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

An extended definition of Anosov representation for relatively hyperbolic groups

Theodore Weisman Department of Mathematics, University of Michigan, Ann Arbor MI 48109, USA [email protected]
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 Γ\Gamma is a hyperbolic group, then the Anosov representations from Γ\Gamma into a semisimple Lie group GG form an open subset of the representation variety Hom(Γ,G)\operatorname{Hom}(\Gamma,G).

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 Γ\Gamma is a hyperbolic group, GG is a semisimple Lie group, and PGP\subset G is a symmetric parabolic subgroup, a representation ρ:ΓG\rho:\Gamma\to G is PP-Anosov if there is a ρ\rho-equivariant embedding ξ:ΓG/P\xi:\partial\Gamma\to G/P of the Gromov boundary Γ\partial\Gamma of Γ\Gamma satisfying certain dynamical properties.

Existing notions of relative Anosov representations (see e.g. [KL18], [Zhu21], [ZZ22]) replace ξ\xi with an embedding of the Bowditch boundary of a relatively hyperbolic group. This condition forces the limit set in G/PG/P 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 G/PG/P, 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 ρ:ΓG\rho:\Gamma\to G is relatively stable: any small deformation of ρ\rho in Hom(Γ,G)\operatorname{Hom}(\Gamma,G) 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 Γ\Gamma is a relatively hyperbolic group, relative to a collection \mathcal{H} of peripheral subgroups, then Γ\Gamma acts as a convergence group on the Bowditch boundary (Γ,)\partial(\Gamma,\mathcal{H}). We recall the definition here.

Definition 1.1.

Let Γ\Gamma act on a topological space MM. The group Γ\Gamma is said to act as a convergence group if for every infinite sequence of distinct elements γnΓ\gamma_{n}\in\Gamma, there exist points a,bMa,b\in M and a subsequence γmΓ\gamma_{m}\in\Gamma such that γm\gamma_{m} converges uniformly on compacts in M{a}M-\{a\} to the constant map bb.

When γn\gamma_{n} is a sequence of distinct elements in a relatively hyperbolic group Γ\Gamma, then γn\gamma_{n} converges to bb uniformly on compacts in (Γ,){a}\partial(\Gamma,\mathcal{H})-\{a\} if and only if γn\gamma_{n} converges to bb in the compactification Γ¯=Γ(Γ,)\overline{\Gamma}=\Gamma\sqcup\partial(\Gamma,\mathcal{H}) and the inverse sequence γn1\gamma_{n}^{-1} converges to aa.

Recall that if a group Γ\Gamma acts by homeomorphisms on a Hausdorff space XX, the pair (Γ,X)(\Gamma,X) is called a topological dynamical system. We say that an extension of (Γ,X)(\Gamma,X) is a topological dynamical system (Γ,Y)(\Gamma,Y) together with a Γ\Gamma-equivariant surjective map ϕ:YX\phi:Y\to X.

In this paper, when (Γ,)(\Gamma,\mathcal{H}) is a relatively hyperbolic pair (i.e. Γ\Gamma is hyperbolic relative to a collection \mathcal{H} of peripheral subgroups), we will consider embedded extensions of the topological dynamical system (Γ,(Γ,))(\Gamma,\partial(\Gamma,\mathcal{H})). We want these embedded extensions to respect the convergence group action of (Γ,(Γ,))(\Gamma,\partial(\Gamma,\mathcal{H})) in some sense, so we introduce the following definition: {restatable}definitiondynamicsPreserving Let (Γ,)(\Gamma,\mathcal{H}) be a relatively hyperbolic pair, with Γ\Gamma acting on a connected compact metrizable space MM by homeomorphisms. Let ΛM\Lambda\subset M be a closed Γ\Gamma-invariant set.

We say that a continuous equivariant surjective map ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}) extends the convergence action of Γ\Gamma if for each z(Γ,)z\in\partial(\Gamma,\mathcal{H}), there exists an open set CzMC_{z}\subset M containing Λϕ1(z)\Lambda-\phi^{-1}(z), satisfying the following:

If γn\gamma_{n} is a sequence in Γ\Gamma with γn±1z±\gamma_{n}^{\pm 1}\to z_{\pm} for z±(Γ,)z_{\pm}\in\partial(\Gamma,\mathcal{H}), then for any compact set KCzK\subset C_{z_{-}} and any open set UU containing ϕ1(z+)\phi^{-1}(z_{+}), for sufficiently large nn, γnK\gamma_{n}\cdot K lies in UU.

Now let GG denote a semisimple Lie group with no compact factor. The central definition of the paper is the following: {restatable}definitionEGF Let Γ\Gamma be a relatively hyperbolic group, let ρ:ΓG\rho:\Gamma\to G be a representation, and let PGP\subset G be a symmetric parabolic subgroup. We say that ρ\rho is extended geometrically finite (EGF) with respect to PP if there exists a closed ρ(Γ)\rho(\Gamma)-invariant set ΛG/P\Lambda\subset G/P and a continuous ρ\rho-equivariant surjective antipodal map ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}) extending the convergence action of Γ\Gamma.

The map ϕ\phi is called a boundary extension of the representation ρ\rho, and the closed invariant set Λ\Lambda 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 ρ\rho is an EGF representation, then certain small relative deformations of ρ\rho must also be EGF.

To state the theorem, we define a notion of a peripherally stable subspace of Hom(Γ,G)\operatorname{Hom}(\Gamma,G). The precise definition is given in Section 9, but roughly speaking, a subspace 𝒲Hom(Γ,G)\mathcal{W}\subseteq\operatorname{Hom}(\Gamma,G) is peripherally stable if the large-scale dynamical behavior of the peripheral subgroups of Γ\Gamma is in some sense preserved by small deformations inside of 𝒲\mathcal{W}. 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 ρ:ΓG\rho:\Gamma\to G be EGF with respect to PP, let ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}) be a boundary extension, and let 𝒲Hom(Γ,G)\mathcal{W}\subseteq\operatorname{Hom}(\Gamma,G) be peripherally stable at (ρ,ϕ)(\rho,\phi). For any compact subset ZZ of (Γ,)\partial(\Gamma,\mathcal{H}) and any open set VG/PV\subset G/P containing ϕ1(Z)\phi^{-1}(Z), there is an open subset 𝒲𝒲\mathcal{W}^{\prime}\subset\mathcal{W} containing ρ\rho such that each ρ𝒲\rho^{\prime}\in\mathcal{W}^{\prime} is EGF with respect to PP, and has an EGF boundary extension ϕ\phi^{\prime} satisfying ϕ1(Z)V\phi^{\prime-1}(Z)\subset V.

Remark 1.2.

When ρ:ΓG\rho:\Gamma\to G is a PP-Anosov representation of a hyperbolic group Γ\Gamma, then the associated boundary embedding ΓG/P\partial\Gamma\to G/P also varies continuously with ρ\rho in the compact-open topology on maps ΓG/P\partial\Gamma\to G/P. 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 ρ:ΓG\rho:\Gamma\to G 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 (Γ,)(\Gamma,\mathcal{H}) be a relatively hyperbolic pair, and let ρ:ΓG\rho:\Gamma\to G be a representation. The space of cusp-preserving representations Homcp(Γ,G,,ρ)\operatorname{Hom}_{\mathrm{cp}}(\Gamma,G,\mathcal{H},\rho) is the set of representations ρ:ΓG\rho^{\prime}:\Gamma\to G such that for each peripheral subgroup HH\in\mathcal{H}, we have ρ|H=gρ|Hg1\rho^{\prime}|_{H}=g\cdot\rho|_{H}\cdot g^{-1} for some gGg\in G (which may depend on HH).

Corollary 1.4.

Let ρ:ΓG\rho:\Gamma\to G be an EGF representation. Then there is a neighborhood of ρ\rho in Homcp(Γ,G,,ρ)\operatorname{Hom}_{\mathrm{cp}}(\Gamma,G,\mathcal{H},\rho) consisting of EGF representations.

1.4 gives a very restrictive example of a peripherally stable subspace of Hom(Γ,G)\operatorname{Hom}(\Gamma,G). 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 ρ:ΓPGL(d,)\rho:\Gamma\to\operatorname{PGL}(d,\mathbb{R}) which change the Jordan block decomposition of elements in the peripheral subgroups. For instance, one can deform an EGF representation in PGL(d,)\operatorname{PGL}(d,\mathbb{R}) 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 ΓPO(d,1)\Gamma\subset\operatorname{PO}(d,1) is convex cocompact if and only if the action of Γ\Gamma on its limit set Λ\Lambda in d\partial\mathbb{H}^{d} satisfies an expansion property. Then, he used this expansion property to give a symbolic coding for infinite quasigeodesic rays in Γ\Gamma. This coding gives a way to see that the correspondence between geodesic rays in Γ\Gamma and points in d\partial\mathbb{H}^{d} 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 G/PG/P 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 (Γ,)\partial(\Gamma,\mathcal{H}) of a relatively hyperbolic group Γ\Gamma, in a way which is compatible with the “topological” expansion of the Γ\Gamma-action on both (Γ,)\partial(\Gamma,\mathcal{H}) and G/PG/P. 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 Γ\Gamma, so that paths in the graph are in rough correspondence with quasi-geodesics in the coned-off Cayley graph for the pair (Γ,)(\Gamma,\mathcal{H}). Each vertex vv of the graph is assigned an open subset WvW_{v} of some compact metrizable space MM on which Γ\Gamma acts as an extended convergence group. Each edge uvu\to v in the graph corresponds to a set of inclusions of the form

(1) αNε(Wv)¯Wu,\alpha\cdot\overline{N_{\varepsilon}(W_{v})}\subset W_{u},

where α\alpha is an element in the subset of Γ\Gamma labeling the edge uvu\to v, and Nε(Wv)N_{\varepsilon}(W_{v}) is a small neighborhood of WvW_{v} with respect to some fixed metric on MM. 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 C0C^{0} topology on actions of Γ\Gamma on MM, so when the set of elements labeling an edge uvu\to v is finite, the corresponding set of inclusions are stable under small perturbations of the Γ\Gamma-action. If the label of uvu\to v 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 A,BA,B are arbitrary finitely generated groups, then the group ABA*B is hyperbolic relative to the collection of conjugates of AA and BB. 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 ABA*B 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 G/PG/P, 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 G/PG/P, 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 PGL(d,)\operatorname{PGL}(d,\mathbb{R}). 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 ΓPGL(d,)\Gamma\subset\operatorname{PGL}(d,\mathbb{R}) is relatively hyperbolic and projectively convex cocompact in the sense of [DGK17], then the inclusion ΓPGL(d,)\Gamma\hookrightarrow\operatorname{PGL}(d,\mathbb{R}) is EGF with respect to the parabolic subgroup stabilizing a flag of type (1,d1)(1,d-1) in d\mathbb{R}^{d}. If Γ\Gamma 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 PO(d,1)\operatorname{PO}(d,1) into PGL(d+1,)\operatorname{PGL}(d+1,\mathbb{R}) 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 MM (with strictly convex boundary) which is a union of a compact piece and finitely many ends homeomorphic to N×[0,)N\times[0,\infty), where NN 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.

We do not (yet) have a general result asserting that all strictly convex compact projective manifolds with generalized cusps have EGF holonomy, but there are indications that this should be true; see for example [Cho10], [Wol20] and the general setup in [IZ22], [BV23].

After a version of this paper originally appeared as a preprint, Blayac-Viaggi [BV23] also produced still more general examples of convex projective nn-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 N×S1×[0,)N\times S^{1}\times[0,\infty), where NN is a closed hyperbolic manifold of dimension n2n-2. 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 ρ:ΓPGL(V)\rho:\Gamma\to\operatorname{PGL}(V) with the symmetric representation τk:PGL(V)PGL(SymkV)\tau_{k}:\operatorname{PGL}(V)\to\operatorname{PGL}(\operatorname{Sym}^{k}V). We show that, assuming the peripheral subgroups in Γ\Gamma are all virtually abelian, then the composition τkρ\tau_{k}\circ\rho 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 Hom(Γ,PGL(SymkV))\operatorname{Hom}(\Gamma,\operatorname{PGL}(\operatorname{Sym}^{k}V)) is peripherally stable about τkρ\tau_{k}\circ\rho. 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:

{restatable}

theoremrelAsympBdryDynamics Let (Γ,)(\Gamma,\mathcal{H}) be a relatively hyperbolic pair, and let PGP\subset G be a symmetric parabolic subgroup. A representation ρ:ΓG\rho:\Gamma\to G is relatively PP-Anosov (in the sense given above) if and only if ρ\rho is EGF with respect to PP, and has an injective boundary extension ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}).

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 \mathcal{H} is trivial (meaning that Γ\Gamma is a hyperbolic group and (Γ,)\partial(\Gamma,\mathcal{H}) is identified with the Gromov boundary Γ\partial\Gamma of Γ\Gamma), 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 ϕ\phi 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 ϕ\phi 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 (Γ,)(\Gamma,\mathcal{H}) be a relatively hyperbolic pair, let ρ:ΓG\rho:\Gamma\to G be a relative PP-Anosov representation, and let 𝒲Hom(Γ,G)\mathcal{W}\subset\operatorname{Hom}(\Gamma,G) be a peripherally stable subspace, such that for each HH\in\mathcal{H} and each ρ𝒲\rho^{\prime}\in\mathcal{W}, the restriction ρ|H\rho^{\prime}|_{H} is PP-divergent with PP-limit set a singleton. Then an open neighborhood of ρ\rho in 𝒲\mathcal{W} consists of relative PP-Anosov representations of Γ\Gamma.

If we restrict the allowable peripheral deformations to conjugacies, this result reduces to:

Corollary 1.9.

Let (Γ,)(\Gamma,\mathcal{H}) be a relatively hyperbolic pair, and let ρ:ΓG\rho:\Gamma\to G be a relative PP-Anosov representation. There is an open neighborhood of ρ\rho in Homcp(Γ,G,,ρ)\operatorname{Hom}_{\mathrm{cp}}(\Gamma,G,\mathcal{H},\rho) consisting of relative PP-Anosov representations.

In the special case where Γ\Gamma 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 ρ\rho with PP-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 Γ\Gamma is a relatively hyperbolic group, the Bowditch boundary of Γ\Gamma (and thus, the definition of an EGF representation) depends on the choice of peripheral structure \mathcal{H} for Γ\Gamma. In general, there might be more than one possible choice: for instance, if a group Γ\Gamma is hyperbolic relative to a collection \mathcal{H} of hyperbolic subgroups, then Γ\Gamma 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 (Γ,)(\Gamma,\mathcal{H}) be a relatively hyperbolic pair, and suppose that each HH\in\mathcal{H} is hyperbolic. If ρ:ΓG\rho:\Gamma\to G is an EGF representation with respect to PP for the peripheral structure \mathcal{H}, and ρ\rho restricts to a PP-Anosov representation on each HH\in\mathcal{H}, then ρ\rho is a PP-Anosov representation of Γ\Gamma.

A potential application of Theorem 1.11 is the construction of new examples of Anosov representations: one could start with an EGF representation ρ:ΓG\rho:\Gamma\to G which is not Anosov, and then attempt to find a peripherally stable deformation of ρ\rho which restricts to an Anosov representation on peripheral subgroups. Section 1.4 and Theorem 1.11 would then imply that the original representation ρ\rho 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 ρ:Γp,q,rSL(3,)\rho:\Gamma_{p,q,r}\to\operatorname{SL}(3,\mathbb{R}), where Γp,q,r\Gamma_{p,q,r} 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 Γp,q,r\partial\Gamma_{p,q,r} into the space of full flags in 3\mathbb{R}^{3}, 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 ΓPO(d,1)\Gamma\subset\operatorname{PO}(d,1) 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 Λ(Γ)\Lambda(\Gamma). 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 Γ\Gamma 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 G/PG/P, which we use to relate the results of the previous sections to relatively hyperbolic group actions on G/PG/P. 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 XX is a metric space, AA is a subset of XX, and r0r\geq 0, we let NX(A,r)N_{X}(A,r) denote the open rr-neighborhood in XX about AA. For a point xXx\in X, we let BX(x,r)B_{X}(x,r) denote the open rr-ball about xx.

When the metric space XX is implied from context, we will often just write N(A,r)N(A,r) or B(x,r)B(x,r).

2.1. Geometrically finite actions

Recall that a finitely generated group Γ\Gamma is hyperbolic (or word-hyperbolic or δ\delta-hyperbolic or Gromov-hyperbolic) if and only if it acts properly discontinuously and cocompactly on a δ\delta-hyperbolic proper geodesic metric space YY.

A relatively hyperbolic group is also a group with an action by isometries on a δ\delta-hyperbolic proper geodesic metric space YY, 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 YY has a Γ\Gamma-invariant decomposition into a thick part YthY_{\mathrm{th}} and a countable collection \mathcal{B} of horoballs. For a horoball BB, we let ctr(B)\mathrm{ctr}(B) denote the center of BB in Y\partial Y, and we let Γp\Gamma_{p} denote the stabilizer of any pYp\in\partial Y.

Definition 2.2.

Let Γ\Gamma be a finitely generated group acting on a hyperbolic metric space YY, and let \mathcal{B} be a countable collection of horoballs in YY, invariant under the action of Γ\Gamma on YY. If:

  1. (1)

    The action of Γ\Gamma on the closure of Yth=YBBY_{\mathrm{th}}=Y-\bigcup_{B\in\mathcal{B}}B is cocompact, and

  2. (2)

    for each BB\in\mathcal{B}, the stabilizer of ctr(B)\mathrm{ctr}(B) in Γ\Gamma is finitely generated and infinite,

then we say that Γ\Gamma is a relatively hyperbolic group, relative to the collection ={StabΓ(p):p=ctr(B) for B}\mathcal{H}=\{\operatorname{Stab}_{\Gamma}(p):p=\mathrm{ctr}(B)\textrm{ for }B\in\mathcal{B}\}.

Definition 2.3.

Let Γ\Gamma be a relatively hyperbolic group, relative to a collection of subgroups \mathcal{H}.

  • The centers of the horoballs in \mathcal{B} are called parabolic points for the Γ\Gamma-action on Y\partial Y. The set of parabolic points in Y\partial Y is denoted parY\partial_{\mathrm{par}}Y.

  • The parabolic point stablizers ={StabΓ(p):pparY}\mathcal{H}=\{\operatorname{Stab}_{\Gamma}(p):p\in\partial_{\mathrm{par}}Y\} are called peripheral subgroups. We often write Γp\Gamma_{p} for StabΓ(p)\operatorname{Stab}_{\Gamma}(p).

A group Γ\Gamma might be hyperbolic relative to different collections \mathcal{H}, \mathcal{H}^{\prime} of peripheral subgroups. The collection \mathcal{H} of peripheral subgroups is sometimes called a peripheral structure for Γ\Gamma.

Definition 2.4.

Let Γ\Gamma be a finitely generated group, and let \mathcal{H} be a collection of subgroups. We say that (Γ,)(\Gamma,\mathcal{H}) is a relatively hyperbolic pair if Γ\Gamma is hyperbolic relative to \mathcal{H}.

2.2. The Bowditch boundary

Definition 2.5.

Let (Γ,)(\Gamma,\mathcal{H}) be a relatively hyperbolic pair, so that \mathcal{H} is the set of stabilizers of parabolic points for an action of Γ\Gamma on a metric space YY as in Definition 2.2. We say that YY is a Gromov model for the pair (Γ,)(\Gamma,\mathcal{H}).

In general there is not a unique choice of Gromov model for a given relatively hyperbolic pair (Γ,)(\Gamma,\mathcal{H}), 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 YY, YY^{\prime} for (Γ,)(\Gamma,\mathcal{H}), there is always a Γ\Gamma-equivariant homeomorphism YY\partial Y\to\partial Y^{\prime} [Bow12]. The Γ\Gamma-space Y\partial Y is the Bowditch boundary of (Γ,)(\Gamma,\mathcal{H}). We will denote it by (Γ,)\partial(\Gamma,\mathcal{H}), or sometimes just Γ\partial\Gamma when the collection of peripheral subgroups is understood from context. Since a Gromov model YY is a proper hyperbolic metric space, (Γ,)\partial(\Gamma,\mathcal{H}) is always compact and metrizable.

Definition 2.6.

We say a relatively hyperbolic pair (Γ,)(\Gamma,\mathcal{H}) is elementary if Γ\Gamma is finite or virtually cyclic, or if ={Γ}\mathcal{H}=\{\Gamma\}.

Whenever (Γ,)(\Gamma,\mathcal{H}) is nonelementary, its Bowditch boundary contains at least three points. The convergence properties of the action of Γ\Gamma on (Γ,)\partial(\Gamma,\mathcal{H}) (see below) imply that in this case, (Γ,)\partial(\Gamma,\mathcal{H}) is perfect (i.e. contains no isolated points).

2.2.1. Cocompactness on pairs

Let YY be a Gromov model for a relatively hyperbolic pair (Γ,)(\Gamma,\mathcal{H}). Since YY is hyperbolic, proper, and geodesic, for any compact subset KYK\subset Y, the space of bi-infinite geodesics passing through KK is compact.

Given any distinct pair of points u,vYu,v\in\partial Y, there is a bi-infinite geodesic cc in YY joining uu to vv. Since a horoball in a hyperbolic metric space has just one point in its ideal boundary, this geodesic must pass through the thick part YthY_{\mathrm{th}} of YY, so up to the action of Γ\Gamma it passes through a fixed compact subset KYthK\subset Y_{\mathrm{th}}.

This implies:

Proposition 2.7.

The action of Γ\Gamma on the space of distinct pairs in (Γ,)\partial(\Gamma,\mathcal{H}) is cocompact.

2.3. Convergence group actions

If a group Γ\Gamma acts on a proper geodesic hyperbolic metric space YY, we can characterize the geometrical finiteness of the action entirely in terms of the topological dynamics of the action on Y\partial Y. 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 Γ\Gamma act as a convergence group (see Definition 1.1) on a topological space ZZ.

  1. (1)

    A point zZz\in Z is a conical limit point if there exists a sequence γnΓ\gamma_{n}\in\Gamma and distinct points a,bZa,b\in Z such that γnza\gamma_{n}z\to a and γnyb\gamma_{n}y\to b for any yzy\neq z.

  2. (2)

    An infinite subgroup HH is a parabolic subgroup if it fixes a point pZp\in Z, and every infinite-order element of HH fixes exactly one point in ZZ.

  3. (3)

    A point pZp\in Z is a parabolic point if it is the fixed point of a parabolic subgroup.

  4. (4)

    A parabolic point pp is bounded if its stabilizer Γp\Gamma_{p} acts cocompactly on Z{p}Z-\{p\}.

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 YY be a hyperbolic metric space, and let zYz\in\partial Y. We say that a sequence ynYy_{n}\in Y limits conically to zz if there is a geodesic ray c:+Yc:\mathbb{R}^{+}\to Y limiting to zz and a constant D>0D>0 such that

dY(yn,c(tn))<Dd_{Y}(y_{n},c(t_{n}))<D

for some sequence tnt_{n}\to\infty.

A bounded neighborhood of a geodesic in a hyperbolic metric space looks like a “cone,” hence “conical limit.”

Proposition 2.10 ([Tuk94], [Tuk98]).

Let Γ\Gamma be a group acting properly discontinuously by isometries on a proper geodesic hyperbolic metric space YY, and fix a basepoint y0Yy_{0}\in Y.

Then Γ\Gamma acts on Y\partial Y as a convergence group. Moreover, a point zYz\in\partial Y is a conical limit point (in the dynamical sense of Definition 2.8) if and only if there is a sequence γny0\gamma_{n}\cdot y_{0} limiting conically to zz (in the geometric sense of Definition 2.9). In this case, there are distinct points a,bYa,b\in\partial Y such that γn1za\gamma_{n}^{-1}\cdot z\to a and γn1zb\gamma_{n}^{-1}z^{\prime}\to b for any zzz^{\prime}\neq z in Y\partial Y.

If γny0\gamma_{n}\cdot y_{0} limits conically to a point zYz\in\partial Y for some (hence any) basepoint y0Yy_{0}\in Y, we just say that γn\gamma_{n} limits conically to zz.

Theorem 2.11 ([Bow12]).

Let Γ\Gamma be a group acting by isometries on a hyperbolic metric space YY. Then Γ\Gamma is a relatively hyperbolic group, acting on YY as in Definition 2.2, if and only if:

  1. (1)

    The induced action of Γ\Gamma on Y\partial Y is a convergence group action.

  2. (2)

    Every point zYz\in\partial Y is either a conical limit point or a bounded parabolic point.

Whenever a group Γ\Gamma acts as a convergence group on a perfect compact metrizable space ZZ, every point in ZZ is either a conical limit point or a bounded parabolic point, and the stabilizer of each parabolic point is finitely generated, we say the Γ\Gamma-action on ZZ 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 YY whose boundary is equivariantly homeomorphic to ZZ. We can then identify the space ZZ with the Bowditch boundary (Γ,)\partial(\Gamma,\mathcal{H}). The set of parabolic points in ZZ coincides exactly with the set of fixed points of peripheral subgroups.

Definition 2.12.

Let (Γ,)(\Gamma,\mathcal{H}) be a relatively hyperbolic pair. We write

(Γ,)=con(Γ,)par(Γ,),\partial(\Gamma,\mathcal{H})=\partial_{\mathrm{con}}(\Gamma,\mathcal{H})\sqcup\partial_{\mathrm{par}}(\Gamma,\mathcal{H}),

where con(Γ,)\partial_{\mathrm{con}}(\Gamma,\mathcal{H}) and par(Γ,)\partial_{\mathrm{par}}(\Gamma,\mathcal{H}) are respectively the conical limit points and parabolic points in (Γ,)\partial(\Gamma,\mathcal{H}).

2.3.1. Compactification of Γ\Gamma and divergent sequences

When (Γ,)(\Gamma,\mathcal{H}) is a relatively hyperbolic pair, there is a natural topology on the set

Γ¯=Γ(Γ,)\overline{\Gamma}=\Gamma\sqcup\partial(\Gamma,\mathcal{H})

making it into a compactification of Γ\Gamma (i.e. Γ¯\overline{\Gamma} is compact, (Γ,)\partial(\Gamma,\mathcal{H}) and Γ\Gamma are both embedded in Γ¯\overline{\Gamma}, and Γ\Gamma is an open dense subset of Γ¯\overline{\Gamma}). Specifically, we view Γ\Gamma as a subset of (any) Gromov model YY, via an orbit map γγy0\gamma\mapsto\gamma\cdot y_{0} for some basepoint y0Yy_{0}\in Y. Since Γ\Gamma acts properly on YY, this is a proper embedding, so if we compactify YY by adjoining its visual boundary (Γ,)\partial(\Gamma,\mathcal{H}), we compactify Γ\Gamma as well; this does not depend on the choice of basepoint y0y_{0} or even the choice of space YY.

Definition 2.13.

A sequence γnΓ\gamma_{n}\in\Gamma is divergent if it leaves every bounded subset of Γ\Gamma (equivalently, if a subsequence of it consists of pairwise distinct elements).

Up to subsequence, a divergent sequence γnΓ\gamma_{n}\in\Gamma converges to a point z(Γ,)z\in\partial(\Gamma,\mathcal{H}). When (Γ,)(\Gamma,\mathcal{H}) is non-elementary, the point zz is determined solely by the action of Γ\Gamma on (Γ,)\partial(\Gamma,\mathcal{H}): we have γnz\gamma_{n}\to z if and only if γnxz\gamma_{n}\cdot x\to z for all but a single x(Γ,)x\in\partial(\Gamma,\mathcal{H}).

2.4. The coned-off Cayley graph

Whenever (Γ,)(\Gamma,\mathcal{H}) is a relatively hyperbolic pair, there are only finitely many conjugacy classes of groups in \mathcal{H}. We can fix a finite set 𝒫\mathcal{P} of conjugacy representatives for the groups in \mathcal{H}. The set 𝒫\mathcal{P} corresponds to a finite set ΠparΓ\Pi\subset\partial_{\mathrm{par}}\Gamma of parabolic points, such that

𝒫={Γp:pΠ}.\mathcal{P}=\{\Gamma_{p}:p\in\Pi\}.

Then Π\Pi contains exactly one point in each Γ\Gamma-orbit in parΓ\partial_{\mathrm{par}}\Gamma.

Definition 2.14.

Let (Γ,)(\Gamma,\mathcal{H}) be a relatively hyperbolic pair, and fix a finite generating set SS for Γ\Gamma and finite collection of conjugacy representatives 𝒫\mathcal{P} for \mathcal{H}.

The coned-off Cayley graph Cay(Γ,S,𝒫)\mathrm{Cay}(\Gamma,S,\mathcal{P}) is a metric space obtained from the Cayley graph Cay(Γ,S)\mathrm{Cay}(\Gamma,S) as follows: for each coset gPigP_{i} for Pi𝒫P_{i}\in\mathcal{P}, we add a vertex v(gPi)v(gP_{i}). Then, we add an edge of length 1 from each hgPih\in gP_{i} to v(gPi)v(gP_{i}).

The quasi-isometry class of Cay(Γ,S,𝒫)\mathrm{Cay}(\Gamma,S,\mathcal{P}) is independent of the choice of generating set SS. When (Γ,)(\Gamma,\mathcal{H}) is a relatively hyperbolic pair, Cay(Γ,S,𝒫)\mathrm{Cay}(\Gamma,S,\mathcal{P}) is a hyperbolic metric space. It is not a proper metric space if \mathcal{H} is nonempty. The Gromov boundary of Cay(Γ,S,𝒫)\mathrm{Cay}(\Gamma,S,\mathcal{P}) is equivariantly homeomorphic to the set conΓ\partial_{\mathrm{con}}\Gamma of conical limit points in (Γ,)\partial(\Gamma,\mathcal{H}).

