Induced subgraphs and tree decompositions
XV. Even-hole-free graphs with bounded clique number have logarithmic treewidth
Abstract.
We prove that for every integer there exists an integer such that every -vertex even-hole-free graph with no clique of size has treewidth at most . 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 , -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 such that every even-hole-free graph has a balanced separator which is contained in the (closed) neighborhood of at most 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.
1. Introduction
All graphs in this paper are finite and simple. For a graph , we denote by the vertex set of , and by the edge set of . For graphs , we say that is an induced subgraph of if , and are adjacent in if and only if . 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, -Coloring and Coloring.
For a graph , a tree decomposition of consists of a tree and a map with the following properties:
-
(i)
For every vertex , there exists such that .
-
(ii)
For every edge , there exists such that .
-
(iii)
For every vertex , the subgraph of induced by is connected.
For each , we refer to as a bag of . The width of a tree decomposition , denoted by , is . The treewidth of , denoted by , is the minimum width of a tree decomposition of .
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 -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 , there exists a constant such that every even-hole-free graph with no induced diamond and no clique of size satisfies .
Conjecture 1.2 (Sintiari and Trotignon [36]).
For every integer , there exists a constant such that every even-hole-free graph with no clique of size satisfies .
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 , -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 be a graph. For a set , we denote by the subgraph of induced by . For , we denote by the subgraph induced by . In this paper, we use induced subgraphs and their vertex sets interchangeably.
For graphs and , we say that contains if some induced subgraph of is isomorphic to . For a family of graphs, contains if contains a member of , and we say that is -free if does not contain . 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 . The open neighborhood of , denoted by , is the set of all vertices in adjacent to . The closed neighborhood of , denoted by , is . Let . The open neighborhood of , denoted by , is the set of all vertices in with at least one neighbor in . The closed neighborhood of , denoted by , is . If is an induced subgraph of and , then and . Let be disjoint from . We say is complete to if all possible edges with an end in and an end in are present in , and is anticomplete to if there are no edges between and .
Given a graph , a path in is an induced subgraph of that is a path. If is a path in , we write to mean that , and is adjacent to if and only if . We call the vertices and the ends of , and say that is a path from to . The interior of , denoted by , is the set . The length of a path is the number of edges in . We denote by a cycle with vertices.
Next, we describe a few types of graphs that we will need (see Figure 1).

A theta is a graph consisting of three internally vertex-disjoint paths , , and , each of length at least 2, such that are pairwise anticomplete. In this case we call and the ends of the theta.
A prism is a graph consisting of three vertex-disjoint paths , , and , each of length at least 1, such that and are triangles, and no edges exist between the paths except those of the two triangles.
A pyramid is a graph consisting of three paths , , and such that are pairwise disjoint, at least two of the three paths have length at least two, is triangle (called the base of the pyramid), and no edges exist between except those of the triangle . The vertex is called the apex of the pyramid.
A wheel in is a pair where is a hole and is a vertex with at least three neighbors in . A wheel is even if has an even number of neighbors on .
Let be the class of (, theta, prism, even wheel)-free graphs (sometimes called “-free odd-signable” graphs). For every integer , let be the class of all graphs in with no clique of size . It is easy to see that every even-hole-free graph is in .
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 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 be an integer and let . We say that is a -banana if is non-adjacent to , and there exist pairwise internally-vertex-disjoint paths from to . Note that if is a -banana in , then is an -banana in for every . We prove:
Theorem 1.3.
For every integer , there exists a constant such that if , then contains no -banana.
The next step in the proof of Conjecture 1.2 is the following. Let and let be a normal weight function on , that is, satisfies . A set is a -balanced separator if for every component of . The set is a -balanced separator if is a -balanced separator. We show:
Theorem 1.4.
There is an integer with the following property. Let and let be a normal weight function on . Then there exists such that
-
•
, and
-
•
is a -balanced separator in .
Theorem 1.5.
For every integer , there exists a constant such that every satisfies .
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 be non-adjacent. Recall that a separation of a graph is a triple of pairwise disjoint subsets of with such that is anticomplete to . Similarly to [6], we use the fact that graphs in 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 (which yields a partition of all the natural separations) of an -vertex graph in into collections , where each is “non-crossing” (this property is captured in Lemma 4.9), (where only depends on and works for all ) and each vertex of has at most (where depends on ) neighbors in . We prove a strengthening of Theorem 1.3, which asserts that the size of the largest -banana is bounded by a linear function of .
We observe that a result of [3] implies that there exists a an integer (that depends on ) such that if has no hubs other than , then is not a -banana in . More precisely, the result of [3] states that if and are joined by internally-vertex-disjoint paths , then for at least one , the neighbor of in is a hub.
We now proceed as follows. Let (where comes from the degeneracy partition and from the previous paragraph); suppose that is a -banana in and let be the set of all paths of with ends .
Let denote the partition of hubs as decribed above. We proceed by induction on and describe a process below that finds an induced subgraph of in which:
-
•
Vertices in are no longer hubs;
-
•
If there are not many internally disjoint -paths in , we can “lift” this to .
We first consider a so-called -lean tree decomposition of (discussed in Section 2). By examining the intersection graphs of subtrees of a tree we deduce that there exist two (not necessarily distinct) vertices such that for every path , the interior of meets . We also argue that . A vertex of is bad if has large degree (at least ) in the torso of , or has large degree in the torso of , or is adjacent to both and . We show that there are at most three bad vertices in .
We would like to bound the size of the largest -banana in the union of the two torsos. Unfortunately, the torso of may not be a graph in . Instead, we find an induced subgraph of , which we call , that consists of together with a collection of disjoint vertex sets for and except vertices on the -path in , where each “remembers” the component of that meets . Importantly, no vertex of is a hub of , and all but at most three vertices of have bounded degree in the union of the torso of and the torso of .
We next decompose , simultaneously, by all the separations corresponding to the hubs in that are not bad, and delete the set of (at most three) bad vertices of . We denote the resulting graph by and call it the “central bag” for . The parameter is smaller for than it is for , and so we can use induction to obtain a bound on the largest size of an -banana in . Since our goal is to bound the size of an -banana in by a linear function of , we now need to show that the size of the largest -banana does not grow by more than an additive constant when we move from to .
We start with a smallest subset of that separates from in (and whose size is bounded by Menger’s theorem) and show how to transform in into a set separating from in , while increasing the size of its intersection with by at most an additive constant, and ensuring the number of sets that meets is bounded by a constant. Then we repeat a similar procedure to obtain a set of vertices separating from in , 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 , we add to the neighbor sets of the vertices of . Since while constructing we have deleted all the bad vertices of , the number of vertices of added to is at most , and meets at most of the sets . Note that the bound on the size of that we have depends on , which may be close to , so another argument is needed to obtain a constant bound on and on the number of sets that meet . This is a consequence of Theorem 6.3, because no vertex of is a hub in (this is proved in Theorem 4.10), and no vertex of is a hub in . (The proof of Theorem 6.3 analyzes the structure of minimal separators in graphs in using tools developed in Section 5.)
In the second growing step, we start with the set obtained in the previous paragraph. Then we add to the following subsets (here we describe the cases when ; if the argument is similar).
-
(1)
where is the unique neighbor of in the path in from to .
-
(2)
where is the unique neighbor of in the path in from to .
-
(3)
for every such that .
-
(4)
for every such that .
-
(5)
The set of all bad vertices of .
One of the properties of -lean tree decompositions is that the size of each adhesion (intersection of “neighboring” bags) is less than . The number of adhesions added to is again bounded since the number of distinct sets that meet 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 such that for every graph and every normal weight function on , there is a -balanced separator in such that is contained in the union of the neighborhoods of at most vertices of . 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 . We then use the two kinds of separations (the kind that come from hubs and the kind that come from pyramids) as follows.
