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

Connectedness properties and splittings of groups with isolated flats

G. Christopher Hruska Department of Mathematical Sciences
University of Wisconsin–Milwaukee
PO Box 413
Milwaukee, WI 53211
USA
chruska@uwm.edu
 and  Kim Ruane Department of Mathematics
Tufts University
Medford, MA 02155
USA
kim.ruane@tufts.edu
(Date: August 12, 2025)
Abstract.

In this paper we study CAT(0)\operatorname{CAT}(0) groups and their splittings as graphs of groups. For one-ended CAT(0)\operatorname{CAT}(0) groups with isolated flats we prove a theorem characterizing exactly when the visual boundary is locally connected. This characterization depends on whether the group has a certain type of splitting over a virtually abelian subgroup. In the locally connected case, we describe the boundary as a tree of metric spaces in the sense of Świątkowski.

A significant tool used in the proofs of the above results is a general convex splitting theorem for arbitrary CAT(0)\operatorname{CAT}(0) groups. If a CAT(0)\operatorname{CAT}(0) group splits as a graph of groups with convex edge groups, then the vertex groups are also CAT(0)\operatorname{CAT}(0) groups.

Key words and phrases:
Nonpositive curvature, isolated flats, locally connected, tree of metric compacta
2010 Mathematics Subject Classification:
20F67, 20E08

1. Introduction

A major theme of geometric group theory over the last few decades has been the study of a group using various boundaries at infinity attached to spaces on which the group acts. There has been a fruitful connection between topological properties of these boundaries and algebraic properties of the group. Our main theorem provides another example of such a connection for the groups that act geometrically on CAT(0)\operatorname{CAT}(0) spaces with isolated flats.

This article is concerned with the following problem: describe the topological spaces that can arise as the boundary of a CAT(0)\operatorname{CAT}(0) space with isolated flats that admits a geometric group action. Kapovich–Kleiner posed an analogous question for word hyperbolic groups in [KK00]. In order to describe boundaries, one first needs to know which topological properties boundaries can have.

If one can prove that the boundary is connected and locally connected, then it is a Peano continuum. This is a class of compact metric spaces with a rich structure theory that is substantially more powerful than general continuum theory. Much of this paper focuses on determining which groups can have such a boundary.

In the hyperbolic setting, all connected boundaries are locally connected, by a deep, nontrivial theorem (see [BM91, Lev98, Bow99, Swa96]) with far reaching consequences such as [Bow98, KK00]. Groups acting on CAT(0)\operatorname{CAT}(0) spaces with isolated flats [KL95, HK05] are, in a sense, the simplest generalization of hyperbolicity in the CAT(0)\operatorname{CAT}(0) setting. By Hruska–Kleiner, a CAT(0)\operatorname{CAT}(0) group with isolated flats is relatively hyperbolic with respect to the collection \mathbb{P} of maximal virtually abelian subgroups of higher rank [HK05]. A virtually abelian group has higher rank if its rank over \mathbb{Q} is at least two.

By a theorem of Mihalik–Ruane [MR99, MR01], many one-ended CAT(0)\operatorname{CAT}(0) groups with isolated flats have connected but non–locally connected boundary. These examples occur when the group exhibits a particular kind of splitting as a graph of groups. A simple example of the non–locally connected type is the fundamental group of the space obtained from a closed genus two surface and a torus by gluing together an essential simple closed curve from each surface (see Example 2.1).

In the general CAT(0)\operatorname{CAT}(0) setting, determining local connectedness seems to be quite a delicate issue. One-ended hyperbolic groups and fundamental groups of closed Hadamard manifolds all have locally connected boundary, while all CAT(0)\operatorname{CAT}(0) boundaries of one-ended nonabelian right-angled Artin groups are non–locally connected. Very few natural classes of CAT(0)\operatorname{CAT}(0) groups are known to admit boundaries of both types, and among those none had been completely characterized in terms of local connectivity. Indeed this question is not even completely understood for right-angled Coxeter groups, a family that includes many groups with both types of boundary.

Let GG act geometrically on a one-ended CAT(0)\operatorname{CAT}(0) space XX with isolated flats. Our main theorem determines exactly when GG has locally connected boundary.

Theorem 1.1 (Locally connected).

Let GG be a one-ended CAT(0)\operatorname{CAT}(0) group with isolated flats. The boundary G\partial G is non–locally connected if and only if GG contains a pair of virtually abelian subgroups B<AB<A with the following properties:

  1. (1)

    GG splits (nontrivially) over BB,

  2. (2)

    AA has higher rank, and

  3. (3)

    The \mathbb{Q}–rank of BB is strictly less than the \mathbb{Q}–rank of AA.

The reverse implication of Theorem 1.1 follows immediately from the Mihalik–Ruane splitting theorem. This article concerns the forward implication of Theorem 1.1, which was not known previously.

A compactum is semistable if it is shape-theoretically equivalent to a locally connected continuum [GS]. A well-known conjecture of Geoghegan–Mihalik would imply that every boundary of every one-ended CAT(0)\operatorname{CAT}(0) group is semistable [Mih83, GS]. This conjecture has been proven in the CAT(0)\operatorname{CAT}(0) with isolated flats setting by combining work of Mihalik–Swenson and the authors [MS, HR]. In other words, all torsion-free one-ended CAT(0)\operatorname{CAT}(0) groups with isolated flats have semistable boundary. Thus the conclusion of Theorem 1.1 gives finer information than one could derive from semistability alone. When considered up to shape equivalence, all isolated flats boundaries are equivalent to Peano continua, but when considered up to homeomorphism it turns out that many are not Peano continua. In a sense, semistability is too weak to see the precise homeomorphism type of a given boundary.

The conclusion of Theorem 1.1 has already been used as an essential ingredient in constructing examples of non-hyperbolic CAT(0)\operatorname{CAT}(0) groups with boundary homeomorphic to the Menger curve [Hau18, HHS, DHW], which were not previously known to exist. Although groups with Menger boundary in the hyperbolic setting are well-known [Ben92, Cha95, Bou97, KK00, DGP11], substantially different techniques are needed in the non-hyperbolic setting.

In the locally connected case we also obtain a detailed description of the boundary of GG as the tree of metric compacta in the sense of Świątkowski [Świ]. This structure explicitly determines the boundary as a tree of spaces built in the same fractal manner as the trees of manifolds of Jakobsche and Ancel–Siebenmann [Jak80, AS85, Jak91], which generalize the classical Pontryagin mod-22 surface [Pon30].

The building blocks in this construction are boundaries of vertex groups in a natural graph of groups splitting of GG. These vertex groups are atomic in the following sense. Suppose GG is a one-ended CAT(0)\operatorname{CAT}(0) group with isolated flats. Let 𝒜\mathcal{A} be the family of subgroups of GG that are contained in higher rank virtually abelian subgroups. We say that GG is atomic if GG does not split over any subgroup in 𝒜\mathcal{A}. The methods used in the proof of Theorem 1.1 also give the following theorem explicitly describing the topology of the boundary.

Theorem 1.2 (Tree of spaces).

Let GG be a one-ended CAT(0)\operatorname{CAT}(0) group GG with isolated flats. If G\partial G is locally connected, then GG is the fundamental group of a graph of groups such that all vertex groups are atomic CAT(0)\operatorname{CAT}(0) groups with isolated flats, and all edge groups are higher rank virtually abelian. Furthermore G\partial G is homeomorphic to a tree of metric compacta where the compacta are the boundaries of the vertex groups of the splitting.

As suggested by Świątkowski, the structure of a tree of metric compacta may prove useful in classifying which topological spaces arise as visual boundaries of CAT(0)\operatorname{CAT}(0) groups.

In a general setting, attempts to understand a group by splitting into indecomposable pieces often require a study of hierarchies of splittings (as in the settings of one-relator groups, 33–manifold groups, cubulated groups, etc.). For example in [HR], the authors prove semistability in the isolated flats setting using a nontrivial theorem of Louder–Touikan [LT17] on the termination of slender hierarchies of relatively hyperbolic groups. The study of hierarchies is technically much more elaborate than the study of a single splitting.

In contrast with [HR], in the locally connected case of Theorem 1.2 it is notable that we reach atomic vertex groups after only a single splitting. We show that the hierarchy terminates after one splitting in Theorem 6.4. In Corollary 7.8 we show that the resulting vertex groups are CAT(0)\operatorname{CAT}(0) with isolated flats.

1.1. Methods of proof

As mentioned above, a group GG acting properly, cocompactly on a CAT(0)\operatorname{CAT}(0) space with isolated flats is hyperbolic relative to the family \mathbb{P} of all maximal virtually abelian subgroups of higher rank; ie, rank at least two. Bowditch introduced a boundary associated to the pair (G,)(G,\mathbb{P}), now known as the Bowditch boundary (G,)\partial(G,\mathbb{P}).

In the isolated flats setting, the Bowditch boundary is always locally connected [Bow01]. Theorem 1.1 does not follow directly from Bowditch’s theorem for the following reason. Hung Cong Tran has shown that the Bowditch boundary is a quotient space of the CAT(0)\operatorname{CAT}(0) boundary [Tra13]. It is well-known that the continuous image of a locally connected space is also locally connected. However to prove Theorem 1.1 we would need the converse, which is simply not true in general.

The proof in [Bow01] that Bowditch’s boundary is locally connected involves three main steps. The first step gives the existence of a maximal peripheral splitting of GG (see Section 6). The vertex groups of this peripheral splitting are relatively hyperbolic groups that admit only a trivial peripheral splitting. The second step involves showing that groups for which this splitting is trivial have locally connected Bowditch boundary. In the third step Bowditch shows that the boundary is composed of many copies of boundaries of the vertex groups glued along points in the pattern of the Bass–Serre tree.

Our proof of the forward direction of Theorem 1.1 is via a contrapositive argument. We show that if GG does not have an “infinite index” splitting as in the statement of the theorem, then the CAT(0)\operatorname{CAT}(0) boundary of GG is locally connected. This proof follows a similar outline as in Bowditch’s proof. But the details of each step are substantially different from Bowditch’s methods.

The first step begins by examining Bowditch’s maximal peripheral splitting. We can use this splitting result since our groups are relatively hyperbolic. However we need the much stronger conclusions that the vertex groups of Bowditch’s splitting are atomic and again CAT(0)\operatorname{CAT}(0) with isolated flats. The algebraic assumption of no infinite index splitting gives atomic vertex groups (see Theorem 6.4).

In order to conclude that the vertex groups are CAT(0)\operatorname{CAT}(0) groups, we introduce the following “Convex Splitting Theorem.” We note that this theorem does not involve the notion of isolated flats, and therefore could prove useful in other CAT(0)\operatorname{CAT}(0) situations. Suppose GG acts geometrically on any CAT(0)\operatorname{CAT}(0) space XX. A subgroup HGH\leq G is convex if HH stabilizes a closed convex subspace YY of XX, and HH acts cocompactly on YY.

Theorem 1.3 (Convex Splitting Theorem).

Let GG act geometrically on any CAT(0)\operatorname{CAT}(0) space XX. Suppose GG splits as the fundamental group of a graph of groups 𝒢\mathcal{G} such that each edge group of 𝒢\mathcal{G} is convex. Then each vertex group is also convex. In particular, each vertex group is a CAT(0)\operatorname{CAT}(0) group itself.

To see that this theorem is not obvious, recall first that a similar result for hyperbolic groups and quasiconvex subgroups is well-known and very straightforward to prove (see, for example [Bow98, Prop. 1.2]). The basic strategy in the hyperbolic case is to map the Cayley graph of GG to the Bass–Serre tree TT and to use this map to cut the Cayley graph into pieces that one proves are quasiconvex.

However in the CAT(0)\operatorname{CAT}(0) setting, it is not obvious how to map XX to the tree. To do this, we rely on a powerful theorem of Ontaneda that produces an equivariant simplicial nerve for a CAT(0)\operatorname{CAT}(0) space XX with a cocompact group action [Ont05]. We map XX to the tree TT using this nerve, and use this map to cut XX into pieces (typically not convex). To complete the proof of Theorem 1.3, much more care is needed than in the hyperbolic case since convexity is more delicate to establish than quasiconvexity.

The second step of the proof of Theorem 1.1 is to prove the following special case involving atomic CAT(0)\operatorname{CAT}(0) groups with isolated flats.

Theorem 1.4.

Let XX be a CAT(0)\operatorname{CAT}(0) space with isolated flats that admits a geometric group action by a group GG. Suppose GG is atomic. Then the CAT(0)\operatorname{CAT}(0) boundary X\partial X is locally connected.

The main technical difficulty in the proof of Theorem 1.4 is proving local connectedness at a point that lies in the limit set of a flat Euclidean subspace. The examination of such points requires the development of new techniques not used in the study of the Bowditch boundary (see Section 9).

The third and final step in the proof of Theorem 1.1 is to use the hypothesis of “no infinite index splittings” to conclude that the CAT(0)\operatorname{CAT}(0) boundary of GG is locally connected. Most of the work in this step is proving that GG has the structure of a tree of metric compacta where the compacta are boundaries of atomic CAT(0)\operatorname{CAT}(0) groups with isolated flats.

The precise topology on this compactification is given as an inverse limit. Using this description of the boundary of GG as an inverse limit, we deduce local connectivity by applying a theorem of Capel on inverse limits of locally connected spaces [Cap54].

1.2. Organization of the paper

Section 2 contains an informal discussion of several examples of CAT(0)\operatorname{CAT}(0) groups with isolated flats, some with locally connected boundary and some with non–locally connected boundary. Sections 3, 4, and 5 summarize necessary background on CAT(0)\operatorname{CAT}(0) spaces, relative hyperbolicity, and groups with isolated flats. In Section 6, we examine Bowditch’s work on peripheral splittings of relatively hyperbolic groups and deduce several consequences in the setting of isolated flats. Theorem 1.3 is proved in Section 7. In Section 8 we show that the visual boundary is locally connected at each point that is not in the boundary of a flat, which is the first part of the proof of Theorem 1.4. The remaining part of this theorem is proved in Section 9. Section 10 summarizes definitions, terminology, and key facts about tree systems of metric compacta. Finally in Section 11 we show that the boundary has the structure of a tree system as described in Theorem 1.2, and we complete the proof of Theorem 1.1.

1.3. Acknowledgements

During their work on this project, the authors benefited from many conversations about this work with Ric Ancel, Mladen Bestvina, Craig Guilbault, Matthew Haulmark, Mike Mihalik, Boris Okun, Eric Swenson, Hung Cong Tran, and Genevieve Walsh. We are grateful for the advice and feedback received during these conversations.

This work was partially supported by a grant from the Simons Foundation (#318815 to G. Christopher Hruska).

2. Examples with and without locally connected boundary

In this section we illustrate both directions of Theorem 1.1 with examples. The first example shows a group whose visual boundary is not locally connected, but whose Bowditch boundary is locally connected. The remaining examples illustrate various constructions of non-hyperbolic groups with locally connected visual boundaries, some illustrating the “indecomposable” case of Theorem 1.4 and others constructed using “locally finite” amalgams of indecomposable groups.

Example 2.1 (A non–locally connected boundary).

Consider the following amalgam of surface groups whose visual boundary is not locally connected. Let Σ\Sigma be a closed hyperbolic surface, and T2T^{2} be a 22–dimensional torus with a fixed Euclidean metric. Fix a simple closed geodesic loop γ\gamma in Σ\Sigma. Choose a closed geodesic γT\gamma^{\prime}\subset T such that γ\gamma and γ\gamma^{\prime} have equal lengths. Let XX be the result of gluing Σ\Sigma to T2T^{2} along γ=γ\gamma=\gamma^{\prime}. Then the universal cover X~\tilde{X} of XX is a CAT(0)\operatorname{CAT}(0) space with isolated flats. The group G=π1(X)G=\pi_{1}(X) is clearly an amalgam of the subgroups A=π1(T2)A=\pi_{1}(T^{2}) and C=π1(Σ)C=\pi_{1}(\Sigma) amalgamated over the subgroup B=γB=\langle\gamma\rangle. Since this splitting ABCA*_{B}C satisfies the conditions of Theorem 1.1, we see that the visual boundary X~\partial\tilde{X} is not locally connected.

In order to illustrate the difference between the non–locally connected CAT(0)\operatorname{CAT}(0) boundary of GG and the locally connected Bowditch boundary (G,)\partial(G,\mathbb{P}) in Example 2.1, we briefly sketch a significant non–locally connected subset of the boundary. We then describe its locally connected image in the Bowditch boundary.

Each flat FF in X~\tilde{X} is a lift of the torus T2T^{2}. Inside FF are infinitely many lifts of the geodesic γ\gamma, along each of which there is a copy of 2\mathbb{H}^{2} attached. The lifts of γ\gamma in the flat FF are all parallel and thus all share the same pair of endpoints a,ba,b in F\partial F. Let YY be the convex subcomplex of X~\tilde{X} consisting of the flat FF along with the countably many hyperbolic planes glued to FF along the lifts of γ\gamma. Then Y\partial Y is a suspension of a set KK that is countably infinite with two limit points. The suspension points are aa and bb. Note that this suspension itself is not locally connected. Furthermore, each of the points in F{a,b}\partial F-\{a,b\} turns out to be a point of non–local connectivity in X~\partial\tilde{X}.

