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

On Ramsey-minimal infinite graphs

Jordan Mitchell Barrett
School of Mathematics and Statistics
Victoria University of Wellington
Wellington, New Zealand
[email protected]
   Valentino Vito
Department of Mathematics
Universitas Indonesia
Depok, Indonesia
[email protected]
Abstract

For fixed finite graphs GG, HH, a common problem in Ramsey theory is to study graphs FF such that F(G,H)F\to(G,H), i.e. every red-blue coloring of the edges of FF produces either a red GG or a blue HH. We generalize this study to infinite graphs GG, HH; in particular, we want to determine if there is a minimal such FF. This problem has strong connections to the study of self-embeddable graphs: infinite graphs which properly contain a copy of themselves. We prove some compactness results relating this problem to the finite case, then give some general conditions for a pair (G,H)(G,H) to have a Ramsey-minimal graph. We use these to prove, for example, that if G=SG=S_{\infty} is an infinite star and H=nK2H=nK_{2}, n1n\geqslant 1 is a matching, then the pair (S,nK2)(S_{\infty},nK_{2}) admits no Ramsey-minimal graphs.

1 Introduction

Let FF, GG and HH be possibly infinite, simple graphs with no isolated vertices. We follow some notation in [11]. We say that FF arrows (G,H)(G,H) or that F(G,H)F\to(G,H) if for every red-blue coloring of the edges of FF, there exists either a red GG or a blue HH contained in FF. In this case, we say that FF is an (G,H)(G,H)-arrowing graph. A red-blue coloring of FF is called (G,H)(G,H)-good if FF does not contain a red GG or a blue HH with respect to the coloring. An alternate definition for F(G,H)F\to(G,H) would then be that the graph FF admits no (G,H)(G,H)-good coloring.

A (G,H)(G,H)-arrowing graph FF is said to be (G,H)(G,H)-minimal if there is no proper subgraph FFF^{\prime}\subset F such that F(G,H)F^{\prime}\to(G,H). In other words, FF is (G,H)(G,H)-minimal if it arrows (G,H)(G,H) and Fe↛(G,H)F-e\not\to(G,H) for every eE(F)e\in E(F). The collection of all (G,H)(G,H)-minimal graphs is denoted as (G,H)\mathcal{R}(G,H), and it satisfies the symmetric property (G,H)=(H,G)\mathcal{R}(G,H)=\mathcal{R}(H,G).

The problem involving (G,H)(G,H)-minimal graphs is classically done for finite GG and HH, as introduced in [5]. One of the major problems that arose was determining whether (G,H)\mathcal{R}(G,H) is finite or infinite. Following the studies done by Beardon [1] on magic labelings, Cáceres et al. [6] on metric dimensions, and Stein [13] on extremal graph theory, we attempt to extend this finite problem to an infinite one. To our knowledge, this is the first serious attempt to do so. It appears that some properties which are expected to be true for finite graphs do not hold in the scope of infinite graphs.

For finite graphs GG and HH, it is known that (G,H)\mathcal{R}(G,H) is nonempty. This is because we can obtain a (G,H)(G,H)-minimal graph from an arbitrary (G,H)(G,H)-arrowing graph by iteratively deleting enough edges. However, if one of GG or HH is infinite, then (G,H)\mathcal{R}(G,H) might be empty. As we shall see in Example 13, the ray PP_{\infty} and K2K_{2} as a pair do not admit any minimal graph. If we consider the double ray P2P_{2\infty} instead of PP_{\infty}, we have that (P2,K2)={P2}\mathcal{R}(P_{2\infty},K_{2})=\{P_{2\infty}\}, and thus a minimal graph exists. An intriguing but difficult problem in general would be to classify which pairs (G,H)(G,H) induce an empty (resp., nonempty) (G,H)\mathcal{R}(G,H).

Problem 1.

For which pairs of graphs (G,H)(G,H) is (G,H)\mathcal{R}(G,H) empty?

The study of Ramsey-minimal properties of infinite graphs is naturally related to graphs which are isomorphic to some proper subgraph of themselves. We will call such graphs self-embeddable. Note that if FF is a self-embeddable graph, then we can pick an FFF^{\prime}\subset F isomorphic to FF. Thus, if F(G,H)F\to(G,H) is self-embeddable, then it is not (G,H)(G,H)-minimal since we can choose a proper subgraph F(G,H)F^{\prime}\to(G,H).

Observation 2.

A (G,H)(G,H)-arrowing graph FF that is self-embeddable cannot be minimal.

The notion of a self-embeddable graph differs from that of a self-contained graph [12], which is one isomorphic to a proper induced subgraph of itself. While self-contained graphs have applications in other problems, such as the tree alternative conjecture [3], we do not require the proper subgraph to be induced in our case, hence the differing vocabulary.

The general outline of this paper is as follows. In Section 2, we present the common notation and conventions that we use for this paper. We then give some compactness results for Ramsey-minimal graphs in Section 3. In Section 4, we obtain some general progress on Problem 1. In Section 5, we turn our attention to the case where GG is an infinite graph and HH is a matching nK2nK_{2}. For an example of previous work on Ramsey-minimal finite graphs involving matchings, see Burr et al. [4].

2 Preliminaries

In this paper, we exclusively work with simple graphs G=(V(G),E(G))G=(V(G),E(G)) with no isolated vertices (i.e. every vertex of GG is adjacent to another vertex). Our graphs are taken to be countable (including finite), with the exception of the graphs of Section 3 which may be uncountable. Let ={1,2,}\mathbb{N}=\{1,2,\ldots\} be the set of natural numbers. For nn\in\mathbb{N}, nGnG denotes the graph consisting of nn disjoint copies of GG.

We say that HH is a subgraph of GG (or simply HGH\subseteq G) if V(H)V(G)V(H)\subseteq V(G) and E(H)E(G)E(H)\subseteq E(G). A subgraph HH of GG is proper, and written as HGH\subset G, if E(H)E(H) is a proper subset of E(G)E(G). Also, we say that GG (properly) contains HH, or that HH (properly) embeds into GG, if there is a (proper) subgraph HH^{\prime} of GG such that HHH^{\prime}\cong H.

1122kk
Figure 1: A kk-ray PkP_{k\infty}.

