1: \documentclass[pre,epsfig,aps]{revtex4}
2: \usepackage{epsfig}
3: \usepackage{fancyheadings}
4: \usepackage{amstext,amsmath,amssymb}
5:
6: \begin{document}
7: \title{Introduction to Phase Transitions in Random Optimization Problems}
8: \author{R. Monasson}
9: \address{Laboratoire de Physique Th\'eorique de l'ENS\\ 24 rue
10: Lhomond, 75005 Paris, France}
11: \begin{abstract}
12: Notes of the lectures delivered in Les Houches during
13: the Summer School on Complex Systems (July 2006).
14: \end{abstract}
15: \maketitle
16:
17: \section{Introduction}
18:
19: \subsection{Preamble}
20:
21: The connection between the statistical physics of disordered systems
22: and optimization problems
23: in computer science dates back from twenty years at least
24: \cite{Me87}. After all zero temperature statistical physics is
25: simply the search for the state with minimal energy, while
26: the main problem in combinatorial optimization is to look for the
27: configurations of parameters minimizing some cost function (the length
28: of a tour in the traveling salesman problem (TSP), the number of violated
29: constraints in constrained satisfaction problems, ...) \cite{Pa83}.
30: Yet, despite the beautiful studies of the average
31: properties of the TSP, Graph partitioning, Matching, ..., based on
32: the recently developed mean-field spin glass theory \cite{Me87},
33: a methodological gap between the fields could not be bridged\cite{Fu85}.
34: In statistical physics statements are usually made on the
35: properties of samples given some quenched disorder
36: distribution such as the typical number of solutions, minimal
37: energy ...
38: In optimization, however, one is interested in solving one (or
39: several) particular instances of a problem, and needs efficient
40: ways to do so, that is, requiring a computational effort growing
41: not too quickly with the number of data defining the instance.
42: Knowing precisely the typical properties for a given,
43: academic distribution of instances does not help much to
44: solve practical cases.
45:
46:
47: At the beginning of the nineties practitionners in artificial
48: intelligence realized that classes of random constrained satisfaction problems
49: used as artificial benchmarks for search algorithms exhibited abrupt
50: changes of behaviour when some control parameter were finely tuned
51: \cite{Mi92}. The most celebrated example was random K-Satisfiability,
52: where one looks for a solution to a set of random logical constraints
53: over a set of Boolean variables. It appeared that, for large sets of
54: variables, there was a critical
55: value of the number of constraints per variable below which there
56: almost surely existed solutions, and above which solutions were
57: absent. An important feature was that search algorithms performances
58: drastically worsened in the vicinity of this critical ratio.
59:
60: This phenomenon, strongly reminiscent of phase transitions in
61: condensed matter physics, led to a revival of the
62: interface between statistical physics and computer science, which has
63: not vanished yet. The purpose of the present lecture is to introduce
64: the non specialist reader to the concepts and techniques required to
65: understand the literature in the field. For the sake of simplicity
66: the presentation will be limited to one
67: computational problem, namely, linear systems of Boolean equations.
68: A good reason to do so is that this problem concentrates most of the features
69: encountered in other optimization problems, while being
70: technically simpler to study. In addition it is closely related to
71: error-correcting codes in communication theory, see lectures by
72: A. Montanari and R. Urbanke in the present book.
73: Extension to other problems will be mentioned in the conclusions.
74:
75: The lecture is divided into three parts. Sections 1 and 2 are
76: devoted to the presentation of the model and of elementary concepts
77: related to phase transitions e.g. finite-size scaling, large
78: deviations, critical exponents, symmetry breaking, ... Sections 3
79: and 4 expose the specific statistical mechanics techniques and concepts
80: developed in disordered systems to deal with highly interacting and
81: random systems, namely the replica and cavity approaches. Finally
82: Section 5 focuses on dynamics and the study of search
83: algorithms.
84:
85:
86: \subsection{Linear systems of Boolean equations}
87:
88: Linear systems of Boolean equations look very much like their
89: well known counterparts for integer-valued variables, except that
90: equalities are defined modulo two.
91: Consider a set of $N$ Boolean variables $x_i$ with indices $i=1,\ldots
92: ,N$. Any variable shall be False (F) or True (T). The sum of two variables,
93: denoted by $+$, corresponds to the logical exclusive OR between these
94: variables defined through,
95: \begin{eqnarray} \label{sumrule}
96: F + T &=& T + F = T \quad , \nonumber \\
97: F + F &=& T + T = F \quad .
98: \end{eqnarray}
99: In the following we shall use an alternative representation of the above
100: sum rule. Variables will be equal to 0 or 1, instead of $F$ or $T$
101: respectively. Then the $+$ operation
102: coincides with the addition between integer numbers modulo two.
103:
104: The following is a linear equation involving three variables,
105: \begin{equation} \label{xoreq01}
106: x_1 + x_2 + x_3 = 1 \quad .
107: \end{equation}
108: Four among the $2^3=8$ assignments of $(x_1,x_2,x_3)$ satisfy
109: the equation: $(1,0,0)$, $(0,1,0)$, $(0,0,1)$ and $(1,1,1)$.
110: A Boolean system of equations is a set of Boolean equations
111: that have to be satisfied together. For instance,
112: the following Boolean system involving four variables
113: \begin{equation} \label{xoreq02}
114: \left\{ \begin{array} {l}
115: x_1 + x_2 + x_3 = 1 \\
116: x_2 + x_4 = 0 \\
117: x_1 + x_4 = 1
118: \end{array} \right.
119: \end{equation}
120: has two solutions: $(x_1,x_2,x_3,x_4)=(1,0,0,0)$ and $(0,1,0,1)$.
121: A system with one or more solutions is called satisfiable. A trivial
122: example of an unsatisfiable Boolean system is
123: \begin{equation} \label{xoreq03}
124: \left\{ \begin{array} {l}
125: x_1 + x_2 + x_3 = 1 \\x_1 + x_2 + x_3 = 0
126: \end{array} \right. \qquad .
127: \end{equation}
128: Determining whether a Boolean system admits
129: an assignment of the Boolean variables satisfying all the
130: equations constitutes the XORSAT
131: (exclusive OR Satisfaction) problem. In the following,
132: we shall restrict for some reasons to be clarified in Section~\ref{secbasic}
133: to K-XORSAT, a variant of XORSAT where each Boolean equation include $K$
134: variables precisely.
135:
136: K-XORSAT belongs to the class P of polynomial problems \cite{Pa83}.
137: Determining whether a system is satisfiable or not
138: can be achieved by the standard Gaussian
139: elimination algorithm in a time (number of elementary
140: operations) bounded from above by some constant times
141: the cube of the number of bits necessary to store the
142: system\footnote{The storage space is $K$ times the number of equations
143: times the number of bits necessary to label a variable, that is, the
144: logarithm of the number of variables appearing in the system.}\cite{Pa83}.
145:
146: If the decision version of K-XORSAT is easy its optimization version
147: is not. Assume you are given a system $F$, run the Gauss procedure and
148: find that it is not satisfiable. Determining the maximal number
149: $M_S(F)$ of satisfiable equations is a very hard problem. Even
150: approximating this number is very hard. It is known that there is no
151: approximation algorithm (unless P=NP) for XORSAT with ratio $r>\frac
152: 12$, that is, guaranteed to satisfy at least $r\times M_S(F)$
153: equations for any $F$. But $r=\frac 12$ is achieved, on average, by
154: making a random guess\footnote{Any equation is satisfied by half of
155: the configurations of a variables, so a randomly chosen
156: configuration satisfies on average $\frac M2 \ge \frac {M_S(F)}2$
157: equations.}!
158:
159: \subsection{Models for random systems}\label{secmodel}
160:
161: There are many different ways of generating random Boolean systems.
162: Perhaps the simplest one is the following, called
163: {\em fixed-size ensemble}. To build an equation we pick up
164: uniformly at random $K$ distinct indices among the $N$ ones, say,
165: ${i_1},{i_2}$ and ${i_k}$. Then we consider the equation
166: \begin{equation}
167: x_{i_1} + x_{i_2} + \ldots + x_{i_k} = v \ .
168: \end{equation}
169: The second member, $v$, is obtained by tossing a coin:
170: $v=0$ or $v=1$ with equal probabilities (one half) and independently of the
171: indices of the variables in the first member.
172: The process is repeated $M$ times, without correlation between
173: equations to obtain a system with $M$ equations.
174:
175: Another statistical ensemble is the {\em fixed-probability ensemble}.
176: One scans the set of all $H=2 {N\choose K}$ equations one after the other.
177: Each equation is added to the system with probability $p$, discarded with
178: probability $1-p$. Then a system with, on average, $p\,H$ equations
179: (without repetition) is obtained. In practice one chooses $p=\frac MH$ to
180: have the same (average) number of equations as in the fixed-size
181: ensemble.
182:
183: The above distributions are not the only possible
184: ones. However they are easy to implement on a computer, are
185: amenable to mathematical studies, and last but not least,
186: lead to a surprisingly rich phenomenology. One of the key
187: quantities which exhibits an interesting behaviour is
188: \begin{eqnarray}
189: P_{SAT} (N,\alpha) &=& \mbox{Probability that a system of random
190: K-XORSAT with} \nonumber \\
191: && \mbox{ $N$ variables and $M=\alpha\, N$ equations is satisfiable}
192: \nonumber \ ,
193: \end{eqnarray}
194: which obviously depends on $K$ and the statistical ensemble.
195: Given $N$ $P_{SAT}$ is a decreasing function of $\alpha$.
196: We will see that, in the infinite size limit (and for $K\ge 2$),
197: the decrease is abrupt
198: at some well defined ratio, defining a phase transition between Satisfiable
199: and Unsatisfiable phase \cite{Cr99}.
200: The scope of the lecture is to give some tools
201: to understand this transition and some related phenomena.
202:
203: % basics
204:
205: \section{Basic concepts: overview of static phase transitions in K-XORSAT}
206: \label{secbasic}
207:
208: In this Section we introduce the basic concepts necessary to the
209: study of random K-XORSAT. It turns out that even the $K=1$ case,
210: trivial from a computer science point of view (each equation
211: contains a single variable!), can be used as an illustration to
212: important concepts such as scaling and self-averageness. Ideas
213: related to the percolation phase transition and random graphs are
214: illustrated on the $K=2$ case. Finally the
215: solution space of 3-XORSAT model exemplifies the notion of clusters
216: and glassy states.
217:
218: \subsection{Finite-size scaling (I): scaling function}\label{secsf1}
219:
220: Figure \ref{proba1-fig}(left) shows the probability $P_{SAT}$ that a randomly
221: extracted 1-XORSAT formula is satisfiable as a function of the ratio
222: $\alpha$, and for sizes $N$ ranging from 100 to 1000. We see that
223: $P_{SAT}$ is a decreasing function of $\alpha$ and $N$.
224:
225: Consider the subformula made of the $n_i$ equations with first member equal to
226: $x_i$. This formula is always satisfiable if $n_i=0$ or $n_i=1$. If
227: $n_i\ge 2$ the formula is
228: satisfiable if and only if all second members are equal (to 0, or to
229: 1), an event with probability $(\frac 12)^{n_i-1}$
230: decreasing exponentially with the number
231: of equations. Hence we have to consider the following variant of the celebrated
232: Birthday problem\footnote{The Birthday problem is a classical
233: elementary probability problem: given a class with $M$ students, what is the
234: probability that at least two of them have the same birthday date?
235: The answer for $M=25$ is $p\simeq 57\%$, while a much lower value is
236: expected on intuitive grounds when $M$ is much smaller than the number
237: $N=365$ of days in a year.}. Consider a year with a number $N$
238: of days, how should scale the number $M$ of students in a
239: class to be sure that no two students have the same birthday date?
240: \begin{equation}
241: \bar p =\prod_{i=0}^{M-1}\left(1-\frac iN\right) = \exp
242: \left(-\frac{M(M-1)}{2N} + O(M^3/N^2) \right) \ .
243: \end{equation}
244: Hence we expect a cross-over from large to small $\bar p$ when $M$
245: crosses the scaling regime $\sqrt N$. Going back to the 1-XORSAT model
246: we expect $P_{SAT}$ to have a non zero limit value when the number of
247: equations and variables are both sent to infinity at a fixed ratio
248: $y=M/\sqrt N$. In other words, random 1-XORSAT formulas with $N$
249: variables, $M$ equations or with, say, $100\times N$ variables,
250: $10\times M$ equations should have roughly the same probabilities of
251: being satisifiable. To check this hypothesis we replot the data in
252: Figure~\ref{proba1-fig} after multiplication of the abscissa of each
253: point by $\sqrt N$ (to keep $y$ fixed instead of $\alpha$). The
254: outcome is shown in the right panel of Figure~\ref{proba1-fig}. Data
255: obtained for various sizes nicely collapse on a single limit curve
256: function of $y$.
257:
258: The calculation of this limit function, usually called scaling
259: function, is done hereafter in the fixed-probability
260: 1-XORSAT model where the number of equations
261: is a Poisson variable of mean value $\bar M=y\sqrt N$. We will
262: discuss the equivalence between the fixed-probability
263: and the fixed-size ensembles later.
264: In the fixed-probability ensemble the numbers $n_i$ of occurence of each
265: variable $x_i$ are independent Poisson variables with average value
266: $\bar M/N=y/\sqrt N$. Therefore the probability of satisfaction is
267: \begin{eqnarray} \label{psatp}
268: P_{SAT} ^{p} (N,\alpha=\frac y{\sqrt N})
269: &=& \left[ e^{- y/\sqrt N} \left( 1+ \sum _{n\ge 1}
270: \frac {(y/\sqrt N)^n}{n!} \left(\frac 12\right)^{n-1}\right) \right] ^N
271: \nonumber \\ &=& \left[ 2 e^{-y/(2\sqrt N)}-e^{-y/\sqrt N}\right]^N \ ,
272: \end{eqnarray}
273: where the $p$ subscript denotes the use of the fixed-probability ensemble.
274: We obtain the desired scaling function
275: \begin{equation} \label{sca1}
276: \Phi _1(y)\equiv \lim _{N\to\infty} \ln P^p_{SAT} (N,\alpha=\frac y{\sqrt
277: N}) = -\frac{y^2}4 \ ,
278: \end{equation}
279: in excellent agreement with the rescaled data of
280: Figure~\ref{proba1-fig} (right) \cite{Cr03a}.
281:
282: \begin{figure}[t]
283: \includegraphics*[width=5.truecm,angle=0]{fig211.eps}
284: \includegraphics*[width=5.truecm,angle=0]{fig212.eps}
285: \caption{Left: Probability that a random 1-XORSAT formula is satisfiable as
286: a function of the ratio $\alpha$ of equations per variable, and for
287: various sizes $N$. Right: same data as in the left panel after the
288: horizontal rescaling $\alpha \to \alpha \times \sqrt N = y$; note the use of
289: a log scale for the vertical axis. The dashed line
290: shows the scaling function $\Phi_1(y)$ (\ref{sca1}).}
291: \label{proba1-fig}
292: \end{figure}
293:
294: \subsection{Self-averageness of energy and entropy}\label{secselfav}
295:
296: Let us now consider random 1-XORSAT formulas at a finite ratio
297: $\alpha$, and ask for the distribution of the minimal fraction of
298: unsatisfied equations, hereafter called ground state (GS) energy $e_{GS}$.
299: For simplicity we work in the fixed-probability ensemble again. The
300: numbers $n_i^0, n_i^1$ of, respectively, $x_i=0, x_i=1$ are
301: independent Poisson variables with mean $\frac \alpha 2$. The
302: minimal number of unsatisfied equations is clearly $\min (n_i^0,
303: n_i^1)$. The GS energy is the sum (divided by $M$) of $N$ such
304: i.i.d. variables; from the law of large number it almost surely
305: converges towards the average value
306: \begin{equation}
307: e_{GS} (\alpha) = \frac 12 \left( 1 - e^{-\alpha} I_0(\alpha) -
308: e^{-\alpha} I_1(\alpha) \right) \ ,
309: \end{equation}
310: where $I_\ell$ denotes the $\ell^{th}$ modified Bessel function.
311: In other words almost all formulas have the same GS energy in the
312: infinite $N$ limit, a property called self-averageness in physics, and
313: concentration in probability.
314:
315: How many configurations of variables realize have minimal energy?
316: Obviously a variable is free (to take 0 or 1 value) if $n_i^0=n_i^1$,
317: and is frozen otherwise. Hence the number of GS configurations is
318: ${\cal N} = 2^{N_f}$ where $N_f$ is the number of free variables.
319: Call
320: \begin{equation}
321: \rho = \sum _{n\ge 0} e^{-\alpha} \left(
322: \frac{\alpha} 2\right)^n \frac 1{(n!)^2} = e^{-\alpha}\; I_0(\alpha)
323: \end{equation}
324: the probability that a variable is free. Then $N_f$ is a binomial
325: variable with parameter $\rho$ among $N$; it is sharply concentrated
326: around $\overline{N_f}=\rho\, N$ with typical fluctuations of the
327: order of $N^{1/2}$. As a consequence, the GS entropy per variable,
328: $s_{GS} = (\log {\cal N})/N$, is self-averaging and almost surely
329: equal to its average value $s_{GS} = \rho \log 2$.
330:
331:
332: Self-averageness is the very useful property. It allows us to study
333: the average value of a random variable, instead of its full
334: distribution. We shall use it in Section \ref{secreplicas}
335: and also in the analysis of algorithms of Section \ref{secuc}.
336: This property is not restricted to XORSAT but was proven to
337: hold for the GS energy \cite{Br93} and entropy \cite{Mo07} of other
338: optimization problems.
339:
340: Not all variables are self-averaging of course. A straightforward
341: example is the number ${\cal N}$ of GS configurations itself.
342: Its $q^{th}$ moment reads
343: $\overline{{\cal N}^q} = \left(1-\rho + \rho \; 2^q\right)^N$
344: where the overbar denotes the average over the formulas. We see that
345: $\overline{{\cal N}^q} \gg (\overline{{\cal N}})^q$: ${\cal N}$
346: exhibits large fluctuations and is not concentrated around its
347: average. Very rare formulas with atypically large number
348: $N_f$ of free variables contribute more to the $q^{th}$ moment than
349: the vast majority of formulas, and spoil the output. This is the very
350: reason we will need the introduction of the replica approach in
351: Section \ref{secreplicas}.
352:
353:
354: \subsection{Large deviations for $P_{SAT}$ (I): 1-XORSAT}\label{seclargedev}
355:
356: As we have seen in the previous sections 1-XORSAT formulas with a
357: finite ratio $\alpha$ are unsatifiable with high probability {\em
358: i.e.} equal to unity in the infinite $N$ limit. For finite but large
359: $N$ there is a tiny probability that a randomly extracted formula is
360: actually satisifiable. A natural question is to characterize the
361: `rate' at which $P_{SAT}$ tends to zero as $N$ increases (at fixed
362: $\alpha$). Answering to such questions is the very scope of large
363: deviation theory (see \ref{applargedev} for an elementary
364: introduction). Looking for events with very small probabilities
365: is not only interesting from an academic point of view, but can also
366: be crucial in practical applications. We will see in Section
367: \ref{secescape} that
368: the behaviour of some algorithms is indeed dominated by rare events.
369:
370: Figure~\ref{proba1b-fig} shows minus the logarithm of $P_{SAT}$, divided by
371: $N$, as a function of the ratio $\alpha$ and for various sizes
372: $N$. Once again the data corresponding to different sizes collapse on
373: a single curve, meaning that
374: \begin{equation}
375: P_{SAT} (N,\alpha) = e^{-N\; \omega _1(\alpha) + o(N)} \ .
376: \end{equation}
377: Decay exponent $\omega_1$ is called rate function in probability
378: theory. We can derive its value in the fixed-probability ensemble from
379: (\ref{psatp}) with $y=\alpha\times \sqrt N$, with the immediate result
380: \begin{equation} \label{sca2p}
381: \omega_1 ^p (\alpha) = \alpha - \ln \big( 2\; e^{\alpha/2} -1 \big) \ .
382: \end{equation}
383: The agreement with numerics is very good for small ratios,
384: but deteriorates as $\alpha$ increases. The reason is simple. In the
385: fixed-probability ensemble the number $M$ of equations is not fixed but may
386: fluctuate around the average value $\bar M =\alpha N$. The
387: ratio $\tilde \alpha =M/N$, is with high probability equal to
388: $\alpha$, but large deviations ($\tilde \alpha \ne \alpha)$ are
389: possible and described by the rate
390: function\footnote{$M$ obeys a Poisson law with parameter $\bar
391: M$. Using Stirling formula, $$e^{-\bar M} \frac{\bar M ^M}{M!}
392: \simeq e^{-\alpha N } (\tilde \alpha N)^{\alpha N}{\sqrt {2\pi N}}
393: \left( \frac e{\alpha N}\right)^{\alpha N} = e^{-N \, \Omega
394: (\tilde\alpha | \alpha) + o(N)}\ ,$$ where $\Omega$ is defined in
395: (\ref{defOmega}). },
396: \begin{equation}\label{defOmega}
397: \Omega (\tilde \alpha | \alpha ) = \tilde \alpha -\alpha - \alpha\;
398: \ln (\alpha/\tilde \alpha ) \ .
399: \end{equation}
400: However the probability that a random 1-XORSAT formula with $M$
401: equations is satisfiable is also exponentially small in $N$, with a
402: rate function $\omega_1(\alpha)$ increasing with $\alpha$. Thus, in
403: the fixed-probability ensemble, a trade-off is found between ratios $\tilde
404: \alpha$ close to $\alpha$ (formulas likely to be generated) and close
405: to 0 (formulas likely to be satisfiable). As a result
406: the fixed-probability rate function is
407: \begin{equation} \label{sca2}
408: \omega_1 ^p (\alpha) = \min _{\tilde \alpha } \big[ \omega _1 (\tilde
409: \alpha) + \Omega (\tilde \alpha | \alpha )\big] \ ,
410: \end{equation}
411: and is smaller than $\omega_1(\alpha)$.
412: It is an easy check that the optimal ratio
413: $\tilde \alpha ^*=\alpha /(2-e^{-\alpha/2}) < \alpha$ as
414: expected. Inverting (\ref{sca2}) we deduce the rate function
415: $\omega_1$ in the
416: fixed-size ensemble, in excellent agreement with numerics
417: (Figure~\ref{proba1b-fig}). This example underlines that
418: thermodynamically equivalent ensembles
419: have to be considered with care as far as rare events are concerned.
420:
421: Remark that, when $\alpha \to 0$, $\tilde \alpha =\alpha +
422: O(\alpha ^2)$, and $\omega _1 ^p(\alpha)=\omega_1 (\alpha)
423: +O(\alpha^3)$. This common value coincides with the scaling
424: function $- \Phi_1(\alpha)$ (\ref{sca1}). This identity is expected on
425: general basis (Section \ref{secld2}) and justifies the agreement between the
426: fixed-probability scaling function and the numerics based on the fixed-size
427: ensemble (Figure~\ref{proba1-fig}, right).
428:
429: \begin{figure}[t]
430: \includegraphics*[width=5.truecm,angle=0]{fig221.eps}
431: \caption{Same data as Figure\ref{proba1-fig} (left) with: logarithmic
432: scale on the vertical axis, and rescaling by $-1/N$. The scaling
433: functions $\omega_1$ (\ref{sca2})
434: and $\omega_1 ^p$ (\ref{sca2p}) for, respectively, the fixed-size
435: and fixed-probability ensembles are shown.}
436: \label{proba1b-fig}
437: \end{figure}
438:
439: \subsection{Percolation in random graphs}\label{secrg}
440:
441: Though 1-XORSAT allowed us to understand some general features of random
442: optimization problems it is very limited due to the absence of interactions
443: between variables. A more interesting problem is 2-XORSAT where
444: every equation define a joint constraint on two variables.
445: Formulas of 2-XORSAT can be represented by a graph with $N$ vertices
446: (one for each variable), and $\alpha N$ edges.
447: To each equation of the type $x_i+x_j=e$
448: corresponds an edge linking vertices $i$ and $j$, and carrying
449: 0 or 1 label (the value $e$ of the second member).
450: Depending on the input model chosen (Section \ref{secmodel}) multiple edges
451: are present or not.
452:
453: As the formula is random so is graph.
454: Figure~\ref{graphealeatoire} shows examples of graphs obtained
455: for various values of $\alpha$. Notice the
456: qualitative change of structure of graphs when
457: the ratio $\alpha$ varies from low values (graphs are mostly made
458: of small isolated trees) to higher ones (a large part of vertices
459: are now connected together). This change is known as the percolation
460: transition in physics, or the appearance of a giant component in
461: mathematics literature.
462:
463: \begin{figure}
464: \begin{center}
465: \includegraphics[width=70pt,angle=0]{fig241.eps}
466: \hskip 1cm
467: \includegraphics[width=70pt,angle=0]{fig242.eps}
468: \hskip 1cm
469: \includegraphics[width=70pt,angle=0]{fig243.eps}
470: \caption{Examples of random graphs generated at fixed number $M=\alpha N$ of
471: edges (fixed-size model without repetition). All graph include $N=20$
472: vertices (grey dots). The average degree of valency, $2\alpha$,
473: is equal to $0.5$ ({\bf A}), $1$ ({\bf B}), and
474: $2$ ({\bf C}). The labels of the vertices have been permuted to
475: obtain planar graphs, {\em i.e.} avoid crossing of edges.}
476: \label{graphealeatoire}
477: \end{center}
478: \end{figure}
479:
480: Before reviewing some of the aspects of the percolation transition
481: let us mention an important fact on the valency of vertices.
