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

Clique-factors in graphs with sublinear \ell-independence number

Jie Han School of Mathematics and Statistics and Center for Applied Mathematics, Beijing Institute of Technology, Beijing, China. Email: han.jie@bit.edu.cn.    Ping Hu School of Mathematics, Sun Yat-sen University, Guangzhou, China. Email: huping9@mail.sysu.edu.cn. Supported in part by National Key Research and Development Program of China (2021YFA1002100) and National Natural Science Foundation of China (11931002).    Guanghui Wang School of Mathematics, Shandong University, Jinan, China, Email: ghwang@sdu.edu.cn. Research supported by Natural Science Foundation of China (11871311,12231018) and Young Taishan Scholars probgram of Shandong Province( 201909001).    Donglei Yang Data Science Institute, Shandong University, Shandong, China. Email: dlyang@sdu.edu.cn. Supported by the China Postdoctoral Science Foundation (2021T140413), Natural Science Foundation of China (12101365) and Natural Science Foundation of Shandong Province (ZR2021QA029).
Abstract

Given a graph GG and an integer 2\ell\geq 2, we denote by α(G)\alpha_{\ell}(G) the maximum size of a KK_{\ell}-free subset of vertices in V(G)V(G). A recent question of Nenadov and Pehova asks for determining the best possible minimum degree conditions forcing clique-factors in nn-vertex graphs GG with α(G)=o(n)\alpha_{\ell}(G)=o(n), which can be seen as a Ramsey–Turán variant of the celebrated Hajnal–Szemerédi theorem. In this paper we find the asymptotical sharp minimum degree threshold for KrK_{r}-factors in nn-vertex graphs GG with α(G)=n1o(1)\alpha_{\ell}(G)=n^{1-o(1)} for all r2r\geq\ell\geq 2.

1 Introduction

Let HH be an hh-vertex graph and GG be an nn-vertex graph. An HH-tiling is a collection of vertex-disjoint copies of HH in GG. An HH-factor is an HH-tiling which covers all vertices of GG. The celebrated Hajnal–Szemerédi theorem [18] states that for all integers n,rn,r with r2r\geq 2 and r|nr|n, any nn-vertex graph GG with δ(G)(11r)n\delta(G)\geq(1-\frac{1}{r})n contains a KrK_{r}-factor. Since then there have been many developments in several directions. From the insight of equitable coloring, Kierstead and Kostochka proved the Hajnal–Szemerédi theorem with an Ore-type degree condition [23]. For a general graph HH, Alon and Yuster [2] first gave an asymptotic result by showing that if δ(G)(11χ(H))n+o(n)\delta(G)\geq\left(1-\frac{1}{\chi(H)}\right)n+o(n), then GG contains an HH-factor, where χ(H)\chi(H) is the chromatic number of HH. Later, Kühn and Osthus [30] managed to characterise, up to an additive constant, the minimum degree condition that forces an HH-factor. There are also several significant generalisations in the setting of partite graphs [22], directed graphs [39] and hypergraphs [36].

1.1 Motivation

Erdős and Sós [14] initiated the study of a variant of Turán problem which excludes all graphs with large independence number. More generally, for an integer 2\ell\geq 2 and a graph GG, the \ell-independence number of GG, denoted by α(G)\alpha_{\ell}(G), is the maximum size of a KK_{\ell}-free subset of vertices. Given integers n,rn,r and a function f(n)f(n), we use RT(n,Kr,f(n))\textbf{RT}_{\ell}(n,K_{r},f(n)) to denote the maximum number of edges of an nn-vertex KrK_{r}-free graph GG with α(G)f(n)\alpha_{\ell}(G)\leq f(n). In particular, the Ramsey–Turán density of KrK_{r} is defined as ϱ(r):=limα0limnRT(n,Kr,αn)(n2)\varrho_{\ell}(r):=\lim\limits_{\alpha\to 0}\lim\limits_{n\to\infty}\frac{\textbf{RT}_{\ell}(n,K_{r},\alpha n)}{\binom{n}{2}}. Szemerédi [38] first showed that ϱ2(4)14\varrho_{2}(4)\leq\frac{1}{4}. This turned out to be sharp as Bollobás and Erdős [7] provided a matching lower bound using an ingenious geometric construction. There are some recent exciting developments in this area [3, 4, 16, 24, 31, 33]. For further information on Ramsey–Turán theory the reader is referred to a comprehensive survey [37] by Simonovits and Sós.

Note that the extremal example that achieves the optimality of the bound on δ(G)\delta(G) in the Hajnal–Szemerédi theorem also has large independence number [6], which makes it far from being typical. Following the spirit of the Ramsey–Turán theory, a natural question on the Hajnal–Szemerédi theorem is whether the minimum degree condition can be weakened when the host graph has sublinear independence number. The following Ramsey–Turán type problem was proposed by Balogh, Molla and Sharifzadeh [6].

Problem 1.1.

[6] Let r3r\geq 3 be an integer and GG be an nn-vertex graph with α(G)=o(n)\alpha(G)=o(n). What is the minimum degree condition on GG that guarantees a KrK_{r}-factor?

Balogh, Molla and Sharifzadeh [6] studied K3K_{3}-factors and showed that if the independence number of an nn-vertex graph GG is o(n)o(n) and δ(G)n2+εn\delta(G)\geq\frac{n}{2}+\varepsilon n for any ε>0\varepsilon>0, then GG contains a triangle factor. Recently Knierim and Su [26] resolved the case r4r\geq 4 by determining the asymptotically tight minimum degree bound (12r)n+o(n)(1-\frac{2}{r})n+o(n).

The following problem was proposed by Nenadov and Pehova [35].

Problem 1.2.

For all r,r,\ell\in\mathbb{N} with r2r\geq\ell\geq 2, let GG be an nn-vertex graph with nrn\in r\mathbb{N} and α(G)=o(n)\alpha_{\ell}(G)=o(n). What is the best possible minimum degree condition on GG that guarantees a KrK_{r}-factor?

Nenadov and Pehova [35] also provided upper and lower bounds on the minimum degree condition. In particular, they solved Problem 1.2 for r=+1r=\ell+1 and proved that n/2+o(n)n/2+o(n) is the correct minimum degree threshold. Knierim and Su [26] reiterated Problem 1.2 in their paper and proposed a minimum degree condition as follows.

Problem 1.3.

[26, 35] Is it true that for every r,r,\ell\in\mathbb{N} with r2r\geq\ell\geq 2 and μ>0\mu>0, there exists α>0\alpha>0 such that for sufficiently large nrn\in r\mathbb{N}, every nn-vertex graph GG with

δ(G)max{rr+μ,12+μ}nandα(G)αn\delta(G)\geq\max\left\{\frac{r-\ell}{r}+\mu,\frac{1}{2}+\mu\right\}n\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ \alpha_{\ell}(G)\leq\alpha n

contains a KrK_{r}-factor?

Very recently, Chang, Han, Kim, Wang and Yang [8] determines the asymptotically optimal minimum degree condition for 34r\ell\geq\frac{3}{4}r, which solves Problem 1.2 for this range, and indeed provides a negative answer to Problem 1.3.

Theorem 1.4.

[8] Let r,r,\ell\in\mathbb{N} such that r>34rr>\ell\geq\frac{3}{4}r. For any μ>0\mu>0, there exists α>0\alpha>0 such that for sufficiently large nrn\in r\mathbb{N}, every nn-vertex graph GG with

δ(G)(12ϱ(r1)+μ)nandα(G)αn\delta(G)\geq\left(\frac{1}{2-\varrho_{\ell}(r-1)}+\mu\right)n\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ \alpha_{\ell}(G)\leq\alpha n

contains a KrK_{r}-factor. Moreover, the minimum degree condition is asymptotically best possible.

Based on this result, Problem 1.3 should be revised as follows.

Problem 1.5.

Is it true that δ(G)max{rr+μ,12ϱ(r1)+μ}n\delta(G)\geq\max\{\frac{r-\ell}{r}+\mu,\frac{1}{2-\varrho_{\ell}(r-1)}+\mu\}n suffices in Problem 1.3?

1.2 Main results and discussions

By the aformentioned results, Problem 1.5 is solved for =2\ell=2 [26] and for 3r/4\ell\geq 3r/4 [8], both done by quite involved proofs. It seems to us that a complete resolution of Problem 1.5 is a quite challenging task.

The purpose of this paper is to extend the discussion on the problem to sublinear \ell-independence numbers. We first state a simplified version of our main result which says that the answer to Problem 1.5 is yes if we assume a slightly stronger assumption α(G)n1o(1)\alpha_{\ell}(G)\leq n^{1-o(1)}.

Theorem 1.6.

For μ,c(0,1)\mu,c\in(0,1), r,r,\ell\in\mathbb{N} such that r>2r>\ell\geq 2, the following holds for sufficiently large nrn\in r\mathbb{N}. Every nn-vertex graph GG with δ(G)max{rr+μ,12ϱ(r1)+μ}n\delta(G)\geq\max\{\frac{r-\ell}{r}+\mu,\frac{1}{2-\varrho_{\ell}(r-1)}+\mu\}n and α(G)nc\alpha_{\ell}(G)\leq n^{c} contains a KrK_{r}-factor.

To state our main result, we define more general versions of the Ramsey–Turán densities as follows.

Definition 1.7.

Let f(n)f(n) be a monotone increasing function, and r,,α(0,1)r,\ell\in\mathbb{N},\alpha\in(0,1).

  1. (ii)

    Let RT(n,Kr,f(αn))\textbf{RT}_{\ell}(n,K_{r},f(\alpha n)) be the maximum integer mm for which there exists an nn-vertex KrK_{r}-free graph GG with e(G)=me(G)=m and α(G)f(αn)\alpha_{\ell}(G)\leq f(\alpha n). Then let

    ϱ(r,f):=limα0lim supnRT(n,Kr,f(αn))(n2).\varrho_{\ell}(r,f):=\lim\limits_{\alpha\to 0}\limsup\limits_{n\to\infty}\frac{\textbf{RT}_{\ell}(n,K_{r},f(\alpha n))}{\binom{n}{2}}.
  2. (iiii)

    Let RT(n,Kr,f(αn))\textbf{RT}^{*}_{\ell}(n,K_{r},f(\alpha n)) be the maximum integer δ\delta for which there exists an nn-vertex KrK_{r}-free graph GG with δ(G)=δ\delta(G)=\delta and α(G)f(αn)\alpha_{\ell}(G)\leq f(\alpha n). Then let

    ϱ(r,f):=limα0lim supnRT(n,Kr,f(αn))n.\varrho^{*}_{\ell}(r,f):=\lim\limits_{\alpha\to 0}\limsup\limits_{n\to\infty}\frac{\textbf{RT}^{*}_{\ell}(n,K_{r},f(\alpha n))}{n}.

By definition trivially it holds that ϱ(r,f)ϱ(r,f)\varrho^{*}_{\ell}(r,f)\leq\varrho_{\ell}(r,f). It is proved in [8] that ϱ(r,f)=ϱ(r,f)\varrho^{*}_{\ell}(r,f)=\varrho_{\ell}(r,f) if there exists c(0,1)c\in(0,1) such that xf(n)f(x1/cn)xf(n)\leq f(x^{1/c}n) for every x(0,1)x\in(0,1) and nn\in\mathbb{N}.

The full version of our result is stated as follows.

Theorem 1.8.

For μ>0\mu>0, r,r,\ell\in\mathbb{N} such that r>2r>\ell\geq 2, there exists α>0\alpha>0 such that the following holds for sufficiently large nrn\in r\mathbb{N}. Let λ=1/r+1\lambda=1/\lfloor\frac{r}{\ell}+1\rfloor and f(n)n1ω(n)logλnf(n)\leq n^{1-\omega(n)\log^{-\lambda}n} be a monotone increasing function, where ω(n)\omega(n)\rightarrow\infty slowly.111The strange-looking function n1ω(n)logλnn^{1-\omega(n)\log^{-\lambda}n} satisfies n1ε<n1ω(n)logλn<nlognn^{1-\varepsilon}<n^{1-\omega(n)\log^{-\lambda}n}<\frac{n}{\log n} for any constant ε>0\varepsilon>0. Then every nn-vertex graph GG with

δ(G)max{rr+μ,12ϱ(r1,f)+μ}nandα(G)<f(αn)\delta(G)\geq\max\left\{\frac{r-\ell}{r}+\mu,\frac{1}{2-\varrho^{*}_{\ell}(r-1,f)}+\mu\right\}n\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ \alpha_{\ell}(G)<f(\alpha n)

contains a KrK_{r}-factor.

