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

Coarse Kernels of Group Actions

Tejas Mittal
Abstract.

In this paper, we study the coarse kernel of a group action, namely the normal subgroup of elements that translate every point by a uniformly bounded amount. We give a complete algebraic characterization of this object. We specialize to CAT(0)\mathrm{CAT}(0) spaces and show that the coarse kernel must be virtually abelian, characterizing when it is finite or cyclic in terms of the curtain model. As an application, we characterize the relation between the coarse kernels of the action on a CAT(0)\mathrm{CAT}(0) space and the induced action on its curtain model. Along the way, we study weakly acylindrical actions on quasi-lines.

1. Introduction

The goal of this paper is to study certain stabilizers of group actions. Within the framework of groups acting on metric spaces, one is often interested in the properties that depend solely on the quasi-isometry type of the action. In other words, the object of interest must remain invariant under GG-equivariant quasi-isometries. Consequently, the kernels of actions are not ideal objects of investigation. Instead, it is more appropriate to consider their coarse versions, which possess this stability feature. In the current work, we study the coarse kernel of a group action, namely the subgroup consisting of elements that translate every point by a uniformly bounded amount.

Definition 1.1.

Let GG be a group acting on a metric space XX. Define the coarse kernel of the action to be GX:={gGC>0 such that d(gx,x)C xX}.{G_{X}:=\{g\in G\mid\exists C>0\text{ such that }d(g\cdot x,x)\leq C\text{ }\forall x\in X\}.}

This coarse kernel then serves as a suitable analog to the kernel of a group action. We remark that our coarse kernels are different from the ones of [LV23], where the focus is on the analogue of kernels for coarse groups. Our first result is to show that, quite surprisingly, for a geometric action, the coarse kernel GXG_{X} can be described in purely algebraic terms. In the theorem below, Z(H)Z(H) denotes the center of the group HH.

Theorem 1.2.

Let GG be a group acting properly discontinuously and co-boundedly on a metric space XX. Then, GX={Z(H)|G/H|<}G_{X}=\bigcup\{Z(H)\mid|G/H|<\infty\}.

In particular, if GG is abelian, then GX=GG_{X}=G, which can be shown to hold for all co-bounded isometric actions. Similarly, in the presence of negative curvature, for instance, when a hyperbolic group acts on its Cayley graph, one expects the coarse kernel to be finite. We focus on CAT(0)\mathrm{CAT}(0) spaces, which interpolate between the two cases above, and show the following.

Theorem 1.3.

Let XX be a co-bounded CAT(0)\mathrm{CAT}(0) space, and let GG act on XX properly discontinuously. Suppose that GG has an unbounded orbit in the curtain model XD.X_{D}. Then, GXG_{X} is virtually cyclic. Moreover, if an orbit of GG in XDX_{D} is not a quasi-line either, then GXG_{X} is finite.

The curtain model was introduced by Petyt, Spriano, and Zalloum in [PSZ24], and is a hyperbolic space XDX_{D} associated to a CAT(0)\mathrm{CAT}(0) space XX that \sayencodes the hyperbolicity of XX. This allows us to make precise the intuition that a group acting on a CAT(0)\mathrm{CAT}(0) space that is hyperbolic enough needs to have a small coarse kernel. We remark that GG does not need to act co-compactly on XX. When we add that assumption, we obtain a significantly stronger characterization of the coarse kernel. Specifically, we are able to identify the algebraic structure of the coarse kernel GXG_{X} based on certain geometric constrains on XX.

Theorem 1.4.

Let XX be a CAT(0)\mathrm{CAT}(0) space, and let GG act on XX geometrically. Then,
(i) GXG_{X} is virtually n\mathbb{Z}^{n} for some n0n\in\mathbb{Z}_{\geq 0}.
(ii) If moreover XDX_{D} is unbounded, then GXG_{X} is virtually cyclic.
(iii) If furthermore XDX_{D} is not a quasi-line, then GXG_{X} is the largest finite normal subgroup of GG.

After the first draft of this paper was completed, it was pointed out to us that item (iii) also follows from [BJ24, Corollary 1.6]. However, our proof and their proof uses different techniques.

Our second line of investigation is concerned more directly on the curtain model, and follows on the program started in [PSZ24] of better understanding the group GG via its action on the curtain model. A natural question to ask is how much the action of GG on XDX_{D} might forget information. We show that, from the perspective of coarse kernels, the action of GG on XDX_{D} retains as much information as the action of GG on XX.

Theorem 1.5.

Let XX be a proper CAT(0)\mathrm{CAT}(0) space with XDX_{D} being unbounded, and let GG act on XX co-compactly. Then, GX=GXDG_{X}=G_{X_{D}}.

Further, for a group acting geometrically the result above gives a complete characterization of when GXG_{X} and GXDG_{X_{D}} coincide.

Corollary 1.6.

Let GG be a group acting geometrically on a CAT(0)\mathrm{CAT}(0) space XX. Then, exactly one of the following holds.

  1. (1)

    XDX_{D} is unbounded and hence GX=GXDG_{X}=G_{X_{D}};

  2. (2)

    XDX_{D} is bounded, |G:Z(G)|<|G:Z(G)|<\infty, and GX=G=GXDG_{X}=G=G_{X_{D}};

  3. (3)

    XDX_{D} is bounded, Z(G)Z(G) has infinite index in GG and GXGXDG_{X}\neq G_{X_{D}}.

1.1. Weak acylindricity

In order to prove Theorem 1.4, we will need to study weakly acylindrical actions on a quasi-line which is a topic of independent interest. We recall the definition of weak acylindrical action.

Definition 1.7.

Let GG be a group acting on a metric space XX. The action of GG is said to be weakly acylindrical if for each ϵ>0\epsilon>0 there exists RR such that for any x,yXx,y\in X with d(x,y)>Rd(x,y)>R, only finitely many gGg\in G satisfy max{d(x,gx),d(y,gy)}<ϵmax\{d(x,g\cdot x),d(y,g\cdot y)\}<\epsilon.

In this setting, we extend a well-known result for acylindrical actions on a quasi-line.

Theorem 1.8.

Let XX be a roughly geodesic hyperbolic space. Let GG act on XX weakly acylindrically with a quasi-line orbit. Then GG is virtually \mathbb{Z}.

As a corollary, we establish the following result, which can be viewed as a counterpart to the ping-pong lemmas for CAT(0)\mathrm{CAT}(0) spaces, thereby helping us better understand the curtain model towards the Tits Alternative.

Theorem 1.9.

Let GG be a group acting geometrically on a CAT(0)\mathrm{CAT}(0) space XX. Let HH be any subgroup of GG. If |XDH|=2|\partial_{X_{D}}H|=2, then HH is virtually \mathbb{Z}.

1.2. Coarse stabilizers of sets

We, in fact, develop the proof of Theorem 1.2 in the broader context of coarse stabilizers of subsets YXY\subseteq X, which generalise the coarse kernel GXG_{X}.

Definition 1.10.

Let GG be a group acting on a metric space XX and let YXY\subseteq X. Define the coarse stabilizer of Y to be GY:={gGC such that d(gy,y)C yY}G_{Y}:=\{g\in G\mid\exists C\text{ such that }d(g\cdot y,y)\leq C\text{ }\forall y\in Y\}.

The coarse stabilizers are a very well-studied object in geometric group theory. Typically, the set YY is unbounded; otherwise GY=GG_{Y}=G. In this sense, one can see that by varying YY, one interpolates between GG and the coarse kernel GXG_{X}. We now give a complete algebraic characterization of the coarse stabilizer GYG_{Y} when YY coarsely coincides with an orbit of a finitely generated subgroup HH of GG. In particular, this shows that when GG acts geometrically on XX, we recover Theorem 1.2. In the theorem below, CG(H)C_{G}(H) denotes the centralizer of HH in GG.

Theorem 1.11.

Let XX be a metric space, and suppose GG acts on XX properly discontinuously and co-boundedly. Let HGH\leq G be finitely generated, and let YXY\subseteq X coarsely coincide with an orbit of HH. Then, GY={CG(H)|H/H|<}G_{Y}=\bigcup\{C_{G}(H^{\prime})\mid|H/H^{\prime}|<\infty\}.

