Ramsey numbers of sparse digraphs
Abstract
Burr and Erdős in 1975 conjectured, and Chvátal, Rödl, Szemerédi and Trotter later proved, that the Ramsey number of any bounded degree graph is linear in the number of vertices. In this paper, we disprove the natural directed analogue of the Burr–Erdős conjecture, answering a question of Bucić, Letzter, and Sudakov. If is an acyclic digraph, the oriented Ramsey number of , denoted , is the least such that every tournament on vertices contains a copy of . We show that for any and any sufficiently large , there exists an acyclic digraph with vertices and maximum degree such that
This proves that is not always linear in the number of vertices for bounded-degree . On the other hand, we show that is nearly linear in the number of vertices for typical bounded-degree acyclic digraphs , and obtain linear or nearly linear bounds for several natural families of bounded-degree acyclic digraphs.
For multiple colors, we prove a quasi-polynomial upper bound for all bounded-degree acyclic digraphs on vertices, where is the least such that every -edge-colored tournament on vertices contains a monochromatic copy of . For and , we exhibit an acyclic digraph with vertices and maximum degree such that , showing that these Ramsey numbers can grow faster than any polynomial in the number of vertices.
1 Introduction
The -color Ramsey number of a (simple undirected) graph , denoted , is the minimum such that every -edge-coloring of the complete graph contains a monochromatic copy of . Broadly speaking, the main question in graph Ramsey theory is to understand how depends on and . The most well-studied case is that of two colors, . For on vertices, it is known [10, 36] that grows exponentially in if and only if has edges.
However, it has long been observed that the Ramsey number of a sparse graph is much smaller than exponential in . In their foundational paper on the topic, Burr and Erdős [6] conjectured that this phenomenon is quite general and that any sparse graph has linear Ramsey number. Here, the appropriate notion of sparsity is degeneracy: is said to be -degenerate if every subgraph of has a vertex of degree at most , and the degeneracy of is the minimum such that is -degenerate. Burr and Erdős conjectured that for any -vertex graph with degeneracy . Here and throughout we use the standard asymptotic notation where the implicit constant is allowed to depend only on the subscripts of . Major progress towards this conjecture was made by Chvátal, Rödl, Szemerédi, and Trotter [8], who proved the Burr–Erdős conjecture under the stronger assumption that has bounded degree (rather than bounded degeneracy), that is, that for any -vertex graph with maximum degree . Finally, building on many prior developments (e.g. [19, 28]), the full Burr–Erdős conjecture was proved by Lee [30] in 2017.
There are many analogous questions and results for directed graphs (henceforth digraphs). We assume all digraphs are simple and oriented, so they do not contain self-loops, parallel edges or anti-parallel edges. For a digraph , define the -color oriented Ramsey number to be the minimum such that any -edge-colored tournament on vertices contains a monochromatic copy of . Note that if contains a directed cycle, then does not appear in any transitive tournament, so only exists for acyclic . Henceforth, we work exclusively with acyclic digraphs .
Unlike undirected Ramsey numbers, oriented Ramsey numbers are interesting even in the case of one color, , where is simply the minimum such that any tournament on vertices contains a copy of . Let denote the transitive tournament on vertices. The study of oriented Ramsey numbers was initiated by Stearns [35] in 1959 and Erdős and Moser [18] in 1964, who showed the upper and lower bounds, respectively, in
(1.1) |
The exponential constants in these bounds have not been improved, similar to the classical case of the diagonal undirected Ramsey number [10, 33, 34]. This may not be surprising, given that . Thus, improving the lower bound in (1.1) is at least as difficult as improving the lower bound on .
Somewhat more is known about the oriented Ramsey number when is sparse. When is the directed path on vertices, Chvátal [7] and Gyárfás and Lehel [22] determined that using the Gallai–Hasse–Roy–Vitaver theorem [20, 24, 32, 38]. In the case of one color, it was more generally conjectured by Sumner in 1971 that for any oriented tree on vertices, . Sumner’s conjecture has received a considerable amount of attention over the years (see e.g. [15, 16, 23, 25, 26, 29, 37]); it was proven for sufficiently large by Kühn, Mycroft, and Osthus [29], and Dross and Havet [15] showed that for all . In more colors, it was shown by Bucić, Letzter, and Sudakov [4] that for any oriented tree and all . In the same paper, they asked a natural directed analogue of the classical Burr–Erdős problem.
Problem 1.1 ([4]).
Is it true that for every acyclic digraph with vertices and maximum degree ?
Here we write for the out-neighborhood and for the in-neighborhood of a vertex , and say that a digraph has maximum degree if .
Yuster [40] recently initiated the study of the special case of Problem 1.1 when is the -th power of a directed path , which is the digraph on vertex set whose edges are the ordered pairs satisfying . This case was recently settled by Draganić et al. [14], who showed that . Letting the bandwidth of an acyclic digraph on vertices be the minimum such that contains a copy of , this aforementioned result implies that if has vertices and bandwidth .
In this paper, we answer Problem 1.1 in the negative, and show that in fact can grow faster than any fixed power of , as long as the maximum degree is a sufficiently large constant.
Theorem 1.2.
For any and sufficiently large in terms of , there exists an acyclic digraph on vertices and maximum degree at most for which
The power seems to be best possible with our method, but the power of can be improved. Although the answer to Problem 1.1 is negative, we prove an almost linear upper bound on for almost all , in the following sense. Define to be the orientation of the random regular graph on vertex set with all edges pointing to the right.
Theorem 1.3.
If is fixed and , then w.h.p.111As usual, we say that an event happens with high probability (w.h.p.) if as . (as )
It is not difficult to extend Theorem 1.3 to show w.h.p. if is the forward acyclic orientation of a uniformly random graph with any fixed degree sequence , and therefore also for the forward acyclic orientation of a uniformly random bounded-degree graph. We also prove a similar bound when is the forward acyclic orientation of an Erdős–Rényi random graph of constant average degree .
Although we are able to show that is w.h.p. almost linear for a random bounded-degree acyclic digraph, we have not determined the worst-case behavior of this Ramsey number. For general acyclic digraphs on vertices with maximum degree , the best upper bound we are able to prove is (see Theorem 1.7 below). Nonetheless, we are able to prove stronger (and in some cases linear) upper bounds on in case lies in certain natural families. We now give two examples of such families.
Let the height of be the number of vertices on the longest directed path in . Equivalently, the height can be seen as a directed analogue of the chromatic number: has height at most if and only if can be partitioned into independent sets such that every edge between and is oriented from to , for every . For acyclic digraphs of bounded height and bounded degree, we are able to prove the following linear bound on .
Theorem 1.4.
If is an acyclic digraph on vertices with maximum degree and height , then
In particular, .
Note that this theorem also implies the aforementioned upper bound on for any bounded-degree acyclic digraph , since the height of an acyclic digraph is at most its vertex count.
Next, we say that an acyclic digraph of height is graded if its vertex set can be partitioned into independent sets such that every edge in is directed from some to . Equivalently, is graded if for every pair of vertices , all directed paths from to have the same length (the equivalence of the definitions follows e.g. from [31, Proposition 4.4]). A natural example of a graded digraph is a grid (in any dimension) with all edges oriented towards the first orthant. In general, a graded digraph can be obtained from any graded lattice (in the sense of partially ordered sets) by orienting every edge of the Hasse diagram of from to . For graded digraphs of bounded degree, we are able to prove a polynomial bound on .
Theorem 1.5.
If is a graded digraph on vertices with maximum degree and height , then
In particular, since , we have that .
Our methods are motivated by those used by Conlox, Fox, Lee, and Sudakov [11] to prove bounds on ordered Ramsey numbers, and the two problems are especially closely related when the number of colors is at least . Using this connection, we are able to give a super-polynomial lower bound for when .
Theorem 1.6.
For any , there exists an acyclic digraph on vertices with maximum degree for which
for all .
Thus, for acyclic digraphs of bounded degree, can grow super-polynomially if . In the other direction, for any number of colors we have the following quasi-polynomial upper bound.
Theorem 1.7.
If and is any acyclic digraph with vertices and maximum degree , then
For one color, there is still a gap between the polynomial lower bound and the super-polynomial upper bound.
