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

Ramsey numbers of sparse digraphs

Jacob Fox Department of Mathematics, Stanford University, Stanford, CA 94305, USA. Email: [email protected]. Research supported by a Packard Fellowship and by NSF award DMS-185563.    Xiaoyu He Department of Mathematics, Stanford University, Stanford, CA 94305, USA. Email: [email protected]. Research supported by NSF GRFP Grant DGE-1656518.    Yuval Wigderson Department of Mathematics, Stanford University, Stanford, CA 94305, USA. Email: [email protected]. Research supported by NSF GRFP Grant DGE-1656518.
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 HH is an acyclic digraph, the oriented Ramsey number of HH, denoted r1(H)\overrightarrow{r_{1}}(H), is the least NN such that every tournament on NN vertices contains a copy of HH. We show that for any Δ2\Delta\geq 2 and any sufficiently large nn, there exists an acyclic digraph HH with nn vertices and maximum degree Δ\Delta such that

r1(H)nΩ(Δ2/3/log5/3Δ).\overrightarrow{r_{1}}(H)\geq n^{\Omega(\Delta^{2/3}/\log^{5/3}\Delta)}.

This proves that r1(H)\overrightarrow{r_{1}}(H) is not always linear in the number of vertices for bounded-degree HH. On the other hand, we show that r1(H)\overrightarrow{r_{1}}(H) is nearly linear in the number of vertices for typical bounded-degree acyclic digraphs HH, 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 rk(H)=2(logn)Ok(1)\overrightarrow{r_{k}}(H)=2^{(\log n)^{O_{k}(1)}} for all bounded-degree acyclic digraphs HH on nn vertices, where rk(H)\overrightarrow{r_{k}}(H) is the least NN such that every kk-edge-colored tournament on NN vertices contains a monochromatic copy of HH. For k2k\geq 2 and n4n\geq 4, we exhibit an acyclic digraph HH with nn vertices and maximum degree 33 such that rk(H)nΩ(logn/loglogn)\overrightarrow{r_{k}}(H)\geq n^{\Omega(\log n/\log\log n)}, showing that these Ramsey numbers can grow faster than any polynomial in the number of vertices.

1 Introduction

The kk-color Ramsey number of a (simple undirected) graph HH, denoted rk(H)r_{k}(H), is the minimum NN such that every kk-edge-coloring of the complete graph KNK_{N} contains a monochromatic copy of HH. Broadly speaking, the main question in graph Ramsey theory is to understand how rk(H)r_{k}(H) depends on HH and kk. The most well-studied case is that of two colors, k=2k=2. For HH on nn vertices, it is known [10, 36] that r2(H)r_{2}(H) grows exponentially in nn if and only if HH has Ω(n2)\Omega(n^{2}) edges.

However, it has long been observed that the Ramsey number of a sparse graph HH is much smaller than exponential in |V(H)||V(H)|. 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: HH is said to be dd-degenerate if every subgraph of HH has a vertex of degree at most dd, and the degeneracy of HH is the minimum dd such that HH is dd-degenerate. Burr and Erdős conjectured that rk(H)=Ok,d(n)r_{k}(H)=O_{k,d}(n) for any nn-vertex graph HH with degeneracy dd. Here and throughout we use the standard asymptotic notation where the implicit constant is allowed to depend only on the subscripts of O()O(\cdot). 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 HH has bounded degree (rather than bounded degeneracy), that is, that rk(H)=Ok,Δ(n)r_{k}(H)=O_{k,\Delta}(n) for any nn-vertex graph HH with maximum degree Δ\Delta. 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 HH, define the kk-color oriented Ramsey number rk(H)\overrightarrow{r_{k}}(H) to be the minimum NN such that any kk-edge-colored tournament on NN vertices contains a monochromatic copy of HH. Note that if HH contains a directed cycle, then HH does not appear in any transitive tournament, so rk(H)\overrightarrow{r_{k}}(H) only exists for acyclic HH. Henceforth, we work exclusively with acyclic digraphs HH.

Unlike undirected Ramsey numbers, oriented Ramsey numbers are interesting even in the case of one color, k=1k=1, where r1(H)\overrightarrow{r_{1}}(H) is simply the minimum NN such that any tournament on NN vertices contains a copy of HH. Let TTn\textnormal{TT}_{n} denote the transitive tournament on nn 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

2n/21r1(TTn)2n1.2^{n/2-1}\leq\overrightarrow{r_{1}}(\textnormal{TT}_{n})\leq 2^{n-1}. (1.1)

The exponential constants in these bounds have not been improved, similar to the classical case of the diagonal undirected Ramsey number r2(Kn)r_{2}(K_{n}) [10, 33, 34]. This may not be surprising, given that r1(TTn)r2(Kn)\overrightarrow{r_{1}}(\textnormal{TT}_{n})\leq r_{2}(K_{n}). Thus, improving the lower bound in (1.1) is at least as difficult as improving the lower bound on r2(Kn)r_{2}(K_{n}).

Somewhat more is known about the oriented Ramsey number rk(H)\overrightarrow{r_{k}}(H) when HH is sparse. When H=PnH=P_{n} is the directed path on nn vertices, Chvátal [7] and Gyárfás and Lehel [22] determined that rk(Pn)=(n1)k+1\overrightarrow{r_{k}}(P_{n})=(n-1)^{k}+1 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 TT on n2n\geq 2 vertices, r1(T)2n2\overrightarrow{r_{1}}(T)\leq 2n-2. 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 nn sufficiently large by Kühn, Mycroft, and Osthus [29], and Dross and Havet [15] showed that r1(T)218n4716\overrightarrow{r_{1}}(T)\leq\frac{21}{8}n-\frac{47}{16} for all n2n\geq 2. In more colors, it was shown by Bucić, Letzter, and Sudakov [4] that rk(T)=Ok(|V(T)|k)\overrightarrow{r_{k}}(T)=O_{k}(|V(T)|^{k}) for any oriented tree TT and all k1k\geq 1. In the same paper, they asked a natural directed analogue of the classical Burr–Erdős problem.

Problem 1.1 ([4]).

Is it true that r1(H)=OΔ(n)\overrightarrow{r_{1}}(H)=O_{\Delta}(n) for every acyclic digraph HH with nn vertices and maximum degree Δ\Delta?

Here we write N+(v)N^{+}(v) for the out-neighborhood and N(v)N^{-}(v) for the in-neighborhood of a vertex vV(H)v\in V(H), and say that a digraph HH has maximum degree Δ\Delta if maxvV(H)(|N+(v)|+|N(v)|)=Δ\max_{v\in V(H)}(|N^{+}(v)|+|N^{-}(v)|)=\Delta.

Yuster [40] recently initiated the study of the special case of Problem 1.1 when H=PnH=P_{n}^{\ell} is the \ell-th power of a directed path PnP_{n}, which is the digraph on vertex set [n][n] whose edges are the ordered pairs (i,j)(i,j) satisfying 1ji1\leq j-i\leq\ell. This case was recently settled by Draganić et al. [14], who showed that r1(Pn)=O(n)\overrightarrow{r_{1}}(P_{n}^{\ell})=O_{\ell}(n). Letting the bandwidth of an acyclic digraph HH on nn vertices be the minimum \ell such that PnP_{n}^{\ell} contains a copy of HH, this aforementioned result implies that r1(H)=O(n)\overrightarrow{r_{1}}(H)=O_{\ell}(n) if HH has nn vertices and bandwidth \ell.

In this paper, we answer Problem 1.1 in the negative, and show that in fact r1(H)\overrightarrow{r_{1}}(H) can grow faster than any fixed power of nn, as long as the maximum degree is a sufficiently large constant.

Theorem 1.2.

For any Δ2\Delta\geq 2 and nn sufficiently large in terms of Δ\Delta, there exists an acyclic digraph HH on nn vertices and maximum degree at most Δ\Delta for which

r1(H)nΩ(Δ2/3/log5/3Δ).\overrightarrow{r_{1}}(H)\geq n^{\Omega(\Delta^{2/3}/\log^{5/3}\Delta)}.

The power Δ2/3\Delta^{2/3} seems to be best possible with our method, but the power of logΔ\log\Delta can be improved. Although the answer to Problem 1.1 is negative, we prove an almost linear upper bound on r1(H)\overrightarrow{r_{1}}(H) for almost all HH, in the following sense. Define G(n,d)\overrightarrow{G}(n,d) to be the orientation of the random regular graph G(n,d)G(n,d) on vertex set [n][n] with all edges pointing to the right.

Theorem 1.3.

If d2d\geq 2 is fixed and H=G(n,d)H=\overrightarrow{G}(n,d), then w.h.p.111As usual, we say that an event \mathcal{E} happens with high probability (w.h.p.) if Pr()1\Pr(\mathcal{E})\to 1 as nn\to\infty. (as nn\rightarrow\infty)

r1(H)n(logn)4logd.\overrightarrow{r_{1}}(H)\leq n(\log n)^{4\log d}.

It is not difficult to extend Theorem 1.3 to show r1(H)=n(logn)OΔ(1)\overrightarrow{r_{1}}(H)=n(\log n)^{O_{\Delta}(1)} w.h.p. if HH is the forward acyclic orientation of a uniformly random graph with any fixed degree sequence Δ=d1d2dn\Delta=d_{1}\geq d_{2}\geq\cdots\geq d_{n}, and therefore also for the forward acyclic orientation of a uniformly random bounded-degree graph. We also prove a similar bound r1(H)n(logn)Od(1)\overrightarrow{r_{1}}(H)\leq n(\log n)^{O_{d}(1)} when H=G(n,p)H=\overrightarrow{G}(n,p) is the forward acyclic orientation of an Erdős–Rényi random graph of constant average degree d=pnd=pn.

Although we are able to show that r1(H)\overrightarrow{r_{1}}(H) 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 HH on nn vertices with maximum degree Δ\Delta, the best upper bound we are able to prove is r1(H)nOΔ(logn)\overrightarrow{r_{1}}(H)\leq n^{O_{\Delta}(\log n)} (see Theorem 1.7 below). Nonetheless, we are able to prove stronger (and in some cases linear) upper bounds on r1(H)\overrightarrow{r_{1}}(H) in case HH lies in certain natural families. We now give two examples of such families.

Let the height of HH be the number of vertices on the longest directed path in HH. Equivalently, the height can be seen as a directed analogue of the chromatic number: HH has height at most hh if and only if V(H)V(H) can be partitioned into independent sets S1,,ShS_{1},\dotsc,S_{h} such that every edge between SiS_{i} and SjS_{j} is oriented from SiS_{i} to SjS_{j}, for every i<ji<j. For acyclic digraphs of bounded height and bounded degree, we are able to prove the following linear bound on r1(H)\overrightarrow{r_{1}}(H).

Theorem 1.4.

If HH is an acyclic digraph on nn vertices with maximum degree Δ\Delta and height hh, then

r1(H)(Δh)10Δloghn.\overrightarrow{r_{1}}(H)\leq(\Delta h)^{10\Delta\log h}n.

In particular, r1(H)=Oh,Δ(n)\overrightarrow{r_{1}}(H)=O_{h,\Delta}(n).

Note that this theorem also implies the aforementioned nOΔ(logn)n^{O_{\Delta}(\log n)} upper bound on r1(H)\overrightarrow{r_{1}}(H) for any bounded-degree acyclic digraph HH, since the height of an acyclic digraph is at most its vertex count.

Next, we say that an acyclic digraph HH of height hh is graded if its vertex set can be partitioned into hh independent sets S1,,ShS_{1},\dotsc,S_{h} such that every edge in HH is directed from some SiS_{i} to Si+1S_{i+1}. Equivalently, HH is graded if for every pair of vertices (u,v)(u,v), all directed paths from uu to vv 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) LL by orienting every edge x<yx<y of the Hasse diagram of LL from xx to yy. For graded digraphs of bounded degree, we are able to prove a polynomial bound on r1(H)\overrightarrow{r_{1}}(H).

Theorem 1.5.

If HH is a graded digraph on nn vertices with maximum degree Δ\Delta and height hh, then

r1(H)h10ΔlogΔn.\overrightarrow{r_{1}}(H)\leq h^{10\Delta\log\Delta}n.

In particular, since hnh\leq n, we have that r1(H)nO(ΔlogΔ)\overrightarrow{r_{1}}(H)\leq n^{O(\Delta\log\Delta)}.

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 22. Using this connection, we are able to give a super-polynomial lower bound for rk(H)\overrightarrow{r_{k}}(H) when k2k\geq 2.

Theorem 1.6.

For any n4n\geq 4, there exists an acyclic digraph HH on nn vertices with maximum degree 33 for which

rk(H)nlogn/20loglogn\overrightarrow{r_{k}}(H)\geq n^{\log n/20\log\log n}

for all k2k\geq 2.

Thus, for acyclic digraphs of bounded degree, rk(H)\overrightarrow{r_{k}}(H) can grow super-polynomially if k2k\geq 2. In the other direction, for any number of colors we have the following quasi-polynomial upper bound.

Theorem 1.7.

If k1k\geq 1 and HH is any acyclic digraph with nn vertices and maximum degree Δ\Delta, then

rk(H)2Ok,Δ((logn)22k1).\overrightarrow{r_{k}}(H)\leq 2^{O_{k,\Delta}\left((\log n)^{2^{2k-1}}\right)}.

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 rk(H)\overleftrightarrow{r_{k}}(H), introduced by Bermond [2]. This is defined as the minimum NN such that a monochromatic copy of HH exists in every kk-coloring of the edges of KN\overleftrightarrow{K_{N}}, 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 HH, the degeneracy and the number of vertices are the main parameters that determine the growth of r2(H)r_{2}(H) (we focus on the two-color case to be concrete). This is easiest to see in the lower bounds: if HH has nn vertices, then certainly r2(H)nr_{2}(H)\geq n. Additionally, if HH has degeneracy dd, then it is a simple exercise to show that a random 22-edge-coloring on 2d/22^{d/2} vertices does not contain a monochromatic copy of HH w.h.p., implying that r2(H)2d/2r_{2}(H)\geq 2^{d/2}. Putting these two facts together, we find that

logr2(H)=Ω(d+logn).\log r_{2}(H)=\Omega(d+\log n).

Conlon, Fox, and Sudakov [12, Conjecture 2.16] conjectured that this bound is tight up to the implicit constant, namely that

logr2(H)=Θ(d+logn)\log r_{2}(H)=\Theta(d+\log n)

for any graph HH with nn vertices and degeneracy dd. Moreover, this conjecture is known to be true up to a multiplicative factor of log2d\log^{2}d [19, Theorem 3.1]. Because of these results, one can say that the degeneracy and the vertex count of HH roughly determine its Ramsey number.

For acyclic digraphs, we do not know what parameters determine the growth order of r1(H)\overrightarrow{r_{1}}(H) (indeed, we don’t even know if r1(H)\overrightarrow{r_{1}}(H) is polynomial or super-polynomial in nn when HH has bounded degree). Nonetheless, our results indicate that one parameter, which we call “multiscale complexity,” is relevant. Namely, suppose we order the vertices of HH as v1,,vnv_{1},\dotsc,v_{n} so that every edge is oriented to the right, that is vivjv_{i}\to v_{j} is an edge only if i<ji<j (such an ordering is often called a topological sorting of HH). Under this ordering we may assign every edge vivjv_{i}\to v_{j} a length jij-i. Here, if u,vV(H)u,v\in V(H), we write uvu\rightarrow v to signify that there is a directed edge from uu to vv, and similarly uvu\leftarrow v for an edge in the other direction.

In graphs of bounded bandwidth, every edge is short and has length O(1)O(1). At the other extreme, if HH has bounded height, then most edges of HH are long and have length Ω(n)\Omega(n). The same is true in the random model G(n,d)\overrightarrow{G}(n,d), where most edges have length Ω(n)\Omega(n). Some other acyclic digraphs have intermediate edge-length statistics, such as the directed grid whose vertex set is [n]2[\sqrt{n}]^{2} 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 11 and Ω(n)\Omega(\sqrt{n}).

Loosely, let us say that HH has high multiscale complexity if, for most dyadic intervals It=[2t,2t+1)I_{t}=[2^{t},2^{t+1}) with 0tlogn0\leq t\leq\log n, there are many edges in HH whose length is in ItI_{t}. Conversely, if most edge lengths of HH are concentrated in a small number of dyadic intervals, then we loosely say that HH has low multiscale complexity. At a high level, all of our upper bound results prove upper bounds on r1(H)\overrightarrow{r_{1}}(H) in terms of n,Δn,\Delta, and the multiscale complexity of HH; if the multiscale complexity is low, then these bounds are stronger, and one can prove linear, nearly linear, or polynomial bounds on r1(H)\overrightarrow{r_{1}}(H) (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 Δ\Delta.

