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

Characterizations of Stability via Morse Limit Sets

Jacob Garcia
(Date: November 2024)
Abstract.

Subgroup stability is a strong notion of quasiconvexity that generalizes convex cocompactness in a variety of settings. In this paper, we characterize stability of a subgroup by properties of its limit set on the Morse boundary. Given H<GH<G, both finitely generated, HH is stable exactly when all the limit points of HH are conical, or equivalently when all the limit points of HH are horospherical, as long as the limit set of HH is a compact subset of the Morse boundary for GG. We also demonstrate an application of these results in the settings of the mapping class group for a finite type surface, Mod(S)\text{Mod}(S).

1. Introduction

An important example of Kleinian groups are called convex cocompact groups. These are exactly the discrete subgroups H<Iso+(3)H<\text{Iso}^{+}(\mathbb{H}^{3}) whose orbit in 3\mathbb{H}^{3} is convex cocompact. Additionally the quotient of 3\mathbb{H}^{3} by these groups are compact Kleinian manifolds, and every infinite order element of a convex cocompact group is loxodromic. We highlight some of the other interesting properties of convex cocompact groups in the following theorem.

Theorem 1.1.

([Mar74, Sul85]) A Kleinian group H<Iso+(3)PSL2()H<\text{Iso}^{+}(\mathbb{H}^{3})\cong PSL_{2}(\mathbb{C}) is called convex cocompact if one of the following equivalent conditions hold:

  1. (1)

    HH acts cocompactly on the convex hull of its limit set ΛH\Lambda H.

  2. (2)

    Any HH-orbit in 3\mathbb{H}^{3} is quasiconvex.

  3. (3)

    Every limit point of HH is conical.

  4. (4)

    HH acts cocompactly on 3Ω\mathbb{H}^{3}\cup\Omega, where Ω=3ΛH\Omega=\partial\mathbb{H}^{3}\setminus\Lambda H. \square

However other more recent versions of this relationship have been shown. Swenson showed a generalization of this theorem for Gromov hyperbolic groups equipped with their visual boundaries [Swe01], and there has been recent interest in generalizing these relationships beyond the setting of word-hyperbolic groups. For example, convex cocompact subgroups of mapping class groups acting on Teichmüller space, equipped with the Thurston compactification, have been characterized by Farb-Mosher [FM01] as exactly the subgroups which determine Gromov hyperbolic surface group extensions. There has also been recent work done in this direction for subgroups of Out(Fn)Out(F_{n}), relating convex cocompact subgroups to hyperbolic extensions of free groups [DT17, HH18, ADT17].

There has also been interest in creating generalizations which are applicable for any finitely generated group. An important generalization comes from [DT15], where Durham and Taylor introduced stability (see Definition 2.32) to characterize convex cocompact subgroups of a mapping class group in a way which is intrinsic to the geometry of the mapping class group, and in fact, generalizes the notions of convex cocompactness to any finitely generated group. A subgroup of isometries is stable when the orbit map HXH\rightarrow X is a quasi-isometric embedding into a hyperbolic subset of XX. The concept of stability was later generalized to strongly quasiconvex subgroup, introduced in [Tra19]. We note that a subgroup is stable when it is undistorted and strongly quasiconvex.

In the Kleinian, hyperbolic, and mapping class group settings, convex cocompactness is characterized by properties of the limit set on an appropriate boundary, as show by Kent and Leininger in [KL08], and independently by Hamenstädt in [Ham05]. For an arbitrary finitely generated group, it is possible to construct a (quasi-isometric invariant) boundary called the Morse boundary, which was introduced by Cordes in [Cor17] and expanded by Cordes and Hume in [CH17]. A generalization of convex cocompactness developed by Cordes and Durham [CD17], called boundary convex cocompactness, is an exact generalization of item (1) of Theorem 1.1, in the case where HH is a proper action on an arbitrary proper geodesic metric space with a non-empty and compact limit set in the Morse boundary, see Definition 2.33.

The purpose of this paper is fully generalize item (3) of Theorem 1.1 to the setting of finitely generated groups, thereby answering [CD17, Question 1.15]. In fact, we additionally generalize some other characterizations from the hyperbolic setting found in [Swe01]. We summarize the results of this paper in the following theorem:

Theorem 1.2.

Let HH be a finitely generated group acting by isometries on a proper geodesic metric space XX. The following are equivalent:

  1. (1)

    Any HH-orbit in XX is a stable embedding of HXH\rightarrow X.

  2. (2)

    HH acts boundary convex cocompactly on XX.

  3. (3)

    Every point in ΛH\Lambda H is a conical limit point of HH, ΛH\Lambda H\neq\emptyset, and ΛH\Lambda H is a compact subset of the Morse boundary of HH.

  4. (4)

    Every point in ΛH\Lambda H is a horospherical limit point of HH, ΛH\Lambda H\neq\emptyset, and ΛH\Lambda H is a compact subset of the Morse boundary of HH.

Remark 1.3.

The result (1)(2)(1)\Leftrightarrow(2) is found in the main theorem of [CD17]. We show (3)(4)(3)\Rightarrow(4) in a combination of Proposition 3.4 and Theorem 3.8, using methods similar to [Swe01]. We show (4)(2)(4)\Rightarrow(2) in Theorem 4.3, by first showing that non-cobounded actions on the weak convex hull of ΛH\Lambda H admit a sequence of points pnp_{n} which diverge quickly from the orbit (see Lemma 4.1), but then showing that the pnp_{n} converge to an element of ΛH\Lambda H, which ultimately contradicts the conical assumption. We give an alternate proof to (2)(3)(2)\Rightarrow(3) in Proposition 4.5 which does not use the main theorem from [CD17].

A limit point in ΛH\Lambda H is conical if the limit point is accumulated by the orbit in a strong way: every geodesic ray representing the limit point gets boundedly close to the orbit. See Definitions 2.9 and 3.2. In general, a geodesic ray which is constructed from geodesic segments [x,hx][x,hx] need not stay close to the orbit of HH, even when HH is stably embedded in the hyperbolic setting. For an example, see [Swe01, Lemma 3]. A limit point in ΛH\Lambda H is horospherical if it is accumulated by the orbit in a similar way: every horoball around a geodesic ray representing the limit point intersects the orbit, see Definition 3.2.

We take a moment to provide a broad overview of stability in the recent literature. In addition to results for the mapping class group from above in [FM01, KL08, Ham05, DT15], it is also known that infinite index Morse subgroups of the mapping class group exactly coincide with stable subgroups [Kim19], and stable subgroups of mapping class groups (and more generally, stable subgroups of Morse local-to-global groups) have interesting combination theorems [RST21]. Stability has also been studied in the context of Morse local-to-global groups [Cor+22], relatively hyperbolic groups [ADT17], and hierarchically hyperbolic groups [ABD21, RST23]. It is also known that stable subgroups admit finite height [Ant+19] and that the growth series of a stable subgroup is rational [RST21]. There has also been recent work on recognizing spaces, i.e. spaces where the orbit map induces a quasi-isometric embedding, for stable subgroups [Bal+23, Zbi24].

Comparing Theorem 1.2 to Theorem 1.1, we see a cocompact action involving a domain of discontinuity in Theorem 1.1 which does not appear in Theorem 1.2. This is because the standard methods used for showing this property rely on the fact that the (Gromov-)hyperbolic boundary for a word hyperbolic group is a compactification, and thus finding the requisite compact set needed for a cocompact action boils down to finding an appropriate closed subset. In contrast, the Morse boundary usually does not compactify the underlying group, in fact the Morse boundary compactifies a finitely generated group HH if and only if HH is word hyperbolic, see [Cor17, Theorem 3.10] and [CD17, Lemma 4.1]. This leads to an open question:

Open Question 1.

Does there exist an appropriate classification of boundary convex cocompactness via an appropriate action on a domain of discontinuity analog?

For other properties in Theorem 1.2, we are able to address the need for some compactness in the Morse Boundary by assuming that the limit set of the group, ΛH\Lambda H, is compact. See Definition 2.25 and Corollary 2.27. It is not possible to remove the compactness condition in either point (3) or (4) of Theorem 1.2. For example, consider the group G=2=a,bcdG=\mathbb{Z}^{2}*\mathbb{Z}*\mathbb{Z}=\langle a,b\rangle*\langle c\rangle*\langle d\rangle with subgroup H=a,b,cH=\langle a,b,c\rangle. As discussed in [CD17, Remark 1.8], HH is isometrically embedded and convex in GG, and so every point of ΛH\Lambda H is conical with respect to HH. In fact all rays representing a point in ΛH\Lambda H travel through HH infinitely often. However HH is not hyperbolic, so HH is not stable. See [CD17, Section 1.2] for a complete discussion.

1.1. Applications

Convex cocompact subgroups of mapping class groups have been well studied, see [FM01, Ham05], but in particular conical limit point characterizations have been analyzed before. Let SS be a finite type surface, Mod(S)\text{Mod}(S) its associated mapping class group, and let 𝒯(S)\mathcal{T}(S) be its associated Teichmüller space. In [KL08, Theorem 1.2], it is shown that a subgroup HH of Mod(S)\text{Mod}(S) is convex cocompact if and only if all the limit points of HH in the Thurston compactification of 𝒯(S)\mathcal{T}(S) are conical. A combination of Theorem 1.2, [DT15, Theorem 1.1], and [CD17, Theorem 1.18] gives the following direct comparison, which uses the intrinsic geometry of Mod(S)\textup{Mod}(S) instead of the geometry of 𝒯(S)\mathcal{T}(S).

Theorem 1.4.

Let SS be a finite type surface, and let H<Mod(S)H<\textup{Mod}(S) be finitely generated. Then HH is a convex-cocompact subgroup of Mod(S)\textup{Mod}(S) if and only if every point in ΛH\Lambda H is a conical limit point of HMod(S)H\curvearrowright\textup{Mod}(S), ΛH\Lambda H\neq\emptyset, and ΛH\Lambda H is compact in the Morse boundary of Mod(S)\textup{Mod}(S).

This theorem, combined with the above result of [KL08], gives the following immediate corollary, which shows that conicality is a strong condition in the setting of mapping class groups:

Corollary 1.5.

Let SS be a finite type surface, and let H<Mod(S)H<\textup{Mod}(S) be finitely generated. The following are equivalent:

  1. (1)

    Every limit point of HH in the Morse boundary of Mod(S)\textup{Mod}(S) is a conical limit point of HMod(S)H\curvearrowright\textup{Mod}(S) and ΛH\Lambda H is compact.

  2. (2)

    Every limit point of HH in the Thurston compactification of 𝒯(S)\mathcal{T}(S) is a conical limit point of H𝒯(S)H\curvearrowright\mathcal{T}(S).

We also show that there exists a natural Mod(S)\textup{Mod}(S)-equivariant map from Mod(S)\textup{Mod}(S) to 𝒯(S)\mathcal{T}(S) which sends conical limit points of H<Mod(S)H<\textup{Mod}(S) in the Morse boundary of Mod(S)\text{Mod}(S) to conical limit point of HH in the the Thurston compactification of 𝒯(S)\mathcal{T}(S). This directly proves the implication (1)(2)(1)\Rightarrow(2) in Corollary 1.5 without requiring results of [KL08], and in fact, does not require HH to be a convex cocompact subgroup. See Theorem 5.2 for details.

Recall that Out(Fn)\text{Out}(F_{n}) denotes the group of outer automorphisms on the free group FnF_{n} with nn generators. Hamenstädt and Hensel defined convex cocompact subgroups of Out(Fn)\text{Out}(F_{n}) as subgroups which have quasi-convex orbits on the free factor graph [HH18, Definition 2]. In [DT15, Theorem 1.3], it is shown that if HOut(Fn)H\leq\text{Out}(F_{n}) is convex cocompact then HH is a stable subgroup of Out(Fn)\text{Out}(F_{n}). Combining this fact with Theorem 1.2, we get the following relationship.

Theorem 1.6.

Let n3n\geq 3. Suppose HH is a convex cocompact subgroup of Out(Fn)\textup{Out}(F_{n}) in the sense of [HH18, Definition 2]. Then every limit point of HH in the Morse boundary of Out(Fn)\textup{Out}(F_{n}) is a conical limit point of HOut(Fn)H\curvearrowright\textup{Out}(F_{n}) and ΛH\Lambda H is compact.

However, in contrast of Theorem 1.4, it is unlikely that the converse holds. Due to an announcement by Hamenstädt [Ham15], there is a classification of stable subgroups of Out(Fn)\textup{Out}(F_{n}) which shows the converse of [DT15, Theorem 1.3] does not hold.

1.2. Acknowledgments

I would like to thank my advisor Matthew Gentry Durham for their guidance and support during this project. Thanks to Elliott Vest for many conversations and for his comments on an earlier draft of this paper. I would like to thank Sam Taylor for a helpful conversation regarding Out(Fn)\text{Out}(F_{n}). I would also like to thank the referees for their helpful comments.

2. Background

We first begin by setting some notation and basic definitions. We recall that a metric space XX is proper if closed balls are compact. A path α:IX\alpha:I\rightarrow X is a geodesic if II\subseteq\mathbb{R} is a closed (potentially unbounded) interval and α\alpha preserves distances, i.e., if for all s,tIs,t\in I, d(s,t)=dX(α(s),α(t))d_{\mathbb{R}}(s,t)=d_{X}(\alpha(s),\alpha(t)). If I=[a,b]I=[a,b], we call α\alpha a geodesic segment, if I=[a,)I=[a,\infty), we call α\alpha a geodesic ray, and if I=(,)I=(-\infty,\infty) then we call α\alpha a geodesic line. Given two points x,yXx,y\in X, we use [x,y][x,y] to denote a geodesic segment starting at xx and ending at yy. If there exists a geodesic segment between any pair of points in XX, we say XX is a geodesic metric space.

Given two geodesic segments α=[x,y]\alpha=[x,y] and β=[y,z]\beta=[y,z], we denote the (speed preserving) concatenation between then as [x,y][y,z][x,y]*[y,z]. Formally, given α:[0,a]X\alpha:[0,a]\rightarrow X and β:[0,b]X\beta:[0,b]\rightarrow X with α(a)=β(0)\alpha(a)=\beta(0), we have αβ:[0,a+b]X\alpha*\beta:[0,a+b]\rightarrow X given by

αβ(t)={α(t)t[0,a],β(ta)t[a,a+b].\alpha*\beta(t)=\begin{cases}\alpha(t)&t\in[0,a],\\ \beta(t-a)&t\in[a,a+b].\\ \end{cases}

We define the concatenation analogously in the case where α\alpha is a geodesic segment and β\beta is a geodesic ray.

