heightadjust=object, valign=c
Size-Ramsey numbers of structurally sparse graphs
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 -vertex graphs with constant maximum degree . 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 to 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 ). As a special case, this yields a bound of 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 , there exists a finite number such that every complete graph on vertices whose edges are colored in red or in blue contains a monochromatic clique of order [26]. Motivated by this, we say that a host graph is -Ramsey for a graph if any -coloring of the edges of yields a monochromatic copy of , and we write .
Given a graph , its size-Ramsey number is defined as the smallest number of edges in a graph such that . This notion measures the minimality of the host graph 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 , answering a -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 on vertices fulfilling for a universal constant . Their lower bound has been improved very recently to 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 -vertex graphs with for some absolute constant remains open.
Turning to general upper bounds, Kohayakawa, Schacht, Rödl and Szemerédi [23] showed that for graphs which have maximum degree it holds that , leaving a wide gap between the best known upper and lower bounds. Recently, Allen and Böttcher [1] announced an improvement of this bound to for . For the case , the first and fourth authors [14] showed that . 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 [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 to grow with , then thus far no better upper bounds on 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 and let be an -vertex graph of constant maximum degree and treewidth . Then
-
•
-
•
Furthermore, if then we have111We use to denote all functions which are in for some absolute constant . .
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 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 , and hence we have the following consequence of Theorem 1.1.
Theorem 1.2.
Let and let be a proper minor-closed class. Then for every -vertex graph of maximum degree , we have
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 -vertices was , which was improved to 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 , we use to indicate that is sufficiently large as a function of for our proofs to work out. For example, we often use inequality chains like , which also implies that in particular . We make use of the strong product operator, which is defined as follows. The strong product of graphs and is the graph with vertex set such that for and , we have that is adjacent to in if and , or and , or and .
2 Structural results
In this section, we prove the announced structural lemma for graphs of bounded treewidth. It says that every bounded-degree -vertex graph of treewidth at most can be embedded as a subgraph into the strong product of a bounded-degree tree of size and a clique of size . In the case that belongs to a structurally sparse class of graphs, such as planar graphs or a proper minor-closed class, we can get rid of the 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 . There exists depending solely on such that the following hold.
-
(i)
For every -vertex graph of maximum degree at most and treewidth at most there exists a tree such that , where , and .
-
(ii)
Let be a class of graphs closed under taking subgraphs, and let be an increasing function in such that every -vertex graph has treewidth at most . Suppose further that there exists a constant such that holds for every . Then for every -vertex graph of maximum degree at most there is a tree such that , where , and .
We remark that a very similar embedding lemma, without the factor, has been proved before by Ding and Oporowski [10] and later on improved by Wood [30]. These results state that every graph of maximum-degree and treewidth at most is contained in , where is some tree and . However, for our purposes, it is very important to be able to carefully control both the size and the maximum degree of the tree . In particular, we need 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 and let be a proper minor-closed class. There exist constants such that every -vertex graph of maximum degree at most is a subgraph of , where and is a tree with and .
- Proof.
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 of treewidth 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 be a graph of treewidth at most and order . Then there exists a set of size and a partition of into sets and such that no edge in connects a vertex in to a vertex in and such that .
To prove Lemma 2.1, we first establish an extension of Lemma 2.3 which works when the graph comes with a given coloring of its vertices with colors, and we wish to find a small separator in the graph such that the two sides and 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 , let be a (partial) -coloring of the first integers. Then can be partitioned into pairwise disjoint intervals such that for some it holds that (denoting )
for .
Lemma 2.5.
Let be a graph and let be an increasing function such that every subgraph of with vertices has treewidth at most .
Then there exists a linear ordering of the vertices in and for each a set including such that
and such that there are no edges between and in .
-
Proof.
We prove the statement by induction on . The claim is trivially true for by setting , so moving on suppose that and the claim is true for all subgraphs of on fewer vertices. Apply Lemma 2.3 to find a set of size at most and a partition of into parts and with no edges in between them, such that . By the inductive assumption, there exist linear orderings and of and and for each and sets and such that the following hold. separates and in , and separates and in , and with we have
Let be an arbitrary linear ordering of , and consider the following linear order on : . Furthermore let for , for , and for . It is easily verified that each of these sets separates their predecessors and successors in the ordering in and since , each of these sets is of size at most . This concludes the proof. ∎
Using the previous statement and Lemma 2.4, we immediately arrive at the following corollary.