We stress that we have not fully solved the problem of which natural parameters determine the growth order of r1(H)\overrightarrow{r_{1}}(H), 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 1.31.4, 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 HTH\hookrightarrow T, we mean an injective function V(H)V(T)V(H)\to V(T) such that edges of HH are mapped to edges of TT, with edge orientations preserved. We say that a digraph TT is HH-free if there is no embedding HTH\hookrightarrow T. All logarithms are to base 22. 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 Δ1\Delta\geq 1 there exists a family of acyclic digraphs {Hn}n1\{H_{n}\}_{n\geq 1} with maximum degree Δ\Delta for which |V(Hn)|=n|V(H_{n})|=n and r1(Hn)nΩ(Δ2/3/log5/3Δ)\overrightarrow{r_{1}}(H_{n})\geq n^{\Omega(\Delta^{2/3}/\log^{5/3}\Delta)}. The lower bound consists of three ingredients: constructing a bounded degree acyclic digraph HH that is hard to embed, constructing a Ramsey tournament TT that is hard to embed HH into, and proving that there is no embedding HTH\hookrightarrow T.

The next three subsections separately provide these three ingredients. Section 2.1 defines a class of digraphs HH with edges “at all scales,” which we call interval meshes, and proves the existence of bounded-degree HH with this property. Section 2.2 defines the Ramsey tournament TT as a lexicographic power of a tournament RR with no large transitive subtournament, and shows that embeddings HTH\hookrightarrow T correspond to certain highly constrained walks on RR which we call (R,f,s)(R,f,s)-walks. Section 2.3 completes the proof by showing that long (R,f,s)(R,f,s)-walks do not exist, and therefore large interval meshes HH cannot be embedded into small powers T=RmT=R^{m}.

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 [n][n]; 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 HH.

Definition 2.1.

If f:>0f:\mathbb{N}\rightarrow\mathbb{R}_{>0} is a nondecreasing function, we define an ff-interval mesh to be an acyclic digraph HH whose vertex set is an interval II\subseteq\mathbb{N} such that any edge (i,j)E(H)(i,j)\in E(H) has i<ji<j and for all pairs of intervals (a1,b1],(a2,b2]I(a_{1},b_{1}],(a_{2},b_{2}]\subseteq I with b1a2b_{1}\leq a_{2} and

a2b1f(min(b1a1,b2a2)),a_{2}-b_{1}\leq f(\min(b_{1}-a_{1},b_{2}-a_{2})), (2.1)

there exists an edge in HH between (a1,b1](a_{1},b_{1}] and (a2,b2](a_{2},b_{2}]. When the function ff is clear from context, we simply call HH 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 f:>0f:\mathbb{N}\rightarrow\mathbb{R}_{>0} with S=m0f(2m+2)22m<S=\sum_{m\geq 0}f(2^{m+2})\cdot 2^{-2m}<\infty, there exists an ff-interval mesh HH on vertex set \mathbb{N} with maximum degree at most 2S+172S+17.

Proof.

Starting from an empty digraph on \mathbb{N}, we construct HH by using a greedy algorithm to add edges. All edges introduced point to the right, so the resulting digraph must be acyclic. Define Im,jI_{m,j} to be the dyadic interval (j2m,(j+1)2m](j\cdot 2^{m},(j+1)\cdot 2^{m}] with length 2m2^{m}.

Let m0m\geq 0 range through the nonnegative integers. On subroutine mm, we iterate through all pairs (i,j)02(i,j)\in\mathbb{N}_{\geq 0}^{2} satisfying

1jif(2m+2)2m+41\leq j-i\leq f(2^{m+2})\cdot 2^{-m}+4 (2.2)

and add an edge between the (currently) lowest degree vertex of Im,iI_{m,i} and the (currently) lowest degree vertex of Im,jI_{m,j}, if an edge does not exist between Im,iI_{m,i} and Im,jI_{m,j} already. Writing d(U)d(U) for the sum of the degrees of the vertices in a set UU, we see that for any given ii, subroutine mm increases d(Im,i)d(I_{m,i}) by a total of at most

2[f(2m+2)2m+4]=2m+1f(2m+2)+8.2[f(2^{m+2})\cdot 2^{-m}+4]=2^{-m+1}\cdot f(2^{m+2})+8.

Let HH be the digraph produced from this infinite process. Explicitly, HH is the edge union of all the graphs H(m)H^{(m)}, where H(m)H^{(m)} is the graph produced after subroutine mm. It has the property that every pair of dyadic intervals Im,iI_{m,i}, Im,jI_{m,j} satisfying (2.2) has an edge between them.

We first check that HH has maximum degree at most 2S+172S+17. Since Im,iI_{m,i} is a union of 2mk2^{m-k} dyadic intervals of length 2k2^{k}, we have that after subroutine mm,

d(Im,i)k=0m2mk(2k+1f(2k+2)+8)2m+1S+2m+18=2m(2S+16).d(I_{m,i})\leq\sum_{k=0}^{m}2^{m-k}(2^{-k+1}\cdot f(2^{k+2})+8)\leq 2^{m+1}\cdot S+2^{m+1}\cdot 8=2^{m}(2S+16).

In particular, the average degree in Im,iI_{m,i} is less than 2S+172S+17 at the end of subroutine mm. However, subroutine mm only adds edges incident to vertices of Im,iI_{m,i} which have at most average degree, so no new vertex of degree at least 2S+182S+18 can appear on subroutine mm. Thus, no vertex of degree at least 2S+182S+18 ever appears, as desired.

Next, we check that HH is an ff-interval mesh. Suppose two intervals (a1,b1](a_{1},b_{1}], (a2,b2](a_{2},b_{2}] with b1a2b_{1}\leq a_{2} satisfy (2.1) and let mm be the largest positive integer such that (a1,b1](a_{1},b_{1}] and (a2,b2](a_{2},b_{2}] both contain dyadic intervals of length 2m2^{m}. Note that in any interval (a,b](a,b] of integers, a longest dyadic subinterval I,iI_{\ell,i} has length 2[ba4,ba]2^{\ell}\in[\frac{b-a}{4},b-a], so in particular 2m+2min(b1a1,b2a2).2^{m+2}\geq\min(b_{1}-a_{1},b_{2}-a_{2}). Next, pick ii maximal and jj minimal such that Im,i(a1,b1]I_{m,i}\subseteq(a_{1},b_{1}] and Im,j(a2,b2]I_{m,j}\subseteq(a_{2},b_{2}].

Using (2.1), we find

ji4+a2b12m4+f(min(b1a1,b2a2))2m4+f(2m+2)2m,j-i\leq 4+\frac{a_{2}-b_{1}}{2^{m}}\leq 4+\frac{f(\min(b_{1}-a_{1},b_{2}-a_{2}))}{2^{m}}\leq 4+\frac{f(2^{m+2})}{2^{m}},

and so there is an edge in HH between Im,iI_{m,i} and Im,jI_{m,j}, and therefore between (a1,b1](a_{1},b_{1}] and (a2,b2](a_{2},b_{2}] as well. ∎

The acyclic digraphs HnH_{n} we use are induced subgraphs on intervals of the infinite interval mesh constructed in Lemma 2.2, for an appropriate choice of ff.

2.2 Walks in tournaments

Next, we construct the large tournament TT which is difficult to embed HH 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 GHG\cdot H of two digraphs GG and HH is the digraph on vertex set V(G)×V(H)V(G)\times V(H) with an edge (g1,h1)(g2,h2)(g_{1},h_{1})\rightarrow(g_{2},h_{2}) iff g1g2g_{1}\rightarrow g_{2} in GG or g1=g2g_{1}=g_{2} and h1h2h_{1}\rightarrow h_{2} in HH. Henceforth, write GmG^{m} for the mm-fold lexicographic product of GG with itself. Note that lexicographic powers of tournaments are tournaments. By (1.1), there exists for any r3r\geq 3 a tournament RR on rr vertices with no transitive subtournament of size 2logr+22\log r+2. We take T=RmT=R^{m} and show that an interval mesh HH is difficult to embed into TT.

To this end, let HH be an interval mesh. We relate embeddings HRmH\hookrightarrow R^{m} to certain constrained walks on the tournament RR.

Definition 2.3.

For a tournament RR, a nondecreasing function f:>0f:\mathbb{N}\rightarrow\mathbb{R}_{>0}, and s1s\geq 1, define an (R,f,s)(R,f,s)-walk to be a sequence of ordered pairs {(vi,ai)}i=1\{(v_{i},a_{i})\}_{i=1}^{\ell} ((\ell possibly infinite) where for each ii, viV(R)v_{i}\in V(R), 1ais1\leq a_{i}\leq s, vivi+1v_{i}\neq v_{i+1}, and for any pair i<ji<j for which vivjv_{i}\leftarrow v_{j} in RR, we have

a(i,j)>f(min(ai,aj)),a_{(i,j)}>f(\min(a_{i},a_{j})),

where aIkIaka_{I}\coloneqq\sum_{k\in I}a_{k} if II is an interval of integers and the empty sum equals 0. We define the length of an (R,f,s)(R,f,s)-walk to be a[1,]a_{[1,\ell]}.

Let LR,f(s)L_{R,f}(s) be the length of the longest (R,f,s)(R,f,s)-walk if such a maximum exists, and ++\infty otherwise. The next lemma reduces Theorem 1.2 to showing asymptotic upper bounds on LR,f(s)L_{R,f}(s).

Lemma 2.4.

Suppose there exists r1r\geq 1, a nondecreasing f:>0f:\mathbb{N}\rightarrow\mathbb{R}_{>0}, and a tournament RR on rr vertices for which lim supsLR,f(s)s1=α\limsup_{s\rightarrow\infty}L_{R,f}(s)s^{-1}=\alpha. If HH is an ff-interval mesh on nn vertices, then

r1(H)nlogrlogαo(1).\overrightarrow{r_{1}}(H)\geq n^{\frac{\log r}{\log\alpha}-o(1)}.
Proof.

For each n1n\geq 1, let m=m(n)m=m(n) be the minimum positive integer for which there exists an ff-interval mesh HH with vertex set [n][n] and an embedding ϕ:HRm\phi:H\hookrightarrow R^{m}.

Let π:RmR\pi:R^{m}\rightarrow R be the projection onto the first coordinate. Consider the sequence {πϕ(j)}j=1n\{\pi\circ\phi(j)\}_{j=1}^{n}. Consecutive terms of this sequence may repeat, so we block the sequence into consecutive runs of identical vertices. Suppose there are \ell total runs I1I=[n]I_{1}\sqcup\cdots\sqcup I_{\ell}=[n] and run IiI_{i} consists of aia_{i} repetitions of vertex viV(R)v_{i}\in V(R). We claim that {(vi,ai)}i=1\{(v_{i},a_{i})\}_{i=1}^{\ell} is an (R,f,s)(R,f,s)-walk, where smax(ai)s\coloneqq\max(a_{i}). It is easy to see that 1ais1\leq a_{i}\leq s for all ii, and that vivi+1v_{i}\neq v_{i+1} since we already blocked consecutive identical vertices together. It remains to check the key condition, that if i<ji<j and vivjv_{i}\leftarrow v_{j} in RR, we must have

a(i,j)>f(min(ai,aj)).a_{(i,j)}>f(\min(a_{i},a_{j})).

Suppose this is not true, so there exists some i<ji<j with vivjv_{i}\leftarrow v_{j} and a(i,j)f(min(ai,aj))a_{(i,j)}\leq f(\min(a_{i},a_{j})). By the definition of vi,vjv_{i},v_{j}, we have that π(ϕ(Ii))={vi}\pi(\phi(I_{i}))=\{v_{i}\}, π(ϕ(Ij))={vj}\pi(\phi(I_{j}))=\{v_{j}\}. By the definition of the lexicographic power, if vivjv_{i}\leftarrow v_{j} then wiwjw_{i}\leftarrow w_{j} for all wiπ1(vi),wjπ1(vj)w_{i}\in\pi^{-1}(v_{i}),w_{j}\in\pi^{-1}(v_{j}). Thus, all edges between ϕ(Ii)\phi(I_{i}) and ϕ(Ij)\phi(I_{j}) are oriented from ϕ(Ij)\phi(I_{j}) to ϕ(Ii)\phi(I_{i}). For ϕ\phi to be a homomorphism, this means that HH has no edge oriented from IiI_{i} to IjI_{j}. On the other hand, |Ii|=ai|I_{i}|=a_{i}, |Ij|=aj|I_{j}|=a_{j}, and the distance between the two intervals is a(i,j)f(min(ai,aj))a_{(i,j)}\leq f(\min(a_{i},a_{j})), so since HH is an ff-interval mesh there is such an edge. This is a contradiction, and we see that {(vi,ai)}i=1\{(v_{i},a_{i})\}_{i=1}^{\ell} is an (R,f,s)(R,f,s)-walk of length nn.

Fix any ε>0\varepsilon>0. We are given that LR,f(s)(α+ε)sL_{R,f}(s)\leq(\alpha+\varepsilon)s for sufficiently large ss, so we get s(α+ε)1ns\geq(\alpha+\varepsilon)^{-1}n using the fact that nLR,f(s)n\leq L_{R,f}(s) since we have found an (R,f,s)(R,f,s)-walk of length nn. Since s=max(ai)s=\max(a_{i}), this means that there is some subinterval Ii[n]I_{i}\subseteq[n] of length at least (α+ε)1n(\alpha+\varepsilon)^{-1}n for which πϕ\pi\circ\phi is constant on IiI_{i}, i.e. the image ϕ(Ii){vi}×Rm1\phi(I_{i})\subseteq\{v_{i}\}\times R^{m-1} lies in a copy of Rm1R^{m-1}. Putting this together, we have shown that for large enough nn, the existence of an embedding ϕ:HRm\phi:H\hookrightarrow R^{m} implies the existence of an embedding ϕ:HRm1\phi^{\prime}:H^{\prime}\hookrightarrow R^{m-1} for some ff-interval mesh HH^{\prime} on (α+ε)1n(\alpha+\varepsilon)^{-1}n vertices. In other words,

m(n)m((α+ε)1n)+1m(n)\geq m((\alpha+\varepsilon)^{-1}n)+1

for all nn sufficiently large. Applying this recursively, we obtain that m(n)lognlog(α+ε)Oε(1)m(n)\geq\frac{\log n}{\log(\alpha+\varepsilon)}-O_{\varepsilon}(1). By the definition of m(n)m(n), any ff-interval mesh HH on nn vertices has no embedding into Rm(n)1R^{m(n)-1}, and so

r1(H)>|V(Rm(n)1)|rlognlog(α+ε)Oε(1)=Ωε(nlogrlog(α+ε)),\overrightarrow{r_{1}}(H)>|V(R^{m(n)-1})|\geq r^{\frac{\log n}{\log(\alpha+\varepsilon)}-O_{\varepsilon}(1)}=\Omega_{\varepsilon}\left(n^{\frac{\log r}{\log(\alpha+\varepsilon)}}\right),

which proves the lemma. ∎

To finish the proof of Theorem 1.2, it remains to bound the asymptotic growth rate of LR,f(s)L_{R,f}(s), 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 (R,f,s)(R,f,s)-walks.

Lemma 2.5.

If RR is a tournament without a transitive tt-subtournament, then any sequence v1,,vtv_{1},\ldots,v_{t} of tt vertices either contains a back-edge vivjv_{i}\leftarrow v_{j} with i<ji<j or two consecutive elements vi=vi+1v_{i}=v_{i+1}.

Proof.

If the vertices are all distinct, then since RR has no transitive tt-subtournament there must exist a back-edge. Suppose vi=vjv_{i}=v_{j} and i<ji<j but ji+1j\neq i+1. Then either vivi+1v_{i}\leftarrow v_{i+1} or vi+1vjv_{i+1}\leftarrow v_{j} is a back-edge. ∎

We now prove a recursive upper bound on LR,f(s)L_{R,f}(s). Given an implicit parameter t3t\geq 3, tournament RR, and nondecreasing function f:>0f:\mathbb{N}\rightarrow\mathbb{R}_{>0}, we say that a positive integer ss is short if LR,f(s)2stL_{R,f}(s)\leq 2st and LR,f(s′′)f(s′′)L_{R,f}(s^{\prime\prime})\leq f(s^{\prime\prime}) for all s′′ss^{\prime\prime}\leq s.

Lemma 2.6.

Suppose s1s\geq 1, t3t\geq 3, RR is a tournament without a transitive tt-subtournament, and f:>0f:\mathbb{N}\rightarrow\mathbb{R}_{>0} is a nondecreasing function satisfying f(s)>6st2f(s)>6st^{2}. If ss is short, then every s[2st,4st]s^{\prime}\in[2st,4st] is short.

Proof.