We use BK(p)B_{K}(p) to denote the closed ball of radius KK centered at pp, i.e. BK(p)={xX:d(p,x)K}B_{K}(p)=\{x\in X:d(p,x)\leq K\}. Given AXA\subseteq X and K0K\geq 0, we denote the K-neighborhood of AA by 𝒩K(A)={xX:d(x,A)K}\mathcal{N}_{K}(A)=\{x\in X:d(x,A)\leq K\}. Given two closed sets A,BXA,B\subseteq X, we denote the Hausdorff distance between AA and BB as

dHaus(A,B)=min{K:A𝒩K(B) and B𝒩K(A)}.d_{Haus}(A,B)=\min\{K:A\subseteq\mathcal{N}_{K}(B)\text{ and }B\subseteq\mathcal{N}_{K}(A)\}.

Finally, given a closed set AXA\subseteq X, and a point pXp\in X, we denote the closest point projection of xx to AA as

πA(p)={aA:d(a,p)=d(A,p)}.\pi_{A}(p)=\{a\in A:d(a,p)=d(A,p)\}.

We now take a moment to give the definition of a quasi-geodesic, since this term will appear frequently.

Definition 2.1.

Let II\subseteq\mathbb{R} be a closed interval XX be a metric space, and let φ:IX\varphi:I\rightarrow X. Let K1K\geq 1 and C0C\geq 0. We call φ\varphi a (K,C)(K,C)-quasi-geodesic if, for every s,tIs,t\in I, we have

1Kd(s,t)Cd(φ(s),φ(t))Kd(s,t)+C.\frac{1}{K}d(s,t)-C\leq d(\varphi(s),\varphi(t))\leq Kd(s,t)+C.

We call φ\varphi a quasi-geodesic if there exists a pair (K,C)(K,C) so that φ\varphi is a (K,C)(K,C)-quasi-geodesic.

For a more thorough treatment of quasi-geodesics and their properties, we refer the reader to [CM17].

2.1. Hyperbolic geometry

Here we provide a brief overview of the main result of [Swe01], which is a direct analog of Theorem 1.1 in the setting (Gromov)-hyperbolic geometry. Although our main results are not in the setting of hyperbolicity, many of the tools and constructions we use are inspired by the results in this setting. We begin with the definition of a δ\delta-hyperbolic space.

Definition 2.2.

Let XX be a geodesic metric space. We call XX a δ\delta-hyperbolic metric space if every geodesic triangle is δ\delta-slim, i.e., if for every x,y,zXx,y,z\in X, [x,z]𝒩δ([x,y][y,z])[x,z]\subseteq\mathcal{N}_{\delta}([x,y]\cup[y,z]). We call XX a hyperbolic space if XX is δ\delta-hyperbolic for some δ0\delta\geq 0.

One of the most useful facts in a δ\delta-hyperbolic space is that quasi-geodesics fellow-travel geodesics. This is known as the Morse lemma. A detailed proof of this lemma can be found in [BH09, Theorem III.H.1.7].

Lemma 2.3 (Morse Lemma).

Let XX be a proper, geodesic δ\delta-hyperbolic space. There exists a (non-decreasing) function N:[1,)×[0,)[0,)N:[1,\infty)\times[0,\infty)\rightarrow[0,\infty) such that, for any geodesic α\alpha and any (K,C)(K,C)-quasi-geodesic φ:[a,b]X\varphi:[a,b]\rightarrow X such that φ(a),φ(b)α\varphi(a),\varphi(b)\in\alpha, we have that φ𝒩N(K,C)(α)\varphi\subseteq\mathcal{N}_{N(K,C)}(\alpha).

An important construction associated with δ\delta-hyperbolicity is the visual boundary. For more information on the visual boundary of a hyperbolic space and it’s uses, we direct the reader to [BH09] and[KB02].

Definition 2.4.

Let XX be a proper geodesic space, and let 𝔬X{\mathfrak{o}}\in X. Let R𝔬(X)R_{\mathfrak{o}}(X) be the collection of all geodesic rays α:[a,)X\alpha:[a,\infty)\rightarrow X such that α(a)=𝔬\alpha(a)={\mathfrak{o}}. Then we can define an equivalence relation on R𝔬(X)R_{\mathfrak{o}}(X) by setting αβ\alpha\sim\beta whenever the Hausdorff distance between α\alpha and β\beta is bounded. The visual boundary of XX based at 𝔬{\mathfrak{o}} is defined to be X𝔬=R𝔬(X)/\partial_{\infty}X_{\mathfrak{o}}=R_{\mathfrak{o}}(X)/\sim. We use α()\alpha(\infty) to refer to the equivalence class of α\alpha in X𝔬\partial_{\infty}X_{\mathfrak{o}}. We equip X𝔬\partial_{\infty}X_{\mathfrak{o}} with the topology generated by the neighborhood basis for α\alpha,

U(α,r,n)={βX𝔬:d(α(t),β(t)r for all tn}.U(\alpha,r,n)=\{\beta\in\partial_{\infty}X_{\mathfrak{o}}:d(\alpha(t),\beta(t)\leq r\text{ for all }t\leq n\}.

We also present another, equivalent definition for two rays to be in the same equivalence class α()\alpha(\infty). As a note, this definition does not require either α\alpha or β\beta to be based at 𝔬{\mathfrak{o}}.

Definition 2.5.

Let (X,d)(X,d) be a proper, geodesic metric space, and let α:[a,)X\alpha:[a,\infty)\rightarrow X and β:[b,)X\beta:[b,\infty)\rightarrow X be two geodesic rays. We say α\alpha and β\beta KK-asymptotically fellow-travel, denoted by αKβ\alpha\sim_{K}\beta, if there exists TT\in\mathbb{R} so that whenever tTt\geq T, we have d(α(t),β(t))Kd(\alpha(t),\beta(t))\leq K.

Importantly, in the context of a δ\delta-hyperbolic space, Definition 2.5 classifies the tail-end fellow traveling distance in terms of only δ\delta, as expressed in the following lemma of Swenson, and is important for the definition of a horoball in a δ\delta-hyperbolic space.

Lemma 2.6.

([Swe01, Lemma 4]) Suppose α\alpha and β\beta are geodesic rays with dHaus(α,β)<d_{Haus}(\alpha,\beta)<\infty, i.e., with α()=β()\alpha(\infty)=\beta(\infty), then there exists an isometry ρ:\rho:\mathbb{R}\rightarrow\mathbb{R} so that α6δβρ\alpha\sim_{6\delta}\beta\circ\rho.

Definition 2.7.

([Swe01]) Let XX be a proper, geodesic, δ\delta-hyperbolic metric space, and let α:[a,)X\alpha:[a,\infty)\rightarrow X and β:[b,)X\beta:[b,\infty)\rightarrow X be geodesic rays.

  • We denote the horoball about α\alpha by H(α)H(\alpha) and define it as H(α)={β([b,)):β6δα,ba}H(\alpha)=\bigcup\{\beta([b,\infty)):\beta\sim_{6\delta}\alpha,~{}b\geq a\}.

  • We denote the funnel about α\alpha by F(α)F(\alpha) and define it as F(α)={xX:d(x,α)d(πα(x),α(a))}F(\alpha)=\{x\in X:d(x,\alpha)\leq d(\pi_{\alpha}(x),\alpha(a))\}.

Remark 2.8.

By Lemma 2.6, H(α)H(\alpha) is well defined.

Using these definitions, we now construct what it means for a point xXx\in\partial_{\infty}X to be a horospherical limit point or a funneled limit point of some subset AXA\subseteq X. Heuristically, xx is a horospherical limit point if every horoball around xx intersects AA. The corresponding statement is true for funneled limit points. We also take a moment to define a conical limit point.

Definition 2.9.

Let XX be a proper, geodesic δ\delta-hyperbolic space.

  • Given a point xX𝔬x\in\partial_{\infty}X_{\mathfrak{o}} and a subset AXA\subseteq X, we say xx is a horospherical limit point of AA if, for every geodesic ray α\alpha with α()=x\alpha(\infty)=x, we have H(α)AH(\alpha)\cap A\not=\emptyset.

  • Given a point xX𝔬x\in\partial_{\infty}X_{\mathfrak{o}} and a subset AXA\subseteq X, we say xx is a funneled limit point of AA if, for every geodesic ray α\alpha with α()=x\alpha(\infty)=x, we have F(α)AF(\alpha)\cap A\not=\emptyset.

  • Given a point xX𝔬x\in\partial_{\infty}X_{\mathfrak{o}} and a subset AXA\subseteq X, we say xx is a conical limit point of AA if there exists K>0K>0 such that, for every geodesic ray α\alpha with α()=x\alpha(\infty)=x, we have 𝒩K(α)A\mathcal{N}_{K}(\alpha)\cap A\not=\emptyset.

We present here for completeness a relaxed version of a claim in [Swe01, pg 125] which shows that every horoball of a geodesic ray contains a funnel of an equivalent geodesic ray in a δ\delta-hyperbolic space.

Lemma 2.10.

Let (X,d)(X,d) be a proper, geodesic, δ\delta-hyperbolic metric space, and let α:[0,)X\alpha:[0,\infty)\rightarrow X be a geodesic ray. Define α:[0,)X\alpha^{\prime}:[0,\infty)\rightarrow X by α(t)=α(t+6δ)\alpha^{\prime}(t)=\alpha(t+6\delta). Then F(α)H(α)F(\alpha^{\prime})\subseteq H(\alpha).

Proof.

See Figure 2.1. Let pF(α)p\in F(\alpha^{\prime}). We construct a geodesic ray β:[b,)X\beta:[b,\infty)\rightarrow X such that β6δα\beta\sim_{6\delta}\alpha and β(b)=p\beta(b)=p: let βn\beta_{n} be a geodesic segment which begins at pp and ends at α(n)\alpha(n). Then, after potentially passing to a subsequence, the βn\beta_{n} converge to a geodesic ray β\beta by the Arzelà–Ascoli theorem. Then as shown in [BH09, pg 427-428], β6δα\beta\sim_{6\delta}\alpha.

Let qπα(p)q\in\pi_{\alpha^{\prime}}(p) such that d(α(0),q)=min{d(α(0),x):xπα(p)}d(\alpha(0),q)=\min\{d(\alpha(0),x):x\in\pi_{\alpha^{\prime}}(p)\}, i.e., so that qq is the point in πα(p)\pi_{\alpha^{\prime}}(p) closest to α(0)\alpha(0). Notice that since pF(α)p\in F(\alpha^{\prime}) we have that d(p,q)d(q,α(0))d(p,q)\leq d(q,\alpha^{\prime}(0)). Choose T6δT\geq 6\delta so that q[α(0),α(T)]q\in[\alpha^{\prime}(0),\alpha^{\prime}(T)] and so that for all tTt\geq T, we have d(α(t),β(t))<6δd(\alpha(t),\beta(t))<6\delta. Then

Tb=d(β(T),p)\displaystyle T-b=d(\beta(T),p) d(β(T),α(T))+d(α(T),q)+d(q,p)\displaystyle\leq d(\beta(T),\alpha(T))+d(\alpha(T),q)+d(q,p)
6δ+d(α(T6δ),q)+d(q,α(0))=6δ+(T6δ).\displaystyle\leq 6\delta+d(\alpha^{\prime}(T-6\delta),q)+d(q,\alpha^{\prime}(0))=6\delta+(T-6\delta).

This shows that b0b\geq 0, and so p=β(b)H(α)p=\beta(b)\in H(\alpha). ∎

α\alphaα(0)\alpha(0)p=β(b)p=\beta(b)α(6δ)\alpha(6\delta)=α(0)=\alpha^{\prime}(0)α(T)\alpha(T)=α(T6δ)=\alpha^{\prime}(T-6\delta)β(T)\beta(T)β\betaπα(p)\pi_{\alpha^{\prime}}(p)qqF(α)F(\alpha^{\prime})H(α)H(\alpha)6δ\leq 6\delta
Figure 2.1. Diagram for Lemma 2.10

We also include the complementary statement that every funnel of a geodesic ray contains a horoball of an equivalent geodesic ray.

Lemma 2.11.

([Swe01, Lemma 5]) Let (X,d)(X,d) be a proper, geodesic, δ\delta-hyperbolic metric space, and let α:[0,)X\alpha:[0,\infty)\rightarrow X be a geodesic ray. Define α:[0,)X\alpha^{\prime}:[0,\infty)\rightarrow X by α(t)=α(t+12δ)\alpha^{\prime}(t)=\alpha(t+12\delta). Then H(α)F(α)H(\alpha^{\prime})\subseteq F(\alpha). \square

The combination of Lemmas 2.10 and 2.11 give the following relationship, which was originally stated as a corollary in [Swe01].

Corollary 2.12.

In a proper, geodesic, δ\delta-hyperbolic metric space, the funneled limit points are exactly the horospherical limit points. \square

2.2. Morse Boundary and Morse Rays

The fact that XX is δ\delta-hyperbolic is an important part of the definition of a horoball in Definition 2.7, as we note in Remark 2.8. Since our main goal of Theorem 1.2 does not have XX as a δ\delta-hyperbolic space, we will need to develop some analog to Lemma 2.6 which does not use hyperbolicity. Or strategy for creating such an analog will be to use properties of Morse rays. We begin by recalling the definition.

Definition 2.13.

([Cor17, Definition 1.3]) A (quasi)-geodesic γ\gamma in a metric space is called NN-Morse, where NN is a function [1,)×[0,)[0,)[1,\infty)\times[0,\infty)\rightarrow[0,\infty), if for any (K,C)(K,C)-quasi-geodesic φ\varphi with endpoints on γ\gamma, we have φ𝒩N(K,C)(γ)\varphi\subset\mathcal{N}_{N(K,C)}(\gamma). We call the function NN a Morse gauge. We say γ\gamma is Morse if there exists a Morse gauge NN so that γ\gamma is NN-Morse.

Comparing this definition with Lemma 2.3 shows that Morse rays are the rays in XX which have hyperbolic like properties. In the next definition, Cordes uses the visual boundary (see Definition 2.4 above) to construct a boundary on proper, geodesic XX without requiring XX to be hyperbolic.

Definition 2.14.