482: As a result of the randomness of the graph generation process,
483: each node share edges with a variable number of neighboring vertices.
484: In the large $N$ limit the degree $v$ of a vertex, {\em i.e.}
485: the number of its neighbors,
486: is a Poisson variable with mean $2\alpha$,
487: \begin{equation} \label{poissonconnec}
488: \mbox{Proba}[v] = e^{-2\alpha}\; \frac{(2\alpha)^v}{v!} \ .
489: \end{equation}
490: For instance the fraction of isolated vertices is $e^{-2\alpha}$.
491: The average degree of a vertex, $c=2\alpha$, is called
492: connectivity.
493:
494: It is natural to decompose the graphs into its connected subgraphs,
495: called components. Erd\"os and R\'enyi were able in 1960 to characterize the
496: distribution of sizes of the largest component \cite{Bo89},
497: \begin{itemize}
498: \item When $c<1$, the largest component includes
499: $\sim \ln N/(c-1-\ln c)$ vertices with high probability.
500: Most components include only
501: a finite number of vertices, and are trees {\em i.e.} contain no circuit.
502: \item For $c=1$ the largest component contain $O(N^{2/3})$ vertices.
503: \item When $c>1$ there is one giant component containing $\sim \gamma
504: (c) N$ vertices; the others components are small {\em i.e.} look
505: like the components in the $c<1$ regime.
506: The fraction of vertices in the giant component is the unique
507: positive solution of
508: \begin{equation} \label{sizegiant}
509: 1-\gamma = e^{-c \, \gamma} \ .
510: \end{equation}
511: It is a non analytic function of $c$, equal to 0 for $c\le 1$, and
512: positive above, tending to unity when $c$ increases.
513: \end{itemize}
514:
515: The phenomenon taking place at $c=1$ is an example of (mean-field)
516: percolation transition. We now give a hand-waving derivation of
517: (\ref{sizegiant}).
518: Consider a random graph $G$ over $N$ vertices, with connectivity
519: $c$. Add a new vertex $A$ to the graph to obtain $G'$. If we want $G'$
520: to be drawn from the same distribution as $G$, a number $v$ of edges
521: must be attached to $A$, where $v$ an integer--valued random number
522: following the Poisson distribution (\ref{poissonconnec}).
523: After addition of $A$, some connected components of $G$ will merge
524: in $G'$. In particular, with some probability $p_v$, $A$ will be part
525: of the giant component of $G'$. To estimate $p_v$, we note that this
526: event will not happen if and only if none of the $v$ neighbors of $A$ in
527: $G'$ belongs to the giant component of $G$. Thus,
528: \begin{equation} \label{processusgraphe}
529: 1-p_v = (1-\gamma ) ^v \qquad ,
530: \end{equation}
531: where $\gamma$ is the size (fraction of vertices) of the giant component.
532: Summing both sides of (\ref{processusgraphe})
533: over the distribution (\ref{poissonconnec}) for $v$, and
534: asserting that the change in size of the giant component between
535: $G$ and $G'$ is $o(1)$ for large $N$, we obtain (\ref{sizegiant}).
536:
537:
538: The above derivation illustrates an ubiquitous idea in
539: probability and
540: statistical physics, which could be phrased as follows: `if a system
541: is very large, its statistical properties should be, in some sense,
542: unaffected by a small increase in size'. This idea will be useful, in
543: a more sophisticated context, in Section \ref{seccavity}.
544:
545: \subsection{Sat/Unsat transition in 2-XORSAT} \label{secsatunsat2}
546:
547: \begin{figure}[t]
548: \includegraphics*[width=5.truecm,angle=0]{fig251.eps}
549: \caption{Probability that a random 2-XORSAT formula is satisfiable as
550: a function of the ratio $\alpha$ of equations per variable, and for
551: various sizes $N$. The full line is the asymptotic analytical
552: formula (\ref{psaty2}).}
553: \label{proba2-fig}
554: \end{figure}
555:
556: Figure \ref{proba2-fig} shows the probability $P_{SAT}$
557: that a randomly extracted
558: 2-XORSAT formula is satisfiable as function of $\alpha$, and for
559: various sizes $N$. It appears that $P_{SAT}$ drops quickly
560: to zero for large $N$ when $\alpha$ reaches
561: the percolation threshold $\alpha_c = \frac 12$. For ratios smaller than
562: $\alpha _c$ the probability of satisfaction is positive, but smaller
563: than unity.
564:
565: Take $\alpha < \frac 12$. Then the random graph $G$ associated to a random
566: 2-XORSAT formula is non percolating, and made of many small components.
567: Identical components (differing only by a relabelling of
568: the variables) may appear several times, depending on their topology.
569: For instance consider a connected graph $G'$ made of $E$ edges and $V$
570: vertices.
571: The average number of times $G'$ appears in $G$ is a function of $E$ and
572: $V$ only,
573: \begin{equation}
574: N_{E,V} = {N \choose V} \left( \frac{2\alpha}N \right)^E
575: \left( 1 - \frac{2\alpha}N \right)^{\frac {V(V-1)}2 + V(N-V)}
576: \end{equation}
577: since any vertex in $G'$ can establish edges with other vertices in
578: $G'$, but is not allowed to be connected to any of the $N-V$ outside vertices.
579: When $N$ is very large compared to $E,V$ we have
580: \begin{equation} \label{avtree}
581: N_{E,V} \simeq N^{V-E} \; \frac{ (2\alpha)^E}{V!}\;
582: e^{-2\alpha\, V} \ .
583: \end{equation}
584: Three cases should distinguished, depending on the value of $V-E$:
585: \begin{itemize}
586: \item{$V-E=1$}: this is the largest value compatible with connectedness,
587: and corresponds to the case of trees. From (\ref{avtree}) every
588: finite tree has of the order of $N$ copies in $G$.
589: \item {$V-E=0$}: this correspond to trees with one additional edge, that is, to graphs
590: having one cycle (closed loop). The average number of unicyclic graphs is,
591: from (\ref{avtree}), finite when $N\to\infty$.
592: \item{$V-E\le -1$}: the average number of components with more than one cycle
593: vanishes in the large $N$ limit; those graphs are unlikely to be
594: found and can be ignored\footnote{The probability that such a graph exists is
595: bounded from above by the average number, see \ref{appmoment}.}.
596: \end{itemize}
597: Obviously a 2-XORSAT formula with tree
598: structure is always satisfiable\footnote{Start from one leaf, assign the
599: attached variable to 0, propagate to the next variable according to
600: the edge value, and so on, up to the completion of the tree.}. Hence
601: dangerous subformulas, as far as satisfiability is concerned, are
602: associated to unicyclic graphs. A simple thought shows that a unicyclic
603: formula is
604: satisfiable if and only if the number of edges carrying label 1
605: along the cycle is
606: even. Since the values attached to the edges (second members in the
607: formula) are uncorrelated
608: with the topology of the subgraph (first members)
609: each cycle is satisfiable with probability one half.
610: We end up with the simple formula
611: \begin{equation} \label{psatx2}
612: P_{SAT} (N,\alpha) = \langle 2 ^{- C (G)} \rangle
613: \end{equation}
614: where $C(G)$ denotes the number of cycles in $G$, and
615: $\langle .\rangle$ the average over $G$. For a reason which will become
616: clear below let us classify cycles according to their length $L$.
617: How many cycles of length $L$ can we construct? We have to choose first
618: $L$ vertices among $N$, and join them one after the order according
619: to some order. As
620: neither the starting vertex nor the direction along the cycle matter,
621: the average number of $L$-cycles is
622: \begin{equation} \label{deflaml}
623: N_L =\frac{ N(N-1)\ldots (N-L+1)}{2L} \times \left( \frac{2\alpha}N \right)^L
624: \to \Lambda _L = \frac{(2\alpha)^L}{2L}\ .
625: \end{equation}
626: when $N\to\infty$. As the emergence of a cycle between $L$ vertices is a
627: local event (independent of the environment) we expect the number of
628: $L$-cycles to be Poisson distributed in the large $N$ limit
629: with parameter $\Lambda_L$. This statement can actually be proven, and
630: extended to any finite collection of cycles of various lengths\cite{Bo89}:
631: in the infinite size limit, the joint distribution of the numbers of cycles
632: of lengths $1,2, \ldots, L$ is the product of Poisson laws with
633: parameters $\Lambda_1, \Lambda _2, \ldots,\Lambda _L$
634: calculated in (\ref{deflaml}). The probability of satisfaction
635: (\ref{psatx2}) therefore converges to
636: \begin{equation}
637: \lim _{N\to\infty} P_{SAT} (N,\alpha)
638: = \prod_{L\ge L_0} \left\{ \sum _{C\ge 0}
639: e^{-\Lambda _L} \; \frac{{(\Lambda _L/2)}^C}{C!}
640: \right\} = \prod _{L\ge L_0} e^{-\Lambda_L/2}
641: \end{equation}
642: where $L_0$ is the minimal cycle length. In normal random graphs $L_0=3$
643: since triangles are the shortest cycles. However in our 2-XORSAT model
644: any equation, or more precisely, any first member can appear twice or
645: more, hence $L_0=2$. We conclude that \cite{Cr03a}
646: \begin{equation} \label{psaty2}
647: \lim _{N\to\infty} P_{SAT} (N,\alpha)
648: = e^{\alpha/2}\; (1-2\alpha) ^{\frac 14} \qquad
649: \mbox{when} \qquad \alpha < \alpha_c =\frac 12 \ .
650: \end{equation}
651: The agreement of this result
652: with the large size trend coming out from numerical
653: simulations is visible in Figure \ref{proba2-fig}. As $P_{SAT}$ is a
654: decreasing function of $\alpha$ it remains null for all ratios larger than
655: $\alpha_c$. The non analyticity of $P_{SAT}$ at $\alpha_c$
656: locates the Sat/Unsat phase transition of 2-XORSAT.
657:
658: It is an implicit assumption of statistical physics that asymptotic
659: results of the kind of (\ref{psaty2}), rigorously valid in the $N\to
660: \infty$ limit, should reflect with good accuracy the finite but large
661: $N$ situation. An inspection of Figure \ref{proba2-fig} shows this is
662: indeed the case. For instance, for ratio $\alpha =.3$, (\ref{psaty2})
663: cannot be told from the probability of satisfaction
664: measured for formulas with $N=100$ variables. This statement
665: does not hold for $\alpha =.4$, where the agreement between infinite size
666: theory and numerics sets in when $N = 1000$ at least. It appears
667: that such finite-size effects become bigger and bigger as $\alpha$
668: gets closer and closer to the Sat/Unsat threshold. This issue, of broad
669: importance in the
670: context of phase transitions and the pratical application of asymptotic
671: results, is studied in Section \ref{secfs2}.
672:
673: \subsection{Large deviations for $P_{SAT}$ (II): bounds in
674: the Unsat phase of 2-XORSAT.} \label{secld2}
675:
676: Consider ratios $\alpha > \alpha_c$.
677: The giant components of the corresponding formulas contain an
678: extensively large number of independent cycles, so we expect
679: from (\ref{psatx2}) that the probability of satisfaction is
680: exponentially small in $N$, $P_{SAT} =\exp (-N \omega_2(\alpha)+o(N))$.
681: Lower and upper bounds to the rate function $\omega _2$ can be obtained
682: from, respectively, the first and second moment inequalities
683: described in \ref{appmoment}. Denoting by ${\cal N}$ the number of solutions
684: of a formula $P_{SAT}$ is the probability that ${\cal N}\ge 1$, and
685: is bracketed according to (\ref{ineq}).
686:
687: To calculate the first moment of ${\cal N}$ remark that an equation
688: is satisfied by one half of the configurations. This result remains true
689: for a restricted set of configurations when we average over the possible
690: choices of (the second member of) the equation. The average number
691: of solutions is thus $2^N/2^M$, from which we get
692: \begin{equation} \label{lowerom2}
693: \omega _2 (\alpha) \ge (\alpha -1) \ln 2 \ .
694: \end{equation}
695: This lower bound is useless for $\alpha <1$ since $\omega_2$ is positive
696: by definition. As for the upper bound we need to calculate the
697: second moment $\langle {\cal N}^2\rangle$ of ${\cal N}$. As equations are
698: independently drawn
699: \begin{equation} \label{n2}
700: \langle {\cal N} ^2\rangle = \sum _{X,Y} q (X,Y) ^M \,
701: \end{equation}
702: where the sum is carried out over the pairs $X,Y$ of configurations
703: of the $N$ variables, and $q(X,Y)$ is the
704: probability that both $X$ and $Y$ satisfies the same randomly
705: drawn equation. $q$ can be easily expressed in terms of
706: the Hamming distance $d$ between
707: $X$ and $Y$, defined as the fraction of variables having opposite
708: values in $X$ and $Y$. The general
709: expression for K-XORSAT is\footnote{The
710: equation is satisfied if the number of its variables taking opposite
711: values in $Y$ as in $X$ is even.
712: definition of $d$ the probability (over its index $i$) that a variable
713: takes different value in $X$ and $Y$ is $d$. Hence expression
714: (\ref{qk})
715: for $q(d)$. Beware of the $O(\frac 1N)$ corrections to this expression
716: e.g. if variable $x_1\ne y_1$ (which happens with probability $d$) then
717: the probability that $x_2\ne y_2$ is $(dN-1)/(N-1) = d + (1-d)/-N-1)$.
718: Those corrections are relevant for the calculation of Gaussian
719: fluctuations around the saddle-point (\ref{appfluctu}).}
720: \begin{equation} \label{qk}
721: q (d)=\frac 12( 1 - (1-2d)^K) \,
722: \end{equation}
723: and we specialize in this section to $K=2$.
724: Going back to (\ref{n2}) we can sum over $Y$ at fixed $X$, that is,
725: over the distances $d$ taking multiple values of $\frac 1N$ with the
726: appropriate binomial multiplicity, and then sum over $X$ with the result
727: \begin{equation} \label{n2p}
728: \langle {\cal N} ^2\rangle = 2^{N} \sum _{d} {N \choose N\,d} \; q(d)^M
729: = \exp( N\, \max_{d \in [0;1]}A(d,\alpha) +o(N))
730: \end{equation}
731: in the large $N$ limit, where
732: \begin{equation} \label{upperom2}
733: A(d,\alpha) = (2\alpha-1) \; \ln 2
734: -d \ln d -(1-d) \ln (1-d) +\alpha \ln q(d) \ .
735: \end{equation}
736: For $\alpha < \frac 12$ the maximum of $A$ is located in $d^*=\frac 12$,
737: and equal to $A^*=0$. When $\alpha > \frac 12$, $A$ has two global
738: maxima located in $d^*(\alpha) <\frac 12$ and $1-d^*(\alpha)$, with
739: equal value $A^*(\alpha) >0$.
740:
741: We plot in Figure \ref{proba2b-fig} the lower (\ref{lowerom2}) and
742: upper bounds to the rate function,
743: \begin{equation} \label{om2}
744: \omega _2 (\alpha) \le 2(1-\alpha)\ln 2 - \max_{d \in [0;1]}A(d,\alpha)
745: \end{equation}
746: from (\ref{ineq}).
747: At large ratio both bounds asymptotically match, proving that
748: $\omega_2 (\alpha) = (\alpha-1) \ln 2 + O(e^{-2\alpha})$.
749: As the ratio departs from its threshold value by
750: $\epsilon = \alpha - \alpha _c$ the upper bound grows quadratically,
751: $A^*(\alpha_c + \epsilon )\simeq \frac 34
752: \epsilon^2 + O(\epsilon ^3)$. Numerics suggest that the increase of
753: the rate function is slower,
754: \begin{equation} \label{om2clth}
755: \omega _2 (\alpha_c + \epsilon) \simeq \Omega\; \epsilon ^3 +
756: O(\epsilon ^4)\ ,
757: \end{equation}
758: for some constant $\Omega\simeq 1$ (Figure \ref{proba2b-fig}).
759: We will see in Section \ref{secreplicas} that
760: a sophisticated statistical physics technique, called the replica
761: method, actually predict this scaling with $\Omega = \frac {32}{27}$.
762: Actually the rate function can be estimated with the replica approach
763: for any ratio $\alpha$ with the result shown in Figure \ref{proba2b-fig}.
764:
765: \begin{figure}[t]
766: \includegraphics*[width=5.truecm,angle=0]{fig261.eps}
767: \caption{Rate function $\omega_2 (\alpha)$ associated to the probability
768: of satisfaction of 2-XORSAT formulas with ratio $\alpha$. The dotted line
769: is the upper bound (\ref{upperom2}) and the dot-dashed line the lower bound
770: (\ref{lowerom2}). The full line is the output of the replica calculation of
771: Section \ref{secresu}, squares and circles represent numerical results for
772: $N=200$, 100 respectively from $10^6$ formulas.
773: Inset: $P_{SAT}$ as a function of the size $N$ at the
774: Sat/Unsat ratio. The slope $-\frac 1{12}$ (\ref{predic112}) is shown for
775: comparison.}
776: \label{proba2b-fig}
777: \end{figure}
778:
779: \subsection{Order parameter and symmetry breaking}\label{secop}
780:
781: What is the meaning of the Hamming distance $d^*(\alpha)$ appearing
782: in the calculation of the second moment of the number of solutions?
783: An easy guess would be the average distance between pairs
784: of solutions
785: \begin{equation} \label{dav}
786: d_{av}(\alpha) = \lim _{N\to\infty} \langle \frac{\sum _{X,Y\ \mbox{solutions
787: of}\ F} d(X,Y)}{{\cal N}(F)^2}\rangle_F
788: \end{equation}
789: where the average is taken over the satisfiable
790: formulas $F$ with ratio $\alpha$, and
791: $d(X,Y)$ denotes the (intensive) Hamming distance between two
792: solutions $X,Y$. However an inspection of the calculation of Section
793: \ref{secld2} shows that
794: \begin{equation} \label{davstar}
795: d^*(\alpha) = \lim _{N\to\infty} \frac{\langle \sum _{X,Y\ \mbox{solutions
796: of}\ F} d(X,Y)\rangle _F}{\langle {\cal N}(F)^2 \rangle_F} \ne
797: d_{av}(\alpha)\ .
798: \end{equation}
799: Actually, though $d^*(\alpha)$ is not the average distance
800: between solutions with the unbiased distribution over formulas, it
801: is the average distance for a biased distribution where
802: each formula is weighted with
803: \begin{equation}\label{biased}
804: w(F) = \frac{ {\cal N}(F)^2}{\sum_{F'} {\cal N}(F')^2}
805: \end{equation}
806: as can be readily checked upon insertion of $w(F)$ in the numerator
807: of (\ref{dav}). We will see in Section (\ref{secreplicas}) how
808: to calculate average properties with the unbiased measure.
809:
810: Even so definition (\ref{davstar}) (and (\ref{dav}) too) is sloppy.
811: If $X$ is a solution so is $-X$, the configuration where variables
812: values are flipped. Thus the average distance, whatever the weights
813: over formulas, is equal $\frac 12$ for any $N$! The difficulty comes from
814: the ambiguity in how the thermodynamic limit is taken, and
815: is the signature of spontaneous symmetry breaking. In the
816: low temperature phase of the Ising model
817: the magnetization is either $m^*>0$ or $-m^*<0$
818: if an external field $h$ with, respectively, positive or negative vanishing
819: amplitude is added prior to taking the infinite size limit.
820: In the present case what plays the role of the field is a coupling
821: between solutions as is well-known in splin-glass theory
822: \cite{Pa86}. Inserting $\exp[- N \, h\, d(X,Y)]$ in the numerator
823: of (\ref{davstar}) we obtain, when $N\to\infty$, $d^*$ if $h \to 0^+$
824: and $1-d^*$ if $h\to 0^-$. The density $\mu$ of probability
825: of distances $d$ between solutions, with the biased measure (\ref{biased}),
826: is concentrated below the Sat/Unsat threshold,
827: \begin{equation}
828: \mu (d) = \delta\big( d - \frac 12 \big) \qquad \mbox{for} \qquad
829: \alpha <\alpha _c \ ,
830: \end{equation}
831: and split into two symmetric peaks above the critical ratio,
832: \begin{equation}
833: \mu (d) = \frac 12 \delta\big( d - d^* \big) +
834: \frac 12 \delta\big( d - (1-d^*) \big)\qquad \mbox{for} \qquad
835: \alpha >\alpha _c \ .
836: \end{equation}
837: The concept of spontaneous symmetry breaking will play a key role
838: in our study of 3-XORSAT (Section \ref{secpin}).
839:
840: \subsection{Finite-size scaling (II): critical exponents}
841: \label{secfs2}
842:
843: Let us summarize what we have found about the probability of satisfying
844: random 2-XORSAT formulas in Section \ref{secsatunsat2}
845: and \ref{secld2}. Close to the transition we have from (\ref{psaty2}) and
846: (\ref{om2clth}),
847: \begin{equation} \label{bornesp2}
848: \ln P_{SAT} (N,\alpha_c +\epsilon) \simeq \left\{ \begin{array} {c c c}
849: \frac 14 \ln (-\epsilon) & \mbox{when} & \epsilon<0, N\to \infty\\
850: - \Omega\; N \; \epsilon ^3 & \mbox{when} & \epsilon>0, N\gg 1 \nonumber
851: \end{array}\right . \quad .
852: \end{equation}
853: The lesson of Section \ref{secsf1} is that $\ln P_{SAT}$ may
854: have a non trivial limit when $N\to \infty$, $\epsilon \to 0$ provided we
855: keep $y=\epsilon\, N^{\psi}$ constant. For 1-XORSAT the exponent
856: $\psi$ was found to be equal to $\frac 12$, and $\ln P_{SAT}$ to
857: converge to the scaling function $\Phi _1(y)$ (\ref{sca1}).
858: The situation is similar but slightly more
859: involved for 2-XORSAT. A natural assumption is to look for
860: the existence of a scaling function such that
861: \begin{equation}\label{sca2a}
862: \ln P_{SAT} ( N, \epsilon) \simeq N^\rho \;
863: \Phi _2 ( \epsilon \, N^\psi) \ .
864: \end{equation}
865: Let us see if (\ref{sca2a}) is compatible with the limiting
866: behaviours (\ref{bornesp2}). Fixing $\epsilon<0$ and sending $N\to\infty$
867: we obtain, for $y=\epsilon N^\psi \to -\infty$,
868: $\frac 14 \ln |y| - \frac \psi 4 \ln N$ for the l.h.s, and
869: $N^\rho \times \Phi_2(y)$ for the r.h.s. Hence $\rho=0$ as in the
870: 1-XORSAT case, but an additive correction is necessary, and we modify
871: scaling Ansatz (\ref{sca2a}) into
872: \begin{equation}\label{sca2b}
873: \ln P_{SAT} ( N, \epsilon) \simeq
874: \Phi _2 ( y= \epsilon \, N^\psi) - \frac{\psi}{4}\; \ln N .
875: \end{equation}
876: The above equation is now compatible with (\ref{bornesp2}) if
877: $\Phi_2(y) \sim \frac 14 \ln |y|$ when $y\to -\infty$.
878: Fixing now $\epsilon >0$ and sending $N$ to infinity we see that
879: (\ref{bornesp2}) is fulfilled if $\Phi_2(y) \sim
880: - \Omega\, y^3$ when $y\to +\infty$ and
881: \begin{equation} \label{valpsi}
882: \psi = \frac 13 \ .
883: \end{equation}
884: The above value for $\psi$ is expected from the study of random graphs
885: \cite{Bo89} and is related to the size
886: $N^{1-\psi}=N^{\frac 23}$ of the largest components at the percolation
887: threshold (Section \ref{secrg}).
888: $\psi$ is called
889: critical exponent and characterize the width of the critical region
890: of 2-XORSAT. Loosely speaking it means that a formula of with $N$
891: variables and $\frac N2+\Delta$ equations is 'critical' when $\Delta \sim
892: N^{\frac 23}$. This information will be useful for the analysis of
893: search algorithms in Section \ref{secuni}.
894:
895: A consequence of (\ref{sca2b},\ref{valpsi}) is that, right at the
896: threshold, the probability of satisfaction decays as\footnote{This
897: scaling is correct provided there is no diverging
898: e.g. $O(\ln \ln N)$ corrections to (\ref{sca2b}).}
899: \begin{equation} \label{predic112}
900: P_{SAT}\left(N, \alpha_c \right) \sim N^{-\frac 1{12}} \ .
901: \end{equation}
902: This scaling agrees with numerical experiments, though the
903: small value of the decay exponent makes an accurate check delicate
904: (Inset of Figure \ref{proba2b-fig}).
905:
906: \subsection{First and second moments inequalities for
907: the 3-XORSAT threshold}
908:
909: \begin{figure}[t]
910: \includegraphics*[width=5.truecm,angle=0]{fig291.eps}
911: \caption{Probability that a random 3-XORSAT formula is satisfiable as
912: a function of the ratio $\alpha$ of equations per variable, and for
913: various sizes $N$. The dotted line locates the threshold $\alpha_c
914: \simeq 0.918$.}
915: \label{proba3-fig}
916: \end{figure}
917:
918: Figure \ref{proba3-fig} shows the probability that a random 3-XORSAT
919: formula is satisfiable as a function of $\alpha$ for increasing sizes
920: $N$. It appears that formulas with ratio $\alpha < \alpha _c \simeq
921: 0.92$ are very likely to be satisfiable in the large $N$ limit,
922: while formulas with ratios beyond this critical value are almost
923: surely unsatisfiable. This behaviour is different from the 2-XORSAT
924: case (Figure \ref{proba2-fig})
925: in that $P_{SAT}$ seems to tend to unity below threshold.
