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

On a generalization of Cannon’s conjecture for cubulated hyperbolic groups

Corey Bregman Department of Mathematics, Tufts University, Medford, MA 02155 USA [email protected] https://sites.google.com/view/cbregman  and  Merlin Incerti-Medici Universität Wien, Fakultät für Mathematik, 1090 Wien, Austria [email protected] https://www.merlinmedici.ch/
Abstract.

We show that cubulated hyperbolic groups with spherical boundary of dimension 3 or at least 5 are virtually fundamental groups of closed, orientable, aspherical manifolds, provided that there are sufficiently many quasi-convex, codimension-1 subgroups whose limit sets are locally flat subspheres. The proof is based on ideas used by Markovic in his work on Cannon’s conjecture for cubulated hyperbolic groups with 2-sphere boundary.

2020 Mathematics Subject Classification:
57N16, 20F67 (primary), 20F65, 57N45, 57N35 (secondary)

1. Introduction

The Cannon conjecture asserts that if GG is a (Gromov) hyperbolic group whose boundary G\partial_{\infty}G is homeomorphic to S2S^{2}, then GG is commensurable up to finite quotients with a cocompact lattice Γ\Gamma in Isom(3)\operatorname{Isom}(\mathbb{H}^{3}), the isometry group of real hyperbolic 3-space. More precisely, if the Cannon conjecture is true, then there exists a cocompact lattice ΓIsom(3)\Gamma\leq\operatorname{Isom}(\mathbb{H}^{3}) and a short exact sequence

1FGΓ11\to F\to G\to\Gamma\to 1

where FF is a finite group equal to the kernel of the action of GG on G\partial_{\infty}G.

Markovic [Mar13] showed that the Cannon conjecture holds under a further assumption, namely that GG has enough quasi-convex surface subgroups to separate pairs of points in G\partial_{\infty}G. This assumption implies in particular that GG is cubulated, i.e. GG acts geometrically on a CAT(0) cube complex. Under this action the hyperplane stabilizers are surface subgroups. Later Haïssinsky [Hai15] strengthened this result by showing that if GG is cubulated, then GG has enough quasi-convex surface subgroups to separate pairs of points at infinity. Conversely, since every cocompact lattice ΓIsom(3)\Gamma\leq\operatorname{Isom}(\mathbb{H}^{3}) is cubulated by results of Kahn–Markovic [KM12] and Bergeron–Wise [BW12], it follows that the Cannon conjecture is equivalent to the statement that every hyperbolic group GG with GS2\partial_{\infty}G\cong S^{2} is cubulated. The goal of the present work is to prove the following extension of Markovic’s result:

Theorem 1.1.

Let GG be a hyperbolic group such that GS3\partial_{\infty}G\cong S^{3} and suppose GG contains quasi-convex, codimension-1 subgroups {H1,,Hk}\{H_{1},\ldots,H_{k}\} satisfying

  1. (1)

    The GG-translates of the limit sets Hi\partial_{\infty}H_{i} separate pairs of points in G\partial_{\infty}G.

  2. (2)

    Hi\partial_{\infty}H_{i} is a locally flat 2-sphere in S3GS^{3}\cong\partial_{\infty}G for i=1,,ki=1,\ldots,k.

Then there exists a finite-index, torsion-free subgroup G^G\hat{G}\leq G such that G^π1(M)\hat{G}\cong\pi_{1}(M), where MM is a closed, orientable 44-manifold covered by 4\mathbb{R}^{4}.

We recall that an embedding e:MNe:M\hookrightarrow N of a kk-manifold into an nn-manifold is locally flat if for every xMx\in M, there exists a neighbourhood UU of e(x)e(x) in NN such that (U,Ue(M))(U,U\cap e(M)) is homeomorphic to (n,k)(\mathbb{R}^{n},\mathbb{R}^{k}). As in Markovic’s case, hypothesis (1) implies that GG is cubulated with respect to a subset of {H1,,Hk}\{H_{1},\ldots,H_{k}\}. In particular, each HiH_{i} is itself cubulated, hence by Markovic’s result, each HiH_{i} is virtually a lattice in Isom(3)\operatorname{Isom}(\mathbb{H}^{3}). Since cubulated hyperbolic groups are residually finite, we may pass to a finite index subgroup of GG that is torsion-free and cubulated with respect to orientable, hyperbolic 3-manifold groups. Hypothesis (2) does not have a parallel in Markovic’s setting because the Jordan curve theorem implies that the boundary of each surface subgroup is a locally flat circle in S2S^{2}. However, Apanasov–Tetenov [AT88] have constructed examples of discrete subgroups in Isom(4)\operatorname{Isom}(\mathbb{H}^{4}) that are isomorphic to closed hyperbolic 3-manifold groups but whose limit sets in S3S^{3} are wild 2-spheres. As far as the authors are aware, such subgroups do not arise as quasi-convex subgroups of uniform lattices in Isom(4)\operatorname{Isom}(\mathbb{H}^{4}), but the necessity of this assumption remains unclear.

The direct analog of Markovic’s result is false when GS3\partial_{\infty}G\cong S^{3}. Indeed, for all n4n\geq 4, Gromov–Thurston [GT87] construct examples of closed, negatively curved nn-manifolds that are not homotopy equivalent to real hyperbolic nn-manifolds, but which were shown to be cubulated by Giralt [Gir17]. In fact, these manifolds contain enough totally geodesic real hyperbolic nn-manifolds to cubulate.

The Gromov–Thurston examples underscore that for n4n\geq 4, if GG is a torsion-free hyperbolic group with GSn1\partial_{\infty}G\cong S^{n-1}, the most one can hope for is that Gπ1(M)G\cong\pi_{1}(M) for some closed aspherical nn-manifold MM. Remarkably, even without the cubulation assumption, Bartels–Lück–Weinberger [BLW10] show exactly this: if GG is a torsion-free hyperbolic group with GSn1\partial_{\infty}G\cong S^{n-1} for n6n\geq 6, then Gπ1(M)G\cong\pi_{1}(M) for some closed aspherical nn-manifold MM. They remark that their proof ought to extend to the n=5n=5 case as well, assuming certain surgery results hold. Therefore, Theorem 1.1 fills in a dimension where surgery techniques and the ss-cobordism theorem are unavailable. Nevertheless, using results of Bartels–Lück [BL12] on the Borel conjecture, our proof works equally well for all n5n\neq 5, hence we will prove the main theorem in this generality:

Theorem 1.2.

Let GG be a torsion-free hyperbolic group such that GSn1\partial_{\infty}G\cong S^{n-1} and suppose GG contains quasi-convex, codimension-1 subgroups {H1,,Hk}\{H_{1},\ldots,H_{k}\} satisfying:

  1. (1)

    The GG-translates of the Hi\partial_{\infty}H_{i} separate pairs of points in G\partial_{\infty}G.

  2. (2)

    Hi\partial_{\infty}H_{i} is a locally flat (n2)(n-2)-sphere in Sn1GS^{n-1}\cong\partial_{\infty}G for i=1,,ki=1,\ldots,k.

If n5n\neq 5, then there exists a finite-index, torsion-free subgroup G^G\hat{G}\leq G such that G^π1(M)\hat{G}\cong\pi_{1}(M), where MM is a closed orientable nn-manifold covered by n\mathbb{R}^{n}.

In the final section of the paper, we explore the local flatness condition in more depth. When G=Sn1\partial_{\infty}G=S^{n-1} and HGH\leq G is a quasi-convex, codimension-1 subgroup such that H=Sn2\partial_{\infty}H=S^{n-2}, we show in Proposition 6.6 that H\partial_{\infty}H is locally flat in G\partial_{\infty}G if and only if both components of GH\partial_{\infty}G\setminus\partial_{\infty}H are simply connected.

We then give two applications of this result, both of which may be regarded as generalizations of analogous results for hyperbolic 3-manifolds. A classical result, essentially due to Stallings [Sta61] asserts that if VV is the interior of a compact, orientable 3-manifold that is homotopy equivalent to an orientable surface Σ\Sigma, then MM is homeomorphic to Σ×\Sigma\times\mathbb{R}. In particular, it follows that if MM is a closed, orientable hyperbolic 3-manifold and VV is the cover associated to a quasi-convex surface subgroup, then VΣ×V\cong\Sigma\times\mathbb{R} for some surface Σ\Sigma. In our setting, with HGH\leq G as above, we say that HH is 2-sided if the action of HH on G\partial_{\infty}G preserves each component of GH\partial_{\infty}G\setminus\partial_{\infty}H. We prove:

Theorem 6.7.

Let MM be a closed, orientable, aspherical nn-manifold with G=π1(M)G=\pi_{1}(M) hyperbolic, GSn1\partial_{\infty}G\cong S^{n-1} and n6n\geq 6. Suppose HGH\leq G is a quasi-convex, codimension-1, 2-sided subgroup such that HSn2\partial_{\infty}H\cong S^{n-2}. If CHMC_{H}\rightarrow M denotes the cover associated to HH, then HG\partial_{\infty}H\subset\partial_{\infty}G is locally flat if and only if there exists a closed, orientable (n1)(n-1)-manifold NN such that CHN×C_{H}\cong N\times\mathbb{R}.

Our second application further assumes that G=π1(M)G=\pi_{1}(M) is cubulated, and that Hπ1(N)H\cong\pi_{1}(N) for some closed orientable manifold NN. We prove that there exist finite covers of M^\widehat{M} and N^\widehat{N} of MM and NN, respectively, such that N^\widehat{N} embeds in M^\widehat{M} as a submanifold:

Theorem 6.11.

Let MM be a closed, orientable aspherical nn-manifold with G=π1(M)G=\pi_{1}(M) cubulated hyperbolic and n6n\geq 6. Let HGH\leq G be a quasi-convex, codimension-1 subgroup such that Hπ1(N)H\cong\pi_{1}(N) for some closed aspherical (n1)(n-1)-manifold NN. Then there exist finite covers M^M\widehat{M}\rightarrow M, N^N\widehat{N}\rightarrow N, and an embedding N^M^\widehat{N}\hookrightarrow\widehat{M}.

We regard this result as a generalization of the fact that a quasi-convex, π1\pi_{1}-injective surface subgroup of a hyperbolic 3-manifold group lifts to embedding in a finite cover. This is in turn a consequence of Agol’s theorem [Ago13] that cubulated hyperbolic groups are virtually special, and thus QCERF. However, in the case of surface subgroups of hyperbolic 3-manifold groups, one automatically has an immersed surface to try to lift. In the general case, we exploit a splitting theorem due to Cappell [Cap76], which allows us to find an embedded, 2-sided submanifold after passing to a cover. A similar technique (following Cappell’s result in spirit) appeared in a paper of Kar–Niblo [KN13].

1.1. Overview of the paper

In Section 2, we review some results on quasi-convex subgroups of cubulated hyperbolic groups, the work of Bartels–Lück on the Borel conjecture for hyperbolic groups, and the topological generalized Schoenflies theorem of Mazur and Brown.

In Section 3, we introduce the concepts of generalized cell decompositions and GG-complexes, first defined by Markovic. Although many of the results of this section are straightforward generalizations of the results in [Mar13], we have provided many of the details for the convenience of the reader. Our exposition differs from Markovic in places, specifically with regard to condition ()(\star) (see Remark 3.3), which is an attempt to recast some notions in purely topological terms. This paper arose in part from a desire to understand all of the ideas in [Mar13], which we believe may be applicable more generally. We therefore hope that this alternative viewpoint will prove useful. Some of the more lengthy proofs have been relegated to the Appendix, when we felt that the techniques of the proof were not vital to understanding the remainder of the paper.

In Section 4, we review nn-dimensional Sierpinski spaces and collect uniformization results about their embeddings in spheres. In 5, we prove Theorem 1.2. Finally, in Section 6, we discuss a condition which guarantees local flatness of boundaries of quasi-convex subgroups, and use it to prove the theorems about closed aspherical manifolds with cubulated hyperbolic fundamental group discussed above.

1.2. Acknowledgements

We would like to thank Vlad Markovic for several helpful comments. The first author was supported by NSF grant DMS-2401403. The second author was supported by the Austrian Science Fund (FWF) grant 10.55776/ESP124. For open access purposes, the authors have applied a CC BY public copyright license to any author-accepted manuscript version arising from this submission.

2. Preliminary reductions

Suppose GG is cubulated hyperbolic. By results of Agol [Ago13], we know that GG is virtually special, hence virtually torsion-free. For simplicity, from now on we will assume that GG is a torsion-free hyperbolic group.

2.1. Malnormal cubulations

Recall that if GG is a hyperbolic group and HGH\leq G is a quasi-convex subgroup, then there is a closed, topological embedding HG\partial_{\infty}H\hookrightarrow\partial_{\infty}G. If G\partial_{\infty}G is connected, then HH has codimension-1 in GG if and only if GH\partial_{\infty}G\setminus\partial_{\infty}H is disconnected.

Definition 2.1.

Let ={H1,,Hk}\mathcal{H}=\{H_{1},\ldots,H_{k}\} be a collection of quasi-convex, codimension-1 subgroups of GG. We say that \mathcal{H} separates points at infinity if for any two points pqGp\neq q\in\partial_{\infty}G, there exists a conjugate HigH_{i}^{g} such that pp and qq lie in different connected components of GHig\partial_{\infty}G\setminus\partial_{\infty}H_{i}^{g}.

Bergeron–Wise [BW12] prove that if \mathcal{H} separates points at infinity then there exists a finite-dimensional CAT(0) cube complex XX on which GG acts geometrically, and whose hyperplane stabilizers are conjugates of elements of \mathcal{H}. For this reason, if ={H1,,Hk}\mathcal{H}=\{H_{1},\ldots,H_{k}\} is a collection of quasi-convex, codimension-1 subgroups that separates points at infinity, we will say that \mathcal{H} cubulates GG.

Definition 2.2.

A subgroup HGH\leq G is called malnormal if HHg={1}H\cap H^{g}=\{1\} for every gHg\notin H.

Markovic showed that after passing to a cover, one can assume that each element of \mathcal{H} is malnormal:

Theorem 2.3 (Theorem 2.1, [Mar13]).

Suppose GG is hyperbolic and cubulated with respect to a collection of quasi-convex, codimension-1 subgroups \mathcal{H}. Then there exists a finite-index subgroup G^G\widehat{G}\leq G that is cubulated with respect to a collection of malnormal, quasi-convex, codimension-1 subgroups.

2.2. The Borel conjecture for hyperbolic groups

Let GG be a finitely generated group and let MM be a closed, aspherical MM manifold such that Gπ1(M)G\cong\pi_{1}(M). In particular, MM is a finite-dimensional K(G,1)K(G,1), and therefore GG is torsion-free. The Borel conjecture asks whether MM is unique up to homeomorphism. More precisely, if NN is another nn-manifold and f:MNf\colon M\rightarrow N is a homotopy equivalence, is ff necessarily homotopic to a homeomorphism?

The Borel conjecture holds in low dimensions as a consequence of the classification of manifolds of dimensions n3n\leq 3. In dimension 1, the circle is the only closed 1-manifold. In dimension 2, it follows from classification of closed surfaces. In dimension 3, it is now a corollary of the Geometrization Theorem due to Perelman [Per02, Per03b, Per03a], relying on previous work of Waldhausen, for Haken 3-manifolds, Scott [Sco83] for Seifert-fibered 3-manifolds, and Mostow rigidity for hyperbolic 3-manifolds (earlier, Gabai–Meyerhoff–Thurston [GMT03] had shown that a homotopy hyperbolic 3-manifold is indeed homeomorphic to a hyperbolic 3-manifold). In high dimensions n5n\geq 5, Farrell–Jones [FJ89a, FJ89c] proved the Borel conjecture when MM admits a metric of non-positive curvature. However, they also proved that the analogous conjecture is false in the smooth category [FJ89b].

The Borel conjecture has been solved in high dimensions for torsion-free hyperbolic groups by work of Bartels–Lück [BL12]. In fact, they prove it for a much more general class of groups, but we will only need the following special case.

Theorem 2.4 (Bartels–Lück).

Let MM be a closed, aspherical manifold of dimension n5n\geq 5 such that π1(M)\pi_{1}(M) is hyperbolic. If NN is an nn-manifold and f:MNf\colon M\rightarrow N is a homotopy equivalence, then ff is homotopic to a homeomorphism.

Combining this with the low-dimensional results mentioned above we have:

Corollary 2.5.

Let MM be a closed, aspherical manifold of dimension n4n\neq 4 such that π1(M)\pi_{1}(M) is hyperbolic. If NN is an nn-manifold and f:MNf\colon M\rightarrow N is a homotopy equivalence, then ff is homotopic to a homeomorphism.

2.3. Local flatness and the generalized Schoenflies theorem

In contrast to the low-dimensional case where GS2\partial_{\infty}G\cong S^{2} the properties of the embedding of HiG\partial_{\infty}H_{i}\hookrightarrow\partial_{\infty}G become more delicate in higher dimensions. To deal with this, we need some terminology and results from the theory of embeddings.

Definition 2.6.

Let k<nk<n, MkM^{k} and NnN^{n} a kk- and nn-dimensional manifold respectively and let xMx\in M. An embedding e:MNe:M\hookrightarrow N is called locally flat at xx, if there exists a neighborhood UU of e(x)e(x) in NN such that (U,Ue(M))(U,U\cap e(M)) is homeomorphic to (n,k)(\mathbb{R}^{n},\mathbb{R}^{k}). We say that ee is locally flat, if it is locally flat at every point in MM.

Definition 2.7.

