. The results for benchmark models, the
left-right symmetric model (LRM), the sequential standard model (SSM where couplings
are the same as the SM $W$ and $Z$ bosons), and the top-flavor model, are also shown.
The uncertainties are the quadratic sum of the uncertainties from statistics and theory and
are shown as black and red bands for integrated luminosities of $10~{\rm fb}^{-1}$ and
$100~{\rm fb}^{-1}$, respectively.
\label{fig:dop}}
\end{center}
\end{figure}
Applying our method to the three benchmark NP models described in Sec.~II, we obtain the
results shown in Fig.~\ref{fig:dop} for $m_{W^\prime} = 1~{\rm TeV}$. In the SSM and top-flavor
models $m_{Z^\prime} = m_{W^\prime}$, while in the LRM $m_{Z^\prime}\simeq 1.2~m_{W^\prime}$ for
$\alpha_{LR}=1.6$.
All the fitted central values are very close to the true values in each model.
Because SM backgrounds are relatively small, the polarization of the top quark in $W^\prime$
decay can be measured precisely. The uncertainty is $\lesssim 5\%$ in $\epsilon_L^{W^\prime}$
for $10~{\rm fb}^{-1}$ of integrated luminosity. It decreases to $\lesssim 3\%$ for $100~{\rm fb}^{-1}$ of integrated luminosity. The polarization is measured less accurately in the $t\bar{t}$ channel, however,
because the SM $t\bar{t}$ background is large. For a $Z^\prime$ in the SSM and top-flavor models, the uncertainties are about $30\%$ and $24\%$
with $10~{\rm fb}^{-1}$ of integrated luminosity, respectively.
They are reduced to about $20\%$ and $17\%$ when the integrated luminosity is
increased to $100~{\rm fb}^{-1}$.
For a $Z^\prime$ in the LRM, the uncertainty is large,
$\sim 70\%$ for $10~{\rm fb}^{-1}$ and $\sim 40\%$ for $100~{\rm fb}^{-1}$ integrated luminosities.
The larger uncertainties arise because $Z^\prime$ mass is heavier ($1.2$ TeV) and the coupling
of the $Z^\prime$ to $t\bar{t}$ is smaller than in the SSM and top-flavor models, as noted in
Table~\ref{table:coup}. The statistics are too low to obtain a good fit.
The fitted uncertainties are generally larger for a left-handed $Z^\prime_L$ compared with a right-handed $Z^\prime_R$ if the the masses and couplings to the top quark are the same, because the charged lepton from the decay of a left-handed top quark is softer than from a right-handed top quark. Therefore, the statistics are usually greater for a $Z^\prime_R$. However, the situation switches when the $Z^\prime$ is heavier, because the separation $\Delta R$ computed from the charged lepton and the $b$ from top quark decay will more easily fail the kinematic cuts, as shown in Sec.~\ref{sec:lhc_z}.
\section{Discussion and Summary}
Extra gauge bosons may be among the new states produced at the LHC. Depending on how the
gauge group is broken in the BSM scenario and how the SM fermions are charged, the couplings of
the heavy gauge bosons $W^\prime$ and $Z^\prime$ to the SM fermions could have
the same or different handedness.
While it may be more straightforward to discover the $W^\prime/Z^\prime$ through their dijet or leptonic decays, we emphasize that decays into top quark final states allow us to study the nature of the
coupling between the new gauge bosons and the SM quarks. These channels are also
complementary to the dijet and leptonic channels, especially for leptophobic $W^\prime/Z^\prime$ bosons.
We focus on final states in which the $W^{\prime+}$ boson decays into a top quark and a bottom quark and
the $Z^\prime$ boson decays into a top-antitop quark pair. The collider signatures we examine are
{$\ell^+ \slash{\!\!\!\!E}_T b\bar{b}$} for the $W^{\prime +}$ and the semileptonic channel $\ell^\pm \slash{\!\!\!\!E}_T j j b \bar{b}$ for the $Z^\prime$. The missing energy $\slash{\!\!\!\!E}_T$ is carried off by a neutrino in the top quark decay.
After an event simulation of the signals and backgrounds, and imposing a set of kinematic cuts, we reconstruct the momentum of the missing neutrino using on-shell conditions for the $W$ boson
and the top quark. Adopting a coupling strength $g_V=\sqrt{g_L^2+g_R^2}=0.4$
($g_L$ and $g_R$ are the left-handed and right-handed $Z^\prime t\bar{t}$ couplings)
consistent with data from the Tevatron
dijet search, we find that a $1$ TeV $Z^\prime$ resonance could be seen above the SM
$t\bar{t}$ background at $14$ TeV with a statistical significance of more than $5 \sigma$ for
$10~{\rm fb}^{-1}$ of integrated luminosity. The $W^\prime$ signal can be much larger than the
SM background if the coupling strength of the $W^\prime$ to the SM quarks is not too small ( $\sim 0.4$).
Even in the presence of SM backgrounds and despite distortions associated with cuts, we
show that the left-right handedness of the top quark can be measured from the angular distribution
in $\cos\theta_{\ell}$ of the charged lepton in the top quark rest frame. This observable is most interesting because it reflects the coupling structure of the $W^\prime/Z^\prime$ to the top quark.
When performing $\chi^2$ fits to our simulated $\cos\theta_{\ell}$ distributions, we allow $5\%$ fluctuations in each bin of $\cos\theta_{\ell}$. In the $W^\prime$ case, the SM backgrounds are
negligible and our fits result in uncertainties of $\lesssim 5\%$ for the fraction of left-handed
coupling for $10~{\rm fb}^{-1}$ of accumulated luminosity. The $Z^\prime$ situation is less good because the statistics are smaller and the SM backgrounds are larger. The uncertainty for the
$Z^\prime$ decreases from $\sim 15\%$ to $\sim10\%$ when the integrated luminosity is increased
from $10~{\rm fb}^{-1}$ to $100~{\rm fb}^{-1}$.
We apply our approach to three benchmark models, the SSM, LRM, and top-flavor models.
The central values from our fits are very close to the true values in these models, with a deviation
about $2\%\sim 5\%$, and they are insensitive to luminosity. The uncertainties, again, are small for $W^\prime$ and larger for $Z^\prime$, depending mainly on statistics.
Owing to the larger uncertainty in
the $Z^\prime$ case and the similarity in the handedness of couplings (true values $\epsilon_L\simeq 84\%$ for SSM and $\epsilon_L = 100\%$ for top-flavor model), the SSM and top-flavor models can be separated
only marginally with $100~{\rm fb}^{-1}$ of integrated luminosity when the $Z^\prime$ and
$W^\prime$ results are combined, as shown in Fig.~\ref{fig:dop}. However, the situation is better if the $Z^\prime$ is leptophobic and the coupling strength to SM quarks is as large as $0.4$. This is the case we show first using toy models with the couplings $\epsilon_L = 30\%$ and $70\%$ in Fig. 7.
Since the uncertainty can be reduced to about $10\%$, one could distinguish models with handedness of couplings to SM quarks differing by about $10\%$.
We remark that the coupling strength of a $W^\prime/Z^\prime$ to the top quark cannot be
determined from our study since our signal cross sections also depend on the couplings of
the $W^\prime/Z^\prime$ to light quarks. Finally, our approach relies on reliable calculations
for the backgrounds as well as the signals, which are crucial for normalizations and the kinematic distributions. Next-to-leading order (NLO) QCD contributions enhance the cross section
significantly~\cite{Sullivan:2002jt}, possibly improving our results. The higher order QCD contributions to the SM backgrounds are also not negligible. A study that consistently includes higher order contributions to both the backgrounds and the signals is needed, and we leave it for future work.
\begin{acknowledgments}
We thank Jiang-Hao Yu for fruitful discussions.
The work by E.L.B. and Q.H.C. is supported in part by the U.S. DOE
under Grants No.~DE-AC02-06CH11357. Q.H.C. is also
supported in part by the Argonne National Laboratory and University
of Chicago Joint Theory Institute Grant 03921-07-137. C.R.C. is supported
by World Premier International Initiative, MEXT, Japan. H.Z. is supported
in part by the National Natural Science Foundation of China under
Grants 10975004 and the China Scholarship Council File No. 2009601282. C.R.C. thanks NCTS and Institute
of Physics, Academia Sinica in Taiwan for the hospitality during the final stages
of this work.
\end{acknowledgments}
|
\section{Motivation}
There is compelling evidence that the collapse of a massive star to a
black hole is the primary engine of long-duration Gamma Ray Bursts
(GRBs). The phenomenology that successfully accommodates the
astronomical observations is that of the creation of a hot fireball of
electrons, photons and protons that is initially opaque to
radiation. The hot plasma therefore expands by radiation pressure and
particles are accelerated to a Lorentz factor $\Gamma$ that grows
until the plasma becomes optically thin and produces the GRB
display. From this point the fireball is coasting with a Lorentz
factor that is constant and depends on its baryonic load. The baryonic
component carries the bulk of the fireball's kinetic energy. The
energetics and rapid time structure of the burst can be successfully
explained by shocks generated in the expanding
fireball~\cite{Shemi:1990rv,Rees:1992ek,Meszaros:1993tv}. Here, the
temporal variation of the $\gamma$-ray burst of the order of
milliseconds can be interpreted as the collision of internal shocks
with a varying baryonic load leading to differences in the bulk
Lorentz factor. Electrons accelerated by first order Fermi
acceleration radiate synchrotron $\gamma$-rays in the strong internal
magnetic field and thus produce spikes in the burst spectra of the
order of seconds~\cite{Rees:1994nw,Paczynski:1994uv}. The collision of
the fireball with interstellar gas forms external shocks that can
explain the GRB afterglow ranging from X-ray to the
optical~\cite{Meszaros:1993ft,Meszaros:1994sd}. (See
also~\cite{Meszaros:2006rc} for a recent review on theoretical
models.).
It has been pointed out that fireball baryons may be the source of
ultra-high energy (UHE) cosmic rays (CRs) with energies extending to
at least $3\times10^{20}$~eV. Baryons are inevitably accelerated along
with the electrons in the expanding fireball with their energies
boosted by the bulk Lorentz factor. It has been shown that a typical
GRB environment can naturally satisfy the requirements to produce UHE
CRs, most likely protons. It is not easy to conceive of a
mechanism where nuclei survive acceleration in a GRB fireball
environment. (Heavy nuclei could be synthesized in
magnetically-dominated jets as in the proto-magnetar model of GRBs,
see {\it e.g.}~\cite{Metzger:2011xs}.) The results of CR observatories disagree with
HiRes~\cite{Abbasi:2009nf} claiming protons and
Auger~\cite{Abraham:2010yv} heavier primaries. These conflicting
results may just illustrate that, given the poor knowledge of hadronic
interactions more than one order of magnitude above LHC energies, is
not sufficiently known to derive a definite
result~\cite{Ulrich:2009yq}.
Though GRBs satisfy the necessary conditions for accelerating
protons to UHE, it is problematic how these protons may eventually be
ejected as CRs: protons are magnetically confined to the expanding
fireball and its adiabatic cooling will reduce the maximum proton
energy significantly~\cite{Rachen:1998ir,Rachen:1998fd}. However, this
does not concern neutrons that are frequently produced in
$p\gamma$-interactions of accelerated protons with fireball photons by
processes like $p\gamma \to \Delta^+ \to n\pi^+ $. Cosmic ray
protons could thus be identified as neutrons from
$p\gamma$-interactions that can escape from the magnetic environment
and $\beta$-decay back to protons at a safe distance. A smoking-gun
test of this scenario is the production of PeV neutrinos from the
decay of the charged pions inevitably produced along with the neutrons
~\cite{Waxman:1997ti,Bahcall:1999yr,Murase:2005hy,Anchordoqui:2007tn}. This neutrino flux
is in reach of large-scale neutrino telescopes like
IceCube~\cite{Ahrens:2003ix,Abbasi:2011qc}. (Possible emission of
neutrinos associated to the GRB progenitor and prior to the
$\gamma$-ray burst have been discussed
in~\cite{Razzaque:2002kb}. Neutrinos from proton interactions in
late internal or external shocks during the afterglow phase have been considered
in~\cite{Murase:2006dr,Murase:2007yt,Razzaque:2009zz}.)
In this paper we calculate the diffuse flux of neutrinos produced in
association with GRB cosmic rays by directly fitting the proton
spectra from the decay of fireball neutrons to CR data from HiRes. As
a byproduct we also obtain the flux of so-called GZK neutrinos
produced when the cosmic rays interact with microwave and
infro-red/optical background
photons~\cite{Greisen:1966jv,Zatsepin:1966jv}. The accompanying photon
flux resulting from the electromagnetic cascading of the neutral pions
peaks in the MeV-GeV energy region; we verify that it does not exceed
the extra-galactic diffuse $\gamma$-ray flux inferred by
Fermi-LAT~\cite{Abdo:2010nz}. Our main conclusion is that the
predicted flux exceeds the upper bound on a diffuse flux of cosmic
neutrinos obtained by IceCube from a year of data taken with half the
instrument during construction. Facing this negative conclusion, we
subsequently investigate the dependence of the predicted neutrino flux
on the cosmological evolution of the sources as well as on the
parameters describing the fireball. Although the latter are
constrained by the electromagnetic observation as well as by the the
requirement that the fireball must accommodate the observed cosmic ray
spectrum, the predictions can be stretched to the point that it will
take 3 years of data with the now completed instrument to conclusively
rule out the GRB origin of UHE CRs.
We work throughout in natural Heaviside-Lorentz units with
$\hbar=c=\epsilon_0=\mu_0=1$, $\alpha=e^2/(4\pi)\simeq1/137$ and
$1~{\rm G} \simeq 1.95\times10^{-2}{\rm eV}^2$.
\section{Fireball Model}
In the fireball model the GRB central engine produces a heated
optically thick plasma of leptons, photons and baryons which is
initially at rest. The fireball expands adiabatically by radiation
pressure until it becomes optically thin. From this point on the
fireball is coasting with a Lorentz factor $\Gamma$ which depends on
its baryonic load. The expanding plasma flow is shocked producing
shells with varying baryonic load; this results in internal shells
with different velocities that in their collision produce internal
shocks. Although in the baryon-rich fireballs that would produce EHE
cosmic rays most of the kinetic energy of the fireball is carried by
the baryons, the internal shocks will convert a fraction
$\varepsilon_e = U'_e/U'_{\rm kin}$ to leptons and $\varepsilon_B =
U'_B/U'_{\rm kin}$ to magnetic fields supported by the plasma. Here
and in the following, primed quantities refer to values in the
comoving plasma frame, whereas unprimed quantities are reserved for
the observer's frame.
Study of the emission of GRB has resulted in ``benchmark'' parameters
describing GRB fireball with the efficiency for lepton acceleration
from bulk energy taken to be 10\%, {\it i.e.}~$\varepsilon_e =
0.1$. The fraction of magnetic field energy and hence the value of the
turbulent magnetic field strength can be inferred from the observed
peak of the $\gamma$-ray synchrotron spectrum with $\varepsilon_B =
0.1$, corresponding to equipartition of energy in leptons and magnetic
field. The ``benchmark'' parameters are required to produce the
observed peak emission at MeV energies resulting from the synchrotron
radiation of GRB electrons and must be consistent with afterglow
observations. The typical radius of on internal shock $r_i$ is given
by the speed-of-light distance implied by the duration of the spikes
observed in the GRB display. Because the shells are boosted toward the
observer by a Lorentz factor $\Gamma_i$, $r_i \simeq
t_v/(1-\beta_i)\simeq 2\Gamma_i^2t_v$. We will take the variability
scale to be $t_v = 0.01$~s. Although variations in GRB spectra down to
milliseconds have been observed, what is important here is the Fourier
strength on the variability, and the choice can be debated. In any
case, when obtaining our final results we will return to this problem
and vary this as well as other critical benchmark parameters over a
wide range.
Long-duration GRB are beamed; their observed isotropically-equivalent
$\gamma$-ray luminosity is on average $L_\gamma =10^{52}$~erg/s. In
the fireball model it results from synchrotron radiation by electrons
accelerated in internal shocks. The spectral photon density of the
burst (GeV${}^{-1}$ cm${}^{-3}$) in the observer's frame can be
adequately parametrized as~\cite{Band:1993eg}
\begin{equation}
n_\gamma \propto
\begin{cases}(\epsilon/\epsilon_0)^\alpha
e^{-\epsilon/\epsilon_0}&\epsilon<(\alpha-\beta)\epsilon_0\\
(\alpha-\beta)^{\alpha-\beta}e^{\beta-\alpha}(\epsilon/\epsilon_0)^\beta
&\epsilon>(\alpha-\beta)\epsilon_0
\end{cases}
\end{equation}
with $\alpha\simeq-1$, $\beta\simeq-2.2$ and
$\epsilon_0\simeq1$~MeV. The normalization is given by $U_\gamma =
\int {\rm d}\epsilon\,\epsilon n_\gamma(\epsilon) = L_\gamma/4\pi
r_i^2$. The corresponding photon density in the comoving frame is then
given by $n'_\gamma(\epsilon') = n_\gamma(\Gamma_i\epsilon')$.
After Fermi acceleration the electron population follows a power-law
spectrum with minimum energy $E'_{e,{\rm min}} \simeq
(\varepsilon_e/\zeta_e) m_p$ in the comoving frame. Here $\zeta_e<1$
is the fraction of electrons that achieve
equipartition~\cite{Meszaros:2006rc}. This spectrum yields by
synchrotron radiation a photon spectrum peaking at
\begin{equation}\label{eq:synpeak}
\epsilon_{0} \simeq\Gamma_i \frac{3}{2}\frac{e B'}
{m^3_e}\left(\frac{\varepsilon_e}{\zeta_e} m_p\right)^2 \simeq
0.8\,\left( \frac{\varepsilon_{e,-1}^3 \varepsilon_{B,-1}
L_{\gamma,52}} {\zeta_{e,-1}^4\Gamma_{i,2.5}^4 t_{v,-2}^2}
\right)^{1/2}\!{\rm MeV}\,,
\end{equation}
where $L_\gamma = L_{\gamma,52}10^{52}$~erg/s, $\varepsilon_{B}
=\varepsilon_{B,-1}0.1$, $\varepsilon_{e} =\varepsilon_{e,-1}0.1$,
$\zeta_{e} =\zeta_{e,-1}0.1$, $\Gamma_{i} = \Gamma_{i,2.5}10^{2.5}$
and $t_{v,-2} = t_{v,-2}0.01$~s. This is close to the observed peak of
the GRB burst spectrum at $\mathcal{O}({\rm MeV})$.
This concludes our description of the GRB fireball and its
electromagnetic spectrum. There is nothing new here. We will next
discuss the production of cosmic rays and neutrinos by applying the
fireball phenomenology just described.
\section{Cosmic Ray and Neutrino Emission}
Although it is straightforward to argue that GRB fireballs represent
an environment that can yield CRs of very high energy, it is important
to take into account that the acceleration competes against energy
loss due to synchrotron radiation and pion production. In other words,
the fireball phenomenology is subject to the conditions that (i) the
time to accelerate protons to the highest energy does not exceed the
lifetime of the fireball and (ii) that the energy gained is not lost
to synchrotron radiation and pion production. We discuss these
constraints sequentially.
For efficient acceleration of UHE CR protons their gyroradius must be
contained within the acceleration region which is related to the size
of the shock. In the comoving frame the Larmor radius of a proton is
$r'_{\rm L} = E'/eB'$ and the corresponding acceleration time
$c/r'_{\rm L}$ is given by $t'_{\rm acc} = \eta r'_L$ with
$\eta\gtrsim1$ or
\begin{equation}
t'_{\rm acc} \simeq 8\times10^{10}\left(\frac{\eta^2\varepsilon_{e,-1}
E_{p,20.5}^2t_{v,-2}^2\Gamma^4_{i,2.5}}{\varepsilon_{B,-1}
L_{\gamma,52}}\right)^{1/2}{\rm
cm}\,,
\end{equation}
where $E_p =E_{p,20.5}10^{20.5}$~eV, corresponding to the upper end of
the UHE CR spectrum. The size of the accelerator is set by the size of
the shock, $t'_{\rm dyn} \simeq r_i/2\Gamma_i = t_v\Gamma_i$, or
\begin{equation}\label{eq:tdyn}
t'_{\rm dyn} \simeq 1\times10^{11} \Gamma_{i,2.5}t_{v,-2}\,{\rm cm}\,.
\end{equation}
From this we derive the maximal proton energy in the observer's frame
for which $t'_{\rm dyn}=t'_{\rm acc}$,\
\begin{equation}\label{eq:Epmax1}
E_{p,\rm max} \simeq 4\times 10^{20} \left( \frac{\varepsilon_{B,-1}
L_{\gamma,52}}{\eta^2\varepsilon_{e,-1} \Gamma_{i,2.5}^{2}}
\right)^{1/2}\!\!{\rm eV}
\end{equation}
We next consider the energy losses that compete with the
acceleration. In the comoving frame the time scale associated with
synchrotron radiation is $t'_{\rm sync} = 9\pi m^4/E'e^4B'^2$ or
\begin{align}\label{eq:tsync}
t'_{\rm sync}
\simeq 7\times10^{10}\left(\frac{\varepsilon_{e,-1}
\Gamma^7_{i,2.5}t_{v,-2}^2}{\varepsilon_{B,-1}E_{p,20.5}
L_{\gamma,52}}\right) {\rm cm}\,.
\end{align}
The maximal proton energy is reduced when $t'_{\rm syn}=t'_{\rm acc}$ or is smaller; in this case, using (\ref{eq:Epmax1})
\begin{equation}\label{eq:Epmax2}
E_{p,\rm max} \simeq 3\times 10^{20} \left( \frac{\varepsilon_{e,-1} \Gamma_{i,2.5}^{10} t_{v,-2}^2}{\eta^2\varepsilon_{B,-1} L_{\gamma,52}} \right)^{1/4}\!\!{\rm eV}\,.
\end{equation}
The photo-pion energy loss rate is determined by the $p\gamma$ cross
section, the photon density $n_{\gamma}$ and the average energy loss
of the protons $\langle x_{p\gamma}\rangle$ in each interaction. In
the $\Delta^+$-resonance approximation
\begin{equation}
t'^{-1}_{\Delta^+} \simeq \sigma_\Delta\langle
x_{p\to\Delta^+}\rangle\Gamma_{\Delta^+} \frac{\pi}{2}\frac{m_p}{E'_p}
n_\gamma\left(\frac{\Gamma_im^2_{\Delta^+}}{4E'_p}\right)\,,
\end{equation}
where $\Gamma_{\Delta^+} \simeq 120$~MeV, $m_{\Delta^+} = 1232$~MeV
and $\sigma_{\Delta^+} \simeq 420\,\mu$b and $\langle
x_{p\to\Delta^+}\rangle\simeq 0.2$. Hence, for a spectral slope
$\alpha\simeq-1$ the optical depth becomes constant above a proton
energy
\begin{equation}\label{eq:Epbreak}
E_{p,{\rm b}} \simeq 0.4\frac{\Gamma_i^2}{\epsilon_{0}}{\rm GeV}^2
\simeq 4\times10^{16}\frac{\Gamma_{i,2.5}^2}{\epsilon_{0,6}}{\rm eV}\,,
\end{equation}
with $\epsilon_0 = \epsilon_{0,6}$~MeV.
Photo-pion production therefore introduces a break in the neutron
production spectrum at (\ref{eq:Epbreak}). Above this energy the
energy loss length in the comoving frame is
\begin{equation}\label{eq:tdelta}
t'_{\Delta} \simeq 4\times10^{12} \left(\frac{\epsilon_{0,6}\Gamma^5_{i,2.5}t_{v,-2}^2}{L_{\gamma, 52}}\right){\rm cm}\,,
\end{equation}
and the maximal proton energy satisfying $t'_{\Delta} = t'_{\rm acc}$ is
\begin{equation}\label{eq:Epmax3}
E_{p,\rm max} \simeq 2\times 10^{22} \left( \frac{\varepsilon_{B,-1}
\Gamma_{i,2.5}^{6} t_{v,-2}^2}{\eta^2\varepsilon_{e,-1}L_{\gamma,52}
} \right)^{1/2}\!\!{\rm eV}\,.
\end{equation}
For GRB fireball parameters considered in this work the scale of
photo-pion losses (\ref{eq:tdelta}) is always larger than the
synchrotron scale (\ref{eq:tsync}) or the dynamical scale
(\ref{eq:tdyn}). The maximal proton energy is hence given by the
smaller of Eqs.~(\ref{eq:Epmax1}) and (\ref{eq:Epmax2}).
There is an additional consideration as mentioned in the introduction:
because protons are magnetically coupled to the expanding fireball
they will lose energy adiabatically if they remain confined in the
expanding shock. Adiabatic cooling, however, is not important for neutrons that are
produced at scales $\langle x_{p\to\Delta}\rangle t'_\Delta$ and can
escape the source~\cite{Rachen:1998ir}.
We will assume that the spectrum of observable CRs results from
neutrons escaping the source. The competition between various energy loss mechanisms in internal shocks can introduce various features in the injection spectrum of individual GRBs~\cite{Protheroe:1998pj}. Within our approximation of a homogeneous source distribution we assume that the {\it effective} CR emission rate can be approximated as
\begin{equation}\label{eq:Qp}
Q_{\rm CR}(E) \simeq Q_0 \frac{(E/E_{p,{\rm
b}})^{-\gamma}}{1+(E/E_{p,{\rm b}})^{\beta-\alpha}}e^{-E/E_{p,{\rm
max}}}\,.
\end{equation}
For $E\gg E_{p,{\rm b}}$, this reduces to the typical $Q_0
E^{-\gamma}\exp(-E/E_{p,{\rm max}})$ approximation of the CR injection
spectrum that we are going to test against HiRes data in the
following. If energy loss of pions and muons prior to decay were
negligible, we can relate the neutrino (per flavor) and CR neutron
emission rates as
\begin{equation}
Q^0_{\nu}(E_\nu) \simeq \frac{1}{\epsilon}Q_{\rm CR}(E_\nu/\epsilon)\,,
\end{equation}
where $\epsilon = \langle E_\nu/E_n\rangle \simeq 0.06$. However, synchrotron losses of secondary pions and muons in the background magnetic field are important. With muon and pion lifetimes of $t_\mu^{\rm dec} = 2.2\,\mu{\rm s}$
and \mbox{$t_\pi^{\rm dec} = 26\,{\rm ns}$}, respectively, this will introduce a synchrotron break in the respective spectrum at
\begin{equation}
E'_{\pi/\mu,{\rm s}} = \frac{3}{4}\sqrt{\frac{m_{\pi/\mu}^5}{\pi\alpha^2 B'^2t_{\pi/\mu}^{\rm dec}}}
\end{equation}
The corresponding break in the neutrino spectra from $\pi^+\to\mu^+\nu_\mu$ and $\mu^+\to e^+\bar\nu_\mu\nu_e$, respectively, is $E'_{\nu,{\rm s}}\simeq E'_{\pi/\mu,{\rm s}}/4$ or
\begin{equation}E_{\nu,{\rm s}} = \left( \frac{\varepsilon_{e,-1} \Gamma_{i,2.5}^8 t_{v,-2}^2} {\varepsilon_{B,-1} L_{\gamma,52}} \right)^{1/2}\!\!\!\!\times\begin{cases} 2\times 10^{17} ~{\rm eV} & (\nu_{\mu})\,,\\ 1\times 10^{16} ~{\rm eV} & ({\bar \nu}_{\mu}, \nu_e)\,. \end{cases}
\end{equation}
The total diffuse neutrino flux can hence be approximated as
\begin{equation}
Q_{{\rm all}\,\nu}(E_\nu) = \sum_\alpha\frac{Q^0_{\nu}(E_\nu)}{1+(E_\nu/E_{\nu_\alpha,{\rm s}})^2}\,,
\end{equation}
where the sum runs over the neutrino flavors $\nu_\alpha = \nu_\mu$, $\bar\nu_\mu$ or $\nu_e$\,.
\section{Diffuse Cosmic Spectra}
For the calculation of the spectrum of UHE CR protons we assume that
the cosmic source distribution is spatially homogeneous and
isotropic. The comoving number density $Y_i = n_i/(1+z)^3$ of nuclei
of type $i$ is then governed by a set of Boltzmann equations of the
form:
\begin{equation}\label{eq:boltzmann}
\dot Y_i = \partial_E(HEY_i) + \partial_E(b_iY_i)-\Gamma_{i}\,Y_i
+\sum_j\int{\rm d} E_j\,\gamma_{j\to i}Y_j+\mathcal{L}_i\,,
\end{equation}
together with the Friedman-Lema\^{\i}tre equations describing the
cosmic expansion rate $H(z)$ as a function of the redshift $z$:
\mbox{$H^2 (z) = H^2_0\,[\Omega_{\rm m}(1 + z)^3 +
\Omega_{\Lambda}]$}, normalized to its present value of $H_0 \sim70$
km\,s$^{-1}$\,Mpc$^{-1}$, in the usual ``concordance model'' dominated
by a cosmological constant with $\Omega_{\Lambda} \sim 0.7$ and a
(cold) matter component, $\Omega_{\rm m} \sim
0.3$~\cite{Nakamura:2010zzi}. The time-dependence of the redshift is
${\rm d}z = -{\rm d} t\,(1+z)H$. The first and second term in the
r.h.s.~of Eq.~(\ref{eq:boltzmann}) describe continuous energy losses
(CEL) due to red-shift and $e^{+}e^{-}$ pair
production~\cite{Blumenthal:1970nn} on the cosmic photon backgrounds,
respectively. The third and fourth terms describe more general
interactions involving particle losses ($i \to$ anything) with
interaction rate $\Gamma_i$, and particle generation $j\to i$. We use
the Monte Carlo package {\tt SOPHIA}~\cite{Mucke:1999yb} to calculate
$p\gamma$ interaction rates and spectra. The last term,
$\mathcal{L}_i$, accounts for the CR emission rate per co-moving
volume.
We describe the cosmic evolution of the CR sources by the ansatz
$\mathcal{L}_{\rm CR}(z,E) = \mathcal{H}_{\rm GRB}(z)Q_{\rm CR}(E)$.
Given their supernova association it is natural to assume that the
comoving density of GRBs follows the star formation rate (SFR). We will use
in the following the approximation~\cite{Hopkins:2006bw,Yuksel:2008cu}
\begin{equation}
\label{eq:HSFR}
\mathcal{H}_{\rm SFR}(z) = \begin{cases}(1+z)^{3.4}&z<1\,,\\
N_1\,(1+z)^{-0.3}&1<z<4\,,\\N_1\,N_4\,(1+z)^{-3.5}&z>4\,,
\end{cases}
\end{equation}
with normalization factors, $N_1 = 2^{3.7}$ and $N_4 = 5^{3.2}$. It
has been suggested that the GRB population may not directly follow the
SFR and may have been stronger in the past. This possibility is interesting
also in the context of UHE CR models: a strong evolution of CR proton sources
can account for the spectrum of extra-galactic UHE CRs even below the
ankle~\cite{Berezinsky:2002nc}. These ``low-crossover'' models predict
significantly larger neutrino fluxes from the sources of CR protons and are
already limited by upper bounds on diffuse neutrino
fluxes~\cite{Ahlers:2005sn,Ahlers:2009rf}. For the illustration of the effect
of a strong GRB rate at high redshift we will use in the following a comparatively
strong evolution of the from~\cite{Yuksel:2006qb}
\begin{equation}\label{eq:Hstrong}
\mathcal{H}_{\rm strong}(z) = (1+z)^{1.4}\, \mathcal{H}_{\rm SFR}(z)\,.
\end{equation}
This is consistent with a more recent study~\cite{Kistler:2009mv}. However,
the sample of high-redshift GRBs where the enhanced evolution (\ref{eq:Hstrong})
becomes strongest is small and plagued by systematics. It has also been argued
that the evolution at low redshift is consistent with the SFR whereas at high
redshift it can be approximated as an almost constant
rate~\cite{Le:2006pt,Guetta:2007zz,Wanderman:2009es}.
The density of CR sources at high redshift has two effects on the
analysis. Firstly, a stronger evolution of the CR sources requires in
general lower values of the power-law index $\gamma$ to reproduce the
CR data. Secondly, a higher density of sources in the past tend to
produce larger energy densities of secondary neutrinos and
$\gamma$-rays. However, we will see later on that the contribution
of UHE CR proton sources to the diffuse extra-galactic $\gamma$-ray
background is already close to maximal if we assume a source evolution
following the SFR. A stronger evolution as in Eq.~(\ref{eq:Hstrong})
can not significantly enhance the prompt neutrino fluxes without violating
the $\gamma$-ray bound. We will estimate the effect of source evolution on
the bolometric electro-magnetic energy density in the following.
Particles with electromagnetic (EM) interactions produced in association
with the cosmic rays, $\gamma$-rays, electrons and positrons, will
cascade in the universal photon background and magnetic fields on
time-scales much shorter than their production rates. The relevant
processes with background photons are inverse Compton scattering (ICS), $e^\pm+\gamma_{\rm
bgr}\to e^\pm+\gamma$, pair production (PP), $\gamma+\gamma_{\rm
bgr}\to e^++e^-$, double pair production (DPP) $\gamma+\gamma_{\rm
bgr}\to e^++e^-+e^++e^-$, and triple pair production (TPP),
$e^\pm+\gamma_{\rm bgr}\to
e^\pm+e^++e^-$~\cite{Blumenthal:1970nn,Blumenthal:1970gc}. High energy
electrons and positrons can also lose energy via synchrotron radiation
on the intergalactic magnetic field with strength limited
to be $\sim 10^{-9}$G~\cite{Kronberg:1993vk}.
The evolution of the diffuse $\gamma$-ray and $e^\pm$ spectra follows
the Boltzmann equations~(\ref{eq:boltzmann}). For recent studies
see~\cite{Berezinsky:2010xa,Ahlers:2010fw}. After cascading the EM
flux accumulates into $\gamma$-rays of GeV-TeV energy with a
characteristic and essentially universal spectrum. Its normalization
is determined by the total energy density of EM radiation from the
propagation loss of CR nuclei. For this purpose, we define the
comoving energy density at redshift $z$ as
\begin{equation}
\label{eq:omegacasz}
\omega_{\rm cas}(z) \equiv \int {\rm d} E\,E\!\sum\limits_{i=\gamma,e^\pm}\!Y_i(z,E)\,,
\end{equation}
which follows the evolution equation
\begin{equation}
\dot\omega_{\rm cas} + H\omega_{\rm cas} = \int{\rm d} E\, b(z,E)Y_{\rm CR}(z,E)\,.
\end{equation}
The energy loss rate $b\equiv{\rm d}E/{\rm d}t$ was already
defined by Eq.~(\ref{eq:boltzmann}), but here comprises the combined
energy loss of CR protons into EM radiation by Bethe-Heitler
production and photo-pion interactions. The energy density (eV
cm${}^{-3}$) of the electromagnetic background observed today is
therefore given by
\begin{equation}
\label{eq:omegacas}
\omega_{\rm cas} = \int{\rm
d}t\int{\rm d}E \,\frac{b(z,E)}{(1+z)}\,Y_{\rm CR}(z,E)\,.
\end{equation}
Typically the energy density (\ref{eq:omegacasz}) today obtained
by a detailed calculation of the EM spectra agrees with the (quicker)
bolometric calculation (\ref{eq:omegacas}) within a few percent and we
will use the latter in our statistical analysis.
\begin{figure}[t]
\centering
\includegraphics[width = 0.85\linewidth]{cont_hires.pdf}
\caption[]{Results of the goodness of fit test of the HiRes
data~\cite{Abbasi:2007sv}. We show the 99\% (magenta) confidence
levels of the injection index $\gamma$ and the maximal energy $E_{\rm
max}$. In this analysis we consider only minimal $\gamma$-ray
production coming from Bethe-Heitler pair production and photo-pion
production of CR protons during propagation (``Min''). The dashed line
(``Max'') assumes that also $\gamma$-rays from $\pi^0$ production
survive the internal shock environment and cascade during propagation
in the inter-galactic background.}\label{fig:gof}
\end{figure}
The influence of cosmic evolution on the energy density of the cascade
can be estimated in the following way. The UHE CR interactions with
background photons are rapid compared to cosmic time-scales. The
energy threshold of these processes scales with redshift as $E_{\rm
th}/(1+z)$ where $E_{\rm th}$ is the (effective) threshold today. We
can therefore approximate the evolution of the energy density of the
secondaries as
\begin{equation}
\dot\omega_{\rm cas} + H\omega_{\rm cas}
\simeq \eta_{\rm cas}\mathcal{H}(z)\!\!\!\!\!\!
\int\limits_{E_{\rm th}/(1+z)}\!\!\!\!\!\!{\rm d} E\,E\,Q_{\rm CR}(E)\,,
\end{equation}
where $\eta_{\rm cas}$ denotes the energy fraction of the CR
luminosity converted to the electromagnetic cascade. Assuming a
power-law injection $Q_{\rm CR}(E)\propto E^{-\gamma}$ with
sufficiently large cutoff $E_{\rm max}\gg E_{\rm th}$ we obtain that
strong cosmic evolution (\ref{eq:Hstrong}) enhances the diffuse
$\gamma$-spectrum as $\omega_{\rm cas} \propto \int{\rm
d}t(1+z)^{1.4+\gamma-2}$. For the proton spectrum $\gamma\simeq2.3$
this corresponds to a relative increase of $\sim4$, which agrees with
numerical results.
Photo-hadronic interactions in internal shocks produce EM radiation
from the decay of pions. Most of this additional EM radiation is
expected to cascade in the high background densities of $\gamma$-rays
or $e^\pm$ and strong magnetic fields of the fireball and should
eventually contribute to the luminosity of the burst. However, it is
conceivable that part of it decouples from the fireball to contribute
an additional source term $Q_{\rm EM}$~\cite{Razzaque:2004cx} to the
diffuse $\gamma$-ray background. Assuming energy conservation in the
cascade this additional contribution is smaller than
\begin{equation}\label{eq:cassource}
\omega_{\rm cas, source} = \frac{\xi}{H_0}\int{\rm d}E\, E\, Q_{\rm
EM}(E)
\end{equation}
with $\xi/H_0 \equiv \int{\rm d}t\mathcal{H}(z)/(1+z)$. Assuming the
GRB evolution (\ref{eq:HSFR}) and (\ref{eq:Hstrong}) this gives
$\xi_{\rm SFR} \simeq 2.4$ and $\xi_{\rm strong} \simeq 7.3$,
respectively.
In the following we will assume two limiting cases for the
contribution of GRBs to the diffuse GeV-TeV background. The minimal
model (``Min'') assumes that the only contribution to the cascade
results from the propagation from CR protons. In a maximal model
(``Max'') we assume that additionally all $\gamma$-rays from neutral
pion production in the GRB contribute maximally to the cascade in the
form of the term (\ref{eq:cassource}) with $Q_{\rm EM} = Q_{\pi^0}$.
\begin{figure}[t]
\centering \includegraphics[width = 0.85\linewidth]{fnor_hires.pdf}
\caption[]{The allowed proton range of proton flux at the 99\%
confidence level. The color-coding is as in Fig.~\ref{fig:gof}. Each
fit of the proton spectrum is marginalized with respect to the
experimental energy uncertainty and we show the shifted predictions in
comparison to the HiRes central values~\cite{Abbasi:2007sv}. For
comparison we also show the Auger data~\cite{Abraham:2010mj} which has
{\em not} been included in the fit.}\label{fig:pspectra}
\end{figure}
\section{Goodness of Fit Test}
In this section we present the details and results of our
statistical analysis. We perform a goodness of fit (GOF) test of the
compatibility of cosmic ray data with a given model, characterized by
the spectral index $\gamma$ and the maximal energy $E_{\rm max}$ of
the injected cosmic ray and by the two evolution models
(\ref{eq:HSFR}) and (\ref{eq:Hstrong}). We use CR data from HiRes I
and II \cite{Abbasi:2007sv} above the ``ankle'' at 4~EeV. We
additionally impose the requirement that the electromagnetic flux
accompanying the cosmic rays does not exceed the Fermi-LAT
measurements of the diffuse extra-galactic $\gamma$-ray background in
the GeV-TeV energy range.
Given the acceptance $A_i$ (in units of area per unit time per unit
solid angle) of the experiment in bin $i$ centered at energy $E_i$
with bin width $\Delta_i$ and energy scale uncertainty $\sigma_{E_s}$,
the number of events expected in the bin is
\begin{equation}
N_i(\gamma,E_{\rm max},{\cal N},\delta)= A_i\!\!\!\!
\int\limits_{E_i(1+\delta)-\Delta_i/2}^{E_i(1+\delta)+\Delta_i/2}\!\!\!\!
{\rm d}E\, J^p_{{\cal N},\gamma,E_{\rm max}}(E)\,,
\label{eq:nev}
\end{equation}
where $J^p_{{\cal N},\gamma,E_{\rm max}}(E)=n_p(0,E)/4\pi$ is the
proton flux arriving at the detector. It is determined by the source
luminosity of Eq.~\eqref{eq:Qp} with $E_{p,b}\ll E_{\rm min} = 4$~EeV
and cosmic evolution given by Eqs.~\eqref{eq:HSFR}
or~\eqref{eq:Hstrong}. The parameter $\delta$ in Eq.~\eqref{eq:nev}
is a fractional energy-scale shift that takes into account the
uncertainty in the energy-scale and ${\cal N}$ is the normalization of
the proton source luminosity.
The probability distribution of events in the $i$-th bin follows a
Poisson ditribution with mean $N_i$. Correspondingly, the
$r$-dimensional probability distribution for a set of non-negative
integer numbers ${\vec k}=\{k_1,...k_r\}$, $P_{\vec k}(n,\gamma,{\cal
N},\delta)$ is just the product of the individual Poisson
distributions with $r$ the number of bins with $E_i\geq E_{\rm
min}$. Given a model, the experimental result ${\vec N^{\rm
exp}}=\{N^{\rm exp}_1,...,N^{\rm exp}_r\}$ has a probability $P_{\vec
N^{\rm exp}}(\gamma,E_{\rm max},{\cal N},\delta)$ which, after
marginalizing over the uncertainty in the energy scale and
normalization, is given by
\begin{equation}
P_{\rm exp}(\gamma,E_{\rm max})={\rm max}_{\delta,{\cal N}}\left(
P_{\vec N^{\rm exp}}(\gamma,E_{\rm max},{\cal N},\delta)\right)\,.
\label{eq:margi}
\end{equation}
Here the maximization is for some prior for $\delta$ and ${\cal N}$.
As a prior the energy shift $\delta$ we used a top hat
of width $\sigma_{E_s}$.
For ${\cal N}$ we impose two priors. The first is associated with the
upper bound on the total EM of Eq.~\eqref{eq:omegacas} that should not
exceed the Fermi-LAT measurements~\cite{Abdo:2010nz}, or following
Ref.~\cite{Berezinsky:2010xa}
\begin{equation}
w_{\rm cas}({\cal N} ,\gamma,E_{\rm max})
\leq 5.8\times 10^{-7}\; {\rm eV}/{\rm cm}^3\,.
\label{eq:fermilat}
\end{equation}
The second prior on the normalization is imposed by requiring that the
proton spectra do not exceed the HiRes I and II data below $E_{\rm
min}$ by more than three standard deviations.
The marginalization in Eq.~(\ref{eq:margi}) also determines ${\cal
N}_{\rm best}$ and $\delta_{\rm best}$ for the model, which are the
values of the energy shift and normalization that yield the best
description of the experimental CR data, subject to the constraint
imposed by the Fermi-LAT measurement.
In the end the model is compatible with the experimental
results at a given GOF if
\begin{equation}
\sum\limits_{P_{\vec k}\,>\,P_{\rm exp}}
P_{\vec k}(\gamma,E_{\rm max},{\cal N}_{\rm best},\delta_{\rm best})
\leq {\rm GOF}\,.
