When entropy meets Turán: new proofs and hypergraph Turán results
Abstract.
In this paper, we provide a new proof of a density version of Turán’s theorem. We also rephrase both the theorem and the proof using entropy. With the entropic formulation, we show that some naturally defined entropic quantity is closely connected to other common quantities such as Lagrangian and spectral radius. In addition, we also determine the Turán density for a new family of hypergraphs, which we call tents. Our result can be seen as a new generalization of Mubayi’s result on the extended cliques.
1. Introduction
For any -graph (i.e. -uniform hypergraph) , its Turán number is the maximum number of edges in an -free -graph on vertices. Here, is -free if it contains no subgraph (not necessarily induced) isomorphic to . The study of Turán numbers was initiated by Turán [62], who first considered the case where and is the complete graph on vertices. There, Turán showed that is maximized by the balanced complete -partite graph , which we now refer to as the Turán graph. Turán’s foundational work has motivated subsequent works on related problems, driving continuing research in extremal graph theory.
The general Turán problem is fairly understood when . Although the exact value of is not known for general graphs , the celebrated Erdős–Stone theorem asserts that if , where is an asymptotic extremizer. If we define the Turán density to be
for a -graph , then the Erdős–Stone theorem can be rephrased as when is a graph. It is worth pointing out that when , Erdős–Stone gives that , showing that is subquadratic but does not determine the asymptotic behavior of . Despite lots of effort, there are still many interesting open problems regarding the asymptotic behavior of when is bipartite. However, in this paper, we will focus on the non-degenerate case where .
Given how much we know about Turán numbers and Turán densities of graphs, it might be surprising how little we know about hypergraph Turán problems. In fact, the exact value of is still unknown even for , the -uniform clique on vertices. Turán showed that and conjectured that it is actually an equality. However, proving this conjecture still seems hard to date, and the current best upper bound was obtained by Razborov [53] using flag-algebraic computation, which was later verified by [3] and [18]. The difficulty comes from the fact that hypergraph Turán problems have drastically different behaviors from the graph case. For example, there is a large family of constructions all showing given in [35] (also see [20]). In comparison, the Erdős–Simonovits theorem states that any asymptotic extremizer of should be close to . We will discuss other interesting phenomena for hypergraph Turán problems in Section 1.3.
The aim of this paper is to find inspiration for new ways to approach hypergraph Turán problems by examining our new proof of the density Turán theorem, i.e. . This leads to new hypergraph Turán results regarding hypergraphs that we call “tents”, which generalizes Mubayi’s result [41] on the extended cliques. We will introduce our results and related work in more detail in Section 1.3.
Before diving into hypergraph Turán problems, we will first give a quick overview of known proofs of Turán’s theorem. We will then introduce the entropy method, which we use to rephrase both the theorem statement and our proof. Then we will mention our hypergraph Turán results that can be obtained using the new perspective, which can be thought of as one of our main results.
1.1. Proofs of Turán’s theorem
Turán’s original proof [62] works by a clever induction on the number of vertices by removing a from the graph. Erdős [17] later provided another proof that modified the graph step by step, maintaining the -freeness and making the graph complete multipartite at the end. This method has the benefit that it is easier to see that the Turán graph is the extremizer. A proof of the same spirit is a folklore proof that proceeds with symmetrization (also known now as Zykov Symmetrization as this trick was used by Zykov [65, 66] in his work). The proof modifies the graph by taking two non-adjacent vertices, and replacing one with another (see [1, Chapter 41]). Unfortunately, all those proofs do not easily generalize to hypergraphs as they all use properties of graphs crucially.
One proof that looks entirely different from the previous proofs is by applying the Caro–Wei theorem, which is due to Alon and Spencer [2]. The Caro–Wei theorem, independently proven by Caro [8] and Wei [63], gives a lower bound on the independence number of a graph based on its degree sequence. The standard proof of the Caro–Wei theorem is a nice probabilistic argument, which can be found in [2]. By taking the complement and an application of Cauchy–Schwarz, the density Turán theorem immediately follows from Caro–Wei. However, this argument does not generalize well to higher uniformities—although the Caro–Wei theorem can be extended to hypergraphs (see [9]), applying the inequality on the complement no longer gives tight hypergraph Turán results.
Another proof that is seemingly different from all the above is a proof due to Motzkin and Straus [40]. Their proof relies crucially on a quantity called Lagrangian. The Lagrangian of a graph is defined as
Despite its somewhat long definition, it is a natural quantity to consider in the context of Turán problems. To see this, let be some large positive integers. Consider the blowup of obtained by putting in copies of each vertex so that there are vertices in total, where is the extremizer for the Lagrangian. Then there are edges in the blowup. On the other hand, it is clear that , which shows that the density Turán theorem is equivalent to that for every -free graph . Motzkin and Straus’ idea is that if and are not adjacent, then there is an extremizer with either or for . Therefore if is -free, then there is an extremizer with support of size at most . A simple application of Cauchy–Schwarz then concludes the proof. Despite its algebraic look, this proof is actually similar to Zykov Symmetrization in spirit.
It is natural to generalize graph Lagrangian to hypergraph Lagrangian. For any -graph , its hypergraph Lagrangian is defined as the maximum of under the same condition. As before, when each is blown-up to vertices where is the extremizer for the Lagrangian, there are edges in the blowup. As we will mostly talk about the density of a hypergraph rather than the number of edges, it is convenient to define to be the blowup density of . Intuitively, it is the largest edge density of the blowups of . As it turns out, hypergraph Lagrangian is indeed useful for some hypergraph Turán problems, and we will discuss some of those later in Section 1.3 and Section 8.
A lesser-known but nonetheless interesting algebraic argument was discovered by Li and Li [38]. There, they considered the polynomial
for any graph . The key observation is that if is -free, then vanishes whenever of the variables are equal to one another. In light of this, let be the ideal of polynomials that vanish whenever of the variables are equal. Then , and Turán’s theorem follows from an explicit description of the generators of that Li and Li worked out.
Our proof looks different from all the proofs mentioned above. For graphs, our proof can be seen as a double-counting argument that, peculiarly, counts infinitely many objects. In particular, we will lower bound the number of stars of each size, and show that -freeness actually imposes an upper bound on the numbers. An interesting feature our proof has is that in order to get the tight bound on the Turán density, it is necessary to take stars of any size into account. Despite the distinctive look of our proof, our proof is closely related to the standard probabilistic proof of the Caro–Wei theorem. In fact, if one runs the standard proof on the blowup of the graph, and take the size of the blowup to infinity, then the limit of the argument becomes our argument (we thank Maya Sankar for pointing this out to us).
In spite of the similarity to the proof of the Caro–Wei theorem, our counting argument has the advantage that it can be easily rephrased in terms of entropy. This will be crucial as it will inform us how we should adapt the proof for hypergraphs. We will therefore give an introduction to the entropy method in the next subsection.
1.2. The entropy method
The concept of entropy in the context of information theory was first formulated by Shannon in his seminal work in 1948 on the noisy-channel coding theorem [56]. Roughly speaking, the entropy of a random variable measures how much information the random variable carries. Using entropy, Shannon determined the best efficiency of a code transmitted through a noisy channel that can be corrected with high probability. This has become the foundation of information theory, and many other definitions of entropy have been made as well. However, in this paper, we will only use Shannon’s definition of entropy.
