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

Property (QT) for 3-manifold groups

Suzhen Han Beijing International Center for Mathematical Research
Peking University
Beijing 100871, China P.R.
[email protected]
Hoang Thanh Nguyen Department of Mathematics
The University of Danang - University of Science and Education
459 Ton Duc Thang, Da Nang, Vietnam & International Centre for Research and Postgraduate Training in Mathematics (ICRTM), Institute of Mathematics, VAST, 18 Hoang Quoc Viet Road, Cau Giay District, Hanoi, Vietnam
[email protected]
 and  Wenyuan Yang Beijing International Center for Mathematical Research
Peking University
Beijing 100871, China P.R.
[email protected]
Abstract.

According to Bestvina-Bromberg-Fujiwara, a finitely generated group is said to have property (QT) if it acts isometrically on a finite product of quasi-trees so that orbital maps are quasi-isometric embeddings. We prove that the fundamental group π1(M)\pi_{1}(M) of a compact, connected, orientable 3-manifold MM has property (QT) if and only if no summand in the sphere-disc decomposition of MM supports either Sol or Nil geometry. In particular, all compact, orientable, irreducible 3-manifold groups with nontrivial torus decomposition and not supporting Sol geometry have property (QT). In the course of our study, we establish property (QT) for the class of Croke-Kleiner admissible groups and of relatively hyperbolic groups under natural assumptions has property (QT).

1. Introduction

1.1. Background and Motivation

The study of group actions on quasi-trees has recently received a great deal of interest. A quasi-tree means here a possibly locally infinite connected graph that is quasi-isometric to a simplicial tree. Groups acting on (simplicial) trees have been well-understood thanks to the Bass-Serre theory. On the one hand, quasi-trees have the obvious advantage of being more flexible; hence, many groups can act unboundedly on quasi-trees but act on any trees with global fixed points. Many hyperbolic groups with Kazhdan’s property (T), mapping class groups among many other examples belong to this category (see [Man05, Man06] for other examples). In effect, these are sample applications of a powerful axiomatic construction of quasi-trees proposed in the work of Bestvina, Bromberg and Fujiwara [BBF15]. This construction will be fundamental in this paper.

We say that a finitely generated group G has property (QT) if it acts isometrically on a finite product X=T1×T2××TnX=T_{1}\times T_{2}\times\cdots\times T_{n} of quasi-trees with L2L^{2}-metric so that for any basepoint oXo\in X, the induced orbit map

gGgoXg\in G\longmapsto go\in X

is a quasi-isometric embedding of GG equipped with some (or any) word metric dGd_{G} to XX. Informally speaking, property (QT) gives an undistorted picture of the group under consideration in a reasonably good space. Here, the direct product structure usually comes from the independence of several negatively curved layers endowed on the group. Such a hierarchy structure has emerged from the study of mapping class groups since Masur-Minsky [MM00]. In addition, property (QT) is a commensurability invariant in [BBF19, But20] and could be thought of as a stronger property than the finiteness of asymptotic dimension.

Extending their earlier results of [BBF15], Bestvina, Bromberg and Fujiwara [BBF19] recently showed that residually finite hyperbolic groups and mapping class groups have property (QT). It is known that Coxeter groups have property (QT) (see [DJ99]), and thus every right-angled Artin group has property (QT) (see [BBF19, Induction 2.2]).

In 3-manifold theory, the study of the fundamental groups of 3-manifolds is one of the most central topics. Determining property (QT) of finitely generated 3-manifold groups is the main task of the present paper.

1.2. Property (QT) of 3-manifold groups

Let MM be a 3-manifold with finitely generated fundamental group. Since property (QT) is a commensurability invariant, we can assume that MM is compact and orientable by considering the Scott core of MM and a double cover of MM (if MM is non-orientable).

In recent years, the theory of special cube complexes [HW08] has led to a deep understanding of 3-manifold groups [Wis20] [Ago13]. By definition, the fundamental group of a compact special cube complex is undistorted in a right-angled Artin group, and then has property (QT) by [DJ99]. However, 3–manifolds without non-positively curved Riemannian metrics cannot be cubulated by [PW18] and certain cubulated 33–manifold groups are not virtually cocompact special (see [HP15], [Tid18]). Thus it was left open to determine the property (QT) for all 3-manifold groups.

By the sphere-disc decomposition, a compact oriented 3-manifold MM is a connected sum of prime summands MiM_{i} (1in)(1\leq i\leq n) with incompressible boundary. It is an easy observation that if a group has property (QT) then every non-trivial element is undistorted (see Lemma 2.5), and hence if MiM_{i} supports SolSol or NilNil from the eight Thurston geometries, then π1(Mi)\pi_{1}(M_{i}) fails to have property (QT). Our first main result is the following characterization of property (QT) for all 3-manifolds.

Theorem 1.1.

Let MM be a connected, compact, orientable 3-manifold. Then π1(M)\pi_{1}(M) has property (QT) if and only if no summand in its sphere-disk decomposition supports either SolSol or NilNil geometry.

By standard arguments, we are reduced to the case where MM is a compact, connected, orientable, irreducible 3-manifold with empty or tori boundary. Theorem 1.1 actually follows from the following theorem.

Theorem 1.2.

Let MM be a compact orientable irreducible 3-manifold with empty or tori boundary, with nontrivial torus decomposition and that does not support the Sol geometry. Then π1(M)\pi_{1}(M) has property (QT).

A 3-manifold MM as in Theorem 1.2 is called a graph manifold if all the pieces in its torus decomposition are Seifert fibered spaces; otherwise MM is called a mixed manifold. It is well-known that the fundamental group of a mixed 3-manifold is hyperbolic relative to a collection of abelian groups and graph manifolds groups. To prove Theorem 1.2, we actually determine the property (QT) of Croke-Kleiner admissible groups, and of relatively hyperbolic groups that will be discussed in detail in the following subsections. These results include but are much more general than the fundamental groups of graph manifolds and mixed manifolds.

1.3. Property (QT) of Croke-Kleiner admissible groups.

We first address property (QT) of graph manifolds. Our approach is based on a study of a particular class of graph of groups introduced by Croke and Kleiner [CK02] which they called admissible groups and generalized the fundamental groups of graph manifolds. We say that an admissible group GG is a Croke-Kleiner admissible group or a CKA group if it acts properly discontinuous, cocompactly and by isometries on a complete proper CAT(0) space XX. Such action GXG\curvearrowright X is called a CKA action and the space XX is called a CKA space. The CKA action is modeled on the JSJ structure of graph manifolds where the Seifert fibration is replaced by the following central extension of a general hyperbolic group HvH_{v}:

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

where Z(Gv)=.Z(G_{v})=\mathbb{Z}. It is worth pointing out that CKA groups encompass a much more general class of groups and can be used to produce interesting groups by a “flip” trick from any finite number of hyperbolic groups (see Example 2.14).

The notion of an omnipotent group was introduced by Wise in [Wis00] and has found many applications in subgroup separability. We refer the reader to Definition 4.6 for its definition and note here that free groups [Wis00], surface groups [Baj07], and the more general class of virtually special hyperbolic groups [Wis20] are omnipotent. In [NY], Nguyen-Yang proved property (QT) for a special class of CKA actions under flip conditions (see Definition 2.18). One of the main contributions of this paper is to remove this assumption and prove the following result in full generality.

Theorem 1.3.

Let GXG\curvearrowright X be a CKA action where for every vertex group the central extension (1) has omnipotent hyperbolic quotient group. Then GG has property (QT).

Remark 1.4.

It is a long-standing problem whether every hyperbolic group is residually finite. Wise noted that if every hyperbolic group is residually finite, then any hyperbolic group is omnipotent (see Remark 3.4 in [Wis00]).

Let us comment on the relation of this work with the previous [NY]. As in [NY], the ultimate goal is to utilize Bestvina-Bromberg-Fujiwara’s projection complex machinery to obtain actions on quasi-trees. The common starting point is the class of special paths developed in [NY] that record the distances of XX. However, the flip assumption (see Definition 2.18) on CKA actions was crucially used there: the fiber lines coincide with boundary lines in adjacent vertex pieces when crossing the boundary plane, roughly speaking. Hence, a straightforward gluing construction works in that case but fails in our general setting. In this paper, we use a completely different projection system to achieve the same purpose with a more delicate analysis.

It is worth mentioning the following fact frequently invoked by many authors: any graph manifolds are quasi-isometric to some flip ones (see [KL98]). This simplification, however, is useless to address property (QT), as such a quasi-isometry does not respect the group actions. Conversely, our direct treatment of any graph manifolds (closed or with nonempty boundary) is new, and we believe it will potentially allow for further applications.

We now explain how we apply Theorem 1.3 to graph manifolds. If MM is a graph manifold with nonempty boundary then it always admit a Riemannian metric of nonpositive curvature (see [Lee95]). In particular, π(M)M~\pi(M)\curvearrowright\tilde{M} is a CKA action, and thus property (QT) of π1(M)\pi_{1}(M) follows immediately from Theorem 1.3. However, closed graph manifolds may not support any Riemannian metric of nonpositive curvature (see [Lee95]), so property (QT) in this case does not follow immediately from Theorem 1.3. We have to make certain modifications on some steps to run the proof of Theorem 1.3 for closed graph manifolds (see Section 8.2.1 for details).

1.4. Property (QT) of relatively hyperbolic groups.

When MM is a mixed 3–manifold, then π1(M)\pi_{1}(M) is hyperbolic relative to the finite collection 𝒫\mathcal{P} of fundamental groups of maximal graph manifold components, isolated Seifert components, and isolated JSJ tori (see [BW13], [Dah03]). Therefore, we need to study property (QT) for relatively hyperbolic groups.

Our main result in this direction is a characterization of property (QT) for residually finite relatively hyperbolic groups, which generalizes the corresponding results of [BBF19] on Gromov-hyperbolic groups.

Theorem 1.5.

Suppose that a finitely generated group HH is hyperbolic relative to a finite set of subgroups \mathbb{P}. Assume that each PP\in\mathbb{P} acts by isometry on finitely many quasi-trees TiT_{i} (1inP)(1\leq i\leq n_{P}) so that the induced diagonal action on i=1nPTi\prod_{i=1}^{n_{P}}T_{i} has property (QT). If HH is residually finite, then HH has property (QT).

Remark 1.6.

Since maximal parabolic subgroups are undistorted, each P𝒫P\in\mathcal{P} obviously has property (QT) if GG has property (QT). A non-equivariant version of this result was proven by Mackay-Sisto [MS13].

Remark 1.7.

It is well-known that mixed 3-manifold groups G=π1(M)G=\pi_{1}(M) are hyperbolic relative to a collection \mathbb{P} of abelian groups and graph manifold groups P=π1(Mi)P=\pi_{1}(M_{i}). However, it is still insufficient to derive directly via Theorem 1.5 the property (QT) of GG from that of graph manifold groups PP asserted in Theorem 1.3, since PP may not preserve factors in the finite product of quasi-trees. Of course, passing to an appropriate finite index subgroup P<PP^{\prime}<P preserves the factors, but it is not clear at all whether PP^{\prime} are peripheral subgroups of a finite index subgroup GG^{\prime} of GG. In order to find such a GG^{\prime}, a stronger assumption must be satisfied so that every finite index subgroup of each PP is separable in GG. This requires the notion of a full profinite topology induced on subgroups (see the precise definition before Theorem 3.5 and a relevant discussion in [Rei18]). See Theorem 3.5 for the precise statement. In the setting of a mixed 3-manifold, Lemma 8.5 verifies that each peripheral subgroup PP\in\mathbb{P} of π1(M)\pi_{1}(M) satisfies this assumption. Therefore, all mixed 3-manifolds are proven to have property (QT).

We now explain a few algebraic and geometric consequences for groups with property (QT).

Similar to trees, any isometry on quasi-trees must be either elliptic or loxodromic ([Man05]). Hence, if a finitely generated group acts properly (in a metric sense) on finite products of quasi-trees, then every non-trivial element is undistorted (Lemma 2.5). Moreover, property (QT) allows to characterize virtually abelian groups among sub-exponential growth groups and solvable groups.

Theorem 1.8.

Let GG be a finitely generated group. Then the following statements hold.

  1. (1)

    Assume that GG has sub-exponential growth. Then GG has property (QT) if and only if GG is virtually abelian.

  2. (2)

    Suppose that GG is solvable with finite virtual cohomological dimension. Then GG has property (QT) if and only if it is virtually abelian.

By Theorem 1.5, this yields as a consequence that non-uniform lattices in SU(n,1)SU(n,1) and Sp(n,1)Sp(n,1) for n2n\geq 2 fail to act properly on finite products of quasi-trees.

Corollary 1.9.

A non-uniform lattice in SU(n,1)SU(n,1) for n2n\geq 2 or Sp(n,1)Sp(n,1) for n1n\geq 1 does not have property (QT), while any lattice of SO(n,1)SO(n,1) has property (QT) for n2n\geq 2.

Overview

The paper is structured as follows. In Section 2, we recall the preliminary materials about Croke-Kleiner admissible groups, axiomatic constructions of quasi-trees, and collect a few preliminary observations to disprove property (QT) for Sol and Nil geometries and to prove Theorem 1.8. Section 3 contains a proof of Theorem 1.5 and its variant Theorem 3.5. The next four sections aim to prove Theorem 1.3: Section 4 first recalls a cone-off construction of CKA actions from [NY] and then outlines the steps executed in Sections 5, 6, and 7 to prove property (QT) for CKA actions. Sections 5 and 6 explain in detail the construction of projection systems of fiber lines and then prove the corresponding distance formula. We finish the proof of Theorem 1.3 in Section 7. In Section 8, we present the applications of the previous results for 3-manifold groups and prove Theorem 1.2 and Theorem 1.1.

Acknowledgments

H.T.Nguyen is partially supported by Project ICRTM04_2021.07 of the International Centre for Research and Postgraduate Training in Mathematics, Vietnam. W. Y. is supported by National Key R & D Program of China (SQ2020YFA070059) and National Natural Science Foundation of China (No. 12131009). We are also grateful to the anonymous referee for many very helpful comments.