\end{equation}
Technically, this calculation is performed by generating a large
number $N_{\rm rep}$ of replica experiments following the
probability distribution $P_{\vec k}$ and by imposing that the
fraction $F$ with $P_{\vec k}>P_{\rm exp}$ satisfy $F\leq{\rm GOF}$.
With this method we determine the range of values of $(\gamma,E_{\rm
max})$ that are compatible with the HiRes I and HiRes II
data~\cite{Abbasi:2007sv}. We show in Fig.~\ref{fig:gof} the regions
with GOF 99\% for the two evolution models (\ref{eq:HSFR}) (left
panel) and (\ref{eq:Hstrong}) (right panel). In order to illustrate
the impact of the constraint imposed by the Fermi-LAT measurements, we
also show the GOF regions without imposing it. The region bounded by
the dotted line shows the reduced parameter space resulting from the
condition that secondary EM radiation from proton propagation does not
exceed the Fermi-LAT measurement (``Min''). This bound can become even
stronger if we consider UHE $\gamma$-ray emission from $\pi^0$ decay
in the source with diffuse energy density (\ref{eq:cassource}) as
indicated by the region bounded by the dashed line
(``Max''). Figure~\ref{fig:pspectra} shows the range of the
corresponding proton fits to the data.
We have not included in the analysis the results from the Auger
Collaboration \cite{Abraham:2008ru,Abraham:2010mj}, which are shown
in Fig.\ref{fig:pspectra} for illustration only. As
described in Refs.~\cite{Abraham:2008ru,Abraham:2010mj}, besides the
energy scale uncertainty there is also an (energy-dependent) energy
resolution uncertainty which implies that bin-to-bin migrations
influence the reconstruction of the flux and spectral shape. Since the
form of the corresponding error matrix is not public, this data
\cite{Abraham:2008ru,Abraham:2010mj} cannot be analysed outside the
Auger Collaboration.
\begin{figure}[p]
\centering
\includegraphics[width =0.9\linewidth]{nuspectra.pdf}
\caption[]{Prompt neutrino spectra for SFR (left) and strong (right)
evolution corresponding to the 99\% C.L.~of the ``Min'' model shown
in Fig.~\ref{fig:gof} and assuming the luminosity range
$0.1<(\varepsilon_B/\varepsilon_e)L_{\gamma,52}<10$. We show the
prompt spectra separately for the branches $t_{\rm dyn}<t_{\rm syn}$
(green right-hatched) and $t_{\rm syn}<t_{\rm dyn}$ (blue
left-hatched). The preliminary IceCube limits~\cite{IceCubeEHE,Abbasi:2011ji} (90\% C.L.) on the total
neutrino flux from the analysis of high-energy (HE) and extremely-high energy (EHE)
neutrinos with the 40 string sub-array (IC-40) assume 1:1:1 flavor
composition after oscillation. We also show the sensitivity of the
full IceCube detector (IC-86) to EHE neutrinos after three years
of observation. The gray solid area shows the range of GZK neutrinos
expected at the 99\% C.L.}\label{fig:nuspectra}
\end{figure}
\section{Results and Their Dependence on Model Parameters}
Our final results are summarized in Fig.~\ref{fig:nuspectra} that
shows the range of prompt neutrino spectra corresponding to the UHE CR
proton spectra at the 99\% C.L.~of the GOF test with the HiRes data
assuming SFR evolution (left panels) and the stronger evolution of
Eq.~(\ref{eq:Hstrong}) (right panels). We have assumed that proton
acceleration in the shocks is efficient ($\eta\simeq1$) and have fixed
the maximum $\epsilon_0$ of the prompt $\gamma$-ray emission
(Eq.~(\ref{eq:synpeak})) at 1~MeV. For each GRB model the maximum
proton energy $E_{\rm max}$ corresponds to the smaller of
Eq.~(\ref{eq:Epmax1}) or (\ref{eq:Epmax2}), {\it i.e.}~$t_{\rm dyn} <
t_{\rm syn}$ or $t_{\rm syn}< t_{\rm dyn}$, respectively. For given
values of $E_{\rm max}$, $\Gamma_i$ and $t_v$ we can in general derive
two solutions of $(\varepsilon_B/\varepsilon_e)L_{\gamma}$, which we
allow to vary within the range
$0.1<(\varepsilon_B/\varepsilon_e)L_{\gamma,52}<10$. The corresponding
branches are indicated in the plots.
The upper three rows of Fig.~\ref{fig:nuspectra} shows prompt neutrino
spectra with a Doppler factor $\Gamma_i = 10^{2.5}$ of the internal
shock. The top panels show the results for the variation
time-scale $t_v=0.01$. This includes the range of neutrino fluxes for
typical ``benchmark'' values of the GRB environment with $L_{\gamma} =
10^{52}$~erg/s and $E_{\rm max}=10^{20.5}$~eV, which is shown separately
as the cross-hatched area. The range of prompt neutrino spectra for
the benchmark model corresponds to the range of the power-law index
$\gamma$ at the 99\% C.L.~and the normalization of each particular
model. Note, that for both cases, SFR or strong evolution, the range
of the prompt neutrino flux exceeds the present limits of IC-40.
The energy density of the prompt neutrino spectrum peaks at an energy
$E\simeq \epsilon E_{p,b}\simeq1$~PeV and is marginally consistent
with recent upper limits on the diffuse neutrino flux by
IC-40~\cite{IceCubeEHE,Abbasi:2011ji}. Only for very short variation time-scales of
$t_v\simeq10^{-3}$~s (third row of Fig.~\ref{fig:nuspectra}) and
for synchrotron-loss-dominated GRBs ($t_{\rm syn} < t_{\rm dyn}$) can
the predicted flux avoid present upper limits. Note that since
$E_{p,b}\propto\Gamma_i^2/\epsilon_0$ we can shift the peak to higher
energies assuming a stronger Doppler boost of the internal
shocks. However, the condition $E_{p,b}\lesssim E_{\rm min}$ imposed
for the fit requires that $\Gamma_i\lesssim10^{3.5}$ and hence
$\epsilon E_{p,b}\lesssim100$~PeV.
The bottom row of Fig.~\ref{fig:nuspectra} we show the prompt neutrino
spectra for $\Gamma_i=10^3$ and the full range of variability
$10^{-3}<t_v/{\rm s}<0.1$. Due to the strong dependence of the
synchrotron time scale on the Doppler factor, $t'_{\rm syn}\propto
\Gamma_i^7$, the GRB models at the 99\% C.L.~correspond to proton
sources where the maximal energy is limited by the dynamical time. The
range of the flux is consistent with present IC-40 limits. Note,
however, that the opacity of of the GRBs for UHE $p\gamma$
interactions and subsequent emission of CR neutrons is very sensitive
to the Doppler factor,
\begin{equation}
\tau_{p\gamma} \simeq \frac{t'_{\rm dyn}}{t'_\Delta\langle
x_{p\to\Delta}\rangle}\simeq0.1\left(\frac{L_{\gamma,52}}
{\Gamma_{i,2.5}^4t_{v,-2}\epsilon_{0,6}}\right)\,.
\end{equation}
Increasing the average Doppler factor of the GRB from $10^{2.5}$ to
$10^3$ reduces the efficiency of UHE CR emission by two orders of
magnitude. It is hence more likely that the successful candidates of
CR sources are GRBs with lower $\Gamma_i$.
The range of prompt neutrino spectra shown in Fig.~\ref{fig:nuspectra}
spans about two orders of magnitude, a range larger than a previous
study in Ref.~\cite{Guetta:2001cd}. The difference of the results can
be traced back to different assumptions. Firstly, we here assumed that
only neutrons can escape the GRB shock environment and form the
spectrum of UHE CRs. Hence, both, CR and neutrino spectra depend on
the opacity of the source. Secondly, we apply a statistical fit to the
data assuming a wide range of spectral indices. Statistically allowed
CR spectra with a steep injection index $\gamma\gg2$ correspond to GRB
models with a higher bolometric energy density and hence larger prompt
neutrino emission. For large maximal proton energies, $E_{\rm max}\gg
10^{19}$~eV, we can approximate the total local power density in UHE CRs
derived from the best fit to the HiRes data as
\begin{equation}
\dot \varepsilon_{\rm CR} \equiv \int\limits_{E_{p,{\rm b}}}{\rm d}EE
{\mathcal L}_{\rm CR}(0,E) \simeq (1-2)\times 10^{44}~{\rm erg}\,{\rm Mpc}^{-3}
\,{\rm yr}^{-1}\times\begin{cases}\frac{1}{\gamma-2}\left(
\frac{10^{19}{\rm eV}}{E_{p,{\rm b}}}\right)^{\gamma-2}&\gamma>2\,,\\
\ln\left(\frac{E_{\rm max}}{E_{p,{\rm b}}}\right)&\gamma=2\,.\end{cases}
\end{equation}
This value is consistent with the original study by Waxman and
Bahcall~\cite{Waxman:1997ti} assuming a total power density in UHE CRs
of the order of $4\times10^{44}~{\rm erg}\,{\rm Mpc}^{-3}\,{\rm yr}^{-1}$
between $10^{19}$~eV and $10^{21}$~eV assuming a flat spectrum with
power-law index $\gamma=2$. In our approach we allow steeper injection
spectra $\gamma\gg2$ extending down to the break energy $E_{p,{\rm b}}\ll
10^{19}$~eV ({\rm cf}.~Eq.~(\ref{eq:Epbreak})) which is set by the GRB
fireball environment. For very steep injection spectra with $\gamma\simeq2.6$
consistent with our fit at the 99\% C.L.~({\rm cf.}~left panel in
Fig.~\ref{fig:gof}) and $E_{p,{\rm b}} = 4\times10^{16}$~eV this corresponds
to a maximal local power density of the order of $\dot \varepsilon_{\rm CR, {\rm max}}
\simeq 10^{46}~{\rm erg}\,{\rm Mpc}^{-3}\,{\rm yr}^{-1}$.
These large local power densities of steep UHE CR proton spectra are
also consistent with the results of Ref.~\cite{Wick:2003ex}. In contrast
to Ref.~\cite{Wick:2003ex} we include here the limits implied by the
extra-galactic diffuse $\gamma$-ray background inferred by Fermi LAT.
This bound has the strongest effect on strong evolution models with
large power-law index as can be seen in the right panel of
Fig.~\ref{fig:gof}. The corresponding proton fluxes at the 99\% C.L.~are
shown in the right panels of Fig.~\ref{fig:pspectra} and the prompt
neutrino fluxes in the right panels of Fig.~\ref{fig:nuspectra}.
The range of spectra is comparable to the moderate evolution following
SFR (left panels) since the contribution of UHE CRs to the diffuse
extra-galactic $\gamma$-ray background is already close to maximal.
\section{Conclusion}
We have discussed prompt neutrino emission associated with the
production of UHE CR protons in GRBs. Our analysis assumes that UHE
CRs above the ``ankle'' at 4~EeV consists of neutrons emitted from
$p\gamma$ interactions in internal shocks of the GRB fireball
model. We determined the CR emission density by a fit to HiRes data
and used this information to determine the corresponding neutrino
fluxes for a wide range of parameters for the shock environment. We
have also carefully studied the effect of secondary production during
CR propagation in the form of GZK neutrinos as well as the GeV-TeV
$\gamma$-ray background.
The main results are shown in Fig.~\ref{fig:nuspectra}. We have
shown that typical (``benchmark'') fireball environments predict
prompt neutrino fluxes that exceed present diffuse neutrino limits
from IceCube (IC-40) if the associated proton spectrum is fitted to
actual CR data. This remains partially true if we relax the condition
of internal shock parameters. Consistency with the diffuse neutrino
limits requires that the CR acceleration takes place in GRB fireballs
with small internal shock radii (corresponding to variation
time-scales of milli-seconds) and/or large Lorentz boost
$\Gamma_i\simeq1000$. Alternatively, it is possible that UHE CRs at
the ankle might still receive a significant contribution from galactic
CRs, which lowers the required power density in extra-galactic CRs
and the corresponding prompt neutrino spectra from GRB
sources~\cite{Katz:2008xx}. However, one has to worry that galactic
CRs at these energies will reveal themselves via an anisotropy toward
the galactic center. In any case, the sensitivity of the full IceCube observatory
(IC-86) after three years of observation will improve present diffuse
neutrino limits by a factor five and will serve as a crucial test of
the GRB scenario of UHE CRs.
Finally, we would also like to mention additional constraints of
this CR scenario. A fraction of UHE $\gamma$-rays produced in neutral
pion production may escape the fireball shocks and contribute to the
GeV-TeV background as well. As an estimate, we have tested a
pessimistic scenario (``Max'' model) where all $\gamma$-rays from $\pi^0$ production
escape the source. We found that this additional contribution may
become important in the case of source evolution much stronger than
the SFR ({\it cf.}~dashed regions in Figs.~\ref{fig:gof} and \ref{fig:pspectra}).
It is also possible to constrain UHE CR emission from
internal shocks of GRBs by the integrated fluence of all bursts as has
been done recently in~\cite{Eichler:2010ky}. On the other hand, the
contribution of the burst spectrum to the diffuse $\gamma$-ray
background from unidentified cosmic GRBs has been discussed
in~\cite{Dermer:2006vd} and has been shown to contribute a relatively
small fraction.
\section*{Acknowledgments}
The authors would like to thank Peter Meszaros, Kohta Murase, Eli Waxman
and Walter Winter for valuable comments on the manuscript.
This work is supported by US National Science Foundation Grant No
PHY-0969739 and by the Research Foundation of SUNY at
Stony Brook. F.H. is supported by
U.S. National Science Foundation-Office of Polar Program,
U.S. National Science Foundation-Physics Division, and the University
of Wisconsin Alumni Research Foundation. M.C.G-G acknowledges further
support from Spanish MICCIN grants 2007-66665-C02-01,
consolider-ingenio 2010 grant CSD2008-0037 and by CUR Generalitat de
Catalunya grant 2009SGR502.
|
\section{Introduction}
\subsection{Problem}
This paper describes a modification to the kernel of John Harrison's
HOL Light \cite{har:96:3,har:00} proof assistant.
Proof assistants are the best route to
\emph{complete} reliability, both
in abstract mathematics as well as for verification of computer systems.
Among the proof assistants HOL \cite{gor:mel:93} is one of the more popular
systems (next to Coq, Isabelle, PVS and ACL2), and among the
HOL implementations HOL
Light is one of the most interesting ones.
HOL Light has both been used for extensive verification of floating
point algorithms at Intel \cite{har:99,har:06:1}, as well as for impressive formalizations in
mathematics \cite{hal:07,har:08}.
Furthermore, a quite precise model of the HOL Light kernel code
has been formally proved correct \cite{har:06}.
HOL is a direct descendant of the pioneering LCF \cite{gor:mil:wad:79} system from the seventies.
In both LCF and HOL the user is not interacting with the proof
assistant through a
system specific language, but instead interacts directly with
the interpreter of the ML language in which the system has been programmed.
In the case of HOL Light this is the OCaml \cite{ler:08} language of Xavier Leroy.
For this reason in HOL there is no one keeping track of which theorems
still are valid.
Once a statement has been presented to the user as \emph{proved} -- by giving
the user
an element of the abstract datatype `\smalltt{thm}'
as a token of its being proven -- it unavoidably will stay available
to the user as a proved statement.
This approach has the advantage that the abstract datatypes of ML
make it easy for the system to have a small proof checking \emph{kernel}
-- also called \emph{logical core} -- with the property that if
the code of that kernel can be trusted (and it implements
a consistent logic), then it is certain that
the system will be mathematically sound, i.e., it will not be possible
to prove the statement of falsity $\bot$ in it.
However, this approach has the disadvantage that it is difficult to
\emph{undo} state-modifying actions in the system.
In particular, in the existing HOL Light system it is not possible
to change the definition of a defined constant:
\begin{alltt}\small
# let X0 = new_definition `X = 0`;;
val ( X0 ) : thm = |- X = 0
# let X1 = new_definition `X = 1`;;
Exception: Failure "new_basic_definition".
\end{alltt}
\noindent
In current practice, a user of HOL Light
who wants to modify a definition will just reload the whole formalization.
This is generally not a big problem, but can become quite slow.
Also, reloading a formalization that is not finished yet and consists
of bits and pieces that have been loaded manually, can be cumbersome.
There is a good reason that HOL does not have a way of undoing definitions.
Let us suppose it would have a function
\smalltt{undo\char`\_definition}.
Then we could have the following session:
\label{sec:problem}
\begin{alltt}\small
# {let X0 = new_definition `X = 0`;;}
val ( X0 ) : thm = {|- X = 0}
{{#}} {{{undo_definition "X";;}
val it : unit = ()}}
# {let X1 = {new_definition `X = 1`};;}
val ( X1 ) : thm = |- X = 1
# {TRANS (SYM X0) X1;;}
val it : thm = |- 0 = 1
\end{alltt}
\noindent
Of course the `theorem' \smalltt{X0} will no longer be valid
after we undo the definition of \smalltt{X} as \smalltt{0},
but there is no way for us to take away
\smalltt{X0} from the user once he or she has it.
Hence, the system with \smalltt{undo\char`\_definition} clearly
is inconsistent, as one can prove \smalltt{0} \smalltt{=} \smalltt{1} in it.
The problem is that the kernel of HOL Light keeps track of
which constants have been defined already.
It has \emph{state}.
Or, stated differently, it is not purely functional.
To be able to undo definitions of constants in the HOL system,
we will switch to a \emph{stateless} kernel for the system.
\subsection{Approach}
The approach that we will follow is quite simple.
In HOL traditionally constants are \emph{named}, i.e.,
they are identified by a string of OCaml type \smalltt{string}.
We will change this to identifying the constants by the
definition itself.\footnote{
In several systems the definition of a constant consists
of two parts:
first the constant is introduced and then the value of the constant is set.
The approach from this paper does not allow such a separation.
}
That way, we do not need a state anymore to find the properties
of the constant from the name.
Then the properties will \emph{be} the name.
Actually, in our approach we will also include a string in the name
of the constant, both for convenience and for backward compatibility.
This means that the `names' that we will use for constants will
consist of a \emph{pair} of a string and a definition.
It is essential for efficiency reasons that when comparing constants
the strings will be compared first.
For this reason we put the string as the first component of the pair.
We will now look at what this means for the data structures in memory.
Here is what the constant \smalltt{`X`} from the above example
looks like in memory in the traditional stateful HOL Light system:
\begin{center}
\begingroup
\setlength{\unitlength}{1.8pt}
\small
\begin{picture}(135,63)(0,17)
{\put(0,60){\framebox(40,10){}}%
\put(0,60){\makebox(20,10){{\normalsize Const}}}%
\put(20,60){\line(0,1){10}}%
\put(30,60){\line(0,1){10}}}%
\put(25,65){{\circle*{1}}}
\put(35,65){{\circle*{1}}}
\thinlines
\put(5,80){{\vector(0,-1){10}}}
\put(25,65){\vector(0,-1){11}}
\put(25,65){{\circle*{1}}}
\put(15,45){\makebox(20,10){\texttt{"X"}}}
\put(35,65){{\vector(0,-1){25}}}
\put(35,65){{\circle*{1}}}
{\put(30,30){\framebox(40,10){}}%
\put(30,30){\makebox(20,10){{\normalsize Tyapp}}}%
\put(50,30){\line(0,1){10}}%
\put(60,30){\line(0,1){10}}}%
\thinlines
\put(55,35){{\circle*{1}}}
\put(55,35){\vector(0,-1){11}}
\put(55,35){{\circle*{1}}}
\put(44.8,15){\makebox(20,10){\texttt{"num"}}}
\put(60,30){\makebox(10,10){\texttt{[]}}}%
\end{picture}
\endgroup
\end{center}
\noindent
This data structure takes 40 bytes in memory (24 bytes for the two blocks
and 16 for the two strings), so this is not a large data structure.
By hiding the detail of the \smalltt{`:num`} type we can
abbreviate this as:
\begin{center}
\begingroup
\setlength{\unitlength}{1.8pt}
\small
\begin{picture}(135,40)(0,30)
\thinlines
\put(5,68){{\vector(0,-1){8}}}
{\put(0,50){\framebox(40,10){}}%
\put(0,50){\makebox(20,10){{\normalsize Const}}}%
\put(20,50){\line(0,1){10}}%
\put(30,50){\line(0,1){10}}}%
\thinlines
\put(25,55){{\vector(0,-1){11}}}
\put(20,35){\makebox(10,10){{\texttt{"X"}}}}
\put(35,55){{\vector(0,-1){19}}}
\put(28.8,27){\makebox(10,10){{\texttt{`:num`}}}}
\put(25,55){{\circle*{1}}}
\put(35,55){{\circle*{1}}}
\end{picture}
\endgroup
\end{center}
\noindent
Now here is the corresponding data structure for the constant \smalltt{`X`} in
our stateless HOL Light implementation:\footnote{%
We are simplifying reality slightly here.
In HOL Light the number \texttt{0} actually is the
term \texttt{`NUMERAL \char`\_0`}, so it is not a constant as drawn in the picture.
The term \texttt{\char`\_0} however \emph{is} a constant with
definition \texttt{`\char`\_0 = mk\char`\_num IND\char`\_0`}.
Tracing all further definitions from \texttt{mk\char`\_num} and \texttt{IND\char`\_0}
is an interesting exercise that we will not pursue here.}
\begin{center}
\begingroup
\setlength{\unitlength}{1.8pt}
\small
\begin{picture}(135,75)(0,-5)
\thinlines
\put(5,68){{\vector(0,-1){8}}}
{\put(0,50){\framebox(50,10){}}%
\put(0,50){\makebox(20,10){{\normalsize Const}}}%
\put(20,50){\line(0,1){10}}%
\put(30,50){\line(0,1){10}}%
\put(40,50){\line(0,1){10}}}%
\thicklines
\put(40,50){\framebox(10.2,10){}}%
\thinlines
\put(25,55){{\vector(0,-1){11}}}
\put(20,35){\makebox(10,10){{\texttt{"X"}}}}
\put(35,55){{\vector(0,-1){19}}}
\put(28.8,27){\makebox(10,10){{\texttt{`:num`}}}}
\put(25,55){{\circle*{1}}}
\put(35,55){{\circle*{1}}}
\thinlines
\put(45.5,55){{\vector(0,-1){31}}}
\put(45.5,55){{\circle*{1}}}
\put(32.5,9){\makebox(77,15)[l]{{\Large\texttt{`X} \texttt{=} \hspace{6.25em} \texttt{`}}}}
{{\put(50,11){\framebox(50,10){}}}
\thicklines
\put(90,11){\framebox(10.2,10){}}%
\thinlines
\put(50,11){\makebox(20,10){{\normalsize Const}}}%
\put(70,11){\line(0,1){10}}%
\put(80,11){\line(0,1){10}}%
\put(90,11){\line(0,1){10}}%
\thinlines
\put(75,16){{\vector(0,-1){11}}}
\put(70,-4){\makebox(10,10){{\texttt{"0"}}}}
\put(85,16){{\vector(0,-1){13}}}
\put(95,16){{\vector(0,-1){15}}}
\put(75,16){{\circle*{1}}}
\put(85,16){{\circle*{1}}}
\put(95,16){{\circle*{1}}}}
\put(93,-8){\makebox(42,10)[l]{\textbf{{\normalsize definition of }\texttt{`0`}}}}
\end{picture}
\endgroup
\end{center}
\noindent
We added an extra field to the \smalltt{Const} block that points
to the definition of the \smalltt{X} constant.\footnote{%
Note that the `\texttt{X}' that occurs on the left hand side
of the defining equation is a {variable} and \emph{not}
the defined constant.
This equation was the argument to
\texttt{new\char`\_basic\char`\_definition}.
An alternative approach would be to have the new field
just point to the right hand side of the equation.
However, we considered it to be more elegant to have the field point
to the exact data that was given to the function that introduced
the constant to the system.
In this case this is the argument to the function \texttt{new\char`\_basic\char`\_definition}.
}
However, this addition is recursive:
the constant \smalltt{0} occurring in the definition of \smalltt{X}
\emph{also} has this pointer which points to the definition of \smalltt{0}.
This continues all the way until we get to pure lambda terms that
do not involve constants at all.
Only then will this chain of references end.
Clearly, the constant \smalltt{`X`} now is a large
graph consisting of blocks connected by pointers.
Of course for different constants these graphs will not be disjoint.
Therefore, all constants together should be considered to be part
of a single large data structure in memory.\footnote{
Note that although the amount of data for a single constant becomes
much larger, the amount of data for the whole system
stays roughly the same, as the data structures of the
constants will be shared.
}
Despite the size of their data structures, comparison of constants still turns out to be rather cheap.
In our modified system
different constants actually still will have different strings in their names.
Therefore, if constants are \emph{not} equal, the comparison will fail
exactly like it would fail before, i.e., while comparing the strings.
However, if the constants \emph{are} equal, their definitions in fact
also will be equal, and even will be given by equal pointers.
Now OCaml can be made to decide that two things are equal
if the pointers to them are equal.
Hence in the normal use of the HOL Light system
constant comparison will never follow the pointers to the definitions.
However, if we use the \smalltt{undo\char`\_definition} function from
the above discussion, suddenly this does not hold
anymore.
The two \smalltt{X}s, that we defined in the example will then \emph{not} be equal
(despite the fact that their strings and types are equal) because
the definitions will be different.
This then will prevent the `proof' of \smalltt{0} \smalltt{=} \smalltt{1} from working.
Our stateless system is almost identical to the stateful one.
We did not so much replace the kernel, as slightly change it.
In particular we moved the stateful part
outside the kernel.
In a picture:
\begin{center}
\medskip
\begin{tabular}{ccc}
\begin{picture}(120,130)
\thinlines
{\put(0,0){\framebox(120,130){}}%
\put(0,0){\makebox(120,0)[lb]{\strut\ \emph{system}}}}%
\thicklines
{\put(20,20){\framebox(80,100){}}}%
{\put(20,20){\makebox(80,0)[lb]{\strut\ \emph{kernel}}}}%
\put(60,30){{\circle*{3}}}
\put(65,30){\makebox(55,0)[l]{\strut\emph{state}}}
\end{picture}%
&$\hspace{3em}$&
\begin{picture}(120,130)
{\put(0,0){\framebox(120,130){}}%
\put(0,0){\makebox(120,0)[lb]{\strut\ \emph{system}}}}%
{\put(20,20){\framebox(80,100){}}}%
\thicklines
{\put(30,40){\framebox(60,70){}}}%
\thinlines
{\put(30,40){\makebox(60,0)[lb]{\strut\ \emph{kernel}}}}%
\put(60,30){{\circle*{3}}}
\put(65,30){\makebox(55,0)[l]{\strut\emph{state}}}
\end{picture} \\
\noalign{\medskip}
\emph{Stateful {HOL} Light} && \emph{{Stateless {HOL} Light}}
\end{tabular}%
\smallskip
\end{center}
\noindent
Specifically, we split the implementation \smalltt{fusion.ml} of the kernel
(the thick box in the left picture)
into a stateless part \smalltt{core.ml} (the thick box in the right picture) and a stateful part \smalltt{state.ml} (the part outside that box),
as further described in Section~\ref{sec:code}.
This means that our system still presents a fully compatible kernel
to the rest of the system.
Therefore all existing HOL Light developments will still work with the stateless
kernel.
The only code that needs to be adapted is code that probes
the representation of constants.
This hardly every happens, and even then the changes needed are small.
We will discuss these changes in more detail below.
\subsection{Related Work}
The idea from this paper of having definitional information be part of the names of constants
is applied to type theory in \cite{geu:kre:mck:wie:10}.
Many systems use an approach similar to the one that we describe here.
For example the HOL4 system also uses a pointer internally to
distinguish different versions of the `same' constant.
However the HOL4 system currently is not purely functional.
The Matita and Epigram 1 systems use an approach similar to ours
in that they
\emph{reconstruct} the context of a term from the term itself.
Many theorem provers have a purely functional kernel.
This holds of course for all systems written in Haskell,
but for example (since version 7) also for Coq \cite{fil:00}.
However, one might argue that the kernels of these systems are not really without a notion of state,
as in those systems \emph{state} is an object that the
kernel operates on.
In Coq the state object is called an \emph{environment}, while
in HOL-based systems like ProofPower and Isabelle it is called a \emph{theory}.
In Coq, one does {not} have a separate type for well-formed terms:
to be purely functional it departs from the traditional LCF architecture in this respect.
Only a type of \emph{preterm} exists, that can be used to
\emph{extend} a state (at which time the preterm will be type checked).
Specifically in the case of Coq the basic function of the kernel essentially is:\footnote{%
In the actual Coq source code the \texttt{string} type is called \texttt{dir\char`\_path} \texttt{*} \texttt{label},
the \texttt{preterm} type is called \texttt{constr}, and
the \texttt{state} type is called \texttt{safe\char`\_environment}.}
\begin{alltt}\small
add_constant : string -> preterm -> state -> state
\end{alltt}
This function adds a new constant with a given name and expansion to the state.
Clearly the Coq kernel is functional by carrying the state
around in a monadic fashion.
In contrast, in our approach there does not exist a type corresponding to a state.
In ProofPower and Isabelle, \emph{theorems} instead of constants
are tagged, where the tag indicates which theory the theorem belongs to.
Only theorems from compatible theories are accepted by the inference rules
of those systems.
ProofPower has stateful theories, and is quite similar to HOL4.
Isabelle has a purely functional architecture, but for efficiency
reasons implements it in a non-functional way \cite{wen:02:1}.
It uses unique ids to provide an efficient
approximation of the inclusion relation on theory content -- both for
efficiency and for decidability, since a theory may contain arbitrary data
(including ML functions or fully abstract stuff).
The approaches of the various
systems show a trade-off between easy access to the implementation of the
system
(systems like HOL) and ease of navigation of the
formalizations (systems like Coq).
We show that one does not need to sacrifice
undo to get the accessibility of HOL's LCF architecture.
The trick of making the definitions part of the names of constants
occurs in logic regularly.
For instance, in Leisenring's book on the epsilon choice operator \cite{lei:69}
the term $\varepsilon x. P[x]$ takes the place of a constant name
for a witness of $\exists x. P[x]$.
That way the completeness theorem can be proved without having to
Skolemize first nor without having to add new constants.
\subsection{Contribution}
We present a version of the HOL Light system with the following
properties:
\begin{itemize}
\item
The kernel of the system is purely functional.
\item
The system supports undoing definitions in a logically sound way.
\item
The system is fully compatible with the existing HOL Light system:
all existing developments can be run with minor changes.
\item
The system runs at almost the speed of the existing HOL Light system.
\item
The kernel of the system is theoretically easier to analyze.
\end{itemize}
\subsection{Outline}
The paper is structured as follows.
In Section~\ref{sec:data} we explain how we modified the HOL data structures
and changed the kernel accordingly.
In Section~\ref{sec:code} we describe how the code of
the system further had to be modified.
In Section~\ref{sec:undo} we describe how to undo definitions.
In Section~\ref{sec:axioms} we discuss how to also have the kernel
track in a functional way which axioms have been used for which theorems.
In Section~\ref{sec:perf} we present the performance of our system.
Finally, in Section~\ref{sec:concl} we conclude.
\section{The datatypes of the logical core}\label{sec:data}
Here are the datatypes of the stateful HOL Light system (these
are the traditional HOL datatypes):
\begingroup
\def\\{\char`\\}
\def\{{\char`\{}
\def\}{\char`\}}
\begin{alltt}\small
type hol_type =
| Tyvar of string
| Tyapp of string * hol_type list\medskip
type term =
| Var of string * hol_type
| Const of string * hol_type
| Comb of term * term
| Abs of term * term\medskip
type thm =
| Sequent of term list * term
\end{alltt}
\endgroup
\noindent
Of course, these datatypes are \emph{abstract}, i.e., only
the kernel can create data of these types.
This is the essence of the LCF architecture.
In our system, we changed these definitions to the following:
\begingroup
\def\\{\char`\\}
\def\{{\char`\{}
\def\}{\char`\}}
\begin{alltt}\small
type hol_type =
| Tyvar of string
| Tyapp of hol_type_op * hol_type list\medskip
and term =
| Var of string * hol_type
| Const of string * hol_type * const_tag
| Comb of term * term
| Abs of term * term\medskip
and thm =
| Sequent of term list * term\medskip
and hol_type_op =
| Typrim of string * int
| Tydefined of string * int * \underbar{thm}\medskip
and const_tag =
| Prim
| Defined of \underbar{term}
| Mk_abstract of string * int * \underbar{thm}
| Dest_abstract of string * int * \underbar{thm}
\end{alltt}
\endgroup
\noindent
That is, we \emph{tagged} both the constants and the
defined types by the information that was used to introduce
them to the system.
In the case of the types it turned out to be more convenient
to integrate these tags with the string and arity into
a \smalltt{hol\char`\_type\char`\_op}.
The simplest example of this kind of tagging is the \smalltt{Defined}
tag for defined constants.
However the constants for the functions that map between an `abstract' defined
type and the `concrete' type from which it was carved have to be tagged
too.
(Those functions go in opposite directions: see \cite{har:00} for a detailed
description of the HOL type definition architecture.)
For this we have the \smalltt{Mk\char`\_abstract} and \smalltt{Dest\char`\_abstract} tags.
In our implementation as a sanity check we temporarily replaced the underlined types by function types,
to check that the system never followed the pointers corresponding
to these fields
(with this hack we did not have to change any other code in the system).
That is, instead of \smalltt{\underbar{term}} we used
\smalltt{unit} \smalltt{->} \smalltt{term}
and instead of \smalltt{\underbar{thm}} we used
\smalltt{unit} \smalltt{->} \smalltt{thm}.
This works because OCaml will return \smalltt{true} if a function is compared to
itself, while it will throw the \smalltt{Invalid\char`\_argument} \smalltt{"equal:} \smalltt{functional value"} exception if it is compared to any other function.
There are various alternatives for the `tagging data' in these type
definitions.
We chose to use the exact arguments of the functions that are used
to define constants and types.
However, more economical variants are possible.
For example, instead of tagging the definition of a constant \smalltt{X}
with the defining equation \smalltt{`X} \smalltt{=} \dots\smalltt{`}, we could just
have used the body of that definition.
Similarly, in \smalltt{Tydefined}, \smalltt{Mk\char`\_abstract} and \smalltt{Dest\char`\_abstract} we could have used a \smalltt{term} instead of a \smalltt{thm}.
In that way we could have lost the dependency
of \smalltt{term} on \smalltt{thm}.
However, all these variants are really more or less equivalent,
and we chose for the one where it is most obvious where the
`tagging information' originates.
One could also consider using structure-less nonces (\smalltt{unit} \smalltt{ref}s, say) as tags.
However, this makes the system stateful again.
Also, a formal analysis of the system will be harder that way, as in that case the
tags will not be semantically meaningful.
We will now present how we adapted and reorganized the kernel according
to these new datatype definitions.
We only will do this for constants.
Analogous changes had to made for defined types.
We will start by presenting the code before we changed it.
The relevant functions for constants from the original, stateful HOL Light kernel
are:
\begingroup
\def\\{\char`\\}
\def\{{\char`\{}
\def\}{\char`\}}
\begin{alltt}\small
constants : unit -> (string * hol_type) list
definitions : unit -> thm list
get_const_type : string -> hol_type
new_constant : string * hol_type -> unit
new_basic_definition : term -> thm
mk_const : string * (hol_type * hol_type) list -> term
\end{alltt}
\endgroup
\noindent
With \smalltt{new\char`\_basic\char`\_definition} one introduces
a new constant definition to the system,
and with \smalltt{mk\char`\_const} one creates constant terms.
The implementation of these functions is completely straightforward:
\begingroup
\def\\{\char`\\}
\def\{{\char`\{}
\def\}{\char`\}}
\begin{alltt}\small
let bool_ty = Tyapp("bool",[]);;\medskip
let aty = Tyvar "A";;\medskip
let the_term_constants =
ref ["=",Tyapp("fun",[aty;Tyapp("fun",[aty;bool_ty])])];;\medskip
let constants() = !the_term_constants;;\medskip
let the_definitions = ref ([]:thm list);;\medskip
let definitions() = !the_definitions;;\medskip
let get_const_type s = assoc s (!the_term_constants);;\medskip
let new_constant(name,ty) =
if can get_const_type name then
failwith ("new_constant: constant "^name^" has already been declared")$\hspace{-.8cm}$
else the_term_constants := (name,ty)::(!the_term_constants);;\medskip
let new_basic_definition tm =
match tm with
Comb(Comb(Const("=",_),(Var(cname,ty) as l)),r) ->
if not(freesin [] r) then failwith "new_definition: term not closed"$\hspace{-.8cm}$
else if not (subset (type_vars_in_term r) (tyvars ty))
then failwith "new_definition: Type variables not reflected in constant"$\hspace{-.8cm}$
else let c = new_constant(cname,ty); Const(cname,ty) in
let dth = Sequent([],safe_mk_eq c r) in
the_definitions := dth::(!the_definitions); dth
| _ -> failwith "new_basic_definition";;\medskip
let mk_const(name,theta) =
let uty = try get_const_type name with Failure _ ->
failwith "mk_const: not a constant name" in
Const(name,type_subst theta uty);;
\end{alltt}
\endgroup
\noindent
We will now show the corresponding code in our system.
In our stateless version of HOL Light,
the kernel interface becomes simpler.
The only relevant functions now are:
\begingroup
\def\\{\char`\\}
\def\{{\char`\{}
\def\}{\char`\}}
\begin{alltt}\small
new_prim_const : string * hol_type -> term
eq_term : hol_type -> term
new_defined_const : term -> term * thm
inst_const : term * (hol_type * hol_type) list -> term
\end{alltt}
\endgroup
\noindent
with implementation:
\begingroup
\def\\{\char`\\}
\def\{{\char`\{}
\def\}{\char`\}}
\begin{alltt}\small
let bool_tyop = Typrim("bool",0);;\medskip
let bool_ty = Tyapp(bool_tyop,[]);;\medskip
let new_prim_const(name,ty) =
Const(name,ty,Prim);;\medskip
let eq_term ty =
Const("=",Tyapp(Typrim("fun",2),[ty;Tyapp(Typrim("fun",2),[ty;bool_ty])]),$\hspace{-.8cm}$
Prim);;\medskip
let new_defined_const tm =
match tm with
Comb(Comb(Const("=",_,Prim),(Var(cname,ty) as l)),r) ->
if not(freesin [] r) then failwith "new_definition: term not closed"$\hspace{-.8cm}$
else if not (subset (type_vars_in_term r) (tyvars ty))
then failwith "new_definition: Type variables not reflected in constant"$\hspace{-.8cm}$
else let c = Const(cname,ty,Defined tm) in
let dth = Sequent([],safe_mk_eq c r) in
c,dth
| _ -> failwith "new_basic_definition";;\medskip
let inst_const(tm,theta) =
match tm with
| Const(name,uty,tag) -> Const(name,type_subst theta uty,tag)
| _ -> failwith "inst_const: not a constant";;
\end{alltt}
\endgroup
\noindent
The remainder of the code was moved out of the kernel.
These are the following functions:
\begingroup
\def\\{\char`\\}
\def\{{\char`\{}
\def\}{\char`\}}
\begin{alltt}\small
the_term_constants : (string * term) list ref
the_definitions : thm list ref
get_const_type : string -> hol_type
new_constant' : string * term -> unit
new_constant : string * hol_type -> unit
new_basic_definition : term -> thm
mk_const : string * (hol_type * hol_type) list -> term
\end{alltt}
\endgroup
\noindent
(We did not need to hide the stateful variables \smalltt{the\char`\_term\char`\_constants} and \smalltt{the\char`\_def\-ini\-tions} anymore, as
changing them can no longer compromise the logic.
For this reason we no longer need the \smalltt{constants} and
\smalltt{definitions} functions for inspecting them.)
The implementation of these functions again is straightforward:
\begingroup
\def\\{\char`\\}
\def\{{\char`\{}
\def\}{\char`\}}
\begin{alltt}\small
let the_term_constants = ref ["=",eq_term aty];;\medskip
let the_definitions = ref ([]:thm list);;\medskip
let get_const_type s = type_of (assoc s (!the_term_constants));;\medskip
let new_constant'(name,c) =
if can get_const_type name then
failwith ("new_constant: constant "^name^" has already been declared")$\hspace{-.8cm}$
else the_term_constants := (name,c)::(!the_term_constants);;\medskip
let new_constant(name,ty) =
new_constant'(name,new_prim_const(name,ty));;\medskip
let new_basic_definition tm =
let c,dth = new_defined_const tm in
match c with
| Const(name,_,_) ->
new_constant'(name,c); the_definitions := dth::(!the_definitions); dth;;$\hspace{-.8cm}$\medskip
let mk_const(name,theta) =
let tm = try assoc name (!the_term_constants) with Failure _ ->
failwith "mk_const: not a constant name" in
inst_const(tm,theta);;
\end{alltt}
\endgroup
\noindent
Clearly, the code in our system is slightly more complicated,
but essentially it is a reorganized version of the original code.
Our code has the property that makes it easy to distinguish calls
to \smalltt{new\char`\_con\-st\-ant} that add `primitive' constants
to the system -- which in practice only is used for the Hilbert
epsilon choice operator \smalltt{`(@)`} --
from other calls to \smalltt{new\char`\_constant}.
In our system the first kind corresponds to the kernel
function \smalltt{new\char`\_prim\char`\_{\penalty 100}const}.\footnote{%
If one wants to be pedantic, one might keep track of calls to \texttt{new\char`\_prim\char`\_const} (and
to its counterpart for types) in
the \texttt{context} data structure described in Section~\ref{sec:axioms} below.
In some sense these are `axiomatic' too.}
\section{Modifications to the HOL source code}\label{sec:code}
The sizes of the kernel files are compared in the following
table:
\begin{center}
\medskip
\def\omit#1{}
\begin{tabular}{llrr}
& \emph{source file} & $\hspace{.5em}$\emph{all lines} & $\hspace{.5em}$\emph{content} \\
\noalign{\smallskip}
\strut{{\emph{Stateful {HOL} Light}}$\hspace{2em}$} \\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
kernel & {\texttt{\small fusion.ml$\hspace{1em}$}} & {669} & {394} \\
\noalign{\bigskip\medskip}
\strut{{\emph{Stateless {HOL} Light}}} \\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
kernel & {\texttt{\small core.ml}} & \omit{465} & {383} \\
state & {\texttt{\small state.ml}} & \omit{91} & {64} \\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
{total} && \omit{556} & {447}
\end{tabular}
\end{center}
\noindent
The last column counts the number of non-blank non-comment lines.
(In our version of the code we removed all the comments, which means
that the total line count of our files is not meaningful in comparison
to the original system.)
The kernel of the stateful HOL Light is called \smalltt{fusion.ml} (it used to
be three files, \smalltt{type.ml}, \smalltt{term.ml} and \smalltt{thm.ml},
which at some point were fused).
We split this file in our kernel, called \smalltt{core.ml}, and the remainder
of the code, \smalltt{state.ml}.
In the rest of the system only two kinds of changes had to be made.
First, the system would be impractically slow if we compared
objects using the default OCaml equality.
Recently OCaml has been changed to consider Not-A-Number floating point numbers not to be
equal to themselves, and for this reason the default equality test
never uses pointer equality as an optimization.
To get around this problem, we added the following lines at the start of
the first HOL Light source file \smalltt{lib.ml}:
\begin{alltt}\small
let (=) = fun x y -> Pervasives.compare x y = 0;;
let (<>) = fun x y -> Pervasives.compare x y <> 0;;
let (<) = fun x y -> Pervasives.compare x y < 0;;
let (<=) = fun x y -> Pervasives.compare x y <= 0;;
let (>) = fun x y -> Pervasives.compare x y > 0;;
let (>=) = fun x y -> Pervasives.compare x y >= 0;;
\end{alltt}
\noindent
Second, pattern matching on kernel datatypes had to
be changed occasionally.
