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

Croke-Kleiner admissible groups: Property (QT) and quasiconvexity

Hoang Thanh Nguyen Beijing International Center for Mathematical Research
Peking University
Beijing 100871, China P.R.
[email protected]
 and  Wenyuan Yang Beijing International Center for Mathematical Research
Peking University
Beijing 100871, China P.R.
[email protected]
Abstract.

Croke-Kleiner admissible groups firstly introduced by Croke-Kleiner in [CK02] belong to a particular class of graph of groups which generalize fundamental groups of 33–dimensional graph manifolds. In this paper, we show that if GG is a Croke-Kleiner admissible group, acting geometrically on a CAT(0) space XX, then a finitely generated subgroup of GG has finite height if and only if it is strongly quasi-convex. We also show that if GXG\curvearrowright X is a flip CKA action then GG is quasi-isometric embedded into a finite product of quasi-trees. With further assumption on the vertex groups of the flip CKA action GXG\curvearrowright X, we show that GG satisfies property (QT) that is introduced by Bestvina-Bromberg-Fujiwara in [BBF19].

1. Introduction

In [CK02], Croke and Kleiner study a particular class of graph of groups which they call admissible groups and generalize fundamental groups of 33–dimensional graph manifolds and torus complexes (see [CK00]). If GG is an admissible group that acts geometrically on a Hadamard space XX then the action GXG\curvearrowright X is called Croke-Kleiner admissibe (see Definition 2.1) termed by Guilbault-Mooney [GM14]. The CKA action is modeling on the JSJ structure of graph manifolds where the Seifert fibration is replaced by the following central extension of a general hyperbolic group:

(1) 1Z(Gv)=GvHv11\to Z(G_{v})=\mathbb{Z}\to G_{v}\to H_{v}\to 1

However, CKA groups can encompass much more general class of groups and can actually serve as one of simplest algebraic means to produce interesting groups from any finite number of hyperbolic groups.

Let 𝒢\mathcal{G} be a finite graph with nn vertices, each of which are associated with a hyperbolic group HiH_{i}. We then pick up an independent set of primitive loxodromic elements in HiH_{i} which crossed with \mathbb{Z} are the edge groups 2\mathbb{Z}^{2}. We identify 2\mathbb{Z}^{2} in adjacent Hi×H_{i}\times\mathbb{Z}’s by flipping \mathbb{Z} and loxodromic elements as did in flip graph manifolds by Kapovich and Leeb [KL98]. These are motivating examples of flip CKA groups and actions, for the precise definition of flip CKA actions, we refer the reader to Section 4.2.

The class of CKA actions has manifested a variety of interesting features in CAT(0) groups. For instance, the equivariant visual boundaries of admissible actions are completely determined in [CK02]. Meanwhile, the non-homeomorphic visual boundaries of torus complexes were constructed in [CK00] and have sparked an intensive research on boundaries of CAT(0) spaces. So far, the most of research on CKA groups is centered around the boundary problem (see [GM14], [Gre16]). In the rest of Introduction, we shall explain our results on the coarse geometry of Croke-Kleiner admissible groups and their subgroups.

1.1. Proper actions on finite products of quasi-trees

A quasi-tree is a geodesic metric space quasi-isometric to a tree. Recently, Bestvina, Bromberg and Fujiwara [BBF19] introduced a (QT) property for a finitely generated group: GG acts properly on a finite product of quasi-trees so that the orbital map from GG with word metrics is a quasi-isometric embedding. This is a stronger property of the finite asymptotic dimension by recalling that a quasi-isometric embedding implies finite asdim of GG. It is known that Coxeter groups have property (QT) (see [DJ99]), and thus every right-angled Artin group has property (QT) (see Induction 2.2 in [BBF19]). Furthermore, the fundamental group of a compact special cube complex is undistorted in RAAGS (see [HW08]) and then has property (QT). As a consequence, many 3-manifold groups have property (QT), among which we wish to mention chargeless (including flip) graph manifolds [HP15] and finite volume hyperbolic 3-manifolds [Wis20]. In [BBF19], residually finite hyperbolic groups and mapping class groups are proven to have property (QT). It is natural to ask which other groups have property (QT) rather than these groups above.

The main result of this paper adds flip CKA actions into the list of groups which have property (QT). The notion of an omnipotent group is introduced by Wise in [Wis00] and has found many applications in subgroup separability. We refer the reader to Definition 5.10 for its definition and note here that free groups [Wis00], surfaces groups [Baj07], and the more general class of virtually special hyperbolic groups [Wis20] are omnipotent.

Theorem 1.1.

Let GXG\curvearrowright X be a flip admissible action where for every vertex group the central extension (1) has omnipotent hyperbolic quotient group. Then GG acts properly on a finite product of quasi-trees so that the orbital map is a quasi-isometric embedding.

Remark 1.2.

It is an open problem whether every hyperbolic group is residually finite. In [Wis00, Remark 3.4], Wise noted that if every hyperbolic group is residually finite, then any hyperbolic group is omnipotent.

Remark 1.3.

As a corollary, Theorem 1.1 gives another proof that flip graph manifold groups have property (QT). This was indeed one of motivations of this study (without noticing [HP15]).

In [HS13], Hume-Sisto prove that the universal cover of any flip graph manifold is quasi-isometrically embedded in the product of three metric trees. However, it does not follow from their proof that the fundamental group of a flip graph manifold has property (QT).

We now give an outline of the proof of Theorem 1.1 and explain some intermediate results, which we believe are of independent interest.

Proposition 1.4.

Let GXG\curvearrowright X be a flip CKA action. Then there exists a quasi-isometric embedding from XX to a product 𝒳1×𝒳2\mathcal{X}_{1}\times\mathcal{X}_{2} of two hyperbolic spaces.

If Gv=Hv×Z(Gv)G_{v}=H_{v}\times Z(G_{v}) for every vertex vT0v\in T^{0} and Ge=Z(Ge)×Z(Ge+)G_{e}=Z(G_{e_{-}})\times Z(G_{e_{+}}) for every edge eT1e\in T^{1}, then there exists a subgroup G˙<G\dot{G}<G of finite index at most 22 such that the above Q.I. embedding is G˙\dot{G}-equivariant.

Let us describe briefly the construction of 𝒳{𝒳1,𝒳2}\mathcal{X}\in\{\mathcal{X}_{1},\mathcal{X}_{2}\}. By Bass-Serre theory, GG acts on the Bass-Serre tree TT with vertex groups GvG_{v} and edge groups GeG_{e}. Let 𝒱\mathcal{V} be one of the two sets of vertices in TT with pairwise even distance. Note that GvG_{v} is the central extension of a hyperbolic group HvH_{v} by \mathbb{Z}, so acts geometrically on a metric product Yv=Y¯v×Y_{v}=\overline{Y}_{v}\times\mathbb{R} where HvH_{v} acts geometrically on Y¯v\overline{Y}_{v} and Z(Gv)Z(G_{v}) acts by translation on \mathbb{R}-lines. Roughly, the space 𝒳\mathcal{X} is obtained by isometric gluing of the boundary lines of Y¯v\overline{Y}_{v}’s over vertices vv in the link of every wT0𝒱w\in T^{0}-\mathcal{V}. In proving Proposition 1.4, the main tool is the construction of a class of quasi-geodesic paths called special paths between any two points in XX. See Section 3 for the details and related discussion after Theorem 1.6 below.

To endow an action on 𝒳\mathcal{X}, we pass to an index at most 2 subgroup G˙\dot{G} preserving 𝒱\mathcal{V} and the stabilizer in G˙\dot{G} of v𝒱v\in\mathcal{V} is GvG_{v} by Lemma 4.6. Under the assumptions on GvG_{v} and GeG_{e}’s, G˙\dot{G} acts by isometry on 𝒳\mathcal{X} and the Q.I. embedding is G˙\dot{G}-equivariant.

To prove Theorem 1.1, we exploit the strategy as [BBF19] to produce a proper action on products of quasi-trees. By Lemma 4.17, we first produce enough quasi-lines

𝔸=v𝒱𝔸v\mathbb{A}=\cup_{v\in\mathcal{V}}\mathbb{A}_{v}

for the hyperbolic space 𝒳\mathcal{X} where 𝔸v\mathbb{A}_{v} is a HvH_{v}-finite set of quasi-lines in Y¯v\overline{Y}_{v} so that the so-called distance formula follows in Proposition 4.20. On the other hand, the “crowd” quasi-lines in 𝔸\mathbb{A} may fail to satisfy the projection axioms in [BBF15] with projection constant required from the distance formula. Thus, we have to partition 𝔸\mathbb{A} into finite sub-collections of sparse quasi-lines: γγ\forall\gamma\neq\gamma^{\prime}, d(γ,γ)>θd(\gamma,\gamma^{\prime})>\theta for a uniform constant θ\theta.

Using local finiteness, we can partition quasi-lines without respecting group action and prove the following general result. This generalize the results of Hume-Sisto in [HS13] to flip CKA actions.

Theorem 1.5.

Let GXG\curvearrowright X be a flip CKA action. Then GG is quasi-isometric embedded into a finite product of quasi-trees.

However, the difficulty in establishing property (QT) is to partition all quasi-lines 𝔸=i=1n𝔸i\mathbb{A}=\cup_{i=1}^{n}\mathbb{A}_{i} so that each 𝔸i\mathbb{A}_{i} is G˙\dot{G}-invariant and sparse. In §5.1, we cone off the boundary lines of YvY_{v}’s so that 𝔸v\mathbb{A}_{v}’s from different pieces YvY_{v} are “isolated”. The gives the coned-off space 𝒳˙\dot{\mathcal{X}} with new distance formula in Proposition 5.9 so that 𝒳\mathcal{X} is quasi-isometric embedded into the product of 𝒳˙\dot{\mathcal{X}} with a quasi-tree from the boundary lines. See Proposition 5.7.

The goal is then to find a finite index subgroup G¨<G˙<G\ddot{G}<\dot{G}<G so that each orbit in 𝔸\mathbb{A} is sparse. This is done in the following two steps:

By residual finiteness of HvH_{v}, we first find a finite index subgroup Kv<HvK_{v}<H_{v} whose orbit in 𝔸v\mathbb{A}_{v} is sparse. This follows the same argument in [BBF19]. Secondly, we need to reassemble those finite index subgroups KvK_{v} as a finite index group G¨\ddot{G} so that the orbit in 𝔸\mathbb{A} is sparse. This step uses crucially the omnipotence, with details given in §5.4. The projection axioms thus fulfilled for each G¨\ddot{G}-orbit produce a finite product of actions on quasi-trees, and finally, the distance formula finishes the proof of Theorem 1.1.

1.2. Strongly quasi-convex subgroups

The height of a finitely generated subgroup HH in a finitely generated group GG is the maximal nn\in\mathbb{N} such that there are distinct cosets g1H,,gnHG/Hg_{1}H,\dots,g_{n}H\in G/H such that the intersection g1Hg11gnHgn1g_{1}Hg_{1}^{-1}\cap\dots\cap g_{n}Hg_{n}^{-1} is infinite. The subgroup HH is called strongly quasi-convex in GG if for any L1L\geq 1, C0C\geq 0 there exists R=R(L,C)R=R(L,C) such that every (L,C)(L,C)–quasi-geodesic in GG with endpoints in HH is contained in the RR–neighborhood of HH. We note that strong quasiconvexity does not depend on the choice of finite generating set of the ambient group and it agrees with quasiconvexity when the ambient group is hyperbolic. In [GMRS98], the authors prove that quasi-convex subgroups in hyperbolic groups have finite height. It is a long-standing question asked by Swarup that whether or not the converse is true (see Question 1.8 in [Bes]). Tran in [Tra19] generalizes the result of [GMRS98] by showing that strongly quasi-convex subgroups in any finitely generated group have finite height. It is natural to ask whether or not the converse is true in this setting (i.e, finite height implies strong quasiconvexity). If the converse is true, then we could characterize strongly quasi-convex subgroup of a finitely generated group purely in terms of group theoretic notions.

In [NTY], the authors prove that having finite height and strong quasiconvexity are equivalent for all finitely generated 33–manifold groups except the only ones containing the Sol command in its sphere-disk decomposition, and the graph manifold case was an essential case treated there. More precisely, Theorem 1.7 in [NTY] states that finitely generated subgroups of the fundamental group of a graph manifold are strongly quasi-convex if and only if they have finite height. The second main result of this paper is to generalize this result to Croke-Kleiner admissible action GXG\curvearrowright X.

Theorem 1.6.

Let GXG\curvearrowright X be a CKA action. Let KK be a nontrivial, finitely generated infinite index subgroup of GG. Then the following are equivalent.

  1. (1)

    KK is strongly quasi-convex.

  2. (2)

    KK has finite height in GG.

  3. (3)

    KK is virtually free and every infinite order elements are Morse.

  4. (4)

    Let GTG\curvearrowright T be the action of GG on the associated Bass-Serre tree. KK is virtually free and the action of KK on the tree TT induces a quasi-isometric embedding of KK into TT.

We prove Theorem 1.6 by showing that (1)(2)(3)(1)(\ref{thm1:item1})\Rightarrow(\ref{thm1:item2})\Rightarrow(\ref{thm1:item3})\Rightarrow(\ref{thm1:item1}) and (3)(4)(\ref{thm1:item3})\Leftrightarrow(\ref{thm1:item4}). Similarly as in [NTY], the heart part of Theorem 1.6 is the implication (3)(1)(\ref{thm1:item3})\Rightarrow(\ref{thm1:item1}). We briefly review ideas in the proof of Theorem 1.7 in [NTY]. Suppose that KK is a finitely generated finite height subgroup of π1(M)\pi_{1}(M) where MM is a graph manifold. Let MKMM_{K}\to M be the covering space of MM corresponding to KK. The authors in [NTY] prove that KK is strongly quasi-convex in π1(M)\pi_{1}(M) by using Sisto’s notion of path system 𝒫𝒮(M~)\mathcal{PS}(\tilde{M}) in the universal cover M~\tilde{M} of MM, and prove that the preimage of the Scott core of MKM_{K} in M~\tilde{M} is 𝒫𝒮(M~)\mathcal{PS}(\tilde{M})–contracting in the sense of Sisto. In this paper, the strategy of the proof of Theorem 1.6 is similar to the proof of Theorem 1.7 in [NTY] where we still use Sisto’s path system in XX but details are different. Sisto’s construction of special paths are carried out only in flip graph manifolds. Our construction of (X,𝒫𝒮(X))(X,\mathcal{PS}(X)) relies on the work of Croke-Kleiner [CK02] and applies to any admissible space XX (so any nonpostively curved graph manifold). We then construct a subspace CKXC_{K}\subset X on which KK acts geometrically and show that CKC_{K} is contracting in XX with respect to the path system (X,𝒫𝒮(X))(X,\mathcal{PS}(X)). As a consequence, KK is strongly quasi-convex in GG.

To conclude the introduction, we list a few questions and problems.

Quasi-isometric classification of graph manifolds has been studied by Kapovich-Leeb [KL98] and a complete quasi-isometric classification for fundamental groups of graph manifolds is given by Behrstock-Neumann [BN08]. Kapovich-Leeb prove that for any graph manifold MM, there exists a flip graph manifold NN such that their fundamental groups are quasi-isometric. We would like to know that whether or not such a result holds for admissible groups.

Question 1.7.

Let GG be an admissible group such that each vertex group is the central extension of a omnipotent hyperbolic CAT(0) group by \mathbb{Z}. Does there exist flip CKA action GXG^{\prime}\curvearrowright X so that GG and GG^{\prime} are quasi-isometric?

Question 1.8 (Quasi-isometry rigidity).

Let GXG\curvearrowright X be a flip CKA action, and QQ be a finitely generated group which is quasi-isometric to GG. Does there exist a finite index subgroup Q<QQ^{\prime}<Q such that QQ^{\prime} is a flip CKA group?

With a positive answer to the above questions, we hope one can try to follow the strategy described in [BN08] to attack the following.

Problem 1.9.

Under the assumption of Theorem 1.1, give a quasi-isometric classification of admissible actions.

In [Liu13], Liu showed that the fundamental group of a non-positively curved graph manifold MM is virtually special (the case M\partial M\neq\varnothing was also obtained independently by Przytycki–Wise [PW14]). Thus, it is natural to ask the following.

Question 1.10.

Let GXG\curvearrowright X be a CKA action where vertex groups are the central extension of a virtually special hyperbolic group by the integer group. Is GG virtually special?

As above, a positive answer to the question (with virtual compact specialness) would give an other proof of Theorem 1.1 under the same assumption.

Overview

In Section 2, we review some concepts and results about Croke-Kleiner admissible groups. In Section 3, we construct special paths in admissible spaces and give some results that will be used in the later sections. The proof of Theorem 1.5 and Proposition 1.4 is given in Section 4. We prove Theorem 1.1 and Theorem 1.6 in Section 5 and Section 6 respectively.

Acknowledgments

We would like to thank Chris Hruska, Hongbin Sun and Dani Wise for helpful conversations. W. Y. is supported by the National Natural Science Foundation of China (No. 11771022).

2. Preliminary

Admissible groups firstly introduced in [CK02]. This is a particular class of graph of groups that includes fundamental groups of 33–dimensional graph manifolds (i.e, compact 33–manifolds are obtained by gluing some circle bundles). In this section, we review admissible groups and their properties that will used throughout in this paper.

Definition 2.1.

A graph of group 𝒢\mathcal{G} is admissible if

  1. (1)

    𝒢\mathcal{G} is a finite graph with at least one edge.

  2. (2)

    Each vertex group Gv{G}_{v} has center Z(Gv)Z({G}_{v})\cong\mathbb{Z}, Hv:=Gv/Z(Gv){H}_{v}\colon={G}_{v}/Z({G}_{v}) is a non-elementary hyperbolic group, and every edge subgroup Ge{G}_{e} is isomorphic to 2\mathbb{Z}^{2}.

  3. (3)

    Let e1e_{1} and e2e_{2} be distinct directed edges entering a vertex vv, and for i=1,2i=1,2, let KiGvK_{i}\subset{G}_{v} be the image of the edge homomorphism GeiGv{G}_{e_{i}}\to{G}_{v}. Then for every gG¯vg\in{\overline{G}}_{v}, gK1g1gK_{1}g^{-1} is not commensurable with K2K_{2}, and for every gGvKig\in G_{v}-K_{i}, gKig1gK_{i}g^{-1} is not commensurable with KiK_{i}.

  4. (4)

    For every edge group Ge{G}_{e}, if αi:GeGvi\alpha_{i}\colon{G}_{e}\to{G}_{v_{i}} are the edge monomorphism, then the subgroup generated by α11(Z(Gv1))\alpha_{1}^{-1}(Z({G}_{v_{1}})) and α21(Z(Gv1))\alpha_{2}^{-1}(Z({G}_{v_{1}})) has finite index in Ge{G}_{e}.

A group GG is admissible if it is the fundamental group of an admissible graph of groups.

Definition 2.2.

We say that the action GXG\curvearrowright X is an Croke-Kleiner admissible (CKA) if GG is an admissible group, and XX is a Hadamard space, and the action is geometrically (i.e, properly and cocompactly by isometries)

Examples of admissible actions:

  1. (1)

    Let MM be a nongeometric graph manifold that admits a nonpositively curve metric. Lift this metric to the universal cover M~\tilde{M} of MM, and we denote this metric by dd. Then the action π1(M)(M~,d)\pi_{1}(M)\curvearrowright(\tilde{M},d) is a CKA action.

  2. (2)

    Let TT be the torus complexes constructed in [CK00]. Then π1(T)T~\pi_{1}(T)\curvearrowright\tilde{T} is a CKA action.

  3. (3)

    Let H1H_{1} and H2H_{2} be two torsion-free hyperbolic groups such that they act geometrically on CAT(0)CAT(0) spaces X1X_{1} and X2X_{2} respectively. Let Gi=Hi×G_{i}=H_{i}\times\mathbb{Z} (with i=1,2i=1,2), then GiG_{i} acts geometrically on the CAT(0)CAT(0) space Yi=Xi×Y_{i}=X_{i}\times\mathbb{R}. A primitive hyperbolic element in HiH_{i} gives a totally geodesic torus TiT_{i} in the quotient space Yi/GiY_{i}/G_{i}. Choose a basis on each torus TiT_{i}. Let f:T1T2f\colon T_{1}\to T_{2} be a flip map. Let MM be the space obtained by gluing Y1Y_{1} to Y2Y_{2} along the homemorphism ff. We note that there exists a metric on MM such that with respect to this metric, MM is a locally CAT(0)CAT(0) space. Then GM~G\curvearrowright\tilde{M} is a CKA action.

Let GXG\curvearrowright X be an admissible action, and let GTG\curvearrowright T be the action of GG on the associated Bass-Serre tree. Let T0=Vertex(T)T^{0}=Vertex(T) and T1=Edge(T)T^{1}=Edge(T) be the vertex and edge sets of TT. For each σT0T1\sigma\in T^{0}\cup T^{1}, we let GσGG_{\sigma}\leq G be the stabilizer of σ\sigma. For each vertex vT0v\in T^{0}, let Yv:=Minset(Z(Gv)):=gZ(Gv)Minset(g)Y_{v}:=Minset(Z(G_{v})):=\cap_{g\in Z(G_{v})}Minset(g) and for every edge eEe\in E we let Ye:=Minset(Z(Ge)):=gZ(Ge)Minset(g)Y_{e}:=Minset(Z(G_{e})):=\cap_{g\in Z(G_{e})}Minset(g). We note that the assignments vYvv\to Y_{v} and eYee\to Y_{e} are GG–equivariant with respect to the natural GG actions.

The following lemma is well-known.

Lemma 2.3.

If H=kH=\mathbb{Z}^{k} for some k1k\geq 1 then Minset(H)=hHMinset(h)Minset(H)=\cap_{h\in H}Minset(h) splits isometrically as a metric product C×𝔼kC\times\mathbb{E}^{k} so that HH acts trivially on CC and as a translation lattice on 𝔼k\mathbb{E}^{k}. Moreover, Z(H,G)Z(H,G) acts cocompactly on C×𝔼kC\times\mathbb{E}^{k}.

As a corollary, we have

  1. (1)

    GvG_{v} acts co-compactly on Yv=Y¯v×Y_{v}=\overline{Y}_{v}\times\mathbb{R} and Z(Gv)Z(G_{v}) acts by translation on the \mathbb{R}–factor and trivially on Y¯v\overline{Y}_{v} where Y¯v\overline{Y}_{v} is a Hadamard space.

  2. (2)

    Ge=2G_{e}=\mathbb{Z}^{2} acts co-compactly on Ye=Y¯e×2YvY_{e}=\overline{Y}_{e}\times\mathbb{R}^{2}\subset Y_{v} where Y¯e\overline{Y}_{e} is a compact Hadamard space.

  3. (3)

    if t1=Z(Gv1),t2=Z(Gv2)\langle t_{1}\rangle=Z(G_{v_{1}}),\langle t_{2}\rangle=Z(G_{v_{2}}) then t1,t2\langle t_{1},t_{2}\rangle generates a finite index subgroup of GeG_{e}.

We summarize results in Section 3.2 of [CK02] that will be used in this paper.

Lemma 2.4.

Let GXG\curvearrowright X be an CKA action. Then there exists a constant D>0D>0 such that the following holds.

  1. (1)

    vT0ND(Yv)=eT1ND(Ye)=X\cup_{v\in T^{0}}{N}_{D}(Y_{v})=\cup_{e\in T^{1}}{N}_{D}(Y_{e})=X. We define Xv:=ND(Yv)X_{v}:={N}_{D}(Y_{v}) and Xe:=ND(Ye)X_{e}:={N}_{D}(Y_{e}) for all vT0v\in T^{0}, eT1e\in T^{1}.

  2. (2)

    If σ,σT0T1\sigma,\sigma^{\prime}\in T^{0}\cup T^{1} and XσXσX_{\sigma}\cap X_{\sigma^{\prime}}\neq\varnothing then dT(σ,σ)<Dd_{T}(\sigma,\sigma^{\prime})<D.

Strips in admissible spaces: (see Section 4.2 in [CK02]). We first choose, in a GG–equivariant way, a plane FeYeF_{e}\subset Y_{e} for each edge eT1e\in T^{1}. Then for every pair of adjacent edges e1e_{1}, e2e_{2}. we choose, again equivariantly, a minimal geodesic from Fe1F_{e_{1}} to Fe2F_{e_{2}}; by the convexity of Yv=Y¯v×Y_{v}=\overline{Y}_{v}\times\mathbb{R}, v:=e1e2v:=e_{1}\cap e_{2}, this geodesic determines a Euclidean strip 𝒮e1e2:=γe1e2×\mathcal{S}_{e_{1}e_{2}}:=\gamma_{e_{1}e_{2}}\times\mathbb{R} (possibly of width zero) for some geodesic segment γe1e2Y¯v\gamma_{e_{1}e_{2}}\subset\overline{Y}_{v}. Note that 𝒮e1e2Fei\mathcal{S}_{e_{1}e_{2}}\cap F_{e_{i}} is an axis of Z(Gv)Z(G_{v}). Hence if e1,e2,eEe_{1},e_{2},e\in E, eie=viVe_{i}\cap e=v_{i}\in V are distinct vertices, then the angle between the geodesics 𝒮e1eFe\mathcal{S}_{e_{1}e}\cap F_{e} and 𝒮e2eFe\mathcal{S}_{e_{2}e}\cap F_{e} is bounded away from zero.

Remark 2.5.
  1. (1)

    We note that it is possible that γe1,e2\gamma_{e_{1},e_{2}} is just a point. The lines Se1,e2Fe1S_{e_{1},e_{2}}\cap F_{e_{1}} and Se1,e2Fe2S_{e_{1},e_{2}}\cap F_{e_{2}} are axes of Z(Gv)Z(G_{v}).

  2. (2)

    There exists a uniform constant such that for any edge ee, the Hausdorff distance between two spaces FeF_{e} and XeX_{e} is no more than this constant.

Remark 2.6.

There exists a GG–equivariant coarse LL–Lipschitz map ρ:XT0\rho\colon X\to T^{0} such that xXρ(x)x\in X_{\rho(x)} for all xXx\in X. The map ρ\rho is called indexed map. We refer the reader to Section 3.3 in [CK02] for existence of such a map ρ\rho.

Definition 2.7 (Templates, [CK02]).

A template is a connected Hadamard space 𝒯\mathcal{T} obtained from disjoint collection of Euclidean planes {W}WWall𝒯\{W\}_{W\in Wall_{\mathcal{T}}} (called walls) and directed Euclidean strips {𝒮}𝒮Strip𝒯\{\mathcal{S}\}_{\mathcal{S}\in Strip_{\mathcal{T}}} (a direction for a strip 𝒮\mathcal{S} is a direction for its RR–factor 𝒮I×\mathcal{S}\simeq I\times\mathbb{R}) by isometric gluing subject to the following conditions.

  1. (1)

    The boundary geodesics of each strip 𝒮Strip𝒯\mathcal{S}\in Strip_{\mathcal{T}} , which we will refer to as singular geodesics, are glued isometrically to distinct walls in Wall𝒯Wall_{\mathcal{T}}.

