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

Induced subgraphs and tree decompositions
XV. Even-hole-free graphs with bounded clique number have logarithmic treewidth

Maria Chudnovsky Peter Gartland Sepehr Hajebi § Daniel Lokshtanov  and  Sophie Spirkl§∥
Abstract.

We prove that for every integer t1t\geq 1 there exists an integer ct1c_{t}\geq 1 such that every nn-vertex even-hole-free graph with no clique of size tt has treewidth at most ctlognc_{t}\log{n}. This resolves a conjecture of Sintiari and Trotignon, who also proved that the logarithmic bound is asymptotically best possible. It follows that several NP-hard problems such as Stable Set, Vertex Cover, Dominating Set and Coloring admit polynomial-time algorithms on this class of graphs. As a consequence, for every positive integer rr, rr-Coloring can be solved in polynomial time on even-hole-free graphs without any assumptions on clique size.

As part of the proof, we show that there is an integer dd such that every even-hole-free graph has a balanced separator which is contained in the (closed) neighborhood of at most dd vertices. This is of independent interest; for instance, it implies the existence of efficient approximation algorithms for certain NP-hard problems while restricted to the class of all even-hole-free graphs.

Department of Computer Science, University of California Santa Barbara, Santa Barbara, CA, USA. Supported by NSF grant CCF-2008838.
§Department of Combinatorics and Optimization, University of Waterloo, Waterloo, Ontario, Canada
Princeton University, Princeton, NJ, USA. Supported by NSF-EPSRC Grant DMS-2120644 and by AFOSR grant FA9550-22-1-0083.
We acknowledge the support of the Natural Sciences and Engineering Research Council of Canada (NSERC), [funding reference number RGPIN-2020-03912]. Cette recherche a été financée par le Conseil de recherches en sciences naturelles et en génie du Canada (CRSNG), [numéro de référence RGPIN-2020-03912]. This project was funded in part by the Government of Ontario. This research was completed while Spirkl was an Alfred P. Sloan fellow.

1. Introduction

All graphs in this paper are finite and simple. For a graph GG, we denote by V(G)V(G) the vertex set of GG, and by E(G)E(G) the edge set of GG. For graphs H,GH,G, we say that HH is an induced subgraph of GG if V(H)V(G)V(H)\subseteq V(G), and x,yV(H)x,y\in V(H) are adjacent in HH if and only if xyE(G)xy\in E(G). A hole in a graph is an induced cycle with at least four vertices. The length of a hole is the number of vertices in it. A hole is even it it has even length, and odd otherwise. The class of even-hole-free graphs has attracted much attention in the past due to its somewhat tractable, yet very rich structure (see the survey [37]). In addition to stuctural results, much effort was put into designing efficient algorithms for even-hole-free graphs (to solve problems that are NP-hard in general). This is discussed in [37], while [1, 12, 15, 28] provide examples of more recent work. Nevertheless, many open questions remain. Among them is the complexity of several algorithmic problems: Stable Set, Vertex Cover, Dominating Set, kk-Coloring and Coloring.

For a graph G=(V(G),E(G))G=(V(G),E(G)), a tree decomposition (T,χ)(T,\chi) of GG consists of a tree TT and a map χ:V(T)2V(G)\chi:V(T)\to 2^{V(G)} with the following properties:

  1. (i)

    For every vertex vV(G)v\in V(G), there exists tV(T)t\in V(T) such that vχ(t)v\in\chi(t).

  2. (ii)

    For every edge v1v2E(G)v_{1}v_{2}\in E(G), there exists tV(T)t\in V(T) such that v1,v2χ(t)v_{1},v_{2}\in\chi(t).

  3. (iii)

    For every vertex vV(G)v\in V(G), the subgraph of TT induced by {tV(T)vχ(t)}\{t\in V(T)\mid v\in\chi(t)\} is connected.

For each tV(T)t\in V(T), we refer to χ(t)\chi(t) as a bag of (T,χ)(T,\chi). The width of a tree decomposition (T,χ)(T,\chi), denoted by width(T,χ)\operatorname{width}(T,\chi), is maxtV(T)|χ(t)|1\max_{t\in V(T)}|\chi(t)|-1. The treewidth of GG, denoted by tw(G)\operatorname{tw}(G), is the minimum width of a tree decomposition of GG.

Treewidth, first introduced by Robertson and Seymour in the context of graph minors, is an extensively studied graph parameter, mostly due to the fact that graphs of bounded treewidth exhibit interesting structural [35] and algorithmic [11] properties. It is easy to see that large complete graphs and large complete bipartite graphs have large treewidth, but there are others (in particular a subdivided k×kk\times k-wall, which is a planar graph with maximum degree three, and which we will not define here). A theorem of [35] characterizes precisely excluding which graphs as minors (and in fact as subgraphs) results in a class of bounded treewidth.

Bringing tree decompositions into the world of induced subgraphs is a relatively recent endeavor. It began in [36], where the authors observed yet another evidence of the structural complexity of the class of even-hole-free graphs: the fact that there exist even-hole-free graphs of arbitrarily large treewidth (even when large complete subgraphs are excluded). Closer examination of their constructions led the authors of [36] to make the following two conjectures (the diamond is the unique simple graph with four vertices and five edges):

Conjecture 1.1 (Sintiari and Trotignon [36]).

For every integer t1t\geq 1, there exists a constant ctc_{t} such that every even-hole-free graph GG with no induced diamond and no clique of size tt satisfies tw(G)ct\operatorname{tw}(G)\leq c_{t}.

Conjecture 1.2 (Sintiari and Trotignon [36]).

For every integer t1t\geq 1, there exists a constant CtC_{t} such that every even-hole-free graph GG with no clique of size tt satisfies tw(G)Ctlog|V(G)|\operatorname{tw}(G)\leq C_{t}\log|V(G)|.

Conjecture 1.1 was resolved in [8]. Here we prove Conjecture 1.2, thus closing the first line of research on induced subgraphs and tree decompositions.

We remark that the construction of [36] shows that the logarithmic bound of Conjecture 1.2 is asymptotically best possible. Furthermore, our main result implies that many algorithmic problems that are NP-hard in general (among them that Stable Set, Vertex Cover, Dominating Set and Coloring) admit polynomial-time algorithms in the class of even-hole-free graphs with bounded clique number. As a consequence, for every positive integer rr, rr-Coloring can be solved in polynomial time on even-hole-free graphs without any assumptions on clique size.

Before we proceed, we introduce some notation and definitions. Let G=(V(G),E(G))G=(V(G),E(G)) be a graph. For a set XV(G)X\subseteq V(G), we denote by G[X]G[X] the subgraph of GG induced by XX. For XV(G)X\subseteq V(G), we denote by GXG\setminus X the subgraph induced by V(G)XV(G)\setminus X. In this paper, we use induced subgraphs and their vertex sets interchangeably.

For graphs GG and HH, we say that GG contains HH if some induced subgraph of GG is isomorphic to HH. For a family \mathcal{H} of graphs, GG contains \mathcal{H} if GG contains a member of \mathcal{H}, and we say that GG is \mathcal{H}-free if GG does not contain \mathcal{H}. A clique in a graph is a set of pairwise adjacent vertices, and a stable set is a set of vertices no two of which are adjacent. Let vV(G)v\in V(G). The open neighborhood of vv, denoted by N(v)N(v), is the set of all vertices in V(G)V(G) adjacent to vv. The closed neighborhood of vv, denoted by N[v]N[v], is N(v){v}N(v)\cup\{v\}. Let XV(G)X\subseteq V(G). The open neighborhood of XX, denoted by N(X)N(X), is the set of all vertices in V(G)XV(G)\setminus X with at least one neighbor in XX. The closed neighborhood of XX, denoted by N[X]N[X], is N(X)XN(X)\cup X. If HH is an induced subgraph of GG and XV(G)X\subseteq V(G), then NH(X)=N(X)HN_{H}(X)=N(X)\cap H and NH[X]=NH(X)(XH)N_{H}[X]=N_{H}(X)\cup(X\cap H). Let YV(G)Y\subseteq V(G) be disjoint from XX. We say XX is complete to YY if all possible edges with an end in XX and an end in YY are present in GG, and XX is anticomplete to YY if there are no edges between XX and YY.

Given a graph GG, a path in GG is an induced subgraph of GG that is a path. If PP is a path in GG, we write P=p1--pkP=p_{1}\hbox{-}\cdots\hbox{-}p_{k} to mean that V(P)={p1,,pk}V(P)=\{p_{1},\dots,p_{k}\}, and pip_{i} is adjacent to pjp_{j} if and only if |ij|=1|i-j|=1. We call the vertices p1p_{1} and pkp_{k} the ends of PP, and say that PP is a path from p1p_{1} to pkp_{k}. The interior of PP, denoted by PP^{*}, is the set V(P){p1,pk}V(P)\setminus\{p_{1},p_{k}\}. The length of a path PP is the number of edges in PP. We denote by CkC_{k} a cycle with kk vertices.

Next, we describe a few types of graphs that we will need (see Figure 1).

Refer to caption
Figure 1. Theta, prism and an even wheel. Dashed lines represent paths of length at least one.

A theta is a graph consisting of three internally vertex-disjoint paths P1=a--bP_{1}=a\hbox{-}\cdots\hbox{-}b, P2=a--bP_{2}=a\hbox{-}\cdots\hbox{-}b, and P3=a--bP_{3}=a\hbox{-}\cdots\hbox{-}b, each of length at least 2, such that P1,P2,P3P_{1}^{*},P_{2}^{*},P_{3}^{*} are pairwise anticomplete. In this case we call aa and bb the ends of the theta.

A prism is a graph consisting of three vertex-disjoint paths P1=a1--b1P_{1}=a_{1}\hbox{-}\cdots\hbox{-}b_{1}, P2=a2--b2P_{2}=a_{2}\hbox{-}\cdots\hbox{-}b_{2}, and P3=a3--b3P_{3}=a_{3}\hbox{-}\cdots\hbox{-}b_{3}, each of length at least 1, such that a1a2a3a_{1}a_{2}a_{3} and b1b2b3b_{1}b_{2}b_{3} are triangles, and no edges exist between the paths except those of the two triangles.

A pyramid is a graph consisting of three paths P1=a--b1P_{1}=a\hbox{-}\cdots\hbox{-}b_{1}, P2=a--b2P_{2}=a\hbox{-}\cdots\hbox{-}b_{2}, and P3=a--b3P_{3}=a\hbox{-}\cdots\hbox{-}b_{3} such that P1{a},P2{a},P3{a}P_{1}\setminus\{a\},P_{2}\setminus\{a\},P_{3}\setminus\{a\} are pairwise disjoint, at least two of the three paths P1,P2,P3P_{1},P_{2},P_{3} have length at least two, b1b2b3b_{1}b_{2}b_{3} is triangle (called the base of the pyramid), and no edges exist between P1{a},P2{a},P3{a}P_{1}\setminus\{a\},P_{2}\setminus\{a\},P_{3}\setminus\{a\} except those of the triangle b1b2b3b_{1}b_{2}b_{3}. The vertex aa is called the apex of the pyramid.

A wheel (H,x)(H,x) in GG is a pair where HH is a hole and xx is a vertex with at least three neighbors in HH. A wheel (H,x)(H,x) is even if xx has an even number of neighbors on HH.

Let 𝒞\mathcal{C} be the class of (C4C_{4}, theta, prism, even wheel)-free graphs (sometimes called “C4C_{4}-free odd-signable” graphs). For every integer t1t\geq 1, let 𝒞t\mathcal{C}_{t} be the class of all graphs in 𝒞\mathcal{C} with no clique of size tt. It is easy to see that every even-hole-free graph is in 𝒞\mathcal{C}.

The reader may be familiar with [3] where a special case of Conjecture  1.2 was proved; moreover, only one Lemma of [3] uses the fact that the set-up there is not the most general case. At the time, the authors of [3] thought that the full proof of Conjecture  1.2 would follow the general outline of [3], fixing the one missing lemma. That is not what happened. The proof in the present paper takes a different path: while it relies on insights and a general understanding of the class of even-hole-free graphs gained in [3], and uses several theorems proved there, a significant detour is needed.

The first part of the paper is not concerned with treewidth at all. Instead, we focus on the following question: given two non-adjacent vertices in a graph in 𝒞\mathcal{C} of bounded clique number, how many internally vertex-disjoint paths can there be between them? Given that if instead of “internally vertex-disjoint” we say “with pairwise anticomplete interiors”, then the answer is “two”, this is somewhat related to the recent work on the induced Menger theorem [21, 25, 31]. Let k1k\geq 1 be an integer and let a,bV(G)a,b\in V(G). We say that abab is a kk-banana if aa is non-adjacent to bb, and there exist kk pairwise internally-vertex-disjoint paths from aa to bb. Note that if abab is a kk-banana in GG, then abab is an ll-banana in GG for every lkl\leq k. We prove:

Theorem 1.3.

For every integer t1t\geq 1, there exists a constant ctc_{t} such that if G𝒞tG\in\mathcal{C}_{t}, then GG contains no ctlog|V(G)|c_{t}\log|V(G)|-banana.

The next step in the proof of Conjecture 1.2 is the following. Let c[0,1]c\in[0,1] and let ww be a normal weight function on GG, that is, w:V(G)0w:V(G)\rightarrow\mathbb{R}_{\geq 0} satisfies vV(G)w(v)=1\sum_{v\in V(G)}w(v)=1. A set XV(G)X\subseteq V(G) is a (w,c)(w,c)-balanced separator if w(D)cw(D)\leq c for every component DD of GXG\setminus X. The set XX is a ww-balanced separator if XX is a (w,12)(w,\frac{1}{2})-balanced separator. We show:

Theorem 1.4.

There is an integer dd with the following property. Let G𝒞G\in\mathcal{C} and let ww be a normal weight function on GG. Then there exists YV(G)Y\subseteq V(G) such that

  • |Y|d|Y|\leq d, and

  • N[Y]N[Y] is a ww-balanced separator in GG.

We then use Theorem 1.3 and Theorem 1.4 to prove our main result:

Theorem 1.5.

For every integer t1t\geq 1, there exists a constant ctc_{t} such that every G𝒞tG\in\mathcal{C}_{t} satisfies tw(G)ctlog|V(G)|\operatorname{tw}(G)\leq c_{t}\log|V(G)|.

1.1. Proof outline and organization

We only include a few definitions in this section; all the necessary definitions appear in later parts of the paper. Our first goal is to prove Theorem 1.3. Let a,bV(G)a,b\in V(G) be non-adjacent. Recall that a separation of a graph GG is a triple (X,Y,Z)(X,Y,Z) of pairwise disjoint subsets of GG with XYZ=GX\cup Y\cup Z=G such that XX is anticomplete to ZZ. Similarly to [6], we use the fact that graphs in 𝒞t\mathcal{C}_{t} admit a natural family of separations that correspond to special vertices of the graph called “hubs” and are discussed in Section 3. Unfortunately, the interactions between these separations may be complex, and, similarly to [5], we use degeneracy to partition the set of all hubs other than a,ba,b (which yields a partition of all the natural separations) of an nn-vertex graph GG in 𝒞t\mathcal{C}_{t} into collections S1,,SpS_{1},\dots,S_{p}, where each SiS_{i} is “non-crossing” (this property is captured in Lemma 4.9), pC(t)lognp\leq C(t)\log n (where C(t)C(t) only depends on tt and works for all G𝒞tG\in\mathcal{C}_{t}) and each vertex of SiS_{i} has at most dd (where dd depends on tt) neighbors in j=ipSj\bigcup_{j=i}^{p}S_{j}. We prove a strengthening of Theorem 1.3, which asserts that the size of the largest abab-banana is bounded by a linear function of pp.

We observe that a result of [3] implies that there exists a an integer kk (that depends on tt) such that if G𝒞tG\in\mathcal{C}_{t} has no hubs other than a,ba,b, then abab is not a kk-banana in GG. More precisely, the result of [3] states that if aa and bb are joined by kk internally-vertex-disjoint paths P1,,PkP_{1},\dots,P_{k}, then for at least one i{1,,k}i\in\{1,\dots,k\}, the neighbor of aa in PiP_{i} is a hub.

We now proceed as follows. Let m=2d+km=2d+k (where dd comes from the degeneracy partition and kk from the previous paragraph); suppose that abab is a (4m+2)(m1)(4m+2)(m-1)-banana in GG and let 𝒫\mathcal{P} be the set of all paths of GG with ends a,ba,b.

Let S1,,SpS_{1},\dots,S_{p} denote the partition of hubs as decribed above. We proceed by induction on pp and describe a process below that finds an induced subgraph HH of GG in which:

  • Vertices in S1S_{1} are no longer hubs;

  • If there are not many internally disjoint abab-paths in HH, we can “lift” this to GG.

We first consider a so-called mm-lean tree decomposition (T,χ)(T,\chi) of GG (discussed in Section 2). By examining the intersection graphs of subtrees of a tree we deduce that there exist two (not necessarily distinct) vertices t1,t2V(T)t_{1},t_{2}\in V(T) such that for every abab path PP, the interior of PP meets χ(t1)χ(t2)\chi(t_{1})\cup\chi(t_{2}). We also argue that a,bχ(t1)χ(t2)a,b\in\chi(t_{1})\cup\chi(t_{2}). A vertex uu of S1S_{1} is bad if uu has large degree (at least D=2m(m1)D=2m(m-1)) in the torso of χ(t1)\chi(t_{1}), or uu has large degree in the torso of χ(t2)\chi(t_{2}), or uu is adjacent to both aa and bb. We show that there are at most three bad vertices in S1S_{1}.

We would like to bound the size of the largest abab-banana in the union of the two torsos. Unfortunately, the torso of χ(ti)\chi(t_{i}) may not be a graph in 𝒞t\mathcal{C}_{t}. Instead, we find an induced subgraph of GG, which we call β\beta, that consists of χ(t1)χ(t2)\chi(t_{1})\cup\chi(t_{2}) together with a collection of disjoint vertex sets Conn(ti,t)\operatorname{Conn}(t_{i},t) for tV(T){t1,t2}t\in V(T)\setminus\{t_{1},t_{2}\} and i{1,2}i\in\{1,2\} except vertices tt on the t1-t2t_{1}\hbox{-}t_{2}-path in TT, where each Conn(ti,t)\operatorname{Conn}(t_{i},t) “remembers” the component of G(χ(t1)χ(t2))G\setminus(\chi(t_{1})\cup\chi(t_{2})) that meets χ(t)\chi(t). Importantly, no vertex of β(χ(t1)χ(t2))\beta\setminus(\chi(t_{1})\cup\chi(t_{2})) is a hub of β\beta, and all but at most three vertices of S1S_{1} have bounded degree in the union of the torso of χ(t1)\chi(t_{1}) and the torso of χ(t2)\chi(t_{2}).

We next decompose β\beta, simultaneously, by all the separations corresponding to the hubs in S1S_{1} that are not bad, and delete the set of (at most three) bad vertices of S1S_{1}. We denote the resulting graph by βA(S1)\beta^{A}(S_{1}) and call it the “central bag” for S1S_{1}. The parameter pp is smaller for βA(S1)\beta^{A}(S_{1}) than it is for GG, and so we can use induction to obtain a bound on the largest size of an abab-banana in βA(S1)\beta^{A}(S_{1}). Since our goal is to bound the size of an abab-banana in GG by a linear function of pp, we now need to show that the size of the largest abab-banana does not grow by more than an additive constant when we move from βA(S1)\beta^{A}(S_{1}) to GG.

We start with a smallest subset XX of βA(S1)\beta^{A}(S_{1}) that separates aa from bb in βA(S1)\beta^{A}(S_{1}) (and whose size is bounded by Menger’s theorem) and show how to transform in into a set YY separating aa from bb in β\beta, while increasing the size of its intersection with χ(t1)χ(t2)\chi(t_{1})\cup\chi(t_{2}) by at most an additive constant, and ensuring the number of sets Conn(ti,t)\operatorname{Conn}(t_{i},t) that YY meets is bounded by a constant. Then we repeat a similar procedure to obtain a set ZZ of vertices separating aa from bb in GG, making sure that the increase in size is again bounded by an additive constant.

Let us now discuss how we obtain the bound on the growth of the separator. In the first step, to obtain YY, we add to XX the neighbor sets of the vertices of S1XS_{1}\cap X. Since while constructing βA(S1)\beta^{A}(S_{1}) we have deleted all the bad vertices of S1S_{1}, the number of vertices of χ(t1)χ(t2)\chi(t_{1})\cup\chi(t_{2}) added to XX is at most 2|S1X|D2|S_{1}\cap X|D, and YXY\setminus X meets at most 2|S1X|D2|S_{1}\cap X|D of the sets Conn(ti,t)\operatorname{Conn}(t_{i},t). Note that the bound on the size of XX that we have depends on pp, which may be close to log|V(G)|\log|V(G)|, so another argument is needed to obtain a constant bound on |S1X||S_{1}\cap X| and on the number of sets Conn(ti,t)\operatorname{Conn}(t_{i},t) that meet YY. This is a consequence of Theorem 6.3, because no vertex of S1S_{1} is a hub in βA(S1)\beta^{A}(S_{1}) (this is proved in Theorem 4.10), and no vertex of Conn(ti,t)\operatorname{Conn}(t_{i},t) is a hub in β\beta. (The proof of Theorem 6.3 analyzes the structure of minimal separators in graphs in 𝒞\mathcal{C} using tools developed in Section 5.)

In the second growing step, we start with the set YY obtained in the previous paragraph. Then we add to YY the following subsets (here we describe the cases when t1t2t_{1}\neq t_{2}; if t1=t2t_{1}=t_{2} the argument is similar).

  1. (1)

    χ(t1)χ(t1)\chi(t_{1})\cap\chi(t_{1}^{\prime}) where t1t_{1}^{\prime} is the unique neighbor of t1t_{1} in the path in TT from t1t_{1} to t2t_{2}.

  2. (2)

    χ(t2)χ(t2)\chi(t_{2})\cap\chi(t_{2}^{\prime}) where t2t_{2}^{\prime} is the unique neighbor of t2t_{2} in the path in TT from t1t_{1} to t2t_{2}.

  3. (3)

    χ(t1)χ(t)\chi(t_{1})\cap\chi(t) for every tN(t1){t1}t\in N(t_{1})\setminus\{t_{1}^{\prime}\} such that YConn(t1,t)Y\cap\operatorname{Conn}(t_{1},t)\neq\emptyset.

  4. (4)

    χ(t2)χ(t)\chi(t_{2})\cap\chi(t) for every tN(t2){t2}t\in N(t_{2})\setminus\{t_{2}^{\prime}\} such that YConn(t2,t)Y\cap\operatorname{Conn}(t_{2},t)\neq\emptyset.

  5. (5)

    The set of all bad vertices of S1S_{1}.

One of the properties of mm-lean tree decompositions is that the size of each adhesion (intersection of “neighboring” bags) is less than mm. The number of adhesions added to YY is again bounded since the number of distinct sets Conn(ti,t)\operatorname{Conn}(t_{i},t) that meet YY is bounded. This completes the proof of Theorem 1.3.

The next key ingredient in the proof of Theorem 1.5 is Theorem 1.4, asserting that there is an integer CC such that for every graph G𝒞G\in\mathcal{C} and every normal weight function ww on GG, there is a ww-balanced separator XX in GG such that XX is contained in the union of the neighborhoods of at most CC vertices of GG. To prove that, we first prove a decomposition theorem, similar to the one giving us the natural separations associated with hubs, but this time the separations come from pyramids in GG. We then use the two kinds of separations (the kind that come from hubs and the kind that come from pyramids) as follows.

For XV(G)X\subseteq V(G) we say that a set YV(G)Y\subseteq V(G) dominates XX if XN[Y]X\subseteq N[Y]. By Lemma 8.5 (from [14]) there is a path P=p1--pkP=p_{1}\hbox{-}\cdots\hbox{-}p_{k} in GG such that w(D)12w(D)\leq\frac{1}{2} for every component DD of GN[P]G\setminus N[P]. We now use a “sliding window” argument: we divide PP into subpaths, each with the property that its neighborhood can be dominated by a small, but not too small, set of vertices. Using the decomposition theorems above, we find a bounded-size set X(W)X(W) such that, except for our window WW, the path PP is disjoint from N[X(W)]N[X(W)], and N[X(W)]N[X(W)] separates the subpath of PP before our current window from that after our current window. This means (unless N[X]N[X] is a balanced separator) that the big component of GN[X]G\setminus N[X] does not contain either the subpath of PP before our window, or the subpath after our window. By looking at the point in the path where this answer changes from “before” to ”after” (and showing that such a point exists), and by combining the separators for the two windows before and after this point as well as small dominating sets for the neighborhood of those two windows, we are able to find a ww-balanced separator with bounded domination number. This completes the proof of Theorem 1.4. We remark that Theorem 1.4 applies to all graphs in 𝒞\mathcal{C} and does not assume a bound on the clique number.

The next step in the proof of Theorem 1.5 is to prove Theorem 9.2, asserting that for every integer LL, if a graph GG contains no LL-banana, then for every normal weight function ww on GG, if GG has a ww-balanced separator N[X]N[X], then GG has an ww-balanced separator YY and a clique TXT\subseteq X such that |YT|3|X|L|Y\setminus T|\leq 3|X|L. This step uses 3L3L-lean tree decompositions of GG and works for all graphs GG.

Now Theorem 9.2 and Theorem 1.3 together imply that for every tt, there exists an integer qq, depending on tt, such that for every G𝒞tG\in\mathcal{C}_{t}, GG has a balanced separator of size at most qlog|V(G)|+tq\log|V(G)|+t; that is Theorem 9.1. By Lemma 10.1 this immediately implies the desired bound on the treewidth of GG.

The paper is organized as follows. In Section 2, we discuss lean tree decompositions and their properties, along with other classical results in graph theory. For some of them we prove variations tailored specifically to our needs. In Section 3 we summarize results guaranteeing the existence of separations associated with hubs. We also establish a stronger version of Theorem 1.3 for the case when the set of hubs in GG is very restricted. In Section 4 we discuss the construction of the graphs β\beta and βA(S1)\beta^{A}(S_{1}), and how to use abab-separators there to obtains abab-separators in GG. In Section 5 we analyze the structure of minimal separators in graphs of 𝒞\mathcal{C}. In Section 6 we use the results of Section 5 to obtain a bound on the size of a stable set of non-hubs in an abab-separator of smallest size. Section 7 puts together the results of all the previous sections to prove Theorem 1.3. The goal of Section 8 is to prove Theorem 1.4. We start with Lemmas 8.2 and 8.3 to establish the existence of certain decompositions in graphs of 𝒞\mathcal{C}, and then proceed with the sliding window argument. Section 9 is devoted to the proof of Theorem 9.1. The proof of Theorem 1.5 is completed in Section 10. Finally, Section 11 discusses algorithmic consequences of Theorem 1.4 and Theorem 1.5.

2. Special tree decompositions and connectivity

In this section we have collected several results from the literature that we need; for some of them we have also proved our own versions, tailored specifically to our needs. Along the way we also introduce some notation.

2.1. Connectivity

We start with a result on connectivity. For two vertices u,vGu,v\in G and a set XG{u,v}X\subseteq G\setminus\{u,v\} we say that XX separates uu from vv if PXP^{*}\cap X\neq\emptyset for every path PP of GG with ends uu and vv. The following is a well-known variant of a classical theorem due to Menger [30]:

Theorem 2.1 (Menger [30]).

Let k1k\geq 1 be an integer, let GG be a graph and let u,vGu,v\in G be distinct and non-adjacent. Then either there exists a set MG{u,v}M\subseteq G\setminus\{u,v\} with |M|<k|M|<k such that MM separates uu and vv in GG, or uvuv is a kk-banana in GG.