As an example, in \smalltt{basics.ml} the line
\begin{alltt}\small
Tyapp("fun",[ty1;ty2]) -> (ty1,ty2)
\end{alltt}
had to be changed to
\begin{alltt}\small
Tyapp(Typrim("fun",2),[ty1;ty2]) -> (ty1,ty2)
\end{alltt}
\noindent
In the basic library of the system (which consists of 26,602
lines of source code) there were only 74 lines
that had to be changed like this.
These changes were systematic and could be made with
a few global replacements.
\section{Undoing definitions}\label{sec:undo}
With a stateless kernel implementing safe removal of definitions
becomes trivial.
We just add the following implementation of \smalltt{undo\char`\_definition} to the source file
\smalltt{pair.ml} (right after the implementation of \smalltt{new\char`\_definition}):
\begin{alltt}\small
let undo_definition cname =
the_term_constants := filter ((<>) cname o fst) !the_term_constants;
the_core_definitions := filter ((<>) cname o fst o dest_const o fst o
strip_comb o fst o dest_eq o snd o strip_forall o concl)
!the_core_definitions;
the_definitions := filter ((<>) cname o fst o dest_const o fst o
strip_comb o fst o dest_eq o snd o strip_forall o concl)
!the_definitions;;
\end{alltt}
\noindent
This code has to be in (or after) \smalltt{pair.ml}, because only there the
variable \smalltt{the\char`_definitions} is introduced.
In fact HOL Light has \emph{two} variables with that name,
one in the kernel (in our version of course in \smalltt{state.ml} outside
the kernel), and another one in \smalltt{pair.ml}.
We renamed the first one to \smalltt{the\char`\_core\char`\_defini\-tions},
and update both variables simultaneously.
Now with this function, our motivating example session from Section~\ref{sec:problem} runs as follows:
\begin{alltt}\small
# let X0 = new_definition `X = 0`;;
val ( X0 ) : thm = |- X = 0
# undo_definition "X";;
val it : unit = ()
# let X1 = new_definition `X = 1`;;
val ( X1 ) : thm = |- X = 1
# TRANS (SYM X0) X1;;
Exception: Failure "TRANS".
\end{alltt}
\noindent
As expected the system considers the two \smalltt{X}s to be different,
and does not allow the transitivity step anymore.
However, there still is a subtle issue.
If we now print \smalltt{X0} and \smalltt{X1},
the system will do this in the following way:
\begin{alltt}\small
# X0;;
val it : thm = |- X = 0
# X1;;
val it : thm = |- X = 1
\end{alltt}
\noindent
I.e., the system prints out what appears to be contradictory
judgements.
Of course these judgements are not \emph{actually} contradictory,
the system is perfectly sound.
The \smalltt{X} in the first \smalltt{thm} is the `old' \smalltt{X},
while the second is the `new' \smalltt{X}.
It therefore will \emph{not} be possible to prove from this that
\begin{alltt}\small
val it : thm = |- 0 = 1
\end{alltt}
\noindent
However one might ask, from a pragmatic point of view, how much
difference that makes with the confusing printout of \smalltt{X0}
and \smalltt{X1}.
This is not a problem with the consistency of the system, but
with what in \cite{wie:10} is called \emph{Pollack-consistency}.
There is nothing wrong with the kernel of the system, but with
its printing/parsing code.
The statement of theorem \smalltt{X0} is printed in a way that does \emph{not}
parse back to the same statement.
That is (using the terminology from \cite{wie:10}) the printing/parsing functions are not \emph{well-behaved}.
Of course, in \cite{wie:10} it is pointed out that the
stateful version of HOL Light already was Pollack-inconsistent.
Apparently this was not considered a serious problem, and the problem shown here
might for the same reason be ignored.
However (although we did not pursue this)
in \cite{wie:10} a simple strategy is given
to make a system Pollack-consistent, which can easily be applied here.
A simple variant of this would be to insert in the printing code
some extra lines that check whether a constant that is being printed
is equal to the `current' value of that constant, and if not to throw
an exception.
In that way it is probably easy to have the system print \smalltt{X0}
as `\smalltt{<obsolete} \smalltt{theorem>}' (or something like that)
after the definition of \smalltt{X} has been undone.
For this paper we were mainly interested in
how to minimally modify the \emph{kernel} of the system
to find out what the performance of our approach would be (and not so much
to further develop
the result into a `better' system),
therefore we have not pursued implementing this yet.
\section{Tracking the axioms}\label{sec:axioms}
The stateful HOL Light system keeps track of the axioms
that have been introduced by the user in the variable
\begin{alltt}\small
the_axioms : thm list ref
\end{alltt}
We moved this variable out of the kernel, and therefore
the system described thus far does not keep track of the
axioms that have been used for the theorems.
The whole system only uses three axioms, so one might
not consider this to be a serious problem.
However, we also investigated a variant of the system where
each \smalltt{thm} was `tagged' with the set of axioms
from which it was derived.
In that case each basic inference rule of the system had to
take the union of this set of axioms for each of the \smalltt{thm}s
that it got as arguments.
If implemented naively this would become expensive, computationally.
To get this to run with acceptable speed, we used the following
data structure\footnote{%
One of the referees of this paper pointed out that the use of the \texttt{array} type introduces
state to the kernel again, and that
this undermines the point of the paper a bit.
However, note that we use the arrays in a `purely functional' way.
We never update arrays, and only use them to be able to get to a specific index
without having to walk a list.
}.
The \smalltt{thm} type now is defined as
\begin{alltt}\small
type context =
int * term list array\medskip
type thm =
| Sequent of context * term list * term
\end{alltt}
\noindent
The context type represents the axioms used in proving the \smalltt{thm}.
It consists of an array holding the \emph{history} of the axiom lists
during the execution of the system.
Specifically it consists of an array of lists of axioms of decreasing length,
prefixed with the length of the array minus one.
The function that introduces an axiom now is:
\begin{alltt}\small
let axiom_sequent ((n,axa) as ctx) tm =
let axl = Array.get axa 0 in
let ctx' = (n + 1),Array.of_list ((tm::axl)::Array.to_list axa) in
let ax = Sequent(ctx',[],tm) in
ax,ctx';;
\end{alltt}
Here the \smalltt{(n,axa)} argument represents the set of axioms thus far.
This is given by the stateful outside of the kernel.
The code to merge contexts is:
\begin{alltt}\small
let empty_context = 0,[|[]|];;\medskip
let merge_contexts ((n1,axa1) as ctx1) ((n2,axa2) as ctx2) =
if ctx1 == ctx2 then ctx1 else
if n1 < n2 then
if Array.get axa1 0 = Array.get axa2 (n2 - n1) then ctx2 else
failwith "merge_contexts" else
if n1 > n2 then
if Array.get axa1 (n1 - n2) = Array.get axa2 0 then ctx1 else
failwith "merge_contexts" else
failwith "merge_contexts";;
\end{alltt}
\noindent
This code, when given two contexts, does not
have to walk those contexts to see whether one is a prefix
of the other (which would cost linear time), but instead uses
the array data structure together with pointer equality
to check whether the two contexts match (taking constant time).
With this code only `compatible' contexts, where one is a subset of the
other, can be merged.
Of course a more refined version of this code, that also is able to merge sets of axioms
that are incompatible, could be written.
With this code, the `set of axioms' for theorems always will be subsets
of each other.
We call this version of the system \emph{with linear tracking of the axioms}.
We were curious whether maybe there was a theorem that, for instance,
only needed the first and third axioms but not the second one.
For this reason, we made yet another modification to the code,
that kept track of the \emph{exact} set of axioms used.
This version is called \emph{with precise tracking of axioms}.
In this version of the system we represented the set of axioms as a bit string
in a 32 bit integer.
(This version of the kernel therefore only can handle 32 axioms.
As the actual system only uses 3 axioms, for the experiment this
was sufficient.)
Now the \smalltt{context} type is:
\begin{alltt}\small
type context =
(int * term list array) * int32
\end{alltt}
and the \smalltt{merge\char`\_context} code used OCaml's
\smalltt{Int32.logor} function on the \smalltt{int32} bitstrings.
The result of this experiment however turned out to be that for \emph{none} of the
theorems in the basic library of HOL Light that are named by a global
OCaml variable, a set of axioms is used that is not
just a prefix of the list of the three axioms in the system.
Therefore this refinement of the kernel turned out not to be very useful.
\section{Performance}\label{sec:perf}
We measured the performance of our modified HOL Light versions,
but only using wall clock time.
Specifically we used the following code in an OCaml session:
\begin{alltt}\small
#load "unix.cma";;
let starttime = ref (Unix.time());;
#use "hol.ml";;
Unix.time() -. !starttime;;
\end{alltt}
Here are the results on an unloaded Debian Etch system, using a computer with
a single $\mathop{1.86}\mathord{\mbox{GHz}}$ Intel Pentium M processor.
\begin{center}
\begin{tabular}{lrr}
\emph{version} & \emph{running time} & $\hspace{.8em}$\emph{increase} \\
\noalign{\smallskip}
\hline
\noalign{\smallskip}
Stateful & 113$\,$s \\
Stateless, without tracking of axioms & 130$\,$s & ${} + 15\%$ \\
Stateless, with linear tracking of axioms & 131$\,$s & ${} + 16\%$ \\
Stateless, with precise tracking of axioms & 132$\,$s & ${} + 17\%$ \\
\end{tabular}
\end{center}
\noindent
These are the times needed to load the basic HOL Light library.
Of course, these numbers not only represent the time spend by the
HOL Light system.
For instance, displaying the output of the system in a terminal window
already takes around 10 seconds.
Still, the table gives a reasonable impression of the performance
of the approach promoted in this paper.
\section{Conclusion}\label{sec:concl}
\subsection{Discussion}
Switching HOL Light to our stateless kernel architecture
has advantages and disadvantages:
The advantages are:
\begin{itemize}
\item
One gets the possibility to implement a function \smalltt{undo\char`\_definition}
in a logically sound way (and a similar function for type definitions).
\item
One gets a system that is probably easier to analyze theoretically.
John Harrison's HOL in HOL formalization \cite{har:06}, in which he proves
his kernel source code sound, currently leaves out the
definitions.
We expect that it will be easier to extend that work to include definitions
for our version of the system, than it would be to do this for the stateful version
of HOL Light.
It might seem that a system with 3 non-mutual datatypes is easier
to analyze than a system with 5 mutually defined datatypes.
However, it is much less difficult to reason
about a purely functional program than about a stateful program.
This more than compensates for the slightly more involved datatypes.
In the stateful version of HOL Light the semantics of data from the kernel depend
on the state of the system, and often data has to be interpreted
in a different state than in which it was created.
This makes it hard to give the semantics in a compositional
way.
In contrast, in the stateless version all data has a direct interpretation,
making analysis much simpler.
One subtlety with proving correctness of our stateless kernel
is that it might be difficult to represent pointer equality in such a proof.
Pointer equality is rarely considered in mechanized correctness proofs of functional
programs.
However, it is clear that pointer equality is just an optimization of structural
equality and
that proving the correctness of the kernel using structural equality could be
considered equivalent.
\end{itemize}
\noindent
The disadvantages of our system compared to the stateful HOL Light are:
\begin{itemize}
\item
The system runs at about 85\% of the normal speed in daily use.
\item
The kernel is more complicated.
In particular the kernel datatype
definitions are more involved.
\end{itemize}
\noindent
We are undecided whether the slowness and added complexity in the
kernel outweighs the nicety of having a purely functional kernel
that supports undo.
\subsection{Availability}
A version of the system described in this paper can be
downloaded on the web at the web address:
\begin{alltt}\small
http://www.cs.ru.nl/~freek/notes/hol_light-stateless.tar.gz
\end{alltt}
This tar file contains just the basic library of the system,
adapted for the stateless kernel.
For reference, the tar file also contains the unmodified source code
of the HOL Light version that we used for the experiment.
\subsection{Future work}
We mainly did this experiment to satisfy our curiosity,
to find out whether the approach was viable.
We were surprised that it all worked as painlessly as it did.
We did not argue here why our changes in the implementation
are sound.
Although this seems rather obvious, it would be good to have a formal
analysis of this.
An interesting way to do this would be to adapt John Harrison's
HOL in HOL proof to also include definitions
using our stateless variant of the code.
We would like the main version of
the HOL Light system to adopt our stateless variant.
In that sense, this article can be considered an open letter
to John Harrison, asking him to consider doing this.
\paragraph{Acknowledgments.}
Thanks to Pierre Corbineau for interesting discussions on state in
the HOL Light kernel.
Thanks to Jean-Christophe Filli\^atre, Rob Arthan and Makarius Wenzel
for details of the
architecture of the Coq, ProofPower and Isabelle kernels.
Thanks to Randy Pollack for pointing out the Pollack-inconsistency
issue addressed in Section~\ref{sec:undo}.
Thanks to anonymous referees for helpful comments.
|
\section{Introduction} This note is based on the following observation regarding the existence of a non-vanishing genus zero Gromov-Witten invariant with a least one point as constraint, namely \emph{symplectic uniruledness} (see Definition \ref{sympunirul}), of closed connected Hamiltonian fibrations over symplectic bases,
$$(F^{2n_F},\omega)\stackrel{\iota}{\hookrightarrow} P^{2n_P}\stackrel{\pi}{\rightarrow} (B^{2n_B},\omega_B),$$
which are \emph{rationally cohomologically split}, i.e. where the map $\iota$ induces an injective map in rational cohomology.
The observation can be stated as follows:
\vspace{0.3cm}
\begin{itemize}\item[\textbf{(O)}:] \emph{For $c$-splitting Hamiltonian fibrations, symplectic uniruledness of the base is sufficient to ensure symplectic uniruledness of the total space.}
\end{itemize}
\vspace{0.3cm}
The $c$-splitting hypothesis actually holds in many cases \cite{LM},\cite{Bl},\cite{Kedra},\cite{H}. It is in fact conjectured by McDuff and Lalond, \cite{LM}, that every Hamiltonian fibration is rationally cohomology split.
Although we think \textbf{(O)} should hold in full generality, we only provide a proof under the following technical assumptions. First, we assume that $(F,\omega)$ is \emph{semi-positive relative to the total space}, i.e. $(F,\omega)$ verifies condition \eqref{ssp} described in \textsection 2.3.2. This condition, verified by Hamiltonian fibrations of real dimension at most six, is a fibered analog of the "standard" semi-positivity condition used in \cite{MS}.
Secondly, we assume that the base is symplectically uniruled for a class $\sigma_B\in H_2(B;\mathbb{Z})$ which \emph{only admits simple decompositions for some $\omega_B$-tame almost complex structure $J_B$ on $B$}. This condition, explicited in Definition \ref{irred}, guaranties that, for suitable almost complex structure $J_P$ on $P$ lifting $J_B$, there are no simple stable $J_P$-holomorphic maps in $P$ projecting to a non-simple stable $J_B$-holomorphic maps in $B$ representing $\sigma_B$. This is in particular realized by primitive classes admitting pseudo-holomorphic representatives. Here is our main result:
\begin{theorem}\label{mainobservation} Let $(F,\omega)\stackrel{\iota}{\hookrightarrow} P\stackrel{\pi}{\rightarrow} (B,\omega_B)$ be a $c$-splitting Hamiltonian fibration. Assume $(F,\omega)$ is semi-positive relatively to $P$ and that $(B,\omega_B)$ is symplectically uniruled for some class $\sigma_B\in H_2(B;\mathbb{Z})$ admitting only simple decompositions, then $P$ is also symplectically uniruled.
\end{theorem}
This generalizes a result in \cite{H} stating that symplectic uniruling holds for total spaces of Hamiltonian fibrations over symplectically rationally connected base (for a class $\sigma_B$ as above), i.e. when the base has a non-vanishing genus zero Gromov-Witten invariant with a least two points as constraints. However, in that latter case, no $c$-splitting assumption is required.
One of the main ingredients of the proof of Theorem \ref{mainobservation} is to use \emph{fibered almost complex structure on $P$} in order to relate pseudo-holomorphic maps in $P$ with pseudo-holomorphic maps in $B$. Roughly speaking, a fibered almost complex structure $J_P$ on $P$ is an almost complex structure lifting some $\omega_B$-tame almost complex structure $J_B$ on $B$, and preserving both horizontal and vertical directions in some splitting of $TP$ induced by a Hamiltonian connection on $P$ (cf Definition \ref{fiberedacs}). This geometrical setup enables to relate, generically, some $GW$-invariant of $P$ with some $GW$-invariant of $B$ via a product-type formula involving $GW$-invariants of some Hamiltonian fibration over $S^2$. The technical assumptions of the theorem are there to ensure that transversality can be realized within the realm of fibered almost complex structures.
In a sense, \textbf{(O)} is complementary to the following result of Ruan and Li:
\begin{theorem}\label{SympDiv}(\cite{RuanLi}, Proposition 2.10) If $(F,\omega)\stackrel{\iota}{\hookrightarrow}P\stackrel{\pi}{\rightarrow} B$ is a $c$-splitting Hamiltonian fibration, then $P$ is symplectically uniruled for a class $\sigma\in H_2(P;\mathbb{Z})$ such that $\pi_*\sigma=0$ if and only if $(F,\omega)$ is symplectically uniruled.
\end{theorem}
Note that the theorem above holds in great generality. It is believed that the \emph{ad hoc} hypothesis of Theorem \ref{mainobservation} can be ruled out using virtual perturbations (see \cite{CL}, \cite{LT}, \cite{Rvirt}, amongst others). The problem is to show that one can make perturbations compatibly with the fiber bundle projection $\pi$. Removing these assumptions is part of a joint work in progress with Shengda Hu.
In some circumstances, we can remove the assumption on $\sigma_B$. For instance, it can be removed if the base is a smooth projective manifold $(B,J_B,\omega_B)$. In such case, Ruan and Koll\`{a}r showed that \emph{strong symplectic uniruledness} is equivalent to \emph{projective uniruledness}, i.e. that there is a rational curve passing through every point of $B$. The following proposition then follows from Theorem \ref{mainobservation}.
\begin{cor} \label{cormainobs} Let $(F,\omega)\stackrel{\iota}{\hookrightarrow} P\stackrel{\pi}{\rightarrow} B$ be a $c$-splitting Hamiltonian fibration over a smooth projective manifold $(B,J_B,\omega_B)$. Assume $(F,\omega)$ is semi-positive relatively to $P$ and that $(B,J_B,\omega_B)$ is uniruled, then $P$ is strongly symplectically uniruled.
\end{cor}
In view of Theorem \ref{mainobservation} and Theorem \ref{SympDiv}, it is natural to ask the following question: "If neither $F$ or $B$ is symplectically uniruled, can we conclude that $P$ is not symplectically uniruled ?". We give a partial answer:
\begin{theorem} \label{nouniruling} Assume $(B,J_B,\omega_B)$ is a smooth projective manifold and that $P$ c-splits. Suppose that neither $(B,J_B,\omega_B)$, nor $(F,\omega)$, is symplectically uniruled, then $P$ is not symplectically uniruled for any (generic) fibered almost complex structure $J_P$ on $P$ lifting $J_B$.
\end{theorem}
\noindent\emph{Proof:}\,\,\,\,\,
Assume by contradiction that $P$ is symplectically uniruled for some class $\sigma\in H_2(P;\mathbb{Z})$. Two possibilities may occur: $\pi_*\sigma$ is zero or not. In case $\pi_*(\sigma)=0$, we can use the c-splitting hypothesis to apply Theorem \ref{SympDiv} and obtain a contradiction. Now, let's consider the case where $\pi_*\sigma\neq 0$. Since $P$ is symplectically uniruled, through every point $p\in P$ there exists a $\sigma$-rational $J_P$ pseudo-holomorphic map $u$ passing through $p$. Since $J_P$ is fibered, the map $u_B:=\pi\circ u$ is $J_B$ pseudo-holomorphic. Moreover, $u_B$ passes through $\pi(p)$. Consequently, $B$ is covered by rational curves. Since $B$ is projective this implies that $B$ is symplectically uniruled; thus we have a contradiction.
\qed\\
Before proceeding to the applications, we make a remark concerning possible generalizations of Theorem \ref{mainobservation}, when we replace symplectic uniruledness by \emph{k-symplectic rational connectedness} with $k>1$ (cf Definition \ref{sympunirul}). Roughly speaking, a symplectic manifold is said to be $k$-symplectically rationally connected if there exists a non-vanishing Gromov-Witten invariant with $k$ points as constraints.
It is easy to find examples where Theorem \ref{SympDiv} and observation \textbf{(O)} do not generalize to $k>1$, for instance when $P$ is a trivial symplectic fibration. Nevertheless, it is possible to combine symplectic rational connectedness of the fiber and of the base in order to obtain symplectic rational connectedness of the total space. Before making a precise statement, recall that the relative semi-positivity condition on $(F,\omega)$ implies that the fiber is actually semi-positive. It is well-known that under such hypothesis, the \emph{Quantum homology} of $(F,\omega)$ is generically well-defined. This quantum homology is endowed with a ring structure called \emph{quantum product}, which can be seen as a deformation of the intersection product in homology. We introduce the following notation: let $[pt]$ be the homology class of a point in $F$, we write $[pt]_Q$ to denote the class of the point seen as an element of the Quantum homology of $F$. Also, let $[pt]^k_Q$ denote the $k$'th power of $[pt]_Q$ with respect to the quantum product.
\begin{theorem}\label{generalization} Suppose $(F,\omega)$ is semi-positive relatively to $P$, and assume that $[pt]_Q^{l}\neq0$, $l>1$. Furthermore, suppose that $(B,\omega_B)$ is $(l+1)$-symplectically rationally connected for some class $\sigma_B\in H_2(B,\mathbb{Z})$ admitting only simple decompositions. Then $P$ is $l$-symplectically rationally connected.
\end{theorem}
Next, we give some applications of Theorem \ref{mainobservation} to the Weinstein Conjecture for separating hypersurfaces, or in shorter terms the \emph{$shW$-Conjecture}.
A. Weinstein conjectured in 1979 that: "\emph{Every closed hypersurface $S$ of contact type in a given symplectic manifold $(X,\omega)$ carries a closed characteristic}." \cite{Wein}. Since Viterbo's proof of the conjecture for $(\mathbb{R}^{2n},\omega_0)$ in 1986 \cite{V}, many results followed. In particular, H. Hofer and C. Viterbo highlighted in \cite{HV} the strong interplay between genus zero Gromov-Witten invariants and the conjecture, in cases where the hypersurface $S$ \emph{separates} $X$, i.e. when there exist submanifolds $X^+$ and $X^-$ of $X$ having common boundary and such that $X=X^+\cup X^-$ and $S=X^+\cap X^-$. Note that this latter condition is realized whenever $H^1(X,\mathbb{Z}_2)=0$. Shortly after, these results were extended by G. Liu and G. Tian in \cite{LiuT}, to any symplectic manifold and any genus. In 2000, using the results of Liu and Tian, G. Lu proved the following:
\begin{theorem}\label{LuTh}(G. Lu \cite{Lu}, Corollary 3) Any separating hypersurface of contact type in a symplectically uniruled closed symplectic manifold $(B,\omega_B)$ admits a closed characteristic.
In particular, the $shW$-Conjecture holds in products $(B\times F,\omega_B\oplus\omega)$ of a symplectically uniruled closed symplectic manifold $(B,\omega_B)$ with any symplectic manifold $(F,\omega)$.
\end{theorem}
As a direct consequence of the first assertion of Theorem \ref{LuTh} and of Theorem \ref{mainobservation} we obtain the following generalization to non-trivial Hamiltonian fibrations of the second assertion of Theorem \ref{LuTh}:
\begin{cor}\label{WeinFib} The $shW$-Conjecture holds in $c$-splitting Hamiltonian fibrations with relatively semi-positive fiber and with symplectically uniruled base for some class admitting only simple decompositions.
\end{cor}
As a consequence, we obtain the following which proof is given in Section 2:
\begin{cor}\label{WeinFib2} The $shW$-Conjecture holds in Hamiltonian fibrations with relatively semi-positive fiber and with 2-symplectically rationnally connected base for some class admitting only simple decompositions.
\end{cor}
In particular, the $shW$-Conjecture holds in Hamiltonian fibrations (with relatively semi-positive fiber) over $(\mathbb{C} P^n,\omega_{FS})$ where $\omega_{FS}$ denotes the Fubini-Study Kahler form, and over $(S^2\times S^2,\omega\oplus\omega)$ where $\omega$ is an area form on $S^2$. It also holds for $c$-splitting Hamiltonian fibrations over ruled surfaces. Finally it should also hold for $c$-splitting Hamiltonian fibrations over blow-ups of such bases, since symplectic uniruledness is a symplectic cobordism invariant \cite{HLR}.\\
This note is organized as follows: in Section 1 we introduce the notions and notations needed for the proofs of the results, and in Section 2 we proceed with the proofs. \\
\noindent\textbf{Aknowledgement.} I would like to thank Fran\c{c}ois Lalonde and Octav Cornea for their useful comments and their encouragements. I would also like to thank Fran\c{c}ois for his constant support and generosity. Finally, I thank R\'emi Lelclercq for his suggestions and comments on an earlier version of this note, that helped improve the presentation of the paper.
\section{framework and tools}
\subsection{Gromov-Witten invariants} Let $J$ be an $\omega$-tame almost complex structure on $(X^{2n_X},\omega)$, i.e. $J$ is a smooth endomorphism of $TX$ such that
$$J^2=-Id_{TX} \hspace{0.5cm}{0.2cm} \text{and}\hspace{0.5cm}{0.2cm}\forall v\in TX, \omega(v ,J v)>0.$$
The set of all such almost complex structures is non-empty and contractible. We denote by $c_1^{TX}$ the associated first Chern class of $TX$, and say that $(X,\omega)$ is \emph{semi-positive} if and only if there are no spherical class $A\in H_2(X;\mathbb{Z})$ with positive area $\omega(A)$ such that
\begin{equation*}3-n_X\leq c_1^{TX}(A)< 0.
\end{equation*}
We begin by recalling the definition of genus zero Gromov-Witten invariant for $(X,\omega)$ assuming it is semi-positive. We follow the expositions of \cite{MS}, Chapter 7, or \cite{RT}. At the end we introduce the definition of simple decomposability.
\subsubsection{Gromov-Witten invariants} Recall that a genus zero $J$-holomorphic map of $X$ is a smooth map $u:S^2\rightarrow X$ solution to the Cauchy-Riemann equation:
$$\ov{\partial}_Ju:=1/2(du+J\circ du\circ j_0)=0,$$
where $j_0$ stands for the standard complex structure on $S^2\cong\mathbb{C}\cup\{\infty\}$.
Roughly speaking, a genus zero Gromov-Witten invariant of $X$ is a count, up to reparametrizations of the domain, of $J$-holomorphic rational maps with marked points with values in $X$, satisfying some prescribed constraints at the marked points. More precisely, consider the moduli space of unparametrized genus zero $J$- holomorphic maps in $X$ with $l$ marked points and representing the class $A$
\begin{eqnarray*}\mathcal{M}_{0,l}(X,A,J)&:=& \Big\{(u,\mathbf{x}):=(u,x_1,...,x_l)\in C^{\infty}(S^2,X)\times (S^2)^l | x_i\neq x_j\hspace{0.5cm}{0.2cm}\text{if}\hspace{0.5cm}{0.2cm} i\neq j, \\
& &\,\,\,\,\left. \ov{\partial}_Ju=0\hspace{0.5cm}{0.2cm},\hspace{0.5cm}{0.2cm} [u(S^2)]=A \Big\}\right/PSL_2(\mathbb{C}).
\end{eqnarray*}
Let $\overline{\mathcal{M}}_{0,l}(X,A,J)$ denote its compactification, in the sense of Gromov, which coincides with the moduli space of isomorphism classes $[\Sigma,u,\mathbf{x}]$ of stable $J$-holomorphic maps with $l$ marked points that represent $A$.
This is a stratified space with finitely many strata. Its top stratum is the subset $\mathcal{M}_{0,l}^{*}(X,A,J)\subset \mathcal{M}_{0,l}(X,A,J)$ consisting of \emph{simple $J$-holomorphic maps}, i.e. maps that are somewhere injective. This stratum is, for generic $J$, a naturally oriented manifold with dimension $2n_X+2c_1^{TX}(A)-6+2l$.
There are two natural maps defined on $\overline{\mathcal{M}}_{0,l}(X,A,J)$, namely the \emph{evaluation at the marked points map}
$$ev^X_{l,J}:\overline{\mathcal{M}}_{0,l}(X,A,J)\rightarrow X^l,\hspace{0.5cm}{0.2cm} [\Sigma,u,x_1,\ldots,x_l]\mapsto (u(x_1),\ldots,u(x_l)),$$
and, for $l\geq3$, the \emph{forgetful-map map}
$$\mathfrak{f}:\overline{\mathcal{M}}_{0,l}(X,A,J)\rightarrow \overline{\mathcal{M}}_{0,l}$$
assigning to every stable map the underlying reduced stable curve.
Formally, when $l\geq3$, Gromov-Witten invariants are the values of a multilinear homomorphism
\begin{equation*}\langle\cdot\rangle^X_{A,l}:H_*(\overline{\mathcal{M}}_{0,l},\mathbb{Q})\otimes (H_*(X,\mathbb{Q}))^{\otimes^{l}}\rightarrow \mathbb{Q}
\end{equation*}
where
\begin{equation*}\langle D;\beta_1,...,\beta_{l}\rangle^X_{A,l}:=\int_{\overline{\mathcal{M}}_{0,l}(X,A,J)}(ev^X_{l,J})^*\left(PD_{X^l}(\beta_1\otimes ...\otimes \beta_l)\right)\cup \mathfrak{f}^*PD_{\overline{\mathcal{M}}_{0,l}}(D)
\end{equation*}
which is set to be zero unless:
\begin{equation*}\label{mathdim} 2n_X+2c_1^{TX}(A)=\sum_{i=1}^l (2n_X-\deg(\beta_i))-\deg(D).
\end{equation*}
When $l<3$, there is no forgetful map map and we simply "integrate" the pull-back under $ev^X_{l,J}$ of the product of the $PD(\beta_i)$; in that case we use the notation $\langle \beta_1,...,\beta_{l}\rangle^X_{A,l}$, which could be viewed as $\langle [\overline{\mathcal{M}}_{0,l}];\beta_1,...,\beta_{l}\rangle^X_{A,l}$ if we consider $\overline{\mathcal{M}}_{0,l}$ as a manifold of negative dimension $2l-6$.
The integration above has to be understood as evaluation of the cohomology class with the fundamental cycle of $\overline{\mathcal{M}}_{0,l}(X,A,J)$. However, this space may not carry a fundamental class. In fact, lower strata in $\overline{\mathcal{M}}_{0,l}(X,A,J)$ may have dimensions greater than $\mathcal{M}_{0,l}^{*}(X,A,J)$, due to the possible presence of stable maps with multiply covered components. This problem does not show up when the manifold is semi-positive. Concretely, this condition imposes that the "boundary component" $\overline{\mathcal{M}}_{0,l}(X,A,J)\backslash\mathcal{M}^*_{0,l}(X,A,J)$ is generically of codimension at least two with respect to the top stratum. Then, consider cycles $V_1,...,V_l$ in $X$ representing the $\beta_i$'s, and a cycle $D$ in $\overline{\mathcal{M}}_{0,l}$ representing $D$. Assume that these cycles are in general position, and such that
$\mathfrak{f}\times ev^X_{l,J}$ is strongly transverse to the product $D\times V_1\times ...\times V_l$. The corresponding \emph{Gromov-Witten invariant} is realized as the intersection number $(\mathfrak{f}\times ev^X_{l,J}). (D\times V_1\times ...\times V_l)$ (which can be seen to be $\mathbb{Z}$-valued in that case). In particular, $GW$-invariants generically count simple maps.
\subsubsection{On the semi-positivity assumption.} We should mention that the semi-positivity assumption can be removed using virtual perturbations of the Cauchy-Riemann equation as in \cite{LT},\cite{RT},\cite{Rvirt}, amongst others. However, we will not work in such generality. Also, let us emphasize that the semi-positivity assumption can be dropped if the boundary of $\overline{\mathcal{M}}_{0,l}(X,A,J)$ only consists of \emph{simple $J$-holomorphic stable maps}. Namely, we say that a stable $J$-holomorphic map is \emph{simple} if and only if it has no non-constant multiply covered component and no two non-constant components having the same image in $X$. This leads to the following definition,
\begin{defn}\label{irred} Let $A\in H_2(X;\mathbb{Z})$ be a spherical class representable by $J$-holomophic maps. We say that $A$ \emph{only admits simple decompositions} if every stable $J$-holomorphic map in $\overline{\mathcal{M}}_{0,l}(X,A,J)$ has no non-constant multiply covered component and no two non-constant components having the same image in $X$.
\end{defn}
This is what we request from $\sigma_B$ in Theorem \ref{mainobservation}. Note that the set of $\omega$-tame almost complex structures $J$ with respect to which a given class $A$
only admits simple decompositions is open in the set of $\omega$-tame almost complex structures. However, we cannot make sure that this set is connected or non-empty.
\subsection{Symplectic rational connectedness.} Using the notations introduced in the preceding paragraph, we define the notion of \emph{$k$-symplectic rational connectedness}, following \cite{HLR} and \cite{Lu}.
\begin{defn}\label{sympunirul} Let $k>0$ be an integer, and let $\sigma\in H_2(P;\mathbb{Z})$ be a spherical homology class. A symplectic manifold $(X,\omega)$ is \emph{$k$-symplectically rationally connected for $\sigma$}, or simply \emph{k-SRC} for $\sigma$, if and only if there exist classes $\beta_{k+1},...,\beta_l\in H_*(X,\mathbb{Q})$, and $D\in H_*(\overline{\mathcal{M}}_{0,l},\mathbb{Z})$ such that:
\begin{equation*}\langle D; [pt],...,[pt],\beta_{k+1},...,\beta_l\rangle_{A,l}^X\neq 0.\end{equation*}
we will say that $(X,\omega)$ is \emph{$k$-symplectically rationally connected} if there exists $\sigma$ such that it is \emph{k-SRC} for $\sigma$.
If $k=1$, we will say that $(X,\omega)$ is \emph{symplectically uniruled} or \emph{SU}.
Furthermore, if $(X,\omega)$ is symplectically uniruled and $l=3$, we say that $(X,\omega)$ is \emph{strongly symplectically uniruled} or \emph{SSU}.
\end{defn}
In particular, $(X,\omega)$ is $k$-SRC for $A$ only if, through every $k$ generic points of $X$, there is a genus zero pseudo-holomorphic map representing $A$. The converse may not be true in general. Nevertheless, the equivalence holds in smooth projective varieties.
\begin{theorem}(Ruan \cite{Rvirt},Koll\`{a}r \cite{Kollar1})\label{unirulingproj} A smooth projective variety is symplectically uniruled if and only if it is \emph{projectively uniruled}, i.e. through every point of the manifold there is a holomorphic map.
\end{theorem}
Actually, it follows from the \emph{splitting axiom for Gromov-Witten invariants} that a projective manifold is (projectively) uniruled if and only if it is strongly symplectically uniruled, as pointed out by Ruan \cite{Rvirt}. The main ingredient of the proof is the following property of rational curves in projectively uniruled manifolds $(X,J,\omega)$ due to J. Koll\`{a}r, Y. Miyaoka, and S. Mori (see Koll\`{a}r \cite{Kollar}, Theorem 3.11):
for a very general point $b\in X$, if $u:\mathbb{C} P^1\rightarrow X$ is a morphism such that $[u(\mathbb{C} P^1)]\neq 0$ and $u(0)=b$, then $H^1(\mathbb{C} P^1,u^*TX)=0$. Such morphism is said to be \emph{free (over 0)}. Now, for $b\in X$ and a spherical class $\sigma$, let $\mathcal{M}(X,\sigma,J;b)$ denote the moduli space of $\sigma$ rational $J$-holomorphic maps $u:\mathbb{C} P^1\rightarrow X$ such that $b\in \text{Im}(u)$. We will say that $\sigma$ is \emph{free} if, for every general point $b\in X$, the moduli space
$\mathcal{M}(X,\sigma,J;b)$ only consists of free morphisms. For our purpose, we mildly refine the statement of Theorem \ref{unirulingproj} by the following straightforward observation.
\begin{lem}\label{projuni} If a smooth projective variety $(X,J,\omega)$ is uniruled, it is strongly symplectically uniruled for a free class $\sigma$ admitting only simple decompositions.
\end{lem}
\noindent\emph{Proof:}\,\,\,\,\, Since $X$ is uniruled, for every $b\in X$ there exists a morphism $u:\mathbb{C} P^1\rightarrow X$ such that $b\in \text{Im}(u)$. Now, fix a sufficiently general point $b$. In the proof of Theorem \ref{unirulingproj} given in \cite{Rvirt}, Ruan shows the equality between
\begin{eqnarray*}N_1:=\min\left\{\omega(\sigma)>0| \mathcal{M}(X,J,\sigma;b)\neq \emptyset\right\}
\end{eqnarray*}
and
\begin{eqnarray*}N_2:=\min\left\{\omega(\sigma)>0| \exists \alpha_1,...,\alpha_l\in H_*(X), \langle [pt];[pt],\alpha_1,...,\alpha_l\rangle^X_{\sigma,l+1}\neq 0\right\}.
\end{eqnarray*}
He also points out that $N_2$ is realized by classes with non trivial three point $GW$-invariant. Suppose $\sigma\in H_2(X;\mathbb{Z})$ realizes this minimum. Then every rational curve passing through $b$ and representing $\sigma$ is irreducible, i.e. any stable holomorphic map through $b$ is simple. It follows that $\mathcal{M}(X,J,\sigma;b)$ is compact. Furthermore, by choosing $b$ general enough, for any $u\in \mathcal{M}(X,J,\sigma;b)$ the obstruction $H^1(\mathbb{C} P^1,u^*TX)$ vanishes; thus, the moduli space $\mathcal{M}(X,J,\sigma;b)$ is a smooth oriented manifold, as desired.
\qed
\subsection{Hamiltonian fibrations and the product formula} In this section we introduce the notion of Hamiltonian fibrations which provides natural framework to study Gromov-Witten invariants.
\subsubsection{Hamiltonian fibrations.} By definition, a \emph{symplectic fibration} is a locally trivial smooth fibration with symplectic reference fiber $(F,\omega)$,
$$(F,\omega)\stackrel{\iota}{\hookrightarrow} P\stackrel{\pi}{\rightarrow} B,$$
and which structure group lies in the group of symplectic diffeomorphisms of the fiber, denoted $\mathrm{Symp}(F,\omega)$. It follows that each fiber $F_b:=\pi^{-1}(b)$ is naturally equipped with a symplectic form $\omega_b$. A symplectic fibration is \emph{Hamiltonian} if the structure group can be reduced to the group $\mathrm{Ham}(F,\omega)$ of Hamiltonian diffeomorphisms. Hamiltonian fibrations are characterized as follows:
\begin{theorem}(\cite{MS2}, Theorem 6.36.) A symplectic fibration $P$ as above is Hamiltonian if and only if the following two conditions are verified:
\begin{itemize}
\item[($H_1$)] $P$ is symplectically trivial over the $1$-skeleton of $B$;
\item[($H_2$)] there exists a unique closed connection 2-form $\tau\in \Omega^2(P)$ extending the family $\{\omega_b\}_{b\in B}$ such that the integration of $\tau^{n_F+1}$ over the fibers of $P$ vanishes.
\end{itemize}
\end{theorem}
The closed $2$-form in the theorem above is usually refered to as the \emph{coupling form}. The coupling form defines a connection on $P$, i.e. a splitting at each $p\in P$,
$$T_pP=Hor_{\tau,p}\oplus \ker d\pi(p).$$
The theorem states that the corresponding holonomy around any loop in $B$ is in $\mathrm{Ham}(F,\omega)$. Any other closed extension $\tau'$ of $\omega$ generating the same horizontal distribution $Hor_{\tau}$ is actually uniquely obtained from $\tau$ via the equation:
$$\tau'=\tau+\pi^*\varrho,\hspace{0.5cm}{0.2cm}\varrho\in\Omega^2(B).$$
One of the main features of Hamiltonian fibrations over closed symplectic bases is that they can be given symplectic structures compatibly with the family $\{\omega_b\}_{b\in B}$.
Such symplectic structures are obtained as follows:
$$\omega_{P,\kappa}:=\tau+\kappa \pi^*\omega_B,$$ where $\kappa>0$ is a large enough real number such that $\omega_{P,\kappa}$ is non-degenerate. Hence, this class of fibrations provide a nice framework to define Gromov-Witten invariants.
We will also use $\tau$ to denote the deRham cohomology class corresponding to $\tau$. There is another canonical cohomology class of $P$ that will play an important part in the next paragraphs, namely the vertical Chern class $c_v\in H^2(P;\mathbb{Z})$. This class is defined as the first Chern class of the vertical subbundle $\ker d\pi$.
As a specific example, let us mention the case of Hamiltonian fibrations over $S^2$ with fiber $(F,\omega)$, which will play an important part in this note. This class of examples is particularly important in symplectic topology due to the correspondence between these fibrations and the fundamental group of $Ham(F,\omega)$. More precisely, for $\gamma\in\pi_1(Ham(F,\omega)) $ one defines a Hamiltonian bundle $\pi:P_{\gamma}\rightarrow S^2$ with fiber $F$ via the clutching construction: choose any representative $\tilde{\gamma}:[0,1]\rightarrow Ham(F,\omega)$ for $\gamma$, then
$$P_{\gamma}:=\frac{(D^+\times F)\sqcup (D^-\times F)}{(e^{2\pi i\theta},x)\sim (e^{-2\pi i\theta},\widetilde{\gamma}(\theta).x),\hspace{0.5cm}{0.2cm}\text{on}\hspace{0.5cm}{0.2cm} S^1\times F}$$
where $D^{\pm}\subset \mathbb{C}$ denotes the closed unit disc. This construction is independent of the representative $\widetilde{\gamma}$; moreover, any Hamiltonian fibration over $S^2$ with $(F,\omega)$ as fiber can be constructed in this way (see \cite{Se} or \cite{LMP}). In this context, the coupling class will be denoted by $\tau_{\gamma}$ and $c_{\gamma}$ will denote the vertical Chern class associated to $P_{\gamma}$.
\subsubsection{The product formula.} Following \cite{H}, we give a product formula for $GW$-invariants of a Hamiltonian fibration, assuming the fiber $(F,\omega)$ is \emph{semi-positive relative to the total space}, i.e.
\begin{equation}\tag{$\star$}\label{ssp}\forall A\in H_2^S(F):\hspace{0.5cm}{0.2cm} \omega(A)>0,\hspace{0.5cm}{0.2cm} c^v(\iota(A))\geq 3-n_P\hspace{0.5cm}{0.3cm}\Longrightarrow c^v(\iota(A))\geq 0,
\end{equation}
where $H_2^S(F)$ is the spherical homology subgroup of $H_2(F,\mathbb{Z})$ (i.e. the image of $\pi_2(F)$ under the Hurewicz map) and $\iota$ denotes the map in homology induced from the natural embedding of the fiber. For instance, this implies that the fiber is semi-positive.
For this purpose we equip $P$ with an \emph{(almost) complex structure} $J_P$ which is compatible with the fibration structure and a Hamiltonian connection in the sense given below:
\begin{defn}\label{fiberedacs} An almost complex structure $J_P$ on $P$ is said to be \emph{compatible with $\pi$ and $\tau$}, or just \emph{fibered}, if and only if there exists an $\omega_B$-tame complex structure $J_B$ on $B$ and a family of $\omega_b$-tame almost complex structures $J_b$ in $F_b$ such that:
\begin{itemize}
\item $d\pi \circ J_P=J_B\circ d\pi$,
\item $J_b:=\left.J_P\right|_{F_b}$ for all $b\in B$,
\item $J_P$ preserves the horizontal distribution induced by $\tau$.