Corollary 2.6.
Let be a graph given with a (partial) coloring of its vertices for some and let be an increasing function such that every subgraph of with vertices has treewidth at most . Then there is a set and a partition of such that
there are no connections between and in , and
for each .
-
Proof.
We apply Lemma 2.5 to to get the linear ordering of . Next, we apply Theorem 2.4 to the coloring . Let be an interval partition of as guaranteed by the lemma. Let be a list of elements in such that any two distinct intervals in are separated by at least one of them. We now define , which by Lemma 2.5 is of size at most , and let . It now follows directly from Lemma 2.5 that there are no edges in connecting and and directly from the properties guaranteed by Theorem 2.4 that for each . In order to reduce from the above bound to the desired for each , we may simply move for each color a vertex of that color from either or into , thereby increasing the size of by at most , 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 be a graph of maximum degree at most and let be an increasing function such that every -vertex subgraph of has treewidth at most . Let
Then there exists a binary tree and a partition of such that for every non-leaf , for every leaf , and such that any two adjacent vertices in are contained in sets for vertices at distance at most in .
-
Proof.
To enable induction, we prove a slight strengthening of the statement in the lemma as follows:
Claim.
Given any sequence of disjoint subsets of such that for each , we can find a tree and a partition as in the lemma with the following additional property: For each and every vertex in at distance at least from the root, we have .
Let us now prove this claim by induction on . It trivially holds true if , so suppose that and the claim is satisfied by every subgraph of with fewer than vertices. Interpret the disjoint sets as the color classes of a partial coloring of , and apply Corollary 2.6 to find a set of size at most in and a partition of such that for each . Pick a set of size exactly in such that (this is possible since and ). Now, let , . For each , define , . Further, let and . From the above we have for , as well as since is of maximum degree at most and by our choice of .
Therefore, we can apply induction to with the sets and to with the sets to find binary trees and , and partitions of and satisfying the inductive claim. We now form a binary tree by adding a new root vertex and connecting it to the roots of the binary trees and , and setting . We claim that this tree and the partition satisfy the claim. Let us first verify the condition with respect to the sets . For , and any vertex at distance at least from in , we have that is disjoint from . Further, for , every vertex at distance at least from in has distance at least from the root in either or , thus implying that is disjoint from respectively , and therefore in each case.
Hence, all that remains to be verified is that any two adjacent vertices in are contained in sets for vertices at distance at most in . So consider any two vertices in such that there exists an edge in with endpoints in and . If then either both of are contained in or both in , and then it follows from the validity of the inductive claim for and that and are at distance at most in . Next consider the case , and w.l.o.g. assume . Then , so in particular intersects or at least one of , let’s say for some . But then by our inductive call to , has to be at distance at most from the root in . But this means that is at distance at most from the root in , showing that and are at distance at most also in this last case. Thus, satisfies the inductive assertion with respect to , 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 on vertices has treewidth at most , where we let for each if we are in case (i) of the lemma. Let and be defined as in Lemma 2.7 and note that we obtain in case (i) and
hence in case (ii).
We can now apply Lemma 2.7 to find a binary tree and a partition of the vertex set of such that for every non-leaf of , for every leaf of , and such that whenever and have a connecting edge in , then and are at distance at most in . Note that the number of non-leaves in is at most , and since a binary tree of size at least has at most one more leaf than non-leaf, we find that . Furthermore, we can see from the above that is a subgraph of , where is the graph on the same vertex set as in which two vertices are adjacent iff they are at distance at most from each other in . Notice that since is a binary tree, it is not hard to show that where is a tree such that and . All in all, we find that . Since , the assertion of the lemma in both cases now follows from the above asymptotic estimates of . ∎
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 , and for every large enough, there exists a graph with vertices and constant maximum degree such that for every -vertex graph of maximum degree and treewidth at most , it holds that .
The theorem above is stated in [4] without the maximum degree condition on , 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 and . For a graph and disjoint sets , we say that is -dense if for any with and , we have
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 -uniformity as follows.
Definition 3.3.
A graph is -uniform if for all disjoint with , we have . If is bipartite with bipartition , then we say that is -uniform if the same holds for all and of size at least and 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 and , there exists such that the following holds for every . Let be a graph on at least two vertices with and let be obtained by replacing every with an independent set of sufficiently large order and every by a -uniform bipartite graph between and . Then, for every -colouring of the edges of , there exists a -colouring of the edges of and, for every , a subset of order such that is -dense for each where stands for the edges of colour between and in .