2.2. Lean tree decompositions

We adopt notation from [3]: For a tree TT and vertices t,tV(T)t,t^{\prime}\in V(T), we denote by tTttTt^{\prime} the unique path in TT from tt to tt^{\prime}. Let (T,χ)(T,\chi) be a tree decomposition of a graph GG. For every uvE(T)uv\in E(T), the adhesion at uvuv, denoted by adh(u,v)\operatorname{adh}(u,v), is the set χ(u)χ(v)\chi(u)\cap\chi(v). For u,vTu,v\in T such that uvE(T)uv\not\in E(T) (in particular, if u=vu=v), we set adh(u,v)=\operatorname{adh}(u,v)=\emptyset. We define adh(T,χ)=maxuvE(T)|adh(u,v)|\operatorname{adh}(T,\chi)=\max_{uv\in E(T)}|\operatorname{adh}(u,v)|. For every xV(T)x\in V(T), the torso at xx, denoted by χ^(x)\hat{\chi}(x), is the graph obtained from the bag χ(x)\chi(x) by, for each yNT(x)y\in N_{T}(x), adding an edge between every two non-adjacent vertices u,vadh(x,y)u,v\in\operatorname{adh}(x,y).

In the proof of Theorem 1.3 and Theorem 1.5, we will use a special kind of a tree decomposition that we present next. Let k>0k>0 be an integer. A tree decomposition (T,χ)(T,\chi) is called kk-lean if the following hold:

  • adh(T,χ)<k\operatorname{adh}(T,\chi)<k; and

  • for all t,tV(T)t,t^{\prime}\in V(T) and sets Zχ(t)Z\subseteq\chi(t) and Zχ(t)Z^{\prime}\subseteq\chi(t^{\prime}) with |Z|=|Z|k|Z|=|Z^{\prime}|\leq k, either GG contains |Z||Z| disjoint paths from ZZ to ZZ^{\prime}, or some edge ssss^{\prime} of tTttTt^{\prime} satisfies |adh(s,s)|<|Z||\operatorname{adh}(s,s^{\prime})|<|Z|.

For a tree TT and an edge tttt^{\prime} of TT, we denote by TttT_{t\rightarrow t^{\prime}} the component of TtT\setminus t containing tt^{\prime}. Let Gtt=G[vTttχ(v)]G_{t\rightarrow t^{\prime}}=G[\bigcup_{v\in T_{t\rightarrow t^{\prime}}}\chi(v)]. A tree decomposition (T,χ)(T,\chi) is tight if for every edge ttE(T)tt^{\prime}\in E(T) there is a component DD of Gttχ(t)G_{t\rightarrow t^{\prime}}\setminus\chi(t) such that χ(t)χ(t)N(D)\chi(t)\cap\chi(t^{\prime})\subseteq N(D) (and therefore χ(t)χ(t)=N(D)\chi(t)\cap\chi(t^{\prime})=N(D)).

Next, we need a definition from [9]. Given a tree decomposition (T,χ)(T,\chi) of an nn-vertex graph GG, its fatness is the vector (an,,a0)(a_{n},\dots,a_{0}) where aia_{i} denotes the number of bags of TT of size ii. A tree decomposition (T,χ)(T,\chi) of GG is kk-atomic if adh(T,χ)<k\operatorname{adh}(T,\chi)<k and the fatness of (T,χ)(T,\chi) is lexicographically minimum among all tree decompositions of GG with adhesion less than kk.

It was observed in [13] that [9] contains a proof of the following:

Theorem 2.2 (Bellenbaum and Diestel [9], see Carmesin, Diestel, Hamann, Hundertmark [13], see also Weißauer [38]).

Every kk-atomic tree decomposition is kk-lean.

We also need:

Theorem 2.3 (Weißauer [38]).

Every kk-atomic tree decomposition is tight.

Using ideas similar to those of Weißauer [38] and using Theorems 2.1 and 2.2, we prove:

Theorem 2.4.

Let GG be a graph, let k1k\geq 1 and let (T,χ)(T,\chi) be 3k3k-atomic tree decomposition of GG. Let t0V(T)t_{0}\in V(T) and let u,vGu,v\in G. Assume that N[u]N[u] is not separated from χ(t0)u\chi(t_{0})\setminus u by a set of size less than 3k3k, and that N[v]N[v] is not separated from χ(t0)v\chi(t_{0})\setminus v by a set of size less than 3k3k. Then uu is not separated from vv by a set of size less than kk, and consequently uvuv is a kk-banana.

Proof.

By Theorems 2.2 and 2.3, we have that (T,χ)(T,\chi) is tight and 3k3k-lean. Suppose that uu is separated from vv by a set of size less than kk. By Theorem 2.1, there exists a set 𝒫u\mathcal{P}_{u} of pairwise vertex-disjoint (except uu) paths, each with one end uu and the other end in χ(t0)u\chi(t_{0})\setminus u, and such that |𝒫u|=3k|\mathcal{P}_{u}|=3k. Let ZuZ_{u} be the set of the ends of members of 𝒫u\mathcal{P}_{u} in χ(t0)\chi(t_{0}). Similarly, there exists a collection 𝒫v\mathcal{P}_{v} of pairwise vertex-disjoint (except vv) paths, each with one end vv and the other end in χ(t0)v\chi(t_{0})\setminus v, and such that |𝒫v|=3k|\mathcal{P}_{v}|=3k. Let ZvZ_{v} be the set of the ends of members of 𝒫v\mathcal{P}_{v} in χ(t0)\chi(t_{0}). Since (T,χ)(T,\chi) is 3k3k-lean, there is a collection 𝒬\mathcal{Q} of pairwise vertex-disjoint paths from ZuZ_{u} to ZvZ_{v}. Let XX be a set with |X|<k|X|<k separating uu from vv. Then, for every uu-vv path PP in GG we have PXP^{*}\cap X\neq\emptyset. But since |Zu|=|Zv|=|𝒬|=3k|Z_{u}|=|Z_{v}|=|\mathcal{Q}|=3k, there is a path Q𝒬Q\in\mathcal{Q} with ends zuZuz_{u}\in Z_{u} and zvZvz_{v}\in Z_{v}, a path Pu𝒫uP_{u}\in\mathcal{P}_{u} from uu to zuz_{u}, and a path Pv𝒫vP_{v}\in\mathcal{P}_{v} from vv to ZvZ_{v} such that X(Q(Puu)(Pvv))=X\cap(Q\cup(P_{u}\setminus u)\cup(P_{v}\setminus v))=\emptyset, contrary to the fact that u-Pu-zu-Q-zv-Pv-vu\hbox{-}P_{u}\hbox{-}z_{u}\hbox{-}Q\hbox{-}z_{v}\hbox{-}P_{v}\hbox{-}v is a path from uu to vv in GG. This proves the first statement of the theorem. The second statement follows immediately by Theorem 2.1. ∎

We finish this subsection with a theorem about tight tree decompositions in theta-free graphs that was proved in [3]. Note that by Theorem 2.3, the following result applies in particular to kk-atomic tree decompositions for every kk.

A cutset CV(G)C\subseteq V(G) of GG is a (possibly empty) set of vertices such that GCG\setminus C is disconnected. A clique cutset is a cutset that is a clique.

Theorem 2.5 (Abrishami, Alecu, Chudnovsky, Hajebi, Spirkl [3]).

Let GG be a theta-free graph and assume that GG does not admit a clique cutset. Let (T,χ)(T,\chi) be a tight tree decomposition of GG. Then for every edge t1t2t_{1}t_{2} of TT the graph Gt1t2χ(t1)G_{t_{1}\rightarrow t_{2}}\setminus\chi(t_{1}) is connected and N(Gt1t2χ(t1))=χ(t1)χ(t2)N(G_{t_{1}\rightarrow t_{2}}\setminus\chi(t_{1}))=\chi(t_{1})\cap\chi(t_{2}). Moreover, if t0,t1,t2V(T)t_{0},t_{1},t_{2}\in V(T) and t1,t2NT(t0)t_{1},t_{2}\in N_{T}(t_{0}), then χ(t0)χ(t1)χ(t0)χ(t2)\chi(t_{0})\cap\chi(t_{1})\neq\chi(t_{0})\cap\chi(t_{2}).

2.3. Catching a banana

In this subsection we discuss another important feature of tree decompositions that is needed in the proof of Theorem 1.3.

Theorem 2.6.

Let GG be a theta-free graph and let a,bV(G)a,b\in V(G). Let (T,χ)(T,\chi) be a tree decomposition of GG. Then there exist t1,t2V(T)t_{1},t_{2}\in V(T) (not necessarily distinct) such that for every path PP of GG with ends a,ba,b we have that (χ(t1)χ(t2))P.(\chi(t_{1})\cup\chi(t_{2}))\cap P^{*}\neq\emptyset. Moreover, if DD is a component of G(χ(t1)χ(t2))G\setminus(\chi(t_{1})\cup\chi(t_{2})), then |N(D){a,b}|1|N(D)\cap\{a,b\}|\leq 1.

Proof.

Let 𝒫\mathcal{P} be the set of all paths of GG with ends a,ba,b. For every P𝒫P\in\mathcal{P} let T(P)={tT:χ(t)P}T(P)=\{t\in T\ \;:\;\chi(t)\cap P^{*}\neq\emptyset\}. Since PP^{*} is a connected subgraph of GG, it follows that T[T(P)]T[T(P)] is a subtree of TT.

Let HH be a graph with vertex set 𝒫\mathcal{P} and such that P1P2E(H)P_{1}P_{2}\in E(H) if and only if T(P1)T(P2)T(P_{1})\cap T(P_{2})\neq\emptyset. Then, by the main result of [22], HH is a chordal graph.

(1) If P1,P2HP_{1},P_{2}\in H are non-adjacent in HH, then P1P_{1}^{*}, P2P_{2}^{*} are disjoint and anticomplete to each other in GG.

Suppose that there exists vPiPjv\in P_{i}^{*}\cap P_{j}^{*}. Then for every tTt\in T such that vχ(T)v\in\chi(T), we have tT(P1)T(P2)t\in T(P_{1})\cap T(P_{2}), contrary to the fact that P1,P2P_{1},P_{2} are non-adjacent in HH. Next suppose that there is an edge v1v2v_{1}v_{2} of GG such that viPiv_{i}\in P_{i}^{*}. Let tTt\in T be such that v1,v2χ(t)v_{1},v_{2}\in\chi(t). Then tT(P1)T(P2)t\in T(P_{1})\cap T(P_{2}), again contrary to the fact that P1P_{1} is non-adjacent to P2P_{2} in HH. This proves (2.3).

(2) HH has no stable set of size three.

Suppose P1,P2,P3P_{1},P_{2},P_{3} is a stable set in HH. By (2.3) the sets P1,P2,P3P_{1}^{*},P_{2}^{*},P_{3}^{*} are pairwise disjoint and anticomplete to each other in GG. But now P1P2P3P_{1}\cup P_{2}\cup P_{3} is a theta with ends a,ba,b in GG, a contradiction. This proves (2.3).

Since HH is chordal, it follows from (2.3) and a result from [23] that there exist cliques K1,K2K_{1},K_{2} of HH such that V(H)=K1K2V(H)=K_{1}\cup K_{2}. Again by [23] we deduce that for each i{1,2}i\in\{1,2\}, there exists tiKit_{i}\in K_{i} such that tiT(P)t_{i}\in T(P) for every PKiP\in K_{i}. Consequently, (χ(t1)χ(t2))P(\chi(t_{1})\cup\chi(t_{2}))\cap P^{*}\neq\emptyset for every P𝒫P\in\mathcal{P}, and the first statement of the theorem holds.

To prove the second statement, suppose a,bN(D)a,b\in N(D) for some component DD of G(χ(t1)χ(t2))G\setminus(\chi(t_{1})\cup\chi(t_{2})). Then D{a,b}D\cup\{a,b\} is connected, and so there is a path PP from aa to bb with PDP^{*}\subseteq D. But now P(χ(t1)χ(t2))=P^{*}\cap(\chi(t_{1})\cup\chi(t_{2}))=\emptyset, a contradiction. This completes the proof. ∎

2.4. Centers of tree decompositions.

We finish this section with a well-known theorem about tree decompositions. Recall that for a graph GG, a function w:V(G)[0,1]w:V(G)\rightarrow[0,1] is a normal weight function on GG if w(V(G))=1w(V(G))=1, where we use the notation w(X)w(X) to mean xXw(x)\sum_{x\in X}w(x) for a set XX of vertices.

Let GG be a graph and let (T,χ)(T,\chi) be a tree decomposition of GG. Let w:V(G)[0,1]w:V(G)\rightarrow[0,1] be a normal weight function on GG. A vertex t0t_{0} of TT is a center of (T,χ)(T,\chi) if for every tNT(t0)t^{\prime}\in N_{T}(t_{0}) we have w(Gt0tχ(t0))12w(G_{t_{0}\rightarrow t^{\prime}}\setminus\chi(t_{0}))\leq\frac{1}{2}.

The following is well-known; a proof can be found in [3], for example.

Theorem 2.7.

Let (T,χ)(T,\chi) be a tree decomposition of a graph GG. Then (T,χ)(T,\chi) has a center.

3. Wheels and star cutsets

Recall that a wheel in GG is a pair (H,x)(H,x) such that HH is a hole and xx is a vertex that has at least three neighbors in HH. Wheels play an important role in the study of even-hole-free and odd-signable graphs. Graphs in this classes that contain no wheels are much easier to handle; on the other hand, the presence of a wheel forces the graph to admit a certain kind of decomposition.

A sector of (H,x)(H,x) is a path PP of HH whose ends are distinct and adjacent to xx, and such that xx is anticomplete to PP^{*}. A sector PP is a long sector if PP^{*} is non-empty. We now define several types of wheels that we will need.

A wheel (H,x)(H,x) is a universal wheel if xx is complete to HH. A wheel (H,x)(H,x) is a twin wheel if N(x)HN(x)\cap H induces a path of length two. A wheel (H,x)(H,x) is a short pyramid if |N(x)H|=3|N(x)\cap H|=3 and xx has exactly two adjacent neighbors in HH. A wheel is proper if it is not a twin wheel or a short pyramid. We say that xV(G)x\in V(G) is a hub if there exists HH such that (H,x)(H,x) is a proper wheel in GG. We denote by Hub(G)\operatorname{Hub}(G) the set of all hubs of GG.

We need the following result, which was observed in [6]:

Theorem 3.1 (Abrishami, Chudnovsky, Vušković [6]).

Let G𝒞G\in\mathcal{C} and let (H,v)(H,v) be a proper wheel in GG. Then there is no component DD of GN[v]G\setminus N[v] such that HN[D]H\subseteq N[D].

We will revisit this result in Section 8. A star cutset in a graph GG is a cutset SV(G)S\subseteq V(G) such that either S=S=\emptyset or for some xSx\in S, SN[x]S\subseteq N[x]. A large portion of this paper is devoted to dealing with hubs and star cutsets arising from them in graphs in 𝒞\mathcal{C}, but in the remainder of this section we focus on the case when Hub(G)\operatorname{Hub}(G) is very restricted. To do that, we use a result from [3]:

Theorem 3.2 (Abrishami, Alecu, Chudnovsky, Hajebi, Spirkl [3]).

For every integer t1t\geq 1 there exists an integer k=k(t)k=k(t) such that if G𝒞tG\in\mathcal{C}_{t} and xyxy is a kk-banana in GG, then N(x)Hub(G)N(x)\cap\operatorname{Hub}(G)\neq\emptyset and N(y)Hub(G)N(y)\cap\operatorname{Hub}(G)\neq\emptyset.

We immediately deduce:

Theorem 3.3.

For every integer t1t\geq 1, there exists a constant k=k(t)k=k(t) with the following property. Let G𝒞tG\in\mathcal{C}_{t} and let a,bV(G)a,b\in V(G) be non-adjacent. Assume that Hub(G){a,b}Hub(G)\subseteq\{a,b\}. Then abab is not a kk-banana in GG. In particular, if Hub(G)=Hub(G)=\emptyset, there is no kk-banana in GG.

We will also use the following:

Theorem 3.4 (Abrishami, Alecu, Chudnovsky, Hajebi, Spirkl [3]).

For every integer t1t\geq 1, there exists an integer γ=γ(t)\gamma=\gamma(t) such that every G𝒞tG\in\mathcal{C}_{t} with Hub(G)=\operatorname{Hub}(G)=\emptyset satisfies tw(G)γ1\operatorname{tw}(G)\leq\gamma-1.

4. Stable sets of safe hubs

As we discussed in Section 1, in the course of the proof of Theorem 1.3, we will repeatedly decompose the graph by star cutsets arising from a stable set of appropriately chosen hubs (using Theorem 3.1). In this section, we prepare the tools for handling one such step: a stable set of safe hubs.

Throughout this section, we fix the following: Let t,dt,d\in\mathbb{N} and let G𝒞tG\in\mathcal{C}_{t} be a graph with |V(G)|=n|V(G)|=n. Let m=k+2dm=k+2d where k=k(t)k=k(t) is as in Theorem 3.2. Let a,bV(G)a,b\in V(G) such that abab is a (4m+2)(m1)(4m+2)(m-1)-banana in GG, and let 𝒫\mathcal{P} be the set of all paths in GG with ends a,ba,b. Let (T,χ)(T,\chi) be an mm-atomic tree decomposition of GG. By Theorems 2.2 and 2.3, we have that (T,χ)(T,\chi) is tight and mm-lean. By Theorem 2.6, there exist t1,t2Tt_{1},t_{2}\in T such that P(χ(t1)χ(t2))P^{*}\cap(\chi(t_{1})\cup\chi(t_{2}))\neq\emptyset for every P𝒫P\in\mathcal{P}. We observe:

Lemma 4.1.

We have a,bχ(t1)χ(t2)a,b\in\chi(t_{1})\cup\chi(t_{2}).

Proof.

Suppose that aχ(t1)χ(t2)a\not\in\chi(t_{1})\cup\chi(t_{2}). Let DD be the component of G(χ(t1)χ(t2))G\setminus(\chi(t_{1})\cup\chi(t_{2})) such that aDa\in D. Then N(D)N(D) is contained in the union of at most two adhesions of (T,χ)(T,\chi), at most one for each of χ(t1)\chi(t_{1}) and χ(t2)\chi(t_{2}). Since (T,χ)(T,\chi) is mm-lean, it follows that |N(D)|2(m1)|N(D)|\leq 2(m-1). Since P(χ(t1)χ(t2))P^{*}\cap(\chi(t_{1})\cup\chi(t_{2}))\neq\emptyset for every P𝒫P\in\mathcal{P}, it follows that PN(D)P^{*}\cap N(D)\neq\emptyset for every P𝒫P\in\mathcal{P}. But then 𝒫\mathcal{P} contains at most |N(D)||N(D)| pairwise internally vertex-disjoint paths, contrary to the fact that abab is a (4m+2)(m1)(4m+2)(m-1)-banana in GG. ∎

We say that a vertex vv is dd-safe if |N(v)Hub(G)|d|N(v)\cap\operatorname{Hub}(G)|\leq d. The goal of the next two lemmas is to classify dd-safe vertices into “good ones” and “bad ones,” and show that the bad ones are rare. Let t0V(T)t_{0}\in V(T). A vertex vV(G)v\in V(G) is t0t_{0}-cooperative if either vχ(t0)v\not\in\chi(t_{0}), or degχ^(t0)(v)<2m(m1)\deg_{\hat{\chi}(t_{0})}(v)<2m(m-1). It was shown in [3] that t0t_{0}-cooperative vertices have the following important property (note that [3] assumed that t0t_{0} is a center of (T,χ)(T,\chi), but this was not used in the proof of the following lemma):

Lemma 4.2 (Abrishami, Alecu, Chudnovsky, Hajebi, Spirkl [3]).

Let t0V(T)t_{0}\in V(T). If u,vχ(t0)u,v\in\chi(t_{0}) are dd-safe and not t0t_{0}-cooperative, then uu is adjacent to vv.

A vertex vGv\in G is abab-cooperative if there exists a component DD of GN[v]G\setminus N[v] such that a,bN[D]a,b\in N[D]. We show:

Lemma 4.3.

If vGv\in G is dd-safe and not abab-cooperative, then vv is adjacent to both aa and bb. In particular, the set of vertices that are not abab-cooperative is a clique.

Proof.

Suppose vv is non-adjacent to aa and vv is not abab-cooperative. Since abab is a (4m+2)(m1)(4m+2)(m-1)-banana and (4m+2)(m1)k+d(4m+2)(m-1)\geq k+d, we can choose P1,,Pk+d𝒫P_{1},\dots,P_{k+d}\in\mathcal{P} such that the sets P1,,Pk+dP_{1}^{*},\dots,P_{k+d}^{*} are pairwise vertex-disjoint. Since vv is not abab-cooperative, it follows that vv has a neighbors in each of P1,,Pk+dP_{1}^{*},\dots,P_{k+d}^{*}, and so there exist pairwise vertex-disjoint paths Q1,Qk+dQ_{1},\dots Q_{k+d}, each with ends a,va,v, and such that QiPiQ_{i}^{*}\subseteq P_{i}^{*} for every i{1,,k+d}i\in\{1,\dots,k+d\}. Since vv is dd-safe, we may assume (by renumbering the paths Q1,,Qk+dQ_{1},\dots,Q_{k+d} if necessary), that N(v)Hub(G)Qi=N(v)\cap\operatorname{Hub}(G)\cap Q_{i}=\emptyset for i{1,,k}i\in\{1,\dots,k\}. But now we get a contradiction applying Theorem 3.2 to the graph H=Q1QkH=Q_{1}\cup\dots\cup Q_{k} and the kk-banana avav in HH. This proves the first statement of 4.3. Since GG is C4C_{4}-free, the second statement follows. ∎

In view of Theorem 2.6, it would be convenient if we could work with the graph χ^(t1)χ^(t2)\hat{\chi}(t_{1})\cup\hat{\chi}(t_{2}), but unfortunately this graph may not be in 𝒞t\mathcal{C}_{t}, or even 𝒞\mathcal{C}. In [3], a tool was designed to construct a safe alternative to torsos by adding a set of vertices for each adhesion, rather than turning the adhesion into a clique:

Theorem 4.4 (Abrishami, Alecu, Chudnovsky, Hajebi, Spirkl [3]).

Let t0Tt_{0}\in T. Assume that GG does not admit a clique cutset. For every tNT(t0)t\in N_{T}(t_{0}), there exists Conn(t0,t)Gt0t\operatorname{Conn}(t_{0},t)\subseteq G_{t_{0}\rightarrow t} such that

  1. (1)

    adh(t,t0)=χ(t)χ(t0)Conn(t0,t)\operatorname{adh}(t,t_{0})=\chi(t)\cap\chi(t_{0})\subseteq\operatorname{Conn}(t_{0},t).

  2. (2)

    Conn(t0,t)χ(t0)\operatorname{Conn}(t_{0},t)\setminus\chi(t_{0}) is connected and N(Conn(t0,t)χ(t0))=χ(t0)χ(t)N(\operatorname{Conn}(t_{0},t)\setminus\chi(t_{0}))=\chi(t_{0})\cap\chi(t).

  3. (3)

    No vertex of Conn(t0,t)χ(t0)\operatorname{Conn}(t_{0},t)\setminus\chi(t_{0}) is a hub in the graph (GGt0t)Conn(t0,t)(G\setminus G_{t_{0}\rightarrow t})\cup\operatorname{Conn}(t_{0},t).

Similarly to the construction in [3], we now define a graph that is a safe alternative to χ^(t1)χ^(t2)\hat{\chi}(t_{1})\cup\hat{\chi}(t_{2}). Let SS^{\prime} be a stable set of hubs of GG with S{a,b}=S^{\prime}\cap\{a,b\}=\emptyset, and assume that every sSs\in S^{\prime} is dd-safe. Let SbadS_{bad} denote the set of all vertices in SS^{\prime} that are common neighbors of aa and bb, or are not t1t_{1}-cooperative, or are not t2t_{2}-cooperative. Since SS^{\prime} is stable, Lemma 4.2 implies that |Sbad|3|S_{bad}|\leq 3. By Lemma 4.3, every vertex in SSbadS^{\prime}\setminus S_{bad} is abab-cooperative. If t1t2t_{1}\neq t_{2}, for i{1,2}i\in\{1,2\}, let tit_{i}^{\prime} be the neighbor of tit_{i} in the (unique) path in TT from t1t_{1} to t2t_{2}. If t1=t2t_{1}=t_{2}, set t1=t2=t1t_{1}^{\prime}=t_{2}^{\prime}=t_{1}. Let S=SSbadS=S^{\prime}\setminus S_{bad} and set

β(S)=(χ(t1)χ(t2)(i{1,2}tNT(ti){ti}Conn(ti,t)))Sbad.\beta(S^{\prime})=\left(\chi(t_{1})\cup\chi(t_{2})\cup\left(\bigcup_{i\in\{1,2\}}\bigcup_{t\in N_{T}(t_{i})\setminus\{t_{i}^{\prime}\}}\operatorname{Conn}(t_{i},t)\right)\right)\setminus S_{bad}.

Write β=β(S)\beta=\beta(S^{\prime}). We fix S,S,SbadS^{\prime},S,S_{bad} and β\beta throughout this section. It follows that for every i{1,2}i\in\{1,2\} and for every tNT(ti){ti}t\in N_{T}(t_{i})\setminus\{t_{i}^{\prime}\}, we have that β(GGtit)Conn(ti,t)\beta\subseteq(G\setminus G_{t_{i}\rightarrow t})\cup\operatorname{Conn}(t_{i},t). Let XβX\subseteq\beta. For i{1,2}i\in\{1,2\}, we define δi(X)\delta_{i}(X) to be the set of all vertices tNT(ti){ti}t\in N_{T}(t_{i})\setminus\{t_{i}^{\prime}\} such that X(Gtit(χ(t1)χ(t2)))X\cap(G_{t_{i}\rightarrow t}\setminus(\chi(t_{1})\cup\chi(t_{2})))\neq\emptyset; let δ(X)=δ1(X)δ2(X)\delta(X)=\delta_{1}(X)\cup\delta_{2}(X). Write

Δ(X)=tδ1(X)adh(t1,t)tδ2(X)adh(t2,t).\Delta(X)=\bigcup_{t\in\delta_{1}(X)}\operatorname{adh}(t_{1},t)\cup\bigcup_{t\in\delta_{2}(X)}\operatorname{adh}(t_{2},t).

Next we summarize several key properties of β\beta.

Lemma 4.5.

Suppose that GG does not admit a clique cutset and let sSHub(β)s\in S\cap\operatorname{Hub}(\beta). Then the following hold.

  1. (1)

    sχ(t1)χ(t2)s\in\chi(t_{1})\cup\chi(t_{2}).

  2. (2)

    For i{1,2}i\in\{1,2\}, we have |Nχ^(ti)(s)|<2m(m1)|N_{\hat{\chi}(t_{i})}(s)|<2m(m-1).

  3. (3)

    |Nχ(t1)χ(t2)(s)|<4m(m1)|N_{\chi(t_{1})\cup\chi(t_{2})}(s)|<4m(m-1).

  4. (4)

    There is a component B(s)B(s) of βN[s]\beta\setminus N[s] such that a,bN[B(s)]a,b\in N[B(s)].

Proof.

Let sSHub(β)s\in S\cap\operatorname{Hub}(\beta). It follows from Theorem 4.4(3) that sχ(t1)χ(t2)s\in\chi(t_{1})\cup\chi(t_{2}). Since ss is tit_{i}-cooperative for i{1,2}i\in\{1,2\}, we have that |Nχ^(ti)(s)|<2m(m1)|N_{\hat{\chi}(t_{i})}(s)|<2m(m-1), and the second assertion of the theorem holds. Since χ(ti)\chi(t_{i}) is a subgraph of χ^(ti)\hat{\chi}(t_{i}), the third assertion follows immediately from the second.

We now prove the fourth assertion. Suppose there is no component DD of βN[s]\beta\setminus N[s] such that a,bN[D]a,b\in N[D]. Write

Δ^(s)=Nχ(t1)χ(t2)[s]adh(t1,t1)adh(t2,t2)Δ(Nβ[s]).\hat{\Delta}(s)=N_{\chi(t_{1})\cup\chi(t_{2})}[s]\cup\operatorname{adh}(t_{1},t_{1}^{\prime})\cup\operatorname{adh}(t_{2},t_{2}^{\prime})\cup\Delta(N_{\beta}[s]).