We remark that there is another well-studied analogue of ordinary Ramsey numbers in the directed setting, namely the directed Ramsey number , introduced by Bermond [2]. This is defined as the minimum such that a monochromatic copy of exists in every -coloring of the edges of , the digraph with edges in both directions between all pairs of vertices. There are similarities and differences between the two theories (see e.g. [4]), and several of our results on oriented Ramsey numbers can be extended to the setting of directed Ramsey numbers. We expand on these connections in our concluding remarks, Section 6.
To conclude this introduction, we remark on an interesting phenomenon brought to light by our results. An important “meta-question” driving many advances in Ramsey theory asks which graph parameters roughly determine a Ramsey number. In the Ramsey theory of undirected graphs, this question has been more or less resolved, at least in a qualitative sense. Namely, for an undirected graph , the degeneracy and the number of vertices are the main parameters that determine the growth of (we focus on the two-color case to be concrete). This is easiest to see in the lower bounds: if has vertices, then certainly . Additionally, if has degeneracy , then it is a simple exercise to show that a random -edge-coloring on vertices does not contain a monochromatic copy of w.h.p., implying that . Putting these two facts together, we find that
Conlon, Fox, and Sudakov [12, Conjecture 2.16] conjectured that this bound is tight up to the implicit constant, namely that
for any graph with vertices and degeneracy . Moreover, this conjecture is known to be true up to a multiplicative factor of [19, Theorem 3.1]. Because of these results, one can say that the degeneracy and the vertex count of roughly determine its Ramsey number.
For acyclic digraphs, we do not know what parameters determine the growth order of (indeed, we don’t even know if is polynomial or super-polynomial in when has bounded degree). Nonetheless, our results indicate that one parameter, which we call “multiscale complexity,” is relevant. Namely, suppose we order the vertices of as so that every edge is oriented to the right, that is is an edge only if (such an ordering is often called a topological sorting of ). Under this ordering we may assign every edge a length . Here, if , we write to signify that there is a directed edge from to , and similarly for an edge in the other direction.
In graphs of bounded bandwidth, every edge is short and has length . At the other extreme, if has bounded height, then most edges of are long and have length . The same is true in the random model , where most edges have length . Some other acyclic digraphs have intermediate edge-length statistics, such as the directed grid whose vertex set is and all edges are oriented towards the lexicographically larger ordered pair. For every acyclic ordering of such a grid, there are many edges in most dyadic length scales between and .
Loosely, let us say that has high multiscale complexity if, for most dyadic intervals with , there are many edges in whose length is in . Conversely, if most edge lengths of are concentrated in a small number of dyadic intervals, then we loosely say that has low multiscale complexity. At a high level, all of our upper bound results prove upper bounds on in terms of , and the multiscale complexity of ; if the multiscale complexity is low, then these bounds are stronger, and one can prove linear, nearly linear, or polynomial bounds on (depending on the precise context). Conversely, our lower bound construction for Theorem 1.2 is a family of digraphs which we call interval meshes. We delay the precise definition to Section 2, but interval meshes are in some sense designed to maximize multiscale complexity among all graphs of maximum degree at most .
We stress that we have not fully solved the problem of which natural parameters determine the growth order of , and we think this problem deserves further research. Nonetheless, our results do make it clear that some notion of multiscale complexity is one of these parameters, and we believe this notion is fundamental. As such, we state and prove many of our technical upper bounds in greater generality than is needed to deduce our main theorems, in order to demonstrate how notions of multiscale complexity arise naturally from our techniques.
The rest of the paper is laid out as follows. Section 2 carries out the construction of interval meshes to prove the lower bound Theorem 1.2. Section 3 uses the greedy embedding technique to prove the main technical lemmas needed for the upper bounds Theorems 1.3, 1.4, and 1.5. Section 4 completes the proofs of these results, as well as a more general upper bound in terms of “multiscale complexity.” Using the connection to ordered Ramsey numbers, Section 5 proves Theorems 1.6 and 1.7. Finally, in our concluding remarks, Section 6, we collect a few of the tantalizing open problems that remain in this area.
Notation and terminology.
By an embedding , we mean an injective function such that edges of are mapped to edges of , with edge orientations preserved. We say that a digraph is -free if there is no embedding . All logarithms are to base . For the sake of clarity of presentation, we sometimes omit floor and ceiling signs when they are not crucial.
2 Lower bounds
In this section, we prove the lower bound Theorem 1.2, which states that for any there exists a family of acyclic digraphs with maximum degree for which and . The lower bound consists of three ingredients: constructing a bounded degree acyclic digraph that is hard to embed, constructing a Ramsey tournament that is hard to embed into, and proving that there is no embedding .
The next three subsections separately provide these three ingredients. Section 2.1 defines a class of digraphs with edges “at all scales,” which we call interval meshes, and proves the existence of bounded-degree with this property. Section 2.2 defines the Ramsey tournament as a lexicographic power of a tournament with no large transitive subtournament, and shows that embeddings correspond to certain highly constrained walks on which we call -walks. Section 2.3 completes the proof by showing that long -walks do not exist, and therefore large interval meshes cannot be embedded into small powers .
2.1 Interval meshes
Our proof of Theorem 1.2 is motivated by the lower bound construction for ordered Ramsey numbers proved by Conlon, Fox, Lee, and Sudakov [11]. They prove a lower bound on the ordered Ramsey number of a random matching on ; see Theorem 5.1 and the surrounding discussion for details. The main property they need of the random matching is that most pairs of long intervals have an edge between them. We need the following stronger property of the same form for our acyclic digraph .
Definition 2.1.
If is a nondecreasing function, we define an -interval mesh to be an acyclic digraph whose vertex set is an interval such that any edge has and for all pairs of intervals with and
(2.1) |
there exists an edge in between and . When the function is clear from context, we simply call an interval mesh.
The notion of an interval mesh is one way of formalizing the notion from the introduction of “high multiscale complexity”, since interval meshes have many edges in every length scale. We construct interval meshes of bounded degree using a greedy algorithm.
Lemma 2.2.
For any nondecreasing function with , there exists an -interval mesh on vertex set with maximum degree at most .
Proof.
Starting from an empty digraph on , we construct by using a greedy algorithm to add edges. All edges introduced point to the right, so the resulting digraph must be acyclic. Define to be the dyadic interval with length .
Let range through the nonnegative integers. On subroutine , we iterate through all pairs satisfying
(2.2) |
and add an edge between the (currently) lowest degree vertex of and the (currently) lowest degree vertex of , if an edge does not exist between and already. Writing for the sum of the degrees of the vertices in a set , we see that for any given , subroutine increases by a total of at most
Let be the digraph produced from this infinite process. Explicitly, is the edge union of all the graphs , where is the graph produced after subroutine . It has the property that every pair of dyadic intervals , satisfying (2.2) has an edge between them.
We first check that has maximum degree at most . Since is a union of dyadic intervals of length , we have that after subroutine ,
In particular, the average degree in is less than at the end of subroutine . However, subroutine only adds edges incident to vertices of which have at most average degree, so no new vertex of degree at least can appear on subroutine . Thus, no vertex of degree at least ever appears, as desired.
Next, we check that is an -interval mesh. Suppose two intervals , with satisfy (2.1) and let be the largest positive integer such that and both contain dyadic intervals of length . Note that in any interval of integers, a longest dyadic subinterval has length , so in particular Next, pick maximal and minimal such that and .
The acyclic digraphs we use are induced subgraphs on intervals of the infinite interval mesh constructed in Lemma 2.2, for an appropriate choice of .
2.2 Walks in tournaments
Next, we construct the large tournament which is difficult to embed into. Again, the construction is motivated by the lower bound argument of Conlon, Fox, Lee, and Sudakov [11] for ordered Ramsey numbers, although its analysis requires new techniques.
Recall that the lexicographic product of two digraphs and is the digraph on vertex set with an edge iff in or and in . Henceforth, write for the -fold lexicographic product of with itself. Note that lexicographic powers of tournaments are tournaments. By (1.1), there exists for any a tournament on vertices with no transitive subtournament of size . We take and show that an interval mesh is difficult to embed into .
To this end, let be an interval mesh. We relate embeddings to certain constrained walks on the tournament .
Definition 2.3.
For a tournament , a nondecreasing function , and , define an -walk to be a sequence of ordered pairs possibly infinite) where for each , , , , and for any pair for which in , we have
where if is an interval of integers and the empty sum equals . We define the length of an -walk to be .