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 be positive integers and let be an -vertex graph with , such that for some and a tree with and . Then .
Lemma 3.6.
Let be positive integers and let be an -vertex graph with , such that for some and a tree with and . Then .
-
Proof of Theorem 1.1.
Let be an -vertex graph of maximum degree and treewidth . By Lemma 2.1, we know that for and , with . Now, by Lemma 3.5, the size-Ramsey number of is , which gives the first part of the theorem.
For the second part, assume . Note that by Lemma 3.6 and since , we have , 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 we know our graph has. Namely, we take a graph that is Ramsey for a constant blow-up of and blow it up by a factor of roughly . 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 . Next, using that we picked to be Ramsey for a constant blow-up of , we find a monochromatic copy of the latter in the auxiliary coloring, which corresponds to a monochromatic -shaped structure of lower-regular pairs in our host graph. We can then use an approach similar to the embedding lemma to embed into that monochromatic structure, being guided by the containment of in to map each vertex to an appropriate set.
-
Proof of Lemma 3.5.
Set . Let and note that has treewidth at most by the definition of treewidth. Let be the graph given by Theorem 3.1 which is -Ramsey for and which has vertices and constant maximum degree. The host graph for which we will consider is . Note that has edges. For each , denote the copy of corresponding to by . Consider an arbitrary -coloring of the edges of .
We apply Lemma 3.4 with and to the graph , which is a blow-up of . Hence, we get a coloring of the edges of and, for each , a subset of size , such that for every edge in , the edges between and form an -dense pair in the color of the edge (meaning that the density between every two sets and of sizes and is at least ).
Now, since is -Ramsey for , there is a monochromatic copy of in , which in turn means that for every edge of that copy, the pair and forms a -dense pair in the same color, say red. As a last step before describing the embedding process, for each , denote with the subgraph of which lives inside of the copy of that corresponds to the vertex (recall that ). Since has maximum degree , it is -partite, so denote by one such partition of , for each . To summarize, we now want to embed into the red subgraph of which is obtained from by replacing the complete bipartite graphs between the relevant copies of by -dense pairs.
The embedding of into the found red subgraph of is done by a standard procedure, which is a variant of the well-known embedding lemma. Briefly, for each , we will want to embed into , so that each is embedded into its own copy of in . This embedding can be done vertex by vertex, crucially maintaining the following invariant. For each vertex which has been assigned to the copy of but has not yet been embedded, we consider the ‘candidate set’ that consists of all the vertices in that are in the common neighbourhood of all already embedded neighbours of in . While embedding , we ensure that for each such , the size of the candidate set is at least , where is the number of neighbours of in which are already embedded at the observed point of the embedding process. When it is time for vertex to be embedded, we choose a vertex for that purpose, in such a way that has enough neighbours in the candidate set of each neighbour of in 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 and , at most of the vertices in have degree less than into . Additionally, at most many vertices in are already taken by vertices previously embedded in , leaving us with at least
viable candidates for the embedding of , where we used that and . Thus, we can continue doing this until we embed the whole graph . 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 into a certain blow-up of 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 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 and be large enough. Let and . Set and . Then w.h.p. is such that the following holds. Let be an -vertex graph of maximum degree that is a subgraph of for some -vertex graph with . Let and be a graph with vertex set such that for each and edge set consisting of an -dense bipartite graph between and for each . Then can be embedded in .
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 . This drop in density then requires a much more careful treatment of the embedding process of , for which the embedding lemma no longer suffices. This is where Lemma 3.7 proves to be helpful.
-
Proof of Lemma 3.6.
Let be as given above. We shall show that there exists a graph with many edges such that .
It will be convenient to define the following constants. Let and let
Let . Let be as given by Lemma 3.4 with , , and .
Let be the graph that is -Ramsey for all -vertex graphs of maximum degree and maximum treewidth given by Theorem 3.1. In particular, , since by the definition of treewidth, has treewidth at most . Note that has vertices and constant maximum degree, say . We define our host graph to be obtained by replacing each with an independent set on vertices and each edge with a random bipartite graph between and in which each edge is added independently at random with probability .