Since Δ^(s)Nχ^(t1)[s]Nχ^(t2)[s]adh(t1,t1)adh(t2,t2)\hat{\Delta}(s)\subseteq N_{\hat{\chi}(t_{1})}[s]\cup N_{\hat{\chi}(t_{2})}[s]\cup\operatorname{adh}(t_{1},t_{1}^{\prime})\cup\operatorname{adh}(t_{2},t_{2}^{\prime}), and adh(T,χ)m1\operatorname{adh}(T,\chi)\leq m-1, and by the second assertion of the theorem, we have that |Δ^(s)|<(4m+2)(m1)|\hat{\Delta}(s)|<(4m+2)(m-1).

(3) Δ^(s)Q\hat{\Delta}(s)\cap Q^{*}\neq\emptyset for every path in β\beta with ends a,ba,b.

Suppose there is a path QQ in β\beta with ends a,ba,b such that Δ^(s)Q=\hat{\Delta}(s)\cap Q^{*}=\emptyset. Since there is no component DD of βN[s]\beta\setminus N[s] such that a,bN[D]a,b\in N[D], it follows that Nβ(s)QN_{\beta}(s)\cap Q^{*}\neq\emptyset. Consequently, there is an i{1,2}i\in\{1,2\} and tNT(ti){ti}t\in N_{T}(t_{i})\setminus\{t_{i}^{\prime}\} such that Nβ(s)QConn(ti,t)N_{\beta}(s)\cap Q^{*}\cap\operatorname{Conn}(t_{i},t)\neq\emptyset. Since QQ is a path from aa to bb in GG, it follows that Q(χ(t1)χ(t2))Q^{*}\cap(\chi(t_{1})\cup\chi(t_{2}))\neq\emptyset. Since QQ^{*} is connected and adh(ti,t)\operatorname{adh}(t_{i},t) separates Gtitχ(ti)G_{t_{i}\rightarrow t}\setminus\chi(t_{i}) from (χ(t1)χ(t2))χ(t)(\chi(t_{1})\cup\chi(t_{2}))\setminus\chi(t), we deduce that Qadh(ti,t)Q^{*}\cap\operatorname{adh}(t_{i},t)\neq\emptyset. But since ss has a neighbor in Conn(ti,t)\operatorname{Conn}(t_{i},t), it follows that adh(ti,t)Δ^(s)\operatorname{adh}(t_{i},t)\subseteq\hat{\Delta}(s), a contradiction. This proves (4).

(4) PΔ^(s)P^{*}\cap\hat{\Delta}(s)\neq\emptyset for every path PP of GG with ends a,ba,b.

Let PP be a path in GG with ends a,ba,b. We prove by induction on |Pβ||P\setminus\beta| that Δ^(s)P\hat{\Delta}(s)\cap P^{*}\neq\emptyset. If PβP\subseteq\beta, the claim follows from (4), and this is the base case.

Let pPβp\in P^{*}\setminus\beta and let t0Tt_{0}\in T such that pχ(t0)p\in\chi(t_{0}). Suppose first that t0t_{0} belongs to the component TT^{\prime} of T{t1,t2}T\setminus\{t_{1},t_{2}\} such that t1Tt_{1}^{\prime}\in T^{\prime}. Then t2Tt_{2}^{\prime}\in T^{\prime}. Since PP^{*} is connected and P(χ(t1)χ(t2))P^{*}\cap(\chi(t_{1})\cup\chi(t_{2}))\neq\emptyset, it follows that P(adh(t1,t1)adh(t2,t2))P^{*}\cap(\operatorname{adh}(t_{1},t_{1}^{\prime})\cup\operatorname{adh}(t_{2},t_{2}^{\prime}))\neq\emptyset, and so PΔ^(s)P^{*}\cap\hat{\Delta}(s)\neq\emptyset. Thus we may assume that pGt1tχ(t1)p\in G_{t_{1}\rightarrow t}\setminus\chi(t_{1}) for some tNT(t1){t1}t\in N_{T}(t_{1})\setminus\{t_{1}^{\prime}\}. Since PP^{*} is connected and P(χ(t1)χ(t2))P^{*}\cap(\chi(t_{1})\cup\chi(t_{2}))\neq\emptyset, it follows that Padh(t1,t)P^{*}\cap\operatorname{adh}(t_{1},t)\neq\emptyset. Write P=p1--plP=p_{1}\hbox{-}\cdots\hbox{-}p_{l} where p1=ap_{1}=a and pl=bp_{l}=b. Let ii be minimum and jj maximum such that piadh(t1,t)p_{i}\in\operatorname{adh}(t_{1},t). Since pGt1tχ(t1)p\in G_{t_{1}\rightarrow t}\setminus\chi(t_{1}), it follows that i<j1i<j-1. By Theorem 4.4(2) there is a path RR form pip_{i} to pjp_{j} with RConn(t1,t)R^{*}\subseteq\operatorname{Conn}(t_{1},t). Then P=p1--pi-R-pj--plP^{\prime}=p_{1}\hbox{-}\cdots\hbox{-}p_{i}\hbox{-}R\hbox{-}p_{j}\hbox{-}\dots\hbox{-}p_{l} is a path from aa to bb in GG with |Pβ|<|Pβ||P^{\prime}\setminus\beta|<|P\setminus\beta|. Inductively, Δ^(s)P\hat{\Delta}(s)\cap P^{\prime*}\neq\emptyset. But Δ^(s)χ(t1)χ(t2)\hat{\Delta}(s)\subseteq\chi(t_{1})\cup\chi(t_{2}) and P(χ(t1)χ(t2))P(χ(t1)χ(t2))P^{\prime*}\cap(\chi(t_{1})\cup\chi(t_{2}))\subseteq P^{*}\cap(\chi(t_{1})\cup\chi(t_{2})), and so Δ^(s)P\hat{\Delta}(s)\cap P^{*}\neq\emptyset. This proves (4).

Since abab is a (4m+2)(m1)(4m+2)(m-1)-banana in GG, and since Δ^(s)<(4m+2)(m1)\hat{\Delta}(s)<(4m+2)(m-1), (4) leads to a contradiction. This completes the proof.

Recall that a separation of β\beta is a triple (X,Y,Z)(X,Y,Z) of pairwise disjoint subsets of β\beta with XYZ=βX\cup Y\cup Z=\beta such that XX is anticomplete to ZZ. We are now ready to move on to star cutsets. We will work in the graph β\beta and take advantage of its special properties. At the end of the section we will explain how to connect our results back to the graph GG that we are interested in.

As in other papers in the series, we associate a certain unique star separation to every vertex of SHub(β)S\cap\operatorname{Hub}(\beta). The separation here is chosen differently from the way it was done in the past, but the behavior we observe is similar. The reason for the difference is that unlike in [2] or [6], our here goal is to disconnect two given vertices, rather than find a “balanced separator” in the graph (more on this in Section 10).

Let vSHub(β)v\in S\cap\operatorname{Hub}(\beta). By Theorem 4.5, there is a component DD of βN[v]\beta\setminus N[v] such that a,bN[D]a,b\in N[D]. Since sSbads\not\in S_{bad}, it follows that ss is not complete to {a,b}\{a,b\}; consequently D{a,b}D\cap\{a,b\}\neq\emptyset, and so DD is unique. Let B(v)B(v) be the unique component of βN[v]\beta\setminus N[v] such that a,bN[B(v)]a,b\in N[B(v)] (this choice of B(v)B(v) is different from what we have done in earlier papers). Let C(v)=N(B(v)){v}C(v)=N(B(v))\cup\{v\}, and finally, let A(v)=β(B(v)C(v))A(v)=\beta\setminus(B(v)\cup C(v)). Then (A(v),C(v),B(v))(A(v),C(v),B(v)) is the canonical star separation of β\beta corresponding to vv. The next lemma is a key step in the proof of the fact that decomposing by canonical star separations makes the graph simpler. We show:

Lemma 4.6.

The vertex vv is not a hub of βA(v)\beta\setminus A(v).

Proof.

Suppose that (H,v)(H,v) is a proper wheel in βA(v)\beta\setminus A(v). Then HN[B(v)]H\subseteq N[B(v)], contrary to Theorem 3.1. This proves Lemma 4.6. ∎

We need just a little more set up. Let 𝒪\mathcal{O} be a linear order on SHub(β)S\cap\operatorname{Hub}(\beta). Following [4], we say that two vertices of SHub(β)S\cap\operatorname{Hub}(\beta) are star twins if B(u)=B(v)B(u)=B(v), C(u){u}=C(v){v}C(u)\setminus\{u\}=C(v)\setminus\{v\}, and A(u){u}=A(v){v}A(u)\cup\{u\}=A(v)\cup\{v\}. (Note that every two of these conditions imply the third.)

Let A\leq_{A} be a relation on SHub(β)S\cap\operatorname{Hub}(\beta) defined as follows:

xAy if{x=y;x and y are star twins and 𝒪(x)<𝒪(y); orx and y are not star twins and yA(x).\hskip 65.44142ptx\leq_{A}y\ \ \ \text{ if}\ \ \ \begin{cases}x=y;\\ \text{$x$ and $y$ are star twins and $\mathcal{O}(x)<\mathcal{O}(y)$; or}\\ \text{$x$ and $y$ are not star twins and }y\in A(x).\\ \end{cases}

Note that if xAyx\leq_{A}y, then either x=yx=y, or yA(x).y\in A(x).

The conclusion of the next lemma is the same as of Lemma 6.2 from [5], but the assumptions are different.

Lemma 4.7.

If yA(x)y\in A(x), then A(y){y}A(x){x}A(y)\cup\{y\}\subseteq A(x)\cup\{x\}.

Proof.

Since C(y)N[y]C(y)\subseteq N[y] and yy is anticomplete to B(x)B(x), we have B(x)GN[y]B(x)\subseteq G\setminus N[y]. Since B(x)B(x) is connected, there is a component DD of GN[y]G\setminus N[y] such that B(x)DB(x)\subseteq D. Since xSbadx\not\in S_{bad}, it follows that {a,b}B(x)\{a,b\}\cap B(x)\neq\emptyset, and so D=B(y)D=B(y). Let vC(x){x}v\in C(x)\setminus\{x\}. Then vv has a neighbor in B(x)B(x) and thus in B(y)B(y). If vN[y]v\in N[y], then vC(y)v\in C(y). If vN[y]v\not\in N[y], then vB(y)v\in B(y). It follows that C(x){x}C(y)B(y)C(x)\setminus\{x\}\subseteq C(y)\cup B(y). But now A(y){x}A(x)A(y)\setminus\{x\}\subseteq A(x), as required. This proves Lemma 4.7. ∎

The proofs of the next two lemmas are identical to the proofs of Lemmas 6.3 and Lemma 6.4 in [5] (using Lemma 4.7 instead of Lemma 6.2 of [5]), and we omit them.

Lemma 4.8.

A\leq_{A} is a partial order on SHub(β)S\cup\operatorname{Hub}(\beta).

In view of Lemma 4.8, let Core(S)\operatorname{Core}(S^{\prime}) be the set of all A\leq_{A}-minimal elements of SHub(β)S\cap\operatorname{Hub}(\beta).

Lemma 4.9.

Let u,vCore(S)u,v\in\operatorname{Core}(S^{\prime}). Then A(u)C(v)=C(u)A(v)=A(u)\cap C(v)=C(u)\cap A(v)=\emptyset.

We have finally reached our goal: we can define a subgraph of GG that is simpler than GG itself, but carries all the information we need. Define

βA(S)=vCore(S)(B(v)C(v)).\beta^{A}(S^{\prime})=\bigcap_{v\in\operatorname{Core}(S^{\prime})}(B(v)\cup C(v)).

The next theorem summarizes the important aspects of the behavior of βA(S)\beta^{A}(S^{\prime}).

Theorem 4.10.

The following hold:

  1. (1)

    For every vCore(S)v\in\operatorname{Core}(S^{\prime}), we have C(v)βA(S)C(v)\subseteq\beta^{A}(S^{\prime}).

  2. (2)

    For every vCore(S)v\in\operatorname{Core}(S^{\prime}), |C(v)(χ(t1)χ(t2))|4m(m1)|C(v)\cap(\chi(t_{1})\cup\chi(t_{2}))|\leq 4m(m-1).

  3. (3)

    For every vCore(S)v\in\operatorname{Core}(S^{\prime}), |Δ(C(v))|4m(m1)|\Delta(C(v))|\leq 4m(m-1).

  4. (4)

    For every component DD of ββA(S)\beta\setminus\beta^{A}(S^{\prime}), there exists vCore(S)v\in\operatorname{Core}(S^{\prime}) such that DA(v)D\subseteq A(v). Further, if DD is a component of ββA(S)\beta\setminus\beta^{A}(S^{\prime}) and vCore(S)v\in\operatorname{Core}(S^{\prime}) such that DA(v)D\subseteq A(v), then Nβ(D)C(v)N_{\beta}(D)\subseteq C(v).

  5. (5)

    SHub(βA(S))=S^{\prime}\cap\operatorname{Hub}(\beta^{A}(S^{\prime}))=\emptyset.

Proof.

(1) is immediate from Lemma 4.9, and (2) follows from Lemma 4.5.

Next we prove (3). By Lemma 4.5(1), it follows that vχ(t1)χ(t2)v\in\chi(t_{1})\cup\chi(t_{2}). Observe that if tT{t1,t2,t1,t2}t\in T\setminus\{t_{1},t_{2},t_{1}^{\prime},t_{2}^{\prime}\} and (Gtit(χ(t1)χ(t2))C(v)(G_{t_{i}\rightarrow t}\setminus(\chi(t_{1})\cup\chi(t_{2}))\cap C(v)\neq\emptyset for some i{1,2}i\in\{1,2\}, then vχ(t1)χ(t)v\in\chi(t_{1})\cap\chi(t) or vχ(t2)χ(t)v\in\chi(t_{2})\cap\chi(t). It follows that for i{1,2}i\in\{1,2\}, adh(ti,t)Δ(C(v))\operatorname{adh}(t_{i},t)\subseteq\Delta(C(v)) only if vadh(ti,t)v\in\operatorname{adh}(t_{i},t). Consequently, |Δ(C(v))|degχ^(t1)(v)+degχ^(t2)(v)|\Delta(C(v))|\leq\deg_{\hat{\chi}(t_{1})}(v)+\deg_{\hat{\chi}(t_{2})}(v), and so (3) follows from Lemma 4.5(3).

Next we prove (4). Let DD be a component of ββA(S)\beta\setminus\beta^{A}(S^{\prime}). Since ββA(S)=vCore(S)A(v)\beta\setminus\beta^{A}(S^{\prime})=\bigcup_{v\in\operatorname{Core}(S^{\prime})}A(v), there exists vCore(S)v\in\operatorname{Core}(S^{\prime}) such that DA(v)D\cap A(v)\neq\emptyset. If DA(v)D\setminus A(v)\neq\emptyset, then, since DD is connected, it follows that DN(A(v))D\cap N(A(v))\neq\emptyset; but then DC(v)D\cap C(v)\neq\emptyset, contrary to (1). Since Nβ(D)βA(S)N_{\beta}(D)\subseteq\beta^{A}(S^{\prime}) and Nβ(D)A(v)C(v)N_{\beta}(D)\subseteq A(v)\cup C(v), it follows that Nβ(D)C(v)N_{\beta}(D)\subseteq C(v). This proves (4).

To prove (5), let uSHub(βA(S))u\in S^{\prime}\cap\operatorname{Hub}(\beta^{A}(S^{\prime})). Since βA(S)β\beta^{A}(S^{\prime})\subseteq\beta, we deduce that uSbadu\not\in S_{bad}, and so uSHub(β)u\in S\cap\operatorname{Hub}(\beta). By Theorem 4.5(1), we have that uχ(t1)χ(t2)u\in\chi(t_{1})\cup\chi(t_{2}). By Lemma 4.6, it follows that βA(S)B(u)C(u)\beta^{A}(S^{\prime})\not\subseteq B(u)\cup C(u), and therefore uCore(S)u\not\in\operatorname{Core}(S^{\prime}). But then uA(v)u\in A(v) for some vCore(S)v\in\operatorname{Core}(S^{\prime}), and so uβA(S)u\not\in\beta^{A}(S^{\prime}), a contradiction. This proves (5) and completes the proof of Theorem 4.10. ∎

We now explain how βA(S)\beta^{A}(S^{\prime}) is used. In the course of the proof of Theorem 1.3, we will inductively obtain a small cutset separating aa from bb in βA(S)\beta^{A}(S^{\prime}). The next theorem allows us to transform this cutset into a cutset separating aa from bb in β\beta.

Theorem 4.11.

Let (X,Y,Z)(X,Y,Z) be a separation of βA(S)\beta^{A}(S^{\prime}) such that aXa\in X and bZb\in Z. Let Y=(YsYCore(S)C(s)){a,b}Y^{\prime}=(Y\cup\bigcup_{s\in Y\cap\operatorname{Core}(S^{\prime})}C(s))\setminus\{a,b\}. Then

  1. (1)

    YY^{\prime} separates aa from bb in β\beta.

  2. (2)

    |Y(χ(t1)χ(t2))||Y|+4m(m1)|YCore(S)||Y^{\prime}\cap(\chi(t_{1})\cup\chi(t_{2}))|\leq|Y|+4m(m-1)|Y\cap\operatorname{Core}(S^{\prime})|.

  3. (3)

    |Δ(Y)Δ(Y)|4m(m1)|YCore(S)||\Delta(Y^{\prime})\setminus\Delta(Y)|\leq 4m(m-1)|Y\cap\operatorname{Core}(S^{\prime})|.

Refer to caption
Figure 2. Proof of Theorem 4.11.
Proof.

Suppose that YY^{\prime} does not separate aa from bb in β\beta. Let PP be a path from aa to bb in βY\beta\setminus Y^{\prime}. Let Da,DbD_{a},D_{b} be the components of βA(S)Y\beta^{A}(S^{\prime})\setminus Y such that aDaa\in D_{a} and bDbb\in D_{b}. Since the first vertex of PP is in DaD_{a} and the last vertex of PP is in DbD_{b}, it follows that there is a subpath QQ of PP such that QQ has ends u,vu,v and:

  • The vertex uu is in DaD_{a};

  • QβA(S)=Q^{*}\cap\beta^{A}(S^{\prime})=\emptyset;

  • The vertex vv is in βA(S)Da\beta^{A}(S^{\prime})\setminus D_{a}.

See Figure 2. Since YYY\subseteq Y^{\prime}, it follows that vv is in a component DcD_{c} of βA(S)Y\beta^{A}(S^{\prime})\setminus Y with DcDaD_{c}\neq D_{a} (possibly Dc=DbD_{c}=D_{b}). This implies that uvE(G)uv\not\in E(G), and so QQ^{*}\neq\emptyset.

By Theorem 4.10(4), there is an sCore(S)βA(S)s\in\operatorname{Core}(S^{\prime})\subseteq\beta^{A}(S^{\prime}) such that QA(s)Q^{*}\subseteq A(s) and u,vC(s)u,v\in C(s). Since Da,DcD_{a},D_{c} are distinct components of βA(S)Y\beta^{A}(S^{\prime})\setminus Y, and since NβA(S)(s)N_{\beta^{A}(S^{\prime})}(s) meets both DaD_{a} and DbD_{b}, it follows that sYs\in Y. Therefore C(s){a,b}YC(s)\setminus\{a,b\}\subseteq Y^{\prime}, and so u,v{a,b}u,v\in\{a,b\}. Since uvu\neq v, it follows that u=au=a and b=vb=v, and so a,bN(s)a,b\in N(s), contrary to the fact that sSbads\not\in S_{bad}. This proves the first assertion of the theorem.

The second assertion follows from Theorem 4.10(2), and the third assertion follows from Theorem 4.10(3). ∎

We finish this section with a theorem that allows us to transform a cutset separating aa from bb in β\beta into a cutset separating aa from bb in GG.

Theorem 4.12.

Assume that GG does not admit a clique cutset. Let (X,Y,Z)(X,Y,Z) be a separation of β\beta such that aXa\in X and bZb\in Z. Let

Y=(Sbadadh(t1,t1)adh(t2,t2)(Y(χ(t1)χ(t2)))Δ(Y)){a,b}.Y^{\prime}=(S_{bad}\cup\operatorname{adh}(t_{1},t_{1}^{\prime})\cup\operatorname{adh}(t_{2},t_{2}^{\prime})\cup(Y\cap(\chi(t_{1})\cup\chi(t_{2})))\cup\Delta(Y))\setminus\{a,b\}.

Then YY^{\prime} separates aa from bb in GG.

Proof.

Suppose not. Let Da,DbD_{a},D_{b} be the components of βY\beta\setminus Y such that aDaa\in D_{a} and bDbb\in D_{b}. Since YY^{\prime} does not separate aa from bb in GG, and Y{a,b}=Y^{\prime}\cap\{a,b\}=\emptyset, there is a component DD of GYG\setminus Y^{\prime} such that a,bDa,b\in D. Let PP be a path from aa to bb with PDP\subseteq D.

As before, it follows that there is a subpath QQ of PP such that QQ has ends u,vu,v and:

  • The vertex uu is in DaD_{a};

  • Qβ=Q^{*}\cap\beta=\emptyset;

  • The vertex vv is in βDa\beta\setminus D_{a}.

We first consider the case that vYv\in Y (and so vPv\in P^{*}). Then vConn(ti,t)(χ(t1)χ(t2))v\in\operatorname{Conn}(t_{i},t)\setminus(\chi(t_{1})\cup\chi(t_{2})) for some i{1,2}i\in\{1,2\} and tNT(ti){ti}t\in N_{T}(t_{i})\setminus\{t_{i}^{\prime}\}. It follows that tδi(Y)t\in\delta_{i}(Y), and so adh(ti,t){a,b}Δ(Y){a,b}Y\operatorname{adh}(t_{i},t)\setminus\{a,b\}\subseteq\Delta(Y)\setminus\{a,b\}\subseteq Y^{\prime}. It follows that Padh(ti,t)=P^{*}\cap\operatorname{adh}(t_{i},t)=\emptyset, and from Theorem 2.6, we know that P(χ(t1)χ(t2))P^{*}\cap(\chi(t_{1})\cup\chi(t_{2}))\neq\emptyset. Therefore, P(Gtit(χ(t1)χ(t2)))=P^{*}\cap(G_{t_{i}\rightarrow t}\setminus(\chi(t_{1})\cup\chi(t_{2})))=\emptyset, contrary to the fact that vConn(ti,t)(χ(t1)χ(t2))v\in\operatorname{Conn}(t_{i},t)\setminus(\chi(t_{1})\cup\chi(t_{2})). This is a contradiction, and proves that vYv\not\in Y.

Therefore, there is a component DcDaD_{c}\neq D_{a} (possibly Db=DcD_{b}=D_{c}) of βY\beta\setminus Y such that vDcv\in D_{c}. Since there are no edges between DaD_{a} and DcD_{c}, it follows that QQ^{*}\neq\emptyset.

Consequently, there is a component T~\tilde{T} of T{t1,t2}T\setminus\{t_{1},t_{2}\} such that Q(tT~χ(t))(χ(t1)χ(t2))Q^{*}\subseteq(\bigcup_{t\in\tilde{T}}\chi(t))\setminus(\chi(t_{1})\cup\chi(t_{2})). Let D~\tilde{D} be the component of G(χ(t1)χ(t2))G\setminus(\chi(t_{1})\cup\chi(t_{2})) such that QD~Q^{*}\subseteq\tilde{D}. By Theorem 2.6, it follows that |N(D~){a,b}|1|N(\tilde{D})\cap\{a,b\}|\leq 1.

Suppose first that t1T~t_{1}^{\prime}\in\tilde{T}. Then t1t2t_{1}^{\prime}\neq t_{2}, and t2t1t_{2}^{\prime}\neq t_{1}, and t2T~t_{2}^{\prime}\in\tilde{T}. Since NG(D~)adh(t1,t1)adh(t2,t2)N_{G}(\tilde{D})\subseteq\operatorname{adh}(t_{1},t_{1}^{\prime})\cup\operatorname{adh}(t_{2},t_{2}^{\prime}), it follows that NGY(D~){a,b}N_{G\setminus Y^{\prime}}(\tilde{D})\subseteq\{a,b\}, and so |NGY(D~)|1|N_{G\setminus Y^{\prime}}(\tilde{D})|\leq 1. Since PP is a path with ends a,bD~a,b\not\in\tilde{D} and with PGYP\subseteq G\setminus Y^{\prime}, it follows that PD~=P\cap\tilde{D}=\emptyset, and so QD~=Q\cap\tilde{D}=\emptyset. Similarly, t2T~t_{2}^{\prime}\not\in\tilde{T}.

We deduce that exactly one of t1,t2t_{1},t_{2} has a neighbor in T~\tilde{T} (unless t1=t2t_{1}=t_{2}), and this neighbor is unique; denote it by tt. By symmetry, we may assume that tt1E(T)tt_{1}\in E(T), and so T~=Tt1t\tilde{T}=T_{t_{1}\rightarrow t}. By Theorem 2.5, we deduce that D~=Gt1tχ(t1)\tilde{D}=G_{t_{1}\rightarrow t}\setminus\chi(t_{1}). Since uDaN(Q)βN[D~]u\in D_{a}\cap N(Q^{*})\subseteq\beta\cap N[\tilde{D}], it follows that uConn(t1,t)u\in\operatorname{Conn}(t_{1},t). Similarly, we have vConn(t1,t)v\in\operatorname{Conn}(t_{1},t). Since Conn(t1,t)χ(t1)\operatorname{Conn}(t_{1},t)\setminus\chi(t_{1}) is connected and N(Conn(t1,t)χ(t1))=χ(t1)χ(t)N(\operatorname{Conn}(t_{1},t)\setminus\chi(t_{1}))=\chi(t_{1})\cap\chi(t) by Theorem 4.4(2), it follows that Conn(t1,t)\operatorname{Conn}(t_{1},t) contains a path QQ^{\prime} with ends u,vu,v and interior in Conn(t1,t)χ(t1)\operatorname{Conn}(t_{1},t)\setminus\chi(t_{1}). Since QQ^{\prime} is a path from DaD_{a} to DcD_{c} in β\beta, it follows that YQY\cap Q^{\prime*}\neq\emptyset, and so Y(Conn(t1,t)χ(t1))Y\cap(\operatorname{Conn}(t_{1},t)\setminus\chi(t_{1}))\neq\emptyset. Therefore, Y(Gt1tχ(t1))Y\cap(G_{t_{1}\rightarrow t}\setminus\chi(t_{1}))\neq\emptyset, which implies that adh(t1,t){a,b}Y\operatorname{adh}(t_{1},t)\setminus\{a,b\}\subseteq Y^{\prime}. It follows that Padh(t1,t)=P^{*}\cap\operatorname{adh}(t_{1},t)=\emptyset, and from Theorem 2.6, we know that P(χ(t1)χ(t2))P^{*}\cap(\chi(t_{1})\cup\chi(t_{2}))\neq\emptyset. Therefore, P(Gt1tχ(t1))=P^{*}\cap(G_{t_{1}\rightarrow t}\setminus\chi(t_{1}))=\emptyset, contrary to the fact that QGt1tχ(t1)\emptyset\neq Q^{*}\subseteq G_{t_{1}\rightarrow t}\setminus\chi(t_{1}). This is a contradiction, and concludes the proof. ∎

5. Connectifiers

We start this section by describing minimal connected subgraphs containing the neighbors of a large number of vertices from a given stable set. We then use this result to deduce what a pair of two such subgraphs can look like (for the same stable set) assuming that they are anticomplete to each other and the graph in question is in 𝒞\mathcal{C}. This generalizes results from [2] and [3].

