Hostname: page-component-8448b6f56d-gtxcr Total loading time: 0 Render date: 2024-04-23T23:16:31.370Z Has data issue: false hasContentIssue false

A bipartite version of the Erdős–McKay conjecture

Published online by Cambridge University Press:  09 December 2022

Eoin Long*
Affiliation:
School of Mathematics, University of Birmingham, UK
Laurenţiu Ploscaru
Affiliation:
School of Mathematics, University of Birmingham, UK
*
*Corresponding author. Email: e.long@bham.ac.uk
Rights & Permissions [Opens in a new window]

Abstract

An old conjecture of Erdős and McKay states that if all homogeneous sets in an $n$ -vertex graph are of order $O(\!\log n)$ then the graph contains induced subgraphs of each size from $\{0,1,\ldots, \Omega \big(n^2\big)\}$ . We prove a bipartite analogue of the conjecture: if all balanced homogeneous sets in an $n \times n$ bipartite graph are of order $O(\!\log n)$ , then the graph contains induced subgraphs of each size from $\{0,1,\ldots, \Omega \big(n^2\big)\}$ .

Type
Paper
Creative Commons
Creative Common License - CCCreative Common License - BYCreative Common License - NCCreative Common License - SA
This is an Open Access article, distributed under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike licence (https://creativecommons.org/licenses/by-nc-sa/4.0/), which permits non-commercial re-use, distribution, and reproduction in any medium, provided the same Creative Commons licence is included and the original work is properly cited. The written permission of Cambridge University Press must be obtained for commercial re-use.
Copyright
© The Author(s), 2022. Published by Cambridge University Press

1. Introduction

Given a graph $G$ we write $\mbox{hom}(G)$ to denote the homogeneous number of $G$ , given by

\begin{align*} \operatorname{hom}\!(G) \,:\!=\, \max\! \big \{t \in{\mathbb N}\,:\, \exists \ U\subset V(G) \mbox{ with } |U| = t \mbox{ such that } G[U] \mbox{ is complete or empty} \big \}. \end{align*}

In its simplest form, Ramsey’s theorem [Reference Ramsey30], [Reference Erdős, Szckeres, Gessel and Rota15] states that any $n$ -vertex graph $G$ satisfies $\operatorname{hom}(G) = \Omega (\!\log n)$ and a classical result of Erdős [Reference Erdős14] shows that this behaviour is essentially optimal; there are $n$ -vertex Ramsey graphs $G_0$ with $\operatorname{hom}(G_0) = O (\!\log n)$ . However, the existence of such Ramsey graphs has only been demonstrated indirectly via probabilistic methods and finding explicit constructions of graphs exhibiting such behaviour remains a tantalising open problem (a $ \$ $ 1000 Erdős problem [Reference Chung and Graham11]).

Despite the challenges in constructing Ramsey graphs, and perhaps influenced by them, there has been much success in understanding the intrinsic properties possessed by these graphs. For example, Ramsey graphs have been shown to (roughly) exhibit similar behaviour to the Erdős-Renyi random graph w.h.p. with respect to edge density [Reference Erdős and Szemerédi16], non-isomorphic induced subgraphs [Reference Shelah31], universality of small induced subgraphs [Reference Prömel and Rödl29], and the possible edge sizes and degrees appearing in induced subgraphs [Reference Kwan and Sudakov23], [Reference Kwan and Sudakov22], [Reference Jenssen, Long, Keevash and Yepremyan19].

A challenging remaining problem in this context is the Erdős–McKay conjecture [Reference Erdős17]. Informally, the conjecture asks whether every Ramsey graph must contain (essentially) the entire interval of possible induced subgraph sizes. More precisely the conjecture asks whether every $n$ -vertex graph $G$ with $\operatorname{hom}(G) \leq C \log n$ satisfies $\left\{0,\ldots, \Omega _C\!\left(n^2\right)\right\} \subset \{e(G[U])\,:\, U \subset V(G)\}$ . The best known bound for the conjecture is due to Alon, Krivelevich and Sudakov [Reference Alon, Krivelevich and Sudakov1] who proved that such graphs necessarily contain induced subgraphs of each size in $\left\{0, \ldots, n^{\Omega _C(1)}\right\}$ . The conjecture was also proved for random graphs (in a strong form) by Calkin, Frieze and McKay in [Reference Calkin, Frieze and McKay10]. More recently, Kwan and Sudakov [Reference Kwan and Sudakov23] gave further support proving that such graphs necessarily contain induced subgraphs of $\Omega _C\big(n^2\big)$ different sizes, which improved an earlier almost quadratic bound of Narayanan, Sahasrabuhde and Tomon [Reference Narayanan, Sahasrabudhe and Tomon25].

Our aim here is to study the natural analogue of this conjecture for bipartite graphs. Recall that the bipartite analogue of Ramsey’s theorem states that any balanced bipartite graph $G = (V_1, V_2, E)$ with $|V_1| = |V_2| = n$ contains either $K_{t,t}$ or $\overline{K_{t,t}}$ as an induced subgraph with $t = \Omega (\!\log n)$ . This type of behaviour is again known to be the best possible in general, though explicit constructions of ‘bipartite Ramsey graphs’ are also unknown. In fact, these are more challenging in a sense as such constructions would lead to constructions in the usual Ramsey setting (see e.g. [Reference Barak, Rao, Shaltiel and Wigderson7]). In part, this has contributed to significant interest in Ramsey results in the bipartite setting, e.g. see [Reference Conlon13], [Reference Collins, Riasanovsky, Wallace and Radziszowski12], [Reference Axenovich, Sereni, Snyder and Weber4] [Reference Axenovich, Tompkins and Weber5], [Reference Bucić, Letzter and Sudakov8], [Reference Souza32].

Given the context above, and the difficulties in the Erdős–McKay conjecture, it is natural to ask what can be said about the edge sizes of induced subgraphs in balanced $n$ -vertex bipartite Ramsey graphs? A general result of Narayanan, Sahasrabuhde and Tomon [Reference Narayanan, Sahasrabudhe and Tomon27] gives some information here. These authors also studied a generalisation of the Erdős ‘multiplication table problem’ and showed that any bipartite graph with $m$ edges has induced subgraphs of $\widetilde{\Omega }(m)$ distinct sizes. Recently, Baksys and Chen [Reference Baksys and Chen6] raised the bipartite Ramsey question and proved an analogue of Kwan and Sudakov’s theorem: any balanced bipartite Ramsey graph on vertex classes of order $n$ has induced subgraphs of $\Omega \big(n^2\big)$ different edge sizes.

Our main theorem extends this line of research, proving an analogue of the Erdős–McKay conjecture in the bipartite setting. Before stating it, we give a more precise definition of the Ramsey property for bipartite graphs, in a slightly more general setting.

Definition 1.1. Given $C\gt 0$ , a graph $G = (V_1, V_2, E)$ is called $C$ -bipartite-Ramsey if for any $t_1\geq C\log _2|V_1|$ and $t_2\geq C\log _2|V_2|$ there is no induced copy of $K_{t_1,t_2}$ or $\overline{K_{t_1,t_2}}$ in $G$ .

We will often simply say that $G$ is a $C$ -Ramsey graph when it is clear that $G$ is bipartite.

We can now state our main result, which gives an analogue of the Erdős–McKay conjecture in the bipartite Ramsey setting.

Theorem 1.2. Given $C\gt 0$ there is $\alpha \gt 0$ such that the following holds. Suppose that $G = (V_1, V_2,E)$ is a $C$ -bipartite-Ramsey graph. Then $\{0,\ldots, \alpha |V_1||V_2|\} \subset \{e(G[U])\,:\, U \subset V(G)\}$ .

Our proof of Theorem 1.2 follows a similar line of approach to [Reference Narayanan, Sahasrabudhe and Tomon25], [Reference Kwan and Sudakov23] and [Reference Baksys and Chen6], in which we first show that one can get close to the desired edge sizes and then refine this to show that certain perturbations are typically available to allow one to adjust to the exact size. Anti-concentration estimates are a key tool in ensuring that the desired perturbations are ‘sufficiently rich’ here. We prove such bounds using diversity of vertices and diversity for pairs of vertices, introduced by Bukh and Sudakov [Reference Bukh and Sudakov9] and Kwan and Sudakov [Reference Kwan and Sudakov23], respectively, though a different notion of pair diversity was key here in obtaining the required behaviour. We were also able to keep track of the perturbation structure using certain sumset techniques.

Update: After this paper was submitted, Kwan, Sah, Sauermann and Sawhney [Reference Kwan, Sah, Sauermann and Sawhney21] uploaded a remarkable paper, which has completely resolved the Erdős $-$ McKay conjecture. The proof is a tour de force, combining a wide range of techniques, and it is significantly different from our approach here.

Notation. Given disjoint sets $V_1$ and $V_2$ we write $G = (V_1, V_2,E)$ to represent a bipartite graph $G$ with vertex set $V(G)=V_1\sqcup V_2$ and edge set $E(G)=E\subset V_1\times V_2$ . The edge density of $G$ is given by $e(G)\large{/}|V_1||V_2|$ . Given $U_i \subset V_i$ for $i = 1,2$ we write $G[U_1, U_2]$ to denote the induced subgraph $G[U_1, U_2] = (U_1, U_2, E \cap (U_1 \times U_2))$ .

Given a graph $G$ and $u,v\in V(G)$ , we write $u\sim v$ if $u$ and $v$ are adjacent vertices in $G$ . The neighbourhood of $u$ is given by $N_G(u) = \{v\in V(G)\,:\, u\sim v\}$ and given $S \subset V(G)$ we let $N^S_G(u) \,:\!=\, N_G(u) \cap S$ ; we will omit the subscript $G$ when the graph is clear from the context. We write $d^S_G(u) = \big|N^S_G(u)\big|$ .

Given vertices $u,v\in V(G)$ we write $\text{div}_G(u,v)$ for the symmetric difference $N(u)\triangle N(v)$ . The biased diversity of $u$ and $v$ , denoted by $\text{divb}_G(u,v)$ , is going to be the largest of the two sets $N(u)\setminus N(v)$ and $N(v)\setminus N(u)$ . If these have the same size, then we arbitrarily pick one of the sets to be $\text{divb}_G(u,v)$ . Clearly $|\text{divb}_G(u,v)|\geq |\text{div}_G(u,v)|/2$ .