  2. (2)

    Each wall WWall𝒯W\in Wall_{\mathcal{T}} is glued to at most two strips, and the gluing lines are not parallel.

Notations: We use the notion aKba\preceq_{K}b if the exists C=C(K)C=C(K) such that aCb+Ca\leq Cb+C, and we use the notion aKba\sim_{K}b if aKba\preceq_{K}b and bKab\preceq_{K}a. Also, when we write aKba\asymp_{K}b we mean that a/CbCaa/C\leq b\leq Ca.

Denote by Len1(γ)Len^{1}(\gamma) and Len(γ)Len(\gamma) the L1L^{1}–length and L2L^{2}–length of a path γ\gamma in a metric product space A×BA\times B. These two lengths are equal for a path if it is parallel to a factor; in general, they are bilipschitz.

3. Special paths in CKA action GXG\curvearrowright X

Let GXG\curvearrowright X be a CKA action. In this section, we are going to define special paths (see Definition 3.6) in XX that will be used on the latter sections. Roughly speaking, each special path in XX is a concatenation of geodesics in consecutive pieces YvY_{v}’s of XX and they are uniform quasi-geodesic in the sense that there exists a constant μ=μ(X)\mu=\mu(X) such that every special path is (μ,μ)(\mu,\mu)–quasi-geodesic.

We first introduce the class of special paths in a template which shall be mapped to special paths in XX up to a finite Hausdorff distance.

3.1. Special paths in a template

Definition 3.1.

Let 𝒯\mathcal{T} be the template given by Definition 2.7. A (connected) path γ\gamma in 𝒯\mathcal{T} is called special path if γ\gamma is a concatenation γ0γ1γn\gamma_{0}\gamma_{1}\cdots\gamma_{n} of geodesics γi\gamma_{i} such that each γi\gamma_{i} lies on the strip 𝒮i\mathcal{S}_{i} adjacent to WiW_{i} and Wi+1W_{i+1}.

Remark 3.2.

By the construction of the template, the endpoints of γi\gamma_{i} (1i<n)(1\leq i<n) must be the intersection points of singular geodesics on walls Wi,Wi+1W_{i},W_{i+1}.

We use Lemma 3.3 in the proof of Proposition 3.8.

Lemma 3.3.

Assume that the angles between the singular geodesics on walls are between β\beta and πβ\pi-\beta for a universal constant β(0,π)\beta\in(0,\pi). There exists a constant μ1\mu\geq 1 such that any special path is a (μ,0)(\mu,0)–quasi-geodesic.

Proof.

Let γ\gamma be a special path with endpoints x,yx,y. We are going to prove that Len(γ)μd(x,y)Len(\gamma)\leq\mu d(x,y) for a constant μ1\mu\geq 1. Since any subpath of a special path is special, this proves the conclusion.

Let α\alpha be the unique CAT(0)(0) geodesic between xx and yy. By the construction of the template, if α\alpha does not pass through the intersection point zWz_{W} of the singular geodesics on a wall WW, then it passes through a point xWx_{W} on one singular geodesic LL_{-} and then a point yWy_{W} on the other singular geodesic L+L_{+}. Recall that the angle between the singular geodesics on walls are uniformly between β\beta and πβ\pi-\beta. There exists a constant μ1\mu\geq 1 depending on β\beta only such that d(xW,zW)+d(yW,zW)μd(xW,yW)d(x_{W},z_{W})+d(y_{W},z_{W})\leq\mu d(x_{W},y_{W}). We thus replace [xW,yW][x_{W},y_{W}] by [xW,z][z,yW][x_{W},z][z,y_{W}] for every possible triangle Δ(xWyWzW)\Delta(x_{W}y_{W}z_{W}) on each wall WW. The resulted path then connects consecutively the points zWz_{W} on the walls WW in the order of their intersection with α\alpha, so it is the special path γ\gamma from xx to yy satisfying the following inequality

Len(γ)μd(x,y)Len(\gamma)\leq\mu d(x,y)

Thus, we proved that γ\gamma is a (μ,0)(\mu,0)–quasi-geodesic. ∎

We are going to define a template associated to a geodesic in the Bass-Serre tree as the following.

Definition 3.4 (Standard template associated to a geodesic γT\gamma\subset T).

Let γ\gamma be a geodesic segment in the Bass-Serre tree TT. We begin with a collection of walls WeW_{e} and an isometry ϕe:WeFe\phi_{e}\colon W_{e}\to F_{e} for each edge eγe\subset\gamma. For every pair ee, ee^{\prime} of adjacent edges of γ\gamma, we let 𝒮^e,e\mathcal{\hat{S}}_{e,e^{\prime}} be a strip which is isometric to 𝒮e,e\mathcal{S}_{e,e^{\prime}} if the width of 𝒮e,e\mathcal{S}_{e,e^{\prime}} is at least 11, and isometric to [0,1]×[0,1]\times\mathbb{R} otherwise; we let ϕe,e:𝒮^e,e𝒮e,e\phi_{e,e^{\prime}}\colon\mathcal{\hat{S}}_{e,e^{\prime}}\to\mathcal{S}_{e,e^{\prime}} be an affine map which respects product structure (ϕe,e\phi_{e,e^{\prime}} is an isometry if the width of 𝒮e,e\mathcal{S}_{e,e^{\prime}} is greater than or equal to 11 and compresses the interval otherwise). We construct 𝒯γ\mathcal{T}_{\gamma} by gluing the strips and walls so that the maps ϕe\phi_{e} and ϕe,e\phi_{e,e^{\prime}} descend to continuous maps on the quotient, we denote the map from 𝒯γX\mathcal{T}_{\gamma}\to X by ϕγ\phi_{\gamma}.

The following lemma is cited from Lemma 4.5 and Proposition 4.6 in [CK02].

Lemma 3.5.
  1. (1)

    There exists β=β(X)>0\beta=\beta(X)>0 such that the following holds. For any geodesic segment γT\gamma\in T, the angle function αγ:Wall𝒯o(0,π)\alpha_{\gamma}\colon Wall^{o}_{\mathcal{T}}\to(0,\pi) satisfies 0<βαγπβ<π0<\beta\leq\alpha_{\gamma}\leq\pi-\beta<\pi.

  2. (2)

    There are constants L,A>0L,A>0 such that the following holds. Let γ\gamma be a geodesic segment in TT, and let ϕγ:𝒯γX\phi_{\gamma}\colon\mathcal{T}_{\gamma}\to X be the map given by Definition 3.4. Then ϕγ\phi_{\gamma} is a (L,A)(L,A)–quasi-isometric embedding. Moreover, for any x,y[eγXe][e,eγ𝒮e,e]x,y\in[\cup_{e\subset\gamma}X_{e}]\cup[\cup_{e,e^{\prime}\subset\gamma}\mathcal{S}_{e,e^{\prime}}], there exists a continuous map α:[x,y]𝒯\alpha\colon[x,y]\to\mathcal{T} such that d(ϕγα,id|[x,y])Ld(\phi_{\gamma}\circ\alpha,id|_{[x,y]})\leq L.

3.2. Special paths in the admissible space XX

In this subsection, we are going to define special paths in XX.

Recall that we choose a GG–equivariant family of Euclidean planes {Fe:FeYe}eT1\{F_{e}:F_{e}\subset Y_{e}\}_{e\in T^{1}}. For every pair of planes (Fe,Fe)(F_{e},F_{e^{\prime}}) so that v=eev=e\cap e^{\prime}, a minimal geodesic between Fe,FeF_{e},F_{e^{\prime}} in YvY_{v} determines a strip 𝒮ee=γee×\mathcal{S}_{ee^{\prime}}=\gamma_{ee^{\prime}}\times\mathbb{R} for some geodesic γeeY¯v\gamma_{ee^{\prime}}\subset\overline{Y}_{v}. It is possible that γee\gamma_{ee^{\prime}} is trivial so the width of the strip is zero. Let xXvx\in X_{v} and ee an edge with an endpoint vv. The minimal geodesic from xx to FeF_{e} (possibly not belong to YvY_{v}) also define a strip 𝒮xe=γxe×\mathcal{S}_{xe}=\gamma_{xe}\times\mathbb{R} where the geodesic γxeY¯v\gamma_{xe}\subset\overline{Y}_{v} is the projection to Y¯v\overline{Y}_{v} of the intersection of this minimal geodesic with YvY_{v}. Thus, xx is possibly not in the strip 𝒮xe\mathcal{S}_{xe} but within its DD-neighborhood by Lemma 2.4.

Definition 3.6 (Special paths in XX).

Let ρ:XT0\rho\colon X\to T^{0} be the indexed map given by Remark 2.6. Let xx and yy be two points in XX. If ρ(x)=ρ(y)\rho(x)=\rho(y) then we define a special path in XX connecting xx to yy is the geodesic [x,y][x,y]. Otherwise, let e1ene_{1}\cdots e_{n} be the geodesic edge path connecting ρ(x)\rho(x) to ρ(y)\rho(y) and let piFeip_{i}\in F_{e_{i}} be the intersection point of the strips 𝒮ei1ei\mathcal{S}_{e_{i-1}e_{i}} and 𝒮eiei+1\mathcal{S}_{e_{i}e_{i+1}}, where e0:=xe_{0}:=x and en+1:=ye_{n+1}:=y. The special path connecting xx to yy is the concatenation of the geodesics

[x,p1][p1,p2][pn1,pn][pn,y].[x,p_{1}][p_{1},p_{2}]\cdots[p_{{n-1}},p_{n}][p_{n},y].
Remark 3.7.

By definition, the special path except the [x,p1][x,p_{1}] and [pn,y][p_{n},y] depends only on the geodesic e1ene_{1}\cdots e_{n} in TT and the choice of planes FeF_{e}.

Refer to caption
Figure 1. Special path γ\gamma: the dotted and blue path from xx to yy

3.3. Special paths in the admissible space XX are uniform quasi-geodesic

In this section, we are going to prove the following proposition.

Proposition 3.8.

There exists a constant μ>0\mu>0 such that every special path γ\gamma in XX is a (μ,μ)(\mu,\mu)–quasi-geodesic.

To get into the proof of Proposition 3.8, we need several lemmas (see Lemma 3.9 and Lemma 3.10).

Lemma 3.9.

[CK02, Lemma 3.17] There exists a constant C>0C>0 with the following property. Let [x,y][x,y] be a geodesic in XX with ρ(x)ρ(y)\rho(x)\neq\rho(y) and e1ene_{1}\cdots e_{n} be the geodesic edge path connecting ρ(x)\rho(x) to ρ(y)\rho(y). Then there exists a sequence of points zi[x,y]NC(Fei)z_{i}\in[x,y]\cap N_{C}(F_{e_{i}}) such that d(x,zi)d(x,zj)d(x,z_{i})\leq d(x,z_{j}) for any iji\leq j.

Let [x,y][x,y] be a geodesic in Yv=Y¯v×Y_{v}=\overline{Y}_{v}\times\mathbb{R} with yFey\in F_{e}. We apply a minimizing horizontal slide of the endpoint yFey\in F_{e} to obtain a point zFez\in F_{e} so that [y,z][y,z] is parallel to Y¯v\overline{Y}_{v} and the projection of [x,z][x,z] on Y¯v\overline{Y}_{v} is orthogonal to FeY¯vF_{e}\cap\overline{Y}_{v}.

Lemma 3.10.

Let xXv0,yXvnx\in X_{v_{0}},y\in X_{v_{n}} where v0,vnv_{0},v_{n} are the endpoints of a geodesic e1ene_{1}\cdots e_{n} in TT. Then there exists a universal constant C>0C>0 depending on XX such that for each 1in1\leq i\leq n, we have

d(x,pi)+d(pi,y)Cd(x,y)+Cd(x,p_{i})+d(p_{i},y)\leq Cd(x,y)+C

where pi=𝒮ei1ei𝒮eiei+1p_{i}=\mathcal{S}_{e_{i-1}e_{i}}\cap\mathcal{S}_{e_{i}e_{i+1}} and e0:=xe_{0}:=x and en+1:=ye_{n+1}:=y.

Proof.

We use the notion aba\asymp b if there exists K=K(X)K=K(X) such that a/KbKaa/K\leq b\leq Ka.

Let DD be the constant given by Lemma 2.4 and satisfying Lemma 3.9 such that Xv=ND(Yv)X_{v}={N}_{D}(Y_{v}). Without loss of generality, we can assume xYv0,yYvnx\in Y_{v_{0}},y\in Y_{v_{n}}.

Denote e=eie=e_{i} and e=ei1,e+=ei+1e^{-}=e_{i-1},e^{+}=e_{i+1} in this proof. By Lemma 3.9, there exists a point qFeq\in F_{e} such that d(q,[x,y])Dd(q,[x,y])\leq D. For ease of computation, we will consider the mixed length xq1+d(q,y)||x-q||_{1}+d(q,y) of the path [x,q][q,y][x,q][q,y] which satisfies

(2) xq1+d(q,y)d(x,q)+d(q,y)d(x,y)+2D||x-q||_{1}+d(q,y)\asymp d(x,q)+d(q,y)\leq d(x,y)+2D

where ||||1||\cdot||_{1} is the L1L^{1}–metric on the metric product Yv=Y¯v×Y_{v}=\overline{Y}_{v^{-}}\times\mathbb{R}.

Note that the Euclidean plane FeYvYvF_{e}\subset Y_{v}\cap Y_{v^{\prime}} for e=vv¯e=\overline{vv^{\prime}} contains two non-parallel lines le:=𝒮eeFel_{e}^{-}:=\mathcal{S}_{e^{-}e}\cap F_{e} and le+:=𝒮ee+Fel_{e}^{+}:=\mathcal{S}_{ee^{+}}\cap F_{e}. So we can apply a minimizing horizontal slide of the endpoint qq of [x,q][x,q] in YvY_{v} to a point zz on lel_{e}^{-}. On the one hand, since the line lel_{e}^{-} on FeF_{e} is \mathbb{R}–factor of YeY_{e}, this slide decreases the L1L_{1}-distance xp1||x-p||_{1} by d(q,z)+Cd(q,z)+C for a constant CC depending on hyperbolicity constant of Y¯v\overline{Y}_{v}. On the other hand, by the triangle inequality, this slide increases d(q,y)d(q,y) by at most d(q,z)d(q,z). Hence, we obtain

|(xq1+d(q,y))(xz1+d(z,y))|C|(||x-q||_{1}+d(q,y))-(||x-z||_{1}+d(z,y))|\leq C

Similarly, by a minimizing horizontal slide of the endpoint zz of [z,y][z,y] in YeY_{e^{\prime}} to pp,

|(xp1+d(p,y))(xz1+d(z,y))|C|(||x-p||_{1}+d(p,y))-(||x-z||_{1}+d(z,y))|\leq C

yielding

|(xq1+d(q,y))(xp1+d(p,y))|2C|(||x-q||_{1}+d(q,y))-(||x-p||_{1}+d(p,y))|\leq 2C

Together with (2) this completes the proof of the lemma. ∎

Proof of Proposition 3.8.

Let γ\gamma be the special path from xx to yy for x,yXx,y\in X so that ρ(x)ρ(y)\rho(x)\neq\rho(y); otherwise it is a geodesic, and thus there is nothing to do. Let e1ene_{1}\cdots e_{n} be the geodesic in TT from ρ(x)\rho(x) to ρ(y)\rho(y). With notations as above (see Definition 3.6),

γ=[x,p1][p1,p2][pn1,pn][pn,y]\gamma=[x,p_{1}][p_{1},p_{2}]\cdots[p_{{n-1}},p_{n}][p_{n},y]

By Lemma 3.10, there exists a constant C1C\geq 1 such that

d(x,p1)+d(p1,pn)+d(pn,y)Cd(x,p1)+Cd(p1,y)+CC2d(x,y)+C2+C\begin{array}[]{rl}&d(x,p_{1})+d(p_{1},p_{n})+d(p_{n},y)\\ \leq&Cd(x,p_{1})+Cd(p_{1},y)+C\\ \leq&C^{2}d(x,y)+C^{2}+C\end{array}

Denoting α=[p1,p2][pn1,pn]\alpha=[p_{1},p_{2}]\cdots[p_{n-1},p_{n}], it remains to give a linear bound on (α)\ell(\alpha) in terms of d(p1,pn)d(p_{1},p_{n}).

By Lemma 3.5, there exists a KK–template (𝒯,f,ϕ)(\mathcal{T},f,\phi) for the e1ene_{1}\cdots e_{n} such that ϕ\phi is a (L,A)(L,A)–quasi-isometric map from the template 𝒯\mathcal{T} to the union of the planes {Fei:1in}\{F_{e_{i}}:1\leq i\leq n\} with the strips {𝒮ei1ei:1in}\{\mathcal{S}_{e_{i-1}e_{i}}:1\leq i\leq n\}. Moreover, ϕ\phi sends walls and strips of 𝒯\mathcal{T} to the KK–neighborhood of planes FeiF_{e_{i}} and strips 𝒮ei1ei\mathcal{S}_{e_{i-1}e_{i}} of 𝒯\mathcal{T} accordingly. Hence, ϕ\phi maps the intersection point on WeiW_{e_{i}} of the singular geodesics of two strips in 𝒯\mathcal{T} to a finite KK–neighborhood of pip_{i} (1in)(1\leq i\leq n). Since the map ϕ\phi is affine on strips and isometric on walls of 𝒯\mathcal{T}, we conclude that there exists a special path α~\tilde{\alpha} in 𝒯\mathcal{T} such that ϕ(α~)\phi(\tilde{\alpha}) is sent to a finite neighborhood of the special path α\alpha. Lemma 3.3 then implies that α~\tilde{\alpha} is a (C1,C1)(C_{1},C_{1})–quasi-geodesic for some C1>1C_{1}>1 so α\alpha is a (μ,μ)(\mu,\mu)–quasi-geodesic for some μ\mu depending on L,A,K,C1L,A,K,C_{1}. The proof is complete. ∎

4. Quasi-isometric embedding of admissible groups into product of trees

A quasi-tree is a geodesic metric space quasi-isometric to a tree. In this section, we are going to prove Theorem 1.5 that states if GXG\curvearrowright X is a flip CKA action (see Definition 4.1) then GG is quasi-isometric embedded into a finite product of quasi-trees. The strategy is that we first show that the space XX is quasi-isometric embedded into product of two hyperbolic spaces 𝒳1\mathcal{X}_{1}, 𝒳2\mathcal{X}_{2} (see Subsection 4.2). We then show that each hyperbolic space 𝒳i\mathcal{X}_{i} is quasi-isometric embedded into a finite product of quasi-trees (see Subsection 4.4).

4.1. Flip CKA actions and constructions of two hyperbolic spaces

Let GXG\curvearrowright X be a CKA action. Recall that each YvY_{v} decomposes as a metric product of a hyperbolic Hadamard space Y¯v\overline{Y}_{v} with the real line \mathbb{R} such that Y¯v\overline{Y}_{v} admits a geometric action of HvH_{v}. Recall that we choose a GG–equivariant family of Euclidean planes {Fe:FeYe}eT1\{F_{e}:F_{e}\subset Y_{e}\}_{e\in T^{1}}.

Definition 4.1 (Flip CKA action).

If for each edge e:=[v,w]T1e:=[v,w]\in T^{1}, the boundary line =Y¯vFe\ell=\overline{Y}_{v}\cap F_{e} is parallel to the \mathbb{R}–line in Yw=Y¯w×Y_{w}=\overline{Y}_{w}\times\mathbb{R}, then the CKA action is called flip in sense of Kapovich-Leeb.

Let 𝕃v\mathbb{L}_{v} be the set of boundary lines of Y¯v\overline{Y}_{v} which are intersections of Y¯v\overline{Y}_{v} with FeF_{e} for all edges ee issuing from vv. Thus, there is a canonical one-to-one correspondence between 𝕃v\mathbb{L}_{v} and the link of vv denoted by Lk(v)Lk(v).

Definition 4.2.

A flat link is the countable union of (closed) flat strips of width 1 glued along a common boundary line called the binding line.

Construction of hyperbolic spaces 𝒳1\mathcal{X}_{1} and 𝒳2\mathcal{X}_{2}: We first partition the vertex set T0T^{0} of the Bass-Serre tree into two disjoint class of vertices 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2} such that if vv and vv^{\prime} are in 𝒱i\mathcal{V}_{i} then dT(v,v)d_{T}(v,v^{\prime}) is even.

Given 𝒱{𝒱1,𝒱2}\mathcal{V}\in\{\mathcal{V}_{1},\mathcal{V}_{2}\}, we shall build a geodesic (non-proper) hyperbolic space 𝒳\mathcal{X} by glueing Y¯v\overline{Y}_{v} for all v𝒱v\in\mathcal{V} along the boundary lines via flat links.

Consider the set of vertices in 𝒱\mathcal{V} such that their pairwise distance in TT equals 2. Equivalently, it is the union of the links of every vertex wT0𝒱w\in T^{0}-\mathcal{V}. For any v1v2Lk(w)v_{1}\neq v_{2}\in Lk(w), the edges e1=[v1,w]e_{1}=[v_{1},w] and e2=[v2,w]e_{2}=[v_{2},w] determine two corresponding boundary lines 1𝕃(v1)\ell_{1}\in\mathbb{L}(v_{1}) and 2𝕃(v2)\ell_{2}\in\mathbb{L}(v_{2}) which are the intersections of Y¯vi\overline{Y}_{v_{i}} with FeiF_{e_{i}} for i=1,2i=1,2 respectively. There exists a canonical identification between 1\ell_{1} and 2\ell_{2} so that their \mathbb{R}–coordinates equal in the metric product Yw=Y¯w×Y_{w}=\overline{Y}_{w}\times\mathbb{R}.

Note the link Lk(w)Lk(w) determines a flat link Fl(w)Fl(w) so that the flat strips are one-to-one correspondence with Lk(w)Lk(w). In equivalent terms, it is a metric product Lk(w)×Lk(w)\times\mathbb{R}, where \mathbb{R} is parallel to the binding line.

For each wT0𝒱w\in T^{0}-\mathcal{V}, the set of hyperbolic spaces Y¯v\overline{Y}_{v} where vLk(w)v\in Lk(w) are glued to the flat links Fl(w)Fl(w) along the boundary lines of flat strips and of hyperbolic spaces with the identification indicated above. Therefore, we obtain a metric space 𝒳\mathcal{X} from the union of {Y¯v:v𝒱}\{\overline{Y}_{v}:v\in\mathcal{V}\} and flat links {Fl(w),wT0𝒱}\{Fl(w),w\in T^{0}-\mathcal{V}\}.

Remark 4.3.

By construction, Y¯v\overline{Y}_{v} and Y¯v\overline{Y}_{v^{\prime}} are disjoint in 𝒳\mathcal{X} for any two vertices v,v𝒱v,v^{\prime}\in\mathcal{V} with dT(v,v)>2d_{T}(v,v^{\prime})>2. Endowed with induced length metric, 𝒳\mathcal{X} is a hyperbolic geodesic space but not proper since each Y¯v\overline{Y}_{v} is glued via flat links with infinitely many Y¯v\overline{Y}_{v^{\prime}}’s where dT(v,v)=2d_{T}(v^{\prime},v)=2.

Definition 4.4.

Let gg be an element in GG. The translation length of gg is defined to be |g|:=infxTd(x,gx)|g|:=\inf_{x\in T}d(x,gx). Let Axis(g)={xT|d(x,gx)=|g|}\mathrm{Axis}(g)=\{x\in T|d(x,gx)=|g|\}. If Axis(g)\mathrm{Axis}(g)\neq\varnothing and |g|=0|g|=0 then gg is called elliptic. If Axis(g)\mathrm{Axis}(g)\neq\varnothing and |g|>0|g|>0, it is called loxodromic (or hyperbolic).

Remark 4.5.

We note that Axis(g)\mathrm{Axis}(g)\neq\varnothing for any gGg\in G. If gg is loxodromic, Axis(g)\mathrm{Axis}(g) is isometric to \mathbb{R}, and gg acts on Axis(g)\mathrm{Axis}(g) as translation by |g||g|.

Lemma 4.6.

There exists a subgroup G˙\dot{G} of index at most 2 in GG so that G˙\dot{G} preserves 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2} respectively and GvG˙G_{v}\subset\dot{G} for any vT0v\in T^{0}.

Proof.

Observe first that if dT(go,o)=0(mod2)d_{T}(go,o)=0\pmod{2} for some oT0o\in T^{0} and gGg\in G, then dT(gv,v)=0(mod2)d_{T}(gv,v)=0\pmod{2} holds for any vT0v\in T^{0}. Indeed, if gg is elliptic and thus rotates about a point oo, the geodesic [gv,v][gv,v] for any vv is contained in the union [o,v][o,gv][o,v]\cup[o,gv] and thus has even length. Otherwise, gg must be a hyperbolic element and leaves invariant a geodesic γ\gamma acted upon by translation. By a similar reasoning, if gg moves the points on γ\gamma with even distance, then dT(gv,v)=0(mod2)d_{T}(gv,v)=0\pmod{2} for any vT0v\in T^{0}.

Consider now the set G˙\dot{G} of elements gGg\in G such that dT(gv,v)=0(mod2)d_{T}(gv,v)=0\pmod{2} for any vT0v\in T^{0}. Using the tree TT again, if g,hG˙g,h\in\dot{G}, then dT(gv,hv)=0(mod2)d_{T}(gv,hv)=0\pmod{2} for any vT0v\in T^{0}. Thus, G˙\dot{G} is a group of finite index 2. ∎

Let G˙\dot{G} be the subgroup of GG given by the lemma. By Bass-Serre theory, it admits a finite graph of groups where the underlying graph 𝒢˙=T/G˙\dot{\mathcal{G}}=T/\dot{G} is bipartite with vertex sets V=𝒱/G˙V=\mathcal{V}/\dot{G} and W=𝒲/G˙W=\mathcal{W}/\dot{G} where 𝒲:=T0𝒱\mathcal{W}:=T^{0}-\mathcal{V}, and the vertex groups are isomorphic to those of GG.

Lemma 4.7.

The space 𝒳\mathcal{X} is a δ\delta–hyperbolic Hadamard space where δ>0\delta>0 only depends on the hyperbolicity constants of Y¯v\overline{Y}_{v} (v𝒱v\in\mathcal{V}).

If for every vT0v\in T^{0}, Gv=Hv×Z(Gv)G_{v}=H_{v}\times Z(G_{v}) and for each edge e:=[v,w]T1e:=[v,w]\in T^{1}, Ge=Z(Gv)×Z(Gw)G_{e}=Z(G_{v})\times Z(G_{w}) then the subgroup G˙<G\dot{G}<G given by Lemma 4.6 acts on 𝒳\mathcal{X} with the following properties:

  1. (1)

    for each v𝒱v\in\mathcal{V}, the stabilizer of Y¯v\overline{Y}_{v} is isomorphic to GvG_{v} and HvH_{v} acts geometrically on Y¯v\overline{Y}_{v}, and

  2. (2)

    for each w𝒲w\in\mathcal{W}, the flat link Fl(w)Fl(w) admits an isometric group action of GwG_{w} so that GwG_{w} acts by translation on the line parallel to the binding line and on the set of flat strips by the action on the link Lk(w)Lk(w).