([Cor17, CH17]) Given a Morse gauge NN and a basepoint 𝔬X{\mathfrak{o}}\in X, the NN-Morse stratum, denoted X𝔬NX_{{\mathfrak{o}}}^{N}, is defined as the set of all points xx such that [𝔬,x][{\mathfrak{o}},x] is an NN-Morse geodesic. Each such stratum is δ\delta-hyperbolic for δ\delta depending only on NN [CH17, Proposition 3.2], and thus has a well defined visual boundary X𝔬N\partial_{\infty}X^{N}_{{\mathfrak{o}}}. If \mathcal{M} is the set of all Morse gauges, then there is a natural partial order on \mathcal{M}: NNN\leq N^{\prime} if N(K,C)N(K,C)N(K,C)\leq N^{\prime}(K,C) for all KK and CC. Note the natural inclusion X𝔬NX𝔬N\partial_{\infty}X_{{\mathfrak{o}}}^{N}\hookrightarrow\partial_{\infty}X_{{\mathfrak{o}}}^{N^{\prime}} is continuous whenever NNN\leq N^{\prime} by [Cor17, Corollary 3.2]. We define the Morse boundary based at 𝔬{\mathfrak{o}} as

X𝔬=limX𝔬N\partial X_{{\mathfrak{o}}}=\varinjlim_{\mathcal{M}}\partial_{\infty}X_{{\mathfrak{o}}}^{N}

with the induced direct limit topology. Given a Morse geodesic ray α\alpha, we denote the associated point in X𝔬\partial X_{{\mathfrak{o}}} as α()\alpha(\infty).

Remark 2.15.

Often when studying the Morse boundary, the basepoint is suppressed from the notation, as the Morse boundary is basepoint independent [Cor17, Proposition 2.5]. However, we will often make use of the basepoint explicitly in the arguments to come, thus we keep it in the notation.

Remark 2.16.

When XX is a δ\delta-hyperbolic space, X𝔬=X𝔬\partial_{\infty}X_{\mathfrak{o}}=\partial X_{\mathfrak{o}}. This is because, by the Morse Lemma (Lemma 2.3) there exists a maximum Morse gauge NN so that X=X𝔬NX=X_{\mathfrak{o}}^{N}. See [Cor17] for details.

The following fact states that subrays of Morse rays are also Morse. This will be especially useful in Section 3, as many of the arguments which describe the relationships between horoballs, funnels, and cones require restriction to a subray, as illustrated in the proof of Lemma 2.10.

Lemma 2.17.

([Liu21, Lemma 3.1]) Let XX be a geodesic metric space. Let α:IX\alpha:I\rightarrow X be an NN-Morse (λ,ϵ)(\lambda,\epsilon)-quasi-geodesic where II is an interval of \mathbb{R}. Then for any interval III^{\prime}\subseteq I, the (λ,ϵ)(\lambda,\epsilon)-quasi-geodesic α=α|I\alpha^{\prime}=\alpha|_{I^{\prime}} is NN^{\prime}-Morse where NN^{\prime} depends only on λ,ϵ\lambda,\epsilon, and NN. \square

We now present a combination of statements which will show that, given one Morse ray and another ray which fellow-travels with the first, then eventually the fellow-travelling constant is determined only by the Morse gauge of the first ray. We begin by recalling two relevant facts from [Cor17].

Proposition 2.18.

([Cor17, Proposition 2.4]) Let XX be a geodesic metric space. Let α:[0,)X\alpha:[0,\infty)\rightarrow X be an NN-Morse geodesic ray. Let β:[0,)\beta:[0,\infty)\rightarrow be a geodesic ray such that d(α(t),β(t))<Kd(\alpha(t),\beta(t))<K for t[A,A+D]t\in[A,A+D] for some A[0,)A\in[0,\infty) and D6KD\geq 6K. Then for all t[A+2K,A+D2K]t\in[A+2K,A+D-2K],

d(α(t),β(t))<4N(1,2N(5,0))+2N(5,0)+d(α(0),β(0)).\hfill d(\alpha(t),\beta(t))<4N(1,2N(5,0))+2N(5,0)+d(\alpha(0),\beta(0)).
Corollary 2.19.

([Cor17, Corollary 2.6]) Let XX be a geodesic metric space. Let α:[0,)X\alpha:[0,\infty)\rightarrow X be an NN-Morse geodesic ray. Let β:[0,)\beta:[0,\infty)\rightarrow be a geodesic ray such that d(α(t),β(t))<Kd(\alpha(t),\beta(t))<K for all t[0,)t\in[0,\infty) (i.e. β()=α()\beta(\infty)=\alpha(\infty)). Then for all t[2K,)t\in[2K,\infty),

d(α(t),β(t))<max{4N(1,2N(5,0))+2N(5,0),8N(3,0)}+d(α(0),β(0)).d(\alpha(t),\beta(t))<\max\{4N(1,2N(5,0))+2N(5,0),8N(3,0)\}+d(\alpha(0),\beta(0)).

The proof of Proposition 2.18, as presented in [Cor17], shows the following additional fact:

Corollary 2.20.

Let XX be a geodesic metric space. Let α:[0,)X\alpha:[0,\infty)\rightarrow X be an NN-Morse geodesic ray. Let β:[0,)\beta:[0,\infty)\rightarrow be a geodesic ray such that d(α(t),β(t))<Kd(\alpha(t),\beta(t))<K for t[A,A+D]t\in[A,A+D] for some A[0,)A\in[0,\infty) and D6KD\geq 6K. Then there exists x,y[0,A+2K]x,y\in[0,A+2K] such that d(α(y),β(x))<N(5,0)d(\alpha(y),\beta(x))<N(5,0).

Proof.

The value xx from the first paragraph of the proof in [Cor17, Proposition 2.4] satisfies x[max{0,A2K},A+2K]x\in[\max\{0,A-2K\},A+2K], and as [max{0,A2K},A+2K][0,A+2K][\max\{0,A-2K\},A+2K]\subseteq[0,A+2K], we get x[0,A+2K]x\in[0,A+2K]. The third to last paragraph defines yy so that α(y)πα(β(x))\alpha(y)\in\pi_{\alpha}(\beta(x)), and shows yA+2Ky\leq A+2K and that d(α(y),β(x))<N(5,0)d(\alpha(y),\beta(x))<N(5,0). ∎

Corollaries 2.20 and 2.19 combine to give the following generalization of [BH09, Chapter 3, Lemma 3.3].

Proposition 2.21.

Let XX be a geodesic metric space. Let α:[0,)X\alpha:[0,\infty)\rightarrow X be an NN-Morse geodesic ray. Let β:[0,)X\beta:[0,\infty)\rightarrow X be a geodesic ray such that d(α(t),β(t))<Kd(\alpha(t),\beta(t))<K for all t[0,)t\in[0,\infty) (i.e. β()=α()\beta(\infty)=\alpha(\infty)). Then there exists T1,T2>0T_{1},T_{2}>0 such that for all t[0,)t\in[0,\infty),

d(α(T1+t),β(T2+t))<max{4N(1,2N(5,0))+2N(5,0),8N(3,0)}+N(5,0).d(\alpha(T_{1}+t),\beta(T_{2}+t))<\max\{4N(1,2N(5,0))+2N(5,0),8N(3,0)\}+N(5,0).
Proof.

By Corollary 2.20, there exists x,y0x,y\geq 0 so that d(α(x),β(y))<N(5,0)d(\alpha(x),\beta(y))<N(5,0). Define α(t)=α(x+t)\alpha^{\prime}(t)=\alpha(x+t) and β(t)=β(y+t)\beta^{\prime}(t)=\beta(y+t), and note in particular that α(0)=α(x)\alpha^{\prime}(0)=\alpha(x) and β(0)=β(y)\beta^{\prime}(0)=\beta(y). Applying Corollary 2.19 to α\alpha^{\prime} and β\beta^{\prime} produces the desired result. ∎

For convenience, we will denote

δN=max{4N(1,2N(5,0))+2N(5,0),8N(3,0)}+N(5,0).\delta_{N}=\max\{4N(1,2N(5,0))+2N(5,0),8N(3,0)\}+N(5,0).

Using this notation, Proposition 2.21 leads to the following generalization of [Swe01, Lemma 4].

Corollary 2.22.

Let XX be a geodesic metric space. Let α:[0,)X\alpha:[0,\infty)\rightarrow X be an NN-Morse geodesic ray. Let β:[0,)X\beta:[0,\infty)\rightarrow X be a geodesic ray such that β()=α()\beta(\infty)=\alpha(\infty). Then there exists aa\in\mathbb{R} and an isometry ρ:[a,)[0,)\rho:[a,\infty)\rightarrow[0,\infty) so that αδNβρ\alpha\sim_{\delta_{N}}\beta\circ\rho.

Proof.

Apply Proposition 2.21 to find T1,T2>0T_{1},T_{2}>0 so that for all t[0,)t\in[0,\infty), d(α(T1+t),β(T2+t))<δNd(\alpha(T_{1}+t),\beta(T_{2}+t))<\delta_{N}. Then let ρ:[a,)[0,)\rho:[a,\infty)\rightarrow[0,\infty) be the unique isometry such that ρ(T1)=T2\rho(T_{1})=T_{2}. ∎

Proposition 2.23.

Suppose α:[a,)X\alpha:[a,\infty)\rightarrow X is an NN-Morse geodesic ray and β:[b,)X\beta:[b,\infty)\rightarrow X is a geodesic ray such that βδNα\beta\sim_{\delta_{N}}\alpha and α(a)=β(b)\alpha(a)=\beta(b). Then β\beta is MM-Morse where MM depends only on NN.

Proof.

It suffices to show that dHaus(α,β)Kd_{Haus}(\alpha,\beta)\leq K where K0K\geq 0 depends only on NN. Choose T>0T>0 so that d(α(t),β(t))δNd(\alpha(t),\beta(t))\leq\delta_{N} for all tTt\geq T. Note that [β(b),β(t)][β(t),α(t)][\beta(b),\beta(t)]*[\beta(t),\alpha(t)] is a (1,2δN)(1,2\delta_{N}) quasi-geodesic, so by [Cor17, Lemma 2.1], dHaus([α(a),α(t)],[β(b),β(t)][β(t),α(t)])Ld_{Haus}([\alpha(a),\alpha(t)],[\beta(b),\beta(t)]*[\beta(t),\alpha(t)])\leq L for some LL depending only on NN. But since d(α(t),β(t))δNd(\alpha(t),\beta(t))\leq\delta_{N}, we have dHaus([α(a),α(t)],[β(b),β(t)])L+δNd_{Haus}([\alpha(a),\alpha(t)],[\beta(b),\beta(t)])\leq L+\delta_{N}. ∎

The above statement leads to the following generalization, which is very similar to [Cor17, Lemma 2.8]. This statement will be useful for showing a generalization of Corollary 2.12, since our horoballs and funnels will be restricted to a single Morse stratum. See Theorem 3.8.

Proposition 2.24.

Suppose xX𝔬Nx\in\partial X_{{\mathfrak{o}}}^{N} for a Morse gauge NN. Then any geodesic ray α:[a,)X\alpha:[a,\infty)\rightarrow X with α()=x\alpha(\infty)=x is MM-Morse, where MM depends only on NN and the Morse gauge of [α(a),𝔬][\alpha(a),{\mathfrak{o}}].

Proof.

See Figure 2.2. Let β:[b,)X\beta:[b,\infty)\rightarrow X be NN-Morse with β(b)=𝔬,β()=α()\beta(b)={\mathfrak{o}},\beta(\infty)=\alpha(\infty), and let NN^{\prime} be the Morse gauge of [α(a),β(b)][\alpha(a),\beta(b)]. For each nn\in\mathbb{N}, let γn=[α(a),β(b+n)]\gamma_{n}=[\alpha(a),\beta(b+n)]. Note that β|[b,b+n]\beta|_{[b,b+n]} is Morse for some Morse gauge depending only on NN by Lemma 2.17, and so by [Cor17, Lemma 2.3], γn\gamma_{n} in N′′N^{\prime\prime}-Morse for N′′N^{\prime\prime} depending only on max{N,N}\max\{N,N^{\prime}\}. By potentially restricting to a subsequence for Arzelà-Ascoli and by [Cor17, Lemma 2.10], there exists an N′′N^{\prime\prime}-Morse geodesic ray γ\gamma with γnγ\gamma_{n}\rightarrow\gamma (uniformly on compact sets) and γ()=β()\gamma(\infty)=\beta(\infty). Then Proposition 2.23 shows that α\alpha is Morse for an appropriate Morse gauge. ∎

β\betaβ(b)\beta(b)α(a)\alpha(a)α\alphaβ(b+1)\beta(b+1)β(b+2)\beta(b+2)γ2\gamma_{2}γ1\gamma_{1}NN^{\prime}-Morseγ\gamma
Figure 2.2. Diagram for Proposition 2.24

2.3. Limit Sets and Weak Convex Hulls

We now introduce limit sets and weak convex hulls, and give some useful properties that these sets have. We use these constructions to turn subsets of XX into subsets of the Morse boundary, and vice versa.

Definition 2.25.

([CD17, Definition 3.2]) Let XX be a proper, geodesic metric space and let AXA\subseteq X. The limit set of AA, denoted as ΛA\Lambda A, is the set of points in X𝔬\partial X_{{\mathfrak{o}}} such that, for some Morse gauge NN, there exists a sequence of points (ak)AX𝔬N(a_{k})\subset A\cap X_{{\mathfrak{o}}}^{N} such that [𝔬,ak][{\mathfrak{o}},a_{k}] converges (uniformly on compact sets) to a geodesic ray α\alpha with α()=x\alpha(\infty)=x. (Note α\alpha is NN-Morse by [Cor17, Lemma 2.10].) In the case where HH acts properly by isometries on XX, we use ΛH\Lambda H to denote the limit set of H𝔬H{\mathfrak{o}}.

Remark 2.26.

By [CD17, Lemma 3.3], ΛH\Lambda H can be defined as the limit set of any orbit of HH, we merely choose the orbit H𝔬H{\mathfrak{o}} for convenience and simplicity in future arguments.

We also prove a small fact about limit sets, which is similar to [CD17, Lemma 4.1, Proposition 4.2].

Corollary 2.27.

Let XX be a proper, geodesic metric space and suppose AXA\subseteq X. If ΛAX𝔬N\Lambda A\subseteq\partial X^{N}_{{\mathfrak{o}}} for some Morse gauge NN, then ΛA\Lambda A is compact.