Now consider an arbitrary colouring of the edges of in colours. We will show that can be embedded into one of the colour classes of . By a standard application of Chernoff bounds and a union bound, with high probability each random bipartite graph between and with is -uniform. We fix an outcome of for which this is the case. In particular, this also implies that .
We can now apply Lemma 3.4 with , , , and . This yields a -colouring of the edges of and, for every , a subset of size such that for each , the graph is -dense, where are the edges in colour .
Next, note that by Theorem 3.1, the -colouring of the edges of contains a monochromatic copy of . This implies that contains a monochromatic subgraph with vertex set with for each and edge set consisting of an -dense bipartite graph between and for each .
We can now apply Lemma 3.7 with and , concluding that the graph contains as a subgraph. Since the copy of in 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, and , where density of pairs is not inherited on one- respectively two-sided neighborhoods.222The definition slightly deviates from Definition 13 in [23], where the pairs and are both -dense. Here, for technical reasons, we include the more general case of -dense pairs whereby can differ from .
Definition A.1.
Let integers and reals , , and be given.
-
1.
Let be the family of tripartite graphs with vertex set , where , , and , satisfying
-
(a)
is a -dense pair,
-
(b)
is a -dense pair, and
-
(c)
there exists with such that for every the pair is not -dense.
-
(a)
-
2.
Let be the family of tripartite graphs with vertex set , where , , and , satisfying
-
(a)
and are -dense pairs,
-
(b)
is a -dense pair, and
-
(c)
there exists with such that for every the pair is not -dense.
-
(a)
Graphs that do not contain large subgraphs — induced or not — from the two above ‘bad’ families are then said to have a denseness property , which we make precise in the next definition.
Definition A.2.
For integers and and reals and we say that a graph with has the denseness property , if contains no member from
with and 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 enjoy this denseness property at various parameter scales, whenever . It will not be enough for our purposes that this fails to hold with any probability . 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 and all reals , there exist , , and satisfying such that if , then .
-
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 where the two parts and are of fixed size, namely only a -fraction of the size of (the more general case can be reduced to this one). The bipartite subgraphs and are quite dense by assumption. Now consider the violating set , whose members all fail to produce -dense pairs . We restrict our attention to such that each has ‘large’ degree into . Because is -dense, contains at least half of the vertices in . Now one can show that each such vertex produces a set of size which, together with produces a pair of density strictly less than . Any superset of in with size then produces a violating pair, that is, the pair is not -dense. By repeating this procedure for all , we could produce an entire family of violating pairs, which is unlikely to occur in . The latter can be proven by a union bound via counting the ways of choosing the vertices in the involved sets and the edges in the bipartite graph as well as the choices of the sets , which are limited by sparse regularity inheritance from (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 , add ”there exists ” and change
to
-
–
In the proof of Corollary 16, add the argument to .
-
–
Change the definition of to include as an argument, in agreement with .
-
–
In Proposition 17, add ”for each ” and change the probability condition to . In particular, ensures that is large enough, see the change below on the lower bound at the top of page 5052.
-
–
In the proof of Proposition 17:
-
*
Choose , which was originally set to , to be, say, 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 , which depends linearly on , enters as in the exponent. See also the change on page 5051 indicated below.
- *
-
*
On page 5050, ”Because of the assumption on , the bipartite subgraphs and of contain at least edges each” — this continues to hold for but is no longer true of , instead it contains at least edges.
-
*
On page 5050, ”From the -denseness …” towards the bottom of the page changes to ”From the -denseness …” and changes to .
-
*
On page 5051, we get that the probability that appears in is at most as a result of the new choice of .
-
*
On top of page 5052, it is no longer true that and have at least edges each, they have instead at least edges each.
-
*
On top of page 5052, in the lower bound for , we get instead , from the -denseness of and .
-
*
On page 5052, we get that the probability that a graph from is contained in is at most .
-
*
-
–
In the proof of Proposition 15:
-
*
Change the third sentence to ”Indeed, roughly speaking, we show that each ’bad’ tripartite graph with arbitrary contains a subgraph for some appropriate and ”.
-
*
In Claim 18, add to the list of given positive reals, add ”there exists ”, and add as an argument to and as an argument to .
-
*
In the paragraph after Claim 18, add ”Pick for the application of Claim 18”, and add as an argument to and ; also change the probability to .