Proof.

On one hand, GeG_{e} acts on the boundary line e=FeY¯v\ell_{e}=F_{e}\cap\overline{Y}_{v} through GeGe/Z(Gv)=Z(Gw)G_{e}\rightarrowtail G_{e}/Z(G_{v})=Z(G_{w}). On the other hand, Z(Gw)Z(G_{w}) acts on the boundary line of the flat strip corresponding to the edge ee. Since these two actions are compatible with glueing of YvY_{v}’s where vLk(w)v\in Lk(w), we can extend the actions on Y¯v\overline{Y}_{v}’s and flat links Fl(w)Fl(w)’s to get the desired action of G˙\dot{G} on 𝒳\mathcal{X}. ∎

4.2. Q.I. embedding into the product of two hyperbolic spaces

Proposition 4.8.

Let GXG\curvearrowright X be a flip CKA action and G˙\dot{G} the subgroup in GG of index at most 22 given by Lemma 4.6. Let 𝒳i\mathcal{X}_{i} (i=1,2i=1,2) be the hyperbolic space constructed in Section 4.1 with respect to 𝒱i\mathcal{V}_{i}. Then there exists a quasi-isometric embedding map ϕ\phi from XX to 𝒳1×𝒳2\mathcal{X}_{1}\times\mathcal{X}_{2}.

If for every vT0v\in T^{0}, Gv=Hv×Z(Gv)G_{v}=H_{v}\times Z(G_{v}) and for each edge e:=[v,w]T1e:=[v,w]\in T^{1}, Ge=Z(Gv)×Z(Gw)G_{e}=Z(G_{v})\times Z(G_{w}) then the above map ϕ\phi can be made G˙\dot{G}-equivariant.

Proof.

Let ρ:XT0\rho\colon X\to T^{0} be the indexed map given by Remark 2.6. Choose a vertex vT0v\in T^{0} and a point x0Yvx_{0}\in Y_{v} such that ρ(x0)=v\rho(x_{0})=v. Note that ρ\rho is a GG–equivariant, hence ρ(g(x0))=gρ(x0)=gv\rho(g(x_{0}))=g\rho(x_{0})=gv. Without loss of generality, we can assume that v𝒱1v\in\mathcal{V}_{1}. Let X~=G˙(x0)\tilde{X}=\dot{G}(x_{0}) be the orbit of x0x_{0} in XX.

For any xX~=G˙(x0)x\in\tilde{X}=\dot{G}(x_{0}) then x=g(x0)x=g(x_{0}) for some gG˙g\in\dot{G}. We remark here that in general it is possible that g(x0)g(x_{0}) belong to several YwY_{w}’s for some wT0w\in T^{0}. However, recall that we have the given indexed map ρ:XT0\rho\colon X\to T^{0}. This indexed function will tell us exactly which space we should project g(x0)g(x_{0}) into, i.e, g(x0)g(x_{0}) should project to Y¯ρ(g(x0))=Y¯gv\overline{Y}_{\rho(g(x_{0}))}=\overline{Y}_{gv}.

We recall that G˙\dot{G} has finite index in GG and it preserves 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2}.

Step 1 : Construct a quasi-isometric embedding map ϕ:X~𝒳1×𝒳2\phi\colon\tilde{X}\to\mathcal{X}_{1}\times\mathcal{X}_{2}

We are going to define the map ϕ=ϕ1×ϕ2:X~𝒳1×𝒳2\phi=\phi_{1}\times\phi_{2}:\tilde{X}\to\mathcal{X}_{1}\times\mathcal{X}_{2} where ϕi:X~𝒳i\phi_{i}:\tilde{X}\to\mathcal{X}_{i}. We first define a map ϕ1:X~𝒳1\phi_{1}:\tilde{X}\to\mathcal{X}_{1}.

For any xX~x\in\tilde{X} then x=g(x0)x=g(x_{0}) for some gG˙g\in\dot{G}, and thus g(x0)Ygvg(x_{0})\in Y_{gv}. Since we assume that v𝒱1v\in\mathcal{V}_{1} and G˙\dot{G} preserves 𝒱1\mathcal{V}_{1}, it follows that gv𝒱1gv\in\mathcal{V}_{1}. We define ϕ1(x):=πY¯gv(x)\phi_{1}(x):=\pi_{\overline{Y}_{gv}}(x) where πY¯gv\pi_{\overline{Y}_{gv}} is the projection of Ygv=Y¯gv×Y_{gv}=\overline{Y}_{gv}\times\mathbb{R} to the factor Y¯v\overline{Y}_{v}. We define ϕ2(x)\phi_{2}(x) to be the point on the binding line of the flat link Fl(v)Fl(v) so that its \mathbb{R}–coordinate is the same as that of xx in the metric product Ygv=Y¯gv×Y_{gv}=\overline{Y}_{gv}\times\mathbb{R}.

Step 2: Verifying ϕ\phi is a quasi-isometric embedding.

We are now going to show that ϕ=ϕ1×ϕ2:X~𝒳1×𝒳2\phi=\phi_{1}\times\phi_{2}\colon\tilde{X}\to\mathcal{X}_{1}\times\mathcal{X}_{2} is a quasi-isometric embedding. Before getting into the proof, we clarify here an observation that will be used later on.

Observation: By the tree-like construction of 𝒳\mathcal{X}, any geodesic α\alpha in 𝒳\mathcal{X} with endpoints αY¯v1,α+Y¯v2\alpha_{-}\in\overline{Y}_{v_{1}},\alpha_{+}\in\overline{Y}_{v_{2}} crosses Y¯v\overline{Y}_{v} for alternating vertices v[v1,v2]v\in[v_{1},v_{2}] in their order appearing in the interval, where [v1,v2][v_{1},v_{2}] is the geodesic in the tree TT. Using the convexity of boundary lines in a hyperbolic CAT(0) space, we see that the intersection αY¯v\alpha\cap\overline{Y}_{v} connects two boundary lines ,\ell,\ell^{\prime} in Y¯v\overline{Y}_{v} so that the projection π()\pi_{\ell}(\ell^{\prime}) is uniformly close to α\alpha. We can then construct a quasi-geodesic β\beta in 𝒳\mathcal{X} with the same endpoints as α\alpha so that βY¯v\beta\cap\overline{Y}_{v} connects π()\pi_{\ell}(\ell^{\prime}) to π()\pi_{\ell^{\prime}}(\ell).

Claim: There exists a constant C2C\geq 2 such that d(x,y)/CCd(ϕ(x),ϕ(y))Cd(x,y)+Dd(x,y)/C-C\leq d(\phi(x),\phi(y))\leq Cd(x,y)+D for any x,yX~x,y\in\tilde{X}.

Indeed, let gG˙g\in\dot{G} and gG˙g^{\prime}\in\dot{G} be two elements such that x=g(x0)x=g(x_{0}) and y=g(x0)y=g^{\prime}(x_{0}). Note that xYgvx\in Y_{gv} and yYgvy\in Y_{g^{\prime}v}. We consider the following cases.

Case 1: gv=gvgv=g^{\prime}v. In this case, it is easy to see the claim holds since

Len1[x,y]=Len1(ϕ1([x,y]))+Len1(ϕ2([x,y]))Len^{1}[x,y]=Len^{1}(\phi_{1}([x,y]))+Len^{1}(\phi_{2}([x,y]))

Case 2: gvgvgv\neq g^{\prime}v. Note that they both belong to 𝒱1\mathcal{V}_{1}, so dT(gv,gv)d_{T}(gv,g^{\prime}v) is even. Let 2n=dT(gv,gv)2n=d_{T}(gv,g^{\prime}v). We write

[gv,gv]=e1e2n[gv,g^{\prime}v]=e_{1}\cdots e_{2n}

as the edge paths in TT. Let vi1v_{i-1} be the initial vertex of eie_{i} with i=1,,2ni=1,\dots,2n and v2nv_{2n} be the terminal vertex of e2ne_{2n}. We note that v0,v2,,v2n𝒱1v_{0},v_{2},\dots,v_{2n}\in\mathcal{V}_{1} and v1,v3,,v2n1𝒱2v_{1},v_{3},\dots,v_{2n-1}\in\mathcal{V}_{2}.

With notations in in Definition 3.6, the special path γ\gamma between x,yx,y decomposes as the concatenation of geodesics:

γ=[x,p1][p1,p2][p2n,y].\gamma=[x,p_{1}][p_{1},p_{2}]\cdots[p_{2n},y].

Denote (x1,x2)=(ϕ1(x),ϕ2(x))𝒳1×𝒳2(x_{1},x_{2})=(\phi_{1}(x),\phi_{2}(x))\in\mathcal{X}_{1}\times\mathcal{X}_{2} and (y1,y2)=(ϕ1(y),ϕ2(y))𝒳1×𝒳2(y_{1},y_{2})=(\phi_{1}(y),\phi_{2}(y))\in\mathcal{X}_{1}\times\mathcal{X}_{2}. By the above observation, we connect x1x_{1} to y1y_{1} by a quasi-geodesic β1\beta_{1} in 𝒳1\mathcal{X}_{1} so that whenever β1\beta_{1} passes through Y¯v2,Y¯v4,,Y¯v2n2\overline{Y}_{v_{2}},\,\overline{Y}_{v_{4}},\,\cdots,\overline{Y}_{v_{2n-2}}, it is orthogonal to the boundary lines. In this way, we can write β1\beta_{1} as the concatenation of geodesic segments β10,β11,β12,,β12n\beta_{1}^{0},\beta_{1}^{1},\beta_{1}^{2},\cdots,\beta_{1}^{2n}, where β12i+1\beta_{1}^{2i+1} are maximal segments contained in the flat links. The first β10\beta_{1}^{0} and last β12n\beta_{1}^{2n} may have overlap with boundary lines, and the other β12i\beta_{1}^{2i} are orthogonal to the boundary lines of Y¯v2i\overline{Y}_{v_{2i}}.

Similarly, let β2\beta_{2} be a quasi-geodesic from x2x_{2} to y2y_{2} in 𝒳2\mathcal{X}_{2} as the concatenation of geodesic segments β20,β21,β22,,β22n\beta_{2}^{0},\beta_{2}^{1},\beta_{2}^{2},\cdots,\beta_{2}^{2n}.

We relabel xx by p0p_{0} and relabel yy by p2n+1p_{2n+1}. For each vertex viv_{i}, let πY¯i\pi_{\overline{Y}_{i}} and πi\pi_{\mathbb{R}_{i}} denote the projections of Yvi=Y¯vi×Y_{v_{i}}=\overline{Y}_{v_{i}}\times\mathbb{R} to the factor Y¯i\overline{Y}_{i} and \mathbb{R} respectively.

By the construction of ϕ1\phi_{1} and ϕ2\phi_{2} we note that there exists a constant A=A(𝒳)A=A(\mathcal{X}) such that

Len𝒳1(β1)Ai=0nLenX(πY¯2i[p2i,p2i+1])+i=0n1LenX(π2i+1[p2i+1,p2i+2])Len_{\mathcal{X}_{1}}(\beta_{1})\sim_{A}\,\,\,\sum_{i=0}^{n}Len_{X}\bigl{(}\pi_{\overline{Y}_{2i}}[p_{2i},p_{2i+1}]\bigr{)}+\sum_{i=0}^{n-1}Len_{X}\bigl{(}\pi_{\mathbb{R}_{2i+1}}[p_{2i+1},p_{2i+2}]\bigr{)}

and

Len𝒳2(β2)Ai=0nLenX(π2i[p2i,p2i+1])+i=0n1LenX(πY¯2i+1[p2i+1,p2i+2])Len_{\mathcal{X}_{2}}(\beta_{2})\sim_{A}\,\,\,\sum_{i=0}^{n}Len_{X}\bigl{(}\pi_{\mathbb{R}_{2i}}[p_{2i},p_{2i+1}]\bigr{)}+\sum_{i=0}^{n-1}Len_{X}\bigl{(}\pi_{\overline{Y}_{2i+1}}[p_{2i+1},p_{2i+2}]\bigr{)}

Summing over two equations above, we obtain

(3) d(x,y)B(Len𝒳1(β1)+Len𝒳2(β2))d(x,y)\sim_{B}\bigl{(}Len_{\mathcal{X}_{1}}(\beta_{1})+Len_{\mathcal{X}_{2}}(\beta_{2})\bigr{)}

for some constant B=B(A)B=B(A).

Since βi\beta_{i} is a (κ,κ)(\kappa,\kappa)–quasi-geodesic connecting two points ϕi(x)\phi_{i}(x) and ϕi(y)\phi_{i}(y) (for some uniform constant κ\kappa that does not depend on xx, yy), we have that (βi)κd(xi,yi)\ell(\beta_{i})\sim_{\kappa}d(x_{i},y_{i}) with i=1,2i=1,2. This fact together with formula (3) and the fact d(ϕ(x),ϕ(y))2d(x1,y1)+d(x2,y2)d(\phi(x),\phi(y))\asymp_{\sqrt{2}}d(x_{1},y_{1})+d(x_{2},y_{2}) give a constant c=C(B,κ)c=C(B,\kappa) such that d(x,y)Cd(ϕ(x),ϕ(y))d(x,y)\sim_{C}d(\phi(x),\phi(y)). The claim is verified. Therefore, ϕ\phi is a quasi-isometric embedding. ∎

4.3. Q.I. embedding into a finite product of trees

In this section, we first recall briefly the work of Bestvina-Bromberg-Fujiwara [BBF15] on constructing a quasi-tree of spaces. The reminder is then to produce a collection of quasi-lines to establish a distance formula for geometric actions of hyperbolic groups, see Lemma 4.17. The result is not new (cf. [BBF19, Prop. 3.3]), but the construction of quasi-lines is new and generalizes to certain non-proper actions, see Lemma 5.5.

We shall make use of the work of . Their theory applies to any collection of spaces 𝕐\mathbb{Y} equipped with a family of projection maps

{πY:𝕐{Y}×𝕐{Y}Y}Y𝕐\{\pi_{Y}:\mathbb{Y}-\{Y\}\times\mathbb{Y}-\{Y\}\to Y\}_{Y\in\mathbb{Y}}

satisfying the so-called projection axioms with projection constant ξ0\xi\geq 0. The precise formulation of projection axioms is irrelevant here. We only mention that their results apply to a collection of quasi-lines 𝔸\mathbb{A} with bounded projection property in a (not necessarily proper) hyperbolic space YY, where the projection maps πγ\pi_{\gamma} for γ𝔸\gamma\in\mathbb{A} are shortest point projections to γ\gamma in YY. Then (𝔸,πγ)(\mathbb{A},{\pi_{\gamma}}) satisfies projection axioms with projection constant ξ\xi (see [BBF19, Proposition 2.4]).

Fix K>0K>0. Following [BBF15], a quasi-tree of spaces 𝒞K(𝔸)\mathcal{C}_{K}(\mathbb{A}) is constructed with a underlying quasi-tree (graph) structure where every vertex represents a quasi-line in 𝔸\mathbb{A} and two quasi-lines γ,γ\gamma,\gamma^{\prime} are connected by an edge of length 1 from πγ(γ)\pi_{\gamma}(\gamma^{\prime}) to πγ(γ)\pi_{\gamma^{\prime}}(\gamma). If K>4ξK>4\xi, then 𝒞K(𝔸)\mathcal{C}_{K}(\mathbb{A}) is a unbounded quasi-tree.

If 𝔸\mathbb{A} admits a group action of GG so that πgγ=gπγ\pi_{g\gamma}=g\pi_{\gamma} for any gGg\in G and Y𝔸Y\in\mathbb{A}, then GG acts by isometry on 𝒞K(𝔸)\mathcal{C}_{K}(\mathbb{A}).

By [BBF15, Lemma 4.2, Corollary 4.10], every quasi-line γ𝔸\gamma\in\mathbb{A} with induced metric from YY is totally geodesically embedded in 𝒞K(𝔸)\mathcal{C}_{K}(\mathbb{A}) and the shortest projection maps from γ\gamma^{\prime} to γ\gamma in the quasi-tree 𝒞K(𝔸)\mathcal{C}_{K}(\mathbb{A}) coincides with the projection maps πγ(γ)\pi_{\gamma}(\gamma^{\prime}) up to uniform finite Hausdorff distance.

By abuse of language, for both x,yYx,y\in Y and x,y𝒞K(𝔸)x,y\in\mathcal{C}_{K}(\mathbb{A}), we denote

dγ(x,y)=diam(πγ({x,y})d_{\gamma}(x,y)=diam(\pi_{\gamma}(\{x,y\})

where the projections in the right-hand are understood in YY and 𝒞K(𝔸)\mathcal{C}_{K}(\mathbb{A}) accordingly. The above discussion implies that the two projections gives the same value up to a uniform bounded error.

Set [t]K=t[t]_{K}=t if tKt\geq K otherwise [t]K=0[t]_{K}=0. We now summarize what we need from [BBF15, BBF19] in the present paper.

Proposition 4.9.

Let 𝔸\mathbb{A} be a collection of quasi-lines in a δ\delta–hyperbolic space YY. If there is θ>0\theta>0 such that diam(πβ(α))θdiam(\pi_{\beta}(\alpha))\leq\theta for all αβ𝔸\alpha\neq\beta\in\mathbb{A}, then (𝔸,πγ)(\mathbb{A},{\pi_{\gamma}}) satisfies the projection axioms with projection constant ξ\xi depending only on θ\theta. Moreover, for any x,y𝒞K(𝔸)x,y\in\mathcal{C}_{K}(\mathbb{A}),

d𝒞K(𝔸)(x,y)Kγ𝔸[dγ(x,y)]Kd_{\mathcal{C}_{K}(\mathbb{A})}(x,y)\sim_{K}\sum_{\gamma\in\mathbb{A}}[d_{\gamma}(x,y)]_{K}

for all K4ξK\geq 4\xi.

Remark 4.10.

In [BBF19, Proposition 2.4], the above formula is stated for (𝔸,πγ)(\mathbb{A},{\pi_{\gamma}}) with the strong projection axioms. However, by [BBF19, Theorem 2.2], a modification πγ\pi_{\gamma}^{\prime} of projection maps πγ\pi_{\gamma} within finite Hausdorff distance can always be done so that (𝔸,πγ)(\mathbb{A},{\pi_{\gamma}^{\prime}}) satisfies the strong projection axioms. Thus, the same formula still holds with original projection maps.

As a corollary, the distance formula still works when the points x,yx,y are perturbed up to bounded error.

Corollary 4.11.

Under the assumption of Theorem 4.9, if d(x,x),d(y,y)Rd(x,x^{\prime}),d(y,y^{\prime})\leq R for some R>0R>0, then exists K0=K0(R,ξ,δ)K_{0}=K_{0}(R,\xi,\delta) such that

d𝒞K(𝔸)(x,y)Kγ𝔸[dγ(x,y)]Kd_{\mathcal{C}_{K}(\mathbb{A})}(x,y)\sim_{K}\sum_{\gamma\in\mathbb{A}}[d_{\gamma}(x^{\prime},y^{\prime})]_{K}

for all K2K0K\geq 2K_{0}.

Proof.

If d(x,x),d(y,y)Rd(x,x^{\prime}),d(y,y^{\prime})\leq R then there exists a constant K0=K0(R,ξ,δ)K_{0}=K_{0}(R,\xi,\delta) such that |dγ(x,y)dγ(x,y)|K0|d_{\gamma}(x,y)-d_{\gamma}(x^{\prime},y^{\prime})|\leq K_{0} for any γ𝔸\gamma\in\mathbb{A}. Assuming dγ(x,y)>K2K0d_{\gamma}(x,y)>K\geq 2K_{0} then dγ(x,y)K0d_{\gamma}(x^{\prime},y^{\prime})\geq K_{0} we see that

12[dγ(x,y)]K[dγ(x,y)]K2[dγ(x,y)]K{1\over 2}[d_{\gamma}(x^{\prime},y^{\prime})]_{K}\leq[d_{\gamma}(x,y)]_{K}\leq 2[d_{\gamma}(x^{\prime},y^{\prime})]_{K}

yielding the desired formula. ∎

Definition 4.12 (Acylindrical action).

[Bow08][Osi16] Let GG be a group acting by isometries on a metric space (X,d)(X,d). The action of GG on XX is called acylindrical if for any r0r\geq 0, there exist constants R,N0R,N\geq 0 such that for any pair a,bXa,b\in X with d(a,b)Rd(a,b)\geq R then we have

#{gG|d(ga,a)randd(gb,b)r}N\#\bigl{\{}g\in G\,|\,d(ga,a)\leq r\,\,\textup{and}\,\,d(gb,b)\leq r\bigr{\}}\leq N

By [Bow08], any nontrivial isometry of acylindrical group action on a hyperbolic space is either elliptic or loxodromic. A (λ,c)(\lambda,c)-quasi-geodesic γ\gamma for some λ,c>0\lambda,c>0 is referred to as a quasi-axis for a loxodromic element gg, if γ,gγ\gamma,g\gamma have (uniform) finite Hausdorff distance.

The following property in hyperbolic groups is probably known to experts, but is referred to a more general result [Yan19, Lemma 2.14] since we could not locate a precise statement as follows. A group is called non-elementary if it is neither finite nor virtually cyclic.

Lemma 4.13.

Let HH be a non-elementary group admitting a co-bounded and acylindrical action on a δ\delta–hyperbolic space (Y¯,d)(\overline{Y},d). Fix a basepoint oo. Then there exist a set FHF\subset H of three loxodromic elements and λ,c>0\lambda,c>0 with the following property.

For any hHh\in H there exists fFf\in F so that hfhf is a loxodromic element and the bi-infinite path

γ=i(hf)i([o,ho][ho,hfo])\gamma=\cup_{i\in\mathbb{Z}}(hf)^{i}([o,ho][ho,hfo])

is a (λ,c)(\lambda,c)–quasi-geodesic.

Sketch of the proof.

This follows from the result [Yan19, Lemma 2.14] which applies to any isometric action of HH on a metric space with a set FF of three pairwise independent contracting elements (loxodromic elements in hyperbolic spaces). If 𝕏\mathbb{X} denotes the set of GG–translated quasi-axis of all elements in FF, the pairwise independence condition is equivalent (defined) to be the bounded projection property of 𝕏\mathbb{X}. Thus, the existence of such FF is clear in a proper action of a non-elementary group. For acylindrical actions, this is also well-known, see [BBF19, Proposition 3.4] recalled below. ∎

Proposition 4.14.

[BBF19] Assume that a hyperbolic group HH acts acylindrically on a hyperbolic space Y¯\overline{Y}. For a loxodromic element gHg\in H, consider the set 𝔸\mathbb{A} of all HH-translates of a fixed (λ,c)(\lambda,c)-quasi-axis of gg for given λ,c>0\lambda,c>0. Then there exist θ=θ(λ,c)>0\theta=\theta(\lambda,c)>0 and N=N(λ,c)>0N=N(\lambda,c)>0 such that for any γ𝔸\gamma\in\mathbb{A}, the set

{hG:diam(πγ(hγ))θ}\{h\in G:diam(\pi_{\gamma}(h\gamma))\geq\theta\}

consists of at most NN double E(g)E(g)-cosets.

In particular, there are at most NN distinct pairs (γ,γ)𝔸×𝔸(\gamma,\gamma^{\prime})\in\mathbb{A}\times\mathbb{A} satisfying diam(πγ(γ))>θdiam(\pi_{\gamma}(\gamma^{\prime}))>\theta up to the action of HH.

The following corollary can be derived from the “in particular” statement using hyperbolicity.

Corollary 4.15.

Under the assumption of Proposition 4.14, for any R>0R>0, there exist constants θ=θ(λ,c,R),N=N(λ,c,R)>0\theta=\theta(\lambda,c,R),N=N(\lambda,c,R)>0 such that for any geodesic segment pp of length θ\theta, we have

{γ𝔸:pNR(γ)}N.\sharp\{\gamma\in\mathbb{A}:p\subset N_{R}(\gamma)\}\leq N.
Convention 4.16.

When speaking of quasi-lines in hyperbolic spaces with actions satisfying Lemma 4.13, we always mean (λ,c)(\lambda,c)–quasi-geodesics where λ,c>0\lambda,c>0 depend on FF and δ\delta.

Lemma 4.17.

Let HH be a non-elementary group admitting a proper and cocompact action on a δ\delta–hyperbolic space (Y¯,d)(\overline{Y},d). Assume that 𝕃\mathbb{L} is a HH–finite collection of quasi-lines. Then for any sufficiently large K>0K>0, there exist a HH–finite collection of quasi-lines 𝕃𝔸\mathbb{L}\subset\mathbb{A} in Y¯\overline{Y} and a constant N=N(K,δ,𝔸)>0N=N(K,\delta,\mathbb{A})>0, such that for any x,yY¯x,y\in\overline{Y}, the following holds

1Nγ𝔸[dγ(x,y)]Kd(x,y)2γ𝔸[dγ(x,y)]K+2K.\frac{1}{N}\sum_{\gamma\in\mathbb{A}}[d_{\gamma}(x,y)]_{K}\leq d(x,y)\leq 2\sum_{\gamma\in\mathbb{A}}[d_{\gamma}(x,y)]_{K}+2K.
Remark 4.18.

The statement of Lemma 4.17 (with torsion allowed here) is a re-package of Proposition 3.3 and Theorem 3.5 in [BBF19]. Our proof is different and generalizes to certain co-bounded and acylindrical actions coming from relatively hyperbolic groups on their relative Cayley graphs. See Lemma 5.5 for details.

Proof.

Fixing a point oY¯o\in\overline{Y}, the co-bounded action of HH on (Y¯,d)(\overline{Y},d) gives a constant R>0R>0 such that NR(Ho)=Y¯N_{R}(Ho)=\overline{Y}. By hyperbolicity, if γ\gamma is a (λ,c)(\lambda,c)–quasi-geodesic, then there exists a constant C=C(λ,c,R)>0C=C(\lambda,c,R)>0 such that diam([x,y]NR(γ))>Cdiam([x,y]\cap N_{R}(\gamma))>C implies

(4) |dγ(x,y)diam([x,y]NR(γ))|C.|d_{\gamma}(x,y)-diam([x,y]\cap N_{R}(\gamma))|\leq C.

Let θ=θ(R),N0=N0(R)\theta=\theta(R),N_{0}=N_{0}(R) be the constant given by Corollary 4.15 for the geometric (so acylindrical) action of HH on Y¯\overline{Y}.

Fix K>2C+2θK>2C+2\theta and denote K~=K+C\tilde{K}=K+C. Since the action is proper, the set S={hH:|d(o,ho)K~|2R}S=\{h\in H:|d(o,ho)-\tilde{K}|\leq 2R\} is finite. Let us consider the set S~\tilde{S} of loxodromic elements hfhf where hSh\in S and fFf\in F is provided by Lemma 4.13. Note that S~=S\sharp\tilde{S}=\sharp S. Let 𝔸\mathbb{A} be the set of all HH–translated axis of hfS~hf\in\tilde{S}. It is possible that 𝔸/HS~\sharp\mathbb{A}/H\leq\sharp\tilde{S} since two elements in S~\tilde{S} may be conjugate.

