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

Tree independence number
II. Three-path-configurations.

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

A three-path-configuration is a graph consisting of three pairwise internally-disjoint paths the union of every two of which is an induced cycle of length at least four. A graph is 3PC-free if no induced subgraph of it is a three-path-configuration. We prove that 3PC-free graphs have poly-logarithmic tree independence number. More explicitly, we show that there exists a constant cc such that every nn-vertex 3PC-free graph graph has a tree decomposition in which every bag has stability number at most c(logn)2c(\log n)^{2}. This implies that the Maximum Weight Independent Set problem, as well as several other natural algorithmic problems that are known to be NP-hard in general, can be solved in quasi-polynomial time if the input graph is 3PC-free.

§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.
Department of Computer Science, University of California Santa Barbara, Santa Barbara, CA, USA. Supported by NSF grant CCF-2008838.
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 conducted while Spirkl was an Alfred P. Sloan fellow.

1. Introduction

All graphs in this paper are finite and simple and all logarithms are base 22. We include the following standard definitions for the reader’s convenience (see, for example, [4]). Let G=(V(G),E(G))G=(V(G),E(G)) be a graph. For XV(G)X\subseteq V(G), we denote by G[X]G[X] the subgraph of GG induced by XX, and by GXG\setminus X the subgraph of GG induced by V(G)XV(G)\setminus X. We use induced subgraphs and their vertex sets interchangeably. For graphs G,HG,H we say that GG contains HH if HH is isomorphic to G[X]G[X] for some XV(G)X\subseteq V(G); otherwise, we say that GG is HH-free. For a family \mathcal{H} of graphs, GG is said to be \mathcal{H}-free if GG is HH-free for every HH\in\mathcal{H}.

Let vV(G)v\in V(G). The open neighborhood of vv, denoted by NG(v)N_{G}(v), is the set of all vertices in V(G)V(G) adjacent to vv. The closed neighborhood of vv, denoted by NG[v]N_{G}[v], is N(v){v}N(v)\cup\{v\}. Let XV(G)X\subseteq V(G). The (open) neighborhood of XX, denoted NG(X)N_{G}(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 NG[X]N_{G}[X], is NG(X)XN_{G}(X)\cup X. When there is no danger of confusion, we often omit the subscript GG. Let YV(G)Y\subseteq V(G) with XY=X\cap Y=\emptyset. 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.

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.

The width of a tree decomposition (T,χ)(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 was first introduced by Robertson and Seymour in their work on graph minors. A bound on the treewidth of a graph provides important information about its structure [12]; it is also useful from the algorithmic perspective [3]. As a result treewidth has been extensively studied in both structural and algorithmic graph theory.

A stable (or independent) set in a graph GG is a set of pairwise non-adjacent vertices of GG. The stability (or independence) number α(G)\alpha(G) of GG is the size of a maximum stable set in GG. Given a graph GG with weights on its vertices, the Maximum Weight Independent Set (MWIS) problem is the problem of finding a stable set in GG of maximum total weight. This problem is known to be NP-hard [8], but it can be solved efficiently (in polynomial time) in graphs of bounded treewidth. Closer examination of the algorithm motivated Dallard, Milanič and Štorgel [6] to define a related graph width parameter, specifically targeting the complexity of the MWIS problem. The independence number of a tree decomposition (T,χ)(T,\chi) of GG is maxtV(T)α(G[χ(t)])\max_{t\in V(T)}\alpha(G[\chi(t)]). The tree independence number of GG, denoted treeα(G)\operatorname{tree-\alpha}(G), is the minimum independence number of a tree decomposition of GG. Graphs with large treewidth and small treeα\operatorname{tree-\alpha} are graphs whose large treewidth can be explained by the presence of a large clique. It is shown in [6] that if a graph is given together with a tree decomposition with bounded independence number, then the MWIS problem can be solved in polynomial time. Moreover, [5] presents an algorithm that constructs such tree decompositions efficiently in graphs of bounded treeα\operatorname{tree-\alpha}, yielding an efficient algorithm for the MWIS problem for graphs of bounded treeα\operatorname{tree-\alpha}.

We need the following standard definitions (see, for example, [2, 4]). A hole in a graph is an induced cycle of length at least four. A path in a graph is an induced subgraph that is a path. The length of a path or a hole is the number of edges in it. Given a path PP with ends a,ba,b, the interior of PP, denoted by PP^{*}, is the set P{a,b}P\setminus\{a,b\}.

A theta is a graph consisting of two distinct vertices a,ba,b and three paths P1,P2,P3P_{1},P_{2},P_{3} from aa to bb, such that PiPjP_{i}\cup P_{j} is a hole for every i,j{1,2,3}i,j\in\{1,2,3\}. It follows that aa is non-adjacent to bb and the sets P1,P2,P3P_{1}^{*},P_{2}^{*},P_{3}^{*} are pairwise disjoint and anticomplete to each other. If a graph GG contains an induced subgraph HH that is a theta, and a,ba,b are the two vertices of degree three in HH, then we say that GG contains a theta with ends aa and bb.

A pyramid is a graph consisting of a vertex aa and a triangle {b1,b2,b3}\{b_{1},b_{2},b_{3}\}, and three paths PiP_{i} from aa to bib_{i} for 1i31\leq i\leq 3, such that PiPjP_{i}\cup P_{j} is a hole for every i,j{1,2,3}i,j\in\{1,2,3\}. It follows that P1a,P2a,P3aP_{1}\setminus a,P_{2}\setminus a,P_{3}\setminus a are pairwise disjoint, and the only edges between them are of the form bibjb_{i}b_{j}. It also follows that at most one of P1,P2,P3P_{1},P_{2},P_{3} has length exactly one. We say that aa is the apex of the pyramid and that b1b2b3b_{1}b_{2}b_{3} is the base of the pyramid.

A generalized prism is a graph consisting of two triangles {a1,a2,a3}\{a_{1},a_{2},a_{3}\} and {b1,b2,b3}\{b_{1},b_{2},b_{3}\}, and three paths PiP_{i} from aia_{i} to bib_{i} for 1i31\leq i\leq 3, and such that PiPjP_{i}\cup P_{j} is a hole for every i,j{1,2,3}i,j\in\{1,2,3\}. It follows that P1,P2,P3P_{1}^{*},P_{2}^{*},P_{3}^{*} are pairwise disjoint and anticomplete to each other, |{a1,a2,a3}{b1,b2,b3}|1|\{a_{1},a_{2},a_{3}\}\cap\{b_{1},b_{2},b_{3}\}|\leq 1, and if a1=b1a_{1}=b_{1}, then P2P_{2}^{*}\neq\emptyset and P3P_{3}^{*}\neq\emptyset. Moreover, the only edges between PiP_{i} and PjP_{j} are aiaja_{i}a_{j} and bibjb_{i}b_{j}. A prism is a generalized prism whose triangles are disjoint. A pinched prism is a generalized prism whose triangles meet.

A three-path-configuration (3PC) is a graph that is either a theta, or pyramid, or a generalized prism (see Figure 1). It is easy to check that this definition is equivalent to the one in the abstract. Let 𝒞\mathcal{C} be the class of (theta, pyramid, generalized prism)-free graphs; 𝒞\mathcal{C} is also known as the class of 3PC-free graphs.

Refer to caption
Figure 1. The three-path-configurations. From left to right: A theta, a pyramid, a prism and a pinched prism (dashed lines depict paths of non-zero length).

The following is the main result of [2]:

Theorem 1.1 ([2]).

For every integer t>0t>0 there exists a constant c(t)c(t) such that for every nn-vertex graph G𝒞G\in\mathcal{C} that contains no clique of size tt, tw(G)c(t)logn\operatorname{tw}(G)\leq c(t)\log n.

This is a strengthening of a conjecture of [14] that theta-free graphs with no 33-vertex clique have logarithmic treewidth It was also shown in [14] that there exist triangle-free graphs in 𝒞\mathcal{C} with arbitrarily large treewidth (in fact, treewidth logarithmic in the number of vertices), and so the bound of Theorem 1.1 is asymptotically best possible. A consequence of Theorem 1.1 is that the MWIS problem (as well as many others) can be solved in polynomial time on 3PC-free graphs with bounded clique number.

It is now natural to ask about 3PC-free graphs with no bound on the clique number. Since the complete bipartite graph K2,3K_{2,3} is a theta, and therefore is forbidden in graphs in 𝒞\mathcal{C}, one would expect these graphs to behave well with respect to treeα\operatorname{tree-\alpha}. Our main result here confirms this. We prove:

Theorem 1.2.

There exists a constant cc such that for every integer n>1n>1 every nn-vertex graph G𝒞G\in\mathcal{C} has tree independence number at most c(logn)2c(\log n)^{2}.

Note that since the class of theta-free graphs is “χ\chi-bounded” (see [13] for details), Theorem 1.2 yields a weakening of Theorem 1.1, that for every integer t>0t>0, there exists a constant c(t)c(t) such that for every nn-vertex graph G𝒞G\in\mathcal{C} that contains no clique of size tt, tw(G)c(t)(logn)2\operatorname{tw}(G)\leq c(t)(\log n)^{2}. On the other hand, since the only construction of 3PC-graphs with large treewidth known to date is the construction of [14] where all graphs have clique number at most four, we do not know if the bound of Theorem 1.2 is asymptotically tight, or whether it can be made linear in logn\log n (in which case, it would imply Theorem 1.1).

Another result in this paper that may be of independent interest is the following:

Theorem 1.3.

Let G𝒞G\in\mathcal{C} with |V(G)|=n|V(G)|=n, and let a,bV(G)a,b\in V(G) be non-adjacent. Then there is a set XV(G){a,b}X\subseteq V(G)\setminus\{a,b\} with α(X)32logn\alpha(X)\leq 32\log n and such that every component of GXG\setminus X contains at most one of a,ba,b.

1.1. Proof outline and organization

The proof of Theorem 1.2 follows an outline similar to [4], but requires several new techniques and ideas. We sketch it in this subsection, postponing the precise definitions for later. We begin by exploring the effect that the presence of “useful wheels” has on 3PC-free graphs, and show that every useful wheel can be broken by a cutset that is contained in the union of the neighborhoods of three vertices. This is done in Section 2.

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. Let c[0,1]c\in[0,1] and let ww be a normal weight function on GG. 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.

This is done in Section 3; the proof is similar to the proof of an analogous statement in [4].

In Section 4 we prove Theorem 1.3. The key insight here is that a stronger result can (and should) be proved, showing that every two “cooperative subgraphs”, disjoint and anticomplete to each other, can be separated by removing a set with logarithmic stability number. The proof of this strengthening follows by relatively standard structural analysis.

In Section 6 we develop a technique that uses results of Section 3 and Section 4 and produces a balanced separator of small stability number in a graph. This technique does not depend on the particular graph-class in question, but only on the validity of statements similar to Theorems  3.1 and 4.1. We also rely on a lemma from Section 5, which is proved here for theta-free graphs, but can be generalized in several ways. Section 6 is completely different from [4], and requires several new ideas.