Acknowledgement

This work was done as part of an undergraduate research project under the supervision of Davide Spriano. Davide suggested the initial problems which evolved into this paper, and his insights helped me resolve many queries. His dedicated time also helped me acquire the necessary background knowledge and assisted in refining several of the proofs. I am extremely grateful to Davide for taking the time to supervise this project and would like to thank him for all his guidance and support. I would also like to thank Harry Petyt for his kind help, which enhanced my understanding of the topic, and for his helpful input on an earlier draft of the paper.

2. Background

In this section, we present definitions and established results necessary to prove our main theorems later. The references for this material are [PSZ24] and [CCMT15].

2.1. Curtain Model

We briefly discuss the curtain model XDX_{D} associated with any CAT(0)\mathrm{CAT}(0) space XX, as defined in [PSZ24]. Essentially, the curtain model XD:=(X,D)X_{D}:=(X,D) is the set XX equipped with a new metric D:X×XD:X\times X\rightarrow\mathbb{R}. We aim to choose the metric DD so that XDX_{D} becomes a hyperbolic space.
The main ingredient in defining XDX_{D} is the notions of curtains and that of separation.

Definition 2.1 (Definition 2.1 in [PSZ24]).

Let XX be a CAT(0)\mathrm{CAT}(0) space, and let α:IX\alpha:I\rightarrow X be a geodesic. For a number rr with [r1/2,r+1/2][r-1/2,r+1/2] in the interior of II, the curtain dual to α\alpha at rr is

hα,r=πα1(α[r1/2,r+1/2]).h_{\alpha,r}=\pi_{\alpha}^{-1}(\alpha[r-1/2,r+1/2]).

Here πα\pi_{\alpha} denotes the closest-point projection onto α\alpha.

Definition 2.2 (Definition 2.2 in [PSZ24]).

Let XX be a CAT(0)\mathrm{CAT}(0) space, and let h=hα,rh=h_{\alpha,r} be a curtain. The halfspaces determined by hh are h=πα1α(I(,r1/2))h^{-}=\pi_{\alpha}^{-1}\alpha(I\cup(-\infty,r-1/2)) and h+=πα1α(I(r+1/2,))h^{+}=\pi_{\alpha}^{-1}\alpha(I\cup(r+1/2,\infty)). Note that {h,h,h+}\{h^{-},h,h^{+}\} is a partition of XX. If AA and BB are subsets of XX such that AhA\subseteq h^{-} and Bh+B\subseteq h^{+}, then we say that hh separates AA from BB.

Next, we define chains and the chain distance.

Definition 2.3 (Definition 2.9 in [PSZ24]).

A set {hi}\{h_{i}\} of curtains is a chain if hih_{i} separates hi1h_{i-1} from hi+1h_{i+1} for all ii. We say that {hi}\{h_{i}\} separates A,BXA,B\subseteq X if every hih_{i} does. The chain distance from xx to yxy\neq x is d(x,y)=1+max{|c|:c is a chain separating x, y}d_{\infty}(x,y)=1+max\{|c|:\text{c is a chain separating x, y}\}.

This allows us to define the notion of an LL-chain.

Definition 2.4 (Definition 2.11 in [PSZ24]).

Let LL\in\mathbb{N}. Disjoint curtains hh and hh^{\prime} are said to be LL-separated if every chain meeting both hh and hh^{\prime} has cardinality at most LL. If cc is a chain of curtains such that each pair is LL-separated, then we refer to cc as an LL-chain.

Now, we use the notion of LL-chains to define a family of metric spaces XL:=(X,dL)X_{L}:=(X,d_{L}) corresponding to LL\in\mathbb{N}.

Definition 2.5 (Definition 2.15 in [PSZ24]).

Given distinct points x,yXx,y\in X, set dL(x,x)=0d_{L}(x,x)=0 and define dL(x,y)=1+max{|c|:c is an L–chain separating x from y}d_{L}(x,y)=1+max\{|c|:\text{c is an L–chain separating x from y}\}.

Remark.

One can check that dL:X×Xd_{L}:X\times X\rightarrow\mathbb{R} is indeed a metric on XX and that dL(x,y)d(x,y)1+d(x,y)d_{L}(x,y)\leq d_{\infty}(x,y)\leq 1+d(x,y) for all LL\in\mathbb{N}.

Finally, we define the curtain model XDX_{D} by combining all the LL-metrics dLd_{L} on XX.

Definition 2.6.

Fix a sequence λL(0,1)\lambda_{L}\in(0,1) such that

n=1λL<n=1LλL<n=1L2λL=Λ<.\sum_{n=1}^{\infty}\lambda_{L}<\sum_{n=1}^{\infty}L\lambda_{L}<\sum_{n=1}^{\infty}L^{2}\lambda_{L}=\Lambda<\infty.

Then, XD:=(X,D)X_{D}:=(X,D) where D(x,y):=L=1λLdL(x,y)D(x,y):=\sum_{L=1}^{\infty}\lambda_{L}d_{L}(x,y) for any x,yXx,y\in X.

We record the following basic fact, which tells us how distances in the curtain model XDX_{D} are bounded in terms of the distances in XX.

Lemma 2.7.

D(x,y)Λ(1+d(x,y))D(x,y)\leq\Lambda(1+d(x,y)) for all x,yXx,y\in X.

Proof.

D(x,y):=L=1λLdL(x,y)L=1λL(1+d(x,y))<Λ(1+d(x,y))D(x,y):=\sum_{L=1}^{\infty}\lambda_{L}d_{L}(x,y)\leq\sum_{L=1}^{\infty}\lambda_{L}(1+d(x,y))<\Lambda(1+d(x,y)). ∎

Now, we state some properties of the curtain model.

Firstly, we note that XDX_{D} is δ\delta-hyperbolic in the sense of Gromov’s four point condition, for some δ>0\delta>0. We also examine how the properness of a group action on a CAT(0)\mathrm{CAT}(0) space XX is revised when we move to the curtain model XDX_{D}.
This is summarised in Theorem 2.8 below, which follows directly from Proposition 9.5, Theorem 9.10, and Proposition 9.16 of [PSZ24].

Theorem 2.8.

Let XX be a CAT(0)\mathrm{CAT}(0) space. Then, XDX_{D} is a roughly geodesic hyperbolic space. Moreover, if GG is a group acting properly discontinuously on XX, then the induced action of GG on XDX_{D} is weakly acylindrical.

Now, we look at how the Gromov boundary XD\partial X_{D} of XDX_{D} embeds in the visual boundary X\partial X of XX. This will be needed in section 5 to prove Theorem 1.5.

Theorem 2.9 (Theorem L in [PSZ24]).

Let XX be a proper CAT(0)\mathrm{CAT}(0) space. Then the space XD\partial X_{D} embeds homeomorphically as an Isom\mathrm{Isom} XX-invariant subspace of X\partial X, and every point in the image of XD\partial X_{D} is a visibility point of X\partial X. The embedding is induced by the change-of-metric map XDXX_{D}\rightarrow X. Moreover, if XD\partial X_{D} is nonempty and XX is cobounded, then XD\partial X_{D} is dense in X\partial X.

2.2. Actions on a Hyperbolic Space

Here, we briefly review Gromov’s classification of group actions on hyperbolic spaces as given in Section 3.A of [CCMT15]. We will use this in Section 5 to prove Theorem 1.3.

Firstly, we recall Gromov’s classification of isometries of hyperbolic spaces in terms of the translation length.

Definition 2.10.

Let XX be a metric space and ϕ\phi be an isometry of XX. Then, the translation length of ϕ\phi is τ(ϕ):=limnd(x,ϕn(x))/n\tau(\phi):=\lim_{n\rightarrow\infty}d(x,\phi^{n}(x))/n.

Remark.

One can check that τ(ϕ)\tau(\phi) is well-defined and is independent of the choice of xXx\in X.

Let XX be a hyperbolic space. An isometry ϕ\phi of XX is called -
Elliptic if ϕ\phi has bounded orbits.
Parabolic if ϕ\phi has unbounded orbits and τ(ϕ)=0\tau(\phi)=0.
Hyperbolic if τ(ϕ)>0\tau(\phi)>0.

Next, we recall the notion of the Gromov boundary of a hyperbolic space XX and that of the limit set of a group in XX.

Fix a base point xXx\in X. Define the Gromov product of points y,zXy,z\in X with respect to xx as (y|z)x:=(d(x,y)+d(x,z)d(y,z))/2(y|z)_{x}:=(d(x,y)+d(x,z)-d(y,z))/2. A sequence (xn)(x_{n}) in XX is said to be Cauchy-Gromov if (xn|xm)x(x_{n}|x_{m})_{x}\rightarrow\infty as n,mn,m\rightarrow\infty. For convenience, we will write (xn|xm)(x_{n}|x_{m}) to mean (xn|xm)x(x_{n}|x_{m})_{x}.