Assume that d(x,y)>K~d(x,y)>\tilde{K}. Consider a geodesic α\alpha from xx to yy and choose points xix_{i} on α\alpha for 0in+10\leq i\leq n+1 such that d(xi,xi+1)=K~d(x_{i},x_{i+1})=\tilde{K} for 0in10\leq i\leq n-1 and d(xn,xn+1)K~d(x_{n},x_{n+1})\leq\tilde{K} where x0=x,xn+1=yx_{0}=x,x_{n+1}=y. Since NR(Ho)=Y¯N_{R}(Ho)=\overline{Y}, there exists hiHh_{i}\in H so that d(xi,hio)Rd(x_{i},h_{i}o)\leq R. It implies that K~2Rd(o,hi1hi+1o)K~+2R\tilde{K}-2R\leq d(o,h_{i}^{-1}h_{i+1}o)\leq\tilde{K}+2R, and thus we have hi1hi+1Sh_{i}^{-1}h_{i+1}\in S for 0in10\leq i\leq n-1. Noting that [hio,hi+1o][h_{i}o,h_{i+1}o] is contained in a HH-translated axis of some loxodromic element in S~\tilde{S}, we thus obtain nn axis γ0,,γn1𝔸\gamma_{0},\cdots,\gamma_{n-1}\in\mathbb{A} (with possible multiplicities: γi=γj\gamma_{i}=\gamma_{j} for iji\neq j) satisfying diam(NR(γi)α)K~diam(N_{R}(\gamma_{i})\cap\alpha)\geq\tilde{K} so that

α[xn,xn+1]0in1NR(γi)α\alpha-[x_{n},x_{n+1}]\subset\bigcup_{0\leq i\leq n-1}N_{R}(\gamma_{i})\cap\alpha

which yields

Len(α)0in1diam(NR(γi)α)+K~Len(\alpha)\leq\sum_{0\leq i\leq n-1}diam(N_{R}(\gamma_{i})\cap\alpha)+\tilde{K}

where the constant K~\tilde{K} bounds the length of the last segment [xn,xn+1][x_{n},x_{n+1}].

By Equation (4), dγi(x,y)diam(NR(γi)α)CK>2Cd_{\gamma_{i}}(x,y)\geq diam(N_{R}(\gamma_{i})\cap\alpha)-C\geq K>2C and then diam(NR(γi)α)dγi(x,y)+C2dγi(x,y)diam(N_{R}(\gamma_{i})\cap\alpha)\leq d_{\gamma_{i}}(x,y)+C\leq 2d_{\gamma_{i}}(x,y). Thus, we obtain

Len(α)0in12[dγi(x,y)]K+K~,Len(\alpha)\leq\sum_{0\leq i\leq n-1}2[d_{\gamma_{i}}(x,y)]_{K}+\tilde{K},

which implies the upper bound over γ𝔸\gamma\in\mathbb{A}. Of course, the upper bound holds as well after adjoining 𝕃\mathbb{L} into 𝔸\mathbb{A}.

The remainder of the proof is to prove the lower bound. Let 𝔹\mathbb{B} be the set of quasi-lines γ𝔸𝕃\gamma\in\mathbb{A}\cup\mathbb{L} satisfying dγ(x,y)>K=2θ+2Cd_{\gamma}(x,y)>K=2\theta+2C. Note that the set of axis γ0,,γn1\gamma_{0},\cdots,\gamma_{n-1} obtained as above is included into 𝔹\mathbb{B}.

By the proper action of HH on Y¯\overline{Y}, the HH–finite 𝕃\mathbb{L} has bounded intersection so does 𝕃𝔸\mathbb{L}\cup\mathbb{A}. Thus, there is D=D(𝔸,𝕃,K~,R)>0D=D(\mathbb{A},\mathbb{L},\tilde{K},R)>0 so that diam(NR(γ)NR(γ))<Ddiam(N_{R}(\gamma)\cap N_{R}(\gamma^{\prime}))<D for any γγ𝔸𝕃\gamma\neq\gamma^{\prime}\in\mathbb{A}\cup\mathbb{L}. In particular, different NR(γ)αN_{R}(\gamma)\cap\alpha’s have overlaps bounded above by DD.

By Eq. (4), we obtain diam(NR(γ)α)dγ(x,y)C12dγ(x,y)θdiam(N_{R}(\gamma)\cap\alpha)\geq d_{\gamma}(x,y)-C\geq{1\over 2}d_{\gamma}(x,y)\geq\theta. As mentioned-above in the second paragraph, by Corollary 4.15, any segment of length θ\theta is covered at most N0N_{0} times by the RR-neighborhood of quasi-lines in 𝔹\mathbb{B}. Thus, there exists a constant N=N(D,N0)>0N=N(D,N_{0})>0 such that

NLen(α)γ𝔹diam(NR(γ)α)12γ𝔹[dγ(x,y)]K.N\cdot Len(\alpha)\geq\sum_{\gamma\in\mathbb{B}}diam(N_{R}(\gamma)\cap\alpha)\geq{1\over 2}\sum_{\gamma\in\mathbb{B}}[d_{\gamma}(x,y)]_{K}.

The proof is completed by renaming 𝔸:=𝔸𝕃.\mathbb{A}:=\mathbb{A}\cup\mathbb{L}.

4.4. Q.I. embedding into a finite product of trees

This subsection is devoted to the proof of Theorem 1.5. The results obtained here are not used in other places, and so can be skipped if the reader is interested in the stronger conclusion, the property (QT), under stronger assumption.

We start by explaining the choice of the constants and the collection of quasi-lines 𝔸\mathbb{A} in 𝒳1\mathcal{X}_{1} that will be used in the rest of this subsection.

The constants DD and θ\theta and ξ=ξ(θ)\xi=\xi(\theta): Let 𝒳1\mathcal{X}_{1} and 𝒳2\mathcal{X}_{2} be two δ\delta–hyperbolic spaces given by Lemma 4.7 where δ>0\delta>0 depends on the hyperbolicity constants of Y¯v\overline{Y}_{v}.

Note that each Y¯v\overline{Y}_{v} for v𝒱1v\in\mathcal{V}_{1} is isometrically embedded into 𝒳1\mathcal{X}_{1} and thus δ\delta–hyperbolic. We follow the Convention 4.16 on the quasi-lines which are (λ,c)(\lambda,c)–quasi-geodesics in Y¯v\overline{Y}_{v} and 𝒳1\mathcal{X}_{1}.

By the δ\delta–hyperbolicity of 𝒳1\mathcal{X}_{1}, there exist constants D,θ>0D,\theta>0 depending on δ\delta (and also λ,c\lambda,c) such that if any ((λ,c)(\lambda,c)–)quasi-lines αβ\alpha\neq\beta have a distance at least DD then diam(πβ(α))θdiam(\pi_{\beta}(\alpha))\leq\theta.

We then obtain the projection constant ξ=ξ(θ)\xi=\xi(\theta) by Proposition 4.9.

The collection of quasi-lines 𝔸\mathbb{A} in 𝒳1\mathcal{X}_{1}: Fix any sufficiently large number K>max{4ξ,θ,2}K>\max\{4\xi,\theta,2\} depending on 𝕃v\mathbb{L}_{v}, where 𝕃v\mathbb{L}_{v} is the collection of boundary lines of Y¯v\overline{Y}_{v}. By Lemma 4.17, there exist a locally finite collection of quasi-lines 𝕃v𝔸v\mathbb{L}_{v}\subset\mathbb{A}_{v} in Y¯v\overline{Y}_{v} and a constant N=N(K,𝔸v,δ)>0N=N(K,\mathbb{A}_{v},\delta)>0 such that

(5) 1Nγ𝔸v[dγ(x,y)]KdY¯v(x,y)2γ𝔸v[dγ(x,y)]K+2K\frac{1}{N}\sum_{\gamma\in\mathbb{A}_{v}}[d_{\gamma}(x,y)]_{K}\leq d_{\overline{Y}_{v}}(x,y)\leq 2\sum_{\gamma\in\mathbb{A}_{v}}[d_{\gamma}(x,y)]_{K}+2K

for any x,yY¯vx,y\in\overline{Y}_{v}. Since there are only finitely many G˙\dot{G}–orbits of (Hv,Y¯v)(H_{v},\overline{Y}_{v}) we assume furthermore 𝔸w=g𝔸v\mathbb{A}_{w}=g\mathbb{A}_{v} if w=gvw=gv for gG˙g\in\dot{G}. Then

𝔸:=v𝒱1𝔸v\mathbb{A}:=\cup_{v\in\mathcal{V}_{1}}\mathbb{A}_{v}

is a locally finite collection of quasi-lines in 𝒳1\mathcal{X}_{1}, preserved by the group G˙\dot{G}.

We use the following lemma in the proof of Proposition 4.20 that gives us a distance formula for 𝒳1\mathcal{X}_{1}.

Lemma 4.19.

There exists a constant L>0L>0 depending only on KK with the following properties.

  1. (1)

    For any γ𝔸\ell\neq\gamma\in\mathbb{A}, we have diam(π(γ))Ldiam(\pi_{\ell}(\gamma))\leq L.

  2. (2)

    For any v𝒱1v\in\mathcal{V}_{1} and x,yY¯vx,y\in\overline{Y}_{v}, there are at most LL quasi-lines γ\gamma in 𝔸𝔸v\mathbb{A}-\mathbb{A}_{v} such that Ldγ(x,y)KL\geq d_{\gamma}(x,y)\geq K.

Proof.

Since there are only finitely many Y¯w\overline{Y}_{w}’s up to isometry, and 𝔸w=g𝔸v\mathbb{A}_{w}=g\mathbb{A}_{v} if w=gvw=gv, the union 𝔸\mathbb{A} of quasi-lines containing w𝒱1𝕃w\cup_{w\in\mathcal{V}_{1}}\mathbb{L}_{w} is uniformly locally finite: any ball of a fixed radius in 𝒳1\mathcal{X}_{1} intersects a uniform number of quasi-lines depending only on the radius. By the hyperbolicity of 𝒳1\mathcal{X}_{1}, the local finiteness implies the bounded projection property, so gives the desired constant LL in the assertion (1).

By the construction of 𝒳1\mathcal{X}_{1}, the shortest projection of a point xY¯vx\in\overline{Y}_{v} to γ𝔸w\gamma\in\mathbb{A}_{w} for wvw\neq v has to pass through a boundary line 𝕃w\ell\in\mathbb{L}_{w} of Y¯w\overline{Y}_{w}, so is contained in the projection of \ell to γ\gamma. By the assertion (1) we have dγ(x,y)diam(πγ())Ld_{\gamma}(x,y)\leq diam(\pi_{\gamma}(\ell))\leq L. If dγ(x,y)Kθd_{\gamma}(x,y)\geq K\geq\theta for x,yY¯vx,y\in\overline{Y}_{v} and γ𝔸w\gamma\in\mathbb{A}_{w} with wvw\neq v, then d(γ,)Dd(\gamma,\ell)\leq D by the above defining property of DD and θ\theta. By local finiteness, there are at most L=L(D)L=L(D) quasi-lines \ell with this property, proving the assertion (2). ∎

Proposition 4.20 (Distance formula for 𝒳1\mathcal{X}_{1}).

For any x,y𝒳1x,y\in\mathcal{X}_{1}, there exists a constant μ=μ(L,K)>0\mu=\mu(L,K)>0 such that

(6) 1μγ𝔸[dγ(x,y)]K+dT(ρ(x),ρ(y))L2d𝒳1(x,y)μγ𝔸[dγ(x,y)]K+4KdT(ρ(x),ρ(y)).\begin{array}[]{cc}{1\over\mu}\sum_{\gamma\in\mathbb{A}}[d_{\gamma}(x,y)]_{K}+d_{T}(\rho(x),\rho(y))-L^{2}\\ \\ \leq d_{\mathcal{X}_{1}}(x,y)\leq\\ \\ \mu\sum_{\gamma\in\mathbb{A}}\par[d_{\gamma}(x,y)]_{K}+4K\cdot d_{T}(\rho(x),\rho(y)).\end{array}
Proof.

Since the 11–neighborhood of the union v𝒱1Y¯v\cup_{v\in\mathcal{V}_{1}}\overline{Y}_{v} is 𝒳1\mathcal{X}_{1}, assume for simplicity xY¯v1x\in\overline{Y}_{v_{1}} and yY¯vny\in\overline{Y}_{v_{n}} where v1=ρ(x),vn=ρ(y)𝒱1v_{1}=\rho(x),v_{n}=\rho(y)\in\mathcal{V}_{1}. By the construction of 𝒳1\mathcal{X}_{1}, a geodesic [x,y][x,y] travels through Y¯vi\overline{Y}_{v_{i}} and then flat links Fl(wi)Fl(w_{i}), where {vi𝒱1:1in}\{v_{i}\in\mathcal{V}_{1}:1\leq i\leq n\} and {wi𝒱2:1in1}\{w_{i}\in\mathcal{V}_{2}:1\leq i\leq n-1\} appear alternatively on [v1,vn]T[v_{1},v_{n}]\subset T. Thus, dT(v1,vn)=2n2d_{T}(v_{1},v_{n})=2n-2.

Let us denote the exit point on the boundary line i\ell_{i} of Y¯vi\overline{Y}_{v_{i}} and entry point on i+1\ell_{i+1}^{\prime} of Y¯vi+1\overline{Y}_{v_{i+1}} by yiy_{i} and xi+1x_{i+1} respectively for 1in1\leq i\leq n where x1:=xx_{1}:=x and yn:=yy_{n}:=y by convention. Thus,

(7) d𝒳1(x,y)1in1d𝒳1(yi,xi+1)=1indY¯vi(xi,yi).d_{\mathcal{X}_{1}}(x,y)-\sum_{1\leq i\leq n-1}d_{\mathcal{X}_{1}}(y_{i},x_{i+1})=\sum_{1\leq i\leq n}d_{\overline{Y}_{v_{i}}}(x_{i},y_{i}).

Therefore, we shall derive (6) from (7) which requires to apply the formula (5) for dY¯vi(xi,yi)d_{\overline{Y}_{v_{i}}}(x_{i},y_{i}). To that end, we need the following estimates. Recall that L,K\asymp_{L,K} means the equality holds up to a multiplicative constant depending on L,KL,K.

Claim 1.
  1. (1)

    If there is γ𝔸vi\gamma\in\mathbb{A}_{v_{i}} such that dγ(xi,yi)Kd_{\gamma}(x_{i},y_{i})\geq K then

    (8) [dγ(xi,yi)]KL,K[dγ(x,y)]K[d_{\gamma}(x_{i},y_{i})]_{K}\asymp_{L,K}[d_{\gamma}(x,y)]_{K}
  2. (2)

    d𝒳1(yi,xi+1)2d_{\mathcal{X}_{1}}(y_{i},x_{i+1})\geq 2. If d𝒳1(yi,xi+1)>K+2d_{\mathcal{X}_{1}}(y_{i},x_{i+1})>K+2, then

    [di(yi,xi+1)]KL,K[di(x,y)]K,di+1(yi,xi+1)L,K[di+1(x,y)]K[d_{\ell_{i}}(y_{i},x_{i+1})]_{K}\asymp_{L,K}[d_{\ell_{i}}(x,y)]_{K},\;\;d_{\ell_{i+1}^{\prime}}(y_{i},x_{i+1})\asymp_{L,K}[d_{\ell_{i+1}^{\prime}}(x,y)]_{K}
Proof of the Claim 1.

If there is γ𝔸vi\gamma\in\mathbb{A}_{v_{i}} such that dγ(xi,yi)Kd_{\gamma}(x_{i},y_{i})\geq K we then have

|dγ(x,y)dγ(xi,yi)|diam(πγ(i))+diam(πγ(i))2L,|d_{\gamma}(x,y)-d_{\gamma}(x_{i},y_{i})|\leq diam(\pi_{\gamma}(\ell_{i}))+diam(\pi_{\gamma}(\ell_{i}^{\prime}))\leq 2L,

where Lemma 4.19 is applied, and after taking the cutoff function []K[\cdot]_{K},

|[dγ(x,y)]K[dγ(xi,yi)]K|2L+K.|[d_{\gamma}(x,y)]_{K}-[d_{\gamma}(x_{i},y_{i})]_{K}|\leq 2L+K.

This in turn implies (8).

Recall that [yi,xi+1][y_{i},x_{i+1}] is contained in the union of two flat strips with width 11 in a flat link, and is from one boundary line i\ell_{i} to the other i+1\ell_{i+1}^{\prime}. Thus, d𝒳1(yi,xi+1)2d_{\mathcal{X}_{1}}(y_{i},x_{i+1})\geq 2. If d𝒳1(yi,xi+1)>K+2d_{\mathcal{X}_{1}}(y_{i},x_{i+1})>K+2 is assumed, then di(yi,xi+1)>Kd_{\ell_{i}}(y_{i},x_{i+1})>K and di+1(yi,xi+1)>Kd_{\ell_{i+1}^{\prime}}(y_{i},x_{i+1})>K. The assertion (2) follows similarly as above. ∎

Recalling K2K\geq 2, the assertion (2) of the Claim 1 implies a constant μ1=μ1(L,K)>1\mu_{1}=\mu_{1}(L,K)>1 such that

(9) dT(ρ(x),ρ(y))1in1d𝒳1(yi,xi+1)μ1𝔸[d(x,y)]K+2KdT(ρ(x),ρ(y)).d_{T}(\rho(x),\rho(y))\leq\sum_{1\leq i\leq n-1}d_{\mathcal{X}_{1}}(y_{i},x_{i+1})\leq\mu_{1}\sum_{\ell\in\mathbb{A}}[d_{\ell}(x,y)]_{K}+2Kd_{T}(\rho(x),\rho(y)).

Using (8), we now replace [dγ(xi,yi)]K[d_{\gamma}(x_{i},y_{i})]_{K} by [dγ(x,y)]K[d_{\gamma}(x,y)]_{K} in the formula (5) for dY¯vi(xi,yi)d_{\overline{Y}_{v_{i}}}(x_{i},y_{i}). Hence, there exists a constant μ2=μ2(K,L)>1\mu_{2}=\mu_{2}(K,L)>1 so that

1μ2γ𝔸vi[dγ(x,y)]KdY¯vi(xi,yi)μ2γ𝔸vi[dγ(x,y)]K+2K.\frac{1}{\mu_{2}}\sum_{\gamma\in\mathbb{A}_{v_{i}}}[d_{\gamma}(x,y)]_{K}\leq d_{\overline{Y}_{v_{i}}}(x_{i},y_{i})\leq\mu_{2}\sum_{\gamma\in\mathbb{A}_{v_{i}}}[d_{\gamma}(x,y)]_{K}+2K.

Noting dT(ρ(x),ρ(y))=2n2d_{T}(\rho(x),\rho(y))=2n-2, we deduce from Eq. (7) and (9) that

d𝒳1(x,y)(μ1+μ2)1inγ𝔸vi[dγ(x,y)]K+4KdT(ρ(x),ρ(y))d_{\mathcal{X}_{1}}(x,y)\leq(\mu_{1}+\mu_{2})\sum_{1\leq i\leq n}\sum_{\gamma\in\mathbb{A}_{v_{i}}}[d_{\gamma}(x,y)]_{K}+4K\cdot d_{T}(\rho(x),\rho(y))

so the upper bound in (6) follows by setting μ:=μ1+μ2\mu:=\mu_{1}+\mu_{2}.

We now derive the lower bound from those of Eq. (7) and (9):

d𝒳1(x,y)1indY¯vi(xi,yi)+dT(ρ(x),ρ(y))1μ21inγ𝔸vi[dγ(x,y)]K+dT(ρ(x),ρ(y))\begin{array}[]{rl}d_{\mathcal{X}_{1}}(x,y)&\geq\sum_{1\leq i\leq n}d_{\overline{Y}_{v_{i}}}(x_{i},y_{i})+d_{T}(\rho(x),\rho(y))\\ &\geq\frac{1}{\mu_{2}}\sum_{1\leq i\leq n}\sum_{\gamma\in\mathbb{A}_{v_{i}}}[d_{\gamma}(x,y)]_{K}+d_{T}(\rho(x),\rho(y))\end{array}

By the Claim 1, there are at most LL quasi-lines γ{𝔸v:v𝒱1[ρ(x),ρ(y)]0}\gamma\in\cup\{\mathbb{A}_{v}:v\in\mathcal{V}_{1}-[\rho(x),\rho(y)]^{0}\} satisfying L[dγ(x,y)]K>0L\geq[d_{\gamma}(x,y)]_{K}>0. Hence, the following holds

d𝒳1(x,y)1μ2v𝒱1γ𝔸v[dγ(x,y)]K+dT(ρ(x),ρ(y))L2\begin{array}[]{rl}d_{\mathcal{X}_{1}}(x,y)&\geq\frac{1}{\mu_{2}}\sum_{v\in\mathcal{V}_{1}}\sum_{\gamma\in\mathbb{A}_{v}}[d_{\gamma}(x,y)]_{K}+d_{T}(\rho(x),\rho(y))-L^{2}\end{array}

completing the proof of the lower bound. ∎

Lemma 4.21.

The collection 𝔸\mathbb{A} can be written as a union (possibly non-disjoint) 𝔸1𝔸n\mathbb{A}_{1}\cup\cdots\cup\mathbb{A}_{n} with the following properties for each 𝔸i\mathbb{A}_{i}:

  1. (1)

    for any two quasi-lines αβ𝔸i\alpha\neq\beta\in\mathbb{A}_{i} we have d(α,β)Dd(\alpha,\beta)\geq D,

  2. (2)

    the (D+R)(D+R)–neighborhood of the union γ𝔸iγ\cup_{\gamma\in\mathbb{A}_{i}}\gamma contains 𝒳1\mathcal{X}_{1},

  3. (3)

    for any K>4ξK>4\xi the quasi-tree of quasi-lines (𝒞K(𝔸i),d𝒞i)(\mathcal{C}_{K}(\mathbb{A}_{i}),d_{\mathcal{C}_{i}}) is a quasi-tree.

Proof.

Since HvH_{v} acts geometrically on Y¯v\overline{Y}_{v} for v𝒱1v\in\mathcal{V}_{1} and 𝒱1\mathcal{V}_{1} is G˙\dot{G}–finite, there exists a constant R>0R>0 such that the RR–neighborhood of the union γ𝔸vγ\cup_{\gamma\in\mathbb{A}_{v}}\gamma contains Y¯v\overline{Y}_{v} for each v𝒱1v\in\mathcal{V}_{1}. Since 𝔸\mathbb{A} is locally finite and G˙\dot{G}–invariant, the DD–neighborhood of any quasi-line in 𝔸\mathbb{A} intersects nn quasi-lines from 𝔸\mathbb{A} for some n=n(D)1n=n(D)\geq 1.

We can now write 𝔸\mathbb{A} as the (possibly non-disjoint) union 𝔸1𝔸n\mathbb{A}_{1}\cup\cdots\cup\mathbb{A}_{n} with the following two properties for each 𝔸i\mathbb{A}_{i}:

  1. (1)

    for any two quasi-lines αβ𝔸i\alpha\neq\beta\in\mathbb{A}_{i} we have d(α,β)Dd(\alpha,\beta)\geq D,

  2. (2)

    the (D+R)(D+R)–neighborhood of the union γ𝔸iγ\cup_{\gamma\in\mathbb{A}_{i}}\gamma contains 𝒳1\mathcal{X}_{1}.

Indeed, by definition of RR, any ball of radius RR intersects a quasi-line so for each α𝔸\alpha\in\mathbb{A}, there exists β𝔸\beta\in\mathbb{A} such that Dd(α,β)D+2RD\leq d(\alpha,\beta)\leq D+2R. Starting from a quasi-line γ1\gamma_{1}, we inductively choose the quasi-lines which intersect the (D+R)(D+R)–neighborhood of the already chosen ones, and by the axiom of choice, a collection 𝔸1\mathbb{A}_{1} of quasi-lines containing γ1\gamma_{1} is obtained so that the properties (1) and (2) are true. The other collections 𝔸i\mathbb{A}_{i} for ni1n\geq i\geq 1 is obtained similarly from the other n1n-1 quasi-lines intersecting the DD–neighborhood of γ1\gamma_{1}. The property (2) guarantees 𝔸1in𝔸i\mathbb{A}\subseteq\cup_{1\leq i\leq n}\mathbb{A}_{i} from the choice of RR. We do allow 𝔸i𝔸j\mathbb{A}_{i}\cap\mathbb{A}_{j}\neq\varnothing, but any γ𝔸\gamma\in\mathbb{A} would appear at most once in each 𝔸i\mathbb{A}_{i}.

By the defining property of DD, the collection 𝔸i\mathbb{A}_{i} of quasi-lines in the hyperbolic space 𝒳1\mathcal{X}_{1} satisfies diam(πβ(α))θdiam(\pi_{\beta}(\alpha))\leq\theta for all αβ𝔸i\alpha\neq\beta\in\mathbb{A}_{i}. By Proposition 4.9, (𝔸i,πγ)(\mathbb{A}_{i},\pi_{\gamma}) satisfies projection axioms with projection constant ξ=ξ(θ)\xi=\xi(\theta). For given K>4ξK>4\xi, the quasi-tree of quasi-lines (𝒞K(𝔸i),d𝒞i)(\mathcal{C}_{K}(\mathbb{A}_{i}),d_{\mathcal{C}_{i}}) is a quasi-tree by [BBF15]. ∎

Proof of Theorem 1.5.

Let 𝒳i\mathcal{X}_{i} (i=1,2i=1,2) be the hyperbolic space constructed in Section 4.1 with respect to 𝒱i\mathcal{V}_{i}. By Proposition 4.8, the admissible group GG admits a quasi-isometric embedding into 𝒳1×𝒳2\mathcal{X}_{1}\times\mathcal{X}_{2}. Thus, to complete the proof of Theorem 1.5, we only need to show that each hyperbolic space 𝒳i\mathcal{X}_{i} is quasi-isometric embedded into a finite product of quasi-trees. We give the proof for 𝒳1\mathcal{X}_{1} and the proof for 𝒳2\mathcal{X}_{2} is symmetric.

Let ρ:XT0\rho\colon X\to T^{0} be the indexed map given by Remark 2.6. Let 𝔸1,,𝔸n\mathbb{A}_{1},\dots,\mathbb{A}_{n} be the collection of quasi-lines given by Lemma 4.21.

Let 𝒳1^:=v𝒱1Y¯v\hat{\mathcal{X}_{1}}:=\cup_{v\in\mathcal{V}_{1}}\overline{Y}_{v}. Since the 11–neighborhood of 𝒳1^\hat{\mathcal{X}_{1}} is 𝒳1\mathcal{X}_{1}, it suffices define a quasi-isometric embedding map from 𝒳1^\hat{\mathcal{X}_{1}} to a finite product of quasi-trees.