We will also be interested in ordered pairs of vertices $\boldsymbol{{p}}= (u,v) \in \binom{V(G)}{2}$ . Naturally, given such a pair $\boldsymbol{{p}}$ we can define $\text{div}_G(\boldsymbol{{p}})\,:\!=\,\text{div}_G(u,v)$ . We will also make the convention throughout that all pairs are ordered so that $\text{divb}_G(\boldsymbol{{p}})=N(u) \setminus N(v)$ , which implies $d(u) \geq d(v)$ . Moreover, for later diversity purposes, we will need to study $\operatorname{deg-diff}^{S}({\bf p})\,:\!=\,d_G^S(u)-d_G^S(v)$ .

Given integers $m \leq n$ we will write $[m,n]$ to denote the interval $\{m, m+1, \ldots, n\}$ and given $n \in{\mathbb N}$ we write $[n]$ for the interval $\{1,\ldots, n\}$ . All logarithms in the paper will be base $2$ unless otherwise stated. Floor and ceiling signs are omitted throughout for clarity of presentation.

2. Collected tools

Before beginning in earnest on the proof of the theorem, we make two simple observations on the relation between vertex classes and the size of a $C$ -bipartite-Ramsey graph $G = (V_1, V_2, E)$ , which are useful for future reference.

  • Given any $\alpha \in (0,1)$ and $U_i \subset V_i$ with $|U_i| \geq |V_i|^{\alpha }$ for $i = 1,2$ , the induced graph $G[U_1, U_2]$ is $(C \alpha ^{-1})$ -Ramsey.

  • Suppose that $|V_1| \geq |V_2|$ and let $W \subset V_2$ of size $|W| = 2C \log |V_2|$ . By the pigeonhole principle there is a set $U \subset V_1$ with $|U| \geq |V_1|2^{-|W|}$ such that for all $u\in U$ one has $N_G^W(u) = W^{\prime}$ . It follows that $G$ contains an induced $K_{t_1, t_2}$ or $\overline{K_{t_1,t_2}}$ where $t_1 = |V_1|2^{-|W|}$ and $t_2 = C \log |V_2|$ . As $G$ is $C$ -Ramsey it follows that $t_1 \leq C \log |V_1|$ , which in particular gives $|V_1| \leq |V_2|^{O_C(1)}$ . Thus the vertex classes of $C$ -bipartite-Ramsey graphs are necessarily polynomially related.

2.1 Density control

We start by proving a density result for bipartite Ramsey graphs, which is the analogue of the Erdős-Szemerédi theorem [Reference Erdős and Szemerédi16] for Ramsey graphs in general. The proof uses a well-known argument due to Kovari $-$ Sós $-$ Turán [Reference Kovari, Sós and Turán20].

Lemma 2.1. Given $C\gt 1$ there is $n_C\in \mathbb{N}$ such that the following holds. Suppose that $G = (V_1, V_2,E)$ is a $C$ -bipartite-Ramsey graph with $|V_1|,|V_2|\geq n_C$ . Then $G$ has edge density between $(16C)^{-1}$ and $1-(16C)^{-1}$ .

Proof. To see this let $\varepsilon =(16C)^{-1}, t_1=C\log |V_1|$ , $t_2=C\log |V_2|$ and assume that $t_1\leq t_2$ . It is enough to show that our graph cannot have density larger than $1-\varepsilon$ , as the other statement follows by looking at the (bipartite) complement $\overline{G}$ .

For the sake of contradiction, suppose the density $d(G) \gt 1-\varepsilon$ , and let us count in two ways the number $M$ of stars $K_{t_1,1}$ which are formed by taking a single vertex in $V_2$ and $t_1$ of its neighbours. On one hand, each vertex $v\in V_2$ contributes $\binom{d(v)}{t_1}$ stars, giving us:

(1) \begin{align} M\geq \sum _{v\in V_2}\binom{d(v)}{t_1}\geq |V_2|\cdot \binom{e(G)/|V_2|}{t_1}\geq |V_2|\cdot \binom{(1-\varepsilon )|V_1|}{t_1} \geq |V_2| \cdot (1-2\varepsilon )^{t_1} \binom{|V_1|}{t_1}. \end{align}

Here we have used Jensen’s inequality for the map $x\to \binom{x}{t_1}$ in the second inequality, and that $t_1 \leq \varepsilon |V_1|$ for the final step, as $|V_1| \geq n_C$ .

On the other hand, if $G$ is $K_{t_1,t_2}$ -free then for each subset $S\subset V_1$ of $t_1$ vertices there are at most $t_2-1$ vertices in $V_2$ that can form a star with $S$ , and so $M\leq (t_2-1)\binom{|V_1|}{t_1}$ . Combined with (1) this gives $|V_2| \leq (t_2-1) (1-2\varepsilon )^{-t_1} \leq |V_2|^{1/2} e^{4\varepsilon t_1}$ , using that $|V_2| \geq n_C$ and that $1 - x \geq e^{-2x}$ for $x \in [0, 1/2]$ . It follows that $|V_2| \leq e^{8\varepsilon t_2} = |V_1|^{8C\varepsilon } = |V_2|^{1/2}$ , a contradiction.

Corollary 2.2. For any $C\gt 1$ there exists $n_C\in \mathbb{N}$ such that the following holds true. Any $C$ -bipartite-Ramsey graph $G = (V_1,V_2,E)$ with $|V_1|,|V_2|\geq n_C$ contains at least $2|V_1|/3$ vertices in $V_1$ which all have degrees between $(32C)^{-1}|V_2|$ and $(1-(32C)^{-1})|V_2|$ .

Proof. Suppose for the sake of contradiction that the conclusion is not true. Let $\varepsilon \,:\!=\,1/32C$ and suppose at most $2|V_1|/3$ vertices $v\in V_1$ have degrees between $\varepsilon |V_2|$ and $(1-\varepsilon )|V_2|$ . Then, by the pigeonhole principle, without loss of generality we can assume that there is a set $U_1 \subset V_1$ with $|U_1| \geq |V_1|/6$ such that all vertices in $U_1$ have degree less than $\varepsilon |V_2|$ . But then the edge density of the induced bipartite graph $G[U_1,V_2]$ is less than $\varepsilon$ . On the other hand, it is easy to see that if $G$ is $C$ -bipartite-Ramsey, since $6|U_1| \geq |V_1| \geq n_C$ , the graph $G[U_1,V_2]$ is $2C$ -bipartite-Ramsey. This is a contradiction by Lemma 2.1.

In particular, we can repeatedly make use of Corollary 2.2 inside a bipartite graph to obtain the following result, which will be useful later.

Lemma 2.3. Given $C\gt 1$ and a natural number $L$ there is $n_0\in \mathbb{N}$ such that the following holds. Suppose that $G = (V_1, V_2, E)$ is a $C$ -bipartite-Ramsey graph with $|V_i|\geq n_0$ . Then, taking $\varepsilon _i\,:\!=\,(64C)^{-i}$ for all $i\in [L]$ , one can find vertices $U\,:\!=\,\{u_1,u_2,\ldots,u_L\}\subset V_1$ such that $ \Big|N(u_i)\setminus \underset{j\lt i}{\bigcup }N(u_j) \Big|\geq \varepsilon _i|V_2|$ and $ \Big|V_2\setminus \underset{j\leq i}{\bigcup } N(u_j) \Big|\geq \varepsilon _i|V_2|$ for all $i\in [L]$ .

Proof. By Corollary 2.2 we can pick $u_1\in V_1$ such that $d(u_1)\in [\varepsilon _1|V_2|,(1-\varepsilon _1)|V_2|]$ . Our requirement is clearly satisfied for $i=1$ . Now let us assume that we have found $u_1,u_2,\ldots,u_i$ with $i \lt L$ that satisfy our requirements and let us look for a vertex $u_{i+1}$ .

Let $S_i\,:\!=\,V_1\setminus \{u_1,u_2,\ldots,u_i\}$ and $T_i\,:\!=\,V_2\setminus \underset{j\leq i}{\cup } N(u_j)$ . Note that $|S_i| \geq |V_1| - L \geq |V_1|^{1/2}$ and that $|T_i| \geq (64C)^{-L}|V_2| \geq |V_2|^{1/2}$ since $|V_1|, |V_2| \geq n_0$ . Therefore, the subgraph $G[S_i,T_i]$ is $(2C)$ -bipartite-Ramsey. Thus, by Corollary 2.2, it follows that we can find a vertex $u_{i+1}\in S_1$ such that $d^{T_i}_G(u_{i+1})\in [\varepsilon _1|T_i|,(1-\varepsilon _1)|T_i|]$ . But this is precisely the vertex we were looking for, as then we have $\Big|N(u_{i+1})\setminus \underset{j\leq i}{\cup }N(u_j)\Big|\geq d^{T_i}_G(u_{i+1})\geq \varepsilon _1|T_i|\geq \varepsilon _{i+1}|V_2|$ and we also get that $\Big|V_2\setminus \underset{j\leq i+1}{\cup } N(u_j)\Big|\geq \Big|T_i\setminus N^{T_i}_{G}(u_{i+1})\Big|\geq \varepsilon _1|T_i| \geq \varepsilon _{i+1}|V_2|$ .

By repeating this step $L$ times we reach our conclusion.

2.2 Richness and diversity

We will use the notion of richness, which was introduced by Kwan and Sudakov [Reference Kwan and Sudakov23].

Definition 2.4. Given $\delta,\varepsilon \gt 0$ , a bipartite graph $G = (V_1, V_2,E)$ is $(\delta,\varepsilon )$ -bipartite-rich if for each $i\in \{1,2\}$ the following holds: for every set $W \subset V_i$ with $|W| \geq \delta |V_i|$ there are at most $|V_{3-i}|^{1/5}$ vertices $v \in V_{3-i}$ such that $|N(v) \cap W| \leq \varepsilon |V_{3-i}|$ or $|\overline{N}(v) \cap W| \lt \varepsilon |V_{3-i}|$ .