For we say that a set dominates if . By Lemma 8.5 (from [14]) there is a path in such that for every component of . We now use a “sliding window” argument: we divide 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 such that, except for our window , the path is disjoint from , and separates the subpath of before our current window from that after our current window. This means (unless is a balanced separator) that the big component of does not contain either the subpath of 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 -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 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 , if a graph contains no -banana, then for every normal weight function on , if has a -balanced separator , then has an -balanced separator and a clique such that . This step uses -lean tree decompositions of and works for all graphs .
Now Theorem 9.2 and Theorem 1.3 together imply that for every , there exists an integer , depending on , such that for every , has a balanced separator of size at most ; that is Theorem 9.1. By Lemma 10.1 this immediately implies the desired bound on the treewidth of .
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 is very restricted. In Section 4 we discuss the construction of the graphs and , and how to use -separators there to obtains -separators in . In Section 5 we analyze the structure of minimal separators in graphs of . 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 -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 , 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 and a set we say that separates from if for every path of with ends and . The following is a well-known variant of a classical theorem due to Menger [30]:
Theorem 2.1 (Menger [30]).
Let be an integer, let be a graph and let be distinct and non-adjacent. Then either there exists a set with such that separates and in , or is a -banana in .
2.2. Lean tree decompositions
We adopt notation from [3]: For a tree and vertices , we denote by the unique path in from to . Let be a tree decomposition of a graph . For every , the adhesion at , denoted by , is the set . For such that (in particular, if ), we set . We define . For every , the torso at , denoted by , is the graph obtained from the bag by, for each , adding an edge between every two non-adjacent vertices .
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 be an integer. A tree decomposition is called -lean if the following hold:
-
•
; and
-
•
for all and sets and with , either contains disjoint paths from to , or some edge of satisfies .
For a tree and an edge of , we denote by the component of containing . Let . A tree decomposition is tight if for every edge there is a component of such that (and therefore ).
Next, we need a definition from [9]. Given a tree decomposition of an -vertex graph , its fatness is the vector where denotes the number of bags of of size . A tree decomposition of is -atomic if and the fatness of is lexicographically minimum among all tree decompositions of with adhesion less than .
Theorem 2.2 (Bellenbaum and Diestel [9], see Carmesin, Diestel, Hamann, Hundertmark [13], see also Weißauer [38]).
Every -atomic tree decomposition is -lean.
We also need:
Theorem 2.3 (Weißauer [38]).
Every -atomic tree decomposition is tight.
Theorem 2.4.
Let be a graph, let and let be -atomic tree decomposition of . Let and let . Assume that is not separated from by a set of size less than , and that is not separated from by a set of size less than . Then is not separated from by a set of size less than , and consequently is a -banana.
Proof.
By Theorems 2.2 and 2.3, we have that is tight and -lean. Suppose that is separated from by a set of size less than . By Theorem 2.1, there exists a set of pairwise vertex-disjoint (except ) paths, each with one end and the other end in , and such that . Let be the set of the ends of members of in . Similarly, there exists a collection of pairwise vertex-disjoint (except ) paths, each with one end and the other end in , and such that . Let be the set of the ends of members of in . Since is -lean, there is a collection of pairwise vertex-disjoint paths from to . Let be a set with separating from . Then, for every - path in we have . But since , there is a path with ends and , a path from to , and a path from to such that , contrary to the fact that is a path from to in . 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 -atomic tree decompositions for every .
A cutset of is a (possibly empty) set of vertices such that is disconnected. A clique cutset is a cutset that is a clique.
Theorem 2.5 (Abrishami, Alecu, Chudnovsky, Hajebi, Spirkl [3]).
Let be a theta-free graph and assume that does not admit a clique cutset. Let be a tight tree decomposition of . Then for every edge of the graph is connected and . Moreover, if and , then .
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 be a theta-free graph and let . Let be a tree decomposition of . Then there exist (not necessarily distinct) such that for every path of with ends we have that Moreover, if is a component of , then .
Proof.
Let be the set of all paths of with ends . For every let . Since is a connected subgraph of , it follows that is a subtree of .
Let be a graph with vertex set and such that if and only if . Then, by the main result of [22], is a chordal graph.
(1) If are non-adjacent in , then , are disjoint and anticomplete to each other in .
Suppose that there exists . Then for every such that , we have , contrary to the fact that are non-adjacent in . Next suppose that there is an edge of such that . Let be such that . Then , again contrary to the fact that is non-adjacent to in . This proves (2.3).
(2) has no stable set of size three.
Suppose is a stable set in . By (2.3)
the sets are pairwise disjoint and anticomplete to
each other in . But now is a theta with ends
in , a contradiction. This proves (2.3).
Since is chordal, it follows from (2.3) and a result from [23] that there exist
cliques of such that . Again by [23] we deduce that for each , there exists such that
for every . Consequently,
for every
, and the first statement of the theorem holds.
To prove the second statement, suppose for some component of . Then is connected, and so there is a path from to with . But now , 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 , a function is a normal weight function on if , where we use the notation to mean for a set of vertices.
Let be a graph and let be a tree decomposition of . Let be a normal weight function on . A vertex of is a center of if for every we have .
The following is well-known; a proof can be found in [3], for example.
Theorem 2.7.
Let be a tree decomposition of a graph . Then has a center.
3. Wheels and star cutsets
Recall that a wheel in is a pair such that is a hole and is a vertex that has at least three neighbors in . 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 is a path of whose ends are distinct and adjacent to , and such that is anticomplete to . A sector is a long sector if is non-empty. We now define several types of wheels that we will need.
A wheel is a universal wheel if is complete to . A wheel is a twin wheel if induces a path of length two. A wheel is a short pyramid if and has exactly two adjacent neighbors in . A wheel is proper if it is not a twin wheel or a short pyramid. We say that is a hub if there exists such that is a proper wheel in . We denote by the set of all hubs of .
We need the following result, which was observed in [6]:
Theorem 3.1 (Abrishami, Chudnovsky, Vušković [6]).
Let and let be a proper wheel in . Then there is no component of such that .
We will revisit this result in Section 8. A star cutset in a graph is a cutset such that either or for some , . A large portion of this paper is devoted to dealing with hubs and star cutsets arising from them in graphs in , but in the remainder of this section we focus on the case when is very restricted. To do that, we use a result from [3]:
Theorem 3.2 (Abrishami, Alecu, Chudnovsky, Hajebi, Spirkl [3]).
For every integer there exists an integer such that if and is a -banana in , then and .
We immediately deduce:
Theorem 3.3.
For every integer , there exists a constant with the following property. Let and let be non-adjacent. Assume that . Then is not a -banana in . In particular, if , there is no -banana in .
We will also use the following:
Theorem 3.4 (Abrishami, Alecu, Chudnovsky, Hajebi, Spirkl [3]).
For every integer , there exists an integer such that every with satisfies .
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 and let be a graph with . Let where is as in Theorem 3.2. Let such that is a -banana in , and let be the set of all paths in with ends . Let be an -atomic tree decomposition of . By Theorems 2.2 and 2.3, we have that is tight and -lean. By Theorem 2.6, there exist such that for every . We observe:
Lemma 4.1.
We have .
Proof.
Suppose that . Let be the component of such that . Then is contained in the union of at most two adhesions of , at most one for each of and . Since is -lean, it follows that . Since for every , it follows that for every . But then contains at most pairwise internally vertex-disjoint paths, contrary to the fact that is a -banana in . ∎
We say that a vertex is -safe if . The goal of the next two lemmas is to classify -safe vertices into “good ones” and “bad ones,” and show that the bad ones are rare. Let . A vertex is -cooperative if either , or . It was shown in [3] that -cooperative vertices have the following important property (note that [3] assumed that is a center of , but this was not used in the proof of the following lemma):
Lemma 4.2 (Abrishami, Alecu, Chudnovsky, Hajebi, Spirkl [3]).