Remark.

Note that |(y|z)x(y|z)w|=|d(x,y)d(w,y)+d(x,z)d(w,z)|/2d(x,w)|(y|z)_{x}-(y|z)_{w}|=|d(x,y)-d(w,y)+d(x,z)-d(w,z)|/2\leq d(x,w). So the notion of a Cauchy-Gromov sequence is independent of the base point xXx\in X.

Define an equivalence relation on the Cauchy-Gromov sequences as follows: two sequences (yn)(y_{n}) and (zn)(z_{n}) are equivalent, denoted (yn)(zn)(y_{n})\sim(z_{n}), if (yn|zn)(y_{n}|z_{n})\rightarrow\infty as nn\rightarrow\infty.

Definition 2.11.

Let XX be a hyperbolic space. The gromov boundary X\partial X of XX is X:={(xn)(xn) is a cauchy-gromov sequence}/\partial X:=\{(x_{n})\mid(x_{n})\text{ is a cauchy-gromov sequence}\}/{\sim}.

Definition 2.12.

Let GG be a group acting on a hyperbolic space XX via isometries. The limit set of GG in XX is X(G):={(yn)X(yn)(gnx) for some xX,gnG}\partial_{X}(G):=\{(y_{n})\in\partial X\mid(y_{n})\sim(g_{n}\cdot x)\text{ for some }x\in X,g_{n}\in G\}.

Now, we state Gromov’s classification of group actions on hyperbolic spaces.

Theorem 2.13 (Gromov’s Classification).

Let GG be a group acting on a hyperbolic geodesic metric space XX. Then exactly one of the following holds, and the action of GG is said to be -
(i) bounded if the orbits of GG in XX are bounded.
(ii) horocyclic if the orbits are unbounded and GG contains no hyperbolic isometry.
(iii) lineal if GG contains a hyperbolic isometry, and any two hyperbolic elements have the same endpoints.
(iv) focal if it is not lineal, GG contains a hyperbolic isometry, and all hyperbolic elements have one common endpoint.
(v) general type if GG contains two hyperbolic isometries which have no common end-point.

In fact, the above classification can be described in terms of the limit set X(G)\partial_{X}(G).

Theorem 2.14 (Proposition 3.1 in [CCMT15]).

Let GG be a group acting on a hyperbolic geodesic metric space XX. Then, the action of GG is -
(i) bounded X(G)\iff\partial_{X}(G) is empty.
(ii) horocyclic |X(G)|=1\iff|\partial_{X}(G)|=1; then X(G)\partial_{X}(G) is the unique finite orbit of GG in X\partial X.
(iii) lineal |X(G)|=2\iff|\partial_{X}(G)|=2; then X(G)\partial_{X}(G) contains all the finite orbits of GG in X\partial X.
(iv) focal X(G)\iff\partial_{X}(G) is uncountable and GG has a fixed point ξ\xi in X(G)\partial_{X}(G); then ξ\xi is the unique finite orbit of GG in X\partial X.
(v) general type X(G)\iff\partial_{X}(G) is uncountable and GG has no finite orbit in X\partial X.

In proving Theorem 1.3 using this classification, it will be important for us to be able to rule out horocyclic and lineal actions. We record two lemmas regarding this.

Lemma 2.15.

Let XX be a hyperbolic geodesic metric space and suppose the action of GG on XX is lineal. Then GG has a quasi-line orbit in XX.

Proof.

This is well-known, but we include a proof for completeness.
Let {ξ+,ξ}\{\xi^{+},\xi^{-}\} be the limit set of GG. Since the Gromov boundary is a visibility space, there exists a geodesic γ\gamma with endpoints ξ\xi^{-} and ξ+\xi^{+}. Since GG fixes X(G)\partial_{X}(G), it fixes the endpoints of γ\gamma. Thus, gγg\cdot\gamma is a geodesic with the same endpoints as γ\gamma for every gGg\in G. By hyperbolicity of XX, the Hausdorff distance between γ\gamma and gγg\cdot\gamma is uniformly bounded. In particular, there exists RR such that gγNR(γ)g\cdot\gamma\subseteq N_{R}(\gamma) for all gGg\in G.
Now, fix a basepoint x0x_{0} on γ\gamma. We will show that the orbit of x0x_{0} is a quasi-line. The previous argument yields that Gx0NR(γ)G\cdot x_{0}\subseteq N_{R}(\gamma). Moreover, GG contains a hyperbolic isometry by Theorem 2.13, and hence we get that the orbit Gx0G\cdot x_{0} coarsely coincides with a neighbourhood of γ\gamma. ∎

Lemma 2.16.

Let GG be a group acting co-boundedly on a hyperbolic geodesic metric space XX. Then the action of GG on XX is not horocyclic.

Proof.

Since the action of GG on XX is co-bounded, the orbit Gx0G\cdot x_{0} is quasi-dense in XX. Consequently, the orbit Gx0G\cdot x_{0} is quasi-convex. By Proposition 3.2 of [CCMT15], it follows that the action of GG cannot be horocyclic. ∎

Remark.

Note that Theorems 2.13, 2.14, and Lemmas 2.15, 2.16 also hold for roughly geodesic hyperbolic spaces XX. This is because we may replace XX by the injective hull E(X)E(X) of XX which now becomes a geodesic hyperbolic space. Moreover, there is an isometric embedding i:XE(X)i:X\rightarrow E(X) which is coarsely surjective and GG-equivariant.

3. Coarse stabilizer and the coarse fixed set

We work with a general metric space XX and give an algebraic characterisation of the coarse stabilizer GYG_{Y} of YY. In particular, we show that if GG acts geometrically on XX and YY coarsely coincides with an orbit of a finitely generated subgroup HH of GG, then GYG_{Y} is the union of centralisers CG(H)C_{G}(H^{\prime}) where HH^{\prime} runs over finite index subgroups of HH.
In Section 5, we use this to show that for a CAT(0)\mathrm{CAT}(0) group, GXG_{X} is virtually n\mathbb{Z}^{n}.

The difficult part of the theorem lies in showing that if gGYg\in G_{Y}, then gCG(H)g\in C_{G}(H^{\prime}) for some finite index subgroup HH^{\prime} of HH.
For this, our strategy will be to show that CG(g)C_{G}(\langle g\rangle) acts geometrically on a subset CFixX(g)CFix_{X}(\langle g\rangle) of XX containing YY. Then, it will follow that CG(g)C_{G}(\langle g\rangle) must contain a finite index subgroup of HH, giving us the desired result.

To this end, we define the following.

Definition 3.1.

Let GG be a group acting on a metric space XX and let HGH\leq G. Define the coarse fixed set of H to be CFixX(H):={xXd(hx,x)C(h) hH}CFix_{X}(H):=\{x\in X\mid d(h\cdot x,x)\leq C(h)\text{ }\forall h\in H\}. Here C(h)C(h) is any constant depending on hh, chosen so that CFixX(H)CFix_{X}(H)\neq\emptyset.
(Note that we can always find such a function C:HC:H\rightarrow\mathbb{R}. Indeed, fix x0Xx_{0}\in X, then defining C(h):=d(x0,hx0)C(h):=d(x_{0},h\cdot x_{0}) ensures that x0CFixX(H)x_{0}\in CFix_{X}(H).)

A consequence of Theorem 3.2 below is that if GG is a group acting geometrically on XX and HH is a finitely generated subgroup of GG then, up to quasi-isometries, CFixX(H)CFix_{X}(H) does not depend on the choice of the function C:HC:H\rightarrow\mathbb{R}.

Theorem 3.2.

Let XX be a metric space and suppose GG acts on XX properly discontinuously and co-boundedly. Let HGH\leq G be finitely generated, and let YXY\subseteq X coarsely coincide with an orbit of HH. Then,
(i) CG(H)C_{G}(H) acts geometrically on CFixX(H)CFix_{X}(H)
(ii) GY={CG(H)|H/H|<}G_{Y}=\bigcup\{C_{G}(H^{\prime})\mid|H/H^{\prime}|<\infty\}.

Proof.