The adaptation of Shannon entropy in combinatorics and outside the context of information theory came much later in comparison. Some early examples include Chung, Frankl, Graham and Shearer’s work on triangle-intersecting family of graphs [12] (where Shearer’s inequality was introduced), Radhakrishnan’s entropic proof of the Brégman’s theorem [52], and Friedgut and Kahn’s theorem on the number of copies of a fixed hypergraph in another hypergraph with a given number of edges [26]. There is nonetheless a significant growth in work using the entropy method in the past decade or two. Two recent exciting, and perhaps unexpected, examples are Gilmer’s breakthrough on the union-closed set conjecture [27] and the work of Gowers, Green, Manners and Tao resolving Marton’s conjecture (also known as the polynomial Freimann–Ruzsa conjecture over ) [28].
In the context of extremal graph theory, the entropy method is particularly useful when dealing with counts of homomorphisms or homomorphism densities. Here, for any that are graphs or general -graphs, a homomorphism from to is a function that sends edges of to edges of . In particular, must be injective on any edge of . The homomorphism density is the probability that a uniformly random chosen function from is actually a homomorphism. In this terminology, a corollary of the Kruskal–Katona theorem says that , which follows immediately from Shearer’s inequality (see also [11] for an entropic proof of a slightly stronger result). In the last decade, the entropy method has been applied to show that various bipartite graphs are Sidorenko, i.e. . This was first formalized by Szegedy [60] building on a previous work [37], and this was later adapted to attack Sidorenko’s conjecture [48, 15, 14, 13] and related problems [19, 36, 29, 5]. In fact, we will also prove some Sidorenko-type result using arguments similar to Szegedy’s in our entropic proofs.
Given how much the entropy method has been utilized to understand relations between homomorphism densities, it should be surprising that no entropic proof for Turán’s theorem was known. Indeed, an equivalent formulation of the density Turán theorem is that if then . In this paper, we give the first entropic proof of the density Turán theorem. To do so, we rephrase the density Turán theorem in the following way, and we will later show the equivalence between the two formulations. Below, and throughout the paper, we use to denote the Shannon entropy of a random variable (see Section 3 for definitions and basic properties).
Theorem 1.1 (Entropic Turán theorem).
Let be a positive integer, and let be a -free graph. Let be random variables distributed on so that is always an edge in . Assume are symmetric, i.e. the distribution of and the one of are the same. Then
We make a brief remark that the equivalence is shown via an entropic reinterpretation of blowup density and Langrangian. Indeed, it turns out that for a given graph , the maximum of the quantity for symmetric -valued random variables with is related to the blowup density of . More surprisingly, the maximum of is related to the spectral radius of . Those connections will be made precise and proven in Section 5, where we also generalize the connections to hypergraphs. One benefit is that as an immediate corollary of our entropic Turán theorem, we can generalize spectral Turán theorems established by Wilf [64] and Nikiforov [44, 45].
Theorem 1.2.
Let and be a tree with vertices. For any -free graph , we have
To see that this is indeed a generalization of Wilf’s and Nikiforov’s results, we can take to be the path on vertices. Wilf’s result corresponds to , whereas Nikiforov’s results correspond to and general .
1.3. Hypergraph Turán densities
Using the idea from our entropic proof of the density Turán theorem, we can determine the Turán densities for some new family of hypergraphs. Before presenting our results, let us first introduce some definitions and previous work that are relevant.
For any family of -graphs , its Turán number is defined to be the maximum number of edges in a -graph that is -free for every . The Turán density is defined analogously by . For any family of -graphs and a -graph , we say that is -hom-free if there does not exist any homomorphism for every . A -hom-free -graph is simply a -graph that is -hom-free.
It is a standard result in the field that is the supremum of where runs through all -hom-free -graphs (see [32, Section 2] or [55, Lemma 2.2] for example). Notice that a single edge has blowup density , showing that if is not empty. This immediately shows that either or for any family of -graphs . We see that among the possible values of Turán density, there is a “jump” going from to . When , this is indeed the behavior of Turán densities: the Erdős–Stone theorem shows that all possible values are , showing that there are only jumps in the case of graphs. However, for hypergraphs, the set of possible Turán densities has a different behavior. It was first discovered by Frankl and Rödl [24] that for each , there are infinitely many non-jumps , where for every there exists a family of -graphs with . On the other hand, Baber and Talbot [3] showed that jumps do exist above when . However, our understanding in jumps and non-jumps is still limited, and we do not even know whether is a jump.
A standard argument shows that is a jump if and only if there exists a finite family of -graph with and for each (see [24]). The fact that we do not know whether is a jump can thus be seen as a result of not having sufficient understanding in the families with . Indeed, known families with Turán densities equal to are so few that we can list them here. For general , Mubayi [41] showed that the -uniform extended clique of size has Turán density . Here, the extension of a hypergraph is another hypergraph with higher uniformity obtained by adding different vertices into the edges, and an extended clique is an extension of a complete graph. In particular, is obtained by adding extra vertices to each edge of , where no two edges share any extra vertices. This was later generalized by Mubayi and Pikhurko [42], who showed that the hypergraph with edges
also has Turán density . Here, and later whenever the vertex set is not explicitly described, the vertex set consists of vertices that appear in the description of the edges. Mubayi and Pikhurko’s result is indeed an improvement as is homomorphic to , showing that -hom-free graphs are also -hom-free and so .
We remark that both Mubayi’s [41] and Mubayi and Pikhurko’s [42] results are stronger—the exact Turán numbers were determined for sufficiently many vertices. If we only care about the Turán density, then an argument of Sidorenko [57] based on hypergraph Lagrangian can be modified to show that as well—this is an observation by Keevash [32, Theorem 3.1].
For smaller ’s, slightly more is known. When , Bollobás [6] showed that where and . This was improved by Frankl and Füredi [25], who showed that is already equal to . Using flag algebra, Baber and Talbot [4] improved this further by showing that . Finally, when , Pikhurko [49] showed that .
As shown above, not a lot is known about families of -graphs with . As an application of our entropic proof of the density Turán theorem, we will generalize our argument to show for a new family of -graphs. Our method has a benefit that we may first come up with an argument and then see what family of -graphs need to be forbidden in order for the argument to work. We believe that this advantage can help discovering more families with minimum positive Turán densities.
To state our result, for any partition of , let where is the length of , and . We also denote by (which is equal to by definition). For any with , we define the -tent, denoted by , to be the following -graph. The -tent comes with an edge that is the base and a vertex that is the apex. Setting to be the length of , for each we also have an edge containing such that . Moreover, we require that for any . It is clear that this determines uniquely up to isomorphism—in fact, we must have partition . It is easy to check that this definition matches the definition of above, (with base and being the apex) and Pikhurko’s result can be rephrased as . Our result can now be stated as follows.
Theorem 1.4.
Let be a positive integer, and let be the family of -tents with and . Then .
Note that this is a stronger statement than Mubayi’s and Mubayi and Pikhurko’s results. In fact, admits a homomorphism to for every and , which shows that . Using the same argument, we can transform Theorem 1.4 into a Turán result of a single -graph.
Theorem 1.5.
Let be a positive integer, and let be a partition of such that and for all . Then .
Although when and , Theorem 1.5 is subsumed by the known results mentioned above, this gives a new Turán result for larger ’s. To show that this should be a nontrivial result for larger ’s, we prove the following result in the opposite direction.
Theorem 1.6.
There exists a constant so that for all sufficiently large and any partition of with , if then .
Theorem 1.5 shows that the constant in Theorem 1.6 cannot be smaller than , and it seems like an interesting question to determine the best possible value of . It might help us understand the -graphs with as well. We leave this as a future direction for interested readers.
Beyond showing for various families of -graphs, our method also applies to some other scenarios where the extremizers are blowups of complete hypergraphs. Unfortunately, we have not been able to find an argument that proves a new and clean statement in those settings. We nonetheless include the arguments later in Section 8 in the hope that they will be enlightening for readers interested in adapting our arguments. The relevant background will also be introduced there.