Let . If are -safe and not -cooperative, then is adjacent to .
A vertex is -cooperative if there exists a component of such that . We show:
Lemma 4.3.
If is -safe and not -cooperative, then is adjacent to both and . In particular, the set of vertices that are not -cooperative is a clique.
Proof.
Suppose is non-adjacent to and is not -cooperative. Since is a -banana and , we can choose such that the sets are pairwise vertex-disjoint. Since is not -cooperative, it follows that has a neighbors in each of , and so there exist pairwise vertex-disjoint paths , each with ends , and such that for every . Since is -safe, we may assume (by renumbering the paths if necessary), that for . But now we get a contradiction applying Theorem 3.2 to the graph and the -banana in . This proves the first statement of 4.3. Since is -free, the second statement follows. ∎
In view of Theorem 2.6, it would be convenient if we could work with the graph , but unfortunately this graph may not be in , or even . 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 . Assume that does not admit a clique cutset. For every , there exists such that
-
(1)
.
-
(2)
is connected and .
-
(3)
No vertex of is a hub in the graph .
Similarly to the construction in [3], we now define a graph that is a safe alternative to . Let be a stable set of hubs of with , and assume that every is -safe. Let denote the set of all vertices in that are common neighbors of and , or are not -cooperative, or are not -cooperative. Since is stable, Lemma 4.2 implies that . By Lemma 4.3, every vertex in is -cooperative. If , for , let be the neighbor of in the (unique) path in from to . If , set . Let and set
Write . We fix and throughout this section. It follows that for every and for every , we have that . Let . For , we define to be the set of all vertices such that ; let . Write
Next we summarize several key properties of .
Lemma 4.5.
Suppose that does not admit a clique cutset and let . Then the following hold.
-
(1)
.
-
(2)
For , we have .
-
(3)
.
-
(4)
There is a component of such that .
Proof.
Let . It follows from Theorem 4.4(3) that . Since is -cooperative for , we have that , and the second assertion of the theorem holds. Since is a subgraph of , the third assertion follows immediately from the second.
We now prove the fourth assertion. Suppose there is no component of such that . Write
Since , and , and by the second assertion of the theorem, we have that .
(3) for every path in with ends .
Suppose there is a path in with ends such that . Since there is no component of such that , it follows that . Consequently, there is an and such that . Since is a path from to in , it follows that . Since is connected and separates from , we deduce that . But since has a neighbor in , it follows that , a contradiction. This proves (4).
(4) for every path of with ends .
Let be a path in with ends . We prove by induction on that . If , the claim follows from (4), and this is the base case.
Let and let such that .
Suppose first that belongs to the component of such that . Then . Since is connected and , it follows that
,
and so .
Thus we may assume that for some . Since is connected and , it follows that
.
Write where and .
Let be minimum and maximum such that
. Since ,
it follows that . By Theorem 4.4(2) there is a
path form to with .
Then
is a path from to in with
.
Inductively, .
But and ,
and so .
This proves (4).
Since is a -banana in , and since ,
(4) leads to a contradiction. This completes the proof.
∎
Recall that a separation of is a triple of pairwise disjoint subsets of with such that is anticomplete to . We are now ready to move on to star cutsets. We will work in the graph 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 that we are interested in.
As in other papers in the series, we associate a certain unique star separation to every vertex of . 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 . By Theorem 4.5, there is a component of such that . Since , it follows that is not complete to ; consequently , and so is unique. Let be the unique component of such that (this choice of is different from what we have done in earlier papers). Let , and finally, let . Then is the canonical star separation of corresponding to . 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 is not a hub of .
We need just a little more set up. Let be a linear order on . Following [4], we say that two vertices of are star twins if , , and . (Note that every two of these conditions imply the third.)
Let be a relation on defined as follows:
Note that if , then either , or
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 , then .
Proof.
Since and is anticomplete to , we have . Since is connected, there is a component of such that . Since , it follows that , and so . Let . Then has a neighbor in and thus in . If , then . If , then . It follows that . But now , 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.
is a partial order on .
In view of Lemma 4.8, let be the set of all -minimal elements of .
Lemma 4.9.
Let . Then .
We have finally reached our goal: we can define a subgraph of that is simpler than itself, but carries all the information we need. Define
The next theorem summarizes the important aspects of the behavior of .
Theorem 4.10.
The following hold:
-
(1)
For every , we have .
-
(2)
For every , .
-
(3)
For every , .
-
(4)
For every component of , there exists such that . Further, if is a component of and such that , then .
-
(5)
.
Proof.
Next we prove (3). By Lemma 4.5(1), it follows that . Observe that if and for some , then or . It follows that for , only if . Consequently, , and so (3) follows from Lemma 4.5(3).
We now explain how is used. In the course of the proof of Theorem 1.3, we will inductively obtain a small cutset separating from in . The next theorem allows us to transform this cutset into a cutset separating from in .
Theorem 4.11.
Let be a separation of such that and . Let . Then
-
(1)
separates from in .
-
(2)
.
-
(3)
.

Proof.
Suppose that does not separate from in . Let be a path from to in . Let be the components of such that and . Since the first vertex of is in and the last vertex of is in , it follows that there is a subpath of such that has ends and:
-
•
The vertex is in ;
-
•
;
-
•
The vertex is in .
See Figure 2. Since , it follows that is in a component of with (possibly ). This implies that , and so .
We finish this section with a theorem that allows us to transform a cutset separating from in into a cutset separating from in .
Theorem 4.12.
Assume that does not admit a clique cutset. Let be a separation of such that and . Let
Then separates from in .
Proof.
Suppose not. Let be the components of such that and . Since does not separate from in , and , there is a component of such that . Let be a path from to with .
As before, it follows that there is a subpath of such that has ends and:
-
•
The vertex is in ;
-
•
;
-
•
The vertex is in .
We first consider the case that (and so ). Then for some and . It follows that , and so . It follows that , and from Theorem 2.6, we know that . Therefore, , contrary to the fact that . This is a contradiction, and proves that .
Therefore, there is a component (possibly ) of such that . Since there are no edges between and , it follows that .
Consequently, there is a component of such that . Let be the component of such that . By Theorem 2.6, it follows that .
Suppose first that . Then , and , and . Since , it follows that , and so . Since is a path with ends and with , it follows that , and so . Similarly, .
We deduce that exactly one of has a neighbor in (unless ), and this neighbor is unique; denote it by . By symmetry, we may assume that , and so . By Theorem 2.5, we deduce that . Since , it follows that . Similarly, we have . Since is connected and by Theorem 4.4(2), it follows that contains a path with ends and interior in . Since is a path from to in , it follows that , and so . Therefore, , which implies that . It follows that , and from Theorem 2.6, we know that . Therefore, , contrary to the fact that . 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 . This generalizes results from [2] and [3].
What follows is mostly terminology from [3], but there are also some new notions. Let be a graph, let be a path in and let . We say that is an alignment if , , every vertex of has a neighbor in , and there exist such that for . We also say that is the order on given by the alignment . An alignment is wide if each of has at least two non-adjacent neighbors in , spiky if each of has a unique neighbor in and triangular if each of has exactly two neighbors in and they are adjacent. An alignment is consistent if it is wide, spiky or triangular.
By a caterpillar we mean a tree with maximum degree three such that there exists a path of such that all vertices of degree in belong to . We call a minimal such path the spine of . By a subdivided star we mean a graph isomorphic to a subdivision of the complete bipartite graph for some . 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 , a vertex of is said to be simplicial if is a clique. We denote by the set of all simplicial vertices of . Note that for every tree , is the set of all leaves of . An edge of a tree is said to be a leaf-edge of if is incident with a leaf of . It follows that if is the line graph of a tree , then is the set of all vertices in corresponding to leaf-edges of .