By adapting a result of Kwan and Sudakov (Lemma 4 in [Reference Kwan and Sudakov23]) to the bipartite context, we show that the Ramsey setting guarantees richness. Perhaps the most striking difference here is that in a general Ramsey setting one needs to move to a subgraph to find richness, whereas in the bipartite setting Ramsey graphs already possess it.

Lemma 2.5. Given $C,\delta \gt 0$ there is $\varepsilon \gt 0$ and $n_0\in \mathbb{N}$ such that the following holds true. Every $C$ -bipartite-Ramsey graph $G = (V_1, V_2,E)$ with $|V_1|, |V_2|\geq n_0$ is $(\delta,\varepsilon )$ -bipartite-rich.

Proof. It is enough to prove that the bipartite-richness condition holds when $i = 1$ . So set $\varepsilon \,:\!=\,(200C)^{-1}$ and suppose there is a set $U_1\subset V_1$ with $|U_1|\geq \delta |V_1|$ which contradicts the bipartite-richness condition – more precisely that there is a set $W_2\subset V_2$ with $|W_2|\geq |V_2|^{1/5}$ such that $|N(v)\cap U_1|\lt \varepsilon |U_1|$ or $|\overline{N(v)}\cap U_1|\lt \varepsilon |U_1|$ for all $v\in W_2$ . Without loss of generality, we can assume that there is a subset $U_2\subset W_2$ of size $|V_2|^{1/5}/2$ such that $|N(v)\cap U_1|\lt \varepsilon |U_1|$ for all $v\in W_2$ . But this means that the edge density of $G[U_1,U_2]$ is less than $\varepsilon$ .

On the other hand, as $|U_i| \geq |V_i|^{1/5}/2 \geq |V_i|^{1/6}$ as $|V_i| \geq n_C$ , and since $G = G[V_1,V_2]$ is $C$ -Ramsey, it follows $G[U_1, U_2]$ is $6C$ -Ramsey. But by Lemma 2.1 such a graph must have edge density at least $(196C)^{-1}\gt \varepsilon,$ which is a contradiction, thus proving our result.

Next, we discuss the notion of diversity, which was introduced by Bukh and Sudakov [Reference Bukh and Sudakov9] (in the non-bipartite setting).

Definition 2.6. A bipartite graph $G = (V_1, V_2,E)$ is said to be $c$ -bipartite-diverse if for each $i\in \{1,2\}$ the following holds: every vertex $v\in V_i$ has $|\text{div}_G(v,w)|\leq c|V_{3-i}|$ for at most $|V_i|^{1/5}$ vertices $w \in V_i$ .

We also introduce a useful diversity notion for pairs. We note that diversity for pairs was also considered by Kwan and Sudakov in [Reference Kwan and Sudakov23], who considered multisets of neighbourhoods, but we require a different notion suitable for our later applications.

Definition 2.7. A bipartite graph $G = (V_1, V_2,E)$ is said to be $(c,\alpha )$ -pair-diverse if for each $i\in \{1,2\}$ the following holds true for both $i = 1,2$ . For each ordered pair $\boldsymbol{{p}}\in \binom{V_i}{2}$ with $|\text{div}(\boldsymbol{{p}})|\geq \alpha |V_{3-i}|$ there are at most $|V_i|^{1/5}$ pairwise vertex disjoint pairs $\boldsymbol{{p}}^\prime =(x,y) \in \binom{V_i}{2}$ such that $|\text{divb}_G(\boldsymbol{{p}})\setminus N(x)|\leq c|V_{3-i}|$ or $|\text{divb}_G(\boldsymbol{{p}})\setminus N(y)|\leq c|V_{3-i}|$ .

Lemma 2.8. Let $G = (V_1, V_2,E)$ be a $(\delta,\varepsilon )$ -bipartite-rich graph with $\delta \leq 1/2$ . Then:

  1. (i) $G$ is $\varepsilon/2$ -diverse;

  2. (ii) $G$ is $(\alpha \varepsilon/2,\alpha )$ -pair-diverse for all $\alpha \geq 2\delta$ .

Proof. It is enough to show that each property holds when taking $i = 1$ .

(i) For each $v\in V_1$ either $|N(v)| \geq |V_2|/2$ or $|\overline{N(v)}|\geq |V_2|/2$ . In the former case, all but at most $|V_1|^{1/5}$ vertices $w\in V_1$ we have $|N(v)\cap \overline{N(w)}|\geq \varepsilon |N(v)|\geq \varepsilon/2\cdot |V_2|$ , and in the latter, for all but at most $|V_1|^{1/5}$ vertices $w\in V_1$ we have $|\overline{N(v)}\cap{N(w)}|\geq \varepsilon |\overline{N(v)}|\geq \varepsilon/2\cdot |V_2|$ . In either case, there are at most $|V_1|^{1/5}$ vertices $w\in V_1$ with $|\text{div}_G(v,w)| \lt \varepsilon/2\cdot |V_2|$ , as desired.

(ii) Suppose now there is an ordered pair $\boldsymbol{{p}}\,:\!=\,\{x,y\}\in \binom{V_1}{2}$ and a collection $Y$ of $|V_1|^{1/5}$ vertex disjoint pairs $\boldsymbol{{p}}_i\,:\!=\,(x_i,y_i) \in \binom{V_1}{2}$ such that $|\text{divb}(\boldsymbol{{p}})\setminus N(x_i)|\leq \alpha \varepsilon/2\cdot |V_2|$ for each $i$ . Note that if $|\text{div}(\boldsymbol{{p}})|\geq \alpha |V_2|$ then $|\text{divb}(\boldsymbol{{p}})|\geq \alpha/2\cdot |V_2|$ , as $\text{divb}(\boldsymbol{{p}})$ is simply the largest of the two sets $N(x)\setminus N(y)$ and $N(y)\setminus N(x)$ , whose union is $\text{div}(\boldsymbol{{p}})$ . But in this case there are $|V_1|^{1/5}$ distinct vertices $z_i \in \{x_i, y_i\}$ such that $|\text{divb}(\boldsymbol{{p}})\cap \overline{N(z_i)}|\leq \alpha \varepsilon/2\cdot |V_2|\leq \varepsilon |\text{divb}(\boldsymbol{{p}})|$ , which contradicts the richness of the set $\text{divb}(\boldsymbol{{p}})$ .

2.3 Probabilistic tools

Throughout the proof, we will use Markov’s inequality, Chebyshev’s inequality, the Chernoff bound and Turán’s theorem. Statements and proofs of all of these can be found, for example, in [Reference Alon and Spencer3]. We will also need a probabilistic variation of the Erdős–Littlewood–Offord theorem, a proof of which can be found, for instance in [Reference Jenssen, Long, Keevash and Yepremyan19], or more generally as a consequence of the Doeblin-Kolmogorov-Levy-Rogozin theorem.

Theorem 2.9. Fix some non-zero parameters $a_1,a_2,\ldots,a_n\in \mathbb{R}$ and $\alpha \in (0,0.5]$ , then let $p_1,p_2,\ldots,p_n\in [\alpha,1-\alpha ]$ . Suppose that $X_1,X_2,\ldots, X_n$ are independent Bernoulli random variables with $X_i\sim Be(p_i)$ . Then the following holds:

\begin{equation*}\displaystyle \max _{x\in \mathbb {R}} {\mathbb {P}} \bigg ( \displaystyle \sum _{i=1}^n a_iX_i=x \bigg ) =O_{\alpha }(n^{-1/2}).\end{equation*}

The following natural proposition will be useful in our proof of Theorem 1.2.

Proposition 2.10. Given an integer $d\geq 2$ , there is an integer $n_0\,:\!=\,n_0(d)$ such that if $X$ is a uniformly chosen subset of $[n]$ , where $n\geq n_0$ , then for any integer $k$ with $0\leq k\lt d$ one has:

\begin{equation*}\frac {1}{d+1}\leq \mathbb P\big (|X|\equiv k\pmod {d} \big )\leq \frac {1}{d-1}.\end{equation*}

Comparable statements already exist in the literature (see for example Lemma 2.3 in [Reference Ferber, Hardiman and Krivelevich18] for a more quantitative behaviour), but to keep the paper contained we outline a simple proof, showing that it can be easily deduced from a standard result about stochastic processes. The next few paragraphs represent a brief introduction to this topic and, for more details, the reader can refer to [Reference Norris28].

A Markov chain is a sequence $\boldsymbol{{X}}\,:\!=\,(X_n)_{n\geq 0}$ of random variables taking values in some common ground set $I$ such that, for all $n\geq 1$ and $i_0,i_1,\ldots,i_{n}\in I$ , one has:

\begin{equation*}\mathbb P\big (X_{n}=i_n\ |\ X_0=i_0;\ X_1=i_1;\ \ldots ;\ X_{n-1}=i_{n-1}\big )=\mathbb P\big (X_{n}=i_n\ |\ X_{n-1}=i_{n-1}\big ).\end{equation*}

A Markov chain is called homogeneous if, in addition, $p_{i,j}\,:\!=\,\mathbb P(X_n=j|X_{n-1}=i)$ depends only on the states $i$ and $j$ , not on the time $n$ . These quantities are known as the transition probabilities of the chain. We are only interested in homogeneous chains, and we can observe that in order to describe such a chain it is enough to have the initial distribution $\lambda$ of $X_0$ , given by $\lambda _i\,:\!=\,\mathbb P(X_0=i)$ , and the transition matrix $P\,:\!=\,(p_{i,j})_{i,j\in I}$ . Moreover, if we write $p_{i,j}^{(n)}$ for $\mathbb P\!\left(X_{k+n}=j|X_k=i\right)$ , then it is easy to see that $p_{i,j}^{(n)}=(P^n)_{i,j}$ .

