1: \documentclass[aps,pre,amsmath,amsfonts,amssymb,11pt,nofootinbib,showpacs]{revtex4}
2: \usepackage{color,epsfig,graphics,psfrag,rotating,theorem}
3: %\draft
4:
5: \newtheorem{lemma}{Lemma}
6: \newtheorem{remark}{Remark}
7: \newtheorem{propo}{Proposition}
8: \newtheorem{conj}{Conjecture}
9:
10: \newcommand{\eps}{\varepsilon}
11: \newcommand{\s}{\sigma}
12: \newcommand{\<}{\langle}
13: \renewcommand{\>}{\rangle}
14: \newcommand{\sign}{\text{sign}}
15:
16: \def\Tree{{\mathbb T}}
17: \def\T{{\sf T}}
18:
19: \def\E{\mathbb E}
20: \def\prob{{\mathbb P}}
21: \def\var{\text{Var}}
22: \def\ed{\stackrel{\rm d}{=}}
23: \def\ind{{\mathbb I}}
24: \def\P{{\sf P}}
25: \def\M{{\mathfrak M}_q}
26: \def\Q{\widehat{Q}}
27: \def\QQ{\widetilde{Q}}
28: \def\W{\widehat{W}}
29: \def\F{{\sf F}}
30: \def\cA{{\cal A}}
31: \def\pih{\widehat{\pi}}
32: \def\phih{\widehat{\varphi}}
33: \def\eo{\overline{\eta}}
34: \def\bX{{\bf X}}
35:
36: \def\ux{\underline{x}}
37: \def\uX{\underline{X}}
38: \def\uY{\underline{Y}}
39: \def\uy{\underline{y}}
40: \def\de{{\rm d}}
41: \def\cG{{\cal G}}
42: \def\cN{{\cal N}}
43:
44: \def\eh{\hat{\eta}}
45:
46: \def\prooft{\hspace{0.5cm}{\bf Proof:}\hspace{0.1cm}}
47: \def\endproof{\hfill$\Box$\vspace{0.4cm}}
48:
49:
50: \begin{document}
51:
52: \title{Reconstruction on trees and spin glass transition}
53:
54: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
55: \author{Marc M\'ezard}
56:
57: \affiliation{Laboratoire de Physique Th\'eorique et Mod\`eles Statistiques,
58: Universit\'e de Paris-Sud, b\^atiment 100, 91405, Orsay Cedex, France}
59:
60: \author{Andrea Montanari}
61:
62: \affiliation{Laboratoire de Physique Th\'eorique de l'Ecole Normale
63: Sup\'erieure,\\
64: 24 rue Lhomond 75231 Paris Cedex 05, France}
65:
66:
67: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
68:
69: \begin{abstract}
70: Consider an information source generating a symbol at the
71: root of a tree network whose links correspond to noisy
72: communication channels, and broadcasting it through the network.
73: We study the problem of reconstructing the transmitted
74: symbol from the information received at the leaves.
75: In the large system limit, reconstruction is possible when the
76: channel noise is smaller than a threshold.
77:
78: We show that this threshold coincides with the dynamical
79: (replica symmetry breaking) glass transition for an associated
80: statistical physics problem. Motivated by this correspondence,
81: we derive a variational principle which implies new rigorous bounds
82: on the reconstruction threshold. Finally, we apply a standard
83: numerical procedure used in statistical physics,
84: to predict the reconstruction thresholds in various channels.
85: In particular, we prove a bound on the
86: reconstruction problem for the antiferromagnetic ``Potts'' channels, which
87: implies, in the noiseless limit, new results on random proper colorings of
88: infinite regular trees.
89:
90: This relation to the reconstruction problem also offers interesting perspective for
91: putting on a clean mathematical basis the theory of glasses on random graphs.
92: \end{abstract}
93:
94: \pacs{02.50.-r
95: (Probability theory, stochastic processes, and statistics),
96: 64.70.Pf (Glass transitions), 89.75.Hc
97: (Networks and genealogical trees)}
98:
99: \maketitle
100:
101: \section{Introduction}
102: \label{sec:intro}
103:
104: Consider the following broadcast problem~\cite{Mossel03}.
105: An information source at the root of a tree network
106: produces a letter taken from a $q$-ary alphabet
107: $x\in \{1,\dots,q\}$ (we shall sometimes refer to a letter
108: from this alphabet as to a `color').
109: The symbol is propagated along the edges of the tree.
110: For simplicity we start with a regular $k$-ary tree $\Tree_k$,
111: cf. Fig.~\ref{fig:broadcast},
112: in which every vertex has exactly $k$ descendants
113: (every vertex has degree $k+1$ except the root which
114: has degree $k$), a more general setting is described in~\cite{EvaKenPerSch} and in Sect.~\ref{se:gene}.
115: Each edge of the tree is an instance of the
116: same noisy communication channel: If the letter $x$ is transmitted
117: through the channel, $y\in\{1,\dots,q\}$ is received with
118: probability $\pi(y|x)$ (with $\pi(y|x)\ge 0$ and $\sum_y\pi(y|x)=1$).
119: The problem of reconstruction is the following: consider all the
120: symbols received at the vertices of the $\ell^{\rm th}$ generation. Does this
121: configuration contain a non-vanishing information on the letter transmitted
122: by the root, in the large $\ell$ limit?
123:
124: Beyond its fundamental interest in
125: probability, this problem is relevant to genetics (propagation of genes
126: from an ancestor) \cite{MosselSteel05},
127: to statistical physics (models on Bethe lattices),
128: and information theory (the problem being equivalent to computing
129: the information capacity of the tree network)
130: \cite{CoverThomas}.
131:
132: An important general bound was obtained by Kesten and Stigum (KS)
133: \cite{KestenStigum1,KestenStigum2}.
134: Consider the matrix $\pi$ with entries $\pi(y|x)$,
135: $x,y\in\{1,\dots,q\}$ and let $\lambda_2(\pi)$ be its
136: eigenvalue with the second largest absolute value. Then, if $k
137: |\lambda_2(\pi) |^2 >1$, the reconstruction problem is solvable: the leaves
138: asymptotically contain some information on the letter sent by the root.
139: In fact in this case the census of the variables in the $\ell^{\rm th}$
140: generation (the number of leaves which have received each letter)
141: contains some information on the root.
142: Conversely, if $k | \lambda_2(\pi)|^2 <1$,
143: the census contains asymptotically no information on the root
144: \cite{MosselPeres01}.
145: Therefore, the
146: KS condition $k \vert \lambda_2(\pi)\vert ^2=1$ defines a
147: threshold for the maximum amount of noise allowing census
148: reconstruction. For larger noise ( $k \vert \lambda_2(\pi)\vert ^2 <1$)
149: one may wonder whether reconstruction is possible exploiting the
150: whole set of symbols received at the
151: $\ell^{\rm th}$ generation, through a clever use of the correlations
152: between the symbols received on the leaves. The answer depends
153: on the channel.
154:
155: %
156: \begin{figure}
157: \includegraphics[width=0.4\linewidth,angle=-0]{broadcast.eps}\hspace{1.5 cm}
158: \includegraphics[width=0.38\linewidth,angle=-0]{reconstr.eps}
159: \caption{Left: Broadcast on a tree. The signal is sent from the root. Each
160: edge is a noisy communication channel broadcasting upwards. Right: The reconstruction problem asks to find what
161: signal was sent from the root, given the signals received on the leaves }
162: \label{fig:broadcast}
163: \end{figure}
164: %
165:
166: In most of this paper we shall focus onto
167: transition kernels $\pi(\,\cdot\,|\,\cdot\,)$ satisfying
168: the detailed balance condition (reversible)
169: with respect to the uniform distribution $\eo(x) = 1/q$.
170: In other words $\pi(y|x) = \pi(x|y)$. With a slight
171: abuse of notation we shall write $\pi(y|x) = \pi(y,x)$.
172: For the problem to be non-trivial, we also assume
173: $\pi(\,\cdot\,|\,\cdot\,)$ to be irreducible and aperiodic.
174: A particularly important example in this family is provided by
175: $q$-ary symmetric channels (or, borrowing from the statistical mechanics
176: terminology, `Potts' channels)
177: %
178: \begin{equation}
179: %
180: \pi(y|x) = \left\{\begin{array}{ll} 1-\eps & \;\; \mbox{if $y=x$.} \\
181: \eps/(q-1) & \;\; \mbox{otherwise}\end{array}
182: \right. \ .\label{eq:PottsChannelDef}
183: %
184: \end{equation}
185: %
186: If $\eps<1-1/q$, $y=x$ is the most likely channel output when the input
187: is $x$: we shall refer to this case as the `ferromagnetic' Potts channel.
188: If $\eps>1-1/q$, the opposite happens and we shall speak of
189: `antiferromagnetic' Potts channel. The particular case
190: $\eps=1$ is of special interest, since the broadcast
191: process provides a uniformly random proper coloring of the
192: $\ell$-generations $k$-ary tree $\Tree_k(\ell)$.
193:
194: It is intuitively clear that the channel (\ref{eq:PottsChannelDef})
195: `gets worse' as $\eps$ increases from $0$ to $1-1/q$ (ferromagnetic channel) and
196: `improves' as $\eps$ goes from $1-1/q$ to $1$ (antiferromagnetic
197: channel).
198: A result by Mossel~\cite{Mossel01} implies that there exist
199: a ferromagnetic and an antiferromagnetic threshold,
200: respectively $\eps^+_{\rm r}(k,q)\in[0,1-1/q]$ and
201: $\eps^-_{\rm r}(k,q)\in[1-1/q,1]$,
202: such that the reconstruction problem is solvable when
203: $\eps \in [0,\eps^+_{\rm r}[\; \cup\; ]\eps^-_{\rm r},1]$ and insolvable
204: if $\eps \in \; ]\eps^+_{\rm r},\eps^-_{\rm r}[$.
205: Hereafter we shall drop the $\pm$ superscripts whenever they are clear from the context.
206:
207: The KS condition $k|\lambda_2(\pi)|^2>1$ is satisfied
208: (and the problem is census-solvable) for the channel
209: (\ref{eq:PottsChannelDef}) if and only if
210: $\eps\in [0,\eps_{\rm KS}^+(k,q)[\; \cup\; ]\eps_{\rm KS}^-(k,q),1]$,
211: where:
212: %
213: \begin{eqnarray}
214: %
215: \eps_{\rm KS}^{\pm}(k,q) &=& \frac{q-1}{q} \; \left(1\mp
216: \frac{1}{\sqrt{k}}\right)\, .
217: %
218: \end{eqnarray}
219: %
220: Notice that the above formula yields $\eps_{\rm KS}^{-}(k,q)>1$ for some pairs
221: of $(k,q)$. In fact, for the antiferromagnetic channel, the
222: census-reconstruction problem (as well as the general reconstruction problem)
223: is not necessarily solvable for $\eps=1$.
224:
225: It is known \cite{BleherRuizZagrebnov} that, for $q=2$ (the ``binary
226: symmetric'' channel, also known as the ``symmetric Ising'' case), the
227: reconstruction threshold is equal to the KS one:
228: $\eps_{\rm r}(k,2)=\eps_{\rm KS}(k,2)$ (for $q=2$ the ferromagnetic
229: and antiferromagnetic cases are equivalent via the mapping
230: $\eps\mapsto 1-\eps$). In general, the KS bound implies
231: $\eps^{+}_{\rm KS}(k,q)\ge \eps_{\rm r}^{+}(k,q)$, and
232: $\eps^{-}_{\rm r}(k,q)\ge \eps_{\rm KS}^{-}(k,q)$.
233: Furthermore, in \cite{Mossel01} it was shown that, for all $k$, when
234: $q$ is large enough, $\eps^{+}_{\rm KS}(k,q)> \eps_{\rm r}^{+}(k,q)$
235: strictly: reconstruction is possible at noise levels where
236: census reconstruction does not work.
237: However, several fundamental
238: questions remain open even for simple Potts channels:
239: Is there any pair $(k,q)$, with $q>2$, such that
240: $\eps_{\rm r}(k,q)=\eps_{\rm KS}(k,q)$? How to distinguish systematically
241: between $\eps_{\rm r}(k,q)$ and $\eps_{\rm KS}(k,q)$? How to determine
242: $\eps_{\rm r}(k,q)$ accurately when it does not coincide with
243: $\eps_{\rm KS}(k,q)$? We shall address these issues in the following.
244:
245: The reconstruction problem is intimately related to statistical physics.
246: Consider a model of Potts spins $y_i \in \{1,\dots,q\}$,
247: on a finite rooted tree with $\ell$ generations, to be denoted by
248: $\Tree_k(\ell)$.
249: Suppose that the energy of a configuration $\uy^{\ell}\equiv \{y_i :
250: i\in \Tree_k(\ell)\}$ is given by:
251: %
252: \begin{equation}
253: %
254: E(\uy^{\ell})= - J \!\!\sum_{(i,j)\in \Tree_k(\ell)}\delta_{y_i, y_j} \ ,
255: \label{eq:PottsEnergy}
256: %
257: \end{equation}
258: %
259: where $(i,j)$ denotes pairs of spins connected by an edge of the
260: tree. Let $\uY^{\ell}$ be the random configuration produced by the broadcast
261: process with channel (\ref{eq:PottsChannelDef}) {\it up to} generation
262: $\ell$,
263: when the transmitted symbol is uniformly random in $\{1,\dots,q\}$.
264: Then
265: %
266: \begin{eqnarray}
267: %
268: \prob\left\{\uY^{\ell}=\uy^{\ell}\right\}
269: =\frac{1}{Z}\, \exp\{-\beta E(\uy^{\ell})\}
270: %
271: \end{eqnarray}
272: %
273: provided we make the identification
274: %
275: \begin{equation}
276: %
277: e^{-\beta J}=\frac{\eps}{(q-1) (1-\eps)}\ .
278: %
279: \end{equation}
280: %
281: In other words, the broadcast process allows to construct one
282: particular Gibbs measure (state) associated to the energy
283: function (\ref{eq:PottsEnergy}): in the
284: statistical physics terminology this is the free-boundary measure.
285: In general this is not the unique Gibbs measure for
286: this energy function. For instance, if $\eps< \frac{q-1}{q} \;
287: \left(1-\frac{1}{k}\right)$, one can construct $q$ `ferromagnetic' states
288: as well. Even if more than one Gibbs state exists, the free-boundary
289: state can be extremal (or `pure'). It turns out that
290: the reconstruction problem is
291: solvable if and only if the Gibbs state with free boundary conditions
292: is not extremal.
293:
294: Given the strong connection between extremality of Gibbs states and
295: spatial decay of correlations \cite{Georgii}, the last remark is not surprising.
296: What is more surprising (and constitutes the main theme of this paper)
297: is the relation of the reconstructibility with the existence of
298: a dynamical glass phase. In recent years, an ongoing effort has been
299: devoted to the study of glassy models on sparse random graphs.
300: These are graphs which contain cycles but locally `look like' a tree
301: (e.g. uniformly random graphs with given degree).
302: One of the most widespread features of these models, is the
303: occurrence of glass phases in which the Boltzmann measure gets split
304: into an exponential number of `lumps' (also referred to as clusters or
305: pure states). This phenomenon is usually studied by solving
306: some `one-step replica symmetry breaking' (1RSB) distributional equations.
307: In the following we show that these equations, as well as the criterion
308: used to detect glass phases, do indeed coincide
309: with the solvability of an appropriate reconstruction problem.
310:
311: In spin glass theory, one can encounter
312: two types of transitions to a glass phase. In the first
313: case the transition is continuous in a properly defined order parameter.
314: In spin glass jargon this leads to a
315: phase with `full replica symmetry breaking' (FRSB).
316: In the second it is discontinuous, leading to 1RSB.
317: Both situations occur in the reconstruction problem, depending on the alphabet
318: and the channel. In the continuous case, the phase transition
319: location is given by a local instability which coincides with the
320: KS threshold, and one has $\eps_{\rm r}(k,q)=\eps_{\rm KS}(k,q)$.
321: This happens, for instance, when $q=2$.
322: In the opposite case, the `dynamical' glass transition is discontinuous and its
323: location (which still coincides with the reconstruction threshold)
324: is distinct from the KS one. In the ferromagnetic Potts model one has,
325: for instance, $\eps_r(k,q)>\eps_{KS}(k,q)$ at large enough $q$.
326:
327: The coincidence of the reconstruction threshold with the dynamical
328: glass transition, apart from being interesting in itself, allows us
329: to adapt several techniques developed within the theory of spin glasses
330: in order to study the reconstruction problem.
331: On the one hand, importing a numerical procedure currently used in this field, we
332: determine the threshold for several pairs $k,q$. These results
333: lead us to conjecture that $\eps_{\rm r}(k,q)=\eps_{\rm KS}(k,q)$,
334: for $k$ not too large and $q\le 4$ (in the ferromagnetic case) or $q\le 3$
335: (in the antiferromagnetic case).
336:
337: Furthermore, we derive a variational principle for the reconstruction problem.
338: In the antiferromagnetic case, this implies
339: a rigorous bound on the reconstruction threshold, which allows to confirm
340: the strict inequality $\eps^{-}_{\rm r}(k,q)<\eps^{-}_{\rm KS}(k,q)$
341: in most of the cases in which this was found to be the case numerically.
342: Although we conjecture such a bound to hold in much greater generality,
343: we weren't able to prove it, and we leave it as a conjecture.