We now define a map

Φ:𝒳1^T×1in𝒞K(𝔸i),\Phi:\hat{\mathcal{X}_{1}}\to T\times\prod_{1\leq i\leq n}\mathcal{C}_{K}(\mathbb{A}_{i}),

where TT is the Bass-Serre tree of GG.

Let x𝒳1^=v𝒱1Y¯vx\in\hat{\mathcal{X}_{1}}=\cup_{v\in\mathcal{V}_{1}}\overline{Y}_{v} and assume xY¯vx\in\overline{Y}_{v}. By the property (2) of Lemma 4.21, we choose a point Φi(x)γ𝔸iγ\Phi_{i}(x)\in\cup_{\gamma\in\mathbb{A}_{i}}\gamma for each 1in1\leq i\leq n such that d(x,Φi(x))R+Dd(x,\Phi_{i}(x))\leq R+D. Denote R~=R+D\tilde{R}=R+D. Let Φ(x)=(ρ(x),Φ1(x),,Φn(x))\Phi(x)=(\rho(x),\Phi_{1}(x),\cdots,\Phi_{n}(x)).

We now verify that Φ\Phi is a quasi-isometric embedding. Since d(x,(Φi(x))R~d(x,(\Phi_{i}(x))\leq\tilde{R} and d(y,Φi(y))R~d(y,\Phi_{i}(y))\leq\tilde{R}, let K0=K0(R~,ξ,δ)K_{0}=K_{0}(\tilde{R},\xi,\delta) be given by Corollary 4.11 so that for K>K0K>K_{0}, the following distance formula holds

d𝒞i(Φi(x),Φi(y))Kγ𝔸i[dγ(x,y)]Kd_{\mathcal{C}_{i}}(\Phi_{i}(x),\Phi_{i}(y))\sim_{K}\sum_{\gamma\in\mathbb{A}_{i}}[d_{\gamma}(x,y)]_{K}

for each 1in1\leq i\leq n. Therefore,

d(Φ(x),Φ(y))=dT(ρ(x),ρ(y))+i=1nd𝒞i(Φi(x),Φi(y))KdT(ρ(x),ρ(y))+i=1nγ𝔸i[dγ(x,y)]K.\begin{array}[]{rl}d(\Phi(x),\Phi(y))&=d_{T}(\rho(x),\rho(y))+\sum_{i=1}^{n}d_{\mathcal{C}_{i}}(\Phi_{i}(x),\Phi_{i}(y))\\ &\\ &\sim_{K}d_{T}(\rho(x),\rho(y))+\sum_{i=1}^{n}\sum_{\gamma\in\mathbb{A}_{i}}[d_{\gamma}(x,y)]_{K}.\end{array}

Recall that 𝔸=1in𝔸i\mathbb{A}=\cup_{1\leq i\leq n}\mathbb{A}_{i} is a possibly non-disjoint union, but any quasi-line γ\gamma with [dγ(x,y)]K>0[d_{\gamma}(x,y)]_{K}>0 in the above sum appears at most once in each 𝔸i\mathbb{A}_{i}. We thus obtain

d(Φ(x),Φ(y))K,ndT(ρ(x),ρ(y))+γ𝔸[dγ(x,y)]K\begin{array}[]{c}d(\Phi(x),\Phi(y))\sim_{K,n}d_{T}(\rho(x),\rho(y))+\sum_{\gamma\in\mathbb{A}}[d_{\gamma}(x,y)]_{K}\end{array}

which together with distance formula (6) for 𝒳1\mathcal{X}_{1} concludes the proof of Theorem. ∎

5. Proper action on a finite product of quasi-trees

In this section, under a stronger assumption on vertex groups as stated in Theorem 1.1, we shall promote the quasi-isometric embedding to be an orbital map of an action of the admissible group on a (different) finite product of quasi-trees.

By [BBF19, Induction 2.2], if H<GH<G has finite index and acts on a finite product of quasi-trees, then so does GG. We are thus free to pass to finite index subgroups in the proof.

Recall that T0=𝒱1𝒱2T^{0}=\mathcal{V}_{1}\cup\mathcal{V}_{2} where 𝒱i\mathcal{V}_{i} consists of vertices in TT with pairwise even distances, and 𝒳{\mathcal{X}} is the hyperbolic space constructed from 𝒱{𝒱1,𝒱2}\mathcal{V}\in\{\mathcal{V}_{1},\mathcal{V}_{2}\}. By Lemma 4.6, let G˙<G\dot{G}<G be the subgroup of index at most 22 preserving 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2}.

5.1. Construct cone-off spaces: preparation

In this preparatory step, we first introduce another hyperbolic space 𝒳˙\dot{\mathcal{X}} which is the cone-off of the previous hyperbolic space 𝒳\mathcal{X} over boundary lines of flat links. We then embed 𝒳{\mathcal{X}} into a product of 𝒳˙\dot{\mathcal{X}} and a quasi-tree built from the set of binding lines from the flat links.

Definition 5.1 (Hyperbolic cones).

[BH99, Part I, Ch. 5] For a line \ell and a constant r>0r>0, a hyperbolic rr–cone denoted by coner()cone_{r}(\ell) is the quotient space of ×[0,r]\ell\times[0,r] by collapsing ×0\ell\times 0 as a point called apex. A metric is endowed on coner()cone_{r}(\ell) so that it is isometric to the metric completion of the universal covering of a closed disk of radius rr in the real hyperbolic plane 2\mathbb{H}^{2} punctured at the center.

A hyperbolic multicones of radius rr is the countable wedge of hyperbolic rr–cones with apex identified.

If ξ\xi is an isometry on \ell then ξ\xi extends to a natural isometric action on the hyperbolic cone coner()cone_{r}(\ell) which rotates around the apex and sends the radicals to radicals.

Similar to the flat links, the link Lk(w)Lk(w) for wT0w\in T^{0} determines a hyperbolic multicones of radius rr denoted by Conesr(w)Cones_{r}(w) so that the set of hyperbolic cones is bijective to the set of vertices adjacent to ww. And Gw=Hw×Z(Gw)G_{w}=H_{w}\times Z(G_{w}) acts on Conesr(w)Cones_{r}(w) so that the center of GwG_{w} rotates each hyperbolic cone around the apex and Gw/Z(Gw)G_{w}/Z(G_{w}) permutes the set of hyperbolic cones by the action of GwG_{w} on Lk(w)Lk(w).

Construction of the cone-off space 𝒳˙\dot{\mathcal{X}}. Let 𝒱{𝒱1,𝒱2}\mathcal{V}\in\{\mathcal{V}_{1},\mathcal{V}_{2}\} and r>0r>0. Let 𝒳˙\dot{\mathcal{X}} be the disjoint union of {Y¯v:v𝒱}\{\overline{Y}_{v}:v\in\mathcal{V}\} and hyperbolic multicones {Conesr(w),wT0𝒱}\{Cones_{r}(w),w\in T^{0}-\mathcal{V}\} glued by isometry along the boundary lines of vLk(w)Y¯v\cup_{v\in Lk(w)}\overline{Y}_{v} and those of hyperbolic multicones Cones(w)Cones(w).

Remark 5.2.

We note that 𝒳˙\dot{\mathcal{X}} is obtained from 𝒳\mathcal{X} by replacing each flat links by hyperbolic multicones. However, the identification in 𝒳˙\dot{\mathcal{X}} between boundary lines of Y¯v\overline{Y}_{v} and multicones Cones(w)Cones(w) is only required to be isometric, while the \mathbb{R}–coordinates of the boundary lines in constructing 𝒳{\mathcal{X}} have to be matched up.

We now give an alternative way to construct the cone-off space 𝒳˙\dot{\mathcal{X}}, which shall be convenient in the sequel.

Relatively hyperbolic structure of cone-off spaces. Let Y˙v\dot{Y}_{v} be the disjoint union of Y¯v\overline{Y}_{v} and hyperbolic cones coner()cone_{r}(\ell) glued along boundary lines 𝕃v\ell\in\mathbb{L}_{v}. If E()E(\ell) denotes the stabilizer in HvH_{v} of the boundary 𝕃v\ell\in\mathbb{L}_{v}, then E()E(\ell) is virtually cyclic and almost malnormal. Since {E():𝕃v}\{E(\ell):\ell\in\mathbb{L}_{v}\} is HvH_{v}-finite by conjugacy, choose a complete set 𝔼v\mathbb{E}_{v} of representatives. By [Bow12], HvH_{v} is hyperbolic relative to peripheral subgroups 𝔼v\mathbb{E}_{v}.

Let 𝒢^(Hv)\hat{\mathcal{G}}(H_{v}) be the coned-off Cayley graph (after choosing a finite generating set) by adding a cone point for each peripheral coset of E()𝔼vE(\ell)\in\mathbb{E}_{v} and joining the cone point by half edges to each element in the peripheral coset. The union of two half edges with two endpoints in a peripheral coset shall be referred to as an peripheral edge. See the relevant details in [Far98].

By [Bow08, Lemma 3.3], [Osi16, Prop. 5.2], the action of HvH_{v} on 𝒢^(Hv)\hat{\mathcal{G}}(H_{v}) is acylindrical. There exists a HvH_{v}-equivariant quasi-isometry between the coned-off Cayley graph 𝒢^(Hv)\hat{\mathcal{G}}(H_{v}) and the coned-off space Y˙v\dot{Y}_{v} which sends peripheral coset of E()𝔼vE(\ell)\in\mathbb{E}_{v} to \ell. Thus, the action of HvH_{v} on Y˙v\dot{Y}_{v} is co-bounded and acylindrical as well.

Alternatively, the cone-off space 𝒳˙\dot{\mathcal{X}} could be obtained from the disjoint union of {Y˙v}v𝒱\{\dot{Y}_{v}\}_{v\in\mathcal{V}} by identifying the apex of hyperbolic cones from the same link Lk(w)Lk(w) where wT0𝒱{w\in T^{0}-\mathcal{V}}.

Since 𝕃v\mathbb{L}_{v} has the bounded intersection property, by [DGO17, Corollary 5.39], for a sufficiently large constant rr, the space Y˙v\dot{Y}_{v} is a hyperbolic space with constant depending only on the original one. Thus, the space 𝒳˙\dot{\mathcal{X}} is a hyperbolic space.

By Lemma 4.6, a subgroup G˙\dot{G} of index at most 2 in GG leaves invariant 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2}. The following lemma is proved similarly as Lemma 4.7.

Lemma 5.3.

Fix a sufficiently large r>0r>0. The space 𝒳˙\dot{\mathcal{X}} is a δ\delta–hyperbolic space where δ>0\delta>0 only depends on the hyperbolicity constants of Y˙v\dot{Y}_{v} (v𝒱v\in\mathcal{V}).

If Gv=Hv×Z(Gv)G_{v}=H_{v}\times Z(G_{v}) for every vT0v\in T^{0} and Ge=Z(Gv)×Z(Gw)G_{e}=Z(G_{v})\times Z(G_{w}) for every edge e=[v,w]T1e=[v,w]\in T^{1}, then a subgroup G˙\dot{G} of index at most 2 in GG acts on 𝒳˙\dot{\mathcal{X}} with the following properties:

  1. (1)

    for each v𝒱v\in\mathcal{V}, the stabilizer of Y˙v\dot{Y}_{v} is isomorphic to GvG_{v} and HvH_{v} acts coboundedly on Y˙v\dot{Y}_{v}, and

  2. (2)

    for each wT0𝒱w\in T^{0}-\mathcal{V}, the stabilizer of the apex of Conesr(w)Cones_{r}(w) is isomorphic to GwG_{w} so that HwH_{w} acts on the set of hyperbolic cones by the action on the link Lk(w)Lk(w) and Z(Gw)Z(G_{w}) on acts by rotation on each hyperbolic cone.

From now on and until the end of the next subsection, we assume that Gv=Hv×Z(Gv)G_{v}=H_{v}\times Z(G_{v}) for every vT0v\in T^{0} and Ge=Z(Gv)×Z(Gw)G_{e}=Z(G_{v})\times Z(G_{w}) for every edge e=[v,w]T1e=[v,w]\in T^{1}. After passage to a finite index subgroup of GG, this assumption will be guaranteed by Corollary 5.13 below, where HvH_{v} are assumed to omnipotent.

The remainder of this section is to relate the metric geometry of (𝒳˙,d𝒳˙)(\dot{\mathcal{X}},d_{\dot{\mathcal{X}}}) and (𝒳,d𝒳)(\mathcal{X},d_{\mathcal{X}}). This is achieved by a series of lemmas, accumulating in Lemma 5.6.

Let π\pi_{\ell} denote the shortest projection to a quasi-line \ell in 𝒳\mathcal{X} or in 𝒳˙\dot{\mathcal{X}}. To be clear, we use d(x,y)d_{\ell}(x,y) (resp. d˙(x,y)\dot{d}_{\ell}(x,y)) to denote the d𝒳d_{\mathcal{X}}-diameter (resp. d𝒳˙d_{\dot{\mathcal{X}}}-diameter) of the projection of the points x,yx,y to \ell.

Let 𝕃\mathbb{L} be the set of all binding lines in the flat links Fl(w)Fl(w) over wT0𝒱w\in T^{0}-\mathcal{V}. Note that 𝕃\mathbb{L} is disjoint with the union v𝒱𝕃v\cup_{v\in\mathcal{V}}\mathbb{L}_{v}, although the binding lines in flat links Fl(w)Fl(w) are parallel to the boundary lines of (the flat strips and) Y¯v\overline{Y}_{v} for vLk(w)v\in Lk(w).

We first give an analogue of Lemma 4.17 for acylindrical action on coned-off spaces Y˙v\dot{Y}_{v}.

Up the equivariant quasi-isometry mentioned above, we shall identify (Y¯v,d)(\overline{Y}_{v},d) with the Cayley graph of HvH_{v}, and (Y˙v,d˙)(\dot{Y}_{v},\dot{d}) with the coned-off Cayley graph 𝒢^(Hv)\hat{\mathcal{G}}(H_{v}), and 𝕃v\mathbb{L}_{v} with the collection of left peripheral cosets of 𝔼v\mathbb{E}_{v}.

A geodesic edge path β\beta in the coned-off Cayley graph Y˙v\dot{Y}_{v} is KK-bounded for K>0K>0 if every peripheral edge has two endpoints within Y¯v\overline{Y}_{v}-distance at most KK.

By definition, a geodesic β=[x,y]\beta=[x,y] can be subdivided into maximal KK-bounded non-trivial segments αi\alpha_{i} (0in0\leq i\leq n) separated by peripheral edges eje_{j} (0jm0\leq j\leq m) where dY¯v((ej),(ej)+)>Kd_{\overline{Y}_{v}}((e_{j})_{-},(e_{j})_{+})>K. It is possible that n=0n=0: β\beta consists of only peripheral edges.

Define

|β|K:=0in[Len(αi)]K,|\beta|_{K}:=\sum_{0\leq i\leq n}[Len(\alpha_{i})]_{K},

which sums up the lengths of KK-bounded subpaths of length at least KK. It is possible that n=0n=0, so |β|K=0|\beta|_{K}=0. Define dY˙vK(x,y)d_{\dot{Y}_{v}}^{K}(x,y) to be the maximum of |β|K|\beta|_{K} over all relative geodesics β\beta between x,yx,y. Thus, dY˙vK(x,y)d_{\dot{Y}_{v}}^{K}(x,y) is HvH_{v}-invariant. We remark that the “thick” distance dY˙vK(x,y)d_{\dot{Y}_{v}}^{K}(x,y) is inspired by the corresponding [BBF19, Def. 4.8] in mapping class groups.

Lemma 5.4.

For any sufficiently large K>0K>0,

(10) dY¯v(x,y)KdY˙vK(x,y)+γ𝕃v[dγ(x,y)]Kd_{\overline{Y}_{v}}(x,y)\sim_{K}d_{\dot{Y}_{v}}^{K}(x,y)+\sum_{\gamma\in\mathbb{L}_{v}}[d_{\gamma}(x,y)]_{K}
Proof.

Let β\beta be a geodesic between x,yx,y in Y˙v\dot{Y}_{v} so that dY˙v(x,y)=Len(β)d_{\dot{Y}_{v}}(x,y)=Len(\beta). We obtain the lifted path β^\hat{\beta} by replacing each peripheral edge ee with endpoints in a peripheral coset e\ell_{e} by a geodesic in Y¯v\overline{Y}_{v} with same endpoints. The well-known fact ([DS05], [GP16]) that β\beta is a uniform quasi-geodesic gives the so-called distance formula [Sis13, Theorem 0.1]: for any K0K\gg 0,

dY¯v(x,y)KdY˙v(x,y)+γ𝕃v[dγ(x,y)]K.d_{\overline{Y}_{v}}(x,y)\sim_{K}d_{\dot{Y}_{v}}(x,y)+\sum_{\gamma\in\mathbb{L}_{v}}[d_{\gamma}(x,y)]_{K}.

Indeed, the additional ingredient (to quasi-geodesicity of β^\hat{\beta}) is the following fact: dγ(x,y)Kd_{\gamma}(x,y)\geq K if and only if β\beta contains a peripheral edge ee with two endpoints in γ\gamma with dY¯v(e,e+)>Kd_{\overline{Y}_{v}}(e_{-},e_{+})>K^{\prime}, where KK depends on KK^{\prime} and vice versa.

By definition, Len(β)Len(\beta) differs from |β|K=dY˙vK(x,y)|\beta|_{K}=d_{\dot{Y}_{v}}^{K}(x,y) in the sum of lengths of at most (m+2)(m+2) segments αi\alpha_{i} and (m+1)(m+1) edges eje_{j}, where Len(αi)<KLen(\alpha_{i})<K and dY¯v((ej),(ej)+)Kd_{\overline{Y}_{v}}((e_{j})_{-},(e_{j})_{+})\geq K. Therefore, we can replace dY˙v(x,y)d_{\dot{Y}_{v}}(x,y) with dY˙vK(x,y)d_{\dot{Y}_{v}}^{K}(x,y) by worsening the multiplicative constant in the above distance formula. This shows (10). ∎

Lemma 5.5.

For any sufficiently large K>max{4ξ,θ}K>\max\{4\xi,\theta\}, there exist a HvH_{v}–finite collection 𝔸v\mathbb{A}_{v} of quasi-lines in Y˙v\dot{Y}_{v} and a constant N=N(K,δ,𝔸v)>0N=N(K,\delta,\mathbb{A}_{v})>0, such that for any two vertices x,yY˙vx,y\in\dot{Y}_{v}, the following holds

(11) dY˙vK(x,y)Nγ𝔸v[d˙γ(x,y)]Kd_{\dot{Y}_{v}}^{K}(x,y)\sim_{N}\sum_{\gamma\in\mathbb{A}_{v}}[\dot{d}_{\gamma}(x,y)]_{K}
Proof.

For simplicity, we suppress indices vv from Hv,Yv,Y˙v,𝕃vH_{v},Y_{v},\dot{Y}_{v},\mathbb{L}_{v} in the proof.

Let SS be the set of elements hHh\in H so that d˙(1,h)=K\dot{d}(1,h)=K and some geodesic [1,h][1,h] in Y˙\dot{Y} is KK-bounded. The definition of KK-boundedness implies that SS is a finite set, since d(1,h)K2d(1,h)\leq K^{2} and the word metric dd is proper.

Let β\beta be a geodesic path between x,yx,y so that dY˙vK(x,y)=|β|Kd_{\dot{Y}_{v}}^{K}(x,y)=|\beta|_{K}. Let {αi:0in}\{\alpha_{i}:0\leq i\leq n\} be the set of maximal KK-bounded segments of β\beta in Y˙\dot{Y}. Then

(12) |β|K=i=0n[LenY˙(αi)]K.\begin{array}[]{rl}|\beta|_{K}=\;\sum_{i=0}^{n}[Len_{\dot{Y}}(\alpha_{i})]_{K}.\end{array}

We now follow the argument in the proof of Lemma 4.17 to produce the collection of quasi-lines 𝔸\mathbb{A} to approximate LenY˙(αi)Len_{\dot{Y}}(\alpha_{i}).

The argument below considers every αi\alpha_{i} in the above sum of (12), but for simplicity, denote α=αi\alpha=\alpha_{i} with endpoints α,α+\alpha_{-},\alpha_{+}.

Since the action of HH on Y˙\dot{Y} is acylindrical, consider the set S~\tilde{S} of loxodromic elements hfhf on Y˙\dot{Y} where hSh\in S and fFf\in F is provided by Lemma 4.13. Since SS is finite, S~\tilde{S} is finite as well. Let 𝔸\mathbb{A} be the set of all HH–translated axis of hfS~hf\in\tilde{S}.

If LenY˙(α)KLen_{\dot{Y}}(\alpha)\geq K, we subdivide further α=γ0γ1γn\alpha=\gamma_{0}\cdot\gamma_{1}\cdots\gamma_{n}, so that subsegments γi\gamma_{i} (0i<n)(0\leq i<n) have Y˙\dot{Y}-length KK and the last one γn\gamma_{n} have Y˙\dot{Y}-length less than KK. Each γi\gamma_{i} is KK-bounded since the KK-boundedness is preserved by taking subpaths. Similarly as in Lemma 4.17, a set of axes of loxodromic elements γif\gamma_{i}f is then produced to give the upper bound of LenY˙(α)Len_{\dot{Y}}(\alpha). The lower bound uses a certain local finiteness of 𝔸\mathbb{A} due to the proper action in Lemma 4.17, which is now guaranteed by Corollary 4.15 in an acylindrical action. Thus,

(13) 1Nγ𝔸[dγ(α,α+)]KLenY˙(α)2γ𝔸[dγ(α,α+)]K+2K\frac{1}{N}\sum_{\gamma\in\mathbb{A}}[d_{\gamma}(\alpha_{-},\alpha_{+})]_{K}\leq Len_{\dot{Y}}(\alpha)\leq 2\sum_{\gamma\in\mathbb{A}}[d_{\gamma}(\alpha_{-},\alpha_{+})]_{K}+2K

where the constant NN depends on K,δ,𝔸K,\delta,\mathbb{A}.

Recall that each γ𝔸\gamma\in\mathbb{A} is a uniform quasi-line, i.e.: a (λ,c)(\lambda,c)-quasi-geodesic in Y˙\dot{Y}. Since α=αi\alpha=\alpha_{i} is a subsegment of the geodesic β=[x,y]\beta=[x,y], the hyperbolicity of Y˙\dot{Y} implies that πγ(α)\pi_{\gamma}(\alpha_{-}) and πγ(α+)\pi_{\gamma}(\alpha_{+}) are contained in a CC-neighborhood of πγ(x)\pi_{\gamma}(x) and πγ(y)\pi_{\gamma}(y) respectively, where CC depends on λ,c,δ\lambda,c,\delta. Thus, we have

(14) [d˙γ(α,α+)]KK,C[d˙γ(x,y)]K[\dot{d}_{\gamma}(\alpha_{-},\alpha_{+})]_{K}\asymp_{K,C}[\dot{d}_{\gamma}(x,y)]_{K}

for every γ𝔸\gamma\in\mathbb{A}. So for every α=αi\alpha=\alpha_{i}, we obtained (13) and (14), and the conclusion thus follows from (12). ∎

Recall that a geodesic γ=[x,y]\gamma=[x,y] in 𝒳˙\dot{\mathcal{X}} is the union of a sequence of maximal geodesics βv\beta_{v} in Y˙v\dot{Y}_{v}’s. As dY˙vK(x,y)d_{\dot{Y}_{v}}^{K}(x,y) is related with dY˙v(x,y)d_{\dot{Y}_{v}}(x,y), we define

(15) d𝒳˙K(x,y):=vdY˙vK((βv),(βv)+).d_{\dot{\mathcal{X}}}^{K}(x,y):=\sum_{v}d_{\dot{Y}_{v}}^{K}((\beta_{v})_{-},(\beta_{v})+).

Since dY˙vK(,)d_{\dot{Y}_{v}}^{K}(\cdot,\cdot) is HvH_{v}-invariant, we see that d𝒳˙Kd_{\dot{\mathcal{X}}}^{K} is G˙\dot{G}-invariant (even though it is not a metric anymore).

Lemma 5.6.

There exists K0>0K_{0}>0 such that for any two points x,y𝒳x,y\in\mathcal{X} and K>K0K>K_{0}, we have

(16) d𝒳(x,y)Kd𝒳˙K(x,y)+𝕃[dγ(x,y)]Kd_{\mathcal{X}}(x,y)\sim_{K}d_{\dot{\mathcal{X}}}^{K}(x,y)+\sum_{\ell\in\mathbb{L}}[d_{\gamma}(x,y)]_{K}

Recall that the notation AKBA\sim_{K}B means that AA equals BB up to multiplicative and additive constants depending on KK.

Proof.

Let γ=[x,y]\gamma=[x,y] be a 𝒳˙\dot{\mathcal{X}}-geodesic with endpoints x,y𝒳x,y\in\mathcal{X}. Let us consider the generic case that ρ(x)ρ(y)\rho(x)\neq\rho(y); the case ρ(x)ρ(y)\rho(x)\neq\rho(y) is much easier and left to the reader.

Assume that x,y𝒳x,y\in\mathcal{X} are not contained in any hyperbolic rr-cones, up to moving x,yx,y with a distance at most rr. We can then write the geodesic γ\gamma as the following union

(17) γ=(v𝒱[ρ(x),ρ(y)]βv)(w(T0𝒱)[ρ(x),ρ(y)]cw)\gamma=\left(\cup_{v\in\mathcal{V}\cap[\rho(x),\rho(y)]}\beta_{v}\right)\bigcup\left(\cup_{w\in(T^{0}-\mathcal{V})\cap[\rho(x),\rho(y)]}c_{w}\right)

where βv\beta_{v} is a geodesic in the cone-off space Y˙v\dot{Y}_{v} with endpoints on boundary lines of YvY˙vY_{v}\subset\dot{Y}_{v}, and cwc_{w} is contained in the hyperbolic multicones Conesr(w)Cones_{r}(w) passing the apex. Thus, Len(cw)Len(c_{w}) has length 2r2r.

We replace c:=cwc:=c_{w} by an 𝒳\mathcal{X}-geodesic cc^{\prime} with the same endpoints c,c+c_{-},c_{+}: cc^{\prime} is contained in the corresponding flat links, whose binding line is denoted by c\ell_{c}. Thus, cN1(c)c^{\prime}\subset N_{1}(\ell_{c}). This replacement becomes non-unique when different cc’s have overlap (in the subspace Y¯vYv˙\overline{Y}_{v}\subset\dot{Y_{v}}). However, the bounded intersection of 𝕃\mathbb{L} gives a uniform upper bound on the overlap. Let KK be any constant sufficiently bigger than this bound. We then number those subpaths cc of γ\gamma with d𝒳(c,c+)>Kd_{\mathcal{X}}(c_{-},c_{+})>K in a fixed order (eg. from left to right): c1,,cnc_{1},\dots,c_{n}. Up to bounded modifications, we obtain a well-defined notion of lifted path γ^\hat{\gamma} with same endpoints of γ\gamma.