In fact, Theorem 1.8 implies Theorem 1.6 by the observation that ϱ(r1,f)ϱ(r1)=ϱ(r1)\varrho^{*}_{\ell}(r-1,f)\leq\varrho^{*}_{\ell}(r-1)=\varrho_{\ell}(r-1) holds for any function f(n)=ncf(n)=n^{c} with c(0,1)c\in(0,1). We shall supply lower bound constructions (see next subsection) which show that the minimum degree condition in Theorem 1.8 is asymptotically best possible for α(G)(n1γ,n1ε)\alpha_{\ell}(G)\in(n^{1-\gamma},n^{1-\varepsilon}), for any γ(0,12+2)\gamma\in(0,\frac{\ell-1}{\ell^{2}+2\ell}) and any ε>0\varepsilon>0. This can be seen as a stepping stone towards a full understanding of the Ramsey–Turán tiling thresholds for cliques where α(G)[2,o(n)]\alpha_{\ell}(G)\in[2,o(n)].

Here we also provide some concrete thresholds that we could spell out from Theorem 1.8. Recall that Nenadov and Pehova [35] solved Problem 1.2 for r=+1r=\ell+1 whilst Knierim and Su [26] solved the case =2,r4\ell=2,r\geq 4. Now we consider the first open case r=+2,3r=\ell+2,\ell\geq 3. Note that ϱ(+1,f)=ϱ(+1)=0\varrho^{*}_{\ell}(\ell+1,f)=\varrho_{\ell}(\ell+1)=0. Then Theorem 1.4 says that for 6\ell\geq 6 (that is, 34(+2)\ell\geq\frac{3}{4}(\ell+2)) and α(G)=o(n)\alpha_{\ell}(G)=o(n), n2+o(n)\frac{n}{2}+o(n) is the minimum degree threshold forcing a K+2K_{\ell+2}-factor. For the remaining cases {3,4,5}\ell\in\{3,4,5\}, Theorem 1.8 implies that the minimum degree threshold is n2+o(n)\frac{n}{2}+o(n) under a stronger condition α(G)n1ε\alpha_{\ell}(G)\leq n^{1-\varepsilon} for any fixed ε>0\varepsilon>0.

Our proof of Theorem 1.8 uses the absorption method and the regularity method. In particular, we use dependent random choice for embedding cliques in regular tuples with sublinear independence number, which is closely related to the Ramsey–Turán problem.

1.3 Sharpness of the minimum degree condition

We note that both terms in the minimum degree condition in Theorem 1.8 are asymptotically best possible. First, we show that the first term cannot be weakened when α(G)(n1γ,o(n))\alpha_{\ell}(G)\in(n^{1-\gamma},o(n)) for some constant γ\gamma as follows.

Proposition 1.9.

Given integers r,r,\ell\in\mathbb{N} with r>2r>\ell\geq 2 and constants η,γ\eta,\gamma with η(0,rr)\eta\in(0,\frac{r-\ell}{r}), γ(0,12+2)\gamma\in(0,\frac{\ell-1}{\ell^{2}+2\ell}), the following holds for all sufficiently large nn\in\mathbb{N} and constant μ:=rr(rrη)\mu:=\tfrac{r}{r-\ell}(\tfrac{r-\ell}{r}-\eta). There exists an nn-vertex graph GG with δ(G)ηn\delta(G)\geq\eta n and α(G)<n1γ\alpha_{\ell}(G)<n^{1-\gamma} such that every KrK_{r}-tiling in GG covers at most (1μ)n(1-\mu)n vertices.

Indeed, Proposition 1.9 also gives, in the setting that α(G)n1o(1)\alpha_{\ell}(G)\leq n^{1-o(1)}, a lower bound construction for the minimum degree condition forcing an almost KrK_{r}-tiling that leaves a constant number of vertices uncovered. More results on almost graph tilings can be found in a recent comprehensive paper [19].

The second term 12ϱ(r1,f)\frac{1}{2-\varrho^{*}_{\ell}(r-1,f)} is also asymptotically tight, which is given by a cover threshold construction as follows.
Cover Threshold. To have a KrK_{r}-factor in GG, a naive necessary condition is that every vertex vV(G)v\in V(G) is covered by a copy of KrK_{r} in GG. The cover threshold has been first discussed in [20] and appeared in a few different contexts [8, 9, 12].

Now we give a construction that shows the optimality of the term 12ϱ(r1,f)\frac{1}{2-\varrho^{*}_{\ell}(r-1,f)} for the function f(n)f(n) as in Theorem 1.8. A similar construction can be found in [8]. Given integers r,r,\ell and constants ε,α,x(0,1)\varepsilon,\alpha,x\in(0,1), we construct (for large nrn\in r\mathbb{N}) an nn-vertex graph GG by

  1. (i)(i)

    first fixing a vertex vv such that N(v)=xnN(v)=xn and G[N(v)]=:GG[N(v)]=:G^{\prime} is a Kr1K_{r-1}-free subgraph with δ(G)ϱ(r1,f)xnεn\delta(G^{\prime})\geq\varrho^{*}_{\ell}(r-1,f)xn-\varepsilon n and α(G)f(αxn)\alpha_{\ell}(G^{\prime})\leq f(\alpha xn);

  2. (ii)(ii)

    and then adding a clique of size nxn1n-xn-1 that is complete to N(v)N(v).

There exists no copy of KrK_{r} covering vv and thus GG contains no KrK_{r}-factor; moreover, by choosing x=12ϱ(r1,f)x=\tfrac{1}{2-\varrho^{*}_{\ell}(r-1,f)}, we obtain δ(G)12ϱ(r1,f)nεn\delta(G)\geq\frac{1}{2-\varrho^{*}_{\ell}(r-1,f)}n-\varepsilon n and α(G)=α(G)f(αn)\alpha_{\ell}(G)=\alpha_{\ell}(G^{\prime})\leq f(\alpha n).

Notation. Throughout the paper we follow standard graph-theoretic notation [11]. For a graph G=(V,E)G=(V,E), let v(G)=|V|v(G)=|V| and e(G)=|E|e(G)=|E|. For UVU\subseteq V, G[U]G[U] denotes the induced subgraph of GG on UU. The notation GUG-U is used to denote the induced subgraph after removing UU, that is, GU:=G[VU]G-U:=G[V\setminus U]. For two subsets A,BV(G)A,B\subseteq V(G), we use e(A,B)e(A,B) to denote the number of edges joining AA and BB. Given a vertex vv(G)v\in v(G) and XV(G)X\subseteq V(G), denote by NX(v)N_{X}(v) the set of neighbors of vv in XX and let dX(v):=|NX(v)|d_{X}(v):=|N_{X}(v)|. In particular, we write NG(v)N_{G}(v) for the set of neighbors of vv in GG. We omit the index GG if the graph is clear from the context. Given a set VV and an integer kk, we write (Vk)\binom{V}{k} for the family of all kk-subsets of VV. For all integers a,ba,b with aba\leq b, let [a,b]:={i:aib}[a,b]:=\{i\in\mathbb{Z}:a\leq i\leq b\} and [a]:={1,2,,a}[a]:=\{1,2,\ldots,a\}.

When we write αβγ\alpha\ll\beta\ll\gamma, we always mean that α,β,γ\alpha,\beta,\gamma are constants in (0,1)(0,1), and βγ\beta\ll\gamma means that there exists β0=β0(γ)\beta_{0}=\beta_{0}(\gamma) such that the subsequent arguments hold for all 0<ββ00<\beta\leq\beta_{0}. Hierarchies of other lengths are defined analogously. In the remaining proofs, we always take λ=1/r+1\lambda=1/\lfloor\frac{r}{\ell}+1\rfloor and f(n)n1ω(n)logλnf(n)\leq n^{1-\omega(n)\log^{-\lambda}n} unless otherwise stated.

2 Proof strategy and Preliminaries

Our proof uses the absorption method, pioneered by the work of Rödl, Ruciński and Szemerédi [36] on perfect matchings in hypergraphs, though similar ideas already appeared implicitly in previous works, e.g. Krivelevich [29]. A key step in the absorption method for HH-factor problem is to show that for every set of h:=|V(H)|h:=|V(H)| vertices, the host graph GG contains Ω(nb)\Omega(n^{b}) bb-vertex absorbers (to be defined shortly). However, as pointed out in [6], in our setting this is usually impossible because when we construct the absorbers using the independence number condition, it does not give such a strong counting. Instead, a much weaker notion has been used in this series of works, that is, we aim to show that for (almost) every set of hh vertices, the host graph GG contains Ω(n)\Omega(n) vertex-disjoint absorbers. Note that this weak notion of absorbers have been successfully used in our setting [35, 26] and the randomly perturbed setting [9].

2.1 The absorption method

Following typical absorption strategies, our main work is to establish an absorbing set (see Lemma 2.2) and find an almost-perfect tiling (see Lemma 2.3). We first introduce the following notions of absorbers and absorbing sets from [35].

Definition 2.1.

Let HH be a graph with hh vertices and GG be a graph with nn vertices.

  1. 1.

    We say that a subset AV(G)A\subseteq V(G) is a ξ\xi-absorbing set for some ξ>0\xi>0 if for every subset RV(G)AR\subseteq V(G)\setminus A with |R|ξn|R|\leq\xi n and |AR|h|A\cup R|\in h\mathbb{N}, G[AR]G[A\cup R] contains an HH-factor.

  2. 2.

    Given a subset SV(G)S\subseteq V(G) of size hh and an integer tt, we say that a subset ASV(G)SA_{S}\subseteq V(G)\setminus S is an (S,t)(S,t)-absorber if |AS|ht|A_{S}|\leq ht and both G[AS]G[A_{S}] and G[ASS]G[A_{S}\cup S] contain an HH-factor.

Now we are ready to state our first crucial lemma, whose proof can be found in Section 4.

Lemma 2.2 (Absorbing Lemma).

Given positive integers r,r,\ell with r>2r>\ell\geq 2 and constants μ,γ\mu,\gamma with 0<γ<μ20<\gamma<\frac{\mu}{2}, there exist α,ξ>0\alpha,\xi>0 such that the following holds for sufficiently large nrn\in r\mathbb{N}. Let GG be an nn-vertex graph with δ(G)max{rr+μ,12ϱ(r1,f)+μ}n\delta(G)\geq\max\{\frac{r-\ell}{r}+\mu,\frac{1}{2-\varrho^{*}_{\ell}(r-1,f)}+\mu\}n and α(G)<f(αn)\alpha_{\ell}(G)<f(\alpha n). Then GG contains a ξ\xi-absorbing set AA of size at most γn\gamma n.

Our second crucial lemma is on almost KrK_{r}-factor as follows, whose proof will be given in Section 3.

Lemma 2.3 (Almost perfect tiling).

Given positive integers r,r,\ell such that r>2r>\ell\geq 2 and positive constants μ,δ\mu,\delta, the following statement holds for sufficiently large nrn\in r\mathbb{N}. Every nn-vertex graph GG with δ(G)(rr+μ)n\delta(G)\geq\left(\frac{r-\ell}{r}+\mu\right)n and α(G)<f(n)\alpha_{\ell}(G)<f(n) contains a KrK_{r}-tiling that leaves at most δn\delta n vertices in GG uncovered.

Now we are ready to prove Theorem 1.8 using Lemma 2.2 and Lemma 2.3.

Proof of Theorem 1.8.