-
*
-
–
In the proof of Claim 18:
-
*
Add to the list of given constants, and take .
-
*
On page 5053, in the definition of , add as an argument to .
-
*
On page 5053, replace with everywhere in the sentence ”Owing to the choice of …” (we can do this since ).
-
*
On the bottom of page 5053, change the last sentence of the penultimate paragraph to ”We shall show that with positive probability is from .
-
*
In the last paragraph of page 5053, replace with when it comes to pairs and .
-
*
On the bottom of page 5054, add as an argument to .
-
*
∎
-
–
Let be a graph and be an integer. We define an auxiliary bipartite graph by
The next definition and lemma will help us conclude that if is the random graph , then has no ”dense parts”.
Definition A.4.
Let integers and and reals and be given. A graph with has the congestion property if for every and every family of pairwise disjoint -sets with
-
•
and
-
•
we have
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 and real , there exists such that if , then .
-
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 to .
-
–
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 .
-
–
At the end of Case 2 in the proof of Proposition 11, one can upper bound the probability of the bad event by , for large enough.
∎
-
–
-
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 for its embedding. Here we use the assumption that is a subgraph of . We first partition the vertices of according to the copy of , indexed by the vertices of , into which they are mapped. This ordered partition is then refined using a coloring (of the third power of ), ensuring that no two vertices at distance at most with respect to are mapped to the same class. The actual embedding of into is then constructed inductively, maintaining at each step for every not yet embedded vertex of a large enough candidate set which takes into account all the already embedded neighbors of , and ensuring that the candidate sets for pairs of not yet embedded vertices that are neighbors in have candidate sets which induce an appropriately regular pair in . Given a vertex and a candidate set , a naive embedding of 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 for each in the current vertex class of simultaneously.
Notation and Constants. We first fix some notation. In , for each , we denote by the copy of that corresponds to , and by the copy of in that corresponds to .
Next, we choose the constants. Let , , and . Let be as given by Lemma A.3 applied with , , and
be as given by Lemma A.3 applied with , , , , . The value of 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 will ensure that the candidate sets are large enough to form dense pairs. The choice of will enable us to verify Hall’s condition.
For each distinct such that , apply Lemma A.3 with , , , , , and to , which is a subgraph of . By a union bound over all at most such triples , we conclude that w.h.p. for each , we have , since
From now on we condition on this being the case.
Similarly, for each , we apply Lemma A.5 with , and to , which we can do since it is a subgraph of . We next apply a union bound over all such vertices , to show that the probability that for some , we have is at most
From here on we condition on this event not occurring.
Preparation of . We start by preparing for the embedding. For this, consider an embedding . We partition into classes with for if belongs to the copy of in that corresponds to . We denote . Let be an arbitrary ”root” of and consider a breadth-first ordering of starting from , namely . We will embed into for each in this order. For each , let be the ”parents” of , namely those of its neighbours that come before in the order . Note that . With slight abuse of notation, denote .
Consider the third power of , that is, if and only if and there is a path between and in consisting of at most edges. Note that and therefore . Fix some and let be a -vertex colouring of . Then partitions into classes such that if two vertices are in the same class, they have distance at least in , and so there are no edges between and . Next, we refine this partition depending on the ”left-degrees” of the vertices. We say that are equivalent if and
Note that vertices of will be embedded in order of increasing colour according to . Thus, two vertices in are equivalent if they have the same colour by and the same number of neighbours that will be embedded before them. This equivalence relation partitions into at most many classes. Denote the partition classes by , some of which may be empty, and let be the corresponding partition function, that is, for , we have if and only if . Note that are ordered in increasing order of . Each will be embedded into a different with , but for simplicity we now rename these vertex classes so that we can treat all of them at the same time.
Set . Then let be obtained by concatenating the sequences in this order, and again let be the corresponding partition function, that is, if and only if for . Thus, if , then
Let and . Then we define the left-degree of with respect to and as
Inductively embedding into . We proceed to embed into . For this, we first relabel the vertex classes by ordering them as , where in the order of that we fixed above, for each we have that are relabelled as (in arbitrary order). We will proceed by induction, embedding into for each . Note that even though is not a complete graph, due to the fact that is a subgraph of and the way we picked and , we have that if and and , then there is an -dense bipartite graph between and .