2. Preliminary

This section reviews concepts property (QT), Croke-Kleiner admissible actions, and the construction of quasi-trees. Several observations are made to determine property (QT) of 3-manifolds with Sol and Nil geometry. This includes the fact that every elements are undistorted in groups with property (QT) and some attempts to characterize by property (QT) the class of virtually abelian groups in solvable/sub-exponential growth groups.

In the sequel, we use the notion aKba\preceq_{K}b if the exists C=C(K)>0C=C(K)>0 such that aCb+Ca\leq Cb+C, and 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. If the constant CC is universal from context, the sub-index K\preceq_{K} shall be omitted.

2.1. Property (QT)

Definition 2.1.

We say that a finitely generated group G has property (QT) if it acts isometrically on finite products X=T1×T2××TnX=T_{1}\times T_{2}\times\cdots\times T_{n} of quasi-trees with L2L^{2}-metric so that for any basepoint oXo\in X, the induced orbit map

gGgoXg\in G\mapsto go\in X

is a quasi-isometric embedding of GG equipped with some (or any) word metric dGd_{G} to XX with the product metric dd.

Remark 2.2.

A group with property (QT) acts properly on finite products of quasi-trees in a metric sense: d(o,go)d(o,go)\to\infty as dG(1,g)d_{G}(1,g)\to\infty. We would emphasize that all consequences of the property (QT) in this paper use merely the existence of a metric proper action.

By definition, a quasi-tree is assumed to be a graph quasi-isometric to a simplicial tree. This does not loss generality as any geodesic metric space (with an isometric action) is quasi-isometric to a graph (with an equivariant isometric action) by taking the 1-skeleton of its Rips complex: the vertex set consists of all points and two points with distance less than 1 are connected by an edge.

The first part of the following lemma allows one to pass to finite index subgroups in the study Property (QT) of groups, as explained in Section 2.2 of [BBF19]. The second part of Lemma 2.3 is an immediate consequence of the definition of property (QT).

Lemma 2.3.
  1. (1)

    Let HGH\leq G be a finite index subgroup of GG. Then GG has property (QT) if and only if HH has property (QT).

  2. (2)

    Let HGH\leq G be an undistorted subgroup of GG. Suppose that GG has property (QT) then HH has property (QT).

Below is a corollary of the de Rham decomposition theorem (see [FL08, Theorem 1.1]) that will be used for the next discussions.

Corollary 2.4.

A finite product X=T1×T2××TnX=T_{1}\times T_{2}\times\cdots\times T_{n} of quasi-trees must have de Rham decomposition

X=k×Tk+1××TnX=\mathbb{R}^{k}\times T_{k+1}\times\cdots\times T_{n}

if the first kk quasi-trees (k0k\geq 0) are all real lines among {Ti:1in}\{T_{i}:1\leq i\leq n\}.

A finite product i=1nTi\prod_{i=1}^{n}T_{i} of quasi-trees has no \mathbb{R}-factor if no TiT_{i} is isometric to \mathbb{R} or a point. In this case, the Euclidean factor k\mathbb{R}^{k} will disappear. In what follows, we give some general results about groups with property (QT).

Lemma 2.5.

Assume that GG has property (QT). Then the subgroup generated by an element gGg\in G is undistorted in GG.

Proof.

Let X=k×Tk+1××TnX=\mathbb{R}^{k}\times T_{k+1}\times\cdots\times T_{n} be the de Rham decomposition of a finite product of quasi-trees. By [FL08, Corollary 1.3], up to passage to finite index subgroups, GG acts by isometry on each factor k,\mathbb{R}^{k}, and TiT_{i} for k+1ink+1\leq i\leq n. Let gGg\in G be an infinite order element. If the image of gg is an isometry on the Euclidean space k\mathbb{R}^{k}, then it either fixes a point or preserves an axe. If the image of gg is an isometry on a quasi-tree TiT_{i} then by [Man06, Corollary 3.2], it has either a bounded orbit or a quasi-isometrically embedded orbit.

Fix a basepoint o=(ok,ok+1,,on)Xo=(o_{k},o_{k+1},\cdots,o_{n})\in X. If the action of GG on XX is proper, by the first paragraph, there must exist a unbounded action of g\langle g\rangle on some factor Y=kY=\mathbb{R}^{k} or Y=TiY=T_{i}, so we have mλ|okgmok|Y+cm\leq\lambda|o_{k}-g^{m}o_{k}|_{Y}+c for some λ,c>0\lambda,c>0. Since any isometric orbital map is Lipschitz, we have |ogmo|XC|1gm|G|o-g^{m}o|_{X}\leq C|1-g^{m}|_{G} for some C>0C>0. Noting that |ogmo|Y|ogmo|X|o-g^{m}o|_{Y}\leq|o-g^{m}o|_{X}, we have mgmm\mapsto g^{m} is a quasi-isometric embedding of g\langle g\rangle into GG. ∎

Note that the Sol group embeds quasi-isometrically into a product of two hyperbolic planes (for example, see [dC08, Section 9]). However, the Sol lattice contains exponentially distorted elements by [NS20, Lemma 5.2].

Corollary 2.6.

The fundamental group of a 3-manifold with Sol geometry does not have property (QT).

Corollary 2.7.

The Baumslag-Solitar group BS(1,n)BS(1,n) for n>1n>1 does not have property (QT).

2.2. Sub-exponential growth and solvable groups with property (QT)

The fundamental group of a 3-manifold MM with Nil geometry also fails to have property (QT) since it contains quadratically distorted elements (for example, see Proposition 1.2 in [NS20]). Generalizing results about property (QT) of 3-manifolds with Sol or Nil geometry, in the rest of this subsection, we provide a characterization of sub-exponential growth groups/ solvable groups with property (QT) and give the proof of Theorem 1.8.

In next results, we apply the general results in [CCMT15] about the isometric actions on hyperbolic spaces to quasi-trees. By Gromov, unbounded isometric group actions can be classified into the following four types:

  1. (1)

    horocyclic if it has no loxodromic element;

  2. (2)

    lineal if it has a loxodromic element and any two loxodromic elements have the same fixed points in the Gromov boundary;

  3. (3)

    focal if it has a loxodromic element, is not lineal and any two loxodromic elements have one common fixed point;

  4. (4)

    general type if it has two loxodromic elements with no common fixed point.

Proposition 2.8.

Assume that GG has property (QT). Then there exist a finite index subgroup G˙\dot{G} of GG which acts on a Euclidean space k\mathbb{R}^{k} with k0k\geq 0 and finitely many quasi-trees TiT_{i} for 1in1\leq i\leq n with lineal or focal or general type action so that the orbital map of G˙\dot{G} into k×i=1nTi\mathbb{R}^{k}\times\prod_{i=1}^{n}T_{i} is a quasi-isometric embedding.

Moreover, the action on each TiT_{i} can be chosen to be cobounded.

Proof.

By Corollary 2.4, the finite product of quasi-trees given by property (QT) has the above form of de Rham decomposition. By [FL08, Corollary 1.3],

1Isom(k)×i=k+1nIsom(Yi)Isom(X)F11\to\operatorname{Isom}(\mathbb{R}^{k})\times\prod_{i=k+1}^{n}\operatorname{Isom}(Y_{i})\to\operatorname{Isom}(X)\to F\to 1

where FF is a subgroup of the permutation group on the indices {k+1,,n}\{k+1,\cdots,n\}. Thus, there exists a finite index subgroup G˙\dot{G} of GG acting on each de Rham factor such that G˙Isom(k)×i=1nIsom(Yi)\dot{G}\subset\operatorname{Isom}(\mathbb{R}^{k})\times\prod_{i=1}^{n}\operatorname{Isom}(Y_{i}) for k0k\geq 0 and ik+1i\geq k+1.

First of all, we can assume that the actions of G˙\dot{G} on k\mathbb{R}^{k} and each TiT_{i} is unbounded. Otherwise, we can remove k,\mathbb{R}^{k}, and TiT_{i} with bounded actions from the product without affecting property (QT).

We now consider the action on TiT_{i} for k+1ink+1\leq i\leq n. We then need verify that the action of G˙\dot{G} on TiT_{i} cannot be horocyclic. By way of contradiction, assume that the action of G˙\dot{G} on given TiT_{i} is horocyclic.

Note that the proof of [CCMT15, Prop 3.1] shows that the intersection of any orbit of G˙\dot{G} on TiT_{i} with any quasi-geodesic is bounded. By [Man06, Corollary 3.2], any isometry on a quasi-tree TiT_{i} has either bounded orbits or a quasi-geodesic orbit. Thus, we conclude that any orbit of h\langle h\rangle for every hG˙h\in\dot{G} on TiT_{i} is bounded. We are then going to prove that the action of G˙\dot{G} on TiT_{i} has bounded orbits. This is a well-known fact and we present the proof for completeness.

By δ\delta-hyperbolicity of TiT_{i}, each hG˙h\in\dot{G} (with bounded orbits) has a quasi-center chTic_{h}\in T_{i}: there exists a constant D>0D>0 depending only on δ\delta such that |chhich|TiD|c_{h}-h^{i}c_{h}|_{T_{i}}\leq D for ii\in\mathbb{Z}. Moreover, for any xchx\in c_{h} and any yTiy\in T_{i}, the Gromov product y,hyx\langle y,hy\rangle_{x} is bounded by a constant CC depending only on DD. As a consequence, the union ZZ of quasi-centers {ch:hG˙}\{c_{h}:h\in\dot{G}\} has finite diameter. Indeed, note that y,h1yx\langle y,h_{1}y\rangle_{x} and x,h21xy\langle x,h_{2}^{-1}x\rangle_{y} are bounded by CC for any xch1,ych2x\in c_{h_{1}},y\in c_{h_{2}}. If for two elements h1,h2h_{1},h_{2}, the distance |ch1ch2|Ti|c_{h_{1}}-c_{h_{2}}|_{T_{i}} is sufficiently large relative to CC, the path connecting dots (h1h2)nx(h_{1}h_{2})^{n}x for nn\in\mathbb{Z} would be a sufficiently long local quasi-geodesic, so it is a global quasi-geodesic. By the previous paragraph, we obtain a contradiction so the G˙\dot{G}-invariant set ZZ is bounded. Since the action on TiT_{i} is assumed to be unbounded, we thus proved that the action on TiT_{i} cannot be horocyclic.

At last, it remains to prove the “moreover” statement. By Manning’s bottleneck criterion [Man06], any geodesic is contained in a uniform neighborhood of every path with the same endpoints. Thus, any connected subgraph of a quasi-tree is uniform quasiconvex and so is a uniform quasi-tree. Since GG is a finitely generated group, by taking the image of the Cayley graph, we can thus construct a connected subgraph on each quasi-tree TiT_{i} so that the action on the subgraph (quasi-tree) is co-bounded. Thus, the proposition is proved. ∎

We are able to characterize sub-exponential groups with property (QT) as follows.

Proposition 2.9.

Let GG be a finitely generated group with sub-exponential growth. Then GG has property (QT) if and only if GG is virtually abelian.

Proof.

We first observe that k\mathbb{R}^{k} in Proposition 2.8 can be replaced by a finite product of real lines. Indeed, consider the action of G˙\dot{G} on Euclidean space k\mathbb{R}^{k}. By assumption, G˙\dot{G} is of sub-exponential growth. It is well-known that the growth of any finitely generated group dominates that of quotients, so the image ΓIsom(k)\Gamma\subset\operatorname{Isom}(\mathbb{R}^{k}) of G˙\dot{G} acting on k\mathbb{R}^{k} has sub-exponential growth. Since finitely generated linear groups do not have intermediate growth, Γ\Gamma must be virtually nilpotent. It is well-known that virtually nilpotent subgroups in Isom(k)\operatorname{Isom}(\mathbb{R}^{k}) must be virtually abelian. Thus, Γ\Gamma contains a finite index subgroup l\mathbb{Z}^{l} for 1lk1\leq l\leq k. By taking the preimage of l\mathbb{Z}^{l} in G˙\dot{G}, we can assume further that G˙\dot{G} acts on k\mathbb{R}^{k} through l\mathbb{Z}^{l}. It is clear that l\mathbb{Z}^{l} acts on ll real lines 1,2,l\mathbb{R}_{1},\mathbb{R}_{2}\cdots,\mathbb{R}_{l} so that the product action is geometric. We thus replace k\mathbb{R}^{k} by the product 1ili\prod_{1\leq i\leq l}\mathbb{R}_{i} where G˙\dot{G} admits a lineal action on each i\mathbb{R}_{i} by translation.

By Proposition 2.8 the action of G˙\dot{G} on TiT_{i} is either lineal or focal or general type. In the latter two cases, G˙\dot{G} contains a free (semi-)group by [CCMT15, Lemma 3.3], contradicting the sub-exponential growth of G˙\dot{G}. Thus, the action of G˙\dot{G} on each TiT_{i} is lineal. By Proposition 2.8, we can assume that TiT_{i} is a quasi-line.

By [Man06, Lemma 3.7], a quasi-line TT admits a (1,C)(1,C)-quasi-isometry ϕ\phi (with a quasi-inverse ψ\psi) to \mathbb{R} for some C>0C>0. A lineal action of GG on TT is then conjugated to a quasi-action of GG on \mathbb{R} sending gGg\in G to a (1,C)(1,C^{\prime})-quasi-isometry ϕgψ\phi g\psi on \mathbb{R} for some C=C(C)>0C^{\prime}=C^{\prime}(C)>0. By taking an index at most 2 subgroup, we can assume that every element in GG fixes pointwise the two ends of TT. Note that a (1,C)(1,C^{\prime})-quasi-isometry ϕgψ\phi g\psi on \mathbb{R} fixing the two ends of \mathbb{R} is uniformly bounded away from a translation on \mathbb{R}. So, for any xx\in\mathbb{R}, the orbital map gϕgψ(x)g\mapsto\phi g\psi(x) is a quasi-homomorphism GG\to\mathbb{R}. It is well-known that for any amenable group, any quasi-homomorphism must be a homomorphism up to bounded error. We conclude that any [G,G][G,G]-orbit on TT stays in a bounded set.