In Section 7 we deduce Theorem 1.2 from the building blocks developed so far. We finish with Section 8 discussing the algorithmic implications of Theorem 1.2.

2. Structural results

In this section we prove a theorem asserting the existence of certain cutsets in graphs in 𝒞\mathcal{C}.

Let GG be a graph. Let X,Y,ZV(G)X,Y,Z\subseteq V(G). We say that XX separates YY from ZZ if no component of GXG\setminus X meets both YY and ZZ. Let WW be a hole in GG and vGWv\in G\setminus W. A sector of (W,v)(W,v) is a path PP of WW of length at least one, such that both ends of PP are adjacent to vv, and vv is anticomplete to PP^{*}. A sector PP is long if PP^{*}\neq\emptyset. A useful wheel in GG is a pair (W,v)(W,v) where WW is a hole of length at least seven and (W,v)(W,v) has at least two long sectors. We prove:

Theorem 2.1.

Let G𝒞G\in\mathcal{C} and let (W,v)(W,v) be a useful wheel in GG. Let SS be a long sector of WW with ends s1,s2s_{1},s_{2}. Then ((N(s1)N(s2))W)N(v)((N(s_{1})\cup N(s_{2}))\setminus W)\cup N(v) separates SS^{*} from WSW\setminus S.

Proof.

Let X=((N(s1)N(s2))W)N(v)X=((N(s_{1})\cup N(s_{2}))\setminus W)\cup N(v). Suppose for a contradiction that there is a component of GXG\setminus X intersecting both SS^{*} and WSW\setminus S. It follows that there is a path P=p1--pkP=p_{1}\hbox{-}\dots\hbox{-}p_{k} in GXG\setminus X, possibly with k=1k=1, such that p1p_{1} has a neighbor in SS^{*} and pkp_{k} has a neighbor in WSW\setminus S. In particular, PP is disjoint from and anticomplete to {s1,s2,v}\{s_{1},s_{2},v\}.

Choose PP with |P|=k|P|=k as small as possible. It follows that

  • we have PG(WX)P\subseteq G\setminus(W\cup X);

  • PP^{*} is anticomplete to W{v}W\cup\{v\};

  • if k>1k>1, then p1p_{1} is anticomplete to WSW\setminus S^{*} and pkp_{k} is anticomplete to SS.

Let t1t_{1} and t2t_{2} be the (unique) neighbors of s1s_{1} and s2s_{2} in WSW\setminus S^{*}, respectively. Since (W,v)(W,v) has at least two long sectors, it follows that s1,s2,t1,t2s_{1},s_{2},t_{1},t_{2} are all distinct, and that WN[S]W\setminus N[S]\neq\emptyset. In particular, since W{v}W\cup\{v\} is not a pyramid, pinched prism, or theta, it follows that vv has a neighbor in wWN[S]w\in W\setminus N[S].

Traversing SS from s1s_{1} to s2s_{2}, let u1u_{1} and u2u_{2} be the first and the last neighbor of p1p_{1} in SS, respectively. It follows that u1,u2Su_{1},u_{2}\in S^{*}. We deduce:

(1) N(pk)(WS){t1,t2}N(p_{k})\cap(W\setminus S)\subseteq\{t_{1},t_{2}\}.

Suppose not. Then there is a path QQ in GG from pkp_{k} to vv such that QWN[S]Q^{*}\subseteq W\setminus N[S]. Assume that u1=u2u_{1}=u_{2}. Then there is a theta in GG with ends u1,vu_{1},v and paths u1-S-s1-v,u1-S-s2-v,u1-p1-P-pk-Q-vu_{1}\hbox{-}S\hbox{-}s_{1}\hbox{-}v,u_{1}\hbox{-}S\hbox{-}s_{2}\hbox{-}v,u_{1}\hbox{-}p_{1}\hbox{-}P\hbox{-}p_{k}\hbox{-}Q\hbox{-}v. Next, assume that u1u_{1} and u2u_{2} are distinct and non-adjacent. Then there is a theta in GG with ends p1,vp_{1},v and paths p1-u1-S-s1-v,p1-u2-S-s2-v,p1-P-pk-Q-vp_{1}\hbox{-}u_{1}\hbox{-}S\hbox{-}s_{1}\hbox{-}v,p_{1}\hbox{-}u_{2}\hbox{-}S\hbox{-}s_{2}\hbox{-}v,p_{1}\hbox{-}P\hbox{-}p_{k}\hbox{-}Q\hbox{-}v. Since GG is theta-free, it follows that u1,u2u_{1},u_{2} are distinct and adjacent. But now there is a pyramid in GG with apex vv, base p1u1u2p_{1}u_{1}u_{2} and paths p1-P-pk-Q-v,u1-S-s1-v,u2-S-s2-vp_{1}\hbox{-}P\hbox{-}p_{k}\hbox{-}Q\hbox{-}v,u_{1}\hbox{-}S\hbox{-}s_{1}\hbox{-}v,u_{2}\hbox{-}S\hbox{-}s_{2}\hbox{-}v, a contradiction. This proves (2).

(2) We have u1=u2u_{1}=u_{2}.

Suppose not. By (2) and without loss of generality, we may assume that pkp_{k} is adjacent to t1t_{1}. Assume first that t1t_{1} and vv are not adjacent. If u1u_{1} and u2u_{2} are not adjacent either, then there is a theta in GG with ends p1,s1p_{1},s_{1} and paths p1-u1-S-s1,p1-u2-S-s2-v-s1,p1-P-pk-t1-s1p_{1}\hbox{-}u_{1}\hbox{-}S\hbox{-}s_{1},p_{1}\hbox{-}u_{2}\hbox{-}S\hbox{-}s_{2}\hbox{-}v\hbox{-}s_{1},p_{1}\hbox{-}P\hbox{-}p_{k}\hbox{-}t_{1}\hbox{-}s_{1}, and if u1,u2u_{1},u_{2} are adjacent, then there is a pyramid in GG with apex s1s_{1}, base p1u1u2p_{1}u_{1}u_{2} and paths p1-P-pk-t1-s1,u1-S-s1,u2-S-s2-v-s1p_{1}\hbox{-}P\hbox{-}p_{k}\hbox{-}t_{1}\hbox{-}s_{1},u_{1}\hbox{-}S\hbox{-}s_{1},u_{2}\hbox{-}S\hbox{-}s_{2}\hbox{-}v\hbox{-}s_{1}. Since GG is (theta, pyramid)-free, it follows that t1t_{1} and vv are adjacent. Assume that u1u_{1} and u2u_{2} are not adjacent. Then there is a pyramid in GG with apex p1p_{1}, base s1t1vs_{1}t_{1}v and paths paths s1-S-u1-p1,t1-pk-P-p1,v-s2-S-u2-p1s_{1}\hbox{-}S\hbox{-}u_{1}\hbox{-}p_{1},t_{1}\hbox{-}p_{k}\hbox{-}P\hbox{-}p_{1},v\hbox{-}s_{2}\hbox{-}S\hbox{-}u_{2}\hbox{-}p_{1}. Again, since GG is pyramid-free, it follows that u1,u2u_{1},u_{2} are adjacent. But now there is a prism in GG with triangles u1p1u2,s1t1vu_{1}p_{1}u_{2},s_{1}t_{1}v and paths u1-S-s1,p1-P-pk-t1,u2-S-s2-vu_{1}\hbox{-}S\hbox{-}s_{1},p_{1}\hbox{-}P\hbox{-}p_{k}\hbox{-}t_{1},u_{2}\hbox{-}S\hbox{-}s_{2}\hbox{-}v, a contradiction. This proves (2).

Henceforth, let u=u1=u2u=u_{1}=u_{2}. It follows that:

(3) We have k=1k=1.

Suppose that k>1k>1. Since W{pk}W\cup\{p_{k}\} is not a theta with ends t1,t2t_{1},t_{2}, we may assume by (2) and without loss of generality, that N(pk)(WS)={t1}N(p_{k})\cap(W\setminus S)=\{t_{1}\}. But then WPW\cup P is a theta in GG with ends u,t1u,t_{1}, a contradiction. This proves (2).

Henceforth, let p=p1=pkp=p_{1}=p_{k}. Since W{p}W\cup\{p\} is not a theta with one end uu and the other end in {t1,t2}\{t_{1},t_{2}\}, (2) implies that N(p)W={t1,t2,u}N(p)\cap W=\{t_{1},t_{2},u\}. Recall that vv has a neighbor wWN[S]w\in W\setminus N[S].

(4) We have t1wE(G)t_{1}w\in E(G) and t2vE(G)t_{2}v\notin E(G). Similarly, we have t2wE(G)t_{2}w\in E(G) and t1vE(G)t_{1}v\notin E(G).

Suppose not. Then we may assume, without loss of generality, that either t1t_{1} and ww are not adjacent, or t2t_{2} and vv are adjacent. In either case, it follows that there is a path QQ in GG from pp to vv such that t2PW(SN(t1))t_{2}\in P^{*}\subseteq W\setminus(S\cup N(t_{1})). Now there is a theta in GG with ends p,s1p,s_{1} and paths p-u-S-s1,p-t1-s1,p-Q-v-s1p\hbox{-}u\hbox{-}S\hbox{-}s_{1},p\hbox{-}t_{1}\hbox{-}s_{1},p\hbox{-}Q\hbox{-}v\hbox{-}s_{1}, a contradiction. This proves (2).

We will now finish the proof. From (2), it follows that WS=t1-w-t2W\setminus S=t_{1}\hbox{-}w\hbox{-}t_{2} and N(v)W={s1,s2,w}N(v)\cap W=\{s_{1},s_{2},w\}. Recall also that N(p)W={t1,t2,u}N(p)\cap W=\{t_{1},t_{2},u\}. Since |W|>6|W|>6, it follows either s1-S-us_{1}\hbox{-}S\hbox{-}u or s2-S-us_{2}\hbox{-}S\hbox{-}u, say the former, has non-empty interior. But then there is a theta in GG with ends s1,us_{1},u and paths s1-S-u,s1-t1-p-us_{1}\hbox{-}S\hbox{-}u,s_{1}\hbox{-}t_{1}\hbox{-}p\hbox{-}u and s1-v-s2-S-us_{1}\hbox{-}v\hbox{-}s_{2}\hbox{-}S\hbox{-}u, a contradiction. ∎

3. Dominated balanced separators

The goal of this section is to prove the following:

Theorem 3.1.

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 follow the outline of the proof of Theorem 8.1 in [4]. First we repeat several definitions from [4]. Let GG be a graph, let P=p1--pnP=p_{1}\hbox{-}\dots\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 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.

The first step in the proof of Theorem 3.1 is the following:

Theorem 3.2.

