Sidorenko's conjecture

Last updated

Sidorenko's conjecture is a conjecture in the field of graph theory, posed by Alexander Sidorenko in 1986. Roughly speaking, the conjecture states that for any bipartite graph and graph on vertices with average degree , there are at least labeled copies of in , up to a small error term. Formally, it provides an intuitive inequality about graph homomorphism densities in graphons. The conjectured inequality can be interpreted as a statement that the density of copies of in a graph is asymptotically minimized by a random graph, as one would expect a fraction of possible subgraphs to be a copy of if each edge exists with probability .

Contents

Statement

Let be a graph. Then is said to have Sidorenko's property if, for all graphons , the inequality

is true, where is the homomorphism density of in .

Sidorenko's conjecture (1986) states that every bipartite graph has Sidorenko's property. [1]

If is a graph , this means that the probability of a uniform random mapping from to being a homomorphism is at least the product over each edge in of the probability of that edge being mapped to an edge in . This roughly means that a randomly chosen graph with fixed number of vertices and average degree has the minimum number of labeled copies of . This is not a surprising conjecture because the right hand side of the inequality is the probability of the mapping being a homomorphism if each edge map is independent. So one should expect the two sides to be at least of the same order. The natural extension to graphons would follow from the fact that every graphon is the limit point of some sequence of graphs.

The requirement that is bipartite to have Sidorenko's property is necessary — if is a bipartite graph, then since is triangle-free. But is twice the number of edges in , so Sidorenko's property does not hold for . A similar argument shows that no graph with an odd cycle has Sidorenko's property. Since a graph is bipartite if and only if it has no odd cycles, this implies that the only possible graphs that can have Sidorenko's property are bipartite graphs.

Equivalent formulation

Sidorenko's property is equivalent to the following reformulation:

For all graphs , if has vertices and an average degree of , then .

This is equivalent because the number of homomorphisms from to is twice the number of edges in , and the inequality only needs to be checked when is a graph as previously mentioned.

In this formulation, since the number of non-injective homomorphisms from to is at most a constant times , Sidorenko's property would imply that there are at least labeled copies of in .

Examples

As previously noted, to prove Sidorenko's property it suffices to demonstrate the inequality for all graphs . Throughout this section, is a graph on vertices with average degree . The quantity refers to the number of homomorphisms from to . This quantity is the same as .

Elementary proofs of Sidorenko's property for some graphs follow from the Cauchy–Schwarz inequality or Hölder's inequality. Others can be done by using spectral graph theory, especially noting the observation that the number of closed paths of length from vertex to vertex in is the component in the th row and th column of the matrix , where is the adjacency matrix of .

Cauchy–Schwarz: The 4-cycle C4

By fixing two vertices and of , each copy of that have and on opposite ends can be identified by choosing two (not necessarily distinct) common neighbors of and . Letting denote the codegree of and (i.e. the number of common neighbors), this implies:

by the Cauchy–Schwarz inequality. The sum has now become a count of all pairs of vertices and their common neighbors, which is the same as the count of all vertices and pairs of their neighbors. So:

by Cauchy–Schwarz again. So:

as desired.

Spectral graph theory: The 2k-cycle C2k

Although the Cauchy–Schwarz approach for is elegant and elementary, it does not immediately generalize to all even cycles. However, one can apply spectral graph theory to prove that all even cycles have Sidorenko's property. Note that odd cycles are not accounted for in Sidorenko's conjecture because they are not bipartite.

Using the observation about closed paths, it follows that is the sum of the diagonal entries in . This is equal to the trace of , which in turn is equal to the sum of the th powers of the eigenvalues of . If are the eigenvalues of , then the min-max theorem implies that:

where is the vector with components, all of which are . But then:

because the eigenvalues of a real symmetric matrix are real. So:

as desired.

Entropy: Paths of length 3

J.L. Xiang Li and Balázs Szegedy (2011) introduced the idea of using entropy to prove some cases of Sidorenko's conjecture. Szegedy (2015) later applied the ideas further to prove that an even wider class of bipartite graphs have Sidorenko's property. [2] While Szegedy's proof wound up being abstract and technical, Tim Gowers and Jason Long reduced the argument to a simpler one for specific cases such as paths of length . [3] In essence, the proof chooses a nice probability distribution of choosing the vertices in the path and applies Jensen's inequality (i.e. convexity) to deduce the inequality.

Partial results

Here is a list of some bipartite graphs which have been shown to have Sidorenko's property. Let have bipartition .

However, there are graphs for which Sidorenko's conjecture is still open. An example is the "Möbius strip" graph , formed by removing a -cycle from the complete bipartite graph with parts of size .

László Lovász proved a local version of Sidorenko's conjecture, i.e. for graphs that are "close" to random graphs in a sense of cut norm. [11]

Forcing conjecture

A sequence of graphs is called quasi-random with density for some density if for every graph :

The sequence of graphs would thus have properties of the Erdős–Rényi random graph .

If the edge density is fixed at , then the condition implies that the sequence of graphs is near the equality case in Sidorenko's property for every graph .