A Markov chain $\boldsymbol{{X}}$ on the set $I$ with transition matrix $P\,:\!=\,(p_{i,j})_{i,j\in I}$ is called irreducible if for all $i,j\in I$ there is $n\geq 0$ such that $p_{i,j}^{(n)}\gt 0$ . The period of a state $i\in I$ is defined to be the greatest common divisor of the set $\left\{n\geq 1\,:\,p_{i,i}^{(n)}\gt 0\right\}$ . The Markov chain $\boldsymbol{{X}}$ is called aperiodic if all its states have period $1$ .

We say $\pi = (\pi _i)_{j\in I}$ is a stationary distribution for a Markov chain $\boldsymbol{{X}}$ if starting the chain from $X_0$ with distribution $\pi$ implies that $X_n$ has distribution $\pi$ for all $n\geq 1$ . As the distribution of $X_n$ is given by $\pi P^n$ , where $P$ is the transition matrix, we deduce that $\pi$ is a stationary distribution if and only if $\pi P=\pi$ .

Theorem 2.11. Suppose $\boldsymbol{{X}}$ is an irreducible and aperiodic Markov chain on a ground set $I$ with transition matrix $P$ , stationary distribution $\pi$ and any initial distribution. Then, for all $j\in I$ , one has $\mathbb P(X_n=j)\to \pi _j$ as $n\to \infty$ .

Proof of Proposition 2.10. We can view choosing $X$ as going through each number from $1$ to $n$ and independently tossing a fair coin for each to decide whether it is an element of $X$ or not. Thus $|X|\ (\text{mod }d)$ can be viewed as a Markov chain on $\{0,1,2,\ldots,d-1\}$ starting at $0$ and with transition probabilities $p_{i,i}=p_{i,i-1}=0.5$ for each $0\leq i\lt d$ , where indices ar taken modulo $d$ . This chain is aperiodic as $p_{i,i}^{(1)}=0.5$ , and it is also irreducible as we can reach state $i$ from state $j$ in $i-j(\text{mod }d)$ steps with positive probability. It is easy to see that the stationary distribution $\pi$ is given by $\pi _i=1/d$ . The conclusion now follows from Theorem 2.11.

2.4 Progressions in sumsets

Given sets $A_1,\ldots, A_K \subset{\mathbb Z}$ , the sumset of $A_1,\ldots, A_K$ is the set $A_1 + \ldots + A_K$ given by

\begin{align*} A_1 + \cdots + A_K = \big \{a_1 + \cdots + a_K\,:\, a_i\in A_i \mbox{ for all } i\in [K] \big \}. \end{align*}

Much research has focused on the topic of estimating the size or understanding the structure of sumsets under certain assumptions on the sets (see for example [Reference Szemerédi and Vu33] or Chapter 2 of [Reference Tao and Vu34]). For our purposes, the following elementary estimate will suffice.

Lemma 2.12. Given $\delta, B\gt 0$ there are $C, d_0 \geq 1$ such that the following holds. Suppose that $A_1,\ldots, A_K \subset [{-}M,M]$ with $K \geq CM$ and that $|A_i| \geq \delta M \geq 2$ for all $i\in [K]$ . Then there is $a\in{\mathbb Z}$ and $d \in{\mathbb N}$ with $1 \leq d \leq d_0$ such that:

(2) \begin{align} \big \{ a + i d\,:\, 0 \leq i \leq BM^2 \big \} \subset A_1 + A_2 + \cdots + A_K. \end{align}

Proof. We will prove the statement under the assumption that $A_i \subset [0,M]$ for all $i\in [K]$ . Note that the general case follows immediately by taking translations of the sets, i.e. replacing $A_i$ with $A^{\prime}_i = A_i-c_i$ for $c_i = \min \{a\,:\, a\in A_i\}$ . Indeed if (2) holds for $A^{\prime}_1,\ldots, A^{\prime}_K$ (which may now lie in $[0,2M]$ instead of $[{-}M,M]$ ) then (2) also holds for $A_1,\ldots, A_K$ (possibly with a different value of $a$ ). The same argument also allows us to assume that $0 \in A_i$ for all $i \in [K]$ .

Next, we double count pairs $(m,j)\in [M]\times [K]$ for which $m$ is among the largest $\delta M/2$ elements in $A_j$ . Counting by each set $A_j$ , we deduce that there are $\delta KM/2$ such pairs. Therefore, there is $M^\prime \in [M]$ which appears in at least $\delta K/2$ of these pairs. Let $J\,:\!=\,\{j\in [K]\,:\,M^\prime \in A_j\}$ , so that $|J|\geq \delta K/2$ , and note that $|A_j\cap [0,M^\prime ]|\geq \delta M/2$ as $M^\prime$ is one of the largest $\delta M/2$ elements in $A_j$ . Recalling that also $0\in A_j$ , by restricting all the sets to $[0,M^{\prime}]$ , then eventually reordering them and slightly adjusting the parameter $C$ , we can now assume that $A_i\subset [0,M]$ for all $i\in [K]$ , but with $\{0,M\} \subset A_i$ as well.

Given a set $S \subset{\mathbb Z}$ , we will let $\overline{S} \,:\!=\, \{a \in [0,M-1]\,:\, a \equiv s \mod M \mbox{ for some } s \in S\}$ . For each $i\in [K]$ let $S_i \,:\!=\, \overline{A_1 + \cdots + A_i}$ . By reordering the sets $A_1,\ldots, A_K$ , we may assume that $|S_{i+1}| \gt |S_{i}|$ for $i \leq K^{\prime}$ and that $S_i = S_{K^{\prime}}$ for all $i \geq K^{\prime}$ . Observe that $S_{K^{\prime}}\subset [0,M-1]$ and that $|S_{K^{\prime}}| \geq |S_1| + (K^{\prime}-1) \geq K^{\prime}+ 1$ , therefore $K^{\prime} \lt M$ .

Now, as $0 \in A_i$ for all $i\in [K]$ , we have $0 \in S_{K^{\prime}} \subset \overline{S_{K^{\prime}} + A_{K^{\prime}+1}} = S_{K^{\prime}+1}$ . As $|S_{K^{\prime}+1}| = |S_{K^{\prime}}|$ , it follows that the set $S_{K^{\prime}}$ contains the subgroup of ${\mathbb Z}_M$ generated by $A_{K^{\prime}+1}$ . However, recalling that $|S_{K^{\prime}}| = |S_{K^{\prime}+1}| \geq |A_{K^{\prime}+1}| \geq \delta M$ , we obtain that there is some $d \leq d_0(\delta )$ with $d|M$ such that $\{id \,:\, 0\leq i \leq M/d\} \subset S_{K^{\prime}} \subset S_M$ .

To proceed with the last step, recall that $\{0,M\} \in A_i$ for all $i\in [K]$ . Using this, it is easy to see that as $\{id \,:\, 0\leq i \leq M/d\} \subset S_{K^{\prime}} \subset S_M$ , we have $\{M^2 + id\,:\, 0 \leq i \leq M/d\} \subset A_1 + \cdots + A_{2M}$ . More generally, as $d | M$ , we have:

\begin{equation*} \big \{M^2 + id + jM\,:\, 0 \leq i \leq M/d, \mbox { } 0 \leq j \leq K-2M \big \} \subset A_1 + \cdots + A_{K}, \end{equation*}

and so (2) holds by taking $a = M^2$ and $K \geq CM \geq (B + 2)M$ . This completes the proof.

3. Proof of Theorem 1.2

3.1 Pair-stars and pair-matchings

We start this section by defining two constructions which will be of central importance in our attempt to find induced subgraphs of many sizes. Let $G = (V_1, V_2,E)$ be a bipartite graph.

An $\varepsilon$ -pair-star of size $k$ rooted at $x_0$ associated to $V_1$ is a set ${\mathcal{P}}_S = \{x_0, x_1,x_2,\ldots,x_k\} \subset V_1$ which satisfies the following properties:

  1. (i) $|d_G(x_j)-d_G(x_0)|\leq{|V_2|}^{0.5}$ for all $j\in [k]$ ;

  2. (ii) $|\text{div}(x_i,x_j)|\geq \varepsilon |V_2|$ for all $i\neq j$ in $[0,k]$ .

We define the head of ${\mathcal{P}}_S$ to be the set $H({\mathcal{P}}_S) = \{x_0\}$ .

An $\varepsilon$ -pair-matching of size $k$ associated to $V_1$ is a collection ${\mathcal{P}}_M = \{\boldsymbol{{p}}_1,\boldsymbol{{p}}_2,\ldots,\boldsymbol{{p}}_k\}$ of vertex disjoint ordered pairs of vertices in $V_1$ which satisfy the following properties:

  1. (i) ${\bf p}_i = (x_i, y_i)$ for all $i\in [k]$ and the pairs ${\bf p}_1, \ldots,{\bf p}_k$ are vertex disjoint;

  2. (ii) $\operatorname{deg-diff}^{}({\boldsymbol{{p}}_i})\leq |V_2|^{0.5}$ for all $i\in [k]$ ;

  3. (iii) $|\text{divb}(\boldsymbol{{p}}_i)\setminus N(x_j)|\geq \varepsilon |V_2|$ and $|\text{divb}(\boldsymbol{{p}}_i)\setminus N(y_j)|\geq \varepsilon |V_2|$ for all $i \neq j$ in $[k]$ .

We define the head of ${\mathcal{P}}_M$ to be $H({\mathcal{P}}_M) = \{x_i\,:\, i \in [k]\}$ .

By swapping $V_1$ and $V_2$ above, we can define pair-stars and pair-matchings associated to $V_2$ . Our result in this subsection gives large pair-stars or large pair-matchings in $C$ -Ramsey graphs.

Lemma 3.1. Given $C\gt 1$ there is $\varepsilon \gt 0$ and $n_C\in \mathbb{N}$ such that the following holds. Suppose that $G = (V_1, V_2,E)$ is a $C$ -bipartite-Ramsey graph with $|V_1|\geq |V_2|/2 \geq n_C$ . Then either:

  • $V_1$ contains an $\varepsilon$ -pair-star of size $|V_1|^{0.5}$ , or

  • $V_1$ contains an $\varepsilon$ -pair-matching of size $|V_1|^{0.5}$ .