1.4. Structure of the paper
We will first present our new proof of the density Turán theorem in Section 2. We will then introduce the necessary entropic tools in Section 3, which will set us up for Section 4, where we rephrase our proof in terms of entropy. In Section 5, we will show how our entropic formulation captures quantities such as hypergrpah Lagrangian and spectral radius. We will use the connection to prove the spectral Turán theorems and the equivalence between the entropic Turán theorem and the density Turán theorem. In Section 6, we set up some notations and propositions that will be useful in the later sections. In Section 7, we will apply the entropic argument in Section 4 to show Theorem 1.4 in two different ways, and we will also prove Theorems 1.5 and 1.6. Some further generalization of our arguments is included in Section 8, where we also introduce some related known results. Finally, we will end with some concluding remarks in Section 9.
2. A new proof of the density Turán theorem
In this section, we give a new proof to the density Turán theorem. The key idea is to lower bound the density of stars of each size in terms of edge density by their Sidorenko property. If the densities are large, then we shall find a large clique. The main difference of this proof from all the previous ones is that we consider stars of all sizes at once.
Proof of the density Turán theorem.
For any two graphs , let be the homomorphism density of in . That is, is the probability that a function chosen uniformly at random is a homomorphism from to . We will need the following lemma about lower bounding the homomorphism density of stars in terms of edge density, which is a special case of Sidorenko’s conjecture. We include the proof here since the proof is short.
Lemma 2.1.
For , let be the star with vertices. Then
holds for any graph .
Proof.
Assume and . Note that has vertices, and hence
where the inequality follows from the convexity of . ∎
Now we assume the graph is -free. We sample a sequence of i.i.d. random vertices from uniformly at random. For , let be the event that the induced graph on vertices contains as a subgraph centered at . In particular, is the true event. Note that there can only be at most events happening at the same time. Otherwise, assume are all true for some . Then form an -clique in . Therefore, by double counting, we may conclude that
On the other hand, we know that for all . Thus, we have
After rearranging, we get
and we are done. ∎
3. Shannon entropy
In this section, we introduce the definition of Shannon entropy and some of the properties we will use from the literature. We refer the readers to [2, Section 14.6] for a more detailed introduction. We will also prove a lemma which upper bounds the entropies of random variables by the entropy of their mixture. This lemma will be one of the key ingredients of many of the proofs in the rest of this paper.
3.1. Preliminaries
For any discrete random variable , we write . Also, we denote by the support of , i.e. the set of all such that . Through out this paper, the random variables we will consider are always discrete with finite support, i.e. . For any such random variable, we may define its Shannon entropy.
Definition 3.1.
For any random variable , we define its Shannon entropy
For any sequence of random variables , we use to denote the entropy of the random tuple .
We also define the conditional entropy of given .
Definition 3.2.
For any two random variables , the conditional entropy of given is given by
Equivalently, we have
Using the definition of conditional entropy, we have the following chain rule.
Proposition 3.3 (Chain rule).
For any random variables , we have
The following proposition says that on a fixed support, the entropy is maximized by the uniform distribution on that support.
Proposition 3.4 (Uniform bound).
For any random variable , we have
where the equality holds if and only if is uniform.
We will also need the following two propositions about entropy.
Proposition 3.5 (Subadditivity).
For any three random variables , we have
Proposition 3.6 (Dropping condition).
For any three random variables , we have
3.2. Mixture and the mixture bound
In this subsection, the concern is what is called the mixture of random variables.
Definition 3.7.
For random variables and weights with , we say that is the mixture of with weight if is obtained from the following procedure. We first pick an independent random index with probability . Then we set .
In our applications, we will consider mixtures of random variables whose supports do not overlap too much.
Definition 3.8.
Let be a positive integer. We say that the random variables have -wise disjoint supports if for any element , there are at most many indices such that .
With the definitions above, we may state our lemma about an upper bound on the entropies of random variables with -wise disjoint supports, in terms of the entropy of their mixture.
Lemma 3.9 (Mixture bound).
Let be random variables with -wise disjoint supports. Then there exists a mixture of , say , such that
Proof.
Let and we define
Let be an independent random index with probability and let be the mixture. By chain rule, we have . Therefore,
By the definition of entropy and conditional entropy, we have
and
We may upper bound by uniform bound. For any , when conditioning on , there are at most possible indices as an outcome of . Thus, we have
Combining all above, we get
and we are done after rearranging. ∎
The following example shows that Lemma 3.9 resembles a double counting on -wise disjoint sets. Thus, the mixture bound can be viewed as an entropic version of this double counting.
Example 3.10.
Let be an integer and let be some sets that are -wise disjoint. Assume is a random element chosen from uniform at random for each , and let be the mixture of provided by Lemma 3.9. We have , and by uniform bound we have . Hence, Lemma 3.9 implies that
which gives the same bound as the double counting argument on pairs with .
4. Reformulation using the entropy method
In this subsection, we reformulate the proof in Section 2 using entropy to prove Theorem 1.1. As expected, we shall sample the stars in the same way as Szegedy did [60], and we will use Lemma 3.9 to replace the double counting argument.
Proof of Theorem 1.1.
Recall that we have a -free graph and symmetric random variables distributed on with always holding. We first fix an integer , and we will take goes to infinity later.
Claim 4.1.
For each , there exists a random tuple such that
-
(1)
there is always an edge between for all ,
-
(2)
the marginal distributions of and are the same for all , and
-
(3)
.
Proof.
For , it is easy to check that i.i.d. random vertices with the law of satisfy the condition.
For , we first sample an edge using the law of . Next, we condition on and resample times conditionally independently to get . Finally, we sample independently using the law of .
Note that the first two conditions are true from the way we sample the random variables. It remains to compute . Note that since we sampled independently. By chain rule, we have
Therefore, . ∎
Now, we may apply Lemma 3.9 to the random tuples in 4.1. Since is -free, similar to the proof in Section 2, any tuple of vertices is in at most supports . Therefore, the supports of are -wise disjoint. Thus, there is a mixture of such that
Note that the marginal distribution of is also the same as the marginal distribution of , so we may upper bound by by subadditivity. By using , we get
where . By taking to infinity, we conclude that . Therefore,
Let and . If we pick uniformly at random from all the oriented edges, Theorem 1.1 and the uniform bound give
That is, , which recovers the density Turán theorem. In the next section, we will see that Theorem 1.1 is in fact equivalent to the density Turán theorem by relating entropy to blowup densities.
5. Connecting entropy to Lagrangian and spectral radius
In this section, we will show that Theorem 1.1 is equivalent to the density Turán theorem. We will actually generalize this equivalence in many ways: we will show it for hypergraphs, and we will also go much beyond Lagrangian and blowup densities. This will be useful later to draw connection to the spectral radius of graphs.
We first observe that in Theorem 1.1, the quantity that we care about is actually the maximum of when ranges over all possible symmetric distributions on the oriented edges of . This quantity turns out to be related to the blowup density . To extend this to hypergraphs, we make the following definitions.
Definition 5.1 (Random edge with uniform ordering).
Let be a -graph, we say that a tuple of random vertices is a random edge with uniform ordering on if is symmetric and is always an edge of . Here, being symmetric means the distribution of is always the same for any permutation of .
Definition 5.2 (Entropic density).
For any -graph , define its entropic density to be the largest possible value of for any random edge with uniform ordering .
Note that exists as the space of random edge with uniform ordering is compact. We will show that is equal to , which immediately shows that Theorem 1.1 is equivalent to the density Turán theorem. We will actually show a stronger statement. To that end, we make the following notations. For any -graph , let be the set of oriented edges. For each , let be the maximum of for subject to (the same definition was made by Keevash, Lenz and Mubayi [33] where they called the quantity the -spectral radius). Also let be the largest possible value of for any random edge with uniform ordering . Note that and both exist by compactness.
Example 5.3.
When , we clearly have and . When is a graph and , it is not hard to see that is the maximum
where is the adjacency matrix of . It is a standard fact that this is exactly the spectral radius of . In this case, is the largest possible value of for any random edge with uniform ordering .
For general , if , then corresponds to the spectral radius of the adjacency -tensor of , which was proven in [51]. The quantity is the largest possible value of . Once we prove , this would provide a nice alternative interpretation of the spectral radius for hypergraphs.
Now we will show that and are equal to each other. The proof uses Lagrange multiplier in a crucial way.
Proposition 5.4.
For any -graph and any , .
Proof.
For any , let be the oriented link of , i.e. the set such that .
We start with the following claim that helps us simplify when is in a certain form.
Claim 5.5.
For any tuple , we consider a random edge with uniform ordering on given by
We also define
Then we have
Proof.
First, we have
Combining this with , we get
Now, we may prove the proposition. We first show that .
Let be the tuple that achieves the maximum in the definition of . Define , , and in the same way as in 5.5. Note that and . From 5.5, we have
where the inequality follows from the Jensen’s inequality and the concavity of . Therefore .
For the opposite direction, let be a random edge with uniform ordering achieving the maximum of . For any unoriented edge , let be the probability . Also let . Then
and
Therefore, is a maximizer of
subject to for all and . Note that is nonzero only if , and if that is the case we have . By Lagrange multiplier, we know that
is constant for all with . Therefore
is the same for all with . Notice that for any , and for any , we have
Therefore, using 5.5 with , we see that
where, in this case, . Thus, . Note that . Therefore by the fact that
we have . ∎
Corollary 5.6.
For any family of -graphs, is the supremum of for any random edge with uniform ordering on any -hom-free -graph .
Proof.
Since is the supremum of for all -hom-free -graphs , we know that is the supremum of for all -hom-free -graphs as well. The statement follows from the definition of entropic density . ∎
Corollary 5.7.
The entropic Turán theorem (Theorem 1.1) is equivalent to the density Turán theorem.
Proof.
By Corollary 5.6, it suffices to show that if is -free, then is -hom-free. This is clear as any homomorphic image of is . ∎
Remark.
In the previous section, we showed that Theorem 1.1 implies the density Turán theorem using a simpler argument. This turns out to be the direction we care about in this paper. For all the Turán-type results proven later in this paper using entropy and Proposition 5.4, we may also avoid the use of Proposition 5.4 by a similar simpler argument. However, we think Proposition 5.4 is interesting on its own, so we establish the proposition here and will freely use it from now on.
Setting , we can now prove Theorem 1.2 by combining Theorem 1.1 and Szegedy’s method of sampling a random homomorphic image of the tree .
Proof of Theorem 1.2.
From Proposition 5.4 and the observation in Example 5.3, there exists a random edge with uniform ordering on such that . By Theorem 1.1, we have
Let be an ordering of the vertices of where for every , the vertex is adjacent to exactly one with . Now, we sample random vertices in as follows. Let be a random vertex sampled using the law of . Assume we have already sampled , and assume is the neighbor of with . We sample conditionally independently such that . It follows that is always a homomorphic image of in . Also, from the way we sample, we know that . Thus, we have
and we are done by combining this with the previous inequality and rearranging. ∎
For general , recall that our definition of matches the definition of -spectral radius given by Keevash, Lenz and Mubayi. Thus, by combining Proposition 5.4 with Theorem 1.1, we recover the following theorem for graphs by Kang and Nikiforov [31].
Theorem 5.8 ([31]).
Let be a positive integer and be a real number. For any -free graph with vertices and edges, we have
and
Proof.
We also remark that, by utilizing Proposition 5.4, we can translate Theorem 7.1 and also results in Section 8 into spectral results using arguments in the proofs of Theorem 1.2 and Theorem 5.8.
6. Partial hypergraphs
In this section, we introduce some notations and an entropic lemma that will be useful in the later sections. Those notations are non-standard and are set for our own notational convenience when describing hypergraphs and homomorphisms.
A partial -graph is a simplicial complex whose faces have size at most . Its set of vertices is denoted by , and its set of faces, or partial edges, is denoted by . A homomorphism from a partial -graph to a -graph is a map such that for any partial edge , is injective on and is contained in some edge in . Now for any partial -graph , its extension is the -graph obtained as follows: first let be the set of maximal partial edges in , and then extend each partial edge in to a -edge by adding in extra vertices, where two different edges do not share any extra vertices. Notice that if is a simplicial complex generated by edges of some -graph with , then is the extension of as defined in the introduction.
Example 6.1 (Definition of partial tents).
In Section 7, the partial -graphs and the corresponding extensions of concern would be the following. For any partition of with , the partial -tent is the partial -graph obtained by taking the simplicial complex generated by , and then restricting it to where is the base and is the apex. It is easy to verify that is the extension of the partial -graph .
Those definitions are useful as for any partial -graph , a homomorphism is essentially the same as a homomorphism . This would be helpful later as instead of considering homomorphisms from , we can consider homomorphisms from , which are easier to describe.
Proposition 6.2.
Let be a partial -graph, and let be a -graph. Then there is a homomorphic image of in if and only if there is a homomorphic image of in .
Proof.
Let be the simplicial complex generated by the edges in , which we will think of as a partial -graph. Then is the restriction of on . For any homomorphism from to , we also have that it is an homomorphism from to . It is then easy to check that is a homomorphism from to .
Conversely, suppose that is a homomorphism from to . Note that for each , we know that and so is in as well. By the definition, this implies that for every , we have that is injective on and is contained in some edge in . As any vertex in is in exactly one edge in , it is possible to extend to so that is an edge in for each . The extended map is indeed a homomorphism from to . ∎
Later on, as in the proof in Section 4, we will need to show that we can sample random homomorphisms from some tree-like structures with high entropy. Before we can do so, we need to first describe what the tree-like structures are.
Definition 6.3 (Partial forest and forest sequence).
For any partial -graph , any linear order on , and any vertex , let be the set of partial edges whose maximum vertex is . A partial -graph is a partial forest with respect to a linear order on if for every , there is exactly one maximal partial edge in . In this case, the forest sequence of is a sequence where for each , is the number of vertices with .
We also define quantities that are analogs of the quantity we used in Section 4.
Definition 6.4 (Ratio sequence).
Let be a random edge with uniform ordering on a -graph . We define the ratio sequence of by for each .
We are now ready to apply Szegedy’s argument to sample homomorphisms from partial forests with high entropy.
Lemma 6.5.
Let be a random edge with uniform ordering on a -graph and let be its ratio sequence. For any partial forest with a linear order , if is its forest sequence, then one can sample a random homomorphism from to with entropy equal to
Moreover, the random homomorphism can be sampled such that for any partial edge , the distribution of is the same as .
Proof.
We will induct on . The case is vacuously true. Now suppose that it holds for partial forest of size . Let be the maximum vertex in . Then is also a partial forest, and so we may sample a random homomorphism with the prescribed properties. Let be the maximal partial edge in , and let . By the inductive hypothesis, is identically distributed as . Therefore, we may sample given conditionally independently so that is identically distributed as . This way,
where we use that for any . It remains to show that for any partial edge containing , the distribution of is the same as . This is true as by the definition of and , and the distribution is symmetric. ∎
7. Proof of Theorems 1.4 and 1.5
In this section, we will first give two proofs of Theorem 1.4. We will then show how Theorem 1.4 implies Theorem 1.5. Finally, we will conclude this section with a proof of Theorem 1.6.
Throughout this section, we will fix a -graph and a random edge with uniform ordering on . We will also set to be its ratio sequence. We make an observation that to upper bound , it suffices to upper bound by the chain rule. Therefore, the upper bound of Theorem 1.4 follows from the following statement.
Theorem 7.1.
If is -tent-hom-free for every and , then we have
We first show that Theorem 1.4 indeed follows from Theorem 7.1.
Proof of Theorem 1.4 using Theorem 7.1.
First, it is clear that as a single edge does not contain any homomorphic image of any tents, and it has blowup density . To show the reverse inequality, if is -hom-free, then by Theorem 7.1, we have . This shows that . ∎
7.1. First proof of Theorem 7.1
To prove Theorem 7.1, we will apply Lemma 6.5 and Lemma 3.9 to obtain several inequalities involving . Then we will solve for the maximum of subject to the inequalities.
Lemma 7.2.
If is -tent-hom-free for every and , then for any with , we have .
Proof.
We will consider two partial forests and both on . Let be spanned by the two partial edges and . Let be spanned by the two partial edges and . Then both partial -graphs are indeed partial forests with respect to the linear order . It is clear that in with the forest sequence , the vertices contribute one to and contributes to . Similarly, the forest sequence of is all-one except for .
Let be the random homomorphism from given by Lemma 6.5, respectively. Note that if some tuple of vertices is in the supports of both and , then this tuple corresponds to a homomorphism from to . As clearly contains a partial -tent with base and apex , we know that the two random homomorphisms have disjoint support. Suppose that is the mixture given by Lemma 3.9, then by Lemmas 3.9 and 6.5 we know
Observe that both and contains the partial edges and . Therefore and both have the same distributions as by Lemma 6.5, which shows that has the same distribution as as well. Using a similar argument, we can show that has the same distribution as . As a consequence,
This shows that
and so the desired statement follows. ∎
Our next goal is to upper bound . To upper bound the product, we prove the following auxiliary inequality.
Lemma 7.3.
Suppose that are some non-negative real numbers with for any with . Then
Proof.
We will prove this by induction. It clearly holds when . Now suppose that and the statement holds for . Then by the inductive hypothesis,
by AM-GM. Since
we know
and so
as desired. ∎
Combining Lemma 7.2 and Lemma 7.3, we are now ready to prove Theorem 7.1.
Proof of Theorem 7.1.
7.2. Second proof of Theorem 7.1
Here, we give an alternative proof using much more complicated partial forests. Although the proof is more involved, this proof would be the one we generalize later in Section 8.
Lemma 7.4.
If is -tent-hom-free for every and , then for every , we have for each and
Proof.
We will fix throughout this proof. As in what we did in Section 4, we will temporarily fix an integer that will later be taken to infinity. For any , we will define a partial forest on The partial forest is spanned by the partial edges for every , where is set to be . This is indeed a partial forest with respect to the linear order with . We can compute the forest sequence with respect to the linear order as follows: each contributes one to for each , and each with contributes to . Therefore the forest sequence is . Now let be the random homomorphism produced by Lemma 6.5. This gives
(7.1) |
We will now show that the supports of are disjoint for different choices of . Suppose for the sake of contradiction that for some there is a tuple of vertices from lying in the supports of and Then this tuple witnesses a homomorphism sending to . We will show a contradiction by demonstrating that contains a homomorphic image of some partial -tent with .
Let be the minimum index in which and differ, and without loss of generality, suppose that . Then we can find partial edges , in and in . By the minimality of , we know Note that form a partial -tent with base and apex , showing that contains a partial -tent, which is a contradiction.
Therefore we may now apply Lemma 3.9 with . Suppose that is the resulting mixture of for all possible . By Lemma 6.5 and the fact that is present in all partial forests we take for any , we know that has the same distribution as for each . Hence
(7.2) |
Thus Lemmas 3.9, 7.1 and 7.2 now gives
and so
Note that we may replace by in the product. This way, when we take to approach infinity, we must have for each in order for the left hand side to converge. Moreover, the left hand side becomes
as desired. ∎
Once again, to prove Theorem 7.1, we need to upper bound given the inequalities in Lemma 7.4. We will prove a slightly stronger statement, which will also be useful in the next section.
Lemma 7.5.
Let be a positive integer. Fix real numbers . Let be real numbers with
for any . Then
Proof.
We will prove by induction on . When this is clearly true. Now suppose that and the statement is true for all smaller . Then we have
for all by the inductive hypothesis. Now let
for any . Note that for any , we have
Here, we are using that as . Multiplying these up for , and we get
Thus
(weighted AM-GM) | ||||
completing the inductive step.
∎
Alternative Proof of Theorem 7.1.
7.3. Proof of Theorems 1.5 and 1.6
As mentioned in the introduction, Theorem 1.5 is an immediate corollary of Theorem 1.4. We give a detailed argument of how Theorem 1.5 follows from Theorem 1.4 below.
Proof of Theorem 1.5.
Let be a partition of with and for all . Again, it is clear that , so it suffices to show that . To show this, it suffices to show that any -hom-free -graph is also -hom-free for any with and . This will follow immediately if we show that admits a homomorphism to for any such . By Proposition 6.2, it is sufficient to show that admits a homomorphism from for any with and . This is now simple: suppose that has base and apex , and are two edges such that for . We also suppose that has base and apex , and are partial edges such that for . As , we can take so that , and . This is a homomorphism from to as any vertex in shares an edge with in . ∎
Finally, we give a proof of Theorem 1.6 by demonstrating a -graph that has and is -free for large . Similar to an earlier lower-bound construction by Frankl and Füredi [23] for , we will do so by constructing a -graph so that the intersection of any two edges is small.
Proof.
Let be some constant that is close to . In particular, assume that . Let be an auxiliary graph with vertices , and two vertices are connected if the corresponding subsets have intersection at least . Then is a regular graph with degree
where and we use that
when .
By the Caro–Wei theorem, there exists an independent set of size
This corresponds to a -graph on with edges so that any two edges have intersection less than .
Now if contains a homomorphic image of where , let be its base and let be the edge with . Also let be a homomorphism from to . Then , and so . This shows if is the apex of , then for some . However, is contained in some edge in , which is a contradiction. Thus is at least , which is at least the density of . The density of is
which is strictly greater than for sufficiently large as long as . As is continuous on and , this is true for sufficiently close to . ∎
The proof roughly gives . Although our proof is not fully optimized, we believe that it would not give the correct upper bound for even after being fully optimized. Therefore we do not pursue in this direction.
8. Other applications of our method
Recall from the introduction that Mubayi [41] showed where is the extended clique of size , and Mubayi and Pikhurko [42] strengthened it to . In fact they both proved more general results than this: Mubayi showed that for each ,
and Mubayi and Pikhurko strengthened it as follows: consider the partial -graph on vertices generated by and all the -subsets of , and then take its extension . Then as well. Note that is the extension of as a partial -graph, and there is a homomorphism from to . Therefore , and so is indeed a stronger statement. We remark that Keevash’s adaptation [32, Theorem 3.1] of Sidorenko’s argument [57] gives a much more general result than Mubayi and Pikhurko’s result in this case, and we refer the readers to Keevash’s survey for the statement.
We are able to prove as well, though our proof is considerably more complicated, and it seems hard to produce a clean stronger statement. We nonetheless outline the argument here for readers interested in improving our argument.
Theorem 8.1.
Let be positive integers with . Let be a family of -graphs such that the following holds. For any , if we take the union of any different partial forests as in the proof of Lemma 7.4, then its extension is not -hom-free.
Then .
Proof.
Suppose that is -hom-free. Let be any random edge with uniform ordering on and let be its ratio sequence. We first fix some . Temporarily fix some large positive integer as in the proof of Lemma 7.4. For any , let be the random homomorphism from to sampled via Lemma 6.5 as in the proof of Lemma 7.4. Then by the assumption on and that is -hom-free, we know that the supports of the random homomorphisms are -wise disjoint. Therefore, if is the mixture of the ’s provided by Lemma 3.9, we have
Using what we have computed in the proof of Lemma 7.4, when is taken to infinity, we get
Now let for each . Then it is easy to verify that
for each . Therefore, by Lemma 7.5, we get that
This shows that for any -hom-free -graph , and so we have . ∎
Corollary 8.2.
Let be the partial -graph on vertices generated by and all the -subsets of . Let be its extension. Then .
Proof.
First of all, it is clear that is -hom-free. Therefore, by Proposition 6.2, is also -hom-free, and so .
To show that , it now suffices to show that the assumption of Theorem 8.1 holds for any . Indeed, for any collection of different possible ’s, we may construct with size that satisfies the following: for each there exists such that , and there exists a with . Indeed, set . Then , which shows that . Now simply take of size while containing some for some . Label this as .
Now we need to show that there is a homomorphic image of in the extension of . By Proposition 6.2, it suffices to construct a homomorphism from to . To do so, we will simply map to , map to , and then map the rest of the vertices into bijectively. To show that this is indeed a homomorphism, notice first that is a partial edge in . Therefore it remains to check that and are both in for any and . Indeed, if and for some , then is indeed a partial edge in , which shows that both and are partial edges in as well. ∎
We remark that Theorem 8.1 seems much stronger than Corollary 8.2, though we do not see a clean way to extract a stronger statement from Theorem 8.1. We leave this as a potential future direction for interested readers.
With a completely different method, we can improve Mubayi’s result in a slightly different way, and this is closer to what Sideorenko actually did in his paper [57] using hypergraph Lagrangian. In that paper, Sidorenko showed that many extensions of partial -graphs on vertices have Turán density equal to , as long as is at least some threshold that depends on . One special case related to our result is the -graph that can be obtained as follows: consider the partial -graph on spanned by the edges and all the -subsets of , and then take the extension of the partial -graph. For example, is the tent Sidorenko’s result is more general and relies on trees that satisfy the Erdős–Sós conjecture , and we refer the readers to Sidorenko’s original paper [57] for more details (also see [59, Section 2] or [61] for some families of trees where the Erdős–Sós conjecture is known to hold).
With a slightly different choice of partial forests, we can also prove that for sufficiently large with respect to . Our argument actually gives a more general statement: for any , let be the extension of the partial -graph spanned by and all the -subsets of . Then we obtain a sufficient condition for .
Theorem 8.3.
Let be positive integers with and
(8.1) |
Then .
Proof.
It is clear that is -hom-free. Therefore, .
To prove the other direction , we may fix a -hom-free -graph and a random with uniform ordering on . Let be the ratio sequence of . We will solve for the maximum of under the constraints given by the following lemma.
Lemma 8.4.
For any integers with , we have
Proof.
We will fix throughout this proof. As in what we did in Section 4, we will temporarily fix an integer that will later be taken to infinity. For any , we will define a partial forest on . The partial forest is spanned by the partial edges , for every , and for every . With the linear order given by , we know that is indeed a partial forest. We can compute the forest sequence with respect to the linear order as follows: each contributes one to for each . For the contribution of , if it contributes one to ; if it contributes one to ; otherwise it contributes one to . Therefore the forest sequence is , where are the vectors in the standard basis. Now let be the random homomorphism produced by Lemma 6.5. This gives
(8.2) |
Now, we show that the random tuples have -wise disjoint supports. Note that, for any , the extension of the union contains a homomorphic image of , given by the partial edges for and for . Since , this is also a homomorphic image of . Thus, no sequence of vertices is in .
Therefore we may now apply Lemma 3.9 with . Suppose that is the resulting mixture of for all . Note that the partial edge is present in all partial forests, so by Lemma 6.5 we know that has the same distribution as . Similarly, for each , since the partial edge is present in all partial forests, we know that has the same distribution as . Hence
(8.3) |
Thus Lemmas 3.9, 8.2 and 8.3 now give
and so
By rearranging and taking goes to infinity, we obtain
and the lemma follows. ∎
Once again, to prove Theorem 8.3, we need to upper bound given the inequalities in Lemma 8.4. We start with the following inequality similar to Lemma 7.3.
Lemma 8.5.
Suppose that and are some non-negative real numbers. Then
Proof.
We will prove this by inducting on . For , the inequality is trivial.
Assume the statement is true for . From the inductive hypothesis and AM-GM inequality, we have
Now, by using this lemma with and , it is sufficient to upper bound right hand side using the conditions from Lemma 8.4.
Claim 8.6.
We have
Proof.
Let be the largest integer such that
holds. In particular, we have . Set to be the real number such that
From the definition of , we have
and
Therefore, . By replacing the coefficient of using the definition of and rearranging, we may rewrite the inequality we want to show as the following.
(8.4) |
By combining Lemmas 8.5 and 8.6, we get
To give a sense of what the inequality in Theorem 8.3 means, with some standard computation, we can show the following. If are growing positive integers such that for some , then the largest positive integer satisfying 8.1 is . In a different regime where for some positive integers , we can get that the smallest positive integer satisfying the inequality is . We include those computations in the appendix (Propositions A.2 and A.3).
We briefly remark that the threshold Sidorenko deduced on is the same as ours when . However, Sidorenko’s argument works for a more general family of hypergraphs. It is also possible that by modifying Sidorenko’s argument appropriately, we may get a statement analogous to Theorem 8.3 with the extra parameter .
9. Concluding remarks
9.1. Exact result and stability
In this paper, we mostly focus on the Turán density rather than the Turán number. However, we believe that with more work, it is possible to extract the exact Turán number for sufficiently many vertices from our density Turán theorems Theorems 1.4 and 8.3 at least when we also forbid all homomorphic images. More specifically, we believe that there is a corresponding stability result for Theorems 1.4 and 8.3, which is usually helpful to deduce the exact Turán number for sufficiently many vertices. Indeed, many exact results were deduced using stability results in a crucial way. For some examples, we refer the readers to [34, 42, 49, 50, 7, 46, 47, 39, 55].
9.2. Other extremizers
All the Turán results we are able to prove in this paper have blowups of as their asymptotic extremizers, and this is not a coincidence. We find it much easier to construct partial forests that would give tight inequalities on the ratio sequences with equality holding when is a uniform oriented edge in . However, as mentioned in the introduction, many difficulties of hypergraph Turán problems come from the potential complicated structures in the extremizers. It would thus be more exciting if our method can be applied to problems with extremizers not as simple as .
The first step would probably be to extend this to other Turán problems where the extremizers are blowups of some other hypergraphs. Two candidates are the complete bipartite -graph where , and the complete oddly bipartite -graph where is even, and is the -edges such that is odd. Although they are not formally blowups of some smaller hypergraphs, one can think of the complete bipartite -graphs as the blowups of , and the completely oddly bipartite -graphs are the blowups of some -vertex “degenerate” hypergraphs as well.
There are many known Turán results where the two hypergraphs are (asymptotic) extremizers. For example, a classical result of De Caen and Füredi [16] shows that the complete bipartite -graph is an asymptotic extremizer for the Fano plane. This was later extended by Mubayi–Rödl [43] and Baber–Talbot [4]. On the other hand, Keevash and Sudakov [34] showed that the complete oddly bipartite -graph is the extremizer for expanded triangle. A very recent breakthrough of Sankar [55] showed that the complete oddly bipartite -graph is an asymptotic extremizer for tight cycles of sufficiently large length not divisible by .
We are unable to construct any partial forests that give tight inequalities when is the complete bipartite -graph. For being complete oddly -graphs, it is possible to construct such partial forests following the argument in Theorem 1.1 and Sidorenko’s [58] and Frankl’s [21] ideas, which used auxiliary -graphs to show that the Turán densities of expanded triangles are . However, we have not found any other partial forests that use essentially different ideas. It would be interesting to see if there are ways to obtain tight inequalities for those two candidates of in the hope that they would give rise to new Turán results.
Let us close this discussion by mentioning that our method seems to capture a little structure in the conjectured extremizer for , the -graph on vertices with edges. Let be a -graph on vertices with edges so that any -subset is in exactly edges—it turns out that does exist and is unique up to isomorphism. The iterated blowup of is constructed inductively by replacing each vertex in with . Then is -free, and by taking to infinity, we get that . This is a construction of Frankl and Füredi [22], and the construction is conjectured to be optimal. The current best upper bound is obtained by Baber and Talbot [3] using flag algebra. Though we cannot say anything new about the Turán problem of itself, our method seems to capture some structure in . Indeed, by the partial forests for , we can show that if is -free and is a random edge with uniform ordering on , then
This is indeed achieved when is a uniformly chosen oriented edge in .
9.3. Entropic spectral radius
In Section 5, we showed that for any -graph , its spectral radius is related to the maximum of for symmetric distribution on the oriented edges of . It would be interesting if this connection can be utilized to deduce some properties of spectral radius. One possible candidate is a result of Kang, Liu and Shan [30] that showed that
for any -graph , where is the spectral radius of .
9.4. Entropic flag algebra
As one may have observed, many upper bounds on Turán densities, especially for those that are still open, were obtained using flag algebra. Such upper bounds using flag algebra, roughly speaking, are obtained via carefully chosen sum-of-squares inequalities, enumeration of possible small configurations, and numerical computationg of positive semidefinite programs. See [54] for a more detailed discussion of the method.
The inequalities obtained using our argument seem to be really different from the inequalities obtained by sum-of-squares. This suggests a possibility that maybe the flag algebra bounds can be improved with this new idea and some enumeration of possible partial forests to use in the argument. However, aside from the time complexity enumerating through the possible partial forests, there seem to be several technicalities to overcome for this to work. The first is that in most of our proofs, we need to look at infinitely many partial forests in order to get a tight bound. In addition, the inequalities we get, unlike the ones in flag-algebraic arguments, are highly non-linear. However, if we are just aiming for some numerical upper bound that is close to the truth, then hopefully finite but sufficiently many partial forests together with an approximation of the supremum of subject to the inequalities would be enough.
The most serious issue is probably that there has not been a framework for automated entropic computation. So far, the flag-algebraic tools are developed to keep track of the homomorphism densities of labeled graphs. Unfortunately, it seems that all our arguments for hypergrpah Turán problems cannot be rephrased using homomorphism densities as we also crucially use the marginal distributions of the random homomorphisms sampled by Lemma 6.5. It would thus be necessary to come up with an “entropic flag algebra” framework and implement corresponding software to execute the idea in this subsection. We refer the readers to [10] for another entropic argument that motivates this idea of “entropic flag algebra”.
Acknowledgement
The project was motivated when the first author was visiting Hong Liu at Institute for Basic Science, and the first author would like to thank his hospitality. We would also like to thank Ryan Alweiss and Freddie Manners for discussions during the early stage of this project, Dhruv Mubayi and Maya Sankar for pointing us to references for hypergraph Turán problems, Yongtao Li for pointing us to references for spectral Turán problems, and Noga Alon for pointing us to other useful references. Last but not least, we would like to thank Zeev Dvir, Xiaoyu He, Cosmin Pohoata and Maya Sankar for helpful comments on an earlier draft.
References
- [1] Martin Aigner and Günter M. Ziegler, Proofs from The Book, sixth ed., Springer, Berlin, 2018.
- [2] Noga Alon and Joel H. Spencer, The probabilistic method, second ed., Wiley-Interscience Series in Discrete Mathematics and Optimization, Wiley-Interscience [John Wiley & Sons], New York, 2000, With an appendix on the life and work of Paul Erdős.
- [3] Rahil Baber and John Talbot, Hypergraphs do jump, Combin. Probab. Comput. 20 (2011), 161–171.
- [4] Rahil Baber and John Talbot, New Turán densities for 3-graphs, Electron. J. Combin. 19 (2012), Paper 22, 21.
- [5] Natalie Behague, Natasha Morrison, and Jonathan A. Noel, Off-diagonal commonality of graphs via entropy, SIAM J. Discrete Math. 38 (2024), 2335–2360.
- [6] Béla Bollobás, Three-graphs without two triples whose symmetric difference is contained in a third, Discrete Math. 8 (1974), 21–24.
- [7] Axel Brandt, David Irwin, and Tao Jiang, Stability and Turán numbers of a class of hypergraphs via Lagrangians, Combin. Probab. Comput. 26 (2017), 367–405.
- [8] Yair Caro, New results on the independence number, Tech. report, Technical Report, Tel-Aviv University, 1979.
- [9] Yair Caro and Zsolt Tuza, Improved lower bounds on -independence, J. Graph Theory 15 (1991), 99–107.
- [10] Ting-Wei Chao and Hung-Hsun Hans Yu, A purely entropic approach to the rainbow triangle problem, arXiv:2407.14084.
- [11] Ting-Wei Chao and Hung-Hsun Hans Yu, Kruskal–Katona-type problems via the entropy method, J. Combin. Theory Ser. B 169 (2024), 480–506.
- [12] F. R. K. Chung, R. L. Graham, P. Frankl, and J. B. Shearer, Some intersection theorems for ordered sets and graphs, J. Combin. Theory Ser. A 43 (1986), 23–37.
- [13] David Conlon, Jeong Han Kim, Choongbum Lee, and Joonkyung Lee, Sidorenko’s conjecture for higher tree decompositions, arXiv:1805.02238.
- [14] David Conlon, Jeong Han Kim, Choongbum Lee, and Joonkyung Lee, Some advances on Sidorenko’s conjecture, J. Lond. Math. Soc. (2) 98 (2018), 593–608.
- [15] David Conlon and Joonkyung Lee, Finite reflection groups and graph norms, Adv. Math. 315 (2017), 130–165.
- [16] Dominique De Caen and Zoltán Füredi, The maximum size of 3-uniform hypergraphs not containing a Fano plane, J. Combin. Theory Ser. B 78 (2000), 274–276.
- [17] Pál Erdős, On the graph theorem of Turán, Mat. Lapok 21 (1970), 249–251.
- [18] Victor Falgas-Ravry and Emil R. Vaughan, Applications of the semi-definite method to the Turán density problem for 3-graphs, Combin. Probab. Comput. 22 (2013), 21–54.
- [19] Matthew Fitch, Applications of entropy to extremal problems, Ph.D. thesis, University of Warwick, 2018.
- [20] D. G. Fon-Der-Flaass, A method for constructing -graphs, Mat. Zametki 44 (1988), 546–550, 559.
- [21] P. Frankl, Asymptotic solution of a Turán-type problem, Graphs Combin. 6 (1990), 223–227.
- [22] P. Frankl and Z. Füredi, An exact result for -graphs, Discrete Math. 50 (1984), 323–328.