\end{itemize}
\end{defn}
For fibered $J_P$, the projection $\pi$ induces a map between moduli spaces:
$$ \overline{\pi}:{\mathcal{M}}_{0,l}(P,\sigma,J_P)\rightarrow{\mathcal{M}}_{0,l}(B,\sigma_B,J_B),\hspace{0.5cm}{0.2cm} [\Sigma,u,\mathbf{x}]\mapsto [\Sigma, \pi(u),\mathbf{x}]$$
where $\sigma_B:=\pi_*\sigma$. In what follows we assume that $\sigma_B$ is non-zero. The fiber of $\overline{\pi}$ over $[\Sigma, u_B,\mathbf{x}]$ can be described as follows.
Let $C$ denote the image of $u_B$ in $B$, and let $P_C$ denote the restriction of $P$ along $C$; $P_C$ defines a Hamiltonian fibration over $S^2$ with coupling form given by the pull-back of $\tau$ under the natural inclusion $\iota_{P_C}:P_C\hookrightarrow P$. If $J_C$ denotes the fibered almost complex structure on $P_C$ given by the restriction of $J_P$ to $P_C$ we have the following identification:
$$\overline{\pi}^{-1}([\Sigma,u_B,\mathbf{x}])\equiv\bigsqcup_{B_{\sigma}:=\{\sigma'\in H_2(P_C;\mathbb{Z})|\iota_{P_C}\sigma'=\sigma\}} \left(\mathcal{M}_{0,l}(P_C,J_C, \sigma')\cap \mathfrak{f}^{-1}([\Sigma,\mathbf{x}])\right).$$
Regarding evaluation at the marked points maps we have the commutative diagram:
\begin{equation}\label{evdiagram}
\xymatrix{\overline{\pi}^{-1}([\Sigma,u_B,\mathbf{x}]) \ar[r]\ar[d]^{ev_{(u_B,\mathbf{x})}}&{\mathcal{M}}_{0,l}(P,\sigma,J_P)\ar[d]^{ev^P_{l,J_P}}\ar[r]^{\overline{\pi}}& {\mathcal{M}}_{0,l}(B,\sigma_B,J_B) \ar[d]^{ev^B_{l,J_B}}\\
F^l\ar[r]^{(\iota)^l}&P^l\ar[r]^{\pi^l} & B^l}
\end{equation}
where
$$ev_{(u_B,\mathbf{x})}:\overline{\pi}^{-1}(u_B,\mathbf{x})\rightarrow F^l,\hspace{0.5cm}{0.5cm} u\mapsto (u(x_1),...,u(x_l))\in\prod_{i=1}^lF_{u_B(x_i)}.$$
The product formula is obtained by considering the (respective) intersections of $ev_{(u_B,\mathbf{x})}$, $ev^B_{l,J_B}$, and $ev^P_{l,J_P}$, with the product cycles:
\begin{equation*} \mathcal{C}^F:=\prod_{i=1}^{l}V^F_i,\hspace{0.5cm}{0.3cm}\mathcal{C}^B:=\prod_{i=1}^{l}V^B_i,\hspace{0.5cm}{0.3cm} \mathcal{C}^P:=\prod_{i=1}^{l}V^P_i,
\end{equation*}
where, $V^F_i$, $V^B_i$, and $V^P_i$, respectively represent homology classes, $c_i^F$, $c_i^B$, and $c_i^P$, verifying that for some integer $0\leq m\leq l$:
\begin{equation}\tag{$\star\star$}\label{cond11}
\begin{cases} c_i^B=pt, \hspace{0.5cm}{0.2cm} c_i^P=\iota(c_i^F) & \text{for $i=1,...,m$} \\
c_i^F=[F],\hspace{0.5cm}{0.2cm} c_i^P=\pi^!(c_i^B)& \text{for $i=m+1,...,l$.}
\end{cases}
\end{equation}
where $\pi^!$ stands for the shriek map:
\begin{equation*}\pi^{!}:H_*(B)\rightarrow H_{2n_F+*}(P), \hspace{0.5cm}{0.2cm} \alpha\mapsto PD_P^{-1}\pi^*PD_B(\alpha).
\end{equation*}
In other words, we consider cycles in $P$ that are either images of cycles in $F$ under $\iota$, or preimages under $\pi$ of cycles in $B$. Let $\iota_C$ denote the map in homology induced from the natural inclusion of $F$ in $P_C$. We have,
\begin{theorem}(\cite{H}, Theorem A)\label{productformula0}
Let $\pi: P\rightarrow B$ be a Hamiltonian fibration with relative semi-positive fiber $(F,\omega)$. Let $\sigma\in H_2(P,\mathbb{Z})$ and
suppose $\sigma_B:=\pi_*(\sigma)\neq 0$ only admits simple decompositions for some $\omega_B$-tame almost complex structure $J_B$ in $B$.
Let $c_i^P,c_i^B,c_i^F$ be as in \eqref{cond11}. For generic fibered almost complex structure lifting $J_B$, we have:
\begin{equation}\label{simpversionPF}\langle D;c^P_1,...,c_l^P\rangle_{\sigma,l}^P=\langle D;c^B_1,...,c_l^B\rangle_{\sigma_B,l}^B\cdot\sum_{\sigma'\in B_{\sigma}} \langle [pt]; \iota_C(c^F_1),...,\iota_C(c^F_l)\rangle_{\sigma',l}^{P_C}
\end{equation}
where $C$ is a curve counted in $\langle D; c_1^B,...,c_l^B\rangle^B_{\sigma_B,l}$ and $D\in H_*(\overline{\mathcal{M}}_{0,l})$.
\end{theorem}
This formula in particular states that the sum in the right handside of the formula is independent of the chosen $C$. Note that this sum is well-defined due to Gromov 's compactness.
The core of the proof consists in establishing that the $GW$-invariants involved are generically and simultanuously well-defined, in other words the problem is to realize transversality while preserving the map $\overline{\pi}$ defined above. The proof of this is based on the relation:
$$\pi_*\circ D^P=D^B\circ \pi_*,$$
where $D^P$ and $D^B$ respctively stand for the Fredholm operators obtained by linearizing the Cauchy Riemann operators $\ov{\partial}_{J_P}$ and $\ov{\partial}_{J_B}$. As a consequence, we derive an exact sequence
\begin{equation*}
0\rightarrow \ker D^{v} \rightarrow \ker D^P \rightarrow \ker D^B\rightarrow {\text{coker }} D^{v}\rightarrow {\text{coker }} D^P\rightarrow{\text{coker }} D^B\rightarrow 0,
\end{equation*}
where $D^v$ denotes the restriction of $D^P$ to vector fields along the curves that are vertically valued, i.e. with values in $\ker d\pi$. The vanishing, at least at the level of the universal moduli spaces, of the obstructions in the sequence above is provided by: 1) the irreducibility hypothesis on $\sigma_B$ for the vanishing of the last term of the sequence; 2) perturbing the Hamiltonian connection for the vanishing of ${\text{coker }} D^v$. It follows from standard arguments that for generic fibered almost complex structure the following holds:
\begin{itemize}
\item the subset ${\mathcal{M}}^{**}_{0,l}(P,\sigma,J_P)$ of ${\mathcal{M}}_{0,l}(P,\sigma,J_P)$ consisting of simple maps that project to simple maps under $\overline{\pi}$ and the moduli space ${\mathcal{M}}^*_{0,l}(B,\sigma_B,J_B)$
are oriented manifolds;
\item
for countably many $(u,\mathbf{x})\in {\mathcal{M}}^*_{0,l}(B,\sigma_B,J_B)$, the preimage $\overline{\pi}^{-1}(u,\mathbf{x})$ is an oriented manifold.
\end{itemize}
More generally, $\overline{\pi}$ extends to a map, using stabilization, between the compactifications $\overline{{\mathcal{M}}}_{0,l}(P,\sigma,J_P)$ and $\overline{{\mathcal{M}}}_{0,l}(B,\sigma_B,J_B)$. We can repeat the arguments above for each stratum in $\overline{{\mathcal{M}}}_{0,l}(P,\sigma,J_P)$ projecting to some stratum in $\overline{{\mathcal{M}}}_{0,l}(B,\sigma_B,J_B)$. The irreducibility hypothesis on the decompositions of $\sigma_B$ ensures that the ${\text{coker }} D^B$ term always vanishes and that the image under the evaluation map $ev^B_{l,J_B}$ of the lower strata in $\overline{{\mathcal{M}}}_{0,l}(B,\sigma_B,J_B)$ have codimension at least 2 with respect to the top stratum $\mathcal{M}^*(B,\sigma_B,J_B)$. Furthemore, condition \eqref{ssp} ensures that the image under the evaluation map $ev^P_{l,J_P}$ of the lower strata in $\overline{{\mathcal{M}}}_{0,l}(P,\sigma,J_P)$ have codimension at least two with respect to ${\mathcal{M}}^{**}_{0,l}(P,\sigma,J_P)$. Once transversality is achieved, one recovers the formula using diagram \eqref{evdiagram} and the following observations:
\begin{itemize}
\item since all cokernels vanish, it follows from the exact sequence above that:
$$\det (\ker D^P)\cong \det (\ker D^B) \otimes \det (\ker D^v).$$
Thus, the orientation of ${\mathcal{M}}^{**}_{0,l}(P,\sigma,J_P)$ is given by the product of the orientations of $\mathcal{M}^*_{0,l}(B,\sigma_B,J_B)$ and $\overline{\pi}^{-1}(u_B,\mathbf{x})$, where $(u_B,\mathbf{x})\in \mathcal{M}^*_{0,l}(B,\sigma_B,J_B)$. Moreover, the orientations of the product pseudo-cycles are also given by a product:
\begin{equation*}
\det T{\mathcal{C}}_P\cong \det T{\mathcal{C}}_B\otimes \det T{\mathcal{C}}_F.
\end{equation*}
\item It follows from symplectic triviality of $P$ over the $1$-squeleton of $B$, that for any two $J_B$-holomorphic maps representing the same class $\sigma_B$, the restrictions of $P$ to the images of the two maps are isomorphic as Hamiltonian fibrations. Hence, $\langle\phantom\cdots\rangle^{P_C}_{\sigma',l}$ does not depend on $C$.
\end{itemize}
The proof is concluded by showing, using standard arguments, the independence with respect to the generic fibered almost complex structure of the Gromov-Witten invariants involved in the formula. See \cite{H} for the details.
\subsection{Quantum Homology and Seidel elements} We start with the definition of (small) Quantum homology with universal Novikov ring. As a module the Quantum homology of $(X,\omega)$ is given by:
\begin{equation*} QH_*(X)\equiv QH_*(X,\Lambda):=H_*(X,\mathbb{Q})\otimes \Lambda
\end{equation*}
where $\Lambda$ denotes some coefficient ring that we specify. We will take
$\Lambda:=\Lambda^{univ}[q^{-1},q]$
where $\Lambda^{univ}$ is the ring of generalized Laurent series in variable $t^{-1}$, i.e. an element $\lambda\in\Lambda^{univ} $ can be written as a formal sum
\begin{equation*} \lambda=\sum_{i\geq 0} \lambda_it^{r_i},\hspace{0.5cm}{0.3cm} \lambda_i\in \mathbb{Q},r_i\in \mathbb{R},\hspace{0.5cm}{0.2cm} r_i>r_{i+1},\hspace{0.5cm}{0.2cm} \lim_{i\rightarrow\infty}r_i=-\infty.
\end{equation*}
The grading on $\Lambda$ is given by imposing that $\deg(q)=2$. Let $\Lambda_j$ denote the set of elements of degree $2j$ in $\Lambda$; then we give quantum homology the following grading:
\begin{equation*} QH_k(X):=\bigoplus_{i+2j=k} H_i(X,\mathbb{Q})\otimes \Lambda_j
\end{equation*}
Next, we introduce the (small) quantum product:
$$\star :QH_i(X)\otimes QH_j(X)\rightarrow QH_{i+j-2n}(X),\hspace{0.5cm}{0.2cm} (a,b)\mapsto a\star b.$$
First, let $\{e_{\nu}\}$ be a basis of homology of $H_*(X)$, and let $\{e^*_{\nu}\}$ denote the corresponding dual basis with respect to the intersection pairing.
Then, the quantum product is defined as follows: for $a\in H_i(X,\mathbb{Q})$ and $b\in H_j(X,\mathbb{Q})$, set
\begin{equation} a\star b=\sum_{B\in H^S_2(X),\nu}\langle a,b,e_{\nu}\rangle^X_{B,3} e^*_{\nu}\otimes q^{-c_1(B)}t^{-\omega(B)}
\end{equation}
We extend this by linearity with respect to $\Lambda$. Note that $[X]=1$ is the identity for this ring structure. Now, consider the $\Lambda$-submodule:
\begin{equation*} \mathcal{Q}_-:=\bigoplus_{i<2n} H_i(X)\otimes \Lambda.
\end{equation*}
The following lemma of McDuff (\cite{MDuniruled}, Lemma 2.1) will be essential when proving Theorem \ref{mainobservation}. We give its proof for the reader's convenience:
\begin{lem}\label{explicitSeidelelement} If $(X,\omega)$ is not strongly uniruled then $\mathcal{Q}_-$ is an ideal in $QH_*(X)$. Furthermore, if $\mathcal{Q}_-$ is an ideal $a\in QH_*(X)$ is a unit, then there exists $x\in\mathcal{Q}_-$ and $\lambda\neq 0$ in $\Lambda$ such that:
$$a=1\otimes\lambda+x.$$
\end{lem}
\noindent\emph{Proof:}\,\,\,\,\, Let $c\in QH_*(X)$ and assume by contradiction that there exists $b\in\mathcal{Q}_-$ such that $c\star b\notin \mathcal{Q}_- $. By definition of the quantum product this means that there exists $B\in H_2^S(X)$ such that
$$\langle c,b,pt\rangle^X_{B,3}\neq 0;$$
hence $(X,\omega)$ is strongly uniruled which is impossible. Next, if $a$ is a unit, then $a\notin \mathcal{Q}_-$, for if it was we would have that $1$ belongs to $\mathcal{Q}_-$. Therefore, we can write
$a=1\otimes\lambda+x$ with $x\in\mathcal{Q}_-$ and $\lambda\neq 0$.
\qed\\
\noindent\emph{Seidel's representation.} This is a representation of $\pi_1(Ham(X,\omega))$ in the subring $QH^{\times}_*(X,\Lambda)$ of units of the quantum homology of $(X,\omega)$. Recall that $\pi:P_{\gamma}\rightarrow S^2$ denotes the Hamiltonian fibration obtained from $\gamma\in\pi_1(Ham(X,\omega)) $ via the clutching construction. Also, let $\mathcal{H}_{\gamma}\subset H^S_2(P_{\gamma})$ denote the subset of section classes, i.e. of spherical classes that project to $[S^2]$ under the fibration projection.
\begin{defn} The Seidel representation map
$$S:\pi_1(Ham(X,\omega))\rightarrow QH^{\times}_*(X,\Lambda),\hspace{0.5cm}{0.2cm} \gamma\mapsto S(\gamma)$$
is defined as follows:
\begin{equation} S(\gamma):=\sum_{\sigma\in \mathcal{H}_{\gamma} ,\nu}\langle \iota(e_{\nu})\rangle^{P_{\gamma}}_{\sigma,1} e^*_{\nu}\otimes q^{-c_{\gamma}(\sigma)}t^{-\tau_{\gamma}(\sigma)}
\end{equation}
\end{defn}
Geometrically, $S(\gamma)$ "counts" holomorphic sections of $P_{\gamma}$ intersecting the cycles $e_{\mu}$ in the fiber above the north pole of $S^2$. The \emph{splitting axiom}, \cite{MS} Theorem 11.4.1, for fibrations over $S^2$ gives that for all section classes $\sigma \in H_2^S(P_{\gamma})$ and for classes $\alpha_1,...,\alpha_l\in H_*(F)$ the following holds for every integer $0\leq k\leq l$:
\begin{eqnarray*}
& &\langle[pt];\iota(\alpha_1),...,\iota(\alpha_l)\rangle^{P_{\gamma}}_{\sigma,l} =
\\&& \sum_{A\in H_2^S(F;\mathbb{Z}),\nu}
\langle[pt];\iota(\alpha_1),...,\iota(\alpha_{k}),\iota(e_{\nu})\rangle^{P_{\gamma}}_{\sigma-\iota(A),k+1}
\langle[pt];e^*_{\nu},\alpha_{k+1},...,\alpha_l\rangle^{F}_{A,l-k+1}
\end{eqnarray*}
It follows easily that the $\Lambda$ linear action of $S(\gamma)$ on $a\in H_*(F)$ is given by:
\begin{equation}\label{Seidelformula} S(\gamma)(a):=S(\gamma)\star a=\sum_{\sigma\in \mathcal{H}_{\gamma} ,\nu}\langle \iota(a), \iota(e_{\nu})\rangle^{P_{\gamma}}_{\sigma,2}e^*_{\nu}\otimes q^{-c_{\gamma}(\sigma)}t^{-\tau_{\gamma}(\sigma)}
\end{equation}
Note that in the above notations $S(\gamma)=S(\gamma)(1)$. Let us define the following equivalence class on section classes: we say that $\sigma_1$ is equivalent to $\sigma_2$ if and only if $$\tau_{\gamma}(\sigma_1-\sigma_2)=0=c_{\gamma}(\sigma_1-\sigma_2).$$
We will use the notation $[\sigma]$ to denote the equivalence class of $\sigma$. From Theorem 3.A in \cite{LMP}, we have
\begin{lem}\label{LemmeSeidelinverse} For every $\gamma\in \pi_1(Ham(F,\omega))$ and every non zero $a\in H_*(F)$, there exists an equivalence class of section classes $[\sigma]$ and $b\in H_*(F)$ such that:
$$\sum_{\sigma'\in [\sigma]}\langle\iota (a),\iota (b) \rangle^{P_{\gamma}}_{2,\sigma'}\neq 0.$$
In particular there is a section class $\sigma\in \mathcal{H}_{\gamma}$ such that $\langle\iota(a),\iota(b) \rangle^{P_{\gamma}}_{2,\sigma}\neq 0$.
\end{lem}
\noindent\emph{Proof:}\,\,\,\,\, Assume it is not true. Equation \eqref{Seidelformula} and linearity of Gromov-Witten invariants then imply that $S(\gamma)(a)=0$. But since $S(\gamma)$ is invertible, this is only possible for $a=0$; hence the contradiction. \qed \\
\section{proofs of the results}
We are now ready to prove the results. We begin by proving Theorem \ref{mainobservation} and Theorem \ref{generalization}. More precisely, we prove the following:
\begin{theorem}\label{unirulingfib2} Assume $(F,\omega)$ verifies \eqref{ssp} and is not strongly uniruled. If $(B,\omega_B)$ is symplectically uniruled for some class $\sigma_B$ admitting only simple decompositions, then $P$ is also symplectically uniruled. Moreover, if $[pt]_Q^{l}\neq0$, and if $(B,\omega_B)$ is $(l+1)$-SRC for $\sigma_B$ admitting only simple decompositions, then $P$ is at least $l$-SRC.
\end{theorem}
\noindent\emph{Proof:}\,\,\,\,\, Since $B$ is symplectically uniruled there exists $\sigma_B\in H_2^S(B)$ and classes $c_1^B,...,c^B_l\in H_*(B)$ such that: \begin{equation}\label{uniB} \langle D; [pt],c_1^B,...,c^B_l\rangle_{\sigma_B,l+1}^B\neq 0.
\end{equation}
Let $C$ be the image of a map counted in \eqref{uniB}. The restriction $P_C$ of $P$ to $C$ is a Hamiltonian fibration over $S^2$. Let $\phi\in \pi_1(Ham(F,\omega))$ be a Hamiltonian loop corresponding to this fibration, and let $S(\phi)\in QH^{\times}(F;\Lambda)$ denote the corresponding Seidel element. Since $(F,\omega)$ is not strongly uniruled, it follows from Lemma \ref{explicitSeidelelement} that there exists $0\neq \lambda\in \Lambda$ and $x\in \mathcal{Q}_- $ such that:
$$S(\phi)=1\otimes \lambda+x.$$
In particular, this directly implies that there is an equivalence class $[\sigma]$ of section classes in $P_C$ such that
$$0\neq \sum_{\sigma'\in[\sigma]} \langle[pt]\rangle^{P_C}_{\sigma',1}=\sum_{\sigma'\in[\sigma]}\langle[pt]; [pt],\iota_C([F]),...,\iota_C([F])\rangle^{P_C}_{\sigma',l+1}. $$
Let $\iota_{P_C}$ denote the inclusion in homology induced by the inclusion of $P_C$ into $P$. It is easy to see that the image under $\iota_{P_C}$ of $[\sigma]$ defines an equivalence class of spherical classes projecting on $\sigma_B$, in the sense that for any $\sigma_1,\sigma_2\in \iota_{P_C}([\sigma])$ we have $$\tau(\sigma_1-\sigma_2)=0=c_v(\sigma_1-\sigma_2).$$
By the product formula for Gromov-Witten invariants \eqref{simpversionPF}:
\begin{eqnarray*}&&\sum_{\sigma'\in \iota_{P_C}([\sigma])}\langle D; [pt],\pi^{!}(c_1^B),..., \pi^{!}(c_l^B)\rangle_{\sigma',l+1}^P\\
&&=\sum_{\sigma'\in \iota_{P_C}([\sigma])}\langle D; [pt], c^B_1,...,c_l^B\rangle_{\sigma_B,l+1}^B\sum_{\sigma''\in B_{ \sigma'}}\langle[pt]; [pt],\iota_C([F]),...,\iota_C([F])\rangle^{P_C}_{\sigma'',l+1}.\\
&&=\langle D; [pt], c^B_1,...,c_l^B\rangle_{\sigma_B,l+1}^B \sum_{\sigma'\in[\sigma]}\langle[pt]; [pt],\iota_C([F]),...,\iota_C([F])\rangle^{P_C}_{\sigma',l+1}\neq 0\\
\end{eqnarray*}
In particular, there is at least one $\sigma'\in \iota_{P_C}([\sigma])$ such that $$\langle D; [pt],\pi^{!}(c_1^B),..., \pi^{!}(c_l^B)\rangle_{\sigma',l+1}^P\neq 0.$$
Now we prove the second assertion of the theorem. Since for every loop $\phi$ of Hamiltonian diffeomorphisms of $F$ the Seidel element $S(\phi)$ is an unit in $QH_*(F;\Lambda)$, we have that
$$0\neq S(\phi)\star ([pt]_Q^l).$$
Using the splitting axiom for Gromov-Witten inviariants, and from the definition of the Seidel element, one obtains that
\begin{eqnarray*}
S(\phi)\star ([pt]_Q^l)&=&\sum_{ A\in H_2^S(F),
\nu}\langle [pt]; [pt],...,[pt], e_{\nu}\rangle^{F}_{A,l+1}S(\phi)(e^*_{\nu})\otimes q^{-c^{TF}_1(A)}t^{-\omega(A)}\\
&=& \sum_{\sigma\in \mathcal{H}_{\phi},
A\in H_2^S(F),
\nu,\mu}\langle [pt]; [pt],...,[pt], e_{\nu}\rangle^{F}_{A,l+1} \langle \iota(e^*_{\nu}),\iota(e_{\mu})\rangle^{P_{\phi}}_{\sigma,2} \\
&&\hspace{0.5cm}{3.5cm}e^*_{\mu}\otimes q^{-c_{\phi}(\sigma+\iota(A))}t^{-\tau_{\phi}(\sigma+\iota(A))} \\
& =&\sum_{
\widetilde{\sigma}\in \mathcal{H}_{\phi},
\mu}\langle [pt]; [pt],...,[pt], \iota(e_{\mu})\rangle^{P_{\phi}}_{\widetilde{\sigma},l+1}e^*_{\mu}\otimes q^{-c_{\phi}(\widetilde{\sigma})}t^{-\tau_{\phi}(\widetilde{\sigma})}
\end{eqnarray*}
Hence, there is a class $a\in H_*(F)$ and a class of section classes $[\sigma]$ such that
$$\sum_{\sigma'\in[\sigma]}\langle[pt]; [pt],...,[pt],\iota(a)\rangle^{P_{\phi}}_{\sigma',l+1}\neq 0.$$
We conclude by the use of the product formula as before.
\qed\\
Now Theorem \ref{mainobservation} follows easily:\\
\noindent\textbf{Proof of Theorem \ref{mainobservation}:} In case where $(F,\omega)$ is strongly uniruled, the conclusion follows from Theorem \ref{SympDiv}. Otherwise, we simply use Theorem \ref{unirulingfib2} to conclude.
\qed\\
As for Theorem \ref{mainobservation}, Corollary \ref{cormainobs} follows directly from Theorem \ref{SympDiv} and the proposition below. This result is a simple consequence of Theorem \ref{unirulingproj} asserting that projectively uniruled manifolds are strongly symplectically uniruled.
\begin{prop}Assume $(F,\omega)$ verifies \eqref{ssp} and is not strongly uniruled. Also assume that $(B,J_B,\omega_B)$ is a uniruled projective manifold. Then $P$ is SSU.
\end{prop}
\noindent\emph{Proof:}\,\,\,\,\, From lemma \ref{projuni}, $(B,J_B,\omega_B)$ is uniruled for a class $\sigma_B$ verifying the assumptions of Theorem \ref{unirulingfib2}, and which is obstruction free. Hence, the product formula can be applied and the proof follows.
\qed\\
Now, we proceed to the proof of Corollary \ref{WeinFib2}:\\
\noindent\textbf{Proof of Corollary \ref{WeinFib2}:} By definition, $B$ is 2-SRC. We will show that $(F,\omega)$ homologically injects in $P$. Assume this is not true, then there exists a non-zero element $\overline{a}\in H_*(F,\mathbb{Z})\cap \ker\iota$. It follows by linearity that any Gromov-Witten invariant with entry $\iota(\overline{a})$ must vanish. Now, let $C$ be the image of a map counted in $$\langle D;[pt],[pt],c_3^B,...,c^B_l\rangle_{\sigma_B,l}^B\neq 0.$$
Since $P_C$ is a Hamiltonian fibration over $S^2$, we can apply Lemma \ref{LemmeSeidelinverse}, taking $a=[pt]$, and find that
there is an equivalence class $[\sigma]$ of section classes in $P_C$ and an element $b\in H_*(F)$ such that:
$$0\neq \sum_{\sigma'\in[\sigma]}\langle \iota_C(a),\iota_C(b)\rangle_{\sigma',2}^{P_C}=\sum_{\sigma'\in [\sigma]}\langle [pt]; \iota_C(a),\iota_C(b),\iota_C([F]),...,\iota_C([F])\rangle_{\sigma',l}^{P_C},$$
where the last equality follows from the Divisor axiom for Gromov-Witten invariants.
Applying the product formula \eqref{simpversionPF} as in the proof of Theorem \ref{unirulingfib2} we conclude that for every $a\in H_*(F)$ there is a non-vanishing Gromov-Witten invariant in $P$, namely:
\begin{equation}\label{preuveunireglage}
\langle D; \iota(a),\iota(b),\pi^{-1}(c_3^B),...,\pi^{-1}(c^B_l)\rangle_{\iota_{P_C}(\sigma),l}^P\neq 0.
\end{equation}
Taking $a=\overline{a}$ gives a contradiction, hence $\ker\iota$ is trivial. Now, the Corollary follows directly from Corollary \ref{WeinFib}.
\qed\\
\bibliographystyle{plain}
|
\section{Introduction}
There are many situations in which it is necessary to consider
topological relations among one-dimensional objects that are homeomorphic
to rings. The most significant examples are provided by long flexible
polymers and biopolymers, whose trajectories may close themselves
and form what in the polymer scientific literature are called
{\it catenanes} \cite{wasser}--\cite{ff}. The latter are able to
entangle themeselves giving rise to
complex links involving two
or more interlocked chains. Additionally, each catenanae may be in the
configuration of a nontrivial knot.
Two cases of polymer links
are shown in Fig.~\ref{loop}.
\begin{figure}[ht]
\includegraphics[scale=0.2]{2loops.eps}
\caption{Entangled polymers rings $P_1$ and $P_2$ with linked trajectories
$C_1$ and $C_2$. In
a) polymer $P_2$ is in a nontrivial knot configuration, while in b)
both trajectories are unknots. } \label{loop}
\end{figure}
Besides polymers, other examples in which topological relations
among a system of one-dimensional objects
become relevant
can be found in
condensed matter physics (paths around defects in melted crystals)
\cite{chaikin,pieranski} or in particle physics (loops in quantum
gravity and the so-called hopfions) \cite{ash,smolin,hopfions}.
In order to specify the topological states of a given system of this
kind one uses {\it knots} or {\it link invariants}. In the
following, we will be interested in the topological relations of a
system of a linked rings without taking into account the fact that
these rings could be also in a nontrivial knot configuration as for
example in Fig.~\ref{loop}~a). For this reason, we will discuss here
only link invariants.
It is well known that the correlation functions
of the observables of a
topological field theory are topological invariants. Moreover,
the coefficients of the perturbative expansion of those
correlation functions are topological invariants too.
In practice, this means that to a finite set of Feynman diagrams
it is possible to associate a given topological invariant.
Our purpose is to solve the inverse problem.
This means that, starting from a given topological
invariant, we would like to obtain a topological
field theory with a finite set of Feynman diagrams and
a correlation function which is
a function of that invariant.
This is the program of topological engineering
that has been stated in Ref.~\cite{ffbookch}.
In the last few decades topological theories with the above characteristics
have been
extensively applied in the statistical mechanics of polymers, see for
instance
\cite{edwards}--\cite{ferrariTFT} and \cite{ffbookch,leal}.
The most popular approach used
in order to distinguish the different topological
configurations of the one-dimensional objects
is based on the {\it Gauss linking number}
(GLN).
The corresponding topological field theory is an abelian BF model
discussed in Ref.~\cite{blau}.
The goal of this work is to extend this approach based
on the GLN to the case of polymer dynamics,
in which the shape of the linked trajectories is not static, but
changes in time.
\section{The topological engineering program}
The program of topological engineering in the case of links
may be summarized
as follows:\\
{\it Let $\cal{T}(\ell)$ be a link invariant, which describes the
topological properties of a $N$--component link $\ell$. It is
required that:
\begin{itemize}
\item[a)] the invariant $\cal{T}(\ell)$ is explicitly written
as a functional of trajectories $C_1,\ldots,C_N$ of knots
composing the link.\\
Given a link invariant of this kind, find a topological
field theory with observables ${\cal{O}}_1,\ldots,{\cal{O}}_n$
such that $\cal{T}(\ell)$, or equivalently a function
$F[\cal{T}(\ell)]$ of it, can be expressed as the correlator
of these observables
\beq{
F({\cal{T}}) = \int\!{\cal{D}}\lbrace \phi \rbrace e^{-S(\lbrace \phi \rbrace )}
{\cal{O}}_1(\lbrace \phi\rbrace),\ldots,{\cal{O}}_n(\lbrace \phi\rbrace)
}{corr}
where $S(\lbrace\phi \rbrace)$ is the action of a system and $\lbrace \phi\rbrace $
is a set of fields that can be scalars, vectors or higher order tensors.
\end{itemize}
The topological field theory and its observables should satisfy the following
conditions:
\begin{itemize}
\item[b)] Each observable ${\cal{O}}_i$, $i=1,\ldots,n$,
must depend on the trajectory of only one knot
\item[c)] No further regularization should be necessary in order
to compute the correlator $\langle{\cal O}_1,\ldots,{\cal O}_n\rangle$,
apart from the usual regularization schemes required by the possible
presence of ultraviolet divergences.
\end{itemize} }
An example of topological engineering is based on the GLN and the
abelian BF field theory. The GLN is given by:
\beq{
\chi(C_1,C_2) = \frac 1{4\pi}\epsilon_{\mu\nu\rho}
\oint_{C_1}\!dx_1^{\mu}(s_1) \oint_{C_2}\!dx_2^{\nu}(s_2)
\frac {(x_1(s_1) - x_2(s_2) )^{\rho} }{|x_1(s_1) - x_2(s_2)|^3}
}{gln}
where $x_1(s_1)^{\mu}$ and $x_2(s_2)^{\nu}$ are spatial curves in
three dimensions
that represent respectively the closed trajectories $C_1$ and $C_2$ of
two polymers $P_1$ and $P_2$.
The Greek indexes $\mu,\nu,\rho=1,2,3$ denote the spatial components.
Here $s_1$ and $s_2$ represent the arc-lengths on the curves $C_1$ and $C_2$.
$s_1$ and $s_2$ are defined in a such a way that
$0\leq s_1\leq L$ and $0\leq s_2\leq L$.
To find a field theory which is associated to the invariant
$\chi(C_1,C_2)$, we rewrite \ceq{gln}
as follows
\beq{
\chi(C_1,C_2) = \int\!d^3x \int\!d^3y
\xi_1^{\mu}(x)G_{\mu\nu}(x-y)\xi_2^{\nu}(y)
}{gln_alt}
where
\beq{
\xi_1^{\mu}(x) = \oint_{C_1}\!dx_1^{\mu}\delta(x-x_1) \qquad
\xi_2^{\nu}(x) = \kappa\oint_{C_2}\!dx_2^{\nu}\delta(x-x_2)}
{currents}
are called the bond vectors densities and
\beq{G_{\mu\nu}(x-y) = \frac 1{2\pi\kappa}\epsilon_{\mu\nu\rho}
\frac{(x-y)^{\rho}}{|x-y|^3}
}{propagator}
Let us note that
$G_{\mu\nu}(x-y)$ coincides with the propagator of the abelian BF
model discussed in Ref.~\cite{blau}. To make the connection with the
BF model even more explicit,
we have introduced a new parameter $\kappa$, which
will play later the role of the coupling constant of that model.
Clearly, the addition of this parameter is irrelevant. As a matter of
fact, the right hand side of Eq.~\ceq{gln_alt} does not depend
on $\kappa$.
Now the quantity
\beq{
e^{i\chi(C_1,C_2)} = e^{i\int\!d^3x \int\!d^3y \xi_1^{\mu}(x)G_{\mu\nu}(x-y)\xi_2^{\nu}(y)}
}{genfun}
can be regarded as the generating functional of a Gaussian field
theory with
propagator $G_{\mu\nu}(x-y)$ for the very special choice of currents
\ceq{currents}.
It is easy to recognize that the underlaying
field theory is an Abelian BF model
with action
\beq{
S_{\mbox{\tiny BF}} = i\kappa \epsilon^{\mu\nu\rho} \int\!d^3x A_{\mu}\partial_{\nu}B_{\rho}
}{BFaction}
It is possible to show that the abelian version of the BF model is
actually equivalent to two
Abelian C-S field theories.
If we quantize the above topological field theory using the Lorentz
gauge fixing,
in which both fields
$A_{\mu}$ and $B_{\mu}$ are completely transverse,
we obtain the
following relation
\beq{
e^{i\chi(C_1,C_2)} =
\int\!{\cal D}A_{\mu}{\cal D}B_{\mu} e^{-S_{\mbox{\tiny BF}}}
e^{i\int\!d^3x \xi_1^{\mu} A_{\mu}}e^{i\kappa\int\! d^3x \xi_2^{\mu}B_{\mu}}
\delta(\partial^{\mu}A_{\mu})\delta(\partial^{\mu}B_{\mu})
}{genfunCS}
The above equation is the analog of Eq.~\ceq{corr} in the present case. There are
just two observables ${\cal O}_1$ and ${\cal O}_2$, namely the
two Abelian Wilson loops given below:
\beq{
{\cal O}_1 = e^{i\int\!d^3x \xi_1^{\mu} A_{\mu} } \qquad
{\cal O}_2 = e^{i\kappa\int\!d^3x \xi_2^{\mu} B_{\mu} }
}{wilson}
\section{The case of dynamics}
In this Section we would like to extend the program of topological
engineering
to the case of two trajectories
whose configurations are changing during time. This problem is very
important
to study the dynamics of two entangled polymers.
Once again, we choose the Gauss linking invariant in order to impose
topological conditions on two closed trajectories $C_1$ and $C_2$.
The only difference from the previous static example is that now the
curves $x_1$ and $x_2$ depend on time, i.e. $x_1=x_1(t,s_1)$
and $x_2=x_2(t,s_2)$.
The GLN can still be defined, but will be a time dependent quantities:
\beq{
\chi(t,C_1,C_2) = \frac 1{4\pi} \epsilon_{\mu\nu\rho}
\oint_{C_1}\!dx_1^{\mu}(t,s_1) \oint_{C_2}\!dx_2^{\nu}(t,s_2)
\frac {(x_1(t,s_1) - x_2(t,s_2) )^{\rho} }{|x_1(t,s_1) - x_2(t,s_2)|^3}
}{chitime}
Of course, if the trajectories would be impenetrable, then $\chi$
would be a
constant, since it is not possible to change the topological
configuration of a system
of knots if their trajectories are not allowed to cross
themselves. However, in
the absence of excluded volume interactions
models of polymer physics are phantom, i.e. crossings are allowed.
For this reason, we will require that only the
time average of the GLN
is fixed. As a consequence, we will consider a time averaged version
of the GLN on the
time interval
$[0,t_f]$:
\beq{
\langle \chi (t,C_1,C_2) \rangle =
\int_0^{t_f} \frac {dt}{t_f} \chi (t,C_1,C_2)
}{timeaGLN}
Next, we generalize Eq.~\ceq{genfun} to the case of dynamics. To this
purpose, we introduce the following field theory
\beq{
S = \frac 1{t_f} \epsilon_{\mu\nu\rho} \int\!d\eta d^3x
A^{\mu}(\eta,x)\partial_x^{\nu}B^{\rho}(\eta,x)
}{actiondyn}
The above action differs from that of Eq.~\ceq{BFaction} by the addition
of the fourth dimension represented by variable $\eta$, with
$-\infty<\eta<+\infty$. Note that $S$ is not invariant under
diffeomorphism on the whole dimensional space spanned by the
coordinates $x^1,x^2,x^3$ and $\eta$, but only on its
three dimensional spatial section.
As a consequence, strictly speaking $S$ does not describe a
topological field theory.
The propagator corresponding to the action \ceq{actiondyn}
in the Lorentz gauge is given by
\beq{
G_{\mu\nu}(\eta,\eta';x,x') = \frac {t_f}{2\pi} \epsilon_{\mu\nu\rho}
\frac{(x-x')^{\rho}}{|x-x'|^3} \delta(\eta - \eta')
}{propdyn}
The analog of Eq.~\ceq{genfun} is
\beq{
e^{-i\lambda\chi(C_1,C_2)} = \int\!{\cal{D}} A_{\mu}{\cal{D}}B_{\nu} e^{-iS}
e^{-i\int\! d\eta d^{3}x ( J_1^{\mu}(\eta,x)A_{\mu}(\eta,x) + J_2^{\mu}(\eta,x)B_{\mu}(\eta,x) ) }
}
{genfungen}
where
\beq{J_1^{\mu}(\eta,x) = \frac{1}{2t_f} \int_0^{t_f}\!\frac{dt}{t_f}\delta(\eta - t)
\int_0^{L_1}\!ds_1 \frac{\partial }{\partial s_1}x_1^{\mu}(t_1,s_1)
\delta^{(3)}(x-x_1(t,s_1))
}{prad1}
and
\beq{
J_2^{\mu}(\eta,x) = \lambda \int_0^{t_f}\!\frac{dt}{t_f}\delta(\eta - t)
\int_0^{L_2}\!ds_2 \frac{\partial }{\partial s_2}x_2^{\mu}(t_1,s_2)
\delta^{(3)}(x-x_2(t,s_2))
}{prad2}
The right hand side of Eq.~\ceq{genfungen} can be seen as the amplitude of the two observables
\beq{
{\cal O}_1 = e^{-i\int\! d\eta d^3x J_1^{\mu}(\eta,x)A_{\mu}(\eta,x)} \qquad
{\cal O}_2 = e^{-i\int\! d\eta d^3x J_2^{\mu}(\eta,x)B_{\mu}(\eta,x)}
}{canbeseen}
To prove Eq.~\ceq{genfungen} it is sufficient to perform the Gaussian
integration in the fields $A^{\mu}$ and $B^{\mu}$. The result of that
operation is
\beq{
e^{-i\int\! d\eta d^{3}x ( J_1^{\mu}(\eta,x)A_{\mu}(\eta,x) + J_2^{\mu}(\eta,x)B_{\mu}(\eta,x) ) }=
e^{-i\int\!d\eta d^3x\int\!d\eta' d^3x' J_1^{\mu}(\eta,x) G_{\mu\nu}(\eta,\eta';x,x') J_2^{\nu}(\eta',x') }
}{gaussintegration}
Using the explicit expression of the propagator $G_{\mu\nu}(\eta,\eta';x,x')$
given in Eq.~\ceq{propdyn} it is possible to verify Eq.~\ceq{genfungen}
after eliminating the spurious variables $\eta,\eta'$ and $x,x'$:
\begin{eqnarray}
&&e^{-i\int\!d\eta d^3x\int\!d\eta' d^3x' J_1^{\mu}(\eta,x) G_{\mu\nu}(\eta,\eta';x,x') J_2^{\nu}(\eta',x') } =\\
\nonumber
&&\!\!\!\!\!\!\exp{\left[-\frac {i\lambda}{4\pi}\int_0^{t_f}\!\frac{dt}{t_f} \int_0^{L_1}\!ds_1\int_0^{L_2}\!ds_2
\epsilon_{\mu\nu\rho}
\frac{\partial}{\partial s_1}x_1^{\mu}(t,s_1) \frac{\partial}{\partial s_2}x_2^{\nu}(t,s_2)
\frac{(x_1(t,s_1) - x_2(t,s_2))^{\rho}}{|(x_1(t,s_1) - x_2(t,s_2)|^3}
\right]}
\label{endofstory}
\end{eqnarray}
The right hand side of above equation coincides with
$e^{-i\lambda\chi(C_1,C_2)}$. This
completes our proof.
\section{Concluding remarks}
In this work the program of topological engineering has been extended to the
case of the dynamics of two polymer chains. In particular, the Gauss linking
invariant has been considered. It has been shown that a time average version
of this topological invariant can be reproduced from an amplitude of
a field theory in the form of Eq.~\ceq{corr}. This amplitude is given in
Eq.~\ceq{genfungen}. Due to the fact that the conformations of the chains
change during time, the underlying field theory is four dimensional
and it is topological only with respect to diffeomorphisms of
the spatial section of four dimensional space.
\section{Acknowledgments}
One of us -- M. Pi\c{a}tek -- would like to thank the University of
Szczecin and the Faculty of Mathematics and Physics of that
University for the kind hospitality.
|
\section{Introduction}
\setcounter{equation}{0}
\renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
Onsager calculated the free energy of the two-dimensional square-lattice
Ising model in 1944.{\bl \cite{Onsager1944}} He did this by showing that the
transfer matrix is a product of two matrices, which together
generate (by successive commutations) a finite-dimensional Lie algebra
(the ``Onsager algebra''). In 1949 Bruria Kaufman
gave a simpler
derivation{\bl \cite{Kaufman1949}} of this result,
using anti-commuting spinor (free-fermion) operators, i.e. a
Clifford algebra.
Onsager was the Josiah Willard Gibbs Professor of
Theoretical Chemistry at Yale University.
Kaufman had recently completed her PhD at Columbia University
in New York, and was a research associate at the Institute
for Advanced Study in Princeton.