Proof. Let us start by dividing the interval $[0,|V_2|]$ into $\ell \,:\!=\,{|V_2|}^{1/2}$ disjoint intervals $I_1,I_2,\ldots I_l$ of length ${|V_2|}^{1/2}$ and let, for each interval $I_j$ , $n_j$ denote the number of vertices in $V_1$ whose degree lie in $I_j$ . Now any vertices $x, y \in I_j$ with $d(x) \geq d(y)$ give an ordered pair $\boldsymbol{{p}} = (x,y)$ with $\operatorname{deg-diff}^{}({{\bf p}}) \leq{|V_2|}^{1/2}$ . Obviously $n_1+n_2+\ldots + n_{\ell }=|V_1|$ , so by Jensen’s inequality the collection $\mathcal{S}$ of such ordered pairs $\boldsymbol{{p}} = (x,y) \in \binom{V_1}{2}$ has size:

\begin{align*} |{\mathcal{S}}| \geq \sum _{j=1}^{\ell }\binom{n_j}{2}\geq \ell \cdot \binom{|V_1|/\ell }{2} \geq \frac{|V_1||\big (|V_1|/\ell -1\big )}{2} \geq \frac{|V_1|^{1.5}}{4}, \end{align*}

using $\ell = |V_2|^{0.5} \leq (2|V_1| )^{0.5}$ and $|V_1| \geq n_C$ . By Lemmas 2.8 and 2.5, we deduce that our graph $G$ is $\varepsilon _0$ -diverse, where $\varepsilon _0\,:\!=\,(400C)^{-1}$ . Therefore, at most $|V_1|^{1.2}=o(|\mathcal{S}|)$ pairs from $\mathcal{S}$ fail to be ${\varepsilon }_0$ -diverse. By removing such pairs, we obtain a collection $\mathcal{S}_0$ of size at least $|V_1|^{1.5}/8$ such that $\operatorname{deg-diff}\!({\bf p}) \leq |V_2|^{1/2}$ and $|\text{div}(\boldsymbol{{p}})|\geq \varepsilon _0|V_2|$ for all $\boldsymbol{{p}}\in \mathcal{S}_0$ .

Next, we view the elements of $\mathcal{S}_0$ as the (unordered) edges of a graph $H$ on $V(G)$ . We claim that either $H$ has a matching of size $m\,:\!=\, |V_1|^{0.75}/4$ or a set of $m+1$ edges that have a common vertex. To see why this is true, suppose that neither of these two events hold and let $M$ be a largest matching in $H$ . Then $|M|\lt m$ and so all the edges in $e(H)\setminus M$ must be adjacent to an edge of $M$ . But because of the other condition, each edge of $M$ has at most $2m$ other edges adjacent to it. Thus, $e(H)\lt 2m^2 \leq |\mathcal{S}_0|$ , which is a contradiction.

Let us now observe that this set of edges will create our pair-star or pair-matching. Indeed, assume first that we have obtained $m+1$ edges $x_0x_1,x_0x_2,\ldots,x_0x_{m+1}$ in $\mathcal{S}_0$ . Again, by diversity, each vertex $x_j$ has at most $|V_1|^{0.2}$ other vertices $x_i$ such that $|\text{div}(x_j,x_i)|\lt \varepsilon _0|V_2|$ . Thus, by Turán theorem there is a set $I\subset [0, m+1]$ , with $0 \in I$ , of size at least $(m+1)/(1+|V_1|^{0.2}) \geq |V_1|^{0.5}$ such that $|\text{div}(x_j,x_i)|\geq \varepsilon _0|V_2|$ for all $i\neq j$ in $I$ . Taking $x_0$ as the root gives our pair-star.

In the other case, assume that we have obtained $m$ disjoint ordered pairs ${\bf p}_1,\ldots,{\bf p}_m$ in ${\mathcal{S}}_0$ , with ${\bf p}_i = (x_i, y_i)$ . Then, by Lemmas 2.5 and 2.8 our graph $G$ is $((\varepsilon _0/2)^2, \varepsilon _0/2)$ -pair-diverse, so for each pair $(x_j,y_j)$ there are at most $|V_1|^{0.2}$ other pairwise disjoint edges $x_iy_i$ such that $|\text{divb}(x_j,y_j)\setminus N(x_i)|\lt (\varepsilon _0/2)^2 |V_2|$ or $|\text{divb}(x_j,y_j)\setminus N(y_i)|\lt (\varepsilon _0/2)^2 |V_2|$ . Setting $\varepsilon = (\varepsilon _0/2)^2$ , by Turán, there is a set $I\subset [m]$ of size at least $m (1+|V_1|^{0.2} )^{-1} \geq |V_1|^{0.5}$ such that $|\text{divb}({\bf p}_j)\setminus N(x_i)|\geq \varepsilon |V_2|$ and $|\text{divb}({\bf p}_j)\setminus N(y_i)|\geq \varepsilon |V_2|$ for all $i\neq j$ in $I$ . This represents our pair-matching.

3.2 Degree control from pair-stars and pair-matchings

Our next aim is to show that when picking a subset $W \subset V_2$ uniformly at random, there is a very good chance that pair-stars and pair-matchings associated to $V_1$ produce well distributed degree sets in $W$ . The following lemma will be useful in this context.

Lemma 3.2. Let $G = (V_1, V_2, E)$ be a bipartite graph and suppose the subset $W \subset V_2$ is selected uniformly at random. Then:

  1. (i) if $x, y \in V_1$ with $|{\text{div}}(x,y)|\geq \delta |V_2|$ then $\mathbb P (d^{W}(x)=d^{W}(y) ) =O_{\delta } (|V_2|^{-0.5} ).$

  2. (ii) if $\boldsymbol{{p}}_1=(x_1,y_1)\in \binom{V_1}{2}$ and $\boldsymbol{{p}}_2=(x_2,y_2)\in \binom{V_1}{2}$ with $|{\text{divb}}(\boldsymbol{{p}}_1)\setminus N(x_2)|\geq \delta |V_2|$ and $|{\text{divb}}(\boldsymbol{{p}}_1)\setminus N(y_2)|\geq \delta |V_2|$ then $ \mathbb P ( \operatorname{deg-diff}^{ }({{\bf p}_1}){W}=\operatorname{deg-diff}^{ }({{\bf p}_2}){W} )=O_\delta (|V_2|^{-0.5} ).$

Proof. (i) We will use a classical randomness exposure argument. Suppose we reveal the random set $W$ on $V_2\setminus \text{div}(x,y)$ . Then, given such a choice, the difference $d^{W}(x)-d^{W}(y)$ becomes $d^{U}(x)-d^{U}(y)+\text{constant}$ , where $U \,:\!=\,W \cap \text{div}(x,y)$ . Now, $d^{U}(x)-d^{U}(y)\,:\!=\,\sum _{v\in \text{div}(x,y)}\theta _vX_v$ , where for each $v\in \text{div}(x,y)$ we have $\theta _v\in \{-1,1\}$ and $X_v\sim \text{Bern}(0.5)$ . Since $|\text{div}(x,y)|\geq \delta |V_2|$ , by Theorem 2.9, the random variable $d^{W^{\prime}}(x)-d^{W^{\prime}}(y)$ hits any particular value with probability $O_\delta (|V_2|^{-1/2} )$ . The conclusion follows from the law of total probability.

(ii) The argument here is similar, but a little more involved. By the choice of the ordering we have $\mbox{divb}({\bf p}_1) = N(x_1) \setminus N(y_1)$ . Note that the condition in the lemma now gives that the set $T = \mbox{divb}({\bf p}_1) \setminus N(x_2) = N(x_1) \setminus ( N(x_2) \cup N(y_1) )$ satisfies $|T| \geq \delta |V_2|$ . Letting $X$ denote the random variable $X = \operatorname{deg-diff}^{W}\!({{\bf p}_1}) - \operatorname{deg-diff}^{W}\!({{\bf p}_2})$ we have:

\begin{align*} X = d^{W}(x_1)-d^{W}(y_1) -d^{W}(x_2) + d^{W}(y_2). \end{align*}

Now note that if we expose the random set $W \cap (V_2 \setminus T)$ , then the random variable $X$ reduces to $d^{W \cap T}(x_1)+d^{W \cap T}(y_2)-C$ , where $C$ is some constant depending only on $W\cap (V_2 \cap T)$ – crucially, here we use that $y_1$ and $x_2$ have no neighbour in $T$ . But $d^{W \cap T}(x_1)+d^{W \cap T}(y_2)=\sum _{v\in T}\theta _vX_v$ where $\theta _v\in \{1,2\}$ is a constant for each $v\in T$ and $X_v\sim \text{Bern}(0.5)$ so that $\{X_v\}_{v\in T}$ are independent. By Theorem 2.9, since $|T|\geq \delta |V_2|$ , the random variable $d^{W \cap T}(x_1)+d^{W\cap T}(y_2)$ takes any particular value with probability at most $O_\delta (|V_2|^{-0.5} )$ . The conclusion again follows from the law of total probability.

Let $G = (V_1, V_2, E)$ be a bipartite graph and select a vertex set $W \subset V_2$ . Given an $\varepsilon$ -pair star ${\mathcal{P}} = \{x_0,x_1,\ldots, x_k\}$ rooted at $x_0$ and associated to $V_1$ , we write:

\begin{align*} A_{\mathcal{P}}^W \,:\!=\, \big \{d^W(x_i) - d^W(x_0)\,:\, i\in [k] \big \} \cap \big [{-}3|V_2|^{0.5},3|V_2|^{0.5} \big ]. \end{align*}

The following lemma provides a useful estimate on the size of $A_{\mathcal{P}}^W$ .

Lemma 3.3. Given $\varepsilon \gt 0$ there is $\delta \gt 0$ such that the following holds. Let $G = (V_1, V_2, E)$ be a bipartite graph with $|V_1|\geq |V_2|/2$ and let ${\mathcal{P}} = \{x_0,\ldots, x_{|V_2|^{0.5}}\}$ be a $\varepsilon$ -pair-star of size ${|V_2|}^{0.5}$ rooted at $x_0$ and associated to $V_1$ . Suppose that a set $W \subset B$ is selected uniformly at random, then ${\mathbb{P}}\! \left(|A_{\mathcal{P}}^W| \geq \delta |V_2|^{0.5} \right) \geq 3/4$ .