To understand what the Bowditch boundary is in this example, note that there are two types of circles in the visual boundary X~\partial\tilde{X}—those that occur as the boundary of a hyperbolic plane and those that occur as the boundary of a flat. By [Tra13], the Bowditch boundary is obtained from the visual boundary by collapsing the circles arising as boundaries of flats. In the quotient, each such circle becomes a global cut point of the Bowditch boundary. Each of these cut points is incident to a countable family of circles whose union forms a Hawaiian earring.

Example 2.2 (Some locally connected boundaries).

In this example we show three different groups with locally connected boundary, formed by gluing hyperbolic 33–manifolds along cusps.

First consider the figure eight knot KS3K\subset S^{3}, and let NN be a closed regular neighborhood of KK. Let M3M^{3} be the compact knot complement, S3S^{3} minus the interior of NN. It is well-known that G=π1(M3)G=\pi_{1}(M^{3}) is one-ended with isolated flats and does not split over any subgroup of the cusp group. Thus by Theorem 1.4 the visual boundary of the CAT(0)\operatorname{CAT}(0) space M3~\tilde{M^{3}} is locally connected. In this simple example, it was known previously that the boundary was locally connected, since it is homeomorphic to a Sierpinski carpet by [Rua05]. (In this case the Bowditch boundary is a 22–sphere formed by collapsing each peripheral circle of the Sierpinski carpet to a point.)

If we double M3M^{3} along its boundary torus N\partial N, we get a closed 33–manifold consisting of two hyperbolic pieces glued along the torus T2=NT^{2}=\partial N. The fundamental group DD of the double does not split over any cyclic subgroup of π1(T2)\pi_{1}(T^{2}), and thus by Theorem 1.1 its boundary is also locally connected. Once again, we knew this already because the visual boundary of any closed nonpositively curved 33–manifold group is a 22–sphere. Applying the proof of Theorem 1.1 to this example recovers the classical decomposition of S2S^{2} as a tree of Sierpinski carpets glued in pairs along peripheral circles.

If we form a “triple” of M3M^{3} instead of a double, we get a less familiar example, whose visual boundary was not previously known to be locally connected. As with the double, the tripled space is formed by gluing three copies of M3M^{3} along the boundary torus T2T^{2}. As above its fundamental group does not split over any cyclic subgroup of π1(T2)\pi_{1}(T^{2}). Thus we establish that its visual boundary is locally connected. We also obtain a description of this boundary as a 22–dimensional compactum formed as the limit of a tree of Sierpinski carpets, this time glued in triples along peripheral circles.

Example 2.3.

We conclude this section with an example of a group with isolated flats having Serre’s Property FA—i.e., no splittings at all. Consider any Coxeter group WW on 55 generators sis_{i} of order two with defining Coxeter relations (sisj)mij=1(s_{i}s_{j})^{m_{ij}}=1 such that 3mij<3\leq m_{ij}<\infty for all iji\neq j. The Davis complex Σ\Sigma of each such WW is a piecewise Euclidean CAT(0)\operatorname{CAT}(0) 22–complex whose 22–cells are isometric to regular Euclidean 2mij2m_{ij}–gons with at least 66 sides. By an observation of Wise, such complexes have isolated flats (see [Hru04]). Since each mijm_{ij} is finite, WW has Property FA by [Ser77]. Therefore WW is one-ended with locally connected visual boundary by Theorem 1.4. Haulmark–Hruska–Sathaye show in [HHS] that each such WW has W\partial W homeomorphic to the Menger curve, using the local connectivity established above as a key step.

3. The visual boundary of a CAT(0)\operatorname{CAT}(0) space

We refer the reader to [Bal95, BH99] for introductions to the theory of CAT(0)\operatorname{CAT}(0) spaces. Throughout this section XX is assumed to be a proper CAT(0)\operatorname{CAT}(0) space, a condition that holds whenever XX admits a proper, cocompact, isometric group action.

The CAT(0)\operatorname{CAT}(0) geometry XX gives rise to the visual boundary X\partial X, which is a compact metrizable space. We first, define the boundary X\partial X as a set as follows:

Definition 3.1 (Visual boundary as a set).

Two geodesic rays c,c:[0,)Xc,c^{\prime}\colon[0,\infty)\to X are said to be asymptotic if there exists a constant KK such that d(c(t),c(t))Kd\bigl{(}c(t),c^{\prime}(t)\bigr{)}\leq K for all t>0t>0—this is an equivalence relation. The boundary of XX, denoted X\partial X, is then the set of equivalence classes of geodesic rays. The equivalence class of a ray cc is denoted by c()c(\infty).

Since XX is complete, then for each basepoint qXq\in X and each ξX\xi\in\partial X there is a unique geodesic cc such that c(0)=qc(0)=q and c()=ξc(\infty)=\xi. Thus we may identify X\partial X with the set qX\partial_{q}X of all rays emanating from qq. We use the notation X¯=XX\overline{X}=X\cup\partial X.

Definition 3.2 (The cone topology on X¯\overline{X}).

There is a natural topology on X¯\overline{X} called the cone topology, which is defined in terms of the following neighborhood basis. Let cc be a geodesic segment or ray, let q=c(0)q=c(0), and choose any r>0r>0 and D>0D>0. Also, let B¯(q,r)\overline{B}(q,r) denote the closed ball of radius rr centered at qq with πr:X¯B¯(q,r)\pi_{r}\colon\overline{X}\to\overline{B}(q,r) denoting projection. Define

U(c,r,D)={xX¯|d(x,q)>r,d(πr(x),c(r))<D}U(c,r,D)=\bigl{\{}\,{x\in\overline{X}}\bigm{|}{d(x,q)>r,\ d\bigl{(}\pi_{r}(x),c(r)\bigr{)}<D}\,\bigr{\}}

This consists of all points in X¯\overline{X} such that when projected back to B¯(q,r)\overline{B}(q,r), this projection is not more than DD away from the intersection of the sphere with cc. These sets along with the metric balls in XX form a basis for the cone topology on X¯\overline{X}. The induced topology on X\partial X is also called the cone topology on X\partial X, and the resulting topological space is the visual boundary of XX. We occasionally use the notation U(c,r,D)U(c,r,D) to refer to basic neighborhoods in the visual boundary. Since XX is proper, both X¯\overline{X} and the visual boundary are compact.

It is a well-known result that for any proper CAT(0)\operatorname{CAT}(0) space, both XX and X¯\overline{X} are ANR’s (Absolute Neighborhood Retracts). We refer the reader to [Ont05] for a proof that XX is an ANR, using a theorem of [Hu65]. See also [Gui14, §2.9] for a proof that X¯\overline{X} is an ANR using work of [Han51].

The main consequence of being an ANR that we will use is that X¯\overline{X} is locally connected. This consequence is much easier to prove directly, which we do below.

Proposition 3.3.

X¯\overline{X} is locally connected. Furthermore, each point ξX\xi\in\partial X has a connected neighborhood N¯\overline{N} such that N=N¯XN=\overline{N}\cap X is a connected set in XX and each point of Λ=N¯X\Lambda=\overline{N}\cap\partial X is a limit point of NN.

Let us pause for a moment to warn the reader that the notation N¯\overline{N} does not refer to a closed set of X¯\overline{X}, but rather refers to the fact that each point of N¯\overline{N} is a limit point of NN.

Proof.

Metric balls in XX are connected. Furthermore the basic open sets N¯=U(c,r,D)\overline{N}=U(c,r,D) are also connected. Indeed a point pXp\in X lies in N=U(c,r,D)XN=U(c,r,D)\cap X if and only if the geodesic segment c=[c(0),p]c^{\prime}=\bigl{[}c(0),p\bigr{]} intersects B(c(r),D)B\bigl{(}c(r),D\bigr{)}. For any such point p=c(t)Np=c^{\prime}(t)\in N, we can choose ss so that d(c(s),c(r))<Dd\bigl{(}c^{\prime}(s),c(r)\bigr{)}<D. It follows that the entire subpath of cc^{\prime} from c(s)c^{\prime}(s) to p=c(t)p=c^{\prime}(t) is contained in NN. Therefore every point of NN lies in a connected subset of NN that intersects the connected ball B(c(r),D)B\bigl{(}c(r),D\bigr{)}. In other words, NN is connected. The entire set N¯=U(c,r,D)\overline{N}=U(c,r,D) is contained in the closure of NN within the space X¯\overline{X}, so N¯\overline{N} is also connected. ∎

Throughout this paper we will often need to compare the sizes of different subsets of the visual boundary. To make such a comparison, it is useful to consider explicit metrics on the visual boundary. The following definition due to Osajda introduces a family of such metrics.

Definition 3.4 (Boundary as a metric space).

Fix a basepoint qXq\in X. For each D>0D>0 and r<r<\infty we define a metric dDd_{D} on qX\partial_{q}X as follows. Given distinct rays cc and cc^{\prime} based at qq, the distance function td(c(t),c(t))t\mapsto d\bigl{(}c(t),c^{\prime}(t)\bigr{)} monotonically increases from 0 to \infty. Thus there exists a unique r(0,)r\in(0,\infty) such that d(c(r),c(r))=Dd\bigl{(}c(r),c^{\prime}(r)\bigr{)}=D. We set dD(c,c)=1/rd_{D}(c,c^{\prime})=1/r. The function dDd_{D} is a metric compatible with the cone topology by Osajda–Świątkowski [OS15]. (See also [Mor16].)

4. Relatively hyperbolic groups and their boundaries

In this section we define the notions of relative hyperbolicity and the Bowditch boundary. The definitions we use are due to Yaman [Yam04] and are given in terms of dynamical properties of an action on a compact space, which turns out to be the Bowditch boundary.

A convergence group action is an action of a finitely generated group GG on a compact, metrizable space MM satisfying the following conditions, depending on the cardinality of MM:

  • If MM is the empty set, then GG is finite.

  • If MM has exactly one point, then GG is infinite.

  • If MM has exactly two points, then GG is virtually cyclic.

  • If MM has at least three points, then the action of GG on the space of distinct (unordered) triples of points of MM is properly discontinuous.

In the first three cases the convergence group action is elementary, and in the final case the action is nonelementary.

Suppose GG has a convergence group action on MM. An element gGg\in G is loxodromic if it has infinite order and fixes exactly two points of MM. A subgroup PGP\leq G is a parabolic subgroup if it is infinite and contains no loxodromic element. A parabolic subgroup PP has a unique fixed point in MM, called a parabolic point. The stabilizer of a parabolic point is always a maximal parabolic group. A parabolic point pp with stabilizer P:=StabG(p)P:=\operatorname{Stab}_{G}(p) is bounded parabolic if PP acts cocompactly on M{p}M-\{p\}. A point ξM\xi\in M is a conical limit point if there exists a sequence (gi)(g_{i}) in GG and distinct points ζ0,ζ1M\zeta_{0},\zeta_{1}\in M such that gi(ξ)ζ0g_{i}(\xi)\to\zeta_{0}, while for all ηM{ξ}\eta\in M-\{\xi\} we have gi(η)ζ1g_{i}(\eta)\to\zeta_{1}.

Definition 4.1 (Relatively hyperbolic).

A convergence group action of GG on MM is geometrically finite if every point of MM is either a conical limit point or a bounded parabolic point. If \mathbb{P} is a collection of subgroups of GG, then the pair (G,)(G,\mathbb{P}) is relatively hyperbolic if GG admits a geometrically finite convergence group action on a compact, metrizable space MM such that \mathbb{P} is equal to the collection of all maximal parabolic subgroups.

Definition 4.2 (Bowditch boundary).

By [Yam04] and [Bow12] the space MM is uniquely determined by (G,)(G,\mathbb{P}) in the following sense: Any two spaces MM and MM^{\prime} arising from the previous definition are GG–equivariantly homeomorphic. The compactum MM is the Bowditch boundary of (G,)(G,\mathbb{P}), and will be denoted by (G,)\partial(G,\mathbb{P}).

We remark that by work of Yaman, the Bowditch boundary can be obtained as the Gromov boundary of a certain δ\delta–hyperbolic space on which GG acts. Although we will not use this point of view in the present article, the construction of this δ\delta–hyperbolic space is the basis of Yaman’s proof that Definition 4.1 is equivalent to other definitions appearing in the literature (such as those in [Bow12]).

Suppose GG acts properly, cocompactly, and isometrically on a CAT(0)\operatorname{CAT}(0) space XX. Suppose also that GG has a family of subgroups \mathbb{P} such that (G,)(G,\mathbb{P}) is relatively hyperbolic. In this case, we have introduced two different types of boundary that one may associate with GG: the visual boundary X\partial X of the CAT(0)\operatorname{CAT}(0) space XX and the Bowditch boundary (G,)\partial(G,\mathbb{P}). These boundaries are closely related by a theorem of Hung Cong Tran, as was mentioned in the introduction. We give a precise statement here.

Theorem 4.3 ([Tra13]).

Let (G,)(G,\mathbb{P}) and XX be as above. The quotient space formed from X\partial X by collapsing the limit set of each PP\in\mathbb{P} to a point is GG–equivariantly homeomorphic to the Bowditch boundary (G,)\partial(G,\mathbb{P}).

5. Isolated flats

A kk–flat in a CAT(0)\operatorname{CAT}(0) space XX is an isometrically embedded copy of Euclidean space 𝔼k\mathbb{E}^{k} for some k2k\geq 2. In particular, note that a geodesic line is not considered to be a flat.

Definition 5.1.

Let XX be a CAT(0)\operatorname{CAT}(0) space, GG a group acting geometrically on XX, and \mathcal{F} a GG–invariant set of flats in XX. We say that XX has isolated flats with respect to \mathcal{F} if the following two conditons hold.

  1. (1)

    There is a constant DD such that every flat FXF\subset X lies in a DD–neighborhood of some FF^{\prime}\in\mathcal{F}.

  2. (2)

    For each positive r<r<\infty there is a constant ρ=ρ(r)<\rho=\rho(r)<\infty so that for any two distinct flats F,FF,F^{\prime}\in\mathcal{F} we have

    diam(𝒩r(F)𝒩r(F))<ρ.\operatorname{diam}\bigl{(}\mathcal{N}_{r}({F})\cap\mathcal{N}_{r}({F^{\prime}})\bigr{)}<\rho.

We say XX has isolated flats if it has isolated flats with respect to some GG–invariant set of flats.

Theorem 5.2 ([HK05]).

Suppose XX has isolated flats with respect to \mathcal{F}. For each FF\in\mathcal{F} the stabilizer StabG(F)\operatorname{Stab}_{G}(F) is virtually abelian and acts cocompactly on FF. The set of stabilizers of flats FF\in\mathcal{F} is precisely the set of maximal virtually abelian subgroups of GG of rank at least two. These stabilizers lie in only finitely many conjugacy classes.

Theorem 5.3 ([HK05]).

Let XX have isolated flats with respect to \mathcal{F}. Then the following properties hold.

  1. (1)

    GG is relatively hyperbolic with respect to the collection of all maximal virtually abelian subgroups of rank at least two.

  2. (2)

    The connected components of the Tits boundary TX\partial_{\text{T}}X are isolated points together with the boundary spheres TF\partial_{\text{T}}F for all FF\in\mathcal{F}.

The previous theorem also has the following converse.

Theorem 5.4 ([HK05]).

Let GG be a group acting geometrically on a CAT(0)\operatorname{CAT}(0) space XX. Suppose GG is relatively hyperbolic with respect to a family of virtually abelian subgroups. Then XX has isolated flats.

A group GG that admits an action on a CAT(0)\operatorname{CAT}(0) space with isolated flats has a “well-defined” visual boundary, often denoted by G\partial G, by the following theorem.

Theorem 5.5 ([HK05]).

Let GG act properly, cocompactly, and isometrically on two CAT(0)\operatorname{CAT}(0) spaces XX and YY. If XX has isolated flats, then so does YY, and there is a GG–equivariant homeomorphism XY\partial X\to\partial Y.

Finally, we point out a key geometric fact about CAT(0)\operatorname{CAT}(0) spaces with isolated flats that will be used several times throughout this paper.

Theorem 5.6.

Suppose XX is a CAT(0)\operatorname{CAT}(0) space with isolated flats with respect to \mathcal{F}. There exists a constant κ>0\kappa>0, such that the following holds: Given a point xx, a flat FF\in\mathcal{F}, with c:[a,b]Xc\colon[a,b]\to X the shortest path from xx to FF, we have cFc\cup F is κ\kappa–quasiconvex in XX. More precisely, if cc^{\prime} is any geodesic joining a point of cc to a point of FF, then cc^{\prime} intersects B(c(b),κ)B\bigl{(}c(b),\kappa\bigr{)}.

Proof.