For every integer x6x\geq 6, there exists an integer σ=σ(x)1\sigma=\sigma(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|=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 and a subpath H1H_{1} of D1D_{1} as well as H2D2H_{2}\subseteq D_{2} such that

  1. (1)

    (H1,X)(H_{1},X) is a consistent alignment, and for every vertex xx in XX except for at most two of them, ND1(x)=NH1(x)N_{D_{1}}(x)=N_{H_{1}}(x).

  2. (2)

    One of the following holds.

    • We have |H2|=1|H_{2}|=1 (so H2XH_{2}\cup X is a star), and (H1,X)(H_{1},X) is wide.

    • (H2,X)(H_{2},X) is a consistent alignment, the orders given on X by (H1,X)(H_{1},X) and by (H2,X)(H_{2},X) are the same, and at least one of (H1,X)(H_{1},X) and (H2,X)(H_{2},X) is wide.

The proof of Theorem 3.2 requires two preliminary results. The first one is Theorem 3.3 below from [4]. Following [4], by a caterpillar we mean a tree CC with maximum degree three such that there exists a path PP of CC where all branch vertices of CC belong to PP. (Our definition of a caterpillar is non-standard for two reasons: a caterpillar is often allowed to be of arbitrary maximum degree, and a spine often contains all vertices of degree more than one.) A claw is the graph K1,3K_{1,3}. For a graph HH, a vertex vv of HH is said to be simplicial if NH(v)N_{H}(v) is a clique.

Theorem 3.3 (Chudnovsky, Gartland, Hajebi, Lokshtanov, Spirkl; Theorem 5.2 in [4]).

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.

  • 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 every vertex of HN(Y)H\cap N(Y^{\prime}) is simplicial in HH.

The second one is:

Lemma 3.4.

Let c,x1c,x\geq 1 be integers. Let GG be a theta-free graph and assume that V(G)=D1D2YV(G)=D_{1}\cup D_{2}\cup Y where

  • YY is a stable set with |Y|=(3x+2)(c+2)|Y|=(3x+2)(c+2);

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

  • N(D1)=N(D2)=YN(D_{1})=N(D_{2})=Y;

  • D1D_{1} is a path; and

  • for every dD1d\in D_{1}, we have |N(d)Y|c|N(d)\cap Y|\leq c.

Then there exist XYX\subseteq Y with |X|=x+2|X|=x+2 and a subpath H1H_{1} of D1D_{1} such that:

  • (H1,X)(H_{1},X) is a consistent alignment.

  • For all but at most two vertices of XX, all their neighbors in D1D_{1} are contained in H1H_{1}.

Proof.

For every vertex yYy\in Y, let PyP_{y} be the path in D1D_{1} such that yy is complete to the ends of PyP_{y} and anticomplete to D1PyD_{1}\setminus P_{y}. Let II be the graph with V(I)=YV(I)=Y, such that two distinct vertices y,yYy,y^{\prime}\in Y are adjacent in II if and only if PyPyP_{y}\cap P_{y^{\prime}}\neq\emptyset. Then II is an interval graph, and so by [9], II is perfect. Since |V(I)|=(3x+2)(c+2)|V(I)|=(3x+2)(c+2), we deduce that II contains either a clique of cardinality c+2c+2 of a stable set of cardinality 3x+23x+2.

Assume that II contains a clique of cardinality c+2c+2. Then there exists CYC\subseteq Y with |C|=c+2|C|=c+2 and dD1d\in D_{1} such that dPyd\in P_{y} for every yCy\in C. It follows that for every yCy\in C, either yy is adjacent to dd, or D1dD_{1}\setminus d has two components and yy has a neighbor in each of them. Since |N(d)Y|c|N(d)\cap Y|\leq c, we deduce that there are two vertices y,yCYy,y^{\prime}\in C\subseteq Y as well as two paths P1P_{1} and P2P_{2} from yy to yy^{\prime} with disjoint and anticomplete interiors contained in D1D_{1}. On the other hand, since D2D_{2} is connected and N(D2)=YN(D_{2})=Y, it follows that there is a path P3P_{3} in GG from yy to yy^{\prime} whose interior is contained in D2D_{2}. But now there is a theta in GG with ends y,yy,y^{\prime} and paths P1,P2,P3P_{1},P_{2},P_{3}, a contradiction.

We deduce that II contains a stable set SS of cardinality 3x+23x+2. From the definition of II, it follows that there is a subpath H1H_{1} of D1D_{1} such that (H1,S)(H_{1},S) is an alignment. Hence, there exists XSYX\subseteq S\subseteq Y with |X|=x|X|=x such that (H1,X)(H_{1},X) is a consistent alignment. This completes the proof of Lemma 3.4. ∎

We are now ready to prove Theorem 3.2:

Proof of Theorem 3.2.

Let σ(x)=18μ(3((x+2)2+1)(x+1))\sigma(x)=18\mu(3((x+2)^{2}+1)(x+1)), where μ()\mu(\cdot) comes from Theorem 3.3. We begin with the following:

(5) Every vertex in D1D_{1} has at most four neighbors in YY.

Suppose for a contradiction that for some i{1,,k}i\in\{1,\dots,k\}, there is a subset ZYZ\subseteq Y of cardinality five such that did_{i} is complete to ZZ. It follows that for every yZy\in Z, we have i(y)ii(y)\neq i, and so there is 33-subset TT of ZZ such that either i(y)<ii(y)<i for all yTy\in T or i<i(y)i<i(y) for all yTy\in T. Consequently, there are two distinct vertices y,yTZYy,y^{\prime}\in T\subseteq Z\subseteq Y for which did_{i} is disjoint from and anticomplete to di(y)-D1-di(y)d_{i(y)}\hbox{-}D_{1}\hbox{-}d_{i(y^{\prime})}. On the other hand, since D2D_{2} is connected and N(D2)=YN(D_{2})=Y, it follows that there is a path QQ in GG from yy to yy^{\prime} whose interior is contained in D2D_{2}. But now there is a theta in GG with ends y,yy,y^{\prime} and paths y-di(y)-D1-di(y)-y,y-di-y,Qy\hbox{-}d_{i(y)}\hbox{-}D_{1}\hbox{-}d_{i(y^{\prime})}\hbox{-}y^{\prime},y\hbox{-}d_{i}\hbox{-}y^{\prime},Q, a contradiction. This proves (3).

From (3), Lemma 3.4 and the choice of σ(x)\sigma(x), it follows that:

(6) There exists Y1YY_{1}\subseteq Y with |Y1|=μ(3((x+2)2+1)(x+1))|Y_{1}|=\mu(3((x+2)^{2}+1)(x+1)) and a subpath H1H_{1} of D1D_{1} such that (H1,Y1)(H_{1},Y_{1}) is consistent alignment.

Henceforth, let Y1Y_{1} be as in (3). Since G1=G[Y1D2]G_{1}=G[Y_{1}\cup D_{2}] and G1Y1=D2G_{1}\setminus Y_{1}=D_{2} are both connected, we can apply Theorem 3.3 to G1G_{1} and Y1Y_{1}. It follows that there is a set YY1Y^{\prime}\subseteq Y_{1} with |Y|=3((x+2)2+1)(x+1)|Y^{\prime}|=3((x+2)^{2}+1)(x+1) and an induced subgraph HH of G1Y1=D2G_{1}\setminus Y_{1}=D_{2} for which one of the following holds.

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

  • 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 every vertex of HN(Y)H\cap N(Y^{\prime}) is simplicial in HH.

Assume that the second bullet above holds. By (3), (H1,Y)(H_{1},Y^{\prime}) is a consistent alignment. But then it is straightforward to observe that GG contains either a theta, a prism or a pyramid, a contradiction. It follows that HH is indeed a path and every vertex of YY^{\prime} has a neighbor in HH.

Now, assume that some vertex in zHz\in H has at least xx neighbors in YY^{\prime}. Choose XN(z)YYX\subseteq N(z)\cap Y^{\prime}\subseteq Y with |X|=x|X|=x. Let H2={z}H_{2}=\{z\}. By (3), (H1,X)(H_{1},X) is a consistent alignment. Note that if (H1,X)(H_{1},X) is spiky, then H1X{z}H_{1}\cup X\cup\{z\} contains a theta, and if (H1,X)(H_{1},X) is triangular, then H1X{z}H_{1}\cup X\cup\{z\} contains a pyramid. Therefore, (H1,X)(H_{1},X) is wide. But now XX and H2H_{2} satisfy Theorem 3.2.

Therefore, we may assume that every vertex in HH has fewer than xx neighbors in YY^{\prime}. Let H2=HH_{2}=H. Since |Y|=3((x+2)2+1)(x+1)|Y^{\prime}|=3((x+2)^{2}+1)(x+1), it follows from Lemma 3.4 that there exists XYX^{\prime}\subseteq Y^{\prime} with |X|=(x+2)2|X^{\prime}|=(x+2)^{2} such that (H2,X)(H_{2},X^{\prime}) is a consistent alignment. Also, by (3), (D1,X)(D_{1},X^{\prime}) is a consistent alignment. This, along with the Erdős-Szekeres Theorem [7], implies that there exists XXYYX\subseteq X^{\prime}\subseteq Y^{\prime}\subseteq Y with |X|=x+2|X|=x+2 such that both (H1,X)(H_{1},X) and (H2,X)(H_{2},X) are consistent alignments, and the orders given on XX by (H1,X)(H_{1},X) and (H2,X)(H_{2},X) are the same. Moreover, since GG is (theta, pyramid, pinched prism)-free, it follows that at least one of (H1,X)(H_{1},X) and (H2,X)(H_{2},X) is wide. Hence, XX and H2H_{2} satisfy Theorem 3.2. This completes the proof. ∎

Now, as in [4], we will show that the class 𝒞\mathcal{C} is “amiable” and “amicable”, and then use Theorem 8.5 of [4] to complete to the proof. The details are below. In [4], a graph class 𝒢\mathcal{G} is said to be amiable if, under the same assumptions as that of Theorem 3.2 for a graph G𝒢G\in\mathcal{G}, there exists XYX\subseteq Y with |X|=x+2|X|=x+2, H1D1H_{1}\subseteq D_{1} and H2D2H_{2}\subseteq D_{2} satisfying one of several possible outcomes. In particular, the outcome of Theorem 3.2 is one of the possible outcomes in the definition of an amiable class. Therefore, from Theorem 3.2, we deduce that:

Corollary 3.5.

The class 𝒞\mathcal{C} is amiable.

Following [4], for an integer m>0m>0, a graph class 𝒢\mathcal{G} is said to be 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 (and so as in Theorem 3.2) 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 assumptions of Theorem 3.2 with |Y|=σ(7)|Y|=\sigma(7). Let XYX\subseteq Y, H1D1H_{1}\subseteq D_{1} and H2D2H_{2}\subseteq D_{2} be as Theorem 3.2 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}. Consequently, 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 prove:

Theorem 3.6.

The class 𝒞\mathcal{C} is 33-amicable.

Proof.