Let be the length of the longest -walk if such a maximum exists, and otherwise. The next lemma reduces Theorem 1.2 to showing asymptotic upper bounds on .
Lemma 2.4.
Suppose there exists , a nondecreasing , and a tournament on vertices for which . If is an -interval mesh on vertices, then
Proof.
For each , let be the minimum positive integer for which there exists an -interval mesh with vertex set and an embedding .
Let be the projection onto the first coordinate. Consider the sequence . Consecutive terms of this sequence may repeat, so we block the sequence into consecutive runs of identical vertices. Suppose there are total runs and run consists of repetitions of vertex . We claim that is an -walk, where . It is easy to see that for all , and that since we already blocked consecutive identical vertices together. It remains to check the key condition, that if and in , we must have
Suppose this is not true, so there exists some with and . By the definition of , we have that , . By the definition of the lexicographic power, if then for all . Thus, all edges between and are oriented from to . For to be a homomorphism, this means that has no edge oriented from to . On the other hand, , , and the distance between the two intervals is , so since is an -interval mesh there is such an edge. This is a contradiction, and we see that is an -walk of length .
Fix any . We are given that for sufficiently large , so we get using the fact that since we have found an -walk of length . Since , this means that there is some subinterval of length at least for which is constant on , i.e. the image lies in a copy of . Putting this together, we have shown that for large enough , the existence of an embedding implies the existence of an embedding for some -interval mesh on vertices. In other words,
for all sufficiently large. Applying this recursively, we obtain that . By the definition of , any -interval mesh on vertices has no embedding into , and so
which proves the lemma. ∎
To finish the proof of Theorem 1.2, it remains to bound the asymptotic growth rate of , which we do in the next section.
2.3 Completing the proof
The next lemma is a simple observation that is helpful for analyzing the structure of -walks.
Lemma 2.5.
If is a tournament without a transitive -subtournament, then any sequence of vertices either contains a back-edge with or two consecutive elements .
Proof.
If the vertices are all distinct, then since has no transitive -subtournament there must exist a back-edge. Suppose and but . Then either or is a back-edge. ∎
We now prove a recursive upper bound on . Given an implicit parameter , tournament , and nondecreasing function , we say that a positive integer is short if and for all .
Lemma 2.6.
Suppose , , is a tournament without a transitive -subtournament, and is a nondecreasing function satisfying . If is short, then every is short.
Proof.
Suppose , is an -walk, and let be the sequence of all indices where .
Our first goal is to show that . If not, by Lemma 2.5 there is either a back-edge with or some for which . We show that neither of these situations is possible.
In the first case, there is an edge with . By the definition of an -walk,
(2.3) |
On the other hand,
since for each , the subsequence is an -walk with length at most . But , so this contradicts (2.3) and the back-edge cannot exist.
Next, suppose for some that . The subsequence is an -walk where is the maximum value of in this subsequence. Pick some for which . Either or is a back-edge, and without loss of generality assume it is the former. Applying the definition of the -walk on the two indices and ,
On the other hand, this sum is bounded above by . We know because is an -walk and is short, so we have another contradiction. Thus .
We obtain
(2.4) |
for all and any -walk . Here we let and for convenience.
Inequality (2.4) implies that for all . It also implies that for all . We have verified both conditions for to be short for every , as desired. ∎
It remains to pick a function for which Lemma 2.6 bootstraps successfully. Define
(2.5) |
where the values of are chosen just to make nondecreasing. Recall that all logarithms are to base .
Lemma 2.7.
If , is defined by (2.5), and is a tournament without a transitive -subtournament, then for all sufficiently large.
Proof.
We apply Lemma 2.6 inductively to show that is short for all sufficiently large. This implies the desired result.
For the base case, we claim that is short for all . Indeed, suppose and is an -walk. By Lemma 2.5, if then there is either a back-edge with or two consecutive elements . The latter contradicts the definition of an -walk, so assume the former holds. But , so this contradicts the definition of an -walk again. We have shown that , so for all . This proves is short for every .
Next, we check that for all . Indeed, is increasing for , and since . We get
proving that the conditions of Lemma 2.6 are satisfied. By induction, Lemma 2.6 then implies that is short for for all . All sufficiently large integers lie in some such interval, so for all sufficiently large , as desired. ∎
Putting everything together, we have a proof of the general lower bound.
Proof of Theorem 1.2..
We may assume that is sufficiently large as we always have , which proves the theorem for small by picking the implicit constant factor in the exponent appropriately. Let ; we may assume . Define by (2.5). We have
Lemma 2.2 implies that there exists an -interval mesh on with maximum degree at most . For any , let be the induced subgraph of on the interval , so that has vertices and is also an -interval mesh of maximum degree at most .
We remark that the polylogarithmic factor in can be easily improved. Indeed, the growth rate of is chosen so that
and we may take for any fixed instead, leading to a slightly smaller .
3 Greedy embedding
In this section we prove the main lemmas needed for all of the upper bounds in this paper. We use the greedy embedding technique, motivated by similar arguments for ordered graphs from [11].
Framework. Let be an acyclic digraph on vertices . We would like to find an embedding into an ambient tournament . In addition we specify sets and aim to satisfy for all . Embedding then proceeds in rounds, where round determines the image . After round , we keep track of the shrinking sets of “valid candidates” for each vertex, where initially for all . On round , is a carefully chosen vertex in , and . The other candidate sets are updated as follows:
The process fails if there is an empty , as there would be no valid choice for . Otherwise, it succeeds if is chosen successfully for each . Note that after round , is an embedded copy of in , and these vertices have been removed from the other candidate sets, so the update rule guarantees that remains a singleton when . If the greedy embedding process succeeds, it exhibits the existence of a copy of in . See Figure 3.1 for a schematic illustration of the greedy embedding process.
For the upper bounds described in the introduction (Theorems 1.3, 1.4, and 1.5), we apply greedy embedding in two separate stages, which we call the outer stage and the inner stage. Roughly speaking, in the outer stage, we run the greedy embedding procedure many times to show that if is -free, then contains a large “approximate blowup of .” In the inner stage, we use greedy embedding one final time within this “approximate blowup” to guarantee the existence of a copy of in . In either case, we can conclude that contains a copy of —either it is found directly by the greedy embedding strategy, or else the failure of the greedy embedding yields the “approximate blowup” of , in which a copy of can be found directly.
This section is split into three subsections. The first covers the basic results that follow from the greedy embedding framework described above, namely how a failure to greedily embed in a tournament implies that contains a certain nice structure, namely a pair of large vertex sets such that most edges between them have the same orientation. The tools built in this first subsection are then used as basic building blocks and iterated in the subsequent subsections. In the second subsection, we use them to build the outer stage of the embedding. In the third subsection, we explain how to use this outer stage construction of an “approximate blowup of ” to finally embed itself.
In some of these greedy embedding arguments, we are concerned with partitioning an acyclic digraph into a number of parts and embedding the parts one at a time, so the following definition will be useful.
Definition 3.1.
If is an acyclic digraph, we say that a collection of vertex subsets of is a directed partition of if , and any edge of between two distinct parts with is oriented from to .
In particular, the height of is exactly the least for which there exists a directed partition of into independent sets.
3.1 The basic greedy embedding building blocks
Recall that an undirected graph is said to be -degenerate if there exists an ordering of the vertices of such that each , and such an ordering is called a -degenerate ordering of . A -degenerate ordering is a natural order for greedily embedding an undirected graph , since each candidate set in the greedy embedding strategy only shrinks in size by more than one at most times. We say that a digraph is -degenerate if its underlying undirected graph is -degenerate. Note that if has maximum degree then it is -degenerate, but -degenerate digraphs can have arbitrarily large maximum degree.
Define a -dense pair in a tournament to be a pair of vertex subsets such that at least of the edges between and point from to . The size of the pair is defined to be . We do not require and to be disjoint, although the assumption of -density implies that cannot be too large if is close to .
The first lemma in this subsection uses greedy embedding to show that if doesn’t contain a copy of a given -degenerate , then contains some large dense pair. The undirected analogue of this lemma is well-known, and goes back at least to work of Erdős and Hajnal [17, Lemma 1.5].
Lemma 3.2.