If the claim were false, there would be sequences of flats FiF_{i}\in\mathcal{F} and points xiXx_{i}\in X and qi,yiFiq_{i},y_{i}\in F_{i} such that [xi,qi][x_{i},q_{i}] is a shortest path from xix_{i} to FiF_{i} and d(qi,[xi,yi])d\bigl{(}q_{i},[x_{i},y_{i}]\bigr{)} tends to infinity.

Pass to a subsequence and translate by the action of GG so that Fi=FF_{i}=F is constant. After passing to a further subsequence, the points qiq_{i}, xix_{i}, and yiy_{i} converge respectively to qFq\in F, ξxX\xi_{x}\in\partial X, and ξyF\xi_{y}\in\partial F. furthermore, ξxF\xi_{x}\notin\partial F since the ray from qq to ξx\xi_{x} meets FF orthogonally. Since d(qi,[xi,yi])d\bigl{(}q_{i},[x_{i},y_{i}]\bigr{)} tends to infinity, it follows from [HK09, Corollary 7] that dT(ξx,ξy)πd_{T}(\xi_{x},\xi_{y})\leq\pi, contradicting Theorem 5.3(2). ∎

6. Peripheral splittings

In this section, we give the definition of a peripheral splitting of a relatively hyperbolic group and examine some of their basic properties. We introduce the notion of a locally finite peripheral splitting, which plays a key role in the proof of Theorem 1.1. The goal of this section is to prove Theorem 6.4, which roughly states that in the locally finite case many features of a relatively hyperbolic group are inherited by the vertex groups of its maximal peripheral splitting.

Definition 6.1 (Peripheral splittings).

Suppose (G,)(G,\mathbb{P}) is a relatively hyperbolic group. A peripheral splitting of (G,)(G,\mathbb{P}) is a splitting of GG as a finite bipartite graph of groups 𝒢\mathcal{G}, whose vertices have two types that we call peripheral vertices and component vertices. We require that the collection of subgroups of GG conjugate to the peripheral vertex groups is identical to the collection \mathbb{P} of all peripheral subgroups. We also require that 𝒢\mathcal{G} does not contain a component vertex of degree one that is contained in the adjacent peripheral group.

Such a splitting is called trivial if one of the vertex groups is equal to GG and nontrivial otherwise. One peripheral splitting \mathcal{H} is a refinement of another 𝒢\mathcal{G} if 𝒢\mathcal{G} can be obtained from \mathcal{H} by a finite sequence of foldings of edges that preserve the colors of vertices. A refinement is trivial if it involves no folds, i.e., the refinement is an isomorphism of graphs. A peripheral splitting of (G,)(G,\mathbb{P}) is maximal if it admits only trivial refinements.

Let 𝒜\mathcal{A} be the collection of all subgroups of the peripheral subgroups of GG. If (G,)(G,\mathbb{P}) admits only a trivial peripheral splitting, then GG does not split over 𝒜\mathcal{A} relative to \mathbb{P} in the following sense: let GG act without inversions on a simplicial tree TT such that each peripheral subgroup has a fixed point in TT. Then the action of GG on TT has a global fixed point.

The notion of an atomic group, mentioned in the introduction, is a much stronger condition that does not include the phrase “relative to \mathbb{P}.”

Definition 6.2 (Atomic).

We say that (G,)(G,\mathbb{P}) is atomic if GG is one-ended, each PP\in\mathbb{P} is one-ended, and GG cannot be expressed as an HNN extension or a nontrivial amalgam over any subgroup in 𝒜\mathcal{A}.

Theorem 6.3 (Bowditch).

Suppose (G,)(G,\mathbb{P}) is relatively hyperbolic, GG is one-ended and each PP\in\mathbb{P} is finitely presented, does not contain an infinite torsion group, and is either one-ended or two-ended.

  1. (1)

    [Bow12, Prop. 10.1] (G,)\partial(G,\mathbb{P}) is connected.

  2. (2)

    [Bow01, Thm. 1.5] (G,)\partial(G,\mathbb{P}) is locally connected.

  3. (3)

    [Bow01, Thm. 1.4] (G,)(G,\mathbb{P}) has a unique maximal peripheral splitting, which could be trivial.

  4. (4)

    [Bow01, §9] If the maximal peripheral splitting is trivial, then (G,)\partial(G,\mathbb{P}) does not contain a global cut point.

If (G,)(G,\mathbb{P}) is atomic then the last conclusion of this theorem holds. In general the component vertex groups of the maximal peripheral splitting inherit a natural relatively hyperbolic structure, but they do not need to be atomic. See [HR] for examples of relatively hyperbolic groups whose component groups are not atomic, and which require a nontrivial hierarchy of splittings over parabolic subgroups in order to reach atomic groups.

In order to ensure that the maximal peripheral splitting has atomic component vertex groups, we need to focus on the special case of locally finite splittings, which are better behaved than the general case. A peripheral splitting 𝒢\mathcal{G} is locally finite if for each peripheral vertex group PP the adjacent edge groups include as finite index subgroups of PP. Equivalently, in the Bass–Serre tree TT for 𝒢\mathcal{G}, the vertex vv stabilized by PP has finite valence. The following theorem summarizes key properties of the component vertex groups in the unique maximal peripheral splitting in the special case that this splitting is locally finite. For simplicity we have stated this result only in the isolated flats setting, which is the only case needed in this paper.

Theorem 6.4.

Let (G,)(G,\mathbb{P}) be relatively hyperbolic such that GG is one-ended and each PP\in\mathbb{P} is virtually abelian of \mathbb{Q}–rank at least two. Suppose the maximal peripheral splitting 𝒢\mathcal{G} of (G,)(G,\mathbb{P}) is locally finite.

For each component vertex group HH of 𝒢\mathcal{G}, let 𝕆\mathbb{O} be the collection of infinite groups of the form HPH\cap P for all PP\in\mathbb{P}. Then

  1. (1)

    (H,𝕆)(H,\mathbb{O}) is relatively hyperbolic.

  2. (2)

    Each O𝕆O\in\mathbb{O} is virtually abelian of \mathbb{Q}–rank at least two.

  3. (3)

    (H,𝕆)(H,\mathbb{O}) is atomic.

Proof.

Let TT be the Bass–Serre tree of the splitting 𝒢\mathcal{G}. Then HH is stabilizer of a vertex vv in TT. Since the graph of 𝒢\mathcal{G} is bipartite, the other end of each such edge ee is a peripheral vertex ww. It follows that the family 𝕆\mathbb{O} is the collection of stabilizers of edges ee adjacent to vv. By the locally finite hypothesis, each O𝕆O\in\mathbb{O} is virtually abelian of \mathbb{Q}–rank at least two. By [Bow01], the pair (H,𝕆)(H,\mathbb{O}) is relatively hyperbolic and (H,𝕆)\partial(H,\mathbb{O}) is connected.

Since each O𝕆O\in\mathbb{O} is one-ended and the Bowditch boundary is connected, it follows that HH is a one-ended group by Proposition 10.1 of [Bow12].

By Lemma 4.6 of [Bow01], the pair (H,𝕆)(H,\mathbb{O}) admits only a trivial peripheral splitting. In general that lemma does not imply that HH has no splittings over subgroups of members of 𝕆\mathbb{O}. However since the members of 𝕆\mathbb{O} are one-ended, we may apply [Bow01, Proposition 5.2] to conclude that (H,𝕆)(H,\mathbb{O}) is atomic. ∎

7. Convex splittings of CAT(0)\operatorname{CAT}(0) groups

In this section we prove Theorem 1.3, which is a convex splitting theorem for CAT(0)\operatorname{CAT}(0) groups. This splitting theorem does not use the notion of isolated flats, and could potentially be a useful tool in many other CAT(0)\operatorname{CAT}(0) settings.

We apply this splitting theorem to peripheral splittings of CAT(0)\operatorname{CAT}(0) groups with isolated flats, in order to prove Theorem 7.7, which states that the vertex groups of such a splitting are also CAT(0)\operatorname{CAT}(0) groups with isolated flats. Although this article is mainly focused on locally finite splittings, Theorem 7.7 has no such restriction and applies to all CAT(0)\operatorname{CAT}(0) groups with isolated flats. Ben-Zvi uses Theorem 7.7 to show that all one-ended CAT(0)\operatorname{CAT}(0) groups with isolated flats have globally path connected boundary [BZ].

The next two results are key tools used in the proof of Theorem 1.3.

Lemma 7.1.

Suppose GG acts geometrically on a CAT(0)\operatorname{CAT}(0) space XX, and GG also acts on a simplicial tree TT. Then there is a GG–equivariant map π:XT\pi\colon X\to T that is continuous with respect to the CW–topology on TT.

Proof.

The proof depends on the following result of Ontaneda [Ont05]: Let XX be a CAT(0)\operatorname{CAT}(0) space on which GG acts properly, cocompactly, and isometrically. Then there exists a locally finite, finite dimensional simplicial complex KK on which GG acts properly, cocompactly, and simplicially. Furthermore there is a GG–equivariant continuous map XKX\to K.

To complete the proof, we need a GG–equivariant continuous map KTK\to T. After replacing TT with its barycentric subdivision, we may assume that GG acts on TT without inversions. Choose a representative σ\sigma for each GG–orbit of 0–simplices. Since GG acts properly on KK, the GG–stabilizer of σ\sigma is a finite group KσK_{\sigma}. Let vσv_{\sigma} be a vertex of TT fixed by KσK_{\sigma}. We define the map K(0)TK^{(0)}\to T by mapping σvσ\sigma\mapsto v_{\sigma} and extending equivariantly. Since TT is contractible, we may extend this map to the higher skeleta of KK in an equivariant fashion. ∎

The following folk result has been used implicitly in many places throughout the literature (see for example [HK05]). We have decided to include the (short) proof for the benefit of the reader. Dani Wise has described the proof as a “pigeonhole principle” for cocompact group actions.

Proposition 7.2 (Pigeonhole).

Suppose a group GG acts cocompactly and isometrically on a metric space XX. Let 𝒜\mathcal{A} be a family of closed subspaces of XX. Suppose 𝒜\mathcal{A} is GG–equivariant and locally finite, in the sense that each compact set of XX intersects only finitely many members of 𝒜\mathcal{A}. Then the stabilizer of each A𝒜A\in\mathcal{A} acts cocompactly on AA. Furthermore the members of 𝒜\mathcal{A} lie in finitely many GG–orbits.

Proof.

Let KK be a compact set whose GG–translates cover XX. Since KK intersects only finitely members of 𝒜\mathcal{A}, the sets of 𝒜\mathcal{A} lie in finitely many orbits. Thus we only need to establish that for each A𝒜A\in\mathcal{A} the group H=StabG(A)H=\operatorname{Stab}_{G}(A) acts cocompactly on AA. Let {gi}\{g_{i}\} be a set of group elements such that the translates gi(K)g_{i}(K) cover AA and each gi(K)g_{i}(K) intersects AA. If the sets gi1(A)g_{i}^{-1}(A) and gj1(A)g_{j}^{-1}(A) coincide, then gjgi1g_{j}g_{i}^{-1} lies in HH. It follows that the gig_{i} lie in only finitely many right cosets HgiHg_{i}. In other words, the sets gi(K)g_{i}(K) lie in only finitely many HH–orbits. But any two HH–orbits Hgi(K)Hg_{i}(K) and Hgj(K)Hg_{j}(K) lie at a finite Hausdorff distance from each other. Thus any gi(K)g_{i}(K) can be increased to a larger compact set KK^{\prime} so that the translates of KK^{\prime} under HH cover AA. Since AA is closed, it follows that HH acts cocompactly on AA, as desired. ∎

The proof of Theorem 1.3 will be developed over the course of the next several lemmas and definitions.

Let π:XT\pi\colon X\to T be the GG–equivariant continuous map given by Lemma 7.1. For each edge ee of TT, let mem_{e} denote the midpoint of ee. For each vertex vv of TT, let S(v)S(v) be the union of all segments of the form [v,me][v,m_{e}] where ee is an edge adjacent to vv; in other words, S(v)S(v) is the union of all half-edges emanating from vv. Let Q(e)XQ(e)\subset X be the preimage π1(me)\pi^{-1}(m_{e}), and let Q(v)XQ(v)\subset X be the preimage π1(S(v))\pi^{-1}\bigl{(}S(v)\bigr{)}.

Lemma 7.3.

Let GeG_{e} be the GG–stabilizer of the edge ee of TT. Then GeG_{e} acts cocompactly on Q(e)Q(e).

Proof.

The family 𝒜={Q(e)|e an edge of T}\mathcal{A}=\bigl{\{}\,{Q(e)}\bigm{|}{\text{$e$ an edge of $T$}}\,\bigr{\}} is clearly GG–equivariant. Each Q(e)Q(e) is closed in XX since it is the preimage of the closed set {me}\{m_{e}\} under a continuous map. We will show that 𝒜\mathcal{A} is locally finite, and that the GG–stabilizer of each Q(e)Q(e) is equal to the group GeG_{e}.

In order to see local finiteness, let KK be a compact set in XX. Since π:XT\pi\colon X\to T is continuous with respect to the CW topology on TT, the image π(K)\pi(K) intersects at most finitely many open edges of TT. In particular π(K)\pi(K) contains only finitely many midpoints mem_{e} of edges. It follows that KK intersects only finitely many sets Q(e)Q(e).

The equivariance of π\pi implies that an element gGg\in G satisfies g(Q(e))=Q(e)g\bigl{(}Q(e)\bigr{)}=Q(e) if and only if g(me)=meg(m_{e})=m_{e}. Thus Stab(Q(e))=Stab(me)=Ge\operatorname{Stab}\bigl{(}Q(e)\bigr{)}=\operatorname{Stab}(m_{e})=G_{e}. The result now follows from Proposition 7.2. ∎

Lemma 7.4.

Let GvG_{v} be the GG–stabilizer of the vertex vv of TT. Then GvG_{v} acts cocompactly on Q(v)Q(v).

Proof.

We only need to observe that when projecting a compact set KK to the tree TT, the image π(K)\pi(K) intersects only finitely many sets S(v)S(v) since it intersects only finitely many open edges of TT. The rest of the proof is identical to the proof of Lemma 7.3. ∎

In the proof of Theorem 1.3 we assume that we have a splitting whose edge groups are convex, in other words each edge group GeG_{e} stabilizes a closed convex subspace YeY_{e} on which it acts cocompactly. Although GeG_{e} also acts cocompactly on Q(e)Q(e), it is unlikely that Q(e)Q(e) itself is a convex subspace of XX. In the next lemma, we show that we can enlarge Q(e)Q(e) to its convex hull and preserve cocompactness.

Lemma 7.5.

Suppose the edge group GeG_{e} cocompactly stabilizes a convex subspace YeY_{e} of the CAT(0)\operatorname{CAT}(0) space XX. Let C(e)C(e) be the closure of the convex hull of Q(e)Q(e). Then GeG_{e} acts cocompactly on C(e)C(e).

Proof.

Let KK be a compact set of XX whose GeG_{e}–translates cover Q(e)Q(e). Then KK is contained in the closed neighborhood 𝒩D(Ye)¯\overline{\mathcal{N}_{D}({Y_{e}})} for some DD. Since XX is proper, any compact set KK^{\prime} whose translates cover YeY_{e} can be increased to a larger compact set 𝒩D(K)¯\overline{\mathcal{N}_{D}({K^{\prime}})} whose GeG_{e}–translates cover the closed neighborhood 𝒩D(Ye)¯\overline{\mathcal{N}_{D}({Y_{e}})}. This neighborhood is a closed, convex, GeG_{e}–cocompact set containing Q(e)Q(e). Thus it contains C(e)C(e), which is also GeG_{e}–cocompact. ∎

In the next proposition, we deal with the vertex groups. We replace Q(v)Q(v) with a larger set C(v)C(v). We first show that GvG_{v} acts cocompactly on C(v)C(v). Afterwards we will complete the proof of Theorem 1.3 by showing that C(v)C(v) is a convex subspace of XX.

Proposition 7.6.

For each vertex vTv\in T, let C(v)C(v) be the union of Q(v)Q(v) together with all sets C(e)C(e) such that ee is adjacent to vv in TT. Then GvG_{v} acts cocompactly on C(v)C(v).

Proof.

Since GG is finitely generated, we could have chosen TT to be a tree on which GG acts with quotient a finite graph (using standard Bass–Serre theory techniques). Instead, we will use the proof of Lemma 7.3 to show that each vertex in the quotient graph G\TG\backslash T has finite valence. For each vertex vTv\in T, the action of GvG_{v} permutes the subspaces Q(e)Q(e) such that ee is adjacent to vv. Since these subspaces form a locally finite family, Proposition 7.2 implies that they lie in finitely many GvG_{v}–orbits. It follows that there are finitely many GvG_{v}–orbits of sets C(e)C(e) where ee is adjacent to vv.