Observe that γ^\hat{\gamma} is a uniform quasi-geodesic. This is well-known and one proof proceeds as follows: since the set of binding lines 𝕃\mathbb{L} has bounded intersection in 𝒳\mathcal{X}, the above construction implies that γ^\hat{\gamma} is an efficient semi-polygonal path in the sense of [Bow12, 6, Section 7]. The observation then follows as a consequence of [Bow12, Lemma 7.3].

If d𝒳(c,c+)>Kd_{\mathcal{X}}(c_{-},c_{+})>K so cN1(c)γ^c^{\prime}\subset N_{1}(\ell_{c})\cap\hat{\gamma}, then d𝒳(c,c+)[dc(x,y)]Kd_{\mathcal{X}}(c_{-},c_{+})\asymp[d_{\ell_{c}}(x,y)]_{K} follows from the quasi-geodesicity of γ^\hat{\gamma} and the hyperbolicity of 𝒳\mathcal{X}. We incorporate the cwc_{w}’s in Eq. (17) with d𝒳((cw),(cw)+)Kd_{\mathcal{X}}((c_{w})_{-},(c_{w})_{+})\leq K into the multiplicative constant, and thus the following holds

Len𝒳(γ^)Kv𝒱[ρ(x),ρ(y)]Len𝒳(βv)+i=1n[dci(x,y)]K.Len_{\mathcal{X}}(\hat{\gamma})\sim_{K}\sum_{v\in\mathcal{V}\cap[\rho(x),\rho(y)]}Len_{\mathcal{X}}(\beta_{v})+\sum_{i=1}^{n}[d_{\ell_{c_{i}}}(x,y)]_{K}.

Combining the formula (10) and (15), we get

Len𝒳(βv)|βv|K+𝕃v[dγ((βv),(βv)+)]K.Len_{\mathcal{X}}(\beta_{v})\sim|\beta_{v}|_{K}+\sum_{\ell\in\mathbb{L}_{v}}[d_{\gamma}((\beta_{v})_{-},(\beta_{v})_{+})]_{K}.

Using the local finiteness and bounded intersection 𝕃\mathbb{L}, for each ci\ell_{c_{i}} with [dγci(x,y)]K>0[d_{\gamma_{c_{i}}}(x,y)]_{K}>0, there are only finitely many 𝕃\ell\in\mathbb{L} such that [dγ(x,y)]K>0[d_{\gamma_{\ell}}(x,y)]_{K}>0. Hence, by worsening the multiplicative constant, we have

Len𝒳(γ^)Kv𝒱[ρ(x),ρ(y)]|βv|K+𝕃[d(x,y)]K.Len_{\mathcal{X}}(\hat{\gamma})\asymp_{K}\sum_{v\in\mathcal{V}\cap[\rho(x),\rho(y)]}|\beta_{v}|_{K}+\sum_{\ell\in\mathbb{L}}[d_{\ell}(x,y)]_{K}.

Recall that Len𝒳(γ^)d𝒳(x,y)Len_{\mathcal{X}}(\hat{\gamma})\sim d_{\mathcal{X}}(x,y) from the quasi-geodesic γ^\hat{\gamma}. Then the desired (16) follows. ∎

Recall that 𝕃\mathbb{L} is the set of all binding lines in Fl(w)Fl(w) over wT0𝒱w\in T^{0}-\mathcal{V}. Since 𝕃\mathbb{L} has bounded projection property in 𝒳\mathcal{X}, it satisfies the projection axioms and we can construct the quasi-tree 𝒞K(𝕃)\mathcal{C}_{K}(\mathbb{L}) for K0K\gg 0, equipped with the length metric d𝒞d_{\mathcal{C}}.

By the following result, we shall be reduced to embed the action of G˙\dot{G} on (𝒳˙,d𝒳˙K)(\dot{\mathcal{X}},d_{\dot{\mathcal{X}}}^{K}) into finite product of quasi-trees. We warn the reader that d𝒳˙Kd_{\dot{\mathcal{X}}}^{K} is not a metric on 𝒳˙\dot{\mathcal{X}}. However, we can still talk about the product space 𝒳˙×𝒞K(𝕃)\dot{\mathcal{X}}\times\mathcal{C}_{K}(\mathbb{L}) equipped with the symmetric non-negative function d𝒳˙K×d𝒞d_{\dot{\mathcal{X}}}^{K}\times d_{\mathcal{C}}. The quasi-isometric embedding of 𝒳\mathcal{X} into the product only means the coarse bilipschitz inequality.

Proposition 5.7.

For any KK0K\gg K_{0}, there exists a G˙\dot{G}–equivariant quasi-isometric embedding from 𝒳{\mathcal{X}} to the product (𝒳˙×𝒞K(𝕃),d𝒳˙K×d𝒞)(\dot{\mathcal{X}}\times\mathcal{C}_{K}(\mathbb{L}),d_{\dot{\mathcal{X}}}^{K}\times d_{\mathcal{C}}).

Proof.

We first define the map Φ=Φ1×Φ2:𝒳𝒳˙×𝒞K(𝕃)\Phi=\Phi_{1}\times\Phi_{2}:\mathcal{X}\to\dot{\mathcal{X}}\times\mathcal{C}_{K}(\mathbb{L}).

If xY¯vx\in\overline{Y}_{v} define Φ1(x)=x\Phi_{1}(x)=x in 𝒳˙\dot{\mathcal{X}}. To define Φ2(x)\Phi_{2}(x), choose a (non-unique) closest quasi-line \ell in 𝕃v\mathbb{L}_{v} to xx and define Φ2(x)=π(x)\Phi_{2}(x)=\pi_{\ell}(x). The choice of \ell is not important and the important thing is the distance from oo to any such chosen \ell is uniformly bounded, thanks to co-compact action of HvH_{v} on Y¯v\overline{Y}_{v}. Extend by GG–equivariance Φ2(gx)=gπ(x)\Phi_{2}(gx)=g\pi_{\ell}(x) for all gGg\in G.

If xx lies in the flat links Fl(w)Fl(w) define Φ1(x)\Phi_{1}(x) to be the apex of Conesr(w)Cones_{r}(w) and Φ2(x)=π(x)\Phi_{2}(x)=\pi_{\ell}(x) where \ell is the binding line of Fl(w)Fl(w).

By Corollary 4.11, we have that

𝕃[dγ(x,y)]KKd𝒞(Φ2(x),Φ2(y)).\sum_{\ell\in\mathbb{L}}[d_{\gamma}(x,y)]_{K}\sim_{K}d_{\mathcal{C}}(\Phi_{2}(x),\Phi_{2}(y)).

The quasi-isometric embedding then follows from the distance formula (16) in Lemma 5.6. ∎

5.2. Construct the collection of quasi-lines in 𝒳˙\dot{\mathcal{X}}

The goal of this subsection is to introduce a collection 𝔸\mathbb{A} of quasi-lines so that a distance formula holds for (𝒳˙,d𝒳˙K)(\dot{\mathcal{X}},d_{\dot{\mathcal{X}}}^{K}).

Recall that 𝒳˙\dot{\mathcal{X}} is the hyperbolic cone-off space constructed from 𝒱{𝒱1,𝒱2}\mathcal{V}\in\{\mathcal{V}_{1},\mathcal{V}_{2}\}. According to Proposition 5.7, we are working in the cone-off space 𝒳˙\dot{\mathcal{X}} endowed with length metric d˙\dot{d}. In particular, all quasi-lines are understood with this metric, and boundary lines 𝕃v\mathbb{L}_{v} of Y¯v\overline{Y}_{v} are of bounded diameter so are not quasi-lines anymore in 𝒳˙\dot{\mathcal{X}}.

First of all, let us fix the constants used in the sequel.

Recall that for every v𝒱v\in\mathcal{V}, the cone-off space Y˙v\dot{Y}_{v} admits a co-bounded and acylindrical action of HvH_{v}. Thus, when talking about quasi-lines, we follow the Convention 4.16: quasi-lines are (λ,c)(\lambda,c)–quasi-geodesics in 𝒳i˙\dot{\mathcal{X}_{i}} and Y˙v\dot{Y}_{v}’s (isometrically embedded into the former), where λ,c>0\lambda,c>0 are given by Lemma 4.13 applied to those actions of HvH_{v} on Y˙v\dot{Y}_{v}.

If γ\gamma is a quasi-line in 𝒳˙\dot{\mathcal{X}}, denote by d˙γ(x,y)\dot{d}_{\gamma}(x,y) the d˙\dot{d}–diameter of the shortest projection of x,y𝒳˙x,y\in\dot{\mathcal{X}} to γ\gamma in 𝒳˙\dot{\mathcal{X}}. By Lemma 5.3, 𝒳˙\dot{\mathcal{X}} is δ\delta–hyperbolic for a constant δ>0\delta>0. The coning-off construction is crucial to obtain the uniform constant θ\theta in the next lemma.

Lemma 5.8.

There exists a constant θ>0\theta>0 depending on δ\delta (and also λ,c\lambda,c) with the following property: for any ((λ,c)(\lambda,c)–)quasi-lines α\alpha in Y˙v\dot{Y}_{v} and β\beta in Y˙v\dot{Y}_{v^{\prime}} with vv𝒱v\neq v^{\prime}\in\mathcal{V} we have diam𝒳1˙(πβ(α))θdiam_{\dot{\mathcal{X}_{1}}}(\pi_{\beta}(\alpha))\leq\theta.

Proof.

By the construction of 𝒳˙\dot{\mathcal{X}}, any geodesic from α\alpha to β\beta has to pass through the apex between Y˙v\dot{Y}_{v} and Y˙v\dot{Y}_{v^{\prime}}, and thus the shortest projection πβ(α)\pi_{\beta}(\alpha) is contained in the projection of the apex to β\beta. By hyperbolicity, there exists a constant θ\theta depending only on λ,c,δ\lambda,c,\delta such that the diameter of the projection of any point to every quasi-line is bounded above by θ\theta. The conclusion then follows. ∎

Fix K>max{4ξ,θ}K>\max\{4\xi,\theta\}. For each v𝒱v\in\mathcal{V}, there exist a HvH_{v}-finite collection of quasi-lines 𝔸v\mathbb{A}_{v} in Y˙v\dot{Y}_{v} and a constant N=N(𝔸v,K,δ)N=N(\mathbb{A}_{v},K,\delta) such that (11) holds.

Since G˙\dot{G} preserves 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2}, by Lemma 5.3, there are only finitely many G˙\dot{G}–orbits in {Y˙v:v𝒱}\{\dot{Y}_{v}:v\in\mathcal{V}\}, so we can assume furthermore 𝔸w=g𝔸v\mathbb{A}_{w}=g\mathbb{A}_{v} if w=gvw=gv for gG˙g\in\dot{G}. Then the collection 𝔸=v𝒱𝔸v\mathbb{A}=\cup_{v\in\mathcal{V}}\mathbb{A}_{v} is a G˙\dot{G}–invariant collection of quasi-lines in 𝒳˙\dot{\mathcal{X}}.

Recall that rr is the radius of the multicones in constructing 𝒳i˙\dot{\mathcal{X}_{i}} for i=1,2i=1,2, and TT is the Bass-Serre tree for admissible group GG.

Proposition 5.9.

For any x,y𝒳˙x,y\in\dot{\mathcal{X}}, the following holds

(18) d𝒳˙K(x,y)N,rγ𝔸[d˙γ(x,y)]K+dT(ρ(x),ρ(y)).\begin{array}[]{cc}d_{\dot{\mathcal{X}}}^{K}(x,y)\;\sim_{N,r}\;\sum_{\gamma\in\mathbb{A}}[\dot{d}_{\gamma}(x,y)]_{K}+d_{T}(\rho(x),\rho(y)).\end{array}
Proof.

The proof proceeds similarly as that of Proposition 4.20, so only the differences are spelled out. Assume that x,y𝒳˙x,y\in\dot{\mathcal{X}} are not in any hyperbolic multicones. We can then write the geodesic [x,y][x,y] as in (17) and keep notations there.

By the formula (11), summing up the lengths of geodesics γv\gamma_{v} in Y˙v\dot{Y}_{v} yields the upper bound in (18). The term dT(ρ(x),ρ(y))d_{T}(\rho(x),\rho(y)) appears since the notation \sim involves additive errors.

By the choice of K>θK>\theta and Lemma 5.8, we have [d˙γ(x,y)]K=0[\dot{d}_{\gamma}(x,y)]_{K}=0 for any quasi-line γ\gamma in 𝒱[ρ(x),ρ(y)]0\mathcal{V}-[\rho(x),\rho(y)]^{0}. The lower bound of (18) is obtained as well by summing up distances in the formula (11). A term rdT(ρ(x),ρ(y))r\cdot d_{T}(\rho(x),\rho(y)) is added, since [x,y][x,y] goes through dT(ρ(x),ρ(y))/2d_{T}(\rho(x),\rho(y))/2 hyperbolic cones with radius rr and each cwc_{w} is of length 2r2r. ∎

5.3. Reassembling finite index vertex groups

By Bass-Serre theory, the finite index subgroup G˙<G\dot{G}<G from Lemma 4.6 acts on the Bass-Serre tree of GG and can be represented as a finite graph 𝒢=T/G˙\mathcal{G}=T/\dot{G} of groups where the vertex subgroups are isomorphic to those of GG.

Let ee be an oriented edge in 𝒢\mathcal{G} from ee_{-} to e+e_{+} (it is possible that e=e+e_{-}=e_{+} because ee could be a loop) and e¯\overline{e} be the oriented edge with reversed orientation. A collection of finite index subgroups {Ge<Ge,Gv<Gv:v𝒢0,e𝒢1}\{G_{e}^{\prime}<G_{e},G_{v}^{\prime}<G_{v}:v\in\mathcal{G}^{0},e\in\mathcal{G}^{1}\} is called compatible if whenever v=ev=e_{-}, we have

GvGe=GvGe.G_{v}\cap G_{e}^{\prime}=G_{v}^{\prime}\cap G_{e}.

We shall make use of [DK18, Theorem 7.51] to obtain a finite index subgroup in G˙\dot{G} from a compatible collection of finite index subgroups. For this purpose, we assume the quotient HvH_{v} of each vertex group GvG_{v} for v𝒢0v\in\mathcal{G}^{0} is omnipotent in the sense of Wise.

In a group two elements are independent if they do not have conjugate powers. (see Definition 3.2 in [Wis00]).

Definition 5.10.

A group HH is omnipotent if for any set of pairwise independent elements {h1,,hr}\{h_{1},\cdots,h_{r}\} (r1r\geq 1) there is a number p1p\geq 1 such that for every choice of positive natural numbers {n1,,nr}\{n_{1},\cdots,n_{r}\} there is a finite quotient HH^H\to\hat{H} such that h^i\hat{h}_{i} has order nipn_{i}p for each ii.

Remark 5.11.

If HH is hyperbolic, two loxodromic elements h,hh,h^{\prime} are usually called independent if the collection of HH–translated quasi-axis of h,hh,h^{\prime} has the bounded projection property. When HH is torsion-free, it is equivalent to the notion of independence in the above sense.

Let gg be a loxodromic element in a hyperbolic group and E(g)E(g) be the maximal elementary group containing g\langle g\rangle. By [BH99, Ch. II Theorem 6.12], GvG_{v} contains a subgroup KvK_{v} intersecting trivially with Z(Gv)Z(G_{v}) so that the direct product Kv×Z(Gv)K_{v}\times Z(G_{v}) is a finite index subgroup. Thus, the image of KvK_{v} in Gv/Z(Gv)G_{v}/Z(G_{v}) is of finite index in HvH_{v} and KvK_{v} acts geometrically on hyperbolic spaces Y¯v\overline{Y}_{v}.

The following result will be used in the next subsection to obtain desired finite index subgroups.

Lemma 5.12.

Let {K˙v<Kv:v𝒢0}\{\dot{K}_{v}<K_{v}:v\in\mathcal{G}^{0}\} be a collection of finite index subgroups. Then there exists a compatible collection of finite index subgroups {Ge<Ge,Gv<Gv:v𝒢0,e𝒢1}\{G_{e}^{\prime}<G_{e},G_{v}^{\prime}<G_{v}:v\in\mathcal{G}^{0},e\in\mathcal{G}^{1}\} such that Gv=K¨v×G_{v}^{\prime}=\ddot{K}_{v}\times\mathbb{Z} is of finite index in K¨v×Z(Gv)\ddot{K}_{v}\times Z(G_{v}) for each v𝒢0v\in\mathcal{G}^{0}, where K¨v\ddot{K}_{v} is of finite index in K˙v\dot{K}_{v}.

Proof.

Let ee be an oriented edge in 𝒢\mathcal{G} from ee_{-} to e+e_{+} (it is possible that e=e+e_{-}=e_{+}) and e¯\overline{e} be the oriented edge with reversed orientation.

If v=ev=e_{-}, then the abelian group KvGeK_{v}\cap G_{e} is a nontrivial cyclic group contained in a maximal elementary E(be)E(b_{e}) in KvK_{v} where beb_{e} is a primitive loxodromic element. Similarly, for w=e+w=e_{+}, let be¯Kwb_{\overline{e}}\in K_{w} be a primitive loxodromic element in E(be¯)E(b_{\overline{e}}) containing KwGeK_{w}\cap G_{e}. Then beb_{e} and be¯b_{\overline{e}} preserve two lines respectively which are orthogonal in the Euclidean plane FeF_{e} and thus generate an abelian group G^e:=be,be¯\hat{G}_{e}:=\langle b_{e},b_{\overline{e}}\rangle of rank 2 so that GeG^eG_{e}\subset\hat{G}_{e} is of finite index.

Consider the collection of all oriented edges e1,,ere_{1},\dots,e_{r} in 𝒢1\mathcal{G}^{1} such that (ei)=v(e_{i})_{-}=v. Let {be1,,ber}\{b_{e_{1}},\dots,b_{e_{r}}\} be the set of primitive loxodromic elements in KvK_{v} obtained as above in correspondence with {e1,,er}\{e_{1},\dots,e_{r}\}. Note that {be1,,ber}\{b_{e_{1}},\dots,b_{e_{r}}\} are pairwise independent in HvH_{v}. This follows from the item (3) in Definition 2.1 of admissible group, otherwise GeiG_{e_{i}} and GejG_{e_{j}} would be commensurable for eieje_{i}\neq e_{j}.

By the finite index of GeG_{e} in G^e\hat{G}_{e}, there exists a set of powers of beib_{e_{i}}’s in K˙vGei\dot{K}_{v}\cap G_{e_{i}} denoted by {he1,,her}\{h_{e_{1}},\cdots,h_{e_{r}}\}. Since {he1,,her}\{h_{e_{1}},\cdots,h_{e_{r}}\} are still pairwise independent in HvH_{v}, the omnipotence of HvH_{v} gives the constant pvp_{v} by Definition 5.10. Let

s=v𝒢0pvs=\prod_{v\in\mathcal{G}^{0}}p_{v}

Define ni=spvn_{i}=\frac{s}{p_{v}} with i{1,r}i\in\{1,\dots r\}. By the omnipotence of HvH_{v} and restricting to K˙vHv\dot{K}_{v}\subset H_{v}, there is a finite index subgroup K¨v\ddot{K}_{v} of K˙v\dot{K}_{v} such that heipvni=heisK¨vh_{e_{i}}^{p_{v}n_{i}}=h_{e_{i}}^{s}\in\ddot{K}_{v}.

For each vertex vv in 𝒢\mathcal{G} and for each edge ee in 𝒢\mathcal{G}, we define

Gv:=K¨v×he¯sG^{\prime}_{v}:=\ddot{K}_{v}\times\langle h_{\overline{e}}^{s}\rangle

and

Ge:=hes×he¯s=s×sG^{\prime}_{e}:=\langle h_{e}^{s}\rangle\times\langle h_{\overline{e}}^{s}\rangle=s\mathbb{Z}\times s\mathbb{Z}

To conclude the proof, it remains to note the collection {Gv,Ge|v𝒢0,e𝒢1}\{G^{\prime}_{v},G^{\prime}_{e}\,\,|\,v\in\mathcal{G}^{0},e\in\mathcal{G}^{1}\} is compatible. It is obvious that GeiGvG^{\prime}_{e_{i}}\subset G^{\prime}_{v}, so GeiGvGeiG^{\prime}_{e_{i}}\leq G^{\prime}_{v}\cap G_{e_{i}}. Conversely, GvGeiGvG^ei(K¨vhes)×he¯sGeiG^{\prime}_{v}\cap G_{e_{i}}\subset G^{\prime}_{v}\cap\hat{G}_{e_{i}}\subset(\ddot{K}_{v}\cap\langle h_{e}^{s}\rangle)\times\langle h_{\overline{e}}^{s}\rangle\subset G^{\prime}_{e_{i}}. ∎

Corollary 5.13.

There is a finite index admissible group G0<G˙<GG_{0}<\dot{G}<G in the sense of Definition 2.1 where every vertex group are direct products of a hyperbolic group and \mathbb{Z}.

Proof.

By [DK18, Theorem 7.50], the compatible collection of finite index subgroups from Lemma 5.12 determines a finite index group G0<G˙<GG_{0}<\dot{G}<G. Indeed, G0G_{0} is the fundamental group of a finite covering space which are obtained from finite many copies of finite coverings in correspondence to Gv<GvG_{v}^{\prime}<G_{v} glueeing along edge spaces in correspondence to GeG_{e}^{\prime}. Thus, G0G_{0} splits over edge groups 2\mathbb{Z}^{2} as a finite graph of groups where the vertex groups are conjugates of Gv=K¨v×G_{v}^{\prime}=\ddot{K}_{v}\times\mathbb{Z}. In view of Definition of 2.1, it suffices to certify the non-commensurable edge groups adjacent to the same vertex group. This follows from the above proof of Lemma 5.12, where the edge groups are direct products of \mathbb{Z} with pairwise independent loxodromic elements. Thus, different edge groups are not commensurable. ∎

5.4. Partition 𝔸\mathbb{A} into sub-collections with good projection constants: completion of the proof

With purpose to prove Theorem 1.1, it suffices to prove property (QT) for a finite index subgroup of GG. By Corollary 5.13, we can assume that the CKA flip action of GG on XX satisfies that every vertex group are direct products Kv×K_{v}\times\mathbb{Z} where KvK_{v} is of finite index in HvH_{v}. Since HvH_{v} is omnipotent and then residually finite, without loss of generality we can assume that KvK_{v} is torsion-free.

Since the assumption of Lemma 5.3 is fulfilled, the results in Sections 5.1 and 5.2 hold: a finite index at most 2 subgroup G˙<G\dot{G}<G acts on the cone-off spaces 𝒳i˙\dot{\mathcal{X}_{i}} for i=1,2i=1,2 with distance formula.

Let 𝒱{𝒱1,𝒱2}\mathcal{V}\in\{\mathcal{V}_{1},\mathcal{V}_{2}\}. Let us recall the data we have now:

  1. (1)

    For every v𝒱v\in\mathcal{V}, 𝔸v\mathbb{A}_{v} is a KvK_{v}-finite collection of quasi-lines in Y˙v\dot{Y}_{v} so that the distance formula (11) holds for Y˙v\dot{Y}_{v}. (Lemma 5.5)

  2. (2)

    Let 𝔸=v𝒱𝔸v\mathbb{A}=\cup_{v\in\mathcal{V}}\mathbb{A}_{v} be the G˙\dot{G}-invariant collection of quasi-lines so that the formula (18) holds. (Proposition 5.9)

The first step is passing to a further finite index subgroup K˙v\dot{K}_{v} of KvK_{v} so that 𝔸v\mathbb{A}_{v} is partitioned into Kv˙\dot{K_{v}}-invariant sub-collections with projection constants ξ\xi. It follows closely the argument in [BBF19] which is presented below for completeness.

The constants θ\theta and ξ\xi: The constant θ>0\theta>0 is chosen so that it satisfies Proposition 4.14 and Lemma 5.8 simultaneously. Then ξ=ξ(θ)\xi=\xi(\theta) is given by Proposition 4.9.

Lemma 5.14.

Let 𝔸v\mathbb{A}_{v} be a KvK_{v}-finite collection of quasi-lines obtained as above by Lemma 4.17. Then there exists a finite index subgroup Kv˙<Kv\dot{K_{v}}<K_{v} such that any two distinct quasi-lines in the same K˙v\dot{K}_{v}-orbit have θ\theta-bounded projection.

Proof.

By construction, the quasi-lines in 𝔸v\mathbb{A}_{v} are quasi-axis of loxodromic elements whose maximal elementary groups are virtually cyclic. Recalling that KvK_{v} is torsion-free, the maximal elementary group is cyclic and thus E(g)E(g) is the centralizer C(g):={hKv:hg=gh}C(g):=\{h\in K_{v}:hg=gh\} of gg. By [BBF19, Lemma 2.1], since KvK_{v} is residually finite, then the centralizer of any element gKvg\in K_{v} is separable, i.e. the intersection of all finite index subgroups containing C(g)C(g).

Proposition 4.14 implies that E:={hKv:diam(πγ(hγ))θ}E:=\{h\in K_{v}:diam(\pi_{\gamma}(h\gamma))\geq\theta\} consists of finite double C(g)C(g)-cosets. Since C(g)C(g) is separable, we use the remark after Lemma 2.1 in [BBF19] to get a finite index K˙v<Kv\dot{K}_{v}<K_{v} such that EK˙v=E\cap\dot{K}_{v}=\varnothing. The proof is complete. ∎

The next step is re-grouping appropriately the collections of quasi-lines v𝒱𝔸v\cup_{v\in\mathcal{V}}\mathbb{A}_{v} in Lemma 5.14.

By [DK18, Theorem 7.51], the compatible collection of finite index subgroups K¨v<K˙v\ddot{K}_{v}<\dot{K}_{v} from Lemma 5.12 determines a finite index group G0<G˙<GG_{0}<\dot{G}<G such that G0Gv=Gv,G0Ge=GeG_{0}\cap G_{v}=G_{v}^{\prime},G_{0}\cap G_{e}=G_{e}^{\prime} and GvK¨v×Z(Gv)K˙v×Z(Gv)G_{v}^{\prime}\subset\ddot{K}_{v}\times Z(G_{v})\subset\dot{K}_{v}\times Z(G_{v}) for every vertex vv and edge ee.

By Bass-Serre theory, G0G_{0} acts on the Bass-Serre tree TT of GG with finitely many vertex orbits. To be precise, let {v0,,vm}\{v_{0},\cdots,v_{m}\} be the full set of vertex representatives.

Since for each 1im1\leq i\leq m, K¨viK˙vi\ddot{K}_{v_{i}}\subset\dot{K}_{v_{i}} is of finite index, Lemma 5.14 implies that 𝔸vi\mathbb{A}_{v_{i}} consists of finitely many K¨v\ddot{K}_{v}-orbits, say 𝔸ˇij(1jli)\check{\mathbb{A}}_{i}^{j}(1\leq j\leq l_{i}),

𝔸vi=j=1li𝔸ˇij,\mathbb{A}_{v_{i}}=\cup_{j=1}^{l_{i}}\check{\mathbb{A}}_{i}^{j},

each of which satisfies projection axioms with projection constant ξ\xi.

