1: \documentclass[11pt]{article}
2: %\usepackage{latexsym,amsmath,amsfonts}
3:
4: \usepackage{times}
5: %\usepackage{palatino}
6:
7: \usepackage{amssymb}
8:
9: \usepackage{epsfig}
10: %\usepackage{psfig}
11: \setlength{\textheight}{9.2 in}
12: \setlength{\textwidth}{6.55 in}
13: \setlength{\oddsidemargin}{0 in}
14: \setlength{\evensidemargin}{0 in}
15: \setlength{\topmargin}{-0.5 in}
16: \setlength{\parskip}{2pt}
17:
18: \newcommand{\tinyspacing}{\baselineskip = 0.7\normalbaselineskip}
19: \newcommand{\smallspacing}{\baselineskip = \normalbaselineskip}
20: \newcommand{\morespacing}{\baselineskip = 1.1\normalbaselineskip}
21: \newcommand{\newspacing}{\baselineskip=1.04\normalbaselineskip}
22: \newcommand{\oldspacing}{\baselineskip= 1.0\normalbaselineskip}
23:
24:
25: \renewcommand*{\Pr}{\mathop{\mathrm{Prob}}}
26:
27:
28:
29: \def \qedbox{\hfill\vbox{\hrule\hbox{\vrule
30: height1.3ex\hskip0.8ex\vrule}\hrule}}
31:
32:
33:
34: \newcommand{\goesto}{\rightarrow}
35: \newtheorem{theorem}{Theorem}
36: \newtheorem{definition}{Definition}
37: \newtheorem{proposition}{Proposition}
38: %\newtheorem{lemma}{Lemma}
39: \newcommand*{\proof}{\noindent {\bf Proof}.\,\,}
40:
41:
42:
43: \def\NP{\rm NP}
44: \def\AND{\wedge}
45: \def\OR{\vee}
46: \def\oper{\circ}
47: \def\goesto{\rightarrow}
48: \def\implies{\Rightarrow}
49: \def\zeroone{\{0,1\}}
50: \def\sstar{\zeroone^{*}}
51: \def\L{\langle}
52: \def\R{\rangle}
53: \def\HYP{\hbox{-}}
54: \def\IFF{\leftrightarrow}
55: \def\Ldef{\buildrel \rm def \over \leftrightarrow}
56: \def\Edef{\buildrel \rm def \over =}
57: \def\almostall{\hbox{\rlap{$_{\thinspace\forall}$}{$^{^\infty}$}}}
58: \def\infoften{\hbox{\rlap{$_{\thinspace\exists}$}{$^{^\infty}$}}}
59: \def\N{{\bf N}}
60: \def\setdelta{\bigtriangleup}
61: \def\cminus{\dot{-}}
62: \def\plusminus{\pm}
63: \def\PR{{\rm Pr}}
64: \def\HSAT{{\rm HORN}\hbox{-}{\rm SAT}}
65: \def\PUR{{\rm PUR}}
66: \def\beginproof{\noindent{\bf Proof.}\quad}
67: \def\endproof{}
68:
69:
70:
71:
72:
73: \newtheorem{lemma}{Lemma}[section]
74: \newtheorem{theo}[lemma]{Theorem}
75: \newtheorem{cor}{Corollary}
76: \newtheorem{prop}[lemma]{Proposition}
77: %\newtheorem{fact}[lemma]{Fact}
78: \newtheorem{example}{Example}
79: \newtheorem{obs}{Observation}
80: \newtheorem{claim}{Claim}
81: \newtheorem{defi}{Definition}
82: %\newtheorem{obs}[lemma]{Observation}
83: \def\qed{\hfill$\Box$\newline\vspace{5mm}}
84: \newenvironment{PROOF}{\noindent{\bf Proof:}}{{\qed}}
85: \newtheorem{conj}{Conjecture}
86:
87: \bibliographystyle{unsrt}
88:
89: %\setpapersize{USletter}
90: %\setmarginsrb{1in}{1in}{1in}{1in}{0pt}{0mm}{0pt}{0mm}
91:
92: \begin{document}
93:
94: \pagestyle{empty}
95:
96: \title{Phase transitions and {\em all that}}
97: \author{Gabriel Istrate\thanks{e-mail: istrate@lanl.gov,
98: NISAC, National Infrastructure Simulation Analysis Center,
99: Los Alamos National Laboratory,
100: Mail Stop M 997, Los Alamos, NM 87545, U.S.A.}}
101: \date{}
102:
103: \maketitle
104: \thispagestyle{empty}
105: \section{Introduction}
106:
107: Since the experimental paper of Cheeseman, Kanefsky and Taylor
108: \cite{cheeseman-kanefsky-taylor}
109: {\em phase transitions in combinatorial problems} held the promise to
110: shed
111: light on the ``practical'' algorithmic complexity of combinatorial
112: problems. However, the connection conjectured in
113: \cite{cheeseman-kanefsky-taylor} was easily seen to be inaccurate. A much more realistic
114: possible connection has been highlighted by the results (based on
115: experimental evidence and nonrigorous arguments from Statistical Mechanics)
116: of Monasson et
117: al. \cite{2+p:nature} (see also \cite{2+p:rsa}). These results
118: supported
119: the conjecture that it is {\em first-order phase transitions} that
120: have algorithmic implications for the complexity of restricted
121: classes of algorithms, including the important class of
122: {\em Davis-Putnam-Longman-Loveland (DPLL) algorithms} \cite{beame:dp}.
123:
124: There exists, indeed, a nonrigorous argument supporting this
125: conjecture:
126: phase transitions amount to nonanalytical behavior of a certain
127: {\em order parameter}; the phase transition is {\em first order} if the
128: order parameter is actually
129: discontinuous. At least for random $k$-SAT \cite{monasson:zecchina} the
130: order parameter suggested by Statistical Mechanics considerations has a
131: purely combinatorial interpretation: it is the {\em backbone} of the
132: formula, the set of
133: literals that assume the same value in all optimal assignments.
134: But intuitively one can relate (see e.g.
135: the presentation of this argument by Achlioptas, Beame and Molloy
136: \cite{achlioptas:beame:molloy:slides})
137: the size of the backbone to the complexity of DPLL algorithms,
138: when run on random $k$-SAT instances slightly above the phase
139: transition: All literals in the
140: backbone require well-defined values in order to satisfy the
141: formula. But a DPLL algorithm has very few ways to know what those
142: ``right'' values are. If w.h.p. the backbone of formulas above the
143: transition contains a positive
144: fraction of the literals that is bounded away from zero as we approach
145: the transition (which happens in a case of a first-order phase
146: transition) then,
147: intuitively, DPLL will misassign a variable having $\Omega(n)$ height
148: in the
149: tree representing the behavior of the algorithm, and
150: will be forced to backtrack on the given variable.
151: {\em The conclusion of this intuitive argument is that a
152: first-order phase transition implies a $2^{\Omega(n)}$ lower bound for
153: the
154: running time of any DPLL algorithm, valid with high probability for
155: random instances located slightly above the transition. }
156:
157: While previous rigorous results
158: \cite{2+psat:ralcom97,scaling:window:2sat,achlioptas:beame:molloy}, supported these intuitions, to date, the
159: extent of a connection between first-order phase transitions and
160: algorithmic complexity was unclear.
161:
162: {\bf The goals of this paper are
163: \begin{enumerate}
164: \item To remedy this, and formally establish a connection between
165: first-order phase
166: transitions and the resolution complexity of random satisfiability
167: problems, and
168: \item To take steps towards obtaining a complete classification of the
169: order of phase transition in generalized satisfiability problems.
170: \end{enumerate}
171: }
172:
173: To accomplish these goals
174:
175: \begin{enumerate}
176: \item we obtain (Theorem~\ref{dichotomy:threshold}) a complete
177: characterization of sharp/coarse thresholds in the random generalized
178: satisfiability model due to Molloy \cite{molloy-stoc2002}.
179: ``Phyisical'' arguments (see discussion below) imply that it makes no
180: sense to study the order of the phase transition unless the problem
181: has a sharp threshold.
182: \item we rigorously prove (Theorem~\ref{3sat:first-order}) that random
183: 3-SAT has a first-order phase transition. We extend this result in
184: several ways: first (Theorem~\ref{2+p-sat:first-order}) to
185: random $(2+p)$-satisfiability, the original problem from
186: \cite{2+p:nature}, obtaining further theoretical support to the
187: heuristic results of \cite{2+p:nature}. Second we give a sufficient
188: condition (Theorem~\ref{sufficient:first-order}) for the existence of
189: a first-order phase transition. We then show
190: (Theorem~\ref{implicates:first-order}) that all problems whose
191: constraints have no implicates of size at most two satisfy this
192: condition.\item we show that in all the cases where
193: we can prove the
194: existence of a first-order phase transition, such problems have
195: a $2^{\Omega(n)}$ lower bound on their resolution complexity
196: (and hence the complexity of DPLL algorithms as well \cite{beame:dp}).
197: Indeed, the two phenomena ($2^{\Omega(n)}$ resolution complexity and
198: the existence of a first-order phase transition) have
199: common causes.
200: \item in contrast, we show (Theorem~\ref{second:order}) that, for {\em
201: any generalized
202: satisfiability problem}, a second-order phase transition implies, for
203: every $\alpha >0$, a $O(2^{\alpha \cdot n})$ upper
204: bound on the resolution complexity of their random instances (in the
205: region where most formulas are unsatisfiable).
206: \end{enumerate}
207:
208: \section{Preliminaries}
209:
210: Throughout the paper we will assume familiarity with the general
211: concepts of phase transitions in combinatorial problems (see e.g.
212: \cite{martin:monasson:zecchina}),
213: random structures \cite{bol:b:random-graphs}, proof complexity
214: \cite{beame:proof:survey}. Some papers whose concepts and methods we use in
215: detail (and we assume greater familiarity with)
216: include \cite{friedgut:k:sat}, \cite{chvatal:szemeredi:resolution},
217: \cite{ben-sasson:resolution:width}.
218:
219: Consider a monotonically increasing problem $A=(A_{n})$,
220: under the constant probability model $\Gamma(n,p)$, that independently
221: sets to 1 with probability $p$
222: each bit of the random string. As usual, for
223: $\epsilon >0$ let $p_{\epsilon}= p_{\epsilon}(n)$ define the canonical
224: probability such that $\Pr_{x \in \Gamma(n,p_{\epsilon}(n))}[x \in
225: A]= \epsilon$.
226:
227: The probability that a random sample $x$ satisfies property $A$ (i.e.
228: $x\in A$) is a monotonically increasing function of $p$. {\em Sharp
229: thresholds} are those for which this function has a ``sudden jump'' from
230: value 0
231: to 1:
232:
233: \begin{definition} \label{sharp}
234: Problem $A$ has a {\em sharp threshold} iff for every $0<\epsilon <
235: 1/2$, we have $\lim_{n\goesto \infty} \frac{p_{1-\epsilon}(n)-
236: p_{\epsilon}(n)}{p_{1/2}(n)} = 0$.
237: $A$ has {\em a coarse threshold} if for some $\epsilon > 0$ it holds
238: that
239: $\underline{\lim}_{n\goesto \infty} \frac{p_{1-\epsilon}(n)-
240: p_{\epsilon}(n)}{p_{1/2}(n)} > 0$.