344:
345: The paper is organized as follows. In Section~\ref{sec:Recursion}
346: we define the main objects studied in the paper and prove the coincidence
347: between reconstruction and dynamical glass transition.
348: In Section \ref{se:varprinc} we state our variational principle
349: and prove that it provides a rigorous bound for a class of
350: kernels $\pi(\,\cdot\,|\,\cdot\,)$ including the antiferromagnetic model.
351: In Secs. \ref{sec:Ferro} and \ref{sec:AntiFerro} we apply this principle
352: as well as a numerical procedure to the determination of thresholds for
353: the Potts channel, respectively in the ferromagnetic and antiferromagnetic case.
354: Section \ref{se:spinglass} discusses the physical meaning of
355: the relation between reconstruction and glass transitions.
356: Section \ref{se:gene} explains how our methods (and the glass - reconstruction correpondance)
357: can be generalized to a broad
358: category of broadcast and reconstruction problems on trees, going much beyond the Potts channel.
359: We conclude in Section \ref{sec:Conclusion} by summarizing a few
360: conjectures and pointing out some interesting open problems.
361: %
362: %***************************************************************************
363: %
364: \section{Distributional recursion}
365: \label{sec:Recursion}
366:
367: \subsection{Definitions}
368:
369: We denote by $V$ and $E$ the vertex and edge sets of the infinite
370: $k$-ary tree $\Tree_k$, by $0$ its root and by $V_{\ell}$ the set of
371: generation-$\ell$ vertices ($|V_{\ell}|= k^{\ell}$). The broadcast process
372: generates a random color
373: configuration $\uX\equiv\{ X_i\; :\;i\in V\}$ with $X_i\in \{1,\dots, q\}$.
374: The root color $X_0 \in \{1,\dots, q\}$, which we also call the transmitted color, is uniformly random.
375: Then, given the values
376: of $X$ up to the $\ell$-th generation, the values at the $(\ell+1)$-th
377: generation are conditionally independent. If a vertex in the $\ell$-th
378: generation has color $y$, the probability that a vertex connected to it in the
379: $(\ell+1)$-th generation has color $z$ is $\pi(z|y)$.
380:
381: We shall denote by $\uX_{\ell}$ the configurations of
382: colors at the $\ell^{\rm th}$ generation, and by
383: $\uY^{\ell}$ the configuration {\em up to} the $\ell^{\rm th}$ generation
384: (i.e. $\uY^{\ell} = \{\uX_0,\uX_1,\dots,\uX_{\ell}\}$).
385: The probability distribution of $\uX_{\ell}$, conditioned to the choice
386: $X_0=x$ of the root color will be denoted by
387: $B^{(\ell)}_x(\ux_\ell)\equiv \prob\{\uX_{\ell} = \ux_{\ell}|X_0=x\}$.
388:
389: Suppose now that the configuration of colors at the $\ell$-th generation,
390: $\ux_{\ell}$, is given. We denote by $\eta_{\ell}(y)$
391: the probability that the root had sent the color $y$, given $\uX_{\ell}$:
392: %
393: \begin{eqnarray}
394: %
395: \eta_{\ell}(y) = \prob[X_0 = y\, |\uX_\ell=\ux_{\ell}]\, .\label{eq:DefEta}
396: %
397: \end{eqnarray}
398: %
399: $\eta_{\ell}(\cdot)$ is a probability distribution over $\{1,\dots, q\}$ (i.e.
400: $\eta_{\ell}(y)\ge 0$ and $\sum_y\eta_{\ell}(y) = 1$). We shall denote the
401: space of such distributions as $\M$.
402: In order to emphasize the dependency of $\eta_{\ell}$ upon the configuration
403: received in shell $\ell$, $\ux_{\ell}$,
404: we shall sometimes write $\eta_{\ell}(y) =\eta_{\ux_\ell}(y) $.
405: It is easy to realize that, given the colors received
406: at the $\ell^{\rm th}$ generation, $\eta_{\ell}(\cdot)$ constitutes
407: a sufficient statistics for the root color $x$. In other terms,
408: given $\uX_{\ell}=\ux_\ell$, there is no loss of information in computing
409: $\eta_{\ux_\ell}(\cdot)$ and then guessing $X_0$ from $\eta_{\ux_\ell}(\cdot)$.
410:
411: Since $\uX_\ell$ is chosen randomly
412: according to the broadcast process, $\eta_{\ell}(\cdot)$ is a
413: random probability distribution, i.e. a random point in $\M$. We
414: denote by $Q^{(\ell)}_{x}(\eta)$ its
415: distribution\footnote{Notice that we adopt here the standard physicists
416: convention: we carelessly denote probability distributions by their densities
417: even if such densities do not exist. We shall also denote by
418: $\int \!f(\eta)\, \de Q^{(\ell)}_x(\eta)$ the expectation with respect to
419: such a distribution. The fussy reader can easily translate
420: all the formulae below in the standard probability language.} conditional to
421: the broadcast being started from $X_0=x$, and call it the `distribution at the root'.
422: Hereafter a distribution $Q$ over $\M$ will be said {\em trivial} if it is a
423: singleton on the uniform measure $\eo$ (defined by $\forall x:\ \eo(x) = 1/q$).
424: Clearly the reconstruction problem is solvable if and only if the large
425: $\ell$ limit of $Q_{x}^{(\ell)}$ is non trivial.
426:
427: There are several ways of characterizing quantitatively the large
428: $\ell$ behavior of $Q_{x}^{(\ell)}$. We shall consider below two parameters
429: $I_{\ell} \equiv I(X_0;\uX_\ell)$ and
430: $\Psi_{\ell}$, which are defined by:
431: %
432: \begin{eqnarray}
433: %
434: I_{\ell} = \frac{1}{q}\sum_x\int\log_2\frac{\eta(x)}{\eo(x)}\;
435: \de Q_x^{(\ell)}(\eta)\, ,\;\;\;\;
436: \Psi_{\ell} = \frac{1}{q}\sum_x\int [\eta(x)-\eo(x)]\;\de
437: Q_x^{(\ell)}(\eta)\, .\
438: %
439: \end{eqnarray}
440: %
441:
442: $I_{\ell}$ gives the number of information bits that can be transmitted reliably
443: per network use. $\Psi_{\ell}$ is the probability that the reconstruction
444: is successful when the receiver guesses color $y$ with probability
445: $\eta_{\ell}(y)$, minus the the same probability when the receiver
446: guesses uniformly.
447: These are
448: non-negative quantities and can be shown to be non-increasing
449: functions of $\ell$. We furthermore let $I_{\infty}\equiv\lim_{\ell\to\infty} I_{\ell}$,
450: and $\Psi_{\infty}\equiv\lim_{\ell\to\infty} \Psi_{\ell}$.
451:
452: The tree reconstruction problem can be rephrased by saying that
453: the problem is solvable if and only if $I_{\infty}>0$ (or, equivalently,
454: $\Psi_{\infty}>0$). For instance, for the ferromagnetic
455: Potts channel, the threshold
456: $\eps^+_{\rm r}(k,q)$ is the
457: supremum of the values of $\eps$ such that $I_{\infty}>0$.
458: %
459: %*************************************************************************
460: %
461: \subsection{Merging rooted trees}
462: \label{se:merging}
463:
464: How does one compute the distribution $\eta(y)$ on the root, given a boundary
465: $\uX_{\ell}=\ux_{\ell}$?
466: Using the tree-structure, this can be done iteratively by a
467: dynamical programming procedure starting from the leaves. Suppose that at some
468: point in this iteration we have determined the probability distributions
469: $\eta_1(\,\cdot\,),\dots,\eta_k(\,\cdot\,)$ of the $k$ vertices in the
470: tree which lie above a given vertex (see Fig.~\ref{fig:iter}, left).
471: Then the probability $\eta(y)$ that this vertex had color $y$
472: during the broadcast is given by:
473: %
474: \begin{eqnarray}
475: %
476: \eta(y) = \frac{1}{z(\{\eta_i\})}\,
477: \prod_{i=1}^k\left(\sum_{y_i=1}^q\pi(y_i|y)\, \eta_i(y_i)\right)\, ,\;\;
478: \;\;\;\;\;\; z(\{\eta_i\}) \equiv \sum_{y=1}^q \prod_{i=1}^k\left(\sum_{y_i}\pi(y_i|y)\, \eta_i(y_i)\right)\, .
479: \label{eq:Site}
480: %
481: \end{eqnarray}
482: %
483: This equation defines a mapping between distributions in $\M$: given $k$
484: distributions $\eta_1,\dots,\eta_k $, one generates a new one $\eta =
485: \F(\eta_1,\dots,\eta_k)$. Iterating this mapping
486: downwards from the leaves down to the root, one can derive the the conditional
487: distribution of the transmitted symbol.
488: %
489: %
490: \begin{figure}
491: \includegraphics[width=0.3\linewidth,angle=-0]{iter.eps}\hspace{1.8 cm}
492: \includegraphics[width=0.5\linewidth,angle=-0]{funct_rec.eps}
493: \caption{Left: a pictorial representation of the mapping $\eta =
494: \F(\eta_1,\dots,\eta_k)$ defined in (\ref{eq:Site}). Here $k=2$,
495: a straight line corresponds to the channel $\pi$, a wiggly line arriving on
496: a vertex $y_j$ corresponds to a weight $\eta_j(y_j)$. Right: a pictorial representation of the
497: recursion (\ref{eq:Iteration}). A triangle of depth $r$ rooted on variable $x_j$ denotes $Q_{x_j}^{(r)}(\eta_r)$}
498: \label{fig:iter}
499: \end{figure}
500: %
501:
502: Equation (\ref{eq:Site}) naturally induces a recursion equation for the
503: distribution $Q_{x}^{(\ell)}$.
504: Consider the reconstruction of the root in a rooted tree with $\ell+1$
505: generations (see Fig.~\ref{fig:iter}, right). This graph is formed by $k$ subtrees rooted in the vertices
506: $1,\dots,k$, which are all joined to the root $0$.
507: Each of these subtrees gives an instance of the reconstruction with $\ell$
508: generations.
509: Therefore:
510: %
511: \begin{eqnarray}
512: %
513: Q_{x}^{(\ell+1)} (\eta)= \sum_{x_1\dots x_k}\prod_{i=1}^k\pi(x_i|x)
514: \int \; \delta\left[\eta- \F(\eta_1,\dots,\eta_k)\right]\;
515: \prod_{i=1}^k \de Q^{(\ell)}_{x_i}(\eta_i)
516: \, ,\label{eq:Iteration}
517: %
518: \end{eqnarray}
519: %
520: where $\delta[\cdots]$ represents a Dirac delta function on $\M$.
521: In words, in order to generate $\eta(\,\cdot\, )$ with distribution
522: $Q_{x}^{(\ell+1)}$, one can proceed as follows.
523: First draw $k$ independent colors $x_1,\dots,x_k$
524: from the distribution $\pi(\,\cdot\,|x)$.
525: Then, draw $\eta_1,\dots,\eta_k$ independently with distribution, respectively,
526: $Q_{x_1}^{(\ell)},\dots,Q_{x_k}^{(\ell)}$. Finally,
527: let $\eta = \F(\eta_1,\dots,\eta_k)$.
528:
529: The initial condition is
530: %
531: \begin{eqnarray}
532: %
533: Q^{(0)}_x(\eta) = \delta\left[\eta-\delta_x\right]\, ,\label{eq:Initial}
534: %
535: \end{eqnarray}
536: %
537: where $\delta_x$ is the distribution in $\M$ which has weight unity on color
538: $x$ (it is given by $\delta_x(y)=1$ if $y=x$, and $\delta_x(y)=0$ otherwise).
539: The equations (\ref{eq:Iteration}) and (\ref{eq:Initial}) fully characterize
540: the distributions $Q^{(\ell)}_x$. The whole reconstruction problem amounts to
541: understanding the large $\ell$ properties of these recursions.
542:
543: %
544: %*************************************************************************
545: %
546: \subsection{Unconditional distribution and symmetry properties}
547:
548: While $Q^{(\ell)}_x$ gives the distribution of $\eta_{\ell}(\, \cdot\, )$
549: (defined in Eq.~(\ref{eq:DefEta})) {\em conditional} on the
550: transmitted color being equal to $x$, it is equally interesting to consider the
551: {\em unconditional} distribution. We will denote it by $\Q^{(\ell)}$.
552: Bayes theorem implies the following relation between
553: $Q^{(\ell)}_x$ and $\Q^{(\ell)}$
554: %
555: \begin{eqnarray}
556: %
557: Q^{(\ell)}_{x}(\eta) = q\, \eta(x)\, \Q^{(\ell)}(\eta)\, .
558: \label{eq:SymmetricRepresentation}
559: %
560: \end{eqnarray}
561: %
562: This is in fact a rephrasing of the identity
563: %
564: \begin{eqnarray}
565: %
566: \prob\{ \eta_{\uX_\ell}=\eta |X_0=x\} =
567: \frac{\prob\{X_0=x|\eta_{\uX_\ell}=\eta\}
568: \prob\{\eta_{\uX_\ell}=\eta\}}{\prob\{X_0=x\}}\, .
569: %
570: \end{eqnarray}
571: %
572:
573: An alternative (analytic) proof can be obtained writing $Q^{(\ell)}_x$
574: in terms of $B^{(\ell)}_x(\ux_\ell)$, the probability that the output of the
575: broadcast process at generation $\ell$ is $\ux_{\ell}$, given that the
576: transmitted color is $x$:
577: %
578: \begin{eqnarray}
579: %
580: Q^{(\ell)}_x(\eta) = \sum_{\ux_{\ell}} B^{(\ell)}_x(\ux_\ell)\;\; \delta\!\left[
581: \eta(\cdot)-
582: \frac{B^{(\ell)}_\cdot(\ux_\ell)}{\sum_z B^{(\ell)}_z(\ux_\ell)}\right]
583: \, .\label{RepresentationSymm}
584: %
585: \end{eqnarray}
586: %
587: It is then easy to show that, if $\lambda_0+\lambda_1+\dots +\lambda_{q-1}=1$,
588: then the expectation value
589: %
590: \begin{eqnarray}
591: %
592: \int \; \frac{\eta(0)^{\lambda_0}\cdots \eta(q-1)^{\lambda_{q-1}}}
593: {\eta(x)} \; \de Q^{(\ell)}_x(\eta) \,
594: %
595: \end{eqnarray}
596: %
597: does not depend upon $x$. This in turns imply that $Q_x^{(\ell)}$
598: can be written in the form (\ref{eq:SymmetricRepresentation})
599: where $\Q^{(\ell)}$ is a distribution which does not depend on $x$
600: (the normalization can be found by summing over $x$).
601:
602: If the channel is symmetric with respect to permutations
603: of the colors (as is the case for Potts channels), the distributions
604: $Q_x^{(\ell)}$ and $Q^{(\ell)}$ inherit the same symmetry.
605: More precisely, given a permutation acting on the colors
606: $\sigma\in S_q$, and a distribution $\eta\in\M$, let
607: $\eta^{\sigma}$ be the permuted distribution defined by
608: $\eta^{\sigma}(x) \equiv \eta(\sigma(x))$. Then, for any permutation $\sigma$,
609: $Q^{(\ell)}_x(\eta)=Q^{(\ell)}_{\sigma(x)}(\eta^{\sigma})$, and
610: %
611: \begin{eqnarray}
612: %
613: \Q^{(\ell)}(\eta^{\sigma}) = \Q^{(\ell)}(\eta)\, . \label{eq:Symmetry}
614: %
615: \end{eqnarray}
616: %
617: A distribution satisfying
618: condition (\ref{eq:Symmetry}) will be called `symmetric'.
619:
620: Let us finally notice that the parameters introduced in
621: Sec.~\ref{sec:Recursion} to measure the amount of information on the transmitted color
622: available at the $\ell^{\rm th}$ generation can be
623: expressed in terms of the distribution $\Q^{(\ell)}$
624: %
625: \begin{eqnarray}
626: %
627: I_{\ell} = \int D(\eta||\eo)\; \de \Q^{(\ell)}(\eta)\, ,\;\;\;\;
628: \Psi_{\ell} = \sum_x\int [\eta(x)-\eo(x)]^2\;\de
629: \Q^{(\ell)}(\eta)\, .
630: %
631: \end{eqnarray}
632: %
633: Here we use the standard notation for the Kullback-Leibler distance
634: \cite{CoverThomas} $D(\eta||\eo) \equiv \sum_x \eta(x)\log_2[\eta(x)/\eo(x)]$.
635: In deriving the second of these expressions, we used the fact that
636: $\int \eta(x)\, \de\Q^{(\ell)}(\eta) = \eo(x)=\frac{1}{q}$ which follows
637: from (\ref{eq:SymmetricRepresentation}).
638: %
639: %******************************************************************
640: %
641: \subsection{Recursion for the unconditional distribution and
642: spin glass correspondence}
643:
644: The recursion relation (\ref{eq:Iteration}) on $Q^{(\ell)}_x $
645: implies the following recursion for the unconditional
646: distribution:
647: %
648: \begin{eqnarray}
649: %
650: \Q^{(\ell+1)} (\eta)= q^{k-1}\;
651: \int \;
652: z(\{\eta_i\})\;\; \delta\left[\eta- \F(\eta_1,\dots,\eta_k)\right]\;
653: \prod_{i=1}^k \de \Q^{(\ell)} (\eta_i)
654: \, ,\label{eq:1RSB}
655: %
656: \end{eqnarray}
657: %
658: where $z(\{\eta_i\} )$ is defined as in Eq.~(\ref{eq:Site}).
