This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

When entropy meets Turán: new proofs and hypergraph Turán results

Ting-Wei Chao*  and  Hung-Hsun Hans Yu
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.

*Department of Mathematics, Massachusetts Institute of Technology, Cambridge, MA, USA. Email: [email protected]
Department of Mathematics, Princeton University, Princeton, NJ 08544. Email: [email protected]

1. Introduction

For any kk-graph (i.e. kk-uniform hypergraph) FF, its Turán number ex(n,F)\operatorname{ex}(n,F) is the maximum number of edges in an FF-free kk-graph GG on nn vertices. Here, GG is FF-free if it contains no subgraph (not necessarily induced) isomorphic to FF. The study of Turán numbers was initiated by Turán [62], who first considered the case where k=2k=2 and FF is the complete graph Kr+1K_{r+1} on (r+1)(r+1) vertices. There, Turán showed that ex(n,F)\operatorname{ex}(n,F) is maximized by the balanced complete rr-partite graph Tn,rT_{n,r}, 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 k=2k=2. Although the exact value of ex(n,F)\operatorname{ex}(n,F) is not known for general graphs FF, the celebrated Erdős–Stone theorem asserts that ex(n,F)=(11r+o(1))(n2)\operatorname{ex}(n,F)=\left(1-\frac{1}{r}+o(1)\right)\binom{n}{2} if χ(F)=r+1\chi(F)=r+1, where Tn,rT_{n,r} is an asymptotic extremizer. If we define the Turán density to be

π(F)=limnex(n,F)(nk)\pi(F)=\lim_{n\to\infty}\frac{\operatorname{ex}(n,F)}{\binom{n}{k}}

for a kk-graph FF, then the Erdős–Stone theorem can be rephrased as π(F)=11χ(F)1\pi(F)=1-\frac{1}{\chi(F)-1} when FF is a graph. It is worth pointing out that when χ(F)=2\chi(F)=2, Erdős–Stone gives that π(F)=0\pi(F)=0, showing that ex(n,F)\operatorname{ex}(n,F) is subquadratic but does not determine the asymptotic behavior of ex(n,F)\operatorname{ex}(n,F). Despite lots of effort, there are still many interesting open problems regarding the asymptotic behavior of ex(n,F)\operatorname{ex}(n,F) when FF is bipartite. However, in this paper, we will focus on the non-degenerate case where π(F)>0\pi(F)>0.

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 π(F)\pi(F) is still unknown even for F=K4(3)F=K_{4}^{(3)}, the 33-uniform clique on 44 vertices. Turán showed that π(K4(3))59\pi(K_{4}^{(3)})\geq\frac{5}{9} and conjectured that it is actually an equality. However, proving this conjecture still seems hard to date, and the current best upper bound π(F)0.561666\pi(F)\leq 0.561666 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 π(K4(3))59\pi(K_{4}^{(3)})\geq\frac{5}{9} given in [35] (also see [20]). In comparison, the Erdős–Simonovits theorem states that any asymptotic extremizer of π(Kr+1)\pi(K_{r+1}) should be close to Tn,rT_{n,r}. 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. π(Kr+1)=11r\pi(K_{r+1})=1-\frac{1}{r}. 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 KrK_{r} from the graph. Erdős [17] later provided another proof that modified the graph step by step, maintaining the Kr+1K_{r+1}-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 Tn,rT_{n,r} 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 GG 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 L(G)L(G) of a graph G=(V,E)G=(V,E) is defined as

max{u,v}Exuxv subj. to xv0vV and vVxv=1.\max\sum_{\{u,v\}\in E}x_{u}x_{v}\textup{ subj. to }x_{v}\geq 0\quad\forall v\in V\textup{ and }\sum_{v\in V}x_{v}=1.

Despite its somewhat long definition, it is a natural quantity to consider in the context of Turán problems. To see this, let NN be some large positive integers. Consider the blowup of GG obtained by putting in (xv+o(1))N(x_{v}+o(1))N copies of each vertex vVv\in V so that there are NN vertices in total, where (xv)vV(x_{v})_{v\in V} is the extremizer for the Lagrangian. Then there are (L(G)+o(1))N2(L(G)+o(1))N^{2} edges in the blowup. On the other hand, it is clear that |E|L(G)|V|2\left\lvert E\right\rvert\leq L(G)\left\lvert V\right\rvert^{2}, which shows that the density Turán theorem is equivalent to that L(G)12(11r)L(G)\leq\frac{1}{2}\left(1-\frac{1}{r}\right) for every Kr+1K_{r+1}-free graph GG. Motzkin and Straus’ idea is that if uu and vv are not adjacent, then there is an extremizer with either xu=0x_{u}=0 or xv=0x_{v}=0 for L(G)L(G). Therefore if GG is Kr+1K_{r+1}-free, then there is an extremizer with support of size at most rr. 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 kk-graph G=(V,E)G=(V,E), its hypergraph Lagrangian L(G)L(G) is defined as the maximum of {u1,,uk}Exu1xuv\sum_{\{u_{1},\ldots,u_{k}\}\in E}x_{u_{1}}\cdots x_{u_{v}} under the same condition. As before, when each vVv\in V is blown-up to (xv+o(1))N(x_{v}+o(1))N vertices where (xv)vV(x_{v})_{v\in V} is the extremizer for the Lagrangian, there are (L(G)+o(1))Nk(L(G)+o(1))N^{k} 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 b(G)=k!L(G)b(G)=k!L(G) to be the blowup density of GG. Intuitively, it is the largest edge density of the blowups of GG. 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

f((xv)vV(G))=uvE(xuxv)f\left((x_{v})_{v\in V(G)}\right)=\prod_{uv\not\in E}(x_{u}-x_{v})

for any graph GG. The key observation is that if GG is Kr+1K_{r+1}-free, then ff vanishes whenever r+1r+1 of the variables (xv)vV(G)(x_{v})_{v\in V(G)} are equal to one another. In light of this, let II be the ideal of polynomials that vanish whenever r+1r+1 of the variables are equal. Then fIf\in I, and Turán’s theorem follows from an explicit description of the generators of II 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 Kr+1K_{r+1}-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 𝔽2\mathbb{F}_{2}) [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 F,GF,G that are graphs or general kk-graphs, a homomorphism from FF to GG is a function f:V(F)V(G)f:V(F)\to V(G) that sends edges of FF to edges of GG. In particular, ff must be injective on any edge of FF. The homomorphism density t(F,G)t(F,G) is the probability that a uniformly random chosen function from V(F)V(G)V(F)\to V(G) is actually a homomorphism. In this terminology, a corollary of the Kruskal–Katona theorem says that t(K3,G)t(K2,G)32t(K_{3},G)\leq t(K_{2},G)^{\frac{3}{2}}, 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 FF are Sidorenko, i.e. t(F,G)t(K2,G)e(F)t(F,G)\geq t(K_{2},G)^{e(F)}. 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 t(Kr+1,G)=0t(K_{r+1},G)=0 then t(K2,G)11rt(K_{2},G)\leq 1-\frac{1}{r}. 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 (X)\mathbb{H}(X) to denote the Shannon entropy of a random variable XX (see Section 3 for definitions and basic properties).

Theorem 1.1 (Entropic Turán theorem).

Let rr be a positive integer, and let GG be a Kr+1K_{r+1}-free graph. Let X,YX,Y be random variables distributed on V(G)V(G) so that {X,Y}\{X,Y\} is always an edge in GG. Assume X,YX,Y are symmetric, i.e. the distribution of (X,Y)(X,Y) and the one of (Y,X)(Y,X) are the same. Then

(X,Y)2(X)+log2(11r).\mathbb{H}(X,Y)\leq 2\mathbb{H}(X)+\log_{2}\left(1-\frac{1}{r}\right).

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 GG, the maximum of the quantity (X,Y)2(X)\mathbb{H}(X,Y)-2\mathbb{H}(X) for symmetric V(G)V(G)-valued random variables X,YX,Y with {X,Y}E(G)\{X,Y\}\in E(G) is related to the blowup density b(G)b(G) of GG. More surprisingly, the maximum of (X,Y)(X)\mathbb{H}(X,Y)-\mathbb{H}(X) is related to the spectral radius ρ(G)\rho(G) of GG. 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 r2r\geq 2 and TT be a tree with 1\ell\geq 1 vertices. For any Kr+1K_{r+1}-free graph GG, we have

ρ(G)(11r)#{homomorphisms from T to G}.\rho(G)^{\ell}\leq\left(1-\frac{1}{r}\right)\#\{\text{homomorphisms from $T$ to $G$}\}.

To see that this is indeed a generalization of Wilf’s and Nikiforov’s results, we can take TT to be the path PP_{\ell} on \ell vertices. Wilf’s result corresponds to =1\ell=1, whereas Nikiforov’s results correspond to =2\ell=2 and general \ell.

Theorem 1.3 ([64, 44, 45]).

Let r2r\geq 2. For any Kr+1K_{r+1}-free graph GG with nn vertices and mm edges, we have

ρ(G)(11r)n,\rho(G)\leq\left(1-\frac{1}{r}\right)n,
ρ(G)2(11r)2m,\rho(G)^{2}\leq\left(1-\frac{1}{r}\right)\cdot 2m,

and

ρ(G)(11r)w(G),\rho(G)^{\ell}\leq\left(1-\frac{1}{r}\right)w_{\ell}(G),

where w(G)w_{\ell}(G) denotes the number of \ell-walks in GG.

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 kk-graphs \mathcal{F}, its Turán number ex(n,)\textup{ex}(n,\mathcal{F}) is defined to be the maximum number of edges in a kk-graph GG that is FF-free for every FF\in\mathcal{F}. The Turán density is defined analogously by π()=limnex(n,)/(nk)\pi(\mathcal{F})=\lim_{n\to\infty}\textup{ex}(n,\mathcal{F})/\binom{n}{k}. For any family of kk-graphs \mathcal{F} and a kk-graph GG, we say that GG is \mathcal{F}-hom-free if there does not exist any homomorphism FGF\to G for every FF\in\mathcal{F}. A FF-hom-free kk-graph is simply a kk-graph that is {F}\{F\}-hom-free.

It is a standard result in the field that π()\pi(\mathcal{F}) is the supremum of b(G)b(G) where GG runs through all \mathcal{F}-hom-free kk-graphs (see [32, Section 2] or [55, Lemma 2.2] for example). Notice that a single edge has blowup density k!/kkk!/k^{k}, showing that b(G)k!/kkb(G)\geq k!/k^{k} if GG is not empty. This immediately shows that either π()=0\pi(\mathcal{F})=0 or π()k!/kk\pi(\mathcal{F})\geq k!/k^{k} for any family of kk-graphs \mathcal{F}. We see that among the possible values of Turán density, there is a “jump” going from 0 to k!/kkk!/k^{k}. When k=2k=2, this is indeed the behavior of Turán densities: the Erdős–Stone theorem shows that all possible values are 0,12,23,34,0,\frac{1}{2},\frac{2}{3},\frac{3}{4},\ldots, 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 k3k\geq 3, there are infinitely many non-jumps δ\delta, where for every ε>0\varepsilon>0 there exists a family \mathcal{F} of kk-graphs with π()(δ,δ+ε)\pi(\mathcal{F})\in(\delta,\delta+\varepsilon). On the other hand, Baber and Talbot [3] showed that jumps do exist above k!/kkk!/k^{k} when k=3k=3. However, our understanding in jumps and non-jumps is still limited, and we do not even know whether k!/kkk!/k^{k} is a jump.

A standard argument shows that k!/kkk!/k^{k} is a jump if and only if there exists a finite family \mathcal{F} of kk-graph with π()=k!/kk\pi(\mathcal{F})=k!/k^{k} and b(F)>k!/kkb(F)>k!/k^{k} for each FF\in\mathcal{F} (see [24]). The fact that we do not know whether k!/kkk!/k^{k} is a jump can thus be seen as a result of not having sufficient understanding in the families \mathcal{F} with π()=k!/kk\pi(\mathcal{F})=k!/k^{k}. Indeed, known families with Turán densities equal to k!/kkk!/k^{k} are so few that we can list them here. For general kk, Mubayi [41] showed that the kk-uniform extended clique Ek+1(k)E^{(k)}_{k+1} of size k+1k+1 has Turán density k!/kkk!/k^{k}. 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, Ek+1(k)E^{(k)}_{k+1} is obtained by adding k2k-2 extra vertices to each edge of Kk+1K_{k+1}, where no two edges share any extra vertices. This was later generalized by Mubayi and Pikhurko [42], who showed that the hypergraph Δ(1,1,,1)\Delta_{(1,1,\ldots,1)} with edges

{v1,,vk} and {w,vi,u1(i),,uk2(i)} for i[k]\left\{v_{1},\ldots,v_{k}\right\}\text{ and }\{w,v_{i},u^{(i)}_{1},\ldots,u^{(i)}_{k-2}\}\text{ for }i\in[k]

also has Turán density k!/kkk!/k^{k}. 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 Ek+1(k)E^{(k)}_{k+1} is homomorphic to Δ(1,1,,1)\Delta_{(1,1,\ldots,1)}, showing that Ek+1(k)E^{(k)}_{k+1}-hom-free graphs are also Δ(1,1,,1)\Delta_{(1,1,\ldots,1)}-hom-free and so π(Ek+1(k))π(Δ(1,1,,1))\pi(E^{(k)}_{k+1})\leq\pi(\Delta_{(1,1,\ldots,1)}).

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 π(Δ(1,,1))=k!/kk\pi(\Delta_{(1,\ldots,1)})=k!/k^{k} as well—this is an observation by Keevash [32, Theorem 3.1].

For smaller kk’s, slightly more is known. When k=3k=3, Bollobás [6] showed that π({K4,F5})=29\pi(\{K_{4}^{-},F_{5}\})=\frac{2}{9} where K4={123,124,134}K_{4}^{-}=\{123,124,134\} and F5={123,124,345}F_{5}=\{123,124,345\}. This was improved by Frankl and Füredi [25], who showed that π(F5)\pi(F_{5}) is already equal to 29\frac{2}{9}. Using flag algebra, Baber and Talbot [4] improved this further by showing that π({123,124,345,156})=29\pi(\{123,124,345,156\})=\frac{2}{9}. Finally, when k=4k=4, Pikhurko [49] showed that π({1234,1235,4567})=332\pi(\{1234,1235,4567\})=\frac{3}{32}.

As shown above, not a lot is known about families \mathcal{F} of kk-graphs with π()=k!/kk\pi(\mathcal{F})=k!/k^{k}. As an application of our entropic proof of the density Turán theorem, we will generalize our argument to show π()=k!/kk\pi(\mathcal{F})=k!/k^{k} for a new family \mathcal{F} of kk-graphs. Our method has a benefit that we may first come up with an argument and then see what family of kk-graphs need to be forbidden in order for the argument to work. We believe that this advantage can help discovering more families \mathcal{F} with minimum positive Turán densities.

BaseApex
Figure 1. (3,2)(3,2)-tent

To state our result, for any partition λ\lambda of kk, let λ=(λ1,,λ)\lambda=(\lambda_{1},\ldots,\lambda_{\ell}) where =(λ)\ell=\ell(\lambda) is the length of λ\lambda, and λ1λ\lambda_{1}\geq\cdots\geq\lambda_{\ell}. We also denote i=1λi\sum_{i=1}^{\ell}\lambda_{i} by |λ|\left\lvert\lambda\right\rvert (which is equal to kk by definition). For any λ\lambda with (λ)2\ell(\lambda)\geq 2, we define the λ\lambda-tent, denoted by Δλ\Delta_{\lambda}, to be the following kk-graph. The λ\lambda-tent comes with an edge ee that is the base and a vertex vv that is the apex. Setting =(λ)\ell=\ell(\lambda) to be the length of λ\lambda, for each i[]i\in[\ell] we also have an edge eie_{i} containing vv such that |eie|=λi\left\lvert e_{i}\cap e\right\rvert=\lambda_{i}. Moreover, we require that eiej={v}e_{i}\cap e_{j}=\{v\} for any ij[]i\neq j\in[\ell]. It is clear that this determines Δλ\Delta_{\lambda} uniquely up to isomorphism—in fact, we must have ee1,,eee\cap e_{1},\ldots,e\cap e_{\ell} partition ee. It is easy to check that this definition matches the definition of Δ(1,1,,1)\Delta_{(1,1,\ldots,1)} above, F5=Δ(2,1)F_{5}=\Delta_{(2,1)} (with base 123123 and 44 being the apex) and Pikhurko’s result can be rephrased as π(Δ(3,1))=332\pi(\Delta_{(3,1)})=\frac{3}{32}. Our result can now be stated as follows.

Theorem 1.4.