Let be a -degenerate digraph with vertices and maximum degree , and let . If is an -free tournament on vertices, then contains a -dense pair with size at least .
Proof.
We use the greedy embedding framework described above. Let us label the vertices of according to the -degenerate ordering as We initialize for all . We now attempt to embed the vertices of one at a time in , in the -degenerate ordering. For , let denote the set of vertices with such that and are connected by an edge (in some direction). We inductively pick the values of and maintain vertex sets with the following properties.
-
1.
For every , we have .
-
2.
For every , we have .
-
3.
For , if then for every , and if then for every .
From these properties, we see that if the process continues through step , then we will have embedded a copy of in , contradicting our assumption that is -free. Moreover, all three properties are vacuously true for . Now, suppose we have maintained this process up through step . Suppose there exists such that for every with (resp. ), we have (resp. ). Then we may define , , and update the remaining sets as
for all . Properties 1 and 3 above continue to hold automatically after round , and all that remains to check is Property 2. For those for which is not adjacent to , and at most one vertex is removed from to obtain . Therefore,
On the other hand, if is adjacent to , then , and therefore
Thus, all three properties are maintained if such a exists.
Since we assumed that was -free, this process cannot continue until step , and therefore must fail at some step . Let . Since the process fails at this step, for every , we can assign some such that either and , or and . Since has at most neighbors in total, at least choices of are assigned the same by the pigeonhole principle. Fix such a “popular” .
Suppose first that . Let , and let be the set of all which have fewer than out-neighbors in . Then is a -dense pair. Similarly, if , then we would similarly find that is a -dense pair.
It remains to verify the lower bound on the sizes of and . Recall that the greedy embedding process succeeded up through step , meaning that
and similarly for , where we use the -degeneracy assumption to conclude that , and our assumption that . Since and , this completes the proof. ∎
The second lemma proves a much stronger bound when is -degenerate and weakly connected, i.e. some orientation of a tree (recall that a digraph is called weakly connected if its underlying undirected graph is connected). Note that a -dense pair in a tournament is just a pair of sets with all edges directed from to .
Lemma 3.3.
Let be a weakly connected -degenerate digraph with maximum degree on vertices , and let be an arbitrary tournament. If there exist sets , each of size such that if there is no embedding satisfying , then contains a -dense pair with size at least .
Proof.
We begin by picking subsets , each of size , such that are pairwise disjoint. We can do this greedily, by first picking an arbitrary subset of size , then picking an arbitrary subset of size , then picking , and so on. At the th step, we have deleted at most elements from , so at least elements remain, from which we pick arbitrarily.
The remainder of the proof is very similar to that of the previous lemma, though it will be more convenient to work with a slightly different setup than before. Let the vertices of be ordered so that each has at most one neighbor with (this is the reverse of the usual degenerate ordering). For each , let denote the set of vertices with that are connected to by a path of vertices whose indices are monotonically increasing, including itself. In other words, we set , let denote the set of vertices adjacent to (in either orientation) with , and set
We define to be the set of all for which there exists an embedding mapping each into for , and mapping to . We know , since otherwise there is an embedding of into the sets , contradicting our assumption. Let be the minimum index such that ; note that since has size . Let . Our maximum degree assumption implies , while if , then , contradicting our assumption that . Thus, .
By definition, is precisely the set of adjacent in the appropriate orientation to at least one vertex in each of . Let denote the choices of which do not have any edge in the appropriate orientation to , respectively. We get . Since and , we see that there exists some for which . Moreover, since was taken to be minimal, we have that as well. This yields a pair of sets where all edges between them are oriented the same way, and both sets have size at least . This is the desired -dense pair. ∎
We remark that we expect the dependence in Lemma 3.3 to be unnecessary, and proving this would improve our results for random digraphs; for details, see Conjecture 6.5 and the surrounding discussion.
We do not use Lemma 3.3 directly, but only the following simple corollary. It allows us to find a large -dense pair in whenever does not contain a copy of some fixed oriented forest with small maximum degree and small weakly connected components.
Lemma 3.4.
Let be a -degenerate digraph with maximum degree and vertices , and suppose that every weakly connected component of has at most vertices. Let be an arbitrary tournament. For any collection of sets , each of size , either there is an embedding satisfying , or contains a -dense pair with size at least .
Proof.
Let the weakly connected components of be . We prove the result by induction on . The base case, , follows immediately from Lemma 3.3, since and . For the inductive step, let be the induced subgraph of consisting of the weakly connected components . By the inductive hypothesis, either contains a -dense pair of size at least , in which case we are done, or else there is an embedding of into satisfying that for all . Let be the tournament obtained by deleting the image of from , and similarly let be obtained by deleting the image of from . Since we have deleted at most vertices, each has size at least . Therefore, by Lemma 3.3, either contains a -dense pair of size at least , in which case so does , or else there is an embedding of into mapping each into . Since we deleted the image of from to define , such an embedding yields an embedding of into , completing the induction. ∎
Our final basic greedy embedding lemma is the following, which shows that if we assume appropriate density conditions on the sets in which we are trying to greedily embed , then we are guaranteed not to fail in the embedding. In Subsection 3.3, we use this lemma as a basic building block. Again, the undirected analogue of this lemma is well-known, e.g. [21, Lemma 2].
Lemma 3.5.
Let be an acyclic digraph with maximum degree on vertices , and let be a tournament containing subsets , each of size at least . If for every edge in , is a -dense pair, then contains a copy of .
Note that the sets are allowed to overlap222However, if , then the assumption that is -dense implies that cannot be too large. or even be identical. This is crucial; for instance, if has height then our choice of ’s will take only distinct values.
Proof.
Every acyclic digraph has a vertex ordering where all edges point forward, so we may reorder the vertices to assume all edges satisfy . We now run the greedy embedding process for into with the given , using essentially the same framework as we used in Lemma 3.2. The key difference is that before, the greedy embedding process could fail and terminate prematurely, whereas the additional assumptions here guarantee that greedy embedding runs to completion.
We begin by refining the sets . Namely, for every edge in , let be the set of vertices with . Since the pair is -dense, there are at most edges directed from to in total, so
and thus . Each has at most out-neighbors , which implies that . Therefore, if , then , and every vertex in has at most in-neighbors in for every such that .
Now, we exhibit an embedding by picking inductively for each , as follows. Having picked , we claim that there is at least one valid candidate for which is consistent with the previous choices. Indeed, there are at most choices of such that . For every such , we have that , which means that has at most in-neighbors in . Hence, at least vertices in are out-neighbors of for all such that . Moreover, at most vertices of have been picked as outputs of , and thus the number of possible candidates for is at least
by our assumption that . This shows that at every step we can pick vertex such that is a copy of in , and so contains a copy of the entirety of , as desired. ∎
We remark that one can prove a strengthening of Lemma 2.4, where the assumption is weakened to each pair being merely -dense. This can be done by replacing the greedy embedding argument by a random embedding technique, using the Lovász local lemma. For details, we refer the reader to [13, Lemma 4.5], where the analogous undirected result is proved using this technique.
3.2 The outer stage
In this subsection, we show how two of our basic building blocks, Lemmas 3.2 and 3.4, can be iterated to construct, in any -free tournament , an “approximate blowup” of . This will be a large collection of vertex sets which correspond to the vertices of , such that the edges between sets are mostly oriented in the correct direction. Before stating the result precisely, we need some definitions.
Let denote the set of all finite binary strings. Recall that a prefix code is a set with the property that no element of is a prefix of another element of . The depth of is defined as the maximum length of an element of . Let denote the lexicographic ordering on , namely the ordering in which if is a proper prefix of or if where is the first index for which .
Definition 3.6.
Given an acyclic digraph , a prefix labeling of is a surjective map for some prefix code , with the property that if is an edge of , then either or . The map naturally defines a graph structure on , where we say that two codewords are adjacent under if there exists some edge between the sets and . By the maximum degree of , we mean the maximum degree of this graph on . If is an independent set for every , then we call a prefix coloring. Less stringently, we call a forest prefix labeling if is a directed forest (or equivalently, a -degenerate digraph) for every . By the maximum component size of , we mean the maximum number of vertices of any weakly connected component in , over all . Thus, is a prefix coloring if and only if its maximum component size is .