Let NnN^{n} be an nn-dimensional manifold, knk\leq n, and DkD^{k} the closed unit ball in k\mathbb{R}^{k}. We call an embedding e:DkNe:D^{k}\hookrightarrow N a tame kk-ball, if the restriction of ee to Dk\partial D^{k} is a locally flat embedding.

Definition 2.8.

Let Sn={(x0,,xn)i=0nxi2=1}n+1S^{n}=\{(x_{0},\ldots,x_{n})\mid\sum_{i=0}^{n}x_{i}^{2}=1\}\subset\mathbb{R}^{n+1} be the nn-sphere. For knk\leq n, the standard SkS^{k} in SnS^{n} is the subspace formed as the intersection of SnS^{n} with each of the hyperplanes Hi={xi=0}H_{i}=\{x_{i}=0\} for k+1ink+1\leq i\leq n.

The generalized Schoenflies theorem due to Mazur [Maz59] and Brown [Bro60] relates the notion of local flatness with standard embeddings in codimension 1.

Theorem 2.9 (Generalized Schoenflies theorem).

Let e:Sn1Sne:S^{n-1}\hookrightarrow S^{n} be a locally flat embedding and let ZZ be one of the two connected components of Sne(Sn1)S^{n}\setminus e(S^{n-1}). Then there exists a homeomorphism (Z¯,Z)(Dn,Dn)(\overline{Z},Z)\rightarrow(D^{n},\partial D^{n}).

Remark 2.10.

The generalized Schoenflies theorem implies that locally flat embeddings Sn1SnS^{n-1}\hookrightarrow S^{n} are conjugate by a homeomorphism to the standard embedding. Indeed, let Z+Z^{+} and ZZ^{-} be the two connected components of Sne(Sn1)S^{n}\setminus e(S^{n-1}). By the generalized Schoenflies theorem, both Z+¯\overline{Z^{+}} and Z¯\overline{Z^{-}} are homeomorphic to the closed disk DnnD^{n}\subset\mathbb{R}^{n}. In particular the glueing Z+¯ΛZ¯\overline{Z^{+}}\coprod_{\Lambda}\overline{Z^{-}} is homeomorphic to the glueing DnSn1DnD^{n}\coprod_{S^{n-1}}D^{n} that produces an nn-dimensional sphere with a standard Sn1S^{n-1} inside. Therefore, the homeomorphisms Z±¯Dn\overline{Z^{\pm}}\rightarrow D^{n} allow us to construct a homeomorphism (G,Λ)(Sn,Sn1)(\partial_{\infty}G,\Lambda)\rightarrow(S^{n},S^{n-1}), where Sn1SnS^{n-1}\hookrightarrow S^{n} is standard.

3. Review of GG-complexes

For n3n\geq 3, let DnD^{n} denote the closed unit ball in n\mathbb{R}^{n} and Sn1=DnS^{n-1}=\partial D^{n} its boundary. We denote the interior by int(Dn)\mathrm{int}(D^{n}). In this section, we review and generalize the notion of GG-complexes used in [Mar13]. Our main-objective is to show that all the key results which Markovic proved for GG-complexes on the pair (D3,S2)(D^{3},S^{2}) hold for the pair (Dn,Sn1)(D^{n},S^{n-1}) as well. For the convenience of the reader, we have reproduced many of the proofs here or in Appendix A. We feel that some of the ideas and techniques herein may be useful more generally; for that reason we have erred on the side of providing all the details in our exposition. However, when the proofs truly do not differ significantly from those of [Mar13] apart from superficial changes, we will simply refer to that paper.

Before we start, we deal with some basic notation, definitions, and conventions. Let XX be a compact metric space, CXC\subset X and aXa\in X. Throughout this section, we use the notation

d(C,x):=sup{d(c,x)|cC}.d(C,x):=\sup\{d(c,x)|c\in C\}.

We will use the following terminology from [KMZ24].

Definition 3.1.

Let XX be as above and (Um)m(U_{m})_{m} a sequence of subsets. We define the limit limmUm\lim_{m\rightarrow\infty}U_{m} to be the collection of points obtained as limits of convergent sequences (xm)m(x_{m})_{m}, where xmUmx_{m}\in U_{m} for every mm.

Note that if XX is a compact metric space, then a sequence (Um)m(U_{m})_{m} converges to a singleton set {p}\{p\}, pXp\in X if and only if for every neighbourhood UU of pp, there exists MM such that UmUU_{m}\subset U for all mMm\geq M.

Let H<Homeo(X)H<\operatorname{Homeo}(X) be a subgroup. We say that HH is a convergence group if for every sequence (hm)m(h_{m})_{m} of distinct hmh_{m} in GG, there exists a subsequence (hmk)k(h_{m_{k}})_{k} and points a,bXa,b\in X such that hmkah_{m_{k}}\rightarrow a uniformly on compact subsets of X{b}X\setminus\{b\}. (We denote the constant map that sends all of XX to aa by aa as well.) Note that in this case the sequence (hnk1)k(h_{n_{k}}^{-1})_{k} converges to bb uniformly on compact subsets of X{a}X\setminus\{a\}.

Let GG be a group. A GG-action on XX is a monomorphism GHomeo(X)G\rightarrow\operatorname{Homeo}(X) (all our actions are faithful). Regarding group multiplication, we use the convention that μ(gh)=μ(h)μ(g)\mu(gh)=\mu(h)\circ\mu(g).

3.1. Generalized cell decompositions and GG-complexes

Definition 3.2.

Let KDnK\subset D^{n} be a closed subset and let 𝒰\mathcal{U} denote the collection of connected components of DnKD^{n}\setminus K. We say the pair (K,𝒰)(K,\mathcal{U}) is a generalized cell decomposition of DnD^{n} if the following holds:

  • Sn1KS^{n-1}\subset K.

  • Every open set U𝒰U\in\mathcal{U} is homeomorphic to int(Dn)\mathrm{int}(D^{n}) and its boundary U\partial U is homeomorphic to Sn1S^{n-1}.

  • Every sequence (Um)m(U_{m})_{m} of distinct elements in 𝒰\mathcal{U} has a subsequence (Umk)k(U_{m_{k}})_{k} such that limkUmk\lim_{k\rightarrow\infty}U_{m_{k}} consists of a single point.

Remark 3.3.

Our definition of a generalized cell decomposition differs from the one given in [Mar13]. Instead of our third condition, Markovic requires the following:

  • ()(\star)

    For every δ>0\delta>0, there exists N(δ)N(\delta)\in\mathbb{N} such that there are at most N(δ)N(\delta) many elements in 𝒰\mathcal{U}^{\prime} that have diameter greater than δ\delta.

It is an easy exercise to see that this is equivalent to our third condition (using the fact that DnD^{n} is compact). Therefore, this is an equivalent definition to the one given in [Mar13] and we exploit both ()(\star) and the third property in our definition. For technical reasons that will become apparent later in this section, we prefer the third condition to be expressed in purely topological terms.

The following two elementary examples of generalized cell decompositions will be useful. First, we may consider K=Sn1K=S^{n-1}. Then 𝒰\mathcal{U} consists of one element, the set int(Dn)\mathrm{int}(D^{n}). Second, we may consider K=DnK=D^{n}. In this case, 𝒰\mathcal{U} is the empty set. These two examples are the minimal and maximal generalized cell decompositions with respect to the following partial order.

Definition 3.4.

Let (K,𝒰)(K,\mathcal{U}), (K,𝒰)(K^{\prime},\mathcal{U}^{\prime}) be two generalized cell decompositions of DnD^{n}. We say that (K,𝒰)(K,\mathcal{U}) is a refinement of (K,𝒰)(K^{\prime},\mathcal{U}^{\prime}) if KKK^{\prime}\subset K. This defines a partial order on the collection of all generalized cell decompositions of DnD^{n}.

We now introduce the notion of an action of a group GG on a generalized cell decomposition.

Definition 3.5.

Let GG be a group and (K,𝒰)(K,\mathcal{U}) a generalized cell decomposition of DnD^{n}. A GG-action on (K,𝒰)(K,\mathcal{U}) is a GG-action μ:GHomeo(K)\mu:G\rightarrow\operatorname{Homeo}(K) such that

  1. (1)

    For every gGg\in G, μ(g)(Sn1)=Sn1\mu(g)(S^{n-1})=S^{n-1}.

  2. (2)

    For all gGg\in G, U𝒰U\in\mathcal{U}, there exists U𝒰U^{\prime}\in\mathcal{U} such that μ(g)(U)=U\mu(g)(\partial U)=\partial U^{\prime}.

Note that every GG-action on (K,𝒰)(K,\mathcal{U}) induces a GG-action on Sn1S^{n-1}. We will assume for all our GG-actions that they act by orientation preserving homeomorphisms on Sn1S^{n-1}.

A GG-action induces an action on the collection of ‘cells’ 𝒰\mathcal{U}. This follows from the following Lemma.

Lemma 3.6.

Let V1,V2𝒰V_{1},V_{2}\in\mathcal{U} such that V1=V2\partial V_{1}=\partial V_{2}. Then V1=V2V_{1}=V_{2}.

In particular, if μ\mu is a GG-action on a generalized cell decomposition (K,𝒰)(K,\mathcal{U}), then for every gGg\in G, U𝒰U\in\mathcal{U}, there exists a unique U𝒰U^{\prime}\in\mathcal{U} such that μ(g)(U)=U\mu(g)(\partial U)=\partial U^{\prime}. We write

μ(g)(U)=U,\mu(g)(U)=U^{\prime},

which defines an action of GG on 𝒰\mathcal{U}.

Proof.

Let V1,V2𝒰V_{1},V_{2}\in\mathcal{U} such that V1=V2\partial V_{1}=\partial V_{2}. Since V1,V2V_{1},V_{2} are connected components of DnKD^{n}\setminus K, they are either disjoint or equal. Suppose they are disjoint. Since V1,V2V_{1},V_{2} are open subsets of DnD^{n} such that (V1¯,V1)(\overline{V_{1}},\partial V_{1}) and (V2¯,V2)(\overline{V_{2}},\partial V_{2}) are homeomorphic to (Dn,Sn1)(D^{n},S^{n-1}), the union V1¯V2¯\overline{V_{1}}\cup\overline{V_{2}} is homeomorphic to the closed manifold DnSn1DnSnD^{n}\coprod_{S^{n-1}}D^{n}\approx S^{n}, where the left-hand-side denotes a glueing of two closed unit balls along a homeomorphism between their boundaries. Since V1¯V2¯\overline{V_{1}}\cup\overline{V_{2}} is compact, it is an embedded copy of SnS^{n} in DnD^{n}. However, closed nn-dimensional manifolds cannot be embedded into n\mathbb{R}^{n} (see for example [Hat02, Corollary 2B.4]). We obtain a contradiction and conclude that V1=V2V_{1}=V_{2}. ∎

Definition 3.7.

For U𝒰U\in\mathcal{U}, we define its stabilizer to be

Stabμ(U):={gG|μ(g)(U)=U}.\operatorname{Stab}_{\mu}(U):=\{g\in G|\mu(g)(\partial U)=\partial U\}.
Definition 3.8.

Let GG be a group, (K,𝒰)(K,\mathcal{U}) a generalized cell decomposition of DnD^{n}, and μ\mu a GG-action on (K,𝒰)(K,\mathcal{U}). We call the triple (μ,K,𝒰)(\mu,K,\mathcal{U}) a GG-complex, if for every U𝒰U\in\mathcal{U} there exists a homeomorphism ΦU:(Dn,Sn1)(U¯,U)\Phi^{U}:(D^{n},S^{n-1})\rightarrow(\overline{U},\partial U) such that the following two conditions are satisfied:

  1. (1)

    ΦU\Phi^{U} fixes every point in Sn1US^{n-1}\cap\partial U.

  2. (2)

    The collection {ΦU}U𝒰\{\Phi^{U}\}_{U\in\mathcal{U}} is GG-equivariant on the boundary, that is for every gGg\in G,

    μ(g)ΦU=Φμ(g)(U)μ(g)on Sn1.\mu(g)\circ\Phi^{U}=\Phi^{\mu(g)(U)}\circ\mu(g)\quad\text{on $S^{n-1}$}.

If we remember the choices of ΦU\Phi^{U}, we call the collection (μ,K,𝒰,Φ𝒰)(\mu,K,\mathcal{U},\Phi^{\mathcal{U}}) a marked GG-complex.

Definition 3.9.

Let μ:GHomeo(K)\mu:G\rightarrow\operatorname{Homeo}(K) be a GG-action on (K,𝒰)(K,\mathcal{U}).

  • We say that μ\mu is free if for every gG{1}g\in G\setminus\{1\}, μ(g)\mu(g) has no fixed points in Kint(Dn)K\cap\mathrm{int}(D^{n}).

  • A GG-complex (μ,K,𝒰)(\mu,K,\mathcal{U}) is called free if and only if the GG-action μ\mu is free.

  • We say (μ,K,𝒰)(\mu,K,\mathcal{U}) is a convergence GG-complex if and only if μ(G)\mu(G) acts as a convergence group on KK.

Remark 3.10.

Let (gm)m(g_{m})_{m} be a sequence of distinct elements in GG. If (μ,K,𝒰)(\mu,K,\mathcal{U}) is a convergence GG-complex, then by definition we find two points a,bKa,b\in K and a subsequence (gmk)k(g_{m_{k}})_{k} such that (μ(gmk))ka(\mu(g_{m_{k}}))_{k}\rightarrow a uniformly on K{b}K\setminus\{b\} and (μ(gmk1))kb(\mu(g_{m_{k}}^{-1}))_{k}\rightarrow b uniformly on K{a}K\setminus\{a\}. Since μ(G)\mu(G) preserves Sn1KS^{n-1}\subset K, we conclude that a,bSn1a,b\in S^{n-1}. This holds for any sequence of distinct elements in GG.

Remark 3.11.

If GG is torsion free, then any convergence GG-complex is also free for the following reason: Suppose (μ,K,𝒰)(\mu,K,\mathcal{U}) is a convergence GG-complex and let gG{1}g\in G\setminus\{1\}. Since GG has no torsion, (gm)m(g^{m})_{m} is a sequence of distinct elements in GG to which we can apply the convergence-property. We conclude that μ(g)\mu(g) cannot have any fixed points in Kint(Dn)K\cap\mathrm{int}(D^{n}).

Definition 3.12.

Let (μ,K,𝒰)(\mu,K,\mathcal{U}) be a GG-complex with marking Φ𝒰\Phi^{\mathcal{U}} and let f:DnDnf\colon D^{n}\rightarrow D^{n} be a homeomorphism. Define f(μ)(g):=fμ(g)f1f_{*}(\mu)(g):=f\circ\mu(g)\circ f^{-1} for all gGg\in G, and f(𝒰):={f(U)U𝒰}f(\mathcal{U}):=\{f(U)\mid U\in\mathcal{U}\}. The pushforward of (μ,K,𝒰)(\mu,K,\mathcal{U}) is

f(μ,K,𝒰):=(f(μ),f(K),f(𝒰))f_{*}(\mu,K,\mathcal{U}):=(f_{*}(\mu),f(K),f(\mathcal{U}))

with marking f(Φ𝒰):={Φf(U)=fΦUf1f(U)f(𝒰)}f_{*}(\Phi^{\mathcal{U}}):=\{\Phi^{f(U)}=f\circ\Phi^{U}\circ f^{-1}\mid f(U)\in f(\mathcal{U})\}.

That f(μ,K,𝒰)f_{*}(\mu,K,\mathcal{U}) satisfies (1) and (2) of Definition 3.8 is an exercise, noting that ff is actually a homeomorphism of pairs (Dn,Dn1)(Dn,Dn1)(D^{n},\partial D^{n-1})\smash{\xrightarrow{\simeq}}(D^{n},\partial D^{n-1}). Similarly, one can define a pullback of a GG-complex, but these are completely interchangeable for our purposes.

Proposition 3.13.

Being a convergence GG-complex or a free GG-complex is preserved under pushforward.

Proof.

Both of these are defined by properties of the GG-action in the compact-open topology and the fixed points of elements of GG in the boundary Sn1S^{n-1}. They are both preserved by pushforward since ff is a homeomorphism of pairs (Dn,Dn1)(Dn,Dn1)(D^{n},\partial D^{n-1})\xrightarrow{\simeq}(D^{n},\partial D^{n-1}). ∎

Proposition 3.14.

Set K0:=DnK_{0}:=D^{n} and let (μ,K0,)(\mu,K_{0},\emptyset) be a free convergence GG-complex. Then the quotient M:=\faktorint(Dn)μ(G)M:=\faktor{\mathrm{int}(D^{n})}{\mu(G)} is an aspherical nn-dimensional manifold whose fundamental group is isomorphic to GG.

Proof.

Since (μ,K0,)(\mu,K_{0},\emptyset) is free, μ(g)\mu(g) has no fixed point on K0int(Dn)=int(Dn)K_{0}\cap\mathrm{int}(D^{n})=\mathrm{int}(D^{n}) for every gG{1}g\in G\setminus\{1\}. In other words, the action of μ(G)\mu(G) on int(Dn)\mathrm{int}(D^{n}) is free. Since (μ,K0,)(\mu,K_{0},\emptyset) is a convergence GG-complex, μ(G)\mu(G) is a convergence group on DnD^{n} and thus its action on int(Dn)\mathrm{int}(D^{n}) is proper (w. r. t. the euclidean metric). Therefore, the quotient \faktorint(Dn)μ(G)\faktor{\mathrm{int}(D^{n})}{\mu(G)} is an nn-dimensional manifold. Since int(Dn)\mathrm{int}(D^{n}) is contractible, this quotient is aspherical and its fundamental group is isomorphic to GG. ∎

We see that Theorems 1.1 and 1.2 follow if we are able to construct a free convergence GG-complex (μ,Dn,)(\mu,D^{n},\emptyset) from the assumptions made about GG in these theorems.