241: \end{definition}
242:
243: For satisfiability problems (whose complements are monotonically
244: increasing) the constant probability model amounts to adding every constraint
245: (among those
246: allowed by the syntactic specification of the model) to the random
247: formula independently
248: with probability $p$. Related definitions can be given for the other
249: two models for
250: generating random structures,
251: the {\em counting model} and {\em the multiset model}
252: \cite{bol:b:random-graphs}. Under reasonable
253: conditions \cite{bol:b:random-graphs}
254: these models are equivalent, and we will liberally switch
255: between them. In particular, for
256: satisfiability problem $A$, and an instance $\Phi$ of $A$,
257: $c_{A}(\Phi)$ will denote its {\em constraint density}, the ratio
258: between
259: the number of clauses and the number of variables of $\Phi$. To specify
260: the random model in this
261: latter cases we have to specify the constraint density as a function of
262: $n$, the number of variables.
263: We will use $c_{A}$ to denote the value of the constraint density
264: $c_{A}(\Phi)$ (in the counting/multiset
265: models) corresponding to taking $p=p_{1/2}$ in the constant probability
266: model.
267: $c_{A}$ is a function on $n$ that is believed to tend to a constant
268: limit as $n\goesto \infty$. However, Friedgut's proof
269: \cite{friedgut:k:sat} of a sharp threshold in $k$-SAT (and our results) leave this issue
270: open.
271:
272:
273: The original investigation of the order of the phase transition in
274: $k$-SAT used an order parameter called {\em the backbone}.
275: Bollob\'{a}s et al. \cite{scaling:window:2sat} have investigated the
276: order of the phase transition in 2-SAT under a different order parameter,
277: a ``monotonic version'' of the
278: backbone called {\em the spine}.
279:
280: \begin{equation}\label{spine:initial}
281: Spine(\Phi) = \{ x\in Lit | (\exists) \Xi \subseteq \Phi, \Xi \in SAT,
282: \Xi \AND \{\overline{x}\}\in \overline{SAT}\}.
283: \end{equation}
284:
285: They showed that random 2-SAT has a continuous (second-order) phase
286: transition: the size of the spine, normalized by dividing it by the
287: number of
288: variables, approaches zero (
289: as $n\goesto \infty$) for $c<c_{2-SAT}=1$, and is continuous
290: at $c=c_{2-SAT}$. By contrast, nonrigorous
291: arguments from Statistical Mechanics \cite{monasson:zecchina} imply the
292: fact
293: that for $3-SAT$ the spine jumps discontinuously from zero to
294: positive values at the transition point $c=c_{3-SAT}$
295: (a first-order phase transition).
296:
297: It is easy to see that the intuition concerning the connection between
298: the
299: complexity of DPLL algorithms and the size of the backbone (discussed
300: briefly in the introduction) extends to the spine as well. In this
301: paper
302: whenever we will discuss the order of a phase transition we will do it
303: with respect to this latter order parameter.
304:
305: We would like to obtain a complete classification of
306: the order of the phase transition in random satisfiability
307: problems.
308: A preliminary problem we have to deal with is characterizing those
309: problems that
310: have a sharp threshold: indeed, Physics considerations require that,
311: in order that the study of the (order of the) phase transition to be
312: meaningful, the order parameter (in the case of $k$-SAT the spine) has
313: to be, w.h.p., concentrated around its expected value
314: (in Physics parlance it is a {\em self averaging quantity}), and it is
315: zero up to a certain value of the control parameter
316: (in our case constraint density $c$) and positive above it.
317: In the case of $k$-SAT these conditions imply the fact that
318: $k$-SAT has a sharp threshold. The argument (a ``folklore'' one)
319: can be formally
320: expressed by the following
321:
322: \begin{lemma}\label{spine-threshold} Let $c$ be an arbitrary {\em
323: constant} value for the constraint
324: density function.
325: \begin{enumerate}
326: \item If $c< \underline{lim}_{n\goesto \infty} c_{k-SAT}(n)$ then
327: $\lim_{n \goesto \infty} \frac{|Spine(\Phi)|}{n} =0$.
328: \item If for some $c$ there exists $\delta > 0$ such that w.h.p. (as
329: $n\goesto \infty$) $\frac{|Spine(\Phi)|}{n} > \delta$ then
330: $\lim_{n\goesto \infty}
331: \Pr[\Phi \in SAT]= 0$, that is $c> \overline{lim}_{n\goesto \infty}
332: c_{k-SAT}(n)$.
333: \end{enumerate}
334: \end{lemma}
335:
336: The argument is generic enough to extend to {\em all} constraint
337: satisfaction problems. So a necessary condition for the study of
338: the phase transition to be meaningful is that the problem have a
339: sharp threshold.
340:
341: \section{Coarse and sharp thresholds of random generalized
342: satisfiability
343: problems}
344:
345:
346: In this section we obtain a complete classification of thresholds of
347: random satisfiability problems, under
348: Molloy's recent model of random constraint satisfaction problems from
349: \cite{molloy-stoc2002} (specialized to satisfiability problems, i.e.
350: problems with domain $\{0,1\}$). This affirmatively solves
351: an open problem raised in \cite{creignou:daude:sat2002}.
352:
353:
354: \begin{definition}\label{model}
355: Consider the set of all $2^{2^{k}}-1$ potential nonempty binary
356: constraints on $k$ variables
357: $X_{1}, \ldots, X_{k}$. We specify a probability distribution ${\cal
358: P}$ which selects a single
359: random constraint, and let ${\cal C}= supp({\cal P})$ be the set of
360: constraints on which ${\cal P}$
361: assigns positive probability.
362:
363: A random formula from $SAT_{n,M}({\cal P})$ is specified by the
364: following procedure:
365: \begin{itemize}
366: \item $n$ is the number of variables.
367: \item $M$ is the number of clauses, chosen by the following procedure:
368: first select, uniformly at
369: random and with repetition $m$ hyperedges of the uniform hypergraph on
370: $n$ variables.
371: \item for each hyperedge choose a random ordering of the variables
372: involved. Choose a random
373: constraint according to ${\cal P}$ and apply it on the list of
374: (ordered) variables.
375: \end{itemize}
376: $SAT({\cal C})$ refers to the random model corresponding to ${\cal P}$
377: being the uniform distribution on ${\cal C}$.
378: \end{definition}
379:
380: It turns out we face a technical difficulty when studying sharp and
381: coarse thresholds in Molloy's model; it cannot be directly mapped onto
382: the
383: constant probability model for which the notion of a sharp threshold in
384: Definition~\ref{sharp} works. The definition of a sharp threshold we
385: need
386: to employ is the one from \cite{molloy-stoc2002}
387:
388: \begin{definition}\label{sharp:2}
389: $SAT({\cal P})$ is said to have {\em a sharp threshold of
390: satisfiability} if there exists a function $c(n)$ bounded away from 0 such that, for
391: any $\epsilon >0$ if $M<(c(n)-\epsilon) n$
392: then $SAT_{n,M}({\cal P})$ is a.s. satisfiable and if
393: $M>(c(n)+\epsilon) n$ then $SAT_{n,M}({\cal P})$ is a.s. unsatisfiable. On the other
394: hand, if there exist two functions $M_{1}(n), M_{2}(n)$
395: such that $M_{1}(n)/M_{2}(n)$ is bounded away from zero, and the
396: satisfaction probability of random instances from $SAT_{n,M_{1}}({\cal P})$,
397: $SAT_{n,M_{2}}({\cal P})$ is bounded away from
398: both 0 and 1 then $SAT({\cal P})$ is said to have {\em a coarse
399: threshold.}
400: \end{definition}
401:
402: However, just as in \cite{molloy-stoc2002} (where this was done in the
403: case when ${\cal P}$ is the uniform distribution), for $k\geq 3$
404: one can map Molloy's model onto a modified version of the constant
405: probability model, defined as follows:
406: Let $p_{1}, \ldots p_{r}$ be positive numbers between zero and 1. A
407: random sample $x$ from the model $\Gamma_{p_{1},\ldots p_{r}}(n,p)$ is
408: obtained in the
409: following way: divide the bits of $x$ into $r$ equal groups. Set each
410: of the bits in the $i$'th group to 1 independently with probability
411: $p\cdot p_{i}$. For this model the definitions of $p_{\epsilon}$ and sharp/coarse threshold from Definition~\ref{sharp} carry over, and are equivalent to
412: those from Definition~\ref{sharp:2}.
413:
414: Indeed, let $r$ be the cardinality of the support of distribution ${\cal P}$,
415: and $p_{1}, \ldots, p_{r}$ be the associated positive
416: probabilities.
417:
418: In its general setting Molloy's model is specified as follows: divide
419: the potential constraints into
420: groups of $rk!$ constraints, corresponding to all possible applications
421: of the $r$ constraint templates on a fixed set of $k$ variables.
422: For each such group, independently with probability $p$, we make the
423: decision to include at least one of the constraints in the group
424: with probability $p=\frac{M}{rk!{{n}\choose {k}}}$ (going from $M$
425: clauses to including each potential edge independently with probability $p$
426: can be done just as in the uniform case from \cite{molloy-stoc2002}).
427:
428: Each realization of constraint constraint template $i$ is chosen with
429: probability probability $p_{i}/k!$. Denote this model by $M(n,p,p_{1},
430: \ldots, p_{r})$.
431:
432: Defining $f(x) = [(1+x pp_{1}/k!)\cdot (1+xpp_{2}/k!) \cdot \ldots
433: \cdot (1+xpp_{r})/k!]^{k!}- x$ we have $f(1)>0 $ and, since (by a simple
434: calculus argument)
435: the minimum of $f(x)$ over the choices of $p_{i}\geq 0$, $\sum p_{i}=1$
436: is obtained when one of them
437: is 1 and the others are zero,
438: \[
439: f(\frac{1}{1-p})\geq [1+\frac{p}{k!(1-p)}]^{k!}-\frac{1}{1-p}\geq 0.
440: \]
441:
442:
443: Let $\alpha= \alpha(n) >0$ be the smallest solution of the equation
444: $f(\alpha) =0$. Thus $\frac{1}{1-p} \leq \alpha$.
445:
446:
447:
448: Define, for $i=1,r$,
449: \[
450: p^{\prime}_{i}= \frac{1/k! \cdot \alpha \cdot p_{i}}{1+1/k! \cdot
451: \alpha pp_{i}}
452: \]
453:
454:
455:
456: \begin{claim}
457: The following hold for any $p=\theta(n^{1-k})$:
458: \begin{enumerate}
459: \item For every formula $\Phi$ such that no two constraints on the same
460: set of variables appear in it,
461: \[
462: \Pr_{M(n,p,p_{1}, \ldots, p_{r})}(\Phi)\geq
463: \Pr_{\Gamma_{p^{\prime}_{1}, \ldots, p^{\prime}_{r}}(n,p)}[\Phi].
464: \]
465: Consequently
466: \begin{eqnarray*}
467: \Pr_{M(n,p,p_{1}, \ldots, p_{r})}[SAT({\cal P})]\geq
468: \Pr_{\Gamma_{p^{\prime}_{1}, \ldots, p^{\prime}_{r}}(n,p)}[SAT({\cal P})| \\
469: \mbox{ no two constraints on the same set of variables appear in }\Phi
470: ].
471: \end{eqnarray*}
472: \item On the other hand, there exists $f(n) = 1+o(1)$ such that for
473: every formula $\Phi$ such that no two constraints on the same set of
474: variables appear in it,
475: \[
476: \Pr_{M(n,p,p_{1}, \ldots, p_{r})}(\Phi)\leq f(n)
477: \Pr_{\Gamma_{p^{\prime}_{1}, \ldots, p^{\prime}_{r}}(n,p)}[\Phi].
