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

Optimally Reconfiguring
List and Correspondence Colourings

Stijn Cambie Department of Mathematics, Radboud University Nijmegen, Netherlands and Mathematics Institute, University of Warwick, UK. E-mail: [email protected], supported by a Vidi grant (639.032.614) of the Netherlands Organisation for Scientific Research (NWO) and the UK Research and Innovation Future Leaders Fellowship MR/S016325/1. Current affiliation: Department of Computer Science, KU Leuven Campus Kulak-Kortrijk, 8500 Kortrijk, Belgium.    Wouter Cames van Batenburg Delft University of Technology, Delft Institute of Applied Mathematics; [email protected]    Daniel W. Cranston Virginia Commonwealth University, Department of Computer Science; [email protected]
Abstract

The reconfiguration graph 𝒞k(G)\mathcal{C}_{k}(G) for the kk-colourings of a graph GG has a vertex for each proper kk-colouring of GG, and two vertices of 𝒞k(G)\mathcal{C}_{k}(G) are adjacent precisely when those kk-colourings differ on a single vertex of GG. Much work has focused on bounding the maximum value of diam𝒞k(G)\operatorname{diam}\mathcal{C}_{k}(G) over all nn-vertex graphs GG. We consider the analogous problems for list colourings and for correspondence colourings. We conjecture that if LL is a list-assignment for a graph GG with |L(v)|d(v)+2|L(v)|\geq d(v)+2 for all vV(G)v\in V(G), then diam𝒞L(G)n(G)+μ(G)\operatorname{diam}\mathcal{C}_{L}(G)\leq n(G)+\mu(G). We also conjecture that if (L,H)(L,H) is a correspondence cover for a graph GG with |L(v)|d(v)+2|L(v)|\geq d(v)+2 for all vV(G)v\in V(G), then diam𝒞(L,H)(G)n(G)+τ(G)\operatorname{diam}\mathcal{C}_{(L,H)}(G)\leq n(G)+\tau(G). (Here μ(G)\mu(G) and τ(G)\tau(G) denote the matching number and vertex cover number of GG.) For every graph GG, we give constructions showing that both conjectures are best possible, which also hints towards an exact form of Cereceda’s Conjecture for regular graphs. Our first main result proves the upper bounds (for the list and correspondence versions, respectively) diam𝒞L(G)n(G)+2μ(G)\operatorname{diam}\mathcal{C}_{L}(G)\leq n(G)+2\mu(G) and diam𝒞(L,H)(G)n(G)+2τ(G)\operatorname{diam}\mathcal{C}_{(L,H)}(G)\leq n(G)+2\tau(G). Our second main result proves that both conjectured bounds hold, whenever all vv satisfy |L(v)|2d(v)+1|L(v)|\geq 2d(v)+1. We conclude by proving one or both conjectures for various classes of graphs such as complete bipartite graphs, subcubic graphs, cactuses, and graphs with bounded maximum average degree.

1 Introduction

In this paper we study questions of transforming one proper colouring of a graph GG into another, by a sequence of recolouring steps. Each step recolours a single vertex, and we require that each resulting intermediate colouring is also proper. Our work fits into the broader context of reconfiguration, in which some object (in our case a proper colouring) is transformed into another object of the same type, via a sequence of small changes, and we require that after each change we again have an object of the prescribed type (for us a proper colouring). It is natural to pose reconfiguration questions for a wide range of objects: proper colourings, independent sets, dominating sets, maximum matchings, spanning trees, and solutions to 3-SAT, to name a few. For each type of object, we must define allowable changes between successive objects in the sequence (in our case, recolouring a single vertex). Typically, we ask four types of questions. (1) Given objects α\alpha and β\beta, is there a reconfiguration sequence from α\alpha to β\beta? (2) If the first question is answered yes, what is the length of a shortest such sequence? (3) Is the first question answered yes for every pair of objects α\alpha and β\beta? (4) If yes, what is the maximum value of dist(α,β)\operatorname{dist}(\alpha,\beta) over all α\alpha and β?\beta? For an introduction to reconfiguration, we recommend surveys by van den Heuvel [16] and Nishimura [15]. Most of our definitions and notation are standard, but for completeness we include many of them at the start of Section 2.

For a graph GG and a positive integer kk, the kk-colouring reconfiguration graph of GG, denoted 𝒞k(G)\mathcal{C}_{k}(G), has as its vertices all proper kk-colourings of GG and two vertices of 𝒞k(G)\mathcal{C}_{k}(G) are adjacent if their corresponding colourings differ on exactly one vertex of GG. Our goal here is to study the diameter of this reconfiguration graph, denoted diam𝒞k(G)\operatorname{diam}\mathcal{C}_{k}(G). In general, 𝒞k(G)\mathcal{C}_{k}(G) may be disconnected, in which case its diameter is infinite. For example 𝒞2(C2s)\mathcal{C}_{2}(C_{2s}) consists of two isolated vertices (here C2sC_{2s} is a cycle of length 2s2s). More strongly, for each integer k2k\geq 2 there exist kk-regular graphs GG such that 𝒞k+1(G)\mathcal{C}_{k+1}(G) has isolated vertices. The simplest example is the clique Kk+1K_{k+1}, but this is also true for every kk-regular graph with a (k+1)(k+1)-colouring α\alpha such that all k+1k+1 colours appear on each closed neighbourhood; such colourings are called frozen.

To avoid a colouring α\alpha being frozen, some vertex vv must have some colour unused by α\alpha on its closed neighbourhood, N[v]N[v]. And to avoid 𝒞k(G)\mathcal{C}_{k}(G) being disconnected, every induced subgraph HH must contain such a vertex vv with some colour unused by α\alpha on N[v]V(H)N[v]\cap V(H). Thus, it is natural to consider the degeneracy of GG, denoted degen(G)\operatorname{degen}(G).

Our examples of frozen colourings above show that, if we aim to have 𝒞k(G)\mathcal{C}_{k}(G) connected, then in general it is not enough to require kdegen(G)+1k\geq\operatorname{degen}(G)+1. However, an easy inductive argument shows that a slightly stronger condition is sufficient: If kdegen(G)+2k\geq\operatorname{degen}(G)+2, then 𝒞k(G)\mathcal{C}_{k}(G) is connected. (Earlier, Jerrum [12] proved that kΔ(G)+2k\geq\Delta(G)+2 suffices, but the arguments are similar.) This inductive proof only yields that diam𝒞k(G)2|V(G)|\operatorname{diam}\mathcal{C}_{k}(G)\leq 2^{|V(G)|}. But Cereceda [7, Conjecture 5.21] conjectured something much stronger.

Cereceda’s Conjecture.

For an nn-vertex graph GG with kdegen(G)+2k\geq\operatorname{degen}(G)+2, the diameter of 𝒞k(G)\mathcal{C}_{k}(G) is O(n2)O(n^{2}).

Bousquet and Heinrich [6] proved that diam𝒞k(G)=Od(ndegen(G)+1)\operatorname{diam}\mathcal{C}_{k}(G)=O_{d}(n^{\operatorname{degen}(G)+1}), which is the current best known bound. When kΔ(G)+2k\geq\Delta(G)+2, Cereceda [7, Proposition 5.23] proved that diam𝒞k(G)=O(nΔ)=O(n2)\operatorname{diam}\mathcal{C}_{k}(G)=O(n\Delta)=O(n^{2}). In particular, Cereceda’s conjecture is true for regular graphs. But, as we show here, if kΔ(G)+2k\geq\Delta(G)+2, then in fact we have the stronger bound diam𝒞k(G)2n\operatorname{diam}\mathcal{C}_{k}(G)\leq 2n, and we conjecture diam𝒞k(G)3n/2\operatorname{diam}\mathcal{C}_{k}(G)\leq\lfloor 3n/2\rfloor.

A folklorish result observes that diam𝒞k(G)=O(n)\operatorname{diam}\mathcal{C}_{k}(G)=O(n) when kk is large relative to Δ(G)\Delta(G) (for example, see [14, Section 3.1]). But until now, it seems that no one has investigated exact values of the diameter (say, when kΔ(G)k\geq\Delta(G)). This is the main goal of our paper. Before stating our two main conjectures, and our results supporting them, we present an easy lower bound on 𝒞k(G)\mathcal{C}_{k}(G) in terms of the matching number μ(G)\mu(G).

Proposition 1.

For a graph GG, if k2Δ(G)k\geq 2\Delta(G), then diam𝒞k(G)n(G)+μ(G)\operatorname{diam}\mathcal{C}_{k}(G)\geq n(G)+\mu(G).

Proof.

Let MM be a maximum matching in GG. Form G^\widehat{G} from GG as follows. If vw,xyMvw,xy\in M and wxE(G)wx\in E(G), then add vyvy to G^\widehat{G}. Note that Δ(G^)2Δ(G)1\Delta(\widehat{G})\leq 2\Delta(G)-1. So χ(G^)2Δ(G)\chi(\widehat{G})\leq 2\Delta(G). Let α\alpha be a 2Δ(G)2\Delta(G)-colouring of G^\widehat{G}. Form β\beta from α\alpha by swapping colours on endpoints of each edge in MM and, for each vV(G)v\in V(G) not saturated by MM, picking β(v)\beta(v) outside of {α(v)}wN(v)β(w)\{\alpha(v)\}\cup\bigcup_{w\in N(v)}\beta(w). To recolour GG from α\alpha to β\beta, every vertex must be recoloured. Further, for each edge eMe\in M, the first endpoint vv of ee to be recoloured must initially receive a colour other than β(v)\beta(v), so must be recoloured at least twice. Thus, diam𝒞k(G)n(G)+μ(G)\operatorname{diam}\mathcal{C}_{k}(G)\geq n(G)+\mu(G), as desired. ∎

At the end of Section 2, we mention various special cases in which the hypothesis of Proposition 1 can be weakened. We pose it as an open question whether the bound in Proposition 1 holds for all kΔ(G)+2k\geq\Delta(G)+2.

We study this reconfiguration problem in the more general contexts of list colouring and correspondence colouring, both of which we define in Section 2. These more general contexts offer the added advantage of enabling us to naturally prescribe fewer allowed colours for vertices of lower degree. Analogous to 𝒞k(G)\mathcal{C}_{k}(G), for a graph GG and a list-assignment LL or correspondence cover (L,H)(L,H) for GG, we define the LL-reconfiguration graph 𝒞L(G)\mathcal{C}_{L}(G) or (L,H)(L,H)-reconfiguration graph 𝒞(L,H)(G)\mathcal{C}_{(L,H)}(G) of GG. Generalizing our constructions of frozen kk-colourings above, it is easy to show that 𝒞L(G)\mathcal{C}_{L}(G) and 𝒞(L,H)(G)\mathcal{C}_{(L,H)}(G) can contain frozen LL-colourings (and thus be disconnected) if we require only that all vV(G)v\in V(G) satisfy |L(v)|=d(v)+1\lvert L(v)\rvert=d(v)+1. Thus, we adopt a slightly stronger hypothesis: all vV(G)v\in V(G) satisfy |L(v)|d(v)+2\lvert L(v)\rvert\geq d(v)+2. Now we can state our two main conjectures.

Conjecture 1 (List Colouring Reconfiguration Conjecture).

For a graph GG, if LL is a list-assignment such that |L(v)|d(v)+2\lvert L(v)\rvert\geq d(v)+2 for every vV(G)v\in V(G), then diam𝒞L(G)n(G)+μ(G)\operatorname{diam}\mathcal{C}_{L}(G)\leq n(G)+\mu(G).

Conjecture 2 (Correspondence Colouring Reconfiguration Conjecture).

For a graph GG, if (L,H)(L,H) is a correspondence cover such that |L(v)|d(v)+2\lvert L(v)\rvert\geq d(v)+2 for every vV(G)v\in V(G), then diam𝒞(L,H)(G)n(G)+τ(G)\operatorname{diam}\mathcal{C}_{(L,H)}(G)\leq n(G)+\tau(G).

Here τ(G)\tau(G) denotes the vertex cover number of GG. For brevity, we often call these conjectures the List Conjecture and the Correspondence Conjecture. Our aim in this paper is to provide significant evidence for both conjectures. Due to Proposition 1, the List Conjecture is best possible, when the lists are large enough. We will soon give easy constructions showing that both conjectures are best possible whenever all vv satisfy |L(v)|d(v)+2|L(v)|\geq d(v)+2. But we defer these constructions until Section 2, where we formally define list and correspondence colourings.

Proposition 2.

For every graph GG (i) there exists a list-assignment LL such that |L(v)|=d(v)+2|L(v)|=d(v)+2 for all vv for which diam𝒞L(G)=n(G)+μ(G)\operatorname{diam}\mathcal{C}_{L}(G)=n(G)+\mu(G) and (ii) there exists a correspondence cover (L,H)(L,H) such that |L(v)|=d(v)+2|L(v)|=d(v)+2 for all vv for which diam𝒞(L,H)(G)=n(G)+τ(G)\operatorname{diam}\mathcal{C}_{(L,H)}(G)=n(G)+\tau(G).

For both the List Conjecture and the Correspondence Conjecture, it is trivial to construct examples that need at least n(G)n(G) recolourings. We simply require that α(v)β(v)\alpha(v)\neq\beta(v) for all n(G)n(G) vertices vv. So we view these conjectured upper bounds as consisting of a “trivial” portion, n(G)n(G) recolourings, and a “non-trivial” portion, μ(G)\mu(G) or τ(G)\tau(G) recolourings. We give two partial results toward each conjecture. Our first result proves both conjectures up to a factor of 2 on the non-trivial portions of these upper bounds.

Theorem 1.

(i) For every graph GG and list-assignment LL with |L(v)|d(v)+2|L(v)|\geq d(v)+2 for all vV(G)v\in V(G) we have diam𝒞L(G)n(G)+2μ(G)\operatorname{diam}\mathcal{C}_{L}(G)\leq n(G)+2\mu(G). (ii) For every graph GG and correspondence cover (L,H)(L,H) with |L(v)|d(v)+2|L(v)|\geq d(v)+2 for all vV(G)v\in V(G) we have diam𝒞(L,H)(G)n(G)+2τ(G)\operatorname{diam}\mathcal{C}_{(L,H)}(G)\leq n(G)+2\tau(G).

Bousquet, Feuilloley, Heinrich, and Rabie [5, following Question 1.3] asked about the diameter of 𝒞k(G)\mathcal{C}_{k}(G) when k=Δ(G)+2k=\Delta(G)+2. In this case, a result of Bonamy and Bousquet [3, Theorem 1] implies, for an nn-vertex graph GG and k=Δ(G)+2k=\Delta(G)+2, that diam𝒞k(G)=O(Δ(G)n)\operatorname{diam}~{}\mathcal{C}_{k}(G)=O(\Delta(G)n). The authors of [5] asked whether it is possible to remove this dependency on Δ(G)\Delta(G). We answer their question affirmatively. Always μ(G)n(G)/2\mu(G)\leq n(G)/2, so Theorem 1(i) implies that diam𝒞k(G)2n(G)\operatorname{diam}\mathcal{C}_{k}(G)\leq 2n(G) when k=Δ(G)+2k=\Delta(G)+2.