Thus, we see that prefix colorings of correspond to colorings of the underlying undirected graph of , with the property that the palette of colors is a prefix code, and that the lexicographic order on is consistent with the edge directions in . Similarly, a forest prefix labeling is in particular a partition of the underlying undirected graph into sets which induce forests, which corresponds to the undirected problem of vertex arboricity. However, for both concepts, we will crucially use the additional structure given both by the edge directions of and by the structure of the prefix code .
For a binary string , let us denote by and the strings obtained by appending a or , respectively, to the end of . For a prefix code , we denote by the set of all elements which have as a prefix, and by the set of elements that have as a prefix. Suppose that is a prefix labeling of an acyclic digraph . For binary string , let us denote by the number of codewords which are adjacent under to some codeword . Similarly, is the number of codewords that are adjacent under to some codeword . Finally, we let
In particular, if is not a proper prefix of any element of . With this notation, we can now define the key parameters of that we later use to bound Ramsey numbers.
Definition 3.7.
Let be an acyclic digraph and some prefix labeling of . We define the dyadic complexity of to be the quantity
Additionally, the depth of , denoted , is defined as the depth of , i.e. the maximum length of an element of .
To understand these definitions, it is helpful to think of as the vertices of the infinite binary tree. In this setup, the elements of a prefix code correspond to the leaves of some subtree. A prefix labeling is then a partition of the vertices of into sets labeled by the leaves of this subtree. Two codewords (leaves) are adjacent under if there is an edge between the corresponding vertex subsets of . For a binary string , which should be thought of as a non-leaf vertex of the subtree, roughly records the “cost” of separating the descendants of : it measures how many pairs of codewords adjacent under there are between its descendants on the left and on the right. Because of the structure of the proof, this cost function is somewhat unnatural: it is the minimum of two quantities, each of which is the number of descendants on one side which are adjacent under to any number of descendants on the other side. Moreover, this cost function should be thought of as multiplicative, so that the ultimate cost of the whole labeling—namely the dyadic complexity —is the product of the costs of all ancestors of , maximized over all .
We remark that the dyadic complexity of a prefix coloring of is one possible formalization of the notion of multiscale complexity discussed in the introduction. Indeed, if every prefix coloring of has high dyadic complexity, then has many edges at “many different dyadic scales”. All of our upper bounds on oriented Ramsey numbers depend on the dyadic complexity of some prefix labeling of , making formal the intuition that the strength of our upper bound results depends on whether has high or low multiscale complexity.
In order to embed some acyclic digraph in a tournament , we first build a certain structure of vertex subsets of with high forward edge density between many of the pairs. Then, in the inner stage of the embedding process, we use such a structure to find a copy of . The structure we build depends on a parameter , as well as on a prefix labeling of . We now define this structure, and next prove a lemma showing how to find such a structure.
Definition 3.8.
Let be some parameter, let be an acyclic digraph, and let be some prefix labeling of , for some prefix code . Let be any tournament. A -skeleton is a collection of (not necessarily disjoint) vertex subsets of , indexed by the codewords in , with the property that if are elements of that are adjacent under , then is a -dense pair. We define the size of a -skeleton to be .
Our next lemma shows how to iterate Lemma 3.2 to construct a -skeleton in any sufficiently large -free tournament . Roughly speaking, since the structure of a -skeleton is based on the binary tree structure of , we may construct such a skeleton by performing a depth-first search, and applying Lemma 3.2 every time we need to split an existing node into two daughter nodes in this binary tree.
Lemma 3.9.
Let be some parameter, let be a -degenerate acyclic digraph with maximum degree , and let be some prefix labeling of , for some prefix code . Suppose that is an -free tournament on vertices, with
Then contains a -skeleton of size at least .
Proof.
For every binary string which is a prefix of some codeword in , let denote the subgraph of induced by the vertices for which is a prefix of . We will construct, for every such , a vertex subset , with
We initialize , where denotes the empty string, and observe that this property holds vacuously for since has no proper prefixes. In order to construct these vertex sets for other , we proceed via a depth-first search along the binary tree, as follows. Recall that for any binary string , the numbers and are the number of codewords in beginning with and , respectively, such that some vertex labeled by that codeword has an edge to and , respectively. If , then we define , and observe that our desired inequality holds, since in this case. In particular, if is a codeword, then we stop the inductive process, since we only wished to define such vertex subsets for strings that are the prefix of some codeword. Now, suppose that , and assume without loss of generality333If , then we swap the roles of and and the roles of in- and out-neighbors in this construction. that . We will first show how to define satisfying the desired inequality on its cardinality. We will then proceed to recursively define vertex subsets for every binary string prefixed by . Note that we have not yet defined the set : we are proceeding in a depth-first fashion, so we will not define until we have defined for every prefixed by . This will eventually happen, since we already described above how to define if is a codeword of ; therefore, we eventually reach the bottom of our depth-first search (namely a codeword ), at which point we stop going down the tree, and begin to retrace our steps and traverse back up the tree.
Recall that we assumed that , and let Since
(3.1) |
and since contains no copy of , we may apply Lemma 3.2 with parameter . Then this lemma says that contains a -dense pair , where . Let denote the set of vertices in whose in-degree to is at most ; since there are at most edges directed from to , we see that . Therefore,
(3.2) |
since the proper prefixes of are just , in addition to all the proper prefixes of itself.
As discussed above, we can now recursively define for all prefixed by . It thus only remains to define , under the assumption that we have defined for all prefixed by .
To do so, let denote the set of codewords prefixed by with the property that some vertex in is adjacent to some vertex in , noting that there are exactly such codewords by the definition of . Let denote the subset of consisting of vertices in whose out-degree to is at least . Since every vertex in has at most in-neighbors in , we see that the total number of edges directed from to is at most , and therefore . Because of this, we have that
Therefore, we can define , and we have that . By the same computation as in equation (3.2), we see that this definition of satisfies our desired lower bound on the size of .
We claim that in this construction, if are codewords that are adjacent under , then the pair is -dense. To see this, let be the longest common prefix of and . In the construction at level , we first proceeded to either or in the depth-first search, depending on the relative sizes of and . In the former case, we ensured that every vertex in had out-degree at most to , while in the latter case, we ensured that every vertex in had in-degree at most to . In either case, we see that is -dense, since and . Additionally, by the same computation as in equation (3.1), we see that for every . This verifies all the properties of a -skeleton, and concludes the proof. ∎
Our next two lemmas are very similar to Lemmas 3.2 and 3.9. Namely, the first shows us how to find a -dense pair in an -free tournament , and the second then iterates the first to form a skeleton of -dense pairs. The main difference between these and the previous results are that for these lemmas, we need strengthened assumptions on (namely that it has a directed partition into forests). Moreover, the first step in the proof is an application of Lemma 3.9, and we find the -dense pair by failing to greedily embed in the skeleton given by Lemma 3.9.
Lemma 3.10.
Let , and suppose that is a -degenerate acyclic digraph on vertices with maximum degree , and suppose that is some forest prefix labeling with maximum degree and maximum component size . Let be an -free tournament on vertices. Then contains a -dense pair of size at least .
Proof.
We first apply Lemma 3.9 to the prefix labeling and with parameter . This lemma outputs a -skeleton in of size at least
For every pair of codewords that are adjacent under , let denote the set of vertices in whose in-degree to is at least . Since at most of the edges between and are directed from to , we see that . Therefore, if we define , then we see that , since there are at most choices for with adjacent under . Additionally, every vertex in has in-degree at most from any with such that are adjacent under . We now attempt to greedily embed in these sets .
Let the codewords of be , sorted so that if and only if . Let , so that is an oriented forest, each of whose weakly connected components has at most vertices. Additionally, forms a directed partition of , which we recall means that every edge of is oriented from to where . For every vertex , we initialize a set of candidates . Inductively, having defined for each and each , we attempt to pick an embedding , such that for every , . If such a exists, then for each and each , we let
Note that by the structure of the sets , we only change as follows. First, in at most steps, we embed an in-neighbor of , and we decrease by at most . Additionally, we remove at most additional vertices from , corresponding to vertices that were picked as images of . In total, we remove at most vertices. We thus see that for all .