478: \]
479: Consequently
480: \begin{eqnarray*}
481: \Pr_{M(n,p,p_{1}, \ldots, p_{r})}[SAT({\cal P})]\leq
482: (1+o(1))\Pr_{\Gamma_{p^{\prime}_{1}, \ldots, p^{\prime}_{r}}(n,p)}[SAT({\cal P})| \\
483: \mbox{ no two constraints on the same set of variables appear in }\Phi
484: ].
485: \end{eqnarray*}
486: \end{enumerate}
487: \end{claim}
488:
489: Indeed, consider the set of constraints on a fixed set of given
490: variables. The probability (under $\Gamma_{p^{\prime}_{1}, \ldots,
491: p^{\prime}_{r}}$) that a given clause of type $i$ is
492: included, and none of the others are is equal to
493: $\frac{pp^{\prime}_{i}}{1- pp^{\prime}_{i}}\cdot
494: [(1-pp^{\prime}_{1})\ldots (1-pp^{\prime}_{r})]^{k!} $.
495: But
496: \[
497: \frac{pp^{\prime}_{i}}{1-pp^{\prime}_{i}}= \alpha pp_{i}/k!.
498: \]
499:
500: Also
501: \[
502: 1-pp^{\prime}_{i}= \frac{1}{1+\alpha pp_{i}/k!},
503: \]
504:
505: so, by the definition of $\alpha$,
506:
507: \[
508: [(1-pp^{\prime}_{1})\ldots (1-pp^{\prime}_{r})]^{k!}= \frac{1}{\alpha}.
509: \]
510:
511: This means that the probability that in a given set of constraints
512: exactly one constraint (of type $i$) is chosen is equal to $pp_{i}/k!$,
513: the same as in model $M$. On the other
514: hand the probability that {\em no} constraint is chosen is equal to
515: $[(1-pp^{\prime}_{1})\ldots (1-pp^{\prime}_{r})]^{k!}= \frac{1}{\alpha}$.
516: But the same probability in model $M$
517: is $1-p$, and we know that $1-p \geq \frac{1}{\alpha}$. In both model
518: decisions on different sets of $k$ variables are independent. The
519: conclusion is that $M$ assigns a larger probability
520: than $\Gamma_{p^{\prime}_{1}, \ldots, p^{\prime}_{r}}$ to any sample
521: $x$ to which it assigns positive probability. Point (1) follows.
522:
523: Point (2) has a similar proof: by calculus the maximum value of $f(x)$
524: is obtained when the $p_{i}$'s are equal, so
525: \[
526: 0 = f(\alpha) \leq (1+\frac{\alpha p}{rk!})^{rk!} - \alpha \leq
527: e^{\alpha p} - \alpha.
528: \]
529:
530: Since $p=o(1)$, for large enough $n$ $e^{\alpha p} \leq 1+ \alpha p +
531: (\alpha p)^{2}$, so $(\alpha p)^{2}+ \alpha (p-1) +1 > 0$, in other
532: words
533: \[
534: \alpha (1-p) \leq 1+ (\alpha p)^{2}.
535: \]
536:
537: But the ratio of the probabilities associated to any given $\Phi$ by
538: $M$ and $\Gamma_{p^{\prime}_{1}, \ldots, p^{\prime}_{r}}(n,p)$ verifies
539: \[
540: \frac{ \Pr_{M(n,p,p_{1}, \ldots,
541: p_{r})}(\Phi)}{\Pr_{\Gamma_{p^{\prime}_{1}, \ldots, p^{\prime}_{r}}(n,p)}[\Phi]}\leq [\alpha (1-p)]^{rk!
542: {{n}\choose {k}}}\leq (1+ (\alpha p)^{2})^{rk! {{n}\choose {k}}}.
543: \]
544: Since $p=\theta(n^{1-k})$ and $k\geq 3$ the right-hand side is a
545: function of $n$ that is $1+o(1)$.
546:
547: To prove the result it is enough to observe that for $k\geq 3$ the
548: expected number of times a random formula in $\Gamma_{p^{\prime}_{1},
549: \ldots, p^{\prime}_{r}}(n,p)$ contains two different
550: clauses on the same set of variables is $o(1)$, since that will imply
551: that the satisfaction probabilities in the two models are related by a
552: $1-o(1)$ factor.
553: Indeed, this number is
554: \[
555: {{n}\choose {k}} \cdot [\sum_{\alpha,\beta}
556: \frac{pp^{\prime}_{\alpha}pp^{\prime}_{\beta}}{(k!)^{2}}],
557: \]
558:
559: where indices $\alpha,\beta$ span the set of different pairs of clauses
560: from a group.
561: Since each group is finite (contains $rk!$ clauses) and
562: $p=\theta(n^{1-k})$, this expected value is $\theta(n^{2-k})$, which is $o(1)$ for
563: $k\geq 3$.
564: \qed
565:
566:
567: \begin{definition}
568: Constraint $C_{2}$ is {\em an implicate of $C_{1}$}
569: iff every satisfying assignment for $C_{1}$ satisfies $C_{2}$.
570: \end{definition}
571:
572:
573: \begin{definition}
574: A boolean constraint $C$ {\em strongly depends on a literal} if
575: it has an unit clause as an implicate.
576: \end{definition}
577:
578: \begin{definition}
579: A boolean constraint $C$ {\em strongly depends on a 2-XOR relation} if
580: $\exists i,j\in \overline{1,k}$
581: such that constraint ``$x_{i}\neq x_{j}$'' is an implicate of $C$.
582: \end{definition}
583:
584: Our result is:
585:
586: \begin{theorem}\label{dichotomy:threshold}
587: Consider a generalized satisfiability problem $SAT({\cal P})$ (that is
588: not
589: trivially satisfiable by the ``all zeros'' or ``all ones''
590: assignment). Let ${\cal C}= supp({\cal P})$.
591: \begin{enumerate}
592: \item if some constraint in ${\cal C}$ strongly depends on one
593: component then $SAT({\cal P})$ has a coarse threshold.
594: \item if some constraint in ${\cal C}$ strongly depends on a
595: 2XOR-relation then $SAT({\cal P})$ has a coarse threshold.
596: \item in all other cases $SAT({\cal P})$ has a sharp threshold.
597: \end{enumerate}
598: \end{theorem}
599:
600: \begin{proof}
601:
602: \begin{enumerate}
603: \item
604: Suppose some clause $C$ implies a unit clause. We claim that $SAT({\cal
605: P})$ has a coarse threshold in the region where the expected number of
606: clauses is $\theta(\sqrt{n})$.
607:
608: That the probability that such a formula is bounded away from zero in
609: this region it is easy to see: consider a random formula with $c\sqrt n$
610: constraints, and let $H$ be the $k$-uniform formula hypergraph.
611: The expected number of pairs of edges $C_{i}$, $C_{j}$ that have
612: nonempty intersection is
613: \[
614: {{c\cdot \sqrt n}\choose {2}}\cdot (1 - \frac{{{n-k}\choose {k}
615: }}{{{n}\choose {k} }})\leq \frac{(ck)^{2}}{2}
616: \]
617:
618: Therefore with constant positive probability (that depends on $c$), all
619: vertices will have degree at most 1 in the hypergraph $H_{n}$, and the
620: formula will be satisfiable.
621:
622: If, on the other hand, both positive and negative unit clauses
623: are implicates of constraints in ${\cal P}$ then one can adapt the
624: well-known lower bound on the probability of intersection of two
625: random sets of size $\theta(\sqrt{n})$ to show that, with constant
626: probability a random formula will contain two contradictory unit
627: clauses as implicates, and be unsatisfiable.
628:
629: The proof is similar in the case when only one type of unit clauses
630: (w.l.o.g
631: assume it's the positive unit clauses) are implicates of constraints in
632: ${\cal C}$. Since $SAT({\cal C})$ is not trivial there exists a
633: constraint $C_{1}\in {\cal C}$ with an implicate of the type
634: $\overline{x_{1}}\OR \ldots \OR \overline{x_{b}}$, $b\geq 2$. We deal first with the
635: case when there exists a constraint
636: $C_{2}\neq C_{1}$ such that $C_{2}$ has an unit clause as implicate.
637: Then it is easy to construct a formula $F$ consisting of $b$ copies of
638: $C_{1}$and one copy of $C_{2}$ that implies the (unsatisfiable) formula
639: $\{x_{1}, \ldots, x_{b},\overline{x_{1}}\OR \ldots \OR
640: \overline{x_{b}}\}$.
641: It is easy to see that the expected number of copies of $F$ in a random
642: instance of $SAT({\cal P})$ with $\theta(\sqrt{n})$ clauses is
643: constant,
644: so the probability that the instance is unsatisfiable is bounded away
645: from zero.
646:
647: Finally, in the case when the only constraint in ${\cal C}$ that has an
648: unit
649: implicate is $C_{1}$. In this case one can use a trick similar to that
650: used
651: in the last paragraph of subsection~\ref{all:together}: we use half of
652: the
653: random copies of $C_{1}$ to imply (random) unit clauses, and the other
654: half
655: to imply (random) copies of $\overline{x_{1}}\OR \ldots \OR
656: \overline{x_{b}}$. This way we can produce, with constant probability, a copy of
657: the formula $F$.
658:
659: \item
660: Suppose now that $C$ does not fall in the first case but
661: has a 2XOR implicate. In this case Creignou and
662: Daud\'{e} have shown when ${\cal P}$ is the uniform distribution
663: (and this extends directly to the case of a general
664: probability distribution as well) that $p_{1/2}= \Omega(n^{1-k}))$ and
665: the expected number of constraints is $\theta{n}$. Let $c_{SAT({\cal
666: P})}\cdot n$ be the expected number of constraints corresponding to
667: $p_{1/2}$. Then there exists $\delta > 0 $ such that, for every $n$,
668: $c_{SAT({\cal P})}= c_{SAT({\cal P})}(n) > \delta$.
669:
670: Let us consider a random formula with $c\cdot n$ of constraints.
671: By the well known result on triangles in random graphs it follows that
672: with positive probability one can use $C$ to create a ``contradictory
673: triangle''. Therefore it is easy to see that for every $c>0$
674: the satisfaction probability is bounded away from 1. It is easy to see
675: than this statement, together with the fact that $c_{SAT({\cal P})}(n)
676: > \delta$ together imply that $SAT({\cal P})$ has a coarse threshold.
677:
678: \item
679: We will concentrate in
680: the sequel on the last one. As discussed previously, for $k\geq 3$
681: Molloy's model can be mapped onto
682: a version of the constraint probability model. In the case $k=2$ we can
683: establish the existence of a sharp
684: threshold in a direct manner, by the same method as the one used by
685: Chv\'{a}tal and Reed for 2-SAT \cite{mickgetssome}
686: (the complete proof of this case will be presented in the full
687: version). Indeed, by the first two points of the Theorem, and the assumption
688: $k=2$
689: the only possible constraints in ${\cal P}$ can be constraints $x\OR
690: y$, $\overline{x} \OR \overline{y}$, $\overline{x}\OR y$, $x \OR
691: \overline{y}$, $x=y$, and
692: the first two are always present.
693:
694: Let us now consider the case $k\geq 3$, using the modified version of
695: the constant probability model. We note first that there exists a simple
696: observation
697: that allows us to reduce the problem to the case when ${\cal P}$ is
698: the uniform probability: the Friedgut-Bourgain
699: result on sharp/coarse threshold properties in monotone problems
700: \cite{friedgut:k:sat} uses the following result, an easy consequence of the
701: Mean Value Theorem:
702: if a monotonic property $A$ does {\em not} have a sharp threshold
703: (under model $\Gamma(n.p)$) then there exists $p^{*}=p(n)$ and a constant
704: $C>0$ such that (for
705: infinitely many $n$)
706:
707: \begin{equation}\label{coarse}
708: p^{*}\cdot I(p^{*}) < C,
709: \end{equation}
710:
711: where $I(p^{*})=\frac{d\mu_{p}(A)}{dp}|_{p=p^{*}}$.