659: The initial condition (\ref{eq:Initial}) converts into
660: $\Q^{(0)}(\eta)=\frac{1}{q}
661: \sum_{y=1}^q \delta\left[\eta(\,\cdot\,)-\delta_y(\,\cdot\,)\right]$.
662: It is also interesting to study the fixed points of this recursion,
663: i.e. the distributions $\Q^{*}$ satisfying:
664: %
665: \begin{eqnarray}
666: %
667: \Q^{*} (\eta)= q^{k-1}\;
668: \int \;
669: z(\{\eta_i\})\;\; \delta\left[\eta- \F(\eta_1,\dots,\eta_k)\right]\;
670: \prod_{i=1}^k \de \Q^{*} (\eta_i)
671: \, ,\label{eq:1RSB_fixed}
672: %
673: \end{eqnarray}
674: %
675: Notice that any solution of this equation has necessarily
676: expectation $\int\,\eta(x)\,\de\Q^*(\eta) = \eo(x)$ (this is proved by
677: taking expectation on both sides). Any probability distribution
678: over $\M$ satisfying this condition will be hereafter said to be
679: `consistent'.
680:
681: The distributional equation (\ref{eq:1RSB_fixed}) is well known in
682: spin glass theory and usually referred to as `1RSB equation with Parisi parameter $m=1$'
683: (in the general 1RSB scheme the factor $z(\{\eta_i\})$ is raised to a power $m \in [0,1]$).
684: It is used to determine whether an associated statistical mechanics model
685: is in a glass phase. We shall return to the definition of the
686: associated model in Sec.~\ref{se:spinglass}. For the time being,
687: we shall adopt the usual physicists criterion as a {\em definition}:
688: We will say that the statistical mechanics model associated
689: to the reconstruction problem (characterized by a degree/kernel pair $k,\pi$) {\em admits a glass phase}
690: if and only if Eq.~(\ref{eq:1RSB_fixed}) has a non-trivial solution.
691:
692: When considering a continuous family of kernels $\pi(\,\cdot\,|\, \cdot\,)$,
693: parametrized by a noise level $\eps$,
694: the value of $\eps$ where a non-trivial solution appears is
695: called a dynamical glass transition.
696: The result below implies that this coincides indeed with the
697: reconstruction threshold (i.e. with the extremality threshold for the free
698: boundary Gibbs measure on the infinite tree).
699: %
700: \begin{propo}\label{propo:Reco}
701: The statistical mechanics model associated
702: with the degree/kernel pair $k,\pi$ {\em admits a glass phase}, if
703: and only if the corresponding reconstruction problem is solvable.
704: \end{propo}
705: %
706: \prooft
707: As noticed for instance in \cite{Winkler03}, the sequence of
708: random variables $\eta_{\ell}(\, \cdot\, )$ (not conditioned on the root
709: color), converges almost surely to a limit $\eta_{\infty}(\, \cdot\, )$.
710: As a consequence, the sequence of distributions $\Q^{(\ell)}$ converges
711: weakly to the distribution $\Q^{(\infty)}$ of $\eta_{\infty}(\,\cdot\,)$.
712: By taking the limit of Eq.~(\ref{eq:1RSB}) (and noticing that
713: $\F(\eta_1,\cdots,\eta_k)$ and $z(\{\eta_i\})$ are continuous
714: and bounded) we find that $\Q^{(\infty)}$ must satisfy the fixed
715: point condition (\ref{eq:1RSB_fixed}).
716: If the reconstruction problem is solvable, then $\Q^{(\infty)}$ is
717: non-trivial and therefore, according to our definition,
718: the pair $k,\pi$ {\em admits a glass phase}.
719:
720: Conversely\footnote{The idea of the converse is due to James Martin
721: who kindly agreed to let us publish it here.}, let $\Q^*$
722: be a non-trivial solution of (\ref{eq:1RSB_fixed}). Following
723: (\ref{eq:SymmetricRepresentation}), define the distribution
724: $Q^*_x(\eta) = q\, \eta(x) \Q^*(\eta)$. Because of the above calculations,
725: the $q$ distributions $Q^*_x$, $x\in\{1,\dots,q\}$
726: are a fixed point of the recursion (\ref{eq:Iteration}).
727: We will now show that they
728: can be used to reconstruct the transmitted color from the
729: output at generation $\ell$, with probability of success independent
730: of $\ell$ and strictly larger than $1/q$.
731:
732: The reconstruction procedure goes as follows.
733: Suppose that the broadcast has generated the values $X_i=x_i$ for $i\in
734: V_\ell$. For each vertex $i$, generate $\eta_i$ from the
735: distribution $Q^*_{x_i}$. Consider now a vertex $a \in
736: V_{\ell-1}$,connected to
737: $a_1,\dots,a_k$ in $V_\ell$.
738: Compute $\eta_a= \F(\eta_{a_1},\dots,\eta_{a_k})$, where $\F(\cdots)$ is
739: defined as in Eq.~(\ref{eq:Site}). Proceeding downwards from the leaves
740: to the root, this allows to construct $\eta_0$. At this point
741: the transmitted symbol can be guessed, for instance, by choosing
742: $X_0=y$ with probability proportional to $\eta_0(y)$.
743:
744: We claim that for each vertex $j\in \Tree_k(\ell)$, and conditional to the broadcast having produced
745: $X_j=x_j$, the $\eta_j(\,\cdot\,)$ provided by the above procedure is
746: distributed according to $Q^*_{x_j}$. This in particular implies
747: that the probability of guessing correctly the root color is
748: %
749: \begin{eqnarray}
750: %
751: \frac{1}{q}\sum_x\int \eta(x)\; \de Q_x^*(\eta) =
752: \sum_x\int \eta(x)^2\; \de \Q^*(\eta) >\frac{1}{q}\, .
753: %
754: \end{eqnarray}
755: %
756: The claim is proved by induction starting from the leaves
757: and proceeding downwards to the root. It is true by construction for
758: the vertices of the last generation. Assume it to be true
759: up to generation $r$ and consider a site $a$ in generation
760: $r-1$ connected to $a_1,\dots,a_k$ in $V_r$, under the condition
761: $X_a=x_a$. It is clear that the distribution of $\eta_a$ is obtained
762: through the recursion (\ref{eq:Iteration}) (with $Q^{(\ell)}_{x_i}$
763: replaced by $Q^{*}_{x_{a_i}}$), and since $Q^{*}_{x_{a_i}}$ is a fixed
764: point of this recursion, this proves the claim.
765: \endproof
766: %
767: %***********************************************************
768: %
769: \section{Variational principle}
770: \label{se:varprinc}
771: %
772: %*****************************
773: %
774: \subsection{The general principle}
775:
776: Here we establish a variational principle from which the fixed point equation
777: (\ref{eq:1RSB_fixed}) for the distribution at the root
778: can be deduced. We shall not try
779: to explain here its physical origin, which is related
780: to spin glass theory \cite{MP_Bethe}, but just discuss the relation
781: with the reconstruction problem. Throughout this
782: section we use the notation $\pi(x|y) = \pi(y|x) \equiv
783: \pi(x,y)$. Given a distribution $\Q$ over $\M$, we define its {\em complexity}
784: as
785: %
786: \begin{eqnarray}
787: %
788: \Sigma(\Q) = -\frac{k+1}{2}\int \W_{\rm e}(\eta_1,\eta_2)\;\;
789: \de\Q(\eta_1)\,\de\Q(\eta_2)+ \int \W_{\rm v}(\eta_1,\dots,\eta_{k+1})\;\;
790: \prod_{i=1}^{k+1}\de\Q(\eta_i)\, ,
791: %
792: \end{eqnarray}
793: %
794: where
795: %
796: \begin{eqnarray}
797: %
798: \W_{\rm e} &\equiv & - \left[\frac{\sum_{x_1,x_2}\eta_1(x_1)\eta_2(x_2)\pi(x_1,x_2) }
799: {\sum_{x_1,x_2} \eo(x_1)\eo(x_2)\pi(x_1,x_2)}\right]
800: \log\left[\frac{\sum_{x_1,x_2}\eta(x_1)\eta(x_2)\pi(x_1,x_2)}
801: {\sum_{x_1,x_2}\eo(x_1)\eo(x_2)\pi(x_1,x_2)}\right] \, ,\\
802: \W_{\rm v} &\equiv & - \left[\frac{\sum_x\prod_i
803: \sum_{x_i}\eta_i(x_i)\pi(x,x_i)}
804: {\sum_x\prod_i\sum_{x_i}\eo(x_i)\pi(x,x_i)}\right]
805: \log\left[\frac{\sum_x\prod_i\sum_{x_i}\eta_i(x_i)\pi(x,x_i)}
806: {\sum_x\prod_i\sum_{x_i}\eo(x_i)\pi(x,x_i)}\right] \, .
807: %
808: \end{eqnarray}
809: %
810: The complexity is interesting for the reconstruction problem because of the
811: following
812: remark:
813: %
814: \begin{propo}\label{propo:Stat}
815: %
816: Let $\Q^*$ be a distribution over $\M$ which satisfies the fixed point equation
817: (\ref{eq:1RSB_fixed}). Then $\Q^*$ is a stationary point of the complexity
818: $\Sigma(\, \cdot\,)$. More precisely, given any consistent distribution
819: $\Q$ over $\M$,
820: define $\Sigma^*(t)\equiv\Sigma((1-t)\Q^*+t\Q)$. Then
821: %
822: \begin{eqnarray}
823: %
824: \left.\frac{\de\Sigma^*}{\de t}\right|_{t=0} = 0\, .
825: %
826: \end{eqnarray}
827: %
828: \end{propo}
829: %
830: \prooft
831: This proposition is a direct consequence of Lemma \ref{remark1} in Appendix \ref{app:Variational}
832: (the proof consists in explicitly computing the derivative
833: of $\Sigma^*(t)$ and checking that it vanishes under the fixed point
834: conditions).
835: \endproof
836:
837: The complexity $\Sigma$ can also be written in terms of the
838: conditional distributions $Q_x(\eta) = q\eta(x)\Q(\eta)$.
839: Define $p(x_1,x_2)$ to be the marginal distribution of two
840: neighboring variables on the tree: $p(x_1,x_2) = \pi(x_1,x_2)/q$.
841: Similarly, let $p(x_1\dots x_{k+1})= [\sum_x \pi(x,x_1)\dots\pi(x,x_{k+1})]/q$,
842: the distribution of $k+1$ variables with one common neighbor.
843: The complexity is then given, in terms of $Q_x(\eta)$, by:
844: %
845: \begin{eqnarray}
846: %
847: \Sigma(Q) &=& -\frac{k+1}{2}\sum_{x_1,x_2}p(x_1,x_2)
848: \int W_{\rm e}(\eta_1,\eta_2)\; \de Q_{x_1}(\eta_1)
849: \,\de Q_{x_2}(\eta_2) +\nonumber\\
850: && +\sum_{\{x_i\}}p(x_1\dots x_{k+1})\int
851: W_{\rm v}(\eta_1,\dots,\eta_{k+1})\,\prod_{i=1}^{k+1} \de Q_{x_i}(\eta_i)\, ,
852: %
853: \end{eqnarray}
854: %
855: where
856: %
857: \begin{eqnarray}
858: %
859: W_{\rm e} &\equiv & -
860: \log\left[\frac{\sum_{x_1,x_2}\eta_1(x_1)\eta_2(x_2)\pi(x_1,x_2)}
861: {\sum_{x_1,x_2}\eo(x_1)\eo(x_2)\pi(x_1,x_2)}\right] \, ,\\
862: W_{\rm v} &\equiv & -
863: \log\left[\frac{\sum_x\prod_i\sum_{x_i}\eta_i(x_i)\pi(x,x_i)}
864: {\sum_x\prod_i\sum_{x_i}\eo(x_i)\pi(x,x_i)}\right] \, .
865: %
866: \end{eqnarray}
867: %
868: %*****************************
869: %
870: \subsection{Implications on reconstructibility}
871:
872:
873: Experience from spin glass theory, and the physical interpretation of the complexity,
874: suggests the following conjecture:
875: %
876: \begin{conj}\label{conj:Sigma}
877: Consider the reconstruction problem for the $k$-ary tree and a reversible
878: channel $\pi(y|x) = \pi(x|y)$.
879: If there exists a consistent distribution $\Q^{\rm tr}$ over $\M$, such that $\Sigma(\Q^{\rm tr})<0$,
880: then the reconstruction problem is solvable.
881: \end{conj}
882: %
883: Let us give here a few comments in favor of the plausibility of this
884: conjecture. Notice first that, if $\Q$ is trivial, then $\Sigma(\Q)=0$. Let
885: $\P(\M)$ denote the space of consistent probability distributions over $\M$.
886: Suppose that there exists $\Q^{\rm tr}$ with $\Sigma(\Q^{\rm tr})<0$.
887: Consider now the distribution $\Q^*\in \P(\M)$ such that the complexity is
888: minimal. Of course $\Sigma(\Q^*)\le \Sigma(\Q^{\rm tr})<0$ and therefore
889: $\Q^*$ is non-trivial. If $(i)$ $\Q^*$ is a stationary point of the complexity,
890: and $(ii)$ the stationary points of $\Sigma(\Q)$ in $\P(\M)$ coincide with the
891: solutions of the fixed point equation (\ref{eq:1RSB_fixed}), then the
892: existence of $\Q^{\rm tr}$ implies that the
893: reconstruction problem is solvable. Point $(i)$ amounts to banishing the
894: possibility that $\Q^*$ is on the `border' of $\P(\M)$. Point $(ii)$ is a
895: stronger version of Proposition \ref{propo:Stat}.
896:
897: Notice that a priori one
898: could formulate a similar conjecture with a $\Q^{\rm tr}$ having
899: $\Sigma(\Q^{\rm tr})>0$, replacing `minimum' with `maximum' and
900: `negative' with `positive' in the above. It is easy to find
901: counterexamples showing that this `reverse' conjecture is false.
902: The reason is probably that the distribution $\Q^*$ maximizing
903: $\Sigma(\Q)$ is on the border of $\P(\M)$, and therefore $(i)$ does not hold.
904:
905: Assuming Conjecture \ref{conj:Sigma} to hold, it implies a simple variational technique
906: for proving that reconstruction is possible. Just consider an explicit finite-dimensional family of
907: distributions $\Q_{\mu}$ depending on some parameters $\mu\in{\mathbb R}^d$, and
908: minimize $\Sigma(\Q_{\mu})$ over $\mu$. If the minimum is negative,
909: then reconstruction is possible.
910: We will apply the {\em variational principle} in this form in the next
911: Sections. In the rest of this Section (and in Appendix
912: \ref{app:Variational}) we shall prove the
913: principle for a special family of kernels $\pi(\,\cdot\,|\,\cdot\,)$
914: including the antiferromagnetic Potts channel.
915:
916: We define a kernel $\pi(y|x) = \pi(x,y)$, $x,y\in\{1,\dots,q\}$
917: to be `frustrated' if it can be decomposed as $\pi(x,y) = \pi_*-\pih(x,y)$
918: where $\pi_*\in {\mathbb R}$ is a constant and $\pih(x,y)$,
919: $x,y\in\{1,\dots,q\}$ is a positive-definite matrix.
920: The antiferromagnetic Potts kernel is a particular instance of this family, with
921: $\pi_* = \eps/(q-1)$, and $\pih(x,y)= |\lambda_2|\,
922: \delta_{x,y}$ where $\lambda_2 = 1-q\eps/(q-1)$.
923:
924: Our basic result is the following.
925: %
926: \begin{lemma}\label{lemma:Variational}
927: Let $\pi(\, \cdot\, ,\,\cdot\,)$ be a frustrated kernel and $\Q^*$
928: a consistent distribution over $\M$ which is not a solution of the
929: associated fixed point equation (\ref{eq:1RSB_fixed}).
930: Then
931: there exists a consistent distribution $\Q$ over $\M$ such that:
932: %
933: \begin{eqnarray}
934: %
935: \left.\frac{\de\phantom{t}}{\de t}\Sigma((1-t)\Q^*+t\Q)\right|_0<0\, .
936: %
937: \end{eqnarray}
938: \end{lemma}
939: %
940: The proof of this statement is postponed to Appendix \ref{app:Variational}.
941: Here we limit ourselves to proving that it implies the desired
942: principle.
943: %
944: \begin{propo}\label{propo:VariationalAnti}
945: %
946: Conjecture \ref{conj:Sigma} holds true in the case of frustrated kernels:
947: Let $\pi(\, \cdot\, ,\,\cdot\,)$ be a frustrated kernel, and
948: $\Sigma(\,\cdot\,)$ the associated complexity function. If
949: there exists a consistent distribution $\Q$ over $\M$ such that
950: $\Sigma(\Q^{\rm tr})<0$,
951: then the reconstruction problem is solvable.
952: %
953: \end{propo}
954: %
955: \prooft
956: Let $\Sigma_{\min}\equiv\inf \Sigma(\Q)$, the $\inf$ being taken in $\P(\M)$. Since this
957: space is subsequentially compact with respect to the weak
958: topology~\cite{Shiryaev},
959: and $\Sigma$ is continuous with respect to this topology, there exists
960: a consistent distribution $\Q^*$, such that $\Sigma(\Q^*) =\Sigma_{\rm min}$.