Suppose s<s4sts<s^{\prime}\leq 4st, {(vi,ai)}i=1\{(v_{i},a_{i})\}_{i=1}^{\ell} is an (R,f,s)(R,f,s^{\prime})-walk, and let i1<<iui_{1}<\dotsb<i_{u} be the sequence of all indices where aijsa_{i_{j}}\geq s.

Our first goal is to show that u<tu<t. If not, by Lemma 2.5 there is either a back-edge vixviyv_{i_{x}}\leftarrow v_{i_{y}} with x<ytx<y\leq t or some xt1x\leq t-1 for which vix=vix+1v_{i_{x}}=v_{i_{x+1}}. We show that neither of these situations is possible.

In the first case, there is an edge vixviyv_{i_{x}}\leftarrow v_{i_{y}} with x<ytx<y\leq t. By the definition of an (R,f,s)(R,f,s^{\prime})-walk,

a(ix,iy)>f(min(aix,aiy))f(s).a_{(i_{x},i_{y})}>f(\min(a_{i_{x}},a_{i_{y}}))\geq f(s). (2.3)

On the other hand,

a(ix,iy)=j(x,y)aij+j[x,y)a(ij,ij+1)st+tLR,f(s)6st2,a_{(i_{x},i_{y})}=\sum_{j\in(x,y)}a_{i_{j}}+\sum_{j\in[x,y)}a_{(i_{j},i_{j+1})}\leq s^{\prime}t+t\cdot L_{R,f}(s)\leq 6st^{2},

since for each j=x,,y1j=x,\ldots,y-1, the subsequence {(vi,ai)}i=ij+1ij+11\{(v_{i},a_{i})\}_{i=i_{j}+1}^{i_{j+1}-1} is an (R,f,s)(R,f,s)-walk with length at most LR,f(s)2stL_{R,f}(s)\leq 2st. But f(s)>6st2f(s)>6st^{2}, so this contradicts (2.3) and the back-edge vixviyv_{i_{x}}\leftarrow v_{i_{y}} cannot exist.

Next, suppose for some xt1x\leq t-1 that vix=vix+1v_{i_{x}}=v_{i_{x+1}}. The subsequence {(vi,ai)}i=ix+1ix+11\{(v_{i},a_{i})\}_{i=i_{x}+1}^{i_{x+1}-1} is an (R,f,s′′)(R,f,s^{\prime\prime})-walk where s′′s^{\prime\prime} is the maximum value of aia_{i} in this subsequence. Pick some zz for which az=s′′a_{z}=s^{\prime\prime}. Either vixvzv_{i_{x}}\leftarrow v_{z} or vzvix+1v_{z}\leftarrow v_{i_{x+1}} is a back-edge, and without loss of generality assume it is the former. Applying the definition of the (R,f,s′′)(R,f,s^{\prime\prime})-walk on the two indices ixi_{x} and zz,

a(ix,z)>f(min(aix,az))=f(s′′).a_{(i_{x},z)}>f(\min(a_{i_{x}},a_{z}))=f(s^{\prime\prime}).

On the other hand, this sum is bounded above by a(ix,ix+1)a_{(i_{x},i_{x+1})}. We know a(ix,ix+1)LR,f(s′′)f(s′′)a_{(i_{x},i_{x+1})}\leq L_{R,f}(s^{\prime\prime})\leq f(s^{\prime\prime}) because {(vi,ai)}i=ix+1ix+11\{(v_{i},a_{i})\}_{i=i_{x}+1}^{i_{x+1}-1} is an (R,f,s′′)(R,f,s^{\prime\prime})-walk and ss′′s\geq s^{\prime\prime} is short, so we have another contradiction. Thus u<tu<t.

We obtain

a[1,]j=1uaij+j=0ua(ij,ij+1)st+tLR,f(s)st+2st2,a_{[1,\ell]}\leq\sum_{j=1}^{u}a_{i_{j}}+\sum_{j=0}^{u}a_{(i_{j},i_{j+1})}\leq s^{\prime}t+t\cdot L_{R,f}(s)\leq s^{\prime}t+2st^{2}, (2.4)

for all s<s4sts<s^{\prime}\leq 4st and any (R,f,s)(R,f,s^{\prime})-walk {(vi,ai)}i=1\{(v_{i},a_{i})\}_{i=1}^{\ell}. Here we let i0=0i_{0}=0 and iu+1=+1i_{u+1}=\ell+1 for convenience.

Inequality (2.4) implies that LR,f(s)2stL_{R,f}(s^{\prime})\leq 2s^{\prime}t for all s[2st,4st]s^{\prime}\in[2st,4st]. It also implies that LR,f(s′′)6st2<f(s)f(s′′)L_{R,f}(s^{\prime\prime})\leq 6st^{2}<f(s)\leq f(s^{\prime\prime}) for all s′′(s,4st]s^{\prime\prime}\in(s,4st]. We have verified both conditions for ss^{\prime} to be short for every s[2st,4st]s^{\prime}\in[2st,4st], as desired. ∎

It remains to pick a function ff for which Lemma 2.6 bootstraps successfully. Define

f(s){10s2t3/2logtlog2ss440t3/2logts<4,f(s)\coloneqq\begin{cases}\frac{10s^{2}t^{3/2}\log t}{\log^{2}s}&s\geq 4\\ 40t^{3/2}\log t&s<4,\end{cases} (2.5)

where the values of f(1),f(2),f(3)f(1),f(2),f(3) are chosen just to make ff nondecreasing. Recall that all logarithms are to base 22.

Lemma 2.7.

If t106t\geq 10^{6}, f:>0f:\mathbb{N}\rightarrow\mathbb{R}_{>0} is defined by (2.5), and RR is a tournament without a transitive tt-subtournament, then LR,f(s)2stL_{R,f}(s)\leq 2st for all ss sufficiently large.

Proof.

We apply Lemma 2.6 inductively to show that ss is short for all ss sufficiently large. This implies the desired result.

For the base case, we claim that ss is short for all ss040t1/2logts\leq s_{0}\coloneqq 40t^{1/2}\log t. Indeed, suppose ss0s\leq s_{0} and {(vi,ai)}i=1\{(v_{i},a_{i})\}_{i=1}^{\ell} is an (R,f,s)(R,f,s)-walk. By Lemma 2.5, if t\ell\geq t then there is either a back-edge vivjv_{i}\leftarrow v_{j} with i<jti<j\leq t or two consecutive elements vi=vi+1v_{i}=v_{i+1}. The latter contradicts the definition of an (R,f,s)(R,f,s)-walk, so assume the former holds. But a(i,j)<stf(1)f(min(ai,aj))a_{(i,j)}<st\leq f(1)\leq f(\min(a_{i},a_{j})), so this contradicts the definition of an (R,f,s)(R,f,s)-walk again. We have shown that <t\ell<t, so LR,f(s)<stf(s)L_{R,f}(s)<st\leq f(s) for all ss0s\leq s_{0}. This proves ss is short for every ss0s\leq s_{0}.

Next, we check that f(s)>6st2f(s)>6st^{2} for all ss0s\geq s_{0}. Indeed, f(s)/sf(s)/s is increasing for ss0s\geq s_{0}, and logs0logt\log s_{0}\leq\log t since t106t\geq 10^{6}. We get

f(s0)=10s02t3/2logtlog2s0>1000t5/2logt>6s0t2,f(s_{0})=\frac{10s_{0}^{2}t^{3/2}\log t}{\log^{2}s_{0}}>1000t^{5/2}\log t>6s_{0}t^{2},

proving that the conditions of Lemma 2.6 are satisfied. By induction, Lemma 2.6 then implies that ss is short for s[(2t)ks0,(4t)ks0]s\in[(2t)^{k}s_{0},(4t)^{k}s_{0}] for all k0k\geq 0. All sufficiently large integers lie in some such interval, so LR,f(s)2stL_{R,f}(s)\leq 2st for all sufficiently large ss, as desired. ∎

Putting everything together, we have a proof of the general lower bound.

Proof of Theorem 1.2..

We may assume that Δ\Delta is sufficiently large as we always have r1(Hn)n\overrightarrow{r_{1}}(H_{n})\geq n, which proves the theorem for small Δ\Delta by picking the implicit constant factor in the exponent appropriately. Let t=Δ2/3/(200log2/3Δ)t=\Delta^{2/3}/(200\log^{2/3}\Delta); we may assume t106t\geq 10^{6}. Define f:>0f:\mathbb{N}\rightarrow\mathbb{R}_{>0} by (2.5). We have

S=m0f(2m+2)22mf(1)m1m280t3/2logt.S=\sum_{m\geq 0}f(2^{m+2})\cdot 2^{-2m}\leq f(1)\cdot\sum_{m\geq 1}m^{-2}\leq 80t^{3/2}\log t.

Lemma 2.2 implies that there exists an ff-interval mesh HH on \mathbb{N} with maximum degree at most 2S+17200t3/2logtΔ2S+17\leq 200t^{3/2}\log t\leq\Delta. For any n1n\geq 1, let HnH_{n} be the induced subgraph of HH on the interval [n][n], so that HnH_{n} has nn vertices and is also an ff-interval mesh of maximum degree at most Δ\Delta.

There exists a tournament RR on r=2Ω(t)r=2^{\Omega(t)} vertices with no transitive tt-subtournament. By Lemmas 2.4 and 2.7 applied with these choices of R,fR,f and HnH_{n}, we find that

r1(Hn)>nlogrlog(2t)o(1)nΩ(t/logt),\overrightarrow{r_{1}}(H_{n})>n^{\frac{\log r}{\log(2t)}-o(1)}\geq n^{\Omega(t/\log t)},

which proves the theorem, by our choice of tt. ∎

We remark that the polylogarithmic factor in Δ=Ω(t2/3/log2/3Δ)\Delta=\Omega(t^{2/3}/\log^{2/3}\Delta) can be easily improved. Indeed, the growth rate of f(s)=Θ(s2/log2s)f(s)=\Theta(s^{2}/\log^{2}s) is chosen so that

m0f(2m+2)22m<,\sum_{m\geq 0}f(2^{m+2})\cdot 2^{-2m}<\infty,

and we may take f(s)=Θ(s2/log1+εs)f(s)=\Theta(s^{2}/\log^{1+\varepsilon}s) for any fixed ε>0\varepsilon>0 instead, leading to a slightly smaller Δ\Delta.

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 HH be an acyclic digraph on nn vertices v1,,vnv_{1},\ldots,v_{n}. We would like to find an embedding ϕ:HT\phi:H\hookrightarrow T into an ambient tournament TT. In addition we specify nn sets U1,,UnV(T)U_{1},\ldots,U_{n}\subseteq V(T) and aim to satisfy ϕ(vi)Ui\phi(v_{i})\in U_{i} for all ii. Embedding then proceeds in nn rounds, where round tt determines the image ϕ(vt)\phi(v_{t}). After round tt, we keep track of the shrinking sets U1(t),,Un(t)U_{1}^{(t)},\ldots,U_{n}^{(t)} of “valid candidates” for each vertex, where initially Ui(0)=UiU_{i}^{(0)}=U_{i} for all ii. On round tt, ϕ(vt)\phi(v_{t}) is a carefully chosen vertex in Ut(t1)U_{t}^{(t-1)}, and Ut(t){ϕ(vt)}U_{t}^{(t)}\coloneqq\{\phi(v_{t})\}. The other candidate sets are updated as follows:

Uj(t){[Uj(t1)N+(vt)]{ϕ(vt)}if vtvj,[Uj(t1)N(vt)]{ϕ(vt)}if vjvt,Uj(t1){ϕ(vt)}else.U_{j}^{(t)}\coloneqq\begin{cases}[U_{j}^{(t-1)}\cap N^{+}(v_{t})]\setminus\{\phi(v_{t})\}&\text{if }v_{t}\rightarrow v_{j},\\ {}[U_{j}^{(t-1)}\cap N^{-}(v_{t})]\setminus\{\phi(v_{t})\}&\text{if }v_{j}\rightarrow v_{t},\\ U_{j}^{(t-1)}\setminus\{\phi(v_{t})\}&\text{else}.\end{cases}

The process fails if there is an empty Ut(t1)U_{t}^{(t-1)}, as there would be no valid choice for ϕ(vt)\phi(v_{t}). Otherwise, it succeeds if ϕ(vt)\phi(v_{t}) is chosen successfully for each tnt\leq n. Note that after round tt, {ϕ(v1),,ϕ(vt)}\{\phi(v_{1}),\ldots,\phi(v_{t})\} is an embedded copy of H[{v1,,vt}]H[\{v_{1},\ldots,v_{t}\}] in TT, and these vertices have been removed from the other candidate sets, so the update rule guarantees that Uj(t)={ϕ(vj)}U_{j}^{(t)}=\{\phi(v_{j})\} remains a singleton when tjt\geq j. If the greedy embedding process succeeds, it exhibits the existence of a copy of HH in TT. See Figure 3.1 for a schematic illustration of the greedy embedding process.

Target graph HHv1v_{1}v2v_{2}v3v_{3}v4v_{4}Step 0U1(0)U_{1}^{(0)}U2(0)U_{2}^{(0)}U3(0)U_{3}^{(0)}U4(0)U_{4}^{(0)}Step 1ϕ(v1)\phi(v_{1})U2(1)U_{2}^{(1)}U3(1)U_{3}^{(1)}U4(1)U_{4}^{(1)}Step 2ϕ(v1)\phi(v_{1})ϕ(v2)\phi(v_{2})U3(2)U_{3}^{(2)}U4(2)U_{4}^{(2)}Step 3ϕ(v1)\phi(v_{1})ϕ(v2)\phi(v_{2})ϕ(v3)\phi(v_{3})U4(3)U_{4}^{(3)}Step 4ϕ(v1)\phi(v_{1})ϕ(v2)\phi(v_{2})ϕ(v3)\phi(v_{3})ϕ(v4)\phi(v_{4})
Figure 3.1: Illustration of the greedy embedding process for an acyclic orientation of the four-cycle. A directed arrow from a vertex to a set indicates that the vertex is complete to the set. All edges point forward in this example, but we do not always make this assumption.

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 TT is HH-free, then TT contains a large “approximate blowup of HH.” In the inner stage, we use greedy embedding one final time within this “approximate blowup” to guarantee the existence of a copy of HH in TT. In either case, we can conclude that TT contains a copy of HH—either it is found directly by the greedy embedding strategy, or else the failure of the greedy embedding yields the “approximate blowup” of HH, in which a copy of HH 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 HH in a tournament TT implies that TT 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 HH” to finally embed HH itself.

In some of these greedy embedding arguments, we are concerned with partitioning an acyclic digraph HH into a number of parts and embedding the parts one at a time, so the following definition will be useful.

Definition 3.1.

If HH is an acyclic digraph, we say that a collection {Pi}i=1r\{P_{i}\}_{i=1}^{r} of vertex subsets of HH is a directed partition of HH if P1P2Pr=V(H)P_{1}\sqcup P_{2}\sqcup\cdots\sqcup P_{r}=V(H), and any edge of HH between two distinct parts Pi,PjP_{i},P_{j} with i<ji<j is oriented from PiP_{i} to PjP_{j}.

In particular, the height of HH is exactly the least hh for which there exists a directed partition of HH into hh independent sets.

3.1 The basic greedy embedding building blocks

Recall that an undirected graph GG is said to be dd-degenerate if there exists an ordering v1,,vnv_{1},\ldots,v_{n} of the vertices of GG such that each |N(vi){v1,,vi1}|d|N(v_{i})\cap\{v_{1},\ldots,v_{i-1}\}|\leq d, and such an ordering is called a dd-degenerate ordering of GG. A dd-degenerate ordering is a natural order for greedily embedding an undirected graph GG, since each candidate set Uj(t)U_{j}^{(t)} in the greedy embedding strategy only shrinks in size by more than one at most dd times. We say that a digraph HH is dd-degenerate if its underlying undirected graph is dd-degenerate. Note that if HH has maximum degree Δ\Delta then it is Δ\Delta-degenerate, but dd-degenerate digraphs can have arbitrarily large maximum degree.

Define a δ\delta-dense pair (W1,W2)(W_{1},W_{2}) in a tournament TT to be a pair of vertex subsets such that at least δ|W1||W2|\delta|W_{1}||W_{2}| of the edges between W1W_{1} and W2W_{2} point from W1W_{1} to W2W_{2}. The size of the pair is defined to be min(|W1|,|W2|)\min(|W_{1}|,|W_{2}|). We do not require W1W_{1} and W2W_{2} to be disjoint, although the assumption of δ\delta-density implies that W1W2W_{1}\cap W_{2} cannot be too large if δ\delta is close to 11.

The first lemma in this subsection uses greedy embedding to show that if TT doesn’t contain a copy of a given dd-degenerate HH, then TT 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 HH be a dd-degenerate digraph with nn vertices and maximum degree Δ\Delta, and let 0<c120<c\leq\frac{1}{2}. If TT is an HH-free tournament on N2ΔcdnN\geq 2\Delta c^{-d}n vertices, then TT contains a (1c)(1-c)-dense pair (W1,W2)(W_{1},W_{2}) with size at least cdN/(2Δ)c^{d}N/(2\Delta).