926:
927: It is important to realize that, contrary to the 2-XORSAT case, the
928: Sat/Unsat transition is not related to connectivity percolation.
929: Consider indeed a variable, say, $x_1$. This variable appear, on
930: average, in $3\alpha$ equations. Each of those equations contain
931: other 2 variables. Hence the `connectivity' of $x_1$ is $c=6\alpha$,
932: which is larger than unity for $\alpha _p=\frac 16$. In the range
933: $[\alpha_p,\alpha_c]$ the formula is percolating but still satisfiable
934: with high probability. The reason is that cycles do not
935: hinder satisfiability as much as in the 2-XORSAT case.
936:
937: Use of the first and second moment inequalities (\ref{appmoment}) for
938: the number ${\cal N}$
939: of solutions provides us with upper and lower bounds to the
940: Sat/Unsat ratio $\alpha_c$. The calculation follows the same line as
941: the one of the 2-XORSAT case (Section \ref{secld2}). The
942: first moment $\langle {\cal N} \rangle =2 ^{N(1-\alpha)}$ vanishes for
943: ratios larger than unity, showing that
944: \begin{equation}
945: \alpha _c \le \alpha _1 =1 \ .
946: \end{equation}
947: This upper bound is definitely larger than the true threshold from the
948: numerical findings of Figure~\ref{proba3-fig}. We have already
949: encountered
950: this situation in 2-XORSAT: in the $\frac 12 < \alpha < 1$ range
951: formulas are unsatisfiable with probability one (when $N\to\infty$),
952: yet the average number of solutions is exponentially large! The reason
953: is, once more, that the average result is spoiled by rare, satisfiable
954: formulas with many solutions.
955:
956: As for the second moment expression (\ref{n2p},\ref{upperom2}) still holds
957: with $q(d)$ given by (\ref{qk}) with $K=3$. The absolute maximum of the
958: corresponding function $A(d,\alpha)$ is located in
959: $d^* = \frac 12$ when $\alpha < \alpha _2 \simeq 0.889$, and
960: $d^* < \frac 12$ when $\alpha > \alpha_2$. In the latter case
961: $\langle {\cal N}^2\rangle$ is exponentially larger than $\langle
962: {\cal N}\rangle ^2$, and the second moment inequality
963: (\ref{ineq}) does not give any information about $P_{SAT}$. In
964: the former case $\langle {\cal N}^2\rangle$ and
965: $\langle {\cal N}\rangle ^2$ are equivalent to exponential-in-$N$
966: order. It is shown in \ref{appfluctu} that their ratio actually tends to
967: one as $N\to\infty$. We conclude that formulas with
968: ratios of equations per variable
969: less than $\alpha_2$ are satisfiable with high probability in
970: the infinite size limit, or, equivalently \cite{Cr03b},
971: \begin{equation}
972: \alpha _c \ge \alpha _2 \simeq 0.889 \ .
973: \end{equation}
974: Unfortunately the lower and upper bounds do not match and the precise
975: value of the threshold remains unknown at this stage. We explain in
976: the next section how a simple preprocessing of the formula, before the
977: application of the first and second moment inequalities, can close the
978: gap, and shed light on the structure of the space of solutions.
979:
980: \subsection{Space of solutions and clustering}\label{seccluster}
981:
982: We start from a simple observation. Assume we have a formula $F$
983: of 3-XORSAT where a variable, say, $x$, appears only once, that is, in
984: one equation, say, $E:x+y+z=0$. Let us call $F'$ the subformula obtained
985: from $F$ after removal of equation $E$. Then the following statement
986: is true: {\em $F$ is satisfiable if and only if $F'$ is satisfiable.}
987: The proof is obvious: whatever the values of $y,z$ required to satisfy
988: $F'$ equation $E$ can be satisfied by an adequate choice of $x$, and
989: so can be the whole formula $F$.
990:
991: In a random 3-XORSAT formula $F$ with ratio $\alpha$ there are about
992: $N\times 3\alpha \, e^{-3\alpha}$ variables appearing only once in the
993: formula. Removal of those variables (and their equations) produces
994: a shorter formula with $O(N)$ less equations. Furthermore it may
995: happen that variables with multiple occurrences in the original
996: formula have disappeared from the output formula, or appear only
997: once. Hence the procedure can be iterated until no single-occurrence
998: variables are present. We are left with $F_2$, the largest
999: subformula (of the original formula) where every variable
1000: appears at least twice.
1001:
1002: Many questions can be asked: how many equations are left in
1003: $F_2$? how many variables does it involve? how many solutions does it
1004: have? Giving the answers requires a thorough
1005: analysis of the removal procedure, with the techniques exposed in Section
1006: \ref{secuc} \cite{Co03,Me03,Du02}.
1007: The outcome depends on the value of the ratio
1008: compared to
1009: \begin{equation}
1010: \alpha _d = \min _b -\frac{\log (1-b)}{3\, b^2} \simeq 0.8184\ldots
1011: \end{equation}
1012: hereafter called clustering threshold. With
1013: high probability when $N\to\infty$ $F_2$ is empty if $\alpha <
1014: \alpha_d$, and contains an extensive number
1015: of equations, variables when $\alpha > \alpha_d$. In the latter case
1016: calculation of the first and second moments
1017: of the number of solutions of $F_2$ shows that this number does
1018: not fluctuate around the value $e^{N\,
1019: s_{cluster}(\alpha)+o(N)}$ where
1020: \begin{equation}\label{scluster}
1021: s_{cluster} (\alpha) = (b - 3 \alpha \,b^2 + 2 \alpha \, b^3)\ln 2
1022: \end{equation}
1023: and $b$ is the strictly positive solution of the self-consistent equation
1024: \begin{equation} \label{eqb}
1025: 1-b=e^{-3\, \alpha\, b^2} \ .
1026: \end{equation}
1027: Hence $F_2$ is satisfiable if and only if $\alpha < \alpha _c$ defined
1028: through $s_{cluster} (\alpha_c)=0$, that is,
1029: \begin{equation}
1030: \alpha_c \simeq 0.9179 \ldots \ .
1031: \end{equation}
1032: This value is, by virtue of the equivalence between $F$ and $F_2$
1033: the Sat/Unsat threshold for 3-XORSAT, in excellent agreement with
1034: Figure~\ref{proba3-fig}.
1035:
1036: How can we reconstruct the solutions of $F$ from the ones of $F_2$?
1037: The procedure is simple. Start from one solution of $F_2$ (empty
1038: string if $\alpha < \alpha_d$). Then introduce back the last
1039: equation which was removed since it contained $n\ge 1$ single-occurrence
1040: variable. If $n=1$ we fix the value of this variable in a unique
1041: way. If $n=2$ (respectively $n=3$) there are 2 (respectively, 4) ways
1042: of assigning the reintroduced variables, defining as many solutions
1043: from our initial, partial solution. Reintroduction of equations one after
1044: the other according to the Last In -- First Out order gives us more and
1045: more solutions from the initial one, until we get a bunch of solutions
1046: of the original formula $F$. It turns out that the number of solutions
1047: created this way is $e^{N\, s_{in}(\alpha)+o(N)}$ where
1048: \begin{equation}\label{sin}
1049: s_{in} (\alpha) = (1-\alpha)\,\ln 2 - s_{cluster}(\alpha) \ .
1050: \end{equation}
1051: The above formula is true for $\alpha> \alpha_d$, and should be
1052: intended as $s_{in}(\alpha) =(1 - \alpha)\,\ln 2$ for $\alpha< \alpha_d$.
1053: These two entropies are shown in Figure \ref{entroxor}.
1054: The total entropy, $s^*(\alpha) = s_{in}(\alpha)+s_{cluster}(\alpha)$,
1055: is simply
1056: $(1-\alpha)\,\ln 2$ for all ratios smaller than the
1057: Sat/Unsat threshold. It shows no singularity at the clustering threshold.
1058: However a drastic change in the structure of the space of solutions
1059: takes place, symbolized in the phase diagram of Figure \ref{cluster-fig}:
1060: \begin{itemize}
1061: \item For ratios $\alpha<\alpha_d$ the intensive
1062: Hamming distance between two solutions
1063: is, with high probability, equal to $d=1/2$. Solutions thus differ
1064: on $N/2 +o(N)$ variables, as if they were statistically unrelated
1065: assignments of the $N$ Boolean variables.
1066: In addition the space of solutions enjoys some connectedness
1067: property. Any two solutions are connected by
1068: a path (in the space of solutions) along which successive solutions
1069: differ by a bounded number of variables. Losely speaking
1070: one is not forced to cross a big region prived of solutions
1071: when going from one solution to another.
1072: \item For ratios $\alpha > \alpha_d$
1073: the space of solutions is not connected any longer. It is made
1074: of an exponentially large (in $N$) number ${\cal N}_{clu}
1075: =e^{N\, s_{cluster}}$ of connected
1076: components, called clusters, each containing an exponentially large
1077: number ${\cal N}_{in}=e^{N\, s_{in}}$ of solutions. Two solutions belonging to
1078: different clusters lie apart at a Hamming distance $d_{clu}=1/2$ while,
1079: inside a cluster, the distance is $d_{in} < d_{clu}$. $b$ given by
1080: (\ref{eqb}) is the fraction of variables having the same value in all
1081: the solutions of a cluster (defined as the backbone).
1082: \end{itemize}
1083:
1084: We present in Sections \ref{secreplicas} and \ref{seccavity}
1085: statistical physics
1086: tools developed to deal with the scenario of Figure \ref{cluster-fig}.
1087:
1088: \begin{figure}
1089: \begin{center}
1090: \includegraphics[width=120pt,angle=0]{fig2101.eps}
1091: \end{center}
1092: \caption{Entropies (base 2 logarithms divided by
1093: size $N$) of the numbers of solutions and clusters as a function of the
1094: ratio $\alpha$. The entropy of solutions
1095: equals $1-\alpha$ for $\alpha<\alpha_c\simeq 0.918$.
1096: For $\alpha<\alpha_d\simeq 0.818$, solutions are uniformly
1097: scattered on the $N$-dimensional hypercube.
1098: At $\alpha_d$ the solution space discontinuously
1099: breaks into disjoint clusters. The entropies of clusters, $s_{cluster}$,
1100: and of solutions in each cluster, $s_{in}$, are such that
1101: $s_{cluster}+s_{in}=s$.
1102: At $\alpha_c$ the number of clusters stops being exponentially large
1103: ($s_{cluster}=0$). Above $\alpha_c$ there is almost surely no solution.}
1104: \label{entroxor}
1105: \end{figure}
1106:
1107:
1108: \begin{figure}
1109: \begin{center}
1110: \includegraphics[width=150pt]{fig2102.eps}
1111: \end{center}
1112: \caption{Phase diagram of 3-XORSAT. A `geometrical' phase
1113: transition takes place in the satisfiable phase at
1114: $\alpha_d\simeq 0.818$. At small ratios $\alpha<\alpha_d$
1115: solutions are uniformely
1116: scattered on the $N$-dimensional hypercube, with a typical normalized
1117: Hamming distance $d=\frac 12$. At $\alpha_d$ the solution space discontinuously
1118: breaks into disjoint clusters: the Hamming distance $d_{in}\simeq 0.14$
1119: between solutions inside a cluster is much smaller than
1120: the typical distance $d_{clu}=\frac 12$ between two clusters.}
1121: \label{cluster-fig}
1122: \end{figure}
1123:
1124: % replicas
1125:
1126: \section{Advanced methods (I): replicas} \label{secreplicas}
1127:
1128: \subsection{From moments to large deviations for the entropy}
1129: \label{secmoment}
1130:
1131: The analysis of Section \ref{secld2} has shown that the first, and second
1132: moments of the number ${\cal N}$ of solutions
1133: are dominated by rare formulas with a lot of solutions.
1134: Let us define the intensive entropy $s$ through ${\cal N} = e^{N\,
1135: s}$. As ${\cal N}$ is random (at fixed $\alpha,N$) so is $s$. We
1136: assume that the distribution of $s$ can be described, in the large
1137: size limit, by a rate function $\omega(s)$ (which depends on $\alpha$).
1138: Hence,
1139: \begin{equation}
1140: \langle {\cal N}^q \rangle = \int ds\; e^{-N \, \omega(s)} \times
1141: \left(e^{N\, s}\right)^q \sim \exp\big[N \max _{s} \big(q\, s -
1142: \omega (s) \big)\big]
1143: \end{equation}
1144: using the Laplace method. If we are
1145: able to estimate the leading behaviour of the $q^{th}$ moment of
1146: the number of solutions when $N$ gets large at fixed $\alpha$,
1147: \begin{equation}
1148: \langle {\cal N}^q \rangle \sim e^{N\, g(q)} \ ,
1149: \end{equation}
1150: then $\omega$ can be easily
1151: calculated by taking the Legendre transform of $g$. In particular
1152: the typical entropy is obtained by $s^* = \frac{dg}{dq}(q\to 0)$.
1153: This is the road we will follow below. We will show how $g(q)$ can be
1154: calculated when $q$ takes integer values, and then perform an analytic
1155: continuation to non integer $q$. The continuation leads to substantial
1156: mathematical difficulties, but is not uncommon in statistical physics
1157: e.g. the $q\to 1$ limit of the $q$-state Potts model to recover
1158: percolation, or
1159: the $n\to 0$ limit of the $O(n)$ model to describe self-avoiding walks.
1160:
1161: To calculate the $q^{th}$ moment we will have to average over the
1162: random components of formulas $F$, that is, the $K$-uplets of index
1163: variables in the first members and the $v=0,1$ second
1164: members. Consider now homogeneous formulas $F_h$ whose first
1165: members are randomly drawn in the same way as for $F$, but with all
1166: second members $v=0$. The number ${\cal N}_h$ of solutions of a
1167: homogeneous formula is always larger or equal to one. It is a simple
1168: exercise to show that
1169: \begin{equation} \label{momentxorequality}
1170: \langle {\cal N} ^{q+1} \rangle =
1171: 2^{N(1-\alpha)}\times \langle \big({\cal N}_h\big)^q \rangle \quad ,
1172: \end{equation}
1173: valid for any positive integer $q$\footnote{Actually the identity
1174: holds for $q=0$ too, and is known under the name of harmonic
1175: mean formula \cite{Al89}.}. Therefore it is sufficient to calculate
1176: the moments of ${\cal N}_h=e^{N\, g_h(q)}$ since
1177: (\ref{momentxorequality}) gives a simple identity between
1178: $g(q+1)$ and $g_h(q)$. This technical simplification has a
1179: deep physical meaning we will comment in Section \ref{secpin}.
1180:
1181: \subsection{Free energy for replicated variables}
1182:
1183: The $q^{th}$ power of the number of solutions to a homogeneous system
1184: reads
1185: \begin{equation}\label{qxor1}
1186: \big({\cal N}_h\big)^q = \left[ \sum _X \prod _{\ell=1}^M
1187: e_\ell ( X) \right] ^q = \sum _{X^1,X^2, \ldots , X^q}
1188: \prod _{\ell=1}^M \prod_{a=1}^q e_\ell ( X^a) \ ,
1189: \end{equation}
1190: where $e_\ell(X)$ is 1 if equation $\ell$ is satisfied by assignment
1191: $X$. The last sum runs over $q$ assignments $X^a$, with $a=1,2, \ldots ,
1192: q$ of the Boolean variables, called replicas of the original
1193: assignment $X$.
1194: It will turn useful to denote by $\vec x_i = (x_i^1,x_i^2, \ldots ,
1195: x_i^q)$ the $q$-dimensional vector whose components are the values of
1196: variable $x_i$ in the $q$ replicas. To simplify notations we consider
1197: the case $K=3$ only here, but extension to other values of $K$ is
1198: straightforward.
1199: Averaging over the instance, that is, the triplets of integers labelling
1200: the variables involved in each equation $\ell$, leads to the following
1201: expression for the $q^{th}$ moment,
1202: \begin{eqnarray} \label{qxor2}
1203: \langle\big({\cal N}_h\big)^q \rangle&=& \sum _{X^1,X^2, \ldots , X^q}
1204: \langle \prod_{a=1}^q e ( X^a) \rangle ^M \nonumber \\
1205: &=& \sum _{X^1,X^2, \ldots , X^q} \left[ \frac 1{N^3}
1206: \sum _{1\le i,j,k\le N} \delta _{\vec x_i + \vec x_j + \vec x_k }
1207: + O\bigg(\frac 1N\bigg)\right] ^M
1208: \end{eqnarray}
1209: where $\delta _{\vec x} =1$ if the compoments of $\vec x$ are all
1210: null mod. 2, and 0 otherwise.
1211: We now procede to some formal manipulations of the above equation
1212: (\ref{qxor2}).
1213:
1214: {\bf First step.}
1215: Be ${\cal X} = \{X^1,X^2,\ldots , X^q\}$ one of the $2^{\,qN}$ replica
1216: assignment. Focus on variable $i$, and its attached assignment
1217: vector, $\vec x_i$. The latter may be any of the $2^q$ possible vectors
1218: e.g. $\vec x_i = (1,0,1,0,0, \ldots , 0)$ if variable $x_i$ is equal
1219: to $0$ in all but the first and third replicas. The histogram of
1220: the assignments vectors given replica assignment ${\cal X}$,
1221: \begin{equation} \label{qxor2b}
1222: \rho \big( \vec x |{\cal X}\big) =
1223: \frac 1N \sum _{i=1}^N \delta _{\vec x- \vec x_i}
1224: \quad ,
1225: \end{equation}
1226: counts the fraction of assignments vectors $\vec x _i$ having value
1227: $\vec x$ when $i$ scans the whole set of variables from 1 to $N$.
1228: Of course, this histogram is normalised to unity,
1229: \begin{equation}\label{qxor2c}
1230: \sum _{\vec x} \rho \big( \vec x\big) = 1
1231: \quad ,
1232: \end{equation}
1233: where the sum runs over all $2^q$ assignment vectors.
1234: An simple but essential observation is that the r.h.s. of
1235: (\ref{qxor2}) may be rewritten in terms of the above histogram,
1236: \begin{equation} \label{qxor3}
1237: \frac 1{N^3}
1238: \sum _{1\le i,j,k\le N} \delta _{ \vec x_i + \vec x_j + \vec x_k }
1239: = \sum _{\vec x , \vec x'}
1240: \rho \big( \vec x \big) \;\rho \big( \vec x'\big) \;\rho \big( \vec
1241: x + \vec x' \big)
1242: \ .
1243: \end{equation}
1244: Keep in mind that $\rho$ in (\ref{qxor2b},\ref{qxor3}) depends on the
1245: replica assignement ${\cal X}$ under consideration.\\
1246:
1247: {\bf Second step.} According to (\ref{qxor3}), two
1248: replica assignments ${\cal X}_1$ and ${\cal X}_2$ defining the same
1249: histogram $\rho$ will give equal contributions to
1250: $\langle\big({\cal N}_h\big)^q\rangle$.
1251: The sum over replica assignments ${\cal X}$ can therefore be replaced
1252: over the sum over possible histograms provided the multiplicity ${\cal
1253: M}$ of the latter is taken properly into account.
1254: This multiplicity is also equal to the number of combinations
1255: of $N$ elements (the $\vec x_i$ vectors) into $2^q$ sets labelled
1256: by $\vec x$ and of cardinalities $N\, \rho(\vec x)$. We obtain
1257: \begin{equation}\label{qxor5}
1258: \langle \big({\cal N}_h\big)^q \rangle= \sum _{\{ \rho\}} ^{(norm)}
1259: e^{\displaystyle{\; N \; {\cal G}_h \big( \{\rho\},\alpha \big) + o(N) }}
1260: \quad ,
1261: \end{equation}
1262: where the $(norm)$ subscript indicates that the sum runs
1263: over histograms $\rho$ normalized according to (\ref{qxor2c}), and
1264: \begin{equation}\label{qxor6}
1265: {\cal G}_h \big( \{\rho\},\alpha \big) = - \sum _x \rho(x) \, \ln \rho(x) +
1266: \alpha\; \ln \bigg[ \sum _{\vec x , \vec x'}
1267: \rho \big( \vec x \big) \;\rho \big( \vec x'\big) \;
1268: \rho \big( \vec x +\vec x' \big)
1269: \bigg]
1270: \ .
1271: \end{equation}
1272: In the large $N$ limit, the sum in (\ref{qxor5}) is
1273: dominated by the histogram $\rho ^*$ maximizing the functional ${\cal G}_h$.
1274:
1275: {\bf Third step.}
1276: Maximisation of function ${\cal G}_h$
1277: over normalized histograms can be done within the Lagrange multiplier
1278: formalism. The procedure consists in considering the modified function
1279: \begin{equation}
1280: {\cal G}^{LM}_h \big( \{\rho\} , \lambda , \alpha \big) =
1281: {\cal G}_h \big( \{\rho\} , \alpha \big) + \lambda \; \bigg( 1 -
1282: \sum _{\vec x} \rho \big( \vec x\big) \bigg) \quad ,
1283: \end{equation}
1284: and first maximise ${\cal G}^{LM}_h$ with respect to
1285: histograms $\rho$ without caring about the normalisation constraint,
1286: and then optimise the result with respect to $\lambda$.
1287: We follow this procedure with ${\cal G}_h$ given by (\ref{qxor6}).
1288: Requiring that ${\cal G}^{LM}_h$ be maximal provides us with a set
1289: of $2^q$ coupled equations for $\rho^*$,
1290: \begin{equation} \label{spxor}
1291: \ln \rho ^*(\vec x ) + 1 +\lambda -3\;\alpha\ \frac
1292: {\displaystyle{ \sum _{\vec x' }
1293: \rho ^* \big( \vec x'\big) \;
1294: \rho ^*\big( \vec x+\vec x' \big) }}
1295: {\displaystyle{ \sum _{\vec x' , \vec x''}
1296: \rho ^*\big( \vec x' \big) \;\rho ^*\big( \vec x''\big) \;
1297: \rho ^*\big( \vec x'+\vec x '' \big) }}
1298: = 0 \ ,
1299: \end{equation}
1300: one for each assignment vector $\vec x$. The optimisation equation
1301: over $\lambda$ implies that $\lambda$ in (\ref{spxor}) is such that
1302: $\rho^*$ is normalised.
1303: At this point of the above and rather abstract calculation it may help
1304: to understand the interpretation of
1305: the optimal histogram $\rho^*$.
1306:
1307: \subsection{The order parameter}
1308:
1309: We have already addressed a similar question at the end
1310: of the second moment calculation in Section \ref{secop}.
1311: The parameter $d^*$ coming out from
1312: the calculation was the (weighted) average Hamming distance
1313: (\ref{davstar}) between two solutions of the same random instance.
1314: The significance of $\rho^*$ is identical. Consider $q'$ solutions
1315: labelled by $a=1,2,\ldots, q'$ of the same random and homogeneous
1316: instance and a variable, say, $x_i$. What is the probability,
1317: over instances and solutions, that this variable takes, for instance,
1318: value 0 in the first and fourth solutions, and 1 in all other solutions?
1319: In other words, what is the probability that the assignment vector
1320: $\vec x _i = (x_i^1, x_i^2, \ldots , x_i^{q'})$
1321: is equal to $\vec x' = (0,1,1,0,1,1,\ldots,1)$? The answer is
1322: \begin{equation}\label{defpxor}
1323: p(\vec x') = \left\langle \frac 1{({\cal N}_h)^{q'}}
1324: \sum _{X^1,X^2, \ldots, X^{q'}}\;\delta _{\vec x_i - \vec x} \;
1325: \prod _{l=1}^M \prod _{a=1}^q e_\ell ( X^a) \right\rangle
1326: \end{equation}
1327: where the dependence on $i$ is wiped out by the average over the instance.
1328: The above probability is an interesting quantity; it provides us information
1329: about the `microscopic' nature of solutions. Setting $q'=1$ gives
1330: us the probabilities $p(0),p(1)$ that a variable is false or true
1331: respectively, that is, takes the same value as in the null assignment or not.
1332: For generic $q'$ we may think of two extreme situations:
1333: \begin{itemize}
1334: \item a flat $p$ over assignment vectors, $p(\vec x ')=1/2^{q'}$, corresponds
1335: to essentially orthogonal solutions;
1336: \item on the opposite, a concentrated probability e.g. $p(\vec x ')=
1337: \delta_{\vec x'}$ implies that variables
1338: are extremely constrained, and that the (almost) unique solution
1339: is the null assignment.
1340: \end{itemize}
1341:
1342: The careful reader will have already guessed that our calculation
1343: of the $q^{th}$ moment gives access to a weighted counterpart of $p$.
1344: The order parameter
1345: \begin{equation} \label{defrhoetoilexor}
1346: \rho ^*(\vec x) = \frac 1{\langle ({\cal N}_h)^q\rangle}
1347: \sum _{X^1,X^2, \ldots, X^q}\; \delta _{\vec x_i - \vec x} \;
1348: \left\langle \prod _{l=1}^M \prod_{a=1}^q e_\ell ( X^a) \right\rangle \quad ,
1349: \end{equation}
1350: is not equal to $p$
1351: even when $q=q'$. However, at the price
1352: of mathematical rigor, the exact probability $p$ over
1353: vector assignments of integer length $q'$ can be reconstructed from
1354: the optimal histogram $\rho ^*$ associated to moments of
1355: order $q$ when $q$ is real-valued and sent to $0$.