Let k2k\geq 2 be a positive integer, and let k\mathcal{F}_{k} be the family of λ\lambda-tents with |λ|=k\left\lvert\lambda\right\rvert=k and (λ)=2\ell(\lambda)=2. Then π(k)=k!/kk\pi(\mathcal{F}_{k})=k!/k^{k}.

Note that this is a stronger statement than Mubayi’s and Mubayi and Pikhurko’s results. In fact, Δ(1,1,,1)\Delta_{(1,1,\ldots,1)} admits a homomorphism to Δλ\Delta_{\lambda} for every |λ|=k\left\lvert\lambda\right\rvert=k and (λ)=2\ell(\lambda)=2, which shows that π(Δ(1,1,,1))π(k)\pi(\Delta_{(1,1,\ldots,1)})\leq\pi(\mathcal{F}_{k}). Using the same argument, we can transform Theorem 1.4 into a Turán result of a single kk-graph.

Theorem 1.5.

Let k2k\geq 2 be a positive integer, and let λ\lambda be a partition of kk such that λ1k/2\lambda_{1}\leq\lceil k/2\rceil and λi=1\lambda_{i}=1 for all 1<i(λ)1<i\leq\ell(\lambda). Then π(Δλ)=k!/kk\pi(\Delta_{\lambda})=k!/k^{k}.

Although when k=3k=3 and 44, Theorem 1.5 is subsumed by the known results mentioned above, this gives a new Turán result for larger kk’s. To show that this should be a nontrivial result for larger kk’s, we prove the following result in the opposite direction.

Theorem 1.6.

There exists a constant α<1\alpha<1 so that for all sufficiently large kk\in\mathbb{N} and any partition λ\lambda of kk with (λ)2\ell(\lambda)\geq 2, if λ1>αk\lambda_{1}>\alpha k then π(Δλ)>k!/kk\pi(\Delta_{\lambda})>k!/k^{k}.

Theorem 1.5 shows that the constant in Theorem 1.6 cannot be smaller than 1/21/2, and it seems like an interesting question to determine the best possible value of α\alpha. It might help us understand the kk-graphs FF with π(F)=k!/kk\pi(F)=k!/k^{k} as well. We leave this as a future direction for interested readers.

Beyond showing π()=k!/kk\pi(\mathcal{F})=k!/k^{k} for various families \mathcal{F} of kk-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 H,GH,G, let t(H,G)t(H,G) be the homomorphism density of HH in GG. That is, t(H,G)t(H,G) is the probability that a function f:V(H)V(G)f:V(H)\rightarrow V(G) chosen uniformly at random is a homomorphism from HH to GG. 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 i0i\geq 0, let Si=K1,iS_{i}=K_{1,i} be the star with i+1i+1 vertices. Then

t(Si,G)t(K2,G)it(S_{i},G)\geq t(K_{2},G)^{i}

holds for any graph GG.

Proof.

Assume n=|V(G)|n=\left\lvert V(G)\right\rvert and m=|E(G)|m=\left\lvert E(G)\right\rvert. Note that SiS_{i} has i+1i+1 vertices, and hence

t(Si,G)=vV(G)deg(v)ini+11ni(vV(G)deg(v)n)i=(2m)in2i=t(K2,G)i,t(S_{i},G)=\frac{\sum_{v\in V(G)}\deg(v)^{i}}{n^{i+1}}\geq\frac{1}{n^{i}}\left(\frac{\sum_{v\in V(G)}\deg(v)}{n}\right)^{i}=\frac{(2m)^{i}}{n^{2i}}=t(K_{2},G)^{i},

where the inequality follows from the convexity of xix^{i}. ∎

Now we assume the graph GG is Kr+1K_{r+1}-free. We sample a sequence of i.i.d. random vertices v0,v1,v_{0},v_{1},\dots from V(G)V(G) uniformly at random. For i0i\geq 0, let AiA_{i} be the event that the induced graph on vertices v0,,vi1,viv_{0},\dots,v_{i-1},v_{i} contains SiS_{i} as a subgraph centered at viv_{i}. In particular, A0A_{0} is the true event. Note that there can only be at most rr events happening at the same time. Otherwise, assume Ai0,Ai1,,AirA_{i_{0}},A_{i_{1}},\dots,A_{i_{r}} are all true for some 0=i0<i1<<ir0=i_{0}<i_{1}<\dots<i_{r}. Then vi0,,virv_{i_{0}},\dots,v_{i_{r}} form an (r+1)(r+1)-clique in GG. Therefore, by double counting, we may conclude that

(A0)+(A1)+r.\mathbb{P}(A_{0})+\mathbb{P}(A_{1})+\dots\leq r.

On the other hand, we know that (Ai)=t(Si,G)t(K2,G)i\mathbb{P}(A_{i})=t(S_{i},G)\geq t(K_{2},G)^{i} for all ii. Thus, we have

11t(K2,G)(A0)+(A1)+r.\frac{1}{1-t(K_{2},G)}\leq\mathbb{P}(A_{0})+\mathbb{P}(A_{1})+\dots\leq r.

After rearranging, we get

2mn2=t(K2,G)11r,\frac{2m}{n^{2}}=t(K_{2},G)\leq 1-\frac{1}{r},

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 XX, we write pX(x)=def(X=x)p_{X}(x)\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\mathbb{P}(X=x). Also, we denote by supp(X)\operatorname{supp}(X) the support of XX, i.e. the set of all xx such that pX(x)>0p_{X}(x)>0. Through out this paper, the random variables we will consider are always discrete with finite support, i.e. |supp(X)|<\left\lvert\operatorname{supp}(X)\right\rvert<\infty. For any such random variable, we may define its Shannon entropy.

Definition 3.1.

For any random variable XX, we define its Shannon entropy

(X)=defxsupp(X)pX(x)log2pX(x).\mathbb{H}(X)\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\sum_{x\in\operatorname{supp}(X)}-p_{X}(x)\log_{2}p_{X}(x).

For any sequence of random variables X1,,XnX_{1},\dots,X_{n}, we use (X1,,Xn)\mathbb{H}(X_{1},\dots,X_{n}) to denote the entropy of the random tuple (X1,,Xn)(X_{1},\dots,X_{n}).

We also define the conditional entropy of XX given YY.

Definition 3.2.

For any two random variables X,YX,Y, the conditional entropy of XX given YY is given by

(XY)=def(X,Y)(Y).\mathbb{H}(X\mid Y)\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\mathbb{H}(X,Y)-\mathbb{H}(Y).

Equivalently, we have

(XY)=\displaystyle\mathbb{H}(X\mid Y)= ysupp(Y)pY(y)(XY=y)\displaystyle\quad\smashoperator[]{\sum_{y\in\operatorname{supp}(Y)}^{}}p_{Y}(y)\mathbb{H}(X\mid Y=y)
=\displaystyle= (x,y)supp(X,Y)pX,Y(x,y)log2(pX,Y(x,y)pY(y)).\displaystyle\quad\smashoperator[]{\sum_{(x,y)\in\operatorname{supp}(X,Y)}^{}}-p_{X,Y}(x,y)\log_{2}\left(\frac{p_{X,Y}(x,y)}{p_{Y}(y)}\right).

Using the definition of conditional entropy, we have the following chain rule.

Proposition 3.3 (Chain rule).

For any random variables X1,,XnX_{1},\dots,X_{n}, we have

(X1,,Xn)=(X1)+(X2X1)++(XnX1,,Xn1).\mathbb{H}(X_{1},\dots,X_{n})=\mathbb{H}(X_{1})+\mathbb{H}(X_{2}\mid X_{1})+\dots+\mathbb{H}(X_{n}\mid X_{1},\dots,X_{n-1}).

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 XX, we have

(X)log2|supp(X)|,\mathbb{H}(X)\leq\log_{2}\left\lvert\operatorname{supp}(X)\right\rvert,

where the equality holds if and only if XX is uniform.

We will also need the following two propositions about entropy.

Proposition 3.5 (Subadditivity).

For any three random variables X,Y,ZX,Y,Z, we have

(X,Y)(X)+(Y),\mathbb{H}(X,Y)\leq\mathbb{H}(X)+\mathbb{H}(Y),
(X,YZ)(XZ)+(YZ).\mathbb{H}(X,Y\mid Z)\leq\mathbb{H}(X\mid Z)+\mathbb{H}(Y\mid Z).
Proposition 3.6 (Dropping condition).

For any three random variables X,Y,ZX,Y,Z, we have

(XY)(X),\mathbb{H}(X\mid Y)\leq\mathbb{H}(X),
(XY,Z)(XZ).\mathbb{H}(X\mid Y,Z)\leq\mathbb{H}(X\mid Z).

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 X1,,XnX_{1},\dots,X_{n} and weights w1,,wn0w_{1},\dots,w_{n}\geq 0 with i=1nwi=1\sum_{i=1}^{n}w_{i}=1, we say that ZZ is the mixture of X1,,XnX_{1},\dots,X_{n} with weight w1,,wnw_{1},\dots,w_{n} if ZZ is obtained from the following procedure. We first pick an independent random index 𝐢\mathbf{i} with probability (𝐢=i)=wi\mathbb{P}(\mathbf{i}=i)=w_{i}. Then we set Z=X𝐢Z=X_{\mathbf{i}}.

In our applications, we will consider mixtures of random variables whose supports do not overlap too much.

Definition 3.8.

Let aa be a positive integer. We say that the random variables X1,,XnX_{1},\dots,X_{n} have (a+1)(a+1)-wise disjoint supports if for any element xi=1nsupp(Xi)x\in\cup_{i=1}^{n}\operatorname{supp}(X_{i}), there are at most aa many indices ii such that xsupp(Xi)x\in\operatorname{supp}(X_{i}).

With the definitions above, we may state our lemma about an upper bound on the entropies of random variables with (a+1)(a+1)-wise disjoint supports, in terms of the entropy of their mixture.

Lemma 3.9 (Mixture bound).

Let X1,,XnX_{1},\dots,X_{n} be random variables with (a+1)(a+1)-wise disjoint supports. Then there exists a mixture of X1,,XnX_{1},\dots,X_{n}, say ZZ, such that

i=1n2(Xi)a2(Z).\sum_{i=1}^{n}2^{\mathbb{H}(X_{i})}\leq a2^{\mathbb{H}(Z)}.
Proof.

Let si=2(Xi)s_{i}=2^{\mathbb{H}(X_{i})} and we define

wi=sij=1nsj.w_{i}=\frac{s_{i}}{\sum_{j=1}^{n}s_{j}}.

Let 𝐢\mathbf{i} be an independent random index with probability (𝐢=i)=wi\mathbb{P}(\mathbf{i}=i)=w_{i} and let Z=X𝐢Z=X_{\mathbf{i}} be the mixture. By chain rule, we have (Z,𝐢)=(𝐢)+(Z𝐢)=(Z)+(𝐢Z)\mathbb{H}(Z,\mathbf{i})=\mathbb{H}(\mathbf{i})+\mathbb{H}(Z\mid\mathbf{i})=\mathbb{H}(Z)+\mathbb{H}(\mathbf{i}\mid Z). Therefore,

(Z)=(𝐢)+(Z𝐢)(𝐢Z).\mathbb{H}(Z)=\mathbb{H}(\mathbf{i})+\mathbb{H}(Z\mid\mathbf{i})-\mathbb{H}(\mathbf{i}\mid Z).

By the definition of entropy and conditional entropy, we have

(𝐢)=i=1nwilog2wi=sii=1nsilog2(sii=1nsi)\mathbb{H}(\mathbf{i})=\sum_{i=1}^{n}-w_{i}\log_{2}w_{i}=\frac{-s_{i}}{\sum_{i=1}^{n}s_{i}}\log_{2}\bigl{(}\frac{s_{i}}{\sum_{i=1}^{n}s_{i}}\bigr{)}

and

(Z𝐢)=i=1nwi(Xi)=silog2sii=1nsi.\mathbb{H}(Z\mid\mathbf{i})=\sum_{i=1}^{n}w_{i}\mathbb{H}(X_{i})=\frac{s_{i}\log_{2}s_{i}}{\sum_{i=1}^{n}s_{i}}.

We may upper bound (iZ)\mathbb{H}(i\mid Z) by uniform bound. For any xi=1nsupp(Xi)x\in\cup_{i=1}^{n}\operatorname{supp}(X_{i}), when conditioning on Z=xZ=x, there are at most aa possible indices as an outcome of 𝐢\mathbf{i}. Thus, we have

(𝐢Z)log2a.\mathbb{H}(\mathbf{i}\mid Z)\leq\log_{2}a.

Combining all above, we get

(Z)\displaystyle\mathbb{H}(Z)\geq sii=1nsilog2(sii=1nsi)+silog2sii=1nsilog2a\displaystyle\frac{-s_{i}}{\sum_{i=1}^{n}s_{i}}\log_{2}\bigl{(}\frac{s_{i}}{\sum_{i=1}^{n}s_{i}}\bigr{)}+\frac{s_{i}\log_{2}s_{i}}{\sum_{i=1}^{n}s_{i}}-\log_{2}a
=\displaystyle= log2(i=1nsi)log2a,\displaystyle\log_{2}\left(\sum_{i=1}^{n}s_{i}\right)-\log_{2}a,

and we are done after rearranging. ∎

The following example shows that Lemma 3.9 resembles a double counting on (a+1)(a+1)-wise disjoint sets. Thus, the mixture bound can be viewed as an entropic version of this double counting.

Example 3.10.

Let aa be an integer and let S1,,SnS_{1},\dots,S_{n} be some sets that are (a+1)(a+1)-wise disjoint. Assume XiX_{i} is a random element chosen from SiS_{i} uniform at random for each i[n]i\in[n], and let ZZ be the mixture of X1,,XnX_{1},\ldots,X_{n} provided by Lemma 3.9. We have 2(Xi)=|Si|2^{\mathbb{H}(X_{i})}=\left\lvert S_{i}\right\rvert, and by uniform bound we have 2(Z)|i=1nSi|2^{\mathbb{H}(Z)}\leq\left\lvert\cup_{i=1}^{n}S_{i}\right\rvert. Hence, Lemma 3.9 implies that

i=1n|Si|a2(Z)a|i=1nSi|,\sum_{i=1}^{n}\left\lvert S_{i}\right\rvert\leq a2^{\mathbb{H}(Z)}\leq a\left\lvert\bigcup_{i=1}^{n}S_{i}\right\rvert,

which gives the same bound as the double counting argument on pairs (x,i)(x,i) with xSix\in S_{i}.

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 Kr+1K_{r+1}-free graph GG and symmetric random variables X,YX,Y distributed on V(G)V(G) with {X,Y}E(G)\{X,Y\}\in E(G) always holding. We first fix an integer NN\in\mathbb{N}, and we will take NN goes to infinity later.

Claim 4.1.

For each i=0,1,,Ni=0,1,\dots,N, there exists a random tuple Ti=(v0(i),,vN(i))V(G)N+1T_{i}=(v_{0}^{(i)},\dots,v_{N}^{(i)})\in V(G)^{N+1} such that

  1. (1)

    there is always an edge between vj(i),vi(i)v_{j}^{(i)},v_{i}^{(i)} for all j=0,,i1j=0,\dots,i-1,

  2. (2)

    the marginal distributions of vj(i)v_{j}^{(i)} and XX are the same for all j=0,1,Nj=0,1\dots,N, and

  3. (3)

    (Ti)=i(YX)+(N+1i)(X)\mathbb{H}(T_{i})=i\mathbb{H}(Y\mid X)+(N+1-i)\mathbb{H}(X).

Proof.

For i=0i=0, it is easy to check that N+1N+1 i.i.d. random vertices v0(0),,vN(0)v_{0}^{(0)},\dots,v_{N}^{(0)} with the law of XX satisfy the condition.

For i1i\geq 1, we first sample an edge (v0(i),vi(i))(v_{0}^{(i)},v_{i}^{(i)}) using the law of (X,Y)(X,Y). Next, we condition on vi(i)v_{i}^{(i)} and resample v0(i)v_{0}^{(i)} (i1)(i-1) times conditionally independently to get v1(i),,vi1(i)v_{1}^{(i)},\dots,v_{i-1}^{(i)}. Finally, we sample vi+1(i),,vN(i)v_{i+1}^{(i)},\dots,v_{N}^{(i)} independently using the law of XX.