If we are able to run this process for all , then we have embedded a copy of in , so we may assume that the process fails at some step . This means that there is no embedding such that every vertex is mapped to . Therefore, by applying Lemma 3.4 to , we conclude that contains a -dense pair of size at least
Our next result shows how we can iterate the previous lemma to construct many -dense pairs. The iteration is nearly identical to the one in Lemma 3.9, where we iterated the construction of one dense pair to the construction of a -skeleton. This lemma takes as input two (not necessarily distinct) prefix labelings on , one of which is a forest labeling. The reason to have two separate labelings is that it may be useful to use the failure of embedding of according to one labeling to construct a good embedding structure for the other labeling.
Lemma 3.11.
Let , and suppose that is a -degenerate acyclic digraph on vertices with maximum degree . Let and be two prefix labelings, such that is a forest prefix labeling of maximum degree and maximum component size . Let
If is an -free tournament on vertices, then contains a -skeleton of size at least .
Proof.
As in the proof of Lemma 3.9, we construct our skeleton by assigning a set for every that is a prefix of some codeword in , with the property that are both subsets of . We guarantee inductively that
(3.3) |
where is the length of . To begin the induction, we set , which satisfies our size hypothesis since . Inductively, suppose we’ve defined . If , we stop. If not, we apply Lemma 3.10 to the induced subtournament on , which we may do since
This allows us to find a -dense pair of size at least . We then set and , which we see satisfy (3.3) inductively. We continue in this way until we define for every . To conclude, suppose that are adjacent under , and let be their longest common prefix. Then and , and we know that every edge between and is oriented from to , which implies that is -dense, as claimed. ∎
3.3 The inner stage
In this subsection, we will see how to use the various structures built in the previous subsection to successfully embed a copy of in any sufficiently large tournament . The basic idea is that the -skeletons constructed in Lemmas 3.9 and 3.11 are precisely the structures we need in order to apply Lemma 3.5 and find a copy of .
Recall that a prefix coloring is a prefix labeling where the preimage of every codeword is an independent set. Our first result here shows a general upper bound on in terms of the depth and complexity of a prefix coloring of .
Theorem 3.12.
Let be an acyclic digraph on vertices with maximum degree . Then for any prefix coloring , we have , where
Proof.
Let be a tournament on vertices, and suppose for contradiction that is -free. By Lemma 3.9 applied with and , we can find in a -skeleton of size at least
Now, for any vertex , let . Then since is an independent set for any , we see that if is an edge of , then is a -dense pair. Therefore, by Lemma 3.5, we conclude that contains a copy of . ∎
The second result of this subsection uses the -skeletons we constructed in case we are given a forest prefix labeling. Using this skeleton, we are able to prove the following result, which takes as input a forest prefix labeling and any arbitrary prefix labeling (which may be the same as the forest prefix labeling). The output is a bound on the Ramsey number, in terms of the depth and complexity of the first labeling, and in terms of the maximum Ramsey number of the parts in the partition induced by the second prefix labeling.
Theorem 3.13.
Let , and let be a -degenerate acyclic digraph on vertices with maximum degree . Suppose that is a forest prefix labeling of and that is any prefix labeling. Let and be the maximum degree and maximum component size of , respectively. Then , where
Proof.
Let be a tournament on vertices, and suppose that is -free. We apply Lemma 3.11 to , which allows us to find a -skeleton where
for all . In other words, this is a collection of disjoint sets such that if precedes in the lexicographic ordering, then every edge is oriented from to . By the definition of the Ramsey number, we see that the induced subtournament must contain a copy of for all . We pick such a copy arbitrarily for each , and observe that their union forms a copy of in . ∎
4 Upper bounds on oriented Ramsey numbers
4.1 Upper bounds in terms of height
Theorem 3.12, which bounds in terms of the dyadic complexity and depth of a prefix coloring, allows us to prove bounds on in terms of other, more natural, parameters. For instance, the next lemma relates the height of an acyclic digraph to its dyadic complexity and depth.
Lemma 4.1.
If is an acyclic digraph of height , then there exists a prefix coloring with and .
Proof.
Recall that if has height , then has a directed partition into independent sets. This partition naturally yields a prefix coloring using the prefix code consisting of all binary strings of length . Namely, we can label each vertex in by the base- representation of , and this yields a prefix coloring with depth . To estimate the dyadic complexity of this prefix coloring, note that for any binary string of length , we have that , since there are at most codewords in prefixed by . Therefore,
Then Theorem 1.4 follows as a simple corollary.
Proof of Theorem 1.4.
Similarly, given a graded digraph, one can find a prefix coloring with small depth and dyadic complexity.
Lemma 4.2.
If is a graded acyclic digraph of height , then there exists a prefix coloring with and .
Proof.
The proof is identical to that of Lemma 4.1, except that for every binary string , since the only codeword prefixed by that can have edges to a codeword prefixed by is the codeword . ∎
As before, Theorem 1.5 follows as a corollary.
Proof of Theorem 1.5.
Recall from the introduction that the bandwidth of an -vertex acyclic digraph is the least so that contains . Using the same argument, one can obtain a bound of for any -vertex acyclic digraph of bandwidth at most , using the fact that the same binary-representation prefix coloring have dyadic complexity at most . However, since the Ramsey number of bounded-bandwidth acyclic digraphs is known to be linear [14], we omit the proof of this weaker result.
4.2 Upper bounds for random digraphs
In this section, we prove Theorem 1.3, showing that if is bounded, then w.h.p. is nearly linear when . Additionally, we prove a nearly-linear upper bound when and .
We will need the following result of Dross and Havet [15] mentioned in the introduction. They only state this result for orientations of trees, but one can easily extend it to orientations of forests by adding edges to join distinct connected components.
Theorem 4.3.
Let be a -degenerate digraph (i.e. an orientation of an undirected forest) on vertices. If is any tournament on at least vertices, then contains a copy of .
We recall that the vertex arboricity of an undirected graph is the minimum number of subsets needed to cover the vertices of the graph, such that each subset induces a forest (see e.g. [5] for more on this undirected graph parameter). We first define a natural directed analogue of this quantity, though for technical reasons we also keep track of the maximum component size in such a partition.
Definition 4.4.
The directed vertex arboricity of an acyclic digraph is the minimum size of a directed partition of into sets where is -degenerate for all .
Additionally, if has a directed partition into sets such that is -degenerate for all , and every weakly connected component in has at most vertices, then we say that has an -forest partition.
Both of our upper bound results for random digraphs follow from the following theorem, which yields an upper bound on in terms of the degeneracy, maximum degree, and forest partition size of .
Theorem 4.5.
Let be an acyclic digraph on vertices with maximum degree , degeneracy , and with an -forest partition. Then .
Proof.
Let , and let denote the prefix code consisting of all strings of length . Fix a partition of into directed forests such that every edge between and is oriented from to for all , and such that has weakly connected components of size at most . Let be the forest prefix labeling mapping to the binary representation of . Then the maximum degree of is at most , the maximum component size of is at most , and . Moreover, we can bound the dyadic complexity of as , since every binary string is the prefix of at most codewords in . We now now apply Theorem 3.13 with . We recall that since is a forest for every , we have that by Theorem 4.3. Therefore,
It remains to check that both and satisfy the conditions of Theorem 4.5 for appropriate and . Maximum degree and degeneracy are both easy to control for these graphs, so the nontrivial part is bounding their directed vertex arboricity, or more precisely the parameters of a forest partition.
Consider the random digraph , where . The idea for bounding is to equitably divide into intervals , and note that is in the subcritical regime of the Erdős–Rényi random graph process, so w.h.p. is a union of trees and unicyclic components, each comprising vertices. Each interval can be further divided in two to break the unicyclic components, which shows that w.h.p. we have a -forest partition. The analysis is broadly similar for but somewhat more technical.
Recall that if and is even, a uniformly random undirected -regular graph can be generated using the “pairings model” (also known as the “configurations model”) of Bollobás [3], see the survey of Wormald [39]. The pairings model of generates a random -regular multi-graph (with self-loops allowed) by taking a uniformly random perfect matching (a “pairing”) on points divided into -sets, and then contracting each -set into a single vertex. There are a total of
such pairings, and any simple -regular graph is equally likely to be generated. If is fixed and , it is known that the probability a pairing generates a simple graph is asymptotic to . To generate an honest , repeatedly sample from the above model (expected constant number of samples) until a simple graph is found.
Lemma 4.6.
If is fixed, is even, and is the induced subgraph of on a fixed set of vertices, then w.h.p. every connected component of has order at most and contains at most one cycle.