712:
713: The same argument works when $A$ is considered under model
714: $\Gamma_{p_{1},\ldots p_{r}}(n,p)$. Moreover, it is an easy consequence of Russo's
715: Lemma for $\Gamma_{p_{1},\ldots p_{r}}(n,p)$
716: that if equation~\ref{coarse} holds for $p^{*}$ and some tuple
717: ${p_{1},\ldots p_{r}}$, then it also holds (with a different constant $C$) for
718: $p^{*}$ and
719: tuple $p_{1}= \ldots p_{r}=1/r$. In other words it is enough to obtain
720: a contradiction to the assumption that $SAT({\cal P})$ did not have a
721: sharp
722: threshold in the case when ${\cal P}$ is the uniform probability, which
723: is what we show next.
724:
725: \subsection{A base case}
726:
727: To prove the theorem in the uniform case we will first consider a
728: ``base case'' that is
729: easier to
730: explain, and will be of use in solving the general case: let $a,b$ be
731: two integers
732: (not necessarily equal), both greater or equal to 2. Let $S$ be a set
733: consisting of two constraints $C_{1}, C_{2}$ of arity $a$,
734: respectively $b$, specified by
735: $C_{1}= \overline{x_{1}}\OR \ldots \overline{x_{a}}$, $C_{2}= x_{1}\OR
736: \ldots x_{b}$. One can represent $SAT(S)$ in the framework of
737: Definition~\ref{model}
738: by ``simulating'' $C_{1}$, $C_{2}$ by suitable constraints of arity
739: $\max\{a,b\}$.
740:
741: We first outline how to prove that $SAT(S)$ has a sharp threshold:
742: we apply the Friedgut-Bourgain result \cite{friedgut:k:sat} and infer
743: that if
744: $SAT(S)$ did not have a sharp threshold than, for some $\epsilon,
745: \delta_{0}, K>0$ and some probability
746: $p=p(n)\in [p_{\epsilon}, p_{1-\epsilon}]$
747:
748: \begin{enumerate}
749: \item either $
750: \Pr_{p=p(n)} [\Phi \mbox{ contains some } F\in \overline{SAT}\mbox{
751: with }|F|\leq K] > \delta_{0}$, or
752: \item there exists a fixed satisfiable formula $F_{0}$, $|F_{0}|\leq K$
753: such that $
754: \Pr_{p=p(n)} [\Phi \AND F_{0} \in \overline{SAT}] - \Pr [ \Phi \in
755: \overline{SAT}] >
756: \delta_{0}$.
757:
758: \end{enumerate}
759:
760: One easy observation is that in the second alternative we can always
761: assume
762: that $F$ consists of a conjunction of
763: unit clauses: if $F$ is satisfiable and satisfies (2), then so does
764: the conjunction of unit clauses specifying
765: one satisfying assignment of $F$. The first alternative is eliminated
766: by a result (Proposition 4.6) from
767: \cite{creignou:daude:sat2002}.
768: The key to disproving the second alternative, in the case of $k$-SAT,
769: is
770: a geometric result, Lemma 5.7 in \cite{friedgut:k:sat}.
771: We restate it here for completeness.
772:
773:
774: \begin{lemma}\label{friedgut}
775: For a sequence $A=(A_{n})$ of subsets of the $n$-dimensional hypercube,
776: $A\subseteq \{0,1\}^{n}$, define $A$ to be {\em
777: $(d,m,\epsilon)$-coverable} if the probability for a union
778: of a random choice of $d$ subcubes (hyperplanes) of codimension $m$ to
779: cover $A$ is greater than $\epsilon$ for large enough $n$.
780:
781: Let $f(n)$ be any function that tends to infinity as $n\goesto \infty$.
782: For fixed $k$, $d$, and $\epsilon$ any $A$ that is
783: $(d,1,\epsilon)$-coverable is $(f(n),k, \epsilon)$-coverable.
784: \end{lemma}
785:
786: The connection with satisfiability can be explained as follows: the set
787: $A$
788: in the application of the Lemma~\ref{friedgut} will (intuitively) refer
789: to the set of satisfying assignments of random formula $\Phi$.
790: Hyperplanes
791: of codimension 1 are associated to unit clauses, more precisely to the
792: set
793: of assignments {\em forbidden} by a given unit clause. The fact that
794: $A$
795: can be covered with probability $\epsilon$ by a union of $d$ random
796: hyperplanes of codimension 1 parallels the fact that with probability
797: $\epsilon$,
798: $\Phi \AND F_{0}$ becomes unsatisfiable. This is what the result of
799: Friedgut-Bourgain gives us (for $\epsilon = \delta_{0}$, under the
800: assumption that $k$-SAT does not have a sharp
801: threshold). Hyperplanes of
802: codimension $k$ correspond to the set of assignments forbidden by a
803: given $k$-clause, and the conclusion of the geometric lemma is that adding
804: any small (but unbounded) number $f(n)$ of random
805: $k$-CNF clauses to random formula $\Phi$ boosts the probability of {\em
806: not} being satisfiable at least as much as the addition of the
807: (constantly many) unit clauses in $F_{0}$.
808:
809: For small enough $f(n)$ this statement can be directly refuted, by
810: concentration results for the binomial distribution (Lemma 5.6 in
811: \cite{friedgut:k:sat}). A simpler
812: and more general way to derive it is given as
813: Lemma 3.1 in \cite{achlioptas:friedgut:kcol}.
814:
815: The same outline works for the case we consider. To state the geometric
816: result we need, however, to
817: work with two types of hyperplanes:
818:
819: \begin{definition}
820: Let $H_{n}=\{0,1\}^{n}$ be the $n$-dimensional hypercube, and let
821: $w_{i}$ denote the value of the $i$'th
822: bit of element $w\in H_{n}$.
823: A {\em positive hyperplane of codimension $d$} is a subset of points
824: of $H_{n}$ defined by a system of equations $
825: x_{i_{1}}= \ldots = x_{i_{d}}=1$, where the $x_{i}$'s are distinct
826: variables. Negative hyperplanes have
827: a similar definition.
828: \end{definition}
829:
830: Our version of the geometric Lemma is
831:
832: \begin{lemma}\label{geometric}
833: For a sequence $A=(A_{n})$ of subsets of the $n$-dimensional hypercube,
834: $A_{n}\subseteq \{0,1\}^{n}$ define $A$ to be {\em $(n_{1}, d_{1},
835: d_{2},m_{1}, m_{2},\epsilon)$-coverable} if the probability of a union
836: of a random choice of $d_{1}$ negative hyperplanes of codimension
837: $m_{1}$
838: and $d_{2}$ positive hyperplanes of codimension $m_{2}$ to cover
839: $A_{n}$ is at least $\epsilon$ if
840: $n\geq n_{1}$. Let $f(n)$, $g(n)$ be any
841: functions that tends to infinity as $n\goesto \infty$. For fixed
842: $k_{1}$,$k_{2}$, $d$, and $0<\delta <\epsilon$,
843: there exists $n_{2}$ that depends on $k_{1}, k_{2}, d,\epsilon,\Delta,
844: n_{1}$ (but {\em not} $A$) such that for any
845: $n\geq n_{2}$ any $A_{n}\subseteq \{0,1\}^{n}$ that is
846: $(n_{1},d_{1},d_{2},1,1,\epsilon)$-coverable is
847: $(n_{2},f(n),g(n),k_{1},k_{2}, \epsilon - \delta)$-coverable.
848: \end{lemma}
849: We will in fact prove a stronger version of the Lemma:
850:
851: \begin{lemma}\label{geometric:2}
852: For a sequence $A=(A_{n})$ of subsets of the $n$-dimensional hypercube,
853: assume that
854: \[
855: \Pr[A\subseteq H_{1}\cup \ldots H_{d}]\geq \epsilon
856: \]
857:
858: for all $n\geq n_{1}$,
859: where the $H_{i}$'s are random hyperplanes of codimension 1,
860: $d_{1}$ of them negative, $d_{2}$ of them positive.
861:
862:
863: Let $f(n)$, $g(n)$ be any
864: functions that tends to infinity as $n\goesto \infty$. For fixed
865: $k_{1}$,$k_{2}$, $d$, and $\delta >0$,
866: there exists $n_{*}=n(n_{1},d_{1},d_{2},k_{1}, k_{2},
867: \epsilon,\delta,f,g)$, (however it does {\em not} depend on $A$)
868: such that for any
869: $n\geq n_{*}$ and any $i$, $0\leq i\leq d$
870: \begin{equation}\label{conclusion}
871: Pr[A\subseteq P_{1}\cup \ldots \cup P_{\frac{if(n)}{d}}\cup N_{1}\ldots
872: \cup N_{\frac{ig(n)}{d}}\cup H_{i+1}\cup
873: \ldots \cup H_{d}]\geq
874: Pr[A\subseteq H_{1}\cup \ldots \cup H_{d}]-\frac{i\delta}{d},
875: \end{equation}
876:
877: where the $N_{i}$'s are random negative hyperplanes of codimension
878: $k_{1}$ and the
879: $P_{i}$'s are random positive hyperplanes of codimension $k_{2}$.
880: \end{lemma}
881:
882: \begin{proof}
883:
884: It is easy to see that one can assume that $d|f(n)$, $d|g(n)$ (since it
885: is enough to prove the lemma for
886: $\overline{f}(n)=d\lfloor f(n)/d \rfloor$, $\overline{g}(n)=d\lfloor
887: g(n)/d \rfloor$).
888:
889: We will prove the lemma by double induction on $d_{1}, d_{2}$. By
890: symmetry we only need to consider two ``base cases:'' \\
891:
892: {\bf Case 1: $d_{1}=0$, $d_{2}=1$} \\
893:
894: In this case (and the dual, $d_{1}=1$, $d_{2}=0$) we can replace, for
895: $i=1$,
896: equation~\ref{conclusion}
897: by the stronger:
898: \begin{equation}\label{conclusion:d=1}
899: \Pr[A\subseteq P_{1}\cup \ldots \cup P_{f(n)}\cup N_{1}\ldots \cup
900: N_{g(n)}]\geq
901: 1-\delta.
902: \end{equation}
903:
904: The hypothesis implies that for $n\geq n_{1}$
905: there exist $n\cdot \epsilon$ positive hyperplanes of codimension $1$
906: such that
907: \[
908: A_{n} \subseteq P^{(n)}_{1}\cap \ldots \cap P^{(n)}_{n\cdot \epsilon}.
909: \]
910:
911: We will assume, w.l.o.g., in what follows that $A$ is in fact {\em
912: equal} to
913: the right hand side.
914: If $KP$ is a random positive hyperplane of
915: codimension $k_{1}$
916: then
917: \[
918: \Pr[A \not \subseteq KP] = 1-\frac{{{n\cdot \epsilon}\choose
919: {k_{1}}}}{{{n}\choose {k_{1}}}}.