Proof.

By [Cor17, Proposition 3.12], this is equivalent to the condition that ΛH\Lambda H is closed. ∎

Remark 2.28.

By Corollary 2.27 and by [CD17, Lemma 4.1], the requirement that ΛH\Lambda H is compact is equivalent to the requirement that ΛH\Lambda H is contained in the boundary of a single Morse stratum.

Definition 2.29.

([Swe01, CD17]) Let XX be a proper, geodesic metric space, and let AXX𝔬A\subseteq X\cup\partial X_{{\mathfrak{o}}}. Then the weak convex hull of AA, denoted WCH(A)WCH(A), is the union of all geodesic (segments, rays, or lines) of XX which have both endpoints in AA.

We take a moment to highlight some nice interactions between the weak convex hull of a compact limit set with the Morse boundary.

Lemma 2.30.

([CD17, Proposition 4.2]) Let XX be a proper geodesic metric space and let AXA\subseteq X such that ΛAX𝔬N\Lambda A\subseteq\partial X_{{\mathfrak{o}}}^{N} for some Morse gauge NN. Then there exists a Morse gauge NN^{\prime}, depending only on NN, such that WCH(ΛA)X𝔬NWCH(\Lambda A)\subset X_{{\mathfrak{o}}}^{N^{\prime}}. \square

Lemma 2.31.

Let XX be a proper geodesic metric space and let AXA\subseteq X such that ΛAX𝔬N\Lambda A\subseteq\partial X_{{\mathfrak{o}}}^{N} for some Morse gauge NN. Then Λ(WCH(ΛA))ΛA\Lambda(WCH(\Lambda A))\subseteq\Lambda A.

Proof.

We may assume |ΛA|>1|\Lambda A|>1. Let xΛ(WCH(ΛA))x\in\Lambda(WCH(\Lambda A)). By Definition 2.25, there exists xnWCH(ΛA)x_{n}\in WCH(\Lambda A) such that [𝔬,xn][{\mathfrak{o}},x_{n}] converges to a geodesic ray γ\gamma with γ()=x\gamma(\infty)=x. We show that there exists K>0K>0 so that for all nn there exists ana_{n} with [𝔬,xn]𝒩K([𝔬,an])[{\mathfrak{o}},x_{n}]\subseteq\mathcal{N}_{K}([{\mathfrak{o}},a_{n}]). Thus, (a subsequence of) the geodesics [𝔬,an][{\mathfrak{o}},a_{n}] converge to a geodesic ray α:[0,)X\alpha:[0,\infty)\rightarrow X with α(0)=𝔬\alpha(0)={\mathfrak{o}}, and α()=γ()=x\alpha(\infty)=\gamma(\infty)=x and so xΛAx\in\Lambda A. It remains to find KK so that [𝔬,xn]𝒩K([𝔬,an])[{\mathfrak{o}},x_{n}]\subseteq\mathcal{N}_{K}([{\mathfrak{o}},a_{n}]).

Fix nn. Since xnWCH(ΛA)x_{n}\in WCH(\Lambda A), xηx\in\eta where η:(,):X\eta:(-\infty,\infty):\rightarrow X is a geodesic with η(±)ΛA\eta(\pm\infty)\in\Lambda A. So, by Definition 2.25, there exists ak+,akAX𝔬Na_{k}^{+},a_{k}^{-}\in A\cap X^{N}_{{\mathfrak{o}}} so that [𝔬,ak+][{\mathfrak{o}},a_{k}^{+}] and [𝔬,ak][{\mathfrak{o}},a_{k}^{-}] converge to geodesics β+\beta^{+} and β\beta^{-}, respectively, with β+()=η()\beta^{+}(\infty)=\eta(\infty) and β()=η()\beta^{-}(-\infty)=\eta(-\infty). Since ΛAX𝔬N\Lambda A\subseteq\partial X_{{\mathfrak{o}}}^{N}, the triangle ηβ+β\eta\cup\beta^{+}\cup\beta^{-} is LL-slim for LL depending only on NN by [CD17, Proposition 3.6], and as xηx\in\eta, there exists yβ+βy\in\beta^{+}\cup\beta^{-} so that d(xn,y)Ld(x_{n},y)\leq L. Without loss of generality, assume yβ+y\in\beta^{+}. Since [𝔬,ak+][{\mathfrak{o}},a_{k}^{+}] converges to β+\beta^{+} uniformly on compact sets, choose mm large enough so that d(y,[𝔬,am+])1d(y,[{\mathfrak{o}},a_{m}^{+}])\leq 1. Let z[𝔬,am+]z\in[{\mathfrak{o}},a_{m}^{+}] so that d(y,z)1d(y,z)\leq 1. See Figure 2.3.

Note that the concatenation [𝔬,xn][xn,z][{\mathfrak{o}},x_{n}]*[x_{n},z] is a (1,L+1)(1,L+1)-quasi-geodesic with endpoints on [𝔬,am+][{\mathfrak{o}},a_{m}^{+}]. Since [𝔬,am+][{\mathfrak{o}},a_{m}^{+}] is NN-Morse, we have that [𝔬,xn][𝔬,xn][xn,z]𝒩N(1,L+1)([𝔬,am+])[{\mathfrak{o}},x_{n}]\subseteq[{\mathfrak{o}},x_{n}]*[x_{n},z]\subseteq\mathcal{N}_{N(1,L+1)}([{\mathfrak{o}},a_{m}^{+}]). Since K:=N(1,L+1)K:=N(1,L+1) did not depend on the choice of nn, this completes the proof. ∎

η\etaβ+\beta^{+}β\beta^{-}xnx_{n}yyzzL\leq L1\leq 1am+a_{m}^{+}𝔬{\mathfrak{o}}
Figure 2.3. Diagram for Lemma 2.31

Finally, we finish this section by stating the definitions of stability and boundary convex cocompactness here for reference.

Definition 2.32.

([DT15] [CD17, Definition 1.3]) If f:XYf:X\rightarrow Y is a quasi-isometric embedding between geodesic metric spaces, we say XX is a stable subspace of YY if there exists a Morse Gauge NN such that every pair of points in XX can be connected by an NN-Morse quasigeodesic in YY; we call ff a stable embedding.

If H<GH<G are finitely generated groups, we say HH is stable in GG if the inclusion map i:HGi:H\hookrightarrow G is a stable embedding.

Definition 2.33.

([CD17, Definition 1.4]) We say that HH acts boundary convex cocompactly on XX if the following conditions hold:

  1. (1)

    HH acts properly on XX,

  2. (2)

    ΛH\Lambda H is nonempty and compact,

  3. (3)

    The action of HH on WCH(ΛH)WCH(\Lambda H) is cobounded.

3. Limit point characterizations in the Morse Boundary

The goal of this section is show that, given a set AXA\subseteq X, if xX𝔬x\in\partial X_{{\mathfrak{o}}} is a Morse conical limit point of AA, then xx is a Morse horospherical limit point of AA. This was first shown in the hyperbolic case in [Swe01], here we generalize this fact into the setting of proper geodesic metric spaces. We begin by introducing definitions which generalize horospheres and funnels for Morse rays.

Definition 3.1 (Horoballs, Funnels).

Let XX be a proper, geodesic metric space and let 𝔬X{\mathfrak{o}}\in X be some designated point. Let α:[a,)X\alpha:[a,\infty)\rightarrow X be an NN^{\prime}-Morse geodesic ray, and let NN be some, potentially different, Morse gauge. We define the NN-Morse horoball around α\alpha based at 𝔬{\mathfrak{o}} as

H𝔬N(α)={xX𝔬N|β:[b,)X with βδNα and ba and β(b)=x}.H^{N}_{{\mathfrak{o}}}(\alpha)=\{x\in X^{N}_{{\mathfrak{o}}}~{}|~{}\exists\beta:[b,\infty)\rightarrow X\text{ with }\beta\sim_{\delta_{N^{\prime}}}\alpha\text{ and }b\geq a\text{ and }\beta(b)=x\}.

We define the NN-Morse funnel around α\alpha based at 𝔬{\mathfrak{o}} as

F𝔬N(α)={xX𝔬N|d(x,πα(x))d(α(a),πα(x))}.F^{N}_{{\mathfrak{o}}}(\alpha)=\{x\in X^{N}_{{\mathfrak{o}}}~{}|~{}d(x,\pi_{\alpha}(x))\leq d(\alpha(a),\pi_{\alpha}(x))\}.

Comparing these definitions to Definition 2.7 shows that a Morse horoball is a horoball about a Morse geodesic intersected with an appropriate Morse stratum, and similarly, a Morse funnel is a funnel about a Morse geodesic intersected with an appropriate Morse stratum. The following three definitions classify points on the Morse boundary by asking if every horoball, funnel, or cone intersects a given subset of XX.

Definition 3.2.

Let XX be a proper, geodesic metric space and let 𝔬X{\mathfrak{o}}\in X be some designated point. Let AXA\subset X.

  • We say that xX𝔬x\in\partial X_{{\mathfrak{o}}} is a Morse horospherical limit point of AA if for every Morse geodesic α\alpha with α()=x\alpha(\infty)=x, there exists a Morse gauge NN such that H𝔬N(α)AH^{N}_{{\mathfrak{o}}}(\alpha)\cap A\neq\emptyset.

  • We say that xX𝔬x\in\partial X_{{\mathfrak{o}}} is a Morse funneled limit point of AA if for every Morse geodesic α\alpha with α()=x\alpha(\infty)=x, there exists a Morse gauge NN such that F𝔬N(α)AF^{N}_{{\mathfrak{o}}}(\alpha)\cap A\neq\emptyset.

  • We say that xX𝔬x\in\partial X_{{\mathfrak{o}}} is a Morse conical limit point of AA if there exists K>0K>0 such that, for every Morse geodesic α\alpha with α()=x\alpha(\infty)=x, we have that 𝒩K(α)A\mathcal{N}_{K}(\alpha)\cap A\neq\emptyset.

Remark 3.3.

Notice that, in the case where XX is a δ\delta-hyperbolic space, these definitions agree with the definitions given in Definition 2.9, as every geodesic in a δ\delta-hyperbolic space is NN-Morse for NN depending only on δ\delta. In light of this, we will use “conical limit point” instead of “Morse conical limit point” for the rest of this paper, except in cases where the difference between these definitions causes confusion. We similarly reduce “Morse horospherical limit point” and “Morse funneled limit point” to “horospherical limit point” and “funneled limit point,” respectively.

We now begin proving the new implications found in Theorem 1.2. We will first show that every conical limit point of AA is a funneled limit point of AA, and then we will show that the funneled limit points of AA exactly coincide with the horospherical limit points of AA. These arguments generalize the arguments found in [Swe01].

Proposition 3.4.

Let XX be a proper, geodesic metric space and let 𝔬X{\mathfrak{o}}\in X. Let AXA\subseteq X. If xX𝔬x\in\partial X_{{\mathfrak{o}}} is a conical limit point of AA, then xx is a funneled limit point of AA.

Proof.

See Figure 3.1. Let xX𝔬x\in\partial X_{{\mathfrak{o}}} be a conical limit point of AXA\subseteq X. Let α:[0,)X\alpha:[0,\infty)\rightarrow X be an NN-Morse geodesic with α()=x\alpha(\infty)=x. By Lemma 2.17, there exists a Morse gauge MM so that every geodesic sub-ray of α\alpha is MM-Morse. Thus by Definition 3.2, there exists K0K\geq 0 so that every subray of α\alpha gets at least KK close to AA.

Now define α=α|[3K,)\alpha^{\prime}=\alpha|_{[3K,\infty)}, and let aAa\in A such that a𝒩K(α)a\in\mathcal{N}_{K}(\alpha^{\prime}). Then note that d(a,πα(a))d(a,πα(a))Kd(a,\pi_{\alpha}(a))\leq d(a,\pi_{\alpha^{\prime}}(a))\leq K, and so d(πα(a),πα(a))2Kd(\pi_{\alpha}(a),\pi_{\alpha^{\prime}}(a))\leq 2K. By the triangle inequality, πα(a)α|[K,)\pi_{\alpha}(a)\subseteq\alpha|_{[K,\infty)}. Therefore, d(πα(a),a)K=d(α(0),α(K))d(α(0),πα(a))d(\pi_{\alpha}(a),a)\leq K=d(\alpha(0),\alpha(K))\leq d(\alpha(0),\pi_{\alpha^{\prime}}(a)). It remains to show that aX𝔬Na\in X_{{\mathfrak{o}}}^{N^{\prime}} for a Morse gauge NN^{\prime} which is independent of the choice of aAa\in A.

Let L=d(𝔬,α(0))L=d({\mathfrak{o}},\alpha(0)), and let pπα(a)p\in\pi_{\alpha^{\prime}}(a), and note that d(p,a)Kd(p,a)\leq K. Thus, [𝔬,α(0)][{\mathfrak{o}},\alpha(0)] and [p,a][p,a] are both N′′N^{\prime\prime}-Morse depending only on max{K,L}\max\{K,L\}, and [𝔬,p][{\mathfrak{o}},p] is N′′′N^{\prime\prime\prime}-Morse depending only on NN by Lemma 2.17. Since [𝔬,a][{\mathfrak{o}},a] is one side of a quadrilateral whose other three sides are max{N′′,N′′′}\max\{N^{\prime\prime},N^{\prime\prime\prime}\}-Morse, [𝔬,a][{\mathfrak{o}},a] is NN^{\prime}-Morse where NN^{\prime} does not depend on choice of aAa\in A by [Cor17, Lemma 2.3]. ∎

α()=x\alpha(\infty)=xα(0)\alpha(0)α(3K)\alpha(3K)πα(a)\pi_{\alpha^{\prime}}(a)πα(a)\pi_{\alpha}(a)𝒩K(α)\mathcal{N}_{K}(\alpha^{\prime})aaK\leq K𝔬{\mathfrak{o}}LLpp
Figure 3.1. Diagram for Proposition 3.4

Our next goal is to show that the funneled limit points of AA coincide with the horospherical limit points of AA. Towards this end, we show that, given a point xx in a horoball of a subray, the projection of xx to the subray is coarsely the same as the projection to the base ray.

Lemma 3.5.