To complement Theorem 1, we prove both conjectured upper bounds when |L(v)||L(v)| is sufficiently large.

Theorem 2.

(i) For every graph GG and list-assignment LL with |L(v)|2d(v)+1|L(v)|\geq 2d(v)+1 for all vV(G)v\in V(G) we have diam𝒞L(G)n(G)+μ(G)\operatorname{diam}\mathcal{C}_{L}(G)\leq n(G)+\mu(G). (ii) For every graph GG and correspondence cover (L,H)(L,H) with |L(v)|2d(v)+1|L(v)|\geq 2d(v)+1 for all vV(G)v\in V(G) we have diam𝒞(L,H)(G)n(G)+τ(G)\operatorname{diam}\mathcal{C}_{(L,H)}(G)\leq n(G)+\tau(G).

In Section 2 we present definitions and notation, as well as some easy lower bounds on diameters of reconfiguration graphs. In Section 3 we prove the list colouring portions of Theorems 1 and 2; and in Section 4 we prove the correspondence colouring portions. In Section 5 we prove more precise results when GG is a tree (in this case the correspondence problem is identical to the list problem).

In Sections 6 and 7, we conclude by proving one or both conjectures exactly for various graph classes, such as complete bipartite graphs, subcubic graphs, cactuses, and graphs with low maximum average degree. These graphs are all sparse or have low chromatic number. On the other end of the sparsity spectrum, the List Conjecture is also true for all complete graphs, due to an argument of Bonamy and Bousquet [3, Lemma 5].

2 Definitions, Notation, and Easy Lower Bounds

For a graph G=(V,E)G=(V,E), we denote its order by n(G)n(G). A matching is a set of vertex disjoint edges in GG. A matching MM of GG is perfect if it saturates every vertex of GG, and MM is near-perfect if it saturates every vertex of GG but one. A vertex cover SS is a vertex subset such that every edge of GG has at least one endpoint in SS. The matching number μ(G)\mu(G) of GG is the size of a largest matching. The vertex cover number τ(G)\tau(G) of GG is the size of a smallest vertex cover. (Recall that if GG is bipartite, then μ(G)=τ(G)\mu(G)=\tau(G).)

The distance dist(v,w)\operatorname{dist}(v,w) between two vertices v,wV(G)v,w\in V(G) is the length of a shortest path in GG between vv and ww. The eccentricity of a vertex vv, denoted ecc(v)\operatorname{ecc}(v), is maxwVdist(v,w).\max_{w\in V}\operatorname{dist}(v,w). The diameter diam(G)\operatorname{diam}(G) of GG is maxv,wVdist(v,w)\max_{v,w\in V}\operatorname{dist}(v,w), which is equal to maxvVecc(v)\max_{v\in V}\operatorname{ecc}(v), while the radius rad(G)\operatorname{rad}(G) of GG is minvVecc(v)\min_{v\in V}\operatorname{ecc}(v). A diameter can also refer to a shortest path between vv and ww for which dist(v,w)=diam(G).\operatorname{dist}(v,w)=\operatorname{diam}(G). A graph GG is kk-degenerate if every non-empty subgraph HH contains a vertex vv such that dH(v)kd_{H}(v)\leq k. The degeneracy of GG is the minimum kk such that GG is kk-degenerate.

A directed graph or digraph D=(V,A)D=(V,A) is analogous to a graph, except that each edge is directed. The directed edges in AA are ordered pairs of vertices, and are called arcs.

Definition 3.

For a digraph DD, the vertex cover number τ(D)\tau(D) and matching number μ(D)\mu(D) are defined to be τ(HD)\tau(H_{D}) and μ(HD)\mu(H_{D}), where HDH_{D} is the undirected graph with vertex set V(D)V(D) whose edges correspond to the bidirected edges in DD.

Colourings are mappings from VV to {\mathbb{N}}, and we denote them by greek letters such as α,β\alpha,\beta, and γ\gamma. A colouring α\alpha of GG is proper if for every edge vwE(G)vw\in E(G) we have α(v)α(w)\alpha(v)\neq\alpha(w). Let [k]:={1,2,,k}[k]:=\{1,2,\ldots,k\}. A kk-colouring is a colouring using at most kk colours, which are generally from the set [k][k]. The chromatic number χ(G)\chi(G) of GG is the smallest kk such that GG has a proper kk-colouring.

A list-assignment LL, for a graph GG, assigns to each vV(G)v\in V(G) a set L(v)L(v) of natural numbers (“list” of allowable colours). A proper LL-colouring is a proper colouring α:V(G)\alpha\colon V(G)\to{\mathbb{N}} such that α(v)L(v)\alpha(v)\in L(v) for every vV(G)v\in V(G). The list chromatic number (or choice number) χ(G)\chi_{\ell}(G) is the least kk such that GG admits a proper LL-colouring whenever every vV(G)v\in V(G) satisfies |L(v)|k\lvert L(v)\rvert\geq k.

A correspondence cover (L,H)(L,H) for a graph GG assigns to each vertex vV(G)v\in V(G) a set L(v)L(v) of colours {(v,1),,\{(v,1),\ldots, (v,f(v))}(v,f(v))\} and to each edge vwE(G)vw\in E(G) a matching between L(v)L(v) and L(w)L(w). (In this paper, we typically consider either f(v)=d(v)+2f(v)=d(v)+2 or f(v)=2d(v)+1f(v)=2d(v)+1 for all vV(G)v\in V(G).) Given a correspondence cover (L,H)(L,H), an (L,H)(L,H)-colouring of GG is a function α\alpha such that αL(v)\alpha\in L(v) for all vv and whenever vwE(G)vw\in E(G) the edge α(v)α(w)\alpha(v)\alpha(w) is not an edge of the matching assigned to vwvw. Here HH denotes the union of the matchings assigned to all edges of GG. (It is easy to check that correspondence colouring generalises list colouring.)

The reconfiguration graph 𝒞k(G)\mathcal{C}_{k}(G) has as its vertices the proper kk-colourings of GG, and two kk-colourings α\alpha and β\beta are adjacent in 𝒞k(G)\mathcal{C}_{k}(G) if they differ on exactly one vertex of GG. In particular, if k<χ(G)k<\chi(G), then 𝒞k(G)\mathcal{C}_{k}(G) has no vertices. (The reconfiguration graph was first defined in [8], where it was called the kk-colour graph.) The distance between kk-colourings α\alpha and β\beta is at most jj if we can form β\beta, starting from α\alpha, by recolouring at most jj vertices, one at a time, so that after each recolouring the current kk-colouring of GG is proper. Similarly, for a graph GG and list-assignment LL, the reconfiguration graph 𝒞L(G)\mathcal{C}_{L}(G), or reconfiguration graph of the LL-colourings of GG, is the graph whose vertices are the proper LL-colourings of GG. Again, two LL-colourings α\alpha and β\beta are adjacent in 𝒞L(G)\mathcal{C}_{L}(G) if they differ on exactly one vertex. This extension was first defined in [4]. For a correspondence cover (L,H)(L,H), the reconfiguration graph 𝒞(L,H)(G)\mathcal{C}_{(L,H)}(G) is defined analogously; its vertices are the correspondence colourings of GG and two vertices of 𝒞(L,H)(G)\mathcal{C}_{(L,H)}(G) are adjacent precisely when their colourings differ on exactly one vertex of GG. We will also use the associated colour-shift digraph for a graph GG and two proper colourings, which is defined as follows.

Definition 4.

Let G=(V,E)G=(V,E) be a graph and α\alpha and β\beta be two proper colourings of GG. We define an associated colour-shift digraph Dα,β=(V,A)D_{\alpha,\beta}=(V,A) where vwA\overrightarrow{vw}\in A if and only if vwEvw\in E and β(v)=α(w)\beta(v)=\alpha(w).

2.1 Improved Lower Bounds

In this subsection we prove lower bounds which show that our upper bounds in the rest of the paper are sharp or nearly sharp. These will not be explicitly needed elsewhere, so the impatient reader should feel free to skip to Section 3, where we prove Theorems 1(i) and 2(i).

Observation 5.

If GG is a graph with list colourings α\alpha and β\beta, then dist(α,β)μ(Dα,β)+vV𝟏α(v)β(v)\operatorname{dist}(\alpha,\beta)\geq\mu(D_{\alpha,\beta})+\sum_{v\in V}\mathbf{1}_{\alpha(v)\not=\beta(v)}.

Proof.

Whenever α(v)β(v)\alpha(v)\not=\beta(v), vertex vv must be recoloured. Further, if vv and ww are neighbours for which α(v)=β(w)\alpha(v)=\beta(w) and α(w)=β(v)\alpha(w)=\beta(v), then we cannot recolour both vv and ww only once, since after every recolouring step the resulting colouring must be proper. So at least vV𝟏α(v)β(v)\sum_{v\in V}\mathbf{1}_{\alpha(v)\not=\beta(v)} vertices must be recoloured, and at least μ(Dα,β)\mu(D_{\alpha,\beta}) vertices must be recoloured at least twice. ∎

Does the lower bound in Observation 5 always hold with equality? Our next example shows that it does not; see Figure 1. Specifically, this example exhibits, for the 4-cycle, colourings α\alpha and β\beta such that μ(Dα,β)<μ(G)\mu(D_{\alpha,\beta})<\mu(G) but still dist(α,β)=n(G)+μ(G)\operatorname{dist}(\alpha,\beta)=n(G)+\mu(G).

Example 6.

Let G=C4G=C_{4} and denote V(G)V(G) by [4][4]. If α(i)=i\alpha(i)=i and β(i)i+1(mod4)\beta(i)\equiv i+1\pmod{4} and the colour set is precisely [4][4], then transforming α\alpha to β\beta uses at least 6 recolourings, i.e. dist(α,β)=6\operatorname{dist}(\alpha,\beta)=6. Part of this reconfiguration graph is shown in Figure 1.

Proof.

Starting from α\alpha, each vertex ii has a single colour with which it can be recoloured. But each possible recolouring creates a colouring α\alpha^{\prime} for which Observation 5 gives dist(α,β)5\operatorname{dist}(\alpha^{\prime},\beta)\geq 5. ∎

Figure 1: Part of the reconfiguration graph 𝒞4(C4)\mathcal{C}_{4}(C_{4})

Next we present the (previously promised) proof of Proposition 2. For easy reference, we restate it.

Proposition 2.

For every graph GG (i) there exists a list-assignment LL such that |L(v)|=d(v)+2|L(v)|=d(v)+2 for all vv for which diam𝒞L(G)=n(G)+μ(G)\operatorname{diam}\mathcal{C}_{L}(G)=n(G)+\mu(G) and (ii) there exists a correspondence cover (L,H)(L,H) such that |L(v)|=d(v)+2|L(v)|=d(v)+2 for all vv for which diam𝒞(L,H)(G)=n(G)+τ(G)\operatorname{diam}\mathcal{C}_{(L,H)}(G)=n(G)+\tau(G).

Proof.

For a graph GG, fix a maximum matching MM. Assign to each vertex vv a list of d(v)+2d(v)+2 colours such that |L(v)L(w)|=2|L(v)\cap L(w)|=2 if vwMvw\in M and otherwise |L(v)L(w)|=0|L(v)\cap L(w)|=0. Pick α\alpha and β\beta such that for each vwMvw\in M, we have α(v)=β(w)\alpha(v)=\beta(w) and α(w)=β(v)\alpha(w)=\beta(v), and for each vV(G)v\in V(G) we have α(v)β(v).\alpha(v)\not=\beta(v). The lower bound holds by Observation 5. The upper bound is trivial, since L(x)L(y)=L(x)\cap L(y)=\emptyset for all xyE(G)Mxy\in E(G)\setminus M. This proves (i).

Let (L,H)(L,H) be a correspondence cover of GG such that L(v)={(v,1)(v,d(v)+2)}L(v)=\{(v,1)\ldots(v,d(v)+2)\} for every vV(G)v\in V(G). For every vwE(G)vw\in E(G), let the matching (from HH) consist of the two edges (v,1)(w,2)(v,1)(w,2) and (v,2)(w,1)(v,2)(w,1); otherwise, (v,i)(v,i) and (w,j)(w,j) are unmatched. Let α(v):=(v,1)\alpha(v):=(v,1) and β(v):=(v,2)\beta(v):=(v,2) for all vV(G)v\in V(G). Starting from α\alpha, we can recolour a vertex vv with β(v)\beta(v) only if all neighbours of vv have already been recoloured. Thus, the set of vertices recoloured only once must be an independent set; equivalently, the set of vertices recoloured at least twice must be a vertex cover. So we must use at least n(G)+τ(G)n(G)+\tau(G) recolourings, which proves the lower bound. For the upper bound, fix a minimum vertex cover SS. First recolour each vSv\in S with (v,3)(v,3); afterward, recolour each vertex with (v,2)(v,2), first the vertices in V(G)\SV(G)\backslash S and then those in SS. So n(G)+τ(G)n(G)+\tau(G) recolourings suffice. This proves (ii). ∎

It is helpful to note that the proof of Proposition 2 actually works (for both list colouring and correspondence colouring) whenever all vv satisfy |L(v)|3|L(v)|\geq 3.

Clearly, the graph G^\widehat{G} constructed in Proposition 1 depends on our choice of a perfect matching MM. It is interesting to note that some choices of MM work quite well, while others work rather poorly.

Example 7.

Figure 2 shows a graph GG built from two cliques KpK_{p} and a complete bipartite graph Kp,pK_{p,p}, with parts UU and WW, by adding a perfect matching MM between one copy of KpK_{p} and the vertices of UU and a perfect matching between WW and the other copy of KpK_{p}. Now χ(G^)=2p\chi(\widehat{G})=2p if we choose our perfect matching to be MM. If pp is even, then we can instead choose a perfect matching MM^{\prime} that is the disjoint union of perfect matchings in the two copies of KpK_{p} and in the Kp,pK_{p,p} so that the resulting G^\widehat{G} instead satisfies χ(G^)=p+2\chi(\widehat{G})=p+2.

Figure 2: A graph GG (shown on both left and right) consisting of a complete bipartite graph Kp,pK_{p,p} in the center, two copies of KpK_{p} on the sides, and a matching from each copy of KpK_{p} to one part of Kp,pK_{p,p}. The left and right figures specify different perfect matchings MM (in bold). On the left, χ(G^)ω(G^)=|V(Kp,p)|=2p=2Δ(G)2\chi(\widehat{G})\geq\omega(\widehat{G})=|V(K_{p,p})|=2p=2\Delta(G)-2. On the right, ω(G^)=ω(G)=p\omega(\widehat{G})=\omega(G)=p, and in fact χ(G^)p+2=Δ(G)+1\chi(\widehat{G})\leq p+2=\Delta(G)+1. (We use colours 1,,p1,\ldots,p on each clique and colours p+1p+1 and p+2p+2 in the center.)