From Chung, Graham, and Wilson's 1989 paper about quasi-random graphs, it suffices for the count to match what would be expected of a random graph (i.e. the condition holds for ). [12] The paper also asks which graphs have this property besides . Such graphs are called forcing graphs as their count controls the quasi-randomness of a sequence of graphs.

The forcing conjecture states the following:

A graph is forcing if and only if it is bipartite and not a tree.

It is straightforward to see that if is forcing, then it is bipartite and not a tree. Some examples of forcing graphs are even cycles (shown by Chung, Graham, and Wilson). Skokan and Thoma showed that all complete bipartite graphs that are not trees are forcing. [13]

Sidorenko's conjecture for graphs of density follows from the forcing conjecture. Furthermore, the forcing conjecture would show that graphs that are close to equality in Sidorenko's property must satisfy quasi-randomness conditions. [14]

See also

Related Research Articles

In graph theory, an expander graph is a sparse graph that has strong connectivity properties, quantified using vertex, edge or spectral expansion. Expander constructions have spawned research in pure and applied mathematics, with several applications to complexity theory, design of robust computer networks, and the theory of error-correcting codes.

<span class="mw-page-title-main">Expected value</span> Average value of a random variable

In probability theory, the expected value is a generalization of the weighted average. Informally, the expected value is the arithmetic mean of a large number of independently selected outcomes of a random variable. Since it is obtained through arithmetic, the expected value sometimes may not even be included in the sample data set; it is not the value you would "expect" to get in reality.

In probability theory, Chebyshev's inequality guarantees that, for a wide class of probability distributions, no more than a certain fraction of values can be more than a certain distance from the mean. Specifically, no more than 1/k2 of the distribution's values can be k or more standard deviations away from the mean. The rule is often called Chebyshev's theorem, about the range of standard deviations around the mean, in statistics. The inequality has great utility because it can be applied to any probability distribution in which the mean and variance are defined. For example, it can be used to prove the weak law of large numbers.

In mathematics, a low-discrepancy sequence is a sequence with the property that for all values of N, its subsequence x1, ..., xN has a low discrepancy.

<span class="mw-page-title-main">Extremal graph theory</span>

Extremal graph theory is a branch of combinatorics, itself an area of mathematics, that lies at the intersection of extremal combinatorics and graph theory. In essence, extremal graph theory studies how global properties of a graph influence local substructure. Results in extremal graph theory deal with quantitative connections between various graph properties, both global and local, and problems in extremal graph theory can often be formulated as optimization problems: how big or small a parameter of a graph can be, given some constraints that the graph has to satisfy? A graph that is an optimal solution to such an optimization problem is called an extremal graph, and extremal graphs are important objects of study in extremal graph theory.

In the mathematical field of spectral graph theory, a Ramanujan graph is a regular graph whose spectral gap is almost as large as possible. Such graphs are excellent spectral expanders. As Murty's survey paper notes, Ramanujan graphs "fuse diverse branches of pure mathematics, namely, number theory, representation theory, and algebraic geometry". These graphs are indirectly named after Srinivasa Ramanujan; their name comes from the Ramanujan–Petersson conjecture, which was used in a construction of some of these graphs.

<span class="mw-page-title-main">Szemerédi regularity lemma</span>

Szemerédi's regularity lemma is one of the most powerful tools in extremal graph theory, particularly in the study of large dense graphs. It states that the vertices of every large enough graph can be partitioned into a bounded number of parts so that the edges between different parts behave almost randomly.

<span class="mw-page-title-main">Kneser graph</span> Graph whose vertices correspond to combinations of a set of n elements

In graph theory, the Kneser graphK(n, k) (alternatively KGn,k) is the graph whose vertices correspond to the k-element subsets of a set of n elements, and where two vertices are adjacent if and only if the two corresponding sets are disjoint. Kneser graphs are named after Martin Kneser, who first investigated them in 1956.

In graph theory, a graph is said to be a pseudorandom graph if it obeys certain properties that random graphs obey with high probability. There is no concrete definition of graph pseudorandomness, but there are many reasonable characterizations of pseudorandomness one can consider.

In mathematics, the Grothendieck inequality states that there is a universal constant with the following property. If Mij is an n × n matrix with

In mathematics, the second moment method is a technique used in probability theory and analysis to show that a random variable has positive probability of being positive. More generally, the "moment method" consists of bounding the probability that a random variable fluctuates far from its mean, by using its moments.

In number theory, specifically the study of Diophantine approximation, the lonely runner conjecture is a conjecture about the long-term behavior of runners on a circular track. It states that runners on a track of unit length, with constant speeds all distinct from one another, will each be lonely at some time—at least units away from all others.

In extremal graph theory, the forbidden subgraph problem is the following problem: given a graph , find the maximal number of edges an -vertex graph can have such that it does not have a subgraph isomorphic to . In this context, is called a forbidden subgraph.