3. Lie theory notation and background

For the rest of the paper, we let GG be a connected semisimple Lie group with no compact factor and finite center. We will be concerned with representations ρ:ΓG\rho:\Gamma\to G, where Γ\Gamma is a relatively hyperbolic group. We want to consider the action of ρ(Γ)\rho(\Gamma) on the flag manifold G/PG/P, where PP is a parabolic subgroup of GG.

In this section, we give an overview of the definitions and notation we will use to describe the dynamical behavior of the Γ\Gamma-action on G/PG/P. 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 GG to reduce to the case G=PGL(n,)G=\operatorname{PGL}(n,\mathbb{R}). The most important part of this section is 3.5, which identifies the connection between PP-divergence (or equivalently τmod{\tau_{\mathrm{mod}}}-regularity) and contracting dynamics in GG.

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 KK be a maximal compact subgroup of the semisimple Lie group GG, and let XX be the Riemannian symmetric space G/KG/K. A subgroup PGP\subset G is a parabolic subgroup if it is the stabilizer of a point in the visual boundary X\partial_{\infty}X of XX. Two parabolic subgroups P,QP,Q are opposite if there is a bi-infinite geodesic cc in XX so that PP is the stabilizer of c()c(\infty) and QQ is the stabilizer of c()c(-\infty).

The compact homogeneous GG-space G/PG/P is called a flag manifold. If PP and QQ are parabolic subgroups, then we say that two flags ξ+G/P\xi^{+}\in G/P and ξG/Q\xi^{-}\in G/Q are opposite if the stabilizers of ξ+\xi^{+}, ξ\xi^{-} are opposite parabolic subgroups. (In particular a conjugate of QQ must be opposite to PP).

3.2. Root space decomposition

Let 𝔤\mathfrak{g} be the Lie algebra of GG, and let 𝔨\mathfrak{k} be the Lie algebra of the maximal compact KK. We can decompose 𝔤\mathfrak{g} as 𝔨𝔭\mathfrak{k}\oplus\mathfrak{p}, and fix a maximal abelian subalgebra 𝔞𝔭\mathfrak{a}\subset\mathfrak{p}. The restriction of the Killing form BB to 𝔭\mathfrak{p} is positive definite, so any maximal abelian 𝔞𝔭\mathfrak{a}\subset\mathfrak{p} is naturally endowed with a Euclidean structure.

Each element of the abelian subalgebra 𝔞\mathfrak{a} acts semisimply on 𝔤\mathfrak{g}, with real eigenvalues. So we let Σ𝔞\Sigma\subset\mathfrak{a}^{*} denote the set of roots for this choice of 𝔞\mathfrak{a}, i.e. the set of nonzero linear functionals α𝔞\alpha\in\mathfrak{a}^{*} such that the linear map 𝔤𝔤\mathfrak{g}\to\mathfrak{g} given by Xα(X)IX-\alpha(X)I has nonzero kernel for every X𝔞X\in\mathfrak{a}. We have a restricted root space decomposition

𝔤=𝔤0αΣ𝔤α,\mathfrak{g}=\mathfrak{g}_{0}\oplus\bigoplus_{\alpha\in\Sigma}\mathfrak{g}_{\alpha},

where X𝔞X\in\mathfrak{a} acts on 𝔤α\mathfrak{g}_{\alpha} by multiplication by α(X)\alpha(X).

We choose a set of simple roots ΔΣ\Delta\subset\Sigma so that each αΣ\alpha\in\Sigma can be uniquely written as a linear combination of elements of Δ\Delta with coefficients either all nonnegative or all nonpositive. We let Σ+\Sigma_{+} denote the positive roots, i.e. roots which are nonnegative linear combinations of elements of Δ\Delta.

The simple roots Δ\Delta determine a Euclidean Weyl chamber

𝔞+={x𝔞:α(x)0, for all αΔ}.\mathfrak{a}^{+}=\{x\in\mathfrak{a}:\alpha(x)\geq 0,\textrm{ for all }\alpha\in\Delta\}.

The kernels of the roots αΔ\alpha\in\Delta are the walls of the Euclidean Weyl chamber.

Choosing a maximal compact KK, a maximal abelian 𝔞𝔭\mathfrak{a}\subset\mathfrak{p}, and a Euclidean Weyl chamber 𝔞+\mathfrak{a}^{+} determines a Cartan projection

μ:G𝔞+,\mu:G\to\mathfrak{a}^{+},

uniquely determined by the equation g=kexp(μ(g))kg=k\exp(\mu(g))k^{\prime}, where k,kKk,k^{\prime}\in K and μ(g)𝔞+\mu(g)\in\mathfrak{a}^{+}.

3.3. PP-divergence

Fix a subset θ\theta of the simple roots Δ\Delta. We define a standard parabolic subgroup Pθ+P^{+}_{\theta} to be the normalizer of the Lie algebra

αΣθ+𝔤α,\bigoplus_{\alpha\in\Sigma^{+}_{\theta}}\mathfrak{g}_{\alpha},

where Σθ+\Sigma^{+}_{\theta} is the set of positive roots which are not in the span of Δθ\Delta-\theta. The opposite subgroup PP^{-} is the normalizer of

αΣθ+𝔤α.\bigoplus_{\alpha\in\Sigma^{+}_{\theta}}\mathfrak{g}_{-\alpha}.

Every parabolic subgroup PGP\subset G is conjugate to a unique standard parabolic subgroup Pθ+P^{+}_{\theta}, and every pair of opposite parabolics (P+,P)(P^{+},P^{-}) is simultaneously conjugate to a unique pair (Pθ+,Pθ)(P^{+}_{\theta},P^{-}_{\theta}).

For a fixed θΔ\theta\subset\Delta, the group Pθ+P^{+}_{\theta} is the stabilizer of the endpoint of a geodesic ray exp(tZ)p\exp(tZ)\cdot p, where pXp\in X is the image of the identity in G/KG/K, and for any αΔ\alpha\in\Delta, the element Z𝔞+Z\in\mathfrak{a}^{+} satisfies

α(Z)=0αΔθ.\alpha(Z)=0\iff\alpha\in\Delta-\theta.
Definition 3.1.

Let gng_{n} be a sequence in GG. The sequence gng_{n} is Pθ+P^{+}_{\theta}-divergent if for every αθ\alpha\in\theta, we have

α(μ(gn)).\alpha(\mu(g_{n}))\to\infty.

That is, the Cartan projections of the sequence gng_{n} drift away from the walls of 𝔞\mathfrak{a} determined by the subset θΔ\theta\subset\Delta.

For a general parabolic subgroup PGP\subset G, we say that gng_{n} is PP-divergent if gng_{n} is Pθ+P^{+}_{\theta}-divergent for Pθ+P^{+}_{\theta} conjugate to PP.

3.4. Affine charts

Definition 3.2.

Let P+P^{+}, PP^{-} be opposite parabolic subgroups in GG. Given a flag ξG/P\xi\in G/P^{-}, we define

Opp(ξ)={ηG/P+:ξ is opposite to η}.\operatorname{Opp}(\xi)=\{\eta\in G/P^{+}:\xi\textrm{ is opposite to }\eta\}.

We call a set of the form Opp(ξ)\operatorname{Opp}(\xi) for some ξG/P\xi\in G/P^{-} an affine chart in G/P+G/P^{+}.

An affine chart is the unique open dense orbit of StabG(ξ)\operatorname{Stab}_{G}(\xi) in G/P+G/P^{+}. When G=PGL(d,)G=\operatorname{PGL}(d,\mathbb{R}) and P+P^{+} is the stabilizer of a line d\ell\subset\mathbb{R}^{d}, G/P+G/P^{+} is identified with (d)\mathbb{P}(\mathbb{R}^{d}) and this notion of affine chart agrees with the usual one in (d)\mathbb{P}(\mathbb{R}^{d}).

3.5. Dynamics in flag manifolds

There is a close connection between PP-divergence in the group GG and the topological dynamics of the action of GG on the associated flag manifold G/PG/P. Kapovich-Leeb-Porti frame this connection in terms of a contraction property for PP-divergent sequences.

Definition 3.3 ([KLP17], Definition 4.1).

Let gng_{n} be a sequence of group elements in GG. We say that gng_{n} is P+P^{+}-contracting if there exist ξG/P+\xi\in G/P^{+}, ξG/P\xi_{-}\in G/P^{-} such that gng_{n} converges uniformly to ξ\xi on compact subsets of Opp(ξ)\operatorname{Opp}(\xi_{-}).

The flag ξ\xi is the uniquely determined limit of the sequence gng_{n}.

Definition 3.4.

For an arbitrary sequence gnGg_{n}\in G, a P+P^{+}-limit point of gng_{n} in G/P+G/P^{+} is the limit point of some P+P^{+}-contracting subsequence of gng_{n}.

The P+P^{+}-limit set of a group ΓG\Gamma\subset G is the set of P+P^{+}-limit points of sequences in Γ\Gamma.

The importance of contracting sequences is captured by the following:

Proposition 3.5 ([KLP17], Proposition 4.15).

A sequence gnGg_{n}\in G is P+P^{+}-divergent if and only if every subsequence of gng_{n} has a P+P^{+}-contracting subsequence.

3.5 implies in particular that if gnGg_{n}\in G is P+P^{+}-divergent, then up to subsequence there is an open subset UG/P+U\subset G/P^{+} such that gnUg_{n}\cdot U converges to a singleton in G/P+G/P^{+}. It turns out that this “weak contraction property” is enough to characterize P+P^{+}-divergence.

{restatable}

proplocalContractingImpliesGlobalContracting Let gng_{n} be a sequence in GG, and suppose that for some nonempty open subset UG/P+U\subset G/P^{+}, we have gnU{ξ}g_{n}\cdot U\to\{\xi\} for ξG/P+\xi\in G/P^{+}. Then gng_{n} is P+P^{+}-divergent, and has a unique P+P^{+}-limit point ξG/P+\xi\in G/P^{+}.

We provide a proof of this fact in Appendix A.

3.5.1. Dynamics of inverses of P+P^{+}-divergent sequences

When gng_{n} is a P+P^{+}-divergent sequence, the inverse sequence is PP^{-}-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 gnGg_{n}\in G and flags ξG/P,ξ+G/P+\xi_{-}\in G/P^{-},\xi_{+}\in G/P^{+}, the following are equivalent:

  1. (1)

    gng_{n} is P+P^{+}-contracting and gn|Opp(ξ)ξ+g_{n}|_{\operatorname{Opp}(\xi_{-})}\to\xi_{+} uniformly on compacts.

  2. (2)

    gng_{n} is P+P^{+}-divergent, gng_{n} has unique P+P^{+}-limit point ξ+\xi_{+}, and gn1g_{n}^{-1} has unique PP^{-}-limit point ξ\xi_{-}.

3.6. τmod{\tau_{\mathrm{mod}}}-regularity

PP-divergent sequences are equivalent to the τmod{\tau_{\mathrm{mod}}}-regular sequences discussed in the work of Kapovich-Leeb-Porti, where τmod{\tau_{\mathrm{mod}}} is the unique face corresponding to PP in a spherical model Weyl chamber. We explain the connection here.

Remark 3.7.

The language of τmod{\tau_{\mathrm{mod}}}-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 pXp\in X, we let 𝔭\mathfrak{p} be the uniquely determined subspace of 𝔤\mathfrak{g} such that 𝔤=𝔨𝔭\mathfrak{g}=\mathfrak{k}\oplus\mathfrak{p}, where 𝔨\mathfrak{k} is the Lie algebra of the stabilizer of pp in GG.

Let zXz\in\partial_{\infty}X. There is a point pXp\in X, a maximal abelian subalgebra 𝔞𝔭\mathfrak{a}\subset\mathfrak{p}, a Euclidean Weyl chamber 𝔞+𝔞\mathfrak{a}^{+}\subset\mathfrak{a}, and a unit-length Z𝔞+Z\in\mathfrak{a}^{+} such that zz is the endpoint of the geodesic ray c(t)=exp(tZ)pc(t)=\exp(tZ)\cdot p.

Up to the action of the stabilizer of zz, the point pp, the maximal abelian subalgebra 𝔞\mathfrak{a}, the Euclidean Weyl chamber 𝔞+\mathfrak{a}^{+}, and the unit vector Z𝔞+Z\in\mathfrak{a}^{+} are uniquely determined. In addition, the stabilizer in GG of the triple (p,𝔞,𝔞+)(p,\mathfrak{a},\mathfrak{a}^{+}) acts trivially on 𝔞+\mathfrak{a}^{+}.

This means that we can identify the space X/G\partial_{\infty}X/G with the set of unit vectors in any Euclidean Weyl chamber 𝔞+\mathfrak{a}^{+}. This set has the structure of a spherical simplex. We let σmod\sigma_{\mathrm{mod}} denote the model spherical Weyl chamber X/G\partial_{\infty}X/G.

We let π:Xσmod\pi:\partial_{\infty}X\to\sigma_{\mathrm{mod}} be the type map to the model spherical Weyl chamber. For fixed zXz\in\partial_{\infty}X, we let PzP_{z} denote the parabolic subgroup stabilizing zz.

After choosing a maximal compact KK, a maximal abelian 𝔞𝔭\mathfrak{a}\subset\mathfrak{p}, and a Euclidean Weyl chamber 𝔞+\mathfrak{a}^{+}, the data of a face τmod{\tau_{\mathrm{mod}}} of the spherical simplex σmod\sigma_{\mathrm{mod}} is the same as the data of a subset of the simple roots of GG: the set of roots identifies a collection of walls of the Euclidean Weyl chamber 𝔞+\mathfrak{a}^{+}. The intersection of those walls with the unit sphere in 𝔞\mathfrak{a} is uniquely identified with a face of σmod\sigma_{\mathrm{mod}}.

Definition 3.8.

Let τmod{\tau_{\mathrm{mod}}} be a face of the model spherical Weyl chamber σmod\sigma_{\mathrm{mod}}. We say that a sequence gnGg_{n}\in G is τmod{\tau_{\mathrm{mod}}}-regular if gng_{n} is PzP_{z}-divergent for some zXz\in\partial_{\infty}X such that π(z)τmod\pi(z)\in{\tau_{\mathrm{mod}}}.

For a fixed model face τmodσmod{\tau_{\mathrm{mod}}}\subset\sigma_{\mathrm{mod}}, we let PτmodP_{\tau_{\mathrm{mod}}} denote any parabolic subgroup which is the stabilizer of a point zπ1(τmod)z\in\pi^{-1}({\tau_{\mathrm{mod}}}). All such parabolic subgroups are conjugate, so as a GG-space the flag manifold G/PτmodG/P_{\tau_{\mathrm{mod}}} depends only on the model face τmod{\tau_{\mathrm{mod}}}.

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 Γ\Gamma to a semisimple Lie group GG 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 G=PGL(d,)G=\operatorname{PGL}(d,\mathbb{R}) or SL(d,)\operatorname{SL}(d,\mathbb{R}) and the parabolic subgroup PP is the stabilizer of a flag of type (1,d1)(1,d-1) in d\mathbb{R}^{d}.

Definition 4.1.

Let PP be a parabolic subgroup of GG. We say that PP is symmetric if P=P+P=P^{+} is conjugate to a subgroup PP^{-} opposite to PP.

When P=P+P=P^{+} is symmetric, we can identify G/P+G/P^{+} with G/PG/P^{-}, so that it makes sense to say that two flags ξ1,ξ2G/P\xi_{1},\xi_{2}\in G/P are opposite.

Definition 4.2.

Let PP be symmetric, and let A,BA,B be two subsets of G/PG/P. We say that AA and BB are opposite if every ξA\xi\in A is opposite to every νB\nu\in B.

Definition 4.3.

Let (Γ,)(\Gamma,\mathcal{H}) be a relatively hyperbolic pair, and let ΛG/P\Lambda\subset G/P for a symmetric parabolic PP. We say that a continuous surjective map ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}) is antipodal if for every pair of distinct points z1,z2(Γ,)z_{1},z_{2}\in\partial(\Gamma,\mathcal{H}), ϕ1(z1)\phi^{-1}(z_{1}) is opposite to ϕ1(z2)\phi^{-1}(z_{2}).

We recall the main definition of the paper here:

\EGF

*

Remark 4.4.

Unfortunately, the boundary set ΛG/P\Lambda\subset G/P is not necessarily uniquely determined by the representation ρ\rho. 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 ρ:ΓG\rho:\Gamma\to G is EGF, the action of ρ(Γ)\rho(\Gamma) on the boundary set Λ\Lambda is by definition an extension of the topological dynamical system (Γ,(Γ,))(\Gamma,\partial(\Gamma,\mathcal{H})). When Γ\Gamma is non-elementary, convergence dynamics imply that the homomorphism ΓHomeo((Γ,))\Gamma\to\operatorname{Homeo}(\partial(\Gamma,\mathcal{H})) has finite kernel and discrete image. So the map ΓHomeo(Λ)\Gamma\to\operatorname{Homeo}(\Lambda) must also have discrete image and finite kernel, and therefore so does the representation ρ:ΓG\rho:\Gamma\to G. The case where Γ\Gamma is elementary can be verified directly.

4.2. Shrinking the sets CzC_{z}

Let ρ:ΓG\rho:\Gamma\to G be an EGF representation with boundary map ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}). By assumption, we know there exists an open subset CzG/PC_{z}\subset G/P for each z(Γ,)z\in\partial(\Gamma,\mathcal{H}), satisfying the extended convergence dynamics conditions (Definition 1.1). In general, there is not a canonical choice for the set CzC_{z}. We are able to make some assumptions about the properties of the CzC_{z}, however.

Proposition 4.5.

Let ρ:ΓG\rho:\Gamma\to G be an EGF representation with boundary extension ϕ\phi. For any z(Γ,)z\in\partial(\Gamma,\mathcal{H}), we can choose the set CzC_{z} to be a subset of

Opp(ϕ1(z)):={ξG/P:ξ is opposite to ν for every νϕ1(z)}.\operatorname{Opp}(\phi^{-1}(z)):=\{\xi\in G/P:\xi\textrm{ is opposite to }\nu\textrm{ for every }\nu\in\phi^{-1}(z)\}.
Proof.

Since ϕ1(z)\phi^{-1}(z) is closed, Opp(ϕ1(z))\operatorname{Opp}(\phi^{-1}(z)) is an open subset of G/PG/P. And, transversality of ϕ\phi implies that Opp(ϕ1(z))\operatorname{Opp}(\phi^{-1}(z)) contains Λϕ1(z)\Lambda-\phi^{-1}(z). So the intersection CzOpp(ϕ1(z))C_{z}\cap\operatorname{Opp}(\phi^{-1}(z)) is open and nonempty, meaning we can replace CzC_{z} 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]).

{restatable}

propconicalPeripheralEGF Let ρ:ΓG\rho:\Gamma\to G be a representation of a relatively hyperbolic group, and let ΛG/P\Lambda\subset G/P be a closed ρ(Γ)\rho(\Gamma)-invariant set, where PGP\subset G is a symmetric parabolic subgroup. Suppose that ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}) is a continuous surjective ρ\rho-equivariant antipodal map.

Then ρ\rho is an EGF representation with EGF boundary extension ϕ\phi if and only if both of the following conditions hold:

  1. (a)

    For any sequence γnΓ\gamma_{n}\in\Gamma limiting conically to some point in (Γ,)\partial(\Gamma,\mathcal{H}), ρ(γn±1)\rho(\gamma_{n}^{\pm 1}) is PP-divergent and every PP-limit point of ρ(γn±1)\rho(\gamma_{n}^{\pm 1}) lies in Λ\Lambda.

  2. (b)

    For every parabolic point ppar(Γ,)p\in\partial_{\mathrm{par}}(\Gamma,\mathcal{H}), there exists an open set CpG/PC_{p}\subset G/P, with Λϕ1(p)Cp\Lambda-\phi^{-1}(p)\subset C_{p}, such that for any compact KCpK\subset C_{p} and any open set UU containing ϕ1(p)\phi^{-1}(p), for all but finitely many γΓp\gamma\in\Gamma_{p}, we have ρ(γ)KU\rho(\gamma)\cdot K\subset U.

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 Λ\Lambda

Proposition 4.6.

Let (Γ,)(\Gamma,\mathcal{H}) be a relatively hyperbolic pair, and let ρ:ΓG\rho:\Gamma\to G be a representation which is EGF with respect to a symmetric parabolic PP, with boundary extension ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}). Then Λ\Lambda contains the PP-limit set of ρ(Γ)\rho(\Gamma).

Proof.

Let ξG/P\xi\in G/P be a flag in the PP-limit set of ρ(Γ)\rho(\Gamma). Then there is a PP-contracting sequence ρ(γn)\rho(\gamma_{n}) for γnΓ\gamma_{n}\in\Gamma and a flag ξG/P\xi_{-}\in G/P such that ρ(γn)η\rho(\gamma_{n})\eta converges to ξ\xi for any η\eta in Opp(ξ)\operatorname{Opp}(\xi_{-}). Up to subsequence γn±1\gamma_{n}^{\pm 1} converges to z±(Γ,)z_{\pm}\in\partial(\Gamma,\mathcal{H}), so for any flag ηCz\eta\in C_{z_{-}}, the sequence ρ(γn)η\rho(\gamma_{n})\eta subconverges to a point in ϕ1(z+)\phi^{-1}(z_{+}). But since Opp(ξ)\operatorname{Opp}(\xi_{-}) is open and dense, for some ηCz\eta\in C_{z_{-}} we have ρ(γn)ηξ\rho(\gamma_{n})\eta\to\xi and hence ξϕ1(z+)\xi\in\phi^{-1}(z_{+}). ∎

In particular, 4.6 implies that the EGF boundary set ΛG/P\Lambda\subset G/P of an EGF representation ρ:ΓG\rho:\Gamma\to G must always contain the PP-proximal limit set of ρ(Γ)\rho(\Gamma). (Recall that gGg\in G is PP-proximal if it has a unique attracting fixed point in G/PG/P; the PP-proximal limit set of a subgroup of GG is the closure of the set of attracting fixed points of PP-proximal elements).

We will see that most of the power of EGF representations lies in the fact that their associated boundary extensions ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}) do not have to be homeomorphisms (so the Bowditch boundary of Γ\Gamma 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 ϕ\phi so that it has a well-defined inverse on conical limit points in (Γ,)\partial(\Gamma,\mathcal{H}). In fact, we can even get a somewhat precise description of all the fibers of ϕ\phi. Concretely, we have the following:

Proposition 4.7.

Let ρ:ΓG\rho:\Gamma\to G be an EGF representation, with boundary extension ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}). There is a ρ(Γ)\rho(\Gamma)-invariant closed subset ΛG/P\Lambda^{\prime}\subset G/P and a ρ\rho-equivariant map ϕ:Λ(Γ,)\phi^{\prime}:\Lambda^{\prime}\to\partial(\Gamma,\mathcal{H}) such that:

  1. (1)

    ϕ:Λ(Γ,)\phi^{\prime}:\Lambda^{\prime}\to\partial(\Gamma,\mathcal{H}) is also a boundary extension for ρ\rho,

  2. (2)

    for every zcon(Γ,)z\in\partial_{\mathrm{con}}(\Gamma,\mathcal{H}), ϕ1(z)\phi^{\prime-1}(z) is a singleton, and

  3. (3)

    for every ppar(Γ,)p\in\partial_{\mathrm{par}}(\Gamma,\mathcal{H}), ϕ1(p)\phi^{\prime-1}(p) is the closure of the set of all accumulation points of orbits γnx\gamma_{n}\cdot x for γn\gamma_{n} a sequence of distinct elements in Γp\Gamma_{p} and xCpx\in C_{p}.

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 Γ\Gamma be a subgroup of GG and suppose (Γ,)(\Gamma,\mathcal{H}) is a relatively hyperbolic pair. Let PGP\subset G be a symmetric parabolic subgroup.

The subgroup Γ\Gamma is relatively PP-Anosov if it is PP-divergent, and there is a Γ\Gamma-equivariant antipodal embedding (Γ,)G/P\partial(\Gamma,\mathcal{H})\to G/P whose image Λ\Lambda is the PP-limit set of Γ\Gamma.

Here, we say an embedding ψ:(Γ,)G/P\psi:\partial(\Gamma,\mathcal{H})\to G/P is antipodal if for every distinct ξ1,ξ2\xi_{1},\xi_{2} in (Γ,)\partial(\Gamma,\mathcal{H}), ψ(ξ1)\psi(\xi_{1}) and ψ(ξ2)\psi(\xi_{2}) are opposite flags.

Remark 4.9.

Several remarks on the definition are in order:

  1. (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.

  2. (b)

    When Γ\Gamma is a hyperbolic group (and the collection of peripheral subgroups \mathcal{H} is empty), then the Bowditch boundary (Γ,)\partial(\Gamma,\mathcal{H}) is identified with the Gromov boundary Γ\partial\Gamma. In this case, Definition 4.8 coincides with the usual definition of an Anosov representation.

  3. (c)

    In general, it is possible to define (relatively) PP-Anosov representations for a non-symmetric parabolic subgroup PP. However, there is no loss of generality in assuming that PP is symmetric: a representation ρ:ΓG\rho:\Gamma\to G is PP-Anosov if and only if it is PP^{\prime}-Anosov for a symmetric parabolic subgroup PGP^{\prime}\subset G depending only on PP.

Proposition 4.10.

Let ρ:ΓG\rho:\Gamma\to G be an EGF representation with respect to PP, and suppose that the boundary extension ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}) is a homeomorphism. Then:

  1. (1)

    ρ(Γ)\rho(\Gamma) is PP-divergent, and Λ\Lambda is the PP-limit set of ρ(Γ)\rho(\Gamma).

  2. (2)

    The sets CzC_{z} for z(Γ,)z\in\partial(\Gamma,\mathcal{H}) can be taken to be

    Opp(ϕ1(z))={νG/P:ν is opposite to ϕ1(z)}.\operatorname{Opp}(\phi^{-1}(z))=\{\nu\in G/P:\nu\textrm{ is opposite to }\phi^{-1}(z)\}.
Proof.

(1). Let γn\gamma_{n} be any infinite sequence of elements in Γ\Gamma. After extracting a subsequence, we have γn±1z±\gamma_{n}^{\pm 1}\to z_{\pm}, and since ϕ\phi is a homeomorphism, ρ(γn)\rho(\gamma_{n}) converges to the point ϕ1(z+)\phi^{-1}(z_{+}) uniformly on compacts in the open set CzC_{z_{-}}. Then 3.5 implies that ρ(γn)\rho(\gamma_{n}) is PP-divergent, with unique PP-limit point ϕ1(z+)Λ\phi^{-1}(z_{+})\in\Lambda.

(2). The fact that ϕ\phi is antipodal is exactly the statement that the sets Opp(ϕ1(z))\operatorname{Opp}(\phi^{-1}(z)) contain Λϕ1(z)\Lambda-\phi^{-1}(z) for every z(Γ,)z\in\partial(\Gamma,\mathcal{H}), so we just need to see that the appropriate dynamics hold for these sets. Let γn\gamma_{n} be an infinite sequence in Γ\Gamma with γn±1z±\gamma_{n}^{\pm 1}\to z_{\pm} for z±(Γ,)z_{\pm}\in\partial(\Gamma,\mathcal{H}).

We know that for open subsets U±G/PU_{\pm}\subset G/P, we have ρ(γn)U+ϕ1(z+)\rho(\gamma_{n})\cdot U_{+}\to\phi^{-1}(z_{+}) and ρ(γn1)Uϕ1(z)\rho(\gamma_{n}^{-1})U_{-}\to\phi^{-1}(z_{-}), uniformly on compacts. 3.5 implies that ρ(γn)\rho(\gamma_{n}) and ρ(γn1)\rho(\gamma_{n}^{-1}) are both PP-divergent with unique PP-limit points ϕ1(z+)\phi^{-1}(z_{+}), ϕ1(z)\phi^{-1}(z_{-}). So in fact by 3.6 ρ(γn)\rho(\gamma_{n}) converges to ϕ1(z+)\phi^{-1}(z_{+}) uniformly on compacts in Opp(ϕ1(z))\operatorname{Opp}(\phi^{-1}(z_{-})). ∎

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 (Γ,)(\Gamma,\mathcal{H}) be a relatively hyperbolic pair, and let PGP\subset G be a symmetric parabolic subgroup. A representation ρ:ΓG\rho:\Gamma\to G is relatively PP-Anosov in the sense of Definition 4.8 if and only if ρ\rho is EGF with respect to PP, and has an injective boundary extension ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}).

Proof.

4.10 ensures that if ρ\rho is an EGF representation, and the boundary extension ϕ\phi is a homeomorphism, then ρ\rho is PP-divergent and ϕ1\phi^{-1} is an antipodal embedding whose image is the PP-limit set.

On the other hand, if ρ\rho is relatively PP-Anosov, with boundary embedding ψ:(Γ,)Λ\psi:\partial(\Gamma,\mathcal{H})\to\Lambda, for each z(Γ,)z\in\partial(\Gamma,\mathcal{H}), we can take

Cz=Opp(ψ(z)).C_{z}=\operatorname{Opp}(\psi(z)).