What follows is mostly terminology from [3], but there are also some new notions. Let GG be a graph, let P=p1--pnP=p_{1}\hbox{-}\cdots\hbox{-}p_{n} be a path in GG and let X={x1,,xk}GPX=\{x_{1},\dots,x_{k}\}\subseteq G\setminus P. We say that (P,X)(P,X) is an alignment if NP(x1)={p1}N_{P}(x_{1})=\{p_{1}\}, NP(xk)={pn}N_{P}(x_{k})=\{p_{n}\}, every vertex of XX has a neighbor in PP, and there exist 1<j2<<jk1<jk=n1<j_{2}<\dots<j_{k-1}<j_{k}=n such that NP(xi)pji-P-pji+11N_{P}(x_{i})\subseteq p_{j_{i}}\hbox{-}P\hbox{-}p_{j_{i+1}-1} for i{2,,k1}i\in\{2,\dots,k-1\}. We also say that x1,,xkx_{1},\dots,x_{k} is the order on XX given by the alignment (P,X)(P,X). An alignment (P,X)(P,X) is wide if each of x2,,xk1x_{2},\dots,x_{k-1} has at least two non-adjacent neighbors in PP, spiky if each of x2,,xk1x_{2},\dots,x_{k-1} has a unique neighbor in PP and triangular if each of x2,,xk1x_{2},\dots,x_{k-1} has exactly two neighbors in PP and they are adjacent. An alignment is consistent if it is wide, spiky or triangular.

By a caterpillar we mean a tree CC with maximum degree three such that there exists a path PP of CC such that all vertices of degree 33 in CC belong to PP. We call a minimal such path PP the spine of CC. By a subdivided star we mean a graph isomorphic to a subdivision of the complete bipartite graph K1,δK_{1,\delta} for some δ3\delta\geq 3. In other words, a subdivided star is a tree with exactly one vertex of degree at least three, which we call its root. For a graph HH, a vertex vv of HH is said to be simplicial if NH(v)N_{H}(v) is a clique. We denote by 𝒵(H)\mathcal{Z}(H) the set of all simplicial vertices of HH. Note that for every tree TT, 𝒵(T)\mathcal{Z}(T) is the set of all leaves of TT. An edge ee of a tree TT is said to be a leaf-edge of TT if ee is incident with a leaf of TT. It follows that if HH is the line graph of a tree TT, then 𝒵(H)\mathcal{Z}(H) is the set of all vertices in HH corresponding to leaf-edges of TT.

Let HH be a graph that is either a path, or a caterpillar, or the line graph of a caterpillar, or a subdivided star with root rr, or the line graph of a subdivided star with root rr. We define an induced subgraph of HH, denoted by P(H)P(H), which we will use throughout the paper. If HH is a path, we let P(H)=HP(H)=H. If HH is a caterpillar, we let P(H)P(H) be the spine of HH. If HH is the line graph of a caterpillar CC, let P(H)P(H) be the path in HH consisting of the vertices of HH that correspond to the edges of the spine of CC. If HH is a subdivided star with root rr, let P(H)={r}P(H)=\{r\}. It HH is the line graph of a subdivided star SS with root rr, let P(H)P(H) be the clique of HH consisting of the vertices of HH that correspond to the edges of SS incident with rr. The legs of HH are the components of HP(H)H\setminus P(H).

Next, let HH be a caterpillar or the line graph of a caterpillar and let SS be a set of vertices disjoint from HH such that every vertex of SS has a unique neighbor in HH and HN(S)=𝒵(H)H\cap N(S)=\mathcal{Z}(H). Let XX be the set of vertices of HP(H)H\setminus P(H) that have neighbors in P(H)P(H). Then the neighbors of elements of XX appear in P(H)P(H) in order (there may be ties at the ends of P(H)P(H), which we break arbitrarily); let x1,,xkx_{1},\dots,x_{k} be the corresponding order on XX. Now, order the vertices of SS as s1,,sks_{1},\dots,s_{k} where sis_{i} has a neighbor in the leg of HH containing xix_{i} for i{1,,k}i\in\{1,\dots,k\}. We say that s1,,sks_{1},\dots,s_{k} is the order on SS given by (H,S)(H,S).

Next, let HH be an induced subgraph of GG that is a caterpillar, or the line graph of a caterpillar, or a subdivided star or the line graph of a subdivided star and let XGHX\subseteq G\setminus H be such that every vertex of XX has a unique neighbor in HH and HN(X)=𝒵(H)H\cap N(X)=\mathcal{Z}(H). We say that (H,X)(H,X) is a consistent connectifier and it is spiky, triangular, stellar, or clique respectively. If HH is a single vertex and XN(H)X\subseteq N(H), we also call (H,X)(H,X) a stellar connectifer. If HH is a subdivided star, a singleton or the line graph of a subdivided star, we say that (H,X)(H,X) is a concentrated connectifier.

The following was proved in [3]:

Theorem 5.1 (Abrishami, Alecu, Chudnovsky, Hajebi, Spirkl [3]).

For every integer h1h\geq 1, there exists an integer ν=ν(h)1\nu=\nu(h)\geq 1 with the following property. Let GG be a connected graph with no clique of cardinality hh. Let SGS\subseteq G such that |S|ν|S|\geq\nu, GSG\setminus S is connected and every vertex of SS has a neighbor in GSG\setminus S. Then there is a set S~S\tilde{S}\subseteq S with |S~|=h|\tilde{S}|=h and an induced subgraph HH of GSG\setminus S for which one of the following holds.

  • HH is a path and every vertex of S~\tilde{S} has a neighbor in HH; or

  • HH is a caterpillar, or the line graph of a caterpillar, or a subdivided star. Moreover, every vertex of S~\tilde{S} has a unique neighbor in HH and HN(S~)=𝒵(H)H\cap N(\tilde{S})=\mathcal{Z}(H).

We now prove a version of Theorem 5.1 that does not assume a bound on the clique number. This result may be of independent use in the future (See Figure 3 for a depiction of the outcomes).

Refer to caption
Figure 3. Outcomes of Theorem 5.2. Circled nodes are the vertices in HN(Y)H\cap N(Y^{\prime}).
Theorem 5.2.

For every integer h1h\geq 1, there exists an integer μ=μ(h)1\mu=\mu(h)\geq 1 with the following property. Let GG be a connected graph. Let YGY\subseteq G such that |Y|μ|Y|\geq\mu, GYG\setminus Y is connected and every vertex of YY has a neighbor in GYG\setminus Y. Then there is a set YYY^{\prime}\subseteq Y with |Y|=h|Y^{\prime}|=h and an induced subgraph HH of GYG\setminus Y for which one of the following holds.

  • HH is a path and every vertex of YY^{\prime} has a neighbor in HH; or

  • HH is a caterpillar, or the line graph of a caterpillar, or a subdivided star or the line graph of a subdivided star. Moreover, every vertex of YY^{\prime} has a unique neighbor in HH and HN(Y)=𝒵(H)H\cap N(Y^{\prime})=\mathcal{Z}(H).

Proof.

Let μ(h)=ν(h+1)\mu(h)=\nu(h+1) where ν\nu is as in Theorem 5.1. Let FF be a minimal connected subset of GYG\setminus Y such that every yYy\in Y has a neighbor in FF.

(5) If FF contains a clique of size hh, then there exists HFH\subseteq F and YYY^{\prime}\subseteq Y with |Y|=h|Y^{\prime}|=h such that (H,Y)(H,Y^{\prime}) is a clique connectifier.

Suppose that there is a clique KK of size hh in FF. It follows from the minimality of FF that for every kKk\in K one of the following holds:

  • There is y(k)Yy(k)\in Y such that y(k)y(k) is anticomplete to FkF\setminus k; in this case set C(k,y(k))=C(k,y(k))=\emptyset.

  • F{k}F\setminus\{k\} is not connected, and for every component CC of FkF\setminus k there is yYy\in Y such that yy is anticomplete to F(C{k})F\setminus(C\cup\{k\}). Since KK is a clique, there is a component CC of FkF\setminus k such that KC=K\cap C=\emptyset. Let y(k)y(k) be a vertex of YY that is anticomplete to F(C{k})F\setminus(C\cup\{k\}); write C(k,y(k))=CC(k,y(k))=C.

Let YY^{\prime} be the set of all vertices y(k)y(k) as above. For every kKk\in K, let HkH_{k} be a path from kk to y(k)y(k) with HkC(k,y(k))H_{k}^{*}\subseteq C(k,y(k)). Write H=kKHkH=\bigcup_{k\in K}H_{k}. We will show that |Y|=h|Y^{\prime}|=h and (H,Y)(H,Y^{\prime}) is a clique connectifier. It follows from the definition of HkH_{k} that every vertex of YY^{\prime} has a neighbor in HH.

Next we claim that if kkk\neq k^{\prime}, then HkH_{k} is disjoint from and anticomplete to HkH_{k^{\prime}}. Recall that HkC(k,y(k)){k}H_{k}\subseteq C(k,y(k))\cup\{k\}, and y(k)y(k) is anticomplete to F(C(k,y(k)){k})F\setminus(C(k,y(k))\cup\{k\}). It follows that HkK={k}H_{k}\cap K=\{k\}, and so kHkk\not\in H_{k^{\prime}} and kHkk^{\prime}\not\in H_{k}. Let DD be the component of FkF\setminus k such that kDk^{\prime}\in D. By the definition of C(k,y(k))C(k,y(k)), we have that DC(k,y(k))D\neq C(k,y(k)). Since HkH_{k^{\prime}} is connected and kHkk\not\in H_{k^{\prime}}, it follows that HkDH_{k^{\prime}}\subseteq D. Consequently, HkH_{k} and HkH_{k^{\prime}} are disjoint and anticomplete to each other. Since y(k)y(k) has no neighbor in DD, we deduce that y(k)y(k) is anticomplete to HkH_{k^{\prime}}, and in particular y(k)y(k)y(k)\neq y(k^{\prime}). Similarly, y(k)y(k^{\prime}) is anticomplete to HkH_{k}. This proves the claim.

Now it follows from the claim that if kkk\neq k^{\prime}, then y(k)y(k)y(k)\neq y(k^{\prime}). Consequently, |Y|=|K|=h|Y^{\prime}|=|K|=h, and (H,Y)(H,Y^{\prime}) is a clique connectifier. This proves (5).

By (5), we may assume that FF is KhK_{h}-free. Since SS is a stable set, it follows that FSF\cup S is Kh+1K_{h+1}-free. Now the result follows from Theorem 5.1 applied to FSF\cup S. ∎

Now we move to two anticomplete connected subgraphs and prove the following:

Theorem 5.3.

For every integer x1x\geq 1, there exists an integer ϕ=ϕ(x)1\phi=\phi(x)\geq 1 with the following property. Let G𝒞G\in\mathcal{C} and assume that V(G)=D1D2YV(G)=D_{1}\cup D_{2}\cup Y where

  • YY is a stable set with |Y|=ϕ|Y|=\phi,

  • D1D_{1} and D2D_{2} are components of GYG\setminus Y, and

  • N(D1)=N(D2)=YN(D_{1})=N(D_{2})=Y.

Then there exist XYX\subseteq Y with |X|=x|X|=x, H1D1H_{1}\subseteq D_{1} and H2D2H_{2}\subseteq D_{2} (possibly with the roles of D1D_{1} and D2D_{2} reversed) such that either:

  1. (1)

    Not both (H1,X)(H_{1},X) and (H2,X)(H_{2},X) are alignments, and

    • (H1,X)(H_{1},X) is a triangular connectifier or a clique connectifier or a triangular alignment; and

    • (H2,X)(H_{2},X) is a stellar connectifier, or a spiky connectifier, or a spiky alignment or a wide alignment.

    or

  2. (2)

    Both (H1,X)(H_{1},X) and (H2,X)(H_{2},X) are alignments, and at least one of (H1,X)(H_{1},X) and (H2,X)(H_{2},X) is not a spiky alignment.

Moreover, if neither of (H1,X),(H2,X)(H_{1},X),(H_{2},X) is a concentrated connectifier, then the orders given on XX by (H1,X)(H_{1},X) and by (H2,X)(H_{2},X) are the same.

In this paper we do not need the full generality of Theorem 5.3; we are only interested in two special cases: when D1D_{1} is a path and when the clique number of GG is bounded. However, it is easier to prove the more general result first using the symmetry between D1D_{1} and D2D_{2}, and then use it to handle the special cases. We also believe that Theorem 5.3 will be useful in the future.

We start by recalling a well known theorem of Erdős and Szekeres [20].

Theorem 5.4 (Erdős and Szekeres [20]).

Let x1,,xn2+1x_{1},\dots,x_{n^{2}+1} be a sequence of distinct reals. Then there exists either an increasing or a decreasing (n+1)(n+1)-sub-sequence.

We start with two lemmas.

Lemma 5.5.

Let G𝒞G\in\mathcal{C} and assume that V(G)=H1H2XV(G)=H_{1}\cup H_{2}\cup X where XX is a stable set with |X|3|X|\geq 3 and H1H_{1} is anticomplete to H2H_{2}. Suppose that for i{1,2}i\in\{1,2\}, the pair (Hi,X)(H_{i},X) is a consistent alignment, or a consistent connectifier. Assume also that if neither of (H1,X),(H2,X)(H_{1},X),(H_{2},X) is concentrated, then the orders given on XX by (H1,X)(H_{1},X) and by (H2,X)(H_{2},X) are the same. Then (possibly switching the roles of H1H_{1} and H2H_{2}), we have that either:

  1. (1)

    Not both (H1,X)(H_{1},X) and (H2,X)(H_{2},X) are alignments, and

    • (H1,X)(H_{1},X) is a triangular connectifier or a clique connectifier or a triangular alignment; and

    • (H2,X)(H_{2},X) is a stellar connectifier, or a spiky connectifier, or a spiky alignment or a wide alignment.

    or

  2. (2)

    Both (H1,X)(H_{1},X) and (H2,X)(H_{2},X) are alignments, and at least one of (H1,X)(H_{1},X) and (H2,X)(H_{2},X) is not a spiky alignment, and at least one of (H1,X)(H_{1},X) and (H2,X)(H_{2},X) is not a triangular alignment.

Proof.

If at least one (Hi,X){(H1,X),(H2,X)}(H_{i},X)\subseteq\{(H_{1},X),(H_{2},X)\} is not a concentrated connectifier, we let x1,,xkx_{1},\dots,x_{k} be the order given on XX by (Hi,X)(H_{i},X). If both of (Hi,X)(H_{i},X) are concentrated connectifiers, we let x1,,xkx_{1},\dots,x_{k} be an arbitrary order on XX. Let HH be the unique hole contained in H1H2{x1,xk}H_{1}\cup H_{2}\cup\{x_{1},x_{k}\}. For j{1,2}j\in\{1,2\} and i{1,,k}i\in\{1,\dots,k\}, if HjH_{j} is a connectifier, let DijD_{i}^{j} be the leg of HjH_{j} containing a neighbor of xix_{i}; and if HjH_{j} is an alignment let Dij=D_{i}^{j}=\emptyset.

Suppose first that (H1,X)(H_{1},X) is a triangular alignment, a clique connectifier or a triangular connectifier. If (H2,X)(H_{2},X) is a triangular alignment, a clique connectifier or a triangular connectifier, then for every i{2,,k1}i\in\{2,\dots,k-1\}, the graph HDi1Di2{xi}H\cup D_{i}^{1}\cup D_{i}^{2}\cup\{x_{i}\} is either a prism or an even wheel with center xix_{i}, a contradiction. This proves that (H2,X)(H_{2},X) is a stellar connectifier, or a spiky connectifier, or a spiky alignment, or a wide alignment, as required.

Thus we may assume that for i{1,2}i\in\{1,2\}, the pair (Hi,X)(H_{i},X) is a stellar connectifier, or a spiky connectifier, or a spiky alignment, or a wide alignment. If (H1,X)(H_{1},X) is a stellar connectifier or a spiky connectifier, then for every xiX{x1,xk}x_{i}\in X\setminus\{x_{1},x_{k}\}, the graph HDi1Di2{xi}H\cup D_{i}^{1}\cup D_{i}^{2}\cup\{x_{i}\} contains a theta, a contradiction.

It follows that for i{1,2}i\in\{1,2\}, the pair (Hi,X)(H_{i},X) is an alignment. We may assume that (H1,X)(H_{1},X) and (H2,X)(H_{2},X) are either both spiky alignments or both triangular alignments, for otherwise the second outcome of the theorem holds. But now for every xiX{x1,xk}x_{i}\in X\setminus\{x_{1},x_{k}\}, the graph H{xi}H\cup\{x_{i}\} is either a theta or an even wheel, a contradiction. ∎

Refer to caption
Figure 4. Proof of Lemma 5.6.
Lemma 5.6.

Let GG be a theta-free graph and let P=p1--pkP=p_{1}\hbox{-}\cdots\hbox{-}p_{k} be a path in GG. Let DD be a connected subset of GG such that DD is anticomplete to PP. Let x,yN(P)N(D)x,y\in N(P)\cap N(D) such that xyxy is not an edge. Assume that each of xx and yy has at least two non-adjacent neighbors in PP. Let ixi_{x} be minimum and jxj_{x} be maximum such that xx is adjacent to pixp_{i_{x}} and pjxp_{j_{x}} and write Px=pix-P-pjxP_{x}=p_{i_{x}}\hbox{-}P\hbox{-}p_{j_{x}}. Suppose that yy has a neighbor in PxP_{x}. Then yy has a neighbor in NPx[pix]NPx[pjx]N_{P_{x}}[p_{i_{x}}]\cup N_{P_{x}}[p_{j_{x}}].

Proof.

Suppose that yy is anticomplete to NP[pix]NP[pjx]N_{P}[p_{i_{x}}]\cup N_{P}[p_{j_{x}}]. Let RR be a path from xx to yy with RDR^{*}\subseteq D. (See Figure 4.) Let i{ix,,jx}i\in\{{i_{x}},\dots,{j_{x}}\} be minimum and j{ix,,jx}j\in\{{i_{x}},\dots,{j_{x}}\} be maximum such that yy is adjacent to pi,pjp_{i},p_{j}. Then there there is a path PiP_{i} from xx to yy with interior in pix-P-pip_{i_{x}}\hbox{-}P\hbox{-}p_{i}, and a path PjP_{j} from xx to yy with interior in pj-P-pjxp_{j}\hbox{-}P\hbox{-}p_{j_{x}}. If j>i+1j>i+1, then PiPjRP_{i}\cup P_{j}\cup R is a theta with ends x,yx,y, a contradiction; therefore ji+1j\leq i+1. Since yy has at least two non-adjacent neighbors in PP, it follows that yy has a neighbor pzp_{z} in PPxP\setminus P_{x}. We may assume that z>jxz>j_{x}, and therefore there is a path QQ from xx to yy with Qpjx-P-pzQ^{*}\subseteq p_{j_{x}}\hbox{-}P\hbox{-}p_{z}. Since yy is anticomplete to NPx[pix]NPx[pjx]N_{P_{x}}[p_{i_{x}}]\cup N_{P_{x}}[p_{j_{x}}], we have that j<jx1j<j_{x}-1, and so PiP_{i}^{*} is anticomplete to QQ^{*}. But now PiQRP_{i}\cup Q\cup R is a theta with ends x,yx,y, a contradiction. ∎

We can now prove Theorem 5.3. The proof follows along the lines of [3], but the assumptions are different.

Proof.

Let z=122362x6z=12^{2}36^{2}x^{6} and let ϕ(x)=μ(μ(z))\phi(x)=\mu(\mu(z)), where μ\mu is as in Theorem 5.2. Applying Theorem 5.2 twice, we obtain a set ZYZ\subseteq Y with |Z|=z|Z|=z and HiDiH_{i}\subseteq D_{i} such that either

  • HiH_{i} is a path and every vertex of ZZ has a neighbor in HiH_{i}; or

  • (Hi,X)(H_{i},X) is a consistent connectifier

for every i{1,2}i\in\{1,2\}.

(6) Let i{1,2}i\in\{1,2\} and yy\in\mathbb{N}. If HiH_{i} is a path and every vertex of ZZ has a neighbor in HiH_{i}, then either some vertex of HiH_{i} has yy neighbors in ZZ, or there exists ZZZ^{\prime}\subseteq Z with |Z||Z|12y|Z^{\prime}|\geq\frac{|Z|}{12y} and a subpath HiH_{i}^{\prime} of HiH_{i} such that (Hi,Z)(H_{i}^{\prime},Z^{\prime}) is a consistent alignment.

Let Hi=h1--hkH_{i}=h_{1}\hbox{-}\cdots\hbox{-}h_{k}. We may assume that HiH_{i} is chosen minimal satisfying Theorem 5.2, and so there exist z1,zkZz_{1},z_{k}\in Z such that NHi(zj)={hj}N_{H_{i}}(z_{j})=\{h_{j}\} for j{1,k}j\in\{1,k\}.

We may assume that |NZ(h)|<y|N_{Z}(h)|<y for every hHih\in H_{i}. Let Z1Z_{1} be the set of vertices in ZZ with exactly one neighbor in HiH_{i}. Suppose that |Z1||Z|3|Z_{1}|\geq\frac{|Z|}{3}. It follows that Z1Z_{1} contains a set ZZ^{\prime} with |Z||Z1|y|Z|3y|Z^{\prime}|\geq\frac{|Z_{1}|}{y}\geq\frac{|Z|}{3y} such that no two vertices in ZZ^{\prime} have a common neighbor in HiH_{i}. We may assume that z1,zkZz_{1},z_{k}\in Z^{\prime}. Therefore, (Hi,Z)(H_{i},Z^{\prime}) is a spiky alignment, as required. Thus we may assume that |Z1|<|Z|3|Z_{1}|<\frac{|Z|}{3}.

Next, let Z2Z_{2} be the set of vertices in zZz\in Z such that either z{z1,zk}z\in\{z_{1},z_{k}\} or zz has exactly two neighbors in HiH_{i}, and they are adjacent. Suppose that |Z2||Z|3|Z_{2}|\geq\frac{|Z|}{3}. By choosing ZZ^{\prime} greedily, it follows that Z2Z_{2} contains a subset ZZ^{\prime} with the following specifications:

  • z1,zkZz_{1},z_{k}\in Z^{\prime};

  • |Z||Z2|2y|Z|6y|Z^{\prime}|\geq\frac{|Z_{2}|}{2y}\geq\frac{|Z|}{6y}; and

  • no two vertices in ZZ^{\prime} have a common neighbor in HiH_{i}.

But then (Hi,Z)(H_{i},Z^{\prime}) is a triangular alignment, as required. Thus we may assume that |Z2|<|Z|3|Z_{2}|<\frac{|Z|}{3}.

Finally, let Z3={z1,zk}(Z(Z1Z2))Z_{3}=\{z_{1},z_{k}\}\cup(Z\setminus(Z_{1}\cup Z_{2})). It follows that |Z3||Z|3|Z_{3}|\geq\frac{|Z|}{3}. Let RR be a path from z1z_{1} to zkz_{k} with RH3iR^{*}\subseteq H_{3-i}, and let HH be the hole z1-Hi-zk-R-z1z_{1}\hbox{-}H_{i}\hbox{-}z_{k}\hbox{-}R\hbox{-}z_{1}. Let zZ3{z1,zk}z\in Z_{3}\setminus\{z_{1},z_{k}\}. Define Hi(z)H_{i}(z) to be the minimal subpath of HiH_{i} containing NHi(z)N_{H_{i}}(z). Let the ends of Hi(z)H_{i}(z) be aza_{z} and bzb_{z}. Next let Bad(z)=NHi(z)[az,bz]Bad(z)=N_{H_{i}(z)}[a_{z},b_{z}]. Since HiH_{i} is a path, it follows that |Bad(z)|4|Bad(z)|\leq 4 for every zz. Since NZ(h)<yN_{Z}(h)<y for every hHih\in H_{i}, we can greedily choose ZZ3Z^{\prime}\subseteq Z_{3} with |Z||Z3|4y|Z|12y|Z^{\prime}|\geq\frac{|Z_{3}|}{4y}\geq\frac{|Z|}{12y}, where z1,zkZz_{1},z_{k}\in Z^{\prime} and such that if z,zZz,z^{\prime}\in Z^{\prime}, then zz^{\prime} is anticomplete to Bad(z)Bad(z). It follows from Lemma 5.6 that Hi(z)Hi(z)=H_{i}(z)\cap H_{i}(z^{\prime})=\emptyset for every z,zZz,z^{\prime}\in Z^{\prime}, and so (Hi,Z)(H_{i},Z^{\prime}) is a wide alignment. This proves (5).

(7) There exist ZZZ^{\prime}\subseteq Z with |Z|x2|Z^{\prime}|\geq x^{2}, and HiHiH_{i}^{\prime}\subseteq H_{i} for i=1,2i=1,2 such that (Hi,Z)(H_{i}^{\prime},Z^{\prime}) is a consistent alignment or a consistent connectifier.

If both (H1,Z)(H_{1},Z) and (H2,Z)(H_{2},Z) are consistent connectifiers, (5) holds. Thus we may assume that H1H_{1} is a path and every vertex of ZZ has a neighbor in H1H_{1}. Suppose first that some hH1h\in H_{1} has at least 36x236x^{2} neighbors in ZZ. Let H1={h}H_{1}^{\prime}=\{h\}, and let Z′′ZN(h)Z^{\prime\prime}\subseteq Z\cap N(h) with |Z′′|=36x2|Z^{\prime\prime}|=36x^{2}. If (H2,Z)(H_{2},Z) is a connectifier, we can clearly choose H2H2H_{2}^{\prime}\subseteq H_{2} such that (5) holds. So we may assume that H2H_{2} is a path and every vertex of ZZ has a neighbor in H2H_{2}. Moreover, no vertex ff of H2H_{2} has three or more neighbors in ZZ^{\prime}, for otherwise {f,h}(N(f)Z)\{f,h\}\cup(N(f)\cap Z^{\prime}) contains a theta with ends h,fh,f. Now by (5) applied with y=3y=3, there is a set Z′′ZZ^{\prime\prime}\subseteq Z^{\prime} with |Z′′|x2|Z^{\prime\prime}|\geq x^{2} such that (H2,Z′′)(H_{2},Z^{\prime\prime}) is a consistent alignment, and (5) holds. Similarly, we may assume that every hH2h\in H_{2} has fewer than 36x236x^{2} neighbors in ZZ.

Therefore, we may assume that every vertex hH1h\in H_{1} has strictly fewer than 36x236x^{2} neighbors in ZZ. Applying (5) with y=36x2y=36x^{2}, we conclude that there exists Z1ZZ_{1}\subseteq Z with |Z1|36×12x4|Z_{1}|\geq 36\times 12x^{4} and a path H1H1H_{1}^{\prime}\subseteq H_{1} such that (H1,Z1)(H_{1}^{\prime},Z_{1}) is a consistent alignment. If (H2,Z)(H_{2},Z) is a consistent connectifier, then (5) holds, so we may assume that H2H_{2} is a path and every vertex of ZZ has a neighbor in H2H_{2}. Since every vertex hH2h\in H_{2} has fewer than 36x236x^{2} neighbors in ZZ, we apply (5) with y=36x2y=36x^{2} again, and we conclude that there exists ZZ1Z^{\prime}\subseteq Z_{1} with |Z|x2|Z^{\prime}|\geq x^{2} and a path H2H2H_{2}^{\prime}\subseteq H_{2} such that (H2,Z)(H_{2}^{\prime},Z^{\prime}) is a consistent alignment. This proves (5).

(8) There exist Z^Z\hat{Z}\subseteq Z with |Z^|x|\hat{Z}|\geq x, and Hi^Hi\hat{H_{i}}\subseteq H_{i} for i=1,2i=1,2 such that (H^i,Z^)(\hat{H}_{i},\hat{Z}) is a consistent alignment or a consistent connectifier. Moreover, if neither of (H1^,Z)^,(H2^,Z^)(\hat{H_{1}},\hat{Z)},(\hat{H_{2}},\hat{Z}) is a concentrated connectifier, then the order given on Z^\hat{Z} by (H1^,Z^)(\hat{H_{1}},\hat{Z}) and (H2^,Z^)(\hat{H_{2}},\hat{Z}) is the same.