Given any positive integers ,r\ell,r with r>2r>\ell\geq 2 and a constant μ>0\mu>0. Choose 1nαδξγμ\frac{1}{n}\ll\alpha\ll\delta\ll\xi\ll\gamma\ll\mu. Let GG be an nn-vertex graph with

δ(G)max{rr+μ,12ϱ(r1,f)+μ}nandα(G)<f(αn).\delta(G)\geq\max\left\{\frac{r-\ell}{r}+\mu,\frac{1}{2-\varrho^{*}_{\ell}(r-1,f)}+\mu\right\}n\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ \alpha_{\ell}(G)<f(\alpha n).

By Lemma 2.2 with γμ2\gamma\leq\frac{\mu}{2}, we find a ξ\xi-absorbing set AV(G)A\subseteq V(G) of size at most γn\gamma n for some ξ>0\xi>0. Let G1:=GAG_{1}:=G-A. Then we have δ(G1)(rr+μ)nγn(rr+μ2)n\delta(G_{1})\geq\left(\tfrac{r-\ell}{r}+\mu\right)n-\gamma n\geq\left(\tfrac{r-\ell}{r}+\tfrac{\mu}{2}\right)n. Therefore by applying Lemma 2.3 on G1G_{1} with δ\delta, we obtain a KrK_{r}-tiling \mathcal{M} that covers all but a set RR of at most δn\delta n vertices in G1G_{1}. Since δξ\delta\ll\xi, the absorbing property of AA implies that G[AR]G[A\cup R] contains a KrK_{r}-factor \mathcal{R}, which together with \mathcal{M} forms a KrK_{r}-factor in GG. ∎

3 Finding almost perfect tilings

In this section we address Lemma 2.3. The proof of Lemma 2.3 uses the regularity method, a tiling result of Komlós (Theorem 3.5), and dependent random choice (Lemma 3.6). We shall first give the crucial notion of regularity and then introduce the powerful Szemerédi’s Regularity Lemma.

3.1 Regularity

Given a graph GG and a pair (X,Y)(X,Y) of vertex-disjoint subsets in V(G)V(G), the density of (X,Y)(X,Y) is defined as

d(X,Y)=e(X,Y)|X||Y|.d(X,Y)=\frac{e(X,Y)}{|X||Y|}.

For constants ε,d>0\varepsilon,d>0, we say that (X,Y)(X,Y) is an ε\varepsilon-regular pair with density at least dd (or (X,Y)(X,Y) is (ε,d)(\varepsilon,d)-regular) if d(X,Y)dd(X,Y)\geq d and for all XXX^{\prime}\subseteq X, YYY^{\prime}\subseteq Y with |X|ε|X||X^{\prime}|\geq\varepsilon|X|, |Y|ε|Y||Y^{\prime}|\geq\varepsilon|Y|, we have

|d(X,Y)d(X,Y)|ε.|d(X^{\prime},Y^{\prime})-d(X,Y)|\leq\varepsilon.

Moreover, a pair (X,Y)(X,Y) is called (ε,d)(\varepsilon,d)-supersuper-regularregular if (X,Y)(X,Y) is (ε,d)(\varepsilon,d)-regular, dY(x)d|Y|d_{Y}(x)\geq d|Y| for all xXx\in X and dX(y)d|X|d_{X}(y)\geq d|X| for all yYy\in Y. The following fact is an easy consequence of the definition of regularity.

Fact 3.1.

Given constants d,η>ε>0d,\eta>\varepsilon>0 and a bipartite graph G=(XY,E)G=(X\cup Y,E), if (X,Y)(X,Y) is (ε,d)(\varepsilon,d)-regular, then for all X1XX_{1}\subseteq X and Y1YY_{1}\subseteq Y with |X1|η|X||X_{1}|\geq\eta|X| and |Y1|η|Y||Y_{1}|\geq\eta|Y|, we have that (X1,Y1)(X_{1},Y_{1}) is (ε,dε)(\varepsilon^{\prime},d-\varepsilon)-regular in GG for any εmax{εη,2ε}\varepsilon^{\prime}\geq{\rm max}\{\frac{\varepsilon}{\eta},2\varepsilon\}.

Given a family of vertex-disjoint sets in V(G)V(G) which are pairwise ε\varepsilon-regular, we can find in each set a large subset such that every pair of resulting subsets is super-regular.

Proposition 3.2 (see Proposition 2.6 in [8]).

Given a constant ε>0\varepsilon>0 and integers m,tm,t with t<12εt<\frac{1}{2\varepsilon}, let GG be an nn-vertex graph and V1,V2,,Vt+1V_{1},V_{2},\ldots,V_{t+1} be vertex-disjoint subsets each of size mm in GG such that every pair (Vi,Vj)(V_{i},V_{j}) is ε\varepsilon-regular with density dij:=d(Vi,Vj)d_{ij}:=d(V_{i},V_{j}). Then there exists for each i[t+1]i\in[t+1] a subset ViViV^{\prime}_{i}\subseteq V_{i} of size at least (1tε)m(1-t\varepsilon)m such that every pair (Vi,Vj)(V_{i}^{\prime},V_{j}^{\prime}) is (2ε,dij(t+1)ε)(2\varepsilon,d_{ij}-(t+1)\varepsilon)-super-regular.

We now state a degree form of the regularity lemma (see [27, Theorem 1.10]).

Lemma 3.3 (Degree form of the Regularity Lemma [27]).

For every ε>0\varepsilon>0 there is an N=N(ε)N=N(\varepsilon) such that the following holds for any real number d[0,1]d\in[0,1] and nn\in\mathbb{N}. Let G=(V,E)G=(V,E) be a graph with nn vertices. Then there exists a partition 𝒫={V0,,Vk}\mathcal{P}=\{V_{0},\ldots,V_{k}\} of VV and a spanning subgraph GGG^{\prime}\subseteq G with the following properties:

  1. (a)(a)

    1εkN\frac{1}{\varepsilon}\leq k\leq N;

  2. (b)(b)

    |Vi|εn|V_{i}|\leq\varepsilon n for 0ik0\leq i\leq k and |V1|=|V2|==|Vk|=m|V_{1}|=|V_{2}|=\cdots=|V_{k}|=m for some mm\in\mathbb{N};

  3. (c)(c)

    dG(v)>dG(v)(d+ε)nd_{G^{\prime}}(v)>d_{G}(v)-(d+\varepsilon)n for every vV(G)v\in V(G);

  4. (d)(d)

    every ViV_{i} is an independent set in GG^{\prime};

  5. (e)(e)

    each pair (Vi,Vj)(V_{i},V_{j}), 1i<jk1\leq i<j\leq k is ε\varepsilon-regular in GG^{\prime} with density 0 or at least dd.

A widely-used auxiliary graph accompanied with the regular partition is the reduced graph. The dd-reduced graph RdR_{d} of 𝒫\mathcal{P} is a graph defined on the vertex set {V1,,Vk}\{V_{1},\ldots,V_{k}\} such that ViV_{i} is connected to VjV_{j} by an edge if (Vi,Vj)(V_{i},V_{j}) has density at least dd in GG^{\prime}. So if ViV_{i} is not connected to VjV_{j}, then (Vi,Vj)(V_{i},V_{j}) has density 0 by property (e)(e) above. To ease the notation, we use dR(Vi)d_{R}(V_{i}) to denote the degree of ViV_{i} in RdR_{d} for each i[k]i\in[k]. Note that RdR_{d} also can be regarded as a weighted graph in which the weight for each edge ViVjV_{i}V_{j}, denoted by dijd_{ij} for simplicity, is exactly the density of the pair (Vi,Vj)(V_{i},V_{j}) in GG^{\prime}.

Fact 3.4.

For positive constants d,εd,\varepsilon and cc, let G=(V,E)G=(V,E) be a graph on nn vertices with δ(G)cn\delta(G)\geq cn. Let GG^{\prime} and 𝒫\mathcal{P} be obtained by applying Lemma 3.3 on GG with constants dd and ε\varepsilon. Let RdR_{d} be the dd-reduced graph as given above. Then for every ViV(Rd)V_{i}\in V(R_{d}) we have dR(Vi)(c2εd)k.d_{R}(V_{i})\geq(c-2\varepsilon-d)k.

Proof.

Note that |V0|εn|V_{0}|\leq\varepsilon n and |Vi|=m|V_{i}|=m for each ViV(Rd)V_{i}\in V(R_{d}). Thus we have

VjVidij|Vi||Vj|=eG(Vi,jiVj)(δ(G)|V0|)|Vi|(c2εd)nm,\sum_{V_{j}\sim V_{i}}d_{ij}|V_{i}||V_{j}|=e_{G^{\prime}}(V_{i},\cup_{j\neq i}V_{j})\geq(\delta(G^{\prime})-|V_{0}|)|V_{i}|\geq\left(c-2\varepsilon-d\right)nm,

which implies

dR(Vi)=VjVi1VjVidij(c2εd)nmm2(c2εd)k.d_{R}(V_{i})=\sum_{V_{j}\sim V_{i}}1\geq\sum_{V_{j}\sim V_{i}}d_{ij}\geq\frac{(c-2\varepsilon-d)nm}{m^{2}}\geq(c-2\varepsilon-d)k.

To find an almost perfect KrK_{r}-tiling, we shall also make use of the following result of Komlós [28] on graph tilings. Given a graph HH on rr vertices, the critical chromatic number of HH is defined as χcr(H)=(k1)rrσ,\chi_{cr}(H)=\frac{(k-1)r}{r-\sigma}, where k=χ(H)k=\chi(H) and σ=σ(H)\sigma=\sigma(H) denotes the smallest size of a color class over all kk-colorings of HH.

Theorem 3.5 (Komlós [28]).

Given any graph HH and a constant γ>0\gamma>0, there exists an integer n0=n0(γ,H)n_{0}=n_{0}(\gamma,H) such that every graph GG of order nn0n\geq n_{0} with δ(G)(11χcr(H))n\delta(G)\geq\left(1-\frac{1}{\chi_{cr}(H)}\right)n contains an HH-tiling covering all but at most γn\gamma n vertices.

Based on this result, we will first apply the regularity lemma to GG to get a reduced graph R:=RdR:=R_{d} for a constant d>0d>0, and then apply Theorem 3.5 to get an HH-tiling of RR covering almost all vertices for a suitably-chosen auxiliary graph HH. To get an almost KrK_{r}-tiling of GG from this almost HH-tiling of RR, we will use the following lemma which says if we can find a KqK_{q} in a copy of HH in RR, then we can find a KpqK_{pq} in GG under certain conditions on αp(G)\alpha_{p}(G). Its proof follows from that of Claim 6.1 in [5], where a similar assumption on α(G)\alpha(G) (instead of αp(G)\alpha_{p}(G)) is used. For completeness we include a proof of Lemma 3.6 in the appendix.

Lemma 3.6.

Given a constant d>0d>0 and integers p,q2p,q\geq 2, there exist C,εC,\varepsilon such that for any constant η>0\eta>0 the following holds for every sufficiently large nn\in\mathbb{N} and g(n):=n1Clog1/qng(n):=n^{1-C\log^{-1/q}n}. Let GG be an nn-vertex graph with αp(G)<g(n)\alpha_{p}(G)<g(n) and V1,V2,,VqV_{1},V_{2},\ldots,V_{q} be pairwise vertex-disjoint sets of vertices in GG with |Vi|ηn|V_{i}|\geq\eta n for each i[q]i\in[q] and every pair (Vi,Vj)(V_{i},V_{j}) being (ε,d)(\varepsilon,d)-regular. Then there exists a copy of KpqK_{pq} in GG which contains exactly pp vertices in each ViV_{i} for i[q]i\in[q].

3.2 Proof of Lemma 2.3

Proof of Lemma 2.3.

Given r,r,\ell\in\mathbb{N} such that r>2r>\ell\geq 2 and μ>0,δ>0\mu>0,\delta>0, we choose

1nεμ,δ,1r.\tfrac{1}{n}\ll\varepsilon\ll\mu,\delta,\tfrac{1}{r}.