Suppose α\alpha is an NN-Morse geodesic ray and let α\alpha^{\prime} be a subray. Suppose xH𝔬N(α)x\in H_{{\mathfrak{o}}}^{N^{\prime}}(\alpha^{\prime}). If α()X𝔬N′′\alpha(\infty)\in\partial X_{{\mathfrak{o}}}^{N^{\prime\prime}}, then dHaus(πα(x),πα(x))Kd_{Haus}(\pi_{\alpha}(x),\pi_{\alpha^{\prime}}(x))\leq K, where K0K\geq 0 depends only on NN, NN^{\prime}, and N′′N^{\prime\prime}.

Proof.

See Figure 3.2. Let α:[0,)X\alpha:[0,\infty)\rightarrow X be an NN-Morse geodesic ray and let α=α|[a,)\alpha^{\prime}=\alpha|_{[a,\infty)} for some a0a\geq 0. By Lemma 2.17, α\alpha^{\prime} is MM-Morse for MM depending only on NN. Let xH𝔬N′′(α)x\in H_{{\mathfrak{o}}}^{N^{\prime\prime}}(\alpha^{\prime}), thus there exists β:[b,)X\beta:[b,\infty)\rightarrow X a geodesic ray with bab\geq a, β(b)=x\beta(b)=x, and βδMα\beta\sim_{\delta_{M}}\alpha^{\prime}. By Proposition 2.24, β\beta is MM^{\prime}-Morse for MM^{\prime} depending only on NN, NN^{\prime}, and N′′N^{\prime\prime}. We note that if πα(x)α\pi_{\alpha}(x)\subseteq\alpha^{\prime}, then πα(x)=πα(x)\pi_{\alpha}(x)=\pi_{\alpha^{\prime}}(x). So, we assume that πα(x)α\pi_{\alpha}(x)\not\subseteq\alpha^{\prime}. We shall show that in this case, d(x,πα(x))d(x,\pi_{\alpha}(x)) and d(x,πα(x))d(x,\pi_{\alpha^{\prime}}(x)) are both bounded above by an appropriate constant, and this gives the desired result.

Let pπα(x)αp\in\pi_{\alpha}(x)\setminus\alpha^{\prime}, and let qπα(x)q\in\pi_{\alpha^{\prime}}(x). Without loss of generality, let TT be large enough so that q[α(a),α(T)]q\in[\alpha^{\prime}(a),\alpha^{\prime}(T)] and d(α(T),β(T))δNd(\alpha^{\prime}(T),\beta(T))\leq\delta_{N}.

Put γ=[β(b),α(a)][α(a),α(T)][α(T),β(T)]\gamma=[\beta(b),\alpha^{\prime}(a)]*[\alpha^{\prime}(a),\alpha^{\prime}(T)]*[\alpha^{\prime}(T),\beta(T)], and note that γ\gamma is a (3,4δN)(3,4\delta_{N}) quasi-geodesic. Thus there exists w[β(b),β(T)]w\in[\beta(b),\beta(T)] and L0L\geq 0 such that d(α(a),w)Ld(\alpha^{\prime}(a),w)\leq L, where LL depends only on MM^{\prime} by [Cor17, Lemma 2.1]. Notice now that |d(α(a),α(T))d(w,β(T))|δN+L|d(\alpha^{\prime}(a),\alpha^{\prime}(T))-d(w,\beta(T))|\leq\delta_{N}+L. However, since bab\geq a and w[β(b),β(T)]w\in[\beta(b),\beta(T)], we know |d(α(a),α(T))d(w,β(T))|=d(α(a),α(T))d(w,β(T))|d(\alpha^{\prime}(a),\alpha^{\prime}(T))-d(w,\beta(T))|=d(\alpha^{\prime}(a),\alpha^{\prime}(T))-d(w,\beta(T)). But then by the definition of the nearest point projection and the triangle inequality, we have

d(x,p)\displaystyle d(x,p) d(x,q)d(x,α(a))d(x,w)+d(w,α(a))=d(x,β(T))d(w,β(T))+d(w,α(a))\displaystyle\leq d(x,q)\leq d(x,\alpha^{\prime}(a))\leq d(x,w)+d(w,\alpha^{\prime}(a))=d(x,\beta(T))-d(w,\beta(T))+d(w,\alpha^{\prime}(a))
=d(α(b),α(T))d(w,β(T))+d(w,α(a))d(α(a),α(T))d(w,β(T))+LδN+L+L.\displaystyle=d(\alpha^{\prime}(b),\alpha^{\prime}(T))-d(w,\beta(T))+d(w,\alpha^{\prime}(a))\leq d(\alpha^{\prime}(a),\alpha^{\prime}(T))-d(w,\beta(T))+L\leq\delta_{N}+L+L.

Therefore, d(πα(x),x)d(\pi_{\alpha}(x),x) and d(πα(x),x)d(\pi_{\alpha^{\prime}}(x),x) are both bounded above by LL, which is a constant depending only on NN, NN^{\prime}, and N′′N^{\prime\prime}, as desired. ∎

α\alphaα(0)\alpha(0)H𝔬N(α)H^{N^{\prime}}_{{\mathfrak{o}}}(\alpha^{\prime})πα(x)\pi_{\alpha^{\prime}}(x)πα(x)\pi_{\alpha}(x)α(a)\alpha(a)=α(a)=\alpha^{\prime}(a)β(T)\beta(T)α(T)\alpha(T)=α(T)=\alpha^{\prime}(T)δN\leq\delta_{N}wwqqβ(b)=x\beta(b)=xppL\leq L
Figure 3.2. Diagram for Lemma 3.5

We’re now ready to show that Morse funneled limit points are exactly Morse horospherical limit points. We proceed using the same overall strategy as the one found in [Swe01], by showing direct generalizations of Lemma 2.10 and Lemma 2.11 for the Morse case.

Proposition 3.6.

Let xX𝔬Nx\in\partial X_{{\mathfrak{o}}}^{N}. Let α:[0,)X\alpha:[0,\infty)\rightarrow X be an NN^{\prime}-Morse geodesic with α()=x\alpha(\infty)=x. Then for every Morse gauge N′′N^{\prime\prime}, there exists T0T\geq 0 such that, for any subray α\alpha^{\prime} of α\alpha with d(α(0),α)Td(\alpha(0),\alpha^{\prime})\geq T, we have H𝔬N′′(α)F𝔬N′′(α)H_{{\mathfrak{o}}}^{N^{\prime\prime}}(\alpha^{\prime})\subseteq F_{{\mathfrak{o}}}^{N^{\prime\prime}}(\alpha).

Proof.

See Figure 3.3. Let α=α|[a,)\alpha^{\prime}=\alpha|_{[a,\infty)} be a subray of α\alpha. By Lemma 2.17, α\alpha^{\prime} is MM-Morse where MM depends only on NN^{\prime}. Let yH𝔬N′′(α)y\in H^{N^{\prime\prime}}_{{\mathfrak{o}}}(\alpha^{\prime}). Thus there exists β:[b,)X\beta:[b,\infty)\rightarrow X be a geodesic ray such that bab\geq a, β(b)=y\beta(b)=y, and βδMα\beta\sim_{\delta_{M}}\alpha^{\prime}. Note that β\beta is MM^{\prime}-Morse where MM^{\prime} depends only on NN, NN^{\prime}, and N′′N^{\prime\prime} by Proposition 2.24. Choose zπα(y)z\in\pi_{\alpha}(y) such that d(α(0),z)=d(α(0),πα(y))d(\alpha(0),z)=d(\alpha(0),\pi_{\alpha}(y)), i.e., so that zz is closest to α(0)\alpha(0). By Lemma 3.5, there exists pπα(x)p\in\pi_{\alpha^{\prime}}(x) so that d(z,p)Ld(z,p)\leq L for some LL depending only on NN, NN^{\prime}, and N′′N^{\prime\prime}. Choose tt large enough so that d(α(t),β(t))=d(α(t),β(t))δMd(\alpha^{\prime}(t),\beta(t))=d(\alpha(t),\beta(t))\leq\delta_{M} and p,α(b)[α(a),α(t)]p,\alpha(b)\in[\alpha(a),\alpha(t)]. Note that [y,p][p,α(t)][α(t),β(t)][y,p]*[p,\alpha(t)]*[\alpha(t),\beta(t)] is a (3,4δM)(3,4\delta_{M})-quasi-geodesic, thus there exists q[β(b),β(t)]q\in[\beta(b),\beta(t)] and λ0\lambda\geq 0 such that d(p,q)λd(p,q)\leq\lambda where λ\lambda depends only on MM^{\prime} by [Cor17, Lemma 2.1]. It suffices to show that d(y,z)d(α(0),z)d(y,z)\leq d(\alpha(0),z).

Using the triangle inequality and the definition of πα\pi_{\alpha}, we find

d(y,z)\displaystyle d(y,z) d(y,p)d(y,q)+d(q,p)d(y,q)+λ\displaystyle\leq d(y,p)\leq d(y,q)+d(q,p)\leq d(y,q)+\lambda
=d(y,β(t))d(q,β(t))+λ=d(α(b),α(t))d(q,β(t))+λ\displaystyle=d(y,\beta(t))-d(q,\beta(t))+\lambda=d(\alpha(b),\alpha(t))-d(q,\beta(t))+\lambda
d(α(a),α(t))d(p,α(t))+λ+δM+λ=d(α(a),p)+2λ+δM\displaystyle\leq d(\alpha(a),\alpha(t))-d(p,\alpha(t))+\lambda+\delta_{M}+\lambda=d(\alpha(a),p)+2\lambda+\delta_{M}
d(α(a),z)+L+2λ+δM.\displaystyle\leq d(\alpha(a),z)+L+2\lambda+\delta_{M}.

So, if aL+2λ+δMa\geq L+2\lambda+\delta_{M}, we have

d(y,z)d(α(a),z)+L+2λ+δMd(α(a),z)+d(α(0),α(a))=d(α(0),z).d(y,z)\leq d(\alpha(a),z)+L+2\lambda+\delta_{M}\leq d(\alpha(a),z)+d(\alpha(0),\alpha(a))=d(\alpha(0),z).\qed
α\alphaα(0)\alpha(0)F𝔬N(α)F^{N^{\prime}}_{{\mathfrak{o}}}(\alpha)H𝔬N(α)H^{N^{\prime}}_{{\mathfrak{o}}}(\alpha^{\prime})πα(y)\pi_{\alpha^{\prime}}(y)πα(y)\pi_{\alpha}(y)α(a)\alpha(a)=α(a)=\alpha^{\prime}(a)β(T)\beta(T)α(T)\alpha(T)=α(T)=\alpha^{\prime}(T)δM\leq\delta_{M}qqppβ(b)=y\beta(b)=yzzλ\leq\lambda
Figure 3.3. Diagram for Proposition 3.6
Proposition 3.7.

Let xX𝔬Nx\in\partial X^{N}_{{\mathfrak{o}}}. Let α:[0,)X\alpha:[0,\infty)\rightarrow X be an NN^{\prime}-Morse geodesic with α()=x\alpha(\infty)=x. Suppose S=δNS=\delta_{N^{\prime}}. Define α=α|[S,)\alpha^{\prime}=\alpha|_{[S,\infty)}. Then F𝔬N′′(α)H𝔬N′′(α)F^{N^{\prime\prime}}_{{\mathfrak{o}}}(\alpha^{\prime})\subseteq H^{N^{\prime\prime}}_{{\mathfrak{o}}}(\alpha) for any Morse gauge N′′N^{\prime\prime}.

Proof.

Let yF𝔬N′′(α)y\in F_{{\mathfrak{o}}}^{N^{\prime\prime}}(\alpha^{\prime}). By definition, d(y,πα(y))d(α(S),πα(y))d(y,\pi_{\alpha^{\prime}}(y))\leq d(\alpha(S),\pi_{\alpha^{\prime}}(y)). Let pπα(y)p\in\pi_{\alpha^{\prime}}(y) such that d(α(S),p)=d(α(S),πα(y))d(\alpha(S),p)=d(\alpha(S),\pi_{\alpha^{\prime}}(y)), i.e., let pp be the element of πα(y)\pi_{\alpha^{\prime}}(y) which is closest to α(S)\alpha(S). Then d(y,p)d(α(S),p)d(y,p)\leq d(\alpha(S),p). Construct β:[b,)X\beta:[b,\infty)\rightarrow X such that β(b)=y\beta(b)=y and βδNα\beta\sim_{\delta_{N^{\prime}}}\alpha. We want to show that b0b\geq 0. Choose T0T\geq 0 so that d(β(T),α(T))δNd(\beta(T),\alpha(T))\leq\delta_{N^{\prime}}. Then

Tb\displaystyle T-b =d(y,β(T))d(y,p)+d(p,α(T))+d(α(T),β(T))\displaystyle=d(y,\beta(T))\leq d(y,p)+d(p,\alpha(T))+d(\alpha(T),\beta(T))
d(α(S),p)+d(p,α(T))+δN=d(α(S),α(T))+δN\displaystyle\leq d(\alpha(S),p)+d(p,\alpha(T))+\delta_{N^{\prime}}=d(\alpha(S),\alpha(T))+\delta_{N^{\prime}}
=d(α(0),α(T))d(α(0),α(S))+δN=TS+δN=T.\displaystyle=d(\alpha(0),\alpha(T))-d(\alpha(0),\alpha(S))+\delta_{N^{\prime}}=T-S+\delta_{N^{\prime}}=T.

In summary, TbTT-b\leq T, but this immediately shows that 0b0\leq b, as desired. ∎

α\alphaα(0)\alpha(0)y=β(b)y=\beta(b)α(S)\alpha(S)=α(S)=\alpha^{\prime}(S)α(T)\alpha(T)β(T)\beta(T)β\betaπα(y)\pi_{\alpha^{\prime}}(y)ppF𝔬N′′(α)F^{N^{\prime\prime}}_{{\mathfrak{o}}}(\alpha^{\prime})H𝔬N′′(α)H^{N^{\prime\prime}}_{{\mathfrak{o}}}(\alpha)δN\leq\delta_{N^{\prime}}
Figure 3.4. Diagram for Proposition 3.7
Theorem 3.8.

Let xX𝔬x\in\partial X_{{\mathfrak{o}}}. Then xx is a Morse horospherical limit point of AXA\subseteq X if and only if xx is a Morse funneled limit point of AA.

Proof.

