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

\definecolor

BrickRedrgb0.714,0.196,0.11 \definecolorBluergb0.176,0.184,0.573 Univ Lyon, CNRS, ENS de Lyon, Université Claude Bernard Lyon 1, LIP UMR5668, France and \urlhttps://kristofhuszar.github.io/[email protected]://orcid.org/0000-0002-5445-5057Supported by the French National Research Agency through the 3IA Côte d’Azur (ANR-19-P3IA-0002), AlgoKnot (ANR-20-CE48-0007), GrR (ANR-18-CE40-0032), and TWIN-WIDTH (ANR-21-CE48-0014) projects. A previous version of this manuscript was written during the author’s visit at Institut Henri Poincaré (IHP). The author acknowledges the hospitality of IHP (UAR 839 CNRS–Sorbonne Université) and LabEx CARMIN (ANR-10-LABX-59-01). School of Mathematics and Statistics F07, The University of Sydney, NSW 2006 Australia and \urlhttps://sites.google.com/view/jonathan-spreer/[email protected]://orcid.org/0000-0001-6865-9483Supported by the Australian Research Council’s Discovery funding scheme (project no. DP220102588). \CopyrightKristóf Huszár and Jonathan Spreer \ccsdesc[500]Mathematics of computing Geometric topology \ccsdesc[300]Theory of computation Problems, reductions and completeness \ccsdesc[300]Theory of computation Fixed parameter tractability

Acknowledgements.
We are grateful to Arnaud de Mesmay and to Clément Maria for their interest in our work and for inspiring discussions. We thank the anonymous referees for several suggestions to improve this article. \nolinenumbers\hideLIPIcs\EventEditorsErin W. Chambers and Joachim Gudmundsson \EventNoEds2 \EventLongTitle39th International Symposium on Computational Geometry (SoCG 2023) \EventShortTitleSoCG 2023 \EventAcronymSoCG \EventYear2023 \EventDateJune 12–15, 2023 \EventLocationDallas, Texas, USA \EventLogosocg-logo.pdf \SeriesVolume258 \ArticleNoXX

On the Width of Complicated JSJ Decompositions

Kristóf Huszár    Jonathan Spreer
Abstract

Motivated by the algorithmic study of 3-dimensional manifolds, we explore the structural relationship between the JSJ decomposition of a given 3-manifold and its triangulations. Building on work of Bachman, Derby-Talbot and Sedgwick, we show that a “sufficiently complicated” JSJ decomposition of a 3-manifold enforces a “complicated structure” for all of its triangulations. More concretely, we show that, under certain conditions, the treewidth (resp. pathwidth) of the graph that captures the incidences between the pieces of the JSJ decomposition of an irreducible, closed, orientable 3-manifold \mathscr{M} yields a linear lower bound on its treewidth tw()\operatorname{tw}(\mathscr{M}) (resp. pathwidth pw()\operatorname{pw}(\mathscr{M})), defined as the smallest treewidth (resp. pathwidth) of the dual graph of any triangulation of \mathscr{M}.

We present several applications of this result. We give the first example of an infinite family of bounded-treewidth 3-manifolds with unbounded pathwidth. We construct Haken 3-manifolds with arbitrarily large treewidth—previously the existence of such 3-manifolds was only known in the non-Haken case. We also show that the problem of providing a constant-factor approximation for the treewidth (resp. pathwidth) of bounded-degree graphs efficiently reduces to computing a constant-factor approximation for the treewidth (resp. pathwidth) of 3-manifolds.

keywords:
computational 3-manifold topology, fixed-parameter tractability, generalized Heegaard splittings, JSJ decompositions, pathwidth, treewidth, triangulations

1 Introduction

Manifolds in geometric topology are often studied through the following two-step process. Given a piecewise linear dd-dimensional manifold \mathscr{M}, first find a “suitable” triangulation of it, i.e., a decomposition of \mathscr{M} into dd-simplices with “good” combinatorial properties. Then apply algorithms on this triangulation to reveal topological information about \mathscr{M}.

The work presented in this article is motivated by this process in dimension d=3d=3. Here every manifold can be triangulated [43] and questions about them typically admit algorithmic solutions [34, 38, 53].111In higher dimensions none of these statements is true in general. See, e.g., [40], [42] or [45, Section 7]. At the same time, the feasibility of a particular computation can greatly depend on structural properties of the triangulation in use. Over the past decade, this phenomenon was recognized and exploited in various settings, leading to fixed-parameter tractable (FPT) algorithms for several problems in low-dimensional topology, some of which are even known to be NP-hard [14, 15, 16, 17, 18].222See [8] for an FPT algorithm checking tightness of (weak) pseudomanifolds in arbitrary dimensions. Although these algorithms have exponential running time in the worst case, for input triangulations with dual graph of bounded treewidth they always terminate in polynomial (in most cases, linear) time.333The running times are given in terms of the size of the input triangulation, i.e., its number of tetrahedra. Moreover, some of them have implementations that are highly effective in practice, providing useful tools for researchers in low-dimensional topology [12, 13].

The theoretical efficiency of the aforementioned FPT algorithms crucially depends on the assumption that the dual graph of the input triangulation has small treewidth. To understand their scope, it is thus instructive to consider the treewidth tw()\operatorname{tw}(\mathscr{M}) of a compact 33-manifold \mathscr{M}, defined as the smallest treewidth of the dual graph of any triangulation of \mathscr{M}. Indeed, the relation between the treewidth and other quantities associated with 3-manifolds has recently been investigated in various contexts [25, 26, 27, 28, 41]. For instance, in [28] together with Wagner we have shown that the treewidth of a non-Haken 3-manifold is always bounded below in terms of its Heegaard genus. Combined with earlier work of Agol [1]—who constructed non-Haken 3-manifolds with arbitrary large Heegaard genus—this implies the existence of 3-manifolds with arbitrary large treewidth. Despite the fact that, asymptotically, most triangulations of most 3-manifolds must have dual graph of large treewidth [28, Appendix A], this collection described by Agol has remained, to this date, the only known family of 3-manifolds with arbitrary large treewidth.

The main result

In this work we unravel new structural links between the triangulations of a given 3-manifold and its JSJ decomposition [30, 31, 32]. Employing the machinery of generalized Heegaard splittings [49], the results developed in [28], and building on the work of Bachman, Derby-Talbot and Sedgwick [5, 6], we show that, under suitable conditions, the dual graph of any triangulation of a given 3-manifold \mathscr{M} inherits structural properties from the decomposition graph that encodes the incidences between the pieces of the JSJ decomposition of \mathscr{M}. More precisely, in \Crefsec:main we prove the following theorem.

Theorem 1.1 (Width inheritance).

For any closed, orientable and irreducible 33-manifold \mathscr{M} with sufficiently complicated444The notion of “sufficiently complicated” under which we establish Theorem 1.1 is discussed in \Crefsec:main. torus gluings in its JSJ decomposition 𝒟\mathscr{D}, the treewidth and pathwidth of \mathscr{M} and that of the decomposition graph Γ(𝒟)\Gamma(\mathscr{D}) of 𝒟\mathscr{D} satisfy

tw(Γ(𝒟))18(tw()+1)\operatorname{tw}(\Gamma(\mathscr{D}))\leq 18\cdot(\operatorname{tw}(\mathscr{M})+1) (1)

and

pw(Γ(𝒟))4(3pw()+1).\operatorname{pw}(\Gamma(\mathscr{D}))\leq 4\cdot(3\operatorname{pw}(\mathscr{M})+1). (2)
An algorithmic construction

Much work in 3-dimensional topology has been devoted to the study of 3-manifolds constructed by pasting together simpler pieces along their boundary surfaces via “sufficiently complicated” gluing maps, and to understand how different decompositions of the same 3-manifold interact under various conditions, see, e.g., [4, 5, 6, 7, 35, 39, 47, 52]. \Crefthm:jsj-width allows us to leverage these results to construct 3-manifolds, where we have tight control over the treewidth and pathwidth of their triangulations [25].

By combining \Crefthm:jsj-width and [4, Theorem 5.4] (cf. [6, Appendix]), in \Crefsec:const we prove the following result.

Theorem 1.2.

There is a polynomial-time algorithm that, given an nn-node graph GG with maximum node-degree Δ\Delta, produces a triangulation 𝒯G\mathscr{T}_{G} of a closed 33-manifold G\mathscr{M}_{G}, such that

  1. 1.

    the triangulation 𝒯G\mathscr{T}_{G} contains OΔ(pw(G)n)O_{\Delta}(\operatorname{pw}(G)\cdot n) tetrahedra,555Similar to the standard big-O notation, OΔ(x)O_{\Delta}(x) means “a quantity bounded above by xx times a constant depending on Δ\Delta.” To ensure that \colorlipicsGray3a is satisfied, but not necessarily \colorlipicsGray3b, OΔ(tw(G)n)O_{\Delta}(\operatorname{tw}(G)\cdot n) tetrahedra suffice.

  2. 2.

    the JSJ decomposition 𝒟\mathscr{D} of G\mathscr{M}_{G} satisfies Γ(𝒟)=G\Gamma(\mathscr{D})=G, and

  3. 3.

    there exist universal constants c,c>0c,c^{\prime}>0, such that

    1. (a)

      (c/Δ)tw(G)tw(G)18(tw(G)+1)(c/\Delta)\operatorname{tw}(\mathscr{M}_{G})\leq\operatorname{tw}(G)\leq 18\cdot(\operatorname{tw}(\mathscr{M}_{G})+1), and

    2. (b)

      (c/Δ)pw(G)pw(G)4(3pw(G)+1)(c^{\prime}/\Delta)\operatorname{pw}(\mathscr{M}_{G})\leq\operatorname{pw}(G)\leq 4\cdot(3\operatorname{pw}(\mathscr{M}_{G})+1).

Applications

In \Crefsec:appl we present several applications of \Crefthm:construction. First, we construct a family of bounded-treewidth 3-manifolds with unbounded pathwidth (\Crefcor:unbounded-pathwidth). Second, we exhibit Haken 3-manifolds with arbitrary large treewidth (\Crefcor:unbounded-treewidth). To our knowledge, no such families of 3-manifolds had been known before. Third, we show that the problem of providing a constant-factor approximation for the treewidth (resp. pathwidth) of bounded-degree graphs reduces in polynomial time to computing a constant-factor approximation for the treewidth (resp. pathwidth) of 3-manifolds (\Crefcor:hardness-treewidth). This reduction, together with previous results [46, 55, 56], suggests that this problem may be computationally hard.

Proof 1.3 (Outline of the proof of Theorem 1.1).

We now give a preview of the proof of our main result. As the arguments for showing (1) and (2) are analogous, we only sketch the proof of (1). To show that tw(Γ(𝒟))18(tw()+1)\operatorname{tw}(\Gamma(\mathscr{D}))\leq 18(\operatorname{tw}(\mathscr{M})+1), we start with a triangulation 𝒯\mathscr{T} of \mathscr{M} whose dual graph has minimal treewidth, i.e., tw(Γ(𝒯))=tw()\operatorname{tw}(\Gamma(\mathscr{T}))=\operatorname{tw}(\mathscr{M}). Following [28, Section 6], we construct from 𝒯\mathscr{T} a generalized Heegaard splitting \mathscr{H} of \mathscr{M}, together with a sweep-out Σ={Σx:xH}\Sigma=\{\Sigma_{x}:x\in H\} along a tree HH, such that the genus of any level surface Σx\Sigma_{x} is at most 18(tw(Γ(𝒯))+1)18\cdot(\operatorname{tw}(\Gamma(\mathscr{T}))+1). If \mathscr{H} is not already strongly irreducible, we repeatedly perform weak reductions until we get a strongly irreducible generalized Heegaard splitting \mathscr{H}^{\prime} with associated sweep-out Σ={Σx:xH}\Sigma^{\prime}=\{\Sigma^{\prime}_{x}:x\in H\} along the same tree HH. Crucially, weak reductions do not increase the genera of level surfaces [49, Section 5.2], thus 18(tw(Γ(𝒯))+1)18\cdot(\operatorname{tw}(\Gamma(\mathscr{T}))+1) is still an upper bound on those in Σ\Sigma^{\prime}. Now, by the assumption of the JSJ decomposition of \mathscr{M} being “sufficiently complicated,” each JSJ torus can be isotoped in \mathscr{M} to coincide with a connected component of some thin level of \mathscr{H}^{\prime}. This implies that, after isotopy, each level set Σx\Sigma^{\prime}_{x} is incident to at most 18(tw(Γ(𝒯))+1)+118\cdot(\operatorname{tw}(\Gamma(\mathscr{T}))+1)+1 JSJ pieces of \mathscr{M}. Sweeping along HH, we can construct a tree decomposition of Γ(𝒟)\Gamma(\mathscr{D}) where each bag contains at most 18(tw(Γ(𝒯))+1)+118\cdot(\operatorname{tw}(\Gamma(\mathscr{T}))+1)+1 nodes of Γ(𝒟)\Gamma(\mathscr{D}). Hence tw(Γ(𝒟))18(tw(Γ(𝒯))+1)=18(tw()+1)\operatorname{tw}(\Gamma(\mathscr{D}))\leq 18\cdot(\operatorname{tw}(\Gamma(\mathscr{T}))+1)=18\cdot(\operatorname{tw}(\mathscr{M})+1).

Organization of the paper

In \Crefsec:prelims we review the necessary background on graphs and 3-manifolds. \Crefsec:gen contains a primer on generalized Heegaard splittings, which provides us with the indispensable machinery for proving our main result (\Crefthm:jsj-width) in \Crefsec:main. In \Crefsec:const we describe the algorithmic construction of 3-manifolds that “inherit” their combinatorial width from that of their JSJ decomposition graph (\Crefthm:construction). Then, in \Crefsec:appl we present the aforementioned applications of this construction (\Crefcor:unbounded-pathwidth,cor:unbounded-treewidth,cor:hardness-treewidth). The paper is concluded with a discussion and some open questions in \Crefsec:discussion. Selected results from 3-manifold topology we rely on are collected in the \hyperref[app]Appendix.

2 Preliminaries

2.1 Graphs

A (multi)graph G=(V,E)G=(V,E) is a finite set V=V(G)V=V(G) of nodes666Throughout this paper we use the terms edge and vertex to refer to an edge or vertex in a 33-manifold triangulation, whereas the terms arc and node denote an edge or vertex in a graph, respectively. together with a multiset E=E(G)E=E(G) of unordered pairs of not necessarily distinct nodes, called arcs. The degree dvd_{v} of a node vVv\in V equals the number of arcs containing it, where loop arcs are counted twice. GG is kk-regular, if dv=kd_{v}=k for all vVv\in V. A tree is a connected graph with nn nodes and n1n-1 arcs.