Therefore, any [G,G][G,G]-orbit on (1ili)×(1inTi)(\prod_{1\leq i\leq l}\mathbb{R}_{i})\times(\prod_{1\leq i\leq n}T_{i}) is bounded, so the proper action on XX implies that [G˙,G˙][\dot{G},\dot{G}] is a finite group. It is well-known that if a group has finite commutator subgroup, then it is virtually abelian ([BH99, Lemma II.7.9]). The lemma is proved. ∎

It would be interesting to ask whether Proposition 2.9 holds within the class of solvable groups. In Proposition 2.11 below, we are able to give a positive answer to the previous question when the solvable group has finite virtual cohomological dimension. To this end, we need the following fact.

Lemma 2.10.

Any unbounded isometric action of a meta-abelian group on a quasi-tree must be lineal.

Recall that a meta-abelian group is a group whose commutator subgroup is abelian.

Proof.

Indeed, the abelian group Γ=[G,G]\Gamma=[G,G] (of possibly infinite rank) cannot contain free semi-groups, so by [CCMT15, Lemma 3.3], the action of Γ\Gamma on a quasi-tree TT must be bounded or lineal.

Assume first that Γ\Gamma has a bounded orbit KK in TT. Since G/ΓG/\Gamma is abelian, we have that gmhnK=hngmKg^{m}h^{n}K=h^{n}g^{m}K for any n,mn,m\in\mathbb{Z} and g,hGg,h\in G, and thus ghnK=hngKgh^{n}K=h^{n}gK has finite Hausdorff distance to hnKh^{n}K for any nn\in\mathbb{Z}. Assume that g,hg,h are loxodromic. Then {hnK,n}\{h^{n}K,n\in\mathbb{Z}\} is quasi-isometric to a line. Hence, we obtain that the fixed points of g,hg,h at the Gromov boundary must be coincide. This means the action of GG on TT is lineal.

In the lineal case, Γ\Gamma preserves some bi-infinite quasi-geodesic γ\gamma up to finite Hausdorff distance. Since Γ\Gamma is a normal subgroup in GG, we see that every loxodromic element in GG also preserves γ\gamma up to a finite Hausdorff distance. Thus, the action of GG on TT is also lineal. ∎

By Lemma 2.5, a group with property (QT) is translation proper in the sense of Conner [Con00]: the translation length of any non-torsion element is positive. If GG is solvable and has finite v.c.d., then Conner shows that GG is virtually meta-abelian.

Proposition 2.11.

Suppose that a solvable group GG has finite virtual cohomological dimension. If GG has property (QT) then it is virtually abelian.

Proof.

Passing to finite index subgroups, assume that GG is meta-abelian so any quotient of GG is meta-abelian. By Lemma 2.10, the action of GG on each TiT_{i} is lineal.

After possibly passing to an index 2 subgroup, a lineal action of any amenable group GG on a quasi-line TT can be quasi-conjugated to be an isometric action on \mathbb{R}. Indeed, by the proof of Lemma 2.9, conjugating the original action by almost isometry gives a quasi-action of GG on \mathbb{R} so that any orbital map induces a quasi-homomorphism of GG to \mathbb{R}. For amenable groups, any quasi-homomorphism differs from a homomorphism by a uniform bounded constant. Thus, up to quasi-conjugacy, the lineal action of GG on TT can be promoted to be an isometric action on \mathbb{R}.

Consequently, we can quasi-conjugate the action of a solvable group GG on a finite product of quasi-trees to a proper action on a Euclidean space. Thus, GG must be virtually abelian. ∎

Proof of Theorem 1.8.

The proof is a combination of Proposition 2.9 and Proposition 2.11. ∎

2.3. CKA groups

Admissible groups firstly introduced in [CK02] are a particular class of graph of groups that includes fundamental groups of 33–dimensional graph manifolds. In this section, we review admissible groups and their properties that will used throughout the paper.

Let 𝒢\mathcal{G} be a connected graph. We often consider oriented edges from ee_{-} to e+e_{+} and denote e=[e,e+]e=[e_{-},e_{+}]. Then e¯=[e+,e]\overline{e}=[e_{+},e_{-}] denotes the oriented edge with reversed orientation. Denote by 𝒢0\mathcal{G}^{0} the set of vertices and by 𝒢1\mathcal{G}^{1} the set of all oriented edges.

Definition 2.12.

A graph of groups 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 gGvg\in{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}} is 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 2.13.

We say that an admissible group GG is a Croke-Kleiner admissible group or CKA group if it acts properly discontinuous, cocompactly and by isometries on a complete proper CAT(0) space XX. Such action GXG\curvearrowright X is called a CKA action and the space XX is called a CKA space.

Example 2.14.
  1. (1)

    Let MM be a nongeometric graph manifold that admits a nonpositively curved 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)

    One may build Croke-Kleiner admissible groups algebraically from any finite number of hyperbolic CAT(0) groups. The following example is for n=2n=2 but the same principle works for any n2n\geq 2. Let H1H_{1} and H2H_{2} be two torsion-free hyperbolic groups that act geometrically on CAT(0)CAT(0) spaces X1X_{1} and X2X_{2} respectively. Then Gi=Hi×tiG_{i}=H_{i}\times\langle t_{i}\rangle (with i=1,2i=1,2) acts geometrically on the CAT(0)CAT(0) space Yi=Xi×Y_{i}=X_{i}\times\mathbb{R}. Any primitive hyperbolic element hih_{i} in HiH_{i} gives rise to a totally geodesic torus TiT_{i} in the quotient space Yi/GiY_{i}/G_{i} with basis ([hi],[ti])([h_{i}],[t_{i}]). We re-scale YiY_{i} so that the translation length of hih_{i} is equal to that of tit_{i} for each ii. Let f:T1T2f\colon T_{1}\to T_{2} be a flip isometry respecting these lengths, that is, an orientation-reversing isometry mapping [h1][h_{1}] to [t2][t_{2}] and [t1][t_{1}] to [h2][h_{2}]. Let MM be the space obtained by gluing Y1Y_{1} to Y2Y_{2} by the isometry ff. There is a metric on MM which makes MM into a locally CAT(0)CAT(0) space (see e.g. [BH99, Proposition II.11.6]). By the Cartan-Hadamard Theorem, the universal cover M~\widetilde{M} with the induced length metric from MM is a CAT(0) space. Let GG be the fundamental group of MM. The action GM~G\curvearrowright\widetilde{M} is geometric, and GG is an example of a Croke-Kleiner admissible group.

Remark 2.15.

All graph 3-manifold groups are admissible, but there are closed graph 3-manifold groups that are not CAT(0) groups (see [KL96]), and thus are not CKA groups. The following is another example. Take two non-virtually split central extensions of hyperbolic groups by \mathbb{Z} (e.g. SL(2,)~\widetilde{SL(2,\mathbb{R})} lattices) and amalgamate them over 2\mathbb{Z}^{2} to get an admissible group. This group cannot act properly on CAT(0) spaces, since central extensions acting on CAT(0) spaces must virtually split as direct products ([BH99, Thm. II.7.1]).

A collection of subgroup {K1,,Kn}\{K_{1},\cdots,K_{n}\} in a group HH is called almost malnormal if (gKig1Kj)=\sharp(gK_{i}g^{-1}\cap K_{j})=\infty implies i=ji=j and gKig\in K_{i}. It is well-known that a hyperbolic group is hyperbolic relative to any almost malnormal collection of quasi-convex subgroups ([Bow12]).

Lemma 2.16.

Let KeK_{e} be the image of an edge group GeG_{e} into GvG_{v} and K¯e\overline{K}_{e} be its projection in HvH_{v} under GvHv=Gv/Z(Gv)G_{v}\to H_{v}=G_{v}/Z(G_{v}). Then :={K¯e:e=v,e𝒢1}\mathbb{P}:=\{\overline{K}_{e}:e_{-}=v,e\in\mathcal{G}^{1}\} is an almost malnormal collection of virtually cyclic subgroups in HvH_{v}.

In particular, HvH_{v} is hyperbolic relative to \mathbb{P}.

Proof.

Since Z(Gv)Ke2Z(G_{v})\subset K_{e}\cong\mathbb{Z}^{2}, we have K¯e=Ke/Z(Gv)\overline{K}_{e}=K_{e}/Z(G_{v}) is virtually cyclic. The almost malnormality follows from non-commensurability of KeK_{e} in GvG_{v}. Indeed, assume that K¯ehK¯eh1\overline{K}_{e}\cap h\overline{K}_{e^{\prime}}h^{-1} contains an infinite order element by the hyperbolicity of HvH_{v}. If gGvg\in G_{v} is sent to hh, then KegKeg1K_{e}\cap gK_{e^{\prime}}g^{-1} is sent to K¯ehK¯eh1\overline{K}_{e}\cap h\overline{K}_{e^{\prime}}h^{-1}. Thus, KegKeg1K_{e}\cap gK_{e^{\prime}}g^{-1} contains an abelian group of rank 2. The non-commensurability of KeK_{e} in GvG_{v} implies that e=ee=e^{\prime} and gKeg\in K_{e}. This shows that \mathbb{P} is almost malnormal. ∎

Let GXG\curvearrowright X be a CKA action where GG is the fundamental group of an admissible graph of groups 𝒢\mathcal{G}, and let GTG\curvearrowright T be the action of GG on the associated Bass-Serre tree TT of 𝒢\mathcal{G} (we refer the reader to Section 2.5 in [CK02] for a brief discussion). Let T0T^{0} and T1T^{1} be the vertex and edge sets of TT. By CAT(0) geometry,

  1. (1)

    for every vertex vT0,v\in T^{0}, the minimal set Yv:=gZ(Gv)Minset(g)Y_{v}:=\cap_{g\in Z(G_{v})}Minset(g) of XX splits as metric product Y¯v×\overline{Y}_{v}\times\mathbb{R} where Z(Gv)Z(G_{v}) acts by translation on the \mathbb{R}–factor and Hv=Gv/Z(Gv)H_{v}=G_{v}/Z(G_{v}) acts geometrically on the Hadamard space Y¯v\overline{Y}_{v}. Since HvH_{v} is a hyperbolic group, it follows that Y¯v\overline{Y}_{v} is a hyperbolic space.

  2. (2)

    for every edge eT1e\in T^{1}, the minimal set Ye:=gGeMinset(g)Y_{e}:=\cap_{g\in G_{e}}Minset(g) of XX splits as Y¯e×2Yv\overline{Y}_{e}\times\mathbb{R}^{2}\subset Y_{v} where Y¯e\overline{Y}_{e} is a compact Hadamard space and Ge=2G_{e}=\mathbb{Z}^{2} acts cocompactly on the Euclidean plane 2\mathbb{R}^{2}.

We note that the assignments vYvv\to Y_{v} and eYee\to Y_{e} are GG–equivariant with respect to the natural GG actions.

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

Lemma 2.17.

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

  1. (1)

    X=vT0ND(Yv)=eT1ND(Ye)X=\cup_{v\in T^{0}}{N}_{D}(Y_{v})=\cup_{e\in T^{1}}{N}_{D}(Y_{e}).

  2. (2)

    If σ,σT0T1\sigma,\sigma^{\prime}\in T^{0}\cup T^{1} and ND(Yσ)ND(Yσ)N_{D}(Y_{\sigma})\cap N_{D}(Y_{\sigma^{\prime}})\neq\varnothing then |σσ|T<D|\sigma-\sigma^{\prime}|_{T}<D.

We shall refer Y~v=ND(Yv)\tilde{Y}_{v}=N_{D}(Y_{v}) and Y~e=ND(Ye)\tilde{Y}_{e}=N_{D}(Y_{e}) to as vertex and edge spaces for XX.

2.3.1. Strips in CKA spaces

(Section 4.2 in [CK02]) We first choose, in a GG–equivariant way, a plane FeYeF_{e}\subset Y_{e} (which we will call boundary plane) for each edge eT1e\in T^{1}. 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} where 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. If f1=Z(Gv1),f2=Z(Gv2)\langle f_{1}\rangle=Z(G_{v_{1}}),\langle f_{2}\rangle=Z(G_{v_{2}}) then f1,f2\langle f_{1},f_{2}\rangle generates a finite index subgroup of GeG_{e}. We remark that the intersection of two strips 𝒮e1e\mathcal{S}_{e_{1}e} and 𝒮e2e\mathcal{S}_{e_{2}e} is a point. Indeed, we have 𝒮e1e𝒮e2e=(𝒮e1eFe)(𝒮e2eFe)\mathcal{S}_{e_{1}e}\cap\mathcal{S}_{e_{2}e}=(\mathcal{S}_{e_{1}e}\cap F_{e})\cap(\mathcal{S}_{e_{2}e}\cap F_{e}). As two lines 𝒮e1eFe\mathcal{S}_{e_{1}e}\cap F_{e} and 𝒮e2eFe\mathcal{S}_{e_{2}e}\cap F_{e} in the plane FeF_{e} are axes of fv1=Z(Gv1)\langle f_{v_{1}}\rangle=Z(G_{v_{1}}), fv1=Z(Gv2)\langle f_{v_{1}}\rangle=Z(G_{v_{2}}) respectively and f1,f2\langle f_{1},f_{2}\rangle generates a finite index subgroup of GeG_{e}, it follows that these two lines are non-parallel, and hence their intersection must be a point.

We note that the intersection of a boundary plane FeF_{e} of YvY_{v} with the hyperbolic space Y¯v\overline{Y}_{v} is a line. The boundary lines 𝕃v\mathbb{L}_{v} of the hyperbolic space Y¯v\overline{Y}_{v} are the following collection of lines: 𝕃v={e:=FeY¯v|e=v}\mathbb{L}_{v}=\{\ell_{e}:=F_{e}\cap\overline{Y}_{v}\,|\,e_{-}=v\}.

Definition 2.18.

If for each edge e:=[v,w]Te:=[v,w]\in T, 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 the sequel, it will be useful to choose.

Definition 2.19.

An indexed map ρ:XT0\rho\colon X\to T^{0} is a GG–equivariant coarsely Lipschitz map such that xXρ(x)x\in X_{\rho(x)} for all xXx\in X.