The ray PP_{\infty} is an infinite graph of the form ({x0,x1,},{x0x1,x1x2,})(\{x_{0},x_{1},\ldots\},\{x_{0}x_{1},x_{1}x_{2},\ldots\}), where x0x_{0} is its endpoint. The double ray P2P_{2\infty}, on the other hand, is of the form ({xn:n},{xnxn+1:n})(\{x_{n}:n\in\mathbb{Z}\},\{x_{n}x_{n+1}:n\in\mathbb{Z}\}). In general, a kk-ray PkP_{k\infty} (shown in Figure 1), k1k\geqslant 1, is formed by identifying the endpoints of kk distinct rays.

A family of graphs of particular interest is the family of comb graphs. Let :\ell\colon\mathbb{N}\to\mathbb{N} be a function. The comb CC^{\ell} is a graph obtained from a base ray PP_{\infty} (called the spine) by attaching, for every nn, a path P(n)P_{\ell(n)} of order (n)\ell(n) by one of its endpoints to the vertex xnx_{n} of PP_{\infty}. Other infinite graphs of interest include the (countably) infinite complete graph KK_{\infty} and the (countably) infinite star S=K1,S_{\infty}=K_{1,\infty}.

We use some terminologies of embeddings from [2, 9, 10]. Recall that a graph homomorphism GHG\to H is a map φ:V(G)V(H)\varphi\colon V(G)\to V(H) such that if vwE(G)vw\in E(G), then φ(v)φ(w)E(H)\varphi(v)\varphi(w)\in E(H). If φ\varphi is injective, then the homomorphism is called an embedding GHG\hookrightarrow H. An embedding GGG\hookrightarrow G is said to be a self-embedding of GG. A self-embedding is nontrivial if its image, seen as a graph with vertex set φ(V(G))\varphi(V(G)) and edge set {φ(v)φ(w):vwE(G)}\{\varphi(v)\varphi(w):vw\in E(G)\}, is a proper subgraph of GG.

A graph GG is said to be self-embeddable if it has a nontrivial self-embedding. In other words, a self-embeddable graph is a graph that properly embeds into itself. We say that GG is strongly self-embeddable if it admits an embedding into GvG-v for every vV(G)v\in V(G). A strongly self-embeddable graph is clearly self-embeddable, but the converse does not hold in general, as shown in the following example.

Example 3.

The infinite star SS_{\infty} is self-embeddable but not strongly so (since SS_{\infty} does not embed into the null graph ScS_{\infty}-c, where cc is its center vertex). Another example can be found in the graph GG of Figure 2. It is self-embeddable, with a right translation as its nontrivial self-embedding. However, GG does not embed into the disconnected graph GvG-v, where vv is the vertex indicated in the figure.

vv
Figure 2: A self-embeddable graph GG that is not strongly self-embeddable. The red subgraph illustrates the image of a nontrivial self-embedding of GG in the form of a right translation by 11.

By induction, we easily obtain the following stronger properties for strongly self-embeddable graphs.

Proposition 4.

Let GG be strongly self-embeddable. Then:

  1. (i)

    GVG-V contains GG for every finite VV(G)V\subset V(G);

  2. (ii)

    GEG-E contains GG for every finite EE(G)E\subset E(G).

Proof.

(i) Suppose that V={v1,,vn}V=\{v_{1},\ldots,v_{n}\} and that H=G{v1,,vn1}H=G-\{v_{1},\ldots,v_{n-1}\} contains GG. In other words, there is a subgraph GHG^{\prime}\subseteq H isomorphic to GG. Since GG^{\prime} and GG are isomorphic, GG^{\prime} is strongly self-embeddable. Thus, GvnG^{\prime}-v_{n} must contain GGG^{\prime}\cong G. Since GvnHvn=GVG^{\prime}-v_{n}\subseteq H-v_{n}=G-V, we can conclude that GvG-v contains GG.

(ii) Suppose that e=uve=uv. Since GvG-v contains a copy of GG by hypothesis, we have that GeGvG-e\supseteq G-v contains GG as well. The statement for any GEG-E then follows by induction. ∎

3 Compactness results

The aim of this section is, for infinite (possibly uncountable) graphs FF, GG, HH, to express what F(G,H)F\to(G,H) means in terms of their finite subgraphs F^\hat{F}, G^\hat{G}, H^\hat{H}.

Theorem 5 confirms that all (G,H)(G,H)-minimal graphs are finite if GG and HH are finite. This result assures us that we only need to deal with the case where one of GG and HH is infinite, since taking both as finite graphs would produce a completely finite problem. We prove the theorem by topological means using Tychonoff’s theorem.

Recall that, for a family of topological spaces SiS_{i}, iIi\in I, the product topology on Si\prod S_{i} is generated by basic open sets of the form Ui\prod U_{i}, where each UiU_{i} is open in SiS_{i}, and Ui=SiU_{i}=S_{i} except for finitely many values of ii. Tychonoff’s theorem states that whenever each SiS_{i} is compact, Si\prod S_{i} is also compact.

Theorem 5.

Let FF be a graph, and GG and HH be finite graphs. If F(G,H)F\to(G,H), then there is a finite F^F\hat{F}\subseteq F such that F^(G,H)\hat{F}\to(G,H).

Proof.

Let XX be the product of |E(F)||E(F)|-many copies of the discrete space S={red,blue}S=\{\text{red},\text{blue}\}, equipped with the product topology. We can identify XX with the set of all functions E(F)SE(F)\to S, i.e. the set of all red-blue colorings of FF’s edges. By Tychonoff’s theorem, XX is compact.

By assumption, F(G,H)F\to(G,H), so for every coloring cXc\in X, we can pick a finite set of edges DcD_{c} forming either a red GG or a blue HH. Then, each c|Dcc|_{D_{c}} determines a basic open set

Oc:={d:E(F)S:d|Dc=c|Dc}.O_{c}:=\{d\colon E(F)\to S:d|_{D_{c}}=c|_{D_{c}}\}.

Since cOcc\in O_{c} for all cXc\in X, the collection {Oc:cX}\{O_{c}:c\in X\} covers XX. By compactness, there is a finite sequence c1,,cnXc_{1},\ldots,c_{n}\in X so that Oc1Ocn=XO_{c_{1}}\cup\cdots\cup O_{c_{n}}=X.