Proof.
We first show that w.h.p. each component contains at most one cycle.
It is not difficult to show that any minimal graph on vertices with at least two cycles is formed from a path of length by adding edges from each of its ends, see e.g. [27, Theorem 5.5]. Thus, there are at most labelled graphs on vertices with this property. For a fixed such , we bound the probability it appears among fixed vertices in the pairings model for . The total number of pairings giving such an (without giving rise to multi-edges) can be bounded above by
since there are at most ways to choose the edges between -sets that correspond to the edges of , and then at most ways to pair the remaining points. Let be the event that the pairings model generating a simple graph, which occurs with probability . We get that
Thus,
which completes the proof that every component of contains at most one cycle w.h.p.
Next, we show that w.h.p. every connected component has order at most with a similar computation. For a given set size , the number of labelled trees on vertices is by Cayley’s theorem. We bound the probability that a fixed such appears among fixed vertices . Similarly to before, we obtain that the total number of pairings giving such an without multi-edges is bounded above by
since there are at most ways to choose the edges between -sets that correspond to the edges of the tree , and at most ways to pair the remaining points. Conditioning on the pairings model generating a simple graph, which occurs with probability , we get
Taking a union bound over all choices of ,
which is for . This implies that w.h.p. every component of has order at most , and completes the proof. ∎
We can now prove Theorem 1.3.
Corollary 4.7.
For any , we have that w.h.p., as .
Proof.
By Lemma 4.6, we know that w.h.p., any fixed vertices of span a disjoint union of trees and unicyclic components, each of which has order at most . Therefore, by applying the union bound to events, we conclude that w.h.p., has a directed partition into parts all with this property. Dividing each part in two to split every cycle, we conclude that w.h.p. has a -forest partition. Additionally, since the underlying undirected graph of is -regular, we see that is -degenerate and has maximum degree . So by Theorem 4.5, we conclude that w.h.p. as ,
which is upper-bounded by for any fixed and sufficiently large . ∎
Let denote the orientation of the Erdős–Rényi random graph on vertex set where all edges are oriented to the right. Similarly to the above, we can prove a nearly-linear upper bound on .
Corollary 4.8.
For any , we have that w.h.p., , where .
Proof.
It is easy to show that if , then is -degenerate (see [19, Theorem 4.8] for a proof of a stronger result). Additionally, it is well-known that the maximum degree of is for any fixed . Finally, by the easier version of Lemma 4.6 (e.g. [27, Theorem 5.5]), we see that has a -forest partition. Hence, Theorem 4.5 implies that . ∎
5 Multiple colors and ordered Ramsey numbers
In this section we study oriented Ramsey numbers of more than one color, proving Theorems 1.6 and 1.7.
We define an ordered graph to be an undirected graph whose vertex set comes with a total order. If the vertex set is a subset of , then the total order is assumed to be that of . If are ordered graphs, then the ordered Ramsey number is the minimum such that any edge-coloring of the complete graph on in colors contains a monochromatic copy of in color for some . Here an ordered copy of is a subgraph isomorphic to with vertices appearing in the same order. We write when all the graphs are the same.
5.1 The lower bound
To prove Theorem 1.6, we need the following theorem from Conlon, Fox, Lee, and Sudakov [11, Theorem 2.3].
Theorem 5.1.
If is a random matching on vertex set , then w.h.p.,
We remark that the existence of ordered matchings with was proven independently by Balko, Cibulka, Král, and Kynčl [1]. We will only need this weaker result.
If is an acyclic digraph with a Hamiltonian (directed) path, we say is Hamiltonian. It has a unique vertex ordering where consecutive vertices are adjacent and all edges point forwards. We assign to a natural ordered graph on where if and only if in .
Lemma 5.2.
If and is an acyclic Hamiltonian digraph, then
Proof.
If , there exists a -edge-coloring of the complete graph on in which there is no monochromatic copy of . Let be the transitive tournament on with all edges oriented forwards (i.e. if and only if ), and the edge between also colored . Since has a Hamiltonian path and all edges in point forwards in , any copy of in corresponds to an ordered copy of in . By construction, there is no monochromatic copy of in , as desired. ∎
The two results above together can be used to prove Theorem 1.6.
Proof of Theorem 1.6.
Since is nondecreasing in , it suffices to prove the result for . By Theorem 5.1, there exists a matching on which satisfies . Define to be the acyclic digraph on where if and either or is an edge of . By Lemma 5.2, . On the other hand, is a ordered subgraph of , so . It follows that
as desired. Note that is the edge union of a path and a matching, so it has maximum degree . ∎
5.2 The upper bound
To prove Theorem 1.7, we upper bound -color oriented Ramsey numbers by -color ordered Ramsey numbers. Let be the ordered graph obtained from by reversing the vertex order.
Lemma 5.3.
If and is an acyclic digraph, then
Proof.
Let be a tournament on
vertices, with an edge-coloring using colors . Arbitrarily identify with and define a -edge-coloring of by
By the definition of , there is either some color where has an ordered copy of in color , or some color where has an ordered copy of in . In the former case, contains a monochromatic copy of in color with all vertices pointed forwards in the arbitrary ordering, and in the latter case contains a monochromatic copy of in color with all edges pointed backwards. In either case, contains a monochromatic copy of , as desired. ∎
To prove an upper bound on , it remains to generalize the following upper bound of [11] on -color ordered Ramsey numbers to more colors.
Theorem 5.4 ([11, Theorem 3.6]).
If is an ordered graph on at most vertices with degeneracy , then
The multicolor bound can now be obtained by iterating the above theorem.
Theorem 5.5.
If and are -degenerate ordered graphs on at most vertices, then
Proof.
We prove the theorem by induction on . The base case is just Theorem 5.4. For the inductive step, note that if then one obtains
by combining the last colors into one “super-color.” It follows by applying the base case that
as desired. ∎
6 Concluding Remarks
In this section we collect a few appealing open problems on the Ramsey numbers of digraphs. For and , let be an acyclic digraph with vertices and maximum degree maximizing the value of . Much of this paper was devoted to understanding the growth rate of for fixed and .
We first consider the one-color case. Theorem 1.2, Lemma 5.3, and Theorem 5.4 together show
(6.1) |
While we do not know whether the above Ramsey number grows polynomially or super-polynomially in for , we also showed that for and , is at least . We conjecture that a super-polynomial growth rate is also possible for .
Conjecture 6.1.
There exists an absolute constant such that
In the case of colors, Theorems 1.6 and 1.7 together imply
(6.2) |
It would be interesting to determine even the logarithmic order of .
Problem 6.2.
For any fixed and , determine the order of growth of .
To our knowledge, the only solved case of Problem 6.2 is , . Here, must be a disjoint union of arbitrarily oriented paths and non-strongly oriented cycles (an acyclic cannot contain a strongly oriented cycle). Thomason [37] proved that if is large enough, for any non-strongly oriented cycle on vertices, which can be used to show .
It would be interesting to improve either side of (6.2) substantially. We tentatively conjecture that neither side is close to the truth, in the following quantitative form.
Conjecture 6.3.
There exist and constants and such that
In the one-color case, another open problem we find interesting is to determine more precisely the Ramsey number of a sparse random digraph. Theorem 1.3 shows that for fixed , the Ramsey number of is w.h.p. bounded above by . We expect that the answer is in fact linear.
Conjecture 6.4.
If is fixed and , then w.h.p.
This would follow from our techniques if one could prove the following strengthening of Lemma 3.3, in which the size of the -dense pair depends only on the maximum degree (and not on the number of vertices).
Conjecture 6.5.
For every , there exists such that the following holds. Let be a -degenerate digraph with maximum degree on vertices , and let be an arbitrary tournament. If there exist sets , each of size , such that there is no embedding satisfying , then contains a -dense pair with size at least .
Indeed, if one could prove Conjecture 6.5, then it would imply that Lemma 3.10, Lemma 3.11, Theorem 3.13, and Theorem 4.5 would all no longer depend on , the maximum size of a tree component in a directed partition of into oriented forests. In particular, this would imply Conjecture 6.4. Moreover, it is entirely possible that Conjecture 6.5 could be proven using similar greedy embedding arguments, since the conjecture is not hard to prove if the injectivity constraint on is removed.