By Corollary 3.5, 𝒞\mathcal{C} is an amiable class. With same notation as in the definition of a 33-amicable class, our goal is to show that 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|3|Z|\leq 3 such that N[Z]N[Z] separates did_{i} from djd_{j}. Consequently, 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 l{1,,k}l\in\{1,\ldots,k\} be minimum such that x4x_{4} is adjacent to dld_{l}, and let m{1,,k}m\in\{1,\ldots,k\} be maximum such that x4x_{4} is adjacent to dmd_{m}. It follows that i+2<lm<j2i+2<l\leq m<j-2.

Let RR be the (unique) path in H2H_{2} with ends r1,r2r_{1},r_{2} (possibly r1=r2r_{1}=r_{2}) such that x1x_{1} is adjacent to r1r_{1} and anticomplete to R{r1}R\setminus\{r_{1}\} and x7x_{7} is adjacent to r2r_{2} and anticomplete to R{r2}R\setminus\{r_{2}\}. Then x4x_{4} has a neighbor in RR. Traversing RR from r1r_{1} to r2r_{2}, let z1z_{1} and z2z_{2} be the first and the last neighbor of x4x_{4} in RR.

Let W=di-D1-dj-x7-r2-R-r1-x1-diW=d_{i}\hbox{-}D_{1}\hbox{-}d_{j}\hbox{-}x_{7}\hbox{-}r_{2}\hbox{-}R\hbox{-}r_{1}\hbox{-}x_{1}\hbox{-}d_{i}. Then WW is a hole in GG and |W|7|W|\geq 7. It follows that (W,x4)(W,x_{4}) is a useful wheel in GG. In particular, S=dl-D1-di-x1-r1-R-z1S=d_{l}\hbox{-}D_{1}\hbox{-}d_{i}\hbox{-}x_{1}\hbox{-}r_{1}\hbox{-}R\hbox{-}z_{1} and S=dm-D1-dj-x7-r2-R-z2S^{\prime}=d_{m}\hbox{-}D_{1}\hbox{-}d_{j}\hbox{-}x_{7}\hbox{-}r_{2}\hbox{-}R\hbox{-}z_{2} are two long sectors of (W,x4)(W,x_{4}).

Note that d1d_{1} and z1z_{1} are the ends of the sector SS from (W,x4)(W,x_{4}). Let Z={x4,dl,z1}Z=\{x_{4},d_{l},z_{1}\}. Then we have ZD2{di+2,,dj2}{x4}Z\subseteq D_{2}\cup\{d_{i+2},\dots,d_{j-2}\}\cup\{x_{4}\}, diSN[Z]d_{i}\in S^{*}\setminus N[Z] and djW(SN[Z])d_{j}\in W\setminus(S\cup N[Z]). Hence, by Theorem 2.1, N[Z]N[Z] separates did_{i} from djd_{j}, as desired. ∎

The following is a restatement of Theorem 8.5 of [4]:

Theorem 3.7 (Chudnovsky, Gartland, Hajebi, Lokshtanov, Spirkl [4]).

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 𝒢\mathcal{G} be a graph class which is mm-amicable. 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.

Now Theorem 3.1 is immediate from Theorems 3.6 and 3.7.

4. Separating a pair of vertices

The goal of this section is to prove the following:

Theorem 4.1.

Let G𝒞G\in\mathcal{C} with |V(G)|=n|V(G)|=n, and let a,bV(G)a,b\in V(G) be non-adjacent. Then there is a set XV(G){a,b}X\subseteq V(G)\setminus\{a,b\} with α(X)16×2logn\alpha(X)\leq 16\times 2\log n and such that every component of GXG\setminus X contains at most one of a,ba,b.

We need the following result from [1].

Lemma 4.2.

Let x1,x2,x3x_{1},x_{2},x_{3} be three distinct vertices of a graph GG. Assume that HH is a connected induced subgraph of G{x1,x2,x3}G\setminus\{x_{1},x_{2},x_{3}\} such that V(H)V(H) contains at least one neighbor of each of x1x_{1}, x2x_{2}, x3x_{3}, and that V(H)V(H) is minimal subject to inclusion. Then, one of the following holds:

  1. (i)

    For some distinct i,j,k{1,2,3}i,j,k\in\{1,2,3\}, there exists PP that is either a path from xix_{i} to xjx_{j} or a hole containing the edge xixjx_{i}x_{j} such that

    • V(H)=V(P){xi,xj}V(H)=V(P)\setminus\{x_{i},x_{j}\}; and

    • either xkx_{k} has two non-adjacent neighbors in HH or xkx_{k} has exactly two neighbors in HH and its neighbors in HH are adjacent.

  2. (ii)

    There exists a vertex aV(H)a\in V(H) and three paths P1,P2,P3P_{1},P_{2},P_{3}, where PiP_{i} is from aa to xix_{i}, such that

    • V(H)=(V(P1)V(P2)V(P3)){x1,x2,x3}V(H)=(V(P_{1})\cup V(P_{2})\cup V(P_{3}))\setminus\{x_{1},x_{2},x_{3}\};

    • the sets V(P1){a}V(P_{1})\setminus\{a\}, V(P2){a}V(P_{2})\setminus\{a\} and V(P3){a}V(P_{3})\setminus\{a\} are pairwise disjoint; and

    • for distinct i,j{1,2,3}i,j\in\{1,2,3\}, there are no edges between V(Pi){a}V(P_{i})\setminus\{a\} and V(Pj){a}V(P_{j})\setminus\{a\}, except possibly xixjx_{i}x_{j}.

  3. (iii)

    There exists a triangle a1a2a3a_{1}a_{2}a_{3} in HH and three paths P1,P2,P3P_{1},P_{2},P_{3}, where PiP_{i} is from aia_{i} to xix_{i}, such that

    • V(H)=(V(P1)V(P2)V(P3)){x1,x2,x3}V(H)=(V(P_{1})\cup V(P_{2})\cup V(P_{3}))\setminus\{x_{1},x_{2},x_{3}\};

    • the sets V(P1)V(P_{1}), V(P2)V(P_{2}) and V(P3)V(P_{3}) are pairwise disjoint; and

    • for distinct i,j{1,2,3}i,j\in\{1,2,3\}, there are no edges between V(Pi)V(P_{i}) and V(Pj)V(P_{j}), except aiaja_{i}a_{j} and possibly xixjx_{i}x_{j}.

For a graph GG and two subsets X,YV(G)X,Y\subseteq V(G) we define the distance in GG between XX and YY as the length (number of edges) of the shortest path of GG with one end in XX and the other in YY. We denote the distance between XX and YY by distG(X,Y)\operatorname{dist}_{G}(X,Y). Thus XX and YY are disjoint if and only if distG(X,Y)>0\operatorname{dist}_{G}(X,Y)>0, and X,YX,Y are anticomplete to each other if and only if distG(X,Y)>1\operatorname{dist}_{G}(X,Y)>1. In order to prove Theorem 4.1 we will prove a stronger statement. Let HGH\subseteq G. We denote by δG(H)\delta_{G}(H) the set of vertices of HH that have a neighbor in GHG\setminus H (so δ(H)=N(GH)\delta(H)=N(G\setminus H)). We say that HH is cooperative if one of the following holds:

  • HH is a clique, or

  • NH(Hδ(H))=δ(H)N_{H}(H\setminus\delta(H))=\delta(H) and Hδ(H)H\setminus\delta(H) is connected.

The following lemma summarizes the property of cooperative subgraphs that is of interest to us.

Lemma 4.3.

Let HGH\subseteq G be cooperative and let {n1,n2,n3}\{n_{1},n_{2},n_{3}\} be a stable set in N(H)N(H). Assume that there exist distinct h1,h2,h3δ(H)h_{1},h_{2},h_{3}\in\delta(H) such that nihjn_{i}h_{j} is an edge if and only if i=ji=j. Then there is KH{n1,n2,n3}K\subseteq H\cup\{n_{1},n_{2},n_{3}\} such that

  1. (1)

    KK is a subdivided claw or the line graph of a subdivided claw

  2. (2)

    {n1,n2,n3}\{n_{1},n_{2},n_{3}\} is the set of simplicial vertices of KK.

Proof.

If {h1,h2,h3}\{h_{1},h_{2},h_{3}\} is a triangle, then {n1,n2,n3,h1,h2,h3}\{n_{1},n_{2},n_{3},h_{1},h_{2},h_{3}\} is the line graph of a subdivided claw and the lemma holds. This we may assume that at least one pair hihjh_{i}h_{j} is non-adjacent, and in particular HH is not a clique. It follows that NH(Hδ(H))=δ(H)N_{H}(H\setminus\delta(H))=\delta(H) and Hδ(H)H\setminus\delta(H) is connected.

Next suppose that h1h2h_{1}h_{2} and h2h3h_{2}h_{3} are edges. Then h1h3h_{1}h_{3} is not an edge, and {n1,n2,n3,h1,h2,h3}\{n_{1},n_{2},n_{3},h_{1},h_{2},h_{3}\} is a subdivided claw and the lemma holds. Thus we may assume that at most one of the pairs hihjh_{i}h_{j} is an edge.

Suppose h1h3h_{1}h_{3} is an edge. Let P=p1--pkP=p_{1}\hbox{-}\dots\hbox{-}p_{k} be path such that p1=h2p_{1}=h_{2}, pkp_{k} has a neighbor in {h1,h3}\{h_{1},h_{3}\}, and Pp1Hδ(H)P\setminus p_{1}\subseteq H\setminus\delta(H); choose PP with kk as small as possible. If pkp_{k} is adjacent to exactly one of h1,h3h_{1},h_{3}, then P{h1,h2,h3,n1,n2,n3}P\cup\{h_{1},h_{2},h_{3},n_{1},n_{2},n_{3}\} is a subdivided claw; and if pkp_{k} is adjacent to both h1h_{1} and h3h_{3}, then P{h1,h2,h3,n1,n2,n3}P\cup\{h_{1},h_{2},h_{3},n_{1},n_{2},n_{3}\} is the line graph of a subdivided claw; in both cases the lemma holds. Thus we may assume that {h1,h2,h3}\{h_{1},h_{2},h_{3}\} is a stable set.

Let RR be a minimal connected subgraph of Hδ(H)H\setminus\delta(H) such that each of h1,h2,h3h_{1},h_{2},h_{3} has a neighbor in RR. We apply Lemma 4.2. Suppose that the first outcome holds; we may assume that RR is path from h1h_{1} to h2h_{2}. If h3h_{3} has two non-adjacent neighbors in RR, then R{h1,h2,h3,n1,n2,n3}R\cup\{h_{1},h_{2},h_{3},n_{1},n_{2},n_{3}\} contains a subdivided claw; and if h3h_{3} has exactly two neighbors in RR and they are adjacent, then R{h1,h2,h3,n1,n2,n3}R\cup\{h_{1},h_{2},h_{3},n_{1},n_{2},n_{3}\} is the line graph of a subdivided claw; in both cases the theorem holds. If the second outcome of Lemma 4.2 holds, then R{h1,h2,h3,n1,n2,n3}R\cup\{h_{1},h_{2},h_{3},n_{1},n_{2},n_{3}\} is a subdivided claw; and if the third outcome of Lemma 4.2 holds, then R{h1,h2,h3,n1,n2,n3}R\cup\{h_{1},h_{2},h_{3},n_{1},n_{2},n_{3}\} is the line graph of a subdivided claw. Thus in all cases the lemma holds. ∎

We also need the following:

Lemma 4.4.

Let G𝒞G\in\mathcal{C} and let H1,H2H_{1},H_{2} be cooperative subgraphs of GG, disjoint and anticomplete to each other. Then α(N(H1)N(H2))<17\alpha(N(H_{1})\cap N(H_{2}))<17.

Proof.

Suppose there is a stable set NN(H1)N(H2)N\subseteq N(H_{1})\cap N(H_{2}) with |N|=17|N|=17. Suppose first that some vertex h1H1h_{1}\in H_{1} has at least five neighbors in NN; let n1,,n5NN(h1)n_{1},\dots,n_{5}\in N\cap N(h_{1}). If some h2H2h_{2}\in H_{2} has three three neighbors in {n1,,n5}\{n_{1},\dots,n_{5}\}, say {n1,n2,n3}\{n_{1},n_{2},n_{3}\}, then {h1,h2,n1,n2,n3}\{h_{1},h_{2},n_{1},n_{2},n_{3}\} is a theta with ends h1,h2h_{1},h_{2}, a contradiction. So no such h2h_{2} exists. It follows that there exist h1,h2,h3H2h_{1}^{\prime},h_{2}^{\prime},h_{3}^{\prime}\in H_{2} and n1,n2,n3{n1,,n5}n_{1}^{\prime},n_{2}^{\prime},n_{3}^{\prime}\in\{n_{1},\dots,n_{5}\} such that hinjh_{i}^{\prime}n_{j}^{\prime} is an edge if and only if i=ji=j. By Lemma 4.3 there exists KH2{n1,n2,n3}K\subseteq H_{2}\cup\{n_{1}^{\prime},n_{2}^{\prime},n_{3}^{\prime}\} such that KK is a subdivided claw or the line graph of a subdivided claw, and {n1,n2,n3}\{n_{1}^{\prime},n_{2}^{\prime},n_{3}^{\prime}\} is the set of simplicial vertices of KK. But now Kh1K\cup h_{1} is a theta or a pyramid in GG, a contradiction.

It follows that for every h1H1h_{1}\in H_{1}, |N(h1)N|4|N(h_{1})\cap N|\leq 4. Since |N|=17|N|=17 and NN(H1)N\subseteq N(H_{1}), there exist h1,,h5Hh_{1},\dots,h_{5}\in H, and n1,,n5Nn_{1},\dots,n_{5}\in N such that hinjh_{i}n_{j} is an edge if and only if i=ji=j.

By renumbering n1,,n5n_{1},\dots,n_{5} if necessary we may assume that one of the following holds:

  • there exists kH2k\in H_{2} such that n1,n2,n3N(k)n_{1},n_{2},n_{3}\in N(k); in this case set K2={k,n1,n2,n3}K_{2}=\{k,n_{1},n_{2},n_{3}\}, or

  • there exist k1,k2,k3H2k_{1},k_{2},k_{3}\in H_{2} such that kinjk_{i}n_{j} is an edge if and only if i=ji=j. In this case let K2H2{n1,n2,n3}K_{2}\subseteq H_{2}\cup\{n_{1},n_{2},n_{3}\} be such that K2K_{2} is a subdivided claw or the line graph of a subdivided claw and {n1,n2,n3}\{n_{1},n_{2},n_{3}\} is the set of simplicial vertices of K2K_{2} (such K2K_{2} exists by Lemma 4.3).

By Lemma 4.3 there exists K1H1{n1,n2,n3}K_{1}\subseteq H_{1}\cup\{n_{1},n_{2},n_{3}\} such that K1K_{1} is a subdivided claw or the line graph of a subdivided claw, and {n1,n2,n3}\{n_{1},n_{2},n_{3}\} is the set of simplicial vertices of K1K_{1}. But now K1K2K_{1}\cup K_{2} is a theta, a pyramid or a prism in GG, a contradiction. ∎

For XV(G)X\subseteq V(G), a component DD of GXG\setminus X is full for XX if N(D)=XN(D)=X. XV(G)X\subseteq V(G) is a minimal separator in GG if there exist two distinct full components for XX. We will now prove the following strengthening of Theorem 4.1:

Theorem 4.5.

Let G𝒞G\in\mathcal{C} with |V(G)|=n|V(G)|=n, and let H1,H2H_{1},H_{2} be cooperative subgraphs of GG, disjoint and anticomplete to each other. Then there is a set XV(G)(H1H2)X\subseteq V(G)\setminus(H_{1}\cup H_{2}) with α(X)16×2log(n+1|H1||H2|)\alpha(X)\leq 16\times 2\log(n+1-|H_{1}|-|H_{2}|) and such that XX separates H1H_{1} from H2H_{2}.

Proof.

Write G1=GG_{1}=G, H21=H2H_{2}^{1}=H_{2} and N1=NG1(H1)NG1(H2)N_{1}=N_{G_{1}}(H_{1})\cap N_{G_{1}}(H_{2}). Define G2=G1N1G_{2}=G_{1}\setminus N_{1}, H22=NG2[H21]H_{2}^{2}=N_{G_{2}}[H_{2}^{1}] and N2=NG2(H1)NG2(H22)N_{2}=N_{G_{2}}(H_{1})\cap N_{G_{2}}(H_{2}^{2}). Let G3=G2N2G_{3}=G_{2}\setminus N_{2}.

(7) distG3(H1,H2)>3\operatorname{dist}_{G_{3}}(H_{1},H_{2})>3.

Let P=p1--pkP=p_{1}\hbox{-}\dots\hbox{-}p_{k} be a shortest path in G3G_{3} from H1H_{1} to H2H_{2}. Then PP is a path in GG, p1H1p_{1}\in H_{1}, pkH2p_{k}\in H_{2}, Pp1P\setminus p_{1} is anticomplete to H1H_{1}, and PpkP\setminus p_{k} is anticomplete to H2H_{2}. Since H1H_{1} is anticomplete to H2H_{2}, it follows that k3k\geq 3. If k=3k=3, then p1N1p_{1}\in N_{1}, a contradiction. Suppose k=4k=4. Then p2,p3N1p_{2},p_{3}\not\in N_{1}. It follows that p3NG2(H2)=H22p_{3}\in N_{G_{2}}(H_{2})=H_{2}^{2}, and therefore p2N2p_{2}\in N_{2}; again a contradiction. This proves that k>4k>4, and (4) follows.

It follows immediately from the definition of a cooperative subgraph that:

(8) For i{1,2}i\in\{1,2\}, H1H_{1} and H2iH_{2}^{i} are both cooperative in GiG_{i}.

Now Lemma 4.3 implies:

(9) α(Ni)<17\alpha(N_{i})<17 for every i{1,2}i\in\{1,2\}.

If H1H_{1} and H2H_{2} belong to different components of G3G_{3}, then N1N2N_{1}\cup N_{2} separates H1H_{1} from H2H_{2} in GG. Since by (4), α(N1N2)16×2\alpha(N_{1}\cup N_{2})\leq 16\times 2, the theorem holds. Thus we may assume that there is a component FF of G3G_{3} such that H1H2FH_{1}\cup H_{2}\subseteq F.

(10) There is a minimal separator ZZ in FF such that distF(Z,Hi)2\operatorname{dist}_{F}(Z,H_{i})\geq 2 for i{1,2}i\in\{1,2\}, and there exist distinct full components F1,F2F_{1},F_{2} for ZZ such that HiFiH_{i}\subseteq F_{i}.

Let X=NF2(H1)X=N_{F}^{2}(H_{1}). It follows that XX separates H1H_{1} from H2H_{2} in FF, and since distF(H1,H2)4\operatorname{dist}_{F}(H_{1},H_{2})\geq 4, we have distF(H2,X)2\operatorname{dist}_{F}(H_{2},X)\geq 2. For i{1,2}i\in\{1,2\}, let DiD_{i} be the component of FXF\setminus X such that HiDiH_{i}\subseteq D_{i}. Let Y=N(D1)Y=N(D_{1}), let D2D_{2}^{\prime} be the component of FYF\setminus Y such that D2D2D_{2}\subseteq D_{2}^{\prime}, and let Z=N(D2)Z=N(D_{2}^{\prime}). Then ZYZ\subseteq Y, and D1,D2D_{1},D_{2}^{\prime} are full components for ZZ. Setting F1=D1F_{1}=D_{1} and F2=D2F_{2}=D_{2}^{\prime}, (4) follows.

Let Z,F1,F2Z,F_{1},F_{2} be as in (4). We are now ready to complete the proof of the theorem. The proof is by induction on n|H1||H2|n-|H_{1}|-|H_{2}|. If n|H1||H2|=0n-|H_{1}|-|H_{2}|=0, then X=X=\emptyset works. Likewise, if Z=Z=\emptyset, we set X=X=\emptyset. Since F1F2=F_{1}\cap F_{2}=\emptyset, we may assume that |F1H1|<n|H1||H2|2|F_{1}\setminus H_{1}|<\frac{n-|H_{1}|-|H_{2}|}{2}. Let F=F1F2ZF^{\prime}=F_{1}\cup F_{2}\cup Z. Let H2=F2ZH_{2}^{\prime}=F_{2}\cup Z. Then δF(H2)=Z\delta_{F^{\prime}}(H_{2}^{\prime})=Z and H2H_{2}^{\prime} is cooperative in FF^{\prime}. Since distF(H1,Z)2\operatorname{dist}_{F^{\prime}}(H_{1},Z)\geq 2, we have that H1H_{1} is anticomplete to H2H_{2}^{\prime}. Also

|F||H1||H2||F1||H1|n|H1||H2|12.|F^{\prime}|-|H_{1}|-|H_{2}^{\prime}|\leq|F_{1}|-|H_{1}|\leq\frac{n-|H_{1}|-|H_{2}|-1}{2}.

Inductively, there exists XF(H1H2)X^{\prime}\subseteq F^{\prime}\setminus(H_{1}\cup H_{2}^{\prime}) with

α(X)16×2log(n|H1||H2|12+1)16×2log(n|H1||H2|+1)16×2\alpha(X^{\prime})\leq 16\times 2\log\left(\frac{n-|H_{1}|-|H_{2}|-1}{2}+1\right)\leq 16\times 2\log\left(n-|H_{1}|-|H_{2}|+1\right)-16\times 2

such that XX^{\prime} separates H1H_{1} from H2=F2ZH_{2}^{\prime}=F_{2}\cup Z in FF^{\prime}. Let X=XN1N2X=X^{\prime}\cup N_{1}\cup N_{2}. Then XX separates H1H_{1} from H2H_{2} in GG. By (4), α(X)16×2log(n|H1||H2|+1)\alpha(X)\leq 16\times 2\log(n-|H_{1}|-|H_{2}|+1) as required. ∎

5. Large stable subsets in neighborhoods

In this section, we prove a statement about theta-free graphs which we expect to use in future papers.

Lemma 5.1.

Let GG be a theta-free graph. Let c2c\geq 2 be integer, and let YY be a set with α(Y)>24c2\alpha(Y)>24c^{2}. Let ZZ be the set of all vertices such that α(N(z)Y)α(Y)c\alpha(N(z)\cap Y)\geq\frac{\alpha(Y)}{c}. Then α(Z)<2c\alpha(Z)<2c.

Proof.

Suppose not, and let IZI\subseteq Z be a stable set of size 2c2c. For every zIz\in I, let J(z)J^{\prime}(z) be a stable set in N(z)YN(z)\cap Y with |J(z)|=α(Y)c|J^{\prime}(z)|=\left\lceil\frac{\alpha(Y)}{c}\right\rceil.

(11) For all distinct z,zIz,z^{\prime}\in I, α(N[z]N[z])2\alpha(N[z]\cap N[z^{\prime}])\leq 2.

Suppose that α(N[z]N[z])3\alpha(N[z]\cap N[z^{\prime}])\geq 3. Since zz is non-adjacent to zz^{\prime}, there exists a stable set of size three in N(z)N(z)N(z)\cap N(z^{\prime}), and we get a theta with ends z,zz,z^{\prime}, a contradiction. This proves (5).

(12) For all distinct z,zIz,z^{\prime}\in I, |J(z)N(J(z)N(z))|4|J^{\prime}(z)\cap N(J^{\prime}(z^{\prime})\setminus N(z))|\leq 4.

Suppose not, and let {n1,..,n5}J(z)N(J(z))\{n_{1},..,n_{5}\}\subseteq J^{\prime}(z)\cap N(J^{\prime}(z^{\prime})). Then {n1,,n5}\{n_{1},\dots,n_{5}\} is a stable set. If some hJ(z)h\in J^{\prime}(z^{\prime}) has three neighbors in {n1,..,n5}\{n_{1},..,n_{5}\}, then we get a theta with ends z,hz,h; so no such hh exists. It follows that there exist h1,h2,h3J(z)h_{1},h_{2},h_{3}\in J^{\prime}(z^{\prime}) such that (permuting n1,,n5n_{1},\dots,n_{5} if necessary) nihjn_{i}h_{j} is an edge if and only if i=ji=j. But now {z,n1,n2,n3,h1,h2,h3,z}\{z,n_{1},n_{2},n_{3},h_{1},h_{2},h_{3},z^{\prime}\} is a theta with ends z,zz,z^{\prime}, again a contradiction. This proves (5).

Let J(z)=J(z)zI{z}(N[z](N(J(z)N(z))))J(z)=J^{\prime}(z)\setminus\bigcup_{z^{\prime}\in I\setminus\{z\}}(N[z^{\prime}]\cup(N(J^{\prime}(z^{\prime})\setminus N(z)))). By (5) and (5) it follows that

|J(z)||J(z)|6|I|α(Y)c12c.|J(z)|\geq|J^{\prime}(z)|-6|I|\geq\frac{\alpha(Y)}{c}-12c.

But for all distinct z,zIz,z^{\prime}\in I the sets J(z),J(z)J(z),J(z^{\prime}) are disjoint and anticomplete to each other; it follows that zIJ(z)\bigcup_{z\in I}J(z) is a stable set of size 2c(α(Y)c12c)2c\left(\frac{\alpha(Y)}{c}-12c\right). Consequently,

2c(α(Y)c12c)α(Y)2c\left(\frac{\alpha(Y)}{c}-12c\right)\leq\alpha(Y)

and so α(Y)24c2\alpha(Y)\leq 24c^{2}, a contradiction. ∎

6. From domination to stability

The last step in the proof of Theorem 1.2 is to transform balanced separators with small domination number into balanced separators with small stability number.

The results in this section are more general than what we need in this paper; again they are to be used in future papers in the series. Let L,d,rL,d,r be integers. We say that an nn-vertex graph GG is (L,d,r)(L,d,r)-breakable if

  1. (1)

    for every two disjoint and anticomplete cliques H1,H2H_{1},H_{2} of GG with |H1|r|H_{1}|\leq r and |H2|r|H_{2}|\leq r, there is a set XG(H1H2)X\subseteq G\setminus(H_{1}\cup H_{2}) with α(X)L\alpha(X)\leq L separating H1H_{1} from H2H_{2}, and

  2. (2)

    for every normal weight function ww on GG and for every induced subgraph GG^{\prime} of GG there exists a set YV(G)Y\subseteq V(G^{\prime}) with |Y|d|Y|\leq d such that for every component DD of GN[Y]G^{\prime}\setminus N[Y], w(D)12w(D)\leq\frac{1}{2}.

We prove:

Theorem 6.1.

Let d>0d>0 be an integer and let C(d)=100d2C(d)=100d^{2}. Let L,d,n,r>0L,d,n,r>0 be integers such that rd(2+logn)r\leq d(2+\log n) and let GG be an nn-vertex (L,d,r)(L,d,r)-breakable theta-free graph. Then there exists a ww-balanced separator YY in GG such that α(Y)C(d)d(2+logn)r(2+logn)L\alpha(Y)\leq C(d)\left\lceil\frac{d(2+\log n)}{r}\right\rceil(2+\log n)L.

We start by proving a variant of Theorem 3.1 for (L,d,1)(L,d,1)-breakable graphs.

Theorem 6.2.

Let L,dL,d be integers, and let GG be an (L,d,1)(L,d,1)-breakable graph. Let ww be a normal weight function on GG. Then there exist a clique KK in GG and a set Y(K)V(G)KY(K)\subseteq V(G)\setminus K such that

  • |K|d|K|\leq d,

  • α(Y(K))d2L\alpha(Y(K))\leq d^{2}L and

  • N[K]Y(K)N[K]\cup Y(K) is a ww-balanced separator in GG.

Proof.

Since GG is (L,d,1)(L,d,1)-breakable, there exists XV(G)X\subseteq V(G) with |X|d|X|\leq d such that for every component DD of GN[X]G\setminus N[X], w(D)12w(D)\leq\frac{1}{2}. For every pair x,xx,x^{\prime} of non-adjacent vertices of XX, let Y(x,x)Y(x,x^{\prime}) be a set with α(Y(x,x))L\alpha(Y(x,x^{\prime}))\leq L and Y(x,x){x,x}=Y(x,x^{\prime})\cap\{x,x^{\prime}\}=\emptyset separating xx form xx^{\prime} in GG (such a set exists since GG (L,d,1)(L,d,1)-breakable). Now let

Y=Xx,xX non-adjacent Y(x,x).Y=X\cup\bigcup_{x,x^{\prime}\in X\text{ non-adjacent }}Y(x,x^{\prime}).

Then α(Y)(d2)L+dd2L\alpha(Y)\leq\binom{d}{2}L+d\leq d^{2}L. If YY is a ww-balanced separator of GG, set K=K=\emptyset and Y(K)=YY(K)=Y, and the theorem holds. Thus we may assume that there is a component DD of GYG\setminus Y with w(D)>12w(D)>\frac{1}{2}. Let KXK\subseteq X be the set of vertices of XX with a neighbor in DD. Since every two non-adjacent vertices of XX are separated by YY, it follows that KK is a clique.

We claim that N[K]YN[K]\cup Y is a ww-balanced separator in GG. Suppose not, and let DD^{\prime} be the component of G(N[K]Y)G\setminus(N[K]\cup Y) with w(D)>12w(D^{\prime})>\frac{1}{2}. Then DDD^{\prime}\subseteq D. But DN(X)DN(K)YD\cap N(X)\subseteq D\cap N(K)\subseteq Y, and consequently DN[X]=D^{\prime}\cap N[X]=\emptyset, contrary to the fact that N[X]N[X] is a ww-balanced separator in GG. Thus, setting Y(K)=YY(K)=Y, the theorem holds. ∎

We are now ready to prove Theorem 6.1. Let us briefly describe the idea of the proof. Throughout, we have Zi1Z_{i-1}, part of the separator we are building, and a clique Li1L_{i-1} and cliques K1,,Ki1K_{1},\dots,K_{i-1} such that N(KjLi1)N(K_{j}\cap L_{i-1}) is a balanced separator in GZi1G\setminus Z_{i-1}.

This means that in each step, we add at least one vertex vv to create LiL_{i} from Li1L_{i-1} (or LiKiL_{i}\cap K_{i} is empty, and then ZiZ_{i} is the balanced separator we are looking for), which means that vv has a neighbor in each previous KjK_{j} (since LiL_{i} remains a clique).

Now our goal becomes controlling vertices with a neighbor in each KjK_{j}; we call them Badi1Bad_{i-1} and we may assume they have big stability number. This tells us via Lemma 5.1 that we can remove a set of small stability number, and for all remaining vertices, their neighbors in Badi1Bad_{i-1} have stability number only a small fraction of α(Badi1)\alpha(Bad_{i-1}).

Theorem 6.2 gives us a balanced separator, but it consists of a clique (KiK_{i}), its neighbours, and some other vertices (Y(Ki)Y(K_{i})). We add KiK_{i} and Y(Ki)Y(K_{i}) to our separator; both have bounded stability number. Since KiK_{i} is small, with Badi=N(Ki)Badi1Bad_{i}=N(K_{i})\cap Bad_{i-1}, we have α(Badi)α(Badi1)/2\alpha(Bad_{i})\leq\alpha(Bad_{i-1})/2, which means that after logarithmically many steps, BadiBad_{i} will be empty and the process terminates.

It remains to build LiL_{i}. To do so, we find a small set of separators for each vertex vKiv\in K_{i} and its non-neighbors in Li1L_{i-1} and add it to ZiZ_{i}. Now, for the big component DD^{\prime} of GZiG\setminus Z_{i}, the parts of KiK_{i} and Li1L_{i-1} that attach to DD^{\prime} are complete to each other, and it turns out that we can restrict ourselves to those parts for LiL_{i}.

Proof.

Let GG be an (L,d,r)(L,d,r)-breakable graph on nn vertices. Let C(d)=200d2C(d)=200d^{2}. We define several sequences of subgraphs of GG and subsets of V(G)V(G). Let G0=GG_{0}=G; let K0=Y(K0)=L0=Z0=K_{0}=Y(K_{0})=L_{0}=Z_{0}=\emptyset and let Bad0=V(G)Bad_{0}=V(G). We iteratively define Gi,Ki,Li,Zi,BadiG_{i},K_{i},L_{i},Z_{i},Bad_{i} with the following properties:

  1. (I)

    If i>0i>0, then KiZi1=K_{i}\cap Z_{i-1}=\emptyset.

  2. (II)

    α(Zi)C(d)Ldiri/2\alpha(Z_{i})\leq C(d)L\cdot\lceil\frac{di}{r}\rceil i/2.

  3. (III)

    Gi=GZiG_{i}=G\setminus Z_{i}.

  4. (IV)

    If vGiv\in G_{i} has a neighbor in KjK_{j} for every j{1,,i}j\in\{1,\dots,i\}, then vBadiv\in Bad_{i}.

  5. (V)

    α(Badi)n2i\alpha(Bad_{i})\leq\frac{n}{2^{i}}.

  6. (VI)

    LiL_{i} is a clique and |Li|di|L_{i}|\leq di.

  7. (VII)

    For all 1ji1\leq j\leq i, ZiN(KjLi)Z_{i}\cup N(K_{j}\cap L_{i}) is a ww-balanced separator in GG.

We proceed as follows. Suppose that we have defined Gi1,Ki1,Li1,Zi1,Badi1G_{i-1},K_{i-1},L_{i-1},Z_{i-1},Bad_{i-1} satisfying the properties above. If Zi1Z_{i-1} is a ww-balanced separator of GG, we stop the construction. Otherwise, we construct Gi,Ki,Li,Zi,BadiG_{i},K_{i},L_{i},Z_{i},Bad_{i} and show that the properties above continue to hold. If α(Badi1)>96d2\alpha(Bad_{i-1})>96d^{2}, let ZZ be obtained by applying Lemma 5.1 to GG with Y=Badi1Y=Bad_{i-1} and c=2dc=2d; then α(Z)4d\alpha(Z)\leq 4d. If α(Badi1)96d2\alpha(Bad_{i-1})\leq 96d^{2}, let Z=Badi1Z=Bad_{i-1}.

In both cases, α(Z)96d2\alpha(Z)\leq 96d^{2}. Let G=Gi1ZG^{\prime}=G_{i-1}\setminus Z. Let

w(v)=w(v)ΣvV(G)w(v).w^{\prime}(v)=\frac{w(v)}{\Sigma_{v\in V(G^{\prime})}w(v)}.

Then ww^{\prime} is a normal weight function on GG^{\prime}, and for every vGv\in G^{\prime}, w(v)w(v)w^{\prime}(v)\geq w(v). Let Ki,Y(Ki)K_{i},Y(K_{i}) be as in Theorem 6.2 applied to GG^{\prime} and ww^{\prime}. It follows that KiK_{i} is a clique of size at most dd, and KGGZi1K\subseteq G^{\prime}\subseteq G\setminus Z_{i-1}, so (I) holds (for ii).

Let vKiv\in K_{i}. By (VI) (for i1i-1), Li1N(v)L_{i-1}\setminus N(v) can be partitioned into at most d(i1)r\left\lceil\frac{d(i-1)}{r}\right\rceil cliques each of size at most rr. Since GG is (L,d,r)(L,d,r)-breakable, this implies that there exists a set Z(v)Z(v) with α(Z(v))d(i1)rL\alpha(Z(v))\leq\left\lceil\frac{d(i-1)}{r}\right\rceil L, such that Z(v)Z(v) separates {v}\{v\} from Li1N(v)L_{i-1}\setminus N(v) in GG. Let Z=vKiZ(v)Z^{\prime}=\bigcup_{v\in K_{i}}Z(v); then α(Z)dd(i1)rL\alpha(Z^{\prime})\leq d\left\lceil\frac{d(i-1)}{r}\right\rceil L.

Let Zi=Zi1ZZKiY(Ki)Z_{i}=Z_{i-1}\cup Z^{\prime}\cup Z\cup K_{i}\cup Y(K_{i}). Next, we have:

α(ZiZi1)dd(i1)rL+96d2+1+d2LC(d)d(i1)rL/2.\alpha(Z_{i}\setminus Z_{i-1})\leq d\left\lceil\frac{d(i-1)}{r}\right\rceil L+96d^{2}+1+d^{2}L\leq C(d)\left\lceil\frac{d(i-1)}{r}\right\rceil L/2.

It follows that α(Zi)C(d)dirL×i/2\alpha(Z_{i})\leq C(d)\left\lceil\frac{di}{r}\right\rceil L\times i/2, and (II) holds.

Let Gi=Gi1Zi=GZiG_{i}=G_{i-1}\setminus Z_{i}=G\setminus Z_{i}; now (III) holds. Let Badi=Badi1NG(Ki)V(Gi)Bad_{i}=Bad_{i-1}\cap N_{G^{\prime}}(K_{i})\cap V(G_{i}); now (IV) holds. Since G=Gi1ZG^{\prime}=G_{i-1}\setminus Z and either Z=Badi1Z=Bad_{i-1} or ZZ is the set of all vertices vv such that α(N(v)Badi)α(Badi)/2d\alpha(N(v)\cap Bad_{i})\geq\alpha(Bad_{i})/2d and by (V) for i1i-1, we have that for every vKiv\in K_{i}, α(NG(v)Badi1)n2id\alpha(N_{G^{\prime}}(v)\cap Bad_{i-1})\leq\frac{n}{2^{i}d}. It follows that α(Badi)n2i\alpha(Bad_{i})\leq\frac{n}{2^{i}}, and (V) holds for ii.

If ZiZ_{i} is a ww^{\prime}-balanced separator in Gi1G_{i-1}, let Li=Li1L_{i}=L_{i-1}; now (VI) and (VII) hold. Thus we may assume not, and let DD^{\prime} be the maximal connected subset of Gi=GZiG_{i}=G\setminus Z_{i} with w(D)>12w(D^{\prime})>\frac{1}{2}.

We now define LiL_{i} and check that (VI) and (VII) hold. Since N[Ki]Y(Ki)N[K_{i}]\cup Y(K_{i}) is a ww^{\prime}-balanced separator in GG^{\prime} and since GiGKiY(Ki)G_{i}\subseteq G^{\prime}\setminus K_{i}\cup Y(K_{i}), it follows that DN(Ki)D^{\prime}\cap N(K_{i})\neq\emptyset, and so N(D)KiN(D^{\prime})\cap K_{i}\neq\emptyset. Since ZGi=Z^{\prime}\cap G_{i}=\emptyset, and since for every vKiv\in K_{i}, Z(v)ZZ(v)\subseteq Z^{\prime} separates vv from Li1N(v)L_{i-1}\setminus N(v), it follows that N(D)KiN(D^{\prime})\cap K_{i} is complete to N(D)Li1N(D^{\prime})\cap L_{i-1}; let

Li=(N(D)Li1)(N(D)Ki).L_{i}=(N(D^{\prime})\cap L_{i-1})\cup(N(D^{\prime})\cap K_{i})\mbox{.}

Then LiL_{i} is a clique, and no vertex of KiLiK_{i}\setminus L_{i} has a neighbor in DD^{\prime}. Since LiLi1KiL_{i}\setminus L_{i-1}\subseteq K_{i}, (VI) holds.

In order to prove (VII), let us consider first the case j=ij=i. Suppose for a contradiction that DD is a component of G(ZiN(KiLi))G\setminus(Z_{i}\cup N(K_{i}\cap L_{i})) with w(D)>1/2w(D)>1/2. Then DDD\subseteq D^{\prime}. Moreover, KiLi=N(D)KiK_{i}\cap L_{i}=N(D^{\prime})\cap K_{i}, and so NGi(KiLi)D=NGi(Ki)DN_{G_{i}}(K_{i}\cap L_{i})\cap D^{\prime}=N_{G_{i}}(K_{i})\cap D^{\prime}. Therefore, DDNGi(Ki)GiN(Ki)D\subseteq D^{\prime}\setminus N_{G_{i}}(K_{i})\subseteq G_{i}\setminus N(K_{i}). However, since N[Ki]Y(Ki)N[K_{i}]\cup Y(K_{i}) is a ww^{\prime}-balanced separator in GG^{\prime}, and since KiY(Ki)ZiK_{i}\cup Y(K_{i})\subseteq Z_{i}, it follows that w(D)w(D)1/2w(D)\leq w^{\prime}(D)\leq 1/2.

Next, we consider the case when j<ij<i. By (VII) for i1i-1, we know that ZiN(KjLi1)Z_{i}\cup N(K_{j}\cap L_{i-1}) is a ww-balanced separator in GG. Suppose for a contradiction that DD is a component of G(ZiN(KjLi))G\setminus(Z_{i}\cup N(K_{j}\cap L_{i})) with w(D)>1/2w(D)>1/2. Then DDD\subseteq D^{\prime}. We have that N(D)KjLi1LiKjN(D^{\prime})\cap K_{j}\cap L_{i-1}\subseteq L_{i}\cap K_{j}, and therefore, ND(Li1Kj)ND(LiKj)N_{D^{\prime}}(L_{i-1}\cap K_{j})\subseteq N_{D^{\prime}}(L_{i}\cap K_{j}). It follows that DDN(Li1Kj)G(ZiN(Li1Kj))D\subseteq D^{\prime}\setminus N(L_{i-1}\cap K_{j})\subseteq G\setminus(Z_{i}\cup N(L_{i-1}\cap K_{j})), a contradiction as ZiN(Li1Kj)Z_{i}\cup N(L_{i-1}\cap K_{j}) is a ww-balanced separator in GG.

We have shown that properties (I)–(VII) are maintained at each step of the construction.

We can now complete the proof of the theorem. It follows immediately from (V) that there exists k1+lognk\leq 1+\log n such that Badk=Bad_{k}=\emptyset. We claim that ZkZ_{k} is a balanced separator in GG (and in particular the construction stops). Suppose not. Then the construction continues and the sets Gk+1,Kk+1,Lk+1,Zk+1,Badk+1G_{k+1},K_{k+1},L_{k+1},Z_{k+1},Bad_{k+1} are defined. Also, there exists a component DD of GZk=GkG\setminus Z_{k}=G_{k} such that w(D)>12w(D)>\frac{1}{2}. We apply (VII) with i=k+1i=k+1. It follows that there is a vertex vLk+1Kk+1v\in L_{k+1}\cap K_{k+1}. By (I) and (III), it follows that vGkv\in G_{k}. Since Lk+1KjL_{k+1}\cap K_{j}\neq\emptyset for every 1jk1\leq j\leq k (again by (VII)), we deduce from (IV) and (VI) that vBadkv\in Bad_{k}, a contradiction.

Now by (II) we have

α(Zk+1)C(d)d(k+1)rL×(k+1)/2C(d)d(logn+2)r(2+logn)L,\alpha(Z_{k+1})\leq C(d)\left\lceil\frac{d(k+1)}{r}\right\rceil L\times(k+1)/2\leq C(d)\frac{d(\log n+2)}{r}(2+\log n)L,

as required.

7. The proof of Theorem 1.2

We start with a lemma. (There are many versions of this lemma; we chose one with a simple proof, without optimizing the constants.)

Lemma 7.1.

Let GG be a graph, let c[12,1)c\in[\frac{1}{2},1), and let dd be a positive integer. If for every normal weight function ww on GG, there is a (c,w)(c,w)-balanced separator XwX_{w} with α(Xw)d\alpha(X_{w})\leq d, then the tree independence number of GG is at most 3c1cd\frac{3-c}{1-c}d.

Proof.

We will prove that for every set ZV(G)Z\subseteq V(G) with α(Z)21cd\alpha(Z)\leq\frac{2}{1-c}d there is a tree decomposition (T,χ)(T,\chi) of GG such that α(χ(v))3c1cd\alpha(\chi(v))\leq\frac{3-c}{1-c}d for every vTv\in T, and there exists tTt\in T such that Zχ(t)Z\subseteq\chi(t).

The proof is by induction on |V(G)||V(G)|. Observe that every induced subgraph of GG satisfies the assumption of the theorem.

Let ZV(G)Z\subseteq V(G) with α(Z)21cd\alpha(Z)\leq\frac{2}{1-c}d. Let II be a stable set of ZZ with |I|=α(Z)|I|=\alpha(Z). Define a function ww where w(v)=1|I|w(v)=\frac{1}{|I|} if vIv\in I, and w(v)=0w(v)=0 if vIv\not\in I. Then ww is a normal weight function on GG. Let XX be a (c,w)(c,w)-balanced separator with α(X)d\alpha(X)\leq d. Then V(G)X=V1VqV(G)\setminus X=V_{1}\cup\dots\cup V_{q}, where V1,,VqV_{1},\dots,V_{q} are the components of GXG\setminus X, and |IVi|c|I||I\cap V_{i}|\leq c|I| for i{1,,q}i\in\{1,\dots,q\}. Let i{1,,q}i\in\{1,\dots,q\}. Define Zi=(ZVi)XZ_{i}=(Z\cap V_{i})\cup X. Since |IVi|c|I||I\cap V_{i}|\leq c|I|, it follows that |I(GVi)|(1c)|I||I\cap(G\setminus V_{i})|\geq(1-c)|I|. Since α(X)d\alpha(X)\leq d, we have that |I(G(ViX))|(1c)|I|d|I\cap(G\setminus(V_{i}\cup X))|\geq(1-c)|I|-d. It follows that α(ZVi)c|I|+d\alpha(Z\cap V_{i})\leq c|I|+d. Consequently,

α(Zi)α(ZVi)+α(X)c|I|+2d2c1cd+2(1c)1cd=21cd.\alpha(Z_{i})\leq\alpha(Z\cap V_{i})+\alpha(X)\leq c|I|+2d\leq\frac{2c}{1-c}d+\frac{2(1-c)}{1-c}d=\frac{2}{1-c}d.

Inductively, for i{1,,q}i\in\{1,\dots,q\}, there is a tree decomposition (Ti,χi)(T_{i},\chi_{i}) of ViXV_{i}\cup X such that α(χi(t))3c1cd\alpha(\chi_{i}(t))\leq\frac{3-c}{1-c}d for every vTiv\in T_{i}, and there exists tiTit_{i}\in T_{i} such that Ziχ(t)Z_{i}\subseteq\chi(t). Now let TT be obtained from the disjoint union of T1,,TqT_{1},\dots,T_{q} by adding a new vertex t0t_{0} adjacent to t1,,tqt_{1},\dots,t_{q} (and with no other neighbors). Define χ(t)=χi(t)\chi(t)=\chi_{i}(t) for every tTit\in T_{i}, and let χ(t0)=ZX\chi(t_{0})=Z\cup X. Since

α(ZX)α(Z)+α(X)21cd+d=3c1cd,\alpha(Z\cup X)\leq\alpha(Z)+\alpha(X)\leq\frac{2}{1-c}d+d=\frac{3-c}{1-c}d,

(T,χ)(T,\chi) satisfies the conclusion of the theorem. ∎

Next we restate and prove Theorem 1.2:

Theorem 7.2.

There exists a constant cc such that for every integer n>1n>1 every nn-vertex graph G𝒞G\in\mathcal{C} has tree independence number at most at most c(logn)2c(\log n)^{2}.

Proof.

Let G𝒞G\in\mathcal{C}. Let dd be as in Theorem 3.1. Let C(d)C(d) be as in Theorem 6.1, and let c=165C(d)c=165C(d). We may assume that ncn\geq c. Let r=d(2+logn)r=d(2+\log n), and let L=32lognL=32\log n. By Theorems 3.1 and 4.5, and since every clique is cooperative, it follows that GG is (L,d,r)(L,d,r)-breakable. Now by Theorem 6.1, for every normal weight function ww on GG, there exists a ww-balanced separator YY in GG such that α(Y)C(d)d(2+logn)r(2+logn)L33C(d)(logn)2\alpha(Y)\leq C(d)\frac{d(2+\log n)}{r}(2+\log n)L\leq 33C(d)(\log n)^{2}. Now Theorem 7.2 follows from Lemma 7.1. ∎

8. Algorithmic consequences

Theorem 7.2 implies quasi-polynomial time (namely 2(logn)O(1)2^{(\log n)^{O(1)}} time) algorithms for a number of problems. In particular Dallard et al. [6] gave nO(k)n^{O(k)} time algorithms for Maximum Weight Independent Set and Maximum Weight Induced Matching assuming that a tree decomposition with independence number at most kk is given as input. Subsequently Dallard et al. [5] gave an algorithm that takes as input a graph GG and integer kk, runs in time 2O(k2)nO(k)2^{O(k^{2})}n^{O(k)} and either outputs a tree decomposition of GG with independence number at most 8k8k, or determines that the tree independence number of GG is larger than kk. Theorem 7.2, together with these results (setting k=clog2nk=c\log^{2}n), immediately imply the following theorem.

Theorem 8.1.

Maximum Weight Independent Set and Maximum Weight Induced Matching admit algorithms with running time nO((logn)3)n^{O((\log n)^{3})} on graphs in 𝒞\mathcal{C}.

It is worth mentioning that the nO(k)n^{O(k)} time algorithm of Dallard et al. [6] works for a slightly more general packing problem (see their Theorem 7.2 for a precise statement) that simultaneously generalizes Maximum Weight Independent Set and Maximum Weight Induced Matching. Thus we could have stated Theorem 8.1 for this even more general problem.

Lima et al. [11] observed that the algorithm of Dallard et al. [6] can be generalized to a much more general class of problems. In particular they show that for every integer \ell and CMSO2 formula ϕ\phi, there exists an algorithm that takes as input a graph GG of tree independence at most kk, and a weight function w:V(G)w:V(G)\rightarrow\mathbb{N}, runs in time f(k,ϕ,)nO(k)f(k,\phi,\ell)n^{O(\ell k)} and outputs a maximum weight vertex subset SS such that G[S]G[S] has treewidth at most \ell and G[S]ϕG[S]\models\phi. This formalism captures Maximum Weight Independent Set, Maximum Weight Induced Matching as well as Maximum Weight Induced Forest, recognition of many well-studied graph classes (including 𝒞\mathcal{C}) and a host of other problems. We remark that their result (Theorem 6.2 of [11]) is stated in terms of clique number rather than treewidth, however at the very beginning in the proof they show that in this context bounded clique number implies treewidth at most \ell and then proceed to prove the theorem as stated here.

Unfortunately the algorithm of [11] does not give any meaningful results when combined with Theorem 7.2. The reason is that the function f(k,ϕ,)f(k,\phi,\ell) bounding the running time of the algorithm is a tower of exponentials, which leads to super-exponential running time bounds even when k=clog2nk=c\log^{2}n. However it turns out that the algorithm of [11] can be modified to run in time (f(,ϕ)n)O(k)(f(\ell,\phi)n)^{O(\ell k)} [10], which is quasi-polynomial for every fixed ,ϕ\ell,\phi when k=O(log2n)k=O(\log^{2}n). This improvement immediately leads to the following theorem.

Theorem 8.2.

For every integer \ell and CMSO2 formula ϕ\phi, there exists an algorithm that takes as input a graph G𝒞G\in\mathcal{C} and a weight function w:V(G)w:V(G)\rightarrow\mathbb{N}, runs in time (f(ϕ,)n)O(log2n)(f(\phi,\ell)n)^{O(\ell\log^{2}n)} and outputs a maximum weight vertex subset SS such that G[S]G[S] has treewidth at most \ell and G[S]ϕG[S]\models\phi.

We refer to [11] for a discussion of the set of problems that are captured by Theorem 8.2.

9. Acknowledgments

Peter Gartland proved Theorem 3.1 independently using a different method. We thank him for sharing his proof with us, and for many useful discussions.

References

  • [1] T. Abrishami, M. Chudnovsky, C. Dibek, and K. Vušković. Submodular functions and perfect graphs. to appear in Mathematics of Operations Research.
  • [2] T. Abrishami, M. Chudnovsky, S. Hajebi, and S. Spirkl. Induced subgraphs and tree decompositions III. Three-path-configurations and logarithmic treewidth. Adv. Comb., pages Paper No. 6, 29, 2022.
  • [3] H. L. Bodlaender. Dynamic programming on graphs with bounded treewidth. In Automata, languages and programming (Tampere, 1988), volume 317 of Lecture Notes in Comput. Sci., pages 105–118. Springer, Berlin, 1988.
  • [4] M. Chudnovsky, P. Gartland, S. Hajebi, D. Lokshtanov, and S. Spirkl. Induced subgraphs and tree decompositions XV. Even-hole-free graphs have logarithmic treewidth. Preprint available at https://arxiv.org/abs/2402.14211.
  • [5] C. Dallard, F. Fomin, T. Korhonen, P. Golovach, and M. Milanič. Computing Tree Decompositions with Small Independence Number. Preprint available at https://arxiv.org/abs/2207.09993.
  • [6] C. Dallard, M. Milanič, and K. Štorgel. Treewidth versus clique number. II. Tree independence number. Preprint available at https://arxiv.org/abs/2111.04543.
  • [7] P. Erdös and G. Szekeres. A combinatorial problem in geometry. Compositio Math., 2:463–470, 1935.
  • [8] M. R. Garey and D. S. Johnson. “Strong” NP-completeness results: motivation, examples, and implications. J. Assoc. Comput. Mach., 25(3):499–508, 1978.
  • [9] M. C. Golumbic. Algorithmic aspects of perfect graphs. In Topics on perfect graphs, volume 88 of North-Holland Math. Stud., pages 301–323. North-Holland, Amsterdam, 1984.
  • [10] P. T. Limma, M. Milanič, P. Muršič, K. Okrasa, P. Rzążewski, and K. Štorgel. Private communication.
  • [11] P. T. Limma, M. Milanič, P. Muršič, K. Okrasa, P. Rzążewski, and K. Štorgel. Tree decompositions meet induced matchings: beyond max weight independent set. Preprint available at https://arxiv.org/abs/2402.15834.
  • [12] N. Robertson and P. D. Seymour. Graph minors. XVI. Excluding a non-planar graph. J. Combin. Theory Ser. B, 89(1):43–76, 2003.
  • [13] A. Scott and P. Seymour. A survey of χ\chi-boundedness. J. Graph Theory, 95(3):473–504, 2020.
  • [14] 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):475–509, 2021.