Let F^\hat{F} be the subgraph of FF induced by Dc1DcnD_{c_{1}}\cup\cdots\cup D_{c_{n}}. We claim F^(G,H)\hat{F}\to(G,H). Pick a coloring d^:E(F^)S\hat{d}\colon E(\hat{F})\to S, and extend it arbitrarily to a coloring d:E(F)Sd\colon E(F)\to S. Then, there is an ini\leqslant n such that dOcid\in O_{c_{i}}, so DciE(F^)D_{c_{i}}\subseteq E(\hat{F}) either forms a red GG or a blue HH under d^\hat{d}, as required. ∎

Now, we want to be able to characterize embeddability of a graph GG into another graph FF in terms of embeddability of finite subgraphs G^G\hat{G}\subseteq G into FF. We might want to prove something such as:

GG embeds into FF if and only if every finite subgraph G^G\hat{G}\subseteq G embeds into FF.

GGFF
Figure 3: Every finite subgraph G^G\hat{G}\subseteq G embeds into FF, but GG itself does not.

However, this statement is not true; a counterexample is shown in Figure 3. The problem is that the embeddings G^F\hat{G}\hookrightarrow F are incompatible in some sense—larger finite subgraphs G^G\hat{G}\subseteq G must be embedded further down FF, and so there is no way to “stitch together” these embeddings to get an embedding GFG\hookrightarrow F. To ensure compatibility between partial embeddings, we instead work with the following notion of a pointed graph. We also need to ensure that our graphs are locally finite: that is, deg(v)<\deg(v)<\infty for each vertex vv.

A pointed graph is a triple G=(V,E,)G=(V,E,*), where (V,E)(V,E) is a graph, and V*\in V is a specified vertex of GG, called the basepoint of GG. A pointed subgraph HH is a subgraph of GG such that H=G*_{H}=*_{G}. A pointed homomorphism is a graph homomorphism mapping basepoints to basepoints—we call it a pointed embedding if it is injective. A pointed graph is locally finite, connected, etc. if the underlying graph is.

Using pointed graphs, we can obtain a version of our desired result. We first provide the following form of Kőnig’s infinity lemma to pave way for the compactness argument used in the proof of Proposition 7.

Lemma 6 ([7, Lemma 8.1.2]).

Let V0,V1,V_{0},V_{1},\ldots be an infinite sequence of disjoint nonempty finite sets, and let KK be a graph on their union. Assume that every vertex in Vn+1V_{n+1} has a neighbor in VnV_{n}. Then, KK contains a ray v0v1v_{0}v_{1}\ldots such that vnVnv_{n}\in V_{n} for all nn.

Proposition 7.

Let FF and GG be locally finite, pointed graphs, and suppose GG is connected. If every connected, finite, pointed subgraph G^\hat{G} of GG admits a pointed embedding into FF, then GG admits a pointed embedding into FF.

Proof.

Since GG is locally finite and connected, it must be countable (see Exercise 1, Chapter 8 of [7]). The lemma clearly holds if GG is finite, so assume that GG is countably infinite. Enumerate V(G)V(G) as follows: let v0=v_{0}=*, let v1,,vkv_{1},\ldots,v_{k} be the neighbors of *, let vk+1,,vv_{k+1},\ldots,v_{\ell} be the vertices of distance 2 to *, etc. This indeed enumerates GG by the assumption that GG is locally finite and connected. Then, the induced subgraphs G^n:=G[v0,,vn]\hat{G}_{n}:=G[v_{0},\ldots,v_{n}] are all connected, finite, pointed subgraphs of GG.

For each nn, let VnV_{n} be the set of pointed embeddings G^nF\hat{G}_{n}\hookrightarrow F. Each VnV_{n} is nonempty by assumption. Inductively, we show each VnV_{n} is finite. We see that V0V_{0} is finite, since there is a unique embedding G^0F\hat{G}_{0}\hookrightarrow F mapping G*_{G} to F*_{F}. Now assume VnV_{n} is finite, and pick fVnf\in V_{n}. Since G^n+1\hat{G}_{n+1} is connected, vn+1v_{n+1} is adjacent to some vjv_{j} for jnj\leqslant n. Since FF is locally finite, f(vj)f(v_{j}) has finite degree, so there are only finitely many ways to define f(vn+1)f(v_{n+1}) and extend ff to an embedding in Vn+1V_{n+1}. Since there are also only finitely many ways to choose fVnf\in V_{n}, it follows that Vn+1V_{n+1} is finite, and so all the VnV_{n} are finite by induction.

Similarly to before, let KK be the graph on n=0Vn\bigcup_{n=0}^{\infty}V_{n}, where we insert all edges between fVn+1f\in V_{n+1} and f|G^nVnf|_{\hat{G}_{n}}\in V_{n}. By Lemma 6, KK contains an infinite ray f0f1f_{0}f_{1}\ldots such that fnVnf_{n}\in V_{n} for all nn. Define f:V(G)V(F)f\colon V(G)\to V(F) by f(vn)=fn(vn)f(v_{n})=f_{n}(v_{n}). We claim ff is a pointed embedding GFG\hookrightarrow F.

  1. (i)

    ff is a graph homomorphism: suppose vnvmE(G)v_{n}v_{m}\in E(G), where nmn\leqslant m. We have f(vn)=fn(vn)=fm(vn)f(v_{n})=f_{n}(v_{n})=f_{m}(v_{n}) and f(vm)=fm(vm)f(v_{m})=f_{m}(v_{m}). Since fmf_{m} is a graph homomorphism, f(vn)f(vm)=fm(vn)fm(vm)E(F)f(v_{n})f(v_{m})=f_{m}(v_{n})f_{m}(v_{m})\in E(F).

  2. (ii)

    ff is injective: suppose f(vn)=f(vm)f(v_{n})=f(v_{m}) for nmn\leqslant m. Then, fm(vn)=fn(vn)=fm(vm)vn=vmf_{m}(v_{n})=f_{n}(v_{n})=f_{m}(v_{m})\implies v_{n}=v_{m} since fmf_{m} is injective.

  3. (iii)

    ff is pointed: f(G)=f0(G)=Ff(*_{G})=f_{0}(*_{G})=*_{F} since f0f_{0} is pointed.∎

Interestingly, Proposition 7 actually generalizes Kőnig’s lemma, in its more standard, graph-theoretical form (as found in [7, Proposition 8.2.1]).

Corollary 8 (Kőnig’s lemma).