Let GG be an nn-vertex graph with δ(G)(rr+μ)n\delta(G)\geq\left(\frac{r-\ell}{r}+\mu\right)n and α(G)<f(n)\alpha_{\ell}(G)<f(n). By applying Lemma 3.3 on GG with constants ε>0\varepsilon>0 and d:=μ4d:=\frac{\mu}{4}, we obtain a partition 𝒫={V0,,Vk}\mathcal{P}=\{V_{0},\ldots,V_{k}\} for some 1εkN\frac{1}{\varepsilon}\leq k\leq N and a spanning subgraph GGG^{\prime}\subseteq G with properties (a)-(e) as stated. Let m:=|Vi|m:=|V_{i}| for all i[k]i\in[k] and RdR_{d} be the corresponding dd-reduced graph of 𝒫\mathcal{P}. Then it follows from Fact 3.4 that δ(Rd)(rr+μ4)k\delta(R_{d})\geq(\frac{r-\ell}{r}+\frac{\mu}{4})k.

Let r=x+yr=x\ell+y for some integers x,yx,y with x1,1yx\geq 1,1\leq y\leq\ell. Note that the complete (x+1)(x+1)-partite graph H:=Ky,,,H:=K_{y,\ell,\ldots,\ell} has χcr(H)=r\chi_{cr}(H)=\frac{r}{\ell}. Now we apply Theorem 3.5 on RdR_{d} with γ=δ2\gamma=\frac{\delta}{2} and H=Ky,,,H=K_{y,\ell,\ldots,\ell} to obtain a family \mathcal{H} of vertex-disjoint copies of HH that cover all but at most δ2k\frac{\delta}{2}k vertices of RdR_{d}.

Given a copy of HH in \mathcal{H}, without loss of generality, we may assume that its vertex set is {V1,,Vr}\{V_{1},\ldots,V_{r}\} together with the parts denoted by

𝒲1={V1,,Vy}and𝒲s+1={Vy+1+(s1),,Vy+s}fors[x].\mathcal{W}_{1}=\{V_{1},\ldots,V_{y}\}\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ \mathcal{W}_{s+1}=\{V_{y+1+(s-1)\ell},\ldots,V_{y+s\ell}\}\leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ s\in[x].

Note that every pair of clusters Vi,VjV_{i},V_{j} from distinct parts forms an ε\varepsilon-regular pair with density at least dd.

We shall greedily embed in the original graph GG vertex-disjoint copies of KrK_{r} that together cover almost all the vertices in i=1rVi\cup_{i=1}^{r}V_{i}. Now for each i[y]i\in[y] we divide ViV_{i} arbitrarily into \ell subclusters Vi,1,,Vi,V_{i,1},\ldots,V_{i,\ell} of (almost) equal size. For each j[y+1,r]j\in[y+1,r] we divide VjV_{j} into yy subclusters Vj,1,,Vj,yV_{j,1},\ldots,V_{j,y} of (almost) equal size. Here for simplicity we may further assume that |Vi,i|=m|V_{i,i^{\prime}}|=\frac{m}{\ell} for i[y],i[]i\in[y],i^{\prime}\in[\ell] and |Vj,j|=my|V_{j,j^{\prime}}|=\frac{m}{y} for every j[y+1,r],j[y]j\in[y+1,r],j^{\prime}\in[y]. We call a family {Vis,js}s=1x+1\{V_{i_{s},j_{s}}\}_{s=1}^{x+1} of x+1x+1 subclusters legal if Vis𝒲sV_{i_{s}}\in\mathcal{W}_{s} for each s[x+1]s\in[x+1], i.e., {Vis}s=1x+1\{V_{i_{s}}\}_{s=1}^{x+1} forms a copy of Kx+1K_{x+1} in RdR_{d}. Note that each 𝒲s\mathcal{W}_{s} (s[x+1]s\in[x+1]) contains exactly yy\ell subclusters in total. Therefore we can greedily partition the set of all subclusters into yy\ell pairwise disjoint legal families.

Now if we have a KrK_{r}-tiling in GG for every legal family {Vis,js}s=1x+1\{V_{i_{s},j_{s}}\}_{s=1}^{x+1}, that covers all but at most rδ4ym\frac{r\delta}{4y\ell}m vertices of s=1x+1Vis,js\bigcup_{s=1}^{x+1}V_{i_{s},j_{s}}, then we can find a KrK_{r}-tiling covering all but at most rδ4m\frac{r\delta}{4}m vertices of V1V2VrV_{1}\cup V_{2}\cup\cdots\cup V_{r}. Applying this to all copies of HH from \mathcal{H} would give us a KrK_{r}-tiling in GG covering all but at most

|V0|+δ2km+||rδ4m<εn+δ2n+δ4n<δn|V_{0}|+\tfrac{\delta}{2}km+|\mathcal{H}|\tfrac{r\delta}{4}m<\varepsilon n+\tfrac{\delta}{2}n+\tfrac{\delta}{4}n<\delta n

vertices. So to complete the proof of Lemma 2.3, it is sufficient to prove the following claim.

Claim 3.7.

Given any legal family {Vis,js}s=1x+1\{V_{i_{s},j_{s}}\}_{s=1}^{x+1}, G[s=1x+1Vis,js]G[\bigcup_{s=1}^{x+1}V_{i_{s},j_{s}}] admits a KrK_{r}-tiling covering all but at most rδ4ym\frac{r\delta}{4y\ell}m vertices of s=1x+1Vis,js\bigcup_{s=1}^{x+1}V_{i_{s},j_{s}}.

Proof of claim.

For convenience, we write Ys:=Vis,jsY_{s}:=V_{i_{s},j_{s}} with s[x+1]s\in[x+1]. Recall that |Y1|=m|Y_{1}|=\frac{m}{\ell} and |Ys|=my|Y_{s}|=\frac{m}{y} for s[2,x+1]s\in[2,x+1]. If we can greedily pick vertex-disjoint copies of KrK_{r} such that each contains exactly yy vertices in Y1Y_{1} and \ell vertices in YsY_{s} for each s[2,x+1]s\in[2,x+1], then almost all vertices in s=1x+1Ys\cup_{s=1}^{x+1}Y_{s} can be covered in this way. Now it suffices to show that for any YsYsY_{s}^{\prime}\subseteq Y_{s} with s[x+1]s\in[x+1], each of size at least δ4ym\frac{\delta}{4y\ell}m, there exists a copy of KrK_{r} with exactly yy vertices inside Y1Y_{1}^{\prime} and \ell vertices inside each YsY_{s}^{\prime}.

For any distinct s,t[x+1]s,t\in[x+1], the pair (Vis,Vit)(V_{i_{s}},V_{i_{t}}) is ε\varepsilon-regular with density at least dd. Then Fact 3.1 implies that every two sets from Y1,,Yx+1Y_{1}^{\prime},\ldots,Y_{x+1}^{\prime} form an ε\varepsilon^{\prime}-regular pair with density at least dεd-\varepsilon, where ε=4yδε\varepsilon^{\prime}=\frac{4y\ell}{\delta}\varepsilon. Therefore as 1nεδ,1r\frac{1}{n}\ll\varepsilon\ll\delta,\frac{1}{r}, by applying Lemma 3.6 on GG with Vi=Yi,p=,q=x+1,η=δ4ymnV_{i}=Y_{i}^{\prime},p=\ell,q=x+1,\eta=\frac{\delta}{4y\ell}\frac{m}{n} and the fact that f(n)g(n)f(n)\leq g(n), we obtain a copy of K(x+1)K_{(x+1)\ell} which contains exactly \ell vertices in each YsY_{s}^{\prime} for s[x+1]s\in[x+1]. Thus we obtain a desired copy of KrK_{r} by discarding arbitrary y\ell-y vertices from Y1Y_{1}^{\prime} from the clique above. ∎

4 Building an absorbing set

The construction of an absorbing set is now known via a novel idea of Montgomery [34], provided that (almost) every set of hh vertices has linearly many vertex-disjoint absorbers as aforementioned. Such an approach is summarized as the following result by Nenadov and Pehova [35].

Lemma 4.1.

[35] Let HH be a graph with hh vertices and let γ>0\gamma>0 and tt\in\mathbb{N} be constants. Then there exist ξ=ξ(h,t,γ)\xi=\xi(h,t,\gamma) and n0n_{0}\in\mathbb{N} such that the following statement holds. Suppose that GG is a graph with nn0n\geq n_{0} vertices such that every S(V(G)h)S\in\tbinom{V(G)}{h} has a family of at least γn\gamma n vertex-disjoint (S,t)(S,t)-absorbers. Then GG contains a ξ\xi-absorbing set of size at most γn\gamma n.

4.1 Finding absorbers

In order to find linearly many vertex-disjoint absorbers for (almost) every hh-subset, we shall use a notion of reachability introduced by Lo and Markström [32]. Here we introduce a slightly different version in our setup. Let GG be a graph of nn vertices and HH be a graph of hh vertices. For any two vertices u,vV(G)u,v\in V(G), a set SV(G)S\subset V(G) is called an HH-reachable set for {u,v}\{u,v\} if both G[{u}S]G[\{u\}\cup S] and G[{v}S]G[\{v\}\cup S] have HH-factors. For t1t\geq 1 and β>0\beta>0, we say that two vertices uu and vv are (H,β,t)(H,\beta,t)-reachable (in GG) if there are βn\beta n vertex-disjoint HH-reachable sets SS in GG, each of size at most ht1ht-1. Moreover, we say that a vertex set UV(G)U\subseteq V(G) is (H,β,t)(H,\beta,t)-closed if any two vertices in UU are (H,β,t)(H,\beta,t)-reachable in GG. Note that the corresponding HH-reachable sets for u,vu,v may not be included in UU. We say UU is (H,β,t)(H,\beta,t)-inner-closed if UU is (H,β,t)(H,\beta,t)-closed and additionally the corresponding HH-reachable sets for every pair u,vu,v also lie inside UU.

The following result from [19] builds a sufficient condition to ensure that every hh-subset SS has linearly many vertex-disjoint absorbers.

Lemma 4.2 ([19]).

Given h,th,t\in\mathbb{N} with h3h\geq 3 and β>0\beta>0, the following holds for any hh-vertex graph HH and sufficiently large nn\in\mathbb{N}. Let GG be an nn-vertex graph such that V(G)V(G) is (H,β,t)(H,\beta,t)-closed. Then every S(V(G)h)S\in\tbinom{V(G)}{h} has a family of at least βh3tn\frac{\beta}{h^{3}t}n vertex-disjoint (S,t)(S,t)-absorbers.

Based on this lemma, it suffices to show that V(G)V(G) is closed. However, we shall show a slightly weaker result, namely, there exists a small vertex set BB such that the induced subgraph GBG-B is inner-closed. The proof of Lemma 4.3 can be found in Section 4.2.

Lemma 4.3.

Given r,r,\ell\in\mathbb{N} with r>2r>\ell\geq 2 and τ,μ\tau,\mu with 0<τ<μ0<\tau<\mu, there exist α,β>0\alpha,\beta>0 such that the following holds for sufficiently large nn\in\mathbb{N}. Let GG be an nn-vertex graph with

δ(G)max{rr+μ,12ϱ(r1,f)+μ}nandα(G)<f(αn).\delta(G)\geq\max\left\{\frac{r-\ell}{r}+\mu,\frac{1}{2-\varrho^{*}_{\ell}(r-1,f)}+\mu\right\}n\leavevmode\nobreak\ \text{and}\leavevmode\nobreak\ \alpha_{\ell}(G)<f(\alpha n).

Then GG admits a partition V(G)=BUV(G)=B\cup U such that |B|τn|B|\leq\tau n and UU is (Kr,β,4)(K_{r},\beta,4)-inner-closed.

Then by Lemma 4.2 applied on G[U]G[U], we can easily get the following corollary.

Corollary 4.4.

Given positive integers r,r,\ell with r>2r>\ell\geq 2 and τ,μ\tau,\mu with 0<τ<μ0<\tau<\mu, there exist 0<α<β<μ30<\alpha<\beta<\mu^{3} such that the following holds for sufficiently large nn\in\mathbb{N}. Let GG be an nn-vertex graph with δ(G)(12ϱ(r1,f)+μ)n\delta(G)\geq\left(\frac{1}{2-\varrho^{*}_{\ell}(r-1,f)}+\mu\right)n and α(G)<f(αn)\alpha_{\ell}(G)<f(\alpha n). Then GG admits a partition V(G)=BUV(G)=B\cup U such that |B|τn|B|\leq\tau n and every S(Ur)S\in\tbinom{U}{r} has a family of at least β4r3n\frac{\beta}{4r^{3}}n vertex-disjoint (S,4)(S,4)-absorbers in UU.

To deal with the exceptional vertex set BB, we shall pick mutually vertex-disjoint copies of KrK_{r} each containing a vertex in BB. To achieve this, one has to make sure that every vertex vV(G)v\in V(G) is covered by many copies of KrK_{r} in GG (the aforementioned cover threshold). The following result enables us to find linearly many copies of KrK_{r} covering any given vertex.

Proposition 4.5.

Given r,r,\ell\in\mathbb{N} and a constant μ>0\mu>0, there exists α>0\alpha>0 such that for all sufficiently large nn the following holds. Let GG be an nn-vertex graph with δ(G)(12ϱ(r1,f)+μ)n\delta(G)\geq\left(\frac{1}{2-\varrho^{*}_{\ell}(r-1,f)}+\mu\right)n and α(G)f(αn)\alpha_{\ell}(G)\leq f(\alpha n). If WW is a subset of V(G)V(G) with |W|μ2n|W|\leq\frac{\mu}{2}n, then for each vertex uV(G)Wu\in V(G)\setminus W, G[V(G)W]G[V(G)\setminus W] contains a copy of KrK_{r} covering uu.

Proof.

We choose 1nαμ,1\frac{1}{n}\ll\alpha\ll\mu,\frac{1}{\ell} and let G1:=G[V(G)W]G_{1}:=G[V(G)\setminus W]. It suffices to show that for each vertex uV(G1)u\in V(G_{1}), there is a copy of Kr1K_{r-1} in NG1(u)N_{G_{1}}(u). Note that for every vertex uu in G1G_{1}, we have |NG1(u)|δ(G1)δ(G)|W|(12ϱ(r1,f)+μ2)n|N_{G_{1}}(u)|\geq\delta(G_{1})\geq\delta(G)-|W|\geq\left(\frac{1}{2-\varrho^{*}_{\ell}(r-1,f)}+\frac{\mu}{2}\right)n. Given any vertex vNG1(u)v\in N_{G_{1}}(u) with dG1(u,v):=|NG1(u)NG1(v)|d_{G_{1}}(u,v):=|N_{G_{1}}(u)\cap N_{G_{1}}(v)|, we have

dG1(u,v)(ϱ(r1,f)+μ4)dG1(u)dG1(u)+dG1(v)n(ϱ(r1,f)+μ4)dG1(u)(2ϱ(r1,f)μ4)δ(G1)n>μ8n>0.\begin{split}d_{G_{1}}(u,v)-\left(\varrho^{*}_{\ell}(r-1,f)+\tfrac{\mu}{4}\right)d_{G_{1}}(u)&\geq d_{G_{1}}(u)+d_{G_{1}}(v)-n-\left(\varrho^{*}_{\ell}(r-1,f)+\tfrac{\mu}{4}\right)d_{G_{1}}(u)\\ &\geq\left(2-\varrho^{*}_{\ell}(r-1,f)-\tfrac{\mu}{4}\right)\delta(G_{1})-n>\tfrac{\mu}{8}n>0.\end{split}

Thus δ(G[NG1(u)])>(ϱ(r1,f)+μ4)|NG1(u)|\delta(G[N_{G_{1}}(u)])>(\varrho^{*}_{\ell}(r-1,f)+\frac{\mu}{4})|N_{G_{1}}(u)|. Therefore by the definition of ϱ(r1,f)\varrho^{*}_{\ell}(r-1,f) and the choice that 1nαμ\frac{1}{n}\ll\alpha\ll\mu, G[NG1(u)]G[N_{G_{1}}(u)] contains a copy of Kr1K_{r-1}, which together with uu yields a copy of KrK_{r} in G1G_{1}. ∎

Now we are ready to prove Lemma 2.2 using Corollary 4.4 and Proposition 4.5.

Proof of Lemma 2.2.

Given positive integers ,r\ell,r with r>2r>\ell\geq 2 and μ,γ\mu,\gamma with 0<γμ20<\gamma\leq\frac{\mu}{2}, we choose 1nαξβ,τγ,1r\frac{1}{n}\ll\alpha\ll\xi\ll\beta,\tau\ll\gamma,\frac{1}{r}. Let GG be an nn-vertex graph with δ(G)(12ϱ(r1,f)+μ)n\delta(G)\geq\left(\frac{1}{2-\varrho^{*}_{\ell}(r-1,f)}+\mu\right)n, α(G)<f(αn)\alpha_{\ell}(G)<f(\alpha n) and nrn\in r\mathbb{N}. Then Corollary 4.4 implies that GG admits a partition V(G)=BUV(G)=B\cup U such that |B|τn|B|\leq\tau n and every S(Ur)S\in\tbinom{U}{r} has a family of at least β4r3n\frac{\beta}{4r^{3}}n vertex-disjoint (S,4)(S,4)-absorbers in UU. Let G1:=G[U]G_{1}:=G[U]. Then by applying Lemma 4.1 on G1G_{1}, we obtain in G1G_{1} a ξ\xi-absorbing subset A1A_{1} of size at most β4r3n\frac{\beta}{4r^{3}}n.

Now, we shall iteratively pick vertex-disjoint copies of KrK_{r} each covering at least a vertex in BB whilst avoiding using any vertex in A1A_{1}, and we claim that every vertex in BB can be covered in this way.

Let G2:=GA1G_{2}:=G-A_{1}. For uBu\in B, we apply Proposition 4.5 iteratively to find a copy of KrK_{r} covering uu in G2G_{2}, while avoiding A1BA_{1}\cup B and all copies of KrK_{r} found so far. Because of the fact that β,τγ,1r\beta,\tau\ll\gamma,\frac{1}{r}, this is possible as during the process, the number of vertices that we need to avoid is at most |A1|+r|B|β4r3n+rτnμ2n|A_{1}|+r|B|\leq\frac{\beta}{4r^{3}}n+r\tau n\leq\frac{\mu}{2}n. Let KK be the union of the vertex sets over all copies of KrK_{r} covering BB and A:=A1KA:=A_{1}\cup K. Recall that A1A_{1} is a ξ\xi-absorbing set for G1=GBG_{1}=G-B, and BKAB\subseteq K\subseteq A. Then it is easy to check that AA is a ξ\xi-absorbing set for GG and

|A|=|A1|+|K|β4r3n+rτnγn,|A|=|A_{1}|+|K|\leq\tfrac{\beta}{4r^{3}}n+r\tau n\leq\gamma n,

where the last inequality follows since β,τγ,1r\beta,\tau\ll\gamma,\frac{1}{r}. ∎

Now it remains to prove Lemma 4.3, which is done in the next subsection.

4.2 Proof of Lemma 4.3

The proof of Lemma 4.3 makes use of Szemerédi’s Regularity Lemma and a result in [8].

Lemma 4.6.

[8, Lemma 5.1] Given n,r,n,r,\ell\in\mathbb{N} with r>2r>\ell\geq 2 and a monotone increasing function f(n)f(n), for all τ,μ\tau,\mu with 0<τ<μ0<\tau<\mu, there exist positive constants β1,γ1\beta_{1},\gamma_{1} and α>0\alpha>0 such that the following holds for sufficiently large nn\in\mathbb{N}. Let GG be an nn-vertex graph with δ(G)(12ϱ(r1,f)+μ)n\delta(G)\geq\left(\frac{1}{2-\varrho^{*}_{\ell}(r-1,f)}+\mu\right)n and α(G)f(αn)\alpha_{\ell}(G)\leq f(\alpha n). Then GG admits a partition V(G)=BUV(G)=B\cup U such that |B|τn|B|\leq\tau n and every vertex in UU is (Kr,β1,1)(K_{r},\beta_{1},1)-reachable to at least γ1n\gamma_{1}n other vertices in UU with all the corresponding KrK_{r}-reachable sets belonging to UU.

Proof of Lemma 4.3.

Given r,r,\ell\in\mathbb{N} with r>2r>\ell\geq 2 and τ,μ\tau,\mu with 0<τ<μ0<\tau<\mu, we choose constants

1nαβεβ1,γ1τ,μ,1\tfrac{1}{n}\ll\alpha\ll\beta\ll\varepsilon\ll\beta_{1},\gamma_{1}\ll\tau,\mu,\tfrac{1}{\ell}

and let GG be an nn-vertex graph with δ(G)max{rr+μ,12ϱ(r1,f)+μ}n\delta(G)\geq\max\{\frac{r-\ell}{r}+\mu,\frac{1}{2-\varrho^{*}_{\ell}(r-1,f)}+\mu\}n and α(G)<f(αn)\alpha_{\ell}(G)<f(\alpha n). Then by applying Lemma 4.6, we obtain a partition V(G)=BUV(G)=B\cup U such that |B|τn|B|\leq\tau n and every vertex in UU is (Kr,β1,1)(K_{r},\beta_{1},1)-reachable (inG[U])({\rm in}\ G[U]) to at least γ1n\gamma_{1}n other vertices in UU. For any two vertices u,vUu,v\in U, we shall prove in G[U]G[U] that u,vu,v are (Kr,β,4)(K_{r},\beta,4)-reachable.

Let XX and YY be the sets of vertices that are (Kr,β1,1)(K_{r},\beta_{1},1)-reachable to uu and vv, respectively. By taking subsets from them and renaming if necessary, we may further assume that XY=X\cap Y=\emptyset and |X|=|Y|=γ1n2|X|=|Y|=\frac{\gamma_{1}n}{2}. Then by applying Lemma 3.3 on GG with positive constants εβ1,γ1\varepsilon\ll\beta_{1},\gamma_{1} and d:=μ4d:=\frac{\mu}{4}, we obtain a refinement 𝒫:={V0,Vi,,Vk}\mathcal{P}:=\{V_{0},V_{i},\ldots,V_{k}\} of the original partition {X,Y,V(G)XY}\{X,Y,V(G)-X-Y\} and a spanning subgraph GGG^{\prime}\subseteq G with properties (a)-(e), where we let m:=|Vi|m:=|V_{i}| for all i[k]i\in[k] and RdR_{d} be the corresponding dd-reduced graph. Without loss of generality, we may assume that V1XV_{1}\subseteq X and V2YV_{2}\subseteq Y. Note that by Fact 3.4, we can observe that δ(Rd)max{rr+μ2,12ϱ(r1,f)+μ2}k(12+μ2)k\delta(R_{d})\geq\max\{\frac{r-\ell}{r}+\frac{\mu}{2},\frac{1}{2-\varrho^{*}_{\ell}(r-1,f)}+\frac{\mu}{2}\}k\geq(\frac{1}{2}+\frac{\mu}{2})k. Let V3V_{3} be a common neighbor of V1V_{1} and V2V_{2} in RdR_{d}.

Now we shall show that uu and vv are (Kr,β,4)(K_{r},\beta,4)-reachable. We write r=x+yr=\ell x+y for some integers x,yx,y with x>0,0y1x>0,0\leq y\leq\ell-1. Note that δ(Rd)(rr+μ2)k(x1x+μ2)k\delta(R_{d})\geq(\frac{r-\ell}{r}+\frac{\mu}{2})k\geq(\frac{x-1}{x}+\frac{\mu}{2})k. Thus every xx vertices in RdR_{d} have at least μ2xk\frac{\mu}{2}xk common neighbors and we can greedily pick two copies of Kx+1K_{x+1} in RdR_{d} that contain the edge V1V3V_{1}V_{3} and V2V3V_{2}V_{3} respectively and overlap only on the vertex V3V_{3}. We use 𝒜={V1,V3,Va1,Va2,,Vax1}\mathcal{A}=\{V_{1},V_{3},V_{a_{1}},V_{a_{2}},\ldots,V_{a_{x-1}}\} and 𝒯={V2,V3,Vb1,Vb2,,Vbx1}\mathcal{T}=\{V_{2},V_{3},V_{b_{1}},V_{b_{2}},\ldots,V_{b_{x-1}}\} to denote the two family of clusters related to the two copies of Kx+1K_{x+1} in RdR_{d}. Applying Lemma 3.6 on GG with η=12mn,q=x+1,p=\eta=\frac{1}{2}\frac{m}{n},q=x+1,p=\ell and V1,V3,Va1,,Vax1V_{1},V_{3},V_{a_{1}},\ldots,V_{a_{x-1}} playing the role of V1,,VqV_{1},\ldots,V_{q}, we can iteratively take m2\frac{m}{2\ell} vertex-disjoint copies of Kr+1K_{r+1} (since r+1pqr+1\leq pq) which are denoted by S1,S2,,Sm2S_{1},S_{2},\ldots,S_{\frac{m}{2\ell}}, such that each SiS_{i} has exactly y+1y+1 vertices in V3V_{3}, \ell vertices in V1V_{1} and VaiV_{a_{i}}, i[x1]i\in[x-1]. Let V3V3V_{3}^{\prime}\subseteq V_{3} be a subset obtained by taking exactly one vertex from each such SiS_{i}. Then |V3|=m2|V_{3}^{\prime}|=\frac{m}{2\ell} and again by applying Lemma 3.6 on GG with η=14mn,q=x+1,p=\eta=\frac{1}{4\ell}\frac{m}{n},q=x+1,p=\ell and V2,V3,Vb1,,Vbx1V_{2},V_{3}^{\prime},V_{b_{1}},\ldots,V_{b_{x-1}} playing the role of V1,,VqV_{1},\ldots,V_{q}, we can greedily pick m42\frac{m}{4\ell^{2}} vertex-disjoint copies of Kr+1K_{r+1}, denoted by T1,T2,,Tm42T_{1},T_{2},\ldots,T_{\frac{m}{4\ell^{2}}}, such that each TiT_{i} has exactly y+1y+1 vertices in V3V_{3}^{\prime}, \ell vertices in V2V_{2} and VbiV_{b_{i}}, i[x1]i\in[x-1].

Refer to caption
Figure 1: SiS_{i} and TiT_{i}: here we take =3,y=1\ell=3,y=1 for instance.

Now it remains to show the following statement with βnm42\beta n\leq\frac{m}{4\ell^{2}}. Recall that mεnm\leq\varepsilon n.

Claim 4.7.

There exist βn\beta n vertex-disjoint KrK_{r}-reachable sets for u,vu,v, each of size 4r14r-1.

Refer to caption
Figure 2: Constructions of KrK_{r}-reachable sets EiE_{i}.
Proof of claim.

Here the main idea is to extend all such TiT_{i}’s to pairwise vertex-disjoint KrK_{r}-reachable sets. Note that for each TiT_{i}, there exists SiS_{i^{\prime}} such that the two copies of Kr+1K_{r+1} intersect on exactly one vertex in V3V_{3}^{\prime}, denoted by wiw_{i}. Let uiu_{i} be an arbitray vertex chosen from SiS_{i^{\prime}} that lies in V1V_{1}, and viv_{i} be chosen from TiT_{i} that lies in V2V_{2}. Then by the assumption that uu is (Kr,β1,1)(K_{r},\beta_{1},1)-reachable to u1u_{1}, there exist at least β1n\beta_{1}n vertex-disjoint KrK_{r}-reachable sets for uu and uiu_{i} (resp. vv and viv_{i}). Therefore by the fact that εβ1\varepsilon\ll\beta_{1}, we can greedily choose two vertex-disjoint KrK_{r}-reachable sets, say CiC_{i} and DiD_{i}, for u,uiu,u_{i} and v,viv,v_{i}, respectively, which are also disjoint from all cliques SiS_{i} or TjT_{j} for i[m2],j[m42]i\in[\frac{m}{2\ell}],j\in[\frac{m}{4\ell^{2}}]. It is easy to check that the set

Ei:=V(Si)V(Ti)CiDiE_{i}:=V(S_{i^{\prime}})\cup V(T_{i})\cup C_{i}\cup D_{i}

has size 4r14r-1 and G[Ei{u}]G[E_{i}\cup\{u\}] (similarly for Ei{v}E_{i}\cup\{v\}) contains 44 copies of KrK_{r}, which are induced on the sets {u}Ci,V(Si){wi},V(Ti){vi}\{u\}\cup C_{i},V(S_{i^{\prime}})-\{w_{i}\},V(T_{i})-\{v_{i}\} and {vi}Di\{v_{i}\}\cup D_{i}, respectively. Thus by definition EiE_{i} is a KrK_{r}-reachable set for uu and vv. A desired number of mutually disjoint KrK_{r}-reachable set EiE_{i} can be chosen by extending each TiT_{i} as above. ∎

5 A construction

In this section, we shall use a construction of K+1K_{\ell+1}-free graphs GG with small α(G)\alpha_{\ell}(G) to prove Proposition 1.9. An explicit construction was firstly obtained by Erdős and Rogers [13] in the setting that α(G)=o(n)\alpha_{\ell}(G)=o(n). Here we give a probabilistic construction as follows, whose proof is very similar to that of a result of Nenadov and Pehova (see Proposition 4.1 in [35]).

Lemma 5.1.

For any \ell\in\mathbb{N} with 2\ell\geq 2, a constant γ(0,12+2)\gamma\in(0,\frac{\ell-1}{\ell^{2}+2\ell}) and sufficiently large integer nn, there exists an nn-vertex K+1K_{\ell+1}-free graph GG_{\ell} such that α(G)n1γ\alpha_{\ell}(G_{\ell})\leq n^{1-\gamma}.

Now we firstly give a short proof of Proposition 1.9, and then present the proof of Lemma 5.1 at the end of this section.

Proof of Proposition 1.9.

Fix r>2r>\ell\geq 2 and constants η,γ\eta,\gamma as in the statement. Let nn be sufficiently large and define μ=rr(rrη)>0\mu=\tfrac{r}{r-\ell}(\tfrac{r-\ell}{r}-\eta)>0. Then by Lemma 5.1, we choose GG_{\ell} to be a (1η)n(1-\eta)n-vertex K+1K_{\ell+1}-free graph with α(G)|G|1γ\alpha_{\ell}(G_{\ell})\leq|G_{\ell}|^{1-\gamma}.

Let GG be an nn-vertex graph with vertex partition V(G)=X1X2V(G)=X_{1}\cup X_{2} such that

  1. (i)

    G[X1]G[X_{1}] is a clique with |X1|=ηn|X_{1}|=\eta n;

  2. (ii)

    G[X1,X2]G[X_{1},X_{2}] is a complete bipartite graph;

  3. (iii)

    G[X2]G[X_{2}] induces a copy of GG_{\ell}.

Now we claim that GG has the desired properties. Indeed it is easy to see that δ(G)ηn\delta(G)\geq\eta n and α(G)α(G)<n1γ\alpha_{\ell}(G)\leq\alpha_{\ell}(G_{\ell})<n^{1-\gamma}. Since G[X2]G[X_{2}] is K+1K_{\ell+1}-free, every copy of KrK_{r} must intersect X1X_{1} on at least rr-\ell vertices. Thus every KrK_{r}-tiling in GG contains at most |X1|r=ηrn\tfrac{|X_{1}|}{r-\ell}=\tfrac{\eta}{r-\ell}n vertex-disjoint copies of KrK_{r}, which together cover at most rηrn=(1μ)n\tfrac{r\eta}{r-\ell}n=(1-\mu)n vertices. ∎

Proof of Lemma 5.1.

We only consider 3\ell\geq 3 as the case =2\ell=2 follows from a celebrated result on Ramsey number that R(3,n)=Θ(n2/logn)R(3,n)=\Theta(n^{2}/\log n) [25]. We choose 1nγ,1\frac{1}{n}\ll\gamma,\frac{1}{\ell} and let x=2γ+1x=\tfrac{2-\gamma}{\ell+1}. Considering the random graph G=G(n,p)G=G(n,p) with p=nxp=n^{-x}, we shall verify that with positive probability, GG is K+1K_{\ell+1}-free and α(G)n1γ\alpha_{\ell}(G)\leq n^{1-\gamma}. By applying the FKG inequality [15], we have that

[G is K+1-free]\displaystyle\mathbb{P}[\text{$G$ is $K_{\ell+1}$-free}] S(V(G)+1)[G[S]K+1]\displaystyle\geq\prod_{S\in\binom{V(G)}{\ell+1}}\mathbb{P}[G[S]\neq K_{\ell+1}]
(1p(+12))(n+1)exp(2p(+12)n+1)=exp(2n1+γ2),\displaystyle\geq\left(1-p^{\binom{\ell+1}{2}}\right)^{\binom{n}{\ell+1}}\geq\exp\left(-2p^{\binom{\ell+1}{2}}n^{\ell+1}\right)=\exp\left(-2n^{1+\frac{\gamma}{2}\ell}\right),

where we bound 1xe2x1-x\geq e^{-2x} for x(0,12)x\in(0,\frac{1}{2}). Now it remains to determine the probability of the event that α(G)n1γ\alpha_{\ell}(G)\leq n^{1-\gamma}.

Let II be the random variable counting all sets AA such that |A|=n1γ|A|=n^{1-\gamma} and G[A]G[A] is KK_{\ell}-free. Then

𝔼(I)=|A|=n1γ[G[A] is K-free].\mathbb{E}(I)=\sum_{|A|=n^{1-\gamma}}\mathbb{P}[\text{$G[A]$ is $K_{\ell}$-free}].

Here we shall use a powerful inequality of Janson [21], where for each \ell-set SAS\subseteq A we denote by XSX_{S} the indicator variable for the event that G[S]=KG[S]=K_{\ell}. Let X=S(A)XSX=\sum_{S\in\binom{A}{\ell}}X_{S}. Then by Janson’s inequality, we obtain that

[G[A] is K-free]=[X=0]exp(𝔼(X)+Δ2),\mathbb{P}[\text{$G[A]$ is $K_{\ell}$-free}]=\mathbb{P}[X=0]\leq\exp\left(-\mathbb{E}(X)+\frac{\Delta}{2}\right),

where 𝔼(X)=(|A|)p(2)=Θ(n(1γ)2γ+1(2))\mathbb{E}(X)=\binom{|A|}{\ell}p^{\binom{\ell}{2}}=\Theta\left(n^{(1-\gamma)\ell-\frac{2-\gamma}{\ell+1}\binom{\ell}{2}}\right) and Δ=SS,|SS|2[XS=1,XS=1]\Delta=\sum\limits_{S\neq S^{\prime},\leavevmode\nobreak\ |S\cap S^{\prime}|\geq 2}\mathbb{P}[X_{S}=1,X_{S^{\prime}}=1]. Note that

Δ\displaystyle\Delta =SS,|SS|2[XS=1,XS=1]\displaystyle=\sum\limits_{S\neq S^{\prime},\leavevmode\nobreak\ |S\cap S^{\prime}|\geq 2}\mathbb{P}[X_{S}=1,X_{S^{\prime}}=1]
(|A|)p(2)2s1(s)(|A|s)p(2)(s2)\displaystyle\leq\binom{|A|}{\ell}p^{\binom{\ell}{2}}\sum_{2\leq s\leq\ell-1}\binom{\ell}{s}\binom{|A|-\ell}{\ell-s}p^{\binom{\ell}{2}-\binom{s}{2}}
𝔼(X)2s1(s)n(s)(1γx+s12)\displaystyle\leq\mathbb{E}(X)\sum_{2\leq s\leq\ell-1}\binom{\ell}{s}n^{(\ell-s)(1-\gamma-x\frac{\ell+s-1}{2})}
=o(𝔼(X)),\displaystyle=o(\mathbb{E}(X)),

where the last equality follows because x=2γ+1x=\frac{2-\gamma}{\ell+1} and thus 1γx+s12γ21-\gamma-x\frac{\ell+s-1}{2}\leq-\frac{\gamma}{2} holds for any s2s\geq 2. Therefore 𝔼(I)2nexp(Θ(n(1γ)2γ+1(2)))\mathbb{E}(I)\leq 2^{n}\exp\left(-\Theta\left(n^{(1-\gamma)\ell-\frac{2-\gamma}{\ell+1}\binom{\ell}{2}}\right)\right) and by Markov’s inequality, with probability at least 12nexp(Θ(n(1γ)2γ+1(2)))1-2^{n}\exp\left(-\Theta\left(n^{(1-\gamma)\ell-\frac{2-\gamma}{\ell+1}\binom{\ell}{2}}\right)\right), we have I=0I=0, that is, α(G)n1γ\alpha_{\ell}(G)\leq n^{1-\gamma}.

By the inclusive-exclusive principle, the probability of the event that GG is K+1K_{\ell+1}-free and α(G)n1γ\alpha_{\ell}(G)\leq n^{1-\gamma} is at least

exp(2n1+γ2)2nexp(Θ(n(1γ)2γ+1(2)))\exp\left(-2n^{1+\frac{\gamma}{2}\ell}\right)-2^{n}\exp\left(-\Theta\left(n^{(1-\gamma)\ell-\frac{2-\gamma}{\ell+1}\binom{\ell}{2}}\right)\right)

and it is positive for sufficiently large nn as long as

1+γ2<(1γ)2γ+1(2),1+\frac{\gamma}{2}\ell<(1-\gamma)\ell-\frac{2-\gamma}{\ell+1}\binom{\ell}{2},

which follows easily as γ<12+2\gamma<\frac{\ell-1}{\ell^{2}+2\ell}. ∎

6 Concluding remarks

In this paper we study the minimum degree condition for KrK_{r}-factors in graphs with sublinear \ell-independence number. Our result is asymptotically sharp when α(G)(n1γ,n1ω(n)logλn)\alpha_{\ell}(G)\in(n^{1-\gamma},n^{1-\omega(n)\log^{-\lambda}n}) for any constant 0<γ<12+20<\gamma<\frac{\ell-1}{\ell^{2}+2\ell}.

This leads to the following question: What is the general behavior of the minimum degree condition forcing a clique factor when the condition of \ell-independence number is imposed within the range (1,n)(1,n)? We formulate this as follows. Given integers n>r>2n>r>\ell\geq 2 with nrn\in r\mathbb{N}, a constant α>0\alpha>0 and a monotone increasing function g(n)[n]g(n)\in[n], we denote by RTT(n,Kr,g(αn))\textbf{RTT}_{\ell}(n,K_{r},g(\alpha n)) the maximum integer δ\delta such that there exists an nn-vertex graph GG with δ(G)δ\delta(G)\geq\delta and α(G)g(αn)\alpha_{\ell}(G)\leq g(\alpha n) which does not contain a KrK_{r}-factor. Here we try to understand when and how the value RTT(n,Kr,g(αn))\textbf{RTT}_{\ell}(n,K_{r},g(\alpha n)) changes sharply when the magnitude of g(n)g(n) varies. This can be seen as a degree version of the well-known phase transition problem for RT2(n,Kr,g(n))\textbf{RT}_{2}(n,K_{r},g(n)) in Ramsey–Turán theory (see [5, 3, 24]). It is worth noting that many open questions on the phase transition problem of RT2(n,Kr,g(n))\textbf{RT}_{2}(n,K_{r},g(n)) are essentially related to Ramsey theory.

Here we consider the basic case RTT2(n,Kr,g(n))\textbf{RTT}_{2}(n,K_{r},g(n)). Recall that Knierim and Su [26] resolved Problem 1.1 for r4r\geq 4 by giving an asymptotically tight minimum degree bound (12r)n+o(n)(1-\frac{2}{r})n+o(n). In our context of g(n)=ng(n)=n, this can be roughly reformulated as

RTT2(n,Kr,o(n))=r2rn+o(n)forr4.\textbf{RTT}_{2}(n,K_{r},o(n))=\frac{r-2}{r}n+o(n)\leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ r\geq 4.

Also, for integers r,r,\ell with r>34rr>\ell\geq\frac{3}{4}r, Theorem 1.4 can be stated as

RTT(n,Kr,o(n))=12ϱ(r1)n+o(n).\textbf{RTT}_{\ell}(n,K_{r},o(n))=\frac{1}{2-\varrho_{\ell}(r-1)}n+o(n).

In this paper, our main theorem combined with Proposition 1.9 and the cover threshold implies that for r>2,γ(0,12+2)r>\ell\geq 2,\gamma\in(0,\frac{\ell-1}{\ell^{2}+2\ell}) and n1γf(n)n1ω(n)logλnn^{1-\gamma}\leq f(n)\leq n^{1-\omega(n)\log^{-\lambda}n},

RTT(n,Kr,f(o(n)))=max{rrn,12ϱ(r1,f)n}+o(n).\textbf{RTT}_{\ell}(n,K_{r},f(o(n)))=\max\left\{\frac{r-\ell}{r}n,\frac{1}{2-\varrho^{*}_{\ell}(r-1,f)}n\right\}+o(n).

This provides an insight into the general behavior of RTT(n,Kr,f(n))\textbf{RTT}_{\ell}(n,K_{r},f(n)) but the asymptotic behavior of RTT(n,Kr,g(n))\textbf{RTT}_{\ell}(n,K_{r},g(n)) for a general g(n)g(n) seems to be out of reach. It will be interesting to study the case g(n)=ncg(n)=n^{c} for any constant c(0,112+2)c\in(0,1-\tfrac{\ell-1}{\ell^{2}+2\ell}).

References

  • [1] Noga Alon, Michael Krivelevich, and Benny Sudakov. Turán numbers of bipartite graphs and related Ramsey-type questions. volume 12, pages 477–494. 2003. Special issue on Ramsey theory.
  • [2] Noga Alon and Raphael Yuster. HH-factors in dense graphs. Journal of Combinatorial Theory, Series B, 66(2):269–282, 1996.
  • [3] J. Balogh and J. Lenz. Some Exact Ramsey–Turán Numbers. Bulletin of the London Mathematical Society, 44:1251–1258, 2012.
  • [4] J. Balogh and J. Lenz. On Ramsey–Turán numbers of graphs and hypergraphs. Israel Journal of Mathematics, 194:45–68, 2013.
  • [5] József Balogh, Ping Hu, and Miklós Simonovits. Phase transitions in Ramsey–Turán theory. Journal of Combinatorial Theory, Series B, 114:148–169, 2015.
  • [6] József Balogh, Theodore Molla, and Maryam Sharifzadeh. Triangle factors of graphs without large independent sets and of weighted graphs. Random Structures and Algorithms, 49(4):669–693, 2016.
  • [7] B. Bollobás and P. Erdős. On a Ramsey–Turán type problem. Journal of Combinatorial Theory, Series B, 21(2):166–168, 1976.
  • [8] Fan Chang, Jie Han, Jaehoon Kim, Guanghui Wang, and Donglei Yang. Embedding clique-factors in graphs with low \ell-independence number. arXiv preprint arXiv:2111.10512, 2021.
  • [9] Yulin Chang, Jie Han, Yoshiharu Kohayakawa, Patrick Morris, and Guilherme Oliveira Mota. Factors in randomly perturbed hypergraphs. Random Structures and Algorithms, 60(2):153–165, 2022.
  • [10] David Conlon, Jacob Fox, and Benny Sudakov. Ramsey numbers of sparse hypergraphs. Random Structures and Algorithms, 35(1):1–14, 2009.
  • [11] R. Diestel. Graph Theory, volume 173 of. Graduate texts in mathematics, 2017.
  • [12] Laihao Ding, Jie Han, Shumin Sun, Guanghui Wang, and Wenling Zhou. FF-factors in quasi-random hypergraphs. Journal of the London Mathematical Society, 106(3):1810–1843, 2022.
  • [13] P. Erdős and C. A. Rogers. The construction of certain graphs. Canadian Journal of Mathematics. Journal Canadien de Mathématiques, 14:702–707, 1962.
  • [14] P. Erdős and Vera T. Sós. Some remarks on Ramsey’s and Turán’s theorem. In Combinatorial theory and its applications, II (Proc. Colloq., Balatonfüred, 1969), pages 395–404, 1970.
  • [15] C. M. Fortuin, P. W. Kasteleyn, and J. Ginibre. Correlation inequalities on some partially ordered sets. Communications in Mathematical Physics, 22:89–103, 1971.
  • [16] Jacob Fox, Po-Shen Loh, and Yufei Zhao. The critical window for the classical Ramsey-Turán problem. Combinatorica, 35(4):435–476, 2015.
  • [17] Jacob Fox and Benny Sudakov. Dependent random choice. Random Structures and Algorithms, 38(1-2):68–99, 2011.
  • [18] András Hajnal and Endre Szemerédi. Proof of a conjecture of P. Erdős. Combinatorial theory and its applications, 2(4):601–623, 1970.
  • [19] J. Han, Patrick Morris, Guanghui Wang, and Donglei Yang. A Ramsey–Turán theory for tilings in graphs. arXiv preprint arXiv:2106.09688, 2021.
  • [20] J. Han, C. Zang, and Y. Zhao. Minimum vertex degree thresholds for tiling complete 3-partite 3-graphs. Journal of Combinatorial Theory, Series A, 149:115–147, 2017.
  • [21] Svante Janson, Tomasz Łuczak, and Andrzej Ruciński. An exponential bound for the probability of nonexistence of a specified subgraph in a random graph. In Random graphs ’87 (Poznań, 1987), pages 73–87. Wiley, Chichester, 1990.
  • [22] Peter Keevash and Richard Mycroft. A multipartite Hajnal–Szemerédi theorem. Journal of Combinatorial Theory, Series B, 114:187–236, 2015.
  • [23] H. A. Kierstead and A. V. Kostochka. An Ore-type theorem on equitable coloring. Journal of Combinatorial Theory, Series B, 98(1):226–234, 2008.
  • [24] Jaehoon Kim, Younjin Kim, and Hong Liu. Two conjectures in Ramsey-Turán theory. SIAM Journal on Discrete Mathematics, 33(1):564–586, 2019.
  • [25] Jeong Han Kim. The Ramsey number R(3,t)R(3,t) has order of magnitude t2/logtt^{2}/\log t. Random Structures and Algorithms, 7(3):173–207, 1995.
  • [26] Charlotte Knierim and Pascal Su. KrK_{r}-factors in graphs with low independence number. Journal of Combinatorial Theory, Series B, 148:60–83, 2021.
  • [27] J. Komlós and M. Simonovits. Szemerédi’s Regularity Lemma and its applications in graph theory. Combinatorics, Paul Erdős is eighty, Vol. 2 (Keszthely, 1993), 1996.
  • [28] János Komlós. Tiling Turán theorems. Combinatorica, 20(2):203–218, 2000.
  • [29] M. Krivelevich. Triangle factors in random graphs. Combinatorics, Probability and Computing, 6:337–347, 1997.
  • [30] D Kühn and D. Osthus. The minimum degree threshold for perfect graph packings. Combinatorica, 29(1):65–107, 2009.
  • [31] Hong Liu, Christian Reiher, Maryam Sharifzadeh, and Katherine Staden. Geometric constructions for Ramsey–Turán theory. arXiv preprint arXiv:2103.10423, 2021.
  • [32] Allan Lo and Klas Markström. FF-factors in hypergraphs via absorption. Graphs and Combinatorics, 31(3):679–712, 2015.
  • [33] Clara Marie Lüders and Christian Reiher. The Ramsey-Turán problem for cliques. Israel Journal of Mathematics, 230(2):613–652, 2019.
  • [34] Richard Montgomery. Spanning trees in random graphs. Advances in Mathematics, 356:106–793, 2019.
  • [35] R. Nenadov and Y. Pehova. On a Ramsey–Turán variant of the Hajnal–Szemerédi theorem. SIAM Journal on Discrete Mathematics, 34(2):1001–1010, 2020.
  • [36] Vojtech Rödl, Andrzej Ruciński, and Endre Szemerédi. Perfect matchings in large uniform hypergraphs with large minimum collective degree. Journal of Combinatorial Theory, Series A, 116(3):613–636, 2009.
  • [37] M. Simonovits and V. T. Sós. Ramsey–Turán theory. Discrete Mathematics, 229(1-3):293–340, 2001.
  • [38] E. Szemerédi. On graphs containing no complete subgraphs with 4 vertices. Matematikai Lapok. Bolyai János Matematikai Társulat, 23:113–116, 1972.
  • [39] A. Treglown. On directed versions of the Hajnal–Szemerédi theorem. Combinatorics, Probability and Computing, 24:873–928, 2015.