1356: The underlying idea is the following. Consider
1357: (\ref{defrhoetoilexor}) and an integer $q'<q$. From any assignment
1358: vector $\vec x$ of length $q$, we define two assignment vectors
1359: $\vec x', \vec x''$ of respective lengths $q', q-q'$ corresponding to
1360: the first $q'$ and the last $q-q'$ components of $\vec x$ respectively.
1361: Summing (\ref{defrhoetoilexor}) over the $2^{q-q'}$ assignment
1362: vectors $\vec x''$ gives,
1363: \begin{equation} \label{defrhoetoilexor2}
1364: \sum _{\vec x''} \rho ^*(\vec x', \vec x'')
1365: = \frac 1{\langle ({\cal N}_h)^q \rangle}
1366: \sum _{\{X^a\}}\; \delta _{\vec x'_i - \vec x'} \;
1367: \left\langle \big({\cal N}_h\big) ^{q-q'}
1368: \prod _{l,a} e_\ell ( X^a) \right\rangle \ .
1369: \end{equation}
1370: As $q$ now appears in the powers of ${\cal N}_h$ in the numerator and
1371: denominator only, it can be formally send to zero at fixed $q'$,
1372: yielding
1373: \begin{equation} \label{replicaess}
1374: \lim _{q\to 0}\ \sum _{\vec x''} \rho ^*(\vec x', \vec x'') = p(\vec x ')
1375: \end{equation}
1376: from (\ref{defpxor}).
1377: This identity justifies the denomination order parameter given to
1378: $\rho^*$.
1379:
1380: Having understood the significance of $\rho^*$ helps us to find
1381: appropriate solutions to (\ref{spxor}). Intuitively and from
1382: the discussion of the first moment case $q=1$,
1383: $p$ is expected to reflect both the special role of the null assignment
1384: (which is a solution to all homogeneous systems) and the ability of
1385: other solutions of a random system to be essentially orthogonal to
1386: this special assignment. A possible guess is thus
1387: \begin{equation} \label{rhoetoilexor2}
1388: p(\vec x ') = \frac {1-b}{2^{q'}}+ b \; \delta
1389: _{\vec x'} \quad ,
1390: \end{equation}
1391: where $b$ expresses some degree of `correlation' of solutions with
1392: the null one. Hypothesis (\ref{rhoetoilexor2}) interpolates between the
1393: fully concentrated ($b=1$) and flat ($b=0$) probabilities.
1394: $b$ measures the fraction of variables (among the $N$ ones)
1395: that take the 0 values in all $q'$ solution, and coincides with the
1396: notion of backbone introduced in Section \ref{seccluster}. Hypothesis
1397: (\ref{rhoetoilexor2}) is equivalent, from the connection (\ref{replicaess})
1398: between $p$ and the annealed histogram $\rho ^*$ to the following
1399: guess for the solution of the maximisation condition (\ref{spxor}),
1400: \begin{equation} \label{rhoetoilexor}
1401: \rho ^*(\vec x ) = \frac {1-b}{2^q}+ b \; \delta
1402: _{\vec x} \ .
1403: \end{equation}
1404: Insertion of Ansatz (\ref{rhoetoilexor}) in (\ref{spxor}) shows
1405: that it is indeed a solution provided $b$ is shrewdly chosen
1406: as a function of $q$ and $\alpha$, $b=b^*(q,\alpha)$.
1407: Its value can be either found from direct resolution of (\ref{spxor}),
1408: or from insertion of histogram (\ref{rhoetoilexor})
1409: in ${\cal G}_h$ (\ref{qxor6}) and maximisation over $b$, with the
1410: result,
1411: \begin{equation} \label{xormaxgh}
1412: g_h(q,\alpha) = \max _{0\le b\le 1} A_h (b,q,\alpha)
1413: \end{equation}
1414: where
1415: \begin{eqnarray} \label{xormaxgh2}
1416: A _h (b,q,\alpha) &=& -\left( 1 -\frac 1{2^q} \right)
1417: \; (1-b) \; \ln \left(\frac{1-b}{2^q}\right) \\
1418: &-& \left( b + \frac{1-b}{2^q} \right) \; \ln
1419: \left( b + \frac{1-b}{2^q} \right) + \alpha\;
1420: \ln \left( b^3 + \frac{1-b^3}{2^q} \right) \ ,
1421: \nonumber
1422: \end{eqnarray}
1423: where the maximum is precisely reached in $b^*$. Notice that, since
1424: $\rho^*$ in (\ref{rhoetoilexor}) is entirely known from the value
1425: of $b^*$, we shall indifferently call order parameter $\rho^*$, or
1426: $b^*$ itself.
1427:
1428: \subsection{Results}\label{secresu}
1429:
1430: Numerical investigation of $A _h$ (\ref{xormaxgh2}) shows
1431: that: for $\alpha<\alpha_M(q)$ the only local maximum of $A _h$
1432: is located in $b^*=0$, and $A_h(q,\alpha)=q(1-\alpha)\ln 2$;
1433: when $\alpha_M(q) < \alpha<\alpha^*(q)$, there exists another local maximum
1434: in $b>0$ but the global maximum is still reached in $b^*=0$;
1435: when $\alpha>\alpha^*(q)$, the global maximum is located in $b^*>0$.
1436: This scenario extends to generic $q$ the findings of the second moment
1437: calculation carried out in Section
1438: \ref{secld2}. The $\alpha_M$ and $\alpha^*$ lines divide the
1439: $q,\alpha$ plane as
1440: shown in Figure~\ref{qetoilexor}. Notice that, while the black dots in
1441: Figure~\ref{qetoilexor} correspond to integer-valued $q$, the continuous
1442: lines are the output of the implicit analytic continuation to real $q$
1443: done by the replica calculation.
1444:
1445: \begin{figure}
1446: \begin{center}
1447: \includegraphics[width=110pt,angle=-90]{fig341.eps}
1448: \end{center}
1449: \caption{The $q,\alpha$
1450: plane and the critical lines $\alpha_M(q)$ (dashed),
1451: $\alpha^*(q)$ (full tick), and $\alpha_s(q)$ (full thin)
1452: appearing in the calculation
1453: of the $q^{th}$ moment for homogeneous 3-XORSAT systems.
1454: Full dots correspond to integer $q$ values, while
1455: continuous curves result from the analytic continuation to
1456: real $q$. The fraction of variables in the backbone, $b^*$, vanishes below the
1457: line $\alpha_M(q)$; the global maximum of $A_h$ in
1458: (\ref{xormaxgh2}) is located in $b^*>0$ for ratios $\alpha>\alpha^*(q)$.
1459: Ansatz (\ref{rhoetoilexor}) is locally unstable in the hardly visible
1460: domain $q<0,\alpha_M(q) <\alpha <\alpha _s(q)$.}
1461: \label{qetoilexor}
1462: \end{figure}
1463:
1464: Taking the derivative of (\ref{xormaxgh}) with respect to $q$ and
1465: sending $q\to 0$ we obtain the typical entropy of a homogeneous
1466: 3-XORSAT system at ratio $\alpha$,
1467: \begin{equation}\label{commentxor1}
1468: s_h ^*(\alpha)= \ln 2 \times \max _{0\le b\le 1}
1469: \big[ (1-b) \big(1 - \ln(1-b)\big) - \alpha \, (1-b^3) \big] \ .
1470: \end{equation}
1471: The optimal value for $b$ coincides with the solution of (\ref{eqb}).
1472: The typical entropy is plotted in Figure~\ref{xorcurve}, and is equal to:
1473: \begin{itemize}
1474: \item $(1-\alpha) \ln 2$ when $\alpha<\alpha_c\simeq 0.918$
1475: (Figure~\ref{qetoilexor}); in this range of ratios, homogeneous and
1476: full (with random second members)
1477: systems have essentially the same properties, with the same cluster
1478: organisation of solutions, and identical entropies of solutions.
1479: \item a positive but rapidly decreasing function given by
1480: (\ref{commentxor1}) when $\alpha>\alpha_c$; above the
1481: critical ratio, a full system has no solution any more, while a homogeneous
1482: instance still enjoys a positive entropy. The expression for
1483: $s^*_h(\alpha)$ coincides with the continuation to $\alpha>\alpha_c$ of
1484: the entropy $s_{in}(\alpha)$ (\ref{sin}) of solutions in
1485: a single cluster for a full system.
1486: In other words, a single cluster of solutions, the one with the
1487: null solution, survive for ratios $\alpha>\alpha_S$ in homogeneous systems.
1488: \end{itemize}
1489:
1490: Atypical instances can be studied and the large deviation rate
1491: function for the entropy can be derived from (\ref{xormaxgh}) for
1492: homogeneous systems, and using equivalence (\ref{momentxorequality}),
1493: for full systems. Minimizing over the entropy we obtain the rate
1494: function $\omega_3(\alpha)$ associated to the probability that a
1495: random 3-XORSAT system is satisfiable,
1496: with the result shown in Figure~\ref{xorcurve}. As expected we find
1497: $\omega_3 =0$ for $\alpha < \alpha_c$ and $\omega_3 >0 $ for
1498: $\alpha>\alpha_c$, allowing us to locate the Sat/Unsat threshold.
1499:
1500: Notice that the emergence of clustering can be guessed from Figure
1501: \ref{qetoilexor}. It coincides with the appearance of a local
1502: maximum of $A_h$ (\ref{xormaxgh2}) with a non vanishing
1503: backbone $b$. While in the intermediate phase $\alpha_d<\alpha<
1504: \alpha_c$, the height of the
1505: global maximum equals the total entropy $s^*$, the height of the local
1506: maximum coincides with the entropy of clusters $s _{cluster}$
1507: (\ref{scluster}).
1508:
1509: \subsection{Stability of the replica Ansatz}
1510:
1511: The above results rely on Ansatz (\ref{rhoetoilexor}). A necessary criterion
1512: for its validity is that $\rho^*$ locates a true local maximum
1513: of ${\cal G}_h$ , and not merely a saddle-point. Hence we have to calculate
1514: the Hessian matrix of ${\cal G}_h$ in $\rho^*$, and check that the eigenvalues
1515: are all negative \cite{De79}.
1516: Differentiating (\ref{qxor6}) with respect to $\rho(\vec x)$
1517: and $\rho (\vec x')$ we obtain the Hessian matrix
1518: \begin{equation} \label{hvalue}
1519: H(\vec x,\vec x') = -\frac{\delta _{\vec x+\vec x'}}{\rho^* (\vec x)}
1520: +6\alpha\; \frac{\rho^* (\vec x+\vec x')}D - 9\alpha \;\frac{N(\vec x)}D
1521: \, \frac{N(\vec x')}D\ ,
1522: \end{equation}
1523: where $D=\frac {1-b^3}{2^q} + b^3$, $N(\vec x) = \frac {1-b^2}{2^q}
1524: + b^2\,\delta _{\vec x}$. We use $b$ instead of $b^*$ to
1525: ligthen the notations, but it is intended that $b$ is the backbone
1526: value which maximizes $A_h$ (\ref{xormaxgh2}) at fixed $q,\alpha$.
1527: To take into account the global constraint over the histogram
1528: (\ref{qxor2c}) one can express one fraction, say, $\rho (\vec 0)$, as
1529: a function of the other fractions $\rho(\vec x)$, $\vec x\ne \vec 0$.
1530: ${\cal G}_H$ is now a fonction of $2^q-1$ independent variables,
1531: with a Hessian matrix $\tilde H$ simply related to $H$,
1532: \begin{equation} \label{htilde}
1533: \tilde H(\vec x,\vec x') = H(\vec x,\vec x') - H(\vec x,\vec 0) -
1534: H(\vec 0,\vec x') + H(\vec 0,\vec 0) \ .
1535: \end{equation}
1536: Plugging expression (\ref{hvalue}) into (\ref{htilde}) we obtain
1537: \begin{eqnarray}
1538: \tilde H(\vec x,\vec x') &=& \lambda _R\; \delta_{\vec x+\vec x'}
1539: + \frac 1{2^q-1}\big( \lambda _L - \lambda _R) \qquad \mbox{where}
1540: \nonumber \\
1541: \lambda_R &=& 6\alpha \; \frac bD - \frac{2^q}{1-b} \\
1542: \lambda _L &=& 2^q \left( 6\alpha \; \frac bD - \frac{2^q}
1543: {(1-b)(1-b+2^qb)} - 9\alpha (1-2^{-q})\frac{b^4}{D^2}\right) \ .
1544: \nonumber
1545: \end{eqnarray}
1546: Diagonalization of $\tilde H$ is immediate, and we find two
1547: eigenvalues:
1548: \begin{itemize}
1549: \item $\lambda _L$ (non degenerate). The eigenmode corresponds
1550: to a uniform infinitesimal
1551: variation of $\rho (\vec x$) for all $\vec x\ne \vec 0$, that is, a
1552: change of $b$ in (\ref{rhoetoilexor}). It is an easy check that
1553: \begin{equation}
1554: \lambda _L = \frac{2^q}{1-2^{-q}}\; \frac{\partial ^2 A_h}{\partial
1555: b^2} (b,q,\alpha) \ ,
1556: \end{equation}
1557: where $A_h$ is defined in (\ref{xormaxgh2}).
1558: As we have chosen $b$ to maximize $A_h$ this mode, called longitudinal
1559: in replica literature \cite{De79}, is stable\footnote{Actually $b^*$
1560: is chosen to {\em minimize} $A_h$ when $q<0$, thus $\lambda_L$ has
1561: always the right negative sign.}.
1562:
1563: \item $\lambda _R$ ($2^q-2$-fold degenerate): the eigenmodes correspond
1564: to fluctuations of the order parameter $\rho$ transverse to
1565: the replica subspace described by (\ref{rhoetoilexor}), and are
1566: called replicon in spin-glass theory\cite{De79}. Inspection of
1567: $\lambda_R$ as a function
1568: of $\alpha,q$ shows that it is always negative when $q>0$. For $q<0$
1569: the replicon mode is stable if
1570: \begin{equation}
1571: \alpha > \alpha_s (q) = \frac{1- b^3 + 2^q\, b^3}{6 \, b(1-b)} \ .
1572: \end{equation}
1573: which is a function of $q$ only once we have chosen $b=b^*(q,\alpha_s)$.
1574: \end{itemize}
1575: The unstable region $q<0, \alpha_M (q) < \alpha <\alpha _s(q)$ is
1576: shown in Figure \ref{qetoilexor} and is hardly visible when $q>-3$.
1577: In this region a continuous symmetry breaking is expected \cite{Me87}.
1578: In particular $\alpha_s$ stay below the $\alpha^*$ line for
1579: small (in absolute value) and negative $q$. We conclude
1580: that our Ansatz (\ref{rhoetoilexor}) defines a maximum
1581: of ${\cal G}_h$.
1582:
1583: Is it the global maximum of ${\cal G}_h$? There is no simple way to
1584: answer this question. Local stability does not rule out the possibility
1585: for a discontinuous transition to another maximum in the replica order
1586: parameter space not described by (\ref{rhoetoilexor}).
1587: A final remark is that a similar calculation can be done for any value
1588: of $K$. The outcome for $K=2$ is the rate function $\omega_2$ plotted
1589: in Figure~\ref{proba2b-fig}, in good agreement with numerics
1590: close to the threshold.
1591:
1592: \begin{figure}
1593: \begin{center}
1594: \includegraphics[width=120pt,angle=0]{fig342.eps}
1595: \end{center}
1596: \caption{Rate function $\omega _3$ for the probability of satisfaction
1597: of full (bottom curve) and entropy $s_h^*$ of solutions
1598: for homogeneous (top curve) 3-XORSAT systems vs. $\alpha$.
1599: The vertical dotted lines indicate the critical
1600: Sat/Unsat threshold, $\alpha_c\simeq 0.918$. For $\alpha <\alpha_c$
1601: $\omega _3=0$, and $s_h^*=(1-\alpha)\,\ln 2$ is the same as for full systems.
1602: Above the threshold $\omega ^* <0$. Homogeneous
1603: systems are, of course, always satisfiable: the entropy $s_h^*$ is a positive
1604: but quickly decreasing function of $\alpha$.}
1605: \label{xorcurve}
1606: \end{figure}
1607:
1608: % cavity
1609:
1610: \section{Advanced methods (II): cavity}\label{seccavity}
1611:
1612: The cavity method, in the context of disordered systems, was
1613: historically developed as an alternative to the the replica method
1614: \cite{Me87}. Its application to spin systems on random graphs is
1615: extensively explained in \cite{Me01y}, and we limit ourselves here
1616: to briefly show how it gives back the 3-XORSAT scenario of Section
1617: \ref{seccluster} \cite{Me03}.
1618:
1619: Let us consider a system $F$ involving variables $x_i$, $i=1,\ldots,
1620: N$. In the following we will indifferently
1621: use the variable $x_i=0,1$ or its spin representation
1622: $S_i=(-1)^{x_i}=\pm 1$ when convenient.
1623: Let us define the GS energy $E^F _i (S_i)$
1624: of the system when the $i^{th}$ spin is kept fixed, that is, the
1625: minimal number of violated equations in $F$, taken over the $2^{N-1}$
1626: configurations. We may always write
1627: \begin{equation} \label{defec}
1628: E^F_i (S_i) = -w_i-h_i\, S_i\ ,
1629: \end{equation}
1630: where $h_i$ is called `field' acting on spin $S_i$.
1631: For a homogeneous system $E^F_i(+1)=0$, and $E^F_i(-1)=n_i$ for some integer
1632: $n_i$. Hence $h_i=\frac {n_i}2$ takes half-integer values.
1633:
1634: The above definition can be extended to the case of
1635: $\ell > 1$ fixed spins. Let $I\subset \{1,2,\ldots, N\}$ be a subset
1636: of the indices of cardinal $|I|\ge 2$, and $S_I$ denote one of the $2^{|I|}$
1637: configurations of the spins $S_i, i\in I$. The GS energy of $F$ for given
1638: $S_I$ can in general be written as
1639: \begin{equation} \label{defec2}
1640: E^F_{I} (S_I) = -w_{I} - \sum _{i \in I} h_{i} \, S_{i}
1641: -\sum _{I'\subset I : |I'|\ge 2} J_{I'}\,\prod
1642: _{i\in I' } S_i
1643: \end{equation}
1644: where the $h_i$s are the fields and the $J_{I'}$s are
1645: effective couplings between subsets of spins.
1646:
1647: The basic cavity assumption is that effective couplings are
1648: vanishingly small: $J_{I'}=0$ for every subset $I'$. This apparently bold
1649: hypothesis critically relies on a general property of random graphs
1650: (from which our system is built on). Define the distance between two
1651: vertices as the minimal number of edges on pathes linking these two
1652: points. Then vertices in a finite subset are, with high probability
1653: when $N\to\infty$, typically at infinite distance from each
1654: other\footnote{An alternative formulation is, for finite size $N$,
1655: that the shortest loops (in extensive number) have lengths of the order of
1656: $\log N$ \cite{Bo89}.}. When correlations
1657: between variables in GS extinguish with the distance {\em i.e.} when
1658: the correlation length is finite the cavity assumption is
1659: correct in the large $N$ limit \cite{Me87,Mo05}. The assumption will break
1660: down when correlations subsist on infinite distance, which
1661: happens to be the case in the clustered phase.
1662:
1663: \subsection{Self-consistent equation for the fields} \label{seccav1}
1664:
1665: Under the assumption that couplings between randomly picked up
1666: spins are null we are left with the fields only. The goal of this
1667: section is to show how to calculate those fields, or more precisely,
1668: their probability distribution. The derivation is based on the
1669: addition procedure
1670: already used in the calculation of the size of the giant component in
1671: random graphs (Section \ref{secrg}).
1672:
1673: Consider a system $F$ over $N$ variables to which we want
1674: to add one
1675: equation involving one new variable $S$, and two variables $S_1,S_2$
1676: appearing in $F$. The energy function
1677: associated to this equation is
1678: \begin{equation}
1679: e(S,S_1,S_2) = \frac 12 \big( 1 - \sigma\, S\, S_1\, S_2 \big)
1680: \end{equation}
1681: where $\sigma=+1$, respectively $-1$, when the second member of the
1682: equation is 0, resp. 1. Let us calculate the GS
1683: energy of the new system $F'=F$ + added equation when the new
1684: variable $S$ is kept fixed,
1685: \begin{equation}
1686: E^{F'} (S) = \min _{S_1,S_2}\big[ e(S,S_1,S_2) + E^F_{1,2}
1687: (S_1,S_2) \big] = - w - u \, S \ .
1688: \end{equation}
1689: With the cavity hypothesis the couplings between spins $S_1,S_2$ is
1690: null and the minimization is straightforward. We deduce the
1691: following explicit
1692: expression for the field acting on $S$ (called bias in the cavity
1693: literature \cite{Me01y}),
1694: \begin{equation}\label{expressu}
1695: u=\frac \sigma2\; \mbox{sign} \big(h_1\; h_2\big) \ .
1696: \end{equation}
1697: Suppose we now add $\ell \ge 1$ (and not only one) equations. The above
1698: calculation can be easily repeated. The absence of couplings make the
1699: total field acting on $S$ a linear combination of the fields coming
1700: from each new equation,
1701: \begin{equation}\label{uj}
1702: h = \sum _{j=1}^\ell u ^{j} \ ,
1703: \end{equation}
1704: where $u^{j}$ is calculated from (\ref{expressu}) and each pair of fields
1705: $(h_1^{j}, h_2^{j})$ acting on the spins in the $j^{th}$ equation,
1706: $j=1,\ldots, \ell$.
1707:
1708: How many equations should we add for our new system over $N+1$
1709: variables to have the same statistical features as old one over $N$
1710: variables? First $\ell$ should be Poisson distributed
1711: with parameter $3\alpha$. Then, given $\ell$, we randomly chose $\ell$
1712: pairs of variables; for each pair the corresponding bias
1713: $u$ can be calculated from (\ref{expressu}). Assume the output is a
1714: set of $\ell$ independent biases, taking values
1715: \begin{equation}\label{uk}
1716: u = \left\{ \begin{array}{c c c}
1717: + \frac 12 & \mbox{with probability} & a_+ \\
1718: 0 & \mbox{with probability} & a_0 \\
1719: - \frac 12 & \mbox{with probability} & a_-
1720: \end{array} \right. \
1721: \end{equation}
1722: Obviously $a_++a_0+a_-=1$.
1723: Summing over the equations as in (\ref{uj}) we obtain the distribution of the
1724: field $h$ acting on the new spin at fixed $\ell$,
1725: \begin{equation}
1726: p(h | \ell )=\sum _{\ell_+,\ell_0,\ell_-} {\ell \choose
1727: \ell+,\ell_0,\ell_-}
1728: a_+^{\ell_+}\, a_0^{\ell_0}\, a_-^{\ell_-}\, \delta
1729: _{h-\frac 12(\ell_+-\ell_-)} \ .
1730: \end{equation}
1731: Finally we sum over the Poisson distribution for $\ell$ to obtain the
1732: distribution of fields $h$,
1733: \begin{equation} \label{defph}
1734: p(h)=e^{-3\alpha (1-a_0)} \sum _{\ell_+,\ell_-} \frac{
1735: (3\alpha)^{\ell_++\ell _-}}{\ell_+!\ell_-!} a_+^{\ell_+}\,
1736: a_-^{\ell_-}\, \delta
1737: _{h-\frac 12(\ell_+-\ell_-)} \ .
1738: \end{equation}
1739: In turn we calculate the distribution of the biases from the one of
1740: the fields through (\ref{expressu}). The outcome are the values of the
1741: probabilities (\ref{uk}) in terms of $p$,
1742: \begin{eqnarray}
1743: a_+ &=& \sum _{h_1,h_2: h_1 h_2 > 0} p(h_1)\, p(h_2) \ ,
1744: \quad
1745: a_- = \sum _{h_1,h_2: h_1 h_2 < 0} p(h_1)\, p(h_2) \ ,\nonumber \\
1746: a_0 &=& \sum _{h_1,h_2: h_1 h_2 = 0} p(h_1)\, p(h_2) = 2\, p(0) -p(0)^2 \ .
1747: \label{conda0}
1748: \end{eqnarray}
1749: The above equations together with (\ref{defph}) define three
1750: self-consistent conditions for $a_0,a_+,a_-$. Notice that the free
1751: energy can be calculated along the same lines \cite{Me01y}.
1752:
1753: \subsection{Application to homogeneous and full systems}\label{seccavapp}
1754:
1755: In the case of homogeneous systems ($\sigma = +1$)
1756: we expect all the fields to be
1757: positive, and look for a solution of (\ref{conda0}) with $a_-=0$. Then
1758: $p(h)$ (\ref{defph})
1759: is a Poisson distribution for the integer-valued variable $2h$,
1760: with parameter $3\,\alpha\, a_+$.
1761: The self-consistent equation (\ref{conda0}) reads
1762: \begin{equation}
1763: a_0=1-a_+= 2 \; e^{-3 \alpha a_+} - e^{-6\, \alpha a_+} = 1 -
1764: \big( 1 - e^{-3\, \alpha a_+} \big)^2
1765: \end{equation}
1766: which coincides with (\ref{eqb}) with the definition $b=\sqrt{a_+}$.