Let be a graph that is either a path, or a caterpillar, or the line graph of a caterpillar, or a subdivided star with root , or the line graph of a subdivided star with root . We define an induced subgraph of , denoted by , which we will use throughout the paper. If is a path, we let . If is a caterpillar, we let be the spine of . If is the line graph of a caterpillar , let be the path in consisting of the vertices of that correspond to the edges of the spine of . If is a subdivided star with root , let . It is the line graph of a subdivided star with root , let be the clique of consisting of the vertices of that correspond to the edges of incident with . The legs of are the components of .
Next, let be a caterpillar or the line graph of a caterpillar and let be a set of vertices disjoint from such that every vertex of has a unique neighbor in and . Let be the set of vertices of that have neighbors in . Then the neighbors of elements of appear in in order (there may be ties at the ends of , which we break arbitrarily); let be the corresponding order on . Now, order the vertices of as where has a neighbor in the leg of containing for . We say that is the order on given by .
Next, let be an induced subgraph of 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 be such that every vertex of has a unique neighbor in and . We say that is a consistent connectifier and it is spiky, triangular, stellar, or clique respectively. If is a single vertex and , we also call a stellar connectifer. If is a subdivided star, a singleton or the line graph of a subdivided star, we say that is a concentrated connectifier.
The following was proved in [3]:
Theorem 5.1 (Abrishami, Alecu, Chudnovsky, Hajebi, Spirkl [3]).
For every integer , there exists an integer with the following property. Let be a connected graph with no clique of cardinality . Let such that , is connected and every vertex of has a neighbor in . Then there is a set with and an induced subgraph of for which one of the following holds.
-
•
is a path and every vertex of has a neighbor in ; or
-
•
is a caterpillar, or the line graph of a caterpillar, or a subdivided star. Moreover, every vertex of has a unique neighbor in and .
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).

Theorem 5.2.
For every integer , there exists an integer with the following property. Let be a connected graph. Let such that , is connected and every vertex of has a neighbor in . Then there is a set with and an induced subgraph of for which one of the following holds.
-
•
is a path and every vertex of has a neighbor in ; or
-
•
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 has a unique neighbor in and .
Proof.
Let where is as in Theorem 5.1. Let be a minimal connected subset of such that every has a neighbor in .
(5) If contains a clique of size , then there exists and with such that is a clique connectifier.
Suppose that there is a clique of size in . It follows from the minimality of that for every one of the following holds:
-
•
There is such that is anticomplete to ; in this case set .
-
•
is not connected, and for every component of there is such that is anticomplete to . Since is a clique, there is a component of such that . Let be a vertex of that is anticomplete to ; write .
Let be the set of all vertices as above. For every , let be a path from to with . Write . We will show that and is a clique connectifier. It follows from the definition of that every vertex of has a neighbor in .
Next we claim that if , then is disjoint from and anticomplete to . Recall that , and is anticomplete to . It follows that , and so and . Let be the component of such that . By the definition of , we have that . Since is connected and , it follows that . Consequently, and are disjoint and anticomplete to each other. Since has no neighbor in , we deduce that is anticomplete to , and in particular . Similarly, is anticomplete to . This proves the claim.
Now we move to two anticomplete connected subgraphs and prove the following:
Theorem 5.3.
For every integer , there exists an integer with the following property. Let and assume that where
-
•
is a stable set with ,
-
•
and are components of , and
-
•
.
Then there exist with , and (possibly with the roles of and reversed) such that either:
-
(1)
Not both and are alignments, and
-
•
is a triangular connectifier or a clique connectifier or a triangular alignment; and
-
•
is a stellar connectifier, or a spiky connectifier, or a spiky alignment or a wide alignment.
or
-
•
-
(2)
Both and are alignments, and at least one of and is not a spiky alignment.
Moreover, if neither of is a concentrated connectifier, then the orders given on by and by 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 is a path and when the clique number of is bounded. However, it is easier to prove the more general result first using the symmetry between and , 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 be a sequence of distinct reals. Then there exists either an increasing or a decreasing -sub-sequence.
We start with two lemmas.
Lemma 5.5.
Let and assume that where is a stable set with and is anticomplete to . Suppose that for , the pair is a consistent alignment, or a consistent connectifier. Assume also that if neither of is concentrated, then the orders given on by and by are the same. Then (possibly switching the roles of and ), we have that either:
-
(1)
Not both and are alignments, and
-
•
is a triangular connectifier or a clique connectifier or a triangular alignment; and
-
•
is a stellar connectifier, or a spiky connectifier, or a spiky alignment or a wide alignment.
or
-
•
-
(2)
Both and are alignments, and at least one of and is not a spiky alignment, and at least one of and is not a triangular alignment.
Proof.
If at least one is not a concentrated connectifier, we let be the order given on by . If both of are concentrated connectifiers, we let be an arbitrary order on . Let be the unique hole contained in . For and , if is a connectifier, let be the leg of containing a neighbor of ; and if is an alignment let .
Suppose first that is a triangular alignment, a clique connectifier or a triangular connectifier. If is a triangular alignment, a clique connectifier or a triangular connectifier, then for every , the graph is either a prism or an even wheel with center , a contradiction. This proves that is a stellar connectifier, or a spiky connectifier, or a spiky alignment, or a wide alignment, as required.
Thus we may assume that for , the pair is a stellar connectifier, or a spiky connectifier, or a spiky alignment, or a wide alignment. If is a stellar connectifier or a spiky connectifier, then for every , the graph contains a theta, a contradiction.
It follows that for , the pair is an alignment. We may assume that and are either both spiky alignments or both triangular alignments, for otherwise the second outcome of the theorem holds. But now for every , the graph is either a theta or an even wheel, a contradiction. ∎

Lemma 5.6.
Let be a theta-free graph and let be a path in . Let be a connected subset of such that is anticomplete to . Let such that is not an edge. Assume that each of and has at least two non-adjacent neighbors in . Let be minimum and be maximum such that is adjacent to and and write . Suppose that has a neighbor in . Then has a neighbor in .
Proof.
Suppose that is anticomplete to . Let be a path from to with . (See Figure 4.) Let be minimum and be maximum such that is adjacent to . Then there there is a path from to with interior in , and a path from to with interior in . If , then is a theta with ends , a contradiction; therefore . Since has at least two non-adjacent neighbors in , it follows that has a neighbor in . We may assume that , and therefore there is a path from to with . Since is anticomplete to , we have that , and so is anticomplete to . But now is a theta with ends , a contradiction. ∎
We can now prove Theorem 5.3. The proof follows along the lines of [3], but the assumptions are different.
Proof.
Let and let , where is as in Theorem 5.2. Applying Theorem 5.2 twice, we obtain a set with and such that either
-
•
is a path and every vertex of has a neighbor in ; or
-
•
is a consistent connectifier
for every .
(6) Let and . If is a path and every vertex of has a neighbor in , then either some vertex of has neighbors in , or there exists with and a subpath of such that is a consistent alignment.
Let . We may assume that is chosen minimal satisfying Theorem 5.2, and so there exist such that for .
We may assume that for every . Let be the set of vertices in with exactly one neighbor in . Suppose that . It follows that contains a set with such that no two vertices in have a common neighbor in . We may assume that . Therefore, is a spiky alignment, as required. Thus we may assume that .
Next, let be the set of vertices in such that either or has exactly two neighbors in , and they are adjacent. Suppose that . By choosing greedily, it follows that contains a subset with the following specifications:
-
•
;
-
•
; and
-
•
no two vertices in have a common neighbor in .
But then is a triangular alignment, as required. Thus we may assume that .
Finally, let . It follows that . Let be a path from to with , and let be the hole . Let . Define to be the minimal subpath of containing . Let the ends of be and . Next let . Since is a path, it follows that for every . Since for every , we can greedily choose with , where and such that if , then is anticomplete to . It follows from Lemma 5.6 that for every , and so is a wide alignment. This proves (5).
(7) There exist with , and for such that is a consistent alignment or a consistent connectifier.
If both and are consistent connectifiers, (5) holds. Thus we may assume that is a path and every vertex of has a neighbor in . Suppose first that some has at least neighbors in . Let , and let with . If is a connectifier, we can clearly choose such that (5) holds. So we may assume that is a path and every vertex of has a neighbor in . Moreover, no vertex of has three or more neighbors in , for otherwise contains a theta with ends . Now by (5) applied with , there is a set with such that is a consistent alignment, and (5) holds. Similarly, we may assume that every has fewer than neighbors in .
Therefore, we may assume that every vertex has strictly fewer than neighbors in . Applying (5) with , we conclude that there exists with and a path such that is a consistent alignment. If is a consistent connectifier, then (5) holds, so we may assume that is a path and every vertex of has a neighbor in . Since every vertex has fewer than neighbors in , we apply (5) with again, and we conclude that there exists with and a path such that is a consistent alignment. This proves (5).
(8) There exist with , and for such that is a consistent alignment or a consistent connectifier. Moreover, if neither of is a concentrated connectifier, then the order given on by and is the same.
We now refine Theorem 5.3 in the two special cases we need here. The first one is when is a path. It is useful to state this result in terms of the following definition.
Given a graph class , we say is amiable if, for every integer , there exists an integer such that the following holds for every graph . Assume that where
-
•
is a stable set with ,
-
•
and are components of ,
-
•
,
-
•
is a path, and
-
•
for every there exists such that .
Then there exist with , and such that
-
(1)
One of the following holds:
-
•
is a triangular alignment and is a stellar connectifier or a spiky connectifier;
-
•
is a spiky alignment, and is a triangular connectifier or a clique connectifier;
-
•
is a wide alignment and is a triangular connectifier or a clique connectifier; or
-
•
Both and are alignments, and at least one of and is not a spiky alignment, and at least one of and is not a triangular alignment.
-
•
-
(2)
If is not a concentrated connectifier, then the orders given on by and by are the same.
-
(3)
For every except at most two, we have that .
Theorem 5.7.
The class is amiable.
Proof.
Let , and let be as in the definition of an amiable class. Let where is as in Theorem 5.3. We start with the following.
(9) For every , we have that .
Suppose there exists and .
We may assume that .
By reversing if necessary, we may assume that , and so .
Let be a path with ends and with interior in
. Let be a path from
to with interior in . Now there is a theta in
with ends and paths , and ,
a contradiction. This proves (5).
Now we apply Theorem 5.3 and obtain a set with
satisfying one of the outcomes of Theorem 5.3.
Let and
be as in Theorem 5.3 where
(and so Theorem 5.3 may hold with the roles of and reversed).
It follows from
(5) that every vertex in has degree at most 6 in and degree 2 in , and so is an alignment.
Let
be the order on given by .
(10) There exist at most two values of for which .
Suppose there exist three such values of . Since is a path,
there exist such that ,
and we may assume (by reversing if necessary) that for two values
, both and have neighbors
in . Let be a path from to with .
Since , there is a path from to with
. But now we get a theta with ends
and path , , and a path from to with interior in
, a contradiction. This proves (5).
By (5), there exists
with and such that for every
. We obtain that and
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 .
∎
The second special case is when the clique number of is bounded, and ; it is a result from [3].
Theorem 5.8 (Abrishami, Alecu, Chudnovsky, Hajebi, Spirkl [3]).
For every pair of integers , there exists an integer with the following property. Let and assume that where
-
•
is a stable set with ,
-
•
and are components of , and
-
•
.
Assume that . Then there exist with , and (possibly with the roles of and reversed) such that
-
•
is a triangular connectifier or a triangular alignment;
-
•
is a stellar connectifier, or a spiky connectifier, or a spiky alignment or a wide alignment; and
-
•
if is a triangular alignment, then is not a wide alignment.
Moreover, if neither of is a stellar connectifier, then the orders given on by and by 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 rounds (rather than just constantly many), which is what we need in the proof. We need the following definition and result. Given a graph , a -subdivision of is a graph that arises from by replacing each edge of by a path from to of length at least 1 and at most ; the interiors of these paths are pairwise disjoint and anticomplete.
Theorem 6.1 (Lozin and Razgon [29]).
For all positive integers and , there exists a positive integer such that every graph containing a ()-subdivision of as a subgraph contains either as a subgraph or a proper ()-subdivision of as an induced subgraph.
Since graphs in contain neither nor as an induced subgraph, we conclude that graphs in do not contain as a subgraph. Likewise, graphs in do not contain subdivisions of (in other words, thetas) as an induced subgraph. Therefore, letting , we conclude:
Lemma 6.2.
Let . Then there exists an such that the following holds: Let in . Then does not contain a -subdivision of as a subgraph.
Recall that the Ramsey number is the minimum integer such that every graph on at least vertices contains either a clique of size or a stable set of size .
Theorem 6.3.
For every integer there exists with the following property. Let and assume that where
-
•
and are components of .
-
•
is a stable set.
-
•
.
-
•
.
-
•
There exist and such that is a -banana in .
Then .
Proof.
Let be as in Theorem 3.2. We may assume that . Let as in Lemma 6.2. Let Let be as in Theorem 5.8. Set . Applying Theorem 5.8, we deduce that there exist , and with such that:
-
•
is a triangular connectifier or a triangular alignment;
-
•
is a stellar connectifier, or a spiky connectifier, or a spiky alignment, or a wide alignment; and
-
•
if is a triangular alignment, then is not a wide alignment.
(11) Suppose is a wide alignment, and let be the order on given by . Then for every , has exactly three neighbors in , and two of them are adjacent.
Let . Let be a path from to
with interior in . Then
is a hole. Since is a wide alignment, has two non-adjacent
neighbors in . Since is not a wheel in , and is non-adjacent to and , (6)
follows.
Next we show:
(12) Every vertex in has at most two neighbors in .
Suppose that there is a vertex such that has three distinct neighbors in ; without loss of generality, we may assume that . We consider two cases. First, if is a wide alignment, then contains a wheel where is a cycle consisting of and the path from to with interior in . By (6), has four neighbors in : , as well as three neighbors in . But is not a hub in , a contradiction.
It follows that is not a wide alignment. Now, for , let be a path from to with interior in . Then is a theta in (see Figure 5). This proves (6).

(13) There exists with , such that for every and for every path from to with interior in we have that .
Define a graph with vertex set such that if and only if there is a path of length two with ends and interior in . Since , there is with such that is either a clique or a stable set in .
If is a stable set, then
(6) holds, so we may assume that is a clique of size in . Write . For with , we let be a common neighbor of and in (which exists, since is a clique in ). By (6), no vertex has three neighbors in , and so the vertices are pairwise distinct. But now contains a -subdivision of as a subgraph, contrary to Lemma 6.2. This is proves (6).
From now on let be as in (6). Let
be the graph obtained from by adding a new vertex
with . We need the following two facts about :
(14) .
Suppose not, let and let be a wheel in . Since , it follows that . Let ; then is a path from to with . Since is a wheel, it follows that has two non-adjacent neighbors in . Consequently, there is a path from to and a path form to , both with interior in , such that is disjoint from and anticomplete to .
Let be a path from to with interior in . Then is a hole. Since , it follows that is not a wheel in , and so is anticomplete to .
Suppose first that is a wide alignment. Since is anticomplete to , it follows that (reversing the order on if necessary, and exchanging the roles of if necessary) appear in this order in the order given by on . Now we get a theta with ends and paths , , and , a contradiction. This proves that is not a wide alignment.
Since is not a wide alignment, there is a vertex , and three paths all with interior in , where is from to , is from to , and is from to , and the sets , and are pairwise disjoint and anticomplete to each other. Note that . Since is anticomplete to , it follows that is non-adjacent to . But now we get a theta with ends and path , , and , a contradiction. This proves (6).

(15) .
Suppose not. Let be such that is a , a , a theta, a prism or an even wheel. Then . Since is a stable set, it follows that is not a . Suppose first that ; then . Suppose that is a . Then there is such that , contrary to the fact that was chosen as in (6). It follows that is a theta, a prism or an even wheel. Let be a path from to with interior in . Now is an induced subgraph of . Moreover, if is a theta then is a theta, and if is a prism then is a prism. Since , it follows that is an even wheel; denote it by . Then . By (6), . It follows that , and so is an even wheel, contrary to the fact that .
This proves that . Since is a stable set, it follows that is not a prism. Suppose that is a theta. Then the ends of are and ; let the paths of be where . By renumbering , we may assume that appear in this order in the order on given by . Then for every , there is a unique path from to with interior in . Similarly, for all there is a unique path from to with interior in . Observe that contains a unique pyramid with the following specifications:
-
•
The paths of are , where .
-
•
Denote the apex of by , and let the base of be where is an end of . Then either , or is a wide alignment and .
-
•
, and .
Let . If is non-adjacent to , then is a theta with ends and paths , contrary to the fact that . This proves that is adjacent to . Consequently, and is a wide alignment. But now is a hole, and, since is adjacent to , is a wheel, contrary to the fact that . This proves that is not a theta.
Consequently, is an even wheel . Since , it follows that either , or is adjacent to . If is adjacent to , then , contrary to (6). It now follows that , and so is a hole in , where . Let where appear in this order in the order given on by . If is a spiky alignment or a wide alignment, or a spiky connectifier, then we get a theta with ends two of whose paths are contained in , and the third is the path from to with interior in . It follows that is a stellar connectifier. Therefore there exist paths , all with a common end , such that is from to , , and the sets are pairwise disjoint and anticomplete to each other. Since is a not an even wheel in , we deduce that is not complete to ; let be non-adjacent to . Since , there exist distinct and paths and of such that
-
•
is from to ;
-
•
is from to ;
-
•
; and
-
•
.
Now
we get a theta with ends and path ,
and .
This proves (6).
Observe that is a -banana in . But by (6),
, and by (6), , 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 , and let be (theta, )-free with . There exist an integer and a partition of with the following properties:
-
(1)
.
-
(2)
is a stable set for every .
-
(3)
For every and we have .
Let be a graph and let . A hub-partition with respect to of is a partition of as in Theorem 7.1; we call the order of the partition. We call the hub-dimension of (denoting it by ) the smallest such that has a hub-partition of order with respect to .
For the remainder of this section, let us fix . Let be as in Theorem 7.1. Let where is as in Theorem 3.2. Let . Let where is as in Theorem 6.3.
Since in view of Theorem 7.1, we have for all , Theorem 1.3 follows immediately from the next result:
Theorem 7.2.
Let and with and let . Then is not a -banana in .
Proof.
Suppose that is a -banana in . We will get a contradiction by induction on , and for fixed by induction on . Suppose that . Then and by Theorem 3.3, we have that there is no -banana in . Now the statement holds since . Thus we may assume .
(16) We may assume that admits no clique cutset.
Suppose is a clique cutset in . Since is -free, it
follows that there is a component of such that
is -banana in .
But and , consequently we get
contradiction
(inductively on ). This proves (7).
Let be a hub-partition of with respect to
and with .
We now use notation and terminology from Section 4 (note that the definitions of and agree; we use , , , as defined above).
It follows from the definition of that every vertex in
is -safe.
Let be an -atomic
tree decomposition of . Let be as in Theorem 2.6.
and let and be as in Section 4. Then by Lemma 4.1, we have .
By Theorem 4.10(5), we have that and
is a hub-partition of
with respect to .
It follows that .
Inductively (on ), we have that
is not a -banana in .
Our first goal is to prove:
(17) There is a set such that
-
(1)
separates from in ;
-
(2)
; and
-
(3)
.
Since is not a -banana in , Theorem 2.1 implies that there is a separation of such that and and . We may assume that is chosen with as small as possible. Let be the component of such that , and and let be the component of such that . It follows from the minimality of that and is a -banana in .
We claim that . Let . Again by the minimality of , it follows that is a -banana in . If is not a -banana in , then and the claim follows immediately since ; thus we may assume that is a -banana in . Now since is a stable set and since no vertex of is a hub in , applying Theorem 6.3 to we deduce that , as required. This proves the claim.
Next we claim that , and . Write . Observe that for distinct , say with and for , the sets and are disjoint and anticomplete to each other. Let be a subset of such that for all and , we have . Then is stable set and . Let . If follows from the minimality of that is a -banana in . If is not a -banana in , then and the claim follows immediately since ; thus we may assume that is a -banana in . By Theorem 4.4(3), . Now since is a stable set, applying Theorem 6.3 to we deduce that , as required. Since , it follows that . This proves the claim.
Now let be as in Theorem 4.11. Then separates from in . Moreover, by Theorem 4.11,
Since , it follows that
Finally, by Theorem 4.11,
. Since and ,
it follows that .
This proves (7).
Let be as in (7), and let
be a separation of such that
and .
Let be obtained from as in Theorem 4.12.
Then separates from in .
Recall that by (7), we have .
Now since and since ,
it follows from Theorem 4.12 that
Since by (7), we deduce that 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 with the following property. Let and let be a normal weight function on . Then there exists such that
-
•
, and
-
•
is a -balanced separator in .
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 be a -free odd-signable graph that contains a proper wheel that is not a universal wheel. Let and be the endpoints of a long sector of . Let be the set of all vertices in such that the subpath of from to contains an even number of neighbors of , and let . Let . Then, is a cutset of that separates from .
The second is a similar kind of theorem, but we start with a pyramid rather than a wheel.
Theorem 8.3.
Let . Let be a pyramid with paths , apex , and base in and let be distinct. For , let be the neighbor of in . Let be a path from a vertex of to a vertex of . Then at least one one of has a neighbor in .
Proof.
It follows from the defintion of that there exist distinct and a subpath of such that has a neighbor in , and has a neighbor in . Let be chosen minimal with this property. It follows that , and so we may assume that is anticomplete to . We may also assume and . See Figure 7. It follows that and , .

(18) is anticomplete to .
Suppose not; then some vertex in has a neighbor in . It follows from the minimality of that . Then has a neighbor in each of the paths and . Now we get a theta with ends whose paths are subpaths of , a contradiction. This proves (8).
(19) If and is adjacent to , then has a neighbor in .
Suppose that is adjacent to and anticomplete to . Since is not a in , and by (8), it follows that is anticomplete to . If has a neighbor in , then we get a theta with ends , two of whose paths are contained in the hole , and third one has interior on , a contradiction. Consequently, is anticomplete to . Next suppose that has a neighbor in . Since is not a in , it follows that is not adjacent to . Now we get a theta with ends and path , a path with interior in and a path with interior in . This proves that is anticomplete to .
Let be the neighbor of in . Suppose that has a neighbor in . Then there is a path from to with and we get a theta with ends and paths and , and a path with interior in , a contradiction. It follows that is anticomplete to and therefore is adjacent to . Since is anticomplete to we deduce that . Now we get a theta with ends and paths , and , a contradiction. This proves (8).
(20) Either or .
Suppose not. Then . If both and have a neighbor in , then there is a theta in with ends , and paths and , a contradiction. By symmetry we may assume that is anticomplete to ; now by (8), it follows that is adjacent to both and . Since is not a , we deduce that is non-adjacent to . Similarly is non-adjacent to . It follows that has a neighbor in . Let be a path from to with interior in . Now we get a theta with ends and path , and a path with interior in , a contradiction. This proves (8).
(21) If has a neighbor in , then is anticomplete to .
Suppose that has a neighbor in and has a neighbor in . Then . If both and have a neighbor in , then there is a theta in with ends , and paths and , a contradiction. This proves that .
Let be a path from to with . Since is not a , it follows that is non-adjacent to . But now we get a theta with ends and paths , and a path from to with interior in , a contradiction. This proves (8).
(22) and .
Suppose has a neighbor in . Then . By (8), is anticomplete to , and by (8), is anticomplete to . Let be the hole , and let be the hole ( is a hole since has a neighbor in ). Then . Let be the neighbor of and let be the path . Observe that , and . Since neither of and is an even wheel, it follows that has exactly two neighbors in and they are adjacent. Let be such that . Since has a neighbor in , it follows that . Now we get a prism with triangles and and paths , and , a contradiction. This proves (8).
(23) is anticomplete to .
Suppose not. By (8), has a neighbor in . Then . Let be the hole and let be the hole ( is a hole since is anticomplete to ). Since , we have that . Since since neither of , is an even wheel, it follows that there exists such that . Now we get a prism with triangles and and paths , and , a contradiction. This proves (8).
(24) has at least two neighbors in , and has at least two neighbors in .
By symmetry it is enough to prove the first assertion of (8). Suppose that has a unique neighbor in . It follows that . By (8), is anticomplete to . Now we get a theta with ends and paths , and , a contradiction. This proves (8).
(25) has two non-adjacent neighbors in , and has two non-adjacent two neighbors in .
By symmetry it is enough to prove the first statement.
By (8), we may assume that has exactly two neighbors
in , and is adjacent to . We may assume that traverses
in this order.
By (8), is anticomplete to .
Let be the path
from to with interior in . It follows from the defintion of
that is anticomplete to . Now we get a prism with triangles
and and paths ,
and , a contradiction. This proves (8).
By (8), there exist paths from to and
from to , both with interior in ,
and such that is anticomplete to
. Let be a path from to
with interior in .
By (8), is anticomplete to
. Now we get a theta with ends
and paths ,
and , a contradiction.
∎
We need a technical definition. Given an integer and a graph class , we say is -amicable if is amiable, and the following holds for every graph . Let be as in the definition of an amiable class for and let such that and satisfy the first five bullets from the definition of an amiable class with . Let , and be as in the definition of an amiable class with , and let such that:
-
•
is the order given on by ; and
-
•
For every , we have .
Let be maximum such that is adjacent to , and let be minimum such that is adjacent to . Then there exists a subset with such that separates from . It follows that separates from .
We deduce:
Theorem 8.4.
The class is -amicable.
Proof.
From Theorem 5.7, we know that is amiable. With the notation as in the definition of an amicable class, we need to show that there exists a subset with such that separates from . Let be the path from to with interior in .
Let be the hole . Let be the path in such that and has a neighbor in . Thus if is an alignment then , and if is a connectifier then is the leg of containing a neighbor of . See Figure 8. (Note that may be empty if is a clique connectifier or a stellar connectifier.) We consider different possibilities for the behavior of and .

(26) If is a wide alignment, then the theorem holds.
Suppose first that is an alignment. Then is a proper wheel, and we are done by Theorem 8.2 and setting . Thus we may assume that is a connectifier. It now follows from Theorem 5.7 that is a triangular connectifier or a clique connectifier. Let . Then contains a pyramid with apex and base . Since is anticomplete to , it follows that belongs to the interior of the path of with ends , and belongs to the interior of the path of with ends (switching the roles of if necessary), and that are not adjacent to the apex of . Since , Theorem 8.3 implies that separates from . Since , we are done. This proves (8).
(27) If is a wide alignment, then the theorem holds.
Since is an alignment, we deduce that is a proper wheel. Now by Theorem 8.2 and setting , we are done. This proves (8).
(28) If is a triangular alignment, then the theorem holds.
Let be such that
.
Since has a neighbor in the path , it
follows that . Since has a neighbor in the path , it follows that .
By Theorem 5.7 and in view of (8),
we may assume that is a spiky alignment, a stellar connectifier
or a spiky connectifier.
In all cases, has a unique neighbor in ; denote it by (where is as defined in Section 5).
Now is a pyramid with apex and base .
Since and since is non-adjacent to , it
follows that belongs to the interior of
the path of with ends , and belongs to the interior of
the path of with ends , and are not adjacent to the apex of .
Since , Theorem 8.3 implies
that separates from . Since
and , we are done. This proves (8).
By Theorem 5.7 and in view of (8) and (8), we deduce that is a spiky alignment. Let
be the unique neighbor of in .
By Theorem 5.7 and in view of (8), it follows that is either
a triangular alignment or a triangular connectifier or a clique connectifier. Let be the
neighbors of in . Now is a pyramid with apex and base .
Since has a neighbor in the path ,
it follows from the definition of a spiky alignment that .
Similarly, .
Since , we deduce that
belongs to the interior of
the path of with ends , and belongs to the interior of
the path of with ends (possibly switching the roles of ), and are non-adjacent to the apex of since . Therefore, Theorem 8.3 implies
that separates from . Since
and , we conclude that is 4-amicable.
We now prove that separates from . Let be the component of such that , and let be the component of such that . Then . Since , it follows that is disjoint from and . Therefore, and as required. ∎
We need the following lemma:
Lemma 8.5 (Chudnovsky, Pilipczuk, Pilipczuk, Thomassé [14], Lemma 5.3).
Let be a connected graph with a weight function . Then there is an induced path in such that is a -balanced separator.
For a graph and a set we denote by the minimum size of a set such that . 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 and every -amicable graph class , there is an integer with the following property. Let and let be a normal weight function on . Then there exists such that
-
•
, and
-
•
is a -balanced separator in .
Proof.
It suffices to consider the unique component of with weight greater than ; if there is no such component, we set . Therefore, we may assume that is connected.
Since is amicable, it is also amiable. Let where is as in the definition of an amiable class for . Let . We claim that satisfies the conclusion of the theorem.
By Lemma 8.5, there is a path in such that for every component of . Let be such a path chosen with as small as possible. It follows from the minimality of that there is a component of with . Let .
(29) Let be a connected subset of such that . Then .
Since and is connected,
it follows that either
or . If , then
. Thus we may assume that
. Since , we deduce that
is contained in a component of , and
so . This proves (8).
Let be minimum such that .
(30) We may assume that .
Suppose . Then
. Let be such that
and with .
Let .
Then every component of is disjoint from
and (8) implies that
satisfies the conclusion of the theorem. This proves (8).
Let be such that
and with .
Let .
For every , let be the minimum
such that .
Let be a set of size at most such that
.
Our next goal is, for every , to define two sets and . To that end, we fix . Let . Let be such that , and assume that is chosen with as small as possible. Order the vertices of as in the order that they appear in when is traversed starting from . It follows from the minimality of that there is such that . Let and let and . So far we have defined and . Suppose that we have defined sequences of subsets , , , and and a sequence of integers such that and with the following properties (for every ):
-
•
is a stable set.
-
•
If and , then is anticomplete to .
-
•
for every .
-
•
If , then is anticomplete to , and .
-
•
For every , .

See Figure 9. Now we either terminate the construction or construct the sets , , , and define , and verify that the properties above continue to hold. Let . If , let and ; the construction terminates here. Now suppose that . Let be a minimal subset of such that . It follows from the second bullet and the minimality of that if and , then is anticomplete to . Since and , we deduce that if , then . Let be minimum such that and set . Then , and the third bullet continues to hold. By the minimality of , the second bullet continues to hold. It follows from the minimality of that there exists such that . Set and . Since is anticomplete to , the first bullet continues to hold. Since , and since , 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 and , and we have:
(31) The following statements hold.
-
•
is a stable set.
-
•
For every , .
-
•
If , then .
-
•
.
To simplify notation, we renumber the elements of , replacing each index by . In the new notation we have , and for every , . Observe that appear in this order in when is traversed from to . Let , and .
(32) For every , there exists and with such that separates from .
By (8), is a stable set.
Since ,
and , it follows from (8) that .
Let with . Recall that .
Now (8)
follows from the definition of an -amicable class with
, and . This proves (8).
We may assume that for every there is a component
of with , for
otherwise the conclusion of Theorem 8.1 holds
setting . By (8), either is anticomplete
to , or is anticomplete to .
Let us say that is of type if is anticomplete to
, and that is of type if is anticomplete to
. By (8), every is of at least one type.
(33) is of type .
By (8), . It follows that contains a vertex such that or . Since , and since is a subpath of , it follows that contains a vertex of . Consequently, is not of type , and so is of type . This proves (8).
(34) We may assume that is of type .
Suppose that is of type . Then
is anticomplete to . We claim that the set
satisfies the conclusion of the theorem.
Let be a component of , and suppose
that .
By (8) there exists .
Since ,
we have that . It follows that is anticomplete
to , and so . Since
,
it follows that .
We deduce that .
But then , contrary to the fact that .
This proves the claim, and (8) follows.
From now on we make the assumption of (8).
By (8) we can choose minimum such that is
of type . By (8), . Let
Then . We claim that for every component of . Suppose not, and let be a component of with . Since , we have that . Since is of type , it follows that is anticomplete to . Since , we have that . Since is of type , it follows that is anticomplete to . Since and since , we deduce that . But then 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.
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 be an integer, let be a graph and let be a weight function on . Assume that there is no -banana in . There exists a clique in with the following property. Let be such that for every component of . Then there exists a balanced separator in such that .
Proof.
Let be a -atomic tree decomposition of . By Theorems 2.2 and 2.3, we have that is tight and -lean. By Theorem 2.7, there exists such that is a center for . A vertex is -friendly if is not separated from by a set of size . We show that -friendly vertices have the following important property.
(35) If are not -friendly, then is adjacent to .
Suppose that and are not -friendly and that
is non-adjacent to . Since there is no -banana in ,
Theorem 2.1 implies that there exists a
set with such separates from . But this contradicts
Theorem 2.4. This proves (9).
Let be the set of all the vertices of that are not -friendly.
By (9), is a clique.
We show that satisfies the conclusion of Theorem 9.2.
Let be such that
for
every component
of .
For every , define the set as follows. Let be such that separates from , chosen with minimum. Let . Then and . Let
It follows that .
To complete the proof it is enough to show that is a -balanced separator of . Suppose not, and let be a component of with .
(36) If for some , then is anticomplete to .
Suppose that has a neighbor in . Let be a separation such that and . Since and is adjacent to , it follows that . But then, since is a component of and , it follows that . Consequently, . It follows that , and so , a contradiction. This proves (9).
(37) .
We now prove Theorem 9.1.
Proof.
Let be as in Theorem 9.2. Let be as in Theorem 1.3 and let . Let be as Theorem 8.1. Define as . Then is a normal weight function on . By Theorem 8.1, there exists such that
-
•
, and
-
•
is a -balanced separator in .
It follows that for every component of . Now Theorem 9.2 applied with implies that there exists a -balanced separator in such that . Since , it follows that . ∎
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 -balanced separators in an earlier paper of this series [4].
Theorem 10.1 (Robertson, Seymour [33], see also [4, 24]).
Let be a graph, let , and let be a positive integer. If has a -balanced separator of size at most for every normal weight function , then .
We prove the following, which immediately implies Theorem 1.5.
Theorem 10.2.
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 are the most immediate. In particular, using the factor approximation algorithm of Korhonen [26] (or the simpler factor approximation algorithm of Robertson and Seymour [34]) for treewidth, we can compute a tree decomposition of width at most in time . A number of well-studied graph problems, including Stable Set, Vertex Cover, Feedback Vertex Set Dominating Set and -Coloring (for fixed ) admit algorithms with running time when a tree decomposition of of width is provided as input (See, for example, [16] chapters 7 and 11, as well as [10]). Since this immediately leads to polynomial time algorithms for these problems in .
Theorem 11.1.
Let be fixed and P be a problem which admits an algorithm running in time on graphs with a given tree decomposition of width at most . Then P is solvable in time in . In particular, Stable Set, Vertex Cover, Feedback Vertex Set, Dominating Set and -Coloring (with fixed ) are all polynomial-time solvable in .
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 -Coloring (with fixed ) in (without any assumptions on max clique size) because every graph that contains a clique of size can not be -colored. Thus, to solve -Coloring we can first check for the existence of an -clique in time . If an -clique is present, report that no -coloring exists, otherwise invoke the algorithm of Theorem 11.1 with . This yields the following result.
Theorem 11.2.
For every positive integer , -Coloring is polynomial-time solvable in .
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 there exists a number such that all graphs in have chromatic number at most . Also, for each fixed , by Theorem 11.1, -Coloring is polynomial-time solvable in . Now by solving -Coloring for every , we also obtain a polynomial-time algorithm for Coloring in .
Theorem 11.1 quite directly leads to a polynomial-time approximation scheme (PTAS) for several problems on graphs in . We illustrate this using Vertex Cover as an example.
Theorem 11.3.
There exists an algorithm that takes as input a graph and , runs in time and outputs a vertex cover of size at most , where is the size of the minimum vertex cover of .
Proof.
Check in time whether contains as input a clique of size at least . If does not have a clique of size at least then compute an optimal vertex cover of in time using the algorithm of Theorem 11.1. If contains such a clique then run the algorithm recursively on to obtain a vertex cover of of size at most . Return ; clearly is a vertex cover of . Furthermore, every vertex cover of (and in particular a minimum one) contains at least vertices of . It follows that and that therefore
Here the last inequality follows because for it holds that . ∎
The PTAS of Theorem 11.3 generalizes without modification (except for the constant in the ) 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 on graphs of treewidth . 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 . On the other hand, a QPTAS for Stable Set in follows from Theorem 1.4 together with an argument of Chudnovsky, Pilipczuk, Pilipczuk and Thomassé [14], who gave a QPTAS for Stable Set in -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 and , runs in time and outputs a stable set in of size at least , where is the size of the maximum size stable set in .
Proof.
The algorithm is recursive, taking as input the graph and also an integer . Initially, the algorithm is called with , in all subsequent recursive calls the value of remains the same. If is a single vertex the algorithm returns . If 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 be the constant of Theorem 1.4. If is a connected graph on at least two vertices the algorithm iterates over all subsets of size at most and all subsets of size at most such that each connected component of has at most vertices. For each such pair the algorithm calls itself recursively on 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 which results in a recursive call of the algorithm, each connected component of the graph that the algorithm is called on has at most vertices. Therefore the running time of the algorithm is governed by the recurrence which solves to . Since the algorithm is initially called with the running time of the algorithm is upper bounded by .
It remains to argue that the size of the stable set output by the algorithm is at least . To that end, we show that the size of the independent set output by a recursive call on is at least . Let be a maximum size stable set of . By a standard coupon-collector argument (see e.g. Lemma 4.1 in [14]) there exists a choice of of size at most such that no vertex of has at least neighbors in . Let now be the set (of size at most ) given by Lemma 1.4 applied to . Since no vertex in has at least neighbors in it follows that . Furthermore, each connected component of has at most vertices. By the inductive hypothesis the recursive call on outputs a stable set of size at least
Since the recursive call on outputs a stable set of size at least , 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. “-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 -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 -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 -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, )-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.