<span class="mw-page-title-main">Graphon</span>

In graph theory and statistics, a graphon is a symmetric measurable function , that is important in the study of dense graphs. Graphons arise both as a natural notion for the limit of a sequence of dense graphs, and as the fundamental defining objects of exchangeable random graph models. Graphons are tied to dense graphs by the following pair of observations: the random graph models defined by graphons give rise to dense graphs almost surely, and, by the regularity lemma, graphons capture the structure of arbitrary large dense graphs.

In probability theory, concentration inequalities provide bounds on how a random variable deviates from some value. The law of large numbers of classical probability theory states that sums of independent random variables are, under very mild conditions, close to their expectation with a large probability. Such sums are the most basic examples of random variables concentrated around their mean. Recent results show that such behavior is shared by other functions of independent random variables.

In mathematics, dependent random choice is a probabilistic technique that shows how to find a large set of vertices in a dense graph such that every small subset of vertices has many common neighbors. It is a useful tool to embed a graph into another graph with many edges. Thus it has its application in extremal graph theory, additive combinatorics and Ramsey theory.

In the mathematical field of extremal graph theory, homomorphism density with respect to a graph is a parameter that is associated to each graph in the following manner:

In graph theory, a forcing graph is one whose density determines whether a graph sequence is quasi-random. The term was first coined by Chung, Graham, and Wilson in 1989., and forcing graphs play an important role in the study of pseudorandomness in graph sequences.

Flag algebras are an important computational tool in the field of graph theory which have a wide range of applications in homomorphism density and related topics. Roughly, they formalize the notion of adding and multiplying homomorphism densities and set up a framework to solve graph homomorphism inequalities with computers by reducing them to semidefinite programming problems. Originally introduced by Alexander Razborov in a 2007 paper, the method has since come to solve numerous difficult, previously unresolved graph theoretic questions. These include the question regarding the region of feasible edge density, triangle density pairs and the maximum number of pentagons in triangle free graphs.

In graph theory, an area of mathematics, common graphs belong to a branch of extremal graph theory concerning inequalities in homomorphism densities. Roughly speaking, is a common graph if it "commonly" appears as a subgraph, in a sense that the total number of copies of in any graph and its complement is a large fraction of all possible copies of on the same vertices. Intuitively, if contains few copies of , then its complement must contain lots of copies of in order to compensate for it.

References

  1. Sidorenko, Alexander (1993), "A correlation inequality for bipartite graphs", Graphs and Combinatorics, 9 (2–4): 201–204, doi:10.1007/BF02988307, S2CID   12233056
  2. Szegedy, Balázs (2015), An information theoretic approach to Sidorenko's conjecture, arXiv: 1406.6738
  3. Gowers, Tim. "Entropy and Sidorenko's conjecture — after Szegedy". Gowers's Weblog. Retrieved 1 December 2019.
  4. Mulholland, H.P.; Smith, Cedric (1959), "An inequality arising in genetical theory", American Mathematical Monthly, 66 (8): 673–683, doi:10.1080/00029890.1959.11989387
  5. Sidorenko, Alexander (1991), "Inequalities for functionals generated by bipartite graphs", Diskretnaya Matematika, 2 (3): 50–65, doi:10.1515/dma.1992.2.5.489, S2CID   117471984
  6. Hatami, Hamed (2010), "Graph norms and Sidorenko's conjecture", Israel Journal of Mathematics , 175: 125–150, arXiv: 0806.0047 , doi: 10.1007/s11856-010-0005-1
  7. Conlon, David; Fox, Jacob; Sudakov, Benny (2010), "An approximate version of Sidorenko's conjecture", Geometric and Functional Analysis , 20 (6): 1354–1366, arXiv: 1004.4236 , doi:10.1007/s00039-010-0097-0, S2CID   1872674
  8. Conlon, David; Lee, Joonkyung (2018), Sidorenko's conjecture for blow-ups, arXiv: 1809.01259
  9. Li, J.L. Xiang; Szegedy, Balázs (2011), On the logarithimic calculus and Sidorenko's conjecture, arXiv: 1107.1153
  10. Kim, Jeong Han; Lee, Choongbum; Lee, Joonkyung (2016), "Two Approaches to Sidorenko's Conjecture", Transactions of the American Mathematical Society, 368 (7): 5057–5074, arXiv: 1310.4383 , doi: 10.1090/tran/6487
  11. Lovász, László (2010), Subgraph densities in signed graphons and the local Sidorenko conjecture, arXiv: 1004.3026
  12. Chung, Fan; Graham, Ronald; Wilson, Richard (1989), "Quasi-random graphs", Combinatorica, 9 (4): 345–362, doi:10.1007/BF02125347
  13. Skokan, Jozef; Thoma, Lubos (2004), "Bipartite Subgraphs and Quasi-Randomness", Graphs and Combinatorics, 20 (2): 255–262, doi:10.1007/s00373-004-0556-1, S2CID   2154492
  14. Conlon, David; Fox, Jacob; Sudakov, Benny (2010), "An approximate version of Sidorenko's conjecture", Geometric and Functional Analysis, 20 (6): 1354–1366, arXiv: 1004.4236 , doi:10.1007/s00039-010-0097-0, S2CID   1872674