1767: As expected
1768: \begin{equation}
1769: b= \sum _{h\ge \frac 12} p(h)
1770: \end{equation}
1771: is the fraction of frozen variables (which cannot be flipped from 0 to
1772: 1 in GS assignments), in agreement with the notion of backbone of Section
1773: \ref{seccluster}.
1774:
1775: The energy is zero at all ratio $\alpha$ by construction. As for the
1776: entropy consider adding a new equation to the system $F$ (but with no
1777: new variable). With probability
1778: $1-b^3$ at least one of the three variables in the new equation $e$ was
1779: not frozen prior to addition, and the number of solutions of the new
1780: system $F+e$ is half the one of $F$. With probability $b^3$ all three
1781: variables are frozen in $F$ (to the zero value) and the number of
1782: solutions of $F+e$ is the same as the one of $F$. Hence the average
1783: decrease in entropy is
1784: \begin{equation} \label{evolsh}
1785: N\, s_h^* (\alpha + \frac 1N) - N \, s_h^*(\alpha) \simeq
1786: \frac{ds ^*_h}{d\alpha} = -(1-b^3) \; \ln 2 \ .
1787: \end{equation}
1788: The same differential equation can be obtained by differentiating
1789: (\ref{commentxor1}). With the limit condition
1790: $s_h^*(\alpha\to\infty)=0$ we obtain back the correct expression for
1791: the average entropy of homogeneous systems. The entropy is equal to
1792: $(1-\alpha)\ln 2$ at $\alpha=\alpha_c$, and becomes smaller when the
1793: ratio decreases. This shows that the solution $b=0$ must be
1794: preferred in this regime to the metastable $b>0$ solution.
1795: We conclude that the cavity assumption leads to sensible results
1796: for homogeneous systems at all ratios
1797: $\alpha$.
1798:
1799:
1800:
1801: In full systems the sign $\sigma$ entering (\ref{expressu}) takes $\pm 1$
1802: values with equal probabilities. We thus expect $p(h)$ to be an even
1803: distribution, and $a_+=a_-=\frac 12 (1-a_0)$.
1804: Remark that a solution with $a_0<1$ cannot exist in the satifiable
1805: phase. It would allow two added equations to impose opposite non zero biases
1806: to the new variable {\em i.e.} to constraint this variable to take
1807: opposite values at the same time.
1808: Given $a_0$ we calculate from (\ref{defph}) the probability that the
1809: field vanishes,
1810: \begin{equation}\label{e1a}
1811: p(0)= e^{-3\,\alpha (1-a_0)} \sum _{\ell =0} ^\infty \left[
1812: \frac {3\alpha}2 (1-a_0)\right]^{2\ell} \frac 1{\ell !^2}
1813: \end{equation}
1814: and, in turn, derive from (\ref{conda0}) a self-consistent equation
1815: for $a_0$.
1816: Numerical investigations
1817: show that $a_0=1$ is the unique solution for $\alpha < \alpha_T=
1818: 1.167$. When $\alpha > \alpha_T$ there appears another solution with
1819: $a_0<1$.
1820: The clustering and Sat/Unsat transitions are totally absent.
1821: This result, incompatible with the exact picture of random 3-XORSAT
1822: exposed in Section \ref{seccluster}, shows that the simple cavity
1823: hypothesis does not hold for full systems.
1824:
1825: \subsection{Spontaneous symmetry breaking between clusters}
1826: \label{secpin}
1827:
1828: In the clustered phase variables are known to be strongly correlated
1829: and the cavity assumption has to be modified. Actually from what we
1830: have done above in the homogeneous case we guess that the independence
1831: condition still holds if we can in some way restrict the whole space
1832: of solutions to one cluster. To do so we explicitely
1833: break the symmetry between clusters as follows\cite{Pa86,Mon95}.
1834:
1835: Let $S_i^*,i=1,\ldots, N$ be a reference solution of a full
1836: satisfiable system $F$, and $F_h$ the corresponding homogeneous
1837: system. We define the local gauge transform $S_i \to \hat S_i= S_i \times
1838: S_i^*$. $\{S\}$ is a solution of $F$ if and only if $\{\hat S\}$ is a
1839: solution of $F_h$. As the cavity assumption is correct for the
1840: homogeneous system we obtain the distribution of fields $\hat
1841: h_i \ge 0$ from (\ref{defph}). Gauging back to the original spin
1842: configuration gives us the fields
1843: \begin{equation}\label{gauge}
1844: h_i = S_i^*\times \hat h_i \ .
1845: \end{equation}
1846: It turns out that the above fields depend only on the cluster to which
1847: belong the reference solution. Indeed
1848: for the fraction $1-b$ of the non frozen spins, $\hat h_i =
1849: h_i=0$. For the remaining fraction $b$ of spins in the backbone
1850: $\hat h_i > \frac 12$ and $S_i^*$ has a unique value for all
1851: solutions in the cluster (Section \ref{seccluster}). Hence
1852: the fields $h_i$ are a function of cluster $(c)$ containing $\{S^*\}$,
1853: and will be denoted by $h_i ^{(c)}$.
1854:
1855: What modification has to be brought to the cavity assumption of
1856: Section \ref{seccav1} is now clear. Given a subset $I$ of the spins
1857: with configuration $S_I$ we define $E_F^{(c)}(S_I)$ as the GS
1858: energy over configurations in the cluster $(c)$. Then the cavity
1859: assumption is correct (spins in $I$ are uncorrelated) and $E_F^{(c)}$
1860: define the fields $h_i^{(c)}$. How do we perform this restriction in
1861: practice? A natural procedure is to break the symmetry between
1862: clusters in an explicit manner by adding a small coupling to the
1863: reference solution \cite{Pa86,Mo95}.
1864: Remark that symmetry was broken (naturally but explicitly!) in the case of
1865: homogeneous systems when we looked for a distribution $p(h)$ with
1866: support on positive fields only. It is a remarkable feature of XORSAT
1867: (rather unique among disordered systems) that symmetry between
1868: disordered clusters can be broken in a constructive and simple way.
1869:
1870: The main outcome of the above discussion is that the field attached to
1871: variable $i$ is not unique, but depends on the cluster $(c)$. We define
1872: the distribution $p_i(h)$ of the fields attached to variable $i$ over
1873: the clusters (with uniform weights since all clusters contain the same
1874: number of solutions) \cite{Mon98}. The naive cavity assumption
1875: corresponds to
1876: \begin{equation}
1877: p_i(h) = \delta _{h-h_i} \ .
1878: \end{equation}
1879: In presence of many
1880: clusters $p_i(h)$ is not highly concentred. From (\ref{gauge}) and the fact
1881: that $S_i^*=\pm 1$ depending on the cluster from which we pick up the
1882: reference solution we find that
1883: \begin{equation}\label{ef}
1884: p_i(h) = \frac 12 \left[ \delta _{h-\hat h_i} + \delta _{h+\hat h_i}\right] \ .
1885: \end{equation}
1886: As $\hat h_i$ is itself randomly distributed we are led to introduce
1887: the distribution ${\cal P}$
1888: of the field distributions $p_i(h)$. This mathematical object, ${\cal
1889: P}(p(h))$, is the order parameter of the cavity theory in the
1890: clustered phase \cite{Mon98,Me01y}.
1891:
1892: \subsection{Distribution of field distributions} \label{seccavrsb}
1893:
1894: Let us see how ${\cal P}$ can be obtained within the one-more variable
1895: approach of Section \ref{seccav1}. A new equation
1896: contains two variables $S_i,S_j$ from $F$, with fields
1897: $h_i^{(c)},h_j^{(c)}$ in each cluster $(c)$. The bias $u$
1898: is a deterministic
1899: function of those two fields for each cluster (\ref{expressu}). We
1900: define its distribution over clusters $\rho$. As $u$ can take three
1901: values only and $\rho$ is an even distribution due to the randomness of
1902: the second member of the new equation we may write
1903: \begin{equation} \label{db}
1904: \rho (u ) = (1-\varphi_{ij}) \; \delta_u + \frac{\varphi_{ij} }2\;
1905: \big( \delta _{u-\frac 12} + \delta _{u+\frac 12} \big) \ .
1906: \end{equation}
1907: The weight $\varphi$ is a random variable which varies from pair
1908: $(ij)$ to pair.
1909:
1910: What is the probability distribution {\sc p}$(\varphi)$ of $\varphi$?
1911: Either the two variables in the pair belong to the backbone and they
1912: are frozen in all clusters; then $u$ will be non zero and
1913: $\varphi=1$. Or one (at least) of the two variables is not frozen and $u=0$
1914: in all clusters, giving $\varphi=0$. We may write
1915: \begin{equation}
1916: \mbox{\sc p}(\varphi ) = (1-w)\; \delta _{\varphi} + w\;
1917: \delta_{\varphi-1} \ .
1918: \end{equation}
1919: From the above argument we expect $w=b^2$. Let us derive this result.
1920:
1921: Assume we add $\ell\ge 1$ equations to our system. For each one of
1922: those equations a bias $u^{j}$ is drawn randomly according to distribution
1923: (\ref{db}). Denote by $m(\le
1924: \ell)$ the number of those equations with parameter $\varphi=1$; $m$
1925: is binomially distributed with probability $w$ among $\ell$.
1926: Then $\ell-m$ biases are null, and $m$ biases are not equal to zero.
1927: For the formula to remain satisfiable the non-zero biases must be all
1928: positive or negative \cite{Me01y}, see Section
1929: \ref{seccavapp}. Hence the distribution of the field on the new variable
1930: is
1931: \begin{equation}\label{dispm}
1932: p^m(h) = \frac 12 \left[ \delta_{h-\frac m2} + \delta _{h+\frac m2}\right] \ ,
1933: \end{equation}
1934: in agreement with the expected form (\ref{ef}).
1935: The upperscript $m$ underlines that field distributions with non zero
1936: probability are can be labelled by an integer $m$; they define
1937: a countable set and the distribution ${\cal P}$ can be defined
1938: as a discrete probability ${\cal P}_m$ over the set of positive
1939: integers $m$.
1940: The probability ${\cal P}_m$ of distribution (\ref{dispm})
1941: is the convolution of
1942: binomial distribution for $m$ at fixed $\ell$ with the Poisson
1943: distribution over $\ell$,
1944: \begin{equation}\label{dispmp}
1945: {\cal P} ^m = e^{-3\,\alpha\, w} \frac{(3\,\alpha\, w)^m
1946: }{m!} \ .
1947: \end{equation}
1948: Identities (\ref{dispm},\ref{dispmp}) fully determine the distribution of field
1949: distributions in term of a single parameter, $w$.
1950:
1951: To close the self-consistency argument consider the two variables in
1952: $F$ in, say, the first added equation. Call $h,h'$ their fields,
1953: distributed according to $p_m(h), p_{m'}(h')$ for
1954: some $m,m'$. The bias created onto the new variable will be non
1955: zero if $h$ and $h'$ may both take non zeros value in some clusters,
1956: that is, if $m$ and $m'$ are not equal to zero.
1957: This translates into the mathematical identity
1958: \begin{equation}
1959: w = \sum _{m\ge \frac 12,m'\ge \frac 12}
1960: {\cal P}_m\; {\cal P}_{m'} = \big(1-{\cal
1961: P}_0 \big)^2 = \big(1 - e^{-3\,\alpha\, w}\big)^2
1962: \end{equation}
1963: from (\ref{dispmp}).
1964: The above equation coincides with (\ref{eqb}) for $w=b^2$. Notice that
1965: $w$ is equal to the probability $a_+$
1966: that the bias is non zero in the homogeneous
1967: case (\ref{uk}), in agreement with the discussion of Section \ref{secpin}.
1968:
1969: It is easy to find back the expressions for the entropies of clusters,
1970: $s_{cluster}$, and solutions in a cluster, $s_{in}$, given in Section
1971: \ref{seccluster}. As for the latter entropy the argument leading to
1972: (\ref{evolsh}) can be repeated, with the modification that the second
1973: member of the added equation is not necessarily zero but the value it
1974: should have for the equation to be satisfied when all three variables
1975: are frozen. Hence (\ref{evolsh}) holds with $s_{h}^*$ replaced with
1976: $s_{in}$. As for the entropy of clusters the same argument again tells us
1977: that, on average, half of the clusters will disappear when the
1978: three variables are frozen and the second member of the equation is
1979: randomly chosen. Therefore
1980: \begin{equation}\label{sc2}
1981: \frac{d s_{cluster}}{d\alpha} = -b^3\, \ln 2 \ ,
1982: \end{equation}
1983: in agreement with equations (\ref{scluster},\ref{eqb}).
1984: Summing differential equations (\ref{sc2}) and (\ref{evolsh})
1985: for $s_{cluster}$ and $s_{in}$ respectively shows that the total
1986: entropy of solutions is $(1-\alpha)\, \ln 2$ (Section \ref{seccluster}).
1987:
1988:
1989:
1990: % algorithms
1991:
1992: \section{Dynamical phase transitions and search algorithms}
1993:
1994: The deep understanding of the statistical properties of
1995: 3-XORSAT makes this problem a valuable benchmark for assessing the
1996: performances of
1997: various combinatorial search algorithms. At first sight, the idea
1998: seems rather odd since 3-XORSAT is a polynomial problem.
1999: Interestingly most of the search procedures devised to deal with
2000: NP-complete problems e.g. SAT have poor performances {\em i.e.}
2001: take exponentially long average running times
2002: on XORSAT above some algorithmic-dependent critical ratio ...
2003: The purpose of this Section is to present two algorithms exhibiting
2004: such a dynamical phase transition, and the techniques required for
2005: their analysis.
2006:
2007: \subsection{Random WalkSAT (RWSAT): definition, worst-case bound}
2008:
2009: The first algorithm we consider is the Random WalkSAT (RWSAT)
2010: algorithm introduced by Papadimitriou \cite{Pa92}.
2011: RWSAT is based on the observation that a violated equation can be satisfied
2012: through negation of one of its variables:
2013:
2014: {\sc
2015: \begin{itemize}
2016: \item Start from a randomly chosen configuration of the
2017: variables. Call energy the number $E$ of unsatisfied equations.
2018: \item While $E\ge 1$;
2019: \begin{itemize}
2020: \item pick up uniformly at random one of the $E$ unsatisfied equations;
2021: \item pick up uniformly at random one of its 3 variables;
2022: \item negate the value of this variable, update $E$;
2023: \end{itemize}
2024: \item Print 'Satisfiable', and Halt.
2025: \end{itemize}
2026: }
2027:
2028: \vskip .3cm
2029: Notice that, as a result of the negation, some equations
2030: that were satisfied may become violated. Therefore the energy is
2031: not guaranteed to decrease with the number of steps of the algorithm.
2032: RWSAT is able to escape from local minima
2033: of the energy landscape, and is {\em a priori} capable of better
2034: performances. From the other hand, RWSAT may run forever...
2035: A major question is how long should the algorithm be running
2036: before we stop thinking that the studied system has solutions
2037: hard to find and get some confidence that there is really no
2038: solution.
2039:
2040: This question was addressed by Sch\"oning \cite{Sc00}, who
2041: showed that RWSAT could easily be used as a one-sided
2042: randomized algorithm \cite{Mo95}\footnote{Sch\"oning's
2043: original work was devoted to
2044: the analysis of RWSAT on K-SAT, but his result holds for K-XORSAT
2045: too.}. Consider one instance of 3-XORSAT and run RWSAT for $3N$ steps
2046: from a randomly chosen configuration of variables. Choose again a
2047: random initial configuration and run RWSAT another $3N$ steps, and so on
2048: ... The probability that no solution has been found after $T$
2049: repetitions of this procedure though the formula is satisfiable is
2050: \begin{equation} \label{ressch}
2051: p_{SAT} \le \exp \left( - T \times \left(\frac 34\right)^{N +o(N)} \right) \ .
2052: \end{equation}
2053: Hence we obtain a probabilistic proof that the instance is not
2054: satisfiable if the algorithm has run unsuccessfully for more than
2055: $(\frac 43)^N$ sets of $3N$ steps. It must be clear that this result
2056: holds for any instance, no assumption being made on the distribution
2057: of formulas. The probability appearing in (\ref{ressch}) is on the
2058: random choices done by RWSAT and the choices of the restart
2059: configurations for a fixed formula.
2060:
2061: The proof of (\ref{ressch}) can be sketched as
2062: follows. Assume that the formula is satisfiable, and called $X^*$ one
2063: of its solutions. Consider now the (extensive) Hamming
2064: distance between the solution and the configuration $X$ of variables
2065: produced by RWSAT at some instant. After each step only one variable is
2066: changed so $D$ changes into $D+1$ (bad move) or $D-1$ (good move).
2067: Call $x,y,z$ the variables in the equation which was not satisfied by
2068: $X$. One or three of those variables have opposite values in $X^*$.
2069: In the latter case the flip is always a good move;
2070: in the former case the good move happens with probability $\frac 13$ and
2071: a bad move with probability $\frac 23$. On the overall the probability of
2072: a good move is $\frac 13$ at least.
2073:
2074: Think of $D$ has the position of a random walker on the $[0;N]$
2075: segment. Initially the position of the walker is a binomial variable,
2076: centered in $\frac N2$. At each step the walker moves to the left
2077: with probability $\frac 13$, and to the right with probability $\frac 23$.
2078: We look for the probability $\rho$ that the walker is absorbed by the
2079: boundary $D=0$ after $S$ steps. A standard calculation shows that
2080: $\rho$ is maximal for $S=3N$, with the value $\rho\simeq
2081: (\frac 34)^N$. After $T$ repetitions the probability of not having been
2082: absorbed is $(1-\rho)^T<\exp(-\rho\,T)$, hence (\ref{ressch}).
2083: The proof can be easily extended to $K$-XORSAT with higher values of $K$.
2084: The number of repetitions necessary to prove unsatisifiability
2085: scales as $(\frac{2(K-1)}K)^N$; it is essentially equal to $2^N$ for large
2086: $K$, showing that RWSAT does not beat exhaustive search in this limit.
2087:
2088: \subsection{Dynamical transition of RWSAT on random XORSAT instances}
2089: \label{secrwsat}
2090:
2091: \begin{figure}
2092: \begin{center}
2093: \includegraphics[width=110pt,angle=0]{fig521.eps}
2094: \hskip.5cm
2095: \includegraphics[width=120pt,angle=0]{fig522.eps}
2096: \end{center}
2097: \caption{Fraction $e$ of unsatisfied equations as a function of time $t$
2098: (number of steps divided by $N$) during the operation of RWSAT on a
2099: random 3-XORSAT formula at ratio $\alpha =0.2$ (left)
2100: and $\alpha =0.4$ (right) with $N=10^3$ (dotted), $10^4$ (dashed), $10^5$
2101: (full curve) variables. Note the difference
2102: of horizontal scales between the two figures. Inset: blow
2103: up of the $t\in[9.5;10.5]$ region; the amplitude of fluctuations around
2104: the plateau decreases with increasing size $N$. }
2105: \label{wsat_phen}
2106: \end{figure}
2107:
2108: Result (\ref{ressch}) is true for any instance; what is the typical
2109: situation for random systems? Numerical experiments
2110: indicate that there is critical value of the ratio of
2111: equations per variables, $\alpha_E\simeq 0.33$, hereafter referred to
2112: as dynamical threshold, separating two regimes:
2113: \begin{itemize}
2114: \item for $\alpha < \alpha _E$, RWSAT generally finds a solution very
2115: quickly, namely with a number of flips growing linearly with the
2116: number of variables $N$\footnote{A proof of this statement was
2117: obtained by \cite{Al02} for the random SAT model.}.
2118: Figure~\ref{wsat_phen} shows the plot of
2119: the fraction $e$ of unsatisfied clauses as a function of the time
2120: (number of steps) $T$ for one randomly drawn system with ratio
2121: $\alpha=0.2$ and $N=500$ variables. The curve shows a fast decrease from
2122: the initial value ($e(T=0)=\frac 12$
2123: independently of $\alpha$ for large values of $N$, but deviations can be
2124: found at small sizes, see Figure~\ref{wsat_phen}) down to zero on a time scale
2125: of the order of $N$\footnote{This decrease characterises the overall operation
2126: of RWSAT. A precise look at the $e(T)$ curve reveals that the energy $e$
2127: may occasionally increase.}. The resolution time $T_{res}$ depends
2128: both on the system of equations under consideration and the choices of
2129: the algorithm; its average value scales as
2130: \begin{equation}
2131: \langle T_{res} \rangle = N\, t _{res} + o(N) \ .
2132: \end{equation}
2133: where $t_{res}$ is an increasing function of
2134: $\alpha$\footnote{On intuitive grounds, as a step of the algorithm can satisfy
2135: $\theta(1)$ equations at a time, we expect the average value of
2136: $T_{res}$ to be of the order of the number $M$ of equations at least.
2137: Thus $t_{res}$ should grow at least linearly with $\alpha$. Experiments
2138: shows that the growth is in fact more than linear.}.
2139:
2140: \item for systems with ratios of equations per variable in the
2141: $\alpha_E<\alpha <\alpha_c$ range, the initial relaxation regime taking place
2142: on the $O(N)$ time scale does not allow RWSAT
2143: to reach a solution (Figure~\ref{wsat_phen}B). The fraction
2144: $e$ of unsat equations then fluctuates around some plateau value
2145: $e_{plateau}$
2146: for a very long time. Fluctuations are smaller and smaller (and the height of
2147: the plateau better and better defined) as the size $N$ increases.
2148: As a result of fluctuations, the fraction $e$ of unsatisfied equations
2149: may temporarily either increase or decrease. When a fluctuation happens to
2150: drive RWSAT to $e=0$, a solution is found and the algorithm stops.
2151: The corresponding resolution time, $T_{res}$, is stochastic;
2152: numerical experiments for different sizes $N$ indicate that its
2153: expectation value scale as
2154: \begin{equation} \label{deftaures}
2155: \langle T_{res} \rangle= \exp( N\, \tau _{res}+o(N)) \ .
2156: \end{equation}
2157: where the coefficient $\tau_{res}$ is an increasing function of $\alpha$.
2158: The plateau energy $e_{plateau}$ and the logarithm $\tau _{res}$ of
2159: the resolution time are shown in Figure~\ref{wsat_plateau}.
2160: \end{itemize}
2161:
2162: \begin{figure}
2163: \begin{center}
2164: A\includegraphics[width=110pt,angle=-90]{fig523.eps}
2165: B\includegraphics[width=110pt,angle=-90]{fig524.eps}
2166: \end{center}
2167: \caption{Fraction $e _{plateau}$ of unsatisfied equations on
2168: the plateau ({\bf A}) and logarithm
2169: $\tau _{res}$ of the average resolution time divided by $N$ ({\bf B}) as a
2170: function of the ratio $\alpha$ of equations per variable.
2171: Diamonds are the output of numerical experiments, and have been obtained
2172: through average of data from simulations over 1,000 systems and runs of RWSAT
2173: for various sizes $N$, and extrapolation to $N\to\infty$ \cite{Se03}.
2174: Full lines are theoretical approximations (\ref{eplateau}),(\ref{trestheo}).}
2175: \label{wsat_plateau}
2176: \end{figure}
2177:
2178: Notice that the dynamical threshold $\alpha_E$
2179: above which the plateau energy is positive is strictly smaller than the
2180: critical threshold $\alpha_c\simeq 0.918$, where systems go from
2181: satisfiable with
2182: high probability to unsatisfiable with high probability. In the
2183: intermediate range $\alpha_E<\alpha<\alpha_c$, systems are almost
2184: surely satisfiable but RWSAT needs an exponentially large time to
2185: prove so. The reason is that RWSAT remains trapped at a high energy level
2186: (plateau of Figure~\ref{wsat_plateau}) for an exponentially large
2187: time. The emergence of metastability can be qualitatively studied with simple
2188: tools we now expose.
2189:
2190: \subsection{Approximate theory for the metastable plateau and the escape time}
2191: \label{secescape}
2192:
2193: Assume that after $T$ steps of the algorithm the energy (number of unsatisfied
2194: equations) is $E_T\ge 1$. Then pick up an unsatisfied equation, say,
2195: $C$, and a variable in $C$, say, $x$, and flip it. The energy after the flip is
2196: \begin{equation} \label{evole}
2197: E_{T+1} = E_T -1 -U+S \ ,
2198: \end{equation}
2199: where $S$ (respectively $U$) is the number of equations including $x$
2200: which were satisfied (resp. unsatisfied after exclusion of equation $C$)
2201: prior to the flip. $S$ and $U$ are random variables with binomial
2202: distributions,
2203: \begin{eqnarray}
2204: \mbox{Proba}[U] &=& {E_T-1 \choose U} \left( \frac 3N\right)^U
2205: \left( 1-\frac 3N\right)^{E_T-1-U} \ , \nonumber \\
2206: \mbox{Proba}[S] &=& {M-E_T \choose S} \left( \frac 3N\right)^S
2207: \left( 1-\frac 3N\right)^{M-E_T-S} \ .
2208: \end{eqnarray}
2209: where the probabilities are intended over the formula content.
2210: Taking the average evolution equation (\ref{evole}) we obtain
2211: \begin{equation} \label{evole2}
2212: \langle E_{T+1}\rangle =\langle E_T\rangle -1 -\frac 3N \big(
2213: \langle E_T\rangle -1\big) +\frac 3N \big( M - \langle E_T\rangle\big) \ .