(i) This can be proven in a similar way to Theorem 3.2 in [Rua01].
We include the proof for completeness.
Since HH is finitely generated, so let h1,h2,,hnh_{1},h_{2},...,h_{n} generate HH. It is straightforward to see that CG(H)CFixX(H)CFixX(H)C_{G}(H)\cdot CFix_{X}(H)\subset CFix_{X}(H). Since GG acts on XX properly discontinuously, we have that CG(H)C_{G}(H) acts on CFixX(H)CFix_{X}(H) properly discontinuously.
So it remains to prove that CG(H)C_{G}(H) acts on CFixX(H)CFix_{X}(H) co-boundedly. Let x0CFixX(H)x_{0}\in CFix_{X}(H) and suppose for a contradiction that CG(H)C_{G}(H) does not act co-boundedly on CFixX(H)CFix_{X}(H). Then there exists znCFixX(H)z_{n}\in CFix_{X}(H) with d(zn,CG(H)x0)d(z_{n},C_{G}(H)\cdot x_{0})\rightarrow\infty as nn\rightarrow\infty. And since Gx0G\cdot x_{0} is quasi-dense in XX, so there exists gnGg_{n}\in G and L>0L>0 so that d(gnx0,zn)<Ld(g_{n}\cdot x_{0},z_{n})<L for all nn\in\mathbb{N}. Hence, we have

d(gnx0,CG(H)x0)d(zn,CG(H)x0)d(gnx0,zn)d(zn,CG(H)x0)L as n.\begin{split}d(g_{n}\cdot x_{0},C_{G}(H)\cdot x_{0})&\geq d(z_{n},C_{G}(H)\cdot x_{0})-d(g_{n}\cdot x_{0},z_{n})\\ &\geq d(z_{n},C_{G}(H)\cdot x_{0})-L\\ &\rightarrow\infty\text{ as }n\rightarrow\infty.\end{split} (1)

Let K=maxi=1,2,..,nC(hi)+2LK=max_{i=1,2,..,n}C(h_{i})+2L. So,

d(x0,gm1higmx0)=d(gmx0,higmx0)d(zm,hizm)+2LC(hi)+2LK.d(x_{0},g_{m}^{-1}h_{i}g_{m}\cdot x_{0})=d(g_{m}\cdot x_{0},h_{i}g_{m}\cdot x_{0})\leq d(z_{m},h_{i}\cdot z_{m})+2L\leq C(h_{i})+2L\leq K.

Since GG acts on XX properly discontinuously, so there are only finitely many gGg\in G with gx0BK(x0)g\cdot x_{0}\in B_{K}(x_{0}). Thus, {gm1higm:m,1in}\{g_{m}^{-1}h_{i}g_{m}:m\in\mathbb{N},1\leq i\leq n\} is a finite set.

So, we may restrict to a subsequence of (gn)(g_{n}) to get that the equation gm1h1gm=gn1h1gn{g_{m}^{-1}h_{1}g_{m}=g_{n}^{-1}h_{1}g_{n}} holds for all gm,gng_{m},g_{n} in the subsequence. Doing the same for h2,h3,..,hnh_{2},h_{3},..,h_{n} we get a further subsequence gmng_{m_{n}} of (gn)n1(g_{n})_{n\geq 1} so that gmn1higmn=gmn1higmng_{m_{n^{\prime}}}^{-1}h_{i}g_{m_{n^{\prime}}}=g_{m_{n}}^{-1}h_{i}g_{m_{n}} for all i,n,ni,n,n^{\prime}. Thus, gmngm11CG(H){g_{m_{n}}g_{m_{1}}^{-1}\in C_{G}(H)} for all nn\in\mathbb{N} and a fixed gm1g_{m_{1}}.
So d(gmnx0,CG(H)x0)d(gmnx0,gmngm11x0)=d(x0,gm11x0)d(g_{m_{n}}\cdot x_{0},C_{G}(H)\cdot x_{0})\leq d(g_{m_{n}}\cdot x_{0},g_{m_{n}}g_{m_{1}}^{-1}\cdot x_{0})=d(x_{0},g_{m_{1}}^{-1}\cdot x_{0}) is bounded as nn tends to infinity, contradicting equation 1.

(ii) If gGYg\in G_{Y}, then YCFixX(g)Y\subseteq CFix_{X}(\langle g\rangle). Note that CFixX(g)CFix_{X}(\langle g\rangle) depends on a function f:gf\colon\langle g\rangle\rightarrow\mathbb{R}. But we can choose ff to satisfy f(gn)=nCf(g^{n})=nC where d(gy,y)Cd(g\cdot y,y)\leq C for all yYy\in Y. This ensures YCFixX(g)Y\subseteq CFix_{X}(\langle g\rangle).
By (i), CG(g)C_{G}(g) acts geometrically on CFixX(g)CFix_{X}(\langle g\rangle). Thus, if y0Yy_{0}\in Y, then CG(g)y0NR(Y){C_{G}(g)\cdot y_{0}\cap N_{R}(Y)} is quasi-dense in NR(Y)N_{R}(Y) for some neighborhood NR(Y)N_{R}(Y) of YY. Since YY coarsely coincides with some (and hence every) orbit of HH, the orbit Hy0H\cdot y_{0} is quasi-dense in some neighborhood NR(Y)N_{R^{\prime}}(Y) of YY. Therefore, there is a constant DD such that for all hHh\in H, there exists zCG(g)z\in C_{G}(g) with d(hy0,zy0)<Dd(h\cdot y_{0},z\cdot y_{0})<D. In other words, d(z1hy0,y0)<Dd(z^{-1}h\cdot y_{0},y_{0})<D, and thus z1h{g1,g2,,gn}z^{-1}h\in\{g_{1},g_{2},...,g_{n}\}, a finite set. This is because the action of GG on XX is properly discontinuous, meaning that there are only finitely many gGg\in G with gy0BD(y0)g\cdot y_{0}\in B_{D}(y_{0}).
So, HCG(g)g1CG(g)g2CG(g)gnH\subseteq C_{G}(g)\cdot g_{1}\cup C_{G}(g)\cdot g_{2}\cup...\cup C_{G}(g)\cdot g_{n}. Thus, [CG(g)H:CG(g)]n[C_{G}(g)H:C_{G}(g)]\leq n and hence [H:CG(g)H]=[CG(g)H:CG(g)]n[H:C_{G}(g)\cap H]=[C_{G}(g)H:C_{G}(g)]\leq n. The last equation is using the fact that for two subgroups A,BA,B it holds [A:AB]=[BA:B][A:A\cap B]=[BA:B], where BABA is the set {babB,aA}\{ba\mid b\in B,a\in A\}. Thus, H:=CG(g)HH^{\prime}:=C_{G}(g)\cap H is a finite index subgroup of HH. Also, gCG(H)g\in C_{G}(H^{\prime}) because HCG(g)H^{\prime}\subseteq C_{G}(g) and so gg commutes with all hHh^{\prime}\in H^{\prime}.
Hence, g{CG(H)|H/H|<}g\in\bigcup\{C_{G}(H^{\prime})\mid|H/H^{\prime}|<\infty\}.

Conversely, if gCG(H)g\in C_{G}(H^{\prime}) for some finite index subgroup HH^{\prime} of HH, then we would have d(g(hy0),hy0)=d(h(gy0),hy0)=d(gy0,y0)d(g\cdot(h\cdot y_{0}),h\cdot y_{0})=d(h\cdot(g\cdot y_{0}),h\cdot y_{0})=d(g\cdot y_{0},y_{0}) for all hHh\in H^{\prime}. Therefore, gGHy0g\in G_{H^{\prime}\cdot y_{0}}. Since Hy0H^{\prime}\cdot y_{0} is quasi-dense in YY (as HH acts co-boundedly on YY, and |H/H|<|H/H^{\prime}|<\infty), it follows that gGYg\in G_{Y}. Indeed, let C>0C>0 be a constant such that for all yYy\in Y, there exists hHh\in H^{\prime} so that d(hy0,y)Cd(h\cdot y_{0},y)\leq C. Then, for all yYy\in Y, we have

d(gy,y)d(g(hy0),gy)+d(g(hy0),hy0)+d(hy0,y)2C+d(gy0,y0).d(g\cdot y,y)\leq d(g\cdot(h\cdot y_{0}),g\cdot y)+d(g\cdot(h\cdot y_{0}),h\cdot y_{0})+d(h\cdot y_{0},y)\leq 2C+d(g\cdot y_{0},y_{0}).

Remark.

In particular, for a finitely generated subgroup HGH\leq G we can tell that CG(H)C_{G}(H) is infinite just by showing CFixX(H)CFix_{X}(H) is unbounded for some metric space XX on which GG acts geometrically.