961: Because of Lemma \ref{lemma:Variational}, $\Q_*$ is a solution
962: of Eq.~(\ref{eq:1RSB_fixed}).
963: Furthermore $\Sigma(\Q_*)\le \Sigma(\Q^{\rm tr})<0$ and therefore
964: $\Q_*$ is non-trivial. The result is a consequence of Proposition \ref{propo:Reco}.
965: \endproof
966: %
967: %*****************************
968: %
969: \subsection{An application to Potts channels}
970: \label{eq:VariationalBound}
971: Here we describe
972: a simple family of distributions which can be used variationally
973: when studying the Potts channels. We will show in the next sections that,
974: in spite of its simplicity, it leads to rather accurate results.
975:
976: The family is indexed by a single real parameter
977: $\mu\in[0,1]$. We shall denote by $\Q_{\mu}$ the corresponding distribution
978: and will write, with some abuse of notation,
979: $\Sigma(\mu)\equiv\Sigma(\Q_{\mu})$.
980: The distribution $\Q_{\mu}$ attributes equal weight $1/q$ to the
981: $q$ points in $\M$ denoted by $\gamma^{(x)}$, $x\in\{1,\dots,q\}$, defined as follows
982: %
983: \begin{eqnarray}
984: %
985: \gamma^{(x)}(y) = \left\{ \begin{array}{ll}
986: 1-\mu & \;\;\;\mbox{if $y=x$},\\
987: \mu/(q-1) & \;\;\;\mbox{otherwise}.
988: \end{array}
989: \right.
990: %
991: \label{eq:Qmudef}
992: \end{eqnarray}
993: %
994: Some calculus shows that $\Sigma(\mu) = -\frac{k+1}{2}\, w_{\rm e}(\mu)+
995: w_{\rm v}(\mu)$, where
996: %
997: \begin{eqnarray}
998: %
999: w_{\rm e}(\mu) & = & -\frac{1}{q}\, A\log A-
1000: \frac{q-1}{q}\, B\log B\, ,\\
1001: A & = & q\left\{\frac{\eps}{q-1}+\left(1-\frac{q\eps}{q-1}\right)
1002: \left[(1-\mu)^2+\frac{\mu^2}{q-1}\right]\right\}\, ,\\
1003: B & = & q\left\{\frac{\eps}{q-1}+\left(1-\frac{q\eps}{q-1}\right)
1004: \left[\frac{2\mu(1-\mu)}{q-1}+\frac{(q-2)\mu^2}{(q-1)^2}\right]
1005: \right\}\, ,
1006: %
1007: \end{eqnarray}
1008: %
1009: and
1010: %
1011: \begin{eqnarray}
1012: %
1013: w_{\rm v}(\mu) & = & -\frac{1}{q^{k+1}}\sum_{n_1,\dots, n_q}
1014: \binom{k+1}{n_1,\dots,n_q}\, z[n]\log z[n]\, ,\\
1015: z[n] & = & q^k\left[\frac{\eps+\mu}{q-1}-\frac{q\eps\mu}{(q-1)^2}\right]^{k+1}
1016: \sum_{x=1}^q\left[\frac{\eps + (q-1-q\eps)(1-\mu)}{\eps + (q-1-q\eps)\mu/(q-1)
1017: }\right]^{n_x}\, ,
1018: %
1019: \end{eqnarray}
1020: %
1021: the first sum being restricted to $n_1,\dots,n_q\ge 0 $ and
1022: $n_1+\cdot+n_q = k+1$.
1023:
1024: Let us briefly discuss how these formulae are used in the following.
1025: To be definite, we refer here to the ferromagnetic case, the antiferromagnetic
1026: one being completely analogous.
1027: Given $k$, $q$ and $\epsilon$, we compute $\Sigma(\mu)$,
1028: and minimize it numerically for $\mu\in[0,1]$.
1029: The largest value (more precisely, the supremum)
1030: of $\eps$ such that the minimum value is negative, is denoted by
1031: $\eps_{\rm var}(k,q)$. According to conjecture \ref{conj:Sigma}, we expect $\eps_{\rm r}(k,q)\ge\eps_{\rm var}(k,q)$.
1032: Although we have proved it only for frustrated kernels
1033: (which do not include the ferromagnetic Potts channel), we shall loosely
1034: use the term `variational bound' also in the other cases..
1035:
1036: One can show that the variational bound is always at least as good
1037: as the KS one:
1038: $\eps_{\rm var}(k,q)\ge \eps_{\rm KS}(k,q)$ by looking
1039: at the behavior of $\Sigma(\mu)$ near to $\mu = (1-1/q)$.
1040: By Taylor expanding $\Sigma(\mu)$
1041: for $\mu = (1-1/q)+\delta\mu$, we obtain
1042: $\Sigma(\mu) = c_{k,q}(\eps) \delta\mu^4+O(\delta\mu^5)$.
1043: Furthermore $c_{k,q}(\eps)<0$ for $\eps<\eps_{\rm KS}(k,q)$ and
1044: $c_{k,q}(\eps)>0$ for $\eps>\eps_{\rm KS}(k,q)$.
1045: In Fig.~\ref{fig:q7} we plot $\Sigma(\mu)$ for the ferromagnetic Potts channel with $k=2,q=7$,
1046: showing that the variational bound $\eps_{\rm var}(k,q)$ is strictly larger than the KS one.
1047: We shall discuss in the next section for which values of $k,q$ this happens.
1048: If the variational principle were proved for the ferromagnetic
1049: channel, this
1050: would prove $\eps_{\rm r}(k,q)>\eps_{\rm KS}(k,q)$
1051: in these cases.
1052: \begin{figure}
1053: \begin{tabular}{cc}
1054: \includegraphics[width=0.4\linewidth,angle=-0]{q7_var1.eps}\hspace{1 cm}
1055: \includegraphics[width=0.4\linewidth,angle=-0]{q7_var2.eps}
1056: \put(-90,-10){$\mu$}
1057: \put(-310,-10){$\mu$}
1058: \put(-200,70){$\Sigma(\mu)$}
1059: \put(-420,70){$\Sigma(\mu)$}
1060: \end{tabular}
1061: \caption{The complexity for the ferromagnetic Potts channel with
1062: $k=2$ and $q=7$ within the variational ansatz $\Q_{\mu}$
1063: described in Sec.~\ref{eq:VariationalBound}. A negative complexity
1064: implies that the reconstruction problem is solvable. The three curves
1065: correspond (from bottom to top) to $\eps=0.250$, $0.253$, $0.256$.
1066: The right plot is a zoom near $\mu=6/7$.
1067: The KS threshold for $k=2,q=7$ is $\eps_{\rm KS}\approx 0.2510$. For
1068: $\eps=0.250<\eps_{\rm KS}$, $\Sigma(\mu)$ is negative in
1069: the neighborhood of $\mu=6/7$. For $\eps=0.253>\eps_{\rm KS}$,
1070: as $\mu$ decreases from its maximum value $6/7$, $\Sigma(\mu)$ is
1071: first positive, but then becomes negative with a minimum for
1072: $\mu\approx 0.6$, implying $\eps_{\rm r}>0.253$.
1073: This behavior is typical of a first order phase
1074: transition. For $\eps=0.256$, $\Sigma(\mu)$ is always positive,
1075: and one cannot draw any conclusion. }
1076: \label{fig:q7}
1077: \end{figure}
1078:
1079: Let us notice that we do not expect the
1080: variational lower bound to be tight. More precisely, even minimizing it over
1081: the space of distributions over $\M$, $\min \Sigma(\Q)$ becomes
1082: negative only below a threshold $\eps_{\rm c}(k,q)$
1083: with $\eps_{\rm KS}(k,q)<\eps_{\rm c}(k,q)<\eps_{\rm r}(k,q)$
1084: (in the case where $\eps_{\rm KS}(k,q)<\eps_{\rm r}(k,q)$). Our numerical
1085: simulations confirm this expectation which is motivated by the physical
1086: interpretation of the complexity.
1087:
1088: %
1089: %********************************************************************
1090: %
1091: \section{Thresholds for the ferromagnetic Potts channel}
1092: \label{sec:Ferro}
1093:
1094: %
1095: \begin{figure}
1096: \includegraphics[width=0.45\linewidth,angle=0]{onerun.eps}
1097: \put(-100,-10){$\ell$}
1098: \put(-210,75){$I_\ell$}
1099: \caption{The information (in bits) that can be transmitted reliably through
1100: a $k$-ary tree network of $q$-ary symmetric channels (ferromagnetic
1101: Potts channels), as determined with the population dynamics algorithm.
1102: Here $k=2$, $q=15$ and the noise parameter is (from top to bottom)
1103: $\eps = 0.333082$, $0.308057$, $0.282444$, $0.256422$.
1104: We used populations of size $M=10^5$, and averaged over $10$ runs.}
1105: \label{Onerun}
1106: \end{figure}
1107: %
1108: In order to determine reconstruction thresholds numerically,
1109: we simulate the recursion (\ref{eq:Iteration}),
1110: by representing the distributions $Q^{(\ell)}_x$ through
1111: a large enough sample. We will estimate reconstruction to be
1112: possible if the sample does not concentrate, for $\ell$ large
1113: around the point $\eo$.
1114:
1115: This procedure is very similar to the `population dynamics' method
1116: used to solve similar equations in spin glass theory~
1117: \cite{Abou-Chacra,MP_Bethe}.
1118: We work with $q$ samples (`populations')
1119: $P^{(\ell)}_1,\dots P^{(\ell)}_q$, each containing $M$ points
1120: $\eta_i\in \M$, $i\in\{1,\dots,M\}$
1121: (i.e. $M$ vectors $\eta_i(x)$, $x \in\{1,\dots,q\}$ with $\eta_i(x)\ge 0$
1122: and $\sum_x\eta_i(x) = 1$).
1123: The population $P^{(\ell)}_x$ represents an i.i.d. sample
1124: from the distribution $Q^{(\ell)}_x$. The population
1125: $P^{(\ell+1)}_x$, $x\in\{1,\dots,q\}$ is computed, for each
1126: $\ell\ge 0$ as follows.
1127: %
1128: \begin{itemize}
1129: %
1130: \item Choose $k$ iid colors $x_1,\dots,x_k$ with distribution
1131: $\pi(\, \cdot\,\vert x)$.
1132: \item Choose $k$ vectors $\eta_1,\dots\eta_k$, with $\eta_i$ uniformly
1133: random in $P_{x_i}$.
1134: \item Compute $\eta= \F(\eta_1,\dots,\eta_k)$ according to (\ref{eq:Site}).
1135: \item Store this new $\eta$ in the population $P^{(\ell+1)}_{x}$,
1136: and repeat until the population contains $M$ elements.
1137: %
1138: \end{itemize}
1139: %
1140: This whole cycle is repeated until the populations $P_x^{(\ell)}$ become
1141: stationary (by this we mean that their moments no longer depend on $\ell$)
1142: within some prescribed accuracy.
1143:
1144: Reconstructibility can be monitored by computing the parameters
1145: $I_{\ell}$ and $\Psi_{\ell}$ on in the populations
1146: $P_x^{(\ell)}$. If $I_\ell,\Psi_\ell\to 0$ as $\ell\to\infty$,
1147: then we estimate that reconstruction is not possible. If they
1148: instead converge to a finite value, we take this value
1149: as an estimate of $I_{\infty}$, $\Psi_{\infty}$. Figure
1150: \ref{Onerun} shows an example of such a calculation.
1151: Reconstructibility thresholds are determined by repeating the same
1152: experiment for several values of the channel noise $\eps$.
1153:
1154: %
1155: \begin{figure}
1156: \includegraphics[width=0.5\linewidth,angle=0]{sumk2zoom.eps}
1157: \put(-205,90){\includegraphics[width=0.2\linewidth,angle=0]{sumk2.eps}}
1158: \put(-100,-5){$q$}
1159: \put(-260,80){$\eps_{\rm r}/\eps_{\rm KS}$}
1160: %\put(-200,85){$0$}
1161: %\put(-135,85){$100$}
1162: %\put(-112,95){$1$}
1163: %\put(-112,145){$1.2$}
1164: \caption{Ratio between the reconstructibility and the KS thresholds
1165: $\eps_{\rm r}(k,q)/\eps_{\rm KS}(k,q)$,
1166: for the ferromagnetic Potts channel and $k=2$.
1167: Squares correspond to the numerical determination of $\eps_{\rm r}(k,q)$
1168: and crosses to the variational lower bound $\eps_{\rm var}(k,q)$.
1169: The inset refer to larger number of colors (up to $100$).}
1170: \label{ExampleFig}
1171: \end{figure}
1172:
1173: \begin{table}
1174: \begin{tabular}{|c|c|c|c|c|c|c|c|c|}
1175: \hline
1176: $q$ & $k$ & $\eps_{\rm r}$ & $\eps_{\rm KS}$ & $\eps_{\rm var}$
1177: & $\eps_{\rm alg}$ & $\eps_{\rm MP}$ &
1178: $I_{*}$ & $\Psi_{*}$ \\
1179: \hline
1180: \hline
1181: $5$ & $2$ & $0.2348(1)$ & $0.2343146$ & $0.23491$ & $---$ &
1182: $0.30264$ & $0.052(5)$ & $0.0152(16)$ \\
1183: $5$ & $3$ & $0.33881(5)$ & $0.3381198$ & $0.33887$ & $0.19047$ &
1184: $0.41712$ & $0.06(2)$ & $0.016(4)$ \\
1185: $5$ & $4$ & $0.4008(1)$ & $0.4$ & $0.40081$ & $0.29046$ &
1186: $0.48$ & $0.06(1)$ & $0.020(4)$ \\
1187: $5$ & $7$ & $0.4986(1)$ & $0.4976284$ & $0.49847$ & $0.41114$ &
1188: $0.57143$ & $0.07(1)$ & $0.020(4)$ \\
1189: $5$ & $15$ & $0.5955(1)$ & $0.5934409$ & $0.59422$ & $0.53965$ &
1190: $0.65238$ & $0.14(1)$ & $0.040(8)$ \\
1191: \hline
1192: $7$ & $2$ & $0.25432(5)$ & $0.2510513$ & $0.25369$ & $---$ &
1193: $0.34577$ & $0.14(1)$ & $0.028(4)$ \\
1194: $7$ & $4$ & $0.43325(5)$ & $0.4285714$ & $0.43250$ & $0.30769$ &
1195: $0.53909$ & $0.195(5)$ & $0.045(2)$\\
1196: \hline
1197: $10$ & $2$ & $0.2716(2)$ & $0.2636039$ & $0.26977$ & $---$ &
1198: $0.38325$ & $0.23(2)$ & $0.040(5)$ \\
1199: \hline
1200: $15$ & $2$ & $0.2881(1)$ & $0.2733670$ & $0.28472$ & $---$ &
1201: $0.41652$ & $0.37(3)$ & $0.053(4)$ \\
1202: \hline
1203: \end{tabular}
1204: \caption{Thresholds (numerical results and bounds) for the ferromagnetic
1205: Potts channel.
1206: The reconstruction threshold $\eps_{\rm r} $,
1207: whose numerical estimate is shown in the first column,
1208: satisfies the rigorous bounds
1209: $\eps_{\rm r}\ge \eps_{\rm KS}$, $\eps_{\rm r}\ge\eps_{\rm alg} $, and
1210: $\eps_{\rm r}\le \eps_{\rm MP}^-$.
1211: The `algorithmic bound' $\epsilon_{\rm alg}$ is computed by analyzing
1212: reconstruction through recursive majority along the lines of
1213: Ref.~\cite{MosselRec}.
1214: The variational principle (that is not proven for this ferromagnetic channel
1215: would imply $\eps_{\rm r}\ge \eps_{\rm var}$.
1216: The symbol $--$ means that the corresponding bound does not provide any
1217: information.
1218: }
1219: \label{TableFerro}
1220: \end{table}
1221:
1222: Numerical simulations clearly show that the reconstructibility
1223: and Kesten Stigum threshold coincide for $q=3$ and $q=4$.
1224: We checked this to be the case for $q=3$ and $k=2$--$7,10,15,20,30,50$,
1225: and $q=4$ and $k=2,3,5,10,15,30$ and expect it to be the case generically, at
1226: least for $k$ not too large. When this is the case, the order
1227: parameters $I_{\infty}$, $\Psi_{\infty}$ decrease continuously and vanish at
1228: $\eps_{\rm r}(k,q)=\eps_{\rm KS}(k,q)$.
1229:
1230: For $q\ge 5$ we always find $\eps_{\rm r}(k,q)>\eps_{\rm KS}(k,q)$.
1231: In these cases $I_{\infty}(\eps)\downarrow I_*>0$,
1232: $\Psi_{\infty}(\eps)\downarrow \Psi_*>0$ as $\eps\uparrow \eps_{\rm r}(k,q)$.
1233: In spin glass language, the transition is discontinuous:
1234: we refer to next Section for some illustrations.
1235: We report our numerical results in Table
1236: \ref{TableFerro}. This table also contains
1237: the variational lower bound $\eps_{\rm var}(k,q)\le\eps_{\rm r}(k,q)$,
1238: as well as the upper bound derived in
1239: \cite{MosselPeres01}:
1240: $\eps_{\rm r}(k,q)\le \eps_{\rm MP}^{+}(k,q)$,
1241: where
1242: %
1243: \begin{eqnarray}
1244: %
1245: \eps_{\rm MP}^{\pm}(k,q) = (q-1)\frac{(2-q+2kq)\mp\sqrt{(2-q+2kq)^2
1246: -4k(k-1)q^2}}{2kq^2}\, .