2214: \end{equation}
2215: The above equation is exact. It is now tempting to iterate it with time, from
2216: the initial condition $\langle E_{T=0}\rangle = \frac M2$. This is what we do
2217: hereafter but one should realize that this procedure is not
2218: correct from a mathematical standpoint. The catch is that one is allowed to
2219: average over the formula only once, and certainly not at each time step
2220: of the algorithm. Evolution
2221: equation (\ref{evole2}) amounts to redraw randomly the instance
2222: at each time step, conditioned to the energy. This approximation
2223: nevertheless allows
2224: us to write down a simple equation for $\langle E_T\rangle$, which
2225: captures much of the true behaviour of RWSAT.
2226:
2227: The next step in our analysis is the large size, large time limit. As
2228: the energy can typically change by a quantity of the order of unity
2229: in one time step we expect the fraction of unsatisfied equations to
2230: vary of a time scale of the order of $N$,
2231: \begin{equation}\label{evola}
2232: \langle E_T\rangle = M\; e\left( \frac TM =t\right) \ ,
2233: \end{equation}
2234: for some smooth function $e(t)$ of the reduced time $t$.
2235: Finite difference equation (\ref{evole2}) turns into a differential equation
2236: after insertion of (\ref{evola}),
2237: \begin{equation}\label{evole3}
2238: \frac{de}{dt} = - 1 + 3\alpha \, ( 1 -2\, e) \ ,
2239: \end{equation}
2240: with the initial condition $e(0)=\frac 12$. Clearly (\ref{evole3}) makes
2241: sense as long as $e>0$; if $e$ vanishes the algorithm stops.
2242: Resolution of (\ref{evole3}) shows the following scenario.
2243: If $\alpha$ is smaller than
2244: \begin{equation}
2245: \alpha _E = \frac 13 \ ,
2246: \end{equation}
2247: the fraction $e$ of unsatisfied equations quickly decreases, and
2248: vanishes at some time $t_{res}(\alpha)$. This regime corresponds to
2249: a successfull action of RWSAT in a $O(N)$ number of steps.
2250: $t_{res}$ is an increasing function of $\alpha$ which diverges as
2251: $\alpha \to\alpha_E$. Above this critical ratio $e$ shows a different
2252: behaviour: after a decreasing transient regime $e$ saturates to a
2253: positive plateau value
2254: \begin{equation}\label{eplateau}
2255: e_{plateau} (\alpha) = \frac 12 \left( 1 - \frac {\alpha_E}{\alpha}\right)
2256: \ .
2257: \end{equation}
2258: The value of the plateau energy is compared to numerics in Figure
2259: \ref{wsat_plateau}A. The agreement on the location of the dynamical
2260: threshold $\alpha_E$ as well as the plateau energy are satisfactory.
2261:
2262: The remaining point is to understand how RWSAT finally finds a solution
2263: when $\alpha >\alpha_E$. The above theory, based on taking the
2264: $N\to\infty$ limit first, washes out the fluctuations of the energy around
2265: its metastable value, of crucial importance for resolution
2266: \cite{Se03}. To take into account these fluctuations let us define
2267: the probability $Q_{plateau}(E)$ that the energy takes value $E$
2268: in the plateau regime of Figure \ref{wsat_phen}B. A stationary
2269: distribution is well defined if we discard the initial transient
2270: regime (choose large $t$) and collect values for $E$ on exponentially
2271: large--in--$N$ time scales. The procedure is standard in the
2272: study of long-time metastable states.
2273:
2274: Within our draw-instance-at-each--step approximation we may write
2275: a self-consistent equation for the stationary distribution
2276: of energies,
2277: \begin{equation} \label{statq}
2278: Q_{plateau}(E) = \sum _{U,S} \mbox{Proba}[U]\, \mbox{Proba}[S]\,
2279: \, Q_{plateau}(E+1+U-S)
2280: \end{equation}
2281: where the meaning of $U,S$ was explained right
2282: after (\ref{evole}). From Section \ref{secrwsat} we expect fluctuations
2283: to decreases sharply with the system size. A reasonable guess for the
2284: scaling of the distribution with $M$ is
2285: \begin{equation}
2286: Q_{plateau}(E) =
2287: \exp\left[ -M\; \omega \left(\frac EM =e\right) +o(M)\right]
2288: \end{equation}
2289: where $\omega$ is the rate function associated to the fraction
2290: of unsatisfied equations. Plugging the above Ansatz into (\ref{statq})
2291: and taking the large $M$ limit we find that $\omega$ fulfills the
2292: following differential equation
2293: \begin{equation}\label{pde1}
2294: F\left( \frac{\partial \omega}{\partial e},e\right) =0\ ,
2295: \end{equation}
2296: where $F(x,y)=3\alpha y(e^{-x}-1) +3\alpha (1-y) (e^x-1) -x$.
2297: This equation has to be solved with the condition
2298: $\omega(e_{plateau}) =0$.
2299:
2300: An analytical solution can be found for (\ref{pde1}) when we restrict
2301: to the vicinity of the dynamical transition {\em i.e.} to small
2302: values of $\omega$.
2303: Expanding $F$ to the second order in its first argument and solving
2304: (\ref{pde1}) we obtain
2305: \begin{equation} \label{ome}
2306: \omega(e) \simeq 2 \, (e - e_{plateau})^2\ ,
2307: \end{equation}
2308: where $e_{plateau}$ is defined in (\ref{eplateau}).
2309:
2310: What happens when time increases is now clear. Assume we have run RWSAT
2311: up to time $t\sim e^{M\tau}$. Then configurations with energy $e$ such that
2312: $\omega (e) <\tau$ have been visited many times and are 'equilibrated'
2313: with probability (\ref{statq}), (\ref{ome}). Configurations with
2314: energies outside the band $e_{plateau} \pm \sqrt{\tau /2}$ are not
2315: accessible. When the time scales reaches
2316: \begin{equation} \label{trestheo}
2317: \tau _{res} = \omega(0) \simeq \frac 12 \left( 1 - \frac 1{3\alpha}\right)^2
2318: \ ,
2319: \end{equation}
2320: zero energy configurations are encountered, and RWSAT comes to a stop.
2321: The agreement between the theoretical estimate (\ref{trestheo}) and
2322: the numerical findings (\ref{deftaures}) visible in Figure \ref{wsat_plateau}B
2323: is acceptable in regard to the crudeness of the approximation done.
2324:
2325: \subsection{Davis-Putnam-Loveland-Logemann (DPLL) algorithm}\label{secdpll}
2326:
2327: The second procedure is the Davis-Putnam-Loveland-Logemann
2328: (DPLL) algorithm \cite{dpll}.
2329: Contrary to RWSAT DPLL can provide exact proofs for
2330: unsatisfiability. The procedure, widely used in practice, is based on
2331: the trial-and-error principle. Variables are assigned
2332: according to some heuristic rule (split step), and
2333: equations involving those variables simplified. If an
2334: equation involving a single variable (unit-equation) appears
2335: its variable is chosen accordingly prior to any other
2336: heuristic assignment
2337: (unit-propagation). If a contradiction is found (two opposite
2338: unit-equations) DPLL backtracks to the last heuristically assigned
2339: variable, flips it, and resumes the search process.
2340: The procedure halts either when all equations have been
2341: satisfied (a solution is then found),
2342: or when all possible values for the variables have been tried in vane and
2343: found to be contradictory (a proof of unsatisfiability
2344: is then obtained).
2345:
2346: DPLL can be described as a recursive
2347: function of the variable assignment $A$.
2348: Given a system $S$ DPLL is first called
2349: with the empty assignment $A=\emptyset$:
2350: \vskip .3cm
2351: {\sc Procedure DPLL[$A$]
2352: \begin{itemize}
2353: \item Let $S_A$ be what is left from $S$ given variable assignment $A$;
2354: \item if $S_A$ is empty, Print `Satisfiable'; Halt;
2355: \item If $S_A$ contains a violated equation, Print `Contradiction', Return;
2356: {\em (backtracking)}
2357: \item Otherwise, let $U$ be the set of unit-equations in $S_A$;
2358: \begin{itemize}
2359: \item If $U \ne \emptyset$, pick-up one of the equations in
2360: $U$, say, $e$, and call DPLL[A$\cup \{e\}$];
2361: {\em (unit-propagation)}
2362: \item if $U=\emptyset$, choose a not-yet-assigned
2363: variable, say, $x$, and its value $v$ according to some heuristic rule, and
2364: call DPLL[A$\cup \{x=v\}$], then
2365: DPLL[A$\cup \{x=\bar v\}$]; {\em (variable splitting)}
2366: \end{itemize}
2367: \end{itemize}
2368: }
2369:
2370: \vskip .3cm
2371: Rules for assigning variables in the absence of unit-equations are heuristic
2372: in that they aim at doing good assumptions {\em i.e.} diminishing as much
2373: as possible the search process to come from limited information
2374: about the current system of equations. Of course, perfect heuristic
2375: do exist: trying all possible values for not-yet-assigned variables
2376: would ensure that no wrong guess is ever done! But the time required
2377: would be exponentially long. In practice, heuristics have to make
2378: their decision in polynomial time. Two simple splitting heuristics
2379: are:
2380: \begin{itemize}
2381: \item[$\diamond$] \underline{UC:}
2382: choose at random and uniformly any unset variable,
2383: and assign it to 0 or 1 with equal probabilities ($\frac 12$).
2384: \item[$\diamond$] \underline{GUC:}
2385: choose at random and uniformly any equation with minimal length {\em i.e.}
2386: involving 2 variables if any, or 3 variables otherwise. Pick up
2387: at random and uniformly one its variable, and assign it to 0 or 1
2388: with equal probabilities ($\frac 12$).
2389: \end{itemize}
2390: UC, which stands for unit-clause \cite{Ch90},
2391: amounts to make a random guess and is the simplest possible heuristic.
2392: GUC (Generalized UC)
2393: is more clever: each time a split is done from an equation with 2
2394: variables, this equation is turned into a unit-equation, and
2395: eliminated through unit-propagation.
2396: In the following, we call DPLL-UC and DPLL-GUC the variants of
2397: DPLL based on the UC and GUC heuristics respectively.
2398:
2399: \begin{figure}
2400: \begin{center}
2401: \includegraphics[width=110pt,angle=-90]{fig541.eps}
2402: \hskip .3cm
2403: \includegraphics[width=110pt,angle=-90]{fig542.eps}
2404: \end{center}
2405: \caption{Median number of splits required by DPLL
2406: with the UC (left) and GUC (right) heuristics as a function of the ratio
2407: $\alpha$, and for $N=20,40,60$ variables (from bottom to
2408: top). Data have been extracted from the resolution
2409: of 10,000 randomly drawn systems; continuous lines are guidelines for the
2410: eye. Note the difference of (logarithmic) scale between UC and GUC
2411: curves showing that DPLL-GUC is much more efficient than DPLL-UC.
2412: The polynomial/exponential transition is located at ratios
2413: $\alpha_E=\frac 23$ and $\alpha_E=0.7507...$ for UC and GUC respectively.}
2414: \label{dpllxorsat}
2415: \end{figure}
2416:
2417: A measure of the computational effort required
2418: by DPLL is the number $T_{split}$
2419: of variable splittings. This number varies from
2420: system to system (at fixed number $N$ of variables and ratio $\alpha$), and
2421: from run to run of DPLL due to the stochasticity introduced by
2422: the heuristic rule. The outcome of numerical experiments for
2423: the median number of splits\footnote{The median is more
2424: representative of the typical value of the number of splits than the
2425: expectation value, since the latter may be dominated by huge and
2426: unlikely samples, see discussion of Section \ref{secselfav}.}.
2427: For a given size $N$ $T_{split}$ shows a maximum located
2428: around $\alpha\simeq \alpha_c$. If one fixes $\alpha$
2429: $T_{split}$ is an increasing function of the size $N$;
2430: numerical data
2431: support the existence of a dynamical threshold, $\alpha_E$, separating
2432: linear and exponential scalings in $N$,
2433: \begin{equation} \label{mesure}
2434: T_{split}\sim \left\{\begin{array} {c c c}
2435: N\, t _{split} + o(N) &\hbox{\rm if} & \alpha < \alpha_E \\
2436: \exp( N\, \tau _{split}+o(N) ) &\hbox{\rm if} & \alpha > \alpha_E
2437: \end{array} \right. \ ,
2438: \end{equation}
2439: where $t_{split}$ and $\tau _{split}$ are functions of the ratio $\alpha$.
2440: The value of the dynamical threshold can be derived from theoretical
2441: calculations shown in Section \ref{secuc} and is equal to $\alpha_E=\frac 23$
2442: and $\alpha_E\simeq 0.7507...$ for UC and GUC heuristics respectively.
2443: Three dynamical regimes are therefored identified \cite{Co01,Ac02}:
2444: \begin{itemize}
2445: \item \underline{Linear \& satisfiable phase ($\alpha<\alpha_E$):} systems with
2446: small ratios are solved with essentially no backtracking. A solution
2447: is found after $O(N)$ splits.
2448: \item \underline{Exponential\& satisfiable phase ($\alpha_E<\alpha
2449: <\alpha_c$):}
2450: systems with ratios slightly below threshold have solutions, but
2451: DPLL generally requires an exponential number of splits to find
2452: one of them. An explanation for this drastic breakdown of performances
2453: will be given in Section \ref{secuc}.
2454: \item \underline{Exponential \& unsatisfiable phase ($\alpha>
2455: \alpha_c$):} finally,
2456: finding a proof of unsatisfiability typically requires an exponentially
2457: large number of splits \cite{Ch88}. Note that, as $\alpha$ gets higher
2458: and higher, each variable assignment affects more and more equations
2459: (of the order of $\alpha$), and contradictions
2460: are detected earlier and earlier. Rigorous calculations show that
2461: $\tau_{split} \sim \frac 1\alpha$\cite{beame},
2462: and the computational effort decreases
2463: with increasing $\alpha$ (Figure~\ref{dpllxorsat}). The median number
2464: of splits
2465: is considerably smaller for DPLL-GUC than for DPLL-UC, a result expected
2466: from the advantages of GUC against UC discussed above.
2467: \end{itemize}
2468:
2469: \subsection{Linear phase: resolution trajectories in the $2+p$-XORSAT
2470: phase diagram}\label{secuc}
2471:
2472: Action of DPLL on an instance of 3-XORSAT causes changes to the
2473: numbers of variables and equationses, and thus to the ratio
2474: $\alpha$. Furthermore DPLL turns equations with 3 variables into
2475: equation with 2 variables. A mixed $2+p$-XORSAT distribution, where $p$
2476: is the fraction of 3-equations and $\alpha$ the ratio of the
2477: total number of 2- and 3- equations over the number of variables
2478: can be used to model what remains of the input
2479: system\footnote{Equations with a single variable
2480: are created too, but are eliminated through unit-propagation. When
2481: a heuristic assignment has to be made the system is a mixture of
2482: equations with 2 and 3 variables only.}. Repeating the calculations of
2483: Section \ref{secreplicas} for the $2+p$-XORSAT models we derive the
2484: phase diagram of Figure \ref{diag2+p}. The Sat/Unsat critical
2485: line $\alpha_c(p)$ separates the satisfiable from the unsatisfiable
2486: phases. For $p \le p_0 = \frac 14$ {\em i.e.} to the left of
2487: point T, the threshold line coincides with the percolation
2488: transition as in the 2-XORSAT model, and is given by
2489: $\alpha _c(p)=\frac 1{2(1-p)}$.
2490: For $p>p_0$ an intermediate clustered phase is found as in the
2491: 3-XORSAT model, and the threshold coincides with the vanishing of the
2492: cluster entropy $s_{cluster}$ (Section \ref{seccluster}).
2493:
2494:
2495: The phase diagram of 2+p-XORSAT is the natural space in which DPLL
2496: dynamic takes place. An input 3-XORSAT instance with ratio $\alpha$ shows
2497: up on the right vertical boundary of Figure~\ref{diag2+p} as a point of
2498: coordinates $(p=1,\alpha )$. Under the action of DPLL the
2499: representative point moves aside from the 3-XORSAT axis and follows a
2500: trajectory, very much alike real-space renormalization, which
2501: depends on the splitting heuristic. Trajectories enjoy two essential
2502: features \cite{Ac01}. First the representative point
2503: of the system treated by DPLL does not `leave' the 2+p-XORSAT phase
2504: diagram. In other words, the instance is, at any stage of the search
2505: process, uniformly distributed from the 2+p-XORSAT
2506: distribution conditioned to its equation per variable ratio $\alpha$ and
2507: fraction $p$ of 3-equations. This assumption is not true for
2508: all heuristics of split, but holds for UC and
2509: GUC\cite{Ch90}\footnote{Analysis of more sophisticated heuristics
2510: e.g. based on the number of occurences of variables require to handle
2511: more complex instance distributions \cite{Ka02}.}.
2512: Secondly, the trajectory followed by an instance in the course of
2513: resolution is a stochastic object, due to the randomness of
2514: the instance and of the assignments done by DPLL.
2515: In the large size limit ($N\to\infty$) the trajectory becomes
2516: self-averageing {\em i.e.}
2517: concentrated around its average locus in the 2+p-XORSAT phase
2518: diagram \cite{Wo95}. We will come back below on this concentration
2519: phenomenon.
2520:
2521:
2522: \begin{figure}
2523: \begin{center}
2524: \includegraphics[width=160pt,angle=0]{fig551.eps}
2525: \caption{Phase diagram of $2+p$-XORSAT and dynamical trajectories of DPLL.
2526: The threshold line $\alpha_c (p)$ (bold full line) separates sat
2527: from unsat phases. Departure points for DPLL trajectories are located on the
2528: 3-XORSAT vertical axis with ratios $.4,\frac 23,.8,1.$ from
2529: bottom to top. The arrow indicates the direction of motion along trajectories
2530: parametrized by the fraction $t$ of variables set by DPLL. For small
2531: ratios $\alpha < \alpha _E$ ($=\frac 23$ for the UC heuristic)
2532: trajectories remain
2533: confined in the sat phase, end in S of coordinates $(0,0)$, where a
2534: solution is found. At $\alpha_E$ the trajectory hits
2535: tangentially the threshold line in T of coordinates $(\frac 14,\frac 23)$.
2536: When $\alpha >\alpha _E$ the
2537: trajectories intersect the threshold line at some point G (which depends
2538: on $\alpha$), and stops before hitting the
2539: $\alpha _D (p)$ dotted line (\ref{defcon}).
2540: After massive backtracking DPLL will find a solution; G corresponds
2541: to the highest node in the search tree.}
2542: \label{diag2+p}
2543: \end{center}
2544: \end{figure}
2545:
2546: Let $\alpha_0$ denote the equation per variable
2547: ratio of the 3-XORSAT instance to be solved.
2548: We call $E_j (T)$ the number of $j$--equations (including $j$ variables)
2549: after $T$ variables have been assigned by the solving procedure.
2550: $T$ will be called hereafter `time', not to be confused with the
2551: computational effort. At time $T=0$ we
2552: have $E_3 (0)= \alpha_0 N$, $E_2 (0)=E_1 (0)=0$.
2553: Assume that the variable $x$ assigned at time $T$ is chosen through
2554: unit-propagation, that is, independently of the $j$-equation content.
2555: Call $n_j(T)$ the number of occurrences of $x$ in
2556: $j$-equations ($j=2,3$). The evolution equations for
2557: the populations of 2-,3-equations read
2558: \begin{equation}
2559: E_3 (T+1) = E_3 (T) - n_3(T)\ , \quad E_2(T+1)=E_2(T) - n_2(T)+ n_3(T) \ .
2560: \label{evolsto}
2561: \end{equation}
2562: Flows $n_2,n_3$ are of course random variables that depend
2563: on the instance under consideration at time $T$, and on the choice of
2564: variable done by DPLL. What are their distributions? At time $T$
2565: there remain $N-T$ untouched variables; $x$ appears in any of
2566: the $E_j(T)$ $j$-equation with probability $p_j = \frac j{N-T}$,
2567: independently of the other equations. In the
2568: large $N$ limit and at fixed fraction of assigned variables,
2569: $t=\frac TN$, the binomial distribution
2570: converges to a Poisson law with mean
2571: \begin{equation} \label{avpois}
2572: \langle n_j \rangle _T=\frac{j\, e_j}{1-t} \qquad \mbox{where} \qquad
2573: e_j= \frac{E_j(T)}N
2574: \end{equation}
2575: is the density of $j$-equations at time $T$. The key remark is that,
2576: when $N\to\infty$, $e_j$ is a slowly varying and non stochastic
2577: quantity and is a function of the fraction
2578: $t=\frac TN$ rather than $T$ itself. Let us iterate (\ref{evolsto})
2579: between times $T_0=t\,N$ and $T_0+ \Delta T$ where $1\ll \Delta T
2580: \ll N$ e.g. $\Delta T =O(\sqrt N)$. Then the change $\Delta E_3$ in the number
2581: of $3$-equations is (minus) the sum of the stochastic variables
2582: $n_j(T)$ for $T=T_0,T_0+1, \ldots,T_0+\Delta T$. As these variables
2583: are uncorrelated Poisson variables with $O(1)$ mean (\ref{avpois})
2584: $\Delta E_3$ will be of the order of $\Delta T$, and the change in
2585: the density $e_3$ will be of order of $\Delta T/N \to 0$. Applying
2586: central limit theorem $\Delta E_3/\Delta T$ will be almost surely
2587: equal to $-\langle n_3\rangle_t$
2588: given by (\ref{avpois}) and with the
2589: equation density measured at reduced time $t$. The argument can be
2590: extended to 2-equations, and we conclude that
2591: $e_2,e_3$ are deterministic (self-averaging) quantities obeying
2592: the two coupled differential equations\cite{Ch90}
2593: \begin{equation}
2594: \frac{de_3}{dt}(t) = -\frac{3\, e_3}{1-t} \quad
2595: , \qquad \frac{de_3}{dt}(t) = \frac{3\,
2596: e_3}{1-t} - \frac{2\, e_2}{1-t} \ .
2597: \end{equation}
2598: Those equations, together with the initial condition $e_3(0)=\alpha
2599: _0$, $e_2(0)=0$ can be easily solved,
2600: \begin{equation}\label{sole23}
2601: e_3(t) = \alpha_0 (1-t)^3 \quad , \qquad e_2(t)= 3\,\alpha_0
2602: \,t\,(1-t)^2 \ .
2603: \end{equation}
2604: To sum up, the dynamical evolution of the equation populations may be
2605: seen as a slow and deterministic evolution of the
2606: equation densities to which are superimposed fast, small fluctuations.
2607: The distribution of the fluctuations adiabatically follows
2608: the slow trajectory. This scenario is pictured in Figure~\ref{deterstoch}.
2609:
2610: \begin{figure}
2611: \begin{center}
2612: \includegraphics[width=160pt,angle=0]{fig552.eps}
2613: \caption{Deterministic versus stochastic dynamics of the equation
2614: population $E$ as a function of the number of steps $T$ of the algorithm.
2615: On the slow time scale (fraction $t=T/N$) the density
2616: $e=E/N$ of (2- or 3-) equations varies smoothly according to a deterministic
2617: law. Blowing up of the dynamics around some point $t',e'$
2618: shows the existence of small and fast fluctuations around this trajectory.
2619: Fluctuations are stochastic but their probability distribution
2620: depends upon the slow variables $t',e'$ only.}
2621: \label{deterstoch}
2622: \end{center}
2623: \end{figure}
2624:
2625: Expressions (\ref{sole23}) for the equation densities allow us to draw
2626: the resolution trajectories corresponding to the action of DPLL on a
2627: 3-XORSAT instance. Initially the instance is represented by a point
2628: with coordinates $(p=1,\alpha=\alpha_0)$ in Figure \ref{diag2+p}. As
2629: more and more variables are assigned the representative point moves
2630: away from the rightmost vertical axis. After a fraction $t$ of
2631: variables have been assigned the coordinates of the point are
2632: \begin{equation}
2633: p(t) = \frac{e_3}{e_2+e_3}=\frac{1-t}{1+2t}\ , \quad
2634: \alpha(t) = \frac{e_2+e_3}{1-t}=\alpha_0(1-t)(1+2t) \ .
2635: \end{equation}
2636: Trajectories corresponding to various initial ratios are shown in
2637: Figure \ref{diag2+p}.
2638: For small ratios $\alpha_0 < \alpha _E$ trajectories remain
2639: confined in the sat phase, end in S of coordinates $(0,0)$, where a
2640: solution is found. At $\alpha_E$ ($=\frac 23$ for the UC heuristic),
2641: the single branch trajectory hits
2642: tangentially the threshold line in T of coordinates $(\frac 14,\frac 23)$.
2643: When $\alpha_0 > \alpha _E$ the trajectories enter the Unsat phase,
2644: meaning that DPLL has turned a satisfiable instance (if $\alpha _0 <
2645: \alpha _c$) into an unsatisfiable one as a result of poor assignments.
2646: It is natural to expect that $\alpha_E$ is the highest ratio at which
2647: DPLL succeeds in finding a solution without resorting to much backtracking.
2648:
2649:
2650:
2651: \subsection{Dynamics of unit-equations and universality}\label{secuni}
2652:
2653: The trajectories we have derived in the previous Section are correct
2654: provided no contradiction emerges. But
2655: contradictions may happen as soon as there are $E_1=2$
2656: unit-equations, and are all the more likely than $E_1$ is large.