Every locally finite, connected, infinite graph FF contains a ray.

Proof.

Pick an arbitrary basepoint F*\in F. For every nn, we claim that PnP_{n} (with Pn*_{P_{n}} chosen as an endpoint) admits a pointed embedding into FF. If PnP_{n} does not admit a pointed embedding into FF, then there is no vertex vV(F)v\in V(F) such that d(F,v)nd(*_{F},v)\geqslant n. Since FF is connected and locally finite, we can enumerate V(F)V(F) as follows: let v0=v_{0}=*, let v1,,vkv_{1},\ldots,v_{k} be the neighbors of *, let vk+1,,vv_{k+1},\ldots,v_{\ell} be the vertices of distance 2 to *, etc. By stage nn, we will have enumerated all of FF, hence FF is finite; contradiction. The result now follows from Proposition 7, with G=PG=P_{\infty} and P*_{P_{\infty}} as its endpoint. ∎

For pointed graphs FF, GG, and a non-pointed graph HH, we write Fpoi(G,H)F\xrightarrow{poi}(G,H) if for every red-blue coloring of the edges of FF, there exists either a red GG as a pointed subgraph of FF or a blue HH in the underlying graph of FF. This definition gives a stronger condition for FF than F(G,H)F\to(G,H), and, unlike regular arrowing, Fpoi(G,H)F\xrightarrow{poi}(G,H) and Fpoi(H,G)F\xrightarrow{poi}(H,G) are not necessarily equivalent. We also note that Fpoi(G,H)F\xrightarrow{poi}(G,H) only if HK1,deg(F)H\subseteq K_{1,\deg(*_{F})}.

Theorem 9.

Let FF, GG be locally finite, pointed graphs and let HH be a graph. Suppose GG is connected and for every connected, finite, pointed G^G\hat{G}\subseteq G, we have Fpoi(G^,H)F\xrightarrow{poi}(\hat{G},H). Then, Fpoi(G,H)F\xrightarrow{poi}(G,H).

Proof.

Take an arbitrary red-blue coloring of FF such that FF does not contain a blue HH. Denote FF^{\prime} as the pointed subgraph of FF induced by all the red edges. By assumption, all connected, finite, pointed G^G\hat{G}\subseteq G admits a pointed embedding into FF^{\prime}. By Proposition 7, GG admits a pointed embedding into FF^{\prime}, so a red GG exists as a pointed subgraph of FF. ∎

Proposition 7 is precisely Theorem 9 for the case H=K2H=K_{2}. Thus, Theorem 9 generalizes both Proposition 7 and Kőnig’s lemma.

4 General progress on Problem 1

Throughout this section and the next, we fix a pair of (potentially infinite) graphs GG and HH. We provide a sufficient condition under which (G,H)\mathcal{R}(G,H) is empty by first finding some suitable family of graphs \mathcal{F} for the pair (G,H)(G,H).

Theorem 10.

Suppose that \mathcal{F} is a (possibly infinite) collection of graphs such that:

  1. (1)

    F(G,H)F\to(G,H) for every FF\in\mathcal{F};

  2. (2)

    Every (G,H)(G,H)-arrowing graph Γ\Gamma contains some graph FF\in\mathcal{F}.

We have the following:

  1. (i)

    (G,H){F:F is not self-embeddable}\mathcal{R}(G,H)\subseteq\{F\in\mathcal{F}:\text{$F$ is not self-embeddable}\};

  2. (ii)

    If every FF\in\mathcal{F} is self-embeddable, then (G,H)\mathcal{R}(G,H) is empty.

Proof.

(i) Fix a (G,H)(G,H)-minimal graph Γ\Gamma. By condition (2), there is an FF\in\mathcal{F} such that FF is contained in Γ\Gamma. Since F(G,H)F\to(G,H) by condition (1), we must have F=ΓF=\Gamma as Γ\Gamma is (G,H)(G,H)-minimal. Therefore, Γ\Gamma\in\mathcal{F}. By Observation 2, Γ\Gamma is not self-embeddable, and we are done.

(ii) follows directly from (i). ∎

Conditions (1) and (2) are not sufficient to ensure that (G,H)\mathcal{R}(G,H) and {F:F is not self-embeddable}\{F\in\mathcal{F}:\text{$F$ is not self-embeddable}\} coincide. For example, take G=PG=P_{\infty}, H=K2H=K_{2} and ={P,P2}\mathcal{F}=\{P_{\infty},P_{2\infty}\}. While conditions (1) and (2) hold, (P,K2)\mathcal{R}(P_{\infty},K_{2}) can be shown to be empty while P2P_{2\infty} is not self-embeddable. Hence, (G,H){F:F is not self-embeddable}\mathcal{R}(G,H)\subset\{F\in\mathcal{F}:\text{$F$ is not self-embeddable}\}. That said, we can create an extra condition to make both sets equal.

Theorem 11.

Let \mathcal{F} be a collection of graphs such that conditions (1) and (2) of Theorem 10 hold. Suppose that we also have the following condition:

  1. (3)

    F1F_{1} and F2F_{2} do not contain each other for every different F1,F2F_{1},F_{2}\in\mathcal{F}.

We have the following:

  1. (i)

    (G,H)={F:F is not self-embeddable}\mathcal{R}(G,H)=\{F\in\mathcal{F}:\text{$F$ is not self-embeddable}\};

  2. (ii)

    (G,H)\mathcal{R}(G,H) is empty if and only if every FF\in\mathcal{F} is self-embeddable.

Proof.

(i) We prove that {F:F is not self-embeddable}(G,H)\{F\in\mathcal{F}:\text{$F$ is not self-embeddable}\}\subseteq\mathcal{R}(G,H). Suppose FF\in\mathcal{F} is not self-embeddable. Since F(G,H)F\to(G,H) by condition (1), it remains to show that no proper FFF^{\prime}\subset F arrows (G,H)(G,H). If we assume the contrary, then we have an F′′F^{\prime\prime}\in\mathcal{F} contained in FF^{\prime} by condition (2). This implies that FF contains F′′F^{\prime\prime} and contradicts condition (3).

(ii) follows easily from (i). ∎

The preceding theorems have a few applications. For example, in order to prove Theorem 19 of the next section, we will need to use Theorem 10(ii). Also, we can consider the special case where \mathcal{F} is chosen as {G}\{G\}. This yields the following results:

Theorem 12.