Let C(e1),,C(e)C(e_{1}),\dots,C(e_{\ell}) contain one from each GvG_{v}–orbit of the sets C(e)C(e). Let KiK_{i} be a compact set whose GeiG_{e_{i}}–translates cover C(ei)C(e_{i}). Let KK be a compact set whose GvG_{v}–translates cover Q(v)Q(v). Then C(v)C(v) is covered by the GvG_{v}–translates of the compact set KK1KK\cup K_{1}\cup\cdots\cup K_{\ell}. ∎

We now complete the proof of Theorem 1.3 by showing that the subspace C(v)C(v) constructed above is convex, which is the most delicate part of the argument.

Proof of Theorem 1.3.

Fix a vertex vTv\in T. We will show that C(v)C(v) is convex in XX. Let ee be any edge adjacent to vv. Recall that Q(e)Q(e) is defined to be π1(me)\pi^{-1}(m_{e}), where mem_{e} is the midpoint of the edge ee. Notice that mem_{e} splits the tree TT into two halfspaces. Thus Q(e)Q(e) can be considered as a “wall” in XX that separates XX into two halfspaces, each a preimage of a halfspace of TT. The key property of this separation that we need is that any path in XX from one halfspace of Q(e)Q(e) to the other must intersect their separating wall Q(e)Q(e). (These halfspaces in XX are preimages of connected sets, so they need not be connected. This will not cause any problems for our proof.)

Let H(e)H(e) be the closed halfspace of TT bounded by the midpoint mem_{e} and pointing away from the vertex vv. Let O(e)XO(e)\subset X be the intersection C(e)π1(H(e))C(e)\cap\pi^{-1}\bigl{(}H(e)\bigr{)}.

Our strategy is to first prove that C(v)C^{\prime}(v) is convex, where C(v)C^{\prime}(v) is the union of Q(v)Q(v) with all sets O(e)O(e) such that ee is adjacent to vv. We will complete the proof by showing that C(v)C^{\prime}(v) is actually equal to C(v)C(v).

Choose two points p,qC(v)p,q\in C^{\prime}(v). We claim that the geodesic cc in XX from pp to qq lies inside C(v)C^{\prime}(v). We have several cases depending on which subspaces they are chosen from.

Case 1. Suppose p,qQ(v)p,q\in Q(v). Any maximal subsegment cc^{\prime} of cc outside Q(v)Q(v) has its endpoints in Q(e)Q(e) for some ee adjacent to vv. By convexity, cc^{\prime} must be contained in C(e)C(e). But cc^{\prime} is outside Q(v)Q(v), so it must lie inside O(e)O(e). Therefore cc lies entirely inside C(v)C^{\prime}(v).

Case 2. Suppose p,qO(e)p,q\in O(e) for some ee adjacent to vv. Any maximal subsegment cc^{\prime} of cc outside O(e)O(e) has its endpoints in Q(e)Q(e). Therefore cc^{\prime} is a path with both endpoints in Q(v)Q(v). By Case 1, the path cc^{\prime} lies in C(v)C^{\prime}(v). Therefore cc lies in C(v)C^{\prime}(v) as well.

Case 3. Suppose pO(e)p\in O(e) and qQ(v)q\in Q(v). By separation, cc contains a point rQ(e)r\in Q(e). By the first two cases, the subsegments [p,r][p,r] and [r,q][r,q] each lie in C(v)C^{\prime}(v). Therefore we are done.

Case 4. Suppose pO(e)p\in O(e) and qO(e)q\in O(e^{\prime}) for eee\neq e^{\prime} two edges adjacent to vv. In this case, pp and qq are separated by two walls, Q(e)Q(e) and Q(e)Q(e^{\prime}). As in the previous case, we can choose rr and rr^{\prime} on cc such that rQ(e)r\in Q(e) and rQ(e)r^{\prime}\in Q(e^{\prime}). By the first two cases, the subsegments [p,r][p,r], [r,r][r,r^{\prime}] and [r,q][r^{\prime},q] each lie in C(v)C^{\prime}(v).

We have shown above that C(v)C^{\prime}(v) is convex. We observe that C(v)C^{\prime}(v) is closed, since it is equal to the union of a locally finite family of closed sets. Now let us see that C(v)C^{\prime}(v) is equal to our original C(v)C(v). Clearly C(v)C(v)C^{\prime}(v)\subseteq C(v). Recall that C(e)C(e) is the closure of the convex hull on Q(e)Q(e). By the above argument each C(e)C(e) is contained in the closed, convex set C(v)C^{\prime}(v). Therefore C(v)C(v)C(v)\subseteq C^{\prime}(v), as needed. ∎

Theorem 7.7.

Let GG be a one-ended group acting geometrically on a CAT(0)\operatorname{CAT}(0) space with isolated flats. Let 𝒢\mathcal{G} be any peripheral splitting of GG with respect to the natural peripheral structure \mathbb{P}. Then each component vertex group HH of the splitting 𝒢\mathcal{G} acts geometrically on a CAT(0)\operatorname{CAT}(0) space with isolated flats. Furthermore, the inclusion HGH\hookrightarrow G induces a topological embedding HG\partial H\hookrightarrow\partial G.

Proof.

Let \mathbb{Q} be the family of infinite subgroups of HH of the form HPH\cap P for PP\in\mathbb{P}. By a result of Bowditch [Bow01], we get that (H,)(H,\mathbb{Q}) is relatively hyperbolic with respect to a family of virtually abelian subgroups. In any peripheral splitting of GG, the edge groups are virtually abelian. By the Flat Torus Theorem, it follows that all edge groups of 𝒢\mathcal{G} are convex subgroups of GG. It follows from Theorem 1.3 that the vertex group HH acts geometrically on a CAT(0)\operatorname{CAT}(0) space YY, which is obtained as a closed convex subspace of XX. The space YY has isolated flats by Theorem 5.4. Since YY is convex in XX, the visual boundary H=Y\partial H=\partial Y embeds in X=G\partial X=\partial G

We obtain the following corollary by combining Theorem 7.7 with Theorem 6.4.

Corollary 7.8.

Let GG be a one-ended group acting geometrically on a CAT(0)\operatorname{CAT}(0) space with isolated flats. Suppose the maximal peripheral splitting 𝒢\mathcal{G} is locally finite. Then each component vertex group HH of 𝒢\mathcal{G} is a one-ended CAT(0)\operatorname{CAT}(0) group with isolated flats that is atomic. ∎

8. Local connectivity at rank one points

In this section, we begin the proof of Theorem 1.1. Our goal is to study the boundary of a one-ended CAT(0)\operatorname{CAT}(0) group with isolated flats. We need to show that if the maximal peripheral splitting is locally finite, then the boundary of the CAT(0)\operatorname{CAT}(0) space is locally connected. The first step of the proof, which is the main goal of the current section, is to show that the visual boundary X\partial X is locally connected at any point not in the boundary of a flat, i.e., the rank one points.

We use decomposition theory to prove Corollary 8.13, which states that the map X(G,)\partial X\to\partial(G,\mathbb{P}) given by Theorem 4.3 from the visual boundary to the Bowditch boundary is upper semicontinuous. Using this structure we are able to pull back local connectivity from the Bowditch boundary to the CAT(0)\operatorname{CAT}(0) boundary, but only at rank one points (see Corollary 8.14).

Many of our techniques in the proof of the main theorem depend on elementary results from decomposition theory, which we summarize below. We refer the reader to [Dav86] for more details. Recall that a decomposition 𝒟\mathcal{D} of a topological space MM is a partition of MM. Decompositions are equivalent to quotient maps in the following sense. A decomposition 𝒟\mathcal{D} of a space MM has a natural quotient map π:MM/𝒟\pi\colon M\to M/\mathcal{D} obtained by collapsing each equivalence class of 𝒟\mathcal{D} to a point and endowing the result with the quotient topology. Conversely, any quotient map MNM\to N gives rise to an associated decomposition of MM consisting of the family of point preimages.

Definition 8.1.

Let 𝒟\mathcal{D} be a decomposition of a space MM, and let SS be a subset of MM. The 𝒟\mathcal{D}–saturation of SS is the union of SS together with all sets D𝒟D\in\mathcal{D} that intersect SS.

Definition 8.2.

A decomposition 𝒟\mathcal{D} of a Hausdorff space MM is upper semicontinuous if each D𝒟D\in\mathcal{D} is compact, and if, for each D𝒟D\in\mathcal{D} and each open subset UU of MM containing DD, there exists another open subset VV of MM containing DD such that the 𝒟\mathcal{D}–saturation of VV is contained in UU. In other words, every D𝒟D^{\prime}\in\mathcal{D} intersecting VV is contained in UU. A quotient map is upper semicontinuous if its associated decomposition is upper semicontinuous.

The proof of the following result is left as a routine exercise to the reader (cf. Proposition I.1.1 of [Dav86]).

Proposition 8.3.

Let 𝒟\mathcal{D} be a decomposition of a Hausdorff space MM. The following are equivalent:

  1. (1)

    𝒟\mathcal{D} is upper semicontinuous.

  2. (2)

    For each open set UU of MM, let UU^{*} equal the union of all members of 𝒟\mathcal{D} contained in UU. Then UU^{*} is open.

  3. (3)

    For each closed set CC of MM, the 𝒟\mathcal{D}–saturation of CC is closed. ∎

Proposition 8.4.

Let 𝒟\mathcal{D} be an upper semicontinuous decomposition of a space MM. Let π:MM/𝒟\pi\colon M\to M/\mathcal{D} be the natural quotient map.

  1. (1)

    [Dav86, Prop. I.3.1] π\pi is a proper map; i.e., a set CM/𝒟C\subseteq M/\mathcal{D} is compact if and only if π1(C)\pi^{-1}(C) is compact.

  2. (2)

    [Dav86, Prop; I.4.1] Suppose each member of 𝒟\mathcal{D} is connected. A set CM/𝒟C\subseteq M/\mathcal{D} is connected if and only if π1(C)\pi^{-1}(C) is connected.

Proposition 8.5.

Let 𝒟\mathcal{D} be an upper semicontinuous decomposition of MM. Suppose each member of 𝒟\mathcal{D} is connected. Let {x}\{x\} be a singleton member of 𝒟\mathcal{D}. If M/𝒟M/\mathcal{D} is locally connected at the point π(x)\pi(x), then MM is locally connected at xx.

Proof.

Let UU be a neighborhood of xx in MM. By Proposition 8.3 there is a saturated neighborhood UU^{*} of xx contained in UU. Since M/𝒟M/\mathcal{D} is locally connected at π(x)\pi(x), the open set π(U)\pi(U^{*}) contains a connected open neighborhood V¯\overline{V} of π(x)\pi(x). By Proposition 8.4, the preimage π1(V¯)\pi^{-1}(\overline{V}) is a connected open neighborhood of xx contained in UU. ∎

Definition 8.6 (Subspace decomposition).

Let 𝒟\mathcal{D} be a decomposition of a Hausdorff space MM. Let WW be an open 𝒟\mathcal{D}–saturated subspace of MM. The induced subspace decomposition of WW is the decomposition consisting of all members of 𝒟\mathcal{D} that are contained in WW. If 𝒟\mathcal{D} is upper semicontinuous, then the induced subspace decomposition is as well.

Proposition 8.7 ([Dav86], Prop. I.2.2).

If MM is compact, metrizable and 𝒟\mathcal{D} is an upper semicontinuous decomposition of MM, then M/𝒟M/\mathcal{D} is metrizable.

Definition 8.8.

A collection of subsets 𝒜\mathcal{A} in a metric space is a null family if, for each ϵ>0\epsilon>0, only finitely many of the sets A𝒜A\in\mathcal{A} have diameter greater than ϵ\epsilon.

Note that if MM is compact and metrizable, then being a null family does not depend on the choice of metric on MM.

The following property of null families is related to the definition of upper semicontinuous, but here we do not require that the members of 𝒜\mathcal{A} be disjoint.

Proposition 8.9.

Let 𝒜\mathcal{A} be a null family of compact sets in a metric space MM. Suppose qMq\in M is not contained in any member of the family 𝒜\mathcal{A}. Then each neighborhood UU of qq contains a smaller neighborhood VV of qq such that each A𝒜A\in\mathcal{A} intersecting VV is contained in UU.

Proof.

Let UU be a neighborhood of qq, and suppose B(q,ϵ)UB(q,\epsilon)\subseteq U. Choose δ\delta such that 0<δ<ϵ/20<\delta<\epsilon/2 and such that d(q,A)>δd(q,A)>\delta for each of the finitely many A𝒜A\in\mathcal{A} with diameter greater than ϵ/2\epsilon/2. The result follows if we set V=B(q,δ)V=B(q,\delta). ∎

Remark 8.10.

The notion of a null family can be formulated in terms of the cone topology on X\partial X as follows. Fix a basepoint x0Xx_{0}\in X. A collection 𝒜\mathcal{A} of subspaces of X\partial X is a null family provided that there exists D>0D>0 such that for each r<r<\infty only finitely many members of the collection 𝒜\mathcal{A} are not contained in any set of the form U(,r,D)U(\cdot,r,D). It follows from Definition 3.4 that only finitely many members of 𝒜\mathcal{A} have diameter at least 1/r1/r with respect to the metric dDd_{D} on x0X\partial_{x_{0}}X.

A similar condition can be used to characterize null families in the cone topology on X¯=XX\overline{X}=X\cup\partial X. We leave the proof as an exercise for the reader.

When the collection 𝒜\mathcal{A} of subsets of MM is disjoint, there is an associated decomposition of MM consisting of the sets in 𝒜\mathcal{A} together with all singletons {x}\{x\} such that xM𝒜x\in M-\bigcup\mathcal{A}. By a slight abuse of notation, we let M/𝒜M/\mathcal{A} denote the corresponding quotient in which each member of 𝒜\mathcal{A} is collapsed to a point.

Decompositions arising from null families play a central role in the proof of the main theorem. The following result is stated as an exercise in [Dav86]. The proof is nearly identical to the proof of Proposition 8.9.

Proposition 8.11 ([Dav86], Prop. I.2.3).

Let 𝒜\mathcal{A} be a null family of disjoint compact subsets in a metric space MM. Then the associated decomposition of MM is upper semicontinuous.

Proposition 8.12.

Let XX be a CAT(0)\operatorname{CAT}(0) space that has isolated flats with respect to the family of flats \mathcal{F}. Let 𝒜\mathcal{A} be the family of spheres {FF}\{\,{\partial F}\mid{F\in\mathcal{F}}\,\}. Then 𝒜\mathcal{A} is a null family of disjoint compact subsets in X\partial X.

Proof.

Each F𝒜\partial F\in\mathcal{A} is a sphere, which is compact. The definition of isolated flats immediately implies that the members of 𝒜\mathcal{A} are pairwise disjoint. Thus we only need to show they are a null family.

Choose a basepoint x0Xx_{0}\in X. Let κ\kappa be the constant from Theorem 5.6. We will prove the following claim below: for any flat FF\in\mathcal{F} satisfying d(x0,F)r+3κd(x_{0},F)\geq r+3\kappa for some constant rr, there exists a geodesic ray cc based at x0x_{0} such that FU(c,r,7κ)\partial F\subseteq U(c,r,7\kappa).

Since the collection of flats \mathcal{F} is locally finite, there are only finitely many within a distance r+3κr+3\kappa of x0x_{0} for each r<r<\infty. Thus it will follow from Remark 8.10 that 𝒜\mathcal{A} is a null family.

In order to prove the claim, let qq be the nearest point in FF to x0x_{0}. Then d(x0,q)r+3κd(x_{0},q)\geq r+3\kappa. Let c,cc,c^{\prime} be geodesic segments from x0x_{0} to FF. By Theorem 5.6 the set [x0,q]F[x_{0},q]\cup F is κ\kappa–quasiconvex. Thus there exist s,ss,s^{\prime} with c(s),c(s)c(s),c^{\prime}(s^{\prime}) both contained in 𝒩κ([x0,q])𝒩κ(F)\mathcal{N}_{\kappa}\bigl{(}{[x_{0},q]}\bigr{)}\cap\mathcal{N}_{\kappa}({F}). It follows that d(c(s),q)d\bigl{(}c(s),q\bigr{)} and d(c(s),q)d\bigl{(}c^{\prime}(s^{\prime}),q\bigr{)} are each less than 3κ3\kappa. Thus d(c(s),c(s))<6κd\bigl{(}c(s),c^{\prime}(s^{\prime})\bigr{)}<6\kappa. By the Law of Cosines d(c(r),c(r))<6κd\bigl{(}c(r),c^{\prime}(r)\bigr{)}<6\kappa. In particular cU(c,r,6κ)c^{\prime}\in U(c,r,6\kappa).

Now suppose cc and cc^{\prime} are geodesic rays asymptotic to FF (possibly not intersecting FF). Then each is a limit of geodesic segments that intersect FF. In this case, we conclude that cU(c,r,7κ)c^{\prime}\in U(c,r,7\kappa). Therefore FU(c,r,7κ)\partial F\subseteq U(c,r,7\kappa) for any cc with c()Fc(\infty)\in\partial F, establishing the claim. ∎

Corollary 8.13.

Let GG act geometrically on a CAT(0)\operatorname{CAT}(0) space XX with isolated flats. Let \mathbb{P} be the standard relatively hyperbolic structure on GG. Then the quotient map XX/𝒜(G,)\partial X\to\partial X/\mathcal{A}\to\partial(G,\mathbb{P}) given by Theorem 4.3 is upper semicontinuous. ∎

Corollary 8.14.