Proof.

We use the greedy embedding framework described above. Let us label the vertices of HH according to the dd-degenerate ordering as v1,v2,,vn.v_{1},v_{2},\dotsc,v_{n}. We initialize Ui(0)=V(T)U_{i}^{(0)}=V(T) for all 1in1\leq i\leq n. We now attempt to embed the vertices of HH one at a time in TT, in the dd-degenerate ordering. For j>tj>t, let Nt(vj)N_{t}(v_{j}) denote the set of vertices viv_{i} with iti\leq t such that viv_{i} and vjv_{j} are connected by an edge (in some direction). We inductively pick the values of ϕ(vt)V(T)\phi(v_{t})\in V(T) and maintain vertex sets Ui(t)U_{i}^{(t)} with the following properties.

  1. 1.

    For every iti\leq t, we have Ui(t)={ϕ(vi)}U_{i}^{(t)}=\{\phi(v_{i})\}.

  2. 2.

    For every j>tj>t, we have |Uj(t)|c|Nt(vj)|Nt|U_{j}^{(t)}|\geq c^{|N_{t}(v_{j})|}N-t.

  3. 3.

    For iti\leq t, if vivjv_{i}\to v_{j} then ϕ(vi)x\phi(v_{i})\to x for every xUj(t)x\in U_{j}^{(t)}, and if vivjv_{i}\leftarrow v_{j} then ϕ(vi)x\phi(v_{i})\leftarrow x for every xUj(t)x\in U_{j}^{(t)}.

From these properties, we see that if the process continues through step t=nt=n, then we will have embedded a copy of HH in TT, contradicting our assumption that TT is HH-free. Moreover, all three properties are vacuously true for t=0t=0. Now, suppose we have maintained this process up through step t1t-1. Suppose there exists wtUt(t1)w_{t}\in U_{t}^{(t-1)} such that for every j>tj>t with vtvjv_{t}\to v_{j} (resp. vtvjv_{t}\leftarrow v_{j}), we have |N+(wt)Uj(t1)|c|Uj(t1)||N^{+}(w_{t})\cap U_{j}^{(t-1)}|\geq c|U_{j}^{(t-1)}| (resp. |N(wt)Uj(t1)|c|Uj(t1)||N^{-}(w_{t})\cap U_{j}^{(t-1)}|\geq c|U_{j}^{(t-1)}|). Then we may define ϕ(vt)=wt\phi(v_{t})=w_{t}, Ut(t)={ϕ(vt)}U_{t}^{(t)}=\{\phi(v_{t})\}, and update the remaining sets as

Uj(t){[Uj(t1)N+(vt)]{ϕ(vt)}if vtvj,[Uj(t1)N(vt)]{ϕ(vt)}if vjvt,Uj(t1){ϕ(vt)}else,U_{j}^{(t)}\coloneqq\begin{cases}[U_{j}^{(t-1)}\cap N^{+}(v_{t})]\setminus\{\phi(v_{t})\}&\text{if }v_{t}\rightarrow v_{j},\\ {}[U_{j}^{(t-1)}\cap N^{-}(v_{t})]\setminus\{\phi(v_{t})\}&\text{if }v_{j}\rightarrow v_{t},\\ U_{j}^{(t-1)}\setminus\{\phi(v_{t})\}&\text{else},\end{cases}

for all j>tj>t. Properties 1 and 3 above continue to hold automatically after round tt, and all that remains to check is Property 2. For those j>tj>t for which vjv_{j} is not adjacent to vtv_{t}, Nt(vj)=Nt1(vj)N_{t}(v_{j})=N_{t-1}(v_{j}) and at most one vertex is removed from Uj(t1)U_{j}^{(t-1)} to obtain Uj(t)U_{j}^{(t)}. Therefore,

|Uj(t)||Uj(t1)|1c|Nt1(vj)|N(t1)1=c|Nt(vj)|t.|U_{j}^{(t)}|\geq|U_{j}^{(t-1)}|-1\geq c^{|N_{t-1}(v_{j})|}N-(t-1)-1=c^{|N_{t}(v_{j})|}-t.

On the other hand, if vjv_{j} is adjacent to vtv_{t}, then |Nt(vj)|=|Nt1(vj)|+1|N_{t}(v_{j})|=|N_{t-1}(v_{j})|+1, and therefore

|Uj(t)|c|Uj(t1)|1c(c|Nt1(vj)|N(t1))1c|Nt(vj)|Nt.|U_{j}^{(t)}|\geq c|U_{j}^{(t-1)}|-1\geq c\cdot(c^{|N_{t-1}(v_{j})|}N-(t-1))-1\geq c^{|N_{t}(v_{j})|}N-t.

Thus, all three properties are maintained if such a wtw_{t} exists.

Since we assumed that TT was HH-free, this process cannot continue until step t=nt=n, and therefore must fail at some step 1tn11\leq t\leq n-1. Let W0=Ut(t1)W_{0}=U_{t}^{(t-1)}. Since the process fails at this step, for every wW0w\in W_{0}, we can assign some j>tj>t such that either vtvjv_{t}\to v_{j} and |N+(w)Uj(t1)|<c|Uj(t1)||N^{+}(w)\cap U_{j}^{(t-1)}|<c|U_{j}^{(t-1)}|, or vtvjv_{t}\leftarrow v_{j} and |N(w)Uj(t1)|<c|Uj(t1)||N^{-}(w)\cap U_{j}^{(t-1)}|<c|U_{j}^{(t-1)}|. Since vtv_{t} has at most Δ\Delta neighbors in total, at least |W0|/Δ|W_{0}|/\Delta choices of ww are assigned the same j>tj>t by the pigeonhole principle. Fix such a “popular” jj.

Suppose first that vtvjv_{t}\to v_{j}. Let W2=Uj(t1)W_{2}=U_{j}^{(t-1)}, and let W1W_{1} be the set of all wW0w\in W_{0} which have fewer than c|W2|c|W_{2}| out-neighbors in W2W_{2}. Then (W1,W2)(W_{1},W_{2}) is a (1c)(1-c)-dense pair. Similarly, if vjN(vt)v_{j}\in N^{-}(v_{t}), then we would similarly find that (W2,W1)(W_{2},W_{1}) is a (1c)(1-c)-dense pair.

It remains to verify the lower bound on the sizes of W1W_{1} and W2W_{2}. Recall that the greedy embedding process succeeded up through step t1t-1, meaning that

|Ut(t1)|c|Nt1(vt)|N(t1)cdNncdN2,|U_{t}^{(t-1)}|\geq c^{|N_{t-1}(v_{t})|}N-(t-1)\geq c^{d}N-n\geq\frac{c^{d}N}{2},

and similarly for Uj(t1)U_{j}^{(t-1)}, where we use the dd-degeneracy assumption to conclude that |Nt1(vt)|d|N_{t-1}(v_{t})|\leq d, and our assumption that t<ncdN/2t<n\leq c^{d}N/2. Since |W2|=|Uj(t1)||W_{2}|=|U_{j}^{(t-1)}| and |W1||Ut(t1)|/Δ|W_{1}|\geq|U_{t}^{(t-1)}|/\Delta, this completes the proof. ∎

The second lemma proves a much stronger bound when HH is 11-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 11-dense pair in a tournament is just a pair of sets W1,W2W_{1},W_{2} with all edges directed from W1W_{1} to W2W_{2}.

Lemma 3.3.

Let HH be a weakly connected 11-degenerate digraph with maximum degree Δ\Delta on mm vertices v1,,vmv_{1},\ldots,v_{m}, and let TT be an arbitrary tournament. If there exist sets U1,,UmV(T)U_{1},\ldots,U_{m}\subseteq V(T), each of size M2mΔM\geq 2m\Delta such that if there is no embedding ϕ:HT\phi:H\hookrightarrow T satisfying ϕ(vi)Ui\phi(v_{i})\in U_{i}, then TT contains a 11-dense pair with size at least M/(m(Δ+1))M/(m(\Delta+1)).

Proof.

We begin by picking subsets V1U1,,VmUmV_{1}\subseteq U_{1},\dotsc,V_{m}\subseteq U_{m}, each of size M/mM/m, such that V1,,VmV_{1},\dotsc,V_{m} are pairwise disjoint. We can do this greedily, by first picking an arbitrary subset V1U1V_{1}\subseteq U_{1} of size M/mM/m, then picking an arbitrary subset V2U2V1V_{2}\subseteq U_{2}\setminus V_{1} of size M/mM/m, then picking V3U3(V1V2)V_{3}\subseteq U_{3}\setminus(V_{1}\cup V_{2}), and so on. At the iith step, we have deleted at most (i1)M/m(m1)M/m(i-1)M/m\leq(m-1)M/m elements from UiU_{i}, so at least M/mM/m elements remain, from which we pick ViV_{i} 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 HH be ordered so that each viv_{i} has at most one neighbor vjv_{j} with j>ij>i (this is the reverse of the usual degenerate ordering). For each tt, let StS_{t} denote the set of vertices viv_{i} with iti\leq t that are connected to viv_{i} by a path of vertices whose indices are monotonically increasing, including vtv_{t} itself. In other words, we set S1={v1}S_{1}=\{v_{1}\}, let Nt(vt)N_{t}(v_{t}) denote the set of vertices viv_{i} adjacent to vtv_{t} (in either orientation) with i<ti<t, and set

St={vt}iNt(vt)Si.S_{t}=\{v_{t}\}\cup\bigcup_{i\in N_{t}(v_{t})}S_{i}.

We define WtVtW_{t}\subseteq V_{t} to be the set of all wVtw\in V_{t} for which there exists an embedding ϕ:H[St]T\phi:H[S_{t}]\hookrightarrow T mapping each viv_{i} into ViV_{i} for viStv_{i}\in S_{t}, and mapping vtv_{t} to ww. We know Wn=W_{n}=\varnothing, since otherwise there is an embedding of HH into the sets V1,,VmV_{1},\dotsc,V_{m}, contradicting our assumption. Let tnt\leq n be the minimum index such that |Wt|<M/(m(Δ+1))|W_{t}|<M/(m(\Delta+1)); note that t>1t>1 since W1=V1W_{1}=V_{1} has size M/mM/m. Let Nt(vt)={vi1,,vis}N_{t}(v_{t})=\{v_{i_{1}},\dotsc,v_{i_{s}}\}. Our maximum degree assumption implies sΔs\leq\Delta, while if s=0s=0, then Wt=VtW_{t}=V_{t}, contradicting our assumption that |Wt|<M/(m(Δ+1))|W_{t}|<M/(m(\Delta+1)). Thus, 1sΔ1\leq s\leq\Delta.

By definition, WtW_{t} is precisely the set of wVtw\in V_{t} adjacent in the appropriate orientation to at least one vertex in each of Wi1,,WisW_{i_{1}},\dotsc,W_{i_{s}}. Let X1,,XsVtX_{1},\dotsc,X_{s}\subseteq V_{t} denote the choices of ww which do not have any edge in the appropriate orientation to Wi1,,WisW_{i_{1}},\dotsc,W_{i_{s}}, respectively. We get Vt=WtX1XsV_{t}=W_{t}\cup X_{1}\cup\dotsb\cup X_{s}. Since sΔs\leq\Delta and |Wt|M/(m(Δ+1))|W_{t}|\leq M/(m(\Delta+1)), we see that there exists some jj for which |Xj|M/(m(Δ+1))|X_{j}|\geq M/(m(\Delta+1)). Moreover, since tt was taken to be minimal, we have that |Wij|M/(m(Δ+1))|W_{i_{j}}|\geq M/(m(\Delta+1)) as well. This yields a pair of sets (Xj,Wij)(X_{j},W_{i_{j}}) where all edges between them are oriented the same way, and both sets have size at least M/(m(Δ+1))M/(m(\Delta+1)). This is the desired 11-dense pair. ∎

We remark that we expect the mm 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 11-dense pair in TT whenever TT does not contain a copy of some fixed oriented forest with small maximum degree and small weakly connected components.

Lemma 3.4.

Let HH be a 11-degenerate digraph with maximum degree Δ\Delta and vertices v1,,vnv_{1},\dotsc,v_{n}, and suppose that every weakly connected component of HH has at most mm vertices. Let TT be an arbitrary tournament. For any collection of sets V1,,VnV(T)V_{1},\ldots,V_{n}\subseteq V(T), each of size M3nΔM\geq 3n\Delta, either there is an embedding ϕ:HT\phi:H\hookrightarrow T satisfying ϕ(vi)Vi\phi(v_{i})\in V_{i}, or TT contains a 11-dense pair with size at least M/(4mΔ)M/(4m\Delta).

Proof.

Let the weakly connected components of HH be C1,,CrC_{1},\dotsc,C_{r}. We prove the result by induction on rr. The base case, r=1r=1, follows immediately from Lemma 3.3, since nmn\geq m and Δ+12Δ\Delta+1\leq 2\Delta. For the inductive step, let HH^{\prime} be the induced subgraph of HH consisting of the weakly connected components C1,,Cr1C_{1},\dotsc,C_{r-1}. By the inductive hypothesis, either TT contains a 11-dense pair of size at least M/(4mΔ)M/(4m\Delta), in which case we are done, or else there is an embedding ϕ\phi of HH^{\prime} into TT satisfying that ϕ(vi)Vi\phi(v_{i})\in V_{i} for all viV(H)v_{i}\in V(H^{\prime}). Let TT^{\prime} be the tournament obtained by deleting the image of ϕ\phi from TT, and similarly let V1,,VnV_{1}^{\prime},\dotsc,V_{n}^{\prime} be obtained by deleting the image of ϕ\phi from V1,,VnV_{1},\dotsc,V_{n}. Since we have deleted at most nnΔn\leq n\Delta vertices, each ViV_{i}^{\prime} has size at least M2nΔ2mΔM^{\prime}\geq 2n\Delta\geq 2m\Delta. Therefore, by Lemma 3.3, either TT^{\prime} contains a 11-dense pair of size at least M/(2mΔ)M/(4mΔ)M^{\prime}/(2m\Delta)\geq M/(4m\Delta), in which case so does TT, or else there is an embedding of H[Cr]H[C_{r}] into TT^{\prime} mapping each viCrv_{i}\in C_{r} into ViV_{i}^{\prime}. Since we deleted the image of ϕ\phi from TT to define TT^{\prime}, such an embedding yields an embedding of HH into TT, 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 HH, 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 HH be an acyclic digraph with maximum degree Δ\Delta on nn vertices v1,,vnv_{1},\ldots,v_{n}, and let TT be a tournament containing subsets V1,,VnV(T)V_{1},\ldots,V_{n}\subseteq V(T), each of size at least 4n4n. If for every edge vivjv_{i}\rightarrow v_{j} in HH, (Vi,Vj)(V_{i},V_{j}) is a (118Δ2)(1-\frac{1}{8\Delta^{2}})-dense pair, then TT contains a copy of HH.

Note that the sets ViV_{i} are allowed to overlap222However, if vivjv_{i}\to v_{j}, then the assumption that (Vi,Vj)(V_{i},V_{j}) is (118Δ2)(1-\frac{1}{8\Delta^{2}})-dense implies that ViVjV_{i}\cap V_{j} cannot be too large. or even be identical. This is crucial; for instance, if HH has height hh then our choice of ViV_{i}’s will take only hh 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 vivjv_{i}\rightarrow v_{j} satisfy i<ji<j. We now run the greedy embedding process for HH into TT with the given UiU_{i}, 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 V1,,VnV_{1},\dotsc,V_{n}. Namely, for every edge vivjv_{i}\to v_{j} in HH, let Vi,jViV_{i,j}\subseteq V_{i} be the set of vertices wViw\in V_{i} with |N(w)Vj||Vj|/(4Δ)|N^{-}(w)\cap V_{j}|\geq|V_{j}|/(4\Delta). Since the pair (Vi,,Vj)(V_{i,},V_{j}) is (118Δ2)(1-\frac{1}{8\Delta^{2}})-dense, there are at most |Vi||Vj|/(8Δ2)|V_{i}||V_{j}|/(8\Delta^{2}) edges directed from VjV_{j} to ViV_{i} in total, so

|Vi,j||Vj|4Δ|Vi||Vj|8Δ2,|V_{i,j}|\cdot\frac{|V_{j}|}{4\Delta}\leq\frac{|V_{i}||V_{j}|}{8\Delta^{2}},

and thus |Vi,j||Vi|/(2Δ)|V_{i,j}|\leq|V_{i}|/(2\Delta). Each viv_{i} has at most Δ\Delta out-neighbors vjv_{j}, which implies that |jVi,j||Vi|/2|\bigcup_{j}V_{i,j}|\leq|V_{i}|/2. Therefore, if UiVi(jVi,j)U_{i}\coloneqq V_{i}\setminus(\bigcup_{j}V_{i,j}), then |Ui||Vi|/2|U_{i}|\geq|V_{i}|/2, and every vertex in UiU_{i} has at most |Vj|/(4Δ)|Uj|/(2Δ)|V_{j}|/(4\Delta)\leq|U_{j}|/(2\Delta) in-neighbors in UjU_{j} for every jj such that vivjv_{i}\to v_{j}.