Let xX𝔬x\in\partial X_{\mathfrak{o}}. Then there exists a Morse gauge NN so that xX𝔬Nx\in\partial X^{N}_{\mathfrak{o}}. Let α\alpha be any Morse geodesic with α()=x\alpha(\infty)=x, and let H𝔬N′′(α)H^{N^{\prime\prime}}_{\mathfrak{o}}(\alpha) and F𝔬N′′(α)F^{N^{\prime\prime}}_{\mathfrak{o}}(\alpha) be a Morse horoball about α\alpha and a Morse funnel about α\alpha, respectively. By Propositions 3.6 and 3.7, there exists a subray α\alpha^{\prime} so that F𝔬N′′(α)H𝔬N′′(α)F^{N^{\prime\prime}}_{\mathfrak{o}}(\alpha^{\prime})\subseteq H^{N^{\prime\prime}}_{\mathfrak{o}}(\alpha) and H𝔬N′′(α)F𝔬N′′(α)H^{N^{\prime\prime}}_{\mathfrak{o}}(\alpha^{\prime})\subseteq F^{N^{\prime\prime}}_{\mathfrak{o}}(\alpha).

Now suppose xx is horospherical. Then there exists aAa\in A so that xH𝔬N′′(α)F𝔬N′′(α)x\in H^{N^{\prime\prime}}_{\mathfrak{o}}(\alpha^{\prime})\subseteq F^{N^{\prime\prime}}_{\mathfrak{o}}(\alpha), and as the funnel F𝔬N′′(α)F^{N^{\prime\prime}}_{\mathfrak{o}}(\alpha) was arbitrary, xx is funneled. Similarly, suppose xx is funneled. Then there exists aAa\in A so that xF𝔬N′′(α)H𝔬N′′(α)x\in F^{N^{\prime\prime}}_{\mathfrak{o}}(\alpha^{\prime})\subseteq H^{N^{\prime\prime}}_{\mathfrak{o}}(\alpha), and as the funnel F𝔬N′′(α)F^{N^{\prime\prime}}_{\mathfrak{o}}(\alpha) was arbitrary, xx is horospherical. ∎

4. Limit Set Conditions For Stability

In this section, we show that the horospherical limit point condition, combined with the limit set being compact, is enough for to show that the group action on the weak convex hull is cobounded. The main idea behind this argument is to show the contrapositive: when the group action is not cobounded, then geodesic rays in the space eventually end up very far from the orbit of the group. We begin by showing the following helpful fact, which states that if a group acts non-coboundedly on the weak convex hull of its limit set, there exists a sequence of points pnp_{n} in the weak convex hull that “maximally avoids” the orbit.

Lemma 4.1.

Suppose XX is a proper geodesic metric space and suppose that HH acts properly on XX by isometries. Assume that ΛH\Lambda H\neq\emptyset. If the action HWCH(ΛH)H\curvearrowright WCH(\Lambda H) is not cobounded, then there exists an increasing sequence of positive integers, (ni)i(n_{i})_{i}, such that for each i1i\in\mathbb{Z}_{\geq 1} there exists piWCH(ΛH)p_{i}\in WCH(\Lambda H) satisfying

  1. (1)

    Bni(pi)H𝔬=B_{n_{i}}(p_{i})\cap H{\mathfrak{o}}=\emptyset,

  2. (2)

    d(pi,𝔬)ni+1d(p_{i},{\mathfrak{o}})\leq n_{i}+1.

Proof.

Set n0=1n_{0}=1. We define qiq_{i} and nin_{i} for i1i\geq 1 via an inductive process. Since the action of HWCH(ΛH)H\curvearrowright WCH(\Lambda H) is not cobounded, there exists a point qiWCH(ΛH)q_{i}\in WCH(\Lambda H) such that ni1+1<d(H𝔬,qi)n_{i-1}+1<d(H{\mathfrak{o}},q_{i}). By the definition of WCH(Λ)WCH(\Lambda), there exists a bi-infinite Morse geodesic γ\gamma with γ(±)ΛH\gamma(\pm\infty)\in\Lambda H such that qiγq_{i}\in\gamma. Set nin_{i} to be the unique positive integer such that ni<d(H𝔬,qi)ni+1n_{i}<d(H{\mathfrak{o}},q_{i})\leq n_{i}+1. Note that the sequence (ni)i(n_{i})_{i} is increasing because ni1+1nin_{i-1}+1\leq n_{i}.

Since d(H𝔬,qi)ni+1d(H{\mathfrak{o}},q_{i})\leq n_{i}+1 there exists hiHh_{i}\in H so that d(qi,hi𝔬)ni+1d(q_{i},h_{i}{\mathfrak{o}})\leq n_{i}+1. Recalling that the action of HH on XX is by isometries, we define pi=hi1qip_{i}=h_{i}^{-1}q_{i}, and so Bni(pi)H𝔬=B_{n_{i}}(p_{i})\cap H{\mathfrak{o}}=\emptyset, and d(𝔬,pi)ni+1d({\mathfrak{o}},p_{i})\leq n_{i}+1. Finally, by [CD17, Lemma 3.3], hi1γh_{i}^{-1}\gamma is a bi-infinite Morse geodesic with endpoints in ΛH\Lambda H, and so piWCH(ΛH)p_{i}\in WCH(\Lambda H). ∎

We note that, under the additional assumption that ΛH\Lambda H is compact and that every point in ΛH\Lambda H is conical, we get a stronger conclusion to this lemma, namely, we can take ni=in_{i}=i for large ii. We formally state and prove this observation.

Lemma 4.2 (Sliding Spheres).

Suppose XX is a proper geodesic metric space, and suppose that HH acts properly on XX by isometries. Assume that ΛH\Lambda H\neq\emptyset, every point of ΛH\Lambda H is a conical limit point of H𝔬H{\mathfrak{o}}, and that ΛHX𝔬N\Lambda H\subseteq\partial X_{{\mathfrak{o}}}^{N} for some Morse gauge NN. If the action HWCH(ΛH)H\curvearrowright WCH(\Lambda H) is not cobounded, there exists a sequence of points pnWCH(ΛH)p_{n}\in WCH(\Lambda H) such that, for sufficiently large nn, Bn(pn)H𝔬=B_{n}(p_{n})\cap H{\mathfrak{o}}=\emptyset and 𝔬Bn+1(pn){\mathfrak{o}}\in B_{n+1}(p_{n}).

Proof.

Let K>0K>0 be the conical limit point constant. Let nn\in\mathbb{N} with n>K+1n>K+1. By [Liu21, Corollary 5.8], we may assume that ΛH\Lambda H has at least two distinct points. Since HWCH(ΛH)H\curvearrowright WCH(\Lambda H) is not cobounded, there exists pWCH(ΛH)p\in WCH(\Lambda H) with d(p,H𝔬)>nd(p,H{\mathfrak{o}})>n. By definition, pγp\in\gamma for some bi-infinite geodesic γ\gamma with γ(±)ΛH\gamma(\pm\infty)\in\Lambda H. Since ΛHX𝔬N\Lambda H\subseteq\partial X_{{\mathfrak{o}}}^{N}, we have by [CD17, Proposition 4.2] that γ\gamma is is Morse for some Morse gauge depending only on NN. Since every point in ΛH\Lambda H is a conical limit point of H𝔬H{\mathfrak{o}}, there exists hHh^{\prime}\in H such that d(h𝔬,γ)<Kd(h^{\prime}{\mathfrak{o}},\gamma)<K. Put qπγ(h𝔬)q\in\pi_{\gamma}(h^{\prime}{\mathfrak{o}}).

We may assume that γ(s)=q\gamma(s)=q and γ(s)=p\gamma(s^{\prime})=p with s<ss<s^{\prime}. Let A={r[s,s]:n<d(γ(r),H𝔬)}.A=\{r\in[s,s^{\prime}]~{}:~{}n<d(\gamma(r),H{\mathfrak{o}})\}. (Equivalently, one may define A={r[s,s]:Bn(γ(r))H𝔬=}A=\{r\in[s,s^{\prime}]~{}:~{}B_{n}(\gamma(r))\cap H{\mathfrak{o}}=\emptyset\}.) Note that sAs^{\prime}\in A. Put t=infAt=\inf A. By the definition of tt, we have nd(γ(t),H𝔬)n\leq d(\gamma(t),H{\mathfrak{o}}). See Figure 4.1. We now claim that d(γ(t),H𝔬)<n+1d(\gamma(t),H{\mathfrak{o}})<n+1.

Suppose for contradiction that n+1d(γ(t),H𝔬)n+1\leq d(\gamma(t),H{\mathfrak{o}}). By the triangle inequality, nd(γ(t1),H𝔬)n\leq d(\gamma(t-1),H{\mathfrak{o}}). So if t1[s,s]t-1\in[s,s^{\prime}], then t1At-1\in A, however t=infAt=\inf A. Thus t1[s,s]t-1\not\in[s,s^{\prime}]. Therefore, t[s,s+1]t\in[s,s+1], and so

n+1d(γ(t),h𝔬)d(γ(t),h𝔬)d(γ(t),q)+d(q,h𝔬)=d(γ(t),γ(s))+d(q,h𝔬)1+Kn,n+1\leq d(\gamma(t),h{\mathfrak{o}})\leq d(\gamma(t),h^{\prime}{\mathfrak{o}})\leq d(\gamma(t),q)+d(q,h^{\prime}{\mathfrak{o}})=d(\gamma(t),\gamma(s))+d(q,h^{\prime}{\mathfrak{o}})\leq 1+K\leq n,

a contradiction.

Therefore, there exists hHh\in H such that h𝔬Bn+1(γ(t))h{\mathfrak{o}}\in B_{n+1}(\gamma(t)), but Bn(γ(t))H𝔬=B_{n}(\gamma(t))\cap H{\mathfrak{o}}=\emptyset. Put pn=h1(γ(t))p_{n}=h^{-1}(\gamma(t)). By [CD17, Lemma 3.3], hγh\gamma is a bi-infinite Morse geodesic with endpoints in ΛH\Lambda H, and since the action of HH on XX is by isometries, Bn(pn)H𝔬=B_{n}(p_{n})\cap H{\mathfrak{o}}=\emptyset, and 𝔬Bn+1(pn){\mathfrak{o}}\in B_{n+1}(p_{n}), as desired. ∎

nnnnγ(A)\gamma(A)γ(s)=p\gamma(s^{\prime})=pγ(t)\gamma(t)h𝔬h{\mathfrak{o}}γ(s)=q\gamma(s)=qh𝔬h^{\prime}{\mathfrak{o}}H𝔬H{\mathfrak{o}}
Figure 4.1. Diagram for Lemma 4.2. We can think of this proof as sliding the ball on the right towards the left until it is “up against” the orbit H𝔬H{\mathfrak{o}}, such as the ball centered at γ(t)\gamma(t).

We now prove that (4) implies (2) in the language of Theorem 1.2. We show that, if the action is not cobounded on the weak convex hull, then using Lemma 4.1 we can find a sequence of points pip_{i} which maximally avoid the orbit of HH. However, this sequence of points defines a new ray γ\gamma with γ()ΛH\gamma(\infty)\in\Lambda H. Then using the horospherical point assumption, we find an orbit point close to pip_{i}, a contradiction.

Theorem 4.3.

Suppose XX is a proper geodesic metric space and suppose HH acts properly on XX by isometries. Assume that ΛH\Lambda H\neq\emptyset, every point of ΛH\Lambda H is a horospherical limit point of H𝔬H{\mathfrak{o}}, and that there exists a Morse gauge NN such that ΛHX𝔬N\Lambda H\subset\partial X^{N}_{{\mathfrak{o}}}. Then the action of HWCH(ΛH)H\curvearrowright WCH(\Lambda H) is cobounded.

Proof.

For contradiction, assume that HWCH(ΛH)H\curvearrowright WCH(\Lambda H) is not a cobounded action. By Lemma 4.1, there exists a sequence of points piWCH(ΛH)p_{i}\in WCH(\Lambda H) and an increasing sequence of positive integers (ni)i(n_{i})_{i} such that Bni(pi)H𝔬=B_{n_{i}}(p_{i})\cap H{\mathfrak{o}}=\emptyset, and 𝔬Bni+1(pi){\mathfrak{o}}\in B_{n_{i}+1}(p_{i}). Let γi:[0,d(0,pi)]X\gamma_{i}:[0,d(0,p_{i})]\rightarrow X be a geodesic connecting 𝔬{\mathfrak{o}} and pip_{i} with γi(0)=𝔬\gamma_{i}(0)={\mathfrak{o}}. Notice that since ΛHX𝔬N\Lambda H\subset\partial X^{N}_{{\mathfrak{o}}}, we have that γi\gamma_{i} is NN^{\prime}-Morse for some NN^{\prime} depending only on NN. By restricting to a subsequence, we may assume that γi\gamma_{i} converges, uniformly on compact subsets, to an NN^{\prime}-Morse geodesic ray γ\gamma with γ(0)=𝔬\gamma(0)={\mathfrak{o}}.

By construction and by Lemma 2.31, γ()Λ(WCH(ΛH))ΛH\gamma(\infty)\in\Lambda(WCH(\Lambda H))\subseteq\Lambda H. So, by [Cor17, Corollary 2.6], there exists an NN-Morse geodesic ray α\alpha with α(0)=𝔬\alpha(0)={\mathfrak{o}} and d(α(t),γ(t))<Dd(\alpha(t),\gamma(t))<D for all t0t\geq 0, where D0D\geq 0 is a constant that depends only on NN. Let T=2D+4T=2D+4, and put α=α[T,)\alpha^{\prime}=\alpha_{[T,\infty)}. Since α()ΛH\alpha^{\prime}(\infty)\in\Lambda H, and so by Theorem 3.8, α()\alpha^{\prime}(\infty) is a funneled limit point of HH. Thus there exists hHh\in H so that h𝔬F𝔬N(α)h{\mathfrak{o}}\in F_{{\mathfrak{o}}}^{N}(\alpha^{\prime}). Let t0=min{s:α(s)πα(h𝔬)}t_{0}=\min\{s:\alpha^{\prime}(s)\in\pi_{\alpha^{\prime}}(h{\mathfrak{o}})\}. Since the sequence γi\gamma_{i} converges uniformly on compact sets to γ\gamma, we may choose ii large enough so that d(γi(t0),γ(t0))1d(\gamma_{i}(t_{0}),\gamma(t_{0}))\leq 1. See Figure 4.2.