Antipodality means that CzC_{z} contains Λψ(z)\Lambda-\psi(z), and PP-divergence and 3.6 imply that ρ(Γ)\rho(\Gamma) has the appropriate convergence dynamics. ∎

4.6. Relativization

We now turn to the situation where we have an EGF representation of a hyperbolic group Γ\Gamma with a nonempty collection of peripheral subgroups. That is, for some invariant set ΛG/P\Lambda\subset G/P, we have an EGF boundary extension ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}), where (Γ,)\partial(\Gamma,\mathcal{H}) is the Bowditch boundary of Γ\Gamma with peripheral structure \mathcal{H}.

We want to prove Theorem 1.11, which says that in this situation, if ρ\rho restricts to a PP-Anosov representation on each HH\in\mathcal{H}, then ρ\rho is a PP-Anosov representation of Γ\Gamma. For the rest of this section, we assume that Γ\Gamma is a hyperbolic group, and \mathcal{H} is a collection of subgroups of Γ\Gamma so that the pair (Γ,)(\Gamma,\mathcal{H}) is relatively hyperbolic. We let ρ:ΓG\rho:\Gamma\to G be an EGF representation for the pair (Γ,)(\Gamma,\mathcal{H}) with respect to a symmetric parabolic subgroup PGP\subset G, and we assume that for each HH\in\mathcal{H}, ρ|H:HG\rho|_{H}:H\to G is PP-Anosov, with Anosov limit map ψH:HG/P\psi_{H}:\partial H\to G/P.

The main step in the proof is to observe that it is always possible to choose the boundary extension ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}) so that Λ\Lambda is equivariantly homeomorphic to the Gromov boundary of Γ\Gamma (which we here denote Γ\partial\Gamma).

Whenever Γ\Gamma is a hyperbolic group and \mathcal{H} is a collection of subgroups so that (Γ,)(\Gamma,\mathcal{H}) is a relatively hyperbolic pair, there is an explicit description of the Bowditch boundary (Γ,)\partial(\Gamma,\mathcal{H}) in terms of the Gromov boundary Γ\partial\Gamma of Γ\Gamma—see [Ger12], [GP13], or [Tra13]. Specifically, we can say:

Proposition 4.12.

There is an equivariant surjective continuous map ϕΓ:Γ(Γ,)\phi_{\Gamma}:\partial\Gamma\to\partial(\Gamma,\mathcal{H}) such that for each conical limit point zz in (Γ,)\partial(\Gamma,\mathcal{H}), ϕΓ1(z)\phi_{\Gamma}^{-1}(z) is a singleton, and for each parabolic point p(Γ,)p\in\partial(\Gamma,\mathcal{H}) with H=StabΓ(p)H=\operatorname{Stab}_{\Gamma}(p), ϕΓ1(p)\phi_{\Gamma}^{-1}(p) is an embedded copy of H\partial H in Γ\partial\Gamma.

In our situation, we can see that the boundary extension ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}) satisfies similar properties.

Lemma 4.13.

There is a closed ρ(Γ)\rho(\Gamma)-invariant subset ΛG/P\Lambda^{\prime}\subset G/P and an EGF boundary extension ϕ:Λ(Γ,)\phi^{\prime}:\Lambda^{\prime}\to\partial(\Gamma,\mathcal{H}) such that:

  1. (1)

    For each conical limit point z(Γ,)z\in\partial(\Gamma,\mathcal{H}), ϕ1(z)\phi^{\prime-1}(z) is a singleton.

  2. (2)

    For each parabolic point p(Γ,)p\in\partial(\Gamma,\mathcal{H}), with H=StabΓ(p)H=\operatorname{Stab}_{\Gamma}(p), we have ϕ1(p)=ψH(H)\phi^{\prime-1}(p)=\psi_{H}(\partial H).

Proof.

We choose Λ\Lambda^{\prime} as in 4.7. The only thing we need to check is that for H=StabΓ(p)H=\operatorname{Stab}_{\Gamma}(p), the set ψH(H)\psi_{H}(\partial H) is exactly the closure of the set of accumulation points of ρ(H)\rho(H)-orbits in CpC_{p}. But since we may assume CpC_{p} is contained in Opp(ψH(H))\operatorname{Opp}(\psi_{H}(\partial H)), this follows immediately from the fact that ρ(H)\rho(H) is PP-divergent and the closed set ψH(H)\psi_{H}(\partial H) is the PP-limit set of ρ(H)\rho(H).

Next we need a lemma which will allow us to characterize the Gromov boundary of Γ\Gamma as an extension of the Bowditch boundary (Γ,)\partial(\Gamma,\mathcal{H}). First recall that if Γ\Gamma acts as a convergence group on a space ZZ, the limit set of Γ\Gamma is the set of points zZz\in Z such that for some yZy\in Z and some sequence γnΓ\gamma_{n}\in\Gamma, we have

γn|Z{y}z\gamma_{n}|_{Z-\{y\}}\to z

uniformly on compacts.

Lemma 4.14.

Let Γ\Gamma act on compact metrizable spaces XX and YY, and let ϕX:X(Γ,)\phi_{X}:X\to\partial(\Gamma,\mathcal{H}), ϕY:Y(Γ,)\phi_{Y}:Y\to\partial(\Gamma,\mathcal{H}) be continuous equivariant surjective maps such that for every conical limit point z(Γ,)z\in\partial(\Gamma,\mathcal{H}), ϕX1(z)\phi_{X}^{-1}(z) and ϕY1(z)\phi_{Y}^{-1}(z) are both singletons, and for every parabolic point p(Γ,)p\in\partial(\Gamma,\mathcal{H}), H=StabΓ(p)H=\operatorname{Stab}_{\Gamma}(p) acts as a convergence group on XX and YY, with limit sets ϕX1(p)\phi_{X}^{-1}(p), ϕY1(p)\phi_{Y}^{-1}(p) equivariantly homeomorphic to H\partial H.

Then for any sequences zn,zncon(Γ,)z_{n},z_{n}^{\prime}\in\partial_{\mathrm{con}}(\Gamma,\mathcal{H}), we have

limnϕX1(zn)=limnϕX1(zn)\lim_{n\to\infty}\phi_{X}^{-1}(z_{n})=\lim_{n\to\infty}\phi_{X}^{-1}(z_{n}^{\prime})

if and only if

limnϕY1(zn)=limnϕY1(zn).\lim_{n\to\infty}\phi_{Y}^{-1}(z_{n})=\lim_{n\to\infty}\phi_{Y}^{-1}(z_{n}^{\prime}).
Proof.

We proceed by contradiction, and suppose that for a pair of sequences zn,zncon(Γ,)z_{n},z_{n}^{\prime}\in\partial_{\mathrm{con}}(\Gamma,\mathcal{H}), we have

limnϕX1(zn)=limnϕX1(zn)=x,\lim_{n\to\infty}\phi_{X}^{-1}(z_{n})=\lim_{n\to\infty}\phi_{X}^{-1}(z_{n}^{\prime})=x,

but

limnϕY1(zn)limnϕY1(zn).\lim_{n\to\infty}\phi_{Y}^{-1}(z_{n})\neq\lim_{n\to\infty}\phi_{Y}^{-1}(z_{n}^{\prime}).

After taking a subsequence we may assume znz_{n} converges to z(Γ,)z\in\partial(\Gamma,\mathcal{H}), and that yn=ϕY1(zn)y_{n}=\phi_{Y}^{-1}(z_{n}) converges to yy and yn=ϕY1(zn)y_{n}^{\prime}=\phi_{Y}^{-1}(z_{n}^{\prime}) converges to yy^{\prime} for yyy\neq y^{\prime}. By continuity, we have

ϕY(y)=ϕY(y)=ϕX(x)=z.\phi_{Y}(y)=\phi_{Y}(y^{\prime})=\phi_{X}(x)=z.

Since ϕX\phi_{X} and ϕY\phi_{Y} are bijective on ϕX1(con(Γ,))\phi_{X}^{-1}(\partial_{\mathrm{con}}(\Gamma,\mathcal{H})) and ϕY1(con(Γ,))\phi_{Y}^{-1}(\partial_{\mathrm{con}}(\Gamma,\mathcal{H})) respectively, we must have z=pz=p for a parabolic point ppar(Γ,)p\in\partial_{\mathrm{par}}(\Gamma,\mathcal{H}). Let H=StabΓ(p)H=\operatorname{Stab}_{\Gamma}(p).

Since pp is a bounded parabolic point, we can find sequences of group elements hn,hnHh_{n},h_{n}^{\prime}\in H so that for a fixed compact subset K(Γ,){p}K\subset\partial(\Gamma,\mathcal{H})-\{p\}, we have

(2) hnznK,hnznK.h_{n}z_{n}\in K,\quad h_{n}^{\prime}z_{n}^{\prime}\in K.

This implies that no subsequence of hnynh_{n}y_{n} or hnynh_{n}^{\prime}y_{n}^{\prime} converges to a point in ϕY1(p)\phi_{Y}^{-1}(p).

Then, since HH acts as a convergence group on YY with limit set ϕY1(p)\phi_{Y}^{-1}(p), up to subsequence there are points u,uϕY1(p)u,u^{\prime}\in\phi_{Y}^{-1}(p) so that hnh_{n} converges to a point in ϕY1(p)\phi_{Y}^{-1}(p) uniformly on compacts in Y{u}Y-\{u\}, and hnh_{n}^{\prime} converges to a point in ϕY1(p)\phi_{Y}^{-1}(p) uniformly on compacts in Y{u}Y-\{u^{\prime}\}. So, we must have u=yu=y and u=yu^{\prime}=y^{\prime}.

This means that the sequences hn1h_{n}^{-1} and hn1h_{n}^{\prime-1} have distinct limits in the compactification H¯=HH\overline{H}=H\sqcup\partial H. So, there are distinct points v,vϕX1(p)v,v^{\prime}\in\phi_{X}^{-1}(p) so that (again up to subsequence) hnh_{n} converges to a point in ϕX1(p)\phi_{X}^{-1}(p) uniformly on compacts in X{v}X-\{v\}, and hnh_{n}^{\prime} converges to a point in ϕX1(p)\phi_{X}^{-1}(p) uniformly on compacts in X{v}X-\{v^{\prime}\}. Without loss of generality, we can assume xvx\neq v.

But then ϕX1(zn)\phi_{X}^{-1}(z_{n}) lies in a compact subset of X{v}X-\{v\}, so hnϕX1(zn)h_{n}\phi_{X}^{-1}(z_{n}) converges to a point in ϕX1(p)\phi_{X}^{-1}(p) and hnznh_{n}z_{n} converges to pp. But this contradicts (2) above. ∎

Proposition 4.15.

If the set Λ\Lambda satisfies the conclusions of 4.13, then Λ\Lambda is equivariantly homeomorphic to the Gromov boundary of Γ\Gamma.

Proof.

Let ϕΓ:Γ(Γ,)\phi_{\Gamma}:\partial\Gamma\to\partial(\Gamma,\mathcal{H}) denote the quotient map identifying the limit set of each HH\in\mathcal{H} to the parabolic point pp with H=StabΓ(p)H=\operatorname{Stab}_{\Gamma}(p). For each conical limit point z(Γ,)z\in\partial(\Gamma,\mathcal{H}), the fiber ϕΓ1(z)\phi_{\Gamma}^{-1}(z) is a singleton. So, there is an equivariant bijection ff from ϕΓ1(con(Γ,))\phi_{\Gamma}^{-1}(\partial_{\mathrm{con}}(\Gamma,\mathcal{H})) to ϕ1(con(Γ,))\phi^{-1}(\partial_{\mathrm{con}}(\Gamma,\mathcal{H})).

Moreover, since ϕΓ1(con(Γ,))\phi_{\Gamma}^{-1}(\partial_{\mathrm{con}}(\Gamma,\mathcal{H})) is Γ\Gamma-invariant, and the action of Γ\Gamma on its Gromov boundary Γ\partial\Gamma is minimal, ϕΓ1(con(Γ,))\phi_{\Gamma}^{-1}(\partial_{\mathrm{con}}(\Gamma,\mathcal{H})) is dense in Γ\partial\Gamma. We claim that ff extends to a continuous injective map ΓΛ\partial\Gamma\to\Lambda by defining f(x)=limf(xn)f(x)=\lim f(x_{n}) for any sequence xnxx_{n}\to x.

To see this, we can apply 4.14, taking Γ=X\partial\Gamma=X and Λ=Y\Lambda=Y. We know that Γ\Gamma always acts on its own Gromov boundary as a convergence group (so in particular each HH\in\mathcal{H} acts on Γ\partial\Gamma as a convergence group with limit set H\partial H). And, since ρ\rho restricts to a PP-Anosov representation on each HH\in\mathcal{H}, for any infinite sequence hnHh_{n}\in H, up to subsequence there are u,uψH(H)u,u_{-}\in\psi_{H}(\partial H) so that ρ(hn)\rho(h_{n}) converges to uu uniformly on compacts in Opp(u)\operatorname{Opp}(u_{-}). Antipodality of ϕ\phi implies that ρ(hn)\rho(h_{n}) converges to uu uniformly on compacts in ΛψH(H)\Lambda-\psi_{H}(\partial H). The other hypotheses of 4.14 follow from 4.12 and 4.13.

We still need to check that ff is actually surjective. We know that ff restricts to a bijection on ϕΓ1(con(Γ,))\phi_{\Gamma}^{-1}(\partial_{\mathrm{con}}(\Gamma,\mathcal{H})), and that ff takes ϕΓ1(p)\phi_{\Gamma}^{-1}(p) to ϕ1(p)\phi^{-1}(p) for each parabolic point pp in (Γ,)\partial(\Gamma,\mathcal{H}). So we just need to check that for every HH\in\mathcal{H}, ff restricts to a surjective map HψH(H)\partial H\to\psi_{H}(\partial H). If HH is non-elementary, this must be the case because the action of HH on H\partial H is minimal and ff maps H\partial H into ψH(H)\psi_{H}(\partial H) as an invariant closed subset. Otherwise, HH is virtually cyclic and H\partial H, ψH(H)\psi_{H}(\partial H) both contain exactly two points. Then injectivity of ff implies surjectivity.

So we conclude that there is a continuous bijection f:ΓΛf:\partial\Gamma\to\Lambda, and since Γ\partial\Gamma is compact and Λ\Lambda is metrizable, ff is a homeomorphism. ∎

We let f:ΛΓf:\Lambda\to\partial\Gamma 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 f:ΛΓf:\Lambda\to\partial\Gamma extends the convergence action of Γ\Gamma on its Gromov boundary Γ\partial\Gamma.

Proof.

By Section 4.3, we just need to show that if γnΓ\gamma_{n}\in\Gamma is a conical limit sequence with γn±1z±\gamma_{n}^{\pm 1}\to z_{\pm} for z±Γz_{\pm}\in\partial\Gamma, then every PP-limit point of ρ(γn±1)\rho(\gamma_{n}^{\pm 1}) lies in Λ\Lambda.

We consider two cases:

Case 1: ϕf(z+)\phi\circ f(z_{+}) is a parabolic point pp in (Γ,)\partial(\Gamma,\mathcal{H}). In this case, γn\gamma_{n} lies along a quasigeodesic ray in Γ\Gamma limiting to some z+Hz_{+}\in\partial H, with H=StabΓ(p)H=\operatorname{Stab}_{\Gamma}(p). This means that for a bounded sequence bnΓb_{n}\in\Gamma, we have γnbnH\gamma_{n}b_{n}\in H. Since ρ\rho restricts to a PP-Anosov representation on HH, this means that ρ(γnbn)\rho(\gamma_{n}b_{n}) is PP-divergent and every PP-limit point of ρ(γnbn)\rho(\gamma_{n}b_{n}) lies in ψH(z+)\psi_{H}(z_{+}). For the same reason, every PP-limit point of ρ(bn1γn1)\rho(b_{n}^{-1}\gamma_{n}^{-1}) lies in ψH(z+)\psi_{H}(z_{+}).

Up to subsequence bnb_{n} is a constant bb. We can use 3.5 to see that ρ(γn)\rho(\gamma_{n}) has the same PP-limit set as ρ(γnb)\rho(\gamma_{n}b). This PP-limit set lies in Λ\Lambda. And, every PP-limit point of ρ(γn1)\rho(\gamma_{n}^{-1}) is a bb-translate of a PP-limit point of ρ(b1γn1)\rho(b^{-1}\gamma_{n}^{-1}). This PP-limit set also lies in Λ\Lambda.

Case 2: ϕf(z+)\phi\circ f(z_{+}) is a conical limit point in (Γ,)\partial(\Gamma,\mathcal{H}). In this case, a subsequence of γn\gamma_{n} is a conical limit sequence for the action of Γ\Gamma on (Γ,)\partial(\Gamma,\mathcal{H}), and the desired result follows from the “only if” part of Section 4.3.

Proof of Theorem 1.11.

Let Γ\Gamma be hyperbolic, let \mathcal{H} be a collection of subgroups such that (Γ,)(\Gamma,\mathcal{H}) is a relatively hyperbolic pair, and let ρ:ΓG\rho:\Gamma\to G be an EGF representation with respect to PP, for the peripheral structure \mathcal{H}.

Suppose that ρ\rho restricts to a PP-Anosov representation on each HH\in\mathcal{H}. 4.16 implies that ρ\rho is also an EGF representation of Γ\Gamma for its empty peripheral structure, whose boundary extension can be chosen to be a homeomorphism. Then Section 1.7 says that ρ\rho is relatively PP-Anosov (again for the empty peripheral structure on Γ\Gamma). This ensures that ρ\rho is actually (non-relatively) PP-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 Γ\Gamma acting on a flag manifold G/PG/P, and a system of open sets in G/PG/P which is in some sense compatible with both the relative quasigeodesic automaton and the action of Γ\Gamma on G/PG/P.

The basic idea is motivated by the computational theory of hyperbolic groups. Given a hyperbolic group Γ\Gamma with finite generating set SS, it is always possible to find a finite directed graph 𝒢\mathcal{G}, with edges labeled by elements of SS, so that directed paths on 𝒢\mathcal{G} starting at a fixed vertex vid𝒢v_{\mathrm{id}}\in\mathcal{G} are in one-to-one correspondence with geodesic words in Γ\Gamma. The graph 𝒢\mathcal{G} is called a geodesic automaton for Γ\Gamma.

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 𝒢\mathcal{G} which encodes the behavior of quasigeodesics in the coned-off Cayley graph of a relatively hyperbolic group Γ\Gamma. Eventually, our goal is to build such an automaton by looking at the dynamics of the action of Γ\Gamma on its Bowditch boundary (Γ,)\partial(\Gamma,\mathcal{H}). 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 (Γ,)(\Gamma,\mathcal{H}) using an open covering of the Bowditch boundary (Γ,)\partial(\Gamma,\mathcal{H}) 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 Γ\Gamma acting by homeomorphisms on a connected compact metrizable space MM, before returning to the case where MM is a flag manifold G/PG/P for the rest of the paper.

Throughout the rest of this section, we fix a non-elementary relatively hyperbolic pair (Γ,)(\Gamma,\mathcal{H}), and let Πpar(Γ,)\Pi\subset\partial_{\mathrm{par}}(\Gamma,\mathcal{H}) be a finite set, containing exactly one point from each Γ\Gamma-orbit in par(Γ,)\partial_{\mathrm{par}}(\Gamma,\mathcal{H}). We also fix a finite generating set SS for Γ\Gamma, which allows us to refer to the coned-off Cayley graph Cay(Γ,S,𝒫)\mathrm{Cay}(\Gamma,S,\mathcal{P}) (Definition 2.14).

Definition 5.1.

A Γ\Gamma-graph is a finite directed graph 𝒢\mathcal{G} where each vertex vv is labelled with a subset TvΓT_{v}\subset\Gamma, which is either:

  • A singleton {γ}\{\gamma\}, with γid\gamma\neq\mathrm{id}, or

  • A cofinite subset of a coset gΓpg\Gamma_{p} for some pΠp\in\Pi, gΓg\in\Gamma.

A sequence {αn}Γ\{\alpha_{n}\}\subset\Gamma is a 𝒢\mathcal{G}-path if αnTvn\alpha_{n}\in T_{v_{n}} for a vertex path {vn}\{v_{n}\} in 𝒢\mathcal{G}.

Remark 5.2.

We will often refer to “the” vertex path {vn}\{v_{n}\} corresponding to a 𝒢\mathcal{G}-path {αn}\{\alpha_{n}\}, although we will never actually verify that such a vertex path is uniquely determined by the sequence of group elements {αn}\{\alpha_{n}\} in Γ\Gamma.

A vertex of a Γ\Gamma-graph which is labeled by a cofinite subset of a (necessarily unique) coset gΓpg\Gamma_{p} is a parabolic vertex. If vv is a parabolic vertex, we let pv=gpp_{v}=g\cdot p denote the corresponding parabolic point in par(Γ,)\partial_{\mathrm{par}}(\Gamma,\mathcal{H}).

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 Γ\Gamma-graphs using the convergence dynamics of the Γ\Gamma-action on (Γ,)\partial(\Gamma,\mathcal{H}).

Definition 5.4.

Let z(Γ,)z\in\partial(\Gamma,\mathcal{H}). We say that a 𝒢\mathcal{G}-path {αn}\{\alpha_{n}\} limits to zz if either:

  • zcon(Γ,)z\in\partial_{\mathrm{con}}(\Gamma,\mathcal{H}), {αn}\{\alpha_{n}\} is infinite, and the sequence

    {γn=α1αn}n=1\{\gamma_{n}=\alpha_{1}\cdots\alpha_{n}\}_{n=1}^{\infty}

    limits to zz in the compactification Γ¯=Γ(Γ,)\overline{\Gamma}=\Gamma\sqcup\partial(\Gamma,\mathcal{H}), or

  • zpar(Γ,)z\in\partial_{\mathrm{par}}(\Gamma,\mathcal{H}), {αn}\{\alpha_{n}\} is a finite 𝒢\mathcal{G}-path whose corresponding vertex path {vn}\{v_{n}\} ends at a parabolic vertex vNv_{N}, and

    z=α1αN1pvN.z=\alpha_{1}\cdots\alpha_{N-1}p_{v_{N}}.
Definition 5.5.

Let 𝒢\mathcal{G} be a Γ\Gamma-graph. The endpoint of a finite 𝒢\mathcal{G}-path {αn}n=1N\{\alpha_{n}\}_{n=1}^{N} is

α1αN.\alpha_{1}\cdots\alpha_{N}.
Definition 5.6.

A Γ\Gamma-graph 𝒢\mathcal{G} is a relative quasigeodesic automaton if:

  1. (1)

    There is a constant D>0D>0 so that for any infinite 𝒢\mathcal{G}-path αn\alpha_{n}, the sequence

    {γn=α1αn}Γ\{\gamma_{n}=\alpha_{1}\cdots\alpha_{n}\}\subset\Gamma

    lies Hausdorff distance at most DD from a geodesic ray in Cay(Γ,S,𝒫)\mathrm{Cay}(\Gamma,S,\mathcal{P}), based at the identity.

  2. (2)

    For every z(Γ,)z\in\partial(\Gamma,\mathcal{H}), there exists a 𝒢\mathcal{G}-path limiting to zz.

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 𝒢\mathcal{G} be a relative quasigeodesic automaton. There is a constant R>0R>0 so that set of endpoints of 𝒢\mathcal{G}-paths is RR-dense in Γ\Gamma.

Proof.

If Γ\Gamma is hyperbolic and \mathcal{H} 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 Γ\Gamma is coarsely dense in Γ\Gamma (see [Bog97]).

If \mathcal{H} is nonempty, there is some R>0R>0 so that the union of all of the cosets gΓpg\cdot\Gamma_{p} for pΠp\in\Pi is RR-dense in Γ\Gamma. So it suffices to show that for each pΠp\in\Pi, there is some R>0R>0 so that all but RR elements in any coset gΓpg\cdot\Gamma_{p} are the endpoints of a 𝒢\mathcal{G}-path.

For any such coset gΓpg\cdot\Gamma_{p}, we can find a finite 𝒢\mathcal{G}-path {αn}n=1N1\{\alpha_{n}\}_{n=1}^{N-1} limiting to the vertex gpg\cdot p. That is,

gp=α1αN1pvN.g\cdot p=\alpha_{1}\cdots\alpha_{N-1}p_{v_{N}}.

By definition pvN=gpp_{v_{N}}=g^{\prime}\cdot p with TvNT_{v_{N}} a cofinite subset of the coset gΓpg^{\prime}\Gamma_{p} That is,

gΓp=α1αN1gΓp,g\cdot\Gamma_{p}=\alpha_{1}\cdots\alpha_{N-1}g^{\prime}\Gamma_{p},

so for all but finitely many γgΓp\gamma\in g\cdot\Gamma_{p} (depending only on the size of the complement of TvNT_{v_{N}} in gΓpg^{\prime}\cdot\Gamma_{p}), we can find αNgΓp\alpha_{N}\in g^{\prime}\Gamma_{p} with

α1αN=γ.\alpha_{1}\cdots\alpha_{N}=\gamma.

Remark 5.8.

In general, we do not require the set of elements in Γ\Gamma labelling the vertices of a relative quasigeodesic automaton 𝒢\mathcal{G} to generate the group Γ\Gamma (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 Γ\Gamma acts on a metrizable space MM by homeomorphisms, and let 𝒢\mathcal{G} be a Γ\Gamma-graph. A 𝒢\mathcal{G}-compatible system of open sets for the action of Γ\Gamma on MM is an assignment of an open subset UvMU_{v}\subset M to each vertex vv of 𝒢\mathcal{G} such that for each edge e=(v,w)e=(v,w) in 𝒢\mathcal{G}, for some ε>0\varepsilon>0, we have

(3) αNM(Uw,ε)Uv\alpha\cdot N_{M}(U_{w},\varepsilon)\subset U_{v}

for all αTv\alpha\in T_{v}.

Remark 5.10.

If 𝒢\mathcal{G} has no parabolic vertices (so each set TvT_{v} contains a single group element αvΓ\alpha_{v}\in\Gamma), then (3) is equivalent to requiring αvUw¯Uv\alpha_{v}\cdot\overline{U_{w}}\subset U_{v} for every edge (v,w)(v,w) in 𝒢\mathcal{G}. When 𝒢\mathcal{G} has parabolic vertices (so TvT_{v} may be infinite), (3) may be a stronger condition.

Proposition 5.11.

Let 𝒢\mathcal{G} be a Γ\Gamma-graph, and let {Uv:v vertex of 𝒢}\{U_{v}:v\textrm{ vertex of $\mathcal{G}$}\} be a 𝒢\mathcal{G}-compatible system of subsets of (Γ,)\partial(\Gamma,\mathcal{H}) for the action of Γ\Gamma on (Γ,)\partial(\Gamma,\mathcal{H}).

There is a constant D>0D>0 satisfying the following: let {αn}\{\alpha_{n}\} be an infinite 𝒢\mathcal{G}-path, corresponding to a vertex path {vn}\{v_{n}\}, and suppose the sequence {γn=α1αn}\{\gamma_{n}=\alpha_{1}\cdots\alpha_{n}\} is divergent in Γ\Gamma. Then for any point zz in the intersection

U=n=1α1αnUvn+1,U_{\infty}=\bigcap_{n=1}^{\infty}\alpha_{1}\cdots\alpha_{n}U_{v_{n+1}},

the sequence γn\gamma_{n} lies within Hausdorff distance DD of a geodesic ray in Cay(Γ,S,𝒫)\mathrm{Cay}(\Gamma,S,\mathcal{P}) tending towards zz.

Proof.

Fix a point zUz\in U_{\infty}, and write z=z+z=z_{+} and Un=UvnU_{n}=U_{v_{n}}. We first claim that there is a uniform ε>0\varepsilon>0 and a point z(Γ,)z_{-}\in\partial(\Gamma,\mathcal{H}) such that

(4) d(γn1z+,γn1z)>εd(\gamma_{n}^{-1}z_{+},\gamma_{n}^{-1}z_{-})>\varepsilon

for all n0n\geq 0.

To prove the claim, choose a uniform ε>0\varepsilon>0 so that for every vertex vv in 𝒢\mathcal{G}, we have N(Uv,ε)(Γ,)N(U_{v},\varepsilon)\neq\partial(\Gamma,\mathcal{H}), and for every edge (v,w)(v,w) in 𝒢\mathcal{G} and every αTv\alpha\in T_{v}, we have αN(Uw,ε)Uv\alpha\cdot N(U_{w},\varepsilon)\subset U_{v}. Then we choose some z(Γ,)N¯(U1,ε)z_{-}\in\partial(\Gamma,\mathcal{H})-\overline{N}(U_{1},\varepsilon).

By the 𝒢\mathcal{G}-compatibility condition, we know that for any nn, γnUn+1γ1U2U1\gamma_{n}U_{n+1}\subset\ldots\subset\gamma_{1}U_{2}\subset U_{1}, so we know that d(z+,z)>εd(z_{+},z_{-})>\varepsilon.

Then, for any n1n\geq 1, we have

γn1z+Un+1.\gamma_{n}^{-1}z_{+}\in U_{n+1}.

Moreover since γnN(Un+1,ε)U1\gamma_{n}N(U_{n+1},\varepsilon)\subset U_{1}, we also have

γn1z(Γ,)N(Un+1,ε).\gamma_{n}^{-1}z_{-}\in\partial(\Gamma,\mathcal{H})-N(U_{n+1},\varepsilon).