Although Example 7 implies that our choice of MM can effect χ(G^)\chi(\widehat{G}) dramatically, for many graphs, any choice of MM is fine. To conclude this section, we present various cases in which (for all choices of MM) we can weaken the hypothesis of Proposition 1.

Proposition 3.

For a graph GG and positive integer kk we have

diam𝒞k(G)n(G)+μ(G)\operatorname{diam}~{}\mathcal{C}_{k}(G)\geq n(G)+\mu(G)

whenever (a) χ(G^)k1\chi(\widehat{G})\leq k-1 or (b) χ(G^)k\chi(\widehat{G})\leq k and kΔ(G)+2k\geq\Delta(G)+2, where G^\widehat{G} is as in Proposition 1. In particular, this is true in the following cases.

  1. (1)(1)

    1+χ(G)(χ(G)1)k1+\chi(G)(\chi(G)-1)\leq k (in particular, if GG is bipartite and k3k\geq 3).

  2. (2)(2)

    Δ(G)=3\Delta(G)=3 and kΔ(G)+2=5k\geq\Delta(G)+2=5.

  3. (3)(3)

    χ(G)3\chi(G)\leq 3 and kΔ(G)+2k\geq\Delta(G)+2 (in particular, if GG is outerplanar and kΔ(G)+2k\geq\Delta(G)+2).

  4. (4)(4)

    kΔ(G)+3k\geq\Delta(G)+3 and GG is triangle-free, if Reed’s Conjecture111Recall that Reed conjectured that χ(G)(Δ(G)+1+ω(G))/2\chi(G)\leq\left\lceil(\Delta(G)+1+\omega(G))/2\right\rceil for every graph GG. holds for every graph G^\widehat{G} with ω(G^)5\omega(\widehat{G})\leq 5.

  5. (5)(5)

    kΔ(G)+2k\geq\Delta(G)+2 and Δ(G)>f(ω(G))\Delta(G)>f(\omega(G)) for some function ff.

Proof.

Recall the definition of G^\widehat{G} from Proposition 1. Let MM be a maximum matching in GG. Form G^\widehat{G} from GG as follows. If vw,xyMvw,xy\in M and wxE(G)wx\in E(G), then add vyvy to G^\widehat{G}. To prove each case of the proposition, we construct the graph G^\widehat{G}, let α\alpha be a χ(G^)\chi(\widehat{G})-colouring of G^\widehat{G} (which is also a χ(G^)\chi(\widehat{G})-colouring of GG), and form a kk-colouring β\beta from α\alpha by swapping the colours on the endpoints of each edge ee in the specified maximum matching MM; for each vv not saturated by MM, we choose β(v)\beta(v) to avoid α(v)xN(v)β(x)\alpha(v)\cup\bigcup_{x\in N(v)}\beta(x). The latter is possible because we assumed that kk is bounded from below by χ(G^)+1\chi(\widehat{G})+1 or by Δ(G)+2\Delta(G)+2. By Observation 5, we have dist(α,β)μ(Dα,β)+vV𝟏α(v)β(v)=μ(G)+n(G)\operatorname{dist}(\alpha,\beta)\geq\mu(D_{\alpha,\beta})+\sum_{v\in V}\mathbf{1}_{\alpha(v)\not=\beta(v)}=\mu(G)+n(G). Note that Δ(G^)2Δ(G)1\Delta(\widehat{G})\leq 2\Delta(G)-1. So χ(G^)1+Δ(G^)2Δ(G)\chi(\widehat{G})\leq 1+\Delta(\widehat{G})\leq 2\Delta(G). Now we show that in each case listed above we have χ(G^)k\chi(\widehat{G})\leq k or (in the first case) χ(G^)k1\chi(\widehat{G})\leq k-1.

  1. (1)(1)

    Fix a χ(G)\chi(G)-colouring α\alpha of GG. We assume that MM is a perfect matching; if not, then we add a pendent edge at each vertex that is unsaturated by MM and add all these new edges to MM. To colour G^\widehat{G}, we give each vertex vv the colour (i,j)(i,j), where α(v)=i\alpha(v)=i, α(w)=j\alpha(w)=j, and vwMvw\in M. It is easy to check that the resulting colouring of G^\widehat{G} is proper.

  2. (2)(2)

    Suppose that Δ(G)=3\Delta(G)=3. As noted above, we have Δ(G^)2Δ(G)1=5\Delta(\widehat{G})\leq 2\Delta(G)-1=5. If Δ(G^)4\Delta(\widehat{G})\leq 4, then χ(G^)5\chi(\widehat{G})\leq 5 as desired. So assume instead that Δ(G^)=5\Delta(\widehat{G})=5. By Brooks’ Theorem it suffices to show that no component of G^\widehat{G} is K6K_{6}. Suppose, to the contrary, that there exists SV(G)S\subseteq V(G) such that G^[S]=K6\widehat{G}[S]=K_{6}. It is easy to check that G[S]G[S] must be 3-regular, and every vertex of SS must be saturated by MM. Observe that if two edges of MM, say v1v2,v3v4v_{1}v_{2},v_{3}v_{4} lie on a 4-cycle in GG, then dG^(vi)4d_{\widehat{G}}(v_{i})\leq 4, for each i{1,2,3,4}i\in\{1,2,3,4\}. Now we need only to check that such a configuration occurs for every 3-regular graph GG on 6 vertices and every perfect matching MM. This is straightforward to verify: GMG-M is either a 6-cycle or two 3-cycles; in the first case, MM can be added in two (non-isomorphic) ways and in the second, MM can be added in only one way.

  3. (3)(3)

    If Δ(G)4\Delta(G)\geq 4, then we are done by condition (1)(1). If Δ(G)=3\Delta(G)=3, then we are done by condition (2)(2). If Δ(G)=2\Delta(G)=2, then we are done by Proposition 1. And if Δ(G)1\Delta(G)\leq 1, then we are done trivially.

  4. (4)(4)

    We show that if GG is triangle-free, then ω(G^)5\omega(\widehat{G})\leq 5. Suppose, to the contrary, that ω(G^)6\omega(\widehat{G})\geq 6. Consider SV(G)S\subseteq V(G) such that G^[S]=K6\widehat{G}[S]=K_{6}, and colour each edge of G^[S]\widehat{G}[S] with 1 if it appears in GG and with 2 otherwise. Since GG is triangle-free, G^[S]\widehat{G}[S] has no triangle coloured 1. Let R(t,t)R(t,t) denote the diagonal Ramsey number for cliques of order tt. Since R(3,3)=6R(3,3)=6, we must have a triangle in G^[S]\widehat{G}[S] that is coloured 2; denote its vertices by v,w,xv,w,x. Let vv,ww,xxvv^{\prime},ww^{\prime},xx^{\prime} denote the edges of MM incident with v,w,xv,w,x. Since vw,vx,wxE(G^)E(G)vw,vx,wx\in E(\widehat{G})\setminus E(G), we conclude that vw,vx,wxE(G)v^{\prime}w^{\prime},v^{\prime}x^{\prime},w^{\prime}x^{\prime}\in E(G). This contradicts that GG is triangle-free. Hence, ω(G^)5\omega(\widehat{G})\leq 5, as claimed. Finally, by Reed’s Conjecture, χ(G^)(Δ(G^)+ω(G^)+1)/2(2Δ(G)1+5+1)/2Δ(G)+3k\chi(\widehat{G})\leq\left\lceil(\Delta(\widehat{G})+\omega(\widehat{G})+1)/2\right\rceil\leq\left\lceil(2\Delta(G)-1+5+1)/2\right\rceil\leq\Delta(G)+3\leq k.

  5. (5)(5)

    Let tt be a positive integer and consider any graph GG with ω(G)=t\omega(G)=t. Note that G^\widehat{G} can be decomposed into two (edge-disjoint) copies of GG; here G^\widehat{G} is formed from the two copies of GG by identifying, for each edge vwMvw\in M, the instance of vv in one copy of GG with the instance of ww in the other copy. This implies that ω(G^)<R(t+1,t+1)\omega(\widehat{G})<R(t+1,t+1). We also know that Δ(G^)2Δ(G).\Delta(\widehat{G})\leq 2\Delta(G). Now by e.g. [13, Theorem 2] we have

    χ(G^)200R(t+1,t+1)2Δ(G)lnln(2Δ(G))ln(2Δ(G)).\chi(\widehat{G})\leq 200R(t+1,t+1)\frac{2\Delta(G)\ln\ln(2\Delta(G))}{\ln(2\Delta(G))}.

    This is smaller than Δ(G)+2\Delta(G)+2 whenever Δ(G)\Delta(G) is sufficiently large.

We also consider the graph G~\tilde{G} formed from GG by contracting each edge of a maximum matching MM.

Proposition 4.

For a graph GG and positive integer kk we have

diam𝒞k(G)n(G)+μ(G)\operatorname{diam}~{}\mathcal{C}_{k}(G)\geq n(G)+\mu(G)

whenever k2χ(G~)k\geq 2\chi(\tilde{G}) for some maximum matching of GG. In particular, this is

  1. (1)(1)

    true a.a.s. for Gn,pG_{n,p}, for fixed p(0,1)p\in(0,1), and kΔ(Gn,p)+2.k\geq\Delta(G_{n,p})+2.

  2. (2)(2)

    true if GG is Kt+1K_{t+1}-minor free and k2tk\geq 2t, provided Hadwiger’s Conjecture222Recall that Hadwiger conjectured that χ(G)t\chi(G)\leq t for every graph GG with no Kt+1K_{t+1}-minor. holds for Kt+1K_{t+1}-minor free graphs. The latter is true for t5t\leq 5.

Proof.

Let α~\tilde{\alpha} be a χ(G~)\chi(\tilde{G})-colouring of G~\tilde{G}. We construct a 2χ(G~)2\chi(\tilde{G})-colouring α\alpha of GG such that if vV(G~)v\in V(\tilde{G}) and vv arises from contracting edge wxMwx\in M and α~(v)=i\tilde{\alpha}(v)=i, then {α(w),α(x)}={2i1,2i}\{\alpha(w),\alpha(x)\}=\{2i-1,2i\}; if vv does not arise from contracting an edge, then we choose α(v)=2α~(v)=2i\alpha(v)=2\tilde{\alpha}(v)=2i. Form β\beta from α\alpha by swapping the colours on the endpoints of each edge of MM and letting β(v):=α(v)1=2i1\beta(v):=\alpha(v)-1=2i-1 for each vV(G)\Mv\in V(G)\backslash M.

Hence diam𝒞k(G)dist(α,β)n(G)+μ(G)\operatorname{diam}\mathcal{C}_{k}(G)\geq\operatorname{dist}(\alpha,\beta)\geq n(G)+\mu(G) by Observation 5. We now prove that in the two mentioned cases, we (a.a.s.) have k2χ(G~)k\geq 2\chi(\tilde{G}), regardless of the maximum matching of GG we choose to form G~\tilde{G}.

  1. (1)(1)

    Let G:=Gn,pG:=G_{n,p}. By [2], the Hadwiger number hh (or contraction clique number) of GG for fixed p(0,1)p\in(0,1) is Θ(nlnn)\Theta\left(\frac{n}{\sqrt{\ln n}}\right) a.a.s., so GG is Kh+1K_{h+1}-minor free. Note that G~\tilde{G} is also Kh+1K_{h+1}-minor free. By recent progress towards Hadwiger’s conjecture, culminating in [9], this implies that (still a.a.s.) χ(G~)=O(hlnlnh)=O(nlnlnnlnn)\chi(\tilde{G})=O(h\ln\ln h)=O\left(\frac{n\ln\ln n}{\sqrt{\ln n}}\right). So χ(G~)\chi(\tilde{G}) is a.a.s. smaller than Δ(Gn,p)+2\Delta(G_{n,p})+2, which is np(1+o(1))=Θ(n).np(1+o(1))=\Theta(n).

  2. (2)(2)

    Note that G~\tilde{G} is also Kt+1K_{t+1}-minor free. So, by assumption, χ(G~)t\chi(\tilde{G})\leq t.

Question 1.

Is it true, for every graph GG and every kΔ(G)+2k\geq\Delta(G)+2, that diam𝒞k(G)n(G)+μ(G)\operatorname{diam}~{}\mathcal{C}_{k}(G)\geq n(G)+\mu(G)?

If the answer to this question is yes, then the List Conjecture (if true) would imply that diam𝒞k(G)=n(G)+μ(G)\operatorname{diam}~{}\mathcal{C}_{k}(G)=n(G)+\mu(G) for every graph GG and every kΔ(G)+2k\geq\Delta(G)+2. In particular, this would constitute a precise version for Cereceda’s Conjecture in the case of regular graphs.

3 Proving the Main Results for Lists

In this section, we prove the list colouring portions of our main results: Theorems 1(i) and 2(i). Recall that a graph GG is factor-critical if |V(G)||V(G)| is odd and GvG-v has a perfect matching for each vV(G)v\in V(G). Gallai showed that if every vertex is unsaturated by some maximum matching, i.e., μ(Gv)=μ(G)\mu(G-v)=\mu(G) for every vertex vv, then GG is factor-critical. So, in each proof we begin with a lemma to handle factor-critical graphs, and thereafter proceed to the general case. The following lemma serves as the base case in an inductive proof of Theorem 1(i). (We will only need it when GG is factor-critical, but we prove it for all GG.)

Lemma 8.

Let GG be graph and let LL be a list-assignment for GG such that |L(v)|d(v)+2|L(v)|\geq d(v)+2 for all vV(G)v\in V(G). If α\alpha and β\beta are LL-colourings of GG, then we can recolour GG from α\alpha to β\beta in at most 2n(G)12n(G)-1 steps.

Proof.

We use induction on |β(V(G))||\beta(V(G))|. The case |β(V(G))|=1|\beta(V(G))|=1, is trivial, since then GG is an independent set; starting from α\alpha, we can recolour each vertex vv to β(v)\beta(v). We use at most n(G)n(G) recolouring steps, which suffices since n(G)2n(G)1n(G)\leq 2n(G)-1.

Now assume |β(V(G))|2|\beta(V(G))|\geq 2. For each colour cc, let α1(c):={vV(G) s.t. α(v)=c}\alpha^{-1}(c):=\{v\in V(G)\mbox{ s.t. }\alpha(v)=c\} and β1(c):={vV(G) s.t. β(v)=c}\beta^{-1}(c):=\{v\in V(G)\mbox{ s.t. }\beta(v)=c\}. Since cα(V(G))|α1(c)|=n(G)=cβ(V(G))|β1(c)|\sum_{c\in\alpha(V(G))}|\alpha^{-1}(c)|=n(G)=\sum_{c\in\beta(V(G))}|\beta^{-1}(c)|, there exists cc such that |α1(c)||β1(c)||\alpha^{-1}(c)|\leq|\beta^{-1}(c)|. Starting from α\alpha, for each vα1(c)v\in\alpha^{-1}(c), recolour vv to avoid colour cc. Now for each vβ1(c)v\in\beta^{-1}(c), recolour vv with cc. After at most 2|β1(c)|2|\beta^{-1}(c)| steps, we have |β1(c)||\beta^{-1}(c)| more vertices that are coloured to agree with β\beta. Now we delete these |β1(c)||\beta^{-1}(c)| vertices, and delete the colour cc from L(v)L(v) whenever vv had some neighbours in β1(c)\beta^{-1}(c), and finish by induction. ∎

Corollary 9.

For every graph GG and every kΔ(G)+2k\geq\Delta(G)+2, the diameter of 𝒞k(G)\mathcal{C}_{k}(G) is at most 2n(G)12n(G)-1.

Now we prove Theorem 1(i). For easy reference, we restate it below.

Theorem 10.

Let GG be an nn-vertex graph and LL be a list-assignment with |L(v)|d(v)+2|L(v)|\geq d(v)+2 for all vV(G)v\in V(G). If α\alpha and β\beta are LL-colourings of GG, then we can recolour GG from α\alpha to β\beta in at most n(G)+2μ(G)n(G)+2\mu(G) steps.

Proof.

We use induction on n(G)n(G). We assume GG is connected; otherwise, we handle each component separately. (This suffices because, for a graph GG with components G1,,GsG_{1},\ldots,G_{s}, we have n(G)=i=1sn(Gi)n(G)=\sum_{i=1}^{s}n(G_{i}) and μ(G)=i=1sμ(Gi)\mu(G)=\sum_{i=1}^{s}\mu(G_{i}).) If μ(Gv)=μ(G)\mu(G-v)=\mu(G) for all vV(G)v\in V(G), then GG is factor-critical. This is an easy result of Gallai, but also follows from Theorem 12(2). Now n(G)+2μ(G)=2n(G)1n(G)+2\mu(G)=2n(G)-1, so we are done by Lemma 8.

Assume instead that some vertex vv is saturated by every maximum matching. Note that 2d(v)<2|L(v)|2d(v)<2|L(v)|. By Pigeonhole there exists cL(v)c\in L(v) such that |β1(c)N(v)|+|α1(c)N(v)|1|\beta^{-1}(c)\cap N(v)|+|\alpha^{-1}(c)\cap N(v)|\leq 1. (We handle the case when equality holds, since the other case is easier.) Assume there exists wN(v)w\in N(v) such that α(w)=c\alpha(w)=c and cxN(v){w}{α(x),β(x)}c\notin\cup_{x\in N(v)\setminus\{w\}}\{\alpha(x),\beta(x)\}. (If instead β(w)=c\beta(w)=c, then we swap the roles of α\alpha and β\beta.) Form α~\tilde{\alpha} from α\alpha by recolouring ww from L(w)(α(N(w)){c})L(w)\setminus(\alpha(N(w))\cup\{c\}) and recolouring vv with cc. Let G:=GvG^{\prime}:=G-v. Let L(x):=L(x)cL^{\prime}(x):=L(x)-c for all xN(v)x\in N(v) and L(x):=L(x)L^{\prime}(x):=L(x) for all other xV(G)x\in V(G^{\prime}). Denote the restrictions to GG^{\prime} of α~\tilde{\alpha} and β\beta by α~\tilde{\alpha}^{\prime} and β\beta^{\prime}. By induction, we can recolour GG^{\prime} from α~\tilde{\alpha}^{\prime} to β\beta^{\prime} in at most n(G)+2μ(G)n(G^{\prime})+2\mu(G^{\prime}) steps. After this, we can finish by recolouring vv with β(v)\beta(v). Thus, distL(α,β)2+distL(α~,β)+12+n(G)+2μ(G)+1=2+n(G)1+2(μ(G)1)+1=n(G)+2μ(G)\operatorname{dist}_{L}(\alpha,\beta)\leq 2+\operatorname{dist}_{L^{\prime}}(\tilde{\alpha}^{\prime},\beta^{\prime})+1\leq 2+n(G^{\prime})+2\mu(G^{\prime})+1=2+n(G)-1+2(\mu(G)-1)+1=n(G)+2\mu(G). ∎

To prove Theorem 2(i), we will use the following lemma to handle the case when GG is factor-critical.

Lemma 11.

Let GG be a graph and let LL be a list-assignment for GG such that |L(v)|2d(v)+1|L(v)|\geq 2d(v)+1 for all vV(G)v\in V(G). If α\alpha and β\beta are LL-colourings of GG, then we can recolour GG from α\alpha to β\beta in at most 3n(G)2\left\lfloor\frac{3n(G)}{2}\right\rfloor steps.

Proof.

Let α\alpha and β\beta be arbitrary LL-colourings of GG. (Recall that all colours are positive integers.) Let V1V_{1} and V2V_{2} be the subsets of V(G)V(G), respectively, for which α(v)>β(v)\alpha(v)>\beta(v) and α(v)<β(v)\alpha(v)<\beta(v). Let n1:=|V1|n_{1}:=|V_{1}| and n2:=|V2|n_{2}:=|V_{2}|. Since n1+n2n(G)n_{1}+n_{2}\leq n(G), we assume by symmetry that n1n(G)2n_{1}\leq\frac{n(G)}{2}; otherwise, we interchange α\alpha and β\beta. We show how to recolour GG from α\alpha to β\beta using at most n(G)+n1n(G)+n(G)2n(G)+n_{1}\leq n(G)+\left\lfloor\frac{n(G)}{2}\right\rfloor steps.

First, iteratively for every vV1v\in V_{1} we recolour vv from L(v)wN(v){β(w),γ(w)}L(v)\setminus\cup_{w\in N(v)}\{\beta(w),\gamma(w)\}, where γ\gamma is the current colouring of GG. By definition this recolouring is proper (we actually do not need to recolour vv if γ(v)β(w)\gamma(v)\not=\beta(w) for all wN(v)w\in N(v)). Next, iteratively for jj from max{vVL(v)}\max\{\cup_{v\in V}L(v)\} to min{vVL(v)}\min\{\cup_{v\in V}L(v)\}, we recolour with jj all vertices vV(G)v\in V(G) for which β(v)=j\beta(v)=j. In these steps, we only have proper colourings, as we will now show. Consider a neighbour ww of vv. If ww has already been recoloured with some γ(w)\gamma(w), then γ(w)β(v)\gamma(w)\neq\beta(v), by construction. If ww has not been recoloured, then α(w)β(w)<β(v)\alpha(w)\leq\beta(w)<\beta(v). ∎

The following classical result is helpful in our proof of Theorem 2(i).

Theorem 12 (Edmonds–Gallai Decomposition Theorem).

For a graph GG, let

  1.     

    V1(G):={vV(G)V_{1}(G):=\{v\in V(G) such that some maximum matching avoids v}v\},

  2.     

    V2(G):={vV(G)V_{2}(G):=\{v\in V(G) such that vv has a neighbour in V1(G)V_{1}(G), but vV1(G)}v\notin V_{1}(G)\},

  3.     

    V3(G):=V(G)(V1(G)V2(G))V_{3}(G):=V(G)\setminus(V_{1}(G)\cup V_{2}(G)).

Now the following statements hold.333We prefer the names V1(G)V_{1}(G), V2(G)V_{2}(G), V3(G)V_{3}(G) over the standard terminology D(G)D(G), A(G)A(G), and C(G)C(G), since the latter terms conflict with our use of D=(V,A)D=(V,A) for a digraph DD with vertex set VV and arc set AA.

  1. (1)(1)

    The subgraph induced by V3(G)V_{3}(G) has a perfect matching.

  2. (2)(2)

    The components of the subgraph induced by V1(G)V_{1}(G) are factor-critical.

  3. (3)(3)

    If MM is any maximum matching, then MM contains a perfect matching of each component of V3(G)V_{3}(G), and MM contains a near-perfect matching of each component of V1(G)V_{1}(G), and MM matches all vertices of V2(G)V_{2}(G) with vertices in distinct components of V1(G)V_{1}(G).

  4. (4)(4)

    μ(G)=12(|V(G)|c(V1(G))+|V2(G)|)\mu(G)=\frac{1}{2}\left(|V(G)|-c(V_{1}(G))+|V_{2}(G)|\right), where c(V1(G))c(V_{1}(G)) denotes the number of components of the graph spanned by V1(G)V_{1}(G).

Now we prove Theorem 2(i). For easy reference, we restate it below.

Theorem 13.

Let GG be a graph and LL be a list-assignment for GG with |L(v)|2d(v)+1|L(v)|\geq 2d(v)+1 for all vV(G)v\in V(G). If α\alpha and β\beta are LL-colourings of GG, then we can recolour GG from α\alpha to β\beta in at most n(G)+μ(G)n(G)+\mu(G) steps.

Proof.

We refer to V1(G)V_{1}(G), V2(G)V_{2}(G), and V3(G)V_{3}(G), as defined in Theorem 12. Starting from α\alpha, recolour each vertex vV2(G)v\in V_{2}(G) to avoid each colour used currently on N(v)N(v) and also to avoid each colour used on N(v)N(v) in β\beta; call the resulting colouring α~(G)\tilde{\alpha}(G). Let G:=GV2(G)G^{\prime}:=G-V_{2}(G). For each vV(G)v\in V(G^{\prime}), let L(v):=L(v)(wN(v)V2(G)α~(w))L^{\prime}(v):=L(v)\setminus(\cup_{w\in N(v)\cap V_{2}(G)}\tilde{\alpha}(w)). Note that |L(v)|2dG(v)+1|L^{\prime}(v)|\geq 2d_{G^{\prime}}(v)+1. Thus, for each component HH of GG^{\prime}, we can recolour HH from α\alpha to β\beta in at most 3n(H)/2\lfloor 3n(H)/2\rfloor steps, by Lemma 11. Finally, we recolour each vertex of V2(G)V_{2}(G) to its colouring β\beta. Let comp(G)\operatorname{comp}(G^{\prime}) denote the set of components of GG^{\prime}. The number of recolouring steps used is at most 2|V2(G)|+Hcomp(G)3n(H)/2=2|V2(G)|+32(|V(G)||V2(G)|)12c(V1(G))=|V(G)|+12(|V(G)|c(V1(G))+|V2(G)|)=|V(G)|+μ(G)2|V_{2}(G)|+\sum_{H\in\operatorname{comp}(G^{\prime})}\lfloor 3n(H)/2\rfloor=2|V_{2}(G)|+\frac{3}{2}(|V(G)|-|V_{2}(G)|)-\frac{1}{2}c(V_{1}(G))=|V(G)|+\frac{1}{2}(|V(G)|-c(V_{1}(G))+|V_{2}(G)|)=|V(G)|+\mu(G). The final equality uses Theorem 12(4). The first equality uses that the vertices of a component Hcomp(G)H\in\operatorname{comp}(G^{\prime}) are contained either in V1(G)V_{1}(G) or in V3(G)V_{3}(G). If V(H)V3(G)V(H)\subseteq V_{3}(G) then n(H)n(H) is even by Theorem 12(3). On the other hand, if V(H)V1(G)V(H)\subseteq V_{1}(G) then n(H)n(H) is odd by Theorem 12(2). ∎

By the comment after Proposition 2, Theorem 13 is sharp. For (non-list) colouring, we get the following.

Corollary 14.

For every graph GG and every k2Δ(G)+1k\geq 2\Delta(G)+1, we have diam𝒞k(G)=n(G)+μ(G)\operatorname{diam}\mathcal{C}_{k}(G)=n(G)+\mu(G).

Proof.

The upper and lower bounds follow, respectively, from Theorem 13 and Proposition 1. ∎

4 Proving the Main Results for Correspondences

In this section, we prove the correspondence colouring portions of our main results: Theorems 1(ii) and 2(ii).

Theorem 15.

Let GG be a graph and (L,H)(L,H) be a correspondence cover with |L(v)|2d(v)+1|L(v)|\geq 2d(v)+1 for all vV(G)v\in V(G). If α\alpha and β\beta are (L,H)(L,H)-colourings of GG, then GG can be recoloured from α\alpha to β\beta in at most n(G)+τ(G)n(G)+\tau(G) steps.

Proof.

Let SS be a minimum vertex cover. For every vertex vSv\in S, we recolour vv with a colour that, for every wN(v)w\in N(v), conflicts (under cover (L,H)(L,H)) with neither the current colour γ(w)\gamma(w) nor the final colour β(w)\beta(w). This is possible because |L(v)|2d(v)+1|L(v)|\geq 2d(v)+1. Now we recolour every vertex vV(G)\Sv\in V(G)\backslash S with β(v)\beta(v) and then do this also for every vertex in SS. Note that we use at most n(G)+τ(G)n(G)+\tau(G) recolourings. ∎

Theorem 16.

Let GG be a graph and (L,H)(L,H) be a correspondence cover for GG such that |L(v)|d(v)+2|L(v)|\geq d(v)+2 for all vV(G)v\in V(G). If α\alpha and β\beta are (L,H)(L,H)-colourings of GG, then we can recolour GG from α\alpha to β\beta in at most n(G)+2τ(G)n(G)+2\tau(G) steps.

Proof.

We use induction on τ(G)\tau(G). If τ(G)=0\tau(G)=0, i.e., GG is an independent set, then we simply recolour every vertex vv to β(v)\beta(v). So assume the theorem is true for all graphs GG^{\prime} with 0τ(G)<τ(G)0\leq\tau(G^{\prime})<\tau(G).

Let SS be a minimum vertex cover of GG and pick vSv\in S. Consider the union over all wN(v)w\in N(v) of the edges in HH from α(w)\alpha(w) and β(w)\beta(w) to L(v)L(v). By Pigeonhole, since 2d(v)<2|L(v)|2d(v)<2|L(v)|, some colour cL(v)c\in L(v) has at most one incident edge in this union. (We handle the case when cc has exactly one incident edge, since the other case is easier.) Assume there exists w0N(v)w_{0}\in N(v) such that α(w0)\alpha(w_{0}) is matched to cc and no other α(w)\alpha(w) or β(w)\beta(w) is matched to cc. (If instead β(w0)\beta(w_{0}) is matched to cc, then we swap the roles of α\alpha and β\beta.) Form α~\tilde{\alpha} from α\alpha by first recolouring w0w_{0} with a colour different from α(w0)\alpha(w_{0}), that still gives a proper (L,H)(L,H)-colouring, and afterward recolouring vv with cc. Let G:=GvG^{\prime}:=G-v. For every wN(v)w\in N(v), remove from L(w)L(w) the colour cc^{\prime} matched with cc; call the resulting correspondence cover (L,H)(L^{\prime},H^{\prime}). Denote the restrictions to GG^{\prime} of α~\tilde{\alpha} and β\beta by α~\tilde{\alpha}^{\prime} and β\beta^{\prime}. By induction, since τ(G)<τ(G)\tau(G^{\prime})<\tau(G), we can recolour GG^{\prime} from α~\tilde{\alpha}^{\prime} to β\beta^{\prime} in at most n(G)+2τ(G)n(G^{\prime})+2\tau(G^{\prime}) steps. After this, we can finish by recolouring vv with β(v)\beta(v). Thus, dist(L,H)(α,β)2+dist(L,H)(α~,β)+12+n(G)+2τ(G)+1=2+n(G)1+2(τ(G)1)+1=n(G)+2τ(G)\operatorname{dist}_{(L,H)}(\alpha,\beta)\leq 2+\operatorname{dist}_{(L^{\prime},H^{\prime})}(\tilde{\alpha}^{\prime},\beta^{\prime})+1\leq 2+n(G^{\prime})+2\tau(G^{\prime})+1=2+n(G)-1+2(\tau(G)-1)+1=n(G)+2\tau(G). This proves the theorem. ∎

5 Distances in Reconfiguration Graphs of Trees

For each tree TT, it is straightforward to check that reconfiguration of correspondence colouring is no harder than reconfiguration of list colouring. Given a tree TT and a correspondence cover (L,H)(L,H), it is easy to construct a list-assignment LL^{\prime} for TT such that LL^{\prime}-colourings of TT are in bijection with (L,H)(L,H)-colourings of TT. We can do this by induction on n(G)n(G), by deleting a leaf vv and extending the list-assignment LL^{\prime}, given by hypothesis, for TvT-v. So, for simplicity, we phrase all results in this section only in terms of list colourings.

Theorem 17.

Fix a tree TT and a list-assignment LL with |L(v)|d(v)+2\lvert L(v)\rvert\geq d(v)+2 for every vV(T).v\in V(T). If α\alpha and β\beta are proper LL-colourings of TT, then the distance between α\alpha and β\beta in the reconfiguration graph 𝒞L(T)\mathcal{C}_{L}(T) is equal to μ(Dα,β)+vV𝟏α(v)β(v).\mu(D_{\alpha,\beta})+\sum_{v\in V}\mathbf{1}_{\alpha(v)\not=\beta(v)}.

Proof.

The lower bound holds by Observation 5. Now we prove that this lower bound is also an upper bound.

We use induction on n(T)n(T). The base case, n(T)=1n(T)=1, is trivial. Assume the theorem is true whenever TT has order at most s1s-1. We will prove it for an arbitrary tree TT on ss vertices. If α(v)=β(v)\alpha(v)=\beta(v) for some vV(T)v\in V(T), then we use induction on the components of TvT-v, with α(v)\alpha(v) removed from the lists of the neighbours of vv.

Suppose instead there are neighbours vv and ww for which vwA(Dα,β)\overrightarrow{vw}\not\in A(D_{\alpha,\beta}). Now TvwT-vw has two components; so let CvC_{v} and CwC_{w} be the components, respectively, containing vv and ww. By deleting α(w)\alpha(w) from L(v)L(v), we can first recolour CvC_{v}. Afterwards, we delete β(v)\beta(v) from L(w)L(w) and recolour CwC_{w}. By using the induction hypothesis (twice), we get the desired upper bound, since μ(Dα,β)=μ(Dα,β[Cv])+μ(Dα,β[Cw]).\mu(D_{\alpha,\beta})=\mu(D_{\alpha,\beta}[C_{v}])+\mu(D_{\alpha,\beta}[C_{w}]).

In the remaining case, we have two colour classes, say with colours 11 and 22, which need to be swapped. We pick a smallest vertex cover SS, which has size τ(Dα,β)=μ(Dα,β).\tau(D_{\alpha,\beta})=\mu(D_{\alpha,\beta}). Iteratively, we recolour every vertex vv in SS with a colour different from 11, 22, and all colours used to recolour neighbours of vv in SS. Note that vv has at most d(v)1d(v)-1 neighbours in SS, since otherwise SS would not be a minimal vertex cover. Since |L(v)|d(v)+2\lvert L(v)\rvert\geq d(v)+2, we can recolour as desired. Next, for each wV(T)Sw\in V(T)\setminus S, we recolour ww with β(w).\beta(w). Finally, we recolour each vSv\in S with β(v)\beta(v). This proves the induction step, which finishes the proof. ∎

Corollary 18.

If TT is a tree and LL is a list-assignment with |L(v)|d(v)+2\lvert L(v)\rvert\geq d(v)+2 for every vV(T)v\in V(T), then

diam𝒞L(T)n(T)+μ(T)3n(T)/2.\operatorname{diam}\mathcal{C}_{L}(T)\leq n(T)+\mu(T)\leq\lfloor 3n(T)/2\rfloor.
Proof.

Theorem 17 implies that diam𝒞L(T)n(T)+μ(T)\operatorname{diam}~{}\mathcal{C}_{L}(T)\leq n(T)+\mu(T), and the bound μ(T)n(T)2\mu(T)\leq\lfloor\frac{n(T)}{2}\rfloor holds trivially, since each edge of a matching saturates two vertices. ∎

Recall that [k][k] denotes {1,,k}\{1,\ldots,k\}. Thus, we write [k][k]-colouring to mean a kk-colouring from the colours [k][k].

Proposition 5.

For every tree TT and kΔ(T)+2k\geq\Delta(T)+2, we have

diam𝒞k(T)=n(T)+μ(T) and rad𝒞k(T)n(T)+μ(T)2.\operatorname{diam}\mathcal{C}_{k}(T)=n(T)+\mu(T)\mbox{ and }\operatorname{rad}\mathcal{C}_{k}(T)\geq n(T)+\left\lceil\frac{\mu(T)}{2}\right\rceil.
Proof.

The inequality diam𝒞k(T)n(T)+μ(T)\operatorname{diam}\mathcal{C}_{k}(T)\leq n(T)+\mu(T) holds by Corollary 18; and this inequality actually holds with equality, since the two [2][2]-colourings of TT are at distance exactly n(T)+μ(T)n(T)+\mu(T). That is, diam𝒞k(T)=n(T)+μ(T)\operatorname{diam}\mathcal{C}_{k}(T)=n(T)+\mu(T). Now we consider rad𝒞k(T)\operatorname{rad}\mathcal{C}_{k}(T). If we contract all edges of a maximum matching MM in TT, the result is also a tree, TT^{\prime}. When we 2-colour TT^{\prime}, one colour is used on at least half of the contracted edges. Denote the set of these contracted edges by EE^{\prime}. Note that EE^{\prime} is an induced matching in TT. Fix an arbitrary proper kk-colouring α\alpha of TT. Now we construct a proper kk-colouring β\beta of TT by swapping the colours on the endpoints of each edge vwEvw\in E^{\prime}, i.e., let β(v):=α(w)\beta(v):=\alpha(w) and β(w):=α(v)\beta(w):=\alpha(v) for every edge vwE.vw\in E^{\prime}. For every vertex vv not belonging to an edge in EE^{\prime}, choose a colour in [k][k] different from the colours already assigned (in β\beta) to neighbours of vv and different from α(v).\alpha(v). Now dist(α,β)n(T)+|E|n(T)+μ(T)2\operatorname{dist}(\alpha,\beta)\geq n(T)+\lvert E^{\prime}\rvert\geq n(T)+\left\lceil\frac{\mu(T)}{2}\right\rceil by Theorem 17. ∎

We construct trees TT for which certain colourings are more “central” in the reconfiguration graph than others. That is, we construct trees TT for which rad𝒞k(T)diam𝒞k(T)\operatorname{rad}\mathcal{C}_{k}(T)\not=\operatorname{diam}\mathcal{C}_{k}(T). We also study the maximum possible diameter and minimum possible radius of reconfiguration graphs of trees of given order, and the maximum possible difference of these quantities.

Proposition 6.

For every k4k\geq 4, the path PnP_{n} satisfies

diam𝒞k(Pn)=3n2 and rad𝒞k(Pn)=4n13.\operatorname{diam}\mathcal{C}_{k}(P_{n})=\left\lfloor\frac{3n}{2}\right\rfloor\mbox{ and }\operatorname{rad}\mathcal{C}_{k}(P_{n})=\left\lceil\frac{4n-1}{3}\right\rceil.

Furthermore, there exist nn-vertex trees TT, with maximum degree 3, such that for every k5k\geq 5 we have

diam𝒞k(T)=3n2 and rad𝒞k(T)=5n14.\operatorname{diam}\mathcal{C}_{k}(T)=\left\lfloor\frac{3n}{2}\right\rfloor\mbox{ and }\operatorname{rad}\mathcal{C}_{k}(T)=\left\lceil\frac{5n-1}{4}\right\rceil.

All such TT maximise, over all nn-vertex trees, the difference diam𝒞k(T)rad𝒞k(T)\operatorname{diam}\mathcal{C}_{k}(T)-\operatorname{rad}\mathcal{C}_{k}(T).

Proof.

Consider PnP_{n} with vertex set {v1,v2,,vn}\{v_{1},v_{2},\ldots,v_{n}\}. Fix k4k\geq 4, and let α\alpha and β\beta be the colourings with ranges [3][3] and [2][2], respectively, such that

α(vi)i(mod 3) and β(vi)i(mod 2)i[n].\alpha(v_{i})\equiv i\ (\mathrm{mod}\ 3)\mbox{~{}~{}~{} and~{}~{}~{} }\beta(v_{i})\equiv i\ (\mathrm{mod}\ 2)~{}~{}~{}~{}~{}\forall i\in[n].

Note that ecc(β)=3n2\operatorname{ecc}(\beta)=\lfloor\frac{3n}{2}\rfloor by Theorem 17, since switching colours 11 and 22 in β\beta requires 3n2\left\lfloor\frac{3n}{2}\right\rfloor recolouring steps. Furthermore, Corollary 18 implies that diam𝒞k(Pn)=ecc(β)=3n2\operatorname{diam}\mathcal{C}_{k}(P_{n})=\operatorname{ecc}(\beta)=\lfloor\frac{3n}{2}\rfloor. On the other hand, for every proper colouring γ\gamma, we know that Dα,γD_{\alpha,\gamma} cannot have a pair of bidirected edges of the form vivi+1v_{i}v_{i+1} and vi+2vi+3,v_{i+2}v_{i+3}, since α(vi)=α(vi+3).\alpha(v_{i})=\alpha(v_{i+3}). So μ(Dα,γ)n13\mu(D_{\alpha,\gamma})\leq\lceil\frac{n-1}{3}\rceil; hence, ecc(α)n+n13\operatorname{ecc}(\alpha)\leq n+\lceil\frac{n-1}{3}\rceil, which implies that rad𝒞k(Pn)4n13\operatorname{rad}\mathcal{C}_{k}(P_{n})\leq\lceil\frac{4n-1}{3}\rceil.

Next we prove the lower bound on rad𝒞k(Pn)\operatorname{rad}\mathcal{C}_{k}(P_{n}). For every colouring α\alpha of PnP_{n}, we can choose at least n13\lceil\frac{n-1}{3}\rceil disjoint edges such that swapping the colours in α\alpha on the endpoints of each edge (and possibly recolouring vertices not in any of these edges) yields another proper colouring. To see this, first select edge v1v2.v_{1}v_{2}. Whenever vivi+1v_{i}v_{i+1} has been selected, there exists j{i+3,i+4}j\in\{i+3,i+4\} for which α(vi)α(vj)\alpha(v_{i})\not=\alpha(v_{j}). Now add edge vj1vjv_{j-1}v_{j} to the set of selected edges. When our selection ends, the set EE^{\prime} contains at least n13\lceil\frac{n-1}{3}\rceil edges. We can now construct a [k][k]-colouring β\beta where β(w)=α(v)\beta(w)=\alpha(v) and β(v)=α(w)\beta(v)=\alpha(w) for every edge vwEvw\in E^{\prime}, and β(v)α(v)\beta(v)\neq\alpha(v) for every vV(Pn).v\in V(P_{n}). Theorem 17 gives dist(α,β)n+n13\operatorname{dist}(\alpha,\beta)\geq n+\lceil\frac{n-1}{3}\rceil.

Now let TnT_{n} be the nn-vertex comb graph; see Figure 3. Here TnT_{n} is formed from a path v1v2vtv_{1}v_{2}\ldots v_{t}, where t=n2t=\lceil\frac{n}{2}\rceil, by adding, for each viv_{i} (except vtv_{t} when nn is odd), an additional neighbour wiw_{i}. Since μ(Tn)=n2\mu(T_{n})=\left\lfloor\frac{n}{2}\right\rfloor, by Proposition 5 the diameter of 𝒞k(Tn)\mathcal{C}_{k}(T_{n}) is 3n2\lfloor\frac{3n}{2}\rfloor. Let α\alpha be the colouring shown and described in Figure 3.

1213212312132123α(vi)={1,for i3(mod 4)2,for i1(mod 4)3,for i0,2(mod 4) α(wi)={1,for i1,2(mod 4)2,for i3,4(mod 4).\alpha(v_{i})=\left.\begin{cases}1,&\text{for }i\equiv 3\ (\mathrm{mod}\ 4)\\ 2,&\text{for }i\equiv 1\ (\mathrm{mod}\ 4)\\ 3,&\text{for }i\equiv 0,2\ (\mathrm{mod}\ 4)\end{cases}\right.\mbox{~{}~{}~{}~{}}\alpha(w_{i})=\left.\begin{cases}1,&\text{for }i\equiv 1,2\ (\mathrm{mod}\ 4)\\ 2,&\text{for }i\equiv 3,4\ (\mathrm{mod}\ 4).\end{cases}\right.
Figure 3: The comb graph T16T_{16} with colouring α\alpha. The shaded region denotes four edges no two of which, for any colouring β\beta, yield disjoint bidirected edges of Dα,βD_{\alpha,\beta}.

For any proper colouring β\beta of TnT_{n}, note that μ(Dα,β)n14\mu(D_{\alpha,\beta})\leq\left\lceil\frac{n-1}{4}\right\rceil. This holds because, for each odd ii, the subgraph Dα,β[{vi,vi+1,vi+2,wi,wi+1}]D_{\alpha,\beta}[\{v_{i},v_{i+1},v_{i+2},w_{i},w_{i+1}\}] contains no two disjoint bidirected edges. Thus, ecc(α)5n14\operatorname{ecc}(\alpha)\leq\left\lceil\frac{5n-1}{4}\right\rceil, by Theorem 17. Since μ(Tn)=n2\mu(T_{n})=\left\lfloor\frac{n}{2}\right\rfloor, Proposition 5 implies that rad𝒞k(Tn)=n+n/2/2=5n14\operatorname{rad}\mathcal{C}_{k}(T_{n})=n+\left\lceil\left\lfloor n/2\right\rfloor/2\right\rceil=\left\lceil\frac{5n-1}{4}\right\rceil. Thus, diam𝒞k(Tn)rad𝒞k(Tn)=3n25n14=n4\operatorname{diam}\mathcal{C}_{k}(T_{n})-\operatorname{rad}\mathcal{C}_{k}(T_{n})=\left\lfloor\frac{3n}{2}\right\rfloor-\left\lceil\frac{5n-1}{4}\right\rceil=\left\lfloor\frac{n}{4}\right\rfloor. For every nn-vertex tree TT, Proposition 5 implies that diam𝒞k(T)rad𝒞k(T)(n+μ(T))(n+μ(T)2)=μ(T)2n4\operatorname{diam}\mathcal{C}_{k}(T)-\operatorname{rad}\mathcal{C}_{k}(T)\leq(n+\mu(T))-(n+\left\lceil\frac{\mu(T)}{2}\right\rceil)=\left\lfloor\frac{\mu(T)}{2}\right\rfloor\leq\left\lfloor\frac{n}{4}\right\rfloor. Thus, TnT_{n} maximises this difference. ∎

For list colourings, the difference diam𝒞L(T)rad𝒞L(T)\operatorname{diam}\mathcal{C}_{L}(T)-\operatorname{rad}\mathcal{C}_{L}(T) can be even larger.

Proposition 7.

For every tree TT, there exists a list-assignment LL, with |L(v)|d(v)+2\lvert L(v)\rvert\geq d(v)+2 for all vV(T)v\in V(T), such that diam𝒞L(G)=n(T)+μ(T)\operatorname{diam}\mathcal{C}_{L}(G)=n(T)+\mu(T) and rad𝒞L(G)=n(T)\operatorname{rad}\mathcal{C}_{L}(G)=n(T). These values are, respectively, the maximum and minimum possible over all such list-assignments LL.

Proof.

Choose a maximum matching MM of TT and fix LL such that

|L(v)L(w)|={0,when vwM2,when vwM.\lvert L(v)\cap L(w)\rvert=\left.\begin{cases}0,&\text{when }vw\not\in M\\ 2,&\text{when }vw\in M.\end{cases}\right.

Let α\alpha and β\beta be LL-colourings such that, for every edge vwMvw\in M, we have α(v)=β(w)\alpha(v)=\beta(w) and α(w)=β(v)\alpha(w)=\beta(v). For every vertex vv not saturated by MM, we pick α(v)\alpha(v) and β(v)\beta(v) to be arbitrary distinct colours in L(v)L(v). By Theorem 17, we have dist(α,β)=n(T)+μ(T)\operatorname{dist}(\alpha,\beta)=n(T)+\mu(T). Corollary 18 implies that diam𝒞L(T)=n(T)+μ(T)\operatorname{diam}\mathcal{C}_{L}(T)=n(T)+\mu(T). Further, this diameter is the maximum possible over all such list-assignments LL.

To show that rad𝒞L(T)n(T)\operatorname{rad}\mathcal{C}_{L}(T)\leq n(T), we construct an LL-colouring γ\gamma such that dist(γ,δ)n(T)\operatorname{dist}(\gamma,\delta)\leq n(T) for every LL-colouring δ\delta. Since |L(v)(wN(v)L(w))|32=1|L(v)\setminus\left(\cup_{w\in N(v)}L(w)\right)|\geq 3-2=1 for every vertex vv, we can form an LL-colouring γ\gamma such that every vV(T)v\in V(T) satisfies γ(v)wN(v)L(w)\gamma(v)\notin\cup_{w\in N(v)}L(w). For every proper colouring δ\delta, by our construction of γ\gamma, we can recolour δ\delta into γ\gamma greedily; hence, dist(δ,γ)n(T)\operatorname{dist}(\delta,\gamma)\leq n(T).

Now we show that radCL(T)n(T)\operatorname{rad}C_{L}(T)\geq n(T). For every LL-colouring γ\gamma, there exists an LL-colouring δ\delta such that δ(v)γ(v)\delta(v)\neq\gamma(v) for all vv. To see this, let L(v):=L(v)γ(v)L^{\prime}(v):=L(v)\setminus\gamma(v). Since |L(v)|d(v)+1|L^{\prime}(v)|\geq d(v)+1 for all vv, we form δ\delta by colouring GG greedily (in any order) from LL^{\prime}. Clearly, dist(γ,δ)n(G)\operatorname{dist}(\gamma,\delta)\geq n(G). Thus, rad𝒞L(T)=n(T)\operatorname{rad}\mathcal{C}_{L}(T)=n(T). In fact, this lower bound does not depend on the specific choice of LL, but only uses that |L(v)|d(v)+2|L(v)|\geq d(v)+2 for all vv. Thus, this radius is the minimum over all such list-assignments LL. ∎

6 Proving the List Conjecture for Complete Bipartite Graphs and Cactuses

A cactus is a connected graph in which each edge lies on at most one cycle. In this section, we prove the List Conjecture for all complete bipartite graphs and for all cactuses. Both proofs are by induction, and rely on the following helpful lemma.

Lemma 19.

Fix a positive integer bb. Let G=(V,E)G=(V,E) be a graph for which there exists a list-assignment LL and two proper LL-colourings α,β\alpha,\beta such that |L(v)|d(v)+b\lvert L(v)\rvert\geq d(v)+b for every vV(G)v\in V(G), and distL(α,β)>n(G)+μ(G)\operatorname{dist}_{L}(\alpha,\beta)>n(G)+\mu(G), but, for every proper induced subgraph of GG, no such list-assignment exists. Now for every partition V(G)=V1V2V(G)=V_{1}\cup V_{2} into two non-empty subsets of vertices, Dα,βD_{\alpha,\beta} has at least one arc from V1V_{1} to V2V_{2} and at least one arc from V2V_{2} to V1,V_{1}, i.e., Dα,βD_{\alpha,\beta} is strongly connected.

Proof.

Assume the lemma is false; by symmetry, assume Dα,βD_{\alpha,\beta} has no arcs from V1V_{1} to V2V_{2}. Let G1:=G[V1]G_{1}:=G[V_{1}] and G2:=G[V2]G_{2}:=G[V_{2}]. Let γ\gamma be the LL-colouring where γ(v):=β(v)\gamma(v):=\beta(v) when vV1v\in V_{1} and γ(v):=α(v)\gamma(v):=\alpha(v) when vV2v\in V_{2}.

For every vV1v\in V_{1}, let L(v):=L(v)\{α(w)wN(v)V2}L^{\prime}(v):=L(v)\backslash\{\alpha(w)\mid w\in N(v)\cap V_{2}\}. Note that |L(v)|dG1(v)+b\lvert L^{\prime}(v)\rvert\geq d_{G_{1}}(v)+b. Also note that still γ(v)L(v)\gamma(v)\in L^{\prime}(v) for every vV1v\in V_{1}; this is where we use that there is no arc from V1V_{1} to V2V_{2}. Since G1G_{1} is a proper induced subgraph of GG, by hypothesis diam𝒞L(G1)n(G1)+μ(G1)\operatorname{diam}\mathcal{C}_{L^{\prime}}(G_{1})\leq n(G_{1})+\mu(G_{1}); thus distL(α,γ)n(G1)+μ(G1)\operatorname{dist}_{L}(\alpha,\gamma)\leq n(G_{1})+\mu(G_{1}). Now for every vV2v\in V_{2}, let L(v):=L(v)\{β(w)wN(v)V1}L^{\prime}(v):=L(v)\backslash\{\beta(w)\mid w\in N(v)\cap V_{1}\}. Similarly to before, |L(v)|dG2(v)+b\lvert L^{\prime}(v)\rvert\geq d_{G_{2}}(v)+b, so diam𝒞L(G2)n(G2)+μ(G2)\operatorname{diam}\mathcal{C}_{L^{\prime}}(G_{2})\leq n(G_{2})+\mu(G_{2}); thus distL(γ,β)n(G2)+μ(G2)\operatorname{dist}_{L}(\gamma,\beta)\leq n(G_{2})+\mu(G_{2}). By the bounds above and the triangle inequality, we have

distL(α,β)\displaystyle\operatorname{dist}_{L}(\alpha,\beta) distL(α,γ)+distL(γ,β)\displaystyle\leq\operatorname{dist}_{L}(\alpha,\gamma)+\operatorname{dist}_{L}(\gamma,\beta)
n(G1)+μ(G1)+n(G2)+μ(G2)\displaystyle\leq n(G_{1})+\mu(G_{1})+n(G_{2})+\mu(G_{2})
n(G)+μ(G).\displaystyle\leq n(G)+\mu(G).

The last step uses that the union of a matching in G1G_{1} and a matching in G2G_{2} is a matching in GG. So GG is not a counterexample, and the lemma is true. ∎

It is helpful to note that an analogous statement holds for correspondence colouring. In fact, its proof is nearly identical to that given above. We use this observation in our proof of Theorem 27.

Using Lemma 19, we prove that the List Conjecture holds for all complete bipartite graphs.

Theorem 20.

The List Conjecture is true for all complete bipartite graphs.

Proof.

Suppose the theorem is false and choose a counterexample Kp,qK_{p,q} minimizing p+qp+q. By symmetry, we assume that pqp\leq q. Denote the parts of GG by UU and WW, with |U|=p|U|=p and |W|=q|W|=q.

By Lemma 19, we have α(U)β(W)\alpha(U)\subseteq\beta(W) and α(W)β(U)\alpha(W)\subseteq\beta(U). By swapping the roles of α\alpha and β\beta, we also have β(U)α(W)\beta(U)\subseteq\alpha(W) and β(W)α(U)\beta(W)\subseteq\alpha(U); thus, α(U)=β(W)\alpha(U)=\beta(W) and α(W)=β(U)\alpha(W)=\beta(U). Note that α(U)β(U)=α(U)α(W)=\alpha(U)\cap\beta(U)=\alpha(U)\cap\alpha(W)=\emptyset, because the graph is complete bipartite.

Case 1: |W||α(U)|+|β(U)|𝟏\bm{|W|\geq|\alpha(U)|+|\beta(U)|-1}. For each uUu\in U, we have |L(u)||W|+2|α(U)|+|β(U)|+1=|β(W)|+|α(W)|+1|L(u)|\geq|W|+2\geq|\alpha(U)|+|\beta(U)|+1=|\beta(W)|+|\alpha(W)|+1. First recolour each uUu\in U from L(u)(α(W)β(W))L(u)\setminus(\alpha(W)\cup\beta(W)). Now recolour each wWw\in W with β(w)\beta(w). Finally, recolour each uUu\in U with β(u)\beta(u). The number of steps that we use is at most 2|U|+|W|=n(G)+μ(G)2|U|+|W|=n(G)+\mu(G).

Case 2: |W||α(U)|+|β(U)|𝟐\bm{|W|\leq|\alpha(U)|+|\beta(U)|-2}. If |W|2|α(U)||W|\geq 2|\alpha(U)| and |W|2|β(U)||W|\geq 2|\beta(U)|, then 2|W|2|α(U)|+2|β(U)|2|W|\geq 2|\alpha(U)|+2|\beta(U)|, which contradicts the case. So assume, by symmetry, that |W|2|β(U)|1|W|\leq 2|\beta(U)|-1; if not, then simply interchange the roles of α\alpha and β\beta. Now, by Pigeonhole, there exists cβ(U)c\in\beta(U) such that |α1(c)|=|α1(c)W|1|\alpha^{-1}(c)|=|\alpha^{-1}(c)\cap W|\leq 1. So, recolour wα1(c)w\in\alpha^{-1}(c), if such ww exists, to avoid α(U){c}\alpha(U)\cup\{c\}, and then recolour every uβ1(c)u\in\beta^{-1}(c) with cc. Now we delete every uu such that β(u)=c\beta(u)=c, we delete cc from L(w)L(w) for every wWw\in W, and we finish on the resulting smaller graph G2G_{2} (with an assignment of smaller lists) by the minimality of GG. In total, the number of recolouring steps we use is at most |α1(c)|+|β1(c)|+|V(G2)|+μ(G2)|α1(c)|+|β1(c)|+(n(G)|β1(c)|)+(μ(G)|β1(c)|)n(G)+μ(G)|\alpha^{-1}(c)|+|\beta^{-1}(c)|+|V(G_{2})|+\mu(G_{2})\leq|\alpha^{-1}(c)|+|\beta^{-1}(c)|+(n(G)-|\beta^{-1}(c)|)+(\mu(G)-|\beta^{-1}(c)|)\leq n(G)+\mu(G). ∎

Using Lemma 19, we prove that the List Conjecture holds for all cycles.

Lemma 21.

Let GG be a cycle v1v2vnv_{1}v_{2}\cdots v_{n} and LL be a list-assignment such that, for all viv_{i}, we have |L(vi)|4\lvert L(v_{i})\rvert\geq 4. If α\alpha and β\beta are proper LL-colourings of GG, then we can recolour GG from α\alpha to β\beta in at most 3n/2\lfloor 3n/2\rfloor steps.

Proof.

By Lemma 19, with b=2b=2, if dist(α,β)>3n/2\operatorname{dist}(\alpha,\beta)>\lfloor 3n/2\rfloor, then Dα,βD_{\alpha,\beta} is strongly connected. This implies that Dα,βD_{\alpha,\beta} contains a directed cycle CnC_{n} as a subdigraph. (If Dα,βD_{\alpha,\beta} contains a bidirected path PP, then |α(V(P))β(V(P))|=2|\alpha(V(P))\cup\beta(V(P))|=2. So, if PP is spanning, then its order must be even, to avoid a conflict between the colours of its endpoints. But then Dα,βD_{\alpha,\beta} contains an additional arc, so Dα,βD_{\alpha,\beta} contains a directed cycle, as claimed.)

If Dα,βD_{\alpha,\beta} is a bidirected cycle, then nn must be even, as in the previous paragraph, and 3n2\frac{3n}{2} recolouring steps suffice. Suppose instead that Dα,βD_{\alpha,\beta} is precisely a directed cyle. When n=3n=3, we recolour one vertex vv in a colour absent from α(V(G))β(V(G))\alpha(V(G))\cup\beta(V(G)), recolour the other two vertices in order (to match β\beta), and finally recolour vv with β(v)\beta(v). When n4n\geq 4, we recolour one vertex vv in a colour absent from α(v)α(N(v))\alpha(v)\cup\alpha(N(v)). Now we can recolour correctly (in order) all vertices of GG except for vv and one of its neighbours. Correctly colouring these final two vertices takes at most 3 recolouring steps. Now we are done, since 1+(n2)+33n/21+(n-2)+3\leq\lfloor 3n/2\rfloor.

In the remaining case, we have n4n\geq 4 and some directed edge is adjacent to a bidirected edge. Without loss of generality, we assume v1v2,v2v3,v2v1A(Dα,β)\overrightarrow{v_{1}v_{2}},\overrightarrow{v_{2}v_{3}},\overrightarrow{v_{2}v_{1}}\in A(D_{\alpha,\beta}) but v3v2A(Dα,β)\overrightarrow{v_{3}v_{2}}\notin A(D_{\alpha,\beta}). Recolour v1v_{1} with a colour cc different from α(v1)\alpha(v_{1}), α(v2)\alpha(v_{2}), and α(vn)\alpha(v_{n}). Now delete cc from L(v2)L(v_{2}) and L(vn)L(v_{n}), and let G:=Gv1G^{\prime}:=G-v_{1}. By Theorem 17, we can recolour GG^{\prime} from α\alpha to β\beta (both restricted to GG^{\prime}) using at most (n1)+n22(n-1)+\left\lfloor\frac{n-2}{2}\right\rfloor recolouring steps. Here we use that v3v2\overrightarrow{v_{3}v_{2}} is not an arc, so μ(Dα,β[V(G)])n22\mu(D_{\alpha,\beta}[V(G^{\prime})])\leq\left\lfloor\frac{n-2}{2}\right\rfloor. Finally, we recolour v1v_{1} with β(v1)\beta(v_{1}). ∎

Using Lemmas 19 and 21, we can prove that the List Conjecture is true for every cactus.

Theorem 22.

The List Conjecture is true for every cactus.

Proof.

Instead assume the theorem is false. Let GG be a counterexample minimizing n(G)n(G) and let α\alpha and β\beta be LL-colourings with dist(α,β)>n(G)+μ(G)\operatorname{dist}(\alpha,\beta)>n(G)+\mu(G). Every proper induced subgraph HH of GG is a disjoint union of cactuses, say H1,,HrH_{1},\ldots,H_{r}; since n(H)=i=1rn(Hi)n(H)=\sum_{i=1}^{r}n(H_{i}) and μ(H)=i=1rμ(Hi)\mu(H)=\sum_{i=1}^{r}\mu(H_{i}), the List Conjecture must hold for HH, by the minimality of GG. Lemma 19 implies that Dα,βD_{\alpha,\beta} is strongly connected. And Lemma 21 implies that GG is not a cycle. The following claim restricts the structure of GG, α\alpha, and β\beta.

Claim 23.

If vV(G)v\in V(G) and μ(Gv)=μ(G)1\mu(G-v)=\mu(G)-1, then L(v)wN(v){α(w),β(w)}L(v)\subseteq\cup_{w\in N(v)}\{\alpha(w),\beta(w)\}. In particular, each block of GG containing vv gives rise to exactly two arcs in Dα,βD_{\alpha,\beta} incident to vv, and vv has no adjacent leaf in GG.

Proof.

Instead assume there exists cL(v)wN(v){α(w),β(w)}c\in L(v)\setminus\cup_{w\in N(v)}\{\alpha(w),\beta(w)\}. Recolour vv with cc, let G:=GvG^{\prime}:=G-v, let L(w):=L(w){c}L^{\prime}(w):=L(w)\setminus\{c\} for all wN(v)w\in N(v), and otherwise let L(w):=L(w)L^{\prime}(w):=L(w). Denote by α\alpha^{\prime} and β\beta^{\prime} the restrictions to GG^{\prime} of α\alpha and β\beta. Since GG is minimal, we can recolour GG^{\prime} from α\alpha^{\prime} to β\beta^{\prime}, using LL^{\prime}, in at most n(G)+μ(G)=n(G)1+μ(G)1n(G^{\prime})+\mu(G^{\prime})=n(G)-1+\mu(G)-1 recolouring steps. We finish by recolouring vv with β(v)\beta(v). This proves the first statement.

Since Dα,βD_{\alpha,\beta} is strongly connected, each block BB of GG containing vv gives rise to at least two arcs in Dα,βD_{\alpha,\beta} incident to vv, one in each direction. Since GG is a cactus, each such BB contains at most 2 neighbours of vv. So at least half of the neighbours of vv are coloured β(v)\beta(v) by α\alpha, and also at least half of the neighbours of vv are coloured α(v)\alpha(v) by β\beta. Therefore |wN(v){α(w),β(w)}|2(d(v)/2+1))=d(v)+2|\cup_{w\in N(v)}\{\alpha(w),\beta(w)\}|\leq 2\cdot(d(v)/2+1))=d(v)+2. If any such BB is either K2K_{2} or gives rise to at least three arcs in Dα,βD_{\alpha,\beta} incident with vv, then |wN(v){α(w),β(w)}|d(v)+1<|L(v)||\cup_{w\in N(v)}\{\alpha(w),\beta(w)\}|\leq d(v)+1<\lvert L(v)\rvert, which contradicts the first statement. This proves the second statement. ∎

It is easy to check that every neighbour vv of a leaf ww in GG satisfies μ(Gv)=μ(G)1\mu(G-v)=\mu(G)-1. Thus, Claim 23 implies that GG has no leaf vertex. This implies that all endblocks of GG are cycles. Let CsC_{s} be an endblock. Denote its vertices by w1,,wsw_{1},\ldots,w_{s}, such that wsw_{s} is the unique cut-vertex in the block.

Claim 24.

The endblock CsC_{s} is not an even cycle.

Proof.

Assume instead that CsC_{s} is an even cycle. It is easy to check that μ(Gwi)=μ(G)1\mu(G-w_{i})=\mu(G)-1 whenever ii is even. That is, every maximum matching saturates wiw_{i} whenever ii is even. By Claim 23, every wiw_{i} is incident in Dα,βD_{\alpha,\beta} to exactly two arcs arising from CsC_{s}. Since Dα,βD_{\alpha,\beta} is strongly connected, this implies that Dα,β[V(Cs)]D_{\alpha,\beta}[V(C_{s})] is a directed cycle; by symmetry, we assume that it is oriented as wsw1,w1w2,,ws1ws\overrightarrow{w_{s}w_{1}},\overrightarrow{w_{1}w_{2}},\ldots,\overrightarrow{w_{s-1}w_{s}}.

We first recolour wsw_{s} with a colour absent from α(N[ws])=(xN(ws)α(x))β(ws1)\alpha(N[w_{s}])=\left(\cup_{x\in N(w_{s})}\alpha(x)\right)\cup\beta(w_{s-1}). Now we recolour each wiw_{i} to β(wi)\beta(w_{i}), with ii decreasing from s1s-1 to 22. Let G:=G{w1,,ws1}G^{\prime}:=G-\{w_{1},\ldots,w_{s-1}\}, let L(ws):=L(ws){α(w1),β(ws1)}L^{\prime}(w_{s}):=L(w_{s})\setminus\{\alpha(w_{1}),\beta(w_{s-1})\}, and otherwise let L(v):=L(v)L^{\prime}(v):=L(v). Since GG is minimal, we can recolour GG^{\prime} from its current colouring to β\beta (restricted to GG^{\prime}) using at most n(G)+μ(G)(n(G)(s1))+(μ(G)s2+1)n(G^{\prime})+\mu(G^{\prime})\leq(n(G)-(s-1))+(\mu(G)-\frac{s}{2}+1) steps. Finally, recolour w1w_{1} to β(w1)\beta(w_{1}). The total number of recolouring steps is at most 1+(s2)+(n(G)(s1)+μ(G)s2+1)+1=n(G)+μ(G)+2s21+(s-2)+(n(G)-(s-1)+\mu(G)-\frac{s}{2}+1)+1=n(G)+\mu(G)+2-\frac{s}{2}. Since s4s\geq 4, this is at most n(G)+μ(G)n(G)+\mu(G). ∎

Claim 25.

The endblock CsC_{s} is not an odd cycle.

Proof.

Assume instead that CsC_{s} is an odd cycle. If Dα,β[Csws]D_{\alpha,\beta}[C_{s}-w_{s}] contains fewer than s2\left\lfloor\frac{s}{2}\right\rfloor disjoint digons (for example, this is true when s=3s=3), then we can easily finish, as follows. We recolour wsw_{s} in a colour different from α(N(ws))β({w1,ws1})\alpha(N(w_{s}))\cup\beta(\{w_{1},w_{s-1}\}). This is possible because β(ws)\beta(w_{s}) is used by α\alpha on at least two neighbours of wsw_{s}, since Dα,βD_{\alpha,\beta} is strongly connected. By Theorem 17, we recolour CswsC_{s}-w_{s} to β\beta in at most (s1)+s21(s-1)+\left\lfloor\frac{s}{2}\right\rfloor-1 steps, and we then recolour G\(Csws)G\backslash(C_{s}-w_{s}) to β\beta in at most n(G)(s1)+μ(G)s2n(G)-(s-1)+\mu(G)-\left\lfloor\frac{s}{2}\right\rfloor steps. Thus, we recolour GG from α\alpha to β\beta in at most n(G)+μ(G)n(G)+\mu(G) steps.

Now we consider the other case, when wiwi+1w_{i}w_{i+1} is a digon of Dα,βD_{\alpha,\beta} for every odd ii with i<si<s. Since Dα,β[Cs]D_{\alpha,\beta}[C_{s}] is strongly connected, we assume Dα,βD_{\alpha,\beta} contains arc wiwi+1w_{i}w_{i+1} for every 0is10\leq i\leq s-1. Similar to the proof of Lemma 21, here Dα,β[Cs]D_{\alpha,\beta}[C_{s}] cannot contain a spanning bidirected path, since ss is odd. This implies, for some even jj, that Dα,βD_{\alpha,\beta} does not contain wj+1wjw_{j+1}w_{j}. Now we will recolour some set of wiw_{i}’s to reach a new colouring α~\tilde{\alpha}, such that Dα~,β[Cs]D_{\tilde{\alpha},\beta}[C_{s}] is either acyclic or contains a single directed cycle, ws2ws1w_{s-2}w_{s-1}. From α~\tilde{\alpha}, we can recolour G(Csws)G\setminus(C_{s}-w_{s}) by the minimality of GG, and afterward finish on CswsC_{s}-w_{s}. The details follow.

First suppose that β(ws)=α(ws1)\beta(w_{s})=\alpha(w_{s-1}). For every ii such that either (a) i<ji<j and ii is odd or (b) i>ji>j and ii is even, recolour wiw_{i} with a colour different from those in {α(wi1),α(wi),α(wi+1)}\{\alpha(w_{i-1}),\alpha(w_{i}),\alpha(w_{i+1})\}. Here we take the indices modulo ss, i.e., w0=wsw_{0}=w_{s}. If β(ws)α(ws1)\beta(w_{s})\neq\alpha(w_{s-1}), then we recolour the same set of wiw_{i}’s, except for ws1w_{s-1}. Let G:=G{w1,,ws1}G^{\prime}:=G-\{w_{1},\ldots,w_{s-1}\}. By the minimality of GG, we can recolour GG^{\prime} in at most n(G)+μ(G)=(n(G)(s1))+(μ(G)s2)n(G^{\prime})+\mu(G^{\prime})=(n(G)-(s-1))+(\mu(G)-\left\lfloor\frac{s}{2}\right\rfloor) steps. In the case that β(ws)α(ws1)\beta(w_{s})\neq\alpha(w_{s-1}), we now recolour ws1w_{s-1} with a colour different from those in {α(ws2),α(ws1),β(ws)}\{\alpha(w_{s-2}),\alpha(w_{s-1}),\beta(w_{s})\}. For every odd ii such that j+1is2j+1\leq i\leq s-2, we recolour wiw_{i} with β(wi)\beta(w_{i}). Next, for every even ii such that 2ij2\leq i\leq j, we recolour wiw_{i} with β(wi).\beta(w_{i}). Finally, the remaining vertices in CsC_{s} can also be recoloured with their colour in β\beta. This process uses at most n(G)+μ(G)n(G)+\mu(G) steps. ∎

Recall that every endblock of GG is a cycle, as observed following Claim 23. Thus, Claims 24 and 25 yield a contradiction, which proves the theorem. ∎

7 Proving the Correspondence Conjecture for Cactuses, Subcubic Graphs, and Graphs with Low Maximum Average Degree

The maximum average degree of a graph GG, denoted mad(G)\operatorname{mad}(G), is the maximum, taken over all subgraphs HH, of the average degree of HH. That is, mad(G):=maxHG2|E(H)|/|V(H)|\operatorname{mad}(G):=\max_{H\subseteq G}2|E(H)|/|V(H)|. Let d1(v)d^{1}(v) denote the number of neighbours ww of vv such that d(w)=1d(w)=1. In this section we prove the Correspondence Conjecture for all subcubic graphs, cactuses, and graphs GG with mad(G)<2.4\operatorname{mad}(G)<2.4. To do so, we often implicitly use the following observation. We omit its proof, which is easy.

Observation 26.

If GG is a graph with a minimum vertex cover SS and vSv\in S, then τ(Gv)=τ(G)1\tau(G-v)=\tau(G)-1.

Theorem 27.

Let GG be a graph and (L,H)(L,H) a correspondence cover for GG such that, for all vV(G)v\in V(G) we have |L(v)|d(v)+2|L(v)|\geq d(v)+2. Now diam𝒞(L,H)(G)n(G)+τ(G)\operatorname{diam}\mathcal{C}_{(L,H)}(G)\leq n(G)+\tau(G) if at least one of the following holds.

  • (a)

    Δ(G)3\Delta(G)\leq 3.

  • (b)

    GG is a cactus.

  • (c)

    mad(G)<2.4\operatorname{mad}(G)<2.4.

Proof.

Fix GG satisfying (a), (b), or (c) and (L,H)(L,H) as in the theorem. Fix arbitrary (L,H)(L,H)-colourings α\alpha and β\beta. Our proof is by a double induction: primarily on τ(G)\tau(G) and secondarily on |V(G)||V(G)|. The base case, τ(G)=0\tau(G)=0, is trivial, since GG is an independent set and we can greedily recolour GG from α\alpha to β\beta. For the induction step, assume τ(G)1\tau(G)\geq 1.

We first show that GG contains either (i) a vertex vv such that d(v)3d(v)\leq 3 and vv lies in some minimum vertex cover or (ii) a vertex vv such that d(v)4d(v)\geq 4 and d(v)d1(v)2d(v)-d^{1}(v)\leq 2. Next we show how to proceed by induction in each of cases (i) and (ii).

If GG satisfies (a), then clearly GG contains an instance of (i). Suppose GG satisfies (b), and consider an endblock BB of GG. If BB is a cycle, then BB contains adjacent non-cut vertices, vv and ww; note that d(v)=d(w)=2d(v)=d(w)=2. Further, every vertex cover of GG contains vv or ww, so GG contains (i). Assume instead that every endblock of GG is an edge.

Form a graph JJ with a vertex for every block in GG, where two vertices of JJ are adjacent if their corresponding blocks share a vertex in GG. Now the endpoints of a diameter in JJ correspond with pendent edges in GG. For the leaf vv corresponding to such a pendent edge, call its neighbour ww. Clearly ww lies in some minimum vertex cover of GG. If d(w)3d(w)\leq 3, then GG contains (i). Otherwise, d(w)d1(w)2d(w)-d^{1}(w)\leq 2, since by the choice of ww at most one block incident to ww is not an endblock; so GG contains (ii). This concludes the case that GG satisfies (b).

Assume instead that GG satisfies (c). Suppose, to reach a contradiction, that GG contains neither (i) nor (ii). Form GG^{\prime} from GG by deleting all vertices with degree at most 1. We will show that 2|E(G)|/|V(G)|2.42|E(G^{\prime})|/|V(G^{\prime})|\geq 2.4, a contradiction. Note that GG^{\prime} has minimum degree (at least) 2. If an arbitrary vertex vv satisfies dG(v)=2d_{G^{\prime}}(v)=2, then dG(v)=2d_{G}(v)=2, since otherwise vv is an instance of (i) or (ii) in GG. Further, if dG(v)=2d_{G^{\prime}}(v)=2, then d(w)3d(w)\geq 3 for all vwE(G)vw\in E(G^{\prime}), since otherwise vv or ww is an instance of (i) in GG. Now we will reach a contradiction with discharging. Give each vertex vv in GG^{\prime} charge ch(v):=dG(v)\operatorname{ch}(v):=d_{G^{\prime}}(v). We use a single discharging rule: Each 2-vertex in GG^{\prime} takes 0.20.2 from each neighbour. Now each 2-vertex vv in GG^{\prime} finishes with charge ch(v)=2+2(0.2)=2.4\operatorname{ch}^{*}(v)=2+2(0.2)=2.4. And each vertex vv with dG(v)3d_{G^{\prime}}(v)\geq 3 finishes with charge ch(v)dG(v)0.2dG(v)=0.8dG(v)2.4\operatorname{ch}^{*}(v)\geq d_{G^{\prime}}(v)-0.2d_{G^{\prime}}(v)=0.8d_{G^{\prime}}(v)\geq 2.4. This yields the desired contradiction, which proves that GG contains either (i) or (ii) if mad(G)<2.4\operatorname{mad}(G)<2.4.

Now we prove the induction step in the cases that GG contains (i) or (ii).

Suppose GG contains (i). Suppose at most one neighbour, say ww, of vv has α(w)\alpha(w) matched with β(v)\beta(v). Now recolour ww with a colour not matched to α(x)\alpha(x), for every xN(w)x\in N(w), and not matched to β(v)\beta(v). Next, recolour vv with β(v)\beta(v), and proceed on GvG-v by induction (with the colour matched to β(v)\beta(v) deleted from the lists of all vertices in N(v)N(v)). So instead assume there exist at least two such neighbours, say w1w_{1} and w2w_{2}. By interchanging the roles of α\alpha and β\beta (and repeating this argument), we see that also there exist two neighbours, say x1x_{1} and x2x_{2}, such that β(x1)\beta(x_{1}) and β(x2)\beta(x_{2}) are matched to α(v)\alpha(v). The total number of colours in L(v)L(v) matched to α(w)\alpha(w) or β(w)\beta(w) for some wN(v)w\in N(v) is at most 2d(v)2<d(v)+22d(v)-2<d(v)+2, since d(v)3d(v)\leq 3. Hence, there exists cL(v)c\in L(v) that is not matched to α(w)\alpha(w) or β(w)\beta(w) for all wN(v)w\in N(v). Recolour vv with cc. Proceed on GvG-v by induction, with that colour that cc is matched to deleted from the list L(w)L(w) for each wN(v)w\in N(v). After finishing on GvG-v, recolour vv with β(v)\beta(v). The number of steps we use is at most 1+n(Gv)+τ(Gv)+1=n(G)+τ(G)1+n(G-v)+\tau(G-v)+1=n(G)+\tau(G).

Suppose instead GG contains (ii). The proof is almost the same as for (i). If there exists wN(v)w\in N(v) such that d(w)=1d(w)=1 and α(v)\alpha(v) is not matched to β(w)\beta(w), then simply recolour ww with β(w)\beta(w), and proceed by induction; here we use the secondary induction hypothesis, since possibly τ(Gw)=τ(G)\tau(G-w)=\tau(G). So assume that no such ww exists. By interchanging the roles of α\alpha and β\beta, we also assume that β(v)\beta(v) is matched to α(w)\alpha(w) for each wN(v)w\in N(v) with d(w)=1d(w)=1. Further, if vv has a non-leaf neighbour ww, then at least one such neighbour has α(v)\alpha(v) matched to β(w)\beta(w) and at least one (possibly the same one) has α(w)\alpha(w) matched to β(v)\beta(v). (This follows from the correspondence colouring analog of Lemma 19.) But now there exists (v,i)L(v)(v,i)\in L(v) such that for all wN(v)w\in N(v), colour (v,i)(v,i) is not matched to either α(w)\alpha(w) or β(w)\beta(w). Recolour vv to (v,i)(v,i), and let G:=GvG^{\prime}:=G-v. Form (L,H)(L^{\prime},H) from (L,H)(L,H) by deleting from L(w)L(w), for every wN(v)w\in N(v), the colour matched with (v,i)(v,i). By induction, we can recolour GG^{\prime} from α\alpha to β\beta (both restricted to GG^{\prime}) using at most n(G)+τ(G)=n(G)+τ(G)2n(G^{\prime})+\tau(G^{\prime})=n(G)+\tau(G)-2 steps. Finally, recolour vv to β(v)\beta(v). This uses at most n(G)+τ(G)n(G)+\tau(G) steps, which finishes the proof. ∎

Corollary 28.

The List Conjecture holds for all bipartite graphs GG with Δ(G)3\Delta(G)\leq 3.

Proof.

When GG is bipartite, recall that τ(G)=μ(G)\tau(G)=\mu(G). ∎

8 Concluding Remarks

In this paper, we give evidence for both the List Conjecture and the Correspondence Conjecture. We also give evidence for an affirmative answer to Question 1. The List Conjecture and Question 1 would together determine the precise diameter for 𝒞Δ(G)+2(G)\mathcal{C}_{\Delta(G)+2}(G) and, as such, a precise bound when Cereceda’s Conjecture is restricted to regular graphs. So we explicitly conjecture the following.

Conjecture 3 (Regular Cereceda’s Conjecture).

For a dd-regular graph GG, if k=d+2k=d+2, then diam𝒞k(G)=n(G)+μ(G)\operatorname{diam}~{}\mathcal{C}_{k}(G)=n(G)+\mu(G).

In Theorem 10 we prove, when |L(v)|d(v)+2\lvert L(v)\rvert\geq d(v)+2 for all vV(G)v\in V(G), that diam𝒞L(G)n(G)+2μ(G)2n(G)\operatorname{diam}\mathcal{C}_{L}(G)\leq n(G)+2\mu(G)\leq 2n(G). In a similar vein, it would be interesting to show that diam𝒞L(G)\operatorname{diam}\mathcal{C}_{L}(G) is linear when |L(v)|mad(G)+2.\lvert L(v)\rvert\geq\lceil\operatorname{mad}(G)+2\rceil. The following conjecture can be viewed as a “balanced” version of the List Conjecture.

Conjecture 4 (Mad Colouring Reconfiguration Conjecture).

For a graph GG with mad(G)=d\operatorname{mad}(G)=d, if LL is a list-assignment such that |L(v)|d+2\lvert L(v)\rvert\geq\lceil d+2\rceil for every vV(G)v\in V(G), then diam𝒞L(G)=Od(n).\operatorname{diam}\mathcal{C}_{L}(G)=O_{d}(n).

Feghali [11] proved that if ϵ>0\epsilon>0 and kd+1+ϵk\geq d+1+\epsilon, then diam𝒞k(G)=Od(n(2lnn)d)\operatorname{diam}\mathcal{C}_{k}(G)=O_{d}(n(2\ln{n})^{d}). Conjecture 4 aims to prove a similar result for list colouring, with one more colour available for each vertex, and with a somewhat stronger bound on diameter. All planar graphs GG have mad(G)<6\operatorname{mad}(G)<6, and all triangle-free planar graphs GG have mad(G)<4\operatorname{mad}(G)<4, so we note that Conjecture 4 would imply stronger forms of [10, Conjecture 22]. If GG is regular, then mad(G)=degen(G)\operatorname{mad}(G)=\operatorname{degen}(G), so Conjecture 4 is true by Theorem 10. Conjecture 4 is also related to444If mad(G)\operatorname{mad}(G) is not an integer, then degen(G)+3mad(G)+2\operatorname{degen}(G)+3\leq\left\lceil\operatorname{mad}(G)+2\right\rceil. But if GG has a regular subgraph of degree mad(G)\operatorname{mad}(G), then this inequality fails. a conjecture of Bartier et al. [1, Conjecture 1.6] that graphs GG with degen(G)=d\operatorname{degen}(G)=d satisfy diam𝒞d+3(G)=Od(n)\operatorname{diam}\mathcal{C}_{d+3}(G)=O_{d}(n).

We do not know if Conjecture 4 might be true with a linear bound of the form O(n)O(n) instead of Od(n).O_{d}(n). But we do note, for this version of the conjecture and every constant cc, that no bound of the form n(G)+cμ(G)n(G)+c\mu(G) can hold. This is shown by the star K1,3c+3=({u}{w1,,w3c+3},E)K_{1,3c+3}=(\{u\}\cup\{w_{1},\ldots,w_{3c+3}\},E) and the colourings α,β\alpha,\beta with α(u)=1,β(u)=4\alpha(u)=1,\beta(u)=4, β(wi)=1\beta(w_{i})=1 and α(wi)=1+i/(c+1)\alpha(w_{i})=1+\left\lceil i/(c+1)\right\rceil. Here mad(G)+2=4\lceil\operatorname{mad}(G)+2\rceil=4 and dist(α,β)>n+c.\operatorname{dist}(\alpha,\beta)>n+c.

For a correspondence cover (L,H)(L,H) such that |L(v)|=d(v)+2|L(v)|=d(v)+2 for all vV(G)v\in V(G), in Theorem 1(ii) we proved that diam𝒞(L,H)(G)n(G)+2τ(G)\operatorname{diam}~{}\mathcal{C}_{(L,H)}(G)\leq n(G)+2\tau(G). We view this as modest evidence that Cereceda’s Conjecture might remain true in the more general context of correspondence colourings.

8.1 Open Problems

The three focuses of this paper are the List Conjecture, the Correspondence Conjecture, and (to a lesser extent) Question 1. All of these remain open. However, each of them seems rather hard. So, to motivate further research, below we identify some specific classes of graphs for which we believe that each conjecture may be approachable. We begin with some graph classes for which it would be particularly interesting to make further progress on the List Conjecture.

  1. (1)(1)

    Complete rr-partite graphs for each r3r\geq 3. (We know the List Conjecture is true for both complete graphs and complete bipartite graphs, so this is a natural common generalization.)

  2. (2)(2)

    Bipartite graphs, not necessarily complete.

  3. (3)(3)

    Outerplanar graphs and, more generally, planar graphs.

  4. (4)(4)

    Subcubic graphs. (We already proved the Correspondence Conjecture for this class; nonetheless, the List Conjecture remains open.)

Conversely, the Correspondence Conjecture remains open for the following basic graph classes.

  1. (1)(1)

    Complete graphs. (The argument of Bonamy and Bousquet for complete graphs directly yields the List Conjecture, but does not yield the Correspondence Conjecture.)

  2. (2)(2)

    Complete bipartite graphs.

  3. (3)(3)

    Bipartite graphs, not necessarily complete. (This would imply the List Conjecture for the same class.)

  4. (4)(4)

    Outerplanar graphs and, more generally, planar graphs.

Finally, we stress that it will be interesting to improve on Theorems 1 and 2. For example, find the smallest ϵ\epsilon such that for every graph GG and list-assignment LL with |L(v)|d(v)+2|L(v)|\geq d(v)+2 for all vV(G)v\in V(G), we have diam𝒞L(G)n(G)+(1+ϵ)μ(G)\operatorname{diam}\mathcal{C}_{L}(G)\leq n(G)+(1+\epsilon)\mu(G). We proved ϵ=1\epsilon=1 suffices, while ϵ=0\epsilon=0 would resolve the List Conjecture.

Acknowledgement

We thank the organisers of the online workshop Graph Reconfiguration of the Sparse Graphs Coalition555For more information, visit https://sparse-graphs.mimuw.edu.pl/doku.php., where this project started. We also thank the two referees for their careful reading and valuable feedback.

References

  • [1] V. Bartier, N. Bousquet, C. Feghali, M. Heinrich, B. Moore, and T. Pierron. Recoloring planar graphs of girth at least five. SIAM J. Discrete Math., 37(1):332–350, 2023. doi:10.1137/21M1463598.
  • [2] B. Bollobás, P. A. Catlin, and P. Erdős. Hadwiger’s conjecture is true for almost every graph. European J. Combin., 1(3):195–199, 1980. doi:10.1016/S0195-6698(80)80001-1.
  • [3] M. Bonamy and N. Bousquet. Recoloring graphs via tree decompositions. European J. Combin., 69:200–213, 2018, arXiv:1403.6386.
  • [4] P. Bonsma and L. Cereceda. Finding paths between graph colourings: PSPACE-completeness and superpolynomial distances. Theoret. Comput. Sci., 410(50):5215--5226, 2009. doi:10.1016/j.tcs.2009.08.023.
  • [5] N. Bousquet, L. Feuilloley, M. Heinrich, and M. Rabie. Short and local transformations between (Δ\Delta+1)-colorings. 2022, arXiv:2203.08885.
  • [6] N. Bousquet and M. Heinrich. A polynomial version of Cereceda’s conjecture. J. Combin. Theory Ser. B, 155:1--16, 2022. doi:10.1016/j.jctb.2022.01.006.
  • [7] L. Cereceda. Mixing graph colourings. PhD thesis, The London School of Economics and Political Science (LSE), 2007.
  • [8] L. Cereceda, J. van den Heuvel, and M. Johnson. Connectedness of the graph of vertex-colourings. Discrete Math., 308(5-6):913--919, 2008. doi:10.1016/j.disc.2007.07.028.
  • [9] M. Delcourt and L. Postle. Reducing linear Hadwiger’s conjecture to coloring small graphs. 2021, arXiv:2108.01633.
  • [10] Z. Dvořák and C. Feghali. A Thomassen-type method for planar graph recoloring. European J. Combin., 95:Paper No. 103319, 12, 2021. doi:10.1016/j.ejc.2021.103319.
  • [11] C. Feghali. Reconfiguring colorings of graphs with bounded maximum average degree. J. Combin. Theory Ser. B, 147:133--138, 2021. doi:10.1016/j.jctb.2020.11.001.
  • [12] M. Jerrum. A very simple algorithm for estimating the number of kk-colorings of a low-degree graph. Random Structures Algorithms, 7(2):157--165, 1995. doi:10.1002/rsa.3240070205.
  • [13] M. Molloy. The list chromatic number of graphs with small clique number. J. Combin. Theory Ser. B, 134:264--284, 2019. doi:10.1016/j.jctb.2018.06.007.
  • [14] C. M. Mynhardt and S. Nasserasr. Reconfiguration of colourings and dominating sets in graphs. In 50 years of combinatorics, graph theory, and computing, Discrete Math. Appl. (Boca Raton), pages 171--191. CRC Press, Boca Raton, FL, 2020. doi:10.1201/9780429280092-10.
  • [15] N. Nishimura. Introduction to reconfiguration. Algorithms (Basel), 11(4):Paper No. 52, 25, 2018.
  • [16] J. van den Heuvel. The complexity of change. In Surveys in combinatorics 2013, volume 409 of London Math. Soc. Lecture Note Ser., pages 127--160. Cambridge Univ. Press, Cambridge, 2013, arXiv:1312.2816.