1247: %
1248: \end{eqnarray}
1249: %
1250:
1251: In Fig.~\ref{ExampleFig} we plot the thresholds as a function of
1252: $q$ for $k=2$.
1253: %
1254: %***********************************************************************
1255: %
1256: \section{Thresholds for the antiferromagnetic Potts channel}
1257: \label{sec:AntiFerro}
1258:
1259: \begin{table}
1260: \begin{tabular}{|c|c|c|c|c|c|c|c|c|c|}
1261: \hline
1262: $q$ & $k$ & $\eps_{\rm r}$ & $\eps_{\rm KS}$ & $\eps_{\rm var}$
1263: & $\eps_{\rm alg}$ & $\eps_{\rm MP}^-$
1264: & $I_{*}$ & $\Psi_{*}$ & $\Sigma_{*}$ \\
1265: \hline
1266: \hline
1267: $4$ & $8$ & $0.99953(4)$ & $--$ & $--$ & $--$ & $0.91552$ &
1268: $1.56(4)$ & $0.56(1)$ & $0.026(3)$\\
1269: $4$ & $9$ & $0.9908(4)$ & $1$ & $0.99298$ & $--$ & $0.90717$ &
1270: $1.31(2)$ & $0.47(2)$ & $0.009(1)$\\
1271: $4$ & $10$ & $0.9820(8)$ & $0.9871708$ & $0.98304$ & $--$ & $0.9$ &
1272: $1.2(2)$ & $0.42(4)$ & $0.005(4)$\\
1273: $4$ & $11$ & $0.9725(3)$ & $0.9761335$ & $0.97363$ & $0.99736$ & $0.89376$ &
1274: $1.07(5)$ & $0.39(1)$ & $\lesssim 0.005$\\
1275: $4$ & $12$ & $0.9643(3)$ & $0.9665063$ & $0.96498$ & $0.98946$ & $0.88826$ &
1276: $0.26(3)$ & $4.2(5)$ & $\lesssim 0.005$ \\
1277: $4$ & $15$ & $0.9431(3)$ & $0.9436492$ & $0.94338$ & $0.96903$ & $0.875$ &
1278: $0.5(1)$ & $0.16(3)$ & $\lesssim 0.001$ \\
1279: $4$ & $18$ & $0.9267(2)$ & $0.9267766$ & $0.92686$ & $0.95264$ & $0.86502$ &
1280: $0.3(1)$ & $0.11(4)$ & $\lesssim 0.001$ \\
1281: \hline
1282: $5$ & $13$ & $0.99741(5)$ & $--$ & $0.99982$ & $--$ & $0.92308$ &
1283: $1.76(4)$ & $0.59(1)$ & $0.042(5)$\\
1284: $5$ & $14$ & $0.9932(1)$ & $--$ & $0.99555$ & $--$ & $0.91916$ &
1285: $1.7(1)$ & $0.54(2)$ & $0.03(1)$\\
1286: $5$ & $15$ & $0.9888(1)$ & $--$ & $0.99092$ & $--$ & $0.91561$ &
1287: $1.48(5)$ & $0.48(2)$ & $0.03(1)$\\
1288: $5$ & $20$ & $0.9685(3)$ & $0.9788854$ & $0.96991$ & $0.98581$ & $0.90177$ &
1289: $1.1(5)$ & $0.36(2)$ & $0.01(1)$\\
1290: \hline
1291: $6$ & $17$ & $0.999924(5)$ & $--$ & $--$ & $--$ & $0.93482$ &
1292: $2.20(4)$ & $0.667(15)$ & $0.095(5)$ \\
1293: $6$ & $20$ & $0.9932(3)$ & $--$ & $0.99546$ & $--$ & $0.92792$ &
1294: $1.87(6)$ & $0.569(15)$ & $0.04(2)$\\
1295: \hline
1296: \end{tabular}
1297: \caption{Thresholds (numerical results and bounds) for the antiferromagnetic
1298: Potts channel.
1299: The reconstruction threshold $\eps_{\rm r} $,
1300: whose numerical estimate is shown in the first column,
1301: satisfies the rigorous bounds
1302: $\eps_{\rm r}\le \eps_{\rm KS}$ (from \cite{KestenStigum2}),
1303: $\eps_{\rm r}\le\eps_{\rm alg}$ (cf. \cite{MosselRec}),
1304: $\eps_{\rm r}\le\eps_{\rm var} $ (from Proposition \ref{propo:VariationalAnti}),
1305: and
1306: $\eps_{\rm r}\ge \eps_{\rm MP}^-$ (from \cite{MosselPeres01}).
1307: The symbol $--$ means that the corresponding bound does not
1308: provide any information.}
1309: \label{TableAnti}
1310: \end{table}
1311: %
1312: %
1313: \begin{figure}
1314: \includegraphics[width=0.45\linewidth,angle=0]{q3k6.eps}
1315: \hspace{1cm}
1316: \includegraphics[width=0.45\linewidth,angle=0]{q4k9.eps}
1317: \put(-465,95){\includegraphics[width=0.475\linewidth,angle=0]{q3k6cplx.eps}}
1318: \put(-221,95){\includegraphics[width=0.475\linewidth,angle=0]{q4k9cplx.eps}}
1319: \put(-100,-10){$\eps$}
1320: \put(-350,-10){$\eps$}
1321: \put(-234,50){$\Psi_{\infty},I_{\infty}$}
1322: \put(-480,50){$\Psi_{\infty},I_{\infty}$}
1323: \put(-225,138){$\Sigma_{\infty}$}
1324: \put(-474,138){$\Sigma_{\infty}$}
1325: \caption{Asymptotic complexity $\Sigma_{\infty}$, information capacity
1326: $I_{\infty}$ and conditional variance $\Psi_{\infty}$ as a function
1327: of the noise parameter for the antiferromagnetic Potts channel.
1328: On the left: typical continuous reconstructibility transition,
1329: $q=3$, $k=6$. On the right: typical discontinuous transition,
1330: $q=4$, $k=9$. }
1331: \label{DisContinuousFig}
1332: \end{figure}
1333: %
1334:
1335: In Table \ref{TableAnti} we present our numerical results for the
1336: reconstruction thresholds of the antiferromagnetic
1337: Potts channel in the cases in which it differs from $\eps_{\rm KS}$, together with the
1338: bounds.
1339:
1340: One distinctive feature of this channel is that, even
1341: in the limit $\eps\to 1$ reconstruction may be impossible. For any
1342: given $q\ge 3$ reconstruction becomes possible only for
1343: $k\ge k_*(q)$. Numerically we found $k_*(3) = 5$, $k_*(4)=8$, $k_*(5)=13$,
1344: $k_*(6)=17$. In fact the case $\eps=1$ has a special
1345: interest. In this case the configuration produced by the broadcast process
1346: is distributed according to the free boundary
1347: Gibbs measure for proper colorings of the (infinite) tree $\Tree_k$.
1348: Our numerical results imply that this measure is extremal only for
1349: $k<k_*(q)$, with $k_*(q)$ as above. Using the variational principle
1350: (which in this case is proved, cf. Proposition \ref{propo:VariationalAnti}),
1351: we can show that $k_*(3)\le 5$, $k_*(4)\le 9$, $k_*(5)\le 13$,
1352: $k_*(6)\le 17$\dots
1353:
1354: For $q=3$ we found the reconstructibility threshold to coincide
1355: always with the KS threshold. This was checked for
1356: $k=4$--$7$, $10$, $20$. The parameters $I_{\infty}$ and $\Psi_{\infty}$ are
1357: continuous functions of $\eps$ vanishing at $\eps_{\rm KS}$.
1358: An example is provided in Fig.~\ref{DisContinuousFig}, left frame.
1359: For $q\ge 4$ and $k\ge k_*(q)$ the transition at the reconstructibility
1360: threshold is discontinuous, cf. Fig.~\ref{DisContinuousFig}, right frame.
1361: Table \ref{TableAnti} gives the values of $I_{*}$, $\Psi_{*}$
1362: and $\Sigma_{*}\equiv \lim_{\eps\to \eps_r}\Sigma(\Q^{(\infty)})$
1363: (in the ferromagnetic channel, this number is so small that it cannot be measured reliably in the numerics).
1364: Most of the remarks made for the ferromagnetic channel apply to this case.
1365:
1366: %
1367: %************************************************************************
1368: %
1369: \section{Relation to spin glass theory}
1370: \label{se:spinglass}
1371:
1372: In this section we explore the link between reconstruction and spin glass
1373: theory.
1374: For simplicity we keep to the Potts channel, but the discussion
1375: can be easily generalized.
1376: Consider a configuration $\uY^{L}$ generated by the
1377: broadcast along a finite rooted tree $\Tree_k(L)$ with $L$
1378: generations, starting from a uniformly random symbol in
1379: $\{1,\dots,q\}$ at the root.
1380: As we saw in the introduction, $\uY^L$ is an equilibrium
1381: configuration of the Potts model with free boundary conditions
1382: on $\Tree_k(L)$, i.e. is distributed according to the
1383: Boltzmann law for the energy function (\ref{eq:PottsEnergy}).
1384: The coupling $J$ of this model is given by
1385: $e^{-\beta J}=\frac{\eps}{(q-1) (1-\eps)}$,
1386: and is ferromagnetic (resp. antiferromagnetic) if $J>0$ (resp $J<0$).
1387:
1388: Once the broadcast process has fixed the variables on
1389: the boundary at distance $L$ from the root, the reconstruction
1390: problem can be phrased in terms of the conditional distribution
1391: $\prob\{X_0=x|\uX_{L}=\ux_{L}\}$.
1392: The distribution of the first $L-1$ generations given the received
1393: symbols, $\prob\{\uY^{L-1}| \uX_{L}=\ux_{L}\}$,
1394: is also given by Boltzmann law for the energy function (\ref{eq:PottsEnergy}).
1395: However the boundary condition is now given by the received symbols.
1396: One fundamental reason why reconstruction is related to a spin glass
1397: problem is that this boundary condition tends
1398: to frustrate the system, in the sense of creating conflicting constraints.
1399:
1400: It is well known that on trees, frustration comes only through the choice
1401: of boundary conditions. Here we discuss the spin glass phase induced in the
1402: Potts model on the tree by various possible choices. We shall first show how a
1403: `naive' choice of boundary conditions leads to a simple replica symmetric
1404: recursion relation. Then we show how well-chosen self-consistent boundary
1405: conditions lead to the correct 1RSB fixed point equation of
1406: reconstruction.
1407: Finally we discuss the explicit realization of the corresponding spin
1408: glass
1409: model as a model of Potts spins on a random graph (not a tree).
1410: %
1411: %************************************************************************
1412: %
1413: \subsection{Boundary conditions with independent spins}
1414: %
1415: As before, we call $\uX_{L}$ the set of all spins in the $L^{\rm th}$
1416: generation. A boundary condition (BC) is a probability distribution on
1417: these spins. One first possibility is when spins on the boundary are
1418: independent random variables: a given spin $X_i$, $i\in V_{L}$ takes
1419: value $x_i$ with probability $\eta_i(x_i)$.
1420: The overall distribution is therefore $\prod_{i \in V_{L}}\eta_i(x_i)$.
1421: Once a set of $\eta_i(\, \cdot\, )$'s is given, the partition function of
1422: the tree is obtained as:
1423: %
1424: \begin{equation}
1425: %
1426: Z_L(\{\eta_i\}_{i\in V_L})=\sum_{\uy^L} \prod_{(i,j)\in \Tree_k(L)}
1427: \pi(x_i,x_j) \prod_{i\in V_L} \eta_i(x_i)\, ,
1428: \label{Zdef}
1429: %
1430: \end{equation}
1431: %
1432: and the Boltzmann distribution is
1433: %
1434: \begin{equation}
1435: %
1436: \prob^L_{\{\eta_i\} }(\uy^{L}) =\frac{1}{Z_L(\{\eta_i\}}\;
1437: \prod_{(ij)\in \Tree_k(L)}
1438: \pi(x_i,x_j) \prod_{i\in V_L} \eta_i(x_i)\, .\label{eq:BoltzmannBC}
1439: %
1440: \end{equation}
1441: %
1442:
1443: We have still the freedom of chosing the $\eta_i(\, .\,)$.
1444: One simple possibility would be to take them identical:
1445: $\eta_i(\, \cdot\, )= \overline{\eta}(\, \cdot\, )$
1446: \cite{Peruggi83,Peruggi84a,Peruggi84b}, but in order to have a
1447: disordered and frustrated problem one can choose to sample
1448: the $\eta_i$'s independently from a {\it symmetric} distribution
1449: $P^{(0)}(\eta)$. This definition of a spin glass problem on a tree
1450: was adopted for instance in \cite{Chayes86} in the Ising case ($q=2$),
1451: where each of the boundary spins was fixed to
1452: $\pm 1$ independently. In our formulation, this corresponds to the choice
1453: $P^{(0)}(\eta)=\frac{1}{2}\left(\delta\left[\eta(\, \cdot\,),\delta_{+1}\right]
1454: +\delta\left[\eta(\,\cdot\,),\delta_{-1}\right]\right)$.
1455:
1456: The recursive procedure for merging rooted trees applies in the same way
1457: as in Sec.~\ref{se:merging}. Consider the marginal distribution on the
1458: first $L-\ell\le L$ generations, $\prob(\uy^{L-\ell})$. It is clear that
1459: this has the same form as in Eq.~(\ref{eq:BoltzmannBC}),
1460: with some new $\eta'_i(\,\cdot\, )$, $i\in V_{L-\ell}$.
1461: When one generate BCs randomly as described above, the $\eta'_i$ are
1462: iid random variables with common distribution
1463: $P^{(\ell)}(\eta)$. A little thought shows
1464: that $P^{(\ell)}(\eta)$ is related to the one in the shell just
1465: above by:
1466: %
1467: \begin{equation}
1468: P^{(\ell+1)} (\eta)=
1469: \int
1470: \delta\left[\eta- \F(\eta_1,\dots,\eta_k)\right]\;\; \prod_{i=1}^k
1471: \de P^{(\ell)}(\eta_i)
1472: \, .
1473: \label{eq:iterRS}
1474: \end{equation}
1475: %
1476: Notice that in each shell, $P^{(\ell)}$ is symmetric.
1477:
1478: The marginal distribution at the root of the tree $\Tree_k(L)$,
1479: under the Boltzmann law (\ref{eq:BoltzmannBC}),
1480: is a random variable with distribution $P^{(L)}(\eta)$.
1481: In this model, the existence of a spin glass phase is characterized by
1482: a non trivial limit of $P^{(L)}(\eta)\to P^{(\infty)}(\eta)$ as $L\to\infty$.
1483: Such a limit solves a fixed point equation corresponding to
1484: (\ref{eq:iterRS}).
1485:
1486: The reader will notice that Eq.~(\ref{eq:iterRS}) is similar to
1487: the reconstruction equation (\ref{eq:1RSB}), with one crucial difference:
1488: the `reweighting' factor $z(\{\eta_i\})$ in
1489: (\ref{eq:1RSB}) is absent here. In the spin-glass jargon,
1490: Eq.~(\ref{eq:iterRS}) is the `replica symmetric' (RS) equation,
1491: while Eq.~(\ref{eq:1RSB}) is the 1RSB
1492: equation with Parisi parameter $m=1$.
1493:
1494:
1495: We want to argue that the model defined by Eq.~(\ref{eq:BoltzmannBC}), with iid $\eta_i$'s
1496: in not a `good' model of spin glass on a tree.
1497: Technically this is seen from the fact that its glass
1498: phase is a RS one, while the spin glass models on graphs with loops typically show RSB.
1499: Fundamentally, the drawback of this model is precisely
1500: that it neglects correlations between spins on the boundary.
1501: As we will discuss in Sec.~\ref{Bethe_lattice}, such correlations are necessary in order
1502: to study the existence of many pure states, a distinguished mark of spin glasses on graphs
1503: with loops.
1504:
1505:
1506: A minimalistic way of introducing such correlations is
1507: to keep uniquely those correlations induced by the tree itself;
1508: this is precisely what is done in reconstruction,
1509: as we now discuss.
1510: %
1511: %**************************************************************
1512: %
1513: \subsection{Self-consistent boundary conditions}
1514: %
1515: It is clear that the broadcast/reconstruction process generates correlated BCs.
1516: By this we mean that the conditional distribution
1517: $\prob(\uY^{L-1}=\uy^{L-1}|\uX_L)$ has still the form
1518: (\ref{eq:BoltzmannBC})
1519: but the $\eta_i$ are no longer independent.
1520: More explicitely, the $\eta_i$'s, $i\in V_L$
1521: have distribution:
1522: %
1523: \begin{equation}
1524: \prob(\{\eta_i\}_{i\in V_L})= \frac{1}{\Xi_L}\,
1525: Z_L(\{\eta_i\}_{i\in V_L}) \,
1526: \prod_{i\in V_L} \QQ^{(0)}(\eta_i) \ ,
1527: \label{scBC}
1528: \end{equation}
1529: %
1530: where $Z(\{\eta_i\})$ is the partition function
1531: (\ref{Zdef}) of the tree $\Tree_k(L)$ with BC $\{\eta_i\}$,
1532: and $\QQ^{(0)}(\eta)$ is the uniform distribution on the $q$
1533: `corners' of the simplex
1534: $\eta(x)=\delta_{x,r}$, $r\in\{1,\dots,q\}$.