Now, we exhibit an embedding ϕ:HT\phi:H\hookrightarrow T by picking ϕ(vt)Ut\phi(v_{t})\in U_{t} inductively for each t[n]t\in[n], as follows. Having picked ϕ(v1),,ϕ(vt1)\phi(v_{1}),\ldots,\phi(v_{t-1}), we claim that there is at least one valid candidate for ϕ(vt)Ut\phi(v_{t})\in U_{t} which is consistent with the previous choices. Indeed, there are at most Δ\Delta choices of i<ti<t such that vivtv_{i}\to v_{t}. For every such ii, we have that ϕ(vi)Ui\phi(v_{i})\in U_{i}, which means that ϕ(vi)\phi(v_{i}) has at most |Ut|/(2Δ)|U_{t}|/(2\Delta) in-neighbors in UtU_{t}. Hence, at least |Ut|/2|U_{t}|/2 vertices in UtU_{t} are out-neighbors of ϕ(vi)\phi(v_{i}) for all ii such that vivjv_{i}\to v_{j}. Moreover, at most t1t-1 vertices of UjU_{j} have been picked as outputs of ϕ\phi, and thus the number of possible candidates for ϕ(vt)\phi(v_{t}) is at least

|Ut|2(t1)|Vt|4n+11,\frac{|U_{t}|}{2}-(t-1)\geq\frac{|V_{t}|}{4}-n+1\geq 1,

by our assumption that |Vj|4n|V_{j}|\geq 4n. This shows that at every step tnt\leq n we can pick vertex ϕ(vt)\phi(v_{t}) such that {ϕ(v1),,ϕ(vt)}\{\phi(v_{1}),\ldots,\phi(v_{t})\} is a copy of H[{v1,,vt}]H[\{v_{1},\ldots,v_{t}\}] in TT, and so TT contains a copy of the entirety of HH, as desired. ∎

We remark that one can prove a strengthening of Lemma 2.4, where the assumption is weakened to each pair (Vi,Vj)(V_{i},V_{j}) being merely (1Ω(1Δ))(1-\Omega(\frac{1}{\Delta}))-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 HH-free tournament TT, an “approximate blowup” of HH. This will be a large collection of vertex sets which correspond to the vertices of HH, such that the edges between sets are mostly oriented in the correct direction. Before stating the result precisely, we need some definitions.

Let {0,1}\{0,1\}^{*} denote the set of all finite binary strings. Recall that a prefix code is a set C{0,1}C\subset\{0,1\}^{*} with the property that no element of CC is a prefix of another element of CC. The depth of CC is defined as the maximum length of an element of CC. Let \prec denote the lexicographic ordering on {0,1}\{0,1\}^{*}, namely the ordering in which xyx\prec y if xx is a proper prefix of yy or if xi<yix_{i}<y_{i} where ii is the first index for which xiyix_{i}\neq y_{i}.

Definition 3.6.

Given an acyclic digraph HH, a prefix labeling of HH is a surjective map ρ:V(H)C\rho:V(H)\to C for some prefix code C{0,1}C\subset\{0,1\}^{*}, with the property that if vivjv_{i}\to v_{j} is an edge of HH, then either ρ(vi)=ρ(vj)\rho(v_{i})=\rho(v_{j}) or ρ(vi)ρ(vj)\rho(v_{i})\prec\rho(v_{j}). The map ρ\rho naturally defines a graph structure on CC, where we say that two codewords x,yx,y are adjacent under ρ\rho if there exists some edge between the sets ρ1(x)\rho^{-1}(x) and ρ1(y)\rho^{-1}(y). By the maximum degree of ρ\rho, we mean the maximum degree of this graph on CC. If ρ1(x)\rho^{-1}(x) is an independent set for every xCx\in C, then we call ρ\rho a prefix coloring. Less stringently, we call ρ\rho a forest prefix labeling if ρ1(x)\rho^{-1}(x) is a directed forest (or equivalently, a 11-degenerate digraph) for every xCx\in C. By the maximum component size of ρ\rho, we mean the maximum number of vertices of any weakly connected component in ρ1(x)\rho^{-1}(x), over all xCx\in C. Thus, ρ\rho is a prefix coloring if and only if its maximum component size is 11.

Thus, we see that prefix colorings of HH correspond to colorings of the underlying undirected graph of HH, with the property that the palette of colors CC is a prefix code, and that the lexicographic order on CC is consistent with the edge directions in HH. 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 HH and by the structure of the prefix code CC.

For a binary string xx, let us denote by x0x0 and x1x1 the strings obtained by appending a 0 or 11, respectively, to the end of xx. For a prefix code CC, we denote by C0(x)C_{0}(x) the set of all elements yCy\in C which have x0x0 as a prefix, and by C1(x)C_{1}(x) the set of elements that have x1x1 as a prefix. Suppose that ρ:V(H)C\rho:V(H)\to C is a prefix labeling of an acyclic digraph HH. For binary string xx, let us denote by a0(x)a_{0}(x) the number of codewords yC0(x)y\in C_{0}(x) which are adjacent under ρ\rho to some codeword zC1(x)z\in C_{1}(x). Similarly, a1(x)a_{1}(x) is the number of codewords zC1(x)z\in C_{1}(x) that are adjacent under ρ\rho to some codeword yC0(x)y\in C_{0}(x). Finally, we let

a(x)={1if a0(x)=0 or a1(x)=0,min(a0(x),a1(x))otherwise.a(x)=\begin{cases}1&\text{if }a_{0}(x)=0\text{ or }a_{1}(x)=0,\\ \min(a_{0}(x),a_{1}(x))&\text{otherwise.}\end{cases}

In particular, a(x)=1a(x)=1 if xx is not a proper prefix of any element of CC. With this notation, we can now define the key parameters of ρ\rho that we later use to bound Ramsey numbers.

Definition 3.7.

Let HH be an acyclic digraph and ρ:V(H)C\rho:V(H)\to C some prefix labeling of HH. We define the dyadic complexity of ρ\rho to be the quantity

comp(ρ):=maxyCx a prefix of ya(x).\operatorname{comp}(\rho):=\max_{y\in C}\prod_{x\text{ a prefix of }y}a(x).

Additionally, the depth of ρ\rho, denoted depth(ρ)\operatorname{depth}(\rho), is defined as the depth of CC, i.e. the maximum length of an element of CC.

To understand these definitions, it is helpful to think of {0,1}\{0,1\}^{*} as the vertices of the infinite binary tree. In this setup, the elements of a prefix code CC correspond to the leaves of some subtree. A prefix labeling ρ:V(H)C\rho:V(H)\to C is then a partition of the vertices of HH into sets labeled by the leaves of this subtree. Two codewords (leaves) are adjacent under ρ\rho if there is an edge between the corresponding vertex subsets of HH. For a binary string xx, which should be thought of as a non-leaf vertex of the subtree, a(x)a(x) roughly records the “cost” of separating the descendants of xx: it measures how many pairs of codewords adjacent under ρ\rho 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 ρ\rho 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 comp(ρ)\operatorname{comp}(\rho)—is the product of the costs of all ancestors of yy, maximized over all yCy\in C.

We remark that the dyadic complexity of a prefix coloring of HH is one possible formalization of the notion of multiscale complexity discussed in the introduction. Indeed, if every prefix coloring of HH has high dyadic complexity, then HH 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 HH, making formal the intuition that the strength of our upper bound results depends on whether HH has high or low multiscale complexity.

In order to embed some acyclic digraph HH in a tournament TT, we first build a certain structure of vertex subsets of TT 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 HH. The structure we build depends on a parameter δ[0,1]\delta\in[0,1], as well as on a prefix labeling ρ\rho of HH. We now define this structure, and next prove a lemma showing how to find such a structure.

Definition 3.8.

Let δ[0,1]\delta\in[0,1] be some parameter, let HH be an acyclic digraph, and let ρ:V(H)C\rho:V(H)\to C be some prefix labeling of HH, for some prefix code CC. Let TT be any tournament. A (ρ,δ)(\rho,\delta)-skeleton is a collection {Vx}xC\{V_{x}\}_{x\in C} of (not necessarily disjoint) vertex subsets of TT, indexed by the codewords in CC, with the property that if xyx\prec y are elements of CC that are adjacent under ρ\rho, then (Vx,Vy)(V_{x},V_{y}) is a δ\delta-dense pair. We define the size of a (ρ,δ)(\rho,\delta)-skeleton to be minxC|Vx|\min_{x\in C}|V_{x}|.

Our next lemma shows how to iterate Lemma 3.2 to construct a (ρ,δ)(\rho,\delta)-skeleton in any sufficiently large HH-free tournament TT. Roughly speaking, since the structure of a (ρ,δ)(\rho,\delta)-skeleton is based on the binary tree structure of CC, 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 c(0,1)c\in(0,1) be some parameter, let HH be a dd-degenerate acyclic digraph with maximum degree Δ\Delta, and let ρ:V(H)C\rho:V(H)\to C be some prefix labeling of HH, for some prefix code CC. Suppose that TT is an HH-free tournament on NN vertices, with

N(4d+1Δcd)depth(ρ)comp(ρ)dn.N\geq\left(\frac{4^{d+1}\Delta}{c^{d}}\right)^{\operatorname{depth}(\rho)}\operatorname{comp}(\rho)^{d}n.

Then TT contains a (ρ,1c)(\rho,1-c)-skeleton of size at least (cd4d+1Δ)depth(ρ)comp(ρ)dN\left(\frac{c^{d}}{4^{d+1}\Delta}\right)^{\operatorname{depth}(\rho)}\operatorname{comp}(\rho)^{-d}N.

Proof.

For every binary string x{0,1}x\in\{0,1\}^{*} which is a prefix of some codeword in CC, let HxH_{x} denote the subgraph of HH induced by the vertices vv for which xx is a prefix of ρ(v)\rho(v). We will construct, for every such x{0,1}x\in\{0,1\}^{*}, a vertex subset VxV(T)V_{x}\subseteq V(T), with

|Vx|(z a proper prefix of xcd4d+1Δa(z)d)N.|V_{x}|\geq\left(\prod_{z\text{ a proper prefix of }x}\frac{c^{d}}{4^{d+1}\Delta a(z)^{d}}\right)N.

We initialize V=V(T)V_{\varnothing}=V(T), where {0,1}\varnothing\in\{0,1\}^{*} denotes the empty string, and observe that this property holds vacuously for VV_{\varnothing} since \varnothing has no proper prefixes. In order to construct these vertex sets for other xx, we proceed via a depth-first search along the binary tree, as follows. Recall that for any binary string xx, the numbers a0(x)a_{0}(x) and a1(x)a_{1}(x) are the number of codewords in CC beginning with x0x0 and x1x1, respectively, such that some vertex labeled by that codeword has an edge to Hx1H_{x1} and Hx0H_{x0}, respectively. If min(a0(x),a1(x))=0\min(a_{0}(x),a_{1}(x))=0, then we define Vx0=Vx1=VxV_{x0}=V_{x1}=V_{x}, and observe that our desired inequality holds, since a(x)=1a(x)=1 in this case. In particular, if xCx\in C is a codeword, then we stop the inductive process, since we only wished to define such vertex subsets for strings xx that are the prefix of some codeword. Now, suppose that min(a0(x),a1(x))1\min(a_{0}(x),a_{1}(x))\geq 1, and assume without loss of generality333If a1(x)<a0(x)a_{1}(x)<a_{0}(x), then we swap the roles of 0 and 11 and the roles of in- and out-neighbors in this construction. that a0(x)a1(x)a_{0}(x)\leq a_{1}(x). We will first show how to define Vx0VxV_{x0}\subseteq V_{x} satisfying the desired inequality on its cardinality. We will then proceed to recursively define vertex subsets VyVx0V_{y}\subseteq V_{x0} for every binary string yy prefixed by x0x0. Note that we have not yet defined the set Vx1V_{x1}: we are proceeding in a depth-first fashion, so we will not define Vx1V_{x1} until we have defined VyV_{y} for every yy prefixed by x0x0. This will eventually happen, since we already described above how to define VyV_{y} if yy is a codeword of CC; therefore, we eventually reach the bottom of our depth-first search (namely a codeword yCy\in C), 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 a0(x)a1(x)a_{0}(x)\leq a_{1}(x), and let cx=c/(4a0(x)).c_{x}=c/(4a_{0}(x)). Since

|Vx|\displaystyle|V_{x}| (z a proper prefix of xcd4d+1Δa(z)d)N\displaystyle\geq\left(\prod_{z\text{ a proper prefix of }x}\frac{c^{d}}{4^{d+1}\Delta a(z)^{d}}\right)N
=4Δ(4a(x))dcd(z a proper prefix of x0cd4d+1Δa(z)d)N\displaystyle=\frac{4\Delta(4a(x))^{d}}{c^{d}}\left(\prod_{z\text{ a proper prefix of }x0}\frac{c^{d}}{4^{d+1}\Delta a(z)^{d}}\right)N
(cd/(4d+1Δ))depth(ρ)comp(ρ)dcd/(4Δ(4a(x))d)N\displaystyle\geq\frac{\left(c^{d}/(4^{d+1}\Delta)\right)^{\operatorname{depth}(\rho)}\operatorname{comp}(\rho)^{-d}}{c^{d}/(4\Delta(4a(x))^{d})}N
4Δcxdn,\displaystyle\geq 4\Delta c_{x}^{-d}n, (3.1)

and since TT contains no copy of HH, we may apply Lemma 3.2 with parameter cxc_{x}. Then this lemma says that VxV_{x} contains a (1cx)(1-c_{x})-dense pair (W0,W1)(W_{0},W_{1}), where min(|W0|,|W1|)cxd|Vx|/(2Δ)\min(|W_{0}|,|W_{1}|)\geq c_{x}^{d}|V_{x}|/(2\Delta). Let Vx0W0V_{x0}\subseteq W_{0} denote the set of vertices in W0W_{0} whose in-degree to W1W_{1} is at most 2cx|W1|2c_{x}|W_{1}|; since there are at most (1cx)|W0||W1|(1-c_{x})|W_{0}||W_{1}| edges directed from W1W_{1} to W0W_{0}, we see that |Vx0||W0|/2|V_{x0}|\geq|W_{0}|/2. Therefore,

|Vx0|cxd4Δ|Vx|=cd4d+1Δa(x)d|Vx|(z a proper prefix of x0cd4d+1Δa(z)d)N,|V_{x0}|\geq\frac{c_{x}^{d}}{4\Delta}|V_{x}|=\frac{c^{d}}{4^{d+1}\Delta a(x)^{d}}|V_{x}|\geq\left(\prod_{z\text{ a proper prefix of }x0}\frac{c^{d}}{4^{d+1}\Delta a(z)^{d}}\right)N, (3.2)

since the proper prefixes of x0x0 are just xx, in addition to all the proper prefixes of xx itself.

As discussed above, we can now recursively define VyV_{y} for all yy prefixed by x0x0. It thus only remains to define Vx1VxV_{x1}\subseteq V_{x}, under the assumption that we have defined VyVx0W0V_{y}\subseteq V_{x0}\subseteq W_{0} for all yy prefixed by x0x0.

To do so, let y1,,ya0(x)y_{1},\dotsc,y_{a_{0}(x)} denote the set of codewords prefixed by x0x0 with the property that some vertex in ρ1(yi)\rho^{-1}(y_{i}) is adjacent to some vertex in Hx1H_{x1}, noting that there are exactly a0(x)a_{0}(x) such codewords by the definition of a0(x)a_{0}(x). Let W1(i)W_{1}^{(i)} denote the subset of W1W_{1} consisting of vertices in W1W_{1} whose out-degree to VyiV_{y_{i}} is at least c|Vyi|c|V_{y_{i}}|. Since every vertex in VyiVx0V_{y_{i}}\subseteq V_{x0} has at most 2cx|W1|2c_{x}|W_{1}| in-neighbors in W1W_{1}, we see that the total number of edges directed from W1W_{1} to VyiV_{y_{i}} is at most 2cx|W1||Vyi|2c_{x}|W_{1}||V_{y_{i}}|, and therefore |W1(i)|(2cx/c)|W1||W_{1}^{(i)}|\leq(2c_{x}/c)|W_{1}|. Because of this, we have that

|i=1a0(x)W1(i)|i=1a0(x)|W1(i)|a0(x)2cxc|W1|=|W1|2.\left|\bigcup_{i=1}^{a_{0}(x)}W_{1}^{(i)}\right|\leq\sum_{i=1}^{a_{0}(x)}|W_{1}^{(i)}|\leq a_{0}(x)\frac{2c_{x}}{c}|W_{1}|=\frac{|W_{1}|}{2}.