If GG acts freely on XX, such a map ρ\rho can be constructed as follows. Choose a fundamental set Σ\Sigma so that Σ\Sigma contains exactly one point from each orbit. Define ρ:ΣT0\rho:\Sigma\to T^{0} so that ρ(x)=Xρ(x)\rho(x)=X_{\rho(x)}, and extend equivariantly ρ\rho to the whole space XX. By Lemma 2.17.(2), one can show that ρ\rho is a coarsely Lipschitz map: |ρ(x)ρ(y)|TL|xy|X+L|\rho(x)-\rho(y)|_{T}\leq L|x-y|_{X}+L for some L>0L>0. See [CK02, Section 3.3] for more details.

If GG acts only geometrically on XX, we could replace XX with a GG-orbit GoGo for a basepoint oo with trivial stabilizer. This does not matter much as we are only interested in the coarse geometry hereafter. By modifying XX, we could always assume such a basepoint oo exists. Indeed, attach a Euclidean cone to a point oo so that its nontrivial but finite stabilizer acts freely on its boundary circle. We do the modification equivariantly for all translates in GoGo.

2.3.2. Special paths in CKA spaces

Let GXG\curvearrowright X be a CKA action. We now introduce the class of special paths in XX.

Definition 2.20 (Special paths in XX).

Let ρ:XT0\rho\colon X\to T^{0} be the indexed map given by Definition 2.19. Let xx and yy be two points in XX. If ρ(x)=ρ(y)\rho(x)=\rho(y), 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 pi=𝒮ei1ei𝒮eiei+1p_{i}=\mathcal{S}_{e_{i-1}e_{i}}\cap\mathcal{S}_{e_{i}e_{i+1}} be the intersection point of adjacent strips, where e0:=xe_{0}:=x and en+1:=ye_{n+1}:=y. A 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 2.21.

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, the choice of planes FeF_{e} and the indexed map ρ\rho.

Refer to caption
Figure 1. The dotted and blue path from xx to yy is a special path, and the red path is one L1L^{1}-version of it.
Proposition 2.22.

[NY, Prop. 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.

Assume that v0=ρ(x),v2n=ρ(y)𝒱v_{0}=\rho(x),v_{2n}=\rho(y)\in\mathcal{V} so that d(v0,v2n)=2nd(v_{0},v_{2n})=2n for n0n\geq 0. If γ\gamma is a special path between xx and yy, we then define

(2) |xy|Xhor:=i=02n|pipi+1|Yvihor,|xy|Xver:=i=02n|pipi+1|Yvver\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}:=\sum_{i=0}^{2n}\bigl{|}p_{i}-p_{i+1}\bigr{|}_{Y_{v_{i}}}^{\operatorname{hor}},\quad\bigl{|}x-y\bigr{|}_{X}^{\operatorname{ver}}:=\sum_{i=0}^{2n}\bigl{|}p_{i}-p_{i+1}\bigr{|}_{Y_{v}}^{\operatorname{ver}}

where p0:=xp_{0}:=x and pn+1:=yp_{n+1}:=y. By Proposition 2.22, we have

|xy|X|xy|Xhor+|xy|Xver.|x-y|_{X}\sim\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}+\bigl{|}x-y\bigr{|}_{X}^{\operatorname{ver}}.

By definition, the system of special paths is GG-invariant, so the symmetric functions dh(x,y)d^{h}(x,y) and dv(x,y)d^{v}(x,y) are GG-invariant for any x,yXx,y\in X.

We partition the vertex set T0T^{0} of the Bass-Serre tree into two disjoint classes 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.

Lemma 2.23.

[NY, Lemma 4.6] There exists a subgroup G˙\dot{G} of index at most 2 in GG preserving 𝒱i\mathcal{V}_{i} for i=1,2i=1,2 so that GvG˙G_{v}\subset\dot{G} for any vT0v\in T^{0}.

2.4. Projection axioms

In this subsection, we briefly recall the work of Bestvina-Bromberg-Fujiwara [BBF15] on constructing a quasi-tree of spaces.

Definition 2.24 (Projection axioms).

Let 𝕐\mathbb{Y} be a collection of geodesic spaces equipped with projection maps

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

Denote dY(X,Z)=diam(πY(X)πY(Z))d_{Y}(X,Z)=diam(\pi_{Y}(X)\cup\pi_{Y}(Z)) for XYZ𝕐X\neq Y\neq Z\in\mathbb{Y}. The pair (𝕐,{πY}Y𝕐)(\mathbb{Y},\{\pi_{Y}\}_{Y\in\mathbb{Y}}) satisfies projection axioms for a projection constant ξ0\xi\geq 0 if

  1. (1)

    diam(πY(X))ξdiam(\pi_{Y}(X))\leq\xi when XYX\neq Y.

  2. (2)

    if X,Y,ZX,Y,Z are distinct and dY(X,Z)>ξd_{Y}(X,Z)>\xi then dX(Y,Z)ξd_{X}(Y,Z)\leq\xi.

  3. (3)

    for XZX\neq Z, the set {Y𝕐:dY(X,Z)>ξ}\{Y\in\mathbb{Y}\,:\,d_{Y}(X,Z)>\xi\} is finite.

The following is a useful example to keep in mind throughout the paper. For further details, we refer the reader to the introduction of [BBF15]. In this example, the collection of metric spaces 𝕐\mathbb{Y} consists of subspaces of a singe metric space; however, we emphasize that this need not be the case in general.

Example 2.25.

Let GG be a discrete group of isometries of 2\mathbb{H}^{2}, and γG\gamma\in G a loxodromic element with axis γ\gamma. Let 𝕐\mathbb{Y} be the set of all GG–translates of γ\gamma. Given Y𝕐Y\in\mathbb{Y}, let πY\pi_{Y} denote the closest point projection map in 2\mathbb{H}^{2}. Since all translates of γ\gamma are convex, this is a well-defined 11–Lipschitz map. One may check that (𝕐,πY)(\mathbb{Y},\pi_{Y}) satisfies the projection axioms for some constant ξ\xi.

Remark 2.26.

Let (𝕐,{πY}Y𝕐)(\mathbb{Y},\{\pi_{Y}\}_{Y\in\mathbb{Y}}) satisfy projection axioms. By [BBFS19, Thm 4.1 and Lem 4.13], there exists a variant πY\pi_{Y}^{\prime} of πY\pi_{Y} so that πY\pi_{Y} and πY\pi^{\prime}_{Y} are uniformly close in Hausdorff distance, and (𝕐,{πY}Y𝕐)(\mathbb{Y},\{\pi_{Y}^{\prime}\}_{Y\in\mathbb{Y}}) satisfies strong projection axioms, i.e, axioms are the same as projection axioms execpt for replacing (2) in Definition 2.24 with the following stronger statement: if X,Y,ZX,Y,Z are distinct and dY(X,Z)>ξd_{Y}(X,Z)>\xi then πX(Y)=πX(Z)\pi_{X}(Y)=\pi_{X}(Z) for a projection constant ξ\xi^{\prime} depending only on ξ\xi.

The following results from [BBF15] will be used in this paper.

  • Fix K>0K>0. In [BBF15], a quasi-tree of spaces 𝒞K(𝕐)\mathcal{C}_{K}(\mathbb{Y}) is constructed for given (𝕐,{πY}Y𝕐)(\mathbb{Y},\{\pi_{Y}\}_{Y\in\mathbb{Y}}) satisfying projection axioms with constant ξ\xi.

  • If K>4ξK>4\xi and 𝕐\mathbb{Y} is a collection of uniform quasi-lines, then 𝒞K(𝕐)\mathcal{C}_{K}(\mathbb{Y}) is a unbounded quasi-tree. If 𝕐\mathbb{Y} admits a group action of GG so that πgY=gπY\pi_{gY}=g\pi_{Y} for any gGg\in G and Y𝕐Y\in\mathbb{Y}, then GG acts by isometry on 𝒞K(𝕐)\mathcal{C}_{K}(\mathbb{Y}).

Set [t]K=t[t]_{K}=t if tKt\geq K, otherwise [t]K=0[t]_{K}=0. Let xX,zZ𝕐x\in X,z\in Z\in\mathbb{Y}. If XYZX\neq Y\neq Z define dY(x,z)=dY(X,Z)d_{Y}(x,z)=d_{Y}(X,Z). If Y=X,YZY=X,Y\neq Z, define dY(x,z)=diam(πY(x,Z))d_{Y}(x,z)=diam(\pi_{Y}(x,Z)). If X=Y=ZX=Y=Z, let dY(x,z)d_{Y}(x,z) be the distance in YY. The following distance formula from [BBF15] is crucial in what follows.

Proposition 2.27.

[BBF19, Proposition 2.4] Let (𝕐,{πY}Y𝕐)(\mathbb{Y},\{\pi_{Y}\}_{Y\in\mathbb{Y}}) satisfy the strong projection axioms with constant ξ\xi. Then for any x,y𝒞K(𝕐)x,y\in\mathcal{C}_{K}(\mathbb{Y}),

14Y𝕐[dY(x,y)]K|xy|𝒞K(𝕐)2Y𝕐[dY(x,y)]K+3K\frac{1}{4}\sum_{Y\in\mathbb{Y}}[d_{Y}(x,y)]_{K}\leq\bigl{|}x-y\bigr{|}_{\mathcal{C}_{K}(\mathbb{Y})}\leq 2\sum_{Y\in\mathbb{Y}}[d_{Y}(x,y)]_{K}+3K

for all K4ξK\geq 4\xi.

Definition 2.28 (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 |ab|XR|a-b|_{X}\geq R then we have

#{gG||gaa|Xrand|gbb|Xr}N.\#\bigl{\{}g\in G\,|\,|ga-a|_{X}\leq r\,\,\textup{and}\,\,|gb-b|_{X}\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 finite Hausdorff distance depending only on λ,c\lambda,c.

A group is called elementary if it is neither finite nor virtually cyclic.

Proposition 2.29.

[BBF19] Assume that a non-elementary 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 given (λ,c)(\lambda,c)-quasi-axis of gg for given λ,c>0\lambda,c>0. Then there exists a constant θ=θ(λ,c)>0\theta=\theta(\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\}

is a finite union of double E(g)E(g)-cosets.

In particular, there are only finitely many 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.

Lemma 2.30.

[Yan19, Lemma 2.14] 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=\bigcup_{i\in\mathbb{Z}}(hf)^{i}\left([o,ho][ho,hfo]\right)

is a (λ,c)(\lambda,c)–quasi-geodesic.

Convention 2.31.

When speaking of quasi-lines in hyperbolic spaces with actions satisfying Lemma 2.30, we always mean (λ,c)(\lambda,c)–quasi-geodesics where λ,c>0\lambda,c>0 depend on FF and δ\delta.

3. Property (QT) of relatively hyperbolic groups

In this section, we are going to prove Theorem 1.5. The notion of relatively hyperbolic groups can be formulated from a number of equivalent ways. Here we shall present a quick definition due to Bowditch [Bow12] and recall the relevant facts we shall need without proofs.

Let HH be a finitely generated group with a finite collection of subgroups 𝒫\mathcal{P}. Fixing a finite generating set SS, we consider the corresponding Cayley graph Cay(H,S)\operatorname{Cay}(H,S) equipped with the word metric ||H||_{H}.

Denote by ={hP:hH,P𝒫}\mathbb{P}=\{hP:h\in H,P\in\mathcal{P}\} the collection of peripheral cosets. Let H^()\hat{H}(\mathbb{P}) be the coned-off Cayley graph obtained from Cay(H,S)\operatorname{Cay}(H,S) as follows. A cone point denoted by c(P)c({P}) is added for each peripheral coset PP\in\mathbb{P} and is joined by half edges to each element in PP. The union of two half edges at a cone point is called a peripheral edge. Denote by ||H^|\cdot|_{\hat{H}} the induced length metric after coning-off.

The pair (G,𝒫)(G,\mathcal{P}) is said to be relatively hyperbolic if the coned-off Cayley graph H^()\hat{H}(\mathbb{P}) is hyperbolic and fine: any edge is contained in finitely many simple circles with uniformly bounded length.

By [Bow08, Lemma 3.3], [Osi16, Prop. 5.2], the action of HH on H^()\hat{H}(\mathbb{P}) is acylindrical.

Let πP\pi_{P} denote the shortest projection in word metric to PP\in\mathbb{P} in HH and dP(x,y)d_{P}(x,y) the ||H|\cdot|_{H}-diameter of the projections of the points x,yx,y to PP. Since \mathbb{P} has the strongly contracting property with bounded intersection property, the projection axioms with a constant ξ>0\xi>0 hold for \mathbb{P} (see [Sis13]).

3.1. Thick distance formula

A geodesic edge path β\beta in the coned-off Cayley graph H^()\hat{H}(\mathbb{P}) is KK-bounded for K>0K>0 if the end points of every peripheral edge have dd-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 |(ej)(ej)+|H>K|(e_{j})_{-}-(e_{j})_{+}|_{H}>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 the KK-thick distance

(3) |xy|H^K=max{|β|K}\bigl{|}x-y\bigr{|}_{\hat{H}}^{K}=\max\{|\beta|_{K}\}

over all relative geodesics β\beta between x,yx,y. Thus, |xy|H^K\bigl{|}x-y\bigr{|}_{\hat{H}}^{K} is HvH_{v}-invariant.

A relative path without backtracking in H^()\hat{H}(\mathbb{P}) admits non-unique lifts in Cay(H,S)\operatorname{Cay}(H,S) which are obtained by replacing the peripheral edge by a geodesic in Cay(H,S)\operatorname{Cay}(H,S) with the same endpoints. The distance formula follows from the fact that the lift of a relative quasi-geodesic is a quasi-geodesic (see [DS05], [GP16, Prop. 6.1]). The following formula is made explicitly in [Sis13, Theorem 0.1].

Lemma 3.1.

For any sufficiently large K>0K>0 and for any x,yHx,y\in H,

(4) |xy|HK|xy|H^K+P[dP(x,y)]K.|x-y|_{H}\sim_{K}\bigl{|}x-y\bigr{|}_{\hat{H}}^{K}+\sum_{P\in\mathbb{P}}[d_{P}(x,y)]_{K}.

The following result is proved in [NY, Lemma 5.5] under the assumption that HH is hyperbolic relative to a set of virtually cyclic subgroups. However, the same proof works for any relatively hyperbolic group.

Lemma 3.2.