3.2. Refinements of GG-complexes

Definition 3.15.

Let (μ,K,𝒰)(\mu,K,\mathcal{U}) and (μ,K,𝒰)(\mu^{\prime},K^{\prime},\mathcal{U}^{\prime}) be two GG-complexes on DnD^{n}. We say (μ,K,𝒰)(\mu,K,\mathcal{U}) is a refinement of (μ,K,𝒰)(\mu^{\prime},K^{\prime},\mathcal{U}^{\prime}) if KKK^{\prime}\subset K and μ(g)=μ(g)\mu(g)=\mu^{\prime}(g) on KK^{\prime}.

The notion of refinement defines a partial order on the collection of all GG-complexes. Note that if (μ,K,𝒰)(\mu,K,\mathcal{U}) is a refinement of (μ,K,𝒰)(\mu^{\prime},K^{\prime},\mathcal{U}^{\prime}), then (K,𝒰)(K,\mathcal{U}) is a refinement of (K,𝒰)(K^{\prime},\mathcal{U}^{\prime}).

The next two lemmas are key technical results concerning refinements of GG-complexes, but we feel their proofs are lengthy and not essential for understanding the remainder of the paper. In order to avoid interrupting the flow of this section, we have relegated their proofs to Appendix A. The first concerns the relationship between stabilizers under refinement, and is stated without proof in [Mar13, p. 1047]:

Lemma 3.16.

Let (μ,K,𝒰)(\mu,K,\mathcal{U}) be a refinement of (μ,K,𝒰)(\mu^{\prime},K^{\prime},\mathcal{U}^{\prime}) and suppose UUU\subset U^{\prime}, where U𝒰U\in\mathcal{U} and U𝒰U^{\prime}\in\mathcal{U}^{\prime}. Then Stabμ(U)<Stabμ(U)\operatorname{Stab}_{\mu}(U)<\operatorname{Stab}_{\mu^{\prime}}(U^{\prime}).

The second key result concerns the behavior of convergence GG-complexes under refinement. The proof of this lemma is very similar to the proof in the D3D^{3}-case, but due to its technical nature we have reproduced it in full to make sure it does not rely on dimension 3.

Lemma 3.17.

Let (μ,K,𝒰)(\mu,K,\mathcal{U}) be a GG-complex which is a refinement of a convergence GG-complex (μ,K,𝒰)(\mu^{\prime},K^{\prime},\mathcal{U}^{\prime}). If for every U𝒰U^{\prime}\in\mathcal{U}^{\prime}, the group μ(Stabμ(U))\mu(\operatorname{Stab}_{\mu^{\prime}}(U^{\prime})) is a convergence group on U¯K\overline{U^{\prime}}\cap K, then (μ,K,𝒰)(\mu,K,\mathcal{U}) is a convergence GG-complex as well.

3.3. Refining one GG-complex by another

Consider two marked GG-complexes (μ1,K1,𝒰1,Φ1𝒰1)(\mu_{1},K_{1},\mathcal{U}_{1},\Phi_{1}^{\mathcal{U}_{1}}) and (μ2,K2,𝒰2,Φ2𝒰2)(\mu_{2},K_{2},\mathcal{U}_{2},\Phi_{2}^{\mathcal{U}_{2}}) that agree on the boundary, i.e. suppose that for all gGg\in G,

μ1(g)|Sn1=μ2(g)|Sn1.\mu_{1}(g)|_{S^{n-1}}=\mu_{2}(g)|_{S^{n-1}}.

Markovic [Mar13] constructs the refinement of (μ1,K1,𝒰1,Φ1𝒰1)(\mu_{1},K_{1},\mathcal{U}_{1},\Phi_{1}^{\mathcal{U}_{1}}) induced by (μ2,K2,𝒰2,Φ2𝒰2)(\mu_{2},K_{2},\mathcal{U}_{2},\Phi_{2}^{\mathcal{U}_{2}}) in the D3D^{3}-case. The construction works the same way in the DnD^{n}-case and it goes as follows. We define

𝒰:={Φ1V(W)|V𝒰1,W𝒰2},\mathcal{U}:=\{\Phi_{1}^{V}(W)|V\in\mathcal{U}_{1},W\in\mathcal{U}_{2}\},
K:=K1V𝒰1Φ1V(K2).K:=K_{1}\cup\bigcup_{V\in\mathcal{U}_{1}}\Phi_{1}^{V}(K_{2}).

For every U=Φ1V(W)𝒰U=\Phi_{1}^{V}(W)\in\mathcal{U}, the map ΦU:=Φ1VΦ2W:(Dn,Sn1)(U¯,U)\Phi^{U}:=\Phi_{1}^{V}\circ\Phi_{2}^{W}:(D^{n},S^{n-1})\rightarrow(\overline{U},\partial U) is a homeomorphism and we define ΦU\Phi^{U} as the marking for U𝒰U\in\mathcal{U}. One easily checks that KK is the complement of the union of all elements in 𝒰\mathcal{U} and thus we obtain that (K,𝒰)(K,\mathcal{U}) satisfies the first two conditions of a generalized cell decomposition of DnD^{n} (see figure 1).

Refer to caption
(μ1,K1,𝒰1)(\mu_{1},K_{1},\mathcal{U}_{1})
(μ2,K2,𝒰2)(\mu_{2},K_{2},\mathcal{U}_{2})
(μ,K,𝒰)(\mu,K,\mathcal{U})
Refer to caption
Figure 1. This figure shows the refinement of (μ1,K1,𝒰1)(\mu_{1},K_{1},\mathcal{U}_{1}) induced by (μ2,K2,𝒰2)(\mu_{2},K_{2},\mathcal{U}_{2}). The blue lines in the bottom figure are the parts of the refinement induced by the second refinement.

To conclude that (K,𝒰)(K,\mathcal{U}) is a generalized cell decomposition, we are left to show that every sequence of distinct elements (Um)m(U_{m})_{m} in 𝒰\mathcal{U} has a subsequence that converges to a point. Consider a sequence of distinct elements (Um)m(U_{m})_{m} and suppose Um=Φ1Vm(Wm)U_{m}=\Phi_{1}^{V_{m}}(W_{m}) for Vm𝒰1V_{m}\in\mathcal{U}_{1}, Wm𝒰2W_{m}\in\mathcal{U}_{2}. There are two cases to consider.

If the sequence (Vm)m(V_{m})_{m} contains infinitely many distinct elements, we may choose a subsequence (Umk)k(U_{m_{k}})_{k} such that all VmkV_{m_{k}} are pairwise distinct. Since (K1,𝒰1)(K_{1},\mathcal{U}_{1}) is a generalized cell decomposition, (Vmk)k(V_{m_{k}})_{k} admits a subsequence (which we again denote by (Vmk)k(V_{m_{k}})_{k} again) that converges to a point. Since UmkVmkU_{m_{k}}\subset V_{m_{k}}, the sequence (Umk)k(U_{m_{k}})_{k} converges to that same point.

Now suppose the sequence (Vm)m(V_{m})_{m} contains only finitely many distinct elements. Then there exists some VMV_{M} that UmVMU_{m}\subset V_{M} for infinitely many mm. Consider the subsequence (Umk)k(U_{m_{k}})_{k} of all elements UmkVMU_{m_{k}}\subset V_{M}. Since all UmkU_{m_{k}} are pairwise distinct, we obtain a sequence ((Φ1VM)1(Umk))k((\Phi_{1}^{V_{M}})^{-1}(U_{m_{k}}))_{k} of pairwise distinct elements of 𝒰2\mathcal{U}_{2}. Since (K2,𝒰2)(K_{2},\mathcal{U}_{2}) is a generalized cell decomposition, this sequence contains a subsequence (which we again denote by ((Φ1VM)1(Umk))k((\Phi_{1}^{V_{M}})^{-1}(U_{m_{k}}))_{k}) that converges to a point pp. Since convergence to a point is preserved under homeomorphisms, we conclude that (Umk)k(U_{m_{k}})_{k} converges to the point Φ1VM(p)\Phi_{1}^{V_{M}}(p). Thus we see that (K,𝒰)(K,\mathcal{U}) is a generalized cell decomposition of DnD^{n}.

Remark 3.18.

When the refinement of one GG-complex by another is introduced in [Mar13], the fact that the refinement still satisfies property ()(\star) is not addressed. Since we are unable to prove directly that the refinement satisfies ()(\star), we use an equivalent topological formulation of ()(\star) that we can work with: the third property in our definition of generalized cell decompositions.

We are left to define the GG-action on (K,𝒰)(K,\mathcal{U}) and show that it satisfies all the properties of a GG-complex. Let gGg\in G. On the subset K1KK_{1}\subset K, we set μ(g)=μ1(g)\mu(g)=\mu_{1}(g). The remainder of KK can be written as the union V𝒰1KV\coprod_{V\in\mathcal{U}_{1}}K\cap V and we define μ(g)\mu(g) on each of these sets individually. Let V𝒰V\in\mathcal{U}. On KV¯K\cap\overline{V}, we define

μ(g):=Φ1μ1(g)(V)μ2(g)(Φ1V)1.\mu(g):=\Phi_{1}^{\mu_{1}(g)(V)}\circ\mu_{2}(g)\circ\left(\Phi_{1}^{V}\right)^{-1}.

We need to show that μ(g)\mu(g) is well-defined and a homeomorphism. Indeed, on the set KV=VK1K\cap\partial V=\partial V\subset K_{1}, we have overlapping definitions of μ(g)\mu(g) and we need to check the following equality.

xV:μ1(g)=Φ1μ1(g)(V)μ2(g)(Φ1V)1.\forall x\in\partial V:\mu_{1}(g)=\Phi_{1}^{\mu_{1}(g)(V)}\circ\mu_{2}(g)\circ\left(\Phi_{1}^{V}\right)^{-1}.

This is equivalent to the equation

xSn1:μ1(g)Φ1V=Φ1μ1(g)(V)μ2(g).\forall x\in S^{n-1}:\mu_{1}(g)\circ\Phi_{1}^{V}=\Phi_{1}^{\mu_{1}(g)(V)}\circ\mu_{2}(g).

By assumption, μ1(g)\mu_{1}(g) and μ2(g)\mu_{2}(g) coincide on Sn1S^{n-1} and the marking Φ1𝒰1\Phi_{1}^{\mathcal{U}_{1}} commutes with the action μ1\mu_{1} on Sn1S^{n-1}. Therefore, this equality holds and we conclude that μ(g)\mu(g) is well-defined.

Next, we show that μ(g)\mu(g) is a homeomorphism on KK. By construction, μ(g)\mu(g) is defined as a continuous map on various closed subsets of KK that cover KK. Thus, μ(g)\mu(g) is continuous. Furthermore, we know that μ(g)\mu(g) sends K1K_{1} to itself and KVK\cap V bijectively to Kμ1(g)(V)K\cap\mu_{1}(g)(V). Since all KVK\cap V are pairwise disjoint and μ1(g)\mu_{1}(g) acts bijectively on 𝒰1\mathcal{U}_{1}, we see that μ(g)\mu(g) is bijective. Since KK is compact and Hausdorff, μ(g)\mu(g) is a homeomorphism.

We are left to show that the map μ:GHomeo(K)\mu:G\rightarrow\operatorname{Homeo}(K) is a homomorphism. Recall that we are using the convention that actions by homeomorphisms satisfy the formula μ(gh)=μ(h)μ(g)\mu(gh)=\mu(h)\circ\mu(g). Let g,hGg,h\in G. On K1K_{1}, we simply have

μ(gh)=μ1(gh)=μ1(h)μ1(g)=μ(h)μ(g)\mu(gh)=\mu_{1}(gh)=\mu_{1}(h)\circ\mu_{1}(g)=\mu(h)\circ\mu(g)

For V𝒰1V\in\mathcal{U}_{1}, we compute

μ(gh)=Φ1μ1(gh)(V)μ2(gh)(Φ1V)=Φ1μ1(h)(μ1(g)(V))μ2(h)μ2(g)(Φ1V)1=Φ1μ1(h)(μ1(g)(V))μ2(h)(Φ1μ1(g)(V))1Φ1μ1(g)(V)μ2(g)(Φ1V)1=μ(h)μ(g).\begin{split}\mu(gh)&=\Phi_{1}^{\mu_{1}(gh)(V)}\circ\mu_{2}(gh)\circ\left(\Phi_{1}^{V}\right)\\ &=\Phi_{1}^{\mu_{1}(h)(\mu_{1}(g)(V))}\circ\mu_{2}(h)\circ\mu_{2}(g)\circ\left(\Phi_{1}^{V}\right)^{-1}\\ &=\Phi_{1}^{\mu_{1}(h)(\mu_{1}(g)(V))}\circ\mu_{2}(h)\circ\left(\Phi_{1}^{\mu_{1}(g)(V)}\right)^{-1}\circ\Phi_{1}^{\mu_{1}(g)(V)}\circ\mu_{2}(g)\circ\left(\Phi_{1}^{V}\right)^{-1}\\ &=\mu(h)\circ\mu(g).\end{split}

We conclude that μ\mu is a GG-action, turning (μ,K,𝒰,Φ𝒰)(\mu,K,\mathcal{U},\Phi^{\mathcal{U}}) into a marked GG-complex. One immediately sees that (μ,K,𝒰)(\mu,K,\mathcal{U}) is a refinement of (μ1,K1,𝒰1)(\mu_{1},K_{1},\mathcal{U}_{1}).

Remark 3.19.

Let (μi,Ki,𝒰i,Φi𝒰i)(\mu_{i},K_{i},\mathcal{U}_{i},\Phi_{i}^{\mathcal{U}_{i}}) for i=1,2,3i=1,2,3 be three marked GG-complexes. For short, we simply denote them 𝒦1,𝒦2,𝒦3\mathcal{K}_{1},\mathcal{K}_{2},\mathcal{K}_{3}. Let 𝒦i,j\mathcal{K}_{i,j} denote the refinement of 𝒦i\mathcal{K}_{i} induced by 𝒦j\mathcal{K}_{j}. It is an easy exercise to show that the refinement of 𝒦1,2\mathcal{K}_{1,2} induced by 𝒦3\mathcal{K}_{3} is the same as the refinement of 𝒦1\mathcal{K}_{1} induced by 𝒦2,3\mathcal{K}_{2,3}. In other words, the refinement of GG-complexes is associative. We never use this, but think it is worth pointing out.

The following proposition summarizes and generalizes Propositions 3.2, 3.3, and 3.4 from [Mar13] to the extent that we need them. Their proofs are identical to the D3D^{3}-case, short, and do not use the topology or geometry of DnD^{n} at all, which is why we give minimal comment and refer to the proofs given by Markovic.

Proposition 3.20.

Let (μi,Ki,𝒰i)(\mu_{i},K_{i},\mathcal{U}_{i}) for i{1,2}i\in\{1,2\} be two GG-complexes on DnD^{n} and let (μ,K,𝒰)(\mu,K,\mathcal{U}) be the refinement of (μ1,K1,𝒰1)(\mu_{1},K_{1},\mathcal{U}_{1}) induced by (μ2,K2,𝒰2)(\mu_{2},K_{2},\mathcal{U}_{2}). Then the following three statements hold:

  1. (1)

    If (μ1,K1,𝒰1)(\mu_{1},K_{1},\mathcal{U}_{1}) and (μ2,K2,𝒰2)(\mu_{2},K_{2},\mathcal{U}_{2}) are convergence GG-complexes, then (μ,K,𝒰)(\mu,K,\mathcal{U}) is a convergence GG-complex.

  2. (2)

    If (μ1,K1,𝒰1)(\mu_{1},K_{1},\mathcal{U}_{1}) and (μ2,K2,𝒰2)(\mu_{2},K_{2},\mathcal{U}_{2}) are free GG-complexes, then (μ,K,𝒰)(\mu,K,\mathcal{U}) is a free GG-complex.

  3. (3)

    For every U𝒰U\in\mathcal{U}, there exist U1𝒰1U_{1}\in\mathcal{U}_{1}, U2𝒰2U_{2}\in\mathcal{U}_{2} such that

    Stabμ(U)=Stabμ1(U1)Stabμ2(U2).\operatorname{Stab}_{\mu}(U)=\operatorname{Stab}_{\mu_{1}}(U_{1})\cap\operatorname{Stab}_{\mu_{2}}(U_{2}).

We need one final proposition, which generalizes to DnD^{n} due to the fact that radial extensions exist in higher dimensions as well.

Proposition 3.21.

Let (μ,K,𝒰)(\mu,K,\mathcal{U}) be a free convergence GG-complex on DnD^{n} such that Stabμ(U)\operatorname{Stab}_{\mu}(U) is trivial for every U𝒰U\in\mathcal{U}. Let K0:=DnK_{0}:=D^{n}. Then there exists a free convergence GG-complex (μ0,K0,)(\mu_{0},K_{0},\emptyset) that is a refinement of the GG-complex (μ,K,𝒰)(\mu,K,\mathcal{U}).

That is, the action of the convergence group μ(G)\mu(G) on Sn1S^{n-1} can be extended to a free convergence action on DnD^{n}.

Proof.