Let GG be a one-ended group acting geometrically on a CAT(0)\operatorname{CAT}(0) space XX with isolated flats. Then X\partial X is locally connected at any point ξ\xi not in the boundary of any flat.

Proof.

By Theorem 6.3, the Bowditch boundary (G,)\partial(G,\mathbb{P}) is locally connected at every point. Each member of the decomposition 𝒜\mathcal{A} is either a point or a sphere SkS^{k} with k>0k>0. Thus all members of 𝒜\mathcal{A} are connected. The result follows immediately from Proposition 8.5. ∎

9. Local connectivity on the boundary of a flat

Recall our main goal is to establish local connectivity of the visual boundary of a CAT(0)\operatorname{CAT}(0) group with isolated flats in the setting where the maximal peripheral splitting is locally finite.

In this section we focus on the special case where the maximal peripheral splitting is trivial, in other words, we study atomic groups. The main goal of this section is to prove Theorem 1.4, which states that the boundary of an atomic group is always locally connected.

In the previous section, we showed that the boundary of such a group is locally connected at any point not in the boundary of a flat subspace. To reach our goal, it suffices to show that the boundary X\partial X is weakly locally connected at points of F\partial F, where FF is a flat subspace. Recall that a space is weakly locally connected at a point ξ\xi if ξ\xi has a local base of (not necessarily open) connected neighborhoods. A space is locally connected if it is weakly locally connected at every point.

In order to understand the topology of X\partial X near a point of F\partial F, we partition X\partial X into F\partial F and its complement Υ=XF\Upsilon=\partial X-\partial F. If PP is the stabilizer of FF, then we will see that PP acts properly and cocompactly on FF, on Υ\Upsilon, and also on (G,){ρ}\partial(G,\mathbb{P})-\{\rho\}, where ρ\rho is the parabolic point corresponding to F\partial F. Our main strategy is to exploit similarities between these three spaces. Many of these similarities do not not depend on the extra hypothesis that GG does not peripherally split. After developing some features of this similarity, we will add the extra “atomic” hypothesis, which implies that both FF and Υ\Upsilon are 0–connected spaces. Since they share proper and cocompact group actions by the same group, we are able to transfer 0–connectedness properties between them. In particular, the local connectedness of X=ΥF\partial X=\Upsilon\cup\partial F at a point ξF\xi\in\partial F will follow from the local connectedness of F¯=FF\overline{F}=F\cup\partial F at ξ\xi.

The similarity between FF and Υ\Upsilon was first introduced and extensively studied by Haulmark [Hau18] in the general situation of groups acting on CAT(0)\operatorname{CAT}(0) spaces with isolated flats (without the “atomic” hypothesis). This similarity depends heavily on the following two lemmas that play a significant role in Haulmark’s work.

Lemma 9.1 ([Hau18]).

Let GG be a group acting geometrically on a CAT(0)\operatorname{CAT}(0) space XX with isolated flats with respect to the family \mathcal{F}. Let FF\in\mathcal{F} be a flat with stabilizer PP, and let Υ=XF\Upsilon=\partial X-\partial F. For each compact set KΥK\subset\Upsilon there exists a compact set CFC\subset F such that for each ηK\eta\in K there is a geodesic ray cc^{\prime} with c(0)Cc^{\prime}(0)\in C and c()=ηc^{\prime}(\infty)=\eta such that cc^{\prime} meets FF orthogonally. Furthermore the compact set CC can be chosen PP–equivariantly in the sense that if pPp\in P then pCpC is the compact set of FF corresponding to pKpK.

We occasionally apply the previous lemma in the following special case: each point of Υ\Upsilon is the endpoint of a geodesic ray meeting FF orthogonally.

The following corollary of Theorem 5.6 was first observed by Haulmark.

Lemma 9.2 ([Hau18]).

Let κ\kappa be the constant given by Theorem 5.6. Let Υ=XF\Upsilon=\partial X-\partial F, and suppose cc^{\prime} is a geodesic ray meeting FF orthogonally. Suppose cc is a geodesic ray contained in FF. If c(0)U(c,r,D)c^{\prime}(0)\in U(c,r,D) for some constants rr and DD then c()U(c,r,D+κ)c^{\prime}(\infty)\in U(c,r,D+\kappa). Conversely if c()U(c,r,D)c^{\prime}(\infty)\in U(c,r,D) then c(0)U(c,r,D+κ)c^{\prime}(0)\in U(c,r,D+\kappa).

In any CAT(0)\operatorname{CAT}(0) space XX with a geometric group action, the family of translates of a compact fundamental domain is a null family in the following sense.

Proposition 9.3 ([Bes96]).

Let HH be any group acting geometrically on a CAT(0)\operatorname{CAT}(0) space YY. Let CYC\subset Y be any compact set. Then the collection of HH–translates of CC is a null family in the compact space Y¯\overline{Y}.

We will use the previous proposition in the case when YY is a flat subspace FF of a CAT(0)\operatorname{CAT}(0) space with isolated flats, and HH is its stabilizer PP.

The similarity between FF and Υ\Upsilon allows us to transfer this version of the null condition to the action of PP on Υ\Upsilon as follows:

Proposition 9.4.

Let GG act geometrically on a space XX with isolated flats. Choose FF\in\mathcal{F} with stabilizer PP, and let Υ=XF\Upsilon=\partial X-\partial F. If KΥK\subset\Upsilon is compact, then the collection of PP–translates of KK is a null family in Υ¯=X\overline{\Upsilon}=\partial X.

Proof.

Choose a compact set KΥK\subset\Upsilon. Our strategy is to exploit the similarity between Υ\Upsilon and FF as follows: use Lemma 9.1 to pull KK back to a compact set CC in FF, use Proposition 9.3 to see that almost all translates of CC are “small” in F¯\overline{F}, and then apply Lemma 9.2 to conclude that the corresponding translates of KK are similarly small in Υ¯=X\overline{\Upsilon}=\partial X.

For our given compact set KΥK\subset\Upsilon, let CC be the corresponding compact set of FF given by Lemma 9.1. Fix a positive constant DD, and let κ\kappa be the constant from Theorem 5.6. By Remark 8.10, it suffices to show that for each r<r<\infty only finitely many PP–translates of KK are not contained in any set of the form U(,r,D+κ)U(\cdot,r,D+\kappa). According to Proposition 9.3, the PP–translates of CC are a null family. Thus only finitely many PP–translates of CC are not contained in a set of the form U(,r,D)U(\cdot,r,D). For any pPp\in P, if pCpC lies in a set of the form U(,r,D)U(\cdot,r,D) then the corresponding set pKpK of Υ\Upsilon lies in a set of the form U(,r,D+κ)U(\cdot,r,D+\kappa) by Lemma 9.2. So {pK}\{pK\} is a null family in Υ¯\overline{\Upsilon}. ∎

Since PP acts cocompactly on FF, Haulmark exploited the similarity between FF and Υ\Upsilon to show that PP also acts cocompactly on Υ\Upsilon [Hau18].

For the rest of this section, we focus on the special setting where GG is atomic. In this special situation, we can improve the conclusion of Haulmark’s cocompactness theorem to get a compact, connected fundamental domain for the action of PP on Υ\Upsilon.

Proposition 9.5.

Let GG be a one-ended group acting geometrically on a CAT(0)\operatorname{CAT}(0) space XX that has isolated flats with respect to the family of flats \mathcal{F}. Suppose GG is atomic. For each flat FF\in\mathcal{F} with stabilizer PP, there is a compact connected set KK in Υ=XF\Upsilon=\partial X-\partial F whose PP–translates cover Υ\Upsilon.

Furthermore if 𝒯\mathcal{T} is any finite generating set for the group PP, we can choose KK large enough that KK intersects tKtK for all t𝒯t\in\mathcal{T}.

The previous results of this section are proved using the similarity between Υ\Upsilon and FF. However for the proof of Proposition 9.5, we use the other similarity between Υ\Upsilon and (G,){ρ}\partial(G,\mathbb{P})-\{\rho\}, where ρ\rho is the parabolic point corresponding to F\partial F. The proof of the proposition relies on the fact that each parabolic point ρ\rho of the Bowditch boundary is bounded parabolic; i.e., there is a compact fundamental domain for the action of its stabilizer on (G,){ρ}\partial(G,\mathbb{P})-\{\rho\}. (See Definition 4.1.)

We also need the following lemma, which allows us to increase any such compact fundamental domain to a connected one, provided that GG is atomic.

Lemma 9.6.

Let (G,)(G,\mathbb{P}) be relatively hyperbolic. Suppose GG is atomic and each PP\in\mathbb{P} is finitely presented and does not contain an infinite torsion subgroup. Let ρ(G,)\rho\in\partial(G,\mathbb{P}) be a parabolic point with stabilizer PP. Let C0C_{0} be any compact fundamental domain for the action of PP on (G,){ρ}\partial(G,\mathbb{P})-\{\rho\}. Then C0C_{0} is contained in a compact connected fundamental domain CC.

Proof.

By Theorem 6.3 the Bowditch boundary (G,)\partial(G,\mathbb{P}) is connected and locally connected, and the parabolic point ρ\rho is not a global cut point. Thus (G,){ρ}\partial(G,\mathbb{P})-\{\rho\} is an open, connected subset of the compact, locally connected, metrizable space (G,)\partial(G,\mathbb{P}). It follows that (G,){ρ}\partial(G,\mathbb{P})-\{\rho\} is path connected by [Wil70, 31C.1].

Let dd be a metric on (G,)\partial(G,\mathbb{P}), and let ϵ=d(ρ,C0)\epsilon=d(\rho,C_{0}). We can cover the compact set C0C_{0} by finitely many open connected sets with diameter less then ϵ/2\epsilon/2. In (G,)\partial(G,\mathbb{P}) the union of the closures of these sets is a compact set C1C_{1} containing C0C_{0} and having only finitely many components. By our choice of ϵ\epsilon, the compact set C1C_{1} is contained in (G,){ρ}\partial(G,\mathbb{P})-\{\rho\}.

Finally we form CC from C1C_{1} by attaching finitely many compact paths in (G,){ρ}\partial(G,\mathbb{P})-\{\rho\} that connect the finitely many components of C1C_{1}. ∎

Proof of Proposition 9.5.

Our strategy is to use the quotient map π:X(G,)\pi\colon\partial X\to\partial(G,\mathbb{P}) given by Theorem 4.3. We will find an appropriate fundamental domain in the Bowditch boundary and then pull it back via π\pi to get a compact connected fundamental domain in X\partial X.

Let ρ\rho be the parabolic point of (G,)\partial(G,\mathbb{P}) stabilized by PP; i.e., {ρ}\{\rho\} is the image of F\partial F in the Bowditch boundary. By the definition of relative hyperbolicity, the action of PP on (G,){ρ}\partial(G,\mathbb{P})-\{\rho\} has a compact fundamental domain C0C_{0}. Increasing the size of C0C_{0}, we may assume without loss of generality that C0C_{0} intersects the finitely many translates tC0tC_{0} for all t𝒯t\in\mathcal{T}. By Lemma 9.6, we can increase C0C_{0} to a compact, connected fundamental domain CC intersecting its translates tCtC for all t𝒯t\in\mathcal{T}.

Recall that the quotient π:X(G,)\pi\colon\partial X\to\partial(G,\mathbb{P}) collapses connected sets to points; i.e. each member of the associated decomposition of X\partial X is either a point or the boundary of a one-ended peripheral subgroup (in our case this boundary is a sphere). By Corollary 8.13, the quotient π\pi is upper semicontinuous. It follows from Proposition 8.4 that the preimage K=π1(C)K=\pi^{-1}(C) is compact and connected. Theorem 4.3 implies that π\pi is GG–equivariant. Thus KK is a fundamental domain for the action of PP on Υ\Upsilon, and KK intersects its translates tKtK for each t𝒯t\in\mathcal{T}. ∎

Our goal for the rest of this section is to prove the following proposition.

Proposition 9.7.

Let GG be a one-ended group acting geometrically on a CAT(0)\operatorname{CAT}(0) space with isolated flats. Assume GG is atomic. Then X\partial X is weakly locally connected at any point in the boundary of any flat.

The proof of Proposition 9.7 depends on three lemmas. Before discussing the lemmas, we outline the broad strategy that leads to the proof. Recall that X=Υ¯\partial X=\overline{\Upsilon} is similar in many ways to F¯\overline{F}. We know that F¯\overline{F} is locally connected at each point ξF\xi\in\partial F by Proposition 3.3. In order to prove that Υ¯\overline{\Upsilon} is also locally connected at ξ\xi, we will describe a procedure for transferring small connected neighborhoods of ξ\xi from F¯\overline{F} to Υ¯\overline{\Upsilon}, which is valid when GG does not peripherally split.

The atomic hypothesis implies that Υ\Upsilon is connected. Obviously FF is also connected. The foundation of our strategy is the following lemma, which allows us to transfer 0–connectedness from FF to Υ\Upsilon using the fact that both FF and Υ\Upsilon are 0–connected spaces on which the same group PP acts properly and cocompactly.

Lemma 9.8.

Assume GG is atomic. There exist compact, connected fundamental domains CFC\subset F and KΥK\subset\Upsilon for the actions of PP on each, such that the following holds. Let 𝒫\mathcal{P} be any subset of the group PP.

If p𝒫pC\bigcup_{p\in\mathcal{P}}pC is connected in FF, then p𝒫pK\bigcup_{p\in\mathcal{P}}pK is connected in Υ\Upsilon.
Proof.

Choose a compact connected fundamental domain CC for the action of PP on FF. Let 𝒯\mathcal{T} be the set of elements tPt\in P such that CC intersects tCtC. Then 𝒯\mathcal{T} is a finite generating set for PP. Choose a compact connected fundamental domain KK for the action of PP on Υ\Upsilon as given by Proposition 9.5 such that KK intersects tKtK for all t𝒯t\in\mathcal{T}. This intersection property immediately implies the following condition that holds for all pPp\in P:

If CpCC\cap pC is nonempty, then KpKK\cap pK is nonempty.

This condition easily implies our conclusion. ∎

The following terminology and notation will be used throughout the rest of this section and the eventual proof of Proposition 9.7. Let CC and KK be the compact, connected fundamental domains given by the previous lemma. By Lemma 9.1 there exists a geodesic ray cc^{\prime} in XX meeting FF orthogonally. We will treat the points q0=c(0)Fq_{0}=c^{\prime}(0)\in F and q=c()Υq_{\infty}=c^{\prime}(\infty)\in\Upsilon as basepoints in FF and Υ\Upsilon respectively. Translating CC and KK by the cocompact group actions, we can also assume that q0Cq_{0}\in C and qKq_{\infty}\in K.

Suppose ξF\xi\in\partial F. As mentioned above, our strategy for proving Proposition 9.7 is to transfer small connected neighborhoods of ξ\xi in F¯\overline{F} to small connected neighborhoods of ξ\xi in Υ¯\overline{\Upsilon}. To facilitate this transfer, we assume that the given neighborhood N¯\overline{N} of ξ\xi in F¯\overline{F} is clean in the sense that N=N¯FN=\overline{N}\cap F is connected, and each point of N¯\overline{N} is a limit point of NN. Recall that ξ\xi has a local base of clean connected neighborhoods by Proposition 3.3.

For each clean connected neighborhood N¯\overline{N}, we will define a corresponding set Z¯\overline{Z} in Υ¯=X\overline{\Upsilon}=\partial X. In the two subsequent lemmas, we will show that Z¯\overline{Z} is a connected neighborhood of ξ\xi in Υ\Upsilon, and that Z¯\overline{Z} can be chosen arbitrarily small. We begin with the construction of Z¯\overline{Z}.

Definition 9.9 (Associated neighborhoods).

Suppose ξF\xi\in\partial F and N¯\overline{N} is a clean connected neighborhood of ξ\xi in F¯\overline{F}. Let N=N¯FN=\overline{N}\cap F, and let Λ=N¯F\Lambda=\overline{N}\cap\partial F. Let 𝒫\mathcal{P} be the set of all pPp\in P such that pCpC intersects NN. The corresponding set ZΥZ\subset\Upsilon is the union p𝒫pK\bigcup_{p\in\mathcal{P}}pK. Finally the Υ¯\overline{\Upsilon}–neighborhood associated to N¯\overline{N} is the set Z¯=ZΛ\overline{Z}=Z\cup\Lambda.

For this definition to make sense, we must verify that Z¯\overline{Z} is actually a neighborhood of ξ\xi in Υ¯\overline{\Upsilon}. The next lemma establishes that Z¯\overline{Z} is, in fact, a (clean) connected neighborhood.

Lemma 9.10.

For any clean, connected neighborhood N¯\overline{N} of ξ\xi in F¯\overline{F}, the Υ¯\overline{\Upsilon}–neighborhood Z¯\overline{Z} associated to N¯\overline{N}, defined above, is a connected neighborhood of ξ\xi in Υ¯\overline{\Upsilon}.