Note that the first two conditions are true from the way we sample the random variables. It remains to compute (Ti)\mathbb{H}(T_{i}). Note that (Ti)=(v0(i),,vi(i))+(Ni)(X)\mathbb{H}(T_{i})=\mathbb{H}(v_{0}^{(i)},\dots,v_{i}^{(i)})+(N-i)\mathbb{H}(X) since we sampled vi+1(i),,vN(i)v_{i+1}^{(i)},\dots,v_{N}^{(i)} independently. By chain rule, we have

(v0(i),,vi(i))=\displaystyle\mathbb{H}(v_{0}^{(i)},\dots,v_{i}^{(i)})= (v0(i),,vi1(i)vi(i))+(vi(i))\displaystyle\mathbb{H}(v_{0}^{(i)},\dots,v_{i-1}^{(i)}\mid v_{i}^{(i)})+\mathbb{H}(v_{i}^{(i)})
=\displaystyle= i(v0(i)vi(i))+(vi(i))\displaystyle i\mathbb{H}(v_{0}^{(i)}\mid v_{i}^{(i)})+\mathbb{H}(v_{i}^{(i)})
=\displaystyle= i(YX)+(X).\displaystyle i\mathbb{H}(Y\mid X)+\mathbb{H}(X).

Therefore, (Ti)=i(YX)+(N+1i)(X)\mathbb{H}(T_{i})=i\mathbb{H}(Y\mid X)+(N+1-i)\mathbb{H}(X). ∎

Now, we may apply Lemma 3.9 to the random tuples T0,,TNT_{0},\dots,T_{N} in 4.1. Since GG is Kr+1K_{r+1}-free, similar to the proof in Section 2, any tuple of N+1N+1 vertices is in at most rr supports supp(Ti)\operatorname{supp}(T_{i}). Therefore, the supports of T0,,TNT_{0},\dots,T_{N} are (r+1)(r+1)-wise disjoint. Thus, there is a mixture T=(v0,,vN)T=(v_{0},\dots,v_{N}) of T0,,TNT_{0},\dots,T_{N} such that

i=0N2(Ti)r2(T).\sum_{i=0}^{N}2^{\mathbb{H}(T_{i})}\leq r2^{\mathbb{H}(T)}.

Note that the marginal distribution of viv_{i} is also the same as the marginal distribution of XX, so we may upper bound (T)\mathbb{H}(T) by (N+1)(X)(N+1)\mathbb{H}(X) by subadditivity. By using (Ti)=i(YX)+(N+1i)(X)\mathbb{H}(T_{i})=i\mathbb{H}(Y\mid X)+(N+1-i)\mathbb{H}(X), we get

i=0Nxir,\sum_{i=0}^{N}x^{i}\leq r,

where x=def2(YX)(X)x\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}2^{\mathbb{H}(Y\mid X)-\mathbb{H}(X)}. By taking NN to infinity, we conclude that 1/(1x)r1/(1-x)\leq r. Therefore,

(YX)(X)=log2xlog2(11r).\mathbb{H}(Y\mid X)-\mathbb{H}(X)=\log_{2}x\leq\log_{2}\left(1-\frac{1}{r}\right).\qed

Let |V(G)|=n\left\lvert V(G)\right\rvert=n and |E(G)|=m\left\lvert E(G)\right\rvert=m. If we pick (X,Y)(X,Y) uniformly at random from all the oriented edges, Theorem 1.1 and the uniform bound give

log2(2m)=(X,Y)2(X)+log2(11r)2log2n+log2(11r).\log_{2}(2m)=\mathbb{H}(X,Y)\leq 2\mathbb{H}(X)+\log_{2}\left(1-\frac{1}{r}\right)\leq 2\log_{2}n+\log_{2}\left(1-\frac{1}{r}\right).

That is, m(11r)n22m\leq\left(1-\frac{1}{r}\right)\frac{n^{2}}{2}, 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 (X,Y)2(X)\mathbb{H}(X,Y)-2\mathbb{H}(X) when (X,Y)(X,Y) ranges over all possible symmetric distributions on the oriented edges of GG. This quantity turns out to be related to the blowup density b(G)b(G). To extend this to hypergraphs, we make the following definitions.

Definition 5.1 (Random edge with uniform ordering).

Let GG be a kk-graph, we say that a tuple of random vertices (X1,,Xk)V(G)k(X_{1},\dots,X_{k})\in V(G)^{k} is a random edge with uniform ordering on GG if (X1,,Xk)(X_{1},\dots,X_{k}) is symmetric and {X1,,Xk}\{X_{1},\dots,X_{k}\} is always an edge of GG. Here, (X1,,Xk)(X_{1},\dots,X_{k}) being symmetric means the distribution of (Xσ(1),,Xσ(k))(X_{\sigma(1)},\dots,X_{\sigma(k)}) is always the same for any permutation σ\sigma of [n][n].

Definition 5.2 (Entropic density).

For any kk-graph GG, define its entropic density bentropy(G)b_{\textup{entropy}}(G) to be the largest possible value of 2(X1,,Xk)k(X1)2^{\mathbb{H}(X_{1},\ldots,X_{k})-k\mathbb{H}(X_{1})} for any random edge with uniform ordering (X1,,Xk)(X_{1},\ldots,X_{k}).

Note that bentropy(G)b_{\textup{entropy}}(G) exists as the space of random edge with uniform ordering is compact. We will show that bentropy(G)b_{\textup{entropy}}(G) is equal to b(G)b(G), 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 kk-graph GG, let E(G)\vec{E}(G) be the set of oriented edges. For each p>0p>0, let bp(G)b_{p}(G) be the maximum of (v1,,vk)E(G)xv1xvk\prod_{(v_{1},\ldots,v_{k})\in\vec{E}(G)}x_{v_{1}}\cdots x_{v_{k}} for (xv)vV(G)(x_{v})_{v\in V(G)} subject to xvp=1\left\lVert x_{v}\right\rVert_{\ell^{p}}=1 (the same definition was made by Keevash, Lenz and Mubayi [33] where they called the quantity the pp-spectral radius). Also let bp,entropy(G)b_{p,\textup{entropy}}(G) be the largest possible value of 2(X1,,Xk)kp(X1)2^{\mathbb{H}(X_{1},\ldots,X_{k})-\frac{k}{p}\mathbb{H}(X_{1})} for any random edge with uniform ordering (X1,,Xk)(X_{1},\ldots,X_{k}). Note that bp(G)b_{p}(G) and bp,entropy(G)b_{p,\textup{entropy}}(G) both exist by compactness.

Example 5.3.

When p=1p=1, we clearly have bp(G)=b(G)b_{p}(G)=b(G) and bp,entropy(G)=bentropy(G)b_{p,\textup{entropy}}(G)=b_{\textup{entropy}}(G). When GG is a graph and p=2p=2, it is not hard to see that bp(G)b_{p}(G) is the maximum

maxxAGx subject to (xv)vV(G)2=1\max\vec{x}^{\intercal}A_{G}\vec{x}\textup{ subject to }\left\lVert(x_{v})_{v\in V(G)}\right\rVert_{\ell^{2}}=1

where AGA_{G} is the adjacency matrix of GG. It is a standard fact that this is exactly the spectral radius of GG. In this case, b2,entropy(G)b_{2,\textup{entropy}}(G) is the largest possible value of 2(X,Y)(X)=2(YX)2^{\mathbb{H}(X,Y)-\mathbb{H}(X)}=2^{\mathbb{H}(Y\mid X)} for any random edge with uniform ordering (X,Y)(X,Y).

For general kk, if p=kp=k, then bp(G)b_{p}(G) corresponds to the spectral radius of the adjacency kk-tensor of GG, which was proven in [51]. The quantity bk,entropy(G)b_{k,\textup{entropy}(G)} is the largest possible value of 2(X1,,Xk)(X1)=2(X2,,XkX1)2^{\mathbb{H}(X_{1},\ldots,X_{k})-\mathbb{H}(X_{1})}=2^{\mathbb{H}(X_{2},\ldots,X_{k}\mid X_{1})}. Once we prove bk(G)=bk,entropy(G)b_{k}(G)=b_{k,\textup{entropy}(G)}, this would provide a nice alternative interpretation of the spectral radius for hypergraphs.

Now we will show that bp(G)b_{p}(G) and bp,entropy(G)b_{p,\textup{entropy}}(G) are equal to each other. The proof uses Lagrange multiplier in a crucial way.

Proposition 5.4.

For any kk-graph GG and any p>0p>0, bp,entropy(G)=bp(G)b_{p,\textup{entropy}}(G)=b_{p}(G).

Proof.

For any vV(G)v\in V(G), let Lv(G)\vec{L}_{v}(G) be the oriented link of vv, i.e. the set (v2,,vk)(v_{2},\ldots,v_{k}) such that (v,v2,,vk)E(G)(v,v_{2},\ldots,v_{k})\in\vec{E}(G).

We start with the following claim that helps us simplify (X1,,Xk)kp(X1)\mathbb{H}(X_{1},\dots,X_{k})-\frac{k}{p}\mathbb{H}(X_{1}) when (X1,,Xk)(X_{1},\dots,X_{k}) is in a certain form.

Claim 5.5.

For any tuple (xv)vV(G)0V(G)(x_{v})_{v\in V(G)}\in\mathbb{R}_{\geq 0}^{V(G)}, we consider a random edge with uniform ordering (X1,,Xk)(X_{1},\dots,X_{k}) on GG given by

((X1,,Xk)=(v1,,vk))=1βi=1kxvi, where β=def(v1,,vk)E(G)i=1kxvi.\mathbb{P}((X_{1},\dots,X_{k})=(v_{1},\dots,v_{k}))=\frac{1}{\beta}\prod_{i=1}^{k}x_{v_{i}},\text{ where }\beta\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\sum_{(v_{1},\dots,v_{k})\in\vec{E}(G)}\prod_{i=1}^{k}x_{v_{i}}.

We also define

yv=def(X1=v)=xvβ(v2,,vk)Lv(G)i=2kxvi.y_{v}\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\mathbb{P}(X_{1}=v)=\frac{x_{v}}{\beta}\sum_{(v_{2},\dots,v_{k})\in\vec{L}_{v}(G)}\prod_{i=2}^{k}x_{v_{i}}.

Then we have

(X1,,Xk)kp(X1)=log2βkpvV(G)yvlog2(xvpyv).\mathbb{H}(X_{1},\dots,X_{k})-\frac{k}{p}\mathbb{H}(X_{1})=\log_{2}\beta-\frac{k}{p}\sum_{v\in V(G)}y_{v}\log_{2}\left(\frac{x_{v}^{p}}{y_{v}}\right).
Proof.

First, we have

(X1,,Xk)=\displaystyle\mathbb{H}(X_{1},\dots,X_{k})= (v1,,vk)E(G)1βi=1kxvilog2(1βi=1kxvi)\displaystyle\sum_{(v_{1},\dots,v_{k})\in\vec{E}(G)}-\frac{1}{\beta}\prod_{i=1}^{k}x_{v_{i}}\log_{2}\left(\frac{1}{\beta}\prod_{i=1}^{k}x_{v_{i}}\right)
=\displaystyle= (v1,,vk)E(G)1βi=1kxvi(log2βi=1klog2xvi)\displaystyle\sum_{(v_{1},\dots,v_{k})\in\vec{E}(G)}\frac{1}{\beta}\prod_{i=1}^{k}x_{v_{i}}\left(\log_{2}\beta-\sum_{i=1}^{k}\log_{2}x_{v_{i}}\right)
=\displaystyle= log2βkvV(G)yvlog2xv\displaystyle\log_{2}\beta-k\sum_{v\in V(G)}y_{v}\log_{2}x_{v}

Combining this with (X1)=vV(G)yvlog2yv\mathbb{H}(X_{1})=\sum_{v\in V(G)}-y_{v}\log_{2}y_{v}, we get

(X1,,Xk)kp(X1)=\displaystyle\mathbb{H}(X_{1},\dots,X_{k})-\frac{k}{p}\mathbb{H}(X_{1})= log2βkpvV(G)(pyvlog2xvyvlog2yv)\displaystyle\log_{2}\beta-\frac{k}{p}\sum_{v\in V(G)}\left(py_{v}\log_{2}x_{v}-y_{v}\log_{2}y_{v}\right)
=\displaystyle= log2βkpvV(G)yvlog2(xvpyv).\displaystyle\log_{2}\beta-\frac{k}{p}\sum_{v\in V(G)}y_{v}\log_{2}\left(\frac{x_{v}^{p}}{y_{v}}\right).\qed

Now, we may prove the proposition. We first show that bp,entropy(G)bp(G)b_{p,\textup{entropy}}(G)\geq b_{p}(G).

Let (xv)vV(G)0V(G)(x_{v})_{v\in V(G)}\in\mathbb{R}_{\geq 0}^{V(G)} be the tuple that achieves the maximum in the definition of bp(G)b_{p}(G). Define (X1,,Xk)(X_{1},\dots,X_{k}), β\beta, and (yv)vV(G)(y_{v})_{v\in V(G)} in the same way as in 5.5. Note that β=bp(G)\beta=b_{p}(G) and vV(G)xvp=1\sum_{v\in V(G)}x_{v}^{p}=1. From 5.5, we have

(X1,,Xk)kp(X1)=\displaystyle\mathbb{H}(X_{1},\dots,X_{k})-\frac{k}{p}\mathbb{H}(X_{1})= log2βkpvV(G)yvlog2(xvpyv)\displaystyle\log_{2}\beta-\frac{k}{p}\sum_{v\in V(G)}y_{v}\log_{2}\left(\frac{x_{v}^{p}}{y_{v}}\right)
\displaystyle\geq log2βkplog2(vV(G)xvp)=log2β,\displaystyle\log_{2}\beta-\frac{k}{p}\log_{2}\left(\sum_{v\in V(G)}x_{v}^{p}\right)=\log_{2}\beta,

where the inequality follows from the Jensen’s inequality and the concavity of log2x\log_{2}x. Therefore bp,entropy(G)bp(G)b_{p,\textup{entropy}}(G)\geq b_{p}(G).

For the opposite direction, let (X1,,Xk)(X_{1},\ldots,X_{k}) be a random edge with uniform ordering achieving the maximum of bp,entropy(G)b_{p,\textup{entropy}}(G). For any unoriented edge eE(G)e\in E(G), let qeq_{e} be the probability ({X1,,Xk}=e)\mathbb{P}(\{X_{1},\ldots,X_{k}\}=e). Also let xv=(1kevqe)1/px_{v}=\left(\frac{1}{k}\sum_{e\ni v}q_{e}\right)^{1/p}. Then

(X1,,Xk)=(X1,,Xk{X1,,Xk})+({X1,,Xk})=log2k!eE(G)qelog2qe\mathbb{H}(X_{1},\ldots,X_{k})=\mathbb{H}(X_{1},\ldots,X_{k}\mid\{X_{1},\ldots,X_{k}\})+\mathbb{H}(\{X_{1},\ldots,X_{k}\})=\log_{2}k!-\sum_{e\in E(G)}q_{e}\log_{2}q_{e}

and

(X1)=vVxvplog2xvp.\mathbb{H}(X_{1})=\sum_{v\in V}-x_{v}^{p}\log_{2}x_{v}^{p}.

Therefore, (qe)eE(G)(q_{e})_{e\in E(G)} is a maximizer of

eE(G)qelog2qe+kpvV(G)xvplog2xvp-\sum_{e\in E(G)}q_{e}\log_{2}q_{e}+\frac{k}{p}\sum_{v\in V(G)}x_{v}^{p}\log_{2}x_{v}^{p}

subject to qe0q_{e}\geq 0 for all eE(G)e\in E(G) and eE(G)qe=1\sum_{e\in E(G)}q_{e}=1. Note that xvp/qe\partial x_{v}^{p}/\partial q_{e} is nonzero only if vev\in e, and if that is the case we have xvp/qe=1/k\partial x_{v}^{p}/\partial q_{e}=1/k. By Lagrange multiplier, we know that

log2qe1+1pve(1+log2xvp)-\log_{2}q_{e}-1+\frac{1}{p}\sum_{v\in e}\left(1+\log_{2}x_{v}^{p}\right)

is constant for all eE(G)e\in E(G) with qe>0q_{e}>0. Therefore

α=defqevexv\alpha\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\frac{q_{e}}{\prod_{v\in e}x_{v}}

is the same for all eE(G)e\in E(G) with qe>0q_{e}>0. Notice that (X1=v)=xvp\mathbb{P}(X_{1}=v)=x_{v}^{p} for any vV(G)v\in V(G), and for any (v1,,vk)E(G)(v_{1},\ldots,v_{k})\in\vec{E}(G), we have

((X1,,Xk)=(v1,,vk))=qek!=αk!i=1kxvi.\mathbb{P}((X_{1},\ldots,X_{k})=(v_{1},\ldots,v_{k}))=\frac{q_{e}}{k!}=\frac{\alpha}{k!}\prod_{i=1}^{k}x_{v_{i}}.

Therefore, using 5.5 with β=k!/α\beta=k!/\alpha, we see that

(X1,,Xk)kp(X1)=\displaystyle\mathbb{H}(X_{1},\dots,X_{k})-\frac{k}{p}\mathbb{H}(X_{1})= log2βkpvV(G)yvlog2(xvpyv),\displaystyle\log_{2}\beta-\frac{k}{p}\sum_{v\in V(G)}y_{v}\log_{2}\left(\frac{x_{v}^{p}}{y_{v}}\right),

where, in this case, yv=xvpy_{v}=x_{v}^{p}. Thus, (X1,,Xk)kp(X1)=log2β\mathbb{H}(X_{1},\dots,X_{k})-\frac{k}{p}\mathbb{H}(X_{1})=\log_{2}\beta. Note that vV(G)xvp=1\sum_{v\in V(G)}x_{v}^{p}=1. Therefore by the fact that

β=(v1,,vk)E(G)i=1kxvi,\beta=\sum_{(v_{1},\ldots,v_{k})\in\vec{E}(G)}\prod_{i=1}^{k}x_{v_{i}},

we have bp,entropy(G)bp(G)b_{p,\textup{entropy}}(G)\leq b_{p}(G). ∎

Corollary 5.6.

For any family \mathcal{F} of kk-graphs, π()\pi(\mathcal{F}) is the supremum of 2(X1,,Xk)k(X1)2^{\mathbb{H}(X_{1},\ldots,X_{k})-k\mathbb{H}(X_{1})} for any random edge with uniform ordering (X1,,Xk)(X_{1},\ldots,X_{k}) on any \mathcal{F}-hom-free kk-graph GG.

Proof.

Since π()\pi(\mathcal{F}) is the supremum of b(G)b(G) for all \mathcal{F}-hom-free kk-graphs GG, we know that π()\pi(\mathcal{F}) is the supremum of bentropy(G)b_{\textup{entropy}}(G) for all \mathcal{F}-hom-free kk-graphs GG as well. The statement follows from the definition of entropic density bentropy(G)b_{\textup{entropy}}(G). ∎

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 GG is Kr+1K_{r+1}-free, then GG is Kr+1K_{r+1}-hom-free. This is clear as any homomorphic image of Kr+1K_{r+1} is Kr+1K_{r+1}. ∎

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 p=2p=2, 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 TT.

Proof of Theorem 1.2.

From Proposition 5.4 and the observation in Example 5.3, there exists a random edge with uniform ordering (X,Y)(X,Y) on GG such that log2ρ(G)=(YX)\log_{2}\rho(G)=\mathbb{H}(Y\mid X). By Theorem 1.1, we have

log2ρ(G)=(YX)(X)+(1)(YX)+log2(11r).\ell\log_{2}\rho(G)=\ell\mathbb{H}(Y\mid X)\leq\mathbb{H}(X)+(\ell-1)\mathbb{H}(Y\mid X)+\log_{2}\left(1-\frac{1}{r}\right).

Let v1,,vv_{1},\ldots,v_{\ell} be an ordering of the vertices of TT where for every i{2,,}i\in\{2,\ldots,\ell\}, the vertex viv_{i} is adjacent to exactly one vjv_{j} with j<ij<i. Now, we sample random vertices X1,,XX_{1},\dots,X_{\ell} in GG as follows. Let X1X_{1} be a random vertex sampled using the law of XX. Assume we have already sampled X1,,Xi1X_{1},\dots,X_{i-1}, and assume vjv_{j} is the neighbor of viv_{i} with j<ij<i. We sample XiX_{i} conditionally independently such that XiXjYXX_{i}\mid X_{j}\sim Y\mid X. It follows that X1,,XX_{1},\dots,X_{\ell} is always a homomorphic image of TT in GG. Also, from the way we sample, we know that (X1,,X)=(X)+(1)(YX)\mathbb{H}(X_{1},\dots,X_{\ell})=\mathbb{H}(X)+(\ell-1)\mathbb{H}(Y\mid X). Thus, we have

(X)+(1)(YX)=(X1,,X)log2#{homomorphisms from T to G},\mathbb{H}(X)+(\ell-1)\mathbb{H}(Y\mid X)=\mathbb{H}(X_{1},\dots,X_{\ell})\leq\log_{2}\#\{\text{homomorphisms from $T$ to $G$}\},

and we are done by combining this with the previous inequality and rearranging. ∎

For general pp, recall that our definition of bp(G)b_{p}(G) matches the definition of pp-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 r2r\geq 2 be a positive integer and p1p\geq 1 be a real number. For any Kr+1K_{r+1}-free graph GG with nn vertices and mm edges, we have

bp(G)(11r)n22/p,b_{p}(G)\leq\left(1-\frac{1}{r}\right)n^{2-2/p},

and

bp(G)(11r)1/p(2m)11/p.b_{p}(G)\leq\left(1-\frac{1}{r}\right)^{1/p}(2m)^{1-1/p}.
Proof.

From Proposition 5.4, there exists a random edge with uniform ordering (X,Y)(X,Y) on GG such that log2bp(G)=(X,Y)2p(X)\log_{2}b_{p}(G)=\mathbb{H}(X,Y)-\frac{2}{p}\mathbb{H}(X). We have

(X,Y)2p(X)(22p)(X)+log2(11r)(22p)log2n+(11r),\mathbb{H}(X,Y)-\frac{2}{p}\mathbb{H}(X)\leq\left(2-\frac{2}{p}\right)\mathbb{H}(X)+\log_{2}\left(1-\frac{1}{r}\right)\leq\left(2-\frac{2}{p}\right)\log_{2}n+\left(1-\frac{1}{r}\right),

and

(X,Y)2p(X)(11p)(X,Y)+1plog2(11r)(11p)log2(2m)+1p(11r).\mathbb{H}(X,Y)-\frac{2}{p}\mathbb{H}(X)\leq\left(1-\frac{1}{p}\right)\mathbb{H}(X,Y)+\frac{1}{p}\log_{2}\left(1-\frac{1}{r}\right)\leq\left(1-\frac{1}{p}\right)\log_{2}(2m)+\frac{1}{p}\left(1-\frac{1}{r}\right).\qed

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 kk-graph FF is a simplicial complex whose faces have size at most kk. Its set of vertices is denoted by V(F)V(F), and its set of faces, or partial edges, is denoted by E(F)E(F). A homomorphism from a partial kk-graph FF to a kk-graph GG is a map f:V(F)V(G)f:V(F)\to V(G) such that for any partial edge eE(F)e\in E(F), ff is injective on ee and f(e)f(e) is contained in some edge in E(G)E(G). Now for any partial kk-graph FF, its extension F~\tilde{F} is the kk-graph obtained as follows: first let EE^{\prime} be the set of maximal partial edges in E(F)E(F), and then extend each partial edge in EE^{\prime} to a kk-edge by adding in extra vertices, where two different edges do not share any extra vertices. Notice that if FF is a simplicial complex generated by edges of some kk^{\prime}-graph FF^{\prime} with k<kk^{\prime}<k, then F~\tilde{F} is the extension of FF^{\prime} as defined in the introduction.

Example 6.1 (Definition of partial tents).

In Section 7, the partial kk-graphs and the corresponding extensions of concern would be the following. For any partition λ\lambda of kk with =def(λ)2\ell\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}\ell(\lambda)\geq 2, the partial λ\lambda-tent Δλp\Delta^{p}_{\lambda} is the partial kk-graph obtained by taking the simplicial complex generated by Δλ\Delta_{\lambda}, and then restricting it to e{v}e\cup\{v\} where ee is the base and vv is the apex. It is easy to verify that Δλ\Delta_{\lambda} is the extension of the partial kk-graph Δλp\Delta^{p}_{\lambda}.

extension
Figure 2. Partial (3,2)(3,2)-tent and its extension. Note that for the partial tent, only the maximal edges are shown.

Those definitions are useful as for any partial kk-graph FF, a homomorphism FGF\to G is essentially the same as a homomorphism F~G\tilde{F}\to G. This would be helpful later as instead of considering homomorphisms from Δλ\Delta_{\lambda}, we can consider homomorphisms from Δλp\Delta^{p}_{\lambda}, which are easier to describe.

Proposition 6.2.

Let FF be a partial kk-graph, and let GG be a kk-graph. Then there is a homomorphic image of FF in GG if and only if there is a homomorphic image of F~\tilde{F} in GG.

Proof.

Let (F~)cpx(\tilde{F})_{\textup{cpx}} be the simplicial complex generated by the edges in F~\tilde{F}, which we will think of as a partial kk-graph. Then FF is the restriction of (F~)cpx(\tilde{F})_{\textup{cpx}} on V(F)V(F). For any homomorphism f:V(F~)V(G)f:V(\tilde{F})\to V(G) from F~\tilde{F} to GG, we also have that it is an homomorphism from (F~)cpx(\tilde{F})_{\textup{cpx}} to GG. It is then easy to check that f|V(F)f|_{V(F)} is a homomorphism from FF to GG.

Conversely, suppose that g:V(F)V(G)g:V(F)\to V(G) is a homomorphism from FF to GG. Note that for each eE(F~)e\in E(\tilde{F}), we know that eV(F)E((F~)cpx)e\cap V(F)\in E((\tilde{F})_{\textup{cpx}}) and so eV(F)e\cap V(F) is in E(F)E(F) as well. By the definition, this implies that for every eE(F~)e\in E(\tilde{F}), we have that gg is injective on eV(F)e\cap V(F) and g(eV(F))g(e\cap V(F)) is contained in some edge in GG. As any vertex in V(F~)\V(F)V(\tilde{F})\backslash V(F) is in exactly one edge in E(F~)E(\tilde{F}), it is possible to extend gg to g~:V(F~)V(G)\tilde{g}:V(\tilde{F})\to V(G) so that g(e)g(e) is an edge in GG for each eE(F~)e\in E(\tilde{F}). The extended map g~\tilde{g} is indeed a homomorphism from F~\tilde{F} to GG. ∎

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 kk-graph FF, any linear order << on V(F)V(F), and any vertex vV(F)v\in V(F), let MF,<(v)M_{F,<}(v) be the set of partial edges whose maximum vertex is vv. A partial kk-graph FF is a partial forest with respect to a linear order << on V(F)V(F) if for every vV(F)v\in V(F), there is exactly one maximal partial edge eve_{v} in MF,<(v)M_{F,<}(v). In this case, the forest sequence of (F,<)(F,<) is a sequence (n1,,nk)(n_{1},\ldots,n_{k}) where for each i[k]i\in[k], nin_{i} is the number of vertices vV(F)v\in V(F) with |ev|=i\left\lvert e_{v}\right\rvert=i.

We also define quantities that are analogs of the quantity 2(YX)(X)2^{\mathbb{H}(Y\mid X)-\mathbb{H}(X)} we used in Section 4.

Definition 6.4 (Ratio sequence).

Let (X1,,Xk)V(G)k(X_{1},\dots,X_{k})\in V(G)^{k} be a random edge with uniform ordering on a kk-graph GG. We define the ratio sequence 0<x1xk=10<x_{1}\leq\dots\leq x_{k}=1 of (X1,,Xk)(X_{1},\dots,X_{k}) by xi=2(XiXi+1,,Xk)(Xi)x_{i}=2^{\mathbb{H}(X_{i}\mid X_{i+1},\dots,X_{k})-\mathbb{H}(X_{i})} for each i[k]i\in[k].

We are now ready to apply Szegedy’s argument to sample homomorphisms from partial forests with high entropy.

Lemma 6.5.

Let (X1,,Xk)(X_{1},\ldots,X_{k}) be a random edge with uniform ordering on a kk-graph GG and let x1,,xkx_{1},\ldots,x_{k} be its ratio sequence. For any partial forest FF with a linear order <<, if (n1,,nk)(n_{1},\ldots,n_{k}) is its forest sequence, then one can sample a random homomorphism (Yv)vV(F)(Y_{v})_{v\in V(F)} from FF to GG with entropy equal to

v(F)H(X1)+log2(i=1kxink+1i).v(F)H(X_{1})+\log_{2}\left(\prod_{i=1}^{k}x_{i}^{n_{k+1-i}}\right).

Moreover, the random homomorphism can be sampled such that for any partial edge eE(F)e\in E(F), the distribution of (Yv)ve(Y_{v})_{v\in e} is the same as (Xi)k|e|+1ik(X_{i})_{k-\left\lvert e\right\rvert+1\leq i\leq k}.

Proof.

We will induct on v(F)v(F). The case v(F)=0v(F)=0 is vacuously true. Now suppose that it holds for partial forest of size v(F)1v(F)-1. Let vmaxv_{\max} be the maximum vertex in V(F)V(F). Then F\{vmax}F\backslash\{v_{\max}\} is also a partial forest, and so we may sample a random homomorphism (Yv)vV(F)\{vmax}(Y_{v})_{v\in V(F)\backslash\{v_{\max}\}} with the prescribed properties. Let ee be the maximal partial edge in MF,<(vmax)M_{F,<}(v_{\textup{max}}), and let j=k+1|e|j=k+1-\left\lvert e\right\rvert. By the inductive hypothesis, (Yv)ve\vmax(Y_{v})_{v\in e\backslash v_{\max}} is identically distributed as (Xi)j+1ik(X_{i})_{j+1\leq i\leq k}. Therefore, we may sample YvmaxY_{v_{\max}} given (Yv)ve\vmax(Y_{v})_{v\in e\backslash v_{\max}} conditionally independently so that (Yv)ve(Y_{v})_{v\in e} is identically distributed as (Xi)jik(X_{i})_{j\leq i\leq k}. This way,

((Yv)vV(F))=\displaystyle\mathbb{H}\left((Y_{v})_{v\in V(F)}\right)= ((Yv)vV(F)\{vmax})+(Yvmax(Yv)ve\{vmax})\displaystyle\mathbb{H}\left((Y_{v})_{v\in V(F)\backslash\{v_{\max}\}}\right)+\mathbb{H}\left(Y_{v_{\max}}\mid(Y_{v})_{v\in e\backslash\{v_{\max}\}}\right)
=\displaystyle= (v(F)1)H(X1)+log2(xj1i=1kxink+1i)+H(XjXj+1,,Xk)\displaystyle\left(v(F)-1\right)H(X_{1})+\log_{2}\left(x_{j}^{-1}\prod_{i=1}^{k}x_{i}^{n_{k+1-i}}\right)+H(X_{j}\mid X_{j+1},\ldots,X_{k})
=\displaystyle= v(F)H(X1)+log2(i=1kxink+1i)\displaystyle v(F)H(X_{1})+\log_{2}\left(\prod_{i=1}^{k}x_{i}^{n_{k+1-i}}\right)

where we use that (Xi)=(X1)\mathbb{H}(X_{i})=\mathbb{H}(X_{1}) for any i[k]i\in[k]. It remains to show that for any partial edge ee^{\prime} containing vmaxv_{\max}, the distribution of (Yv)ve(Y_{v})_{v\in e^{\prime}} is the same as (Xi)k|e|+1ik(X_{i})_{k-\left\lvert e^{\prime}\right\rvert+1\leq i\leq k}. This is true as eee^{\prime}\subseteq e by the definition of ee and vmaxv_{\max}, and the distribution (X1,,Xk)(X_{1},\ldots,X_{k}) 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 kk-graph GG and a random edge with uniform ordering (X1,,Xk)(X_{1},\ldots,X_{k}) on GG. We will also set 0<x1xk=10<x_{1}\leq\cdots\leq x_{k}=1 to be its ratio sequence. We make an observation that to upper bound b(G)=bentropy(G)b(G)=b_{\textup{entropy}}(G), it suffices to upper bound2(X1,,Xk)(Xk)=x1xk12^{\mathbb{H}(X_{1},\ldots,X_{k})-\mathbb{H}(X_{k})}=x_{1}\cdots x_{k-1} by the chain rule. Therefore, the upper bound of Theorem 1.4 follows from the following statement.

Theorem 7.1.

If GG is λ\lambda-tent-hom-free for every |λ|=k\left\lvert\lambda\right\rvert=k and (λ)=2\ell(\lambda)=2, then we have

(X1,,Xk)k(X1)=log2(x1xk)log2k!kk.\mathbb{H}(X_{1},\dots,X_{k})-k\mathbb{H}(X_{1})=\log_{2}(x_{1}\cdots x_{k})\leq\log_{2}\frac{k!}{k^{k}}.

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 π(k)k!/kk\pi(\mathcal{F}_{k})\geq k!/k^{k} as a single edge does not contain any homomorphic image of any tents, and it has blowup density k!/kkk!/k^{k}. To show the reverse inequality, if GG is k\mathcal{F}_{k}-hom-free, then by Theorem 7.1, we have b(G)=bentropy(G)k!/kkb(G)=b_{\textup{entropy}}(G)\leq k!/k^{k}. This shows that π(k)k!/kk\pi(\mathcal{F}_{k})\leq k!/k^{k}. ∎

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 x1,,xkx_{1},\ldots,x_{k}. Then we will solve for the maximum of x1xk1x_{1}\cdots x_{k-1} subject to the inequalities.

Lemma 7.2.

If GG is λ\lambda-tent-hom-free for every |λ|=k\left\lvert\lambda\right\rvert=k and (λ)=2\ell(\lambda)=2, then for any i,j[k]i,j\in[k] with i+jki+j\leq k, we have xi+xjxi+jx_{i}+x_{j}\leq x_{i+j}.

Proof.

We will consider two partial forests F(1)F^{(1)} and F(2)F^{(2)} both on V={v1,,vk,w}V=\{v_{1},\ldots,v_{k},w\}. Let F(1)F^{(1)} be spanned by the two partial edges {v1,,vk}\{v_{1},\ldots,v_{k}\} and {vi+1,,vk,w}\{v_{i+1},\ldots,v_{k},w\}. Let F(2)F^{(2)} be spanned by the two partial edges {v1,,vk}\{v_{1},\ldots,v_{k}\} and {v1,,vkj,w}\{v_{1},\ldots,v_{k-j},w\}. Then both partial kk-graphs are indeed partial forests with respect to the linear order v1<<vk<wv_{1}<\cdots<v_{k}<w. It is clear that in F(1)F^{(1)} with the forest sequence (n1,,nk)(n_{1},\ldots,n_{k}), the vertices v1,,vkv_{1},\ldots,v_{k} contribute one to n1,,nkn_{1},\ldots,n_{k} and ww contributes to nki+1n_{k-i+1}. Similarly, the forest sequence of F(2)F^{(2)} is all-one except for nkj+1=2n_{k-j+1}=2.

Let (Yv(1))vV,(Yv(2))vV(Y^{(1)}_{v})_{v\in V},(Y^{(2)}_{v})_{v\in V} be the random homomorphism from F(1),F(2)F^{(1)},F^{(2)} given by Lemma 6.5, respectively. Note that if some tuple of vertices is in the supports of both (Yv(1))vV(Y^{(1)}_{v})_{v\in V} and (Yv(2))vV(Y^{(2)}_{v})_{v\in V}, then this tuple corresponds to a homomorphism from F(1)F(2)F^{(1)}\cup F^{(2)} to GG. As F(1)F(2)F^{(1)}\cup F^{(2)} clearly contains a partial (i,ki)(i,k-i)-tent with base {v1,,vk}\{v_{1},\ldots,v_{k}\} and apex ww, we know that the two random homomorphisms have disjoint support. Suppose that (Zv)vV(Z_{v})_{v\in V} is the mixture given by Lemma 3.9, then by Lemmas 3.9 and 6.5 we know

(x1xk1xi+x1xk1xj)2(k+1)(X1)2((Zv)vV).\left(x_{1}\cdots x_{k-1}\cdot x_{i}+x_{1}\cdots x_{k-1}\cdot x_{j}\right)2^{(k+1)\mathbb{H}(X_{1})}\leq 2^{\mathbb{H}((Z_{v})_{v\in V})}.

Observe that both F(1)F^{(1)} and F(2)F^{(2)} contains the partial edges {v1,,vk}\{v_{1},\ldots,v_{k}\} and {vi+1,,vkj,w}\{v_{i+1},\ldots,v_{k-j},w\}. Therefore (Yv1(1),,Yvk(1))(Y^{(1)}_{v_{1}},\ldots,Y^{(1)}_{v_{k}}) and (Yv1(2),,Yvk(2))(Y^{(2)}_{v_{1}},\ldots,Y^{(2)}_{v_{k}}) both have the same distributions as (X1,,Xk)(X_{1},\ldots,X_{k}) by Lemma 6.5, which shows that (Zv1,,Zvk)(Z_{v_{1}},\ldots,Z_{v_{k}}) has the same distribution as (X1,,Xk)(X_{1},\ldots,X_{k}) as well. Using a similar argument, we can show that (Zw,Zvi+1,,Zvkj)(Z_{w},Z_{v_{i+1}},\ldots,Z_{v_{k-j}}) has the same distribution as (Xi+j,,Xk)(X_{i+j},\ldots,X_{k}). As a consequence,

((Zv)vV)\displaystyle\mathbb{H}\left((Z_{v})_{v\in V}\right)\leq (Zv1,,Zvk)+(ZwZvi+1,,Zvkj)\displaystyle\mathbb{H}(Z_{v_{1}},\ldots,Z_{v_{k}})+\mathbb{H}(Z_{w}\mid Z_{v_{i+1}},\ldots,Z_{v_{k-j}})
=\displaystyle= (X1,,Xk)+(Xi+jXi+j+1,,Xk)\displaystyle\mathbb{H}(X_{1},\ldots,X_{k})+\mathbb{H}(X_{i+j}\mid X_{i+j+1},\ldots,X_{k})
=\displaystyle= (k+1)(X1)+log2(x1xk1xi+j).\displaystyle(k+1)\mathbb{H}(X_{1})+\log_{2}(x_{1}\cdots x_{k-1}\cdot x_{i+j}).

This shows that

x1xk12(k+1)(X1)(xi+xj)x1xk12(k+1)(X1)xi+jx_{1}\cdots x_{k-1}2^{(k+1)\mathbb{H}(X_{1})}(x_{i}+x_{j})\leq x_{1}\cdots x_{k-1}2^{(k+1)\mathbb{H}(X_{1})}\cdot x_{i+j}

and so the desired statement follows. ∎

Our next goal is to upper bound x1xk1x_{1}\cdots x_{k-1}. To upper bound the product, we prove the following auxiliary inequality.

Lemma 7.3.

Suppose that y1,,yky_{1},\ldots,y_{k} are some non-negative real numbers with yi+yjyi+jy_{i}+y_{j}\leq y_{i+j} for any i,j[k]i,j\in[k] with i+jki+j\leq k. Then

y1ykk!(y1++yk(k+12))k.y_{1}\cdots y_{k}\leq k!\left(\frac{y_{1}+\cdots+y_{k}}{\binom{k+1}{2}}\right)^{k}.
Proof.

We will prove this by induction. It clearly holds when k=1k=1. Now suppose that k2k\geq 2 and the statement holds for k1k-1. Then by the inductive hypothesis,

y1yk(k1)!(y1++yk1(k2))k1ykk!((k1)y1++yk1(k2)+ykkk)ky_{1}\cdots y_{k}\leq(k-1)!\left(\frac{y_{1}+\cdots+y_{k-1}}{\binom{k}{2}}\right)^{k-1}y_{k}\leq k!\left(\frac{(k-1)\cdot\frac{y_{1}+\cdots+y_{k-1}}{\binom{k}{2}}+\frac{y_{k}}{k}}{k}\right)^{k}

by AM-GM. Since

y1++yk1=12i=1k1(yi+yki)k12yk,y_{1}+\dots+y_{k-1}=\frac{1}{2}\sum_{i=1}^{k-1}(y_{i}+y_{k-i})\leq\frac{k-1}{2}y_{k},

we know

(k1)y1++yk1(k2)+ykk=2k(y1++yk1+yk2)2kkk+1(y1++yk)(k-1)\cdot\frac{y_{1}+\cdots+y_{k-1}}{\binom{k}{2}}+\frac{y_{k}}{k}=\frac{2}{k}\left(y_{1}+\cdots+y_{k-1}+\frac{y_{k}}{2}\right)\leq\frac{2}{k}\cdot\frac{k}{k+1}\left(y_{1}+\cdots+y_{k}\right)

and so

y1ykk!(2k+1(y1++yk)k)k=k!(y1++yk(k+12))k,y_{1}\cdots y_{k}\leq k!\left(\frac{\frac{2}{k+1}(y_{1}+\cdots+y_{k})}{k}\right)^{k}=k!\left(\frac{y_{1}+\cdots+y_{k}}{\binom{k+1}{2}}\right)^{k},

as desired. ∎

Combining Lemma 7.2 and Lemma 7.3, we are now ready to prove Theorem 7.1.

Proof of Theorem 7.1.

By Lemma 7.2, x1,,xkx_{1},\ldots,x_{k} are non-negative reals satisfying the condition of Lemma 7.3. We also know that xk=1x_{k}=1, so x1++xkk12+1=k+12x_{1}+\cdots+x_{k}\leq\frac{k-1}{2}+1=\frac{k+1}{2}. Thus by Lemma 7.3,

x1xk1=x1xkk!(k+12(k+12))k=k!kk,x_{1}\cdots x_{k-1}=x_{1}\cdots x_{k}\leq k!\left(\frac{\frac{k+1}{2}}{\binom{k+1}{2}}\right)^{k}=\frac{k!}{k^{k}},

which is the desired statement ∎

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 GG is λ\lambda-tent-hom-free for every |λ|=k\left\lvert\lambda\right\rvert=k and (λ)=2\ell(\lambda)=2, then for every i[k1]i\in[k-1], we have xj<xi+1x_{j}<x_{i+1} for each jij\leq i and

j=1ixjxi+1xj1.\prod_{j=1}^{i}\frac{x_{j}}{x_{i+1}-x_{j}}\leq 1.
Proof.

We will fix ii throughout this proof. As in what we did in Section 4, we will temporarily fix an integer NN\in\mathbb{N} that will later be taken to infinity. For any 1=t0<t1<t2<<tiN1=t_{0}<t_{1}<t_{2}<\cdots<t_{i}\leq N, we will define a partial forest F(t)F^{(\vec{t})} on V={v1,,vki1,w1,,wN}.V=\{v_{1},\ldots,v_{k-i-1},w_{1},\ldots,w_{N}\}. The partial forest F(t)F^{(\vec{t})} is spanned by the partial edges {wm,wtj+1,,wti}{v1,,vki1}\{w_{m},w_{t_{j+1}},\ldots,w_{t_{i}}\}\cup\{v_{1},\ldots,v_{k-i-1}\} for every tjm<tj+1t_{j}\leq m<t_{j+1}, where ti+1t_{i+1} is set to be N+1N+1. This is indeed a partial forest with respect to the linear order << with v1<<vki1<wN<<w1v_{1}<\cdots<v_{k-i-1}<w_{N}<\cdots<w_{1}. We can compute the forest sequence with respect to the linear order as follows: each vjv_{j} contributes one to njn_{j} for each jki1j\leq k-i-1, and each wmw_{m} with tjm<tj+1t_{j}\leq m<t_{j+1} contributes 11 to nkjn_{k-j}. Therefore the forest sequence (n1,,nk)(n_{1},\ldots,n_{k}) is (t1t0,,ti+1ti,1,,1)(t_{1}-t_{0},\ldots,t_{i+1}-t_{i},1,\ldots,1). Now let (Yv(t))vV(Y^{(\vec{t})}_{v})_{v\in V} be the random homomorphism produced by Lemma 6.5. This gives

((Yv(t))vV)=(N+ki1)(X1)+log2(xi+2xkji+1xjtjtj1).\displaystyle\mathbb{H}\left((Y^{(\vec{t})}_{v})_{v\in V}\right)=(N+k-i-1)\mathbb{H}(X_{1})+\log_{2}\left(x_{i+2}\cdots x_{k}\cdot\prod_{j\leq i+1}x_{j}^{t_{j}-t_{j-1}}\right). (7.1)

We will now show that the supports of (Yv(t))vV(Y^{(\vec{t})}_{v})_{v\in V} are disjoint for different choices of t\vec{t}. Suppose for the sake of contradiction that for some tt\vec{t}\neq\vec{t}^{\prime} there is a tuple of vertices from V(G)V(G) lying in the supports of (Yv(t))vV(Y^{(\vec{t})}_{v})_{v\in V} and (Yv(t))vV.(Y^{(\vec{t}^{\prime})}_{v})_{v\in V}. Then this tuple witnesses a homomorphism sending F(t)F(t)F^{(\vec{t})}\cup F^{(\vec{t}^{\prime})} to GG. We will show a contradiction by demonstrating that F(t)F(t)F^{(\vec{t})}\cup F^{(\vec{t}^{\prime})} contains a homomorphic image of some partial λ\lambda-tent with (λ)=2\ell(\lambda)=2.

Let j1j\geq 1 be the minimum index in which t\vec{t} and t\vec{t}^{\prime} differ, and without loss of generality, suppose that tj<tjt_{j}^{\prime}<t_{j}. Then we can find partial edges e={v1,,vki1,wt0,wt1,,wti}e=\{v_{1},\ldots,v_{k-i-1},w_{t_{0}},w_{t_{1}},\ldots,w_{t_{i}}\},e1={v1,,vki1,wtj,wtj,,wti}e_{1}=\{v_{1},\ldots,v_{k-i-1},w_{t_{j}^{\prime}},w_{t_{j}},\ldots,w_{t_{i}}\} in F(t)F^{(\vec{t})} and e2={wt0,,wtj}e_{2}=\{w_{t_{0}^{\prime}},\ldots,w_{t_{j}^{\prime}}\} in F(t)F^{(\vec{t}^{\prime})}. By the minimality of jj, we know e2={wt0,,wtj1,wtj}.e_{2}=\{w_{t_{0}},\ldots,w_{t_{j-1}},w_{t_{j}^{\prime}}\}. Note that e,e1,e2e,e_{1},e_{2} form a partial (kj,j)(k-j,j)-tent with base ee and apex wtjw_{t_{j}^{\prime}}, showing that F(t)F(t)F^{(\vec{t})}\cup F^{(\vec{t}^{\prime})} contains a partial (kj,j)(k-j,j)-tent, which is a contradiction.

Therefore we may now apply Lemma 3.9 with a=1a=1. Suppose that (Zv)vV(Z_{v})_{v\in V} is the resulting mixture of (Yv(t))vV(Y_{v}^{(\vec{t})})_{v\in V} for all possible t\vec{t}. By Lemma 6.5 and the fact that {v1,,vki1,wm}\{v_{1},\ldots,v_{k-i-1},w_{m}\} is present in all partial forests we take for any m[N]m\in[N], we know that (Zv1,,Zvki1,Zwm)(Z_{v_{1}},\ldots,Z_{v_{k-i-1}},Z_{w_{m}}) has the same distribution as (Xi+1,,Xk)(X_{i+1},\ldots,X_{k}) for each m[N]m\in[N]. Hence

((Zv)vV)\displaystyle\mathbb{H}\left((Z_{v})_{v\in V}\right)\leq (Zv1,,Zvki1)+m=1N(ZwmZv1,,Zvki1)\displaystyle\mathbb{H}(Z_{v_{1}},\ldots,Z_{v_{k-i-1}})+\sum_{m=1}^{N}\mathbb{H}(Z_{w_{m}}\mid Z_{v_{1}},\ldots,Z_{v_{k-i-1}})
=\displaystyle= (Xi+2,,Xk)+N(Xi+1Xi+2,,Xk)\displaystyle\mathbb{H}(X_{i+2},\ldots,X_{k})+N\mathbb{H}(X_{i+1}\mid X_{i+2},\ldots,X_{k})
=\displaystyle= (N+ki1)(X1)+log2(xi+2xkxi+1N).\displaystyle(N+k-i-1)\mathbb{H}(X_{1})+\log_{2}(x_{i+2}\cdots x_{k}\cdot x_{i+1}^{N}). (7.2)

Thus Lemmas 3.9, 7.1 and 7.2 now gives

2(N+ki1)(X1)1=t0<t1<<ti+1=N+1xi+2xkji+1xjtjtj1xi+2xkxi+1N2(N+ki1)(X1),2^{(N+k-i-1)\mathbb{H}(X_{1})}\sum_{1=t_{0}<t_{1}<\cdots<t_{i+1}=N+1}x_{i+2}\cdots x_{k}\cdot\prod_{j\leq i+1}x_{j}^{t_{j}-t_{j-1}}\leq x_{i+2}\cdots x_{k}\cdot x_{i+1}^{N}\cdot 2^{(N+k-i-1)\mathbb{H}(X_{1})},

and so

1=t0<t1<<ti+1=N+1ji+1(xjxi+1)tjtj11.\sum_{1=t_{0}<t_{1}<\cdots<t_{i+1}=N+1}\prod_{j\leq i+1}\left(\frac{x_{j}}{x_{i+1}}\right)^{t_{j}-t_{j-1}}\leq 1.

Note that we may replace ji+1j\leq i+1 by j<i+1j<i+1 in the product. This way, when we take NN to approach infinity, we must have xj<xi+1x_{j}<x_{i+1} for each j[i]j\in[i] in order for the left hand side to converge. Moreover, the left hand side becomes

δ1,,δiji(xjxi+1)δj=jixjxi+1xj,\sum_{\delta_{1},\ldots,\delta_{i}\in\mathbb{N}}\prod_{j\leq i}\left(\frac{x_{j}}{x_{i+1}}\right)^{\delta_{j}}=\prod_{j\leq i}\frac{x_{j}}{x_{i+1}-x_{j}},

as desired. ∎

Once again, to prove Theorem 7.1, we need to upper bound x1xk1x_{1}\cdots x_{k-1} 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 kk be a positive integer. Fix real numbers 0<z1<<zk0<z_{1}<\cdots<z_{k}. Let 0<y1<<yk0<y_{1}<\ldots<y_{k} be real numbers with

jiyjyi+1yjjizjzi+1zj\prod_{j\leq i}\frac{y_{j}}{y_{i+1}-y_{j}}\leq\prod_{j\leq i}\frac{z_{j}}{z_{i+1}-z_{j}}

for any i=1,,k1i=1,\ldots,k-1. Then

y1yk1z1zk1zkk1ykk1.y_{1}\cdots y_{k-1}\leq\frac{z_{1}\cdots z_{k-1}}{z_{k}^{k-1}}y_{k}^{k-1}.
Proof.

We will prove by induction on kk. When k=1k=1 this is clearly true. Now suppose that k2k\geq 2 and the statement is true for all smaller kk. Then we have

y1yiz1ziyi+1izi+1i\frac{y_{1}\cdots y_{i}}{z_{1}\cdots z_{i}}\leq\frac{y_{i+1}^{i}}{z_{i+1}^{i}}

for all i<k1i<k-1 by the inductive hypothesis. Now let

αi=1ijizjzkzj\alpha_{i}=\frac{1}{i}\sum_{j\leq i}\frac{z_{j}}{z_{k}-z_{j}}

for any ik1i\leq k-1. Note that for any i<k1i<k-1, we have

(y1yi+1z1zi+1)αi+1(y1yiz1zi)αi(yi+1izi+1i)(αi+1αi)(yi+1zi+1)αi+1=(y1yiz1zi)αi(yi+1zi+1)zi+1zkzi+1.\left(\frac{y_{1}\cdots y_{i+1}}{z_{1}\cdots z_{i+1}}\right)^{\alpha_{i+1}}\leq\left(\frac{y_{1}\cdots y_{i}}{z_{1}\cdots z_{i}}\right)^{\alpha_{i}}\left(\frac{y_{i+1}^{i}}{z_{i+1}^{i}}\right)^{(\alpha_{i+1}-\alpha_{i})}\left(\frac{y_{i+1}}{z_{i+1}}\right)^{\alpha_{i+1}}=\left(\frac{y_{1}\cdots y_{i}}{z_{1}\cdots z_{i}}\right)^{\alpha_{i}}\left(\frac{y_{i+1}}{z_{i+1}}\right)^{\frac{z_{i+1}}{z_{k}-z_{i+1}}}.

Here, we are using that αi+1αi0\alpha_{i+1}-\alpha_{i}\geq 0 as z1zkz1<<zi+1zkzi+1\frac{z_{1}}{z_{k}-z_{1}}<\cdots<\frac{z_{i+1}}{z_{k}-z_{i+1}}. Multiplying these up for i=1,,k2i=1,\ldots,k-2, and we get

(y1yk1z1zk1)αk1(y1z1)z1zkz1(yk1zk1)zk1zkzk1.\left(\frac{y_{1}\cdots y_{k-1}}{z_{1}\cdots z_{k-1}}\right)^{\alpha_{k-1}}\leq\left(\frac{y_{1}}{z_{1}}\right)^{\frac{z_{1}}{z_{k}-z_{1}}}\cdots\left(\frac{y_{k-1}}{z_{k-1}}\right)^{\frac{z_{k-1}}{z_{k}-z_{k-1}}}.

Thus

(y1yk1z1zk1)αk1+1\displaystyle\left(\frac{y_{1}\cdots y_{k-1}}{z_{1}\cdots z_{k-1}}\right)^{\alpha_{k-1}+1}\leq i=1k1(yizi)zizkzi(ykyizkzi)\displaystyle\prod_{i=1}^{k-1}\left(\frac{y_{i}}{z_{i}}\right)^{\frac{z_{i}}{z_{k}-z_{i}}}\left(\frac{y_{k}-y_{i}}{z_{k}-z_{i}}\right)
=\displaystyle= i=1k1[(yizi)zizk(ykyizkzi)zkzizk]zkzkzi\displaystyle\prod_{i=1}^{k-1}\left[\left(\frac{y_{i}}{z_{i}}\right)^{\frac{z_{i}}{z_{k}}}\left(\frac{y_{k}-y_{i}}{z_{k}-z_{i}}\right)^{\frac{z_{k}-z_{i}}{z_{k}}}\right]^{\frac{z_{k}}{z_{k}-z_{i}}}
\displaystyle\leq i=1k1(zizkyizi+zkzizkykyizkzi)zkzkzi\displaystyle\prod_{i=1}^{k-1}\left(\frac{z_{i}}{z_{k}}\cdot\frac{y_{i}}{z_{i}}+\frac{z_{k}-z_{i}}{z_{k}}\cdot\frac{y_{k}-y_{i}}{z_{k}-z_{i}}\right)^{\frac{z_{k}}{z_{k}-z_{i}}} (weighted AM-GM)
=\displaystyle= i=1k1(ykzk)zkzkzi=(ykzk)(k1)(αk1+1),\displaystyle\prod_{i=1}^{k-1}\left(\frac{y_{k}}{z_{k}}\right)^{\frac{z_{k}}{z_{k}-z_{i}}}=\left(\frac{y_{k}}{z_{k}}\right)^{(k-1)(\alpha_{k-1}+1)},

completing the inductive step.

Alternative Proof of Theorem 7.1.

Suppose GG is k\mathcal{F}_{k}-hom-free. Set zi=iz_{i}=i for each i[k]i\in[k]. By Lemma 7.4, we know that

jixjxi+1xj1=jij(i+1)j=jizjzi+1zj.\prod_{j\leq i}\frac{x_{j}}{x_{i+1}-x_{j}}\leq 1=\prod_{j\leq i}\frac{j}{(i+1)-j}=\prod_{j\leq i}\frac{z_{j}}{z_{i+1}-z_{j}}.

Therefore by Lemma 7.5 and the fact that xk=1x_{k}=1,

x1xk1(k1)!kk1=k!kk,x_{1}\cdots x_{k-1}\leq\frac{(k-1)!}{k^{k-1}}=\frac{k!}{k^{k}},

as desired. ∎

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 λ\lambda be a partition of kk with λ1k/2\lambda_{1}\leq\lceil k/2\rceil and λi=1\lambda_{i}=1 for all 1<i(λ)1<i\leq\ell(\lambda). Again, it is clear that π(Δλ)k!/kk\pi(\Delta_{\lambda})\geq k!/k^{k}, so it suffices to show that π(Δλ)k!/kk\pi(\Delta_{\lambda})\leq k!/k^{k}. To show this, it suffices to show that any Δλ\Delta_{\lambda}-hom-free kk-graph GG is also Δλ\Delta_{\lambda^{\prime}}-hom-free for any λ\lambda^{\prime} with |λ|=k\left\lvert\lambda^{\prime}\right\rvert=k and (λ)=2\ell(\lambda^{\prime})=2. This will follow immediately if we show that Δλ\Delta_{\lambda} admits a homomorphism to Δλ\Delta_{\lambda^{\prime}} for any such λ\lambda^{\prime}. By Proposition 6.2, it is sufficient to show that Δλ\Delta_{\lambda^{\prime}} admits a homomorphism from Δλp\Delta^{p}_{\lambda} for any λ\lambda^{\prime} with |λ|=k\left\lvert\lambda^{\prime}\right\rvert=k and (λ)=2\ell(\lambda^{\prime})=2. This is now simple: suppose that Δλ\Delta_{\lambda^{\prime}} has base ee^{\prime} and apex vv^{\prime}, and e1,e2e_{1}^{\prime},e_{2}^{\prime} are two edges such that |eie|=λi\left\lvert e_{i}^{\prime}\cap e^{\prime}\right\rvert=\lambda_{i}^{\prime} for i[2]i\in[2]. We also suppose that Δλp\Delta^{p}_{\lambda} has base ee and apex vv, and e1,,ee_{1},\ldots,e_{\ell} are partial edges such that |eie|=λi\left\lvert e_{i}\cap e\right\rvert=\lambda_{i} for i[]i\in[\ell]. As λ1k/2λ1\lambda_{1}^{\prime}\geq\lceil k/2\rceil\geq\lambda_{1}, we can take f:e{v}V(Δλ)f:e\cup\{v\}\to V(\Delta_{\lambda^{\prime}}) so that f(v)=vf(v)=v^{\prime}, f(e)=ef(e)=e^{\prime} and f(ee1)ee1f(e\cap e_{1})\subseteq e^{\prime}\cap e_{1}^{\prime}. This is a homomorphism from Δλp\Delta^{p}_{\lambda} to Δλ\Delta_{\lambda^{\prime}} as any vertex in ee^{\prime} shares an edge with vv^{\prime} in Δλ\Delta_{\lambda^{\prime}}. ∎

Finally, we give a proof of Theorem 1.6 by demonstrating a kk-graph GG that has b(G)>k!/kkb(G)>k!/k^{k} and is Δλ\Delta_{\lambda}-free for large λ1\lambda_{1}. Similar to an earlier lower-bound construction by Frankl and Füredi [23] for Δ(k1,1)\Delta_{(k-1,1)}, we will do so by constructing a kk-graph GG so that the intersection of any two edges is small.

Proof.

Let α<1\alpha<1 be some constant that is close to 11. In particular, assume that α>1/2\alpha>1/2. Let GauxG_{\textup{aux}} be an auxiliary graph with vertices ([2k]k)\binom{[2k]}{k}, and two vertices are connected if the corresponding subsets have intersection at least αk\alpha k. Then GauxG_{\textup{aux}} is a regular graph with degree

i(1α)k(ki)2<k(k(1α)k)2=2(2h(α)+o(1))k,\sum_{i\leq(1-\alpha)k}\binom{k}{i}^{2}<k\binom{k}{\lfloor(1-\alpha)k\rfloor}^{2}=2^{(2h(\alpha)+o(1))k},

where h(α)=αlog2α(1α)log2αh(\alpha)=-\alpha\log_{2}\alpha-(1-\alpha)\log_{2}\alpha and we use that

(k(1α+o(1))k)=2(h(α)+o(1))k\binom{k}{(1-\alpha+o(1))k}=2^{(h(\alpha)+o(1))k}

when α>1/2\alpha>1/2.

By the Caro–Wei theorem, there exists an independent set of size

(2kk)2(2h(α)+o(1))k=2(22h(α)+o(1))k.\frac{\binom{2k}{k}}{2^{(2h(\alpha)+o(1))k}}=2^{(2-2h(\alpha)+o(1))k}.

This corresponds to a kk-graph GG on [2k][2k] with 2(22h(α)+o(1))k2^{(2-2h(\alpha)+o(1))k} edges so that any two edges have intersection less than αk\alpha k.

Now if GG contains a homomorphic image of Δλ\Delta_{\lambda} where λ1>αk\lambda_{1}>\alpha k, let ee be its base and let e1e_{1} be the edge with |ee1|=λ1\left\lvert e\cap e_{1}\right\rvert=\lambda_{1}. Also let ff be a homomorphism from Δλ\Delta_{\lambda} to GG. Then |f(e)f(e1)|>αk\left\lvert f(e)\cap f(e_{1})\right\rvert>\alpha k, and so f(e)=f(e1)f(e)=f(e_{1}). This shows if vv is the apex of Δλ\Delta_{\lambda}, then f(v)=f(u)f(v)=f(u) for some ueu\in e. However, {uv}\{uv\} is contained in some edge in Δλ\Delta_{\lambda}, which is a contradiction. Thus π(Δλ)\pi(\Delta_{\lambda}) is at least b(G)b(G), which is at least the density of GG. The density of GG is

k!2(22h(α)+o(1))k(2k)k=2(12h(α)+o(1))kk!kk,\frac{k!\cdot 2^{(2-2h(\alpha)+o(1))k}}{(2k)^{k}}=2^{(1-2h(\alpha)+o(1))k}\cdot\frac{k!}{k^{k}},

which is strictly greater than k!/kkk!/k^{k} for sufficiently large kk as long as h(α)<1/2h(\alpha)<1/2. As hh is continuous on [1/2,1][1/2,1] and h(1)=0h(1)=0, this is true for α\alpha sufficiently close to 11. ∎

The proof roughly gives α0.89\alpha\approx 0.89. Although our proof is not fully optimized, we believe that it would not give the correct upper bound for α\alpha 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 π(Ek+1(k))=k!/kk\pi(E^{(k)}_{k+1})=k!/k^{k} where Ek+1(k)E^{(k)}_{k+1} is the extended clique of size k+1k+1, and Mubayi and Pikhurko [42] strengthened it to π(Δ(1,1,,1))=k!/kk\pi(\Delta_{(1,1,\ldots,1)})=k!/k^{k}. In fact they both proved more general results than this: Mubayi showed that for each rkr\geq k,

π(Er+1(k))=b(Kr(k))=i=1k1(1ir)\pi(E^{(k)}_{r+1})=b(K^{(k)}_{r})=\prod_{i=1}^{k-1}\left(1-\frac{i}{r}\right)

and Mubayi and Pikhurko strengthened it as follows: consider the partial kk-graph FF on r+1r+1 vertices generated by [k][k] and all the 22-subsets of [r+1][r+1], and then take its extension F~\tilde{F}. Then π(F~)=b(Kr(k))\pi(\tilde{F})=b(K^{(k)}_{r}) as well. Note that Er+1(k)E^{(k)}_{r+1} is the extension of Kr+1K_{r+1} as a partial kk-graph, and there is a homomorphism from Kr+1K_{r+1} to F~\tilde{F}. Therefore π(Er+1(k))π(F~)\pi(E^{(k)}_{r+1})\leq\pi(\tilde{F}), and so π(F~)=b(Kr(k))\pi(\tilde{F})=b(K^{(k)}_{r}) 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 π(F~)=b(Kr(k))\pi(\tilde{F})=b(K^{(k)}_{r}) 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 k,rk,r be positive integers with rkr\geq k. Let \mathcal{F} be a family of kk-graphs such that the following holds. For any i=1,,k1i=1,\ldots,k-1, if we take the union of any (rk+ii)+1\binom{r-k+i}{i}+1 different partial forests F(t)F^{(\vec{t})} as in the proof of Lemma 7.4, then its extension is not \mathcal{F}-hom-free.

Then π()b(Kr(k))\pi(\mathcal{F})\leq b(K^{(k)}_{r}).

Proof.

Suppose that GG is \mathcal{F}-hom-free. Let (X1,,Xk)(X_{1},\ldots,X_{k}) be any random edge with uniform ordering on GG and let x1,,xkx_{1},\dots,x_{k} be its ratio sequence. We first fix some i[k1]i\in[k-1]. Temporarily fix some large positive integer NN as in the proof of Lemma 7.4. For any 1=t0<t1<<tiN1=t_{0}<t_{1}<\cdots<t_{i}\leq N, let (Yv(t))vV(Y_{v}^{(\vec{t})})_{v\in V} be the random homomorphism from F(t)F^{(\vec{t})} to GG sampled via Lemma 6.5 as in the proof of Lemma 7.4. Then by the assumption on \mathcal{F} and that GG is \mathcal{F}-hom-free, we know that the supports of the random homomorphisms (Yv(t))vV(Y_{v}^{(\vec{t})})_{v\in V} are ((rk+ii)+1)\left(\binom{r-k+i}{i}+1\right)-wise disjoint. Therefore, if (Zv)vV(Z_{v})_{v\in V} is the mixture of the (Yv(t))vV(Y_{v}^{(\vec{t})})_{v\in V}’s provided by Lemma 3.9, we have

1=t0<t1<<tiN2((Yv(t))vV)(rk+ii)2((Zv)vV).\sum_{1=t_{0}<t_{1}<\cdots<t_{i}\leq N}2^{\mathbb{H}\left((Y_{v}^{(\vec{t})})_{v\in V}\right)}\leq\binom{r-k+i}{i}2^{\mathbb{H}\left((Z_{v})_{v\in V}\right)}.

Using what we have computed in the proof of Lemma 7.4, when NN is taken to infinity, we get

jixjxi+1xj(rk+ii).\prod_{j\leq i}\frac{x_{j}}{x_{i+1}-x_{j}}\leq\binom{r-k+i}{i}.

Now let zi=rk+iz_{i}=r-k+i for each i=1,,ki=1,\ldots,k. Then it is easy to verify that

(rk+ii)=jizjzi+1zj\binom{r-k+i}{i}=\prod_{j\leq i}\frac{z_{j}}{z_{i+1}-z_{j}}

for each i[k1]i\in[k-1]. Therefore, by Lemma 7.5, we get that

x1xk1z1zk1zkk1=(rk+1)(r1)rk1=i=1k1(1ir)=b(Kr(k)).x_{1}\cdots x_{k-1}\leq\frac{z_{1}\cdots z_{k-1}}{z_{k}^{k-1}}=\frac{(r-k+1)\cdots(r-1)}{r^{k-1}}=\prod_{i=1}^{k-1}\left(1-\frac{i}{r}\right)=b(K_{r}^{(k)}).

This shows that b(G)=bentropy(G)b(Kr(k))b(G)=b_{\textup{entropy}}(G)\leq b(K_{r}^{(k)}) for any \mathcal{F}-hom-free kk-graph GG, and so we have π()b(Kr(k))\pi(\mathcal{F})\leq b(K_{r}^{(k)}). ∎

Corollary 8.2.

Let FF be the partial kk-graph on r+1r+1 vertices generated by [k][k] and all the 22-subsets of [r+1][r+1]. Let F~\tilde{F} be its extension. Then π(F~)=b(Kr(k))\pi(\tilde{F})=b(K_{r}^{(k)}).

Proof.

First of all, it is clear that Kr(k)K_{r}^{(k)} is FF-hom-free. Therefore, by Proposition 6.2, Kr(k)K_{r}^{(k)} is also F~\tilde{F}-hom-free, and so π(F~)b(Kr(k))\pi(\tilde{F})\geq b(K_{r}^{(k)}).

To show that π(F~)b(Kr(k))\pi(\tilde{F})\leq b(K_{r}^{(k)}), it now suffices to show that the assumption of Theorem 8.1 holds for any i[k1]i\in[k-1]. Indeed, for any collection TT of (rk+ii)+1\binom{r-k+i}{i}+1 different possible t\vec{t}’s, we may construct SS\subseteq\mathbb{N} with size rk+i+1r-k+i+1 that satisfies the following: for each sSs\in S there exists tT\vec{t}\in T such that s{t1,,ti}s\in\{t_{1},\ldots,t_{i}\}, and there exists a tT\vec{t}\in T with {t1,,ti}S\{t_{1},\ldots,t_{i}\}\subseteq S. Indeed, set S=tT{t1,,ti}S^{\prime}=\bigcup_{\vec{t}\in T}\{t_{1},\ldots,t_{i}\}. Then |T|(|S|i)\left\lvert T\right\rvert\leq\binom{\left\lvert S^{\prime}\right\rvert}{i}, which shows that |S|rk+i+1\left\lvert S^{\prime}\right\rvert\geq r-k+i+1. Now simply take SSS\subseteq S^{\prime} of size rk+i+1r-k+i+1 while containing some {t1,,ti}\{t_{1},\ldots,t_{i}\} for some tT\vec{t}\in T. Label this t\vec{t} as t\vec{t^{*}}.

Now we need to show that there is a homomorphic image of F~\tilde{F} in the extension of tTF(t)\bigcup_{\vec{t}\in T}F^{(\vec{t})}. By Proposition 6.2, it suffices to construct a homomorphism from FF to tTF(t)\bigcup_{\vec{t}\in T}F^{(\vec{t})}. To do so, we will simply map 1,,ki11,\ldots,k-i-1 to v1,,vki1v_{1},\ldots,v_{k-i-1}, map ki,,kk-i,\ldots,k to wt0,,wtiw_{t^{*}_{0}},\ldots,w_{t^{*}_{i}}, and then map the rest of the vertices into S\{t1,,ti}S\backslash\{t^{*}_{1},\ldots,t^{*}_{i}\} bijectively. To show that this is indeed a homomorphism, notice first that {v1,,vki1,wt0,,wti}\{v_{1},\ldots,v_{k-i-1},w_{t_{0}^{*}},\ldots,w_{t_{i}^{*}}\} is a partial edge in F(t)F^{(\vec{t^{*}})}. Therefore it remains to check that {ws1,ws2}\{w_{s_{1}},w_{s_{2}}\} and {vm,ws1}\{v_{m},w_{s_{1}}\} are both in tTF(t)\bigcup_{\vec{t}\in T}F^{(\vec{t})} for any s1s2Ss_{1}\neq s_{2}\in S and m[ki1]m\in[k-i-1]. Indeed, if s1<s2s_{1}<s_{2} and s2=tjs_{2}=t_{j} for some tT\vec{t}\in T, then {vm,ws1,ws2}\{v_{m},w_{s_{1}},w_{s_{2}}\} is indeed a partial edge in F(t)F^{(\vec{t})}, which shows that both {ws1,ws2}\{w_{s_{1}},w_{s_{2}}\} and {vm,ws1}\{v_{m},w_{s_{1}}\} are partial edges in F(t)F^{(\vec{t})} 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 kk-graphs on r+1r+1 vertices have Turán density equal to b(Kr(k))b(K_{r}^{(k)}), as long as rr is at least some threshold MkM_{k} that depends on kk. One special case related to our result is the kk-graph Fr+1(k,k1)F_{r+1}^{(k,k-1)} that can be obtained as follows: consider the partial kk-graph on [r+1][r+1] spanned by the edges {[k1]i:i=k,,r+1}\{[k-1]\cup i:i=k,\ldots,r+1\} and all the 22-subsets of [r+1][r+1], and then take the extension of the partial kk-graph. For example, Fk+1(k,k1)F_{k+1}^{(k,k-1)} is the tent Δ(k1,1).\Delta_{(k-1,1)}. Sidorenko’s result is more general and relies on trees TT that satisfy the Erdős–Sós conjecture ex(T,n)12(v(T)2)n\textup{ex}(T,n)\leq\frac{1}{2}(v(T)-2)n, 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 π(Fr+1(k,k1))=b(Kr(k))\pi(F^{(k,k-1)}_{r+1})=b(K^{(k)}_{r}) for sufficiently large rr with respect to kk. Our argument actually gives a more general statement: for any s<krs<k\leq r, let Fr+1(k,s)F^{(k,s)}_{r+1} be the extension of the partial kk-graph spanned by {[s]i:i=s+1,,r+1}\{[s]\cup i:i=s+1,\ldots,r+1\} and all the 22-subsets of [r+1][r+1]. Then we obtain a sufficient condition for π(Fr+1(k,s))=b(Kr(k))\pi(F^{(k,s)}_{r+1})=b(K_{r}^{(k)}).

Theorem 8.3.

Let k,r,sk,r,s be positive integers with krk\leq r and

ksi=1s1iri.k-s\geq\sum_{i=1}^{s-1}\frac{i}{r-i}. (8.1)

Then π(Fr+1(k,s))=b(Kr(k))\pi(F^{(k,s)}_{r+1})=b(K_{r}^{(k)}).

Proof.

It is clear that Kr(k)K_{r}^{(k)} is Fr+1(k,s)F^{(k,s)}_{r+1}-hom-free. Therefore, π(Fr+1(k,s))b(Kr(k))\pi(F^{(k,s)}_{r+1})\geq b(K_{r}^{(k)}).

To prove the other direction π(Fr+1(k,s))b(Kr(k))\pi(F^{(k,s)}_{r+1})\leq b(K_{r}^{(k)}), we may fix a Fr+1(k,s)F^{(k,s)}_{r+1}-hom-free kk-graph GG and a random with uniform ordering (X1,,Xk)(X_{1},\dots,X_{k}) on GG. Let x1,,xkx_{1},\dots,x_{k} be the ratio sequence of (X1,,Xk)(X_{1},\dots,X_{k}). We will solve for the maximum of x1xk1x_{1}\dots x_{k-1} under the constraints given by the following lemma.

Lemma 8.4.

For any integers i,ji,j with i[ks],ij<ki\in[k-s],i\leq j<k, we have

xirk+ixj+1xj.\frac{x_{i}}{r-k+i}\leq x_{j+1}-x_{j}.
Proof.

We will fix i,ji,j throughout this proof. As in what we did in Section 4, we will temporarily fix an integer NN\in\mathbb{N} that will later be taken to infinity. For any t[N]t\in[N], we will define a partial forest F(t)F^{(t)} on V={v1,,vki,w1,,wN}V=\{v_{1},\dots,v_{k-i},w_{1},\dots,w_{N}\}. The partial forest F(t)F^{(t)} is spanned by the partial edges {v1,,vki,wt}\{v_{1},\dots,v_{k-i},w_{t}\}, {v1,,vkj1,wm,wt}\{v_{1},\dots,v_{k-j-1},w_{m},w_{t}\} for every m<tm<t, and {v1,,vkj1,wm}\{v_{1},\dots,v_{k-j-1},w_{m}\} for every m>tm>t. With the linear order << given by v1<<vki<wN<<w1v_{1}<\dots<v_{k-i}<w_{N}<\cdots<w_{1}, we know that F(t)F^{(t)} is indeed a partial forest. We can compute the forest sequence with respect to the linear order as follows: each vmv_{m} contributes one to nmn_{m} for each mkim\leq k-i. For the contribution of wmw_{m}, if m>tm>t it contributes one to nkjn_{k-j}; if m=tm=t it contributes one to nki+1n_{k-i+1}; otherwise it contributes one to nkj+1n_{k-j+1}. Therefore the forest sequence (n1,,nk)(n_{1},\ldots,n_{k}) is e1++eki+(Nt)ekj+eki+1+(t1)ekj+1\vec{e}_{1}+\dots+\vec{e}_{k-i}+(N-t)\vec{e}_{k-j}+\vec{e}_{k-i+1}+(t-1)\vec{e}_{k-j+1}, where e1,,ek\vec{e}_{1},\dots,\vec{e}_{k} are the vectors in the standard basis. Now let (Yv(t))vV(Y^{(t)}_{v})_{v\in V} be the random homomorphism produced by Lemma 6.5. This gives

((Yv(t))vV)=(N+ki)(X1)+log2(xixkxjt1xj+1Nt).\displaystyle\mathbb{H}\left((Y^{(t)}_{v})_{v\in V}\right)=(N+k-i)\mathbb{H}(X_{1})+\log_{2}\left(x_{i}\cdots x_{k}\cdot x_{j}^{t-1}x_{j+1}^{N-t}\right). (8.2)

Now, we show that the random tuples (Yv(1))vV,,(Yv(N))vV(Y^{(1)}_{v})_{v\in V},\dots,(Y^{(N)}_{v})_{v\in V} have (rk+i+1)(r-k+i+1)-wise disjoint supports. Note that, for any t1<<trk+i+1t_{1}<\dots<t_{r-k+i+1}, the extension of the union =1rk+i+1F(t)\cup_{\ell=1}^{r-k+i+1}F^{(t_{\ell})} contains a homomorphic image of Fr+1(k,ki)F^{(k,k-i)}_{r+1}, given by the partial edges {v1,,vki,wt}\{v_{1},\dots,v_{k-i},w_{t_{\ell}}\} for [rk+i+1]\ell\in[r-k+i+1] and {wt,wt}\{w_{t_{\ell^{\prime}}},w_{t_{\ell}}\} for 1<rk+i+11\leq\ell^{\prime}<\ell\leq r-k+i+1. Since kisk-i\leq s, this is also a homomorphic image of Fr+1(k,s)F^{(k,s)}_{r+1}. Thus, no sequence of vertices is in =1rk+i+1supp((Yv(t))vV)\cap_{\ell=1}^{r-k+i+1}\operatorname{supp}((Y^{(t_{\ell})}_{v})_{v\in V}).

Therefore we may now apply Lemma 3.9 with a=rk+ia=r-k+i. Suppose that (Zv)vV(Z_{v})_{v\in V} is the resulting mixture of (Yv(t))vV(Y_{v}^{(t)})_{v\in V} for all t[N]t\in[N]. Note that the partial edge {v1,,vki}\{v_{1},\dots,v_{k-i}\} is present in all partial forests, so by Lemma 6.5 we know that (Zv1,,Zvki)(Z_{v_{1}},\ldots,Z_{v_{k-i}}) has the same distribution as (Xi+1,,Xk)(X_{i+1},\ldots,X_{k}). Similarly, for each m[N]m\in[N], since the partial edge {v1,,vkj1,wm}\{v_{1},\dots,v_{k-j-1},w_{m}\} is present in all partial forests, we know that (Zv1,,Zvkj1,Zwm)(Z_{v_{1}},\ldots,Z_{v_{k-j-1}},Z_{w_{m}}) has the same distribution as (Xj+1,,Xk)(X_{j+1},\ldots,X_{k}). Hence

((Zv)vV)\displaystyle\mathbb{H}\left((Z_{v})_{v\in V}\right)\leq (Zv1,,Zvki)+m=1N(ZwmZv1,,Zvkj1)\displaystyle\mathbb{H}(Z_{v_{1}},\ldots,Z_{v_{k-i}})+\sum_{m=1}^{N}\mathbb{H}(Z_{w_{m}}\mid Z_{v_{1}},\ldots,Z_{v_{k-j-1}})
=\displaystyle= (Xi+1,,Xk)+N(Xj+1Xj+2,,Xk)\displaystyle\mathbb{H}(X_{i+1},\ldots,X_{k})+N\mathbb{H}(X_{j+1}\mid X_{j+2},\ldots,X_{k})
=\displaystyle= (N+ki)(X1)+log2(xi+1xkxj+1N).\displaystyle(N+k-i)\mathbb{H}(X_{1})+\log_{2}(x_{i+1}\cdots x_{k}\cdot x_{j+1}^{N}). (8.3)

Thus Lemmas 3.9, 8.2 and 8.3 now give

t=1Nxixkxjt1xj+1Nt2(N+ki)(X1)(rk+i)xi+1xkxj+1N2(N+ki)(X1),\sum_{t=1}^{N}x_{i}\cdots x_{k}\cdot x_{j}^{t-1}x_{j+1}^{N-t}\cdot 2^{(N+k-i)\mathbb{H}(X_{1})}\leq(r-k+i)x_{i+1}\cdots x_{k}\cdot x_{j+1}^{N}\cdot 2^{(N+k-i)\mathbb{H}(X_{1})},

and so

t=1Nxixjt1xj+1trk+i.\sum_{t=1}^{N}x_{i}x_{j}^{t-1}x_{j+1}^{-t}\leq r-k+i.

By rearranging and taking NN goes to infinity, we obtain

xixj+111xjxj+1=t=1xixj+1(xjxj+1)t1rk+i,\frac{x_{i}}{x_{j+1}}\cdot\frac{1}{1-\frac{x_{j}}{x_{j+1}}}=\sum_{t=1}^{\infty}\frac{x_{i}}{x_{j+1}}\left(\frac{x_{j}}{x_{j+1}}\right)^{t-1}\leq r-k+i,

and the lemma follows. ∎

Once again, to prove Theorem 8.3, we need to upper bound x1xk1x_{1}\cdots x_{k-1} given the inequalities in Lemma 8.4. We start with the following inequality similar to Lemma 7.3.

Lemma 8.5.

Suppose that y1,,yty_{1},\ldots,y_{t} and zz are some non-negative real numbers. Then

y1yt(i=1tyiz+i)t(z+1)(z+t)tt.y_{1}\cdots y_{t}\leq\left(\sum_{i=1}^{t}\frac{y_{i}}{z+i}\right)^{t}\frac{(z+1)\cdots(z+t)}{t^{t}}.
Proof.

We will prove this by inducting on tt. For t=1t=1, the inequality is trivial.

Assume the statement is true for t1t-1. From the inductive hypothesis and AM-GM inequality, we have

y1yt\displaystyle y_{1}\cdots y_{t}\leq yt(i=1t1yiz+i)t1(z+1)(z+t1)(t1)t1\displaystyle y_{t}\left(\sum_{i=1}^{t-1}\frac{y_{i}}{z+i}\right)^{t-1}\frac{(z+1)\cdots(z+t-1)}{(t-1)^{t-1}}
=\displaystyle= (t1z+tyt)(i=1t1yiz+i)t1(z+1)(z+t)(t1)t\displaystyle\left(\frac{t-1}{z+t}y_{t}\right)\left(\sum_{i=1}^{t-1}\frac{y_{i}}{z+i}\right)^{t-1}\frac{(z+1)\cdots(z+t)}{(t-1)^{t}}
\displaystyle\leq (t1ti=1tyiz+i)t(z+1)(z+t)(t1)t\displaystyle\left(\frac{t-1}{t}\sum_{i=1}^{t}\frac{y_{i}}{z+i}\right)^{t}\frac{(z+1)\cdots(z+t)}{(t-1)^{t}}
=\displaystyle= (i=1tyiz+i)t(z+1)(z+t)tt.\displaystyle\left(\sum_{i=1}^{t}\frac{y_{i}}{z+i}\right)^{t}\frac{(z+1)\cdots(z+t)}{t^{t}}.\qed

Now, by using this lemma with t=k1,yi=xit=k-1,y_{i}=x_{i} and z=rkz=r-k, it is sufficient to upper bound right hand side using the conditions from Lemma 8.4.

Claim 8.6.

We have

x1rk+1++xk1r1k1rxk.\frac{x_{1}}{r-k+1}+\dots+\frac{x_{k-1}}{r-1}\leq\frac{k-1}{r}x_{k}.
Proof.

Let ss^{\prime} be the largest integer such that

ksi=1s1irik-s^{\prime}\geq\sum_{i=1}^{s^{\prime}-1}\frac{i}{r-i}

holds. In particular, we have ss<ks\leq s^{\prime}<k. Set cc to be the real number such that

k1r=(1c)1rs+1rs+1++1r1.\frac{k-1}{r}=(1-c)\frac{1}{r-s^{\prime}}+\frac{1}{r-s^{\prime}+1}+\dots+\frac{1}{r-1}.

From the definition of ss^{\prime}, we have

k1s1+s1rs+1++1r1=rrs+1++rr1k-1\geq s^{\prime}-1+\frac{s^{\prime}-1}{r-s^{\prime}+1}+\dots+\frac{1}{r-1}=\frac{r}{r-s^{\prime}+1}+\dots+\frac{r}{r-1}

and

k1<s+srs++1r1=rrs++rr1.k-1<s^{\prime}+\frac{s^{\prime}}{r-s^{\prime}}+\dots+\frac{1}{r-1}=\frac{r}{r-s^{\prime}}+\dots+\frac{r}{r-1}.

Therefore, c(0,1]c\in(0,1]. By replacing the coefficient of xkx_{k} using the definition of cc and rearranging, we may rewrite the inequality we want to show as the following.

x1rk+1++xks1rs1+cxksrs\displaystyle\frac{x_{1}}{r-k+1}+\dots+\frac{x_{k-s-1}}{r-s-1}+c\frac{x_{k-s}}{r-s}
\displaystyle\leq (1c)srsxkxkss+s1rs+1xkxks+1s1++1r1xkxk11.\displaystyle(1-c)\frac{s^{\prime}}{r-s^{\prime}}\frac{x_{k}-x_{k-s^{\prime}}}{s^{\prime}}+\frac{s^{\prime}-1}{r-s^{\prime}+1}\frac{x_{k}-x_{k-s^{\prime}+1}}{s-1}+\dots+\frac{1}{r-1}\frac{x_{k}-x_{k-1}}{1}. (8.4)

Note that Lemma 8.4 implies that

xirk+ixkxkjj\frac{x_{i}}{r-k+i}\leq\frac{x_{k}-x_{k-j}}{j}

holds for all iksji\leq k-s^{\prime}\leq j. Thus, to prove 8.4, it is sufficient to check

ks1+c(1c)srs+s1rs+1++1r1.k-s-1+c\leq(1-c)\frac{s^{\prime}}{r-s^{\prime}}+\frac{s^{\prime}-1}{r-s^{\prime}+1}+\dots+\frac{1}{r-1}.

Actually, the equality holds because, by the choice of cc, we have

ks1+c=\displaystyle k-s-1+c= (1c)rrs+rrs+1+rrs+2++rr1s+c\displaystyle(1-c)\frac{r}{r-s^{\prime}}+\frac{r}{r-s^{\prime}+1}+\frac{r}{r-s^{\prime}+2}+\dots+\frac{r}{r-1}-s+c
=\displaystyle= (1c)srs+s1rs+1+s2rs+2++1r1.\displaystyle(1-c)\frac{s^{\prime}}{r-s^{\prime}}+\frac{s^{\prime}-1}{r-s^{\prime}+1}+\frac{s^{\prime}-2}{r-s^{\prime}+2}+\dots+\frac{1}{r-1}.\qed

By combining Lemmas 8.5 and 8.6, we get

x1xk1(k1rxk)k1(rk+1)(r1)(k1)k1=(rk+1)(r1)rk1=b(Kr(k)).x_{1}\dots x_{k-1}\leq\left(\frac{k-1}{r}x_{k}\right)^{k-1}\frac{(r-k+1)\cdots(r-1)}{(k-1)^{k-1}}=\frac{(r-k+1)\cdots(r-1)}{r^{k-1}}=b(K_{r}^{(k)}).\qed

To give a sense of what the inequality in Theorem 8.3 means, with some standard computation, we can show the following. If r,kr,k are growing positive integers such that r=(C+ok(1))kr=(C+o_{k\to\infty}(1))k for some C1C\geq 1, then the largest positive integer ss satisfying 8.1 is (C(1exp(C1))+ok(1))k(C(1-\exp(-C^{-1}))+o_{k\to\infty}(1))k. In a different regime where s=kds=k-d for some positive integers dd, we can get that the smallest positive integer rr satisfying the inequality is ((2d)1+ok(1))k2((2d)^{-1}+o_{k\to\infty}(1))k^{2}. We include those computations in the appendix (Propositions A.2 and A.3).

We briefly remark that the threshold MkM_{k} Sidorenko deduced on rr is the same as ours when s=k1s=k-1. 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 ss.

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 Kr(k)K^{(k)}_{r} 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 x1,,xkx_{1},\ldots,x_{k} with equality holding when (X1,,Xk)(X_{1},\ldots,X_{k}) is a uniform oriented edge in Kr(k)K^{(k)}_{r}. 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 Kr(k)K^{(k)}_{r}.

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 33-graph (AB,E)(A\sqcup B,E) where E=(A2)×BA×(B2)E=\binom{A}{2}\times B\cup A\times\binom{B}{2}, and the complete oddly bipartite kk-graph (AB,E)(A\sqcup B,E) where kk is even, and EE is the kk-edges ee such that |Ae|\left\lvert A\cap e\right\rvert is odd. Although they are not formally blowups of some smaller hypergraphs, one can think of the complete bipartite 33-graphs as the blowups of ({1,2},{{1,1,2},{1,2,2}})(\{1,2\},\{\{1,1,2\},\{1,2,2\}\}), and the completely oddly bipartite kk-graphs are the blowups of some 22-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 33-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 kk-graph is the extremizer for expanded triangle. A very recent breakthrough of Sankar [55] showed that the complete oddly bipartite 44-graph is an asymptotic extremizer for tight cycles of sufficiently large length not divisible by 44.

We are unable to construct any partial forests that give tight inequalities when GG is the complete bipartite 33-graph. For GG being complete oddly kk-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 22-graphs to show that the Turán densities of expanded triangles are 1/21/2. 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 GG 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 K4(3)K_{4}^{(3)-}, the 33-graph on 44 vertices with 33 edges. Let G1G_{1} be a 33-graph on 66 vertices with 1010 edges so that any 22-subset is in exactly 22 edges—it turns out that G1G_{1} does exist and is unique up to isomorphism. The iterated blowup GmG_{m} of G1G_{1} is constructed inductively by replacing each vertex in G1G_{1} with Gm1G_{m-1}. Then GmG_{m} is K4(3)K_{4}^{(3)-}-free, and by taking mm to infinity, we get that π(K4(3))27\pi(K_{4}^{(3)-})\geq\frac{2}{7}. This is a construction of Frankl and Füredi [22], and the construction is conjectured to be optimal. The current best upper bound π(K4(3))0.2871\pi(K_{4}^{(3)-})\leq 0.2871 is obtained by Baber and Talbot [3] using flag algebra. Though we cannot say anything new about the Turán problem of K4(3)K_{4}^{(3)-} itself, our method seems to capture some structure in G1G_{1}. Indeed, by the partial forests F(i)=([4],{[3],[4]\{i}})F^{(i)}=([4],\{[3],[4]\backslash\{i\}\}) for i=1,2,3i=1,2,3, we can show that if GG is K4(3)K_{4}^{(3)-}-free and (X1,X2,X3)(X_{1},X_{2},X_{3}) is a random edge with uniform ordering on GG, then

x1=def2(X1X2,X3)(X1)13.x_{1}\stackrel{{\scriptstyle\mbox{\tiny def}}}{{=}}2^{\mathbb{H}(X_{1}\mid X_{2},X_{3})-\mathbb{H}(X_{1})}\leq\frac{1}{3}.

This is indeed achieved when (X1,X2,X3)(X_{1},X_{2},X_{3}) is a uniformly chosen oriented edge in G1G_{1}.

9.3. Entropic spectral radius

In Section 5, we showed that for any kk-graph GG, its spectral radius is related to the maximum of (X2,,XkX1)\mathbb{H}(X_{2},\ldots,X_{k}\mid X_{1}) for symmetric distribution (X1,,Xk)(X_{1},\ldots,X_{k}) on the oriented edges of GG. 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

ρ(G)(1v(G)vV(G)deg(v)kk1)k1k\rho(G)\geq\left(\frac{1}{v(G)}\sum_{v\in V(G)}\deg(v)^{\frac{k}{k-1}}\right)^{\frac{k-1}{k}}

for any kk-graph GG, where ρ(G)\rho(G) is the spectral radius of GG.

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 x1xk1x_{1}\cdots x_{k-1} 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 kk-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 (3,4)(3,4)-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 33-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 pp-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 (3, 4)(3,\,4)-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 kk-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 rr-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 r,sr,s and kk in Theorem 8.3

In this appendix, we will relate positive integers k,r,sk,r,s with krk\leq r satisfying the inequality

ksi=1s1iri.k-s\geq\sum_{i=1}^{s-1}\frac{i}{r-i}. (A.1)

We first compute the right hand side.

Lemma A.1.

Suppose that k,r,sk,r,s are positive integers satisfying A.1. Then r(ks)=Ω(s2)r(k-s)=\Omega(s^{2}), rs=Ω(r)r-s=\Omega(r) and

i=1s1iri=rlog(r1rs)(s1)+O(sr).\sum_{i=1}^{s-1}\frac{i}{r-i}=r\log\left(\frac{r-1}{r-s}\right)-(s-1)+O\left(\frac{s}{r}\right).
Proof.

We first show r(ks)=Ω(s2)r(k-s)=\Omega(s^{2}). This is clear as

ksi=1s1iris12s12rs12=Ω(s2r1).k-s\geq\sum_{i=1}^{s-1}\frac{i}{r-i}\geq\left\lfloor\frac{s-1}{2}\right\rfloor\frac{\left\lceil\frac{s-1}{2}\right\rceil}{r-\left\lceil\frac{s-1}{2}\right\rceil}=\Omega(s^{2}r^{-1}).

Now we show that rs=Ω(r)r-s=\Omega(r). This is clear when r2kr\geq 2k, so it suffices to check the case when r<2kr<2k. In this case, we have 2k(ks)>r(ks)Ω(s2)2k(k-s)>r(k-s)\geq\Omega(s^{2}). This forces scks\leq ck for some constant c<1c<1, and so rs=Ω(r)r-s=\Omega(r) as s<krs<k\leq r.

Now let \mathcal{E} be the error term defined by

=i=1s1iri1sxrx dx=1s(f(x)f(x)) dx\mathcal{E}=\sum_{i=1}^{s-1}\frac{i}{r-i}-\int_{1}^{s}\frac{x}{r-x}\textup{ d}x=\int_{1}^{s}\left(f(\lfloor x\rfloor)-f(x)\right)\textup{ d}x

where we set f(x)=x/(rx)f(x)=x/(r-x). Note that f(x)=r(rx)2f^{\prime}(x)=r(r-x)^{-2} is positive and increasing in xx when x[1,s][1,r1]x\in[1,s]\subseteq[1,r-1]. Therefore

0f(x)f(x)(xx)f(x)>r(rx)20\geq f(\lfloor x\rfloor)-f(x)\geq(\lfloor x\rfloor-x)f^{\prime}(x)>-\frac{r}{(r-x)^{2}}

for any x[1,s]x\in[1,s]. This shows that

0(s1)r(rs)2,0\geq\mathcal{E}\geq-\frac{(s-1)r}{(r-s)^{2}},

which shows that =O(s/r)\mathcal{E}=O(s/r). Therefore

i=1s1iri=1sxrx dx+O(sr)=rlog(r1rs)(s1)+O(sr).\sum_{i=1}^{s-1}\frac{i}{r-i}=\int_{1}^{s}\frac{x}{r-x}\textup{ d}x+O\left(\frac{s}{r}\right)=r\log\left(\frac{r-1}{r-s}\right)-(s-1)+O\left(\frac{s}{r}\right).\qed
Proposition A.2.

Let rkr\geq k be a positive integer growing with kk so that r=(C+ok(1))kr=(C+o_{k\to\infty}(1))k for some constant C1C\geq 1. Then the largest positive integer ss satisfying A.1 also satisfies s=C(1exp(C1)+ok(1))ks=C(1-\exp(-C^{-1})+o_{k\to\infty}(1))k.

Proof.

By the choice of ss, we know

ksi=1s1irik-s\geq\sum_{i=1}^{s-1}\frac{i}{r-i}

and

k(s1)<i=1s2iri.k-(s-1)<\sum_{i=1}^{s-2}\frac{i}{r-i}.

Therefore

ks+O(1)=i=1s1iri+O(srs).k-s+O(1)=\sum_{i=1}^{s-1}\frac{i}{r-i}+O\left(\frac{s}{r-s}\right).

By Lemma A.1, we know that this implies

ks+O(1)=rlog(r1rs)(s1)+O(sr).k-s+O(1)=r\log\left(\frac{r-1}{r-s}\right)-(s-1)+O\left(\frac{s}{r}\right).

Rearranging, we get

r1rs=exp(k1r+O(r1)),\frac{r-1}{r-s}=\exp\left(\frac{k-1}{r}+O(r^{-1})\right),

and so

s=\displaystyle s= 1+(r1)(1exp(k1r+O(r1)))\displaystyle 1+(r-1)\left(1-\exp\left(-\frac{k-1}{r}+O(r^{-1})\right)\right)
=\displaystyle= 1+(C+ok(1))k(1exp(C1+ok(1))),\displaystyle 1+(C+o_{k\to\infty}(1))k\cdot\left(1-\exp\left(-C^{-1}+o_{k\to\infty}(1)\right)\right),

where we use the fact that r1=O(k1)r^{-1}=O(k^{-1}). The desired statement thus follows. ∎

Proposition A.3.

Let k,dk,d be positive integers with d<kd<k, and let s=kds=k-d. Then the smallest positive integer rr satisfying A.1 also satisfies r=(12d+ok(1))k2r=(\frac{1}{2d}+o_{k\to\infty}(1))k^{2}.

Proof.

By the choice of rr, we also have

di=1s1irid\geq\sum_{i=1}^{s-1}\frac{i}{r-i}

and

d<i=1s1ir1i.d<\sum_{i=1}^{s-1}\frac{i}{r-1-i}.

Note that

iriir1i=O(ir2)\frac{i}{r-i}-\frac{i}{r-1-i}=O(ir^{-2})

for every is1i\leq s-1 as we know that rs=Ω(r)r-s=\Omega(r) by Lemma A.1. Therefore

d=i=1s1iri+O(s2r2).d=\sum_{i=1}^{s-1}\frac{i}{r-i}+O(s^{2}r^{-2}).

By Lemma A.1, we know that r=Ω(d1s2)=Ω(s2)r=\Omega(d^{-1}s^{2})=\Omega(s^{2}). Therefore by Lemma A.1,

d=rlog(r1rs)(s1)+O(sr)=rlog(r1rs)(s1)+ok(1).d=r\log\left(\frac{r-1}{r-s}\right)-(s-1)+O\left(\frac{s}{r}\right)=r\log\left(\frac{r-1}{r-s}\right)-(s-1)+o_{k\to\infty}(1).

Note that

rlog(r1rs)=\displaystyle r\log\left(\frac{r-1}{r-s}\right)= rlog(1+s1r+s(s1)r2+O(s3r3))\displaystyle r\log\left(1+\frac{s-1}{r}+\frac{s(s-1)}{r^{2}}+O\left(\frac{s^{3}}{r^{3}}\right)\right)
=\displaystyle= r(s1r+(s+1)(s1)2r2+O(s3r3))\displaystyle r\left(\frac{s-1}{r}+\frac{(s+1)(s-1)}{2r^{2}}+O\left(\frac{s^{3}}{r^{3}}\right)\right)
=\displaystyle= s1+s22r+ok(1).\displaystyle s-1+\frac{s^{2}}{2r}+o_{k\to\infty}(1).

Therefore we get that

rk2=(1+ok(1))rs2=12d+ok(1),\frac{r}{k^{2}}=\left(1+o_{k\to\infty}(1)\right)\frac{r}{s^{2}}=\frac{1}{2d}+o_{k\to\infty}(1),

as desired.