We have the following:

  1. (i)

    (G,K2)\mathcal{R}(G,K_{2}) is empty if and only if GG is self-embeddable. If it is nonempty, then (G,K2)={G}\mathcal{R}(G,K_{2})=\{G\};

  2. (ii)

    If HK2H\neq K_{2}, then G(G,H)G\to(G,H) implies that (G,H)\mathcal{R}(G,H) is empty.

Proof.

(i) Take ={G}\mathcal{F}=\{G\}. Conditions (1)–(3) of Theorems 10 and 11 all hold when H=K2H=K_{2}. The statement directly follows from Theorem 11.

(ii) Again, take ={G}\mathcal{F}=\{G\}. Conditions (1) and (2) of Theorem 10 are both satisfied. Now we just need to show that GG is self-embeddable. Color an arbitrary edge of GG blue and the rest of the edges red. Since HK2H\neq K_{2}, there is no blue HH in GG. So by the fact that G(G,H)G\to(G,H), there is a red copy of GG in GG. This proves that GG properly contains itself, and thus self-embeddable. ∎

The following examples demonstrate some direct applications of Theorem 12 for some pairs of graphs:

Example 13.

(Pk,K2)\mathcal{R}(P_{k\infty},K_{2}) is empty if and only if k=1k=1. When k>1k>1, we have (Pk,K2)={Pk}\mathcal{R}(P_{k\infty},K_{2})=\{P_{k\infty}\}. These observations are obtained directly from Theorem 12(i).

Example 14.

By Theorem 12(ii), (S,K1,n)\mathcal{R}(S_{\infty},K_{1,n}) is empty for all 2n2\leqslant n\leqslant\infty since S(S,K1,n)S_{\infty}\to(S_{\infty},K_{1,n}).

Example 15.

(K,H)\mathcal{R}(K_{\infty},H) is empty for all graphs HH. This follows from Theorem 12 since KK_{\infty} is self-embeddable and K(K,H)K_{\infty}\to(K_{\infty},H) for all HH by the infinite Ramsey theorem.

5 Some results involving matchings

We saw in Theorem 12(i) an answer to Problem 1 whenever H=K2H=K_{2}. Now, let us consider the more general case where H=nK2H=nK_{2}. It becomes apparent that the characteristics of (G,nK2)\mathcal{R}(G,nK_{2}), n2n\geqslant 2, are still related to whether GG is (strongly) self-embeddable.

Theorem 16.

We have the following:

  1. (i)

    If GG is connected and not self-embeddable, then for all n2n\geqslant 2, we have nG(G,nK2)nG\in\mathcal{R}(G,nK_{2}) so that (G,nK2)\mathcal{R}(G,nK_{2}) is nonempty;

  2. (ii)

    If GG is strongly self-embeddable, then (G,nK2)\mathcal{R}(G,nK_{2}) is empty for all n2n\geqslant 2.

Proof.

(i) Fix a red-blue coloring of nGnG which does not create a blue nK2nK_{2}. It follows that there must be a component of nGnG isomorphic to GG which is colored all red. Hence, nG(G,nK2)nG\to(G,nK_{2}).

Now let eE(nG)e\in E(nG) be an arbitrary edge located in some component GG^{\prime} of nGnG. We show that nGe↛(G,nK2)nG-e\not\to(G,nK_{2}) by constructing a (G,nK2)(G,nK_{2})-good coloring of nGenG-e as follows: for every component of nGnG other than GG^{\prime}, color one of its edges blue; color the rest of the edges red. Since GG is connected, there are exactly n1n-1 components in nGnG other than GG^{\prime}, so this coloring only manages to produce a blue (n1)K2(n-1)K_{2}. Also, since GG is not self-embeddable, there cannot be a red GG in any of the components of nGnG. By the connectivity of GG, there cannot be a red GG in all of nGnG either. Therefore, this coloring is indeed (G,nK2)(G,nK_{2})-good.

(ii) By appealing to Theorem 12(ii), it suffices to prove that G(G,nK2)G\to(G,nK_{2}). Fix a red-blue coloring of GG which does not create a blue nK2nK_{2}. We claim that this coloring creates a red GG. We construct a set of vertices VV(G)V\subset V(G) using the following algorithm:

  1. 1.

    initialize V=V=\emptyset

  2. 2.

    while GVG-V contains a blue edge do

  3. 3.

    choose a blue edge e=uve=uv in GVG-V

  4. 4.

    VV{u,v}V\leftarrow V\cup\{u,v\}

  5. 5.

    output VV

This algorithm must terminate after at most n1n-1 while loop iterations since the edge chosen at each iteration must be independent from the edges chosen at previous iterations. It is then clear that the output VV is finite and that GVG-V only contains red edges. By Proposition 4(i), we have a copy of GG in GVG-V. It follows that there exists a red copy of GG in GG, and we are done. ∎

Remark 17.

We note that Theorem 16(i) can fail to hold if GG is disconnected. For example, given G=2P2G=2P_{2\infty}, it can be shown that (n+1)P2(G,nK2)(n+1)P_{2\infty}\to(G,nK_{2}). This implies that (2n)P2(2n)P_{2\infty} is not minimal for n2n\geqslant 2, so (2P2,nK2)\mathcal{R}(2P_{2\infty},nK_{2}) cannot be shown to be nonempty using the previous line of reasoning.

Example 18.

Observe that PP_{\infty} is strongly self-embeddable, while PkP_{k\infty} is connected and not self-embeddable for k>1k>1. By Theorem 16, we have for every n2n\geqslant 2, (Pk,nK2)\mathcal{R}(P_{k\infty},nK_{2}) is empty if and only if k=1k=1.

We note that the converse of Theorem 16(ii) does not necessarily hold. There indeed exists a graph G=SG=S_{\infty} not strongly self-embeddable such that (G,nK2)\mathcal{R}(G,nK_{2}) is empty for all nn.

Theorem 19.

For all n1n\geqslant 1, (S,nK2)\mathcal{R}(S_{\infty},nK_{2}) is empty.

We prove Theorem 19 by first defining a collection of graphs n\mathcal{F}_{n} such that conditions (1) and (2) of Theorem 10 hold.

Lemma 20.

For every n1n\geqslant 1, define n\mathcal{F}_{n} to be the collection of all graphs FF satisfying the following two conditions:

  1. (a)

    FF contains exactly nn vertices of infinite degree forming the set X={x1,,xn}X=\{x_{1},\ldots,x_{n}\};

  2. (b)

    If uvE(F)uv\in E(F), then at least one of uu, vv is an element of XX.

Then, n\mathcal{F}_{n} satisfies conditions (1) and (2) of Theorem 10, with G=SG=S_{\infty} and H=nK2H=nK_{2}.

Proof.

To prove condition (1), we show that F(S,nK2)F\to(S_{\infty},nK_{2}) for all FnF\in\mathcal{F}_{n} by induction on nn. The base case n=1n=1 follows from the fact that 1={S}\mathcal{F}_{1}=\{S_{\infty}\} and S(S,K2)S_{\infty}\to(S_{\infty},K_{2}). Now assume that every graph in n\mathcal{F}_{n} arrows (S,nK2)(S_{\infty},nK_{2}), and let Fn+1F\in\mathcal{F}_{n+1} be arbitrary. Suppose that cc is a red-blue coloring of FF which produces no red SS_{\infty}.

Pick an arbitrary vertex uu adjacent to xn+1Xx_{n+1}\in X such that uXu\notin X and the edge xn+1ux_{n+1}u is colored blue. This can be done since otherwise, xn+1x_{n+1} is incident to infinitely many red edges. Observe that F:=F{xn+1,u}F^{\prime}:=F-\{x_{n+1},u\} is an element of n\mathcal{F}_{n}, so it contains a blue nK2nK_{2} with respect to the coloring c|Fc|_{F^{\prime}}. Since this blue nK2nK_{2} and the blue xn+1ux_{n+1}u form an independent set of edges, there exists a blue (n+1)K2(n+1)K_{2} in FF with respect to cc. This proves that F(S,(n+1)K2)F\to(S_{\infty},(n+1)K_{2}) and completes the induction.

Now we prove condition (2). Suppose that Γ\Gamma arrows (S,nK2)(S_{\infty},nK_{2}). We claim that Γ\Gamma contains at least nn vertices of infinite degree. Assume that YY, where |Y|<n|Y|<n, is the set of vertices of Γ\Gamma having an infinite degree. By coloring all edges incident to a vertex in YY blue and the rest of the edges red, we obtain a (S,nK2)(S_{\infty},nK_{2})-good coloring of Γ\Gamma; contradiction. Now, we can take arbitrary vertices x1,,xnx_{1},\ldots,x_{n} of Γ\Gamma of infinite degree. The subgraph of Γ\Gamma induced by all edges that are incident to at least one of x1,,xnx_{1},\ldots,x_{n} is an element of n\mathcal{F}_{n}, therefore condition (2) holds. ∎

Proof of Theorem 19.

Let n1n\geqslant 1. Take n\mathcal{F}_{n} as defined in Lemma 20. Invoking Theorem 10(ii), we need to show that every FnF\in\mathcal{F}_{n} is self-embeddable.

Let FnF\in\mathcal{F}_{n} be arbitrary. We aim to define a nontrivial self-embedding φ\varphi of FF. Denote UU as the complement of XX (i.e. U:=V(F)XU:=V(F)\setminus X), and for all 1in1\leqslant i\leqslant n, let ρi:U{0,1}\rho_{i}\colon U\to\{0,1\} be such that ρi(u)=1\rho_{i}(u)=1 iff uu is adjacent to xix_{i}. Define ρ(u)=(ρ1(u),,ρn(u))\rho(u)=(\rho_{1}(u),\ldots,\rho_{n}(u)) for every uUu\in U. Since the image of ρ\rho is finite, then by the infinite pigeonhole principle, there exists an x{0,1}nx\in\{0,1\}^{n} such that ρ1(x)\rho^{-1}(x) is an infinite set, say ρ1(x)={u1,u2,}\rho^{-1}(x)=\{u_{1},u_{2},\ldots\}. Define φ:V(F)V(F)\varphi\colon V(F)\to V(F) as

φ(v)={ui+1,v=ui,i1,v,otherwise.\varphi(v)=\begin{cases}u_{i+1},&v=u_{i},\ i\geqslant 1,\\ v,&\text{otherwise}.\end{cases}

It is clear that φ\varphi is injective, and it is nontrivial since u1φ(V(F))u_{1}\notin\varphi(V(F)). It remains to show that φ\varphi is a graph homomorphism. Suppose that uvE(F)uv\in E(F). By condition (b), we can assume without loss of generality that vXv\in X, say v=xjv=x_{j} for some 1jn1\leqslant j\leqslant n. Since vUv\notin U, we must have vρ1(x)v\notin\rho^{-1}(x), so φ(v)=v\varphi(v)=v. If uρ1(x)u\notin\rho^{-1}(x), then φ(u)φ(v)=uvE(F)\varphi(u)\varphi(v)=uv\in E(F), and we are done. So assume that u=uiu=u_{i} for some i1i\geqslant 1. Since uv=uixjuv=u_{i}x_{j} is an edge, we have that ρj(ui)=1\rho_{j}(u_{i})=1. Thus, we also have ρj(ui+1)=1\rho_{j}(u_{i+1})=1 since ρ(ui+1)=x=ρ(ui)\rho(u_{i+1})=x=\rho(u_{i}). It follows that

φ(u)φ(v)=φ(ui)φ(xj)=ui+1xjE(F).\varphi(u)\varphi(v)=\varphi(u_{i})\varphi(x_{j})=u_{i+1}x_{j}\in E(F).

The proof that φ\varphi is a nontrivial self-embedding is then complete. ∎

Now, let CC^{\ell} be a comb with ray x0x1x_{0}x_{1}\ldots as its spine and a path P(n)P_{\ell}(n) of order (n)\ell(n) attached to every xnx_{n}. Suppose that the value of sC:=min(n)>1ns_{C^{\ell}}:=\min_{\ell(n)>1}n, which is equal to the smallest natural number nn such that deg(xn)=3\deg(x_{n})=3, exists (that is, CC^{\ell} is not a ray). We can assume that our combs satisfy sC(sC)1s_{C^{\ell}}\geqslant\ell(s_{C^{\ell}})-1 without loss of generality. Indeed, if s=sCs=s_{C^{\ell}} is such that s<(s)1s<\ell(s)-1, then we can define a function

(n)={1,1n<(s)1s+1,n=(s)1,(n(s)+s+1),n(s),\ell^{*}(n)=\begin{cases}1,&1\leqslant n<\ell(s)-1\\ s+1,&n=\ell(s)-1,\\ \ell(n-\ell(s)+s+1),&n\geqslant\ell(s),\end{cases} (1)

so that CCC^{\ell^{*}}\cong C^{\ell} and s=sCs^{*}=s_{C^{\ell^{*}}} satisfies s(s)1s^{*}\geqslant\ell^{*}(s^{*})-1. Here, CC^{\ell^{*}} is basically the comb obtained from CC^{\ell} by exchanging the positions of x0xsx_{0}\ldots x_{s} and P(s)P_{\ell(s)} in the comb.

The following theorem gives equivalent formulations to the statement that (C,nK2)\mathcal{R}(C^{\ell},nK_{2}) is empty for all n2n\geqslant 2. In particular, the theorem gives an answer for Problem 1 whenever GG is a comb and HH is a matching.

Theorem 21.

Let CC^{\ell} be a comb which is not a ray with x0x1x_{0}x_{1}\ldots as its spine. Suppose that the value of s=min(n)>1ns=\min_{\ell(n)>1}n is at least (s)1\ell(s)-1. The following statements are equivalent:

  1. 1.

    (C,nK2)\mathcal{R}(C^{\ell},nK_{2}) is empty for all n2n\geqslant 2;

  2. 2.

    (C,nK2)\mathcal{R}(C^{\ell},nK_{2}) is empty for some n2n\geqslant 2;

  3. 3.

    CC^{\ell} is self-embeddable;

  4. 4.

    CC^{\ell} is strongly self-embeddable;

  5. 5.

    There exists a p1p\geqslant 1 such that (n)(n+p)\ell(n)\leqslant\ell(n+p) for all nn\in\mathbb{N}.

Proof.

Implication (12)(1\to 2) is trivial, while implications (23)(2\to 3) and (41)(4\to 1) are both consequences of Theorem 16. So we just need to prove (54)(5\to 4) and (35)(3\to 5).

545\to 4: Let vV(C)v\in V(C^{\ell}). If v=x0v=x_{0}, then by positively translating CC^{\ell} by pp (so that xnxn+px_{n}\mapsto x_{n+p} and P(n)P_{\ell(n)} maps into P(n+p)P_{\ell(n+p)}), we obtain an embedding CCvC^{\ell}\hookrightarrow C^{\ell}-v. Otherwise, there exists a kk\in\mathbb{N} such that vv is located in the path P(k)P_{\ell(k)} attached to xkx_{k}. In this case, positively translate CC^{\ell} by apap, where aa is chosen such that ap>kap>k, to obtain the desired embedding into CvC^{\ell}-v.

353\to 5: Let φ:V(C)V(C)\varphi\colon V(C^{\ell})\to V(C^{\ell}) be a nontrivial self-embedding given by the assumption that CC^{\ell} is self-embeddable. We cannot have φ(x0)=x0\varphi(x_{0})=x_{0}, since that would mean φ\varphi is the identity map, hence not nontrivial. Thus, φ(x0)\varphi(x_{0}) is located in the path P(k)P_{\ell(k)}, which is attached to xkx_{k}, for some kk\in\mathbb{N}. Suppose that P(k):=xky1y(k)1P_{\ell(k)}:=x_{k}y_{1}\ldots y_{\ell(k)-1}.

If φ(x0)=xk\varphi(x_{0})=x_{k}, then φ\varphi must be a positive translation by kk. Hence, statement 5 holds by taking p=kp=k since φ\varphi maps each P(n)P_{\ell(n)} into P(n+k)P_{\ell(n+k)}. So assume that φ(x0)=yj\varphi(x_{0})=y_{j} for some 1j(k)11\leqslant j\leqslant\ell(k)-1. This has an implication that (k)>1\ell(k)>1, and thus sks\leqslant k since s=min(n)>1ns=\min_{\ell(n)>1}n. Observe that we cannot have j>sj>s since that would mean φ(xs)=yjs\varphi(x_{s})=y_{j-s} has degree 22, which is less than deg(xs)=3\deg(x_{s})=3; contradiction. In summary, we have the inequality jskj\leqslant s\leqslant k.

We claim that j<kj<k. Assume for the sake of contradiction that j=s=kj=s=k. Since we have established that j(k)1j\leqslant\ell(k)-1 and s(s)1=(k)1s\geqslant\ell(s)-1=\ell(k)-1, we then have (k)1=j=s=k\ell(k)-1=j=s=k. It follows that φ\varphi must be the map

φ(v)={yjn,v=xn, 0ns1,xsm,v=ym, 1m(k)1,v,otherwise,\varphi(v)=\begin{cases}y_{j-n},&v=x_{n},\ 0\leqslant n\leqslant s-1,\\ x_{s-m},&v=y_{m},\ 1\leqslant m\leqslant\ell(k)-1,\\ v,&\text{otherwise},\end{cases}

which is not nontrivial; contradiction.

Now we can define p=kj1p=k-j\geqslant 1. We prove that (n)(n+p)\ell(n)\leqslant\ell(n+p) for all nsn\geqslant s (the case where n<sn<s is trivial since then (n)=1\ell(n)=1).

Case 1. j=sj=s. We have previously established the inequalities j(k)1j\leqslant\ell(k)-1 and s(s)1s\geqslant\ell(s)-1. We thus have

(s)s+1=j+1(k)=(s+p).\ell(s)\leqslant s+1=j+1\leqslant\ell(k)=\ell(s+p).

In addition, φ\varphi necessarily maps each P(n)P_{\ell(n)}, n>sn>s, into P(n+(ks))P_{\ell(n+(k-s))}. Hence, we also have (n)(n+p)\ell(n)\leqslant\ell(n+p) for all n>sn>s.

Case 2. j<sj<s. We can see that φ\varphi necessarily maps each P(n)P_{\ell(n)}, nsn\geqslant s, into P(n+(kj))P_{\ell(n+(k-j))}. It follows that (n)(n+p)\ell(n)\leqslant\ell(n+p) for all nsn\geqslant s.

In both cases, we see that statement 5 holds for the chosen p=kjp=k-j. This completes the proof. ∎

x0x_{0}x1x_{1}x2x_{2}x3x_{3}x4x_{4}x0x_{0}x1x_{1}x2x_{2}x3x_{3}x4x_{4}(a)(b)
Figure 4: Two combs CC^{\ell} such that (a) (n)=n\ell(n)=n, and (b) (1)=3\ell(1)=3 and (n)=2\ell(n)=2 for n>1n>1. For each of the two combs, the red subgraph illustrates the graph image of its nontrivial self-embedding.
Example 22.

If (n)=n\ell(n)=n, then CC^{\ell} satisfies s(s)1s\geqslant\ell(s)-1 (with s=2s=2) as well as statement 5 of Theorem 21 for any choice of p1p\geqslant 1. As such, CC^{\ell} is strongly self-embeddable via a positive translation. Figure 4(a) shows a positive translation of CC^{\ell} by 11. In addition, we have that (C,nK2)\mathcal{R}(C^{\ell},nK_{2}) is empty for all n2n\geqslant 2 by Theorem 21.

Example 23.

Suppose that (1)=3\ell(1)=3 and (n)=2\ell(n)=2 for n>1n>1. We have s<(s)1s<\ell(s)-1 (with s=1s=1). So we define, using formula (1) preceding Theorem 21, a function \ell^{*} such that

(n)={1,n=12,n>1.\ell^{*}(n)=\begin{cases}1,&n=1\\ 2,&n>1.\end{cases}

The comb CC^{\ell^{*}} is illustrated in Figure 4(b) as a red subgraph of CC^{\ell}. We see that (C,nK2)\mathcal{R}(C^{\ell^{*}},nK_{2}) is always empty since CC^{\ell^{*}} satisfies statement 5 of Theorem 21. By the fact that CCC^{\ell}\cong C^{\ell^{*}}, we have that (C,nK2)\mathcal{R}(C^{\ell^{*}},nK_{2}) is always empty as well.

Example 24.

If (1)=2\ell(1)=2 and (n)=1\ell(n)=1 for n>1n>1, then s(s)1s\geqslant\ell(s)-1 (with s=1s=1). However, statement 5 does not hold since (1)>(1+p)\ell(1)>\ell(1+p) for all p1p\geqslant 1. This implies that CC^{\ell} is not self-embeddable and (C,nK2)\mathcal{R}(C^{\ell},nK_{2}) is nonempty for all n2n\geqslant 2. From Theorem 16(i), we can infer that nC(C,nK2)nC^{\ell}\in\mathcal{R}(C^{\ell},nK_{2}) in this case.

6 Concluding remarks

Problem 1, in its full generality, is quite a challenging problem to attack. For this reason, we chose to devote a significant part of this study to the particular case where HH is a matching. Even then, we were not able to completely answer Problem 1. While Theorem 16 managed to get us closer, we still have the following problem involving (G,nK2)\mathcal{R}(G,nK_{2}).

Problem 25.

Is there a sufficient and necessary condition for GG under which (G,nK2)\mathcal{R}(G,nK_{2}) is empty for all n2n\geqslant 2?

Further studies can also be done on other specific cases of the pair (G,H)(G,H). Of course, another avenue of research would be to consider multi-color Ramsey-minimal infinite graphs, as done in [8] for finite graphs.

While the compactness results of Section 3 are interesting in their own right, our only application of them was to eliminate consideration of the case where both graphs GG, HH are finite (Theorem 5). We believe that these results, and other compactness results, could prove extremely useful in future studies of Ramsey-type problems for infinite graphs.

Apart from Section 3, we have only considered countable graphs in this paper. It would be interesting and worthwhile to study (G,H)\mathcal{R}(G,H) when at least one of GG, HH is uncountable. Presumably, this problem is significantly harder, and one would have to consider set-theoretic concerns.

Acknowledgements

We would like to thank Fawwaz Fakhrurrozi Hadiputra for addressing Theorem 12 in the beginning of our study, which inspired us to formulate Theorems 10 and 11. Also, we would like to thank the anonymous referees for their valuable feedback and comments on this paper.

References

  • [1] A. F. Beardon, Magic labellings of infinite graphs, Australas. J. Combin. 30 (2004), 117–132.
  • [2] S. Binns, B. Kjos-Hanssen, M. Lerman, J. H. Schmerl, and R. Solomon, Self-embeddings of computable trees, Notre Dame J. Form. Log. 49 (2008), no. 1, 1–37.
  • [3] A. Bonato and C. Tardif, Mutually embeddable graphs and the tree alternative conjecture, J. Combin. Theory Ser. B 96 (2006), 874–880.
  • [4] S. A. Burr, P. Erdős, R. J. Faudree, C. C. Rousseau, and R. H. Schelp, Ramsey minimal graphs for matchings, The theory and applications of graphs (Kalamazoo, MI, 1980), Wiley, New York, 1981, pp. 159–168.
  • [5] S. A. Burr, P. Erdős, and L. Lovász, On graphs of Ramsey type, Ars Combin. 1 (1976), no. 1, 167–190.
  • [6] J. Cáceres, C. Hernando, M. Mora, I. M. Pelayo, and M. L. Puertas, On the metric dimension of infinite graphs, Discrete Appl. Math. 160 (2012), no. 18, 2618–2626.
  • [7] R. Diestel, Graph theory, 5th ed., Springer-Verlag, Berlin, 2017.
  • [8] J. Fox, A. Grinshpun, A. Liebenau, Y. Person, and T. Szabó, On the minimum degree of minimal Ramsey graphs for multiple colors, J. Combin. Theory Ser. B 120 (2016), 64–82.
  • [9] R. Halin, Automorphisms and endomorphisms of infinite locally finite graphs, Abh. Math. Semin. Univ. Hambg. 39 (1973), no. 1, 251–283.
  • [10] M. Hamann, Self-embeddings of trees, Discrete Math. 342 (2019), no. 12, 111586.
  • [11] M. Schaefer, Graph Ramsey theory and the polynomial hierarchy, J. Comput. System Sci. 62 (2001), no. 2, 290–322.
  • [12] M. H. Shekarriz and M. Mirzavaziri, Self-contained graphs, preprint (2015), arXiv: 1503.00139.
  • [13] M. Stein, Extremal infinite graph theory, Discrete Math. 311 (2011), no. 15, 1472–1496.