920: \]
921:
922: Indeed, suppose KP is specified by the (random set of) equations
923: $x^{(n)}_{1}= \ldots =x^{(n)}_{k_{1}}=1$. The condition that
924: $A\subseteq KP$ is
925: equivalent to
926: \[
927: \{x^{(n)}_{1}, \ldots x^{(n)}_{k_{1}}\} \subseteq \{p^{(n)}_{1},
928: \ldots, p^{(n)}_{n\cdot \epsilon}\},
929: \]
930:
931: where $\{p^{(n)}_{1}, \ldots, p^{(n)}_{n\cdot \epsilon}\}$ are the
932: literals that
933: specify the hyperplanes $P^{(n)}_{1}, \ldots, P^{(n)}_{n\cdot
934: \epsilon}$.
935:
936: Thus the probability that $A$ is included in the union of $g(n)$ random
937: positive hyperplanes $KP_{i}$ of codimension $k_{1}$ is at least $1 -
938: \Pr[ (\forall i): A\not \subseteq KP_{i}]$, which is equal to
939: \[
940: 1 - (1-\frac{{{n\cdot \epsilon}\choose {k_{1}}}}{{{n}\choose
941: {k_{1}}}})^{g(n)} \sim 1- (1-\epsilon^{k_{1}})^{g(n)}\goesto 1 \mbox{
942: as }
943: n\goesto \infty.
944: \]
945:
946: It follows that there exists $n_{*}=n(n_{1},d_{1},d_{2},k_{1}, k_{2},
947: \epsilon,\delta,f,g)$
948: such that for $n\geq n_{*}$ the
949: left-hand side is larger than $1-\delta$.
950:
951: \vspace{5mm}
952: {\bf Case 2: $d_{1}+d_{2}>1$} \\
953:
954: It is enough to prove that there exists
955: $n_{i}=n(n_{1},d_{1},d_{2},k_{1}, k_{2}, \epsilon,\delta,f,g,i)$
956: such that~\ref{conclusion} holds, for a fixed value of $i$, $0\leq i
957: \leq d$, when
958: $n\geq n_{i}$. Then we can take $n_{*}=max\{n_{i}\}$.
959:
960: We prove this on induction on $i$. The claim is clearly true for $i=0$.
961: Assume, therefore, that
962: the claim is true up to $i$; we will prove it for $i+1$.
963:
964: Denote for all $j$
965: \[
966: p_{j}= \Pr[A \subseteq P_{1}\cup \ldots \cup
967: P_{\frac{(j-1)f(n)}{d}}\cup N_{1}\cup \ldots \cup
968: N_{\frac{(j-1)g(n)}{d}}\cup H_{j}\cup \ldots \cup H_{d}].
969: \]
970:
971: To accomplish that it is enough to show that
972: \begin{equation}\label{inductive:step}
973: p_{i+1}\geq p_{i}-\delta,
974: \end{equation}
975:
976: By the Bayes formula:
977:
978: \begin{eqnarray*}\label{bayes}
979: p_{i}=\Pr[A \subseteq P_{1}\cup \ldots \cup P_{\frac{(i-1)f(n)}{d}}\cup
980: N_{1}\cup \ldots \cup
981: N_{\frac{(i-1)g(n)}{d}}\cup H_{i}\cup \ldots \cup H_{d}] = \\
982: = \sum_{B}
983: \ Pr[B\subseteq H_{i}| A\setminus (P_{1}\cup \ldots \cup
984: P_{\frac{(i-1)f(n)}{d}}\cup N_{1}\cup \ldots \cup
985: N_{\frac{(i-1)g(n)}{d}}) = B] \cdot \\
986: \cdot \Pr[A\setminus (P_{1}\cup \ldots \cup P_{\frac{(i-1)f(n)}{d}}\cup
987: N_{1}\cup \ldots \cup
988: N_{\frac{(i-1)g(n)}{d}}) = B]
989: \end{eqnarray*}
990:
991: Assume without loss of generality that $H_{i}$ is a positive
992: hyperplane.
993:
994: Let $\gamma= \frac{\delta}{2d}$. Let
995: \[
996: C_{\gamma}=\{B\subseteq \{0,1\}^{n}: Pr[B\subset P] > \gamma\}.
997: \]
998:
999: Let $\alpha$ be the sum of those terms in~\ref{bayes} corresponding to
1000: sets $B\in C_{\gamma}$,
1001: and let $\beta$ be the sum corresponding to sets $B\in
1002: \overline{C_{\gamma}}$.
1003:
1004: From the definition of $C_{\gamma}$ it follows that
1005: \[
1006: 0\leq \beta \leq \gamma,
1007: \]
1008: therefore
1009: \begin{eqnarray*}\label{inequality}
1010: \Pr[A\setminus (P_{1}\cup \ldots \cup P_{\frac{(i-1)f(n)}{d}}\cup
1011: N_{1}\cup \ldots \cup
1012: N_{\frac{(i-1)g(n)}{d}}) \in C_{\gamma}]\geq \\
1013: \geq \frac{1}{\gamma}\cdot [\Pr[A \subseteq P_{1}\cup \ldots \cup
1014: P_{\frac{(i-1)f(n)}{d}}\cup N_{1}\cup \ldots \cup
1015: N_{\frac{(i-1)g(n)}{d}}\cup H_{i}\cup \ldots \cup H_{d}] -\gamma]= \\
1016: = \frac{1}{\gamma}\cdot [p_{i}-\gamma].
1017: \end{eqnarray*}
1018:
1019: On the other hand
1020:
1021: \begin{eqnarray*}\label{bayes:2}
1022: p_{i+1}=\Pr[A \subseteq P_{1}\cup \ldots \cup P_{\frac{if(n)}{d}}\cup
1023: N_{1}\cup \ldots \cup
1024: N_{\frac{ig(n)}{d}}\cup H_{i+1}\cup \ldots \cup H_{d}] = \\
1025: = \sum_{B}
1026: \ Pr[B\subseteq P_{\frac{(i-1)f(n)}{d}+1}\cup \ldots \cup
1027: P_{\frac{if(n)}{d}}
1028: \cup N_{\frac{(i-1)g(n)}{d}+1}\cup \ldots \cup N_{\frac{ig(n)}{d}}| \\
1029: A\setminus (P_{1}\cup \ldots \cup P_{\frac{(i-1)f(n)}{d}}\cup N_{1}\cup
1030: \ldots \cup
1031: N_{\frac{(i-1)g(n)}{d}}) = B] \cdot \\
1032: \cdot \Pr[A\setminus (P_{1}\cup \ldots \cup P_{\frac{(i-1)f(n)}{d}}\cup
1033: N_{1}\cup \ldots \cup
1034: N_{\frac{(i-1)g(n)}{d}}) = B]
1035: \end{eqnarray*}
1036:
1037: Let $\overline{f}(n)=f(n)/d$, $\overline{g}(n)= g(n)/d$. Since the
1038: $N_{i}$'s, $P_{i}$'s
1039: are random hyperplanes, one can
1040: rewrite the previous recurrence as
1041:
1042: \begin{eqnarray*}\label{bayes:3}
1043: p_{i+1}=\Pr[A \subseteq P_{1}\cup \ldots \cup P_{\frac{if(n)}{d}}\cup
1044: N_{1}\cup \ldots \cup
1045: N_{\frac{ig(n)}{d}}\cup H_{i+1}\cup \ldots \cup H_{d}] = \\
1046: = \sum_{B}
1047: \Pr[B\subseteq P_{1}\cup \ldots \cup P_{\overline{f(n)}}
1048: \cup N_{1}\cup \ldots \cup N_{\overline{g(n)}}| \\
1049: A\setminus (P_{1}\cup \ldots \cup P_{\frac{(i-1)f(n)}{d}}\cup N_{1}\cup
1050: \ldots \cup
1051: N_{\frac{(i-1)g(n)}{d}}) = B] \cdot \\
1052: \cdot \Pr[A\setminus (P_{1}\cup \ldots \cup P_{\frac{(i-1)f(n)}{d}}\cup
1053: N_{1}\cup \ldots \cup
1054: N_{\frac{(i-1)g(n)}{d}}) = B]
1055: \end{eqnarray*}
1056:
1057:
1058: Since all terms are nonnegative, one can obtain a lower bound on the
1059: left-hand size of this
1060: latter terms by only considering those $B\in C_{\gamma}$.
1061:
1062: Applying the induction hypothesis from case one for all $B\in
1063: C_{\gamma}$ and
1064: $n\geq
1065: n_{i}=n_{*}(n_{1},0,1,k_{1},k_{2},\gamma,\delta,\overline{f},\overline{g})$, we
1066: infer that for all such $B$,
1067: \[
1068: \Pr[B\subseteq P_{1}\cup \ldots \cup P_{\overline{f(n)}}
1069: \cup N_{1}\cup \ldots \cup N_{\overline{g(n)}}]\geq (1-\gamma),
1070: \]
1071: therefore
1072: \begin{eqnarray*}
1073: p_{i+1}= \Pr[A \subseteq P_{1}\cup \ldots \cup P_{\frac{if(n)}{d}}\cup
1074: N_{1}\cup \ldots \cup
1075: N_{\frac{ig(n)}{d}}\cup H_{i+1}\cup \ldots \cup H_{d}]\geq \\
1076: (1-\gamma)\cdot
1077: \sum_{B\in C_{\gamma}} Pr[A\setminus (P_{1}\cup \ldots \cup
1078: P_{\frac{(i-1)f(n)}{d}}\cup N_{1}\cup \ldots \cup
1079: N_{\frac{(i-1)g(n)}{d}}) = B] \\
1080: = (1-\gamma) \cdot Pr[A\setminus (P_{1}\cup \ldots \cup
1081: P_{\frac{(i-1)f(n)}{d}}\cup N_{1}\cup \ldots \cup
1082: N_{\frac{(i-1)g(n)}{d}})\in C_{\gamma}] \\
1083: \geq \frac{(1-\gamma)}{\gamma}\cdot [\Pr[A \subseteq P_{1}\cup \ldots
1084: \cup P_{\frac{(i-1)f(n)}{d}}\cup N_{1}\cup \ldots \cup
1085: N_{\frac{(i-1)g(n)}{d}}\cup H_{i}\cup \ldots \cup H_{d}] -\gamma] = \\
1086: \frac{(1-\gamma)}{\gamma}\cdot [p_{i}-\gamma].
1087: \end{eqnarray*}
1088:
1089: Since $\gamma \leq 1$,
1090: \[
1091: p_{i+1}\geq (1-\gamma)\cdot (p_{i}-\gamma)= p_{i} - \gamma \cdot
1092: [1+p_{i}-\gamma]\geq p_{i}-2\gamma.
1093: \]
1094:
1095: which is precisely equation~\ref{inductive:step} (that we wanted to
1096: prove).
1097:
1098: \end{proof}
1099: \qed
1100:
1101: \subsection{How to contradict Lemma~\ref{geometric}}
1102:
1103: It is now easy to infer the fact that $SAT(S)$ has a sharp threshold,
1104: by
1105: using the previous lemma with $d_{1}= |F_{0}\cap Var|$,
1106: $d_{2}=|F_{0}|-d_{1}$,
1107: $k_{1}=b$, $k_{2}=a$, $\epsilon = \delta_{0}$, $\Delta = \delta_{0}/2$
1108: and a refutation of the geometric lemma similar to the
1109: one for 3-SAT.
1110:
1111: The conclusion of the Geometric Lemma (similar to the one for $k$-SAT)
1112: is that, by adding any number $f(n)$ of copies of $x_{1}\OR \ldots \OR
1113: x_{a}$
1114: and $g(n)$ copies of $\overline{x_{1}}\OR \ldots \OR x_{b}$ suffices to
1115: boost the unsatisfiability probability by a constant.