Remark.

Note that under the assumptions as above, we also have that, up to quasi-isometries, CFixX(H)CFix_{X}(H) is independent of the choice of the function C:HC:H\rightarrow\mathbb{R}.
Indeed, suppose CFixX(H,C1)CFix_{X}(H,C_{1}) and CFixX(H,C2)CFix_{X}(H,C_{2}) are two such subsets of XX corresponding to functions C1,C2:HC_{1},C_{2}:H\rightarrow\mathbb{R}. Pick any x0CFixX(H,C1)x_{0}\in CFix_{X}(H,C_{1}). Part (i) of Theorem 3.2 tells us that CG(H)x0C_{G}(H)\cdot x_{0} is quasi-dense in CFixX(H,C)CFix_{X}(H,C) for any function CC. And since x0CFixX(H,C1)CFixX(H,C1+C2)x_{0}\in CFix_{X}(H,C_{1})\subseteq CFix_{X}(H,C_{1}+C_{2}), we have that CG(H)x0C_{G}(H)\cdot x_{0} is quasi-dense in CFixX(H,C1+C2)CFix_{X}(H,C_{1}+C_{2}). If we denote by ABA\sim B the equivalence relation of quasi-isometry, then we have CFixX(H,C1+C2)CG(H)x0CFixX(H,C1)CFix_{X}(H,C_{1}+C_{2})\sim C_{G}(H)\cdot x_{0}\sim CFix_{X}(H,C_{1}). Similarly, we get CFixX(H,C1+C2)CFixX(H,C2)CFix_{X}(H,C_{1}+C_{2})\sim CFix_{X}(H,C_{2}). Therefore, CFixX(H,C2)CFixX(H,C1)CFix_{X}(H,C_{2})\sim CFix_{X}(H,C_{1}).

Corollary 3.3.

Let GG be a group acting properly discontinuously and co-boundedly on a metric space XX. Then, GX={Z(H)|G/H|<}G_{X}=\bigcup\{Z(H)\mid|G/H|<\infty\}.

Proof.

Applying Theorem 3.2 with Y=XY=X gives us GX={CG(H)|G/H|<}G_{X}=\bigcup\{C_{G}(H)\mid|G/H|<\infty\}. Now, gZ(H)g\in Z(H) implies gCG(H)g\in C_{G}(H), and hence {Z(H)|G/H|<}GX\bigcup\{Z(H)\mid|G/H|<\infty\}\subseteq G_{X}. Conversely, if gGXg\in G_{X}, then gCG(H)g\in C_{G}(H) for some finite index subgroup HH of GG. Letting H~\tilde{H} be the subgroup of GG generated by H{g}H\cup\{g\}, we see that gZ(H~)g\in Z(\tilde{H}). Moreover, since HH~GH\subseteq\tilde{H}\subseteq G and |G/H|<|G/H|<\infty, H~\tilde{H} is a finite index subgroup of GG.
Therefore, g{Z(H)|G/H|<}g\in\bigcup\{Z(H)\mid|G/H|<\infty\}. ∎

4. Weakly-acylindrical actions

This is an aside on weakly acylindrical actions on a quasi-line. We show that a group acting weakly acylindrically on a roughly geodesic hyperbolic space with a quasi-line orbit has to be virtually \mathbb{Z}. This will be usefull later while proving Theorem 5.3.

Proposition 4.1.

Let XX be a roughly geodesic hyperbolic space. Let GG act on XX weakly acylindrically with a quasi-line orbit. Then GG is virtually \mathbb{Z}.

Proof.

Without loss of generality, we may assume that XX is a geodesic metric space. This is because we may replace XX by E(X)E(X), the injective hull of XX. Then, there exists i:XE(X)i:X\rightarrow E(X), an isometric embedding, which is coarsely surjective and GG-equivariant. It is straightforward to check that GG acts weakly acylindrically on E(X)E(X). Also, if Gx0G\cdot x_{0} is a quasi-line orbit of GG in XX then i(Gx0)i(G\cdot x_{0}) is a quasi-line orbit of GG acting on E(X)E(X).

So, let XX be a geodesic hyperbolic space and YXY\subseteq X be a quasi-line orbit of GG.

Note that the action of GG on YY extends to an action of GG on Y\partial Y. Thus, we get a homomorphism ϕ:GHomeo(Y)\phi:G\rightarrow\operatorname{Homeo}(\partial Y) with kernel KK. Since |Y|=2|\partial Y|=2, we have |G/K||Homeo(Y)|=2|G/K|\leq|\operatorname{Homeo}(\partial Y)|=2. Thus, KK has finite index in GG. So, an orbit of KK is quasi-dense in YY, and hence is a quasi-line. Thus, it suffices to prove the proposition for KK. It will then follow that KK, and consequently GG, is virtually \mathbb{Z}. So, we assume, without loss of generality, that the action of GG on Y\partial Y is trivial.

Now, |YG|=|Y|=2|\partial_{Y}G|=|\partial Y|=2 and hence, by Theorem 2.14, it follows that the action of GG on YY is lineal. So, GG contains a hyperbolic element gGg\in G i.e. there exists gGg\in G satisfying τ:=limnd(x,gnx)/n>0\tau:=\lim_{n\to\infty}d(x,g^{n}\cdot x)/n>0. Since YY is a quasi-line orbit of GG, we get that for any yYy\in Y, the orbit gy\langle g\rangle\cdot y is quasi-dense in YY. So, ψ:Y\psi:\mathbb{Z}\rightarrow Y given by ψ(n)=gny\psi(n)=g^{n}\cdot y, is a (L,C)(L,C)-quasi-isometry for some C,L>0C,L>0.

Now, we will show that g\langle g\rangle has finitely many right cosets in GG. This implies that GG is virtually isomorphic to g\langle g\rangle which is isomorphic to \mathbb{Z}.

Fix yYy\in Y. Let ga\langle g\rangle a be a right-coset of g\langle g\rangle.
As ψ\psi is a (L,C)(L,C)-quasi-isometry, so d(ay,ψ(m))=d(ay,gmy)<Cd(a\cdot y,\psi(m))=d(a\cdot y,g^{m}\cdot y)<C for some mm\in\mathbb{Z}. Let h=gmah=g^{-m}a. Then,

d(hy,y)=d(gmay,y)=d(ay,gmy)<Cd(h\cdot y,y)=d(g^{-m}a\cdot y,y)=d(a\cdot y,g^{m}\cdot y)<C (2)

Similarly, hgnyhg^{n}\cdot y is CC-close to gbnyg^{b_{n}}\cdot y for some bnb_{n}\in\mathbb{Z}. Moreover, as hh acts trivially on Y\partial Y, so (gbny)n0=h(gny)n0=(gny)n0Y({g^{b_{n}}\cdot y})_{n\geq 0}=h\cdot({g^{n}\cdot y})_{n\geq 0}=({g^{n}\cdot y})_{n\geq 0}\in\partial Y. Thus, bnb_{n}\rightarrow\infty as nn\rightarrow\infty. Hence, bn>0b_{n}>0 for all nMn\geq M for some MM\in\mathbb{N}.

Fix nMn\geq M and let N>max(bn,n)N>max(b_{n},n). Now, ψ([0,N])\psi([0,N]) is a (L,C)(L,C)-quasi-geodesic with endpoints y,gNyy,g^{N}\cdot y. Let γ\gamma be a geodesic from yy to gNyg^{N}\cdot y. By the Morse lemma, the hausdorff distance dH(γ,ψ([0,N]))Kd_{H}(\gamma,\psi([0,N]))\leq K for some K>0K>0 which depends just on CC, LL. Thus, there exists cn,dnγc_{n},d_{n}\in\gamma which satisfy d(cn,gny)Kd(c_{n},g^{n}\cdot y)\leq K and d(dn,gbny)Kd(d_{n},g^{b_{n}}\cdot y)\leq K.