1535:
1536: We can analyze the system with BC
1537: (\ref{scBC}) as we did in the previous section for uncorrelated BCs.
1538: Consider, as before, the marginal distribution of the first $L-\ell\le L$
1539: generations of the tree. It also has the form
1540: (\ref{eq:BoltzmannBC}) and
1541: the new $\eta_i$'s at distance $\ell$ retain the same correlation structure.
1542: Their joint distribution is
1543: %
1544: \begin{equation}
1545: %
1546: \prob(\{\eta_i\}_{i\in V_\ell})= \frac{1}{\Xi_{\ell}}\,
1547: Z_\ell(\{\eta_i\}_{i\in V_\ell}) \,
1548: \prod_{i\in V_{\ell}} \QQ^{(\ell)}(\eta_i) \, .
1549: \label{scBC2}
1550: %
1551: \end{equation}
1552: %
1553: Finally, the distribution
1554: $\QQ^{(\ell+1)}$ is related to $\QQ^{(\ell)}$ through the recursion
1555: (\ref{eq:1RSB}) that
1556: we found when discussing reconstruction, with the correct reweighting
1557: factor. Since we took $\QQ^{(0)}(\eta) = \Q^{(0)}(\eta)$,
1558: this implies $\QQ^{(\ell)}(\eta) = \Q^{(\ell)}(\eta)$ for any
1559: $\ell\ge 0$.
1560:
1561: It is interesting to define the spin glass problem on the tree associated
1562: with the non-trivial fixed point of Eq.~(\ref{eq:1RSB_fixed}).
1563: This just amounts to generating
1564: the BCs from (\ref{scBC}) using $\QQ^{(0)}=\Q^{*}$. This problem has the
1565: virtue of being statistically translation
1566: invariant~\footnote{Provided one replaces the rooted tree
1567: (with a root of degree $k$) with a regular Cayley tree
1568: (with all the vertices of degree $k+1$).}
1569: (although, for any given realization, the resulting Gibbs measure is
1570: not translation invariant).
1571: In particular the properties of a spin don't depend on its distance to the
1572: root.
1573: %
1574: %***********************************************************************
1575: %
1576: \subsection{Spin glass on the Bethe lattice}
1577: %
1578: \label{Bethe_lattice}
1579: While the previous definition of a spin glass on a tree is perfectly
1580: correct, it is clear that a lot of the physics has been put into the choice
1581: of the BC distribution. This is necessary because of the crucial role of BCs
1582: on trees. An alternative definition of the Bethe lattice spin glass, proposed
1583: in \cite{MP_Bethe}, is to use, instead of trees, random graphs which have a
1584: tree structure on finite length scales. Let us consider for instance the
1585: problem of $N$ Potts spins on the vertices of a random regular graph
1586: $\cG_N$ with degree
1587: $k+1$, with pairwise interactions given by the kernel $\pi(x,y)$
1588: The partition function of such a model is
1589: %
1590: \begin{equation}
1591: Z=\sum_{x_1,\dots,x_N} \prod_{(i,j)\in \cG_N} \pi(x_i,x_j)
1592: \label{eq:randGr}
1593: \end{equation}
1594: %
1595: In any finite neighborhood of a randomly chosen node $i$, the local
1596: structure~\footnote{By this we mean the subgraph within any fixed distance
1597: from $i$. The property described here can also be phrased in
1598: terms of {\em local weak convergence}~\cite{AldousSteele}} of $\cG_N$
1599: is (with high probability) the one of a regular tree with degree $k+1$.
1600: In fact, the shortest loop through $i$
1601: is typically of size $\log N$ which diverges when $N\to\infty$.
1602: This setting is interesting for two reasons:
1603: $(i)$ Loops, although large, can create some frustration;
1604: $(ii)$ The system is approximately homogeneous (unlike on a regular tree,
1605: where vertices on the boundary have a neighborhood very
1606: different from the others).
1607:
1608: Spin glasses on random lattices with a local tree-like structure have been
1609: the object of many studies in recent times. The cavity method of
1610: \cite{MP_Bethe,MP_Bethe_T0} is an iterative procedure which exploits the
1611: tree-like structure. Here we shall just mention some of its main results
1612: without justification, the aim being to clarify the correspondence
1613: between the spin glass model on a random graph and the reconstruction
1614: problem.
1615:
1616: %
1617: \begin{figure}
1618: \includegraphics[width=0.25\linewidth,angle=0]{cavrec.eps}
1619: \put(-50,-5){$j$}
1620: \put(-50,40){$i$}
1621: \put(-50,85){$l$}
1622: \put(-90,25){$\eta_{i\to j}$}
1623: \caption{The basic cavity recursion.}
1624: \label{fig:CavityRecursion}
1625: \end{figure}
1626: %
1627: In the cavity method one first considers the graph, rooted in a site $i$,
1628: obtained by cutting the edge $(i,j)$ to one of its neighbors, cf.
1629: Fig.~\ref{fig:CavityRecursion}. The
1630: marginal distribution of the root, $\eta_{i \to j}(x_i)$,
1631: with respect to the model on the `amputated' graph, is then
1632: written in terms of the distributions $\eta_{l\to i}(x_l)$
1633: where $l$ are the neighbors of $i$ different from $j$. It is possible to write
1634: such a recursion only if the variables $x_l$,
1635: in the absence of the edges $(l,j)$, become uncorrelated in the large $N$
1636: limit. Such a property is made possible by the local
1637: tree-like structure, but also requires
1638: a fast decay of correlations in the graph.
1639: This is expected to happen either when the system admits
1640: a single Gibbs state, or when the Boltzmann measure is restricted
1641: to a pure (extremal) Gibbs state~\footnote{The definition of
1642: extremal Gibbs state on a {\em finite} graph goes beyond the scope of this
1643: paper.}. In
1644: the first case, the problem is described by a unique distribution $\eta(x)$,
1645: and $\eta_{i \to j}=\eta$ for all directed edges $i\to j$.
1646: This distribution is a fixed
1647: point of (\ref{eq:Site}), satisfying thus $\eta=\F(\eta,\dots,\eta)$. It
1648: is called the 'paramagnetic' or 'liquid' phase.
1649: In the case where there exist
1650: several pure states, the recursion holds when the measure is restricted to
1651: one pure state $\alpha$: on a given (large) graph one could thus generate a
1652: set of `messages' $\eta_{i\to j}^\alpha(x_i)$ for each state $\alpha$. Notice
1653: that, for a given $\alpha$, the messages now depend explicitely on the
1654: edge: the measure is no longer uniform, but it is modulated. The 1RSB
1655: cavity method assumes that there exist exponentially
1656: many such pure states, the number $\cN(f)$ of states with free energy density
1657: $F^\alpha/N=f$ is written in terms of the complexity function $\Sigma(f)$ as
1658: $\cN(f)=\exp(N \Sigma(f))$. In such a case one can perform a statistics in the space of
1659: pure
1660: states, by introducing, for each edge $(i,j)$, the probability $R_{i\to
1661: j}(\eta)$ that the message $\eta_{i\to j}^\alpha=\eta$, when $\alpha$ is
1662: chosen
1663: randomly with a weight proportional to the total Boltzmann weight of state
1664: $\alpha$. After performing this average over states the various edges
1665: become
1666: again equivalent, and one finds that the distribution $R_{i\to
1667: j}(\eta)=\Q^*$
1668: satisfies exactly the 1RSB fixed point equation (\ref{eq:1RSB_fixed}). So
1669: there exists a 1RSB glass phase if and only if this equation has a non
1670: trivial
1671: symmetric solution. Notice that this equation can also have other
1672: non-symmetric solutions. For instance in the case of the ferromagnetic
1673: Potts
1674: channel, at low enough temperature there is a solution where $Q^*$ is
1675: peaked
1676: on a $\eta$ with a ferromagnetic bias, but it does not satisfy the
1677: symmetry
1678: property that we impose for the study of the glass state. In such a system
1679: the
1680: glass solution exists, but it is not realized on a random graph: the
1681: system
1682: will transit to a ferromagnetic phase. On the contrary in some other cases
1683: the
1684: glass phase will be realized. For instance we expect this to be the case
1685: for
1686: the antiferromagnetic Potts model on the random graph \cite{Mulet02,Brauenstein03}.
1687:
1688:
1689: The tree reconstruction problem on the one hand, and the spin glass on a
1690: random graph on the other, thus naturally lead to the same equations.
1691: Some aspects of this correspondence call for a better understanding.
1692: Consider the model on the random graph defined in
1693: Eq.~(\ref{eq:randGr}). Let us suppose that it has several pure states, and that
1694: the 1RSB cavity solution of the problem is correct. Now isolate around an
1695: arbitrary point the set of all its neighbors up to distance $\ell$.
1696: Generically it is a tree $\Tree_k(\ell)$. The vertices outside this tree create
1697: some boundary condition on the leaves of this tree, depending
1698: on the pure state $\alpha$ that we are considering.
1699: We have found that
1700: the statistics of these BC on the pure states corresponds to the statistics
1701: of the boundaries in the broadcast/reconstruction, and both are described
1702: by the distribution $\Q^*$.
1703: In spin glass theory (within 1RSB) one can count
1704: the pure states through the computation of the complexity function
1705: $\Sigma(f)$. It would be very interesting to have an
1706: interpretation of this function in terms of the reconstruction problem.
1707: %
1708: %********************************************************************
1709: %
1710: \section{Generalizations}
1711: \label{se:gene}
1712:
1713: So far we have focused on $k$-ary trees whose links corresponds
1714: to identical copies of the same $q$-ary channel satisfying the symmetry
1715: condition $\pi(y|x) = \pi(x|y)$. However, none of these hypotheses is
1716: crucial to our approach. In this section we define a considerably more
1717: general context, and sketch how to adapt the above formalism to this case.
1718: Some cases of broadcast through non regular trees, or with asymmetric channels have been considered
1719: for instance in \cite{EvaKenPerSch,Mossel01,MosselPeres01,Martin03}. The present formalism encompasses
1720: all these cases and generalizes them to broadcast through hypergraphs.
1721:
1722: We consider a finite set of kernels
1723: $\{\pi^{(\alpha)}(\, \cdot\, |\,\cdot\,)\, ;\, \alpha=1,\dots,n\}$,
1724: each kernel describing a one-to-many communication channel.
1725: For each $x\in\{1,\dots,q\}$, and each $k_{\alpha}$-uple
1726: $y_1,\dots, y_{k_\alpha}$,
1727: $\pi^{(\alpha)}(y_1,\dots,y_{k_{\alpha}} |x)$ gives is the probability
1728: that users $1,\dots,k_\alpha$ receive outputs $y_1,\dots,y_{k_{\alpha}}$
1729: if the channel input was $x$. These kernels must
1730: satisfy the conditions
1731: %
1732: \begin{eqnarray}
1733: %
1734: \pi^{(\alpha)}(y_1,\dots,y_{k_{\alpha}} |x)\ge 0\, ,
1735: \;\;\;\;\;\;\; \sum_{y_1,\dots,y_{k_{\alpha}}}
1736: \pi^{(\alpha)}(y_1,\dots,y_{k_{\alpha}} |x) = 1\, ,
1737: %
1738: \end{eqnarray}
1739: %
1740: and $x$ is called the parent of $y_1,\dots,y_{k_{\alpha}}$. Such `one-to-$k$' communication
1741: channels can be represented graphically
1742: using factor nodes of degree $k+1$, cf. Fig.~\ref{fig:fnode}.
1743:
1744: %
1745: \begin{figure}
1746: \includegraphics[width=0.3\linewidth,angle=-0]{fnode.eps}\hspace{1 cm}
1747: \includegraphics[width=0.6\linewidth,angle=-0]{fnode2.eps}
1748: \caption{Left: a function node representing the `one-to-$k$' channel $\pi^{(\alpha)}(y_1,\dots,y_{k} |x)$.
1749: Right: An example of a small random tree network, with $l_0=3$ }
1750: \label{fig:fnode}
1751: \end{figure}
1752: %
1753:
1754:
1755: Next, we define a {\em random} tree network ensemble depending
1756: on two probability distributions $q_{\alpha}$, $\alpha\in\{1,\dots,n\}$
1757: ($q_{\alpha}\ge 0$, $\sum q_{\alpha}=1$) and
1758: $p_l$, $l\ge 0$ ($p_{l}\ge 0$, $\sum p_{l}=1$).
1759: One (infinite) random network $\Tree$
1760: from this ensemble is generated as follows
1761: starting from the root $0$.
1762: %
1763: \begin{itemize}
1764: %
1765: \item Draw an integer $l_0$ with distribution $p_l$. This is the degree
1766: of the vertex $0$
1767: %
1768: \item For each $a\in\{1,\dots, l_0\}$, draw $\alpha_a$ independently with
1769: distribution $q_{\alpha}$, and attach a channel of type $\alpha$
1770: to the vertex $0$. The root will transmit through such channels.
1771: %
1772: \item For each $a\in\{1,\dots, l_0\}$, and each $i_a\in\{1,\dots,
1773: k_{\alpha_a}\}$, associate a vertex to the $i_a$-th output of
1774: channel $a$. Repeat the above construction for each of these vertices.
1775: %
1776: \end{itemize}
1777: %
1778: In Fig.~\ref{fig:fnode} we show a small example of such a network.
1779: We denote by $\Tree(i)$ the random (sub)network rooted at $i$, by
1780: $\Tree(i,\ell)$ its first $\ell$ generations (starting from $i$),
1781: and by $\uX_{i,\ell}$ the received colors, $\ell$ generations above
1782: $i$ ( $\uX_{0,\ell}\equiv\uX_{\ell}$).
1783:
1784: The network is used to communicate. A color $x_0\in\{1,\dots,q\}$
1785: is chosen at the root with probability $\varphi_0(x_0)$ and
1786: broadcast through the $l_0$ channels connected to the root itself.
1787: Each of the first generation vertices receives a corrupted
1788: version $x_i\in\{1,\dots,q\}$ of this color, with joint distribution
1789: %
1790: \begin{eqnarray}
1791: %
1792: \prod_{a=1}^{l_0}\pi^{(\alpha)}(x_{a,1},\dots,x_{a,i_a}|x_0)\, ,
1793: %
1794: \end{eqnarray}
1795: %
1796: where $(a,r)$ denotes the $r$-th output vertex of the $a$-th channel.
1797: In other words distinct channels act independently.
1798: The same transmission process is repeated at the first generation and so
1799: forth, through the entire network.
1800: The problem is to reconstruct the transmitted color from the output
1801: at generation $\ell$, denoted as $\ux_{\ell}$.
1802:
1803: Analogously to the case investigated in the previous sections,
1804: we say that the reconstruction problem is solvable if the
1805: conditional mutual information
1806: $I(X_0;\uX_{\ell}|\Tree)$ does not vanish as $\ell\to\infty$.
1807: Equivalently, the problem is solvable if there is a reconstruction
1808: procedure which succeeds with probability strictly larger than
1809: $\max_{x_0}\varphi_0(x_0)$ in the $\ell\to\infty$ limit. In these definitions
1810: we assume the network structure to be known at the
1811: receiver (this is why we consider a mutual information which is {\em conditional} to this structure).
1812:
1813: Unlike for $q$-ary reversible channels, the distribution of the
1814: color $X_i$ received at vertex $i$, is not uniform and depends upon the
1815: vertex. We shall denote it by $\varphi_i(x_i)\equiv \prob[X_i=x_i|\Tree]$.
1816: In fact $\varphi_i(\, \cdot\,)$ can be determined recursively:
1817: if $j$ is the parent of $i$, and $i_1$, $i_{k-1}$ the other vertices
1818: that share this parent, then
1819: %
1820: \begin{eqnarray}
1821: %
1822: \varphi_i(x_i) = \sum_{x_j}\sum_{x_{i_1},\dots,x_{i_{k-1}}}
1823: \pi^{(\alpha)}(x_i,x_{i_1},\dots,x_{i_{k-1}}|x_j)\,\varphi_j(x_j)\, .
1824: %
1825: \end{eqnarray}
1826: %
1827:
1828: Let us consider the reconstruction problem.
1829: The conditional distribution of the transmitted color given the observation
1830: at generation $\ell$, $\prob[X_0=x|\uX_{\ell}=\ux_{\ell}]$ can be computed
1831: recursively proceeding from the leaves downwards to the root
1832: as in Sec.~\ref{se:merging}.
1833: In order to simplify the analysis, it is convenient to `factor out'
1834: the {\em a priori} information $\varphi_i(x_i)$, and define
1835: %
1836: \begin{eqnarray}
1837: %
1838: \eta_{i,\ell}(x) = \prob_{i}[X_i=x|\uX_{i,\ell}]\, .
1839: %
1840: \end{eqnarray}
1841: %
1842: where $\prob_{i}$ denotes probability with respect to a modified
1843: process in which the boundary $ \uX_{i,\ell}$ is obtained from a broadcast
1844: starting from $X_i$ chosen uniformly at random in
1845: $\{1,\dots,q\}$.
1846: Of course we have
1847: %
1848: \begin{eqnarray}
1849: %
1850: \prob[X_i=x|\uX_{i,\ell}] = \frac{1}{z_i}\, \varphi_i(x)\,
1851: \eta_{i,\ell}(x)\, ,
1852: %
1853: \end{eqnarray}
1854: %
1855: where $z_i \equiv \sum_x \varphi_i(x)\eta_{i,\ell}(x)$ ensures the
1856: correct normalization.