Later that year, Kaufman and
Onsager{\bl \cite{KaufmanOnsager1949}}
went on to calculate
some of the two-spin correlations. Let $i$ label the columns of
the square lattice (oriented in the usual manner, with axes
horizontal and vertical), and $j$ label the rows, as in Figure
\ref{sqlatt}. Let the spin at site $(i,j)$ be $\sigma_{i,j}$, with
values $+1$ and $-1$. Then the
total energy is
\begin{displaymath} E \; = \; -J \sum_{ij} \sigma_{i,j} \sigma_{i,j+1} -J'
\sum_{ij} \sigma_{i,j} \sigma_{i+1,j} \end{displaymath}
and the partition function is
\begin{displaymath} Z \; = \; \sum_{\sigma} {\rm e}^{-E/{\kappa T} } \;\; , \end{displaymath}
the sum being over all values of all the spins,
$\kappa$ being Boltzmann's constant and $T$ the temperature.
Onsager defines $H, H', H^*$ by
\begin{equation} \label{defH*}
H = J/{\kappa T} \;\; , \;\; H' = J'/{\kappa T} \;\; , \;\;
{\rm e}^{-2H^*} = \tanh H \;\; . \end{equation}
The specific heat diverges logarithmically at a critical
temperature $T_c$, where
\begin{displaymath} \sinh (2J/\kappa T_c ) \sinh (2J'/\kappa T_c ) = 1
\;\; . \end{displaymath}
\begin{figure}[hbt]
\begin{picture}(320,240) (-14,-40)
\setlength{\unitlength}{1.0pt}
{\color{black}
\put (60,10) {\line(1,0) {250}}
\put (60,11) {\line(1,0) {250}}
\put (60,60) {\line(1,0) {250}}
\put (60,61) {\line(1,0) {250}}
\put (60,110) {\line(1,0) {250}}
\put (60,111) {\line(1,0) {250}}
\put (60,160) {\line(1,0) {250}}
\put (60,161) {\line(1,0) {250}}
\put (85,-15) {\line(0,1) {200}}
\put (86,-15) {\line(0,1) {200}}
\put (135,-15) {\line(0,1) {200}}
\put (136,-15) {\line(0,1) {200}}
\put (185,-15) {\line(0,1) {200}}
\put (186,-15) {\line(0,1) {200}}
\put (235,-15) {\line(0,1) {200}}
\put (236,-15) {\line(0,1) {200}}
\put (285,-15) {\line(0,1) {200}}
\put (286,-15) {\line(0,1) {200}}
\multiput(85,10)(50,0){5}{\circle*{10}}
\multiput(85,60)(50,0){5}{\circle*{10}}
\multiput(85,110)(50,0){5}{\circle*{10}}
\multiput(85,160)(50,0){5}{\circle*{10}}
\put (85,-35){\large{1}}
\put (135,-35){\large{2}}
\put (235,-35){\Large{$i$}}
\put (320,5){\large{1}}
\put (320,55){\large{2}}
\put (320,105){\Large{$j$}}
\put (52,43){\large{(1,2)}}
\put (204,93){\large{$(i,j)$}}
}
\end{picture}
\caption{\small The square lattice.}
\label{sqlatt}
\vspace{1cm}
\end{figure}
The correlation between the two spins at sites
$(1,1)$ and $(i,j)$ is
\begin{displaymath} \langle \sigma_{1,1} \sigma_{i,j} \rangle \; = \;
Z^{-1} \sum_{\sigma} \sigma_{1,1}
\sigma_{i,j} {\rm e}^{-E/{\kappa T }} \;\; , \end{displaymath}
For the isotropic case $H' = H$,
Kaufman and Onsager{\bl \cite{KaufmanOnsager1949}}
give in their equation 43 the formula
for the correlation
between two spins in the same row:{\color{blue} \footnote{ We have
negated their $\Sigma_r$,
which makes them the same as those in Appendix A, and
corrected what appears to be a sign error. It is now the same as
III.43 of the draft paper below.}}
\begin{equation} \label{III.43}
\langle \sigma_{1,1} \sigma_{1,1+j} \rangle
\; = \; \cosh^2 H^{*} \, \Delta_j - \sinh^2 H^* \, \Delta_{-j} \end{equation}
Here $\Delta_j$ and $\Delta_{-j}$ are Toeplitz
determinants:
\begin{displaymath} \Delta_{j} \; = \; \raisebox{-0mm}{$
\left| \begin{array}{lllll}
\Sigma_1 & \Sigma_2 & \Sigma_3 & \cdots & \; \Sigma_{j} \\
\Sigma_{0} & \Sigma_1 & \Sigma_2 & \cdots & \Sigma_{j-1} \\
\cdot & \cdot & \cdot & \cdots & \; \; \cdot \\
\Sigma_{2-j} & \cdot & \cdot & \cdots & \; \Sigma_1\end{array}
\right| $} \end{displaymath}
\begin{equation} \label{Dmk}
\Delta_{-j} \; = \; \raisebox{-0mm}{$
\left| \begin{array}{lllll}
\Sigma_{-1} & \Sigma_{-2} & \Sigma_{-3} & \cdots & \; \Sigma_{-j} \\
\Sigma_{0} & \Sigma_{-1} & \Sigma_{-2} & \cdots & \Sigma_{1-j} \\
\cdot & \cdot & \cdot & \cdots & \; \; \cdot \\
\Sigma_{j-2} & \cdot & \cdot & \cdots & \; \Sigma_{-1}\end{array}
\right| $} \end{equation}
where
\begin{displaymath} \Sigma_r \; = \; \frac{1}{2 \pi} \int_0^{2 \pi}
{\rm e}^{{\rm i } r \omega +{\rm i } \delta' (\omega)} \, {\rm {d}} \omega \end{displaymath}
and
\begin{equation} \label{tandelta}
\tan \delta '(\omega) \; = \; \frac{\sinh 2 H \sin \omega }
{\coth 2 H' -\cosh 2 H \cos \omega } \;\; . \end{equation}
Setting
\begin{equation} \label{deldelp}
\delta ( \omega) = \delta'(\omega) + \omega \;\; , \end{equation}
this implies
\begin{equation} \label{eidelta}
{\rm e}^{{\rm i } \delta(\omega) } \; = \; \left\{
\frac{(1-\coth H' \, {\rm e}^{-2 H + {\rm i } \omega})(1-\tanh H' \, {\rm e}^{-2 H + {\rm i } \omega})}
{(1-\coth H' \, {\rm e}^{-2 H - {\rm i } \omega})(1-\tanh H' \, {\rm e}^{-2 H - {\rm i } \omega})}
\right\}^{1/2} \;\; . \end{equation}
(Equations (\ref{tandelta}), (\ref{eidelta}) follow from (89)
of {\bl \cite{Onsager1944}} and are true for the general case when $H'$, $H$
are not necessarily equal.)
Kaufman and Onsager also give the
formula for the correlation between spins in adjacent rows.
In particular, from their equations 17 and 20, we obtain
\begin{equation} \label{1stform}
\langle \sigma_{1,1} \sigma_{2,2} \rangle \; = \;
\cosh ^2 H^* \, \Sigma_1 + \sinh ^2 H^* \, \Sigma_{-1}
\;\; . \end{equation}
The long-range order, or spontaneous magnetization, can be defined
as
\begin{equation} \label{defM0}
M_0 \; = \; \left( \lim_{j \rightarrow \infty} \langle
\sigma_{1,1} \sigma_{1,1+j} \rangle \right) ^{1/2} \;\; . \end{equation}
It is expected to be zero for $T$ above the critical
temperature, and positive below it, as in Figure
\ref{graphM}.
Kaufman and Onsager were obviously very close to
calculating $M_0$: all they needed to do was to
evaluate $\Delta_j$, $\Delta_{-j}$ in the limit $j \rightarrow
\infty$. Here I shall present the evidence that they devoted
their attention to doing so, and indeed succeeded. They used two
methods: the first is discussed in section 2 and Appendix A,
the second in sections 3, 4, 5. Many years ago (probably
in the early or mid-1990's), John Stephenson (then in Edmonton,
Canada) sent the author
a photocopy of an eight-page typescript, bearing the
names Onsager and Kaufman, that deals with the topic.
Stephenson had copied it about 1965 in Adelaide, Australia,
from a copy owned by Ren Potts: both Potts and
Stephenson were students of
Cyril Domb.{\bl \cite{Domb1974, Domb1990} } It's possible that
Potts's copy had come from Domb when they were together in Oxford
from 1949 to 1951, but more likely that he had been given it by
Elliott Montroll, with whom Potts had collaborated in the early 1960's
(see ref. {\bl \cite{MPW1963}}). A transcript of the author's
copy is given in section 3, and a scanned copy
forms Appendix B.
Both these methods start from the formulae for the pair correlation
of two spins deep within an infinite lattice. There is also a third method
used by the author for the superintegrable chiral Potts model
(which is an $N$--state generalization of the Ising
model){\bl \cite{baxter2008}}:
if one calculates the single-spin expectation
value $\langle \sigma_{1,M} \rangle$ in a lattice of width $L$ and
height $2 M$ with cylindrical
boundary conditions and fixed-spin boundary conditions
on the top and bottom rows, then one can write the result as a determinant
of dimension proportional to $L$. (This method is similar to that of
Yang.{\bl \cite{Yang1952}}) The determinant is {\it not}
Toeplitz, but in the limit $M \rightarrow \infty$ it is a product of Cauchy
determinants, so can be evaluated directly for finite
$L$.{\bl \cite{baxter2010a,baxter2010b}}
\section{The first method}
\setcounter{equation}{0}
In August 1948, Onsager silenced a conference at Cornell
by writing on the blackboard an exact formula for
$M_0.${\bl \cite[p.457] {LongHigginsFisher}}
The following year, in May 1949 at a conference
of the International Union of Physics on statistical mechanics
in Florence, Italy,{\bl \cite[p. 261]{Onsager1949}}
Onsager referred to
the magnetization of the Ising model and announced that
``B. Kaufman and I have recently solved'' this problem.
He gave the result as
\begin{equation} \label{Isingmag}
M_0 \; = \; (1-k^2)^{1/8} \end{equation}
where
\begin{equation} k = 1/(\sinh 2 H \, \sinh 2 H' ) \end{equation}
and the result is true for $0 <k <1$, when $T < T_c$.
For $k >1$ the magnetization
vanishes, i.e. ${\cal M} = 0 $.
Figure \ref{graphM} shows the resulting graph of
$M_0$ for the isotropic case $H' = H$.
\begin{figure}
\begin{picture}(300,300) (5,-57)
\setlength{\unitlength}{0.9pt}
\thicklines
{\color{black}
\put (110,-35) {\line(0,1) {185}}
\put (111,-35) {\line(0,1) {185}}
\put (112,-35) {\line(0,1) {185}}
\put (80,-15) {\line(1,0) {240}}
\put (80,-14) {\line(1,0) {240}}
\put (80,-13) {\line(1,0) {240}}
{\color{red}
\put(110.0, 105.0) { \scriptsize $ \bullet $}
\put(115.7, 105.0) { \scriptsize $ \bullet $}
\put(121.4, 105.0) { \scriptsize $ \bullet $}
\put(127.1, 105.0) { \scriptsize $ \bullet $}
\put(132.8, 105.0) { \scriptsize $ \bullet $}
\put(138.5, 105.0) { \scriptsize $ \bullet $}
\put(144.2, 105.0) { \scriptsize $ \bullet $}
\put(149.9, 105.0) { \scriptsize $ \bullet $}
\put(155.7, 105.0) { \scriptsize $ \bullet $}
\put(161.4, 105.0) { \scriptsize $ \bullet $}
\put(167.1, 104.9) { \scriptsize $ \bullet $}
\put(172.8, 104.7) { \scriptsize $ \bullet $}
\put(178.5, 104.5) { \scriptsize $ \bullet $}
\put(184.2, 104.2) { \scriptsize $ \bullet $}
\put(189.9, 103.7) { \scriptsize $ \bullet $}
\put(195.5, 102.9) { \scriptsize $ \bullet $}
\put(201.1, 101.9) { \scriptsize $ \bullet $}
\put(206.6, 100.5) { \scriptsize $ \bullet $}
\put(212.0, 98.61) { \scriptsize $ \bullet $}
\put(217.0, 96.10) { \scriptsize $ \bullet $}
\put(221.6, 92.87) { \scriptsize $ \bullet $}
\put(225.4, 88.88) { \scriptsize $ \bullet $}
\put(228.3, 84.27) { \scriptsize $ \bullet $}
\put(230.4, 79.27) { \scriptsize $ \bullet $}
\put(231.8, 74.05) { \scriptsize $ \bullet $}
\put(232.7, 68.73) { \scriptsize $ \bullet $}
\put(233.3, 63.37) { \scriptsize $ \bullet $}
\put(233.6, 58.00) { \scriptsize $ \bullet $}
\put(233.8, 52.62) { \scriptsize $ \bullet $}
\put(233.9, 47.23) { \scriptsize $ \bullet $}
\put(233.9, 41.85) { \scriptsize $ \bullet $}
\put(234.0, 36.46) { \scriptsize $ \bullet $}
\put(234.0, 31.08) { \scriptsize $ \bullet $}
\put(234.0, 25.69) { \scriptsize $ \bullet $}
\put(234.0, 20.31) { \scriptsize $ \bullet $}
\put(234.0, 14.92) { \scriptsize $ \bullet $}
\put(234.0, 9.539) { \scriptsize $ \bullet $}
\put(234.0, 4.155) { \scriptsize $ \bullet $}
\put(234.0, -1.230) { \scriptsize $ \bullet $}
\put(234.0, -6.615) { \scriptsize $ \bullet $}
\put(234.0, -12.00) { \scriptsize $ \bullet $}
\put(240.615, -12) { \scriptsize $ \bullet $}
\put(247.23, -12) { \scriptsize $ \bullet $}
\put(253.845, -12) { \scriptsize $ \bullet $}
\put(260.46, -12) { \scriptsize $ \bullet $}
\put(267.075, -12) { \scriptsize $ \bullet $}
\put(273.69, -12) { \scriptsize $ \bullet $}
\put(280.305, -12) { \scriptsize $ \bullet $}
\put(286.92, -12) { \scriptsize $ \bullet $}
\put(293.535, -12) { \scriptsize $ \bullet $}
\put(300.15, -12) { \scriptsize $ \bullet $}
\put(306.765, -12) { \scriptsize $ \bullet $}
\put(313.38, -12) { \scriptsize $ \bullet $}
}
}
{\color{blue}
\put (72,152){\Large {$\cal M$}}
}
{\color{green}
\put (224,-42){\Large {$\cal T$}}
\put (231,-46){ \large {\it c}}
}
{\color{blue}
\put (71,103){\Large {\it 1}}
}
{\color{blue}
\put (92,-40){\Large {\it 0}}
}
{\color{green}
\put (304,-38){\Large {$\cal T$}}
}
\end{picture}
\caption{ ${\cal M}$ as a function of temperature $\cal T$.}
\label{graphM}
\end{figure}
Onsager and Kaufman did not publish their derivation,
which has led to speculation as to why they did not do so.
The first published derivation was not until 1952, by
C. N. Yang,{\bl \cite{Yang1952}} who later described the calculation as
``the longest in my career. Full of local, tactical tricks,
the calculation proceeded by twists and
turns.''{\bl \cite[p.11]{Yang1983}}
Onsager did outline what happened in an article published in
1971.{\bl \cite{Onsager1971a}} He starts by remarking that
correlations along a diagonal are particularly simple, and
gives the formula
\begin{equation} \label{2ndform}
\langle \sigma_{1,1} \sigma_{m,m} \rangle
\; = \; D_{m-1} \end{equation}
where $D_m$, like $\Delta_m$, is an
$m$ by $m$ determinant:
\begin{equation} \label{Toeplitzdet}
D_{m} \; = \; \raisebox{-0mm}{$
\left| \begin{array}{lllll}
c_0 & c_1 & c_2 & \cdots & \; c_{m-1} \\
c_{-1} & c_0 & c_1 & \cdots & c_{m-2} \\
\cdot & \cdot & \cdot & \cdots & \; \; \cdot \\
c_{1-m} & \cdot & \cdot & \cdots & \; c_0\end{array}
\right| $} \end{equation}
the $c_r$ being the coefficients in the Fourier expansion
\begin{equation} \label{genfn}
f( \omega ) \; = \; {\rm e}^{{\rm i } {\widehat \delta} (\omega)} \; = \;
\sum_{r=-\infty}^{\infty} c_r {\rm e}^{{\rm i } r \omega} \end{equation}
of the function{\footnote{I write the $\delta$ of {\bl \cite{Onsager1971a}}
as $\widehat{\delta}$.}}
\begin{equation}
\label{newdet}
{\rm e}^{{\rm i } {\widehat \delta} (\omega)} \; = \; \left( \frac{1- k \, {\rm e}^{{\rm i } \omega }}
{1- k \, {\rm e}^{ -{\rm i } \omega } } \right)^{1/2} \;\; . \end{equation}
It is clear that Onsager knew this when his paper with Kaufman
was submitted in May 1949, because footnote 7 of
{\bl \cite{KaufmanOnsager1949}}
states that ``It can be shown that $\delta' = \pi/2 - \omega/2 $ at
the critical temperature for correlations along a $45^{\circ}$
diagonal of the lattice.'' Indeed, this result does follow immediately
from (\ref{newdet}) when $k = 1$, provided we replace
$\delta$ in (1.4) by $\widehat{\delta}$. Certainly
(\ref{2ndform}) is true, being
the special case $J_3= v_3 = 0$ of equations (2.4), (5.13),
(6.10), (6.12) of Stephenson's pfaffian
calculation{\bl \cite{Stephenson1964}}
of the diagonal correlations of the triangular lattice Ising
model. The formula (\ref{2ndform})
agrees with (\ref{1stform}) when $m=2$.
The fact that the formula
depends on $H, H'$ only via $k$ is a consequence
of the property that the diagonal transfer matrices of two models,
with different values of $H$ and $H'$, but the same value
of $k$, commute.{\bl \cite[\S7.5]{book}}
In {\bl \cite{Onsager1971a}}, Onsager says that he first evaluated
$D_m$ in the limit $m \rightarrow \infty$ by using generating
functions to calculate the characteristic numbers
(eigenvalues) of the matrix
$D_m$ and that this leads to an integral equation
with a kernel of the form
\begin{equation} \label{kernel}
K(u,v) \; = \; K_1(u+v) + K_2(u-v)
\;\; . \end{equation}
He then obtained the determinant by taking the product
of the eigenvalues and says that ``This was the basis for
the first announcements of the result.''.
We show how this can be
done in Appendix A. One does indeed find a kernel of the form
(\ref{kernel}), and go on to obtain
\begin{equation} \label{detD}
\lim_{m \rightarrow \infty} D_m \; = \; (1-k^2)^{1/4}
\;\; , \end{equation}
which is the result (\ref{Isingmag}) that Onsager announced
in Cornell in 1948 and Florence in 1949.
\section{The second method}
\setcounter{equation}{0}
In his 1971 article{\bl \cite{Onsager1971a}} Onsager goes on to say that
after evaluating the particular determinant $D_{\infty}$ by the
integral equation method,
he looked for a method for the evaluation of a
general infinite-dimensional Toeplitz
determinant (\ref{Toeplitzdet}), with arbitrary entries $c_r$.
(The $c_r$ must tend to zero as $r \rightarrow \pm \infty$
sufficiently fast for the sum in (\ref{genfn}) to be uniformly
convergent when $\omega $ is real.)
As soon as he
tried rational functions of the form
\begin{equation} \label{genfn3}
f(\omega) \; = \; \frac{\prod (1-\alpha_j {\rm e}^{{\rm i } \omega })}
{\prod (1-\beta_k {\rm e}^{-{\rm i } \omega} )}
\;\; , \;\; \; \; |\alpha_j|, |\beta_k | <1 \;\; , \end{equation}
``the general result stared me in the face. Only, before I knew what sort
of conditions to impose on the generating function, we talked to Kakutani
and Kakutani talked to Szeg{\H o}, and the mathematicians
got there first.''
In another article in the same book {\bl \cite{Onsager1971b}},
Onsager gives further explanation of that comment,
saying that he had found
``a general formula for the evaluation of Toeplitz
matrices.{\color{blue} \footnote{Refs. {\bl \cite{Onsager1971a}},
{\bl \cite{Onsager1971b}} are reprinted in
Onsager's collected works, pages {\color{blue} { 232 -- 241}}
and {\color{blue} {37 -- 45}},
respectively.{\bl \cite{Onsager1996}} }}
The only thing I did not know was how to fill out the holes
in the mathematics and show the epsilons and the deltas
and all of that''. Onsager adds that six years later
the mathematician Hirschman told him that he could
readily have completed his proof by using
a theorem of Wiener's.
There is contemporary evidence to support these statements
in the form of correspondence in 1950 between Onsager and
Kaufman. There is also the photocopy of a
typescript mentioned in the Introduction, which deals with
the problem of calculating the $\Delta_k$ of
(\ref{Dmk}). Here I present a transcript of it, with approximately
the original layout and
pagination. A scanned copy is in Appendix B.
It seems highly likely that this is Kaufman's initial draft of
paper IV in their series of papers on Crystal Statistics (see points
3 and 4 in section 5).
Hand-written additions are shown in red (for contemporary additions)
and in blue (for probably later additions). Not all additions
are shown. The ``Fig. 1'' mentioned on page 2, after equation III.45,
may be Fig. 4 of {\bl \cite{KaufmanOnsager1949}}. Equation (17)
on page 7 follows from eqn. 89b of {\bl \cite{Onsager1944}} after
interchanging $H'$ with $H^*$, $\delta' $ with $\delta^*$ and
setting $H' = H$. There is an error in the equation between
(20) and (21): the second $e^{-2H}$ in the numerator should
be $e^{2H}$, as should the first $e^{-2H}$ in the denominator.
\vspace{12cm}
\newpage
\begin{figure}[hbt]
\begin{picture}(280,419) (74,250)
\setlength{\unitlength}{1.0pt}
\includegraphics[height=29.75cm]{Page1tr.pdf}
\end{picture}
\end{figure}
\newpage
\begin{figure}[hbt]
\begin{picture}(280,473) (76,280)
\setlength{\unitlength}{1.0pt}
\includegraphics[height=29.8cm]{Page2_revised.pdf}
\end{picture}
\end{figure}
\newpage
\begin{figure}[hbt]
\begin{picture}(280,426) (70,280)
\setlength{\unitlength}{1.0pt}
\includegraphics[height=29.2cm]{Page3tr.pdf}
\end{picture}
\end{figure}
\newpage
\begin{figure}[hbt]
\begin{picture}(280,416) (74,280)
\setlength{\unitlength}{1.0pt}
\includegraphics[height=29.8cm]{Page4tr.pdf}
\end{picture}
\end{figure}
\newpage
\begin{figure}[hbt]
\begin{picture}(280,416) (74,280)
\setlength{\unitlength}{1.0pt}
\includegraphics[height=29.8cm]{Page5tr.pdf}
\end{picture}
\end{figure}
\newpage
\begin{figure}[hbt]
\begin{picture}(280,416) (74,280)
\setlength{\unitlength}{1.0pt}
\includegraphics[height=29.8cm]{Page6tr.pdf}
\end{picture}
\end{figure}
\newpage
\begin{figure}[hbt]
\begin{picture}(300,462) (72,245)
\setlength{\unitlength}{1.0pt}
\includegraphics[height=29.2cm]{Page7tr.pdf}
\end{picture}
\end{figure}
\newpage
\begin{figure}[hbt]
\begin{picture}(260,448) (74,280)
\setlength{\unitlength}{1.0pt}
\includegraphics[height=29.7cm]{Page8tr.pdf}
\end{picture}
\end{figure}
\newpage
\section{Summary of the draft}
\setcounter{equation}{0}
\renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
The names at the top of the first page have been added by hand
to the typescript, so it is not immediately obvious who are the
authors.
However, from the first sentence and their frequent references to
paper III , specifically to III.45 (particularly in conjunction with their
correspondence discussed below) it is clear that the paper is
by Onsager and Kaufman jointly. Footnote 1 in the first line is not
given, but is presumably their paper III.
The paper begins by quoting III.43, modified to (\ref{III.43})
above, for the isotropic case $H' = H$.
It focusses on the problem of calculating a $j$ by $j$
Toeplitz determinant $\Delta_j$,
of the general form
(\ref{Toeplitzdet}), in the limit $j \rightarrow \infty$.
The $c_i$ are the coefficients of the
Fourier expansion (\ref{genfn}) of some function $f( \omega)$,
initially allowed to be arbitrary.
It first takes $f(\omega)$
\begin{equation} \label{gform}
f(\omega ) \; = \; \frac{a_0 + a_1 {\rm e}^{{\rm i } \omega} + \ldots +
a_p {\rm e}^{p {\rm i } \omega}}
{b_0 + b_1 {\rm e}^{-{\rm i } \omega} + \ldots + b_q {\rm e}^{-q {\rm i } \omega}}
\end{equation}
and shows that
\begin{equation} \label{vanishes}
\Delta_j = 0 \; \; {\rm if} \; \; a_0 = 0 \; \; {\rm and } \; \; j > q \;\; . \end{equation}
It then takes $f(\omega)$ to be of the form
(\ref{genfn3}), or more specifically
\begin{equation} \label{genfn4}
f(\omega) \; = \; \frac{g({\rm e}^{{\rm i } \omega})}{h({\rm e}^{-{\rm i } \omega})}
\;\; , \end{equation}
where \begin{displaymath} g({\rm e}^{{\rm i } \omega }) =
\prod_{t=1}^p (1-\alpha_t {\rm e}^{{\rm i } \omega} )
\;\; , \;\; h({\rm e}^{-{\rm i } \omega }) = \prod_{r=1}^q (1-\beta_r {\rm e}^{-{\rm i } \omega} )
\;\; , \end{displaymath}
and goes on to show, using (\ref{vanishes}), that
\begin{displaymath} \Delta_j \; = \; \prod_{t=1}^p
\prod_{r=1}^q (1- \alpha_t \beta_r )
\end{displaymath}
provided that $ j \geq q$. This is an algebraic identity, true for
all
$ \alpha_t, \beta_r$. It is a trivial generalization then to say that
if
\begin{displaymath}
f(\omega ) = \frac{\prod_{j=1}^p (1-\alpha_j {\rm e}^{{\rm i } \omega} )^{m_j}}
{ \prod_{k=1}^q (1-\beta_k {\rm e}^{-{\rm i } \omega} )^{n_k} }\;\; , \end{displaymath}
then
\begin{equation} \label{formoff} \Delta_\infty \; = \; \prod_{j=1}^p
\prod_{k=1}^q (1- \alpha_j \beta_k )^{m_j n_k} \end{equation}
for positive integers $m_j, n_k$.
For $\Delta_j$, $f(\omega) = {\rm e}^{{\rm i } \delta(\omega)}$ is given by
(\ref{eidelta}).
This is of the general form (\ref{formoff}), with
$p = q=2$ and
\begin{displaymath} \alpha_1 = \beta_1= \coth H {\rm e}^{-2H} \;\; , \;\; \alpha_2 =
\beta_2 = \tanh H {\rm e}^{-2H} \;\; , \;\; m_1 = m_2 = n_1 = n_2 = \textstyle \frac{1}{2}
\end{displaymath}
so the $m_j, n_j$ are no longer integers. The paper assumes that
(\ref{formoff}) can be generalized to this case,\footnote{This vital point
is considered by Onsager in the first letter quoted in section 5.} so obtains
(using $\tanh H = {\rm e}^{-2 H^*}$)
\begin{equation} \Delta_{\infty} \; = \; \{ [1 - {\rm e}^{4(H^*-H)}][1-{\rm e}^{-4(H^*+H)}]
[1-{\rm e}^{-4H}]^2 \}^{1/4} \;\; . \end{equation}
If we transpose the second determinant (\ref{Dmk}) in
(\ref{III.43}), then its generating function is not $f( \omega)$
but ${\rm e}^{2{\rm i } \omega } f(\omega)$. This corresponds to the form
(\ref{gform}), but with $a_0 = 0$ (and $a_1 = 0$), so $\Delta_{-\infty}$
should vanish because of (\ref{vanishes}). Thus (\ref{III.43})
gives
\begin{displaymath} \lim_{j \rightarrow \infty} \langle \sigma_{1,1} \sigma_{1, 1+j} \rangle
\; = \; \cosh^2 H^* \, \{ [1 - {\rm e}^{4(H^*-H)}][1-{\rm e}^{-4(H^*+H)}]
[1-{\rm e}^{-4H}]^2 \}^{1/4} \;\; . \end{displaymath}
For the isotropic case $H' = H$, ${\rm e}^{-2 H^*} = \tanh H$, so we obtain,
using (\ref{defM0}),
\begin{displaymath} M_0^2 \; = \; \lim_{j \rightarrow \infty} \langle \sigma_{1,1}
\sigma_{1, 1+j} \rangle
\; = \; (1 - 1/\sinh^4 2 H)^{1/4} \; = \; (1-k^2)^{1/4} \;\; , \end{displaymath}
in agreement with (\ref{Isingmag}).
\subsubsection*{Generalization to the anisotropic case}
In their paper III,{\bl \cite{KaufmanOnsager1949}}
Kaufman and Onsager focus on the isotropic case $H' = H$, as
does the above draft. It does in fact appear that their results
generalize immediately to the
anisotropic case when $H, H'$ are independent, and
$H^*$, $\delta'$ are defined as in {\bl \cite{Onsager1944}} (i.e.
by eqns. (\ref{defH*}), (\ref{deldelp}), (\ref{eidelta}) above),
and that (\ref{III.43}), (\ref{1stform}) are then correct as
written.\footnote{I have verified that (\ref{1stform}) agrees
with (\ref{2ndform}) - (\ref{newdet}). I have not
verified (\ref{III.43}) directly, but
have checked
algebraically, with the aid of Mathematica, that it
agrees with eqn. 56 of {\bl \cite{MPW1963}} for $ j \leq 8$.}
Replacing equation (24) of the draft by (\ref{eidelta}), and
repeating the last few steps, one obtains
\begin{eqnarray} \lim_{j \rightarrow \infty} \langle \sigma_{1,1}
\sigma_{1, 1+j} \rangle
& = & \cosh^2 H^* \, \{ [1 -\coth^2 H' {\rm e}^{-4H}][1- \tanh^2 H' {\rm e}^{-4H}]
[1-{\rm e}^{-4H}]^2 \}^{1/4} \nonumber \\
& = & [1 - 1/(\sinh 2 H \sinh 2 H' )^2 ]^{1/4} = (1-k^2)^{1/4}
\;\; , \end{eqnarray}
which again agrees with (\ref{Isingmag}).
\section{Further comments}
\setcounter{equation}{0}
There are several letters on the Onsager archive in Norway
between Onsager and Kaufman relating to this second
method, at
\noindent {\footnotesize
{\color{blue}
\href{http://www.ntnu.no/ub/spesialsamlingene/tekark/tek5/research/009_0097.html}
{http://www.ntnu.no/ub/spesialsamlingene/%
tekark/tek5/research/009{\_}0097.html}}}
and
\noindent {\footnotesize
{\color{blue}
\href{http://www.ntnu.no/ub/spesialsamlingene/tekark/tek5/research/009_0096.html}
{http://www.ntnu.no/ub/spesialsamlingene/%
tekark/tek5/research/009{\_}0096.html}}}
\noindent We shall refer to these two files as 0097 and 0096, respectively.
\item{{\bf 1.}} On pages 21 -- 24 of 0097, (also in {\bl \cite{JSP1995} })
is a letter dated April 12, 1950 from
Onsager to Kaufman giving the argument of the above draft up to
equation 15 therein. Onsager states that the result admits
considerable
generalization. He defines
\begin{equation} \eta_+ (\omega) = \log g(\e^{{\rm i} \omega}) \;\; , \;\; \eta_- (\omega) = - \log \hom
\end{equation}
so $f( \omega ) = {\rm e}^{ \eta_+ (\omega) + \eta_- (\omega) }$.
He remarks that if $\log f(\omega)$ is analytic in a strip which contains
the real axis,
then these functions may be approximated by polynomials which
have ``no wrong zeros'' in such a manner that the corresponding
determinants converge. This implies that the $\alpha_j, \beta_j$ all
have modulus less than one. He goes on to say that (15) is equivalent
to
\begin{equation} \label{Tform}
\log \Delta_{\infty} \; = \; \frac{\i}{2 \pi} \, \int_{\omega = 0}^{2 \pi} \eta_+ \, {\rm {d}} \eta_- ( \omega )
\;\; , \end{equation}
If
\begin{equation} \log f( \omega ) \; = \; \sum_{n=-\infty} ^{\infty} b_n {\rm e}^{{\rm i } n \omega }
\end{equation}
and $b_0 = 0$, then (\ref{Tform}) is equivalent to
\begin{equation} \label{Szego}
\log \Delta_{\infty} \; = \; \sum_{n=1} ^{\infty} n b_n b_{-n}
\;\; , \end{equation}
which is the result now known as Szeg{\H o}'s
theorem.{\bl \cite{Szego1952,Kac1954,MPW1963}}
Onsager then says ``we get the degree of order from C.S. III without
much trouble. It equals $(1-k^2)^{1/8}$ as before.''
\item{\bf 2. } On pages {{\color{blue} } 32, 33} of 0096 is a letter
from Kaufman to Onsager. It is undated, but was presumably
written after April 12th and before April 18th 1950 (the date
of the next letter we discuss). She thanks him for his
letter and comments that his method is
elegant and simple, far superior to the integral-equations
method for this purpose.
She also mentions the formula (\ref{newdet}), i.e.
\begin{displaymath} f(\omega ) \; = \; \left( \frac{1-k {\rm e}^{{\rm i } \omega}}{1-k {\rm e}^{-{\rm i } \omega}} \right)^{1/2}
\end{displaymath}
for the generating function for long-range order along a diagonal.
She goes on to say that the mathematician Kakutani had
written to her saying that he had spoken to Onsager about
this. He had been very interested in Onsager's letter, which he
saw in her house and immediately copied down.
\item{\bf 3.} On pages
{{\color{blue} } 30, 31} of 0096
is a letter from Kaufman to Onsager dated April 18, 1950,
she says she's glad to hear that Onsager is coming to
Princeton the following Friday and refers to Onsager's 1944
paper I, to her 1949 paper II, to their joint paper III,
and then, significantly to a paper IV ``on long-range order''.
She says she would like Onsager to see her m.s.
\item{\bf 4.} Then on page {{\color{blue} } 34} of 0096
is a letter from Kaufman to Onsager dated May 10,
saying ``Here is a draft of Crystal Statistics IV''.
\vspace{4mm}
All this fits with Onsager's recollections of
1971{\bl \cite{Onsager1971a, Onsager1971b}}.
He and Kaufman had evaluated the determinants in
III.43{\bl \cite{KaufmanOnsager1949}}, or alternatively in
(\ref{Toeplitzdet}) - (\ref{newdet}), by the integral equation method
(Appendix A) and by the general Toeplitz determinant formulae
(15) of the draft, i.e (\ref{Tform}).
He preferred the second method, but apparently was content to let
Kakutani and Szeg{\H o} take it over and put in the
rigorous mathematics. Kaufman's draft of paper IV was never
published, but it seems highly likely that the draft given here
is indeed that.
Szeg{\H o} did publish his resulting
general theorem{\bl \cite{Szego1952}},
{\bl \cite[p.76]{Kac1954}}
on the large-size limit of a Toeplitz
determinant, but not until 1952. He also restricted his attention
to Hermitian forms, where
$f^* (\omega ) = f(\omega )$ and
$b^*_n = b_{-n}$,{\bl \cite[footnote 17] {MPW1963}} so his
result needed further generalization before it could be
applied to to the Ising model magnetization.
The first derivation of $M_0$ published was in 1952 by C.N.
Yang.{\bl \cite{Yang1952} } He used the spinor operator algebra
to write $M_0$
as the determinant of an $L$ by $L$ matrix and evaluated the
determinant by calculating the eigenvalues of the matrix in
the limit $L \rightarrow \infty $. Intriguingly, he mentions Onsager and
Kaufman's papers I, II and III, and in footnote 10 of his paper,
Yang thanks Kaufman for showing him her notes on Onsager's work.
However, his method is quite different from theirs.
Later, combinatorial ways were found of writing the partition function
of the Ising model on a finite lattice directly as a determinant or a
pfaffian (the square root of an anti-symmetric
determinant).{\bl \cite{KacWard, HurstGreen}}. Then it was
realised that the problem could be solved by first expressing it
as one of filling a planar lattice with
dimers.{\bl \cite{Kasteleyn1961, Fisher,TemperleyFisher,
Kasteleyn1963}}
In 1963 Montroll, Potts and Ward{\bl \cite{MPW1963}} used
these pfaffian
methods to show that
$\langle \sigma_{1,1} \sigma_{1,j+1} \rangle $ could be written as a
single Toeplitz determinant.
They evaluated its value in the limit
$j \rightarrow \infty $ limit by using Szeg{\H o}'s theorem.
So there is no reason to doubt that Onsager and Kaufman
had derived
the formula (\ref{Isingmag}) by May 1949, and ample evidence that
they had obtained the result by what is now known as Szeg{\H o}'s
theorem by May 1950. They did not publish the calculation,
perhaps because
the mathematicians beat them to the remaining problem of
``how to fill out the holes
in the mathematics and show the epsilons and the deltas
and all of that''.
\section{Acknowledgements}
The author is most grateful to John Stephenson for giving
him a copy of the draft paper reproduced here, and
thanks Jacques Perk for telling him of
ref.{\bl \cite{JSP1995}} and the material on
the Lars Onsager Online archive under ``Selected research material
and writings'' at
\noindent {\footnotesize
{\color{blue}
\href{http://www.ntnu.no/ub/spesialsamlingene/tekark/tek5/arkiv5.php}
{http://www.ntnu.no/ub/spesialsamlingene/tekark/tek5/arkiv5.php}}}
\noindent in particular items 9.96 and 9.97, and for pointing out a number of
typographical errors in the original form of this paper.
He also thanks Harold Widom for sending him a copy of the letter
{\bl \cite{JSP1995}}, and Richard Askey for alerting him to
page 41 of Onsager's collected works. He is grateful to the reviewers
for a number of helpful comments.
\section*{Appendix A: An integral equation method}
\setcounter{equation}{0}
\renewcommand{\theequation}{A\arabic{equation}}
We regard $D= D_m$ as a matrix and
write $z$ for ${\rm e}^{{\rm i } \omega }$ and $C(z)$ for
${\rm e}^{{\rm i } {\widehat \delta} ( \omega)}$, so (\ref{genfn}), (\ref{newdet}) become
\begin{equation} \label{defC}
C(z) \; = \; { \left( \frac {1-k z}{1-k/z} \right) }^{1/2} =
\sum_{m=-\infty}^{\infty} c_r z^r \;\; . \end{equation}
For $0 < k < 1$, the determinant of $D_{m}$ tends to a non-zero limit as
$m \rightarrow \infty$.
The eigenvalues of $D$ itself lie on an arc in the
complex plane, and appear to tend to a continuous distribution as
$m \rightarrow \infty $. By contrast, in this limit the individual eigenvalues
$\lambda_r$ of $D^T D$ occur in discrete pairs, lying on the real axis,
between 0 and 1. If one orders them so that
$\lambda_r \leq \lambda_{r+1}$, then for given $r$
the eigenvalue $\lambda_r$ appears to tend to a limit as
$m \rightarrow \infty$, and these limiting values tend to
one as $r \rightarrow \infty$.
We assume these properties and seek to calculate the
eigenvalues of $D^T D$ in the limit $m \rightarrow \infty $,
and hence the determinant of $D^T D$, which is
$(\det D)^2$.
Writing them as $\lambda^2$ , we can write the
eigenvalue equation as
\begin{equation} \label{eigveceqn}
\lambda x = D y \;\; , \;\; \lambda y = D^T x \;\; . \end{equation}
For finite $m$, $x = \{ x_0, x_1, \ldots, x_{m-1} \}$,
$y = \{ y_0, y_1, \ldots, y_{m-1} \}$. Let $P$ be the $m$ by
$m$ matrix with entries
\begin{displaymath} P_{ij} \; = \; \delta_{i, m-1-j} \end{displaymath}
so $P x = \{ x_{m-1}, x_{m-2}, \ldots, x_{0} \}$. Then
\begin{displaymath} D^T = PDP \;\; , \;\; D^T D = (PD)^2 \;\; . \end{displaymath}
For finite $m$, the eigenvalues of $D^T D$ are distinct,
so the eigenvectors are those of $PD$. For large $m$
and a given $\lambda$, this means that the elements
$x_i, y_i$ are of order one if $i$ is close to zero or
to $m-1$, but tend to zero in between, when both
$i$ and $m-i$ become large.
However, in the limit $m \rightarrow \infty$, the eigenvalues
occur in equal pairs. One can then choose the two
corresponding eigenvectors so that one has the property
that
\begin{equation} \label{1sttype}
x_i, y_i \rightarrow 0 \; \; \; {\rm as \; \; } i \rightarrow \infty \;\; ,
\end{equation}
while for the other eigenvector $ x_{m-i} , y_{m-i} \rightarrow 0 $.
The eigenvectors are transformed one to another by
replacing $x_i$ and $y_i$ by $y_{m-1-i}$ and $x_{m-1-i}$,
respectively.
Since the eigenvalues are equal, we can and do restrict our
attention to eigenvectors with the property (\ref{1sttype}).
\subsubsection*{Generating functions}
Taking the limit $m \rightarrow \infty$, we can write
(\ref{eigveceqn}) explicitly as
\begin{equation} \label{eigeqn}
\lambda \, x_i = \sum_{j=0}^{\infty} c_{j-i} y_j \;\; , \;\;
\lambda \, y_i = \sum_{j=0}^{\infty} c_{i-j} x_j \end{equation}
for $i = 0, 1,2, \ldots \, $.
We can extend these equations to negative $i$, defining
$x_i, y_i$ to then be given by the left-hand sides of the
equations. Let $X(z), Y(z)$ be the generating functions
\begin{displaymath} X(z) = \sum_{i=-\infty}^{\infty} x_i z^i \;\; , \;\;
Y(z) = \sum_{i=-\infty}^{\infty} y_i z^i \;\; . \end{displaymath}
Then (\ref{eigeqn}) gives
\begin{equation} \label{inteqns}
\lambda \, X(z) = C(1/z) \tilde{Y}(z) \;\; , \;\;
\lambda \, Y(z) = C(z) \tilde{X}(z) \;\; , \end{equation}
where
\begin{displaymath} \tilde{X}(z) = \sum_{i=0}^{\infty} x_i z^i \;\; , \;\;
\tilde{Y}(z) = \sum_{i=0}^{\infty} y_i z^i \;\; . \end{displaymath}
A Wiener--Hopf argument (or simple contour integration on each
term) gives, for $|z| < 1 $,
\begin{equation} \label{WH}
\tilde{X}(z) = \frac{1}{2 \pi {\rm i } } \oint \frac{X(w) }{w-z} \, {\rm {d}} w
\;\; , \;\; \tilde{Y}(z) = \frac{1}{2 \pi {\rm i } } \oint \frac{Y(w) }{w-z} \, {\rm {d}} w
\;\; , \end{equation}
the integrations being round the unit circle in the complex
$w$-plane.
Thus (\ref{inteqns}) is a pair of coupled integral equations
for $\lambda, X(z), Y(z)$.