Proof. To begin, pick $j\in [|V_2|^{0.5}]$ and set $D_j \,:\!=\, d^W(x_j) - d^W(x_0)$ . By Chebyshev’s inequality, we have $ |D_j-\mathbb E[D_j] |\leq 2{|V_2|}^{0.5}$ with probability at least $15/16$ . As $ |\mathbb E[D_j] |\leq |V_2|^{1/2}/2$ , it follows by triangle inequality that $\mathbb P (|d^{W}(x_0)-d^{W}(x_j)|\leq 3|V_2|^{0.5} )\geq 15/16$ .

Next, call a vertex $x_j$ with $j\in [|V_2|^{0.5}]$ bad if $|d^{W}(x_0)-d^{W}(x_j)|\leq 3|V_2|^{0.5}$ and good otherwise. Note $\mathbb P(x_j\text{ is bad})\leq 1/16$ and so $\mathbb E [ |\{j\in [|V_1|^{0.5}]\,:\,x_j\text{ is bad}\} | ]\leq |V_2|^{0.5}/16$ . If $U$ denotes the set of good vertices, then it follows by Markov that $\mathbb P (|U|\geq |V_2|^{0.5}/2 )\geq 7/8$ .

Now we build a graph $H$ on $\{x_1,\ldots, x_{|V_2|^{1/2}}\}$ where we join two vertices by an edge in $H$ if their degrees in $W$ are equal. By Lemma 3.2 (i) there is an edge in $H$ between two vertices $x_i$ and $x_j$ with probability $O_{\varepsilon } (|V_2|^{-1/2} )$ . Therefore, $\mathbb E[e(H)]=O_{\varepsilon } ((|V_2|^{0.5})^2\cdot |V_2|^{-0.5} )$ and by Markov we easily deduce that, with probability at least $7/8$ , one has $e(H)=O_{\varepsilon } (|V_2|^{0.5} )$ .

Combining our estimates, by applying the union bound we find that $|U| \geq |V_2|^{0.5}/2$ and that $e(H[U]) \leq e(H)=O_{\varepsilon } (|V_2|^{0.5} )$ with probability at least $3/4$ . But then the average degree of $H$ is of order $O_{\varepsilon } (1 )$ , hence by Turán $H$ contains an independent set of size $\Omega _{\varepsilon } ({|V_2|^{0.5}} )$ . This set though gives us precisely the vertices with pairwise distinct degrees that we sought.

Let $G = (V_1, V_2, E)$ be a bipartite graph, and set $W \subset V_2$ . Given an $\varepsilon$ -pair-matching ${\mathcal{P}} = \{{\bf p}_1,\ldots,{\bf p}_k\}$ associated to $V_1$ , we write:

\begin{align*} A_{\mathcal{P}}^W \,:\!=\, \big \{\!\operatorname{deg-diff}^{W}\!({{\bf p}_i})\,:\, i\in [k] \big \} \cap \big [{-}3|V_2|^{0.5},3|V_2|^{0.5} \big ]. \end{align*}

The following lemma gives the analogous estimate for $A_{\mathcal{P}}^W$ when ${\mathcal{P}}$ is an pair-matching and can be proven in exactly the same way as in Lemma 3.3, replacing Lemma 3.2 (i) in the proof with Lemma 3.2 (ii) instead.

Lemma 3.4. Given $\varepsilon \gt 0$ there is $\delta \gt 0$ such that the following holds. Let $G = (V_1, V_2, E)$ be a bipartite graph with $|V_1|\geq |V_2|/2$ and let ${\mathcal{P}}$ be a $\varepsilon$ -pair-matching in $V_1$ of size $|V_2|^{0.5}$ . Suppose that a set $W \subset B$ is selected uniformly at random, then ${\mathbb{P}} (|A_{\mathcal{P}}^W| \geq \delta |V_2|^{0.5} ) \geq 3/4$ .

3.3 Breaking modular obstructions

Lemma 3.5. Given a natural number $d \gt 1$ and $\varepsilon \gt 0$ there are $L, D \gt 0$ such that the following holds. Suppose that $G = (V_1, V_2, E)$ is a bipartite graph and that $\{u_1,\ldots, u_L\} \subset V_1$ such that $S_i = N_G(u_i) \setminus \cup _{j \lt i} N_G(u_j)$ satisfies $|S_i| \geq D$ for all $i\in [L]$ . Then, if a subset $W$ is chosen uniformly at random from $V_2$ , with probability at least $1 - \varepsilon$ for every pair $(k,m)$ with $0 \leq k \leq m$ and $2 \leq m \leq d$ there is $i\in [L]$ such that $d_G^{W}(u_i) \equiv k \mod m$ .

Proof. Suppose that $W \subset V_2$ is selected as in the lemma. To begin, we fix a pair $(k,m)$ so that $0 \leq k \leq m$ and $2 \leq m\leq d$ . For $i \in [L]$ let $E_i$ denote the event $E_i \,:\!=\, \left\{d^W_G(u_i) \not \equiv k \mod m\right\}$ . Our first aim is to upper bound ${\mathbb{P}} \left( \cap _{i\in [L]} E_i \right)$ . To do this, we can observe that:

\begin{align*} \mathbb P\big (\cap _{i=1}^L E_i\big )=\prod _{i=1}^L \mathbb P\!\left (E_i|\cap _{j\lt i}E_j\right ). \end{align*}

Now the event $\cap _{j\lt i}E_i$ is entirely determined by the choice of $W^{\prime} = W \cap (V_2 \setminus S_i)$ . Thus, to upper bound $\mathbb P \!\left(E_i|\cap _{j\lt i}E_j \right)$ , it suffices to upper bound $\mathbb P\! \left(E_i|W^{\prime} = W_0 \right)$ for each choice of $W_0$ . However, given such a choice of $W_0$ , the conditional probability $\mathbb P\! \left(E_i|W^{\prime}=W_0 \right)$ becomes $\mathbb P\! \left(d^{W\cap S_i}_G(u_i)\not \equiv k_i\ (\text{mod }m) \right)$ for some $0\leq k_i\lt m$ . Besides this, from our hypothesis we know that $W \cap S_i \sim \mbox{Bin}(|S_i|,0.5)$ and $|S_i| \geq D$ , so by Proposition 2.10 this gives $\mathbb P\! \left(E_i|\cap _{j\lt i}E_j \right) \leq \max _{W_0}{\mathbb{P}}\left(E_i| W^{\prime} = W_0\right) \leq d(d+1)^{-1}$ . Combining all these, we obtain:

\begin{align*} \mathbb P\big (\cap _{i=1}^L E_i\big ) \leq \prod _{i=1}^L \mathbb P\!\left (E_i|\cap _{j\lt i}E_j\right ) \leq \Big ( 1 - \frac{1}{d+1} \Big ) ^{L}\leq \exp \!\left ({-}L(d+1)^{-1}\right ). \end{align*}

To finish the proof, let $F_{(k,m)}$ denote the event that $\{d_G^W(u_i) \not \equiv k \mod m \mbox{ for all } i\in [L]\}$ . We have shown that ${\mathbb{P}}(F_{(k,m)}) \leq \exp\!\left( -L(d+1)^{-1} \right)$ . It follows that the probability that some congruence is not obtained is ${\mathbb{P}}\! \left( \cup _{(k,m)} F_{k,m} \right) \leq \sum _{(k,m)}{\mathbb{P}}\!\left(F_{k,m}\right) \leq d^2 \exp\!\left({-}L(d+1)^{-1} \right) \leq \varepsilon$ , provided that $L$ (and $D$ from above) are sufficiently large.

3.4 Completing the proof of Theorem 1.2

The following lemma is the final key step in our proof.

Lemma 3.6. Given $C\gt 1$ , there are constants $a_C \gt 0$ and $n_C\in \mathbb{N}$ such that the following holds. Every $C$ -bipartite-Ramsey graph $G = (V_1, V_2, E)$ with $|V_1|,|V_2|\geq n_C$ has:

\begin{equation*}\big [a_C|V_1||V_2|, 2a_C|V_1||V_2| \big ] \subset \big \{e(G[U])\,:\, U \subset V(G) \big \}.\end{equation*}

Proof. We start by fixing parameters $1\geq C^{-1} \gg \varepsilon \gg \delta \gg d_0^{-1} \gg L^{-1} \gg C_0^{-1} \gg c \gg n_C^{-1} \gt 0$ . Without loss of generality, we may assume that $|V_1| \geq |V_2|$ . We also fix an arbitrary set $V^{\prime}_2 \subset V_2$ with $|V^{\prime}_2| = c|V_2|$ and a partition $V_1 = U_1 \cup U_2 \cup U_3$ , where $|U_i| \geq |V_1|/4$ for $i\in [3]$ .

To begin, we claim that there are ${\mathcal{P}}_1,\ldots,{\mathcal{P}}_K$ with $K \,:\!=\, C_0|V_2|^{0.5}$ such that each ${\mathcal{P}}_i$ is either an $\varepsilon$ -pair-star or a $\varepsilon$ -pair-matching in $G\left[U_1, V^{\prime}_2\right]$ of size $|V^{\prime}_2|^{0.5}$ and so that the ${\mathcal{P}}_i$ ’s are all vertex disjoint. To see this, suppose that we have found ${\mathcal{P}}_1,\ldots,{\mathcal{P}}_i$ and now seek ${\mathcal{P}}_{i+1}$ . Let $U^{\prime}_1$ denote the subset of $U_1$ with the vertices of $\cup _{j\leq i}{\mathcal{P}}_j$ removed and apply Lemma 3.1 to $G\!\left[U^{\prime}_1, V^{\prime}_2\right]$ , noting that as $|U^{\prime}_1| \geq |U_1| - 2 K |V^{\prime}_2|^{0.5} \geq |U_1| - 2(C_0|V_2|^{0.5}) (c|V_1|)^{0.5} \geq |U_1|/2 \geq |V_1|/8$ , making $G\!\left[U^{\prime}_1, V^{\prime}_2\right]$ a $2C$ -Ramsey graph. By Lemma 3.1, we can find a $\varepsilon$ -star-pair or a $\varepsilon$ -pair-matching ${\mathcal{P}}_{i+1}$ of size $|U^{\prime}_1|^{0.5} \geq |V^{\prime}_2|^{0.5}$ , which gives the desired set (perhaps after removing some of its elements). Thus such a collection ${\mathcal{P}}_1,\ldots,{\mathcal{P}}_K$ exists.