2657: Actually the set of 1-equations form a 1-XORSAT instance which is
2658: unsatisfiable with a finite probability as soon as $E_1$ is of the order
2659: of $\sqrt N$ from the results of Section \ref{secsf1}.
2660: Assume now that $E_1 (T)\ll N$ after $T$ variables have been assigned,
2661: what is the probability $\rho_T$ that no contradiction
2662: emerges when the $T^{th}$ variable is assigned by DPLL? This probability is
2663: clearly one when $E_1=0$. When $E_1\ge 1$ we pick up a 1-equation,
2664: say, $x_{6}=1$, and wonder whether the
2665: opposite 1-equation, $x_6=0$, is present among the $(E_1-1)$ 1-equations
2666: left. As equations are uniformly distributed over the set
2667: of $N-T$ untouched variables
2668: \begin{equation}\label{probanocttot}
2669: \rho_T = \left( 1 - \frac 1{2 (N-T)} \right) ^{\max( E_1(T) -1,0)}
2670: \ .
2671: \end{equation}
2672: The presence of the $\max$ in the above equation ensures it remains
2673: correct even in the absence of unit-equations ($E_1=0$).
2674: $E_1(T)$ is a stochastic variable. However from the
2675: decoupling between fast and slow time scales sketched in Figure
2676: \ref{deterstoch} the probability distribution of $E_1(T)$
2677: depends only on the slow time scale $t$. Let us call $\mu (E_1;t)$
2678: this probability. Multiplying (\ref{probanocttot}) over the times
2679: $T=0$ to $T=N-1$ we deduce the probability that DPLL has successfully
2680: found a solution without ever backtracking,
2681: \begin{equation} \label{rhosuc}
2682: \rho _{success} = \exp \left( - \int _0^1 \frac{dt}{2(1-t)}
2683: \sum _{E_1\ge 1} \mu (E_1;t)\; (E_1-1) \right)
2684: \end{equation}
2685: in the large $N$ limit.
2686:
2687: \begin{figure}
2688: \begin{center}
2689: \includegraphics[width=200pt,angle=0]{fig561.eps}
2690: \caption{Evolution of the number $E_1$ of 1-equations as one more variable
2691: is assigned. $n_2$ denotes the number of 2-equations reduced to
2692: 1-equations, $s_1$ the number of 1-equations satisfied. If $E_1\ge 1$ a
2693: variable is fixed through unit-propagation: $E_1$ decreases by one
2694: plus $s_1$, and increases by
2695: $n_2$. In the absence of unit-equation ($E_1=0$) the number of
2696: 1-equations after the assignment is simply $E_1'=n_2$. }
2697: \label{rw1clause}
2698: \end{center}
2699: \end{figure}
2700:
2701: We are left with the calculation of $\mu$\cite{Fr96}.
2702: Figure \ref{rw1clause} sketches the stochastic evolution of the number
2703: $E_1$ during one step. The number of 1-equations produced
2704: from 2-equations, $n_2$, is a Poisson variable
2705: with average value, from (\ref{sole23}),
2706: \begin{equation}\label{valued}
2707: d(t) = \frac {2 \, e_2 (t)}{1-t} = 6\,\alpha_0\, t(1-t)\,
2708: \end{equation}
2709: when $N\to \infty$. The number of
2710: satisfied 1-equations, $s_1$, is negligible as long as $E_1$ remains
2711: bounded. The probability that the number of 1-equations
2712: goes from $E_1$ to $E_1'$ when $T\to T+1$ defines the entry of the
2713: transition matrix
2714: \begin{equation}\label{matrixmm}
2715: M (E'_1,E_1;t) = \sum _{n_2 \ge 0} e^{-d(t)} \frac{d(t)^{n_2}}{n_2!}
2716: \delta _{E'_1-(E_1+n_2-\delta_{E_1})}\ .
2717: \end{equation}
2718: from which a master equation for the probability of
2719: $E_1$ at time $T$ may be written.
2720: On time scales $1\ll \Delta T\ll N$ this master equation
2721: converges to the equilibrium distribution $\mu$ \cite{Fr96,Co01},
2722: conveniently expressed in terms of the generating function
2723: \begin{equation} \label{defgtstat}
2724: G( x;t ) = \sum _{E_1\ge 0} \mu(E_1;t)\; x^{E_1} =
2725: \frac{(1-d(t))(x-1)}{x\; e^{d(t) \; (1-x)} -1} \;
2726: \ .
2727: \end{equation}
2728: The above is a sensible result for $d(t)\le 1$ but does not make sense when
2729: $d(t)>1$ since a probability cannot be negative! The reason is that we
2730: have derived (\ref{defgtstat}) under the implicit condition that no
2731: contradiction was encountered. This assumption cannot hold when the
2732: average rate of 1-equation production, $d(t)$, is larger that one, the
2733: rate at which 1-equations are satisfed by unit-propagation.
2734: From (\ref{valued}) we see, when $\alpha > \alpha _E=\frac 23$, the
2735: trajectory would cross the
2736: \begin{equation}\label{defcon}
2737: \alpha_D(p) = \frac {1} {2(1-p)}
2738: \end{equation}
2739: on which $d=1$ for some time $t_D<1$. A contradiction is very
2740: likely to emerge before the crossing.
2741:
2742:
2743: When $\alpha <\alpha _E$ $d$ remains smaller than unity at any time.
2744: In this regime the probability of
2745: success reads, using (\ref{rhosuc}) and (\ref{defgtstat}),
2746: \begin{equation} \label{nobtsuc}
2747: \rho_{success} = \exp \left(\frac {3\alpha}{4} - \frac 12
2748: \sqrt{\frac{3\alpha}{2-3\alpha}}\; \tanh^{-1}
2749: \bigg[ \sqrt{\frac{3\alpha}{2-3\alpha}}\bigg] \right)\ .
2750: \end{equation}
2751: $\rho_{success}$ is a decreasing
2752: function of the ratio $\alpha$, down from unity for $\alpha =0$ to zero for
2753: $\alpha =\alpha_{E}$.
2754: The present analysis of the UC heuristic can be easily transposed to the GUC
2755: heuristic. Details are not given here but can be found in
2756: \cite{Ac02,Co01}. The result is an expression
2757: for $\rho_{success}$ larger than its UC counterpart (\ref{nobtsuc}),
2758: and vanishing in $\alpha _E\simeq 0.7507$. Interestingly the way
2759: $\rho_{success}$ vanishes when $\alpha$ reaches $\alpha_E$,
2760: \begin{equation}
2761: - \ln \rho_{success} (\alpha_E - \epsilon) \sim \epsilon ^{-\frac 12}
2762: \qquad (\epsilon \to 0^+)
2763: \end{equation}
2764: is the same for both heuristics. This similarity extends to a whole
2765: class of heuristics which can be described by the flow of equation
2766: densities only and based on unit-propagation \cite{De04}. The
2767: probability that DPLL finds a solution without backtracking
2768: to a 3-XORSAT instance of size $N$ satisfies finite-size scaling at the
2769: dynamical critical point,
2770: \begin{equation}\label{ffss}
2771: - \ln \rho_{success} (\alpha_E - \epsilon, N) \sim N^{\frac 16} \;
2772: \Phi\big( \epsilon \, N^{\frac 13} \big) \ ,
2773: \end{equation}
2774: where the scaling function $\Phi$ is independent of the heuristics
2775: and can be calculated exactly \cite{De04}. The exponent
2776: characterizing the width of the critical region is the one associated
2777: to percolation in random graphs (\ref{valpsi}). A consequence of
2778: (\ref{ffss}) is that, right at $\alpha_E$, $\rho_{success} \sim \exp
2779: (- Cst\times N^{\frac 16})$ decreases as a stretched exponential of
2780: the size. The value of the exponent, and its robustness against the
2781: splitting heuristics can be understood from the following
2782: argument \cite{De04}.
2783:
2784: Let us represent 1- and 2- equations by a graph $G$ over the set
2785: of $N-T$ vertices (one for each variable $x_i$) with $E_1$ marked
2786: vertices (one for each unit-equation $x_i=0,1$), and $E_2$
2787: signed edges ($x_i + x_j=0,1$), see Section \ref{secrg}.
2788: $d$ is simply the average degree of vertices in $G$.
2789: Unit-propagation corresponds to removing
2790: a marked vertex (and its attached edges), after having marked its
2791: neighbours; the process is iterated until the connected component is
2792: entirely removed (no vertex is marked). Meanwhile, new edges
2793: have been created from the reduction of 3-equations into 2-equations.
2794: Then a vertex is picked up according to the heuristic and marked,
2795: and unit-propagation resumes. The
2796: success/failure transition coincides with the percolation
2797: transition on $G$: $d=1$ as expected.
2798: From random graph theory \cite{Bo89} the
2799: percolation critical window is of width
2800: $|d -1| \sim N^{-1/3}$. As $d$ is
2801: proportional to the ratio $\alpha _0$ (\ref{valued})
2802: we find back $\psi= \frac 13$.
2803: The time spent by resolution trajectories in the critical
2804: window is $\Delta t \sim \sqrt {|d -1|} \sim N^{-1/6}$,
2805: corresponding to $\Delta T = N \, \Delta t \sim N^{5/6}$ eliminated
2806: variables. As the largest components have size $S\sim N^{2/3}$ the
2807: number of such components eliminated is $C=\Delta T/S\sim N^{1/6}$.
2808: What is the probability $q$ that a large component is removed without
2809: encountering a contradiction? During the removal of the component the
2810: number of marked vertices `freely' diffuses, and reaches $E_1\sim \sqrt
2811: {S}\sim N^{1/3}$. The probability that no contradiction occurs
2812: is, from (\ref{probanocttot}), $q\sim (1-\frac {Cst}N)^{E_1\times S}$,
2813: a finite quantity. Thus $\rho_{success} \sim q^C \sim
2814: \exp(-N^{1/6})$. The presence of numerous, smaller components does
2815: not affect this scaling.
2816:
2817:
2818:
2819: \subsection{Exponential phase: massive backtracking}
2820:
2821: For ratios $\alpha _0>\alpha _E$ DPLL is very likely to find a
2822: contradiction. Backtracking enters into play, and is responsible for
2823: the drastic slowing down of the algorithm (Figure \ref{dpllxorsat}).
2824:
2825: The history of the search process can be represented by a search tree,
2826: where the nodes represent the variables assigned, and the descending
2827: edges their values (Figure \ref{fig-tree}).
2828: The leaves of the tree correspond to solutions (S), or
2829: to contradictions (C). The analysis of the $\alpha<\alpha_E$ regime
2830: leads us to the conclusion that search trees look like
2831: Figure \ref{fig-tree}A at small ratios\footnote{A small amount of
2832: backtracking may be necessary to find the solution since
2833: $\rho_{success}<1$ \cite{Fr96}, but the overall picture of a single
2834: branch is not qualitatively affected.}.
2835: Consider now the case of unsatisfiable formulas
2836: ($\alpha_0> \alpha_c$) where all leaves carry contradictions after
2837: DPLL halts (Figure \ref{fig-tree}C).
2838: DPLL builds the tree in a sequential manner, adding nodes
2839: and edges one after the other, and completing branches through
2840: backtracking steps. We can think of the same search tree built in a
2841: parallel way\cite{Co01}. At time (depth $T$) our tree is
2842: composed of $L(T)\le 2^T$ branches, each carrying a partial assignment
2843: over $T$
2844: variables. Step $T$ consists in assigning one more variable to each
2845: branch, according to DPLL rules, that is, through unit-propagation or
2846: split. Possible consequences are: emergence of a contradiction and end
2847: of the branch, simplification of the attached formulas and the branch keeps
2848: growing.
2849:
2850: \begin{figure}
2851: \begin{center}
2852: \includegraphics[width=130pt,angle=-90]{fig571.eps}
2853: \caption{Search trees in three regimes of Section \ref{secdpll}:
2854: {\bf A.} linear, satisfiable ($\alpha < \alpha_E$); {\bf B.}
2855: exponential, satisfiable ($\alpha_E<\alpha<\alpha_c$);
2856: {\bf C.} exponential, unsatisfiable ($\alpha > \alpha _c$). Leaves
2857: are marked with S (solutions) or C (contradictions). G is the highest
2858: node to which DPLL backtracks, see Figure \ref{diag2+p}.}
2859: \label{fig-tree}
2860: \end{center}
2861: \end{figure}
2862:
2863: The number of branches $L(T)$ is a stochastic variable. Its average
2864: value can be calculated as follows \cite{Mon05}. Let us define the
2865: average number $L(\vec E;T)$ of branches with equation populations $\vec
2866: E=(E_1,E_2,E_3)$ at depth $T$. Initially $L(\vec E;0)=1$ for
2867: $\vec E=(0,0,\alpha_0 N)$, 0 otherwise.
2868: Call $M(\vec E',\vec E;T)$ the average
2869: number of branches with population $\vec E'$ generated from a branch with
2870: population $\vec E$ once the $T^{th}$ variable is assigned. Transition
2871: matrix $M$ is an extension of (\ref{matrixmm})
2872: to the whole population vector $\vec E$ and not only $E_1$.
2873: We have $0\le M\le 2$, the extreme values corresponding to a
2874: contradiction and to a split respectively. We claim that
2875: \begin{equation}\label{evol5}
2876: L(\vec E';T+1) = \sum _{\vec E} M(\vec E',\vec E;T) \; L(\vec E;T) \ .
2877: \end{equation}
2878: Evolution equation (\ref{evol5}) is somewhat suspicious since it looks like
2879: the approximation (\ref{statq}) we have done in the analysis of
2880: RWSAT. Yet a major difference exists which makes (\ref{evol5})
2881: exact \cite{Mon05}. Drawing randomly many times the same instance, as
2882: we are doing, is in principle forbidden but not along {\em one} branch
2883: for the very reason the analysis of Section \ref{secuc} was
2884: correct. Actually what we have done in Section \ref{secuc} is to draw
2885: randomly at time $T$ the equations containing the $T^{th}$
2886: variable. But this is correct since those equations are immediately
2887: simplified into shorter equations and their remaining content remains
2888: unknown \cite{Ka02}. The situation seems more complicated in the case
2889: of the whole tree since the same equation can appear at different
2890: depth along distinct branches. Indeed the number of branches produced
2891: from two distinct branches after assignment of one variable are correlated
2892: variables. But thanks to the linearity of expectation those
2893: correlations do not matter and (\ref{evol5}) is correct.
2894:
2895: Transition matrix $M$ can be explicitely written down. It is more
2896: convenient to write (\ref{evol5}) for the generating function of the
2897: number of branches, $B(\vec x;T) = \sum _{\vec E} L(\vec E;T)
2898: x_1^{E_1}\,x_2^{E_2}\,x_3^{E_3}$, with the result
2899: \begin{equation} \label{evol6}
2900: B(\vec x;T+1) = \frac 1{f_1} B(\vec f;T) + (2-\frac 1{f_1})
2901: B(0,f_2,f_3;T) - 2 B(\vec 0;T)\ ,
2902: \end{equation}
2903: where $\vec f$ is the vector with components
2904: \begin{equation}
2905: f_1= x_1 + \frac{\frac 12-x_1}{N-T}, f_2 = x_2 + \frac{2(
2906: x_2-x_1)}{N-T}, f_3 = x_3 + \frac{3( x_2-x_3)}{N-T} \ .
2907: \end{equation}
2908: The three terms on the r.h.s. of (\ref{evol6}) correspond, from left
2909: to right: unit-propagation (the branch keeps growing), variable
2910: splitting (2 branches are created from the previous one), branches
2911: carrying empty instances (satisfied instance).
2912: Equation (\ref{evol6}) together with the initial condition $B(\vec
2913: x;0)= x_3^{\alpha _0N}$ completely defines the average dynamics of the
2914: search tree. We sketch the main steps of its resolution below\cite{Co01}:
2915: \begin{enumerate}
2916: \item To count the number of branches irrespectively of the number of
2917: unit-equations we should consider the value $x_1=1$. However, as long
2918: as branches grow the number $E_1$ of unit-equations cannot be large,
2919: and remains bounded. We can therefore choose $x_1=\frac 12$ which
2920: simplifies (\ref{evol6}) without affecting the large size scaling of
2921: $L$ and $B$. This technical trick is reminiscent of Knuth's kernel
2922: method \cite{Kn66}.
2923: \item For large $N$ it is reasonable to expect that the number of
2924: branches grows exponentially with the depth, or, equivalently,
2925: \begin{equation}
2926: \sum _{E_1} L(E_1,E_2,E_3;T) \sim e ^{N \; \lambda (e_2,e_2;t) + o(N)}
2927: \end{equation}
2928: where $e_2,e_3$ are the densities of equations as usual. From point 1
2929: the Legendre transform of $\lambda$
2930: \begin{equation}\label{degpp}
2931: \gamma (x_2,x_3;t) = \max _{e_2,e_3} \big[ \lambda (e_2,e_2;t) +
2932: e_2 \ln x_2 + e_3 \ln x_3 \big]
2933: \end{equation}
2934: fulfills the partial differential equation (PDE)
2935: \begin{equation}
2936: \frac{\partial \gamma} {\partial t} = \ln 2 + \frac{1-2x_2}{1-t}
2937: \frac{\partial \gamma}{\partial x_2} + \frac{3(x_2-x_3)}{1-t}
2938: \frac{\partial \gamma}{\partial x_3} \ .
2939: \end{equation}
2940: with the initial condition $\gamma (x_2,x_3;t) = \alpha _0 \ln x_3$.
2941: \item The first order PDE can be solved exactly with the
2942: characteristic method. The output, after Legendre inversion
2943: through (\ref{degpp}), is the entropy $\lambda (e_2,e_3;t)$
2944: of branches at reduced depth $t$. Let us call $\lambda ^*(t)$ the
2945: maximum value of $\lambda$ over the equation densities for a fixed
2946: fraction $t$ of assigned variables.
2947: \item $\lambda ^*(t)$ is a function growing from $\lambda ^*=0$ at
2948: $t=0$, reaching a maximum value $\lambda ^*_M$ in $t_M$, and
2949: decreasing for larger times $t\le 1$. $t_M$ is the depth in the tree
2950: of Figure \ref{fig-tree}C where most contradictions are found; the
2951: number of C leaves is, to exponential order, $e^{N\lambda^*_M}$. We
2952: conclude that the size of the tree we were looking for is
2953: \begin{equation}
2954: \tau _{split} = \lambda^*_M \ ,
2955: \end{equation}
2956: compare with (\ref{mesure}). For large $\alpha \gg \alpha_c$ one finds
2957: $\tau_{split}\sim \ln 2 /(6\alpha)$ in agreement with \cite{beame}.
2958: The calculation can be extended to highers values of $K$.
2959: \end{enumerate}
2960:
2961: The above calculation holds for the unsatisfiable, exponential phase.
2962: How can we understand the satisfiable but exponential regime
2963: $\alpha_E<\alpha_0<\alpha_c$? The resolution trajectory crosses the
2964: Sat/Unsat critical line at some point G shown in Figure
2965: \ref{diag2+p}. Immediately after $G$ the instance left by DPLL is
2966: unsatisfiable. A subtree with all its leaves carrying contradictions
2967: will develop below G (Figure \ref{fig-tree}B). The size $\tau_{split}
2968: ^G$ of this
2969: subtree can be easily calculated from the above theory. The only
2970: change is the initial condition over $\gamma$:
2971: $\gamma (x_2,x_3;0) = \alpha_G (p_G \ln
2972: x_3 + (1-p_G) \ln x_2)$ where $(p_G,\alpha_G)$ are the coordinates of
2973: G which can be calculated from $\alpha_0$ and the knowledge of the
2974: critical Sat/Unsat line. Once this subtree has been built DPLL
2975: backtracks to G, flips the attached variable and will finally end up
2976: with a solution. Hence the (log of the) number of splits necessary will be
2977: typically equal to $\tau_{split} = (1-t_G)\, \tau_{split}^G$ \cite{Co01}.
2978:
2979: % conclusion
2980:
2981: \section{Conclusions}\label{secconclusion}
2982:
2983: Previous Sections have allowed us to illustrate rather general
2984: techniques and ideas to deal with random systems. It does not come as
2985: a surprise that other problems than XORSAT e.g. the satisfaction of
2986: Boolean constraints, graph coloring, the covering of vertices, ...
2987: have been successfully studied with these tools. Many of those
2988: problems, when given
2989: an input distribution based on random graphs, actually share a lot
2990: of common features with XORSAT. The reader
2991: is referred to
2992: \cite{Mo97,Bi00,Me02,Me03b,Se06,Ac05} (satisfiability),
2993: \cite{Mu02,Se06} (coloring),
2994: \cite{We01b,We01} (vertex cover), ...
2995: for entry points to the literature. Let us also mention that
2996: many other interesting optimization problems, not directly related to
2997: random graphs, have been studied with the techniques of Sections 4
2998: and 5, and the results sometimes rigorously proven
2999: e.g. matching \cite{Or85,Me87b,Al01},
3000: traveling salesman \cite{Me86},
3001: number partitioning \cite{Me98,Me01},
3002: graph partitioning \cite{Fu85}, ...
3003: Finally, from a historical point of view, one should not forget
3004: that statistical mechanics tools have found numerous and beautiful
3005: applications in the study of the learning and storage
3006: properties of neural networks\cite{Am89,Va01}, all the more
3007: so the random satisfiability problem can be recast as an Ising
3008: perceptron problem \cite{Kr89}.
3009:
3010: The study of random optimization problems is obviously interesting
3011: from a probabilistic point of view. As far as computer science is
3012: concerned they can be seen as useful benchmarks for testing and
3013: improving resolution procedures. A successful example is the
3014: traduction of the cavity equations of Section 5 into an algorithm for
3015: solving given instances of the satisfiability problem
3016: \cite{Me02}. This algorithm, called Survey Propagation, extends to
3017: the clustered phase the Belief Propagation procedure of wide-spread use in
3018: statistical inference, and is a very efficient procedure
3019: to find solutions to
3020: 3-Satisfiability slightly below threshold. Another application of
3021: statistical physics ideas is the conception of new heuristics for DPLL
3022: capable of proving the unsatisfiability of formulas with 700 hundreds
3023: variables at threshold \cite{De03}.
3024:
3025: Despite those successes important question remain open. First is there
3026: a relationship between clustering and hardness of resolution?
3027: This question is reminiscent of a very general issue in statistical physics,
3028: namely the relationship between dynamical and static properties of
3029: disordered or glassy systems \cite{Cu93}.
3030: The onset of clustering, or more
3031: precisely of strong correlations between variables over the space of solutions
3032: drastically worsens the performances of
3033: sampling algorithms e.g. Monte Carlo procedures \cite{Mo05,Se06}. However, in
3034: practical applications, one
3035: looks for a solution rather than for the sampling of
3036: the solution space... From this point of view knowing whether
3037: solutions are clustered or not
3038: does not seem to be of crucial relevance. Actually
3039: a local and polynomial search strategy capable
3040: of finding solutions well above the clustering threshold has been
3041: explicitely found for various optimizations problems \cite{Jo07}.
3042:
3043: Another open question is what happens at large
3044: $K$, that is, when constraints involve more and more variables.
3045: The performances of all known algorithms, be they local search
3046: procedures or DPLL solvers, seem to deteriorate. Worst-case bound
3047: indicate that the large $K$ case is very difficult \cite{Im99}.
3048: From statistical
3049: mechanics point of view problems look like more and more the random
3050: energy model \cite{Me87} as $K$ increases, but can we beat the
3051: worst-case bounds on average? Finally let us mention a recent work by
3052: Feige \cite{Fe02} which, for the first time, showed that the
3053: complexity of solving random SAT (or XORSAT) model had a fundamental
3054: interest in worst-case approximation theory. Consider 3-SAT
3055: instances with ratio $\alpha \gg \alpha_c$. Most of them have GS
3056: energy close to $\alpha N/2$, but a very tiny fraction of those
3057: instances have energy smaller than, say, $\epsilon N$ where $\epsilon\ll
3058: \alpha$ is fixed. Is there a polynomial algorithm capable of
3059: recognizing all such atypical formulas from the vast majority of
3060: typical instances? Insights from statistical physics suggest that,
3061: the answer is positive for SAT (if we want most satisfiable instances
3062: to be detected and not all of them) while XORSAT seems to be much
3063: harder\cite{Al07}! Actually, to the knowledge of the author, no local
3064: search algorithm (based on random walk, variable assigment, Monte
3065: Carlo, message-passing, cooling procedure, ...) is efficient for
3066: solving XORSAT. This makes the
3067: study of this problem even more valuable from a computer science point
3068: of view.
3069:
3070:
3071:
3072:
3073: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
3074:
3075: \appendix
3076:
3077: \section{A primer on large deviations}\label{applargedev}
3078:
3079: Large deviation theory is the field of probability which deals with
3080: very unlikely events \cite{De93}.
3081: You are given a fair (unbiased) coin and
3082: toss it $N$ times. The number $H$ of head draws has probability
3083: \begin{equation} \label{ld1}
3084: p_N(H) = \frac 1{2^N} {N \choose H} \ .
3085: \end{equation}
3086: When $N$ gets large $H$ is highly concentrated around $H^*=N/2$
3087: with small relative fluctuations of the order of $O(\sqrt N)$.