So for all nn we have d(γn1z+,γn1z)>εd(\gamma_{n}^{-1}z_{+},\gamma_{n}^{-1}z_{-})>\varepsilon, which establishes that (4) holds for all nn.

Now, consider a bi-infinite geodesic cc in a cusped space YY for Γ\Gamma joining z+z_{+} and zz_{-}. The sequence of geodesics γn1c\gamma_{n}^{-1}\cdot c has endpoints in Y=(Γ,)\partial Y=\partial(\Gamma,\mathcal{H}) lying distance at least ε\varepsilon apart, so each geodesic in the sequence passes within a uniformly bounded neighborhood of a fixed basepoint y0Yy_{0}\in Y. Therefore γny0\gamma_{n}\cdot y_{0} lies in a uniformly bounded neighbood of the geodesic cc.

Since γn\gamma_{n} is divergent, γny0\gamma_{n}y_{0} can only accumulate at either z+z_{+} or zz_{-}. But in fact γny0\gamma_{n}y_{0} can only accumulate at z+z_{+}—for in the construction of cc above, we could have chosen any zz_{-} in the nonempty open set (Γ,)N¯(U1,ε)\partial(\Gamma,\mathcal{H})-\overline{N}(U_{1},\varepsilon), and since (Γ,)\partial(\Gamma,\mathcal{H}) is perfect there is at least one such zzz_{-}^{\prime}\neq z_{-}.

This implies that γn\gamma_{n} is a conical limit sequence in Γ\Gamma, limiting to z+z_{+}. Since the distance between γn\gamma_{n} and γn+1\gamma_{n+1} is bounded in Cay(Γ,S,𝒫)\mathrm{Cay}(\Gamma,S,\mathcal{P}), the desired conclusion follows. ∎

Definition 5.12.

Let 𝒢\mathcal{G} be a Γ\Gamma-graph. An infinite 𝒢\mathcal{G}-path {αn}\{\alpha_{n}\} is divergent if the sequence {γn=α1αn}\{\gamma_{n}=\alpha_{1}\cdots\alpha_{n}\} leaves every bounded subset of Γ\Gamma.

We say that a Γ\Gamma-graph 𝒢\mathcal{G} is divergent if every infinite 𝒢\mathcal{G}-path is divergent.

Whenever {Uv}\{U_{v}\} is a 𝒢\mathcal{G}-compatible system of open sets for a Γ\Gamma-graph 𝒢\mathcal{G}, one can think of a 𝒢\mathcal{G}-path {αn}\{\alpha_{n}\} as giving a symbolic coding of a point in the intersection

α1αnUn+1.\alpha_{1}\cdots\alpha_{n}U_{n+1}.

The following proposition gives a way to construct such a coding for a given point z(Γ,)z\in\partial(\Gamma,\mathcal{H}), given an appropriate pair of open coverings of the Bowditch boundary (Γ,)\partial(\Gamma,\mathcal{H}) compatible with a Γ\Gamma-graph 𝒢\mathcal{G}.

Proposition 5.13.

Let 𝒢\mathcal{G} be a divergent Γ\Gamma-graph. Suppose that for each vertex a𝒢a\in\mathcal{G}, there exist open subsets Va,WaV_{a},W_{a} of (Γ,)\partial(\Gamma,\mathcal{H}) such that the following conditions hold:

  1. (1)

    The sets {Wa}\{W_{a}\} give a 𝒢\mathcal{G}-compatible system of sets for the action of Γ\Gamma on (Γ,)\partial(\Gamma,\mathcal{H}).

  2. (2)

    For all vertices aa, we have VaWaV_{a}\subset W_{a} and Wa¯(Γ,)\overline{W_{a}}\neq\partial(\Gamma,\mathcal{H}).

  3. (3)

    The sets VaV_{a} give an open covering of (Γ,)\partial(\Gamma,\mathcal{H}).

  4. (4)

    For every z(Γ,)z\in\partial(\Gamma,\mathcal{H}) and every non-parabolic vertex aa such that zVaz\in V_{a}, there is an edge (a,b)(a,b) in 𝒢\mathcal{G} such that αa1zVb\alpha_{a}^{-1}\cdot z\in V_{b} for {αa}=Ta\{\alpha_{a}\}=T_{a}.

  5. (5)

    For every z(Γ,)z\in\partial(\Gamma,\mathcal{H}) and every parabolic vertex aa such that zVa{pa}z\in V_{a}-\{p_{a}\}, there is an edge (a,b)(a,b) in 𝒢\mathcal{G} and αTa\alpha\in T_{a} such that α1zVb\alpha^{-1}\cdot z\in V_{b}.

Then 𝒢\mathcal{G} is a relative quasigeodesic automaton for Γ\Gamma.

Refer to caption
znz_{n}
Wan+1W_{a_{n+1}}
Van+1V_{a_{n+1}}
WanW_{a_{n}}
VanV_{a_{n}}
αn\alpha_{n}
Refer to caption
ana_{n}
an+1a_{n+1}
Refer to caption
Figure 1. Illustration for the proof of 5.13. The group element αn\alpha_{n} nests an ε\varepsilon-neighborhood of Wan+1W_{a_{n+1}} inside of WanW_{a_{n}} whenever αnVan+1\alpha_{n}\cdot V_{a_{n+1}} intersects VanV_{a_{n}}.
Refer to caption
W0W_{0}
W1W_{1}
W2W_{2}
Refer to caption
α1\alpha_{1}
α2\alpha_{2}
Refer to caption
z0z_{0}
Refer to caption
z1z_{1}
z2z_{2}
Refer to caption
a0a_{0}
a1a_{1}
a2a_{2}
\ldots
Figure 2. By iterating the nesting procedure backwards, we produce an infinite 𝒢\mathcal{G}-path and a sequence of subsets intersecting in the initial point z=z0z=z_{0}.
Proof.

5.11 implies that any infinite 𝒢\mathcal{G}-path lies finite Hausdorff distance from a geodesic ray in Cay(Γ,S,𝒫)\mathrm{Cay}(\Gamma,S,\mathcal{P}). So, we just need to show that every z(Γ,)z\in\partial(\Gamma,\mathcal{H}) is the limit of a 𝒢\mathcal{G}-path.

The idea behind the proof is to use the fact that the sets VaV_{a} cover (Γ,)\partial(\Gamma,\mathcal{H}) to show that we can keep “expanding” a neighborhood of zz in (Γ,)\partial(\Gamma,\mathcal{H}) to construct a path in 𝒢\mathcal{G} limiting to zz. The {Va}\{V_{a}\} covering tells us how to find the next edge in the path, and the {Wa}\{W_{a}\} cover gives us the 𝒢\mathcal{G}-compatible system we need to show that the path is a geodesic.

We let AA denote the vertex set of 𝒢\mathcal{G}. When aAa\in A is not a parabolic vertex, we write Ta={γa}T_{a}=\{\gamma_{a}\}.

Case 1: zz is a conical limit point. Fix aAa\in A so that zVaz\in V_{a}. We take z0=zz_{0}=z, a0=aa_{0}=a, and define sequences {zn}n=0conΓ\{z_{n}\}_{n=0}^{\infty}\subset\partial_{\mathrm{con}}\Gamma, {an}n=0A\{a_{n}\}_{n=0}^{\infty}\subset A, and {αn}n=1Γ\{\alpha_{n}\}_{n=1}^{\infty}\subset\Gamma as follows:

  • If ana_{n} is not a parabolic vertex, then we choose αn+1=γan\alpha_{n+1}=\gamma_{a_{n}}. Let zn+1=αn+11znz_{n+1}=\alpha_{n+1}^{-1}\cdot z_{n}. Since conical limit points are invariant under the action of Γ\Gamma, zn+1z_{n+1} is a conical limit point. By condition 4, there is a vertex an+1a_{n+1} satisfying zn+1Van+1z_{n+1}\in V_{a_{n+1}} with (an,an+1)(a_{n},a_{n+1}) an edge in 𝒢\mathcal{G}.

  • If ana_{n} is a parabolic vertex, then since znz_{n} is a conical limit point, znpz_{n}\neq p for p=panp=p_{a_{n}}. Then condition 5 implies that there exists some αn+1Tan\alpha_{n+1}\in T_{a_{n}} so that αn+11znVan+1\alpha_{n+1}^{-1}\cdot z_{n}\in V_{a_{n+1}} for an edge (an,an+1)(a_{n},a_{n+1}) in 𝒢\mathcal{G}. Again, zn+1=αn+11znz_{n+1}=\alpha_{n+1}^{-1}\cdot z_{n} must be a conical limit point since conΓ\partial_{\mathrm{con}}\Gamma is Γ\Gamma-invariant.

The sequence {αn}\{\alpha_{n}\} necessarily gives a 𝒢\mathcal{G}-path. By assumption the sequence

γn=α1αn\gamma_{n}=\alpha_{1}\cdots\alpha_{n}

is divergent. And by construction z=γnznz=\gamma_{n}z_{n} lies in γnWan\gamma_{n}W_{a_{n}} for all nn. So, 5.11 implies that γn\gamma_{n} is a conical limit sequence, limiting to zz. See Figure 2.

Case 2: zz is a parabolic point. As before fix aAa\in A so that zVaz\in V_{a}, and take z0=zz_{0}=z, a0=aa_{0}=a. We inductively define sequences znz_{n}, ana_{n}, αn\alpha_{n} as before, but we claim that for some finite NN, aNa_{N} is a parabolic vertex with zN=paNz_{N}=p_{a_{N}}. For if not, we can build an infinite 𝒢\mathcal{G}-path (as in the previous case) limiting to zz. But then, 5.11 would imply that zz is actually a conical limit point. So, we must have

z=γNaN=α1αNaNz=\gamma_{N}a_{N}=\alpha_{1}\cdots\alpha_{N}a_{N}

as required.

Remark 5.14.

In a typical application of 5.13, it will not be possible to construct the open coverings {Va}\{V_{a}\} and {Wa}\{W_{a}\} so that Va=WaV_{a}=W_{a} for all vertices aa. In particular we expect this to be impossible whenever (Γ,)\partial(\Gamma,\mathcal{H}) is connected.

To conclude this section, we make one more observation about systems of 𝒢\mathcal{G}-compatible sets as in 5.13.

Lemma 5.15.

Let Γ\Gamma be a relatively hyperbolic group, let 𝒢\mathcal{G} be a Γ\Gamma-graph, and let {Va}\{V_{a}\}, {Wa}\{W_{a}\} be an assignment of open subsets of (Γ,)\partial(\Gamma,\mathcal{H}) to vertices of 𝒢\mathcal{G} satisfying the hypotheses of 5.13.

Fix zconΓz\in\partial_{\mathrm{con}}\Gamma and NN\in\mathbb{N}. There exists δ>0\delta>0 so that if d(z,z)<δd(z,z^{\prime})<\delta, then there are 𝒢\mathcal{G}-paths {αn},{βn}\{\alpha_{n}\},\{\beta_{n}\} limiting to z,zz,z^{\prime} respectively, with αi=βi\alpha_{i}=\beta_{i} for all i<Ni<N.

Proof.

Let {αn}\{\alpha_{n}\} be a 𝒢\mathcal{G}-path limiting to zz coming from the construction in 5.13, passing through vertices vnv_{n}. We choose δ>0\delta>0 small enough so that if d(z,z)<δd(z,z^{\prime})<\delta, then zz^{\prime} lies in the set

α1αNVvn+1.\alpha_{1}\cdots\alpha_{N}V_{v_{n+1}}.

Then for every i<Ni<N, we have

αi1αi11α11zVvi+1.\alpha_{i}^{-1}\alpha_{i-1}^{-1}\cdots\alpha_{1}^{-1}z^{\prime}\in V_{v_{i+1}}.

As in 5.13, we can then extend {αn}n=1N1\{\alpha_{n}\}_{n=1}^{N-1} to a 𝒢\mathcal{G}-path limiting to zz^{\prime}. ∎

6. Extended convergence dynamics

Let Γ\Gamma be a relatively hyperbolic group acting on a connected compact metrizable space MM. In this section, we will show that if the action of Γ\Gamma on MM extends the convergence dynamics of Γ\Gamma (Definition 1.1), then we can construct a relative quasigeodesic automaton 𝒢\mathcal{G} and a 𝒢\mathcal{G}-compatible system of open subsets of MM which are in some sense reasonably well-behaved with respect to the group action.

To give the precise statement, we let ΛM\Lambda\subset M be a closed Γ\Gamma-invariant subset, and let ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}) be an equivariant, surjective, and continuous map satisfying the following: for each z(Γ,)z\in\partial(\Gamma,\mathcal{H}), there is an open set CzMC_{z}\subset M containing Λϕ1(z)\Lambda-\phi^{-1}(z) such that:

  1. (1)

    For any sequence γnΓ\gamma_{n}\in\Gamma limiting conically to zz, with γn1z\gamma_{n}^{-1}\to z_{-}, any open set UU containing ϕ1(z)\phi^{-1}(z), and any compact KCzK\subset C_{z_{-}}, we have γnKU\gamma_{n}\cdot K\subset U for all sufficiently large nn.

  2. (2)

    For any parabolic point pp, any compact KCpK\subset C_{p}, and any open set UU containing ϕ1(p)\phi^{-1}(p), for all but finitely many γΓp\gamma\in\Gamma_{p}, we have γKU\gamma\cdot K\subset U.

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 ϕ\phi. In this context, we will show:

Proposition 6.1.

For any ε>0\varepsilon>0, there is a relative quasigeodesic automaton 𝒢\mathcal{G} for Γ\Gamma, a 𝒢\mathcal{G}-compatible system of open sets {Uv}\{U_{v}\} for the action of Γ\Gamma on MM, and a 𝒢\mathcal{G}-compatible system of open sets {Wv}\{W_{v}\} for the action of Γ\Gamma on (Γ,)\partial(\Gamma,\mathcal{H}) such that:

  1. (1)

    For every vertex vv, there is some zWvz\in W_{v} so that

    ϕ1(Wv)UvNM(ϕ1(z),ε).\phi^{-1}(W_{v})\subset U_{v}\subset N_{M}(\phi^{-1}(z),\varepsilon).
  2. (2)

    For every pΠp\in\Pi, there is a parabolic vertex aa with pa=pp_{a}=p. Moreover, for every parabolic vertex ww with pw=gpp_{w}=g\cdot p, (a,b)(a,b) is an edge of 𝒢\mathcal{G} if and only if (w,b)(w,b) is an edge of 𝒢\mathcal{G}.

  3. (3)

    If q=gpq=g\cdot p for pΠp\in\Pi, aa is a parabolic vertex with pa=qp_{a}=q, and (a,b)(a,b) is an edge of 𝒢\mathcal{G}, then qWaq\in W_{a} and UbCpU_{b}\subset C_{p}.

Remark 6.2.

By equivariance of ϕ\phi, for each pparΓp\in\partial_{\mathrm{par}}\Gamma, we can replace CpC_{p} with ΓpCp\Gamma_{p}\cdot C_{p} and assume that CpC_{p} is Γp\Gamma_{p}-invariant (and that if q=gpq=g\cdot p, then Cq=gCpC_{q}=g\cdot C_{p}).

The proof of 6.1 involves some technicalities, so we first outline the general approach:

  1. (1)

    For each z(Γ,)z\in\partial(\Gamma,\mathcal{H}), we construct a pair VzV_{z}, WzW_{z} of small open neighborhoods of zz and a subset TzΓT_{z}\subset\Gamma so that for each αTz\alpha\in T_{z}, α1\alpha^{-1} is “expanding” about some point in VzV_{z}. When zz is a conical limit point, then we can choose a single element αzΓ\alpha_{z}\in\Gamma which expands about every point in VzV_{z}. When zz is a parabolic point, we may use a different element of Γ\Gamma to “expand” about each uVz{z}u\in V_{z}-\{z\}.

    We choose VzV_{z}, WzW_{z}, and TzT_{z} so that if α1\alpha^{-1} is “expanding” about uVzu\in V_{z}, and α1uVy\alpha^{-1}u\in V_{y}, then α1WzWy\alpha^{-1}W_{z}\supset W_{y}. See Figure 3.

    Refer to caption
    α1u\alpha^{-1}u
    Refer to caption
    uu
    WzW_{z}
    VzV_{z}
    WyW_{y}
    VyV_{y}
    α1\alpha^{-1}
    Figure 3. The group element α1\alpha^{-1} is “expanding” about uVzu\in V_{z}. We will construct VzV_{z}, WzW_{z} and Vy,WyV_{y},W_{y} so that if α1u\alpha^{-1}u lies in VyV_{y}, then α1Wz\alpha^{-1}W_{z} contains WyW_{y}. Equivalently, we get the containment αWyWz\alpha W_{y}\subset W_{z} illustrated earlier in Figure 1.
  2. (2)

    Using compactness of (Γ,)\partial(\Gamma,\mathcal{H}), we pick a finite set of points a(Γ,)a\in\partial(\Gamma,\mathcal{H}) so that the sets {Va}\{V_{a}\} give an open covering of (Γ,)\partial(\Gamma,\mathcal{H}). These points in (Γ,)\partial(\Gamma,\mathcal{H}) are identified with the vertices of a Γ\Gamma-graph 𝒢\mathcal{G}. We define the edges of 𝒢\mathcal{G} in such a way so that if, for some αTa\alpha\in T_{a}, α1\alpha^{-1} expands about uVau\in V_{a} and α1uVb\alpha^{-1}u\in V_{b}, then there is an edge from aa to bb. This ensures that {Wa}\{W_{a}\} is a 𝒢\mathcal{G}-compatible system of open subsets of (Γ,)\partial(\Gamma,\mathcal{H}).

  3. (3)

    Simultaneously, we construct a 𝒢\mathcal{G}-compatible system {Ua}\{U_{a}\} of open sets in MM by taking UaU_{a} to be a small neighborhood of ϕ1(a)\phi^{-1}(a). The idea is to use the extended convergence dynamics to ensure that if, for some αTz\alpha\in T_{z}, α1\alpha^{-1} “expands” about some uVzu\in V_{z} and the point α1u\alpha^{-1}u lies in VyV_{y}, then α1Uz\alpha^{-1}U_{z} contains UyU_{y}. See Figure 6 below.

  4. (4)

    Finally, we use 5.13 to prove that 𝒢\mathcal{G} is actually a relative quasigeodesic automaton. The open sets Va,WaV_{a},W_{a} are constructed exactly to satisfy the conditions of the proposition, so the main thing to check in this step is that the graph 𝒢\mathcal{G} is actually divergent (using the action of Γ\Gamma on MM).

Throughout the rest of the section, we will work with fixed metrics on both (Γ,)\partial(\Gamma,\mathcal{H}) and MM. 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 αΓ\alpha\in\Gamma “expands” on a small open subset UU of a metric space XX, we just mean that αU\alpha U is quantifiably “bigger” than UU, and not that for any x,yUx,y\in U, we have d(αx,αy)Cd(x,y)d(\alpha\cdot x,\alpha\cdot y)\geq C\cdot d(x,y) for some expansion constant CC. 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 PGL(2,)\operatorname{PGL}(2,\mathbb{Z}). While PGL(2,)\operatorname{PGL}(2,\mathbb{Z}) is virtually a free group (and therefore word-hyperbolic), it is also relatively hyperbolic, relative to the collection \mathcal{H} of conjugates of the parabolic subgroup {(±1t01):t}\left\{\begin{pmatrix}\pm 1&t\\ 0&1\end{pmatrix}:t\in\mathbb{Z}\right\}.

Since PGL(2,)\operatorname{PGL}(2,\mathbb{Z}) acts with finite covolume on the hyperbolic plane 2\mathbb{H}^{2}, the Bowditch boundary of the pair (PGL(2,),)(\operatorname{PGL}(2,\mathbb{Z}),\mathcal{H}) is equivariantly identified with 2\partial\mathbb{H}^{2}, the visual boundary of 2\mathbb{H}^{2}. Given a non-parabolic point w2w\in\partial\mathbb{H}^{2}, we can find an element of PGL(2,)\operatorname{PGL}(2,\mathbb{Z}) which “expands” a neighborhood of ww. There are two distinct possibilities:

  1. (1)

    Suppose ww is in a small neighborhood VzV_{z} of a conical limit point z2z\in\partial\mathbb{H}^{2}. Then choose some loxodromic element γPGL(2,)\gamma\in\operatorname{PGL}(2,\mathbb{Z}) whose attracting fixed point is close to zz. Then, if WzW_{z} is a slightly larger neighborhood of zz, γ1Wz\gamma^{-1}\cdot W_{z} is large enough to contain a uniformly large neighborhood of γ1w\gamma^{-1}\cdot w. See Figure 4.

  2. (2)

    On the other hand, suppose ww is in a small neighborhood VqV_{q} of a parabolic fixed point q2q\in\partial\mathbb{H}^{2}, but wqw\neq q. We can find some element γΓq=StabΓ(q)\gamma\in\Gamma_{q}=\operatorname{Stab}_{\Gamma}(q) so that γ1\gamma^{-1} takes ww into a fundamental domain for the action of Γq\Gamma_{q} on 2{q}\partial\mathbb{H}^{2}-\{q\}. Then, if WqW_{q} is a slightly larger neighborhood of qq, γ1Wq\gamma^{-1}\cdot W_{q} is again large enough to contain a uniformly large neighborhood of γ1w\gamma^{-1}\cdot w. See Figure 5.

There is a slight issue with this approach: in the second case above (when ww is close to a parabolic point qq), it is actually not quite good enough to “expand” a neighborhood of ww by using Γq\Gamma_{q} to push ww into a fundamental domain for Γq\Gamma_{q} on 2{q}\partial\mathbb{H}^{2}-\{q\}. The reason is that there might be no such fundamental domain which is actually far away from 2{q}\partial\mathbb{H}^{2}-\{q\}. We resolve this issue by instead choosing γ\gamma to lie in a coset gΓpg\Gamma_{p}, where q=gpq=gp for some pΠp\in\Pi. Then γ1w\gamma^{-1}\cdot w lies in a fundamental domain for Γp\Gamma_{p} on 2{p}\partial\mathbb{H}^{2}-\{p\}, which allows us to get uniform control on the size of the expanded neighborhood γ1Wq\gamma^{-1}W_{q}.

Refer to caption
γ1\gamma^{-1}
zz
ww
Refer to caption
γ1w\gamma^{-1}w
Figure 4. For any point ww in a sufficiently small neighborhood VzV_{z} (pink) of zz, the expanded neighborhood γ1Wz\gamma^{-1}W_{z} (red) contains a uniform neighborhood of γ1w\gamma^{-1}w.
Refer to caption
γ1\gamma^{-1}
qq
ww
KqK_{q}
Refer to caption
γ1w\gamma^{-1}w
Figure 5. For any point wqw\neq q in a neighborhood VqV_{q} (pink) of the parabolic point qq, we find some γΓq\gamma\in\Gamma_{q} so that γ1w\gamma^{-1}w lies in KqK_{q} (dark gray), a fundamental domain for the action of Γq\Gamma_{q} on 2{q}\partial\mathbb{H}^{2}-\{q\}. The expanded neighborhood γ1Wq\gamma^{-1}W_{q} (red) contains a uniform neighborhood of KqK_{q}, so γ1Wq\gamma^{-1}W_{q} contains a uniform neighborhood of γ1w\gamma^{-1}w.

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 Γ\Gamma and in a neighborhood of the Γ\Gamma-invariant set ΛM\Lambda\subset M. 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 𝒢\mathcal{G}-compatible system of open sets on a Γ\Gamma-graph 𝒢\mathcal{G}.

Lemma 6.4.

There exists ε>0\varepsilon>0 (depending on ϕ\phi and DD) so that for any a,b(Γ,)a,b\in\partial(\Gamma,\mathcal{H}) with d(a,b)>Dd(a,b)>D, the ε\varepsilon-neighborhood of ϕ1(a)\phi^{-1}(a) in MM is contained in CbC_{b}.

Proof.

Since ϕ1(z)\phi^{-1}(z) is closed in MM, such an ε>0\varepsilon>0 exists for any fixed pair of distinct (a,b)(Γ,)2(a,b)\in\partial(\Gamma,\mathcal{H})^{2}. Then the result follows, since the space of pairs (a,b)((Γ,))2(a,b)\in(\partial(\Gamma,\mathcal{H}))^{2} satisfying d(a,b)>Dd(a,b)>D is compact. ∎

Lemma 6.5.

There exists εcon>0,δcon>0\varepsilon_{\mathrm{con}}>0,\delta_{\mathrm{con}}>0 satisfying the following: for any ε>0\varepsilon>0, δ>0\delta>0 with ε<εcon\varepsilon<\varepsilon_{\mathrm{con}}, δ<δcon\delta<\delta_{\mathrm{con}}, and every conical limit point zz, we can find:

  • A group element γzΓ\gamma_{z}\in\Gamma

  • Open subsets Wz,Vz(Γ,)W_{z},V_{z}\subset\partial(\Gamma,\mathcal{H}) with zVzWzz\in V_{z}\subset W_{z}

such that:

  1. (1)

    diam(Wz)<δ\mathrm{diam}(W_{z})<\delta,

  2. (2)

    In (Γ,)\partial(\Gamma,\mathcal{H}), we have

    NΓ(γz1Vz,δ)γz1Wz.N_{\partial\Gamma}(\gamma_{z}^{-1}V_{z},\delta)\subset\gamma_{z}^{-1}W_{z}.
  3. (3)

    In MM we have

    NM(γz1ϕ1(Wz),2ε)γz1NM(ϕ1(z),ε).N_{M}(\gamma_{z}^{-1}\phi^{-1}(W_{z}),2\varepsilon)\subset\gamma_{z}^{-1}N_{M}(\phi^{-1}(z),\varepsilon).
Remark 6.6.

Conditions (1) and (2) together imply that for any y,zconΓy,z\in\partial_{\mathrm{con}}\Gamma, if γz1Vz\gamma_{z}^{-1}V_{z} intersects VyV_{y}, then γzWyWz\gamma_{z}W_{y}\subset W_{z}. Later, we will see that condition (3) implies that if γz1Vz\gamma_{z}^{-1}V_{z} intersects VyV_{y}, then also γzNM(ϕ1(y),2ε)NM(ϕ1(z),ε)\gamma_{z}N_{M}(\phi^{-1}(y),2\varepsilon)\subset N_{M}(\phi^{-1}(z),\varepsilon) (giving us the inclusion indicated by Figure 3).

Refer to caption
2ε2\varepsilon
WzW_{z}
δ\delta
Refer to caption
ϕ\phi
ϕ\phi
γz1\gamma_{z}^{-1}
Refer to caption
<δ<\delta
MM
(Γ,)\partial(\Gamma,\mathcal{H})
Refer to caption
ε\varepsilon
Refer to caption
VzV_{z}
zz
Refer to caption
Figure 6. The group element γz1\gamma_{z}^{-1} is “expanding” about Vz(Γ,)V_{z}\subset\partial(\Gamma,\mathcal{H}): while WzW_{z} has diameter <δ<\delta, γz1Wz\gamma_{z}^{-1}W_{z} contains a δ\delta-neighborhood of γz1Vz\gamma_{z}^{-1}V_{z}. At the same time, γz1\gamma_{z}^{-1} enlarges an ε\varepsilon-neighborhood of ϕ1(z)\phi^{-1}(z) in MM, so that it contains a 2ε2\varepsilon-neighborhood of γz1ϕ1(Wz)\gamma_{z}^{-1}\phi^{-1}(W_{z}).
Proof.

For a conical limit point zz, we choose a sequence γn\gamma_{n} so that for distinct a,b(Γ,)a,b\in\partial(\Gamma,\mathcal{H}), we have γn1za\gamma_{n}^{-1}z\to a and γn1wb\gamma_{n}^{-1}w\to b for any wzw\neq z. That is, γn\gamma_{n} limits conically to zz in Γ¯\overline{\Gamma}, and γn1\gamma_{n}^{-1} converges (not necessarily conically) to bb. Since the Γ\Gamma-action on distinct pairs in (Γ,)\partial(\Gamma,\mathcal{H}) is cocompact (2.7), we may assume that d(a,b)>Dd(a,b)>D for a uniform constant D>0D>0.

We choose εcon>0\varepsilon_{\mathrm{con}}>0 from 6.4 so that if a,b(Γ,)a,b\in\partial(\Gamma,\mathcal{H}) satisfy d(a,b)>D/2d(a,b)>D/2, then a 2εcon2\varepsilon_{\mathrm{con}}-neighborhood of ϕ1(a)\phi^{-1}(a) is contained in CbC_{b}. Let ε>0\varepsilon>0 satisfy ε<εcon\varepsilon<\varepsilon_{\mathrm{con}}, and let δ\delta satisfy δ<δcon:=D/4\delta<\delta_{\mathrm{con}}:=D/4.

By the triangle inequality, we have d(c,b)>D/2d(c,b)>D/2 for all cBΓ(a,2δ)c\in B_{\partial\Gamma}(a,2\delta), so the closed 2ε2\varepsilon-neighborhood of ϕ1(BΓ(a,2δ))\phi^{-1}(B_{\partial\Gamma}(a,2\delta)) is contained in CbC_{b}. This means that we can choose nn large enough so that

γnNM(ϕ1(B(a,2δ)),2ε)\gamma_{n}\cdot N_{M}(\phi^{-1}(B(a,2\delta)),2\varepsilon)

is contained in NM(ϕ1(z),ε)N_{M}(\phi^{-1}(z),\varepsilon) and

γnBΓ(a,2δ)\gamma_{n}\cdot B_{\partial\Gamma}(a,2\delta)

is contained in BΓ(z,δ/2)B_{\partial\Gamma}(z,\delta/2). We let γz=γn\gamma_{z}=\gamma_{n} for this large nn, and take

Wz=γzBΓ(a,2δ)W_{z}=\gamma_{z}\cdot B_{\partial\Gamma}(a,2\delta)

and

Vz=γzBΓ(a,δ).V_{z}=\gamma_{z}\cdot B_{\partial\Gamma}(a,\delta).