The GG-complex (μ0,K0,)(\mu_{0},K_{0},\emptyset) will be the refinement of (μ,K,𝒰)(\mu,K,\mathcal{U}) induced by the following GG-complex:

For every gGg\in G, we define μrad(g):DnDn\mu_{rad}(g):D^{n}\rightarrow D^{n} to be the radial extension of the homeomorphism μ(g):Sn1Sn1\mu(g):S^{n-1}\rightarrow S^{n-1}. Since the radial extension defines a monomorphism Homeo(Sn1)Homeo(Dn)\operatorname{Homeo}(S^{n-1})\rightarrow\operatorname{Homeo}(D^{n}), we obtain a GG-complex (μrad,K0,)(\mu_{rad},K_{0},\emptyset).

Set (μ0,K0,)(\mu_{0},K_{0},\emptyset) to be the refinement of (μ,K,𝒰)(\mu,K,\mathcal{U}) induced by (μrad,K0,)(\mu_{rad},K_{0},\emptyset). By assumption, Stabμ(U)\operatorname{Stab}_{\mu}(U) is trivial for every U𝒰U\in\mathcal{U}. Therefore, it follows from Proposition 3.20 (2) that (μ0,K0,)(\mu_{0},K_{0},\emptyset) is a free GG-complex and from Proposition 3.20 (1) that it is a convergence GG-complex. ∎

4. Uniformizing Sierpinski Spaces

In this section we prove a result concerning embeddings of the Sierpinski space n1n\mathcal{M}_{n-1}^{n} into SnS^{n}. We recall the definition of n1n\mathcal{M}_{n-1}^{n} and some of its properties. Let I=[0,1]I=[0,1] be the unit interval, let 𝒞I\mathcal{C}\subset I denote the middle-thirds Cantor set, and let χ:I{0,1}\chi\colon I\rightarrow\{0,1\} be the indicator function for 𝒞\mathcal{C}.

Definition 4.1.

Let InI^{n} be the nn-cube with coordinates 𝐱=(x1,,xn)\mathbf{x}=(x_{1},\ldots,x_{n}). For each n1n\geq 1 and each kk satisfying 0kn0\leq k\leq n, we define a space kn\mathcal{M}^{n}_{k} as follows. For each (nk)(n-k)-element subset α{1,,n}\alpha\subseteq\{1,\ldots,n\}, consider the function fα(𝐱)=iαχ(xi)f_{\alpha}(\mathbf{x})=\prod_{i\in\alpha}\chi(x_{i}), and let Fk(𝐱)=αfα(𝐱)F_{k}(\mathbf{x})=\sum_{\alpha}f_{\alpha}(\mathbf{x}). By convention, we define Fn(𝐱)1F_{n}(\mathbf{x})\equiv 1. If Vk={𝐱InFk(𝐱)=0}V_{k}=\{\mathbf{x}\in I^{n}\mid F_{k}(\mathbf{x})=0\} then kn=InVk\mathcal{M}_{k}^{n}=I^{n}\setminus V_{k}.

By construction, kn\mathcal{M}^{n}_{k} is the set of points with at least nkn-k coordinates in 𝒞\mathcal{C}. In particular, 0n\mathcal{M}^{n}_{0} is the set of points for which every coordinate lies in 𝒞\mathcal{C}, i.e. 𝒞nIn\mathcal{C}^{n}\subset I^{n}, while nn\mathcal{M}_{n}^{n} is InI^{n}. The topological dimension of kn\mathcal{M}_{k}^{n} is kk, and it satisfies a universal property: any compact kk-dimensional metric space which embeds into n\mathbb{R}^{n}, embeds in kn\mathcal{M}^{n}_{k}. In this sense, kn\mathcal{M}^{n}_{k} is the universal receptor for kk-dimensional compacta in n\mathbb{R}^{n}.

By embedding InI^{n} into SnS^{n}, we can regard kn\mathcal{M}^{n}_{k} as a compact subset of SnS^{n}. When k=n1k=n-1, each complementary component of n1n\mathcal{M}^{n}_{n-1} is an open disk with boundary homeomorphic to Sn1(InS^{n-1}(\cong\partial I^{n}). These spheres are called peripheral. With respect to the standard metric on SnS^{n}, the collection of peripheral spheres form a null collection, defined as follows.

Definition 4.2.

A collection of {σk}k=\{\sigma_{k}\}_{k=\infty}^{\infty} of pairwise disjoint embedded (n1)(n-1)-spheres in SnS^{n} (with its standard metric) is called a null collection if limkdiam(σk)=0.\lim_{k\to\infty}\textrm{diam}(\sigma_{k})=0.

In the case of n1n\mathcal{M}^{n}_{n-1}, if one considers the complement of all the peripheral spheres, then the closure of each component is either InI^{n} or n1n\mathcal{M}^{n}_{n-1}. We will show below that for a dense null collection of locally flat spheres, the closure of each component of the complement is homeomorphic to n1n\mathcal{M}^{n}_{n-1}, and that there are at most countably many such components. For this, we will need the following definition.

Definition 4.3.

An nn-triod Υn\Upsilon^{n} is the topological space formed as the union of the nn-disk DnD^{n} and an interval II such that IDnI\cap D^{n} is one of the endpoints of II and an interior point of DnD^{n}.

Since an Υn1\Upsilon^{n-1} is (n1)(n-1)-dimensional, and clearly embeds in n\mathbb{R}^{n}, it also embeds in n1n\mathcal{M}^{n}_{n-1}. In fact, there exists an embedding of Υn1\Upsilon^{n-1} into the “interior" of n1n\mathcal{M}^{n}_{n-1}.

Lemma 4.4.

There exists an embedding of Υn1\Upsilon^{n-1} into n1n\mathcal{M}^{n}_{n-1} whose image is disjoint from every peripheral sphere.

Proof.

Recall that n1n\mathcal{M}^{n}_{n-1} is the set of points in InI^{n} where at least one coordinate is in 𝒞\mathcal{C}. Let x1x_{1} be a point of 𝒞\mathcal{C} that is not a boundary point (i.e. not in the closure of one of the complementary intervals of 𝒞\mathcal{C}). Then {x1}×int(In1)n1n\{x_{1}\}\times\mathrm{int}(I^{n-1})\subset\mathcal{M}^{n}_{n-1}, and is disjoint from every peripheral sphere. We can then find a closed (n1)(n-1)-disk DD contained in {x1}×int(In1)\{x_{1}\}\times\mathrm{int}(I^{n-1}). Now choose x2,,xn𝒞x_{2},\ldots,x_{n}\in\mathcal{C} that are not boundary points such that in p=(x1,xn)int(D)p=(x_{1},\ldots x_{n})\in\mathrm{int}(D). Out of pp, adjoin a small arc γ={(x1+t,,xn1,xn)0tϵ}\gamma=\{(x_{1}+t,\ldots,x_{n-1},x_{n})\mid 0\leq t\leq\epsilon\} for some ϵ<1x1\epsilon<1-x_{1}. Then γ\gamma is also disjoint from every peripheral sphere, and meets DD only at pp. Thus Υ=Dγ\Upsilon=D\cup\gamma is the desired (n1)(n-1)-triod. ∎

Lemma 4.5.

Let {Σk}k=1\{\Sigma_{k}\}_{k=1}^{\infty} be a dense null sequence of locally flat (n-1)-spheres embedded in SnS^{n}. Then Sn(k=1Σk)S^{n}\setminus(\cup_{k=1}^{\infty}\Sigma_{k}) has countably many connected components, and the closure of each is homeomorphic to n1n\mathcal{M}^{n}_{n-1}.

Proof.

We first prove the second claim. Let 𝒦\mathcal{K} be the closure of a component of Sn(k=1Σk)S^{n}\setminus(\cup_{k=1}^{\infty}\Sigma_{k}). Since the Σk\Sigma_{k} are dense and locally flat in SnS^{n}, the closure 𝒦\mathcal{K} is nowhere dense. Moreover, because each Σk\Sigma_{k} is locally flat, Sn𝒦S^{n}\setminus\mathcal{K} consists of a union of tame n-balls {Di}\{D_{i}\}. The boundary of each DiD_{i} is one of the Σk\Sigma_{k}, hence D¯iD¯j=\overline{D}_{i}\cap\overline{D}_{j}=\emptyset if iji\neq j. The fact that the Σk\Sigma_{k} form a null collection implies that the D¯i\overline{D}_{i} do too. By [Can73, Theorem 1], 𝒦\mathcal{K} is homeomorphic to n1n\mathcal{M}^{n}_{n-1}. (The case n=4n=4 follows from [BR13, Theorem 1.3].)

Now we prove that there are only countably many components of Sn(k=1Σk)S^{n}\setminus(\cup_{k=1}^{\infty}\Sigma_{k}). Let 𝒦i\mathcal{K}_{i}, iIi\in I be the closures of components of Sn(k=1Σk)S^{n}\setminus(\cup_{k=1}^{\infty}\Sigma_{k}). Observe that the intersection 𝒦i𝒦j\mathcal{K}_{i}\cap\mathcal{K}_{j} is either empty or equal to Σk\Sigma_{k} for some kk. By Lemma 4.4, in each 𝒦i\mathcal{K}_{i} we can find an embedded (n1)(n-1)-triod ΥiΥn1𝒦i\Upsilon_{i}\cong\Upsilon^{n-1}\subset\mathcal{K}_{i} which does not meet any peripheral sphere of 𝒦i\mathcal{K}_{i}. Since 𝒦i\mathcal{K}_{i} and 𝒦j\mathcal{K}_{j} meet only in a peripheral sphere if at all, the triods will be disjoint. Young [You44] generalized Moore’s triod theorem to all dimensions, proving that n\mathbb{R}^{n} (and thus SnS^{n}) contains at most countably many pairwise disjoint embedded (n1)(n-1)-triods. Hence there can be at most countably many 𝒦i\mathcal{K}_{i}. This proves the first claim, and the lemma. ∎

Proposition 4.6.

Let {Σk}k=1\{\Sigma_{k}\}_{k=1}^{\infty} be a dense null sequence of locally flat (n1)(n-1)-spheres embedded in SnS^{n}. Then there exists a homeomorphism F:SnSnF\colon S^{n}\rightarrow S^{n} such that F(Σk)F(\Sigma_{k}) is round for each kk.

Proof.

By Lemma 4.5, there are countably many components of Sn(k=1Σk)S^{n}\setminus(\cup_{k=1}^{\infty}\Sigma_{k}), and the closure of each is homeomorphic to n1n.\mathcal{M}_{n-1}^{n}. Enumerate the closures of these complementary components as {𝒦j}j=1\{\mathcal{K}_{j}\}_{j=1}^{\infty}. Clearly we have j=1𝒦j=Sn\cup_{j=1}^{\infty}\mathcal{K}_{j}=S^{n} and as we observed in the proof of Lemma 4.5, for iji\neq j, the intersection 𝒦i𝒦j\mathcal{K}_{i}\cap\mathcal{K}_{j} is empty or some Σk\Sigma_{k}. Conversely, each Σk\Sigma_{k} is the intersection of exactly two of the 𝒦i\mathcal{K}_{i}.

Now let Δr\Delta_{r} be a dense null collection of round (n1)(n-1)-spheres in SnS^{n}. The closures of Sn(m=1Δr)S^{n}\setminus(\cup_{m=1}^{\infty}\Delta_{r}) are also homeomorphic to n1n\mathcal{M}_{n-1}^{n}, and again there are countably many of them, which we denote by m,\mathcal{L}_{m}, m1m\geq 1. We build F:SnSnF\colon S^{n}\rightarrow S^{n} as follows. First set j1=m1=1j_{1}=m_{1}=1 and choose a homeomorphism F1:𝒦11F_{1}\colon\mathcal{K}_{1}\rightarrow\mathcal{L}_{1}. Now suppose we have defined FN:i=1N𝒦jii=1NjiF_{N}:\bigcup_{i=1}^{N}\mathcal{K}_{j_{i}}\rightarrow\bigcup_{i=1}^{N}\mathcal{L}_{j_{i}}. Choose some peripheral sphere of ΣjN+1i=1N𝒦ji\Sigma_{j_{N+1}}\subset\cup_{i=1}^{N}\mathcal{K}_{j_{i}} and let ΔmN+1i=1Nji\Delta_{m_{N+1}}\subset\cup_{i=1}^{N}\mathcal{L}_{j_{i}} be the corresponding peripheral sphere. There is a unique 𝒦jN+1\mathcal{K}_{j_{N+1}} not among 𝒦j1,,𝒦jN\mathcal{K}_{j_{1}},\ldots,\mathcal{K}_{j_{N}} meeting ΣjN+1\Sigma_{j_{N+1}}. Let mN+1\mathcal{L}_{m_{N+1}} be the corresponding element meeting ΔmN+1\Delta_{m_{N+1}}. By [Can73, Lemma 1], there exists a homeomorphism hN+1:𝒦jN+1jN+1h_{N+1}\colon\mathcal{K}_{j_{N+1}}\rightarrow\mathcal{L}_{j_{N+1}} extending FN:ΣjN+1ΔmN+1F_{N}\colon\Sigma_{j_{N+1}}\rightarrow\Delta_{m_{N+1}}. (The proof Cannon gives also works when n=4n=4 case, once one knows that the annulus theorem holds. This was proven by Quinn [Qui82, Edw84].) We use this to define FN+1:i=1N+1𝒦jii=1N+1miF_{N+1}\colon\cup_{i=1}^{N+1}\mathcal{K}_{j_{i}}\rightarrow\cup_{i=1}^{N+1}\mathcal{L}_{m_{i}}, by pasting FNF_{N} to hN+1h_{N+1} along ΣjN+1\Sigma_{j_{N+1}}, where they agree.

Since FNF_{N} agrees with FN+1F_{N+1} on i=1N+1𝒦ji\cup_{i=1}^{N+1}\mathcal{K}_{j_{i}}, in the limit we get a well-defined map F:SnSnF\colon S^{n}\rightarrow S^{n}. Injectivity is clear from the definition, hence FF is an embedding as SnS^{n} is compact and Hausdorff. In particular, the image of FF is closed. On the other hand, the domain and codomain have the same dimension so FF is also open. Since SnS^{n} is connected, FF must be onto, hence a homeomorphism. ∎

5. Proof of the main theorem

In this section we prove Theorem 1.2. Let us first recall the hypotheses. We assume that GG is a torsion-free hyperbolic group whose boundary G\partial_{\infty}G is homeomorphic to Sn1S^{n-1}, n5n\neq 5. Moreover GG contains a finite collection of quasi-convex, codimension-1 subgroups \mathcal{H} satisfying:

  1. (1)

    \mathcal{H} separates pairs of points at infinity.

  2. (2)

    For each HH\in\mathcal{H}, H\partial_{\infty}H is a locally flat Sn2S^{n-2} in Sn1G.S^{n-1}\cong\partial_{\infty}G.

By (1), we know that \mathcal{H} cubulates GG; that is, GG acts geometrically on a finite-dimensional CAT(0) cube complex whose hyperplane stabilizers are conjugates of the HiH_{i}\in\mathcal{H}.

Recall that GG acts on its boundary by homeomorphisms, inducing an action by automorphisms on the cohomology Hn1(G)Hn1(Sn1)H^{n-1}(\partial_{\infty}G)\cong H^{n-1}(S^{n-1})\cong\mathbb{Z}. By passing to a subgroup of index at most 22, we may therefore assume that the action of GG on G\partial_{\infty}G is orientation-preserving. We denote the action of GG on the boundary Sn1S^{n-1} by a representation ρ:GHomeo+(Sn1)\rho\colon G\rightarrow\operatorname{Homeo}^{+}(S^{n-1}).

Lemma 5.1.

After replacing \mathcal{H} by a family of finite index subgroups \mathcal{H}^{\prime}, we may assume that each HH\in\mathcal{H}^{\prime} additionally satisfies:

  1. (3)

    There exists a closed, orientable n1n-1-manifold NHN_{H} such that Hπ1(NH)H\cong\pi_{1}(N_{H}). Additionally, HH preserves each complementary component of GH\partial_{\infty}G\setminus\partial_{\infty}H and acts freely, properly discontinuously cocompactly on each.

Proof.

Let HH\in\mathcal{H}. By (2) and the generalized Schoenflies theorem, the pair (G,H)(\partial_{\infty}G,\partial_{\infty}H) is homeomorphic to the standard pair (Sn1,Sn2)(S^{n-1},S^{n-2}). In particular, GH\partial_{\infty}G\setminus\partial_{\infty}H consists of exactly two components, each homeomorphic to n1\mathbb{R}^{n-1} and each of whose closures is homeomorphic to the closed n1n-1-ball 𝔻n1\mathbb{D}^{n-1}.

Since the action of HH on G\partial_{\infty}G preserves H\partial_{\infty}H it also preserves the complement. Therefore, there exists a subgroup HHH^{\prime}\leq H of index at most 22 such that HH^{\prime} preserves each complementary component of GH\partial_{\infty}G\setminus\partial_{\infty}H. Because HH^{\prime} has finite index in HH, the limit set of HH^{\prime} is the same as that of HH, i.e. H=H\partial_{\infty}H^{\prime}=\partial_{\infty}H. Since the action of GG on Sn1S^{n-1} is orientation-preserving, the action of HH^{\prime} on Sn2S^{n-2} must also be orientation-preserving.

Let UU be one of the two components of GH\partial_{\infty}G\setminus\partial_{\infty}H^{\prime}. By a result of Swenson [Swe01], HH^{\prime} acts freely, properly discontinuously and cocompactly on UU. Since Un1U\cong\mathbb{R}^{n-1}, this implies that HH^{\prime} is isomorphic to the fundamental group of U/H=NHU/H^{\prime}=N_{H^{\prime}}, which is a closed, orientable, aspherical n1n-1-manifold. Now the collection \mathcal{H}^{\prime} obtained by replacing each HH\in\mathcal{H} with its corresponding subgroup HH^{\prime} satisfies (1), (2) and (3). ∎

We now assume \mathcal{H} satisfies conditions (1)–(3). Applying Theorem 2.3, we may assume that \mathcal{H} also satisfies

  1. (4)

    Every HH\in\mathcal{H} is malnormal.

Let Λi\Lambda_{i} denote the limit set of HiH_{i}, which can be identified with Hi\partial_{\infty}H_{i}. The action of each HiH_{i} preserves the limit set Λi\Lambda_{i}, which by assumption is homeomorphic to a locally flat (n2)(n-2)-sphere. Now consider HiG:={Hig=gHig1gG}H_{i}^{G}:=\{H_{i}^{g}=gH_{i}g^{-1}\mid g\in G\}, the set of conjugates of HiH_{i}. The limit set of HigH_{i}^{g} is gΛig\Lambda_{i}, which is disjoint from Λi\Lambda_{i} by malnormality. Each translate gΛig\Lambda_{i} separates Sn1S^{n-1} into two components, hence we may define the dual tree TiT_{i} whose edges are translates gΛig\Lambda_{i} and whose vertices are in 1-1 correspondence with connected components of

Sn1(gGgΛi).S^{n-1}\setminus\left(\bigcup_{g\in G}g\Lambda_{i}\right).

For any H=HigHiGH=H_{i}^{g}\in H_{i}^{G}, we use the orientation on Sn1S^{n-1} to choose a left side LHL_{H} and a right side RHR_{H} of ΛH=gΛi\Lambda_{H}=g\Lambda_{i}. The sides LHL_{H} and RHR_{H} correspond to the two half-edges at the midpoint of each edge of TiT_{i}. By (3)(3), GG acts on TiT_{i} without inversions, hence for any gGg^{\prime}\in G, the left side (resp. right side) of ΛH\Lambda_{H} is taken to the left side (resp. right side) of gΛH=ΛHgg^{\prime}\Lambda_{H}=\Lambda_{H^{g^{\prime}}}.

Lemma 5.2.

For HHiGH\in H_{i}^{G}, there exists a collection of homeomorphisms σH:L¯HR¯H\sigma_{H}\colon\overline{L}_{H}\rightarrow\overline{R}_{H} satisfying

  1. (i)

    σH\sigma_{H} restricts to the identity on ΛH\Lambda_{H}.

  2. (ii)

    For any gGg\in G, HHiGH\in H_{i}^{G} and xLHx\in L_{H} we have

    (σHgρ(g))(x)=(ρ(g)σH)(x).(\sigma_{H^{g}}\circ\rho(g))(x)=(\rho(g)\circ\sigma_{H})(x).
Proof.

By (3), HiH_{i} acts on both sides LHi,RHiL_{H_{i}},\leavevmode\nobreak\ R_{H_{i}} of Sn1ΛiS^{n-1}\setminus\Lambda_{i} properly discontinuously, cocompactly. Define NiL=LHi/HiN_{i}^{L}=L_{H_{i}}/H_{i}, and NiR=RHi/HiN_{i}^{R}=R_{H_{i}}/H_{i}. Choosing basepoints in NiLN_{i}^{L} and NiRN_{i}^{R}, we get an identification between π1(NiL)\pi_{1}(N_{i}^{L}) and π1(NiR)\pi_{1}(N_{i}^{R}). Since both of these are aspherical, we obtain a homotopy equivalence hi:NiLNiRh_{i}\colon N_{i}^{L}\rightarrow N_{i}^{R}. Our choice of hih_{i} yields a commutative diagram

π1(NiL)\textstyle{\pi_{1}(N_{i}^{L})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(hi)\scriptstyle{(h_{i})_{*}}π1(NiR)\textstyle{\pi_{1}(N_{i}^{R})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hi\textstyle{H_{i}}

and in particular the lift h~i:N~iLN~iR\widetilde{h}_{i}\colon\widetilde{N}_{i}^{L}\rightarrow\widetilde{N}_{i}^{R} to universal covers induces the identity map on the boundary. By Corollary 2.5, hih_{i} is homotopic to a homeomorphism fi:NiLNiR.f_{i}\colon N_{i}^{L}\rightarrow N_{i}^{R}. (This is where we need that n5n\neq 5.) Since fif_{i} is homotopic to hih_{i}, the same holds for the lift f~i\widetilde{f}_{i}. Set σHi=f~iidΛHi:L¯HiR¯Hi\sigma_{H_{i}}=\widetilde{f}_{i}\cup\text{id}_{\Lambda_{H_{i}}}\colon\overline{L}_{H_{i}}\rightarrow\overline{R}_{H_{i}}; by construction σHiρ(h)(x)=ρ(h)σHi(x)\sigma_{H_{i}}\circ\rho(h)(x)=\rho(h)\circ\sigma_{H_{i}}(x) for every hHih\in H_{i} and xL¯Hix\in\overline{L}_{H_{i}}.

Now suppose H=Hig0H=H_{i}^{g_{0}} and define σH=ρ(g0)σHiρ(g0)1\sigma_{H}=\rho(g_{0})\circ\sigma_{H_{i}}\circ\rho(g_{0})^{-1}. Since the action of gg takes the left side of HiH_{i} to the left side of HH, σH:L¯HR¯H\sigma_{H}\colon\overline{L}_{H}\rightarrow\overline{R}_{H} is a homeomorphism and σH|ΛH\sigma_{H}|_{\Lambda_{H}} is the identity since the same holds for σHi\sigma_{H_{i}}, which proves (i). Let g1Gg_{1}\in G be another group element satisfying Hig1=HH_{i}^{g_{1}}=H, whence g1=g0hg_{1}=g_{0}h for some hHih\in H_{i}. If we define σH=ρ(g1)σHiρ(g1)1\sigma_{H}^{\prime}=\rho(g_{1})\sigma_{H_{i}}\rho(g_{1})^{-1} then for any xL¯Hx\in\overline{L}_{H}

σH(x)\displaystyle\sigma_{H}^{\prime}(x) =\displaystyle= ρ(g1)σHiρ(g1)1(x)\displaystyle\rho(g_{1})\circ\sigma_{H_{i}}\circ\rho(g_{1})^{-1}(x)
=\displaystyle= ρ(g0)ρ(h)σHiρ(h)1ρ(g0)1(x)\displaystyle\rho(g_{0})\circ\rho(h)\circ\sigma_{H_{i}}\circ\rho(h)^{-1}\circ\rho(g_{0})^{-1}(x)
=\displaystyle= ρ(g0)σHiρ(g0)1(x)\displaystyle\rho(g_{0})\circ\sigma_{H_{i}}\circ\rho(g_{0})^{-1}(x)
=\displaystyle= σH(x)\displaystyle\sigma_{H}(x)

where the third equality follows from HiH_{i}-equivariance of σHi\sigma_{H_{i}}. Thus σH\sigma_{H} depends only the coset gHigH_{i}. The equality in (ii) now follows easily. ∎

Combined with results of §4, we now use the homeomorphisms constructed above to construct a generalized cell decomposition of DnD^{n} for each Hi.H_{i}\in\mathcal{H}.

Lemma 5.3.

Let 𝔻n\mathbb{D}^{n} denote the closed nn-disk with the Euclidean metric, with boundary 𝕊n1\mathbb{S}^{n-1}. There exists a homeomorphism fi:Dn𝔻nf_{i}\colon D^{n}\rightarrow\mathbb{D}^{n} and a collection of pairwise disjoint, properly embedded (n1)(n-1)-disks BH𝔻n1𝔻nB_{H}\cong\mathbb{D}^{n-1}\subset\mathbb{D}^{n}, HHiGH\in H_{i}^{G} satisfying the following properties:

  1. (i)

    BH=fi(ΛH)\partial B_{H}=f_{i}(\Lambda_{H}).

  2. (ii)

    If diam(fi(ΛH))0\textrm{diam}(f_{i}(\Lambda_{H}))\to 0 then diam(BH)0\textrm{diam}(B_{H})\to 0.

  3. (iii)

    Let

    Ki=𝕊n1(HHiGBH),K_{i}=\mathbb{S}^{n-1}\cup\left(\bigcup_{H\in H_{i}^{G}}B_{H}\right),

    and let 𝒰i\mathcal{U}_{i} be the set of complementary components of 𝔻nKi\mathbb{D}^{n}\setminus K_{i}. Then KiK_{i} is closed and each pair (U,U)(U,\partial U), U𝒰iU\in\mathcal{U}_{i} is homeomorphic to (Dn,Sn1)(D^{n},S^{n-1}).

  4. (iv)

    Let U𝒰iU\in\mathcal{U}_{i} and a,bUSn1a,b\in\partial U\cap S^{n-1}. Then for each HHigH\in H_{i}^{g}, fi(ΛH)f_{i}(\Lambda_{H}) does not separate aa and bb.

Proof.

For each ϵ>0\epsilon>0, there are only finitely many ΛH,\Lambda_{H}, HHiGH\in H_{i}^{G} such that diam(ΛH)ϵ.\textrm{diam}(\Lambda_{H})\geq\epsilon. The union HHiGΛH\cup_{H\in H^{G}_{i}}\Lambda_{H} is dense in SnS^{n} because it contains the orbit of the endpoints of one axis of an element of GG. Therefore {ΛHHHiG}\{\Lambda_{H}\mid H\in H_{i}^{G}\} forms a dense null collection. By Proposition 4.6, there exists a homeomorphism fi:Sn1𝕊n1f_{i}\colon S^{n-1}\rightarrow\mathbb{S}^{n-1} such that fi(ΛH)f_{i}(\Lambda_{H}) is a round (n2)(n-2)-sphere, for each HHiGH\in H_{i}^{G}. By the Alexander trick, fif_{i} extends to a homeomorphism fi:Dn𝔻nf_{i}\colon D^{n}\rightarrow\mathbb{D}^{n}.

Regarding int(𝔻n)\mathrm{int}(\mathbb{D}^{n}) as hyperbolic nn-space in the Poincaré model, for each HHiGH\in H_{i}^{G}, fi(ΛH)f_{i}(\Lambda_{H}) bounds a round hemisphere BHB_{H}, where BHBHB_{H}\cap B_{H^{\prime}} are disjoint if HHH\neq H^{\prime}. In the Euclidean metric, diam(BH)=diam(fi(ΛBH))\textrm{diam}(B_{H})=\textrm{diam}(f_{i}(\Lambda_{B_{H}})), proving (i) and (ii). The remainder of the proof is exactly the same as that of [Mar13, Lemma 3.2]. ∎

Thus the pair (Ki,𝒰i)(K_{i},\mathcal{U}_{i}) is a generalized cell decomposition. The next result shows that each HiH_{i}\in\mathcal{H} yields a free convergence GG-complex, and when combined with the results of Section 3, it will be the basis for the induction in the proof of the main theorem.

Lemma 5.4.

There is a free convergence GG-complex (ρi,Ki,𝒰i)(\rho_{i},K_{i},\mathcal{U}_{i}) satisfying:

  1. (5)

    For any U𝒰iU\in\mathcal{U}_{i}, any g1Stabρi(U)g\neq 1\in\operatorname{Stab}_{\rho_{i}}(U) and any HHiGH\in H^{G}_{i}, ΛH\Lambda_{H} does not separate the fixed points of gg in U\partial U.

Proof of Lemma 5.4.

Let (Ki,𝒰i)K_{i},\mathcal{U}_{i}) be the generalized cell decomposition from Lemma 5.3. For each HHiGH\in H^{G}_{i}, choose a homeomorphism ψH:L¯HB¯H\psi_{H}\colon\overline{L}_{H}\rightarrow\overline{B}_{H} which restricts to the identity on ΛH\Lambda_{H}. Let ρ0\rho_{0} be the action of GG on Sn1S^{n-1}. We extend this to an action ρi\rho_{i} on KiK_{i} via the formula:

(\dagger) ρi(g)(x)=(ψgHg1ρ0(g)ψH1)(x), for xBH.\rho_{i}(g)(x)=(\psi_{gHg^{-1}}\circ\rho_{0}(g)\circ\psi_{H}^{-1})(x),\textrm{ for }x\in B_{H}.

This is a homeomorphism of KiK_{i} since it agrees with ρ0\rho_{0} on Sn1S^{n-1} and the interiors int(BH)\mathrm{int}(B_{H}) are pairwise disjoint. A simple calculation using Equation (\dagger5) verifies that ρi\rho_{i} defines a GG-action. The proof that (ρi,𝒰i,Ki)(\rho_{i},\mathcal{U}_{i},K_{i}) is a free, convergence GG-action follows exactly as in [Mar13, Lemma 3.2], substituting Lemma 5.2 for [Mar13, Proposition 3.6] and Lemma 5.3 for [Mar13, Proposition 3.7]. ∎

We now combine the results of this section to deduce Theorem 1.2.

Proof of 1.2.

Let GG be a torsion-free hyperbolic group such that GSn1\partial_{\infty}G\cong S^{n-1}, and let ={H1,,Hk}\mathcal{H}=\{H_{1},\ldots,H_{k}\} be quasi-convex, codimension-1 subgroups which satisfy (1) and (2). By Lemma 5.1 and Theorem 2.3, we may assume each HiH_{i}\in\mathcal{H} satisfies (1)–(4). Now by Lemmas 5.3 and 5.4, for 1ik1\leq i\leq k, there exists a homeomorphism fi:Dn𝔻nf_{i}\colon D^{n}\rightarrow\mathbb{D}^{n} and a free convergence GG-complex (ρi,Ki,𝒰i)(\rho_{i},K_{i},\mathcal{U}_{i}) satisfying (5), such that KiK_{i} is a union of 𝕊n1\mathbb{S}^{n-1} with a disjoint collection of round hemispheres 𝔹H\mathbb{B}_{H}, HHiGH\in H_{i}^{G}.

For 1ik1\leq i\leq k, we now define a sequence of free, convergence GG-complexes (μi,Li,𝒱i)(\mu_{i},L_{i},\mathcal{V}_{i}) that is a refinement of (ρi,Ki,𝒰i)(\rho_{i},K_{i},\mathcal{U}_{i}). To start, set (μ1,L1,𝒱1)=(ρ1,K1,𝒰1)(\mu_{1},L_{1},\mathcal{V}_{1})=(\rho_{1},K_{1},\mathcal{U}_{1}), and then inductively define (μi+1,Li+1,𝒱i+1)(\mu_{i+1},L_{i+1},\mathcal{V}_{i+1}) as the refinement of (ρi+1,Ki+1,𝒰i+1)(\rho_{i+1},K_{i+1},\mathcal{U}_{i+1}) by the pushforward (Fi+1)(μi,Li,𝒱i)(F_{i+1})_{*}(\mu_{i},L_{i},\mathcal{V}_{i}), where Fi+1:=fi+1fi1F_{i+1}:=f_{i+1}\circ f_{i}^{-1}. By construction, (ρi+1,Li+1,𝒱i+1)(\rho_{i+1},L_{i+1},\mathcal{V}_{i+1}) is a GG-complex; moreover, it is a free, convergence GG-complex by Proposition 3.20 (1) and (2).

Now consider (μk,𝒱k,Lk)(\mu_{k},\mathcal{V}_{k},L_{k}). By Proposition 3.20 (3), for any V𝒱kV\in\mathcal{V}_{k} there exist Ui𝒰iU_{i}\in\mathcal{U}_{i} such that Stabμk(V)=i=1kStabρi(Ui)\operatorname{Stab}_{\mu_{k}}(V)=\cap_{i=1}^{k}\operatorname{Stab}_{\rho_{i}}(U_{i}). Suppose that there exists g1Stabμk(V)g\neq 1\in\operatorname{Stab}_{\mu_{k}}(V), and let ξ±\xi^{\pm} be the fixed points of gSn1.g\in S^{n-1}. Since gStabρi(Ui)g\in\operatorname{Stab}_{\rho_{i}}(U_{i}), then by (5) we know that ξ+\xi^{+} and ξ\xi^{-} are not separated by ΛH\Lambda_{H} for any HHiGH\in H_{i}^{G} for each ii. But this holds for every HiH_{i}\in\mathcal{H}, contradicting our assumption that \mathcal{H} separates pairs of points in G\partial_{\infty}G. Therefore Stabμk(V)\operatorname{Stab}_{\mu_{k}}(V) is trivial for every V𝒱kV\in\mathcal{V}_{k}. The proof now follows from Propositions 3.21 and 3.14. ∎

6. Characterization of local flatness for codimension-1 submanifolds

6.1. From simple connectedness to local flatness

Let GG be a torsion-free, hyperbolic group such that GSn1\partial_{\infty}G\approx S^{n-1}. Let H<GH<G be a quasi-convex, codimension-1 subgroup such that its limit set ΛG\Lambda\subset\partial_{\infty}G is homeomorphic to Sn2S^{n-2}. By the Jordan-Brouwer separation theorem, GΛ\partial_{\infty}G\setminus\Lambda has exactly two connected components, which we denote by Z+Z^{+} and ZZ^{-}.

Our goal in this section is to prove that if Z+Z^{+} and ZZ^{-} are simply connected, then the embedding ΛG\Lambda\hookrightarrow\partial_{\infty}G is equivalent to the standard embedding Sn2Sn1S^{n-2}\hookrightarrow S^{n-1}. To show this, we need a series of definitions from topology. We do not give much context to these definitions, as they merely appear as assumptions in established theorems that we need to invoke.

Definition 6.1.

An absolute neighborhood retract (ANR) is a normal topological space XX such that for every normal space ZZ, every closed subset YZY\subset Z and every continuous map f:YXf:Y\rightarrow X there exists an open neighborhood UU of YY such that ff has a continuous extension UXU\rightarrow X.

We will need two important facts about ANRs which are summarised in the following Lemma.

Lemma 6.2 ([Pal66, Theorem 5 & Theorem 14]).

Paracompact Hausdorff spaces which are locally ANRs are ANRs. (This includes open topological manifolds.) Furthermore, any ANR has the homotopy type of a CW-complex.

Definition 6.3.

Let YY be a compact space of dimension kk and ZYZ\subset Y a closed subspace. We say that ZZ is kk-LCC in YY, if for every zZz\in Z and every neighborhood UU of zz, there exists a neighborhood VV of zz such that every continuous map α:SkVZ\alpha:S^{k}\rightarrow V\setminus Z extends continuously to α~:Dk+1UZ\tilde{\alpha}:D^{k+1}\rightarrow U\setminus Z.

Let ZZ denote either Z+Z^{+} or ZZ^{-}. Observe that ZZ is an open subspace of the compact manifold Sn1S^{n-1} and thus an ANR. In particular, ZZ has the homotopy type of a CW complex. We begin with the following result.

Lemma 6.4.

If ZZ is simply connected, then ZZ is contracible.

Proof.

As ΛSn2\Lambda\cong S^{n-2} is locally contractible, Alexander duality implies that H~i(Sn1Λ)H~n2i(Λ)\tilde{H}_{i}(S^{n-1}\setminus\Lambda)\cong\tilde{H}^{n-2-i}(\Lambda). Since Sn1Λ=Z+ZS^{n-1}\setminus\Lambda=Z^{+}\coprod Z^{-}, we know that the homology of ZZ appears as a summand in H~i(Sn1Λ)\tilde{H}_{i}(S^{n-1}\setminus\Lambda). The fact that H~n2i(Sn2)=0\tilde{H}^{n-2-i}(S^{n-2})=0 for all i1i\geq 1 implies that ZZ is acyclic. (For degree zero, note that ZZ is a connected component.) Since ZZ is simply connected, we may use the Hurewicz theorem to conclude that the higher homotopy groups vanish as well. Contractibility now follows from Whitehead’s theorem. ∎

For any finitely generated group GG, there is a family of finite-dimensional, simplicial complexes on which GG acts geometrically, known as Rips complexes.

Definition 6.5.

Let SS be a finite generating set for GG and consider the word metric d=dG,Sd=d_{G,S} for GG with respect to SS. For any r>0r>0 we define a simplicial complex called the Rips complex Pr(G)=Pr(G,S)P_{r}(G)=P_{r}(G,S) as follows. The vertex set of Pr(G)P_{r}(G) is GG, and g0,,gkg_{0},\ldots,g_{k} span a kk-simplex if d(gi,gj)rd(g_{i},g_{j})\leq r for all i,j{0,,k}i,j\in\{0,\ldots,k\}.

The word metric on GG induces a metric on Pr(G)P_{r}(G), and GG acts faithfully, cocompactly on Pr(G)P_{r}(G) with finite stabilizers. By Milnor–Schwarz, GG is quasi-isometric to Pr(G)P_{r}(G). When GG is hyperbolic, then for any generating set and rr sufficiently large, Pr(G)P_{r}(G) is in fact contractible [GdlH90]. In particular, if GG is torsion-free then the action of GG on Pr(G)P_{r}(G) is free and thus for rr sufficienty large Pr(G)/GP_{r}(G)/G is a K(G,1)K(G,1). In this case, Bestvina–Mess show that the proper homotopy type of Pr(G)P_{r}(G) is an invariant of GG [BM91].

Proposition 6.6 (Flatness Criterion).

If Z+Z^{+} and ZZ^{-} are simply connected, then the embedding ΛG\Lambda\hookrightarrow\partial_{\infty}G is locally flat, hence conjugate by a homeomorphism to a standard embedding Sn2Sn1S^{n-2}\hookrightarrow S^{n-1}.

Proof.

Let ZZ be either Z+Z^{+} or ZZ^{-}. Since HH acts on G\partial_{\infty}G and preserves its limit set Λ\Lambda, it also preserves the complement Z+ZZ^{+}\coprod Z^{-}. If the action of HH does not preserve ZZ, we may find an index 2 subgroup (which we also denote by HH and which has the same limit set), whose action preserves ZZ. Since ZZ simply connected, it is contractible by Lemma 6.4. As HH is torsion-free, a result of Swenson [Swe01, Main Theorem (3)] implies that HH acts freely, properly discontinuously, and cocompactly on ZZ. In particular, the quotient \faktorZH\faktor{Z}{H} is aspherical and thus a compact K(H,1)K(H,1).

Since any K(H,1)K(H,1) is unique up to homotopy equivalence, \faktorZH\faktor{Z}{H} is homotopy equivalent to any other K(H,1)K(H,1), notably the quotient of the Rips complex Pd(H)P_{d}(H) by HH for d1d\gg 1. Taking universal coverings, this homotopy equivalence lifts to a proper, HH-equivariant homotopy equivalence between ZZ and Pd(H)P_{d}(H).

Choosing a basepoint z0Zz_{0}\in Z, the orbit map HHz0H\rightarrow H\cdot z_{0} identifies the boundary at infinity of ZZ (regarded as a δ\delta-hyperbolic space) with Λ\Lambda. Since ZZ is an ANR, we can apply [BLW10, Theorem 2.4] to obtain that Λ\Lambda is kk-LCC in Z¯\overline{Z}. Recalling that ZZ was either Z+Z^{+} or ZZ^{-}, we see that Λ\Lambda is kk-LCC in both Z+¯\overline{Z^{+}} and Z¯\overline{Z^{-}}. By [DV09, Theorem 7.6.5], this implies that Λ\Lambda is locally flat in Z+¯Z¯=G\overline{Z^{+}}\cup\overline{Z^{-}}=\partial_{\infty}G. The generalised Schoenflies theorem now implies that the embedding ΛG\Lambda\hookrightarrow\partial_{\infty}G is conjugate by a homeomorphism to a standard embedding Sn2Sn1S^{n-2}\hookrightarrow S^{n-1} by Remark 2.10.∎

Suppose now that Gπ1(M)G\cong\pi_{1}(M) for some closed, orientable, aspherical manifold MM. We now apply the above criterion to give an alternative characterization of local flatness of H\partial_{\infty}H in terms of the cover of MM corresponding to HH. In this setting, we refer to HH as 2-sided if the action of HH on G\partial_{\infty}G preserves both components of GH\partial_{\infty}G\setminus\partial_{\infty}H.

Theorem 6.7.

Let MM be a closed, orientable, aspherical nn-manifold with G=π1(M)G=\pi_{1}(M) hyperbolic, GSn1\partial_{\infty}G\cong S^{n-1} and n6n\geq 6. Suppose HGH\leq G is a quasi-convex, codimension-1, 2-sided subgroup such that HSn2\partial_{\infty}H\cong S^{n-2}. If CHMC_{H}\rightarrow M denotes the cover associated to HH, then HG\partial_{\infty}H\subset\partial_{\infty}G is locally flat if and only if there exists a closed, orientable (n1)(n-1)-manifold NN such that CHN×C_{H}\cong N\times\mathbb{R}.

Proof.

Let Z±Z^{\pm} denote the two complementary components of GH\partial_{\infty}G\setminus\partial_{\infty}H. Since MM is orientable and HH is 2-sided, the action of HH on G\partial_{\infty}G preserves Z+Z^{+} and ZZ^{-} separately, acting on each by orientation-preserving homeomorphisms. The action of HH on V~=M~Z+Z\widetilde{V}=\widetilde{M}\cup Z^{+}\cup Z^{-} is free, properly discontinuous and cocompact by [Swe01]. In particular, p:V~Vp\colon\widetilde{V}\rightarrow V is a covering map onto a compact manifold VV whose interior is CHC_{H} and whose boundary is W+WW^{+}\sqcup W^{-}, where W±=\faktorZ±HW^{\pm}=\faktor{Z^{\pm}}{H} are closed and orientable. Since both W+W^{+} and WW^{-} are collared in VV, we have that Vint(V)=CHV\simeq\mathrm{int}(V)=C_{H}. The latter is aspherical and has fundamental group isomorphic to HH. Hence VV is a K(H,1)K(H,1).

If H\partial_{\infty}H is locally flat, then Z+,ZZ^{+},Z^{-} are both contractible. Since HH acts freely properly discontinuously on Z+,ZZ^{+},Z^{-}, this implies W+W^{+} and WW^{-} are also K(H,1)K(H,1)’s. To verify the inclusions W+,WVW^{+},W^{-}\hookrightarrow V are both homotopy equivalences, it is enough to check that they induce isomorphisms on π1\pi_{1}. The inclusions are surjective because p1(W+)=Z+p^{-1}(W^{+})=Z^{+} and p1(W)=Zp^{-1}(W^{-})=Z^{-} are both connected. They are injective because Z+Z^{+} and ZZ^{-} are simply connected. Thus VV is an hh-cobordism. Since HH is torsion-free hyperbolic, the Whitehead group Wh(H)\textrm{Wh}(H) vanishes by [BL12] so VV is an ss-cobordism. Since dim(W+)5\dim(W^{+})\geq 5, we conclude that VW+×[0,1]V\cong W^{+}\times[0,1]. Thus, CH=int(V)W+×C_{H}=\mathrm{int}(V)\cong W^{+}\times\mathbb{R} and we can take N=W+N=W^{+}.

Conversely, suppose CHN×C_{H}\cong N\times\mathbb{R} for some closed, orientable manifold NN. Thus the inclusion NVN\hookrightarrow V is a homotopy equivalence. Let V+VV^{+}\subset V be the submanifold bounded by N×{0}N\times\{0\} and W+W^{+}, and let RW+×[0,ε]R\cong W^{+}\times[0,\varepsilon] be a collar neighborhood of W+W^{+} in V+V^{+}. Since V+W+N×[0,)V^{+}\setminus W^{+}\cong N\times[0,\infty) and NN is compact, there exists an injective map f:V+V+f\colon V^{+}\rightarrow V^{+} which is the identity on W+W^{+}, and whose image lies in RR. We compose ff with the projection of RR onto W+W^{+} to obtain a retraction r:V+W+r\colon V^{+}\rightarrow W^{+}. Therefore, the inclusion ι:W+V+\iota\colon W^{+}\hookrightarrow V^{+} induces an injection of π1(W+)\pi_{1}(W^{+}) into π1(V+)π1(N)\pi_{1}(V^{+})\cong\pi_{1}(N). On the other hand, since Hπ1(N)H\cong\pi_{1}(N) is the group of deck transformations of the cover Z+W+Z^{+}\rightarrow W^{+} we have a short exact sequence:

1π1(Z+)π1(W+)ιπ1(N)1.1\rightarrow\pi_{1}(Z^{+})\rightarrow\pi_{1}(W^{+})\xrightarrow{\iota_{*}}\pi_{1}(N)\rightarrow 1.

Therefore, ι\iota_{*} is an isomorphism and π1(Z+)=1\pi_{1}(Z^{+})=1. The same argument shows π1(Z)=1\pi_{1}(Z^{-})=1 as well, whence H\partial_{\infty}H is locally flat by Proposition 6.6. ∎

6.2. Embedded submanifolds from codimension-1 subgroups

Let MM be a closed, orientable aspherical nn-manifold with cubulated hyperbolic fundamental group G=π1(M)G=\pi_{1}(M). Suppose there exists a quasi-convex subgroup HGH\leq G such that Hπ1(N)H\cong\pi_{1}(N) for some closed orientable aspherical (n1)(n-1)-manifold NN.

Definition 6.8.

A subgroup HGH\leq G is square-root closed if g2Hg^{2}\in H implies gHg\in H.

Lemma 6.9.

Let GG be a 1-ended torsion-free hyperbolic group. If HGH\leq G is malnormal and quasi-convex, then HH is square-root closed. Moreover, if HH is 1-ended and has codimension-1, then GG splits as an amalgamated product or HNN extension over HH.

Proof.

If g2Hg^{2}\in H then g2gHg1H\langle g^{2}\rangle\leq gHg^{-1}\cap H, and since g2g^{2} has infinite order malnormality implies this is only possible if gHg\in H. This proves the first statement. The second statement follows from a result of Kropholler [NR93, p.146, Theorem 4.9]. ∎

Our interest in GG splitting over a square-root closed subgroup HH stems from a theorem of Cappell [Cap76, Theorem 1] giving homotopical conditions for a manifold MM with π1(M)=G\pi_{1}(M)=G to contain an embedded, 2-sided, codimension-1 submanifold NN with π1(N)H\pi_{1}(N)\cong H.

First we introduce some terminology. Recall that a finite CW-complex is called an orientable Poincaré complex if its cohomology ring satisfies Poincaré duality with respect to the trivial \mathbb{Z}-coefficients. Clearly any space homotopy equivalent to a closed orientable manifold is an orientable Poincaré complex. Consider now the following setup. Let YY be a connected, orientable (n+1)(n+1)-dimensional Poincaré complex and j:XYj\colon X\rightarrow Y an embedding of a connected, orientable nn-dimensional sub-Poincaré complex with n4n\geq 4. Suppose further that j:π1(X)π1(Y)j_{*}\colon\pi_{1}(X)\rightarrow\pi_{1}(Y) is injective.

Definition 6.10.

Let XYX\subset Y be as above and let WW be a manifold. A homotopy equivalence f:WYf\colon W\rightarrow Y is splittable along XX if it is homotopic to a map gg which is transverse regular along XX (whence g1(X)g^{-1}(X) is embedded, codimension-1 submanifold of WW) and if gg restricted to both g1(X)g^{-1}(X) and g1(YX)g^{-1}(Y\setminus X) are homotopy equivalences. If YXY\setminus X has two components we require that the restriction of gg to each component of g1(YX)g^{-1}(Y\setminus X) be a homotopy equivalence.

Given a pair (Y,X)(Y,X) of Poincaré complexes, a manifold WW, and a homotopy equivalence f:WYf\colon W\rightarrow Y, the main result of Cappell [Cap76, Theorem 1] gives homotopical conditions for ff to be splittable. Crucially, one must assume that π1(X)\pi_{1}(X) is square-root closed in π1(Y)\pi_{1}(Y), and that π1(Y)\pi_{1}(Y) splits over π1(X)\pi_{1}(X) either as an amalgamated product or HNN-extension. The remaining assumptions are KK-theoretic in nature, and are satisfied vacuously when both π1(X)\pi_{1}(X) and π1(Y)\pi_{1}(Y) are torsion-free hyperbolic. In the setting where WW and XX are both aspherical manifolds and π1(X)\pi_{1}(X) satisfies the Borel conjecture, the conclusion of Cappell’s result is that WW contains an embedded, 2-sided submanifold homeomorphic to XX. We now apply Cappell’s result to our situation.

Theorem 6.11.

Let MM be a closed, orientable aspherical nn-manifold with G=π1(M)G=\pi_{1}(M) cubulated hyperbolic and n6n\geq 6. Let HGH\leq G be a quasi-convex, codimension-1 subgroup such that Hπ1(N)H\cong\pi_{1}(N) for some closed aspherical (n1)(n-1)-manifold NN. Then there exist finite covers M^M\widehat{M}\rightarrow M, N^N\widehat{N}\rightarrow N and an embedding N^M^\widehat{N}\hookrightarrow\widehat{M}.

Proof.

By Theorem 2.3, there exist finite index subgroups G^G\widehat{G}\leq G and H^H\widehat{H}\leq H such that H^G^\widehat{H}\leq\widehat{G} is a malnormal subgroup. Let M^\widehat{M} and N^\widehat{N} be the corresponding finite covers of MM and NN, respectively. By Lemma 6.9, H^\widehat{H} is square-root closed and G^\widehat{G} splits over H^\widehat{H} as either an amalgamated product or HNN extension. In other words, G^\widehat{G} acts on a tree with edge stabilizers conjugate to H^\widehat{H} and a single edge orbit. Since H^\widehat{H} is quasi-convex, so are the vertex stabilizers [Bow98, Proposition 1.2]. In particular, being torsion-free hyperbolic, they have finite-dimensional classifying spaces (e.g. the quotient of the Rips complex).

We can then build a finite-dimensional classifying space for G^\widehat{G} from classifying spaces for the vertex stabilizers and N^\widehat{N}. Explicitly, if G^=G0H^G1\widehat{G}=G_{0}*_{\widehat{H}}G_{1}, then let X0X_{0} be a K(G0,1)K(G_{0},1) and X1X_{1} be a K(G1,1)K(G_{1},1). Now we obtain XG^X_{\widehat{G}} as the identification space from X1X_{1}, X2X_{2} and N^×[0,1]\widehat{N}\times[0,1] by gluing N^×{i}\widehat{N}\times\{i\} to XiX_{i} by a map inducing the inclusion H^Gi\widehat{H}\hookrightarrow G_{i}, for i=0,1.i=0,1. The construction for an HNN extension is similar. XG^X_{\widehat{G}} is a Poincaré duality complex since is it homotopy equivalent to M^\widehat{M}, and it contains N^\widehat{N} as an embedded, 2-sided submanifold.

Since all groups involved are torsion-free hyperbolic, their Whitehead groups vanish, hence condition (ii) in Theorem 1 of [Cap76] is automatically satisfied. Thus, all hypotheses of Cappell’s theorem are satisfied and any homotopy equivalence f:M^XG^f\colon\widehat{M}\rightarrow X_{\widehat{G}} is splittable. In particular, we find an embedded, 2-sided, codimension-1 manifold NM^N^{\prime}\hookrightarrow\widehat{M} such that π1(N)\pi_{1}(N^{\prime}) injects as H^G^\widehat{H}\leq\widehat{G}. Finally, since dim(N^)5\dim(\widehat{N})\geq 5, we have that NN^{\prime} is homeomorphic to N^\widehat{N} by Theorem 2.4. ∎

Appendix A Proofs of Lemma 3.16 and Lemma 3.17

Lemma A.1 (Lemma 3.16).

Let (μ,K,𝒰)(\mu,K,\mathcal{U}) be a refinement of (μ,K,𝒰)(\mu^{\prime},K^{\prime},\mathcal{U}^{\prime}) and suppose UUU\subset U^{\prime}, where U𝒰U\in\mathcal{U} and U𝒰U^{\prime}\in\mathcal{U}^{\prime}. Then Stabμ(U)<Stabμ(U)\operatorname{Stab}_{\mu}(U)<\operatorname{Stab}_{\mu^{\prime}}(U^{\prime}).

Proof.

Suppose gStabμ(U)g\notin\operatorname{Stab}_{\mu^{\prime}}(U^{\prime}), that is there exists some V𝒰V^{\prime}\in\mathcal{U}^{\prime} such that UV\partial U^{\prime}\neq\partial V^{\prime} and μ(g)(U)=V\mu^{\prime}(g)(\partial U^{\prime})=\partial V^{\prime}. Since μ\mu and μ\mu^{\prime} coincide on KK^{\prime}, we conclude that μ(g)(U)=V\mu(g)(\partial U^{\prime})=\partial V^{\prime}. If U=UU=U^{\prime}, then this implies that μ(g)\mu(g) does not preserve U\partial U and gStabμ(U)g\notin\operatorname{Stab}_{\mu}(U).

Now suppose UUU\subsetneq U^{\prime} and suppose by contradiction that gStabμ(U)g\in\operatorname{Stab}_{\mu}(U). Since μ(g)(U)=V\mu(g)(\partial U^{\prime})=\partial V^{\prime} and UV\partial U^{\prime}\neq\partial V^{\prime}, there exists some pUp_{\infty}\in\partial U^{\prime} such that

μ(g)(p)=μ(g)(p)VU.\mu(g)(p_{\infty})=\mu^{\prime}(g)(p_{\infty})\in\partial V^{\prime}\setminus\partial U^{\prime}.

Choose a point pUp\in U and a path γ\gamma from pp to pp_{\infty} that meets U\partial U^{\prime} only in its endpoint pp_{\infty}. (Such a path may be found by exploiting the homeomorphism ΦU:(Dn,Sn1)(U¯,U)\Phi^{U^{\prime}}:(D^{n},S^{n-1})\rightarrow(\overline{U^{\prime}},\partial U^{\prime}).) The path γ\gamma passes through a sequence of elements U=U0,U1,U2,𝒰U=U_{0},U_{1},U_{2},\dots\in\mathcal{U}. In between these open sets γ\gamma meets KK in a sequence of path-segments S0,S1,KUS_{0},S_{1},\dots\subset K\cap U^{\prime}, where SiS_{i} is the segment of γ\gamma between its time in UiU_{i} and its time in Ui+1U_{i+1}. We obtain a decomposition of the path γ\gamma into segments that are contained in U0,S0,U1,S1,U_{0},S_{0},U_{1},S_{1},\dots in that order. If this sequence is finite, then it ends in SMS_{M} for some MM. However, this sequence may go on forever. We denote the start and end point of SiS_{i} by pip_{i} and qiq_{i} respectively.

Let Vi:=μ(g)(Ui)V_{i}:=\mu(g)(U_{i}) and Ti:=μ(g)(Si)T_{i}:=\mu(g)(S_{i}). We will prove by induction that Vi,TiUV_{i},T_{i}\subset U^{\prime} for all ii (except for the last TMT_{M} which, if it exists, will be contained in U¯\overline{U^{\prime}}). The induction starts with V0V_{0}. Since we assume, by contradiction, that gStabμ(U)g\in\operatorname{Stab}_{\mu}(U), we have that V0=μ(g)(U0)=μ(g)(U)=UV_{0}=\mu(g)(U_{0})=\mu(g)(U)=U which is a subset of UU^{\prime} by assumption.

Now suppose μ(g)(Ui)U\mu(g)(U_{i})\subset U^{\prime} and suppose SiS_{i} is not the last segment in the sequence U0,S0,U_{0},S_{0},\dots. Since γ\gamma meets U\partial U^{\prime} only in its endpoint and SiS_{i} is not the last segment, piUp_{i}\in U^{\prime} and thus piUiUp_{i}\in\partial U_{i}\cap U^{\prime}. By induction assumption, μ(g)(Ui)U\mu(g)(U_{i})\subset U^{\prime} and thus μ(g)(pi)μ(g)(Ui¯)U¯\mu(g)(p_{i})\in\mu(g)(\overline{U_{i}})\subset\overline{U^{\prime}}. Since piKp_{i}\notin K^{\prime}, UK\partial U^{\prime}\subset K^{\prime}, and μ(g)\mu(g) sends KK^{\prime} homeomorphically to itself, we conclude that μ(g)(pi)U\mu(g)(p_{i})\in U^{\prime}.

Since γ\gamma meets U\partial U^{\prime} only in its endpoint and SiS_{i} is not the last segment, we have that SiK=S_{i}\cap K^{\prime}=\emptyset. Since μ(g)\mu(g) sends KK^{\prime} homeomorphically to itself, we conclude that TiK=T_{i}\cap K^{\prime}=\emptyset. Combined with the fact that the starting point μ(g)(pi)\mu(g)(p_{i}) of TiT_{i} lies in UU^{\prime}, we conclude that TiUT_{i}\subset U^{\prime}.

We now use this to show that Vi+1UV_{i+1}\subset U^{\prime}. Since TiUT_{i}\subset U^{\prime}, we have that μ(g)(qi)U\mu(g)(q_{i})\in U^{\prime}. At the same time, μ(g)(qi)Ui+1\mu(g)(q_{i})\in\partial U_{i+1} and thus μ(g)(qi)μ(g)(Ui+1¯)U=Vi+1¯U\mu(g)(q_{i})\in\mu(g)(\overline{U_{i+1}})\cap U^{\prime}=\overline{V_{i+1}}\cap U^{\prime}. Since every element in 𝒰\mathcal{U} is contained in an element of 𝒰\mathcal{U}^{\prime} and Vi+1¯U\overline{V_{i+1}}\cap U^{\prime}\neq\emptyset, we conclude that Vi+1UV_{i+1}\subset U^{\prime}.

This induction shows that ViV_{i} and TiT_{i} are contained in UU^{\prime} for all ii (except for TMT_{M}, if it is the last segment of the sequence). Depending on whether the sequence U0,S0,U_{0},S_{0},\dots is finite or infinite, we now finish the proof in two different ways.

We first deal with the case where the sequence is finite and thus we have a segment TMT_{M}. By induction, we know that VMUV_{M}\subset U^{\prime}. By the same argument as before, we see that pMp_{M} either lies in UU^{\prime}, or is the endpoint of γ\gamma. If it is the endpoint of γ\gamma, then pMUM¯p_{M}\in\overline{U_{M}} and thus μ(g)(pM)μ(g)(UM¯)U¯\mu(g)(p_{M})\in\mu(g)(\overline{U_{M}})\subset\overline{U^{\prime}}. Now suppose pMp_{M} is not the endpoint of γ\gamma. In that case, SMS_{M} is a segment that starts in UU^{\prime}, that only meets UK\partial U^{\prime}\subset K^{\prime} in its endpoint, and whose starting point is sent to a point in

μ(g)(UM¯U)U¯K=U.\mu(g)(\overline{U_{M}}\cap U^{\prime})\subset\overline{U^{\prime}}\setminus K^{\prime}=U^{\prime}.

We conclude that TMT_{M} is contained in UU^{\prime} except for its endpoint which lies in U\partial U^{\prime}. Since the endpoint of TMT_{M} is the endpoint of γ\gamma, which is pp_{\infty}, this implies that μ(g)(p)U\mu(g)(p_{\infty})\in\partial U^{\prime}. However, we chose pp_{\infty} so that μ(g)(p)U\mu(g)(p_{\infty})\notin\partial U^{\prime}, a contradiction.

Now suppose the sequence U0,S0,U_{0},S_{0},\dots is infinite. In that case, Vi,TiV_{i},T_{i} are contained in UU^{\prime} for every ii. Since γ\gamma is continuous, the endpoints qiq_{i} of the segments SiS_{i} converge to the point pp_{\infty}. Since μ(g)\mu(g) is continuous, we conclude that

μ(g)(qi)μ(g)(p)VU.\mu(g)(q_{i})\rightarrow\mu(g)(p_{\infty})\in\partial V^{\prime}\setminus\partial U^{\prime}.

However, μ(g)(qi)U\mu(g)(q_{i})\in U^{\prime} for all ii and thus their limit has to lie in U¯\overline{U^{\prime}}. This is a contradiction. Since we obtain a contradiction in both cases, we conclude that gg cannot lie in Stabμ(U)Stabμ(U)\operatorname{Stab}_{\mu}(U)\setminus\operatorname{Stab}_{\mu^{\prime}}(U^{\prime}), which completes the proof. ∎

Lemma A.2 (Lemma 3.17).

Let (μ,K,𝒰)(\mu,K,\mathcal{U}) be a GG-complex which is a refinement of a convergence GG-complex (μ,K,𝒰)(\mu^{\prime},K^{\prime},\mathcal{U}^{\prime}). If for every U𝒰U^{\prime}\in\mathcal{U}^{\prime}, the group μ(Stabμ(U))\mu(\operatorname{Stab}_{\mu^{\prime}}(U^{\prime})) is a convergence group on U¯K\overline{U^{\prime}}\cap K, then (μ,K,𝒰)(\mu,K,\mathcal{U}) is a convergence GG-complex as well.

Proof.

Let gmGg_{m}\in G be a sequence of distinct elements. Since (μ,K,𝒰)(\mu^{\prime},K^{\prime},\mathcal{U}^{\prime}) is a convergence GG-complex, there exist a,bSn1a,b\in S^{n-1} and a subsequence of gmg_{m} (which we denote by gmg_{m} again) such that μ(gm)a\mu^{\prime}(g_{m})\rightarrow a uniformly on compact subsets of K{b}K^{\prime}\setminus\{b\}. This implies that, for the inverse sequence, we have μ(gm)1=μ(gm1)b\mu^{\prime}(g_{m})^{-1}=\mu^{\prime}(g_{m}^{-1})\rightarrow b uniformly on compact subsets of K{a}K^{\prime}\setminus\{a\}.

Step 1: Using the convergence action on KK^{\prime} to get parts of the convergence behavior on KK.

Since the actions of μ\mu and μ\mu^{\prime} coincide on KKK^{\prime}\subset K, we conclude that μ(gm)a\mu(g_{m})\rightarrow a and μ(gm)1b\mu(g_{m})^{-1}\rightarrow b on compact subsets of K{b}K^{\prime}\setminus\{b\} and K{a}K^{\prime}\setminus\{a\} respectively. We thus obtain the following two properties: For every compact set CK{b}C\in K\setminus\{b\} and every ϵ>0\epsilon>0, there exists m(ϵ,C)m^{\prime}(\epsilon,C)\in\mathbb{N} such that

(1) mm(ϵ,C):d(μ(gm)(CK),a)<ϵ.\forall m\geq m^{\prime}(\epsilon,C):d(\mu(g_{m})(C\cap K^{\prime}),a)<\epsilon.

Similarly, for every compact set DK{a}D\in K\setminus\{a\} and every δ>0\delta>0, there exists m′′(δ,D)m^{\prime\prime}(\delta,D) such that

(2) mm′′(δ,D):d(μ(gm1)(DK),b)<δ.\forall m\geq m^{\prime\prime}(\delta,D):d(\mu(g_{m}^{-1})(D\cap K^{\prime}),b)<\delta.

We need to show that for every compact set CK{b}C\in K\setminus\{b\} and every ϵ>0\epsilon>0, there exists m(ϵ,C)m(\epsilon,C) such that

mm(ϵ,C):d(μ(gm)(C),a)<ϵ.\forall m\geq m(\epsilon,C):d(\mu(g_{m})(C),a)<\epsilon.

Step 2: Exhausting K{b}K\setminus\{b\} by well-behaved compact sets.

For α(0,1)\alpha\in(0,1), we define

Cα:=(KBα(0))(Sn1int(B1α(b))C_{\alpha}:=(K\cap B_{\alpha}(0))\cup(S^{n-1}\setminus\mathrm{int}(B_{1-\alpha}(b))

where Br(x)B_{r}(x) denotes the closed ball in DnD^{n} of radius rr with respect to the euclidean metric centered at xx and int(Br(x))\mathrm{int}(B_{r}(x)) its interior.

The collection CαC_{\alpha} is an exhaustion of K{b}K\setminus\{b\} by compact sets. We may thus assume without loss of generality that C=CαC=C_{\alpha} for some α(0,1)\alpha\in(0,1). Note that, for 1α1-\alpha sufficiently small, CαC_{\alpha} has the following property: If CαUC_{\alpha}\cap U^{\prime}\neq\emptyset for some U𝒰U^{\prime}\in\mathcal{U}^{\prime}, then either CαUint(Dn)C_{\alpha}\cap\partial U^{\prime}\cap\mathrm{int}(D^{n})\neq\emptyset or U=int(Dn)U^{\prime}=\mathrm{int}(D^{n}) and U=Sn1\partial U^{\prime}=S^{n-1}. (This property is crucial to the proof and we highlight that it holds for DnD^{n} for the same reason as it does for D3D^{3}.) If there exists U𝒰U^{\prime}\in\mathcal{U}^{\prime} such that U=int(Dn)U^{\prime}=\mathrm{int}(D^{n}), then K=Sn1K^{\prime}=S^{n-1} and (μ,K,𝒰(\mu,K,\mathcal{U} is a convergence GG-complex by the assumption in the Lemma. We thus assume from now on that KSn1K^{\prime}\neq S^{n-1}. We define compact sets DαD_{\alpha} that form an exhaustion of K{a}K\setminus\{a\} in the analogous way.

Let ϵ>0\epsilon>0. Since (K,𝒰)(K^{\prime},\mathcal{U}^{\prime}) satisfies ()(\star) (see Remark 3.3), there are only finitely many sets W1,,WL𝒰W_{1},\dots,W_{L}\in\mathcal{U}^{\prime} whose diameter is ϵ2\geq\frac{\epsilon}{2}.

Step 3: There exist M(α),N(ϵ,α)M(\alpha),N(\epsilon,\alpha)\in\mathbb{N} and sets Vj𝒰V_{j}\in\mathcal{U}^{\prime}, where j{1,,M(α)}j\in\{1,\dots,M(\alpha)\} such that the following holds:

If V𝒰V\in\mathcal{U}^{\prime} satisfies

  • V¯Cα\overline{V}\cap C_{\alpha}\neq\emptyset,

  • there exists m>N(ϵ,α)m>N(\epsilon,\alpha) such that μ(gm)(V)=Wi\mu(g_{m})(V)=W_{i} for some 1iL1\leq i\leq L,

then V=VjV=V_{j} for some jj.

By construction, CαC_{\alpha} does not intersect the open ball in DnD^{n} of radius 1α1-\alpha centered at bb. Put δ:=1α\delta:=1-\alpha. We obtain the sets ViV_{i} as follows: Since (K,𝒰)(K^{\prime},\mathcal{U}^{\prime}) satisfies ()(\star), there exist only finitely many sets V1,,VM𝒰V_{1},\dots,V_{M}\in\mathcal{U}^{\prime} with diameter δ2\geq\frac{\delta}{2}. We set the number of these sets to be M=M(α)M=M(\alpha).

Choose α\alpha^{\prime} such that DαD_{\alpha^{\prime}} intersects each WiW_{i} (thus α=α(ϵ)\alpha^{\prime}=\alpha^{\prime}(\epsilon), as the WiW_{i} depend only on ϵ\epsilon). Furthermore, choose α\alpha^{\prime} large enough such that, whenever DαUD_{\alpha^{\prime}}\cap U^{\prime}\neq\emptyset, then DαUint(Dn)D_{\alpha^{\prime}}\cap\partial U^{\prime}\cap\mathrm{int}(D^{n})\neq\emptyset. (This depends only on the generalized cell decomposition (K,𝒰)(K^{\prime},\mathcal{U}^{\prime}) and the fact that we already dealt with the case K=Sn1K^{\prime}=S^{n-1}.) Set D:=DαD:=D_{\alpha^{\prime}}. By equation (2), we know that for every m>m′′(δ2,D)m>m^{\prime\prime}(\frac{\delta}{2},D), we have

d(μ(gm1)(DK),b)<δ2.d(\mu(g_{m}^{-1})(D\cap K^{\prime}),b)<\frac{\delta}{2}.

We set N(ϵ,α):=m′′(δ2,D)N(\epsilon,\alpha):=m^{\prime\prime}(\frac{\delta}{2},D), which depends on α\alpha and ϵ\epsilon, as δ\delta and DD are determined by α\alpha and ϵ\epsilon respectively. We observe that, since DD intersects each WiW_{i} by assumption and α\alpha^{\prime} was chosen sufficiently large, it intersects Wiint(Dn)\partial W_{i}\cap\mathrm{int}(D^{n}). Thus we find piWiDp_{i}\in\partial W_{i}\cap D. Since WiK\partial W_{i}\subset K^{\prime}, equation (2) implies for every m>N(ϵ,α)m>N(\epsilon,\alpha) that

(3) d(μ(gm1)(pi),b)<δ2.d(\mu(g_{m}^{-1})(p_{i}),b)<\frac{\delta}{2}.

We claim that these choices of M(α)M(\alpha), N(ϵ,α)N(\epsilon,\alpha), V1,,VM(α)V_{1},\dots,V_{M(\alpha)} have the properties we require. Let V𝒰V\in\mathcal{U}^{\prime} such that μ(gm)(V)=Wi\mu(g_{m})(V)=W_{i} for some ii and some m>N(ϵ,α)m>N(\epsilon,\alpha). Suppose VVjV\neq V_{j} for all 1jM(α)1\leq j\leq M(\alpha). Then

δ2>diam(V)=diam(μ(gm1)(Wi))=diam(μ(gm1)(Wi¯)).\frac{\delta}{2}>\textrm{diam}(V)=\textrm{diam}(\mu(g_{m}^{-1})(W_{i}))=\textrm{diam}(\mu(g_{m}^{-1})(\overline{W_{i}})).

Combined with equation (3), this implies that

d(V¯,b)=d(μ(gm1)(Wi¯),b)<δ=1α.d(\overline{V},b)=d(\mu(g_{m}^{-1})(\overline{W_{i}}),b)<\delta=1-\alpha.

Therefore, V¯\overline{V} cannot intersect CαC_{\alpha} which proves Step 3.

Step 4: Produce a subsequence of (gm)m(g_{m})_{m} that shrinks the diameter of V\partial V for all but finitely many V𝒰V\in\mathcal{U}^{\prime} that intersect CαC_{\alpha}.

Choose some V𝒰V\in\mathcal{U}^{\prime} such that CαVC_{\alpha}\cap V\neq\emptyset. We partition the sequence (gm)m(g_{m})_{m} into L+1L+1 new sequences gmig_{m}^{i} for 0iL0\leq i\leq L as follows: The sequence gm0g_{m}^{0} consists of all elements of (gm)m(g_{m})_{m} such that

diam(μ(gm0)(V))ϵ2.\textrm{diam}(\mu(g_{m}^{0})(\partial V))\leq\frac{\epsilon}{2}.

For i1i\geq 1, (gmi)m(g_{m}^{i})_{m} consists of all elements of (gm)m(g_{m})_{m} that satisfy μ(gmi)(V)=Wi\mu(g_{m}^{i})(V)=W_{i}.

It is important to note that we index the elements gmig_{m}^{i} such that gmi=gmg_{m}^{i}=g_{m} for every mm. In turn, this means that gmig_{m}^{i} is defined only for certain mm\in\mathbb{N}. It is important that we index our elements in this way to formulate the following statement.

We claim that, if VV is not one of the sets VjV_{j} from Step 3, then gmig_{m}^{i} is not defined for any 1iL1\leq i\leq L and any m>N(ϵ,α)m>N(\epsilon,\alpha). Indeed, if VV is not any of the sets VjV_{j} and m>N(ϵ,α)m>N(\epsilon,\alpha), then VV cannot satisfy both properties required in Step 3. But CαVC_{\alpha}\cap V\neq\emptyset by assumption on VV and thus μ(gm)(V)Wi\mu(g_{m})(V)\neq W_{i} for all 1iL1\leq i\leq L.

Step 5: Estimate d(μ(gm0)(CαV),a)d(\mu(g_{m}^{0})(C_{\alpha}\cap V),a) for V𝒰V\in\mathcal{U}^{\prime}.

Let V𝒰V\in\mathcal{U}^{\prime} such that CαVC_{\alpha}\cap V\neq\emptyset. Since CαC_{\alpha} intersects VV and α\alpha was chosen sufficiently large, there exists qVCαint(Dn)q\in\partial V\cap C_{\alpha}\cap\mathrm{int}(D^{n}). Applying inequality (1), we obtain

m>m(ϵ2,Cα):d(μ(gm)(q),a)<ϵ2.\forall m>m^{\prime}\left(\frac{\epsilon}{2},C_{\alpha}\right):d(\mu(g_{m})(q),a)<\frac{\epsilon}{2}.

Restricting to the elements gm0g_{m}^{0}, we also have

diam(μ(gm0)(V¯))ϵ2.\textrm{diam}\left(\mu\left(g_{m}^{0}\right)(\overline{V})\right)\leq\frac{\epsilon}{2}.

Combining these two inequalities, we obtain

(4) m>m(ϵ2,Cα):d(μ(gm0)(CαV),a)<ϵ.\forall m>m^{\prime}\left(\frac{\epsilon}{2},C_{\alpha}\right):d(\mu(g_{m}^{0})(C_{\alpha}\cap V),a)<\epsilon.

This is the equality we desire for all gmg_{m} (and all mm greater than some m0m_{0}). We are left to show that this inequality also holds for the elements gmig_{m}^{i} when mm is sufficiently large.

If VVjV\neq V_{j} for all 1jL1\leq j\leq L, then gmig_{m}^{i} is not defined for any m>N(ϵ,α)m>N(\epsilon,\alpha) by Step 4 and we are done.

Let 1jL1\leq j\leq L and suppose V=VjV=V_{j}. Let mim_{i} be the smallest mm such that gmig_{m}^{i} is defined. We define hmi:=(gmii)1gmih_{m}^{i}:=\left(g_{m_{i}}^{i}\right)^{-1}\cdot g_{m}^{i} and compute

μ(hmi)=μ(gmi)μ(gmii)1.\mu(h_{m}^{i})=\mu(g_{m}^{i})\circ\mu(g_{m_{i}}^{i})^{-1}.

(Recall our convention that μ(gh)=μ(h)μ(g)\mu(gh)=\mu(h)\circ\mu(g).) We conclude that μ(hmi)(Wi)=Wi\mu(h_{m}^{i})(W_{i})=W_{i} and thus hmiStabμ(Wi)h_{m}^{i}\in\operatorname{Stab}_{\mu^{\prime}}(W_{i}). Since μ(gm)a\mu(g_{m})\rightarrow a uniformly on compact sets in K{b}K^{\prime}\setminus\{b\}, we have that μ(hmi)a\mu(h_{m}^{i})\rightarrow a uniformly on compact sets in K{bi}K^{\prime}\setminus\{b_{i}\}, where bi:=μ(gmii)(b)b_{i}:=\mu(g_{m_{i}}^{i})(b).

Suppose, gmig_{m}^{i} did not converge to aa uniformly on CαVjC_{\alpha}\cap V_{j}. Then there exists an infinite subsequence (hmki)k(h_{m_{k}}^{i})_{k} and a sequence of points zmkμ(gmii)(Cα)Wiz_{m_{k}}\in\mu(g_{m_{i}}^{i})(C_{\alpha})\cap W_{i} such that

m:d(μ(hmki)(zmk),a)ϵ.\forall m:d(\mu(h_{m_{k}}^{i})(z_{m_{k}}),a)\geq\epsilon.

We first observe that (zmk)k(z_{m_{k}})_{k} cannot converge to bib_{i}. Indeed, if zmkbiz_{m_{k}}\rightarrow b_{i}, then μ(gmii)1(zmk)b\mu(g_{m_{i}}^{i})^{-1}(z_{m_{k}})\rightarrow b while μ(gmii)1(zmk)Cα\mu(g_{m_{i}}^{i})^{-1}(z_{m_{k}})\in C_{\alpha} for every kk. Since CαC_{\alpha} is closed by construction, this implies that bCαb\in C_{\alpha}. But bCαb\notin C_{\alpha} so this is a contradiction and we conclude that (zmk)k(z_{m_{k}})_{k} cannot converge to bb.

By assumption, μ(Stabμ(Wi))\mu(\operatorname{Stab}_{\mu^{\prime}}(W_{i})) is a convergence group on Wi¯K\overline{W_{i}}\cap K. Therefore, we may pass to a subsequence of (hmki)k(h_{m_{k}}^{i})_{k} (which we denote by (hmki)k(h_{m_{k}}^{i})_{k} again) and find points a,bWi¯Ka^{\prime},b^{\prime}\in\overline{W_{i}}\cap K such that μ(hmki)a\mu(h_{m_{k}}^{i})\rightarrow a^{\prime} uniformly on compact subsets of (Wi¯K){b}(\overline{W_{i}}\cap K)\setminus\{b^{\prime}\}.

On the other hand, we know that μ(hmi)a\mu(h_{m}^{i})\rightarrow a uniformly on compact sets in K{bi}K^{\prime}\setminus\{b_{i}\}, in particular on compact sets in Wi{bi}\partial W_{i}\setminus\{b_{i}\}. Since WiSn1\partial W_{i}\approx S^{n-1} contains infinitely many points, uniform convergence on compact sets in Wi{b,bi}\partial W_{i}\setminus\{b^{\prime},b_{i}\} yields that a=aa=a^{\prime} and bi=bb_{i}=b^{\prime}. Therefore, μ(hmki)a\mu(h_{m_{k}}^{i})\rightarrow a uniformly on compact subsets of (Wi¯K){bi}(\overline{W_{i}}\cap K)\setminus\{b_{i}\}. Since zmkWiKz_{m_{k}}\in W_{i}\cap K and zmkz_{m_{k}} does not converge to bib_{i}, we conclude that there exists mi(ϵ,Cα,Vj)m_{i}(\epsilon,C_{\alpha},V_{j}) such that

m>mi(ϵ,Cα,Vj):d(μ(hmki(zmk),a)<ϵ.\forall m>m_{i}(\epsilon,C_{\alpha},V_{j}):d(\mu(h_{m_{k}}^{i}(z_{m_{k}}),a)<\epsilon.

This is a contradiction and we conclude that there exists mi(ϵ,Cα,Vj)m_{i}(\epsilon,C_{\alpha},V_{j}) such that

m>mi(ϵ,Cα,Vj):d(μ(gmi(CαVj),a)ϵ.\forall m>m_{i}(\epsilon,C_{\alpha},V_{j}):d(\mu(g_{m}^{i}(C_{\alpha}\cap V_{j}),a)\leq\epsilon.

Step 6: Concluding the Lemma.

Set

m(ϵ,α):=max{m(ϵ2,Cα),N(ϵ,α),m1(ϵ,Cα,V1),,mM(α)(ϵ,Cα,VM(α))}.m(\epsilon,\alpha):=\max\left\{m^{\prime}\left(\frac{\epsilon}{2},C_{\alpha}\right),N(\epsilon,\alpha),m_{1}(\epsilon,C_{\alpha},V_{1}),\dots,m_{M(\alpha)}(\epsilon,C_{\alpha},V_{M(\alpha)})\right\}.

By the estimates obtained in Step 5, we have for every V𝒰V\in\mathcal{U} such that CαVC_{\alpha}\cap V\neq\emptyset that

mm(ϵ,α):d(μ(gm(CαV),a)ϵ.\forall m\geq m(\epsilon,\alpha):d(\mu(g_{m}(C_{\alpha}\cap V),a)\leq\epsilon.

Combined with inequality (1) this implies that (μ,K,𝒰)(\mu,K,\mathcal{U}) is a convergence GG-complex, which completes the proof. ∎

References

  • [Ago13] Ian Agol. The virtual Haken conjecture. Doc. Math., 18:1045–1087, 2013. With an appendix by Agol, Daniel Groves, and Jason Manning.
  • [AT88] B. N. Apanasov and A. V. Tetenov. Nontrivial cobordisms with geometrically finite hyperbolic structures. J. Differential Geom., 28(3):407–422, 1988.
  • [BL12] Arthur Bartels and Wolfgang Lück. The Borel conjecture for hyperbolic and CAT(0){\rm CAT}(0)-groups. Ann. of Math. (2), 175(2):631–689, 2012.
  • [BLW10] Arthur Bartels, Wolfgang Lück, and Shmuel Weinberger. On hyperbolic groups with spheres as boundary. Journal of Differential Geometry, 86(1), 2010.
  • [BM91] Mladen Bestvina and Geoffrey Mess. The boundary of negatively curved groups. J. Amer. Math. Soc., 4(3):469–481, 1991.
  • [Bow98] Brian H. Bowditch. Cut points and canonical splittings of hyperbolic groups. Acta Math., 180(2):145–186, 1998.
  • [BR13] Taras Banakh and Dušan Repovš. Universal nowhere dense subsets of locally compact manifolds. Algebr. Geom. Topol., 13(6):3687–3731, 2013.
  • [Bro60] Morton Brown. A proof of the generalized Schoenflies theorem. Bull. Amer. Math. Soc., 66:74–76, 1960.
  • [BW12] Nicolas Bergeron and Daniel T. Wise. A boundary criterion for cubulation. Amer. J. Math., 134(3):843–859, 2012.
  • [Can73] J. W. Cannon. A positional characterization of the (n1)(n-1)-dimensional Sierpiński curve in Sn(n4)S^{n}(n\not=4). Fund. Math., 79(2):107–112, 1973.
  • [Cap76] Sylvain E. Cappell. A splitting theorem for manifolds. Invent. Math., 33(2):69–170, 1976.
  • [DV09] Robert J. Daverman and Gerard Venema. Embeddings in Manifolds. American Mathematical Soc., Heidelberg, 2009. ISBN 978-0-8218-3697-.
  • [Edw84] Robert D. Edwards. The solution of the 44-dimensional annulus conjecture (after Frank Quinn). In Four-manifold theory (Durham, N.H., 1982), volume 35 of Contemp. Math., pages 211–264. Amer. Math. Soc., Providence, RI, 1984.
  • [FJ89a] F. T. Farrell and L. E. Jones. Compact negatively curved manifolds (of dim 3,4\neq 3,4) are topologically rigid. Proc. Nat. Acad. Sci. U.S.A., 86(10):3461–3463, 1989.
  • [FJ89b] F. T. Farrell and L. E. Jones. Negatively curved manifolds with exotic smooth structures. J. Amer. Math. Soc., 2(4):899–908, 1989.
  • [FJ89c] F. T. Farrell and L. E. Jones. A topological analogue of Mostow’s rigidity theorem. J. Amer. Math. Soc., 2(2):257–370, 1989.
  • [GdlH90] Étienne Ghys and Pierre de la Harpe. Espaces métriques hyperboliques. In Sur les groupes hyperboliques d’après Mikhael Gromov (Bern, 1988), volume 83 of Progr. Math., pages 27–45. Birkhäuser Boston, Boston, MA, 1990.
  • [Gir17] Anne Giralt. Cubulation of Gromov-Thurston manifolds. Groups Geom. Dyn., 11(2):393–414, 2017.
  • [GMT03] David Gabai, G. Robert Meyerhoff, and Nathaniel Thurston. Homotopy hyperbolic 3-manifolds are hyperbolic. Ann. of Math. (2), 157(2):335–431, 2003.
  • [GT87] M. Gromov and W. Thurston. Pinching constants for hyperbolic manifolds. Invent. Math., 89(1):1–12, 1987.
  • [Hai15] Peter Haissinsky. Hyperbolic groups with planar boundaries. Invent. Math., 201(1):239–307, 2015.
  • [Hat02] Allen Hatcher. Algebraic topology. Cambridge: Cambridge University Press, 2002.
  • [KM12] Jeremy Kahn and Vladimir Markovic. Immersing almost geodesic surfaces in a closed hyperbolic three manifold. Ann. Math. (2), 175(3):1127–1190, 2012.
  • [KMZ24] Annette Karrer, Babak Miraftab, and Stefanie Zbinden. Subgroups arising from connected components in the morse boundary. ArXiv:2403.03939, 2024.
  • [KN13] Aditi Kar and Graham A. Niblo. A topological splitting theorem for Poincaré duality groups and high-dimensional manifolds. Geom. Topol., 17(4):2203–2221, 2013.
  • [Mar13] Vladimir Markovic. Criterion for Cannon’s conjecture. Geom. Funct. Anal., 23(3):1035–1061, 2013.
  • [Maz59] Barry Mazur. On embeddings of spheres. Bull. Amer. Math. Soc., 65:59–65, 1959.
  • [NR93] Graham A. Niblo and Martin A. Roller, editors. Geometric group theory. Vol. 1, volume 181 of London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge, 1993.
  • [Pal66] R. S. Palais. Homotopy theory of infinite dimensional manifolds. Topology, 5:1–16, 1966.
  • [Per02] G. Perelman. The entropy formula for the Ricci flow and its geometric applications. arXiv:0211159, 2002.
  • [Per03a] G. Perelman. Finite extinction time for the solutions to the Ricci flow on certain three-manifolds. arXiv:0307245, 2003.
  • [Per03b] G. Perelman. Ricci flow with surgery on three-manifolds. arXiv:0303109, 2003.
  • [Qui82] Frank Quinn. Ends of maps. III. Dimensions 44 and 55. J. Differential Geometry, 17(3):503–521, 1982.
  • [Sco83] Peter Scott. There are no fake Seifert fibre spaces with infinite π1\pi_{1}. Ann. of Math. (2), 117(1):35–70, 1983.
  • [Sta61] John Stallings. On fibering certain 33-manifolds. In Topology of 3-manifolds and related topics (Proc. The Univ. of Georgia Institute, 1961), pages 95–100. Prentice-Hall, Inc., Englewood Cliffs, NJ, 1961.
  • [Swe01] Eric L. Swenson. Quasi-convex groups of isometries of negatively curved spaces. volume 110, pages 119–129. 2001. Geometric topology and geometric group theory (Milwaukee, WI, 1997).
  • [You44] Gail S. Young, Jr. A generalization of Moore’s theorem on simple triods. Bull. Amer. Math. Soc., 50:714, 1944.