For any sufficiently large K>0K>0, there exists an HH–finite collection 𝔸\mathbb{A} of quasi-lines in H^\hat{H} and a constant N=N(K,H^,𝔸)>0N=N(K,\hat{H},\mathbb{A})>0, such that for any two vertices x,yH^x,y\in\hat{H}, the following holds

(5) |xy|H^KN𝔸[d^(x,y)]K\bigl{|}x-y\bigr{|}_{\hat{H}}^{K}\sim_{N}\sum_{\ell\in\mathbb{A}}[\hat{d}_{\ell}(x,y)]_{K}

A group HH endowed with the profinite topology is a topological group so that the set of all finite index subgroups is a (close/open) neighborhood base of the identity. A subgroup PP is called separable if it is closed in the profinite topology. Equivalently, it is the intersection of all finite index subgroups containing PP. A group is called residually finite if the trivial subgroup is closed.

A maximal abelian subgroup of a residually finite group is separable (see [Ham01, Proposition 1]). Note that a maximal elementary (i.e. virtually cyclic) group EE in a relatively hyperbolic group HH contains a maximal abelian group (of rank 1) as a finite index subgroup. If HH is residually finite, then EE as a finite union of closed subsets is closed and thus separable.

We will use the following corollary in the proof of Theorem 1.5.

Corollary 3.3.

Assume that HH is a residually finite relatively hyperbolic group. Then for any K0K\gg 0, there exists a finite index subgroup H˙\dot{H} acting on finitely many quasi-trees TiT_{i} (1in)(1\leq i\leq n) such that the orbital map of the H˙\dot{H}-action on i=1nTi\prod_{i=1}^{n}T_{i} is a quasi-isometric embedding from (H˙,||H^K)(\dot{H},|\cdot|^{K}_{\hat{H}}) to i=1nTi\prod_{i=1}^{n}T_{i}.

This corollary is essentially proved in [NY], inspired by the arguments in the setting of mapping class groups [BBF19]. We sketch the proof at the convenience of the reader.

Sketch of proof.

Recall that for any θ>0\theta>0, a set 𝕋\mathbb{T} of (uniform) quasi-lines in a hyperbolic space with θ\theta-bounded projection satisfies the projection axioms with projection constant ξ\xi for a constant ξ=ξ(θ)>0\xi=\xi(\theta)>0. Let λ\lambda and cc be the constants given by Lemma 2.30 with respect to the acylindrical action HH^H\curvearrowright\hat{H}. For our purpose, we will choose θ\theta to be the constant given by Proposition 2.29. Then the distance formula for the quasi-tree 𝒞K(𝕋)\mathcal{C}_{K}(\mathbb{T}) constructed from 𝕋\mathbb{T} holds for any K4ξK\geq 4\xi.

For a fixed large constant KK, Lemma 3.2 provides an HH-finite set of quasi-lines 𝔸\mathbb{A} so that (5) holds. We then use the separability to find a finite index subgroup H˙\dot{H} of HH so that 𝔸\mathbb{A} decomposes as a finite union of H˙\dot{H}-invariant 𝕋i\mathbb{T}_{i}’s each of which satisfies the projection axioms with projection constant ξ\xi. To be precise, the stabilizer EE of a quasi-line \ell in 𝔸\mathbb{A} is a maximal elementary subgroup of HH and thus is separable in HH if HH is residually finite (since a maximal abelian group in a residually finite group is separable). By Proposition 2.29 and the paragraph after Lemma 2.1 in [BBF19], the separability of EE allows one to choose a finite index subgroup H˙\dot{H} containing EE such that any H˙\dot{H}-orbit 𝕋i\mathbb{T}_{i} in the collection of quasi-lines HH\ell satisfies the projection axioms with projection constant ξ\xi. We take a common finite index subgroup H˙\dot{H} for finitely many quasi-lines \ell in 𝔸\mathbb{A} up to HH-orbits and therefore have found all H˙\dot{H}-orbit 𝕋i\mathbb{T}_{i} so that their union covers 𝔸\mathbb{A}.

Finally, it is straightforward to verify that the right-hand term of (5) coincides with the sum of distances over the finitely many quasi-trees Ti:=𝒞K(𝕋i)T_{i}:=\mathcal{C}_{K}(\mathbb{T}_{i}). Thus, the thick distance dH^K(x,y)d_{\hat{H}}^{K}(x,y) is quasi-isometric to the distance on a finite product of quasi-trees. ∎

All our discussion generalizes to the geometric action of HH on a geodesic metric space YY, since there exists a HH-equivariant quasi-isometry between Cay(H,S)\operatorname{Cay}(H,S) and YY. Therefore, replacing HH with YY, we have the same thick distance formula. This is the setup for CKA actions in next sections.

In next subsection, we obtain the property (QT) for relatively hyperbolic groups provided peripheral subgroups do so.

3.2. Proof of property (QT) of relatively hyperbolic groups

Proof of Theorem 1.5.

Recall that 𝒫\mathcal{P} is a finite set of subgroups. For each P𝒫P\in\mathcal{P}, choose a full set EPE_{P} of left PP-coset representatives in HH so that 1EP1\in E_{P}. For given PP and 1inP1\leq i\leq n_{P}, we define the collection of quasi-trees

𝕋Pi:={fTi:fEP}\mathbb{T}_{P}^{i}:=\{fT_{i}:f\in E_{P}\}

where TiT_{i} are quasi-trees associated to PP given in assumption. Then HH preserves 𝕋Pi\mathbb{T}_{P}^{i} by the following action: for any point f(x)fTif(x)\in fT_{i} and hHh\in H,

hf(x):=fp(x)fTih\cdot f(x):=f^{\prime}p(x)\in f^{\prime}T_{i}

where pPp\in P is given by hf=fphf=f^{\prime}p for fEPf^{\prime}\in E_{P}.

We are now going to define projection maps {πfTi}\{\pi_{fT_{i}}\} as follows.

By assumption, we fix an orbital embedding ιPi\iota^{i}_{P} of PP into TiT_{i} so that the induced map i=1nPιPi:Pi=1nPTi\prod_{i=1}^{n_{P}}\iota^{i}_{P}:P\to\prod_{i=1}^{n_{P}}T_{i} is a quasi-isometric embedding. We then define an equivariant family of orbital maps ιfPi:fPfTi\iota_{fP}^{i}:fP\to fT_{i} so that

xfP,ιfPi(x):=fιPi(f1x).\forall x\in fP,\;\iota_{fP}^{i}(x):=f\iota_{P}^{i}(f^{-1}x).

Then for any hHh\in H and xfPx\in fP, hιfPi(x)=ιfPi(hx)h\cdot\iota_{fP}^{i}(x)=\iota_{f^{\prime}P}^{i}(hx) where fEPf^{\prime}\in E_{P} with hf=fphf=f^{\prime}p and pPp\in P.

Let πfP\pi_{fP} be the shortest projection to the coset fPfP in HH with respect to the word metric. For any two distinct fTi,fTi𝕋PifT_{i},f^{\prime}T_{i}\in\mathbb{T}_{P}^{i}, we set

πfTi(fTi):=ιfPi(πfP(fP))\pi_{fT_{i}}(f^{\prime}T_{i}):=\iota_{fP}^{i}(\pi_{fP}(f^{\prime}P))

Recall that ={fP:fH,P𝒫}\mathbb{P}=\{fP:f\in H,P\in\mathcal{P}\} satisfies the projection axioms with shortest projection maps πfP\pi_{fP}’s. It is readily checked that the projection axioms pass to the collection 𝕋Pi\mathbb{T}^{i}_{P} under equivariant Lipschitz maps {ιfPi}fP\{\iota_{fP}^{i}\}_{fP\in\mathbb{P}}.

We can therefore build the projection complex for 𝕋Pi\mathbb{T}_{P}^{i} for a fixed K0K\gg 0. By Proposition 2.27, the following distance holds for any x,y𝒞K(𝕋Pi)x^{\prime},y^{\prime}\in\mathcal{C}_{K}(\mathbb{T}_{P}^{i}):

(6) |xy|𝒞K(𝕋Pi)KT𝕋Pi[dT(x,y)]K.\displaystyle\bigl{|}x^{\prime}-y^{\prime}\bigl{|}_{\mathcal{C}_{K}(\mathbb{T}_{P}^{i})}\sim_{K}\sum_{T\in\mathbb{T}_{P}^{i}}[d_{T}(x^{\prime},y^{\prime})]_{K}.

Note that i=1nPιPi:Pi=1nPTi\prod_{i=1}^{n_{P}}\iota^{i}_{P}:P\to\prod_{i=1}^{n_{P}}T_{i} is a quasi-isometric embedding for each P𝒫P\in\mathcal{P}. Thus, for any x,yGx,y\in G and PP\in\mathbb{P},

(7) dP(x,y)=|πP(x)πP(y)|Pi=1nP|ιPi(πP(x))ιPi(πP(y))|Ti.\displaystyle d_{P}(x,y)=|\pi_{P}(x)-\pi_{P}(y)|_{P}\sim\sum_{i=1}^{n_{P}}\bigl{|}\iota^{i}_{P}(\pi_{P}(x))-\iota^{i}_{P}(\pi_{P}(y))\bigr{|}_{T_{i}}.

Setting x=ιPi(πP(x))x^{\prime}=\iota^{i}_{P}(\pi_{P}(x)) and y=ιPi(πP(y))y^{\prime}=\iota^{i}_{P}(\pi_{P}(y)) in (7), we deduce from (6) that

(8) dP(x,y)Ki=1nP|ιPi(πP(x))ιPi(πP(y))|𝒞K(𝕋Pi).\displaystyle d_{P}(x,y)\preceq_{K}\sum_{i=1}^{n_{P}}\bigl{|}\iota_{P}^{i}(\pi_{P}(x))-\iota^{i}_{P}(\pi_{P}(y))\bigr{|}_{\mathcal{C}_{K}(\mathbb{T}_{P}^{i})}.

Recall from Lemma 3.1 that for any x,yHx,y\in H, we have

|xy|HK|xy|H^K+P[dP(x,y)]K.|x-y|_{H}\sim_{K}\bigl{|}x-y\bigr{|}_{\hat{H}}^{K}+\sum_{P\in\mathbb{P}}[d_{P}(x,y)]_{K}.

Note that the orbital map of any isometric action is Lipschitz. To prove property (QT) of HH, it suffices to give an upper bound of |xy|H|x-y|_{H}. Taking account of (8), it remains to construct a finite product of quasi-trees to bound |xy|H^K\bigl{|}x-y\bigr{|}_{\hat{H}}^{K} as follows.

Since HH is residually finite, by Corollary 3.3, there exists a finite index subgroup, still denoted by HH, and a finite product YY of quasi-trees so that the orbital map Π0\Pi_{0} from HH to YY gives a quasi-isometric embedding of HH equipped with ||H^K|\cdot|^{K}_{\hat{H}}-function into YY.

Recall that πP\pi_{P} is the shortest projection to PP\in\mathbb{P}. For 1inP1\leq i\leq n_{P}, define

Πi:H𝒞K(𝕋Pi)\Pi_{i}:H\to\mathcal{C}_{K}(\mathbb{T}_{P}^{i})

by sending an element hHh\in H to ιPi(πP(h))\iota^{i}_{P}(\pi_{P}(h)). We then have nn equivariant maps Πi\Pi_{i} of HH to quasi-trees after re-indexing, where n:=P𝒫nPn:=\sum_{P\in\mathcal{P}}n_{P}.

Let Π:=Π0×i=1nΠi\Pi:=\Pi_{0}\times\prod_{i=1}^{n}\Pi_{i} be the map from HH to Y×i=1n𝒞K(𝕋Pi)Y\times\prod_{i=1}^{n}\mathcal{C}_{K}(\mathbb{T}_{P}^{i}), where YY is the finite product of quasi-trees as in the previous paragraphs. As fore-mentioned, the product map Π\Pi gives an upper bound on dH(x,y)d_{H}(x,y), so is a quasi-isometric embedding of HH. Therefore, HH has property (QT). ∎

Remark 3.4.

An immediate corollary of Theorem 1.5 is that the fundamental group of a finite volume hyperbolic 3-manifold has property (QT). Alternate proof is that π1(M)\pi_{1}(M) is virtually compact special by deep theorems of Agol and Wise (see [Ago13] [Wis20]), thus π1(M)\pi_{1}(M) has property (QT).

We say that the profinite topology on HH induces a full profinite topology on a subgroup PP if every finite index subgroup of PP contains the intersection of PP with a finite index subgroup in HH.

Theorem 3.5.

Suppose that HH is residually finite and each P𝒫P\in\mathcal{P} is separable. Assume furthermore that HH induces the full profinite topology on each P𝒫P\in\mathcal{P}. If each P𝒫P\in\mathcal{P} acts by isometry on a finite product of quasi-trees without \mathbb{R}-factor such that orbital maps are quasi-isometric embeddings, then HH has property (QT).

Proof.

By [FL08, Corollary 1.3], there is a finite index subgroup P˙\dot{P} of PP acting on each quasi-tree TiT_{i} so that the diagonal action of P˙\dot{P} on i=1nTi\prod_{i=1}^{n}T_{i} induces a quasi-isometric embedding orbital map i=1nιP˙i\prod_{i=1}^{n}\iota_{\dot{P}}^{i}.

By the assumption, HH induces the full profinite topology on P𝒫P\in\mathcal{P}, so every finite index subgroups of a separable subgroup PP is also separable. Thus, there are finite index subgroups H˙P\dot{H}_{P} of HH for P𝒫P\in\mathcal{P} such that P˙=H˙PP.\dot{P}=\dot{H}_{P}\cap P.

Consider the finite index normal subgroup H˙:={hH˙Ph1:P𝒫}\dot{H}:=\cap\{h\dot{H}_{P}h^{-1}:P\in\mathcal{P}\} in HH. Since H˙\dot{H} is normal in HH, we see that H˙hPh1hP˙h1\dot{H}\cap hPh^{-1}\subset h\dot{P}h^{-1} is equivalent to H˙PP˙\dot{H}\cap P\subset\dot{P}. The later holds by the choice of H˙P\dot{H}_{P}. Hence, for every hHh\in H, H˙hPh1\dot{H}\cap hPh^{-1} preserves the factors of the product decomposition. Note that H˙\dot{H} is hyperbolic relative to {H˙hPh1:hH}\{\dot{H}\cap hPh^{-1}:h\in H\}. The conclusion follows from Theorem 1.5. ∎

In next sections (Sections 4,  5,  6 and  7), the proof of property (QT) of CKA groups will be discussed, which may be considered as the technical heart of this paper.