Our next step is to apply Lemma 2.3 to $G\!\left[U_2, V^{\prime}_2\right]$ , which is again $2C$ -Ramsey, to find a set $S_2 = \{s_1,\ldots, s_L\} \subset U_2$ such that $\big|N(s_i) \setminus ( \cup _{j\lt i}N(s_j) )\big| \geq |V_2|^{0.5}$ for all $i\in [L]$ . It is possible to do this since $C^{-1} \gg L^{-1} \gg n_C^{-1}$ .

The final step of preparation is to select a set $W \subset V^{\prime}_2$ uniformly at random. For each $i\in [K]$ let $A_i \,:\!=\, A_{{\mathcal{P}}_i}^W$ . Now consider the following events:

  • Let ${\mathcal{E}}_1$ denote the event that, for at least $K/2$ values $i\in [K]$ , we have $|A_i| \geq \delta |V^{\prime}_2|^{1/2}$ . By Lemmas 3.3 and 3.4, and using that $\varepsilon \gg \delta$ , we see that for any $A_i$ we have $|A_i| \geq \delta |W|^{1/2}$ with probability at least $3/4$ , so a simple application of Markov gives ${\mathbb{P}}({\mathcal{E}}_1) \geq 1/2$ ;

  • Secondly, let ${\mathcal{E}}_2$ denote the event that, for every $0 \leq k \leq m$ and $2 \leq m \leq d_0$ , there exists $s \in S_2$ with $d^W_G(s) \equiv k \mod m$ . By the choice of $S_2$ , Lemma 3.5 gives ${\mathbb{P}}({\mathcal{E}}_2) \geq 7/8$ ;

  • Lastly, let ${\mathcal{E}}_3$ denote the event that $|W| \geq |V^{\prime}_2|/4$ . By Chebyshev’s inequality ${\mathbb{P}}({\mathcal{E}}_3) \geq 7/8$ .

It follows from the union bound that ${\mathbb{P}}({\mathcal{E}}_1 \cap{\mathcal{E}}_2 \cap{\mathcal{E}}_3) \geq 1/2$ . Fix a choice of $W \subset V^{\prime}_2 \subset V_2$ for which all three events occur. By reordering, we get that $|A_i| \geq \delta |V^{\prime}_2|^{1/2} \geq \delta |W|^{1/2}$ for $i\in [K/2]$ .

We are now ready to complete the proof. Take $V_1^{\mbox{init}}$ to be the union of the heads of ${\mathcal{P}}_i$ , i.e. $V_1^{\mbox{init}} \,:\!=\, \cup _{i \leq L/2} H({\mathcal{P}}_i)$ , and $e_0 \,:\!=\, e(G[V_1^{\mbox{init}}, W])$ . Take $a \in A_i$ and note the following:

  1. (i) If ${\mathcal{P}}_i = \{x_0,\ldots, x_{|V^{\prime}_2|^{0.5}}\}$ is a pair-star rooted at $x_0$ , then by definition $d^W(x_j) - d^W(x_0) = a$ for some $j\in \left[\big|V^{\prime}_2\big|^{1/2}\right]$ . Removing $x_0$ from $V_1^{\mbox{init}}$ and adding $x_j$ changes the number of edges in the resulting graph by exactly $a$ .

  2. (ii) Similarly, if ${\mathcal{P}}_i$ is a pair-matching, then there is ${\bf p} = (x,y) \in{\mathcal{P}}_i$ such that $\operatorname{deg-diff}^{W}\!({{\bf p}}) = a$ . Removing $x$ from $V_1^{\mbox{init}}$ and adding $y$ to it changes the number of edges by exactly $a$ .

  3. (iii) The edits from different ${\mathcal{P}}_i$ ’s can be performed independently of one another, resulting in the same changes to the edge size given in (i) and (ii).

From observations (i) $-$ (iii) we can immediately deduce that:

\begin{equation*}\{e_0\} + A_1 + A_2 + \cdots + A_{K/2} \subset \big \{ e(G[U,W])\,:\, U \subset U_1 \big \}.\end{equation*}

By definition of $A_i\,:\!=\,A_{{\mathcal{P}}_i}^W$ , we have $A_i \subset \left[{-}3|V^{\prime}_2|^{1/2}, 3|V^{\prime}_2|^{1/2}\right] \subset \left[{-}6|W|^{1/2}, 6|W|^{1/2}\right]$ and $|A_i| \geq \delta |W|^{1/2}$ . By taking $M = 6|W|^{1/2}$ and $B = 1$ in Lemma 2.12, as $\delta,1 \gg C_0^{-1}, d_0^{-1}$ , it follows that there is $a \in{\mathbb Z}$ and $1 \leq d \leq d_0$ such that (2) holds, which gives:

(3) \begin{align} \left\{e_0 + a + id\,:\, 0 \leq i \leq 3|W|\right\} \subset \{e_0\} + \left\{a + id\,:\, 0 \leq i \leq M^2\right\} & \subset \{e_0\} + A_1 + A_2 + \cdots + A_{L/2} \nonumber \\ & \subset \left\{e(G[U, W])\,:\, U \subset U_1 \right\}. \end{align}

Furthermore, as $d \leq d_0$ , by the choice of $S_2$ there are $s_{i_0}, s_{i_1},\ldots, s_{i_{d-1}} \in S_2$ with $d^W(s_j) \equiv j \mod d$ for each $j\in [0,d-1]$ . Combined with (3), setting $e_1 = e_0 + a + |W|$ , this gives:

(4) \begin{align} \big [e_1, e_1 + |W| \big ] \subset \big \{e(G[U \cup \{s\}, W])\,:\, U \subset U_1, s \in S_2 \big \}. \end{align}

To finish the proof, note that $|U_3| \geq |V_1|/4 \geq |V_1|^{1/2}$ and $|W| \geq |V^{\prime}_2|/4 \geq |V_2|^{1/2}$ . As $G$ is $C$ -Ramsey, we get that $G[U_3, W]$ is $(2C)$ -Ramsey. It follows from Corollary 2.2 that at least $2|U_3|/3$ vertices in $U_3$ have degrees between $(64C)^{-1}|W|$ and $(1-(64C)^{-1})|W|$ in $G[U_3, W]$ . In particular, $\{e(G[U^{\prime},W]) \,:\, U^{\prime} \subset U_3\}$ contains an element from each interval $[|W|i, |W|(i+1)]$ with $i\in [0,(96C)^{-1}|U_3|]$ . However, $e(G[U \cup \{s\} \cup U^{\prime}, W]) = e(G[U \cup \{s\}, W]) + e(G[ U^{\prime}, W])$ for any $U \subset U_1, s \in S_2$ and $U^{\prime} \subset U_3$ . Therefore, we deduce from (4) that:

\begin{align*} \big [ e_1 + |W|, e_1 + |W| + (96C)^{-1}|U_3||W| \big ] \subset \big \{ e(G[U \cup \{s\} \cup U^{\prime}, W])\,:\, U \subset U_1, s \in S_2, U^{\prime} \subset U_3 \big \}. \end{align*}

We can finally claim that the lemma holds with $a_C \,:\!=\, c/4000C$ . To see this, note that:

\begin{align*} e_1 + |W| = e_0 + a + |W| & \leq (K/2)|V^{\prime}_2|^{0.5}|W| + (K/2)(6|W|)^{0.5}|W| + |W|\\ &\leq 4K|W|^{1.5} \leq 4(C_0|V_2|)^{0.5}(c|V_2|)^{1.5} = 4c^{1.5}(C_0)^{0.5}|V_2|^2 \leq a_C|V_1||V_2|, \end{align*}

where we have used that $|W| \leq c|V_2|$ , that $c \ll C_0^{-1}, C^{-1}$ , and that $|V_2| \leq |V_1|$ .

Since $a_C|V_1||V_2| = (c/4000C)|V_1||V_2| \leq (200C)^{-1}|U_3||W|$ , it is quite easy to observe that $[a_C|V_1||V_2|, 2a_C|V_1||V_2|] \subset \{e(G[U])\,:\, U \subset V(G)\}$ , as required. This completes the proof.

With Lemma 3.6 in hand, it is now easy to complete the proof of Theorem 1.2.

Proof of Theorem 1.2. By making $\alpha$ small enough we may assume, without losing generality, that $|V_1|\geq |V_2|$ and that $|V_2|$ is large enough so that our estimates below hold. We can also assume that $C\gt 1$ since the Ramsey condition still holds when we increase $C$ .

First note that as $G = (V_1, V_2, E)$ is $C$ -Ramsey, there is a vertex $v \in V_2$ with degree at least $(32C)^{-1}|V_1|$ in $V_1$ , by Corollary 2.2. But then the induced subgraphs of the form $G[W, \{v\}]$ , where $W \subset V_1$ , give all edge sizes in $[0,(32C)^{-1}|V_1|]$ (and so in particular the analogue of the Alon $-$ Krivelevich $-$ Sudakov theorem from [Reference Alon, Krivelevich and Sudakov1] is easier in the bipartite Ramsey context).

On the other hand, observe that by fixing any $W_1 \subset V_1$ and $W_2 \subset V_2$ with $|W_i| \geq |V_i|^{1/3}$ for $i \in \{1,2\}$ , the graph $H = G[W_1, W_2]$ is $(3C)$ -Ramsey, as $G$ is $C$ -Ramsey. It follows by applying Lemma 3.6 to $H$ that there are induced subgraphs of each size in $[a_{3C}|W_1||W_2|, 2a_{3C}|W_1||W_2|]$ . It follows, taking $M = \{(m_1, m_2)\,:\, |V_i|^{1/3} \leq m_i \leq |V_i| \mbox{ for } i \in \{1,2\}\}$ , that $G$ contains a subgraph of each size in $\cup _{(m_1,m_2) \in M} [a_{3C}m_1m_2, 2a_{3C}m_1m_2]$ . It can be easily seen that these sets together cover the interval $[a_C(|V_1||V_2|)^{1/3}, 2a_C |V_1||V_2|]$ .