1116:
1117: However, Lemma 3.1 from \cite{achlioptas:friedgut:kcol} asserts that
1118: adding up to $o(\sqrt(n))$ random clauses is not enough to boost the
1119: unsatisfiability
1120: probability by more than $o(1)$.
1121:
1122: Because of the nature of the random model, adding $H(n)=o(\sqrt(n))$
1123: random clauses insures that w.h.p. we have $\Theta(H(n))$ copies
1124: of each type of clause, as long as $H(n)$ grows faster than some power
1125: of $n$. Taking the number of such copies to be the functions
1126: $f(n)$, $g(n)$ contradicts the consequence of the Geometric Lemma.
1127:
1128:
1129: Note that we do {\em not} make use of all the details of the random
1130: model (such as the precise number of copies of each clause in random
1131: instances at $p=p(n)$), but only that:
1132: \begin{itemize}
1133: \item the expected number of copies of both $C_{1}$ and $C_{2}$ in a
1134: random formula at $p=p(n)$ is unbounded. This is enough to make the
1135: analog
1136: of Lemma 3.1 from \cite{achlioptas:friedgut:kcol} work.
1137: \item the clauses are independent.
1138: \end{itemize}
1139:
1140: \subsection{Putting it all together} \label{all:together}
1141:
1142: In the previous section we have proved a geometric lemma that is used
1143: to prove that the above-defined set $SAT(S)$ has a sharp threshold.
1144:
1145: Consider now a set ${\cal C}$ of constraints that satisfies the
1146: condition (3)
1147: of the theorem. Since $SAT({\cal C})$ is not trivially satisfied by the
1148: ``all zero'' (all ones) assignment, there exist constraints $C_{1}$,
1149: $C_{2}$ in ${\cal C}$ and $a,b \geq 1$ such that $C_{1}\models x_{1}\OR
1150: \ldots x_{a}$,
1151: $C_{2}\models \overline{x_{1}}\OR \ldots \OR \overline{x_{b}}$. In fact
1152: $a,b \geq 2$, otherwise some constraint in ${\cal C}$ would strongly
1153: depend on one variable.
1154:
1155: Just as in the base case, condition (i) in the Friedgut-Bourgain result
1156: is eliminated by the result of Creignou and Daud\'{e}, and the formula
1157: $F_{0}$ in condition
1158: (ii) can be assumed to consist of a conjunction of unit clauses.
1159: Reflecting this fact, the geometric lemma needed for the
1160: general case has the same hypothesis as
1161: the one of Lemma~\ref{geometric}: the set $A$ can be covered by a union
1162: of random hyperplanes of codimension 1.
1163: However, the covering desired in the conclusion no longer consists
1164: of hyperplanes, but of (general) sets of points in the hyperplane,
1165: corresponding to sets of assignments forbidden by a certain constraint
1166: $C$.
1167:
1168: The critical observation (easiest to make in the case when $C_{1}\neq
1169: C_{2}$ )
1170: is that {\bf Lemma~\ref{geometric} implies the geometric result we need
1171: for this case}:
1172: since $C_{1}\models x_{1}\OR \ldots x_{a}$, the ``forbidden set''
1173: associated to $C_{1}$ contains the ``forbidden set'' associated to
1174: constraint
1175: $x_{1}\OR \ldots x_{a}$ in Lemma~\ref{geometric}, the hyperplane
1176: $x_{1}= \ldots = x_{a}=0$. A similar result holds for $C_{2}$
1177: and the positive hyperplanes.
1178:
1179: Thus each ``covering set'' in a conclusion of the (general case of
1180: the ) geometric lemma contains a corresponding ``covering set'' from
1181: the conclusion of Lemma~\ref{geometric}. In other words {\bf the
1182: geometric lemma for $SAT({\cal C})$ follows from the geometric lemma for the
1183: base case by monotonicity}.
1184:
1185: All is left to show is that the analog of Lemma 3.1 from
1186: \cite{achlioptas:friedgut:kcol} also
1187: works in this general case. We have previously observed that this
1188: amounts
1189: to showing that the expected number of copies of $C_{1}$, $C_{2}$ in
1190: a random formula is unbounded. But Proposition 3.5 in
1191: \cite{creignou:daude:sat2002} (slightly generalized to probability distributions ${\cal
1192: P}$ that are not uniform) and the details of the random model imply
1193: that in fact this number is linear.
1194:
1195: A simple modification of this argument holds when $C_{1}=C_{2}$. To see
1196: this, note that in the proof of the fact that
1197: Lemma~\ref{geometric} holds for {\em some} small enough (but unbounded)
1198: $f(n),g(n)$ is enough to obtain a contradiction.
1199:
1200: The expected number of copies of $C_{1}$ in a random formula is linear,
1201: denote it by $h(n)$. Dividing the set of such copies into two (random)
1202: sets of cardinality $h(n)/2$ yields
1203: infinitely many random copies of $C_{1}$
1204: used to imply $x_{1}\OR \ldots \OR x_{a}$ in the previous argument
1205: and infinitely many copies used to imply
1206: $\overline{x_{1}}\OR \ldots \OR \overline{x_{b}}$. We then apply the
1207: same strategy as in the first case.
1208:
1209: To summarize: the proof follows from the corresponding
1210: argument for the base case by monotonicity. It
1211: critically uses the fact that we are {\em not} in cases (i) or (ii) of
1212: the Theorem, since it is only under these conditions when the first
1213: alternative in the
1214: Friedgut-Bourgain argument can be eliminated.
1215:
1216:
1217: \end{enumerate}
1218: \end{proof}
1219: \qed
1220:
1221: \section{3-SAT has a first-order phase transition}
1222:
1223: \begin{theorem}\label{3sat:first-order}
1224: $k$-SAT, $k\geq 3$ has a first-order phase transition. In other words
1225: there exists
1226: $\eta > 0 $ such that for every sequence $p = p(n)$
1227: we have
1228: \begin{equation}\label{jump}
1229: \lim_{n\goesto \infty} \Pr_{p=p(n)}[\Phi \in SAT] = 0 \implies
1230: \lim_{n\goesto \infty}\Pr_{p=p(n)}[ \frac{|Spine(\Phi)|}{n}\geq \eta]=
1231: 1.
1232: \end{equation}
1233: \end{theorem}
1234:
1235: \begin{proof}
1236:
1237: We start by giving a simple sufficient condition for a literal to
1238: belong to the spine of the formula:
1239:
1240: \begin{claim}\label{spine:unsat}
1241: Let $\Phi$ be a minimally unsatisfiable formula, and let $x$ be a
1242: literal that appears in $\Phi$. Then $x\in Spine(\Phi)$.
1243: \end{claim}
1244: \begin{proof}
1245: Let $C$ be a clause that contains $x$. By the minimal unsatisfiability
1246: of $\Phi$, $\Phi \setminus \{C\} \in SAT$. On the other hand $\Phi
1247: \setminus \{C\}
1248: \AND \{x\} \in \overline{SAT}$, otherwise $\Phi$ would also be
1249: satisfiable.
1250: Thus $x \in Spine(\Phi \setminus \{C\})$.
1251: \end{proof}
1252: \qedbox
1253:
1254: Thus, to show that 3-SAT has a first-order phase transition it is
1255: enough to show that a random unsatisfiable formula contains w.h.p.
1256: a minimally unsatisfiable subformula containing a linear number of
1257: literals. A way to accomplish this is by using the two ingredients of
1258: the Chv\'{a}tal-Szemer\'{e}di proof \cite{chvatal:szemeredi:resolution}
1259: that random 3-SAT has exponential
1260: resolution size w.h.p. They are explicitly stated to make the
1261: argument self-contained:
1262:
1263: \begin{claim}
1264: There exists a constant $\delta>0$ such that for every constant $c>c_{3
1265: SAT}$ with high
1266: probability (as $n\goesto \infty$) a random formula $\Phi$ with $n$
1267: variables and $cn$
1268: clauses has no minimally unsatisfiable subformula of size less than
1269: $\delta \cdot n$.
1270: \end{claim}
1271:
1272: \begin{claim}
1273: There exists $\eta >0$ so that w.h.p. for every $c>c_{3-SAT}$ all
1274: subformulas
1275: of a random formula $\Phi$ having between $(\delta/2)\cdot n$ and
1276: $\delta\cdot n$
1277: clauses contain at least $\eta\cdot n$ (pure) literals (corresponding
1278: to different variables).
1279: \end{claim}
1280:
1281: The argument is now transparent: if $\Phi$ is unsatisfiable then
1282: w.h.p. a minimally unsatisfiable subformula $\Xi$ of $\Phi$ has
1283: size at least $\delta n$. By the second claim, applied to an arbitrary
1284: subformula of $\Xi$ of size $(3\delta n)/4$, we infer that w.h.p. $\Xi$
1285: contains at least many $\eta\cdot n$ different variables.
1286:
1287: \end{proof}
1288:
1289:
1290:
1291: \section{First-order phase transitions and resolution complexity
1292: of random generalized satisfiability problems}
1293:
1294: In this section we extend the previous result (and the connection
1295: between first-order phase transitions and resolution complexity) to
1296: other
1297: classes of satisfiability problems. Interestingly, we find that a
1298: condition
1299: Molloy investigated in \cite{molloy-stoc2002} is a sufficient condition
1300: for the
1301: existence of a first-order phase transition.
1302:
1303: It turns out that there are differences between the case of random
1304: $k$-SAT and the general case that force us to employ an
1305: alternative definition of the spine. The most obvious one is that
1306: formula (~\ref{spine:initial}) involves
1307: negations of variables, whereas Molloy's model does not. But it has
1308: more serious problems: consider for example {\em $k$-uniform hypergraph
1309: 2-coloring}
1310: specified as $SAT(\{C_{0}\})$, where $C_{0}(x_{1}, \ldots, x_{k})$ has
1311: the interpretation ``not all of $x_{1}, \ldots x_{k}$ are equal''.
1312: Because of the built-in symmetry to permuting colors 0 and 1, the
1313: spine of {\em any} instance is empty (under
1314: definition~\ref{spine:initial}).
1315: Similar phenomena have appeared before (and forced a different
1316: definition of the backbone/spine)
1317: in {\em k-coloring \cite{frozen:development}} (symmetry = permutation
1318: of colors) or {\em graph partition} \cite{graph-partition:transition}
1319: (symmetry = permutation of sides).
1320: There are other ways (to be detailed in the journal version of the
1321: paper) in which the original definition of the
1322: spine behaves differently in the general case than in that of random
1323: $k$-SAT. Our solution is to define the concept of spine of a
1324: random instance of a satisfiability problem $SAT(\cal P)$ in a slightly
1325: different way. The definition is consistent with those in
1326: \cite{frozen:development}, \cite{graph-partition:transition}.
1327:
1328: \begin{definition}$
1329: Spine(\Phi) = \{ x\in Var | (\exists) \Xi \subseteq \Phi, \Xi \in SAT,
1330: (\exists) C\in {\cal C}, x\in C,\mbox{ such that }
1331: \Xi \AND C \in \overline{SAT}\}$.
1332:
1333: \end{definition}
1334:
1335: It is easy to see that, for $k$-CNF formulas whose (original) spine
1336: contains at
1337: least three literals a variable $x$ is in the (new version of the)
1338: spine if and only if either $x$ or $\overline{x}$
1339: were present in the old version. In particular the new definition does
1340: not change the order of the
1341: phase transition of random $k$-SAT. Moreover the proof of
1342: Claim~\ref{spine:unsat} carries over to the general case.
1343: The {\em resolution complexity} of an instance $\Phi$ of $SAT({\cal
1344: P})$ is defined as the resolution complexity of the formula obtained by
1345: converting each constraint of $\Phi$ to CNF-form.