Proof.

We first verify that Z¯\overline{Z} is a neighborhood of ξ\xi in Υ¯\overline{\Upsilon}. Let κ\kappa be the constant given by Theorem 5.6, and let D>0D>0 be an arbitrary constant. Choose a geodesic ray cc in FF with c()=ξc(\infty)=\xi. Since N¯\overline{N} is a neighborhood of ξ\xi in F¯\overline{F}, we can choose RR large enough so that U(c,R,D)F¯U(c,R,D)\cap\overline{F} lies inside N¯\overline{N}. By Proposition 9.4, the collection of PP–translates of our compact fundamental domain KK is a null family in Υ¯\overline{\Upsilon}. Therefore there exists a neighborhood VV of ξ\xi in Υ¯\overline{\Upsilon} such that VU(c,R,D+κ)V\subseteq U(c,R,D+\kappa), and such that every pKpK intersecting VV is contained in U(c,R,D+κ)U(c,R,D+\kappa) by Proposition 8.9.

It follows that VZ¯V\subseteq\overline{Z}. Indeed, each element ηV\eta\in V either lies in Υ\Upsilon or in F\partial F. In the first case, by our choice of KK, the point η\eta lies in pKpK for some pPp\in P. Each such pKpK is contained in U(c,R,D+κ)U(c,R,D+\kappa). In particular, p(q)p(q_{\infty}) lies in U(c,R,D+κ)Υ¯U(c,R,D+\kappa)\subset\overline{\Upsilon}. Thus p(q0)U(c,R,D)F¯p(q_{0})\in U(c,R,D)\subset\overline{F} by Lemma 9.2. Since p(q0)pCp(q_{0})\in pC, our choice of RR implies that the PP–translate pCpC intersects NN. By the definition of ZZ, we have ηpKZZ¯\eta\in pK\subset Z\subset\overline{Z}.

In the second case, we have ηVF\eta\in V\cap\partial F, so

ηU(c,R,D+κ)FU(c,R,D)FN¯F=ΛZ¯.\eta\ \in\ U(c,R,D+\kappa)\cap\partial F\ \subseteq\ U(c,R,D)\cap\partial F\ \subseteq\ \overline{N}\cap\partial F\ =\ \Lambda\ \subseteq\ \overline{Z}.

Combining the two cases, we see that VZ¯V\subseteq\overline{Z}, as desired. It follows that Z¯\overline{Z} is a neighborhood of ξ\xi in Υ¯\overline{\Upsilon}.

Next we will see that Z¯\overline{Z} is connected. The cleanliness of N¯\overline{N} means that N=N¯FN=\overline{N}\cap F is connected. Recall that 𝒫\mathcal{P} is the set of all pp such that pCpC intersects NN. Since CC is connected, the union N^=p𝒫pC\widehat{N}=\bigcup_{p\in\mathcal{P}}pC is also connected. By Lemma 9.8, the union Z=p𝒫pKZ=\bigcup_{p\in\mathcal{P}}pK is connected as well.

In order to show that Z¯=ZΛ\overline{Z}=Z\cup\Lambda is connected, it suffices to show that every point of Λ\Lambda is a limit point of ZZ. Since N¯\overline{N} is clean, each point ζΛ\zeta\in\Lambda is a limit of a sequence {xi}\{x_{i}\} in NN. Each xipiCx_{i}\in p_{i}C for some pi𝒫p_{i}\in\mathcal{P}, and by Proposition 9.3 the sequence {pi(q0)}\bigl{\{}p_{i}(q_{0})\bigr{\}} also converges to ζ\zeta. By Lemma 9.2, the sequence {pi(q)}\bigl{\{}p_{i}(q_{\infty})\bigr{\}} converges to ζ\zeta as well. Since qKq_{\infty}\in K, we have pi(q)piKZp_{i}(q_{\infty})\in p_{i}K\subseteq Z. Thus ζ\zeta is a limit point of ZZ. ∎

Lemma 9.11.

The Υ¯\overline{\Upsilon}–neighborhood Z¯\overline{Z} associated to N¯\overline{N} can be made arbitrarily small by choosing N¯\overline{N} to be a sufficiently small neighborhood of ξ\xi.

Proof.

Let UU be a neighborhood of ξ\xi in Υ¯=X\overline{\Upsilon}=\partial X. Our goal is to show that if N¯\overline{N} is chosen appropriately, its associated neighborhood Z¯\overline{Z} will be contained in UU. By Proposition 9.4, there is a neighborhood VV of ξ\xi in Υ¯\overline{\Upsilon} such that VUV\subseteq U and every pKpK intersecting VV is contained in UU. By Lemma 9.2, there is a neighborhood WW of ξ\xi in F¯\overline{F} such that WFVW\cap\partial F\subset V and such that if p(q0)Wp(q_{0})\in W then p(q)Vp(q_{\infty})\in V and hence pKUpK\subset U. Proposition 9.3 gives a neighborhood WW^{\prime} of ξ\xi in F¯\overline{F} such that WWW^{\prime}\subseteq W and every pCpC intersecting WW^{\prime} is contained in WW. Due to Proposition 3.3, there is a clean connected neighborhood N¯\overline{N} of ξ\xi inside WW^{\prime}.

It follows that the associated Υ¯\overline{\Upsilon}–neighborhood Z¯\overline{Z} is contained in UU. Indeed, each element ηZ¯\eta\in\overline{Z} either lies in ZZ or in Λ=N¯F\Lambda=\overline{N}\cap\partial F. Suppose first that ηZ\eta\in Z. Then ηpK\eta\in pK for some pp such that pCpC intersects NN, and NWN\subseteq W^{\prime}. In this case, it is clear from our choices above that ηU\eta\in U. On the other hand, suppose ηΛ\eta\in\Lambda. Since ΛWF\Lambda\subseteq W\cap\partial F, it is contained in VV, which is contained in UU. We have shown that Z¯U\overline{Z}\subseteq U, as needed. ∎

At this point the proof of Proposition 9.7 is nearly complete.

Proof of Proposition 9.7.

Suppose ξF\xi\in\partial F. We must show that X=Υ¯\partial X=\overline{\Upsilon} is weakly locally connected at ξ\xi. The combination of Lemmas  9.10 and 9.11 implies that ξ\xi has arbitrarily small connected neighborhoods (that are not necessarily open). ∎

Combining Corollary 8.14 and Proposition 9.7 completes the proof of Theorem 1.4.

10. The limit of a tree system of spaces

In order to complete the proof of Theorem 1.1, we need to examine the boundary of a one-ended CAT(0)\operatorname{CAT}(0) group GG with isolated flats whose maximal peripheral splitting is nontrivial and locally finite.

By Corollary 7.8, the component vertex groups of this splitting are atomic CAT(0)\operatorname{CAT}(0) groups with isolated flats. Therefore by Theorem 1.4 the boundary of each component vertex group is locally connected. We will see in Section 11 that the visual boundary of GG is obtained by gluing copies of the component group boundaries along spheres in the pattern of the Bass-Serre tree. The tool necessary to make this precise is the notion of a tree system of spaces, introduced by Świątkowski in [Świ].

In this section we present the definition of the limit of a tree system of metric compacta. We also prove Theorem 10.6, which roughly states that a tree system of locally connected spaces has a limit that is also locally connected.

Definition 10.1.

A tree is a connected nonempty graph without circuits. We use Serre’s notation for graphs with oriented edges. A graph has a vertex set 𝒱\mathcal{V}, a set of oriented edges \mathcal{E}, a map 𝒱×𝒱\mathcal{E}\to\mathcal{V}\times\mathcal{V} denoted e(o(e),t(e))e\mapsto\bigl{(}o(e),t(e)\bigr{)}, and an involution \mathcal{E}\to\mathcal{E} denoted ee¯e\mapsto\overline{e}. We require that e¯e\overline{e}\neq e and that o(e)=t(e¯)o(e)=t(\overline{e}). The vertices o(e)o(e) and t(e)t(e) are the origin and terminus of the edge ee. We refer the reader to [Ser77] for more details.

Definition 10.2.

A tree system Θ\Theta of metric compacta consists of the following data:

  1. (1)

    TT is a bipartite tree with a countable vertex set 𝒱=𝒞𝒫\mathcal{V}=\mathcal{C}\coprod\mathcal{P} such that each vertex w𝒫w\in\mathcal{P} has finite valence.

  2. (2)

    To each vertex v𝒱v\in\mathcal{V} there is associated a compact metric space KvK_{v}.

  3. (3)

    To each edge ee\in\mathcal{E} there is associated a compact metric space Σe\Sigma_{e}, a homeomorphism ϕe:ΣeΣe¯\phi_{e}\colon\Sigma_{e}\to\Sigma_{\overline{e}} such that ϕe¯=ϕe1\phi_{\overline{e}}=\phi_{e}^{-1}, and a topological embedding ie:ΣeKt(e)i_{e}\colon\Sigma_{e}\to K_{t(e)}.

  4. (4)

    For each v𝒞v\in\mathcal{C} the family of subspaces {ie(Σe)t(e)=v}\{\,{i_{e}(\Sigma_{e})}\mid{t(e)=v}\,\} is null and consists of pairwise disjoint sets. We will refer to the spaces KvK_{v} for v𝒱v\in\mathcal{V} as component spaces.

  5. (5)

    For each w𝒫w\in\mathcal{P} and each ee with t(e)=wt(e)=w the map ie:ΣeKwi_{e}\colon\Sigma_{e}\to K_{w} is a homeomorphism. We will refer to the spaces KwK_{w} for w𝒫w\in\mathcal{P} as peripheral spaces.

The tree system Θ\Theta is degenerate if the tree TT contains only one vertex, and that vertex is peripheral.

Remark 10.3.

The definition above of tree system is slightly more general than the one used by Świątkowski in [Świ]. In the special case that each vertex w𝒫w\in\mathcal{P} has valence two, our definition is equivalent to Świątkowski’s using a barycentric subdivision of his tree. The proofs in [Świ] generalize to tree systems in the sense of Definition 10.2 with only minor modifications.

Let #Θ\#\Theta denote the quotient (v𝒞𝒫Kv)/\bigl{(}\coprod_{v\in\mathcal{C}\cup\mathcal{P}}K_{v}\bigr{)}/\!\sim by the equivalence relation generated by ie(x)ie¯ϕe(x)i_{e}(x)\sim i_{\overline{e}}\phi_{e}(x) for all edges ee\in\mathcal{E} and all xΣex\in\Sigma_{e} endowed with the quotient topology.

For each subtree SS of TT let 𝒱S=𝒞S𝒫S\mathcal{V}_{S}=\mathcal{C}_{S}\coprod\mathcal{P}_{S} and S\mathcal{E}_{S} denote the set of vertices and the set of edges of SS. Let ΘS\Theta_{S} denote the restriction of the tree system Θ\Theta to the subtree SS. Let NS={eo(e)𝒱S and t(e)𝒱S}N_{S}=\{\,{e\in\mathcal{E}}\mid{o(e)\notin\mathcal{V}_{S}\text{ and }t(e)\in\mathcal{V}_{S}}\,\} be the set of oriented edges adjacent to SS but not contained in SS, oriented towards SS.

Definition 10.4 (Limit of a tree system).

For each finite subtree FF of TT, the partial union KFK_{F} is defined to be #ΘF\#\Theta_{F}. Since FF is finite it follows that KFK_{F} is compact and metrizable. Let 𝒜F={ie(Σe)eNF}\mathcal{A}_{F}=\{\,{i_{e}(\Sigma_{e})}\mid{e\in N_{F}}\,\}. We consider 𝒜F\mathcal{A}_{F} to be a family of subsets of KFK_{F}, and note that this is a null family that consists of pairwise disjoint compact sets. Let KF=KF/𝒜FK_{F}^{*}=K_{F}/\mathcal{A}_{F}, the quotient formed by collapsing each set in 𝒜F\mathcal{A}_{F} to a point. By Propositions 8.11 and 8.7, the quotient KFK_{F}^{*} is metrizable.

For each pair of finite subtrees F1F2F_{1}\subseteq F_{2}, let fF1F2:KF2KF1f_{F_{1}F_{2}}\colon K^{*}_{F_{2}}\to K^{*}_{F_{1}} be the quotient map obtained by collapsing KsK_{s} to a point for each s𝒱F2𝒱F1s\in\mathcal{V}_{F_{2}}-\mathcal{V}_{F_{1}} and identifying the resulting quotient space with KF1K_{F_{1}}^{*}. Since fF1F2fF2F3=fF1F3f_{F_{1}F_{2}}\circ f_{F_{2}F_{3}}=f_{F_{1}F_{3}} whenever F1F2F3F_{1}\subseteq F_{2}\subseteq F_{3}, the system of spaces KFK_{F}^{*} and maps fFFf_{FF^{\prime}} where FFF\subseteq F^{\prime} is an inverse system of metric compacta indexed by the poset of all finite subtrees FF of TT.

The limit limΘ\lim\Theta of the tree system Θ\Theta is the inverse limit of the above inverse system. Observe that limΘ\lim\Theta is compact and metrizable, since it is an inverse limit of a countable system of compact metrizable spaces.

A function f:YZf\colon Y\to Z is monotone if ff is surjective and for each zZz\in Z the preimage f1(z)f^{-1}(z) is compact and connected. The following theorem due to Capel is used in the proof of Theorem 10.6.

Theorem 10.5 ([Cap54]).

Let {Xα}\{X_{\alpha}\} be an inverse system such that each bonding map XαXβX_{\alpha}\to X_{\beta} is monotone. If each factor space XαX_{\alpha} is compact and locally connected then the inverse limit limXα\varprojlim X_{\alpha} is locally connected.

Theorem 10.6.

The limit limΘ\lim\Theta of a nondegenerate tree system Θ\Theta is locally connected, provided that each component vertex space KvK_{v} with v𝒞v\in\mathcal{C} is connected and locally connected and that each peripheral vertex space KwK_{w} with w𝒫w\in\mathcal{P} is nonempty.

Proof.

Since the component vertex spaces are connected and the peripheral vertex spaces are nonempty, KFK_{F} and KFK_{F}^{*} are connected for each finite subtree FF of TT. Recall that any quotient of a locally connected space is locally connected. Since KFK_{F} is obtained by gluing finitely many locally connected spaces, it is locally connected itself. Since KFK_{F}^{*} is a quotient of KFK_{F}, it is also locally connected.

In order to apply Theorem 10.5 to see that limΘ\lim\Theta is locally connected, it suffices to check that the bonding maps KF2KF1K^{*}_{F_{2}}\to K^{*}_{F_{1}} are monotone. Any nontrivial point preimage is a quotient of KFK_{F} for some subtree FF of F2F1F_{2}-F_{1}, which must be compact and connected. Thus limΘ\lim\Theta is locally connected. ∎

11. Putting together the pieces

In this section, we complete the proof of Theorem 1.1. Suppose GG is a one-ended group acting geometrically on a CAT(0)\operatorname{CAT}(0) space XX with isolated flats, and assume that the maximal peripheral splitting 𝒢\mathcal{G} of GG is locally finite. The results of this section lead up to Proposition 11.11, which states that X\partial X is homeomorphic to the limit of a tree system of spaces, whose underlying tree is the Bass–Serre tree TT for the splitting 𝒢\mathcal{G} and whose component spaces are the boundaries of the component vertex groups. The proof of Theorem 1.1 will follow by combining Proposition 11.11 with ingredients established in the previous sections.

Let XX be the given CAT(0)\operatorname{CAT}(0) space with isolated flats on which GG acts geometrically. To simplify some of the geometric arguments, we will replace XX with a quasi-isometric space XTX_{T} obtained by Bridson–Haefliger’s Equivariant Gluing construction for graphs of groups with CAT(0)\operatorname{CAT}(0) vertex groups and convex edge groups (see Theorems II.11.18 and II.11.21 of [BH99]).

The space XTX_{T} is constructed as follows. Recall that the vertices of TT have two types: component vertices and peripheral vertices. We denote the set of component vertices by 𝒞\mathcal{C} and the set of peripheral vertices by 𝒫\mathcal{P}. Each component vertex group GvG_{v} acts geometrically on a CAT(0)\operatorname{CAT}(0) space CvC_{v}. Each peripheral vertex group PwP_{w}, being virtually abelian, acts geometrically on a flat Euclidean space FwF_{w}.

Each edge ee in TT is incident to a unique peripheral vertex w𝒫w\in\mathcal{P}. Let FeF_{e} be equal to the flat FwF_{w}. Each edge group PeP_{e} also comes with a monomorphism ϕe:PeGv\phi_{e}\colon P_{e}\hookrightarrow G_{v}. By the Flat Torus Theorem there is a ϕe\phi_{e}–equivariant isometric embedding of the flat FeF_{e} in the space CvC_{v}.