We would also like to highlight one powerful digraph embedding technique that we have not used in this paper, but are hopeful can be incorporated into our arguments to prove more general results. The median ordering of a tournament is the vertex ordering maximizing the number of forward edges. To see the power of the median ordering, note that for every in this ordering, so this immediately exhibits a Hamiltonian path in . Previous work showing linear upper bounds on when is an oriented tree (e.g. [16]) or an acyclic digraph of bounded bandwidth [14] all depend on embedding into a tournament in some iterative manner according to its median ordering. We were not able to reproduce these upper bounds using greedy embedding arguments, which seem primarily suited for embedding digraphs without long paths.
Finally, we recall from the introduction the directed Ramsey number , which is defined as the least such that every -coloring of the edges of contains a monochromatic copy of . It is easy to see that for any and any acyclic . Indeed, if , then given a -edge-coloring of , we may ignore one edge from each anti-parallel pair to obtain a -edge-colored -vertex tournament, which contains a monochromatic . There is also an inequality in the other direction, whose proof is identical to that given in Lemma 5.3, and which states that , where there are copies of and of , and is obtained from by reversing all the edges. Thanks to these connections, one can convert many of our results on oriented Ramsey numbers to results on directed Ramsey numbers. For example, Theorem 1.7 immediately implies a quasi-polynomial upper bound on for any bounded-degree acyclic digraph . In the other direction, since reversing the edges of any interval mesh yields another interval mesh, Theorem 1.2 shows the existence of a bounded-degree acyclic digraph with which grows faster than any fixed polynomial in the number of vertices of . For colors, we can similarly use Theorem 1.6 to produce a bounded-degree acyclic digraph such that grows super-polynomially. However, there is an interesting intermediate case at colors, and we end with the following conjecture, which may be easier than Conjecture 6.1.
Conjecture 6.6.
There is an absolute constant and an infinite sequence of -vertex acyclic digraphs with maximum degree at most and .
Acknowledgments. We would like to thank Jasmine Yan for producing Figure 3.1, and the anonymous referee for many helpful comments.
References
- [1] M. Balko, J. Cibulka, K. Král, and J. Kynčl, Ramsey numbers of ordered graphs, Electron. J. Comb. 27 (2020), P1.16.
- [2] J.-C. Bermond, Some Ramsey numbers for directed graphs, Discrete Math. 9 (1974), 313–321.
- [3] B. Bollobás, A probabilistic proof of an asymptotic formula for the number of labelled regular graphs, European J. Combin. 1 (1980), 311–316.
- [4] M. Bucić, S. Letzter, and B. Sudakov, Directed Ramsey number for trees, J. Combin. Theory Ser. B 137 (2019), 145–177.
- [5] S. A. Burr, An inequality involving the vertex arboricity and edge arboricity of a graph, J. Graph Theory 10 (1986), 403–404.
- [6] S. A. Burr and P. Erdős, On the magnitude of generalized Ramsey numbers for graphs, in: Infinite and Finite Sets I, Colloq. Math. Soc Janos Bolyai 10, North-Holland, Amsterdam (1975), 214–240.
- [7] V. Chvátal, Monochromatic paths in edge-colored graphs, J. Combin. Theory Ser. B 13 (1972), 69–70.
- [8] V. Chvátal, V. Rödl, E. Szemerédi and W. T. Trotter Jr., The Ramsey number of a graph with bounded maximum degree, J. Combin. Theory Ser. B 34 (1983), 239–243.
- [9] D. Conlon, A new upper bound for diagonal Ramsey numbers, Ann. of Math. 170 (2009), 941–960.
- [10] D. Conlon, The Ramsey number of dense graphs, Bull. Lond. Math. Soc. 45 (2013), 483–496.
- [11] D. Conlon, J. Fox, C. Lee, and B. Sudakov, Ordered Ramsey numbers, J. Combin. Theory Ser. B 122 (2017), 353–383.
- [12] D. Conlon, J. Fox, and B. Sudakov, Recent developments in graph Ramsey theory, London Math. Soc. Lecture Note Ser. 424 (2015), 49–118.
- [13] D. Conlon, J. Fox, and B. Sudakov, Short proofs of some extremal results II, J. Combin. Theory Ser. B 121 (2016), 173–196.
- [14] N. Draganić, F. Dross, J. Fox, A. Girão, F. Havet, D. Korándi, W. Lochet, D. Munhá Correia, A. Scott, and B. Sudakov, Powers of paths in tournaments, Combin. Probab. Comput. 30 (2021), 894–898.
- [15] F. Dross and F. Havet, On the unavoidability of oriented trees, Electron. Notes Theor. Comput. Sci. 346 (2019), 425–436.
- [16] A. El Sahili, Trees in tournaments, J. Combin. Theory Ser. B 92 (2004), 183–187.
- [17] P. Erdős and A. Hajnal, Ramsey-type theorems, Discrete Appl. Math. 25 (1989), 37–52.
- [18] P. Erdős and L. Moser, On the representation of directed graphs as unions of orderings, Publ. Math. Inst. Hung. Acad. Sci., Ser. A 9 (1964), 125–132.
- [19] J. Fox and B. Sudakov, Two remarks on the Burr–Erdős conjecture, European J. Combin. 30 (2009), 1630–1645.
- [20] T. Gallai, On directed paths and circuits, Theory of graphs (Proc. Colloq., Tihany, 1966) (1968), 115–118.
- [21] R. L. Graham, V. Rödl, and A. Ruciński, On graphs with linear Ramsey numbers, J. Graph Theory 35 (2000), 176–192.
- [22] A. Gyárfás and J. Lehel, A Ramsey-type problem in directed and bipartite graphs, Period. Math. Hungar. 3 (1973), 299–304.
- [23] R. Häggkvist and A. Thomason, Trees in tournaments, Combinatorica 11 (1991), 123–130.
- [24] M. Hasse, Zur algebraischen Begründung der Graphentheorie. I, Mathematische Nachrichten 28 (1965), 275–290.
- [25] F. Havet, Trees in tournaments, Discr. Math. 243 (2002), 121–134.
- [26] F. Havet and S. Thomassé, Median orders of tournaments: A tool for the second neighborhood problem and Sumner’s conjecture, J. Graph Theory 35 (2000), 244–256.
- [27] S. Janson, T. Łuczak, and A. Ruciński, Random graphs, Wiley, New York, 2000.
- [28] A. Kostochka and B. Sudakov, On Ramsey numbers of sparse graphs, Combin. Probab. Comput. 12 (2003), 627–641.
- [29] D. Kühn, R. Mycroft, and D. Osthus, A proof of Sumner’s universal tournament conjecture for large tournaments, Proc. Lond. Math. Soc. (3) 102 (2011), 731–766.
- [30] C. Lee, Ramsey numbers of degenerate graphs, Ann. of Math. 185 (2017), 791–829.
- [31] L. Lovász, Graphs and Geometry, American Mathematical Society Colloquium Publications, American Mathematical Society, Providence, RI, 2019.
- [32] B. Roy, Nombre chromatique et plus longs chemins d’un graphe, Rev. Française Informat. Recherche Opérationnelle 1 (1967), 129–132.
- [33] A. Sah, Diagonal Ramsey via effective quasirandomness, 2020. Preprint available at arXiv:2005.09251.
- [34] J. Spencer, Ramsey’s theorem—a new lower bound, J. Combin. Theory Ser. A 18 (1975), 108–115.
- [35] R. Stearns, The voting problem, Amer. Math. Monthly 66 (1959), 761–763.
- [36] B. Sudakov, A conjecture of Erdős on graph Ramsey numbers, Adv. Math. 227 (2011), 601–609.
- [37] A. Thomason, Paths and cycles in tournaments, Trans. Amer. Math. Soc. 296 (1986), 167–180.
- [38] L. M. Vitaver, Determination of minimal coloring of vertices of a graph by means of Boolean powers of the incidence matrix, Dokl. Akad. Nauk SSSR 147 (1962), 728.
- [39] N. C. Wormald, Models of random regular graphs. In Surveys in combinatorics, 1999 (Canterbury), volume 267 of London Math. Soc. Lecture Note Ser., pages 239–298. Cambridge Univ. Press, Cambridge, 1999.
- [40] R. Yuster, Paths with many shortcuts in tournaments, Discrete Math. 334 (2021), 112–168.