4. Coning off CKA spaces

In this section, we recapitulate the content of [NY, Sect. 5] and give an outline of the proof of Theorem 1.3.

Let GXG\curvearrowright X be a CKA action where GG is the fundamental group of an admissible graph of groups 𝒢\mathcal{G} (see Subsection 2.3), and let GTG\curvearrowright T be the action of GG on the associated Bass-Serre tree TT of 𝒢\mathcal{G}. Let T0T^{0} and T1T^{1} be the vertex and edge sets of TT.

Let {Fe}\{F_{e}\} be the collection of boundary planes of the space YvY_{v} (see Subsection 2.3). We note that the intersection of a boundary plane FeF_{e} of YvY_{v} with the hyperbolic space Y¯v\overline{Y}_{v} is a line. We define the collection of lines 𝕃v\mathbb{L}_{v} of the hyperbolic space Y¯v\overline{Y}_{v} as follows:

𝕃v={e:=FeY¯v|e=v}\mathbb{L}_{v}=\{\ell_{e}:=F_{e}\cap\overline{Y}_{v}\,|\,e_{-}=v\}

which shall be referred boundary lines.

4.1. Construction of coned-off spaces

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. Let G˙<G\dot{G}<G be the subgroup of index at most 22 preserving 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2} given by Lemma 2.23.

Fix a large r>0r>0. A hyperbolic rr-cone by definition is the metric completion of the (incomplete) universal cover of a punctured hyperbolic disk of radius rr. Let 𝕐i={Y¯v:v𝒱i}\mathbb{Y}_{i}=\{\overline{Y}_{v}:v\in\mathcal{V}_{i}\} be the collection of hyperbolic spaces and 𝕐˙i={Y˙v:v𝒱i}\dot{\mathbb{Y}}_{i}=\{\dot{Y}_{v}:v\in\mathcal{V}_{i}\} their coned-off spaces (which are uniformly hyperbolic for r0r\gg 0) by attaching hyperbolic rr-cones along the boundary lines of Y¯v\overline{Y}_{v}.

Note that G˙\dot{G} preserves 𝕐i\mathbb{Y}_{i} and 𝕐˙i\dot{\mathbb{Y}}_{i} by the action on the index gYv=YgvgY_{v}=Y_{gv} for any gG˙g\in\dot{G}. For each wT0w\in T^{0}, let St(w)St(w) be the star of ww in TT with adjacent vertices as extremities. Then St(w)St(w) admits the action of GwG_{w} so that the stabilizers of the extremities are the corresponding edge groups.

Define 𝒳i˙\dot{\mathcal{X}_{i}} to be the space obtained from the disjoint union of coned-off spaces Y˙v\dot{Y}_{v} (v𝒱i)(v\in\mathcal{V}_{i}) with cone points identified with the extremities of the stars St(w)St(w) with vLk(w)v\in Lk(w). Endowed with induced length metric, the space 𝒳i˙\dot{\mathcal{X}_{i}} is a Gromov-hyperbolic space.

Lemma 4.1.

Fix a sufficiently large r>0r>0 and i{1,2}i\in\{1,2\}. The space 𝒳i˙\dot{\mathcal{X}_{i}} is a δ\delta–hyperbolic space where δ>0\delta>0 only depends on the hyperbolicity constants of Y˙v\dot{Y}_{v} (v𝒱iv\in\mathcal{V}_{i}).

The subgroup G˙\dot{G} acts on 𝒳i˙\dot{\mathcal{X}_{i}} with the following properties:

  1. (1)

    for each v𝒱iv\in\mathcal{V}_{i}, the stabilizer of Y˙v\dot{Y}_{v} is isomorphic to GvG_{v} and HvH_{v} acts co-boundedly on Y˙v\dot{Y}_{v}, and

  2. (2)

    for each wT0𝒱iw\in T^{0}-\mathcal{V}_{i}, GwG_{w} acts on St(w)St(w) in the same manner of the action on Bass-Serre tree TT.

Proof.

Note that the stabilizers of the cone points of Y˙v\dot{Y}_{v} under the action of GvG_{v} on Y˙v\dot{Y}_{v} are the same as that of the extremities of stars St(w)St(w), which are both the edge groups GeG_{e} for e=[v,w]e=[v,w]. By construction, the cone points of Y˙v\dot{Y}_{v} are identified with the extremities of stars St(w)St(w), so the actions of GvG_{v} on Y˙v\dot{Y}_{v} (v𝒱iv\in\mathcal{V}_{i}) and of GwG_{w} on St(w)St(w) (wT0𝒱iw\in T^{0}-\mathcal{V}_{i}) extend over 𝒳i˙\dot{\mathcal{X}_{i}}, and hence G˙\dot{G} acts by isometries on 𝒳i˙\dot{\mathcal{X}_{i}}. ∎

Remark 4.2.

Our construction of coned-off spaces is slightly different from the one in [NY, Section 5.1], where the cone points are identified directly between different spaces Y˙v\dot{Y}_{v} and Y˙v\dot{Y}_{v^{\prime}}. Thus certain assumption on vertex groups is necessary in [NY] to ensure an action on the coned-off space.

We now define the thick distance on 𝒳i˙\dot{\mathcal{X}_{i}} (i=1,2)(i=1,2) by taking the sum of thick distances through Y˙v\dot{Y}_{v} as follows.

If xx is a point in a coned-off space Y˙v𝒳˙\dot{Y}_{v}\subset\dot{\mathcal{X}}, we denote ρ(x)\rho(x) by vv (by abuse of notations). By the above tree-like construction, any path between x,y𝒳i˙x,y\in\dot{\mathcal{X}_{i}} has to pass through in order a pair of boundary lines v,v+\ell_{v}^{-},\ell_{v}^{+} of Y¯v\overline{Y}_{v} for each v[ρ(x),ρ(y)]v\in[\rho(x),\rho(y)]. By abuse of language, if xx is not contained in a hyperbolic cone, set v=x\ell_{v}^{-}=x for v=ρ(x)v=\rho(x). Similarly, if yy is not contained in a hyperbolic cone, set v+=y\ell_{v}^{+}=y for v=ρ(y)v=\rho(y).

Let (xv,yv)(x_{v},y_{v}) be a pair of points in the boundary lines (v,v+)(\ell_{v}^{-},\ell_{v}^{+}) so that [xv,yv][x_{v},y_{v}] is orthogonal to v\ell_{v}^{-} and v+\ell_{v}^{+}. Recall that |xvyv|Y˙vK\bigl{|}x_{v}-y_{v}\bigr{|}_{\dot{Y}_{v}}^{K} is the KK-cut-off thick distance defined in (3).

Definition 4.3.

For any K0K\geq 0, the KK-thick distance between xx and yy is defined by

(9) |xy|𝒳i˙K:=v[ρ(x),ρ(y)]𝒱i|xvyv|Y˙vK.\bigl{|}x-y\bigr{|}^{K}_{\dot{\mathcal{X}_{i}}}:=\sum_{v\in[\rho(x),\rho(y)]\cap\mathcal{V}_{i}}\bigl{|}x_{v}-y_{v}\bigr{|}_{\dot{Y}_{v}}^{K}.

Since ||Y˙vK|\cdot|_{\dot{Y}_{v}}^{K} is HvH_{v}-invariant, we see that |xy|𝒳i˙K\bigl{|}x-y\bigr{|}^{K}_{\dot{\mathcal{X}_{i}}} is G˙\dot{G}-invariant.

Remark 4.4.

The definition of ||𝒳i˙K|\cdot|^{K}_{\dot{\mathcal{X}_{i}}} is designed to ignore the parts in hyperbolic cones between different pieces. One consequence is that perturbing x,yx,y in hyperbolic cones does not change their KK-thick distance.

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

If E()E(\ell) denotes the stabilizer in HvH_{v} of a boundary line \ell of Y¯v\overline{Y}_{v}, then E()E(\ell) is virtually cyclic and almost malnormal. Since {E()}\{E(\ell)\} is HvH_{v}-finite by conjugacy, let 𝔼v\mathbb{E}_{v} be a complete finite set of conjugacy representatives. By Lemma 2.16, HvH_{v} is hyperbolic relative to peripheral subgroups 𝔼v\mathbb{E}_{v}. Hence, the results in Section 3 apply here.

Let λ,c>0\lambda,c>0 be the universal constants given by Lemma 2.30 applied to the actions of HvH_{v} on Y˙v\dot{Y}_{v} for all vT0v\in T^{0} (since there are only finitely many actions up to conjugacy). By convention, the quasi-lines in coned-off spaces are understood as (λ,c)(\lambda,c)–quasi-geodesics in 𝒳i˙\dot{\mathcal{X}_{i}} and Y˙v\dot{Y}_{v}’s.

The coning-off construction has the following consequence ([NY, Lemma 5.14]): the shortest projection of any quasi-line α\alpha in Y˙v\dot{Y}_{v} to a quasi-line β\beta in Y˙v\dot{Y}_{v^{\prime}} has to pass through the cone point attached to Y˙v\dot{Y}_{v^{\prime}}, and thus has uniformly bounded diameter by θ=θ(λ,c)>0\theta=\theta(\lambda,c)>0.

For simplicity, we also assume that θ=θ(λ,c)>0\theta=\theta(\lambda,c)>0 satisfies the conclusion of Proposition 2.29. Consequently, this determines a constant ξ=ξ(θ)>0\xi=\xi(\theta)>0 such that any set of quasi-lines with θ\theta-bounded projection satisfies the projection axioms with projection constant ξ\xi.

Fix K>max{4ξ,θ}K>\max\{4\xi,\theta\}. For each v𝒱v\in\mathcal{V}, there exists an HvH_{v}-finite collection of quasi-lines 𝔸v\mathbb{A}_{v} in Y˙v\dot{Y}_{v} and a constant N=N(𝔸v,K)N=N(\mathbb{A}_{v},K) such that dH^vKd^{K}_{\hat{H}_{v}}-distance formula holds by Lemma 3.2.

Since G˙\dot{G} acts co-finitely on 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2}, we can assume 𝔸w=g𝔸v\mathbb{A}_{w}=g\mathbb{A}_{v} if w=gvw=gv for gG˙g\in\dot{G}. Let

𝔸i:=v𝒱i𝔸v\mathbb{A}_{i}:=\cup_{v\in\mathcal{V}_{i}}\mathbb{A}_{v}

be the union of 𝔸v\mathbb{A}_{v}’s for i=1,2i=1,2 which are both G˙\dot{G}-invariant. We now equip 𝔸i\mathbb{A}_{i} with projection maps as the shortest projection maps between two quasi-lines in 𝒳i˙\dot{\mathcal{X}_{i}} for i=1,2i=1,2.

If γ\gamma is a quasi-line in 𝒳i˙\dot{\mathcal{X}_{i}} for i=1,2i=1,2, denote by d˙γ(x,y)\dot{d}_{\gamma}(x,y) the ||𝒳i˙|\cdot|_{\dot{\mathcal{X}_{i}}}–diameter of the shortest projection of x,y𝒳i˙x,y\in\dot{\mathcal{X}_{i}} to γ\gamma.

The following result shows that the thick distance is captured by the projections of 𝔸i\mathbb{A}_{i}. Recall that rr is the radius of the hyperbolic cones in constructing 𝒳i˙\dot{\mathcal{X}_{i}}.

Proposition 4.5.

[NY, Prop. 5.9] For any x,y𝒳i˙x,y\in\dot{\mathcal{X}_{i}}, the following holds

(10) |xy|𝒳i˙Kr,Kγ𝔸i[d˙γ(x,y)]K+|ρ(x)ρ(y)|T.\begin{array}[]{cc}\bigl{|}x-y\bigr{|}_{\dot{\mathcal{X}_{i}}}^{K}\;\sim_{r,K}\;\sum_{\gamma\in\mathbb{A}_{i}}[\dot{d}_{\gamma}(x,y)]_{K}+|\rho(x)-\rho(y)|_{T}.\end{array}

In the next subsection, we construct a suitable finite subgroup of GG such that it acts isometrically on a finite product of quasi-trees T1,,TnT_{1},\cdots,T_{n} under some assumptions on vertex groups. This allows rewriting the right-hand side of the distance formula (10) as the product distance of TiT_{i}’s.

4.3. Isometric action of a suitable finite index subgroup of GG

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

Definition 4.6.

A group HH is omnipotent if for any non-empty 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.

Let GXG\curvearrowright X be a CKA action, where GG is the admissible graph of groups 𝒢\mathcal{G} so that every vertex group GvG_{v} is a central extension of an omnipotent hyperbolic group. By Lemma 4.1, the finite index subgroup G˙\dot{G} acts on 𝒳1˙×𝒳2˙×T\dot{\mathcal{X}_{1}}\times\dot{\mathcal{X}_{2}}\times T that is equipped with the G˙\dot{G}-invariant function ||𝒳1˙K×||𝒳2˙K×||T|\cdot|_{\dot{\mathcal{X}_{1}}}^{K}\times|\cdot|_{\dot{\mathcal{X}_{2}}}^{K}\times|\cdot|_{T}. The main result of this subsection is the following.

Proposition 4.7.

The group G˙\dot{G} admits finitely many isometric actions on quasi-trees TiT_{i} for 1in1\leq i\leq n such that there exists a G˙\dot{G}-equivariant quasi-isometric embedding from 𝒳1˙×𝒳2˙×T\dot{\mathcal{X}_{1}}\times\dot{\mathcal{X}_{2}}\times T to T1×T2×Tn×TT_{1}\times T_{2}\cdots\times T_{n}\times T.

We emphasize here that ||𝒳1˙K×||𝒳2˙K×||T|\cdot|_{\dot{\mathcal{X}_{1}}}^{K}\times|\cdot|_{\dot{\mathcal{X}_{2}}}^{K}\times|\cdot|_{T} on the domain for the quasi-isometric embedding is not a distance function, but the target is equipped with product distance.