\subsubsection*{The integral equations in terms of elliptic
functions}
To get rid of the square root in (\ref{defC}) we introduce
Jacobi elliptic functions of modulus $k$ and set
\begin{displaymath} z = k \, {\rm {sn}}^2 u \;\; . \end{displaymath}
If $K, K'$ are the complete elliptic integrals and
$ \label{defalpha}
u \; = \; \alpha - K -{\rm i } K'/2 $,
then from (8.181), (8.191) of {\bl \cite{GR}},
or (15.1.5), (15.1.6) of {\bl \cite{book}},
\begin{displaymath} z \; = \; s \, \prod_{n=1}^{\infty}
\frac{(1+p^{4n-3}/s)^2(1+p^{4n-1}s)^2}
{(1+p^{4n-3} \, s)^2(1+p^{4n-1}/s)^2} \end{displaymath}
where
\begin{displaymath} p = q^{1/2} = {\rm e}^{-\pi K'/2 K } \;\; , \;\; s = {\rm e}^{{\rm i } \pi \alpha/K }
\;\; , \;\; 0 < q < 1 \;\; . \end{displaymath}
If $ 0 < k <1$, then $0 < p,q < 1$, so we see that
$z$ goes round the unit circle as $\alpha$ goes from
$- K $ to $K$ along the real axis, i.e.
$u$ goes along a horizontal line in the complex plane
from $-2K -{\rm i } K'/2$ to $-{\rm i } K'/2$.
From the elliptic function relations
$ {\rm {cn}}^2 u = 1 - {\rm {sn}}^2 u$, ${\rm {dn}}^2 u = 1 - k^2 {\rm {sn}}^2 u$,
\begin{eqnarray} \label{fnC}
C(z) & = & {\rm i } \, \frac{{\rm {sn}} \, u \, {\rm {dn}} \, u}{{\rm {cn}} \, u} \;\; ,
\nonumber \\
& = & \prod_{n=1} ^{\infty}
\frac{(1+(-1)^n p^{2n-1} s)(1-(-1)^n p^{2n-1}/s)}
{(1 -(-1)^n p^{2n-1} s)(1 +(-1)^n p^{2n-1}/s)}
\end{eqnarray}
Replace $w$ in (\ref{WH}) by
\begin{displaymath} w = k \, {\rm {sn}}^2 v \;\; , \end{displaymath}
where
\begin{equation} \label{domain}
{\rm Im} (v) = -K'/2 \;\; , \;\; -K'/2 < {\rm Im} (u) <
K'/2 \;\; . \end{equation}
Then
\begin{equation} {\rm {d}} w \; = \; 2 k \, {\rm {sn}} \, v \, {\rm {cn}} \, v \, {\rm {dn}} \, v \, {\rm {d}} v \;\; . \end{equation}
It is helpful to set
\begin{equation} \label{XYhat}
X(u) \; = \; \frac{- {\rm i } \widehat{X} (u)}{{\rm {sn}} \, u\, {\rm {dn}} \, u } \;\; , \;\;
Y(u) \; = \; \frac{ {\rm i } \widehat{Y} (u)}{{\rm {cn}} \, u } \;\; . \end{equation}
Then the integral equations (\ref{inteqns}) become
\begin{equation} \label{inteq2}
\lambda \widehat{X} (u) = \frac{1}{2\pi} \int M(u,v)
\widehat{Y} (v) {\rm {d}} v \;\; , \;\;
\lambda \widehat{Y} (u) = \frac{1}{2\pi} \int M(v,u)
\widehat{X} (v) {\rm {d}} v
\;\; , \end{equation}
where the integrations are from $-2K -{\rm i } K'/2$ to $-{\rm i } K'/2$
and
\begin{displaymath} M(u,v) \; = \; 2 \, \frac{{\rm {cn}} \, u\, {\rm {sn}} \, v \, {\rm {dn}} \, v }
{{\rm {sn}}^2 \, v - {\rm {sn}}^2 \, u} \;\; . \end{displaymath}
Using the Liouville theorem-type arguments of sections
15.3, 15.4 of {\bl \cite{book}}, one can establish that
\begin{displaymath} M(u,v) \; = \; \frac{{\rm {dn}} (v+u)}{{\rm {sn}} (v+u) } +
\frac{{\rm {dn}} (v-u)}{{\rm {sn}} (v-u) } \;\; . \end{displaymath}
We see that this kernel is indeed of the form (\ref{kernel}).
\vspace{2mm}
\setlength{\jot}{4mm}
We require the functions $\tilde{X}(z), \tilde{Y}(z)$ to be
analytic for $|z| <1$, i.e. for $|{\rm Im }(u) | <K'/2$, and in
particular for real $u$. From (\ref{inteqns}), (\ref{fnC}) and
(\ref{XYhat}),
$ \tilde{X}(z) = \lambda \widehat{Y}(u)/{\rm {sn}} u \, {\rm {dn}} u$,
$ \tilde{Y}(z) = \lambda \widehat{X}(u)/{\rm {cn}} u$.
Since ${\rm {sn}} \, 0 = {\rm {cn}} \, K = 0$, it follows
that
\begin{equation} \label{restrict}
\widehat{X}(K) = \widehat{Y}(0) = 0 \;\; . \end{equation}
Also, $M(u,v)$ is an analytic function of
$u$ within the domain (\ref{domain}), and we can negate
$u$ within the domain. Since $M(u,v)$ is an
even function of $u$, it follows from (\ref{inteq2})
(provided $\lambda$ is not zero)
that $\widehat{X}(u)$ is also an even function.
Similarly, $\widehat{Y}(u)$ is an odd function.
\subsubsection*{Solution by Fourier transforms}
The sum and difference form of the kernel ${M}(\alpha, \beta ) $
suggests solving (\ref{inteq2}) by Fourier transforms.
The functions $\widehat{X}, \widehat{Y}, M$ are anti-periodic of
period $2 K$, while $\widehat{X}(u)$, $\widehat{Y}(u)$
are even and odd, respectively. We therefore try
\begin{equation} \label{guess}
{\widehat{X}}(u) = A \cos \pi (2r-1) u/2K \;\; , \;\;
{\widehat{Y}}(u) = B \sin \pi (2r-1) u/2K \;\; , \end{equation}
$r$ being a positive integer. We note that this ansatz immediately
satisfies (\ref{restrict}).
Integrating round a period rectangle and using Cauchy's theorem,
or using (8.146) of {\bl \cite{GR}}, for all integers $r$ we obtain
\setlength{\jot}{5mm}
\begin{equation} \frac{1 }{2 \pi } \, \int \frac{{\rm {dn}} (v-u)}
{{\rm {sn}} (v-u)} \, {\rm e}^{{\rm i } \pi (2r-1) v/2K} \, {\rm {d}} v =
{\rm i } \, \frac{{\rm e}^{{\rm i } \pi (2r-1) u/2K}}{1+q^{2r-1}} \;\; . \end{equation}
Negating $2r-1$ and/or $u$ (but {\em not} $v$) and
taking appropriate sums and differences, we obtain
the identities, true when both $v_0+u$ and $v_0-u$ have
imaginary parts between $-2K'$ and zero and the integration
is from
$v_0 - 2K$ to $v_0$ :
\begin{equation} \frac{1}{2 \pi} \int M(u,v) \sin {\textstyle \frac {\pi (2r-1) v}{2K} }
{\rm {d}} v = \frac{1-q^{2r-1}}{1+q^{2r-1}} \,
\cos {\textstyle \frac {\pi (2r-1) u}{2K} } \end{equation}
\begin{equation} \frac{1}{2 \pi} \int M(v,u) \cos {\textstyle \frac {\pi (2r-1) v}{2K} }
{\rm {d}} v = \frac{1-q^{2r-1}}{1+q^{2r-1}} \,
\sin {\textstyle \frac {\pi (2r-1) u}{2K} } \;\; .
\end{equation}
We see that the ansatz (\ref{guess}) does indeed satisfy
(\ref{inteq2}), with $A = B= 1$ and
\begin{equation} \label{lambda}
\lambda = \frac{1-q^{2r-1}}{1+q^{2r-1}}
\;\; . \end{equation}
More generally, we can allow $\widehat{X}(u)$,
$\widehat{Y}(u)$ each to be an arbitrary
linear combination of $\exp[{\rm i } \pi (2r-1) u/2K]$ and
$\exp[{ - {\rm i } \pi (2r-1) u/2K}]$. We obtain the above solution, plus
another where $\lambda = 0$. However, this
second solution does not satisfy
the necessary condition (\ref{restrict}).
\subsubsection*{Determinant of $D$}
For every positive integer $r$ there is one and only one
eigenvector of $D^T D$ of the form (\ref{1sttype}), with
eigenvalue $\lambda^2$, $\lambda$ being given by
(\ref{lambda}). However, from the discussion
before (\ref{1sttype}), there is also an equal
eigenvalue whose eigenvector has elements
$x_i, y_i$ of order 1 only when $i$ is close to $m$. Hence,
using (8.197.4) of {\bl \cite{GR}} or
(15.1.4b) of {\bl \cite{book}},
\begin{equation} \det D^T D = ( \det D )^2 = \prod_{r=1}^{\infty}
\left( \frac{1-q^{2r-1}}{1+q^{2r-1}} \right)^4 = k'
= (1-k^2)^{1/2} \;\; . \end{equation}
so
\begin{equation} \det D \; = \; (1-k^2)^{1/4} \end{equation}
as in (\ref{detD}).
\newpage
\section*{Appendix B: The typewritten draft}
\setcounter{equation}{0}
\renewcommand{\theequation}{B\arabic{equation}}
\begin{figure}[hbt]
\begin{picture}(320,340) (14,355)
\setlength{\unitlength}{1.0pt}
\includegraphics[height=25cm]{Onsager1.pdf}
\end{picture}
\end{figure}
\newpage
\begin{figure}[hbt]
\begin{picture}(320,470) (14,200)
\setlength{\unitlength}{1.0pt}
\includegraphics[height=25cm]{Onsager2.pdf}
\end{picture}
\end{figure}
\newpage
\begin{figure}[hbt]
\begin{picture}(320,470) (14,200)
\setlength{\unitlength}{1.0pt}
\includegraphics[height=25cm]{Onsager3.pdf}
\end{picture}
\end{figure}
\newpage
\begin{figure}[hbt]
\begin{picture}(320,470) (14,200)
\setlength{\unitlength}{1.0pt}
\includegraphics[height=25cm]{Onsager4.pdf}
\end{picture}
\end{figure}
\newpage
\begin{figure}[hbt]
\begin{picture}(320,470) (14,200)
\setlength{\unitlength}{1.0pt}
\includegraphics[height=25cm]{Onsager5.pdf}
\end{picture}
\end{figure}
\newpage
\begin{figure}[hbt]
\begin{picture}(320,470) (14,200)
\setlength{\unitlength}{1.0pt}
\includegraphics[height=25cm]{Onsager6.pdf}
\end{picture}
\end{figure}
\newpage
\begin{figure}[hbt]
\begin{picture}(320,470) (14,200)
\setlength{\unitlength}{1.0pt}
\includegraphics[height=25cm]{Onsager7.pdf}
\end{picture}
\end{figure}
\newpage
\begin{figure}[hbt]
\begin{picture}(320,470) (14,200)
\setlength{\unitlength}{1.0pt}
\includegraphics[height=25cm]{Onsager8.pdf}
\end{picture}
\end{figure}
\newpage
|
\section{Introduction}
Flux calibrations in physical units for astronomical instruments are required to
make comparisons to physical models of observed objects (Kent et al. 2009). In
particular, one of the main incentives for accurate absolute flux standards is
the requirement for measuring the relative fluxes of redshifted SN Ia spectra in
the rest frame in order to constrain the parameters of the dark energy. These
constraints depend only on the precision of the ratio of fluxes from one
wavelength to another and not on the absolute flux level. Quantitative
descriptions of dark energy are significantly improved when the relative flux
with wavelength is known to an accuracy of 1\% or better (Aldering et al.
2004).
Absolute flux calibrations of spectrometers and photometers are normally derived
from observations of standard stars with well-known spectral energy
distributions (SEDs). For all \emph{Hubble Space Telescope} instruments, all
flux calibrations are traceable to three primary WD standards, G191B2B, GD71,
and GD153. The slopes of these WD SEDs are determined by non-local thermodynamic
equilibrium (NLTE) model calculations using the Hubeny Tlusty Version-203 code
for pure hydrogen atmospheres (Bohlin 2000, Bohlin et al. 2001, Hubeny \& Lanz
1995, Tremblay \& Bergeron 2009). The effective temperature and gravity are
determined by fitting the models to ground-based observations of the Balmer line
profiles (Finley et~al. 1997).
The absolute flux of the models for these three primary standards is set by
Space Telescope Imaging Spectrograph (STIS) relative spectrophotometry (Bohlin
\& Gilliland 2004, Bohlin 2007a) and the Megessier (1995) absolute flux for
Vega of $3.46\times10^{-9}$~erg cm$^{-2}$ s$^{-1}$~\AA$^{-1}$ at
5556~\AA~(3560~Jy or 3562~Jy for vacuum wavelengths). As discussed in the review
by Hayes (1985), by Megessier (1995), and in \S 4.1.1 below, there is a small
uncertainty in the 5556~\AA\ flux; but this uncertainty just affects the overall
level and $not$ the slope (i.e. ``color'') of the WD models used for \emph{HST}
flux calibrations. Despite suggestions of variations, Hayes (1985) discusses the
evidence for variability of Vega and concludes that the star is likely not
variable. However, Engelke et al. (2010) present evidence for a 0.08 mag
variation of Vega at visible wavelengths.
To compare with previously published calibrations of the \emph{Spitzer Space
Telescope} in the four IRAC bands (Fazio et~al. 2004, Reach et~al. 2005,
hereafter Re05), a set of new observations of white dwarf (WD), A stars, and
solar-analog G stars were made near the end of the cold mission. \emph{Spitzer}
data in the IRS blue peakup channel (Houck et~al. 2004) or the MIPS 24~$\mu$m
band (Rieke et~al. 2004, Engelbracht et~al. 2007) were included to ensure that
debris disks or red companions do not contaminate the results. These new
observations are supplemented by more data sets for the same stars from the
\emph{Spitzer} archive. Table 1 lists the 14 stars with \emph{HST} based SEDs
that are used for the comparison with the absolute flux calibrations of
Re05 for IRAC, the IRS Instrument
Handbook\footnote{http://ssc.spitzer.caltech.edu/irs/irsinstrumenthandbook/IRS\_Instrument\_Handbook.pdf},
and Engelbracht et~al. (2007) for the MIPS 24~$\mu$m channel. The \emph{HST}
flux distributions are all in the
CALSPEC\footnote{http://www.stsci.edu/hst/observatory/cdbs/calspec.html/}
database. The Re05 IRAC calibration is based on four A-star SEDs from an
extension of the original Cohen CWW network (Cohen, et~al. 1992a, Cohen et~al.
1999, Cohen et~al. 2003, Cohen 2007), as validated by Price et~al. (2004). Two
of these four primary A star calibrators, HD165459 and 1812095, are included in
Table 1, while our other five A stars are listed as IRAC candidate primary
calibrators by Re05. For a comparison of the \emph{HST} SEDs with the Cohen flux
distributions, see Bohlin \& Cohen (2008), whose minor revisions include average
fluxes that are $\sim$0.5\% lower in the IRAC wavelength range for the set of
seven A stars in Table 1. The K~star calibrators of Re05 are not utilized,
because of the extra complexity of modeling the molecular absorption and because
Re05 used only A~stars to define their final IRAC calibration constants.
In this paper, \S 2 covers the fundamental equations and concept of photometric
flux calibrations, while \S 3 compares the published calibration constants for
four IRAC imaging modes to those that are derived from the \emph{HST} based
SEDs. \S 4 compares our results with the IRAC calibrations of Re05 and with the
8~$\mu$m fluxes of Rieke et~al. (2008, hereafter Ri08). \S 5 includes
suggestions for future efforts to improve the flux calibration accuracy. Finally
in Appendix A, the \emph{HST} method of photometric flux calibration is
illustrated for six \emph{Spitzer} modes. The Appendix is not absolutely
essential to the main thrust of this paper but does expand on several points as
forward referenced in the main body. In addition, the Appendix attempts to
provide a cohesive mathematical foundation for the student or practicioner
of the art of flux calibration.
\section{Generic Calibration Constants}
\subsection{Equations}
This section compares the \emph{Spitzer} flux calibration derived from the
\emph{HST} flux standard stars using the Re05 methodology and nomenclature. For
comparison, Appendix A presents the traditional \emph{HST} flux calibration
methodology. According to Re05, the \emph{Spitzer} flux calibration is defined
such that a point source flux estimate \begin{equation}\langle
F_{\nu}\rangle=C'N_e=F_{\nu_o}K~,\label{fnu}\end{equation} where $\langle
F_{\nu}\rangle$ is the mean flux over the bandpass and $C'$ is the calibration
constant for a point source. $N_e$ is the number of detected photo-electrons per
second, either $N_e(pred)$ predicted from the stellar flux and the system
fractional throughput $R$ or $N_e(obs)$ observed in an infinite-radius
photometric aperture. Re05 uses a 10 pixel (12\arcsec) reference radius for the
published calibration constants; but these calibrations refer to surface
brightness (see \S A.2 and Equation (\ref{pnucp})). $F_{\nu_o}$ is the flux at
the nominal wavelength $\lambda_o=c/\nu_o$ for a $\nu F_\nu=constant$ flux
spectrum. $N_e(pred)$ is
\begin{equation}N_e(pred)=A\int{F_\nu\over{h\nu}}~R~d\nu={A\over hc}
\int{F_\lambda~\lambda~R~d\lambda}~, \label{ne}\end{equation} where
A=4869~cm$^{-2}$ is the collecting area of the \emph{Spitzer} 85cm primary
mirror with its 14.2\% obscuration (Werner 2004).
$K$ is the color correction
\begin{equation}K={\int (F_\nu/F_{\nu_o})(\nu/\nu_o)^{-1}~R~d\nu\over
{\int(\nu/\nu_o)^{-2}~R~d\nu}}~,\label{k}\end{equation} where $F_\nu$ is the
actual stellar spectral flux distribution. The nominal wavelength is
\begin{equation}\lambda_o={\int\lambda\nu^{-1}~R~d\nu\over{\int\nu^{-1}~R~d\nu}}
={\int R~d\lambda\over{\int\lambda^{-1}~R~d\lambda}}~,
\label{lamo}\end{equation} where the term after the first equal sign is from
Re05 and the term after the second equal sign is the equivalent formulation of
Hora et~al. (2008).
If the SEDs of the \emph{HST} stars in Table~1 are used for the flux $F_\nu$ to
produce a new calibration constant $C'_{ST}$ and corresponding mean flux
${\langle{F_{\nu}}^{ST}\rangle}$, then the ratio of the new to the original
calibration is \begin{equation}{\langle {F_{\nu}^{ST}}\rangle\over \langle
F_{\nu}\rangle}= {C'_{ST}\over C'}= {\int F_\nu/\nu~R~d\nu \over \nu_o~\langle
F_{\nu}\rangle \int \nu^{-2}~R~d\nu}\label{rat}\end{equation} or,
equivalently, in terms of integrals over wavelength
\begin{equation}{C'_{ST}\over C'}= {\lambda_o \int F_\lambda~\lambda ~R~d\lambda
\over c~\langle F_{\nu}\rangle \int R~d\lambda}~, \label{ratl}\end{equation}
where $\langle F_{\nu}\rangle$ is the stellar flux derived with aperture
photometry from the \emph{Spitzer} images, as calibrated with the official
calibration constants that appear in the data-file headers. Re05
quotes an uncertainty of 2\% in the IRAC absolute flux calibration. Tests of the
numerical integrations over $\nu$ per Equation (\ref{rat}) or over $\lambda$ per
Equation (\ref{ratl}) are the same to a few parts in $10^5$.
\subsection{Simplified Concept of a Point Source Calibration}
A specific-intensity calibration for the surface brightness (see \S A.2) of
diffuse sources requires a measure of the total response to a point source in an
infinite aperture, as specified above for $N_e$. However, a flux calibration
that is strictly for point sources has no such requirement and can be explained
simply and elegantly in the case of a stable instrumental configuration with a
linear response. Stability means that repeated observations produce the same
response, measured in terms of say a background-subtracted net count rate N,
while linearity implies that the count rate is directly proportional to the
physical flux F, i.e. the ratio of flux to count rate will be the same ratio of
F/N over the dynamic range of the system. There is no restriction on the
entrance slit or extraction aperture as long as the same choice is made for both
stars and the extracted count rate is repeatable for both stars. The measured
count rate can be in a certain radius aperture for point source photometry or of
a certain height for a resolution element of a spectrophotometer. If one star is
a flux standard with known flux, then the ratio F/N defines a point source
calibration constant P, so that the second star with unknown flux has the same
constant measured ratio; and the unknown flux is simply F=PN. This constant P
might be alluded to as a sensitivity but is really more properly an inverse
sensitivity, because a more sensitive instrument will have a higher count rate
for a source of the same flux.
The main complication of this concept is due to the different spectral
resolutions of the flux standard and the unknown star. A common example is a
standard star with a tabulated medium resolution SED. For broadband photometry,
the average flux of the standard over the bandpass must be calculated as in
Equation (\ref{favl}) or (\ref{favnu}), which is straightforward if the spectral
resolution is much better than the band width. In the case of a spectrometer
calibration with a resolution that is lower than the tabulated resolution of the
standard, the calibration P is defined simply as the known SED, F, binned to the
bandpass of the instrument to be calibrated divided by the response spectrum N
for the same standard star. More properly, P as a function of wavelength is
defined as the convolution of the known SED, F, with the instrumental
line-spread function, LSF, divided by the count rate spectrum, N, of the
standard convolved with the LSF of the standard star spectrum, which brings the
numerator and denominator spectra of P to the same resolution and enables a
pixel-by-pixel division of F by N after resampling to the same wavelength scale.
This procedure may fail for the case where a low resolution standard star SED
must be bootstrapped to a calibration of a much higher resolution spectrometer
where the sensitivity of the high resolution data changes significantly over the
resolution element of the known SED. For example, a single echelle order may
have a variation in sensitivity by a factor of 10 or more over a wavelength
range covered by only one or a few resolution elements of the flux standard.
\section{Spitzer Calibration}
\emph{Spitzer} observations of the sources were taken either as part of a cycle
5 Director's Discretionary Time program (PI: Gordon) or from existing archival
observations. The stars were observed in the four IRAC bands, and as
many as possible were observed in the IRS blue peakup band or MIPS 24~\micron\
band. The main goal of the IRS blue peakup and MIPS 24~\micron\ observations is
to check for dust emission (e.g., a debris disk) or faint red companions.
\subsection{Data Reduction}
All the IRAC and IRS blue peakup data (reduction version S18.7.0) were
downloaded from the \emph{Spitzer} archive. The photometry is measured using a 3
pixel radius aperture and sky annulus with radii of 10 and 20 pixels on each
individual image. As our sources are faint, refined positions are determined by
centroiding on the star in each observation mosaic image. An aperture of three
pixels radius is chosen in order to minimize noise from the sky and
contamination from other sources in our relatively crowded fields. To get the
total stellar flux, $\langle F_{\nu}\rangle$, our 3 pixel radius aperture
photometry is corrected to the standard calibration aperture sizes per Table~2.
Table~2 includes the nominal wavelengths computed from Equation (\ref{lamo}),
the size of the standard reference aperture for each filter (i.e. 10 pixels for
IRAC and infinite for IRSB and MIPS), the aperture correction needed to convert
our three pixel photometry to the reference aperture size, and the references
for the aperture corrections and instrumental-throughput spectral-response R vs.
vacuum wavelengths. Our aperture correction values for the IRAC photometry of
HD165459 agree with the tabulated values from Hora et~al. (2008) to better than
0.25\%, even though the Hora background annulus is 10--20 pixels instead of the
12--20 pixels used by Re05 for the standard 10 pixel photometry. The IRAC
aperture correction from 10 pixels to infinity is discussed in \S A.2. Our
nominal vacuum wavelengths $\lambda_o$ are within 0.4\% of those in Re05 and
within 0.2\% of Hora et~al. (2008).
\begin{figure}
\epsscale{1.2}
\plotone{fig1.eps}
\caption{The IRAC4 photometry for G191B2B for the 1st (circles) and
2nd (squares) AOR. The filled symbols gives those
measurements that were used in computing the averages (dashed lines)
and standard deviations (dotted lines). The open symbols are the
measurements that were iteratively sigma-clipped because of contamination
by cosmic ray hits.
\label{fig1}}
\end{figure}
For G191B2B, a nearby bright star produces an artifact in its sky region; and
the pixels affected are rejected prior to determining the sky flux. For each
independent observation (Astronomical Observation Request, AOR), the photometry
from the multiple image frames is averaged after sigma-clipping rejection of
outlying points. For example, Figure~\ref{fig1} illustrates
the rejected and accepted IRAC4 photometry points for the two AORs for G191B2B.
The weighted average IRAC and IRS fluxes for each star and band are given in
Table~3 after multiplication of our three pixel radius photometry by the
aperture corrections in Table~2. In order to achieve robust results, IRAC
observations with a signal-to-noise ratio (S/N) of less than 7 or with a
location more than 25 pixels from the center are rejected and do not contribute
to the averages in Table~3. The restriction to centrally located sources avoids
any confusion due to possible errors in the flat fielding procedure. The second
line for each star in Table~3 is the synthetic photometry predictions computed
from our standard star SEDs per Equation (\ref{favnu}).
\subsubsection{Uncertainties}
\begin{figure}
\epsscale{1.2}
\plotone{fig2.eps}
\caption{
Observations of two Re05 primary stars in the four IRAC bands, where
each point is the result from one AOR. Each set of points is offset by 0.1 along
the Y-axis from the set below. The points displayed all have a S/N of at
least 7 and are within 25 pixels of image center in each IRAC channel. The
red points are for HD165459, while 1812095 is shown in green. Each of the eight
panels is labeled with the number of observations (AORs), the IRAC band, the
star name, and the rms of the points shown.
\label{fig2}}
\end{figure}
\begin{figure}
\epsscale{1.15}
\plotone{fig3.eps}
\caption{
The continuous lines are SEDs typical of our three spectral categories of
standard stars scaled by $\lambda^4$ on the left axis and by $\lambda^2$ on the
right axis with $\lambda$ in \micron. The absolute fluxes are multiplied by a
factor of two for G191B2B and 1812095. The statistical error bars of
$\pm1\sigma$ are shown for the measured Spitzer photometry at the nominal
wavelengths and at the effective wavelengths. For G191B2B, a Tlusty LTE model
(dash) and a NLTE 60,000K model (dots) with solar CNO (Gianninas 2010) are shown
in addition to the standard NLTE model (solid). \label{fig3}}
\end{figure}
Many ($\sim$100) AORs exist for the four primary Re05 standard stars. For
example, Figure~\ref{fig2} shows the observations of two of our stars,
HD165459 and 1812095, that are also Re05 primary standards. Each point in
Figure~\ref{fig2} represents the sigma-clipped average of multiple image
frames in one AOR. The rms indicated on each panel is the scatter among the
remaining AOR observations after rejecting points that deviate by more than
3$\sigma$ from the average. The brighter star HD165459 has 2, 5, 4, and 1
rejected deviant AORs with an rms for the remaining points of 0.6, 0.7, 0.6, and
0.8\% for channels 1--4, respectively, while the comparable rms dispersions from
Re05 are 1.7, 0.9, 0.9, and 0.5\%. In order to compute the weight for each AOR
included in the average fluxes of Table 3, our repeatabilities for HD165459 are
added in quadrature with the statistical uncertainty for each independent
observation. Figure~\ref{fig3} compares the Spitzer broadband fluxes to
our absolute flux distributions for one example of each of our three stellar
classes. The differences between the flux levels for the nominal and effective
wavelengths quantify the ambiguity associated with assigning monochromatic
wavelengths to the broadband \emph{Spitzer} photometry.
\subsubsection{Comparison with the Re05 Photometry}
Re05 based the final, recommended IRAC calibration on four primary A-star
standards, two of which, HD165459 and 1812095, are among our standard stars.
Table 1 of Re05 contains the photometric fluxes for these two stars along with
their other two primaries, HD180609 and BD+60$^{\circ}$1753. On average, our
extracted IRAC photometric fluxes are 2.5, 2.8, 2.4, and 1.1\% higher than
the corresponding Re05 tabulations for channels 1--4, respectively.
Our IRAC photometry from each image is corrected for distortions and pixel phase
as recommended by Hora et al. (2008). These corrections differ from Re05, who
used a preliminary version of the Hora et al. work. Because of these different
data reduction procedures, our corrected photometry is expected to be
systematically brighter than Re05 by 1.1, 1.4, 0.6, and 0.5\%, for bands 1--4,
respectively. These expected systematic difference between Re05 and our
photometry account for around half of the actual differences. The remaining
unexplained differences of up to 1.8\% for IRAC3 must be due to the changing
IRAC pipeline processing and/or a somewhat different selection of IRAC
observations to include in the average for each star.
\subsubsection{Special Cases}
There are no IRAC4 measurements for 1732526. Two additional \emph{HST} standard
G stars have IRAC observations (C26202 and SNAP-2), yet neither star has high
enough quality observations to be included in this work. C26202 is in the
CDF-S (Smith et~al. 2003) and has a cooler companion at 3--4\arcsec\ that
produces blended IRAC images and precludes accurate photometry of the separate
stars. The second G star, SNAP-2, has IRAC data but lies off-center by $\sim$50
pixels where flat fielding errors might be important.
\subsubsection{MIPS}
The raw MIPS data were downloaded from the Spitzer archive and reduced using the
MIPS Data Reduction Tool (Gordon et al. 2005). In addition, several additional
steps to remove residual instrumental signatures are used (see Engelbracht et
al. 2007 for details). Given the crowded nature of some of the fields, PSF
fitting code is required to extract the MIPS24 photometry of our sources; and
our choice is StarFinder (Diolaiti et al. 2000).
\subsubsection{Predicted Fluxes}
The measured and synthetic fluxes appear in alternate rows in Table~3. For the
three pure hydrogen WDs, the temperature and gravity of the TLusty NLTE models
are derived from fits to ground-based spectra of the Balmer lines. For the pure
He WD LDS749B, the model of Bohlin \& Koester (2008) is used. For the A stars,
Bohlin \& Cohen (2008) fit NICMOS spectrophotometry from 0.8--2.5~\micron\ and
ground-based photometry with Castelli \& Kurucz (2004, hereafter CK04) model
SEDs. Similarly for the G stars, Bohlin (2010, hereafter B10) fit STIS and
NICMOS spectrophotometry. Comparing our Table 3 synthetic fluxes with the
corresponding values in Table~6 of Re05 for the two stars in common confirms
that Re05 used stellar SEDs without modification from the Cohen CWW network. A
generous statistical uncertainty of 1\% is assigned to our Table 3 synthetic
fluxes to account for the effects of the broad band Poisson noise and
repeatability of STIS spectrophotometry (Bohlin 2002) on the fitting of models
to the \emph{HST} spectrophotometry. More details of the \emph{HST} synthetic
fluxes in Table 3 and their systematic uncertainties are discussed in \S 4.1.
\subsection{Linearity}
\begin{figure}
\epsscale{1.1}
\plotone{fig4.eps}
\caption{
Check of the linearity of the observed IRAC and IRSB photometry versus source
brightness. The
plotted ratios per the equivalent Equations (\ref{rat}) and (\ref{ratl}) are a
measure of the mean flux of the \emph{HST} SEDs divided by the measured
photometric flux per the \emph{Spitzer} pipeline calibrations. The stars are
color coded by spectral type, with black, green, and red for WDs, A~stars, and
G~stars, respectively. One sigma error bars are shown on the points that differ
from unity by more than 3$\sigma$. The key for each symbol type is at the right
side of Figure~\ref{fig5}. Notice the progressively more
compressed scales in the upper three panels. \label{fig4}}
\end{figure}
The linearity of five of the \emph{Spitzer} detector systems can be evaluated by
comparing the observed photometry to the actual stellar brightness. A more
common linearity check is a comparison of count rate vs. well depth. For the
sixth system, MIPS24, our three data points are insufficient to reach any
conclusion; see Engelbracht et~al. (2007) for details of the confirmation of
MIPS 24\micron\ linearity. Using the modeled \emph{HST} SEDs as the stellar
flux, Equation (\ref{rat}) is the ratio of predicted to measured fluxes; and
that ratio is shown in Figure~\ref{fig4} as a function of stellar
brightness for our three classes of stars. IRAC3 has the narrowest bandpass and
the lowest countrates. G191B2B is significantly high in the IRAC4 band; but that
one anomalous point is not an indication of non-linearity. However for IRSB,
even over the small dynamic range of 7, the five data points show evidence of
some issue with the photometry. The ratio for the faintest IRSB source 1732526
is 30\% high, while the brightest source P041C is 10\% low. However, Gilliland
\& Rajan (2011) discovered that P041C is a double star with an M star companion
separated by 0.57\arcsec, which could make the observed IRSB flux too high by
$\sim$10\%. These problems preclude an accurate comparison of the \em{HST}
versus \em{Spitzer} IRSB calibration.
\subsection{Dust Rings}
Many stars are encircled by a ring of cool dust that emits strongly at longer
wavelengths. These sources can be identified by small ratios of predicted to
observed flux that fall off-scale on Figure~\ref{fig4}. Our observed
excess flux at 24~\micron\ for HD165459 confirms the evidence for a disk found
by Rieke et~al. (2005), who reported a 40\% excess.
Su et~al. (2006) revised the excess to 46.7\%.
\begin{figure}
\epsscale{1.1}
\plotone{fig5.eps}
\caption{
Ratios as in Figure~\ref{fig4} of calibration constants computed from
the \emph{HST} based fluxes to the IRAC values of Re05. WDs are in the top
panel, solar analogs are in the center panel, and A~stars are at the bottom. One
sigma error bars are shown on the points that differ from unity by more than
3$\sigma$. The key to the individual symbols is at the right side of each
panel.The scales are the same for all three stellar types, but there is an
offset of 0.05 in the top panel. \label{fig5}}
\end{figure}
Our excess of a factor of 2.5 for 1740346 in the 16~\micron\ band suggests the
presence of a contaminating debris disk. While there is no evidence of dust
emission below 8~\micron\ for HD165459, the low values for 1740346 in
Figure~\ref{fig5} suggest the presence of
hotter dust than is around HD165459. Therefore,
1740346 is not included below in the comparison with the Re05 calibration of
IRAC.
\subsection{Results}
Per Equation (\ref{rat}) or (\ref{ratl}), Figure~\ref{fig5}
compares the \emph{HST} calibration constants
to those published by Re05. One $\sigma$ error bars appear on the points that
differ from unity by more than the 3$\sigma$ uncertainty of the ratio. The
problematic IRSB data are not shown; and the MIPS24 ratios for the three stars
without dust rings are insufficient to draw any firm conclusions, although all
three MIPS ratios are within $\sim$10\% of unity.
With a 4$\sigma$ statistical significance, the IRAC4 ratio of 1.12
for the observation of G191B2B is the largest deviation from unity in
Figure~\ref{fig5}. As a check on this problematic case, the
background level in the IRAC4 band was studied by constructing a histogram of
the set of all three-pixel radius photometry in a 120 pixel square, centered on
G191B2B. There are no background regions in the vicinity of the star that are
low enough to bring the observed result to the predicted level with a ratio of
unity in Figure~\ref{fig4}; and there is only $\sim$1\% probability of
a sky level that would reduce the discrepancy to 2$\sigma$.
Table 4 reports the average results separately for the WDs, for two Re05 primary
A stars, and for the group of A and G stars. These values are the weighted
averages of the ratios of the synthetic to the measure fluxes in Table 3, where
the uncertainty for each star is the measurement error in Table~3 combined with
the 1\% uncertainty assigned to the synthetic values. These uncertainties are
the statistical errors and do not include any possible systematic errors that
would affect all stars equally. As a group, the averages for the WDs are
significantly above the other averages, especially with the inclusion of the
IRAC4 ratio for the hottest WD, G191B2B, that is 12\% high in
Figure~\ref{fig5}. Individually, many of the WD results could be
explained as not significant at the $3\sigma$ level or because of the less well
vetted pure helium model for LDS749B (Bohlin \& Koester 2008).
NLTE effects in the IR become more pronounced at higher stellar temperatures
(Bohlin 2000), eg. the pure hydrogen LTE model for G191B2B shown in
Figure~\ref{fig3} falls near the IRAC4 data. For G191B2B, metal
line-blanketing is observed at the 1--2\% level in the FUV (Barstow et~al.
1999). Per Gianninas et~al. (2010), an additional NLTE model spectrum at 60,000
K and $\log g= 7.5$ with CNO metals at the Asplund et~al. (2005) solar abundance
also appears in Figure~\ref{fig3} near the LTE model. This metal
abundance is not the actual metallicity but is only a proxy for the effect of
the metals, which is to reduce the NLTE effects and cool the upper atmosphere,
where the IR continuum is formed. There is no coherent simultaneous
determination of the metal abundances along with $T_\mathrm{eff}$ and $\log g$
from the Balmer lines (Barstow et~al. 2003); however, the abundances suggested
by Barstow et~al. (2003) are considerably less than the solar Asplund et~al.
(2005) values, so that a model with proper trace metallicities for G191B2B in
the IR should fall somewhere between the 60,000K CNO model and the pure hydrogen
NLTE model, potentially within 1--2$\sigma$ of all the IRAC fluxes.
The significant deviation from unity in
Figures~\ref{fig4}--\ref{fig5} for G191B2B should be
explored with a proper model. Meanwhile, the most relevant result in Table 4 is
for the set of eight G and A stars, for which our calibration constants are
lower than Re05 for IRAC1--4 by 2.3, 1.9, 2.0, and 0.5\%, respectively. For the
four Re05 primary stars, our re-reduced photometry is 2.5, 2.8, 2.4, and 1.1\%
higher than Re05. Higher extracted photometry implies lower calibration
constants; and the measured photometry differences correspond to the calibration
constant differences to an accuracy of better than 1\%. Understanding that our
differences with Re05 are mostly due to differences in the photometry extracted
from the IRAC images helps verify that both calibrations have been properly
derived per the adopted common methodology and that our results are expected to
be lower by about the amounts computed in Table 4 for A+G stars.
While the calibration constants themselves depend directly on the accuracy of
the bandpass throughput $R$ per Equation (\ref{favl}) or (\ref{favnu}), our
methodology and that of Re05 involve the same integral of the stellar SED over
$R$, so that errors in $R$ cancel to first order in the comparison of the two
sets of results. However, bandpass errors, such as an overall shift of $R$ in
wavelength, changes the predicted $N_e(pred)$ per Equation (\ref{ne}). For
example, a shift of +0.1~\micron\ in the 2.9~\micron\ wide IRAC4 band, i.e. a
shift in wavelength by 3\% of the band width, would cause a 3.8\% decrease in
$N_e(pred)$ for the WD, A, and G stellar types. Per Equation (\ref{ne}), such a
3.8\% error would be difficult to distinguish from a simple 3.8\% compensating
error in the laboratory measurements of $R$, because all of our standard stars
have nearly the same Rayleigh-Jeans slope in the IR. A bandpass shift would only
be important in the case of strong spectral features within the bandpass or in
the case of a source with an SED much cooler than our A and G standard stars.
\section{Details of Absolute Flux Calibration}
\subsection{Details of the \emph{HST} Calibration}
The discussion of our measures of instrumental response (N) appear above in \S
2.1--2.2, while this section covers the details of the flux (F) in the
calibration constant P=F/N.
The essence of the \emph{HST} flux system is to establish standard star SEDs and
then measure other stars relative to these standard candles. Pure hydrogen WD
stars are chosen as these \emph{HST} fundamental standard candles, because their
atmospheric models are simpler than other stars where metal lines, molecular
lines, and convection add complications. The models of our pure-hydrogen primary
WD standards are specified by two parameters $T_\mathrm{eff}$ and $\log g$, both
of which are defined by the Balmer line profiles. To establish absolute flux
standards, the unreddened WD stars G191B2B, GD71, and GD153 are observed with
STIS, which also observed Vega in the same modes with the same dispersion.
Because STIS is precisely linear even into the regime of many times
over-saturated in its CCD (Gilliland et~al. 1999, Bohlin \& Gilliland 2004), the
relative flux between Vega and each WD is measured (Bohlin \& Gilliland 2004);
and the well-known absolute monochromatic flux at 5556\AA\ for Vega establishes
the absolute flux of the model SED for each WD. Minor complications arise due to
the slowly changing STIS sensitivity with time and the gradual loss of charge
transfer efficiency in the CCD. These corrections are tracked to better than
1\%, and the STIS response is corrected for these effects as a function of
wavelength.
Observations of the three fundamental primary WDs establish the instrumental
flux calibration. For example, a set of observations in the low dispersion modes
of STIS below 1~\micron\ and from 0.8--2.5~\micron\ in the grism modes of
NICMOS establish the flux calibrations and enables the creation of secondary
flux standards. Bohlin \& Koester (2008) demonstrated that such STIS and NICMOS
spectrophotometry of LDS 749B, a pure helium star, could be modeled to the
statistical precision of the data. NICMOS spectrophotometry, supplemented by
ground-based photometry for the A stars used in this paper was modeled by Bohlin
\& Cohen (2008) with $T_\mathrm{eff}$, $\log g$, $\log z$, and color excess
$E(B-V)$ as free parameters. Similarly, B10 modeled STIS and NICMOS fluxes for
the G stars discussed here. The BVR photometry used to define the A-star models
is from Mount Hopkins Observatory (Cohen et~al. 2003) and is referenced to the
9400K Vega model and to Vega photometric values from Maiz Apell\'aniz (2007).
\subsubsection{Uncertainties}
Establishing proper uncertainties is a complex process, involving the division
into categories of systematic and statistical errors, which can each be divided
into sub-categories of absolute level and relative slope of flux vs. wavelength
(i.e. ``color''). In many cases, the statistical scatter can be reduced below
the systematic error bars by repeated observations or by binning spectra.
Systematic uncertainty of the \emph{HST} WD flux scale is estimated as $\sim$1\%
by B10 for the ratio of the flux at 5556\AA\ to fluxes in the 1--2.5 \micron\
range. The total systematic uncertainty in our IR flux scale should be less than
2\%, even when the 0.7\% uncertainty in the absolute 5556\AA\ flux (Megessier
1995) is included. Considerable confidence in the estimate of precision in the
slope of the relative flux with wavelength can be gained by examining the
internal agreement among the \emph{HST} standards: i) Bohlin (2007a)
illustrates the $<<$1\% agreement of the three primary WD models with their
calibrated fluxes, which means that if the slope of any model, which depends
primarily on $T_\mathrm{eff}$, is in error, then the other two model
temperatures must be similarly in error in order to make the same change in flux
vs. wavelength for all three stars from 0.12--2.5~\micron. ii) The NICMOS
photometric calibration is based on the \emph{HST} primary standard G191B2B and
the secondary standard P330E per de Jong (2006), where the consistency of the
absolute spectrophotometry is confirmed for these two SEDs from the
CALSPEC\footnote{http://www.stsci.edu/hst/observatory/cdbs/calspec.html/}
database. In particular, de Jong's Figure 2 shows an agreement of 0.8--1.4\%
between the NICMOS calibrations derived from the SEDs for P330E and G191B2B over
the combined 0.9--2.4 \micron\ range of the three cameras. Finally, the
extrapolation of the \emph{HST} NICMOS fluxes from 2.2~\micron\ to the first
IRAC band at 3.6~\micron\ is rather independent of the particular stellar model.
For example, among the seven A stars in Table 1, Bohlin \& Cohen (2008)
demonstrate that all the measured NICMOS SEDs fit their model SEDs over the
0.8--2.4 \micron\ range within 1\% in broad wavelength bins; and the maximum
difference in the 2.2/3.6~\micron\ ratio is 0.7\% between the minimum
$T_\mathrm{eff}$=7650K model for 1743045 and the maximum 9100K model for
1802271.