Therefore, we can define Vx1=W1(i=1a0(x)W1(i))V_{x1}=W_{1}\setminus\left(\bigcup_{i=1}^{a_{0}(x)}W_{1}^{(i)}\right), and we have that |Vx1||W1|/2|V_{x1}|\geq|W_{1}|/2. By the same computation as in equation (3.2), we see that this definition of Vx1V_{x1} satisfies our desired lower bound on the size of Vx1V_{x1}.

We claim that in this construction, if yzy\prec z are codewords that are adjacent under ρ\rho, then the pair (Vy,Vz)(V_{y},V_{z}) is (1c)(1-c)-dense. To see this, let xx be the longest common prefix of yy and zz. In the construction at level xx, we first proceeded to either Vx0V_{x0} or Vx1V_{x1} in the depth-first search, depending on the relative sizes of a0(x)a_{0}(x) and a1(x)a_{1}(x). In the former case, we ensured that every vertex in Vx1V_{x1} had out-degree at most c|Vy|c|V_{y}| to VyV_{y}, while in the latter case, we ensured that every vertex in Vx0V_{x0} had in-degree at most c|Vz|c|V_{z}| to VzV_{z}. In either case, we see that (Vy,Vz)(V_{y},V_{z}) is (1c)(1-c)-dense, since VyVx0V_{y}\subseteq V_{x0} and VzVx1V_{z}\subseteq V_{x1}. Additionally, by the same computation as in equation (3.1), we see that |Vy|(cd4d+1Δ)depth(ρ)comp(ρ)dN|V_{y}|\geq\left(\frac{c^{d}}{4^{d+1}\Delta}\right)^{\operatorname{depth}(\rho)}\operatorname{comp}(\rho)^{-d}N for every yCy\in C. This verifies all the properties of a (ρ,1c)(\rho,1-c)-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 11-dense pair in an HH-free tournament TT, and the second then iterates the first to form a skeleton of 11-dense pairs. The main difference between these and the previous results are that for these lemmas, we need strengthened assumptions on HH (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 11-dense pair by failing to greedily embed HH in the skeleton given by Lemma 3.9.

Lemma 3.10.

Let d2d\geq 2, and suppose that HH is a dd-degenerate acyclic digraph on nn vertices with maximum degree Δ\Delta, and suppose that ρ:V(H)C\rho:V(H)\to C is some forest prefix labeling with maximum degree AA and maximum component size mm. Let TT be an HH-free tournament on N(210AΔ2)ddepth(ρ)comp(ρ)dnN\geq(2^{10}A\Delta^{2})^{d\operatorname{depth}(\rho)}\operatorname{comp}(\rho)^{d}n vertices. Then TT contains a 11-dense pair (U1,U2)(U_{1},U_{2}) of size at least m1(210AΔ2)ddepth(ρ)comp(ρ)dNm^{-1}(2^{10}A\Delta^{2})^{-d\operatorname{depth}(\rho)}\operatorname{comp}(\rho)^{-d}N.

Proof.

We first apply Lemma 3.9 to the prefix labeling ρ\rho and with parameter c=1/(16AΔ)c=1/(16A\Delta). This lemma outputs a (ρ,1116AΔ)(\rho,1-\frac{1}{16A\Delta})-skeleton {Vx}xC\{V_{x}\}_{x\in C} in TT of size at least

(cd4d+1Δ)depth(ρ)comp(ρ)dN12Δn.\left(\frac{c^{d}}{4^{d+1}\Delta}\right)^{\operatorname{depth}(\rho)}\operatorname{comp}(\rho)^{-d}N\geq 12\Delta n.

For every pair of codewords xyx\prec y that are adjacent under ρ\rho, let Vx,yV_{x,y} denote the set of vertices in VxV_{x} whose in-degree to VyV_{y} is at least 18Δ|Vy|\frac{1}{8\Delta}|V_{y}|. Since at most 116AΔ|Vx||Vy|\frac{1}{16A\Delta}|V_{x}||V_{y}| of the edges between VxV_{x} and VyV_{y} are directed from VyV_{y} to VxV_{x}, we see that |Vx,y|12A|Vx||V_{x,y}|\leq\frac{1}{2A}|V_{x}|. Therefore, if we define Vx=Vx(yxVx,y)V_{x}^{\prime}=V_{x}\setminus\left(\bigcup_{y\succ x}V_{x,y}\right), then we see that |Vx|12|Vx||V_{x}^{\prime}|\geq\frac{1}{2}|V_{x}|, since there are at most AA choices for yxy\succ x with x,yx,y adjacent under ρ\rho. Additionally, every vertex in VxV_{x}^{\prime} has in-degree at most 18Δ|Vy|14Δ|Vy|\frac{1}{8\Delta}|V_{y}|\leq\frac{1}{4\Delta}|V_{y}^{\prime}| from any VyV_{y}^{\prime} with yxy\succ x such that x,yx,y are adjacent under ρ\rho. We now attempt to greedily embed HH in these sets {Vx}xC\{V_{x}^{\prime}\}_{x\in C}.

Let the codewords of CC be x1,,xrx_{1},\dotsc,x_{r}, sorted so that i<ji<j if and only if xixjx_{i}\prec x_{j}. Let Pi=ρ1(xi)P_{i}=\rho^{-1}(x_{i}), so that PiP_{i} is an oriented forest, each of whose weakly connected components has at most mm vertices. Additionally, P1PrP_{1}\sqcup\dotsb\sqcup P_{r} forms a directed partition of V(H)V(H), which we recall means that every edge of HH is oriented from PiP_{i} to PjP_{j} where iji\leq j. For every vertex vPiv\in P_{i}, we initialize a set of candidates Ui,v(0)=VxiU_{i,v}^{(0)}=V_{x_{i}}^{\prime}. Inductively, having defined Ui,v(t)U_{i,v}^{(t)} for each i>ti>t and each vPiv\in P_{i}, we attempt to pick an embedding ϕt:H[Pt]Vxt\phi_{t}:H[P_{t}]\hookrightarrow V_{x_{t}}^{\prime}, such that for every vPtv\in P_{t}, ϕt(v)Ut,v(t1)\phi_{t}(v)\in U_{t,v}^{(t-1)}. If such a ϕt\phi_{t} exists, then for each i>ti>t and each vPiv\in P_{i}, we let

Ui,v(t){uUi,v(t1)\ϕt(Pt):wN(v)Pt,ϕt(w)u}.U_{i,v}^{(t)}\coloneqq\{u\in U_{i,v}^{(t-1)}\backslash\phi_{t}(P_{t}):{\forall w\in N^{-}(v)\cap P_{t},}\phi_{t}(w)\rightarrow u\}.

Note that by the structure of the sets Vx1,,VxrV_{x_{1}}^{\prime},\dotsc,V_{x_{r}}^{\prime}, we only change Ui,v(t)U_{i,v}^{(t)} as follows. First, in at most Δ\Delta steps, we embed an in-neighbor of vv, and we decrease |Ui,v(t)||U_{i,v}^{(t)}| by at most 14Δ|Vxi|\frac{1}{4\Delta}|V_{x_{i}}^{\prime}|. Additionally, we remove at most nn additional vertices from Ui,v(t)U_{i,v}^{(t)}, corresponding to vertices that were picked as images of ϕt\phi_{t}. In total, we remove at most Δ14Δ|Vxi|+n12|Vxi|\Delta\cdot\frac{1}{4\Delta}|V_{x_{i}}^{\prime}|+n\leq\frac{1}{2}|V_{x_{i}}^{\prime}| vertices. We thus see that |Ui,v(t)|12|Vxi||U_{i,v}^{(t)}|\geq\frac{1}{2}|V_{x_{i}}^{\prime}| for all tt.

If we are able to run this process for all 1tr1\leq t\leq r, then we have embedded a copy of HH in TT, so we may assume that the process fails at some step tt. This means that there is no embedding ϕt:H[Pt]T\phi_{t}:H[P_{t}]\hookrightarrow T such that every vertex vPtv\in P_{t} is mapped to Ut,v(t1)U_{t,v}^{(t-1)}. Therefore, by applying Lemma 3.4 to H[Pt]H[P_{t}], we conclude that TT contains a 11-dense pair (U1,U2)(U_{1},U_{2}) of size at least

|Ut,v(t1)|4mΔ|Vxt|16mΔ1m(210AΔ2)ddepth(ρ)comp(ρ)dN.\frac{|U_{t,v}^{(t-1)}|}{4m\Delta}\geq\frac{|V_{x_{t}}|}{16m\Delta}\geq\frac{1}{m}(2^{10}A\Delta^{2})^{-d\operatorname{depth}(\rho)}\operatorname{comp}(\rho)^{-d}N.\qed

Our next result shows how we can iterate the previous lemma to construct many 11-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 (ρ,1c)(\rho,1-c)-skeleton. This lemma takes as input two (not necessarily distinct) prefix labelings on HH, 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 HH according to one labeling to construct a good embedding structure for the other labeling.

Lemma 3.11.

Let d2d\geq 2, and suppose that HH is a dd-degenerate acyclic digraph on nn vertices with maximum degree Δ\Delta. Let ρ1:V(H)C1\rho_{1}:V(H)\to C_{1} and ρ2:V(H)C2\rho_{2}:V(H)\to C_{2} be two prefix labelings, such that ρ1\rho_{1} is a forest prefix labeling of maximum degree A1A_{1} and maximum component size m1m_{1}. Let

γ=(m1(210A1Δ2)ddepth(ρ1)comp(ρ1)d)1.\gamma=\left(m_{1}(2^{10}A_{1}\Delta^{2})^{d\operatorname{depth}(\rho_{1})}\operatorname{comp}(\rho_{1})^{d}\right)^{-1}.

If TT is an HH-free tournament on Nγdepth(ρ2)nN\geq\gamma^{-\operatorname{depth}(\rho_{2})}n vertices, then TT contains a (ρ2,1)(\rho_{2},1)-skeleton of size at least γdepth(ρ2)N\gamma^{\operatorname{depth}(\rho_{2})}N.

Proof.

As in the proof of Lemma 3.9, we construct our skeleton by assigning a set VxV(T)V_{x}\subseteq V(T) for every x{0,1}x\in\{0,1\}^{*} that is a prefix of some codeword in C2C_{2}, with the property that Vx0,Vx1V_{x0},V_{x1} are both subsets of VxV_{x}. We guarantee inductively that

|Vx|γ|x|N,|V_{x}|\geq\gamma^{|x|}N, (3.3)

where |x||x| is the length of xx. To begin the induction, we set V=V(T)V_{\varnothing}=V(T), which satisfies our size hypothesis since ||=0|\varnothing|=0. Inductively, suppose we’ve defined VxV_{x}. If xC2x\in C_{2}, we stop. If not, we apply Lemma 3.10 to the induced subtournament on VxV_{x}, which we may do since

|Vx|\displaystyle|V_{x}| γ|x|Nγdepth(ρ2)1N(210A1Δ2)ddepth(ρ1)comp(ρ1)dn.\displaystyle\geq\gamma^{|x|}N\geq\gamma^{\operatorname{depth}(\rho_{2})-1}N\geq(2^{10}A_{1}\Delta^{2})^{d\operatorname{depth}(\rho_{1})}\operatorname{comp}(\rho_{1})^{d}n.

This allows us to find a 11-dense pair (U1,U2)(U_{1},U_{2}) of size at least γ|Vx|\gamma|V_{x}|. We then set Vx0=U1V_{x0}=U_{1} and Vx1=U2V_{x1}=U_{2}, which we see satisfy (3.3) inductively. We continue in this way until we define VxV_{x} for every xC2x\in C_{2}. To conclude, suppose that y,zC2y,z\in C_{2} are adjacent under ρ2\rho_{2}, and let xx be their longest common prefix. Then VyVx0V_{y}\subseteq V_{x0} and VzVx1V_{z}\subseteq V_{x1}, and we know that every edge between Vx0V_{x0} and Vx1V_{x1} is oriented from Vx0V_{x0} to Vx1V_{x1}, which implies that (Vy,Vz)(V_{y},V_{z}) is 11-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 HH in any sufficiently large tournament TT. The basic idea is that the (ρ,δ)(\rho,\delta)-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 HH.

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 r1(H)\overrightarrow{r_{1}}(H) in terms of the depth and complexity of a prefix coloring of HH.

Theorem 3.12.

Let HH be an acyclic digraph on nn vertices with maximum degree Δ1\Delta\geq 1. Then for any prefix coloring ρ:V(H)C\rho:V(H)\to C, we have r1(H)N\overrightarrow{r_{1}}(H)\leq N, where

N=(25Δ+4Δ2Δ+1)depth(ρ)comp(ρ)Δn.N=\left(2^{5\Delta+4}\Delta^{2\Delta+1}\right)^{\operatorname{depth}(\rho)}\operatorname{comp}(\rho)^{\Delta}n.
Proof.

Let TT be a tournament on NN vertices, and suppose for contradiction that TT is HH-free. By Lemma 3.9 applied with d=Δd=\Delta and c=18Δ2c=\frac{1}{8\Delta^{2}}, we can find in TT a (ρ,118Δ2)(\rho,1-\frac{1}{8\Delta^{2}})-skeleton {Vx}xC\{V_{x}\}_{x\in C} of size at least

((8Δ2)Δ4Δ+1Δ)depth(ρ)comp(ρ)ΔN=(25Δ+2Δ2Δ+1)depth(ρ)comp(ρ)ΔN4n.\left(\frac{(8\Delta^{2})^{-\Delta}}{4^{\Delta+1}\Delta}\right)^{\operatorname{depth}(\rho)}\operatorname{comp}(\rho)^{-\Delta}N=\left(2^{5\Delta+2}\Delta^{2\Delta+1}\right)^{-\operatorname{depth}(\rho)}\operatorname{comp}(\rho)^{-\Delta}N\geq 4n.

Now, for any vertex viV(H)v_{i}\in V(H), let Vi=Vρ(vi)V_{i}=V_{\rho(v_{i})}. Then since ρ1(x)\rho^{-1}(x) is an independent set for any xCx\in C, we see that if vivjv_{i}\to v_{j} is an edge of HH, then (Vi,Vj)(V_{i},V_{j}) is a (118Δ2)(1-\frac{1}{8\Delta^{2}})-dense pair. Therefore, by Lemma 3.5, we conclude that TT contains a copy of HH. ∎

The second result of this subsection uses the (ρ,1)(\rho,1)-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 d2d\geq 2, and let HH be a dd-degenerate acyclic digraph on nn vertices with maximum degree Δ\Delta. Suppose that ρ1:V(H)C1\rho_{1}:V(H)\to C_{1} is a forest prefix labeling of HH and that ρ2:V(H)C2\rho_{2}:V(H)\to C_{2} is any prefix labeling. Let A1A_{1} and m1m_{1} be the maximum degree and maximum component size of ρ1\rho_{1}, respectively. Then r1(H)N\overrightarrow{r_{1}}(H)\leq N, where

N=(m1(210A1Δ2)ddepth(ρ1)comp(ρ1)d)depth(ρ2)max(n,maxxC2r1(H[ρ21(x)])).N=\left(m_{1}(2^{10}A_{1}\Delta^{2})^{d\operatorname{depth}(\rho_{1})}\operatorname{comp}(\rho_{1})^{d}\right)^{\operatorname{depth}(\rho_{2})}\max\left(n,\max_{x\in C_{2}}\overrightarrow{r_{1}}(H[\rho_{2}^{-1}(x)])\right).
Proof.

Let TT be a tournament on NN vertices, and suppose that TT is HH-free. We apply Lemma 3.11 to TT, which allows us to find a (ρ2,1)(\rho_{2},1)-skeleton {Vx}xC2\{V_{x}\}_{x\in C_{2}} where

|Vx|maxyC2r1(H[ρ21(y)])r1(H[ρ21(x)])|V_{x}|\geq\max_{y\in C_{2}}\overrightarrow{r_{1}}(H[\rho_{2}^{-1}(y)])\geq\overrightarrow{r_{1}}(H[\rho_{2}^{-1}(x)])

for all xC2x\in C_{2}. In other words, this is a collection of disjoint sets {Vx}xC2\{V_{x}\}_{x\in C_{2}} such that if xx precedes yy in the lexicographic ordering, then every edge is oriented from VxV_{x} to VyV_{y}. By the definition of the Ramsey number, we see that the induced subtournament T[Vx]T[V_{x}] must contain a copy of H[ρ21(x)]H[\rho_{2}^{-1}(x)] for all xC2x\in C_{2}. We pick such a copy arbitrarily for each xC2x\in C_{2}, and observe that their union forms a copy of HH in TT. ∎

4 Upper bounds on oriented Ramsey numbers

4.1 Upper bounds in terms of height

Theorem 3.12, which bounds r1(H)\overrightarrow{r_{1}}(H) in terms of the dyadic complexity and depth of a prefix coloring, allows us to prove bounds on r1(H)\overrightarrow{r_{1}}(H) 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 HH is an acyclic digraph of height hh, then there exists a prefix coloring ρ:V(H)C\rho:V(H)\to C with depth(ρ)logh\operatorname{depth}(\rho)\leq\lceil\log h\rceil and comp(ρ)hlogh\operatorname{comp}(\rho)\leq h^{\lceil\log h\rceil}.