By [BH99, Theorem II.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}. Since HvH_{v} is omnipotent and then is residually finite, we can assume that KvK_{v} is torsion-free.

Recall the G˙\dot{G}-invariant collection of quasi-lines in Subsection 4.2:

𝔸i=v𝒱i𝔸v\mathbb{A}_{i}=\cup_{v\in\mathcal{V}_{i}}\mathbb{A}_{v}

where 𝔸v\mathbb{A}_{v} is the collection of quasi-lines so that dKd^{K}-distance formula holds by Lemma 3.2. By the residual finiteness of KvK_{v}, there exists a finite index subgroup K˙v\dot{K}_{v} so that 𝔸v\mathbb{A}_{v} is partitioned into Kv˙\dot{K_{v}}-invariant sub-collections with projection constants ξ\xi.

To prepare the proof, we need to introduce a compatible condition of glueing finitely index subgroups. A collection of finite index subgroups {Ge,Gv:v𝒢0,e𝒢1}\{G_{e}^{\prime},G_{v}^{\prime}: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}.

By [DK18, Theorem 7.51], a compatible collection of finite index subgroups gives a finite index subgroup of GG. The following result says that upon taking finite index subgroups, we can assume that each vertex group is a direct product in a CKA group.

Lemma 4.8.

Let {K˙v<Kv:v𝒢0}\{\dot{K}_{v}<K_{v}:v\in\mathcal{G}^{0}\} be a collection of finite index subgroups. Then there exist finite index subgroups K¨v\ddot{K}_{v} of K˙v\dot{K}_{v}, GeG_{e}^{\prime} of GeG_{e} and ZvZ_{v} of Z(Gv)Z(G_{v}) so that the collection of finite index subgroups {Ge,Gv=K¨v×Zv:v𝒢0,e𝒢1}\{G_{e}^{\prime},G_{v}^{\prime}=\ddot{K}_{v}\times Z_{v}:v\in\mathcal{G}^{0},e\in\mathcal{G}^{1}\} is compatible.

Assuming Lemma 4.8, we now complete the proof of Proposition 4.7.

Proof.

We pass to further finite index subgroups K¨v<K˙v\ddot{K}_{v}<\dot{K}_{v} satisfying compatible conditions, which then gives a further index subgroup G¨G˙\ddot{G}\subset\dot{G}. For i=1,2i=1,2, let us partition 𝔸i=k=1ni𝔸ki\mathbb{A}_{i}=\cup_{k=1}^{n_{i}}\mathbb{A}_{k}^{i} into G¨\ddot{G}-obits 𝔸ki\mathbb{A}_{k}^{i}. By the construction of G¨\ddot{G}, we know that G¨\ddot{G} intersects each vertex group GvG_{v} of the Bass-Serre tree in a (conjugate) subgroup K¨v\ddot{K}_{v}. Thus, for each kk, 𝔸ki\mathbb{A}_{k}^{i} are the union of certain Kv¨\ddot{K_{v}}-invariant sub-collections where vv are varied in 𝒱i\mathcal{V}_{i}.

Recall that 𝔸i\mathbb{A}^{i} for i=1,2i=1,2 satisfies the projection axioms with a uniform projection constant ξ\xi in Subsection 4.2. We can then build the quasi-trees Tki:=𝒞K(𝔸ki)T_{k}^{i}:=\mathcal{C}_{K}(\mathbb{A}_{k}^{i}) where 1kni1\leq k\leq n_{i}. Setting n=n1+n2n=n_{1}+n_{2}, this thus yields isometric group actions of G¨\ddot{G} on quasi-trees TiT_{i} (1in)(1\leq i\leq n).

We first construct a G¨\ddot{G}-equivariant map Φ\Phi from 𝒳1˙×𝒳2˙×T\dot{\mathcal{X}_{1}}\times\dot{\mathcal{X}_{2}}\times T to T1×T2×Tn×TT_{1}\times T_{2}\cdots\times T_{n}\times T. By equivariance, it suffices to fix a basepoint in each 𝒳1,𝒳2,T\mathcal{X}_{1},\mathcal{X}_{2},T and TiT_{i}’s so that Φ\Phi sends basepoints to basepoints. The quasi-isometric embedding property follows from the distance formula (10), where the right-hand side is now replaced by the distance in the corresponding quasi-trees.

Note that G¨\ddot{G} is of finite index in G˙\dot{G}. By taking more copies of quasi-trees TiT_{i} in the target, the map Φ\Phi can be made G˙\dot{G}-equivariant. Indeed, if a finite index subgroup HGH\subset G acts on some space XX then GG acts on a finite product of [G:H][G:H] copies of XX without preserving the factors. The map Φ\Phi can be extended to these copies as well. The proof of the Proposition is complete. ∎

We now give the proof of Lemma 4.8.

Proof of Lemma 4.8.

Assume that fv=Z(Gv)\langle f_{v}\rangle=Z(G_{v}) for any v𝒢0v\in\mathcal{G}^{0}. Then for an oriented edge e=[v,w]e=[v,w] from vv to ww, the subgroup fv,fw\langle f_{v},f_{w}\rangle is of finite index in GeG_{e}.

Note that Ge2G_{e}\cong\mathbb{Z}^{2} admits a base {f^v,b^e}\{\hat{f}_{v},\hat{b}_{e}\} where f^v\hat{f}_{v} is primitive so that fvf_{v} is some power of f^v\hat{f}_{v}. Let πv:GvHv=Gv/Z(Gv)\pi_{v}:G_{v}\to H_{v}=G_{v}/Z(G_{v}). Thus, πv(Ge)\pi_{v}(G_{e}) is a direct product of a torsion group with be\langle b_{e}\rangle in HvH_{v}, where be=πv(b^e)b_{e}=\pi_{v}(\hat{b}_{e}) is a loxodromic element.

Similarly, let f^w,b^e¯Ge\hat{f}_{w},\hat{b}_{\overline{e}}\in G_{e} such that f^w,b^e¯=Ge\langle\hat{f}_{w},\hat{b}_{\overline{e}}\rangle=G_{e}. Keep in mind that for any integer n0n\neq 0, f^vn,b^en=b^e¯n,f^wn\langle\hat{f}_{v}^{n},\hat{b}_{e}^{n}\rangle=\langle\hat{b}_{\overline{e}}^{n},\hat{f}_{w}^{n}\rangle is of finite index in GeG_{e}.

We choose an integer m0m\neq 0 such that b^emK˙v\hat{b}_{e}^{m}\in\dot{K}_{v} for every vertex v𝒢0v\in\mathcal{G}^{0} and every oriented edge ee from e=ve_{-}=v. Such an integer mm exists since K˙v\dot{K}_{v} injects into HvH_{v} as a finite index subgroup, and 𝒢\mathcal{G} is a finite graph of groups.

Apply the omnipotence of HvH_{v} to the independent set of elements {be:e=v}\{b_{e}:e_{-}=v\}. Let pvp_{v} be the constant given by Definition 4.6. Set

s:=mv𝒢0pvs:=m\prod_{v\in\mathcal{G}^{0}}p_{v}

Set lv=s/pvl_{v}=s/p_{v}. Thus, for the collection {be:e=v}\{b_{e}:e_{-}=v\}, there exists a finite quotient ξv:HvH¯v\xi_{v}:H_{v}\to\overline{H}_{v} such that ξv(be)\xi_{v}(b_{e}) has order s=lvpvs=l_{v}p_{v} and besker(ξv)b_{e}^{s}\in\ker(\xi_{v}). Then K¨v:=K˙vπv1ker(ξv)\ddot{K}_{v}:=\dot{K}_{v}\cap\pi_{v}^{-1}\ker(\xi_{v}) is of finite index in K˙v\dot{K}_{v}. Recall that πv|Kv:KvHv\pi_{v}|_{K_{v}}:K_{v}\to H_{v} is injective (see the paragraph before Lemma 4.8). Since πv(b^es)=bes\pi_{v}(\hat{b}_{e}^{s})=b_{e}^{s} is loxodromic in HvH_{v} and b^esK˙v\hat{b}_{e}^{s}\in\dot{K}_{v} for m|sm|s, we have b^es\hat{b}_{e}^{s} is a loxodromic element in K¨v\ddot{K}_{v}.

For each oriented edge e=[v,w]𝒢1e=[v,w]\in\mathcal{G}^{1}, define Gv:=f^vs×K¨vG_{v}^{\prime}:=\langle\hat{f}_{v}^{s}\rangle\times\ddot{K}_{v}, Gw:=f^ws×K¨wG_{w}^{\prime}:=\langle\hat{f}_{w}^{s}\rangle\times\ddot{K}_{w} and Ge:=f^vs,b^es=b^e¯s,f^ws<GvG_{e}^{\prime}:=\langle\hat{f}_{v}^{s},\hat{b}_{e}^{s}\rangle=\langle\hat{b}_{\overline{e}}^{s},\hat{f}_{w}^{s}\rangle<G_{v}^{\prime}.

Let gGeGvg\in G_{e}\cap G_{v}^{\prime} be any element so we can write g=f^vsmkg=\hat{f}_{v}^{sm}k for some mm\in\mathbb{Z} and kK¨vk\in\ddot{K}_{v}. Recall that πv(Ge)\pi_{v}(G_{e}) is a direct product of be\langle b_{e}\rangle and a torsion group, and K¨v\ddot{K}_{v} is torsion-free. So πv(g)=πv(k)πv(Ge)πv(K¨v)\pi_{v}(g)=\pi_{v}(k)\in\pi_{v}(G_{e})\cap\pi_{v}(\ddot{K}_{v}) is some power of beb_{e}: πv(k)=bel\pi_{v}(k)=b_{e}^{l} for some ll\in\mathbb{Z}. Note that bel=πv(k)ker(ξv)b_{e}^{l}=\pi_{v}(k)\in\ker(\xi_{v}) so the omnipotence implies that s|ls|l, i.e. l=nsl=ns for some nn\in\mathbb{Z}. Since be=πv(b^e)b_{e}=\pi_{v}(\hat{b}_{e}) and πv:K¨vHv\pi_{v}:\ddot{K}_{v}\to H_{v} is injective, we obtain that k=b^ensk=\hat{b}_{e}^{ns}. Therefore, g=f^vsmb^ensGeg=\hat{f}_{v}^{sm}\hat{b}_{e}^{ns}\in G_{e}^{\prime} which implies

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

Therefore, the collection {Gv,Ge|v𝒢0,e𝒢1}\{G^{\prime}_{v},G^{\prime}_{e}\,\,|\,v\in\mathcal{G}^{0},e\in\mathcal{G}^{1}\} is verified to be compatible. ∎

4.4. Outline of the proof of Theorem 1.3

Let GXG\curvearrowright X be a CKA action where GG is the fundamental group of an admissible graph of groups 𝒢\mathcal{G} such that for every vertex group the central extension (1) has omnipotent hyperbolic quotient group. Recall that property (QT) is preserved undertaking finite index subgroups (see Lemma 2.3). Upon passing to further index subgroups in Lemma 4.8, we can assume that Gv=Hv×G_{v}=H_{v}\times\mathbb{Z}, where HvH_{v} acts geometrically on Y¯v\overline{Y}_{v} and also we can assume G˙=G\dot{G}=G. To show the property (QT) of GG, we must find not only a suitable action on a finite product of quasi-trees, but also ensure the distance of points in the image can recover word distance in the ambient group. We briefly describe here the strategy of the proof. Details are performed in Section 5 and Section 6.

Thanks to Proposition 4.7, we know that there exists a GG-equivariant quasi-isometric embedding (note that G˙=G\dot{G}=G)

𝒳1˙×𝒳2˙×TT1×T2×Tn×T\dot{\mathcal{X}_{1}}\times\dot{\mathcal{X}_{2}}\times T\to T_{1}\times T_{2}\cdots\times T_{n}\times T

Here TiT_{i} (with i{1,2,,n}i\in\{1,2,\cdots,n\}) is a quasi-tree. As the geometry of space 𝒳1˙×𝒳2˙×T\dot{\mathcal{X}_{1}}\times\dot{\mathcal{X}_{2}}\times T does not capture the distance from vertical parts of XX, there is no way finding a quasi-isometric embedding from the orbit GoGo to 𝒳1˙×𝒳2˙×T\dot{\mathcal{X}_{1}}\times\dot{\mathcal{X}_{2}}\times T. To overcome this obstacle, in Section 5, we will construct two additional quasi-trees, denoted by 𝒞K(𝔽1)\mathcal{C}_{K}(\mathbb{F}_{1}) and 𝒞K(𝔽2)\mathcal{C}_{K}(\mathbb{F}_{2}), and will show that there is indeed a GG-equivariant quasi-isometric embedding

Φ:Go𝒞K(𝔽1)×𝒞K(𝔽2)×𝒳1˙×𝒳2˙×T\Phi\colon Go\to\mathcal{C}_{K}(\mathbb{F}_{1})\times\mathcal{C}_{K}(\mathbb{F}_{2})\times\dot{\mathcal{X}_{1}}\times\dot{\mathcal{X}_{2}}\times T

(Section 6 is devoted to constructing Φ\Phi and verifying GG-equivariant quasi-isometric embedding of Φ\Phi). As a consequence, we obtain the desirable GG–equivariant quasi-isometric embedding

Go𝒞K(𝔽1)×𝒞K(𝔽2)×T1×T2×Tn×TGo\to\mathcal{C}_{K}(\mathbb{F}_{1})\times\mathcal{C}_{K}(\mathbb{F}_{2})\times T_{1}\times T_{2}\cdots\times T_{n}\times T

which entails property (QT) of GG.

5. Projection system of fiber lines

Recall we partition T0=𝒱1𝒱2T^{0}=\mathcal{V}_{1}\cup\mathcal{V}_{2} where 𝒱i\mathcal{V}_{i} consists of vertices in TT with pairwise even distances. For convenience, we sometimes write 𝒱=𝒱1\mathcal{V}=\mathcal{V}_{1} and 𝒲=𝒱2\mathcal{W}=\mathcal{V}_{2}. We note that property (QT) of a group is preserved under taking a finite index subgroup (see Lemma 2.3). Thus passing to a finite index subgroup (see Lemma 2.23) if necessary we could assume that GG is torsion-free and preserves 𝒱i\mathcal{V}_{i} with i=1,2i=1,2.