α()\alpha(\infty)γ()\gamma(\infty)𝔬{\mathfrak{o}}pip_{i}γi(t0)\gamma_{i}(t_{0})γ(t0)\gamma(t_{0})α(t0)\alpha(t_{0})h𝔬h{\mathfrak{o}}α(T)\alpha(T)2D+42D+41\leq 1D\leq Dd(α(T),α(t0))\leq d(\alpha(T),\alpha(t_{0}))
Figure 4.2. Diagram for Theorem 4.3

By the triangle inequality we have that d(γi(t0),α(t0))D+1d(\gamma_{i}(t_{0}),\alpha^{\prime}(t_{0}))\leq D+1, and therefore |d(0,γi(t0))d(0,α(t0))|D+1|d(0,\gamma_{i}(t_{0}))-d(0,\alpha(t_{0}))|\leq D+1. Also, by construction we have that d(h𝔬,α(t0))d(α(T),α(t0))d(h{\mathfrak{o}},\alpha^{\prime}(t_{0}))\leq d(\alpha(T),\alpha(t_{0})). Therefore we have

d(pi,h𝔬)\displaystyle d(p_{i},h{\mathfrak{o}}) d(pi,γi(t0))+d(γi(t0),α(t0))+d(α(t0),h0)\displaystyle\leq d(p_{i},\gamma_{i}(t_{0}))+d(\gamma_{i}(t_{0}),\alpha^{\prime}(t_{0}))+d(\alpha(^{\prime}t_{0}),h_{0})
d(0,γi(t0))+(D+1)+d(α(T),α(t0))\displaystyle\leq d(0,\gamma_{i}(t_{0}))+(D+1)+d(\alpha(T),\alpha(t_{0}))
=d(𝔬,pi)d(𝔬,γi(t0))+(D+1)+d(0,α(t0))d(0,α(T))\displaystyle=d({\mathfrak{o}},p_{i})-d({\mathfrak{o}},\gamma_{i}(t_{0}))+(D+1)+d(0,\alpha(t_{0}))-d(0,\alpha(T))
(ni+1)+(D+1)+(D+1)(2D+4)ni1.\displaystyle\leq(n_{i}+1)+(D+1)+(D+1)-(2D+4)\leq n_{i}-1.

However, this contradicts the assumption that Bni(pi)H𝔬=B_{n_{i}}(p_{i})\cap H{\mathfrak{o}}=\emptyset. ∎

We now present an alternate definition of a conical limit point which agrees with Definition 3.2 in the case where ΛA\Lambda A is compact, and requires us to only consider of the geodesic rays which emanate from the given basepoint. By Corollary 2.27 and by [CD17, Lemma 4.1], the requirement that ΛH\Lambda H is compact is equivalent to the requirement that ΛH\Lambda H is contained in the boundary of a single Morse stratum.

Proposition 4.4.

Let XX be a proper, geodesic metric space. Let YXY\subseteq X. Suppose ΛY\Lambda Y\not=\emptyset. Then the following are equivalent:

  1. (1)

    xX𝔬x\in\partial X_{{\mathfrak{o}}} is a conical limit point of YY

  2. (2)

    There exists K>0K>0 such that, for every NN-Morse geodesic ray α:[0,)X\alpha:[0,\infty)\rightarrow X with α(0)=𝔬\alpha(0)={\mathfrak{o}} and α()=x\alpha(\infty)=x, and for every T>0T>0, there exists yYy\in Y such that y𝒩K(α)y\in\mathcal{N}_{K}(\alpha^{\prime}), where α:[0,)X\alpha^{\prime}:[0,\infty)\rightarrow X is defined by α(t)=α(t+T)\alpha^{\prime}(t)=\alpha(t+T).

Proof.

Showing that (1) implies (2) is a direct consequence of Lemma 2.17 and Definition 3.2.

Instead assume (2). Let β:[b,)X\beta:[b,\infty)\rightarrow X be an NN^{\prime}-Morse ray with β()=x\beta(\infty)=x. Let α:[0,)X\alpha:[0,\infty)\rightarrow X an NN-Morse geodesic ray with α(0)=𝔬\alpha(0)={\mathfrak{o}} and α()=x\alpha(\infty)=x. Without loss of generality, by Cor 2.22 there exists T>0T>0 such that d(α(t),β(t))<δNd(\alpha(t),\beta(t))<\delta_{N} for all t>Tt>T. Put α:[0,)X\alpha^{\prime}:[0,\infty)\rightarrow X via α(t)=α(t+T)\alpha^{\prime}(t)=\alpha(t+T). By hypothesis, there exists yYy\in Y such that y𝒩K(α)y\in\mathcal{N}_{K}(\alpha^{\prime}). Say s[0,)s\in[0,\infty) such that d(α(s),y)<Kd(\alpha^{\prime}(s),y)<K, so via the triangle inequality we have

d(β(s+T),y)d(β(s+t),α(s+t))+d(α(s),y)δN+K.d(\beta(s+T),y)\leq d(\beta(s+t),\alpha(s+t))+d(\alpha^{\prime}(s),y)\leq\delta_{N}+K.

Thus, y𝒩K+δN(β)y\in\mathcal{N}_{K+\delta_{N}}(\beta), which shows (1). ∎

We conclude this section by showing that (3)(2)(3)\Rightarrow(2) for Theorem 1.2, which was first shown in [CD17, Corollary 1.14], however here we present a direct proof that does not rely on [CD17, Theorem 1.1].

Proposition 4.5.

Let XX be a proper geodesic metric space and let HH be a finitely generated group of isometries of XX such that the orbit map HXH\rightarrow X via hh𝔬h\mapsto h{\mathfrak{o}} is a stable mapping. If there exists a Morse gauge NN so that ΛHX𝔬N\Lambda H\subseteq\partial X_{{\mathfrak{o}}}^{N}, then every xΛHx\in\Lambda H is a conical limit point of H𝔬H{\mathfrak{o}}.

Proof.

Let xΛHx\in\Lambda H, and let α:[0,)X\alpha:[0,\infty)\rightarrow X be an NN-Morse geodesic ray with α()=x\alpha(\infty)=x, α(0)=𝔬\alpha(0)={\mathfrak{o}}. Let α=α|[a,)\alpha^{\prime}=\alpha|_{[a,\infty)} be a subray of α\alpha. Notice that α\alpha^{\prime} is NN^{\prime}-Morse where NN^{\prime} depends only on NN by Lemma 2.17. By Proposition 4.4, it suffices to show that there exists some K0K\geq 0, depending only on NN^{\prime} and HH, so that H𝔬𝒩K(α)H{\mathfrak{o}}\cap\mathcal{N}_{K}(\alpha^{\prime})\not=\emptyset.

Since HH is a stable subgroup of isometries on XX, we have that for any hHh\in H, there exists a (λ,λ)(\lambda,\lambda)-quasi-geodesic γ\gamma from 𝔬{\mathfrak{o}} to h𝔬h{\mathfrak{o}} such that, for any pγp\in\gamma, B2λ(p)H𝔬B_{2\lambda}(p)\cap H{\mathfrak{o}}\not=\emptyset. (To find such a path γ\gamma, take a geodesic in a Cayley graph for HH and embed it into XX by extending the orbit map along appropriate geodesic segments.)

Now, since xΛHx\in\Lambda H, there exists a sequence hnHh_{n}\in H such that the sequence of geodesic segments, βn=[𝔬,hn𝔬]\beta_{n}=[{\mathfrak{o}},h_{n}{\mathfrak{o}}], converges (uniformly on compact subsets) to a geodesic ray β:[b,)X\beta:[b,\infty)\rightarrow X with β()=x\beta(\infty)=x and β(b)=𝔬\beta(b)={\mathfrak{o}}. Since HH is a stable group of isometries, βn\beta_{n} is N′′N^{\prime\prime}-Morse by Definition 2.32. Up to potentially re-parameterizing β\beta, there exists T>aT>a so that d(β(T),α(T))<δNd(\beta(T),\alpha(T))<\delta_{N} by Corollary 2.22.

Since βn\beta_{n} converges to β\beta uniformly on BT+1(𝔬)¯\overline{B_{T+1}({\mathfrak{o}})}, the ball of radius T+1T+1 centered at 𝔬{\mathfrak{o}}, there exists nn\in\mathbb{N} and pβnp\in\beta_{n} so that d(β(T),p)<1d(\beta(T),p)<1. Since γn\gamma_{n} is an (λ,λ)(\lambda,\lambda)-quasi-geodesic with endpoints on βn\beta_{n}, there exists qγnq\in\gamma_{n} so that d(p,q)N′′(λ,λ)d(p,q)\leq N^{\prime\prime}(\lambda,\lambda). Finally, there exists hHh\in H so that d(h𝔬,q)λd(h{\mathfrak{o}},q)\leq\lambda.

Therefore by the triangle inequality,

d(α(T),h𝔬)d(α(T),β(T))+d(β(T),p)+d(p,q)+d(q,h𝔬)δN+1+N′′(λ,λ)+2λ.d(\alpha(T),h{\mathfrak{o}})\leq d(\alpha(T),\beta(T))+d(\beta(T),p)+d(p,q)+d(q,h{\mathfrak{o}})\leq\delta_{N}+1+N^{\prime\prime}(\lambda,\lambda)+2\lambda.

As α(T)α\alpha(T)\in\alpha^{\prime}, this completes the proof. ∎

5. Applictations to Teichmüller Space

We conclude by illustrating applications of the above work in the setting of Teichmüller space for a finite type surface SS. We begin by setting some notation. Let Mod(S)\text{Mod}(S) denote the mapping class group of SS, i.e. the group of orientation-preserving homeomorphisms on SS up to isotopy equivalence, which may permute punctures but fixes boundaries pointwise. Let 𝒯(S)\mathcal{T}(S) denote the associated Teichmüller space, equipped with the Teichmüller metric. We will denote the set of projective measured foliations on SS by PMF(S)\text{PMF}(S). The Thurston compactification of Teichmüller space is 𝒯(S)¯=𝒯(S)PMF(S)\overline{\mathcal{T}(S)}=\mathcal{T}(S)\cup\text{PMF}(S). For a thorough overview of the mapping class group, it’s associated Teichmüller space, and projective measured foliations, we refer the reader to [MM00, FLP12, FM11, Beh06].

We take a moment to restate Corollary 1.5 using the above notation:

Corollary 5.1.

(Restatement of Corollary 1.5.) Let HH be a finitely generated subgroup of Mod(S)\textup{Mod}(S). The following are equivalent:

  1. (1)

    Every element of ΛHMod(S)\Lambda H\subset\partial\textup{Mod}(S) is a conical limit point of HMod(S)H\curvearrowright\textup{Mod}(S) and ΛH\Lambda H is compact (in the Morse boundary of Mod(S)\textup{Mod}(S)).

  2. (2)

    Every element of ΛHPMF(S)\Lambda H\subset\textup{PMF}(S) is a conical limit point of H𝒯(S)H\curvearrowright\mathcal{T}(S).

By work of Cordes, there exists a homeomorphism g:Mod(S)𝒯(S)g_{\infty}:\partial\text{Mod}(S)\rightarrow\partial\mathcal{T}(S) (where \partial refers to the Morse boundary) [Cor17, Theorem 4.12], and there exists a natural continuous injective map h:𝒯(S)PMF(S)h_{\infty}:\partial\mathcal{T}(S)\hookrightarrow\text{PMF}(S) [Cor17, Proposition 4.14]. We denote the continuous inclusion formed by the composition of gg_{\infty} and hh_{\infty} as f:Mod(S)PMF(S)f_{\infty}:\partial\textup{Mod}(S)\hookrightarrow\text{PMF}(S). The purpose of this section is to prove the following theorem.

Theorem 5.2.

Let HH be a subgroup of Mod(S)\textup{Mod}(S), and let xΛHMod(S)x_{\infty}\in\Lambda H\subseteq\partial\textup{Mod}(S) be a conical limit point of HMod(S)H\curvearrowright\textup{Mod}(S). Then f(x)PMF(S)f_{\infty}(x_{\infty})\in\textup{PMF}(S) is a conical limit point of H𝒯(S)H\curvearrowright\mathcal{T}(S).

Remark 5.3.

This theorem directly proves (1)(2)(1)\Rightarrow(2) of Corollary 1.5.

Our proof of Theorem 5.2 uses several of the tools developed in [Cor17], so we take a moment to recall the construction and definitions presented therein and from [MM00]. The curve graph, denoted 𝒞(S)\mathcal{C}(S), is a locally infinite simplicial graph whose vertices are isotopy classes of simple closed curves on SS. We join two vertices with an edge it there exists representative from each class that are disjoint.

A set of (pairs of) curves μ={(α1,β1),(α2,β2),,(αm,βm)}\mu=\{(\alpha_{1},\beta_{1}),(\alpha_{2},\beta_{2}),\dots,(\alpha_{m},\beta_{m})\} is called a complete clean marking of SS if the {α1,,αm}\{\alpha_{1},\dots,\alpha_{m}\} forms a pants decomposition of SS, if each αi\alpha_{i} is disjoint from βj\beta_{j} whenever iji\not=j, and if each αi\alpha_{i} intersects βi\beta_{i} once if the surface filled by αi\alpha_{i} and βi\beta_{i} is a one-punctured torus. (Otherwise, αi\alpha_{i} and βi\beta_{i} will intersect twice, and the filling surface is a four-punctured sphere.) We call {α1,,αm}\{\alpha_{1},\dots,\alpha_{m}\} the base of μ\mu and we call βi\beta_{i} the transverse curve to αi\alpha_{i} in μ\mu. For the sake of completeness, we also define the marking graph, (S)\mathcal{M}(S), although the definition is not needed in this paper.

(S)\mathcal{M}(S) is the simplicial graph whose vertices are markings as defined above, and two markings are joined by an edge of length one if they differ by an elementary move: either twisting βi\beta_{i} around αi\alpha_{i} by a full, or when possible, by a half twist, or by swapping βi\beta_{i} and αi\alpha_{i}. Note that, after performing an elementary move, one may need to replace the curves with isotopically equivalent curves to create a valid marking again. The marking graph (S)\mathcal{M}(S) is quasi-isometric to the mapping class group Mod(S)\textup{Mod}(S), see [MM00] and [Beh06].

For each σ𝒯(S)\sigma\in\mathcal{T}(S) there is a short marking, which is constructed inductively by picking the shortest curves in σ\sigma for the base and repeating for the transverse curves. Now define a map Υ:(S)𝒯(S)\Upsilon:\mathcal{M}(S)\rightarrow\mathcal{T}(S) by taking a marking μ\mu to the region in the ϵ\epsilon-thick part of 𝒯(S)\mathcal{T}(S), denoted 𝒯ϵ(S)\mathcal{T}_{\epsilon}(S), where μ\mu is a short marking in that region. As stated in [Cor17], it is a well known fact that Υ\Upsilon is a coarsely well defined map which is coarsely Lipschitz. We take a moment to prove that this map is coarsely equivariant.