1857: Notice that $\eta_{i,\ell}(x)$ depends uniquely on the portion
1858: of the tree above $i$, more precisely on $\Tree(i,\ell)$ and $\uX_{i,\ell}$.
1859:
1860: With a slight abuse of notation, let us denote by $\pi^{(a)}=\pi^{\alpha(a)}$,
1861: $a \in\{ 1,\dots, l_i\}$ be the channels whose input is $x_i$,
1862: and by $j_1,\dots,j_{k(a)}$ the corresponding output vertices.
1863: It is easy to derive the following recursion
1864: which generalizes Eq.~(\ref{eq:Site})
1865: %
1866: \begin{eqnarray}
1867: %
1868: \eta_{i,\ell+1}(x) = \frac{1}{z(\{\eta_{j,\ell}\},\{\pi^{(a)}\})}
1869: \,\prod_{a=1}^{l_i}\sum_{x_{1},\dots x_{k(a)}}
1870: \pi^{(a)}(x_{1},\dots x_{k(a)}|x)\;
1871: \eta_{j_1,\ell}(x_1)\cdots\eta_{j_{k(a)},\ell}(x_{k(a)})\, ,
1872: %
1873: \end{eqnarray}
1874: %
1875: where $\ell$ is the distance from the leaves.
1876: The constant $z(\{\eta_{j,\ell}\},\{\pi^{(a)}\})$ is defined
1877: by the constraint $\sum_{x}\eta_{i,\ell+1}(x) =1$.
1878: We shall denote the above mapping synthetically by writing
1879: $\eta_{i,\ell+1} = \F(\{\eta_{j,\ell}\},\{\pi^{(a)}\})$.
1880:
1881: One can define two types of probability distributions
1882: associated to $\eta_{i,\ell}$. We first assume that the
1883: tree $\Tree$ is given, and
1884: consider the distribution of $\eta_{i,\ell}$ conditional to $X_i=x$
1885: and $\Tree$. This will be denoted by $Q^{(i,\ell)}_x(\eta)$.
1886: Arguing as in the case of regular trees, one derives the recursion
1887: %
1888: \begin{eqnarray}
1889: %
1890: Q^{(i,\ell+1)}_x(\eta) =\sum_{\{x_j\}} \prod_{a=1}^{l_i}
1891: \pi^{(a)}(x_{j_1},\dots x_{j_{k(a)}}|x)\;
1892: \int\;\delta [\eta- \F(\{\eta_{j}\},\{\pi^{(a)}\})]\;
1893: \prod_j \de Q^{(j,\ell)}_{x_j}(\eta_j)\, .\label{eq:GeneralIteration}
1894: %
1895: \end{eqnarray}
1896: %
1897:
1898: Next, we consider the distribution of $\eta_{i,\ell}$
1899: {\em unconditional} of $\Tree$, which we denote by $Q^{(\ell)}_x(\eta)$.
1900: This can also be regarded as the expectation of the previous
1901: distribution: $Q^{(\ell)}_x(\eta) = \E_{\Tree} Q^{(i,\ell)}_x(\eta)$.
1902: Notice that $Q^{(i,\ell)}_{x}(\eta)$ depends on
1903: $\Tree$ only through the first $\ell$ generations of the subtree rooted at $i$
1904: $\Tree(i,\ell)$. Since the structures of distinct subtrees are
1905: independent, if we average Eq.~(\ref{eq:GeneralIteration})
1906: with respect to $\Tree$, the averages factorize, yielding
1907: a recursion equation for $Q^{(\ell)}_x(\eta)$:
1908: %
1909: \begin{eqnarray}
1910: %
1911: Q^{(\ell+1)}_x(\eta) = \E\sum_{\{x_j\}} \prod_{a=1}^{l_i}
1912: \pi^{(a)}(x_{j_1},\dots x_{j_{k(a)}}|x)\;
1913: \int\;\delta [\eta- \F(\{\eta_{j}\},\{\pi^{(a)}\})]\;
1914: \prod_j \de Q^{(\ell)}_{x_j}(\eta_j)\, .\label{eq:GeneralIterationSimpl}
1915: %
1916: \end{eqnarray}
1917: %
1918: Here $\E$ denotes expectation with respect to the degree $l_i$ and the
1919: channel types. The last expression is particularly convenient for
1920: numerical simulations and it is not more complex than the iteration
1921: (\ref{eq:Iteration}) studied in the previous Sections.
1922:
1923: We can also consider the distributions unconditional to
1924: the transmitted color: $\Q^{(i,\ell)}(\eta)$ and $\Q^{(\ell)}(\eta)$.
1925: The same relation as for regular trees
1926: hold in this case $Q^{(i,\ell)}_x(\eta) = q\eta(x) \Q^{(i,\ell)}(\eta)$
1927: and $Q^{(\ell)}_x(\eta) = q\eta(x) \Q^{(\ell)}(\eta)$. It is easy to
1928: derive the corresponding recursions. We just
1929: write down the equation for the last (non-random) distribution
1930: %
1931: \begin{eqnarray}
1932: %
1933: \Q^{(\ell+1)}(\eta) = \E\;
1934: \int\;\frac{z(\{\eta_j\}\{\pi^{(a)}\})}{z(\{\eo\}\{\pi^{(a)}\})}\;\;
1935: \delta [\eta- \F(\{\eta_{j}\},\{\pi^{(a)}\})]\;
1936: \prod_j \de \Q^{(\ell)}(\eta_j)\, .\label{eq:SymmetricGeneral}
1937: %
1938: \end{eqnarray}
1939: %
1940:
1941: In order to discuss the correspondence with the dynamical glass transition
1942: in this more general setting, it is necessary to distinguish two cases.
1943: In the simplest one, the RS cavity equations for the associated statistical
1944: mechanics model admit the solution $\eta_{i\to j}(x) =\eo(x)$
1945: for any directed link $i\to j$ in the graph. Under this hypothesis
1946: it is not hard to show that Eq.~(\ref{eq:SymmetricGeneral}) is equivalent
1947: (in the same sense discussed in Sec.~\ref{sec:Recursion}) to the
1948: $m=1$ 1RSB equation~\footnote{The expert will perhaps be surprised
1949: by this remark since it is usually said that the order parameter for
1950: such systems is a `measure over the space of distributions'.
1951: However
1952: it turns out that, for $m=1$, the expectation of this measure
1953: satisfies an equation which is Eq.~(\ref{eq:SymmetricGeneral}).}
1954:
1955: In the general case (i.e. if $\eta_{i\to j}(x) =\eo(x)$ does not solve the
1956: RS equations), the dynamical glass transition still corresponds to
1957: the extremality of the free boundary measure on an infinite tree.
1958: The last problem, however, cannot be formulated in the same framework as
1959: described here. One can still write an equation of the form
1960: (\ref{eq:GeneralIteration}), conditioned to the graph structure, as is usually
1961: done in statistical physics. But the average over the graph structure can only be
1962: performed conditioning upon the value of $\eta_{i\to j}(\,\cdot\,)$
1963: in the RS solution~\cite{ReconstrNext}, and so the relation between the spin glass problem and the
1964: reconstruction problem unconditioned to the structure of the tree, is not as simple as before. We
1965: shall not enter these details here.
1966: %
1967: %********************************************************************
1968: %
1969: \section{Conclusion and open problems}
1970: \label{sec:Conclusion}
1971:
1972: The coincidence between the reconstruction threshold and the dynamical glass
1973: transition explored in this paper is interesting from several points
1974: of view.
1975:
1976: First of all, it provides some perspective for putting on a firmer
1977: basis the theory of glassy systems on locally tree-like graphs
1978: developed in the last few years.
1979: As a concrete example, notice that the population dynamics algorithm
1980: defined in Section \ref{sec:Ferro} presents two important
1981: advantages with respect to the procedure adopted by statistical physicists.
1982: First, in spin glass theory the usual
1983: prescription is to look for a solution of the fixed point
1984: equation (\ref{eq:1RSB_fixed}). This poses the problem of the initial
1985: condition: even if a given initial condition yields a trivial fixed point,
1986: this may not be the case for all the initial conditions. In the present
1987: formulation the iteration is initialized with a very specific initial
1988: condition, and one is guaranteed that, if it converges to a trivial fixed
1989: point, no non-trivial fixed point exists.
1990: Second, simulating Eq.~(\ref{eq:1RSB_fixed})
1991: requires a `reweighting' of the sample which is usually the trickier
1992: part of the calculation. No reweighting is needed in the new approach:
1993: Eq.~(\ref{eq:Iteration}) can be handled easily.
1994:
1995: Also, it provides some indication on the correctness of simple
1996: analytic and algorithmic approaches to these systems,
1997: such as the replica symmetric cavity method, or the belief
1998: propagation algorithm. While it has long been known that their
1999: correctness should be related to a fast correlation decay in a
2000: model on a tree, a precise criterion has never been
2001: formulated
2002: \footnote{A frequently used sufficient criterion is the uniqueness of
2003: the Gibbs state on the infinite tree (see \cite{BanGam} and references
2004: therein). As it emerges from our discussion, this criterion is often much stronger
2005: than needed. It is also interesting to recall that
2006: Tatikonda and Jordan \cite{Tatikonda} first connected the convergence
2007: properties of belief propagation to the extremality of the free boundary Gibbs
2008: on a properly defined tree.}.
2009: Ous work suggests that the extremality of the associated Gibbs measure
2010: on a tree provides such a criterion.
2011:
2012: More broadly, it illustrate some subtleties of the physics of
2013: the glass transition. It is well known that, in a 1RSB
2014: glass transition, point to point correlations (static scattering factors)
2015: do not present any diverging correlation length. This paper
2016: shows that such a length can be derived quite generally from
2017: point-to-set correlations. Indeed the definition considered here is
2018: essentially equivalent to the one of Ref.~\cite{BoBi04}.
2019: It was shown in Ref.~\cite{MoSe05} that this length scale divergence
2020: implies a lower bound on the time scale divergence.
2021:
2022: Finally, statistical physics ideas can inspire new results
2023: on the original reconstruction problem. The most interesting such
2024: idea is, in our view, the complexity functional introduced in
2025: Section \ref{se:varprinc}. Apart from being conceptually innovative
2026: with respect to classical techniques, it seems to provide by
2027: far the best rigorous quantitative estimates of the reconstruction
2028: threshold, cf. Proposition \ref{propo:VariationalAnti}.
2029: This is well illustrated by the results in Table
2030: \ref{TableAnti}. It will certainly be interesting
2031: to find a concrete interpretation of this object in terms of
2032: the original reconstruction problem. Also, the values of the channel parameter
2033: at which the asymptotic complexity $\Sigma(\Q^{(\infty)})$ vanish have
2034: a particularly important role in statistical mechanics, but did not
2035: find any role here.
2036:
2037: There are several results that we have not been able to prove rigorously.
2038: We already formulated one such results as Conjecture \ref{conj:Sigma},
2039: and proved it for a particular class of channel models as Proposition
2040: \ref{propo:VariationalAnti}. Another interesting fact which has emerged from our
2041: numerical simulations is the coincidence of the KS and reconstruction
2042: thresholds when the number of colors is small.
2043: %
2044: \begin{conj}\label{conj:KS}
2045: Consider the reconstruction problem for the $k$-ary tree and the
2046: ferromagnetic Potts channel ($q$-ary symmetric channel) with $q\le 4$,
2047: or the antiferromagnetic Potts channel with $q\le 3$
2048: Then, there exists a $k_{\rm max}\ge 30$ such that, if $k<k_{\rm max}$
2049: the reconstruction threshold coincides with the Kesten-Stigum threshold.
2050: \end{conj}
2051: %
2052: A stronger version of this conjecture would be to require the thesis
2053: to be valid for {\em all} values of $k$ (i.e. to state
2054: that $k_{\rm max} = \infty$). Although we didn't find any $k$
2055: contradicting this stronger version, this might of course be due to
2056: the limitation on the values of $k$ that we can treat numerically.
2057:
2058: Finally, let us single out the case of completely antiferromagnetic
2059: ($\varepsilon=1$) Potts channels:
2060: \begin{conj}\label{conj:Col}
2061: Let $k_*(q)$ be the maximum value of $k$ such that the free boundary
2062: Gibbs measure for uniformly random proper colorings on the infinite $k$-ary
2063: tree is extremal. Then $k_*(3) = 5$, $k_*(4)=8$, $k_*(5)=13$,
2064: $k_*(6)=17$.
2065: \end{conj}
2066: %
2067: %***********************************************************
2068: %
2069: \section*{Acknowledgments}
2070: This work has been carried out mostly during our stay
2071: at the MSRI on the occasion of the program on ``Probability, Algorithms, and
2072: Statistical Physics''. We have benefited from numerous discussions,
2073: suggestions and comments from James Martin, Elchanan Mossel and Yuval Peres;
2074: it is a great pleasure to thank them for these stimulating exchanges.
2075: Useful discussions with Olivier Rivoire and Guilhem Semerjian
2076: are also gratefully acknowledged.
2077:
2078: %
2079: %***********************************************************
2080: %
2081: \appendix
2082:
2083: \section{Variational principle for frustrated kernels}
2084: \label{app:Variational}
2085:
2086:
2087: This appendix is devoted to the proof of Lemma \ref{lemma:Variational}.
2088: It is convenient to introduce some notations. If
2089: $A(x,y)$, $x,y\in\{1,\dots,q\}$ is a symmetric matrix and
2090: $\eta_1(x)$, $\eta_2(x)$, $x\in\{1,\dots,q\}$ are two vectors,
2091: we shall write $A\eta_1(x)\equiv\sum_y A(x,y)\eta_1(y)$ and
2092: $\eta_1 A\eta_2\equiv\sum_y A(x,y)\eta_1(x)\eta_2(y)$ .
2093: Furthermore, if $\Q$ is a distribution over $\M$ we will
2094: denote by $\T\Q$ the distribution obtained
2095: by using Eq.~(\ref{eq:1RSB_fixed}): $\T\Q$ is the left hand side
2096: of Eq.~(\ref{eq:1RSB_fixed}) when in the right hand side $\Q^*$
2097: has been substituted by $\Q$.
2098: Finally, given
2099: $\eta_1,\eta_2\in\M$, we let
2100:
2101: %
2102: \begin{eqnarray}
2103: %
2104: \Delta(\eta_1,\eta_2) \equiv \left(\frac{\eta_1\pi\eta_2}{\eo\pi\eo}\right)
2105: \log\left(\frac{\eta_1\pi\eta_2}{\eo\pi\eo}\right)\, .
2106: %
2107: \end{eqnarray}
2108: %
2109:
2110: We also write $\eta\ed \Q$ when $\eta$ has distribution $\Q$.
2111: We first derive two simple lemmas.
2112: %
2113: \begin{lemma}\label{remark1}
2114: %
2115: Let $\Q^*$ and $\Q$ be two consistent distributions over $\M$ and
2116: $\Sigma^*(t) \equiv\Sigma((1-t)\Q^*+t\Q)$. Then
2117: %
2118: \begin{eqnarray}
2119: %
2120: -\frac{1}{(k+1)}\left.\frac{\de\Sigma^*}{\de t}\right|_0 =
2121: \E\left\{\Delta(\nu,\eta_2')-\Delta(\nu,\eta_2)-\Delta(\eta_1,\eta_2')
2122: +\Delta(\eta_1,\eta_2)\right\}\, ,
2123: %
2124: \label{eq:der_result}
2125: \end{eqnarray}
2126: %
2127: where the expectation is taken with respect to the independent random variables
2128: $\eta_1,\eta_2 \ed \Q^*$, $\eta_2'\ed \T\Q^*$ and $\nu\ed \Q$.
2129: %
2130: \end{lemma}
2131: %
2132: \prooft
2133: Elementary calculus yields
2134: %
2135: \begin{eqnarray}
2136: %
2137: -\frac{1}{(k+1)}\left.\frac{\de\Sigma^*}{\de t}\right|_0 =\psi(1)-\psi(0)\, ,
2138: %
2139: \label{eq:derivee}
2140: \end{eqnarray}
2141: %
2142: where
2143: %
2144: \begin{eqnarray}
2145: %
2146: \psi(t) & \equiv &-\E\left\{\left(\frac{\nu^t\pi\eta}{\eo\pi\eo}\right)
2147: \log\!\left(\frac{\nu^t\pi\eta}{\eo\pi\eo}\right)\right\}+\\
2148: && +\E\left\{ \left(\frac{\sum_x\pi\nu^t(x)\prod_{i=1}^k
2149: \pi\eta_i(x)}{\sum_x\prod_{i=0}^k \pi\eo(x)}\right)\log\!
2150: \left(\frac{\sum_x\pi\nu^t(x)\prod_{i=1}^k
2151: \pi\eta_i(x)}{\sum_x\prod_{i=0}^k \pi\eo(x)}\right)\right\}\, .\nonumber
2152: %
2153: \end{eqnarray}
2154: %
2155: Here $\eta,\eta_1,\dots,\eta_k\ed \Q^*$ and $\nu^t\ed (1-t)\Q^*+t\Q$
2156: are independent random variables. The first term, when integrated on $t$,
2157: gives the contribution $\E\left\{\Delta(\eta_1,\eta_2)-\Delta(\nu,\eta_2)\right\}$
2158: to (\ref{eq:der_result}).