Finally, note that the intervals $[0,(32C)^{-1}|V_1|]$ and $[a_{3C}(|V_1||V_2|)^{1/3}, 2a_{3C} |V_1||V_2|]$ together cover $[0, 2a_{3C} |V_1||V_2|]$ , since $|V_1| \geq |V_2|$ and $|V_2|$ is large enough, thus completing the proof of the theorem.

4. Concluding remarks

In this paper, we have proven a bipartite analogue of the Erdős–McKay conjecture, showing that any $C$ -bipartite-Ramsey graph $G = (V_1, V_2, E)$ must contain induced subgraphs of all sizes in $[0, \Omega _C(|V_1||V_2|)]$ . Of course the Erdős–McKay conjecture itself remains an outstanding problem, and we hope that some of our ideas will be useful in future approaches here.

Another interesting direction is to understand the effect of weakening the Ramsey hypothesis in Theorem 1.2, as considered already in a number of other settings (see e.g. [Reference Alon and Bollobás2], [Reference Narayanan and Tomon26], [Reference Long and Ploscaru24]). That is, what are we able to say about the edge sizes of induced subgraphs of a bipartite graph $G = (V_1, V_2, E)$ , which do not contain induced copies of $K_{t_1,t_2}$ or $\overline{K_{t_1, t_2}}$ , where $t_1$ and $t_2$ are now general parameters?

Lastly, we note that Narayanan, Sahasrabudhe and Tomon in [Reference Narayanan, Sahasrabudhe and Tomon27] showed that any bipartite graph with $m$ edges must have induced subgraphs of $\Omega (m/ \log ^{12}\!(m))$ different sizes. An obvious upper bound here for the number of such sizes is $m$ , though if $m$ is a perfect square $m = k^2$ then the complete bipartite graphs $K_{k,k}$ allows only $m/ (\!\log m)^{0.086\ldots + o(1)}$ edge sizes. The authors of [Reference Narayanan, Sahasrabudhe and Tomon27] conjecture that when $m = k^2$ the graph $K_{k,k}$ is extremal here. It would be interesting to understand whether our approach, perhaps combined with stability arguments, could be applied to improve knowledge here.

Footnotes

The second author is grateful for support through an EPSRC DTP studentship.

References

Alon, N., Krivelevich, M. and Sudakov, B. (2003) Induced subgraphs of prescribed size. J. Graph Theory 43 239251, https://onlinelibrary.wiley.com/doi/abs/10.1002/jgt.10117 CrossRefGoogle Scholar
Alon, N. and Bollobás, B. (1989) Graphs with a small number of distinct induced subgraphs, Discr. Math. 75 2330, ISSN: 0012-365x, https://www.sciencedirect.com/science/article/pii/0012365X89900745 Google Scholar
Alon, N. and Spencer, J. H. (2004) The Probabilistic Method. Second edition. Wiley, ISBN: 0471370460.Google Scholar
Axenovich, M., Sereni, J.-S., Snyder, R. and Weber, L. (2021) Bipartite independence number in graphs with bounded maximum degree. SIAM J. Discr. Math. 35 11361148.CrossRefGoogle Scholar
Axenovich, M., Tompkins, C. and Weber, L. (2021) Large homogeneous subgraphs in bipartite graphs with forbidden induced subgraphs. J. Graph Theory 97 3446.CrossRefGoogle Scholar
Baksys, M. and Chen, X. (2021) On number of different sized induced subgraphs of bipartite-ramsey graphs, arXiv preprint arXiv: 2109.08485, https://arxiv.org/abs/2109.08485 Google Scholar
Barak, B., Rao, A., Shaltiel, R. and Wigderson, A. (2012) 2-source dispersers for $n^{o(1)}$ -entropy, and ramsey graphs beating the frankl-wilson construction. Ann. Math., 14831543. doi: 10.4007/annals.2012.176.3.3 CrossRefGoogle Scholar
Bucić, M., Letzter, S. and Sudakov, B. (2019) 3-color bipartite ramsey number of cycles and paths. J Graph Theory 92 445459.CrossRefGoogle Scholar
Bukh, B. and Sudakov, B. (2007) Induced subgraphs of Ramsey graphs with many distinct degrees. J. Comb. Theory, B 97 612619, ISSN: 0095-8956. https://www.sciencedirect.com/science/article/pii/S0095895606001080 CrossRefGoogle Scholar
Calkin, N., Frieze, A. and McKay, B. D. (1992) On subgraph sizes in random graphs. Comb. Prob. Comp. 1 123134. doi: 10.1017/S0963548300000146.CrossRefGoogle Scholar
Chung, F. and Graham, R. (1998) Erdős on Graphs: His Legacy of Unsolved Problems. AK Peters/CRC Press.CrossRefGoogle Scholar
Collins, A., Riasanovsky, A., Wallace, J. and Radziszowski, S. (2016) Zarankiewicz numbers and bipartite ramsey numbers, arXiv preprint arXiv: 1604.01257.Google Scholar
Conlon, D. (2008) A new upper bound for the bipartite Ramsey problem. J. Graph Theory 58 351356.CrossRefGoogle Scholar
Erdős, P. (1947) Some remarks on the theory of graphs. Bull. Am. Math. Soc. 53 292294, doi: bams/1183510596.CrossRefGoogle Scholar
Erdős, P. and Szckeres, G. (1987) A combinatorial problem in geometry. In Classic Papers in Combinatorics (Gessel, I. and Rota, G.-C.,eds), Birkhäuser Boston, pp. 4956, ISBN: 978-0-8176-4842-8. doi: 10.1007/978-0-8176-4842-8_3.Google Scholar
Erdős, P. and Szemerédi, A. (1972) On a Ramsey-type theorem. Period Math Hung 2 295299, https://link.springer.com/article/10.1007%5C%2FBF02018669#citeas CrossRefGoogle Scholar
Erdős, P. (1992) Some of my favourite problems in various branches of combinatorics. Le Matematiche 47 231240.Google Scholar
Ferber, A., Hardiman, L. and Krivelevich, M. (2021) On subgraphs with degrees of prescribed residues in the random graph. arXiv preprint arXiv: 2107.06977.Google Scholar
Jenssen, M., Long, E., Keevash, P. and Yepremyan, L. (2020) Distinct degrees in induced subgraphs . Proc. Am. Math. Soc. 148 38353846. doi: 10.1090/proc/15060. ISSN:0002-9939.CrossRefGoogle Scholar
Kovari, T., Sós, V. and Turán, P. (1954) On a problem of K. Zarankiewicz. Colloquium Mathematicae 3 5057.CrossRefGoogle Scholar
Kwan, M., Sah, A., Sauermann, L. and Sawhney, M. (2022) Anticoncentration in ramsey graphs and a proof of the erdős–mckay conjecture. arXiv preprint arXiv: 2208.02874.Google Scholar
Kwan, M. and Sudakov, B. (2019) Proof of a conjecture on induced subgraphs of Ramsey graphs. Trans. Am. Math. Soc. 372 55715594. doi: 10.1090/tran/7729.CrossRefGoogle Scholar
Kwan, M. and Sudakov, B. (2017) Ramsey graphs induce subgraphs of quadratically many sizes . Int. Math. Res. Not. 2020, doi: 10.1093/imrn/rny064.CrossRefGoogle Scholar
Long, E. and Ploscaru, L. (2022) Distinct degrees and homogeneous sets. arXiv preprint arXiv: 2204.05932, p. 36, https://arxiv.org/abs/2204.05932.Google Scholar
Narayanan, B., Sahasrabudhe, J. and Tomon, I. (2019) Ramsey graphs induce subgraphs of many different sizes. Combinatorica 39 215237. doi: 10.1007/s00493-017-3755-0.CrossRefGoogle Scholar
Narayanan, B. and Tomon, I. (2018) Induced subgraphs with many distinct degrees . Comb. Prob. Comp. 27 110123. doi: 10.1017/S0963548317000256 CrossRefGoogle Scholar
Narayanan, B.P., Sahasrabudhe, J. and Tomon, I. (2017) The multiplication table problem for bipartite graphs. Combinatorica 37(5) 9911010. DOI 10.1007/s00493-016-3322-0.CrossRefGoogle Scholar
Norris, J. R. (1997) Markov Chains, Cambridge Series in Statistical and Probabilistic Mathematics. Cambridge University Press, doi: 10.1017/CBO9780511810633.CrossRefGoogle Scholar
Prömel, H. J. and Rödl, V. (1999) Non-ramsey graphs are clogn-universal. J. Comb. Theory, A 88 379384, ISSN: 0097-3165. https://www.sciencedirect.com/science/article/pii/S0097316599929722 CrossRefGoogle Scholar
Ramsey, F. P. (1930) On a problem of formal logic. Proc. London Math. Soc. s2-30 264286, https://londmathsoc.onlinelibrary.wiley.com/doi/abs/10.1112/plms/s2-30.1.264 CrossRefGoogle Scholar
Shelah, S. (1998) Erdős and rényi conjecture. J. Comb. Theory, Ser. A 82 179185. doi: 10.1006/jcta.1997.2845.CrossRefGoogle Scholar
Souza, V. (2021) Blowup ramsey numbers. Eur. J. Comb. 92 103238.CrossRefGoogle Scholar
Szemerédi, E. and Vu, V. H. (2005) Long arithmetic progressions in sum-sets and the number of x-sumfree sets. Proc. Lond Math. Soc. 90 273296. doi: 10.1112/S0024611504015059.CrossRefGoogle Scholar
Tao, T. and Vu, V. H. (2006) Additive Combinatorics, Vol. 105. Cambridge University Press.CrossRefGoogle Scholar