The next lemma is a version of 6.5 for parabolic points. As before, we want to show that for a point qq in the Bowditch boundary, we can find a neighborhood WqW_{q} of qq in (Γ,)\partial(\Gamma,\mathcal{H}) with uniformly bounded diameter δ\delta, and group elements γΓ\gamma\in\Gamma so that γ1\gamma^{-1} enlarges WqW_{q} enough to contain a 2δ2\delta-neighborhood of γ1z\gamma^{-1}z, for some zz close to qq. Simultaneously we want to choose γ\gamma so that γ1\gamma^{-1} enlarges an ε\varepsilon-neighborhood of ϕ1(q)\phi^{-1}(q) in a similar manner. This case is more complicated, because we need to allow γ\gamma to depend on the point zWqz\in W_{q}: if qq is a parabolic point in (Γ,)\partial(\Gamma,\mathcal{H}), then in general there is not a single group element in Γ\Gamma which expands distances in a neighborhood of qq.

Lemma 6.7.

For each point pΠp\in\Pi, there exist constants εp>0\varepsilon_{p}>0, δp>0\delta_{p}>0 such that for any q=gpΓpq=g\cdot p\in\Gamma\cdot p, any ε<εp\varepsilon<\varepsilon_{p}, and any δ<δp\delta<\delta_{p}, we can find:

  • A cofinite subset TqT_{q} of the coset gΓpg\Gamma_{p},

  • Open neighborhoods Vq,WqV_{q},W_{q} of (Γ,)\partial(\Gamma,\mathcal{H}), with qVqWqq\in V_{q}\subset W_{q},

  • Open neighborhoods V^q,W^q\hat{V}_{q},\hat{W}_{q} of (Γ,)\partial(\Gamma,\mathcal{H}) with V^qW^q\hat{V}_{q}\subset\hat{W}_{q}

such that:

  1. (1)

    diam(Wq)<δ\mathrm{diam}(W_{q})<\delta, and ϕ1(Wq)N(ϕ1(q),ε)\phi^{-1}(W_{q})\subset N(\phi^{-1}(q),\varepsilon).

  2. (2)

    in (Γ,)\partial(\Gamma,\mathcal{H}), we have

    NΓ(V^q,δ)W^q.N_{\partial\Gamma}(\hat{V}_{q},\delta)\subset\hat{W}_{q}.
  3. (3)

    For every zVq{q}z\in V_{q}-\{q\}, there exists γTq\gamma\in T_{q} with γ1zV^q\gamma^{-1}\cdot z\in\hat{V}_{q}.

  4. (4)

    For every γTq\gamma\in T_{q}, we have

    NM(ϕ1(W^q),2ε)γ1NM(ϕ1(q),ε)N_{M}(\phi^{-1}(\hat{W}_{q}),2\varepsilon)\subset\gamma^{-1}N_{M}(\phi^{-1}(q),\varepsilon)

    and

    W^qγ1Wq.\hat{W}_{q}\subset\gamma^{-1}W_{q}.
  5. (5)

    NM(ϕ1(W^q),2ε)N_{M}(\phi^{-1}(\hat{W}_{q}),2\varepsilon) is contained in CpC_{p} and gΓpV^qg\Gamma_{p}\cdot\hat{V}_{q} contains (Γ,){q}\partial(\Gamma,\mathcal{H})-\{q\}.

Refer to caption
q=gpq=g\cdot p
VqV_{q}
WqW_{q}
zz
pp
V^q\hat{V}_{q}
W^q\hat{W}_{q}
γ1\gamma^{-1}
Figure 7. The behavior of sets in (Γ,)\partial(\Gamma,\mathcal{H}) described by 6.7. Given zVqz\in V_{q}, we pick an element γgΓp\gamma\in g\Gamma_{p} so that a uniformly large neighborhood of γ1z\gamma^{-1}z is contained in γ1Wq\gamma^{-1}W_{q}. We cannot pick γ1\gamma^{-1} to expand the metric everywhere close to qq—some points in VqV_{q} get contracted close to pp.
Remark 6.8.

If zVq{q}z\in V_{q}-\{q\} and γ1zV^q\gamma^{-1}z\in\hat{V}_{q} for some γTq\gamma\in T_{q}, we think of γ1\gamma^{-1} as “expanding” about zz. Conditions (1) and (2) imply that if γ1zVy\gamma^{-1}z\in V_{y} for some γTq\gamma\in T_{q}, then W^q\hat{W}_{q} contains WyW_{y}, and by condition (4), γ1Wq\gamma^{-1}W_{q} contains WyW_{y}. Here Vy,WyV_{y},W_{y} are the sets from either 6.5 or 6.7.

Proof.

Pick a compact set K(Γ,){p}K\subset\partial(\Gamma,\mathcal{H})-\{p\} so that ΓpK\Gamma_{p}\cdot K covers (Γ,){p}\partial(\Gamma,\mathcal{H})-\{p\}. Choose δp\delta_{p} small enough so that the closure of NΓ(K,2δp)N_{\partial\Gamma}(K,2\delta_{p}) does not contain pp. Then, for any δ<δp\delta<\delta_{p}, we can pick

V^q=NΓ(K,δ),W^q=NΓ(K,2δ).\hat{V}_{q}=N_{\partial\Gamma}(K,\delta),\quad\hat{W}_{q}=N_{\partial\Gamma}(K,2\delta).

We can choose εp\varepsilon_{p} sufficiently small so that a 2εp2\varepsilon_{p}-neighborhood of ϕ1(NΓ(K,2δp))\phi^{-1}(N_{\partial\Gamma}(K,2\delta_{p})) is contained in CpC_{p}. Now, fix ε<εp\varepsilon<\varepsilon_{p}. We claim that for a cofinite subset TqgΓpT_{q}\subset g\cdot\Gamma_{p}, for any γTq\gamma\in T_{q}, we have

(5) γW^q\displaystyle\gamma\cdot\hat{W}_{q}\subset BΓ(q,δ/2)\displaystyle B_{\partial\Gamma}(q,\delta/2)
(6) γNM(ϕ1(W^q),2ε)\displaystyle\gamma\cdot N_{M}(\phi^{-1}(\hat{W}_{q}),2\varepsilon)\subset NM(ϕ1(q),ε)\displaystyle N_{M}(\phi^{-1}(q),\varepsilon)

To see that this claim holds, it suffices to verify that for any infinite sequence γn\gamma_{n} of distinct group elements in gΓpg\Gamma_{p}, (5) and (6) both hold for all sufficiently large nn.

We write γn=gγn\gamma_{n}=g\cdot\gamma_{n}^{\prime} for γnΓp\gamma_{n}^{\prime}\in\Gamma_{p}. Then γn\gamma_{n}^{\prime} converges uniformly to pp on compact subsets of (Γ,){p}\partial(\Gamma,\mathcal{H})-\{p\}, so γn\gamma_{n} converges uniformly to qq on compact subsets of (Γ,){p}\partial(\Gamma,\mathcal{H})-\{p\}, implying that (5) eventually holds. And by our assumptions, we know that

γnNM(ϕ1(W^q),2ε)g1NM(ϕ1(q),ε)\gamma_{n}^{\prime}\cdot N_{M}(\phi^{-1}(\hat{W}_{q}),2\varepsilon)\subset g^{-1}\cdot N_{M}(\phi^{-1}(q),\varepsilon)

for sufficiently large nn, implying that (6) also eventually holds.

So we can take WqW_{q} to be the set

{q}γTqγW^q,\{q\}\cup\bigcup_{\gamma\in T_{q}}\gamma\cdot\hat{W}_{q},

and VqV_{q} to be the set

{q}γTqγV^q.\{q\}\cup\bigcup_{\gamma\in T_{q}}\gamma\cdot\hat{V}_{q}.

To see that WqW_{q} and VqV_{q} are open we just need to verify that they each contain a neighborhood of qq. Since V^q\hat{V}_{q} and W^q\hat{W}_{q} each contain KK, and ΓpK\Gamma_{p}\cdot K covers (Γ,){p}\partial(\Gamma,\mathcal{H})-\{p\}, VqV_{q} and WqW_{q} each contain the set

(Γ,)γgΓpTqγK.\partial(\Gamma,\mathcal{H})-\bigcup_{\gamma\in g\Gamma_{p}-T_{q}}\gamma K.

Since TqT_{q} is cofinite in gΓpg\Gamma_{p} this is an open set containing qq. ∎

6.1. Construction of the relative automaton

We will construct the relative automaton 𝒢\mathcal{G} satisfying the conditions of 6.1 by choosing a suitable open covering of (Γ,)\partial(\Gamma,\mathcal{H}), and then using compactness to take a finite subcover. The subsets of this subcover will be the vertices of 𝒢\mathcal{G}.

We choose constants ε>0\varepsilon>0, δ>0\delta>0 so that ε<εcon\varepsilon<\varepsilon_{\mathrm{con}}, δ<δcon\delta<\delta_{\mathrm{con}} (where εcon\varepsilon_{\mathrm{con}}, δcon\delta_{\mathrm{con}} are the constants coming from 6.5) and ε<εp\varepsilon<\varepsilon_{p}, δ<δp\delta<\delta_{p} for each pΠp\in\Pi (where εp,δp\varepsilon_{p},\delta_{p} are the constants coming from 6.7).

Then:

  • For each zconΓz\in\partial_{\mathrm{con}}\Gamma, we define WzW_{z}, VzV_{z}, γz\gamma_{z} as in 6.5, with parameters ε\varepsilon, δ\delta.

  • For each qparΓq\in\partial_{\mathrm{par}}\Gamma, we define Vq,Wq,V^q,W^qV_{q},W_{q},\hat{V}_{q},\hat{W}_{q}, and TqT_{q} as in 6.7, again with parameters ε\varepsilon, δ\delta.

The collections of sets {Vz:zconΓ}\{V_{z}:z\in\partial_{\mathrm{con}}\Gamma\} and {Vq:qparΓ}\{V_{q}:q\in\partial_{\mathrm{par}}\Gamma\} together give an open covering of the Bowditch boundary (Γ,)\partial(\Gamma,\mathcal{H}). So we choose a finite subcover 𝒱\mathcal{V}, which we can write as

𝒱={Va:aA}\mathcal{V}=\{V_{a}:a\in A\}

where AA is a finite subset of (Γ,)\partial(\Gamma,\mathcal{H}). We can in particular ensure that AA contains the finite set Π\Pi.

We identify the vertices of our Γ\Gamma-graph 𝒢\mathcal{G} with AA. For each aAa\in A, the set TaT_{a} is either {γa}\{\gamma_{a}\} (if aa is a conical limit point) or TqT_{q} (if a=qa=q for a parabolic point qq). Then, for each aAa\in A, we define the open sets UaU_{a} by

Ua=NM(ϕ1(a),ε).U_{a}=N_{M}(\phi^{-1}(a),\varepsilon).

The edges of the Γ\Gamma-graph 𝒢\mathcal{G} are defined as follows:

  • For a,bAa,b\in A with aconΓa\in\partial_{\mathrm{con}}\Gamma, there is an edge from aa to bb if (γa1Va)Vb(\gamma_{a}^{-1}\cdot V_{a})\cap V_{b} is nonempty.

  • If a,bAa,b\in A with aparΓa\in\partial_{\mathrm{par}}\Gamma, there is an edge from aa to bb if V^aVb\hat{V}_{a}\cap V_{b} is nonempty.

Since 𝒱\mathcal{V} is an open covering of (Γ,)\partial(\Gamma,\mathcal{H}), and the sets V^a\hat{V}_{a} and γa1Va\gamma_{a}^{-1}V_{a} are nonempty, every vertex of 𝒢\mathcal{G} has at least one outgoing edge. Moreover, for any parabolic point aa, the set V^a\hat{V}_{a} depends only on the orbit of aa in (Γ,)\partial(\Gamma,\mathcal{H}), so 𝒢\mathcal{G} must satisfy condition (2) in 6.1.

Proposition 6.9.

For each aAa\in A, we have

ϕ1(Wa)Ua.\phi^{-1}(W_{a})\subset U_{a}.
Proof.

When aa is not a parabolic vertex, Part (3) of 6.5 implies:

ϕ1(Wa)=γaγa1ϕ1(Wa)γaN(γa1ϕ1(Wa),2ε)NM(ϕ1(a),ε)=Ua.\phi^{-1}(W_{a})=\gamma_{a}\gamma_{a}^{-1}\phi^{-1}(W_{a})\subset\gamma_{a}N(\gamma_{a}^{-1}\phi^{-1}(W_{a}),2\varepsilon)\subset N_{M}(\phi^{-1}(a),\varepsilon)=U_{a}.

When aa is a parabolic vertex, then the claim follows directly from Part (1) of 6.7. ∎

Next we verify:

Proposition 6.10.

The collection of sets {Wv}\{W_{v}\} and {Uv}\{U_{v}\} are both 𝒢\mathcal{G}-compatible systems of open sets for the Γ\Gamma-graph 𝒢\mathcal{G}.

Proof.

First fix an edge (a,b)(a,b) with aconΓa\in\partial_{\mathrm{con}}\Gamma. Since (γa1Va)Vb(\gamma_{a}^{-1}V_{a})\cap V_{b} is nonempty, part 2 of 6.5 implies that γa1Wa\gamma_{a}^{-1}\cdot W_{a} contains the δ\delta-neighborhood of some point zVbz\in V_{b}. Since diam(Wb)<δ\mathrm{diam}(W_{b})<\delta and VbWbV_{b}\subset W_{b}, we can find a small ε>0\varepsilon^{\prime}>0 so that γaNΓ(Wb,ε)Wa\gamma_{a}N_{\partial\Gamma}(W_{b},\varepsilon^{\prime})\subset W_{a}.

In particular, γa1Wa\gamma_{a}^{-1}\cdot W_{a} contains bb, which means that NM(γa1ϕ1(Wa),2ε)N_{M}(\gamma_{a}^{-1}\phi^{-1}(W_{a}),2\varepsilon) contains NM(ϕ1(b),2ε)N_{M}(\phi^{-1}(b),2\varepsilon), which contains NM(Ub,ε)N_{M}(U_{b},\varepsilon). Then, part 3 of 6.5 implies that γaNM(Ub,ε)\gamma_{a}\cdot N_{M}(U_{b},\varepsilon) is contained in NM(ϕ1(a),ε)=UaN_{M}(\phi^{-1}(a),\varepsilon)=U_{a}.

Next fix an edge (q,b)(q,b) with qparΓq\in\partial_{\mathrm{par}}\Gamma. From part 2 of 6.7, we know that W^q\hat{W}_{q} contains the δ\delta-neighborhood of a point zV^qVbz\in\hat{V}_{q}\cap V_{b}. Since diam(Wb)<δ\mathrm{diam}(W_{b})<\delta and VbWbV_{b}\subset W_{b}, this means that W^q\hat{W}_{q} contains an ε\varepsilon^{\prime}-neighborhood of WbW_{b} for some small ε>0\varepsilon^{\prime}>0. So part 4 of 6.7 implies that for any γTq\gamma\in T_{q}, we have γN(Wb,ε)Wq\gamma\cdot N(W_{b},\varepsilon^{\prime})\subset W_{q}.

In particular W^q\hat{W}_{q} contains bb, so NM(ϕ1(W^q),2ε)N_{M}(\phi^{-1}(\hat{W}_{q}),2\varepsilon) contains NM(ϕ1(b),2ε)N_{M}(\phi^{-1}(b),2\varepsilon), which contains NM(Ub,ε)N_{M}(U_{b},\varepsilon). Then, part 4 of 6.7 implies that

γNM(Ub,ε)NM(ϕ1(q),ε)=Uq\gamma N_{M}(U_{b},\varepsilon)\subset N_{M}(\phi^{-1}(q),\varepsilon)=U_{q}

for any γTq\gamma\in T_{q}. ∎

We observe:

Proposition 6.11.

The 𝒢\mathcal{G}-compatible systems of open subsets {Uv}\{U_{v}\} and {Wv}\{W_{v}\} satisfy conditions (1) - (3) in 6.1.

Proof.

Part (1) follows directly from 6.9, and the fact that we defined each UaU_{a} to be the ε\varepsilon-neighborhood of ϕ1(a)\phi^{-1}(a). Part (2) is true by the construction of the open covering 𝒱\mathcal{V} and the graph 𝒢\mathcal{G}. Part (3) is true by construction and part (5) of 6.7. ∎

To finish the proof of 6.1, we now just need to show:

Proposition 6.12.

The Γ\Gamma-graph 𝒢\mathcal{G} is a relative quasigeodesic automaton for the pair (Γ,)(\Gamma,\mathcal{H}).

Proof.

We apply 5.13, using the 𝒢\mathcal{G}-compatible system {Wa}\{W_{a}\} and the sets {Va}\{V_{a}\} we defined in the construction of 𝒢\mathcal{G}.

The first three conditions of 5.13 are satisfied by construction. To see that conditions 4 and 5 hold, first observe that if zVaz\in V_{a} for a non-parabolic vertex aa, then γa1z\gamma_{a}^{-1}\cdot z lies in some VbV_{b} and (a,b)(a,b) is an edge in 𝒢\mathcal{G}. And if zVa{pa}z\in V_{a}-\{p_{a}\} for a parabolic vertex aa, then part (3) of 6.7 says that there is some γTa\gamma\in T_{a} such that γ1zV^a\gamma^{-1}\cdot z\in\hat{V}_{a}. If VbV_{b} contains γ1z\gamma^{-1}\cdot z, the edge (a,b)(a,b) must be in 𝒢\mathcal{G}.

It only remains to check that 𝒢\mathcal{G} is a divergent Γ\Gamma-graph. Let {αn}\{\alpha_{n}\} be an infinite 𝒢\mathcal{G}-path, following a vertex path {vn}\{v_{n}\}. The 𝒢\mathcal{G}-compatibility condition implies that γnU¯vn+1\gamma_{n}\overline{U}_{v_{n+1}} is a subset of γn1Uvn\gamma_{n-1}U_{v_{n}} for every nn. Since MM is connected and each UvU_{v} is a proper subset of MM, this inclusion must be proper. This implies that in the sequence γn\gamma_{n}, no element can appear more than #A\#A times and therefore γn\gamma_{n} is divergent. ∎

Remark 6.13.

This last step is the only part of the proof of 6.1 which uses the connectedness of MM. This hypothesis is likely unnecessary, but omitting it would involve introducing additional technicalities in the construction of the sets VaV_{a}, WaW_{a}—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 M=(Γ,)M=\partial(\Gamma,\mathcal{H})). As stated, the proposition only shows that such an automaton exists when (Γ,)\partial(\Gamma,\mathcal{H}) 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 ZZ of the Bowditch boundary (Γ,)\partial(\Gamma,\mathcal{H}). Then, for any open set UMU\subset M containing ϕ1(Z)\phi^{-1}(Z), there is a relative quasigeodesic automaton 𝒢\mathcal{G} and a pair of 𝒢\mathcal{G}-compatible systems of open sets {Ua}\{U_{a}\}, {Wa}\{W_{a}\} as in 6.1, additionally satisfying the following: any zZz\in Z is the limit of a 𝒢\mathcal{G}-path {αn}\{\alpha_{n}\} (with corresponding vertex path {vn}\{v_{n}\}) such that Uv1UU_{v_{1}}\subset U.

Proof.

We choose ε>0\varepsilon>0 so that UU contains NM(ϕ1(Z),ε)N_{M}(\phi^{-1}(Z),\varepsilon). We then construct our relative quasigeodesic automaton 𝒢\mathcal{G} as in the proof of 6.1, but we also choose a finite subset AZZA_{Z}\subset Z so that the sets VaV_{a} for aAZa\in A_{Z} give a finite open covering of ZZ. We can ensure that the vertex set AA of 𝒢\mathcal{G} contains AZA_{Z}.

Then, for any zZz\in Z, by the construction in 5.13, we can find a 𝒢\mathcal{G}-path limiting to zz whose first vertex is some aAZa\in A_{Z}. The corresponding open set for this vertex is Ua=NM(ϕ1(a),ε)UU_{a}=N_{M}(\phi^{-1}(a),\varepsilon)\subset U. ∎

Proposition 6.15.

For each parabolic point pΠp\in\Pi, let KpK_{p} be a compact subset of (Γ,){p}\partial(\Gamma,\mathcal{H})-\{p\} such that ΓpKp=(Γ,){p}\Gamma_{p}\cdot K_{p}=\partial(\Gamma,\mathcal{H})-\{p\}. Then the relative quasigeodesic automaton in 6.1 can be chosen to satisfy the following:

For every parabolic vertex ww with pw=pΠp_{w}=p\in\Pi, and every zKpz\in K_{p}, there is a 𝒢\mathcal{G}-path limiting to zz whose first vertex uu is connected to ww by an edge (w,u)(w,u).

Proof.

The proof of 6.7 shows that in our construction of the relative automaton, we can ensure that each set V^p\hat{V}_{p} contains KpK_{p}. So if zKpz\in K_{p}, then by definition of the automaton, zz lies in zz lies in VuV_{u} with ww connected to uu by a directed edge. Then, following the proof of 5.13, we can find a 𝒢\mathcal{G}-path limiting to zz whose first vertex is uu. ∎

7. Contracting paths in flag manifolds

Let ΓG\Gamma\subset G be a discrete relatively hyperbolic group, and let 𝒢\mathcal{G} be a Γ\Gamma-graph. Fix a pair of opposite parabolic subgroups P+P^{+}, PP^{-}. Our goal in this section is to show that under certain conditions, if {Uv}\{U_{v}\} is a 𝒢\mathcal{G}-compatible system of open subsets of G/P+G/P^{+} for the action of Γ\Gamma on G/P+G/P^{+}, then the sequence of group elements lying along an infinite 𝒢\mathcal{G}-path is P+P^{+}-divergent.

7.1. Contracting paths in Γ\Gamma-graphs

Definition 7.1.

Let Γ\Gamma be a discrete subgroup of GG, let 𝒢\mathcal{G} be a Γ\Gamma-graph, and let {Uv}vV(𝒢)\{U_{v}\}_{v\in V(\mathcal{G})} be a 𝒢\mathcal{G}-compatible system of open subsets of G/P+G/P^{+}. We say that a 𝒢\mathcal{G}-path {αn}n\{\alpha_{n}\}_{n\in\mathbb{N}} is contracting if the decreasing intersection

(7) n=1α1αnUvn+1\bigcap_{n=1}^{\infty}\alpha_{1}\cdots\alpha_{n}\cdot U_{v_{n+1}}

is a singleton in G/P+G/P^{+}.

Definition 7.2.

We say that an open set ΩG/P+\Omega\subset G/P^{+} is a proper domain if the closure of Ω\Omega lies in an affine chart Opp(ξ)G/P+\operatorname{Opp}(\xi)\subset G/P^{+} for some ξG/P\xi\in G/P^{-}.

Here is the main result in this section:

Proposition 7.3.

Let 𝒢\mathcal{G} be a Γ\Gamma-graph for (Γ,)(\Gamma,\mathcal{H}), and let {Uv}vV(𝒢)\{U_{v}\}_{v\in V(\mathcal{G})} be 𝒢\mathcal{G}-compatible system of open subsets of G/P+G/P^{+}.

If the set UvU_{v} is a proper domain for each vertex vv of the automaton, then every infinite 𝒢\mathcal{G}-path is contracting.

7.2. A metric property for bounded domains in flag manifolds

To prove 7.3, we consider a metric CΩC_{\Omega} defined by Zimmer [Zim18] on any proper domain ΩG/P+\Omega\subset G/P^{+}. CΩC_{\Omega} is defined so that it is invariant under the action of GG on G/P+G/P^{+}: for any x,yx,y in some proper domain ΩG/P+\Omega\subset G/P^{+}, and any gGg\in G, we have

(8) CΩ(x,y)=CgΩ(gx,gy).C_{\Omega}(x,y)=C_{g\Omega}(gx,gy).

In general, CΩC_{\Omega} is not a complete metric. However, CΩC_{\Omega} induces the standard topology on Ω\Omega as an open subset of G/PG/P. We will show that for a 𝒢\mathcal{G}-path {αn}\{\alpha_{n}\}, the diameter of

α1αnUvn+1\alpha_{1}\cdots\alpha_{n}U_{v_{n+1}}

with respect to CUv1C_{U_{v_{1}}} tends to zero as nn\to\infty.

Zimmer’s construction of CΩC_{\Omega} depends on an irreducible representation ζ:GPGL(V)\zeta:G\to\operatorname{PGL}(V) for some real vector space VV. 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 VV, an irreducible representation ζ:GPGL(V)\zeta:G\to\operatorname{PGL}(V), a line V\ell\subset V, and a hyperplane HVH\subset V such that:

  1. (1)

    +H=V\ell+H=V.

  2. (2)

    The stabilizer of \ell in GG is P+P^{+} and the stabilizer of HH in GG is PP^{-}.

  3. (3)

    gP+g1gP^{+}g^{-1} and hPh1hP^{-}h^{-1} are opposite if and only if ζ(g)\zeta(g)\ell and ζ(h)H\zeta(h)H are transverse.

The representation ζ\zeta determines a pair of embeddings ι:G/P+(V)\iota:G/P^{+}\to\mathbb{P}(V) and ι:G/P(V)\iota^{*}:G/P^{-}\to\mathbb{P}(V^{*}) by

ι(gP+)=ζ(g),ι(gP)=ζ(g)H.\iota(gP^{+})=\zeta(g)\ell,\qquad\iota^{*}(gP^{-})=\zeta(g)H.

In this section, we will identify (V)\mathbb{P}(V^{*}) with the space of projective hyperplanes in (V)\mathbb{P}(V), by identifying the projectivization of a functional wVw\in V^{*} with the projectivization of its kernel.

Definition 7.5.

Let Ω\Omega be an open subset of G/P+G/P^{+}. The dual domain ΩG/P\Omega^{*}\subset G/P^{-} is

Ω={νG/P:ν is opposite to ξ for every ξΩ¯}.\Omega^{*}=\{\nu\in G/P^{-}:\nu\textrm{ is opposite to }\xi\textrm{ for every }\xi\in\overline{\Omega}\}.

Note that Ω\Omega^{*} is open if and only if Ω\Omega is a proper domain.

Definition 7.6.

Let w1,w2(V)w_{1},w_{2}\in\mathbb{P}(V^{*}), and let z1,z2(V)z_{1},z_{2}\in\mathbb{P}(V). The cross-ratio [w1,w2;z1,z2][w_{1},w_{2};z_{1},z_{2}] is defined by

w~1(z~2)w~2(z~1)w~1(z~1)w~2(z~2),\frac{\tilde{w}_{1}(\tilde{z}_{2})\tilde{w}_{2}(\tilde{z}_{1})}{\tilde{w}_{1}(\tilde{z}_{1})\tilde{w}_{2}(\tilde{z}_{2})},

where w~i\tilde{w}_{i}, z~i\tilde{z}_{i} are respectively lifts of wiw_{i} and ziz_{i} in VV^{*} and VV.

Remark 7.7.

When VV is two-dimensional, we can identify the projective line (V)\mathbb{P}(V^{*}) with (V)\mathbb{P}(V) by identifying each [w](V)[w]\in\mathbb{P}(V^{*}) with [ker(w)](V)[\ker(w)]\in\mathbb{P}(V). In that case, the cross-ratio defined above agrees with the standard four-point cross-ratio on P1\mathbb{R}\mathrm{P}^{1}, given by

(9) [a,b;c,d]:=(da)(cb)(ca)(db).[a,b;c,d]:=\frac{(d-a)(c-b)}{(c-a)(d-b)}.

The differences in (9) can be measured in any affine chart in P1\mathbb{R}\mathrm{P}^{1} containing a,b,c,da,b,c,d. Our convention is chosen so that if we identify P1\mathbb{R}\mathrm{P}^{1} with {}\mathbb{R}\cup\{\infty\}, we have [0,;1,z]=z[0,\infty;1,z]=z.

Definition 7.8.

Let ΩG/P+\Omega\subset G/P^{+} be a proper domain. We define the function CΩ:Ω×ΩC_{\Omega}:\Omega\times\Omega\to\mathbb{R} by

CΩ(x,y)=supξ1,ξ2Ωlog|[ι(ξ1),ι(ξ2);ι(x),ι(y)]|.C_{\Omega}(x,y)=\sup_{\xi_{1},\xi_{2}\in\Omega^{*}}\log|[\iota^{*}(\xi_{1}),\iota^{*}(\xi_{2});\iota(x),\iota(y)]|.

For any gGg\in G and any proper domain ΩG/P+\Omega\subset G/P^{+}, we have (gΩ)=(gΩ)(g\Omega)^{*}=(g\Omega^{*}). So CΩC_{\Omega} must satisfy the GG-invariance condition (8).

If Ω\Omega is a properly convex subset of (V)\mathbb{P}(V), and ζ\zeta, ι\iota, ι\iota^{*} are the identity maps on PGL(V)\operatorname{PGL}(V), (V)\mathbb{P}(V), and (V)\mathbb{P}(V^{*}) respectively, then CΩC_{\Omega} agrees with the well-studied Hilbert metric on Ω\Omega. More generally we have:

Theorem 7.9 ([Zim18], Theorem 5.2).