A tree decomposition of a graph G=(V,E)G=(V,E) is a pair (𝒳={Bi:iI},T=(I,F))(\mathscr{X}=\{B_{i}:i\in I\},T=(I,F)) with bags BiVB_{i}\subseteq V and a tree T=(I,F)T=(I,F), such that 1. iIBi=V\bigcup_{i\in I}B_{i}=V(node coverage), 2. for all arcs {u,v}E\{u,v\}\in E, there exists iIi\in I such that {u,v}Bi\{u,v\}\subseteq B_{i} (arc coverage), and 3. for all vVv\in V, Tv={iI:vBi}T_{v}=\{i\in I:v\in B_{i}\} spans a connected sub-tree of TT (sub-tree property). The width of a tree decomposition equals maxiI|Bi|1\max_{i\in I}|B_{i}|-1, and the treewidth tw(G)\operatorname{tw}(G) is the smallest width of any tree decomposition of GG. Replacing all occurrences of “tree” with “path” in the definition of treewidth yields the notion of pathwidth pw(G)\operatorname{pw}(G). We have tw(G)pw(G)\operatorname{tw}(G)\leq\operatorname{pw}(G).

2.2 Manifolds

A dd-dimensional manifold is a topological space \mathscr{M}, where each point xx\in\mathscr{M} has a neighborhood homeomorphic to d\mathbb{R}^{d} or to the closed upper half-space {(x1,,xd)d:xd0}\{(x_{1},\ldots,x_{d})\in\mathbb{R}^{d}:x_{d}\geq 0\}. The latter type of points of \mathscr{M} constitute the boundary \partial\mathscr{M} of \mathscr{M}. A compact manifold is said to be closed if it has an empty boundary. We consider manifolds up to homeomorphism (“continuous deformations”) and write 12\mathscr{M}_{1}\cong\mathscr{M}_{2} for homeomorphic manifolds 1\mathscr{M}_{1} and 2\mathscr{M}_{2}.

3-Manifolds and surfaces

The main objects of study in this paper are 3-dimensional manifolds, or 33-manifolds for short. Here we give a brief introduction to 33-manifolds tailored to our purposes. We refer the reader to [51] (and the references therein) for more details. All 3-manifolds and surfaces encountered in this article are compact and orientable. We let 𝕊g\mathbb{S}_{g} denote the closed, connected, orientable surface of genus gg. We also refer to the dd-dimensional torus and sphere as 𝕋d\mathbb{T}^{d} and 𝕊d\mathbb{S}^{d}, respectively (hence 𝕊0=𝕊2\mathbb{S}_{0}=\mathbb{S}^{2} and 𝕊1=𝕋2\mathbb{S}_{1}=\mathbb{T}^{2}). The genus g(𝒮)g(\mathcal{S}) of a (not necessarily connected) surface 𝒮\mathcal{S} is defined to be the sum of the genera of its connected components.

Triangulations and the treewidth of 3-manifolds

A triangulation 𝒯\mathscr{T} of a given 33-manifold \mathscr{M} is a finite collection of abstract tetrahedra glued together along pairs of their triangular faces, such that the resulting space is homeomorphic to \mathscr{M}. Unpaired triangles comprise a triangulation of the boundary of \mathscr{M}. Note that the face gluings may also identify several tetrahedral edges (or vertices) in a single edge (or vertex) of 𝒯\mathscr{T}. Every compact 3-manifold admits a triangulation [43] (cf. [10]). Given a triangulation 𝒯\mathscr{T}, its dual graph Γ(𝒯)\Gamma(\mathscr{T}) is the multigraph whose nodes correspond to the tetrahedra in 𝒯\mathscr{T}, and arcs to face gluings (\Creffig:tetrahedra).

Refer to caption

Figure 1: (i) Example of a triangulation 𝒯\mathscr{T} with two tetrahedra Δ1\Delta_{1} and Δ2\Delta_{2}, and three face gluing maps φ1\varphi_{1}, φ2\varphi_{2} and φ3\varphi_{3}. The map φ1\varphi_{1} is specified to be Δ1(123)Δ2(103)\Delta_{1}(123)\longleftrightarrow\Delta_{2}(103). (ii) The dual graph Γ(𝒯)\Gamma(\mathscr{T}) of the triangulation 𝒯\mathscr{T}. Reproduced from [26, Figure 1].

For a compact 3-manifold \mathscr{M}, its treewidth tw()\operatorname{tw}(\mathscr{M}) (resp. pathwidth pw()\operatorname{pw}(\mathscr{M})) is defined as the smallest treewidth (resp. pathwidth) of the dual graph of any triangulation of \mathscr{M}.

Incompressible surfaces and essential disks

Given a 33-manifold \mathscr{M}, a surface 𝒮\mathcal{S}\subset\mathscr{M} is said to be properly embedded in \mathscr{M} if 𝒮\partial\mathcal{S}\subset\partial\mathscr{M} and (𝒮𝒮)()(\mathcal{S}\setminus\partial\mathcal{S})\subset(\mathscr{M}\setminus\partial\mathscr{M}). Given a properly embedded surface 𝒮\mathcal{S}\subset\mathscr{M}, an embedded disk DD\subset\mathscr{M} with int(D)𝒮=\operatorname{int}(D)\cap\mathcal{S}=\emptyset and D𝒮\partial D\subset\mathcal{S} a curve not bounding a disk on 𝒮\mathcal{S} is called a compressing disk. If such a disk exists, 𝒮\mathcal{S} is said to be compressible in \mathscr{M}, otherwise – and if 𝒮\mathcal{S} is not a 22-sphere – it is called incompressible. A 33-manifold \mathscr{M} is said to be irreducible, if every embedded 22-sphere bounds a 33-ball in \mathscr{M}. A disk DD\subset\mathscr{M} properly embedded in a 3-manifold \mathscr{M} is called inessential if it cuts off a 3-ball from \mathscr{M}, otherwise DD is called essential. A compact, orientable, irreducible 33-manifold is called Haken if it contains an orientable, properly embedded, incompressible surface, and otherwise is referred to as non-Haken.

Heegaard splittings of closed 3-manifolds

A handlebody \mathcal{H} is a connected 33-manifold homeomorphic to a thickened graph. The genus g()g(\mathcal{H}) of \mathcal{H} is defined as the genus of its boundary surface \partial\mathcal{H}. A Heegaard splitting of a closed, orientable 3-manifold \mathscr{M} is a decomposition =𝒮\mathscr{M}=\mathcal{H}\cup_{\mathcal{S}}\mathcal{H}^{\prime} where \mathcal{H} and \mathcal{H}^{\prime} are homeomorphic handlebodies with =\mathcal{H}\cup\mathcal{H}^{\prime}=\mathscr{M} and ===𝒮\mathcal{H}\cap\mathcal{H}^{\prime}=\partial\mathcal{H}=\partial\mathcal{H}^{\prime}=\mathcal{S} called the splitting surface. Introduced in [24], the Heegaard genus 𝔤()\mathfrak{g}\left(\mathscr{M}\right) of \mathscr{M} is the smallest genus g(𝒮)g(\mathcal{S}) over all Heegaard splittings of .\mathscr{M}.

The JSJ decomposition

A central result by Jaco–Shalen [30, 31] and Johannson [32] asserts that every closed, irreducible and orientable 3-manifold \mathscr{M} admits a collection 𝐓\mathbf{T} of pairwise disjoint embedded, incompressible tori, where each piece of the complement 𝐓\mathscr{M}\setminus\mathbf{T} is either Seifert fibered777See [51, Section 3.7] or [23, p. 18] for an introduction to Seifert fibered spaces (cf. \Crefapp:jsj). or atoroidal888An irreducible 3-manifold \mathscr{M} is called atoroidal if every incompressible torus in \mathscr{M} is boundary-parallel.. A minimal such collection of tori is unique up to isotopy and gives rise to the so-called JSJ decomposition (or torus decomposition) of \mathscr{M} [23, Theorem 1.9]. We refer to this collection of incompressible tori as the JSJ tori of \mathscr{M}. The graph with nodes the JSJ pieces, and an arc for each JSJ torus (with endpoints the two nodes corresponding to its two adjacent pieces) is called the dual graph Γ(𝒟)\Gamma(\mathscr{D}) of the JSJ decomposition 𝒟\mathscr{D} of \mathscr{M}.

3 Generalized Heegaard Splittings

A Heegaard splitting of a closed 3-manifold is a decomposition into two identical handlebodies along an embedded surface. Introduced by Scharlemann and Thompson [50], a generalized Heegaard splitting of a compact 3-manifold \mathscr{M} (possibly with boundary) is a decomposition of \mathscr{M} into several pairs of compression bodies along a family of embedded surfaces, subject to certain rules. Following [49, Chapters 2 and 5],999For an open-access version, see [48, Sections 3.1 and 4]. here we give an overview of this framework.

3.1 Compression bodies and forks

A compression body 𝒞\mathcal{C} is a 3-manifold with boundary obtained by the following procedure.

  1. 1.

    Consider the thickening 𝒮×[0,1]\mathcal{S}\times[0,1] of a closed, orientable, possibly disconnected surface 𝒮\mathcal{S},

  2. 2.

    optionally attach some 11-handles, each being of the form D×[0,1]D\times[0,1] (thickened edge, where the disk DD is the cross-section), to 𝒮×{1}\mathcal{S}\times\{1\} along D×{0}D×{1}D\times\{0\}\cup D\times\{1\}, and

  3. 3.

    optionally fill in some collection of 22-sphere components of 𝒮×{0}\mathcal{S}\times\{0\} with 33-balls.

We call 𝒞=𝒮×{0}{filled-in 2-sphere components}\partial_{-}\mathcal{C}=\mathcal{S}\times\{0\}\setminus\{\text{filled-in 2-sphere components}\} the lower boundary of 𝒞\mathcal{C} and +𝒞=𝒞𝒞\partial_{+}\mathcal{C}=\partial\,\mathcal{C}\setminus\partial_{-}\mathcal{C} its upper boundary. By construction, g(+𝒞)g(𝒞)g(\partial_{+}\mathcal{C})\geq g(\partial_{-}\mathcal{C}). Note that, if 𝒞\partial_{-}\mathcal{C} is empty, then 𝒞\mathcal{C} is a handlebody. We allow compression bodies to be disconnected.

A fork, more precisely an nn-fork, FF is a tree with n+2n+2 nodes V(F)={ρ,γ,τ1,,τn}V(F)=\{\rho,\gamma,\tau_{1},\ldots,\tau_{n}\}, where ρ\rho, called the root, is of degree n+1n+1, and the other nodes are leaves. One of them, denoted γ\gamma, is called the grip, and the remaining leaves τ1,,τn\tau_{1},\ldots,\tau_{n} are called tines. A fork can be regarded as an abstraction of a connected compression body 𝒞\mathcal{C}, where the grip corresponds to +𝒞\partial_{+}\mathcal{C} and each tine corresponds to a connected component of 𝒞\partial_{-}\mathcal{C}, see, e.g., \Creffig:compbody.

{subfigure}

[b]0.375 \begin{overpic}[scale={.9}]{compbody-construction-2} \put(3.5,40.0){\small{$h_{1}$}} \put(75.0,47.0){\small{$h_{2}$}} \put(56.0,64.0){\small{$\mathbb{T}^{2}\times[0,1]$}} \put(41.0,-6.0){\small{$\mathbb{S}_{2}\times[0,1]$}} \end{overpic} {subfigure}[b]0.3 Refer to caption {subfigure}[b]0.26 \begin{overpic}[scale={.9}]{fork} \put(5.0,91.0){\small{$\partial_{+}\mathcal{C}$}} \put(46.5,91.0){\small{$\leftarrow\textit{grip}$}} \put(46.5,64.0){\small{$\leftarrow\textit{root}$}} \put(48.0,48.5){\small{\raisebox{-6.5pt}{\rotatebox{40.0}{$\leftarrow$}}\,{tine}}} \put(2.0,25.0){\small{$\partial^{(1)}_{-}\mathcal{C}$}} \put(46.0,25.0){\small{$\partial^{(2)}_{-}\mathcal{C}$}} \put(55.0,13.0){\small{\rotatebox{90.0}{$=$}}} \put(11.0,13.0){\small{\rotatebox{90.0}{$=$}}} \put(1.0,3.0){\small{$\mathbb{T}^{2}\times\{0\}$}} \put(41.25,3.0){\small{$\mathbb{S}_{2}\times\{0\}$}} \end{overpic}

Figure 2: A compression body 𝒞\mathcal{C}
Figure 3: 𝒞\mathcal{C} after some isotopy
Figure 4: A 22-fork representing 𝒞\mathcal{C}
Figure 5: The compression body 𝒞\mathcal{C} is obtained by first thickening the disconnected surface 𝕋2𝕊2\mathbb{T}^{2}\cup\mathbb{S}_{2} to (𝕋2𝕊2)×[0,1](\mathbb{T}^{2}\cup\mathbb{S}_{2})\times[0,1], then attaching two 11-handles (h1h_{1} and h2h_{2}) between 𝕋2×{1}\mathbb{T}^{2}\times\{1\} and 𝕊2×{1}\mathbb{S}_{2}\times\{1\}. For the lower boundary of 𝒞\mathcal{C} we have 𝒞=(𝕋2𝕊2)×{0}\partial_{-}\mathcal{C}=(\mathbb{T}^{2}\cup\mathbb{S}_{2})\times\{0\}, and for its upper boundary +𝒞=𝒞𝒞𝕊4.\partial_{+}\mathcal{C}=\partial\,\mathcal{C}\setminus\partial_{-}\mathcal{C}\cong\mathbb{S}_{4}.
Non-faithful forks

In certain situations (notably, in the proof of \Crefthm:jsj-width, cf. \Crefsec:main) it is useful to also take a simplified view on a generalized Heegaard splitting. To that end, one may represent several compression bodies by a single non-faithful fork, where the grip and the tines may correspond to collections of boundary components. To distinguish faithful forks from non-faithful ones, we color the roots of the latter with gray, see \Creffig:faithful-forks.


\nolinenumbers
\begin{overpic}[scale={.9}]{faithful-forks} \put(57.5,29.85){\Large$\leadsto$} \end{overpic}\captionof

figureTwo faithful forks bundled into a non-faithful fork. The colors show the grouping of the tines.

\begin{overpic}[scale={.9}]{compbody-spine} \put(47.25,40.25){$\Gamma$} \put(-10.0,1.0){$\mathcal{C}$} \put(102.0,1.0){$\mathcal{H}$} \end{overpic}\captionof

figure\nolinenumbersSpines of a compression body 𝒞\mathcal{C} and of a handlebody \mathcal{H}.

Spines and sweep-outs of compression bodies

A graph Γ\Gamma embedded in a compression body 𝒞\mathcal{C} is called a spine of 𝒞\mathcal{C}, if every node of Γ\Gamma that is incident to 𝒞\partial_{-}\mathcal{C} is of degree one, and 𝒞(Γ𝒞)+𝒞×(0,1]\mathcal{C}\setminus(\Gamma\cup\partial_{-}\mathcal{C})\cong\partial_{+}\mathcal{C}\times(0,1], see \Creffig:compbody-spine. Assume first that 𝒞\partial_{-}\mathcal{C}\neq\emptyset. A sweep-out of 𝒞\mathcal{C} along an interval, say [1,1][-1,1], is a continuous map f:𝒞[1,1]f\colon\mathcal{C}\rightarrow[-1,1], such that f1(±1)=±𝒞f^{-1}(\pm 1)=\partial_{\pm}\mathcal{C},

f1(t){+𝒞,ift(0,1),Γ𝒞,ift=0,and𝒞,ift(1,0).\displaystyle f^{-1}(t)\cong\begin{cases}\partial_{+}\mathcal{C},&\text{if}\phantom{-}t\in(0,1),\\ \Gamma\cup\partial_{-}\mathcal{C},&\text{if}\phantom{-}t=0,~{}\text{and}\\ \partial_{-}\mathcal{C},&\text{if}\phantom{-}t\in(-1,0).\end{cases} (3)

Writing Σt=f1(t)\Sigma_{t}=f^{-1}(t), we get a 1-parameter family of surfaces (except for Σ0\Sigma_{0}) “sweeping through” 𝒞\mathcal{C}. For handlebodies, the definition of a sweep-out is similar, but it “ends at 0” with the spine: f1(1)=f^{-1}(1)=\partial\mathcal{H}, f1(t)f^{-1}(t)\cong\partial\mathcal{H} if t(0,1)t\in(0,1), and f1(0)=Γf^{-1}(0)=\Gamma, see \Creffig:sweep-out.

{subfigure}

[b]0.48 \begin{overpic}[scale={.9}]{sweep-out-compbody-2} \put(0.0,32.75){\small{$\phantom{-}1\phantom{-}$}} \put(0.0,17.5){\small{$\phantom{-}0\phantom{-}$}} \put(0.0,3.25){\small{$-1\phantom{-}$}} \put(88.0,32.75){\small{$\Sigma_{1}$}} \put(88.0,17.5){\small{$\Sigma_{0}$}} \put(88.0,3.25){\small{$\Sigma_{-1}$}} \end{overpic} {subfigure}[b]0.48 \begin{overpic}[scale={.9}]{sweep-out-hbody} \put(0.0,32.75){\small{$\phantom{1/}1$}} \put(0.0,17.5){\small{$1/2$}} \put(0.0,3.25){\small{$\phantom{1/}0$}} \put(88.0,32.75){\small{$\Sigma_{1}$}} \put(88.0,17.5){\small{$\Sigma_{1/2}$}} \put(88.0,3.25){\small{$\Sigma_{0}$}} \end{overpic}

Figure 6: Sweep-outs of the compression body 𝒞\mathcal{C} and of the handlebody \mathcal{H} shown in \Creffig:compbody-spine.
Remark 3.1 (Sweep-out along a fork).

It is straightforward to adapt the definition of sweep-out in a way that, instead of an interval, we sweep a compression body 𝒞\mathcal{C} along a (faithful or non-faithful) fork FF. Such a sweep-out is a continuous map f:𝒞Ff\colon\mathcal{C}\rightarrow\|F\| (where F\|F\| denotes a geometric realization of FF) that satisfies very similar requirements to those in (3), however, the components of the lower boundary 𝒞\partial_{-}\mathcal{C} may appear in different level sets, see \Creffig:sweep-out-fork.

{subfigure}

[b]0.48 \begin{overpic}[scale={.9}]{sweep-out-compbody-gray} \put(0.0,32.75){\small{$\phantom{-}1\phantom{-}$}} \put(0.0,17.5){\small{$\phantom{-}0\phantom{-}$}} \put(0.0,3.25){\small{$-1\phantom{-}$}} \put(88.0,32.75){\small{$\Sigma_{1}$}} \put(88.0,17.5){\small{$\Sigma_{0}$}} \put(88.0,3.25){\small{$\Sigma_{-1}$}} \end{overpic} {subfigure}[b]0.48 \begin{overpic}[scale={.9}]{sweep-out-compbody-color} \put(4.25,33.25){\small{grip}} \put(4.0,18.0){\small{root}} \put(2.5,3.75){\small{tines}} \put(88.0,32.75){\small{$\Sigma_{\gamma}$}} \put(88.0,17.5){\small{$\Sigma_{\rho}$}} \put(88.0,3.25){\small{$\Sigma_{\tau_{1}}\cup\Sigma_{\tau_{2}}$}} \end{overpic}

Figure 7: Sweep-out of a compression body along [1,1][-1,1] (left) and along its faithful fork (right).

3.2 Generalized Heegaard splittings and fork complexes

Heegaard splittings revisited

Let \mathscr{M} be a 3-manifold, and {1,2}\{\partial_{1}\mathscr{M},\partial_{2}\mathscr{M}\} be a partition of the components of \partial\mathscr{M}. A Heegaard splitting of (,1,2)(\mathscr{M},\partial_{1}\mathscr{M},\partial_{2}\mathscr{M}) is a triplet (𝒞1,𝒞2,𝒮)(\mathcal{C}_{1},\mathcal{C}_{2},\mathcal{S}), where 𝒞1\mathcal{C}_{1} and 𝒞2\mathcal{C}_{2} are compression bodies with 𝒞1𝒞2=\mathcal{C}_{1}\cup\mathcal{C}_{2}=\mathscr{M}, 𝒞1𝒞1=+𝒞1=+𝒞2=𝒮\mathcal{C}_{1}\cap\mathcal{C}_{1}=\partial_{+}\mathcal{C}_{1}=\partial_{+}\mathcal{C}_{2}=\mathcal{S}, 𝒞1=1\partial_{-}\mathcal{C}_{1}=\partial_{1}\mathscr{M}, and 𝒞2=2\partial_{-}\mathcal{C}_{2}=\partial_{2}\mathscr{M}. The genus of the Heegaard splitting (𝒞1,𝒞2,𝒮)(\mathcal{C}_{1},\mathcal{C}_{2},\mathcal{S}) is the genus g(𝒮)g(\mathcal{S}) of the splitting surface 𝒮\mathcal{S}. The Heegaard genus 𝔤()\mathfrak{g}\left(\mathscr{M}\right) of \mathscr{M} is the smallest genus of any Heegaard splitting of (,1,2)(\mathscr{M},\partial_{1}\mathscr{M},\partial_{2}\mathscr{M}), taken over all partitions {1,2}\{\partial_{1}\mathscr{M},\partial_{2}\mathscr{M}\} of \partial\mathscr{M}.

\begin{overpic}[scale={.7}]{heegaard-splitting-boundary} \put(-1.0,6.0){\small{$\partial_{1}\mathscr{M}$}} \put(13.75,6.0){\small{$\mathscr{M}$}} \put(22.75,6.0){\small{$\partial_{2}\mathscr{M}$}} \put(31.5,6.0){\small{$\partial_{-}\mathcal{C}_{1}$}} \put(38.0,1.0){\scalebox{1.4}{\rotatebox{90.0}{$\Bigg{\{}$}}} \put(42.75,-1.25){\small{$\mathcal{C}_{1}$}} \put(48.5,9.5){\small{$\mathcal{S}$}} \put(49.75,1.0){\scalebox{1.4}{\rotatebox{90.0}{$\Bigg{\{}$}}} \put(54.5,-1.25){\small{$\mathcal{C}_{2}$}} \put(63.0,6.0){\small{$\partial_{-}\mathcal{C}_{2}$}} \put(71.75,8.75){\small{$\partial_{-}\mathcal{C}_{1}$}} \put(74.5,1.0){\scalebox{1.4}{\rotatebox{90.0}{$\Bigg{\{}$}}} \put(79.25,-1.25){\small{$\mathcal{C}_{1}$}} \put(85.0,9.5){\small{$\mathcal{S}$}} \put(86.25,1.0){\scalebox{1.4}{\rotatebox{90.0}{$\Bigg{\{}$}}} \put(91.0,-1.25){\small{$\mathcal{C}_{2}$}} \put(95.25,8.75){\small{$\partial_{-}\mathcal{C}_{2}$}} \end{overpic}
Figure 8: Schematic of a 33-manifold \mathscr{M} with a partition of its boundary components (left). Faithful (center) and non-faithful (right) fork complexes representing a Heegaard splitting of (,1,2)(\mathscr{M},\partial_{1}\mathscr{M},\partial_{2}\mathscr{M}).
Proposition 3.2 ([49, Theorem 2.1.11], cf. [26, Appendix B]).

For any partition {1,2}\{\partial_{1}\mathscr{M},\partial_{2}\mathscr{M}\} of the boundary components of \mathscr{M}, there exists a Heegaard splitting of (,1,2)(\mathscr{M},\partial_{1}\mathscr{M},\partial_{2}\mathscr{M}).

Generalized Heegaard splittings

A generalized Heegaard splitting \mathscr{H} of a 33-manifold \mathscr{M} consists of 1. a decomposition =iIi\mathscr{M}=\bigcup_{i\in I}\mathscr{M}_{i} into submanifolds i\mathscr{M}_{i}\subseteq\mathscr{M} intersecting along closed surfaces (\Creffig:manifold-decomp), 2. for each iIi\in I a partition {1i,2i}\{\partial_{1}\mathscr{M}_{i},\partial_{2}\mathscr{M}_{i}\} of the components of i\partial\mathscr{M}_{i} that together satisfy an acyclicity condition: there is an ordering, i.e., a bijection :I{1,,|I|}\ell\colon I\rightarrow\{1,\ldots,|I|\}, such that, for each iIi\in I the components of 1i\partial_{1}\mathscr{M}_{i} (resp. 2i\partial_{2}\mathscr{M}_{i}) not belonging to \partial\mathscr{M} are only incident to submanifolds j\mathscr{M}_{j} with (j)<(i)\ell(j)<\ell(i) (resp. (j)>(i)\ell(j)>\ell(i)), and 3. a choice of a Heegaard splitting (𝒞1(i),𝒞2(i),𝒮i)(\mathcal{C}^{(i)}_{1},\mathcal{C}^{(i)}_{2},\mathcal{S}_{i}) for each (i,1i,2i)(\mathscr{M}_{i},\partial_{1}\mathscr{M}_{i},\partial_{2}\mathscr{M}_{i}). Such a choice of splitting surfaces 𝒮i\mathcal{S}_{i} (iI)(i\in I) is said to be compatible with \ell, see, e.g., \Creffig:gen-heegaard.

{subfigure}

[b]0.29 \begin{overpic}[scale={.9}]{manifold-decomp} \put(15.0,44.0){\small{$\mathscr{M}_{2}$}} \put(25.0,15.0){\small{$\mathscr{M}_{1}$}} \put(75.0,46.0){\small{$\mathscr{M}_{4}$}} \put(73.0,16.0){\small{$\mathscr{M}_{3}$}} \put(10.0,28.0){\small{$\mathcal{R}_{1}$}} \put(48.5,-10.0){\small{$\mathcal{R}_{2}$}} \put(45.0,61.0){\small{$\mathcal{R}_{3}$}} \put(82.5,31.5){\small{$\mathcal{R}_{4}$}} \end{overpic} {subfigure}[b]0.29 \begin{overpic}[scale={.9}]{gen-heegaard} \put(18.0,45.25){\small{$\mathcal{S}_{2}$}} \put(17.25,17.25){\small{$\mathcal{S}_{1}$}} \put(75.25,47.25){\small{$\mathcal{S}_{4}$}} \put(72.0,16.5){\small{$\mathcal{S}_{3}$}} \put(-4.0,-5.0){\footnotesize{$\partial_{1}\mathscr{M}_{1}=\emptyset$}} \put(-4.0,-15.0){\footnotesize{$\partial_{2}\mathscr{M}_{1}=\mathcal{R}_{1}\cup\mathcal{R}_{2}$}} \put(67.0,-5.0){\footnotesize{$\partial_{1}\mathscr{M}_{3}=\mathcal{R}_{2}$}} \put(67.0,-15.0){\footnotesize{$\partial_{2}\mathscr{M}_{3}=\mathcal{R}_{4}$}} \put(-4.0,75.0){\footnotesize{$\partial_{1}\mathscr{M}_{2}=\mathcal{R}_{1}$}} \put(-4.0,65.0){\footnotesize{$\partial_{2}\mathscr{M}_{2}=\mathcal{R}_{3}$}} \put(47.0,75.0){\footnotesize{$\partial_{1}\mathscr{M}_{4}=\mathcal{R}_{3}\cup\mathcal{R}_{4}$}} \put(47.0,65.0){\footnotesize{$\partial_{2}\mathscr{M}_{4}=\emptyset$}} \end{overpic} {subfigure}[b]0.29 \begin{overpic}[scale={.9}]{fork-comp} \put(19.0,52.0){\small{$\mathcal{S}_{2}$}} \put(19.0,2.5){\small{$\mathcal{S}_{1}$}} \put(72.75,52.0){\small{$\mathcal{S}_{4}$}} \put(72.75,2.5){\small{$\mathcal{S}_{3}$}} \end{overpic}

Figure 9: A decomposition of \mathscr{M} into four submanifolds 1,,4\mathscr{M}_{1},\ldots,\mathscr{M}_{4} intersecting along (possibly disconnected) closed surfaces i\mathcal{R}_{i}.
Figure 10: An admissible choice of splitting surfaces 𝒮i\mathcal{S}_{i} for the i\mathscr{M}_{i} that is compatible with the trivial ordering (i)i\ell(i)\mapsto i.
Figure 11: The faithful fork complex that represents the generalized Heegaard splitting shown in the center (\Creffig:gen-heegaard).
Figure 12: Schematics of a generalized Heegaard splitting (based on figures from [26, Section 2.4]).

Just as compression bodies can be represented by forks, (generalized) Heegaard splittings can be visualized via fork complexes, see \Creffig:heegaard-splitting-boundary,fig:forkcomp (cf. [49, Section 5.1] for details).

Sweep-outs of 3-manifolds

A generalized Heegaard splitting \mathscr{H} of a 33-manifold \mathscr{M} induces a sweep-out f:f\colon\mathscr{M}\rightarrow\|\mathscr{F}\| of \mathscr{M} along any fork complex \mathscr{F} that represents \mathscr{H} (here \|\mathscr{F}\| denotes a drawing, i.e., a geometric realization of the abstract fork complex \mathscr{F}) by concatenating the corresponding sweep-outs of the compression bodies that comprise \mathscr{H} (cf. \Crefrem:sweep-out-fork). We also refer to a sweep-out f:f\colon\mathscr{M}\rightarrow\|\mathscr{F}\| by the ensemble Σ={Σx:x}\Sigma=\{\Sigma_{x}:x\in\|\mathscr{F}\|\} of its level sets, where Σx=f1(x)\Sigma_{x}=f^{-1}(x).

The width of a generalized Heegaard splitting

For a generalized Heegaard splitting \mathscr{H}, the surfaces 𝒮i\mathcal{S}_{i} (iI)(i\in I) are also called the thick levels, and the lower boundaries 𝒞1(i)\partial_{-}\mathcal{C}^{(i)}_{1}, 𝒞2(i)\partial_{-}\mathcal{C}^{(i)}_{2} are called the thin levels of \mathscr{H}. The width 0pt()0pt(\mathscr{H}) of \mathscr{H} is the sequence obtained by taking a non-increasing ordering of the multiset {g(𝒮i):iI}\{g(\mathcal{S}_{i}):i\in I\} of the genera of the thick levels.

A generalized Heegaard splitting \mathscr{H} of a 3-manifold \mathscr{M} for which 0pt()0pt(\mathscr{H}) is minimal with respect to the lexicographic order (<<) among all splittings of \mathscr{M} is said to be in thin position.

3.3 Weak reductions

A Heegaard splitting (𝒞1,𝒞2,𝒮)(\mathcal{C}_{1},\mathcal{C}_{2},\mathcal{S}) of a connected 3-manifold is said to be weakly reducible [19], if there are essential disks Di𝒞iD_{i}\subset\mathcal{C}_{i} (i=1,2)(i=1,2)101010The assumption that DiD_{i} is essential in the compression body 𝒞i\mathcal{C}_{i} implies that Di+𝒞i=𝒮\partial D_{i}\subset\partial_{+}\mathcal{C}_{i}=\mathcal{S} (i=1,2)(i=1,2). with D1D2=\partial D_{1}\cap\partial D_{2}=\emptyset, see \Creffig:weak-reduction-surface. In this case we also say that the splitting surface 𝒮\mathcal{S} is weakly reducible. A generalized Heegaard splitting \mathscr{H} is weakly reducible, if at least one of its splitting surfaces is weakly reducible; otherwise \mathscr{H} is called strongly irreducible. Every 3-manifold possesses a strongly irreducible generalized Heegaard splitting and this fact can be exploited in various contexts (e.g., in the proof of \Crefthm:jsj-width). The usefulness of such splittings is mainly due to the following seminal result.

Theorem 3.3 ([49, Lemma 5.2.4], [50, Rule 5]).

Let \mathscr{H} be a strongly irreducible generalized Heegaard splitting. Then every connected component of every thin level of \mathscr{H} is incompressible.

Given a weakly reducible generalized Heegaard splitting \mathscr{H} with a weakly reducible splitting surface111111For future reference we remind the reader that splitting surfaces, also called thick levels, correspond to grips in the faithful fork complex that represents the generalized Heegaard splitting \mathscr{H}. 𝒮\mathcal{S} and essential disks D1D_{1} and D2D_{2} as above, one can execute a weak reduction at 𝒮\mathcal{S}. This modification amounts to performing particular cut-and-paste operations on 𝒮\mathcal{S} guided by D1D_{1} and D2D_{2}, and decomposes each of the two compression bodies adjacent to 𝒮\mathcal{S} into a pair of compression bodies. Importantly, this operation results in another generalized Heegaard splitting \mathscr{H}^{\prime} of the same 3-manifold with 0pt()<0pt()0pt(\mathscr{H}^{\prime})<0pt(\mathscr{H}).

We illustrate weak reductions via \Crefex:weak-reduction and refer to [49, Proposition 5.2.3] for further details (notably, Figures 5.8–5.13 therein, but also Lemma 5.2.2, Figures 5.6 and 5.7, and Proposition 5.2.4), including an exhaustive list of instances of weak reductions.121212In the open-access version [48] these are Proposition 4.2.3 and Figures 87–92 (as well as Lemma 4.2.2, Figures 85 and 86, and Lemma 4.2.4).

\begin{overpic}[scale={.9}]{weak-reduction-surface} \put(28.25,54.0){\small{$D_{1}$}} \put(25.25,18.5){\small{$\mathcal{S}$}} \put(65.5,21.5){\small{$D_{2}$}} \end{overpic}
Figure 13: Local picture of a portion of a weakly reducible splitting surface 𝒮\mathcal{S}.

4 The Main Result

In this section we prove Theorem 1.1. The inequalities (1) and (2) are deduced in the same way, thus we only show the proof of (1) in detail, and then explain how it can be adapted to that of (2). First, we specify what we mean by a “sufficiently complicated” JSJ decomposition.

Definition 4.1.

Given δ>0\delta>0, the JSJ decomposition of an irreducible 33-manifold \mathscr{M} is δ\delta-complicated, if any incompressible or strongly irreducible Heegaard surface 𝒮\mathcal{S}\subset\mathscr{M} with genus g(𝒮)δg(\mathcal{S})\leq\delta can be isotoped to be simultaneously disjoint from all the JSJ tori of \mathscr{M}.

Proof 4.2 (Proof of inequality (1)).

Our goal is to prove that tw(Γ(𝒟))18(tw()+1)\operatorname{tw}(\Gamma(\mathscr{D}))\leq 18(\operatorname{tw}(\mathscr{M})+1), where Γ(𝒟)\Gamma(\mathscr{D}) is the dual graph of the δ\delta-complicated JSJ decomposition of the irreducible 3-manifold \mathscr{M}. To this end, we set δ=18(tw()+1)\delta=18(\operatorname{tw}(\mathscr{M})+1) and fix a triangulation 𝒯\mathscr{T} of \mathscr{M}, whose dual graph Γ(𝒯)\Gamma(\mathscr{T}) has minimal treewidth, i.e., tw(Γ(𝒯))=tw()\operatorname{tw}(\Gamma(\mathscr{T}))=\operatorname{tw}(\mathscr{M}). Now, (1) is established in four stages.

{claimproof}

[1. Setup] By invoking the construction in [28, Section 6], from 𝒯\mathscr{T} we obtain a generalized Heegaard splitting \mathscr{H} of \mathscr{M} together with a sweep-out f:\colorgrayf\colon\mathscr{M}\rightarrow\|\mathscr{F}_{\color{gray}\bullet}\| along a non-faithful fork complex \colorgray\mathscr{F}_{\color{gray}\bullet} representing \mathscr{H}. By construction, \colorgray\mathscr{F}_{\color{gray}\bullet} is a tree with all of its nodes having degree one or three (\Creffig:tree-ghs), moreover all non-degenerate level surfaces Σx=f1(x)\Sigma_{x}=f^{-1}(x) have genus bounded above by 18(tw()+1)18(\operatorname{tw}(\mathscr{M})+1). Let \mathscr{F}_{\circ} be the faithful fork complex representing \mathscr{H}. Note that \mathscr{F}_{\circ} is obtained from \colorgray\mathscr{F}_{\color{gray}\bullet} by replacing every non-faithful fork F\colorgrayF\in\mathscr{F}_{\color{gray}\bullet} with the collection 𝒞F\mathscr{C}_{F} of faithful forks that accurately represents the (possibly disconnected) compression body 𝒞\mathcal{C} corresponding to FF. The inverse operation, i.e., for each F\colorgrayF\in\mathscr{F}_{\color{gray}\bullet} bundling all faithful forks in 𝒞F\mathscr{C}_{F} into FF (see \Creffig:faithful-forks), induces a projection map π:\colorgray\pi\colon\|\mathscr{F}_{\circ}\|\rightarrow\|\mathscr{F}_{\color{gray}\bullet}\| between the drawings (cf. Figures 15(i)–(ii)). Note that every fork, tine, or grip in \colorgray\mathscr{F}_{\color{gray}\bullet} corresponds to a collection of the corresponding items in \mathscr{F}_{\circ} and so the projection map is well-defined.

\begin{overpic}[scale={.909091}]{tree-ghs} \put(54.0,10.0){\Large{$\mathscr{F}_{\color{gray}\bullet}$}} \end{overpic}
Figure 14: (i) A. The dual graph Γ(𝒯)\Gamma(\mathscr{T}) of some triangulation 𝒯\mathscr{T} of a 3-manifold \mathscr{M}. B. Low-congestion routing of Γ(𝒯)\Gamma(\mathscr{T}) along a host tree with marked root arc rr. C. Drawing of a non-faithful fork complex \colorgray\mathscr{F}_{\color{gray}\bullet} that represents the generalized Heegaard splitting of \mathscr{M} induced by the routing of Γ(𝒯)\Gamma(\mathscr{T}). The genera of all (possibly disconnected) thick levels 𝒮i\mathcal{S}_{i} is bounded above by 18(tw()+1)18(\operatorname{tw}(\mathscr{M})+1). (ii) Color-coded segmentation of \colorgray\|\mathscr{F}_{\color{gray}\bullet}\| in preparation for the next stage of the proof (cf. Figure 15).
{claimproof}

[2. Weak reductions] In case \mathscr{H} is weakly reducible, we repeatedly perform weak reductions until we obtain a strongly irreducible generalized Heegaard splitting \mathscr{H}^{\prime} of \mathscr{M}. Since weak reductions always decrease the width of a generalized Heegaard splitting, this process terminates after finitely many iterations. Throughout, we maintain that the drawings of the associated faithful fork complexes follow \colorgray\|\mathscr{F}_{\color{gray}\bullet}\|. Let \mathscr{F}_{\circ}^{\prime} be the faithful fork complex representing the final splitting \mathscr{H}^{\prime}, f:f^{\prime}\colon\mathscr{M}\rightarrow\|\mathscr{F}_{\circ}^{\prime}\| be the sweep-out of \mathscr{M} induced by \mathscr{H}^{\prime}, and π:\colorgray\pi^{\prime}\colon\|\mathscr{F}_{\circ}^{\prime}\|\rightarrow\|\mathscr{F}_{\color{gray}\bullet}\| be the associated projection map (\Creffig:thm-proof(iii)). Due to the nature of weak reductions, 18(tw()+1)18(\operatorname{tw}(\mathscr{M})+1) is still an upper bound on the genus of any (possibly disconnected) level surface of πf:\colorgray\pi^{\prime}\circ f^{\prime}\colon\mathscr{M}\rightarrow\|\mathscr{F}_{\color{gray}\bullet}\|. As \mathscr{M} is irreducible, we may assume that no component of such a level surface is homeomorphic to a sphere (cf. [36, p. 337]). It follows that the number of components of any level set of πf\pi^{\prime}\circ f^{\prime} is at most 18(tw()+1)18(\operatorname{tw}(\mathscr{M})+1).

{claimproof}

[3. Perturbation and isotopy] We now apply a level-preserving perturbation ππ′′\pi^{\prime}\leadsto\pi^{\prime\prime} on π\pi^{\prime}, after which each thin level of \mathscr{H}^{\prime} lies in different level sets of π′′f:\colorgray\pi^{\prime\prime}\circ f^{\prime}\colon\mathscr{M}\rightarrow\|\mathscr{F}_{\color{gray}\bullet}\| (\Creffig:thm-proof(iv)). Since \mathscr{H}^{\prime} is strongly irreducible, its thick levels are strongly irreducible (by definition), and its thin levels are incompressible (by \Crefthm:scharlemann-thompson). Moreover, all these surfaces have genera at most 18(tw()+1)18(\operatorname{tw}(\mathscr{M})+1). Hence, as the JSJ decomposition of \mathscr{M} is assumed to be δ\delta-complicated with δ=18(tw()+1)\delta=18(\operatorname{tw}(\mathscr{M})+1), we may isotope all JSJ tori to be disjoint from all the thick and thin levels of \mathscr{H}^{\prime}. Then, by invoking [5, Corollary 4.5] we can isotope each JSJ torus of \mathscr{M} to coincide with a component of some thin level of \mathscr{H}^{\prime} (\Creffig:thm-proof(v)). As a consequence, every compression body of the splitting \mathscr{H}^{\prime} is contained in a unique JSJ piece of \mathscr{M}.

Refer to caption
Figure 15: Overview of the proof of inequality (1). The figure shows the circled area in \Creffig:tree-ghs(ii). Stage 1: Construction of initial fork complex (i) and split into faithful fork complex (ii). Stage 2: Weak reductions (iii), see [49, Proposition 5.2.3, Figures 5.8–5.13] for a complete list and their effects on the underlying fork complexes. Stage 3: perturbations and isotopy (iv). Stage 4: Construction of tree decomposition (v) and (vi).
{claimproof}

[4. The tree decomposition of Γ(𝒟)\Gamma(\mathscr{D})] First note that every level set (π′′f)1(x)(\pi^{\prime\prime}\circ f^{\prime})^{-1}(x) is incident to at most 18(tw()+1)+1=18tw()+1918(\operatorname{tw}(\mathscr{M})+1)+1=18\operatorname{tw}(\mathscr{M})+19 JSJ pieces of \mathscr{M}. The ‘plus one’ appears, because if (π′′f)1(x)(\pi^{\prime\prime}\circ f^{\prime})^{-1}(x) contains a JSJ torus, then this torus is incident to two JSJ pieces of \mathscr{M}. Also note that, because of the perturbation performed in the previous stage, each level set (π′′f)1(x)(\pi^{\prime\prime}\circ f^{\prime})^{-1}(x) can contain at most one JSJ torus of \mathscr{M}.

We now construct a tree decomposition (𝒳,T)(\mathscr{X},T) of Γ(𝒟)\Gamma(\mathscr{D}) of width 18(tw()+1)18(\operatorname{tw}(\mathscr{M})+1). Eventually, TT will be a subdivision of \colorgray\mathscr{F}_{\color{gray}\bullet} (which is a tree) with nodes corresponding to the bags in 𝒳\mathscr{X}, which we now describe. By [28, Section 6 (p. 86)], each leaf ll of \colorgray\mathscr{F}_{\color{gray}\bullet} corresponds to a spine of a handlebody l\mathcal{H}_{l}. We define a bag that contains the unique node of Γ(𝒟)\Gamma(\mathscr{D}) associated with the JSJ piece containing l\mathcal{H}_{l}. As we sweep through \colorgray\|\mathscr{F}_{\color{gray}\bullet}\| (cf. the arrows on \Creffig:tree-ghs(2)), whenever we pass through a point x\colorgrayx\in\|\mathscr{F}_{\color{gray}\bullet}\| such that the level set (π′′)1(x)(\pi^{\prime\prime})^{-1}(x) contains a tine of \|\mathscr{F}_{\circ}^{\prime}\|, one of four possible events may occur (illustrated in \Creffig:thm-proof(v)–(vi)): 1. A new JSJ piece appears. In this case we take a copy of the previous bag, and add the corresponding node of Γ(𝒟)\Gamma(\mathscr{D}) into the bag. 2. A JSJ piece disappears. Then we delete the corresponding node of Γ(𝒟)\Gamma(\mathscr{D}) from a copy of the previous bag. 3. Both previous kinds of events happen simultaneously. In this case we introduce two new bags. The first to introduce the new JSJ piece, the second to delete the old one. 4. If neither a new JSJ piece is introduced, nor an old one is left behind, we do nothing. Whenever we arrive at a merging point in the sweep-out (i.e., a degree-three node of \colorgray\|\mathscr{F}_{\color{gray}\bullet}\|), we introduce a new bag, which is the union of the two previous bags. Note that the two previous bags do not necessarily need to be disjoint.

It remains to verify that (𝒳,T)(\mathscr{X},T) is, indeed, a tree decomposition of Γ(𝒟)\Gamma(\mathscr{D}) of width at most 18tw()+1818\operatorname{tw}(\mathscr{M})+18. Node coverage: Every node of Γ(𝒟)\Gamma(\mathscr{D}) must be considered at least once, since we sweep through the entirety of \colorgray\|\mathscr{F}_{\color{gray}\bullet}\|. Arc coverage: we must ensure that all pairs of nodes of Γ(𝒟)\Gamma(\mathscr{D}) with their JSJ pieces meeting at a tine are contained in some bag. This is always the case for \colorlipicsGray1, because a new JSJ piece appears while all other JSJ pieces are still in the bag. It is also the case for \colorlipicsGray2 because a JSJ piece disappears but the previous bag contained all the pieces. In \colorlipicsGray3 a new JSJ piece appears and at the same time another JSJ piece disappears. However, in this case we first introduce the new piece, thus making sure that adjacent JSJ pieces always occur in at least one bag. Sub-tree property: This follows from the fact that a JSJ piece, once removed from all bags, must be contained in the part of \colorgray\|\mathscr{F}_{\color{gray}\bullet}\| that was already swept. Now, every JSJ piece incident to a given level set (π′′f)1(x)(\pi^{\prime\prime}\circ f^{\prime})^{-1}(x) of the sweep-out must contribute a positive number to its genus. Hence, it follows that every bag can contain at most 18tw()+1918\operatorname{tw}(\mathscr{M})+19 elements (with equality only possible where a tine simultaneously introduces and forgets a JSJ piece). This proves inequality (1).

Proof 4.3 (Proof of inequality (2)).

We start with the results from [28, Section 5] yielding a fork-complex \colorgray\mathscr{F}_{\color{gray}\bullet} whose underlying space \colorgray\|\mathscr{F}_{\color{gray}\bullet}\| is a path, and the genus of the level sets of the associated sweep-out is bounded above by 4(3pw()+1)4(3\operatorname{pw}(\mathscr{M})+1). Setting δ=4(3pw()+1)\delta=4(3\operatorname{pw}(\mathscr{M})+1), the remainder of the proof is analogous with the proof of inequality (1).

5 An Algorithmic Construction

Here we establish \Crefthm:construction that paves the way to the applications in \Crefsec:appl. In what follows, Δ\Delta denotes an arbitrary, but fixed, positive integer. Let G=(V,E)G=(V,E) be a graph with |V|=n|V|=n and maximum degree Δ\Delta. \Crefthm:construction asserts that, in poly(n)\operatorname{poly}(n) time one can construct a triangulation 𝒯G\mathscr{T}_{G} of a closed, irreducible 3-manifold G\mathscr{M}_{G}, such that the dual graph Γ(𝒟)\Gamma(\mathscr{D}) of its JSJ decomposition 𝒟\mathscr{D} equals GG, moreover, the pathwidth (resp. treewidth) of GG determines the pathwidth (resp. treewidth) of G\mathscr{M}_{G} up to a constant factor.

Our proof of \Crefthm:construction rests on a synthesis of work by Lackenby [37] and by Bachman, Derby-Talbot and Sedgwick [6]. In [37, Section 3] it is shown that the homeomorphism problem for closed 3-manifolds is at least as hard as the graph isomorphism problem. The proof relies on a simple construction that, given a graph GG, produces a closed, orientable, triangulated 3-manifold whose JSJ decomposition 𝒟\mathscr{D} satisfies Γ(𝒟)=G\Gamma(\mathscr{D})=G. This gives the blueprint for our construction as well. In particular, we use the same building blocks that are described in [37, p. 591]. However, as opposed to Lackenby, we paste together these building blocks via high-distance torus gluings akin to the construction presented in [6, Section 4]. This ensures that we can apply \Crefthm:jsj-width for the resulting 3-manifold G\mathscr{M}_{G} and deduce the right-hand-side inequalities of \colorlipicsGray3a and \colorlipicsGray3b. The left-hand-side inequalities are shown by inspecting 𝒯G\mathscr{T}_{G}. We now elaborate on the ingredients of the proof of \Crefthm:construction.

The building blocks

We first recall the definition of the building blocks from [37, p. 591]. Let kk\in\mathbb{N} be a positive integer. Consider the 2-dimensional torus 𝕋2=𝕊1×𝕊1\mathbb{T}^{2}=\mathbb{S}^{1}\times\mathbb{S}^{1}, and let 𝕋k2\mathbb{T}^{2}_{k} denote the compact surface obtained from 𝕋2\mathbb{T}^{2} by the removal of kk pairwise disjoint open disks. We define 𝒩(k)=𝕋k2×𝕊1\mathscr{N}(k)=\mathbb{T}^{2}_{k}\times\mathbb{S}^{1}. Note that the boundary 𝒩(k)\partial\mathscr{N}(k) of 𝒩(k)\mathscr{N}(k) consists of kk tori, which will be the gluing sites. See \Creffig:building-block for the example of 𝒩(3)\mathscr{N}(3).

We choose a triangulation 𝒯(k)\mathscr{T}(k) for 𝒩(k)\mathscr{N}(k) that induces the minimal 2-triangle triangulation of the torus (cf. \Creffig:torus-trg) at each boundary component. Note that 𝒯(k)\mathscr{T}(k) can be constructed from O(k)O(k) tetrahedra. An explicit description of 𝒯(k)\mathscr{T}(k) is given in \Crefapp:trg.


\nolinenumbers
\begin{overpic}[scale={0.75}]{torus-3-holes} \put(-47.0,37.0){\large{$\mathscr{N}(3)=$}} \put(106.0,37.0){\large{$\times~{}\mathbb{S}^{1}$}} \end{overpic}\captionof

figureIllustration of 𝒩(3)\mathscr{N}(3).

\begin{overpic}[scale={0.75}]{torus-trg} \end{overpic}\captionof

figure\nolinenumbersMinimal triangulation of the torus 𝕋2\mathbb{T}^{2}.

Overview of the construction of 𝓜𝑮\boldsymbol{\mathscr{M}_{G}}

We now give a high-level overview of constructing G\mathscr{M}_{G} from the above building blocks. Given a graph G=(V,E)G=(V,E), for each node vVv\in V we pick a block 𝒩v𝒩(dv)\mathscr{N}_{v}\cong\mathscr{N}(d_{v}), where dvd_{v} denotes the degree of vv. Next, for each arc e={u,v}Ee=\{u,v\}\in E, we pick a homeomorphism ϕe:𝕋2𝕋2\phi_{e}\colon\mathbb{T}^{2}\rightarrow\mathbb{T}^{2} of sufficiently high distance (we discuss this notion below) and use it to glue together a boundary torus of 𝒩u\mathscr{N}_{u} with one of 𝒩v\mathscr{N}_{v}. After performing all of these gluings, we readily obtain the 3-manifold G\mathscr{M}_{G} (cf. \Creffig:schematic-construction(i)–(ii)).

Claim 1 (name=Based on [37, p. 591], label=claim:jsj).

The JSJ decomposition 𝒟\mathscr{D} of G\mathscr{M}_{G} satisfies Γ(𝒟)=G\Gamma(\mathscr{D})=G.

For a proof of \Crefclaim:jsj we refer to \Crefapp:jsj.

\begin{overpic}[scale={0.45128205128}]{schematic-var} \put(30.0,24.5){\small{high-distance}} \put(30.0,21.0){\small{torus gluings}} \put(51.0,40.5){\small{layered}} \put(51.0,37.0){\small{triangulations}} \put(14.5,-4.0){\large{$\mathscr{M}_{G}$}} \put(74.0,-4.0){\large{$\Gamma(\mathscr{T}_{G})$}} \put(-1.5,19.0){\small{$\mathscr{N}(3)\cong\mathscr{N}_{v}$}} \put(-1.5,28.0){\small{$\mathscr{N}(1)\cong\mathscr{N}_{u}$}} \put(-1.0,1.0){\small{$\mathscr{N}_{w}$}} \put(28.75,1.0){\small{$\mathscr{N}_{z}\cong\mathscr{N}(2)$}} \put(74.0,54.0){\scalebox{0.7}{{$\Gamma(\mathscr{T}(1))$}}} \put(73.0,23.75){\small{$\Gamma(\mathscr{T}(3))$}} \put(57.5,5.45){\scalebox{0.8}{{$\Gamma(\mathscr{T}(2))$}}} \put(89.8,5.45){\scalebox{0.8}{{$\Gamma(\mathscr{T}(2))$}}} \put(20.0,52.0){\large{$G$}} \put(12.5,58.0){\small{$u$}} \put(12.5,48.0){\small{$v$}} \put(3.5,38.5){\small{$w$}} \put(27.0,38.5){\small{$z$}} \put(-8.0,50.0){\small{(i)}} \put(-8.0,15.0){\small{(ii)}} \put(92.0,50.0){\small{(iii)}} \end{overpic}
Figure 16: Schematic overview of the construction underlying \Crefthm:construction.
High-distance torus gluings

As already mentioned, to ensure that we can apply \Crefthm:jsj-width to G\mathscr{M}_{G}, we use torus homeomorphisms of “sufficiently high distance” to glue the building blocks together. This notion of distance, which is defined through the Farey distance, is somewhat technical. Thus, we refer to [6, Section 4.1 and Appendix] for details, and rather recall a crucial result that makes the usefulness of distance in the current context apparent.

Theorem 5.1 ([6, Appendix], [4, Theorem 5.4]131313The notation and the statement of \Crefthm:distance have been adapted to match the present context.).

There exists a computable constant KK, depending only on the homeomorphism types of the blocks, so that if any set of blocks are glued with maps of distance at least KδK\delta along their torus boundary components to form a closed 33-manifold \mathscr{M}, then the JSJ decomposition of \mathscr{M} is δ\delta-complicated (cf. \Crefdef:complicated-jsj).

Triangulating the gluing maps

We have already discussed that the block 𝒩(k)\mathscr{N}(k) admits a triangulation 𝒯(k)\mathscr{T}(k) with O(k)O(k) tetrahedra, where 𝒯(k)\mathscr{T}(k) induces a minimal, 1-vertex triangulation at each torus boundary of 𝒩(k)\mathscr{N}(k). It is shown in [6, Section 4.2] that the gluings beading these blocks together can be realized as layered triangulations [29]. These triangulations manifest as “daisy chains” in the dual graph Γ(𝒯G)\Gamma(\mathscr{T}_{G}) of the final triangulation 𝒯G\mathscr{T}_{G}, see \Creffig:schematic-construction(iii).

Lemma 5.2 ([6, Lemma 4.6]).

There exist torus gluings with distance at least DD, that can be realized as layered triangulations using 2D2D tetrahedra.

Claim 2.

There exist universal constants c,c>0c,c^{\prime}>0 such that

(c/Δ)tw(G)tw(G)and(c/Δ)pw(G)pw(G).\displaystyle(c/\Delta)\operatorname{tw}(\mathscr{M}_{G})\leq\operatorname{tw}(G)\quad\text{and}\quad(c^{\prime}/\Delta)\operatorname{pw}(\mathscr{M}_{G})\leq\operatorname{pw}(G).
{claimproof}

Since every node of GG has degree at most Δ\Delta, the construction of G\mathscr{M}_{G} only uses building blocks homeomorphic to 𝒩(1),,𝒩(Δ)\mathscr{N}(1),\ldots,\mathscr{N}(\Delta). Hence, for each vVv\in V the triangulation 𝒯(dv)\mathscr{T}(d_{v}) of the block 𝒩v\mathscr{N}_{v} contains O(Δ)O(\Delta) tetrahedra. This, together with the above discussion on triangulating the gluing maps implies that (upon ignoring multi-arcs and loop arcs, which are anyway not “sensed” by treewidth or pathwidth), the dual graph Γ(𝒯G)\Gamma(\mathscr{T}_{G}) is obtained from GG by 1. replacing each node vVv\in V with a copy of the graph Γ(𝒯(dv))\Gamma(\mathscr{T}(d_{v})) that contains O(Δ)O(\Delta) nodes, and by 2. possibly subdividing each arc eEe\in E several times.

Now, the first operation increases the treewidth (resp. pathwidth) at most by a factor of O(Δ)O(\Delta), while the arc-subdivisions keep these parameters basically the same, cf. \Creflem:subdiv-pw. Hence tw(Γ(𝒯G))O(Δtw(G))\operatorname{tw}(\Gamma(\mathscr{T}_{G}))\leq O(\Delta\operatorname{tw}(G)) and pw(Γ(𝒯G))O(Δpw(G))\operatorname{pw}(\Gamma(\mathscr{T}_{G}))\leq O(\Delta\operatorname{pw}(G)), and the claim follows.

Lemma 5.3 (Folklore, cf. [9, Lemma A. 1]).

Let G=(V,E)G=(V,E) be a graph. If GG^{\prime} is a graph obtained from GG by subdividing a set FEF\subseteq E of arcs an arbitrary number of times. Then

pw(G)pw(G)+2andtw(G)max{tw(G),3}.\displaystyle\operatorname{pw}(G^{\prime})\leq\operatorname{pw}(G)+2\qquad\text{and}\qquad\operatorname{tw}(G^{\prime})\leq\max\{\operatorname{tw}(G),3\}.
Proof 5.4 (Finishing the proof of \Crefthm:construction).

We have already shown \colorlipicsGray2 (\Crefclaim:jsj) and the left-hand-sides of the inequalities \colorlipicsGray3a and \colorlipicsGray3b (\Crefclaim:lhs-inequalities). To prove the remaining parts of \Crefthm:construction, let δ=max{18(tw(G)+1),4(3pw(G)+1)}=O(pw(G))\delta=\max\{18(\operatorname{tw}(G)+1),4(3\operatorname{pw}(G)+1)\}=O(\operatorname{pw}(G)). By \Crefthm:distance, there is a computable constant KΔK_{\Delta} depending only on 𝒩(1),,𝒩(Δ)\mathscr{N}(1),\ldots,\mathscr{N}(\Delta) and hence only on Δ\Delta, so that if we glue together the blocks via maps of distance at least KΔδK_{\Delta}\delta, then the JSJ decomposition of G\mathscr{M}_{G} is δ\delta-complicated. By \Creflem:layering, each such gluing map can be realized as a layered triangulation consisting of 2KΔδ2K_{\Delta}\delta tetrahedra. Since GG has at most Δn/2\Delta n/2 arcs, these layered triangulations contain at most 2KΔδΔn/2=ΔKΔδn=O(ΔKΔpw(G)n)=OΔ(pw(G)n)2K_{\Delta}\delta\Delta n/2=\Delta K_{\Delta}\delta n=O(\Delta K_{\Delta}\operatorname{pw}(G)\cdot n)=O_{\Delta}(\operatorname{pw}(G)\cdot n) tetrahedra altogether. Since the triangulated blocks 𝒯(dv)\mathscr{T}(d_{v}) contain O(Δn)O(\Delta\cdot n) tetrahedra in total, the triangulation 𝒯G\mathscr{T}_{G} of the manifold G\mathscr{M}_{G} can be built from OΔ(pw(G)n)OΔ(n2)O_{\Delta}(\operatorname{pw}(G)\cdot n)\leq O_{\Delta}(n^{2}) tetrahedra. Last, as it follows from [6, Section 4], the construction can be executed in quadratic time.

6 Applications

Corollary 6.1.

There exist 33-manifolds (h)h(\mathscr{M}_{h})_{h\in\mathbb{N}} with tw(h)2\operatorname{tw}(\mathscr{M}_{h})\leq 2 and pw(h)h\operatorname{pw}(\mathscr{M}_{h})\overset{h\to\infty}{\longrightarrow}\infty.

Corollary 6.2.

There exist Haken 33-manifolds (𝒩k)k(\mathscr{N}_{k})_{k\in\mathbb{N}} with tw(𝒩k)k\operatorname{tw}(\mathscr{N}_{k})\overset{k\to\infty}{\longrightarrow}\infty.

Proof 6.3 (Proof of \Crefcor:unbounded-pathwidth,cor:unbounded-treewidth).

Using Theorem 1.2, the construction of bounded-treewidth (Haken) 3-manifolds with arbitrarily large pathwidth follows by taking h=Th\mathscr{M}_{h}=\mathscr{M}_{T_{h}} (with the notation of \Crefthm:construction), where ThT_{h} is the complete binary tree of height hh. The construction of Haken 3-manifolds of arbitrary large treewidth is deduced by setting 𝒩k=Grid(k)\mathscr{N}_{k}=\mathscr{M}_{\operatorname{Grid}(k)}, where Grid(k)\operatorname{Grid}(k) denotes the k×kk\times k grid graph. See \Creffig:pw-tw. Obtained as JSJ decompositions, where the JSJ tori are two-sided, incompressible surfaces, all of these manifolds are Haken.

\begin{overpic}[scale={0.8205128205}]{pw-tw_examples-2} \put(56.5,19.25){\footnotesize{$\operatorname{Grid}(5)$}} \end{overpic}
Figure 17: (i) The complete binary tree ThT_{h} of height hh has pathwidth h/2\lceil h/2\rceil, cf. [11, Theorem 67]. (ii) The k×kk\times k grid graph is known to have pathwidth and treewidth both equal to kk.
Corollary 6.4.

Approximating the treewidth (resp. pathwidth) of closed, orientable 33-manifolds up to a constant factor is at least as hard as giving a constant-factor approximation of the treewidth (resp. pathwidth) of bounded-degree graphs.

Proof 6.5.

The argument for treewidth and pathwidth is the same. Given a graph GG with maximum vertex degree Δ\Delta, we use the polynomial-time procedure from \Crefthm:construction to build a 33-manifold \mathscr{M} with tw()\operatorname{tw}(\mathscr{M}) within a constant factor of tw(G)\operatorname{tw}(G). An oracle for a constant-factor approximation of tw()\operatorname{tw}(\mathscr{M}) hence gives us a constant-factor approximation of tw(G)\operatorname{tw}(G) as well.

Remark 6.6.

Computing a constant-factor approximation of treewidth (resp. pathwidth) for arbitrary graphs is known to be conditionally NP-hard under the Small Set Expansion Hypothesis [46, 55, 56]. For proving \Crefcor:hardness-treewidth, however, we rely on the assumption that the graph has bounded degree. Thus the conditional hardness of approximating the treewidth (resp. pathwidth) of a 33-manifold does not directly follow. Establishing such a hardness result would add to the growing, but still relatively short list of algorithmic problems in low-dimensional topology that are known to be (conditionally) hard, cf. [2, 6, 21, 33, 37].

7 Discussion and Open Problems

We have demonstrated that 33-manifolds with JSJ decompositions with dual graphs of large treewidth (resp. pathwidth) and “sufficiently complicated” gluing maps cannot admit triangulations of low treewidth (resp. pathwidth). This provides a technique to construct a wealth of families of 33-manifolds with unbounded tree- or pathwidth that hopefully will prove to be useful for future research in the field.

One obvious limitation of our construction is a seemingly heavy restriction on the JSJ gluing maps in order to deduce a connection between the treewidth of a 33-manifold and that of the dual graph of its JSJ decomposition. Hence, a natural question to ask is, how much this restriction on gluing maps may be relaxed while still allowing meaningful structural results about treewidth. In particular, we have the following question.

Question 7.1.

Given a 33-manifold \mathscr{M} with JSJ decomposition 𝒟\mathscr{D} and no restrictions on its JSJ gluings. Is there a lower bound for tw()\operatorname{tw}(\mathscr{M}) in terms of tw(Γ(𝒟))\operatorname{tw}(\Gamma(\mathscr{D}))?

Note that the assumption that we are considering the JSJ decomposition of \mathscr{M} is necessary: Consider a graph G=(V,E)G=(V,E) and a collection of 33-manifolds {v}vV\{\mathscr{M}_{v}\}_{v\in V}, where v\mathscr{M}_{v} has deg(v)\deg(v) torus boundary components. Assume that we glue the manifolds v\mathscr{M}_{v} along the arcs of GG to obtain a closed 33-manifold \mathscr{M}. Without restrictions on how these pieces are glued together, this cannot result in a lower bound tw()\operatorname{tw}(\mathscr{M}) in terms of tw(G)\operatorname{tw}(G): we can construct Seifert fibered spaces \mathscr{M} in this way, even if G=(V,E)G=(V,E) is the complete graph with |V||V| arbitrarily large. At the same time, Seifert fibered spaces have constant treewidth, see [27].

Question 7.2.

What is the complexity of computing the treewidth of a 33-manifold?

We believe that this should be at least as hard as computing the treewidth of a graph.

References

  • [1] I. Agol. Small 3-manifolds of large genus. Geom. Dedicata, 102:53–64, 2003. \hrefhttps://doi.org/10.1023/B:GEOM.0000006584.85248.c5 \pathdoi:10.1023/B:GEOM.0000006584.85248.c5.
  • [2] I. Agol, J. Hass, and W. Thurston. The computational complexity of knot genus and spanning area. Trans. Am. Math. Soc., 358(9):3821–3850, 2006. \hrefhttps://doi.org/10.1090/S0002-9947-05-03919-X \pathdoi:10.1090/S0002-9947-05-03919-X.
  • [3] M. A. Armstrong. Basic topology. Undergrad. Texts Math. Springer-Verlag, New York-Berlin, 1983. Corrected reprint of the 1979 original. \hrefhttps://doi.org/10.1007/978-1-4757-1793-8 \pathdoi:10.1007/978-1-4757-1793-8.
  • [4] D. Bachman. Stabilizing and destabilizing Heegaard splittings of sufficiently complicated 3-manifolds. Math. Ann., 355(2):697–728, 2013. \hrefhttps://doi.org/10.1007/s00208-012-0802-4 \pathdoi:10.1007/s00208-012-0802-4.
  • [5] D. Bachman, R. Derby-Talbot, and E. Sedgwick. Heegaard structure respects complicated JSJ decompositions. Math. Ann., 365(3-4):1137–1154, 2016. \hrefhttps://doi.org/10.1007/s00208-015-1314-9 \pathdoi:10.1007/s00208-015-1314-9.
  • [6] D. Bachman, R. Derby-Talbot, and E. Sedgwick. Computing Heegaard genus is NP-hard. In A Journey Through Discrete Mathematics: A Tribute to Jiří Matoušek, pages 59–87. Springer, Cham, 2017. \hrefhttps://doi.org/10.1007/978-3-319-44479-6_3 \pathdoi:10.1007/978-3-319-44479-6_3.
  • [7] D. Bachman, S. Schleimer, and E. Sedgwick. Sweepouts of amalgamated 3-manifolds. Algebr. Geom. Topol., 6:171–194, 2006. \hrefhttps://doi.org/10.2140/agt.2006.6.171 \pathdoi:10.2140/agt.2006.6.171.
  • [8] B. Bagchi, B. A. Burton, B. Datta, N. Singh, and J. Spreer. Efficient algorithms to decide tightness. In 32nd Int. Symp. Comput. Geom. (SoCG 2016), volume 51 of LIPIcs. Leibniz Int. Proc. Inf., pages 12:1–12:15. Schloss Dagstuhl–Leibniz-Zent. Inf., 2016. \hrefhttps://doi.org/10.4230/LIPIcs.SoCG.2016.12 \pathdoi:10.4230/LIPIcs.SoCG.2016.12.
  • [9] R. Belmonte, T. Hanaka, M. Kanzaki, M. Kiyomi, Y. Kobayashi, Y. Kobayashi, M. Lampis, H. Ono, and Y. Otachi. Parameterized complexity of (A,)(A,\ell)-path packing. Algorithmica, 84(4):871–895, 2022. \hrefhttps://doi.org/10.1007/s00453-021-00875-y \pathdoi:10.1007/s00453-021-00875-y.
  • [10] R. H. Bing. An alternative proof that 33-manifolds can be triangulated. Ann. Math. (2), 69:37–65, 1959. \hrefhttps://doi.org/10.2307/1970092 \pathdoi:10.2307/1970092.
  • [11] H. L. Bodlaender. A partial k-arboretum of graphs with bounded treewidth. Theor. Comput. Sci., 209(1–2):1–45, 1998. \hrefhttps://doi.org/10.1016/S0304-3975(97)00228-4 \pathdoi:10.1016/S0304-3975(97)00228-4.
  • [12] B. A. Burton. Computational topology with Regina: algorithms, heuristics and implementations. In Geometry and Topology Down Under, volume 597 of Contemp. Math., pages 195–224. Am. Math. Soc., Providence, RI, 2013. \hrefhttps://doi.org/10.1090/conm/597/11877 \pathdoi:10.1090/conm/597/11877.
  • [13] B. A. Burton, R. Budney, W. Pettersson, et al. Regina: Software for low-dimensional topology, 1999–2022. Version 7.2. URL: \urlhttps://regina-normal.github.io.
  • [14] B. A. Burton and R. G. Downey. Courcelle’s theorem for triangulations. J. Comb. Theory, Ser. A, 146:264–294, 2017. \hrefhttps://doi.org/10.1016/j.jcta.2016.10.001 \pathdoi:10.1016/j.jcta.2016.10.001.
  • [15] B. A. Burton, T. Lewiner, J. Paixão, and J. Spreer. Parameterized complexity of discrete Morse theory. ACM Trans. Math. Softw., 42(1):6:1–6:24, 2016. \hrefhttps://doi.org/10.1145/2738034 \pathdoi:10.1145/2738034.
  • [16] B. A. Burton, C. Maria, and J. Spreer. Algorithms and complexity for Turaev–Viro invariants. J. Appl. Comput. Topol., 2(1–2):33–53, 2018. \hrefhttps://doi.org/10.1007/s41468-018-0016-2 \pathdoi:10.1007/s41468-018-0016-2.
  • [17] B. A. Burton and W. Pettersson. Fixed parameter tractable algorithms in combinatorial topology. In Proc. 20th Int. Conf. Comput. Comb. (COCOON 2014), pages 300–311, 2014. \hrefhttps://doi.org/10.1007/978-3-319-08783-2_26 \pathdoi:10.1007/978-3-319-08783-2_26.
  • [18] B. A. Burton and J. Spreer. The complexity of detecting taut angle structures on triangulations. In Proc. 24th Annu. ACM-SIAM Symp. Discrete Algorithms (SODA 2013), pages 168–183, 2013. \hrefhttps://doi.org/10.1137/1.9781611973105.13 \pathdoi:10.1137/1.9781611973105.13.
  • [19] A. J. Casson and C. McA. Gordon. Reducing Heegaard splittings. Topology Appl., 27(3):275–283, 1987. \hrefhttps://doi.org/10.1016/0166-8641(87)90092-7 \pathdoi:10.1016/0166-8641(87)90092-7.
  • [20] H. S. M. Coxeter. Regular honeycombs in hyperbolic space. In Proc. Int. Congr. Math., Amsterdam, The Netherlands, September 2–9, 1954, volume 3, pages 155–169. Erven P. Noordhoff; North-Holland, 1956. Available at \urlhttps://www.mathunion.org/fileadmin/ICM/Proceedings/ICM1954.3/ICM1954.3.ocr.pdf (accessed: March 12, 2023).
  • [21] A. de Mesmay, Y. Rieck, E. Sedgwick, and M. Tancer. The unbearable hardness of unknotting. Adv. Math., 381:Paper No. 107648, 36, 2021. \hrefhttps://doi.org/10.1016/j.aim.2021.107648 \pathdoi:10.1016/j.aim.2021.107648.
  • [22] Henri Paul de Saint-Gervais. Uniformization of Riemann surfaces. Revisiting a hundred-year-old theorem. Translated from the French by R. G. Burns. Herit. Eur. Math. Eur. Math. Soc. (EMS), Zürich, 2016. \hrefhttps://doi.org/10.4171/145 \pathdoi:10.4171/145.
  • [23] A. Hatcher. Notes on basic 3-manifold topology. Available at \urlhttps://pi.math.cornell.edu/ hatcher/3M/3Mfds.pdf (accessed: February 18, 2023).
  • [24] P. Heegaard. Sur l’”Analysis situs”. Bull. Soc. Math. France, 44:161–242, 1916. \hrefhttps://doi.org/10.24033/bsmf.968 \pathdoi:10.24033/bsmf.968.
  • [25] K. Huszár. Combinatorial width parameters for 33-dimensonal manifolds. PhD thesis, IST Austria, June 2020. \hrefhttps://doi.org/10.15479/AT:ISTA:8032 \pathdoi:10.15479/AT:ISTA:8032.
  • [26] K. Huszár. On the pathwidth of hyperbolic 3-manifolds. Comput. Geom. Topol., 1(1):1–19, 2022. \hrefhttps://doi.org/10.57717/cgt.v1i1.4 \pathdoi:10.57717/cgt.v1i1.4.
  • [27] K. Huszár and J. Spreer. 3-Manifold triangulations with small treewidth. In 35th Int. Symp. Comput. Geom. (SoCG 2019), volume 129 of LIPIcs. Leibniz Int. Proc. Inf., pages 44:1–44:20. Schloss Dagstuhl–Leibniz-Zent. Inf., 2019. \hrefhttps://doi.org/10.4230/LIPIcs.SoCG.2019.44 \pathdoi:10.4230/LIPIcs.SoCG.2019.44.
  • [28] K. Huszár, J. Spreer, and U. Wagner. On the treewidth of triangulated 3-manifolds. J. Comput. Geom., 10(2):70–98, 2019. \hrefhttps://doi.org/10.20382/jogc.v10i2a5 \pathdoi:10.20382/jogc.v10i2a5.
  • [29] W. Jaco and J. H. Rubinstein. Layered-triangulations of 3-manifolds, 2006. 97 pages, 32 figures. \hrefhttp://arxiv.org/abs/math/0603601 \patharXiv:math/0603601.
  • [30] W. Jaco and P. B. Shalen. A new decomposition theorem for irreducible sufficiently-large 33-manifolds. In Algebraic and Geometric Topology, volume 32, part 2 of Proc. Sympos. Pure Math., pages 71–84. Am. Math. Soc., Providence, RI, 1978. \hrefhttps://doi.org/10.1090/pspum/032.2/520524 \pathdoi:10.1090/pspum/032.2/520524.
  • [31] W. H. Jaco and P. B. Shalen. Seifert fibered spaces in 33-manifolds. Mem. Am. Math. Soc., 21(220):viii+192, 1979. \hrefhttps://doi.org/10.1090/memo/0220 \pathdoi:10.1090/memo/0220.
  • [32] K. Johannson. Homotopy equivalences of 33-manifolds with boundaries, volume 761 of Lect. Notes Math. Springer, Berlin, 1979. \hrefhttps://doi.org/10.1007/BFb0085406 \pathdoi:10.1007/BFb0085406.
  • [33] D. Koenig and A. Tsvietkova. NP-hard problems naturally arising in knot theory. Trans. Am. Math. Soc. Ser. B, 8:420–441, 2021. \hrefhttps://doi.org/10.1090/btran/71 \pathdoi:10.1090/btran/71.
  • [34] G. Kuperberg. Algorithmic homeomorphism of 3-manifolds as a corollary of geometrization. Pacific J. Math., 301(1):189–241, 2019. \hrefhttps://doi.org/10.2140/pjm.2019.301.189 \pathdoi:10.2140/pjm.2019.301.189.
  • [35] M. Lackenby. The Heegaard genus of amalgamated 3-manifolds. Geom. Dedicata, 109:139–145, 2004. \hrefhttps://doi.org/10.1007/s10711-004-6553-y \pathdoi:10.1007/s10711-004-6553-y.
  • [36] M. Lackenby. Heegaard splittings, the virtually Haken conjecture and property (τ)(\tau). Invent. Math., 164(2):317–359, 2006. \hrefhttps://doi.org/10.1007/s00222-005-0480-x \pathdoi:10.1007/s00222-005-0480-x.
  • [37] M. Lackenby. Some conditionally hard problems on links and 3-manifolds. Discrete Comput. Geom., 58(3):580–595, 2017. \hrefhttps://doi.org/10.1007/s00454-017-9905-8 \pathdoi:10.1007/s00454-017-9905-8.
  • [38] M. Lackenby. Algorithms in 3-manifold theory. In I. Agol and D. Gabai, editors, Surveys in 33-manifold topology and geometry, volume 25 of Surv. Differ. Geom., pages 163–213. Int. Press Boston, 2020. \hrefhttps://doi.org/10.4310/SDG.2020.v25.n1.a5 \pathdoi:10.4310/SDG.2020.v25.n1.a5.
  • [39] Tao Li. Heegaard surfaces and the distance of amalgamation. Geom. Topol., 14(4):1871–1919, 2010. \hrefhttps://doi.org/10.2140/gt.2010.14.1871 \pathdoi:10.2140/gt.2010.14.1871.
  • [40] C. Manolescu. Lectures on the triangulation conjecture. In Proc. 22nd Gökova Geom.-Topol. Conf. (GGT 2015), pages 1–38. Int. Press Boston, 2016. URL: \urlhttps://gokovagt.org/proceedings/2015/manolescu.html.
  • [41] C. Maria and J. Purcell. Treewidth, crushing and hyperbolic volume. Algebr. Geom. Topol., 19(5):2625–2652, 2019. \hrefhttps://doi.org/10.2140/agt.2019.19.2625 \pathdoi:10.2140/agt.2019.19.2625.
  • [42] A. Markov. The insolubility of the problem of homeomorphy (in Russian). Dokl. Akad. Nauk SSSR, 121:218–220, 1958.
  • [43] E. E. Moise. Affine structures in 33-manifolds. V. The triangulation theorem and Hauptvermutung. Ann. Math. (2), 56:96–114, 1952. \hrefhttps://doi.org/10.2307/1969769 \pathdoi:10.2307/1969769.
  • [44] P. Orlik, E. Vogt, and H. Zieschang. Zur Topologie gefaserter dreidimensionaler Mannigfaltigkeiten. Topology, 6:49–64, 1967. \hrefhttps://doi.org/10.1016/0040-9383(67)90013-4 \pathdoi:10.1016/0040-9383(67)90013-4.
  • [45] B. Poonen. Undecidable problems: a sampler. In Interpreting Gödel, pages 211–241. Cambridge Univ. Press, 2014. \hrefhttps://doi.org/10.1017/CBO9780511756306.015 \pathdoi:10.1017/CBO9780511756306.015.
  • [46] P. Raghavendra and D. Steurer. Graph expansion and the unique games conjecture. In Proc. 2010 ACM Int. Symp. Theor. Comput. (STOC’10), pages 755–764. ACM, New York, 2010.
  • [47] M. Scharlemann and J. Schultens. Comparing Heegaard and JSJ structures of orientable 3-manifolds. Trans. Am. Math. Soc., 353(2):557–584, 2001. \hrefhttps://doi.org/10.1090/S0002-9947-00-02654-4 \pathdoi:10.1090/S0002-9947-00-02654-4.
  • [48] M. Scharlemann, J. Schultens, and T. Saito. Lecture notes on generalized Heegaard splittings, 2005. Open-access version of [49] with slightly different structure. \hrefhttp://arxiv.org/abs/math/0504167 \patharXiv:math/0504167.
  • [49] M. Scharlemann, J. Schultens, and T. Saito. Lecture Notes on Generalized Heegaard Splittings. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2016. \hrefhttps://doi.org/10.1142/10019 \pathdoi:10.1142/10019.
  • [50] M. Scharlemann and A. Thompson. Thin position for 33-manifolds. In Geometric topology (Haifa, 1992), volume 164 of Contemp. Math., pages 231–238. Am. Math. Soc., Providence, RI, 1994. \hrefhttps://doi.org/10.1090/conm/164/01596 \pathdoi:10.1090/conm/164/01596.
  • [51] J. Schultens. Introduction to 33-Manifolds, volume 151 of Grad. Stud. Math. Am. Math. Soc., Providence, RI, 2014. \hrefhttps://doi.org/10.1090/gsm/151 \pathdoi:10.1090/gsm/151.
  • [52] J. Schultens and R. Weidmann. Destabilizing amalgamated Heegaard splittings. In Workshop on Heegaard Splittings, volume 12 of Geom. Topol. Monogr., pages 319–334. Geom. Topol. Publ., Coventry, 2007. \hrefhttps://doi.org/10.2140/gtm.2007.12.319 \pathdoi:10.2140/gtm.2007.12.319.
  • [53] P. Scott and H. Short. The homeomorphism problem for closed 3-manifolds. Algebr. Geom. Topol., 14(4):2431–2444, 2014. \hrefhttps://doi.org/10.2140/agt.2014.14.2431 \pathdoi:10.2140/agt.2014.14.2431.
  • [54] F. Waldhausen. Eine Klasse von 33-dimensionalen Mannigfaltigkeiten. I, II. Invent. Math., 3:308–333; ibid. 4:87–117, 1967. URL: \urlhttps://eudml.org/doc/141878 and \urlhttps://eudml.org/doc/141884. \hrefhttps://doi.org/10.1007/BF01402956 \pathdoi:10.1007/BF01402956.
  • [55] Y. Wu, P. Austrin, T. Pitassi, and D. Liu. Inapproximability of treewidth, one-shot pebbling, and related layout problems. J. Artificial Intelligence Res., 49:569–600, 2014. \hrefhttps://doi.org/10.1613/jair.4030 \pathdoi:10.1613/jair.4030.
  • [56] K. Yamazaki. Inapproximability of rank, clique, Boolean, and maximum induced matching-widths under small set expansion hypothesis. Algorithms (Basel), 11(11):Paper No. 173, 10, 2018. \hrefhttps://doi.org/10.3390/a11110173 \pathdoi:10.3390/a11110173.

Appendix A Selected Results from 3-Manifold Topology

In this appendix we collect important facts about the topology of 3-manifolds we rely on. We do not aim to reproduce their complete proofs; indeed, those are often quite technical or include lengthy case analyses. However, we do aim to give precise statements and references.

A.1 Weak reductions

Example A.1.

The 33-torus 𝕋3=𝕊1×𝕊1×𝕊1\mathbb{T}^{3}=\mathbb{S}^{1}\times\mathbb{S}^{1}\times\mathbb{S}^{1} is a closed 3-manifold built by identifying opposite faces of a solid cube KK (\Creffig:3-torus). It is well-known that 𝕋3\mathbb{T}^{3} has Heegaard genus 𝔤(𝕋3)=3\mathfrak{g}\left(\mathbb{T}^{3}\right)=3. A genus-three Heegaard splitting =(1,2,𝒮)\mathscr{H}=(\mathcal{H}_{1},\mathcal{H}_{2},\mathcal{S}) of 𝕋3\mathbb{T}^{3} can be obtained as follows: Upon the face-identifications, the 1-skeleton of KK becomes a bouquet [Uncaptioned image] of three circles. We let 1=N¯𝕋3([Uncaptioned image])\mathcal{H}_{1}=\overline{N}_{\mathbb{T}^{3}}(\raisebox{-1.5pt}{\includegraphics[scale={.75}]{bouquet}}) be a closed regular neighborhood of [Uncaptioned image] in 𝕋3\mathbb{T}^{3}, 2\mathcal{H}_{2} be the closure of 𝕋31\mathbb{T}^{3}\setminus\mathcal{H}_{1}, and 𝒮=12\mathcal{S}=\mathcal{H}_{1}\cap\mathcal{H}_{2} be the resulting splitting surface (\Creffig:3-torus-splitting).

{subfigure}

[b]0.3 Refer to caption {subfigure}[b]0.3 Refer to caption {subfigure}[b]0.3 \begin{overpic}[width=368.57964pt]{3-torus-weak-reduction} \put(13.0,17.0){\color[RGB]{0,0,0}$D_{1}$} \put(35.0,40.0){\color[RGB]{0,0,0}$D_{2}$} \end{overpic}

Figure 18: The 3-torus 𝕋3=𝕊1×𝕊1×𝕊1\mathbb{T}^{3}=\mathbb{S}^{1}\times\mathbb{S}^{1}\times\mathbb{S}^{1} can be obtained by identifying opposite faces of a solid cube.
Figure 19: A surface 𝒮𝕋3\mathcal{S}\subset\mathbb{T}^{3} defining a Heegaard splitting (1,2,𝒮)(\mathcal{H}_{1},\mathcal{H}_{2},\mathcal{S}) of 𝕋3\mathbb{T}^{3} of minimum genus three.
Figure 20: Two essential disks DiiD_{i}\subset\mathcal{H}_{i} (i=1,2)(i=1,2) witnessing the weak reducibility of (1,2,𝒮)(\mathcal{H}_{1},\mathcal{H}_{2},\mathcal{S}).
Figure 21: A weakly reducible minimum-genus Heegaard splitting of the 3-torus 𝕋3=𝕊1×𝕊1×𝕊1\mathbb{T}^{3}=\mathbb{S}^{1}\times\mathbb{S}^{1}\times\mathbb{S}^{1}.

Note that D1D_{1} and D2D_{2} in \Creffig:3-torus-weak-reduction are essential disks in the handlebodies 1\mathcal{H}_{1} and 2\mathcal{H}_{2}, respectively, and D1D2=\partial D_{1}\cap\partial D_{2}=\emptyset, hence the splitting =(1,2,𝒮)\mathscr{H}=(\mathcal{H}_{1},\mathcal{H}_{2},\mathcal{S}) is weakly reducible. Now, a weak reduction at 𝒮\mathcal{S} may be carried out as presented on \Creffig:schemata-weak-reduction. We refer to [49, p. 108–111] for details (cf. [48, p. 61–62]).

{subfigure}

[b]0.3 \begin{overpic}[scale={1.05}]{heegaard-splitting-schematic} \put(1.0,39.0){\small{$\mathcal{S}$}} \put(46.0,23.0){\small{$\mathcal{H}_{2}$}} \put(46.0,54.0){\small{$\mathcal{H}_{1}$}} \put(102.5,54.0){\LARGE{$\leadsto$}} \end{overpic} {subfigure}[b]0.3 \begin{overpic}[scale={1.05}]{weak-reducing-disks} \put(1.0,39.0){\small{$\mathcal{S}$}} \put(62.0,0.0){\small{$D_{2}$}} \put(60.0,46.0){\small{$\partial D_{2}$}} \put(29.5,76.5){\small{$D_{1}$}} \put(27.5,30.5){\small{$\partial D_{1}$}} \put(102.5,54.0){\LARGE{$\leadsto$}} \end{overpic} {subfigure}[b]0.3 \begin{overpic}[scale={1.05}]{weak-reduction-schematic} \end{overpic}

{subfigure}

[b]1 \begin{overpic}[scale={1.05}]{weak-reduction-schematic-colors} \put(46.75,19.0){\small{$\mathcal{S}$}} \put(42.0,12.0){\small{$\mathcal{H}_{2}$}} \put(51.0,12.0){\small{$\mathcal{H}_{1}$}} \put(74.75,19.0){\small{$\mathcal{S}_{2}$}} \put(70.0,12.0){\small{$\mathcal{C}^{(2)}_{1}$}} \put(79.0,12.0){\small{$\mathcal{C}^{(2)}_{2}$}} \put(92.5,19.0){\small{$\mathcal{S}_{1}$}} \put(87.75,12.0){\small{$\mathcal{C}^{(1)}_{1}$}} \put(96.75,12.0){\small{$\mathcal{C}^{(1)}_{2}$}} \put(37.5,4.0){\small{$0pt(\mathscr{H})=(g(\mathcal{S}))=(3)$}} \put(68.5,4.0){\small{$0pt(\mathscr{H}^{\prime})=(g(\mathcal{S}_{1}),g(\mathcal{S}_{2}))=(2,2)$}} \put(60.0,16.0){\Large{$\leadsto$}} \put(8.5,5.5){\small{$\mathcal{C}^{(2)}_{1}$}} \put(19.25,18.0){\small{$\mathcal{C}^{(1)}_{2}$}} \put(0.0,9.0){\small{$\mathcal{S}_{2}$}} \put(0.0,14.5){\small{$\mathcal{S}_{1}$}} \put(30.25,8.0){\small{$\mathcal{C}^{(2)}_{2}$}} \put(30.25,15.5){\small{$\mathcal{C}^{(1)}_{1}$}} \end{overpic}

Figure 22: Schematic “movie” of a weak reduction of the genus-three Heegaard splitting of 𝕋3\mathbb{T}^{3}.

A.2 Seifert fibered spaces and JSJ decompositions

A key assertion about the manifold G\mathscr{M}_{G} constructed in \Crefsec:const is \Crefclaim:jsj, according to which the dual graph Γ(𝒟)\Gamma(\mathscr{D}) of the JSJ decomposition 𝒟\mathscr{D} of G\mathscr{M}_{G} satisfies Γ(𝒟)=G\Gamma(\mathscr{D})=G. This can be shown through the same argument that is outlined by Lackenby in [37, p. 591]. However, while we use the same building blocks as Lackenby (namely the manifolds 𝒩(k)\mathscr{N}(k)), we glue them together via different (high-distance) maps. Therefore, some explanations are in order. To this end, closely following [23], we recall some basic definitions and important results.

Seifert fiberings

A model Seifert fibering of the solid torus 𝕊1×D\mathbb{S}^{1}\times D is a decomposition of 𝕊1×D\mathbb{S}^{1}\times D into disjoint circles, called fibers, obtained as follows. Consider the solid cylinder [0,1]×D[0,1]\times D and identify its boundary disks {0}×D\{0\}\times D and {1}×D\{1\}\times D through a rotation by 2πp/q2\pi p/q, where pp and qq are relatively prime integers with q>0q>0. After this identification the segment [0,1]×{0}[0,1]\times\{0\} becomes a fiber 𝕊1×{0}\mathbb{S}^{1}\times\{0\}, whereas every other fiber consists of qq segments of the form [0,1]×{x}[0,1]\times\{x\}. Next, a Seifert fibering of 33-manifold \mathscr{M} is a decomposition of \mathscr{M} into disjoint circles (also called fibers), such that each fiber has a neighborhood homeomorphic to that of a fiber in some model Seifert fibering of 𝕊1×D\mathbb{S}^{1}\times D via a fiber-preserving homeomorphism. A 33-manifold that admits a Seifert fibering is called a Seifert fibered space. The following classical result says that most Seifert fibered spaces admit a unique Seifert fibering.

Theorem A.2 ([44, 54], adapted from [23, Theorem 2.3]).

Every orientable Seifert fibered space admits a unique Seifert fibering up to fiber-preserving homeomorphism, except for:

  • 𝕊1×D\mathbb{S}^{1}\times D, the solid torus (various model fiberings; see above),

  • 𝕊1×𝕊1×I\mathbb{S}^{1}\mathbin{\stackrel{{\scriptstyle\sim}}{{\smash{\times}\rule{0.0pt}{3.87495pt}}}}\mathbb{S}^{1}\mathbin{\stackrel{{\scriptstyle\sim}}{{\smash{\times}\rule{0.0pt}{3.87495pt}}}}I, the orientable 𝕊1\mathbb{S}^{1}-bundle over the Möbius band (two fiberings),

  • Seifert fibered spaces over P2\mathbb{R}P^{2} with one exceptional fiber (two fiberings),

  • lens spaces L(p,q)L(p,q) including 𝕊3\mathbb{S}^{3} and 𝕊1×𝕊2\mathbb{S}^{1}\times\mathbb{S}^{2} (various fiberings), and

  • 𝕊1×𝕊1×𝕊1\mathbb{S}^{1}\mathbin{\stackrel{{\scriptstyle\sim}}{{\smash{\times}\rule{0.0pt}{3.87495pt}}}}\mathbb{S}^{1}\mathbin{\stackrel{{\scriptstyle\sim}}{{\smash{\times}\rule{0.0pt}{3.87495pt}}}}\mathbb{S}^{1}, the orientable 𝕊1\mathbb{S}^{1}-bundle over the Klein bottle (two fiberings).

Corollary A.3.

The manifolds 𝒩(k)\mathscr{N}(k) defined in \Crefsec:const admit unique Seifert fiberings.

Proof A.4.

The statement follows by observing that 𝒩(k)\mathscr{N}(k) is a Seifert fibered space over the torus, and so it does not belong to any of the exceptional manifolds mentioned in \Crefthm:seifert-uniqueness.

Proposition A.5 (folklore, see [23, Proposition 1.6]).

If an irreducible 33-manifold ~\widetilde{\mathscr{M}} is a covering space of the interior of another 33-manifold \mathscr{M}, then \mathscr{M} is irreducible as well.

Claim 3 (folklore, see [23, p. 11]).

For any compact, connected and orientable surface FF (possibly with boundary) other than 𝕊2\mathbb{S}^{2}, the interior of F×𝕊1F\times\mathbb{S}^{1} has 3\mathbb{R}^{3} as its universal cover.141414The claim holds for non-orientable surfaces as well with the exception of P2\mathbb{R}P^{2}.

{claimproof}

If F\ncong𝕊2F\ncong\mathbb{S}^{2}, it is well-known that the universal cover of FF is 2\mathbb{R}^{2}. Indeed, for F𝕋2F\cong\mathbb{T}^{2} a covering map 2F\mathbb{R}^{2}\rightarrow F can be defined through the regular tiling of the Euclidean plane via squares, while for F𝕊gF\cong\mathbb{S}_{g} with genus g>1g>1 such a map can be obtained through the regular tiling of the hyperbolic plane via regular 4g4g-gons, with 4g4g such polygons meeting at each vertex [20]. If FF has non-empty boundary, then by the theory of covering spaces [3, Theorem 10.19] the interior of FF has some universal cover 𝒳\mathscr{X}, and by the uniformization theorem [22, Theorem XII.0.1], this space 𝒳\mathscr{X} – being a non-compact, open, simply-connected surface – must be homeomorphic to 2\mathbb{R}^{2}. As the universal cover of 𝕊1\mathbb{S}^{1} is \mathbb{R}, it follows that the universal cover of F×𝕊1F\times\mathbb{S}^{1} is 3\mathbb{R}^{3}.

Claim 4 (folklore, see [23, p. 14]).

A 22-sided torus TT\subset\mathscr{M} in an irreducible 3-manifold \mathscr{M} is compressible if and only if TT either bounds a solid torus 𝕊1×D\mathbb{S}^{1}\times D\subset\mathscr{M} or lies in a ball in \mathscr{M}.

Corollary A.6.

For any kk\in\mathbb{N}, 𝒩(k)\mathscr{N}(k) is irreducible and has incompressible boundary.

Proof A.7.

The irreducibility of 𝒩(k)\mathscr{N}(k) is a consequence of \Crefprop:irreducible-cover and \Crefclaim:surface-cover. Then, the incompressibility of 𝒩(k)\partial\mathscr{N}(k) follows from \Crefclaim:compressible-torus by noting that 𝒩(k)\partial\mathscr{N}(k) is a union of tori and, by construction of 𝒩(k)\mathscr{N}(k), none of them bounds a solid torus or lies in a ball in 𝒩(k)\mathscr{N}(k).

Proposition A.8.

The manifold G\mathscr{M}_{G} constructed in \Crefsec:const is irreducible.

Proof A.9.

Let G=(V,E)G=(V,E) be a multigraph on nn nodes with mm arcs, and 𝐓={T1,,Tm}\mathbf{T}=\{T_{1},\ldots,T_{m}\} be the collection of pairwise disjoint tori TiGT_{i}\subset\mathscr{M}_{G} that are the images of the boundary tori of the manifolds 𝒩v𝒩(dv)\mathscr{N}_{v}\cong\mathscr{N}(d_{v}) after performing the torus-gluings in the construction of G\mathscr{M}_{G}.

Let us assume that G\mathscr{M}_{G} is not irreducible, i.e., there is some reducing 2-sphere SGS\subset\mathscr{M}_{G}. By definition, SS does not bound a ball in G\mathscr{M}_{G}. Note that SS cannot be entirely contained in any of the building blocks 𝒩vG\mathscr{N}_{v}\subset\mathscr{M}_{G}, otherwise it would imply the reducibility of 𝒩v𝒩(dv)\mathscr{N}_{v}\cong\mathscr{N}(d_{v}), contradicting \Crefcor:irreducible. It follows that S(i=1mTi)S\cap(\cup_{i=1}^{m}T_{i})\neq\emptyset. Thus, after an isotopy if necessary, we may assume that SS intersects i=1mTi\cup_{i=1}^{m}T_{i} transversally, and the intersection S(i=1mTi)S\cap(\cup_{i=1}^{m}T_{i}) is a union of pairwise disjoint closed curves 𝜶={α1,,α}\boldsymbol{\alpha}=\{\alpha_{1},\ldots,\alpha_{\ell}\} with \ell being minimal.

Now let αj𝜶\alpha_{j}\in\boldsymbol{\alpha} be a curve that is innermost in SS, i.e., one of the two disks bounded by αj\alpha_{j} in SS, say DSD\subset S, does not contain any other curves from 𝜶\boldsymbol{\alpha}. After relabeling if necessary, let Tj𝐓T_{j}\in\mathbf{T} denote the torus for which αj=STj\alpha_{j}=S\cap T_{j}. Note that, if αj\alpha_{j} also bounds a disk DTjD^{\prime}\subset T_{j}, then DDD\cup D^{\prime} is a sphere entirely contained in one of the building blocks 𝒩w\mathscr{N}_{w} incident to TjT_{j}. Since 𝒩w\mathscr{N}_{w} is irreducible (\Crefcor:irreducible), DDD\cup D^{\prime} bounds a ball B𝒩wB\subset\mathscr{N}_{w}, which can be used to eliminate αj\alpha_{j} by first isotoping DD inside BB to coincide with DD^{\prime} and then pushing DD^{\prime} off the torus TjT_{j}, contradicting the minimality of 𝜶\boldsymbol{\alpha}. Hence the disk DD is essential in 𝒩w\mathscr{N}_{w}. However, this contradicts the incompressibility of TjT_{j} in 𝒩w\mathscr{N}_{w} (\Crefcor:irreducible).

Proof A.10 (Proof of \Crefclaim:jsj).

We follow the notation introduced in the proof of \Crefprop:irreducible-MG. By \Crefcor:seifert-uniqueness-blocks, each building block 𝒩v\mathscr{N}_{v} admits a unique Seifert fibering. In particular, each boundary component of 𝒩v\mathscr{N}_{v} has a uniquely determined fibering. Therefore, when two building blocks 𝒩u\mathscr{N}_{u} and 𝒩u\mathscr{N}_{u^{\prime}} are glued together along boundary tori T𝒩uT\subset\partial\mathscr{N}_{u} and T𝒩uT^{\prime}\subset\partial\mathscr{N}_{u^{\prime}}, the resulting 3-manifold is not Seifert fibered, unless the gluing map aligns the (identical) Seifert fiberings of TT and TT^{\prime}. Now, it follows from the definition of distance [6, Section 4.1] that gluing maps of distance at least one do not align these Seifert fiberings. (In fact, for any two fixed Seifert fiberings of the torus, there are precisely two maps that align them.) Hence, upon the manifold G\mathscr{M}_{G} is constructed, the Seifert fibrations of adjacent building blocks induce different fibrations on each torus in 𝐓={T1,,Tm}\mathbf{T}=\{T_{1},\ldots,T_{m}\}. Now, by construction of the building blocks, none of the tori in 𝐓\mathbf{T} bounds a solid torus (cf. [37, p. 591]), therefore they are all incompressible (cf. \Crefclaim:compressible-torus) and pairwise non-parallel. This implies that the building blocks 𝒩v\mathscr{N}_{v} (vVv\in V) are indeed the JSJ pieces of G\mathscr{M}_{G} and the tori in 𝐓\mathbf{T} comprise its JSJ tori.

Appendix B Triangulating the building blocks

Here we describe the triangulation 𝒯(k)\mathscr{T}(k) of the building block 𝒩(k)\mathscr{N}(k) introduced in \Crefsec:const. Topologically, 𝒩(k)=𝕋k2×𝕊1\mathscr{N}(k)=\mathbb{T}^{2}_{k}\times\mathbb{S}^{1}, where 𝕋k2\mathbb{T}^{2}_{k} is obtained by removing the interiors of kk pairwise disjoint disks from the torus 𝕋2\mathbb{T}^{2}. To construct 𝒯(k)\mathscr{T}(k), first we triangulate 𝕋k2\mathbb{T}^{2}_{k} with 3k+23k+2 triangles as shown in \Creffig:torus+holes. This triangulation of 𝕋k2\mathbb{T}^{2}_{k} naturally lifts to a decomposition of 𝕋k2×[0,1]\mathbb{T}^{2}_{k}\times[0,1] into 3k+23k+2 triangular prisms, each of which can be triangulated with three tetrahedra, see \Creffig:trg_prism. The construction of 𝒯(k)\mathscr{T}(k) is concluded by identifying the triangulations of 𝕋k2×{0}\mathbb{T}^{2}_{k}\times\{0\} and 𝕋k2×{1}\mathbb{T}^{2}_{k}\times\{1\} via the identity map. It immediately follows that 𝒯(k)\mathscr{T}(k) consists of 9k+6=O(k)9k+6=O(k) tetrahedra and induces a two-triangle triangulation at each boundary torus.

Refer to caption
Figure 23: A minimal triangulation of 𝕋k2\mathbb{T}^{2}_{k} with 3k+23k+2 triangles, drawn for k=3k=3. The extension of the drawing for larger kk is straightforward.
Refer to caption
Figure 24: (i) The triangular prism (the shape described as t×[0,1]t\times[0,1], where tt denotes a triangle) can be triangulated with three tetrahedra. (ii) A triangulation of the triangular prism and its corresponding dual graph. (iii) A triangular prism with its top and bottom triangles identified. Reproduced from [27, Figure 13].