On a generalization of Cannon’s conjecture for cubulated hyperbolic groups
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 is a (Gromov) hyperbolic group whose boundary is homeomorphic to , then is commensurable up to finite quotients with a cocompact lattice in , the isometry group of real hyperbolic 3-space. More precisely, if the Cannon conjecture is true, then there exists a cocompact lattice and a short exact sequence
where is a finite group equal to the kernel of the action of on .
Markovic [Mar13] showed that the Cannon conjecture holds under a further assumption, namely that has enough quasi-convex surface subgroups to separate pairs of points in . This assumption implies in particular that is cubulated, i.e. 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 is cubulated, then has enough quasi-convex surface subgroups to separate pairs of points at infinity. Conversely, since every cocompact lattice 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 with is cubulated. The goal of the present work is to prove the following extension of Markovic’s result:
Theorem 1.1.
Let be a hyperbolic group such that and suppose contains quasi-convex, codimension-1 subgroups satisfying
-
(1)
The -translates of the limit sets separate pairs of points in .
-
(2)
is a locally flat 2-sphere in for .
Then there exists a finite-index, torsion-free subgroup such that , where is a closed, orientable -manifold covered by .
We recall that an embedding of a -manifold into an -manifold is locally flat if for every , there exists a neighbourhood of in such that is homeomorphic to . As in Markovic’s case, hypothesis (1) implies that is cubulated with respect to a subset of . In particular, each is itself cubulated, hence by Markovic’s result, each is virtually a lattice in . Since cubulated hyperbolic groups are residually finite, we may pass to a finite index subgroup of 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 . However, Apanasov–Tetenov [AT88] have constructed examples of discrete subgroups in that are isomorphic to closed hyperbolic 3-manifold groups but whose limit sets in are wild 2-spheres. As far as the authors are aware, such subgroups do not arise as quasi-convex subgroups of uniform lattices in , but the necessity of this assumption remains unclear.
The direct analog of Markovic’s result is false when . Indeed, for all , Gromov–Thurston [GT87] construct examples of closed, negatively curved -manifolds that are not homotopy equivalent to real hyperbolic -manifolds, but which were shown to be cubulated by Giralt [Gir17]. In fact, these manifolds contain enough totally geodesic real hyperbolic -manifolds to cubulate.
The Gromov–Thurston examples underscore that for , if is a torsion-free hyperbolic group with , the most one can hope for is that for some closed aspherical -manifold . Remarkably, even without the cubulation assumption, Bartels–Lück–Weinberger [BLW10] show exactly this: if is a torsion-free hyperbolic group with for , then for some closed aspherical -manifold . They remark that their proof ought to extend to the case as well, assuming certain surgery results hold. Therefore, Theorem 1.1 fills in a dimension where surgery techniques and the -cobordism theorem are unavailable. Nevertheless, using results of Bartels–Lück [BL12] on the Borel conjecture, our proof works equally well for all , hence we will prove the main theorem in this generality:
Theorem 1.2.
Let be a torsion-free hyperbolic group such that and suppose contains quasi-convex, codimension-1 subgroups satisfying:
-
(1)
The -translates of the separate pairs of points in .
-
(2)
is a locally flat -sphere in for .
If , then there exists a finite-index, torsion-free subgroup such that , where is a closed orientable -manifold covered by .
In the final section of the paper, we explore the local flatness condition in more depth. When and is a quasi-convex, codimension-1 subgroup such that , we show in Proposition 6.6 that is locally flat in if and only if both components of 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 is the interior of a compact, orientable 3-manifold that is homotopy equivalent to an orientable surface , then is homeomorphic to . In particular, it follows that if is a closed, orientable hyperbolic 3-manifold and is the cover associated to a quasi-convex surface subgroup, then for some surface . In our setting, with as above, we say that is 2-sided if the action of on preserves each component of . We prove:
Theorem 6.7.
Let be a closed, orientable, aspherical -manifold with hyperbolic, and . Suppose is a quasi-convex, codimension-1, 2-sided subgroup such that . If denotes the cover associated to , then is locally flat if and only if there exists a closed, orientable -manifold such that .
Our second application further assumes that is cubulated, and that for some closed orientable manifold . We prove that there exist finite covers of and of and , respectively, such that embeds in as a submanifold:
Theorem 6.11.
Let be a closed, orientable aspherical -manifold with cubulated hyperbolic and . Let be a quasi-convex, codimension-1 subgroup such that for some closed aspherical -manifold . Then there exist finite covers , , and an embedding .
We regard this result as a generalization of the fact that a quasi-convex, -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 -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 (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 -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 is cubulated hyperbolic. By results of Agol [Ago13], we know that is virtually special, hence virtually torsion-free. For simplicity, from now on we will assume that is a torsion-free hyperbolic group.
2.1. Malnormal cubulations
Recall that if is a hyperbolic group and is a quasi-convex subgroup, then there is a closed, topological embedding . If is connected, then has codimension-1 in if and only if is disconnected.
Definition 2.1.
Let be a collection of quasi-convex, codimension-1 subgroups of . We say that separates points at infinity if for any two points , there exists a conjugate such that and lie in different connected components of .
Bergeron–Wise [BW12] prove that if separates points at infinity then there exists a finite-dimensional CAT(0) cube complex on which acts geometrically, and whose hyperplane stabilizers are conjugates of elements of . For this reason, if is a collection of quasi-convex, codimension-1 subgroups that separates points at infinity, we will say that cubulates .
Definition 2.2.
A subgroup is called malnormal if for every .
Markovic showed that after passing to a cover, one can assume that each element of is malnormal:
Theorem 2.3 (Theorem 2.1, [Mar13]).
Suppose is hyperbolic and cubulated with respect to a collection of quasi-convex, codimension-1 subgroups . Then there exists a finite-index subgroup that is cubulated with respect to a collection of malnormal, quasi-convex, codimension-1 subgroups.
2.2. The Borel conjecture for hyperbolic groups
Let be a finitely generated group and let be a closed, aspherical manifold such that . In particular, is a finite-dimensional , and therefore is torsion-free. The Borel conjecture asks whether is unique up to homeomorphism. More precisely, if is another -manifold and is a homotopy equivalence, is necessarily homotopic to a homeomorphism?
The Borel conjecture holds in low dimensions as a consequence of the classification of manifolds of dimensions . 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 , Farrell–Jones [FJ89a, FJ89c] proved the Borel conjecture when 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 be a closed, aspherical manifold of dimension such that is hyperbolic. If is an -manifold and is a homotopy equivalence, then is homotopic to a homeomorphism.
Combining this with the low-dimensional results mentioned above we have:
Corollary 2.5.
Let be a closed, aspherical manifold of dimension such that is hyperbolic. If is an -manifold and is a homotopy equivalence, then is homotopic to a homeomorphism.
2.3. Local flatness and the generalized Schoenflies theorem
In contrast to the low-dimensional case where the properties of the embedding of 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 , and a - and -dimensional manifold respectively and let . An embedding is called locally flat at , if there exists a neighborhood of in such that is homeomorphic to . We say that is locally flat, if it is locally flat at every point in .
Definition 2.7.
Let be an -dimensional manifold, , and the closed unit ball in . We call an embedding a tame -ball, if the restriction of to is a locally flat embedding.
Definition 2.8.
Let be the -sphere. For , the standard in is the subspace formed as the intersection of with each of the hyperplanes for .
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 be a locally flat embedding and let be one of the two connected components of . Then there exists a homeomorphism .
Remark 2.10.
The generalized Schoenflies theorem implies that locally flat embeddings are conjugate by a homeomorphism to the standard embedding. Indeed, let and be the two connected components of . By the generalized Schoenflies theorem, both and are homeomorphic to the closed disk . In particular the glueing is homeomorphic to the glueing that produces an -dimensional sphere with a standard inside. Therefore, the homeomorphisms allow us to construct a homeomorphism , where is standard.
3. Review of -complexes
For , let denote the closed unit ball in and its boundary. We denote the interior by . In this section, we review and generalize the notion of -complexes used in [Mar13]. Our main-objective is to show that all the key results which Markovic proved for -complexes on the pair hold for the pair 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 be a compact metric space, and . Throughout this section, we use the notation
We will use the following terminology from [KMZ24].
Definition 3.1.
Let be as above and a sequence of subsets. We define the limit to be the collection of points obtained as limits of convergent sequences , where for every .
Note that if is a compact metric space, then a sequence converges to a singleton set , if and only if for every neighbourhood of , there exists such that for all .
Let be a subgroup. We say that is a convergence group if for every sequence of distinct in , there exists a subsequence and points such that uniformly on compact subsets of . (We denote the constant map that sends all of to by as well.) Note that in this case the sequence converges to uniformly on compact subsets of .
Let be a group. A -action on is a monomorphism (all our actions are faithful). Regarding group multiplication, we use the convention that .
3.1. Generalized cell decompositions and -complexes
Definition 3.2.
Let be a closed subset and let denote the collection of connected components of . We say the pair is a generalized cell decomposition of if the following holds:
-
•
.
-
•
Every open set is homeomorphic to and its boundary is homeomorphic to .
-
•
Every sequence of distinct elements in has a subsequence such that 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:
-
For every , there exists such that there are at most many elements in that have diameter greater than .
It is an easy exercise to see that this is equivalent to our third condition (using the fact that is compact). Therefore, this is an equivalent definition to the one given in [Mar13] and we exploit both 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 . Then consists of one element, the set . Second, we may consider . In this case, 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 , be two generalized cell decompositions of . We say that is a refinement of if . This defines a partial order on the collection of all generalized cell decompositions of .
We now introduce the notion of an action of a group on a generalized cell decomposition.
Definition 3.5.
Let be a group and a generalized cell decomposition of . A -action on is a -action such that
-
(1)
For every , .
-
(2)
For all , , there exists such that .
Note that every -action on induces a -action on . We will assume for all our -actions that they act by orientation preserving homeomorphisms on .
A -action induces an action on the collection of ‘cells’ . This follows from the following Lemma.
Lemma 3.6.
Let such that . Then .
In particular, if is a -action on a generalized cell decomposition , then for every , , there exists a unique such that . We write
which defines an action of on .
Proof.
Let such that . Since are connected components of , they are either disjoint or equal. Suppose they are disjoint. Since are open subsets of such that and are homeomorphic to , the union is homeomorphic to the closed manifold , where the left-hand-side denotes a glueing of two closed unit balls along a homeomorphism between their boundaries. Since is compact, it is an embedded copy of in . However, closed -dimensional manifolds cannot be embedded into (see for example [Hat02, Corollary 2B.4]). We obtain a contradiction and conclude that . ∎
Definition 3.7.
For , we define its stabilizer to be
Definition 3.8.
Let be a group, a generalized cell decomposition of , and a -action on . We call the triple a -complex, if for every there exists a homeomorphism such that the following two conditions are satisfied:
-
(1)
fixes every point in .
-
(2)
The collection is -equivariant on the boundary, that is for every ,
If we remember the choices of , we call the collection a marked -complex.
Definition 3.9.
Let be a -action on .
-
•
We say that is free if for every , has no fixed points in .
-
•
A -complex is called free if and only if the -action is free.
-
•
We say is a convergence -complex if and only if acts as a convergence group on .
Remark 3.10.
Let be a sequence of distinct elements in . If is a convergence -complex, then by definition we find two points and a subsequence such that uniformly on and uniformly on . Since preserves , we conclude that . This holds for any sequence of distinct elements in .
Remark 3.11.
If is torsion free, then any convergence -complex is also free for the following reason: Suppose is a convergence -complex and let . Since has no torsion, is a sequence of distinct elements in to which we can apply the convergence-property. We conclude that cannot have any fixed points in .
Definition 3.12.
Let be a -complex with marking and let be a homeomorphism. Define for all , and . The pushforward of is
with marking .
That satisfies (1) and (2) of Definition 3.8 is an exercise, noting that is actually a homeomorphism of pairs . Similarly, one can define a pullback of a -complex, but these are completely interchangeable for our purposes.
Proposition 3.13.
Being a convergence -complex or a free -complex is preserved under pushforward.
Proof.
Both of these are defined by properties of the -action in the compact-open topology and the fixed points of elements of in the boundary . They are both preserved by pushforward since is a homeomorphism of pairs . ∎
Proposition 3.14.
Set and let be a free convergence -complex. Then the quotient is an aspherical -dimensional manifold whose fundamental group is isomorphic to .
Proof.
Since is free, has no fixed point on for every . In other words, the action of on is free. Since is a convergence -complex, is a convergence group on and thus its action on is proper (w. r. t. the euclidean metric). Therefore, the quotient is an -dimensional manifold. Since is contractible, this quotient is aspherical and its fundamental group is isomorphic to . ∎
3.2. Refinements of -complexes
Definition 3.15.
Let and be two -complexes on . We say is a refinement of if and on .
The notion of refinement defines a partial order on the collection of all -complexes. Note that if is a refinement of , then is a refinement of .
The next two lemmas are key technical results concerning refinements of -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 be a refinement of and suppose , where and . Then .
The second key result concerns the behavior of convergence -complexes under refinement. The proof of this lemma is very similar to the proof in the -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 be a -complex which is a refinement of a convergence -complex . If for every , the group is a convergence group on , then is a convergence -complex as well.
3.3. Refining one -complex by another
Consider two marked -complexes and that agree on the boundary, i.e. suppose that for all ,
Markovic [Mar13] constructs the refinement of induced by in the -case. The construction works the same way in the -case and it goes as follows. We define
For every , the map is a homeomorphism and we define as the marking for . One easily checks that is the complement of the union of all elements in and thus we obtain that satisfies the first two conditions of a generalized cell decomposition of (see figure 1).
To conclude that is a generalized cell decomposition, we are left to show that every sequence of distinct elements in has a subsequence that converges to a point. Consider a sequence of distinct elements and suppose for , . There are two cases to consider.
If the sequence contains infinitely many distinct elements, we may choose a subsequence such that all are pairwise distinct. Since is a generalized cell decomposition, admits a subsequence (which we again denote by again) that converges to a point. Since , the sequence converges to that same point.
Now suppose the sequence contains only finitely many distinct elements. Then there exists some that for infinitely many . Consider the subsequence of all elements . Since all are pairwise distinct, we obtain a sequence of pairwise distinct elements of . Since is a generalized cell decomposition, this sequence contains a subsequence (which we again denote by ) that converges to a point . Since convergence to a point is preserved under homeomorphisms, we conclude that converges to the point . Thus we see that is a generalized cell decomposition of .
Remark 3.18.
When the refinement of one -complex by another is introduced in [Mar13], the fact that the refinement still satisfies property is not addressed. Since we are unable to prove directly that the refinement satisfies , we use an equivalent topological formulation of that we can work with: the third property in our definition of generalized cell decompositions.
We are left to define the -action on and show that it satisfies all the properties of a -complex. Let . On the subset , we set . The remainder of can be written as the union and we define on each of these sets individually. Let . On , we define
We need to show that is well-defined and a homeomorphism. Indeed, on the set , we have overlapping definitions of and we need to check the following equality.
This is equivalent to the equation
By assumption, and coincide on and the marking commutes with the action on . Therefore, this equality holds and we conclude that is well-defined.
Next, we show that is a homeomorphism on . By construction, is defined as a continuous map on various closed subsets of that cover . Thus, is continuous. Furthermore, we know that sends to itself and bijectively to . Since all are pairwise disjoint and acts bijectively on , we see that is bijective. Since is compact and Hausdorff, is a homeomorphism.
We are left to show that the map is a homomorphism. Recall that we are using the convention that actions by homeomorphisms satisfy the formula . Let . On , we simply have
For , we compute
We conclude that is a -action, turning into a marked -complex. One immediately sees that is a refinement of .
Remark 3.19.
Let for be three marked -complexes. For short, we simply denote them . Let denote the refinement of induced by . It is an easy exercise to show that the refinement of induced by is the same as the refinement of induced by . In other words, the refinement of -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 -case, short, and do not use the topology or geometry of at all, which is why we give minimal comment and refer to the proofs given by Markovic.
Proposition 3.20.
Let for be two -complexes on and let be the refinement of induced by . Then the following three statements hold:
-
(1)
If and are convergence -complexes, then is a convergence -complex.
-
(2)
If and are free -complexes, then is a free -complex.
-
(3)
For every , there exist , such that
We need one final proposition, which generalizes to due to the fact that radial extensions exist in higher dimensions as well.
Proposition 3.21.
Let be a free convergence -complex on such that is trivial for every . Let . Then there exists a free convergence -complex that is a refinement of the -complex .
That is, the action of the convergence group on can be extended to a free convergence action on .
Proof.
The -complex will be the refinement of induced by the following -complex:
For every , we define to be the radial extension of the homeomorphism . Since the radial extension defines a monomorphism , we obtain a -complex .
4. Uniformizing Sierpinski Spaces
In this section we prove a result concerning embeddings of the Sierpinski space into . We recall the definition of and some of its properties. Let be the unit interval, let denote the middle-thirds Cantor set, and let be the indicator function for .
Definition 4.1.
Let be the -cube with coordinates . For each and each satisfying , we define a space as follows. For each -element subset , consider the function , and let . By convention, we define . If then .
By construction, is the set of points with at least coordinates in . In particular, is the set of points for which every coordinate lies in , i.e. , while is . The topological dimension of is , and it satisfies a universal property: any compact -dimensional metric space which embeds into , embeds in . In this sense, is the universal receptor for -dimensional compacta in .
By embedding into , we can regard as a compact subset of . When , each complementary component of is an open disk with boundary homeomorphic to ). These spheres are called peripheral. With respect to the standard metric on , the collection of peripheral spheres form a null collection, defined as follows.
Definition 4.2.
A collection of of pairwise disjoint embedded -spheres in (with its standard metric) is called a null collection if
In the case of , if one considers the complement of all the peripheral spheres, then the closure of each component is either or . 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 , and that there are at most countably many such components. For this, we will need the following definition.
Definition 4.3.
An -triod is the topological space formed as the union of the -disk and an interval such that is one of the endpoints of and an interior point of .
Since an is -dimensional, and clearly embeds in , it also embeds in . In fact, there exists an embedding of into the “interior" of .
Lemma 4.4.
There exists an embedding of into whose image is disjoint from every peripheral sphere.
Proof.
Recall that is the set of points in where at least one coordinate is in . Let be a point of that is not a boundary point (i.e. not in the closure of one of the complementary intervals of ). Then , and is disjoint from every peripheral sphere. We can then find a closed -disk contained in . Now choose that are not boundary points such that in . Out of , adjoin a small arc for some . Then is also disjoint from every peripheral sphere, and meets only at . Thus is the desired -triod. ∎
Lemma 4.5.
Let be a dense null sequence of locally flat (n-1)-spheres embedded in . Then has countably many connected components, and the closure of each is homeomorphic to .
Proof.
We first prove the second claim. Let be the closure of a component of . Since the are dense and locally flat in , the closure is nowhere dense. Moreover, because each is locally flat, consists of a union of tame n-balls . The boundary of each is one of the , hence if . The fact that the form a null collection implies that the do too. By [Can73, Theorem 1], is homeomorphic to . (The case follows from [BR13, Theorem 1.3].)
Now we prove that there are only countably many components of . Let , be the closures of components of . Observe that the intersection is either empty or equal to for some . By Lemma 4.4, in each we can find an embedded -triod which does not meet any peripheral sphere of . Since and 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 (and thus ) contains at most countably many pairwise disjoint embedded -triods. Hence there can be at most countably many . This proves the first claim, and the lemma. ∎
Proposition 4.6.
Let be a dense null sequence of locally flat -spheres embedded in . Then there exists a homeomorphism such that is round for each .
Proof.
By Lemma 4.5, there are countably many components of , and the closure of each is homeomorphic to Enumerate the closures of these complementary components as . Clearly we have and as we observed in the proof of Lemma 4.5, for , the intersection is empty or some . Conversely, each is the intersection of exactly two of the .
Now let be a dense null collection of round -spheres in . The closures of are also homeomorphic to , and again there are countably many of them, which we denote by . We build as follows. First set and choose a homeomorphism . Now suppose we have defined . Choose some peripheral sphere of and let be the corresponding peripheral sphere. There is a unique not among meeting . Let be the corresponding element meeting . By [Can73, Lemma 1], there exists a homeomorphism extending . (The proof Cannon gives also works when case, once one knows that the annulus theorem holds. This was proven by Quinn [Qui82, Edw84].) We use this to define , by pasting to along , where they agree.
Since agrees with on , in the limit we get a well-defined map . Injectivity is clear from the definition, hence is an embedding as is compact and Hausdorff. In particular, the image of is closed. On the other hand, the domain and codomain have the same dimension so is also open. Since is connected, 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 is a torsion-free hyperbolic group whose boundary is homeomorphic to , . Moreover contains a finite collection of quasi-convex, codimension-1 subgroups satisfying:
-
(1)
separates pairs of points at infinity.
-
(2)
For each , is a locally flat in
By (1), we know that cubulates ; that is, acts geometrically on a finite-dimensional CAT(0) cube complex whose hyperplane stabilizers are conjugates of the .
Recall that acts on its boundary by homeomorphisms, inducing an action by automorphisms on the cohomology . By passing to a subgroup of index at most , we may therefore assume that the action of on is orientation-preserving. We denote the action of on the boundary by a representation .
Lemma 5.1.
After replacing by a family of finite index subgroups , we may assume that each additionally satisfies:
-
(3)
There exists a closed, orientable -manifold such that . Additionally, preserves each complementary component of and acts freely, properly discontinuously cocompactly on each.
Proof.
Let . By (2) and the generalized Schoenflies theorem, the pair is homeomorphic to the standard pair . In particular, consists of exactly two components, each homeomorphic to and each of whose closures is homeomorphic to the closed -ball .
Since the action of on preserves it also preserves the complement. Therefore, there exists a subgroup of index at most such that preserves each complementary component of . Because has finite index in , the limit set of is the same as that of , i.e. . Since the action of on is orientation-preserving, the action of on must also be orientation-preserving.
Let be one of the two components of . By a result of Swenson [Swe01], acts freely, properly discontinuously and cocompactly on . Since , this implies that is isomorphic to the fundamental group of , which is a closed, orientable, aspherical -manifold. Now the collection obtained by replacing each with its corresponding subgroup satisfies (1), (2) and (3). ∎
We now assume satisfies conditions (1)–(3). Applying Theorem 2.3, we may assume that also satisfies
-
(4)
Every is malnormal.
Let denote the limit set of , which can be identified with . The action of each preserves the limit set , which by assumption is homeomorphic to a locally flat -sphere. Now consider , the set of conjugates of . The limit set of is , which is disjoint from by malnormality. Each translate separates into two components, hence we may define the dual tree whose edges are translates and whose vertices are in 1-1 correspondence with connected components of
For any , we use the orientation on to choose a left side and a right side of . The sides and correspond to the two half-edges at the midpoint of each edge of . By , acts on without inversions, hence for any , the left side (resp. right side) of is taken to the left side (resp. right side) of .
Lemma 5.2.
For , there exists a collection of homeomorphisms satisfying
-
(i)
restricts to the identity on .
-
(ii)
For any , and we have
Proof.
By (3), acts on both sides of properly discontinuously, cocompactly. Define , and . Choosing basepoints in and , we get an identification between and . Since both of these are aspherical, we obtain a homotopy equivalence . Our choice of yields a commutative diagram
and in particular the lift to universal covers induces the identity map on the boundary. By Corollary 2.5, is homotopic to a homeomorphism (This is where we need that .) Since is homotopic to , the same holds for the lift . Set ; by construction for every and .
Now suppose and define . Since the action of takes the left side of to the left side of , is a homeomorphism and is the identity since the same holds for , which proves (i). Let be another group element satisfying , whence for some . If we define then for any
where the third equality follows from -equivariance of . Thus depends only the coset . 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 for each
Lemma 5.3.
Let denote the closed -disk with the Euclidean metric, with boundary . There exists a homeomorphism and a collection of pairwise disjoint, properly embedded -disks , satisfying the following properties:
-
(i)
.
-
(ii)
If then .
-
(iii)
Let
and let be the set of complementary components of . Then is closed and each pair , is homeomorphic to .
-
(iv)
Let and . Then for each , does not separate and .
Proof.
For each , there are only finitely many such that The union is dense in because it contains the orbit of the endpoints of one axis of an element of . Therefore forms a dense null collection. By Proposition 4.6, there exists a homeomorphism such that is a round -sphere, for each . By the Alexander trick, extends to a homeomorphism .
Regarding as hyperbolic -space in the Poincaré model, for each , bounds a round hemisphere , where are disjoint if . In the Euclidean metric, , proving (i) and (ii). The remainder of the proof is exactly the same as that of [Mar13, Lemma 3.2]. ∎
Thus the pair is a generalized cell decomposition. The next result shows that each yields a free convergence -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 -complex satisfying:
-
(5)
For any , any and any , does not separate the fixed points of in .
Proof of Lemma 5.4.
Let ( be the generalized cell decomposition from Lemma 5.3. For each , choose a homeomorphism which restricts to the identity on . Let be the action of on . We extend this to an action on via the formula:
() |
This is a homeomorphism of since it agrees with on and the interiors are pairwise disjoint. A simple calculation using Equation ( ‣ 5) verifies that defines a -action. The proof that is a free, convergence -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 be a torsion-free hyperbolic group such that , and let be quasi-convex, codimension-1 subgroups which satisfy (1) and (2). By Lemma 5.1 and Theorem 2.3, we may assume each satisfies (1)–(4). Now by Lemmas 5.3 and 5.4, for , there exists a homeomorphism and a free convergence -complex satisfying (5), such that is a union of with a disjoint collection of round hemispheres , .
For , we now define a sequence of free, convergence -complexes that is a refinement of . To start, set , and then inductively define as the refinement of by the pushforward , where . By construction, is a -complex; moreover, it is a free, convergence -complex by Proposition 3.20 (1) and (2).
Now consider . By Proposition 3.20 (3), for any there exist such that . Suppose that there exists , and let be the fixed points of Since , then by (5) we know that and are not separated by for any for each . But this holds for every , contradicting our assumption that separates pairs of points in . Therefore is trivial for every . 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 be a torsion-free, hyperbolic group such that . Let be a quasi-convex, codimension-1 subgroup such that its limit set is homeomorphic to . By the Jordan-Brouwer separation theorem, has exactly two connected components, which we denote by and .
Our goal in this section is to prove that if and are simply connected, then the embedding is equivalent to the standard embedding . 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 such that for every normal space , every closed subset and every continuous map there exists an open neighborhood of such that has a continuous extension .
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 be a compact space of dimension and a closed subspace. We say that is -LCC in , if for every and every neighborhood of , there exists a neighborhood of such that every continuous map extends continuously to .
Let denote either or . Observe that is an open subspace of the compact manifold and thus an ANR. In particular, has the homotopy type of a CW complex. We begin with the following result.
Lemma 6.4.
If is simply connected, then is contracible.
Proof.
As is locally contractible, Alexander duality implies that . Since , we know that the homology of appears as a summand in . The fact that for all implies that is acyclic. (For degree zero, note that is a connected component.) Since 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 , there is a family of finite-dimensional, simplicial complexes on which acts geometrically, known as Rips complexes.
Definition 6.5.
Let be a finite generating set for and consider the word metric for with respect to . For any we define a simplicial complex called the Rips complex as follows. The vertex set of is , and span a -simplex if for all .
The word metric on induces a metric on , and acts faithfully, cocompactly on with finite stabilizers. By Milnor–Schwarz, is quasi-isometric to . When is hyperbolic, then for any generating set and sufficiently large, is in fact contractible [GdlH90]. In particular, if is torsion-free then the action of on is free and thus for sufficienty large is a . In this case, Bestvina–Mess show that the proper homotopy type of is an invariant of [BM91].
Proposition 6.6 (Flatness Criterion).
If and are simply connected, then the embedding is locally flat, hence conjugate by a homeomorphism to a standard embedding .
Proof.
Let be either or . Since acts on and preserves its limit set , it also preserves the complement . If the action of does not preserve , we may find an index 2 subgroup (which we also denote by and which has the same limit set), whose action preserves . Since simply connected, it is contractible by Lemma 6.4. As is torsion-free, a result of Swenson [Swe01, Main Theorem (3)] implies that acts freely, properly discontinuously, and cocompactly on . In particular, the quotient is aspherical and thus a compact .
Since any is unique up to homotopy equivalence, is homotopy equivalent to any other , notably the quotient of the Rips complex by for . Taking universal coverings, this homotopy equivalence lifts to a proper, -equivariant homotopy equivalence between and .
Choosing a basepoint , the orbit map identifies the boundary at infinity of (regarded as a -hyperbolic space) with . Since is an ANR, we can apply [BLW10, Theorem 2.4] to obtain that is -LCC in . Recalling that was either or , we see that is -LCC in both and . By [DV09, Theorem 7.6.5], this implies that is locally flat in . The generalised Schoenflies theorem now implies that the embedding is conjugate by a homeomorphism to a standard embedding by Remark 2.10.∎
Suppose now that for some closed, orientable, aspherical manifold . We now apply the above criterion to give an alternative characterization of local flatness of in terms of the cover of corresponding to . In this setting, we refer to as 2-sided if the action of on preserves both components of .
Theorem 6.7.
Let be a closed, orientable, aspherical -manifold with hyperbolic, and . Suppose is a quasi-convex, codimension-1, 2-sided subgroup such that . If denotes the cover associated to , then is locally flat if and only if there exists a closed, orientable -manifold such that .
Proof.
Let denote the two complementary components of . Since is orientable and is 2-sided, the action of on preserves and separately, acting on each by orientation-preserving homeomorphisms. The action of on is free, properly discontinuous and cocompact by [Swe01]. In particular, is a covering map onto a compact manifold whose interior is and whose boundary is , where are closed and orientable. Since both and are collared in , we have that . The latter is aspherical and has fundamental group isomorphic to . Hence is a .
If is locally flat, then are both contractible. Since acts freely properly discontinuously on , this implies and are also ’s. To verify the inclusions are both homotopy equivalences, it is enough to check that they induce isomorphisms on . The inclusions are surjective because and are both connected. They are injective because and are simply connected. Thus is an -cobordism. Since is torsion-free hyperbolic, the Whitehead group vanishes by [BL12] so is an -cobordism. Since , we conclude that . Thus, and we can take .
Conversely, suppose for some closed, orientable manifold . Thus the inclusion is a homotopy equivalence. Let be the submanifold bounded by and , and let be a collar neighborhood of in . Since and is compact, there exists an injective map which is the identity on , and whose image lies in . We compose with the projection of onto to obtain a retraction . Therefore, the inclusion induces an injection of into . On the other hand, since is the group of deck transformations of the cover we have a short exact sequence:
Therefore, is an isomorphism and . The same argument shows as well, whence is locally flat by Proposition 6.6. ∎
6.2. Embedded submanifolds from codimension-1 subgroups
Let be a closed, orientable aspherical -manifold with cubulated hyperbolic fundamental group . Suppose there exists a quasi-convex subgroup such that for some closed orientable aspherical -manifold .
Definition 6.8.
A subgroup is square-root closed if implies .
Lemma 6.9.
Let be a 1-ended torsion-free hyperbolic group. If is malnormal and quasi-convex, then is square-root closed. Moreover, if is 1-ended and has codimension-1, then splits as an amalgamated product or HNN extension over .
Proof.
If then , and since has infinite order malnormality implies this is only possible if . This proves the first statement. The second statement follows from a result of Kropholler [NR93, p.146, Theorem 4.9]. ∎
Our interest in splitting over a square-root closed subgroup stems from a theorem of Cappell [Cap76, Theorem 1] giving homotopical conditions for a manifold with to contain an embedded, 2-sided, codimension-1 submanifold with .
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 -coefficients. Clearly any space homotopy equivalent to a closed orientable manifold is an orientable Poincaré complex. Consider now the following setup. Let be a connected, orientable -dimensional Poincaré complex and an embedding of a connected, orientable -dimensional sub-Poincaré complex with . Suppose further that is injective.
Definition 6.10.
Let be as above and let be a manifold. A homotopy equivalence is splittable along if it is homotopic to a map which is transverse regular along (whence is embedded, codimension-1 submanifold of ) and if restricted to both and are homotopy equivalences. If has two components we require that the restriction of to each component of be a homotopy equivalence.
Given a pair of Poincaré complexes, a manifold , and a homotopy equivalence , the main result of Cappell [Cap76, Theorem 1] gives homotopical conditions for to be splittable. Crucially, one must assume that is square-root closed in , and that splits over either as an amalgamated product or HNN-extension. The remaining assumptions are -theoretic in nature, and are satisfied vacuously when both and are torsion-free hyperbolic. In the setting where and are both aspherical manifolds and satisfies the Borel conjecture, the conclusion of Cappell’s result is that contains an embedded, 2-sided submanifold homeomorphic to . We now apply Cappell’s result to our situation.
Theorem 6.11.
Let be a closed, orientable aspherical -manifold with cubulated hyperbolic and . Let be a quasi-convex, codimension-1 subgroup such that for some closed aspherical -manifold . Then there exist finite covers , and an embedding .
Proof.
By Theorem 2.3, there exist finite index subgroups and such that is a malnormal subgroup. Let and be the corresponding finite covers of and , respectively. By Lemma 6.9, is square-root closed and splits over as either an amalgamated product or HNN extension. In other words, acts on a tree with edge stabilizers conjugate to and a single edge orbit. Since 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 from classifying spaces for the vertex stabilizers and . Explicitly, if , then let be a and be a . Now we obtain as the identification space from , and by gluing to by a map inducing the inclusion , for The construction for an HNN extension is similar. is a Poincaré duality complex since is it homotopy equivalent to , and it contains 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 is splittable. In particular, we find an embedded, 2-sided, codimension-1 manifold such that injects as . Finally, since , we have that is homeomorphic to by Theorem 2.4. ∎
Appendix A Proofs of Lemma 3.16 and Lemma 3.17
Lemma A.1 (Lemma 3.16).
Let be a refinement of and suppose , where and . Then .
Proof.
Suppose , that is there exists some such that and . Since and coincide on , we conclude that . If , then this implies that does not preserve and .
Now suppose and suppose by contradiction that . Since and , there exists some such that
Choose a point and a path from to that meets only in its endpoint . (Such a path may be found by exploiting the homeomorphism .) The path passes through a sequence of elements . In between these open sets meets in a sequence of path-segments , where is the segment of between its time in and its time in . We obtain a decomposition of the path into segments that are contained in in that order. If this sequence is finite, then it ends in for some . However, this sequence may go on forever. We denote the start and end point of by and respectively.
Let and . We will prove by induction that for all (except for the last which, if it exists, will be contained in ). The induction starts with . Since we assume, by contradiction, that , we have that which is a subset of by assumption.
Now suppose and suppose is not the last segment in the sequence . Since meets only in its endpoint and is not the last segment, and thus . By induction assumption, and thus . Since , , and sends homeomorphically to itself, we conclude that .
Since meets only in its endpoint and is not the last segment, we have that . Since sends homeomorphically to itself, we conclude that . Combined with the fact that the starting point of lies in , we conclude that .
We now use this to show that . Since , we have that . At the same time, and thus . Since every element in is contained in an element of and , we conclude that .
This induction shows that and are contained in for all (except for , if it is the last segment of the sequence). Depending on whether the sequence 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 . By induction, we know that . By the same argument as before, we see that either lies in , or is the endpoint of . If it is the endpoint of , then and thus . Now suppose is not the endpoint of . In that case, is a segment that starts in , that only meets in its endpoint, and whose starting point is sent to a point in
We conclude that is contained in except for its endpoint which lies in . Since the endpoint of is the endpoint of , which is , this implies that . However, we chose so that , a contradiction.
Now suppose the sequence is infinite. In that case, are contained in for every . Since is continuous, the endpoints of the segments converge to the point . Since is continuous, we conclude that
However, for all and thus their limit has to lie in . This is a contradiction. Since we obtain a contradiction in both cases, we conclude that cannot lie in , which completes the proof. ∎
Lemma A.2 (Lemma 3.17).
Let be a -complex which is a refinement of a convergence -complex . If for every , the group is a convergence group on , then is a convergence -complex as well.
Proof.
Let be a sequence of distinct elements. Since is a convergence -complex, there exist and a subsequence of (which we denote by again) such that uniformly on compact subsets of . This implies that, for the inverse sequence, we have uniformly on compact subsets of .
Step 1: Using the convergence action on to get parts of the convergence behavior on .
Since the actions of and coincide on , we conclude that and on compact subsets of and respectively. We thus obtain the following two properties: For every compact set and every , there exists such that
(1) |
Similarly, for every compact set and every , there exists such that
(2) |
We need to show that for every compact set and every , there exists such that
Step 2: Exhausting by well-behaved compact sets.
For , we define
where denotes the closed ball in of radius with respect to the euclidean metric centered at and its interior.
The collection is an exhaustion of by compact sets. We may thus assume without loss of generality that for some . Note that, for sufficiently small, has the following property: If for some , then either or and . (This property is crucial to the proof and we highlight that it holds for for the same reason as it does for .) If there exists such that , then and is a convergence -complex by the assumption in the Lemma. We thus assume from now on that . We define compact sets that form an exhaustion of in the analogous way.
Let . Since satisfies (see Remark 3.3), there are only finitely many sets whose diameter is .
Step 3: There exist and sets , where such that the following holds:
If satisfies
-
•
,
-
•
there exists such that for some ,
then for some .
By construction, does not intersect the open ball in of radius centered at . Put . We obtain the sets as follows: Since satisfies , there exist only finitely many sets with diameter . We set the number of these sets to be .
Choose such that intersects each (thus , as the depend only on ). Furthermore, choose large enough such that, whenever , then . (This depends only on the generalized cell decomposition and the fact that we already dealt with the case .) Set . By equation (2), we know that for every , we have
We set , which depends on and , as and are determined by and respectively. We observe that, since intersects each by assumption and was chosen sufficiently large, it intersects . Thus we find . Since , equation (2) implies for every that
(3) |
We claim that these choices of , , have the properties we require. Let such that for some and some . Suppose for all . Then
Combined with equation (3), this implies that
Therefore, cannot intersect which proves Step 3.
Step 4: Produce a subsequence of that shrinks the diameter of for all but finitely many that intersect .
Choose some such that . We partition the sequence into new sequences for as follows: The sequence consists of all elements of such that
For , consists of all elements of that satisfy .
It is important to note that we index the elements such that for every . In turn, this means that is defined only for certain . It is important that we index our elements in this way to formulate the following statement.
We claim that, if is not one of the sets from Step 3, then is not defined for any and any . Indeed, if is not any of the sets and , then cannot satisfy both properties required in Step 3. But by assumption on and thus for all .
Step 5: Estimate for .
Let such that . Since intersects and was chosen sufficiently large, there exists . Applying inequality (1), we obtain
Restricting to the elements , we also have
Combining these two inequalities, we obtain
(4) |
This is the equality we desire for all (and all greater than some ). We are left to show that this inequality also holds for the elements when is sufficiently large.
If for all , then is not defined for any by Step 4 and we are done.
Let and suppose . Let be the smallest such that is defined. We define and compute
(Recall our convention that .) We conclude that and thus . Since uniformly on compact sets in , we have that uniformly on compact sets in , where .
Suppose, did not converge to uniformly on . Then there exists an infinite subsequence and a sequence of points such that
We first observe that cannot converge to . Indeed, if , then while for every . Since is closed by construction, this implies that . But so this is a contradiction and we conclude that cannot converge to .
By assumption, is a convergence group on . Therefore, we may pass to a subsequence of (which we denote by again) and find points such that uniformly on compact subsets of .
On the other hand, we know that uniformly on compact sets in , in particular on compact sets in . Since contains infinitely many points, uniform convergence on compact sets in yields that and . Therefore, uniformly on compact subsets of . Since and does not converge to , we conclude that there exists such that
This is a contradiction and we conclude that there exists such that
Step 6: Concluding the Lemma.
Set
By the estimates obtained in Step 5, we have for every such that that
Combined with inequality (1) this implies that is a convergence -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 -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 -dimensional Sierpiński curve in . 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 -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 ) 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 and . J. Differential Geometry, 17(3):503–521, 1982.
- [Sco83] Peter Scott. There are no fake Seifert fibre spaces with infinite . Ann. of Math. (2), 117(1):35–70, 1983.
- [Sta61] John Stallings. On fibering certain -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.