If Ω\Omega is open and bounded in an affine chart, then CΩC_{\Omega} is a metric on Ω\Omega which induces the standard topology on Ω\Omega as an open subset of G/P+G/P^{+}.

Remark 7.10.

This particular result in [Zim18] is stated only for noncompact simple Lie groups, but the proof only assumes that GG is semisimple with no compact factor.

Since taking duals of proper domains reverses inclusions, it follows that if Ω1Ω2\Omega_{1}\subset\Omega_{2}, then CΩ1CΩ2C_{\Omega_{1}}\geq C_{\Omega_{2}}. Our goal now is to sharpen this inequality, and show:

Proposition 7.11.

Let Ω1\Omega_{1}, Ω2\Omega_{2} be proper domains in G/P+G/P^{+}, such that Ω1¯Ω2\overline{\Omega_{1}}\subset\Omega_{2}.

There exists a constant λ>1\lambda>1 (depending on Ω1\Omega_{1} and Ω2\Omega_{2}) so that for all x,yΩ1x,y\in\Omega_{1},

CΩ1(x,y)λCΩ2(x,y).C_{\Omega_{1}}(x,y)\geq\lambda\cdot C_{\Omega_{2}}(x,y).

A consequence is the following, which in particular implies 7.3.

Corollary 7.12.

Let 𝒢\mathcal{G} be a Γ\Gamma-graph for a relatively hyperbolic group Γ\Gamma, and let {Uv}\{U_{v}\} be a 𝒢\mathcal{G}-compatible system of open subsets of G/P+G/P^{+}. If each UvU_{v} is a proper domain, then there are constants λ1,λ2>0\lambda_{1},\lambda_{2}>0 so that for any 𝒢\mathcal{G}-path {αn}\{\alpha_{n}\} in the Γ\Gamma-graph 𝒢\mathcal{G}, the diameter of

α1αnUvn+1\alpha_{1}\cdots\alpha_{n}\cdot U_{v_{n+1}}

with respect to CUv1C_{U_{v_{1}}} is at most

λ1exp(λ2n).\lambda_{1}\cdot\exp(-\lambda_{2}\cdot n).
Proof.

For any open set UG/P+U\subset G/P^{+} and AUA\subset U, we let diamU(A)\mathrm{diam}_{U}(A) denote the diameter of AA with respect to the metric CUC_{U}. We choose a uniform ε>0\varepsilon>0 so that in some fixed metric on G/P+G/P^{+}, every edge (v,w)(v,w) in 𝒢\mathcal{G}, and every αTv\alpha\in T_{v}, we have

αN(Uw,ε)Uv.\alpha N(U_{w},\varepsilon)\subset U_{v}.

Then for each vertex set UvU_{v}, we write Uvε=N(Uv,ε)U_{v}^{\varepsilon}=N(U_{v},\varepsilon).

We take

λ1=max{diamUvε(Uv)}.\lambda_{1}=\max\{\mathrm{diam}_{U_{v}^{\varepsilon}}(U_{v})\}.

7.11 implies that there exists λv>0\lambda_{v}>0 such that for all x,yUvx,y\in U_{v}, we have

CUv(x,y)exp(λv)CUvε(x,y).C_{U_{v}}(x,y)\geq\exp(\lambda_{v})\cdot C_{U_{v}^{\varepsilon}}(x,y).

Take λ2=minv{λv}\lambda_{2}=\min_{v}\{\lambda_{v}\}. We claim that for all n1n\geq 1, we have

diamU1ε(α1αnUvn)λ1exp(λ2(n1)).\mathrm{diam}_{U_{1}^{\varepsilon}}(\alpha_{1}\cdots\alpha_{n}U_{v_{n}})\leq\lambda_{1}\exp(-\lambda_{2}\cdot(n-1)).

We prove the claim via induction on the length of the 𝒢\mathcal{G}-path {αn}\{\alpha_{n}\}. For n=1n=1, the claim is true because α1Uv2Uv1\alpha_{1}U_{v_{2}}\subset U_{v_{1}}. For n>1n>1, we can assume

λ1exp(λ2(n2))diamUv2ε(α2αnUvn+1).\lambda_{1}\exp(-\lambda_{2}(n-2))\geq\mathrm{diam}_{U_{v_{2}}^{\varepsilon}}(\alpha_{2}\cdots\alpha_{n}\cdot U_{v_{n+1}}).

Then we have

diamUv2ε(α2αnUvn+1)\displaystyle\mathrm{diam}_{U_{v_{2}}^{\varepsilon}}(\alpha_{2}\cdots\alpha_{n}\cdot U_{v_{n+1}}) =diamα1Uv2ε(α1αnUvn+1)\displaystyle=\mathrm{diam}_{\alpha_{1}U_{v_{2}}^{\varepsilon}}(\alpha_{1}\cdots\alpha_{n}\cdot U_{v_{n+1}})
diamUv1(α1αnUvn+1)\displaystyle\geq\mathrm{diam}_{U_{v_{1}}}(\alpha_{1}\cdots\alpha_{n}\cdot U_{v_{n+1}})
exp(λ2)diamUv1ε(α1αnUvn+1).\displaystyle\geq\exp(\lambda_{2})\cdot\mathrm{diam}_{U_{v_{1}}^{\varepsilon}}(\alpha_{1}\cdots\alpha_{n}\cdot U_{v_{n+1}}).

Finally, the claim implies the corollary because we know that

diamU1(α1αnUn+1)\displaystyle\mathrm{diam}_{U_{1}}(\alpha_{1}\cdots\alpha_{n}U_{n+1}) diamα1U2ε(α1αnUn+1)\displaystyle\leq\mathrm{diam}_{\alpha_{1}U_{2}^{\varepsilon}}(\alpha_{1}\cdots\alpha_{n}U_{n+1})
=diamU2ε(α2αnUn+1)\displaystyle=\mathrm{diam}_{U_{2}^{\varepsilon}}(\alpha_{2}\cdots\alpha_{n}U_{n+1})
λ1exp(λ2(n2)).\displaystyle\leq\lambda_{1}\exp(\lambda_{2}(n-2)).

So, we can replace λ1\lambda_{1} with λ1exp(2λ2)\lambda_{1}\exp(-2\lambda_{2}) to get the desired result. ∎

We now proceed with the proof of 7.11. We first observe that in the special case where Ω1,Ω2\Omega_{1},\Omega_{2} are properly convex subsets of real projective space, one can show the desired result essentially via the following:

Proposition 7.13.

Let a,b,c,da,b,c,d be points in P1\mathbb{R}\mathrm{P}^{1}, arranged so that a<b<c<daa<b<c<d\leq a in a cyclic ordering on P1\mathbb{R}\mathrm{P}^{1}. Then there exists a constant λ>1\lambda>1, depending only on the cross-ratio [a,b;c,d][a,b;c,d], so that for all distinct x,y(b,c)x,y\in(b,c), we have

|log[b,c;x,y]|λ|log[a,d;x,y]|.|\log[b,c;x,y]|\geq\lambda\cdot|\log[a,d;x,y]|.

7.13 is a standard fact in real projective geometry and can be verified by a computation. Note that we allow the degenerate case a=da=d: in this situation the right-hand side is identically zero for distinct x,y(b,c)x,y\in(b,c). We allow no other equalities among a,b,c,da,b,c,d, so the cross-ratio [a,b;c,d][a,b;c,d] lies in {1}\mathbb{R}-\{1\}.

To apply 7.13 to our situation, we need to get some control on the behavior of the embeddings ι:G/P+(V)\iota:G/P^{+}\to\mathbb{P}(V) and ι:G/P(V)\iota^{*}:G/P^{-}\to\mathbb{P}(V^{*}). We do so in the next three lemmas below.

Lemma 7.14.

Let x,yx,y be distinct points in G/P+G/P^{+}. There exists a one-parameter subgroup gtGg_{t}\subset G such that ζ(gt)\zeta(g_{t}) fixes ι(x)\iota(x) and ι(y)\iota(y), and acts nontrivially on the projective line LxyL_{xy} spanned by ι(x)\iota(x) and ι(y)\iota(y).

Proof.

We can write x=gP+x=gP^{+} for some gGg\in G. Let 𝔞\mathfrak{a} denote an abelian subalgebra of the Lie algebra 𝔤\mathfrak{g} of GG, such that for a maximal compact KGK\subset G, the exponential map 𝔞G\mathfrak{a}\to G induces an isometric embedding 𝔞G/K\mathfrak{a}\to G/K whose image is a maximal flat in G/KG/K.

There is a conjugate 𝔞\mathfrak{a}^{\prime} of 𝔞\mathfrak{a} such that the action of exp(𝔞)\exp(\mathfrak{a}^{\prime}) on G/P+G/P^{+} fixes both xx and yy (see [Ebe96], Proposition 2.21.14). So, up to the action of GG on G/P+G/P^{+}, we can assume that xx is fixed by a standard parabolic subgroup Pθ+P^{+}_{\theta} conjugate to P+P^{+}, and that x,yx,y are both fixed by the subgroup exp(𝔞)\exp(\mathfrak{a}).

We choose Z𝔞+Z\in\mathfrak{a}^{+} so that α(Z)0\alpha(Z)\neq 0 for all αθ\alpha\in\theta. Then gt=exp(tZ)g_{t}=\exp(tZ) is a 1-parameter subgroup of GG fixing xx. As t+t\to+\infty, gtg_{t} is Pθ+P^{+}_{\theta}-divergent, with unique attracting fixed point xx.

Then [GGKW17], Lemma 3.7 implies that ζ(gt)\zeta(g_{t}) is P1P_{1}-divergent, where P1P_{1} is the stabilizer of a line in VV, and ι(x)\iota(x) is the unique one-dimensional eigenspace of ζ(gt)\zeta(g_{t}) whose eigenvalue has largest modulus. And, since ζ(gt)\zeta(g_{t}) fixes ι(x)\iota(x) and ι(y)\iota(y), ζ(gt)\zeta(g_{t}) preserves LxyL_{xy}, and acts nontrivially since the eigenvalues of ζ(gt)\zeta(g_{t}) on ι(x)\iota(x) and ι(y)\iota(y) must be distinct. ∎

Lemma 7.15.

Let LL be any projective line in (V)\mathbb{P}(V) tangent to the image of the embedding ι:G/P+(V)\iota:G/P^{+}\to\mathbb{P}(V) at a point ι(x)\iota(x) for xG/P+x\in G/P^{+}. There exists a one-parameter subgroup gtg_{t} of GG so that ζ(gt)\zeta(g_{t}) acts nontrivially on LL with unique fixed point ι(x)\iota(x).

Proof.

Fix a sequence ynG/P+y_{n}\in G/P^{+} such that ynxy_{n}\neq x and the projective line LnL_{n} spanned by ι(x)\iota(x) and ι(yn)\iota(y_{n}) converges to LL. By 7.14, there exists Zn𝔤Z_{n}\in\mathfrak{g} so that ζ(exp(tZn))\zeta(\exp(tZ_{n})) acts nontrivially on LnL_{n}, with fixed points ι(x)\iota(x) and ι(yn)\iota(y_{n}).

In the projectivization (𝔤)\mathbb{P}(\mathfrak{g}), [Zn][Z_{n}] converges to some [Z][Z]. Since ζ:GPGL(V)\zeta:G\to\operatorname{PGL}(V) has finite kernel, there is an induced map ζ:(𝔤)(𝔰𝔩(V))\zeta:\mathbb{P}(\mathfrak{g})\to\mathbb{P}(\mathfrak{sl}(V)), which satisfies

ζ([Zn])ζ([Z]).\zeta([Z_{n}])\to\zeta([Z]).

A continuity argument shows that the one-parameter subgroup ζ(exp(tZ))\zeta(\exp(tZ)) acts nontrivially on the line LL, and has unique fixed point at ι(x)\iota(x). ∎

Lemma 7.16.

Let ΩG/P+\Omega\subset G/P^{+} be a proper domain, and let LL be a projective line in (V)\mathbb{P}(V) which is either spanned by two points in ι(Ω)\iota(\Omega), or is tangent to ι(G/P+)\iota(G/P^{+}) at a point ι(x)\iota(x) for xΩx\in\Omega. Then

WL={vL:v=ι(ξ)L for ξΩ}W_{L}=\{v\in L:v=\iota^{*}(\xi)\cap L\textrm{ for }\xi\in\Omega^{*}\}

is a nonempty open subset of LL.

Proof.

WLW_{L} is nonempty since Ω\Omega^{*} is nonempty. So let vWLv\in W_{L}, and choose ξΩ\xi\in\Omega^{*} so that ι(ξ)L=v\iota^{*}(\xi)\cap L=v. We need to show that an open interval ILI\subset L containing vv is contained in WLW_{L}.

If LL is spanned by x,yι(Ω)x,y\in\iota(\Omega), then 7.14 implies that we can find a one-parameter subgroup gtGg_{t}\in G such that ζ(gt)\zeta(g_{t}) fixes xx and yy, and acts nontrivially on LL. Since Ω\Omega^{*} is open, we can find ε>0\varepsilon>0 so that gtξΩg_{t}\cdot\xi\in\Omega^{*} for t(ε,ε)t\in(-\varepsilon,\varepsilon). Since xx and yy are in ι(Ω)\iota(\Omega), ι(ξ)\iota^{*}(\xi) is transverse to both xx and yy, so we have vxv\neq x, vyv\neq y. Then as tt varies from ε-\varepsilon to ε\varepsilon,

ι(gtξ)L=ζ(gt)v\iota^{*}(g_{t}\cdot\xi)\cap L=\zeta(g_{t})\cdot v

gives an open interval in WLW_{L} containing vv.

A similar argument using 7.15 shows that the claim also holds if LL is tangent to ι(Ω)\iota(\Omega). ∎

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 Ω1,Ω2\Omega_{1},\Omega_{2} be proper domains in G/P+G/P^{+}, with Ω1¯Ω2\overline{\Omega_{1}}\subset\Omega_{2}, and let KΩ1K\subset\Omega_{1} be compact. There exists a constant λ>1\lambda>1 such that for all x,yKx,y\in K,

CΩ1(x,y)λCΩ2(x,y).C_{\Omega_{1}}(x,y)\geq\lambda\cdot C_{\Omega_{2}}(x,y).
Proof.

Since KK is compact, it suffices to show that for fixed xΩ1x\in\Omega_{1}, the ratio

CΩ1(x,y)CΩ2(x,y)\frac{C_{\Omega_{1}}(x,y)}{C_{\Omega_{2}}(x,y)}

is bounded below by some λ>1\lambda>1 as yy varies in K{x}K-\{x\}.

Let yK{x}y\in K-\{x\}, and let LxyL_{xy} denote the projective line spanned by ι(x)\iota(x) and ι(y)\iota(y). Choose ξ,ηΩ2¯\xi,\eta\in\overline{\Omega_{2}^{*}} so that

CΩ2(x,y)=log|[ι(ξ),ι(η);ι(x),ι(y)]|.C_{\Omega_{2}}(x,y)=\log|[\iota^{*}(\xi),\iota^{*}(\eta);\iota(x),\iota(y)]|.

That is, if v=ι(ξ)Lxyv=\iota^{*}(\xi)\cap L_{xy}, w=ι(η)Lxyw=\iota^{*}(\eta)\cap L_{xy}, we have

CΩ2(x,y)=log|[v,w;ι(x),ι(y)]|=log|vι(y)||wι(x)||vι(x)||wι(y)|,C_{\Omega_{2}}(x,y)=\log|[v,w;\iota(x),\iota(y)]|=\log\frac{|v-\iota(y)|\cdot|w-\iota(x)|}{|v-\iota(x)|\cdot|w-\iota(y)|},

where the distances are measured in any identification of LxyL_{xy} with P1={}\mathbb{R}\mathrm{P}^{1}=\mathbb{R}\cup\{\infty\}.

We can choose an identification of LxyL_{xy} with {}\mathbb{R}\cup\{\infty\} so that either v<ι(x)<ι(y)<wv<\iota(x)<\iota(y)<w or v<ι(x)<w<ι(y)v<\iota(x)<w<\iota(y). In either case, for any v(v,ι(x))Lxyv^{\prime}\in(v,\iota(x))\subset L_{xy}, we have

log|[v,w;ι(x),ι(y)]|>log|[v,w;ι(x),ι(y)]|.\log|[v^{\prime},w;\iota(x),\iota(y)]|>\log|[v,w;\iota(x),\iota(y)]|.

We know that Ω2¯Ω1\overline{\Omega_{2}^{*}}\subset\Omega_{1}^{*}, so ξ,η\xi,\eta lie in Ω1\Omega_{1}^{*}. Then 7.16 implies that there exists ξΩ1\xi^{\prime}\in\Omega_{1}^{*} so that v=ι(ξ)Lxyv^{\prime}=\iota^{*}(\xi^{\prime})\cap L_{xy} lies in the interval (v,ι(x))Lxy(v,\iota(x))\subset L_{xy}. See Figure 8.

Refer to caption
ι(x)\iota(x)
ι(y)\iota(y)
ι(ξ)\iota^{*}(\xi)
ι(η)\iota^{*}(\eta)
ι(ξ)\iota^{*}(\xi^{\prime})
Refer to caption
LxyL_{xy}
Refer to caption
vv
ww
vv^{\prime}
ι(Ω1)\iota(\Omega_{1})
ι(Ω2)\iota(\Omega_{2})
Refer to caption
ι(x)\iota(x)
ι(y)\iota(y)
ι(ξ)\iota^{*}(\xi)
ι(η)\iota^{*}(\eta)
ι(ξ)\iota^{*}(\xi^{\prime})
Refer to caption
LxyL_{xy}
Refer to caption
vv
ww
vv^{\prime}
ι(Ω1)\iota(\Omega_{1})
ι(Ω2)\iota(\Omega_{2})
Figure 8. We can always find ξΩ1\xi^{\prime}\in\Omega_{1}^{*} close to ξ\xi so that the absolute value of the cross-ratio [ι(ξ),ι(ν);ι(x),ι(y)][\iota^{*}(\xi),\iota^{*}(\nu);\iota(x),\iota(y)] increases when we replace ξ\xi with ξ\xi^{\prime}. In particular this is possible even when the sets ι(Ω1)\iota(\Omega_{1}), ι(Ω2)\iota(\Omega_{2}) fail to be convex (left) or even connected (right).

Then, we have

CΩ1(x,y)\displaystyle C_{\Omega_{1}}(x,y) log|[ι(ξ),ι(η);ι(x),ι(y)]|\displaystyle\geq\log|[\iota^{*}(\xi^{\prime}),\iota^{*}(\eta);\iota(x),\iota(y)]|
=log|[v,w;ι(x),ι(y)]\displaystyle=\log|[v^{\prime},w;\iota(x),\iota(y)]
>log|[v,w;ι(x),ι(y)]\displaystyle>\log|[v,w;\iota(x),\iota(y)]
=CΩ2(x,y).\displaystyle=C_{\Omega_{2}}(x,y).

This shows that CΩ1(x,y)CΩ2(x,y)>1\frac{C_{\Omega_{1}}(x,y)}{C_{\Omega_{2}}(x,y)}>1 for all yK{x}y\in K-\{x\}. We still need to find some uniform λ>1\lambda>1 so that CΩ1(x,y)CΩ2(x,y)λ\frac{C_{\Omega_{1}}(x,y)}{C_{\Omega_{2}}(x,y)}\geq\lambda for all yK{x}y\in K-\{x\}. To see this, suppose for the sake of a contradiction that for a sequence ynK{x}y_{n}\in K-\{x\}, we have

(10) CΩ1(x,yn)CΩ2(x,yn)1.\frac{C_{\Omega_{1}}(x,y_{n})}{C_{\Omega_{2}}(x,y_{n})}\to 1.

Since KK is compact, yny_{n} must converge to xx. Up to subsequence, the sequence of projective lines LnL_{n} spanned by ι(x)\iota(x) and ι(yn)\iota(y_{n}) converges to a line LL tangent to ι(G/P+)\iota(G/P^{+}) at ι(x)\iota(x).

For each yny_{n}, choose ξn\xi_{n}, ηnΩ2¯\eta_{n}\in\overline{\Omega_{2}^{*}} so that

CΩ2(x,yn)=log|[ι(ξn),ι(ηn);ι(x),ι(yn)]|.C_{\Omega_{2}}(x,y_{n})=\log|[\iota^{*}(\xi_{n}),\iota^{*}(\eta_{n});\iota(x),\iota(y_{n})]|.

Let vn=ι(ξn)Lnv_{n}=\iota^{*}(\xi_{n})\cap L_{n}, wn=ι(ηn)Lnw_{n}=\iota^{*}(\eta_{n})\cap L_{n}. Then up to subsequence ξn\xi_{n} converges to ξΩ2¯\xi\in\overline{\Omega_{2}^{*}}, ηn\eta_{n} converges to ηΩ2¯\eta\in\overline{\Omega_{2}^{*}}, and vnv_{n}, wnw_{n} respectively converge to v=ι(ξ)Lv=\iota^{*}(\xi)\cap L, w=ι(η)Lw=\iota^{*}(\eta)\cap L.

Since xx is in Ω2\Omega_{2}, ι(ξ)\iota^{*}(\xi) and ι(η)\iota^{*}(\eta) are both transverse to ι(x)\iota(x)—so in particular xwx\neq w and xvx\neq v (although a priori we could have v=wv=w).

Since ξΩ2¯Ω1\xi\in\overline{\Omega_{2}^{*}}\subset\Omega_{1}^{*}, 7.16 implies that there exist ξ,ηΩ1\xi^{\prime},\eta^{\prime}\in\Omega_{1}^{*} so that for some identification of LL with {}\mathbb{R}\cup\{\infty\}, we have

v<ι(ξ)L<ι(x)<ι(η)L<w.v<\iota^{*}(\xi^{\prime})\cap L<\iota(x)<\iota^{*}(\eta^{\prime})\cap L<w.

Note that this is possible even if v=wv=w, because then we can just identify both vv and ww with \infty. Let vn=ι(ξ)Lnv_{n}^{\prime}=\iota^{*}(\xi^{\prime})\cap L_{n}, and let wn=ι(η)Lnw_{n}^{\prime}=\iota^{*}(\eta^{\prime})\cap L_{n}. Respectively, vnv_{n}^{\prime} and wnw_{n}^{\prime} converge to v=ι(ξ)Lv^{\prime}=\iota^{*}(\xi^{\prime})\cap L and w=ι(η)Lw^{\prime}=\iota^{*}(\eta^{\prime})\cap L.

This means that the cross-ratios [vn,vn;wn,wn][v_{n},v_{n}^{\prime};w_{n}^{\prime},w_{n}] converge to [v,v;w,w]{1}[v,v^{\prime};w^{\prime},w]\in\mathbb{R}-\{1\}, and in particular are bounded away from both \infty and 11 for all nn.

We choose identifications of LnL_{n} with {}\mathbb{R}\cup\{\infty\} so that vn<vn<ι(x)<wn<wnv_{n}<v_{n}^{\prime}<\iota(x)<w_{n}^{\prime}<w_{n}. Since yny_{n} converges to xx, for all sufficiently large nn, we have vn<ι(yn)<wnv_{n}^{\prime}<\iota(y_{n})<w_{n}^{\prime}. Then, 7.13 implies that for all nn, we have

log|[vn,ι(x),ι(yn),wn]|λlog|[vn,ι(x),ι(yn),wn]|\log|[v_{n}^{\prime},\iota(x),\iota(y_{n}),w_{n}^{\prime}]|\geq\lambda\cdot\log|[v_{n},\iota(x),\iota(y_{n}),w_{n}]|

for some λ>1\lambda>1 independent of nn. But then since

CΩ1(x,yn)log|[ι(ξ),ι(η);ι(x),ι(yn)]|,C_{\Omega_{1}}(x,y_{n})\geq\log|[\iota^{*}(\xi^{\prime}),\iota^{*}(\eta^{\prime});\iota(x),\iota(y_{n})]|,

we have CΩ1(x,yn)/CΩ2(x,yn)λC_{\Omega_{1}}(x,y_{n})/C_{\Omega_{2}}(x,y_{n})\geq\lambda for all nn, contradicting (10) above. ∎

Proof of 7.11.

We fix an open set Ω1.5\Omega_{1.5} such that Ω1¯Ω1.5\overline{\Omega_{1}}\subset\Omega_{1.5} and Ω1.5¯Ω2\overline{\Omega_{1.5}}\subset\Omega_{2}. Since CΩ1(x,y)CΩ1.5(x,y)C_{\Omega_{1}}(x,y)\geq C_{\Omega_{1.5}}(x,y) for all x,yΩ1x,y\in\Omega_{1}, we just need to see that there is some λ>1\lambda>1 so that

CΩ1.5(x,y)CΩ2(x,y)λ\frac{C_{\Omega_{1.5}}(x,y)}{C_{\Omega_{2}}(x,y)}\geq\lambda

for all x,yΩ1¯x,y\in\overline{\Omega_{1}}. This follows from 7.17. ∎

7.3. Contracting paths are P+P^{+}-divergent

Proposition 7.18.

Let 𝒢\mathcal{G} be a Γ\Gamma-graph for a group ΓG\Gamma\subset G, and let {Uv}\{U_{v}\} be a 𝒢\mathcal{G}-compatible system of open sets of G/P+G/P^{+} with each UvU_{v} a proper domain.

If αn\alpha_{n} is a contracting 𝒢\mathcal{G}-path, then the sequence

γn=α1αn\gamma_{n}=\alpha_{1}\cdots\alpha_{n}

is P+P^{+}-divergent with unique limit point ξ\xi, where {ξ}=n=1γnUn+1\{\xi\}=\bigcap_{n=1}^{\infty}\gamma_{n}U_{n+1}.

Proof.

Consider the sequence of open sets

γnUvn+1.\gamma_{n}\cdot U_{v_{n+1}}.

Up to subsequence, Uvn+1U_{v_{n+1}} is a fixed open set UG/P+U\subset G/P^{+}. By assumption {αn}\{\alpha_{n}\} is a contracting path, so γnUvn+1\gamma_{n}\cdot U_{v_{n+1}} converges to a singleton {ξ}\{\xi\}. 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 (Γ,)(\Gamma,\mathcal{H}) be a relatively hyperbolic pair, let ρ:ΓG\rho:\Gamma\to G be a representation, and let PGP\subset G be a symmetric parabolic subgroup. Suppose there exists

  1. (1)

    a Γ\Gamma-invariant closed set ΛG/P\Lambda\subset G/P and a continuous equivariant surjective antipodal map ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}), and

  2. (2)

    a relative quasigeodesic automaton 𝒢\mathcal{G} and a 𝒢\mathcal{G}-compatible system {Uv}\{U_{v}\} of open subsets of G/PG/P, such that the UvU_{v}’s cover Λ\Lambda and each UvU_{v} is a proper domain intersecting Λ\Lambda nontrivially.

Then for every sequence γnΓ\gamma_{n}\in\Gamma which is unbounded in the coned-off Cayley graph Cay(Γ,S,𝒫)\mathrm{Cay}(\Gamma,S,\mathcal{P}), the sequence ρ(γn)\rho(\gamma_{n}) is PP-divergent, and every PP-limit point of ρ(γn)\rho(\gamma_{n}) lies in Λ\Lambda.

Proof.

We will show that every subsequence of γn\gamma_{n} has a PP-contracting subsequence, so take an arbitrary subsequence of γn\gamma_{n}. By 5.7, we may assume that for a bounded sequence bnΓb_{n}\in\Gamma, γnbn\gamma_{n}b_{n} is the endpoint of a finite 𝒢\mathcal{G}-path {αm(n)}m=1Mn\{\alpha_{m}^{(n)}\}_{m=1}^{M_{n}}. Up to subsequence bnb_{n} is a constant bb, independent of nn.

Let {vmn}\{v_{m}^{n}\} be the vertex path associated to {αm(n)}\{\alpha_{m}^{(n)}\}. Up to subsequence vMn+1nv_{M_{n}+1}^{n} is a fixed vertex vv, and v1nv_{1}^{n} is a fixed vertex vv^{\prime}. Let UvεU_{v^{\prime}}^{\varepsilon} be an ε\varepsilon-neighborhood of UvU_{v^{\prime}}, with ε\varepsilon chosen sufficiently small so that UvεU_{v^{\prime}}^{\varepsilon} is still a proper domain.

The sequence MnM_{n} must be unbounded, since the length of γn\gamma_{n} with respect to the coned-off Cayley graph metric is at most a fixed constant times MnM_{n}. 7.12 then implies that the diameter of

ρ(γnb)Uv=ρ(α1(n))ρ(αMn(n))Uv\rho(\gamma_{n}b)\cdot U_{v}=\rho(\alpha_{1}^{(n)})\cdots\rho(\alpha_{M_{n}}^{(n)})U_{v}

with respect to the metric CUvεC_{U_{v^{\prime}}^{\varepsilon}} tends to zero, exponentially in nn. Since this sequence of sets lies in the compact set Uv¯Uvε\overline{U_{v^{\prime}}}\subset U_{v^{\prime}}^{\varepsilon}, up to subsequence it must converge to a singleton {ξ}\{\xi\} in G/PG/P. In fact ξ\xi must lie in Λ\Lambda, because Λ\Lambda is compact and ξ\xi is the limit of a sequence of points in the sequence of nonempty closed sets (ρ(γnb)Uv¯)Λ(\rho(\gamma_{n}b)\cdot\overline{U_{v}})\cap\Lambda. Then, since ρ(γn)ρ(b)Uv\rho(\gamma_{n})\cdot\rho(b)U_{v} converges to {ξ}\{\xi\}, 3.5 implies that ρ(γn)\rho(\gamma_{n}) is PP-divergent with unique PP-limit ξ\xi. ∎

Lemma 8.2.