\subsection{Details of the Re05 Calibration}
As discussed above, our measured IRAC instrumental responses are smaller than
published by Re05 for the same stars. However, Re05 used data from the IRAC
pipeline processing version S10, while version S18.7.0 is utilized for this
paper. To be directly comparable, any two independent calibrations should use
the same pipeline processing version, as well as the same algorithms for
extracting the point source fluxes from the images.
As discussed above and in the Appendix, the equations that define our new
calibrations are exactly equivalent to those of Re05, so that differences arise
only from different measured instrumental photometry or from differences in the
adopted stellar fluxes.
The stellar fluxes used by Re05 are from the CWW grid, while
the revised SEDs from Bohlin \& Cohen (2008) are used here. Some major
differences in the derivation of model SEDs are that Bohlin \& Cohen fitted
models to NICMOS spectrophotometry in the 0.8-2.4~\micron\ range rather than to
the 2MASS plus I band photometry used for determining the CWW SEDs. The
CK04 model grid was used to fit the data;
and rather than specifying the model $T_\mathrm{eff}$ and $\log g$ from spectral
classification spectra, Bohlin \& Cohen found the best fitting model from the
grid, allowing $T_\mathrm{eff}$, $\log g$, $\log z$, and the reddening E(B-V) to
vary as free parameters. Despite these differences, the resulting SEDs differ
only slightly from the CCW fluxes used by Re05. For example over the IRAC
3--9~\micron\ range, our SEDs for the IRAC primary stars HD165469 and 1812095
agree with the CWW SEDs used by Re05 to 0.5 and 1--2\%, respectively.
\subsection{Details of the Ri08 Calibration}
Ri08 start by determining a best calibrated value for an equivalent Vega
photospheric flux at 10.6 \micron\ and use a Vega theoretical model normalized
to this value to predict the photospheric flux density at 2.22 \micron. This
prediction is robust against different models for A0 star photospheres. After
correction for the small contribution for the extended debris ring found in
interferometry, the resulting prediction is compared with measurements of the
absolute flux density from Vega near 2.22 \micron. The agreement is good; and
the two approaches are averaged to generate a 'best' A-star-based calibration at
2.22 \micron. Ri08 independently generated a calibrated solar spectrum as a
combination of the measurements of Thuillier et al. (2003) out to 2.4 \micron\
and an Engelke function from 2.4 to 12 \micron\ (over which range the Engelke
function gives a good fit to a number of accurate calibrated solar
measurements). Ri08 use this spectrum to compare with the observed K-[8] color
of solar-type stars and extrapolated the result to 10.6 \micron\ for an
independent test of the beginning calibration at this wavelength. Their
calibration makes no reference to the visible calibration, although the
predicted color of Vega based on the absolute calibrations at V and K is
consistent with the observed color of the star to within about 2\%. The V-K
colors of solar-type stars are also consistent roughly within this error with
their calibration at 2.22 \micron\ using the Thuillier measurements to translate
to V. The Ri08 results should be understood as a purely infrared-based
calibration with an estimated accuracy in this region of 2\%.
The Ri08 calibration suggests that the IRAC4 8~\micron\ fluxes should be 1.5
$\pm$2\% higher, while our fluxes are $\sim$0.5\% lower than Re05. Thus, our
results agree with Ri08 within the uncertainties. A new extraction of the IRAC
photometry for the 32 Ri08 stars compared with the Ri08 tabulation of predicted
fluxes produces a mean predicted-to-observed ratio that is 2.1~$\pm$0.5\% above
unity, i.e. confirming the Ri08 difference of 1.5\% within the uncertainties.
\subsection{Comparison of Methodologies}
The following points compare the Ri08 technique to the similar \emph{HST} and
CWW methods for establishing standard star SEDs.
i) Ri08 establish two independent IR absolute flux determinations for A stars
(Vega) and for G stars (the Sun), for which agreement provides a
powerful confirmation of the results. Both the \emph{HST} and CWW
absolute levels are traceable only to the visible range for Vega.
ii) In the Ri08 method, a large number of flux standards are more easily and
cheaply established from existing accurate photometric systems, so that
statistical uncertainties can be reduced well below systematic
uncertainties. The other methods are more observationally expensive,
especially in the case of the \emph{HST} flux measurements.
iii) Because of the similarity of SEDs at long wavelengths, errors between
widely separated IR wavelengths are minimal. Ri08 normalized the adopted
SEDs to IR absolute flux measurements, so that errors in the IR are
minimized.
iv) The shape of stellar SEDs short of 1.5~\micron\ becomes increasingly dependent
on the stellar temperature and on the metallicity.
Thus, the uncertainty of the flux calibration below 1.5~\micron\ grows
with decreasing wavelength making the Ri08 technique less useful for
calibration in the short wavelength (0.6-1.5~\micron) range of JWST.
\section{Recommendations}
Several deficiencies in our prototype set of flux standards must be addressed
before \emph{JWST} can be calibrated to our goal of 1\% precision at
0.6--30~\micron. Required improvements include i) better stellar
atmosphere grids and a better understanding of their uncertainties, ii) updated
and expanded lists of potential standard stars in each of our three spectral
type classes, and iii) resolution of the proper normalization of these models to
measures of absolute stellar fluxes in the visible and IR.
\subsection{Model Atmosphere Grids}
Perhaps, the best way to assess systematic errors within a set of similar type
stars is to compare independent sets of models to the measured flux
distributions, which B10 did for the G stars by fitting both MARCS models
(Gustafsson et al. 2008) and those of CK04. In broad continuum bins,
both sets of models agree with the \emph{HST} fluxes to $\sim$0.5\%.
However, both of these sets have serious deficiencies. The MARCS grid does not
extend past 20~\micron, while the CK04 grid has rather coarse wavelength spacing
with a total of 1221 points and only one point between 10 and 40~$\mu$m. A third
independent model grid, an update of the CK04 models with good wavelength
resolution, and an extension of the MARCS grid beyond its current 20~\micron\
limit would be ideal. Shortward of 20~\micron, the best-fit CK04 models agree
with the best-fit MARCS models to 1\% for the G stars in broad continuum bands.
Combining this 1\% with a 1\% systematic uncertainty in the \emph{HST} WD flux
scale results in the B10 estimate of a possible systematic 2\% uncertainty of
the broadband IR fluxes of individual G stars with respect to the V band.
For the A stars, a second independent model grid is needed, because the MARCS
models are limited to maximum temperatures of 8000K. Both the original CWW SEDs
and the newer Bohlin \& Cohen (2008) SEDs are extrapolations of the measured
fluxes into the IR using models that are based on computer code traceable to R.
Kurucz. The best check on these IR fluxes of the A stars was from the two models
provided by T. Lanz at 9400~K and 8020~K. While the agreement of the Lanz SED
with CK04 is within 2\% for the 9400~K model, the Lanz SED is brighter than CK04
by 4\% at 10~\micron\ for the 8020~K model, which raises the question of the
accuracy of model SEDs, in general. Because of the scatter among the WDs and
because only two G stars are included, our results are based
mainly on A stars. Furthermore, the final Re05 IRAC calibration is based
entirely on A stars, so that more work on independent A star grids is
essential.
\subsection{Sample of Stars}
Our goal of comparing results among different stellar types is compromised by
the large scatter and offset of the average results for the WDs and by the poor
significance of the G-star result that is based on only two stars. Perhaps,
warm-mission IRAC1 and IRAC2 observations of our four WD stars could illuminate
the reality of their large scatter in the top panel of
Figure~\ref{fig5}. Among the A stars,
Figure~\ref{fig5} shows the most scatter at 8~\micron, where
1740346 is excluded because of evidence for dust at 16~\micron. However, 1805292
has a 3$\sigma$ deviation without a strong indication for dust. Thus, more A
stars would reduce uncertainties, especially with respect to the pure A-star
calibration of Re05. The remaining two primary and primary-candidate A star
standards of Re05 should be observed with STIS and modeled to achieve a more
complete and precise comparison between the \emph{HST} and Re05 A-star results.
In order to confirm the A star comparison with IRAC, more G stars and more WD
stars with good existing cold mission \emph{Spitzer} observations are needed.
Plenty of observations of brighter stars with both IRAC and MIPS 24\micron\ exist
in the Spitzer archives, except for WDs. Thus, the WD category will be
supplemented by late O or B type stars. This new set of standards is also
needed to cover the large dynamic range and spectral variety required to
calibrate the full suite of JWST instrumentation but will require new STIS
spectrophotometry to provide the link to \emph{HST}. ACS photometry, WFC3
photometry, and WFC3 grism spectrophotometry will also help in firmly
establishing the tie to the \emph{HST} flux scale, especially for the WFC3 data
at wavelengths longward of the STIS cutoff at 1\micron.
\subsection{Absolute Flux Level}
\begin{figure}
\epsscale{1.1}
\plotone{fig6.eps}
\caption{
Ratio to the 9400~K model $continuum$ level from 0.32--30~\micron~in three
segments for the STIS measured fluxes (black), for the Kurucz
9400~K model (green), for the Kurucz 9550~K model (red), and for the Ri08 fluxes
(blue). In the top two panels, the STIS spectrophotometry (black line) is mostly
hidden under the green line. The three monochromatic, absolute IR-flux values of
Ri08 appear on the blue Ri08 SED as filled circles with error bars. The
Megessier flux value at 0.5556~$\mu$m is at unity, where the 0.7\% error bar is
inside the circle. The inset is a blowup of this 5556~\AA\ region that shows
this point along with the robust set of direct absolute flux measurements
summarized by Hayes (1985). The five smaller filled circles are labeled per the
Hayes nomenclature and are averaged to get his estimate of
$3.44\times10^{-9}$~erg cm$^{-2}$ s$^{-1}$~\AA$^{-1}$ at 5556~\AA\ with its
1.5\% error bar that is labeled as Hayes in the inset graph. Also in the inset,
the nine small open circles connected with a line are the Hayes
spectrophotometry on 25~\AA\ centers from 5450--5650~\AA. \label{fig6}}
\end{figure}
While the absolute-flux zero-point of all \emph{HST} standards is tied to the
monochromatic value of flux for Vega of $3.46\times10^{-9}$~erg cm$^{-2}$
s$^{-1}$~\AA$^{-1}$ at 5556~\AA\ (Megessier 1995), there are valid absolute
measures in the IR, as summarized by Ri08. Because Ri08 discuss and present a
SED for Vega, the differences between our preferred Vega SED and the Ri08 SED
are investigated as a possible explanation for the differences between the
\emph{HST} and Ri08 flux calibrations for IRAC. The SED of Vega has been
measured by STIS at 0.17--1~$\mu$m by Bohlin \& Gilliland (2004), who suggested
that the Kurucz $T_{\rm eff}=9550$~K model\footnote{http://kurucz.harvard.edu/}
fits the STIS observations, while Bohlin (2007a) discovered that a change in the
STIS non-linearity correction made a Kurucz 9400~K model$^{4}$ fit the observed
flux much better both below the Balmer jump and in the 0.7--0.8~$\mu$m region,
as illustrated in Figure~\ref{fig6} with red for the 9550~K and green
for the 9400~K models. These two models and the reference continuum level are
all normalized to $3.46\times10^{-9}$~erg cm$^{-2}$ s$^{-1}$~\AA$^{-1}$ at
5556~\AA\ (Megessier 1995). The blue line is the same 9550~K model but is
normalized to the Ri08 IR value of 645Jy $\pm$2.3\% at 2.22~$\mu$m. The Ri08
fluxes below 1~\micron\ are not expected to match the actual stellar flux and
are not shown. The division of all the SEDs by the same smooth theoretical model
continuum in Figure~\ref{fig6} illustrates the differences among the
various flux distributions and also shows where the (mostly hydrogen) line
blanketing complicates the comparisons. The STIS measurements (black line in
Figure~\ref{fig6}) and the 9400~K model (green line) agree to $\sim$1\%
in the unblanketed regions from below the Balmer jump to $\sim$1~$\mu$m.
Fortunately, the discrepancy of 1.1\% at 2.22~$\mu$m between the \emph{HST}
based fluxes and Ri08 is almost negligible and could be reduced by using the
9400~K model with a weighted average of the visible and IR normalizations.
Another complication is that Vega is a pole-on rapid rotator (Peterson et~al.
2006) with temperature zones in the 7900--10150~K range (Aufdenberg et~al.
2006). Even though the line blanketing in the Aufdenberg et~al.
multi-temperature model is occasionally stronger than measured, the continuum
levels of his computed pole-on SED (Aufdenberg, private comm.) track the ratio
of the 5556~\AA\ to the 2.22~\micron\ absolute fluxes to 0.4\%, i.e. much
better than any uncertainties in the absolute flux measurements in the visible
or IR. In summary, there is no reason to question the accuracy of either the
5556~\AA\ flux of Megessier (1995) or the 2.22~\micron\ flux quoted by Ri08.
A robust and straightforward comparison of a model SED for Vega with the
measured absolute visible and IR fluxes is not possible because of the IR
emission from the dust ring and because Vega is a pole-on rapid rotator. As
pointed out by Cohen et~al. (1992b), Sirius is a much better primary IR standard
because of its low rotation speed and lack of a contaminating dust ring. Thus,
one big step forward in resolving any possible offset between visible and IR
absolute fluxes is to observe Sirius with STIS and fit a model, in order to
directly compare the model with IR absolute fluxes, eg. from MSX (Price et~al.
2004), who have $\sim$1\% measurement accuracy in six bands from 4.3 to
21.3~\micron. STIS spectrophotometry of Sirius will be somewhat more saturated
than for Vega; but the same techniques of Bohlin \& Gilliland (2004) for
precisely analyzing saturated data should apply.
On longer time scales, there are on-going programs to measure stellar
spectrophotometric fluxes relative to NIST laboratory standards. The ACCESS
rocket program will establish a few fundamental flux standards in the brightness
range of Sirius to V$\sim$9.5 (Kaiser et~al. 2007). A ground-based program, NIST
STARS, is supported by lidar to measure the real-time atmospheric extinction and
uses NIST calibrated detectors to establish an all-sky set of SEDs for
standard stars with precisions of 0.5\% (Zimmer et~al. 2010).
\section{Summary}
The IRAC calibration constants have been determined from a somewhat different
set of flux standards than utilized by Re05. Small differences between our
results and those of Re05 are explained by differences in the photometry
extracted from the IRAC images. Our results are in agreement with the
independent IRAC4 8~\micron\ calibration of Ri08.
The robustness of our results suffer from a deficiency of G-star flux standards,
from the exclusion of two out of the four Re05 primary flux standards, and from
offsets of up to 4$\sigma$ between our set of WD stars and the cooler standards.
To alleviate these deficiencies, more comparison standards with \emph{HST} fluxes
and better model grids are required to model the measured fluxes and establish
SEDs to the JWST limit of 30~\micron.
\newpage
|
\section{Introduction}
The interpretation presented here completely revives classical ontology: Reality is described completely by a classical trajectory $q(t)$, in the classical configuration space $Q$.
The wave function is also interpreted in a completely classical way -- as a particular description of a classical probability flow, defined by the probability distribution $\rho(q)$ and average velocities $v^i(q)$. The formula which defines this connection has been proposed by Madelung 1926 \cite{Madelung} and is therefore as old as quantum theory itself. It is the polar decomposition
\begin{equation}\label{polar}
\psi(q)=\sqrt{\rho(q)} e^{\frac{i}{\hbar} S(q)},
\end{equation}
with the phase $S(q)$ being a potential of the velocity $v^i(q)$:
\begin{equation}\label{guiding}
v^i(q) = m^{ij} \pd_j S(q),
\end{equation}
so that the flow is a potential one.\footnote{Here, $m^{ij}$ denotes a quadratic form -- a ``mass matrix'' -- on configuration space. I assume in this paper that the Hamiltonian is quadratic in the momentum variables, thus,
\begin{equation}\label{HamiltonFunction}
H(p,q) = \frac{1}{2}m^{ij}p_ip_j + V(q).
\end{equation}
This is quite sufficient for relativistic field theory, see app. \ref{relativity}.
}
This puts the interpretation into the classical realist tradition of interpretation of quantum theory, the tradition of de Broglie-Bohm (dBB) theory \cite{deBroglie}, \cite{Bohm} and Nelsonian stochastics \cite{Nelson}.
But there is a difference -- the probability flow is interpreted as describing incomplete information about the true trajectory $q(t)$. This puts the interpretation into another tradition -- the interpretation of probability theory as the logic of plausible reasoning in situations with incomplete information, as proposed by Jaynes \cite{Jaynes}. This objective variant of the Bayesian interpretation of probability follows the classical tradition of Laplace \cite{Laplace} (to be distinguished from the subjective variant proposed by de Finetti \cite{deFinetti}).
So this interpretation combines two classical traditions -- classical realism about trajectories and the classical interpretation of probability as the logic of plausible reasoning.
There is also some aspect of novelty -- a clear and certain program for development of subquantum theory. Quantum theory makes sense only as an approximation for potential probability flows. It has to be generalized to non-potential flows, described by the flow variables $\rho(q), v^i(q)$. Without a potential $S(q)$ there will be also no wave function $\psi(q)$ in such a subquantum theory. And the flow variables are not fundamental fields themself, they also describe only incomplete information. The fundamental theory has to be one for the classical trajectory $q(t)$ alone.
So this interpretation is also a step toward the development of a subquantum theory. This is an aspect which invalidates prejudices against quantum interpretations as pure philosphy leading nowhere. Nonetheless even this program for subquantum theory has also classical character, making subquantum theory closer to a classical theory.
Thinking about how to name this interpretation of quantum theory, I have played around with ``neoclassical'', but rejected it, for a simple reason: There is nothing ``neo'' in it. Instead, it deserves to be named ``paleo''.\footnote{Association with the ``paleolibertarian'' political direction is welcome and not misleading -- it also revives classical libertarian ideas considered to be out of date for a long time.}
And so I have decided to name this interpretation of quantum theory ``paleoclassical''.
\subsection{The Bayesian character of the paleoclassical interpretation}
The interpretation of the wave function is essentially Bayesian.
It is the objective (information-dependent) variant of Bayesian probability, as proposed by Jaynes \cite{Jaynes}, which is used here. It has to be distinguished from the subjective variant proposed by de Finetti \cite{deFinetti}, which is embraced by the Bayesian interpretation of quantum theory proposed by Caves, Fuchs and Schack \cite{Bayesian}.
The Bayesian interpretation of the wave function is in conflict with the objective character assigned to the wave function in dBB theory. Especially Bell has emphasized that the wave function has to be understood as real object.
But the arguments in favour of the reality of the wave function, even if strong, appear insufficient: The effective wave function of small subsystems depends on the configuration of the environment. This dependence is sufficient to explain everything which makes the wave function similar to a really existing object. It is only the wave function of a closed system, like the whole universe, which is completely interpreted in terms of incomplete information about the system itself.
Complexity and time dependence, even if they are typical properties of real things, are characteristics of incomplete information as well.
The Bayesian aspects essentially change the character of the interpretation: The dBB ``guiding equation'' \eqref{guiding}, instead of guiding the configuration, becomes part of the definition of the information about the configuration contained in the wave function.
This has the useful consequence that there is no longer any ``action without reaction'' asymmetry: While it is completely natural that the real configuration does not have an influence on the information available about it, there is also no longer any ``guiding'' of the configuration by the wave function.
\subsection{The realistic character of the paleoclassical interpretation}
On the other hand, there is also a strong classical realistic aspect of the paleoclassical interpretation: First of all, it is a realistic interpretation. But, even more, its ontology is completely classical -- the classical trajectory $q(t)$ is the only beable.
This defines an important difference between the paleoclassical interpretation and other approaches to interprete the wave function in terms of information: It answers the question ``information about what'' by explicitly defining the ``what'' -- the classical trajectory -- and even the ``how'' -- by giving explicit formulas for the probability distribution $\rho(q)$ and the average velocity $v^i(q)$.
The realistic character leads also to another important difference -- that of motivation. For the paleoclassical interpretation, there is no need to solve a measurement problem -- there is none already in dBB theory, and the dBB solution of this problem -- the explicit non-Schr\"{o}dinger\/ evolution of the conditional wave function of the measured system, defined by the wave function of system and measurement device and the actual trajectory of the measurement device -- can be used without modification.
And there is also no intention to save relativistic causality by getting rid of the non-local collapse -- the paleoclassical interpretation accepts a hidden preferred frame, which is yet another aspect of its paleo character. Anyway, because of Bell's theorem, there is no realistic alternative.
\footnote{Here I, of course, have in mind the precise meaning of ``realistic'' used in Bell's theorem, instead of the metaphorical one used in ``many worlds'', which is, in my opinion, not even a well-defined interpretation.}
\subsection{The justification}
So nor the realistic approach of dBB theory, which remains influenced by the frequentist interpretation of probability (an invention of the positivist Richard von Mises \cite{Mises}) and therefore tends to objectivize the wave function, nor the Bayesian approach, which embraces the anti-realistic rejection of unobservables and therefore rejects the preferred frame, are sufficiently classical to see the possibility of a complete classical picture.
But it is one thing to see it, and possibly even to like it because of its simplicity, and another thing to consider it as justified.
The justification of the paleoclassical interpretation presented here is based on a reformulation of classical mechanics in term of -- a wave function. It is a simple variant of Hamilton-Jacobi theory with a density added, but this funny reformulation of classical mechanics appears to be the key for the paleoclassical interpretation. The point is that its exact equivalence to classical mechanics (in the domain of its validity) and even the very fact that it shortly becomes invalid (because of caustics) almost force us to accept -- for this classical variant -- the interpretation in terms of insufficient information. And it also provides all the details we use.
But, then, why should we change the ontological interpretation if all what changes is the equation of motion? Moreover, if (as it appears to be the case) the Schr\"{o}dinger\/ equation is simply the linear part of the classical equation, so that there is not a single new term which could invalidate the interpretation? And where one equation is the classical limit of the other one?
We present even more evidence of the conceptual equivalence between the two equations: A shared explanation, in terms of information, that there should be a global $U(1)$ phase shift symmetry, a shared explanation of the product rule for independence, a shared explanation for homogenity of the equation. And there is, of course, the shared Born probability interpretation. All these things being equal, why should one even think about giving the two wave functions a different interpretation?
\subsection{Objections}
There are a lot of objections against this interpretation to care about.
Some of them are quite irrelevant, because they are handled already appropriately from point of view of de Broglie-Bohm theory, so I have banned them into appendices: The objection of incompatibility with relativity (app. \ref{relativity}), the Pauli objection that it destroys the symmetry between configuration and momentum variables (app. \ref{Pauli}), some doubts about viability of field-theoretic variants related with overlaps (app. \ref{fields}).
There are the arguments in favour of interpreting the wave function as describing external real beables. These are quite strong but nonetheless insufficient: The wave functions of small subsystems -- and we have no access to different ones -- \emph{depend} on real beables external to the system itself, namely the trajectories of the environment. And complexity and dynamics are properties of incomplete information as well.
And there is the Wallstrom objection \cite{Wallstrom}, the in my opinion most serious one. How to justify, in terms of $\rho(q)$ and $v^i(q)$, the ``quantization condition''
\begin{equation
\oint m_{ij} v^i(q) dq^j = \oint \pd_j S(q) dq^j = 2\pi m\hbar, \qquad m\in\mbox{$\mathbb{Z}$}.
\end{equation}
for closed trajectories around zeros of the wave function, which is, in quantum theory, a trivial consequence of the wave function being uniquely defined globally, but which is completely implausible if formulated in terms of the velocity field?
Fortunately, I have found a solution of this problem in \cite{againstWallstrom}, based on an additional regularity postulate. All what has to be postulated is that \emph{$0< \Delta \rho(q) < \infty$ almost everywhere where $\rho(q)=0$}. This gives the necessary quantization condition. Moreover, there are sufficiently strong arguments that it can be justified by a subquantum theory.
The idea that what we observe are localized wave packets instead of the configurations themself I reject in app. \ref{wavepackets}. So all the objections I know about can be answered in a satisfactory way. Or at least I think so.
\subsection{Directions of future research}
Different from many other interpretations of quantum theory, the paleoclassical interpretation suggests a quite definite program of development of a more fundamental, subquantum theory: It defines the ontology of subquantum theory as well as the equation which can hold only approximately -- the potentiality condition for the velocity $v^i(q)$. The consideration of the Wallstrom objection even identifies the domain where modifications of quantum theory are necessary -- the environment of the zeros of the wave function.
One can identify also another domain where quantum predictions will fail in subquantum theory -- the reliability of quantum computers, in particular their ability to reach exponential speedup in comparison with classical computers.
Another interesting question is what restrictions follow for $\rho(q), v^i(q)$ from the interpretation as a probability flow for a more fundamental theory for the trajectory $q(t)$ alone. An answer may be interesting for finding answers to the ``why the quantum'' question. A question which cannot be answered by an interpretation, which is restricted to the ``what is the quantum'' question.
\section{A wave function variant of Hamilton-Jacobi theory}
Do you know that one can reformulate classical theory in terms of a wave function? With an equation for this wave function which is completely classical, but, nonetheless, quite close to the Schr\"{o}dinger\/ equation?
In fact, this is a rather trivial consequence of the mathematics of Hamilton-Jacobi theory and the insights of Madelung \cite{Madelung}, de Broglie \cite{deBroglie}, and Bohm \cite{Bohm}. All one has to do is to look at them from another point of view. Their aim was to understand quantum theory, by representing quantum theory in a known, more comprehensible, classical form -- a form remembering the classical Hamilton-Jacobi equation
\begin{equation}\label{HamiltonJacobi}
\pd_t S(q) + \frac12 m^{ij} \pd_i S(q) \pd_j S(q) + V(q) = 0.
\end{equation}
So, the real part of the Schr\"{o}dinger\/ equation, divided by the wave function, gives
\begin{equation}\label{Bohm}
\pd_t S(q) + \frac12 m^{ij} \pd_i S(q) \pd_j S(q) + V(q) + Q[\rho] = 0,
\end{equation}
with only one additional term -- the quantum potential
\begin{equation}\label{Qdef}
Q[\rho] = -\frac{\hbar^2}{2} \frac{\Delta \sqrt{\rho}}{\sqrt{\rho}}
\end{equation}
The imaginary part of the Schr\"{o}dinger\/ equation (also divided by $\psi(q)$) is the continuity equation for $\rho$
\begin{equation}\label{continuity}
\pd_t\rho(q,t) + \pd_i(\rho(q,t)v^i(q,t)) = 0
\end{equation}
Now, all what we have to do is to revert the aim -- instead of presenting the Schr\"{o}dinger\/ equation like a classical equation, let's present the classical Hamilton-Jacobi equation like a Schr\"{o}dinger\/ equation. There is almost nothing to do -- to add a density $\rho(q)$ together with a continuity equation \eqref{continuity} is trivial. We use the same polar decomposition formula \eqref{polar} to define the classical wave function. It remains to do the same procedure in the other direction. The difference is the same -- the quantum potential. So we obtain, as the equation for the classical wave function, an equation I have named pre-Schr\"{o}dinger\/ equation:
\begin{equation}\label{pre}
i\hbar \pd_t \psi(q,t) = -\frac{\hbar^2}{2} m^{ij}\pd_i\pd_j\psi(q,t) + (V(q) - Q[\rho]) \psi(q,t) = \hat{H} \psi - Q[|\psi|^2]\psi.
\end{equation}
Of course, the additional term is a nasty, nonlinear one. But that's the really funny point: We can obtain now quantum theory as the linear approximation of classical theory, in other words, as a simplification.
But there is more than this in this classical equation for the classical wave equation. The point is that there is an exact equivalence between the different formulations of classical theory, despite the fact that one is based on a classical trajectory $q(t)$ only, and the other has a wave function $\psi(q,t)$ together with the trajectory $q(t)$. And it is this exact equivalence which can be used to identify the meaning of the classical wave function.
Because of its importance, let's formulate this in form of a theorem:
\begin{theorem}[equivalence]
Assume $\psi(q,t),q(t)$ fulfill the pre-Schr\"{o}dinger\/ equation \eqref{pre} for some Hamilton function $H(p,q)$ of type \eqref{HamiltonFunction}, together with the guiding equation \eqref{guiding}.
Then, whatever the initial values $\psi_0(q),q_0$ for the wave function $\psi_0(q)=\psi(q,t_0)$ and the initial configuration $q_0=q(t_0)$, there exists initial values $q_0,p_0$ so that the Hamilton equation for $H(p,q)$ with these initial values gives $q(t)$.
\end{theorem}
\begin{proof}
The difficult part of this theorem is classical Hamilton-Jacobi theory, so, presupposed here as known. The simple modifications of this theory used here -- adding the density $\rho(q)$, with continuity equation \eqref{continuity}, and rewriting the result in terms of the wave function $\psi(q)$ defined by the polar decomposition \eqref{polar}, copying what has been done by Bohm \cite{Bohm} -- do not endanger the equivalence between Hamiltonian (or Lagrangian) and Hamilton-Jacobi theory.
\end{proof}
\section{The wave function describes incomplete information}
So let's evaluate now what follows from the equivalence theorem about the physical meaning of the (yet classical) wave function.
First, it is the classical (Hamiltonian or Lagrangian) variant which is preferable as a fundamental theory. There are three arguments to justify this:
\begin{itemize}
\item Simplicity of the set of fundamental variables: We need only a single trajectory $q(t)$ instead of trajectory $q(t)$ together with a wave function $\psi(q,t)$.
\item Correspondence between fundamental variables and observables: It is only the trajectory $q(t)$ in configuration space which is observable.
\item Stability in time: The wave function develops caustics after a short period of time and becomes invalid. The classical equations of motion does not have such a problem.
\end{itemize}
So it is the classical variant which is preferable as a fundamental theory. Thus, we can identify the true beable of classical theory with the classical trajectory $q(t)$.
The next observation is that, once $q(t)$ is known and fixed, the wave function contains many degrees of freedom which are unobservable in principle: Many different wave functions define the same trajectory $q(t)$. So we can conclude that these different wave functions do \emph{not} describe physically different states, containing additional beables.
On the other hand, this is true only if $q$ is known. What if $q$ is not known? In this case, the wave function defines simply a subset of all possible classical trajectories, and a probability measure on this subset.
To illustrate this, a particular most important example of a Hamilton-Jacobi function is useful: It is the function $S(q_0,t_0,q_1,t_1)$ defined by
\begin{equation}\label{Minimumproblem}
S(q_0,t_0,q_1,t_1) = \int_{t_0}^{t_1} L(q(t),\dot{q}(t),t) dt,
\end{equation}
where the integral is taken over the classical solutions $q(t)$ of this minimum problem with initial and final values $q(t_0)=q_0$, $q(t_1)=q_1$. This function fulfills the Hamilton-Jacobi equation in the variables $q_0,t_0$ as well as $q_1,t_1$. In above versions, it can be characterized as a Hamilton-Jacobi function $S(q,t)$ which is defined by a subset of trajectories: The function $S(q_0,t_0,q,t)$ describes the subset of trajectories going through $q_0$ at $t_0$, while the function $S(q,t,q_1,t_1)$ describes the subset of trajectories going through $q_1$ at $t_1$.
We can generalize this. The phase $S(q,t)$ tells us the value of $\dot{q}(t)$ given $q(t)$: \emph{If} $q(t)=q_0$ \emph{then} $\dot{q}(t)=v_0$, with $v_0$ defined by the guiding equation \eqref{guiding}. So, $S(q,t)$ always distinguishes a particular subset of classical trajectories.
Even more specific, this subset described by $S(q,t)$ can be uniquely described in terms of a subset of the possible initial values at the moment of time $t$ -- the configuration $q(t_0)$ and the momentum $p(t_0)=\nabla S(q(t_0),t_0)$, that means, as a subset of the possible values for the fundamental beables $q,p$ (or $q,\dot{q}$).
The other part -- the density $\rho(q)$ -- is nothing but a probability density on this particular subset.
Of course, a subset is nothing but a special delta-like probability measure, so that the wave function simply defines a probability density on the set of possible initial values:
\begin{equation}\label{probability}
\rho(p,q)dpdq = \rho(q)\delta(p-\nabla S(q))dpdq
\end{equation}
The pre-Schr\"{o}dinger\/ equation is, therefore, nothing but the evolution equation for this particular probability distribution.
So our classical, Hamilton-Jacobi wave function is nothing mystical, but simply a very particular form of a standard classical probability density $\rho(p,q)$ on the phase space.
In particular, the pre-Schr\"{o}dinger\/ equation for the wave function is nothing but a particular case of the Liouville equation, the standard law of evolution of standard classical probability distributions $\rho(p,q)$, for the particular ansatz \eqref{probability}, and follows from the fundamental law of evolution for the true, fundamental classical beables.
Moreover, the Liouville equation also defines the domain of applicability of the equation for the wave function. This domain is, in fact, restricted. In terms of $\rho(p,q)$, it will always remain a probability distribution on some Lagrangian submanifold. But this Lagrangian submanifold will be, after some time, no longer a graphic of a function $p=\nabla S(q)$ on configuration space -- there may be caustics, and in this case there will be several values of momentum for the same configuration $q$. If this happens, the wave function is no longer an adequate description.
Such an effect -- restricted validity -- is quite natural for the evolution of information, but not for fundamental beables.
So the wave function variant of Hamilton-Jacobi theory almost forces us to accept an interpretation of the wave function in terms of incomplete information. Indeed,
\begin{itemize}
\item The parts of the wave function, $\rho(q)$ as well as $S(q)$, make sense as describing a well-defined type of incomplete information about the classical configuration, namely the probability distribution $\rho(p,q)$ defined by \eqref{probability}.
\item The alternative, to interpret the wave function as describing some other, external beables, does not make sense, given the observational equivalence of the theory with simple Lagrangian classical mechanics, with $q(t)$ as the only observable. Additional beables should influence, at least in some circumstances, the $q(t)$. They don't.
\end{itemize}
So it looks like a incomplete information, it behaves like incomplete information, it becomes invalid like incomplete information -- it is incomplete information.
And, given that we know, in the case of the pre-Schr\"{o}dinger\/ equation, the fundamental law of evolution of the beables themself, it also makes no sense to reify this particular probability distribution as objective. A probability distribution defines a set of incomplete information about the real configuration, that's all.
\subsection{What about the more general case?}
Of course, the considerations above have been based on the equivalence theorem between the classical evolution and the evolution defined by the pre-Schr\"{o}dinger\/ equation. It was this exact equivalence which was able to give us some certainty, to remove all doubts that there is something else, some information about other beables, hidden in the wave function.
But what about the more general case, the case where we do not have an exact equivalence between an equation in terms of $q,\psi(q)$ and an equation purely in terms of the classical trajectory $q(t)$? In such a situation, the case for an interpretation of $\psi(q)$ in terms of incomplete information about the $q(t)$ is, of course, a little bit weaker.
Nonetheless, given such an ideal correspondence for the pre-Schr\"{o}dinger\/ equation, the interpretation remains certainly extremely plausible in a more general situation too. Indeed, why should one change the interpretation, the ontological meaning, of $\psi(q)$, if all what is changed is that we have replaced the pre-Schr\"{o}dinger\/ equation by another equation? The evolution equation is different, that's all. In itself, a change in the evolution equation does not give even a bit of motivation to change the interpretation.
And it should be noted that they have a hard job. The similarity between the two variants of the Schr\"{o}dinger\/ equation does not makes it easier: In fact, the funny observation that the Schr\"{o}dinger\/ equation is the linear part of the pre-Schr\"{o}dinger\/ equation becomes relevant here. If the Schr\"{o}dinger\/ equation would contain some new terms, this would open the door for attempts to show that the new terms do not make sense in the original interpretation. But there are no new terms in the linear part of the pre-Schr\"{o}dinger\/ equation. All terms of the Schr\"{o}dinger\/ equation are already contained in the pre-Schr\"{o}dinger\/ equation. So they all make sense.\footnote{It has to be mentioned in this connection that there is something present in the Schr\"{o}dinger\/ equation which is not present in the pre-Schr\"{o}dinger\/ equation -- the dependence on $\hbar$. So one can at least try to base an argument on this additional dependence. But, as described below, if one incorporates a Nelsonian stochastic process, the dependence on $\hbar$ appears in a natural way as connected with the stochastic process. So to use this difference to justify a modification of the ontology remains quite nontrivial.}
\subsection{The classical limit}
There is also another strong argument for interpreting above theories in the same way -- that classical theory appears as the classical limit of quantum theory.
The immediate consequence of this is that above theories have the same intention -- the description, as accurate as possible, of the same reality.
So this is not a situation where the same mathematical apparatus is applied to quite different phenomena, so that it is natural use a different interpretation even if the same mathematical apparatus is applied. In our case, we use the same mathematical formalism to describe the same thing. At least in the classical limit, quantum theory has to describe the same thing -- with the same interpretation -- as the classical interpretation.
\section{What follows from the interpretation}
But there is not only the similarity between the equations, and that the object is essentially the same, which suggests to use the same interpretation for above variants of the wave function.
There are also some interesting consequences of the interpretation. And these consequences, to be shared by all wave functions which follow this interpretation, are fulfilled by the quantum wave function.
\subsection{The relevance of the information contained in the wave function}
A first consequence of the interpretation can be found considering the \emph{relevance} of the information contained in the wave function, as information about the real beables. In fact, assume that the wave function $\psi(q)$ contains a lot of information which is irrelevant as information about the $q$. This would open the door for some speculation about the nature of this additional information. Once it cannot be interpreted as information about the $q$, it has to contain information about some other real beables. So let's consider which of the information contained in the wave function is really relevant, tells us something about the true trajectory $q(t)$.
Now, knowledge about the probability $\rho(q)$ is certainly relevant if we don't know the real value of $q$. And it is relevant in all of its parts.
Then, as we have seen, $S(q)$ gives us, via the ``guiding equation'' \eqref{guiding}, the clearly relevant information about the value of $\dot{q}$ given the value of $q$. Given that we don't know the true value of $q$, this information is clearly relevant. But, different from $\rho(q)$, the function $S(q)$ contains also a little bit more: Another function $S'=S+c$ would give \emph{exactly} the same information about the real beables.
As a consequence, the wave function $\psi(q)$ also contains a corresponding additional, irrelevant information. The polar decomposition \eqref{polar} defines what it is -- a global constant phase factor.
So we find that all the information contained in $\psi(q)$ -- \emph{except for a global constant phase factor} -- is relevant information about the real beables $q$.
At a first look, this seems trivial, but I think it is a really remarkable property of the paleoclassical interpretation.
The irrelevance of the phase factor is a property of the interpretation of the meaning of the wave function. It follows from this interpretation: We have considered all the ways how to obtain information about the beables from $\psi(q)$, and we have found that all the information is relevant, except this constant phase factor.
And now let's consider the equations. Above equations considered up to now -- the pre-Schr\"{o}dinger\/ equation as well as the Schr\"{o}dinger\/ equation -- have the same symmetry regarding multiplication with a constant phase factor: Together with $\psi(q,t)$, $c\psi(q,t)$ is a solution of the equations too.
This is, of course, how it should be if the wave function describes incomplete information about the $q$. If, instead, it would describe some different, external beables, then there would be no reason at all to assume such a symmetry. In fact, why should the values of some function, values at points far away from each other, be connected with each other by such a global symmetry?
So every interpretation which assigns the status of reality to $\psi(q)$ has, in comparison, a problem to explain this symmetry. A popular solution is to assign reality not to $\psi(q)$, but, instead, to the even more complicate density matrix $|\psi\rangle\langle\psi|$. The flow variable are, obviously, preferable by simplicity.
\subsection{The independence condition}
It is one of the great advantages of Jaynes' information-based approach to plausible reasoning \cite{Jaynes} that it contains common sense principles to be applied in situations with insufficient information. If we have no information which distinguishs the probability of the six possible outcomes of throwing a die, we \emph{have} to assign equal probability to them. Everything else would be irrational.\footnote{The purely subjectivist, de Finetti approach differs here. It does not make prescriptions about the initial probability distributions -- all what is required is that one updates the probabilities following the rules of Bayesian updating. This difference is one of the reasons for my preference for the objective approach proposed by Jaynes \cite{Jaynes}}
Similar principles work if we consider the question of independence. From a frequentist point of view, independence is a quite nontrivial physical assumption, which has to be tested. And, in fact, there is no justification for it, at least none coming from the frequentist interpretation. From point of view of the logic of plausible reasoning, the situation is much better: Once we have no \emph{information} which justifies any hypothesis of a dependence between two statements $A$ and $B$, we \emph{have} to assume their independence. There is no information which makes $A$ more or less plausible given $B$, so we have to assign equal plausibilities to them: $P(A|B)=P(A)$. But this is the condition of independence $P(AB)=P(A)P(B)$.
The different status of plausibilities in comparison with physical hypotheses makes this obligatory character reasonable. Probabilities are not hypotheses about reality, but logical conclusions derived from the available information, derived using the logic of plausible reasoning.
Now, all this is much better explained in \cite{Jaynes}, so why I talk here about this? The point is that I want to obtain here a formula which appropriately describes different independent subsystems. Subsystems are independent if we have no information suggesting their dependence. A quite typical situation, so independence is quite typical too.
So assume we have two subsystems, and we have no information suggesting any dependence between them. What do we have to assume based on the logic of plausible reasoning?
For the probability distribution itself, the answer is trivial:
\begin{equation}
\rho(q_1,q_2)=\rho_1(q_1)\rho_2(q_2).
\end{equation}
But what about the phase function $S(q)$? We have to assume that the velocity of one system is independent of the state of the other one. So the potential of that velocity also should not depend on the state of the other system. So we obtain
\begin{equation}
S(q_1,q_2) = S_1(q_1) + S_2(q_2).
\end{equation}
This gives, combined using the polar decomposition \eqref{polar}, the following rule for the wave function:
\begin{equation}
\psi(q_1,q_2) = \psi_1(q_1) \psi_2(q_2).
\end{equation}
Of course, again, the rule is trivial. Nobody would have proposed any other rule for defining a wave function in case of independence. Nonetheless, I consider it as remarkable that it has been derived from the very logic of plausible reasoning and the interpretation of the wave function.
And, again, this property is shared by above variants, the quantum as well as the classical one. And if one thinks about the applicability of this rule to the classical variant of the wave function, one has to recognize that it is nontrivial: From a classical point of view, the rule of combining $\rho(q)$ and $S(q)$ into a wave function $\psi(q)$ is quite arbitrary, and there is no a priori reason to expect that such an arbitrary combination transforms the rule for defining independent classical states into a simple rule for independent wave functions, moreover, into the one used in quantum theory.
\subsection{Appropriate reduction to a subsystem}
While the pre-Schr\"{o}dinger\/ equation is non-linear, it shares with the Schr\"{o}dinger\/ equation the weaker property of homogenity of degree one: The operator condition
\begin{equation}\label{homogenity}
\Omega(c\psi) = c \Omega(\psi)
\end{equation}
holds not only for the Schr\"{o}dinger\/ operator, but also for the non-linear pre-Schr\"{o}dinger\/ operator, and not only for $|c|=1$, as required by the $U(1)$ symmetry, which we have already justified, but for arbitrary $c$, so that the $U(1)$ symmetry is, in fact, a $GL(1)$ symmetry of the equations.
So one ingredient of linearity is shared by the pre-Schr\"{o}dinger\/ equation. This suggests that for this property there should be a more fundamental explanation, one which is shared by above equations. And, indeed, such an explanation is possible. There should be a principle which allows an appropriate reduction of the equation to subsystems, something like the following
\begin{principle}[splitting principle]
There should a be simple ``no interaction'' conditions for the operators on a system consisting of two subsystems of type $\Omega=\Omega_1+\Omega_2$ such that the equation $\pd_t\psi=\Omega(\psi)$ splits for independent product states $\psi(q_1,q_2)=\psi_1(q_1)\psi_2(q_2)$ into independent equation's for the two subsystems.