Lemma 5.4.

Let Υ:(S)𝒯(S)\Upsilon:\mathcal{M}(S)\rightarrow\mathcal{T}(S) be as above, and let H<Mod(S)H<\textup{Mod}(S) be finitely generated. Then there exists a constant K0K\geq 0 such that, for any marking μ\mu\in\mathcal{M} and for any hHh\in H,

d𝒯(S)(hΥ(μ),Υ(hμ))K.d_{\mathcal{T}(S)}(h\Upsilon(\mu),\Upsilon(h\mu))\leq K.
Proof.

Let μ={(α1,β1),,(αm,βm)}(S)\mu=\{(\alpha_{1},\beta_{1}),\dots,(\alpha_{m},\beta_{m})\}\in\mathcal{M}(S) and hHh\in H be arbitrary. Let σ𝒯(S)\sigma\in\mathcal{T}(S) so that μ\mu is a short marking on σ\sigma. (Equivalently, let σ=Υ(μ)\sigma=\Upsilon(\mu).) Since the action of HH on 𝒯(S)\mathcal{T}(S) permutes the lengths of curves, the length of each pair (αi,βi)(\alpha_{i},\beta_{i}) with respect to σ\sigma is the same as the length of the pair (hαi,hβi)(h\alpha_{i},h\beta_{i}) with respect to hσh\sigma. Therefore as μ\mu was a short marking for σ\sigma, this shows that hμh\mu is a short marking for hσ=hΥ(μ)h\sigma=h\Upsilon(\mu). However, by definition of Υ\Upsilon, hμh\mu is also a short marking for Υ(hμ)\Upsilon(h\mu). As Υ\Upsilon was a coarsely well defined function, this shows that d𝒯(S)(hΥ(μ),Υ(hμ))Kd_{\mathcal{T}(S)}(h\Upsilon(\mu),\Upsilon(h\mu))\leq K for some K0K\geq 0, as desired. ∎

We now prove Theorem 5.2, using the above lemma and several tools from [Cor17] to show that points in conical neighborhoods in (S)\mathcal{M}(S) end up in conical neighborhoods of 𝒯(S)\mathcal{T}(S).

Proof.

Fix μ0(S)\mu_{0}\in\mathcal{M}(S). Let x(S)μ0x\in\partial\mathcal{M}(S)_{\mu_{0}} be a conical limit point of Hμ0H\mu_{0}. Put σ0=Υ(μ0)\sigma_{0}=\Upsilon(\mu_{0}). We shall show that f(x)f_{\infty}(x) is a conical limit point of Hσ0H\sigma_{0} by verifying the condition in Proposition 4.4. Let T0T\geq 0 be arbitrary, and let λ:[0,)𝒯(S)\lambda:[0,\infty)\rightarrow\mathcal{T}(S) be an arbitrary Morse geodesic ray with λ(0)=σ0\lambda(0)=\sigma_{0} and λ()=f(x)\lambda(\infty)=f_{\infty}(x).

Let α:(S)\alpha:\mathbb{N}\rightarrow\mathcal{M}(S) be an NN-Morse geodesic with α(0)=μ0\alpha(0)=\mu_{0} and α()=x\alpha(\infty)=x. By [Cor17, Lemma 4.9], Υ(α)\Upsilon(\alpha) is an NN^{\prime}-Morse (A,B)(A,B)-quasi-geodesic, for some AA, BB, and NN^{\prime} depending only on NN. Put β=Υ(α)\beta=\Upsilon(\alpha). Notice that β(0)=σ0\beta(0)=\sigma_{0} and, by the construction of ff_{\infty}, we have β()=f(x)\beta(\infty)=f_{\infty}(x). (For details on the construction of ff_{\infty}, we refer to [Cor17], specifically Proposition 4.11, Theorem 4.12, and Proposition 4.14.)

Now let γn=[σ0,β(n)]\gamma_{n}=[\sigma_{0},\beta(n)]. Then each γn\gamma_{n} is N′′N^{\prime\prime}-Morse for N′′N^{\prime\prime} depending on NN, and by Arzelá-Ascoli and [Cor17, Lemma 2.10], a subsequence of the γn\gamma_{n} converges to a geodesic ray β\beta which is N′′N^{\prime\prime}-Morse, and by [Cor17, Lemma 4.9], β\beta is bounded Hausdorff distance from γ\gamma, where the bound only depends on NN. Say that dHaus(β,γ)K1d_{Haus}(\beta,\gamma)\leq K_{1} for K10K_{1}\geq 0. By [Cor17, Corollary 2.6], dHaus(γ,λ)K2d_{Haus}(\gamma,\lambda)\leq K_{2} where K20K_{2}\geq 0 depends only on NN. Choose S0S\geq 0 so that, for all sSs\geq S, d𝒯(S)(β(s),λ[T,))K1+K2d_{\mathcal{T}(S)}(\beta(s),\lambda_{[T,\infty)})\leq K_{1}+K_{2}.

By Proposition 4.4, there exists L0L\geq 0 where, for all r0r\geq 0, d(S)(hμ0,α|[r,))Ld_{\mathcal{M}(S)}(h\mu_{0},\alpha|_{[r,\infty)})\leq L for some hHh\in H. Since β=Υ(α)\beta=\Upsilon(\alpha) and Υ\Upsilon is coarse Lipschitz, there exits K30K_{3}\geq 0 and hHh\in H so that d𝒯(S)(Υ(hμ0),β|[S,))K3d_{\mathcal{T}(S)}(\Upsilon(h\mu_{0}),\beta|_{[S,\infty)})\leq K_{3}. Let s0[S,)s_{0}\in[S,\infty) so that d𝒯(S)(Υ(hμ0),β(s0))K3d_{\mathcal{T}(S)}(\Upsilon(h\mu_{0}),\beta(s_{0}))\leq K_{3}. By Lemma 5.4, there exists K40K_{4}\geq 0 such that d𝒯(S)(Υ(hμ0),hΥ(μ0))K4d_{\mathcal{T}(S)}(\Upsilon(h\mu_{0}),h\Upsilon(\mu_{0}))\leq K_{4}.

By the triangle inequality, we have

d𝒯(S)(hσ0,λ|[T,))\displaystyle d_{\mathcal{T}(S)}(h\sigma_{0},\lambda|_{[T,\infty)}) d(hΥ(μ0),Υ(hμ0))+d(Υ(hμ0),β(s0))+d(β(s0),λ|[T,))\displaystyle\leq d(h\Upsilon(\mu_{0}),\Upsilon(h\mu_{0}))+d(\Upsilon(h\mu_{0}),\beta(s_{0}))+d(\beta(s_{0}),\lambda|_{[T,\infty)})
K4+K3+K2+K1\displaystyle\leq K_{4}+K_{3}+K_{2}+K_{1}

By Proposition 4.4, λ()=f(x)\lambda(\infty)=f_{\infty}(x) is a conical limit point of Hσ0H\sigma_{0}. ∎

References

  • [ABD21] Carolyn Abbott, Jason Behrstock and Matthew Durham “Largest acylindrical actions and Stability in hierarchically hyperbolic groups” In Transactions of the American Mathematical Society, Series B 8.3 American Mathematical Society (AMS), 2021, pp. 66–104 DOI: 10.1090/btran/50
  • [ADT17] Tarik Aougab, M. Durham and S.. Taylor “Pulling back stability with applications to Out(Fn)Out(F_{n}) and relatively hyperbolic groups” In J. Lond. Math. Soc. 96, 2017, pp. 565–583
  • [Ant+19] Yago Antolin, Mahan MJ, Alessandro Sisto and Samuel Taylor “Intersection properties of stable subgroups and bounded cohomology” In Indiana University Mathematics Journal 68.1 Indiana University Mathematics Journal, 2019, pp. 179–199 DOI: 10.1512/iumj.2019.68.7592
  • [Bal+23] Sahana Balasubramanya et al. “(Non-)Recognizing Spaces for Stable Subgroups” arXiv, 2023 DOI: 10.48550/ARXIV.2311.15187
  • [Beh06] Jason A Behrstock “Asymptotic geometry of the mapping class group and Teichmüller space” In Geometry & Topology 10.3 Mathematical Sciences Publishers, 2006, pp. 1523–1578 DOI: 10.2140/gt.2006.10.1523
  • [BH09] Martin Bridson and André Haefliger “Metric Spaces of Non-Positive Curvature” In Fundamental Principles of Mathematical Sciences 319 Springer Science & Business Media, 2009 DOI: 10.1007/978-3-662-12494-9
  • [CD17] Matthew Cordes and Matthew Gentry Durham “Boundary Convex Cocompactness and Stability of Subgroups of Finitely Generated Groups” In International Mathematics Research Notices 2019.6 Oxford University Press (OUP), 2017, pp. 1699–1724 DOI: 10.1093/imrn/rnx166
  • [CH17] Matthew Cordes and David Hume “Stability and the Morse boundary” In Journal of the London Mathematical Society 95.3 Wiley, 2017, pp. 963–988 DOI: 10.1112/jlms.12042
  • [CM17] “Office hours with a geometric group theorist” Princeton, NJ: Princeton University Press, 2017
  • [Cor+22] Matthew Cordes, Jacob Russell, Davide Spriano and Abdul Zalloum “Regularity of Morse geodesics and growth of stable subgroups” In Journal of Topology 15.3, 2022, pp. 1217–1247 DOI: https://doi.org/10.1112/topo.12245
  • [Cor17] Matthew Cordes “Morse boundaries of proper geodesic metric spaces” In Groups, Geometry, and Dynamics 11.4 European Mathematical Society - EMS - Publishing House GmbH, 2017, pp. 1281–1306 DOI: 10.4171/ggd/429
  • [DT15] Matthew Durham and Samuel J Taylor “Convex cocompactness and stability in mapping class groups” In Algebraic & Geometric Topology 15.5 MSP, 2015, pp. 2837–2857 DOI: 10.2140/agt.2015.15.2839
  • [DT17] Spencer Dowdall and Samuel Taylor “Hyperbolic extensions of free groups” In Geometry & Topology 22.1 Mathematical Sciences Publishers, 2017, pp. 517–570 DOI: 10.2140/gt.2018.22.517
  • [FLP12] Albert Fathi, François Laudenbach and Valentin Poénaru “Thurston’s Work on Surfaces” Translated from the 1979 French original by Djun M. Kim and Dan Margalit 48, Mathematical Notes Princeton University Press, 2012, pp. xvi+254
  • [FM01] Benson Farb and Lee Mosher “Convex cocompact subgroups of mapping class groups” In Geometry & Topology 6, 2001, pp. 91–152
  • [FM11] Benson Farb and Dan Margalit “A primer on mapping class groups (PMS-49)”, Princeton Mathematical Series Princeton, NJ: Princeton University Press, 2011
  • [Ham05] Ursula Hamenstaedt “Word hyperbolic extensions of surface groups” arXiv, 2005 DOI: 10.48550/ARXIV.MATH/0505244
  • [Ham15] Ursula Hamenstädt “Annoucement” Conference on Mod(S) and Out(F), Austin, TX, 2015
  • [HH18] Ursula Hamenstädt and Sebastian Hensel “Stability in Outer space” In Groups, Geometry, and Dynamics 12.1 European Mathematical Society - EMS - Publishing House GmbH, 2018, pp. 359–398 DOI: 10.4171/ggd/445
  • [KB02] Ilya Kapovich and Nadia Benakli “Boundaries of Hyperbolic Groups” In Combinatorial and Geometric Group Theory 296, Contemporary Mathematics Providence, RI: American Mathematical Society, 2002, pp. 39–94
  • [Kim19] Heejoung Kim “Stable subgroups and Morse subgroups in mapping class groups” In International Journal of Algebra and Computation 29.05 World Scientific Pub Co Pte Lt, 2019, pp. 893–903 DOI: 10.1142/s0218196719500346
  • [KL08] Autumn E. Kent and Christopher J. Leininger “Shadows of mapping class groups: Capturing convex Cocompactness” In Geometric and Functional Analysis 18.4, 2008, pp. 1270–1325 DOI: 10.1007/s00039-008-0680-9
  • [Liu21] Qing Liu “Dynamics on the Morse boundary” In Geometriae Dedicata 214.1, 2021, pp. 1–20 DOI: 10.1007/s10711-021-00600-7
  • [Mar74] Albert Marden “The Geometry of Finitely Generated Kleinian Groups” In Annals of Mathematics 99.3 Annals of Mathematics, 1974, pp. 383–462
  • [MM00] H.A. Masur and Y.N. Minsky “Geometry of the complex of curves II: Hierarchical structure” In Geometric and Functional Analysis 10.4 Springer ScienceBusiness Media LLC, 2000, pp. 902–974 DOI: 10.1007/pl00001643
  • [RST21] Jacob Russell, Davide Spriano and Hung Cong Tran “The local-to-global property for Morse quasi-geodesics” In Mathematische Zeitschrift 300.2 Springer ScienceBusiness Media LLC, 2021, pp. 1557–1602 DOI: 10.1007/s00209-021-02811-w
  • [RST23] Jacob Russell, Davide Spriano and Hung Cong Tran “Convexity in hierarchically hyperbolic spaces” In Algebraic & Geometric Topology 23.3 Mathematical Sciences Publishers, 2023, pp. 1167–1248 DOI: 10.2140/agt.2023.23.1167
  • [Sul85] Dennis Sullivan “Quasiconformal Homeomorphisms and Dynamics I. Solution of the Fatou-Julia Problem on Wandering Domains” In Annals of Mathematics 122.2 Annals of Mathematics, 1985, pp. 401–418
  • [Swe01] Eric L. Swenson “Quasi-convex groups of isometries of negatively curved spaces” In Topology and its Applications 110, 2001, pp. 119–129
  • [Tra19] Hung Tran “On strongly quasiconvex subgroups” In Geometry & Topology 23.3 Mathematical Sciences Publishers, 2019, pp. 1173–1235 DOI: 10.2140/gt.2019.23.1173
  • [Zbi24] Stefanie Zbinden “Hyperbolic spaces that detect all strongly-contracting directions” arXiv, 2024 DOI: 10.48550/ARXIV.2404.12162