Recall that G0G˙G_{0}\subset\dot{G} acts on 𝒳1\mathcal{X}_{1} and 𝒳2\mathcal{X}_{2}. We now set 𝔸ij:=gG0g𝔸ˇij\mathbb{A}_{ij}:=\cup_{g\in G_{0}}g\check{\mathbb{A}}_{i}^{j} so we have

𝔸=i=1mj=1li𝔸ij.\mathbb{A}=\cup_{i=1}^{m}\cup_{j=1}^{l_{i}}\mathbb{A}_{ij}.

We summarize the above discussion as the following.

Proposition 5.15.

For each 𝒳{𝒳1,𝒳2}\mathcal{X}\in\{\mathcal{X}_{1},\mathcal{X}_{2}\}, there exists a finite partition 𝔸=𝔸1𝔸2𝔸n\mathbb{A}=\mathbb{A}_{1}\cup\mathbb{A}_{2}\cdots\cup\mathbb{A}_{n} where n=i=1lin=\sum_{i=1}l_{i} such that for each 1in1\leq i\leq n, 𝔸i\mathbb{A}_{i} is G0G_{0}–invariant and satisfies projection axioms with projection constant ξ\xi.

We are now ready to complete the proof of Theorem 1.1.

Proof of Theorem 1.1.

By [BBF19, Induction 2.2], if a finite index subgroup of GG has property (QT) then so does GG. Thus it suffices to show that G0G_{0} has property (QT) where G0G_{0} is the finite index subgroup of G˙<G\dot{G}<G given by Corollary 5.13. By abuse of notations, we denote G0G_{0} by GG, and we remark here that for the rest of the proof, results in Section 4 and Section 5 will apply for G=G0G=G_{0}, but not for the original GG.

By Proposition 4.8, G˙\dot{G} acts on the product 𝒳1×𝒳2\mathcal{X}_{1}\times\mathcal{X}_{2} so that the orbital map is quasi-isometrically embedded. Furthermore, there exists a G˙\dot{G}–equivariant quasi-isometric embedding of each 𝒳i\mathcal{X}_{i} (i=1,2)(i=1,2) into the product of the cone-off space (𝒳i˙,d𝒳˙K)(\dot{\mathcal{X}_{i}},d_{\dot{\mathcal{X}}}^{K}) and a quasi-tree by Proposition 5.7. Therefore, it suffices to establish a G0G_{0}–equivariant quasi-isometric embedding of (𝒳i˙,d𝒳˙K)(\dot{\mathcal{X}_{i}},d_{\dot{\mathcal{X}}}^{K}) into a finite product of quasi-trees.

By construction, each 𝔸ˇij(1jli)\check{\mathbb{A}}_{i}^{j}(1\leq j\leq l_{i}) is K¨vi\ddot{K}_{v_{i}}–invariant and K¨vi\ddot{K}_{v_{i}} acts co-boundedly on Y˙v\dot{Y}_{v}, so there exists some RR independent of i,ji,j so that the union of quasi-lines in 𝔸ˇij\check{\mathbb{A}}_{i}^{j} is RR–cobounded in Y˙vi\dot{Y}_{v_{i}}.

Let x𝒳x\in\mathcal{X} and xY˙vix\in\dot{Y}_{v_{i}}, we choose a point Φi(x)γ𝔸ˇijγ\Phi_{i}(x)\in\cup_{\gamma\in\check{\mathbb{A}}_{i}^{j}}\gamma for 1in1\leq i\leq n such that d(x,Φi(x))Rd(x,\Phi_{i}(x))\leq R. By G0G_{0}–equivariance we define Φi(gx)=gΦi(x)\Phi_{i}(gx)=g\Phi_{i}(x) for any gG0g\in G_{0}.

By Proposition 5.9, the formula (18) holds for any x,y𝒳˙x,y\in\dot{\mathcal{X}}. Note the sum

γ𝔸[d˙γ(x,y)]K=i=1nγ𝔸i[d˙γ(x,y)]K\sum_{\gamma\in\mathbb{A}}[\dot{d}_{\gamma}(x,y)]_{K}=\sum_{i=1}^{n}\sum_{\gamma\in\mathbb{A}_{i}}[\dot{d}_{\gamma}(x,y)]_{K}

For each 𝔸i\mathbb{A}_{i}, let 𝒞K(𝔸i)\mathcal{C}_{K}(\mathbb{A}_{i}) be the quasi-tree of quasi-lines and by Proposition 4.9

γ𝔸i[d˙γ(x,y)]Kd𝒞i(Φi(x),Φi(y))\sum_{\gamma\in\mathbb{A}_{i}}[\dot{d}_{\gamma}(x,y)]_{K}\sim d_{\mathcal{C}_{i}}(\Phi_{i}(x),\Phi_{i}(y))

Hence the formula (18) implies

(𝒳˙,d𝒳˙K)T×i=1n𝒞K(𝔸i)(\dot{\mathcal{X}},d_{\dot{\mathcal{X}}}^{K})\to T\times\prod_{i=1}^{n}\mathcal{C}_{K}(\mathbb{A}_{i})

is a G0G_{0}–equivariant quasi-isometric embedding. The proof of the Theorem is thus completed. ∎

6. Finite height subgroups in a CKA action GXG\curvearrowright X

In this section, we are going to prove Theorem 1.6 that basically says having finite height and strongly quasiconvexity are equivalent to each other in the context of CKA actions, and both properties can be characterized in term of their group elements. The heart of the proof of this theorem belongs to the implication (3)(1)(\ref{thm1:item3})\Rightarrow(\ref{thm1:item1}) where we use Sisto’s notion of path systems ([Sis18]). We first review some concepts finite height subgroups, strongly quasi-convex subgroups as well as some terminology in [Sis18].

Definition 6.1.

Let GG be a group and HH a subgroup of GG. We say that conjugates g1Hg11,gkHgk1g_{1}Hg_{1}^{-1},\cdots g_{k}Hg_{k}^{-1} are essentially distinct if the cosets g1H,,gkHg_{1}H,\cdots,g_{k}H are distinct. We call HH has height at most nn in GG if the intersection of any (n+1)(n+1) essentially distinct conjugates is finite. The least nn for which this is satisfied is called the height of HH in GG.

Definition 6.2 (Strongly quasiconvex, [Tra19]).

A subset YY of a geodesic space XX is called strongly quasiconvex if for every K1,C0K\geq 1,C\geq 0 there is some M=M(K,C)M=M(K,C) such that every (K,C)(K,C)–quasi–geodesic with endpoints on YY is contained in the MM–neighborhood of YY.

Let GG be a finitely generated group and HH a subgroup of GG. We say HH is strongly quasiconvex in GG if HH is a strongly quasi-convex subset in the Cayley graph Γ(G,S)\Gamma(G,S) for some (any) finite generating set SS. A group element gg in GG is Morse if gg is of infinite order and the cyclic subgroup generated by gg is strongly quasiconvex.

Remark 6.3.

The strong quasiconvexity of a subgroup does not depend on the choice of finite generating sets, and this notion is equivalent to quasiconvexity in the setting of hyperbolic groups. It is shown in [Tra19] (see Theorem 1.2) that strongly quasi-convex subgroups of a finitely generated group are finitely generated and have finite height.

The following proposition is cited from Proposition 2.3 and Proposition 2.6 in [NTY].

Proposition 6.4.
  1. (1)

    Let GG be a group such that the centralizer Z(G)Z(G) of GG is infinite. Let HH be a finite height infinite subgroup of GG. Then HH must have finite index in GG

  2. (2)

    Assume a group GG is decomposed as a finite graph TT of groups that satisfies the following.

    1. (a)

      For each vertex vv of TT each finite height subgroup of vertex group GvG_{v} must be finite or have finite index in GvG_{v}.

    2. (b)

      Each edge group is infinite.

    Then, if HH is a finite height subgroup of G of infinite index, then gHg1GvgHg^{-1}\cap G_{v} is finite for each vertex group GvG_{v} and each group element gg. In particular, if HH is torsion free, then HH is a free group.

Definition 6.5 (Path system, [Sis18]).

Let XX be a metric space. A path system 𝒫𝒮(X)\mathcal{PS}(X) in XX is a collection of (c,c)(c,c)–quasi-geodesic for some c1c\geq 1 such that any subpath of a path in 𝒫𝒮(X)\mathcal{PS}(X) is in 𝒫𝒮(X)\mathcal{PS}(X), and all pairs of points in XX can be connected by a path in 𝒫𝒮(X)\mathcal{PS}(X).

Definition 6.6 (𝒫𝒮\mathcal{PS}–contracting, [Sis18]).

Let XX be a metric space and let 𝒫𝒮(X)\mathcal{PS}(X) be a path system in XX. A subset AA of XX is called 𝒫𝒮(X)\mathcal{PS}(X)–contracting if there exists C>0C>0 and a map π:XA\pi\colon X\to A such that

  1. (1)

    For any xAx\in A, then d(x,π(x))Cd(x,\pi(x))\leq C

  2. (2)

    For any x,yXx,y\in X such that d(π(x),π(y))Cd(\pi(x),\pi(y))\geq C then for any path γ\gamma in 𝒫𝒮(X)\mathcal{PS}(X) connecting xx to yy then d(π(x),γ)Cd(\pi(x),\gamma)\leq C and d(π(y),γ)Cd(\pi(y),\gamma)\leq C.

The map π\pi will be called 𝒫𝒮(X)\mathcal{PS}(X)–projection on AA with constant CC.

Lemma 6.7.

[Sis18, Lemma 2.8] Let AA be a 𝒫𝒮(X)\mathcal{PS}(X)–contracting subset of a metric space XX, then AA is strongly quasi-convex.

Theorem 6.8.

Let GXG\curvearrowright X be a CKA action. Let 𝒫𝒮(X)\mathcal{PS}(X) be the collection of all special paths defined in Definition 3.6. Then (X,𝒫𝒮(X))(X,\mathcal{PS}(X)) is a path system.

Proof.

The proof follows from Proposition 3.8. ∎

For the rest of this section, we fix a CKA action GXG\curvearrowright X and GTG\curvearrowright T the action of GG on the associated Bass-Serre tree. We also fix the path system (X,𝒫𝒮(X))(X,\mathcal{PS}(X)) in Theorem 6.8.

To get into the proof of Theorem 1.6, we need several lemmas. The following lemma tells us that finite height subgroups in the CKA action GXG\curvearrowright X are virtually free.

Lemma 6.9.

Let KGK\leq G be a nontrivial finitely generated infinite index subgroup of GG. If KK has finite height in GG, then KgGvg1K\cap gG_{v}g^{-1} is finite for any vT0v\in T^{0} and gGg\in G. In particular, KK is virtually free.

Proof.

Suppose that KK has finite height in GG. Since the centralizer Z(Gv)Z(G_{v}) each each vertex group is isomorphic to \mathbb{Z}, it follows from Proposition 6.4 that for any gGg\in G and vT0v\in T^{0}, the intersection KgGvg1K\cap gG_{v}g^{-1} is finite. Thus, KK acts properly on the tree TT and the stabilizer in KK of each vertex in TT is finite. It follows from [DK18, Theorem 7.51] that KK is virtually free. ∎

Remark 6.10.

Let KGK\leq G be a nontrivial finitely generated infinite index subgroup of GG. Suppose that KK is a free group of finite rank and every nontrivial element in KK is not conjugate into any vertex group. Then there exists a subspace CKC_{K} of XX such that KK acts geometrically on CKC_{K} with respect to the induced length metric on CKC_{K}. The subspace CKC_{K} is constructed as the following.

Fix a vertex vv in TT, and fix a point x0x_{0} in YvY_{v} such that ρ(x0)=v\rho(x_{0})=v. Let {g1,g2,,gn}\{g_{1},g_{2},\dots,g_{n}\} be a generating set of KK. For each i{1,2,,n}i\in\{1,2,\dots,n\}, let gn+i=gi1g_{n+i}=g_{i}^{-1}. Let γj\gamma_{j} be the geodesic in XX connecting x0x_{0} to gj(x0)g_{j}(x_{0}) with j{1,2,,2n}j\in\{1,2,\dots,2n\}. Let CKC_{K} be the union of segment g(γj)g(\gamma_{j}) where gg varies elements of KK and j{1,,2n}j\in\{1,\dots,2n\}.

The following lemma is well-known (see Lemma 2.9 in [CK02] or Lemma 4.5 in [GM14] for proofs).

Lemma 6.11.

Let XX be a δ\delta–hyperbolic Hadamard space. Let γ1\gamma_{1} and γ2\gamma_{2} be two geodesic lines of XX such that γ1γ2=\partial_{\infty}\gamma_{1}\cap\partial_{\infty}\gamma_{2}=\varnothing. Let η\eta be a minimal geodesic segment between γ1\gamma_{1} to γ2\gamma_{2}. Then any geodesic segment running from γ1\gamma_{1} to γ2\gamma_{2} will pass within distance D=D(γ1,γ2)D=D(\gamma_{1},\gamma_{2}) of both endpoints of η\eta. Moreover, when d(γ1,γ2)>4δd(\gamma_{1},\gamma_{2})>4\delta then we may take D=2δD=2\delta.

Lemma 6.12 and Lemma 6.13 below are used in the proof of Proposition 6.15.

Lemma 6.12.

Given a constant μ>0\mu>0, there exists a constant r>0r>0 such that the following holds. Let ee and ee^{\prime} be two consecutive edges in TT with a common vertex vv. Let AA be a subset of YvY_{v} such that diam(A)μdiam(A)\leq\mu. Suppose that AFeA\cap F_{e}\neq\varnothing and AFeA\cap F_{e^{\prime}}\neq\varnothing.

Let [p,q][p,q] be the shortest path joining two lines :=Y¯vFe\ell:=\overline{Y}_{v}\cap F_{e} to :=Y¯vFe\ell^{\prime}:=\overline{Y}_{v}\cap F_{e^{\prime}} where pp\in\ell and qq\in\ell^{\prime}. For any xFeAx\in F_{e}\cap A and yFeAy\in F_{e^{\prime}}\cap A, let uu and vv be the projections of xx and yy into the lines \ell and \ell^{\prime} respectively. Then d(u,p)rd(u,p)\leq r and d(v,q)rd(v,q)\leq r.

Proof.

We recall that Yv=Y¯v×Y_{v}=\overline{Y}_{v}\times\mathbb{R} and HvH_{v} acts properly and cocompactly on Y¯v\overline{Y}_{v}. Since HvH_{v} is a nonelementary hyperbolic group, it follows that Y¯v\overline{Y}_{v} is a δv\delta_{v}–hyperbolic space for some δv0\delta_{v}\geq 0. Let δ\delta be the maximum of the hyperbolicity constants of the Y¯v\overline{Y}_{v}’s.

Let D=D(,)>0D=D(\ell,\ell^{\prime})>0 be the constant given by Lemma 6.11. Let r(e,e):=4D+μr(e,e^{\prime}):=4D+\mu. Since =\partial_{\infty}\ell\cap\partial_{\infty}\ell^{\prime}=\varnothing and uu\in\ell, vv\in\ell^{\prime}, it follows from Lemma 6.11 that there exist p,q[u,v]p^{\prime},q^{\prime}\in[u,v] such that d(p,p)Dd(p,p^{\prime})\leq D and d(q,q)Dd(q,q^{\prime})\leq D. By the triangle inequality, we have d(u,p)+d(p,q)+d(q,v)4D+d(u,v)d(u,p)+d(p,q)+d(q,v)\leq 4D+d(u,v). Since uu and vv are projection points of xx and yy into the factor Y¯v\overline{Y}_{v} of Yv=Y¯v×Y_{v}=\overline{Y}_{v}\times\mathbb{R} respectively, it follows that d(u,v)d(x,y)d(u,v)\leq d(x,y). Since x,yAx,y\in A and diam(A)μdiam(A)\leq\mu, it follows that d(x,y)μd(x,y)\leq\mu. Hence d(u,v)d(x,y)μd(u,v)\leq d(x,y)\leq\mu. Thus, d(u,p)4D+d(u,v)4D+μ=r(e,e)d(u,p)\leq 4D+d(u,v)\leq 4D+\mu=r(e,e^{\prime}) and d(v,q)4D+d(u,v)4D+μ=r(e,e)d(v,q)\leq 4D+d(u,v)\leq 4D+\mu=r(e,e^{\prime}).

By Lemma 6.11, we note that whenever the distance between two lines Y¯vFe\overline{Y}_{v}\cap F_{e} and Y¯vFe\overline{Y}_{v}\cap F_{e^{\prime}} is at least 4δ4\delta then we can define r(e,e)=8δ+μr(e,e^{\prime})=8\delta+\mu. We remark here that module GG there are only finitely many cases d(Fe,Fe)<4δd(F_{e},F_{e^{\prime}})<4\delta. Thus there are only finitely many r(e,e)r(e,e^{\prime}) up to the action of GG. Let rr be the maximum of these constants. ∎

Lemma 6.13.

Let KGK\leq G be a finitely generated, finite height subgroup of GG of infinite index. Let CKC_{K} be the subspace of XX given by Remark 6.10. Then there exists a constant R>0R>0 such that if γ\gamma is a special path in XX (see Definition 3.6) connecting two points in CKC_{K} then γNR(CK)\gamma\subset{N}_{R}(C_{K}).

Proof.

By the construction of CKC_{K}, we note that there exists a constant μ>0\mu>0 such that diam(CKXv)<μdiam(C_{K}\cap X_{v})<\mu for any vertex v𝒱(T)v\in\mathcal{V}(T). Let rr be the constant given by Lemma 6.12.

Recall that we choose a GG–equivariant family of Euclidean planes {Fe:FeYe}eT1\{F_{e}:F_{e}\subset Y_{e}\}_{e\in T^{1}}. Let {𝒮ee}\{\mathcal{S}_{ee^{\prime}}\} be the collection of strips in XX given by Section 2. For any three consecutive edges e,e,e′′e,e^{\prime},e^{\prime\prime} in the tree TT, two lines 𝒮eeFe\mathcal{S}_{ee^{\prime}}\cap F_{e^{\prime}} and 𝒮ee′′Fe\mathcal{S}_{e^{\prime}e^{\prime\prime}}\cap F_{e^{\prime}} in the plane FeF_{e^{\prime}} determine an angle in (0,π)(0,\pi). However, there are only finitely many angles shown up. We denote these angles by θ1,,θk\theta_{1},\dots,\theta_{k}.

Let DD be the constant given by Lemma 2.4 such that Xv=ND(Yv)X_{v}={N}_{D}(Y_{v}) for every vertex vT0v\in T^{0}. Let

ξ=2μ+r+max{2μ+rsin(θj)+2μ+rsin(πθj)|j{1,,k}andθjπ/2}\xi=2\mu+r+\max\bigl{\{}\,{\frac{2\mu+r}{\sin(\theta_{j})}+\frac{2\mu+r}{\sin(\pi-\theta_{j})}}\bigm{|}{j\in\{1,\dots,k\}\,\textup{and}\,\theta_{j}\neq\pi/2}\,\bigr{\}}

and

R=2r+μ+2ξ+DR=2r+\mu+2\xi+D

Let xx and yy be the initial and terminal points of γ\gamma. We note that xXρ(x)x\in X_{\rho(x)} and yXρ(y)y\in X_{\rho(y)}. We consider the following cases:

Case 1: ρ(x)=ρ(y)\rho(x)=\rho(y). In this case, the special path γ\gamma is the geodesic in XX connecting xx to yy. Since x,yCKXρ(x)x,y\in C_{K}\cap X_{\rho(x)} and diam(CKXρ(x))μdiam(C_{K}\cap X_{\rho(x)})\leq\mu, it follows that Len(γ)=d(x,y)μ<RLen(\gamma)=d(x,y)\leq\mu<R. Thus, γNR(CKXρ(x))NR(CK)\gamma\subset{N}_{R}(C_{K}\cap X_{\rho(x)})\subset{N}_{R}(C_{K}).

Case 2: ρ(x)ρ(y)\rho(x)\neq\rho(y). Since Xu=ND(Yu)X_{u}={N}_{D}(Y_{u}) for any vertex uT0u\in T^{0}, hence without losing of generality, we can assume that xYρ(x)x\in Y_{\rho(x)} and yYρ(y)y\in Y_{\rho(y)}. We recall the construction of the path γ\gamma from Definition 3.6. Let e1ene_{1}\cdots e_{n} be the geodesic edge path connecting ρ(x)\rho(x) to ρ(y)\rho(y) and let piFeip_{i}\in F_{e_{i}} be the intersection point of the strips 𝒮ei1ei\mathcal{S}_{e_{i-1}e_{i}} and 𝒮eiei+1\mathcal{S}_{e_{i}e_{i+1}}, where e0:=xe_{0}:=x and en+1:=ye_{n+1}:=y. Then

γ=[x,p1][p1,p2][pn1,pn][pn,y]\gamma=[x,p_{1}][p_{1},p_{2}]\cdots[p_{{n-1}},p_{n}][p_{n},y]

Let p0:=xp_{0}:=x and pn+1:=yp_{n+1}:=y. In order to prove that γNR(CK)\gamma\subset{N}_{R}(C_{K}), we only need to show that [pi,pi+1]NR(CK)[p_{i},p_{i+1}]\subset{N}_{R}(C_{K}) with i{1,,n}i\in\{1,\dots,n\}.

The proofs for the cases i=0i=0 and i=ni=n and for the cases i=1,,n1i=1,\dots,n-1 are similar, so we only need to give the proofs for the cases i=0,1i=0,1.

Proof of case i=0i=0:

Let viv_{i} be the initial vertex of eie_{i} (with i{1,,n}i\in\{1,\dots,n\}), and vnv_{n} be the terminal vertex of ene_{n}. We recall that two lines 𝒮xe1Fe1\mathcal{S}_{xe_{1}}\cap F_{e_{1}} and Fe1Y¯v0F_{e_{1}}\cap\overline{Y}_{v_{0}} in the plane Fe1F_{e_{1}} are perpendicular. Since CKFe1C_{K}\cap F_{e_{1}}\neq\varnothing, we choose a point O1YFe1O_{1}\in Y\cap F_{e_{1}}.

Claim: d(O1,p1)<r+ξd(O_{1},p_{1})<r+\xi.

Proof of the claim.

Let O¯1\overline{O}_{1} be the projection of O1O_{1} into the line Fe1Y¯v0F_{e_{1}}\cap\overline{Y}_{v_{0}}. Let V¯1\overline{V}_{1} be the projection of O1O_{1} into the line Fe1Y¯v1F_{e_{1}}\cap\overline{Y}_{v_{1}}. By Lemma 6.12, we have

(19) d(V¯1,𝒮e1e2Fe1Y¯v1)rd(\overline{V}_{1},\mathcal{S}_{e_{1}e_{2}}\cap F_{e_{1}}\cap\overline{Y}_{v_{1}})\leq r

(we note that 𝒮e1e2Fe1Y¯v1=γe1e2(0)\mathcal{S}_{e_{1}e_{2}}\cap F_{e_{1}}\cap\overline{Y}_{v_{1}}=\gamma_{e_{1}e_{2}}(0)). Since O1O_{1} and p0=xp_{0}=x belong to Xv0CKX_{v_{0}}\cap C_{K} and diam(Xv0CK)μdiam(X_{v_{0}}\cap C_{K})\leq\mu, it follows that d(O1,p0)μd(O_{1},p_{0})\leq\mu. Let p¯0\overline{p}_{0} be the projection of p0p_{0} into the factor Y¯v0\overline{Y}_{v_{0}} of Yv0=Y¯v0×Y_{v_{0}}=\overline{Y}_{v_{0}}\times\mathbb{R}. We have that d(O¯1,p¯0)d(O1,p0)μd(\overline{O}_{1},\overline{p}_{0})\leq d(O_{1},p_{0})\leq\mu. Since d(p¯0,𝒮xe1Fe1Y¯v0)d(\overline{p}_{0},\mathcal{S}_{xe_{1}}\cap F_{e_{1}}\cap\overline{Y}_{v_{0}}) is the minimal distance from p¯0\overline{p}_{0} to the line Fe1Y¯v0F_{e_{1}}\cap\overline{Y}_{v_{0}} and O¯1Fe1Y¯v0\overline{O}_{1}\in F_{e_{1}}\cap\overline{Y}_{v_{0}} we have that d(p¯0,𝒮xe1Fe1Y¯v0)d(p¯0,O¯1)μd(\overline{p}_{0},\mathcal{S}_{xe_{1}}\cap F_{e_{1}}\cap\overline{Y}_{v_{0}})\leq d(\overline{p}_{0},\overline{O}_{1})\leq\mu. Using the triangle inequality for three points p¯0\overline{p}_{0}, O¯1\overline{O}_{1}, and 𝒮xe1Fe1Y¯v0\mathcal{S}_{xe_{1}}\cap F_{e_{1}}\cap\overline{Y}_{v_{0}}, we have

(20) d(O¯1,𝒮xe1Fe1Y¯v0)2μd(\overline{O}_{1},\mathcal{S}_{xe_{1}}\cap F_{e_{1}}\cap\overline{Y}_{v_{0}})\leq 2\mu

Let AA be the projection of O1O_{1} into the line Fe1𝒮e1e2F_{e_{1}}\cap\mathcal{S}_{e_{1}e_{2}}. Using formula (19), we have

(21) d(O1,A)=d(V1,𝒮e1e2Fe1Y¯v1)rd(O_{1},A)=d(V_{1},\mathcal{S}_{e_{1}e_{2}}\cap F_{e_{1}}\cap\overline{Y}_{v_{1}})\leq r

Let TT be the projection of O1O_{1} into the line 𝒮xe1Fe1\mathcal{S}_{xe_{1}}\cap F_{e_{1}}. Using formula (20), we have d(O1,T)=d(O¯1,𝒮xe1Fe1Y¯v0)2μd(O_{1},T)=d(\overline{O}_{1},\mathcal{S}_{xe_{1}}\cap F_{e_{1}}\cap\overline{Y}_{v_{0}})\leq 2\mu. Thus, we have d(A,T)d(A,O1)+d(O1,T)r+2μd(A,T)\leq d(A,O_{1})+d(O_{1},T)\leq r+2\mu. An easy application of Rule of Sines to the triangle Δ(T,p1,A)\Delta(T,p_{1},A) together with the fact d(A,T)2μ+rd(A,T)\leq 2\mu+r give us that d(p1,A)<ξd(p_{1},A)<\xi and d(p1,T)<ξd(p_{1},T)<\xi. Combining these inequalities with formula (21), we obtain that d(O1,p1)d(O1,A)+d(A,p1)<r+ξd(O_{1},p_{1})\leq d(O_{1},A)+d(A,p_{1})<r+\xi The claim is verified. ∎

Using the facts d(O1,p0)μd(O_{1},p_{0})\leq\mu and d(O1,p1)<r+ξd(O_{1},p_{1})<r+\xi we have d(p0,p1)d(p0,O1)+d(O1,p1)μ+r+ξ<Rd(p_{0},p_{1})\leq d(p_{0},O_{1})+d(O_{1},p_{1})\leq\mu+r+\xi<R. Since p0=xCKp_{0}=x\in C_{K}, it follows that [p0,p1]NR(CK)[p_{0},p_{1}]\subset{N}_{R}(C_{K}).

Proof of the case i=1i=1:

Since CKFe2C_{K}\cap F_{e_{2}}\neq\varnothing, we choose a point O2CKFe2O_{2}\in C_{K}\cap F_{e_{2}}. Since O1,O2O_{1},O_{2} belong to CKYv1C_{K}\cap Y_{v_{1}} and diam(CKYv1)μdiam(C_{K}\cap Y_{v_{1}})\leq\mu, we have d(O1,O2)μd(O_{1},O_{2})\leq\mu. By a similar argument as in the proof of the claim of the case i=0i=0, we can show that d(O2,p2)<r+ξd(O_{2},p_{2})<r+\xi. Thus, d(p1,p2)d(p1,O1)+d(O1,O2)+d(O2,p2)<(r+ξ)+μ+(r+ξ)=2r+μ+2ξd(p_{1},p_{2})\leq d(p_{1},O_{1})+d(O_{1},O_{2})+d(O_{2},p_{2})<(r+\xi)+\mu+(r+\xi)=2r+\mu+2\xi Since O1,O2CKO_{1},O_{2}\in C_{K}, it is easy to see that [p1,p2]N3r+μ+3ξ(Y)NR(Y)[p_{1},p_{2}]\subset{N}_{3r+\mu+3\xi}(Y)\subset{N}_{R}(Y). ∎

We recall that an infinite order element gg in a finitely generated group is Morse if the cyclic subgroup generated by gg is strongly quasi-convex.

Lemma 6.14.

If an infinite order element gg in GG is More, then it is not conjugate into any vertex group of GG.

Proof.

Since gg is Morse, it follows that the infinite cyclic subgroup g\langle g\rangle generated by gg is strongly quasi-convex in GG. We would like to show that gg is not conjugate into any vertex group. Indeed, by way of contradiction, we assume that gxGvx1g\in xG_{v}x^{-1} for some xGx\in G and for some vertex group GvG_{v}. Hence, the cyclic subgroup generated by h=x1gxh=x^{-1}gx is strongly quasi-convex in GG. Since GvG_{v} is undistorted in GG (as GvG_{v} acts geometrically on YvY_{v} and YvY_{v} is undistorted in XX), it follows from Proposition 4.11 in [Tra19] that h\langle h\rangle is strongly quasi-convex in GvG_{v}. By Theorem 1.2 in [Tra19], h\langle h\rangle has finite height in GvG_{v}. Since the centralizer Z(Gv)Z(G_{v}) of GvG_{v} is isomorphic to \mathbb{Z}, it follows from Proposition 6.4 that h\langle h\rangle has finite index in GvG_{v}. This contradicts to the fact that GvG_{v} is not virtually cyclic group. Therefore gg is not conjugate into any vertex group of GG. ∎

Proposition 6.15.

Let KK be a finitely generated free subgroup of GG of infinite index such that all nontrivial elements in KK are Morse in GG. Choose a vertex vv in a minimal KK–invariant subtree TT^{\prime} of TT. Let CKC_{K} be the subspace of XX given by Remark 6.10 with respect to a generating set {g1,g2,,gn}\{g_{1},g_{2},\dots,g_{n}\} of KK. Then CKC_{K} is contracting in (X,𝒫𝒮(X))(X,\mathcal{PS}(X)). As a consequence, KK is strongly quasi-convex in GG.

Proof.

We first recall the construction of CKC_{K} from Remark 6.10. We first fix a point x0x_{0} in YvY_{v} such that ρ(x0)=v\rho(x_{0})=v. For each i{1,2,,n}i\in\{1,2,\dots,n\}, let gn+i=gi1g_{n+i}=g_{i}^{-1}. Let γj\gamma_{j} be the geodesic in XX connecting x0x_{0} to gj(x0)g_{j}(x_{0}) with j{1,2,,2n}j\in\{1,2,\dots,2n\}. Then CKC_{K} is the union of segment g(γj)g(\gamma_{j}) where gg varies elements of KK and j{1,,2n}j\in\{1,\dots,2n\}. Since KK is a free subgroup of GG and all nontrivial elements in KK are Morse in GG, it follows from Lemma 6.14 that every nontrivial element in KK is not conjugate into any vertex group GvG_{v}. Hence, KK acts freely on the Bass-Serre tree TT. To show that CKC_{K} (we note that K(x0)CKK(x_{0})\subset C_{K}) is contracting in (X,𝒫𝒮(X))(X,\mathcal{PS}(X)), we need to define a 𝒫𝒮(X)\mathcal{PS}(X)–projection π:XCK\pi\colon X\to C_{K} satisfying conditions (1) and (2) in Definition 6.6.

Step 1: Constructing 𝒫𝒮(X)\mathcal{PS}(X)–projection π:XCK\pi\colon X\to C_{K} on CKC_{K}.

Let ρ:XT0\rho\colon X\to T^{0} be the indexed GG-map given by Remark 2.6, which is coarsely LL-lipschitez. Let :TT\mathcal{R}\colon T\to T^{\prime} be the nearest point projection from TT to the minimal KK-invariant subtree TT^{\prime}. Since KK acts on the minimal tree TT^{\prime} cocompactly, it follows that there exists a constant δ>0\delta^{\prime}>0 such that TNδ(K(v))T^{\prime}\subset N_{\delta^{\prime}}(K(v)).

Let xx be any point in XX. Choose an element gKg\in K such that (ρ(x))B(g(v),δ)\mathcal{R}(\rho(x))\in B(g(v),\delta^{\prime}). We define π(x):=g(x0)\pi(x):=g(x_{0}).

Step 2: Verifying the condition (1) in Definition 6.6. Recall that γi\gamma_{i} is the geodesic in XX connecting x0x_{0} to gi(x0)g_{i}(x_{0}). Let δ=max{Len(γi)|i{1,,n}}\delta=\max\bigl{\{}\,{Len(\gamma_{i})}\bigm{|}{i\in\{1,\dots,n\}}\,\bigr{\}}. Let μ>0\mu>0 be a constant such that diam(CKYu)μdiam(C_{K}\cap Y_{u})\leq\mu for any vertex uT0u\in T^{0}. Let RR be the constant given by Lemma 6.13. Since KK acts cocompactly both on CKC_{K} and TT^{\prime}, there exists a constant ϵ1\epsilon\geq 1 such that for any hh and hh^{\prime} in KK then

(22) d(h(x0),h(x0))/ϵϵdT(h(v),h(v))ϵd(h(x0),h(x0))+ϵd\bigl{(}h(x_{0}),h^{\prime}(x_{0})\bigr{)}\,\bigl{/}\epsilon-\epsilon\leq d_{T}\bigl{(}h(v),h^{\prime}(v)\bigr{)}\leq\epsilon d\bigl{(}h(x_{0}),h^{\prime}(x_{0})\bigr{)}+\epsilon

Claim 1: Let CC be a sufficiently large constant such that δ+ϵ(ϵ+2(Lδ+L)+δ)+(5+2δ)μ+R<C\delta+\epsilon(\epsilon+2(L\delta+L)+\delta^{\prime})+(5+2\delta^{\prime})\mu+R<C. Then d(x,π(x))Cd(x,\pi(x))\leq C for any xCKx\in C_{K}.

Indeed, since CKNδ(K(x0))C_{K}\subset{N}_{\delta}(K(x_{0})), there exists kKk\in K such that d(x,k(x0))δd(x,k(x_{0}))\leq\delta. Recalling that ρ\rho is a KK-equivariant, coarsely LL-lipschitez map, and v=ρ(x0)v=\rho(x_{0}), we have

dT(ρ(x),k(v))=d(ρ(x),ρ(k(x0))Ld(x,k(x0))+LLδ+L.d_{T}(\rho(x),k(v))=d(\rho(x),\rho(k(x_{0}))\leq Ld(x,k(x_{0}))+L\leq L\delta+L.

The nearest point projection of :TT\mathcal{R}\colon T\to T^{\prime} implies dT(ρ(x),(ρ(x)))dT(ρ(x),k(v))Lδ+Ld_{T}(\rho(x),\mathcal{R}(\rho(x)))\leq d_{T}(\rho(x),k(v))\leq L\delta+L. Hence, dT(ρ(x),g(v))dT(ρ(x),(ρ(x)))+d((ρ(x)),g(v))(Lδ+L)+δd_{T}(\rho(x),g(v))\leq d_{T}(\rho(x),\mathcal{R}(\rho(x)))+d(\mathcal{R}(\rho(x)),g(v))\leq(L\delta+L)+\delta^{\prime}. It implies that dT(k(v),g(v))dT(k(v),ρ(x))+dT(ρ(x),g(v))2(Lδ+L)+δd_{T}(k(v),g(v))\leq d_{T}(k(v),\rho(x))+d_{T}(\rho(x),g(v))\leq 2(L\delta+L)+\delta^{\prime}. Putting the above inequalities together with formula (22), we have

d(x,π(x))=d(x,g(x0))\displaystyle d(x,\pi(x))=d(x,g(x_{0})) d(x,k(x0))+d(k(x0),g(x0))δ+ϵ(ϵ+d(k(v),g(v)))\displaystyle\leq d(x,k(x_{0}))+d(k(x_{0}),g(x_{0}))\leq\delta+\epsilon(\epsilon+d(k(v),g(v)))
δ+ϵ(ϵ+2(Lδ+L)+δ)<C\displaystyle\leq\delta+\epsilon(\epsilon+2(L\delta+L)+\delta^{\prime})<C

Claim 1 is confirmed.

Step 3: Verifying condition (2) in Definition 6.6.

Claim 2: Let CC be the constant given by Claim 1. Then the projection π:XCK\pi\colon X\to C_{K} satisfies condition (2) in Definition 6.6 with respect to this constant CC.

Let xx and yy be two points in XX such that d(π(x),π(y))Cd(\pi(x),\pi(y))\geq C. Let γ\gamma be a special path in XX connecting xx to yy. We would like to show that d(π(x),γ)Cd(\pi(x),\gamma)\leq C and d(π(y),γ)Cd(\pi(y),\gamma)\leq C.

We recall that Xu=ND(Yu)X_{u}={N}_{D}(Y_{u}) for any vertex uT0u\in T^{0}. Thus we assume, without loss of generality that xYρ(x)x\in Y_{\rho(x)} and yYρ(y)y\in Y_{\rho(y)}. Recall that π(x)=g(x0)\pi(x)=g(x_{0}) and π(y)=g(x0)\pi(y)=g^{\prime}(x_{0}) where g,gKg,g^{\prime}\in K such that (ρ(x))\mathcal{R}(\rho(x)) and (ρ(y))\mathcal{R}(\rho(y)) are in the balls B(g(v),δ)B(g(v),\delta^{\prime}) and B(g(v),δ)B(g^{\prime}(v),\delta^{\prime}) respectively. We have ϵ2+(20+4δ)ϵ<Cd(π(x),π(y))=d(g(x0),g(x0))ϵ(ϵ+dT(g(v),g(v))\epsilon^{2}+(20+4\delta^{\prime})\epsilon<C\leq d(\pi(x),\pi(y))=d(g(x_{0}),g^{\prime}(x_{0}))\leq\epsilon(\epsilon+d_{T}(g(v),g^{\prime}(v)). Hence 20+4δ<dT(g(v),g(v))20+4\delta^{\prime}<d_{T}(g(v),g^{\prime}(v)). Since (ρ(x))B(g(v),δ)\mathcal{R}(\rho(x))\in B(g(v),\delta^{\prime}) and (ρ(y))B(g(v),δ)\mathcal{R}(\rho(y))\in B(g^{\prime}(v),\delta^{\prime}), we have dT((ρ(x)),(ρ(y)))>20+2δd_{T}(\mathcal{R}(\rho(x)),\mathcal{R}(\rho(y)))>20+2\delta^{\prime}.

Choose vertices σ\sigma, τ\tau in the geodesic [(ρ(x)),(ρ(y))][\mathcal{R}(\rho(x)),\mathcal{R}(\rho(y))] such that dT(σ,(ρ(x)))d_{T}(\sigma,\mathcal{R}(\rho(x))) and dT(τ,(ρ(y)))d_{T}(\tau,\mathcal{R}(\rho(y))) are the smallest integers bigger than or equal to 3+δ3+\delta^{\prime}. Thus dT(g(v),σ)dT((ρ(x)),σ)dT((ρ(x)),g(v))dT((ρ(x)),σ)δ3d_{T}(g(v),\sigma)\geq d_{T}(\mathcal{R}(\rho(x)),\sigma)-d_{T}(\mathcal{R}(\rho(x)),g(v))\geq d_{T}(\mathcal{R}(\rho(x)),\sigma)-\delta^{\prime}\geq 3. Similarly, we have dT(g(v),τ)3d_{T}(g^{\prime}(v),\tau)\geq 3.

Let α\alpha be a special path in XX connecting g(x0)Xg(v)g(x_{0})\in X_{g(v)} to g(x0)Xg(v)g^{\prime}(x_{0})\in X_{g^{\prime}(v)}. We remark here that there exist a subpath γ\gamma^{\prime} of the special path γ\gamma and a subpath α\alpha^{\prime} of α\alpha such that γ\gamma^{\prime} and α\alpha^{\prime} connect some point in XσX_{\sigma} to some point in XτX_{\tau}. By Remark 3.7, we have that αXu=γXu\alpha^{\prime}\cap X_{u}=\gamma^{\prime}\cap X_{u} for any vertex uu in the geodesic [σ,τ][\sigma,\tau].

Let RR be the constant given by Lemma 6.13. It follows from Lemma 6.13( see Case 1 and Case 2 in this lemma) that αYσNR(CKYσ)\alpha\cap Y_{\sigma}\subset{N}_{R}(C_{K}\cap Y_{\sigma}). Choose zαYσz\in\alpha\cap Y_{\sigma} and zCKXσz^{\prime}\in C_{K}\cap X_{\sigma} such that d(z,z)Rd(z,z^{\prime})\leq R. Since dT(g(v),σ)dT(g(v),(ρ(x)))+dT((ρ(x)),σ)δ+(4+δ)=2δ+4d_{T}(g(v),\sigma)\leq d_{T}(g(v),\mathcal{R}(\rho(x)))+d_{T}(\mathcal{R}(\rho(x)),\sigma)\leq\delta^{\prime}+(4+\delta^{\prime})=2\delta^{\prime}+4, and g(x0)CKXg(v)g(x_{0})\in C_{K}\cap X_{g(v)} and zCKXσz^{\prime}\in C_{K}\cap X_{\sigma}, it follows that d(g(x0),z)(5+2δ)μd(g(x_{0}),z^{\prime})\leq(5+2\delta^{\prime})\mu. Hence,

d(π(x),γ)d(π(x),z)d(π(x),z)+R=d(g(x0),z)+R(5+2δ)μ+R<Cd(\pi(x),\gamma)\leq d(\pi(x),z)\leq d(\pi(x),z^{\prime})+R=d(g(x_{0}),z^{\prime})+R\leq(5+2\delta^{\prime})\mu+R<C

Similarly, we can show that d(π(y),γ)<Cd(\pi(y),\gamma)<C. Thus Claim 2 is established.

Combining Step 1, Step 2, and Step 3 together, we conclude that CKC_{K} is contracting in (X,𝒫𝒮(X)(X,\mathcal{PS}(X). By Lemma 6.7, CKC_{K} is strongly quasi-convex in XX. We conclude that KK is strongly quasi-convex in GG. The proposition is proved. ∎

Proposition 6.15 has the following corollary that characterize Morse elements and contracting element in the admissible group GG.

Corollary 6.16.
  1. (1)

    An infinite order element gg in GG is Morse if and only if gg is not conjugate into any vertex group GvG_{v}.

  2. (2)

    An element of GG is contracting with respect to (X,𝒫𝒮(X)(X,\mathcal{PS}(X) if and only if its acts hyperbolically on the Bass-Serre tree TT.

Proof.

(1). By Lemma 6.14, if gg is Morse in GG then it is not conjugate into vertex group of GG. Conversely, if gg is not conjugate into any vertex group of GG then gg acts hyperbolically on the Bass-Serre tree TT. If gGg\in G acts hyperbolically on the Bass-Serre tree TT. Let KK be the infinite cyclic subgroup generated by gg. Since gGg\in G acts hyperbolically on the Bass-Serre tree TT, it is not conjugate into any vertex subgroup. Fix a point x0x_{0} in XX, by Proposition 6.15 the orbit space K(x0)K(x_{0}) is contracting in (X,𝒫𝒮(X))(X,\mathcal{PS}(X)) (because CKC_{K} is contracting in (X,𝒫𝒮(X))(X,\mathcal{PS}(X))). Thus gg is a contracting element with respect to (X,𝒫𝒮(X))(X,\mathcal{PS}(X)). By Lemma 2.8 in [Sis18], gg must be Morse.

(2). If gGg\in G acts hyperbolically on the Bass-Serre tree TT. Let KK be the infinite cyclic subgroup generated by gg. Since gGg\in G acts hyperbolically on the Bass-Serre tree TT, it is not conjugate into any vertex subgroup. Fix a point x0x_{0} in XX, by Proposition 6.15 the orbit space Kx0K\cdot x_{0} is contracting in (X,𝒫𝒮(X))(X,\mathcal{PS}(X)) (because CKC_{K} is contracting in (X,𝒫𝒮(X))(X,\mathcal{PS}(X))). Thus gg is a contracting element with respect to (X,𝒫𝒮(X))(X,\mathcal{PS}(X)).

Conversely, if gGg\in G is contracting with respect to (X,𝒫𝒮(X))(X,\mathcal{PS}(X)), then gg is Morse. By the assertion (1), gg is not conjugate into any vertex group. Thus it acts hyperbolically on the Bass-Serre tree TT. ∎

We now ready to prove Theorem 1.6.

Proof of Theorem 1.6.

We are going to prove the following implications: (1)(2)(\ref{thm1:item1})\Rightarrow(\ref{thm1:item2}), (2)(3)(\ref{thm1:item2})\Rightarrow(\ref{thm1:item3}), (3)(1)(\ref{thm1:item3})\Rightarrow(\ref{thm1:item1}), and (3)(4)(\ref{thm1:item3})\Longleftrightarrow(\ref{thm1:item4}).

(1)(2)(\ref{thm1:item1})\Rightarrow(\ref{thm1:item2}): The implication just follows from Theorem 1.2 in [Tra19].

(2)(3)(\ref{thm1:item2})\Rightarrow(\ref{thm1:item3}): By Lemma 6.9, KK has a finite index subgroup KK^{\prime} such that KK^{\prime} is a free group of finite rank. Let xx be an infinite order element in KK^{\prime}. By way of contradiction, suppose that xx is not Morse in GG. By Corollary 6.16, xx is conjugate into a vertex group of GG. In other words, xgGvg1x\in gG_{v}g^{-1} for some vertex group GvG_{v} and for some gGg\in G. Hence xKgGvg1KgGvg1x\in K^{\prime}\cap gG_{v}g^{-1}\subset K\cap gG_{v}g^{-1} that is finite by Proposition 6.4. So, xx has finite order that contradicts to our assumption that xx is an infinite order element. Since any infinite order element in KK has a power that belongs to KK^{\prime}, the implication (2)(3)(\ref{thm1:item2})\Rightarrow(\ref{thm1:item3}) is verified.

(3)(1)(\ref{thm1:item3})\Rightarrow(\ref{thm1:item1}): Let KK^{\prime} be a finite index subgroup of KK such that KK^{\prime} is free. It follows from Proposition 6.15 that KK^{\prime} is strongly quasi-convex in GG. Since KK^{\prime} is a finite index subgroup of KK, it follows that KK is also strongly quasi-convex in GG.

(3)(4)(\ref{thm1:item3})\Rightarrow(\ref{thm1:item4}): Let KK^{\prime} be a finite index subgroup of KK such that KK^{\prime} is free. Let TTT^{\prime}\subset T be the minimal subtree of TT that contains K(v)K^{\prime}(v). Since the map KTK^{\prime}\to T^{\prime} given by kk(v)k\to k(v) is quasi-isometric and the inclusion map TTT^{\prime}\to T is also a quasi-isometric embedding, it follows that the composition KTTK^{\prime}\to T^{\prime}\to T is a quasi-isometric embedding. Since KK^{\prime} is a finite index subgroup of KK, it follows that the map KTK\to T given by kk(v)k\to k(v) is a quasi-isometric embedding.

(4)(3)(\ref{thm1:item4})\Rightarrow(\ref{thm1:item3}): By way of contradiction, suppose that there exists an infinite order element gKg\in K such that gg is not Morse in GG. It follows from Corollary 6.16 that gg is conjugate into a vertex group, hence gg fixes a vertex vv of TT. By our assumption, there is a vertex ww in TT such that the map KTK\to T given by kk(w)k\to k(w) is a quasi-isometric embedding. It implies that the map KTK\to T given by kk(v)k\to k(v) is also a (λ,λ)(\lambda,\lambda)–quasi-isometric embedding for some λ>0\lambda>0. Choose an integer nn large enough such that |gn|>λ2|g^{n}|>\lambda^{2}. We have 1/λ|gn|λd(g(v),gn(v))=d(v,v)=01/\lambda|g^{n}|-\lambda\leq d(g(v),g^{n}(v))=d(v,v)=0. Hence |gn|<λ2|g^{n}|<\lambda^{2}. This contradicts to our choice of nn. ∎

References

  • [Baj07] Jitendra Bajpai. Omnipotence of surface groups. Master’s thesis, McGill University, 2007.
  • [BBF15] Mladen Bestvina, Ken Bromberg, and Koji Fujiwara. Constructing group actions on quasi-trees and applications to mapping class groups. Publ. Math. Inst. Hautes Études Sci., 122:1–64, 2015.
  • [BBF19] Mladen Bestvina, Kenneth Bromberg, and Koji Fujiwara. Proper actions on finite products of quasi-trees. arXiv e-prints, page arXiv:1905.10813, May 2019.
  • [Bes] Mladen Bestvina. Questions in geometric group theory. M. Bestvina’s home page, 2004.
  • [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.
  • [BN08] Jason A. Behrstock and Walter D. Neumann. Quasi-isometric classification of graph manifold groups. Duke Math. J., 141(2):217–240, 2008.
  • [Bow08] Brian H. Bowditch. Tight geodesics in the curve complex. Invent. Math., 171(2):281–300, 2008.
  • [Bow12] B. H. Bowditch. Relatively hyperbolic groups. Internat. J. Algebra Comput., 22(3):1250016, 66, 2012.
  • [CK00] Christopher B. Croke and Bruce Kleiner. Spaces with nonpositive curvature and their ideal boundaries. Topology, 39(3):549–556, 2000.
  • [CK02] C. B. Croke and B. Kleiner. The geodesic flow of a nonpositively curved graph manifold. Geom. Funct. Anal., 12(3):479–545, 2002.
  • [DGO17] F. Dahmani, V. Guirardel, and D. Osin. Hyperbolically embedded subgroups and rotating families in groups acting on hyperbolic spaces. Mem. Amer. Math. Soc., 245(1156):v+152, 2017.
  • [DJ99] A. Dranishnikov and T. Januszkiewicz. Every Coxeter group acts amenably on a compact space. In Proceedings of the 1999 Topology and Dynamics Conference (Salt Lake City, UT), volume 24, pages 135–141, 1999.
  • [DK18] Cornelia Druţu and Michael Kapovich. Geometric group theory, volume 63 of American Mathematical Society Colloquium Publications. American Mathematical Society, Providence, RI, 2018. With an appendix by Bogdan Nica.
  • [DS05] Cornelia Druţu and Mark Sapir. Tree-graded spaces and asymptotic cones of groups. Topology, 44(5):959–1058, 2005. With an appendix by Denis Osin and Mark Sapir.
  • [Far98] Benson Farb. Relatively hyperbolic groups. Geometric and Functional Analysis, 8:810–840, 1998.
  • [GM14] Craig R. Guilbault and Christopher P. Mooney. Boundaries of Croke–Kleiner-admissible groups and equivariant cell-like equivalence. J. Topol., 7(3):849–868, 2014.
  • [GMRS98] Rita Gitik, Mahan Mitra, Eliyahu Rips, and Michah Sageev. Widths of subgroups. Trans. Amer. Math. Soc., 350(1):321–329, 1998.
  • [GP16] Victor Gerasimov and Leonid Potyagailo. Quasiconvexity in relatively hyperbolic groups. J. Reine Angew. Math., 710:95–135, 2016.
  • [Gre16] Sebastian Grensing. Virtual boundaries of Hadamard spaces with admissible actions of higher rank. Math. Z., 284(1-2):1–22, 2016.
  • [HP15] Mark F. Hagen and Piotr Przytycki. Cocompactly cubulated graph manifolds. Israel J. Math., 207(1):377–394, 2015.
  • [HS13] David Hume and Alessandro Sisto. Embedding universal covers of graph manifolds in products of trees. Proc. Amer. Math. Soc., 141(10):3337–3340, 2013.
  • [HW08] Frédéric Haglund and Daniel T. Wise. Special cube complexes. Geom. Funct. Anal., 17(5):1551–1620, 2008.
  • [KL98] M. Kapovich and B. Leeb. 33-manifold groups and nonpositive curvature. Geom. Funct. Anal., 8(5):841–852, 1998.
  • [Liu13] Yi Liu. Virtual cubulation of nonpositively curved graph manifolds. J. Topol., 6(4):793–822, 2013.
  • [NTY] Hoang Thanh Nguyen, Hung Cong Tran, and Wenyuan Yang. Quasiconvexity in 3-manifold groups. arXiv:1911.07807. To appear in Mathematische Annalen.
  • [Osi16] D. Osin. Acylindrically hyperbolic groups. Trans. Amer. Math. Soc., 368(2):851–888, 2016.
  • [PW14] Piotr Przytycki and Daniel T. Wise. Graph manifolds with boundary are virtually special. J. Topol., 7(2):419–435, 2014.
  • [Sis13] Alessandro Sisto. Projections and relative hyperbolicity. Enseign. Math. (2), 59(1-2):165–181, 2013.
  • [Sis18] Alessandro Sisto. Contracting elements and random walks. J. Reine Angew. Math., 742:79–114, 2018.
  • [Tra19] Hung Cong Tran. On strongly quasiconvex subgroups. Geom. Topol., 23(3):1173–1235, 2019.
  • [Wis00] Daniel T. Wise. Subgroup separability of graphs of free groups with cyclic edge groups. Q. J. Math., 51(1):107–129, 2000.
  • [Wis20] Daniel T. Wise. The Structure of Groups with a Quasiconvex Hierarchy, volume AMS-209. Annals of Mathematics Studies, 2020.
  • [Yan19] Wen-yuan Yang. Statistically convex-cocompact actions of groups with contracting elements. Int. Math. Res. Not. IMRN, (23):7259–7323, 2019.