Proof.

Recall that if HH has height hh, then HH has a directed partition P0,,Ph1P_{0},\ldots,P_{h-1} into hh independent sets. This partition naturally yields a prefix coloring using the prefix code CC consisting of all binary strings of length logh\lceil\log h\rceil. Namely, we can label each vertex in PiP_{i} by the base-22 representation of ii, and this yields a prefix coloring with depth logh\lceil\log h\rceil. To estimate the dyadic complexity of this prefix coloring, note that for any binary string xx of length logh\ell\leq\lceil\log h\rceil, we have that a(x)2logha(x)\leq 2^{\lceil\log h\rceil-\ell}, since there are at most 2logh2^{\lceil\log h\rceil-\ell} codewords in CC prefixed by xx. Therefore,

comp(ρ)=0logh2logh=2logh2=0loghhlogh.\operatorname{comp}(\rho)\leq\prod_{\ell=0}^{\lceil\log h\rceil}2^{\lceil\log h\rceil-\ell}=2^{\lceil\log h\rceil^{2}-\sum_{\ell=0}^{\lceil\log h\rceil}\ell}\leq h^{\lceil\log h\rceil}.\qed

Then Theorem 1.4 follows as a simple corollary.

Proof of Theorem 1.4.

The result is immediate if Δ=1\Delta=1, so assume Δ2\Delta\geq 2. Let ρ\rho be the prefix coloring from Lemma 4.1, which has depth(ρ)logh\operatorname{depth}(\rho)\leq\lceil\log h\rceil and comp(ρ)hlogh\operatorname{comp}(\rho)\leq h^{\lceil\log h\rceil}. By Theorem 3.12,

r1(H)(25Δ+4Δ2Δ+1)depth(ρ)comp(ρ)Δn27ΔloghΔ3Δloghh2Δloghn(Δh)10Δloghn.\overrightarrow{r_{1}}(H)\leq\left(2^{5\Delta+4}\Delta^{2\Delta+1}\right)^{\operatorname{depth}(\rho)}\operatorname{comp}(\rho)^{\Delta}n\leq 2^{7\Delta\log h}\Delta^{3\Delta\log h}h^{2\Delta\log h}n\leq(\Delta h)^{10\Delta\log h}n.\qed

Similarly, given a graded digraph, one can find a prefix coloring with small depth and dyadic complexity.

Lemma 4.2.

If HH is a graded acyclic digraph of height hh, then there exists a prefix coloring ρ:V(H)C\rho:V(H)\to C with depth(ρ)logh\operatorname{depth}(\rho)\leq\lceil\log h\rceil and comp(ρ)=1\operatorname{comp}(\rho)=1.

Proof.

The proof is identical to that of Lemma 4.1, except that a(x)1a(x)\leq 1 for every binary string xx, since the only codeword prefixed by x0x0 that can have edges to a codeword prefixed by x1x1 is the codeword (x,0,1,1,,1)(x,0,1,1,\dotsc,1). ∎

As before, Theorem 1.5 follows as a corollary.

Proof of Theorem 1.5.

We may again assume that Δ2\Delta\geq 2. Let ρ\rho be the prefix coloring from Lemma 4.2, with depth(ρ)logh\operatorname{depth}(\rho)\leq\lceil\log h\rceil and comp(ρ)=1\operatorname{comp}(\rho)=1. Then from Theorem 3.12,

r1(H)(25Δ+4Δ2Δ+1)depth(ρ)comp(ρ)Δn27ΔloghΔ3Δloghnh10ΔlogΔn.\overrightarrow{r_{1}}(H)\leq\left(2^{5\Delta+4}\Delta^{2\Delta+1}\right)^{\operatorname{depth}(\rho)}\operatorname{comp}(\rho)^{\Delta}n\leq 2^{7\Delta\log h}\Delta^{3\Delta\log h}n\leq h^{10\Delta\log\Delta}n.\qed

Recall from the introduction that the bandwidth of an nn-vertex acyclic digraph HH is the least \ell so that PnP_{n}^{\ell} contains HH. Using the same argument, one can obtain a bound of r1(H)nO(1)\overrightarrow{r_{1}}(H)\leq n^{O_{\ell}(1)} for any nn-vertex acyclic digraph of bandwidth at most \ell, using the fact that the same binary-representation prefix coloring have dyadic complexity at most nO(log)n^{O(\log\ell)}. 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 d1d\geq 1 is bounded, then w.h.p. r1(H)\overrightarrow{r_{1}}(H) is nearly linear when H=G(n,d)H=\overrightarrow{G}(n,d). Additionally, we prove a nearly-linear upper bound when p=d/np=d/n and H=G(n,p)H=\overrightarrow{G}(n,p).

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 HH be a 11-degenerate digraph (i.e. an orientation of an undirected forest) on n2n\geq 2 vertices. If TT is any tournament on at least 218n4716\frac{21}{8}n-\frac{47}{16} vertices, then TT contains a copy of HH.

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 va(H)\overrightarrow{va}(H) of an acyclic digraph HH is the minimum size rr of a directed partition of HH into sets {Pi}i=1r\{P_{i}\}_{i=1}^{r} where H[Pi]H[P_{i}] is 11-degenerate for all ii.

Additionally, if HH has a directed partition into sets {Pi}i=1r\{P_{i}\}_{i=1}^{r} such that H[Pi]H[P_{i}] is 11-degenerate for all ii, and every weakly connected component in H[Pi]H[P_{i}] has at most mm vertices, then we say that HH has an (r,m)(r,m)-forest partition.

Both of our upper bound results for random digraphs follow from the following theorem, which yields an upper bound on r1(H)\overrightarrow{r_{1}}(H) in terms of the degeneracy, maximum degree, and forest partition size of HH.

Theorem 4.5.

Let HH be an acyclic digraph on nn vertices with maximum degree Δ\Delta, degeneracy d2d\geq 2, and with an (r,m)(r,m)-forest partition. Then r1(H)(rΔ)6d(logr)2m2logrn\overrightarrow{r_{1}}(H)\leq(r\Delta)^{6d(\log r)^{2}}m^{2\log r}n.

Proof.

Let s=logrs=\lceil\log r\rceil, and let CC denote the prefix code consisting of all strings of length ss. Fix a partition P0Pr1P_{0}\sqcup\dotsb\sqcup P_{r-1} of V(H)V(H) into directed forests such that every edge between PiP_{i} and PjP_{j} is oriented from PiP_{i} to PjP_{j} for all i<ji<j, and such that H[Pi]H[P_{i}] has weakly connected components of size at most mm. Let ρ:V(H)C\rho:V(H)\to C be the forest prefix labeling mapping PiP_{i} to the binary representation of ii. Then the maximum degree of ρ\rho is at most |C|=2s2r|C|=2^{s}\leq 2r, the maximum component size of ρ\rho is at most mm, and depth(ρ)=s\operatorname{depth}(\rho)=s. Moreover, we can bound the dyadic complexity of ρ\rho as comp(ρ)rs\operatorname{comp}(\rho)\leq r^{s}, since every binary string is the prefix of at most rr codewords in CC. We now now apply Theorem 3.13 with ρ1=ρ2=ρ\rho_{1}=\rho_{2}=\rho. We recall that since ρ1(x)\rho^{-1}(x) is a forest for every xx, we have that r1(H[ρ1(x)])3n\overrightarrow{r_{1}}(H[\rho^{-1}(x)])\leq 3n by Theorem 4.3. Therefore,

r1(H)\displaystyle\overrightarrow{r_{1}}(H) (m1(210A1Δ2)ddepth(ρ1)comp(ρ1)d)depth(ρ2)max(n,maxxC2r1(H[ρ21(x)]))\displaystyle\leq\left(m_{1}(2^{10}A_{1}\Delta^{2})^{d\operatorname{depth}(\rho_{1})}\operatorname{comp}(\rho_{1})^{d}\right)^{\operatorname{depth}(\rho_{2})}\max\left(n,\max_{x\in C_{2}}\overrightarrow{r_{1}}(H[\rho_{2}^{-1}(x)])\right)
(m(211rΔ2)dsrds)s(3n)\displaystyle\leq\left(m(2^{11}r\Delta^{2})^{ds}r^{ds}\right)^{s}\cdot(3n)
(rΔ)6d(logr)2m2logrn.\displaystyle\leq(r\Delta)^{6d(\log r)^{2}}m^{2\log r}n.\qed

It remains to check that both G(n,d)\overrightarrow{G}(n,d) and G(n,p)\overrightarrow{G}(n,p) satisfy the conditions of Theorem 4.5 for appropriate Δ,d,m,\Delta,d,m, and rr. 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 G(n,p)\overrightarrow{G}(n,p), where np=d1np=d\geq 1. The idea for bounding va(H)\overrightarrow{va}(H) is to equitably divide [n][n] into 5d5d intervals IiI_{i}, and note that H[Ii]G(n/5d,p)H[I_{i}]\sim\overrightarrow{G}(n/5d,p) 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 O(logn)O(\log n) vertices. Each interval IiI_{i} can be further divided in two to break the unicyclic components, which shows that w.h.p.  we have a (10d,O(logn))(10d,O(\log n))-forest partition. The analysis is broadly similar for H=G(n,d)H=\overrightarrow{G}(n,d) but somewhat more technical.

Recall that if d1d\geq 1 and ndnd is even, a uniformly random undirected dd-regular graph G(n,d)G(n,d) 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 G(n,d)G(n,d) generates a random dd-regular multi-graph (with self-loops allowed) by taking a uniformly random perfect matching (a “pairing”) on ndnd points divided into nn dd-sets, and then contracting each dd-set into a single vertex. There are a total of

P(nd)=(nd)!(nd/2)!2nd/2P(nd)=\frac{(nd)!}{(nd/2)!\cdot 2^{nd/2}}

such pairings, and any simple dd-regular graph is equally likely to be generated. If d1d\geq 1 is fixed and nn\rightarrow\infty, it is known that the probability a pairing generates a simple graph is asymptotic to e(1d2)/4e^{(1-d^{2})/4}. To generate an honest G(n,d)G(n,d), repeatedly sample from the above model (expected constant number of samples) until a simple graph is found.

Lemma 4.6.

If d2d\geq 2 is fixed, ndnd is even, and GG is the induced subgraph of G(n,d)G(n,d) on a fixed set of n5d\frac{n}{5d} vertices, then w.h.p. every connected component of GG has order at most 2logn2\log n 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 HH on kk vertices with at least two cycles is formed from a path of length kk by adding edges from each of its ends, see e.g. [27, Theorem 5.5]. Thus, there are at most k2k!k^{2}\cdot k! labelled graphs on kk vertices with this property. For a fixed such HH, we bound the probability it appears among kk fixed vertices v1,,vkv_{1},\ldots,v_{k} in the pairings model for G(n,d)G(n,d). The total number of pairings giving such an HH (without giving rise to multi-edges) can be bounded above by

d2(k+1)P(nd2(k+1)),d^{2(k+1)}\cdot P(nd-2(k+1)),

since there are at most (d2)k+1(d^{2})^{k+1} ways to choose the edges between dd-sets that correspond to the k+1k+1 edges of HH, and then at most P(nd2(k+1))P(nd-2(k+1)) ways to pair the remaining points. Let \mathcal{E} be the event that the pairings model generating a simple graph, which occurs with probability (1+o(1))e(1d2)/4(1+o(1))e^{(1-d^{2})/4}. We get that

Pr[{v1,,vk} is a copy of H]\displaystyle\Pr[\{v_{1},\ldots,v_{k}\}\text{ is a copy of }H\mid\mathcal{E}] (1+o(1))e(1d2)/4d2(k+1)P(nd2(k+1))P(nd)\displaystyle\leq(1+o(1))e^{-(1-d^{2})/4}\cdot\frac{d^{2(k+1)}P(nd-2(k+1))}{P(nd)}
(1+o(1))e(1d2)/4dk+1nk+1.\displaystyle\leq(1+o(1))e^{-(1-d^{2})/4}\cdot\frac{d^{k+1}}{n^{k+1}}.

Thus,

Pr[some such H appears in G](1+o(1))e(1d2)/4k4(n/(5d)k)k2k!dk+1nk+1=o(1),\Pr[\text{some such }H\text{ appears in }G]\leq(1+o(1))e^{-(1-d^{2})/4}\sum_{k\geq 4}\binom{n/(5d)}{k}k^{2}\cdot k!\cdot\frac{d^{k+1}}{n^{k+1}}=o(1),

which completes the proof that every component of GG contains at most one cycle w.h.p.

Next, we show that w.h.p. every connected component has order at most 2logn2\log n with a similar computation. For a given set size kk, the number of labelled trees HH on kk vertices is kk2k^{k-2} by Cayley’s theorem. We bound the probability that a fixed such HH appears among kk fixed vertices v1,,vkv_{1},\ldots,v_{k}. Similarly to before, we obtain that the total number of pairings giving such an HH without multi-edges is bounded above by

d2(k1)P(nd2(k1)),d^{2(k-1)}\cdot P(nd-2(k-1)),

since there are at most (d2)k1(d^{2})^{k-1} ways to choose the edges between dd-sets that correspond to the k1k-1 edges of the tree HH, and at most P(nd2(k1))P(nd-2(k-1)) ways to pair the remaining points. Conditioning on the pairings model generating a simple graph, which occurs with probability (1+o(1))e(1d2)/4(1+o(1))e^{(1-d^{2})/4}, we get

Pr[{v1,,vk} is a copy of H]\displaystyle\Pr[\{v_{1},\ldots,v_{k}\}\text{ is a copy of }H\mid\mathcal{E}] (1+o(1))e(1d2)/4d2(k1)P(nd2(k1))P(nd)\displaystyle\leq(1+o(1))e^{-(1-d^{2})/4}\cdot\frac{d^{2(k-1)}P(nd-2(k-1))}{P(nd)}
(1+o(1))e(1d2)/4dk1nk1.\displaystyle\leq(1+o(1))e^{-(1-d^{2})/4}\cdot\frac{d^{k-1}}{n^{k-1}}.

Taking a union bound over all choices of v1,,vkv_{1},\ldots,v_{k},

Pr[a tree on k vertices appears in G]\displaystyle\Pr[\text{a tree on }k\text{ vertices appears in }G] (1+o(1))e(1d2)/4(n/(5d)k)kk2dk1nk1\displaystyle\leq(1+o(1))e^{-(1-d^{2})/4}\binom{n/(5d)}{k}k^{k-2}\cdot\frac{d^{k-1}}{n^{k-1}}
(1+o(1))e(1d2)/4nkkk2dk1(5d)k(k/e)knk1\displaystyle\leq(1+o(1))e^{-(1-d^{2})/4}\frac{n^{k}\cdot k^{k-2}\cdot d^{k-1}}{(5d)^{k}\cdot(k/e)^{k}\cdot n^{k-1}}
n(e/5+o(1))k,\displaystyle\leq n\cdot(e/5+o(1))^{k},

which is o(1)o(1) for k=2lognk=2\log n. This implies that w.h.p. every component of GG has order at most 2logn2\log n, and completes the proof. ∎

We can now prove Theorem 1.3.

Corollary 4.7.

For any d2d\geq 2, we have that w.h.p., r1(G(n,d))n(logn)4logd\overrightarrow{r_{1}}(\overrightarrow{G}(n,d))\leq n(\log n)^{4\log d} as nn\to\infty.

Proof.