2159:
2160: As for the second term,
2161: observing that
2162: $\prod_{i=1}^k \pi\eta_i(x) = z(\{\eta_i\})\, \F(\eta_1,\dots,\eta_k)$,
2163: it can be rewritten as
2164: %
2165: \begin{eqnarray}
2166: %
2167: q^{k-1}\E\left\{z(\{\eta_i\}) \left[ \frac{\nu^t\pi\F(\eta_1\dots\eta_k)}
2168: {\eo\pi\eo}\right]\log\left[ \frac{\nu^t\pi\F(\eta_1\dots\eta_k)}
2169: {\eo\pi\eo}\right] \right\}+\nonumber\\
2170: q^{k-1}\E\left\{z(\{\eta_i\}) \left[ \frac{\nu^t\pi\F(\eta_1\dots\eta_k)}
2171: {\eo\pi\eo}\right]\log\left[ q^{k-1}z(\{\eta_i\})\right] \right\}\, .
2172: %
2173: \end{eqnarray}
2174: %
2175: Since $\E\nu^t(x) =\eo(x)$, the second term is $t$-independent
2176: and does not contribute to (\ref{eq:derivee}). The first term is equal
2177: to $\E\Delta(\nu^t,\eta'_2)$ where $\eta'_2\ed\T\Q^*$.
2178: %
2179: \endproof
2180:
2181: %
2182: \begin{lemma}\label{remark2}
2183: %
2184: Let $\eta_1,\eta_2\in\M$ and define $\delta\eta_i(x) = \eta_i(x)-\eo(x)$.
2185: If $\pi(x,y)=\pi_*-\pih(x,y)$ is a frustrated kernel, then
2186: %
2187: \begin{eqnarray}
2188: %
2189: |\delta\eta_1\pih\delta\eta_2|\le \eo\pi\eo\, .\label{eq:remark2}
2190: %
2191: \end{eqnarray}
2192: %
2193: \end{lemma}
2194: %
2195: \prooft
2196: %
2197: Since $\pih$ is positive definite, $\phi\pih\psi\equiv\sum_{x,y}
2198: \pih(x,y)\phi(x)\psi(y)$ is a well defined scalar product.
2199: Cauchy-Schwarz inequality implies
2200: %
2201: \begin{eqnarray}
2202: %
2203: |\delta\eta_1\pih\delta\eta_2|\le\sqrt{(\delta\eta_1\pih\delta\eta_1)
2204: (\delta\eta_2\pih\delta\eta_2)}\le \max\left\{(\delta\eta_1\pih\delta\eta_1),
2205: (\delta\eta_2\pih\delta\eta_2)\right\}\,.
2206: %
2207: \end{eqnarray}
2208: %
2209: Therefore it is sufficient to prove Eq.~(\ref{eq:remark2})
2210: for $\delta\eta_1=\delta\eta_2=\delta\eta$. Let $\eta(x) = \eo(x)+
2211: \delta\eta(x)$. Since $\pi(x,y),\eta(x)\ge 0$,
2212: and $\sum_x\delta\eta(x) = 0$, we have
2213: %
2214: \begin{eqnarray}
2215: %
2216: 0\le\eta\pi\eta = \eo\pi\eo+\delta\eta\pi\delta\eta=\eo\pi\eo-
2217: \delta\eta\pih\delta\eta\, .
2218: %
2219: \end{eqnarray}
2220: %
2221: \endproof
2222: %
2223:
2224: We can now turn to the proof of Lemma \ref{lemma:Variational}. In the
2225: following, given $\eta\in\M$, we define $\delta\eta(x)\equiv
2226: \eta(x)-\eo(x)$. Obviously we have
2227: %
2228: \begin{eqnarray}
2229: %
2230: \Delta(\eta_1,\eta_2) = \left(1-\frac{\delta\eta_1\pih\delta\eta_2}{\eo\pi\eo}\right)
2231: \log\!\left(1-\frac{\delta\eta_1\pih\delta\eta_2}{\eo\pi\eo}\right)\, .
2232: %
2233: \end{eqnarray}
2234: %
2235: Because of Lemma \ref{remark2} we can expand this expression in an
2236: absolutely convergent series
2237: %
2238: \begin{eqnarray}
2239: %
2240: \Delta(\eta_1,\eta_2) = -\frac{\delta\eta_1\pih\delta\eta_2}{\eo\pi\eo}
2241: +\sum_{n=2}^{\infty} C_n\,
2242: \left(\frac{\delta\eta_1\pih\delta\eta_2}{\eo\pi\eo}\right)^n\, ,
2243: %
2244: \end{eqnarray}
2245: %
2246: where $C_n\equiv 1/n(n-1)>0$. If $\eta_1\ed \Q_1$ and $\eta_2\ed\Q_2$ are
2247: independent random variables with consistent distributions, we get
2248: %
2249: \begin{eqnarray}
2250: %
2251: \E\Delta(\eta_1,\eta_2) = \sum_{n=2}^{\infty} C_n\,q^n
2252: \phi_1^{(n)}\pih^{\otimes n}\phi_2^{(n)}\, ,
2253: \label{eq:Expansion}
2254: %
2255: \end{eqnarray}
2256: %
2257: where $\pih^{\otimes n}(x_1\dots x_n;y_1\dots y_n) \equiv \pih(x_1,y_1)
2258: \cdots \pih(x_n,y_n)$ is the $n$-fold
2259: tensor product of $\pih$, $\phi_a^{(n)}(x_1\dots x_n)
2260: \equiv \E\{\delta\eta_a(x_1)\cdots \delta\eta_a(x_n)\}$
2261: are the moments of the distribution $\Q_a$, and
2262: %
2263: \begin{eqnarray}
2264: %
2265: \phi_1^{(n)}\pih^{\otimes n}\phi_2^{(n)} \equiv \sum_{x_1\dots x_n}
2266: \sum_{x_1\dots x_n} \pih^{\otimes n}(x_1\dots x_n;y_1\dots y_n)
2267: \phi_1^{(n)}(x_1\dots x_n)\phi_2^{(n)}(x_1\dots x_n)\, .
2268: %
2269: \end{eqnarray}
2270: %
2271:
2272: Now consider Remark \ref{remark1} and take $\Q = \T\Q^*$. We get
2273: %
2274: \begin{eqnarray}
2275: %
2276: -\frac{1}{(k+1)}\left.\frac{\de\Sigma^*}{\de t}\right|_0 =
2277: \E\left\{\Delta(\eta_1',\eta_2')-\Delta(\eta_1',\eta_2)-\Delta(\eta_1,\eta_2')
2278: +\Delta(\eta_1,\eta_2)\right\}\, ,
2279: %
2280: \end{eqnarray}
2281: %
2282: where $\eta_1,\eta_2\ed\Q^*$ and $\eta_1',\eta_2'\ed\T\Q^*$. Applying
2283: Eq.~(\ref{eq:Expansion}) we get
2284: %
2285: \begin{eqnarray}
2286: %
2287: -\frac{1}{(k+1)}\left.\frac{\de\Sigma^*}{\de t}\right|_0 =
2288: \sum_{n=2}^{\infty} C_n\,q^n
2289: (\phi_{\T}^{(n)}-\phi^{(n)})\pih^{\otimes n}
2290: (\phi_{\T}^{(n)}-\phi^{(n)})\, ,\label{eq:LastProofVariational}
2291: %
2292: \end{eqnarray}
2293: %
2294: where $\phi^{(n)}$ and $\phi^{(n)}_{\T}$ denote the moments (respectively)
2295: of $\Q^*$ and $\T\Q^*$. Since $\pih$ is positive definite,
2296: $\pih^{\otimes n}$ is positive definite as well and therefore the
2297: right hand side is a sum of non-negative terms. In order for this right hand side
2298: to vanish, each of the terms must vanish, which implies
2299: $\phi_{\T}^{(n)}=\phi^{(n)}$ for each $n$. But, since $\Q^*$ and
2300: $\T\Q^*$ have bounded support, this implies $\Q^* = \T\Q^*$, which
2301: is false by hypothesis. Therefore the right hand side of
2302: Eq.~(\ref{eq:LastProofVariational}) is strictly positive and
2303: $\left.\frac{\de\Sigma^*}{\de t}\right|_0 <0$ as desired. \endproof
2304: %
2305: %***********************************************************
2306: %
2307: \bibliographystyle{alpha}
2308: %\bibliography{bookref}
2309:
2310: \newcommand{\etalchar}[1]{$^{#1}$}
2311: \begin{thebibliography}{MPWZ02}
2312:
2313: \bibitem[AS04]{AldousSteele}
2314: David Aldous and J.~Michael Steele.
2315: \newblock The {O}bjective {M}ethod: {P}robabilistic {C}ombinatorial
2316: {O}ptimization and {L}ocal {W}eak {C}onvergence.
2317: \newblock In H.~Kesten, editor, {\em Probability on discrete structures}.
2318: Springer, New York, 2004.
2319:
2320: \bibitem[BG05]{BanGam}
2321: Antar Bandyopadhyay and David Gamarnik.
2322: \newblock Counting without sampling. New algorithms for enumeration problems using statistical physics.
2323: \newblock {\tt arXiv:math.PR/0510471}
2324:
2325: \bibitem[BB04]{BoBi04}
2326: Jean-Philippe Bouchaud and Giulio Biroli.
2327: \newblock On the {Adam-Gibbs-Kirkpatrick-Thirumalai-Wolynes} scenario for the
2328: viscosity increase in glasses.
2329: \newblock {\em J.~Chem.~Phys.}, 121:7347--7354, 2004.
2330:
2331: \bibitem[BMP{\etalchar{+}}03]{Brauenstein03}
2332: Alfredo Braunstein, Roberto Mulet, Andrea Pagnani, Martin Weigt, and Riccardo
2333: Zecchina.
2334: \newblock Polynomial iterative algorithms for coloring and analyzing random
2335: graphs.
2336: \newblock {\em Phys.~Rev.~E}, 68:036702, 2003.
2337:
2338: \bibitem[BRZ95]{BleherRuizZagrebnov}
2339: Pavel~M. Bleher, Jean Ruiz, and Valentin~A. Zagrebnov.
2340: \newblock On the purity of limiting {G}ibbs state for the {I}sing model on the
2341: {B}ethe lattice.
2342: \newblock {\em J. Stat. Phys.}, 79:473--482, 1995.
2343:
2344: \bibitem[BW03]{Winkler03}
2345: Grahm~R. Brightwell and Peter Winkler.
2346: \newblock Gibbs extremality for the hard-core model on a {B}ethe lattice.
2347: \newblock {\em submitted}, 1, 2003.
2348:
2349: \bibitem[CCST86]{Chayes86}
2350: Jennifer~T. Chayes, Lincoln Chayes, James.~P. Sethna, and David.~J. Thouless.
2351: \newblock A {M}ean {F}ield {S}pin {G}lass with {S}hort-{R}ange {I}nteractions.
2352: \newblock {\em Commun.~Math.~Phys.}, 106:41--89, 1986.
2353:
2354: \bibitem[CT91]{CoverThomas}
2355: Thomas~M. Cover and Joy~A. Thomas.
2356: \newblock {\em Elements of Information Theory}.
2357: \newblock Wiley Interscience (?), New York, 1991.
2358:
2359: \bibitem[EKPS00]{EvaKenPerSch}
2360: William Evans, Claire Kenyon, Yuval Peres, and Leonard~J. Schulman.
2361: \newblock Broadcasting on trees and the {I}sing model.
2362: \newblock {\em Ann. Appl. Probab.}, 10:410--433, 2000.
2363:
2364: \bibitem[FP84]{Peruggi83}
2365: Gabriella~Monroy Fulvio~Peruggi, Francesco~di~Liberto.
2366: \newblock The potts model on bethe lattices. i. general results.
2367: \newblock {\em J.~Math.~Phys.}, 16:811--817, 1984.
2368:
2369: \bibitem[Geo88]{Georgii}
2370: Hans-Otto Georgii.
2371: \newblock {\em {G}ibbs {M}easures and {P}hase {T}ransition}.
2372: \newblock de Gruyter, Berlin, 1988.
2373:
2374: \bibitem[KS66a]{KestenStigum2}
2375: Harry Kesten and Bernt~P. Stigum.
2376: \newblock Additional limit theorems for decomposable multi-dimesional
2377: {G}alton-{W}atson processes.
2378: \newblock {\em Ann. Math. Statist.}, 37:1463--1481, 1966.
2379:
2380: \bibitem[KS66b]{KestenStigum1}
2381: Harry Kesten and Bernt~P. Stigum.
2382: \newblock Limit theorems for decomposable multi-dimesional {G}alton-{W}atson
2383: processes.
2384: \newblock {\em J. Math. Anal. Appl.}, 17:309--338, 1966.
2385:
2386: \bibitem[Mar03]{Martin03}
2387: James B. Martin
2388: \newblock Reconstruction Thresholds on Regualr Trees
2389: \newblock {\em Discrete Mathematics and Theoretical Computer Science }{\bf AC}, 2003, 191-204
2390:
2391: \bibitem[MM05]{ReconstrNext}
2392: Marc M\'ezard and Andrea Montanari.
2393: \newblock Unpublished, 2005.
2394:
2395: \bibitem[Mos98]{MosselRec}
2396: Elchanan Mossel.
2397: \newblock Recursive reconstruction on periodic trees.
2398: \newblock {\em Rand.~Struct.~Alg.}, 13:81--97, 1998.
2399:
2400: \bibitem[Mos01]{Mossel01}
2401: Elchanan Mossel.
2402: \newblock Reconstruction on trees: beating the second eigenvalue.
2403: \newblock {\em Ann. Appl. Probab.}, 11:285--300, 2001.
2404:
2405: \bibitem[Mos04]{Mossel03}
2406: Elchanan Mossel.
2407: \newblock Survey: {I}nformation flows on trees.
2408: \newblock In J.~Nestril and P.~Winkler, editors, {\em Graphs, Morphisms and
2409: Statistical Physics. DIMACS Series in Discrete Mathematics and Theoretical
2410: Computer Science}. American Mathematical Society, 2004.
2411:
2412: \bibitem[MP01]{MP_Bethe}
2413: Marc M\'ezard and Giorgio Parisi.
2414: \newblock The {B}ethe lattice {S}pin glass revisited.
2415: \newblock {\em Eur. Phys. J. B}, 20:217--233, 2001.
2416:
2417: \bibitem[MP03a]{MP_Bethe_T0}
2418: Marc M{\'e}zard and Giorgio Parisi.
2419: \newblock The cavity method at zero temperature.
2420: \newblock {\em J.~Stat.~Phys.}, 111:1--34, 2003.
2421:
2422: \bibitem[MP03b]{MosselPeres01}
2423: Elchanan Mossel and Yuval Peres.
2424: \newblock Information flow on trees.
2425: \newblock {\em Ann. Appl. Probab.}, 13:817--844, 2003.
2426:
2427: \bibitem[MPWZ02]{Mulet02}
2428: Roberto Mulet, Andrea Pagnani, Martin Weigt, and Riccardo Zecchina.
2429: \newblock Coloring random graphs.
2430: \newblock {\em Phys.~Rev.~Lett.}, 89:268701, 2002.
2431:
2432: \bibitem[MS05a]{MoSe05}
2433: Andrea Montanari and Guilhem Semerjian.
2434: \newblock On the dynamics of the glass transition on {B}ethe lattices.
2435: \newblock {\em {\tt cond-mat/0509366}}, 2005.
2436:
2437: \bibitem[MS05b]{MosselSteel05}
2438: Elchanan Mossel and Mike Steel.
2439: \newblock How much can evolved characters tell us about the tree that generated
2440: them?
2441: \newblock In O.~Gascuel, editor, {\em Mathematics of Evolution and Phylogeny}.
2442: Oxford University Press, 2005.
2443:
2444: \bibitem[Per83]{Peruggi84b}
2445: Fulvio Peruggi.
2446: \newblock Probability measures and {H}amiltonian models on {B}ethe lattices:
2447: {II}. {T}he solution of thermal and configurational problems.
2448: \newblock {\em J.~Phys.~A:~Math.~Gen.}, 25:3316--3323, 1983.
2449:
2450: \bibitem[Per84]{Peruggi84a}
2451: Fulvio Peruggi.
2452: \newblock Probability measures and {H}amiltonian models on {B}ethe lattices:
2453: {I}. {P}roperties and construction of {MRT} probability measures.
2454: \newblock {\em J.~Math.~Phys.}, 25:3303--3315, 1984.
2455:
2456: \bibitem[TACA73]{Abou-Chacra}
2457: David~J.~Thouless R.~Abou-Chacra and Philip~W. Anderson.
2458: \newblock A selfconsistent theory of localization.
2459: \newblock {\em J.~Phys.~C}, 6:1734--1752, 1973.
2460:
2461: \bibitem[Shi96]{Shiryaev}
2462: Albert~N. Shiryaev.
2463: \newblock {\em {P}robability}.
2464: \newblock Springer, New York, 1996.
2465:
2466: \bibitem[TJ02]{Tatikonda}
2467: Sekhar Tatikonda and Michael I. Jordan
2468: \newblock {\em Loopy Belief Propagation and Gibbs Measures}.
2469: \newblock 18th Conference on Uncertainty in Artificial Intelligence, August
2470: 2002 (Edmonton, Canada)
2471:
2472: \end{thebibliography}
2473:
2474:
2475: %\end{thebibliography}
2476:
2477:
2478: \end{document}
2479: