An extended definition of Anosov representation for relatively hyperbolic groups
Abstract.
We define a new family of discrete representations of relatively hyperbolic groups which unifies many existing definitions and examples of geometrically finite behavior in higher rank. The definition includes the relative Anosov representations defined by Kapovich-Leeb and Zhu, and Zhu-Zimmer, as well as holonomy representations of various different types of “geometrically finite” convex projective manifolds. We prove that these representations are all stable under deformations whose restriction to the peripheral subgroups satisfies a dynamical condition, in particular allowing for deformations which do not preserve the conjugacy class of the peripheral subgroups.
1. Introduction
1.1. Format of this paper
This paper constitutes the first part of a preprint originally posted under the same title in April 2022, and covers the definition of an EGF representation and the proof of the main stability theorem. The only changes to the contents are minor corrections and some reworked exposition. For the sake of shortening the paper, the second part of the earlier preprint (covering examples of EGF representations) will be posted as a separate article [Wei23b]. The contents of the sequel are also largely unchanged from their form in the original preprint.
1.2. Overview
Anosov representations are a family of discrete representations of hyperbolic groups into reductive Lie groups, and give a natural higher-rank generalization of convex cocompact representations in rank-one. They were originally defined for surface groups by Labourie [Lab06], and the definition was extended to general word-hyperbolic groups by Guichard-Wienhard [GW12].
Many equivalent characterizations of Anosov representations have since been uncovered [KLP14, KLP17, GGKW17, BPS19, Tso20, KP22], and much of the theory of convex cocompact groups in rank-one has been extended to this higher-rank context. For instance, Anosov representations satisfy a stability property: if is a hyperbolic group, then the Anosov representations from into a semisimple Lie group form an open subset of the representation variety .
The purpose of this paper is to introduce a higher-rank notion of geometrical finiteness, which involves generalizing the notion of an Anosov representation to allow for a relatively hyperbolic domain group. Our definition—that of an extended geometrically finite (EGF) representation—encompasses all previous definitions of relative Anosov representation, and additionally covers various forms of higher-rank “geometrical finite” behavior which are not described by other definitions.
Our starting point is a definition of Anosov representation in terms of topological dynamics: if is a hyperbolic group, is a semisimple Lie group, and is a symmetric parabolic subgroup, a representation is -Anosov if there is a -equivariant embedding of the Gromov boundary of satisfying certain dynamical properties.
Existing notions of relative Anosov representations (see e.g. [KL18], [Zhu21], [ZZ22]) replace with an embedding of the Bowditch boundary of a relatively hyperbolic group. This condition forces the limit set in of a peripheral subgroup to be a singleton, so that the peripheral subgroup itself must be both weakly unipotent and regular (see [KL18, Section 5]). These requirements are natural in rank one, but less so in higher rank, and there are a number of interesting examples where they are not satisfied (see Section 1.6).
The main idea behind the definition of an extended geometrically finite representation is to reverse the direction of the boundary map: we characterize geometrical finiteness via the existence of an equivariant map from a closed subset of a flag manifold to the Bowditch boundary of a relatively hyperbolic group, rather than the other way around. Our “backwards” boundary map does not need to be a homeomorphism, so the approach allows for more flexibility in the higher rank setting; in particular, there is no restriction on the limit set of a peripheral subgroup in , or even a requirement that this “limit set” is well-defined. Moreover, even in rank one, this new perspective seems useful for studying deformations of discrete groups which do not preserve the homeomorphism type of the limit set (see Section 1.8.3).
With the additional flexibility afforded by the definition, we can prove a general stability result for EGF representations (see Section 1.4 below). It is not true that an arbitrary (sufficiently small) deformation of an EGF representation is still EGF; indeed, it is possible to find small deformations of geometrically finite representations in rank one which are not even discrete. However, we prove that any EGF representation is relatively stable: any small deformation of in which satisfies a condition on the peripheral subgroups is also EGF. The peripheral condition—which we call peripheral stability—is general enough to hold even in the absence of a topological conjugacy between the limit sets of the original peripheral subgroups and their deformations. In specific cases, it can even be used to deduce absolute stability results for non-hyperbolic groups (see the last section of [Wei23b]).
1.3. The definition
If is a relatively hyperbolic group, relative to a collection of peripheral subgroups, then acts as a convergence group on the Bowditch boundary . We recall the definition here.
Definition 1.1.
Let act on a topological space . The group is said to act as a convergence group if for every infinite sequence of distinct elements , there exist points and a subsequence such that converges uniformly on compacts in to the constant map .
When is a sequence of distinct elements in a relatively hyperbolic group , then converges to uniformly on compacts in if and only if converges to in the compactification and the inverse sequence converges to .
Recall that if a group acts by homeomorphisms on a Hausdorff space , the pair is called a topological dynamical system. We say that an extension of is a topological dynamical system together with a -equivariant surjective map .
In this paper, when is a relatively hyperbolic pair (i.e. is hyperbolic relative to a collection of peripheral subgroups), we will consider embedded extensions of the topological dynamical system . We want these embedded extensions to respect the convergence group action of in some sense, so we introduce the following definition: {restatable}definitiondynamicsPreserving Let be a relatively hyperbolic pair, with acting on a connected compact metrizable space by homeomorphisms. Let be a closed -invariant set.
We say that a continuous equivariant surjective map extends the convergence action of if for each , there exists an open set containing , satisfying the following:
If is a sequence in with for , then for any compact set and any open set containing , for sufficiently large , lies in .
Now let denote a semisimple Lie group with no compact factor. The central definition of the paper is the following: {restatable}definitionEGF Let be a relatively hyperbolic group, let be a representation, and let be a symmetric parabolic subgroup. We say that is extended geometrically finite (EGF) with respect to if there exists a closed -invariant set and a continuous -equivariant surjective antipodal map extending the convergence action of .
The map is called a boundary extension of the representation , and the closed invariant set is called the boundary set.
We refer to Section 4 for the definition of “antipodal map” in this context.
1.4. Stability
Like (relative) Anosov representations, extended geometrically finite representations are always discrete with finite kernel (see 4.1). The central result of this paper says that EGF representations have a relative stability property: if is an EGF representation, then certain small relative deformations of must also be EGF.
To state the theorem, we define a notion of a peripherally stable subspace of . The precise definition is given in Section 9, but roughly speaking, a subspace is peripherally stable if the large-scale dynamical behavior of the peripheral subgroups of is in some sense preserved by small deformations inside of . We emphasize again that the action of a deformed peripheral subgroup does not need to be even topologically conjugate to the action of the original peripheral subgroup.
We prove the following: {restatable}theoremcuspStableStability Let be EGF with respect to , let be a boundary extension, and let be peripherally stable at . For any compact subset of and any open set containing , there is an open subset containing such that each is EGF with respect to , and has an EGF boundary extension satisfying .
Remark 1.2.
When is a -Anosov representation of a hyperbolic group , then the associated boundary embedding also varies continuously with in the compact-open topology on maps . Since EGF representations come with boundary extensions (rather than embeddings), Section 1.4 only gives us a semicontinuity result.
We expect that it is possible to extend the methods of this paper to prove stronger continuity results when the original EGF representation satisfies additional assumptions; see the discussion following 1.9.
While the peripheral stability condition in Section 1.4 is mildly technical, we can also apply it to yield more concrete results.
Definition 1.3.
Let be a relatively hyperbolic pair, and let be a representation. The space of cusp-preserving representations is the set of representations such that for each peripheral subgroup , we have for some (which may depend on ).
Corollary 1.4.
Let be an EGF representation. Then there is a neighborhood of in consisting of EGF representations.
1.4 gives a very restrictive example of a peripherally stable subspace of . But, in general the peripheral stability condition is flexible enough to allow peripheral subgroups to deform in nontrivial ways.
In particular, it is possible to find peripherally stable deformations of an EGF representation which change the Jordan block decomposition of elements in the peripheral subgroups. For instance, one can deform an EGF representation in with unipotent peripheral subgroups into an EGF representation with diagonalizable peripheral subgroups—see Example 9.3.
1.5. Techniques used in the proof
Our proof of Section 1.4 is loosely inspired by Sullivan’s proof of stability for convex cocompact groups in rank-one [Sul85], but with substantial modification and several new ideas needed. Sullivan’s original approach was to show that a discrete group is convex cocompact if and only if the action of on its limit set in satisfies an expansion property. Then, he used this expansion property to give a symbolic coding for infinite quasigeodesic rays in . This coding gives a way to see that the correspondence between geodesic rays in and points in is stable under small perturbations of the representation.
Other authors have successfully used expansion dynamics and symbolic codings to prove stability results in the higher-rank setting (see [KKL19], [BPS19]). In our context, however, we encounter several new challenges. The first problem is that EGF representations do not actually have “expansive” or even “relatively expansive” dynamics on their limit sets in in any metric sense. This means we need a way to understand “relative expansion” (or its inverse, “relative contraction”) purely topologically. Second, we need to come up with a procedure for constructing a relative coding for points in the Bowditch boundary of a relatively hyperbolic group , in a way which is compatible with the “topological” expansion of the -action on both and . In addition, we need to ensure that this compatibility is preserved under a wide variety of possible perturbations of the action of a peripheral subgroup.
The solution for all of these problems is to work almost entirely within the framework of extended convergence group actions. Using this setup, we provide a general construction for a relative quasi-geodesic automaton: a finite directed graph with edges labeled by subsets of , so that paths in the graph are in rough correspondence with quasi-geodesics in the coned-off Cayley graph for the pair . Each vertex of the graph is assigned an open subset of some compact metrizable space on which acts as an extended convergence group. Each edge in the graph corresponds to a set of inclusions of the form
(1) |
where is an element in the subset of labeling the edge , and is a small neighborhood of with respect to some fixed metric on . This strong nesting of subsets turns out to be a reasonable stand-in for a “contraction” property of the group action. Each inclusion of the form (1) is an open condition in the topology on actions of on , so when the set of elements labeling an edge is finite, the corresponding set of inclusions are stable under small perturbations of the -action. If the label of is instead infinite, then the corresponding (infinite) set of inclusions may not be stable under arbitrary deformations—this is where the peripheral stability assumption is needed.
Remark 1.5.
Even in the non-relative case, our procedure for constructing an automaton and our topological viewpoint on “contraction” seem to be useful—see [MMW22] for an application of the idea to abstract word-hyperbolic groups.
In its simplest form, our approach can be thought of as a “generalized ping-pong” argument: if are arbitrary finitely generated groups, then the group is hyperbolic relative to the collection of conjugates of and . In this case, the inclusions in (1) are precisely the inclusions of sets required to set up a ping-pong argument proving that a representation of is discrete and faithful.
1.5.1. Contraction in flag manifolds
To complete our proof of Section 1.4, we also need to work in the specific setting of an extended convergence action on a flag manifold , rather than an arbitrary metrizable space. We prove a useful general result (7.11) about a metric defined by Zimmer [Zim18] on certain open subsets of , which allows us to precisely reinterpret the rough topological “contraction” given by (1) as an actual metric contraction. The technique is similar to the one applied in the context of multicone systems and dominated splittings in [BPS19]. One nice consequence is that the boundary extension of an EGF representation can always be chosen to be injective on preimages of conical limit points in the Bowditch boundary (see 4.7).
1.6. Examples
The related paper [Wei23b] (originally the second part of the current article) is focused on describing examples of EGF representations. Here, we briefly explain the nature of these examples, as well as additional examples described by other authors. See the references for further detail.
1.6.1. Convex projective structures
A host of examples of Anosov representations arise from the theory of convex projective structures; see e.g. [Ben04], [Ben06b], [Kap07], [DGK18], [DGK+21]. In fact, work of Danciger-Guéritaud-Kassel [DGK17] and Zimmer [Zim21] implies that Anosov representations can be essentially characterized as holonomy representations of convex cocompact projective orbifolds with hyperbolic fundamental group. However, convex projective structures also yield a number of interesting examples of discrete non-Anosov subgroups of . In many cases, the groups in question are relatively hyperbolic, and appear to have “geometrically finite” properties.
The theory of EGF representations is well-suited to these examples. For instance, in [Wei23b], we apply some of our previous work [Wei23a] together with work of Islam-Zimmer [IZ22] to see that whenever a subgroup is relatively hyperbolic and projectively convex cocompact in the sense of [DGK17], then the inclusion is EGF with respect to the parabolic subgroup stabilizing a flag of type in . If is not hyperbolic, then these examples are not covered by other definitions of relative Anosov representations (see [Wei23a, Remark 1.14]); such non-hyperbolic examples have been constructed in e.g. [Ben06a], [BDL15], [CLM20], [CLM22], [DGK+21], [BV23].
In [CM14], Crampon-Marquis introduced several definitions of “geometrical finiteness” for strictly convex projective manifolds. Zhu [Zhu21] proved that the manifolds satisfying one of their definitions111Crampon-Marquis originally claimed that all of their definitions of “geometrically finite” were equivalent; this appears to have been an error. have relative Anosov holonomy (see Section 1.7), which means they also have EGF holonomy by Section 1.7 below. Examples can be found by deforming geometrically finite groups in into while keeping the conjugacy classes of cusp groups fixed (see [Bal14], [BM20], [CLT18]), or via Coxeter reflection groups [CLM22].
There are, however, more general notions of “geometrically finite” convex projective structures. In [CLT18], Cooper-Long-Tillmann considered the situation of a convex projective manifold (with strictly convex boundary) which is a union of a compact piece and finitely many ends homeomorphic to , where is a compact manifold with virtually nilpotent fundamental group. The ends of such a manifold are called “generalized cusps,” and the possible “types” of generalized cusps were later classified by Ballas-Cooper-Leitner [BCL20]. Examples of projective manifolds with generalized cusps have been produced by Ballas [Bal21], Ballas-Marquis [BM20], and Bobb [Bob19]. In general the holonomy representations of these manifolds are not relative Anosov, but in [Wei23b] we prove that they do provide additional examples of EGF representations. The proof is an application of the EGF stability theorem: it turns out that peripheral stability is actually flexible enough to allow for deformation between the different Ballas-Cooper-Leitner generalized cusp types.
Remark 1.6.
After a version of this paper originally appeared as a preprint, Blayac-Viaggi [BV23] also produced still more general examples of convex projective -manifolds which decompose into a compact piece and several projective “cusps.” In these examples (which can arise as limits of convex cocompact representations), each cusp is finitely covered by a product , where is a closed hyperbolic manifold of dimension . Consequently, these manifolds do not have “generalized cusps” in the sense of Cooper-Long-Tillmann, and their fundamental groups cannot even admit relative Anosov representations. Nevertheless, Blayac-Viaggi showed that the holonomy representations of their examples are always EGF.
1.6.2. Other examples
In [Wei23b], we construct additional examples of EGF representations by considering compositions of projectively convex cocompact representations with the symmetric representation . We show that, assuming the peripheral subgroups in are all virtually abelian, then the composition is still EGF; this holds even though the compositions are not believed to be convex cocompact.
We are also able to prove that the entire space is peripherally stable about . Via Section 1.4, this gives a new source of examples of stable discrete subgroups of higher-rank Lie groups.
1.7. Comparison with relative Anosov representations
Previously, Kapovich-Leeb [KL18] and Zhu [Zhu21] independently introduced several notions of a relative Anosov representation. Later work of Zhu-Zimmer [ZZ22] showed that Zhu’s definition (that of a relatively dominated representation) is equivalent to one of the Kapovich-Leeb definitions (specifically, the definition of a relatively asymptotically embedded representation). In the special case where the domain group is isomorphic to a Fuchsian group, these definitions also agree with a notion of relative Anosov representation for Fuchsian groups introduced by Canary-Zhang-Zimmer [CZZ21].
Extended geometrically finite representations give a strict generalization of all of these definitions. We can precisely characterize when an EGF representation satisfies the stronger definition as well:
theoremrelAsympBdryDynamics Let be a relatively hyperbolic pair, and let be a symmetric parabolic subgroup. A representation is relatively -Anosov (in the sense given above) if and only if is EGF with respect to , and has an injective boundary extension .
Remark 1.7.
By 4.7, any EGF representation has a boundary extension which is injective on preimages of conical limit points. So, in the case where the peripheral structure is trivial (meaning that is a hyperbolic group and is identified with the Gromov boundary of ), Section 1.7 implies that EGF representations are precisely the same as Anosov representations.
This actually gives a new characterization of Anosov representations, since a priori the EGF boundary extension surjecting onto the Gromov boundary of a hyperbolic group does not need to be a homeomorphism; the theorem tells us that if such a boundary extension exists, then it is possible to replace with an injective boundary extension, whose inverse is the Anosov boundary map.
1.7.1. Stability for relative Anosov representations
In [KL18], Kapovich-Leeb suggested that a relative stability result should hold for relative Anosov representations, but not did not give a precise statement. By applying Section 1.4, 4.7, and Section 1.7, we obtain the following stability theorem:
Theorem 1.8.
Let be a relatively hyperbolic pair, let be a relative -Anosov representation, and let be a peripherally stable subspace, such that for each and each , the restriction is -divergent with -limit set a singleton. Then an open neighborhood of in consists of relative -Anosov representations of .
If we restrict the allowable peripheral deformations to conjugacies, this result reduces to:
Corollary 1.9.
Let be a relatively hyperbolic pair, and let be a relative -Anosov representation. There is an open neighborhood of in consisting of relative -Anosov representations.
In the special case where is isomorphic to a Fuchsian group, 1.9 follows from work of Canary-Zhang-Zimmer [CZZ21]. Further, Zhu-Zimmer [ZZ22] have given an independent proof of 1.9, and additionally showed that in this case the associated relative boundary maps vary continuously (in fact, analytically).
The methods used in this paper are considerably different from those employed in the Canary-Zhang-Zimmer and Zhu-Zimmer stability results. Their results do not cover the additional deformations allowed by Section 1.4 or Theorem 1.8, and as Zhu-Zimmer observe, it seems unlikely that their techniques are applicable for the general study of EGF representations. On the other hand, while we expect that the methods in this paper could be used to generalize the Zhu-Zimmer result regarding continuously varying boundary embeddings for relative Anosov representations, it does not seem easy to use our techniques to yield more precise quantitative results.
Remark 1.10.
More recently, Wang [Wan23b] showed that the situation of an EGF representation with -divergent image can be interpreted in terms of restricted Anosov representations, i.e. representations which are Anosov “along a subflow” of a certain flow space associated to the representation (see [Wan23a]). Using these ideas, Wang proves a version of 1.4 for this special class of EGF representations.
1.8. Further applications, and potential future applications
1.8.1. Anosov relativization
When is a relatively hyperbolic group, the Bowditch boundary of (and thus, the definition of an EGF representation) depends on the choice of peripheral structure for . In general, there might be more than one possible choice: for instance, if a group is hyperbolic relative to a collection of hyperbolic subgroups, then is itself hyperbolic, relative to an empty collection of peripheral subgroups (see [DS05, Corollary 1.14]).
In this paper, we prove the following Anosov relativization theorem:
Theorem 1.11.
Let be a relatively hyperbolic pair, and suppose that each is hyperbolic. If is an EGF representation with respect to for the peripheral structure , and restricts to a -Anosov representation on each , then is a -Anosov representation of .
A potential application of Theorem 1.11 is the construction of new examples of Anosov representations: one could start with an EGF representation which is not Anosov, and then attempt to find a peripherally stable deformation of which restricts to an Anosov representation on peripheral subgroups. Section 1.4 and Theorem 1.11 would then imply that the original representation can be realized as a non-Anosov limit of Anosov representations in the peripherally stable deformation space.
1.8.2. Limits of Anosov representations
In [LLS21], Lee-Lee-Stecker considered the deformation space of Anosov representations , where is a triangle reflection group, and showed that certain components of this space have representations in their boundary which are not Anosov. Interestingly, these limiting representations still have equivariant injective boundary maps from into the space of full flags in , but they fail to be Anosov because the boundary maps fail to be transverse.
The limiting representations constructed by Lee-Lee-Stecker cannot be relatively Anosov, but they do appear to be EGF. Together with the Anosov relativization theorem mentioned above, this provides evidence that EGF representations could serve as a useful tool in the study of boundaries of spaces of Anosov representations. In addition, it gives a potential source of examples of EGF representations which do not directly derive from convex projective structures.
1.8.3. Deformations in rank one
Even in rank one, the deformation theory of geometrically finite representations is not completely understood. In [Bow98], Bowditch described circumstances which guarantee that a small deformation of a geometrically finite group is still geometrically finite, but his criteria do not have an obvious analog in other rank one Lie groups. Moreover, the conditions Bowditch gives are too strict to allow for deformations which change the homeomorphism type of the limit set . Such deformations exist and are often peripherally stable, meaning that the EGF framework could be used to understand them further. It even seems possible that a version of the theory could be applied in circumstances where the isomorphism type of is allowed to change.
1.9. Outline of the paper
We begin by providing some background in Sections 2 and 3, and then give the full formal definition of EGF representations in Section 4. In that section we also prove Section 1.7 (giving the connection between EGF representations and relative Anosov representations) and Theorem 1.11 (the Anosov relativization theorem). Some of these proofs assume the results of later sections, but they are not relied upon anywhere else in the paper.
The rest of the paper is devoted to the proof of our main stablity theorem for EGF representations (Section 1.4). In Section 5 and Section 6, we develop the main technical tool needed for the proof, which involves using the notion of an extended convergence group action to construct the relative quasigeodesic automaton alluded to previously. Then, in Section 7, we prove a key result (7.11) regarding a metric on certain open subsets of flag manifolds , which we use to relate the results of the previous sections to relatively hyperbolic group actions on . Then, we use all of these tools to develop an alternative characterization of EGF representations in Section 8, and finally prove our main theorem in Section 9.
1.10. Acknowledgements
The author thanks his PhD advisor, Jeff Danciger, for encouragement and many helpful conversations—without which this paper could not have been written. The author also thanks Katie Mann and Jason Manning for assistance simplifying some of the arguments in Sections 5 and 6. Further thanks are owed to Daniel Allcock, Dick Canary, Fanny Kassel, Max Riestenberg, Feng Zhu, and Andy Zimmer for providing feedback on various versions of this project.
This work was supported in part by NSF grants DMS-1937215 and DMS-2202770.
2. Relative hyperbolicity
In this section we discuss some of the basic theory of relatively hyperbolic groups, mostly to establish the notation and conventions we will use throughout the paper. We refer to [BH99], [Bow12], [DS05] for background on hyperbolic groups and relatively hyperbolic groups. See also section 3 of [KL18] for an overview (which we follow in part here).
Notation 2.1.
Throughout this paper, if is a metric space, is a subset of , and , we let denote the open -neighborhood in about . For a point , we let denote the open -ball about .
When the metric space is implied from context, we will often just write or .
2.1. Geometrically finite actions
Recall that a finitely generated group is hyperbolic (or word-hyperbolic or -hyperbolic or Gromov-hyperbolic) if and only if it acts properly discontinuously and cocompactly on a -hyperbolic proper geodesic metric space .
A relatively hyperbolic group is also a group with an action by isometries on a -hyperbolic proper geodesic metric space , but instead of asking for the action to cocompact, we ask for the action to be in some sense “geometrically finite.”
To be precise, this means that has a -invariant decomposition into a thick part and a countable collection of horoballs. For a horoball , we let denote the center of in , and we let denote the stabilizer of any .
Definition 2.2.
Let be a finitely generated group acting on a hyperbolic metric space , and let be a countable collection of horoballs in , invariant under the action of on . If:
-
(1)
The action of on the closure of is cocompact, and
-
(2)
for each , the stabilizer of in is finitely generated and infinite,
then we say that is a relatively hyperbolic group, relative to the collection .
Definition 2.3.
Let be a relatively hyperbolic group, relative to a collection of subgroups .
-
•
The centers of the horoballs in are called parabolic points for the -action on . The set of parabolic points in is denoted .
-
•
The parabolic point stablizers are called peripheral subgroups. We often write for .
A group might be hyperbolic relative to different collections , of peripheral subgroups. The collection of peripheral subgroups is sometimes called a peripheral structure for .
Definition 2.4.
Let be a finitely generated group, and let be a collection of subgroups. We say that is a relatively hyperbolic pair if is hyperbolic relative to .
2.2. The Bowditch boundary
Definition 2.5.
Let be a relatively hyperbolic pair, so that is the set of stabilizers of parabolic points for an action of on a metric space as in Definition 2.2. We say that is a Gromov model for the pair .
In general there is not a unique choice of Gromov model for a given relatively hyperbolic pair , even up to quasi-isometry. There are various “canonical” constructions for a preferred quasi-isometry class of Gromov model, with certain desirable metric properties (see e.g. [Bow12], [GM08]).
Given any two Gromov models , for , there is always a -equivariant homeomorphism [Bow12]. The -space is the Bowditch boundary of . We will denote it by , or sometimes just when the collection of peripheral subgroups is understood from context. Since a Gromov model is a proper hyperbolic metric space, is always compact and metrizable.
Definition 2.6.
We say a relatively hyperbolic pair is elementary if is finite or virtually cyclic, or if .
Whenever is nonelementary, its Bowditch boundary contains at least three points. The convergence properties of the action of on (see below) imply that in this case, is perfect (i.e. contains no isolated points).
2.2.1. Cocompactness on pairs
Let be a Gromov model for a relatively hyperbolic pair . Since is hyperbolic, proper, and geodesic, for any compact subset , the space of bi-infinite geodesics passing through is compact.
Given any distinct pair of points , there is a bi-infinite geodesic in joining to . Since a horoball in a hyperbolic metric space has just one point in its ideal boundary, this geodesic must pass through the thick part of , so up to the action of it passes through a fixed compact subset .
This implies:
Proposition 2.7.
The action of on the space of distinct pairs in is cocompact.
2.3. Convergence group actions
If a group acts on a proper geodesic hyperbolic metric space , we can characterize the geometrical finiteness of the action entirely in terms of the topological dynamics of the action on . In particular, we can understand geometrical finiteness by studying properties of convergence group actions. See [Tuk94], [Tuk98], [Bow99] for further detail on such actions, and justifications for the results stated in this section.
Definition 2.8.
Let act as a convergence group (see Definition 1.1) on a topological space .
-
(1)
A point is a conical limit point if there exists a sequence and distinct points such that and for any .
-
(2)
An infinite subgroup is a parabolic subgroup if it fixes a point , and every infinite-order element of fixes exactly one point in .
-
(3)
A point is a parabolic point if it is the fixed point of a parabolic subgroup.
-
(4)
A parabolic point is bounded if its stabilizer acts cocompactly on .
The name “conical limit point” makes more sense in the context of convergence group actions on boundaries of hyperbolic metric spaces.
Definition 2.9.
Let be a hyperbolic metric space, and let . We say that a sequence limits conically to if there is a geodesic ray limiting to and a constant such that
for some sequence .
A bounded neighborhood of a geodesic in a hyperbolic metric space looks like a “cone,” hence “conical limit.”
Proposition 2.10 ([Tuk94], [Tuk98]).
Let be a group acting properly discontinuously by isometries on a proper geodesic hyperbolic metric space , and fix a basepoint .
Then acts on as a convergence group. Moreover, a point is a conical limit point (in the dynamical sense of Definition 2.8) if and only if there is a sequence limiting conically to (in the geometric sense of Definition 2.9). In this case, there are distinct points such that and for any in .
If limits conically to a point for some (hence any) basepoint , we just say that limits conically to .
Theorem 2.11 ([Bow12]).
Let be a group acting by isometries on a hyperbolic metric space . Then is a relatively hyperbolic group, acting on as in Definition 2.2, if and only if:
-
(1)
The induced action of on is a convergence group action.
-
(2)
Every point is either a conical limit point or a bounded parabolic point.
Whenever a group acts as a convergence group on a perfect compact metrizable space , every point in is either a conical limit point or a bounded parabolic point, and the stabilizer of each parabolic point is finitely generated, we say the -action on is geometrically finite. This is justified by a theorem of Yaman [Yam04], which says that any such group action is induced by the action of a relatively hyperbolic group on a Gromov model whose boundary is equivariantly homeomorphic to . We can then identify the space with the Bowditch boundary . The set of parabolic points in coincides exactly with the set of fixed points of peripheral subgroups.
Definition 2.12.
Let be a relatively hyperbolic pair. We write
where and are respectively the conical limit points and parabolic points in .
2.3.1. Compactification of and divergent sequences
When is a relatively hyperbolic pair, there is a natural topology on the set
making it into a compactification of (i.e. is compact, and are both embedded in , and is an open dense subset of ). Specifically, we view as a subset of (any) Gromov model , via an orbit map for some basepoint . Since acts properly on , this is a proper embedding, so if we compactify by adjoining its visual boundary , we compactify as well; this does not depend on the choice of basepoint or even the choice of space .
Definition 2.13.
A sequence is divergent if it leaves every bounded subset of (equivalently, if a subsequence of it consists of pairwise distinct elements).
Up to subsequence, a divergent sequence converges to a point . When is non-elementary, the point is determined solely by the action of on : we have if and only if for all but a single .
2.4. The coned-off Cayley graph
Whenever is a relatively hyperbolic pair, there are only finitely many conjugacy classes of groups in . We can fix a finite set of conjugacy representatives for the groups in . The set corresponds to a finite set of parabolic points, such that
Then contains exactly one point in each -orbit in .
Definition 2.14.
Let be a relatively hyperbolic pair, and fix a finite generating set for and finite collection of conjugacy representatives for .
The coned-off Cayley graph is a metric space obtained from the Cayley graph as follows: for each coset for , we add a vertex . Then, we add an edge of length 1 from each to .
The quasi-isometry class of is independent of the choice of generating set . When is a relatively hyperbolic pair, is a hyperbolic metric space. It is not a proper metric space if is nonempty. The Gromov boundary of is equivariantly homeomorphic to the set of conical limit points in .
3. Lie theory notation and background
For the rest of the paper, we let be a connected semisimple Lie group with no compact factor and finite center. We will be concerned with representations , where is a relatively hyperbolic group. We want to consider the action of on the flag manifold , where is a parabolic subgroup of .
In this section, we give an overview of the definitions and notation we will use to describe the dynamical behavior of the -action on . We mostly follow the notation of [GGKW17], but we will also identify the connection to the language of [KLP17].
The exposition here is fairly brief, since most of this paper does not use much of the technical theory of semisimple Lie groups and their associated Riemannian symmetric spaces. In fact, in nearly every case, our approach will be to use a representation of to reduce to the case . The most important part of this section is 3.5, which identifies the connection between -divergence (or equivalently -regularity) and contracting dynamics in .
Standard references for the general theory are [Ebe96], [Hel01], and [Kna02]. See also section 3 of [Max21] for a careful discussion of the theory as it relates to Anosov representations and the work of Kapovich-Leeb-Porti.
3.1. Parabolic subgroups
Let be a maximal compact subgroup of the semisimple Lie group , and let be the Riemannian symmetric space . A subgroup is a parabolic subgroup if it is the stabilizer of a point in the visual boundary of . Two parabolic subgroups are opposite if there is a bi-infinite geodesic in so that is the stabilizer of and is the stabilizer of .
The compact homogeneous -space is called a flag manifold. If and are parabolic subgroups, then we say that two flags and are opposite if the stabilizers of , are opposite parabolic subgroups. (In particular a conjugate of must be opposite to ).
3.2. Root space decomposition
Let be the Lie algebra of , and let be the Lie algebra of the maximal compact . We can decompose as , and fix a maximal abelian subalgebra . The restriction of the Killing form to is positive definite, so any maximal abelian is naturally endowed with a Euclidean structure.
Each element of the abelian subalgebra acts semisimply on , with real eigenvalues. So we let denote the set of roots for this choice of , i.e. the set of nonzero linear functionals such that the linear map given by has nonzero kernel for every . We have a restricted root space decomposition
where acts on by multiplication by .
We choose a set of simple roots so that each can be uniquely written as a linear combination of elements of with coefficients either all nonnegative or all nonpositive. We let denote the positive roots, i.e. roots which are nonnegative linear combinations of elements of .
The simple roots determine a Euclidean Weyl chamber
The kernels of the roots are the walls of the Euclidean Weyl chamber.
Choosing a maximal compact , a maximal abelian , and a Euclidean Weyl chamber determines a Cartan projection
uniquely determined by the equation , where and .
3.3. -divergence
Fix a subset of the simple roots . We define a standard parabolic subgroup to be the normalizer of the Lie algebra
where is the set of positive roots which are not in the span of . The opposite subgroup is the normalizer of
Every parabolic subgroup is conjugate to a unique standard parabolic subgroup , and every pair of opposite parabolics is simultaneously conjugate to a unique pair .
For a fixed , the group is the stabilizer of the endpoint of a geodesic ray , where is the image of the identity in , and for any , the element satisfies
Definition 3.1.
Let be a sequence in . The sequence is -divergent if for every , we have
That is, the Cartan projections of the sequence drift away from the walls of determined by the subset .
For a general parabolic subgroup , we say that is -divergent if is -divergent for conjugate to .
3.4. Affine charts
Definition 3.2.
Let , be opposite parabolic subgroups in . Given a flag , we define
We call a set of the form for some an affine chart in .
An affine chart is the unique open dense orbit of in . When and is the stabilizer of a line , is identified with and this notion of affine chart agrees with the usual one in .
3.5. Dynamics in flag manifolds
There is a close connection between -divergence in the group and the topological dynamics of the action of on the associated flag manifold . Kapovich-Leeb-Porti frame this connection in terms of a contraction property for -divergent sequences.
Definition 3.3 ([KLP17], Definition 4.1).
Let be a sequence of group elements in . We say that is -contracting if there exist , such that converges uniformly to on compact subsets of .
The flag is the uniquely determined limit of the sequence .
Definition 3.4.
For an arbitrary sequence , a -limit point of in is the limit point of some -contracting subsequence of .
The -limit set of a group is the set of -limit points of sequences in .
The importance of contracting sequences is captured by the following:
Proposition 3.5 ([KLP17], Proposition 4.15).
A sequence is -divergent if and only if every subsequence of has a -contracting subsequence.
3.5 implies in particular that if is -divergent, then up to subsequence there is an open subset such that converges to a singleton in . It turns out that this “weak contraction property” is enough to characterize -divergence.
proplocalContractingImpliesGlobalContracting Let be a sequence in , and suppose that for some nonempty open subset , we have for . Then is -divergent, and has a unique -limit point .
We provide a proof of this fact in Appendix A.
3.5.1. Dynamics of inverses of -divergent sequences
When is a -divergent sequence, the inverse sequence is -divergent. Kapovich-Leeb-Porti show that this can be framed in terms of the dynamical behavior of the inverse sequence.
Lemma 3.6 ([KLP17], Lemma 4.19).
For and flags , the following are equivalent:
-
(1)
is -contracting and uniformly on compacts.
-
(2)
is -divergent, has unique -limit point , and has unique -limit point .
3.6. -regularity
-divergent sequences are equivalent to the -regular sequences discussed in the work of Kapovich-Leeb-Porti, where is the unique face corresponding to in a spherical model Weyl chamber. We explain the connection here.
Remark 3.7.
The language of -regularity is not used anywhere else in this paper, so this part of the background is provided for convenience only and may be safely skipped.
For any point , we let be the uniquely determined subspace of such that , where is the Lie algebra of the stabilizer of in .
Let . There is a point , a maximal abelian subalgebra , a Euclidean Weyl chamber , and a unit-length such that is the endpoint of the geodesic ray .
Up to the action of the stabilizer of , the point , the maximal abelian subalgebra , the Euclidean Weyl chamber , and the unit vector are uniquely determined. In addition, the stabilizer in of the triple acts trivially on .
This means that we can identify the space with the set of unit vectors in any Euclidean Weyl chamber . This set has the structure of a spherical simplex. We let denote the model spherical Weyl chamber .
We let be the type map to the model spherical Weyl chamber. For fixed , we let denote the parabolic subgroup stabilizing .
After choosing a maximal compact , a maximal abelian , and a Euclidean Weyl chamber , the data of a face of the spherical simplex is the same as the data of a subset of the simple roots of : the set of roots identifies a collection of walls of the Euclidean Weyl chamber . The intersection of those walls with the unit sphere in is uniquely identified with a face of .
Definition 3.8.
Let be a face of the model spherical Weyl chamber . We say that a sequence is -regular if is -divergent for some such that .
For a fixed model face , we let denote any parabolic subgroup which is the stabilizer of a point . All such parabolic subgroups are conjugate, so as a -space the flag manifold depends only on the model face .
4. EGF representations and relative Anosov representations
In this section we cover basic properties of the central objects of this paper: extended geometrically finite representations from a relatively hyperbolic group to a semisimple Lie group with no compact factor and trivial center. We also show that they generalize a definition of relative Anosov representation (Section 1.7), and prove our Anosov relativization theorem (Theorem 1.11).
We refer also to Section 2 of the related paper [Wei23b] for an overview of the definition in the special case where or and the parabolic subgroup is the stabilizer of a flag of type in .
Definition 4.1.
Let be a parabolic subgroup of . We say that is symmetric if is conjugate to a subgroup opposite to .
When is symmetric, we can identify with , so that it makes sense to say that two flags are opposite.
Definition 4.2.
Let be symmetric, and let be two subsets of . We say that and are opposite if every is opposite to every .
Definition 4.3.
Let be a relatively hyperbolic pair, and let for a symmetric parabolic . We say that a continuous surjective map is antipodal if for every pair of distinct points , is opposite to .
We recall the main definition of the paper here:
*
Remark 4.4.
Unfortunately, the boundary set is not necessarily uniquely determined by the representation . In many contexts, we will be able to make a natural choice, but we do not give a procedure for doing so in general.
4.1. Discreteness and finite kernel
When is EGF, the action of on the boundary set is by definition an extension of the topological dynamical system . When is non-elementary, convergence dynamics imply that the homomorphism has finite kernel and discrete image. So the map must also have discrete image and finite kernel, and therefore so does the representation . The case where is elementary can be verified directly.
4.2. Shrinking the sets
Let be an EGF representation with boundary map . By assumption, we know there exists an open subset for each , satisfying the extended convergence dynamics conditions (Definition 1.1). In general, there is not a canonical choice for the set . We are able to make some assumptions about the properties of the , however.
Proposition 4.5.
Let be an EGF representation with boundary extension . For any , we can choose the set to be a subset of
Proof.
Since is closed, is an open subset of . And, transversality of implies that contains . So the intersection is open and nonempty, meaning we can replace with this intersection. ∎
4.3. An equivalent characterization of EGF representations
It is often possible to prove properties of relatively hyperbolic groups by first showing that the property holds for conical subsequences in the group, and then showing that the property holds inside of peripheral subgroups. There is a characterization of the EGF property along these lines, which is frequently useful for constructing examples of EGF representations (see [Wei23b]).
propconicalPeripheralEGF Let be a representation of a relatively hyperbolic group, and let be a closed -invariant set, where is a symmetric parabolic subgroup. Suppose that is a continuous surjective -equivariant antipodal map.
Then is an EGF representation with EGF boundary extension if and only if both of the following conditions hold:
-
(a)
For any sequence limiting conically to some point in , is -divergent and every -limit point of lies in .
-
(b)
For every parabolic point , there exists an open set , with , such that for any compact and any open set containing , for all but finitely many , we have .
The proof of Section 4.3 requires the technical machinery of relative quasigeodesic automata, so we defer it to Section 8. At the end of Section 8, we also provide another (weaker) characterization of EGF representations which may be of interest.
4.4. Properties of
Proposition 4.6.
Let be a relatively hyperbolic pair, and let be a representation which is EGF with respect to a symmetric parabolic , with boundary extension . Then contains the -limit set of .
Proof.
Let be a flag in the -limit set of . Then there is a -contracting sequence for and a flag such that converges to for any in . Up to subsequence converges to , so for any flag , the sequence subconverges to a point in . But since is open and dense, for some we have and hence . ∎
In particular, 4.6 implies that the EGF boundary set of an EGF representation must always contain the -proximal limit set of . (Recall that is -proximal if it has a unique attracting fixed point in ; the -proximal limit set of a subgroup of is the closure of the set of attracting fixed points of -proximal elements).
We will see that most of the power of EGF representations lies in the fact that their associated boundary extensions do not have to be homeomorphisms (so the Bowditch boundary of does not need to be equivariantly embedded in any flag manifold). However, it turns out that it is always possible to choose the boundary extension so that it has a well-defined inverse on conical limit points in . In fact, we can even get a somewhat precise description of all the fibers of . Concretely, we have the following:
Proposition 4.7.
Let be an EGF representation, with boundary extension . There is a -invariant closed subset and a -equivariant map such that:
-
(1)
is also a boundary extension for ,
-
(2)
for every , is a singleton, and
-
(3)
for every , is the closure of the set of all accumulation points of orbits for a sequence of distinct elements in and .
We will prove 4.7 at the end of Section 9, where it will follow as a consequence of the proof of the relative stability theorem for EGF representations (Section 1.4)—see Remark 9.18.
We will rely on both Section 4.3 and 4.7 to prove the rest of the results in this section (which are not needed anywhere else in this paper).
4.5. Relatively Anosov representations
EGF representations give a strict generalization of the relative Anosov representations mentioned in the introduction. We give a precise definition here.
Definition 4.8 ([KL18, Definition 7.1] or [ZZ22, Definition 1.1]; see also [ZZ22, Proposition 4.4]).
Let be a subgroup of and suppose is a relatively hyperbolic pair. Let be a symmetric parabolic subgroup.
The subgroup is relatively -Anosov if it is -divergent, and there is a -equivariant antipodal embedding whose image is the -limit set of .
Here, we say an embedding is antipodal if for every distinct in , and are opposite flags.
Remark 4.9.
Several remarks on the definition are in order:
-
(a)
In [KL18], Kapovich-Leeb provide several possible ways to relativize the definition of an Anosov representation; Definition 4.8 agrees with essentially their most general definition, that of a relatively asymptotically embedded representation.
-
(b)
When is a hyperbolic group (and the collection of peripheral subgroups is empty), then the Bowditch boundary is identified with the Gromov boundary . In this case, Definition 4.8 coincides with the usual definition of an Anosov representation.
-
(c)
In general, it is possible to define (relatively) -Anosov representations for a non-symmetric parabolic subgroup . However, there is no loss of generality in assuming that is symmetric: a representation is -Anosov if and only if it is -Anosov for a symmetric parabolic subgroup depending only on .
Proposition 4.10.
Let be an EGF representation with respect to , and suppose that the boundary extension is a homeomorphism. Then:
-
(1)
is -divergent, and is the -limit set of .
-
(2)
The sets for can be taken to be
Proof.
(1). Let be any infinite sequence of elements in . After extracting a subsequence, we have , and since is a homeomorphism, converges to the point uniformly on compacts in the open set . Then 3.5 implies that is -divergent, with unique -limit point .
(2). The fact that is antipodal is exactly the statement that the sets contain for every , so we just need to see that the appropriate dynamics hold for these sets. Let be an infinite sequence in with for .
Using the previous proposition, we can see the relationship between relatively Anosov representations and EGF representations (Section 1.7). We restate this theorem as the following:
Proposition 4.11.
Let be a relatively hyperbolic pair, and let be a symmetric parabolic subgroup. A representation is relatively -Anosov in the sense of Definition 4.8 if and only if is EGF with respect to , and has an injective boundary extension .
Proof.
4.10 ensures that if is an EGF representation, and the boundary extension is a homeomorphism, then is -divergent and is an antipodal embedding whose image is the -limit set.
On the other hand, if is relatively -Anosov, with boundary embedding , for each , we can take
Antipodality means that contains , and -divergence and 3.6 imply that has the appropriate convergence dynamics. ∎
4.6. Relativization
We now turn to the situation where we have an EGF representation of a hyperbolic group with a nonempty collection of peripheral subgroups. That is, for some invariant set , we have an EGF boundary extension , where is the Bowditch boundary of with peripheral structure .
We want to prove Theorem 1.11, which says that in this situation, if restricts to a -Anosov representation on each , then is a -Anosov representation of . For the rest of this section, we assume that is a hyperbolic group, and is a collection of subgroups of so that the pair is relatively hyperbolic. We let be an EGF representation for the pair with respect to a symmetric parabolic subgroup , and we assume that for each , is -Anosov, with Anosov limit map .
The main step in the proof is to observe that it is always possible to choose the boundary extension so that is equivariantly homeomorphic to the Gromov boundary of (which we here denote ).
Whenever is a hyperbolic group and is a collection of subgroups so that is a relatively hyperbolic pair, there is an explicit description of the Bowditch boundary in terms of the Gromov boundary of —see [Ger12], [GP13], or [Tra13]. Specifically, we can say:
Proposition 4.12.
There is an equivariant surjective continuous map such that for each conical limit point in , is a singleton, and for each parabolic point with , is an embedded copy of in .
In our situation, we can see that the boundary extension satisfies similar properties.
Lemma 4.13.
There is a closed -invariant subset and an EGF boundary extension such that:
-
(1)
For each conical limit point , is a singleton.
-
(2)
For each parabolic point , with , we have .
Proof.
We choose as in 4.7. The only thing we need to check is that for , the set is exactly the closure of the set of accumulation points of -orbits in . But since we may assume is contained in , this follows immediately from the fact that is -divergent and the closed set is the -limit set of .
∎
Next we need a lemma which will allow us to characterize the Gromov boundary of as an extension of the Bowditch boundary . First recall that if acts as a convergence group on a space , the limit set of is the set of points such that for some and some sequence , we have
uniformly on compacts.
Lemma 4.14.
Let act on compact metrizable spaces and , and let , be continuous equivariant surjective maps such that for every conical limit point , and are both singletons, and for every parabolic point , acts as a convergence group on and , with limit sets , equivariantly homeomorphic to .
Then for any sequences , we have
if and only if
Proof.
We proceed by contradiction, and suppose that for a pair of sequences , we have
but
After taking a subsequence we may assume converges to , and that converges to and converges to for . By continuity, we have
Since and are bijective on and respectively, we must have for a parabolic point . Let .
Since is a bounded parabolic point, we can find sequences of group elements so that for a fixed compact subset , we have
(2) |
This implies that no subsequence of or converges to a point in .
Then, since acts as a convergence group on with limit set , up to subsequence there are points so that converges to a point in uniformly on compacts in , and converges to a point in uniformly on compacts in . So, we must have and .
This means that the sequences and have distinct limits in the compactification . So, there are distinct points so that (again up to subsequence) converges to a point in uniformly on compacts in , and converges to a point in uniformly on compacts in . Without loss of generality, we can assume .
But then lies in a compact subset of , so converges to a point in and converges to . But this contradicts (2) above. ∎
Proposition 4.15.
If the set satisfies the conclusions of 4.13, then is equivariantly homeomorphic to the Gromov boundary of .
Proof.
Let denote the quotient map identifying the limit set of each to the parabolic point with . For each conical limit point , the fiber is a singleton. So, there is an equivariant bijection from to .
Moreover, since is -invariant, and the action of on its Gromov boundary is minimal, is dense in . We claim that extends to a continuous injective map by defining for any sequence .
To see this, we can apply 4.14, taking and . We know that always acts on its own Gromov boundary as a convergence group (so in particular each acts on as a convergence group with limit set ). And, since restricts to a -Anosov representation on each , for any infinite sequence , up to subsequence there are so that converges to uniformly on compacts in . Antipodality of implies that converges to uniformly on compacts in . The other hypotheses of 4.14 follow from 4.12 and 4.13.
We still need to check that is actually surjective. We know that restricts to a bijection on , and that takes to for each parabolic point in . So we just need to check that for every , restricts to a surjective map . If is non-elementary, this must be the case because the action of on is minimal and maps into as an invariant closed subset. Otherwise, is virtually cyclic and , both contain exactly two points. Then injectivity of implies surjectivity.
So we conclude that there is a continuous bijection , and since is compact and is metrizable, is a homeomorphism. ∎
We let denote the equivariant homeomorphism from 4.15. The final step in the proof of Theorem 1.11 is the following:
Proposition 4.16.
The equivariant homeomorphism extends the convergence action of on its Gromov boundary .
Proof.
By Section 4.3, we just need to show that if is a conical limit sequence with for , then every -limit point of lies in .
We consider two cases:
Case 1: is a parabolic point in . In this case, lies along a quasigeodesic ray in limiting to some , with . This means that for a bounded sequence , we have . Since restricts to a -Anosov representation on , this means that is -divergent and every -limit point of lies in . For the same reason, every -limit point of lies in .
Up to subsequence is a constant . We can use 3.5 to see that has the same -limit set as . This -limit set lies in . And, every -limit point of is a -translate of a -limit point of . This -limit set also lies in .
Case 2: is a conical limit point in . In this case, a subsequence of is a conical limit sequence for the action of on , and the desired result follows from the “only if” part of Section 4.3.
∎
Proof of Theorem 1.11.
Let be hyperbolic, let be a collection of subgroups such that is a relatively hyperbolic pair, and let be an EGF representation with respect to , for the peripheral structure .
Suppose that restricts to a -Anosov representation on each . 4.16 implies that is also an EGF representation of for its empty peripheral structure, whose boundary extension can be chosen to be a homeomorphism. Then Section 1.7 says that is relatively -Anosov (again for the empty peripheral structure on ). This ensures that is actually (non-relatively) -Anosov; see e.g. [KLP17, Theorem 1.1]. ∎
5. Relative quasigeodesic automata
In the next three sections, we develop the technical tools needed to prove the main results of the paper: namely, a relative quasigeodesic automaton for a relatively hyperbolic group acting on a flag manifold , and a system of open sets in which is in some sense compatible with both the relative quasigeodesic automaton and the action of on .
The basic idea is motivated by the computational theory of hyperbolic groups. Given a hyperbolic group with finite generating set , it is always possible to find a finite directed graph , with edges labeled by elements of , so that directed paths on starting at a fixed vertex are in one-to-one correspondence with geodesic words in . The graph is called a geodesic automaton for .
Geodesic automata are really a manifestation of the local-to-global principle for geodesics in hyperbolic metric spaces: the fact that the automaton exists means that it is possible to recognize a geodesic path in a hyperbolic group just by looking at bounded-length subpaths.
In this section of the paper, we consider a relative version of a geodesic automaton. This is a finite directed graph which encodes the behavior of quasigeodesics in the coned-off Cayley graph of a relatively hyperbolic group . Eventually, our goal is to build such an automaton by looking at the dynamics of the action of on its Bowditch boundary . The main result of this section is 5.13, which says that we can construct such a relative quasigeodesic automaton for a relatively hyperbolic pair using an open covering of the Bowditch boundary which satisfies certain technical conditions.
In this section of the paper and the next, we will work in the general context of a relatively hyperbolic group acting by homeomorphisms on a connected compact metrizable space , before returning to the case where is a flag manifold for the rest of the paper.
Throughout the rest of this section, we fix a non-elementary relatively hyperbolic pair , and let be a finite set, containing exactly one point from each -orbit in . We also fix a finite generating set for , which allows us to refer to the coned-off Cayley graph (Definition 2.14).
Definition 5.1.
A -graph is a finite directed graph where each vertex is labelled with a subset , which is either:
-
•
A singleton , with , or
-
•
A cofinite subset of a coset for some , .
A sequence is a -path if for a vertex path in .
Remark 5.2.
We will often refer to “the” vertex path corresponding to a -path , although we will never actually verify that such a vertex path is uniquely determined by the sequence of group elements in .
A vertex of a -graph which is labeled by a cofinite subset of a (necessarily unique) coset is a parabolic vertex. If is a parabolic vertex, we let denote the corresponding parabolic point in .
Remark 5.3.
It will be convenient to allow parabolic vertices to be labeled by cofinite subsets of peripheral cosets (instead of just the entire coset) when we construct -graphs using the convergence dynamics of the -action on .
Definition 5.4.
Let . We say that a -path limits to if either:
-
•
, is infinite, and the sequence
limits to in the compactification , or
-
•
, is a finite -path whose corresponding vertex path ends at a parabolic vertex , and
Definition 5.5.
Let be a -graph. The endpoint of a finite -path is
Definition 5.6.
A -graph is a relative quasigeodesic automaton if:
-
(1)
There is a constant so that for any infinite -path , the sequence
lies Hausdorff distance at most from a geodesic ray in , based at the identity.
-
(2)
For every , there exists a -path limiting to .
One way to think of a relative quasigeodesic automaton is that it gives us a system for finding quasigeodesic representatives of every element in the group. More concretely, we have the following:
Lemma 5.7.
Let be a relative quasigeodesic automaton. There is a constant so that set of endpoints of -paths is -dense in .
Proof.
If is hyperbolic and is empty, then this is a consequence of the Morse lemma and the fact that the union of the images of all infinite geodesic rays based at the identity in is coarsely dense in (see [Bog97]).
If is nonempty, there is some so that the union of all of the cosets for is -dense in . So it suffices to show that for each , there is some so that all but elements in any coset are the endpoints of a -path.
For any such coset , we can find a finite -path limiting to the vertex . That is,
By definition with a cofinite subset of the coset That is,
so for all but finitely many (depending only on the size of the complement of in ), we can find with
∎
Remark 5.8.
In general, we do not require the set of elements in labelling the vertices of a relative quasigeodesic automaton to generate the group (although the proposition above implies that they at least generate a finite-index subgroup).
5.1. Compatible systems of open sets
A relative quasigeodesic automaton always exists for any relatively hyperbolic group (although we will not prove this fact in full generality). We will give a way to construct a relative quasigeodesic automaton using the convergence group action of a group acting on its Bowditch boundary.
Definition 5.9.
Suppose that acts on a metrizable space by homeomorphisms, and let be a -graph. A -compatible system of open sets for the action of on is an assignment of an open subset to each vertex of such that for each edge in , for some , we have
(3) |
for all .
Remark 5.10.
Proposition 5.11.
Let be a -graph, and let be a -compatible system of subsets of for the action of on .
There is a constant satisfying the following: let be an infinite -path, corresponding to a vertex path , and suppose the sequence is divergent in . Then for any point in the intersection
the sequence lies within Hausdorff distance of a geodesic ray in tending towards .
Proof.
Fix a point , and write and . We first claim that there is a uniform and a point such that
(4) |
for all .
To prove the claim, choose a uniform so that for every vertex in , we have , and for every edge in and every , we have . Then we choose some .
By the -compatibility condition, we know that for any , , so we know that .
Then, for any , we have
Moreover since , we also have
So for all we have , which establishes that (4) holds for all .
Now, consider a bi-infinite geodesic in a cusped space for joining and . The sequence of geodesics has endpoints in lying distance at least apart, so each geodesic in the sequence passes within a uniformly bounded neighborhood of a fixed basepoint . Therefore lies in a uniformly bounded neighbood of the geodesic .
Since is divergent, can only accumulate at either or . But in fact can only accumulate at —for in the construction of above, we could have chosen any in the nonempty open set , and since is perfect there is at least one such .
This implies that is a conical limit sequence in , limiting to . Since the distance between and is bounded in , the desired conclusion follows. ∎
Definition 5.12.
Let be a -graph. An infinite -path is divergent if the sequence leaves every bounded subset of .
We say that a -graph is divergent if every infinite -path is divergent.
Whenever is a -compatible system of open sets for a -graph , one can think of a -path as giving a symbolic coding of a point in the intersection
The following proposition gives a way to construct such a coding for a given point , given an appropriate pair of open coverings of the Bowditch boundary compatible with a -graph .
Proposition 5.13.
Let be a divergent -graph. Suppose that for each vertex , there exist open subsets of such that the following conditions hold:
-
(1)
The sets give a -compatible system of sets for the action of on .
-
(2)
For all vertices , we have and .
-
(3)
The sets give an open covering of .
-
(4)
For every and every non-parabolic vertex such that , there is an edge in such that for .
-
(5)
For every and every parabolic vertex such that , there is an edge in and such that .
Then is a relative quasigeodesic automaton for .
Proof.
5.11 implies that any infinite -path lies finite Hausdorff distance from a geodesic ray in . So, we just need to show that every is the limit of a -path.
The idea behind the proof is to use the fact that the sets cover to show that we can keep “expanding” a neighborhood of in to construct a path in limiting to . The covering tells us how to find the next edge in the path, and the cover gives us the -compatible system we need to show that the path is a geodesic.
We let denote the vertex set of . When is not a parabolic vertex, we write .
Case 1: is a conical limit point. Fix so that . We take , , and define sequences , , and as follows:
-
•
If is not a parabolic vertex, then we choose . Let . Since conical limit points are invariant under the action of , is a conical limit point. By condition 4, there is a vertex satisfying with an edge in .
-
•
If is a parabolic vertex, then since is a conical limit point, for . Then condition 5 implies that there exists some so that for an edge in . Again, must be a conical limit point since is -invariant.
The sequence necessarily gives a -path. By assumption the sequence
is divergent. And by construction lies in for all . So, 5.11 implies that is a conical limit sequence, limiting to . See Figure 2.
Case 2: is a parabolic point. As before fix so that , and take , . We inductively define sequences , , as before, but we claim that for some finite , is a parabolic vertex with . For if not, we can build an infinite -path (as in the previous case) limiting to . But then, 5.11 would imply that is actually a conical limit point. So, we must have
as required.
∎
Remark 5.14.
In a typical application of 5.13, it will not be possible to construct the open coverings and so that for all vertices . In particular we expect this to be impossible whenever is connected.
To conclude this section, we make one more observation about systems of -compatible sets as in 5.13.
Lemma 5.15.
Let be a relatively hyperbolic group, let be a -graph, and let , be an assignment of open subsets of to vertices of satisfying the hypotheses of 5.13.
Fix and . There exists so that if , then there are -paths limiting to respectively, with for all .
6. Extended convergence dynamics
Let be a relatively hyperbolic group acting on a connected compact metrizable space . In this section, we will show that if the action of on extends the convergence dynamics of (Definition 1.1), then we can construct a relative quasigeodesic automaton and a -compatible system of open subsets of which are in some sense reasonably well-behaved with respect to the group action.
To give the precise statement, we let be a closed -invariant subset, and let be an equivariant, surjective, and continuous map satisfying the following: for each , there is an open set containing such that:
-
(1)
For any sequence limiting conically to , with , any open set containing , and any compact , we have for all sufficiently large .
-
(2)
For any parabolic point , any compact , and any open set containing , for all but finitely many , we have .
Note that in particular, any map extending convergence dynamics satisfies these conditions. For the rest of this section, however, we only assume that (1) and (2) both hold for our map . In this context, we will show:
Proposition 6.1.
For any , there is a relative quasigeodesic automaton for , a -compatible system of open sets for the action of on , and a -compatible system of open sets for the action of on such that:
-
(1)
For every vertex , there is some so that
-
(2)
For every , there is a parabolic vertex with . Moreover, for every parabolic vertex with , is an edge of if and only if is an edge of .
-
(3)
If for , is a parabolic vertex with , and is an edge of , then and .
Remark 6.2.
By equivariance of , for each , we can replace with and assume that is -invariant (and that if , then ).
The proof of 6.1 involves some technicalities, so we first outline the general approach:
-
(1)
For each , we construct a pair , of small open neighborhoods of and a subset so that for each , is “expanding” about some point in . When is a conical limit point, then we can choose a single element which expands about every point in . When is a parabolic point, we may use a different element of to “expand” about each .
We choose , , and so that if is “expanding” about , and , then . See Figure 3.
Figure 3. The group element is “expanding” about . We will construct , and so that if lies in , then contains . Equivalently, we get the containment illustrated earlier in Figure 1. -
(2)
Using compactness of , we pick a finite set of points so that the sets give an open covering of . These points in are identified with the vertices of a -graph . We define the edges of in such a way so that if, for some , expands about and , then there is an edge from to . This ensures that is a -compatible system of open subsets of .
-
(3)
Simultaneously, we construct a -compatible system of open sets in by taking to be a small neighborhood of . The idea is to use the extended convergence dynamics to ensure that if, for some , “expands” about some and the point lies in , then contains . See Figure 6 below.
-
(4)
Finally, we use 5.13 to prove that is actually a relative quasigeodesic automaton. The open sets are constructed exactly to satisfy the conditions of the proposition, so the main thing to check in this step is that the graph is actually divergent (using the action of on ).
Throughout the rest of the section, we will work with fixed metrics on both and . Critically, none of our “expansion” arguments will depend sensitively on the precise choice of metric. That is, in the sketch above, when we say that some group element “expands” on a small open subset of a metric space , we just mean that is quantifiably “bigger” than , and not that for any , we have for some expansion constant . 6.5 and 6.7 below describe precisely what we mean by “bigger.” The general idea is captured by the following example.
Example 6.3.
We consider the group . While is virtually a free group (and therefore word-hyperbolic), it is also relatively hyperbolic, relative to the collection of conjugates of the parabolic subgroup .
Since acts with finite covolume on the hyperbolic plane , the Bowditch boundary of the pair is equivariantly identified with , the visual boundary of . Given a non-parabolic point , we can find an element of which “expands” a neighborhood of . There are two distinct possibilities:
-
(1)
Suppose is in a small neighborhood of a conical limit point . Then choose some loxodromic element whose attracting fixed point is close to . Then, if is a slightly larger neighborhood of , is large enough to contain a uniformly large neighborhood of . See Figure 4.
-
(2)
On the other hand, suppose is in a small neighborhood of a parabolic fixed point , but . We can find some element so that takes into a fundamental domain for the action of on . Then, if is a slightly larger neighborhood of , is again large enough to contain a uniformly large neighborhood of . See Figure 5.
There is a slight issue with this approach: in the second case above (when is close to a parabolic point ), it is actually not quite good enough to “expand” a neighborhood of by using to push into a fundamental domain for on . The reason is that there might be no such fundamental domain which is actually far away from . We resolve this issue by instead choosing to lie in a coset , where for some . Then lies in a fundamental domain for on , which allows us to get uniform control on the size of the expanded neighborhood .
The two technical lemmas below (6.5 and 6.7) essentially say that one can set up this kind of expansion simultaneously on the Bowditch boundary of our relatively hyperbolic group and in a neighborhood of the -invariant set . The precise formulation of the expansion condition found in these two lemmas is best motivated by the proof of 6.10 below, which shows that the “expanding” open sets we construct give rise to a -compatible system of open sets on a -graph .
Lemma 6.4.
There exists (depending on and ) so that for any with , the -neighborhood of in is contained in .
Proof.
Since is closed in , such an exists for any fixed pair of distinct . Then the result follows, since the space of pairs satisfying is compact. ∎
Lemma 6.5.
There exists satisfying the following: for any , with , , and every conical limit point , we can find:
-
•
A group element
-
•
Open subsets with
such that:
-
(1)
,
-
(2)
In , we have
-
(3)
In we have
Remark 6.6.
Conditions (1) and (2) together imply that for any , if intersects , then . Later, we will see that condition (3) implies that if intersects , then also (giving us the inclusion indicated by Figure 3).
Proof.
For a conical limit point , we choose a sequence so that for distinct , we have and for any . That is, limits conically to in , and converges (not necessarily conically) to . Since the -action on distinct pairs in is cocompact (2.7), we may assume that for a uniform constant .
We choose from 6.4 so that if satisfy , then a -neighborhood of is contained in . Let satisfy , and let satisfy .
By the triangle inequality, we have for all , so the closed -neighborhood of is contained in . This means that we can choose large enough so that
is contained in and
is contained in . We let for this large , and take
and
∎
The next lemma is a version of 6.5 for parabolic points. As before, we want to show that for a point in the Bowditch boundary, we can find a neighborhood of in with uniformly bounded diameter , and group elements so that enlarges enough to contain a -neighborhood of , for some close to . Simultaneously we want to choose so that enlarges an -neighborhood of in a similar manner. This case is more complicated, because we need to allow to depend on the point : if is a parabolic point in , then in general there is not a single group element in which expands distances in a neighborhood of .
Lemma 6.7.
For each point , there exist constants , such that for any , any , and any , we can find:
-
•
A cofinite subset of the coset ,
-
•
Open neighborhoods of , with ,
-
•
Open neighborhoods of with
such that:
-
(1)
, and .
-
(2)
in , we have
-
(3)
For every , there exists with .
-
(4)
For every , we have
and
-
(5)
is contained in and contains .
Remark 6.8.
Proof.
Pick a compact set so that covers . Choose small enough so that the closure of does not contain . Then, for any , we can pick
We can choose sufficiently small so that a -neighborhood of is contained in . Now, fix . We claim that for a cofinite subset , for any , we have
(5) | |||||
(6) |
To see that this claim holds, it suffices to verify that for any infinite sequence of distinct group elements in , (5) and (6) both hold for all sufficiently large .
We write for . Then converges uniformly to on compact subsets of , so converges uniformly to on compact subsets of , implying that (5) eventually holds. And by our assumptions, we know that
for sufficiently large , implying that (6) also eventually holds.
So we can take to be the set
and to be the set
To see that and are open we just need to verify that they each contain a neighborhood of . Since and each contain , and covers , and each contain the set
Since is cofinite in this is an open set containing . ∎
6.1. Construction of the relative automaton
We will construct the relative automaton satisfying the conditions of 6.1 by choosing a suitable open covering of , and then using compactness to take a finite subcover. The subsets of this subcover will be the vertices of .
We choose constants , so that , (where , are the constants coming from 6.5) and , for each (where are the constants coming from 6.7).
Then:
-
•
For each , we define , , as in 6.5, with parameters , .
-
•
For each , we define , and as in 6.7, again with parameters , .
The collections of sets and together give an open covering of the Bowditch boundary . So we choose a finite subcover , which we can write as
where is a finite subset of . We can in particular ensure that contains the finite set .
We identify the vertices of our -graph with . For each , the set is either (if is a conical limit point) or (if for a parabolic point ). Then, for each , we define the open sets by
The edges of the -graph are defined as follows:
-
•
For with , there is an edge from to if is nonempty.
-
•
If with , there is an edge from to if is nonempty.
Since is an open covering of , and the sets and are nonempty, every vertex of has at least one outgoing edge. Moreover, for any parabolic point , the set depends only on the orbit of in , so must satisfy condition (2) in 6.1.
Proposition 6.9.
For each , we have
Proof.
Next we verify:
Proposition 6.10.
The collection of sets and are both -compatible systems of open sets for the -graph .
Proof.
First fix an edge with . Since is nonempty, part 2 of 6.5 implies that contains the -neighborhood of some point . Since and , we can find a small so that .
In particular, contains , which means that contains , which contains . Then, part 3 of 6.5 implies that is contained in .
We observe:
Proposition 6.11.
The -compatible systems of open subsets and satisfy conditions (1) - (3) in 6.1.
Proof.
To finish the proof of 6.1, we now just need to show:
Proposition 6.12.
The -graph is a relative quasigeodesic automaton for the pair .
Proof.
We apply 5.13, using the -compatible system and the sets we defined in the construction of .
The first three conditions of 5.13 are satisfied by construction. To see that conditions 4 and 5 hold, first observe that if for a non-parabolic vertex , then lies in some and is an edge in . And if for a parabolic vertex , then part (3) of 6.7 says that there is some such that . If contains , the edge must be in .
It only remains to check that is a divergent -graph. Let be an infinite -path, following a vertex path . The -compatibility condition implies that is a subset of for every . Since is connected and each is a proper subset of , this inclusion must be proper. This implies that in the sequence , no element can appear more than times and therefore is divergent. ∎
Remark 6.13.
This last step is the only part of the proof of 6.1 which uses the connectedness of . This hypothesis is likely unnecessary, but omitting it would involve introducing additional technicalities in the construction of the sets , —and as stated, the proposition is strong enough for our purposes.
Note that with this hypothesis removed, 6.1 would imply that any non-elementary relatively hyperbolic group has a relative quasigeodesic automaton (by taking ). As stated, the proposition only shows that such an automaton exists when is connected.
We conclude this section by observing that one can slightly refine the construction in 6.1 to obtain some stronger conditions on the resulting automaton.
Proposition 6.14.
Fix a compact subset of the Bowditch boundary . Then, for any open set containing , there is a relative quasigeodesic automaton and a pair of -compatible systems of open sets , as in 6.1, additionally satisfying the following: any is the limit of a -path (with corresponding vertex path ) such that .
Proof.
We choose so that contains . We then construct our relative quasigeodesic automaton as in the proof of 6.1, but we also choose a finite subset so that the sets for give a finite open covering of . We can ensure that the vertex set of contains .
Then, for any , by the construction in 5.13, we can find a -path limiting to whose first vertex is some . The corresponding open set for this vertex is . ∎
Proposition 6.15.
For each parabolic point , let be a compact subset of such that . Then the relative quasigeodesic automaton in 6.1 can be chosen to satisfy the following:
For every parabolic vertex with , and every , there is a -path limiting to whose first vertex is connected to by an edge .
Proof.
The proof of 6.7 shows that in our construction of the relative automaton, we can ensure that each set contains . So if , then by definition of the automaton, lies in lies in with connected to by a directed edge. Then, following the proof of 5.13, we can find a -path limiting to whose first vertex is . ∎
7. Contracting paths in flag manifolds
Let be a discrete relatively hyperbolic group, and let be a -graph. Fix a pair of opposite parabolic subgroups , . Our goal in this section is to show that under certain conditions, if is a -compatible system of open subsets of for the action of on , then the sequence of group elements lying along an infinite -path is -divergent.
7.1. Contracting paths in -graphs
Definition 7.1.
Let be a discrete subgroup of , let be a -graph, and let be a -compatible system of open subsets of . We say that a -path is contracting if the decreasing intersection
(7) |
is a singleton in .
Definition 7.2.
We say that an open set is a proper domain if the closure of lies in an affine chart for some .
Here is the main result in this section:
Proposition 7.3.
Let be a -graph for , and let be -compatible system of open subsets of .
If the set is a proper domain for each vertex of the automaton, then every infinite -path is contracting.
7.2. A metric property for bounded domains in flag manifolds
To prove 7.3, we consider a metric defined by Zimmer [Zim18] on any proper domain . is defined so that it is invariant under the action of on : for any in some proper domain , and any , we have
(8) |
In general, is not a complete metric. However, induces the standard topology on as an open subset of . We will show that for a -path , the diameter of
with respect to tends to zero as .
Zimmer’s construction of depends on an irreducible representation for some real vector space . This is provided by a theorem of Guéritaud-Guichard-Kassel-Wienhard.
Theorem 7.4 ([GGKW17], see also [Zim18], Theorem 4.6).
There exists a real vector space , an irreducible representation , a line , and a hyperplane such that:
-
(1)
.
-
(2)
The stabilizer of in is and the stabilizer of in is .
-
(3)
and are opposite if and only if and are transverse.
The representation determines a pair of embeddings and by
In this section, we will identify with the space of projective hyperplanes in , by identifying the projectivization of a functional with the projectivization of its kernel.
Definition 7.5.
Let be an open subset of . The dual domain is
Note that is open if and only if is a proper domain.
Definition 7.6.
Let , and let . The cross-ratio is defined by
where , are respectively lifts of and in and .
Remark 7.7.
When is two-dimensional, we can identify the projective line with by identifying each with . In that case, the cross-ratio defined above agrees with the standard four-point cross-ratio on , given by
(9) |
The differences in (9) can be measured in any affine chart in containing . Our convention is chosen so that if we identify with , we have .
Definition 7.8.
Let be a proper domain. We define the function by
For any and any proper domain , we have . So must satisfy the -invariance condition (8).
If is a properly convex subset of , and , , are the identity maps on , , and respectively, then agrees with the well-studied Hilbert metric on . More generally we have:
Theorem 7.9 ([Zim18], Theorem 5.2).
If is open and bounded in an affine chart, then is a metric on which induces the standard topology on as an open subset of .
Remark 7.10.
This particular result in [Zim18] is stated only for noncompact simple Lie groups, but the proof only assumes that is semisimple with no compact factor.
Since taking duals of proper domains reverses inclusions, it follows that if , then . Our goal now is to sharpen this inequality, and show:
Proposition 7.11.
Let , be proper domains in , such that .
There exists a constant (depending on and ) so that for all ,
A consequence is the following, which in particular implies 7.3.
Corollary 7.12.
Let be a -graph for a relatively hyperbolic group , and let be a -compatible system of open subsets of . If each is a proper domain, then there are constants so that for any -path in the -graph , the diameter of
with respect to is at most
Proof.
For any open set and , we let denote the diameter of with respect to the metric . We choose a uniform so that in some fixed metric on , every edge in , and every , we have
Then for each vertex set , we write .
We take
7.11 implies that there exists such that for all , we have
Take . We claim that for all , we have
We prove the claim via induction on the length of the -path . For , the claim is true because . For , we can assume
Then we have
Finally, the claim implies the corollary because we know that
So, we can replace with to get the desired result. ∎
We now proceed with the proof of 7.11. We first observe that in the special case where are properly convex subsets of real projective space, one can show the desired result essentially via the following:
Proposition 7.13.
Let be points in , arranged so that in a cyclic ordering on . Then there exists a constant , depending only on the cross-ratio , so that for all distinct , we have
7.13 is a standard fact in real projective geometry and can be verified by a computation. Note that we allow the degenerate case : in this situation the right-hand side is identically zero for distinct . We allow no other equalities among , so the cross-ratio lies in .
To apply 7.13 to our situation, we need to get some control on the behavior of the embeddings and . We do so in the next three lemmas below.
Lemma 7.14.
Let be distinct points in . There exists a one-parameter subgroup such that fixes and , and acts nontrivially on the projective line spanned by and .
Proof.
We can write for some . Let denote an abelian subalgebra of the Lie algebra of , such that for a maximal compact , the exponential map induces an isometric embedding whose image is a maximal flat in .
There is a conjugate of such that the action of on fixes both and (see [Ebe96], Proposition 2.21.14). So, up to the action of on , we can assume that is fixed by a standard parabolic subgroup conjugate to , and that are both fixed by the subgroup .
We choose so that for all . Then is a 1-parameter subgroup of fixing . As , is -divergent, with unique attracting fixed point .
Then [GGKW17], Lemma 3.7 implies that is -divergent, where is the stabilizer of a line in , and is the unique one-dimensional eigenspace of whose eigenvalue has largest modulus. And, since fixes and , preserves , and acts nontrivially since the eigenvalues of on and must be distinct. ∎
Lemma 7.15.
Let be any projective line in tangent to the image of the embedding at a point for . There exists a one-parameter subgroup of so that acts nontrivially on with unique fixed point .
Proof.
Fix a sequence such that and the projective line spanned by and converges to . By 7.14, there exists so that acts nontrivially on , with fixed points and .
In the projectivization , converges to some . Since has finite kernel, there is an induced map , which satisfies
A continuity argument shows that the one-parameter subgroup acts nontrivially on the line , and has unique fixed point at . ∎
Lemma 7.16.
Let be a proper domain, and let be a projective line in which is either spanned by two points in , or is tangent to at a point for . Then
is a nonempty open subset of .
Proof.
is nonempty since is nonempty. So let , and choose so that . We need to show that an open interval containing is contained in .
If is spanned by , then 7.14 implies that we can find a one-parameter subgroup such that fixes and , and acts nontrivially on . Since is open, we can find so that for . Since and are in , is transverse to both and , so we have , . Then as varies from to ,
gives an open interval in containing .
A similar argument using 7.15 shows that the claim also holds if is tangent to . ∎
We can now prove a slightly weaker version of 7.11, which we will then use to show the stronger version.
Lemma 7.17.
Let be proper domains in , with , and let be compact. There exists a constant such that for all ,
Proof.
Since is compact, it suffices to show that for fixed , the ratio
is bounded below by some as varies in .
Let , and let denote the projective line spanned by and . Choose so that
That is, if , , we have
where the distances are measured in any identification of with .
We can choose an identification of with so that either or . In either case, for any , we have
We know that , so lie in . Then 7.16 implies that there exists so that lies in the interval . See Figure 8.
Then, we have
This shows that for all . We still need to find some uniform so that for all . To see this, suppose for the sake of a contradiction that for a sequence , we have
(10) |
Since is compact, must converge to . Up to subsequence, the sequence of projective lines spanned by and converges to a line tangent to at .
For each , choose , so that
Let , . Then up to subsequence converges to , converges to , and , respectively converge to , .
Since is in , and are both transverse to —so in particular and (although a priori we could have ).
Since , 7.16 implies that there exist so that for some identification of with , we have
Note that this is possible even if , because then we can just identify both and with . Let , and let . Respectively, and converge to and .
This means that the cross-ratios converge to , and in particular are bounded away from both and for all .
7.3. Contracting paths are -divergent
Proposition 7.18.
Let be a -graph for a group , and let be a -compatible system of open sets of with each a proper domain.
If is a contracting -path, then the sequence
is -divergent with unique limit point , where .
Proof.
Consider the sequence of open sets
Up to subsequence, is a fixed open set . By assumption is a contracting path, so converges to a singleton . So, we apply 3.5. ∎
8. A weaker criterion for EGF representations
We have now developed enough tools to be able to prove our weaker characterization of EGF representations. We first prove a pair of lemmas.
Lemma 8.1.
Let be a relatively hyperbolic pair, let be a representation, and let be a symmetric parabolic subgroup. Suppose there exists
-
(1)
a -invariant closed set and a continuous equivariant surjective antipodal map , and
-
(2)
a relative quasigeodesic automaton and a -compatible system of open subsets of , such that the ’s cover and each is a proper domain intersecting nontrivially.
Then for every sequence which is unbounded in the coned-off Cayley graph , the sequence is -divergent, and every -limit point of lies in .
Proof.
We will show that every subsequence of has a -contracting subsequence, so take an arbitrary subsequence of . By 5.7, we may assume that for a bounded sequence , is the endpoint of a finite -path . Up to subsequence is a constant , independent of .
Let be the vertex path associated to . Up to subsequence is a fixed vertex , and is a fixed vertex . Let be an -neighborhood of , with chosen sufficiently small so that is still a proper domain.
The sequence must be unbounded, since the length of with respect to the coned-off Cayley graph metric is at most a fixed constant times . 7.12 then implies that the diameter of
with respect to the metric tends to zero, exponentially in . Since this sequence of sets lies in the compact set , up to subsequence it must converge to a singleton in . In fact must lie in , because is compact and is the limit of a sequence of points in the sequence of nonempty closed sets . Then, since converges to , 3.5 implies that is -divergent with unique -limit . ∎
Lemma 8.2.
Let be a relatively hyperbolic pair, let be a representation, let be a closed -invariant set, and let be a continuous equivariant surjective antipodal map.
Suppose that is a sequence converging to , such that is -divergent and every -limit point of lies in . If converges to , then for every compact set and every open containing , for large enough , we have .
Proof.
It suffices to show that every subsequence of has a further subsequence satisfying the desired property. So, we can freely extract subsequences throughout this proof.
We assume symmetric, so is also -divergent and has nonempty -limit set. Let be a pair of flags in the -limit sets of , respectively; by assumption we have . By 3.6, we have a subsequence so that converges to uniformly on compacts in .
Antipodality of implies that every compact subset of is contained in . Then, by equivariance and continuity of , we see that must converge to on compacts in . This uniquely characterizes the points as the limits of in . So, we see that converges uniformly to on every compact in , as required. ∎
We recall the statement of our weaker characterization of EGF representations here:
*
Proof.
To see the “only if” part, observe that if we know that is an EGF boundary extension, we can use the results of Section 6 to construct an automaton satisfying the hypotheses of 8.1, which immediately implies that the first condition holds. The second condition is immediate from the fact that extends convergence dynamics.
So, we focus on the “if” part. For each conical limit point , we let . Each contains by antipodality of . For each , we can replace with : this set is still open, and it again contains by antipodality.
Observe that if is a sequence limiting conically to , with converging to , then 8.2, together with part (b) of our hypotheses, implies that the map satisfies both conditions (1) and (2) given at the beginning of Section 6. So, by 6.1, we know that there is a relative quasigeodesic automaton satisfying the hypotheses of 8.1.
We now want to show that parts (a) and (b) of our hypotheses show that is an EGF boundary extension, so let be a sequence with . We fix an open set containing and a compact . Our goal is to show that for large enough , we have .
We consider two cases:
Case 1: is unbounded in the coned-off Cayley graph . By 8.1, is -divergent, and every -limit point of lies in . Then we are done by 8.2.
Case 2: is bounded in . We can write as an alternating product
where is bounded in , and lies in for a parabolic point . Without loss of generality, the are unbounded in as . Up to subsequence we can assume that and (independent of ). Since contains exactly one representative of each parabolic orbit, we can also assume that for any .
We claim that converges to , converges to , and for any compact and open containing , for large , we have .
Fix such a compact and open . We will prove the claim by inducting on . When , then , and for and fixed. The distance between and is bounded in any word metric on , so converges to in and converges to . We also know that , so eventually lies in a small neighborhood of by part (b) of our hypotheses. Then lies in any small neighborhood of .
When , we consider the sequence
Inductively we can assume that for large , and is a subset of an arbitrarily small neighborhood of . Then since , for large enough , is a compact subset of . So our hypotheses imply that for large ,
∎
The arguments above also imply the following characterization of EGF representations. This result is not needed anywhere else in the paper.
Proposition 8.3.
Let be a relatively hyperbolic pair, let be a representation, and let be a symmetric parabolic subgroup. Suppose that there exists a closed -invariant subset and a surjective equivariant antipodal map .
Then is EGF with boundary extension if and only if for every , there exists an open containing , such that:
-
(a)
For any sequence limiting conically to some point in , with , any open set containing , and any compact , we have for all sufficiently large .
-
(b)
For any parabolic point , any compact , and any open set containing , for all but finitely many , we have .
Proof.
The “only if” direction is immediate, so suppose we have a representation satisfying the hypotheses above. The results of Section 6 imply that there is a relative quasigeodesic automaton satisfying the hypotheses of 8.1. We then apply this lemma together with Section 4.3 to obtain the desired result. ∎
9. Relative stability
In this section we prove the main relative stability property for EGF representations (Section 1.4).
9.1. Deformations of EGF representations
In general, the set of EGF representations is not an open subset of . However, it is relatively open in a subspace of where we restrict the deformations of the peripheral subgroups appropriately. Roughly speaking, we want to consider subspaces where the dynamical behavior of the peripheral subgroups changes continuously under deformation. That is, if is a small deformation of a representation , where attracts points towards at a particular “speed,” then we want to attract points towards a small deformation of at a similar “speed.”
The precise condition is the following:
Definition 9.1.
Let be an EGF representation with boundary extension , and let contain .
We say that is peripherally stable at if for every , every open set containing , every compact set , and every cofinite set such that
there is an open set containing , such that for every , we have
We restate the main result of the paper below:
*
Remark 9.2.
In [Bow98], Bowditch explored the deformation spaces of geometrically finite groups , and gave an explicit discription of semialgebraic subspaces of in which small deformations of are still geometrically finite.
Bowditch’s deformation spaces are peripherally stable, so it seems desirable to find a general algebraic description of peripherally stable subspaces.
Even in , the question is subtle, however. Bowditch also gives examples of geometrically finite representations (for ) and deformations of in such that the restriction of to each cusp group in is discrete, faithful, and parabolic, but is not even discrete; further examples exist where the deformation is discrete, but not geometrically finite.
Example 9.3.
Let be a -dimensional Jordan block with eigenvalue 1 and eigenvector , and let be the block matrix .
Although is not quite an attracting fixed point of , it is still an “attracting subspace” in the sense that if is any compact subset of which does not intersect a fixed hyperplane of , then converges to . Via a ping-pong argument, one can use this “attracting” behavior to show that for some and some , the group generated by and is a discrete free group with free generators . The group is hyperbolic relative to the subgroups , , and the inclusion is EGF with respect to (the stabilizer of a line in a hyperplane in ).
Here, there are peripherally stable deformations of which change the Jordan block decomposition of . For instance, consider a continuous path given by , where and is a diagonalizable matrix in . For small values of , the group generated by and is still discrete and freely generated by and —since the “attracting” fixed points of deform continuously with , the same exact ping-pong setup works for all small . And indeed the path in determined by the path is a peripherally stable subspace.
On the other hand, consider the path , and let . In this case the corresponding subspace of is not peripherally stable: while the group generated by is still discrete, the attracting fixed points of do not deform continuously in . So, there is no way to use the ping-pong setup for to ensure that is a discrete group.
Example 9.4.
Here is a somewhat more interesting example of a non-peripherally stable deformation. Let be a finite-volume noncompact hyperbolic 3-manifold, with holonomy representation (so there is an identification ). Then is hyperbolic relative to the collection of conjugates of cusp groups (each of which is isomorphic to ), and the representation is geometrically finite (in particular, EGF).
In this case, for any sufficiently small nontrivial deformation of in the character variety , the restriction of to some cusp group either fails to be discrete or has infinite kernel. So is not peripherally stable, because any sufficiently small deformation of inside of a peripherally stable subspace must have discrete image and finite kernel on each . This is true despite the fact that arbitrarily small deformations of are holonomy representations of complete hyperbolic structures on Dehn fillings of (so in particular, they are discrete).
The main ingredient in the proof of Section 1.4 is the relative quasigeodesic automaton and the associated -compatible system of open sets we constructed in 6.1. The following proposition is immediate from the definition of peripheral stability:
Proposition 9.5.
Let be an EGF representation with boundary extension , and let be a subspace which is peripherally stable at .
If is a relative quasigeodesic automaton for , and is a -compatible system of open subsets of for , then there is an open subset containing such that for every , is also a -compatible system of open sets for .
Section 1.4 then follows from a kind of converse to 6.1: we will show that we can reconstruct a map extending the convergence dynamics of from the -compatible system .
9.2. An equivariant map on conical limit points
For the rest of this section, we let be a representation which is EGF with respect to a symmetric parabolic subgroup . We let be a boundary extension for , and assume that is peripherally stable at . We also let be a compact subset of , and let be an open subset containing . We again fix a finite subset , containing one point from every -orbit in .
Using 6.1, we can find a relative quasigeodesic automaton and -compatible system of open subsets of for . Using 6.14, we can ensure that for any , there is a -path limiting to (with vertex path ) so that is contained in .
For each , we also fix a compact set such that , and assume that the automaton has been constructed to satisfy 6.15.
Antipodality of the map implies that for each , each fiber is a closed subset of some affine chart in . So, we can also assume that is a proper domain for each vertex of . In fact, by way of the following lemma, we can assume even more:
Lemma 9.6.
Let be an EGF representation with boundary map .
For any , we can find a relative quasigeodesic automaton with -compatible system of open sets in as in 6.1, so that for any with , if and , then and are opposite.
Proof.
We choose so that if for , then the closed -neighborhoods of
are opposite. This is possible for a fixed pair since antipodality is an open condition, and , are opposite compact sets. Then we can pick a uniform for all pairs since the the subset is compact.
Consider -compatible systems of open subsets and for the action of on and , coming from 6.1. We can ensure that for each vertex , the diameter of is at most , and for some .
If satisfy , and , , we have
for all , .
Then, if and , we have
for , with . By our choice of , the closures of and are opposite. ∎
Using cocompactness of the action of on the space of distinct pairs in , we know that there exists some fixed such that for any distinct , we can find some such that . Then, in light of 9.6, we can make the following assumption:
Assumption 9.7.
For any satisfying , if and for vertices of , then and are opposite.
With our relative quasigeodesic automaton and compatible system of open sets fixed, we now choose an open subset so that for any , is also a -compatible system for the action of on . Our main goal for the rest of this section is to show that any is an EGF representation. So, we fix some .
Let denote the set of infinite -paths. 7.3 implies that every path in is contracting for the -action, so we have a map
where the path maps to the unique element of
Lemma 9.8.
The map induces an equivariant map
Proof.
We first need to see that is well-defined, i.e. that if is a conical limit point and , are -paths limiting to , then .
Let
We can use 5.11 to see that and lie within bounded Hausdorff distance of a geodesic in limiting to , so there is a fixed so that for infinitely many pairs ,
in the Cayley graph of . 7.18 implies that and are both -divergent sequences and each have a unique -limit point in , given by , , respectively. Then, Lemma 4.23 in [KLP17] implies that because for a bounded sequence , the -limit points of and must agree and therefore .
Next we observe that is equivariant. Fix a finite generating set for . It suffices to show that for all .
Let be a -path limiting to some , and consider the sequence
Again, 5.11 implies that lies bounded Hausdorff distance from a geodesic in , which must limit to . So if we fix a -path limiting to , the same argument as above shows that . ∎
It will turn out that is also both continuous and injective. However, we do not prove this directly.
9.3. Extending to parabolic points
We want to extend the map to the entire Bowditch boundary . To do so, we need to view as a map to the set of closed subsets of .
The first step is to define on the finite set . For any vertex in , we consider the set
Then, for each , we pick a parabolic vertex so that . We define to be the closure of the set of accumulation points of sequences of the form , for and distinct elements of . Part (3) of 6.1 guarantees that , and -compatiblity of the system for the -action on implies that . By construction, is invariant under the action of .
Next, given a parabolic point , we write for , and then define
Since is -invariant and is exactly the stabilizer of , this does not depend on the choice of coset representative in . Moreover is still -equivariant.
In addition, if is any parabolic vertex with parabolic point for , part (2) of 6.1 ensures that for any parabolic vertex with . So, is exactly the closure of the set of accumulation points of the form for sequences and . Then -compatibility implies that is a subset of .
Remark 9.9.
There is a natural topology on the space of closed subsets of , induced by the Hausdorff distance arising from some (any) choice of metric on . We emphasize that the map is not necessarily continuous with respect to this topology.
Ultimately we want to use to define a map extending the convergence dynamics of , so we will need to also define the sets for each . For now, we only define for : this will be the set
We can immediately observe:
Proposition 9.10.
is -invariant. Moreover, for any infinite sequence , any compact , and any open containing , for sufficiently large , lies in .
Proof.
-invariance follows directly from the definition.
Fix a compact and an open containing . is contained in finitely many sets for , so any accumulation point of for and lies in . In particular, for sufficiently large , lies in , and since is compact we can pick large enough so that for all . ∎
We next want to use to define an antipodal extension from a subset of to .
Lemma 9.11.
For any , if is a -path limiting to with corresponding vertex path , then and are both subsets of .
Proof.
If is a conical limit point, then this follows immediately from 5.11 and the definition of . On the other hand, if is a parabolic point, then , where is a parabolic vertex at the end of the vertex path . By part (3) of 6.1, we have and thus . By -equivariance of we have
so by -compatibility we have . On the other hand, we have constructed so that , so -equivariance of and -compatibility also show that . ∎
Lemma 9.12.
For any two distinct points in , the sets
are opposite (in particular disjoint).
Proof.
We know that for any distinct , we can find so that . So, since is -equivariant, we just need to show that if satisfy , then is opposite to .
9.4. The boundary set of the deformed representation
We define our candidate boundary set by
We then have an equivariant map
where if . 9.12 implies that is well-defined and antipodal. It is necessarily both surjective and -equivariant, and its fibers are either singletons or translates of the sets for .
It now remains to verify the properties of the candidate set and the map needed to show that is an EGF boundary extension.
Lemma 9.13.
For every vertex of , the intersection is nonempty.
Proof.
The construction in Section 6 ensures that every vertex of the automaton has at least one outgoing edge. In particular this means that for a given vertex , there is an infinite -path whose first vertex is . This -path limits to a conical limit point , and 9.11 implies that is a (nonempty) subset of both and . ∎
Lemma 9.14.
For any , we have .
Proof.
Lemma 9.15.
is compact.
Proof.
Fix a sequence , and let . Since is compact, up to subsequence . We want to see that a subsequence of converges to some . We consider two possibilities:
Case 1: is a parabolic point. We can write , where . Let be a parabolic vertex with , and consider the compact set , chosen so that . If for infinitely many , we are done, so assume otherwise, and choose so that .
We have assumed (using 6.15) that the automaton has been constructed so that there is always a -path limiting to whose first vertex is connected to by an edge . 9.11 implies that lies in , which is contained in by definition.
Then using 9.10, we know that up to subsequence,
converges to a compact subset of , which means that
subconverges to a point in .
Case 2: is a conical limit point. We want to show that any sequence in limits to , so fix any . Using 7.12, we can choose so that if is any -path limiting to , with corresponding vertex path , then the diameter of
is less than with respect to a metric on . We fix such a -path . Then, we use 5.15 to see that for sufficiently large , there is a -path limiting to with for . Thus the Hausdorff distance (with respect to ) between and is at most . Since and both lie in the compact set , this proves the claim.
∎
Lemma 9.16.
is continuous and proper.
Proof.
Since is compact, we just need to show continuity. Fix and a sequence approaching . We want to show that approaches .
Suppose otherwise. Since is compact, up to a subsequence approaches . Using the equivariance of , and cocompactness of the -action on distinct pairs in , we may assume that . For sufficiently large , we have as well. Then, as in the proof of 9.12, by 9.7 we know that for any vertices in such that contains and contains , the intersection is empty.
But by definition of , we have
for vertices in . This contradicts the fact that . ∎
9.5. Dynamics on the deformation
To complete the proof of Section 1.4, we just need to show:
Proposition 9.17.
The map extends the convergence group action of .
Proof.
We will apply Section 4.3. The preceding arguments show that the relative quasi-geodesic automaton , the map , and the -compatible system satisfy the hypotheses of 8.1. This immediately implies that the first condition of Section 4.3 is satisfied.
To see that the second condition is also satisfied, let be a parabolic point in , and write for , and then take . 9.10 says that for any , any compact , and any open containing , if is an infinite sequence in , then for sufficiently large . Then, since , the same is true for any parabolic point .
Remark 9.18.
The definition of the set and the map immediately imply that the fibers of the deformed boundary extension satisfy the conclusions of 4.7: the fiber over each conical limit point is a singleton, and the fiber over each parabolic point is the closure of the accumulation sets of -orbits in . So, we obtain 4.7 by taking to be the singleton , and following the proof of Section 1.4 (using for throughout).
Appendix A Contraction dynamics on flag manifolds
Let be a real vector space, and let be a sequence of elements of . It is sometimes possible to determine the global dynamical behavior of on by considering the action of on a small open subset of : if there is an open subset such that converges to a point in , then in fact there is a dense open subset (the complement of a hyperplane) on which converges to the same point, uniformly on compacts.
A similar statement holds for the action of on Grassmannians . These claims can be proved by considering the behavior of the singular value gaps of as .
In this appendix we give a general result along these lines, where we take sequences of group elements for a semisimple Lie group with no compact factor and trivial center, and consider the limiting behavior of on open subsets of some flag manifold , where is a parabolic subgroup.
*
We will prove 3.5 by reducing it to the case where and is the stabilizer of . In this situation, -divergence can be understood in terms of the behavior of the singular value gaps of the sequence :
Proposition A.1.
Suppose that , and let be the stabilizer of a line in . A sequence is -divergent if and only if
where is the th-largest singular value of .
For convenience, we give a proof of 3.5 in this special case.
Lemma A.2.
Let be a sequence in , and suppose that for a nonempty open subset , converges to a point in . Then, the singular value gap
tends to as .
Proof.
It suffices to show that any subsequence of has a subsequence which satisfies the property. Using the Cartan decomposition of , we can write
for and a diagonal matrix whose diagonal entries are . Up to subsequence and converge respectively to . For sufficiently large , is nonempty, so by replacing with we can assume that for all . Furthermore, if converges to a point , then converges to .
So, converges to a point, and since is a diagonal matrix, the gap between the moduli of its largest and second-largest eigenvalues must be unbounded. ∎
To prove the general case of 3.5, we take an irreducible representation coming from Theorem 7.4, so that maps to the stabilizer of a line in , maps to the stabilizer of a hyperplane in , and , are opposite if and only if , are transverse. As in section 7, this determines embeddings and by
The representation additionally has the property that for any sequence , the singular value gaps
are unbounded if and only if is -divergent (see [GGKW17], Lemma 3.7).
Proof of 3.5.
By [Zim18], Lemma 4.7, there exist flags so that lifts of give a basis of . Since converges to a point in , the set
converges to a single point in .
This means that we can fix lifts so that, up to a subsequence, takes the projective -simplex
to a point. This simplex is an open subset of . Now we can apply A.2 to see that the sequence is -divergent.
We now just need to check that is the unique -limit point of . Choose any subsequence of . Then any -contracting subsequence of this subsequence satisfies
uniformly on compacts for some and . But since is open and dense, it intersects nontrivially and thus . ∎
References
- [Bal14] Samuel A. Ballas. Deformations of noncompact projective manifolds. Algebr. Geom. Topol., 14(5):2595–2625, 2014.
- [Bal21] Samuel A. Ballas. Constructing convex projective 3-manifolds with generalized cusps. J. Lond. Math. Soc. (2), 103(4):1276–1313, 2021.
- [BCL20] Samuel A. Ballas, Daryl Cooper, and Arielle Leitner. Generalized cusps in real projective manifolds: classification. J. Topol., 13(4):1455–1496, 2020.
- [BDL15] Samuel Ballas, Jeffrey Danciger, and Gye-Seon Lee. Convex projective structures on non-hyperbolic three-manifolds. Geometry & Topology, 22, 08 2015.
- [Ben04] Yves Benoist. Convexes divisibles. I. In Algebraic groups and arithmetic, pages 339–374. Tata Inst. Fund. Res., Mumbai, 2004.
- [Ben06a] Yves Benoist. Convexes divisibles. IV. Structure du bord en dimension 3. Invent. Math., 164(2):249–278, 2006.
- [Ben06b] Yves Benoist. Convexes hyperboliques et quasiisométries. Geom. Dedicata, 122:109–134, 2006.
- [BH99] Martin R. Bridson and André Haefliger. Metric spaces of non-positive curvature, volume 319 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1999.
- [BM20] Samuel Ballas and Ludovic Marquis. Properly convex bending of hyperbolic manifolds. Groups Geom. Dyn., 14(2):653–688, 2020.
- [Bob19] Martin D. Bobb. Convex projective manifolds with a cusp of any non-diagonalizable type. J. Lond. Math. Soc. (2), 100(1):183–202, 2019.
- [Bog97] O. V. Bogopol’skiĭ. Infinite commensurable hyperbolic groups are bi-Lipschitz equivalent. Algebra i Logika, 36(3):259–272, 357, 1997.
- [Bow98] Brian H. Bowditch. Spaces of geometrically finite representations. Ann. Acad. Sci. Fenn. Math., 23(2):389–414, 1998.
- [Bow99] B. H. Bowditch. Convergence groups and configuration spaces. In Geometric group theory down under (Canberra, 1996), pages 23–54. de Gruyter, Berlin, 1999.
- [Bow12] B. H. Bowditch. Relatively hyperbolic groups. Internat. J. Algebra Comput., 22(3):1250016, 66, 2012.
- [BPS19] Jairo Bochi, Rafael Potrie, and Andrés Sambarino. Anosov representations and dominated splittings. J. Eur. Math. Soc. (JEMS), 21(11):3343–3414, 2019.
- [BV23] Pierre-Louis Blayac and Gabriele Viaggi. Divisible convex sets with properly embedded cones. arXiv e-prints, page arXiv:2302.07177, February 2023.
- [Cho10] Suhyoung Choi. The convex real projective orbifolds with radial or totally geodesic ends: The closedness and openness of deformations. arXiv e-prints, page arXiv:1011.1060, November 2010.
- [CLM20] Suhyoung Choi, Gye-Seon Lee, and Ludovic Marquis. Convex projective generalized Dehn filling. Ann. Sci. Éc. Norm. Supér. (4), 53(1):217–266, 2020.
- [CLM22] Suhyoung Choi, Gye-Seon Lee, and Ludovic Marquis. Deformation spaces of Coxeter truncation polytopes. J. Lond. Math. Soc. (2), 106(4):3822–3864, 2022.
- [CLT18] Daryl Cooper, Darren Long, and Stephan Tillmann. Deforming convex projective manifolds. Geom. Topol., 22(3):1349–1404, 2018.
- [CM14] Mickaël Crampon and Ludovic Marquis. Finitude géométrique en géométrie de Hilbert. Ann. Inst. Fourier (Grenoble), 64(6):2299–2377, 2014.
- [CZZ21] Richard Canary, Tengren Zhang, and Andrew Zimmer. Cusped Hitchin representations and Anosov representations of geometrically finite Fuchsian groups. arXiv e-prints, page arXiv:2103.06588, March 2021.
- [DGK17] Jeffrey Danciger, François Guéritaud, and Fanny Kassel. Convex cocompact actions in real projective geometry. arXiv e-prints, page arXiv:1704.08711, April 2017.
- [DGK18] Jeffrey Danciger, François Guéritaud, and Fanny Kassel. Convex cocompactness in pseudo-Riemannian hyperbolic spaces. Geom. Dedicata, 192:87–126, 2018.
- [DGK+21] Jeffrey Danciger, François Guéritaud, Fanny Kassel, Gye-Seon Lee, and Ludovic Marquis. Convex cocompactness for Coxeter groups. arXiv e-prints, page arXiv:2102.02757, February 2021.
- [DS05] Cornelia Druţu and Mark Sapir. Tree-graded spaces and asymptotic cones of groups. Topology, 44(5):959–1058, 2005. With an appendix by Denis Osin and Mark Sapir.
- [Ebe96] Patrick B. Eberlein. Geometry of nonpositively curved manifolds. Chicago Lectures in Mathematics. University of Chicago Press, Chicago, IL, 1996.
- [Ger12] Victor Gerasimov. Floyd maps for relatively hyperbolic groups. Geom. Funct. Anal., 22(5):1361–1399, 2012.
- [GGKW17] François Guéritaud, Olivier Guichard, Fanny Kassel, and Anna Wienhard. Anosov representations and proper actions. Geom. Topol., 21(1):485–584, 2017.
- [GM08] Daniel Groves and Jason Fox Manning. Dehn filling in relatively hyperbolic groups. Israel J. Math., 168:317–429, 2008.
- [GP13] Victor Gerasimov and Leonid Potyagailo. Quasi-isometric maps and Floyd boundaries of relatively hyperbolic groups. J. Eur. Math. Soc. (JEMS), 15(6):2115–2137, 2013.
- [GW12] Olivier Guichard and Anna Wienhard. Anosov representations: domains of discontinuity and applications. Invent. Math., 190(2):357–438, 2012.
- [Hel01] Sigurdur Helgason. Differential geometry, Lie groups, and symmetric spaces, volume 34 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2001. Corrected reprint of the 1978 original.
- [IZ22] Mitul Islam and Andrew Zimmer. The structure of relatively hyperbolic groups in convex real projective geometry. arXiv e-prints, page arXiv:2203.16596, March 2022.
- [Kap07] Michael Kapovich. Convex projective structures on Gromov-Thurston manifolds. Geom. Topol., 11:1777–1830, 2007.
- [KKL19] Michael Kapovich, Sungwoon Kim, and Jaejeong Lee. Structural stability of meandering-hyperbolic group actions. arXiv e-prints, page arXiv:1904.06921, April 2019.
- [KL18] Michael Kapovich and Bernhard Leeb. Relativizing characterizations of Anosov subgroups, I. arXiv e-prints, page arXiv:1807.00160, June 2018.
- [KLP14] Michael Kapovich, Bernhard Leeb, and Joan Porti. Morse actions of discrete groups on symmetric space. arXiv e-prints, page arXiv:1403.7671, March 2014.
- [KLP17] Michael Kapovich, Bernhard Leeb, and Joan Porti. Anosov subgroups: dynamical and geometric characterizations. Eur. J. Math., 3(4):808–898, 2017.
- [Kna02] Anthony W. Knapp. Lie groups beyond an introduction, volume 140 of Progress in Mathematics. Birkhäuser Boston, Inc., Boston, MA, second edition, 2002.
- [KP22] Fanny Kassel and Rafael Potrie. Eigenvalue gaps for hyperbolic groups and semigroups. J. Mod. Dyn., 18:161–, 2022.
- [Lab06] François Labourie. Anosov flows, surface groups and curves in projective space. Invent. Math., 165(1):51–114, 2006.
- [LLS21] Gye-Seon Lee, Jaejeong Lee, and Florian Stecker. Anosov triangle reflection groups in SL(3,R). arXiv e-prints, page arXiv:2106.11349, June 2021.
- [Max21] J. Maxwell Riestenberg. A quantified local-to-global principle for Morse quasigeodesics. arXiv e-prints, page arXiv:2101.07162, January 2021.
- [MMW22] Kathryn Mann, Jason Fox Manning, and Theodore Weisman. Stability of hyperbolic groups acting on their boundaries. Preprint, arXiv:2206.14914, 2022.
- [Sul85] Dennis Sullivan. Quasiconformal homeomorphisms and dynamics. II. Structural stability implies hyperbolicity for Kleinian groups. Acta Math., 155(3-4):243–260, 1985.
- [Tra13] Hung Cong Tran. Relations between various boundaries of relatively hyperbolic groups. Internat. J. Algebra Comput., 23(7):1551–1572, 2013.
- [Tso20] Konstantinos Tsouvalas. Anosov representations, strongly convex cocompact groups and weak eigenvalue gaps. arXiv e-prints, page arXiv:2008.04462, August 2020.
- [Tuk94] Pekka Tukia. Convergence groups and Gromov’s metric hyperbolic spaces. New Zealand J. Math., 23(2):157–187, 1994.
- [Tuk98] Pekka Tukia. Conical limit points and uniform convergence groups. J. Reine Angew. Math., 501:71–98, 1998.
- [Wan23a] Tianqi Wang. Anosov representations over closed subflows. Trans. Amer. Math. Soc., 376(9):6177–6214, 2023.
- [Wan23b] Tianqi Wang. Notions of Anosov representation of relatively hyperbolic groups. arXiv e-prints, page arXiv:2309.15636, September 2023.
- [Wei23a] Theodore Weisman. Dynamical properties of convex cocompact actions in projective space. Journal of Topology, 16(3):990–1047, 2023.
- [Wei23b] Theodore Weisman. Examples of extended geometrically finite representations. 2023.
- [Wol20] Adva Wolf. Convex Projective Geometrically Finite Structures. ProQuest LLC, Ann Arbor, MI, 2020. Thesis (Ph.D.)–Stanford University.
- [Yam04] Asli Yaman. A topological characterisation of relatively hyperbolic groups. J. Reine Angew. Math., 566:41–89, 2004.
- [Zhu21] Feng Zhu. Relatively dominated representations. Ann. Inst. Fourier (Grenoble), 71(5):2169–2235, 2021.
- [Zim18] Andrew M. Zimmer. Proper quasi-homogeneous domains in flag manifolds and geometric structures. Ann. Inst. Fourier (Grenoble), 68(6):2635–2662, 2018.
- [Zim21] Andrew Zimmer. Projective Anosov representations, convex cocompact actions, and rigidity. J. Differential Geom., 119(3):513–586, 2021.
- [ZZ22] Feng Zhu and Andrew Zimmer. Relatively Anosov representations via flows I: theory. arXiv e-prints, page arXiv:2207.14737, July 2022.