Appendix A Proof of Lemma 3.6

We use the method of dependent random choice to prove Lemma 3.6. The method was developed by Füredi, Gowers, Kostochka, Rödl, Sudakov, and possibly many others. The next lemma is taken from Alon, Krivelevich and Sudakov [1]. Interested readers may check the survey paper on this method by Fox and Sudakov [17].

Lemma A.1 (Dependent Random Choice).

[1] Let a,d,m,n,ra,d,m,n,r be positive integers. Let G=(V,E)G=(V,E) be a graph with nn vertices and average degree d=2e(G)/nd=2e(G)/n. If there is a positive integer tt such that

dtnt1(nr)(mn)ta,\frac{d^{t}}{n^{t-1}}-\binom{n}{r}\left(\frac{m}{n}\right)^{t}\geq a, (1)

then GG contains a subset UU of at least aa vertices such that every rr vertices in UU have at least mm common neighbors.

Conlon, Fox, and Sudakov [10] extended Lemma A.1 to hypergraphs. The weight w(S)w(S) of a set SS of edges in a hypergraph is the number of vertices in the union of these edges.

Lemma A.2 (Hypergraph Dependent Random Choice).

[10] Suppose s,Δs,\Delta are positive integers, ε,δ>0\varepsilon,\delta>0, and Gr=(V1,,Vr;E)G_{r}=(V_{1},\ldots,V_{r};E) is an rr-uniform rr-partite hypergraph with |V1|==|Vr|=N|V_{1}|=\ldots=|V_{r}|=N and at least εNr\varepsilon N^{r} edges. Then there exists an (r1)(r-1)-uniform (r1)(r-1)-partite hypergraph Gr1G_{r-1} on the vertex sets V2,,VrV_{2},\ldots,V_{r} which has at least εs2Nr1\frac{\varepsilon^{s}}{2}N^{r-1} edges and such that for each nonnegative integer w(r1)Δw\leq(r-1)\Delta, there are at most 4rΔεsβswrΔrwNw4r\Delta\varepsilon^{-s}\beta^{s}w^{r\Delta}r^{w}N^{w} dangerous sets of edges of Gr1G_{r-1} with weight ww, where a set SS of edges of Gr1G_{r-1} is dangerous if |S|Δ|S|\leq\Delta and the number of vertices vV1v\in V_{1} such that for every edge eS,e+vGre\in S,e+v\in G_{r} is less than βN\beta N.

Proof of Lemma 3.6.

Given a constant d>0d>0 and integers p,q2p,q\geq 2, we choose

1n1C,εd,1p,1qand in addition1nη.\tfrac{1}{n}\ll\tfrac{1}{C},\varepsilon\ll d,\tfrac{1}{p},\tfrac{1}{q}\leavevmode\nobreak\ \text{and in addition}\leavevmode\nobreak\ \tfrac{1}{n}\ll\eta.

Let GG be an nn-vertex graph with αp(G)<g(n)\alpha_{p}(G)<g(n), where

g(n)=n2Clog11/qn.g(n)=n2^{-C\log^{1-1/q}n}.

Let V1,V2,,VqV_{1},V_{2},\ldots,V_{q} be given such that |Vi|ηn|V_{i}|\geq\eta n, i[q]i\in[q] and every pair (Vi,Vj)(V_{i},V_{j}) is ε\varepsilon-regular with density at least dd. We define a qq-uniform qq-partite hypergraph H0H^{0} whose vertex set is i[q]Vi\cup_{i\in[q]}V_{i} and edge set E(H0)E(H^{0}) is the family of qq-sets that span qq-cliques in GG and contain one vertex from each of V1,,VqV_{1},\ldots,V_{q}. We may assume |Vi|=ηn=:N|V_{i}|=\eta n=:N, then by the counting lemma, |E(H0)|ε0Nq|E(H^{0})|\geq\varepsilon_{0}N^{q}, where ε0>(d/3)(q2)\varepsilon_{0}>\left(d/3\right)^{\binom{q}{2}}. Let

β=g(n)N,s=log1qn,εi=ε0si2si1s1,ri=qi,Δi=priandwi=pri.\beta=\tfrac{g(n)}{N},\quad s=\log^{\frac{1}{q}}n,\quad\varepsilon_{i}=\varepsilon_{0}^{s^{i}}2^{-\frac{s^{i}-1}{s-1}},\quad r_{i}=q-i,\quad\Delta_{i}=p^{r_{i}}\quad\textnormal{and}\quad w_{i}=pr_{i}.

We start from H0H^{0}. For 1iq21\leq i\leq q-2 we apply Lemma A.2 to Hi1H^{i-1} with Δ=Δi,ε=εi1,r=ri1\Delta=\Delta_{i},\varepsilon=\varepsilon_{i-1},r=r_{i-1} and w=wiw=w_{i} to get HiH^{i}. Note that Δ,ε0,r,w\Delta,\varepsilon_{0},r,w are all constants and 1Cd,1p,1q\frac{1}{C}\ll d,\frac{1}{p},\frac{1}{q}. It is easy to check that for 1iq21\leq i\leq q-2, we have

4rΔεsβswrΔrwNw\displaystyle 4r\Delta\varepsilon^{-s}\beta^{s}w^{r\Delta}r^{w}N^{w} =\displaystyle= O(22logi1qnε0logiqn(1/η)log1qn2Clogn(ηn)w)\displaystyle O\left(2^{2\log^{\frac{i-1}{q}}n}\varepsilon_{0}^{-\log^{\frac{i}{q}}n}(1/\eta)^{\log^{\frac{1}{q}}n}2^{-C\log n}(\eta n)^{w}\right)
=\displaystyle= O(nC/2)=o(1)<1.\displaystyle O(n^{-C/2})=o(1)<1.

Then by Lemma A.2 there exists an rir_{i}-uniform rir_{i}-partite hypergraph HiH^{i} on the vertex sets Vi+1,,VqV_{i+1},\ldots,V_{q} that contains at least εiNri\varepsilon_{i}N^{r_{i}} edges and contains no dangerous sets of Δi\Delta_{i} edges on wiw_{i} vertices. (Recall that a set SS of Δi\Delta_{i} edges on wiw_{i} vertices is dangerous if the number of vertices vViv\in V_{i} for which for every edge eS,e+vHi1e\in S,e+v\in H^{i-1} is less than βN\beta N). Now we have a hypergraph sequence {H}=0q2\{H^{\ell}\}_{\ell=0}^{q-2}. We will prove by induction on ii that there is a pp-set AqVqA^{q-\ell}\subset V_{q-\ell} for 0i0\leq\ell\leq i such that G[Aq]=KpG\left[A^{q-\ell}\right]=K_{p} and Hqi1[=0iAq]H^{q-i-1}\left[\bigcup_{\ell=0}^{i}A^{q-\ell}\right] is complete rqi1r_{q-i-1}-partite. Note that if a vertex set TT is an edge of H0H^{0}, then G[T]G[T] is a qq-clique. So G[=0q1Aq]=KpqG\left[\bigcup_{\ell=0}^{q-1}A^{q-\ell}\right]=K_{pq}, which will prove Lemma 3.6.

We first show that the induction hypothesis holds for i=1i=1. Note that rq2r_{q-2} = 2, so Hq2H^{q-2} is a bipartite graph on 2N2N vertices with at least εq2N2\varepsilon_{q-2}N^{2} edges. We now apply Lemma A.1 to Hq2H^{q-2} with

a=2βN,d=εq2N,t=s,r=pandm=βN.a=2\beta N,\qquad d=\varepsilon_{q-2}N,\qquad t=s,\qquad r=p\qquad\textnormal{and}\qquad m=\beta N.

We check condition (1):

(εq2N)s(2N)s1(2Np)(βN2N)s\displaystyle\frac{(\varepsilon_{q-2}N)^{s}}{(2N)^{s-1}}-{\binom{2N}{p}}\left(\frac{\beta N}{2N}\right)^{s} \displaystyle\geq (ε0/2)log11/qnNnp(1/2η)log1qn2Clogn\displaystyle(\varepsilon_{0}/2)^{\log^{1-1/q}n}N-n^{p}(1/2\eta)^{\log^{\frac{1}{q}}n}2^{-C\log n}
=\displaystyle= (ε0/2)log11/qnNo(1)2βN,\displaystyle(\varepsilon_{0}/2)^{\log^{1-1/q}n}N-o(1)\geq 2\beta N,

where the last equality and inequality follow as long as C>max{p,log2ε0}C>\max\{p,\log\frac{2}{\varepsilon_{0}}\}. Therefore we have a subset UU of Vq1VqV_{q-1}\cup V_{q} with |U|=2βN|U|=2\beta N such that every pp vertices in UU have at least βN\beta N common neighbors in Hq2H^{q-2}. Either Vq1V_{q-1} or VqV_{q} contains at least half of the vertices of UU, so w.l.o.g. we may assume that U=UVq1U^{\prime}=U\cap V_{q-1} contains at least βN=m\beta N=m vertices. Because αp(G)<m\alpha_{p}(G)<m, the vertex set UU^{\prime} contains a pp-vertex set Aq1A^{q-1} such that G[Aq1]=KpG\left[A^{q-1}\right]=K_{p}. The vertices of Aq1A^{q-1} have at least mm common neighbors in VqV_{q}, so their common neighborhood also contains a pp-vertex subset AqA^{q} of VqV_{q} such that G[Aq]=KpG[A^{q}]=K_{p}. Now Hq2[Aq1Aq]H^{q-2}\left[A^{q-1}\cup A^{q}\right] is complete bipartite. We are done with the base case i=1i=1.

For the induction step, assume that the induction hypothesis holds for i1i-1, then we can find a complete rqir_{q-i}-partite subhypergraph H~qi\widetilde{H}^{q-i} of HqiH^{q-i} spanned by =0i1Aq\bigcup_{\ell=0}^{i-1}A^{q-\ell}, where G[Aq]=KpG[A^{q-\ell}]=K_{p} for every \ell. The hypergraph HqiH^{q-i} has no dangerous set of Δqi\Delta_{q-i} edges on wqiw_{q-i} vertices, and H~qi\widetilde{H}^{q-i} contains pi=wqipi=w_{q-i} vertices and pi=Δqip^{i}=\Delta_{q-i} edges, so H~qi\widetilde{H}^{q-i} is not dangerous. Then we can find a set BB of βN\beta N vertices in VqiV_{q-i} such that for every edge eH~qie\in\widetilde{H}^{q-i} and every vertex vBv\in B, e+vHqi1e+v\in H^{q-i-1}, which means that Hqi1[B=0i1Aq]H^{q-i-1}\left[B\cup\bigcup_{\ell=0}^{i-1}A^{q-\ell}\right] is complete rqi1r_{q-i-1}-partite. Then, because αp(G)<βN\alpha_{p}(G)<\beta N, we can find a pp-vertex subset AqiA^{q-i} of BB such that G[Aqi]=KpG[A^{q-i}]=K_{p}. ∎