By Lemma 4.6, we know that w.h.p., any fixed n/(5d)n/(5d) vertices of G(n,d)\overrightarrow{G}(n,d) span a disjoint union of trees and unicyclic components, each of which has order at most 2logn2\log n. Therefore, by applying the union bound to 5d5d events, we conclude that w.h.p., G(n,d)\overrightarrow{G}(n,d) has a directed partition into 5d5d parts all with this property. Dividing each part in two to split every cycle, we conclude that w.h.p. G(n,d)\overrightarrow{G}(n,d) has a (10d,2logn)(10d,2\log n)-forest partition. Additionally, since the underlying undirected graph of G(n,d)\overrightarrow{G}(n,d) is dd-regular, we see that G(n,d)\overrightarrow{G}(n,d) is dd-degenerate and has maximum degree dd. So by Theorem 4.5, we conclude that w.h.p. as nn\to\infty,

r1(G(n,d))(10d2)6d(log(10d))2(2logn)2log(10d)n,\overrightarrow{r_{1}}(\overrightarrow{G}(n,d))\leq(10d^{2})^{6d(\log(10d))^{2}}(2\log n)^{2\log(10d)}n,

which is upper-bounded by n(logn)4logdn(\log n)^{4\log d} for any fixed dd and sufficiently large nn. ∎

Let G(n,p)\overrightarrow{G}(n,p) denote the orientation of the Erdős–Rényi random graph G(n,p)G(n,p) on vertex set [n][n] where all edges are oriented to the right. Similarly to the above, we can prove a nearly-linear upper bound on r1(G(n,p))\overrightarrow{r_{1}}(\overrightarrow{G}(n,p)).

Corollary 4.8.

For any d2d\geq 2, we have that w.h.p., r1(G(n,p))n(logn)O(d(logd)2)\overrightarrow{r_{1}}(\overrightarrow{G}(n,p))\leq n\cdot(\log n)^{O(d(\log d)^{2})}, where p=d/np=d/n.

Proof.

It is easy to show that if p=d/np=d/n, then G(n,p)G(n,p) is O(d)O(d)-degenerate (see [19, Theorem 4.8] for a proof of a stronger result). Additionally, it is well-known that the maximum degree of G(n,p)G(n,p) is O(logn/loglogn)O(\log n/\log\log n) for any fixed d2d\geq 2. Finally, by the easier version of Lemma 4.6 (e.g. [27, Theorem 5.5]), we see that G(n,p)\overrightarrow{G}(n,p) has a (10d,2logn)(10d,2\log n)-forest partition. Hence, Theorem 4.5 implies that r1(G(n,p))n(logn)O(d(logd)2)\overrightarrow{r_{1}}(\overrightarrow{G}(n,p))\leq n\cdot(\log n)^{O(d(\log d)^{2})}. ∎

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 GG to be an undirected graph whose vertex set comes with a total order. If the vertex set is a subset of \mathbb{N}, then the total order is assumed to be that of \mathbb{N}. If G1,,GkG_{1},\ldots,G_{k} are ordered graphs, then the ordered Ramsey number r<(G1,,Gk)r_{<}(G_{1},\ldots,G_{k}) is the minimum NN such that any edge-coloring of the complete graph on [N][N] in colors 1,,k1,\ldots,k contains a monochromatic copy of GiG_{i} in color ii for some ii. Here an ordered copy of GiG_{i} is a subgraph isomorphic to GiG_{i} with vertices appearing in the same order. We write r<(G;k)r<(G,,G)r_{<}(G;k)\coloneqq r_{<}(G,\ldots,G) when all the graphs GiG_{i} 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 MM is a random matching on vertex set [n][n], then w.h.p.,

r<(M;2)>nlogn/20loglogn.r_{<}(M;2)>n^{\log n/20\log\log n}.

We remark that the existence of ordered matchings MM with r<(M;2)>nlogn/20loglognr_{<}(M;2)>n^{\log n/20\log\log n} was proven independently by Balko, Cibulka, Král, and Kynčl [1]. We will only need this weaker result.

If HH is an acyclic digraph with a Hamiltonian (directed) path, we say HH is Hamiltonian. It has a unique vertex ordering v1,,vnv_{1},\ldots,v_{n} where consecutive vertices are adjacent and all edges point forwards. We assign to HH a natural ordered graph H+H^{+} on [n][n] where iji\sim j if and only if vivjv_{i}\rightarrow v_{j} in HH.

Lemma 5.2.

If k1k\geq 1 and HH is an acyclic Hamiltonian digraph, then

rk(H)r<(H+;k).\overrightarrow{r_{k}}(H)\geq r_{<}(H^{+};k).
Proof.

If N=r<(H+;k)1N=r_{<}(H^{+};k)-1, there exists a kk-edge-coloring χ\chi of the complete graph on [N][N] in which there is no monochromatic copy of H+H^{+}. Let TT be the transitive tournament on [N][N] with all edges oriented forwards (i.e. iji\to j if and only if i<ji<j), and the edge between iji\to j also colored χ(i,j)\chi(i,j). Since HH has a Hamiltonian path and all edges in TT point forwards in [N][N], any copy of HH in TT corresponds to an ordered copy of H+H^{+} in χ\chi. By construction, there is no monochromatic copy of HH in TT, as desired. ∎

The two results above together can be used to prove Theorem 1.6.

Proof of Theorem 1.6.

Since rk(H)\overrightarrow{r_{k}}(H) is nondecreasing in kk, it suffices to prove the result for k=2k=2. By Theorem 5.1, there exists a matching MM on [n][n] which satisfies r<(M;2)>nlogn/20loglognr_{<}(M;2)>n^{\log n/20\log\log n}. Define HH to be the acyclic digraph on [n][n] where iji\rightarrow j if i<ji<j and either j=i+1j=i+1 or (i,j)(i,j) is an edge of MM. By Lemma 5.2, r2(H)r<(H+;2)\overrightarrow{r_{2}}(H)\geq r_{<}(H^{+};2). On the other hand, MM is a ordered subgraph of H+H^{+}, so r<(H+;2)r<(M;2)r_{<}(H^{+};2)\geq r_{<}(M;2). It follows that

r2(H)r<(M;2)>nlogn/20loglogn,\overrightarrow{r_{2}}(H)\geq r_{<}(M;2)>n^{\log n/20\log\log n},

as desired. Note that HH is the edge union of a path and a matching, so it has maximum degree 33. ∎

5.2 The upper bound

To prove Theorem 1.7, we upper bound kk-color oriented Ramsey numbers by 2k2k-color ordered Ramsey numbers. Let HH^{-} be the ordered graph obtained from H+H^{+} by reversing the vertex order.

Lemma 5.3.

If k1k\geq 1 and HH is an acyclic digraph, then

rk(H)r<(H+,,H+k,H,,Hk).\overrightarrow{r_{k}}(H)\leq r_{<}(\underbrace{H^{+},\ldots,H^{+}}_{k},\underbrace{H^{-},\ldots,H^{-}}_{k}).
Proof.

Let TT be a tournament on

N=r<(H+,,H+k,H,,Hk)N=r_{<}(\underbrace{H^{+},\ldots,H^{+}}_{k},\underbrace{H^{-},\ldots,H^{-}}_{k})

vertices, with an edge-coloring χ\chi using colors 1,,k1,\dotsc,k. Arbitrarily identify V(T)V(T) with [N][N] and define a (2k)(2k)-edge-coloring χ\chi^{\prime} of KNK_{N} by

χ(i,j)={χ(i,j)if i<j and ijχ(i,j)+kelse.\chi^{\prime}(i,j)=\begin{cases}\chi(i,j)&\text{if }i<j\text{ and }i\rightarrow j\\ \chi(i,j)+k&\text{else}.\end{cases}

By the definition of NN, there is either some color ckc\leq k where χ\chi^{\prime} has an ordered copy of H+H^{+} in color cc, or some color c>kc>k where χ\chi^{\prime} has an ordered copy of HH^{-} in cc. In the former case, TT contains a monochromatic copy of HH in color cc with all vertices pointed forwards in the arbitrary ordering, and in the latter case TT contains a monochromatic copy of HH in color ckc-k with all edges pointed backwards. In either case, TT contains a monochromatic copy of HH, as desired. ∎

To prove an upper bound on rk(H)\overrightarrow{r_{k}}(H), it remains to generalize the following upper bound of [11] on 22-color ordered Ramsey numbers to more colors.

Theorem 5.4 ([11, Theorem 3.6]).

If HH is an ordered graph on at most nn vertices with degeneracy d2d\geq 2, then

r<(H,Kn)2O(dlog2(2n/d)).r_{<}(H,K_{n})\leq 2^{O(d\log^{2}(2n/d))}.

The multicolor bound can now be obtained by iterating the above theorem.

Theorem 5.5.

If k,d2k,d\geq 2 and H1,,Hk1H_{1},\ldots,H_{k-1} are dd-degenerate ordered graphs on at most nn vertices, then

r<(H1,,Hk1,Kn)2Ok,d(log2k1n).r_{<}(H_{1},\ldots,H_{k-1},K_{n})\leq 2^{O_{k,d}(\log^{2^{k-1}}n)}.
Proof.

We prove the theorem by induction on kk. The base case k=2k=2 is just Theorem 5.4. For the inductive step, note that if M=r<(H2,,Hk1,Kn)2O(log2k2n)M=r_{<}(H_{2},\ldots,H_{k-1},K_{n})\leq 2^{O(\log^{2^{k-2}}n)} then one obtains

r<(H1,H2,,Hk1,Kn)r<(H1,KM),r_{<}(H_{1},H_{2},\ldots,H_{k-1},K_{n})\leq r_{<}(H_{1},K_{M}),

by combining the last k1k-1 colors into one “super-color.” It follows by applying the base case that

r<(H1,H2,,Hk1,Kn)r<(H1,KM)2O(log2M)2O(log2k1n),r_{<}(H_{1},H_{2},\ldots,H_{k-1},K_{n})\leq r_{<}(H_{1},K_{M})\leq 2^{O(\log^{2}M)}\leq 2^{O(\log^{2^{k-1}}n)},

as desired. ∎

Theorem 1.7 follows by combining Lemma 5.3 with Theorem 5.5.

6 Concluding Remarks

In this section we collect a few appealing open problems on the Ramsey numbers of digraphs. For k,Δ1k,\Delta\geq 1 and n>Δn>\Delta, let Hk,n,ΔH_{k,n,\Delta} be an acyclic digraph HH with nn vertices and maximum degree Δ\Delta maximizing the value of rk(H)\overrightarrow{r_{k}}(H). Much of this paper was devoted to understanding the growth rate of rk(Hk,n,Δ)\overrightarrow{r_{k}}(H_{k,n,\Delta}) for fixed kk and Δ\Delta.

We first consider the one-color case. Theorem 1.2, Lemma 5.3, and Theorem 5.4 together show

nΩ(Δ2/3/log5/3Δ)r1(H1,n,Δ)2O(Δlog2(2n/Δ))n^{\Omega(\Delta^{2/3}/\log^{5/3}\Delta)}\leq\overrightarrow{r_{1}}(H_{1,n,\Delta})\leq 2^{O(\Delta\log^{2}(2n/\Delta))} (6.1)

While we do not know whether the above Ramsey number grows polynomially or super-polynomially in nn for k=1k=1, we also showed that for k2k\geq 2 and Δ3\Delta\geq 3, rk(Hk,n,Δ)\overrightarrow{r_{k}}(H_{k,n,\Delta}) is at least nΩ(logn/loglogn)n^{\Omega(\log n/\log\log n)}. We conjecture that a super-polynomial growth rate is also possible for k=1k=1.

Conjecture 6.1.

There exists an absolute constant Δ\Delta such that

r1(H1,n,Δ)nω(1).\overrightarrow{r_{1}}(H_{1,n,\Delta})\geq n^{\omega(1)}.

In the case of k2k\geq 2 colors, Theorems 1.6 and 1.7 together imply

nΩ(logn/loglogn)rk(Hk,n,Δ)2O(log22k1n).n^{\Omega(\log n/\log\log n)}\leq\overrightarrow{r_{k}}(H_{k,n,\Delta})\leq 2^{O(\log^{2^{2k-1}}n)}. (6.2)

It would be interesting to determine even the logarithmic order of rk(Hk,n,Δ)\overrightarrow{r_{k}}(H_{k,n,\Delta}).

Problem 6.2.

For any fixed k1k\geq 1 and Δ2\Delta\geq 2, determine the order of growth of logrk(Hk,n,Δ)\log\overrightarrow{r_{k}}(H_{k,n,\Delta}).

To our knowledge, the only solved case of Problem 6.2 is k=1k=1, Δ=2\Delta=2. Here, H1,n,2H_{1,n,2} must be a disjoint union of arbitrarily oriented paths and non-strongly oriented cycles (an acyclic HH cannot contain a strongly oriented cycle). Thomason [37] proved that if nn is large enough, r1(Cn)=n\overrightarrow{r_{1}}(C_{n})=n for any non-strongly oriented cycle CnC_{n} on nn vertices, which can be used to show r1(H1,n,2)=n+O(1)\overrightarrow{r_{1}}(H_{1,n,2})=n+O(1).

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 Δ3\Delta\geq 3 and constants ck=ωk(1)c_{k}=\omega_{k}(1) and Ck=2o(k)C_{k}=2^{o(k)} such that

Ω(logckn)logrk(Hk,n,Δ)O(logCkn).\Omega(\log^{c_{k}}n)\leq\log\overrightarrow{r_{k}}(H_{k,n,\Delta})\leq O(\log^{C_{k}}n).

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 dd, the Ramsey number of G(n,d)\overrightarrow{G}(n,d) is w.h.p. bounded above by n(logn)Od(1)n(\log n)^{O_{d}(1)}. We expect that the answer is in fact linear.

Conjecture 6.4.

If d2d\geq 2 is fixed and H=G(n,d)H=\overrightarrow{G}(n,d), then w.h.p. r1(H)=Od(n).\overrightarrow{r_{1}}(H)=O_{d}(n).

This would follow from our techniques if one could prove the following strengthening of Lemma 3.3, in which the size of the 11-dense pair depends only on the maximum degree Δ\Delta (and not on the number mm of vertices).

Conjecture 6.5.

For every Δ1\Delta\geq 1, there exists CΔ>0C_{\Delta}>0 such that the following holds. Let HH be a 11-degenerate digraph with maximum degree Δ\Delta on mm vertices v1,,vmv_{1},\ldots,v_{m}, and let TT be an arbitrary tournament. If there exist sets U1,,UmV(T)U_{1},\ldots,U_{m}\subseteq V(T), each of size MCΔmM\geq C_{\Delta}m, such that there is no embedding ϕ:HT\phi:H\hookrightarrow T satisfying ϕ(vi)Ui\phi(v_{i})\in U_{i}, then TT contains a 11-dense pair with size at least M/CΔM/C_{\Delta}.

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 mm, the maximum size of a tree component in a directed partition of HH 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 ϕ\phi 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 TT is the vertex ordering v1,,vnv_{1},\ldots,v_{n} maximizing the number of forward edges. To see the power of the median ordering, note that vivi+1v_{i}\rightarrow v_{i+1} for every 1in11\leq i\leq n-1 in this ordering, so this immediately exhibits a Hamiltonian path in TT. Previous work showing linear upper bounds on r1(H)\overrightarrow{r_{1}}(H) when HH is an oriented tree (e.g. [16]) or an acyclic digraph of bounded bandwidth [14] all depend on embedding HH into a tournament TT 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 HH without long paths.

Finally, we recall from the introduction the directed Ramsey number rk(H)\overleftrightarrow{r_{k}}(H), which is defined as the least NN such that every kk-coloring of the edges of KN\overleftrightarrow{K_{N}} contains a monochromatic copy of HH. It is easy to see that rk(H)rk(H)\overrightarrow{r_{k}}(H)\geq\overleftrightarrow{r_{k}}(H) for any kk and any acyclic HH. Indeed, if N=rk(H)N=\overrightarrow{r_{k}}(H), then given a kk-edge-coloring of KN\overleftrightarrow{K_{N}}, we may ignore one edge from each anti-parallel pair to obtain a kk-edge-colored NN-vertex tournament, which contains a monochromatic HH. There is also an inequality in the other direction, whose proof is identical to that given in Lemma 5.3, and which states that rk(H)r2k(H,,H,H,,H)\overrightarrow{r_{k}}(H)\leq\overleftrightarrow{r_{2k}}(H,\dots,H,H^{\prime},\dots,H^{\prime}), where there are kk copies of HH and kk of HH^{\prime}, and HH^{\prime} is obtained from HH 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 rk(H)\overleftrightarrow{r_{k}}(H) for any bounded-degree acyclic digraph HH. 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 HH with r2(H)\overleftrightarrow{r_{2}}(H) which grows faster than any fixed polynomial in the number of vertices of HH. For k4k\geq 4 colors, we can similarly use Theorem 1.6 to produce a bounded-degree acyclic digraph HH such that rk(H)\overleftrightarrow{r_{k}}(H) grows super-polynomially. However, there is an interesting intermediate case at k=3k=3 colors, and we end with the following conjecture, which may be easier than Conjecture 6.1.

Conjecture 6.6.

There is an absolute constant Δ\Delta and an infinite sequence {Hn}\{H_{n}\} of nn-vertex acyclic digraphs with maximum degree at most Δ\Delta and r3(Hn)nω(1)\overleftrightarrow{r_{3}}(H_{n})\geq n^{\omega(1)}.

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.