Let Z,H1,H2Z^{\prime},H_{1}^{\prime},H_{2}^{\prime} be as in (5). We may assume that neither of (H1,Z)(H_{1}^{\prime},Z^{\prime}) and (H2,Z)(H_{2}^{\prime},Z^{\prime}) is a concentrated connectifier. For i{1,2}i\in\{1,2\}, let πi\pi_{i} be the order given on ZZ^{\prime} by (Hi,Z)(H_{i}^{\prime},Z^{\prime}). By Theorem 5.4 there exists Z^Z\hat{Z}\subseteq Z^{\prime} such that (possibly reversing HiH_{i}^{\prime}) the orders πi\pi_{i} restricted to Z^\hat{Z} are the same, as required. This proves (5).

Now Theorem 5.3 follows from Lemma 5.5. ∎

We now refine Theorem 5.3 in the two special cases we need here. The first one is when D1D_{1} is a path. It is useful to state this result in terms of the following definition.

Given a graph class 𝒢\mathcal{G}, we say 𝒢\mathcal{G} is amiable if, for every integer x2x\geq 2, there exists an integer σ=σ(x)1\sigma=\sigma(x)\geq 1 such that the following holds for every graph G𝒢G\in\mathcal{G}. Assume that V(G)=D1D2YV(G)=D_{1}\cup D_{2}\cup Y where

  • YY is a stable set with |Y|=σ|Y|=\sigma,

  • D1D_{1} and D2D_{2} are components of GYG\setminus Y,

  • N(D1)=N(D2)=YN(D_{1})=N(D_{2})=Y,

  • D1=d1--dkD_{1}=d_{1}\hbox{-}\cdots\hbox{-}d_{k} is a path, and

  • for every yYy\in Y there exists i(y){1,,k}i(y)\in\{1,\dots,k\} such that N(di(y))Y={y}N(d_{i(y)})\cap Y=\{y\}.

Then there exist XYX\subseteq Y with |X|=x+2|X|=x+2, H1D1H_{1}\subseteq D_{1} and H2D2H_{2}\subseteq D_{2} such that

  1. (1)

    One of the following holds:

    • (H1,X)(H_{1},X) is a triangular alignment and (H2,X)(H_{2},X) is a stellar connectifier or a spiky connectifier;

    • (H1,X)(H_{1},X) is a spiky alignment, and (H2,X)(H_{2},X) is a triangular connectifier or a clique connectifier;

    • (H1,X)(H_{1},X) is a wide alignment and (H2,X)(H_{2},X) is a triangular connectifier or a clique connectifier; or

    • Both (H1,X)(H_{1},X) and (H2,X)(H_{2},X) are alignments, and at least one of (H1,X)(H_{1},X) and (H2,X)(H_{2},X) is not a spiky alignment, and at least one of (H1,X)(H_{1},X) and (H2,X)(H_{2},X) is not a triangular alignment.

  2. (2)

    If (H2,X)(H_{2},X) is not a concentrated connectifier, then the orders given on XX by (H1,X)(H_{1},X) and by (H2,X)(H_{2},X) are the same.

  3. (3)

    For every xXx\in X except at most two, we have that ND1(x)=NH1(x)N_{D_{1}}(x)=N_{H_{1}}(x).

Theorem 5.7.

The class 𝒞\mathcal{C} is amiable.

Proof.

Let x2x\geq 2, and let G𝒞,D1=d1--dk,D2,YG\in\mathcal{C},D_{1}=d_{1}\hbox{-}\cdots\hbox{-}d_{k},D_{2},Y be as in the definition of an amiable class. Let σ(x)=ϕ(x+4)\sigma(x)=\phi(x+4) where ϕ\phi is as in Theorem 5.3. We start with the following.

(9) For every dD1d\in D_{1}, we have that |NY(d)|4|N_{Y}(d)|\leq 4.

Suppose there exists i{1,,k}i\in\{1,\dots,k\} and y1,y2,y3,y4,y5YN(di)y_{1},y_{2},y_{3},y_{4},y_{5}\in Y\cap N(d_{i}). We may assume that i(y1)<<i(y5)i(y_{1})<\dots<i(y_{5}). By reversing D1D_{1} if necessary, we may assume that i(y3)<ii(y_{3})<i, and so i(y2)<i1i(y_{2})<i-1. Let PP be a path with ends y1,y2y_{1},y_{2} and with interior in di(y1)-D1-di(y2)d_{i(y_{1})}\hbox{-}D_{1}\hbox{-}d_{i(y_{2})}. Let RR be a path from y1y_{1} to y2y_{2} with interior in D2D_{2}. Now there is a theta in GG with ends y1,y2y_{1},y_{2} and paths PP,RR and y1-di-y2y_{1}\hbox{-}d_{i}\hbox{-}y_{2}, a contradiction. This proves (5).

Now we apply Theorem 5.3 and obtain a set XX^{\prime} with |X|=x+4|X^{\prime}|=x+4 satisfying one of the outcomes of Theorem 5.3. Let (H1,X)(H_{1},X^{\prime}) and (H2,X)(H_{2},X^{\prime}) be as in Theorem 5.3 where H1D1H_{1}\subseteq D_{1} (and so Theorem 5.3 may hold with the roles of H1H_{1} and H2H_{2} reversed). It follows from (5) that every vertex in D1D_{1} has degree at most 6 in GG and degree 2 in G[D1]G[D_{1}], and so (H1,X)(H_{1},X^{\prime}) is an alignment. Let z1,,zx+4z_{1},\dots,z_{x+4} be the order on XX^{\prime} given by (H1,X)(H_{1},X^{\prime}).

(10) There exist at most two values of i{2,,x+3}i\in\{2,\dots,x+3\} for which N(zi)D1N(zi)H1N(z_{i})\cap D_{1}\neq N(z_{i})\cap H_{1}.

Suppose there exist three such values of ii. Since H1H_{1} is a path, there exist i,j{1,,k}i,j\in\{1,\dots,k\} such that H1=di-D1-djH_{1}=d_{i}\hbox{-}D_{1}\hbox{-}d_{j}, and we may assume (by reversing D1D_{1} if necessary) that for two values p,q{2,,x+3}p,q\in\{2,\dots,x+3\}, both zpz_{p} and zqz_{q} have neighbors in d1-D1-di1d_{1}\hbox{-}D_{1}\hbox{-}d_{i-1}. Let PP be a path from zpz_{p} to zqz_{q} with Pd1-D1-di1P^{*}\subseteq d_{1}\hbox{-}D_{1}\hbox{-}d_{i-1}. Since p,q>1p,q>1, there is a path QQ from zpz_{p} to zqz_{q} with QH1diQ^{*}\subseteq H_{1}\setminus d_{i}. But now we get a theta with ends zp,zqz_{p},z_{q} and path PP, QQ, and a path from zpz_{p} to zqz_{q} with interior in D2D_{2}, a contradiction. This proves (5).

By (5), there exists X{z2,,zx+3}X\subseteq\{z_{2},\dots,z_{x+3}\} with |X|=x|X|=x and such that ND1(z)=NH1(z)N_{D_{1}}(z)=N_{H_{1}}(z) for every zZz\in Z. We obtain that (H1,X{z1,zx+4})(H_{1},X\cup\{z_{1},z_{x+4}\}) and (H2,X{z1,zx+4})(H_{2},X\cup\{z_{1},z_{x+4}\}) satisfy the first statement of Theorem 5.7. The second statement of Theorem 5.7 follows immediately from Theorem 5.3, and the third statement holds by the choice of XX. ∎

The second special case is when the clique number of GG is bounded, and YHub(G)=Y\cap\operatorname{Hub}(G)=\emptyset; it is a result from [3].

Theorem 5.8 (Abrishami, Alecu, Chudnovsky, Hajebi, Spirkl [3]).

For every pair of integers t,x1t,x\geq 1, there exists an integer τ=τ(t,x)1\tau=\tau(t,x)\geq 1 with the following property. Let G𝒞tG\in\mathcal{C}_{t} and assume that V(G)=D1D2YV(G)=D_{1}\cup D_{2}\cup Y where

  • YY is a stable set with |Y|=τ|Y|=\tau,

  • D1D_{1} and D2D_{2} are components of GYG\setminus Y, and

  • N(D1)=N(D2)=YN(D_{1})=N(D_{2})=Y.

Assume that YHub(G)=Y\cap\operatorname{Hub}(G)=\emptyset. Then there exist XYX\subseteq Y with |X|=x|X|=x, H1D1H_{1}\subseteq D_{1} and H2D2H_{2}\subseteq D_{2} (possibly with the roles of D1D_{1} and D2D_{2} reversed) such that

  • (H1,X)(H_{1},X) is a triangular connectifier or a triangular alignment;

  • (H2,X)(H_{2},X) is a stellar connectifier, or a spiky connectifier, or a spiky alignment or a wide alignment; and

  • if (H1,X)(H_{1},X) is a triangular alignment, then (H2,X)(H_{2},X) is not a wide alignment.

Moreover, if neither of (H1,X),(H2,X)(H_{1},X),(H_{2},X) is a stellar connectifier, then the orders given on XX by (H1,X)(H_{1},X) and by (H2,X)(H_{2},X) are the same.

6. Bounding the number of non-hubs in a minimal separator

The goal of this section is to prove Theorem 6.3. This is the second main ingredient of the proof of Theorem 1.3 (in addition to the machinery of Section 4). Theorem 6.3 is what allows us to iterate the construction of Section 4 for 𝒪(log|V(G)|)\mathcal{O}(\log|V(G)|) rounds (rather than just constantly many), which is what we need in the proof. We need the following definition and result. Given a graph HH, a (p)(\leq p)-subdivision of HH is a graph that arises from HH by replacing each edge uvuv of HH by a path from uu to vv of length at least 1 and at most pp; the interiors of these paths are pairwise disjoint and anticomplete.

Theorem 6.1 (Lozin and Razgon [29]).

For all positive integers pp and rr, there exists a positive integer m=m(p,r)m=m(p,r) such that every graph GG containing a (p\leq p)-subdivision of KmK_{m} as a subgraph contains either Kp,pK_{p,p} as a subgraph or a proper (p\leq p)-subdivision of Kr,rK_{r,r} as an induced subgraph.

Since graphs in 𝒞t\mathcal{C}_{t} contain neither Kt+1K_{t+1} nor K2,2K_{2,2} as an induced subgraph, we conclude that graphs in 𝒞t\mathcal{C}_{t} do not contain Kt+1,t+1K_{t+1,t+1} as a subgraph. Likewise, graphs in 𝒞t\mathcal{C}_{t} do not contain subdivisions of K2,3K_{2,3} (in other words, thetas) as an induced subgraph. Therefore, letting m=m(t+1,3)m=m(t+1,3), we conclude:

Lemma 6.2.

Let tt\in\mathbb{N}. Then there exists an m=m(t)m=m(t)\in\mathbb{N} such that the following holds: Let GG in 𝒞t\mathcal{C}_{t}. Then GG does not contain a (2)(\leq 2)-subdivision of KmK_{m} as a subgraph.

Recall that the Ramsey number R(t,s)R(t,s) is the minimum integer such that every graph on at least R(t,s)R(t,s) vertices contains either a clique of size tt or a stable set of size ss.

Theorem 6.3.

For every integer t1t\geq 1 there exists Γ=Γ(t)\Gamma=\Gamma(t) with the following property. Let G𝒞tG\in\mathcal{C}_{t} and assume that V(G)=D1D2YV(G)=D_{1}\cup D_{2}\cup Y where

  • D1D_{1} and D2D_{2} are components of GYG\setminus Y.

  • YY is a stable set.

  • N(D1)=N(D2)=YN(D_{1})=N(D_{2})=Y.

  • YHub(G)=Y\cap\operatorname{Hub}(G)=\emptyset.

  • There exist a1D1a_{1}\in D_{1} and a2D2a_{2}\in D_{2} such that a1a2a_{1}a_{2} is a |Y||Y|-banana in GG.

Then |Y|Γ|Y|\leq\Gamma.

Proof.

Let k=k(t)k=k(t) be as in Theorem 3.2. We may assume that k3k\geq 3. Let m=m(t)m=m(t) as in Lemma 6.2. Let λ=R(m,k).\lambda=R(m,k). Let τ=τ(t,λ)\tau=\tau(t,\lambda) be as in Theorem 5.8. Set Γ=τ\Gamma=\tau. Applying Theorem 5.8, we deduce that there exist H1D1H_{1}\subseteq D_{1}, H2D2H_{2}\subseteq D_{2} and XYX\subseteq Y with |X|=λ|X|=\lambda such that:

  • (H1,X)(H_{1},X) is a triangular connectifier or a triangular alignment;

  • (H2,X)(H_{2},X) is a stellar connectifier, or a spiky connectifier, or a spiky alignment, or a wide alignment; and

  • if (H1,X)(H_{1},X) is a triangular alignment, then (H2,X)(H_{2},X) is not a wide alignment.

(11) Suppose (H2,X)(H_{2},X) is a wide alignment, and let x1,,xλx_{1},\dots,x_{\lambda} be the order on XX given by H2H_{2}. Then for every i{2,,λ1}i\in\{2,\dots,\lambda-1\}, xix_{i} has exactly three neighbors in H2H_{2}, and two of them are adjacent.

Let i{2,,λ1}i\in\{2,\dots,\lambda-1\}. Let RR be a path from x1x_{1} to xλx_{\lambda} with interior in H1H_{1}. Then H=x1-H2-xλ-R-x1H=x_{1}\hbox{-}H_{2}\hbox{-}x_{\lambda}\hbox{-}R\hbox{-}x_{1} is a hole. Since (H2,X)(H_{2},X) is a wide alignment, xix_{i} has two non-adjacent neighbors in HH. Since (H,xi)(H,x_{i}) is not a wheel in GG, and xix_{i} is non-adjacent to x1x_{1} and xλx_{\lambda}, (6) follows.

Next we show:

(12) Every vertex in D1D_{1} has at most two neighbors in XX.

Suppose that there is a vertex dD1d\in D_{1} such that dd has three distinct neighbors xi,xj,xkx_{i},x_{j},x_{k} in XX; without loss of generality, we may assume that i<j<ki<j<k. We consider two cases. First, if H2H_{2} is a wide alignment, then GG contains a wheel (C,xj)(C,x_{j}) where CC is a cycle consisting of xi,d,xkx_{i},d,x_{k} and the path from xkx_{k} to xix_{i} with interior in H2H_{2}. By (6), xjx_{j} has four neighbors in CC: dd, as well as three neighbors in H2H_{2}. But xjx_{j} is not a hub in GG, a contradiction.

It follows that H2H_{2} is not a wide alignment. Now, for r,s{i,j,k}r,s\in\{i,j,k\}, let PrsP_{rs} be a path from xrx_{r} to xsx_{s} with interior in H2H_{2}. Then {d}PijPjkPik\{d\}\cup P_{ij}\cup P_{jk}\cup P_{ik} is a theta in GG (see Figure 5). This proves (6).

Refer to caption
Figure 5. Options in the proof of Theorem 6.3 (except (1) and (6) cannot be the case together).

(13) There exists SXS\subseteq X with |S|=k|S|=k, such that for every s,sSs,s^{\prime}\in S and for every path PP from ss to ss^{\prime} with interior in D1D_{1} we have that |E(P)|>2|E(P)|>2.

Define a graph HH with vertex set XX such that xxE(H)xx^{\prime}\in E(H) if and only if there is a path of length two with ends x,xx,x^{\prime} and interior in D1D_{1}. Since |X|=λ=R(m,k)|X|=\lambda=R(m,k), there is SXS\subseteq X with |S|=k|S|=k such that SS is either a clique or a stable set in HH.

If SS is a stable set, then (6) holds, so we may assume that SS is a clique of size mm in HH. Write S={s1,,sm}S=\{s_{1},\dots,s_{m}\}. For i,j{1,,m}i,j\in\{1,\dots,m\} with iji\neq j, we let dijd_{ij} be a common neighbor of sis_{i} and sjs_{j} in D1D_{1} (which exists, since SS is a clique in HH). By (6), no vertex has three neighbors in SS, and so the vertices dijd_{ij} are pairwise distinct. But now S{dij:i,j{1,,m},ij}S\cup\{d_{ij}:i,j\in\{1,\dots,m\},i\neq j\} contains a (2)(\leq 2)-subdivision of KmK_{m} as a subgraph, contrary to Lemma 6.2. This is proves (6).

From now on let SXS\subseteq X be as in (6). Let GG^{\prime} be the graph obtained from D1SD_{1}\cup S by adding a new vertex vv with N(v)=SN(v)=S. We need the following two facts about GG^{\prime}:

(14) SHub(G)=S\cap\operatorname{Hub}(G^{\prime})=\emptyset.

Suppose not, let sSs\in S and let (H,s)(H,s) be a wheel in GG^{\prime}. Since SHub(G)=S\cap\operatorname{Hub}(G)=\emptyset, it follows that vHv\in H. Let NH(v)={s,s′′}N_{H}(v)=\{s^{\prime},s^{\prime\prime}\}; then HvH\setminus v is a path RR from ss^{\prime} to s′′s^{\prime\prime} with RD1R^{*}\subseteq D_{1}. Since (H,s)(H,s) is a wheel, it follows that ss has two non-adjacent neighbors in RR. Consequently, there is a path PP^{\prime} from ss to ss^{\prime} and a path P′′P^{\prime\prime*} form ss to s′′s^{\prime\prime}, both with interior in RR, such that PP^{\prime*} is disjoint from and anticomplete to P′′P^{\prime\prime*}.

Let TT be a path from ss^{\prime} to s′′s^{\prime\prime} with interior in H2H_{2}. Then H=s-R-s′′-T-sH^{\prime}=s^{\prime}\hbox{-}R\hbox{-}s^{\prime\prime}\hbox{-}T\hbox{-}s^{\prime} is a hole. Since XHub(G)=X\cap\operatorname{Hub}(G)=\emptyset, it follows that (H,s)(H^{\prime},s) is not a wheel in GG, and so ss is anticomplete to TT.

Suppose first that (H2,X)(H_{2},X) is a wide alignment. Since ss is anticomplete to TT, it follows that (reversing the order on XX if necessary, and exchanging the roles of s,s′′s^{\prime},s^{\prime\prime} if necessary) s,s,s′′s,s^{\prime},s^{\prime\prime} appear in this order in the order given by H2H_{2} on XX. Now we get a theta with ends s,ss,s^{\prime} and paths s-P-ss\hbox{-}P^{\prime}\hbox{-}s^{\prime}, s-P′′-s′′-H2-ss\hbox{-}P^{\prime\prime}\hbox{-}s^{\prime\prime}\hbox{-}H_{2}\hbox{-}s^{\prime}, and s-H2-ss\hbox{-}H_{2}\hbox{-}s^{\prime}, a contradiction. This proves that (H2,X)(H_{2},X) is not a wide alignment.

Since (H2,X)(H_{2},X) is not a wide alignment, there is a vertex aH2a\in H_{2}, and three paths Q,Q,Q′′Q,Q^{\prime},Q^{\prime\prime} all with interior in H2H_{2}, where QQ is from aa to ss, QQ^{\prime} is from aa to ss^{\prime}, and Q′′Q^{\prime\prime} is from aa to s′′s^{\prime\prime}, and the sets QaQ\setminus a, QaQ^{\prime}\setminus a and Q′′aQ^{\prime\prime}\setminus a are pairwise disjoint and anticomplete to each other. Note that T=s-Q-a-Q′′-s′′T=s^{\prime}\hbox{-}Q^{\prime}\hbox{-}a\hbox{-}Q^{\prime\prime}\hbox{-}s^{\prime\prime}. Since ss is anticomplete to TT, it follows that ss is non-adjacent to aa. But now we get a theta with ends a,sa,s and path s-Q-as\hbox{-}Q\hbox{-}a, s-P-s-Q-as\hbox{-}P^{\prime}\hbox{-}s^{\prime}\hbox{-}Q^{\prime}\hbox{-}a, and s-P′′-s′′-Q′′-as\hbox{-}P^{\prime\prime}\hbox{-}s^{\prime\prime}\hbox{-}Q^{\prime\prime}\hbox{-}a, a contradiction. This proves (6).

Refer to caption
Figure 6. Proof of (6).

(15) G𝒞tG^{\prime}\subseteq\mathcal{C}_{t}.

Suppose not. Let ZGZ\subseteq G^{\prime} be such that ZZ is a KtK_{t}, a C4C_{4}, a theta, a prism or an even wheel. Then vZv\in Z. Since NZ(v)N_{Z}(v) is a stable set, it follows that ZZ is not a KtK_{t}. Suppose first that |NZ(v)|=2|N_{Z}(v)|=2; then NZ(v)={s,s}SN_{Z}(v)=\{s,s^{\prime}\}\subseteq S. Suppose that ZZ is a C4C_{4}. Then there is dD1d\in D_{1} such that Z=v-s-d-s-vZ=v\hbox{-}s\hbox{-}d\hbox{-}s^{\prime}\hbox{-}v, contrary to the fact that SS was chosen as in (6). It follows that ZZ is a theta, a prism or an even wheel. Let QQ be a path from ss to ss^{\prime} with interior in D2D_{2}. Now Z=(Zv)QZ^{\prime}=(Z\setminus v)\cup Q is an induced subgraph of GG. Moreover, if ZZ is a theta then ZZ^{\prime} is a theta, and if ZZ is a prism then ZZ^{\prime} is a prism. Since G𝒞G\in\mathcal{C}, it follows that ZZ is an even wheel; denote it by (H,w)(H,w). Then vHv\in H. By (6), wSw\not\in S. It follows that NZ(w)=NZ(w)N_{Z}(w)=N_{Z^{\prime}}(w^{\prime}), and so ZZ^{\prime} is an even wheel, contrary to the fact that G𝒞G\in\mathcal{C}.

This proves that |NZ(v)|>2|N_{Z}(v)|>2. Since NZ(v)N_{Z}(v) is a stable set, it follows that ZZ is not a prism. Suppose that ZZ is a theta. Then the ends of ZZ are vv and dD1d\in D_{1}; let the paths of ZZ be Z1,Z2,Z3Z_{1},Z_{2},Z_{3} where NZi(v)=siN_{Z_{i}}(v)=s_{i}. By renumbering Z1,Z2,Z3Z_{1},Z_{2},Z_{3}, we may assume that s1,s2,s3s_{1},s_{2},s_{3} appear in this order in the order on XX given by H1H_{1}. Then for every i,j{1,,3}i,j\in\{1,\dots,3\}, there is a unique path Fij1F^{1}_{ij} from sis_{i} to sjs_{j} with interior in H1H_{1}. Similarly, for all i,j{1,,3}i,j\in\{1,\dots,3\} there is a unique path Fij2F^{2}_{ij} from sis_{i} to sjs_{j} with interior in H2H_{2}. Observe that F121F131F231F122F132F232F^{1}_{12}\cup F^{1}_{13}\cup F^{1}_{23}\cup F^{2}_{12}\cup F^{2}_{13}\cup F^{2}_{23} contains a unique pyramid Σ\Sigma with the following specifications:

  • The paths of Σ\Sigma are P1,P2,P3P_{1},P_{2},P_{3}, where siPis_{i}\in P_{i}.

  • Denote the apex of Σ\Sigma by aa, and let the base of Σ\Sigma be b1b2b3b_{1}b_{2}b_{3} where bib_{i} is an end of PiP_{i}. Then either aH2a\in H_{2}, or (H2,X)(H_{2},X) is a wide alignment and a=s2a=s_{2}.

  • b1,b3H1b_{1},b_{3}\in H_{1}, and b2H1{s2}b_{2}\in H_{1}\cup\{s_{2}\}.

Let Σ2=ΣD1\Sigma_{2}=\Sigma\setminus D_{1}. If dd is non-adjacent to aa, then (Zv)Σ2(Z\setminus v)\cup\Sigma_{2} is a theta with ends a,da,d and paths d-Zi-si-Pi-ad\hbox{-}Z_{i}\hbox{-}s_{i}\hbox{-}P_{i}\hbox{-}a, contrary to the fact that G𝒞tG\in\mathcal{C}_{t}. This proves that dd is adjacent to aa. Consequently, a=s2a=s_{2} and (H2,X)(H_{2},X) is a wide alignment. But now H=s1-Z1-d-Z3-s3-F132-s1H=s_{1}\hbox{-}Z_{1}\hbox{-}d\hbox{-}Z_{3}\hbox{-}s_{3}\hbox{-}F^{2}_{13}\hbox{-}s_{1} is a hole, and, since s2s_{2} is adjacent to dd, (H,s2)(H,s_{2}) is a wheel, contrary to the fact that SHub(G)=S\cap\operatorname{Hub}(G)=\emptyset. This proves that ZZ is not a theta.

Consequently, ZZ is an even wheel (H,w)(H,w). Since |NZ(v)|>2|N_{Z}(v)|>2, it follows that either v=wv=w, or vv is adjacent to ww. If vv is adjacent to ww, then wSw\in S, contrary to (6). It now follows that w=vw=v, and so HH is a hole in D1SD_{1}\cup S, where |SH|=l3|S\cap H|=l\geq 3. Let SH={s1,,sl}S\cap H=\{s_{1},\dots,s_{l}\} where s1,sls_{1},\dots s_{l} appear in this order in the order given on SS by (H2,X)(H_{2},X). If (H2,X)(H_{2},X) is a spiky alignment or a wide alignment, or a spiky connectifier, then we get a theta with ends s1,s2s_{1},s_{2} two of whose paths are contained in HH, and the third is the path from s1s_{1} to s2s_{2} with interior in H2H_{2}. It follows that (H2,X)(H_{2},X) is a stellar connectifier. Therefore there exist paths Q1,,QlQ_{1},\dots,Q_{l}, all with a common end aH2a\in H_{2}, such that QiQ_{i} is from aa to sis_{i}, QiH2Q_{i}^{*}\subseteq H_{2}, and the sets Q1a,,QkaQ_{1}\setminus a,\dots,Q_{k}\setminus a are pairwise disjoint and anticomplete to each other. Since (H,a)(H,a) is a not an even wheel in GG, we deduce that aa is not complete to {s1,,sl}\{s_{1},\dots,s_{l}\}; let si{s1,,sl}s_{i}\in\{s_{1},\dots,s_{l}\} be non-adjacent to aa. Since l3l\geq 3, there exist distinct p,q{1,,l}{i}p,q\in\{1,\dots,l\}\setminus\{i\} and paths RqR_{q} and RpR_{p} of HH such that

  • RpR_{p} is from sis_{i} to sps_{p};

  • RqR_{q} is from sis_{i} to sqs_{q};

  • RpRq=R_{p}^{*}\cap R_{q}^{*}=\emptyset; and

  • Rp{s1,,sl}=Rq{s1,sl}=R_{p}^{*}\cap\{s_{1},\dots,s_{l}\}=R_{q}^{*}\cap\{s_{1}\dots,s_{l}\}=\emptyset.

Now we get a theta with ends si,as_{i},a and path si-Qi-as_{i}\hbox{-}Q_{i}\hbox{-}a, si-Rp-sp-Qp-as_{i}\hbox{-}R_{p}\hbox{-}s_{p}\hbox{-}Q_{p}\hbox{-}a and si-Rq-sq-Qq-as_{i}\hbox{-}R_{q}\hbox{-}s_{q}\hbox{-}Q_{q}\hbox{-}a. This proves (6).

Observe that a1va_{1}v is a kk-banana in GG^{\prime}. But by (6), G𝒞tG^{\prime}\in\mathcal{C}_{t}, and by (6), NG(v)Hub(G)=N_{G^{\prime}}(v)\cap\operatorname{Hub}(G^{\prime})=\emptyset, contrary to Theorem 3.2. ∎

7. The proof of Theorem 1.3

We can now prove our first main result. The “big picture” of the proof is similar to [3] and [5], but the context here is different. For the remainder of the paper, all logarithms are taken in base 2. We start with the following theorem from [5]:

Theorem 7.1 (Abrishami, Chudnovsky, Hajebi, Spirkl [5]).

Let tt\in\mathbb{N}, and let GG be (theta, KtK_{t})-free with |V(G)|=n|V(G)|=n. There exist an integer d=d(t)d=d(t) and a partition (S1,,Sk)(S_{1},\dots,S_{k}) of V(G)V(G) with the following properties:

  1. (1)

    kd4lognk\leq\frac{d}{4}\log n.

  2. (2)

    SiS_{i} is a stable set for every i{1,,k}i\in\{1,\dots,k\}.

  3. (3)

    For every i{1,,k}i\in\{1,\dots,k\} and vSiv\in S_{i} we have degGj<iSj(v)d\deg_{G\setminus\bigcup_{j<i}S_{j}}(v)\leq d.

Let G𝒞tG\in\mathcal{C}_{t} be a graph and let a,bV(G)a,b\in V(G). A hub-partition with respect to abab of GG is a partition S1,,SkS_{1},\dots,S_{k} of Hub(G){a,b}\operatorname{Hub}(G)\setminus\{a,b\} as in Theorem 7.1; we call kk the order of the partition. We call the hub-dimension of (G,ab)(G,ab) (denoting it by hdim(G,ab)\operatorname{hdim}(G,ab)) the smallest kk such that GG has a hub-partition of order kk with respect to abab.

For the remainder of this section, let us fix tt\in\mathbb{N}. Let d=d(t)d=d(t) be as in Theorem 7.1. Let m=k+2dm=k+2d where k=k(t)k=k(t) is as in Theorem 3.2. Let Ct=(4m+2)(m1)C_{t}=(4m+2)(m-1). Let Γ=max{k,Γ(t)}\Gamma=\max\{k,\Gamma(t)\} where Γ(t)\Gamma(t) is as in Theorem 6.3.

Since in view of Theorem 7.1, we have hdim(G,ab)d4logn\operatorname{hdim}(G,ab)\leq\frac{d}{4}\log n for all a,bV(G)a,b\in V(G), Theorem 1.3 follows immediately from the next result:

Theorem 7.2.

Let G𝒞tG\in\mathcal{C}_{t} and with |V(G)|=n|V(G)|=n and let a,bV(G)a,b\in V(G). Then abab is not a Ct+8m2Γhdim(G,ab)C_{t}+8m^{2}\Gamma\operatorname{hdim}(G,ab)-banana in GG.

Proof.

Suppose that abab is a Ct+8m2Γhdim(G,ab)C_{t}+8m^{2}\Gamma\operatorname{hdim}(G,ab)-banana in GG. We will get a contradiction by induction on hdim(G,ab)\operatorname{hdim}(G,ab), and for fixed hdim\operatorname{hdim} by induction on nn. Suppose that hdim(G,ab)=0\operatorname{hdim}(G,ab)=0. Then Hub(G){a,b}\operatorname{Hub}(G)\subseteq\{a,b\} and by Theorem 3.3, we have that there is no kk-banana in GG. Now the statement holds since k<Ctk<C_{t}. Thus we may assume hdim(G,ab)>0\operatorname{hdim}(G,ab)>0.

(16) We may assume that GG admits no clique cutset.

Suppose CC is a clique cutset in GG. Since GG is KtK_{t}-free, it follows that there is a component DD of GCG\setminus C such that abab is Ct+8m2Γhdim(G,ab)C_{t}+8m^{2}\Gamma\operatorname{hdim}(G,ab)-banana in G=DCG^{\prime}=D\cup C. But |V(G)|<n|V(G^{\prime})|<n and hdim(G,ab)hdim(G,ab)\operatorname{hdim}(G^{\prime},ab)\leq\operatorname{hdim}(G,ab), consequently we get contradiction (inductively on nn). This proves (7).

Let S1,,SqS_{1},\dots,S_{q} be a hub-partition of GG with respect to abab and with q=hdim(G,ab)q=\operatorname{hdim}(G,ab). We now use notation and terminology from Section 4 (note that the definitions of kk and mm agree; we use dd, aa, bb, as defined above). It follows from the definition of S1S_{1} that every vertex in S1S_{1} is dd-safe. Let (T,χ)(T,\chi) be an mm-atomic tree decomposition of GG. Let t1,t2t_{1},t_{2} be as in Theorem 2.6. and let β=β(S1)\beta=\beta(S_{1}) and βA(S1)\beta^{A}(S_{1}) be as in Section 4. Then by Lemma 4.1, we have a,bβA(S1)a,b\in\beta^{A}(S_{1}). By Theorem 4.10(5), we have that S1Hub(βA(S1))=S_{1}\cap\operatorname{Hub}(\beta^{A}(S_{1}))=\emptyset and S2Hub(βA(S1)),,SqHub(βA(S1))S_{2}\cap\operatorname{Hub}(\beta^{A}(S_{1})),\dots,S_{q}\cap\operatorname{Hub}(\beta^{A}(S_{1})) is a hub-partition of βA(S1)\beta^{A}(S_{1}) with respect to abab. It follows that hdim(βA(S1),ab)q1\operatorname{hdim}(\beta^{A}(S_{1}),ab)\leq q-1. Inductively (on hdim(,ab)\operatorname{hdim}(\cdot,ab)), we have that abab is not a Ct+8m2Γ(q1)C_{t}+8m^{2}\Gamma(q-1)-banana in βA(S1)\beta^{A}(S_{1}).

Our first goal is to prove:

(17) There is a set YβY^{\prime}\subseteq\beta such that

  1. (1)

    YY^{\prime} separates aa from bb in β\beta;

  2. (2)

    |Y(χ(t1)χ(t2))|Ct+8m2Γ(q1)+4m2Γ|Y^{\prime}\cap(\chi(t_{1})\cup\chi(t_{2}))|\leq C_{t}+8m^{2}\Gamma(q-1)+4m^{2}\Gamma; and

  3. (3)

    |Δ(Y)|(4m+1)(m1)Γ|\Delta(Y^{\prime})|\leq(4m+1)(m-1)\Gamma.

Since abab is not a (Ct+8m2Γ(q1))(C_{t}+8m^{2}\Gamma(q-1))-banana in βA(S1)\beta^{A}(S_{1}), Theorem 2.1 implies that there is a separation (X,Y,Z)(X,Y,Z) of βA(S1)\beta^{A}(S_{1}) such that aXa\in X and bZb\in Z and |Y|Ct+8m2Γ(q1)|Y|\leq C_{t}+8m^{2}\Gamma(q-1). We may assume that (X,Y,Z)(X,Y,Z) is chosen with |Y||Y| as small as possible. Let D1D_{1} be the component of XX such that aD1a\in D_{1}, and and let D2D_{2} be the component of ZZ such that bD2b\in D_{2}. It follows from the minimality of |Y||Y| that N(D1)=N(D2)=YN(D_{1})=N(D_{2})=Y and abab is a |Y||Y|-banana in D1D2YD_{1}\cup D_{2}\cup Y.

We claim that |YS1|Γ|Y\cap S_{1}|\leq\Gamma. Let G=D1D2(YS1)G^{\prime}=D_{1}\cup D_{2}\cup(Y\cap S_{1}). Again by the minimality of YY, it follows that abab is a |YS1||Y\cap S_{1}|-banana in GG^{\prime}. If abab is not a kk-banana in GG^{\prime}, then |YS1|<k|Y\cap S_{1}|<k and the claim follows immediately since kΓk\leq\Gamma; thus we may assume that abab is a kk-banana in GG^{\prime}. Now since S1S_{1} is a stable set and since no vertex of S1S_{1} is a hub in βA(S1)\beta^{A}(S_{1}), applying Theorem 6.3 to GG^{\prime} we deduce that |YS1|Γ|Y\cap S_{1}|\leq\Gamma, as required. This proves the claim.

Next we claim that |δ1(Y)δ2(Y)|Γ|\delta_{1}(Y)\cup\delta_{2}(Y)|\leq\Gamma, and |Δ(Y)|(m1)Γ|\Delta(Y)|\leq(m-1)\Gamma. Write β0=χ(t1)χ(t2)\beta_{0}=\chi(t_{1})\cup\chi(t_{2}). Observe that for distinct t,tδ1(Y)δ2(Y)t,t^{\prime}\in\delta_{1}(Y)\cup\delta_{2}(Y), say with tδi(Y)t\in\delta_{i}(Y) and tδi(Y)t^{\prime}\in\delta_{i^{\prime}}(Y) for i,i{1,2}i,i^{\prime}\in\{1,2\}, the sets Gtitβ0G_{t_{i}\rightarrow t}\setminus\beta_{0} and Gtitβ0G_{t_{i^{\prime}}\rightarrow t^{\prime}}\setminus\beta_{0} are disjoint and anticomplete to each other. Let WW be a subset of YY such that for all i{1,2}i\in\{1,2\} and tδi(Y)t\in\delta_{i}(Y), we have |W(Gtitβ0))|=1|W\cap(G_{t_{i}\rightarrow t}\setminus\beta_{0}))|=1. Then WW is stable set and |W|=|δ1(Y)δ2(Y)||W|=|\delta_{1}(Y)\cup\delta_{2}(Y)|. Let G=D1D2WG^{\prime}=D_{1}\cup D_{2}\cup W. If follows from the minimality of YY that abab is a |W||W|-banana in GG^{\prime}. If abab is not a kk-banana in GG^{\prime}, then |W|<k|W|<k and the claim follows immediately since kΓk\leq\Gamma; thus we may assume that abab is a kk-banana in GG^{\prime}. By Theorem 4.4(3), WHub(β)=W\cap\operatorname{Hub}(\beta)=\emptyset. Now since WW is a stable set, applying Theorem 6.3 to GG^{\prime} we deduce that |W|Γ|W|\leq\Gamma, as required. Since adh(T,χ)m1\operatorname{adh}(T,\chi)\leq m-1, it follows that |Δ(Y)|(m1)Γ|\Delta(Y)|\leq(m-1)\Gamma. This proves the claim.

Now let YY^{\prime} be as in Theorem 4.11. Then YY^{\prime} separates aa from bb in β\beta. Moreover, by Theorem 4.11,

|Y(χ(t1)χ(t2))||Y|+4m(m1)|YCore(S1)|.|Y^{\prime}\cap(\chi(t_{1})\cup\chi(t_{2}))|\leq|Y|+4m(m-1)|Y\cap\operatorname{Core}(S_{1})|.

Since |YCore(S1)||YS1|Γ|Y\cap\operatorname{Core}(S_{1})|\leq|Y\cap S_{1}|\leq\Gamma, it follows that

|Y(χ(t1)χ(t2))|Ct+8m2Γ(q1)+4m2Γ.|Y^{\prime}\cap(\chi(t_{1})\cup\chi(t_{2}))|\leq C_{t}+8m^{2}\Gamma(q-1)+4m^{2}\Gamma.

Finally, by Theorem 4.11, |Δ(Y)Δ(Y)|4m(m1)|YCore(S1)||\Delta(Y^{\prime})\setminus\Delta(Y)|\leq 4m(m-1)|Y\cap\operatorname{Core}(S_{1})|. Since |Δ(Y)|(m1)Γ|\Delta(Y)|\leq(m-1)\Gamma and |YCore(S1)|Γ|Y\cap\operatorname{Core}(S_{1})|\leq\Gamma, it follows that |Δ(Y)|(4m+1)(m1)Γ|\Delta(Y^{\prime})|\leq(4m+1)(m-1)\Gamma. This proves (7).

Let YY^{\prime} be as in (7), and let (X,Y,Z)(X^{\prime},Y^{\prime},Z^{\prime}) be a separation of β\beta such that aXa\in X^{\prime} and bZb\in Z^{\prime}. Let Y′′Y^{\prime\prime} be obtained from YY^{\prime} as in Theorem 4.12. Then Y′′Y^{\prime\prime} separates aa from bb in GG. Recall that by (7), we have |Δ(Y)|(4m+1)(m1)Γ|\Delta(Y^{\prime})|\leq(4m+1)(m-1)\Gamma. Now since adh(T,χ)<m\operatorname{adh}(T,\chi)<m and since |S1bad|3|{S_{1}}_{bad}|\leq 3, it follows from Theorem 4.12 that

|Y′′|\displaystyle|Y^{\prime\prime}| |Y(χ(t1)χ(t2))|+(4m+1)(m1)Γ+2(m1)+3\displaystyle\leq|Y^{\prime}\cap(\chi(t_{1})\cup\chi(t_{2}))|+(4m+1)(m-1)\Gamma+2(m-1)+3
|Y(χ(t1)χ(t2))|+4m2Γ.\displaystyle\leq|Y^{\prime}\cap(\chi(t_{1})\cup\chi(t_{2}))|+4m^{2}\Gamma.

Since |Y(χ(t1)χ(t2))|Ct+8m2Γ(q1)+4m2Γ|Y^{\prime}\cap(\chi(t_{1})\cup\chi(t_{2}))|\leq C_{t}+8m^{2}\Gamma(q-1)+4m^{2}\Gamma by (7), we deduce that |Y′′|Ct+8m2Γq|Y^{\prime\prime}|\leq C_{t}+8m^{2}\Gamma q as required. ∎

8. Dominated balanced separators

Several more steps are needed to complete the proof of Theorem 1.5. We take the first one in this section. The goal of this section is to prove Theorem 1.4, which we restate:

Theorem 8.1.

There is an integer dd with the following property. Let G𝒞G\in\mathcal{C} and let ww be a normal weight function on GG. Then there exists YV(G)Y\subseteq V(G) such that

  • |Y|d|Y|\leq d, and

  • N[Y]N[Y] is a ww-balanced separator in GG.

We start with two decomposition theorems that will become the engine for obtaining separators. The first is a more explicit version of Theorem 3.1; it shows that the presence of a wheel in the graph forces a decomposition that is an extension of what can be locally observed in the wheel (see [4] for detailed treatment of this concept).

Theorem 8.2 (Addario-Berry, Chudnovsky, Havet, Reed, Seymour [7], da Silva, Vušković [17]).

Let GG be a C4C_{4}-free odd-signable graph that contains a proper wheel (H,x)(H,x) that is not a universal wheel. Let x1x_{1} and x2x_{2} be the endpoints of a long sector QQ of (H,x)(H,x). Let WW be the set of all vertices hh in HN(x)H\cap N(x) such that the subpath of H{x1}H\setminus\{x_{1}\} from x2x_{2} to hh contains an even number of neighbors of xx, and let Z=H(QN(x))Z=H\setminus(Q\cup N(x)). Let N=N(x)WN^{\prime}=N(x)\setminus W. Then, N{x}N^{\prime}\cup\{x\} is a cutset of GG that separates QQ^{*} from WZW\cup Z.

The second is a similar kind of theorem, but we start with a pyramid rather than a wheel.

Theorem 8.3.

Let G𝒞G\in\mathcal{C}. Let Σ\Sigma be a pyramid with paths P1,P2,P3P_{1},P_{2},P_{3}, apex aa, and base b1b2b3b_{1}b_{2}b_{3} in GG and let i,j{1,2,3}i,j\in\{1,2,3\} be distinct. For i{1,,3}i\in\{1,\dots,3\}, let aia_{i} be the neighbor of aa in PiP_{i}. Let PP be a path from a vertex of Pi{a,ai,bi}P_{i}\setminus\{a,a_{i},b_{i}\} to a vertex of Pj{a,aj,bj}P_{j}\setminus\{a,a_{j},b_{j}\}. Then at least one one of a,b1,b2,b3a,b_{1},b_{2},b_{3} has a neighbor in PP^{*}.

Proof.

It follows from the defintion of PP that there exist distinct i,j{1,2,3}i,j\in\{1,2,3\} and a subpath P=p1-pkP^{\prime}=p_{1}\hbox{-}\cdots p_{k} of PP such that p1p_{1} has a neighbor in Pi{a,ai,bi}P_{i}\setminus\{a,a_{i},b_{i}\}, and pkp_{k} has a neighbor in Pj{a,aj,bj}P_{j}\setminus\{a,a_{j},b_{j}\}. Let PP^{\prime} be chosen minimal with this property. It follows that PPP^{\prime}\subseteq P^{*}, and so we may assume that {a,b1,b2,b3}\{a,b_{1},b_{2},b_{3}\} is anticomplete to PP^{\prime}. We may also assume i=1i=1 and j=3j=3. See Figure 7. It follows that and a1b1a_{1}\neq b_{1}, a3b3a_{3}\neq b_{3}.

Refer to caption
Figure 7. Proof of Theorem 8.3.

(18) PP^{\prime} is anticomplete to P2a2P_{2}\setminus a_{2}.

Suppose not; then some vertex in PP^{\prime} has a neighbor in P2{a,a2,b2}P_{2}\setminus\{a,a_{2},b_{2}\}. It follows from the minimality of PP^{\prime} that k=1k=1. Then p1p_{1} has a neighbor in each of the paths P1b1,P2b2P_{1}\setminus b_{1},P_{2}\setminus b_{2} and P3b3P_{3}\setminus b_{3}. Now we get a theta with ends p1,ap_{1},a whose paths are subpaths of P1,P2,P3P_{1},P_{2},P_{3}, a contradiction. This proves (8).

(19) If k>1k>1 and a1a_{1} is adjacent to pkp_{k}, then a1a_{1} has a neighbor in PpkP^{\prime}\setminus p_{k}.

Suppose that a1a_{1} is adjacent to pkp_{k} and anticomplete to PpkP^{\prime}\setminus p_{k}. Since a2-a-a1-pk-a2a_{2}\hbox{-}a\hbox{-}a_{1}\hbox{-}p_{k}\hbox{-}a_{2} is not a C4C_{4} in GG, and by (8), it follows that pkp_{k} is anticomplete to P2P_{2}. If a2a_{2} has a neighbor in Pp1P^{\prime}\setminus p_{1}, then we get a theta with ends a1,a2a_{1},a_{2}, two of whose paths are contained in the hole b1-P1-a-P2-b2-b1b_{1}\hbox{-}P_{1}\hbox{-}a\hbox{-}P_{2}\hbox{-}b_{2}\hbox{-}b_{1}, and third one has interior on Pp1P^{\prime}\setminus p_{1}, a contradiction. Consequently, a2a_{2} is anticomplete to Pp1P^{\prime}\setminus p_{1}. Next suppose that a3a_{3} has a neighbor in PP^{\prime}. Since a1-a-a3-pk-a1a_{1}\hbox{-}a\hbox{-}a_{3}\hbox{-}p_{k}\hbox{-}a_{1} is not a C4C_{4} in GG, it follows that a3a_{3} is not adjacent to pkp_{k}. Now we get a theta with ends a3,pka_{3},p_{k} and path pk-a1-a-a3p_{k}\hbox{-}a_{1}\hbox{-}a\hbox{-}a_{3}, a path with interior in PP^{\prime} and a path with interior in P3P_{3}. This proves that a3a_{3} is anticomplete to PP^{\prime}.

Let xx be the neighbor of a1a_{1} in P1aP_{1}\setminus a. Suppose that p1p_{1} has a neighbor in P1{a,a1,x}P_{1}\setminus\{a,a_{1},x\}. Then there is a path TT from p1p_{1} to aa with T(P1P2){a,a1,x}T^{*}\subseteq(P_{1}\cup P_{2})\setminus\{a,a_{1},x\} and we get a theta with ends pk,ap_{k},a and paths pk-a1-ap_{k}\hbox{-}a_{1}\hbox{-}a and pk-P-p1-T-ap_{k}\hbox{-}P^{\prime}\hbox{-}p_{1}\hbox{-}T\hbox{-}a, and a path with interior in P3P_{3}^{*}, a contradiction. It follows that p1p_{1} is anticomplete to P1{a1,a,x}P_{1}\setminus\{a_{1},a,x\} and therefore p1p_{1} is adjacent to xx. Since a1a_{1} is anticomplete to PpkP^{\prime}\setminus p_{k} we deduce that NP1(p1)={x}N_{P_{1}}(p_{1})=\{x\}. Now we get a theta with ends x,pkx,p_{k} and paths x-a1-pkx\hbox{-}a_{1}\hbox{-}p_{k}, x-P-pkx\hbox{-}P^{\prime}\hbox{-}p_{k} and x-P1-b1-b3-P3-pkx\hbox{-}P_{1}\hbox{-}b_{1}\hbox{-}b_{3}\hbox{-}P_{3}\hbox{-}p_{k}, a contradiction. This proves (8).

(20) Either NP(a1){p1}N_{P^{\prime}}(a_{1})\subseteq\{p_{1}\} or NP(a3){pk}N_{P^{\prime}}(a_{3})\subseteq\{p_{k}\}.

Suppose not. Then k>1k>1. If both a1a_{1} and a3a_{3} have a neighbor in PP^{\prime*}, then there is a theta in GG with ends a1,a3a_{1},a_{3}, and paths a1-a-a3,a1-P1-b1-b3-P3-a3a_{1}\hbox{-}a\hbox{-}a_{3},a_{1}\hbox{-}P_{1}\hbox{-}b_{1}\hbox{-}b_{3}\hbox{-}P_{3}\hbox{-}a_{3} and a1-P-a3a_{1}\hbox{-}P^{\prime}\hbox{-}a_{3}, a contradiction. By symmetry we may assume that a1a_{1} is anticomplete to PP^{\prime*}; now by (8), it follows that a1a_{1} is adjacent to both p1p_{1} and pkp_{k}. Since a1-p1-a3-a-a1a_{1}\hbox{-}p_{1}\hbox{-}a_{3}\hbox{-}a\hbox{-}a_{1} is not a C4C_{4}, we deduce that a3a_{3} is non-adjacent to p1p_{1}. Similarly a3a_{3} is non-adjacent to pkp_{k}. It follows that a3a_{3} has a neighbor in PP^{\prime*}. Let SS be a path from a3a_{3} to pkp_{k} with interior in PP^{\prime*}. Now we get a theta with ends a3,pka_{3},p_{k} and path a3-S-pka_{3}\hbox{-}S\hbox{-}p_{k}, a3-a-a1-pka_{3}\hbox{-}a\hbox{-}a_{1}\hbox{-}p_{k} and a path with interior in P3P_{3}, a contradiction. This proves (8).

(21) If a1a_{1} has a neighbor in Pp1P^{\prime}\setminus p_{1}, then a2a_{2} is anticomplete to PP^{\prime}.

Suppose that a1a_{1} has a neighbor in Pp1P^{\prime}\setminus p_{1} and a2a_{2} has a neighbor in PP^{\prime}. Then k>1k>1. If both a1a_{1} and a2a_{2} have a neighbor in Pp1P^{\prime}\setminus p_{1}, then there is a theta in GG with ends a1,a2a_{1},a_{2}, and paths a1-a-a2,a1-P1-b1-b2-P2-a2a_{1}\hbox{-}a\hbox{-}a_{2},a_{1}\hbox{-}P_{1}\hbox{-}b_{1}\hbox{-}b_{2}\hbox{-}P_{2}\hbox{-}a_{2} and a1-P-a2a_{1}\hbox{-}P^{\prime}\hbox{-}a_{2}, a contradiction. This proves that NP(a2)={p1}N_{P^{\prime}}(a_{2})=\{p_{1}\}.

Let S1S_{1} be a path from a1a_{1} to p1p_{1} with S1PS_{1}^{*}\subseteq P^{\prime}. Since a1-p1-a2-a-a1a_{1}\hbox{-}p_{1}\hbox{-}a_{2}\hbox{-}a\hbox{-}a_{1} is not a C4C_{4}, it follows that a1a_{1} is non-adjacent to p1p_{1}. But now we get a theta with ends a1,p1a_{1},p_{1} and paths a1-a-a2-p1a_{1}\hbox{-}a\hbox{-}a_{2}\hbox{-}p_{1}, a1-S1-p1a_{1}\hbox{-}S_{1}\hbox{-}p_{1} and a path from a1a_{1} to p1p_{1} with interior in P1P_{1}, a contradiction. This proves (8).

(22) NP(a1){p1}N_{P^{\prime}}(a_{1})\subseteq\{p_{1}\} and NP(a3){pk}N_{P^{\prime}}(a_{3})\subseteq\{p_{k}\}.

Suppose a1a_{1} has a neighbor in Pp1P^{\prime}\setminus p_{1}. Then k>1k>1. By (8), a3a_{3} is anticomplete to PpkP^{\prime}\setminus p_{k}, and by (8), a2a_{2} is anticomplete to PP^{\prime}. Let H1H_{1} be the hole b1-P1-p1-P-pk-P3-b3-b1b_{1}\hbox{-}P_{1}\hbox{-}p_{1}\hbox{-}P^{\prime}\hbox{-}p_{k}\hbox{-}P_{3}\hbox{-}b_{3}\hbox{-}b_{1}, and let H2H_{2} be the hole p1-P1-b1-b2-P2-a-P3-pk-P-p1p_{1}\hbox{-}P_{1}\hbox{-}b_{1}\hbox{-}b_{2}\hbox{-}P_{2}\hbox{-}a\hbox{-}P_{3}\hbox{-}p_{k}\hbox{-}P^{\prime}\hbox{-}p_{1} (H2H_{2} is a hole since pkp_{k} has a neighbor in P3b3P_{3}\setminus b_{3}). Then NH2(a1)=NH1(a1){a}N_{H_{2}}(a_{1})=N_{H_{1}}(a_{1})\cup\{a\}. Let p0p_{0} be the neighbor of p1H1Pp_{1}\in H_{1}\setminus P^{\prime} and let QQ be the path p0-p1-P-pkp_{0}\hbox{-}p_{1}\hbox{-}P^{\prime}\hbox{-}p_{k}. Observe that QH2Q\subseteq H_{2}, and NH1(a1)QN_{H_{1}}(a_{1})\subseteq Q. Since neither of (H1,a1)(H_{1},a_{1}) and (H2,a1)(H_{2},a_{1}) is an even wheel, it follows that a1a_{1} has exactly two neighbors in QQ and they are adjacent. Let s{0,,k1}s\in\{0,\dots,k-1\} be such that NQ(a1)={ps,ps+1}N_{Q}(a_{1})=\{p_{s},p_{s+1}\}. Since a1a_{1} has a neighbor in P1p1P_{1}\setminus p_{1}, it follows that s>0s>0. Now we get a prism with triangles psa1ps+1p_{s}a_{1}p_{s+1} and b1b2b3b_{1}b_{2}b_{3} and paths a1-a-P2-b2a_{1}\hbox{-}a\hbox{-}P_{2}\hbox{-}b_{2}, ps-P-p1-p0-P1-b1p_{s}\hbox{-}P^{\prime}\hbox{-}p_{1}\hbox{-}p_{0}\hbox{-}P_{1}\hbox{-}b_{1} and ps+1-P-pk-P3-b3p_{s+1}\hbox{-}P^{\prime}\hbox{-}p_{k}\hbox{-}P_{3}\hbox{-}b_{3}, a contradiction. This proves (8).

(23) P2P_{2} is anticomplete to PP^{\prime}.

Suppose not. By (8), a2a_{2} has a neighbor in PP^{\prime}. Then a2b2a_{2}\neq b_{2}. Let H1H_{1} be the hole p1-P-pk-P3-b3-b1-P1-p1p_{1}\hbox{-}P^{\prime}\hbox{-}p_{k}\hbox{-}P_{3}\hbox{-}b_{3}\hbox{-}b_{1}\hbox{-}P_{1}\hbox{-}p_{1} and let H2H_{2} be the hole p1-P-pk-P3-a3-a-a1-P1-p1p_{1}\hbox{-}P^{\prime}\hbox{-}p_{k}\hbox{-}P_{3}\hbox{-}a_{3}\hbox{-}a\hbox{-}a_{1}\hbox{-}P_{1}\hbox{-}p_{1} (H2H_{2} is a hole since {b1,b3}\{b_{1},b_{3}\} is anticomplete to PP^{\prime}). Since a2b2a_{2}\neq b_{2}, we have that NH2(a2)=NH1(a2){a}N_{H_{2}}(a_{2})=N_{H_{1}}(a_{2})\cup\{a\}. Since since neither of (H1,a2)(H_{1},a_{2}), (H2,a2)(H_{2},a_{2}) is an even wheel, it follows that there exists s{1,,k1}s\in\{1,\dots,k-1\} such that NP(a2)={ps,ps+1}N_{P^{\prime}}(a_{2})=\{p_{s},p_{s+1}\}. Now we get a prism with triangles b1b2b3b_{1}b_{2}b_{3} and psa2ps+1p_{s}a_{2}p_{s+1} and paths ps-P-p1-P1-b1p_{s}\hbox{-}P^{\prime}\hbox{-}p_{1}\hbox{-}P_{1}\hbox{-}b_{1}, a2-P2-b2a_{2}\hbox{-}P_{2}\hbox{-}b_{2} and ps+1-P-pk-P3-b3p_{s+1}\hbox{-}P^{\prime}\hbox{-}p_{k}\hbox{-}P_{3}\hbox{-}b_{3}, a contradiction. This proves (8).

(24) p1p_{1} has at least two neighbors in P1P_{1}, and p3p_{3} has at least two neighbors in P3P_{3}.

By symmetry it is enough to prove the first assertion of (8). Suppose that p1p_{1} has a unique neighbor xx in P1P_{1}. It follows that xa1x\neq a_{1}. By (8), P2P_{2} is anticomplete to PP^{\prime}. Now we get a theta with ends x,ax,a and paths x-P1-ax\hbox{-}P_{1}\hbox{-}a, x-P1-b1-b2-P2-ax\hbox{-}P_{1}\hbox{-}b_{1}\hbox{-}b_{2}\hbox{-}P_{2}\hbox{-}a and x-p1-P-pk-P3-ax\hbox{-}p_{1}\hbox{-}P^{\prime}\hbox{-}p_{k}\hbox{-}P_{3}\hbox{-}a, a contradiction. This proves (8).

(25) p1p_{1} has two non-adjacent neighbors in P1P_{1}, and p3p_{3} has two non-adjacent two neighbors in P3P_{3}.

By symmetry it is enough to prove the first statement. By (8), we may assume that p1p_{1} has exactly two neighbors x,yx,y in P1P_{1}, and xx is adjacent to yy. We may assume that P1P_{1} traverses b1,x,y,a1b_{1},x,y,a_{1} in this order. By (8), P2P_{2} is anticomplete to PP^{\prime}. Let SS be the path from pkp_{k} to b3b_{3} with interior in P3P_{3}. It follows from the defintion of PP^{\prime} that aa is anticomplete to SS. Now we get a prism with triangles xyp1xyp_{1} and b1b2b3b_{1}b_{2}b_{3} and paths x-P1-b1x\hbox{-}P_{1}\hbox{-}b_{1}, y-P1-a-P2-b2y\hbox{-}P_{1}\hbox{-}a\hbox{-}P_{2}\hbox{-}b_{2} and p1-P-pk-S-b3p_{1}\hbox{-}P^{\prime}\hbox{-}p_{k}\hbox{-}S\hbox{-}b_{3}, a contradiction. This proves (8).

By (8), there exist paths P1P_{1}^{\prime} from p1p_{1} to aa and P1′′P_{1}^{\prime\prime} from p1p_{1} to b1b_{1}, both with interior in P1P_{1}^{*}, and such that P1p1P_{1}^{\prime}\setminus p_{1} is anticomplete to P1′′p1P_{1}^{\prime\prime}\setminus p_{1}. Let P3P_{3}^{\prime} be a path from pkp_{k} to aa with interior in P3P_{3}^{*}. By (8), P2P_{2} is anticomplete to PP^{\prime}. Now we get a theta with ends p1,ap_{1},a and paths p1-P1-ap_{1}\hbox{-}P_{1}^{\prime}\hbox{-}a, p1-P1′′-b1-b2-P2-ap_{1}\hbox{-}P_{1}^{\prime\prime}\hbox{-}b_{1}\hbox{-}b_{2}\hbox{-}P_{2}\hbox{-}a and p1-P-pk-P3-ap_{1}\hbox{-}P^{\prime}\hbox{-}p_{k}\hbox{-}P_{3}^{\prime}\hbox{-}a, a contradiction. ∎

We need a technical definition. Given an integer m>0m>0 and a graph class 𝒢\mathcal{G}, we say 𝒢\mathcal{G} is mm-amicable if 𝒢\mathcal{G} is amiable, and the following holds for every graph G𝒢G\in\mathcal{G}. Let σ\sigma be as in the definition of an amiable class for 𝒢\mathcal{G} and let V(G)=D1D2YV(G)=D_{1}\cup D_{2}\cup Y such that D1=d1--dk,D2D_{1}=d_{1}\hbox{-}\cdots\hbox{-}d_{k},D_{2} and YY satisfy the first five bullets from the definition of an amiable class with |Y|=σ(7)|Y|=\sigma(7). Let XYX\subseteq Y, H1D1H_{1}\subseteq D_{1} and H2D2H_{2}\subseteq D_{2} be as in the definition of an amiable class with |X|=9|X|=9, and let {x1,,x7}X\{x_{1},\dots,x_{7}\}\subseteq X such that:

  • x1,,x7x_{1},\dots,x_{7} is the order given on {x1,,x7}\{x_{1},\dots,x_{7}\} by (H1,X)(H_{1},X); and

  • For every x{x1,,x7}x\in\{x_{1},\dots,x_{7}\}, we have ND1(x)=NH1(x)N_{D_{1}}(x)=N_{H_{1}}(x).

Let ii be maximum such that x1x_{1} is adjacent to did_{i}, and let jj be minimum such that x7x_{7} is adjacent to djd_{j}. Then there exists a subset ZD2{di+2,,dj2}{x4}Z\subseteq D_{2}\cup\{d_{i+2},\dots,d_{j-2}\}\cup\{x_{4}\} with |Z|m|Z|\leq m such that N[Z]N[Z] separates did_{i} from djd_{j}. It follows that N[Z]N[Z] separates d1-D1-did_{1}\hbox{-}D_{1}\hbox{-}d_{i} from dj-D1-dkd_{j}\hbox{-}D_{1}\hbox{-}d_{k}.

We deduce:

Theorem 8.4.

The class 𝒢\mathcal{G} is 44-amicable.

Proof.

From Theorem 5.7, we know that 𝒢\mathcal{G} is amiable. With the notation as in the definition of an amicable class, we need to show that there exists a subset YD2{di+2,,dj2}{x4}Y\subseteq D_{2}\cup\{d_{i+2},\dots,d_{j-2}\}\cup\{x_{4}\} with |Z|4|Z|\leq 4 such that N[Z]N[Z] separates did_{i} from djd_{j}. Let RR be the path from x1x_{1} to x7x_{7} with interior in H2H_{2}.

Let HH be the hole x1-H1-x7-R-x1x_{1}\hbox{-}H_{1}\hbox{-}x_{7}\hbox{-}R\hbox{-}x_{1}. Let L=l1--lqL=l_{1}\hbox{-}\cdots\hbox{-}l_{q} be the path in H2{x4}H_{2}\cup\{x_{4}\} such that l1=x4l_{1}=x_{4} and lql_{q} has a neighbor in P(H2)P(H_{2}). Thus if (H2,X)(H_{2},X) is an alignment then L=x4L=x_{4}, and if (H2,X)(H_{2},X) is a connectifier then Lx4L\setminus x_{4} is the leg of H2H_{2} containing a neighbor of x4x_{4}. See Figure 8. (Note that Lx4L\setminus x_{4} may be empty if (H2,X)(H_{2},X) is a clique connectifier or a stellar connectifier.) We consider different possibilities for the behavior of (H1,X)(H_{1},X) and (H2,X)(H_{2},X).

Refer to caption
Figure 8. Proof of Theorem 8.4.

(26) If (H1,X)(H_{1},X) is a wide alignment, then the theorem holds.

Suppose first that (H2,X)(H_{2},X) is an alignment. Then (H,x4)(H,x_{4}) is a proper wheel, and we are done by Theorem 8.2 and setting Z={x4}Z=\{x_{4}\}. Thus we may assume that (H2,X)(H_{2},X) is a connectifier. It now follows from Theorem 5.7 that (H2,X)(H_{2},X) is a triangular connectifier or a clique connectifier. Let NH(lq)={h,h}N_{H}(l_{q})=\{h,h^{\prime}\}. Then HLH\cup L contains a pyramid Σ\Sigma with apex x4x_{4} and base lqhhl_{q}hh^{\prime}. Since {di,dj}\{d_{i},d_{j}\} is anticomplete to {x4,h,h,lq}\{x_{4},h,h^{\prime},l_{q}\}, it follows that did_{i} belongs to the interior of the path of Σ\Sigma with ends x4,hx_{4},h, and djd_{j} belongs to the interior of the path of Σ\Sigma with ends x4,hx_{4},h^{\prime} (switching the roles of h,hh,h^{\prime} if necessary), and that di,djd_{i},d_{j} are not adjacent to the apex x4x_{4} of Σ\Sigma. Since {di,dj}N(x4)=\{d_{i},d_{j}\}\cap N(x_{4})=\emptyset, Theorem 8.3 implies that N[x4,h,h,l1]N[x_{4},h,h^{\prime},l_{1}] separates did_{i} from djd_{j}. Since h,h,lqD2h,h^{\prime},l_{q}\in D_{2}, we are done. This proves (8).

(27) If (H2,X)(H_{2},X) is a wide alignment, then the theorem holds.

Since (H1,X)(H_{1},X) is an alignment, we deduce that (H,x4)(H,x_{4}) is a proper wheel. Now by Theorem 8.2 and setting Z={x4}Z=\{x_{4}\}, we are done. This proves (8).

(28) If (H1,X)(H_{1},X) is a triangular alignment, then the theorem holds.

Let l{i,,j}l\in\{i,\dots,j\} be such that N(x4)D1={dl,dl+1}N(x_{4})\cap D_{1}=\{d_{l},d_{l+1}\}. Since x2x_{2} has a neighbor in the path di-H1-dld_{i}\hbox{-}H_{1}\hbox{-}d_{l}, it follows that l>i+1l>i+1. Since x6x_{6} has a neighbor in the path dl+1-H1-djd_{l+1}\hbox{-}H_{1}\hbox{-}d_{j}, it follows that l<j2l<j-2. By Theorem 5.7 and in view of (8), we may assume that (H2,X)(H_{2},X) is a spiky alignment, a stellar connectifier or a spiky connectifier. In all cases, lql_{q} has a unique neighbor in P(H2)P(H_{2}); denote it by ss (where P(H2)P(H_{2}) is as defined in Section 5). Now HLH\cup L is a pyramid Σ\Sigma with apex ss and base x4dldl+1x_{4}d_{l}d_{l+1}. Since sD2s\in D_{2} and since x4x_{4} is non-adjacent to di,djd_{i},d_{j}, it follows that did_{i} belongs to the interior of the path of Σ\Sigma with ends dl,sd_{l},s, and djd_{j} belongs to the interior of the path of Σ\Sigma with ends dl+1,sd_{l+1},s, and di,djd_{i},d_{j} are not adjacent to the apex ss of Σ\Sigma. Since {di,dj}N(s)=\{d_{i},d_{j}\}\cap N(s)=\emptyset, Theorem 8.3 implies that N[x4,s,dl,dl+1]N[x_{4},s,d_{l},d_{l+1}] separates did_{i} from djd_{j}. Since sD2s\in D_{2} and i+1<l<j2i+1<l<j-2, we are done. This proves (8).

By Theorem 5.7 and in view of (8) and (8), we deduce that (H1,X)(H_{1},X) is a spiky alignment. Let dld_{l} be the unique neighbor of x4x_{4} in H1H_{1}. By Theorem 5.7 and in view of (8), it follows that (H2,X)(H_{2},X) is either a triangular alignment or a triangular connectifier or a clique connectifier. Let h,hh,h^{\prime} be the neighbors of lql_{q} in HH. Now HLH\cup L is a pyramid Σ\Sigma with apex dld_{l} and base lqhhl_{q}hh^{\prime}. Since x2x_{2} has a neighbor in the path di-H1-dld_{i}\hbox{-}H_{1}\hbox{-}d_{l}, it follows from the definition of a spiky alignment that i<l1i<l-1. Similarly, j>l+1j>l+1. Since h,hD2h,h^{\prime}\in D_{2}, we deduce that did_{i} belongs to the interior of the path of Σ\Sigma with ends dl,hd_{l},h, and djd_{j} belongs to the interior of the path of Σ\Sigma with ends dl,hd_{l},h^{\prime} (possibly switching the roles of h,hh,h^{\prime}), and di,djd_{i},d_{j} are non-adjacent to the apex dld_{l} of Σ\Sigma since i+1<l<j1i+1<l<j-1. Therefore, Theorem 8.3 implies that N[dl,h,h,lq]N[d_{l},h,h^{\prime},l_{q}] separates did_{i} from djd_{j}. Since h,hD2h,h^{\prime}\in D_{2} and i+1<l<j1i+1<l<j-1, we conclude that 𝒞\mathcal{C} is 4-amicable.

We now prove that N[Z]N[Z] separates d1-D1-did_{1}\hbox{-}D_{1}\hbox{-}d_{i} from dj-D1-dkd_{j}\hbox{-}D_{1}\hbox{-}d_{k}. Let DD be the component of GN[Z]G\setminus N[Z] such that diDd_{i}\in D, and let DD^{\prime} be the component of GN[Z]G\setminus N[Z] such that djDd_{j}\in D^{\prime}. Then DDD\neq D^{\prime}. Since YD2{di+2,,dj2}{x4}Y\subseteq D_{2}\cup\{d_{i+2},\dots,d_{j-2}\}\cup\{x_{4}\}, it follows that N[Z]N[Z] is disjoint from d1-D1-djd_{1}\hbox{-}D_{1}\hbox{-}d_{j} and dj-D1-dkd_{j}\hbox{-}D_{1}\hbox{-}d_{k}. Therefore, d1-D1-djDd_{1}\hbox{-}D_{1}\hbox{-}d_{j}\in D and dj-D1-dkDd_{j}\hbox{-}D_{1}\hbox{-}d_{k}\in D^{\prime} as required. ∎

We need the following lemma:

Lemma 8.5 (Chudnovsky, Pilipczuk, Pilipczuk, Thomassé [14], Lemma 5.3).

Let GG be a connected graph with a weight function ww. Then there is an induced path P=p1--pkP=p_{1}\hbox{-}\dots\hbox{-}p_{k} in GG such that N[P]N[P] is a ww-balanced separator.

For a graph GG and a set XGX\subseteq G we denote by γ(X)\gamma(X) the minimum size of a set YY such that XN[Y]X\subseteq N[Y]. We are now ready to prove Theorem 8.1. We prove the following strengthening, which along with Theorem 8.4 yields Theorem 8.1 at once. This will be used in a future paper.

Theorem 8.6.

For every integer m>0m>0 and every mm-amicable graph class 𝒢\mathcal{G}, there is an integer d>0d>0 with the following property. Let G𝒢G\in\mathcal{G} and let ww be a normal weight function on GG. Then there exists YV(G)Y\subseteq V(G) such that

  • |Y|d|Y|\leq d, and

  • N[Y]N[Y] is a ww-balanced separator in GG.

Proof.

It suffices to consider the unique component of GG with weight greater than 1/21/2; if there is no such component, we set Y=Y=\emptyset. Therefore, we may assume that GG is connected.

Since 𝒢\mathcal{G} is amicable, it is also amiable. Let K=σ(7)K=\sigma(7) where σ\sigma is as in the definition of an amiable class for 𝒞\mathcal{C}. Let d=6K+2m+1d=6K+2m+1. We claim that dd satisfies the conclusion of the theorem.

By Lemma 8.5, there is a path PP in GG such that w(D)12w(D)\leq\frac{1}{2} for every component DD of GN[P]G\setminus N[P]. Let P=p1--pkP=p_{1}\hbox{-}\cdots\hbox{-}p_{k} be such a path chosen with kk as small as possible. It follows from the minimality of kk that there is a component BB of GN[p1-P-pk1]G\setminus N[p_{1}\hbox{-}P\hbox{-}p_{k-1}] with w(B)>12w(B)>\frac{1}{2}. Let T=N(Ppk)N(B)T=N(P\setminus p_{k})\cap N(B).

(29) Let DD be a connected subset of GG such that D(TN[pk])=D\cap(T\cup N[p_{k}])=\emptyset. Then w(D)12w(D)\leq\frac{1}{2}.

Since DT=D\cap T=\emptyset and DD is connected, it follows that either DBD\subseteq B or DB=D\cap B=\emptyset. If DB=D\cap B=\emptyset, then w(D)1w(B)<12w(D)\leq 1-w(B)<\frac{1}{2}. Thus we may assume that DBD\subseteq B. Since DN[pk]=D\cap N[p_{k}]=\emptyset, we deduce that DD is contained in a component of GN[P]G\setminus N[P], and so w(D)12w(D)\leq\frac{1}{2}. This proves (8).

Let rr be minimum such that γ[TN(pr+1,,pk1)]2K\gamma[T\cap N(p_{r+1},\dots,p_{k-1})]\leq 2K.

(30) We may assume that r>0r>0.

Suppose r=0r=0. Then γ(T)<2K\gamma(T)<2K. Let YV(G)Y^{\prime}\subseteq V(G) be such that TN[Y]T\subseteq N[Y^{\prime}] and with |Y|2K|Y^{\prime}|\leq 2K. Let Y=Y{pk}Y=Y^{\prime}\cup\{p_{k}\}. Then every component of GN[Y]G\setminus N[Y] is disjoint from TN[pk]T\cup N[p_{k}] and (8) implies that YY satisfies the conclusion of the theorem. This proves (8).

Let Y0V(G)Y_{0}^{\prime}\subseteq V(G) be such that TN(pr,,pk1)N[Y0]T\cap N(p_{r},\dots,p_{k-1})\subseteq N[Y_{0}^{\prime}] and with |Y0|2K|Y_{0}^{\prime}|\leq 2K. Let Y0=Y0{pk}Y_{0}=Y_{0}^{\prime}\cup\{p_{k}\}. For every i{1,,r}i\in\{1,\dots,r\}, let end(i)\operatorname{end}(i) be the minimum j>ij>i such that γ(TN[{pi,,pend(i)}])=2K\gamma(T\cap N[\{p_{i},\dots,p_{\operatorname{end}(i)}\}])=2K. Let ZiZ_{i} be a set of size at most 2K2K such that TN[{pi,,pend(i)}]N[Zi]T\cap N[\{p_{i},\dots,p_{\operatorname{end}(i)}\}]\subseteq N[Z_{i}].

Our next goal is, for every i{1,,r}i\in\{1,\dots,r\}, to define two sets QiQ_{i} and NiN_{i}. To that end, we fix i{1,,r}i\in\{1,\dots,r\}. Let T1=TN({pi,,pend(i)})T_{1}^{\prime}=T\cap N(\{p_{i},\dots,p_{\operatorname{end}(i)}\}). Let S1{pi,,pend(i)}S_{1}^{\prime}\subseteq\{p_{i},\dots,p_{\operatorname{end}(i)}\} be such that T1N(S1)T_{1}^{\prime}\subseteq N(S_{1}^{\prime}), and assume that S1S_{1}^{\prime} is chosen with |S1||S_{1}^{\prime}| as small as possible. Order the vertices of S1S_{1}^{\prime} as q1,,qlq_{1},\dots,q_{l} in the order that they appear in PP when PP is traversed starting from p1p_{1}. It follows from the minimality of |S1||S_{1}^{\prime}| that there is n1T1n_{1}\in T_{1}^{\prime} such that NS1(n1)={q1}N_{S_{1}^{\prime}}(n_{1})=\{q_{1}\}. Let t(1)=1t(1)=1 and let S1={qt(1)}S_{1}=\{q_{t(1)}\} and T1={n1}T_{1}=\{n_{1}\}. So far we have defined S1,S1,T1,T1S_{1}^{\prime},S_{1},T_{1}^{\prime},T_{1} and t(1)t(1). Suppose that we have defined sequences of subsets S1,,SsS_{1}^{\prime},\dots,S_{s}^{\prime}, S1,,SsS_{1},\dots,S_{s}, T1,,TsT_{1}^{\prime},\dots,T_{s}^{\prime}, and T1,,TsT_{1},\dots,T_{s} and a sequence of integers t(1),,t(s)t(1),\dots,t(s) such that TsN(Ss)T_{s}^{\prime}\subseteq N(S_{s}^{\prime}) and with the following properties (for every l{1,,s}l\in\{1,\dots,s\}):

  • TlT_{l} is a stable set.

  • If t<t(l)t<t(l) and qtSlq_{t}\in S_{l}^{\prime}, then qtq_{t} is anticomplete to TlT_{l}^{\prime}.

  • t(j)<t(l)t(j)<t(l) for every j<lj<l.

  • If j<lj<l, then njn_{j} is anticomplete to SlS_{l}^{\prime}, and NSl(nl)=qt(l)N_{S_{l}^{\prime}}(n_{l})=q_{t(l)}.

  • For every j{1,,l}j\in\{1,\dots,l\}, NSl(nj)=qt(j)N_{S_{l}}(n_{j})=q_{t(j)}.

Refer to caption
Figure 9. Proof of Theorem 8.6.

See Figure 9. Now we either terminate the construction or construct the sets Ts+1T_{s+1}^{\prime}, Ts+1T_{s+1}, Ss+1S_{s+1}^{\prime}, Ss+1S_{s+1} and define t(s+1)t(s+1), and verify that the properties above continue to hold. Let Ts+1=TsN[SsTs]T_{s+1}^{\prime}=T_{s}^{\prime}\setminus N[S_{s}\cup T_{s}]. If Ts+1=T_{s+1}^{\prime}=\emptyset, let Qi=SsQ_{i}=S_{s} and Ni=TsN_{i}=T_{s}; the construction terminates here. Now suppose that Ts+1T_{s+1}^{\prime}\neq\emptyset. Let Ss+1S_{s+1}^{\prime} be a minimal subset of SsS_{s}^{\prime} such that Ts+1N(Ss+1)T_{s+1}^{\prime}\subseteq N(S_{s+1}^{\prime}). It follows from the second bullet and the minimality of t(s)t(s) that if qtSsq_{t}\in S_{s}^{\prime} and t<t(s)t<t(s), then qtq_{t} is anticomplete to TsT_{s}^{\prime}. Since Ss+1SsS_{s+1}^{\prime}\subseteq S_{s}^{\prime} and Ts+1TsT_{s+1}^{\prime}\subseteq T_{s}^{\prime}, we deduce that if qtSs+1q_{t}\in S_{s+1}^{\prime}, then t>t(s)t>t(s). Let tt be minimum such that qtSs+1q_{t}\in S_{s+1}^{\prime} and set t(s+1)=tt(s+1)=t. Then t(s+1)>t(s)t(s+1)>t(s), and the third bullet continues to hold. By the minimality of tt, the second bullet continues to hold. It follows from the minimality of Ss+1S_{s+1}^{\prime} that there exists ns+1Ts+1n_{s+1}\in T_{s+1}^{\prime} such that NSs+1(ns+1)={qt(s+1)}N_{S_{s+1}^{\prime}}(n_{s+1})=\{q_{t(s+1)}\}. Set Ss+1=Ss{qt(s+1)}S_{s+1}=S_{s}\cup\{q_{t(s+1)}\} and Ts+1=Ts{ns+1}T_{s+1}=T_{s}\cup\{n_{s+1}\}. Since TsT_{s} is anticomplete to Ts+1T_{s+1}^{\prime}, the first bullet continues to hold. Since NSs+1(ns+1)=qt(s+1)N_{S_{s+1}^{\prime}}(n_{s+1})=q_{t(s+1)}, and since Ss+1Ss{qt(s)}S_{s+1}^{\prime}\subseteq S_{s}^{\prime}\setminus\{q_{t(s)}\}, it follows that the fourth bullet continues to hold, and consequently the fifth bullet continues to hold. Thus our construction maintains the required properties. When we finish, write Ni={n1,,nm}N_{i}=\{n_{1},\dots,n_{m}\} and Qi={qt(1),,qt(m)}Q_{i}=\{q_{t(1)},\dots,q_{t(m)}\}, and we have:

(31) The following statements hold.

  • NiN_{i} is a stable set.

  • For every j{1,,m}j\in\{1,\dots,m\}, NQi(nj)=qt(j)N_{Q_{i}}(n_{j})=q_{t(j)}.

  • If j<jj<j^{\prime}, then t(j)<t(j)t(j)<t(j^{\prime}).

  • TN({pi,,pend(i)})N[QiNi]T\cap N(\{p_{i},\dots,p_{\operatorname{end}(i)}\})\subseteq N[Q_{i}\cup N_{i}].

To simplify notation, we renumber the elements of QiQ_{i}, replacing each index t(j)t(j) by jj. In the new notation we have Qi={q1,,qm}Q_{i}=\{q_{1},\dots,q_{m}\}, and for every j{1,,m}j\in\{1,\dots,m\}, NQ(nj)={qj}N_{Q}(n_{j})=\{q_{j}\}. Observe that q1,,qmq_{1},\dots,q_{m} appear in this order in PP when PP is traversed from p1p_{1} to pkp_{k}. Let Li=p1-P-q1L_{i}=p_{1}\hbox{-}P\hbox{-}q_{1}, Mi=q1-P-qmM_{i}=q_{1}\hbox{-}P\hbox{-}q_{m} and Ri=qm+1-P-pkR_{i}=q_{m+1}\hbox{-}P\hbox{-}p_{k}.

(32) For every ii, there exists riNir_{i}\in N_{i} and Yi(GN[P])Mi{ri}Y_{i}\subseteq(G\setminus N[P])\cup M_{i}\cup\{r_{i}\} with |Yi|m|Y_{i}|\leq m such that N[Yi]N[Y_{i}] separates LiL_{i} from RiR_{i}.

By (8), NiN_{i} is a stable set. Since TN({pi,,pend(i)})N[QiNi]T\cap N(\{p_{i},\dots,p_{\operatorname{end}(i)}\})\subseteq N[Q_{i}\cup N_{i}], and |Qi|=|Ni||Q_{i}|=|N_{i}|, it follows from (8) that |Ni|K|N_{i}|\geq K. Let NiNiN_{i}^{\prime}\subseteq N_{i} with |Ni|=K|N_{i}^{\prime}|=K. Recall that K=σ(7)K=\sigma(7). Now (8) follows from the definition of an mm-amicable class with D1=PD_{1}=P, D2=BD_{2}=B and Y=NiY=N_{i}^{\prime}. This proves (8).

We may assume that for every ii there is a component DiD_{i} of GN[ZiYi]G\setminus N[Z_{i}\cup Y_{i}] with w(Di)>12w(D_{i})>\frac{1}{2}, for otherwise the conclusion of Theorem 8.1 holds setting Y=ZiYiY=Z_{i}\cup Y_{i}. By (8), either DiD_{i} is anticomplete to LiL_{i}, or DiD_{i} is anticomplete to RiR_{i}. Let us say that QiQ_{i} is of type LL if DiD_{i} is anticomplete to LiL_{i}, and that QiQ_{i} is of type RR if DiD_{i} is anticomplete to RiR_{i}. By (8), every QiQ_{i} is of at least one type.

(33) Q1Q_{1} is of type LL.

By (8), D1(TN[pk])D_{1}\cap(T\cup N[p_{k}])\neq\emptyset. It follows that D1D_{1} contains a vertex tt such that tN[pk]t\in N[p_{k}] or tTN[Z1]t\in T\setminus N[Z_{1}]. Since TN({p1,,pend(1)})N[Z1]T\cap N(\{p_{1},\dots,p_{\operatorname{end}(1)}\})\subseteq N[Z_{1}], and since L1M1L_{1}\cup M_{1} is a subpath of p1-P-pend(1)p_{1}\hbox{-}P\hbox{-}p_{\operatorname{end}(1)}, it follows that D1D_{1} contains a vertex of N[R1]N[R_{1}]. Consequently, Q1Q_{1} is not of type RR, and so Q1Q_{1} is of type LL. This proves (8).

(34) We may assume that QrQ_{r} is of type RR.