1346:
1347:
1348: \begin{definition} Let ${\cal P}$ be such that $SAT({\cal P})$ has a
1349: sharp
1350: threshold.
1351: Problem $SAT({\cal P})$ has a {\em first-order phase transition} if
1352: there exists
1353: $\eta > 0 \mbox{ such that for every sequence } p = p(n)$
1354: we have
1355: \begin{equation}\label{jump:general}
1356: \lim_{n\goesto \infty} \Pr_{p=p(n)}[\Phi \in SAT] = 0 \implies
1357: \lim_{n\goesto \infty}\Pr_{p=p(n)}[ \frac{|Spine(\Phi)|}{n}\geq \eta]=
1358: 1.
1359: \end{equation}
1360: If, on the other hand, for every $epsilon >0$ there exists $p^{\epsilon}(n)$
1361: with
1362: \begin{equation}\label{jump:continuous}
1363: \lim_{n\goesto \infty} \Pr_{p=p^{\epsilon}(n)}[\Phi \in SAT] = 0 \mbox{ and }
1364: \lim_{n\goesto \infty}\Pr_{p=p(n)}[ \frac{|Spine(\Phi)|}{n}\geq \epsilon]=
1365: 0
1366: \end{equation}
1367: we say that $SAT({\cal P})$ has a {\em second-order phase
1368: transition}\footnote{strictly speaking the order of the
1369: phase transition is {\em at least two}.}.
1370: \end{definition}
1371:
1372: A first observation is that a second-order phase transition
1373: has computational implications:
1374:
1375: \begin{theorem}\label{second:order}
1376: Let ${\cal P}$ be such $SAT({\cal P})$ has a second-order phase
1377: transition. Then for every constant
1378: $c>\overline{lim}_{n\goesto \infty} c_{SAT({\cal P})}(n)$, and {\em
1379: every $\alpha>0$}, random formulas of constraint density $c$
1380: have w.h.p. resolution complexity $O(2^{k\cdot \alpha \cdot n})$.
1381: \end{theorem}
1382:
1383: \begin{proof}
1384:
1385: By the analog of Claim~\ref{spine:unsat} for the general case, if
1386: $SAT({\cal P})$ has a second-order phase transition)
1387: then for every $c>c_{SAT({\cal P})}$ and for every $\alpha>0$,
1388: minimally unsatisfiable subformulas of
1389: a random formula $\Phi$ with constraint density $c$ have w.h.p. size at
1390: most $\alpha \cdot n$.
1391: Consider
1392: the backtrack tree of the natural DPLL algorithm (that tries to
1393: satisfies clauses one at a time)
1394: on such a minimally unsatisfiable subformula $F$. By the usual
1395: correspondence between DPLL trees and resolution
1396: complexity (e.g. \cite{beame:dp}, pp. 1) it yields a resolution proof
1397: of the unsatisfiability of $\Phi$
1398: having size at most $2^{k\cdot \alpha \cdot n+1}$.
1399:
1400: \end{proof}
1401: \qed
1402:
1403: \begin{definition}
1404: For a formula $F$ define $
1405: c^{*}(F)= \max\{ \frac{|Constraints(G)|}{|Var(G)|}: \emptyset \neq G
1406: \subseteq F\}$.
1407: \end{definition}
1408:
1409: The next result gives a sufficient condition for a generalized
1410: satisfiability problem
1411: to have a first-order phase transition.
1412:
1413: \begin{theorem} \label{sufficient:first-order}
1414: Let $C$ be a set of constraints such that $SAT({\cal C})$ has a sharp
1415: threshold. If
1416: there exists $\epsilon > 0$ such that for every minimally unsatisfiable
1417: formula $F$ it holds
1418: that
1419: \[
1420: c^{*}(F) > \frac{1+\epsilon}{k-1}
1421: \]
1422: then for every ${\cal P}$ with $supp({\cal P})=C$
1423:
1424: $SAT({\cal P})$ has a first-order phase transition.
1425:
1426: \end{theorem}
1427: \begin{proof}
1428:
1429: We first recall the following concept from \cite{chvatal:szemeredi:resolution}:
1430:
1431: \begin{definition}
1432: Let $x,y>0$. A $k$-uniform hypergraph with $n$ vertices is {\em ($x$,$y$)-sparse} if every set of $s\leq xn$ vertices contains at most $ys$ edges.
1433: \end{definition}
1434:
1435: We also recall Lemma 1 from the same paper.
1436: \begin{lemma}\label{sparsity:hypergraph}
1437: Let $k,c>0$ and $y$ be such that $(k-1)y>1$. Then w.h.p. the random $k$-uniform hypegraph with $n$ vertices and $cn$ edges is $(x,y)$-sparse, where
1438: \begin{eqnarray*}
1439: \epsilon = y-1/(k-1), \\
1440: x = (\frac{1}{2e}(\frac{y}{ce})^{y})^{1/\epsilon}, \\
1441: \end{eqnarray*}
1442: \end{lemma}
1443:
1444: The critical observation is that the existence of a minimally
1445: unsatisfiable formula of size $xn$ and with $c^{*}(F) > \frac{1+\epsilon}{k-1}$
1446: implies that the
1447: $k$-uniform hypergraph associated to the given formula is {\em not}
1448: $(x,y)$-sparse, for $y=
1449: \frac{\epsilon}{k-1}$.
1450:
1451: But, according to Lemma~\ref{sparsity:hypergraph}, w.h.p. a random $k$-uniform hypergraph with $cn$ edges is
1452: $(x_{0},y)$
1453: sparse, for $x_{0}=(\frac{1}{2e}(\frac{y}{ce})^{y})^{1/\epsilon}$ (a
1454: dirrect application of Lemma 1 in their paper). We infer that any formula
1455: with less than $x_{0}\cdot n /K$
1456: constraints is satisfiable, therefore the same is true for any formula
1457: with $x_{0}\cdot n/K$ clauses picked up from the clausal representation
1458: of constraints in $\Phi$.
1459:
1460: The second condition (expansion of the formula hypergraph) can be
1461: proved similarly.
1462:
1463: \end{proof}
1464: \qedbox
1465:
1466: One can give an explicitly defined class of satisfiability
1467: problems for which the previous result applies:
1468:
1469: \begin{theorem}\label{implicates:first-order}
1470: Let ${\cal P}$ be such that $SAT({\cal P})$ has a
1471: sharp
1472: threshold. If {\em no} clause $C\in {\cal C}=supp({\cal P})$ has an
1473: implicate of
1474: length
1475: at most 2 then
1476: \begin{enumerate}
1477: \item for every minimally unsatisfiable formula $F$
1478: \[
1479: c^{*}(F)\geq \frac{2}{2k-3}.
1480: \]
1481: Therefore $SAT({\cal P})$ satisfies the conditions of the previous
1482: theorem, i.e. it has a first-order
1483: phase transition.
1484: \item Moreover $SAT({\cal P})$ also has $2^{\Omega(n)}$
1485: resolution complexity\footnote{this result subsumes some of the recent results
1486: in \cite{mitchell:cp02}}.
1487: \end{enumerate}
1488: \end{theorem}
1489:
1490: \begin{proof}
1491:
1492: \begin{enumerate}
1493: \item
1494: For any real $r \geq 1$, formula $F$ and set of clauses $G\subseteq F$,
1495: define the {\em $r$-deficiency of $G$}, $\delta_{r}(G)=
1496: r|Clauses(G)|-|Vars(G)|$.
1497:
1498: Also define
1499: \begin{equation} \label{max}
1500: \delta^{*}_{r}(F)= \max\{\delta_{r}(G): \emptyset \neq G \subseteq F\}
1501: \end{equation}
1502:
1503: We claim that for any minimally unsatisfiable $F$,
1504: $\delta^{*}_{2k-3}(F)\geq 0$. Indeed,
1505: assume
1506: this was not true. Then there exists such $F$ such that:
1507: \begin{equation}\label{deff}
1508: \delta_{2k-3}(G)\leq
1509: -1\mbox{ for all }\emptyset \neq G\subseteq F.
1510: \end{equation}
1511: \begin{proposition} \label{1-transversal}
1512: Let $F$ be a formula for which condition~\ref{deff} holds. Then there
1513: exists
1514: an ordering $C_{1}, \ldots, C_{|F|}$ of constraints in $F$
1515: such that each constraint $C_{i}$ contains
1516: at least $k-2$ variables that appear in {\em no} $C_{j}$, $j<i$.
1517: \end{proposition}
1518:
1519: \begin{proof}
1520: Denote by $v_{i}$ the number of variables that appear in {\em exactly} $i$ constraints of $F$. We have
1521: \[
1522: \sum_{i\geq 1} i\cdot v_{i} = k\cdot |Constraints(F)|.
1523: \]
1524:
1525: therefore $2|Var(F)|-v_{1}\leq k\cdot |Constraints(F)|$. This can be
1526: rewritten as $v_{1}\geq 2|Var(F)|-k|Constraints(F)|> |Constraints(F)|\cdot
1527: (2k-3 - k)= (k-3)\cdot |Constraints(F)|$ (we have used the upper bound on $c^{*}(F)$. Therefore there exists at least one constraint in $F$ with at least $k-2$ variables that are free in $F$. We set $C_{|F|}=C$ and
1528: apply this argument recursively to $F\setminus C$.
1529: \end{proof}
1530: \qed
1531:
1532: Call the $k-2$ new variables of $C_{i}$ {\em free in $C_{i}$}. Call the
1533: other
1534: two variables {\em bound in $C_{i}$}. Let us show now that $F$ cannot
1535: be
1536: minimally unsatisfiable. Construct a satisfying assignment for $F$
1537: incrementally: Consider constraint $C_{j}$. At most two of the
1538: variables in
1539: $C_{j}$ are bound for $C_{j}$. Since $C$ has no implicates of size at
1540: most two,
1541: one can set the remaining variables in a way that satisfies $C_{j}$.
1542: This yields a satisfying
1543: assignment for $F$, a contradiction with our assumption that $F$ was
1544: minimally unsatisfiable.
1545:
1546: Therefore $\delta^{*}_{2k-3}(F)\geq 0$, a
1547: statement equivalent to our conclusion.
1548:
1549: \item To prove the resolution complexity lower bound we use the size-width
1550: connection for resolution complexity obtained in
1551: \cite{ben-sasson:resolution:width}:
1552: we prove that there exists $\eta >0$ such that w.h.p. random instances
1553: of $SAT({\cal P})$ having constraint density $c$ have resolution
1554: width at least $\eta \cdot n$.
1555:
1556: We use the same strategy as in \cite{ben-sasson:resolution:width}
1557: \begin{enumerate}
1558: \item prove that w.h.p. minimally unsatisfiable subformulas are ``large''.
1559: \item prove that any clause implied by a satisfiable formula of ``intermediate'' size will contain ``many'' literals.
1560: \end{enumerate}
1561:
1562: Indeed, define for a unsatisfiable formula $\Phi$ and (possibly empty) clause
1563: $C$
1564: \[
1565: \mu(C)=\min\{|\Xi|: \Xi\subseteq \Phi, \Xi \models C\}.
1566: \]
1567:
1568: \begin{claim}
1569: There exists $\eta_{1}>0$ such that for any $c>0$, w.h.p.
1570: $\mu(\Box)\geq \eta_{1}\cdot n$ (where $\Phi$ is a random instance
1571: of $SAT({\cal P})$ having constraint density $c$).
