Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

Handbook of Satisfiability 243

Armin Biere, Marijn Heule, Hans van Maaren and Toby Walsch
IOS Press, 2008
c 2008 Dimitris Achlioptas. All rights reserved.

Chapter 8

Random Satisfiability
Dimitris Achlioptas

8.1. Introduction

Satisfiability has received a great deal of study as the canonical NP-complete prob-
lem. In the last twenty years a significant amount of this effort has been devoted
to the study of randomly generated satisfiability instances and the performance
of different algorithms on them. Historically, the motivation for studying random
instances has been the desire to understand the hardness of typical instances. In
fact, some early results suggested that deciding satisfiability is easy on average.
Unfortunately, while easy is easy to interpret, on average is not.
One of the earliest and most often quoted results for satisfiability being easy
on average is due to Goldberg [Gol79]. In [FP83], though, Franco and Paull
pointed out that the distribution of instances used in the analysis of [Gol79]
is so greatly dominated by very satisfiable formulas that if one tries truth
assignments completely at random, the expected number of trials until finding a
satisfying one is O(1). Alternatively, Franco and Paull pioneered the analysis of
random instances of k-SAT, i.e., asking the satisfiability question for random k-
CNF formulas (defined precisely below). Among other things, they showed [FP83]
that for all k 3 the DPLL algorithm needs an exponential number of steps to
report all cylinders of solutions of such a formula, or that no solutions exist.
Random k-CNF formulas. Let Fk (n, m) denote a Boolean formula in Conjunc-
tive Normal Form with m clauses over n variables, where the clauses  are chosen
uniformly, independently and without replacement among all 2k nk non-trivial
clauses of length k, i.e., clauses with k distinct, non-complementary literals.
Typically, k is a (small) fixed integer, while n is allowed to grow. In particular,
we will say that a sequence of random events En occurs with high probability
(w.h.p.) if limn Pr[En ] = 1 and with uniformly positive probability (w.u.p.p.)
if lim inf n Pr[En ] > 0. One of the first facts to be established [FP83] for
random k-CNF formulas is that Fk (n, rn) is w.h.p. unsatisfiable for r 2k ln 2.
A few years later, Chao and Franco [CF86, CF90] showed that, in contrast, if
r < 2k /k, a very simple algorithm will find a satisfying truth assignment w.u.p.p.
Combined, these two facts established m = (n) as the most interesting regime
for considering the satisfiability of random k-CNF formulas.
244 Chapter 8. Random Satisfiability

The study of random k-CNF formulas took off in the late 1980s. In a cel-
ebrated result, Chvatal and Szemeredi [CS88] proved that random k-CNF for-
mulas w.h.p. have exponential resolution complexity, implying that if F is a
random k-CNF formula with r 2k ln 2, then w.h.p. every DPLL-type algo-
rithm needs exponential time to prove its unsatisfiability. A few years later,
Chvatal and Reed [CR92] proved that random k-CNF formulas are satisfiable
w.h.p. for r = O(2k /k), strengthening the w.u.p.p. result of [CF90]. Arguably,
the biggest boost came from the experimental work of Mitchell, Selman and
Levesque [MSL92] and the analysis of these experiments by Kirkpatrick and Sel-
man in [KS94].
In particular, [SML96] gave extensive experimental evidence suggesting that
for k 3, there is a range of the clauses-to-variables ratio, r, within which it seems
hard even to decide if a randomly chosen k-SAT instance is satisfiable or not (as
opposed to reporting all cylinders of solutions or that no solutions exist). The
analysis of these experiments using finite-size scaling methods of statistical physics
in [KS94] showed that this peak in experimental decision complexity coincided
with a precipitous drop in the probability that a random formula is satisfiable.
For example, for k = 3 the analysis drew the followng remarkable picture: for
r < 4, a satisfying truth assignment can be easily found for almost all formulas;
for r > 4.5, almost all formulas are unsatisfiable; for r 4.2, a satisfying truth
assignment can be found for roughly half the formulas and around this point
the computational effort for finding a satisfying truth assignment, whenever one
exists, is maximized.
This potential connection between phase transitions and computational com-
plexity generated a lot of excitement. A particularly crisp question is whether the
probability of satisfiability indeed exhibits a sharp threshold around some critical
density (ratio of clauses to variables).
Satisfiability Threshold Conjecture. For every k 3, there exists a constant
rk > 0 such that,
(
1, if r < rk
lim Pr[Fk (n, rn) is satisfiable] =
n 0, if r > rk .

By now, the Satisfiability Threshold Conjecture has attracted attention in


computer science, mathematics, and statistical physics. It has stimulated numer-
ous interesting results and is at the center of an area of intense research activity,
some of which we will survey in the following sections. As we will see, recent
results have both strengthened the connection between computational complex-
ity and phase transitions and shown that it is far more nuanced than originally
thought. In particular, to the extent that finding satisfying assignments in ran-
dom formulas is hard, this hardness comes from phase transitions in the solution-
space geometry of the formulas, not in their probability of satisfiability. But this
is getting ahead of our story.
To conclude this introduction we note that while a number of other generative
models have been proposed for random SAT instances over the years (see [Fra01]
for an excellent survey) random k-SAT is by far the dominant model. One reason
is that random k-CNF formulas enjoy a number of intriguing mathematical prop-
Chapter 8. Random Satisfiability 245

erties. Another is that random k-SAT instances remain computationally hard for
a certain range of densities, making them a popular benchmark for testing and
tuning satisfiability algorithms. In fact, some of the better practical ideas in use
today come from insights gained by studying the performance of algorithms on
random k-SAT instances [GSCK00].
The rest of this chapter is divided into two major pieces. The first piece is
an overview of the current state of the art regarding mathematical properties of
random k-CNF formulas where the number of clauses m = (n). The second
piece concerns the performance of different algorithms on sparse random k-CNF
formulas. Both pieces are concerned with (and limited to) rigorous mathematical
results. For experimental algorithmic results we refer the reader to the survey by
Cook and Mitchell [CM97]. For results using the methods of statistical physics
we refer the reader to Part 1, Chapter 9.

8.2. The State of the Art

8.2.1. The Satisfiability Threshold Conjecture

The satisfiability threshold conjecture remains open. A big step was made by
Friedgut [Fri99] who showed the existence of a sharp threshold for satisfiability
around a critical sequence of densities (rather than a single density rk ).
Theorem 1 ([Fri99]). For every k 3, there exists a sequence rk (n) such that
for any  > 0,
(
1, if m = (rk (n) )n
lim Pr[Fk (n, m) is satisfiable] =
n 0, if m = (rk (n) + )n.

While it is widely believed that rk (n) rk , a proof remains elusive. A useful


corollary of Friedguts theorem is the following, as it allows one to give lower
bounds for rk by only establishing satisfiability w.u.p.p. rather than w.h.p.
Corollary 2. If Fk (n, r n) is satisfiable w.u.p.p., then for all r < r , Fk (n, rn)
is satisfiable w.h.p.
Although the existence of rk has not been established, we will allow ourselves
the notational convenience of writing rk r to denote that for r > r , Fk (n, rn)
is w.h.p. unsatisfiable (and analogously for rk r ). In this notation, the current
best bounds for the location of the satisfiability threshold are, roughly,

2k ln 2 (k) rk 2k ln 2 (1) .

The precise form of these bounds is given by Theorem 3, which we illustrate for
some small values of k in Table 8.1.
Theorem 3. There exists sequences k , k 0 such that for all k 3,
ln 2 1 + ln 2
2k ln 2 (k + 1) 1 k rk 2k ln 2 + k .
2 2
246 Chapter 8. Random Satisfiability

It is very easy to show that rk 2k ln 2. This is because the probability


that Fk (n, m) is satisfied by at least one assignment is trivially bounded by 2n
times the probability that it is satisfied by any particular assignment, which, in
turn, is bounded by (1 2k )m (indeed, this last expression is exact if we assume
that clauses are chosen with replacement). Since 2(1 2k )r < 1 for r 2k ln 2,
this implies that for all such r the probability of satisfiability is exponentially
small. The sharpening of this upper bound for general k is due, independently,
to Dubois and Boufkhad [DB97] and Kirousis et al. [KKKS98]. It corresponds
to bounding the probability of satisfiability by 2n times the probability that a
particular assignment is a locally maximum satisfying assignment, i.e., one in
which no 0 can be turned into 1 without violating satisfiability (it is clear that
every satisfiable formula has at least one such satisfying assignment, e.g., the
lexicographically greatest one).
The lower bound in Theorem 3 is due to Achlioptas and Peres [AP04] and
was proven via the, so-called, second moment method. It corresponds to the
largest density for which Fk (n, m) w.h.p. has balanced satisfying assignments, i.e.,
satisfying assignments in which the number of satisfied literals is km/2 + O(n).
In other words, balanced satisfying assignments have the property that in spite
of being satisfying their number of satisfied literals is like that of a uniformly
random {0, 1}n . Focusing on balanced assignments is done of technical
necessity for the second moment method to work (we discuss this point in detail
in Section 8.6.3) and there does not appear to be an inherent reason for doing so.
Indeed, as k grows, the upper bound of Theorem 3 coincides with the presults1 for
the location of the threshold (up to a o(1) term in k), giving even more evidence
that the O(k) term in the lower bound is an artifact of the analysis.
The second moment method, used to prove the lower bound, ignores individ-
ual solutions and captures, instead, statistical properties of the entire solution
space. As such, it offers no guidance whatsoever on how to efficiently find satisfy-
ing assignments for those densities for which it establishes their existence. Indeed,
as we will see shortly, no efficient algorithm is known that can find solutions be-
yond density O(2k /k), even w.u.p.p. In the third row of Table 8.1 we indicate
this phenomenon by giving the largest densities for which any efficient algorithm
is known to succeed [KKL06, FS96]).
Finally, we note that while both general bounds above extend to k = 3, better
bounds exist for that case. In particular, the lower bound for r3 in Table 8.1 is
actually algorithmic and due, independently, to Hajiaghayi and Sorkin [HS03] and
Kaporis, Kirousis and Lalas [KKL06]. We discuss its derivation in Section 8.7.2.
The upper bound is due to Dubois, Boufkhad and Mandler [DBM03] and uses
the idea of locally maximum assignments mentioned above in combination with
conditioning on the literal degree sequence, thus curtailing certain large deviations
contributions that inflate the unconditional expectation.

1 We use the term presults as a way of referring to p[hysics] results in this area, the

subject of Part 1, Chapter 9. In general, presults rest on combining mathematically rigorous


arguments with highly non-trivial unproven assumptions, the latter often informed by general
considerations of statistical physics. The term is also motivated by the fact that many presults
eventually became rigorous mathematical results via proofs that were deeply informed by the
physical arguments.
Chapter 8. Random Satisfiability 247

Table 8.1. Best known rigorous bounds for the location of the satisfiability threshold for some
small values of k. The last row gives the largest density for which a polynomial-time algorithm
has been proven to find satisfying assignments.
k 3 4 5 7 10 20
Best upper bound 4.51 10.23 21.33 87.88 708.94 726, 817
Best lower bound 3.52 7.91 18.79 84.82 704.94 726, 809
Algorithmic lower bound 3.52 5.54 9.63 33.23 172.65 95, 263

8.3. Random MAX k-SAT

The methods used to derive bounds for the location of the satisfiability threshold,
also give bounds for the fraction of clauses that can be satisfied above it. Specif-
ically, let us say that a k-CNF formula with m clauses is p-satisfiable,  for some
p [0, 1], if there exists an assignment satisfying at least 1 1p 2k
m clauses.
Observe that every k-CNF formula is 0-satisfiable, since the average number of
satisfied clauses over all assignments is (1 2k )m, while satisfiability is simply
1-satisfiability. Thus, given p (0, 1], the relevant quantity is the largest density,
rk (p), for which a random k-CNF formula is w.h.p. p-satisfiable. Analogously
to the case p = 1, the naive union bound over {0, 1}n for the existence of a
p-satisfying truth assignment implies rk (p) Tk (p), where
2k ln 2
Tk (p) = .
p + (1 p) ln(1 p)
Note that since limp1 Tk (p) = 2k ln 2, this recovers the naive satisfiability up-
per bound. Using the second moment method, in [ANP07] it was shown that
asymptotically, this upper bound is tight up to second order terms.
Theorem 4. There exists a sequence k = O(k2k/2 ), such that for all k 2
and p (0, 1],
(1 k ) Tk (p) rk (p) Tk (p) .
Algorithms fall far short from this bound. Analogously to satisfiability, the
best known algorithm [CGHS04] finds p-satisfying assignments only for r =
O(Tk (p)/k).

8.3.1. The case k [2, 3)

For k = 2, satisfiability can be decided in polynomial time using a very simple


method: tentatively set any unset variable v to 0; repeatedly satisfy any 1-clauses
that result; if a 0-clause is generated set v permanently to 1, otherwise set it per-
manently to 0. A 2-CNF formula is satisfiable iff when this process terminates
no 0-clauses are present. For random 2-CNF formulas, the trees of implications
generated by this process mimick the connected component structure of random
digraphs. Using this connection, Chvatal and Reed [CR92], Goerdt [Goe96] and
Fernandez de la Vega [FdlV92] independently proved r2 = 1. Later, in [BBC+ 01],
Bollobas et al. [BBC+ 01], also using this connection, determined the scaling win-
dow for random 2-SAT, showing that the transition from satisfiability to unsat-
isfiability occurs for m = n + n2/3 as goes from to +.
248 Chapter 8. Random Satisfiability

In [MZK+ , MZK+ 99a, MZ98, MZK+ 99b], Monasson et al., using mathemati-
cally sophisticated but non-rigorous techniques of statistical physics, initiated the
analytical study of random CNF formulas that are mixtures of 2- and 3-clauses,
which they dubbed (2 + p)-CNF. Such formulas arise for a number of reasons. For
example, a frequent observation when converting problems from other domains
into satisfiability problems is that they result into mixed CNF formulas with a
substantial number of clauses of length 2, along with the clauses of length 3.
Another reason is that DPLL algorithms run by recursively solving satisfiability
on residual formulas, restricted versions of their input CNF formula, which are
mixtures of clauses of length at least 2. When given random 3-CNF formulas
as input, many DPLL algorithms produce residual formulas that are mixtures of
random 2- and 3-clauses, making properties of random (2 + p)-CNF crucial for
analyzing their running time.
A random (2 + p)-CNF on n variables with m clauses is formed by choosing
pm clauses of length 3 and (1p)m clauses of length 2, amongst all clauses of each
length, uniformly and independently. Thus, p = 0 corresponds to random 2-SAT,
while p = 1 corresponds to random 3-SAT. Below we will find it convenient to
sidestep this original formulation and state results directly in terms of the number
of 2- and 3-clauses, not of the total number of clauses m and p.
The fact r2 = 1 implies that for any  > 0 a random 2-CNF formula on n
variables with (1 /2)n clauses is w.h.p. satisfiable, but adding n 2-clauses to it
w.h.p. results in an unsatisfiable formula. The presults in [MZK+ 99b] suggested,
rather remarkably, that if instead of adding n random 2-clauses one adds up to
0.703...n random 3-clauses, the formula remains satisfiable. In other words, that
there is no finite exchange rate between 2- and 3-clauses.
Inspired by these claims, Achlioptas et al. [AKKK01] proved that
Theorem 5. A random CNF formula with n variables, (1)n 2-clauses and n
3-clauses is w.h.p. satisfiable for all  > 0 and all 2/3, but w.h.p. unsatisfiable
for  = 0.001 and = 2.28.
In other words, the physical prediction of an infinite exchange ratio is valid
as one can add at least 0.66n 3-clauses (and no more than 2.28n). In [Ach99] it
was conjectured that the inequality 2/3 in Theorem 5 is tight. That is,
Conjecture 6. For all > 2/3, there exists  = () > 0, such that a random
CNF formula with n variables, (1 )n 2-clauses and n 3-clauses is w.h.p.
unsatisfiable.
Conjecture 6 is supported by a presult of Biroli, Monasson and Weigt [BMW00],
subsequent to [MZK+ , MZK+ 99a], asserting that 2/3 is indeed tight. As we will
see, if true, it implies that the running time of a large class of DPLL algorithms
exhibits a sharp threshold behavior: for each algorithm A, there exists a critical
density rA rk such that A takes linear time for r < rA , but exponential time
for r > rA .

8.3.2. Proof Complexity and its Implications for Algorithms

We saw that sparse random k-CNF formulas are hard to prove unsatisfiable us-
ing resolution [CS88]. Ben-Sasson and Impagliazzo [BSI99] and Alekhnovich and
Chapter 8. Random Satisfiability 249

Razborov [AR01] proved that the same is true for the polynomial calculus, while
Alekhnovich [Ale05] proved its hardness for k-DNF resolution. Random k-CNF
formulas are believed to be hard for other proof systems also, such as cutting
planes, and this potential hardness has been linked to hardness of approxima-
tion [Fei02]. Moreover, the hardness of proving their unsatisfiability has been
explored for dense formulas, where for sufficiently high densities it provably dis-
appears [BKPS02, FGK05, GL03, COGLS04, COGL07]. For a more general
discussion of the connections between k-CNF formulas and proof-complexity see
Chapter ??. Here we will focus on the resolution complexity of random k-CNF
formulas as it has immediate and strong implications for the most commonly used
satisfiability algorithm, namely the DPLL procedure.

8.3.2.1. Resolution Complexity of k-CNF formulas


The resolution rule allows one to derive a clause (A B) from two clauses (A x)
and (B x). A resolution derivation of a clause C from a CNF formula F is a
sequence of clauses C1 , . . . , C` = C such that each Ci is either a clause of F or
follows from two clauses Cj , Ck for j, k < i using the resolution rule. A resolution
refutation of an unsatisfiable formula F is a resolution derivation of the empty
clause. The size of a resolution refutation is its number of clauses.
In contrast, the Davis-Putnam/DLL (DPLL) algorithm on a CNF formula
F performs a backtracking search for a satisfying assignment of F by extending
partial assignments until they either reach a satisfying assignment or violate a
clause of F . It is well known that for an unsatisfiable formula F , the tree of nodes
explored by any DPLL algorithm can be converted to a resolution refutation of
F where the pattern of inferences forms the same tree.
For random k-CNF formulas in the unsatisfiable regime, the behavior of
DPLL algorithms, and the more general class of resolution-based algorithms,
is well-understood. Specifically, since every unsatisfiable 2-CNF formula has a
linear-size resolution refutation, if r > 1 then even the simplest DPLL algorithms
w.h.p. run in polynomial time on a random 2-CNF formula. On the other hand,
for k 3 the aforementioned result of Chvatal and Szemeredi [CS88] asserts that
w.h.p. a random k-CNF formula in the unsatisfiable regime requires an exponen-
tially long resolution proof of unsatisfiability. More precisely, let res(F ) be the
size of the minimal resolution refutation a formula F (assume res(F ) = if F is
satisfiable). In [CS88] it was proved that

Theorem 7. For all k 3 and any constant r > 0, w.h.p. res(Fk (n, rn)) = 2(n) .

Corollary 8. Every DPLL algorithm w.h.p. takes exponential time on Fk (n, rn),
for any constant r 2k ln 2.

8.3.2.2. (2 + p)-CNF formulas and their Algorithmic Implications


Since all unsatisfiable 2-CNF formulas have linear-size resolution refutations and
r2 = 1, it follows that adding (1 + )n random 2-clauses to a random 3-CNF
formula w.h.p. causes its resolution complexity to collapse from exponential to
linear. In [ABM04b], Achlioptas, Beame and Molloy proved that, in contrast,
250 Chapter 8. Random Satisfiability

adding (1 )n random 2-clauses w.h.p. has essentially no effect on its proof
complexity.

Theorem 9. For any constants r,  > 0, let F be a random formula with n


variables, (1 )n 2-clauses and rn 3-clauses. W.h.p. res(F ) = 2(n) .

Theorem 9 allows one to readily prove exponential lower bounds for the run-
ning times of DPLL algorithms for satisfiable random k-CNF formulas. This is
because many natural DPLL algorithms when applied to random k-CNF formulas
generate at least one unsatisfiable subproblem consisting of a random mixture of
2- and higher-length clauses, where the 2-clauses alone are satisfiable. (We will
discuss how, when and for which algorithms this happens in greater detail in Sec-
tion 8.9.) By Theorem 9, such a mixture has exponential resolution complexity
(converting k-clauses with k > 3 to 3-clauses arbitrarily can only reduce resolu-
tion complexity) and, as a result, to resolve any such subproblem (and backtrack)
any DPLL algorithm needs exponential time.

8.4. Physical Predictions for Solution-space Geometry

Random k-SAT, along with other random Constraint Satisfaction Problems, such
as random graph coloring and random XOR-SAT have also been studied system-
atically in physics in the past two decades. For a general introduction and ex-
position to the physical methods see Part 1, Chapter 9. In particular, motivated
by ideas developed for the study of materials known as spin glasses, physicists
have put forward a wonderfully complex picture about how the geometry of the
set of satisfying assignments of a random k-CNF formula evolves as clauses are
added. Perhaps the most detailed and sophisticated version of this picture comes
from [KMRT+ 07].
Roughly speaking, statistical physicists have predicted (using non-rigorous
but mathematically sophisticated methods) that while for low densities the set of
satisfying assignments forms a single giant cluster, at some critical density this
cluster shatters into exponentially many clusters, each of which is relatively tiny
and far apart from all other clusters. These clusters are further predicted to be
separated from one another by huge energy barriers, i.e., every path connecting
satisfying assignments in different clusters must pass through assignments that
violate (n) constraints. Moreover, inside each cluster the majority of variables
are predicted to be frozen, i.e., take the same value in all solutions in the cluster;
thus getting even a single frozen variable wrong requires traveling (n) away and
over a huge energy barrier to correct it.
The regime of exponentially many tiny clusters, each one having constant
probability of vanishing every time a clause is added (due to its frozen variables),
is predicted in [KMRT+ 07] to persist until very close to the threshold, namely
for densities up to
3 ln 2
2k ln 2 + o(1) , (8.1)
2
where the o(1) term is asymptotic in k. In comparison, the satisfiability threshold
Chapter 8. Random Satisfiability 251

is predicted [MMZ06] to occur at

1 + ln 2
2k ln 2 + o(1) , (8.2)
2
i.e., fewer than 0.2n k-clauses later. For densities between (8.1) and (8.2) , it is
predicted that nearly all satisfying assignments lie in a small (finite) number of
(atypically large) clusters, while exponentially many clusters still exist.

8.4.1. The Algorithmic Barrier

Perhaps the most remarkable aspect of the picture put forward by physicists is
that the predicted density for the shattering of the set of solutions into exponen-
tially many clusters scales, asymptotically in k, as

2k
ln k , (8.3)
k
fitting perfectly with the fact that all known algorithms fail at some density

2k
cA , (8.4)
k
where the constant cA depends on the algorithm. We will refer to this phe-
nomenon as the algorithmic barrier. We note that so far there does not appear
to be some natural general upper bound for cA for the types of algorithms that
have been analyzed, i.e., more sophisticated algorithms do have a greater con-
stant, but with rapidly diminishing returns in terms of their complexity.

8.4.2. Rigorous Results for Solution-space Geometry

By now a substantial part of the picture put forward by physicists has been made
rigorous, at least for k 8. Success for smaller k stumbles upon the fact that
the rigorous results rely on the second moment method which does not seem to
perform as well for small k.
For the definitions in the rest of this section we assume we are dealing with
an arbitrary CNF formula F defined over variables X = x1 , . . . , xn , and we let
S(F ) {0, 1}n denote its set of satisfying assignments. We say that two assign-
ments are adjacent if their Hamming distance is 1 and we let HF : {0, 1}n N
be the function counting the number of clauses of F violated by each {0, 1}n .

Definition 10. The clusters of a formula F are the connected components of


S(F ). A region is a non-empty union of clusters. The height of any path
0 , 1 , . . . , t {0, 1}n is maxi H(i ).

One can get an easy result regarding the solution-space geometry of very
sparse random formulas by observing that if a formula F is satisfiable by the
pure literal rule alone, then the set S(F ) is connected (recall that the pure literal
rule amounts to permanently satisfying any literal ` whose complement does not
appear in the formula and permanently removing all clauses containing `; the
252 Chapter 8. Random Satisfiability

proof of connectivity is left as an exercise). The pure literal rule alone w.h.p.
finds satisfying assignments in random k-CNF formulas with up to k n clauses,
for some k 0, so this is very far away from the densities up to which the best
known algorithms succeed, namely O(2k /k).
The following theorem of Achlioptas and Coja-Oghlan [ACO08], on the other
hand, asserts that for all k 8, precisely at the asymptotic density predicted
by physics in (8.3), the set of satisfying assignments shatters into an exponential
number of well-separated regions. Given two functions f (n), g(n), let us write
f g if = limn f (n)/g(n) = 1. Recall that the lower bound for the location
of the satisfiability threshold provided by the second moment method is sk
2k ln 2 (k + 1) ln22 1 o(1) 2k ln 2.

Theorem 11. For all k 8, there exists ck < sk such that for all r (ck , sk ),
the set S(Fk (n, rn)) w.h.p. consists of exponentially many regions where:
1. Each region only contains an exponentially small fraction of S.
2. The Hamming distance between any two regions is (n).
3. Every path between assignments in distinct regions has height (n).
In particular, ck ln k 2k /k .

The picture of Theorem 11 comes in even sharper focus for large k. In partic-
ular, for sufficiently large k, sufficiently close to the threshold, the regions become
arbitrarily small and maximally far apart, while still exponentially many.

Theorem 12. For any 0 < < 1/3, fix r = (1 )2k ln 2, and let

1 1 5
k = , k = , k = 3k 2 .
k 2 6 2
For all k k0 (), w.h.p. the set S = S(Fk (n, rn)) consists of 2k n regions where:
1. The diameter of each region is at most k n.
2. The distance between every pair of regions is at least k n.

Thus, the solution-space geometry becomes like that of an correcting code


with a little bit of fuzz around each codeword. This shattering phase transi-
tion is known as a dynamical transition, while the transition in which nearly all
assignments concentrate on an finite number of clusters is known as a condensa-
tion transition. This dynamical phase transition has strong negative repercussions
for the performance of random-walk type algorithms on random k-CNF formulas
and exploring them precisely is an active area of research.

Remark 13. The presults of [KMRT+ 07] state that the picture of Theorem 11
should hold for all k 4, but not for k = 3. In particular for k = 3, the solution-
space geometry is predicted to pass from a single cluster to a condensed phase
directly.

It turns out that after the set of satisfying assignments has shattered into
exponentially many clusters, we can look inside these clusters and make some
statements about their shape. For that, we need the following definition.
Chapter 8. Random Satisfiability 253

Definition 14. The projection of a variable xi over a set of assignments C,


denoted as i (C), is the union of the values taken by xi over the assignments in
C. If i (C) 6= {0, 1} we say that xi is frozen in C.
Theorem 15 below [ACO08] asserts that the dynamical transition is followed,
essentially immediately, by the massive appearance of frozen variables.
Theorem 15. For all k 8, there exists dk < sk such that for all r (dk , sk ), a
random S(Fk (n, rn)) w.h.p. has at least k n frozen variables, where k 1
for all r (dk , sk ). In particular, dk ln k 2k /k .
The first rigorous analysis of frozen variables by Achlioptas and Ricci-Tersenghi [ART06]
also establishes that
Corollary 16. For every k 9, there exists r < sk such that w.h.p. every
cluster of Fk (n, rn) has frozen variables.
It remains open whether frozen variables exist for k < 8.

8.5. The Role of the Second Moment Method

Recall that the best upper bound for the satisfiability threshold scales as (2k ),
whereas all efficient algorithms known work only for densities up to O(2k /k). To
resolve whether this gap was due to a genuine lack of solutions, as opposed to
the difficulty of finding them, Achlioptas and Moore [AM06] introduced an ap-
proach which allows one to avoid the pitfall of computational complexity: namely,
using the second-moment method one can prove that solutions exist in random
instances, without the need to identify any particular solution for each instance
(as algorithms do). Indeed, if random formulas are genuinely hard at some densi-
ties below the satisfiability threshold, then focusing on the existence of solutions
rather than their efficient discovery is essential: one cannot expect algorithms to
provide accurate results on the thresholds location; they simply cannot get there!
Before we delve into the ideas underlying some of the results mentioned above
it is helpful to establish the probabilistic equivalence between a few different
models for generating random k-CNF formulas.

8.6. Generative models

Given a set V of n Boolean variables, let C  k = Ck (V ) denote the set of all


proper k-clauses on V , i.e., the set of all 2k nk disjunctions of k literals involving
distinct variables. As we saw, a random k-CNF formula Fk (n, m) is formed by
selecting a uniformly random m-subset of Ck . While Fk (n, m) is perhaps the most
natural model for generating random k-CNF formulas, there are a number of slight
variations of the model, largely motivated by their amenability to calculations.
For example, it is fairly common to consider the clauses as ordered k-tuples
(rather than as k-sets) and/or to allow replacement in sampling the set Ck .
Clearly, for properties such as satisfiability the issue of ordering is irrelevant.
Moreover, as long as m = O(n), essentially the same is true for the issue of re-
placement. To see that, observe that w.h.p. the number, q of repeated clauses is
254 Chapter 8. Random Satisfiability

o(n), while the set of m q distinct clauses is a uniformly random (m q)-subset


of Ck . Thus, if a monotone decreasing property (such as satisfiability) holds with
probability p for a given m = r n when replacement is allowed, it holds with
probability p o(1) for all r < r when replacement is not allowed.
The issue of selecting the literals of each clause with replacement (which
might result in some improper clauses) is completely analogous. That is, the
probability that a variable appears more than once in a given clause is at most
k 2 /n = O(1/n) and hence w.h.p. there are o(n) improper clauses. Finally, we
note that by standard techniques, e.g., see [FS96], results also transfer between
Fk (n, m) and the model where every clause in Ck is included in the formula
independently of all others with probability p, when pCk m.

8.6.1. The Vanilla Second Moment Method

The second moment method approach [ANP05] rests on the following basic fact:
every non-negative random variable X satisfies Pr[X > 0] E[X]2 /E[X 2 ]. Given
any k-SAT instance F on n variables, let X = X(F ) be its number of satisfying
assignments. By computing E[X 2 ] and E[X]2 for random formulas with a given
density r one can hope to get a lower bound on the probability that X > 0, i.e.,
that Fk (n, rn) is satisfiable. Unfortunately, this direct application fails dramati-
cally as E[X 2 ] is exponentially (in n) greater than E[X]2 for every density r > 0.
Nevertheless, it is worth going through this computation below as it points to
the source of the problem and also helps in establishing the existence of clusters.
We perform all computations below in the model where clauses are chosen with
replacement from Ck .
For a k-CNF formula with m clauses chosen independently with replacement
it is straightforward to show that the number of satisfying assignments X satisfies
n
 
2 n
X
n
E[X ] = 2 fS (z/n)m , (8.5)
z=0
z

where fS () = 121k +2k k is the probability that two fixed truth assignments
that agree on z = n variables, both satisfy a randomly drawn clause. Thus,
(8.5) decomposes the second moment of X into the expected number of pairs of
satisfying assignments at each possible distance.
Observe that f is an increasing function of and that fS (1/2) = (1 2k )2 ,
i.e.,truth assignments at distance n/2 are uncorrelated. Using the approximation
n 1 n
n = ( (1 ) ) poly(n) and letting

2 fS ()r
S () =
(1 )1

we see that
 n
E[X 2 ] = max S () poly(n) ,
01

E[X] = S (1/2)n .
2
Chapter 8. Random Satisfiability 255

Therefore, if there exists some 6= 1/2 such that S () > S (1/2), then the
second moment is exponentially greater than the square of the expectation and we
only get an exponentially small lower bound for Pr[X > 0]. Put differently, unless
the dominant contribution to E[X 2 ] comes from uncorrelated pairs of satisfying
assignments, i.e., pairs with overlap n/2, the second moment method fails.
Unfortunately, this is precisely what happens for all r > 0 since the entropic
factor E() = 1/( (1 )1 ) in S is symmetric around = 1/2, while fS
is increasing in (0, 1), implying 0S (1/2) > 0. That is, S is maximized at some
> 1/2 where the correlation benefit balances the penalty of decreased entropy.
While the above calculation does not give us what we want, it is still useful.
For example, for any real number [0, 1], it would be nice to know the number of
pairs of satisfying truth assignments that agree on z = n variables in a random
formula. Each term in the sum in (8.5) gives us the expected number of such
pairs. While this expectation may overemphasize formulas with more satisfying
assignments (as they contribute more heavily to the expectation), it still gives
valuable information on the distribution of distances among truth assignments in
a random formula. For example, if for some values of z (and k, r) this expectation
tends to 0 with n, we can readily infer that w.h.p. there are no pairs of truth
assignments that agree on z variables in Fk (n, rn). This is because for any integer-
valued random variable Y , Pr[Y > 0]] E[Y ]. Indeed, this simple argument is
enough to provide the existence of clustering for all k 8, at densities in the
(2k ) range (getting down to (2k /k) ln k requires a lot more work).

8.6.2. Proving the Existence of Exponentially Many Clusters

To prove the existence of exponentially many regions one divides a lower bound for
the total number of satisfying assignments with an upper bound for the number
of truth assignments in each region. The lower bound comes from the expected
number of balanced satisfying assignments since the success of the second moment
method for such assignments, which we will see immediately next, implies that
w.h.p. the actual number of balanced assignments is not much lower than its
expectation. For the upper bound, one bounds the total number of pairs of truth
assignments in each region as poly(n) g(k, r)n , where
g(k, r) = max S (, k, r) ,
[0,]

where is the smallest number such that S (, k, r) < 1, i.e., n is a bound


on the diameter of any region. Dividing the lower bound for the total number of
satisfying assignments with the square root of this quantity yields the existence
of exponentially many clusters.

8.6.3. Weighted second moment: the importance of being balanced

An attractive feature of the second moment method is that we are free to apply it
to any random variable X = X(F ) such that X > 0 implies that F is satisfiable.
With this in mind, let us consider random variables of the form
XY
X= w(, c)
c
256 Chapter 8. Random Satisfiability

where w is some arbitrary function. (Eventually, we will require that w(, c) = 0


if falsifies c.) Similarly to (8.5), it is rather straightforward to prove that
n  
2 n
X n
E[X ] = 2 fw (z/n)m ,
z=0
z

where fw (z/n) = E[w(, c) w(, c)] is the correlation, with respect to a single
random clause c, between two truth assignments and that agree on z variables.
It is also not hard to see that fw (1/2) = E[w(, c)]2 , i.e., truth assignments at
distance n/2 are uncorrelated for any function w. Thus, arguing as in the previous
section, we see that E[X 2 ] is exponentially greater than E[X]2 unless fw0 (1/2) = 0.
At this point we observe that since we are interested in random formulas
where literals are drawn uniformly, it suffices to consider functions w such that:
for every truth assignment and every clause c = `1 `k , w(, c) = w(v),
where vi = +1 if `i is satisfied under and 1 if `i is falsified under . (So, we
will require that w(1, . . . , 1) = 0.) Letting A = {1, +1}k and differentiating
fw , yields the geometric condition
X
fw0 (1/2) = 0 w(v)v = 0 . (8.6)
vA

The condition in the r.h.s. of (8.6) asserts that the vectors in A, when scaled
by w(v), must cancel out. This gives us another perspective on the failure of
the vanilla second moment method: when w = wS is the indicator variable for
satisfiability, the condition in the right hand side of (8.6) does not hold since the
vector (1, 1, . . . , 1) has weight 0, while all other v A have weight 1.
Note that the r.h.s. of (8.6) implies that in a successful w each coordinate
must have mean 0, i.e., that each literal must be equally likely to be +1 or 1
when we pick truth assignments with probability proportional to their weight
under w. We will call truth assignments with km/2 O( m) satisfied literal
occurrences balanced. As was shown in [AP04], if X is the number of balanced
satisfying assignments then E[X 2 ] < CE[X]2 , for all r sk , where C = C(k) > 0
is independent of n. Thus, Pr[X > 0] 1/C and by Corollary 2 we get rk sk .
To gain some additional intuition on the success of balanced assignments it
helps to think of Fk (n, m) as generated in two steps: first choose the km lit-
eral occurrences randomly, and then partition them randomly into k-clauses. At
the end of the first step, truth assignments that satisfy many literal occurrences
cleary have significantly greater conditional probability of eventually being sat-
isfying assignments. But such assignments are highly correlated with each other
since in order to satisfy many literal occurrences they tend to agree with the
majority truth assignment on more than half the variables. Focusing on balanced
assignments only, curbs the tendency of satisfying assignments to lean towards
the majority vote assignment.
Finally, we note that it is not hard to prove that for r sk + O(1), w.h.p.
Fk (n, rn) has no satisfying truth assignments that only satisfy km/2 + o(n) literal
occurrences. Thus, any asymptotic improvement over the lower bound for rk pro-
vided by balanced assignments would mean that tendencies toward the majority
assignment become essential as we approach the threshold. By the same token,
Chapter 8. Random Satisfiability 257

we note that as k increases the influence exerted by the majority vote assign-
ment becomes less and less significant as most literals occur very close to their
expected kr/2 times. As a result, as k increases, typical satisfying assignments
get closer and closer to being balanced, meaning that the structure of the space
of solutions for small values of k (e.g., k = 3, 4) might be significantly different
from the structure for large values of k, something also predicted by the physical
methods.

8.7. Algorithms

For the purposes of this chapter we will categorize satisfiability algorithms into
two broad classes: DPLL algorithms and random walk algorithms. Within the for-
mer, we will distinguish between backtracking and non-backtracking algorithms.
More concretely, the DPLL procedure for satisfiability is as follows:

DPLL(F )
1. Repeatedly satisfy any pure literals and 1-clauses.
If the resulting formula F 0 is empty, exit reporting satisfiable.
If a contradiction (0-clause) is generated, exit.
2. Select a variable x F 0 and a value v for x
0
3. DPLL(Fv=x )
0
4. DPLL(Fv=1x )
Clearly, different rules for performing Step 2 give rise to different algorithms
and, in practice, the complexity of these rules can vary from minimal to huge.
Perhaps the simplest possible rule is to consider the variables in a fixed order, e.g.,
x1 , x2 , . . . , xn , always selecting the first variable in the order that is present in the
formula, and always setting x to the same value, e.g., x = 0. If one forgoes the
pure literal rule, something which simplifies the mathematical analysis on random
formulas, the resulting algorithms in known as ordered dll. If one also forgoes
backtracking, the resulting algorithm is known as uc, for Unit Clause propagation,
and was one the first algorithms to be analyzed on random formulas.

8.7.1. Random walk algorithms

In contrast to DPLL algorithms, random walk algorithms always maintain an


entire assignment which they evolve until either it becomes satisfying or the al-
gorithm is exhausted. While, in principle, such an algorithm could switch the
value of arbitrarily many variables at a time, typically only a constant number of
variables are switched and, in fact, the most common case is to switch only one
variable (hence walk). An idea that appears to make a significant difference in
this context is that of focusing [SAO05], i.e., restricting the choice of variable(s)
to switch among those contained in clauses violated by the present assignment.
Unfortunately, while there are numerous experimental results regarding per-
formance of random walk algorithms on random formulas, the only mathemat-
ically rigorous result known is due to Alekhnovich and Ben-Sasson [ABS07] as-
serting that the algorithm select a violated clause at random and among its
258 Chapter 8. Random Satisfiability

violated literals select one at random and flip it, due to Papadimitriou [Pap91],
succeeds in linear time on random 3-CNF for densities as high as 1.63, i.e., up to
the largest density for which the pure literal rule succeeds (recall that the success
of the pure literal implies that the set of satisfying assignments is connected).
It is worth pointing out that, at this point, there exist both very interesting
presults regarding the performance of random walk algorithms on random for-
mulas [CMMS03, SM03] and also very interesting mathematical results [FMV06,
COMV07] regarding their performance on the planted model, i.e., on formulas
generated by selecting uniformly at random among all 2k1 nk k-clauses satis-
fied by a fixed (planted) assignment. Perhaps, our recent understanding of the
evolution of the solution-space geometry of random k-CNF formulas will help in
transferring some of these results.

8.7.2. Non-backtracking Algorithms

On random CNF formulas uc is equivalent to simply repeating the following until


either no clauses remain (success), or a 0-clause is generated (failure): if there
is a clause of length 1 satisfy it; otherwise, select a random unassigned variable
and assign it a random value. What makes the analysis of uc possible is the fact
that if the input is a random k-CNF formula, then throughout the execution of
uc the following ia true. Let V (t) be the set of variables that have not yet been
assigned a value after t steps of the algorithm, and let Ci (t) be the set of clauses
of length i at that time. Then the set Ci (t) is distributed exactly as a random
i-CNF formula on the variables V (t), with exactly |Ci (t)| clauses.
To see the above claim, imagine representing a CNF formula by a column
of k cards for each k-clause, each card bearing the name of one literal. Assume,
further, that originally all the cards are face-down, i.e., the literal on each card
is concealed (and we never had an opportunity to see which literal is on each card).
At the same time, assume that an intermediary with photographic memory knows
precisely which literal is on each card. To interact with the intermediary we are
allowed to either
Point to a particular card, or,
Name a variable that has not yet been assigned a value.
In response, if the card we point to carries literal `, the intermediary reveals (flips)
Similarly, if we name variable v, the intermediary
all the cards carrying `, `.
reveals all the cards carrying v, v. In either case, faced with all the occurrences of
the chosen variable we proceed to decide which value to assign to it. Having done
so, we remove all the cards corresponding to literals dissatisfied by our setting
and all the cards (some of them still concealed) corresponding to satisfied clauses.
As a result, at the end of each step only face-down cards remain, containing
only literals corresponding to unset variables. This card-game representation
immediately suggests our claimed uniform randomness property for uc and this
can be made easily rigorous by appealing to the method of deferred decisions.
In fact, all algorithms that can be carried out via this game enjoy this property.
Lemma 17 (Uniform randomness). If V (t) = X and |Ci (t)| = qi , the set of
i-clauses remaining at time t form Fi (|X|, qi ) on the variables in X.
Chapter 8. Random Satisfiability 259

Armed with such a simple representation of Markovian state, it is not hard to


show that as long as an algorithm takes care of unit clauses whenever they exist,
its success or failure rests on whether the set of 2-clauses ever acquires density
greater than 1. The main tool used to determine whether that occurs it is to
pass to a so-called liquid model and approximate the evolution of the number
of clauses of each length via a system of differential equations. Indeed, both of
these ideas (uniform randomness plus differential equations) were present in the
work of Chao and Franco in the mid-80s [CF86], albeit not fully mathematically
justified.
The card-game described above does not allow one to carry out the pure
literal heuristic, as it offers no information regarding the number of occurrences
of each literal in the formula. To allow for this, we switch to a slightly different
model for generating random k-CNF formulas. First, we generate km literals
independently, each literal drawn uniformly at random among all 2n literals and,
then, we generate a formula by partitioning these km literals into k-clauses uni-
formly at random. (This might result in a few improper clauses; we addressed
this essentially trivial point in Section 8.6.)
One can visualize the second part of this process as a uniformly random
matching between km literals on the left and their km occurrences in k-clauses
on the right. As the matching is uniformly random, similarly to the card game,
it can be exposed on the fly. (Indeed, the card game above is just the result
of concealing the left hand of the matching.) With this representation we can
now execute algorithms such as: if there is a pure literal or a clause of length 1
satisfy it; otherwise, select a random literal of maximum degree and satisfy it; or,
alternatively, select a most polarized variable and assign it its majority value.
To analyze either one of these algorithms, note that it is enough to maintain as
Markovian state the number of clauses of each length and the number of variables
having i positive and j negative literal occurrences for each i, j 0.
With this long introduction in place we can now describe the state of the art:
Historically, the first fully rigorous lower bound for r3 was given by Broder,
Frieze and Upfal [BFU93] who considered the pure literal heuristic. They
showed that for r 1.63, w.h.p. this eventually sets all the variables (and
that for r > 1.7 w.h.p. it does not). The exact threshold for its success was
given later in [LMS98, Mol05].
Following that, only algorithms expressible in the card-game model were
analyzed for a while. In [AS00] Achlioptas and Sorkin showed that the
optimal algorithm expressible in this model works up to 3.26... For a survey
of the state of the art up to that point, along with a unified framework for
analyzing satisfiability algorithms via differential equations, see [Ach01].
The configuration model for analyzing satisfiability algorithms has been
taken up for k = 3. Specifically, the simple algorithm mentioned above,
in the absence of unit clauses satisfy a random literal of maximum degree,
was proposed and analyzed by Kaporis, Kirousis and Lalas [KKL02] who
showed that it succeeds w.h.p. up to density 3.42... More complicated
algorithms that select which literal to satisfy by considering the degree
of each literal and that of its complement, achieve the best known lower
bound, 3.52..., for the 3-SAT threshold, and were analyzed, independently,
260 Chapter 8. Random Satisfiability

by Kaporis, Kirousis and Lalas [KKL06] and Hajiaghayi and Sorkin [HS03].
It seems likely that this bound can be further improved, at least slightly, by
making even more refined considerations in the choice of variable and value,
but it also seems clear that this is well within the regime of diminishing
returns.
For k > 3, the best results come from consider the following very natural
shortest clausealgorithm sc: pick a random clause c of minimum length;
among the literals in c pick one at random and satisfy it. A weakening
of this algorithm which in the absence of 1-, 2-, and 3-clauses selects a
random literal to satisfy was analyzed by Frieze and Suen in [FS96] and
gives the best known algorithmic lower bound for the k-SAT threshold for
k > 3, namely rk k 2k /k, where k 1.817. In [FS96], the authors
give numerical evidence that even the full sc algorithm only succeeds up
to some k 2k /k, where k = O(1).
In a nutshell, the only algorithms that have been proven to find satisfying
assignments efficiently in random k-CNF formulas are extremely limited and only
succeed for densities up to O(2k /k). In particular, both their variable ordering
and their value assignment heuristic can be implemented given very little, and
completely local, information about the variabe-clause interactions. Of course,
this limitation is also what enables their analysis.

Question 18. Is there a polynomial-time algorithm which w.u.p.p. finds satisfying


assignments of random k-CNF formulas of density 2k /k 1 for some  > 0?

As we saw, the geometry of the space of satisfying assignments undergoes a


phase transition at ln k 2k /k. As a result, an affirmative answer to Question 18
might require a significant departure from the type of algorithms that have been
analyzed so far.

8.8. Belief/Survey Propagation and the Algorithmic Barrier

In [MPR02], Mezard, Parisi, and Zecchina proposed a new satisfiability algorithm


called Survey Propagation (SP) which performs extremely well experimentally on
instances of random 3-SAT. This was a big breakthrough and allowed for optimism
that, perhaps, random k-SAT instances might not be so hard, even close to the
threshold. Unfortunately, conducting experiments with random k-CNF formulas
becomes practically harder at a rapid pace as k increases: the interesting densities
scale as (2k ) so, for example, already k = 10 requires extremely large n in order
for the formulas to be plausibly considered sparse.
As mentioned earlier, in [KMRT+ 07] it is predicted that just below the sat-
isfiability threshold there is a small range of densities, scaling as 2k ln 2 (1),
for which although exponentially many clusters exist, almost all satisfying assign-
ments lie in a finite number of (atypically large) clusters. This condensation of
nearly all satisfying assignments to a small number of clusters induces long-range
correlations among the variables, making it difficult to estimate their marginal
distributions by examining only a bounded neighborhood around each variable.
SP is an ingenuous heuristic idea for addressing this problem by considering not
Chapter 8. Random Satisfiability 261

the uniform measure over satisfying assignments but, rather, (an approximation
of) the uniform measure over clusters, where each cluster is represented by the
fixed point of a certain iterative procedure applied to any assignment in the clus-
ter.
That said, for all densities below the condensation transition, SP is not strictly
necessary: if SP can compute variable marginals, then so can a much simpler al-
gorithm called Belief Propagation, i.e., dynamic programming on trees. This
is because when the measure is carried by exponentially many well-scattered
clusters, marginals are expected to decorrelate. Indeed Gershenfeld and Monta-
nari [GM07] gave very strong rigorous evidence that BP succeeds in computing
marginals in the uncondensed regime for the coloring problem. So, although SP
might be useful when working very close to the threshold, it is not readily helpful
in designing an algorithm that can provably find solutions even at much lower
densities, e.g., say at r = 2k2 , roughly in the middle of the satisfiable regime.
One big obstacle is that, currently, to use either BP or SP to find satisfy-
ing assignments one sets variables iteratively. When a constant fraction of the
variables are frozen in each cluster, as is the case after the dynamical transition,
setting a single variable typically eliminates a constant fraction of all clusters. As
a result, very quickly, one can be left with so few remaining clusters that decor-
relation stops to hold. Concretely, in [MRTS07], Montanari, Ricci-Tersenghi and
Semerjian showed that (even with the relatively generous assumptions of statisti-
cal physics computations) the following algorithm fails for densities greater than
ln k 2k /k. That is, step 2 below fails to converge after only a small fraction of
all variables have been assigned a value:
1. Select a variable v at random.
2. Compute the marginal distribution of v using Belief Propagation.
3. Set v to {0, 1} according to the computed marginal distribution; simplify
the formula; go to step 1.

8.9. Backtracking Algorithms

Backtracking satisfiability algorithms operate by building an assignment step by


step. In particular, the choices they make when no unit clauses and pure literals
are present are called free and when those choices lead to contradictions (empty
clauses), backtracking occurs. In contrast, their non-backtracking variants en-
countered in Section 8.7.2 simply give up when that happens. Moreover, the only
non-backtracking algorithms that have been analyzed maintain their residual for-
mula uniformly random conditional on some small notion of state, reflecting
the number of clauses of each length and perhaps, additionally, the number of
occurrences of each literal. Let us call such algorithms myopic.
It is not hard to prove that the largest density, rA , for which a myopic non-
backtracking algorithm A will find satisfying assignments of a random k-CNF
formula is precisely the largest density for which its residual 2-CNF subformula
remains below density 1 throughout As execution (see e.g. [Ach01]). For densities
beyond rA one can endow A with a backtracking scheme and attempt to analyze
its performance. Unfortunately, any non-trivial amount of backtracking makes it
262 Chapter 8. Random Satisfiability

hard to have a compact probabilistic model for the residual formula. As a result,
a probabilistic analysis akin to that possible for r < rA appears beyond the
reach of current mathematical techniques (but see [CM04, CM05, Mon05] for an
analysis using techniques of statistical physics). Nevertheless, for all backtracking
extensions of myopic algorithms it is possible to prove that they take exponential
time when the initial k-CNF formula is above a certain critical density. This is
because of the following immediate implication of Theorem 9.
Corollary 19. If a DPLL algorithm ever generates a residual formula that is an
unsatisfiable mixture of uniformly random clauses in which the 2-clause density
is below 1, then w.h.p. it will spend exponential time before backtracking from it.
That is, by Corollary 19, once a node in the backtracking search is reached
that corresponds to an unsatisfiable random mixture (but where the 2-clauses
alone are satisfiable), the search cannot leave the sub-tree for an exponentially
long time. Standard results (see e.g. [Ach01]) imply that w.u.p.p. this is precisely
what happens for uc started with 3.81n 3-clauses and for sc started with 3.98n 3-
clauses. This is because for such initial densities, at some point, the corresponding
algorithm w.u.p.p. generates a residual (2+p)-CNF formula which is unsatisfiable
w.h.p. per the results of Theorem 5.
Theorem 20. For any constant r 3.81, any backtracking extension of uc
w.u.p.p. takes time 2(n) on F3 (n, rn). Similarly for sc and r 3.98.
We note that the only reason for Theorem 20 is not a high probability result
is that w.u.p.p. each algorithm might generate a contradiction and backtrack,
thus destroying the uniform randomness of the residual formula, before creat-
ing a formula like the one mandated by Corollary 19. It is worth pointing out,
though, that whenever this occurs w.h.p. it is for trivial local reasons. In partic-
ular, Frieze and Suen in [FS96], introduced the following form of backtracking:
when a contradiction is reached, record the portion of the assignment between
the last free choice and the contradiction; these literals become hot. After flip-
ping the variable involved in the last free choice, instead of making the choice
that the original heuristic would suggest, give priority to the complements of the
hot literals in the order that they appeared; once the hot literals are exhausted
continue as with the original heuristic. This backtracking rule is quite natural in
that it is the last part of the partial assignment that got is into trouble in the first
place. Moreover, it appears to be a genuinely good idea. Experiments on random
formulas comparing this backtracking extension of uc with just reversing the last
free choice show that the histogram of run-times is significantly better for a large
range of densities [ABM04b]. Another property of this backtracking rule is that
as long as the value of each variable in a partial assignment has been flipped at
most once, as happens when dealing with trivial, local contradictions, the resid-
ual formula is uniformly random. In particular, for this particular backtracking
extension Theorem 20 holds w.h.p.
Theorem 20 sheds light on a widely-cited observation of Mitchell, Selman and
Levesque [MSL92], based on experiments with ordered-dll on small problems,
stating that random 3-SAT is easy in the satisfiable region up to the 4.2 threshold,
becomes sharply much harder at the threshold and quickly becomes easy again at
Chapter 8. Random Satisfiability 263

larger densities in the unsatisfiable region. The upper end of this easy-hard-easy
characterization is somewhat misleading since, as we saw, the result of [CS88] in
fact asserts that w.h.p. random 3-CNF formulas only have exponential-size proofs
of unsatisfiability above the threshold. By now, the rate of decline in running time
as the density is increased has been analyzed as well by Beame et al. [BKPS02].
Theorem 20 shows that the lower end of this characterization is also misleading in
that the onset of exponential behavior occurs significantly below the (conjectured)
satisfiability threshold at density 4.2. This concurs with experimental evidence
that even the best of current DPLL implementations seem to have bad behavior
below the threshold [CDA+ 03]. Moreover, as we will see shortly, the gap between
the onset of exponential behavior and the satisfiability threshold increases rapidly
with k.
Corollary 19 probably points to a much larger truth than what is specifically
derived for the algorithms and backtracking schemes mentioned above. This is
because the proof of Theorem 9 is quite robust with respect to the probability
distribution of the clauses in the mixture. The essential ingredient seems to be
that the variable-clause incidence graph is an expander, suggesting that, in fact,
random k-CNF formulas are not the only formulas for which one could hope
to prove a result similar to Theorem 20. Moreover, precisely the combinatorial
richness of expanders suggests that restarting a DPLL algorithm on a random
k-CNF formula is unlikely to yield dramatically different results from run to run,
unless, of course, one is willing to restart an exponential number of times.
The bounds on the 3-clause density needed to cause exponential behavior
in satisfiability algorithms will be readily improved with any improvement on
the 2.28n upper bound for unsatisfiability in random (2 + p)-SAT. In particu-
lar, if Conjecture 6 is true, then Theorem 9 implies [ABM04b] that the running
time of every myopic algorithm goes from linear to exponential around a critical,
algorithm-specific density.

8.10. Exponential Running-Time for k > 3

The most obvious drawback of Theorem 20 is that it implies exponential behavior


for densities that are only conjectured (but not proven) to be in the satisfiable
regime. For all k 4, the analogue to Theorem 20 holds for densities that
are provably in the satisfiable regime [ABM04a]. Moreover, the ratio between
the density at which the algorithms begins to require exponential time, and the
greatest density for which formulas are known to be satisfiable is of order 1/k.
Concretely,

Theorem 21. For k = 4 and r 7.5 and for k 5 and r (11/k) 2k2 ,
ordered-dll w.u.p.p. requires time 2(n) on Fk (n, rn).

Analogues of Theorem 21 hold for many other backtracking extensions of


myopic [Ach01] algorithms and, similarly to the results for k = 3, one can get
hight probability results by considering the Frieze-Suen style backtracking.
264 Chapter 8. Random Satisfiability

References

[ABM04a] Dimitris Achlioptas, Paul W. Beame, and Michael Molloy. Expo-


nential bounds for dpll below the satisfiability threshold. In SODA,
pages 139140, 2004.
[ABM04b] Dimitris Achlioptas, Paul W. Beame, and Michael S. O. Molloy. A
sharp threshold in proof complexity yields lower bounds for satisfi-
ability search. J. Comput. Syst. Sci., 68(2):238268, 2004.
[ABS07] Mikhail Alekhnovich and Eli Ben-Sasson. Linear upper bounds for
random walk on small density random 3-CNFs. SIAM J. Comput.,
36(5):12481263, 2007.
[Ach99] Dimitris Achlioptas. Threshold phenomena in random graph colour-
ing and satisfiability. PhD thesis, Toronto, Ont., Canada, Canada,
1999. Adviser-Allan Borodin and Adviser-Michael Molloy.
[Ach01] Dimitris Achlioptas. Lower bounds for random 3-SAT via differential
equations. Theor. Comput. Sci., 265(1-2):159185, 2001.
[ACO08] Dimitris Achlioptas and Amin Coja-Oghlan. Algorithmic barriers
from phase transitions, 2008. preprint.
[AKKK01] Dimitris Achlioptas, Lefteris M. Kirousis, Evangelos Kranakis, and
Danny Krizanc. Rigorous results for random (2 + p)-SAT. Theoret.
Comput. Sci., 265(1-2):109129, 2001.
[Ale05] Mikhail Alekhnovich. Lower bounds for k-dnf resolution on random
3-cnfs. In STOC 05: Proceedings of the thirty-seventh annual ACM
symposium on Theory of computing, pages 251256, New York, NY,
USA, 2005. ACM.
[AM06] Dimitris Achlioptas and Cristopher Moore. Random k-SAT: Two
moments suffice to cross a sharp threshold. SIAM J. Comput.,
36(3):740762, 2006.
[ANP05] Dimitris Achlioptas, Assaf Naor, and Yuval Peres. Rigorous location
of phase transitions in hard optimization problems. Nature, 435:759
764, 2005.
[ANP07] Dimitris Achlioptas, Assaf Naor, and Yuval Peres. On the maximum
satisfiability of random formulas. J. ACM, 54(2), 2007.
[AP04] Dimitris Achlioptas and Yuval Peres. The threshold for random
k-SAT is 2k log 2 O(k). J. Amer. Math. Soc., 17(4):947973, 2004.
[AR01] Mikhail Alekhnovich and Alexander A. Razborov. Lower bounds for
polynomial calculus: Non-binomial case. In FOCS, pages 190199,
2001.
[ART06] Dimitris Achlioptas and Federico Ricci-Tersenghi. On the solution-
space geometry of random constraint satisfaction problems. In
STOC, pages 130139. ACM, 2006.
[AS00] Dimitris Achlioptas and Gregory B. Sorkin. Optimal myopic algo-
rithms for random 3-sat. In FOCS, pages 590600, 2000.
[BBC+ 01] Bela Bollobas, Christian Borgs, Jennifer T. Chayes, Jeong Han Kim,
and David B. Wilson. The scaling window of the 2-SAT transition.
Random Structures Algorithms, 18(3):201256, 2001.
[BFU93] Andrei Z. Broder, Alan M. Frieze, and Eli Upfal. On the satisfiability
Chapter 8. Random Satisfiability 265

and maximum satisfiability of random 3-CNF formulas. In SODA,


pages 322330, 1993.
[BKPS02] Paul W. Beame, Richard Karp, Toniann Pitassi, and Michael Saks.
The efficiency of resolution and Davis-Putnam procedures. SIAM
J. Comput., 31(4):10481075 (electronic), 2002.
[BMW00] Giulio Biroli, Remi Monasson, and Martin Weigt. A variational
description of the ground state structure in random satisfiability
problems. Eur. Phys. J. B, 14:551568, 2000.
[BSI99] Eli Ben-Sasson and Russell Impagliazzo. Random CNFs are hard
for the polynomial calculus. In FOCS, pages 415421, 1999.
[CDA+ 03] Cristian Coarfa, Demetrios D. Demopoulos, Alfonso San Miguel
Aguirre, Devika Subramanian, and Moshe Y. Vardi. Random 3-
SAT: The plot thickens. Constraints, 8(3):243261, 2003.
[CF86] Ming-Te Chao and John Franco. Probabilistic analysis of two heuris-
tics for the 3-satisfiability problem. SIAM J. Comput., 15(4):1106
1118, 1986.
[CF90] Ming-Te Chao and John Franco. Probabilistic analysis of a gen-
eralization of the unit-clause literal selection heuristics for the k-
satisfiability problem. Inform. Sci., 51(3):289314, 1990.
[CGHS04] Don Coppersmith, David Gamarnik, Mohammad Taghi Hajiaghayi,
and Gregory B. Sorkin. Random max sat, random max cut, and
their phase transitions. Random Struct. Algorithms, 24(4):502545,
2004.
[CM97] Stephen A. Cook and David G. Mitchell. Finding hard instances of
the satisfiability problem: a survey. In Satisfiability problem: theory
and applications (Piscataway, NJ, 1996), volume 35 of DIMACS
Ser. Discrete Math. Theoret. Comput. Sci., pages 117. 1997.
[CM04] Simona Cocco and Remi Monasson. Heuristic average-case analy-
sis of the backtrack resolution of random 3-satisfiability instances.
Theoret. Comput. Sci., 320(2-3):345372, 2004.
[CM05] Simona Cocco and Remi Monasson. Restarts and exponential ac-
celeration of the Davis-Putnam-Loveland-Logemann algorithm: a
large deviation analysis of the generalized unit clause heuristic for
random 3-SAT. Ann. Math. Artif. Intell., 43(1-4):153172, 2005.
[CMMS03] Simona Cocco, Remi Monasson, Andrea Montanari, and Guilhem
Semerjian. Approximate analysis of search algorithms with physi-
cal methods. CoRR, cs.CC/0302003, 2003.
[COGL07] Amin Coja-Oghlan, Andreas Goerdt, and Andre Lanka. Strong
refutation heuristics for random k-SAT. Combin. Probab. Comput.,
16(1):528, 2007.
[COGLS04] Amin Coja-Oghlan, Andreas Goerdt, Andre Lanka, and Frank
Schadlich. Techniques from combinatorial approximation algorithms
yield efficient algorithms for random 2k-SAT. Theoret. Comput. Sci.,
329(1-3):145, 2004.
[COMV07] Amin Coja-Oghlan, Elchanan Mossel, and Dan Vilenchik. A spec-
tral approach to analyzing belief propagation for 3-coloring. CoRR,
abs/0712.0171, 2007.
266 Chapter 8. Random Satisfiability