3088: Yet we can ask for the
3089: probability of observing a fraction $h=H/N$ equal to say, 25\%, of
3090: heads, far away from the likely value $h^*=50\%$. To calculate this
3091: probability we use Stirling's asymptotic expression for the binomial
3092: coefficient in (\ref{ld1}) to obtain
3093: \begin{equation} \label{ld2}
3094: p_N(H=h \, N) = e^{ - N \omega(h) +o(N)} \ ,
3095: \end{equation}
3096: where
3097: \begin{equation}
3098: \omega (h) = \ln 2 + h \ln h + (1-h)\ln (1-h)\
3099: \end{equation}
3100: is called rate function. The
3101: meaning of (\ref{ld2}) is that events with value of $h\ne h^*$ are
3102: exponentially rare in $N$, and $\omega(h)$ give the decay (rate) exponent.
3103: The answer to our question is $e^{-N
3104: \omega(.25)} \sim e^{-0.13 \,N}$ when $N$ is large. Some comments are:
3105: \begin{itemize}
3106: \item $\omega(h)$ is strictly positive, except in $h = h^*=\frac 12$
3107: where it vanishes. This is the only value for the fraction of head draws
3108: with non exponentially small--in--$N$ probability.
3109: \item Let $h=h^* + \delta h$ where $\delta h$ is small. Using
3110: $\omega (h^*) = \omega '(h^*) =0$ we have
3111: \begin{equation}
3112: P_N \big( H= (h^* + \delta h) N \big) = \exp \big[ -N\, \frac 12
3113: \omega ''(h^*) \,(\delta h)^2 + \ldots \big] \ ,
3114: \end{equation}
3115: that is, $\delta h$ is Gaussianly distributed with zero mean and
3116: variance $(N\omega''(h^*))^{-1}=(4N)^{-1}$. Hence central limit
3117: theorem is found back from the parabolic behaviour of the
3118: rate function around its minimum\footnote{Non standard behaviour
3119: e.g. fluctuations of the order of $N^{\nu}$ with $\nu\ne \frac 12$
3120: as found in Levy flights correspond to non-analyticies of $\omega$
3121: in $h^*$ or the vanishing of the second derivative.}.
3122: \item $\omega$ is here a convex function of its argument. This
3123: property is true rate functions describing independent
3124: events. Indeed, suppose we have $H$ positive (according to some
3125: criterion e.g. being a head for a coin) events among a set of $N$
3126: events,
3127: then another set of $N'$ events among which $H'$ are positive. If the
3128: two sets are uncorrelated
3129: \begin{equation}
3130: p_{N+N'} (H+H') \ge p_N (H) \times p_{N'} (H')
3131: \end{equation}
3132: since the same total number $H+H'$ of positive events could be
3133: observed in another
3134: combination of $N+N'$ events. Taking the logarithm
3135: and defining $h=H/N$, $h'=H'/N$, $u=N/(N+N')$ we obtain
3136: \begin{equation}
3137: \omega (u\, h+ (1-u)\, h') \le u\, \omega(h) + (1-u)\, \omega(h') \ ,
3138: \end{equation}
3139: for any $u\in [0;1]$. Hence the representative curve of $\omega$ lies
3140: below the chord joining any two points on this curve, and $\omega$ is
3141: convex. Non-convex rate functions are found in presence
3142: of strong correlations\footnote{Consider the following experiment.
3143: You are given three
3144: coins: the first one is fair (coin A), the second and third coins,
3145: respectively denoted by B and C, are biased and
3146: give head with probabilities, respectively, $\frac 14$ and $\frac
3147: 34$. First draw coin A once. If the outcome is head pick up coin B,
3148: otherwise pick up coin C. Then draw your coin $N$ times. What is the
3149: rate function associated to the fraction $h$ of heads?}.
3150: \end{itemize}
3151:
3152:
3153: \section{Inequalities of first and second moments}\label{appmoment}
3154:
3155: Let ${\cal N}$ be a random variable taking values on the positive integers, and
3156: call $p_{\cal N}$ its probability. We denote by $\langle{\cal N}\rangle$ and
3157: $\langle {\cal N} ^2\rangle$ the first and second moments of
3158: ${\cal N}$ (assumed to be finite), and write
3159: \begin{equation}
3160: p({\cal N}\ge 1) =\sum _{{\cal N}=1,2,3, \ldots} p_{\cal N} = 1 -p_0
3161: \end{equation}
3162: the probability that ${\cal N}$ is not equal to zero.
3163: Our aim is to show the inequalities
3164: \begin{equation} \label{ineq}
3165: \frac{\langle {\cal N}\rangle^2}{\langle {\cal N}^2\rangle} \le
3166: p({\cal N}\ge 1) \le {\langle {\cal N}\rangle} \ .
3167: \end{equation}
3168: The right inequality, call 'first moment inequality', is straightforward:
3169: \begin{equation}
3170: {\langle {\cal N}\rangle} = \sum _{\cal N} {\cal N}\; p_{\cal N} =
3171: \sum _{{\cal N}\ge 1} {\cal N} \; p_{\cal N} \ge
3172: \sum _{{\cal N}\ge 1} p_{\cal N} = p({\cal N}\ge 1) .
3173: \end{equation}
3174: Consider now the linear space made of vectors
3175: ${\bf v}=(v_0,v_1,v_2,\ldots\}$ whose components are labelled by positive
3176: integers, with the scalar product
3177: \begin{equation}
3178: {\bf v} \cdot {\bf v}' = \sum _{\cal N} p_{\cal N}\; v_{\cal N}\;
3179: v_{\cal N}' \ .
3180: \end{equation}
3181: Choose now $v_{\cal N}={\cal N}$, and $v'_0=0,v'_{\cal N}=1$ for
3182: ${\cal N}\ge 1$. Then
3183: \begin{equation}
3184: {\bf v} \cdot {\bf v} = \langle {\cal N}^2\rangle \ , \
3185: {\bf v} \cdot {\bf v}' = \langle {\cal N}\rangle \ , \
3186: {\bf v}' \cdot {\bf v}' = p({\cal N}\ge 1) \ .
3187: \end{equation}
3188: The left inequality in (\ref{ineq}) is simply the Cauchy-Schwarz inequality
3189: for ${\bf v},{\bf v}'$: $({\bf v} \cdot {\bf v}')^2 \le ({\bf v} \cdot {\bf v} ) \times ({\bf v}' \cdot {\bf v} ')$.
3190:
3191: \section{Corrections to the saddle-point calculation of $\langle
3192: {\cal N}^2\rangle$} \label{appfluctu}
3193:
3194: In this Appendix we show that $\langle {\cal N}^2\rangle$ is
3195: asymptotically equivalent to $\langle {\cal N}\rangle^2$, where ${\cal N}$ is the number of solutions of a 3-XORSAT formula with ratio $\alpha
3196: < \alpha_2 \simeq 0.889$. This
3197: requires to take care of the finite-size corrections around the
3198: saddle-point calculations of Section \ref{secld2}. Let $Z=(z_1,z_2,
3199: \ldots, z_N)$ denotes a configuration of variables at distance $d$
3200: from the zero configuration {\em i.e.} $dN$ variables $z_i$ are equal
3201: to 1, the other $(1-d)N$ variables are null.
3202: Let $q(d,N)$ be the probability that $Z$ satisifies
3203: the equation $z_i+z_j+z_k=0$ where $(i,j,k)$ is a random triplet of
3204: distinct integers (unbiased distribution):
3205: \begin{eqnarray}
3206: q(d,N)&=& \frac 1{{N\choose 3}} \left[ {(1-d)N \choose 3} + (1-d)N \;
3207: {dN \choose 2} \right] \\
3208: &=& q(d) \left( 1+ \frac{h(d)}N\right) +
3209: \ldots \quad \mbox{where} \quad h(d) =\frac{
3210: 6 d (2d-1)}{3 d^2+(1-d)^2}\ ,\nonumber
3211: \end{eqnarray}
3212: and $q(d)$ is defined in (\ref{qk}) with $K=3$. Terms of the order of $N^{-2}$
3213: have been discarded.
3214:
3215: Using formula (\ref{n2p}) with $q(d)$ substituted with $q(d,N)$
3216: and the Stirling formula for the asymptotic behaviour of
3217: combinatorial coefficients we have
3218: \begin{eqnarray} \label{truc1}
3219: \langle {\cal N}^2\rangle \sim \sum _{d=0,\frac 1N, \frac 2N, \ldots}
3220: \frac{\sqrt{2\pi N}}{\sqrt{2\pi Nd}\sqrt{2\pi N(1-d)}}\;
3221: e^{N\, A(d,\alpha) + \alpha\, h(d)}
3222: \end{eqnarray}
3223: where $A(d,\alpha)$ is defined in (\ref{upperom2}), and $\sim$
3224: indicates a true asymptotic equivalence (no multiplicative factor
3225: omitted). The r.h.s. of (\ref{truc1}) is the Riemann sum associated
3226: to the integral
3227: \begin{eqnarray} \label{truc2}
3228: \langle {\cal N}^2\rangle \sim \int _0^1 \frac{N \, dd}
3229: {\sqrt{2\pi Nd(1-d)}}\;
3230: e^{N\, A(d,\alpha) + \alpha\, h(d)} \ .
3231: \end{eqnarray}
3232: We now estimate the integral through the saddle-point method.
3233: For $\alpha < \alpha _2 \simeq 0.889$ the dominant contribution to the integral comes from the vicinity of
3234: $d^*=\frac 12$. There are quadratic fluctuations
3235: around this saddle-point, with a variance equal to $N$ times the
3236: inverse of (the modulus of)
3237: the second derivative $A_{dd}$ of $A$ with respect to $d$.
3238: Carrying out the Gaussian integral over those fluctuations we obtain
3239: \begin{equation} \label{truc3}
3240: \langle {\cal N}^2\rangle \sim \frac{N \;
3241: e^{N\, A(d^*,\alpha) + \alpha\, h(d^*)} }{\sqrt{2\pi N d^*(1-d^*)}}\;
3242: \sqrt{\frac{2\pi}{N\, |A_{dd}(d^*,\alpha)|}}
3243: \sim \langle {\cal N}\rangle ^2
3244: \end{equation}
3245: since $h(d^*)=0$, $A_{dd}(d^*,\alpha)=-4$. Therefore, from the
3246: second moment inequality, $P_{SAT}\to 1$ when $N\to\infty$ at ratios
3247: smaller than $\alpha_2$.
3248:
3249: %
3250: % ********** End of text entry *************
3251: %
3252: \begin{thebibliography}{99}
3253:
3254: %b1 ###
3255: \bibitem{Ac01}
3256: {{D.} {Achlioptas}},
3257: {\em Theor. Comp. Sci.} {\bf 265} (2001), {159}.
3258:
3259: \bibitem{Ac02}
3260: {{D.} {Achlioptas}},
3261: {{P.} {Beame}}, and {{M.} {Molloy}},
3262: {\em Journal of Computer and System Sciences}
3263: {\bf 68} (2004), {238}.
3264:
3265: \bibitem{Ac05}
3266: {{D.} {Achlioptas}}, {{A.} {Naor}} and
3267: {{Y.} {Perez}},
3268: {\em Nature} {\bf 435} (2005), {759}.
3269:
3270: \bibitem{Al89}
3271: {{D.J.} {Aldous}},
3272: {\em Discrete Math.} {\bf 76} (1989), {167}.
3273:
3274: \bibitem{Al01}
3275: {{D.J.} {Aldous}},
3276: {\em Rand. Struct. Algo.)} {\bf 48} (2001), {381}.
3277:
3278: \bibitem{Al02}
3279: {{M.} {Alekhnovich}} and {{E.} {Ben-Sasson}},
3280: \textit{Analysis of the Random Walk Algorithm on Random 3-CNFs},
3281: preprint (2002).
3282:
3283: \bibitem{Al07}
3284: {{F.} {Altarelli}}, {{R.} {Monasson}}
3285: and {{F.} {Zamponi}},
3286: {\em J. Phys. A} {\bf 40} (2007), {867}.
3287:
3288: \bibitem{Am89}
3289: {{D.J.} {Amit}},
3290: \textit{Modeling Brain Function},
3291: (Cambridge University Press, Cambridge, 1989).
3292:
3293: \bibitem{De79}
3294: {{J.R.L.} {de Almeida}} and {{D.J.}
3295: {Thouless}},
3296: {\em J. Phys. A} {\bf 11} (1978), {983}.
3297:
3298: \bibitem{beame}
3299: {{P.} {Beame}}, {{R.} {Karp}},
3300: {{T.} {Pitassi}}, and {{M.} {Saks}},
3301: (Proceedings of the ACM Symp.\ on Theory of Computing, 1998, pp. 561).
3302:
3303: \bibitem{Bi00}
3304: {{G.} {Biroli}}, {{R.} {Monasson}} and
3305: {{M.} {Weigt}},
3306: {\em Eur. Phys. J. B} {\bf 14} (2000), {551}.
3307:
3308: \bibitem{Bo89}
3309: {{B.} {Bollobas}},
3310: \textit{Random Graphs}
3311: (Cambridge University Press, Cambridge, 2001).
3312:
3313: \bibitem{Br93}
3314: {{A.Z.} {Broder}}, {{A.M.} {Frieze}}
3315: and {{E.} {Upfal}},
3316: (Proceedings of Symposium of Discrete Algorithms (SODA), Austin,
3317: 1993).
3318:
3319: \bibitem{Ch90}
3320: {{M.T.} {Chao}} and {{J.} {Franco}},
3321: {\em Information Science} {\bf 51} (1990),
3322: {289}; {\em SIAM Journal on Computing}
3323: {\bf 15} (1986), {1106}.
3324:
3325: \bibitem{Ch88}
3326: {{V.} {Chv{\`a}tal}} and {{E.} {Szmeredi}},
3327: {\em Journal of the ACM} {\bf 35} (1988), {759}.
3328:
3329: \bibitem{Co01}
3330: {{S.} {Cocco}} and {{R.} {Monasson}},
3331: {\em Phys. Rev. Lett.} {\bf 86} (2001), {1658}; {\em Eur. Phys. J. B} {\bf 22}
3332: (2002) , {505}.
3333:
3334: \bibitem{Co03}
3335: {{S.} {Cocco}}, {{O.}
3336: {Dubois}}, {{J.} {Mandler}} and {{R.}
3337: {Monasson}},
3338: {\em Phys. Rev. Lett.} {\bf 90} (2003), {047205}.
3339:
3340: \bibitem{Cr99}
3341: {{N.} {Creignou}} and {{H.} {Daud\'e}},
3342: {\em Discrete Applied Mathematics} {\bf 96-97} (1999), {41}.
3343:
3344: \bibitem{Cr03a}
3345: {{N.} {Creignou}} and {{H.} {Daud\'e}},
3346: {\em RAIRO: Theoretical Informatics and Applications}
3347: {\bf 37} (2003), {127}.
3348:
3349: \bibitem{Cr03b}
3350: {{N.} {Creignou}}, {{H.} {Daud\'e}}
3351: and {{O.} {Dubois}},
3352: {\em Combinatorics, Probability and Computing}
3353: {\bf 12} (2003), {113}.
3354:
3355: \bibitem{Cu93}
3356: {{L.} {Cugliandolo}} and {{J.} {Kurchan}},
3357: {\em Phys. Rev. Lett.}
3358: {\bf 71} (1993), {173}.
3359:
3360: \bibitem{dpll}
3361: {{M.} {Davis}} and {{H.} {Putnam}},
3362: {\em J.\ Assoc.\ Comput.\ Mach.} {\bf 7} (1960), {201};
3363: {{M.} {Davis}}, {{G.} {Logemann}}
3364: and {{D.} {Loveland}},
3365: {\em Communications of the ACM} {\bf 5} (1962), {394}.
3366:
3367:
3368: \bibitem{De93}
3369: {{A.} {Dembo}} and {{O.} {Zeitouni}},
3370: \textit{Large deviations techniques and applications}
3371: (Springer-Verlag, New York, 1993).
3372:
3373: \bibitem{De03}
3374: {{G.} {Dequen}} and {{O.} {Dubois}},
3375: (Proceedings of Theory and Applications of Satisfiability Testing, 6th
3376: International Conference, SAT 2003. Santa Margherita Ligure, 2003, pp 486).
3377:
3378:
3379: \bibitem{De04}
3380: {{C.} {Deroulers}} and {{R.} {Monasson}},
3381: {\em Eur. Phys. J. B} {\bf 49} (2006), {339}.
3382:
3383: \bibitem{Du02}
3384: {{O.} {Dubois}} and {{J.} {Mandler}},
3385: (Proc. of the 43rd
3386: annual IEEE symposium on Foundations of Computer Science, Vancouver,
3387: 2002).
3388:
3389: \bibitem{Va01}
3390: {{A.} {Engel}} and
3391: {{C.} {Van den Broeck}},
3392: \textit{Statistical Mechanics of Learning},
3393: (Cambridge University Press, Cambridge, 2001).
3394:
3395: \bibitem{Fe02}
3396: {{U.} {Feige}},
3397: (Proceedings of 34th STOC conference, 2002, pp 534).
3398:
3399: \bibitem{Fr96}
3400: {{A.} {Frieze}} and {{S.} {Suen}},
3401: {\em Journal of Algorithms} {\bf 20} (1996), {312}.
3402:
3403: \bibitem{Fu85}
3404: {{Y.} {Fu}} and {{P.W.} {Anderson}},
3405: {\em J. Phys. A} {\bf 19} (1986), {1605}.
3406:
3407: \bibitem{We01}
3408: {{A.K.} {Hartmann}} and {{M.} {Weigt}},
3409: {\em Theor. Comp. Sci.} {\bf 265} (2001), {199}.
3410:
3411: \bibitem{Im99}
3412: {{R.} {Impagliazzo}} and {{R.} {Paturi}},
3413: (Proceedings of the
3414: {IEEE} Conference on Computational Complexity, 1999, pp 237).
3415:
3416: \bibitem{Ka02}
3417: {{A.C.} {Kaporis}}, {{L.M..}
3418: {Kirousis}}, and {{Y.C..} {Stamatiou}},
3419: \textit{How to prove conditional randomness using the principle of deferred
3420: decisions}, technical report, Computer technology Institute, Patras
3421: (2002).
3422:
3423: \bibitem{Kn66}
3424: {{D.} {Knuth}},
3425: \textit{The Art of Computer Programming; vol 1: fundamental
3426: algorithms, section 2.2.1},
3427: (Addison-Wesley, Ney York, 1968).
3428:
3429: \bibitem{Kr89}
3430: {{W.} {Krauth}} and
3431: {{M.} {M\'ezard}},
3432: {\em J. Phys. (France) } {\bf 50} (1989), {3057}.
3433:
3434: \bibitem{Se06}
3435: {{F.} {Krzakala}} {\em at al.}
3436: \textit{Gibbs States and the Set of Solutions of Random Constraint
3437: Satisfaction Problems} (preprint, 2006).
3438:
3439: \bibitem{Jo07}
3440: {{F.} {Krzakala}} and {{J.} {Kurchan}},
3441: \textit{(A landscape analysis of constraint satisfaction problems}
3442: (preprint, 2007).
3443:
3444: \bibitem{Me98}
3445: {{S.} {Mertens}},
3446: {\em Phys. Rev. Lett.} {\bf 81} (1998), {4281};
3447: {\em Phys. Rev. Lett.} {\bf 84} (2000), {1347}.
3448:
3449: \bibitem{Me01}
3450: {{S.} {Mertens}},
3451: {\em Theor. Comp. Sci.} {\bf 265} (2001), {79}.
3452:
3453: \bibitem{Me03b}
3454: {{M.} {Mertens}},
3455: {{M.} {M\'ezard}} and {{R} {Zecchina}},
3456: {\em Rand. Struct. Algo.} {\bf 28} (2006), {340}.
3457:
3458: \bibitem{Me87b}
3459: {{M.} {M\'ezard}} and {{G.} {Parisi}},
3460: {\em J. Phys. (Paris)} {\bf 48} (1987), {1451}.
3461:
3462: \bibitem{Me86}
3463: {{M.} {M\'ezard}} and {{G.} {Parisi}},
3464: {\em J. Phys. (Paris)} {\bf 47} (1986), {1285}.
3465:
3466: \bibitem{Me87}
3467: {{M.} {M\'ezard}}, {{G.} {Parisi}} and
3468: {{M.} {Virasoro}},
3469: \textit{Spin glasses and beyond}
3470: (World Scientific, Singapore, 1987).
3471:
3472: \bibitem{Me01y}
3473: {{M.} {M\'ezard}} and {{G.} {Parisi}},
3474: {\em Eur. Phys. J. B}
3475: {\bf 20} (2001), {217};
3476: {\em J. Stat. Phys}
3477: {\bf 111} (2003), {111}.
3478:
3479:
3480: \bibitem{Me02}
3481: {{M.} {M\'ezard}} and {{R} {Zecchina}},
3482: {\em Phys. Rev. E} {\bf 56} (2002), {066126}.
3483:
3484: \bibitem{Me03}
3485: {{M.} {M\'ezard}}, {{F.}
3486: {Ricci-Tersenghi}}, and {{R.} {Zecchina}},
3487: {\em J. Stat. Phys.} {\bf 111} (2003), {505}.
3488:
3489: \bibitem{Mi92}
3490: {{D.} {Mitchell}}, {{B.} {Selmann}}
3491: and {{H.} {Levesque}},
3492: {\em Proc.\ of the Tenth Natl.\ Conf.\ on Artificial
3493: Intelligence (AAAI-92)} (1992) , {440}.
3494:
3495: \bibitem{Mon95}
3496: {{R.} {Monasson}},
3497: {\em Phys. Rev. Lett.}
3498: {\bf 75} (1995), {2847}.
3499:
3500: \bibitem{Mo97}
3501: {{R.} {Monasson}} and {{R.} {Zecchina}},
3502: {\em Phys. Rev. E} {\bf 56} (1997), {1357}.
3503:
3504: \bibitem{Mon98}
3505: {{R.} {Monasson}},
3506: {\em J. Phys. A} {\bf 31} (1998), {513}.
3507:
3508: \bibitem{Mon05}
3509: {{R.} {Monasson}},
3510: {\em Lecture Notes in Computer Science} {\bf 3624} (2005), {402}.
3511:
3512: \bibitem{Mo07}
3513: {{A.} {Montanari}} and {{D.} {Shah}},
3514: (Proceedings of Symposium of Discrete Algorithms (SODA), New Orleans,
3515: 2007).
3516:
3517: \bibitem{Mo05}
3518: {{A.} {Montanari}} and {{G.} {Semerjian}},
3519: {\em Phys. Rev. Lett.}
3520: {\bf 94} (2005), {247201}.
3521:
3522:
3523: \bibitem{Mo95}
3524: {{R.} {Motwani}} and {{P.} {Raghavan}},
3525: \textit{Randomized algorithms}
3526: (Cambridge University Press, Cambridge, 1995).
3527:
3528: \bibitem{Mu02}
3529: {{R.} {Mulet}}, {{A.} {Pagnani}},
3530: {{M.} {Weigt}} and {{R} {Zecchina}},
3531: {\em Phys. Rev. Lett.} {\bf 89} (2002), {268701}.
3532:
3533: \bibitem{Or85}
3534: {{H.} {Orland}},
3535: {\em J. Phys. (Paris) Lett.} {\bf 46} (1985), {L763};
3536: {{M.} {M\'ezard}} and {{G.} {Parisi}},
3537: {\em J. Phys. (Paris) Lett.} {\bf 46} (1985), {L771}.
3538:
3539: \bibitem{Pa83}
3540: {{C.} {Papadimitriou}} and {{K.}
3541: {Steiglitz}},
3542: \textit{Combinatorial Optimization: Algorithms and Complexity}
3543: (Dover, 1998).
3544:
3545: \bibitem{Pa92}
3546: {{C.H.} {Papadimitriou}},
3547: (Proceedings of the 32nd Annual IEEE Symposium on Foundations of
3548: Computer Science, 1991, pp 163).
3549:
3550: \bibitem{Pa86}
3551: {{G.} {Parisi}} and {{M.} {Virasoro}},
3552: {\em J. Phys. (Paris)} {\bf 50} (1986), {3317}.
3553:
3554: \bibitem{We02}
3555: {{F.} {Ricci-Tersenghi}},
3556: {{M.} {Weigt}} and {{R.} {Zecchina}},
3557: {\em Phys. Rev. E} {\bf 63} (1999), {026702}.
3558:
3559: \bibitem{Sc00}
3560: {{U.} {Sch\"oning}},
3561: {\em Algorithmica} {\bf 32} (2002), {615}.
3562:
3563: \bibitem{Se03}
3564: {{G.} {Semerjian}} and {{R.} {Monasson}},
3565: {\em Phys. Rev. E} {\bf 67} (2003), {066103};
3566: {{W.} {Barthel}}, {{A.} {Hartmann}}
3567: and {{M.} {Weigt}},
3568: {\em Phys. Rev. E} {\bf 67} (2003), {066104}.
3569:
3570: \bibitem{Wo95}
3571: {{N.} {Wormald}},
3572: {\em The Annals of Applied Probability} {\bf 5} (1995), {1217}.
3573:
3574:
3575: \bibitem{We01b}
3576: {{M.} {Weigt}} and {{A.K.} {Hartmann}}
3577: {\em Phys. Rev. Lett.} {\bf 84} (2000), {6118};
3578: {\em Phys. Rev. Lett.} {\bf 86} (2001), {1658}.
3579:
3580: \end{thebibliography}
3581: \end{document}
3582:
3583: