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

\floatsetup

heightadjust=object, valign=c

Size-Ramsey numbers of structurally sparse graphs

Nemanja Draganić Department of Mathematics, ETH, Zürich, Switzerland. Research supported in part by SNSF grant 200021_196965.
Emails: {nemanja.draganic,david.munhacanascorreia}@math.ethz.ch.
   Marc Kaufmann Institute of Theoretical Computer Science, Department of Computer Science, ETH, Zürich, Switzerland. M.K. was supported by SNSF grant 200021_192079. K.P. was supported by grant no. CRSII5 173721 of the Swiss National Science Foundation. R.S. was supported by an ETH Zürich Postdoctoral Fellowship.
Emails: {marc.kaufmann,kalina.petrova,raphaelmario.steiner}@inf.ethz.ch.
   David Munhá Correia11footnotemark: 1    Kalina Petrova22footnotemark: 2    Raphael Steiner22footnotemark: 2
Abstract

Size-Ramsey numbers are a central notion in combinatorics and have been widely studied since their introduction by Erdős, Faudree, Rousseau and Schelp in 1978. Research has mainly focused on the size-Ramsey numbers of nn-vertex graphs with constant maximum degree Δ\Delta. For example, graphs which also have constant treewidth are known to have linear size-Ramsey numbers. On the other extreme, the canonical examples of graphs of unbounded treewidth are the grid graphs, for which the best known bound has only very recently been improved from O(n3/2)O(n^{3/2}) to O(n5/4)O(n^{5/4}) by Conlon, Nenadov and Trujić. In this paper, we study a common generalization of these problems and establish new bounds on the size-Ramsey numbers in terms of treewidth (which may grow as a function of nn). As a special case, this yields a bound of O~(n3/21/2Δ)\tilde{O}(n^{3/2-1/2\Delta}) for proper minor-closed classes of graphs. In particular, this bound applies to planar graphs, addressing a question of Kamcev, Liebenau, Wood and Yepremyan. Our proof combines methods from structural graph theory and classic Ramsey-theoretic embedding techniques, taking advantage of the product structure exhibited by graphs with bounded treewidth.

1 Introduction

The famous theorem of Ramsey states that for every positive integer nn, there exists a finite number NN such that every complete graph on NN vertices whose edges are colored in red or in blue contains a monochromatic clique of order nn [26]. Motivated by this, we say that a host graph GG is kk-Ramsey for a graph HH if any kk-coloring of the edges of GG yields a monochromatic copy of HH, and we write G𝑘HG\xrightarrow{k}H.

Given a graph HH, its size-Ramsey number r^k(H)\hat{r}^{k}(H) is defined as the smallest number of edges in a graph GG such that G𝑘HG\xrightarrow{k}H. This notion measures the minimality of the host graph GG in a precise manner, and was introduced in 1978 by Erdős, Faudree, Rousseau and Schelp [16]. Since then, it has been extensively studied and in particular in recent years it has gained a lot of attention.

Intuitively, sparse graphs are a good candidate for having low size-Ramsey number, but already double stars — trees containing only two non-leaf vertices, each of degree linear in the size of the graph — have quadratic size-Ramsey number [16]. Thus, a natural class of graphs to study in this context is the class of bounded-degree graphs. Among the first instances of this was the result of Beck [3] who proved that for paths it holds that r^(Pn)900n\hat{r}(P_{n})\leq 900n, answering a $100\$100-dollar question of Erdős. Furthermore, Friedman and Pippenger [17] proved that bounded-degree trees have linear size-Ramsey numbers (i.e., bounded from above by a constant times their order). More recently, it has been shown that graphs with bounded degree and bounded treewidth exhibit the same behaviour [21, 4], as well as long subdivisions of bounded-degree graphs [13].

Perhaps surprisingly, not all bounded-degree graphs have this behaviour. In fact, there exist graphs with maximum degree three which serve as counterexamples, as shown by Rödl and Szemerédi [28]. More concretely, they constructed cubic graphs HH on nn vertices fulfilling r^(H)cnlog160(n)\hat{r}(H)\geq c\cdot n\cdot\log^{\frac{1}{60}}(n) for a universal constant c>0c>0. Their lower bound has been improved very recently to r^(H)cneclog(n)\hat{r}(H)\geq c\cdot n\cdot e^{c\sqrt{\log(n)}} by Tikhomirov [29], modifying their construction by a clever randomization trick. In spite of these advances, the conjecture of Rödl and Szemerédi that there exist bounded-degree nn-vertex graphs HH with r^(H)n1+ε\hat{r}(H)\geq n^{1+\varepsilon} for some absolute constant ε>0\varepsilon>0 remains open.

Turning to general upper bounds, Kohayakawa, Schacht, Rödl and Szemerédi [23] showed that for graphs HH which have maximum degree Δ\Delta it holds that r^(H)=O(n21/Δlog1/Δn)\hat{r}(H)=O(n^{2-1/\Delta}\log^{1/\Delta}n), leaving a wide gap between the best known upper and lower bounds. Recently, Allen and Böttcher [1] announced an improvement of this bound to r^(H)=O(n21/(Δ1)+o(1))\hat{r}(H)=O(n^{2-1/(\Delta-1)+o(1)}) for Δ4\Delta\geq 4. For the case Δ=3\Delta=3, the first and fourth authors [14] showed that r^(H)=O(n3/2+o(1))\hat{r}(H)=O(n^{3/2+o(1)}). As mentioned above, the class of bounded-degree graphs with constant treewidth is far better understood, and in fact it is known that the size-Ramsey numbers of these graphs are linear in nn [4, 21]. Unfortunately, the proofs in both [4] and [21] do not generalize beyond the setting of constant treewidth. In particular, if one allows the treewidth of the considered graphs HH to grow with nn, then thus far no better upper bounds on r^(H)\hat{r}(H) than the bound of Kohayakawa et al. (which only takes into account degree bounds, and not the structure of the graphs) were known [23] . In this paper, we generalize the above-mentioned results to graph classes of unbounded treewidth. In particular, leveraging the treewidth gives a substantial improvement over the bound of Kohayakawa et al.

Theorem 1.1.

Let kk\in\mathbb{N} and let HH be an nn-vertex graph of constant maximum degree Δ\Delta and treewidth t=t(n)t=t(n). Then

  • r^k(H)=O(ntlogn)\hat{r}_{k}(H)=O(nt\log n)

  • Furthermore, if t=Ω(elogn)t=\Omega(e^{\sqrt{\log n}}) then we have111We use O~(f(n))\tilde{O}(f(n)) to denote all functions which are in O(f(n)(logn)C)O(f(n)(\log n)^{C}) for some absolute constant C>0C>0. r^k(H)=O~(nt11/Δ)\hat{r}_{k}(H)=\tilde{O}(nt^{1-1/\Delta}).

Treewidth is a key parameter in structural graph theory and has served as the watershed in Robertson and Seymour’s seminal sequence of papers on graph minors, culminating in what is now known as the Graph Minor Theorem. Tree decompositions, which are inherently related to treewidth, have enabled advances in algorithmic graph theory, with many NP-hard problems proving solvable in polynomial time for entire graph classes whose treewidth is known to be bounded.

In fact, Theorem 1.1 is a special case of more general results, namely Lemma 3.5 and Lemma 3.6, which we prove in Section 3. Indeed, there we show that any bounded-degree graph contained in a product of a bounded-degree tree and a clique of certain size has small size-Ramsey number. Since (as we will prove in Section 2) graphs of bounded degree and treewidth tt are contained in such a product with certain parameters, Theorem 1.1 will follow.

As the treewidth of many graph classes is well understood, Theorem 1.1 may provide good upper bounds for the size-Ramsey numbers of graphs in those classes. One such example are proper minor-closed classes of graphs, for which it is known [2, 15] that they have treewidth at most O(n)O(\sqrt{n}), and hence we have the following consequence of Theorem 1.1.

Theorem 1.2.

Let kk\in\mathbb{N} and let 𝒢\mathcal{G} be a proper minor-closed class. Then for every nn-vertex graph G𝒢G\in\mathcal{G} of maximum degree Δ\Delta, we have

r^k(G)=O~(n3/212Δ).\hat{r}_{k}(G)=\tilde{O}(n^{3/2-\frac{1}{2\Delta}}).

A natural and well studied class of minor-closed graphs are planar graphs, and hence the bound above also applies to them. This addresses a question raised by Kamčev, Liebenau, Wood and Yepremyan [21], who asked for general bounds for bounded-degree planar graphs. Until recently, the best known bound for the two-dimensional grid graph on nn-vertices was O(n3/2)O(n^{3/2}), which was improved to O(n5/4)O(n^{5/4}) by Conlon, Nenadov and Trujić [8]. The only other bounded-degree planar graphs which were studied are cycles and bounded-degree trees, which have linear size-Ramsey numbers.

Let us also remark that all our results are of universality type. Namely, in our proofs we construct a host graph such that we can find all the graphs in the relevant class within the same color class. Results of this type were also obtained for size-Ramsey numbers of graphs in other classes [17, 12, 13, 22, 23, 25].

Organization. In Section 2 we prove several results which embed a graph with a given bound on its treewidth into the product of a bounded-degree tree and a clique. The results in this section might be of independent interest. In Section 3 we prove bounds on the size-Ramsey numbers of bounded-degree subgraphs of products of bounded-degree trees and cliques, which, combined with the results from Section 2 then allow us to prove our main results. We decided to defer some of the technical details of the proofs from Section 3 to the appendix (Section A) to ease the readability and intuitive understanding of our proofs.

Notation. For simplicity, we employ the following conventions. We omit rounding of real numbers to nearest integers whenever it is not of vital importance. For two constants a,ba,b, we use aba\ll b to indicate that bb is sufficiently large as a function of aa for our proofs to work out. For example, we often use inequality chains like abcda\gg b\gg c\gg d, which also implies that in particular abcda\gg bcd. We make use of the strong product operator, which is defined as follows. The strong product GHG\boxtimes H of graphs GG and HH is the graph with vertex set V(G)×V(H)V(G)\times V(H) such that for v1,v2V(G)v_{1},v_{2}\in V(G) and u1,u2V(H)u_{1},u_{2}\in V(H), we have that (v1,u1)(v_{1},u_{1}) is adjacent to (v2,u2)(v_{2},u_{2}) in GHG\boxtimes H if v1=v2v_{1}=v_{2} and {u1,u2}E(H)\{u_{1},u_{2}\}\in E(H), or {v1,v2}E(G)\{v_{1},v_{2}\}\in E(G) and u1=u2u_{1}=u_{2}, or {v1,v2}E(G)\{v_{1},v_{2}\}\in E(G) and {u1,u2}E(H)\{u_{1},u_{2}\}\in E(H).

2 Structural results

In this section, we prove the announced structural lemma for graphs of bounded treewidth. It says that every bounded-degree nn-vertex graph GG of treewidth at most tt can be embedded as a subgraph into the strong product of a bounded-degree tree of size O(n/(tlogn))O(n/(t\log n)) and a clique of size O(tlogn)O(t\log n). In the case that GG belongs to a structurally sparse class of graphs, such as planar graphs or a proper minor-closed class, we can get rid of the logn\log n factor in the above estimate, which can be used to further improve Theorem 1.2 by removing additional hidden logarithmic factors. Evidently, the significance of this result does not lie in this minor improvement, but rather in the fact that it is tight (consider, say, the grid graph).

Lemma 2.1.

Let Δ>0\Delta>0. There exists d>0d>0 depending solely on Δ\Delta such that the following hold.

  1. (i)

    For every nn-vertex graph GG of maximum degree at most Δ\Delta and treewidth at most tt there exists a tree TT such that GTKsG\subseteq T\boxtimes K_{s}, where ns=ΘΔ(tlogn)n\geq s=\Theta_{\Delta}(t\log n), v(T)=OΔ(ns)v(T)=O_{\Delta}\big{(}\frac{n}{s}\big{)} and Δ(T)d\Delta(T)\leq d.

  2. (ii)

    Let 𝒢\mathcal{G} be a class of graphs closed under taking subgraphs, and let t:++t:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+} be an increasing function in Ω(logx)\Omega(\log x) such that every nn-vertex graph G𝒢G\in\mathcal{G} has treewidth at most t(n)t(n). Suppose further that there exists a constant α>0\alpha>0 such that t(λx)λαt(x)t(\lambda x)\leq\lambda^{\alpha}t(x) holds for every λ,x>0\lambda,x>0. Then for every nn-vertex graph G𝒢G\in\mathcal{G} of maximum degree at most Δ\Delta there is a tree TT such that GTKsG\subseteq T\boxtimes K_{s}, where ns=ΘΔ,α(t(n))n\geq s=\Theta_{\Delta,\alpha}(t(n)), v(T)=OΔ(ns)v(T)=O_{\Delta}\big{(}\frac{n}{s}\big{)} and Δ(T)d\Delta(T)\leq d.

We remark that a very similar embedding lemma, without the logn\log n factor, has been proved before by Ding and Oporowski [10] and later on improved by Wood [30]. These results state that every graph GG of maximum-degree Δ\Delta and treewidth at most tt is contained in TKsT\boxtimes K_{s}, where TT is some tree and s=O(Δt)s=O(\Delta t). However, for our purposes, it is very important to be able to carefully control both the size and the maximum degree of the tree TT. In particular, we need TT to have constant degree, and this is not the case in the results in [10, 30]. Motivated by Lemma 2.1, very recently Distel and Wood [11] showed a strengthening of it by removing the logarithmic factor in the first part of Lemma 2.1.

Before moving on to the proof of Lemma 2.1, let us note the following immediate consequence of it.

Corollary 2.2.

Let Δ>0\Delta>0 and let 𝒢\mathcal{G} be a proper minor-closed class. There exist constants c,d>0c,d>0 such that every nn-vertex graph G𝒢G\in\mathcal{G} of maximum degree at most Δ\Delta is a subgraph of TKsT\boxtimes K_{s}, where scns\leq c\sqrt{n} and TT is a tree with v(T)cnv(T)\leq c\sqrt{n} and Δ(T)d\Delta(T)\leq d.

  • Proof.

    It is well-known [2, 15] that for every proper minor-closed class 𝒢\mathcal{G} of graphs, there is a constant C>0C>0 such that every nn-vertex graph G𝒢G\in\mathcal{G} has treewidth at most CnC\sqrt{n}. The statement therefore immediately follows from Lemma 2.1 (ii) by setting t(x)=Cxt(x)=C\sqrt{x} and α=12\alpha=\frac{1}{2}. ∎

We now move on to the proof of Lemma 2.1, which follows closely the arguments in [6, 5]. A crucial tool to embed a given graph GG of treewidth tt into the strong product of a tree and a clique is a repeated application of the following well-known result from structural graph theory concerning the existence of so-called balanced separators in graphs of bounded treewidth.

Lemma 2.3 (Robertson and Seymour, cf. statement (2.6) in [27]).

Let GG be a graph of treewidth at most tt and order nn. Then there exists a set SV(G)S\subseteq V(G) of size |S|t+1|S|\leq t+1 and a partition of V(G)SV(G)\setminus S into sets AA and BB such that no edge in GSG-S connects a vertex in AA to a vertex in BB and such that |A|,|B|23n|A|,|B|\leq\frac{2}{3}n.

To prove Lemma 2.1, we first establish an extension of Lemma 2.3 which works when the graph GG comes with a given coloring of its vertices with kk colors, and we wish to find a small separator SS in the graph such that the two sides AA and BB separated by it contain each at most half the vertices in each color. To find such separators, we make use of the celebrated necklace-splitting theorem due to Goldberg and West in the following form.

Theorem 2.4 (Goldberg and West [20]).

For some R{1,,n}R\subseteq\{1,\ldots,n\}, let c:R{1,,k}c:R\rightarrow\{1,\ldots,k\} be a (partial) kk-coloring of the first nn integers. Then {1,,n}\{1,\ldots,n\} can be partitioned into k+1k+1 pairwise disjoint intervals I1,,Ik+1I_{1},\ldots,I_{k+1} such that for some 0rk+10\leq r\leq k+1 it holds that (denoting X:=i=1rIi,Y:=i=r+1k+1IiX:=\bigcup_{i=1}^{r}{I_{i}},Y:=\bigcup_{i=r+1}^{k+1}{I_{i}})

max{|Xc1(i)|,|Yc1(i)|}|c1(i)|2\max\{|X\cap c^{-1}(i)|,|Y\cap c^{-1}(i)|\}\leq\left\lceil\frac{|c^{-1}(i)|}{2}\right\rceil

for i=1,,ki=1,\ldots,k.

Lemma 2.5.

Let GG be a graph and let t:++t:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+} be an increasing function such that every subgraph of GG with xx vertices has treewidth at most t(x)t(x).

Then there exists a linear ordering v1,,vnv_{1},\ldots,v_{n} of the vertices in GG and for each i{1,,n}i\in\{1,\ldots,n\} a set S(vi)V(G)S(v_{i})\subseteq V(G) including viv_{i} such that

|S(vi)|j=0log3/2(n)(t((23)jn)+1)|S(v_{i})|\leq\sum_{j=0}^{\lceil\log_{3/2}(n)\rceil}{\left(t\left(\left(\frac{2}{3}\right)^{j}n\right)+1\right)}

and such that there are no edges between {v1,,vi1}S(vi)\{v_{1},\ldots,v_{i-1}\}\setminus S(v_{i}) and {vi+1,,vn}S(vi)\{v_{i+1},\ldots,v_{n}\}\setminus S(v_{i}) in GG.

  • Proof.

    We prove the statement by induction on nn. The claim is trivially true for n=1n=1 by setting S(v1):={v1}S(v_{1}):=\{v_{1}\}, so moving on suppose that n2n\geq 2 and the claim is true for all subgraphs of GG on fewer vertices. Apply Lemma 2.3 to find a set SV(G)S\subseteq V(G) of size at most t(n)+1t(n)+1 and a partition of V(G)SV(G)\setminus S into parts AA and BB with no edges in GG between them, such that |A|,|B|23n|A|,|B|\leq\frac{2}{3}n. By the inductive assumption, there exist linear orderings v1,,v|A|v_{1},\ldots,v_{|A|} and w1,,w|B|w_{1},\ldots,w_{|B|} of AA and BB and for each i{1,,|A|}i\in\{1,\ldots,|A|\} and j{1,,|B|}j\in\{1,\ldots,|B|\} sets SA(vi)S_{A}(v_{i}) and SB(wj)S_{B}(w_{j}) such that the following hold. SA(vi)viS_{A}(v_{i})\ni v_{i} separates {vk|k<i}\{v_{k}|k<i\} and {vk|k>i}\{v_{k}|k>i\} in G[A]G[A], and SB(wj)wjS_{B}(w_{j})\ni w_{j} separates {wk|k<j}\{w_{k}|k<j\} and {wk|k>j}\{w_{k}|k>j\} in G[B]G[B], and with m:=max{|A|,|B|}23nm:=\max\{|A|,|B|\}\leq\frac{2}{3}n we have

    |SA(vi)|,|SB(vj)|k=0log3/2(m)(t((23)km)+1)|S_{A}(v_{i})|,|S_{B}(v_{j})|\leq\sum_{k=0}^{\lceil\log_{3/2}(m)\rceil}{\left(t\left(\left(\frac{2}{3}\right)^{k}m\right)+1\right)}
    k=1log3/2(m)+1(t((23)kn)+1)k=1log3/2(n)(t((23)kn)+1).\leq\sum_{k=1}^{\lceil\log_{3/2}(m)+1\rceil}{\left(t\left(\left(\frac{2}{3}\right)^{k}n\right)+1\right)}\leq\sum_{k=1}^{\lceil\log_{3/2}(n)\rceil}{\left(t\left(\left(\frac{2}{3}\right)^{k}n\right)+1\right)}.

    Let s1,,s|S|s_{1},\ldots,s_{|S|} be an arbitrary linear ordering of SS, and consider the following linear order on V(G)V(G): v1,,v|A|,s1,,s|S|,w1,,w|B|v_{1},\ldots,v_{|A|},s_{1},\ldots,s_{|S|},w_{1},\ldots,w_{|B|}. Furthermore let S(vi):=SA(vi)SS(v_{i}):=S_{A}(v_{i})\cup S for i=1,,|A|i=1,\ldots,|A|, S(sk):=SS(s_{k}):=S for k=1,,|S|k=1,\ldots,|S|, and S(wj):=SB(wj)SS(w_{j}):=S_{B}(w_{j})\cup S for j=1,,|B|j=1,\ldots,|B|. It is easily verified that each of these sets separates their predecessors and successors in the ordering in GG and since |S|t(n)+1=t((2/3)0n)+1|S|\leq t(n)+1=t((2/3)^{0}n)+1, each of these sets is of size at most h=0log3/2(n)(t((23)hn)+1)\sum_{h=0}^{\lceil\log_{3/2}(n)\rceil}{\left(t\left(\left(\frac{2}{3}\right)^{h}n\right)+1\right)}. This concludes the proof. ∎

Using the previous statement and Lemma 2.4, we immediately arrive at the following corollary.

Corollary 2.6.

Let GG be a graph given with a (partial) coloring c:R{1,,k}c:R\rightarrow\{1,\ldots,k\} of its vertices for some RV(G)R\subseteq V(G) and let t:++t:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+} be an increasing function such that every subgraph of GG with xx vertices has treewidth at most t(x)t(x). Then there is a set SV(G)S\subseteq V(G) and a partition A,BA,B of V(G)SV(G)\setminus S such that

|S|k+ki=0log3/2(n)(t((23)in)+1),|S|\leq k+k\sum_{i=0}^{\lceil\log_{3/2}(n)\rceil}{\left(t\left(\left(\frac{2}{3}\right)^{i}n\right)+1\right)},

there are no connections between AA and BB in GG, and

max{|Ac1(i)|,|Bc1(i)|}|c1(i)|2\max\{|A\cap c^{-1}(i)|,|B\cap c^{-1}(i)|\}\leq\frac{|c^{-1}(i)|}{2}

for each i{1,,k}i\in\{1,\ldots,k\}.

  • Proof.

    We apply Lemma 2.5 to GG to get the linear ordering v1,,vnv_{1},\ldots,v_{n} of V(G)V(G). Next, we apply Theorem 2.4 to the coloring c(i):=c(vi)c(i):=c(v_{i}). Let I1,,Ir,Ir+1,,Ik,Ik+1I_{1},\ldots,I_{r},I_{r+1},\ldots,I_{k},I_{k+1} be an interval partition of {1,,n}\{1,\ldots,n\} as guaranteed by the lemma. Let x1,,xkx_{1},\ldots,x_{k} be a list of kk elements in {1,,n}\{1,\ldots,n\} such that any two distinct intervals in I1,I2,,Ik+1I_{1},I_{2},\ldots,I_{k+1} are separated by at least one of them. We now define S:=S(x1)S(x2)S(xk)S:=S(x_{1})\cup S(x_{2})\cup\cdots\cup S(x_{k}), which by Lemma 2.5 is of size at most ki=0log3/2(n)(t((23)in)+1)k\sum_{i=0}^{\lceil\log_{3/2}(n)\rceil}{\left(t\left(\left(\frac{2}{3}\right)^{i}n\right)+1\right)}, and let A:={vi|iI1Ir}S,B:={vi|iIr+1Ik+1}SA:=\{v_{i}|i\in I_{1}\cup\cdots\cup I_{r}\}\setminus S,B:=\{v_{i}|i\in I_{r+1}\cup\cdots\cup I_{k+1}\}\setminus S. It now follows directly from Lemma 2.5 that there are no edges in GG connecting AA and BB and directly from the properties guaranteed by Theorem 2.4 that max{|Ac1(i)|,|Bc1(i)|}|c1(i)|2\max\{|A\cap c^{-1}(i)|,|B\cap c^{-1}(i)|\}\leq\left\lceil\frac{|c^{-1}(i)|}{2}\right\rceil for each i{1,,k}i\in\{1,\ldots,k\}. In order to reduce from the above bound |c1(i)|2\left\lceil\frac{|c^{-1}(i)|}{2}\right\rceil to the desired |c1(i)|2\frac{|c^{-1}(i)|}{2} for each ii, we may simply move for each color ii a vertex of that color from either AA or BB into SS, thereby increasing the size of SS by at most kk, and the statement follows. ∎

Using Corollary 2.6 repeatedly, we can now prove the following statement, which directly implies Lemma 2.1.

Lemma 2.7.

Let GG be a graph of maximum degree at most Δ\Delta and let t:++t:\mathbb{R}^{+}\rightarrow\mathbb{R}^{+} be an increasing function such that every xx-vertex subgraph of GG has treewidth at most t(x)t(x). Let

k:=log2(Δ)+2,s:=k+ki=0log3/2(n)(t((23)in)+1).k:=\lceil\log_{2}(\Delta)\rceil+2,\,s:=\left\lfloor k+k\sum_{i=0}^{\lceil\log_{3/2}(n)\rceil}{\left(t\left(\left(\frac{2}{3}\right)^{i}n\right)+1\right)}\right\rfloor.

Then there exists a binary tree TT and a partition (Xv)vV(T)(X_{v})_{v\in V(T)} of V(G)V(G) such that |Xv|=2s|X_{v}|=2s for every non-leaf vV(T)v\in V(T), |Xv|2s|X_{v}|\leq 2s for every leaf vV(T)v\in V(T), and such that any two adjacent vertices in GG are contained in sets Xu,XvX_{u},X_{v} for vertices u,vV(T)u,v\in V(T) at distance at most kk in TT.

  • Proof.

    To enable induction, we prove a slight strengthening of the statement in the lemma as follows:

Claim.

Given any sequence C1,,CkC_{1},\ldots,C_{k} of disjoint subsets of V(G)V(G) such that |Ci|2i1s|C_{i}|\leq 2^{i-1}\cdot s for each ii, we can find a tree TT and a partition (Xv)vV(T)(X_{v})_{v\in V(T)} as in the lemma with the following additional property: For each i=1,,ki=1,\ldots,k and every vertex vv in TT at distance at least ii from the root, we have XvCi=X_{v}\cap C_{i}=\varnothing.

Let us now prove this claim by induction on v(G)v(G). It trivially holds true if v(G)2sv(G)\leq 2s, so suppose that v(G)>2sv(G)>2s and the claim is satisfied by every subgraph of GG with fewer than v(G)v(G) vertices. Interpret the disjoint sets C1,,CkC_{1},\ldots,C_{k} as the color classes of a partial coloring of V(G)V(G), and apply Corollary 2.6 to find a set SS of size at most ss in GG and a partition (A,B)(A,B) of V(G)SV(G)\setminus S such that |CiA|,|CiB||Ci|22i2s|C_{i}\cap A|,|C_{i}\cap B|\leq\frac{|C_{i}|}{2}\leq 2^{i-2}\cdot s for each ii. Pick a set XX of size exactly 2s2s in V(G)V(G) such that C1SXC_{1}\cup S\subseteq X (this is possible since |C1|,|S|s|C_{1}|,|S|\leq s and v(G)>2sv(G)>2s). Now, let A:=AXA^{\prime}:=A\setminus X, B:=BXB^{\prime}:=B\setminus X. For each i=1,,k1i=1,\ldots,k-1, define CiA:=Ci+1AC_{i}^{A}:=C_{i+1}\cap A^{\prime}, CiB:=Ci+1BC_{i}^{B}:=C_{i+1}\cap B^{\prime}. Further, let CkA:=(NG(X)A)i=1k1CiAC_{k}^{A}:=(N_{G}(X)\cap A^{\prime})\setminus\bigcup_{i=1}^{k-1}{C_{i}^{A}} and CkB:=(NG(X)B)i=1k1CiBC_{k}^{B}:=(N_{G}(X)\cap B^{\prime})\setminus\bigcup_{i=1}^{k-1}{C_{i}^{B}}. From the above we have |CiA|,|CiB|2(i+1)2s=2i1s|C_{i}^{A}|,|C_{i}^{B}|\leq 2^{(i+1)-2}s=2^{i-1}s for i=1,,k1i=1,\ldots,k-1, as well as |CkA|,|CkB||NG(X)|Δ2s2k1s|C_{k}^{A}|,|C_{k}^{B}|\leq|N_{G}(X)|\leq\Delta\cdot 2s\leq 2^{k-1}s since GG is of maximum degree at most Δ\Delta and by our choice of kk.

Therefore, we can apply induction to G[A]G[A^{\prime}] with the sets C1A,,CkAC_{1}^{A},\ldots,C_{k}^{A} and to G[B]G[B^{\prime}] with the sets C1B,,CkBC_{1}^{B},\ldots,C_{k}^{B} to find binary trees TAT_{A} and TBT_{B}, and partitions (Xv)vV(TA),(Xv)vV(TB)(X_{v})_{v\in V(T_{A})},(X_{v})_{v\in V(T_{B})} of AA^{\prime} and BB^{\prime} satisfying the inductive claim. We now form a binary tree TT by adding a new root vertex rr and connecting it to the roots of the binary trees TAT_{A} and TBT_{B}, and setting Xr:=XX_{r}:=X. We claim that this tree and the partition satisfy the claim. Let us first verify the condition with respect to the sets C1,,CkC_{1},\ldots,C_{k}. For i=1i=1, and any vertex vv at distance at least 11 from rr in TT, we have that XvX_{v} is disjoint from XC1X\supseteq C_{1}. Further, for i{2,,k}i\in\{2,\ldots,k\}, every vertex vv at distance at least ii from rr in TT has distance at least i1i-1 from the root in either TAT_{A} or TBT_{B}, thus implying that XvX_{v} is disjoint from Ci1A=CiAC_{i-1}^{A}=C_{i}\cap A^{\prime} respectively Ci1B=CiBC_{i-1}^{B}=C_{i}\cap B^{\prime}, and therefore XvCi=X_{v}\cap C_{i}=\varnothing in each case.

Hence, all that remains to be verified is that any two adjacent vertices in GG are contained in sets Xu,XvX_{u},X_{v} for vertices u,vu,v at distance at most kk in TT. So consider any two vertices u,vu,v in TT such that there exists an edge in GG with endpoints in XuX_{u} and XvX_{v}. If u,vru,v\neq r then either both of u,vu,v are contained in TAT_{A} or both in TBT_{B}, and then it follows from the validity of the inductive claim for TAT_{A} and TBT_{B} that uu and vv are at distance at most kk in TT. Next consider the case u=ru=r, and w.l.o.g. assume vV(TA)v\in V(T_{A}). Then XvNG(Xu)=XvNG(X)=Xv(NG(X)A)\varnothing\neq X_{v}\cap N_{G}(X_{u})=X_{v}\cap N_{G}(X)=X_{v}\cap(N_{G}(X)\cap A^{\prime}), so in particular XvX_{v} intersects CkAC_{k}^{A} or at least one of C1A,,Ck1AC_{1}^{A},\ldots,C_{k-1}^{A}, let’s say CiAXvC_{i}^{A}\cap X_{v}\neq\varnothing for some i{1,,k}i\in\{1,\ldots,k\}. But then by our inductive call to G[A]G[A^{\prime}], vv has to be at distance at most i1i-1 from the root in TAT_{A}. But this means that vv is at distance at most ii from the root rr in TT, showing that u=ru=r and vv are at distance at most iki\leq k also in this last case. Thus, GG satisfies the inductive assertion with respect to C1,,CkC_{1},\ldots,C_{k}, concluding the proof of the lemma. ∎

We are now ready to show Lemma 2.1.

  • Proof of Lemma 2.1.

    We do the proof of both cases (i) and (ii) of the lemma at once. Every subgraph of GG on xx vertices has treewidth at most t(x)t(x), where we let t(x):=tt(x):=t for each xx if we are in case (i) of the lemma. Let kk and ss be defined as in Lemma 2.7 and note that we obtain s=ΘΔ(tlogn)s=\Theta_{\Delta}(t\log n) in case (i) and

    k(t(n)+1)<sk+ki=0log3/2(n)((23)αit(n)+1)k(t(n)+1)<s\leq k+k\sum_{i=0}^{\lceil\log_{3/2}(n)\rceil}{\left(\left(\frac{2}{3}\right)^{\alpha i}t(n)+1\right)}
    k+k1(2/3)αt(n)+k(log3/2(n)+1)=OΔ,α(t(n))+OΔ(logn)=OΔ,α(t(n)),\leq k+\frac{k}{1-(2/3)^{\alpha}}t(n)+k(\lceil\log_{3/2}(n)\rceil+1)=O_{\Delta,\alpha}(t(n))+O_{\Delta}(\log n)=O_{\Delta,\alpha}(t(n)),

    hence s=ΘΔ,α(t(n))s=\Theta_{\Delta,\alpha}(t(n)) in case (ii).

    We can now apply Lemma 2.7 to find a binary tree TT and a partition (Xv)vT(X_{v})_{v\in T} of the vertex set of GG such that |Xv|=2s|X_{v}|=2s for every non-leaf vv of TT, |Xv|2s|X_{v}|\leq 2s for every leaf vv of TT, and such that whenever XuX_{u} and XvX_{v} have a connecting edge in GG, then uu and vv are at distance at most k=log2(Δ)+2k=\lceil\log_{2}(\Delta)\rceil+2 in TT. Note that the number of non-leaves in TT is at most n2s\frac{n}{2s}, and since a binary tree of size at least 33 has at most one more leaf than non-leaf, we find that v(T)ns+1v(T)\leq\frac{n}{s}+1. Furthermore, we can see from the above that GG is a subgraph of TkK2sT^{k}\boxtimes K_{2s}, where TkT^{k} is the graph on the same vertex set as TT in which two vertices are adjacent iff they are at distance at most kk from each other in TT. Notice that since TT is a binary tree, it is not hard to show that TkTK2kT^{k}\subseteq T^{\prime}\boxtimes K_{2^{k}} where TT^{\prime} is a tree such that v(T)v(T)ns+1v(T^{\prime})\leq v(T)\leq\frac{n}{s}+1 and Δ(T)1+2k=:d\Delta(T^{\prime})\leq 1+2^{k}=:d. All in all, we find that G(TK2k)K2sTK2k+1sG\subseteq(T^{\prime}\boxtimes K_{2^{k}})\boxtimes K_{2s}\simeq T^{\prime}\boxtimes K_{2^{k+1}s}. Since 2k+1s=ΘΔ(s)2^{k+1}s=\Theta_{\Delta}(s), the assertion of the lemma in both cases now follows from the above asymptotic estimates of ss. ∎

3 Proof of main result

In this section we introduce some tools that we need, and give the proofs of our main results. We start with the main result from [4], which states that graphs with constant treewidth and constant degree have linear size-Ramsey numbers, and moreover, that there exists a host graph for this class of graphs which also has constant maximum degree.

Theorem 3.1 (Corollary 2 in [4]).

For any positive integers k,Δ,tk,\Delta,t, and for every nn large enough, there exists a graph GG with O(n)O(n) vertices and constant maximum degree such that for every nn-vertex graph HH of maximum degree Δ\Delta and treewidth at most tt, it holds that GkHG\rightarrow_{k}H.

The theorem above is stated in [4] without the maximum degree condition on GG, but the construction in [4] clearly satisfies it.

3.1 Regularity method

Similarly to previous work on size-Ramsey numbers, to find monochromatic subgraphs in our host graph, we make use of the celebrated sparse regularity method (for a detailed exposition of which we refer the reader to [19]). Instead of working with regular pairs as usual, it is enough for our purposes to only have lower bounds on the density of subgraphs, as expressed in the following definition.

Definition 3.2.

Let α,ε>0\alpha,\varepsilon>0 and 0<p10<p\leq 1. For a graph G=(V,E)G=(V,E) and disjoint sets X,YVX,Y\subseteq V, we say that (X,Y)(X,Y) is (ε,α,p)(\varepsilon,\alpha,p)-dense if for any XX,YYX^{\prime}\subseteq X,Y^{\prime}\subseteq Y with |X|ε|X||X^{\prime}|\geq\varepsilon|X| and |Y|ε|Y||Y^{\prime}|\geq\varepsilon|Y|, we have

dG(X,Y)(αε)p.d_{G}(X^{\prime},Y^{\prime})\geq(\alpha-\varepsilon)p.

To be able to apply the sparse regularity method, we need our host graph to have somewhat ‘uniformly’ distributed edges, typically captured by the concept of (λ,p)(\lambda,p)-uniformity as follows.

Definition 3.3.

A graph G=(V,E)G=(V,E) is (λ,p)(\lambda,p)-uniform if for all disjoint U,WVU,W\subseteq V with |U|,|W|λ|V||U|,|W|\geq\lambda|V|, we have (1λ)pdG(U,W)(1+λ)p(1-\lambda)p\leq d_{G}(U,W)\leq(1+\lambda)p. If GG is bipartite with bipartition V=ABV=A\cup B, then we say that GG is (λ,p)(\lambda,p)-uniform if the same holds for all UAU\subseteq A and WBW\subseteq B of size at least λ|A|\lambda|A| and λ|B|\lambda|B| respectively.

The next lemma, which we will use in the proofs of Lemmas 3.5 and 3.6, allows us to find certain useful structure in colored blow-ups of bounded-degree graphs. Namely, it turns out one can choose a small fraction of the vertices in each part of the blow-up in such a way that the bipartite graph between each pair of these subparts has at least one color with the lower-regularity properties captured by Definition 3.2 and some minimum density.

Lemma 3.4 (Lemma 2.3 in [8]).

For every k,Δ2k,\Delta\geq 2 and ε>0\varepsilon>0, there exists λ>0\lambda>0 such that the following holds for every p(0,1]p\in(0,1]. Let HH be a graph on at least two vertices with Δ(H)Δ\Delta(H)\leq\Delta and let Γ\Gamma be obtained by replacing every xV(H)x\in V(H) with an independent set VxV_{x} of sufficiently large order nn and every xyE(H)xy\in E(H) by a (λ,p)(\lambda,p)-uniform bipartite graph between VxV_{x} and VyV_{y}. Then, for every kk-colouring of the edges of Γ\Gamma, there exists a kk-colouring ϕ\phi of the edges of HH and, for every xV(H)x\in V(H), a subset UxVxU_{x}\subseteq V_{x} of order |Ux|=λn|U_{x}|=\lambda n such that (Ux,Uy;Eϕ(xy))(U_{x},U_{y};E_{\phi(xy)}) is (ε,12k,p)(\varepsilon,\frac{1}{2k},p)-dense for each xyHxy\in H where Eϕ(xy)EΓE_{\phi(xy)}\subseteq E_{\Gamma} stands for the edges of colour ϕ(xy)\phi(xy) between UxU_{x} and UyU_{y} in Γ\Gamma.

3.2 Key lemmas and proof of Theorem 1.1

As announced in the introduction, here we prove Theorem 1.1. To do so, we will show the following two stronger lemmas, each of which implies one of the two statements in Theorem 1.1. Both of them give bounds on the size-Ramsey number of bounded-degree graphs which are contained in a product of a bounded-degree tree and a clique of certain size. We first state the two lemmas, and show how Theorem 1.1 is implied. In the next subsection, we give a proof of the first lemma, which relies on the proof of the classic lemma for embedding graphs into regular partitions. We are succinct with the details, as the embedding lemma is by now a standard tool (for an in-depth survey of it and other foundational regularity tools, see [24]), and only minor modifications are necessary for our application.

After that, we give a proof of the second lemma, which relies on the ideas used for the first result, and on the embedding machinery developed in [23], with several changes needed to adapt their method to our setting. We present those changes in the appendix.

Lemma 3.5.

Let Δ,d,k\Delta,d,k be positive integers and let HH be an nn-vertex graph with Δ(H)Δ\Delta(H)\leq\Delta, such that HTKsH\subset T\boxtimes K_{s} for some s=s(n)s=s(n) and a tree TT with Δ(T)d\Delta(T)\leq d and v(T)=τ=τ(n)v(T)=\tau=\tau(n). Then r^k(H)=O(τs2)\hat{r}_{k}(H)=O(\tau s^{2}).

Lemma 3.6.

Let Δ,d,k\Delta,d,k be positive integers and let HH be an nn-vertex graph with Δ(H)Δ\Delta(H)\leq\Delta, such that HTKsH\subset T\boxtimes K_{s} for some s=Ω(elogn)s=\Omega(e^{\sqrt{\log{n}}}) and a tree TT with Δ(T)d\Delta(T)\leq d and v(T)=τ=τ(n)v(T)=\tau=\tau(n). Then r^k(H)=O(τs2(logss)1/Δ)\hat{r}_{k}(H)=O(\tau s^{2}(\frac{\log s}{s})^{1/\Delta}).

  • Proof of Theorem 1.1.

    Let HH be an nn-vertex graph of maximum degree Δ\Delta and treewidth t=t(n)t=t(n). By Lemma 2.1, we know that HTKsH\subseteq T\boxtimes K_{s} for s=Θ(tlogn)s=\Theta(t\log n) and v(T)=O(n/s)v(T)=O(n/s), with Δ(T)d=d(Δ)\Delta(T)\leq d=d(\Delta). Now, by Lemma 3.5, the size-Ramsey number of HH is O(ns)=O(ntlogn)O(ns)=O(nt\log n), which gives the first part of the theorem.

    For the second part, assume t=Ω(elogn)t=\Omega(e^{\sqrt{\log{n}}}). Note that by Lemma 3.6 and since s=Θ(tlogn)=Ω(elogn)s=\Theta(t\log n)=\Omega(e^{\sqrt{\log{n}}}), we have r^(H)=O(ns(logss)1/Δ)=O~(nt11/Δ)\hat{r}(H)=O(ns(\frac{\log s}{s})^{1/\Delta})=\tilde{O}(nt^{1-1/\Delta}), as required. ∎

3.3 Proof of Lemma 3.5

In this section, we give the proof of our first key lemma. We build our host graph in a way that is convenient for the structure TKsT\boxtimes K_{s} we know our graph HH has. Namely, we take a graph \mathcal{R} that is Ramsey for a constant blow-up of TT and blow it up by a factor of roughly ss. We then make use of Lemma 3.4 to get colorful lower-regular pairs within each random bipartite graph, thus yielding an auxiliary coloring of \mathcal{R}. Next, using that we picked \mathcal{R} to be Ramsey for a constant blow-up of TT, we find a monochromatic copy of the latter in the auxiliary coloring, which corresponds to a monochromatic TT-shaped structure of lower-regular pairs in our host graph. We can then use an approach similar to the embedding lemma to embed HH into that monochromatic structure, being guided by the containment of HH in TKsT\boxtimes K_{s} to map each vertex to an appropriate set.

  • Proof of Lemma 3.5.

    Set C1λεk1,Δ1,d1C^{-1}\ll\lambda\ll\varepsilon\ll k^{-1},\Delta^{-1},d^{-1}. Let F=TKΔ+1F=T\boxtimes K_{\Delta+1} and note that FF has treewidth at most 2Δ+22\Delta+2 by the definition of treewidth. Let \mathcal{R} be the graph given by Theorem 3.1 which is kk-Ramsey for FF and which has O(v(F))=O(τ)O(v(F))=O(\tau) vertices and constant maximum degree. The host graph for HH which we will consider is G=KCsG=\mathcal{R}\boxtimes K_{Cs}. Note that GG has O(τs2)O(\tau s^{2}) edges. For each xV()x\in V(\mathcal{R}), denote the copy of KCsK_{Cs} corresponding to xx by VxV_{x}. Consider an arbitrary kk-coloring of the edges of GG.

    We apply Lemma 3.4 with p=1p=1 and λ\lambda to the graph Γ:=G\Gamma:=G, which is a blow-up of \mathcal{R}. Hence, we get a coloring of the edges of \mathcal{R} and, for each xV()x\in V(\mathcal{R}), a subset UxVxU_{x}\subseteq V_{x} of size λCs\lambda Cs, such that for every edge {x,y}\{x,y\} in \mathcal{R}, the edges between UxU_{x} and UyU_{y} form an (ε,12k,1)(\varepsilon,\frac{1}{2k},1)-dense pair in the color of the edge {x,y}\{x,y\} (meaning that the density between every two sets XUxX\subseteq U_{x} and YUyY\subseteq U_{y} of sizes |X|ε|Ux||X|\geq\varepsilon|U_{x}| and |Y|ε|Uy||Y|\geq\varepsilon|U_{y}| is at least 12kε\frac{1}{2k}-\varepsilon).

    Now, since \mathcal{R} is kk-Ramsey for FF, there is a monochromatic copy of FF in \mathcal{R}, which in turn means that for every edge {x,y}\{x,y\} of that copy, the pair UxU_{x} and UyU_{y} forms a (ε,12k,1)(\varepsilon,\frac{1}{2k},1)-dense pair in the same color, say red. As a last step before describing the embedding process, for each vV(T)v\in V(T), denote with HvH_{v} the subgraph of HH which lives inside of the copy of KsK_{s} that corresponds to the vertex vv (recall that HTKsH\subseteq T\boxtimes K_{s}). Since HH has maximum degree Δ\Delta, it is (Δ+1)(\Delta+1)-partite, so denote by Hv1,,Hvd+1H_{v}^{1},\ldots,H_{v}^{d+1} one such partition of HvH_{v}, for each vV(T)v\in V(T). To summarize, we now want to embed HTKsH\subseteq T\boxtimes K_{s} into the red subgraph of GG which is obtained from TKΔ+1KλCsT\boxtimes K_{\Delta+1}\boxtimes K_{\lambda Cs} by replacing the complete bipartite graphs between the relevant copies of KλCsK_{\lambda Cs} by (ε,12k,1)(\varepsilon,\frac{1}{2k},1)-dense pairs.

    The embedding of HH into the found red subgraph of GG is done by a standard procedure, which is a variant of the well-known embedding lemma. Briefly, for each vV(T)v\in V(T), we will want to embed HvH_{v} into vKΔ+1KλCsv\boxtimes K_{\Delta+1}\boxtimes K_{\lambda Cs}, so that each HviH_{v}^{i} is embedded into its own copy of KλCsK_{\lambda Cs} in vKΔ+1KλCsv\boxtimes K_{\Delta+1}\boxtimes K_{\lambda Cs}. This embedding can be done vertex by vertex, crucially maintaining the following invariant. For each vertex xV(H)x\in V(H) which has been assigned to the copy UU of KλCsK_{\lambda Cs} but has not yet been embedded, we consider the ‘candidate set’ CxC_{x} that consists of all the vertices in UU that are in the common neighbourhood of all already embedded neighbours of xx in HH. While embedding HH, we ensure that for each such xx, the size of the candidate set CxC_{x} is at least (14k)i|U|(\frac{1}{4k})^{i}|U|, where ii is the number of neighbours of xx in HH which are already embedded at the observed point of the embedding process. When it is time for vertex xx to be embedded, we choose a vertex yCxy\in C_{x} for that purpose, in such a way that yy has enough neighbours in the candidate set of each neighbour xx^{\prime} of xx in HH yet to be embedded, in order to preserve the aforementioned invariant. This is possible because, due to the lower-regular properties of the graph between CxC_{x} and CxC_{x^{\prime}}, at most ελCs\varepsilon\lambda Cs of the vertices in CxC_{x} have degree less than 14k|Cx|\frac{1}{4k}|C_{x^{\prime}}| into CxC_{x^{\prime}}. Additionally, at most ss many vertices in CxC_{x} are already taken by vertices previously embedded in UU, leaving us with at least

    |Cx|ΔελCss(14k)ΔλCsΔελCss1|C_{x}|-\Delta\varepsilon\lambda Cs-s\geq\Big{(}\frac{1}{4k}\Big{)}^{\Delta}\lambda Cs-\Delta\varepsilon\lambda Cs-s\gg 1

    viable candidates for the embedding of xx, where we used that Δ(H)Δ\Delta(H)\leq\Delta and (14k)Δε(\frac{1}{4k})^{\Delta}\gg\varepsilon. Thus, we can continue doing this until we embed the whole graph HH. For more details, we refer the reader to [7] where such an embedding lemma is proven; the difference here is that we have more than a constant number of large cliques in which we want to embed, but the proof is essentially the same.∎

3.4 Proof of Lemma 3.6

We make use of the following result, which allows us to embed any bounded-degree subgraph of a blow-up of a bounded-degree graph \mathcal{R} into a certain blow-up of \mathcal{R} where the bipartite graphs between the parts are lower-regular pairs as in Definition 3.2. Although here we only apply the lemma below with \mathcal{R} being a tree, we state it in this more general form, as it might be of independent interest and may prove to be useful in other settings which deal with embedding bounded-degree graphs.

Lemma 3.7.

Let Δ,d,ρ1ε1CC\Delta,d,\rho^{-1}\ll\varepsilon^{-1}\ll C^{\prime}\ll C and nn be large enough. Let s=Ω(elogn)s=\Omega(e^{\sqrt{\log{n}}}) and nsτn\frac{n}{s}\leq\tau\leq n. Set p:=C(logss)1/Δp:=C\big{(}\frac{\log{s}}{s}\big{)}^{1/\Delta} and Δ~:=Δ4+2Δ+1\tilde{\Delta}:=\Delta^{4}+2\Delta+1. Then w.h.p. GG(τΔ~Cs,p)G\sim G(\tau\tilde{\Delta}C^{\prime}s,p) is such that the following holds. Let HH be an nn-vertex graph of maximum degree Δ\Delta that is a subgraph of Ks\mathcal{R}\boxtimes K_{s} for some τ\tau-vertex graph \mathcal{R} with Δ()d\Delta(\mathcal{R})\leq d. Let J:=KΔ~J:=\mathcal{R}\boxtimes K_{\tilde{\Delta}} and FGF\subseteq G be a graph with vertex set V(F)=xV(J)FxV(F)=\bigcup_{x\in V(J)}F_{x} such that |Fx|=Cs|F_{x}|=C^{\prime}s for each xV(J)x\in V(J) and edge set consisting of an (ε,ρ,p)(\varepsilon,\rho,p)-dense bipartite graph between FxF_{x} and FyF_{y} for each xyE(J)xy\in E(J). Then HH can be embedded in FF.

The proof of Lemma 3.7 mostly follows the approach from [23], with some changes needed for our setting. We defer it to the appendix.

We are now ready to present the proof of our second key lemma. It makes use of the same main ideas as the first one, except that to get the tighter bound, the complete bipartite graphs between the sets of the host graph are now replaced by random graphs with smaller than constant density pp. This drop in density then requires a much more careful treatment of the embedding process of HH, for which the embedding lemma no longer suffices. This is where Lemma 3.7 proves to be helpful.

  • Proof of Lemma 3.6.

    Let HH be as given above. We shall show that there exists a graph GG with O(τs21/Δlog1/Δs)O(\tau s^{2-1/\Delta}\log^{1/\Delta}s) many edges such that GkHG\rightarrow_{k}H.

    It will be convenient to define the following constants. Let Δ~:=Δ4+2Δ+1\tilde{\Delta}:=\Delta^{4}+2\Delta+1 and let

    Δ,d,kdε1CC.\Delta,d,k\ll d^{\prime}\ll\varepsilon^{-1}\ll C^{\prime}\ll C.

    Let p=C(logss)1/Δp=C\big{(}\frac{\log{s}}{s}\big{)}^{1/\Delta}. Let λ\lambda be as given by Lemma 3.4 with k:=kk:=k, ε:=ε\varepsilon:=\varepsilon, and Δ:=d\Delta:=d^{\prime}.

    Let \mathcal{R} be the graph that is kk-Ramsey for all τΔ~\tau\tilde{\Delta}-vertex graphs of maximum degree Δ~(d+1)\tilde{\Delta}(d+1) and maximum treewidth 2Δ~2\tilde{\Delta} given by Theorem 3.1. In particular, kTKΔ~\mathcal{R}\rightarrow_{k}T\boxtimes K_{\tilde{\Delta}}, since by the definition of treewidth, TKΔ~T\boxtimes K_{\tilde{\Delta}} has treewidth at most 2Δ~2\tilde{\Delta}. Note that \mathcal{R} has O(τ)O(\tau) vertices and constant maximum degree, say dd^{\prime}. We define our host graph GG to be obtained by replacing each vV()v\in V(\mathcal{R}) with an independent set VvV_{v} on λ1Cs\lambda^{-1}C^{\prime}s vertices and each edge uvE()uv\in E(\mathcal{R}) with a random bipartite graph between VuV_{u} and VvV_{v} in which each edge is added independently at random with probability pp.

    Now consider an arbitrary colouring of the edges of GG in kk colours. We will show that HH can be embedded into one of the colour classes of GG. By a standard application of Chernoff bounds and a union bound, with high probability each random bipartite graph between VuV_{u} and VvV_{v} with uvE()uv\in E(\mathcal{R}) is (λ,p)(\lambda,p)-uniform. We fix an outcome of GG for which this is the case. In particular, this also implies that e(G)=O(τs21/Δlog1/Δs)e(G)=O(\tau s^{2-1/\Delta}\log^{1/\Delta}s).

    We can now apply Lemma 3.4 with Γ:=G\Gamma:=G, H:=H:=\mathcal{R}, n:=λ1Csn:=\lambda^{-1}C^{\prime}s, and p:=pp:=p. This yields a kk-colouring ϕ\phi of the edges of \mathcal{R} and, for every xV()x\in V(\mathcal{R}), a subset UxVxU_{x}\subseteq V_{x} of size |Ux|=λ|Vx|=Cs|U_{x}|=\lambda|V_{x}|=C^{\prime}s such that for each xyE()xy\in E(\mathcal{R}), the graph (Ux,Uy;Eϕ(xy))(U_{x},U_{y};E_{\phi(xy)}) is (ε,12k,p)(\varepsilon,\frac{1}{2k},p)-dense, where Eϕ(xy)EGE_{\phi(xy)}\subseteq E_{G} are the edges in colour ϕ(xy)\phi(xy).

    Next, note that by Theorem 3.1, the kk-colouring ϕ\phi of the edges of \mathcal{R} contains a monochromatic copy of TKΔ~=:JT\boxtimes K_{\tilde{\Delta}}=:J. This implies that GG contains a monochromatic subgraph FF with vertex set V(F)=xV(J)FxV(F)=\bigcup_{x\in V(J)}F_{x} with |Fx|=Cs|F_{x}|=C^{\prime}s for each xV(J)x\in V(J) and edge set consisting of an (ε,12k,p)(\varepsilon,\frac{1}{2k},p)-dense bipartite graph between FxF_{x} and FyF_{y} for each xyE(J)xy\in E(J).

    We can now apply Lemma 3.7 with :=T\mathcal{R}:=T and ρ:=12k\rho:=\frac{1}{2k}, concluding that the graph FF contains HH as a subgraph. Since the copy of FF in GG was monochromatic, this concludes the proof of the lemma. ∎

Acknowledgments

We thank David Wood for asking for the bounds on the size-Ramsey number of bounded-degree planar graphs, which led to this paper [31].

Appendix A Appendix

Here we give the proof of Lemma 3.7, which roughly follows [23]. We begin by establishing certain useful ‘uniformity’ properties, which allow us to proceed with the embedding later. First we define two classes of ‘bad’ tripartite graphs, pI\mathcal{B}^{I}_{p} and pII\mathcal{B}^{II}_{p}, where density of pairs (X,Y),(Y,Z)(X,Y),(Y,Z) is not inherited on one- respectively two-sided neighborhoods.222The definition slightly deviates from Definition 13 in [23], where the pairs (X,Y)(X,Y) and (Y,Z)(Y,Z) are both (ε,α,p)(\varepsilon,\alpha,p)-dense. Here, for technical reasons, we include the more general case of (η,α,p)(\eta,\alpha,p)-dense pairs (X,Y)(X,Y) whereby η\eta can differ from ε\varepsilon.

Definition A.1.

Let integers m1,m2,m3m_{1},m_{2},m_{3} and reals α,ε,ε,μ>0\alpha,\varepsilon^{\prime},\varepsilon,\mu>0, η>0\eta>0, and 0<p10<p\leq 1 be given.

  1. 1.

    Let pI(m1,m2,m3,α,ε,ε,μ,η)\mathcal{B}^{I}_{p}(m_{1},m_{2},m_{3},\alpha,\varepsilon^{\prime},\varepsilon,\mu,\eta) be the family of tripartite graphs with vertex set X˙Y˙ZX\dot{\cup}Y\dot{\cup}Z, where |X|=m1|X|=m_{1}, |Y|=m2|Y|=m_{2}, and |Z|=m3|Z|=m_{3}, satisfying

    1. (a)

      (X,Y)(X,Y) is a (η,α,p)(\eta,\alpha,p)-dense pair,

    2. (b)

      (Y,Z)(Y,Z) is a (ε,α,p)(\varepsilon,\alpha,p)-dense pair, and

    3. (c)

      there exists XXX^{\prime}\subseteq X with |X|μ|X||X^{\prime}|\geq\mu|X| such that for every xXx\in X^{\prime} the pair (N(x)Y,Z)(N(x)\cap Y,Z) is not (ε,α,p)(\varepsilon^{\prime},\alpha,p)-dense.

  2. 2.

    Let pII(m1,m2,m3,α,ε,ε,μ,η)\mathcal{B}^{II}_{p}(m_{1},m_{2},m_{3},\alpha,\varepsilon^{\prime},\varepsilon,\mu,\eta) be the family of tripartite graphs with vertex set X˙Y˙ZX\dot{\cup}Y\dot{\cup}Z, where |X|=m1|X|=m_{1}, |Y|=m2|Y|=m_{2}, and |Z|=m3|Z|=m_{3}, satisfying

    1. (a)

      (X,Y)(X,Y) and (X,Z)(X,Z) are (η,α,p)(\eta,\alpha,p)-dense pairs,

    2. (b)

      (Y,Z)(Y,Z) is a (ε,α,p)(\varepsilon,\alpha,p)-dense pair, and

    3. (c)

      there exists XXX^{\prime}\subseteq X with |X|μ|X||X^{\prime}|\geq\mu|X| such that for every xXx\in X^{\prime} the pair (N(x)Y,N(x)Z)(N(x)\cap Y,N(x)\cap Z) is not (ε,α,p)(\varepsilon^{\prime},\alpha,p)-dense.

Graphs that do not contain large subgraphs — induced or not — from the two above ‘bad’ families are then said to have a denseness property 𝒟N,pΔ\mathcal{D}^{\Delta}_{N,p}, which we make precise in the next definition.

Definition A.2.

For integers NN and Δ2\Delta\geq 2 and reals α,γ,ε,ε,μ,η>0\alpha,\gamma,\varepsilon^{\prime},\varepsilon,\mu,\eta>0 and 0<p10<p\leq 1 we say that a graph G=(V,E)G=(V,E) with V=[N]V=[N] has the denseness property 𝒟N,pΔ(γ,α,ε,ε,μ,η)\mathcal{D}^{\Delta}_{N,p}(\gamma,\alpha,\varepsilon^{\prime},\varepsilon,\mu,\eta), if GG contains no member from

BpI(m1I,m2I,m3I,α,ε,ε,μ,η)BpII(m1II,m2II,m3II,α,ε,ε,μ,η)B^{I}_{p}(m^{I}_{1},m^{I}_{2},m^{I}_{3},\alpha,\varepsilon^{\prime},\varepsilon,\mu,\eta)\cup B^{II}_{p}(m^{II}_{1},m^{II}_{2},m^{II}_{3},\alpha,\varepsilon^{\prime},\varepsilon,\mu,\eta)

with m1I,m3IγpΔ1Nm^{I}_{1},m^{I}_{3}\geq\gamma p^{\Delta-1}N and m2I,m1II,m2II,m3IIγpΔ2Nm_{2}^{I},m_{1}^{II},m_{2}^{II},m_{3}^{II}\geq\gamma p^{\Delta-2}N as a (not necessarily induced) subgraph.

The following key technical lemma which is Corollary 16 in [23] (and follows by repeatedly applying Proposition 15 therein) guarantees that random graphs G(N,p)G(N,p) enjoy this denseness property at various parameter scales, whenever p(logNN)1Δp\gg(\frac{\log N}{N})^{\frac{1}{\Delta}}. It will not be enough for our purposes that this fails to hold with any probability o(1)o(1). However, performing the calculations explicitly allows us to bound the failure probability precisely. The other adjustments in the proof will stem from the changes in the definition of the ‘bad’ graphs.

Lemma A.3 (Corollary 16 in [23]).

For all integers Δ,Δ2\Delta,\Delta^{\prime}\geq 2 and all reals α,μ,γ,ε>0\alpha,\mu,\gamma,\varepsilon^{*}>0, there exist C>1C>1, η=η(Δ,α,μ)\eta=\eta(\Delta,\alpha,\mu), and ε0,,εΔ\varepsilon_{0},\ldots,\varepsilon_{\Delta^{\prime}} satisfying 0<ε0εΔε0<\varepsilon_{0}\leq\ldots\leq\varepsilon_{\Delta^{\prime}}\leq\varepsilon^{*} such that if p>C(logN/N)1/Δp>C(\log{N}/N)^{1/\Delta}, then Pr[G(N,p)k=1Δ𝒟N,pΔ(γ,α,εk,εk1,μ,η)]1NΘ(NpΔ2)Pr[G(N,p)\in\cap_{k=1}^{\Delta^{\prime}}\mathcal{D}^{\Delta}_{N,p}(\gamma,\alpha,\varepsilon_{k},\varepsilon_{k-1},\mu,\eta)]\geq 1-N^{-\Theta(Np^{\Delta-2})}.

  • Proof.

    The proof is almost the same as that of Corollary 16 in [23], which can be summarized roughly as follows. Restricting to the case of one-sided regularity inheritance, one first considers the special case of a bad graph TT where the two parts XX and ZZ are of fixed size, namely only a pp-fraction of the size of YY (the more general case can be reduced to this one). The bipartite subgraphs T[X,Y]T[X,Y] and T[Y,Z]T[Y,Z] are quite dense by assumption. Now consider the violating set XXX^{\prime}\subset X, whose members xXx\in X^{\prime} all fail to produce (ε,α,p)(\varepsilon^{\prime},\alpha,p)-dense pairs (NT(x)Y,Z)(N_{T}(x)\cap Y,Z). We restrict our attention to X′′XX^{\prime\prime}\subseteq X^{\prime} such that each xX′′x\in X^{\prime\prime} has ‘large’ degree into YY. Because T[X,Y]T[X,Y] is (η,α,p)(\eta,\alpha,p)-dense, X′′X^{\prime\prime} contains at least half of the vertices in XX^{\prime}. Now one can show that each such vertex xx produces a set YxNT(x)YY^{\prime}_{x}\subset N_{T}(x)\cap Y of size εαpm2\frac{\varepsilon^{\prime}\alpha pm}{2} which, together with ZZ produces a pair (Yx,Z)(Y^{\prime}_{x},Z) of density strictly less than αε\alpha-\varepsilon^{\prime}. Any superset YxY_{x} of YxY^{\prime}_{x} in YY with size αpm2\frac{\alpha pm}{2} then produces a violating pair, that is, the pair (Yx,Z)(Y_{x},Z) is not (ε,α,p)(\varepsilon^{\prime},\alpha,p)-dense. By repeating this procedure for all xX′′x\in X^{\prime\prime}, we could produce an entire family of violating pairs, which is unlikely to occur in G(N,p)G(N,p). The latter can be proven by a union bound via counting the ways of choosing the vertices in the involved sets X′′,Y,ZX^{\prime\prime},Y,Z and the edges in the bipartite graph T[Y,Z]T[Y,Z] as well as the choices of the sets YxY_{x}, which are limited by sparse regularity inheritance from T[Y,Z]T[Y,Z] (cf. [18] Theorem 3.6).

    In what follows we note the necessary changes to Section 3.3 [23].

    • In the statement of Proposition 15, after ε=ε(Δ,α,ε,μ)>0\varepsilon=\varepsilon(\Delta,\alpha,\varepsilon^{\prime},\mu)>0, add ”there exists η=η(Δ,α,μ)>0\eta=\eta(\Delta,\alpha,\mu)>0” and change

      Pr[G(N,p)𝒟N,pΔ(γ,α,.εk,εk1,μ)]=1o(1)Pr[G(N,p)\in\mathcal{D}^{\Delta}_{N,p}(\gamma,\alpha,.\varepsilon_{k},\varepsilon_{k-1},\mu)]=1-o(1)

      to

      Pr[G(N,p)𝒟N,pΔ(γ,α,εk,εk1,μ,η)]1NΘ(NpΔ2).Pr[G(N,p)\in\mathcal{D}^{\Delta}_{N,p}(\gamma,\alpha,\varepsilon_{k},\varepsilon_{k-1},\mu,\eta)]\geq 1-N^{-\Theta(Np^{\Delta-2})}.
    • In the proof of Corollary 16, add the argument η\eta to 𝒟N,pΔ(γ,α,εk,εk1,μ)\mathcal{D}^{\Delta}_{N,p}(\gamma,\alpha,\varepsilon_{k},\varepsilon_{k-1},\mu).

    • Change the definition of 𝒟^N,pΔ\hat{\mathcal{D}}^{\Delta}_{N,p} to include η\eta as an argument, in agreement with 𝒟N,pΔ\mathcal{D}^{\Delta}_{N,p}.

    • In Proposition 17, add ”for each 0<η<μ100,α1000<\eta<\frac{\mu}{100},\frac{\alpha}{100}” and change the probability condition to Pr[G(N,p)𝒟^N,pΔ(γ,α,ε,ε,μ,η)]1NΘ(NpΔ2)Pr[G(N,p)\in\hat{\mathcal{D}}^{\Delta}_{N,p}(\gamma,\alpha,\varepsilon^{\prime},\varepsilon,\mu,\eta)]\geq 1-N^{-\Theta(Np^{\Delta-2})}. In particular, η<μ100\eta<\frac{\mu}{100} ensures that |X′′||X^{\prime\prime}| is large enough, see the change below on the lower bound at the top of page 5052.

    • In the proof of Proposition 17:

      • *

        Choose CC, which was originally set to C=(4γ)1ΔC=(\frac{4}{\gamma})^{\frac{1}{\Delta}}, to be, say, 100100 times larger. This will ensure the desired rate of decay later on in the proof, where the failure probability is estimated by a bound where pp, which depends linearly on CC, enters as p2p^{2} in the exponent. See also the change on page 5051 indicated below.

      • *

        Add η\eta as an argument to pI\mathcal{B}^{I}_{p} and pII\mathcal{B}^{II}_{p} everywhere since T[X,Y]T[X,Y] is (η,α,p)(\eta,\alpha,p)-dense according to Definition A.1 instead of (ε,α,p)(\varepsilon,\alpha,p)-dense as in [23].

      • *

        On page 5050, ”Because of the assumption on TT, the bipartite subgraphs T[X,Y]T[X,Y] and T[Y,Z]T[Y,Z] of TT contain at least (αε)p2m2(\alpha-\varepsilon)p^{2}m^{2} edges each” — this continues to hold for T[Y,Z]T[Y,Z] but is no longer true of T[X,Y]T[X,Y], instead it contains at least (αη)p2m2(\alpha-\eta)p^{2}m^{2} edges.

      • *

        On page 5050, ”From the (ε,α,p)(\varepsilon,\alpha,p)-denseness …” towards the bottom of the page changes to ”From the (η,α,p)(\eta,\alpha,p)-denseness …” and |X′′|(1ε/μ)|X||X^{\prime\prime}|\geq(1-\varepsilon/\mu)|X^{\prime}| changes to |X′′|(1η/μ)|X||X^{\prime\prime}|\geq(1-\eta/\mu)|X^{\prime}|.

      • *

        On page 5051, we get that the probability that T[X′′,Y,Z]T[X^{\prime\prime},Y,Z] appears in G(N,p)G(N,p) is at most NΘ(NpΔ2)N^{-\Theta(Np^{\Delta-2})} as a result of the new choice of CC.

      • *

        On top of page 5052, it is no longer true that T[X,Y]T[X,Y] and T[X,Z]T[X,Z] have at least (αε)pm2(\alpha-\varepsilon)pm^{2} edges each, they have instead at least (αη)pm2(\alpha-\eta)pm^{2} edges each.

      • *

        On top of page 5052, in the lower bound for |X′′||X^{\prime\prime}|, we get instead (12η/μ)|X||X|/2(1-2\eta/\mu)|X^{\prime}|\geq|X^{\prime}|/2, from the (η,α,p)(\eta,\alpha,p)-denseness of T[X,Y]T[X,Y] and T[X,Z]T[X,Z].

      • *

        On page 5052, we get that the probability that a graph from pII(m,α,ε,ε,μ,η)\mathcal{B}_{p}^{II}(m,\alpha,\varepsilon^{\prime},\varepsilon,\mu,\eta) is contained in G(n,p)G(n,p) is at most NΘ(NpΔ2)N^{-\Theta(Np^{\Delta-2})}.

    • In the proof of Proposition 15:

      • *

        Change the third sentence to ”Indeed, roughly speaking, we show that each ’bad’ tripartite graph TpII(m1,m2,m3,α,ε,ε,μ,η)T\in\mathcal{B}_{p}^{II}(m_{1},m_{2},m_{3},\alpha,\varepsilon^{\prime},\varepsilon,\mu,\eta) with arbitrary m1,m2,m3mm_{1},m_{2},m_{3}\geq m contains a subgraph T^pII(m,α,ε/2,ε^,μ/4,η^)\hat{T}\in\mathcal{B}_{p}^{II}(m,\alpha,\varepsilon^{\prime}/2,\hat{\varepsilon},\mu/4,\hat{\eta}) for some appropriate ε^\hat{\varepsilon} and η^\hat{\eta}”.

      • *

        In Claim 18, add η^\hat{\eta} to the list of given positive reals, add ”there exists η=η(η^,α,μ)\eta=\eta(\hat{\eta},\alpha,\mu)”, and add η\eta as an argument to pII(m1,m2,m3,α,ε,ε,μ)\mathcal{B}_{p}^{II}(m_{1},m_{2},m_{3},\alpha,\varepsilon^{\prime},\varepsilon,\mu) and η^\hat{\eta} as an argument to pII(m,α,ε/2,ε^,μ/4)\mathcal{B}_{p}^{II}(m,\alpha,\varepsilon^{\prime}/2,\hat{\varepsilon},\mu/4).

      • *

        In the paragraph after Claim 18, add ”Pick η^<μ400,α400\hat{\eta}<\frac{\mu}{400},\frac{\alpha}{400} for the application of Claim 18”, and add η^\hat{\eta} as an argument to pI\mathcal{B}^{I}_{p} and pII\mathcal{B}^{II}_{p}; also change the probability to 1NΘ(NpΔ2)1-N^{-\Theta(Np^{\Delta-2})}.

    • In the proof of Claim 18:

      • *

        Add η^\hat{\eta} to the list of given constants, and take η:=min{α400,μ400,ε0(α,β,η^)}\eta:=\min\{\frac{\alpha}{400},\frac{\mu}{400},\varepsilon_{0}(\alpha,\beta,\hat{\eta})\}.

      • *

        On page 5053, in the definition of T=(X˙Y˙Z,ET)T=(X\dot{\cup}Y\dot{\cup}Z,E_{T}), add η\eta as an argument to pII\mathcal{B}_{p}^{II}.

      • *

        On page 5053, replace ε\varepsilon with η\eta everywhere in the sentence ”Owing to the choice of ε\varepsilon …” (we can do this since η<μ100\eta<\frac{\mu}{100}).

      • *

        On the bottom of page 5053, change the last sentence of the penultimate paragraph to ”We shall show that with positive probability T^\hat{T} is from pII(m,α,ε/2,ε^,μ/4,η^)\mathcal{B}_{p}^{II}(m,\alpha,\varepsilon^{\prime}/2,\hat{\varepsilon},\mu/4,\hat{\eta}).

      • *

        In the last paragraph of page 5053, replace ε^\hat{\varepsilon} with η^\hat{\eta} when it comes to pairs (X^,Y^)(\hat{X},\hat{Y}) and (X^,Z^)(\hat{X},\hat{Z}).

      • *

        On the bottom of page 5054, add η^\hat{\eta} as an argument to pII(m,α,ε/2,ε^,μ/4)\mathcal{B}_{p}^{II}(m,\alpha,\varepsilon^{\prime}/2,\hat{\varepsilon},\mu/4).

Let G=(V,E)G=(V,E) be a graph and k1k\geq 1 be an integer. We define an auxiliary bipartite graph Γ(k,G)=((Vk)˙V,EΓ(k,G)\Gamma(k,G)=(\binom{V}{k}\dot{\cup}V,E_{\Gamma(k,G}) by

(K,v)EΓ(k,G){w,v}E(G) for all wK.(K,v)\in E_{\Gamma(k,G)}\Longleftrightarrow\{w,v\}\in E(G)\text{ for all }w\in K.

The next definition and lemma will help us conclude that if GG is the random graph G(N,p)G(N,p), then Γ(k,G)\Gamma(k,G) has no ”dense parts”.

Definition A.4.

Let integers NN and k1k\geq 1 and reals ξ>0\xi>0 and 0<p10<p\leq 1 be given. A graph G=(V,E)G=(V,E) with V=[N]V=[N] has the congestion property 𝒞N,pk(ξ)\mathscr{C}^{k}_{N,p}(\xi) if for every UVU\subseteq V and every family k(VUk)\mathcal{F}_{k}\subseteq\binom{V\setminus U}{k} of pairwise disjoint kk-sets with

  • |k|ξN|\mathcal{F}_{k}|\leq\xi N and

  • |U||k||U|\leq|\mathcal{F}_{k}|

we have

eΓ(k,G)(k,U)pk|k||U|+6ξNpk|k|.e_{\Gamma(k,G)}(\mathcal{F}_{k},U)\leq p^{k}|\mathcal{F}_{k}||U|+6\xi Np^{k}|\mathcal{F}_{k}|.

Similarly to the denseness property from Lemma A.3, we will require better control over the probability that a graph fails to have the congestion property than the original phrasing of Corollary 12 in [23]. We will hence adjust the proof accordingly.

Lemma A.5 (Corollary 12 in [23]).

For every integer Δ1\Delta\geq 1 and real ξ>0\xi>0, there exists C>1C>1 such that if p>C(logN/N)1/Δp>C(log{N}/N)^{1/\Delta}, then Pr[G(N,p)k=1Δ𝒞N,pk(ξ)]=1e100log2NPr[G(N,p)\in\cap_{k=1}^{\Delta}\mathscr{C}_{N,p}^{k}(\xi)]=1-e^{-100\log^{2}{N}}.

  • Proof.

    The proof is almost the same as that of Corollary 12 in [23]. In what follows we note the necessary changes in Section 3.2.

    • In the statement of Proposition 11, change o(1)o(1) to e200log2Ne^{-200\log^{2}{N}}.

    • At the end of the analysis for Case 1 in the proof of Proposition 11, note that the probability of the bad event can in fact be upper bounded by e2Ne^{-2N}.

    • At the end of Case 2 in the proof of Proposition 11, one can upper bound the probability of the bad event by e400log2Ne^{-400\log^{2}{N}}, for CC large enough.

We are now ready to prove Lemma 3.7, which we restate here for convenience of the reader. See 3.7

  • Proof of Lemma 3.7.

    We begin the proof by setting the constants. This will allow for the application of Lemmas A.3 and A.5 at the necessary level. In particular, the subgraphs of our host graph will have denseness and congestion properties which still hold after taking union bounds. The second step will consist of preparing our graph HH for its embedding. Here we use the assumption that HH is a subgraph of Ks\mathcal{R}\boxtimes K_{s}. We first partition the vertices of HH according to the copy of KsK_{s}, indexed by the vertices of \mathcal{R}, into which they are mapped. This ordered partition is then refined using a coloring (of the third power of H3H^{3}), ensuring that no two vertices at distance at most 33 with respect to HH are mapped to the same class. The actual embedding of HH into FF is then constructed inductively, maintaining at each step \ell for every not yet embedded vertex zz of HH a large enough candidate set C(z)C_{\ell}(z) which takes into account all the already embedded neighbors of zz, and ensuring that the candidate sets for pairs of not yet embedded vertices z,zz,z^{\prime} that are neighbors in HH have candidate sets C(z),C(z)C_{\ell}(z),C_{\ell}(z^{\prime}) which induce an appropriately regular pair in FF. Given a vertex yHy\in H and a candidate set C(y)C_{\ell}(y), a naive embedding of yy into it will not work, as previously embedded vertices from the same class may have exhausted the set. Instead, we use Hall’s condition to find one distinct vertex in C(y)C_{\ell}(y) for each yy in the current vertex class of HH simultaneously.

    Notation and Constants. We first fix some notation. In J=KΔ~J=\mathcal{R}\boxtimes K_{\tilde{\Delta}}, for each xV()x\in V(\mathcal{R}), we denote by JxJ_{x} the copy of KΔ~K_{\tilde{\Delta}} that corresponds to xx, and by 𝐅x\mathbf{F}_{x} the copy of KΔ~KCsK_{\tilde{\Delta}}\boxtimes K_{C^{\prime}s} in FF that corresponds to JxJ_{x}.

    Next, we choose the constants. Let ε0:=ε\varepsilon_{0}:=\varepsilon, μ:=14Δ2\mu:=\frac{1}{4\Delta^{2}}, ξ:=1(d+1)Δ~C\xi:=\frac{1}{(d+1)\tilde{\Delta}C^{\prime}} and γ:=1(2/ρ)Δ13Δ~\gamma:=\frac{1}{(2/\rho)^{\Delta-1}3\tilde{\Delta}}. Let η\eta be as given by Lemma A.3 applied with Δ:=Δ\Delta:=\Delta, α:=ρ\alpha:=\rho, μ:=μ\mu:=\mu and

    0<ε0ε2Δε:=min{ρ6Δ~,η}.0<\varepsilon_{0}\leq\ldots\leq\varepsilon_{2\Delta}\leq\varepsilon^{*}:=\min\{\frac{\rho}{6\tilde{\Delta}},\eta\}.

    be as given by Lemma A.3 applied with Δ:=Δ\Delta:=\Delta, Δ:=2Δ\Delta^{\prime}:=2\Delta, α:=ρ\alpha:=\rho, μ:=μ\mu:=\mu, γ:=γ\gamma:=\gamma. The value of ε\varepsilon^{*} ensures in particular that in the embedding which we will construct step by step, the candidate sets for not yet embedded vertices are large enough. The choice of γ\gamma will ensure that the candidate sets are large enough to form dense pairs. The choice of ξ\xi will enable us to verify Hall’s condition.

    For each distinct x,y,zV()x,y,z\in V(\mathcal{R}) such that xy,yzE()xy,yz\in E(\mathcal{R}), apply Lemma A.3 with Δ:=Δ\Delta:=\Delta, Δ:=2Δ\Delta^{\prime}:=2\Delta, α:=ρ\alpha:=\rho, μ:=μ\mu:=\mu, γ:=γ\gamma:=\gamma, ε:=ε\varepsilon^{*}:=\varepsilon^{*} and N:=3Δ~CsN:=3\tilde{\Delta}C^{\prime}s to F[𝐅x𝐅y𝐅z]F[\mathbf{F}_{x}\cup\mathbf{F}_{y}\cup\mathbf{F}_{z}], which is a subgraph of G(N,p)G(N,p). By a union bound over all at most O(t)O(t) such triples x,y,zV()x,y,z\in V(\mathcal{R}), we conclude that w.h.p. for each xy,yzE()xy,yz\in E(\mathcal{R}), we have F[𝐅x𝐅y𝐅z]k=12Δ𝒟N,pΔ(γ,ρ,εk,εk1,μ,η)F[\mathbf{F}_{x}\cup\mathbf{F}_{y}\cup\mathbf{F}_{z}]\in\cap_{k=1}^{2\Delta}\mathcal{D}^{\Delta}_{N,p}(\gamma,\rho,\varepsilon_{k},\varepsilon_{k-1},\mu,\eta), since

    O(t)sΘ(spΔ2)=O(n)eΘ(spΔ2logs)=O(n)eΘ(s2/Δlog(2Δ2)/Δs)=O(n)eω(logn)=o(1).O(t)s^{-\Theta(sp^{\Delta-2})}=O(n)e^{-\Theta(sp^{\Delta-2}\log{s})}=O(n)e^{-\Theta(s^{2/\Delta}\log^{(2\Delta-2)/\Delta}{s})}=O(n)e^{-\omega(\log{n})}=o(1).

    From now on we condition on this being the case.

    Similarly, for each xV()x\in V(\mathcal{R}), we apply Lemma A.5 with Δ:=Δ\Delta:=\Delta, ξ:=ξ\xi:=\xi and N:=(𝖽𝖾𝗀(x)+1)Δ~CsN:=\big{(}\mathsf{deg}_{\mathcal{R}}(x)+1\big{)}\tilde{\Delta}C^{\prime}s to F[yN(x)𝐅y𝐅x]F[\bigcup_{y\in N_{\mathcal{R}}(x)}\mathbf{F}_{y}\cup\mathbf{F}_{x}], which we can do since it is a subgraph of G(N,p)G(N,p). We next apply a union bound over all tt such vertices xV()x\in V(\mathcal{R}), to show that the probability that for some xV()x\in V(\mathcal{R}), we have F[yN(x)𝐅y𝐅x]k=1Δ𝒞N,pk(ξ)F[\bigcup_{y\in N_{\mathcal{R}}(x)}\mathbf{F}_{y}\cup\mathbf{F}_{x}]\notin\cap_{k=1}^{\Delta}\mathscr{C}_{N,p}^{k}(\xi) is at most

    te100log2(2Δ~Cs)ne100log2sne100log2(Ω(elogn))ne25logn=o(1).te^{-100\log^{2}(2\tilde{\Delta}C^{\prime}s)}\leq ne^{-100\log^{2}{s}}\leq ne^{-100\log^{2}(\Omega(e^{\sqrt{\log{n}}}))}\leq ne^{-25\log{n}}=o(1).

    From here on we condition on this event not occurring.

    Preparation of HH. We start by preparing HH for the embedding. For this, consider an embedding ψ:HKs\psi:H\rightarrow\mathcal{R}\boxtimes K_{s}. We partition V(H)V(H) into classes {Sx}xV()\{S_{x}\}_{x\in V(\mathcal{R})} with vSxv\in S_{x} for xV()x\in V(\mathcal{R}) if ψ(v)\psi(v) belongs to the copy of KsK_{s} in Ks\mathcal{R}\boxtimes K_{s} that corresponds to xx. We denote Hx:=H[Sx]H_{x}:=H[S_{x}]. Let qq be an arbitrary ”root” of \mathcal{R} and consider a breadth-first ordering of V()V(\mathcal{R}) starting from qq, namely q=x1,,xtq=x_{1},\ldots,x_{t}. We will embed SxS_{x} into 𝐅x\mathbf{F}_{x} for each x=x1,,xtx=x_{1},\ldots,x_{t} in this order. For each xV()x\in V(\mathcal{R}), let p1(x),,pdx(x)p_{1}(x),\ldots,p_{d_{x}}(x) be the ”parents” of xx, namely those of its neighbours that come before xx in the order x1,,xtx_{1},\ldots,x_{t}. Note that 0dxd0\leq d_{x}\leq d. With slight abuse of notation, denote Sp(x):=i=1dxSpi(x)S_{p(x)}:=\bigcup_{i=1}^{d_{x}}S_{p_{i}(x)}.

    Consider the third power H3H^{3} of HH, that is, uvE(H3)uv\in E(H^{3}) if and only if uvu\neq v and there is a path between uu and vv in HH consisting of at most 33 edges. Note that Δ(H3)Δ+Δ(Δ1)+Δ(Δ1)2=Δ3Δ2+Δ\Delta(H^{3})\leq\Delta+\Delta(\Delta-1)+\Delta(\Delta-1)^{2}=\Delta^{3}-\Delta^{2}+\Delta and therefore χ(H3)Δ3Δ2+Δ+1\chi(H^{3})\leq\Delta^{3}-\Delta^{2}+\Delta+1. Fix some xV()x\in V(\mathcal{R}) and let fxf_{x} be a Δ3Δ2+Δ+1\Delta^{3}-\Delta^{2}+\Delta+1-vertex colouring of H3[Sx]H^{3}[S_{x}]. Then fxf_{x} partitions SxS_{x} into Δ3Δ2+Δ+1\Delta^{3}-\Delta^{2}+\Delta+1 classes such that if two vertices u,vu,v are in the same class, they have distance at least 44 in HH, and so there are no edges between NH(u)N_{H}(u) and NH(v)N_{H}(v). Next, we refine this partition depending on the ”left-degrees” of the vertices. We say that u,vSxu,v\in S_{x} are equivalent if fx(u)=fx(v)f_{x}(u)=f_{x}(v) and

    |NH(u)({wSx:fx(w)<fx(u)}Sp(x))|=|NH(v)({wSx:fx(w)<fx(v)}Sp(x))|.|N_{H}(u)\cap\Big{(}\{w\in S_{x}:f_{x}(w)<f_{x}(u)\}\cup S_{p(x)}\Big{)}|=|N_{H}(v)\cap\Big{(}\{w\in S_{x}:f_{x}(w)<f_{x}(v)\}\cup S_{p(x)}\Big{)}|.

    Note that vertices of SxS_{x} will be embedded in order of increasing colour according to fxf_{x}. Thus, two vertices in SxS_{x} are equivalent if they have the same colour by fxf_{x} and the same number of neighbours that will be embedded before them. This equivalence relation partitions SxS_{x} into at most (Δ3Δ2+Δ+1)(Δ+1)=Δ~(\Delta^{3}-\Delta^{2}+\Delta+1)(\Delta+1)=\tilde{\Delta} many classes. Denote the partition classes by W1x,,WΔ~xW^{x}_{1},\ldots,W^{x}_{\tilde{\Delta}}, some of which may be empty, and let gx:Sx[Δ~]g_{x}:S_{x}\rightarrow[\tilde{\Delta}] be the corresponding partition function, that is, for wSxw\in S_{x}, we have gx(w)=jg_{x}(w)=j if and only if wWjxw\in W^{x}_{j}. Note that W1x,,WΔ~xW^{x}_{1},\ldots,W^{x}_{\tilde{\Delta}} are ordered in increasing order of fxf_{x}. Each WixW^{x}_{i} will be embedded into a different FyF_{y} with yJxy\in J_{x}, but for simplicity we now rename these vertex classes so that we can treat all of them at the same time.

    Set Q:=tΔ~Q:=t\tilde{\Delta}. Then let W1,,WQW_{1},\ldots,W_{Q} be obtained by concatenating the sequences (W1x1,,WΔ~x1),(W^{x_{1}}_{1},\ldots,W^{x_{1}}_{\tilde{\Delta}}), ,(W1xt,,WΔ~xt)\ldots,(W^{x_{t}}_{1},\ldots,W^{x_{t}}_{\tilde{\Delta}}) in this order, and again let g:V(H)[Q]g:V(H)\rightarrow[Q] be the corresponding partition function, that is, g(w)=jg(w)=j if and only if wWjw\in W_{j} for wV(H)w\in V(H). Thus, if g(u)=g(v)g(u)=g(v), then

    |NH(u){wV(H):g(w)<g(u)}|=|NH(v){wV(H):g(w)<g(v)}|.|N_{H}(u)\cap\{w\in V(H):g(w)<g(u)\}|=|N_{H}(v)\cap\{w\in V(H):g(w)<g(v)\}|.

    Let uV(H)u\in V(H) and g(u)\ell\leq g(u). Then we define the left-degree of uu with respect to gg and \ell as

    𝗅𝖽𝖾𝗀g(u):=|NH(u){wV(H):g(w)}|.\mathsf{ldeg}^{\ell}_{g}(u):=|N_{H}(u)\cap\{w\in V(H):g(w)\leq\ell\}|.

    Inductively embedding HH into FF. We proceed to embed HH into FF. For this, we first relabel the vertex classes {Fy}yV(J)\{F_{y}\}_{y\in V(J)} by ordering them as F1,,FQF_{1},\ldots,F_{Q}, where in the order x1,,xtx_{1},\ldots,x_{t} of V()V(\mathcal{R}) that we fixed above, for each xix_{i} we have that {Fy}yJxi\{F_{y}\}_{y\in J_{x_{i}}} are relabelled as F(i1)Δ~+1,,FiΔ~F_{(i-1)\tilde{\Delta}+1},\ldots,F_{i\tilde{\Delta}} (in arbitrary order). We will proceed by induction, embedding WiW_{i} into FiF_{i} for each i=1,,Qi=1,\ldots,Q. Note that even though JJ is not a complete graph, due to the fact that HH is a subgraph of Ks\mathcal{R}\boxtimes K_{s} and the way we picked W1,,WQW_{1},\ldots,W_{Q} and F1,,FQF_{1},\ldots,F_{Q}, we have that if uWiu\in W_{i} and vWjv\in W_{j} and uvE(H)uv\in E(H), then there is an (ε0,ρ,p)(\varepsilon_{0},\rho,p)-dense bipartite graph between FiF_{i} and FjF_{j}.

    Throughout our embedding, we make sure that the following statement (𝒮)(\mathcal{S}_{\ell}) holds for =0,1,,Q\ell=0,1,\ldots,Q:

    (𝒮)\displaystyle(\mathcal{S}_{\ell}) There exists a partial embedding ϕ of H[j=1Wj] into F[j=1Fj] such that for every\displaystyle\text{ There exists a partial embedding }\phi_{\ell}\text{ of }H[\cup_{j=1}^{\ell}W_{j}]\text{ into }F[\cup_{j=1}^{\ell}F_{j}]\text{ such that for every }
    zj=+1QWj there exists a candidate set C(z)V(F) given by\displaystyle z\in\cup_{j=\ell+1}^{Q}W_{j}\text{ there exists a candidate set }C_{\ell}(z)\subseteq V(F)\text{ given by }
    (a) C(z)={NF(ϕ(x)):xNH(z) and g(x)}Fg(z),\displaystyle\text{(a) }C_{\ell}(z)=\cap\{N_{F}(\phi_{\ell}(x)):x\in N_{H}(z)\text{ and }g(x)\leq\ell\}\cap F_{g(z)},
    satisfying
    (b) |C(z)|(ρp2)𝗅𝖽𝖾𝗀g(z)m, where m=|Fg(z)|=Cs, and\displaystyle\text{(b) }|C_{\ell}(z)|\geq\Big{(}\frac{\rho p}{2}\Big{)}^{\mathsf{ldeg}^{\ell}_{g}(z)}m\text{, where }m=|F_{g(z)}|=C^{\prime}s,\text{ and}
    (c) for every edge {z,z}E(H) with g(z),g(z)> the pair (C(z),C(z)) is\displaystyle\text{(c) for every edge }\{z,z^{\prime}\}\in E(H)\text{ with }g(z),g(z^{\prime})>\ell\text{ the pair }(C_{\ell}(z),C_{\ell}(z^{\prime}))\text{ is }
    (ε𝗅𝖽𝖾𝗀g(z)+𝗅𝖽𝖾𝗀g(z),ρ,p)-dense in F.\displaystyle\hskip 17.07182pt(\varepsilon_{\mathsf{ldeg}_{g}^{\ell}(z)+\mathsf{ldeg}_{g}^{\ell}(z^{\prime})},\rho,p)\text{-dense in }F.

    The inductive statement (𝒮)(\mathcal{S}_{\ell}) tells us that the first \ell classes W1,,WW_{1},\ldots,W_{\ell} of HH can be embedded into F1FF_{1}\cup\ldots\cup F_{\ell} in such a way that for every not yet embedded vertex zW+1WQz\in W_{\ell+1}\cup\ldots\cup W_{Q}, there is a large enough candidate set C(z)C_{\ell}(z) that consists of the intersection of Fg(z)F_{g(z)} and the neighbourhoods of all already embedded neighbours of zz. The third condition of (𝒮)(\mathcal{S}_{\ell}) additionally ensures that all pairs of candidate sets inherit regularity.

    Note that from (𝒮Q)(\mathcal{S}_{Q}) it follows that HH can be embedded into FF, which means that showing that (𝒮0),,(𝒮Q)(\mathcal{S}_{0}),\ldots,(\mathcal{S}_{Q}) hold completes the proof of Lemma 3.7. We next prove that (𝒮)(\mathcal{S}_{\ell}) holds for =0,1,,Q\ell=0,1,\ldots,Q by induction.

    Basis of the induction: =0\ell=0. We then have that ϕ0\phi_{0} is the empty embedding and C0(z)=Fg(z)C_{0}(z)=F_{g(z)} for every zV(H)z\in V(H) by (a). This implies that property (b) also holds, as |C0(z)|=m|C_{0}(z)|=m and 𝗅𝖽𝖾𝗀g0(z)=0\mathsf{ldeg}^{0}_{g}(z)=0. Property (c) holds since, as mentioned above, for each zzE(H)zz^{\prime}\in E(H), there is an (ε0,ρ,p)(\varepsilon_{0},\rho,p)-dense graph between Fg(z)=C0(z)F_{g(z)}=C_{0}(z) and Fg(z)=C0(z)F_{g(z^{\prime})}=C_{0}(z^{\prime}).

    Induction step: +1\ell\rightarrow\ell+1. Suppose that (𝒮)(\mathcal{S}_{\ell}) holds for some <Q\ell<Q. We will now extend ϕ\phi_{\ell} to an embedding ϕ+1\phi_{\ell+1} that satisfies (a), (b), and (c) to show that (𝒮+1)(\mathcal{S}_{\ell+1}) holds. We do this in the following way. Firstly, for each yW+1y\in W_{\ell+1}, we identify a subset C(y)C(y)C(y)\subseteq C_{\ell}(y) such that if yy is embedded into some vertex from C(y)C(y), then for every ”right-neighbour” zz of yy in HH, the candidate set C+1(z):=C(z)NF(ϕ+1(y))C_{\ell+1}(z):=C_{\ell}(z)\cap N_{F}(\phi_{\ell+1}(y)) will be such that properties (b) and (c) will still hold.

    However, since if 𝗅𝖽𝖾𝗀g1\mathsf{ldeg}^{\ell}_{g}\geq 1, we have |C(y)||C(y)|=o(s)|W+1||C(y)|\leq|C_{\ell}(y)|=o(s)\ll|W_{\ell+1}|, we cannot greedily pick an embedding of each yy into C(y)C(y), as this way some C(y)C(y) may be entirely occupied by other embeddings before we get to embed it. Thus, in a second step we will use Hall’s condition to find distinct representatives for {C(y):yW+1}\{C(y):y\in W_{\ell+1}\}, and then we will set ϕ+1(y)\phi_{\ell+1}(y) to be the representative of C(y)C(y). We now describe these two steps in detail.

    Let yW+1y\in W_{\ell+1}, and set

    NH+1(y):={zNH(y):g(z)>+1}.N_{H}^{\ell+1}(y):=\{z\in N_{H}(y):g(z)>\ell+1\}.

    We say that vC(y)v\in C_{\ell}(y) is bad (that is, it will not be in C(y)C(y)) if there is some vertex zNH+1(y)z\in N_{H}^{\ell+1}(y) such that the set NF(v)C(z)N_{F}(v)\cap C_{\ell}(z) does not satisfy condition (b) or (c) of (𝒮+1)(\mathcal{S}_{\ell+1}) and therefore cannot become C+1(z)C_{\ell+1}(z).