Let (Γ,)(\Gamma,\mathcal{H}) be a relatively hyperbolic pair, let ρ:ΓG\rho:\Gamma\to G be a representation, let ΛG/P\Lambda\subset G/P be a closed Γ\Gamma-invariant set, and let ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}) be a continuous equivariant surjective antipodal map.

Suppose that γnΓ\gamma_{n}\in\Gamma is a sequence converging to z+con(Γ,)z_{+}\in\partial_{\mathrm{con}}(\Gamma,\mathcal{H}), such that ρ(γn)\rho(\gamma_{n}) is PP-divergent and every PP-limit point of ρ(γn±1)\rho(\gamma_{n}^{\pm 1}) lies in Λ\Lambda. If γn1\gamma_{n}^{-1} converges to z(Γ,)z_{-}\in\partial(\Gamma,\mathcal{H}), then for every compact set KOpp(ϕ1(z))K\subset\operatorname{Opp}(\phi^{-1}(z_{-})) and every open UU containing ϕ1(z+)\phi^{-1}(z_{+}), for large enough nn, we have ρ(γn)KU\rho(\gamma_{n})K\subset U.

Proof.

It suffices to show that every subsequence of γn\gamma_{n} has a further subsequence satisfying the desired property. So, we can freely extract subsequences throughout this proof.

We assume PP symmetric, so ρ(γn1)\rho(\gamma_{n}^{-1}) is also PP-divergent and has nonempty PP-limit set. Let ξ±\xi_{\pm} be a pair of flags in the PP-limit sets of ρ(γn±1)\rho(\gamma_{n}^{\pm 1}), respectively; by assumption we have ξ±Λ\xi_{\pm}\in\Lambda. By 3.6, we have a subsequence so that ρ(γn)\rho(\gamma_{n}) converges to ξ+\xi_{+} uniformly on compacts in Opp(ξ)\operatorname{Opp}(\xi_{-}).

Antipodality of ϕ\phi implies that every compact subset of (Γ,){ϕ(ξ)}\partial(\Gamma,\mathcal{H})-\{\phi(\xi_{-})\} is contained in ϕ(Opp(ξ)Λ)\phi(\operatorname{Opp}(\xi_{-})\cap\Lambda). Then, by equivariance and continuity of ϕ\phi, we see that γn\gamma_{n} must converge to ϕ(ξ+)\phi(\xi_{+}) on compacts in (Γ,){ϕ(ξ)}\partial(\Gamma,\mathcal{H})-\{\phi(\xi_{-})\}. This uniquely characterizes the points ϕ(ξ±)\phi(\xi_{\pm}) as the limits of γn±1\gamma_{n}^{\pm 1} in (Γ,)\partial(\Gamma,\mathcal{H}). So, we see that ρ(γn)\rho(\gamma_{n}) converges uniformly to ξ+ϕ1(z+)\xi_{+}\in\phi^{-1}(z_{+}) on every compact in Opp(ϕ1(z))Opp(ξ)\operatorname{Opp}(\phi^{-1}(z_{-}))\subset\operatorname{Opp}(\xi_{-}), as required. ∎

We recall the statement of our weaker characterization of EGF representations here:

\conicalPeripheralEGF

*

Proof.

To see the “only if” part, observe that if we know that ϕ\phi 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 ϕ\phi extends convergence dynamics.

So, we focus on the “if” part. For each conical limit point zcon(Γ,)z\in\partial_{\mathrm{con}}(\Gamma,\mathcal{H}), we let Cz=Opp(ϕ1(z))C_{z}=\operatorname{Opp}(\phi^{-1}(z)). Each CzC_{z} contains Λϕ1(z)\Lambda-\phi^{-1}(z) by antipodality of ϕ\phi. For each ppar(Γ,)p\in\partial_{\mathrm{par}}(\Gamma,\mathcal{H}), we can replace CpC_{p} with Opp(ϕ1(p))Cp\operatorname{Opp}(\phi^{-1}(p))\cap C_{p}: this set is still open, and it again contains Λ{ϕ1(p)}\Lambda-\{\phi^{-1}(p)\} by antipodality.

Observe that if γn\gamma_{n} is a sequence limiting conically to z+(Γ,)z_{+}\in\partial(\Gamma,\mathcal{H}), with γn1\gamma_{n}^{-1} converging to zz_{-}, then 8.2, together with part (b) of our hypotheses, implies that the map ϕ:ΛG/P\phi:\Lambda\to G/P 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 𝒢\mathcal{G} satisfying the hypotheses of 8.1.

We now want to show that parts (a) and (b) of our hypotheses show that ϕ\phi is an EGF boundary extension, so let γnΓ\gamma_{n}\in\Gamma be a sequence with γn±1z±(Γ,)\gamma_{n}^{\pm 1}\to z_{\pm}\in\partial(\Gamma,\mathcal{H}). We fix an open set UG/PU\subset G/P containing ϕ1(z+)\phi^{-1}(z_{+}) and a compact KCzK\subset C_{z_{-}}. Our goal is to show that for large enough nn, we have ρ(γn)KU\rho(\gamma_{n})K\subset U.

We consider two cases:

Case 1: γn\gamma_{n} is unbounded in the coned-off Cayley graph Cay(Γ,S,𝒫)\mathrm{Cay}(\Gamma,S,\mathcal{P}). By 8.1, ρ(γn±1)\rho(\gamma_{n}^{\pm 1}) is PP-divergent, and every PP-limit point of ρ(γn±1)\rho(\gamma_{n}^{\pm 1}) lies in Λ\Lambda. Then we are done by 8.2.

Case 2: γn\gamma_{n} is bounded in Cay(Γ,S,𝒫)\mathrm{Cay}(\Gamma,S,\mathcal{P}). We can write γn\gamma_{n} as an alternating product

γn=g1(n)h1(n)gk(n)hk(n)gk+1(n),\gamma_{n}=g_{1}^{(n)}h_{1}^{(n)}\cdots g_{k}^{(n)}h_{k}^{(n)}g_{k+1}^{(n)},

where gi(n)g_{i}^{(n)} is bounded in Γ\Gamma, and hi(n)h_{i}^{(n)} lies in Γpin\Gamma_{p_{i}^{n}} for a parabolic point pinΠp_{i}^{n}\in\Pi. Without loss of generality, the hi(n)h_{i}^{(n)} are unbounded in Γ\Gamma as nn\to\infty. Up to subsequence we can assume that gi(n)=gig_{i}^{(n)}=g_{i} and pin=pip_{i}^{n}=p_{i} (independent of nn). Since Π\Pi contains exactly one representative of each parabolic orbit, we can also assume that gi+1pi+1pig_{i+1}p_{i+1}\neq p_{i} for any ii.

We claim that γn\gamma_{n} converges to z+=g1p1z_{+}=g_{1}p_{1}, γn1\gamma_{n}^{-1} converges to z=gk+11pkz_{-}=g_{k+1}^{-1}p_{k}, and for any compact KCzK\subset C_{z_{-}} and open UU containing ϕ1(z+)\phi^{-1}(z_{+}), for large nn, we have γnKU\gamma_{n}\cdot K\subset U.

Fix such a compact KK and open UU. We will prove the claim by inducting on kk. When k=1k=1, then p=p1=pkp=p_{1}=p_{k}, and γn=g1hng2\gamma_{n}=g_{1}h_{n}g_{2} for hnΓph_{n}\in\Gamma_{p} and g1,g2Γg_{1},g_{2}\in\Gamma fixed. The distance between hng2h_{n}g_{2} and hnh_{n} is bounded in any word metric on Γ\Gamma, so hng2h_{n}g_{2} converges to pp in Γ¯\overline{\Gamma} and g1hng2g_{1}h_{n}g_{2} converges to g1p=z+g_{1}p=z_{+}. We also know that KCz=Cg21pK\subset C_{z_{-}}=C_{g_{2}^{-1}p}, so hng2Kh_{n}g_{2}K eventually lies in a small neighborhood of ϕ1(p)\phi^{-1}(p) by part (b) of our hypotheses. Then g1hng2Kg_{1}h_{n}g_{2}K lies in any small neighborhood of ϕ1(g1p)=ϕ1(z+)\phi^{-1}(g_{1}p)=\phi^{-1}(z_{+}).

When k>1k>1, we consider the sequence

γn=g2h2(n)gkhk(n)gk+1.\gamma_{n}^{\prime}=g_{2}h_{2}^{(n)}\cdots g_{k}h_{k}^{(n)}g_{k+1}.

Inductively we can assume that for large nn, γng2p2\gamma_{n}^{\prime}\to g_{2}p_{2} and ρ(γn)K\rho(\gamma_{n}^{\prime})\cdot K is a subset of an arbitrarily small neighborhood of ϕ1(g2p2)\phi^{-1}(g_{2}p_{2}). Then since p1g2p2p_{1}\neq g_{2}p_{2}, for large enough nn, ρ(γn)K\rho(\gamma_{n}^{\prime})\cdot K is a compact subset of Cp1C_{p_{1}}. So our hypotheses imply that for large nn,

ρ(γn)K=ρ(g1h1(n))ρ(γn)KU.\rho(\gamma_{n})\cdot K=\rho(g_{1}h_{1}^{(n)})\rho(\gamma_{n}^{\prime})\cdot K\subset U.

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 (Γ,)(\Gamma,\mathcal{H}) be a relatively hyperbolic pair, let ρ:ΓG\rho:\Gamma\to G be a representation, and let PGP\subset G be a symmetric parabolic subgroup. Suppose that there exists a closed Γ\Gamma-invariant subset ΛG/P\Lambda\subseteq G/P and a surjective equivariant antipodal map ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}).

Then ρ\rho is EGF with boundary extension ϕ\phi if and only if for every z(Γ,)z\in\partial(\Gamma,\mathcal{H}), there exists an open CzG/PC_{z}\subset G/P containing Λϕ1(z)\Lambda-\phi^{-1}(z), such that:

  1. (a)

    For any sequence γnΓ\gamma_{n}\in\Gamma limiting conically to some point zz in (Γ,)\partial(\Gamma,\mathcal{H}), with γn1z\gamma_{n}^{-1}\to z_{-}, any open set UU containing ϕ1(z)\phi^{-1}(z), and any compact KCzK\subset C_{z_{-}}, we have ρ(γn)KU\rho(\gamma_{n})\cdot K\subset U for all sufficiently large nn.

  2. (b)

    For any parabolic point p(Γ,)p\in\partial(\Gamma,\mathcal{H}), any compact KCpK\subset C_{p}, and any open set UU containing ϕ1(p)\phi^{-1}(p), for all but finitely many γΓp\gamma\in\Gamma_{p}, we have ρ(γ)KU\rho(\gamma)\cdot K\subset U.

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 Hom(Γ,G)\operatorname{Hom}(\Gamma,G). However, it is relatively open in a subspace of Hom(Γ,G)\operatorname{Hom}(\Gamma,G) where we restrict the deformations of the peripheral subgroups appropriately. Roughly speaking, we want to consider subspaces 𝒲Hom(Γ,G)\mathcal{W}\subset\operatorname{Hom}(\Gamma,G) where the dynamical behavior of the peripheral subgroups changes continuously under deformation. That is, if ρt\rho_{t} is a small deformation of a representation ρ0\rho_{0}, where ρ0(Γp)\rho_{0}(\Gamma_{p}) attracts points towards Λp\Lambda_{p} at a particular “speed,” then we want ρt(Γp)\rho_{t}(\Gamma_{p}) to attract points towards a small deformation of Λp\Lambda_{p} at a similar “speed.”

The precise condition is the following:

Definition 9.1.

Let ρ0:ΓG\rho_{0}:\Gamma\to G be an EGF representation with boundary extension ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}), and let 𝒲Hom(Γ,G)\mathcal{W}\subset\operatorname{Hom}(\Gamma,G) contain ρ0\rho_{0}.

We say that 𝒲\mathcal{W} is peripherally stable at (ρ0,ϕ)(\rho_{0},\phi) if for every ppar(Γ,)p\in\partial_{\mathrm{par}}(\Gamma,\mathcal{H}), every open set UU containing ϕ1(p)\phi^{-1}(p), every compact set KCpK\subset C_{p}, and every cofinite set TΓpT\subset\Gamma_{p} such that

ρ0(T)KU,\rho_{0}(T)\cdot K\subset U,

there is an open set 𝒲𝒲\mathcal{W}^{\prime}\subset\mathcal{W} containing ρ0\rho_{0}, such that for every ρ𝒲\rho^{\prime}\in\mathcal{W}^{\prime}, we have

ρ(T)KU.\rho^{\prime}(T)\cdot K\subset U.

We restate the main result of the paper below:

\cuspStableStability

*

Remark 9.2.

In [Bow98], Bowditch explored the deformation spaces of geometrically finite groups ΓPO(d,1)\Gamma\subset\operatorname{PO}(d,1), and gave an explicit discription of semialgebraic subspaces of Hom(Γ,PO(d,1))\operatorname{Hom}(\Gamma,\operatorname{PO}(d,1)) in which small deformations of Γ\Gamma 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 PO(d,1)\operatorname{PO}(d,1), the question is subtle, however. Bowditch also gives examples of geometrically finite representations ρ:ΓPO(d,1)\rho:\Gamma\to\operatorname{PO}(d,1) (for d4d\geq 4) and deformations ρt\rho_{t} of ρ\rho in Hom(Γ,PO(d,1))\operatorname{Hom}(\Gamma,\operatorname{PO}(d,1)) such that the restriction of ρt\rho_{t} to each cusp group in Γ\Gamma is discrete, faithful, and parabolic, but ρt\rho_{t} is not even discrete; further examples exist where the deformation is discrete, but not geometrically finite.

Example 9.3.

Let BSL(d,)B\in\operatorname{SL}(d,\mathbb{R}) be a dd-dimensional Jordan block with eigenvalue 1 and eigenvector vv, and let ASL(d+2,)A\in\operatorname{SL}(d+2,\mathbb{R}) be the block matrix (B11)\begin{pmatrix}B\\ &1\\ &&1\end{pmatrix}.

Although [v][v] is not quite an attracting fixed point of AA, it is still an “attracting subspace” in the sense that if KK is any compact subset of Pd+1\mathbb{R}\mathrm{P}^{d+1} which does not intersect a fixed hyperplane of AA, then AnKA^{n}\cdot K converges to {[v]}\{[v]\}. Via a ping-pong argument, one can use this “attracting” behavior to show that for some k1k\geq 1 and some MSL(d+2,)M\in\operatorname{SL}(d+2,\mathbb{R}), the group Γ\Gamma generated by α=Ak\alpha=A^{k} and β=MAkM1\beta=MA^{k}M^{-1} is a discrete free group with free generators α,β\alpha,\beta. The group Γ\Gamma is hyperbolic relative to the subgroups α\langle\alpha\rangle, β\langle\beta\rangle, and the inclusion ΓSL(d+2,)\Gamma\hookrightarrow\operatorname{SL}(d+2,\mathbb{R}) is EGF with respect to P1,d+1P_{1,d+1} (the stabilizer of a line in a hyperplane in d+2\mathbb{R}^{d+2}).

Here, there are peripherally stable deformations of Γ\Gamma which change the Jordan block decomposition of AA. For instance, consider a continuous path At:[0,1]SL(d+2,)A_{t}:[0,1]\to\operatorname{SL}(d+2,\mathbb{R}) given by At=(Bt11)A_{t}=\begin{pmatrix}B_{t}\\ &1\\ &&1\end{pmatrix}, where B0=BB_{0}=B and BtB_{t} is a diagonalizable matrix in SL(d,)\operatorname{SL}(d,\mathbb{R}). For small values of tt, the group Γt\Gamma_{t} generated by αt=Atk\alpha_{t}=A_{t}^{k} and β\beta is still discrete and freely generated by αt\alpha_{t} and β\beta—since the “attracting” fixed points of AtA_{t} deform continuously with tt, the same exact ping-pong setup works for all small t0t\geq 0. And indeed the path in Hom(Γ,SL(d+2,))\operatorname{Hom}(\Gamma,\operatorname{SL}(d+2,\mathbb{R})) determined by the path AtA_{t} is a peripherally stable subspace.

On the other hand, consider the path At=(Betet)A_{t}^{\prime}=\begin{pmatrix}B\\ &e^{t}\\ &&e^{-t}\end{pmatrix}, and let αt=Atk\alpha_{t}^{\prime}=A_{t}^{\prime k}. In this case the corresponding subspace of Hom(Γ,SL(d+2,))\operatorname{Hom}(\Gamma,\operatorname{SL}(d+2,\mathbb{R})) is not peripherally stable: while the group generated by αt\alpha_{t}^{\prime} is still discrete, the attracting fixed points of AtA_{t}^{\prime} do not deform continuously in tt. So, there is no way to use the ping-pong setup for Γ\Gamma to ensure that Γt=αt,β\Gamma_{t}^{\prime}=\langle\alpha_{t}^{\prime},\beta\rangle is a discrete group.

Example 9.4.

Here is a somewhat more interesting example of a non-peripherally stable deformation. Let MM be a finite-volume noncompact hyperbolic 3-manifold, with holonomy representation ρ:π1MPSL(2,)\rho:\pi_{1}M\to\operatorname{PSL}(2,\mathbb{C}) (so there is an identification M=3/ρ(π1M)M=\mathbb{H}^{3}/\rho(\pi_{1}M)). Then π1M\pi_{1}M is hyperbolic relative to the collection 𝒞\mathcal{C} of conjugates of cusp groups (each of which is isomorphic to 2\mathbb{Z}^{2}), and the representation ρ\rho is geometrically finite (in particular, EGF).

In this case, for any sufficiently small nontrivial deformation ρ\rho^{\prime} of ρ\rho in the character variety Hom(π1M,PSL(2,))/PSL(2,)\operatorname{Hom}(\pi_{1}M,\operatorname{PSL}(2,\mathbb{C}))/\operatorname{PSL}(2,\mathbb{C}), the restriction of ρ\rho^{\prime} to some cusp group C𝒞C\in\mathcal{C} either fails to be discrete or has infinite kernel. So Hom(π1M,PSL(2,))\operatorname{Hom}(\pi_{1}M,\operatorname{PSL}(2,\mathbb{C})) is not peripherally stable, because any sufficiently small deformation of ρ\rho inside of a peripherally stable subspace must have discrete image and finite kernel on each C𝒞C\in\mathcal{C}. This is true despite the fact that arbitrarily small deformations of ρ\rho are holonomy representations of complete hyperbolic structures on Dehn fillings of MM (so in particular, they are discrete).

The main ingredient in the proof of Section 1.4 is the relative quasigeodesic automaton 𝒢\mathcal{G} and the associated 𝒢\mathcal{G}-compatible system of open sets {Uv}\{U_{v}\} we constructed in 6.1. The following proposition is immediate from the definition of peripheral stability:

Proposition 9.5.

Let ρ:ΓG\rho:\Gamma\to G be an EGF representation with boundary extension ϕ\phi, and let 𝒲Hom(Γ,G)\mathcal{W}\subset\operatorname{Hom}(\Gamma,G) be a subspace which is peripherally stable at (ρ,ϕ)(\rho,\phi).

If 𝒢\mathcal{G} is a relative quasigeodesic automaton for Γ\Gamma, and {Uv}\{U_{v}\} is a 𝒢\mathcal{G}-compatible system of open subsets of G/PG/P for ρ(Γ)\rho(\Gamma), then there is an open subset 𝒲𝒲\mathcal{W}^{\prime}\subset\mathcal{W} containing ρ\rho such that for every ρ𝒲\rho^{\prime}\in\mathcal{W}^{\prime}, {Uv}\{U_{v}\} is also a 𝒢\mathcal{G}-compatible system of open sets for ρ(Γ)\rho^{\prime}(\Gamma).

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 Γ\Gamma from the 𝒢\mathcal{G}-compatible system {Uv}\{U_{v}\}.

9.2. An equivariant map on conical limit points

For the rest of this section, we let ρ:ΓG\rho:\Gamma\to G be a representation which is EGF with respect to a symmetric parabolic subgroup PGP\subset G. We let ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}) be a boundary extension for ρ\rho, and assume that 𝒲Hom(Γ,G)\mathcal{W}\subset\operatorname{Hom}(\Gamma,G) is peripherally stable at (ρ,ϕ)(\rho,\phi). We also let ZZ be a compact subset of (Γ,)\partial(\Gamma,\mathcal{H}), and let VG/PV\subset G/P be an open subset containing ϕ1(Z)\phi^{-1}(Z). We again fix a finite subset Πpar(Γ,)\Pi\subset\partial_{\mathrm{par}}(\Gamma,\mathcal{H}), containing one point from every Γ\Gamma-orbit in par(Γ,)\partial_{\mathrm{par}}(\Gamma,\mathcal{H}).

Using 6.1, we can find a relative quasigeodesic automaton 𝒢\mathcal{G} and 𝒢\mathcal{G}-compatible system {Uv}\{U_{v}\} of open subsets of G/PG/P for ρ(Γ)\rho(\Gamma). Using 6.14, we can ensure that for any zZz\in Z, there is a 𝒢\mathcal{G}-path {αn}\{\alpha_{n}\} limiting to zz (with vertex path {vn}\{v_{n}\}) so that Uv1U_{v_{1}} is contained in VV.

For each pΠp\in\Pi, we also fix a compact set Kp(Γ,){p}K_{p}\subset\partial(\Gamma,\mathcal{H})-\{p\} such that ΓpKp=(Γ,){p}\Gamma_{p}\cdot K_{p}=\partial(\Gamma,\mathcal{H})-\{p\}, and assume that the automaton 𝒢\mathcal{G} has been constructed to satisfy 6.15.

Antipodality of the map ϕ\phi implies that for each z(Γ,)z\in\partial(\Gamma,\mathcal{H}), each fiber ϕ1(z)\phi^{-1}(z) is a closed subset of some affine chart in G/PG/P. So, we can also assume that UvU_{v} is a proper domain for each vertex vv of 𝒢\mathcal{G}. In fact, by way of the following lemma, we can assume even more:

Lemma 9.6.

Let ρ\rho be an EGF representation with boundary map ϕ:Λ(Γ,)\phi:\Lambda\to\partial(\Gamma,\mathcal{H}).

For any δ>0\delta>0, we can find a relative quasigeodesic automaton 𝒢\mathcal{G} with 𝒢\mathcal{G}-compatible system {Uv}\{U_{v}\} of open sets in G/PG/P as in 6.1, so that for any x,y(Γ,)x,y\in\partial(\Gamma,\mathcal{H}) with d(x,y)>δd(x,y)>\delta, if ϕ1(x)Uv\phi^{-1}(x)\subset U_{v} and ϕ1(y)Uw\phi^{-1}(y)\subset U_{w}, then Uv¯\overline{U_{v}} and Uw¯\overline{U_{w}} are opposite.

Proof.

We choose ε>0\varepsilon>0 so that if d(v,w)>δ/2d(v,w)>\delta/2 for v,w(Γ,)v,w\in\partial(\Gamma,\mathcal{H}), then the closed ε\varepsilon-neighborhoods of

ϕ1(v),ϕ1(w)\phi^{-1}(v),\qquad\phi^{-1}(w)

are opposite. This is possible for a fixed pair v,w(Γ,)v,w\in\partial(\Gamma,\mathcal{H}) since antipodality is an open condition, and ϕ1(v)\phi^{-1}(v), ϕ1(w)\phi^{-1}(w) are opposite compact sets. Then we can pick a uniform ε\varepsilon for all pairs since the the subset {(u,v)((Γ,))2:d(u,v)>δ/2}\{(u,v)\in(\partial(\Gamma,\mathcal{H}))^{2}:d(u,v)>\delta/2\} is compact.

Consider 𝒢\mathcal{G}-compatible systems of open subsets {Uv}\{U_{v}\} and {Wv}\{W_{v}\} for the action of Γ\Gamma on G/PG/P and (Γ,)\partial(\Gamma,\mathcal{H}), coming from 6.1. We can ensure that for each vertex aa, the diameter of WaW_{a} is at most δ/4\delta/4, and UaN(ϕ1(w),ε)U_{a}\subset N(\phi^{-1}(w),\varepsilon) for some wWaw\in W_{a}.

If x,y(Γ,)x,y\in\partial(\Gamma,\mathcal{H}) satisfy d(x,y)>δd(x,y)>\delta, and xWax\in W_{a}, yWby\in W_{b}, we have

d(v,w)>δ/2d(v,w)>\delta/2

for all vWav\in W_{a}, wWbw\in W_{b}.

Then, if ϕ1(x)Ua\phi^{-1}(x)\subset U_{a} and ϕ1(y)Ub\phi^{-1}(y)\subset U_{b}, we have

UaN(ϕ1(v),ε),UbN(ϕ1(w),ε)U_{a}\subset N(\phi^{-1}(v),\varepsilon),\qquad U_{b}\subset N(\phi^{-1}(w),\varepsilon)

for vWav\in W_{a}, wWbw\in W_{b} with d(v,w)>δ/2d(v,w)>\delta/2. By our choice of ε\varepsilon, the closures of N(ϕ1(w),ε)N(\phi^{-1}(w),\varepsilon) and N(ϕ1(v),ε)N(\phi^{-1}(v),\varepsilon) are opposite. ∎

Using cocompactness of the action of Γ\Gamma on the space of distinct pairs in (Γ,)\partial(\Gamma,\mathcal{H}), we know that there exists some fixed δ>0\delta>0 such that for any distinct z1,z2(Γ,)z_{1},z_{2}\in\partial(\Gamma,\mathcal{H}), we can find some γΓ\gamma\in\Gamma such that d(γz1,γz2)>δd(\gamma z_{1},\gamma z_{2})>\delta. Then, in light of 9.6, we can make the following assumption:

Assumption 9.7.

For any z1,z2(Γ,)z_{1},z_{2}\in\partial(\Gamma,\mathcal{H}) satisfying d(z1,z2)>δd(z_{1},z_{2})>\delta, if ϕ1(z1)Uv\phi^{-1}(z_{1})\subset U_{v} and ϕ1(z2)Uw\phi^{-1}(z_{2})\subset U_{w} for v,wv,w vertices of 𝒢\mathcal{G}, then Uv¯\overline{U_{v}} and Uw¯\overline{U_{w}} are opposite.

With our relative quasigeodesic automaton 𝒢\mathcal{G} and compatible system of open sets {Uv}\{U_{v}\} fixed, we now choose an open subset 𝒲𝒲\mathcal{W}^{\prime}\subset\mathcal{W} so that for any ρ𝒲\rho^{\prime}\in\mathcal{W}^{\prime}, {Uv}\{U_{v}\} is also a 𝒢\mathcal{G}-compatible system for the action of ρ(Γ)\rho^{\prime}(\Gamma) on G/PG/P. Our main goal for the rest of this section is to show that any ρ𝒲\rho^{\prime}\in\mathcal{W}^{\prime} is an EGF representation. So, we fix some ρ𝒲\rho^{\prime}\in\mathcal{W}^{\prime}.

Let Path(𝒢)\mathrm{Path}(\mathcal{G}) denote the set of infinite 𝒢\mathcal{G}-paths. 7.3 implies that every path in Path(𝒢)\mathrm{Path}(\mathcal{G}) is contracting for the ρ\rho^{\prime}-action, so we have a map

ψρ:Path(𝒢)G/P,\psi_{\rho^{\prime}}:\mathrm{Path}(\mathcal{G})\to G/P,

where the path {αn}\{\alpha_{n}\} maps to the unique element of

n=1ρ(α1)ρ(αn)Uvn+1.\bigcap_{n=1}^{\infty}\rho^{\prime}(\alpha_{1})\cdots\rho^{\prime}(\alpha_{n})U_{v_{n+1}}.
Lemma 9.8.

The map ψρ:Path(𝒢)G/P\psi_{\rho^{\prime}}:\mathrm{Path}(\mathcal{G})\to G/P induces an equivariant map

ψρ:conΓG/P.\psi_{\rho^{\prime}}:\partial_{\mathrm{con}}\Gamma\to G/P.
Proof.

We first need to see that ψρ\psi_{\rho^{\prime}} is well-defined, i.e. that if zz is a conical limit point and {αn}\{\alpha_{n}\}, {βn}\{\beta_{n}\} are 𝒢\mathcal{G}-paths limiting to zz, then ψρ({αn})=ψρ({βn})\psi_{\rho^{\prime}}(\{\alpha_{n}\})=\psi_{\rho^{\prime}}(\{\beta_{n}\}).

Let

γn=α1αn,ηm=β1βm.\gamma_{n}=\alpha_{1}\cdots\alpha_{n},\qquad\eta_{m}=\beta_{1}\cdots\beta_{m}.