Refer to caption
dnd_{n}gbnyg^{b_{n}}\cdot yhgnyhg^{n}\cdot yyyhyh\cdot ygNyg^{N}\cdot ygnyg^{n}\cdot ycnc_{n}γ\gamma
Figure 1. The proof of Proposition 4.1

We hence obtain the following inequalities.

|d(y,dn)d(y,gny)|=|d(y,dn)d(hy,hgny)|d(y,hy)+d(dn,gbny)+d(gbny,hgny)2C+K.\begin{split}|d(y,d_{n})-d(y,g^{n}\cdot y)|&=|d(y,d_{n})-d(h\cdot y,hg^{n}\cdot y)|\\ &\leq d(y,h\cdot y)+d(d_{n},g^{b_{n}}\cdot y)+d(g^{b_{n}}\cdot y,hg^{n}\cdot y)\\ &\leq 2C+K.\end{split} (3)
|d(y,cn)d(y,gny)|d(cn,gny)K.|d(y,c_{n})-d(y,g^{n}\cdot y)|\leq d(c_{n},g^{n}\cdot y)\leq K. (4)

Thus, we have

d(gbny,gny)d(dn,cn)+d(dn,gbny)+d(cn,gny)d(cn,dn)+2K=|d(cn,y)d(dn,y)|+2K|d(y,cn)d(y,gny)|+|d(y,dn)d(y,gny)|+2K2C+4K.\begin{split}d(g^{b_{n}}\cdot y,g^{n}\cdot y)&\leq d(d_{n},c_{n})+d(d_{n},g^{b_{n}}\cdot y)+d(c_{n},g^{n}\cdot y)\\ &\leq d(c_{n},d_{n})+2K\\ &=|d(c_{n},y)-d(d_{n},y)|+2K\\ &\leq|d(y,c_{n})-d(y,g^{n}\cdot y)|+|d(y,d_{n})-d(y,g^{n}\cdot y)|+2K\\ &\leq 2C+4K.\end{split} (5)

So, d(hgny,gny)d(gbny,gny)+d(gbny,hgny)3C+4Kd(hg^{n}\cdot y,g^{n}\cdot y)\leq d(g^{b_{n}}\cdot y,g^{n}\cdot y)+d(g^{b_{n}}\cdot y,hg^{n}\cdot y)\leq 3C+4K.

As GG acts weakly acylindrically on XX, so there exists R>0R>0 such that for any y1,y2Yy_{1},y_{2}\in Y with d(y1,y2)>Rd(y_{1},y_{2})>R there are only finitely many kGk\in G which satisfy max(d(ky1,y1),d(ky2,y2))3C+4K{max(d(k\cdot y_{1},y_{1}),d(k\cdot y_{2},y_{2}))\leq 3C+4K}. Since, gy\langle g\rangle\cdot y is quasi-dense in a quasi-line, so we can choose nn large enough so that d(y,gny)>Rd(y,g^{n}\cdot y)>R. Applying the condition that GG acts weakly acylindrically on YY with y1=yy_{1}=y, y2=gnyy_{2}=g^{n}\cdot y, and k=h:=gmak=h:=g^{-m}a, we see that there are only finitely many possible values for gmag^{-m}a.

So, there are only finitely many right-cosets ga=ggma\langle g\rangle a=\langle g\rangle g^{-m}a of g\langle g\rangle. ∎

We record a consequence of the above proposition, which can be seen as part of the Tits Alternative for CAT(0)\mathrm{CAT}(0) groups, serving as a counterpart to the ping-pong lemmas.

Corollary 4.2.

Let GG be a group acting geometrically on a CAT(0)\mathrm{CAT}(0) space XX. Let HH be any subgroup of GG. If |XDH|=2|\partial_{X_{D}}H|=2 then HH is virtually \mathbb{Z}.

Proof.

As GG (and hence HH) acts properly discontinuously on XX, so by Theorem 2.8, XDX_{D} is a roughly geodesic hyperbolic space and HH acts weakly acylindrically on XDX_{D}. Also, since |XDH|=2|\partial_{X_{D}}H|=2, by Theorem 2.14 and Lemma 2.15, the action of HH on XDX_{D} is lineal and HH has a quasi-line orbit in XDX_{D}. Hence, by Proposition 4.1, HH is virtually \mathbb{Z}. ∎

5. CAT(0)\mathrm{CAT}(0) Spaces

In this section, we specialise to the case when XX is a CAT(0)\mathrm{CAT}(0) space.

First, we relate the coarse kernels GXG_{X} and GXDG_{X_{D}} corresponding to the action of GG on XX and the induced action of GG on XDX_{D}. In particular, we show that if XX is a proper CAT(0)\mathrm{CAT}(0) space with an unbounded curtain model XDX_{D}, and GG acts co-compactly on XX, then GXG_{X} and GXDG_{X_{D}} coincide.

Our strategy will be to look at the action of GG on the boundaries X\partial X and XD\partial X_{D} of XX and XDX_{D} respectively. In particular, we look at the elements in GG which fix the boundaries of XX and XDX_{D}. Define, KX:={gGg fixes the boundary X pointwise}{K_{X}:=\{g\in G\mid g\text{ fixes the boundary }\partial X\text{ pointwise}\}} and KD:={gGg fixes XD pointwise}K_{D}:=\{g\in G\mid g\text{ fixes }\partial X_{D}\text{ pointwise}\}. We show that GXGXDKDG_{X}\subseteq G_{X_{D}}\subseteq K_{D} and that GX=KX=KDG_{X}=K_{X}=K_{D}. These together give us GX=GXDG_{X}=G_{X_{D}}.

Lemma 5.1.

Let XX be a CAT(0)\mathrm{CAT}(0) space and let GG be a group acting via isometries on XX. If GXG_{X}, GXDG_{X_{D}}, and KDK_{D} are as defined above, then GXGXDKDG_{X}\subseteq G_{X_{D}}\subseteq K_{D}.

Proof.

If gGXg\in G_{X}, then d(x,gx)Cd(x,g\cdot x)\leq C for all xXx\in X, for some fixed constant CC. Hence, by Lemma 2.7, we have D(x,gx)Λ(1+d(x,gx))Λ(C+1)D(x,g\cdot x)\leq\Lambda\cdot(1+d(x,g\cdot x))\leq\Lambda\cdot(C+1) for all xXx\in X. So, gGXDg\in G_{X_{D}}.
If gGXDg\in G_{X_{D}}, then D(x,gx)CD(x,g\cdot x)\leq C for all xXx\in X and some fixed constant CC. Thus, if (xn)(x_{n}) is a Cauchy-Gromov sequence, then (gxn)(g\cdot x_{n}) is a Cauchy-Gromov sequence equivalent to (xn)(x_{n}).
This is because,

(xn|gxn)w\displaystyle(x_{n}|g\cdot x_{n})_{w} =[D(xn,w)+D(gxn,w)D(xn,gxn)]/2\displaystyle=[D(x_{n},w)+D(g\cdot x_{n},w)-D(x_{n},g\cdot x_{n})]/2
[D(xn,w)+D(xn,w)D(xn,gxn)D(xn,gxn)]/2\displaystyle\geq[D(x_{n},w)+D(x_{n},w)-D(x_{n},g\cdot x_{n})-D(x_{n},g\cdot x_{n})]/2
[D(xn,w)C] as n\displaystyle\geq[D(x_{n},w)-C]\rightarrow\infty\text{ as }n\rightarrow\infty

Hence, gg fixes XD\partial X_{D}, implying that gKD.g\in K_{D}.

Theorem 5.2.

Let XX be a proper CAT(0)\mathrm{CAT}(0) space with XDX_{D} being unbounded, and let GG act on XX co-compactly. Then, GX=GXDG_{X}=G_{X_{D}}.

Proof.

We start with showing that GXG_{X} and KXK_{X} coincide.

Claim.

GX=KXG_{X}=K_{X}.

Proof of Claim.

We first show that KXGXK_{X}\subseteq G_{X}. Fix gKXg\in K_{X} and xXx\in X. As XDX_{D} is unbounded, XX is unbounded as well, and hence non-compact, and since GG acts on XX co-compactly, use Corollary 3 in [GO07] to obtain a constant CC such that for each yXy\in X there is a geodesic ray γ:[0,]X\gamma:[0,\infty]\rightarrow X with γ(0)=x\gamma(0)=x and d(y,z)<Cd(y,z)<C for some zIm(γ)z\in\mathrm{Im}(\gamma). We claim there exists CC^{\prime} not depending on yy such that d(y,gy)Cd(y,g\cdot y)\leq C^{\prime}. We have d(y,gy)d(y,z)+d(z,gz)+d(gz,gy)2C+d(z,gz)d(y,g\cdot y)\leq d(y,z)+d(z,g\cdot z)+d(g\cdot z,g\cdot y)\leq 2C+d(z,g\cdot z). Moreover, if d(z,gz)>d(x,gx)d(z,g\cdot z)>d(x,g\cdot x) then by the convexity of CAT(0)\mathrm{CAT}(0) distance we have that d(γ(t),gγ(t))d(\gamma(t),g\cdot\gamma(t))\rightarrow\infty which contradicts gKXg\in K_{X}. Thus, d(z,gz)d(x,gx)d(z,g\cdot z)\leq d(x,g\cdot x) and hence d(y,gy)2C+d(z,gz)2C+d(x,gx)d(y,g\cdot y)\leq 2C+d(z,g\cdot z)\leq 2C+d(x,g\cdot x). Since xXx\in X is fixed, C:=2C+d(x,gx)C^{\prime}:=2C+d(x,g\cdot x) does not depend on yXy\in X. So, gGXg\in G_{X}.

Conversely, if gGXg\in G_{X} then for any γ:[0,]X\gamma:[0,\infty]\rightarrow X, d(γ(t),gγ(t))<Cd(\gamma(t),g\cdot\gamma(t))<C for some constant CC. Thus, gg fixes the point γ()X\gamma(\infty)\in\partial X. So gKXg\in K_{X}. ∎

We claim KX=KDK_{X}=K_{D}. We note that in the case that the action of GG is properly discontinuous as well, the equality follows by Corollary 1.6 of [BJ24].
Observe that GG acts co-compactly on XX, and hence co-boundedly on XDX_{D}. As XDX_{D} is unbounded, so an orbit of GG in XDX_{D} is unbounded as well. So, by Theorem 2.14, XD\partial X_{D} is nonempty. Thus, by Theorem 2.9, XD\partial X_{D} embeds as an Isom XX-invariant, dense subspace of X\partial X. Now, if gKXg\in K_{X}, then gg fixes X\partial X, and hence gg also fixes XDX\partial X_{D}\hookrightarrow\partial X. So gKDg\in K_{D}. Conversely, if gKDg\in K_{D}, then gg fixes XD\partial X_{D}, which is dense in X\partial X. Since X\partial X is hausdorff and gg acts as a homeomorphism on X\partial X, gg must fix X\partial X as well. Thus, gKXg\in K_{X}.

Therefore, GX=KX=KDG_{X}=K_{X}=K_{D}. Thus, by Lemma 5.1 we have that GX=GXDG_{X}=G_{X_{D}}. ∎

Next, we examine the structure of the coarse kernel GXG_{X} of XX when the action of GG on XX is properly discontinuous but not, in general, co-compact. Specifically, we establish mild conditions concerning the orbit of GG in the curtain model XDX_{D} that ensure the coarse kernel GXG_{X} is virtually cyclic, and likewise finite.

Our key strategy will be to look at the action of the coarse kernel GXDG_{X_{D}} on the curtain model XDX_{D} and apply the classification of actions on hyperbolic spaces, as discussed in Section 2.2. This will allow us to deduce that the subgroup has either a bounded or quasi-line orbit in XDX_{D}. We can then exploit the weak acylindricity of the action of GG on XDX_{D} to obtain the desired result.

Notice, firstly, that for any group GG acting on a metric space XX by isometries, GXG_{X} is a normal subgroup of GG. Indeed, if hGXh\in G_{X}, then d(hx,x)Cd(h\cdot x,x)\leq C for all xXx\in X. Therefore, d(g1hgx,x)=d(hgx,gx)C{d(g^{-1}hg\cdot x,x)=d(hg\cdot x,g\cdot x)\leq C} for all xXx\in X, implying that g1hgGXg^{-1}hg\in G_{X}.

Theorem 5.3.

Let XX be a co-bounded CAT(0)\mathrm{CAT}(0) space, and let GG act on XX properly discontinuously (but not necessarily co-boundedly). Suppose that GG has an unbounded orbit in XDX_{D} . Then, GXG_{X} is virtually cyclic. Moreover, if an orbit of GG in XDX_{D} is not a quasi-line either, then GXG_{X} is finite.

Proof.

By assumption, the orbit of GG in XDX_{D} is not bounded. Firstly, we focus on the case it is not a quasi-line either.

Claim.

If an orbit of GG on XDX_{D} is not a quasi-line, then the action of GXDG_{X_{D}} on XDX_{D} is bounded.

Proof of Claim.

Since GXDKDG_{X_{D}}\subseteq K_{D}, we have that GXDG_{X_{D}} fixes the boundary of XDX_{D}. Thus, by Theorem 2.14, the action of GXDG_{X_{D}} can be either bounded, horocyclic, or lineal. Indeed, if the action of GXDG_{X_{D}} is focal or of general type, then |XD|=|\partial X_{D}|=\infty and GXDG_{X_{D}} fixes at most one point in the boundary XD\partial X_{D}. A contradiction.

If the action is horocyclic, then |XD(GXD)|=1|\partial_{X_{D}}(G_{X_{D}})|=1. Since each element of XD\partial X_{D} is fixed by the action of GXDG_{X_{D}}, it follows from Theorem 2.14 that |XD|=|XD(GXD)|=1|\partial X_{D}|=|\partial_{X_{D}}(G_{X_{D}})|=1. But this is impossible since XX, and hence XDX_{D}, is co-bounded. Indeed, let HH act co-boundedly on XDX_{D}. Then, |XDH|=|XD|=1|\partial_{X_{D}}H|=|\partial X_{D}|=1, implying that the action of HH on XDX_{D} is horocyclic. This contradicts Lemma 2.16.

If the action is lineal, then |XD(GXD)|=2|\partial_{X_{D}}(G_{X_{D}})|=2. Since each element of XD\partial X_{D} is fixed by the action of GXDG_{X_{D}}, it follows from Theorem 2.14 that |XD|=|XD(GXD)|=2|\partial X_{D}|=|\partial_{X_{D}}(G_{X_{D}})|=2. Therefore, we have 2=|XD||XDG||XD(GXD)|22=|\partial X_{D}|\geq|\partial_{X_{D}}G|\geq|\partial_{X_{D}}(G_{X_{D}})|\geq 2. So, |XDG|=2|\partial_{X_{D}}G|=2, implying that the action of GG on XDX_{D} is lineal. Hence, by Lemma 2.15, the orbit of GG is a quasi-line, contradicting our assumption.
So the action of GXDG_{X_{D}} on XDX_{D} is bounded. ∎

Fix xXx\in X. By the claim above, GXDG_{X_{D}} acts boundedly on XDX_{D}, so there is a constant C>0C>0 so that D(x,hx)<CD(x,h\cdot x)<C for all hGXDh\in G_{X_{D}}.
Since GG acts properly discontinuously on XX, by Theorem 2.8 we get that the induced action of GG on XDX_{D} is weakly acylindrical. So let R>0R>0 satisfy that for any x,yXx,y\in X with D(x,y)>RD(x,y)>R, only finitely many hGh\in G have max{D(x,hx),D(y,hy)}<Cmax\{D(x,h\cdot x),D(y,h\cdot y)\}<C. Now, since GG has unbounded orbits in XDX_{D}, we can always find a point yy in the orbit GxG\cdot x with D(x,y)>RD(x,y)>R. Let y=gxy=g\cdot x be such a point. Then,

D(y,hy)=D(gx,h(gx))=D(g1hgx,x)<C hGXD.D(y,h\cdot y)=D(g\cdot x,h\cdot(g\cdot x))=D(g^{-1}hg\cdot x,x)<C\text{ }\forall h\in G_{X_{D}}.

The last inequality follows because GXDGG_{X_{D}}\trianglelefteq G, and thus g1hgGXDg^{-1}hg\in G_{X_{D}}.
So, max{D(x,hx),D(y,hy)}<Cmax\{D(x,h\cdot x),D(y,h\cdot y)\}<C for all hGXDh\in G_{X_{D}}, and D(x,y)>RD(x,y)>R. Hence, GXDG_{X_{D}} must be finite. So GXGXDG_{X}\subseteq G_{X_{D}} is finite as well.

Now, suppose that the orbit of GG in XDX_{D} is a quasi-line. From Theorem 2.8, we have that XDX_{D} is roughly geodesic and GG acts weakly acylindrically on XDX_{D}. So, by Proposition 4.1, GG is virtually \mathbb{Z}. Hence, GXGG_{X}\subseteq G is virtually cyclic. ∎

Finally, using Theorem 5.3 above alongside Theorem 3.2, we obtain the following results regarding the algebraic structure of the coarse kernel GXG_{X} for a group GG that acts geometrically on a CAT(0)\mathrm{CAT}(0) space XX. Consequently, we prove Corollary 1.6 characterizing when the coarse kernels GXG_{X} and GXDG_{X_{D}} coincide.

Theorem 5.4.

Let XX be a CAT(0)\mathrm{CAT}(0) space, and let GG act on XX geometrically. Then,
(i) GXG_{X} is virtually n\mathbb{Z}^{n} for some n0n\in\mathbb{Z}_{\geq 0}.
(ii) If moreover XDX_{D} is unbounded, then GXG_{X} is virtually cyclic.
(iii) If furthermore XDX_{D} is not a quasi-line, then GXG_{X} is the largest finite normal subgroup of GG.

Proof.

(i) Since GG acts geometrically on XX, we can apply Theorem 3.2 to Y=XY=X, yielding GX={CG(H)|G/H|<}G_{X}=\bigcup\{C_{G}(H)\mid|G/H|<\infty\}. Assume that this infinite union cannot be realized as the centraliser of a single finite index subgroup.
Then, we may inductively choose finite index subgroups HiH_{i}, ii\in\mathbb{Z} as follows:-
Pick H1=GH_{1}=G. Suppose we have chosen finite index subgroups H1,H2,,HnH_{1},H_{2},...,H_{n} of GG so that we get proper inclusions HnHn1H1H_{n}\subset H_{n-1}\subset...\subset H_{1}, and CG(H1)CG(H2)CG(Hn){C_{G}(H_{1})\subset C_{G}(H_{2})\subset...\subset C_{G}(H_{n})}. Now, there must exist a finite index subgroup Kn+1K_{n+1} of GG so that CG(Kn+1)CG(Hn)C_{G}(K_{n+1})\not\subseteq C_{G}(H_{n}). If not, then we would have {CG(H)|G/H|<}=CG(Hn)\bigcup\{C_{G}(H)\mid|G/H|<\infty\}=C_{G}(H_{n}), contradicting our assumption. Let Hn+1=Kn+1HnH_{n+1}=K_{n+1}\cap H_{n}. Since HnH_{n} and Kn+1K_{n+1} are finite index subgroups of GG, so is Hn+1H_{n+1}. Observe also that Hn+1H_{n+1} is a proper subgroup of HnH_{n} because CG(Kn+1)CG(Hn){C_{G}(K_{n+1})\not\subseteq C_{G}(H_{n})} and CG(Kn+1)CG(Hn+1)C_{G}(K_{n+1})\subseteq C_{G}(H_{n+1}). In particular, we obtain proper chains Hn+1HnH1H_{n+1}\subset H_{n}\subset...\subset H_{1}, and CG(H1)CG(H2)CG(Hn+1)C_{G}(H_{1})\subset C_{G}(H_{2})\subset...\subset C_{G}(H_{n+1}).

Hence, we have an infinite chain CG(H1)CG(H2)C_{G}(H_{1})\subset C_{G}(H_{2})\subset\cdot\cdot\cdot of subgroups of GG. Moreover, all of these are virtually abelian. Indeed, note that the center Z(Hi)Z(H_{i}) of HiH_{i} is an abelian subgroup of CG(Hi)C_{G}(H_{i}) and we have

|CG(Hi)/Z(Hi)|=|CG(Hi)/CG(Hi)Hi|=|CG(Hi)Hi/Hi||G/Hi||C_{G}(H_{i})/Z(H_{i})|=|C_{G}(H_{i})/{C_{G}(H_{i})\cap H_{i}}|=|{C_{G}(H_{i})\cdot H_{i}}/H_{i}|\leq|G/H_{i}| (6)

where |G/Hi||G/H_{i}| is finite as HiH_{i} is a finite index subgroup of GG.
(The third equality in equation 6 follows from the second isomorphism theorem applied to the group NG(Hi)N_{G}(H_{i}) with subgroups CG(Hi),HiNG(Hi)C_{G}(H_{i}),H_{i}\trianglelefteq N_{G}(H_{i}) )
However, this contradicts the Ascending Chain Condition [[BH99], Theorem 7.5].

So, GX=CG(H)G_{X}=C_{G}(H) for some finite index subgroup HH of GG. Now, from equation 6, we know Z(H)Z(H) is a finite index subgroup of CG(H)C_{G}(H). Also, Z(H)Z(H) is an abelian subgroup of GG and hence, by Corollary 7.6 of [BH99], must be finitely generated. Thus, Z(H)Z(H), and hence CG(H)=GXC_{G}(H)=G_{X}, is virtually n\mathbb{Z}^{n} for some nn.

(ii) As GG acts on XX geometrically, so XX is co-bounded. Also, the orbit of GG is quasi-dense in XX and hence the corresponding orbit in XDX_{D} is quasi-dense as well. As XDX_{D} is unbounded, we have that the orbit of GG in XDX_{D} is unbounded. Therefore, by Theorem 5.3, GXG_{X} is virtually cyclic.

(iii) Now, if XDX_{D} is not a quasi-line then the orbit of GG in XDX_{D} isn’t a quasi-line either. So, by Theorem 5.3, GXG_{X} is finite. Also, GXG_{X} is a normal subgroup of GG. Moreover, if NN is a finite normal subgroup of GG, then NN acts uniformly boundedly on an orbit Gx0G\cdot x_{0}, which is quasi-dense in XX. So NN acts uniformly boundedly on XX, implying that NGXN\subseteq G_{X}. Thus, {N|N is a finite normal subgroup of G}GX=\bigcup\{N|N\text{ is a finite normal subgroup of }G\}\subseteq G_{X}= finite normal subgroup. Thus, GXG_{X} must be the largest finite normal subgroup of GG. ∎

Corollary 5.5.

Let GG be a group acting geometrically on a CAT(0)\mathrm{CAT}(0) space XX. Then exactly one of the following holds.

  1. (1)

    XDX_{D} is unbounded and hence GX=GXDG_{X}=G_{X_{D}};

  2. (2)

    XDX_{D} is bounded, |G:Z(G)|<|G:Z(G)|<\infty, and GX=G=GXDG_{X}=G=G_{X_{D}};

  3. (3)

    XDX_{D} is bounded, Z(G)Z(G) has infinite index in GG, and GXGXDG_{X}\neq G_{X_{D}}.

Proof.

If XDX_{D} is unbounded, then by Theorem 5.3, GX=GXDG_{X}=G_{X_{D}}. Assuming XDX_{D} is bounded, we get GXD=GG_{X_{D}}=G. Therefore, GX=GXDG_{X}=G_{X_{D}} if and only if G=GXG=G_{X}.

Now, by the proof of Theorem 5.4 part (i), we have that GX=CG(H)G_{X}=C_{G}(H) for some finite index subgroup HH of GG. Thus, if G=GXG=G_{X}, then G=CG(H)G=C_{G}(H), and hence HZ(G)H\subseteq Z(G). This implies that |G:Z(G)||G:H|<|G:Z(G)|\leq|G:H|<\infty. Conversely, if |G:Z(G)|<|G:Z(G)|<\infty, then G=CG(Z(G))GXG=C_{G}(Z(G))\subseteq G_{X}. Hence, GX=GG_{X}=G. ∎

References

  • [BH99] Martin R. Bridson and André Haefliger. Metric spaces of non-positive curvature, volume 319 of Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1999.
  • [BJ24] Hyungryul Baik and Wonyong Jang. On the kernel of actions on asymptotic cones. arXiv preprint arXiv:2402.09969, 2024.
  • [CCMT15] Pierre-Emmanuel Caprace, Yves Cornulier, Nicolas Monod, and Romain Tessera. Amenable hyperbolic groups. J. Eur. Math. Soc. (JEMS), 17(11):2903–2947, 2015.
  • [GO07] Ross Geoghegan and Pedro Ontaneda. Boundaries of cocompact proper CAT(0){\rm CAT}(0) spaces. Topology, 46(2):129–137, 2007.
  • [LV23] Arielle Leitner and Federico Vigolo. Coarse Kernels, pages 119–134. Springer Nature Switzerland, Cham, 2023.
  • [PSZ24] Harry Petyt, Davide Spriano, and Abdul Zalloum. Hyperbolic models for cat(0) spaces. Advances in Mathematics, 450:109742, 2024.
  • [Rua01] Kim E. Ruane. Dynamics of the action of a CAT(0){\rm CAT}(0) group on the boundary. Geom. Dedicata, 84(1-3):81–99, 2001.