Suppose that QrQ_{r} is of type LL. Then DrD_{r} is anticomplete to LrL_{r}. We claim that the set Y=ZrYrY0Y=Z_{r}\cup Y_{r}\cup Y_{0} satisfies the conclusion of the theorem. Let DD be a component of GN[Y]G\setminus N[Y], and suppose that w(D)>12w(D)>\frac{1}{2}. By (8) there exists tD(TN[pk])t\in D\cap(T\cup N[p_{k}]). Since ZrYrYZ_{r}\cup Y_{r}\subseteq Y, we have that DDrD\subseteq D_{r}. It follows that DD is anticomplete to LrL_{r}, and so tN(Lr)t\not\in N(L_{r}). Since TN({pr,,pend(r)})N[Zr]T\cap N(\{p_{r},\dots,p_{\operatorname{end}(r)}\})\subseteq N[Z_{r}], it follows that t{pr,,pend(r)}t\not\in\{p_{r},\dots,p_{\operatorname{end}(r)}\}. We deduce that tN({pend(r)+1,,pk})t\in N(\{p_{\operatorname{end}(r)+1},\dots,p_{k}\}). But then tN[Y0]t\in N[Y_{0}], contrary to the fact that Y0YY_{0}\subseteq Y. This proves the claim, and (8) follows.

From now on we make the assumption of (8). By (8) we can choose jj minimum such that QjQ_{j} is of type RR. By (8), j>1j>1. Let

Y=Zj1Yj1ZjYjY0.Y=Z_{j-1}\cup Y_{j-1}\cup Z_{j}\cup Y_{j}\cup Y_{0}.

Then |Y|6k+2m+1=d|Y|\leq 6k+2m+1=d. We claim that w(D)12w(D)\leq\frac{1}{2} for every component DD of GN[Y]G\setminus N[Y]. Suppose not, and let DD be a component of GN[Y]G\setminus N[Y] with w(D)>12w(D)>\frac{1}{2}. Since Zj1Yj1YZ_{j-1}\cup Y_{j-1}\subseteq Y, we have that DDj1D\subseteq D_{j-1}. Since Qj1Q_{j-1} is of type LL, it follows that DD is anticomplete to Lj1L_{j-1}. Since ZjYjYZ_{j}\cup Y_{j}\subseteq Y, we have that DDjD\subseteq D_{j}. Since QjQ_{j} is of type RR, it follows that DD is anticomplete to RjR_{j}. Since TN({pj1,,pend(j)})N[Zj1Zj]T\cap N(\{p_{j-1},\dots,p_{\operatorname{end}(j)}\})\subseteq N[Z_{j-1}\cup Z_{j}] and since pkY0p_{k}\in Y_{0}, we deduce that D(TN[pk])=D\cap(T\cup N[p_{k}])=\emptyset. But then w(D)<12w(D)<\frac{1}{2} by (8), a contradiction. ∎

9. From balanced to small

The second step in the proof of Theorem 1.5 is to transform balanced separators with small domination number into balanced separators of small size. We show:

Theorem 9.1.

Let tt be an integer, let dd be as in Theorem 8.1 and let ctc_{t} be as in Theorem 1.3. Let G𝒞tG\in\mathcal{C}_{t} and let ww be a weight function on GG. Then there exists YV(G)Y\subseteq V(G) such that

  • |Y|3ctdlogn+t|Y|\leq 3c_{t}d\log n+t, and

  • YY is a ww-balanced separator in GG.

The proof uses Theorems 1.3 and 8.1. In order to prove Theorem 9.1, we first prove a more general statement (below). We will then explain how to deduce Theorem 9.1 from it.

Theorem 9.2.

Let L>0L>0 be an integer, let GG be a graph and let ww be a weight function on GG. Assume that there is no LL-banana in GG. There exists a clique KK in GG with the following property. Let XV(G)KX\subseteq V(G)\setminus K be such that w(D)12w(D)\leq\frac{1}{2} for every component DD of G(KN[X])G\setminus(K\cup N[X]). Then there exists a balanced separator YY in GG such that |YK|3L|X||Y\setminus K|\leq 3L|X|.

Proof.

Let (T,χ)(T,\chi) be a 3L3L-atomic tree decomposition of GG. By Theorems 2.2 and 2.3, we have that (T,χ)(T,\chi) is tight and 3L3L-lean. By Theorem 2.7, there exists t0Tt_{0}\in T such that t0t_{0} is a center for TT. A vertex vV(G)v\in V(G) is t0t_{0}-friendly if vv is not separated from χ(t0)v\chi(t_{0})\setminus v by a set of size <3L<3L. We show that t0t_{0}-friendly vertices have the following important property.

(35) If u,vχ(t0)u,v\in\chi(t_{0}) are not t0t_{0}-friendly, then uu is adjacent to vv.

Suppose that uu and vv are not t0t_{0}-friendly and that uu is non-adjacent to vv. Since there is no LL-banana in GG, Theorem 2.1 implies that there exists a set XX with |X|<L|X|<L such XX separates uu from vv. But this contradicts Theorem 2.4. This proves (9).

Let KK be the set of all the vertices of GG that are not t0t_{0}-friendly. By (9), KK is a clique. We show that KK satisfies the conclusion of Theorem 9.2. Let XGKX\subseteq G\setminus K be such that w(D)12w(D)\leq\frac{1}{2} for every component DD of G(KN[X])G\setminus(K\cup N[X]).

For every vGKv\in G\setminus K, define the set Δ(v)\Delta(v) as follows. Let XvV(G)X_{v}\subseteq V(G) be such that XvX_{v} separates vv from χ(t0)v\chi(t_{0})\setminus v, chosen with |Xv||X_{v}| minimum. Let Δ(v)=Xv{v}\Delta(v)=X_{v}\cup\{v\}. Then |Δ(v)|3L|\Delta(v)|\leq 3L and Nχ(t0)(v)Δ(v)N_{\chi(t_{0})}(v)\subseteq\Delta(v). Let

Y=vXΔ(v).Y=\bigcup_{v\in X}\Delta(v).

It follows that |Y|3L|X||Y|\leq 3L|X|.

To complete the proof it is enough to show that YKY\cup K is a ww-balanced separator of GG. Suppose not, and let DD be a component of G(YK)G\setminus(Y\cup K) with w(D)>12w(D)>\frac{1}{2}.

(36) If D(Gt0tχ(t0))D\cap(G_{t_{0}\rightarrow t}\setminus\chi(t_{0}))\neq\emptyset for some tNT(t0)t\in N_{T}(t_{0}), then XX is anticomplete to DD.

Suppose that vXv\in X has a neighbor dd in DD. Let (A,Xv,B)(A,X_{v},B) be a separation such that χ(t0)XvB\chi(t_{0})\setminus X_{v}\subseteq B and vAv\in A. Since dXvd\not\in X_{v} and dd is adjacent to vv, it follows that dAd\in A. But then, since DD is a component of G(KY)G\setminus(K\cup Y) and XvYX_{v}\subseteq Y, it follows that DAD\subseteq A. Consequently, Dχ(t0)=D\cap\chi(t_{0})=\emptyset. It follows that DGt0tχ(t0)D\subseteq G_{t_{0}\rightarrow t}\setminus\chi(t_{0}), and so w(D)w(Gt0tχ(t0))12w(D)\leq w(G_{t_{0}\rightarrow t}\setminus\chi(t_{0}))\leq\frac{1}{2}, a contradiction. This proves (9).

(37) DN[X]=D\cap N[X]=\emptyset.

Suppose there is vXv\in X such that vv has a neighbor dDd\in D. By (9), dχ(t0)d\in\chi(t_{0}). But then dΔ(v)Yd\in\Delta(v)\subseteq Y, a contradiction. This proves (9).

By (9), DD is a component of G(KN[X])G\setminus(K\cup N[X]), and therefore w(D)12w(D)\leq\frac{1}{2}, a contradiction. ∎

We now prove Theorem 9.1.

Proof.

Let KK be as in Theorem 9.2. Let ctc_{t} be as in Theorem 1.3 and let L=ctlognL=c_{t}\log n. Let dd be as Theorem 8.1. Define w:V(G)K[0,1]w^{\prime}:V(G)\setminus K\rightarrow[0,1] as w(v)=w(v)1w(K)w^{\prime}(v)=\frac{w(v)}{1-w(K)}. Then ww^{\prime} is a normal weight function on GKG\setminus K. By Theorem 8.1, there exists XV(G)KX\subseteq V(G)\setminus K such that

  • |X|d|X|\leq d, and

  • N[X]N[X] is a ww^{\prime}-balanced separator in GKG\setminus K.

It follows that w(D)12w(D)\leq\frac{1}{2} for every component DD of G(N[X]K)G\setminus(N[X]\cup K). Now Theorem 9.2 applied with L=ctlognL=c_{t}\log_{n} implies that there exists a ww-balanced separator YY in GG such that |YK|3L|X|=3ct(logn)|X|3ctdlogn|Y\setminus K|\leq 3L|X|=3c_{t}(\log n)|X|\leq 3c_{t}d\log n. Since G𝒞tG\in\mathcal{C}_{t}, it follows that |Y|3ctdlogn+t|Y|\leq 3c_{t}d\log n+t. ∎

10. The proof of Theorem 1.5

We are now ready to complete the proof of Theorem 1.5. The following result was originally proved by Robertson and Seymour in [33], and tightened by Harvey and Wood in [24]. It was then restated and proved in the language of (w,c)(w,c)-balanced separators in an earlier paper of this series [4].

Theorem 10.1 (Robertson, Seymour [33], see also [4, 24]).

Let GG be a graph, let c[12,1)c\in[\frac{1}{2},1), and let dd be a positive integer. If GG has a (w,c)(w,c)-balanced separator of size at most dd for every normal weight function w:V(G)[0,1]w:V(G)\to[0,1], then tw(G)11ck\operatorname{tw}(G)\leq\frac{1}{1-c}k.

We prove the following, which immediately implies Theorem 1.5.

Theorem 10.2.

Let tt be an integer. Let let ctc_{t} be as in Theorem 1.3 and let dd be as in Theorem 1.4. Then every G𝒞tG\in\mathcal{C}_{t} satisfies tw(G)6ctdlogn+2t\operatorname{tw}(G)\leq 6c_{t}d\log n+2t.

Proof.

By Theorem 9.1, for every normal weight function ww of GG there exists ww-balanced separator of size at most 3ctdlogn+t3c_{t}d\log n+t. Now the result follows immediately from Lemma 10.1. ∎

11. Algorithmic consequences

We now summarize the algorithmic consequences of our structural results, specifically of Theorems 1.4 and 1.5.

The consequences for graphs in 𝒞t\mathcal{C}_{t} are the most immediate. In particular, using the factor 22 approximation algorithm of Korhonen [26] (or the simpler factor 44 approximation algorithm of Robertson and Seymour [34]) for treewidth, we can compute a tree decomposition of width at most O(ctlogn)O(c_{t}\log n) in time 2O(ctlogn)nO(1)=nO(ct)2^{O(c_{t}\log n)}n^{O(1)}=n^{O(c_{t})}. A number of well-studied graph problems, including Stable Set, Vertex Cover, Feedback Vertex Set Dominating Set and rr-Coloring (for fixed rr) admit algorithms with running time 2O(k)n2^{O(k)}n when a tree decomposition of GG of width kk is provided as input (See, for example, [16] chapters 7 and 11, as well as [10]). Since 2O(ctlogn)=nO(ct)2^{O(c_{t}\log n)}=n^{O(c_{t})} this immediately leads to polynomial time algorithms for these problems in 𝒞t\mathcal{C}_{t}.

Theorem 11.1.

Let t1t\geq 1 be fixed and P be a problem which admits an algorithm running in time 𝒪(2𝒪(k)|V(G)|)\mathcal{O}(2^{\mathcal{O}(k)}|V(G)|) on graphs GG with a given tree decomposition of width at most kk. Then P is solvable in time n𝒪(t)n^{\mathcal{O}(t)} in 𝒞t\mathcal{C}_{t}. In particular, Stable Set, Vertex Cover, Feedback Vertex Set, Dominating Set and rr-Coloring (with fixed rr) are all polynomial-time solvable in 𝒞t\mathcal{C}_{t}.

This list of problems is far from exhaustive, it is worth mentioning the work of Pilipczuk [32] who provides a relatively easy-to-check sufficient condition (expressibility in Existential Counting Modal Logic) for a problem to admit such an algorithm.

Theorem 11.1 implies a polynomial time algorithm for rr-Coloring (with fixed rr) in 𝒞\mathcal{C} (without any assumptions on max clique size) because every graph that contains a clique of size r+1r+1 can not be rr-colored. Thus, to solve rr-Coloring we can first check for the existence of an (r+1)(r+1)-clique in time nr+𝒪(1)n^{r+\mathcal{O}(1)}. If an (r+1)(r+1)-clique is present, report that no rr-coloring exists, otherwise invoke the algorithm of Theorem 11.1 with t=r+1t=r+1. This yields the following result.

Theorem 11.2.

For every positive integer rr, rr-Coloring is polynomial-time solvable in 𝒞\mathcal{C}.

Let us now discuss another important problem, and that is Coloring. It is well-known (and also follows immediately from Theorem 7.1), that for every tt there exists a number d(t)d(t) such that all graphs in 𝒞t\mathcal{C}_{t} have chromatic number at most d(t)d(t). Also, for each fixed rr, by Theorem 11.1, rr-Coloring is polynomial-time solvable in 𝒞t\mathcal{C}_{t}. Now by solving rr-Coloring for every r{1,,d(t)}r\in\{1,\dots,d(t)\}, we also obtain a polynomial-time algorithm for Coloring in 𝒞t\mathcal{C}_{t}.

Theorem 11.1 quite directly leads to a polynomial-time approximation scheme (PTAS) for several problems on graphs in 𝒞\mathcal{C}. We illustrate this using Vertex Cover as an example.

Theorem 11.3.

There exists an algorithm that takes as input a graph G𝒞G\in\mathcal{C} and 0<ϵ10<\epsilon\leq 1, runs in time nO(c2/ϵ)n^{O(c_{2/\epsilon})} and outputs a vertex cover of size at most (1+ϵ)vc(G)(1+\epsilon)\textsf{vc}(G), where vc(G)\textsf{vc}(G) is the size of the minimum vertex cover of GG.

Proof.

Check in time nO(1/ϵ)n^{O(1/\epsilon)} whether GG contains as input a clique CC of size at least 2/ϵ2/\epsilon. If GG does not have a clique of size at least 2/ϵ2/\epsilon then compute an optimal vertex cover of GG in time nO(c2/ϵ)n^{O(c_{2/\epsilon})} using the algorithm of Theorem 11.1. If GG contains such a clique CC then run the algorithm recursively on GCG-C to obtain a vertex cover SS of GCG-C of size at most (1+ϵ)vc(GC)(1+\epsilon)\textsf{vc}(G-C). Return SCS\cup C; clearly SCS\cup C is a vertex cover of GG. Furthermore, every vertex cover of GG (and in particular a minimum one) contains at least |C|1|C|-1 vertices of CC. It follows that vc(GC)vc(G)|C|+1\textsf{vc}(G-C)\leq\textsf{vc}(G)-|C|+1 and that therefore

|SC|\displaystyle|S\cup C| =|S|+|C|\displaystyle=|S|+|C|
(1+ϵ)vc(GC)+|C|\displaystyle\leq(1+\epsilon)\textsf{vc}(G-C)+|C|
(1+ϵ)(vc(G)|C|+1)+|C|\displaystyle\leq(1+\epsilon)(\textsf{vc}(G)-|C|+1)+|C|
(1+ϵ)(vc(G))\displaystyle\leq(1+\epsilon)(\textsf{vc}(G))

Here the last inequality follows because for C2/ϵC\geq 2/\epsilon it holds that (1+ϵ)(|C|1)|C|(1+\epsilon)(|C|-1)\geq|C|. ∎

The PTAS of Theorem 11.3 generalizes without modification (except for the constant 22 in the 2/ϵ2/\epsilon) to Feedback Vertex Set, and more generally, to deletion to any graph class which is closed under vertex deletion, excludes some clique, and admits an algorithm (for the deletion problem) with running time 2O(k)nO(1)2^{O(k)}n^{O(1)} on graphs of treewidth kk. Despite the tight connection between Vertex Cover and Stable Set, Theorem 11.3 does not directly lead to a PTAS for Stable Set on graphs in 𝒞\mathcal{C}. On the other hand, a QPTAS for Stable Set in 𝒞\mathcal{C} follows from Theorem 1.4 together with an argument of Chudnovsky, Pilipczuk, Pilipczuk and Thomassé [14], who gave a QPTAS for Stable Set in PkP_{k}-free graphs (they refer to the problem as Independent Set). For ease of reference, we repeat their argument here.

Theorem 11.4.

There exists an algorithm that takes as input a graph G𝒞G\in\mathcal{C} and 0<ϵ10<\epsilon\leq 1, runs in time nO(log2n/ϵ)n^{O(\log^{2}n/\epsilon)} and outputs a stable set SS in GG of size at least (1ϵ)α(G)(1-\epsilon)\alpha(G), where α(G)\alpha(G) is the size of the maximum size stable set in GG.

Proof.

The algorithm is recursive, taking as input the graph GG and also an integer NN. Initially, the algorithm is called with N=|V(G)|N=|V(G)|, in all subsequent recursive calls the value of NN remains the same. If GG is a single vertex the algorithm returns V(G)V(G). If GG is disconnected, the algorithm runs itself recursively on each of the connected components and returns the union of the stable sets returned for each of them.

Let dd be the constant of Theorem 1.4. If GG is a connected graph on at least two vertices the algorithm iterates over all subsets SV(G)S\subseteq V(G) of size at most 2dlognlogNϵ\frac{2d\log n\log N}{\epsilon} and all subsets YY of size at most dd such that each connected component of G(N(S)N[Y])G-(N(S)\cup N[Y]) has at most |V(G)|/2|V(G)|/2 vertices. For each such pair (S,Y)(S,Y) the algorithm calls itself recursively on G(N(S)N[Y])G-(N(S)\cup N[Y]) and returns the maximum size stable set found over all such recursive calls.

Clearly, the set returned by the algorithm is a stable set. For each pair (S,Y)(S,Y) which results in a recursive call of the algorithm, each connected component of the graph G(N(S)N[Y])G-(N(S)\cup N[Y]) that the algorithm is called on has at most |V(G)|/2|V(G)|/2 vertices. Therefore the running time of the algorithm is governed by the recurrence T(n)n2dlognlogNϵ+dnT(n/2)T(n)\leq n^{\frac{2d\log n\log N}{\epsilon}+d}\cdot n\cdot T(n/2) which solves to T(n)nO(dlog2nlogNϵ)T(n)\leq n^{O(\frac{d\log^{2}n\log N}{\epsilon})}. Since the algorithm is initially called with N=|V(G)|N=|V(G)| the running time of the algorithm is upper bounded by |V(G)|O(dlog3|V(G)|ϵ)|V(G)|^{O(\frac{d\log^{3}|V(G)|}{\epsilon})}.

It remains to argue that the size of the stable set output by the algorithm is at least (1ϵ)α(G)(1-\epsilon)\alpha(G). To that end, we show that the size of the independent set output by a recursive call on (G,N)(G,N) is at least α(G)(1ϵlog|V(G)|logN)\alpha(G)(1-\epsilon\frac{\log|V(G)|}{\log N}). Let II be a maximum size stable set of GG. By a standard coupon-collector argument (see e.g. Lemma 4.1 in [14]) there exists a choice of SIS\subseteq I of size at most dlogNϵ2logn\frac{d\log N}{\epsilon}2\log n such that no vertex of GN(S)G-N(S) has at least |I|ϵdlogN|I|\frac{\epsilon}{d\log N} neighbors in II. Let YY now be the set (of size at most dd) given by Lemma 1.4 applied to GN[S]G-N[S]. Since no vertex in YY has at least ϵdlogN\frac{\epsilon}{d\log N} neighbors in II it follows that |I(N(S)N[Y])||I|(1ϵlogN)|I-(N(S)\cup N[Y])|\geq|I|(1-\frac{\epsilon}{\log N}). Furthermore, each connected component of G(N(S)N[Y])G-(N(S)\cup N[Y]) has at most |V(G)|/2|V(G)|/2 vertices. By the inductive hypothesis the recursive call on (G(N(S)N[Y]),N)(G-(N(S)\cup N[Y]),N) outputs a stable set of size at least

|I|(1ϵlogN)(1ϵ(log|V(G)|1)logN)|I|(1ϵlog|V(G)|logN).|I|\left(1-\frac{\epsilon}{\log N}\right)\left(1-\frac{\epsilon(\log|V(G)|-1)}{\log N}\right)\geq|I|\left(1-\epsilon\frac{\log|V(G)|}{\log N}\right)\mbox{.}

Since α(G)=|I|\alpha(G)=|I| the recursive call on (G,N)(G,N) outputs a stable set of size at least α(G)(1ϵlog|V(G)|logN)\alpha(G)(1-\epsilon\frac{\log|V(G)|}{\log N}), as claimed. ∎

12. Acknowledgments

We are grateful to Tara Abrishami and Bogdan Alecu for their involvement in early stages of this work. We also thank Kristina Vušković for many years of conversations about the structure of even-hole-free graphs.

References

  • [1] P. Aboulker, I. Adler, E. J. Kim, N. L. D. Sintiari, and N. Trotignon. “On the tree-width of even-hole-free graphs.”
  • [2] T. Abrishami, B. Alecu, M. Chudnovsky, S. Hajebi and S. Spirkl. “Induced subgraphs and tree decompositions VIII. Excluding a forest in (theta, prism)-free graphs.” https://arxiv.org/abs/2301.02138 (2023)
  • [3] T. Abrishami, B. Alecu, M. Chudnovsky, S. Hajebi and S. Spirkl. “Induced subgraphs and tree decompositions X. Towards logarithmic treewidth in even-hole-free graphs.” https://arxiv.org/abs/2307.13684 (2023)
  • [4] T. Abrishami, M. Chudnovsky, C. Dibek, S. Hajebi, P. Rzążewski, S. Spirkl, and K. Vušković. “Induced subgraphs and tree decompositions II. Toward walls and their line graphs in graphs of bounded degree.” Journal of Combinatorial Theory. Series B 124, 1 (2024), 371–403.
  • [5] T. Abrishami, M. Chudnovsky, S. Hajebi and S. Spirkl. “Induced subgraphs and tree decompositions III. Three paths configurations and logarithmic treewidth.” Advances in Combinatorics (2022).
  • [6] T. Abrishami, M. Chudnovsky, and K. Vušković. “Induced subgraphs and tree decompositions I. Even-hole-free graphs of bounded degree.” J. Comb. Theory Ser. B, 157 (2022), 144–175.
  • [7] L. Addario-Berry, M. Chudnovsky, F. Havet, B. Reed, and P. Seymour. “Bisimplicial vertices in even-hole-free graphs.” Journal of Combinatorial Theory, Series B 98, 6 (2008), 1119–1164.
  • [8] B. Alecu, M. Chudnovsky, S. Hajebi and S. Spirkl. “Induced subgraphs and tree decompositions XI. Local structure for even-hole-free graphs of large treewidth.” https://arxiv.org/abs/2309.04390 (2023)
  • [9] P. Bellenbaum and R. Diestel. “Two short proofs concerning tree decompositions.” Combinatorics, Probability and Computing, 11 (2002) , 541–547.
  • [10] H. L. Bodlaender, M. Cygan, S. Kratsch and J. Nederlof. “Deterministic single exponential time algorithms for connectivity problems parameterized by treewidth.” Information and Computation, 243, (2015), 86–111.
  • [11] H. L. Bodlaender. “Dynamic programming on graphs with bounded treewidth.” Springer, Berlin, Heidelberg, (1988), pp. 105–118.
  • [12] K. Cameron, M. V. G. da Silva, S. Huang, and K. Vušković, “Structure and algorithms for (cap, even hole)-free graphs”, Discrete Mathematics 341 (2018), 463–473.
  • [13] J. Carmesin, R. Diestel, M. Hamann, and F. Hundertmark. “kk-Blocks: a connectivity invariant for graphs.” SIAM Journal on Discrete Mathematics, 28(4), (2014) 1876–1891.
  • [14] M. Chudnovsky, Marcin Pillipczuk, Mihal Pillipczuk and Stephan Thomassé. “Quasi-polynomial time approximation schemes for the Maximum Weight Independent Set Problem in HH-free graphs.” https://arxiv.org/abs/1907.04585 (2019).
  • [15] M. Conforti, B. Gerards and K. Pashkovich, “Stable sets and graphs with no even holes”, Mathematical Programming: Series A and B, 153 (2015), 13–39.
  • [16] M. Cygan, F. V. Fomin, L. Kowalik, D. Lokshtanov, D. Marx, M. Pilipczuk, M. Pilipczuk and S. Saurabh. “Parameterized Algorithms.” Springer (2015).
  • [17] M. V. da Silva and and K. Vušković. “Decomposition of even-hole-free graphs with star cutsets and 2-joins.” J. Comb. Theory, Ser. B 104 (2013), 144–183.
  • [18] J. Davies. “Vertex-minor-closed classes are χ\chi-bounded.” Combinatorica 42 (2022), 1049-1079.
  • [19] R. Diestel. Graph Theory. Springer-Verlag, Electronic Edition, (2005).
  • [20] P. Erdős and G. Szekeres. “A combinatorial problem in geometry.” Compositio Math 2 (1935), 463–470.
  • [21] P. Gartland, T. Korhonen, and D. Lokshtanov. “On Induced Versions of Menger’s Theorem on Sparse Graphs.” https://arxiv.org/pdf/2309.08169.pdf (2023)
  • [22] F. Gavril. “The intersection graphs of subtrees in trees are exactly the chordal graphs.” Journal of Combinatorial Theory. Series B 16 (1974), 47–56.
  • [23] M. C. Golumbic. “Algorithmic graph theory and perfect graphs.” In Annals of Discrete Mathematics, second edition, 16 Elsevier Science B.V., Amsterdam, (2004).
  • [24] D.J. Harvey and D.R. Wood. “Parameters Tied to Treewidth.” J. Graph Theory 84 (2017), 364–385.
  • [25] K. Hendry, S. Norin, R. Steiner, and J. Turcotte. “On an induced version of Menger’s theorem.” https://arxiv.org/abs/2309.07905 (2023)
  • [26] T. Korhonen. “Grid induced minor theorem for graphs of small degree.” J. Comb. Theory, Ser. B, 160, (2023), 206–214.
  • [27] D. Kühn and D. Osthus. “Induced subgraphs in Ks,sK_{s,s}-free graphs of large average degree.” Combinatorica 24 (2004), 287–304.
  • [28] N.K. Le, “Coloring even-hole-free graphs with no star cutset”, https://arxiv.org/abs/1805.01948
  • [29] V. Lozin, I. Razgon. “Tree-width dichotomy.” European Journal of Combinatorics 103, (2022), 103517.
  • [30] K. Menger, “Zur allgemeinen Kurventheorie.” Fund. Math. 10 (1927) 96 – 115.
  • [31] T. Ngueyn, A. Scott, P. Seymour, “A counterexampe to the coarse Menger conjecture.” preprint (2023)
  • [32] M. Pilipczuk, “Problems Parameterized by Treewidth Tractable in Single Exponential Time: A Logical Approach.” Proceedings of Mathematical Foundations of Computer Science 2011, 6907, (2011), 520–531.
  • [33] N. Robertson and P.D. Seymour. “Graph minors. II. Algorithmic aspects of tree-width.” Journal of Algorithms 7(3) (1986): 309–322.
  • [34] N. Robertson and P.D. Seymour. “Graph Minors. XIII. The Disjoint Paths Problem.” J. Comb. Theory, Ser. B, 63(1) (1995): 65–110.
  • [35] N. Robertson and P.D. Seymour. “Graph minors. XVI. Excluding a non-planar graph.” J. Combin. Theory, Ser. B, 89 (2003), 43–76.
  • [36] N.L.D. Sintiari and N. Trotignon. “(Theta, triangle)-free and (even-hole, K4K_{4})-free graphs. Part 1: Layered wheels.” J. Graph Theory 97 (4) (2021), 475-509.
  • [37] K. Vušković, “Even-hole-free graphs: A survey.” Applicable Analysis and Discrete Mathematics, (2010), 219–240.
  • [38] D. Weißauer, “On the block number of graphs.” SIAM J. Disc. Math. 33, 1 (2019): 346–357.