1572: \end{claim}
1573:
1574: \begin{proof}
1575: In the proof of Theorem~\ref{sufficient:first-order} we have shown that there
1576: exists $\eta_{0}>0$ such that w.h.p. any unsatisfiable subformula of a given
1577: formula has at least $\eta_{0} \cdot n$ constraints. Therefore {\em any}
1578: formula made of {\em clauses} in the CNF-representation of constraints in
1579: $\Phi$, and which has less than $\eta_{0} \cdot n$ clauses is satisfiable,
1580: and the claim follows, by taking $\eta_{1} = \eta_{0}$.
1581: \end{proof}
1582: \qed
1583:
1584: The only (slightly) nontrivial step of the proof, which critically uses the
1585: fact that constraints in ${\cal P}$ do not have implicates of length at most
1586: two, is to prove that clause implicates of subformulas of ``medium'' size
1587: have ``many'' variables. Formally:
1588:
1589: \begin{claim}\label{expansion}
1590: There exists $d>0$ and $\eta_{2}>0$ such that w.h.p., for every clause
1591: $C$ such that $d/2\cdot n <\mu(C)<=dn$, $|C|\geq \eta_{2}\cdot n$.
1592: \end{claim}
1593:
1594: \begin{proof}
1595: Take $0<\epsilon$. It is easy to see that if $c^{*}(F)<\frac{2}{2k-3+\epsilon}$ then w.h.p. for every subformula $G$ of $F$, at least $\frac{\epsilon}{3} \cdot |Constraints(G)|$ have at least $k-2$ private variables: Indeed, since
1596: $c^{*}(G)<\frac{2}{2k-3+\epsilon}$, by a reasoning similar to the one we made
1597: previously $v_{1}(G)\geq (k-3+\epsilon)|Constraints(G)|$. Since constraints in $G$ have arity $k$, at least $\epsilon/3 \cdot |Constraints(G)|$ have at
1598: least $k-2$ ``private'' variables.
1599:
1600: Choose $y=\frac{2}{2k-3+\epsilon}$ in Lemma~\ref{sparsity:hypergraph} for $\epsilon>0$ a small
1601: enough constant. Since the problem has a sharp threshold in the region where
1602: the number of clauses is linear,
1603: \[
1604: d=\inf\{x(y,c): c>= c_{SAT({\cal P})}\} >0.
1605: \]
1606:
1607: W.h.p. all subformulas of $\Phi$ having size less than
1608: $d/k \cdot n$ have a formula hypergraph that is $(x,y)$-sparse,
1609: therefore fall under the scope of the previous argument.
1610:
1611: Let $\Xi$ be a subformula of $\Phi$, having minimal size, such that
1612: $\Xi \models C$. We claim:
1613:
1614: \begin{claim}
1615: For every clause $P$ of $\Xi$ with $k-2$ ``private'' variables, (i.e. one that does not appear in any other clause), at least one of these ``private'' variables appears in $C$.
1616: \end{claim}
1617:
1618: Indeed, supppose there exists a clause $D$ of $\Xi$ such that none of its
1619: private variables appears in $C$.
1620:
1621: Because of the minimality of $\Xi$ there exists an assignment $F$ that satisfies $\Xi \setminus \{D\}$ but does not satisfy $D$ or $C$. Since $D$ has no implicates
1622: of size two, there exists an assignment $G$, that differs from $F$ only on
1623: the private variables of $D$, that satisfies $\Xi$. But since $C$ does not
1624: contain any of the private variables of $D$, $F$ coincides with $G$ on variables in $C$. The conclusion is that $G$ does not satisfy $C$, which contradicts
1625: the fact that $\Xi\models C$.
1626:
1627: The proof of Claim~\ref{expansion} (and of item 2. of Theorem~\ref{implicates:first-order}) follows: since for any clause $K$ of one of the original constraints $\mu(K)=1$, since $\mu(\Box)>\eta_{1}\cdot n$ and since w.l.o.g. $0<d<\eta_{1}$ (otherwise replace $d$ with the smaller value)
1628: there exists a clause $C$ such that
1629: \begin{equation}\label{intermediate}
1630: \mu(C)\in [d/2k \cdot n, d/k \cdot n].
1631: \end{equation}
1632:
1633: Indeed,
1634: let $C^{\prime}$ be a clause in the resolution refutation of $\Phi$ minimal with
1635: the property that $\mu(C^{\prime})> dn$. Then at least one clause $C$, of the
1636: two involved in deriving $C^{\prime}$ satisfies equation~\ref{intermediate}.
1637:
1638: By the previous claim it
1639: $C$ contains at least one ``private'' variable from each clause of $\Xi$. Therefore $|C|\geq \eta_{2}\cdot n$, with $\eta_{2}=d/2k\cdot \epsilon$.
1640:
1641: \end{proof}
1642:
1643: \end{enumerate}
1644:
1645: \end{proof}
1646: \qedbox
1647:
1648: It is instructive to note that the condition in the theorem is violated
1649: (as expected) by random 2-SAT, as well as by random 1-in-$k$ SAT: the
1650: formula $C(x_{1}, x_{2}, \ldots, x_{k-1}, x_{k})\AND
1651: C(\overline{x_{k}},
1652: x_{k+1}, \ldots , x_{2k-2}, x_{1})\AND C(\overline{x_{1}}, x_{2k-1},
1653: \ldots, x_{3k-3}, \overline{x_{k}})
1654: \AND C(x_{k}, x_{3k-2}, \ldots, x_{4k-4}, x_{1})$ (where $C$ is the
1655: constraint ``1-in-$k$'') is minimally unsatisfiable, but
1656: has clause/variable ratio $1/(k-1)$ and implicates
1657: $\overline{x_{1}} \OR \overline{x_{k}}$ and $x_{1}\OR x_{k}$.
1658:
1659:
1660: It would be tempting to speculate that whenever both $x\OR y$ and
1661: $\overline{x}\OR \overline{y}$ are implicates of clauses in ${\cal C}$
1662: then
1663: $SAT({\cal P})$ has a second-order phase transitions for every
1664: distribution ${\cal P}$ with $supp({\cal P})= {\cal C}$. That is,
1665: however, {\em
1666: not} true, at least
1667: for some distributions ${\cal P}$. Consider the random $(2+p)$-SAT
1668: model of Monasson et al.
1669: \cite{2+p:nature}. In this model $p$ is a fixed real in $[0,1]$. A
1670: random instances of $(2+p)$-SAT with $n$
1671: variables and $c\cdot n$ clauses is obtained by choosing $pcn$ random
1672: clauses of length 3 and $(1-p)cn$ random clauses of length 2.
1673: It was shown in \cite{2+p:nature} (using the nonrigorous replica
1674: method) that
1675: \begin{enumerate}
1676: \item $(2+p)$-SAT has a second-order phase transition for $0\leq p \leq
1677: p_{0}\sim 0.413...$.
1678: \item the transition becomes first-order for $p > p_{0}$.
1679: \item when the transition changes from second-order to first-order the
1680: complexity
1681: of a certain DPLL algorithm changes from polynomial to exponential.
1682: \end{enumerate}
1683:
1684:
1685:
1686: Several rigorous results have complemented these findings. Achlioptas
1687: et al. \cite{2+psat:ralcom97}
1688: have shown that for $0\leq p \leq 0.4$ the phase transition in
1689: $(2+p)$-SAT only
1690: depends on the ``2-SAT part''. One can perhaps use the techniques
1691: of \cite{scaling:window:2sat} to confirm statement (i).
1692:
1693:
1694:
1695: It is easily seen that
1696: $(2+p)$-SAT can be represented in Molloy's framework.
1697: On the other hand Achlioptas, Beame and Molloy
1698: \cite{achlioptas:beame:molloy} have shown that for
1699: those $p$ for
1700: which $(2+p)$-SAT does {\em not} behave like 2-SAT the resolution
1701: complexity of
1702: the problem is exponential. Using Theorem~\ref{second:order} and
1703: and results in \cite{achlioptas:beame:molloy} we get:
1704:
1705: \begin{theorem}\label{2+p-sat:first-order}
1706: Let $p\in [0,1]$ be s.t. there exist $\epsilon >0$ and
1707: $c<\frac{(1-\epsilon)}{1-p}$ s.t.
1708: random instances of $(2+p)$-SAT with $n$ variables and $c\cdot n$
1709: clauses are w.h.p. unsatisfiable. Then $(2+p)$-SAT
1710: has a first-order phase transition.
1711: \end{theorem}
1712:
1713:
1714: It would be interesting to obtain a complete characterization of the
1715: order of the phase
1716: transition in an arbitrary problem $SAT({\cal P})$. Such a
1717: characterization, however, requires substantial advances: Exactly locating the
1718: ``tricritical point'' $p_{0}$ in random $2+p$-SAT (or merely deciding
1719: whether it is equal or not to 0.4) is an open problem. A complete
1720: characterization would yield a solution to this problem as a byproduct.
1721:
1722: On the other hand Theorem~\ref{2+p-sat:first-order} suggests an
1723: interesting conjecture: whenever the location of the phase transition
1724: is {\em not} determined by the implicates of size at most two in the
1725: given formula, the phase transition in $SAT({\cal P})$ is first-order.
1726: Perhaps techniques in \cite{achlioptas:beame:molloy} can help settle
1727: this question.
1728:
1729:
1730:
1731: \section{Discussion}
1732:
1733: We have shown that the existence of a first-order phase transition
1734: in a random satisfiability problem is often correlated with a
1735: $2^{\Omega(n)}$ peak in the complexity of resolution/DPLL algorithms at
1736: the
1737: transition point.
1738:
1739: As for the extent of the connection it is easy to see that it does not
1740: extend to a substantially larger class of algorithms: consider random
1741: $k$-XOR-SAT, the problem of testing the satisfiability of random
1742: systems of linear equations of size $k$ over ${\bf Z}_{2}$.
1743: $k$-XOR-SAT is a version of XOR-SAT, one of the polynomial time
1744: cases of satisfiability from Schaefer's Dichotomy Theorem
1745: \cite{schaefer-dich}.
1746: Indeed,
1747: it is easily solved by Gaussian elimination. But Ricci-Tersenghi et al.
1748: \cite{zecchina:kxorsat} have presented
1749: a non-rigorous argument using the replica method that supports
1750: the existence of a first-order phase transition, and we can show
1751: this formally (as a direct consequence of
1752: Theorem~\ref{implicates:first-order}):
1753:
1754: \begin{proposition}
1755: Random $k$-XOR-SAT, $k\geq 3$, has a first-order phase transition.
1756: \end{proposition}
1757:
1758: To sum up: {\bf the intuitive argument states that a first-order phase
1759: transition correlates with a $2^{\Omega(n)}$ lower bound of the
1760: complexity
1761: of DPLL algorithms at the transition. This is true in many cases, and
1762: the underlying reason is that the two phenomena (the jump in the
1763: order parameter and the resolution complexity lower bound) have
1764: common causes. However, at least for satisfiability problems,
1765: this connection does not extend substantially beyond the class of
1766: DPLL algorithms.}
1767:
1768: \section*{Acknowledgments}
1769:
1770: I thank Madhav Marathe, Anil Kumar and Cris Moore for useful comments.
1771: In particular Cris made the observation that led to the realization
1772: that
1773: my previous results implied Theorem~\ref{second:order}.
1774:
1775: This work has been supported by the Department of Energy under contract
1776: W-705-ENG-36.
1777:
1778: %\bibliography{/u1/gistrate/bib/bibtheory}
1779: \bibliography{bibtheory}
1780:
1781:
1782:
1783: \end{document}
1784:
1785:
1786: