IV.19 Extremal and Probabilistic Combinatorics

Noga Alon and Michael Krivelevich


1 Combinatorics: An Introduction

1.1 Examples

It is hard to give a rigorous definition of combinatorics. Instead, let us start with a few examples to illustrate what the area is about.

(i) In the course of an examination of friendship between children some fifty years ago, the Hungarian sociologist Sandor Szalai observed that among any group of about twenty children he checked he could always find four children any two of whom were friends, or else four children no two of whom were friends. Despite the temptation to try to draw sociological conclusions, Szalai realized that this might well be a mathematical phenomenon rather than a sociological one. Indeed, a brief discussion with the mathematicians Erdös, Turán, and Sós convinced him this was the case. If X is any set of size 18 or more, and R is some symmetric RELATION [I.2 §2.3] on X, then there is always a subset S of X of size 4 with the following property: either xRy for any two distinct elements x, y of S, or xRy for no two distinct elements x,y of S. In this case, X is a set of children and R is the relation “is friends with.” This mathematical fact is a special case of Ramsey’s theorem, which was proved by the economist and mathematician Frank Plumpton Ramsey in 1930. Ramsey’s theorem led to the development of Ramsey theory, a branch of extremal combinatorics, which will be discussed in the next section.

(ii) In 1916, Schur was studying FERMAT’S LAST THEOREM [V.10]. It is sometimes possible to prove that a Diophantine equation has no solutions by showing that it has no solutions mod p for some prime p. However, Schur proved that for every integer k and every sufficiently large prime p, there are three integers a, b, and c, none of them congruent to 0 mod p, such that ak + bk is congruent to ck. Although this is a result in number theory, it has a relatively simple and purely combinatorial proof, which is another example of the many applications of Ramsey theory.

(iii) When studying the number of real zeros of random polynomials, LITTLEWOOD [VI.79] and Offord investigated in 1943 the following problem. Let z1, z2, . . ., zn be n not-necessarily-distinct complex numbers, each of modulus at least 1. One can form 2n sums by taking some subset of these numbers and adding them together (with the convention that if one takes the empty set, then the sum is 0). Littlewood and Offord wanted to know how many of these sums there could conceivably be such that the difference between any two of them had modulus less than 1. When n = 2 the answer is easily seen to be at most 2. There are four sums: 0, z1, z2, and z1 + z2. You cannot choose both of the first two or both of the last two or you will have a difference of z1, which has modulus at least 1. Kleitman and Katona proved that in general the maximum is Image. Notice that a simple construction proves that this maximum can be achieved. Indeed, let z1 = z2 = • • • = zn and choose all sums of precisely [n/2] of them. There are Image such sums and they are all equal. The proof that one cannot do better than this uses tools from another area of extremal combinatorics, where the basic objects studied are systems of finite sets.

(iv) Consider a school in which there are m teachers T1, T2, . . ., Tm and n classes C1, C2, . . ., Cn. The teacher Ti has to teach the class Cj for a specified number ρij of lessons. What is the minimum possible number of periods in a complete timetable? Let di denote the total number of lessons the teacher Tí has to teach, and let cj denote the total number of lessons the class Cj has to be taught. Clearly, the number of periods required for a complete schedule is at least as big as any dí or ci, and thus at least as big as the maximum of all these numbers, which we denote by d. It turns out that this obvious lower bound of d is also an upper bound: it is always possible to fit all the lessons that need to be taught into d periods. This is a consequence of König’s theorem, which is a basic result in graph theory. Suppose now that the situation is not so simple: for every teacher Ti and every class Cj there is some specified set of d periods in which the teaching has to take place. Can we always find a feasible timetable with these more complicated constraints? Recent breakthroughs from a subject known as list coloring of graphs imply that it is always possible.

(v) Given a map with several countries represented, how many colors do you need if you want to color the countries without giving any two adjacent countries the same color? Here we assume that each country forms a connected region in the plane. Of course, at least four colors may be necessary: think of Belgium, France, Germany, and Luxembourg, out of which any two have a common border. The FOUR-COLOR THEOREM [V.12], proved by Appel and Haken in 1976, asserts that you never need more than four colors. The study of this problem led to numerous interesting questions and results about graph coloring.

(vi) Let S be an arbitrary subset of the two-dimensional lattice Image2. For any two finite subsets A, BImage we can think of the Cartesian product A × B as a sort of “combinatorial rectangle.” This set has size |A| |B| (where |X| denotes the size of a set X), and we can define an obvious notion of the density ds(A, B) of S in A × B by the formula ds(A, B) = |S(A × B)|/|A| |B|, which measures what proportion of the elements of A × B belong to S. For each k, let d(S, k) be the largest possible value of ds(A,B) if |A| = |B| = k. What can we say about d(S, k) as k tends to infinity? One might guess that almost any behavior is possible, but, remarkably, basic results in extremal graph theory (about the so-called Turán numbers of complete bipartite graphs) imply that d(S, k) must always tend to 0 or 1.

(vii) Suppose that n basketball teams compete in a tournament and any two teams play each other exactly once. The organizers wish to award k prizes at the end of the tournament. It would be embarrassing if there ended up being a team that had not won a prize despite beating all the teams that had won a prize. However, unlikely though it might sound, it is quite possible that this will be the case whatever k teams they choose, at least if n is large enough. To demonstrate this is easy if one uses the probabilistic method, which is one of the most powerful techniques in combinatorics. For any fixed k, and all sufficiently large n, if the results of all the games are chosen randomly (and uniformly and independently), then there is a very high probability that for any k teams there is another team that beats all of them. Probabilistic combinatorics, which is one of the most active areas in modern combinatorics, started with the realization that probabilistic reasoning often provides simple solutions to problems of this type, problems that are often very hard to solve in any other way.

(viii) If G is a finite group of n elements, and H is a subgroup of size k in G, then there are n/k left cosets and n/k right cosets of H. Is there always a set of n/k elements of G that contains a single representative of each right coset and a single representative of each left coset? Hall’s theorem, a basic result in graph theory, implies that there is. In fact, if H′ is another subgroup of size k in G, then there is always a set of n/k elements of G that contains a single representative of each right coset of H and a single representative of each left coset of H′. This may sound like a result in group theory, but it is really a (simple) result in combinatorics.

1.2 Topics

The examples described above illustrate some of the main themes of combinatorics. The subject, sometimes also called discrete mathematics, is a branch of mathematics that focuses on the study of discrete objects (as opposed to continuous ones) and their properties. Although combinatorics is probably as old as the human ability to count, the field has experienced tremendous growth during the last fifty years and has matured into a thriving area with its own set of problems, approaches, and methodology.

The examples above suggest that combinatorics is a basic mathematical discipline that plays a crucial role in the development of many other mathematical areas. In this essay we discuss some of the main aspects of this modern field, focusing on extremal and probabilistic combinatorics. (An account of combinatorial problems with a rather different flavor can be found in ALGEBRAIC AND ENUMERATIVE COMBINATORICS[IV.18].) It is, of course, impossible to cover the area fully in such a short article. A detailed account of the subject can be found in Graham, Grötschel, and Lovász (1995). Our main intention is to give a glimpse of the topics, methods, and applications illustrated by representative examples. The topics we discuss include extremal graph theory, Ramsey theory, the extremal theory of set systems, combinatorial number theory, combinatorial geometry, random graphs, and probabilistic combinatorics. The methods applied in the area include combinatorial techniques, probabilistic methods, tools from linear algebra, spectral techniques, and topological methods. We also discuss the algorithmic aspects and some of the many fascinating open problems in the area.

2 Extremal Combinatorics

Extremal combinatorics deals with the problem of determining or estimating the maximum or minimum possible size of a collection of finite objects that satisfies certain requirements. Such problems are often related to other areas, including computer science, information theory, number theory, and geometry. This branch of combinatorics has developed spectacularly over the last few decades (see, for example, Bollobás (1978), Jukna (2001), and their many references).

2.1 Extremal Graph Theory

A GRAPH[III.34] is one of the very basic combinatorial structures. It consists of a set of points, called vertices, some of which are linked by edges. One can represent a graph visually by drawing the vertices as points in the plane and the edges as lines (or curves). However, formally a graph is more abstract: it is just a set together with a collection of pairs taken from the set. More precisely, it consists of a set V, called the vertex set, and a set E, called the edge set; the elements of E (the edges) are sets of the form {u, v}, where u and v are distinct elements of V. If {u, v} is an edge, we say that u and v are adjacent. The degree d(v) of a vertex v is the number of vertices adjacent to it.

Here are a number of simple definitions associated with graphs that have emerged as important. A path of length k from u to v in G is a sequence of distinct vertices u = v0, v1, . . .,vk = v, where vi and vi+l are adjacent for all i < k. If v0 = vk (but all vertices vi for i < k are distinct), this is called a cycle of length k, and is usually denoted by Ck. A graph G is connected if for any two vertices u, v of G there is a path from u to v. A complete graph Kr is a graph with r vertices such that any two of them are adjacent. A subgraph of a graph G is a graph that contains some of the vertices of G and some of its edges. A clique in G is a set of vertices in G such that any two of them are adjacent. The maximum size of a clique in G is called the clique number of G. Similarly, an independent set in G is a set of vertices in G with no two of them adjacent, and the independence number of G is the maximum size of an independent set in it.

Extremal graph theory deals with quantitative connections between various parameters of a graph, such as its numbers of vertices and edges, its clique number, or its independence number. In many cases a certain optimization problem involving these parameters has to be solved (for example, determining how big one parameter can be if another one is at most some given size), and its optimal solutions are the extremal graphs for this problem. Many important optimization problems that do not explicitly mention graphs can be reformulated, using the definitions above, as problems about extremal graphs.

2.1.1 Graph Coloring

Let us return to the map-coloring example discussed in the introduction. To translate the problem into mathematics, we can describe the map-coloring problem in terms of a graph G, as follows. The vertices of G correspond to the countries on the map, and two vertices are connected by an edge in G if and only if the corresponding countries share a common border. It is not hard to show that one can draw such a graph in such a way that no two edges cross each other: such graphs are called planar. Conversely, any planar graph arises in this way. Therefore, our problem is equivalent to the following: if you want to color the vertices of a planar graph so that no two adjacent vertices receive the same color, then how many colors do you need? (One can make the problem yet more mathematical by removing the nonmathematical notion of color. For example, one can assign to each vertex a positive integer instead.) Such a coloring is called proper. In this language, the four-color theorem states that every planar graph can be properly colored with four colors.

Here is another example of a graph-coloring problem. Suppose we must schedule meetings of several parliament committees. We do not wish to have two committees meeting at the same time if some parliament member belongs to both, so how many sessions do we need?

Again we can model this situation by using a graph G. The vertices of G represent the committees, with two vertices adjacent if and only if the corresponding committees share a member. A schedule is a function f that assigns to each committee one of k time slots. More mathematically, we can think of it as just a function from V to the set {1, 2, . . ., k}. Let us call a schedule valid if no two adjacent vertices are assigned the same number. This corresponds to no two committees being assigned the same time slot if they share a member. The question then becomes, “What is the minimal value of k for which a valid schedule exists?”

The answer is called the chromatic number of the graph G, denoted ξ(G): it is the smallest number of colors in any proper coloring of G. Notice that a coloring of a graph G is proper if and only if for each color the set of vertices of that color is independent. Therefore, ξ(G) can also be defined as the smallest number of independent sets into which it is possible to partition the vertices of G. A graph is called k-colorable if it admits a k-coloring, or, equivalently, if it can be partitioned into k independent sets. Thus, ξ(G) is the minimum k for which G is k-colorable.

Two simple examples are in order. If G is a complete graph Kn on n vertices, then obviously in any coloring of G all vertices get distinct colors, and thus n colors are necessary. Of course, n colors are also sufficient, so ξ (Kn) = n. If G is a cycle C2n+1 on 2n + 1 vertices, then easy parity arguments show that at least three colors are needed, and three colors are enough: color the vertices along the cycle alternately by colors 1 and 2, and then color the last vertex by color 3. Thus, ξ(C2n+1) = 3.

It is not hard to prove that G is 2-colorable if and only if it does not contain a cycle of odd length. Graphs that are 2-colorable are usually called bipartite, since they split into two parts, with all the edges going from one part to the other. The easy characterization ends here, and no simple criterion equivalent to k-colorability is available for k ≥ 3. This is related to the fact that for each fixed k ≤ 3 the computational problem of deciding whether a given graph is k-colorable is NP-hard, a notion discussed in COMPUTATIONAL COMPLEXITY [IV.20].

Coloring is one of the most fundamental notions of graph theory, as a huge array of problems in this field and in related areas like computer science and operations research can be formulated in terms of graph coloring. Finding an optimal coloring of a graph is known to be a very hard task, both theoretically and practically.

There are two simple yet fundamental lower bounds on the chromatic number. First, as every color class in a proper coloring of a graph G forms an independent set, it cannot be bigger than the independence number of G, which is denoted by α(G). Therefore, at least |V(G)|/α(G) colors are necessary. Secondly, if G contains a clique of size k, then k colors are needed to color that clique alone, and thus χ(G) ≥ k. This implies that χ(G)ω(G), where ω(G) is the clique number of G.

What about upper bounds on the chromatic number? One of the simplest approaches to coloring a graph is to do it greedily: put the vertices in some order and color them one by one, assigning to each one the smallest positive integer that has not already been assigned to one of its neighbors. While the greedy algorithm can sometimes be very inefficient (for example, it can color bipartite graphs in an unbounded number of colors, even though two colors are sufficient), it often works quite well. Observe that when applying the greedy algorithm, a color given to a vertex v is at most one more than the number of the neighbors of v preceding it in the chosen order, and is thus at most d(v) + 1, where d(v) is the degree of v in G. It follows that if Δ(G) is the maximum degree of G, then the greedy algorithm uses at most Δ(G) + 1 colors. Therefore χ(G) ≤ Δ(G) + 1. This bound is tight for complete graphs and odd cycles, and, as shown by Brooks in 1941, those are the only cases: if G is a graph of maximum degree Δ, then χ(G) ≤ Δ unless G contains a clique KΔ+1, or Δ = 2 and G contains an odd cycle.

It is also possible to color the edges of a graph, rather than the vertices. In this case a proper coloring is defined to be one where no two edges that meet at a vertex are given the same color. The chromatic index of G, denoted by χ′(G), is the minimum k for which G admits a proper edge-coloring with k colors. For example, if G is the complete graph K2n, then χ′(G) = 2n-1. This turns out to be equivalent to the fact that it is possible to organize a round-robin tournament with 2n teams and fit it into 2n - 1 rounds: just ask the manager of a soccer league. It is also not hard to show that ξ′(K2n - 1) = 2n - 1. Since in any proper edge-coloring of G all edges of G that are incident to a vertex v get distinct colors, the chromatic index is obviously at least as big as the maximum degree. Equality holds for bipartite graphs, as proved by König in 1931, which implies the existence of a complete timetable using d periods in the problem of teachers and classes discussed in the introduction.

Remarkably, this trivial lower bound of χ′(G) ≥ Δ(G) is very close to the true behavior of χ′(G). A fundamental theorem of Vizing from 1964 states that ξ′(G) is always equal either to the maximum degree Δ(G) or to Δ(G) +1. Thus, the chromatic index of G is much easier to approximate than its chromatic number.

2.1.2 Excluded Subgraphs

If a graph G has n vertices and contains no triangle (that is, three vertices all joined to each other) then how many edges can it contain? If n is even, then you can split the vertex set into two equal parts A and B of size n/2 and join every vertex in A to every vertex in B. The resulting graph G contains no triangles and has n2/4 edges. Moreover, adding another edge will automatically create a triangle (in fact, several triangles). But is this the densest possible triangle-free graph? A hundred years ago the answer was shown to be yes by Mantel. (A similar theorem holds when n is odd, but now A and n must have nearly equal sizes (n + 1)/2 and (n - 1)/2.)

Let us look at a more general problem, where the role of the triangle is played by an arbitrary graph. More precisely, let H be any graph, with m vertices, say, and when nm let us define ex(n, H) to be the maximum possible number of edges in a graph with n vertices that does not contain H as a subgraph. (The notation “ex” stands for “exclude.”) The function ex(n, H) is usually called the Turán number of H, for reasons that will become clear, and finding good approximations for it has been a central problem in extremal graph theory.

What kind of examples of graphs that do not contain H can we think of? One observation that gets us started is that if H has chromatic number r, then it cannot be a subgraph of a graph G with chromatic number less than r. (Why not? Because a proper (r - 1)-coloring of G provides us with a proper (r - 1)-coloring of any subgraph of G.) So a promising approach is to look for a graph G with n vertices, chromatic number r - 1, and as many edges as possible. This is easy to find. Our constraint is that the vertices can be partitioned into r - 1 independent sets. Once we have done that, we may as well include all edges between those sets. The result is a complete (r — 1)-partite graph. A routine calculation shows that in order to maximize the number of edges, one should partition into sets that have sizes as nearly equal as possible. (For example, if n = 10 and r = 4, then we would partition into three sets of sizes 3, 3, and 4.)

The graph that satisfies this condition is called the Turán graph Tr-1(n) and the number of edges it contains is denoted by tr-1(n). We have just argued that ex(n, H) ≥ tr-1(n), which can be shown to be at least as big as (1 - 1/(r - 1)) Image.

Turán’s contribution to this area was to give an exact solution, in 1941, for the most important case, when H is the complete graph Kr on r vertices. He proved that ex(n, Kr) is not just at least tr-1(n), but is actually equal to tr-1(n). Moreover, the only Kr-free graph with n vertices and ex(n, Kr) edges is the Turán graph Tr-1(n). Turán’s paper is generally considered the starting point of extremal graph theory.

Later, ErdImages, Stone, and Simonovits extended Turán’s theorem by proving that the above simple lower bound for ex(n, H) is asymptotically tight for any fixed H with chromatic number at least 3. That is, if r is the chromatic number of H, then the ratio of ex(n, H) to tr-1(n) tends to 1 as n tends to infinity

Thus, the function ex(n, H) is well-understood for all nonbipartite graphs. Bipartite graphs are rather different, because their Turán numbers are much smaller: if H is bipartite, then ex(n, H)/n2 tends to zero. Determining the asymptotics of ex(n, H) in this case remains a challenging open problem with many unsettled questions. Indeed, the full story is unknown even for the very simple case when H is a cycle. Partial results obtained so far use a variety of techniques from different fields, including probability theory, number theory, and algebraic geometry.

2.1.3 Matchings and Cycles

Let G be a graph. A matching in G is a collection of edges in G of which no two share a vertex. A matching M in G is called perfect if every vertex belongs to one of the edges in M. (The idea is that the edges determine a “match” for each vertex: the match for x is the vertex y for which xy is an edge of M.) Of course, for G to have a perfect matching it must have an even number of vertices.

One of the best-known theorems in graph theory is Halls theorem, which provides a necessary and sufficient condition for the existence of a perfect matching in a bipartite graph. What kind of condition can this be? It is very easy to write down a trivial necessary condition, as follows. Let G be a bipartite graph with vertex sets A and B of equal size. (If they do not have equal size, then clearly there is no perfect matching.) Given any subset S of A, let N(S) denote the set of all vertices in B that are joined to at least one vertex in S. If there is to be a matching, then it must be possible to assign to each vertex in S a distinct “match,” so obviously N(S) must have at least as many elements as S. Hall’s theorem, proved in 1935, asserts that, remarkably, this obvious necessary condition is also sufficient. That is, if N(S) is always at least as big as S, then there will be a perfect matching. More generally, if A is smaller than B, then the same condition guarantees that one can find a matching that includes every vertex in A (but leaves some vertices in B unmatched).

There is a useful reformulation of Hall’s theorem in terms of set systems. Let S1, S2, . . . , Sn be a collection of sets, and suppose that we would like to find a system of distinct representatives: that is, a sequence x1, x2, . . . , xn such that xi, is an element of Si and no two of the xi are the same. Obviously this cannot be done if the union of some k of the sets Si has size less than k. Again, this obvious necessary condition is sufficient. It is not hard to show that this assertion is equivalent to Hall’s theorem: let S be the union of the Si and define a bipartite graph with vertex sets {1, 2, . . . , n} and S, joining i to x if and only if xSi. Then a matching that includes all of the set {1, 2, . . . , n} picks out a system of distinct representatives: xi is the element of S that is matched with i.

Hall’s theorem can be applied to solve the problem of finding a system of representatives for the right and left cosets of a subgroup H, mentioned in section 1.1. Define a bipartite graph F, whose two sides (of size n/k each) are the left and right cosets of H. A left coset g1H is connected by an edge of F to a right coset Hg2 if they share a common element. It is not difficult to show that F satisfies the Hall condition, and hence it has a perfect matching M. Choosing for each edge (giH,Hgj) of M a common element of giH and Hg1, we obtain the required family of representatives.

There is also a necessary and sufficient condition for the existence of a perfect matching in a general (not-necessarily-bipartite) graph G. This is a theorem of Tutte, which we shall not state here.

Recall that Ck denotes a cycle of length k. A cycle is a very basic graph structure, and, as one might expect, there are many extremal results concerning cycles.

Suppose that G is a connected graph with no cycles. If you pick a vertex and look at its neighbors and then the neighbors of its neighbors, and so on, you will see that it has a tree-like structure. Indeed, such graphs are called trees. An easy exercise shows that any tree with n vertices has exactly n - 1 edges. It follows that every graph G on n vertices with at least n edges has a cycle. If you want to guarantee that this cycle has certain extra properties, then you may need more edges. For example, the theorem of Mantel mentioned earlier implies that a graph G with n vertices and more than n2/4 edges contains a triangle C3 = K3. One can also prove that a graph G = (V, E) with |E| > Image k(|V| - 1) has a cycle of length longer than k (and this is in fact a sharp result).

A Hamilton cycle in a graph G is a cycle that visits every vertex of G. This term originated in a game, invented by HAMILTON [VI.37] in 1857, the objective of which was to complete a Hamilton cycle in the graph of the dodecahedron. A graph containing a Hamilton cycle is called Hamiltonian. This concept is strongly related to the well-known TRAVELING SALESMAN PROBLEM [VII.5 §2]: you are given a graph with positive weights assigned to the edges, and you must find a Hamilton cycle for which the sum of the weights of its edges is minimized There are many sufficient criteria for a graph to be Hamiltonian, quite a few of which are based on the sequence of degrees. For example, Dirac proved in 1952 that a graph on n ≥ 3 vertices all of whose degrees are at least n/2 is Hamiltonian.

2.2 Ramsey Theory

Ramsey theory is a systematic study of the following general phenomenon. Surprisingly often, a large structure of a certain kind has to contain a fairly large highly organized substructure, even if the structure itself is completely arbitrary and apparently chaotic. As succinctly put by the mathematician T. S. Motzkin, “Complete disorder is impossible.” One might expect that the simple and very general form of this paradigm ensures that it has many diverse manifestations in different mathematical areas, and this is indeed the case. (One should, however, bear in mind that some natural statements of this kind are false for nonobvious reasons.)

A very simple statement, which can be regarded as a basic prototype for what follows, is the pigeonhole principle. This states that if a set X of n objects is colored with s colors, then there must be a subset of X of size at least n/s that uses just one color. Such a subset is called monochromatic.

The situation becomes more interesting if the set X has some additional structure. It then becomes natural to ask for a monochromatic subset that keeps some of the structure of X. However, it also becomes much less obvious whether such a subset exists. Ramsey theory consists of problems and theorems of this general kind. Although several Ramsey-type theorems had appeared before, Ramsey theory is traditionally regarded as having started with Ramsey’s theorem, proved in 1930. Ramsey took as his set X the set of all the edges in a complete graph, and the monochromatic subset he obtained consisted of all the edges of some complete subgraph. A precise statement of his theorem is as follows. Let k and 1 be integers greater than 1. Then there exists an integer n such that, however you color the edges of the complete graph with n vertices, using the two colors red and blue, there will either be k vertices such that all edges between them are red or 1 vertices such that all edges between them are blue. That is, a sufficiently large complete graph colored with two colors contains a largish complete subgraph that is monochromatic. Let R(k,l) denote the minimum number n with this property. In this language, the observation of Szalai, mentioned in the introduction, is that R(4,4) ≥ 20 (in fact, R(4,4) = 18). Actually, Ramsey’s theorem was more general, in that he allowed any number of colors, and the objects colored could be r-tuples of elements rather than just pairs, as one has when coloring graphs. The exact computation of small Ramsey numbers turns out to be a notoriously difficult task: even the value of R(5,5) is unknown at present.

The second cornerstone of Ramsey theory was laid by ErdImages and Szekeres, who in 1935 wrote a paper containing several important Ramsey-type results. In particular, they proved the recursion R(k,1)R(k - 1, l) + R(k,l - 1). Combined with the easy boundary conditions R(2, 1) = 1, R(k,2) = k, the recursion leads to the estimate R(k,l) ≤ Image. In particular, for the so-called diagonal case k = l we obtain R(k, k) < 4k. Remarkably, no improvement in the exponent of the latter estimate has been found so far. That is, nobody has found an upper bound of the form Ck for some C < 4. The best lower bound known, which we shall discuss in section 3.2, is roughly R(k,k) ≥ 2k/2, so there is a rather substantial gap.

Another Ramsey-type statement, proved by ErdImages and Szekeres, is of a geometric nature. They showed that for every n ≥ 3 there exists a positive integer N such that, given any configuration of N points in the plane in general position (i.e., no three of them are on a line), there are n that form a convex n-gon. (It is instructive to prove that if n = 4 then N can be taken to be 5.) There are several proofs of this theorem, some using the general Ramsey theorem. It is conjectured that the smallest value of N that will do in order to ensure a convex n-gon is 2n-2 + 1.

The classic ErdImages-Szekeres paper also contains the following Ramsey-type result: any sequence of n2 + 1 distinct numbers contains a monotone (increasing or decreasing) subsequence of length n + 1.

This provides a quick lower bound of Image for a well-known problem of Ulam, asking for the typical length of a longest increasing subsequence of a random sequence of length n. A detailed description of the distribution of this length has recently been given by Baik, Deift, and Johansson.

In 1927 van der Waerden proved what became known as van der Waerden’s theorem: for all positive integers k and r there exists an integer W such that for every coloring of the set of integers {1, 2, . . . , W} using r colors, one of the colors contains an arithmetic progression of length k. The minimum W for which this is true is denoted by W(k,r). Van der Waerden’s bounds for W (k, r) are enormous: they grow like an Ackermanntype function. A new proof of his theorem was found by Shelah in 1987, and yet another proof was given by Gowers in 2000, while he was studying the (much deeper) “density version” of the theorem, which will be described in section 2.4. These recent proofs provided improved upper bounds for W(k,r), but the bestknown lower bound for this number, which is only exponential in k for each fixed r, is much smaller.

Even before van der Waerden, Schur proved in 1916 that for any positive integer r there exists an integer S(r) such that for every r-coloring of {1, . . . ,S(r)} one of the colors contains a solution of the equation x + y = z. The proof can be derived rather easily from the general Ramsey theorem. Schur applied this statement to prove the following result, mentioned in section 1.1: for every k and all sufficiently large primes p, the equation ak + bk = ck has a nontrivial solution in the integers modulo p. To prove this result, assume that pS(k) and consider the FIELD [I.3 §2.2] Imagep of integers mod p. The nonzero elements of Imagep form a GROUP [I.3 §2.1] under multiplication. Let H be the subgroup of this group consisting of all kth powers: that is, H = {xk : xImage}. It is not hard to show that the index r of H is the highest common factor of k and p - 1, and in particular is at most k. The partition of Image into the cosets of H can be thought of as an r-coloring of Image. By Schur’s theorem there exist x, y, z ∈ {1, . . . , p - 1} that all have the same color—that is, they all belong to the same coset of H. In other words, there exists a residue dImage such that x = dak, y = dbk, z = dck, and dak + dbk = dck modulo p. The desired result follows if we multiply both sides by d-1.

Many additional Ramsey-type results can be found in Graham, Rothschild, and Spencer (1990) or in Graham, Grötschel, and Lovász (1995, chapter 25).

2.3 Extremal Theory of Set Systems

Graphs are one of the fundamental structures studied by combinatorialists, but there are others too. An important branch of the subject is the study of set systems. Most often, these are simply collections of subsets of some n-element set. For example, the collection of all subsets of the set {1, 2, . . . , n} of size at most n/3 is a good example of a set system. An extremal problem in this area is any problem where the aim is to determine, or estimate, the maximum number of sets there can be in a set system that satisfies certain conditions. For example, one of the first results in the area was proved by Sperner in 1928. He looked at the following question: how large a collection of subsets can one choose from an n-element set in such a way that no set from the collection is a subset of any other? A simple example of a set system satisfying this condition is the collection of all sets of size r, for some r. From this it immediately follows that we can obtain a collection as large as the largest binomial coefficient, which is Image if n is even and Image if n is odd.

Sperner showed that this is indeed the maximum possible size of such a collection. This result supplies a quick solution to the real analogue of the problem of Little wood and Offord described in section 1.1. Suppose that x1, x2, . . . , xn are n not-necessarily-distinct real numbers, each of modulus at least 1. A first observation is that we may assume that all the xi are positive, since if we replace a negative xi by -xi (which is positive), then we end up with exactly the same set of sums, but shifted by -xi. (To see this, compare a sum that used to involve xi with the corresponding sum that does not involve -xi, and vice versa.) But now, if A is a proper subset of B, then some xi belongs to B and not to A, so

Image

Therefore, the total number of subset sums you can find with any two differing by less than 1 is at most Image, by Sperner’s theorem.

A set system is called an intersecting family if any two sets in the system intersect. Since a set and its complement cannot both belong to an intersecting family of subsets of {1, 2, . . . , n}, we see immediately that such a family can have size at most 2n-1. Moreover, this bound is achieved by, for example, the collection of all sets that contain the element 1. But what happens if we fix a k and assume in addition that all our sets have size k? We may assume that n ≥ 2k, as otherwise the solution is trivial. ErdImages, Ko, and Rado proved that the maximum is Image. Here is a beautiful proof discovered later by Katona. Suppose you arrange the elements randomly around a circle. Then there are n ways of choosing k elements that are consecutive in this arrangement, and it is quite easy to convince yourself that at most k of these can intersect (if n ≥ 2k). So out of these n sets of size k, only k of them can belong to any given intersecting family. Now it is also easy to show that every set has an equal chance of being one of these n sets, and this proves (by a simple double-counting argument) that the largest possible proportion of sets in the family is k/n. Therefore, the family itself has size at most (k/n) Image, which equals Image. The original proof of ErdImages, Ko, and Rado is more complicated than this, but it is important because it introduced a technique known as compression, which was used to solve many other extremal problems.

Let n > 2k be two positive integers. Suppose that you wish to color all subsets of the set {1, 2, . . . , n} of size k in such a way that any two sets with the same color intersect each other. What is the smallest number of colors you can use? It is not difficult to see that n-2k + 2 colors suffice. Indeed, one color class can be the family of all subsets of {1, 2, . . . , 2k - 1}, which is clearly an intersecting family. And then, for each i such that 2kin, you can take the family of all subsets whose largest element is i. There are n - 2k + 1 such families, and any set of size k belongs either to one of them or to the first family. Therefore, n-2k + 2 colors are enough.

Kneser conjectured in 1955 that this bound was tight: in other words, that if you have fewer than n - 2k + 2 colors then you will have to give the same color to some pair of disjoint sets. This conjecture was proved by Lovász in 1978. His proof is topological, and uses the Borsuk-Ulam theorem. Several simpler proofs have been found since, but they are all based on the topological idea in the first proof. Since Lovász’s breakthrough, topological arguments have become an important part of the armory of researchers in combinatorics.

2.4 Combinatorial Number Theory

Number theory is one of the oldest branches of mathematics. At its core are problems about integers, but a sophisticated array of techniques has been developed to deal with those problems, and these techniques have often themselves been the basis for further study (see, for example, ALGEBRAIC NUMBERS [IV.1], ANALYTIC NUMBER THEORY [IV.2], and ARITHMETIC GEOMETRY [IV.5]). However, some problems in number theory have yielded to the methods of combinatorics. Some of these problems are extremal problems with a combinatorial flavor, while others are classical problems in number theory where the existence of a combinatorial solution has been quite surprising. We describe below a few examples. Many more can be found in chapter 20 of Graham, Grötschel, and Lovász (1995), in Nathanson (1996), and in Tao and Vu (2006).

A simple but important notion in the area is that of a sumset. If A and B are two sets of integers, or more generally are two subsets of an ABELIAN GROUP [I.3 §2.1], then the sumset A + B is defined to be {a + b : ab, A, bB}. For instance, if A = {1, 3} and B = {5, 6, 12}, then A + B = {6, 7, 8, 9, 13, 15}. There are many results relating the size and structure of A + B to those of A and B. For example, the Cauchy–Davenport theorem, which has numerous applications in additive number theory, is the statement that if p is a prime, and A, B are two nonempty subsets of Imagep, then the size of A + B is at least the minimum of p and |A| + |B| - 1. (Equality occurs if A and B are arithmetic progressions with the same common difference.) CAUCHY [VI.29] proved this theorem in 1813, and applied it to give a new proof of a lemma that LAGRANGE [VI.22] had proved as part of his well-known 1770 paper that shows that every positive integer is a sum of four squares. Davenport formulated the theorem as a discrete analogue of a related conjecture of Khinchin about densities of sums of sequences of integers. The proofs given by Cauchy and by Davenport are combinatorial, but there is also a more recent algebraic proof, based on some properties of roots of polynomials. The advantage of the latter is that it provides many variants that do not seem to follow from the combinatorial approach. For example, let us define A⊕B to be the set of all a + b such that aA, bB, and ab. Then the smallest possible size of AB, given the sizes of A and B, is the minimum of p and |A| + |B| - 2. Further extensions can be found in Nathanson (1996) and in Tao and Vu (2006).

The theorem of van der Waerden mentioned in section 2.2 implies that, however you color the positive integers with some finite number r of colors, there must be some color that contains arithmetic progressions of every length. ErdImages and Turán conjectured in 1936 that this always holds for the “most popular” color class. More precisely, they conjectured that for any positive integer k and for any real number Image > 0, there is a positive integer n0 such that if n > n0, any set of at least Imagen positive integers between 1 and n contains a k-term arithmetic progression. (Setting Image = r-1 one can easily deduce van der Waerden’s theorem from this.) After several partial results, this conjecture was proved by Szemerédi in 1975. His deep proof is combinatorial, and applies techniques from Ramsey theory and extremal graph theory. Furstenberg gave another proof in 1977, based on techniques of ERGODIC THEORY [V.9]. In 2000 Gowers gave a new proof, combining combinatorial arguments with tools from analytic number theory. This proof supplied a much better quantitative estimate. A related very recent spectacular result of Green and Tao asserts that there are arbitrarily long arithmetic progressions of prime numbers. Their proof combines number-theoretic techniques with the ergodic theory approach. ErdImages conjectured that any infinite sequence ni for which the sum Σi(1/ni) diverges contains arbitrarily long arithmetic progressions. This conjecture would imply the theorem of Green and Tao.

2.5 Discrete Geometry

Let P be a set of points and let L be a set of lines in the plane. Let us define an incidence to be a pair (p, l), where p is a point in P, l is a line in L, and the point p lies on the line l. Suppose that P contains m distinct points and L contains n distinct lines. How many incidences can there be? This is a geometrical problem, but again it has a strong flavor of extremal combinatorics. As such, it is typical of the area known as discrete (or combinatorial) geometry.

Let us write I (m, n) for the maximum number of incidences there can be between m points and n lines. Szemerédi and Trotter determined the asymptotic behavior of this quantity, up to a constant factor, for all possible values of m and n. There are two absolute positive constants c1, c2 such that, for all m, n,

Image

If m > n2 or n > m2 then one can establish the lower bound by taking all m points on a single line, or all n lines through a single point, respectively. In the harder cases when m and n are closer to each other, one can prove it by letting P contain all the points of a Image by Image grid, and by taking the n most “popular” lines: that is, the n lines that contain the most points of P. Establishing the upper bound is more difficult. The most elegant proof of it is due to Székely, and is based on the fact that, however you draw a graph with m vertices and more than 4m edges, you must have many pairs of edges that cross each other. (This is a rather simple consequence of the famous Euler formula connecting the numbers of vertices, edges, and regions in any drawing of a planar graph.) To bound the number of incidences between a set of points P and a set of lines L in the plane, one considers the graph whose vertices are the points P, and whose edges are all segments between consecutive points along a line in L. The desired bound is obtained by observing that the number of crossings in this graph does not exceed the number of pairs of lines in L, and yet should be large if there are many incidences.

Similar ideas can be used to give a partial answer to the following question: if you take n points in the plane, how many pairs (x,y) of these points can there be with the distance from x to y equal to 1? It is not surprising that the two problems are related: the number of such pairs is the number of incidences between the given n points and the n unit circles that are centered at these points. Here, however, there is a large gap between the best known upper bound, which is cn4/3 for some absolute constant c, and the best known lower bound, which is only n1 + c′/loglog n for some constant c′ > 0.

A fundamental theorem of Helly asserts that if you have a finite family Image of at least d + 1 convex sets in Imaged, and if any d + 1 of them have a point in common, then all sets in the family have a common point. Now let us start with a weaker assumption: given any p of the sets, some d + 1 of those p sets have a point in common. (Here p is some integer greater than d + 1.) Can one then find a set X of at most C points such that each set in Image contains a point in X, with C a constant that depends on p but not on the number of convex sets in the family? This question was raised by Hadwiger and Debrunner in 1957 and solved by Kleitman and Alon in 1992. The proof combines a “fractional version” of Hellys theorem with the duality of LINEAR PROGRAMMING [III.84] and various additional geometric results. Unfortunately, it gives a very poor estimate for C: even in two dimensions and with p = 4 it is not known what the best possible value of C is.

This is just a small sample of problems and results in discrete geometry. Such results have been applied extensively in computational geometry and in combinatorial optimization in recent decades. Two good books on the subject are Pach and Agarwal (1995) and Matousek (2002).

2.6 Tools

Many of the basic results in extremal combinatorics were obtained mainly by ingenuity and detailed reasoning. However, the subject has grown out of this early stage: several deep tools have been developed that have been essential to much of the recent progress in the area. In this subsection, we include a very brief description of some of these tools.

Szemerédi’s regularity lemma is a result in graph theory that has numerous applications in various areas, including combinatorial number theory, computational complexity, and, mainly, extremal graph theory. The precise statement of the lemma, which can be found, for example, in Bollobás (1978), is somewhat technical. The rough statement is that the vertex set of any large graph can be partitioned into a constant number of pieces of nearly equal size, so that the bipartite graphs between most pairs of pieces behave like random bipartite graphs. The strength of this lemma is that it applies to any graph, providing a rough approximation of its structure that enables one to extract a lot of information about it. A typical application is that a graph with “few” triangles can be “well-approximated” by a graph with no triangles. More precisely, for any Image > 0 there exists δ > 0 such that if G is a graph with n vertices and at most δn3 triangles, then one can remove at most Imagen2 edges from G and make it triangle free. This innocent-looking statement turns out to imply the case k = 3 of Szemerédi’s theorem that was mentioned earlier.

Tools from linear and multilinear algebra play an essential role in extremal combinatorics. The most fruitful technique of this kind, which is possibly also the simplest, is the so-called dimension argument. In its simplest form, the method can be described as follows. In order to bound the cardinality of a discrete structure A, one maps its elements to distinct vectors in a VECTOR SPACE [I.3 §2.3], and proves that those vectors are linearly independent. It then follows that the size of A is at most the dimension of the vector space in question. An early application of this argument was found by Larman, Rogers, and Seidel in 1977. They wanted to know how many points it was possible to find in Imagen that determine at most two distinct differences. An example of such a system is the set of all points whose coordinates consist of n-2 0s and two 1s. Notice, however, that these points all lie in the hyperplane of points whose coordinates add up to 2. So this actually provides us with an example in Imagen-1. Therefore, we have a simple lower bound of n(n + 1)/2. Larman, Rogers, and Seidel matched this with an upper bound of (n + 1) (n + 4)/2. They did this by associating with each point of such a set a polynomial in n variables, and by showing that these polynomials are linearly independent and all lie in a space of dimension (n + 1) (n + 4)/2. This has been improved by Blokhuis to (n + 1) (n + 2)/2. He did this by finding n + 1 further polynomials that lie in the same space in such a way that the augmented set of polynomials is still linearly independent. More applications of the dimension argument can be found in Graham, Grötschel, and Lovász (1995, chapter 31).

Spectral techniques, that is, an analysis of EIGEN-VECTORS AND EIGENVALUES [I.3 §4.3], have been used extensively in graph theory. The link comes through the notion of an adjacency matrix of a graph G. This is defined to be the matrix A with entries au,v for each pair of (not-necessarily-distinct) vertices u and v, where au,v = 1 if u and v are joined by an edge, and au,v = 0 otherwise. This matrix is symmetric, and therefore, by standard results in linear algebra, it has real eigenvalues and an ORTHONORMAL BASIS [III.37] of eigenvectors. It turns out that there is a tight relationship between the eigenvalues of the adjacency matrix A and several structural properties of the graph G, and these properties can often be useful in the study of various extremal problems. Of particular interest is the second largest eigenvalue of a regular graph. Suppose that every vertex of a graph G has degree d. Then the vector for which every entry is 1 is easily seen to be an eigenvector with eigenvalue d, and this is the largest eigenvalue. If all other eigenvalues have modulus much smaller than d, then it turns out that G behaves in many ways like a random d-regular graph. In particular, the number of edges inside any set of k of the vertices is roughly the same (provided k is not too small) as one would expect with a random graph. It follows easily that any set of vertices that is not too big has many neighbors among the vertices outside that set. Graphs with the latter property are called EXPANDERS [III.24] and have numerous applications in theoretical computer science. Constructing such graphs explicitly is not an easy matter and was at one time a major open problem. Now, however, several constructions are known, based on algebraic tools. See chapter 9 of Alon and Spencer (2000), and its references, for more details.

The application of topological methods in the study of combinatorial objects such as partially ordered sets, graphs, and set systems has already become part of the mathematical machinery commonly used in combinatorics. An early example is Lovász’s proof of Kneser’s conjecture, mentioned in section 2.3. Another example is a result of which the following is a representative special case. Suppose you have a piece of string with 10 red beads, 15 blue beads, and 20 yellow beads on it. Then, no matter what order the beads come in, you can cut the string in at most 12 places and place the resulting segments of beaded string into five piles, each of which contains two red beads, three blue beads, and four yellow beads. The number 12 is obtained by multiplying 4, the number of piles minus 1, by 3, the number of colors. The general case of this result was proved by Alon using a generalization of Borsuk’s theorem. Many additional examples of topological proofs appear in Graham, Grötschel, and Lovász (1995, chapter 34).

3 Probabilistic Combinatorics

A wonderful development took place in twentieth-century mathematics when it was realized that it is sometimes possible to use probabilistic reasoning to prove mathematical statements that do not have an obvious probabilistic nature. For example, in the first half of the century, Paley, Zygmund, ErdImages, Turán, Shannon, and others used probabilistic reasoning to obtain striking results in analysis, number theory, combinatorics, and information theory. It soon became clear that the so-called probabilistic method is a very powerful tool for proving results in discrete mathematics. The early results combined combinatorial arguments with fairly elementary probabilistic techniques, but in recent years the method has been greatly developed, and now it often requires one to apply much more sophisticated techniques. A recent text dealing with the subject is Alon and Spencer (2000).

The applications of probabilistic techniques in discrete mathematics were initiated by Paul ErdImages, who contributed to the development of the method more than anyone else. One can classify them into three groups.

The first deals with the study of certain classes of random combinatorial objects, like random graphs or random matrices. The results here are essentially results in probability theory, although most of them are motivated by problems in combinatorics. A typical problem is the following: if we pick a graph “at random,” what is the probability that it contains a Hamilton cycle?

The second group consists of applications of the following idea. Suppose you want to prove that a combinatorial structure exists with certain properties. Then one possible method is to choose a structure randomly (from a probability distribution that you are free to specify) and estimate the probability that it has the properties you want. If you can show that this probability is greater than 0, then such a structure exists. Surprisingly often it is much easier to prove this than it is to give an example of a structure that works. For instance, is there a graph with large girth (meaning it has no short cycles) and large chromatic number? Even if “large” means “at least 7,” it is very hard to come up with an example of such a graph. But their existence is a fairly easy consequence of the probabilistic method.

The third group of applications is perhaps the most striking of all. There are many examples of statements that appear to be completely deterministic (even when one is used to the idea of using probability to give existence proofs) but that nevertheless yield to probabilistic reasoning. In the remainder of this section we shall briefly describe some typical examples of each of these three kinds of application.

3.1 Random Structures

The systematic study of random graphs was initiated by ErdImages and Rényi in 1960. The most common way of defining a random graph is to fix a probability p and then to join each pair of vertices with an edge with probability p, with all the choices made independently. The resulting graph is denoted G(n, p). (Formally speaking, G(n, p) is not a graph but a probability distribution, but one often talks about it as though it is a graph that has been produced in a random way.) Given any property, such as “contains no triangles,” we can study the probability that G(n, p) has that property.

A striking discovery of ErdImages and Rényi was that many properties of graphs “emerge very suddenly.” Some examples are “contains a Hamilton cycle,” “is not planar,” and “is connected.” These properties are all monotone, which means that if a graph G has the property and you add an edge to G, then the resulting graph still has the property. Let us take one of these properties and define f(p) to be the probability that the random graph G(n, p) has it. Because the property is monotone, f(p) increases as p increases. What ErdImages and Rényi discovered was that almost all of this increase happens in a very short time. That is, f (p) is almost 0 for small p and then suddenly changes very rapidly and becomes almost 1.

Perhaps the most famous and illustrative example of this swift change is the sudden appearance of the so-called giant component. Let us look at G(n, p) when p has the form c/n. If c < 1, then with high probability all the connected components of G(n, p) have size at most logarithmic in n. However, if c > 1, then G(n, p) almost certainly has one component of size linear in n (the giant component), while all the rest have logarithmic size. This is related to the phenomenon of phase transitions in mathematical physics, which are discussed in PROBABILISTIC MODELS OF CRITICAL PHENOMENA [IV.25]. A result of Friedgut shows that the phase transition for a graph property that is “global,” in a sense that can be made precise, is sharper than the one for a “local” property.

Another interesting early discovery in the study of random graphs was that many of the basic parameters of graphs are highly “concentrated.” A striking example that illustrates what this means is the fact that, for any fixed value of p and for most values of n, almost all graphs G(n, p) have the same clique number. That is, there exists some r (depending on p and n) such that with high probability, when n is large, the clique number of G(n, p) is equal to r. Such a result cannot hold for all n, for continuity reasons, but in the exceptional cases there is still some r such that the clique number is almost certainly equal either to r or to r + 1. In both cases, r is roughly 2log n/log(1/p). The proof of this result is based on the so-called second moment method: one estimates the expectation and the variance of the number of cliques of a given size contained in G(n, p), and applies well-known inequalities of Markov and CHEBYSHEV [VI.45].

The chromatic number of the random graph G(n, p) is also highly concentrated. Its typical behavior for values of p that are bounded away from 0 was determined by Bollobás. A more general result, in which p is allowed to tend to 0 as n → ∞, was proved by Shamir, Spencer, Luczak, Alon, and Krivelevich. In particular, it can be shown that for every α < Image and every integervalued function r(n) < nα, there exists a function p(n) such that the chromatic number of G(n, p (n)) is precisely r(n) almost surely. However, determining the precise degree of concentration of the chromatic number of G(n, p), even in the most basic and important case p = Image (in which all labeled graphs on n vertices occur with equal probability), remains an intriguing open problem.

Many additional results on random graphs can be found in Janson, Łuczak, and RuciImageski (2000).