The space XTX_{T} is obtained by gluing all components CvC_{v} and flats FwF_{w} using edge spaces of the form Fe×[0,1]F_{e}\times[0,1] in the pattern of the tree TT. For each edge ee incident to vertices v𝒞v\in\mathcal{C} and w𝒫(T)w\in\mathcal{P}(T), the map from FeF_{e} to FwF_{w} is the identity while the map from FeF_{e} to CvC_{v} is the map given by the Flat Torus Theorem. Our setup is now a special case of the Equivariant Gluing discussed in [BH99, Chapter II.11]. Therefore GG acts geometrically on XTX_{T}.

From this point on, we work in the space XTX_{T} instead of XX. Since CAT(0)\operatorname{CAT}(0) groups with isolated flats have “well-defined boundaries,” the spaces XX and XTX_{T} have GG–equivariantly homeomorphic boundaries (see Theorem 5.5).

Suppose x0XTx_{0}\in X_{T} is a basepoint in a component space C0C_{0}. Let v0𝒞(T)v_{0}\in\mathcal{C}(T) be the vertex corresponding to C0C_{0}. Let ξXT\xi\in\partial X_{T}. Using the terminology in [CK00], we can assign an itinerary to ξ\xi at x0x_{0}. This consists of a sequence of edges {ei}\{e_{i}\} of TT corresponding to the sequence of edge spaces Fi×[0,1]F_{i}\times[0,1] that ξ\xi enters when based at x0x_{0}. We say a ray enters an edge space Fi×[0,1]F_{i}\times[0,1] if the ray reaches a point of the interior Fi×(0,1)F_{i}\times(0,1). Observe that ξ\xi has an empty itinerary if and only if ξC0\xi\in\partial C_{0}. The next result is analogous to Lemma 2 in [CK00].

Lemma 11.1.

If ξC0\xi\notin\partial C_{0}, then the itinerary to ξ\xi at x0x_{0} is the sequence of successive edges of a geodesic segment or geodesic ray beginning at v0v_{0} in the tree TT.

Proof.

The separation properties of edge spaces imply that successive edges in the itinerary must be adjacent in TT, so the itinerary defines a path in TT. A geodesic that enters an edge space F×[0,1]F\times[0,1] through the flat F×{0}F\times\{0\} must exit through the flat F×{1}F\times\{1\} without backtracking. Edge spaces are convex so a geodesic cannot revisit any edge space which it has left. Therefore the corresponding path in TT is a geodesic. ∎

Lemma 11.2.

Let x0x_{0} be a basepoint contained in a component Cv0C_{v_{0}}.

  1. (1)

    If a geodesic ray based at x0x_{0} has an infinite itinerary, then that itinerary is a geodesic ray of TT based at v0v_{0}.

  2. (2)

    Every ray of TT based at v0v_{0} is the itinerary of a geodesic ray based at x0x_{0}.

  3. (3)

    If c,cc,c^{\prime} are geodesic rays based at x0x_{0} in XTX_{T} that have the same infinite itinerary at x0x_{0}, then c=cc=c^{\prime}.

Proof.

In order to show (1), it suffices to verify that the infinite itinerary is based at v0v_{0}, which is clear since the first flat the ray enters must be adjacent to the component Cv0C_{v_{0}}.

Any geodesic ray {ei}i=1\{e_{i}\}_{i=1}^{\infty} in TT based at v0v_{0} corresponds to a sequence of edge spaces Fi×[0,1]F_{i}\times[0,1]. Any geodesic segment cic_{i} from x0x_{0} to Fi×{1/2}F_{i}\times\{1/2\} must enter the edge space Fj×[0,1]F_{j}\times[0,1] for all 1ji1\leq j\leq i since that edge space separates x0x_{0} from Fi×{1/2}F_{i}\times\{1/2\}. After passing to a subsequence the geodesics cic_{i} converge to a ray cc based at x0x_{0} which enters all edge spaces Fi×[0,1]F_{i}\times[0,1] for i=1,2,3,i=1,2,3,\dots since the edge spaces separate XTX_{T}. It follows that the given geodesic in TT is the itinerary of cc at x0x_{0}, establishing (2).

Since the itinerary is infinite, there is a sequence of edge spaces Fi×[0,1]F_{i}\times[0,1] that c,cc,c^{\prime} both enter. Let FiF_{i} denote the flat Fi×{1/2}F_{i}\times\{1/2\} in the iith edge space. Using Theorem 5.6, there exists a constant κ\kappa such that for each ii, if qiq_{i} is the closest point of FiF_{i} to x0x_{0} then [x0,qi]Fi[x_{0},q_{i}]\cup F_{i} is κ\kappa–quasiconvex in XX. Thus both cc and cc^{\prime} come within a distance 3κ3\kappa of qiq_{i}. In particular, c,cc,c^{\prime} come 6κ6\kappa–close to each other arbitrarily far from x0x_{0}. By convexity of the distance function in XTX_{T}, this implies c=cc=c^{\prime}, which proves (3). ∎

Lemma 11.3.

If ξXT\xi\in\partial X_{T}, then exactly one of the following holds:

  1. (1)

    ξC\xi\in\partial C for some component CC. This includes the case that ξF\xi\in\partial F for some flat.

  2. (2)

    ξ\xi has an infinite itinerary with respect to any x0XTx_{0}\in X_{T} and is not contained in C\partial C for any component CC in XTX_{T}.

Proof.

Let cc be a geodesic ray based at x0x_{0} representing ξ\xi. Note that ξ\xi has a finite itinerary if and only if cc enters only finitely many edge spaces, which holds if and only if cc eventually remains in a component CvC_{v} for some v𝒞v\in\mathcal{C}. That component is Cv0C_{v_{0}}, if and only if the itinerary of ξ\xi at x0x_{0} is empty as was previously noted. Otherwise, the itinerary of ξ\xi at x0x_{0} is the geodesic segment [v0,v][v_{0},v] in TT, and ξCv\xi\in\partial C_{v}.

The property of having an infinite itinerary does not depend on the choice of basepoint x0XTx_{0}\in X_{T}. Suppose ξ\xi has an infinite itinerary and CvC_{v} is any component. Then ξCv\xi\notin\partial C_{v}. Indeed this becomes obvious if we choose a basepoint x0x_{0} from the convex set CvC_{v}. ∎

Lemma 11.4.

Let CvC_{v} and CvC_{v^{\prime}} be two distinct components of XTX_{T}. Then one of the following holds:

  1. (1)

    CvCv=\partial C_{v}\cap\partial C_{v^{\prime}}=\varnothing.

  2. (2)

    CvCv=Fw\partial C_{v}\cap\partial C_{v^{\prime}}=\partial F_{w} for some w𝒫(T)w\in\mathcal{P}(T) adjacent to both vv and vv^{\prime}. In this case, there is a copy of Fw×[1,1]F_{w}\times[-1,1] embedded in XTX_{T} with Fw×{1}CvF_{w}\times\{-1\}\subset C_{v} and Fw×{1}CvF_{w}\times\{1\}\subset C_{v^{\prime}}.

Proof.

Suppose ξCvCv\xi\in\partial C_{v}\cap\partial C_{v^{\prime}} for some vvv\neq v^{\prime}. Then dT(v,v)=2d_{T}(v,v^{\prime})=2. Indeed if this distance were greater than two then CvC_{v} and CvC_{v^{\prime}} would be separated by a pair of distinct flats FeFeF_{e}\neq F_{e^{\prime}} from the family of isolated flats \mathcal{F}. The existence of a ray asymptotic to both CvC_{v} and CvC_{v}^{\prime} would then contradict isolated flats.

Let w𝒫w\in\mathcal{P} be the unique peripheral vertex adjacent to both vv and vv^{\prime}. Note that the two edge spaces between vv and vv^{\prime} are each isometric to Fw×[0,1]F_{w}\times[0,1], so their union is isometric to Fw×[1,1]F_{w}\times[-1,1]. Observe that this copy of Fw×[1,1]F_{w}\times[-1,1] separates CvC_{v} from CvC_{v^{\prime}} in XTX_{T}. Thus for any constant rr, the intersection Nr(Cv)Nr(Cv)N_{r}(C_{v})\cap N_{r}(C_{v^{\prime}}) lies in a finite tubular neighborhood of FwF_{w}. It follows that ξFw\xi\in\partial F_{w}. ∎

The results above hold for any one-ended CAT(0)\operatorname{CAT}(0) group with isolated flats. For the rest of the section we are concerned only with the special case in which the maximal peripheral splitting is locally finite.

It is our goal to show that, in this case, XT\partial X_{T} is homeomorphic to the limit of the tree system Θ\Theta defined in the following construction.

Construction (The tree system of the peripheral splitting).

Recall that the one-ended group GG has a maximal peripheral splitting 𝒢\mathcal{G}. By hypothesis, this splitting is assumed to be locally finite, which means that each peripheral vertex w𝒫w\in\mathcal{P} of the Bass–Serre tree TT has finite valence. To each vertex v𝒞v\in\mathcal{C} we associate the subspace Kv=CvK_{v}=\partial C_{v}, and to each w𝒫w\in\mathcal{P} we associate the subspace Kw=FwK_{w}=\partial F_{w}. To each oriented edge ee\in\mathcal{E} we associate the subspace Σe=Fe\Sigma_{e}=\partial F_{e}. Since Σe=Σe¯\Sigma_{e}=\Sigma_{\overline{e}}, we set ϕe\phi_{e} to be the identity map. Since ΣeKt(e)\Sigma_{e}\subseteq K_{t(e)} we set iei_{e} to be the inclusion. For each v𝒞v\in\mathcal{C}, the family of closed subspaces {ie(Σe)|t(e)=v}\bigl{\{}\,{i_{e}(\Sigma_{e})}\bigm{|}{t(e)=v}\,\bigr{\}} is pairwise disjoint by the definition of isolated flats. Proposition 11.6 will imply that this family is null. Thus the data above define a tree system Θ\Theta whose underlying tree is TT.

The following proposition follows easily from the conclusions of Lemmas 11.2, 11.3, and 11.4 about itineraries and their relation to the structure of XT\partial X_{T}. As in Section 10, given a tree system Θ\Theta, we let #Θ\#\Theta denote the quotient space obtained by gluing the vertex spaces of Θ\Theta along edge spaces via the maps iei_{e}.

Proposition 11.5.

There is a map ρ:#ΘTXT\rho\colon\#\Theta\cup\partial T\to\partial X_{T} with the following properties:

  1. (1)

    ρ\rho is a bijection.

  2. (2)

    ρ\rho is continuous on #Θ\#\Theta.

  3. (3)

    For each finite subtree FF of TT, the map ρ\rho restricted to the partial union KFK_{F} is a topological embedding. In particular, ρ\rho is an embedding when restricted to any vertex space KvK_{v}. ∎

If SS is a subtree of TT, we write Ψ(S)=ρ(#ΘSS)\Psi(S)=\rho\bigl{(}\#\Theta_{S}\cup\partial S\bigr{)}. If TeT_{e} is a branch of TT, the set Ψ(Te)\Psi(T_{e}) is a branch of XT\partial X_{T}.

Proposition 11.6.

Choose a basepoint v𝒱v\in\mathcal{V}. The family of all branches Ψ(Te)\Psi(T_{e}) of XT\partial X_{T} such that ee points away from vv is a null family.

The proof of the previous proposition uses the following two lemmas.

Lemma 11.7.

Let x0x_{0} be a basepoint contained in a component Cv0C_{v_{0}}.

  1. (1)

    If ξCv\xi\in\partial C_{v} for some v𝒞v\in\mathcal{C} but not in Fw\partial F_{w} for any ww, then the itinerary of ξ\xi at x0x_{0} is the geodesic segment [v0,v][v_{0},v].

  2. (2)

    If ξFw\xi\in\partial F_{w} for some w𝒫w\in\mathcal{P}, then the itinerary of ξ\xi at x0x_{0} is the geodesic segment [v0,v][v_{0},v] of TT, where vv is the vertex adjacent to ww that is closest to v0v_{0}.

Proof.

In case (1), it follows from Lemma 11.4 that CvC_{v} is the unique component whose boundary contains ξ\xi. The proof of Lemma 11.3 implies that ξ\xi eventually remains in CvC_{v}, and the itinerary of ξ\xi at x0x_{0} equals [v0,v][v_{0},v] as desired.

If ξFw\xi\in\partial F_{w} then ξ\xi also lies in Cv\partial C_{v} for all vv adjacent to ww in TT. Furthermore, ξ\xi does not lie in C\partial C^{\prime} for any other component CC^{\prime}. The component that ξ\xi eventually remains in must therefore be adjacent to FwF_{w}. Observe that ξ\xi cannot enter any edge space Fe×[0,1]F_{e}\times[0,1] adjacent (and parallel) to FwF_{w} since then ξ\xi would fail to be asymptotic to FwF_{w}. ∎

Lemma 11.8.

Let Ψ(Te)\Psi(T_{e}) be a branch of XT\partial X_{T} determined by an oriented edge ee\in\mathcal{E}. Let x0XTx_{0}\in X_{T} be a basepoint contained in CvC_{v} for some vertex vTev\notin T_{e}. Let κ\kappa be the constant from Theorem 5.6. If d(x0,Fe)r+3κd(x_{0},F_{e})\geq r+3\kappa, then there exists a geodesic ray cc based at x0x_{0} such that Ψ(Te)U(c,r,7κ)\Psi(T_{e})\subseteq U(c,r,7\kappa).

Proof.

Let qq be the nearest point in FeF_{e} to x0x_{0}. Then d(x0,q)r+3κd(x_{0},q)\geq r+3\kappa. Let c,cc,c^{\prime} be geodesic rays based at x0x_{0} that both intersect FeF_{e}. By Theorem 5.6 the set [x0,q]Fe[x_{0},q]\cup F_{e} is κ\kappa–quasiconvex. Thus there exist s,ss,s^{\prime} with c(s),c(s)c(s),c^{\prime}(s^{\prime}) both contained in 𝒩κ([x0,q])𝒩κ(Fe)\mathcal{N}_{\kappa}\bigl{(}{[x_{0},q]}\bigr{)}\cap\mathcal{N}_{\kappa}({F_{e}}). It follows that d(c(s),q)d\bigl{(}c(s),q\bigr{)} and d(c(s),q)d\bigl{(}c^{\prime}(s^{\prime}),q\bigr{)} are each less than 3κ3\kappa. Thus d(c(s),c(s))<6κd\bigl{(}c(s),c^{\prime}(s^{\prime})\bigr{)}<6\kappa. By the Law of Cosines d(c(r),c(r))<6κd\bigl{(}c(r),c^{\prime}(r)\bigr{)}<6\kappa. In particular cU(c,r,6κ)c^{\prime}\in U(c,r,6\kappa).

If cc intersects FeF_{e} and cc^{\prime} is asymptotic to FeF_{e} (but does not intersect FeF_{e}), then cc^{\prime} is a limit of geodesics that intersect FeF_{e}. In this case, we conclude that cU(c,r,7κ)c^{\prime}\in U(c,r,7\kappa).

By Lemma 11.7 each ray based at x0x_{0} and asymptotic to Ψ(Te)\Psi(T_{e}) has an itinerary involving FeF_{e}—and hence intersects FeF_{e}—unless the ray is asymptotic to FeF_{e} itself. In all cases we see that Ψ(Te)U(c,r,7κ)\Psi(T_{e})\subseteq U(c,r,7\kappa) for any cc crossing FeF_{e}. ∎

Proof of Proposition 11.6.

Let v𝒞v\in\mathcal{C}. Choose a basepoint x0Cvx_{0}\in C_{v}. Let D=7κD=7\kappa where κ\kappa is as in the previous lemma and let r<r<\infty. Let Ψ(Te)\Psi(T_{e}) be a branch such that ee points away from vv and such that Ψ(Te)\Psi(T_{e}) is not contained in any set of the form U(,r,D)U(\cdot,r,D). Applying Lemma 11.8 in the contrapositive implies d(x0,Fe)<r+3κd(x_{0},F_{e})<r+3\kappa. Since the collection of flats {Fe}\{F_{e}\} in XTX_{T} is locally finite, there are only finitely many edges ee whose corresponding flat FeF_{e} is that close to x0x_{0}. Thus there are only finitely many possibilities for the branch Ψ(Te)\Psi(T_{e}). ∎

Proposition 11.9.

Branches Ψ(Te)\Psi(T_{e}) are closed in XT\partial X_{T}.

Proof.

Let x0XTx_{0}\in X_{T} be a basepoint contained in CvC_{v} for some vertex vTev\notin T_{e}. Suppose {ci}\{c_{i}\} is a sequence of geodesic rays in XTX_{T} based at x0x_{0} asymptotic to Ψ(Te)\Psi(T_{e}) such that cic_{i} converges to the geodesic ray cc based at x0x_{0}. As in the proof of Lemma 11.8, each such ray intersects FeF_{e} or is asymptotic to FeF_{e}. By passing to a subsequence we can assume all cic_{i} intersect FeF_{e} or all cic_{i} are asymptotic to FeF_{e}.

If each cic_{i} is asymptotic to FeF_{e}, then cc is as well, since FeF_{e} is a closed convex subspace of XTX_{T}. On the other hand, if each cic_{i} intersects FeF_{e} then cc either intersects FeF_{e} or is asymptotic to FeF_{e} depending on whether the intersections of the cic_{i} with FeF_{e} remain bounded as ii\to\infty. In all cases we conclude that cc is asymptotic to the branch Ψ(Te)\Psi(T_{e}). ∎

Recall that limΘ\lim\Theta is the inverse limit of an inverse system of spaces KF=KF/𝒜FK_{F}^{*}=K_{F}/\mathcal{A}_{F} for all finite subtrees FF of TT (see Definition 10.4). The collection 𝒜F\mathcal{A}_{F} contains an edge space ie(Σe)i_{e}(\Sigma_{e}) for each edge eNFe\in N_{F}, where NFN_{F} contains all edges whose origin is outside FF and whose terminus is in FF.

Lemma 11.10.

For each finite subtree FF of TT, let XT/𝒟F\partial X_{T}/\mathcal{D}_{F} be the quotient of XT\partial X_{T} formed by collapsing each branch Ψ(Te)\Psi(T_{e}) to a point whenever e¯NF\overline{e}\in N_{F}. Let qF:XX/𝒟Fq_{F}\colon\partial X\to\partial X/\mathcal{D}_{F} denote the natural quotient map. Then XT/𝒟F\partial X_{T}/\mathcal{D}_{F} is homeomorphic to the quotient KF=KF/𝒜FK^{*}_{F}=K_{F}/\mathcal{A}_{F}.

Proof.

The embedding ρ:KFXT\rho\colon K_{F}\to\partial X_{T} induces a continuous map gF:KFXT/𝒟Fg_{F}\colon K_{F}^{*}\to\partial X_{T}/\mathcal{D}_{F}, which is clearly a bijection. It suffices to verify that gFg_{F} is a closed map. We will show that the decomposition 𝒟F\mathcal{D}_{F} is upper semicontinuous. It then follows immediately from Proposition 8.7 that XT/𝒟F\partial X_{T}/\mathcal{D}_{F} is Hausdorff, and since KFK_{F}^{*} is compact, we can conclude that gFg_{F} is closed.

By Proposition 11.5(1), two branches Ψ(Te)\Psi(T_{e}) and Ψ(Te)\Psi(T_{e^{\prime}}) with e¯,e¯NF\overline{e},\overline{e}^{\prime}\in N_{F} intersect precisely when their origin vertices o(e)o(e) and o(e)o(e^{\prime}) are peripheral vertices (lying in 𝒫\mathcal{P}) and are equal. In this case the intersection of the branches is the peripheral vertex space Σe=Ko(e)=Σe\Sigma_{e}=K_{o(e)}=\Sigma_{e^{\prime}}. Therefore each nontrivial member of the decomposition 𝒟F\mathcal{D}_{F} is the union of finitely many branches whose defining edges have a common origin.

By Proposition 11.6 the branches being collapsed are a null family. Therefore 𝒟F\mathcal{D}_{F} is also null. Similarly by Proposition 11.9 we see that the members of 𝒟F\mathcal{D}_{F} are closed. Therefore 𝒟F\mathcal{D}_{F} is upper semicontinuous by Proposition 8.11. ∎

Proposition 11.11.

Suppose GG is a one-ended group acting geometrically on a CAT(0)\operatorname{CAT}(0) space XX with isolated flats. Suppose the maximal peripheral splitting of GG is locally finite. Then the boundary X\partial X is homeomorphic to the limit limΘ\lim\Theta of the associated tree system.

Proof.

By Theorem 5.5, the boundary X\partial X is GG–equivariantly homeomorphic to the boundary XT\partial X_{T} constructed above. In order to prove that XT\partial X_{T} is homeomorphic to limΘ\lim\Theta we will define maps hFh_{F} from XT\partial X_{T} onto each KFK_{F}^{*} in the inverse system, and show that the induced map h:XTlimΘh\colon\partial X_{T}\to\lim\Theta is a homeomorphism.

The map hFh_{F} is the composition gF1qFg_{F}^{-1}\circ q_{F}, where gFg_{F} is the homeomorphism defined in Lemma 11.10 and qFq_{F} is the natural quotient map XTXT/𝒟F\partial X_{T}\to\partial X_{T}/\mathcal{D}_{F}. Since fF1F2hF2=hF1f_{F_{1}F_{2}}\circ h_{F_{2}}=h_{F_{1}} whenever F1F2F_{1}\subseteq F_{2}, the maps hFh_{F} induce a continuous map h:XTlimΘh\colon\partial X_{T}\to\lim\Theta.

Observe that XT\partial X_{T} is compact, each quotient space KFK_{F}^{*} is Hausdorff, and hFh_{F} is surjective for each finite subtree FF. It follows that hh is surjective (see for instance §I.9.6, Corollary 2(b) of [Bou71]). Since limΘ\lim\Theta is Hausdorff, the map hh is closed.

Thus we only need to show that hh is injective. Suppose ξη\xi\neq\eta are two distinct points of XT\partial X_{T}. Recall that by Proposition 11.5(1) each point of XT\partial X_{T} is either contained in a block boundary KvK_{v} for some v𝒞v\in\mathcal{C} or is equal to an ideal point ρ(z)\rho(z) for some zTz\in\partial T. We also know that the map ρ:TXT\rho\colon\partial T\to\partial X_{T} from ends of the tree to ideal points is injective.

Case 1: Suppose ξ\xi and η\eta are both contained in block boundaries. Recall that each point of XT\partial X_{T} lies in at most finitely many block boundaries. Choose a finite subtree FF that contains all block boundaries KvK_{v} that contain either ξ\xi or η\eta. Then ξ\xi and η\eta are contained in the partial union KFK_{F}, but neither ξ\xi nor η\eta is contained in any subspace ie(Σe)i_{e}(\Sigma_{e}) with eNF{e}\in N_{F}. It follows that ξ\xi and η\eta have distinct images hF(ξ)h_{F}(\xi) and hF(η)h_{F}(\eta) in the quotient KFK_{F}^{*}.

Case 2: Suppose ξ\xi and η\eta are equal to distinct ideal points ρ(z)\rho(z) and ρ(z)\rho(z^{\prime}) with zzz\neq z^{\prime} in T\partial T. Let cc be the geodesic in the tree TT from zz to zz^{\prime}. Choose any vertex v𝒞v\in\mathcal{C} of cc, and let FF be the finite subtree {v}\{v\}. Then KF=KvK_{F}=K_{v} consists of only one block boundary. Since cc is a geodesic, there are distinct edges eee\neq e^{\prime} with o(e)=o(e)=vo(e)=o(e^{\prime})=v such that ξ\xi lies in the branch Ψ(Te)\Psi(T_{e}) and η\eta lies in the branch Ψ(Te)\Psi(T_{e^{\prime}}). Therefore ξ\xi and η\eta have distinct images hF(ξ)h_{F}(\xi) and hF(η)h_{F}(\eta) in the quotient KFK_{F}^{*}.

Case 3: Suppose ξ\xi is contained in a block boundary and η\eta is an ideal point ρ(z)\rho(z) for zTz\in\partial T. Choose a finite subtree FF containing all of the finitely many blocks KvK_{v} that contain ξ\xi. Then ξ\xi lies in the partial union KFK_{F} but is not contained in any branch Ψ(Te)\Psi(T_{e}) with e¯NF\overline{e}\in N_{F}. On the other hand, η\eta is not contained in KFK_{F}, so it must lie in some branch Ψ(Te)\Psi(T_{e}) with e¯NF\overline{e}\in N_{F}. It follows that ξ\xi and η\eta have distinct images qF(ξ)q_{F}(\xi) and qF(η)q_{F}(\eta) in the quotient XT/𝒟F\partial X_{T}/\mathcal{D}_{F}. Consequently their images in KFK_{F}^{*} are distinct as well. ∎

We now use Proposition 11.11 to complete the proof of Theorem 1.1.

Proof of Theorem 1.1.

The reverse implication follows from work of Mihalik–Ruane [MR99, MR01].

We prove the forward implication using the contrapositive; i.e., if GG does not split over a virtually abelian subgroup as in the statement of Theorem 1.1, we must show that X\partial X is locally connected.

Let 𝒢\mathcal{G} be the maximal peripheral splitting of GG given by Theorem 6.3. By hypothesis, this splitting is locally finite. Proposition 11.11 implies that X\partial X is homeomorphic to the limit of the associated tree system. The trivial case in which GG is virtually abelian of higher rank is obvious since the boundary is a sphere in that case. In all other cases the tree system is nondegenerate. By Corollary 7.8 each component vertex group is an atomic CAT(0)\operatorname{CAT}(0) group with isolated flats. By Theorem 1.4 each component vertex space is connected and locally connected. Each peripheral vertex space is a sphere of dimension at least one, hence is connected. Therefore we may apply Theorem 10.6 to conclude that X\partial X is locally connected, as desired. ∎

References

  • [AS85] F.D. Ancel and L. Siebenmann. The construction of homogeneous homology manifolds. Abstracts Amer. Math. Soc., 6, 1985. Abstract number 816-57-72.
  • [Bal95] Werner Ballmann. Lectures on spaces of nonpositive curvature, volume 25 of DMV Seminar. Birkhäuser Verlag, Basel, 1995.
  • [Ben92] N. Benakli. Polyèdres hyperboliques: passage du local au global. PhD thesis, Université Paris–Sud, 1992.
  • [Bes96] Mladen Bestvina. Local homology properties of boundaries of groups. Michigan Math. J., 43(1):123–139, 1996.
  • [BH99] Martin R. Bridson and André Haefliger. Metric spaces of non-positive curvature. Springer-Verlag, Berlin, 1999.
  • [BM91] Mladen Bestvina and Geoffrey Mess. The boundary of negatively curved groups. J. Amer. Math. Soc., 4(3):469–481, 1991.
  • [Bou71] N. Bourbaki. Éléments de mathématique. Topologie générale. Chapitres 1 à 4. Hermann, Paris, 1971.
  • [Bou97] M. Bourdon. Immeubles hyperboliques, dimension conforme et rigidité de Mostow. Geom. Funct. Anal., 7(2):245–268, 1997.
  • [Bow98] Brian H. Bowditch. Cut points and canonical splittings of hyperbolic groups. Acta Math., 180(2):145–186, 1998.
  • [Bow99] B.H. Bowditch. Treelike structures arising from continua and convergence groups. Mem. Amer. Math. Soc., 139(662):1–86, 1999.
  • [Bow01] B.H. Bowditch. Peripheral splittings of groups. Trans. Amer. Math. Soc., 353(10):4057–4082, 2001.
  • [Bow12] B.H. Bowditch. Relatively hyperbolic groups. Internat. J. Algebra Comput., 22(3):1250016, 66 pages, 2012.
  • [BZ] Michael Ben-Zvi. Path connectedness of CAT(0)\text{CAT}(0) groups with the isolated flats property. In preparation.
  • [Cap54] C. E. Capel. Inverse limit spaces. Duke Math. J., 21:233–245, 1954.
  • [Cha95] Christophe Champetier. Propriétés statistiques des groupes de présentation finie. Adv. Math., 116(2):197–262, 1995.
  • [CK00] Christopher B. Croke and Bruce Kleiner. Spaces with nonpositive curvature and their ideal boundaries. Topology, 39(3):549–556, 2000.
  • [Dav86] Robert J. Daverman. Decompositions of manifolds, volume 124 of Pure and Applied Mathematics. Academic Press, Inc., Orlando, FL, 1986.
  • [DGP11] François Dahmani, Vincent Guirardel, and Piotr Przytycki. Random groups do not split. Math. Ann., 349(3):657–673, 2011.
  • [DHW] Pallavi Dani, Matthew Haulmark, and Genevieve Walsh. Right-angled Coxeter groups with non-planar boundary. Preprint. arXiv:1902.01029 [math.GT].
  • [GS] Ross Geoghegan and Eric Swenson. On semistability of CAT(0)\operatorname{CAT}(0) groups. Preprint. arXiv:1707.07061.
  • [Gui14] Craig R. Guilbault. Weak Z-structures for some classes of groups. Algebr. Geom. Topol., 14(2):1123–1152, 2014.
  • [Han51] Olof Hanner. Some theorems on absolute neighborhood retracts. Ark. Mat., 1:389–408, 1951.
  • [Hau18] Matthew Haulmark. Boundary classification and two-ended splittings of groups with isolated flats. J. Topol., 11(3):645–665, 2018.
  • [HHS] Matthew Haulmark, G. Christopher Hruska, and Bakul Sathaye. Nonhyperbolic Coxeter groups with Menger curve boundary. To appear in Enseign. Math.. arXiv:1812.04649 [math.GR].
  • [HK05] G. Christopher Hruska and B. Kleiner. Hadamard spaces with isolated flats. Geom. Topol., 9:1501–1538, 2005. With an appendix jointly written by the authors and M. Hindawi.
  • [HK09] G. Christopher Hruska and Bruce Kleiner. Erratum to: “Hadamard spaces with isolated flats”. Geom. Topol., 13(2):699–707, 2009.
  • [HR] G. Christopher Hruska and Kim Ruane. Hierarchies and semistability of relatively hyperbolic groups. Preprint. arXiv:1904.12947 [math.GR].
  • [Hru04] G. Christopher Hruska. Nonpositively curved 2–complexes with isolated flats. Geom. Topol., 8:205–275, 2004.
  • [Hu65] Sze-Tsen Hu. Theory of retracts. Wayne State University Press, Detroit, 1965.
  • [Jak80] W. Jakobsche. The Bing–Borsuk conjecture is stronger than the Poincaré conjecture. Fund. Math., 106(2):127–134, 1980.
  • [Jak91] W. Jakobsche. Homogeneous cohomology manifolds which are inverse limits. Fund. Math., 137(2):81–95, 1991.
  • [KK00] Michael Kapovich and Bruce Kleiner. Hyperbolic groups with low-dimensional boundary. Ann. Sci. École Norm. Sup. (4), 33(5):647–669, 2000.
  • [KL95] M. Kapovich and B. Leeb. On asymptotic cones and quasi-isometry classes of fundamental groups of 33–manifolds. Geom. Funct. Anal., 5(3):582–603, 1995.
  • [Lev98] Gilbert Levitt. Non-nesting actions on real trees. Bull. London Math. Soc., 30(1):46–54, 1998.
  • [LT17] Larsen Louder and Nicholas Touikan. Strong accessibility for finitely presented groups. Geom. Topol., 21(3):1805–1835, 2017.
  • [Mih83] Michael L. Mihalik. Semistability at the end of a group extension. Trans. Amer. Math. Soc., 277(1):307–321, 1983.
  • [Mor16] Molly A. Moran. Metrics on visual boundaries of CAT(0)\operatorname{CAT}(0) spaces. Geom. Dedicata, 183:123–142, 2016.
  • [MR99] Michael Mihalik and Kim Ruane. CAT(0)\operatorname{CAT}(0) groups with non–locally connected boundary. J. London Math. Soc. (2), 60(3):757–770, 1999.
  • [MR01] Michael Mihalik and Kim Ruane. CAT(0)\operatorname{CAT}(0) HNN-extensions with non–locally connected boundary. Topology Appl., 110(1):83–98, 2001. Geometric topology and geometric group theory (Milwaukee, WI, 1997).
  • [MS] M. Mihalik and E. Swenson. Relatively hyperbolic groups have semistable fundamental group at infinity. Preprint, arXiv:1709.02420 [math.GR].
  • [Ont05] Pedro Ontaneda. Cocompact CAT(0)\operatorname{CAT}(0) spaces are almost geodesically complete. Topology, 44:47–62, 2005.
  • [OS15] Damian Osajda and Jacek Świątkowski. On asymptotically hereditarily aspherical groups. Proc. Lond. Math. Soc. (3), 111(1):93–126, 2015.
  • [Pon30] L.S. Pontryagin. Sur une hypothèse fundamentale de la théorie de la dimension. C.R. Acad. Sci. Paris, 190:1105–1107, 1930.
  • [Rua05] Kim Ruane. CAT(0) boundaries of truncated hyperbolic space. Topology Proc., 29(1):317–331, 2005.
  • [Ser77] Jean-Pierre Serre. Arbres, amalgames, SL2{\rm SL}_{2}, volume 46 of Astérisque. Société Mathématique de France, Paris, 1977. Written in collaboration with Hyman Bass.
  • [Swa96] G.A. Swarup. On the cut point conjecture. Electron. Res. Announc. Amer. Math. Soc., 2(2):98–100, 1996.
  • [Świ] J. Świątkowski. Trees of metric compacta and trees of manifolds. Preprint. arXiv:1304.5064.
  • [Tra13] Hung Cong Tran. Relations between various boundaries of relatively hyperbolic groups. Internat. J. Algebra Comput., 23(7):1551–1572, 2013.
  • [Wil70] Stephen Willard. General topology. Addison-Wesley Publishing Co., Reading, Mass.-London-Don Mills, Ont., 1970.
  • [Yam04] Aslı Yaman. A topological characterisation of relatively hyperbolic groups. J. Reine Angew. Math., 566:41–89, 2004.