Throughout our embedding, we make sure that the following statement holds for :
satisfying The inductive statement tells us that the first classes of can be embedded into in such a way that for every not yet embedded vertex , there is a large enough candidate set that consists of the intersection of and the neighbourhoods of all already embedded neighbours of . The third condition of additionally ensures that all pairs of candidate sets inherit regularity.
Note that from it follows that can be embedded into , which means that showing that hold completes the proof of Lemma 3.7. We next prove that holds for by induction.
Basis of the induction: . We then have that is the empty embedding and for every by (a). This implies that property (b) also holds, as and . Property (c) holds since, as mentioned above, for each , there is an -dense graph between and .
Induction step: . Suppose that holds for some . We will now extend to an embedding that satisfies (a), (b), and (c) to show that holds. We do this in the following way. Firstly, for each , we identify a subset such that if is embedded into some vertex from , then for every ”right-neighbour” of in , the candidate set will be such that properties (b) and (c) will still hold.
However, since if , we have , we cannot greedily pick an embedding of each into , as this way some 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 , and then we will set to be the representative of . We now describe these two steps in detail.
Let , and set
We say that is bad (that is, it will not be in ) if there is some vertex such that the set does not satisfy condition (b) or (c) of and therefore cannot become .
We start by focusing on (b) of . Let . We know by (c) of that is an -dense pair. Thus, there are at most vertices in such that
Considering all in the same way, it follows that for all but at most vertices , the following holds for all :
where for the penultimate inequality we used condition (b) of and the final inequality follows from our choice of .
Next, we consider property (c) of . Let with and with at least one of in . The number of these edges is at most . Let be such that , and let . If , we have
because each of has at least two neighbours in . Property (b) of then implies
by our choice of . Since either or (and the same holds for and ), we know that is a subgraph of some graph in . Thus, there are at most vertices in such that is not -dense.
Suppose now that only one of is in , say and . Then we have
Therefore, analogously to above,
Again, since either or (and the same holds for ), is a subgraph of some graph in . Thus, there are at most vertices such that is not -dense.
Note that in both cases, , and so any pair that is -dense is also -dense. For some fixed , set if and if , and define in the same way.
What we have shown so far is that there are at least
vertices with the properties
(b’) (c’) Denote the vertices in that satisfy (b’) and (c’) by . For all , since , and by our choice of the partition , we have that . Let
and note that . Then by our choice of and and from condition (b) of , we have
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 exist. For this, we use Hall’s condition [9], which tells us that in a bipartite graph with bipartitions and , there is a perfect matching if for every , it holds that . Thus, in our case it is enough to show that for every , it holds that
In case , the above holds due to our lower bound on each .
Now let be a set with . As mentioned above, for each , we have , so there is a -tuple of neighbours of that are already embedded. Moreover, since we chose the partition in such a way that the distance in between any two vertices in the same class is at least , we have that for every , the sets and are disjoint. Therefore, the sets of already embedded vertices and are disjoint too, and so is a family of disjoint sets of size in . Furthermore,
Set
and observe that if for some , then and
Assume for contradiction that
Since we conditioned on , and we have , it follows that
From our lower bound on above, it follows that
The last two inequalities together with imply that
which contradicts our assumption that . This shows that Hall’s condition holds, as desired. Therefore, there is a system of representatives for . In other words, there is an injective function such that for each .
We can now extend to get and define for . Firstly, let
Since all pairs of vertices in are at distance at least , each has no more than one neighbour in . Thus, for each such we can set
We now verify that and for each satisfy .
From property (a) of together with for every and the fact that is injective, it follows that is a partial embedding of into .
We now check that parts (a) and (b) of hold. Fix some . In case , we have and , so (a) and (b) of are satisfied for that . Now suppose (recall that this is the only other possible case). Since we defined , property (a) of holds for in this case. Furthermore, by our choice and by (b’), it follows that property (b) of is satisfied.
Lastly, we check property (c) of . Fix an edge of such that . There are three cases to consider, based on the size of and (recall that each of them has size either or ). Suppose first and . Then we have and , as well as and , so part (c) of follows directly from part (c) of . Next, suppose and . Then property (c’) together with the definition of and imply part (c) of . Finally, suppose and . Then it must be the case that , since otherwise and would be two distinct vertices in connected by a path of length in , namely the path , and this is impossible by our choice of . Then in this case also, part (c) of follows from (c’) and our choice of and .
We have shown that (a), (b), and (c) of 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 -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.