- [23] P. Frankl and Z. Füredi, Extremal problems whose solutions are the blowups of the small Witt-designs, J. Combin. Theory Ser. A 52 (1989), 129–147.
- [24] P. Frankl and V. Rödl, Hypergraphs do not jump, Combinatorica 4 (1984), 149–159.
- [25] Peter Frankl and Zoltán Füredi, A new generalization of the Erdős-Ko-Rado theorem, Combinatorica 3 (1983), 341–349.
- [26] Ehud Friedgut and Jeff Kahn, On the number of copies of one hypergraph in another, Israel J. Math. 105 (1998), 251–256.
- [27] Justin Gilmer, A constant lower bound for the union-closed sets conjecture, arXiv:2211.09055.
- [28] W. T. Gowers, Ben Green, Freddie Manners, and Terence Tao, On a conjecture of marton, to appear on Ann. of Math.
- [29] Andrzej Grzesik, Joonkyung Lee, Bernard Lidický, and Jan Volec, On tripartite common graphs, Combin. Probab. Comput. 31 (2022), 907–923.
- [30] Liying Kang, Lele Liu, and Erfang Shan, Sharp lower bounds for the spectral radius of uniform hypergraphs concerning degrees, Electron. J. Combin. 25 (2018), Paper No. 2.1, 13.
- [31] Liying Kang and Vladimir Nikiforov, Extremal problems for the -spectral radius of graphs, Electron. J. Combin. 21 (2014), Paper 3.21, 23.
- [32] Peter Keevash, Hypergraph Turán problems, Surveys in combinatorics 392 (2011), 83–140.
- [33] Peter Keevash, John Lenz, and Dhruv Mubayi, Spectral extremal problems for hypergraphs, SIAM J. Discrete Math. 28 (2014), 1838–1854.
- [34] Peter Keevash and Benny Sudakov, On a hypergraph Turán problem of Frankl, Combinatorica 25 (2005), 673–706.
- [35] A. V. Kostochka, A class of constructions for Turán’s -problem, Combinatorica 2 (1982), 187–192.
- [36] Joonkyung Lee, On some graph densities in locally dense graphs, Random Structures Algorithms 58 (2021), 322–344.
- [37] J.L. Xiang Li and Balázs Szegedy, On the logarithimic calculus and Sidorenko’s conjecture, arXiv:1107.1153.
- [38] Shuo-Yen Robert Li and Wen Ch’ing Winnie Li, Independence numbers of graphs and generators of ideals, Combinatorica 1 (1981), 55–61.
- [39] Xizhi Liu, Dhruv Mubayi, and Christian Reiher, A unified approach to hypergraph stability, J. Combin. Theory Ser. B 158 (2023), 36–62.
- [40] T. S. Motzkin and E. G. Straus, Maxima for graphs and a new proof of a theorem of Turán, Canadian J. Math. 17 (1965), 533–540.
- [41] Dhruv Mubayi, A hypergraph extension of Turán’s theorem, J. Combin. Theory Ser. B 96 (2006), 122–134.
- [42] Dhruv Mubayi and Oleg Pikhurko, A new generalization of Mantel’s theorem to -graphs, J. Combin. Theory Ser. B 97 (2007), 669–678.
- [43] Dhruv Mubayi and Vojtêch Rödl, On the Turán number of triple systems, J. Combin. Theory Ser. A 100 (2002), 136–152.
- [44] V. Nikiforov, Some inequalities for the largest eigenvalue of a graph, Combin. Probab. Comput. 11 (2002), 179–189.
- [45] Vladimir Nikiforov, Walks and the spectral radius of graphs, Linear Algebra Appl. 418 (2006), 257–268.
- [46] S. Norin and L. Yepremyan, Turán number of generalized triangles, J. Combin. Theory Ser. A 146 (2017), 312–343.
- [47] Sergey Norin and Liana Yepremyan, Turán numbers of extensions, J. Combin. Theory Ser. A 155 (2018), 476–492.
- [48] Olaf Parczyk, On Sidorenko’s conjecture, Ph.D. thesis, Master’s thesis, Freie Universität, Berlin, 2014.
- [49] Oleg Pikhurko, An exact Turán result for the generalized triangle, Combinatorica 28 (2008), 187–208.
- [50] Oleg Pikhurko, Exact computation of the hypergraph Turán function for expanded complete 2-graphs, J. Combin. Theory Ser. B 103 (2013), 220–225.
- [51] Liqun Qi, Symmetric nonnegative tensors and copositive tensors, Linear Algebra Appl. 439 (2013), 228–238.
- [52] Jaikumar Radhakrishnan, An entropy proof of Bregman’s theorem, J. Combin. Theory Ser. A 77 (1997), 161–164.
- [53] Alexander A. Razborov, On 3-hypergraphs with forbidden 4-vertex configurations, SIAM J. Discrete Math. 24 (2010), 946–963.
- [54] Alexander A. Razborov, Flag algebras: an interim report, The mathematics of Paul Erdős. II, Springer, New York, 2013, pp. 207–232.
- [55] Maya Sankar, The Turán density of 4-uniform tight cycles, arXiv:2411.01782.
- [56] C. E. Shannon, A mathematical theory of communication, Bell System Tech. J. 27 (1948), 379–423, 623–656.
- [57] A. F. Sidorenko, Asymptotic solution for a new class of forbidden -graphs, Combinatorica 9 (1989), 207–215.
- [58] Alexander Sidorenko, Asymptotic solution of the Turán problem for some hypergraphs, Graphs Combin. 8 (1992), 199–201.
- [59] Maya Stein, Tree containment and degree conditions, Discrete mathematics and applications, Springer Optim. Appl., vol. 165, Springer, Cham, [2020] ©2020, pp. 459–486.
- [60] Balázs Szegedy, An information theoretic approach to Sidorenko’s conjecture, arXiv:1406.6738.
- [61] Gary Tiner and Zachery Tomlin, On the Erdős-Sós conjecture for k = 9, Alabama Journal of Mathematics 45 (2022), 37–45.
- [62] Gy. Turán, On the greedy algorithm for an edge-partitioning problem, Theory of algorithms (Pécs, 1984), Colloq. Math. Soc. János Bolyai, vol. 44, North-Holland, Amsterdam, 1985, pp. 405–423.
- [63] Victor K Wei, A lower bound on the stability number of a simple graph, Bell Lab. Tech. Memor. (1981).
- [64] Herbert S. Wilf, Spectral bounds for the clique and independence numbers of graphs, J. Combin. Theory Ser. B 40 (1986), 113–117.
- [65] A. A. Zykov, On some properties of linear complexes, Mat. Sbornik N.S. 24/66 (1949), 163–188.
- [66] A. A. Zykov, On some properties of linear complexes, Amer. Math. Soc. Translation 1952 (1952), 33.
Appendix A Explicit relation between and in Theorem 8.3
In this appendix, we will relate positive integers with satisfying the inequality
(A.1) |
We first compute the right hand side.
Lemma A.1.
Suppose that are positive integers satisfying A.1. Then , and
Proof.
We first show . This is clear as
Now we show that . This is clear when , so it suffices to check the case when . In this case, we have . This forces for some constant , and so as .
Now let be the error term defined by
where we set . Note that is positive and increasing in when . Therefore
for any . This shows that
which shows that . Therefore
Proposition A.2.
Let be a positive integer growing with so that for some constant . Then the largest positive integer satisfying A.1 also satisfies .
Proof.
By the choice of , we know
and
Therefore
By Lemma A.1, we know that this implies
Rearranging, we get
and so
where we use the fact that . The desired statement thus follows. ∎
Proposition A.3.
Let be positive integers with , and let . Then the smallest positive integer satisfying A.1 also satisfies .