3.2 Probabilistic Constructions

One of the first applications of the probabilistic method in combinatorics was a lower bound given by ErdImages for the Ramsey number R(k, k), which was defined in section 2.2. He proved that if

Image

then R(k, k) > n. That is, there is a red/blue coloring of the edges of the complete graph on n vertices such that no clique of size k is completely red or completely blue. Notice that the number n = Image satisfies the above inequality for all k ≥ 3, so ErdImages’s result gives an exponential lower bound for R(k, k). The proof is simple: if you color the edges randomly and independently, then the probability that any fixed set of k vertices has all its edges of the same color is twice 2Image. Thus, the expected number of cliques with this property is

Image

If this is less than 1, then there must be at least one coloring for which there are no cliques with this property, and the result is proved.

Note that this proof is completely nonconstructive, in the sense that it merely proves the existence of such a coloring, but gives no efficient way of actually constructing one.

A similar computation yields a solution for the tournament problem mentioned in section 1.1. If the results of the tournament are random, then the probability, for any particular k teams, that no other team beats them all is (1 - (1/2k))n-k. From this it follows that if

Image

then there is a nonzero probability that for every choice of k teams, there is another team that beats them all. In particular, it is possible for this to happen. If n is larger than about k2 2k log 2, then the above inequality holds.

Probabilistic constructions have been very powerful in supplying lower bounds for Ramsey numbers. Besides the bound for R (k, k) mentioned above, there is a subtle probabilistic proof, due to Kim, that R(3, k) ≥ ck2/logk, for some c > 0. This is known to be tight up to a constant factor, as proved by Ajtai, Komlós, and Szemerédi, who also used probabilistic methods.

3.3 Proving Deterministic Theorems

Suppose that you color the integers with k colors. Let us call a set S multicolored if all k colors appear in S. Straus conjectured that for every k there is an m with the following property: given any set S with m elements, there is a coloring of the integers with k colors such that all translates of S are multicolored. This conjecture was proved by ErdImages and Lovász. The proof is probabilistic, and applies a tool called the Lovász local lemma, which, unlike many probabilistic techniques, allows one to show that certain events hold with nonzero probability even when this probability is extremely small. The assertion of this lemma, which has numerous additional applications, is, roughly, that for any finite collection of “nearly independent” low-probability events, there is a positive probability that none of the events holds. Note that the statement of Straus’s conjecture has nothing to do with probability, and yet its proof relies on probabilistic arguments.

A graph G is k-colorable, as we have said, if you can properly color its vertices with k colors. Suppose now that instead of trying to use k colors in total, you have a separate list of k colors for each vertex, and this time you want to find a proper coloring of G where each vertex gets a color from its own list. If you can always do so, no matter what the lists are, then G is called k-choosable, and the smallest k for which G is k-choosable is called the choice number ch(G). If all the lists are the same, then one obtains a k-coloring, so ch(G) must be at least as big as χ(G). One might expect ch(G) to be equal to χ(G), since it seems as though using different lists of k colors for different vertices would make it easier to find a proper coloring than using the same k colors for all vertices. However, this turns out to be far from true. It can be proved that for any constant c there is a constant C such that any graph with average degree at least C has choice number at least c. Such a graph might easily be bipartite (and therefore have chromatic number 2), so it follows that ch(G) can be much bigger than χ(G). Somewhat surprisingly, the proof of this result is probabilistic.

An interesting application of this fact concerns a graph that arises in Ramsey theory. Its vertices are all the points in the plane, with two vertices joined by an edge if and only if the distance between them is 1. The choice number of this graph is infinite, by the above result, but the chromatic number is known to be between 4 and 7.

A typical problem in Ramsey theory asks for a substructure of some kind that is entirely colored with one color. Its cousin, discrepancy theory, merely asks that the numbers of times the colors are used are not too close to each other. Probabilistic arguments have proved extremely useful in numerous problems of this general kind. For example, ErdImages and Spencer proved that in any red/blue coloring of the edges of the complete graph Kn there is a subset V0 of vertices such that the difference between the number of red edges inside V0 and the number of blue edges inside V0 is at least cn3/2 for some absolute constant c > 0. This problem is a convincing manifestation of the power of probabilistic methods, since they can be used in the other direction as well, to prove that the result is tight up to a constant factor. Additional examples of such results can be found in Alon and Spencer (2000).

4 Algorithmic Aspects and Future Challenges

As we have seen, it is one matter to prove that a certain combinatorial structure exists, and quite another to construct an example. A related question is whether an example can be generated by means of an EFFICIENT ALGORITHM [IV.20 §2.3], in which case we call it explicit. This question has become increasingly important because of the rapid development of theoretical computer science, which has close connections with discrete mathematics. It is particularly interesting when the structures in question have been proved to exist by means of probabilistic arguments. Efficient algorithms for producing them are not just interesting on their own, but also have important applications in other areas. For example, explicit constructions of error-correcting codes that are as good as random ones are of major interest in CODING AND INFORMATION THEORY [VII.6], and explicit constructions of certain Ramsey-type colorings may have applications in DERANDOMIZATION [IV.20 §7.1.1] (the process of converting randomized algorithms into deterministic ones).

It turns out, however, that the problem of finding a good explicit construction is often very difficult. Even the simple proof of ErdImages, described in section 3.2, that there are red/blue colorings of graphs with Image vertices containing no monochromatic clique of size k leads to an open problem that seems very difficult. Can we construct, explicitly, such a graph with n ≥ (1 + Image)k vertices in time that is polynomial in n? Here we allow Image to be any constant, as long as it is positive. This problem is still wide open, despite considerable efforts from many mathematicians.

The application of other advanced tools, such as algebraic and analytic techniques, spectral methods, and topological proofs, also tends to lead in many cases to nonconstructive proofs. The conversion of these to algorithmic arguments may well be one of the main future challenges of the area.

Another interesting recent development is the increased appearance of computer-aided proofs in combinatorics, starting with the proof of the FOUR-COLOR THEOREM [V.12]. To incorporate such proofs into the area, without threatening its special beauty and appeal, is a further challenge.

These challenges, the fundamental nature of the area, its tight connection to other disciplines, and its many fascinating open problems ensure that combinatorics will continue to play an essential role in the general development of mathematics and science in the future.

Further Reading

Alon, N., and J. H. Spencer. 2000. The Probabilistic Method, 2nd edn. New York: John Wiley.

Bollobás, B. 1978. Extremal Graph Theory. New York: Academic Press.

Graham, R. L., M. Grötschel, and L. Lovász, eds. 1995. Handbook of Combinatorics. Amsterdam: North-Holland.

Graham, R. L., B. L. Rothschild, and J. H. Spencer. 1990. Ramsey Theory, 2nd edn. New York: John Wiley.

Janson, S., T. Luczak, and A. RuciImageski. 2000. Random Graphs. New York: John Wiley.

Jukna, S. 2001. Extremal Combinatorics. New York: Springer.

Matoušek, J. 2002. Lectures on Discrete Geometry. New York: Springer.

Nathanson, M. 1996. Additive Number Theory: Inverse Theorems and the Geometry of Sumsets. New York: Springer. Pach, J., and P. Agarwal. 1995. Combinatorial Geometry. New York: John Wiley.

Tao, T., and V. H. Vu. 2006. Additive Combinatorics. Cambridge: Cambridge University Press.

 

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset