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

Extensions of Veech groups II: Hierarchical hyperbolicity and quasi-isometric rigidity

Spencer Dowdall Department of Mathematics, Vanderbilt University, Nashville, TN spencer.dowdall@vanderbilt.edu Matthew G. Durham Department of Mathematics, University of California, Riverside, CA mdurham@ucr.edu Christopher J. Leininger Department of Mathematics, Rice University, Houston, TX cjl12@rice.edu  and  Alessandro Sisto Department of Mathematics, Heriot-Watt University, Edinburgh, UK a.sisto@hw.ac.uk
Abstract.

We show that for any lattice Veech group in the mapping class group Mod(S)\mathrm{Mod}(S) of a closed surface SS, the associated π1S\pi_{1}S–extension group is a hierarchically hyperbolic group. As a consequence, we prove that any such extension group is quasi-isometrically rigid.

1. Introduction

This paper studies geometric properties of surface group extensions and how these relate to their defining subgroups of mapping class groups. Let SS be a closed, connected, oriented surface of genus at least 22. Recall that a π1S\pi_{1}S–extension of a group GG is a short exact sequence of the form

1π1SΓG1.1\to\pi_{1}S\to\Gamma\to G\to 1.

Such extensions are in bijective correspondence with monodromy homomorphisms from GG to the extended mapping class group Mod±(S)Out(π1S)\mathrm{Mod}^{\pm}(S)\cong\mathrm{Out}(\pi_{1}S) of the surface. Alternatively, these groups Γ\Gamma are precisely the fundamental groups of SS–bundles.

Many advances in the study of mapping class groups have been motivated by a longstanding but incomplete analogy between hyperbolic space n\mathbb{H}^{n} and the Teichmüller space 𝒯(S)\mathcal{T}(S) of a surface. In the theory of Kleinian groups, a discrete group of isometries of n\mathbb{H}^{n} is convex cocompact if it acts cocompactly on an invariant, convex subset. Farb and Mosher [FM02a] adapted this notion to mapping class groups by defining a subgroup GMod±(S)G\leq\mathrm{Mod}^{\pm}(S) to be convex cocompact if it acts cocompactly on a quasi-convex subset of 𝒯(S)\mathcal{T}(S). This has proven to be a fruitful concept with many interesting connections to, for example, the intrinsic geometry of the mapping class group [DT15, BBKL20], and its actions on the curve complex and the boundary of Teichmüller space [KL08a]. Most importantly, the work of Farb–Mosher [FM02a] and Hamenstädt [Ham] remarkably shows that an extension Γ\Gamma as above is word hyperbolic if and only if the associated monodromy GMod±(S)G\to\mathrm{Mod}^{\pm}(S) has finite kernel and convex cocompact image (see also [MS12]).

For Kleinian groups, convex cocompactness is a special case of a more prevalent phenomenon called geometric finiteness, which roughly amounts to acting cocompactly on a convex subset minus horoballs invariant by parabolic subgroups. In [Mos06], Mosher suggested this notion should have an analogous framework in mapping class groups that would extend the geometric connection with surface bundles to a larger class of examples. The prototypical candidates for geometric finiteness are the lattice Veech subgroups; these are special punctured-surface subgroups of Mod(S)\mathrm{Mod}(S) that arise naturally in the context of Teichmüller dynamics and whose corresponding SS–bundles are amenable to study via techniques from flat geometry.

Our prequel paper [DDLS21] initiated an analysis of the π1S\pi_{1}S–extensions associated to lattice Veech subgroups, with the main result being that each such extension Γ\Gamma admits an action on a hyperbolic space E^\hat{E} that captures much of the geometry of Γ\Gamma. Building on that work, the first main result of this paper is the following, which provides a concrete answer to [Mos06, Problem 6.2] for lattice Veech groups.

Theorem 1.1.

For any lattice Veech subgroup G<Mod(S)G<\mathrm{Mod}(S), the associated π1S\pi_{1}S–extension group Γ\Gamma of GG is a hierarchically hyperbolic group.

Hierarchical hyperbolicity means that in fact all the geometry of Γ\Gamma is robustly encoded by hyperbolic spaces. This is exactly the sort of relaxed hyperbolicity for π1S\pi_{1}S–extensions that one hopes should follow from a good definition of geometric finiteness in Mod(S)\mathrm{Mod}(S). Thus Theorem 1.1 suggests a possible general theory of geometric finiteness, which we expound upon in §1.4 below.

Hierarchical hyperbolicity has many strong consequences, some of which are detailed in §1.1 below. It also enables, via tools from [BHS21], the proof of our second main result, which answers [Mos06, Problem 5.4]:

Theorem 1.2.

For any lattice Veech group G<Mod(S)G<\mathrm{Mod}(S), the associated π1S\pi_{1}S–extension group Γ\Gamma of GG is quasi-isometrically rigid.

In [Mos06], Mosher in fact suggests an alternate approach to proving quasi-isometric rigidity that culminates in the formulation of Problem 5.4 of [Mos06] as an equivalent condition in this case. Both our proof and this alternate approach share a common key step of showing that quasi-isometries are coarsely fiber-preserving; for this we use tools from hierarchical hyperbolicity (see Proposition 5.4 below), whereas Mosher’s approach uses ideas from coarse algebraic topology appealing to the fact that GG is virtually free (see [FM00]). In §5.7 we give a rough sketch that carries out Mosher’s approach, drawing partly from his unpublished results in [Mos03], and leading to an alternate proof of Theorem 1.2.

It is our hope that the framework we develop for proving quasi-isometric rigidity for extension groups via hierarchical hyperbolicity will be applicable to a wider class of geometrically finite subgroups (§1.4), including those which may not be virtually free.

The rest of this introduction gives a more in-depth treatment of these results while elaborating on the concepts of, and connections between, hierarchical hyperbolicity, extensions of Veech groups, quasi-isometric rigidity, and geometric finiteness.

1.1. Hierarchical hyperbolicity

The notion of hierarchical hyperbolicity was defined by Behrstock, Hagen, and Sisto [BHS17b] and motivated by the seminal work of Masur and Minsky [MM00]. In short, it provides a framework and toolkit for understanding the coarse geometry of a space/group in terms of interrelated hyperbolic pieces. More precisely, a hierarchically hyperbolic space (HHS) structure on a metric space XX is a collection of hyperbolic spaces {𝒞(W)}W𝔖\{\mathcal{C}(W)\}_{W\in\mathfrak{S}}, arranged in a hierarchical fashion, in which any pair are nested \sqsubseteq, orthogonal \bot, or transverse \pitchfork, along with Lipschitz projections to and between these spaces that together capture the coarse geometry of XX. A hierarchically hyperbolic group (HHG) is then an HHS structure on a group that is equivariant with respect to an appropriate action on the union of hyperbolic spaces 𝒞(W)\mathcal{C}(W). See §4 for details or [BHS17b, BHS19, Sis19] for many examples and further discussion.

Showing that a space/group is a hierarchically hyperbolic gives access to several results regarding, for example, a coarse median structure and quadratic isoperimetric function [Bow18, Bow13], asymptotic dimension [BHS17a], stable and quasiconvex subsets and subgroups [ABD21, RST18], quasiflats [BHS21], bordifications and automorphisms [DHS17], and quasi-isometric embeddings of nilpotent groups [BHS17b]. In particular, the following is an immediate consequence of Theorem 1.1.

Corollary 1.3.

Let G<Mod(S)G<\mathrm{Mod}(S) be any lattice Veech group and Γ\Gamma the associated π1S\pi_{1}S–extension group. Then:

  1. (1)

    Γ\Gamma has quadratic Dehn function [Bow13].

  2. (2)

    Γ\Gamma is acylindrically hyperbolic and, moreover, its action on the \sqsubseteq–maximal hyperbolic space in the hierarchy is a universal acylindrical action [ABD21].

  3. (3)

    Γ\Gamma is semihyperbolic and thus has solvable conjugacy problem [DMS20, HHP20].

As discussed in §1.2 below, further information about Γ\Gamma can be gleaned from the specific HHG structure constructed in proving Theorem 1.1. We note that the \sqsubseteq–maximal hyperbolic space of this structure, and thus the universal acylindrical action indicated in Corollary 1.3(2), is simply the space E^\hat{E} from [DDLS21].

1.2. The HHG structure on Γ\Gamma

In order to describe the HHG structure more precisely and explain its connection to quasi-isometric rigidity in Theorem 1.2, we must first recall some of the structure of Veech groups and their extensions. Let G<Mod(S)G<\mathrm{Mod}(S) be a lattice Veech group and Γ=ΓG\Gamma=\Gamma_{G} the associated extension group. First note that (up to finite index) Γ\Gamma is naturally the fundamental group of an SS–bundle E¯/Γ\bar{E}/\Gamma over a compact surface with boundary (see §2 for details and notation). Each boundary component of E¯/Γ\bar{E}/\Gamma is virtually the mapping torus of a multi-twist on SS, and is thus a graph manifold: the tori in the JSJ decomposition are suspensions of the multi-twist curves.

Graph manifolds admit HHS structures [BHS19] where the maximal hyperbolic space is the Bass–Serre tree dual to the JSJ decomposition, and all other hyperbolic spaces are either quasi-lines or quasi-trees (obtained by coning off the boundaries of the universal covers of the base orbifolds of the Seifert pieces). The stabilizers of the vertices of the Bass–Serre trees are called vertex subgroups, and are precisely the fundamental groups of the Seifert pieces of the JSJ decomposition. We let 𝒱\mathcal{V} denote the disjoint union of the vertices of all Bass–Serre trees associated to the boundary components of the universal cover E¯\bar{E} of this SS–bundle. Given v,w𝒱v,w\in\mathcal{V}, we say that these vertices are adjacent if they are connected by an edge in the same Bass-Serre tree.

The HHG structure on the extension group Γ\Gamma may now be described as follows:

Theorem 1.4.

Suppose G<Mod(S)G<\mathrm{Mod}(S) is a lattice Veech group with extension group Γ\Gamma and let Υ1,,Υk<Γ\Upsilon_{1},\ldots,\Upsilon_{k}<\Gamma be representatives of the conjugacy classes of vertex subgroups. Then Γ\Gamma admits an HHG structure with the following set of hyperbolic spaces and relations among them (ignoring those of diameter 2\leq 2):

  1. (1)

    The maximal hyperbolic space E^\hat{E} is quasi-isometric to the Cayley graph of Γ\Gamma coned off along the cosets of Υ1,,Υk\Upsilon_{1},\ldots,\Upsilon_{k} [DDLS21].

  2. (2)

    There is a quasi-tree vqtv^{qt} and a quasi-line vqlv^{ql}, for each v𝒱v\in\mathcal{V}, and:

    1. (a)

      For all v𝒱v\in\mathcal{V}, vqtvqlv^{qt}\bot v^{ql}.

    2. (b)

      For all v,w𝒱v,w\in\mathcal{V}, if vv and ww are adjacent, then wqlvqlw^{ql}\bot v^{ql} and wqlvqtw^{ql}\sqsubseteq v^{qt}.

    3. (c)

      All other pairs are transverse.

This description of the HHG structure readily leads to further consequences for Γ\Gamma. For example, the maximal number of infinite-diameter pairwise orthogonal hyperbolic spaces is evidently 22. In view of [BHS17b, BHS21], we thus see that Γ\Gamma is as “close to hyperbolic” as possible in that its quasi-flats are at worst 22–dimensional:

Corollary 1.5.

Each top-dimensional quasi-flat in Γ\Gamma has dimension 22 and is contained in a finite-radius neighborhood of finitely many cosets of vertex subgroups.

We note that quasiflats will be crucial for our proof of quasi-isometric rigidity, and we remark that the analogous statement for graph manifolds is due to Kapovich–Leeb [KL97].

Recall that an element of a group is a generalized loxodromic if it acts loxodromically under some acylindrical action on a hyperbolic space, and that a universal acylindrical action on a hyperbolic space is one in which every generalized loxodromic acts loxodromically [ABD21]. It is shown in [Sis16] that a generalized loxodromic element gg of a finitely generated group is necessarily Morse, meaning that in any finite-valence Cayley graph for the group, any (K,C)(K,C)–quasi-geodesic with endpoints in the cyclic subgroup g\langle g\rangle stays within controlled distance M=M(K,C)M=M(K,C) of g\langle g\rangle. While being Morse is, in general, strictly weaker than being generalized loxodromic, these conditions are in fact equivalent in HHGs [ABD21, Theorem B].

In the case of our extension group Γ\Gamma, it follows from Corollary 1.3(2) that the generalized loxodromics and Morse elements are precisely those elements acting loxodromically on E^\hat{E}. In [DDLS21, Theorem 1.1] we characterized these elements in terms of the vertex subgroups of Γ\Gamma, thus yielding the following:

Corollary 1.6.

Let Γ\Gamma be a lattice Veech group extension with vertex subgroups Υ1,,Υk\Upsilon_{1},\dots,\Upsilon_{k} as in Theorem 1.4. The following are equivalent for an infinite order element γΓ\gamma\in\Gamma:

  • γ\gamma is not conjugate into any of the vertex subgroups Υi\Upsilon_{i}

  • γ\gamma is a generalized loxodromic element of Γ\Gamma

  • γ\gamma is a Morse element of Γ\Gamma.

1.3. Quasi-isometric rigidity

To state our rigidity theorem, first recall that Γ\Gamma is (up to finite index) the fundamental group of an SS-bundle E¯/Γ\widebar{E}/\Gamma over a compact surface with boundary. Here E¯\bar{E} is a Γ\Gamma–invariant truncation of the universal S~\tilde{S}–bundle over the Teichmüller disk stabilized by the Veech group GG. In particular, E¯\bar{E} is quasi-isometric to Γ\Gamma. Let Isom(E¯)\operatorname{Isom}(\widebar{E}) and QI(E¯)\operatorname{QI}(\widebar{E}) denote the isometry and quasi-isometry groups of E¯\widebar{E}, respectively, and let Isomfib(E¯)Isom(E¯)\mathrm{Isom_{fib}}(\widebar{E})\leq\operatorname{Isom}(\widebar{E}) denote the subgroup of isometries that map fibers to fibers.

Theorem 1.7.

There is an allowable truncation E¯\bar{E} of EE such that the natural homomorphisms Isomfib(E¯)Isom(E¯)QI(E¯)QI(Γ)\mathrm{Isom_{fib}}(\bar{E})\to\operatorname{Isom}(\bar{E})\to\operatorname{QI}(\bar{E})\cong\operatorname{QI}(\Gamma) are all isomorphisms, and ΓIsom(E¯)QI(Γ)\Gamma\leq\operatorname{Isom}(\bar{E})\cong\operatorname{QI}(\Gamma) has finite index.

This is an analog, and indeed was motivated by, Farb and Mosher’s [FM02b] theorem that in the case of a surface group extension ΓH\Gamma_{H} associated to a Schottky subgroup HH of Mod(S)\mathrm{Mod}(S), the natural homomorphism ΓHQI(ΓH)\Gamma_{H}\to\operatorname{QI}(\Gamma_{H}) is injective with finite cokernel. This rigidity also leads to the following strong algebraic consequence:

Corollary 1.8.

If HH is any finitely generated group quasi-isometric to Γ\Gamma, then HH and Γ\Gamma are weakly commensurable.

In the statement, recall that two groups H1,H2H_{1},H_{2} are weakly commensurable if there are finite normal subgroups NiHiN_{i}\lhd H_{i} so that the quotients Hi/NiH_{i}/N_{i} have a pair of finite-index subgroups that are isomorphic to each other.

1.4. Motivation and Geometric Finiteness

Before outlining the paper and providing some ideas about the proofs, we provide some speculative discussion. For Kleinian groups—that is, discrete groups of isometries of hyperbolic 33–space—the notion of geometric finiteness is important in the deformation theory of hyperbolic 33–manifolds by the work of Ahlfors [Ahl66] and Greenberg [Gre66]. While the definition has many formulations (see [Mar74, Mas70, Thu86, Bow93]), roughly speaking a group is geometrically finite if it acts cocompactly on a convex subset of hyperbolic 33–space minus a collection of horoballs that are invariant by parabolic subgroups. When there are no parabolic subgroups, geometric finiteness reduces to convex cocompactness: a cocompact action on a convex subset of hyperbolic 33–space.

While there is no deformation theory for subgroups of mapping class groups, Farb and Mosher [FM02a] introduced a notion of convex cocompactness for G<Mod(S)G<\mathrm{Mod}(S) in terms of the action on Teichmüller space 𝒯(S)\mathcal{T}(S). Their definition requires that GG acts cocompactly on a quasi-convex subset for the Teichmüller metric, while Kent and Leininger later proved a variety of equivalent formulations analogous to the Kleinian setting [KL07, KL08a, KL08b]. Farb and Mosher proved that convex cocompactness is equivalent to hyperbolicity of the associated extension group ΓG\Gamma_{G} (with monodromy given by inclusion) when GG is virtually free. This equivalence was later proven in general by Hamenstädt [Ham] (see also Mj–Sardar [MS12]), though at the moment the only known examples are virtually free.

The coarse nature of Farb and Mosher’s formulation reflects the fact that the Teichmüller metric is far less well-behaved than that of hyperbolic 33–space. Quasi-convexity in the definition is meant to help with the lack of nice local behavior of the Teichmüller metric. It also helps with the global lack of Gromov hyperbolicity (see Masur–Wolf [MW95]), as cocompactness of the action ensures that the quasi-convex subset in the definition is Gromov hyperbolic (see Kent–Leininger [KL08a], Minsky [Min96b], and Rafi [Raf14]).

The inclusion of reducible/parabolic mapping classes in a subgroup G<Mod(S)G<\mathrm{Mod}(S) brings the thin parts of 𝒯(S)\mathcal{T}(S) into consideration; these subspaces contain higher rank quasi-flats and even exhibit aspects of positive curvature (see Minsky [Min96a]). This is a main reason why extending the notion of convex cocompactness to geometric finiteness is complicated. These complications are somewhat mitigated in the case of lattice Veech groups. Such subgroups are stabilizers of isometrically and totally geodesically embedded hyperbolic planes, called Teichmüller disks, that have finite area quotients. Thus, the intrinsic hyperbolic geometry agrees with the extrinsic Teichmüller geometry, and as a group of isometries of the hyperbolic plane, a lattice Veech group is geometrically finite. This is why these subgroups serve as a test case for geometric finiteness in the mapping class group. This is also why a subgroup of a Veech group is convex cocompact in Mod(S)\mathrm{Mod}(S) if and only if it is convex cocompact as a group of isometries of the hyperbolic plane (which also happens if and only if it is finitely generated and contains no parabolic elements).

The action of Mod(S)\mathrm{Mod}(S) on the curve graph, which is Gromov hyperbolic by work of Masur–Minsky [MM99], provides an additional model for these considerations. Specifically, convex cocompactness is equivalent to the orbit map to the curve graph 𝒞(S)\mathcal{C}(S) being a quasi-isometric embedding with respect to the word metric from a finite generating set (see Kent–Leininger [KL08a] and Hamenstädt [Ham]). Viewing geometric finiteness as a kind of “relative convex cocompactness” for Kleinian groups suggests an interesting connection with the curve complex formulation. The connection is best illustrated by the following theorem of Tang [Tan19].

Theorem 1.9 (Tang).

For any lattice Veech group G<Mod(S)G<\mathrm{Mod}(S) stabilizing a Teichmüller disk D𝒯(S)D\subset\mathcal{T}(S), there is a GG–equivariant quasi-isometric embedding Del𝒞(S)D^{el}\to\mathcal{C}(S), where DelD^{el} is the path metric space obtained from DD by coning off the GG–invariant family of horoballs in which DD ventures into the thin parts of 𝒯(S)\mathcal{T}(S).

Farb [Far98] showed that non-cocompact lattices in the group of isometries of hyperbolic space are relatively hyperbolic relative to the parabolic subgroups. For Veech groups, the space DelD^{el} is quasi-isometric to the (hyperbolic) coned off Cayley graph, illustrating (part of) the relative hyperbolicity of GG. We thus propose a kind of “qualified” notion of geometric finiteness with this in mind:

Definition 1.10 (Parabolic geometric finiteness).

A finitely generated subgroup G<Mod(S)G<\mathrm{Mod}(S) is parabolically geometrically finite if GG is relatively hyperbolic, relative to a (possibly trivial) collection of subgroups ={H1,,Hk}\mathcal{H}=\{H_{1},\ldots,H_{k}\}, and

  1. (1)

    HiH_{i} contains a finite index, abelian subgroup consisting entirely of multitwists, for each 1ik1\leq i\leq k; and

  2. (2)

    the coned off Cayley graph GG–equivariantly and quasi-isometrically embeds into 𝒞(S)\mathcal{C}(S).

When ={{id}}\mathcal{H}=\{\{id\}\}, we note that the condition is equivalent to GG being convex cocompact. By Theorem 1.9, lattice Veech groups are parabolically geometrically finite. In fact, Tang’s result is more general and implies that any finitely generated Veech group satisfies this definition. These examples are all virtually free, but other examples include the combination subgroups of Leininger–Reid [LR06], which are isomorphic to fundamental groups of closed surfaces of higher genus, and free products of higher rank abelian groups constructed by Loa [Loa21].

In view of Theorem 1.1, one might formulate the following.

Conjecture 1.11.

Let G<Mod(S)G<\mathrm{Mod}(S) be parabolically geometrically finite. Then the π1S\pi_{1}S–extension group Γ\Gamma of GG is a hierarchically hyperbolic group.

We view Definition 1.10 as only a qualified formulation because there are many subgroups of Mod(S)\mathrm{Mod}(S) that are not relatively hyperbolic but are nevertheless candidates for being geometrically finite in some sense. It is possible that there are different types of geometric finiteness for subgroups of mapping class groups, with Definition 1.10 being among the most restrictive. Other notions might include an HHS structure on the subgroup which is compatible with the ambient one on Mod(S)\mathrm{Mod}(S) (e.g., hierarchical quasiconvexity [BHS19]). From this perspective, some candidate subgroups that may be considered geometrically finite include:

  • the whole group Mod(S)\mathrm{Mod}(S);

  • multi-curve stabilizers;

  • the right-angled Artin subgroups of mapping class groups constructed in [CLM12, Kob12, Run20];

  • free and amalgamated products of other examples.

Question 1.12.

For each example group GMod(S)G\leq\mathrm{Mod}(S) above, is the associated extension ΓG\Gamma_{G} a hierarchically hyperbolic group?

We note that the answer is ‘yes’ for the first example, since the extension group is the mapping class group of the surface SS with a puncture. Moreover, since our work on this subject first appeared, Russell [Rus21] addressed the second example by proving extensions of multicurve stabilizers are hierarchically hyperbolic groups.


1.5. Outline and proofs

Let us briefly outline the paper and comment on the main structure of the proofs. In §2 we review necessary background material and introduce the objects and notation that will be used throughout the paper. In particular, we define the spaces EE and E¯\bar{E}, the latter being a quasi-isometric model for the Veech group extension Γ\Gamma, as well as the hyperbolic collapsed space E^\hat{E}. All of these were constructed in [DDLS21].

In §§34 we prove that the extension group Γ\Gamma is hierarchically hyperbolic by utilizing a combinatorial criterion from [BHMS20]. Besides hyperbolicity of E^\hat{E}, the other hard part of the criterion is an analogue of Bowditch’s fineness condition from the context of relative hyperbolicity. Its geometric interpretation is roughly that two cosets of vertex subgroups as above have bounded coarse intersection, aside from the “obvious” exception when the cosets correspond to vertices of the same Bass–Serre tree within distance 2 of each other. To this end, in §3 we associate to each vertex v𝒱v\in\mathcal{V} a spine bundle 𝚯vE¯{\bf\Theta}^{v}\subset\bar{E}, which corresponds to a Seifert piece of the JSJ decomposition of the peripheral graph manifold, along with a pair of hyperbolic spaces 𝒦v\mathcal{K}^{v} and 𝚵v{\bf\Xi}^{v} that will figure into the HHS structure on Γ\Gamma. The space 𝒦v\mathcal{K}^{v} is obtained via a quasimorphism constructed using the Seifert fibered structure following ideas in forthcoming work of the fourth author with Hagen, Russell, and Spriano [HRSS21], while 𝚵v{\bf\Xi}^{v} is coarsely obtained by coning off boundary components of the universal covers of the base 22–orbifold of this Seifert fibered manifold. We then appeal to the flat geometry of the fibers of EE to construct and study certain projection maps

E¯\bar{E}𝚯v{\bf\Theta}^{v}𝒦v\mathcal{K}^{v}𝚵v{\bf\Xi}^{v}Πv\Pi^{v}λv\lambda^{v}ivi^{v}Λv\Lambda^{v}ξv\xi^{v}

and prove that various pairs of subspaces of E¯\bar{E} have bounded projection onto each other (Proposition 3.19).

In §4, we begin assembling the combinatorial objects necessary to apply the HHG criterion from [BHMS20], which involves both combinatorial and geometric aspects. The first step involves the construction of a natural flag complex 𝒳\mathcal{X} containing the union of the Bass-Serre trees, together with appropriate “subjoins” with the union of all 𝒦v\mathcal{K}^{v}, over v𝒱v\in\mathcal{V}. Next, we use the geometry of E¯\bar{E} to construct a certain graph 𝒲\mathcal{W} whose vertices are maximal simplices of 𝒳\mathcal{X} and on which Γ\Gamma acts metrically properly and coboundedly. The remainder of this section is devoted to verifying the necessary combinatorial conditions as well as translating the facts about 𝒦v\mathcal{K}^{v} and 𝚵v{\bf\Xi}^{v} and the projections described above into proofs of the necessary geometric conditions. We note that in the combinatorial HHG setup, the complex 𝒳\mathcal{X} comes with its own hierarchy projections between the induced hyperbolic spaces (Definitions 4.94.10), which may be different than the projections to 𝒦v\mathcal{K}^{v} and 𝚵v{\bf\Xi}^{v}.

In §5 we prove our QI-rigidity result Theorem 1.7. The starting point is the hierarchical hyperbolicity of Γ\Gamma provided by Theorem 1.4, as it gives access to the results and arguments in [BHS21] about the preservation of quasi-isometrically embedded flats. Every collection of pairwise orthogonal hyperbolic spaces in an HHG determines a natural product subspace, with the maximal standard quasi-isometrically embedded flats (or orthants) arising inside such subspaces as products of quasi-lines in a maximal collection of pairwise orthogonal hyperbolic spaces of the HHG. Theorem A of [BHS21] states that a quasi-isometry of an HHS preserves the structure of its quasi-flats and takes any maximal quasi-flat within bounded Hausdorff distance of the union of standard maximal orthants. The maximal quasi-flats in the HHG structure on E¯\widebar{E}, namely the 22–dimensional flats indicated in Corollary 1.5, are encoded by certain strip bundles that, roughly, correspond to flats in the peripheral graph manifolds. We use the preservation of the maximal quasi-flats to derive coarse preservation of these strip bundles, which we then upgrade to coarse preservation of the fibers (§5.1). By using tools of flat geometry from [BL18, DELS18], we then show any quasi-isometry induces an affine homeomorphism of any fiber to itself (§§5.25.3) and moreover that this assignment is injective (§5.4). Finally, we show this association is an isomorphism by proving (§5.5) that every affine homeomorphism of a fiber induces an isometry and hence quasi-isometry of E¯\bar{E}. Quasi-isometric rigidity and its algebraic consequence Corollary 1.8 are then easily obtained in §5.6.

Acknowledgments

The authors would like to thank MSRI and its Fall 2016 program on Geometric Group Theory, where this work began. We also gratefully acknowledge NSF grants DMS 1107452, 1107263, 1107367 (the GEAR Network) for supporting travel related to this project. Dowdall was partially supported by NSF grants DMS-1711089 and DMS-2005368. Durham was partially supported by NSF grant DMS-1906487. Leininger was partially supported by NSF grants DMS-1510034, DMS-1811518, and DMS-2106419. Sisto was partially supported by the Swiss National Science Foundation (grant #182186). The authors would like to thank the anonymous referee for their very helpful comments on the first version of this paper.

2. Setup: The groups and spaces

Here we briefly recall the basic set up from [DDLS21] which we will use throughout the remainder of the paper. We refer the reader to Sections 2 and 3 of that paper for details and precise references.

2.1. Flat metrics and Veech groups

Fix a closed surface of genus at least 22, a complex structures X0X_{0} (viewed as a point in the Teichmüller space 𝒯(S)\mathcal{T}(S)), and a nonzero holomorphic quadratic differential qq on (S,X0)(S,X_{0}). Integrating a square root of qq determines preferred coordinates on (S,X0)(S,X_{0}) for qq which defines a translation structure (in the complement of the isolated zeros of qq). We also write qq for the associated flat metric defined by the half-translation structure (though the metric only determines the half-translation structure or quadratic differential up to a complex scalar multiple). This metric is a non-positively curved Euclidean cone metric, with cone singularities at the zeros of qq. The orbit of (X0,q)(X_{0},q) under the natural SL2()\rm SL_{2}(\mathbb{R}) action on quadratic differentials projects to a Teichmüller disk, D=Dq𝒯(S)D=D_{q}\subset\mathcal{T}(S), which we equip with its Poincaré metric ρ\rho. The circle at infinity of DD is naturally identified with the projective space of directions, 1(q)\mathbb{P}^{1}(q), in the tangent space of any nonsingular point of qq. For α1(q)\alpha\in\mathbb{P}^{1}(q), we write (α)\mathcal{F}(\alpha) for the singular foliation by geodesics in direction α\alpha.

We assume that the associated Veech group G=GqG=G_{q} is a lattice—recall that GG can be viewed as the stabilizer in the mapping class group of SS of DD as well as the affine group of qq, and the lattice assumption is equivalent to requiring the quotient orbifold D/GD/G to have finite ρ\rho–area. The parabolic fixed points in the circle at infinity form a subset we denote 𝒫1(q){\mathcal{P}}\subset\mathbb{P}^{1}(q). This subset corresponds precisely to the completely periodic directions for the flat metric qq; that is, the directions α\alpha for which the foliation (α)\mathcal{F}(\alpha) decomposes SS into cylinders foliated by qq–geodesic core circles. The boundaries of these cylinders are qq–saddle connections (qq–geodesic segments connecting pairs of cone points, with no cone points in their interior), and by the Veech Dichotomy, every saddle connection is in a direction in 𝒫{\mathcal{P}}. We let {Bα}α𝒫\{B_{\alpha}\}_{\alpha\in{\mathcal{P}}} denote any GG–invariant, 11–separated set of horoballs in DD and let

D¯=Dα𝒫int(Bα)\bar{D}=D\smallsetminus\bigcup_{\alpha\in{\mathcal{P}}}{\rm{int}}(B_{\alpha})

be the GG–invariant subspace obtained by removing these horoballs. We write ρ¯\bar{\rho} for the induced path metric on D¯\bar{D}. Finally, we let

p:DD^p\colon D\to\hat{D}

be the GG–equivariant quotient obtained by collapsing each horoball α{\mathcal{B}}_{\alpha} to a point, for α𝒫\alpha\in{\mathcal{P}}. There is a natural path metric ρ^\hat{\rho} on D^\hat{D} so that pp is 11–Lipschitz and is a local isometry at every point not in one of the horoballs.

We will also make use of the closest point projection to the horoball

cα:DBαc_{\alpha}\colon D\to B_{\alpha}

for each α𝒫\alpha\in{\mathcal{P}}.

2.2. The bundles EE and E¯\bar{E}.

For each point XDX\in D, we let qXq_{X} denote the associated flat metric or quadratic differential (defined up to scalar multiplication) on SS. The space of interest EE is a bundle over DD,

π:ED,\pi\colon E\to D,

for which the fiber EXE_{X} over XDX\in D is naturally identified with the universal cover S~\widetilde{S} of SS, equipped with the pull-back complex structure XX and quadratic differential/flat metric qXq_{X}. We write α=π1(Bα){\mathcal{B}}_{\alpha}=\pi^{-1}(B_{\alpha}) for α𝒫\alpha\in{\mathcal{P}}.

For any X,YDX,Y\in D, the Teichmüller map between these complex structures has initial and terminal quadratic differentials qXq_{X} and qYq_{Y} (up to scalar multiple) and this map lifts to a canonical affine map between the fibers fY,X:EXEYf_{Y,X}\colon E_{X}\to E_{Y}. These maps satisfy fZ,X=fZ,YfY,Xf_{Z,X}=f_{Z,Y}f_{Y,X} for all X,Y,ZDX,Y,Z\in D, and for any XDX\in D, assemble to a map fX:EEXf_{X}\colon E\to E_{X} defined by fX(y)=fX,π(y)(y)f_{X}(y)=f_{X,\pi(y)}(y). Moreover, for any X,YDX,Y\in D, fY,Xf_{Y,X} is eρ(X,Y)e^{\rho(X,Y)}–bi-Lipschitz. We use the maps fX,X0f_{X,X_{0}} to identify 1(q)1(qX)\mathbb{P}^{1}(q)\cong\mathbb{P}^{1}(q_{X}) for all XDX\in D.

The fiber over X0X_{0} is denoted E0=EX0E_{0}=E_{X_{0}} and the maps f0=fX0:EE0f_{0}=f_{X_{0}}\colon E\to E_{0} and π:ED\pi\colon E\to D are projections on the factors in a product structure ED×E0D×S~E\cong D\times E_{0}\cong D\times\widetilde{S}. For xEx\in E, we write Dx=fπ(x)1(x)D_{x}=f_{\pi(x)}^{-1}(x), which is just the slice D×{f0(x)}D\times\{f_{0}(x)\} in the product structure. The affine maps fY,Xf_{Y,X} sends the cone points ΣX\Sigma_{X} of EXE_{X} to the cone points ΣY\Sigma_{Y} of EYE_{Y}. Consequently, the union of all singular points

Σ=XDΣX\Sigma=\bigcup_{X\in D}\Sigma_{X}

is a locally finite union of disks DxD_{x}, one for each xΣ0=ΣX0x\in\Sigma_{0}=\Sigma_{X_{0}}.

We give the space EE a singular Riemannian metric dd which is the flat metric on each fiber EXE_{X} and the Poincaré metric on each diskDxD_{x} so that at each smooth point of intersection, the tangent planes are orthogonal. The singular locus of this metric is precisely Σ\Sigma. Each disk DxD_{x} is isometrically embedded since π\pi is a 11–Lipschitz map, and hence restricts to an isometry π|Dx:DxD\pi|_{D_{x}}\colon D_{x}\to D. The metric on EΣE\smallsetminus\Sigma is in fact a locally homogeneous metric, modeled on a four-dimensional, Thurston-type geometry; see [DDLS21, §5].

The extension group Γ\Gamma acts on EE by bundle maps with the kernel π1S<Γ\pi_{1}S<\Gamma of the projection to GG acting trivially on DD and by covering transformation on each fiber EXE_{X}. We set E¯=π1(D¯)E\bar{E}=\pi^{-1}(\bar{D})\subset E, and write π¯:E¯D¯\bar{\pi}\colon\bar{E}\to\bar{D}. When convenient to do so, we put “bars” over objects associated to D¯\bar{D} or E¯\bar{E}, e.g. D¯x=DxE¯\bar{D}_{x}=D_{x}\cap\bar{E}, p¯:D¯D^\bar{p}\colon\bar{D}\to\hat{D}, etc. In particular, we write d¯\bar{d} for the induced path metric on E¯E\bar{E}\subset E, induced from the metric on EE described above.

For any α𝒫\alpha\in{\mathcal{P}}, the closest point projection cα:DBαc_{\alpha}\colon D\to B_{\alpha} has a useful “lift” fα:Eαf_{\alpha}\colon E\to{\mathcal{B}}_{\alpha}, defined by

fα(x)=fcα(π(x))(x),f_{\alpha}(x)=f_{c_{\alpha}(\pi(x))}(x),

for any xE¯x\in\bar{E}. That is, fαf_{\alpha} maps each fiber EXE_{X} via the map fY,Xf_{Y,X} to EYE_{Y}, where Y=cα(X)Y=c_{\alpha}(X) is the image of the closest point projection to BαB_{\alpha} of XX in DD.

2.3. The hyperbolic space E^\hat{E}

The quotient p:DD^p\colon D\to\hat{D} is the descent of a quotient P:EE^P\colon E\to\hat{E} which we now describe. First, for each α𝒫\alpha\in{\mathcal{P}}, the foliation (α)\mathcal{F}(\alpha) lifts to a foliation on E0E_{0} in direction α\alpha, and hence on any fiber EXE_{X} by push-forward via the map fX,X0f_{X,X_{0}}, also in direction α\alpha (via the identification 1(q)1(qX)\mathbb{P}^{1}(q)\cong\mathbb{P}^{1}(q_{X})). There is a natural transverse measure coming from the flat metric on XX. Given α𝒫\alpha\in{\mathcal{P}}, we fix some XαBαX_{\alpha}\in\partial B_{\alpha} and let TαT_{\alpha} be the dual simplicial \mathbb{R}–tree to this measured foliation in direction α\alpha on EXαE_{X_{\alpha}}, and we let tα:ETαt_{\alpha}\colon E\to T_{\alpha} be the composition of the leafspace projection EXαTαE_{X_{\alpha}}\to T_{\alpha} with the map fXα:EEXαf_{X_{\alpha}}\colon E\to E_{X_{\alpha}}.

Now we define P:EE^P\colon E\to\hat{E} to be the quotient space obtained by collapsing the subset α{\mathcal{B}}_{\alpha} to TαT_{\alpha} via tα|αt_{\alpha}|_{{\mathcal{B}}_{\alpha}} for each α𝒫\alpha\in{\mathcal{P}}. We also write P¯=P|E¯:E¯E^\bar{P}=P|_{\bar{E}}\colon\bar{E}\to\hat{E}. The maps PP and P¯\bar{P} descend to the maps pp and p¯\bar{p}, and the map π\pi determines maps π^\hat{\pi} and π¯\bar{\pi}, which all fit into the following commutative diagram.

E{E}E¯{\bar{E}}E^{\hat{E}}D{D}D¯{\bar{D}}D^.{\hat{D}.}P\scriptstyle{P}π\scriptstyle{\pi}π¯\scriptstyle{\bar{\pi}}P¯\scriptstyle{\bar{P}}π^\scriptstyle{\hat{\pi}}p\scriptstyle{p}p¯\scriptstyle{\bar{p}}

A metric d^\hat{d} on E^\hat{E} is determined by d¯\bar{d} on E¯\bar{E} and the map P¯\bar{P}. The main facts about this metric are summarized in the following theorem; see [DDLS21, Theorem 1.1, Lemma 3.2].

Theorem 2.1.

There is a Gromov hyperbolic path metric d^\hat{d} on E^\hat{E} so that P¯:E¯E^\bar{P}\colon\bar{E}\to\hat{E} is 11–Lipschitz and is a local isometry at every point xE¯E¯x\in\bar{E}-\partial\bar{E}. Furthermore, for every α𝒫\alpha\in{\mathcal{P}},

  • The induced path metric on P(α)=TαP(\partial{\mathcal{B}}_{\alpha})=T_{\alpha} is the \mathbb{R}–tree metric determined by the transverse measure on the foliation of EXαE_{X_{\alpha}} in direction α\alpha.

  • The subspace topology on TαE^T_{\alpha}\subset\hat{E} agrees with the \mathbb{R}–tree topology on TαT_{\alpha}.

Remark 2.2.

The underlying simplicial tree TαT_{\alpha} is precisely the Bass-Serre tree dual to the splitting of π1S\pi_{1}S defined by the cores of the cylinders of (α)\mathcal{F}(\alpha) on SS.

For each xEx\in E, we denote the image of DxD_{x} in E^\hat{E} by D^x\hat{D}_{x}, which is obtained by collapsing αDx{\mathcal{B}}_{\alpha}\cap D_{x} to a point, for each α𝒫\alpha\in{\mathcal{P}}. Consequently, π^|D^x:D^xD^\hat{\pi}|_{\hat{D}_{x}}\colon\hat{D}_{x}\to\hat{D} is a bijection, and so each D^x\hat{D}_{x}, with its path metric, is isometric to D^\hat{D} and isometrically embedded in E^\hat{E}. We call objects in EE, E¯\bar{E}, and E^\hat{E} vertical if they are contained in a fiber of π\pi, π¯\bar{\pi}, or π^\hat{\pi}, respectively, and horizontal if they are contained in DxD_{x}, D¯x\bar{D}_{x}, or D^x\hat{D}_{x}, for some xE,E¯x\in E,\bar{E}.

2.4. Vertices, spines, and spine bundles

We will write 𝒱E^\mathcal{V}\subset\hat{E} for the union over all α𝒫\alpha\in{\mathcal{P}} of all vertices of TαT_{\alpha}. We will simultaneously view 𝒱\mathcal{V} as both a subset of E^\hat{E} and abstractly as an indexing set that will be used in sections §§34 to develop an HHS structure on E¯\bar{E}. Since each vertex belongs to a unique tree, and since the trees are indexed by α𝒫\alpha\in{\mathcal{P}}, we obtain a map α:𝒱𝒫\alpha\colon\mathcal{V}\to{\mathcal{P}} so that vv is a vertex of Tα(v)T_{\alpha(v)}. For convenience, we also write Bv=Bα(v)B_{v}=B_{\alpha(v)}, Bv=Bα(v)\partial B_{v}=\partial B_{\alpha(v)}, etc for each v𝒱v\in\mathcal{V}, and write cv=cα(v)c_{v}=c_{\alpha(v)} for the ρ\rho–closest point projection DBvD\to B_{v}.

For v,w𝒱v,w\in\mathcal{V}, we write vwv\parallel w if α(v)=α(w)\alpha(v)=\alpha(w). Then define dtree(v,w)0{}d_{\mathrm{tree}}(v,w)\in\mathbb{Z}_{\geq 0}\cup\{\infty\} to be the combinatorial (integer valued) distance in the simplicial tree Tα(v)=Tα(w)T_{\alpha(v)}=T_{\alpha(w)} when vwv\parallel w (as opposed to the distance from the \mathbb{R}–tree metric) and to equal \infty when v∦wv\not\parallel w.

Given α𝒫\alpha\in{\mathcal{P}}, XDX\in D, and vTαv\in T_{\alpha}, the vv–spine in EXE_{X} is the subspace

θXv=(PfXα,X)1(v)=tα1(v)EX.\theta^{v}_{X}=(P\circ f_{X_{\alpha},X})^{-1}(v)=t_{\alpha}^{-1}(v)\cap E_{X}.

The vv–spine θXv\theta^{v}_{X} is the union of the saddle connections on the fiber EXE_{X} in direction α\alpha that project to vv by tαt_{\alpha}. When dtree(v,w)=1d_{\mathrm{tree}}(v,w)=1 (and hence v,wv,w are adjacent in the same tree TαT_{\alpha}) there is a unique component of EX(θXvθXw)E_{X}\smallsetminus(\theta^{v}_{X}\cup\theta^{w}_{X}) whose closure is an infinite strip, ×[a,b]\mathbb{R}\times[a,b], that covers a maximal cylinder in the quotient EX/π1S=(S,X,qX)E_{X}/\pi_{1}S=(S,X,q_{X}) in the direction α\alpha. We let 𝚯Xv{\bf\Theta}^{v}_{X} be the union of θXv\theta^{v}_{X} and all such strips defined by wTαw\in T_{\alpha} with dtree(v,w)=1d_{\mathrm{tree}}(v,w)=1. We call 𝚯Xv{\bf\Theta}^{v}_{X} the thickened vv–spine in EXE_{X}. In the special case X=X0X=X_{0}, we write θ0v=θX0v\theta^{v}_{0}=\theta^{v}_{X_{0}} and 𝚯0v=𝚯X0v{\bf\Theta}^{v}_{0}={\bf\Theta}^{v}_{X_{0}}. Observe that the affine map fY,Xf_{Y,X} maps θXv\theta^{v}_{X} and 𝚯Xv{\bf\Theta}^{v}_{X} to θYv\theta^{v}_{Y} and 𝚯Yv{\bf\Theta}^{v}_{Y}, respectively, for all X,YDX,Y\in D. Finally, we write

θv=XBvθXv𝚯v=XBv𝚯Xv.\theta^{v}=\bigcup_{X\in\partial B_{v}}\theta^{v}_{X}\quad\quad{\bf\Theta}^{v}=\bigcup_{X\in\partial B_{v}}{\bf\Theta}^{v}_{X}.

These spaces are bundles over Bv\partial B_{v} which we call, respectively, the vv–spine bundle and the thickened vv–spine bundle.

2.5. Schematic of the space E¯\bar{E} and its important pieces.

Refer to caption
Figure 1. A schematic of E¯\bar{E} and various key features of it.

Figure 1 is a cartoon of the bundle E¯\bar{E} over the truncated Teichmüller disk D¯\bar{D}. We have tried to highlight some of the key features of E¯\bar{E} which are relevant to this paper.

  • (a)

    The stabilizer of a horoball based at a point β𝒫\beta\in\mathcal{P} is virtually cyclic, generated by a multitwist τβ\tau_{\beta} acting as a parabolic on DD. The base point XβX_{\beta} on the horocycle based at β\beta and its image are shown.

  • (b)

    The bundle over the boundary horocycle based at β\beta is shown. This is the universal cover, β\partial\mathcal{B}_{\beta}, of a graph manifold which is the mapping torus of τβ\tau_{\beta}. Two fibers EXβE_{X_{\beta}} and Eτβ(Xβ)E_{\tau_{\beta}(X_{\beta})} are shown with the effect on a part of a spine (in green) in some other direction illustrating the sheering in strips after applying τβ\tau_{\beta}.

  • (c)

    This is another horoball in some direction α\alpha, with the chosen basepoint XαX_{\alpha} and its horocycle Bα\partial B_{\alpha}.

  • (d)

    The spine θXαv\theta^{v}_{X_{\alpha}} in direction α\alpha is shown in red, corresponding to a vertex vTαv\in T_{\alpha}. The thickened spine 𝚯Xαv{\bf\Theta}^{v}_{X_{\alpha}} is indicated in lavender. Spines for vertices of TαT_{\alpha} adjacent to vv meet 𝚯Xαv{\bf\Theta}^{v}_{X_{\alpha}} along lines in 𝚯Xαv\partial{\bf\Theta}^{v}_{X_{\alpha}} and are shown in various other colors.

  • (e)

    The restriction of tα:ETαt_{\alpha}\colon E\to T_{\alpha} to EXαE_{X_{\alpha}} collapses each spine θXαw\theta^{w}_{X_{\alpha}} or strip in direction α\alpha to the corresponding vertex ww or edge the Bass-Serre tree TαT_{\alpha}. The space E^\hat{E} is formed by collapsing α\mathcal{B}_{\alpha} to TαT_{\alpha} via tαt_{\alpha}.

2.6. Some technical lemmas and coarse geometry

Here we briefly recall some basic facts about the setup above proved in [DDLS21] as well as some useful coarse geometric facts. The first fact is the following; see [DDLS21, Lemma 3.4].

Lemma 2.3.

There exists a constant M>0M>0 such that for each v𝒱v\in\mathcal{V} and XBvX\in\partial B_{v}, every saddle connection in θXv\theta^{v}_{X} has length at most MM and every strip in 𝚯Xv{\bf\Theta}^{v}_{X} has width at most MM. In particular, for points XBαX\in\partial B_{\alpha}, the saddle connections and strips of EXE_{X} in direction α𝒫\alpha\in{\mathcal{P}} have, respectively, uniformly bounded lengths and widths.

Every connected graph can be made into a geodesic metric space by locally isometrically identifying each edge with a unit interval. We will need the following well-known result (for a proof of this version, see [DDLS21, Proposition 2.1]).

Proposition 2.4.

Let Ω\Omega be a path metric space and ΥΩ\Upsilon\subset\Omega an RR–dense subset for some R>0R>0. For any R>3RR^{\prime}>3R, consider a graph 𝒢\mathcal{G} with vertex set Υ\Upsilon such that:

  • all pairs of elements of Υ\Upsilon within distance 3R3R are joined by an edge in 𝒢\mathcal{G},

  • if an edge in 𝒢\mathcal{G} joins points w,wΥw,w^{\prime}\in\Upsilon, then dΩ(w,w)Rd_{\Omega}(w,w^{\prime})\leq R^{\prime}.

Then the inclusion of Υ\Upsilon into Ω\Omega extends to a quasi-isometry 𝒢Ω\mathcal{G}\to\Omega.

The following criterion for a graph to be a quasi-tree is well-known, and an easy consequence of Manning’s bottleneck criterion [Man05]. We include a proof for completeness.

Proposition 2.5.

Let XX be a graph, and suppose that there exists a constant BB with the following property: For each pair of vertices w,ww,w^{\prime} there exists an edge path γ(w,w)\gamma(w,w^{\prime}) from ww to ww^{\prime} so that for any vertex vv on γ(w,w)\gamma(w,w^{\prime}), any path from ww to ww^{\prime} intersects the ball of radius BB around vv. Then XX is quasi-isometric to a tree, with quasi-isometry constants depending on BB only.

Proof.

We check that [Man05, Theorem 4.6] applies; that is, we check the following property. For any two vertices w,wXw,w^{\prime}\in X, there is a midpoint m(w,w)m(w,w^{\prime}) between ww and ww^{\prime} so that any path from ww to ww^{\prime} passes within distance B=B(B)B^{\prime}=B^{\prime}(B) of m(w,w)m(w,w^{\prime}). (The uniformity in the quasi-isometry comes from the proof of Manning’s theorem, see [Man05, page 1170].)

Consider any geodesic α\alpha from ww to ww^{\prime}, and let m=m(w,w)m=m(w,w^{\prime}) be its midpoint. We will show that mm lies within distance 2B+12B+1 of a vertex of γ=γ(w,w)\gamma=\gamma(w,w^{\prime}), so that we can take B=3B+1B^{\prime}=3B+1.

Indeed, suppose by contradiction that this is not the case. Let w=w0,,wn=ww=w_{0},\dots,w_{n}=w^{\prime} be the vertices of γ\gamma (in the order in which they appear along γ\gamma), and let di=d(w,wi)d_{i}=d(w,w_{i}), so that |di+1di|1|d_{i+1}-d_{i}|\leq 1. Each wiw_{i} lies within distance BB of some point pip_{i} on α\alpha which must satisfy d(pi,m)B+1d(p_{i},m)\geq B+1. In particular, we have that every did_{i} satisfies either did(w,w)/21d_{i}\leq d(w,w^{\prime})/2-1 or did(w,w)/2+1d_{i}\geq d(w,w^{\prime})/2+1. Since d0=0d_{0}=0 and dn=d(w,w)d_{n}=d(w,w^{\prime}), we cannot have |di+1di|1|d_{i+1}-d_{i}|\leq 1 for all 0in10\leq i\leq n-1, a contradiction. ∎

We end with a few definitions from coarse geometry which may not be completely standard, but will appear in the next two sections. Given two metrics dd and dd^{\prime} on a set XX, we say that dd is coarsely bounded by dd^{\prime} if there exists a monotone function N:[0,)[0,)N\colon[0,\infty)\to[0,\infty) so that d(x,y)N(d(x,y))d(x,y)\leq N(d^{\prime}(x,y)), for all x,yXx,y\in X. If dd is coarsely bounded by dd^{\prime} and dd^{\prime} is coarsely bounded by dd, we say that dd and dd^{\prime} are coarsely equivalent. An isometric action of a group HH on a metric space YY is metrically proper if for any R>0R>0 and any point yYy\in Y, there are at most finitely many elements hHh\in H for which hB(y,R)B(y,R)h\cdot B(y,R)\cap B(y,R)\neq\emptyset. For proper geodesic spaces, this is equivalent to acting properly discontinuously. If there exists y,Ry,R so that HB(y,R)=YH\cdot B(y,R)=Y, then we say that the action is cobounded, and for proper geodesic metric spaces this is equivalent to acting cocompactly.

3. Projections and vertex spaces

An HHS structure on a metric space consists of certain additional data, most importantly a collection of hyperbolic spaces together with projection maps to each space. For the HHS structure that we will build on (Cayley graphs of) Γ\Gamma, the hyperbolic spaces will (up to quasi-isometry) be the space E^\hat{E} from [DDLS21] (see §2.3) and the spaces 𝒦v\mathcal{K}^{v} and Ξv\Xi^{v} introduced in this section, where vv varies over all vertices of the trees TαT_{\alpha}. Morally, the projections will be given by the maps Λv\Lambda^{v} and ξv\xi^{v} that we study below. However, to prove hierarchical hyperbolicity we will use a criterion from [BHMS20] which does not require actually defining projections, but nevertheless provides them. Still, the maps Λv\Lambda^{v} and ξv\xi^{v} will play a crucial role in proving this criterion applies.

We will establish properties of Λv\Lambda^{v} and ξv\xi^{v} that are reminiscent of subsurface projections or of closest-point projections to peripheral sets in relatively hyperbolic spaces/groups; these are summarized in Proposition 3.19. Essentially, these same properties would be needed if we wanted to construct an HHS structure on Γ\Gamma directly without using [BHMS20].

From a technical point of view, we would like to draw attention to Lemma 3.13, which is the crucial lemma that ensures that the projections behave as desired and that various subspaces have bounded projections. Roughly, the lemma says that closest-point projections to a spine do not vary much under affine deformations.

In what follows, we will write d𝚯vd_{{\bf\Theta}^{v}} and dαd_{\partial{\mathcal{B}}_{\alpha}} for the path metrics on 𝚯v{\bf\Theta}^{v} and α\partial{\mathcal{B}}_{\alpha} induced from d¯\bar{d}. Using the map fα:E¯αf_{\alpha}\colon\bar{E}\to\partial{\mathcal{B}}_{\alpha}, it is straightforward to see that dαd_{\partial{\mathcal{B}}_{\alpha}} is uniformly coarsely equivalent to the subspace metric from d¯\bar{d}: in fact, d¯dαed¯d¯\bar{d}\leq d_{\partial{\mathcal{B}}_{\alpha}}\leq e^{\bar{d}}\bar{d}. The same is true for d𝚯vd_{{\bf\Theta}^{v}}, which follows from the fact that the inclusion of 𝚯v{\bf\Theta}^{v} into α\partial{\mathcal{B}}_{\alpha} is a quasi-isometric embedding with respect to the path metrics (see below).

Associated to each v𝒱v\in\mathcal{V} we will be considering two types of projections. These projections have a single projection Πv:Σ¯𝚯v\Pi^{v}\colon\bar{\Sigma}\to{\bf\Theta}^{v} as a common ingredient. It is convenient to analyze Πv\Pi^{v} via an auxiliary map which serves as a kind of fiberwise closest point projection that survives affine deformations, and which we call the window map. We describe the two types of projections restricted to 𝚯v{\bf\Theta}^{v}, as well as the target spaces of said projections, in §3.1 and §3.3, where we also explain some of their basic features. Next we define the window map and prove what is needed from it. Finally, we define Πv\Pi^{v} and prove the key properties of the associated projections.

3.1. Quasimorphism distances

For each v𝒱v\in\mathcal{V}, we will use ideas from work-in-progress of the fourth author with Hagen, Russell, and Spriano [HRSS21] to define a map

λv:𝚯v𝒦v,\lambda^{v}\colon{\bf\Theta}^{v}\rightarrow\mathcal{K}^{v},

where 𝒦v\mathcal{K}^{v} is a discrete set quasi-isometric to \mathbb{R}. The key properties of this map are given by the next proposition. We note that the proposition and Lemma 3.6 can be used as black-boxes (in particular, the definitions of λv\lambda^{v} and 𝒦v\mathcal{K}^{v} are never used after we prove those results).

Proposition 3.1.

There exists K1>0K_{1}>0 such that, for each vTα(0)𝒱v\in T_{\alpha}^{(0)}\subset\mathcal{V}, there exist a space 𝒦v\mathcal{K}^{v} that is (K1,K1)(K_{1},K_{1})–quasi-isometric to \mathbb{R} and a map λv:𝚯v𝒦v\lambda^{v}\colon{\bf\Theta}^{v}\rightarrow\mathcal{K}^{v} satisfying the following properties:

  1. (1)

    λv\lambda^{v} is K1K_{1}–coarsely Lipschitz with respect to the path metric on 𝚯v{\bf\Theta}^{v}.

  2. (2)

    For any x𝚯vx\in\partial{\bf\Theta}^{v}, if x,α=Dxα\ell_{x,\alpha}=D_{x}\cap\partial\mathcal{B}_{\alpha} then λv(x,α)\lambda^{v}(\ell_{x,\alpha}) is a set of diameter bounded by K1K_{1}.

  3. (3)

    For any v,w𝒱v,w\in\mathcal{V} with dtree(v,w)=1d_{\mathrm{tree}}(v,w)=1, λv×λw:𝚯v𝚯w𝒦v×𝒦w\lambda^{v}\times\lambda^{w}\colon{\bf\Theta}^{v}\cap{\bf\Theta}^{w}\to\mathcal{K}^{v}\times\mathcal{K}^{w} is a K1K_{1}–coarsely surjective (K1,K1)(K_{1},K_{1})–quasi-isometry with respect to the induced path metric on the domain.

  4. (4)

    (Equivariance) For any gΓg\in\Gamma and vTα(0)v\in T_{\alpha}^{(0)} there is an isometry g:𝒦v𝒦gvg\colon\mathcal{K}^{v}\to\mathcal{K}^{gv} and for all xΘvx\in\Theta^{v} we have λgv(gx)=gλv(x)\lambda^{gv}(gx)=g\lambda^{v}(x).

The sets x,α\ell_{x,\alpha} in item (2) are certain lines whose significance is explained below.

Remark 3.2.

An earlier version of this paper used work of Kapovich and Leeb to construct the spaces 𝒦v\mathcal{K}^{v} and maps λv\lambda^{v}, resulting in a weaker version of this proposition which did not include the last, equivariance condition. Consequently Γ\Gamma could only be shown to be an HHS, rather than an HHG. The ideas from [HRSS21] were crucial in this extension.

To explain the proof of the proposition, it is useful to review some background on graph manifolds, which we do now.

Graph manifolds and trees

Recall that a graph manifold is a 33–manifold that contains a canonical finite union of tori (up to isotopy), so that cutting along the tori produces a disjoint union of Seifert fibered 33–manifolds, called the Seifert pieces. Seifert fibered 33–manifolds are compact 33–manifolds foliated by circle leaves; see [JS79].

The universal cover of a graph manifold decomposes into a union of universal covers of the Seifert pieces glued together along 22–planes (covering the tori). The decomposition is dual to a tree, and the universal covers of the Seifert pieces are the vertex spaces. For any Seifert fibered space, its universal cover is foliated by lines, the lift of the foliation by circles, and we refer to the leaves simply as lines in the universal cover.

Horocycles and bundles

Next we describe the specific graph manifolds that are relevant for our purposes.

Let Gα<GG_{\alpha}<G denote the stabilizer of BαB_{\alpha}, for each α𝒫\alpha\in\mathcal{P}. This has a finite index cyclic subgroup Gα0G_{\alpha}^{0} generated by a multitwist, τα=Gα0<Gα\langle\tau_{\alpha}\rangle=G_{\alpha}^{0}<G_{\alpha}; see e.g. [DDLS21, §2.9]. The preimage of GαG_{\alpha} in Γ\Gamma is the π1S\pi_{1}S–extension group Γα\Gamma_{\alpha} of GαG_{\alpha}, and we likewise denote by Γα0<Γα\Gamma_{\alpha}^{0}<\Gamma_{\alpha} the extension group of Gα0G_{\alpha}^{0}. The action of Γα\Gamma_{\alpha} on α\partial\mathcal{B}_{\alpha} is cocompact, and α/Γα\partial\mathcal{B}_{\alpha}/\Gamma_{\alpha} has a finite sheeted (orbifold) covering by α/Γα0\partial\mathcal{B}_{\alpha}/\Gamma_{\alpha}^{0}, which is the graph manifold mentioned in the introduction.

1122112233334444
Figure 2. The surface obtained by gluing sides of the “L-shaped” polygon in pairs by translation according to the numbering has a decomposition into two cylinders (shaded blue and green) in the horizontal direction, α\alpha; τα\tau_{\alpha} is a twist in the bottom cylinder and a square of a twist in top cylinder. The boundaries of these cylinders (drawn in bold) form spines for the complement of core curves (drawn as dotted lines).

Consider the surface SS with the flat metric qXαq_{X_{\alpha}}, so that (S,Xα,qXα)=EXα/π1S(S,X_{\alpha},q_{X_{\alpha}})=E_{X_{\alpha}}/\pi_{1}S. The multitwist τα\tau_{\alpha} is an affine map that preserves the cylinders in direction α\alpha, acting as a power of a Dehn twist in each cylinder and as the identity on their boundaries. The union of the boundaries of the cylinders are spines (deformation retracts) for the subsurfaces that are the complements of the twisting curves (core curves of the cylinders). Consequently, τα\tau_{\alpha} is the identity on these spines. The homeomorphism τα\tau_{\alpha} induces a homeomorphism on the subsurface obtained by cutting open SS along a core curve of each cylinder. Each such induced homeomorphism is the identity on the corresponding spine, and is thus isotopic to the identity relative to the spine; see Figure 2. The mapping torus of each subsurface is a product of the subsurface times a circle, and embeds in the the mapping torus α/Γα0\partial\mathcal{B}_{\alpha}/\Gamma_{\alpha}^{0} of τα\tau_{\alpha}. These sub-mapping tori are the Seifert pieces for the graph manifold structure on α/Γα0\partial\mathcal{B}_{\alpha}/\Gamma_{\alpha}^{0}.

The lifted graph manifold decomposition of α\partial\mathcal{B}_{\alpha} corresponds to TαT_{\alpha}. That is, for each vTα(0)v\in T_{\alpha}^{(0)}, there is a vertex space contained in 𝚯v{\bf\Theta}^{v} and containing θv\theta^{v}. In fact, with respect to the covering group, 𝚯v{\bf\Theta}^{v} is an invariant, bounded neighborhood of the vertex space and θv\theta^{v} is an equivariant deformation retraction of that space. We let Γv<Γα\Gamma^{v}<\Gamma_{\alpha} denote the stabilizer of 𝚯v{\bf\Theta}^{v} in Γα\Gamma_{\alpha} and Γv0<Γα0\Gamma^{v0}<\Gamma_{\alpha}^{0} the stabilizer in Γα0\Gamma_{\alpha}^{0}. The suspension flow on the mapping torus α/Γα0\partial\mathcal{B}_{\alpha}/\Gamma_{\alpha}^{0} restricted to each quotient of the spine bundle, θv/Γv0\theta^{v}/\Gamma^{v0}, defines circle leaves of the corresponding Seifert piece; that is, flow lines through any point on the θv/Γv0\theta^{v}/\Gamma^{v0} are precisely the circle leaves. In the universal covering α\partial\mathcal{B}_{\alpha}, the lifted flowline through a point xαx\in\partial\mathcal{B}_{\alpha} is a lifted horocycle, x,α=Dxα\ell_{x,\alpha}=D_{x}\cap\partial\mathcal{B}_{\alpha}. Thus, for any vertex vv and any xθvx\in\theta^{v}, x,α\ell_{x,\alpha} is a line for the vertex space corresponding to vv. We note that not only does Γv0\Gamma^{v0} preserve this set of lines, but so does Γv\Gamma^{v}.

For any xθvx\in\theta^{v}, the stabilizer in Γv0\Gamma^{v0} of x,α\ell_{x,\alpha} is generated by a lift gvg_{v} of τα\tau_{\alpha}. Therefore, the quotient 𝚯v/Γv0{\bf\Theta}^{v}/\Gamma^{v0} is homeomorphic to a product, 𝚯Xv/π1Sv×S1{\bf\Theta}^{v}_{X}/\pi_{1}S^{v}\times S^{1}, where π1Sv\pi_{1}S^{v} is the stabilizer of vv in π1S<Γ\pi_{1}S<\Gamma and XBαX\subset\partial B_{\alpha} is any point. Indeed, there is a deformation retraction to θv/Γv0=θXv/π1Sv×S1\theta^{v}/\Gamma^{v0}=\theta^{v}_{X}/\pi_{1}S^{v}\times S^{1}. If we do not care about the particular point XX over which we take the fiber, we simply write SvS^{v} for the surface 𝚯Xv/π1Sv{\bf\Theta}^{v}_{X}/\pi_{1}S^{v}, so that 𝚯v/Γv0Sv×S1{\bf\Theta}^{v}/\Gamma^{v0}\cong S^{v}\times S^{1}. Since EXE_{X} is a copy of the universal cover of SS, we can consider SvS^{v} as a subsurface of SS (embedded on the interior) and π1Sv\pi_{1}S^{v} is its fundamental group inside π1S\pi_{1}S (up to conjugacy).

The product structure Sv×S1𝚯v/Γv0S^{v}\times S^{1}\cong{\bf\Theta}^{v}/\Gamma^{v0} can be chosen so that 𝚯v/Γv0𝚯v/Γ{\bf\Theta}^{v}/\Gamma^{v0}\to{\bf\Theta}^{v}/\Gamma is an orbifold cover sending circles to circles making 𝚯v/Γv{\bf\Theta}^{v}/\Gamma^{v} into a Seifert fibered orbifold (some of the Seifert fibers may be part of the orbifold locus) that also (orbifold)-fibers over the circle (with finite order monodromy). Write 𝚯v/Γv𝒪v{\bf\Theta}^{v}/\Gamma^{v}\to\mathcal{O}^{v} for the Seifert fibration to the quotient 22–orbifold. Further write

νv:Γvπ1orb(𝒪v)\nu^{v}\colon\Gamma^{v}\to\pi_{1}^{orb}(\mathcal{O}^{v})

for the induced homomorphism of the Seifert fibration and

ϕv:Γv\phi^{v}\colon\Gamma^{v}\to\mathbb{Z}

for the induced homomorphism from the fibration over the circle. Because gvg_{v} acts as translation on the line x,α\ell_{x,\alpha} for xθvx\in\theta^{v}, it represents a loop that traverses a circle in the Seifert fibration, which is thus also a suspension flowline for the fibration over the circle. Thus we have νv(gv)=0\nu^{v}(g_{v})=0 and ϕv(gv)0\phi^{v}(g_{v})\neq 0. To complete the picture, we note that restricting Sv×S1𝚯v/Γv0𝚯v/Γv𝒪vS^{v}\times S^{1}\cong{\bf\Theta}^{v}/\Gamma^{v0}\to{\bf\Theta}^{v}/\Gamma^{v}\to\mathcal{O}^{v} to SvS^{v} defines an orbifold covering Sv𝒪vS^{v}\to\mathcal{O}^{v}.

Finally, note that for any ww adjacent to vv in TαT_{\alpha}, gv×gw2\langle g_{v}\rangle\times\langle g_{w}\rangle\cong\mathbb{Z}^{2} has finite index in ΓvΓw\Gamma^{v}\cap\Gamma^{w}. Viewing gw<Γv\langle g_{w}\rangle<\Gamma^{v}, we note that νv|gw\nu^{v}|_{\langle g_{w}\rangle} is an isomorphism onto an infinite cyclic subgroup of π1orb𝒪v\pi_{1}^{orb}\mathcal{O}^{v}. In fact, the image νv(gw)\nu^{v}(\langle g_{w}\rangle) is (a conjugate of a power of) the fundamental group of a boundary component of 𝒪v\mathcal{O}^{v}.

Remark 3.3.

One caveat about the lines for the vertex spaces: flowlines through points not on a spine are not lines of any vertex space. In fact, they are not even uniformly close to lines for any vertex space.

Constructing the map

Here we define 𝒦v\mathcal{K}^{v} and λv\lambda^{v} and prove the main properties we will need about them. We require a little more set up first. We choose representatives of the Γ\Gamma–orbits of vertices, 𝒱0={v1,,vk}𝒱\mathcal{V}_{0}=\{v_{1},\ldots,v_{k}\}\subset\mathcal{V}. For each v𝒱0v\in\mathcal{V}_{0}, choose a fundamental domain Δv\Delta^{v} for the action of Γv\Gamma^{v} on 𝚯v{\bf\Theta}^{v}. We assume that Δv\Delta^{v} has compact, connected closure, that gΔvΔv=g\Delta^{v}\cap\Delta^{v}=\emptyset for all gΓv{1}g\in\Gamma^{v}\setminus\{1\}, and that gΓvgΔv=𝚯v\bigcup_{g\in\Gamma^{v}}g\cdot\Delta^{v}={\bf\Theta}^{v}. The set

(1) {gΓvgΔ¯vΔ¯v}\{g\in\Gamma^{v}\mid g\overline{\Delta}^{v}\cap\overline{\Delta}^{v}\neq\emptyset\}

is a finite generating set for Γv\Gamma^{v}. The Γv\Gamma^{v}–translates of Δv\Delta^{v} define a tiling of 𝚯v{\bf\Theta}^{v}, and the map sending every point of gΔvg\Delta^{v} to gΓvg\in\Gamma^{v} is a quasi-isometry by the Milnor-Schwarz Lemma. We denote this map as λ~v:𝚯vΓv\widetilde{\lambda}^{v}\colon{\bf\Theta}^{v}\to\Gamma^{v}.

We note that any word metric on Γv\Gamma^{v} defines a “word metric” on each coset gΓvg\Gamma^{v}, for gΓg\in\Gamma (elements are distance 11 if they differ by right multiplication by an element of the generating set). We can push the tiling forward by gg to a Γgv=gΓvg1\Gamma^{gv}=g\Gamma^{v}g^{-1}–invariant tiling of 𝚯gv{\bf\Theta}^{gv} (if gΓvg\in\Gamma^{v}, this is precisely the given tiling of 𝚯v{\bf\Theta}^{v}). For any element ggΓvg^{\prime}\in g\Gamma^{v}, the map that sends every point in gΔvg^{\prime}\Delta^{v} to gg^{\prime} defines a quasi-isometry λ~gv:𝚯gvgΓv\widetilde{\lambda}^{gv}\colon{\bf\Theta}^{gv}\to g\Gamma^{v} which is Γgv\Gamma^{gv}–equivariant, with the same quasi-isometry constants. If gΓg^{\prime}\in\Gamma and xΔvx^{\prime}\in\Delta^{v}, then for all gΓg\in\Gamma

λ~ggv(ggx)=gg=gλ~gv(gx).\widetilde{\lambda}^{gg^{\prime}v}(gg^{\prime}x^{\prime})=gg^{\prime}=g\widetilde{\lambda}^{g^{\prime}v}(g^{\prime}x^{\prime}).

On the other hand, any wΓvw\in\Gamma\cdot v and x𝚯wx\in{\bf\Theta}^{w} have the form w=gvw=g^{\prime}v and x=gxx=g^{\prime}x^{\prime} for some gΓg^{\prime}\in\Gamma and xΔvx^{\prime}\in\Delta^{v}. Thus, for any gΓg\in\Gamma, the equation above becomes

(2) λ~gw(gx)=g~λw(x)\widetilde{\lambda}^{gw}(gx)=\widetilde{g}\lambda^{w}(x)

Having carried out the construction above for each v𝒱0v\in\mathcal{V}_{0} and each vertex in its orbit, we have maps λ~w\widetilde{\lambda}^{w} from 𝚯w{\bf\Theta}^{w} to a coset of a vertex stabilizer from 𝒱0\mathcal{V}_{0} for every w𝒱w\in\mathcal{V}, so that equation (2) holds for every x𝚯wx\in{\bf\Theta}^{w}, and gΓg\in\Gamma.

Next, recall that a homogeneous quasimorphism (with deficiency DD) from a group HH to \mathbb{R} is a map

ψ:H\psi\colon H\to\mathbb{R}

such that for all h,h1,h2Hh,h_{1},h_{2}\in H and nn\in\mathbb{Z} we have ψ(hn)=nψ(h)\psi(h^{n})=n\psi(h) and

|ψ(h1h2)ψ(h1)ψ(h2)|D.|\psi(h_{1}h_{2})-\psi(h_{1})-\psi(h_{2})|\leq D.
Lemma 3.4.

For any v𝒱v\in\mathcal{V}, there is a homogeneous quasimorphism ψv:Γv\psi^{v}\colon\Gamma^{v}\to\mathbb{R} such that ψv(gv)\psi^{v}(\langle g_{v}\rangle) is unbounded, and ψv(gw)=0\psi^{v}(g_{w})=0 for any adjacent vertex w𝒱w\in\mathcal{V}.

Proof.

Let w1,,wrw_{1},\ldots,w_{r} be Γv\Gamma^{v}–orbit representatives of the vertices adjacent to vv. Here rr is the number of boundary components of 𝒪v\mathcal{O}^{v}, so that νv(gw1),,νv(gwr)\nu^{v}(g_{w_{1}}),\ldots,\nu^{v}(g_{w_{r}}) are peripheral loops around the rr distinct boundary components of 𝒪v\mathcal{O}^{v}. Since π1orb𝒪v\pi_{1}^{orb}\mathcal{O}^{v} is the fundamental group of a hyperbolic 22–orbifold with non-empty boundary, appealing to [HO13, Theorem 4.2], which applies to π1orb𝒪v\pi^{orb}_{1}\mathcal{O}^{v} and its subgroups νv(gwi)\langle\nu^{v}(g_{w_{i}})\rangle in view of [DGO17, Corollary 6.6, Theorem 6.8], one can find a homogeneous quasimorphism ηi:π1orb𝒪v\eta_{i}\colon\pi^{orb}_{1}\mathcal{O}^{v}\to\mathbb{R}, for i=1,,ri=1,\ldots,r, such that ηi(νv(gwi))=1\eta_{i}(\nu^{v}(g_{w_{i}}))=1 and ηi(νv(gwj))=0\eta_{i}(\nu^{v}(g_{w_{j}}))=0 for jij\neq i. (The construction of Epstein–Fujiwara [EF97] should also be applicable to construct such quasimorphisms). Set s0=1/ϕv(gv)s_{0}=1/\phi^{v}(g_{v}), and for each i=1,,ri=1,\ldots,r, set si=s0ϕv(gwi)s_{i}=s_{0}\phi^{v}(g_{w_{i}}), and then define

ψv=s0ϕvi=1rsiηiνv.\psi^{v}=s_{0}\phi^{v}-\sum_{i=1}^{r}s_{i}\eta_{i}\circ\nu^{v}.

As a linear combination of homogeneous quasimorphisms, ψv\psi^{v} is a homogeneous quasimorphism. Since gvker(νv)g_{v}\in\ker(\nu^{v}), it follows that ηiνv(gv)=0\eta_{i}\circ\nu^{v}(g_{v})=0 for all ii, hence ψv(gv)=s0ϕv(gv)=ϕv(gv)/ϕv(gv)=1\psi^{v}(g_{v})=s_{0}\phi^{v}(g_{v})=\phi^{v}(g_{v})/\phi^{v}(g_{v})=1. On the other hand, for any j=1,,rj=1,\ldots,r we have

ψv(gwj)=s0ϕv(gwj)i=1rsiηi(νv(gj))=s0ϕv(gwj)i=1rs0ϕv(gwi)δij=0,\psi^{v}(g_{w_{j}})=s_{0}\phi^{v}(g_{w_{j}})-\sum_{i=1}^{r}s_{i}\eta_{i}(\nu^{v}(g_{j}))=s_{0}\phi^{v}(g_{w_{j}})-\sum_{i=1}^{r}s_{0}\phi^{v}(g_{w_{i}})\delta_{ij}=0,

proving the lemma. ∎

According to [ABO19, Lemma 4.15], there is an (infinite) generating set for Γv\Gamma^{v} so that with respect to the resulting word metric, the quasimorphism ψv:Γv\psi^{v}\colon\Gamma^{v}\to\mathbb{R} from Lemma 3.4 is a quasi-isometry. For v𝒱0v\in\mathcal{V}_{0}, define 𝒦v=Γv\mathcal{K}^{v}=\Gamma^{v} with this choice of word metric and let

λv:𝚯v𝒦v\lambda^{v}\colon{\bf\Theta}^{v}\to\mathcal{K}^{v}

simply be the map λ~v\widetilde{\lambda}^{v} (followed by the identification of Γv\Gamma^{v} with 𝒦v\mathcal{K}^{v}). For any gΓg\in\Gamma, define 𝒦gv\mathcal{K}^{gv} to be the coset gΓvg\Gamma^{v} with this generating set so that λ~gv\widetilde{\lambda}^{gv} defines a map

λgv:𝚯gv𝒦gv.\lambda^{gv}\colon{\bf\Theta}^{gv}\to\mathcal{K}^{gv}.

Carrying this out for every v𝒱0v\in\mathcal{V}_{0}, (2) implies

(3) λgw(gx)=gλw(x)\lambda^{gw}(gx)=g\lambda^{w}(x)

for all w𝒱w\in\mathcal{V} and x𝚯wx\in{\bf\Theta}^{w}, and gΓg\in\Gamma.

Before we proceed to the proof of Proposition 3.1, observe that Γgv=gΓvg1\Gamma^{gv}=g\Gamma^{v}g^{-1} acts isometrically on gΓvg\Gamma^{v} with respect to any generating set, and thus we can use this to define a generating set for the conjugate so that (any) orbit map is an isometry; in fact, this will just be a conjugate of the generating set for Γv\Gamma^{v}. In particular, when convenient we will identify 𝒦gv\mathcal{K}^{gv} isometrically with the conjugate gΓvg1g\Gamma^{v}g^{-1} via such an orbit map. Conjugating the quasimorphisms ψv\psi^{v} from the lemma, for v𝒱0v\in\mathcal{V}_{0}, we obtain uniform quasi-isometries

(4) ψw:𝒦w\psi^{w}\colon\mathcal{K}^{w}\to\mathbb{R}

for all w𝒱w\in\mathcal{V}, which for an appropriate choice of identification of 𝒦w\mathcal{K}^{w} with a conjugate of some Γv\Gamma^{v}, v𝒱0v\in\mathcal{V}_{0}, is a quasimorphism (with uniformly bounded deficiency).

Proof of Proposition 3.1.

From the discussion above and Equation (3), we immediately see that item (4) of the proposition holds.

Next, observe that by adding finitely many generators to the infinite generating set of Γv0\Gamma^{v_{0}} for any v0𝒱0v_{0}\in\mathcal{V}_{0}, changes 𝒦v0\mathcal{K}^{v_{0}} by quasi-isometry. On the other hand, the finite generating set described in Equation (1) for v0𝒱0v_{0}\in\mathcal{V}_{0} makes λ~v0\widetilde{\lambda}^{v_{0}} a quasi-isometry. Thus, adding these generators to the infinite generating set does not change the quasi-isometry type of 𝒦v0\mathcal{K}^{v_{0}}, but clearly makes λv0\lambda^{v_{0}} coarsely Lipschitz. Therefore, λv\lambda^{v} is uniformly coarsely Lipschitz for all v𝒱v\in\mathcal{V}, and hence item (1) holds for all v𝒱v\in\mathcal{V}.

To prove item (2), let v𝒱v\in\mathcal{V} and x𝚯vx\in\partial{\bf\Theta}^{v}. Then xθwx\in\theta^{w}, for some w𝒱w\in\mathcal{V} adjacent to vv. As discussed above, we view 𝒦v\mathcal{K}^{v} and 𝒦w\mathcal{K}^{w} as conjugates Γv\Gamma^{v} and Γw\Gamma^{w} of groups Γv0\Gamma^{v_{0}} and Γw0\Gamma^{w_{0}}, respectively, for v0,w0𝒱0v_{0},w_{0}\in\mathcal{V}_{0}, equipped with their conjugated infinite generating sets. Let ψv:𝒦v\psi^{v}\colon\mathcal{K}^{v}\to\mathbb{R} and ψw:𝒦w\psi^{w}\colon\mathcal{K}^{w}\to\mathbb{R} be the associated uniform quasi-isometric homogeneous quasimorphisms. The element gwΓvg_{w}\in\Gamma^{v} stabilizes x,α\ell_{x,\alpha} acting by translation on it, and by construction, ψv(gw)=ψv(gwn)=0\psi^{v}(g_{w})=\psi^{v}(g_{w}^{n})=0 for all nn\in\mathbb{Z}. It follows that every orbit of gw\langle g_{w}\rangle acting on 𝒦v\mathcal{K}^{v} is uniformly bounded. Indeed, if DD is the deficiency of ψv\psi^{v}, then for any g𝒦vg\in\mathcal{K}^{v}, we have

|ψv(gwng)ψv(g)|=|ψv(gwng)ψv(g)ψv(gwn)|D|\psi^{v}(g_{w}^{n}g)-\psi^{v}(g)|=|\psi^{v}(g_{w}^{n}g)-\psi^{v}(g)-\psi^{v}(g_{w}^{n})|\leq D

and therefore gwngg_{w}^{n}g and gg are uniformly bounded distance apart in 𝒦v\mathcal{K}^{v} (since ψv\psi^{v} is a uniform quasiisometry).

Now, since gwnv=vg_{w}^{n}v=v, by item (4) of the proposition we have

λv(gwnx)=λgwnv(gwnx)=gwnλv(x),\lambda^{v}(g_{w}^{n}x)=\lambda^{g_{w}^{n}v}(g_{w}^{n}x)=g_{w}^{n}\lambda^{v}(x),

and since gwnλv(x)g_{w}^{n}\lambda^{v}(x) is uniformly close to λv(x)\lambda^{v}(x), it follows that λv\lambda^{v} sends the gw\langle g_{w}\rangle–orbit of xx to a uniformly bounded set. Since this orbit is RR–dense in x,α\ell_{x,\alpha} for some uniform R>0R>0, and since λv\lambda^{v} is uniformly coarsely Lipschitz (by item (1)) we see that λv(x,α)\lambda^{v}(\ell_{x,\alpha}) has uniformly bounded diameter. This proves item (2).

For item (3), we continue with the assumptions on v,wv,w as above. Note that since ψv(gvn)=n\psi^{v}(g_{v}^{n})=n, using again the fact that ψv\psi^{v} is a uniform quasi-isometric homogeneous quasimorphism to \mathbb{R}, it follows that for any x𝚯vx\in{\bf\Theta}^{v}, the map nλv(gvnx)n\mapsto\lambda^{v}(g_{v}^{n}x) is a uniformly coarsely surjective, uniform quasiisometry 𝒦v\mathbb{Z}\to\mathcal{K}^{v}. Since every orbit of gw\langle g_{w}\rangle on 𝒦v\mathcal{K}^{v} is uniformly bounded, it follows that for all n,mn,m\in\mathbb{Z}, the two points λv(gvngwmx)=gwmλv(gvnx)\lambda^{v}(g_{v}^{n}g_{w}^{m}x)=g_{w}^{m}\lambda^{v}(g_{v}^{n}x) and λv(gvnx)\lambda^{v}(g_{v}^{n}x) are uniformly close to each other. Likewise, λw(gvngwmx)\lambda^{w}(g_{v}^{n}g_{w}^{m}x) and λw(gwmx)\lambda^{w}(g_{w}^{m}x) are also uniformly close to each other. But this means that

λv×λw(gvngwmx) and (λv(gvnx),λw(gwmx))\lambda^{v}\times\lambda^{w}(g_{v}^{n}g_{w}^{m}x)\mbox{ and }(\lambda^{v}(g_{v}^{n}x),\lambda^{w}(g_{w}^{m}x))

are uniformly close, and thus

(n,m)λv×λw(gvngwmx)(n,m)\mapsto\lambda^{v}\times\lambda^{w}(g_{v}^{n}g_{w}^{m}x)

is a uniformly coarsely surjective, uniform quasiisometry 2𝒦v×𝒦w\mathbb{Z}^{2}\to\mathcal{K}^{v}\times\mathcal{K}^{w}.

On the other hand, the assignment (n,m)gvngwmx(n,m)\mapsto g_{v}^{n}g_{w}^{m}x defines a uniform quasiisometry 2𝚯v𝚯w\mathbb{Z}^{2}\to{\bf\Theta}^{v}\cap{\bf\Theta}^{w} since gw×gv2\langle g_{w}\rangle\times\langle g_{v}\rangle\cong\mathbb{Z}^{2} acts cocompactly on 𝚯v𝚯w{\bf\Theta}^{v}\cap{\bf\Theta}^{w} (with uniformity coming from the fact that there are only finitely many Γ\Gamma–orbits of pairs (v,w)(v,w) of adjacent vertices). Combining these two facts, together with the fact that λv\lambda^{v} and λw\lambda^{w} are uniformly coarsely Lipschitz, it follows that

λv×λw:𝚯v𝚯w𝒦v×𝒦w\lambda^{v}\times\lambda^{w}\colon{\bf\Theta}^{v}\cap{\bf\Theta}^{w}\to\mathcal{K}^{v}\times\mathcal{K}^{w}

is a uniformly coarsely surjective, uniform quasiisometry. This proves item (3), and completes the proof of the proposition. ∎

3.2. A technical lemma

The goal of this subsection is to prove Lemma 3.6, whose relevance will only be clear in §4. We prove it here since we have now established the setup for its proof.

We recall that for each vv, since 𝚯v/Γv{\bf\Theta}^{v}/\Gamma^{v} is a Seifert fibered orbifold, we have have a Γv\Gamma^{v}–equivariant, uniformly biLipschitz homeomorphism μv×ρv:𝚯vS~v×\mu^{v}\times\rho^{v}\colon{\bf\Theta}^{v}\to\widetilde{S}^{v}\times\mathbb{R}, where S~v\widetilde{S}^{v} is the (simply connected) surface-with-boundary 𝚯XvEX{\bf\Theta}^{v}_{X}\subset E_{X} for some XBαX\in\partial B_{\alpha} (and α=α(v)\alpha=\alpha(v)) and the slices {x}×\{x\}\times\mathbb{R} (more precisely, the level sets of (μv)1(x)𝚯v(\mu^{v})^{-1}(x)\subset{\bf\Theta}^{v}) are lines for 𝚯v{\bf\Theta}^{v}. These lines project to circle fibers in 𝚯v/Γv{\bf\Theta}^{v}/\Gamma^{v} and we may assume they contain all the lines x,v\ell_{x,v} for all xθvx\in\theta^{v}.

Lemma 3.5.

The map μv×λv:𝚯vS~v×𝒦v\mu^{v}\times\lambda^{v}\colon{\bf\Theta}^{v}\to\widetilde{S}^{v}\times\mathcal{K}^{v} is a uniform, Γv\Gamma^{v}–equivariant quasi-isometry with uniformly dense image. Moreover, the constant K1K_{1} from Proposition 3.1 can be chosen so that for any v𝒱v\in\mathcal{V} and s𝒦vs\in\mathcal{K}^{v}, the subspace

M(s)=(λv)1(NK1(s))𝚯v,M(s)=(\lambda^{v})^{-1}(N_{K_{1}}(s))\subset{\bf\Theta}^{v},

has the property that μv×λv(M(s))\mu^{v}\times\lambda^{v}(M(s)) has uniformly bounded Hausdorff distance to the slice S~v×{s}\widetilde{S}^{v}\times\{s\}, and furthermore M(s)M(s) nontrivially intersects every line of 𝚯v{\bf\Theta}^{v}.

We note that the intersection of M(s)M(s) with each line of 𝚯v{\bf\Theta}^{v} is necessarily a uniformly bounded diameter set by the uniform bounded Hausdorff distance condition.

Proof.

All constants will be independent of the specific vertex vv, so we drop it from the notation. We write dd for all path-metric distances in what follows (the location of points will determine which metric is being used). Products are given the L1L^{1} metric for convenience. We further let KK be the maximum of the coarse Lipschitz constants of μ,ρ,λ\mu,\rho,\lambda and the biLipschitz constant of μ×ρ\mu\times\rho, and assume, as we may, that K2K\geq 2. From the proof of Proposition 3.1(3), if xx is any point of a line of 𝚯{\bf\Theta}, then nλv(gvnx)n\mapsto\lambda^{v}(g_{v}^{n}x) is a uniformly coarsely surjective, uniform quasi-isometry from \mathbb{Z} to 𝒦\mathcal{K}. Therefore, λ=λv\lambda=\lambda^{v} is a uniformly coarsely surjective, uniform quasi-isometry from any line of 𝚯{\bf\Theta} to 𝒦\mathcal{K}. We further assume that the coarse surjectivity constants and quasi-isometry constants are also all taken to be KK.

Let x,y𝚯x,y\in{\bf\Theta} be any two points. Since μ\mu and λ\lambda are KK–coarsely Lipschitz, μ×λ\mu\times\lambda is (2K,2K)(2K,2K)–coarsely Lipschitz. To prove the required uniform lower bound on μ×λ\mu\times\lambda–distances, we note that since μ×ρ\mu\times\rho is a KK–biLipschitz homeomorphism, it suffices to uniformly coarsely bound d(μ×ρ(x),μ×ρ(y))d(\mu\times\rho(x),\mu\times\rho(y)) from above by d(μ×λ(x),μ×λ(y))d(\mu\times\lambda(x),\mu\times\lambda(y)). For reasons that will become clear shortly, we observe that

(5) d(μ×ρ(x),μ×ρ(y))\displaystyle d(\mu\times\rho(x),\mu\times\rho(y)) =\displaystyle= d(μ(x),μ(y))+d(ρ(x),ρ(y))\displaystyle d(\mu(x),\mu(y))+d(\rho(x),\rho(y))
\displaystyle\leq 2max{K4d(μ(x),μ(y)),d(ρ(x),ρ(y))}.\displaystyle\displaystyle{2\max\left\{K^{4}d(\mu(x),\mu(y)),d(\rho(x),\rho(y))\right\}}.

If the maximum is realized by K4d(μ(x),μ(y))K^{4}d(\mu(x),\mu(y)), then note that

d(μ×ρ(x),μ×ρ(y))\displaystyle d(\mu\times\rho(x),\mu\times\rho(y)) \displaystyle\leq 2K4d(μ(x),μ(y))\displaystyle 2K^{4}d(\mu(x),\mu(y))
\displaystyle\leq 2K4d(μ(x),μ(y))+2K4d(λ(x),λ(y))\displaystyle 2K^{4}d(\mu(x),\mu(y))+2K^{4}d(\lambda(x),\lambda(y))
=\displaystyle= 2K4d(μ×λ(x),μ×λ(y)),\displaystyle 2K^{4}d(\mu\times\lambda(x),\mu\times\lambda(y)),

as required.

We are left to consider the case that the maximum in (5) is realized by d(ρ(x),ρ(y))d(\rho(x),\rho(y)), which thus satisfies

d(ρ(x),ρ(y))K4d(μ(x),μ(y)).d(\rho(x),\rho(y))\geq K^{4}d(\mu(x),\mu(y)).

Let z𝚯z\in{\bf\Theta} be such that ρ(z)=ρ(x)\rho(z)=\rho(x) and μ(z)=μ(y)\mu(z)=\mu(y). Since μ(z)=μ(y)\mu(z)=\mu(y), zz and yy lie on a line, and since the restriction of λ\lambda to this line is a (K,K)(K,K)–quasi-isometry, we have

d(λ(z),λ(y))1Kd(ρ(z),ρ(y))K=1Kd(ρ(x),ρ(y))K.d(\lambda(z),\lambda(y))\geq\tfrac{1}{K}d(\rho(z),\rho(y))-K=\tfrac{1}{K}d(\rho(x),\rho(y))-K.

Since λ\lambda is KK–coarsely Lipschitz and μ×ρ\mu\times\rho is KK–biLipschitz, we have

d(λ(x),λ(z))\displaystyle d(\lambda(x),\lambda(z)) \displaystyle\leq Kd(x,z)+K\displaystyle Kd(x,z)+K
\displaystyle\leq K2(d(μ(x),μ(z))+d(ρ(x),ρ(z)))+K\displaystyle K^{2}(d(\mu(x),\mu(z))+d(\rho(x),\rho(z)))+K
=\displaystyle= K2d(μ(x),μ(y))+K\displaystyle K^{2}d(\mu(x),\mu(y))+K
\displaystyle\leq K2(1K4d(ρ(x),ρ(y)))+K\displaystyle K^{2}\left(\tfrac{1}{K^{4}}d(\rho(x),\rho(y))\right)+K
=\displaystyle= 1K2d(ρ(x),ρ(y))+K\displaystyle\tfrac{1}{K^{2}}d(\rho(x),\rho(y))+K

Combining the previous two sets of inequalities and the triangle inequality, we have

d(λ(x),λ(y))\displaystyle d(\lambda(x),\lambda(y)) \displaystyle\geq d(λ(z),λ(y))d(λ(z),λ(x))\displaystyle d(\lambda(z),\lambda(y))-d(\lambda(z),\lambda(x))
\displaystyle\geq 1Kd(ρ(x),ρ(y))K(1K2d(ρ(x),ρ(y))+K)\displaystyle\tfrac{1}{K}d(\rho(x),\rho(y))-K-\left(\tfrac{1}{K^{2}}d(\rho(x),\rho(y))+K\right)
\displaystyle\geq K1K2d(ρ(x),ρ(y))2K.\displaystyle\tfrac{K-1}{K^{2}}d(\rho(x),\rho(y))-2K.

Combining this inequality with (5) where we have assumed the maximum is realized by d(ρ(x),ρ(y))d(\rho(x),\rho(y)), we obtain

d(μ×ρ(x),μ×ρ(y))\displaystyle d(\mu\times\rho(x),\mu\times\rho(y)) \displaystyle\leq 2d(ρ(x),ρ(y))\displaystyle 2d(\rho(x),\rho(y))
\displaystyle\leq 2K2K1d(λ(x),λ(y))+4K3K1\displaystyle\tfrac{2K^{2}}{K-1}d(\lambda(x),\lambda(y))+\tfrac{4K^{3}}{K-1}
\displaystyle\leq 2K2K1d(μ×λ(x),μ×λ(y))+4K3K1.\displaystyle\tfrac{2K^{2}}{K-1}d(\mu\times\lambda(x),\mu\times\lambda(y))+\tfrac{4K^{3}}{K-1}.

which provides the required upper bound. This completes the proof of the first claim of the lemma.

For the second claim of the lemma, we now increase K1K_{1} from Proposition 3.1 if necessary, so that K1KK_{1}\geq K. Observe that

(6) (μ×λ)(μ×ρ)1(x,t)=(x,λ((μ×ρ)1(x,t))).(\mu\times\lambda)\circ(\mu\times\rho)^{-1}(x,t)=(x,\lambda((\mu\times\rho)^{-1}(x,t))).

That is, (μ×λ)(μ×ρ)1(\mu\times\lambda)\circ(\mu\times\rho)^{-1} sends the line {x}×\{x\}\times\mathbb{R} to {x}×𝒦\{x\}\times\mathcal{K}, for any xS~x\in\widetilde{S}. As already noted at the start of the proof, restricting to this line, λ\lambda is KK–coarsely Lipschitz and KK–coarsely onto. Therefore, for any s𝒦s\in\mathcal{K} and xS~x\in\widetilde{S}, there exists tt so that λ((μ×ρ)1(x,t))\lambda((\mu\times\rho)^{-1}(x,t)) is within K1KK_{1}\geq K of ss. Thus, for any line of 𝚯{\bf\Theta}, the λ\lambda–image nontrivially intersects NK1(s)N_{K_{1}}(s), and hence this line nontrivially intersects M(s)M(s). By definition, μ×λ\mu\times\lambda maps M(s)M(s) into S~v×NK1(s)\widetilde{S}^{v}\times N_{K_{1}}(s), and by the previous sentence, every point of S~×{s}\widetilde{S}\times\{s\} is within K1K_{1} of some point of μ×λ(M(s))\mu\times\lambda(M(s)). Thus, λ×μ(M(s))\lambda\times\mu(M(s)) has Hausdorff distance at most K1K_{1} from S~×{s}\widetilde{S}\times\{s\}, as required. ∎

As mentioned above, the following technical lemma will be needed in §4. In the statement M(s)M(s) is as defined by Lemma 3.5.

Lemma 3.6.

There is a function N=N𝒦:[0,)[0,)N=N_{\mathcal{K}}\colon[0,\infty)\to[0,\infty) with the following property. Suppose that v,v1,v2𝒱v,v_{1},v_{2}\in\mathcal{V} are so that dtree(v,vi)=1d_{\mathrm{tree}}(v,v_{i})=1. Then for each s𝒦vs\in\mathcal{K}^{v} and ti𝒦vit_{i}\in\mathcal{K}^{v_{i}} we have

d¯(M(s)M(t1),M(s)M(t2))N(d¯(M(t1),M(t2))).\bar{d}(M(s)\cap M(t_{1}),M(s)\cap M(t_{2}))\leq N(\bar{d}(M(t_{1}),M(t_{2}))).
Proof.

It suffices to prove the lemma with d¯\bar{d} replaced by the path metric dαd_{\partial\mathcal{B}_{\alpha}} on α\partial\mathcal{B}_{\alpha}, since they are uniformly coarsely equivalent. In fact, it will be convenient to consider the path metric d0d_{0} on the union of the three vertex subspaces

Ω=𝚯v𝚯v1𝚯v2,\Omega={\bf\Theta}_{v}\cup{\bf\Theta}_{v_{1}}\cup{\bf\Theta}_{v_{2}},

which is also uniformly coarsely equivalent since each vertex space uniformly quasi-isometrically embeds in α\partial\mathcal{B}_{\alpha}. In this subspace, we will actually prove that the two distances are uniformly comparable.

Now, for each i=1,2i=1,2 the uniform quasi-isometry μvi×λvi:𝚯viS~vi×𝒦vi\mu^{v_{i}}\times\lambda^{v_{i}}\colon{\bf\Theta}^{v_{i}}\to\widetilde{S}^{v_{i}}\times\mathcal{K}^{v_{i}} from Lemma 3.5 maps the space 𝚯v𝚯vi{\bf\Theta}^{v}\cap{\bf\Theta}^{v_{i}} within bounded Hausdorff distance of a subspace vS~vi×𝒦vi\partial^{v}\widetilde{S}^{v_{i}}\times\mathcal{K}^{v_{i}}, for a boundary component vS~vi\partial^{v}\widetilde{S}^{v_{i}} of S~vi\widetilde{S}^{v_{i}}. Let η:S~vivS~vi\eta\colon\widetilde{S}^{v_{i}}\to\partial^{v}\widetilde{S}^{v_{i}} be the closest point projection, and then set

(μvi×λvi)1(η×id)(μvi×λvi):𝚯vi𝚯vi(\mu^{v_{i}}\times\lambda^{v_{i}})^{-1}\circ(\eta\times id)\circ(\mu^{v_{i}}\times\lambda^{v_{i}})\colon{\bf\Theta}^{v_{i}}\to{\bf\Theta}^{v_{i}}

where (μvi×λvi)1(\mu^{v_{i}}\times\lambda^{v_{i}})^{-1} is a coarse inverse of μvi×λvi\mu^{v_{i}}\times\lambda^{v_{i}} with μvi(μvi×λvi)1(x,s)=x\mu^{v_{i}}\circ(\mu^{v_{i}}\times\lambda^{v_{i}})^{-1}(x,s)=x (c.f. Equation (6)). This map is a uniformly coarsely Lipschitz, coarse retraction of 𝚯vi{\bf\Theta}^{v_{i}} onto 𝚯vi𝚯v{\bf\Theta}^{v_{i}}\cap{\bf\Theta}^{v}. Moreover, this sends M(ti)M(t_{i}), which is uniformly close to the (μvi×λvi)1(\mu^{v_{i}}\times\lambda^{v_{i}})^{-1}–image of S~vi×{ti}\widetilde{S}^{v_{i}}\times\{t_{i}\}, to a uniformly bounded neighborhood of M(ti)𝚯vM(t_{i})\cap{\bf\Theta}^{v}. Consequently,

(7) d0(M(t1),M(t2))d0(M(t1)𝚯v,M(t2)𝚯v)d_{0}(M(t_{1}),M(t_{2}))\asymp d_{0}(M(t_{1})\cap{\bf\Theta}^{v},M(t_{2})\cap{\bf\Theta}^{v})

with uniform constants.

Next, observe that M(ti)𝚯v𝚯v𝚯vi𝚯vM(t_{i})\cap{\bf\Theta}^{v}\subset{\bf\Theta}^{v}\cap{\bf\Theta}^{v_{i}}\subset{\bf\Theta}^{v}.

Claim 3.7.

The quasi-isometry μv×λv\mu^{v}\times\lambda^{v} maps M(ti)𝚯vM(t_{i})\cap{\bf\Theta}^{v} within a uniformly bounded Hausdorff distance of the slice {zi}×𝒦vS~v×𝒦v\{z_{i}\}\times\mathcal{K}^{v}\subset\widetilde{S}^{v}\times\mathcal{K}^{v}, for each i=1,2i=1,2, where (zi,ti)(z_{i},t_{i}^{\prime}) is a point in the μv×λv\mu^{v}\times\lambda^{v}–image of M(s)M(ti)M(s)\cap M(t_{i}).

Assuming the claim, we note then that

d0(M(t1)𝚯v,M(t2)𝚯v)\displaystyle d_{0}(M(t_{1})\cap{\bf\Theta}^{v},M(t_{2})\cap{\bf\Theta}^{v}) \displaystyle\asymp dS~v×𝒦v({z1}×𝒦v,{z2}×𝒦v)\displaystyle d_{\widetilde{S}^{v}\times\mathcal{K}^{v}}(\{z_{1}\}\times\mathcal{K}^{v},\{z_{2}\}\times\mathcal{K}^{v})
=\displaystyle= dS~v(z1,z2)\displaystyle d_{\widetilde{S}^{v}}(z_{1},z_{2})
\displaystyle\asymp d0(M(s)M(t1),M(s)M(t2))\displaystyle d_{0}(M(s)\cap M(t_{1}),M(s)\cap M(t_{2}))

again with uniform constants. Combining this coarse equation with (7) we get the required uniform estimate

d0(M(s)M(t1),M(s)M(t2))d0(M(t1),M(t2)).d_{0}(M(s)\cap M(t_{1}),M(s)\cap M(t_{2}))\asymp d_{0}(M(t_{1}),M(t_{2})).

Fix i=1i=1 or 22 and we prove the claim. Since λv×λvi\lambda^{v}\times\lambda^{v_{i}} is K1K_{1}–coarsely surjective (Proposition 3.1(3), there exists some point yi𝚯v𝚯viy_{i}\in{\bf\Theta}^{v}\cap{\bf\Theta}^{v_{i}} with (λv(yi),λvi(yi))(\lambda^{v}(y_{i}),\lambda^{v_{i}}(y_{i})) within distance K1K_{1} of (s,ti)𝒦v×𝒦vi(s,t_{i})\in\mathcal{K}^{v}\times\mathcal{K}^{v_{i}}. Therefore, yiM(s)M(ti)y_{i}\in M(s)\cap M(t_{i}) and we set μv×λv(yi)=(zi,ti)\mu^{v}\times\lambda^{v}(y_{i})=(z_{i},t_{i}^{\prime}).

Next, we observe that λvi\lambda^{v_{i}} is uniformly coarsely constant on any line of 𝚯v{\bf\Theta}^{v} contained in 𝚯v𝚯vi{\bf\Theta}^{v}\cap{\bf\Theta}^{v_{i}} by Proposition 3.1(2) and uniformly coarsely Lipschitz by Proposition 3.1(1). Hence, the line

(μv)1(zi)=(μv×ρv)1({zi}×)(\mu^{v})^{-1}(z_{i})=(\mu^{v}\times\rho^{v})^{-1}(\{z_{i}\}\times\mathbb{R})

of 𝚯v{\bf\Theta}^{v} maps under λv×λvi:𝚯v𝚯vi𝒦v×𝒦vi\lambda^{v}\times\lambda^{v_{i}}:{\bf\Theta}^{v}\cap{\bf\Theta}^{v_{i}}\to\mathcal{K}^{v}\times\mathcal{K}^{v_{i}} into a neighborhood of uniformly bounded radius of 𝒦v×{ti}\mathcal{K}^{v}\times\{t_{i}\}. Therefore, any point in the image of the line in 𝒦v×𝒦vi\mathcal{K}^{v}\times\mathcal{K}^{v_{i}} lies uniformly close to a point in λv×λvi(M(ti)𝚯v)\lambda^{v}\times\lambda^{v_{i}}(M(t_{i})\cap{\bf\Theta}^{v}) by Proposition 3.1(3) (which guarantees that any point in 𝒦v×{ti}\mathcal{K}^{v}\times\{t_{i}\} is K1K_{1}–close to a point in the image of the subspace 𝚯v𝚯vi{\bf\Theta}^{v}\cap{\bf\Theta}^{v_{i}}). Therefore any point in the line lies uniformly close to some point in M(ti)𝚯vM(t_{i})\cap{\bf\Theta}^{v} since λv×λvi\lambda^{v}\times\lambda^{v_{i}} is a uniform quasi-isometry again by Proposition 3.1(3).

On the other hand, Lemma 3.5 implies μvi×λvi(M(ti))\mu^{v_{i}}\times\lambda^{v_{i}}(M(t_{i})) is uniformly bounded Hausdorff distance from the slice S~vi×{ti}S~vi×𝒦vi\widetilde{S}^{v_{i}}\times\{t_{i}\}\subset\widetilde{S}^{v_{i}}\times\mathcal{K}^{v_{i}}. Moreover, since M(ti)M(t_{i}) meets every line of 𝚯vi{\bf\Theta}^{v_{i}} (Lemma 3.5 again), it follows that

μvi×λvi(M(ti)𝚯v)\mu^{v_{i}}\times\lambda^{v_{i}}(M(t_{i})\cap{\bf\Theta}^{v})

is uniformly bounded Hausdorff distance to the quasi-line vS~vi×{ti}\partial^{v}\widetilde{S}^{v_{i}}\times\{t_{i}\} (see the proof of Lemma 3.5). In particular, M(ti)𝚯vM(t_{i})\cap{\bf\Theta}^{v} is itself a uniform quasi-line and consequently lies within a uniformly bounded neighborhood of the line (μv)1(zi)(\mu^{v})^{-1}(z_{i}). Since this line maps within within a uniformly bounded Hausdorff distance of the slice {zi}×𝒦v\{z_{i}\}\times\mathcal{K}^{v} in S~v×𝒦v\widetilde{S}^{v}\times\mathcal{K}^{v} by μv×λv\mu^{v}\times\lambda^{v}, we see that M(ti)𝚯vM(t_{i})\cap{\bf\Theta}^{v} does as well ∎

3.3. Coned-off surfaces

For v𝒱v\in\mathcal{V}, we define 𝚵v{\bf\Xi}^{v} to be the graph whose vertices are all w𝒱w\in\mathcal{V} so that dtree(v,w)=1d_{\mathrm{tree}}(v,w)=1, and with edges connecting the pairs w,ww,w^{\prime} whenever 𝚯w𝚯w{\bf\Theta}^{w}\cap{\bf\Theta}^{w^{\prime}}\neq\emptyset. As such, vertices w𝒱w\in\mathcal{V} are in bijective correspondence with the boundary components of 𝚯v{\bf\Theta}^{v} and there is an “inclusion” map

iv:𝚯v𝚵vi^{v}\colon\partial{\bf\Theta}^{v}\to{\bf\Xi}^{v}

that sends any point x𝚯vx\in\partial{\bf\Theta}^{v} to the vertex ww for which x𝚯wx\in{\bf\Theta}^{w}. In light of the following lemma, we note that we could alternately define the edges of 𝚵v{\bf\Xi}^{v} in terms of subspaces lying within bounded distance of each other, and produce a space quasi-isometric to 𝚵v{\bf\Xi}^{v}.

Lemma 3.8.

There is a function N=N𝚵:[0,)[0,)N=N_{{\bf\Xi}}\colon[0,\infty)\to[0,\infty) so that whenever v,w1,w2𝒱v,w_{1},w_{2}\in\mathcal{V} satisfy d¯(𝚯w1𝚯v,𝚯w2𝚯v)r\bar{d}({\bf\Theta}^{w_{1}}\cap{\bf\Theta}^{v},{\bf\Theta}^{w_{2}}\cap{\bf\Theta}^{v})\leq r, the sets 𝚯w1𝚯v{\bf\Theta}^{w_{1}}\cap{\bf\Theta}^{v} and 𝚯v𝚯w2{\bf\Theta}^{v}\cap{\bf\Theta}^{w_{2}} may be connected via a concatenation of at most N(r)N(r) paths, each of which is contained in a set of the form 𝚯v𝚯w{\bf\Theta}^{v}\cap{\bf\Theta}^{w}.

Proof.

If d¯(𝚯w1𝚯v,𝚯w2𝚯v)r\bar{d}({\bf\Theta}^{w_{1}}\cap{\bf\Theta}^{v},{\bf\Theta}^{w_{2}}\cap{\bf\Theta}^{v})\leq r, then there are cone points pi𝚯wiθvp_{i}\in{\bf\Theta}^{w_{i}}\cap\theta^{v} within distance r+3Mr+3M, where MM is the bound on the width of a strip and length of a saddle connection from Lemma 2.3. Since the path metric dθvd_{\theta^{v}} on θv\theta^{v} is coarsely equivalent to the subspace metric, dθv(p1,p2)d_{\theta^{v}}(p_{1},p_{2}) is bounded in terms of rr. The path metric on θv\theta^{v} is biLipschitz equivalent to the 1\ell^{1}–metric on the product θvXv×\theta^{v}_{X_{v}}\times\mathbb{R}. Since each edge of θvXv\theta^{v}_{X_{v}} has definite length, there is a path from p1p_{1} to p2p_{2} in θv\theta^{v} obtained by concatenating boundedly many (in terms of rr) paths αi\alpha_{i} with fXv(αi)f_{X_{v}}(\alpha_{i}) a saddle connection in θvXv\theta^{v}_{X_{v}}. Since each αi\alpha_{i} is contained in some 𝚯w{\bf\Theta}^{w}, we are done. ∎

Corollary 3.9.

For any x,y𝚯vx,y\in\partial{\bf\Theta}^{v}, we have d𝚵v(iv(x),iv(y))N(d𝚯v(x,y))+1d_{{\bf\Xi}^{v}}(i^{v}(x),i^{v}(y))\leq N(d_{{\bf\Theta}^{v}}(x,y))+1.

Proof.

Let w=iv(x)w=i^{v}(x) and w=iv(y)w^{\prime}=i^{v}(y). Since d𝚯vd_{{\bf\Theta}^{v}} and d¯\bar{d} are path metrics, we have d¯(x,y)d𝚯v(x,y)\bar{d}(x,y)\leq d_{{\bf\Theta}^{v}}(x,y). By Lemma 3.8, xx and yy may be joined by a concatenation α1αk\alpha_{1}\dotsb\alpha_{k} of kN(d𝚯v(x,y))+2k\leq N(d_{{\bf\Theta}^{v}}(x,y))+2 paths αj\alpha_{j} each of which lies in some 𝚯v𝚯wj{\bf\Theta}^{v}\cap{\bf\Theta}^{w_{j}}, and where w=w1w=w_{1} and w=wkw^{\prime}=w_{k}. For successive paths αj,αj+1\alpha_{j},\alpha_{j+1}, the vertices wj,wj+1w_{j},w_{j+1} are adjacent in 𝚵v{\bf\Xi}^{v} by definition. Therefore d𝚵v(w,w)k1N(d𝚯v(x,y))+1d_{{\bf\Xi}^{v}}(w,w^{\prime})\leq k-1\leq N(d_{{\bf\Theta}^{v}}(x,y))+1. ∎

Lemma 3.10.

Each 𝚵v{\bf\Xi}^{v} is uniformly quasi-isometric to a tree. In particular, there exists δ>0\delta>0 so that each 𝚵v{\bf\Xi}^{v} is δ\delta–hyperbolic. Moreover, 𝚵v{\bf\Xi}^{v} has at least two points at infinity.

Proof.

We appeal to Proposition 2.5 and show that for any vertices w,ww,w^{\prime} of 𝚵v{\bf\Xi}^{v} there exists a path γ(w,w)\gamma(w,w^{\prime}) so that any path from ww to ww^{\prime} passes within distance 33 of each vertex of γ(w,w)\gamma(w,w^{\prime}).

First, note that 𝚵v{\bf\Xi}^{v} is isomorphic to the intersection graph of the collection of strips in 𝚯0v{\bf\Theta}_{0}^{v}. For each strip we have a vertex, and for each saddle connection of the spine θ0v\theta_{0}^{v}, there is an edge of 𝚵v{\bf\Xi}^{v} that connects the vertices corresponding to the strips that contain the saddle connection. For each cone point in the spine θ0v\theta_{0}^{v}, there is also a complete graph on the vertices corresponding to strips that contain this cone point. This accounts for all edges (because intersections of strips either arise along saddle connections or single cone points), and we note that the closure of each edge of the first type separates 𝚵v{\bf\Xi}^{v} into two components.

Suppose w,w𝚵vw,w^{\prime}\in{\bf\Xi}^{v} are two vertices, and let x,xθ0vx,x^{\prime}\in\theta_{0}^{v} be points in the (boundaries of the) strips corresponding to ww and ww^{\prime}, respectively, that are closest in 𝚯0v{\bf\Theta}_{0}^{v}, and consider the geodesic in θ0v\theta_{0}^{v} connecting these points, which is a concatenation of saddle connections σ1σ2σn\sigma_{1}\sigma_{2}\cdots\sigma_{n}. For each 1in1\leq i\leq n, let wi±w_{i}^{\pm} be the vertices corresponding to the two strips Ai±A_{i}^{\pm} that intersect in the saddle connection σi\sigma_{i}. We can form an edge path γ(w,w)\gamma(w,w^{\prime}) in 𝚵v{\bf\Xi}^{v}, containing w,ww,w^{\prime}, and the wi+w_{i}^{+} as vertices, since Ai+Ai+1+A_{i}^{+}\cap A_{i+1}^{+}\neq\emptyset. Observe that any path from xx to xx^{\prime} must pass through the union Ai+AiA_{i}^{+}\cup A_{i}^{-}, for each ii, since xx and xx^{\prime} lie in the closures of distinct components of 𝚯0v(Ai+Ai){\bf\Theta}_{0}^{v}\smallsetminus(A_{i}^{+}\cup A_{i}^{-}).

Now let w=w0,w1,,wk=ww=w_{0},w_{1},\ldots,w_{k}=w^{\prime} be the vertices of an edge path connecting ww to ww^{\prime} in 𝚵v{\bf\Xi}^{v}. For any points in the strips corresponding to ww and ww^{\prime}, respectively, it is easy to construct a path in 𝚯0v{\bf\Theta}_{0}^{v} between these points that decomposes as a concatenation ν1ν2νk\nu_{1}\nu_{2}\cdots\nu_{k} so that νj\nu_{j} is contained entirely in the strip corresponding to wjw_{j}. From the previous paragraph, this path must pass through Ai+AiA_{i}^{+}\cup A_{i}^{-}, for each i=1,,ni=1,\ldots,n. It follows that for each 1in1\leq i\leq n, the edge path must meet the union of the stars star(wi+)star(wi)star(w_{i}^{+})\cup star(w_{i}^{-}). Since these stars intersect, their union has diameter at most 33, and we are done.

We now show that 𝚵v{\bf\Xi}^{v} contains a quasi-geodesic line. Consider strips AiA_{i} of 𝚯0v{\bf\Theta}_{0}^{v}, for ii\in\mathbb{Z}, such that for all ii we have

  • AiA_{i} and Ai+1A_{i+1} share a saddle connection;

  • Ai1A_{i-1} and Ai+2A_{i+2} lie on distinct components of the complement of the interior of AiAi+1A_{i}\cup A_{i+1} in 𝚯0v{\bf\Theta}_{0}^{v}.

The AiA_{i} give a bi-infinite path in 𝚵v{\bf\Xi}^{v}, and we now show that this path is a quasi-geodesic. Fix integers m,nm,n and consider a geodesic γ\gamma in 𝚵v{\bf\Xi}^{v} from AmA_{m} to AnA_{n} (where we think of the strips themselves as vertices of 𝚵v{\bf\Xi}^{v} for convenience). Then for each m<k<n1m<k<n-1 we have that γ\gamma needs to contain a vertex v(k)v(k) which, regarded as a strip, intersects AkA_{k} or Ak+1A_{k+1}. Indeed, the interior of AkAk+1A_{k}\cup A_{k+1} separates AmA_{m} from AnA_{n}, and the sequence of vertices of γ\gamma corresponds to a connected union of strips containing AmA_{m} and AnA_{n}. Moreover, there is no strip intersecting both AkA_{k} and AkA_{k^{\prime}} if |kk|3|k-k^{\prime}|\geq 3, and in particular we have v(k)v(k)v(k)\neq v(k^{\prime}) if |kk|4|k-k^{\prime}|\geq 4. These observations imply that γ\gamma contains at least (nm2)/4\lfloor(n-m-2)/4\rfloor vertices, so that geodesics connecting AmA_{m} to AnA_{n} have length comparable to nmn-m, and the AiA_{i} form a quasi-geodesic line as required. ∎

3.4. Windows and bridges

Recall that ΣE\Sigma\subset E and Σ¯E¯\bar{\Sigma}\subset\bar{E} are the sets of all singular points in all fibers in EE and E¯\bar{E}, respectively; see §2.2. For v𝒱v\in\mathcal{V}, consider the set Σ¯v\bar{\Sigma}^{v} of points in Σ¯\bar{\Sigma} that are inside some vv–spine, as well as those points Σ¯v\bar{\Sigma}^{\notin v} that are outside every vv–spine:

Σ¯v=XD¯(θvXΣ)andΣ¯v=Σ¯Σ¯v.\bar{\Sigma}^{v}=\bigcup_{X\in\bar{D}}(\theta^{v}_{X}\cap\Sigma)\qquad\text{and}\qquad\bar{\Sigma}^{\not\in v}=\bar{\Sigma}\setminus\bar{\Sigma}^{v}.

For each YD¯Y\in\bar{D} we now define a window map ΠvY:Σ¯𝐏(𝚯Yv)\Pi^{v}_{Y}\colon\bar{\Sigma}\to{\bf P}(\partial{\bf\Theta}_{Y}^{v}) from cone points to the set of subsets of the boundary 𝚯Yv𝚯vY\partial{\bf\Theta}_{Y}^{v}\subset{\bf\Theta}^{v}_{Y}. There are two cases. Firstly,

for xΣ¯v,ΠvY(x)={z𝚯vY[fY(x),z]𝚯vY={z}}.\text{for $x\in\bar{\Sigma}^{\notin v}$,}\quad\Pi^{v}_{Y}(x)=\big{\{}z\in\partial{\bf\Theta}^{v}_{Y}\mid[f_{Y}(x),z]\cap{\bf\Theta}^{v}_{Y}=\{z\}\big{\}}.

In words, ΠvY(x)\Pi^{v}_{Y}(x) is the union of entrance points in 𝚯vY{\bf\Theta}^{v}_{Y} of any flat geodesic in EYE_{Y} from fY(x)f_{Y}(x) to 𝚯vY{\bf\Theta}^{v}_{Y} (basically the closest point projection in EYE_{Y}), and we call it the window for xx in 𝚯vY{\bf\Theta}^{v}_{Y}. Observe that for any X,YD¯X,Y\in\bar{D} we have fX,Y(ΠYv(x))=ΠXv(x)f_{X,Y}(\Pi_{Y}^{v}(x))=\Pi_{X}^{v}(x). The second case, that of Σ¯v\bar{\Sigma}^{v}, is handled slightly differently:

for xΣ¯v,ΠvY(x)\displaystyle\text{for $x\in\bar{\Sigma}^{v}$,}\quad\Pi^{v}_{Y}(x) =fY,X(ΠvX(x)),whereX=cv(π(x))Bv\displaystyle=f_{Y,X}(\Pi^{v}_{X}(x)),\quad\text{where}\quad X=c_{v}(\pi(x))\in\partial B_{v}
andΠvX(x)\displaystyle\text{and}\quad\Pi^{v}_{X}(x) ={z𝚯vXz is a closest cone point to fX(x)θvX}.\displaystyle=\big{\{}z\in\partial{\bf\Theta}^{v}_{X}\mid z\text{ is a closest cone point to }f_{X}(x)\in\theta^{v}_{X}\big{\}}.

Thus affine invariance ΠvZ(x)=fZ,Y(ΠvY(x))\Pi^{v}_{Z}(x)=f_{Z,Y}(\Pi^{v}_{Y}(x)) is built directly into the definition.

Now for any YD¯Y\in\bar{D} and x,yΣ¯x,y\in\bar{\Sigma}, we define

dYv(x,y)=diam(ΠvY(x)ΠvY(y)),d_{Y}^{v}(x,y)=\mathrm{diam}(\Pi^{v}_{Y}(x)\cup\Pi^{v}_{Y}(y)),

where the distance is computed in the path metric on EYE_{Y} (or equivalently on 𝚯vY{\bf\Theta}^{v}_{Y}). Finally, we extend window maps to arbitrary subsets by declaring

for UE¯,ΠvY(U)=ΠvY(UΣ¯)=xΣ¯UΠvY(x)\text{for $U\subset\bar{E}$,}\quad\Pi^{v}_{Y}(U)=\Pi^{v}_{Y}(U\cap\bar{\Sigma})=\bigcup_{x\in\bar{\Sigma}\cap U}\Pi^{v}_{Y}(x)

which has the same effect as defining ΠvY(x)=𝚯vY\Pi^{v}_{Y}(x)=\emptyset\subset\partial{\bf\Theta}^{v}_{Y} for xΣ¯x\notin\bar{\Sigma}.

Lemma 3.11.

If x,yΣ¯vx,y\in\bar{\Sigma}^{\not\in v} satisfy f0(x)=f0(y)f_{0}(x)=f_{0}(y), then ΠvY(x)=ΠvY(y)\Pi^{v}_{Y}(x)=\Pi^{v}_{Y}(y) for all YD¯Y\in\bar{D}.

Recall that f0=fX0f_{0}=f_{X_{0}}.

Proof.

This is immediate since f0(x)=f0(y)f_{0}(x)=f_{0}(y) if and only if fY(x)=fY(y)f_{Y}(x)=f_{Y}(y) for all YD¯Y\in\bar{D}, and ΠvY(x)\Pi^{v}_{Y}(x) is defined just in terms of fY(x)=fY(y)f_{Y}(x)=f_{Y}(y). ∎

The following gives a counterpoint to Lemma 3.11 for points in Σ¯v\bar{\Sigma}^{v}.

Lemma 3.12.

There exists K2>0K_{2}>0 such that for any v𝒱v\in\mathcal{V} the following holds: If x,yΣ¯vx,y\in\bar{\Sigma}^{v} satisfy f0(x)=f0(y)f_{0}(x)=f_{0}(y) and either

  1. (1)

    xx and yy are connected by a horizontal geodesic of length 1\leq 1, or

  2. (2)

    xx and yy are contained in w\partial\mathcal{B}_{w} for some w𝒱w\in\mathcal{V} with α(w)α(v)\alpha(w)\neq\alpha(v),

then dXv(x,y)K2d_{X}^{v}(x,y)\leq K_{2}, where X=cv(π(x))X=c_{v}(\pi(x)).

Proof.

Set X=cv(π(x))X=c_{v}(\pi(x)) and Y=cv(π(y))Y=c_{v}(\pi(y)). Since cv:DBvc_{v}\colon D\to\partial B_{v} is 11–Lipschitz and diam(cv(Bw))\mathrm{diam}(c_{v}(\partial B_{w})) is uniformly bounded for all such ww, either condition (1) or (2) gives a uniform bound K>0K>0 on the distance between XX and YY. Hence fX,Yf_{X,Y} is eKe^{K}–biLipschitz. The distance between fY(y)θvYf_{Y}(y)\in\theta^{v}_{Y} and its closest cone points ΠvY(y)\Pi^{v}_{Y}(y) in 𝚯vY\partial{\bf\Theta}^{v}_{Y} is also uniformly bounded by 2M2M, by Lemma 2.3. The same holds for the distance between fX(x)f_{X}(x) and ΠvX(x)\Pi^{v}_{X}(x). It follows that ΠvX(y)=fX,Y(ΠvY(y))\Pi^{v}_{X}(y)=f_{X,Y}(\Pi^{v}_{Y}(y)) lies within distance 2eKM2e^{K}M of fX(x)=fX,Y(fY(y))f_{X}(x)=f_{X,Y}(f_{Y}(y)) and hence within distance 2eKM+2M2e^{K}M+2M of ΠvX(x)\Pi^{v}_{X}(x). ∎

The next lemma explains that the image of ΠvY\Pi^{v}_{Y} is not so far from being a point.

Lemma 3.13 (Window lemma).

For any v𝒱v\in\mathcal{V}, YD¯Y\in\bar{D}, and xΣ¯vx\in\bar{\Sigma}^{\not\in v}, the window ΠvY(x)𝚯vY\Pi^{v}_{Y}(x)\subset\partial{\bf\Theta}^{v}_{Y} is either a cone point or a single saddle connection.

Proof.

Recall that each cone point in the flat surface EYE_{Y} has total angle at least 3π3\pi, and that EYE_{Y} is a unique geodesic space in which a concatenation of saddle connections is geodesic if and only if successive saddle connections subtend an angle of at least π\pi on each side.

If fY(x)𝚯vYf_{Y}(x)\in\partial{\bf\Theta}^{v}_{Y} then clearly ΠvY(x)\Pi^{v}_{Y}(x) is the cone point fY(x)f_{Y}(x) itself. So suppose fY(x)𝚯vYf_{Y}(x)\notin{\bf\Theta}^{v}_{Y} and let \ell be the component of 𝚯vY\partial{\bf\Theta}^{v}_{Y} separating fY(x)f_{Y}(x) from θvY\theta^{v}_{Y} in EYE_{Y}, so that ΠvY(x)\Pi^{v}_{Y}(x)\subset\ell. Take any flat geodesic [fY(x),z][f_{Y}(x),z] in EYE_{Y} from fY(x)f_{Y}(x) to a cone point zz\in\ell. The geodesic [fY(x),z][f_{Y}(x),z] is a concatenation of saddle connections and first meets \ell in some cone point pp. Since the total cone angle at pp is at least 3π3\pi and the angle at pp along the side of \ell containing 𝚯v{\bf\Theta}^{v} is exactly π\pi, the last saddle connection δ\delta in the geodesic [fY(x),p][fY(x),z][f_{Y}(x),p]\subset[f_{Y}(x),z] must make an angle of at least π\pi with one of the two halves of \ell determined by pp. It follows that concatenating [fY(x),p][f_{Y}(x),p] with that half of \ell gives an infinite geodesic ray in EYE_{Y}. Hence, by uniqueness of geodesics, the geodesic from fY(x)f_{Y}(x) to any cone point on that side of pp evidently passes through pp.

The strip between θvY\theta^{v}_{Y} and fY(x)f_{Y}(x)The vv spine, θvY\theta^{v}_{Y}The window ΠvY(x)\Pi^{v}_{Y}(x)π\geq\piπ\geq\pifY(x)f_{Y}(x)
Figure 3. The window (shown in red) for xx in the thickend spine 𝚯vY{\bf\Theta}^{v}_{Y}

If both angles between δ\delta and \ell at pp are at least π\pi, then any geodesic from fY(x)f_{Y}(x) to \ell passes through pp. Hence pp is the unique point in 𝚯vY\partial{\bf\Theta}^{v}_{Y} closest to fY(x)f_{Y}(x) and ΠvY(x)={p}\Pi^{v}_{Y}(x)=\{p\} is a cone point as required. Otherwise, consider the flat geodesic from fY(x)f_{Y}(x) to the adjacent cone point pp^{\prime} on the other side of pp along \ell. The last saddle connection of this geodesic must also make an angle with \ell of at least π\pi on one side. This cannot be the side containing pp, or else the geodesic from fY(x)f_{Y}(x) to pp would pass through pp^{\prime} contradicting our choice of pp. Hence any geodesic from fY(x)f_{Y}(x) to a cone point on the opposite side of pp^{\prime} must pass through pp^{\prime}. Therefore ΠvY(x)\Pi^{v}_{Y}(x) is the saddle connection between pp and pp^{\prime}, and we are done. See Figure 3. ∎

The following lemma gives us partial control over the window for points in adjacent vertex spaces in the same Bass–Serre tree.

Lemma 3.14 (Bridge lemma).

For any v,w𝒱v,w\in\mathcal{V} with dtree(v,w)=1d_{\mathrm{tree}}(v,w)=1, any YD¯Y\in\bar{D}, and any component UEYθYwU\subset E_{Y}\smallsetminus\theta_{Y}^{w} not containing θvY\theta^{v}_{Y}, there exists a (possibly degenerate) saddle connection δU𝚯vY\delta_{U}\subset\partial{\bf\Theta}^{v}_{Y} with the following property: Every xΣ¯x\in\bar{\Sigma} with fY(x)U¯f_{Y}(x)\in\overline{U} satisfies ΠvY(x)δU\Pi^{v}_{Y}(x)\subset\delta_{U}.

We call δU\delta_{U} the bridge for UU in EYE_{Y}. It is clear from the construction in the proof below that fZ,Y(δU)f_{Z,Y}(\delta_{U}) is the bridge for fZ,Y(U)f_{Z,Y}(U), for any ZD¯Z\in\bar{D}.

Proof.

Let UU be as in the statement and WW be the component of EYθwYE_{Y}\smallsetminus\theta^{w}_{Y} containing θvY\theta^{v}_{Y}. Let γU=U¯θwY\gamma_{U}=\overline{U}\cap\theta^{w}_{Y} and γW=W¯θwY𝚯vY\gamma_{W}=\overline{W}\cap\theta^{w}_{Y}\subset\partial{\bf\Theta}^{v}_{Y}, which are both bi-infinite flat geodesics in θwY\theta^{w}_{Y}.

If γUγW=\gamma_{U}\cap\gamma_{W}=\emptyset, then there is a unique geodesic between them in θwY\theta^{w}_{Y}, and we take δU\delta_{U} to be the endpoint of this geodesic which lies along γW\gamma_{W}. On the other hand, if γUγW\gamma_{U}\cap\gamma_{W}\neq\emptyset, then their intersection is contained in the boundary of a strip along θvY\theta^{v}_{Y} and another along θwY\theta^{w}_{Y}. Two distinct strips in the same direction that intersect do so in either a single point or a single saddle connection, and hence γUγW\gamma_{U}\cap\gamma_{W} is a point or single saddle connection, and we call δU\delta_{U}. See Figure 4.

UU^{\prime}UUδU\delta_{U^{\prime}}δU\delta_{U}θYw\theta_{Y}^{w}θvY\theta^{v}_{Y}
Figure 4. Examples of bridges (in red), and the proof of Lemma 3.14.

Now consider any point xΣ¯x\in\bar{\Sigma} with fY(x)U¯f_{Y}(x)\in\overline{U}. Observe that xΣ¯vx\in\bar{\Sigma}^{\notin v} so that ΠvY(x)\Pi^{v}_{Y}(x) falls under the first case of the window definition. Further, any flat geodesic from fY(x)f_{Y}(x) to 𝚯vY{\bf\Theta}^{v}_{Y} must pass through both γU\gamma_{U} and γW\gamma_{W}, and hence must pass through δU𝚯vY\delta_{U}\subset\partial{\bf\Theta}^{v}_{Y}. It follows that ΠvY(x)δU\Pi^{v}_{Y}(x)\subset\delta_{U}, as required. ∎

The following is an easy consequence of the previous lemma.

Corollary 3.15.

For any v,w𝒱v,w\in\mathcal{V} with 2dtree(v,w)<2\leq d_{\mathrm{tree}}(v,w)<\infty and YD¯Y\in\bar{D}, there exists a (possibly degenerate) saddle connection δvY(w)𝚯vY\delta^{v}_{Y}(w)\subset\partial{\bf\Theta}^{v}_{Y} so that if xΣ¯x\in\bar{\Sigma} has fY(x)𝚯wYf_{Y}(x)\in{\bf\Theta}^{w}_{Y}, then ΠvY(x)δvY(w)\Pi^{v}_{Y}(x)\subset\delta^{v}_{Y}(w). In particular, ΠvY(𝚯w)δvY(w)\Pi^{v}_{Y}({\bf\Theta}^{w})\subset\delta^{v}_{Y}(w).

Proof.

There exists uTα(v)(0)u\in T_{\alpha(v)}^{(0)} between vv and ww with dtree(u,v)=1d_{\mathrm{tree}}(u,v)=1, and a component UEYθuYU\subset E_{Y}\setminus\theta^{u}_{Y} whose closure contains 𝚯wY{\bf\Theta}^{w}_{Y}. Setting δvY(w)=δU\delta^{v}_{Y}(w)=\delta_{U} and applying Lemma 3.14 completes the proof. ∎

The next corollary is similar.

Corollary 3.16.

If v,w𝒱v,w\in\mathcal{V} satisfy dtree(v,w)=d_{\mathrm{tree}}(v,w)=\infty and θw0θv0=\theta^{w}_{0}\cap\theta^{v}_{0}=\emptyset, then for each YD¯Y\in\bar{D} there is a connected union δvY(w)𝚯vY\delta^{v}_{Y}(w)\subset\partial{\bf\Theta}^{v}_{Y} of at most two saddle connections such that ΠvY(x)δvY(w)\Pi^{v}_{Y}(x)\subset\delta^{v}_{Y}(w) for all xΣ¯wx\in\bar{\Sigma}^{w}. In particular ΠvY(θw)δvY(w)\Pi^{v}_{Y}(\theta^{w})\subset\delta^{v}_{Y}(w).

Proof.

From the hypotheses, θwY\theta^{w}_{Y} is contained in some component WEYθvYW\subset E_{Y}\smallsetminus\theta^{v}_{Y}. Let u𝒱u\in\mathcal{V} be such that dtree(v,u)=1d_{\mathrm{tree}}(v,u)=1 and θuYW\theta^{u}_{Y}\subset W. Since α(u)=α(v)α(w)\alpha(u)=\alpha(v)\neq\alpha(w), θuY\theta^{u}_{Y} and θwY\theta^{w}_{Y} can intersect in at most one point. If θuYθwY=\theta^{u}_{Y}\cap\theta^{w}_{Y}=\emptyset, then θwY\theta^{w}_{Y} is contained in a component UU of EYθuYE_{Y}\setminus\theta^{u}_{Y} disjoint from θvY\theta^{v}_{Y}; thus ΠvY(θwY)\Pi^{v}_{Y}(\theta^{w}_{Y}) is contained in the bridge δU\delta_{U} for UU by Lemma 3.14. Otherwise θuYθwY\theta^{u}_{Y}\cap\theta^{w}_{Y}\neq\emptyset, and we claim there is a cone point pθuYp\in\theta^{u}_{Y} such that pU¯p\in\overline{U} for every component UU of EYθuYE_{Y}\setminus\theta^{u}_{Y} that intersects θwY\theta^{w}_{Y}. Indeed, if θuYθwY\theta^{u}_{Y}\cap\theta^{w}_{Y} is a cone point, we take pp to be this intersection point, and if not θuYθwY\theta^{u}_{Y}\cap\theta^{w}_{Y} is an interior point of a saddle connection of θuY\theta^{u}_{Y} and we may take pp to be either of its endpoints. For each component UEYθuYU\subset E_{Y}\setminus\theta^{u}_{Y} that intersects θwY\theta^{w}_{Y} we then have ΠvY(p)δU\Pi^{v}_{Y}(p)\subset\delta_{U}, where δU\delta_{U} is the bridge for UU. Since fY(Σ¯w)θwYf_{Y}(\bar{\Sigma}^{w})\subset\theta^{w}_{Y} is contained in the union of the closures of such UU, it follows that ΠvY(Σ¯w)\Pi^{v}_{Y}(\bar{\Sigma}^{w}) is contained in a union of saddle connections along 𝚯vY\partial{\bf\Theta}^{v}_{Y}, all of which contain ΠvY(p)\Pi^{v}_{Y}(p), and hence is a connected union of at most two saddle connections. This completes the proof. ∎

The final case to consider is that of spines in different directions that intersect:

Lemma 3.17.

There exists K3>0K_{3}>0 such that if v,w𝒱v,w\in\mathcal{V} with dtree(v,w)=d_{\mathrm{tree}}(v,w)=\infty and θw0θv0\theta^{w}_{0}\cap\theta^{v}_{0}\neq\emptyset, then diam(ΠvY(θw))<K3\mathrm{diam}(\Pi^{v}_{Y}(\theta^{w}))<K_{3} for all Ycv(Bw)Y\in c_{v}(\partial B_{w}).

x1x_{1}W1W_{1}x2x_{2}W2W_{2}x3x_{3}W3W_{3}x4x_{4}W4W_{4}x5x_{5}W5W_{5}θwY\theta^{w}_{Y}W0W_{0}𝚯vY{\bf\Theta}^{v}_{Y}
Figure 5. The proof of Lemma 3.17: The arcs shown in red are the segments δi\delta_{i}, which contain ΠvY(x)\Pi^{v}_{Y}(x) for all xθwx\in\theta^{w} with fY(x)f_{Y}(x) in the subgraph WiW_{i} of θwY\theta^{w}_{Y}.
Proof.

Let x0=θvYθwYx_{0}=\theta^{v}_{Y}\cap\theta^{w}_{Y} be the unique intersection point of the spines. Let W0W_{0} be the smallest subgraph of θwY\theta^{w}_{Y} containing θwY𝚯vY\theta^{w}_{Y}\cap{\bf\Theta}^{v}_{Y} and let W1,,WkW_{1},\dots,W_{k} be the closures of the components of θwYW0\theta^{w}_{Y}\setminus W_{0}, so that θwY=W0Wk\theta^{w}_{Y}=W_{0}\cup\dots\cup W_{k}. See Figure 5. For 1ik1\leq i\leq k, let xiWix_{i}\in W_{i} be the closest (cone) point to x0x_{0}. Then define pip_{i} to be the intersection of the geodesic [x0,xi]θwY[x_{0},x_{i}]\subset\theta^{w}_{Y} with 𝚯vY\partial{\bf\Theta}^{v}_{Y} and let δi𝚯vY\delta_{i}\subset\partial{\bf\Theta}^{v}_{Y} be the segment consisting of the (11 or 22) saddle connections along 𝚯vY\partial{\bf\Theta}^{v}_{Y} that contain pip_{i}. Define δvY(w)=δ1δk\delta^{v}_{Y}(w)=\delta_{1}\cup\dots\cup\delta_{k}, and note that δvX(w)=fX,Y(δvY(w))\delta^{v}_{X}(w)=f_{X,Y}(\delta^{v}_{Y}(w)) and θwY𝚯vYδvY(w)\theta^{w}_{Y}\cap\partial{\bf\Theta}^{v}_{Y}\subset\delta^{v}_{Y}(w) by construction.

For any xΣ¯x\in\bar{\Sigma} with fY(x)Wif_{Y}(x)\in W_{i}, where i=1,,ki=1,\dots,k, the flat geodesic from fY(x)f_{Y}(x) to x0x_{0} first intersects 𝚯vY{\bf\Theta}^{v}_{Y} at pip_{i}; therefore piΠvY(x)p_{i}\in\Pi^{v}_{Y}(x) by definition of the window. By Lemma 3.13, it follows that ΠvY(x)δi\Pi^{v}_{Y}(x)\subset\delta_{i}. The union i=1kWi\cup_{i=1}^{k}W_{i} contains every cone point of θwY\theta^{w}_{Y} except possibly x0x_{0}. Thus we have proven ΠvY(θwΣ¯v)=ΠvY(θwYΣ¯v)δvY(w)\Pi^{v}_{Y}(\theta^{w}\cap\bar{\Sigma}^{\notin v})=\Pi^{v}_{Y}(\theta^{w}_{Y}\cap\bar{\Sigma}^{\notin v})\subset\delta^{v}_{Y}(w).

Let ZDZ\in D be the closest point on Bv\partial B_{v} to Bw\partial B_{w}, thus ZBvZ\in\partial B_{v} lies on the unique hyperbolic geodesic that intersects Bv\partial B_{v} and Bw\partial B_{w} orthogonally. In EZE_{Z} the directions α(v)\alpha(v) and α(w)\alpha(w) are perpendicular. Therefore the cone points of 𝚯vZ\partial{\bf\Theta}^{v}_{Z} that are closest to θvZθwZ\theta^{v}_{Z}\cap\theta^{w}_{Z} all lie in δvZ(w)\delta^{v}_{Z}(w), since they must be endpoints of saddle connections along 𝚯vZ\partial{\bf\Theta}^{v}_{Z} that intersect θwZ\theta^{w}_{Z}. More generally, for any Ycv(Bw)Y\in c_{v}(\partial B_{w}), the directions α(v)\alpha(v) and α(w)\alpha(w) are nearly perpendicular and thus we have

ΠvY(θwΣ¯v)δvY(w)BK(θvYθwY).\Pi^{v}_{Y}(\theta^{w}\cap\bar{\Sigma}^{\notin v})\subset\delta^{v}_{Y}(w)\subset B_{K}(\theta^{v}_{Y}\cap\theta^{w}_{Y}).

for some uniform constant K>0K>0 that depends only on the length of cv(Bw)c_{v}(\partial B_{w}) and the maximum over Bv\partial B_{v} of the length/width of any saddle connection/strip in the α(v)\alpha(v) direction (Lemma 2.3). Now, for any xθwΣ¯vx\in\theta^{w}\cap\bar{\Sigma}^{v}, the point X=cv(π(x))X=c_{v}(\pi(x)) lies in cv(Bw)c_{v}(\partial B_{w}) and we have that fX(x)=θvXθwXf_{X}(x)=\theta^{v}_{X}\cap\theta^{w}_{X}. The above equation shows there are cone points of 𝚯vX\partial{\bf\Theta}^{v}_{X} within KK of fX(x)f_{X}(x); hence ΠvX(x)=ΠvX(fX(x))\Pi^{v}_{X}(x)=\Pi^{v}_{X}(f_{X}(x)) lies in BK(θvXθwX)B_{K}(\theta^{v}_{X}\cap\theta^{w}_{X}) by definition. Using the fact that the length of cv(Bw)c_{v}(\partial B_{w}) is uniformly bounded, we see that the map fY,Xf_{Y,X} is uniformly biLipschitz and therefore that ΠvY(x)=fY,X(ΠvX(x))\Pi^{v}_{Y}(x)=f_{Y,X}(\Pi^{v}_{X}(x)) lies within bounded distance of θvYθwY=fY,X(θvXθwX)\theta^{v}_{Y}\cap\theta^{w}_{Y}=f_{Y,X}(\theta^{v}_{X}\cap\theta^{w}_{X}). Combining this with the above finding that ΠvY(θwΣ¯v)\Pi^{v}_{Y}(\theta^{w}\cap\bar{\Sigma}^{\notin v}) lies within bounded distance of θvYθwY\theta^{v}_{Y}\cap\theta^{w}_{Y}, we finally conclude that diam(ΠvY(θw))\mathrm{diam}(\Pi^{v}_{Y}(\theta^{w})) is uniformly bounded. ∎

3.5. Projections

Here we define projections Λv:Σ¯𝐏(𝒦v)\Lambda^{v}\colon\bar{\Sigma}\to{\bf P}(\mathcal{K}^{v}) and ξv:Σ¯𝐏(𝚵v)\xi^{v}\colon\bar{\Sigma}\to{\bf P}({\bf\Xi}^{v}). In preparation, we first define Πv:Σ¯𝐏(𝚯v)\Pi^{v}\colon\bar{\Sigma}\to{\bf P}(\partial{\bf\Theta}^{v}) by

Πv(x)=YBvΠvY(x),for any v𝒱 and xΣ¯.\Pi^{v}(x)=\bigcup_{Y\in\partial B_{v}}\Pi^{v}_{Y}(x),\quad\text{for any $v\in\mathcal{V}$ and $x\in\bar{\Sigma}$}.

In words, Πv(x)\Pi^{v}(x) consists of the vv–windows of xx in all fibers over Bv\partial B_{v}. As before, we extend to arbitrary subsets UE¯U\subset\bar{E} by setting Πv(U)=Πv(UΣ¯)\Pi^{v}(U)=\Pi^{v}(U\cap\bar{\Sigma}).

Now, for each v𝒱v\in\mathcal{V} our projections are defined as the compositions

Λv=λvΠv and ξv(x)=ivΠv.\Lambda^{v}=\lambda^{v}\circ\Pi^{v}\quad\mbox{ and }\quad\xi^{v}(x)=i^{v}\circ\Pi^{v}.

A useful observation is that for any two vertices v,w𝒱v,w\in\mathcal{V} with dtree(v,w)=1d_{\mathrm{tree}}(v,w)=1 and XD¯X\in\bar{D}, we have

(8) ξv(θwX)=ξv(θw)=w𝚵v.\xi^{v}(\theta^{w}_{X})=\xi^{v}(\theta^{w})=w\in{\bf\Xi}^{v}.

Mirroring the notation dYv(x,y)d_{Y}^{v}(x,y) above, for x,yΣ¯x,y\in\bar{\Sigma} we define

d𝒦v(x,y)=diam(Λv(x)Λv(y))andd𝚵v(x,y)=diam(ξv(x)ξv(y)).d_{\mathcal{K}^{v}}(x,y)=\mathrm{diam}(\Lambda^{v}(x)\cup\Lambda^{v}(y))\quad\text{and}\quad d_{{\bf\Xi}^{v}}(x,y)=\mathrm{diam}(\xi^{v}(x)\cup\xi^{v}(y)).
Lemma 3.18.

There is a function N:[0,)[0,)N^{\prime}\colon[0,\infty)\to[0,\infty) such that for all v𝒱v\in\mathcal{V}, XBvX\in\partial B_{v}, and x,yΣ¯x,y\in\bar{\Sigma} the quantities d𝒦v(x,y)d_{\mathcal{K}^{v}}(x,y) and d𝚵v(x,y)d_{{\bf\Xi}^{v}}(x,y) are at most N(dXv(x,y))N^{\prime}(d_{X}^{v}(x,y)).

Proof.

For YBvY\in\partial B_{v}, let us compare the images of some subset U𝚯vYU\subset\partial{\bf\Theta}^{v}_{Y} and

Smear(U)=ZBvfZ,Y(U)𝚯v\mathrm{Smear}(U)=\bigcup_{Z\in\partial B_{v}}f_{Z,Y}(U)\subset\partial{\bf\Theta}^{v}

under the maps λv,iv\lambda^{v},i^{v}. Since the boundary components of 𝚯v{\bf\Theta}^{v} are preserved by the maps fZ,Yf_{Z,Y}, the images iv(U)=iv(Smear(U))i^{v}(U)=i^{v}(\mathrm{Smear}(U)) are exactly the same. Moreover, for each x𝚯vYx\in\partial{\bf\Theta}^{v}_{Y} we have λv(Smear({x}))=λv(x,α(v))\lambda^{v}(\mathrm{Smear}(\{x\}))=\lambda^{v}(\ell_{x,\alpha(v)}), and so by Proposition 3.1(2) this is a set with diameter at most K1K_{1}. Therefore λv(U)\lambda^{v}(U) and λv(Smear(U))\lambda^{v}(\mathrm{Smear}(U)) have Hausdorff distance at most 2K12K_{1}.

Now, let x,yΣ¯x,y\in\bar{\Sigma} and XBvX\in\partial B_{v} be as in the statement. Set U=ΠvX(x)ΠvX(y)U=\Pi^{v}_{X}(x)\cup\Pi^{v}_{X}(y) so that diam(U)=dvX(x,y)\mathrm{diam}(U)=d^{v}_{X}(x,y). Since Πv(x)=Smear(ΠvX(x))\Pi^{v}(x)=\mathrm{Smear}(\Pi^{v}_{X}(x)) and similarly for Πv(y)\Pi^{v}(y), we see that

d𝒦v(x,y)=diam(λv(Smear(U)))andd𝚵v(x,y)=diam(iv(Smear(U))).d_{\mathcal{K}^{v}}(x,y)=\mathrm{diam}(\lambda^{v}(\mathrm{Smear}(U)))\quad\text{and}\quad d_{{\bf\Xi}^{v}}(x,y)=\mathrm{diam}(i^{v}(\mathrm{Smear}(U))).

Since λv\lambda^{v} is coarsely K1K_{1}–Lipschitz by Proposition 3.11, the preceding paragraph and the triangle inequality shows that

d𝒦v(x,y)=diam(λv(Smear(U)))diam(λv(U))+2K1(K1dXv(x,y)+K1)+2K1.d_{\mathcal{K}^{v}}(x,y)=\mathrm{diam}(\lambda^{v}(\mathrm{Smear}(U)))\leq\mathrm{diam}(\lambda^{v}(U))+2K_{1}\leq(K_{1}d_{X}^{v}(x,y)+K_{1})+2K_{1}.

Similarly, by Corollary 3.9 we have that

diam𝚵v(x,y)=diam(iv(U))N(dXv(x,y))+1.\mathrm{diam}_{{\bf\Xi}^{v}}(x,y)=\mathrm{diam}(i^{v}(U))\leq N(d_{X}^{v}(x,y))+1.

Setting N(t)=max{N(t)+1,K1t+3K1}N^{\prime}(t)=\max\{N(t)+1,K_{1}t+3K_{1}\} completes the proof. ∎

Proposition 3.19.

There exists K4>0K_{4}>0 so that for any v𝒱v\in\mathcal{V}:

  1. (1)

    Λv\Lambda^{v} and ξv\xi^{v} are K4K_{4}–coarsely Lipschitz;

  2. (2)

    For any w𝒱w\in\mathcal{V}, we have

    1. (a)

      diam(Λv(𝚯w))<K4\mathrm{diam}\left(\Lambda^{v}({\bf\Theta}^{w})\right)<K_{4}, unless dtree(v,w)1d_{\mathrm{tree}}(v,w)\leq 1;

    2. (b)

      diam(ξv(𝚯w))<K4\mathrm{diam}\left(\xi^{v}({\bf\Theta}^{w})\right)<K_{4} unless w=vw=v.

Proof.

For part (1), we first observe that by [DDLS21, Lemma 3.5], there exists an R>0R>0 so that any pair of points x,yΣ¯x,y\in\bar{\Sigma} may be connected by a path of length at most Rd¯(x,y)R\bar{d}(x,y) that is a concatenation of at most Rd¯(x,y)+1R\bar{d}(x,y)+1 pieces, each of which is either a saddle connection of length at most RR in a vertical vertical fiber, or a horizontal geodesic segment in E¯\bar{E}. By the triangle inequality, it thus suffices to assume that xx and yy are the endpoints of either a horizontal geodesic or a vertical saddle connection of length at most RR. Appealing to Lemma 3.18, it further suffices to show that dXv(x,y)d_{X}^{v}(x,y) is linearly bounded by d¯(x,y)\bar{d}(x,y) for some XBvX\in\partial B_{v}. Lemmas 3.11 and 3.12(1) handle the horizontal segment case, since we are free to subdivide such a path into d¯(x,y)\lceil\bar{d}(x,y)\rceil segments of length at most 11.

For the vertical segment case we assume xx and yy lie in the same fiber EYE_{Y} and differ by a saddle connection δ\delta of length at most RR. Let θwY\theta^{w}_{Y}, where w𝒱w\in\mathcal{V}, be the spine containing δ\delta. The fact that δ\delta is bounded means that Y=π(x)=π(y)Y=\pi(x)=\pi(y) is bounded distance from the horocycle Bw\partial B_{w}. Let X=cw(Y)BwX=c_{w}(Y)\in\partial B_{w} and let δ=fX,Y(δ)\delta^{\prime}=f_{X,Y}(\delta) be the saddle connection in θwX\theta^{w}_{X} connecting x=fX,Y(x)x^{\prime}=f_{X,Y}(x) and y=fX,Y(y)y^{\prime}=f_{X,Y}(y). By the triangle inequality and the first part above about bounded length horizontal segments, it suffices to work with the points x,yθwx^{\prime},y^{\prime}\in\theta^{w}. There are three cases to consider: Firstly, if v=wv=w, then x,yθvXx^{\prime},y^{\prime}\in\theta^{v}_{X} so that ΠvX(x)\Pi^{v}_{X}(x^{\prime}) and ΠvX(y)\Pi^{v}_{X}(y^{\prime}) choose the closest cone points in 𝚯vX\partial{\bf\Theta}^{v}_{X} to xx^{\prime} and yy^{\prime}, respectively. Since xx^{\prime} and yy^{\prime} are close, so are ΠvX(x)\Pi^{v}_{X}(x^{\prime}) and ΠvX(y)\Pi^{v}_{X}(y^{\prime}). Secondly, if dtree(v,w)>1d_{\mathrm{tree}}(v,w)>1, then Corollaries 3.15 and 3.16, and Lemma 3.17 give a uniform bound on dvZ(x,y)diam(ΠvZ(θw))d^{v}_{Z}(x^{\prime},y^{\prime})\leq\mathrm{diam}(\Pi^{v}_{Z}(\theta^{w})) for any point Zcv(Bw)Z\in c_{v}(\partial B_{w}). Finally, if dtree(v,w)=1d_{\mathrm{tree}}(v,w)=1 then θwX\theta^{w}_{X} and θvX\theta^{v}_{X} are adjacent non-crossing spines in EXE_{X}. Since θwX\theta^{w}_{X} is totally geodesic, if follows that ΠvX(x)\Pi^{v}_{X}(x^{\prime}) and ΠvX(y)\Pi^{v}_{X}(y^{\prime}) are either equal or connected by a single edge of 𝚯vX\partial{\bf\Theta}^{v}_{X}. But this saddle connection has uniformly bounded length, since XBvX\in\partial B_{v}, which completes the proof of (1).

For (2), first recall that strips/saddle connections in the α(w)\alpha(w) direction have uniformly bounded width/length over Bw\partial B_{w} (Lemma 2.3). Therefore 𝚯wΣ¯{\bf\Theta}^{w}\cap\bar{\Sigma} is contained in a bounded neighborhood of θwΣ¯\theta^{w}\cap\bar{\Sigma}. By part (1) it thus suffices to bound diam(Λv(θw))\mathrm{diam}(\Lambda^{v}(\theta^{w})) and diam(ξv(θw))\mathrm{diam}(\xi^{v}(\theta^{w})). When dtree(v,w)2d_{\mathrm{tree}}(v,w)\geq 2, Corollaries 3.15 and 3.16, and Lemma 3.17, imply that there exists XBvX\in\partial B_{v} so that ΠvX(θw)\Pi^{v}_{X}(\theta^{w}) has bounded diameter in 𝚯vX{\bf\Theta}^{v}_{X}. Appealing to Lemma 3.18 now bounds diam(Λv(θw))\mathrm{diam}(\Lambda^{v}(\theta^{w})) and diam(ξv(θw))\mathrm{diam}(\xi^{v}(\theta^{w})) in these cases. For the remaining case dtree(v,w)=1d_{\mathrm{tree}}(v,w)=1 of 2(b), we note that ξv(θw)\xi^{v}(\theta^{w}) is a single point by (8), and thus 2(b) follows. ∎

4. Hierarchical hyperbolicity of Γ\Gamma

In this section we complete the proof that Γ\Gamma is hierarchically hyperbolic. We will use a criterion from [BHMS20], which we now briefly discuss. For further information and heuristic discussion of this approach to hierarchical hyperbolicity, we refer the reader to [BHMS20, §1.5, “User’s guide and a simple example”].

Consider a simplicial complex 𝒳{\mathcal{X}} and a graph 𝒲{\mathcal{W}} whose vertex set is the set of maximal simplices of 𝒳{\mathcal{X}}. The pair (𝒳,𝒲)({\mathcal{X}},{\mathcal{W}}) is called a combinatorial HHS if it satisfies the requirements listed in Definition 4.8 below, and [BHMS20, Theorem 1.18] guarantees that in this case 𝒲{\mathcal{W}} is an HHS. The main requirement is along the lines of: 𝒳{\mathcal{X}} is hyperbolic, and links of simplices of 𝒳{\mathcal{X}} are also hyperbolic. However, this is rarely the case because co-dimension–1 faces of maximal simplices have discrete links. To rectify this, additional edges (coming from 𝒲{\mathcal{W}}) should be added to 𝒳{\mathcal{X}} and its links as detailed in Definition 4.2. In our case, after adding these edges, 𝒳{\mathcal{X}} will be quasi-isometric to E^\hat{E}, and each other link will be quasi-isometric to either a point or to one of the spaces 𝒦v\mathcal{K}^{v} or 𝚵v{\bf\Xi}^{v} introduced in §3.

There are two natural situations where such pairs arise that the reader might want to keep in mind. First, consider a group HH acting on a simplicial complex 𝒳{\mathcal{X}} so that there is one orbit of maximal simplices, and those have trivial stabilizers. In this case, we take 𝒲{\mathcal{W}} to be (a graph isomorphic to) a Cayley graph of HH. (More generally, if the action is cocompact with finite stabilizers of maximal simplices, then the appropriate 𝒲{\mathcal{W}} is quasi-isometric to a Cayley graph.) For the second situation, 𝒳{\mathcal{X}} is the curve graph of a surface; then maximal simplices are pants decompositions of the surface and 𝒲{\mathcal{W}} can be taken to be the pants graph. We will use this as a working example below, when we get into the details.

Most of the work carried out in §3 will be used (as a black-box) to prove that, roughly, links are quasi-isometrically embedded in a space obtained by removing all the “obvious” vertices that provide shortcuts between vertices of the link. This can be seen as an analogue of Bowditch’s fineness condition in the context of relative hyperbolicity.

This section is organized as follows. In §4.1 we list all the relevant definitions and results from [BHMS20], and we illustrate them using pants graphs. In §4.2 we construct the relevant combinatorial HHS for our purposes. In §4.3 we analyze all the various links and related combinatorial objects; we note that most of the work done in §3 is used here to prove Lemma 4.22. At that point, essentially only one property of combinatorial HHSs will be left to be checked, and we do so in §4.4.

4.1. Basic definitions

We start by recalling some basic combinatorial definitions and constructions. Let 𝒳{\mathcal{X}} be a flag simplicial complex.

Definition 4.1 (Join, link, star).

Given disjoint simplices Δ,Δ\Delta,\Delta^{\prime} of 𝒳{\mathcal{X}}, the join is denoted ΔΔ\Delta\star\Delta^{\prime} and is the simplex spanned by Δ(0)Δ(0)\Delta^{(0)}\cup\Delta^{\prime(0)}, if it exists. More generally, if K,LK,L are disjoint induced subcomplexes of 𝒳{\mathcal{X}} such that every vertex of KK is adjacent to every vertex of LL, then the join KLK\star L is the induced subcomplex with vertex set K(0)L(0)K^{(0)}\cup L^{(0)}.

For each simplex Δ\Delta, the link Lk(Δ)\mathrm{Lk}(\Delta) is the union of all simplices Δ\Delta^{\prime} of 𝒳{\mathcal{X}} such that ΔΔ=\Delta^{\prime}\cap\Delta=\emptyset and ΔΔ\Delta^{\prime}\star\Delta is a simplex of 𝒳{\mathcal{X}}. The star of Δ\Delta is Star(Δ)=Lk(Δ)ΔStar(\Delta)=\mathrm{Lk}(\Delta)\star\Delta, i.e. the union of all simplices of 𝒳{\mathcal{X}} that contain Δ\Delta.

We emphasize that \emptyset is a simplex of 𝒳{\mathcal{X}}, whose link is all of 𝒳{\mathcal{X}} and whose star is all of 𝒳{\mathcal{X}}.

Definition 4.2 (XX–graph, WW–augmented dual complex).

An 𝒳{\mathcal{X}}–graph is any graph 𝒲{\mathcal{W}} whose vertex set is the set of maximal simplices of 𝒳{\mathcal{X}} (those not contained in any larger simplex).

For a flag complex 𝒳{\mathcal{X}} and an 𝒳{\mathcal{X}}–graph 𝒲{\mathcal{W}}, the 𝒲{\mathcal{W}}–augmented dual graph 𝒳+𝒲{{\mathcal{X}}}^{+{{\mathcal{W}}}} is the graph defined as follows:

  • the 0–skeleton of 𝒳+𝒲{{\mathcal{X}}}^{+{{\mathcal{W}}}} is 𝒳(0){\mathcal{X}}^{(0)};

  • if v,w𝒳(0)v,w\in{\mathcal{X}}^{(0)} are adjacent in 𝒳{\mathcal{X}}, then they are adjacent in 𝒳+𝒲{{\mathcal{X}}}^{+{{\mathcal{W}}}};

  • if two vertices in 𝒲{\mathcal{W}} are adjacent, then we consider σ,ρ\sigma,\rho, the associated maximal simplices of 𝒳{\mathcal{X}}, and in 𝒳+𝒲{{\mathcal{X}}}^{+{{\mathcal{W}}}} we connect each vertex of σ\sigma to each vertex of ρ\rho.

We equip 𝒲{\mathcal{W}} with the usual path-metric, in which each edge has unit length, and do the same for 𝒳+𝒲{{\mathcal{X}}}^{+{{\mathcal{W}}}}. Observe that the 11–skeleton of XX is a subgraph 𝒳(1)𝒳+𝒲{\mathcal{X}}^{(1)}\subset{{\mathcal{X}}}^{+{{\mathcal{W}}}}.

We provide a running example to illustrate the various definitions in a familiar situation. This example will not be used in the sequel.

Example 4.3.

If 𝒳{\mathcal{X}} is the curve complex of the surface SS, then an example of the an 𝒳{\mathcal{X}}–graph, 𝒲{\mathcal{W}}, is the pants graph, since a maximal simplex is precisely a pants decomposition. The 𝒲{\mathcal{W}}–augmented dual graph can be thought of as adding to the curve graph, 𝒳(0){\mathcal{X}}^{(0)} an edge between any two curves that fill a one-holed torus or four-holed sphere and intersect once or twice, respectively: indeed, these subsurfaces are precisely those where an elementary move happens as in the definition of adjacency in the pants graph.

Definition 4.4 (Equivalent simplices, saturation).

For Δ,Δ\Delta,\Delta^{\prime} simplices of 𝒳{\mathcal{X}}, we write ΔΔ\Delta\sim\Delta^{\prime} to mean Lk(Δ)=Lk(Δ)\mathrm{Lk}(\Delta)=\mathrm{Lk}(\Delta^{\prime}). We denote by [Δ][\Delta] the equivalence class of Δ\Delta. Let Sat(Δ)\mathrm{Sat}(\Delta) denote the set of vertices v𝒳v\in{\mathcal{X}} for which there exists a simplex Δ\Delta^{\prime} of 𝒳{\mathcal{X}} such that vΔv\in\Delta^{\prime} and ΔΔ\Delta^{\prime}\sim\Delta, i.e.

Sat(Δ)=(Δ[Δ]Δ)(0).\mathrm{Sat}(\Delta)=\left(\bigcup_{\Delta^{\prime}\in[\Delta]}\Delta^{\prime}\right)^{(0)}.

We denote by 𝔖\mathfrak{S} the set of \sim–classes of non-maximal simplices in 𝒳{\mathcal{X}}.

Definition 4.5 (Complement, link subgraph).

Let 𝒲{\mathcal{W}} be an 𝒳{\mathcal{X}}–graph. For each simplex Δ\Delta of 𝒳{\mathcal{X}}, let YΔY_{\Delta} be the subgraph of 𝒳+𝒲{{\mathcal{X}}}^{+{{\mathcal{W}}}} induced by the set of vertices (𝒳+𝒲)(0)Sat(Δ)({{\mathcal{X}}}^{+{{\mathcal{W}}}})^{(0)}-\mathrm{Sat}(\Delta).

Let 𝒞(Δ)\mathcal{C}(\Delta) be the full subgraph of YΔY_{\Delta} spanned by Lk(Δ)(0)\mathrm{Lk}(\Delta)^{(0)}. Note that 𝒞(Δ)=𝒞(Δ)\mathcal{C}(\Delta)=\mathcal{C}(\Delta^{\prime}) whenever ΔΔ\Delta\sim\Delta^{\prime}. (We emphasize that we are taking links in 𝒳{\mathcal{X}}, not in 𝒳+𝒲{\mathcal{X}}^{+{\mathcal{W}}}, and then considering the subgraphs of YΔY_{\Delta} induced by those links.)

We now pause and continue with the illustrative example.

Example 4.6.

Let 𝒳{\mathcal{X}} and 𝒲{\mathcal{W}} be as in Example 4.3. A simplex Δ\Delta is a multicurve which determines two (open) subsurfaces U=U(Δ),U=U(Δ)SU=U(\Delta),U^{\prime}=U^{\prime}(\Delta)\subset S, where UU is the union of the complementary components of the multicurve that are not a pair of pants, and U=SU¯U^{\prime}=S-\overline{U}. Note that UΔ\partial U\subset\Delta is a submulticurve and that ΔU\Delta-\partial U is a pants decomposition of UU^{\prime}. A simplex Δ\Delta^{\prime} is equivalent to Δ\Delta if it defines the same subsurfaces. Thus Sat(Δ)\mathrm{Sat}(\Delta) consists of U(Δ)\partial U(\Delta) together with all essential curves in U(Δ)U^{\prime}(\Delta), while 𝒞(Δ)\mathcal{C}(\Delta) is the join of graphs quasi-isometric to curve graphs of the components of U(Δ)U(\Delta). For components of U(Δ)U(\Delta) which are one-holed tori or four-holed spheres, the corresponding subgraphs are isometric to their curve graphs (since the extra edges in 𝒳+𝒲{{\mathcal{X}}}^{+{{\mathcal{W}}}} precisely give edges for these curve graphs).

Definition 4.7 (Nesting).

Let 𝒳{\mathcal{X}} be a simplicial complex. Let Δ,Δ\Delta,\Delta^{\prime} be non-maximal simplices of 𝒳{\mathcal{X}}. Then we write [Δ][Δ][\Delta]\sqsubseteq[\Delta^{\prime}] if Lk(Δ)Lk(Δ)\mathrm{Lk}(\Delta)\subseteq\mathrm{Lk}(\Delta^{\prime}).

We note that if ΔΔ\Delta^{\prime}\subset\Delta, then [Δ][Δ][\Delta]\sqsubseteq[\Delta^{\prime}]. Also, for Example 4.3,4.6, [Δ][Δ][\Delta]\sqsubseteq[\Delta^{\prime}] if and only if U(Δ)U(Δ)U(\Delta)\subset U(\Delta^{\prime}).

Finally, we are ready for the main definition:

Definition 4.8 (Combinatorial HHS).

A combinatorial HHS (𝒳,𝒲)({\mathcal{X}},{\mathcal{W}}) consists of a flag simplicial complex 𝒳{\mathcal{X}} and an 𝒳{\mathcal{X}}–graph 𝒲{\mathcal{W}} satisfying the following conditions for some nn\in\mathbb{N} and δ1\delta\geq 1:

  1. (1)

    any chain [Δ1][Δ2][\Delta_{1}]\sqsubsetneq[\Delta_{2}]\sqsubsetneq\dots has length at most nn;

  2. (2)

    for each non-maximal simplex Δ\Delta, the subgraph 𝒞(Δ)\mathcal{C}(\Delta) is δ\delta–hyperbolic and (δ,δ)(\delta,\delta)–quasi-isometrically embedded in YΔY_{\Delta};

  3. (3)

    Whenever Δ\Delta and Δ\Delta^{\prime} are non-maximal simplices for which there exists a non-maximal simplex Γ\Gamma such that [Γ][Δ][\Gamma]\sqsubseteq[\Delta], [Γ][Δ][\Gamma]\sqsubseteq[\Delta^{\prime}], and diam(𝒞(Γ))δ\mathrm{diam}(\mathcal{C}(\Gamma))\geq\delta, then there exists a simplex Π\Pi in the link of Δ\Delta^{\prime} such that [ΔΠ][Δ][\Delta^{\prime}\star\Pi]\sqsubseteq[\Delta] and all [Γ][\Gamma] as above satisfy [Γ][ΔΠ][\Gamma]\sqsubseteq[\Delta^{\prime}\star\Pi];

  4. (4)

    if v,wv,w are distinct non-adjacent vertices of Lk(Δ)\mathrm{Lk}(\Delta), for some simplex Δ\Delta of 𝒳{\mathcal{X}}, contained in 𝒲{\mathcal{W}}–adjacent maximal simplices, then they are contained in 𝒲{\mathcal{W}}–adjacent simplices of the form ΔΔ\Delta\star\Delta^{\prime}.

We will see below that combinatorial HHSs give HHSs. The reader not interested in the explicit description of the HHS structure can skip the following two definitions.

Definition 4.9 (Orthogonality, transversality).

Let 𝒳{\mathcal{X}} be a simplicial complex. Let Δ,Δ\Delta,\Delta^{\prime} be non-maximal simplices of 𝒳{\mathcal{X}}. Then we write [Δ][Δ][\Delta]\bot[\Delta^{\prime}] if Lk(Δ)Lk(Lk(Δ))\mathrm{Lk}(\Delta^{\prime})\subseteq\mathrm{Lk}(\mathrm{Lk}(\Delta)). If [Δ][\Delta] and [Δ][\Delta^{\prime}] are not \bot–related or \sqsubseteq–related, we write [Δ][Δ][\Delta]\pitchfork[\Delta^{\prime}].

Definition 4.10 (Projections).

Let (𝒳,𝒲)(\mathcal{X},\mathcal{W}) be a combinatorial HHS.

Fix [Δ]𝔖[\Delta]\in\mathfrak{S} and define a map π[Δ]:𝒲𝐏(𝒞([Δ]))\pi_{[\Delta]}\colon{\mathcal{W}}\to{\bf P}({\mathcal{C}([\Delta])}) as follows. First let p:YΔ𝐏(𝒞([Δ]))p\colon Y_{\Delta}\to{\bf P}({\mathcal{C}([\Delta])}) be the coarse closest-point projection, i.e.

p(x)={y𝒞([Δ])dYΔ(x,y)dYΔ(x,𝒞([Δ]))+1}.p(x)=\big{\{}y\in\mathcal{C}([\Delta])\mid d_{Y_{\Delta}}(x,y)\leq d_{Y_{\Delta}}(x,\mathcal{C}([\Delta]))+1\big{\}}.

Suppose that w𝒲(0)w\in{\mathcal{W}}^{(0)}, so ww corresponds to a unique simplex Δw\Delta_{w} of 𝒳{\mathcal{X}}. Define

π[Δ](w)=p(ΔwYΔ).\pi_{[\Delta]}(w)=p(\Delta_{w}\cap Y_{\Delta}).

We have thus defined π[Δ]:𝒲(0)𝐏(𝒞([Δ]))\pi_{[\Delta]}\colon{\mathcal{W}}^{(0)}\to{\bf P}({\mathcal{C}([\Delta])}). If v,w𝒲(0)v,w\in{\mathcal{W}}^{(0)} are joined by an edge ee of 𝒲{\mathcal{W}}, then Δv,Δw\Delta_{v},\Delta_{w} are joined by edges in 𝒳+𝒲{{\mathcal{X}}}^{+{{\mathcal{W}}}}, and we let π[Δ](e)=π[Δ](v)π[Δ](w)\pi_{[\Delta]}(e)=\pi_{[\Delta]}(v)\cup\pi_{[\Delta]}(w).

Now let [Δ],[Δ]𝔖[\Delta],[\Delta^{\prime}]\in\mathfrak{S} satisfy [Δ][Δ][\Delta]\pitchfork[\Delta^{\prime}] or [Δ][Δ][\Delta^{\prime}]\sqsubsetneq[\Delta]. Let

ρ[Δ][Δ]=p(Sat(Δ)YΔ).\rho^{[\Delta^{\prime}]}_{[\Delta]}=p(\mathrm{Sat}(\Delta^{\prime})\cap Y_{\Delta}).

Let [Δ][Δ][\Delta]\sqsubsetneq[\Delta^{\prime}]. Let ρ[Δ][Δ]:𝒞([Δ])𝒞([Δ])\rho^{[\Delta^{\prime}]}_{[\Delta]}\colon\mathcal{C}([\Delta^{\prime}])\to\mathcal{C}([\Delta]) be the restriction of pp to 𝒞([Δ])YΔ\mathcal{C}([\Delta^{\prime}])\cap Y_{\Delta}, and \emptyset otherwise.

The next theorem from [BHMS20] provides the criteria we will use to prove that Γ\Gamma is a hierarchically hyperbolic group.

Given a combinatorial HHS (𝒳,𝒲)({\mathcal{X}},{\mathcal{W}}), we denote 𝔖\mathfrak{S} the set as in Definition 4.4, endowed with nesting and orthogonality relations as in Definitions 4.7 and 4.9. Also, we associated to 𝔖\mathfrak{S} the hyperbolic spaces as in Definition 4.8, and define projections as in Definition 4.10.

Theorem 4.11.

[BHMS20, Theorem 1.18, Remark 1.19] Let (𝒳,𝒲)({\mathcal{X}},{\mathcal{W}}) be a combinatorial HHS. Then (𝒲,𝔖)({\mathcal{W}},\mathfrak{S}) is a hierarchically hyperbolic space.

Moreover, if a group GG acts by simplicial automorphisms on 𝒳{\mathcal{X}} with finitely many orbits of links of simplices, and the resulting GG-action on maximal simplices extends to a metrically proper cobounded action on 𝒲{\mathcal{W}}, then GG acts metrically properly and coboundedly by HHS automorphisms on (𝒲,𝔖)({\mathcal{W}},\mathfrak{S}), and is therefore a hierarchically hyperbolic group.

4.2. Combinatorial HHS structure

We now define a flag simplicial complex 𝒳{\mathcal{X}}. The vertex set is 𝒳(0)=𝒱𝒦{\mathcal{X}}^{(0)}=\mathcal{V}\sqcup{\mathcal{K}}, where

𝒦=v𝒱𝒦v.{\mathcal{K}}=\bigsqcup_{v\in\mathcal{V}}\mathcal{K}^{v}.

Given a vertex s𝒦s\in{\mathcal{K}}, let v(s)𝒱v(s)\in\mathcal{V} be the unique vertex with s𝒦vs\in{\mathcal{K}}^{v}. We also write α(s)=α(v(s))\alpha(s)=\alpha(v(s)).

There are 3 types of edges (see Figure 6):

  1. (1)

    v,w𝒱v,w\in\mathcal{V} are connected by an edge if and only if dtree(v,w)=1d_{\mathrm{tree}}(v,w)=1.

  2. (2)

    s,t𝒦s,t\in{\mathcal{K}} are connected by an edge if and only if dtree(v(s),v(t))=1d_{\mathrm{tree}}(v(s),v(t))=1.

  3. (3)

    s𝒦s\in{\mathcal{K}} and w𝒱w\in\mathcal{V} are connected by an edge if and only if dtree(v(s),w)1d_{\mathrm{tree}}(v(s),w)\leq 1.

We declare 𝒳{\mathcal{X}} to be the flag simplicial complex with the 1-skeleton defined above.

v1v_{1}t1t_{1}t2t_{2}t3t_{3}v2v_{2}s1s_{1}s2s_{2}s3s_{3}v3v_{3}t1t_{1}t2t_{2}t3t_{3}v4v_{4}t1t_{1}t2t_{2}t3t_{3}v1v_{1}v2v_{2}v3v_{3}v4v_{4}v(si)=v(s_{i})=v(ti)=v(t_{i})=ZZ
Figure 6. The simplicial map ZZ restricted to a part of 𝒳{\mathcal{X}} (on the left) to a part of the union of trees TαT_{\alpha} (on the right). Vertices in in 𝒳{\mathcal{X}} are colored the same as their image vertices in TαT_{\alpha}.

The map 𝒦𝒱{\mathcal{K}}\to\mathcal{V} given by sv(s)s\mapsto v(s) and the identity 𝒱𝒱\mathcal{V}\to\mathcal{V} extends to a surjective simplicial map

Z:𝒳α𝒫Tα.Z\colon{\mathcal{X}}\to\bigsqcup_{\alpha\in{\mathcal{P}}}T_{\alpha}.

We note that we may view the union Tα\bigsqcup T_{\alpha} on the right as a subgraph of 𝒳(1){\mathcal{X}}^{(1)}, making ZZ a retraction.

For any vertex vv in any tree TαT_{\alpha}, Z1(v)Z^{-1}(v) is the join of {v}\{v\} and the set 𝒦v\mathcal{K}^{v}:

(9) Z1(v)={v}𝒦v.Z^{-1}(v)=\{v\}\star\mathcal{K}^{v}.

For any pair of adjacent vertices v,wTαv,w\in T_{\alpha} (so dtree(v,w)=1d_{\mathrm{tree}}(v,w)=1), the preimage of the edge [v,w]Tα[v,w]\subset T_{\alpha} is also a join:

(10) Z1([v,w])=Z1(v)Z1(w)=({v}𝒦v)({w}𝒦w).Z^{-1}([v,w])=Z^{-1}(v)\star Z^{-1}(w)=(\{v\}\star\mathcal{K}^{v})\star(\{w\}\star\mathcal{K}^{w}).
Lemma 4.12.

The maximal simplices of 𝒳{\mathcal{X}} are exactly the 33–simplices with vertex set {s,v(s),t,v(t)}\{s,v(s),t,v(t)\} where s,t𝒦s,t\in{\mathcal{K}} and dtree(v(s),v(t))=1d_{\mathrm{tree}}(v(s),v(t))=1. In this case, we say that (s,t)(s,t) defines a maximal simplex, denoted σ(s,t)\sigma(s,t).

Proof.

Because the map ZZ is simplicial, any simplex of 𝒳{\mathcal{X}} is contained in Z1(v)Z^{-1}(v) or Z1([v,w])Z^{-1}([v,w]) for some vertex vv in some TαT_{\alpha} or some edge [v,w][v,w] in some TαT_{\alpha}. The lemma thus follows from (9) and (10). ∎

Given a vertex s𝒦s\in{\mathcal{K}}, recall from Lemma 3.5 that M(s)=(λv(s))1(NK1(s))𝚯v(s)M(s)=(\lambda^{v(s)})^{-1}(N_{K_{1}}(s))\subset{\bf\Theta}^{v(s)}, for K1K_{1} as in Proposition 3.1 (and Lemma 3.5). Given a pair of vertices (s,t)(s,t) in 𝒦{\mathcal{K}} that define a maximal simplex σ(s,t)\sigma(s,t), we will write M(s,t)=M(s)M(t)M(s,t)=M(s)\cap M(t).

Lemma 4.13.

There exists R>0R>0 with the following properties.

  1. (1)

    For any pair of adjacent vertices s,t𝒦s,t\in{\mathcal{K}} (i.e., defining a maximal simplex σ(s,t)\sigma(s,t)), M(s,t)M(s,t) is a non-empty subset of diameter at most RR.

  2. (2)

    Given v𝒱v\in\mathcal{V}, we have

    𝚯v=v(s)=v,dtree(v(s),v(t))=1M(s,t).{\bf\Theta}^{v}=\bigcup_{v(s)=v,d_{\mathrm{tree}}(v(s),v(t))=1}M(s,t).
  3. (3)

    Fixing ss, we have

    M(s)=dtree(v(s),v(t))=1M(s,t).M(s)=\bigcup_{d_{\mathrm{tree}}(v(s),v(t))=1}M(s,t).
  4. (4)

    The collection of all M(s,t)M(s,t) is RR–dense in E¯\bar{E}.

Proof.

Item (1) follows from Proposition 3.1(3). More precisely, the fact that M(s,t)M(s,t) is non-empty follows from K1K_{1}-coarse-surjectivity of λv(s)×λv(t)\lambda^{v(s)}\times\lambda^{v(t)}, while boundedness follows from the fact that said map is a quasi-isometry.

In order to show item (2), notice that

𝚯v=dtree(v,w)=1𝚯v𝚯w.{\bf\Theta}^{v}=\bigcup_{d_{\mathrm{tree}}(v,w)=1}{\bf\Theta}^{v}\cap{\bf\Theta}^{w}.

That is, every point of 𝚯v{\bf\Theta}^{v} is also in 𝚯w{\bf\Theta}^{w} for some ww adjacent to vv. In view of this, we conclude by noticing that if x𝚯v𝚯wx\in{\bf\Theta}^{v}\cap{\bf\Theta}^{w}, then xM(λv(x),λw(x))x\in M(\lambda^{v}(x),\lambda^{w}(x)). Item (3) follows similarly.

Finally, item (4) follows from item (2) and the fact that the collection of all 𝚯v{\bf\Theta}^{v} is coarsely dense in E¯\bar{E}. ∎

Next we define a graph 𝒲{\mathcal{W}} whose vertex set is the set of maximal simplices of 𝒳{\mathcal{X}}. We would like to just connect maximal simplices when the corresponding subsets M(s,t)M(s,t) are close in E¯\bar{E} (first bullet below); however, in order to arrange item (4) of the definition of combinatorial HHS (and only for that reason) we need different closeness constants for different situations. We fix RR as in Lemma 4.13, and moreover we require R>K12+K1R>K_{1}^{2}+K_{1}, for K1K_{1} as in Proposition 3.1 and Lemma 3.5.

Given maximal simplices σ(s1,t1)\sigma(s_{1},t_{1}) and σ(s2,t2)\sigma(s_{2},t_{2}), we declare them to be connected by an edge in 𝒲{\mathcal{W}} if one of the following holds:

(11) d¯(M(s1,t1),M(s2,t2))10Rs1=s2 and d¯(M(t1),M(t2))10R\begin{array}[]{l}\bullet\,\,\bar{d}(M(s_{1},t_{1}),M(s_{2},t_{2}))\leq 10R\\ \bullet\,\,s_{1}=s_{2}\mbox{ and }\bar{d}(M(t_{1}),M(t_{2}))\leq 10R\end{array}

Here the the d¯\bar{d}–distances are the infimal distances between the sets in E¯\bar{E} (as opposed to the diameter of the union). Note that since M(s,t)=M(t,s)M(s,t)=M(t,s), the second case also implicitly describes a “symmetric case” with sis_{i} and tit_{i} interchanged.

The following is immediate from Lemma 3.6, setting R=max{10R,N𝒦(10R)}R^{\prime}=\max\{10R,N_{\mathcal{K}}(10R)\}.

Lemma 4.14.

There exists R10RR^{\prime}\geq 10R so then the following holds. If s,t1,t2𝒦s,t_{1},t_{2}\in{\mathcal{K}} are vertices with ss connected to both tit_{i} in 𝒳{\mathcal{X}} and d¯(M(t1),M(t2))10R\bar{d}(M(t_{1}),M(t_{2}))\leq 10R, then d¯(M(s,t1),M(s,t2))R\bar{d}(M(s,t_{1}),M(s,t_{2}))\leq R^{\prime}. In particular, whenever σ(s1,t1)\sigma(s_{1},t_{1}) and σ(s2,t2)\sigma(s_{2},t_{2}) are connected in 𝒲{\mathcal{W}}, we have d¯(M(s1,t1),M(s2,t2))R\bar{d}(M(s_{1},t_{1}),M(s_{2},t_{2}))\leq R^{\prime}.

Lemma 4.15.

𝒲{\mathcal{W}} is quasi-isometric to E¯\bar{E}, by mapping each σ(s,t)\sigma(s,t) to (any point in) M(s,t)M(s,t). Moreover, the extension group Γ\Gamma acts by simplicial automorphisms on 𝒳\mathcal{X}, induced by the existing action on 𝒱𝒳(0)\mathcal{V}\subseteq\mathcal{X}^{(0)} and the action on 𝒦𝒳(0)\mathcal{K}\subseteq\mathcal{X}^{(0)} as in Proposition 3.1(4). The resulting action on maximal simplices extends to a metrically proper cobounded action on 𝒲{\mathcal{W}}.

Proof.

In view of Lemma 4.14, the first part follows by combining Lemma 4.13(4) and Proposition 2.4 (applied to any choice of a point in each M(s,t)M(s,t)).

It is immediate to check that the Γ\Gamma-action defined on the 0-skeleton of 𝒳{\mathcal{X}} extends to an action on 𝒳{\mathcal{X}}. That the resulting action on maximal simplices of 𝒳{\mathcal{X}} (that is, the 0-skeleton of 𝒲{\mathcal{W}}) extends to an action on 𝒲{\mathcal{W}} follows from the equivariance property in Proposition 3.1(4) and the definitions of the sets M(s)M(s) and M(s,t)M(s,t).

Moreover, the quasi-isometry 𝒲E¯{\mathcal{W}}\to\bar{E} described in the statement is Γ\Gamma-equivariant, so that the action of Γ\Gamma on 𝒲{\mathcal{W}} is metrically proper and cobounded since the action of Γ\Gamma on E¯\bar{E} has these properties. ∎

The goal for the remainder of this section is to prove the following.

Theorem 4.16.

The pair (𝒳,𝒲)({\mathcal{X}},{\mathcal{W}}) is a combinatorial HHS. Moreover, there is an action of Γ\Gamma on 𝒳{\mathcal{X}} satisfying the properties stated in Theorem 4.11.

4.3. Simplices, links, and saturations

Before giving the proof of Theorem 4.16, we begin by describing explicitly the kinds of simplices of 𝒳{\mathcal{X}} that there are, explain what their links and saturations are, and observe some useful properties.

Lemma 4.17 (Empty simplex).

For the empty simplex, 𝒞()=𝒳+𝒲\mathcal{C}(\emptyset)={\mathcal{X}}^{+{\mathcal{W}}} is quasi-isometric to E^\hat{E}.

Proof.

We define a map Z:𝒳+𝒲E^Z^{\prime}\colon{\mathcal{X}}^{+{\mathcal{W}}}\to\hat{E} that extends the (restricted) simplicial map Z:𝒳(1)TαZ\colon{\mathcal{X}}^{(1)}\to\bigsqcup T_{\alpha} already constructed above. To do that, we must extend over each edge e=[x,y]e=[x,y] of 𝒳+𝒲{\mathcal{X}}^{+{\mathcal{W}}} coming from the edge of 𝒲{\mathcal{W}} connecting σ(s1,t1)\sigma(s_{1},t_{1}) and σ(s2,t2)\sigma(s_{2},t_{2}). Since Z(x),Z(y)𝒱Z(x),Z(y)\in\mathcal{V}, and d¯(M(s1,t1),M(s2,t2))R\bar{d}(M(s_{1},t_{1}),M(s_{2},t_{2}))\leq R^{\prime} (for RR^{\prime} as in Lemma 4.14), we see that dE^(v,w)Rd_{\hat{E}}(v,w)\leq R^{\prime}. We can then define ZZ^{\prime} on ee to be a constant speed parameterization of a uniformly bounded length path from Z(x)Z(x) to Z(y)Z(y). It follows that ZZ^{\prime} is Lipschitz.

The union of the trees Tα\bigsqcup T_{\alpha} is R0R_{0}–dense for some R0>0R_{0}>0 by [DDLS21, Lemma 3.6], so it suffices to find a one-sided inverse to ZZ^{\prime}, from Tα\bigsqcup T_{\alpha} to 𝒳+𝒲{\mathcal{X}}^{+{\mathcal{W}}}, and show that with respect to the subspace metric from E^\hat{E}, it is coarsely Lipschitz. As already noted, ZZ restricts to a retraction of 𝒳(1){\mathcal{X}}^{(1)} onto Tα𝒳(1)𝒳+𝒲\bigsqcup T_{\alpha}\subset{\mathcal{X}}^{(1)}\subset{\mathcal{X}}^{+{\mathcal{W}}}, which is thus the required one-sided inverse. All that remains is to show that it is coarsely Lipschitz.

According to [DDLS21, Lemma 3.8], any vTα,wTβv\in T_{\alpha},w\in T_{\beta} are connected by a combinatorial path of length comparable to d^(v,w)\hat{d}(v,w). Such a path is the concatenation of horizontal jumps, each of which is the P¯\bar{P}–image in E^\hat{E} of a geodesic in D¯z\bar{D}_{z}, for some zΣ¯z\in\bar{\Sigma}, that connects two components of D¯z\partial\bar{D}_{z} and whose interior is disjoint from D¯z\partial\bar{D}_{z}. From that same lemma, we may assume each horizontal jump has length uniformly bounded above and below, and thus has total number of jumps bounded in terms of d^(v,w)\hat{d}(v,w). Therefore, we can reduce to the case that v,wv,w are joined by a single horizontal jump of bounded length. Such a horizontal jump can also be regarded as a path in E¯\bar{E} connecting 𝚯v{\bf\Theta}^{v} to 𝚯w{\bf\Theta}^{w}. Hence, in view of Lemma 4.13(2), there are M(s1,t1)𝚯vM(s_{1},t_{1})\subseteq{\bf\Theta}^{v} and M(s2,t2)𝚯wM(s_{2},t_{2})\subseteq{\bf\Theta}^{w} within uniformly bounded distance of each other in E¯\bar{E}. Lemma 4.15 implies that there exists a path in 𝒲{\mathcal{W}} of uniformly bounded length from σ(s1,t1)\sigma(s_{1},t_{1}) to σ(s2,t2)\sigma(s_{2},t_{2}), which can be easily turned into a path of uniformly bounded length from vv to ww in 𝒳+𝒲{\mathcal{X}}^{+{\mathcal{W}}}, as required. ∎

There is an important type of 11–dimensional simplex, which we call a 𝚵{\bf\Xi}–type simplex, due to the following lemma. See Figure 7. Given w𝒱w\in\mathcal{V}, set

(12) Lk𝚵(w)=dtree(u,w)=1Z1(u)=dtree(u,w)=1{u}𝒦u.\mathrm{Lk}_{{\bf\Xi}}(w)=\bigcup_{d_{\mathrm{tree}}(u,w)=1}Z^{-1}(u)=\bigcup_{d_{\mathrm{tree}}(u,w)=1}\{u\}\star{\mathcal{K}}^{u}.
Lemma 4.18 (𝚵{\bf\Xi}–type simplex).

Let Δ\Delta be a 11–simplex of 𝒳{\mathcal{X}} with vertices s,v(s)s,v(s), for s𝒦s\in{\mathcal{K}}. Then

Lk(Δ)=Lk𝚵(v(s)) and Sat(Δ)={v(s)}𝒦v(s).\mathrm{Lk}(\Delta)=\mathrm{Lk}_{{\bf\Xi}}(v(s))\quad\mbox{ and }\quad\mathrm{Sat}(\Delta)=\{v(s)\}\cup{\mathcal{K}}^{v(s)}.

Moreover, 𝒞(Δ)\mathcal{C}(\Delta) is quasi-isometric to 𝚵v(s){\bf\Xi}^{v(s)}, via a quasi-isometry which is the identity on 𝒱𝒞(Δ)\mathcal{V}\cap\mathcal{C}(\Delta) and maps tt to v(t)v(t) for t𝒦𝒞(Δ)t\in{\mathcal{K}}\cap\mathcal{C}(\Delta).

v1v_{1}t1t_{1}t2t_{2}t3t_{3}v\,\,v\,\,s1s_{1}s\,\,s\,\,s3s_{3}v3v_{3}t1t_{1}t2t_{2}t3t_{3}v4v_{4}t1t_{1}t2t_{2}t3t_{3}....v\,\,v\,\,s1s_{1}s2s_{2}s3s_{3}........Sat(Δ)\mathrm{Sat}(\Delta)v1v_{1}t1t_{1}t2t_{2}t3t_{3}....v3v_{3}t1t_{1}t2t_{2}t3t_{3}v4v_{4}t1t_{1}t2t_{2}t3t_{3}Lk(Δ)=Lk𝚵(v)\mathrm{Lk}(\Delta)=\mathrm{Lk}_{{\bf\Xi}}(v)
Figure 7. 𝚵{\bf\Xi}–type simplex: Part of the link and saturation of a 11–simplex Δ\Delta with vertices s,v=v(s)s,v=v(s). Again, vertices of a common color correspond to the same element of 𝒱\mathcal{V}.
Proof.

It is clear from the definitions that the link of Δ\Delta is as described. Also, any simplex with vertex set of the form {v(s),t}\{v(s),t\} for some t𝒦v(s)t\in{\mathcal{K}}^{v(s)} has the same link as Δ\Delta. Therefore, to prove that the saturation is as described we are left to show if a simplex Δ\Delta^{\prime} has the same link as Δ\Delta, then its vertex set is contained in the set we described. If w𝒱w\in\mathcal{V} is a vertex of Δ\Delta^{\prime}, then dtree(v,w)=1d_{\mathrm{tree}}(v,w)=1 for all neighbors vv of v(s)v(s) in Tα(v)T_{\alpha(v)}. This implies w=v(s)w=v(s). Similarly, if t𝒦t\in{\mathcal{K}} is a vertex of Δ\Delta^{\prime}, then v(t)=v(s)v(t)=v(s), and we are done.

Let us show that the map given in the statement is coarsely Lipschitz. To do so, it suffices to consider www\neq w^{\prime} with dtree(w,v(s))=dtree(w,v(s))=1d_{\mathrm{tree}}(w,v(s))=d_{\mathrm{tree}}(w^{\prime},v(s))=1 and connected by an edge in 𝒳+𝒲{\mathcal{X}}^{+{\mathcal{W}}}, and show that they are connected by a bounded-length path in 𝚵v{\bf\Xi}^{v}. We argue below that d¯(𝚯w𝚯v(s),𝚯w𝚯v(s))R\bar{d}({\bf\Theta}^{w}\cap{\bf\Theta}^{v(s)},{\bf\Theta}^{w^{\prime}}\cap{\bf\Theta}^{v(s)})\leq R^{\prime}. Once we do that, the existence of the required bounded-length path follows directly from Lemma 3.8.

Let us now prove the desired inequality. Notice that w,ww,w^{\prime} cannot be connected by an edge of 𝒳{\mathcal{X}} since they are both distance 1 from v(s)v(s) in the tree Tα(v)T_{\alpha(v)}. Hence, ww and ww^{\prime} are contained respectively in maximal simplices σ(t,u)\sigma(t,u) and σ(t,u)\sigma(t^{\prime},u^{\prime}) connected by an edge in 𝒲\mathcal{W}. Say, up to swapping tt with uu and/or tt^{\prime} with uu^{\prime}, that v(t)=wv(t)=w and v(t)=wv(t^{\prime})=w^{\prime}. In either case of the definition of the edges of 𝒲{\mathcal{W}} we have d¯(M(t),M(t))10R\bar{d}(M(t),M(t^{\prime}))\leq 10R. Since M(t,s)𝚯w𝚯v(s)M(t,s)\subseteq{\bf\Theta}^{w}\cap{\bf\Theta}^{v(s)} and M(t,s)𝚯w𝚯v(s)M(t^{\prime},s)\subseteq{\bf\Theta}^{w^{\prime}}\cap{\bf\Theta}^{v(s)}, using Lemma 4.14 we get

d¯(𝚯w𝚯v(s),𝚯w𝚯v(s))d¯(M(t,s),M(t,s))R,\bar{d}({\bf\Theta}^{w}\cap{\bf\Theta}^{v(s)},{\bf\Theta}^{w^{\prime}}\cap{\bf\Theta}^{v(s)})\leq\bar{d}(M(t,s),M(t^{\prime},s))\leq R^{\prime},

as we wanted.

Conversely, if w,ww,w^{\prime} as above are joined by an edge in 𝚵v(s){\bf\Xi}^{v(s)} we will now show that they are also connected by an edge in 𝒞(Δ)\mathcal{C}(\Delta). By definition of 𝚵v(s){\bf\Xi}^{v(s)}, we have 𝚯w𝚯w{\bf\Theta}^{w}\cap{\bf\Theta}^{w^{\prime}}\neq\emptyset. By Lemma 4.13(2), There exists t,u,t,ut,u,t^{\prime},u^{\prime} with v(t)=wv(t)=w and v(t)=wv(t^{\prime})=w^{\prime}, so that M(t,u)M(t,u)M(t,u)\cap M(t^{\prime},u^{\prime})\neq\emptyset. In particular, d¯(M(t,u),M(t,u))10R\bar{d}(M(t,u),M(t^{\prime},u^{\prime}))\leq 10R. This says that σ(t,u)\sigma(t,u) and σ(t,u)\sigma(t^{\prime},u^{\prime}) are connected in 𝒲{\mathcal{W}}, and hence that ww and ww^{\prime} are connected in 𝒳+𝒲{\mathcal{X}}^{+{\mathcal{W}}}, as required. ∎

There is also an important type of 22–dimensional simplex, which we call a 𝒦\mathcal{K}–type simplex, due to the next lemma. See Figure 8.

Lemma 4.19 (𝒦\mathcal{K}–type simplex).

Let Δ\Delta be a 22–simplex of 𝒳{\mathcal{X}} with vertices s,v(s),ws,v(s),w, for s𝒦s\in{\mathcal{K}} and w𝒱w\in\mathcal{V} with dtree(w,v(s))=1d_{\mathrm{tree}}(w,v(s))=1. Then

Lk(Δ)=𝒦w and Sat(Δ)={w}Lk𝚵(w)(0).\mathrm{Lk}(\Delta)={\mathcal{K}}^{w}\quad\mbox{ and }\quad\mathrm{Sat}(\Delta)=\{w\}\cup\mathrm{Lk}_{{\bf\Xi}}(w)^{(0)}.

Moreover, 𝒞(Δ)\mathcal{C}(\Delta) is quasi-isometric to 𝒦w\mathcal{K}^{w}, the quasi-isometry being the identity at the level of vertices.

v\,\,v\,\,t1t_{1}s\,\,s\,\,t3t_{3}w\,\,w\,\,s1s_{1}s2s_{2}s3s_{3}v3v_{3}t1t_{1}t2t_{2}t3t_{3}v4v_{4}t1t_{1}t2t_{2}t3t_{3}v1v_{1}t1t_{1}t2t_{2}t3t_{3}w\,\,w\,\,...v1v_{1}t1t_{1}t2t_{2}t3t_{3}v1v_{1}t1t_{1}t2t_{2}t3t_{3}Sat(Δ)\mathrm{Sat}(\Delta).....s1s_{1}s2s_{2}s3s_{3}........Lk(Δ)=𝒦w\mathrm{Lk}(\Delta)={\mathcal{K}}^{w}
Figure 8. 𝒦\mathcal{K}–type simplex: Part of the link and saturation of a 22–simplex Δ\Delta with vertices s,v=v(s),ws,v=v(s),w (colors correspond to elements of 𝒱\mathcal{V}).
Proof.

It is clear from the definitions that the link of Δ\Delta is as described. Also, any simplex with vertex set of the form {t,v(t),w}\{t,v(t),w\} for some t𝒦t\in{\mathcal{K}} with dtree(v(t),w)=1d_{\mathrm{tree}}(v(t),w)=1 has the same link as Δ\Delta. Therefore, to prove that the saturation is as described we are left to show if a simplex Δ\Delta^{\prime} has the same link as Δ\Delta, then its vertex set is contained in the set we described. Given any vertex u𝒱u\in\mathcal{V} of Δ\Delta^{\prime}, it has to be connected to all t𝒦t\in{\mathcal{K}} with v(t)=wv(t)=w, implying that either u=wu=w or dtree(u,w)=1d_{\mathrm{tree}}(u,w)=1, as required for vertices in 𝒱\mathcal{V}. Similarly, any vertex u𝒦u\in{\mathcal{K}} of Δ\Delta^{\prime} has to be connected to all t𝒦t\in{\mathcal{K}} with v(t)=wv(t)=w, implying dtree(v(u),w)=1d_{\mathrm{tree}}(v(u),w)=1, and we are done.

To prove that 𝒞(Δ)\mathcal{C}(\Delta) is naturally quasi-isometric to 𝒦w\mathcal{K}^{w}, it suffices to show that if u,t𝒞(Δ)u,t\in\mathcal{C}(\Delta) are connected by an edge in 𝒳+𝒲{\mathcal{X}}^{+{\mathcal{W}}}, then they are uniformly close in 𝒦w\mathcal{K}^{w} and that, vice versa, if d𝒦w(u,t)=1d_{\mathcal{K}^{w}}(u,t)=1, then u,tu,t are connected by an edge in 𝒞(Δ)\mathcal{C}(\Delta).

First, if u,t𝒞(Δ)u,t\in\mathcal{C}(\Delta) are connected by an edge in 𝒳+𝒲{\mathcal{X}}^{+{\mathcal{W}}}, then there exist u,tu^{\prime},t^{\prime} so that d¯(M(u,t),M(u,t))R\bar{d}(M(u,t^{\prime}),M(u^{\prime},t))\leq R^{\prime} (see Lemma 4.14). In particular, d¯(M(u),M(t))R\bar{d}(M(u),M(t))\leq R^{\prime}, which in turn gives a uniform bound on the distance in the path metric of 𝚯w{\bf\Theta}^{w} between M(u)M(u) and M(t)M(t) because the metrics d¯\bar{d} and the path metric on θw\theta^{w} are coarsely equivalent. By Proposition 3.1(1) we must also have a uniform bound on d𝒦w(u,t)d_{\mathcal{K}^{w}}(u,t).

Now suppose that d𝒦w(u,t)=1d_{\mathcal{K}^{w}}(u,t)=1. We can then deduce from Proposition 3.1(3) that d¯(M(s,u),M(s,t))K12+K110R\bar{d}(M(s^{\prime},u),M(s^{\prime},t))\leq K_{1}^{2}+K_{1}\leq 10R for any fixed s𝒦v(s)s^{\prime}\in\mathcal{K}^{v(s)} (the proposition yields the analogous upper bound in the path metric of 𝚯w{\bf\Theta}^{w}, which is a stronger statement). Therefore u,tu,t are connected by an edge in 𝒳+𝒲{\mathcal{X}}^{+{\mathcal{W}}}, whence in 𝒞(Δ)\mathcal{C}(\Delta), as required. ∎

The remaining simplices are not particularly interesting as their links are joins (or points), and hence have diameter at most 22, but we will still need to verify properties for them. We define the type of a simplex in 𝒳{\mathcal{X}} to be the graph isomorphism type of its link. We describe these simplices with finite diameter links in the next lemma. Recall the definition of Lk𝚵(w)\mathrm{Lk}_{{\bf\Xi}}(w) in (12).

Lemma 4.20.

The following is a list of all types of non-empty, nonmaximal, simplices Δ\Delta of 𝒳{\mathcal{X}} that are not of 𝚵{\bf\Xi}–type or 𝒦\mathcal{K}–type, together with their links. In each case, the link is a nontrivial join (or a point), and 𝒞(Δ)\mathcal{C}(\Delta) has diameter at most 22.

In the table below the simplices Δ\Delta have vertices u,w𝒱u,w\in\mathcal{V} with dtree(u,w)=1d_{\mathrm{tree}}(u,w)=1 and s,t𝒦s,t\in{\mathcal{K}} with dtree(v(s),v(t))=1d_{\mathrm{tree}}(v(s),v(t))=1 and dtree(v(s),u)=1d_{\mathrm{tree}}(v(s),u)=1.

Δ\Delta Lk(Δ)\mathrm{Lk}(\Delta)
{u}\phantom{\frac{\int_{1}^{2}}{\int_{1}^{2}}}\{u\}\phantom{\frac{\int_{1}^{2}}{\int_{1}^{2}}} 𝒦uLk𝚵(u){\mathcal{K}}^{u}\star\mathrm{Lk}_{\bf\Xi}(u)
{s}\phantom{\frac{\int_{1}^{2}}{\int_{1}^{2}}}\{s\}\phantom{\frac{\int_{1}^{2}}{\int_{1}^{2}}} {v(s)}Lk𝚵(v(s))\{v(s)\}\star\mathrm{Lk}_{\bf\Xi}(v(s))
{u,w}\phantom{\frac{\int_{1}^{2}}{\int_{1}^{2}}}\{u,w\}\phantom{\frac{\int_{1}^{2}}{\int_{1}^{2}}} 𝒦u𝒦w{\mathcal{K}}^{u}\star{\mathcal{K}}^{w}
{s,t}\phantom{\frac{\int_{1}^{2}}{\int_{1}^{2}}}\{s,t\}\phantom{\frac{\int_{1}^{2}}{\int_{1}^{2}}} {v(s),v(t)}\{v(s),v(t)\}
{s,u}\phantom{\frac{\int_{1}^{2}}{\int_{1}^{2}}}\{s,u\}\phantom{\frac{\int_{1}^{2}}{\int_{1}^{2}}} {v(s)}𝒦u\{v(s)\}\star{\mathcal{K}}^{u}
{s,t,v(s)}\phantom{\frac{\int_{1}^{2}}{\int_{1}^{2}}}\{s,t,v(s)\}\phantom{\frac{\int_{1}^{2}}{\int_{1}^{2}}} {v(t)}\{v(t)\}
Proof.

This is straightforward given the definition of 𝒳{\mathcal{X}} and we leave its verification to the reader. Referring to Figure 6, and comparing with Figures 7 and 8, may be helpful. ∎

The next lemma collects a few additional properties we will need. There are 99 types of nonempty simplices: maximal, 𝚵{\bf\Xi}–type, 𝒦{\mathcal{K}}–type, and the six types listed in Lemma 4.20.

Lemma 4.21.

The following hold in 𝒳{\mathcal{X}}.

  1. (a)(a)

    The link of a simplex with a given type cannot be strictly contained in the link of a simplex with the same type.

  2. (b)(b)

    For all non-maximal simplices Δ\Delta and Δ\Delta^{\prime} so that there is a simplex Ω\Omega with Lk(Ω)Lk(Δ)Lk(Δ)\mathrm{Lk}(\Omega)\subseteq\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta) and diam(𝒞(Ω))>3diam(\mathcal{C}(\Omega))>3, there exists a simplex Π\Pi in the link of Δ\Delta^{\prime} with Lk(ΔΠ)Lk(Δ)\mathrm{Lk}(\Delta^{\prime}*\Pi)\subseteq\mathrm{Lk}(\Delta) so that for any Ω\Omega as above we have Lk(Ω)Lk(ΔΠ)\mathrm{Lk}(\Omega)\subseteq\mathrm{Lk}(\Delta^{\prime}*\Pi).

Proof.

Part (a)(a) follows directly from the descriptions of the simplices given in Lemmas 4.18, 4.19, and 4.20, and we leave this to the reader.

Before we prove (b)(b), we suppose Lk(Δ)Lk(Δ)\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta)\neq\emptyset, and make a few observations. First, Δ\Delta^{\prime} and Δ\Delta must project by ZZ to the same tree: Z(Δ),Z(Δ)TαZ(\Delta^{\prime}),Z(\Delta)\subset T_{\alpha} for some α𝒫\alpha\in\mathcal{P}. Next, note that Z(Lk(Δ)Lk(Δ))Z(\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta)) is contained in the intersection of the stars in TαT_{\alpha} of Z(Δ)Z(\Delta^{\prime}) and Z(Δ)Z(\Delta). Moreover, (as in any tree) the intersection of these two stars is contained in a single edge, or else Z(Δ)=Z(Δ)={w}Tα(0)𝒱Z(\Delta^{\prime})=Z(\Delta)=\{w\}\in T_{\alpha}^{(0)}\subset\mathcal{V} is a single point. In this latter case, by (9), we have

Δ,Δ{w}𝒦w.\Delta,\Delta^{\prime}\subset\{w\}\star{\mathcal{K}}^{w}.

Next, note that for any 𝒦{\mathcal{K}}–type simplex Ω={s,v(s),w}\Omega=\{s,v(s),w\}, Lk(Ω)=𝒦w\mathrm{Lk}(\Omega)={\mathcal{K}}^{w} by Lemma 4.19, and if Lk(Ω)Lk(Δ)Lk(Δ)\mathrm{Lk}(\Omega)\subset\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta), then ww is in the intersection of the stars of Z(Δ)Z(\Delta^{\prime}) and Z(Δ)Z(\Delta). For a 𝚵{\bf\Xi}–type simplex Ω={s,v(s)}\Omega=\{s,v(s)\}, Lk(Ω)=Lk𝚵(v(s))\mathrm{Lk}(\Omega)=\mathrm{Lk}_{{\bf\Xi}}(v(s)) by Lemma 4.18, and together with Lemma 4.20 and the previous paragraph, we see that Lk(Ω)Lk(Δ)Lk(Δ)\mathrm{Lk}(\Omega)\subset\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta) if and only if Z(Δ)=Z(Δ)={v(s)}Z(\Delta^{\prime})=Z(\Delta)=\{v(s)\} in TαT_{\alpha}.

With these observations in hand, we proceed to the proof of (b)(b), which divides into two cases.

Case 1. There is a 𝚵{\bf\Xi}–type simplex Ω={s,v(s)}\Omega=\{s,v(s)\} with Lk(Ω)Lk(Δ)Lk(Δ)\mathrm{Lk}(\Omega)\subset\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta).

In this situation, Z(Δ)=Z(Δ)={v(s)}Z(\Delta^{\prime})=Z(\Delta)=\{v(s)\}, and thus Δ,Δ{v(s)}𝒦v(s)\Delta,\Delta^{\prime}\subset\{v(s)\}\star{\mathcal{K}}^{v(s)}. From Lemmas 4.18 and 4.20, we see that Lk(Δ)Lk(Δ)\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta) must be equal to one of Lk(Ω)\mathrm{Lk}(\Omega), Lk({s})\mathrm{Lk}(\{s\}), or Lk({v(s)})\mathrm{Lk}(\{v(s)\}). Inspection of these links shows that Lk(Ω)\mathrm{Lk}(\Omega) is the only link of a 𝚵{\bf\Xi}–type simplex contained in it. First suppose that Lk(Δ)Lk(Δ)\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta) has the form Lk(Ω)\mathrm{Lk}(\Omega) or Lk({s})\mathrm{Lk}(\{s\}). In this situation, we easily find ΠLk(Δ)\Pi\subset\mathrm{Lk}(\Delta^{\prime}) so that Lk(Ω)=Lk(ΔΠ)\mathrm{Lk}(\Omega)=\mathrm{Lk}(\Delta^{\prime}\star\Pi). Furthermore, for any 𝒦{\mathcal{K}}–type simplex link 𝒦w{\mathcal{K}}^{w} in the intersection, we must have 𝒦wLk(Ω)=Lk(ΔΠ){\mathcal{K}}^{w}\subset\mathrm{Lk}(\Omega)=\mathrm{Lk}(\Delta^{\prime}\star\Pi). Therefore, the link of any 𝚵{\bf\Xi}–or 𝒦{\mathcal{K}}–type simplex contained in Lk(Δ)Lk(Δ)\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta) must be contained in Lk(Ω)=Lk(ΔΠ)\mathrm{Lk}(\Omega)=\mathrm{Lk}(\Delta^{\prime}\star\Pi), as required. Now suppose instead that Lk(Δ)Lk(Δ)=Lk({v(s)})\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta)=\mathrm{Lk}(\{v(s)\}). By Lemma 4.20, we see that Δ=Δ={v(s)}\Delta=\Delta^{\prime}=\{v(s)\}. In this case, setting Π=\Pi=\emptyset trivially completes the proof since then Lk(Δ)Lk(Δ)=Lk(Δ)=Lk(ΔΠ)\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta)=\mathrm{Lk}(\Delta^{\prime})=\mathrm{Lk}(\Delta^{\prime}\star\Pi).

Case 2. No link of a 𝚵{\bf\Xi}–type simplex is contained in Lk(Δ)Lk(Δ)\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta).

From the observations above, Z(Δ)Z(\Delta) and Z(Δ)Z(\Delta^{\prime}) do not consist of the same single point, and hence the stars of Z(Δ)Z(\Delta) and Z(Δ)Z(\Delta^{\prime}) intersect in either a point or an edge in TαT_{\alpha}. Since Z(Lk(Δ)Lk(Δ))Z(\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta)) is contained in the intersection of these stars, there are at most two 𝒦{\mathcal{K}}–type simplices whose links are contained in Lk(Δ)Lk(Δ)\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta). If there are two 𝒦{\mathcal{K}}–type simplices Ω,Ω\Omega,\Omega^{\prime} with 𝒦w=Lk(Ω){\mathcal{K}}^{w}=\mathrm{Lk}(\Omega) and 𝒦u=Lk(Ω){\mathcal{K}}^{u}=\mathrm{Lk}(\Omega^{\prime}) contained in Lk(Δ)Lk(Δ)\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta), then observe that

𝒦u,𝒦wLk({u,w})=𝒦u𝒦wLk(Δ)Lk(Δ).{\mathcal{K}}^{u},{\mathcal{K}}^{w}\subset\mathrm{Lk}(\{u,w\})={\mathcal{K}}^{u}\star{\mathcal{K}}^{w}\subset\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta).

By inspection of the possible links in Lemmas 4.18, 4.19, and 4.20, it must be that Δ\Delta^{\prime} is either {u}\{u\}, {w}\{w\}, or {u,w}\{u,w\}, and so setting Π\Pi to be {w}\{w\}, {u}\{u\}, or \emptyset, respectively, we are done. On the other hand, if there is exactly one 𝒦{\mathcal{K}}–type simplex Ω\Omega with 𝒦w=Lk(Ω)Lk(Δ)Lk(Δ){\mathcal{K}}^{w}=\mathrm{Lk}(\Omega)\subset\mathrm{Lk}(\Delta^{\prime})\cap\mathrm{Lk}(\Delta), then again inspecting all possible situations, we can find ΠLk(Δ)\Pi\subset\mathrm{Lk}(\Delta^{\prime}) with 𝒦w=Lk(Ω)=Lk(ΔΠ){\mathcal{K}}^{w}=\mathrm{Lk}(\Omega)=\mathrm{Lk}(\Delta^{\prime}\star\Pi), and again we are done with this case. This completes the proof. ∎

Lemma 4.22.

There exists L1L\geq 1 so that for every non-maximal simplex, there is an (L,L)(L,L)–coarsely Lipschitz retraction rΔ:YΔ𝐏(𝒞(Δ))r_{\Delta}\colon Y_{\Delta}\to{\bf P}({\mathcal{C}(\Delta)}). In particular, 𝒞(Δ)\mathcal{C}(\Delta) is uniformly quasi-isometrically embedded in YΔY_{\Delta}.

Proof.

By Lemma 4.20, we only have to consider simplices of 𝚵{\bf\Xi}–and 𝒦\mathcal{K}–type.

Consider Δ={s,v}\Delta=\{s,v\} with v=v(s)v=v(s) of 𝚵{\bf\Xi}–type first. Recall from Lemma 4.18 that 𝚵v{\bf\Xi}^{v} naturally includes into 𝒞(Δ)\mathcal{C}(\Delta) by a quasi-isometry. Here we will use make use of the map ξv\xi^{v}, whose relevant properties for our current purpose are stated in Proposition 3.19. For a vertex u𝒱Sat(Δ)u\in\mathcal{V}\smallsetminus\mathrm{Sat}(\Delta) (so, uvu\neq v) we define rΔ(u)=ξv(𝚯u)r_{\Delta}(u)=\xi^{v}({\bf\Theta}^{u}). For t𝒦Sat(Δ)t\in{\mathcal{K}}\smallsetminus\mathrm{Sat}(\Delta) we define rΔ(t)=ξv(𝚯v(t))r_{\Delta}(t)=\xi^{v}({\bf\Theta}^{v(t)}). Notice that the sets rΔ(u)r_{\Delta}(u) are uniformly bounded by Proposition 3.19 (and Lemma 4.18). Also, rΔr_{\Delta} is coarsely the identity on the vertices of Lk(Δ)\mathrm{Lk}(\Delta) in 𝒱\mathcal{V} by Equation (8) and Proposition 3.19(2b). To check that rΔr_{\Delta} is coarsely Lipschitz it suffices to consider 𝒳+𝒲{\mathcal{X}}^{+{\mathcal{W}}}–adjacent vertices of 𝒱\mathcal{V}. Notice that vertices w,w𝒱w,w^{\prime}\in\mathcal{V} that are adjacent in 𝒳+𝒲{\mathcal{X}}^{+{\mathcal{W}}} have corresponding 𝚯w,𝚯w{\bf\Theta}^{w},{\bf\Theta}^{w^{\prime}} within 10R10R of each other in E¯\bar{E}. Indeed, 𝚯v{\bf\Theta}^{v} and 𝚯w{\bf\Theta}^{w} actually intersect if w,ww,w^{\prime} are adjacent in 𝒳{\mathcal{X}}, and they contain subsets M()M(\cdot) within 10R10R of each other if w,ww,w^{\prime} are contained in 𝒲{\mathcal{W}}–adjacent maximal simplices (this is true regardless of which case of the definition (11) for the edges of 𝒲{\mathcal{W}} applies). The fact that rΔr_{\Delta} is coarsely Lipschitz now follows from, Proposition 3.19(1), which says that ξv\xi^{v} is coarsely Lipschitz on E¯\bar{E}.

Next, consider Δ={s,v(s),w}\Delta=\{s,v(s),w\} of 𝒦\mathcal{K}–type. For a vertex u𝒱Sat(Δ)u\in\mathcal{V}\smallsetminus\mathrm{Sat}(\Delta) (so, dtree(u,w)2d_{\mathrm{tree}}(u,w)\geq 2), define rΔ(u)=Λw(𝚯u)r_{\Delta}(u)=\Lambda^{w}({\bf\Theta}^{u}). For a vertex t𝒦Sat(Δ)t\in{\mathcal{K}}\smallsetminus\mathrm{Sat}(\Delta), define rΔ(t)=Λw(M(t))r_{\Delta}(t)=\Lambda^{w}(M(t)). Notice that, by definition of M(t)M(t), if t𝒦wt\in{\mathcal{K}}^{w}, then rΔ(t)r_{\Delta}(t) lies within Hausdorff distance K1K_{1} of tt. Also, since dtree(u,w)2d_{\mathrm{tree}}(u,w)\geq 2, Proposition 3.19(2a) ensures that the diameter of Λw(𝚯u)\Lambda^{w}({\bf\Theta}^{u}) is bounded. Since M(t)𝚯v(t)M(t)\subseteq{\bf\Theta}^{v(t)}, we see that all the sets in the image of rΔr_{\Delta} are bounded, and also we see that in order to prove that rΔr_{\Delta} is Lipschitz it suffices to consider vertices of 𝒦{\mathcal{K}}. But vertices s,t𝒦s,t\in{\mathcal{K}} that are adjacent in 𝒳+𝒲{\mathcal{X}}^{+{\mathcal{W}}} have corresponding M(s),M(t)M(s),M(t) within 10R10R of each other in E¯\bar{E}, so the conclusion follows from Proposition 3.19(1), which states that Λw\Lambda^{w} is coarsely Lipschitz on E¯\bar{E}. ∎

4.4. Final proof

We now have all the tools necessary for the:

Proof of Theorem 4.16.

We must verify each of the conditions from Definition 4.8.

Item (1) (bound on length of \sqsubseteq–chains) follows from Lemma 4.21(a), which implies that any chain Lk(Δ1)\mathrm{Lk}(\Delta_{1})\subsetneq\dots has length bounded by the number of possible types, which is 9.

Let us now discuss item (2) of the definition. The descriptions of the 𝒞(Δ)\mathcal{C}(\Delta) from Lemmas 4.17, 4.18, 4.19, 4.20 yields that all 𝒞(Δ)\mathcal{C}(\Delta) are hyperbolic, since each of them is either bounded or uniformly quasi-isometric to one of E^\hat{E} (which is hyperbolic by Theorem 2.1), some 𝚵v{\bf\Xi}^{v} (which is hyperbolic by Lemma 3.10), or \mathbb{R}. Moreover, any 𝒞(Δ)\mathcal{C}(\Delta) is (uniformly) quasi-isometrically embedded in YΔY_{\Delta} by Lemma 4.22.

Item (3) of the definition (common nesting) is precisely Lemma 4.21(b).

Finally, we show item (4) of the definition (fullness of links), which we recall for the convenience of the reader:

  • If v,wv,w are distinct non-adjacent vertices of Lk(Δ)\mathrm{Lk}(\Delta), for some simplex Δ\Delta of 𝒳{\mathcal{X}}, contained in 𝒲{\mathcal{W}}–adjacent maximal simplices, then they are contained in 𝒲{\mathcal{W}}–adjacent simplices of the form ΔΔ\Delta\star\Delta^{\prime}.

It suffices to consider simplices Δ\Delta of 𝚵{\bf\Xi}–and 𝒦\mathcal{K}–type. Indeed, in all other cases (see Lemma 4.20), the vertices v,wv,w under consideration are contained in the link of a simplex Δ\Delta^{\prime} containing Δ\Delta where Δ\Delta^{\prime} is of 𝚵{\bf\Xi}–or 𝒦\mathcal{K}–type (as can be seen by enlarging Δ\Delta until its link is no longer a join; vv and ww are not 𝒳{\mathcal{X}}–adjacent so they are contained in the same “side” of any join structure). Hence, once we deal with those cases, we know that there are suitable maximal simplices containing the larger simplex, whence Δ\Delta.

Consider first a simplex Δ\Delta of 𝒦\mathcal{K}–type with vertices s,v(s),ws,v(s),w. Consider distinct vertices t1,t2t_{1},t_{2} (necessarily in 𝒦{\mathcal{K}}) of Lk(Δ)\mathrm{Lk}(\Delta), and suppose that there are vertices s1,s2𝒦s_{1},s_{2}\in{\mathcal{K}} so that the maximal simplices σ(s1,t1)\sigma(s_{1},t_{1}) and σ(s2,t2)\sigma(s_{2},t_{2}) are connected in 𝒲{\mathcal{W}}. There are two possibilities:

  • d¯(M(s1,t1),M(s2,t2))10R\bar{d}(M(s_{1},t_{1}),M(s_{2},t_{2}))\leq 10R. In this case, we have d¯(M(t1),M(t2))10R\bar{d}(M(t_{1}),M(t_{2}))\leq 10R. In particular, in view of the second bullet in the definition of the edges of 𝒲{\mathcal{W}}, we have that t1,t2t_{1},t_{2} are contained, respectively, in the 𝒲{\mathcal{W}}–connected maximal simplices Δt1=σ(s,t1)\Delta*t_{1}=\sigma(s,t_{1}) and Δt2=σ(s,t2)\Delta*t_{2}=\sigma(s,t_{2}).

  • s1=s2s_{1}=s_{2} and d¯(M(t1),M(t2))10R\bar{d}(M(t_{1}),M(t_{2}))\leq 10R (notice that t1t2t_{1}\neq t_{2} so that the “symmetric” case cannot occur). Again, we reach the same conclusion as above.

We can now consider a simplex Δ\Delta of 𝚵{\bf\Xi}–type with vertices s,v(s)s,v(s). Consider vertices x1,x2x_{1},x_{2} of Lk(Δ)\mathrm{Lk}(\Delta) that are not 𝒳{\mathcal{X}}–adjacent but are contained in 𝒲{\mathcal{W}}–adjacent maximal simplices. Furthermore, we can assume that x1,x2x_{1},x_{2} are not in the link of a simplex of 𝒦\mathcal{K}–type (the case we just dealt with) which contains Δ\Delta, since in that case we already know that there are suitable maximal simplices containing the larger simplex, whence Δ\Delta. Then, using the structure of Lk(Δ)\mathrm{Lk}(\Delta), we see that there are vertices si,ti𝒦s_{i},t_{i}\in{\mathcal{K}} so that:

  • xi{ti,v(ti)}x_{i}\in\{t_{i},v(t_{i})\},

  • tit_{i} and v(ti)v(t_{i}) all belong to Lk(Δ)\mathrm{Lk}(\Delta), and

  • σ(s1,t1)\sigma(s_{1},t_{1}) and σ(s2,t2)\sigma(s_{2},t_{2}) are connected in 𝒲{\mathcal{W}}.

In turn, the last bullet splits into two cases:

  • d¯(M(s1,t1),M(s2,t2))10R\bar{d}(M(s_{1},t_{1}),M(s_{2},t_{2}))\leq 10R. In this we have d¯(M(t1),M(t2))10R\bar{d}(M(t_{1}),M(t_{2}))\leq 10R, so that x1,x2x_{1},x_{2} are contained, respectively, in the 𝒲{\mathcal{W}}–connected maximal simplices Δ{t1,v(t1)}=σ(s,t1)\Delta*\{t_{1},v(t_{1})\}=\sigma(s,t_{1}) and Δ{t2,v(t2)}=σ(s,t2)\Delta*\{t_{2},v(t_{2})\}=\sigma(s,t_{2}), so this case is fine.

  • s1=s2s_{1}=s_{2} and d¯(M(t1),M(t2))10R\bar{d}(M(t_{1}),M(t_{2}))\leq 10R. But again we reach the same conclusion as before.

We now also have to check the existence of an action of Γ\Gamma with the required properties. The action is constructed in Lemma 4.15, where all properties are checked except finiteness of the number of orbits of links of 𝒳{\mathcal{X}}. The finitely many possible types of links are listed in Lemmas 4.174.20, and for each type of simplex there are only finitely many orbits, so we are done. ∎

5. Quasi-isometric rigidity

In this section, using the HHS structure, we prove a strong form of rigidity for the group Γ\Gamma and the model space E¯\bar{E}. Recall that E¯\bar{E} is defined via a particular truncation D¯\bar{D} of the Teichmüller disk DD obtained by removing 11–separated horoballs. We say that such a truncation E¯\bar{E} is an allowable truncation of EE if Γ\Gamma acts by isometries on it with cocompact quotient. Write Isom(Ω)\operatorname{Isom}(\Omega) and QI(Ω)\operatorname{QI}(\Omega) for the isometry group and quasi-isometry group, respectively, of a metric space Ω\Omega. For E¯\bar{E}, we write Isomfib(E¯)Isom(E¯)\mathrm{Isom_{fib}}(\bar{E})\leq\operatorname{Isom}(\bar{E}) for the subgroup of isometries that map fibers to fibers.

Theorem 1.7.

There is an allowable truncation E¯\bar{E} of EE such that the natural homomorphisms Isomfib(E¯)Isom(E¯)QI(E¯)QI(Γ)\mathrm{Isom_{fib}}(\bar{E})\to\operatorname{Isom}(\bar{E})\to\operatorname{QI}(\bar{E})\cong\operatorname{QI}(\Gamma) are all isomorphisms, and ΓIsom(E¯)QI(Γ)\Gamma\leq\operatorname{Isom}(\bar{E})\cong\operatorname{QI}(\Gamma) has finite index.

The proof is divided up into several steps which we outline here before getting into the details. The first step is to use the HHS structure to identify certain quasi-flats in E¯\bar{E}, and prove that they are coarsely preserved by a quasi-isometry. The maximal quasi-flats are encoded by the strip bundles in E¯\bar{E}, and using the preservation of quasi-flats, we show that a quasi-isometry further preserves strip bundles, and even sends all strip bundles for strips in any fixed direction to strip bundles in some other fixed direction. From there we deduce that a quasi-isometry sends fibers EXE_{X} within a bounded distance of some other fibers EYE_{Y}, and in fact induces a quasi-isometry between the fibers. Fixing attention on E0E_{0} and further appealing to the structure of strip bundles, we show that a self quasi-isometry of E¯\bar{E} induces a special type of quasi-isometry from E0E_{0} to itself sending strips in a fixed direction within a uniformly bounded distance of strips in some other fixed direction. This quasi-isometry is promoted to a piecewise affine biLipschitz map from E0E_{0} to itself, which we then show is in fact affine. This produces a homomorphism to the full affine group of E0E_{0}, QI(E¯)Aff(E0)\operatorname{QI}(\bar{E})\to\rm Aff(E_{0}). Given an affine homeomorphism of E0E_{0}, we construct an explicit fiber preserving isometry associated to it, which via the inclusions Isomfib(E¯)QI(E¯)\mathrm{Isom_{fib}}(\bar{E})\to\operatorname{QI}(\bar{E}) serves as a one-sided inverse. Finally, we prove that the homomorphism QI(E¯)Aff(E0)\operatorname{QI}(\bar{E})\to\rm Aff(E_{0}) is injective, hence the homomorphisms Isomfib(E¯)Isom(E¯)QI(E¯)Aff(E0)\mathrm{Isom_{fib}}(\bar{E})\to\operatorname{Isom}(\bar{E})\to\operatorname{QI}(\bar{E})\to\rm Aff(E_{0}) are all isomorphisms. The fact that Γ\Gamma has finite index in Isomfib(E¯)\mathrm{Isom_{fib}}(\bar{E}), and hence in Isom(E¯)\operatorname{Isom}(\bar{E}), is straightforward using the cocompactness of the action of Γ\Gamma and the singular structure.

5.1. HHS structure and quasi-flats

Denote by 𝔖0\mathfrak{S}_{0} the set 𝒱×{0,1}\mathcal{V}\times\{0,1\}. We denote the element (v,0)(v,0) by vqtv^{qt} (for “quasi-tree”) and (v,1)(v,1) by vqlv^{ql} (for “quasi-line”).

We denote by \mathcal{F} the set of all strip bundles of E¯\bar{E}, that is, subbundles with fiber a strip and base the horocycle corresponding to the direction of the strip. (Roughly, these are the flats of the peripheral graph manifolds.)

Proposition 5.1 (Properties of the HHS structure).

The HHS structure (E¯,𝔖)(\bar{E},\mathfrak{S}) on E¯\bar{E} coming from Theorem 4.16 has the following properties, for some K1K\geq 1.

  1. (1)

    The set of non-\sqsubseteq-maximal Y𝔖Y\in\mathfrak{S} with diam(𝒞(Y))3diam(\mathcal{C}(Y))\geq 3 is in bijection with 𝔖0\mathfrak{S}_{0}. Under said bijection:

  2. (2)

    𝒞(vqt)\mathcal{C}(v^{qt}) is (K,K)(K,K)-quasi-isometric to a quasi-tree with at least two points at infinity, and 𝒞(vql)\mathcal{C}(v^{ql}) is (K,K)(K,K)–quasi-isometric to a line;

  3. (3)

    For all v𝒱v\in\mathcal{V}, we have vqtvqlv^{qt}\bot v^{ql};

  4. (4)

    For all adjacent v,w𝒱v,w\in\mathcal{V}, we have wqlvqlw^{ql}\bot v^{ql} and wqlvqtw^{ql}\sqsubseteq v^{qt};

  5. (5)

    All pairs of elements of 𝔖0\mathfrak{S}_{0} that do not fall into the aforementioned cases are transverse;

  6. (6)

    For each adjacent v,w𝒱v,w\in\mathcal{V} there is FF\in\mathcal{F} so πvql(F)\pi_{v^{ql}}(F) and πwql(F)\pi_{w^{ql}}(F) are KK–coarsely dense, and πY(F)\pi_{Y}(F) has diameter at most KK for all Yvql,wqlY\neq v^{ql},w^{ql}.

Proof.

The second paragraph of the proof of Theorem 4.16 implies 𝒞(vqt)\mathcal{C}(v^{qt}) is quasi-isometric to the quasi-tree 𝚵v{\bf\Xi}^{v} (with at least two points at infinity by Lemma 3.10) and 𝒞(vql)\mathcal{C}(v^{ql}) is quasi-isometric to the quasi-line 𝒦v\mathcal{K}^{v}, and that these are the only non-maximal elements of diameter at least 33. This proves (1) and (2). In view of the combinatorial description of orthogonality and nesting from Definition 4.9, properties (3)-(5) boil down to combinatorial properties of 𝒳\mathcal{X} that are straightforward to check. For example, regarding property (3) note that two (equivalence classes of) simplices are orthogonal if their links form a join. The links of the simplices corresponding to vqtv^{qt} and vqlv^{ql} are Lk𝚵(v)\mathrm{Lk}_{{\bf\Xi}}(v) and 𝒦v\mathcal{K}^{v} (see Lemmas 4.18 and 4.19), which indeed form a join.

Regarding property (6), first of all the projections in the HHS structure on E¯\bar{E} are obtained composing the quasi-isometry E¯𝒲\bar{E}\to\mathcal{W} from Lemma 4.15 and the projections defined in Definition 4.10 (roughly, those are closest-point projections in the complement of saturations).

The required strip bundle is the intersection 𝚯v𝚯w{\bf\Theta}^{v}\cap{\bf\Theta}^{w}, which under the quasi-isometry of Lemma 4.15 corresponds to the set of all maximal simplices of 𝒲\mathcal{W} of the form σ(s,t)\sigma(s,t) for s𝒦vs\in\mathcal{K}^{v}, t𝒦wt\in\mathcal{K}^{w}. In view of the description of the πY\pi_{Y} from Definition 4.10, the coarse density claim follows since the union of the simplices described above contains the links of the simplices corresponding to vqlv^{ql} and wqlw^{ql}, which are 𝒦v\mathcal{K}^{v} and 𝒦w\mathcal{K}^{w}.

Regarding the boundedness claim, it can be checked case-by-case that the set of simplices described above gives a bounded set of YΔY_{\Delta} for [Δ]vql,wql[\Delta]\neq v^{ql},w^{ql} (for example, note that said set is bounded if the saturation of Δ\Delta does not intersect 𝒦v𝒦w\mathcal{K}^{v}\cup\mathcal{K}^{w}, or if it does not contain vv or ww). This implies boundedness of the projections since the projections are coarsely Lipschitz; this follows from Theorem 4.16 since the projection maps of an HHS are required to be coarsely Lipschitz. ∎

From now on we identify 𝔖0\mathfrak{S}_{0} with the set of all Y𝔖Y\in\mathfrak{S} with diam(𝒞(Y))3\mathrm{diam}(\mathcal{C}(Y))\geq 3 as in Proposition 5.1. Notice that the maximal number of pairwise orthogonal elements of 𝔖0\mathfrak{S}_{0} is 2. Therefore, a complete support set as in [BHS21, Definition 5.1] is just a pair of orthogonal elements of 𝔖0\mathfrak{S}_{0}.

Let \mathcal{H} be the set of pairs (Y,p)(Y,p) where Y𝔖Y\in\mathfrak{S} and pC(Y)p\in\partial C(Y) with Y=vqlY=v^{ql} for some v𝒱v\in\mathcal{V}. We say that two such pairs (Y,p)(Y,p) and (W,q)(W,q) are orthogonal if YY and WW are. Any element σ=(Y,p)\sigma=(Y,p)\in\mathcal{H} comes with a quasi-geodesic ray hσh_{\sigma} in E¯\bar{E}, as in [BHS21, Definition 5.3], so that πYhσ\pi_{Y}\circ h_{\sigma} is a quasi-geodesic in 𝒞(Y)\mathcal{C}(Y) and πW(hσ)\pi_{W}(h_{\sigma}) is bounded for all WYW\neq Y.

We recall that given subsets AA and BB of a metric space XX, we say that the subset CC of XX is the coarse intersection of AA and BB if for every sufficiently large RR we have that NR(A)NR(B)N_{R}(A)\cap N_{R}(B) lies within finite Hausdorff distance of CC. If the coarse intersection of two subsets exists, then it is well-defined up to finite Hausdorff distance.

Lemma 5.2.

Let ϕ:E¯E¯\phi\colon\bar{E}\to\bar{E} be a quasi-isometry. Then there is a bijection ϕ:\phi_{\mathcal{H}}\colon\mathcal{H}\to\mathcal{H} preserving orthogonality and so that dHaus(ϕ(hσ),hϕ(σ))<d_{\mathrm{Haus}}(\phi(h_{\sigma}),h_{\phi_{\mathcal{H}}(\sigma)})<\infty.

Proof.

Let \mathcal{H}^{\prime} be the set of pairs (Y,p)(Y,p) where Y𝔖Y\in\mathfrak{S} and pC(Y)p\in\partial C(Y), without the restriction on YY.

The lemma with \mathcal{H}^{\prime} replacing \mathcal{H} would follow directly from [BHS21, Theorem 5.7], except that the HHS structure on E¯\bar{E} does not satisfy one of the 3 required assumptions, namely Assumption 2 (while it does satisfy Assumption 1 by parts (1) and (2) of Proposition 5.1, and it also satisfies Assumption 3 since there are no 3 pairwise orthogonal elements of 𝔖0\mathfrak{S}_{0}, by parts (3)-(5) of Proposition 5.1).

Inspecting the proof of [BHS21, Theorem 5.7], we see that Assumption 2 is used in two places.

The first one is to define the map ϕ\phi_{\mathcal{H}} on a certain pair σ=(Y,p)\sigma=(Y,p)\in\mathcal{H}^{\prime}. The argument applies verbatim if YY satisfies Assumption 2, that is, if and only if YY is the intersection of 2 complete support sets. This is the case if Y=vqlY=v^{ql} for some vv\in\mathcal{H}, that is, if σ\sigma\in\mathcal{H}. Therefore, one can use that argument to define a map ϕ:\phi_{\mathcal{H}}:\mathcal{H}\to\mathcal{H}^{\prime}. What is more, the image of ϕ\phi_{\mathcal{H}} needs to be contained in \mathcal{H}. This can be seen from the fact that hϕ(σ)h_{\phi_{\mathcal{H}}(\sigma)} for σ\sigma\in\mathcal{H} arises as a coarse intersection of standard orthants, which are, essentially, products of rays hσh_{\sigma}, see [BHS21, Definition 4.1] for the precise definition. Notice that [BHS21, Lemma 4.11] says, roughly, that coarse intersections of standard orthants are the expected sub-products. Hence, the failure of Assumption 2 for Y=vqtY=v^{qt} implies that h(Y,p)h_{(Y,p)} cannot be a coarse intersection of standard orthants, and therefore ϕ(σ)\phi_{\mathcal{H}}(\sigma) for σ\sigma\in\mathcal{H} also needs to lie in \mathcal{H}.

The second place where Assumption 2 is mentioned in [BHS21, Theorem 5.7] is the proof that ϕ\phi_{\mathcal{H}} preserves orthogonality. There the assumption is used to say that certain quasi-geodesic rays are of the form hσh_{\sigma}. Such quasi-geodesic rays arise as coarse intersections of standard orthants, so, as mentioned above, they need to be of the form hσh_{\sigma} for σ\sigma\in\mathcal{H}, hence Assumption 2 is not actually needed there.

Thus, the arguments in the proof of [BHS21, Theorem 5.7] give the lemma. ∎

Lemma 5.3.

For every KK there exists CC so that the following holds. Let ϕ:E¯E¯\phi\colon\bar{E}\to\bar{E} be a (K,K)(K,K)–quasi-isometry. Then there is a bijection ϕ:\phi_{\mathcal{F}}\colon\mathcal{F}\to\mathcal{F} so that dHaus(ϕ(F),ϕ(F))Cd_{\mathrm{Haus}}(\phi(F),\phi_{\mathcal{F}}(F))\leq C for all FF\in\mathcal{F}.

Proof.

Let p±p^{\pm} be the two points at infinity of 𝒞(vql)\mathcal{C}(v^{ql}) for some v𝒱v\in\mathcal{V}. We claim that there exists w𝒱w\in\mathcal{V} so that, for q±q^{\pm} the points at infinity of 𝒞(wql)\mathcal{C}(w^{ql}), we have ϕ((vql,p±))=(wql,q±)\phi_{\mathcal{H}}((v^{ql},p^{\pm}))=(w^{ql},q^{\pm}), up to relabeling. We use that ϕ\phi_{\mathcal{H}} preserves orthogonality to show this. Let u1,u2𝒱u_{1},u_{2}\in\mathcal{V} be distinct and adjacent to vv, and let r±1,r±2r^{\pm}_{1},r^{\pm}_{2} be the points at infinity of 𝒞(u1ql)\mathcal{C}(u_{1}^{ql}), 𝒞(u2ql)\mathcal{C}(u_{2}^{ql}). Then (vql,p±)(v^{ql},p^{\pm}) are the only elements of \mathcal{H} that are orthogonal to all the (uiql,ri±)(u_{i}^{ql},r_{i}^{\pm}). Since ϕ\phi_{\mathcal{H}} preserves orthogonality, we see that ϕ((vql,p±))\phi_{\mathcal{H}}((v^{ql},p^{\pm})) are both orthogonal to the same 4 distinct elements of \mathcal{H} with the property that no pair of them is orthogonal. This is easily seen to imply that ϕ((vql,p±))\phi_{\mathcal{H}}((v^{ql},p^{\pm})) must be of the form (wql,q±)(w^{ql},q^{\pm}), since said 44 elements need to be associated to at least 2 distinct vertices of 𝒱\mathcal{V}. This shows the claim.

In view of the claim, we see that [BHS21, Lemma 5.9] applies. (We note that Assumption 2 in said Lemma is only needed to have the map from [BHS21, Theorem 5.7], but our map from Lemma 5.2 has the same defining properties, just with a smaller domain and range.)

The standard flats in [BHS21, Lemma 5.9] coarsely coincide with the elements of \mathcal{F} in view of Proposition 5.1(6) (compare with [BHS21, Definition 4.1]) so the lemma follows. ∎

Denote by 𝒮\mathcal{S} the collection of all strips in E0E_{0}, and for A𝒮A\in\mathcal{S} denote by α(A)𝒫\alpha(A)\in\mathcal{P} the direction of AA. Similarly, for FF\in\mathcal{F} we denote α(F)𝒫\alpha(F)\in\mathcal{P} the direction of the strip defining FF.

Proposition 5.4.

Given K1K\geq 1, there exists C0C\geq 0 so that if ϕ:E¯E¯\phi\colon\bar{E}\to\bar{E} is a (K,K)(K,K)–quasi-isometry, then for all XD¯X\in\bar{D}, there exists YD¯Y\in\bar{D} so that the Hausdorff distance between ϕ(EX)\phi(E_{X}) and EYE_{Y} is at most CC. In particular, dHaus(E0,ϕ(E0))<d_{\mathrm{Haus}}(E_{0},\phi(E_{0}))<\infty. Moreover, there are bijections ϕ𝒫:𝒫𝒫\phi_{\mathcal{P}}:\mathcal{P}\to\mathcal{P} and ϕ𝒮:𝒮𝒮\phi_{\mathcal{S}}:\mathcal{S}\to\mathcal{S} so that:

  1. (1)

    dHaus(ϕ(α),ϕ𝒫(α))Cd_{\mathrm{Haus}}(\phi(\partial\mathcal{B}_{\alpha}),\partial\mathcal{B}_{\phi_{\mathcal{P}}(\alpha)})\leq C for each α𝒫\alpha\in\mathcal{P}.

  2. (2)

    α(ϕ𝒮(A))=ϕ𝒫(α(A))\alpha(\phi_{\mathcal{S}}(A))=\phi_{\mathcal{P}}(\alpha(A)) and dHaus(ϕ(A),ϕ𝒮(A))<d_{\mathrm{Haus}}(\phi(A),\phi_{\mathcal{S}}(A))<\infty for all A𝒮A\in\mathcal{S},

Proof.

First, note that fibers are quantitatively coarse intersections of the sets α\partial\mathcal{B}_{\alpha}, in the sense that exists a function f:f\colon\mathbb{R}\to\mathbb{R} and t00t_{0}\geq 0 such that

  • for any XD¯X\in\bar{D} and any tt0t\geq t_{0} there are two distinct α\partial\mathcal{B}_{\alpha} whose tt–neighborhoods intersect in a set within Hausdorff distance f(t)f(t) of EXE_{X};

  • for tt0t\geq t_{0}, if the tt–neighborhoods of two distinct α\partial\mathcal{B}_{\alpha} intersect, then this intersection lies within Hausdorff distance f(t)f(t) of a fiber.

This follows via the bundle-map π:E¯D¯\pi\colon\bar{E}\to\bar{D} and the corresponding relationship between neighborhoods of distinct horocycles Bα\partial B_{\alpha} in D¯\bar{D}.

We next make three preliminary observations. Firstly, for each α𝒫\alpha\in\mathcal{P} the set α\partial\mathcal{B}_{\alpha} is the union of all FF\in\mathcal{F} with α(F)=α\alpha(F)=\alpha. Secondly, if F1,F2F_{1},F_{2}\in\mathcal{F} have α(F1)α(F2)\alpha(F_{1})\neq\alpha(F_{2}) then the coarse intersection of F1F_{1} and F2F_{2} is bounded. Indeed, FiF_{i} is contained in α(Fi)\partial\mathcal{B}_{\alpha(F_{i})}, and the coarse intersection of these α(Fi)\partial\mathcal{B}_{\alpha(F_{i})} is some (or really, any) fiber EXE_{X}. Since the coarse intersection of FiF_{i} with EXE_{X} is a strip in the corresponding direction, and strips in different directions have bounded coarse intersection, the claim follows. Thirdly, observe that F1,F2F_{1},F_{2}\in\mathcal{F} have α(F1)=α(F2)\alpha(F_{1})=\alpha(F_{2}) if and only if there is a chain of elements in \mathcal{F} from F1F_{1} to F2F_{2} so that consecutive elements have unbounded coarse intersection. The “if” part follows from the previous observation, while the “only if” follows from the fact that elements of \mathcal{F} corresponding to adjacent edges of some TαT_{\alpha} have unbounded coarse intersection.

In view of all this and Lemma 5.3, we see that for each α\alpha there exists a (necessarily unique) ϕ𝒫(α)𝒫\phi_{\mathcal{P}}(\alpha)\in\mathcal{P} so that ϕ(α)\phi(\partial\mathcal{B}_{\alpha}) and ϕ𝒫(α)\partial\mathcal{B}_{\phi_{\mathcal{P}}(\alpha)} have finite Hausdorff distance. In fact, the distance is uniformly bounded by the constant CC, depending only on KK, coming from Lemma 5.3. This is how we define ϕ𝒫\phi_{\mathcal{P}}.

Now for any XD¯X\in\bar{D}, we may choose α1,α2𝒫\alpha_{1},\alpha_{2}\in\mathcal{P} so that the fiber EXE_{X} has Hausdorff distance at most f(t0)f(t_{0}) from the intersection of the t0t_{0}–neighborhoods of αi\partial\mathcal{B}_{\alpha_{i}}. Thus there is some uniform t0t0t^{\prime}_{0}\geq t_{0}, again depending only on KK, so that ϕ(EX)\phi(E_{X}) has Hausdorff distance at most t0t_{0}^{\prime} from the intersection of the t0t_{0}^{\prime}–neighborhoods of ϕ𝒫(αi)\partial\mathcal{B}_{\phi_{\mathcal{P}}(\alpha_{i})}; further, as mentioned above, this intersection of t0t_{0}^{\prime}–neighborhoods has Hausdorff distance at most f(t0)f(t^{\prime}_{0}) to some fiber EYE_{Y}, as claimed.

Finally, we define ϕ𝒮\phi_{\mathcal{S}} via the bijection 𝒮\mathcal{F}\leftrightarrow\mathcal{S} between strip bundles and strips in E0E_{0}. That is, if A𝒮A\in\mathcal{S} corresponds to FF\in\mathcal{F}, then ϕ𝒮(A)\phi_{\mathcal{S}}(A) is the strip corresponding to ϕ(F)\phi_{\mathcal{F}}(F). Since AA is the coarse intersection of FF with E0E_{0}, the desired properties for ϕ𝒮\phi_{\mathcal{S}} then follow from the facts that αϕ(F)=ϕ𝒫(α(F))\alpha_{\phi_{\mathcal{F}}(F)}=\phi_{\mathcal{P}}(\alpha(F)) and that ϕ(F)\phi_{\mathcal{F}}(F) lies within finite Hausdorff distance of ϕ(F)\phi(F). ∎

5.2. From QI(E¯)\operatorname{QI}(\bar{E}) to QI(E0)\operatorname{QI}(E_{0})

The next step is to construct a homomorphism QI(E¯)QI(E0)\operatorname{QI}(\bar{E})\to\operatorname{QI}(E_{0}) by associating a quasi-isometry of E0E_{0} to each quasi-isometry of E¯\bar{E} (see Lemma 5.7). This step requires some preliminaries which we now explain.

To distinguish between two relevant notions of properness, we will call a map f:XYf\colon X\to Y between metric spaces topologically proper if it is continuous and preimages of compact sets are compact, and metrically proper if there exist diverging functions ρ,ρ+:00\rho_{-},\rho_{+}:\mathbb{R}_{\geq 0}\to\mathbb{R}_{\geq 0} (which we will call properness functions) such that for all x,yXx,y\in X we have

ρ(dX(x,y))dY(f(x),f(y))ρ+(dX(x,y)).\rho_{-}(d_{X}(x,y))\leq d_{Y}(f(x),f(y))\leq\rho_{+}(d_{X}(x,y)).

(Both types of maps are just referred to as “proper” in the appropriate contexts, but neither notion implies the other.)

For R>0R>0 and XD¯X\in\bar{D}, we endow NR(EX)N_{R}(E_{X}) with the restriction of the metric of E¯\bar{E}, while EXE_{X} is endowed with its path metric. Then the restriction of fX:E¯EXf_{X}\colon\bar{E}\to E_{X} to NR(EX)N_{R}(E_{X}) is metrically proper. Indeed, fXf_{X} is topologically proper and equivariant with respect to a group acting cocompactly. Note that the properness functions here can be taken independently of the fiber XX (once we fix RR) because there is also a cocompact action on D¯\bar{D}. We also note the following lemma.

Lemma 5.5.

A metrically proper coarsely surjective map between geodesic metric spaces is a quasi-isometry. Moreover, the quasi-isometry constants depend only on the properness functions and the coarse surjectivity constant.

This follows from standard arguments. First, a metrically proper map from a geodesic metric space is coarsely Lipschitz (the proof involves subdividing geodesics into segments of length at most 1, each of which has bounded image). Also, coarse surjectivity allows one to construct a quasi-inverse of the map, which is furthermore metrically proper. As above, the quasi-inverse is coarsely Lipschitz, and we conclude since a coarsely Lipschitz map with a coarsely Lipschitz quasi-inverse is a quasi-isometry.

Given any quasi-isometry ϕ:E¯E¯\phi\colon\bar{E}\to\bar{E} and XD¯X\in\bar{D}, define νϕX:EXEX\nu_{\phi}^{X}\colon E_{X}\to E_{X} to be νϕX=fXϕ|EX\nu_{\phi}^{X}=f_{X}\circ\phi|_{E_{X}}. In the case of the base fiber X=X0X=X_{0} we denote this νϕ=νϕX0\nu_{\phi}=\nu_{\phi}^{X_{0}}. When ϕ\phi is understood, we also write νX=νϕX\nu^{X}=\nu_{\phi}^{X} and ν=νϕ\nu=\nu_{\phi}.

Lemma 5.6.

For any (K,K)(K,K)–quasi-isometry ϕ:E¯E¯\phi\colon\bar{E}\to\bar{E} and XD¯X\in\bar{D}, the map νϕX:EXEX\nu_{\phi}^{X}\colon E_{X}\to E_{X} is a (K,K)(K^{\prime},K^{\prime})–quasi-isometry, where KK^{\prime} depends only on KK and dHaus(EX,ϕ(EX))d_{\mathrm{Haus}}(E_{X},\phi(E_{X})). Furthermore, for any A𝒮A\in\mathcal{S}, dHaus(νϕX(A),ϕ𝒮(A))<d_{\mathrm{Haus}}(\nu_{\phi}^{X}(A),\phi_{\mathcal{S}}(A))<\infty.

Proof.

First note that the restriction of ϕ\phi to EXE_{X} is metrically proper, since the path metric on EXE_{X} and the restricted metric from E¯\bar{E} are coarsely equivalent (that is, the identity on EXE_{X} is a metrically proper map between these metric spaces). Next let R=dHaus(EX,ϕ(EX))R=d_{\mathrm{Haus}}(E_{X},\phi(E_{X})), which is finite by Proposition 5.4, and note that the restriction fX|NR(EX):NR(EX)EXf_{X}|_{N_{R}(E_{X})}\colon N_{R}(E_{X})\to E_{X} is also metrically proper. Therefore the composition νϕX=(fX|NR(EX))(ϕ|EX)\nu_{\phi}^{X}=(f_{X}|_{N_{R}(E_{X})})\circ(\phi|_{E_{X}}) is metrically proper and, moreover, the properness functions depend only on K,RK,R and not on the fiber EXE_{X}.

By [KL12, Theorem 3.8] and the fact that EXE_{X} is uniformly quasi-isometric to 2\mathbb{H}^{2}, any metrically proper map of EXE_{X} to itself is coarsely surjective and, moreover, the coarse surjectivity constant depends only on the properness functions. Therefore νϕX\nu_{\phi}^{X} is coarsely surjective and a uniform quasi-isometry by Lemma 5.5.

Regarding the claim about AA, this follows from Proposition 5.4(2) and the fact that fXf_{X} moves each point of ϕ(A)NR(EX)\phi(A)\subseteq N_{R}(E_{X}) at most RR away. ∎

Lemma 5.7.

The assignment ϕνϕ\phi\mapsto\nu_{\phi}, for any quasi-isometry ϕ:E¯E¯\phi\colon\bar{E}\to\bar{E}, gives a well-defined homomorphism 𝒜0:QI(E¯)QI(E0)\mathcal{A}_{0}\colon\operatorname{QI}(\bar{E})\to\operatorname{QI}(E_{0}).

Proof.

Given any quasi-isometry ϕ:E¯E¯\phi\colon\bar{E}\to\bar{E} and xE0x\in E_{0} we have

d(ϕ(x),νϕ(x))=d(ϕ(x),f0(ϕ(x))dHaus(ϕ(E0),E0)d(\phi(x),\nu_{\phi}(x))=d(\phi(x),f_{0}(\phi(x))\leq d_{\mathrm{Haus}}(\phi(E_{0}),E_{0})

The right hand side is finite by Proposition 5.4, so the left hand side is bounded, independent of xx. From this, the triangle inequality, and the uniform metric properness of the inclusion of E0E_{0} into E¯\bar{E}, it easily follows that if ϕ\phi and ϕ\phi^{\prime} are bounded distance, then so are νϕ\nu_{\phi} and νϕ\nu_{\phi^{\prime}}. Therefore the assignment ϕνϕ\phi\mapsto\nu_{\phi} descends to a well-defined function 𝒜0:QI(E¯)QI(E0)\mathcal{A}_{0}\colon\operatorname{QI}(\bar{E})\to\operatorname{QI}(E_{0}).

To see that 𝒜0\mathcal{A}_{0} is a homomorphism, suppose ϕ,ϕ\phi,\phi^{\prime} are (K,K)(K,K)–quasi-isometries of E¯\bar{E}. Then from the inequality above, for all xE0x\in E_{0} we have

d(ϕϕ(x),ϕνϕ(x))Kd(ϕ(x),νϕ(x))+KKdHaus(ϕ(E0),E0)+K.d(\phi^{\prime}\circ\phi(x),\phi^{\prime}\circ\nu_{\phi}(x))\leq Kd(\phi(x),\nu_{\phi}(x))+K\leq Kd_{\mathrm{Haus}}(\phi(E_{0}),E_{0})+K.

The left-hand side is thus uniformly bounded, independent of xx. From this, the triangle inequality, and Proposition 5.4, it follows that dHaus(ϕϕ(E0),E0)d_{\mathrm{Haus}}(\phi^{\prime}\circ\phi(E_{0}),E_{0}) and dHaus(ϕνϕ(E0),E0)d_{\mathrm{Haus}}(\phi^{\prime}\circ\nu_{\phi}(E_{0}),E_{0}) are bounded by some constant r>0r>0. Then for all xE0x\in E_{0},

d(νϕϕ(x),νϕνϕ(x))=d(f0(ϕϕ(x)),f0(ϕνϕ(x)))erd(ϕϕ(x),ϕνϕ(x)).d(\nu_{\phi^{\prime}\circ\phi}(x),\nu_{\phi^{\prime}}\circ\nu_{\phi}(x))=d(f_{0}(\phi^{\prime}\circ\phi(x)),f_{0}(\phi^{\prime}\circ\nu_{\phi}(x)))\leq e^{r}d(\phi^{\prime}\circ\phi(x),\phi^{\prime}\circ\nu_{\phi}(x)).

Combining this with the previous inequality, we see that the quantity on the right, and hence the left, is uniformly bounded above, independent of xx. Therefore νϕϕ\nu_{\phi^{\prime}\circ\phi} and νϕνϕ\nu_{\phi^{\prime}}\circ\nu_{\phi} are bounded distance apart and 𝒜0\mathcal{A}_{0} is a homomorphism. ∎

5.3. From quasi-isometries to affine homeomorphisms

The flat metric qq on E0E_{0} determines an associated affine group Aff(E0)\rm Aff(E_{0}), and we observe that if ϕΓ\phi\in\Gamma is an element of the extension group (which is an isometry of E¯\bar{E}, and so also a quasi-isometry), we have νϕAff(E0)\nu_{\phi}\in\rm Aff(E_{0}). The next step in the proof of rigidity is the following.

Proposition 5.8.

For any quasi-isometry ϕ:E¯E¯\phi\colon\bar{E}\to\bar{E}, the quasi-isometry νϕ\nu_{\phi} is uniformly close to a unique element νϕaAff(E0)\nu_{\phi}^{a}\in\rm Aff(E_{0}).

The proof of the proposition will take place over the remainder of this subsection. Before getting to the proof, however, we note a useful corollary. Two quasi-isometries ϕ1,ϕ2\phi_{1},\phi_{2} that are a bounded distance apart have νϕ1\nu_{\phi_{1}} and νϕ2\nu_{\phi_{2}} a bounded distance apart, and so by the uniqueness νϕ1a=νϕ2a\nu_{\phi_{1}}^{a}=\nu_{\phi_{2}}^{a}. Thus we have the following.

Corollary 5.9.

The map [ϕ]νϕa[\phi]\mapsto\nu_{\phi}^{a} defines a homomorphism 𝒜:QI(E¯)Aff(E0)\mathcal{A}\colon\mathrm{QI}(\bar{E})\to\rm Aff(E_{0}). Moreover, the homomorphism 𝒜0:QI(E¯)QI(E0)\mathcal{A}_{0}\colon\operatorname{QI}(\bar{E})\to\operatorname{QI}(E_{0}) from Lemma 5.7 factors as the composition of 𝒜\mathcal{A} with the natural inclusion Aff(E0)QI(E0)\rm Aff(E_{0})\to\mathrm{QI}(E_{0}).

Fix a triangulation 𝔱\mathfrak{t} of X0X_{0} so that the vertex set is the set of cone points and all triangles are Euclidean triangles (that is, they are images of triangles by maps that are locally isometric and injective on the interior; see e.g. [DDLS21, Lemma 3.4]). Moreover, we assume that all saddle connections in some direction α0\alpha_{0} appear as edges of the triangulation; see Figure 9. Lift 𝔱\mathfrak{t} to a triangulation 𝔱~\widetilde{\mathfrak{t}} of E0E_{0}. By assumption, all saddle connections in E0E_{0} in direction α0\alpha_{0} are edges of 𝔱~\widetilde{\mathfrak{t}}, and the complement of the union of this subset is a union of all (interiors of) strips in direction α0\alpha_{0}.

1122112233334444
Figure 9. The surface from Figure 2 with a triangulation 𝔱\mathfrak{t} on the left. The edges of 𝔱\mathfrak{t} consist of all saddle connections in the horizontal direction (blue) together with some saddle connections (green), each contained in horizontal cylinder. A piece of the lifted triangulation 𝔱~\tilde{\mathfrak{t}} in the universal cover.
Lemma 5.10.

Given a quasi-isometry ϕ\phi, there is a biLipschitz homeomorphism νϕa:E0E0\nu_{\phi}^{a}\colon E_{0}\to E_{0} a bounded distance from νϕ\nu_{\phi} so that νϕa\nu_{\phi}^{a} restricts to an affine map on each triangle of 𝔱~\widetilde{\mathfrak{t}}. Moreover, if an edge δ\delta of 𝔱~\widetilde{\mathfrak{t}} has direction α𝒫\alpha\in\mathcal{P}, then νϕa(δ)\nu_{\phi}^{a}(\delta) has direction ϕ𝒫(α)\phi_{{\mathcal{P}}}(\alpha).

We will later prove that νϕa\nu_{\phi}^{a} is in fact globally affine, justifying the notation.

Proof.

Given ν=νϕ:E0E0\nu=\nu_{\phi}\colon E_{0}\to E_{0}, let ν:S1S1\partial\nu\colon S^{1}_{\infty}\to S^{1}_{\infty} be the restriction of the extension to the Gromov boundary S1S^{1}_{\infty} of E0E_{0}. (Recall that, since the flat metric (X0,q)(X_{0},q) on SS is biLipschitz to a hyperbolic metric, its universal cover E0E_{0} is quasi-isometric to the standard hyperbolic plane and is, in particular, Gromov hyperbolic.) The space 𝒢\mathcal{G} of (unordered) pairs of distinct points in S1S^{1}_{\infty} is precisely the space of endpoint-pairs at infinity of unoriented biinfinite geodesics (up to the equivalence relation of having finite Hausdorff distance). The map ν\partial\nu induces a map ν:𝒢𝒢\partial\nu_{*}\colon\mathcal{G}\to\mathcal{G}.

Let 𝒢𝒢\mathcal{G}^{*}\subset\mathcal{G} be the closure of the set of endpoint-pairs at infinity of non-singular geodesics (i.e. geodesics that miss every cone point). Observe that all geodesics in a given strip have the same pair of endpoints, and any geodesic with that pair of endpoints is contained in the strip. Given a strip, we are therefore justified in referring to the pair of endpoints of the strip.

It follows from the description of geodesics with endpoints in 𝒢\mathcal{G}^{*} (see [BL18, Proposition 2.4]) together with the Veech Dichotomy (see e.g. [MT02]), that for any {ξ,ζ}𝒢\{\xi,\zeta\}\in\mathcal{G}^{*}, either {ξ,ζ}\{\xi,\zeta\} are the endpoints of a strip, or endpoints of a geodesic meeting at most one cone point.

According to Proposition 5.4, for any strip A𝒮A\in\mathcal{S}, the strip ϕ𝒮(A)\phi_{\mathcal{S}}(A) has finite Hausdorff distance to ϕ(A)\phi(A), and hence it also has finite Hausdorff distance to ν(A)\nu(A). Since ϕ𝒮\phi_{\mathcal{S}} is a bijection, this means that the homeomorphism ν\partial\nu_{*} sends the dense subset of 𝒢\mathcal{G}^{*} consisting of endpoint of strips onto itself, hence ν(𝒢)=𝒢\nu_{*}(\mathcal{G}^{*})=\mathcal{G}^{*}. From this and [BL18, Proposition 4.1] (see also [DELS18, Proposition 11]), it follows that there is a bijection

ϕΣ0:Σ0Σ0\phi_{\Sigma_{0}}\colon\Sigma_{0}\to\Sigma_{0}

from the set of cone points Σ0\Sigma_{0} of E0E_{0} to itself with the following property. If γE0\gamma\subset E_{0} is a geodesic or strip containing xΣ0x\in\Sigma_{0} with endpoints {ξ,ζ}𝒢\{\xi,\zeta\}\in\mathcal{G}^{*}, then ν({ξ,ζ})\partial\nu_{*}(\{\xi,\zeta\}) are the endpoints of a geodesic containing ϕΣ0(x)\phi_{\Sigma_{0}}(x). Given xΣ0x\in\Sigma_{0} consider any two geodesics γ1\gamma_{1} and γ2\gamma_{2} with endpoints in 𝒢\mathcal{G}^{*} (not necessarily contained in strips) passing through xx making an angle at least π/2\pi/2 with each other. We note that ν(x)\nu(x) is contained in ν(γ1)\nu(\gamma_{1}) and ν(γ2)\nu(\gamma_{2}), and is thus some uniform distance r>0r>0 to both of their geodesic representatives. Since γ1\gamma_{1} and γ2\gamma_{2} meet at angle at least π/2\pi/2, the rr–neighborhoods of the geodesic representatives of ν(γ1)\nu(\gamma_{1}) and ν(γ2)\nu(\gamma_{2}) intersect in a uniformly bounded diameter set, which contains ϕΣ0(x)\phi_{\Sigma_{0}}(x). Therefore, ϕΣ0(x)\phi_{\Sigma_{0}}(x) is uniformly close to ν(x)\nu(x), for all xΣ0x\in\Sigma_{0}.

From the properties of ϕΣ0\phi_{\Sigma_{0}} described above, we see that if xΣ0x\in\Sigma_{0} is contained in a strip AA, then ϕΣ0(x)\phi_{\Sigma_{0}}(x) is contained in the strip ϕ𝒮(A)\phi_{\mathcal{S}}(A). For any saddle connection δ\delta in some direction α𝒫\alpha\in{\mathcal{P}} between a pair of points x,yΣ0x,y\in\Sigma_{0}, there is a unique pair of strips A,A0A,A_{0}, also in direction α\alpha, that contain δ\delta; see Figure 10. Since ϕΣ0(x),ϕΣ0(y)\phi_{\Sigma_{0}}(x),\phi_{\Sigma_{0}}(y) are contained in ϕ𝒮(A)\phi_{\mathcal{S}}(A) and ϕ𝒮(A0)\phi_{\mathcal{S}}(A_{0}), it follows that there is a unique saddle connection with endpoints ϕΣ0(x),ϕΣ0(y)\phi_{\Sigma_{0}}(x),\phi_{\Sigma_{0}}(y). For any strip AA the saddle connections whose union makes up one of its boundary components is determined by a collection of strips meeting AA in the given saddle connections. These saddle connections are ordered along this side and thus so are the corresponding strips ,A1,A0,A1,\ldots,A_{-1},A_{0},A_{1},\ldots. The endpoints of the strip AA and strips AnA_{n} appear in a particular order; see Figure 10. Considering the cyclic ordering of the endpoints of these strips (and those of AA) on S1S^{1}_{\infty}, and the fact that ν\partial\nu is a homeomorphism, it follows that ϕΣ0\phi_{\Sigma_{0}} maps the ordered set of cone points along each boundary component of the strip AA by an order preserving (or reversing) bijection to the ordered set of cone points along the boundary components of ϕ𝒮(A)\phi_{\mathcal{S}}(A).

A0A_{0}AAA  1A_{\,\,1}A1A_{1}δ\deltaxxyy
Figure 10. Strips AA and A0A_{0} determine the saddle connection δ\delta connecting points x,yx,y in the universal cover. The ordered set of saddle connections along one side of the strip AA are determined by an ordered set of strips ,A1,A0,A1,\ldots,A_{-1},A_{0},A_{1},\ldots, and the endpoints of all of these strips have a cyclic ordering around S1S^{1}_{\infty} as indicated.

We can now extend the map ϕΣ0\phi_{\Sigma_{0}} to a map νϕa:E0E0\nu_{\phi}^{a}\colon E_{0}\to E_{0} using 𝔱~\widetilde{\mathfrak{t}} as follows. First, recall that any edge of 𝔱~\widetilde{\mathfrak{t}} is a saddle connection δ\delta connecting two points x,yΣ0x,y\in\Sigma_{0}. By the previous paragraph, there is a saddle connection δ\delta^{\prime} connecting ϕΣ0(x)\phi_{\Sigma_{0}}(x) and ϕΣ0(y)\phi_{\Sigma_{0}}(y), and we define νϕa\nu_{\phi}^{a} on δ\delta so that it maps δ\delta by an affine map to δ\delta^{\prime} extending ϕΣ0\phi_{\Sigma_{0}} on the endpoints. This defines νϕa\nu_{\phi}^{a} on the 11–skeleton, 𝔱~1\widetilde{\mathfrak{t}}^{1}, and since ϕΣ0\phi_{\Sigma_{0}} is a bounded distance from ν|Σ0\nu|_{\Sigma_{0}}, it follows that νϕa|𝔱~1\nu_{\phi}^{a}|_{\widetilde{\mathfrak{t}}^{1}} is a bounded distance from ν|𝔱~1\nu|_{\widetilde{\mathfrak{t}}^{1}}.

By our assumptions on 𝔱~\widetilde{\mathfrak{t}}, there is a subset of the edges of 𝔱~\widetilde{\mathfrak{t}} whose union is precisely the union of boundaries of all strips in direction α0\alpha_{0}. The order preserving (or reversing) property described above for the cone points along the boundary of a strip, together with Proposition 5.4, implies that for any boundary component of any strip AA in direction α0\alpha_{0}, νϕa\nu_{\phi}^{a} restricted to its boundary components is a homeomorphism onto the boundary components of ϕ𝒮(A)\phi_{\mathcal{S}}(A). Furthermore, since the sides of any triangle of 𝔱~\widetilde{\mathfrak{t}} are contained in such a strip AA, the νϕa\nu_{\phi}^{a}–image of the sides are contained in ϕ𝒮(A)\phi_{\mathcal{S}}(A). We can now extend νϕa\nu_{\phi}^{a} over the triangles by the unique affine map extending the map on their sides.

Since disjoint strips map to disjoint strips, the map νϕa\nu_{\phi}^{a} is a homeomorphism. By construction, any edge in direction α\alpha is sent to an edge in direction ϕ𝒫(α)\phi_{{\mathcal{P}}}(\alpha). Since 𝔱~\widetilde{\mathfrak{t}} projects to 𝔱\mathfrak{t}, there are only finitely many directions that the sides of a triangle can lie in and so finitely many isometry types of triangles. Each of these finitely many isometry types maps by an affine map to only finitely many types of triangles in the image (because the direction of the images of sides are determined by ϕ𝒫\phi_{{\mathcal{P}}}), and therefore these affine maps are uniformly biLipschitz. Therefore, νϕa\nu_{\phi}^{a} is biLipschitz, completing the proof. ∎

To show that νϕa\nu_{\phi}^{a} is affine, we analyze the effect of using it to conjugate the action of π1S\pi_{1}S on E0E_{0}.

Lemma 5.11.

The action of π1S\pi_{1}S on E0E_{0} obtained by conjugating the isometric action by νϕa\nu_{\phi}^{a} is again an isometric action.

Before proving the lemma, we use it to prove the proposition.

Proof of Proposition 5.8.

By Lemma 5.11, Λ=νϕaπ1S(νϕa)1\Lambda=\nu_{\phi}^{a}\pi_{1}S(\nu_{\phi}^{a})^{-1} acts by isometries, and νϕa\nu_{\phi}^{a} descends to a homeomorphism μϕa:SE0/Λ\mu_{\phi}^{a}\colon S\to E_{0}/\Lambda and is biLipschitz with respect to descent to SS and E0/ΛE_{0}/\Lambda of qq. Since νϕa\nu_{\phi}^{a} and ν\nu are a bounded distance, they have the same boundary maps. Since (νϕa)=ν\partial(\nu_{\phi}^{a})_{*}=\partial\nu_{*} maps 𝒢\mathcal{G}^{*} to 𝒢\mathcal{G}^{*}, the Current Support Theorem of [DELS18] (and its proof) implies that the descent of μϕa:(S,q)(E0/Λ,q)\mu_{\phi}^{a}\colon(S,q)\to(E_{0}/\Lambda,q) is affine. Therefore νϕa\nu_{\phi}^{a} is an affine map which is a bounded distance from ν=νϕ\nu=\nu_{\phi}, as required.

Uniqueness follows from the fact that no two distinct affine maps are a bounded distance apart. ∎

Proof of Lemma 5.11.

We need to show that for all gπ1Sg\in\pi_{1}S, the map

νϕag(νϕa)1:E0E0\nu_{\phi}^{a}\circ g\circ(\nu_{\phi}^{a})^{-1}\colon E_{0}\to E_{0}

is an isometry. For this, fix a triangle τ\tau of 𝔱~\widetilde{\mathfrak{t}} and consider the restriction to νϕa(τ)\nu_{\phi}^{a}(\tau). Let α1,α2,α3𝒫\alpha_{1},\alpha_{2},\alpha_{3}\in{\mathcal{P}} be the directions of the sides. Setting αi=ϕ𝒫(αi)\alpha_{i}^{\prime}=\phi_{{\mathcal{P}}}(\alpha_{i}), for i=1,2,3i=1,2,3, Lemma 5.10 implies that the directions of the sides of νϕa(τ)\nu_{\phi}^{a}(\tau) are α1,α2,α3\alpha_{1}^{\prime},\alpha_{2}^{\prime},\alpha_{3}^{\prime}. The action of π1S\pi_{1}S on E0E_{0} is by isometries, but it also preserves parallelism (i.e. each element induces the identity on 1(q)\mathbb{P}^{1}(q)). Therefore, for any gπ1Sg\in\pi_{1}S, the directions of the sides of g(τ)g(\tau) are also α1,α2,α3\alpha_{1},\alpha_{2},\alpha_{3}, and by Lemma 5.10 again, it follows that the sides of νϕa(g(τ))\nu_{\phi}^{a}(g(\tau)) are α1,α2,α3\alpha_{1}^{\prime},\alpha_{2}^{\prime},\alpha_{3}^{\prime}.

For any gπ1Sg\in\pi_{1}S, since νϕa\nu_{\phi}^{a} is affine on τ\tau, the composition νϕag(νϕa)1\nu_{\phi}^{a}\circ g\circ(\nu_{\phi}^{a})^{-1} is also affine on νϕa(τ)\nu_{\phi}^{a}(\tau). On the other hand, it also preserves the directions of the sides, α1,α2,α3\alpha_{1}^{\prime},\alpha_{2}^{\prime},\alpha_{3}^{\prime}. Therefore, the restriction of νϕag(νϕa)1\nu_{\phi}^{a}\circ g\circ(\nu_{\phi}^{a})^{-1} is a Euclidean similarity. Triangles of νϕa(𝔱~)\nu_{\phi}^{a}(\widetilde{\mathfrak{t}}) that share a side are scaled by the same factor by the similarity νϕag(νϕa)1\nu_{\phi}^{a}\circ g\circ(\nu_{\phi}^{a})^{-1} in each triangle (since this is the scaling factor on the shared side). Therefore, the similarities agree along edges, and hence νϕag(hϕa)1\nu_{\phi}^{a}\circ g\circ(h_{\phi}^{a})^{-1} defines a global similarity of E0E_{0}.

So, the action of π1S\pi_{1}S on E0E_{0} obtained by conjugating by νϕa\nu_{\phi}^{a} is an action by similarities. To see that the action is by isometries, suppose that for some element gπ1Sg\in\pi_{1}S, the similarity g0=νϕag(νϕa)1g_{0}=\nu_{\phi}^{a}\circ g\circ(\nu_{\phi}^{a})^{-1} scales the metric some number λ1\lambda\neq 1. Taking the inverse if necessary, we can assume λ<1\lambda<1. Fix any xE0x\in E_{0} and observe that dq(g0(x),g02(x))=λdq(x,g0(x))d_{q}(g_{0}(x),g_{0}^{2}(x))=\lambda d_{q}(x,g_{0}(x)), where dqd_{q} is the distance function on E0E_{0} determined by qq. Iterating this, it follows that

dq(x,g0n(x))k=1ndq(g0k1(x),g0k(x))=k=1nλk1dq(x,g0(x))dq(x,g0(x))k=1λk1.d_{q}(x,g_{0}^{n}(x))\leq\!\sum_{k=1}^{n}d_{q}(g_{0}^{k-1}(x),g_{0}^{k}(x))\!=\!\!\sum_{k=1}^{n}\lambda^{k-1}d_{q}(x,g_{0}(x))\leq\!d_{q}(x,g_{0}(x))\sum_{k=1}^{\infty}\lambda^{k-1}.

Since the right-hand side is a convergent geometric series, it follows that {g0n(x)}n=1\{g_{0}^{n}(x)\}_{n=1}^{\infty} is a Cauchy sequence. On the other hand, this sequence exits every compact set (since g0g_{0} is an infinite order element of π1S\pi_{1}S), and since qq is a complete metric on E0E_{0}, thus we obtain a contradiction. Therefore, the conjugation action of π1S\pi_{1}S is by isometries. ∎

5.4. Injectivity of 𝒜\mathcal{A}.

Our next goal is to prove that 𝒜:QI(E¯)Aff(E0)\mathcal{A}\colon\operatorname{QI}(\bar{E})\to\rm Aff(E_{0}), the homomorphism from Corollary 5.9, is injective.

In preparation, it will be useful to have the following general fact about quasiisometries of hyperbolic spaces, whose proof we sketch for convenience of the reader:

Lemma 5.12.

For each K,C,δK,C,\delta there exists RR so that the following holds. Suppose that ZZ is δ\delta–hyperbolic and that each zZz\in Z lies within δ\delta of all three sides of a nondegenerate ideal geodesic triangle. Let f:ZZf\colon Z\to Z be a (K,C)(K,C)-quasi-isometry that lies within finite distance of the identity. Then ff lies within distance RR of the identity.

Proof.

Since ff is within bounded distance of the identity, its extension f:ZZ\partial f\colon\partial Z\to\partial Z is the identity. Hence if zZz\in Z and Δ\Delta is an ideal geodesic triangle as in the statement, then f(Δ)f(\Delta) is a (K,C)(K,C)-quasigeodesic ideal triangle with the same endpoints as Δ\Delta. By the Morse lemma, there is a constant κ=κ(K,C,δ)>0\kappa=\kappa(K,C,\delta)>0 such that f(z)f(z) lies within κ\kappa of the three quasi-geodesic sides of f(Δ)f(\Delta), and these sides in turn lie within κ\kappa of the sides of Δ\Delta. Thus zz and f(z)f(z) both lie within 2κ+δ2\kappa+\delta of all three sides of Δ\Delta. Since the set of points within 2κ+δ2\kappa+\delta of all three sides of a nondegenerate geodesic triangle in a δ\delta–hyperbolic space has diameter bounded in terms of κ\kappa and δ\delta, we see that dZ(z,f(z))d_{Z}(z,f(z)) is bounded solely in terms of δ,K,C\delta,K,C, as required. ∎

With this fact in hand, we can now prove:

Proposition 5.13.

Let ϕ:E¯E¯\phi\colon\bar{E}\to\bar{E} be a quasi-isometry with νϕa=idE0\nu_{\phi}^{a}=id_{E_{0}}. There is a constant C=C(ϕ)>0C^{\prime}=C^{\prime}(\phi)>0 such that for all xE¯x\in\bar{E}, d(x,ϕ(x))Cd(x,\phi(x))\leq C^{\prime}. Consequently, 𝒜:QI(E¯)Aff(E0)\mathcal{A}\colon\operatorname{QI}(\bar{E})\to\rm Aff(E_{0}) is injective.

Proof.

We first claim that for any XD¯X\in\bar{D}, ϕ(EX)\phi(E_{X}) lies within the CC^{\prime\prime}–neighborhood of EXE_{X}, where C=C(ϕ)>0C^{\prime\prime}=C^{\prime\prime}(\phi)>0. To see this, observe that since νϕa\nu_{\phi}^{a} is the identity and is bounded distance from νϕ\nu_{\phi}, it follows that ϕ|E0\phi|_{E_{0}} is within bounded distance of the inclusion of E0E_{0} in E¯\bar{E}. Proposition 5.4 then implies that dHaus(A,ϕS(A))<+d_{\mathrm{Haus}}(A,\phi_{S}(A))<+\infty for each strip A𝒮A\in\mathcal{S}. Since strips that lie within finite Hausdorff distance coincide, we have ϕS(A)=A\phi_{S}(A)=A. Combining this fact with Proposition 5.4 it follows that ϕ𝒫(α(A))=α(ϕ𝒮(A))=α(A)\phi_{\mathcal{P}}(\alpha(A))=\alpha(\phi_{\mathcal{S}}(A))=\alpha(A). Hence for each α\alpha, we have that ϕ(α)\phi(\partial\mathcal{B}_{\alpha}) lies within Hausdorff distance CC of α=ϕ𝒫(α)\partial\mathcal{B}_{\alpha}=\partial\mathcal{B}_{\phi_{\mathcal{P}}(\alpha)}, for CC as in Proposition 5.4.

Now let XD¯X\in\bar{D} be any point and choose distinct α,α𝒫\alpha,\alpha^{\prime}\in\mathcal{P} so that XX is contained in the coarse intersection of Bα\partial B_{\alpha} and Bα\partial B_{\alpha^{\prime}}, implying that EXE_{X} lies in the coarse intersection of α\partial\mathcal{B}_{\alpha} and α\partial\mathcal{B}_{\alpha^{\prime}}. By the coarse preservation of the α\partial\mathcal{B}_{\alpha} in the previous paragraph, the coarse intersection of ϕ(α)\phi(\partial\mathcal{B}_{\alpha}) and ϕ(α)\phi(\partial\mathcal{B}_{\alpha^{\prime}}) is within Hausdorff distance CC of the coarse intersection of α\partial\mathcal{B}_{\alpha} and α\partial\mathcal{B}_{\alpha^{\prime}}, and hence EXE_{X} and ϕ(EX)\phi(E_{X}) are within uniform Hausdorff distance. This proves the claim.

Since fX:E¯EXf_{X}\colon\bar{E}\to E_{X} is eCe^{C^{\prime\prime}}-bi-Lipschitz when restricted to fibers in the CC^{\prime\prime}-neighborhood of EXE_{X}, the claim implies that νϕX=fXϕ:EXEX\nu_{\phi}^{X}=f_{X}\circ\phi\colon E_{X}\to E_{X} is a quasi-isometry with constants depending only on ϕ\phi and not XX. Moreover, since each EXE_{X} lies within finite (but not necessarily bounded) Hausdorff distance of E0E_{0}, the fact that ϕ|E0\phi|_{E_{0}} lies within finite distance of the inclusion E0E¯E_{0}\hookrightarrow\bar{E} implies that νϕX\nu_{\phi}^{X} lies within finite distance of the identity EXEXE_{X}\to E_{X}. Since each EXE_{X} is uniformly quasiisometric to 2\mathbb{H}^{2}, it follows that νϕX:EXEX\nu_{\phi}^{X}\colon E_{X}\to E_{X} satisfies the assumptions of Lemma 5.12. We conclude that νϕX\nu_{\phi}^{X} is within uniformly bounded distance of the identity for each XD¯X\in\bar{D}. Since d(νϕX(x),ϕ(x))Cd(\nu_{\phi}^{X}(x),\phi(x))\leq C^{\prime\prime}, it follows that d(x,ϕ(x))d(x,\phi(x)) is uniformly bounded, independent of xx. This proves the first statement of the proposition.

If 𝒜(ϕ)\mathcal{A}(\phi) is the identity for some ϕQI(E¯)\phi\in\operatorname{QI}(\bar{E}), then by the first part of the proposition, ϕ\phi is a bounded distance from the identity. Therefore, ϕ\phi and the identity represent the same class, and 𝒜\mathcal{A} is injective. This completes the proof. ∎

5.5. From affine homeomorphisms to isometries

Next we will choose a particular allowable truncation and construct a homomorphism Aff(E0)Isomfib(E¯)\rm Aff(E_{0})\to\operatorname{Isom}_{{\mathrm{fib}}}(\bar{E}), that we will eventually show is an isomorphism. We first construct such a homomorphism to the fiber-preserving isometry group of the space EE, which avoids the issue of choosing the truncation.

Lemma 5.14.

For any νAff(E0)\nu\in\rm Aff(E_{0}), there exists a isometry ϕ=ϕνIsomfib(E)\phi=\phi_{\nu}\in\mathrm{Isom_{fib}}(E) such that f0ϕν|E0=νf_{0}\circ\phi_{\nu}|_{E_{0}}=\nu. Moreover, this assignment νϕν\nu\mapsto\phi_{\nu} defines an injective homomorphism Aff(E0)Isomfib(E)\rm Aff(E_{0})\to\mathrm{Isom_{fib}}(E).

Proof.

Recall from §2.1 that the projective tangent space at any non-cone point of E0E_{0} is denoted 1(q)\mathbb{P}^{1}(q) and is canonically identified with D\partial D. The derivative of ν:E0E0\nu\colon E_{0}\to E_{0} (which may reverse orientations) is a well-defined projective transformation dνPGL(1(q))d\nu\in\rm PGL(\mathbb{P}^{1}(q)) which, using the preferred coordinates on q=q0q=q_{0} with distinguished vertical and horizontal directions, we canonically identify with PGL2()\rm PGL_{2}(\mathbb{R}). The Teichmüller disk DD is the orbit of qq under the SL2()\rm SL_{2}(\mathbb{R}) action and is identified with 2=SL2()/SO(2)\mathbb{H}^{2}=\rm SL_{2}(\mathbb{R})/SO(2) (see e.g. [DDLS21, §2.8]). As PGL2()\rm PGL_{2}(\mathbb{R}) acts isometrically on 2\mathbb{H}^{2}, we thus obtain an isometry Φ=dν:DD\Phi=d\nu\colon D\to D whose induced map Φ\partial\Phi of the circle at infinity D\partial D agrees with the derivative dνd\nu under the canonical identification D1(q)\partial D\cong\mathbb{P}^{1}(q). In particular, setting X=Φ(X0)X=\Phi(X_{0}), the geodesic ray in DD emanating from X0X_{0} and asymptotic to ξ1(q)\xi\in\mathbb{P}^{1}(q) is sent to the geodesic ray emanating from XX asymptotic to dν(ξ)d\nu(\xi).

We claim the map ϕ0=fX,X0ν:E0EX\phi_{0}=f_{X,X_{0}}\circ\nu\colon E_{0}\to E_{X} is an isometry of fibers. Indeed, any pair ξ,ξ1(q)\xi,\xi^{\perp}\in\mathbb{P}^{1}(q) of orthogonal directions on E0E_{0} are the endpoints of a geodesic ρ\rho in DD containing X0X_{0}. Since Φ\Phi is an isometry with Φ=dν\partial\Phi=d\nu, we have that X=Φ(X0)X=\Phi(X_{0}) lies on the geodesic Φ(ρ)\Phi(\rho) from dν(ξ)d\nu(\xi^{\perp}) to dν(ξ)d\nu(\xi); that is, dν(ξ),dν(ξ)d\nu(\xi),d\nu(\xi^{\perp}) are orthogonal on XX. But since 1(q)\mathbb{P}^{1}(q) and 1(qX)\mathbb{P}^{1}(q_{X}) are canonically identified by the Teichmüller map fX,X0f_{X,X_{0}} (see e.g. [DDLS21, §2.8]), this means ϕ0\phi_{0} is an affine map whose derivative dϕ0=dνd\phi_{0}=d\nu preserves orthogonality of lines; hence ϕ0\phi_{0} is an isometry as claimed.

Now we define ϕ=ϕν:EE\phi=\phi_{\nu}\colon E\to E by the formula:

ϕ(x)=fΦ(π(x)),X0νf0(x).\phi(x)=f_{\Phi(\pi(x)),X_{0}}\circ\nu\circ f_{0}(x).

In words, this maps the fiber over a point YY to the fiber over Φ(Y)\Phi(Y), and the horizontal disk DxD_{x}, for xE0x\in E_{0}, to Dν(x)D_{\nu(x)}. The restriction ϕ|Dx:DxDν(x)\phi|_{D_{x}}\colon D_{x}\to D_{\nu(x)} is an isometry since it covers Φ\Phi. To prove that ϕ\phi is an isometry, it therefore suffices to show that ϕ|EY:EYEΦ(Y)\phi|_{E_{Y}}\colon E_{Y}\to E_{\Phi(Y)} is an isometry for any YDY\in D.

Fix any YDY\in D. For Y=X0Y=X_{0}, we have already seen that ϕ|E0\phi|_{E_{0}} is the isometry ϕ0:E0EX\phi_{0}\colon E_{0}\to E_{X}. If YXY\neq X, there exist unique orthogonal directions α,α(q)\alpha,\alpha^{\perp}\in\mathbb{P}(q) and t>0t>0, so that X0X_{0} and YY both lie on the the geodesic from α\alpha^{\perp} to α\alpha in DD and YY lies distance tt from X0X_{0} in the direction of α\alpha. This means that fY,X0:E0EYf_{Y,X_{0}}\colon E_{0}\to E_{Y} contracts in direction α\alpha by ete^{-t} and stretches in direction α\alpha^{\perp} by ete^{t}. The image Φ(Y)\Phi(Y) lies along the geodesic from dϕ0(α)d\phi_{0}(\alpha^{\perp}) to dϕ0(α)d\phi_{0}(\alpha) at distance tt from Φ(X0)=X\Phi(X_{0})=X; therefore fΦ(Y),X:EXEΦ(Y)f_{\Phi(Y),X}\colon E_{X}\to E_{\Phi(Y)} contracts by ete^{-t} in direction dϕ0(α)d\phi_{0}(\alpha) and stretches by ete^{t} in direction dϕ0(α)d\phi_{0}(\alpha^{\perp}). The restriction ϕ|Y:YΦ(Y)\phi|_{Y}\colon Y\to\Phi(Y) is given by fΦ(Y),Xϕ0fE0,Yf_{\Phi(Y),X}\circ\phi_{0}\circ f_{E_{0},Y}. Since ϕ0\phi_{0} sends αdϕ0(α)\alpha\mapsto d\phi_{0}(\alpha) and αdϕ0(α)\alpha^{\perp}\mapsto d\phi_{0}(\alpha^{\perp}), the description above shows that ϕ|Y\phi|_{Y} is an isometry. Therefore ϕ\phi is an isometry, as required.

To see that νϕν\nu\mapsto\phi_{\nu} is a homomorphism, note that by construction Φν\Phi_{\nu} is the unique isometry of DD for which Φν=dν\partial\Phi_{\nu}=d\nu. Thus the chain rule implies Φνg=ΦνΦg\Phi_{\nu\circ g}=\Phi_{\nu}\circ\Phi_{g} is the unique isometry whose action on D\partial D agrees with d(νg)=dνdgd(\nu\circ g)=d\nu\circ dg. For any xEx\in E we have π(ϕg(x))=Φg(π(x))\pi(\phi_{g}(x))=\Phi_{g}(\pi(x)) and hence by construction

ϕνϕg(x)\displaystyle\phi_{\nu}\circ\phi_{g}(x) =fΦν(π(ϕg(x))),X0νf0(fΦg(x),X0gf0(x))\displaystyle=f_{\Phi_{\nu}(\pi(\phi_{g}(x))),X_{0}}\circ\nu\circ f_{0}\big{(}f_{\Phi_{g}(x),X_{0}}\circ g\circ f_{0}(x)\big{)}
=fΦν(Φg(π(x))),X0νfX0,Φg(x)fΦg(x),X0gf0(x))\displaystyle=f_{\Phi_{\nu}(\Phi_{g}(\pi(x))),X_{0}}\circ\nu\circ f_{X_{0},\Phi_{g}(x)}\circ f_{\Phi_{g}(x),X_{0}}\circ g\circ f_{0}(x))
=fΦνg(π(x)),X0(νg)f0(x)=ϕνg(x)\displaystyle=f_{\Phi_{\nu\circ g}(\pi(x)),X_{0}}\circ(\nu\circ g)\circ f_{0}(x)=\phi_{\nu\circ g}(x)

as needed. Finally, if ϕν=idE\phi_{\nu}=\mathrm{id}_{E} then clearly X=Φ(X0)=X0X=\Phi(X_{0})=X_{0}. Since ϕ0=ϕν|E0\phi_{0}=\phi_{\nu}|_{E_{0}} by construction, we conclude that

idE0=ϕν|E0=ϕ0=fX0,X0ν=ν.\mathrm{id}_{E_{0}}=\phi_{\nu}|_{E_{0}}=\phi_{0}=f_{X_{0},X_{0}}\circ\nu=\nu.

Hence ν\nu is the identity affine map, showing that νϕν\nu\mapsto\phi_{\nu} is injective. ∎

Lemma 5.15.

The subgroup Γ<Isomfib(E)\Gamma<\operatorname{Isom}_{{\mathrm{fib}}}(E) has finite index.

Proof.

By [DDLS21, Proposition 5.5], Isomfib(E)\mathrm{Isom_{fib}}(E) acts properly discontinuously on EE. Therefore E/Isomfib(E)E/\operatorname{Isom}_{\mathrm{fib}}(E) is a topological orbifold with well-defined, positive Riemannian volume. The index of Γ\Gamma in Isomfib(E)\operatorname{Isom}_{\mathrm{fib}}(E) is the degree of the orbifold cover E/ΓE/Isomfib(E)E/\Gamma\to E/\operatorname{Isom}_{\mathrm{fib}}(E) and equals the ratio of the respective volumes. As E/ΓE/\Gamma has finite volume, since the quotient D/GD/G has finite area and the fibers EX/π1SE_{X}/\pi_{1}S all have equal, finite area, we conclude that Γ\Gamma indeed has finite index. ∎

Lemma 5.16.

There is an allowable truncation E¯\bar{E} that is Isomfib(E)\operatorname{Isom}_{\mathrm{fib}}(E)–invariant and for which restricting to E¯\bar{E} induces an injection Isomfib(E)Isomfib(E¯)\operatorname{Isom}_{\mathrm{fib}}(E)\to\operatorname{Isom}_{\mathrm{fib}}(\bar{E}).

Remark 5.17.

Every fiber-preserving isometry of E¯\bar{E} uniquely extends to one of EE (e.g. by following the proof of Lemma 5.14) and thus Isomfib(E)Isomfib(E¯)\operatorname{Isom}_{\mathrm{fib}}(E)\to\operatorname{Isom}_{\mathrm{fib}}(\bar{E}) is in fact an isomorphism.

Proof.

There is a natural map Isomfib(E)Isom(D)\operatorname{Isom}_{\mathrm{fib}}(E)\to\operatorname{Isom}(D) that sends Γ\Gamma onto GG. Hence, by the previous lemma, the image GG^{*} of Isomfib(E)\operatorname{Isom}_{\mathrm{fib}}(E) under this map contains GG with finite index. Therefore GG^{*} acts properly discontinuously on DD and we may choose a collection {Bα}α𝒫\{B_{\alpha}\}_{\alpha\in\mathcal{P}} of 11–separated horoballs as in §2.1 that is GG^{*}–invariant. If E¯\bar{E} denotes the corresponding truncation of EE, it follows that every element of Isomfib(E)\operatorname{Isom}_{\mathrm{fib}}(E) preserves E¯\bar{E}. The map Isomfib(E)Isomfib(E¯)\mathrm{Isom_{fib}}(E)\to\mathrm{Isom_{fib}}(\bar{E}) given by restricting ϕϕ|E¯\phi\mapsto\phi|_{\bar{E}} is injective by [DDLS21, Corollary 5.6] since if ϕ|E¯\phi|_{\bar{E}} is the identity, then ϕ\phi must be the identity on each Teichmüller disk DxD_{x} and fiber EXE¯E_{X}\subset\bar{E}. ∎

Choosing such an allowable truncation E¯\bar{E}, Lemmas 5.14 and 5.16 now give an injective homomorphism Ψ:Aff(E0)Isomfib(E¯)\Psi\colon\rm Aff(E_{0})\to\operatorname{Isom}_{\mathrm{fib}}(\bar{E}) given by Ψ(ν)=ϕν|E¯\Psi(\nu)=\phi_{\nu}|_{\bar{E}}.

Lemma 5.18.

For any νAff(E0)\nu\in\rm Aff(E_{0}), 𝒜(Ψ(ν))=ν\mathcal{A}(\Psi(\nu))=\nu, where we have identified Ψ(ν)\Psi(\nu) with its image in QI(E¯)\operatorname{QI}(\bar{E}) from the homomorphism Isomfib(E¯)QI(E¯)\operatorname{Isom}_{\mathrm{fib}}(\bar{E})\to\operatorname{QI}(\bar{E}).

Proof.

The construction of 𝒜\mathcal{A} in Corollary 5.9 sends the (quasi-)isometry Ψ(ν)=ϕν|E¯:E¯E¯\Psi(\nu)=\phi_{\nu}|_{\bar{E}}\colon\bar{E}\to\bar{E} to the the unique affine homeomorphism of E0E_{0} that is uniformly close to the map f0ϕν|E0:E0E0f_{0}\circ\phi_{\nu}|_{E_{0}}\colon E_{0}\to E_{0}. But by the construction of ϕν\phi_{\nu} in Lemma 5.14, f0ϕν|E0f_{0}\circ\phi_{\nu}|_{E_{0}} is affine itself and equal to ν\nu. Thus evidently 𝒜(Ψ(ν))=ν\mathcal{A}(\Psi(\nu))=\nu as claimed. ∎

Lemma 5.19.

For any ϕIsomfib(E¯)=Isomfib(E)\phi\in\operatorname{Isom}_{\mathrm{fib}}(\bar{E})=\operatorname{Isom}_{\mathrm{fib}}(E), we have Ψ(𝒜(ϕ))=ϕ\Psi(\mathcal{A}(\phi))=\phi. In particular, the natural maps Isomfib(E¯)Isom(E¯)QI(E¯)\mathrm{Isom_{fib}}(\bar{E})\to\operatorname{Isom}(\bar{E})\to\operatorname{QI}(\bar{E}) are both injective.

Proof.

By construction ν=νϕa=𝒜(ϕ)\nu=\nu_{\phi}^{a}=\mathcal{A}(\phi) is the unique affine homeomorphism bounded distance from f0ϕ|E0f_{0}\circ\phi|_{E_{0}}. As this map is itself affine, we have ν=f0ϕ|E0\nu=f_{0}\circ\phi|_{E_{0}}. The isometry Φ:DD\Phi\colon D\to D in the construction of Ψ(ν)\Psi(\nu) is then just the descent of ϕ\phi to DD. Further, for any X,YDX,Y\in D we have ϕ|XfX,Y=fΦ(X),Φ(Y)ϕ|EY\phi|_{X}\circ f_{X,Y}=f_{\Phi(X),\Phi(Y)}\circ\phi|_{E_{Y}}, since if XX lies at distance t>0t>0 from YY along the geodesic from α\alpha^{\perp} to α\alpha, then both maps send (α,α)(Φ(α),Φ(α))(\alpha^{\perp},\alpha)\mapsto(\partial\Phi(\alpha^{\perp}),\partial\Phi(\alpha)) while contracting the first by ete^{-t} and expanding the second by ete^{t}, hence they are the same affine map EYEΦ(X)E_{Y}\to E_{\Phi(X)}. It follows that the restriction Ψ(ν)|EY:EYEΦ(Y)\Psi(\nu)|_{E_{Y}}\colon E_{Y}\to E_{\Phi(Y)} is then the composition

Ψ(ν)|EY\displaystyle\Psi(\nu)|_{E_{Y}} =fΦ(Y),X0(f0ϕ|E0)f0|EY=fΦ(Y),Φ(X0)ϕ|E0fX0,Y\displaystyle=f_{\Phi(Y),X_{0}}\circ(f_{0}\circ\phi|_{E_{0}})\circ f_{0}|_{E_{Y}}=f_{\Phi(Y),\Phi(X_{0})}\circ\phi|_{E_{0}}\circ f_{X_{0},Y}
=fΦ(Y),Φ(X0)fΦ(X0),Φ(Y)ϕ|EY=ϕ|EY.\displaystyle=f_{\Phi(Y),\Phi(X_{0})}\circ f_{\Phi(X_{0}),\Phi(Y)}\circ\phi|_{E_{Y}}=\phi|_{E_{Y}}.

Since this holds for each YY, we conclude Ψ(ν)=ϕ\Psi(\nu)=\phi as claimed.

It follows that Isomfib(E¯)QI(E¯)\mathrm{Isom_{fib}}(\bar{E})\to\operatorname{QI}(\bar{E}) is injective, since if [ϕ][\phi] is the identity in QI(E¯)\operatorname{QI}(\bar{E}), meaning ϕ\phi is finite distance from the identity, then 𝒜(ϕ)=𝒜([ϕ])\mathcal{A}(\phi)=\mathcal{A}([\phi]) and consequently ϕ=Ψ(𝒜(ϕ))\phi=\Psi(\mathcal{A}(\phi)) are both the identity. Finally, Isom(E¯)QI(E¯)\operatorname{Isom}(\bar{E})\to\operatorname{QI}(\bar{E}) is injective since we have Isomfib(E¯)=Isom(E¯)\mathrm{Isom_{fib}}(\bar{E})=\operatorname{Isom}(\bar{E}) by [DDLS21, Corollary 5.4]. ∎

5.6. Rigidity

We are now ready to complete the proof of Theorem 1.7:

Proof of Theorem 1.7.

By Lemma 5.18, the composition

Aff(E0)ΨIsomfib(E¯)Isom(E¯)QI(E¯)QI(Γ)𝒜Aff(E0)\rm Aff(E_{0})\stackrel{{\scriptstyle\Psi}}{{\to}}\operatorname{Isom}_{\mathrm{fib}}(\bar{E})\to\operatorname{Isom}(\bar{E})\to\operatorname{QI}(\bar{E})\cong\operatorname{QI}(\Gamma)\stackrel{{\scriptstyle\mathcal{A}}}{{\to}}\rm Aff(E_{0})

is the identity. Hence the first map Ψ\Psi is injective, and the remaining maps are injective by Lemma 5.19 and Proposition 5.13 . It follows that each map above is an isomorphism, as claimed. The fact that Γ\Gamma has finite index in Isom(E¯)QI(Γ)\operatorname{Isom}(\bar{E})\cong\operatorname{QI}(\Gamma) thus follows from Lemma 5.15. ∎

Standard techniques (see, for example [Sch95, §10.4]) now imply the following:

Corollary 5.20.

If HH is any finitely generated group quasi-isometric to Γ\Gamma, then HH and Γ\Gamma are weakly commensurable, meaning HH has a finite normal subgroup NN so that H/NH/N and Γ\Gamma contain finite index subgroups that are isomorphic.

This proof requires one more lemma.

Lemma 5.21.

For every KK there exists RR^{\prime} such that if ϕ:E¯E¯\phi\colon\bar{E}\to\bar{E} is a (K,K)(K,K)-quasi-isometry that lies within finite distance of the identity, then ϕ\phi lies within distance RR^{\prime} of the identity, meaning d(x,ϕ(x))Rd(x,\phi(x))\leq R^{\prime} for all xE¯x\in\bar{E}.

Proof.

First define a map ϕ¯:D¯D¯\bar{\phi}\colon\bar{D}\to\bar{D} by setting ϕ¯(X)=Y\bar{\phi}(X)=Y, where YY is the point provided by Proposition 5.4 such that dHaus(ϕ(EX),EY)Cd_{\mathrm{Haus}}(\phi(E_{X}),E_{Y})\leq C. Since dD¯(X,Y)=dE¯(EX,EY)d_{\bar{D}}(X,Y)=d_{\bar{E}}(E_{X},E_{Y}) for all X,YX,Y in D¯\bar{D}, we see that ϕ¯\bar{\phi} is a quasi-isometry with constants depending only on KK. It also lies within finite distance of the identity, as it inherits this property from ϕ\phi; thus applying Lemma 5.12 to Z=D¯Z=\bar{D} implies that ϕ¯\bar{\phi} lies within uniformly finite distance of the identity. That is, there exists RR depending only on KK so that dHaus(EX,Φ(EX))Rd_{\mathrm{Haus}}(E_{X},\Phi(E_{X}))\leq R for all XD¯X\in\bar{D}. Hence, Lemma 5.6 implies that for each map νϕX=fXϕ|EX\nu_{\phi}^{X}=f_{X}\circ\phi|_{E_{X}} is a (K,K)(K^{\prime},K^{\prime})–quasi-isometry for some KK^{\prime} depending only on KK. Again by Lemma 5.12, this time with Z=EXZ=E_{X}, we see that each νϕX\nu_{\phi}^{X} moves points uniformly bounded distance, and therefore ϕ|EX\phi|_{E_{X}} lies within uniform distance of the inclusion of EXE_{X} in E¯\bar{E}. Since this holds for all XX, we have that ϕ\phi lies within uniform distance of the identity, as required. ∎

Proof of Corollary 5.20.

If HH is quasi-isometric to Γ\Gamma, there is a quasi-isometry μ:HE¯\mu\colon H\to\bar{E} with a quasi-inverse μ1:E¯H\mu^{-1}\colon\bar{E}\to H. Left multiplication by hHh\in H gives an isometry Lh:HHL_{h}\colon H\to H. In this way, for each hHh\in H we obtain a quasi-isometry (h)=μLhμ1\mathcal{B}(h)=\mu\circ L_{h}\circ\mu^{-1} of E¯\bar{E} with uniformly bounded constants. Let us also set (h)=Ψ(𝒜((h)))Isomfib(E¯)=Isom(E¯)\mathcal{B}^{\prime}(h)=\Psi(\mathcal{A}(\mathcal{B}(h)))\in\operatorname{Isom}_{\mathrm{fib}}(\bar{E})=\operatorname{Isom}(\bar{E}), which is the unique isometry of E¯\bar{E} at finite distance from (h)\mathcal{B}(h). Since the quasi-isometry constants of (h)\mathcal{B}(h) are uniform, depending only on μ\mu, it follows from Lemma 5.21 that there is a constant RR^{\prime} so that d((h)(x),(h)(x))Rd(\mathcal{B}(h)(x),\mathcal{B}^{\prime}(h)(x))\leq R^{\prime} for all xE¯x\in\bar{E} and hHh\in H.

We now claim the homomorphism :HIsom(E¯)\mathcal{B}^{\prime}\colon H\to\operatorname{Isom}(\bar{E}) has finite kernel and cokernel. Indeed, if (h)=IdE¯\mathcal{B}^{\prime}(h)=\mathrm{Id}_{\bar{E}} the above implies (h)\mathcal{B}(h) moves μ(e)\mu(e) (and in fact all points) distance at most RR^{\prime}. But this means LhL_{h} moves the identity eHe\in H uniformly bounded distance, and there are only finitely many such elements of HH. To prove \mathcal{B}^{\prime} has finite cokernel it suffices, as in Lemma 5.15, to show E¯/(H)\bar{E}/\mathcal{B}^{\prime}(H) has finite volume or, better yet, finite diameter. For this, given x,yE¯x,y\in\bar{E} we must find hh so that d((h)(x),y)d(\mathcal{B}^{\prime}(h)(x),y) is uniformly bounded. This is equivalent to bounding d((h)(x),y)=d(μ(hμ1(x)),y)d(\mathcal{B}(h)(x),y)=d(\mu(h\cdot\mu^{-1}(x)),y), which is coarsely dH(hμ1(x),μ1(y))d_{H}(h\cdot\mu^{-1}(x),\mu^{-1}(y)). Since HH acts transitively on itself, this is clearly possible.

We now see that H/ker()H/\ker(\mathcal{B}^{\prime}) and Γ\Gamma are both realized as finite index subgroups of Isom(E¯)\operatorname{Isom}(\bar{E}) and hence that their intersection has finite index in both. ∎

5.7. An alternative proof of quasi-isometric rigidity

Here we sketch another proof of Theorem 1.2, following an approach described by Mosher in [Mos06]; we refer the reader to that paper for a more detailed discussion.

First, we require some additional definitions. Consider the maximal orbifold quotient S𝒪S\to\mathcal{O} so that GG descends to a group G𝒪<Mod(𝒪)G_{\mathcal{O}}<\mathrm{Mod}(\mathcal{O}) in the mapping class group of the orbifold 𝒪\mathcal{O}; that is, GG consists of lifts of elements of G𝒪G_{\mathcal{O}}. Let <Mod(𝒪)\mathfrak{C}<\mathrm{Mod}(\mathcal{O}) be the relative commensurator of G𝒪G_{\mathcal{O}} in Mod(𝒪)\mathrm{Mod}(\mathcal{O}), which consists of the elements gMod(𝒪)g\in\mathrm{Mod}(\mathcal{O}) so that gG𝒪g1G𝒪gG_{\mathcal{O}}g^{-1}\cap G_{\mathcal{O}} has finite index in both gG𝒪g1gG_{\mathcal{O}}g^{-1} and G𝒪G_{\mathcal{O}}. (In fact, we must allow for orientation reversing mapping classes, but continue to denote this group Mod\mathrm{Mod} for simplicity.). Finally, we let Γ\Gamma_{\mathfrak{C}} denote the π1orb𝒪\pi_{1}^{\rm{orb}}\mathcal{O}–extension of \mathfrak{C}. Since S𝒪S\to\mathcal{O} is a finite sheeted cover, ΓG𝒪\Gamma_{G_{\mathcal{O}}} contains Γ\Gamma with finite index, and so is quasi-isometric to it, and thus QI(ΓG𝒪)QI(Γ)\operatorname{QI}(\Gamma_{G_{\mathcal{O}}})\cong\operatorname{QI}(\Gamma). It is also not hard to see that there is a natural injection from ΓQI(ΓG𝒪)QI(Γ)\Gamma_{\mathfrak{C}}\to\operatorname{QI}(\Gamma_{G_{\mathcal{O}}})\cong\operatorname{QI}(\Gamma).

We can now state Mosher’s key reduction of quasi-isometric rigidity from [Mos06].

Theorem 5.22 (Mosher).

The homomorphism ΓQI(Γ)\Gamma_{\mathfrak{C}}\to\operatorname{QI}(\Gamma) is an isomorphism.

Very briefly, the proof of this theorem divides into two key steps. First, Γ\Gamma_{\mathfrak{C}} naturally maps not just to QI(ΓG𝒪)\operatorname{QI}(\Gamma_{G_{\mathcal{O}}}), but to the subgroup of coarsely fiber preserving quasi-isometries, QIfib(ΓG𝒪)<QI(ΓG𝒪)\operatorname{QI}_{{\mathrm{fib}}}(\Gamma_{G_{\mathcal{O}}})<\operatorname{QI}(\Gamma_{G_{\mathcal{O}}}). In [Mos03], Mosher proves that this is in fact an isomorphism, ΓQIfib(ΓG𝒪)\Gamma_{\mathfrak{C}}\cong\operatorname{QI}_{{\mathrm{fib}}}(\Gamma_{G_{\mathcal{O}}}), whenever GG contains a pseudo-Anosov. Second, a general result of Farb and Mosher [FM00, Theorem 7.7(2)], proved using coarse algebraic-topology, ensures that QIfib(ΓG𝒪)\operatorname{QI}_{{\mathrm{fib}}}(\Gamma_{G_{\mathcal{O}}}) has finite index in QI(ΓG𝒪)\operatorname{QI}(\Gamma_{G_{\mathcal{O}}}) when GG is further assumed to be virtually free.

With Theorem 5.22 in hand, we see that proving quasi-isometric rigidity of Γ\Gamma reduces to proving that ΓG𝒪\Gamma_{G_{\mathcal{O}}} in Γ\Gamma_{\mathfrak{C}} has finite index (which is precisely what [Mos06, Problem 5.4] asks). Equivalently, this reduces to the following.

Lemma 5.23.

The subgroup G𝒪<G_{\mathcal{O}}<\mathfrak{C} has finite index.

Proof.

Observe that the defining quadratic differential qq for GG descends to a quadratic differential q𝒪q_{\mathcal{O}} on 𝒪\mathcal{O} (with at worst simple poles at the orbifold points). To see this, note that any pseudo-Anosov element fGf\in G descends to a pseudo-Anosov element f0𝒪f_{0}\in\mathcal{O} (i.e. ff is a lift of f0f_{0}). The stable/unstable foliations for this pseudo-Anosov element ff are vertical/horizontal for (an affine deformation of) qq, and these descend to stable/unstable foliations for f0f_{0} which are thus vertical/horizontal for q𝒪q_{\mathcal{O}}; thus, qq is the pull back of q𝒪q_{\mathcal{O}}. Since the fixed points of pseudo-Anosov elements of G𝒪G_{\mathcal{O}} are dense in (q𝕆)\mathbb{P}(q_{\mathbb{O}}), it follows that \mathfrak{C} is contained in the stabilizer of (q𝕆)\mathbb{P}(q_{\mathbb{O}}). By [GM91], for example, (q𝒪)\mathbb{P}(q_{\mathcal{O}}) determines the associated Teichmüller disk q𝒪\mathbb{H}_{q_{\mathcal{O}}} and thus the stabilizer of (q𝒪)\mathbb{P}(q_{\mathcal{O}}) is the stabilizer of q𝒪\mathbb{H}_{q_{\mathcal{O}}}, and is therefore the maximal Veech group defined by q𝒪q_{\mathcal{O}}. Since GG is a lattice Veech group, so is G𝒪G_{\mathcal{O}}. Therefore, the inclusion of G𝒪G_{\mathcal{O}} into this maximal Veech group of q𝒪q_{\mathcal{O}} must have finite index. Since \mathfrak{C} is a subgroup of this Veech group, G𝒪<G_{\mathcal{O}}<\mathfrak{C} is finite index as well. ∎

Remark 5.24.

Our proof here is similar to the proof for Kleinian groups that the relative commensurator of a (non-lattice, Zariski dense) geometrically finite Kleinian group is equal to the stabilizer of its limit set. In fact, from [KL07, Theorem 2.1], the limit set of G𝒪G_{\mathcal{O}} in 𝒫ML(𝒪)\mathcal{P}ML(\mathcal{O}) is precisely (q𝒪)\mathbb{P}(q_{\mathcal{O}}).

References

  • [ABD21] Carolyn Abbott, Jason Behrstock, and Matthew Gentry Durham. Largest acylindrical actions and Stability in hierarchically hyperbolic groups. Trans. Amer. Math. Soc. Ser. B, 8:66–104, 2021.
  • [ABO19] Carolyn Abbott, Sahana H. Balasubramanya, and Denis Osin. Hyperbolic structures on groups. Algebr. Geom. Topol., 19(4):1747–1835, 2019.
  • [Ahl66] Lars V. Ahlfors. Fundamental polyhedrons and limit point sets of Kleinian groups. Proc. Nat. Acad. Sci. U.S.A., 55:251–254, 1966.
  • [BBKL20] Mladen Bestvina, Kenneth Bromberg, Autumn E. Kent, and Christopher J. Leininger. Undistorted purely pseudo-Anosov groups. J. Reine Angew. Math., 760:213–227, 2020.
  • [BHMS20] Jason Behrstock, Mark F. Hagen, A. Martin, and A. Sisto. A combinatorial take on hierarchical hyperbolicity and applications to quotients of mapping class groups. preprint, arXiv:2005.00567, 2020.
  • [BHS17a] Jason Behrstock, Mark F. Hagen, and Alessandro Sisto. Asymptotic dimension and small-cancellation for hierarchically hyperbolic spaces and groups. Proc. Lond. Math. Soc. (3), 114(5):890–926, 2017.
  • [BHS17b] Jason Behrstock, Mark F. Hagen, and Alessandro Sisto. Hierarchically hyperbolic spaces, I: Curve complexes for cubical groups. Geom. Topol., 21(3):1731–1804, 2017.
  • [BHS19] Jason Behrstock, Mark Hagen, and Alessandro Sisto. Hierarchically hyperbolic spaces II: Combination theorems and the distance formula. Pacific J. Math., 299(2):257–338, 2019.
  • [BHS21] Jason Behrstock, Mark F. Hagen, and Alessandro Sisto. Quasiflats in hierarchically hyperbolic spaces. Duke Math. J., 170(5):–, 2021.
  • [BL18] Anja Bankovic and Christopher J. Leininger. Marked-length-spectral rigidity for flat metrics. Trans. Amer. Math. Soc., 370(3):1867–1884, 2018.
  • [Bow93] B. H. Bowditch. Geometrical finiteness for hyperbolic groups. J. Funct. Anal., 113(2):245–317, 1993.
  • [Bow13] Brian H. Bowditch. Coarse median spaces and groups. Pacific J. Math., 261(1):53–93, 2013.
  • [Bow18] Brian H. Bowditch. Large-scale rigidity properties of the mapping class groups. Pacific J. Math., 293(1):1–73, 2018.
  • [CLM12] Matt T. Clay, Christopher J. Leininger, and Johanna Mangahas. The geometry of right-angled Artin subgroups of mapping class groups. Groups Geom. Dyn., 6(2):249–278, 2012.
  • [DDLS21] Spencer Dowdall, Matthew G. Durham, Christopher J. Leininger, and Allesandro Sisto. Extensions of Veech groups I: A hyperbolic action. arXiv preprint arXiv:2006.16425v2, 2021.
  • [DELS18] Moon Duchin, Viveka Erlandsson, Christopher J. Leininger, and Chandrika Sadanand. You can hear the shape of a billiard table: Symbolic dynamics and rigidity for flat surfaces. Preprint, to appear in Comment. Math. Helv., 2018.
  • [DGO17] F. Dahmani, V. Guirardel, and D. Osin. Hyperbolically embedded subgroups and rotating families in groups acting on hyperbolic spaces. Mem. Amer. Math. Soc., 245(1156):v+152, 2017.
  • [DHS17] Matthew Gentry Durham, Mark F. Hagen, and Alessandro Sisto. Boundaries and automorphisms of hierarchically hyperbolic spaces. Geom. Topol., 21(6):3659–3758, 2017.
  • [DMS20] Matthew G Durham, Yair N Minsky, and Alessandro Sisto. Stable cubulations, bicombings, and barycenters. arXiv preprint arXiv:2009.13647, 2020.
  • [DT15] Matthew Gentry Durham and Samuel J. Taylor. Convex cocompactness and stability in mapping class groups. Algebr. Geom. Topol., 15(5):2839–2859, 2015.
  • [EF97] David B. A. Epstein and Koji Fujiwara. The second bounded cohomology of word-hyperbolic groups. Topology, 36(6):1275–1289, 1997.
  • [Far98] B. Farb. Relatively hyperbolic groups. Geom. Funct. Anal., 8(5):810–840, 1998.
  • [FM00] Benson Farb and Lee Mosher. On the asymptotic geometry of abelian-by-cyclic groups. Acta Math., 184(2):145–202, 2000.
  • [FM02a] Benson Farb and Lee Mosher. Convex cocompact subgroups of mapping class groups. Geom. Topol., 6:91–152, 2002.
  • [FM02b] Benson Farb and Lee Mosher. The geometry of surface-by-free groups. Geometric & Functional Analysis GAFA, 12(5):915–963, 2002.
  • [GM91] F.P. Gardiner and H.A. Masur. Extremal length geometry of Teichmüller space. Complex Variables Theory Appl., 16(2-3):209–237, 1991.
  • [Gre66] L. Greenberg. Fundamental polyhedra for kleinian groups. Ann. of Math. (2), 84:433–441, 1966.
  • [Ham] Ursula Hamenstädt. Word hyperbolic extensions of surface groups. Preprint, arXiv:math.GT/0505244.
  • [HHP20] Thomas Haettel, Nima Hoda, and Harry Petyt. Coarse injectivity, hierarchical hyperbolicity, and semihyperbolicity. arXiv preprint arXiv:2009.14053, 2020. To appear in Geom. Topol.
  • [HO13] Michael Hull and Denis Osin. Induced quasicocycles on groups with hyperbolically embedded subgroups. Algebr. Geom. Topol., 13(5):2635–2665, 2013.
  • [HRSS21] Mark F. Hagen, Jacob Russell, Alessandro Sisto, and Davide Spriano. Equivariant hierarchically hyperbolic structures for 3-manifold groups via quasimorphisms. In preparation, 2021.
  • [JS79] William H. Jaco and Peter B. Shalen. Seifert fibered spaces in 33-manifolds. Mem. Amer. Math. Soc., 21(220):viii+192, 1979.
  • [KL97] Michael Kapovich and Bernhard Leeb. Quasi-isometries preserve the geometric decomposition of Haken manifolds. Invent. Math., 128(2):393–416, 1997.
  • [KL07] Autumn E. Kent and Christopher J. Leininger. Subgroups of mapping class groups from the geometrical viewpoint. In In the tradition of Ahlfors-Bers. IV, volume 432 of Contemp. Math., pages 119–141. Amer. Math. Soc., Providence, RI, 2007.
  • [KL08a] Autumn E. Kent and Christopher J. Leininger. Shadows of mapping class groups: capturing convex cocompactness. Geom. Funct. Anal., 18(4):1270–1325, 2008.
  • [KL08b] Autumn E. Kent and Christopher J. Leininger. Uniform convergence in the mapping class group. Ergodic Theory Dynam. Systems, 28(4):1177–1195, 2008.
  • [KL12] Ilya Kapovich and Anton Lukyanenko. Quasi-isometric co-Hopficity of non-uniform lattices in rank-one semi-simple Lie groups. Conform. Geom. Dyn., 16:269–282, 2012.
  • [Kob12] Thomas Koberda. Right-angled Artin groups and a generalized isomorphism problem for finitely generated subgroups of mapping class groups. Geom. Funct. Anal., 22(6):1541–1590, 2012.
  • [Loa21] Christopher Loa. Free products of abelian groups in mapping class groups. arXiv preprint arXiv:2103.05144, 2021.
  • [LR06] C. J. Leininger and A. W. Reid. A combination theorem for Veech subgroups of the mapping class group. Geom. Funct. Anal., 16(2):403–436, 2006.
  • [Man05] Jason Fox Manning. Geometry of pseudocharacters. Geom. Topol., 9:1147–1185, 2005.
  • [Mar74] Albert Marden. The geometry of finitely generated kleinian groups. Ann. of Math. (2), 99:383–462, 1974.
  • [Mas70] Bernard Maskit. On boundaries of Teichmüller spaces and on Kleinian groups. II. Ann. of Math. (2), 91:607–639, 1970.
  • [Min96a] Yair N. Minsky. Extremal length estimates and product regions in Teichmüller space. Duke Math. J., 83(2):249–286, 1996.
  • [Min96b] Yair N. Minsky. Quasi-projections in Teichmüller space. J. Reine Angew. Math., 473:121–136, 1996.
  • [MM99] H.A. Masur and Y.N. Minsky. Geometry of the complex of curves. I. Hyperbolicity. Invent. Math., 138(1):103–149, 1999.
  • [MM00] H.A. Masur and Y. N. Minsky. Geometry of the complex of curves. II. Hierarchical structure. Geom. Funct. Anal., 10(4):902–974, 2000.
  • [Mos03] Lee Mosher. Fiber respecting quasi-isometries of surface group extensions. preprint arXiv:math/0308067, 2003.
  • [Mos06] Lee Mosher. Problems in the geometry of surface group extensions. In Problems on mapping class groups and related topics, volume 74 of Proc. Sympos. Pure Math., pages 245–256. Amer. Math. Soc., Providence, RI, 2006.
  • [MS12] Mahan Mj and Pranab Sardar. A combination theorem for metric bundles. Geom. Funct. Anal., 22(6):1636–1707, 2012.
  • [MT02] Howard Masur and Serge Tabachnikov. Rational billiards and flat structures. In Handbook of dynamical systems, Vol. 1A, pages 1015–1089. North-Holland, Amsterdam, 2002.
  • [MW95] H.A. Masur and M. Wolf. Teichmüller space is not Gromov hyperbolic. Ann. Acad. Sci. Fenn. Ser. A I Math., 20(2):259–267, 1995.
  • [Raf14] Kasra Rafi. Hyperbolicity in Teichmüller space. Geom. Topol., 18(5):3025–3053, 2014.
  • [RST18] Jacob Russell, Davide Spriano, and Hung Cong Tran. Convexity in hierarchically hyperbolic spaces. arXiv preprint arXiv:1809.09303, 2018. To appear in Algebr. Geom. Topol.
  • [Run20] Ian Runnels. Effective generation of right-angled artin groups in mapping class groups. arXiv preprint arXiv:2004.13585, 2020.
  • [Rus21] Jacob Russell. Extensions of multicurve stabilizers are hierarchically hyperbolic. arXiv preprint arXiv:2107.14116, 2021.
  • [Sch95] Richard Evan Schwartz. The quasi-isometry classification of rank one lattices. Inst. Hautes Études Sci. Publ. Math., (82):133–168 (1996), 1995.
  • [Sis16] Alessandro Sisto. Quasi-convexity of hyperbolically embedded subgroups. Math. Z., 283(3-4):649–658, 2016.
  • [Sis19] Alessandro Sisto. What is a hierarchically hyperbolic space? In Beyond hyperbolicity, volume 454 of London Math. Soc. Lecture Note Ser., pages 117–148. Cambridge Univ. Press, Cambridge, 2019.
  • [Tan19] Robert Tang. Affine diffeomorphism groups are undistorted. preprint arXiv:1912.06537, 2019.
  • [Thu86] W.P. Thurston. Geometry and topology of 33–manifolds. Princeton University Lecture Notes, online at http://www.msri.org/publications/books/gt3m, 1986.