We can use 5.11 to see that γn\gamma_{n} and ηm\eta_{m} lie within bounded Hausdorff distance of a geodesic in Cay(Γ,S,𝒫)\mathrm{Cay}(\Gamma,S,\mathcal{P}) limiting to zz, so there is a fixed DD so that for infinitely many pairs m,nm,n,

d(γn,ηm)<Dd(\gamma_{n},\eta_{m})<D

in the Cayley graph of Γ\Gamma. 7.18 implies that ρ(γn)\rho^{\prime}(\gamma_{n}) and ρ(ηn)\rho^{\prime}(\eta_{n}) are both PP-divergent sequences and each have a unique PP-limit point in G/PG/P, given by ψρ({αn})\psi_{\rho^{\prime}}(\{\alpha_{n}\}), ψρ({βm})\psi_{\rho^{\prime}}(\{\beta_{m}\}), respectively. Then, Lemma 4.23 in [KLP17] implies that because ρ(γn)=ρ(ηn)gn\rho^{\prime}(\gamma_{n})=\rho^{\prime}(\eta_{n})g_{n} for a bounded sequence gnGg_{n}\in G, the PP-limit points of ρ(γn)\rho^{\prime}(\gamma_{n}) and ρ(ηn)\rho^{\prime}(\eta_{n}) must agree and therefore ψρ({αn})=ψρ({βm})\psi_{\rho^{\prime}}(\{\alpha_{n}\})=\psi_{\rho^{\prime}}(\{\beta_{m}\}).

Next we observe that ψρ\psi_{\rho^{\prime}} is equivariant. Fix a finite generating set SS for Γ\Gamma. It suffices to show that ψρ(sz)=ρ(s)ψρ(z)\psi_{\rho^{\prime}}(s\cdot z)=\rho^{\prime}(s)\cdot\psi_{\rho^{\prime}}(z) for all sSs\in S.

Let {αn}\{\alpha_{n}\} be a 𝒢\mathcal{G}-path limiting to some zconΓz\in\partial_{\mathrm{con}}\Gamma, and consider the sequence

γn=sα1αn.\gamma_{n}^{\prime}=s\alpha_{1}\cdots\alpha_{n}.

Again, 5.11 implies that γn\gamma_{n}^{\prime} lies bounded Hausdorff distance from a geodesic in Cay(Γ,S,𝒫)\mathrm{Cay}(\Gamma,S,\mathcal{P}), which must limit to szs\cdot z. So if we fix a 𝒢\mathcal{G}-path βn\beta_{n} limiting to szs\cdot z, the same argument as above shows that ψρ({βn})=ρ(s)ψρ({αn})\psi_{\rho^{\prime}}(\{\beta_{n}\})=\rho^{\prime}(s)\cdot\psi_{\rho^{\prime}}(\{\alpha_{n}\}). ∎

It will turn out that ψρ\psi_{\rho^{\prime}} is also both continuous and injective. However, we do not prove this directly.

9.3. Extending ψρ\psi_{\rho^{\prime}} to parabolic points

We want to extend the map ψρ:conΓG/P\psi_{\rho^{\prime}}:\partial_{\mathrm{con}}\Gamma\to G/P to the entire Bowditch boundary (Γ,)\partial(\Gamma,\mathcal{H}). To do so, we need to view ψρ\psi_{\rho^{\prime}} as a map to the set of closed subsets of G/PG/P.

The first step is to define ψρ\psi_{\rho^{\prime}} on the finite set ΠparΓ\Pi\subset\partial_{\mathrm{par}}\Gamma. For any vertex vv in 𝒢\mathcal{G}, we consider the set

Bv=(v,y) edge in 𝒢Uy.B_{v}=\bigcup_{(v,y)\textrm{ edge in }\mathcal{G}}U_{y}.

Then, for each pΠp\in\Pi, we pick a parabolic vertex ww so that pw=pp_{w}=p. We define Λp\Lambda_{p}^{\prime} to be the closure of the set of accumulation points of sequences of the form ρ(γn)x\rho^{\prime}(\gamma_{n})\cdot x, for xBwx\in B_{w} and γn\gamma_{n} distinct elements of Γp\Gamma_{p}. Part (3) of 6.1 guarantees that BwCpB_{w}\subset C_{p}, and 𝒢\mathcal{G}-compatiblity of the system {Uv}\{U_{v}\} for the ρ(Γ)\rho^{\prime}(\Gamma)-action on G/PG/P implies that ΛpUw\Lambda_{p}^{\prime}\subset U_{w}. By construction, Λp\Lambda_{p}^{\prime} is invariant under the action of ρ(Γp)\rho(\Gamma_{p}).

Next, given a parabolic point qparΓq\in\partial_{\mathrm{par}}\Gamma, we write q=gpq=g\cdot p for pΠp\in\Pi, and then define

ψρ(q):=ρ(g)Λp.\psi_{\rho^{\prime}}(q):=\rho^{\prime}(g)\Lambda_{p}^{\prime}.

Since Λp\Lambda_{p}^{\prime} is Γp\Gamma_{p}-invariant and Γp\Gamma_{p} is exactly the stabilizer of pp, this does not depend on the choice of coset representative in gΓpg\Gamma_{p}. Moreover ψρ\psi_{\rho^{\prime}} is still ρ\rho^{\prime}-equivariant.

In addition, if vv is any parabolic vertex with parabolic point pv=gpp_{v}=g\cdot p for pΠp\in\Pi, part (2) of 6.1 ensures that Bv=BwB_{v}=B_{w} for any parabolic vertex ww with pw=pp_{w}=p. So, ρ(g)Λp\rho^{\prime}(g)\cdot\Lambda_{p}^{\prime} is exactly the closure of the set of accumulation points of the form ρ(gγn)x\rho^{\prime}(g\gamma_{n})\cdot x for sequences γnΓp\gamma_{n}\in\Gamma_{p} and xBvx\in B_{v}. Then 𝒢\mathcal{G}-compatibility implies that ψρ(pv)=ρ(g)Λp\psi_{\rho^{\prime}}(p_{v})=\rho(g)\Lambda_{p}^{\prime} is a subset of UvU_{v}.

Remark 9.9.

There is a natural topology on the space of closed subsets of G/PG/P, induced by the Hausdorff distance arising from some (any) choice of metric on G/PG/P. We emphasize that the map ψρ\psi_{\rho^{\prime}} is not necessarily continuous with respect to this topology.

Ultimately we want to use ψρ\psi_{\rho^{\prime}} to define a map extending the convergence dynamics of Γ\Gamma, so we will need to also define the sets CzC_{z}^{\prime} for each z(Γ,)z\in\partial(\Gamma,\mathcal{H}). For now, we only define CpC_{p}^{\prime} for pΠp\in\Pi: this will be the set

γΓpρ(γ)Bw.\bigcup_{\gamma\in\Gamma_{p}}\rho^{\prime}(\gamma)B_{w}.

We can immediately observe:

Proposition 9.10.

CpC_{p}^{\prime} is ρ(Γp)\rho^{\prime}(\Gamma_{p})-invariant. Moreover, for any infinite sequence γnΓp\gamma_{n}\in\Gamma_{p}, any compact KCpK\subset C_{p}^{\prime}, and any open UG/PU\subset G/P containing Λp\Lambda_{p}^{\prime}, for sufficiently large nn, ρ(γn)K\rho^{\prime}(\gamma_{n})\cdot K lies in UU.

Proof.

Γp\Gamma_{p}-invariance follows directly from the definition.

Fix a compact KCpK\subset C_{p}^{\prime} and an open UG/PU\subset G/P containing Λp\Lambda_{p}^{\prime}. KK is contained in finitely many sets ρ(γ)Bw\rho^{\prime}(\gamma)B_{w} for γΓp\gamma\in\Gamma_{p}, so any accumulation point of ρ(γn)x\rho^{\prime}(\gamma_{n})x for xKx\in K and γnΓp\gamma_{n}\in\Gamma_{p} lies in Λp\Lambda_{p}^{\prime}. In particular, for sufficiently large nn, γnx\gamma_{n}x lies in UU, and since KK is compact we can pick nn large enough so that γnxU\gamma_{n}x\in U for all xKx\in K. ∎

We next want to use ψρ\psi_{\rho^{\prime}} to define an antipodal extension from a subset of G/PG/P to (Γ,)\partial(\Gamma,\mathcal{H}).

Lemma 9.11.

For any z(Γ,)z\in\partial(\Gamma,\mathcal{H}), if {αn}\{\alpha_{n}\} is a 𝒢\mathcal{G}-path limiting to zz with corresponding vertex path {vn}\{v_{n}\}, then ϕ1(z)\phi^{-1}(z) and ψρ(z)\psi_{\rho^{\prime}}(z) are both subsets of Uv1U_{v_{1}}.

Proof.

If zz is a conical limit point, then this follows immediately from 5.11 and the definition of ψρ\psi_{\rho^{\prime}}. On the other hand, if zz is a parabolic point, then z=α1αNpvz=\alpha_{1}\cdots\alpha_{N}p_{v}, where vv is a parabolic vertex at the end of the vertex path {vn}\{v_{n}\}. By part (3) of 6.1, we have pvWvp_{v}\in W_{v} and thus ϕ1(pv)Uv\phi^{-1}(p_{v})\subset U_{v}. By ρ\rho-equivariance of ϕ\phi we have

ϕ1(z)=ρ(α1αN)ϕ1(pv),\phi^{-1}(z)=\rho(\alpha_{1}\cdots\alpha_{N})\phi^{-1}(p_{v}),

so by 𝒢\mathcal{G}-compatibility we have ϕ1(z)Uv1\phi^{-1}(z)\subset U_{v_{1}}. On the other hand, we have constructed ψρ\psi_{\rho^{\prime}} so that ψρ(pv)Uv\psi_{\rho^{\prime}}(p_{v})\subset U_{v}, so ρ\rho^{\prime}-equivariance of ψρ\psi_{\rho^{\prime}} and 𝒢\mathcal{G}-compatibility also show that ψρ(z)Uv1\psi_{\rho^{\prime}}(z)\subset U_{v_{1}}. ∎

Lemma 9.12.

For any two distinct points z1,z2z_{1},z_{2} in (Γ,)\partial(\Gamma,\mathcal{H}), the sets

ψρ(z1),ψρ(z2)\psi_{\rho^{\prime}}(z_{1}),\qquad\psi_{\rho^{\prime}}(z_{2})

are opposite (in particular disjoint).

Proof.

We know that for any distinct z1,z2>0z_{1},z_{2}>0, we can find γΓ\gamma\in\Gamma so that d(γz1,γz2)>δd(\gamma z_{1},\gamma z_{2})>\delta. So, since ψρ\psi_{\rho^{\prime}} is ρ\rho^{\prime}-equivariant, we just need to show that if z1,z2(Γ,)z_{1},z_{2}\in\partial(\Gamma,\mathcal{H}) satisfy d(z1,z2)>δd(z_{1},z_{2})>\delta, then ψρ(z1)\psi_{\rho^{\prime}}(z_{1}) is opposite to ψρ(z2)\psi_{\rho^{\prime}}(z_{2}).

Let {αn}\{\alpha_{n}\}, {βn}\{\beta_{n}\} be 𝒢\mathcal{G}-paths respectively limiting to points z1,z2(Γ,)z_{1},z_{2}\in\partial(\Gamma,\mathcal{H}) with d(z1,z2)>δd(z_{1},z_{2})>\delta, with corresponding vertex paths {vn}\{v_{n}\} and {wn}\{w_{n}\}. By 9.11, we must have ϕ1(z1)Uv1\phi^{-1}(z_{1})\subset U_{v_{1}} and ϕ1(z2)Uw2\phi^{-1}(z_{2})\subset U_{w_{2}}, so under 9.7, we know that Uv1U_{v_{1}} and Uw1U_{w_{1}} are opposite. But then we are done since 9.11 also implies that ψρ(z1)Uv1\psi_{\rho^{\prime}}(z_{1})\subset U_{v_{1}} and ψρ(z2)Uw1\psi_{\rho^{\prime}}(z_{2})\subset U_{w_{1}}. ∎

9.4. The boundary set of the deformed representation

We define our candidate boundary set ΛG/P\Lambda^{\prime}\subset G/P by

Λ=z(Γ,)ψρ(z).\Lambda^{\prime}=\bigcup_{z\in\partial(\Gamma,\mathcal{H})}\psi_{\rho^{\prime}}(z).

We then have an equivariant map

ϕ:Λ(Γ,),\phi^{\prime}:\Lambda^{\prime}\to\partial(\Gamma,\mathcal{H}),

where ϕ(x)=z\phi^{\prime}(x)=z if xψρ(z)x\in\psi_{\rho^{\prime}}(z). 9.12 implies that ϕ\phi^{\prime} is well-defined and antipodal. It is necessarily both surjective and ρ\rho^{\prime}-equivariant, and its fibers are either singletons or translates of the sets Λp\Lambda_{p}^{\prime} for pΠp\in\Pi.

It now remains to verify the properties of the candidate set Λ\Lambda^{\prime} and the map ϕ\phi^{\prime} needed to show that ϕ\phi^{\prime} is an EGF boundary extension.

Lemma 9.13.

For every vertex vv of 𝒢\mathcal{G}, the intersection ΛUv\Lambda^{\prime}\cap U_{v} is nonempty.

Proof.

The construction in Section 6 ensures that every vertex of the automaton 𝒢\mathcal{G} has at least one outgoing edge. In particular this means that for a given vertex vv, there is an infinite 𝒢\mathcal{G}-path whose first vertex is vv. This 𝒢\mathcal{G}-path limits to a conical limit point zz, and 9.11 implies that ψρ(z)\psi_{\rho^{\prime}}(z) is a (nonempty) subset of both UvU_{v} and Λ\Lambda^{\prime}. ∎

Lemma 9.14.

For any zZz\in Z, we have ϕ1(z)V\phi^{\prime-1}(z)\subset V.

Proof.

Recall that we used 6.14 to construct our automaton so that for any zZz\in Z, there is a 𝒢\mathcal{G}-path limiting to zz with vertex path {vn}\{v_{n}\} such that Uv1VU_{v_{1}}\subset V. Then 9.11 implies ϕ1(z)V\phi^{\prime-1}(z)\subset V. ∎

Lemma 9.15.

Λ\Lambda^{\prime} is compact.

Proof.

Fix a sequence ynΛy_{n}\in\Lambda^{\prime}, and let xn=ϕ(yn)x_{n}=\phi^{\prime}(y_{n}). Since (Γ,)\partial(\Gamma,\mathcal{H}) is compact, up to subsequence xnxx_{n}\to x. We want to see that a subsequence of yny_{n} converges to some yΛy\in\Lambda^{\prime}. We consider two possibilities:

Case 1: xx is a parabolic point. We can write x=gpx=g\cdot p, where pΠp\in\Pi. Let ww be a parabolic vertex with pw=pp_{w}=p, and consider the compact set Kp(Γ,){p}K_{p}\subset\partial(\Gamma,\mathcal{H})-\{p\}, chosen so that ΓpK=(Γ,){p}\Gamma_{p}\cdot K=\partial(\Gamma,\mathcal{H})-\{p\}. If xn=qx_{n}=q for infinitely many nn, we are done, so assume otherwise, and choose γnΓp\gamma_{n}\in\Gamma_{p} so that zn=γn1g1xnKpz_{n}=\gamma_{n}^{-1}g^{-1}x_{n}\in K_{p}.

We have assumed (using 6.15) that the automaton 𝒢\mathcal{G} has been constructed so that there is always a 𝒢\mathcal{G}-path limiting to znz_{n} whose first vertex vnv_{n} is connected to ww by an edge (w,vn)(w,v_{n}). 9.11 implies that ϕ1(zn)\phi^{\prime-1}(z_{n}) lies in UnU_{n}, which is contained in CpC_{p}^{\prime} by definition.

Then using 9.10, we know that up to subsequence,

ρ(γn)ϕ1(zn)=ρ(γn)ϕ1(γn1g1xn)\rho^{\prime}(\gamma_{n})\phi^{\prime-1}(z_{n})=\rho^{\prime}(\gamma_{n})\phi^{\prime-1}(\gamma_{n}^{-1}g^{-1}x_{n})

converges to a compact subset of Λp\Lambda_{p}^{\prime}, which means that

ynρ(g)ρ(γn)ϕ1(γn1g1xn)y_{n}\in\rho^{\prime}(g)\rho^{\prime}(\gamma_{n})\phi^{\prime-1}(\gamma_{n}^{-1}g^{-1}x_{n})

subconverges to a point in ρ(g)Λp\rho^{\prime}(g)\Lambda_{p}^{\prime}.

Case 2: xx is a conical limit point. We want to show that any sequence in ϕ1(xn)\phi^{\prime-1}(x_{n}) limits to ϕ1(x)\phi^{\prime-1}(x), so fix any ε>0\varepsilon>0. Using 7.12, we can choose NN so that if {αm}\{\alpha_{m}\} is any 𝒢\mathcal{G}-path limiting to xx, with corresponding vertex path {vm}\{v_{m}\}, then the diameter of

ρ(α1αN)UvN+1\rho^{\prime}(\alpha_{1}\cdots\alpha_{N})U_{v_{N+1}}

is less than ε\varepsilon with respect to a metric on Uv1U_{v_{1}}. We fix such a 𝒢\mathcal{G}-path {αm}\{\alpha_{m}\}. Then, we use 5.15 to see that for sufficiently large nn, there is a 𝒢\mathcal{G}-path {βmn}\{\beta^{n}_{m}\} limiting to xnx_{n} with βi=αi\beta_{i}=\alpha_{i} for iNi\leq N. Thus the Hausdorff distance (with respect to CUv1C_{U_{v_{1}}}) between ϕ1(xn)\phi^{\prime-1}(x_{n}) and ϕ1(x)\phi^{\prime-1}(x) is at most ε\varepsilon. Since ϕ1(xn)\phi^{\prime-1}(x_{n}) and ϕ1(x)\phi^{\prime-1}(x) both lie in the compact set ρ(α1)Uv2¯Uv1\rho^{\prime}(\alpha_{1})\overline{U_{v_{2}}}\subset U_{v_{1}}, this proves the claim.

Lemma 9.16.

ϕ\phi^{\prime} is continuous and proper.

Proof.

Since Λ\Lambda^{\prime} is compact, we just need to show continuity. Fix yΛy\in\Lambda^{\prime} and a sequence ynΛy_{n}\in\Lambda^{\prime} approaching yy. We want to show that ϕ(yn)\phi^{\prime}(y_{n}) approaches ϕ(y)=x\phi^{\prime}(y)=x.

Suppose otherwise. Since (Γ,)\partial(\Gamma,\mathcal{H}) is compact, up to a subsequence zn=ϕ(yn)z_{n}=\phi^{\prime}(y_{n}) approaches zxz\neq x. Using the equivariance of ϕ\phi^{\prime}, and cocompactness of the Γ\Gamma-action on distinct pairs in (Γ,)\partial(\Gamma,\mathcal{H}), we may assume that d(x,z)>δd(x,z)>\delta. For sufficiently large nn, we have d(x,zn)>δd(x,z_{n})>\delta as well. Then, as in the proof of 9.12, by 9.7 we know that for any vertices v,wv,w in 𝒢\mathcal{G} such that UvU_{v} contains ψρ(x)\psi_{\rho^{\prime}}(x) and UwU_{w} contains ψρ(zn)\psi_{\rho^{\prime}}(z_{n}), the intersection Uv¯Uw¯\overline{U_{v}}\cap\overline{U_{w}} is empty.

But by definition of ϕ\phi^{\prime}, we have

yψρ(x)Uv,ynψρ(zn)Uwy\in\psi_{\rho^{\prime}}(x)\subset U_{v},\qquad y_{n}\in\psi_{\rho^{\prime}}(z_{n})\subset U_{w}

for vertices v,wv,w in 𝒢\mathcal{G}. This contradicts the fact that ynyy_{n}\to y. ∎

9.5. Dynamics on the deformation

To complete the proof of Section 1.4, we just need to show:

Proposition 9.17.

The map ϕ\phi^{\prime} extends the convergence group action of Γ\Gamma.

Proof.

We will apply Section 4.3. The preceding arguments show that the relative quasi-geodesic automaton 𝒢\mathcal{G}, the map ϕ:Λ(Γ,)\phi^{\prime}:\Lambda^{\prime}\to\partial(\Gamma,\mathcal{H}), and the 𝒢\mathcal{G}-compatible system {Uv}\{U_{v}\} 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 qq be a parabolic point in (Γ,)\partial(\Gamma,\mathcal{H}), and write q=gpq=g\cdot p for pΠp\in\Pi, and then take Cq=ρ(g)CpC_{q}^{\prime}=\rho^{\prime}(g)\cdot C_{p}^{\prime}. 9.10 says that for any pΠp\in\Pi, any compact KCpK\subset C_{p}^{\prime}, and any open UG/PU\subset G/P containing Λp\Lambda_{p}^{\prime}, if γn\gamma_{n} is an infinite sequence in Γp\Gamma_{p}, then ρ(γn)KU\rho^{\prime}(\gamma_{n})\cdot K\subset U for sufficiently large nn. Then, since Γq=gΓpg1\Gamma_{q}=g\Gamma_{p}g^{-1}, the same is true for any parabolic point qq.

So, we just need to check that for each pΠp\in\Pi, CpC_{p}^{\prime} contains ΛΛp\Lambda^{\prime}-\Lambda_{p}^{\prime}. We consider the compact set Kp(Γ,){p}K_{p}\subset\partial(\Gamma,\mathcal{H})-\{p\} satisfying ΓpKp=(Γ,){p}\Gamma_{p}K_{p}=\partial(\Gamma,\mathcal{H})-\{p\}. We observed in the proof of 9.15 that CpC_{p}^{\prime} contains ϕ1(Kp)\phi^{\prime-1}(K_{p}). But then since CpC_{p}^{\prime} is ρ(Γp)\rho^{\prime}(\Gamma_{p})-invariant (by 9.10), we have

ρ(Γp)ϕ1(Kp)=ϕ1((Γ,){p})Cp.\rho^{\prime}(\Gamma_{p})\cdot\phi^{\prime-1}(K_{p})=\phi^{\prime-1}(\partial(\Gamma,\mathcal{H})-\{p\})\subset C_{p}^{\prime}.

Remark 9.18.

The definition of the set Λ\Lambda^{\prime} and the map ϕ\phi^{\prime} immediately imply that the fibers of the deformed boundary extension ϕ:Λ(Γ,)\phi^{\prime}:\Lambda^{\prime}\to\partial(\Gamma,\mathcal{H}) satisfy the conclusions of 4.7: the fiber over each conical limit point is a singleton, and the fiber over each parabolic point pp is the closure of the accumulation sets of Γp\Gamma_{p}-orbits in CpC_{p}^{\prime}. So, we obtain 4.7 by taking 𝒲\mathcal{W} to be the singleton {ρ}\{\rho\}, and following the proof of Section 1.4 (using CpC_{p} for CpC_{p}^{\prime} throughout).

Appendix A Contraction dynamics on flag manifolds

Let VV be a real vector space, and let AnA_{n} be a sequence of elements of PGL(V)\operatorname{PGL}(V). It is sometimes possible to determine the global dynamical behavior of AnA_{n} on (V)\mathbb{P}(V) by considering the action of AnA_{n} on a small open subset of (V)\mathbb{P}(V): if there is an open subset U(V)U\subset\mathbb{P}(V) such that AnUA_{n}\cdot U converges to a point in (V)\mathbb{P}(V), then in fact there is a dense open subset U(V)U_{-}\subset\mathbb{P}(V) (the complement of a hyperplane) on which AnA_{n} converges to the same point, uniformly on compacts.

A similar statement holds for the action of AnA_{n} on Grassmannians Gr(k,V)\operatorname{Gr}(k,V). These claims can be proved by considering the behavior of the singular value gaps of AnA_{n} as nn\to\infty.

In this appendix we give a general result along these lines, where we take sequences of group elements gnGg_{n}\in G for a semisimple Lie group GG with no compact factor and trivial center, and consider the limiting behavior of gng_{n} on open subsets of some flag manifold G/P+G/P^{+}, where P+P^{+} is a parabolic subgroup.

\localContractingImpliesGlobalContracting

*

We will prove 3.5 by reducing it to the case where G=PGL(d,)G=\operatorname{PGL}(d,\mathbb{R}) and P+=P1P^{+}=P_{1} is the stabilizer of [e1]Pd1G/P1[e_{1}]\in\mathbb{R}\mathrm{P}^{d-1}\simeq G/P_{1}. In this situation, P+P^{+}-divergence can be understood in terms of the behavior of the singular value gaps of the sequence gng_{n}:

Proposition A.1.

Suppose that G=PGL(d,)G=\operatorname{PGL}(d,\mathbb{R}), and let P+=P1GP^{+}=P_{1}\subset G be the stabilizer of a line in d\mathbb{R}^{d}. A sequence gnGg_{n}\in G is P1P_{1}-divergent if and only if

σ1(gn)σ2(gn),\frac{\sigma_{1}(g_{n})}{\sigma_{2}(g_{n})}\to\infty,

where σi(gn)\sigma_{i}(g_{n}) is the iith-largest singular value of gng_{n}.

For convenience, we give a proof of 3.5 in this special case.

Lemma A.2.

Let gng_{n} be a sequence in PGL(d,)\operatorname{PGL}(d,\mathbb{R}), and suppose that for a nonempty open subset UPd1U\subset\mathbb{R}\mathrm{P}^{d-1}, gnUg_{n}U converges to a point in Pd1\mathbb{R}\mathrm{P}^{d-1}. Then, the singular value gap

σ1(gn)σ2(gn)\frac{\sigma_{1}(g_{n})}{\sigma_{2}(g_{n})}

tends to \infty as nn\to\infty.

Proof.

It suffices to show that any subsequence of gng_{n} has a subsequence which satisfies the property. Using the Cartan decomposition of PGL(d,)\operatorname{PGL}(d,\mathbb{R}), we can write

gn=knankn,g_{n}=k_{n}a_{n}k_{n}^{\prime},

for kn,knK=PO(d)k_{n},k_{n}^{\prime}\in K=\operatorname{PO}(d) and ana_{n} a diagonal matrix whose diagonal entries are σ1,,σd\sigma_{1},\ldots,\sigma_{d}. Up to subsequence knk_{n} and knk_{n}^{\prime} converge respectively to k,kKk,k^{\prime}\in K. For sufficiently large nn, knUkUk_{n}^{\prime}U\cap k^{\prime}U is nonempty, so by replacing UU with kUk^{\prime}U we can assume that kn=idk_{n}^{\prime}=\mathrm{id} for all nn. Furthermore, if knanUk_{n}a_{n}U converges to a point zPd1z\in\mathbb{R}\mathrm{P}^{d-1}, then anUa_{n}U converges to k1zk^{-1}z.

So, anUa_{n}U converges to a point, and since ana_{n} 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 ζ:GPGL(V)\zeta:G\to\operatorname{PGL}(V) coming from Theorem 7.4, so that P+P^{+} maps to the stabilizer of a line \ell in VV, PP^{-} maps to the stabilizer of a hyperplane HH in VV, and gP+g1gP^{+}g^{-1}, hPh1hP^{-}h^{-1} are opposite if and only if ζ(g)\zeta(g)\ell, ζ(h)H\zeta(h)H are transverse. As in section 7, this determines embeddings ι:G/P(V)\iota:G/P\to\mathbb{P}(V) and ι:G/P(V)\iota^{*}:G/P^{-}\to\mathbb{P}(V^{*}) by

ι(gP+)=ζ(g),ι(gP)=ζ(g)H.\iota(gP^{+})=\zeta(g)\ell,\qquad\iota^{*}(gP^{-})=\zeta(g)H.

The representation ζ\zeta additionally has the property that for any sequence gnGg_{n}\in G, the singular value gaps

σ1(ζ(gn))/σ2(ζ(gn))\sigma_{1}(\zeta(g_{n}))/\sigma_{2}(\zeta(g_{n}))

are unbounded if and only if gng_{n} is P+P^{+}-divergent (see [GGKW17], Lemma 3.7).

Proof of 3.5.

By [Zim18], Lemma 4.7, there exist flags ξ1,,ξDU\xi_{1},\ldots,\xi_{D}\in U so that lifts of ι(ξi)\iota(\xi_{i}) give a basis of VV. Since gnUg_{n}\cdot U converges to a point in G/PG/P, the set

{ζ(gn)ι(ξi):1iD}\{\zeta(g_{n})\cdot\iota(\xi_{i}):1\leq i\leq D\}

converges to a single point in (V)\mathbb{P}(V).

This means that we can fix lifts ι(ξi)~V\tilde{\iota(\xi_{i})}\in V so that, up to a subsequence, ζ(gn)\zeta(g_{n}) takes the projective (D1)(D-1)-simplex

[i=1Dλiι(ξi)~:λi>0](V)\left[\sum_{i=1}^{D}\lambda_{i}\tilde{\iota(\xi_{i})}:\lambda_{i}>0\right]\subset\mathbb{P}(V)

to a point. This simplex is an open subset of (V)\mathbb{P}(V). Now we can apply A.2 to see that the sequence gng_{n} is P+P^{+}-divergent.

We now just need to check that ξ\xi is the unique P+P^{+}-limit point of gng_{n}. Choose any subsequence of gng_{n}. Then any P+P^{+}-contracting subsequence gmg_{m} of this subsequence satisfies

gm|Opp(ξ)ξg_{m}|_{\operatorname{Opp}(\xi_{-})}\to\xi^{\prime}

uniformly on compacts for some ξG/P\xi_{-}\in G/P^{-} and ξG/P+\xi^{\prime}\in G/P^{+}. But since Opp(ξ)\operatorname{Opp}(\xi_{-}) is open and dense, it intersects UU nontrivially and thus ξ=ξ\xi^{\prime}=\xi. ∎

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.