    We start by focusing on (b) of (𝒮+1)(\mathcal{S}_{\ell+1}). Let zNH+1(y)z\in N_{H}^{\ell+1}(y). We know by (c) of (𝒮)(\mathcal{S}_{\ell}) that (C(y),C(z))(C_{\ell}(y),C_{\ell}(z)) is an (ε𝗅𝖽𝖾𝗀g(y)+𝗅𝖽𝖾𝗀g(z),ρ,p)(\varepsilon_{\mathsf{ldeg}_{g}^{\ell}(y)+\mathsf{ldeg}_{g}^{\ell}(z)},\rho,p)-dense pair. Thus, there are at most ε𝗅𝖽𝖾𝗀g(y)+𝗅𝖽𝖾𝗀g(z)|C(y)|ε2Δ|C(y)|\varepsilon_{\mathsf{ldeg}_{g}^{\ell}(y)+\mathsf{ldeg}_{g}^{\ell}(z)}|C_{\ell}(y)|\leq\varepsilon_{2\Delta}|C_{\ell}(y)| vertices vv in C(y)C_{\ell}(y) such that

    |NF(v)C(z)|<(ρε𝗅𝖽𝖾𝗀g(y)+𝗅𝖽𝖾𝗀g(z))p|C(z)|.|N_{F}(v)\cap C_{\ell}(z)|<\Big{(}\rho-\varepsilon_{\mathsf{ldeg}_{g}^{\ell}(y)+\mathsf{ldeg}_{g}^{\ell}(z)}\Big{)}p|C_{\ell}(z)|.

    Considering all zNH+1(y)z\in N_{H}^{\ell+1}(y) in the same way, it follows that for all but at most Δε2Δ|C(y)|\Delta\varepsilon_{2\Delta}|C_{\ell}(y)| vertices vC(y)v\in C_{\ell}(y), the following holds for all zNH+1(y)z\in N_{H}^{\ell+1}(y):

    |NF(v)C(z)|(ρε2Δ)p|C(z)||N_{F}(v)\cap C_{\ell}(z)|\geq\Big{(}\rho-\varepsilon_{2\Delta}\Big{)}p|C_{\ell}(z)|
    (ρε2Δ)p(ρp2)𝗅𝖽𝖾𝗀g(z)|Fg(z)|(ρp2)𝗅𝖽𝖾𝗀g+1(z)|Fg(z)|,\geq\Big{(}\rho-\varepsilon_{2\Delta}\Big{)}p\Big{(}\frac{\rho p}{2}\Big{)}^{\mathsf{ldeg}^{\ell}_{g}(z)}|F_{g(z)}|\geq\Big{(}\frac{\rho p}{2}\Big{)}^{\mathsf{ldeg}_{g}^{\ell+1}(z)}|F_{g(z)}|,

    where for the penultimate inequality we used condition (b) of (𝒮)(\mathcal{S}_{\ell}) and the final inequality follows from our choice of εε2Δ\varepsilon^{*}\geq\varepsilon_{2\Delta}.

    Next, we consider property (c) of (𝒮+1)(\mathcal{S}_{\ell+1}). Let e={z,z}e=\{z,z^{\prime}\} with g(z),g(z)>+1g(z),g(z^{\prime})>\ell+1 and with at least one of z,zz,z^{\prime} in NH+1(y)N_{H}^{\ell+1}(y). The number of these edges is at most Δ(Δ1)<Δ2\Delta(\Delta-1)<\Delta^{2}. Let y~,z~,z~V()\tilde{y},\tilde{z},\tilde{z}^{\prime}\in V(\mathcal{R}) be such that Fg(y)𝐅y~,Fg(z)𝐅z~,Fg(z)𝐅z~F_{g(y)}\subseteq\mathbf{F}_{\tilde{y}},F_{g(z)}\subseteq\mathbf{F}_{\tilde{z}},F_{g(z^{\prime})}\subseteq\mathbf{F}_{\tilde{z}^{\prime}}, and let j:=𝗅𝖽𝖾𝗀g(z)+𝗅𝖽𝖾𝗀g(z)j:=\mathsf{ldeg}^{\ell}_{g}(z)+\mathsf{ldeg}^{\ell}_{g}(z^{\prime}). If z,zNH(y)z,z^{\prime}\in N_{H}^{\ell}(y), we have

    max{𝗅𝖽𝖾𝗀g(y),𝗅𝖽𝖾𝗀g(z),𝗅𝖽𝖾𝗀g(z)}Δ2,\max\{\mathsf{ldeg}_{g}^{\ell}(y),\mathsf{ldeg}_{g}^{\ell}(z),\mathsf{ldeg}_{g}^{\ell}(z^{\prime})\}\leq\Delta-2,

    because each of y,z,zy,z,z^{\prime} has at least two neighbours in W+1WQW_{\ell+1}\cup\ldots\cup W_{Q}. Property (b) of (𝒮)(\mathcal{S}_{\ell}) then implies

    min{|C(y),C(z),C(z)|}(ρp2)max{𝗅𝖽𝖾𝗀g(y),𝗅𝖽𝖾𝗀g(z),𝗅𝖽𝖾𝗀g(z)}CsγpΔ23Δ~Cs,\min\{|C_{\ell}(y),C_{\ell}(z),C_{\ell}(z^{\prime})|\}\geq\Big{(}\frac{\rho p}{2}\Big{)}^{\max\{\mathsf{ldeg}_{g}^{\ell}(y),\mathsf{ldeg}_{g}^{\ell}(z),\mathsf{ldeg}_{g}^{\ell}(z^{\prime})\}}C^{\prime}s\geq\gamma p^{\Delta-2}3\tilde{\Delta}C^{\prime}s,

    by our choice of γ\gamma. Since either y~=z~\tilde{y}=\tilde{z} or y~z~E()\tilde{y}\tilde{z}\in E(\mathcal{R}) (and the same holds for z~,z~\tilde{z},\tilde{z}^{\prime} and y~,z~\tilde{y},\tilde{z}^{\prime}), we know that F[𝐅y~𝐅z~𝐅z~]F[\mathbf{F}_{\tilde{y}}\cup\mathbf{F}_{\tilde{z}}\cup\mathbf{F}_{\tilde{z}^{\prime}}] is a subgraph of some graph in 𝒟3Δ~Cs,pΔ(γ,ρ,εj+1,εj,μ,η)\mathcal{D}_{3\tilde{\Delta}C^{\prime}s,p}^{\Delta}(\gamma,\rho,\varepsilon_{j+1},\varepsilon_{j},\mu,\eta). Thus, there are at most μ|C(y)|\mu|C_{\ell}(y)| vertices vv in C(y)C_{\ell}(y) such that (NF(v)C(z),NF(v)C(z))\big{(}N_{F}(v)\cap C_{\ell}(z),N_{F}(v)\cap C_{\ell}(z^{\prime})\big{)} is not (εj+1,ρ,p)(\varepsilon_{j+1},\rho,p)-dense.

    Suppose now that only one of z,zz,z^{\prime} is in NH+1(y)N_{H}^{\ell+1}(y), say zNH+1(y)z\in N_{H}^{\ell+1}(y) and zNH+1(y)z^{\prime}\notin N_{H}^{\ell+1}(y). Then we have

    max{𝗅𝖽𝖾𝗀g(y),𝗅𝖽𝖾𝗀g(z)}Δ1 and 𝗅𝖽𝖾𝗀g(z)Δ2.\max\{\mathsf{ldeg}^{\ell}_{g}(y),\mathsf{ldeg}^{\ell}_{g}(z^{\prime})\}\leq\Delta-1\text{ and }\mathsf{ldeg}_{g}^{\ell}(z)\leq\Delta-2.

    Therefore, analogously to above,

    min{|C(y)|,|C(z)|}γpΔ13Δ~Cs and |C(z)|γpΔ23Δ~Cs.\min\{|C_{\ell}(y)|,|C_{\ell}(z^{\prime})|\}\geq\gamma p^{\Delta-1}3\tilde{\Delta}C^{\prime}s\text{ and }|C_{\ell}(z)|\geq\gamma p^{\Delta-2}3\tilde{\Delta}C^{\prime}s.

    Again, since either y~=z~\tilde{y}=\tilde{z} or y~z~E()\tilde{y}\tilde{z}\in E(\mathcal{R}) (and the same holds for z~,z~\tilde{z},\tilde{z}^{\prime}), F[𝐅y~𝐅z~𝐅z~]F[\mathbf{F}_{\tilde{y}}\cup\mathbf{F}_{\tilde{z}}\cup\mathbf{F}_{\tilde{z}^{\prime}}] is a subgraph of some graph in 𝒟3Δ~Cs,pΔ(γ,ρ,εj+1,εj,μ,η)\mathcal{D}_{3\tilde{\Delta}C^{\prime}s,p}^{\Delta}(\gamma,\rho,\varepsilon_{j+1},\varepsilon_{j},\mu,\eta). Thus, there are at most μ|C(y)|\mu|C_{\ell}(y)| vertices vC(y)v\in C_{\ell}(y) such that (NF(v)C(z),C(z))\big{(}N_{F}(v)\cap C_{\ell}(z),C_{\ell}(z^{\prime})\big{)} is not (εj+1,ρ,p)(\varepsilon_{j+1},\rho,p)-dense.

    Note that in both cases, 𝗅𝖽𝖾𝗀g+1(z)+𝗅𝖽𝖾𝗀g+1(z)𝗅𝖽𝖾𝗀g(z)+𝗅𝖽𝖾𝗀g(z)+1=j+1\mathsf{ldeg}^{\ell+1}_{g}(z)+\mathsf{ldeg}^{\ell+1}_{g}(z^{\prime})\geq\mathsf{ldeg}^{\ell}_{g}(z)+\mathsf{ldeg}^{\ell}_{g}(z^{\prime})+1=j+1, and so any pair that is (εj+1,ρ,p)(\varepsilon_{j+1},\rho,p)-dense is also (ε𝗅𝖽𝖾𝗀g+1(z)+𝗅𝖽𝖾𝗀g+1(z),ρ,p)(\varepsilon_{\mathsf{ldeg}^{\ell+1}_{g}(z)+\mathsf{ldeg}^{\ell+1}_{g}(z^{\prime})},\rho,p)-dense. For some fixed vC(y)v\in C_{\ell}(y), set C^(z)=C(z)NF(v)\hat{C}_{\ell}(z)=C_{\ell}(z)\cap N_{F}(v) if zNH+1(y)z\in N_{H}^{\ell+1}(y) and C^(z)=C(z)\hat{C}_{\ell}(z)=C_{\ell}(z) if zNH+1(y)z\notin N_{H}^{\ell+1}(y), and define C^(z)\hat{C}_{\ell}(z^{\prime}) in the same way.

    What we have shown so far is that there are at least

    (1Δε2ΔΔ2μ)|C(y)|(1-\Delta\varepsilon_{2\Delta}-\Delta^{2}\mu)|C_{\ell}(y)|

    vertices vC(y)v\in C_{\ell}(y) with the properties

    (b’) |NH(v)C(z)|(ρp2)𝗅𝖽𝖾𝗀g+1(z)|Fg(z)| for each zNH+1(y) and\displaystyle|N_{H}(v)\cap C_{\ell}(z)|\geq\Big{(}\frac{\rho p}{2}\Big{)}^{\mathsf{ldeg}_{g}^{\ell+1}(z)}|F_{g(z)}|\text{ for each }z\in N_{H}^{\ell+1}(y)\text{ and}
    (c’) (C^(z),C^(z)) is (ε𝗅𝖽𝖾𝗀g+1(z)+𝗅𝖽𝖾𝗀g+1(z),ρ,p)-dense for all edges {z,z} of H with\displaystyle\big{(}\hat{C}_{\ell}(z),\hat{C}_{\ell}(z^{\prime})\big{)}\text{ is }(\varepsilon_{\mathsf{ldeg}^{\ell+1}_{g}(z)+\mathsf{ldeg}^{\ell+1}_{g}(z^{\prime})},\rho,p)\text{-dense for all edges }\{z,z^{\prime}\}\text{ of }H\text{ with }
    g(z),g(z)>+1 and {z,z}NH+1(y).\displaystyle g(z),g(z^{\prime})>\ell+1\text{ and }\{z,z^{\prime}\}\cap N_{H}^{\ell+1}(y)\neq\varnothing.

    Denote the vertices in C(y)C_{\ell}(y) that satisfy (b’) and (c’) by C(y)C(y). For all y,yW+1y,y^{\prime}\in W_{\ell+1}, since g(y)=g(y)=g(y)=g(y^{\prime})=\ell, and by our choice of the partition W1,,WQW_{1},\ldots,W_{Q}, we have that 𝗅𝖽𝖾𝗀g(y)=𝗅𝖽𝖾𝗀g(y)\mathsf{ldeg}_{g}^{\ell}(y)=\mathsf{ldeg}_{g}^{\ell}(y^{\prime}). Let

    k=𝗅𝖽𝖾𝗀g(y) for some yW+1k=\mathsf{ldeg}_{g}^{\ell}(y)\text{ for some }y\in W_{\ell+1}

    and note that 0kΔ0\leq k\leq\Delta. Then by our choice of μ\mu and ε2Δε\varepsilon_{2\Delta}\leq\varepsilon^{*} and from condition (b) of (𝒮)(\mathcal{S}_{\ell}), we have

    |C(y)|(1Δε2ΔΔ2μ)|C(y)|(1Δε2ΔΔ2μ)(ρp2)kCsρk2k+1pkCs.|C(y)|\geq(1-\Delta\varepsilon_{2\Delta}-\Delta^{2}\mu)|C_{\ell}(y)|\geq(1-\Delta\varepsilon_{2\Delta}-\Delta^{2}\mu)\Big{(}\frac{\rho p}{2}\Big{)}^{k}C^{\prime}s\geq\frac{\rho^{k}}{2^{k+1}}p^{k}C^{\prime}s.

    This finishes the first part of our inductive step.

    Next, we move to the second part, in which we show that distinct representatives for the system (C(y))yW+1\big{(}C(y)\big{)}_{y\in W_{\ell+1}} exist. For this, we use Hall’s condition [9], which tells us that in a bipartite graph with bipartitions XX and YY, there is a perfect matching if for every XXX^{\prime}\subseteq X, it holds that |N(X)||X||N(X)|\geq|X|. Thus, in our case it is enough to show that for every YW+1Y\subseteq W_{\ell+1}, it holds that

    |Y||yYC(y)|.|Y|\leq\Big{|}\bigcup_{y\in Y}C(y)\Big{|}.

    In case 1|Y|ρkpkCs/2k+11\leq|Y|\leq\rho^{k}p^{k}C^{\prime}s/2^{k+1}, the above holds due to our lower bound on each |C(y)||C(y)|.

    Now let YW+1Y\subseteq W_{\ell+1} be a set with |Y|>ρkpkCs/2k+1|Y|>\rho^{k}p^{k}C^{\prime}s/2^{k+1}. As mentioned above, for each yW+1y\in W_{\ell+1}, we have 𝗅𝖽𝖾𝗀g(y)=k\mathsf{ldeg}_{g}^{\ell}(y)=k, so there is a kk-tuple K(y)={u1(y),,uk(y)}=NH(y)NH+1(y)K(y)=\{u_{1}(y),\ldots,u_{k}(y)\}=N_{H}(y)\setminus N_{H}^{\ell+1}(y) of neighbours of yy that are already embedded. Moreover, since we chose the partition W1,,WQW_{1},\ldots,W_{Q} in such a way that the distance in HH between any two vertices in the same class WiW_{i} is at least 44, we have that for every y,yW+1y,y^{\prime}\in W_{\ell+1}, the sets K(y)K(y) and K(y)K(y^{\prime}) are disjoint. Therefore, the sets of already embedded vertices ϕ(K(y))\phi_{\ell}(K(y)) and ϕ(K(y))\phi_{\ell}(K(y^{\prime})) are disjoint too, and so k={ϕ(K(y)):yY}(V(F)k)\mathcal{F}_{k}=\{\phi_{\ell}(K(y)):y\in Y\}\subseteq\binom{V(F)}{k} is a family of disjoint sets of size kk in V(F)V(F). Furthermore,

    C(y)vϕ(K(y))NF(v).C(y)\subseteq\bigcap_{v\in\phi(K(y))}N_{F}(v).

    Set

    U=yYC(y)F+1,U=\bigcup_{y\in Y}C(y)\subseteq F_{\ell+1},

    and observe that if F+1𝐅y~F_{\ell+1}\subseteq\mathbf{F}_{\tilde{y}} for some y~V()\tilde{y}\in V(\mathcal{R}), then U𝐅y~U\subseteq\mathbf{F}_{\tilde{y}} and

    k((i=1dy~𝐅pi(y~)˙𝐅y~)Uk).\mathcal{F}_{k}\subseteq\binom{(\bigcup^{d_{\tilde{y}}}_{i=1}\mathbf{F}_{p_{i}(\tilde{y})}\dot{\cup}\mathbf{F}_{\tilde{y}})\setminus U}{k}.

    Assume for contradiction that

    |U|<|Y|=|k|.|U|<|Y|=|\mathcal{F}_{k}|.

    Since we conditioned on 𝒢:=F[n~N(y~)𝐅n~˙𝐅y~]𝒞(𝖽𝖾𝗀(y~)+1)Δ~Cs,pk(ξ)\mathscr{G}:=F[\bigcup_{\tilde{n}\in N_{\mathcal{R}}(\tilde{y})}\mathbf{F}_{\tilde{n}}\dot{\cup}\mathbf{F}_{\tilde{y}}]\in\mathscr{C}^{k}_{(\mathsf{deg}_{\mathcal{R}}(\tilde{y})+1)\tilde{\Delta}C^{\prime}s,p}(\xi), and we have |k||W+1|sξ(𝖽𝖾𝗀(y~)+1)Δ~Cs|\mathcal{F}_{k}|\leq|W_{\ell+1}|\leq s\leq\xi(\mathsf{deg}_{\mathcal{R}}(\tilde{y})+1)\tilde{\Delta}C^{\prime}s, it follows that

    eΓ(k,𝒢)(k,U)pk|k||U|+6ξ(𝖽𝖾𝗀(y~)+1)Δ~Cspk|k|.e_{\Gamma(k,\mathscr{G})}(\mathcal{F}_{k},U)\leq p^{k}|\mathcal{F}_{k}||U|+6\xi(\mathsf{deg}_{\mathcal{R}}(\tilde{y})+1)\tilde{\Delta}C^{\prime}sp^{k}|\mathcal{F}_{k}|.

    From our lower bound on |C(y)||C(y)| above, it follows that

    eΓ(k,𝒢)(k,U)ρk2k+1pkCs|k|.e_{\Gamma(k,\mathscr{G})}(\mathcal{F}_{k},U)\geq\frac{\rho^{k}}{2^{k+1}}p^{k}C^{\prime}s|\mathcal{F}_{k}|.

    The last two inequalities together with CΔ,ρ1,dC^{\prime}\gg\Delta,\rho^{-1},d imply that

    |yYC(y)|=|U|(ρk2k+16(𝖽𝖾𝗀(y~)+1)ξΔ~)Css|W+1||Y|,\Big{|}\bigcup_{y\in Y}C(y)\Big{|}=|U|\geq\Big{(}\frac{\rho^{k}}{2^{k+1}}-6(\mathsf{deg}_{\mathcal{R}}(\tilde{y})+1)\xi\tilde{\Delta}\Big{)}C^{\prime}s\geq s\geq|W_{\ell+1}|\geq|Y|,

    which contradicts our assumption that |U|<|Y|=|k||U|<|Y|=|\mathcal{F}_{k}|. This shows that Hall’s condition |Y||U||Y|\leq|U| holds, as desired. Therefore, there is a system of representatives for (C(y))yW+1\big{(}C(y)\big{)}_{y\in W_{\ell+1}}. In other words, there is an injective function ψ:W+1yW+1C(y)\psi:W_{\ell+1}\rightarrow\cup_{y\in W_{\ell+1}}C(y) such that ψ(y)C(y)\psi(y)\in C(y) for each yW+1y\in W_{\ell+1}.

    We can now extend ϕ\phi_{\ell} to get ϕ+1\phi_{\ell+1} and define C+1(z)C_{\ell+1}(z) for zj=+2Qz\in\cup_{j=\ell+2}^{Q}. Firstly, let

    ϕ+1(w)={ϕ(w),if wj=1Wj,ψ(w),if wW+1.\phi_{\ell+1}(w)=\begin{cases}\phi_{\ell}(w),&\text{if }w\in\cup_{j=1}^{\ell}W_{j},\\ \psi(w),&\text{if }w\in W_{\ell+1}.\end{cases}

    Since all pairs of vertices in W+1W_{\ell+1} are at distance at least 44, each zj=+2QWjz\in\cup_{j=\ell+2}^{Q}W_{j} has no more than one neighbour in W+1W_{\ell+1}. Thus, for each such zj=+2QWjz\in\cup_{j=\ell+2}^{Q}W_{j} we can set

    C+1(z)={C(z),if NH(z)W+1=,C(z)NF(ϕ+1(y)),if NH(z)W+1={y}.C_{\ell+1}(z)=\begin{cases}C_{\ell}(z),&\text{if }N_{H}(z)\cap W_{\ell+1}=\varnothing,\\ C_{\ell}(z)\cap N_{F}(\phi_{\ell+1}(y)),&\text{if }N_{H}(z)\cap W_{\ell+1}=\{y\}.\end{cases}

    We now verify that ϕ+1\phi_{\ell+1} and C+1(z)C_{\ell+1}(z) for each zj=+2QWjz\in\cup_{j=\ell+2}^{Q}W_{j} satisfy (𝒮+1)(\mathcal{S}_{\ell+1}).

    From property (a) of (𝒮)(\mathcal{S}_{\ell}) together with ϕ+1(y)C(y)C(y)\phi_{\ell+1}(y)\in C(y)\subseteq C_{\ell}(y) for every yW+1y\in W_{\ell+1} and the fact that ψ\psi is injective, it follows that ϕ+1\phi_{\ell+1} is a partial embedding of H[j=1+1Wj]H[\bigcup_{j=1}^{\ell+1}W_{j}] into F[j=1+1Fj]F[\bigcup_{j=1}^{\ell+1}F_{j}].

    We now check that parts (a) and (b) of (𝒮+1)(\mathcal{S}_{\ell+1}) hold. Fix some zj=+2QWjz\in\bigcup_{j=\ell+2}^{Q}W_{j}. In case NH(z)W+1=N_{H}(z)\cap W_{\ell+1}=\varnothing, we have C+1(z)=C(z)C_{\ell+1}(z)=C_{\ell}(z) and 𝗅𝖽𝖾𝗀g+1(z)=𝗅𝖽𝖾𝗀g(z)\mathsf{ldeg}_{g}^{\ell+1}(z)=\mathsf{ldeg}_{g}^{\ell}(z), so (a) and (b) of (𝒮+1)(\mathcal{S}_{\ell+1}) are satisfied for that zz. Now suppose NH(z)W+1={y}N_{H}(z)\cap W_{\ell+1}=\{y\} (recall that this is the only other possible case). Since we defined C+1(z)=C(z)NF(ϕ+1(y))C_{\ell+1}(z)=C_{\ell}(z)\cap N_{F}(\phi_{\ell+1}(y)), property (a) of (𝒮+1)(\mathcal{S}_{\ell+1}) holds for zz in this case. Furthermore, by our choice ϕ+1(y)C(y)\phi_{\ell+1}(y)\in C(y) and by (b’), it follows that property (b) of (𝒮+1)(\mathcal{S}_{\ell+1}) is satisfied.

    Lastly, we check property (c) of (𝒮+1)(\mathcal{S}_{\ell+1}). Fix an edge zzzz^{\prime} of HH such that z,zj=+2QWjz,z^{\prime}\in\bigcup_{j=\ell+2}^{Q}W_{j}. There are three cases to consider, based on the size of NH(z)W+1N_{H}(z)\cap W_{\ell+1} and NH(z)W+1N_{H}(z^{\prime})\cap W_{\ell+1} (recall that each of them has size either 0 or 11). Suppose first NH(z)W+1=N_{H}(z)\cap W_{\ell+1}=\varnothing and NH(z)W+1=N_{H}(z^{\prime})\cap W_{\ell+1}=\varnothing. Then we have 𝗅𝖽𝖾𝗀g(z)=𝗅𝖽𝖾𝗀g+1(z)\mathsf{ldeg}^{\ell}_{g}(z)=\mathsf{ldeg}^{\ell+1}_{g}(z) and 𝗅𝖽𝖾𝗀g(z)=𝗅𝖽𝖾𝗀g+1(z)\mathsf{ldeg}^{\ell}_{g}(z^{\prime})=\mathsf{ldeg}^{\ell+1}_{g}(z^{\prime}), as well as C(z)=C+1(z)C_{\ell}(z)=C_{\ell+1}(z) and C(z)=C(z)C_{\ell}(z^{\prime})=C_{\ell}(z^{\prime}), so part (c) of (𝒮+1)(\mathcal{S}_{\ell+1}) follows directly from part (c) of (𝒮)(\mathcal{S}_{\ell}). Next, suppose NH(z)W+1={y}N_{H}(z)\cap W_{\ell+1}=\{y\} and NH(z)W+1=N_{H}(z^{\prime})\cap W_{\ell+1}=\varnothing. Then property (c’) together with the definition of C+1(z)C_{\ell+1}(z) and C+1(z)C_{\ell+1}(z^{\prime}) imply part (c) of (𝒮+1)(\mathcal{S}_{\ell+1}). Finally, suppose NH(z)W+1={y}N_{H}(z)\cap W_{\ell+1}=\{y\} and NH(z)W+1={y}N_{H}(z^{\prime})\cap W_{\ell+1}=\{y^{\prime}\}. Then it must be the case that y=yy=y^{\prime}, since otherwise yy and yy^{\prime} would be two distinct vertices in W+1W_{\ell+1} connected by a path of length 33 in HH, namely the path yzzyyzz^{\prime}y^{\prime}, and this is impossible by our choice of W1,,WQW_{1},\ldots,W_{Q}. Then in this case also, part (c) of (𝒮+1)(\mathcal{S}_{\ell+1}) follows from (c’) and our choice of C+1(z)C_{\ell+1}(z) and C+1(z)C_{\ell+1}(z^{\prime}).

    We have shown that (a), (b), and (c) of (𝒮+1)(\mathcal{S}_{\ell+1}) hold. This finishes the induction step and the proof of Lemma 3.7. ∎

References

  • [1] Peter Allen and Julia Böttcher. Partition universality for graphs of bounded degeneracy and degree. arXiv preprint arXiv:2211.15819, 2022.
  • [2] Noga Alon, Paul D. Seymour, and Robin Thomas. A separator theorem for nonplanar graphs. Journal of the American Mathematical Society, 3(4):801–808, 1990.
  • [3] József Beck. On size Ramsey number of paths, trees, and circuits. I. Journal of Graph Theory, 7(1):115–129, 1983.
  • [4] Sören Berger, Yoshiharu Kohayakawa, Giulia Satiko Maesaka, Taísa Martins, Walner Mendonça, Guilherme Oliveira Mota, and Olaf Parczyk. The size-Ramsey number of powers of bounded degree trees. Journal of the London Mathematical Society, 103(4):1314–1332, 2021.
  • [5] Sandeep N. Bhatt, Fan R. K. Chung, Frank Thomson Leighton, and Arnold L. Rosenberg. Universal graphs for bounded-degree trees and planar graphs. SIAM Journal on Discrete Mathematics, 2(2):145–155, 1989.
  • [6] Fan R.K. Chung. Separator theorems and their applications. Paths, Flows, and VLSI-Layout, 1990.
  • [7] Václav Chvatál, Vojtech Rödl, Endre Szemerédi, and William Thomas Trotter Jr. The Ramsey number of a graph with bounded maximum degree. Journal of Combinatorial Theory, Series B, 34(3):239–243, 1983.
  • [8] David Conlon, Rajko Nenadov, and Miloš Trujić. On the size-Ramsey number of grids. Combinatorics, Probability and Computing, pages 1–7, 2022.
  • [9] Reinhard Diestel. Matching Covering and Packing, pages 35–58. Springer Berlin Heidelberg, Berlin, Heidelberg, 2017.
  • [10] Guoli Ding and Bogdan Oporowski. Some results on tree decompositions of graphs. Journal of Graph Theory, 20(4):481–499, 1995.
  • [11] Marc Distel and David R Wood. Tree-partitions with bounded degree trees. arXiv preprint arXiv:2210.12577, 2022.
  • [12] Nemanja Draganić, Michael Krivelevich, and Rajko Nenadov. The size-Ramsey number of short subdivisions. Random Structures & Algorithms, 59(1):68–78, 2021.
  • [13] Nemanja Draganić, Michael Krivelevich, and Rajko Nenadov. Rolling backwards can move you forward: on embedding problems in sparse expanders. Trans. Am. Math. Soc., 375(7):5195–5216, 2022.
  • [14] Nemanja Draganić and Kalina Petrova. Size-Ramsey numbers of graphs with maximum degree three. arXiv preprint arXiv:2207.05048, 2022.
  • [15] Zdeněk Dvořák and Sergey Norin. Treewidth of graphs with balanced separations. Journal of Combinatorial Theory, Series B, 137:137–144, 2019.
  • [16] Paul Erdős, Ralph J. Faudrée, Cecil C. Rousseau, and Richard H. Schelp. The size Ramsey number. Period. Math. Hungar., 9(1-2):145–161, 1978.
  • [17] Joel Friedman and Nicholas Pippenger. Expanding graphs contain all small trees. Combinatorica, 7:71–76, 1987.
  • [18] Stefanie Gerke, Yoshiharu Kohayakawa, Vojtěch Rödl, and Angelika Steger. Small subsets inherit sparse ε\varepsilon-regularity. Journal of Combinatorial Theory, Series B, 97(1):34–56, 2007.
  • [19] Stefanie Gerke and Angelika Steger. The sparse regularity lemma and its applications. In BCC, pages 227–258, 2005.
  • [20] Charles H. Goldberg and Douglas B. West. Bisection of circle colorings. SIAM Journal on Algebraic Discrete Methods, 6(1):93–106, 1985.
  • [21] Nina Kamčev, Anita Liebenau, David R. Wood, and Liana Yepremyan. The size Ramsey number of graphs with bounded treewidth. SIAM Journal on Discrete Mathematics, 35(1):281–293, 2021.
  • [22] Yoshiharu Kohayakawa, Troy Retter, and Vojtěch Rödl. The size Ramsey number of short subdivisions of bounded degree graphs. Random Structures & Algorithms, 54(2):304–339, 2019.
  • [23] Yoshiharu Kohayakawa, Vojtěch Rödl, Mathias Schacht, and Endre Szemerédi. Sparse partition universal graphs for graphs of bounded degree. Advances in Mathematics, 226(6):5041–5065, 2011.
  • [24] János Komlós, Ali Shokoufandeh, Miklós Simonovits, and Endre Szemerédi. The regularity lemma and its applications in graph theory. Theoretical Aspects of Computer Science: Advanced Lectures, pages 84–112, 2002.
  • [25] Choongbum Lee. Ramsey numbers of degenerate graphs. Annals of Mathematics, 185(3):791–829, 2017.
  • [26] Frank P. Ramsey. On a problem of formal logic. Proc. London Math. Soc, 30(4):264–286, 1929.
  • [27] Neil Robertson and Paul D. Seymour. Graph minors. II. Algorithmic aspects of tree-width. Journal of Algorithms, 7(3):309–322, 1986.
  • [28] Vojtěch Rödl and Endre Szemerédi. On size Ramsey numbers of graphs with bounded degree. Combinatorica, 20:257–262, 02 2000.
  • [29] Konstantin Tikhomirov. On bounded degree graphs with large size-Ramsey numbers. arXiv preprint arXiv:2210.05818, 2022.
  • [30] David R. Wood. On tree-partition-width. European Journal of Combinatorics, 30(5):1245–1253, 2009.
  • [31] David R. Wood. Presentation at the open problem session of the Third Southwestern German Workshop on Graph Theory, University of Heidelberg, 2022.