[CR92] Vasek Chvatal and B. Reed. Mick gets some (the odds are on his
side). In FOCS, pages 620627, 1992.
[CS88] Vasek Chvatal and Endre Szemeredi. Many hard examples for res-
olution. J. Assoc. Comput. Mach., 35(4):759768, 1988.
[DB97] Olivier Dubois and Yacine Boufkhad. A general upper bound for the
satisfiability threshold of random r-SAT formulae. J. Algorithms,
24(2):395420, 1997.
[DBM03] Olivier Dubois, Yacine Boufkhad, and Jacques Mandler. Typical
random 3-sat formulae and the satisfiability threshold. Electronic
Colloquium on Computational Complexity (ECCC), 10(007), 2003.
[FdlV92] Wenceslas Fernandez de la Vega. On random 2-SAT. 1992. Unpub-
lished Manuscript.
[Fei02] Uriel Feige. Relations between average case complexity and approx-
imation complexity. In STOC, pages 534543, 2002.
[FGK05] Joel Friedman, Andreas Goerdt, and Michael Krivelevich. Recog-
nizing more unsatisfiable random k-SAT instances efficiently. SIAM
J. Comput., 35(2):408430 (electronic), 2005.
[FMV06] Uriel Feige, Elchanan Mossel, and Dan Vilenchik. Complete conver-
gence of message passing algorithms for some satisfiability problems.
In APPROX-RANDOM, pages 339350, 2006.
[FP83] John Franco and Marvin Paull. Probabilistic analysis of the Davis
Putnam procedure for solving the satisfiability problem. Discrete
Appl. Math., 5(1):7787, 1983.
[Fra01] John Franco. Results related to threshold phenomena research in
satisfiability: lower bounds. Theoret. Comput. Sci., 265(1-2):147
157, 2001.
[Fri99] Ehud Friedgut. Sharp thresholds of graph properties, and the k-SAT
problem. J. Amer. Math. Soc., 12:10171054, 1999.
[FS96] Alan Frieze and Stephen Suen. Analysis of two simple heuristics on
a random instance of k-SAT. J. Algorithms, 20(2):312355, 1996.
[GL03] Andreas Goerdt and Andre Lanka. Recognizing more random un-
satisfiable 3-SAT instances efficiently. In Typical case complexity
and phase transitions, volume 16 of Electron. Notes Discrete Math.,
page 26 pp. (electronic). Elsevier, Amsterdam, 2003.
[GM07] Antoine Gerschenfeld and Andrea Montanari. Reconstruction for
models on random graphs. In FOCS, pages 194204, 2007.
[Goe96] Andreas Goerdt. A threshold for unsatisfiability. J. Comput. System
Sci., 53(3):469486, 1996.
[Gol79] Allen Goldberg. On the complexity of the satisfiability problem. In
4th Workshop on Automated Deduction (Austin, TX, 1979), pages
16, 1979.
[GSCK00] Carla P. Gomes, Bart Selman, Nuno Crato, and Henry Kautz.
Heavy-tailed phenomena in satisfiability and constraint satisfaction
problems. J. Automat. Reason., 24(1-2):67100, 2000.
[HS03] Mohammad Taghi Hajiaghayi and Gregory B. Sorkin. The satisfia-
bility threshold of random 3-SAT is at least 3.52. volume RC22942
of IBM Research Report. 2003.
Chapter 8. Random Satisfiability 267

[KKKS98] Lefteris M. Kirousis, Evangelos Kranakis, Danny Krizanc, and Yian-


nis Stamatiou. Approximating the unsatisfiability threshold of ran-
dom formulas. Random Structures Algorithms, 12(3):253269, 1998.
[KKL02] Alexis C. Kaporis, Lefteris M. Kirousis, and Efthimios G. Lalas.
The probabilistic analysis of a greedy satisfiability algorithm. In
AlgorithmsESA 2002, volume 2461 of Lecture Notes in Comput.
Sci., pages 574585. Springer, Berlin, 2002.
[KKL06] Alexis C. Kaporis, Lefteris M. Kirousis, and Efthimios G. Lalas. The
probabilistic analysis of a greedy satisfiability algorithm. Random
Structures Algorithms, 28(4):444480, 2006.
[KMRT+ 07] Florent Krzakala, Andrea Montanari, Federico Ricci-Tersenghi,
Guilhem Semerjian, and Lenka Zdeborova. Gibbs states and the
set of solutions of random constraint satisfaction problems. Proc.
Natl. Acad. Sci. USA, 104(25):1031810323 (electronic), 2007.
[KS94] Scott Kirkpatrick and Bart Selman. Critical behavior in the satisfia-
bility of random Boolean expressions. Science, 264(5163):12971301,
1994.
[LMS98] Michael G. Luby, Michael Mitzenmacher, and M. Amin Shokrollahi.
Analysis of random processes via And-Or tree evaluation. In SODA,
pages 364373, 1998.
[MMZ06] Stephan Mertens, Marc Mezard, and Riccardo Zecchina. Threshold
values of random K-SAT from the cavity method. Random Struc-
tures Algorithms, 28(3):340373, 2006.
[Mol05] Michael S. O. Molloy. Cores in random hypergraphs and Boolean
formulas. Random Structures Algorithms, 27(1):124135, 2005.
[Mon05] Remi Monasson. A generating function method for the average-case
analysis of DPLL. In Approximation, randomization and combina-
torial optimization, volume 3624 of Lecture Notes in Comput. Sci.,
pages 402413. Springer, Berlin, 2005.
[MPR02] Marc Mezard, Giorgio Parisi, and Zecchina Ricardo. Analytic and
algorithmic solution of random satisfiability problems. Science,
297:812 815, 2002.
[MRTS07] Andrea Montanari, Federico Ricci-Tersenghi, and Guilhem Se-
merjian. Solving constraint satisfaction problems through belief
propagation-guided decimation. CoRR, abs/0709.1667, 2007.
[MSL92] David G. Mitchell, Bart Selman, and Hector J. Levesque. Hard and
easy distributions of sat problems. In AAAI, pages 459465, 1992.
[MZ98] Remi Monasson and Riccardo Zecchina. Tricritical points in random
combinatorics: the (2 + p)-SAT case. J. Phys. A: Math. and Gen.,
31(46):92099217, 1998.
[MZK+ ] Remi Monasson, Riccardo Zecchina, Scott Kirkpatrick, Bart Sel-
man, and Lidror Troyansky. Phase transition and search cost in the
(2+p)-SAT problem. In 4th Workshop on Physics and Computation,
(Boston, MA, 1996).
[MZK+ 99a] Remi Monasson, Riccardo Zecchina, Scott Kirkpatrick, Bart Sel-
man, and Lidror Troyansky. 2 + p-SAT: relation of typical-case
complexity to the nature of the phase transition. Random Struc-
268 Chapter 8. Random Satisfiability

tures Algorithms, 15(3-4):414435, 1999. Statistical physics meth-


ods in discrete probability, combinatorics, and theoretical computer
science (Princeton, NJ, 1997).
[MZK+ 99b] Remi Monasson, Riccardo Zecchina, Scott Kirkpatrick, Bart Sel-
man, and Lidror Troyansky. Determining computational complexity
from characteristic phase transitions. Nature, 400(6740):133137,
1999.
[Pap91] Christos H. Papadimitriou. On selecting a satisfying truth assign-
ment (extended abstract). In FOCS, pages 163169. IEEE, 1991.
[SAO05] Sakari Seitz, Mikko Alava, and Pekka Orponen. Focused local search
for random 3-satisfiability. Journal of Statistical Mechanics: Theory
and Experiment, 6:6+, June 2005.
[SM03] Guilhem Semerjian and Remi Monasson. A study of pure random
walk on random satisfiability problems with physical methods. In
SAT, pages 120134, 2003.
[SML96] Bart Selman, David G. Mitchell, and Hector J. Levesque. Generating
hard satisfiability problems. Artificial Intelligence, 81(1-2):1729,
1996.

You might also like