\end{principle}
Now, the time derivative splits nicely:
\begin{equation}
\pd_t \psi(q_1,q_2)= \pd_t\psi_1(q_1)\psi_2(q_2) + \psi_1(q_1)\pd_t\psi_2(q_2).
\end{equation}
It remains to insert the equations for the whole system $\pd_t \psi = \Omega(\psi)$ as well as for the two subsystems into this equation. This gives:
\begin{equation}
\Omega(\psi_1(q_1)\psi_2(q_2)) = \Omega_1(\psi_1(q_1))\psi_2(q_2) + \psi_1(q_1)\Omega_2(\psi_2(q_2)),
\end{equation}
where the $\Omega_i$ are subsystem operators acting only on the $q_i$, so that functions of the other variable are simply constants for them. On the other hand, in the splitting principle we have assumed $\Omega = \Omega_1+\Omega_2$. Comparison gives
\begin{equation}
\begin{split}
\Omega_1(\psi_1(q_1)\psi_2(q_2)) &= \Omega_1(\psi_1(q_1))\psi_2(q_2),\\
\Omega_2(\psi_1(q_1)\psi_2(q_2)) &= \psi_1(q_1)\Omega_2(\psi_2(q_2)).
\end{split}
\end{equation}
This follows from the homogenity condition \eqref{homogenity}. The weaker $U(1)$ symmetry is not sufficient, because the values of $\psi_2(q_2)$ may be arbitrary complex numbers.
Something similar to the splitting property is necessary for any equation relevant for us -- we have not enough information to consider the equation of the whole universe, thus, have to restrict ourself to small subsystems. And the equations of these subsystems should be at least approximately independent of the state of the environment.
So the homogenity of the equations -- in itself a quite nontrivial property of the equations, given the definition of the wave function by polar decomposition -- can be explained by this splitting property.
\subsection{Summary}
So we have found three points, each in itself quite trivial, but each containing some nontrivial element of explanation based on the paleoclassical principles: The global $U(1)$ phase shift symmetry of the wave function, explained by the irrelevance of the phase as information about $q(t)$, the product rule of independence, explained by the logic of plausible reasoning, applied to the set of information described by the wave function, and the homogenity of the equations, explained in terms of a splitting property for independent equations for subsystems.
All three points follow the same scheme -- the interpretation in terms of incomplete information allows to derive a rule, and this rule is shared by above theories, quantum theory as well as the wave function variant of classical theory.
So all three points give additional evidence that the simple, straightforward proposal to use the same interpretation of the wave function for above theories is the correct one.
\section{Incorporating Nelsonian stochastics}
The analogy between pre-Schr\"{o}dinger\/ and Schr\"{o}dinger\/ equation is a useful one, but it is that useful especially because the equations are nonetheless very different. And one should not ignore these differences.
One should not, in particular, try to interpret the Schr\"{o}dinger\/ equation as the linear approximation of the pre-Schr\"{o}dinger\/ equation: The pre-Schr\"{o}dinger\/ equation does not depend on $\hbar$. Real physical effects do depend. And so there should be also something in the fundamental theory, the theory in terms of the trajectory $q(t)$, which depends on $\hbar$.
Here, Nelsonian stochastics comes into mind. In Nelsonian stochastics the development of the configuration $q$ in time is described by a deterministic drift term $b^i(q(t),t)dt$ and a stochastic diffusion term $dB^i_t$
\begin{equation}\label{process}
dq^i(t) = b^i(q(t),t) dt + dB^i_t,
\end{equation}
with is a classical Wiener process $B^i_t$ with expectation $0$ and variance
\begin{equation}
\langle dB^i_t m_{ij} dB^j_t \rangle = \hbar dt,
\end{equation}
so that we have a $\hbar$-dependence on the fundamental level. The probability distribution $\rho(q(t),t)$ has, then, to fulfill the Fokker-Planck-equation:
\begin{equation}\label{FokkerPlanck}
\partial_t \rho + \pd_i (\rho b^i) - \frac{\hbar}{2}m^{ij}\pd_i\pd_j\rho = 0
\end{equation}
For the average velocity $v^i$ one obtains
\begin{equation}\label{vdef}
v^i(q(t),t) = b^i(q(t),t) - \frac{\hbar}{2}\frac{m^{ij}\pd_j \rho(q(t),t)}{\rho(q(t),t)},
\end{equation}
which fulfills the continuity equation. The difference between flow velocity $b^i$ and average velocity $v^i$, the osmotic velocity
\begin{equation}\label{osmotic}
u^i(q(t),t) = b^i(q(t),t)-v^i(q(t),t) = \frac{\hbar}{2}\frac{m^{ij}\pd_j\rho(q(t),t)}{\rho(q(t),t)} = \frac{\hbar}{2}m^{ij}\pd_j \ln\rho(q(t),t)
\end{equation}
has a potential $\ln\rho(q)$. The average acceleration is given by
\begin{equation}\label{acceleration}
a^i(q(t),t) = \partial_t v^i + v^j \pd_j v^i - \frac{\hbar^2}{2} m^{ij}\pd_j
\left(\frac{m^{kl}\pd_k\pd_l \sqrt{\rho(q(t),t)}}{\sqrt{\rho(q(t),t)}}\right)
\end{equation}
For this average acceleration the classical Newtonian law
\begin{equation}\label{Newton}
a^i(q(t),t) = - m^{ij} \pd_j V(q(t),t)
\end{equation}
is postulated. Putting this into the equation \eqref{acceleration} gives
\begin{equation}\label{derivativeBohm}
\partial_t v^i + (v^j \pd_j)v^i =
- m^{ij} \pd_j \left(V -\frac{\hbar^2}{2}
\frac{\Delta \sqrt{\rho}}{\sqrt{\rho}}
\right) = - m^{ij}\pd_j(V + Q[\rho]).
\end{equation}
The next postulate is that the average velocity $v^i(q)$ has, at places where $\rho>0$, a potential, that means, a function $S(q)$, so that \eqref{guiding} holds. Then equation \eqref{derivativeBohm} can be integrated (putting the integration constant into $V(q)$), which gives \eqref{Bohm}. Finally, one combines the equations for $S(q)$ and $\rho(q)$ into the Schr\"{o}dinger\/ equation as in de Broglie-Bohm theory.
So far the basic formulas of Nelsonian stochastics. Now, the beables of Nelsonian stochastics are quite different from those of the paleoclassical interpretation. The external flow $b^i(q,t)$ does not define some state of information, but some objective flow which takes away the configuration if it is at $q$. The configuration is guided in a way close to a swimmer in the ocean: He can randomly swim into one or another direction, but whereever he is, he is always driven by the flow of the ocean, a flow which exists independent of the swimmer.
What would be the picture suggested for a stochastic process by the paleoclassical interpretation? It would be different, more close to a spaceship flying in the cosmos. If it changes its place because of some stochastic process, there will be no predefined, independent velocity $b^i(q)$ at the new place which can guide it. Instead, it would have to follow its previous velocity. Indeed, the velocity field $v^i(q)$, and, therefore, also $b^i(q)$, of the new place describes only information, not a physical field which could possibly influence the spaceship in its new position.
But what follows for the information if there is some probability that the stochastic process causes the particle to change its location to the new one? It means that the old average velocity has to be recomputed.
Of course, there is such a necessity to recompute only if the velocity at the new location is different from the old one. So, if $\pd_i v^k=m^{jk}\pd_i p_j=0$, there would be no need for such a recomputation at all. This condition is clearly too strong, it would cause the whole velocity field to become constant. Now there is an interesting subcondition, namely that $\pd_i p_j-\pd_j p_i=0$, the condition that the velocity field is a potential one. So the re-averaging of the average velocity $v^i(q)$ caused by the stochastic process will decrease the curl.
This gives a first advantage in comparison with Nelsonian stochastics: The potentiality assumption does not have to be postulated, without any justification, for the external flow $b^i(q,t)$. (It is postulated for average velocity, but their difference -- the osmotic velocity -- has a potential anyway.) In the paleoclassical interpretation we have an independent motivation for postulating this. But, let's recognize, there is no fundamental reason to postulate potentiality: On the fundamental level, the curl may be nontrivial. All what is reached is that the assumption of potentiality is not completely unmotivated, but that it appears as a natural consequence of the necessary re-averaging.
There is another strangeness connected with the external flow picture of Nelsonian stochastics. The probability distribution $\rho(q)$ already characterizes the information about the configuration itself -- it is a probability for the swimmer in the ocean, not for the ocean. Once they depend on $\rho(q)$, as the average velocity $v^i(q)$, as the average acceleration $a^i(q)$ also describe information about the swimmer, not about the ocean. Then, it is postulated that the average acceleration of the swimmer has to be defined by the Newtonian law. But because this average velocity is, essentially, already defined by the very process -- the flow of the ocean -- together with the initial value for the swimmer, this condition for the swimmer becomes an equation for the ocean. This is conceptually unsound -- as if the ocean has to care that the swimmer is always correctly accelerated.
But this conceptual inconsistency disappears in the paleoclassical interpretation. The drift field is now part of the incomplete information about the configuration itself, as defined by the average velocity and osmotic velocity. There is no longer any external drift field. And, so, it is quite natural that a condition for the average acceleration of the configuration gives an equation for the average velocity $v^i(q)$. So the paleoclassical picture is internally much more consistent.
But is it viable? This is a quite non-trivial question discussed below in sec. \ref{viability}.
\section{The character of the wave function}
Let's start with the consideration of the objections against the paleoclassical interpretation. Given that the basic formulas are not new at all, I do not have to wait for reactions to this paper -- some quite strong arguments are already well-known.
The first one is the evidence in favour of the thesis that the wave function describes the behaviour of real degrees of freedom, degrees of freedom which actually influence the things we can observe immediately. Here, at first Bell's argumentation comes into mind -- an argumentation for a double ontology which, I think, has impressed many of those who support today realistic interpretations:
\begin{quote}
Is it not clear fro the smalleness of the scintillation on the screen that we have to do with a particle? And is it not clear, from the diffraction and interference patterns, that the motion of the particle is directed by a wave?
(\cite{Bell} p. 191).\end{quote}
But what are the points which make this argument that impressive? What is it what motivates us to accept some things as real? Here I see no way to express this better than Brown and Wallace:
\begin{quote}
From the corpuscles' perspective, the wave-function is just a (time-dependent) function on their configuration space, telling them how to behave; it superficially appears similar to the Newtonian or Coulomb potential field, which is again a function on configuration space. No-one was tempted to reify the Newtonian potential; why, then, reify the wave-function?
Because the wave-function is a very different sort of entity. It is contingent (equivalently, it has dynamical degrees of freedom independent of the corpuscles); it evolves over time; it is structurally overwhelmingly more complex (the Newtonian potential can be written in closed form in a line; there is not the slightest possibility of writing a closed form for the wave-function of the Universe.) Historically, it was exactly when the gravitational and electric fields began to be attributed independent dynamics and degrees of freedom that they were reified: the Coulomb or Newtonian `fields' may be convenient mathematical fictions, but the Maxwell field and the dynamical spacetime metric are almost universally accepted as part of the ontology of modern physics.
We don't pretend to offer a systematic theory of which mathematical entities in physical theories should be reified. But we do claim that the decision is not to be made by fiat, and that some combination of contingency, complexity and time evolution seems to be a requirement.
(\cite{BrownWallace} p. 12-13)\end{quote}
So, let's consider the various points in favour of the reality of the wave function:
\subsection{Complexity of the wave function}
The argument of complexity seems powerful. But, in fact, for an interpretation in terms of incomplete information \emph{this} is not a problem at all.
Complexity is, in fact, a natural consequence of incompleteness of information. The complete information about the truth of a statement is a single bit: true or false. The incomplete information is much more complex: it is a real number, the probability.
As well, the complete information about reality in this interpretation is simple: a single trajectory $q(t)$. But incomplete information requires much more information: Essentially, we need probabilities for all possible trajectories.
\subsection{Time dependence of the wave function}
Time dependence is, as well, a natural property of information -- complete or incomplete. The information about where the particle has been yesterday transforms into some other information about where the particle is now.
This transformation is, moreover, quite nontrivial and complex.
It is also worth to note here that the law of transformation of information is a derivative of the real, physical law of the behaviour of the real beables.
So it necessarily has all the properties of a physical law.
We can, in particular, use the standard Popperian scientific method (making hypotheses, deriving predictions from them, testing and falsifying them, and inventing better hypotheses) to find these laws.
This is, conceptually, a quite interesting point: The laws of probability themself are best understood, following Jaynes \cite{Jaynes}, as laws of extended logic, of the logic of plausible reasoning.
But, instead, the laws of \emph{transformation} of probabilities in time follow from the laws of the original beables in time, and, therefore, have the character of physical laws.
Or, in other words, incomplete information develops in time in a way indistinguishable from the development in time of real beables. In particular, we use the same objective scientific methods to find and test them.
So, nor the simple fact that there is a nontrivial time evolution, nor the very physical character of the dynamical laws, in all details, up to the scientific method we use to find them, give any argument against an interpretation in terms of incomplete information. All this is quite natural for the evolution of incomplete information too.
\subsection{The contingent character of the wave function}
There remains the most powerful argument in favour of the reality of the wave function: Its contingent character.
There are different wave functions, and these different wave functions lead to objectively different probability distributions of the observable results.
If we have, in particular, different preparation procedures, leading to different interference pictures, we really observe different interference pictures. It is completely implausible that these different interference pictures -- quite objective pictures -- could be the result of different sets of incomplete information about the same reality. The different interference pictures are, clearly, the result of different things happening in reality.
But, fortunately, there is not even a reason to disagree with this. The very point is that one has to distinguish the wave function of a small subsystem -- we have no access to any other wave functions -- from the wave function of a closed system. The last is, in fact, only an object of purely theoretical speculation, because there is no closed system in nature except the whole universe, but we have no idea about the wave function of the whole universe.
For the wave function of a small subsystem, the situation is quite different. It does not contain only incomplete information about the subsystem. In fact, it is only an effective wave function, and there is a nice formula of dBB theory, which can be used in our paleoclassical interpretation too: The formula which defines the conditional wave function of a subsystem $\psi^S(q_S,t)$ in terms of the wave function of the whole system (say the whole universe) $\psi(q_S,q_E,t)$ and the configuration of the environment $q_E(t)$:
\begin{equation}\label{conditional}
\psi^S(q_S,t) = \psi(q_S,q_E(t),t)
\end{equation}
This is a remarkable formula of dBB theory which contains, in particular, the solution of the measurement problem: The evolution of $\psi^S(q_S,t)$ is, in general, not described by a Schr\"{o}dinger\/ equation -- if there is interaction with the environment, the evolution of $\psi^S(q_S,t)$ is different, but, nonetheless, completely well-defined. And this different evolution is the collapse of the wave function caused by measurement.
Let's note that the paleoclassical interpretation requires to justify this formula in terms of the information about the subsystem. But this is not a problem. Indeed, assume the trajectory of the environment $q_E(t)$ is known -- say by observation of a classical, macroscopic measurement device. Then the combination of the knowledge described by the wave function of the whole system with the knowledge of $q_E(t)$ gives exactly the same knowledge as that described by $\psi^S(q_S,t)$. Indeed, the probability distribution gives
\begin{equation}
\rho^S(q_S,t) = \rho(q_S,q_E(t),t),
\end{equation}
and, similarly, the velocity field defined by $S(q)$ follows the same reduction principle:
\begin{equation}
\nabla S^S(q_S,t) = \nabla S(q_S, q_E(t),t).
\end{equation}
So in the paleoclassical interpretation the dBB formula the conditional wave function of a subsystem is a logical necessity. This provides yet another consistency check for the interpretation.
But the reason for considering this formula here was a different one: The point is that the wave function of the subsystem in fact contains important information about other real beables -- the actual configuration of the whole environment $q_E(t)$. So there are real degrees of freedom, different from the configuration of the system $q_S(t)$ itself, which are distinguished by different wave functions $\psi^S(q_S,t)$.
And we do not have to object at all if one argues that the wave function contains such additional degrees of freedom. That's fine, it really contains them. These degrees of freedom are those of the configuration of the environment $q_E(t)$. And this is not an excuse, but a logical consequence of the interpretation itself, a consequence of the definition of the conditional wave function of the subsystem \eqref{conditional}.
\subsection{Conclusion}
So the consideration of the arguments in favour of a beable status for the wave function, even if they seem to be strong and decisive at a first look, appear to be in no conflict at all with the interpretation of the wave function of a closed system in terms of incomplete information about this system.
The most important point to understand this is, of course, the very fact that the conditional wave function of the small subsystems of the universe we can consider really have a different character -- they depend on the configuration of the environment. And, so, the argumentation that these conditional wave functions describe real degrees of freedom, external to the system itself, is accepted and even derived from the interpretation.
Nonetheless, the point that nor the complexity of the wave function of the whole universe, nor its evolution in time, nor the physical character of the laws of this evolution are in any conflict with an interpretation in terms of incomplete information is an important insight too.
\section{The Wallstrom objection}
Wallstrom \cite{Wallstrom} has made an objection against giving the fields of the polar decomposition $\rho(q)$, $S(q)$ (instead of the wave function $\psi(q)$ itself) a fundamental role.
The first point is that around the zeros of the quantum mechanical wave function, the flow has no longer a potential. The quantum flow is a potential one only where $\rho(q)>0$. But in general there will be submanifolds of dimension $n-2$ where the wave function is zero. And for a closed path $q(s)$ around such a zero submanifold one finds that
\begin{equation}\label{curl}
\oint m_{ij} v^i(q) dq^j \neq 0.
\end{equation}
This, in itself, is unproblematic for the interpretation: The condition of potentiality is not assumed to be a fundamental one -- the fundamental object is not $S(q)$ but the $v^i(q)$. There will be some mechanism in subquantum theory which locally reduces violations of potentiality, so we can assume that the flow is a potential one only as an approximation.
It is also quite natural to assume that such a mechanism works more efficient for higher densities and fails near the zeros of the density.
So having them localized at the zeros of the density is quite nice -- not for a really fundamental equation, which is not supposed to have any infinities, but if we consider quantum theory as being only an approximation.
The really problematic part of the Wallstom objection is a different one: It is that the quantum flow has to fulfill a nontrivial \emph{quantization condition}, namely
\begin{equation}\label{curlquantization}
\oint m_{ij} v^i(q) dq^j = \oint \pd_j S(q) dq^j = 2\pi m\hbar, \qquad m\in\mbox{$\mathbb{Z}$}.
\end{equation}
The point is, in particular, that the equations \eqref{Bohm}, \eqref{continuity} in flow variables are not sufficient to derive this quantization condition. So, in fact, this set of equations is \emph{not} empirically equivalent to the Schr\"{o}dinger\/ equation.
This is, of course, no wonder, given the fact that the equivalence holds only for $\rho(q)=|\psi(q)|^2>0$. But, however natural, empirical inequivalence is empirical inequivalence.
Then, this condition looks quite artificial in terms of the $v^i$. What is a triviality in terms of the wave function -- that it has to globally uniquely defined -- becomes extremely artificial and strange if formulated in terms of the $v^i(q)$. As Wallstrom \cite{Wallstrom} writes, to ``the best of my knowledge, this condition [\eqref{curlquantization}] has not yet found any convincing explanation outside the context of the Schr\"{o}dinger\/ equation''.
\subsection{A solution for this problem}
Fortunately I have found a solution for this problem in \cite{againstWallstrom}. I do not claim that it is a complete one -- there is a part which is beyond the scope of an interpretation, which has to be left to particular proposals for a subquantum theory. One has to check if the assumptions I have made about such a subquantum theory are really fulfilled in that particular theory.
The first step of the solution is to recognize that, for empirical equivalence with quantum theory, it is sufficient to recover only solutions with simple zeros. Such simple zeros give $m=\pm 1$ in the quantization condition \eqref{curlquantization}. This is a consequence of the principles of general position: A small enough modification of the wave function cannot be excluded by observation, but leads to a wave function in general position, and in general position the zeros of the wave function are non-degenerated.
The next step is a look at the actual solutions. For the simple, two dimensional, rotational invariant, zero potential case these solutions are defined by $S(q)=m\varphi$, $\rho(q)=r^{2|m|}$. And this extends to the general situation, where $S(q)=m\varphi+\tilde{S}(q)$, $\rho(q)=r^{2|m|}\tilde{\rho}(q)$, such that $\tilde{S}(q)$ is well-defined in a whole environment of the zero, and $\tilde{\rho}(0)>0$.
But that means we can replace the problem of justifying an integer $m$ in $S(q)=m\varphi$, where all values of $m$ seem equally plausible, by the quite different problem of justifying $\rho(q)=r^2$ (once we need only $m=\pm 1$) in comparison with other $\rho(q)=r^\a$. This is already a quite different perspective.
We make the natural conclusion and invent a criterion which prefers $\rho(q)=r^2$ in comparison with other $r^\a$. This is quite easy:
\begin{postulate}[regularity of $\Delta\rho$]\label{preg}
If $\rho(q)=0$, then $0< \Delta \rho(q) < \infty$ almost everywhere.
\end{postulate}
This postulate already solves the inequivalence argument. The equations \eqref{Bohm}, \eqref{continuity} for $\rho(q)>0$, together with postulate \ref{preg}, already defines a theory empirically equivalent to quantum theory (even if the equivalence is not exact, because only solutions in general position are recovered).
It remains to invent a justification for this postulate.
The next step is to rewrite equation \eqref{Bohm} for stable states in form of a balance of energy densities. In particular, we can rewrite the densitized quantum potential as
\begin{equation}\label{Q}
Q[\rho]\rho = \frac12\rho u^2 - \frac14 \Delta \rho,
\end{equation}
with the ``osmotic velocity'' $u(q) = \frac12 \nabla \ln \rho(q)$. Then the energy balance looks like
\begin{equation}
\frac12\rho v^2 + \frac12 \rho u^2 + V(q) = \frac14\Delta\rho.
\end{equation}
So, the operator we have used in the postulate is not an arbitrary expression, but a meaningful term, which appears in an important equations -- an energy balance. This observation is already sufficient to justify the $\Delta \rho(q) < \infty$ part of the condition. There may be, of course, subquantum theories which allow for infinities in energy densities, but it is hardly a problem for a subquantum theory to justify that expressions which appear like energy densities in energy balances have to be finite.
Last but not least, subquantum theory has to allow for nonzero curl $\nabla\times v$, but has to suppress it to obtain a quantum limit. One way to suppress it is to add a penalty term $U(\rho,\nabla\times v)$ which increases with $|\nabla\times v|$. This would give
\begin{equation}
\frac12\rho v^2 + \frac12 \rho u^2 + V(q) + U(\rho,\nabla\times v) = \frac14\Delta\rho.
\end{equation}
Moreover subquantum theory has to regularize the infinities of $v$ and $u$ at the zeros of the density. One can plausibly expect that this gives finite but large values of $|\nabla\times v|$ at zero which decrease sufficiently fast with $r$. Now, a look at the energy balance shows that, if the classical potential term $V(q)$ is neglected (for example assuming that it changes only smoothly) the only term which can balance $U(\rho,\nabla\times v)$ at zero is the $\Delta \rho$ terms, which, therefore, has to be finite but nonzero. Or, at least, it would not be difficult to modify the definition of $U(\rho,\nabla\times v)$ in such a way that the extremal value $\Delta\rho=0$ (we have necessarily $\Delta\rho\ge 0$ at the minima) has to be excluded.
So the postulate seems nicely justifiable. For some more details I refer to \cite{againstWallstrom}. What remains is the particular job of particular proposals for subquantum theories -- they have to check if the way to justify the postulate really works in this theory, or if it may be justified in some other way. But this is beyond the scope of the interpretation. What has to be done by the interpretation -- in particular, to obtain empirical equivalence with quantum theory -- has been done.
\section{A Popperian argument for preference of an information-based interpretation}
One of Popper's basic ideas was that we should prefer -- as long as possible without conflict with experience -- theories which are more restrictive, make more certain predictions, depend on less parameters. And, while this criterion has been formulated for theories, it should be applied, for the same reasons, to more general principles of constructing theories too.
This gives an argument in preference for an interpretation in terms of incomplete information.
Indeed, let's consider, from this point of view, the difference between interpretations of fields $\rho(q)$, $v^i(q)$ in terms of a probability for some real trajectories $q(t)$, and interpretations which reify them as describing some external reality, different from $q(t)$, which influences the trajectory $q(t)$.
It is quite clear and obvious which of the two approaches is more restrictive. Designing theories of the first type, we are restricted, for the real physics, to theories for single trajectories $q(t)$. Then, given that we have identified the connection between the fields $\rho(q)$, $v^i(q)$ and $q(t)$ as those of a probability flow, everything else follows. There is no longer any freedom of choice for the field equations. If we have fixed the Hamiltonian evolution for $p(t), q(t)$, the Liouville field equation for $\rho(p,q)$ is simply a logical consequence. Similarly, the continuity equation \eqref{continuity} is a law of logic, it cannot be modified, is no longer a subject of theoretical speculation. It is fixed by the interpretation of $\rho(q)$, $v^i(q)$ as a probability flow, in the same way as $\rho(q)\ge 0$ is fixed.
In the second case, we have much more freedom -- the full freedom of speculation about field theories in general. In particular, the continuity equation can be modified, introducing, say, some creation and destruction processes, which are quite natural if $\rho(q)$ describes a density of some external objects.
The derivation of the $U(1)$ global phase shift symmetry is another particular example of such an additional logical law following from the interpretation.
So there are some consequences of the interpretation which have purely logical character, including the continuity equation, $\rho(q)\ge 0$, and the $U(1)$ global phase shift symmetry. But these will not be the only consequences. The other equations will be restricted, in comparison with field theories, too, but in a less clear and obvious way. There is, last but not least, a large freedom of choice for the equations of the real beables $q(t)$, which correspondence to a similarly large freedom of choice of the resulting equations for $\rho(q)$, $v^i(q)$. But this freedom of choice will be, nonetheless, much smaller than the completely arbitrariness of a general field theory.
This consideration strongly indicates that we have to prefer the interpretation in terms of incomplete information until it has been falsified, until it appears incompatible with observation.
The immediate, sufficiently trivial logical consequences we have found yet are compatible with the Schr\"{o}dinger\/ equation and therefore with observation. So we should prefer this interpretation.
\section{Open problems}
Instead of using such a Popperian argumentation, I could have, as well, used simply Ockham's razor: Don't multiply entities without necessity. Once, given this interpretation, there is no necessity for more than a single classical trajectory $q(t)$, one should not introduce other real entities like really existing wave functions.
But the Popperian consideration has the advantage that it implicitly defines an interesting research program.
\subsection{Other restrictions following from the interpretation}\label{viability}
In fact, given the restrictive character of the interpretation, there may be other, additional, more subtle restrictions of the equations for probility flows $\rho(q)$, $v^i(q)$, restrictions which we have not yet identified, but which plausibly exist.
So what are these additional restrictions for equations for probability flows $\rho(q)$, $v^i(q)$ in comparison with four general, unspecific fields fulfilling a continuity equation? I have no answer.
This is clearly an interesting question for future research. It is certainly also a question interesting in itself, interesting from point of view of pure mathematics, for a better understanding of probability theory.
The consequences may be fatal for this approach -- it may be that we find that the Schr\"{o}dinger\/ equation does not fit into this set of restrictions. This possibility of falsifiction is, of course, the very point of the Popperian consideration. I'm nonetheless not afraid that this happens, but this is only a personal opinion.
The situation may be, indeed, much better: That this subclass of theories contains the Schr\"{o}dinger\/ equation, but appears to be heavily restricted by some additional, yet unknown, conditions. Then, all of these additional restrictions give us partial answers to the ``why the quantum'' question.
\subsection{Why is the Schr\"{o}dinger\/ equation linear?}
The most interesting question which remains open is why the Schr\"{o}dinger\/ equation is linear. We have found only some part of the answer -- an explanation for the global $U(1)$ phase symmetry based on the informational content and of the homogeneity based on the reduction of the equation to subsystems.
But, given that the pre-Schr\"{o}dinger\/ equation is non-linear, but interpreted in the same way, the linearity of the Schr\"{o}dinger\/ equation cannot follow from the interpretation taken alone. Some other considerations are necessary to obtain linearity.
One idea is to justify linearity as an approximation. Last but not least, linearization is a standard way to obtain approximations.
The problem of stability in time may be also relevant here. The pre-Schr\"{o}dinger\/ equation becomes invalid after a short period of time, then the first caustic appears. There is no such problem in quantum theory, which has a lot of stable solutions. Then, there should be not only stable solutions, but also slowly changing solutions: It doesn't even matter if the fundamental time scale is Planck time or something much larger -- even if it is only the time scale of strong interactions, all the things changing around us are changing in an extremely slow way in comparison with this fundamental time scale. But the linear character of the Schr\"{o}dinger\/ equation gives us a way to obtain solutions slowly changing in time by combining different stable solutions with close energies.
\section{A theoretical possibility to test: The speedup of quantum computers}\label{computer}
There is an idea which suggests, at least in principle, a way to distinguish observationally the paleoclassical interpretation (or, more accurate, the class of all more fundamental theories compatible with the paleoclassical interpretation) from the minimal interpretation.
The idea is connected with the theory of quantum computers. If quantum computers really work as predicted by quantum theory, these abilities will provide fascinating tests of the accuracy of quantum theory. In the case of Simon's algorithm, the speed-up is exponential over any classical algorithm. It may be a key for the explanation of this speed-up that the state space (phase space) of a composite classical system is the Cartesian product of the state spaces of its subsystems, while the state space of a composite quantum system is the tensor product of the state spaces of its subsystems. For $n$ qubits, the quantum state space has $2^n$ instead of $n$ dimensions. So the information required to represent a general state increases exponentially with $n$ (see, for example, \cite{Bub}). There is also the idea ``that a quantum computation is something like a massively parallel classical computation, for all possible values of a function. This appears to be Deutsch's view, with the parallel computations taking place in parallel universes.'' \cite{Bub}.
It is this \emph{exponential} speedup which suggests that the predictions of standard QM may differ from those of the paleoclassical interpretation. An exact quantum computer would have all beables contained in the wave function as degrees of freedom. A quantum computer in the paleoclassical interpretation has only the resources provided by its beables. But these beables are, essentially, only the classical states of the universe. Given the exponential difference between them, $n$ vs. $2^n$ dimensions for qubits instead of classical bits, an exact quantum computer realizable at least in principle in a laboratory on Earth can have more computational resources than a the corresponding computer of the paleoclassical interpretation, which can use only the classical degrees of freedom, even if these are the classical degrees of freedom of the whole universe.
But if we distort a quantum computer, even slightly, the result will be fatal for the computation. In particular, if this distortion is of the type of the paleoclassical interpretation, which replaces an exact computer with a $2^n$-dimensional state space by an approximate one with only $N$ dimensions, then even for quite large $N\gg n$ the approximate computer will be simply unable to do the exact computations, even in principle. There simply are no parallel universes in the paleoclassical interpretation to make the necessary parallel computations.
So, roughly speaking, the prediction of the paleoclassical interpretation is that a sufficiently large quantum computer will fail to give the promised exponential speedup. The exponential speedup will work only up to a certain limit, defined by the logarithm of the relation between the size of the whole universe and the size of the quantum computer.
Of course, we do not know the size of the universe. It may be much larger than the size of the observable universe, or even infinite. Nonetheless, this argument, applied to any finite model of the universe, shows that the true theory, the theory in the configurational beables alone, cannot be exactly quantum theory. This is in my opinion the most interesting conclusion.
But let's see if we can, nonetheless, make even testable (at least in principle) predictions. So let's presuppose that the universe is finite, and, moreover, let's assume that this size is not too many orders larger than the its observable part. This would be already sufficient to obtain some numbers about the number of qubits such that the $2^n$ exponential speedup is no longer possible. This number will be sufficiently small, small enough that a quantum computer in a laboratory on Earth will be sufficient to reach this limit.
And, given the logarithmic dependence on $N$, increasing $N$ does not help much. If it is possible to build a quantum computer with $n$ qubits, why not with $2n$? This would already move the size of the universe into completely implausible regions.
\subsection{The speed of quantum information as another boundary}
Instead of caring about the size of the universe, it may be more reasonable to care about the size of the region which can causally influence us. Here I do not have in mind the limits given by of relativity, by the speed of light. Given the violation of Bell's inequality, there has to be (from a realist's point of view) a hidden preferred frame where some other sort of information -- quantum information -- is transferred with a speed much larger than the speed of light. But if we assume that the true theory has some locality properties, even if only in terms of a much larger maximal speed, the region which may be used by a quantum computer for its computational speedup decreases in comparison with the size of the universe.
So if we assume that there is such a speed limit for quantum information, then we obtain in the paleoclassical interpretation even more restrictive limits for the speedup reachable by quantum computers, limits which depend logarithmically on the speed limit for quantum information.
Nonetheless, I personally don't believe that quantum computers will really reach large speedups. I think the general inaccuracy of human devices will prevent us from constructing quantum computers which can really use the full $2^n$ power for large enough $n$. I would guess that the accuracy requirements necessary to obtain a full $2^n$ speedup will also grow exponentially. So I guess that quantum computers will fail already on a much smaller scale.
\section{Conclusions}
So it's time to summarize:
\begin{itemize}
\item The unknown, true theory of the whole universe is a theory defined on the classical configuration space $Q$, with the configuration $q(t)$ evolving in absolute time $t$ as a complete description of all beables.
\item The wave function of the whole universe is interpreted as a consistent set of incomplete information about these fundamental beables.
\item In particular, $\rho(q)$ defines not some ``objective'' probability, but an incomplete set of information about the real position $q$, described, as required by the logic of plausible reasoning, by a probability distribution $\rho(q)dq$.
\item The phase $S(q)$ describes, via the ``guiding equation'', the expectation value $\langle\dot{q}\rangle$ of the velocity given the actual configuration $q$ itself. So the ``guiding equation'' is not a physical equation, but has to be interpreted as part of the definition of $S(q)$, which describes which information about $q$ is contained in $S(q)$.
\item Only a constant phase factor of $\psi(q)$ does not contain any relevant information about the trajectory $q(t)$. Therefore, the equations for $\psi(q)$ should not depend on such a factor.
\item The Schr\"{o}dinger\/ equation is interpreted as an approximate equation. More is not to be expected, given that it describes the evolution of an incomplete set of information.
\item The linear character of the Schr\"{o}dinger\/ equation is interpreted as an additional hint that it is only an approximate equation.
\item The interpretation can be used to reinterpret Nelsonian stochastics. The resulting picture is conceptually more consistent than the original proposal.
\item The Wallstrom objection appears much less serious than expected. The quantization condition for simple zeros (which is sufficient because it is the general position) can be derived from the much simpler regularity postulate that $0<\Delta \rho(q)<\infty$ if $\rho(q)=0$. While a final justification of this condition has to be left to a more fundamental theory, it is, as shown in \cite{againstWallstrom}, plausible that this is not a problem for such theories.
\item If the true theory of the universe is defined on classical configurations, and the whole universe is finite, quantum computers can give their promised exponential speedup only up to an upper bound for the number of qubits, which is much smaller than the available qubits of the universe. This argument shows that the Schr\"{o}dinger\/ equation has to be approximate.
\item The dBB problem with of the ``action without reaction'' asymmetry is solved: For effective wave function, the collapse defines the back-reaction, for the wave function of the whole universe there should be no such back-reaction -- it is only an equation about incomplete information about reality, not about reality itself.
\item The wave functions of small subsystems obtain a seemingly objective, physical character only because they, as conditional wave functions, depend on the physical beables of the environment.
\end{itemize}
From point of view of simplicity, the paleoclassical interpretation is superior to all alternatives. The identification of the fundamental beables with the classical configuration space trajectory $q(t)$ is sufficient for this point.
It has also the additional advantage that it leads to strong restrictions of the properties of a more fundamental, sub-quantum theory: It has to be a theory completely defined on the classical configuration space. Moreover, it has to be a theory which, in its statistical variant, leads to a Fokker-Planck-like equations for the probability flow defined by the classical flow variables $\rho(q)$ and $v^i(q)$.
\begin{appendix}
\section{Compatibility with relativity}\label{relativity}
Most physicists consider the problem of compatibility with relativity as the major problem of dBB-like interpretations -- sufficient to reject them completely. But I have different, completely independent reasons for accepting a preferred frame, so that I don't worry about this.
There are two parts of this compatibility problem, a physical and a philosophical one, which should not be mingled:
The physical part is that we need a dBB version of relativistic quantum field theories, in particular of the standard model of particle physics -- versions which do not have to change the fundamental scheme of dBB, and, therefore, may have a hidden preferred frame.
The philosophical part is the incompatibility of a hidden preferred frame with relativistic metaphysics.
The physical problem is heavily overestimated, in part because of the way dBB theory is often presented: As a theory of many particles. I think it should be forbidden to introduce dBB theory in such a way. The appropriate way is to present it as a general theory in terms of an abstract configuration space $Q$, and to recognize that field theories as well as their lattice regularizations fit into this scheme. The fields are, of course, fields on three-dimensional space $\mbox{$\mathbb{R}$}^3$ changing in time, and their lattice regularizations live on three-dimensional spatial lattices $\mbox{$\mathbb{Z}$}^3$, not four-dimensional space-time lattices. But this violation of manifest relativistic symmetry is already part of the second, philosophical problem.
The simple, seemingly non-relativistic Hamiltonian \eqref{HamiltonFunction}, with $p^2$ instead of $\sqrt{p^2 + m^2}$, is also misleading: For relativistic field theories the quadratic Hamiltonian is completely sufficient. Indeed, a relativistic field Lagrangian is of type
\begin{equation}\label{field}
\mathscr{L} = \frac{1}{2}((\pd_t\varphi)^2 - (\pd_i\varphi)^2)-V(\varphi).
\end{equation}
This gives momentum fields $\pi = \pd_t\varphi$ and the Hamiltonian
\begin{equation}
\mathscr{H} = \frac{1}{2}(\pi^2 + (\pd_i\varphi)^2)+V(\varphi) = \frac{1}{2}\pi^2 + \tilde{V}(\varphi)
\end{equation}
quadratic in $\pi$, thus, the straightforward field generalization of the standard Hamiltonian \eqref{HamiltonFunction}. And a for lattice regularization, the Hamiltonian is already exactly of the form \eqref{HamiltonFunction}. So, whatever one thinks about the dBB problems with other relativistic fields, it is certainly not relativity itself which causes the problem.
The problem with fermions and gauge fields is certainly more subtle. Here, my proposal is described in \cite{clm}. It heavily depends on a preferred frame, but for completely different reasons -- interpretations of quantum theory are not even mentioned. Nonetheless, fermion fields are obtained from field theories of type \eqref{field}, and gauge-equivalent states are interpreted as fundamentally different beables, so that no BRST factorization procedure is necessary.
Another part of the physical problem is compatibility with relativistic gravity. Here I argue that it is the general-relativistic concept of background-freedom which is incompatible with quantum theory and has to be given up. I use a quantum variant of the classical hole argument for this purpose \cite{hole}. As a replacement, I propose a theory of gravity with background and preferred frame \cite{glet}.
So there remains only the philosophical part. But here the violation of Bell's inequality gives a strong argument in favour of a preferred frame: Every realistic interpretation needs it. Moreover, the notion of metaphysical realism presupposed by ``realistic interpretation'' is so weak that Norsen \cite{Norsen} has titled a paper ``against realism'', arguing that one should not mention realism at all in this context. The metaphysical notion of realism used there is so weak that to give it up does not save Einstein locality at all -- it is presupposed in this notion too.
\section{Pauli's symmetry argument}\label{Pauli}
There is also another symmetry argument against dBB theory, which goes back to Pauli \cite{Pauli}, which deserves to be mentioned:
\begin{quote}
\ldots the artificial asymmetry introduced in the treatment of the two variables of a canonically conjugated pair
characterizes this form of theory as artificial metaphysics. (\cite{Pauli}, as quoted by \cite{Freire}),
``\ldots the Bohmian corpuscle picks out by fiat a preferred basis (position) \ldots'' \cite{BrownWallace}
\end{quote}
Here my counterargument is presented in \cite{kdv}. I construct there an explicit counterexample, based on the KdV equation, that the Hamilton operator alone, without a specification which operator measures position, is not sufficient to fix the physics. It follows that the canonical operators have to be part of the complete definition of a quantum theory and so have to be distinguished by the interpretation as something special, different from the other similar pairs of operators.
The Copenhagen interpretation makes such a difference -- this is one of the roles played by the classical part. But attempts to get rid of the classical part of the Copenhagen interpretation, without adding something else as a replacement, are not viable \cite{pure}. One has to introduce a replacement.
Recognizing that the configuration space has to be part of the definition of the physics gives more power to an old argument in favour of the pilot wave approach, made already by de Broglie at the Solvay conference 1927:
\begin{quote}
``It seems a little paradoxical to construct a configuration space with the coordinates of points which do not exist.'' \cite{deBroglie}.
\end{quote}
\section{Problems with field theories}\label{fields}
It has been argued that fields are problematic as beables in general for dBB theory, a point which could be problematic for the paleoclassical interpretation too.
In particular, the equivalence proof between quantum theory and dBB theory depends on the fact that the overlap of the wave function for different macroscopic states is irrelevant. But it appeared in field theory that for one-particle states there is always a non-trivial overlap, even if these field states are localized far away from each other.
But, as I have shown in \cite{overlap}, the overlap decreases sufficiently fast (approximately exponentially) with greater particle numbers.
\section{Why we observe configurations, not wave packets}\label{wavepackets}
In the many worlds community there is a quite popular argument against dBB theory -- that it is many worlds in denial (for example, see \cite{BrownWallace}). But this argument depends on the property of dBB theory that the wave function is a beable, a really existing object. So it cannot be applied against the paleoclassical interpretation, where the wave function is no longer a beable.
But in fact it is invalid also as an argument against dBB theory. In fact, already in dBB theory it is the configuration $q(t)$ which is observable and not the wave function.
This fact is sometimes not presented in a clear enough way, so that misrepresentations become possible. For example Brown and Wallace \cite{BrownWallace} find support for another interpretation even in Bohm's original paper \cite{Bohm}:
\begin{quote}
\ldots even in his hidden variables paper II of 1952, Bohm seems to associate the wavepacket chosen by the corpuscles as the representing outcome of the measurement -- the role of the corpuscles merely being to point to it.
(\cite{BrownWallace} p. 15)\end{quote}
and support their claim with the following quote from Bohm
\begin{quote}
Now, the packet entered by the apparatus variable $y$ determines the actual result of the measurement, which the observer will obtain when he looks at the apparatus.
(\cite{Bohm} p. 118)\end{quote}
This quote may, indeed, lead to misunderstandings about this issue. So, maybe we obserse only the wave packet containing the configuration, instead of configuration itself?
My answer is a clear no. I don't believe into the existence of sufficiently localized wave packets to construct some effective reality out of them, as assumed by many worlders.
Today they use decoherence to justify their belief that wave packets will be sufficiently localized. But decoherence presupposes another structure -- a decomposition of the world into systems. Only from point of view of such a subdivision of $q$ into, say, $(x,y)$, a non-localized wave function like $e^{-(x-y)^2/2}$ may look similar to a superposition, for different $a$, of product states localized in $x$ and $y$ like $e^{-(x-a)^2}\cdot e^{-(y-a)^2}$.
But where does this subdivision into systems come from? The systems around us -- observers, planets, measurement devices -- cannot be used for this purpose. They do not exist on the fundamental level, as a predefined structure on the configuration space. But the subdivision into systems has to, once we need it to construct localized objects. Else, the whole construction would be circular.
So one would have to postulate something else as a fundamental subdivision into systems. This something else is undefined in the interpretations considered here, so an interpretation based on it is simply another interpretation, with another, additional fundamental structure -- a fundamental subdivision into systems.
\end{appendix}
|
Subsets and Splits
No community queries yet
The top public SQL queries from the community will appear here once available.