Note that e=[w,v]e=[w,v] is an oriented edge from ww towards vv, and e¯=[v,w]\overline{e}=[v,w] the oriented edge from vv towards ww. For each oriented edge ee, let FeF_{e} be the corresponding boundary plane. It is clear that Fe=Fe¯F_{e}=F_{\overline{e}} does not depend on the orientation.

5.1. Desired quasi-lines

By Lemma 2.17, the CKA space XX decomposes as the union of vertex spaces Y~v=ND(Yv)\tilde{Y}_{v}=N_{D}(Y_{v}) for vT0v\in T^{0}, on which the vertex groups GvG_{v} act geometrically. The center Z(Gv)Z(G_{v})\simeq\mathbb{Z} allows to split YvY_{v} as a metric product Y¯v×\overline{Y}_{v}\times\mathbb{R}. Upon passing to further finite index subgroups in Lemma 4.8, we can assume that Gv=Hv×G_{v}=H_{v}\times\mathbb{Z}, where HvH_{v} acts geometrically on Y¯v\overline{Y}_{v}. If the CKA action GXG\curvearrowright X is not flip (as in [NY]), the system of fiber lines \mathbb{R} in Yv=Y¯v×Y_{v}=\overline{Y}_{v}\times\mathbb{R} does not behave well with respect to the GG-action. Following the recent work [HRSS22], we introduce a better geometric model for vertex subgroups GvG_{v}, still as the metric product of Y¯v\overline{Y}_{v} with a quasi-line, to resolve the GG-action on the original fiber lines.

We first explain the construction of the quasi-line obtained from a quasi-homeomorphism. The following lemma is cited from Lemma 4.2 and the proof of the Corollary 4.3 in [HRSS22]. We present their proof as it is short and crucial for our discussion.

Lemma 5.1.

Let HH be a hyperbolic group relative to a finite collection of virtually cyclic subgroups {Ei:1in}\{E_{i}:1\leq i\leq n\}. Consider G=H×G=H\times\mathbb{Z} and fix a set of elements ciEi×c_{i}\in E_{i}\times\mathbb{Z} for each 1in1\leq i\leq n such that ci\langle c_{i}\rangle has unbounded projection to EiE_{i}. Then there exist a generating set S{S} of GG and a (λ,λ)(\lambda,\lambda)-quasi-isometry φ:Cay(G,S)\varphi:\operatorname{Cay}(G,{S})\to\mathbb{R} such that the following holds.

  1. (1)

    If gci,gcig\langle c_{i}\rangle,g^{\prime}\langle c_{i}\rangle are two ci\langle c_{i}\rangle-cosets for g,gEi×g,g^{\prime}\in E_{i}\times\mathbb{Z}, then

    λ1|gcigci|Gλ|φ(gci)φ(gci)|λ|gcigci|G+λ\lambda^{-1}|g\langle c_{i}\rangle-g^{\prime}\langle c_{i}\rangle|_{G}-\lambda\leq|\varphi(g\langle c_{i}\rangle)-\varphi(g^{\prime}\langle c_{i}\rangle)|\leq\lambda|g\langle c_{i}\rangle-g^{\prime}\langle c_{i}\rangle|_{G}+\lambda

    where |gcigci|G|g\langle c_{i}\rangle-g^{\prime}\langle c_{i}\rangle|_{G} denotes the distance between two subsets in GG equipped with a word metric relative to a finite generating set (so not the distance on Cay(G,S)\operatorname{Cay}(G,{S})).

  2. (2)

    With the natural action of GHG\to H, the diagonal action of G=H×G=H\times\mathbb{Z} on H×Cay(G,S)H\times\operatorname{Cay}(G,{S}) is metrically proper and cobounded, where G\mathbb{Z}\subset G acts loxodromically on Cay(G,S)\operatorname{Cay}(G,{S}) but ci\langle c_{i}\rangle acts boundedly.

In applications, the choice of elements cic_{i} shall come from the fiber generator of the adjacent pieces. See Lemma 5.2 below.

Proof.

Let πH:G=H×H\pi_{H}:G=H\times\mathbb{Z}\to H and π:G=H×\pi_{\mathbb{Z}}:G=H\times\mathbb{Z}\to\mathbb{Z} be the natural projections. Let ti=πH(ci)Eit_{i}=\pi_{H}(c_{i})\in E_{i} be the projection to HH of the element cic_{i}. We then choose a quasi-homomorphism ϕi:H\phi_{i}:H\to\mathbb{R} by [HO13] such that ϕi(ti)=1\phi_{i}(t_{i})=1 but ϕi(Ek)=0\phi_{i}(E_{k})=0 if EkEiE_{k}\neq E_{i}. Define the quasi-homomorphism of GG\to\mathbb{R} as follows: for any xG,x\in G,

φ(x):=π(x)i=1nπ(ci)(ϕiπH(x))\varphi(x):=\pi_{\mathbb{Z}}(x)-\sum_{i=1}^{n}\pi_{\mathbb{Z}}(c_{i})\cdot\bigl{(}\phi_{i}\circ\pi_{H}(x)\bigr{)}

By definition, φ\varphi takes the constant value on each ci\langle c_{i}\rangle-cosets. Moreover, the distance |gcigci|G|g\langle c_{i}\rangle-g^{\prime}\langle c_{i}\rangle|_{G} is bi-Lipschitz to |φ(gci)φ(gci)||\varphi(g\langle c_{i}\rangle)-\varphi(g^{\prime}\langle c_{i}\rangle)| with a constant depending only on ci\langle c_{i}\rangle.

To find the generating set SS, notice that the homogenization of φ\varphi (still denoted by φ\varphi) has a bounded distance to the original one. As φ\varphi is unbounded, there exists hGh\in G so that {φ(hn)=nφ(h):n}\{\varphi(h^{n})=n\varphi(h):n\in\mathbb{Z}\} is an infinite cyclic subgroup. Denote S:=φ1([0,2φ(h)])S:=\varphi^{-1}([0,2\varphi(h)]) a (possibly infinite) subset of GG. One can prove that SS generates GG, and φ:G\varphi:G\to\mathbb{R} induces a desired quasi-isometry φ:Cay(G,S)\varphi:\operatorname{Cay}(G,{S})\to\mathbb{R}. See [ABO19, Lemma 4.15] for details. ∎

5.2. New geometric model for vertex spaces

Recall that GG acts on the Bass-Serre tree TT with finitely many vertex orbits. Let {v1,v2,,vn}T\{v_{1},v_{2},\cdots,v_{n}\}\subset T be the full set of vertex representatives, and let SviS_{v_{i}} be the (infinite) generating set for GviG_{v_{i}} given by Lemma 5.1. Then GviG_{v_{i}} acts on the quasi-line 𝔣𝔩(vi):=Cay(Gvi,Svi)\mathfrak{fl}(v_{i}):=\operatorname{Cay}(G_{v_{i}},S_{v_{i}}). Let vv be an arbitrary vertex in TT, so that v=gviv=gv_{i} for some gGg\in G and i{1,2,,n}i\in\{1,2,\cdots,n\}. By equivariance, we define the quasi-line 𝔣𝔩(v):=g𝔣𝔩(vi)=gCay(Gvi,Svi)\mathfrak{fl}(v):=g\mathfrak{fl}(v_{i})=g\operatorname{Cay}(G_{v_{i}},S_{v_{i}}), and the action of Ggvi=gGvig1G_{gv_{i}}=gG_{v_{i}}g^{-1} on g𝔣𝔩(vi)g\mathfrak{fl}(v_{i}) is induced from the action of GviG_{v_{i}} on 𝔣𝔩(vi)\mathfrak{fl}(v_{i}).

Consider the word metric on GG given by a finite generating set of GG including a finite generating set of GviG_{v_{i}} for each representative vertex viv_{i}. Equipping each vertex group GvG_{v} with a word metric, the inclusion of GvG_{v} into GG is a quasi-isometric embedding since YvY_{v} is quasi-isometric embedded in the CAT(0) space XX.

Write Xv:=Y¯v×𝔣𝔩(v)X_{v}:=\overline{Y}_{v}\times\mathfrak{fl}(v) for the new geometric model for GvG_{v}. By Lemma 5.1, the diagonal action GvXvG_{v}\curvearrowright X_{v} is metrically proper and cobounded, and hence the induced orbital map

GvGvoXvG_{v}\longrightarrow G_{v}o^{\prime}\subset X_{v}

is a GvG_{v}-equivariant quasi-isometry for any basepoint o=(o1,o2)Xvo^{\prime}=(o^{\prime}_{1},o_{2}^{\prime})\in X_{v}.

Let us fix a basepoint o=(o1,o2)Yvo=(o_{1},o_{2})\in Y_{v}. As GvG_{v} acts freely and geometrically on Yv=Y¯v×Y_{v}=\overline{Y}_{v}\times\mathbb{R}, let

GvoGvG_{v}o\longrightarrow G_{v}

be a bijective GvG_{v}-equivariant quasi-isometry, a quasi-inverse to the orbital map of GvYvG_{v}\curvearrowright Y_{v}.

Choose the same first coordinate o1=o1o_{1}=o_{1}^{\prime} for the above basepoints o,oo,o^{\prime}. Define a GvG_{v}-equivariant map Λv:YvXv\Lambda_{v}:{Y}_{v}\to X_{v} as the composite of the above two GG-equivariant maps

Λv:Yv=Y¯v×GvXv=Y¯v×𝔣𝔩(v).\Lambda_{v}:\quad Y_{v}=\overline{Y}_{v}\times\mathbb{R}\longrightarrow G_{v}\longrightarrow X_{v}=\overline{Y}_{v}\times\mathfrak{fl}(v).

Define the horizontal and vertical projection maps

(11) Λvhor:YvY¯v,Λvver:Yv𝔣𝔩(v)\Lambda_{v}^{\operatorname{hor}}\colon Y_{v}\to\overline{Y}_{v},\quad\Lambda_{v}^{\operatorname{ver}}:Y_{v}\to\mathfrak{fl}(v)

as the composites of the map Λv\Lambda_{v} with the projections to the factor Y¯v\overline{Y}_{v} and 𝔣𝔩(v)\mathfrak{fl}(v) respectively. For the product space Xv=Y¯v×𝔣𝔩(v)X_{v}=\overline{Y}_{v}\times\mathfrak{fl}(v), we define similarly the horizontal distance and vertical distances ||Xvhor|\cdot|^{\operatorname{hor}}_{X_{v}} and ||Xvver|\cdot|^{\operatorname{ver}}_{X_{v}}. In terms of these notations, we have for any x,yYvx,y\in Y_{v}:

|Λv(x)Λv(y)|Xvhor=|Λvhor(x)Λvhor(y)|Y¯v\bigl{|}\Lambda_{v}(x)-\Lambda_{v}(y)\bigr{|}_{X_{v}}^{\operatorname{hor}}=\bigl{|}\Lambda_{v}^{\operatorname{hor}}(x)-\Lambda_{v}^{\operatorname{hor}}(y)\bigr{|}_{\overline{Y}_{v}}
|Λv(x)Λv(y)|Xvver=|Λvver(x)Λvver(y)|𝔣𝔩(v)\bigl{|}\Lambda_{v}(x)-\Lambda_{v}(y)\bigr{|}_{X_{v}}^{\operatorname{ver}}=\bigl{|}\Lambda_{v}^{\operatorname{ver}}(x)-\Lambda_{v}^{\operatorname{ver}}(y)\bigr{|}_{\mathfrak{fl}(v)}

We now derive a few important facts from Lemma 5.1 about Λv\Lambda_{v}.

Lemma 5.2.

There exists a uniform constant λ>0\lambda>0 such that Λv\Lambda_{v} is a (λ,λ)(\lambda,\lambda)–quasi-isometry: for any x,yYvx,y\in Y_{v} then

1λ|Λv(x)Λv(y)|Xvλ|xy|Yvλ|Λv(x)Λv(y)|Xv+λ\frac{1}{\lambda}\bigl{|}\Lambda_{v}(x)-\Lambda_{v}(y)\bigr{|}_{X_{v}}-\lambda\leq\bigl{|}x-y\bigr{|}_{Y_{v}}\leq\lambda\bigl{|}\Lambda_{v}(x)-\Lambda_{v}(y)\bigr{|}_{X_{v}}+\lambda

Moreover, let YwY_{w} be the adjacent piece of YvY_{v} in the CKA space XX. Let ,\ell,\ell^{\prime} be a line in the plane P=YvYwP=Y_{v}\cap Y_{w} such that ,\ell,\ell^{\prime} are fibers in YwY_{w}. Then

  1. (1)

    diam(Λvver())λ\operatorname{diam}(\Lambda_{v}^{\operatorname{ver}}(\ell))\leq\lambda. In other words, Λv()Y¯vB(a,λ)\Lambda_{v}(\ell)\subset\overline{Y}_{v}\cap B(a,\lambda) in Y¯v×𝔣𝔩(v)\overline{Y}_{v}\times\mathfrak{fl}(v) for some a𝔣𝔩(v)a\in\mathfrak{fl}(v).

  2. (2)

    Let pYv=Y¯v×p\in Y_{v}=\overline{Y}_{v}\times\mathbb{R} be any point and πv(p)\pi_{v}(p) be the projection of pp into the factor Y¯v\overline{Y}_{v}. Then |πv(p)Λvhor(p)|Y¯vλ\bigl{|}\pi_{v}(p)-\Lambda_{v}^{\operatorname{hor}}(p)\bigr{|}_{\overline{Y}_{v}}\leq\lambda.

  3. (3)

    Denote by ||Yv\bigl{|}\ell-\ell^{\prime}\bigr{|}_{Y_{v}} the distance of \ell and \ell^{\prime} in YvY_{v}. Then

    λ1||Yvλdiam𝔣𝔩(v)(Λvver()Λvver())λ||Yv+λ\lambda^{-1}\bigl{|}\ell-\ell^{\prime}\bigr{|}_{Y_{v}}-\lambda\leq\operatorname{diam}_{\mathfrak{fl}(v)}\bigl{(}\Lambda_{v}^{\operatorname{ver}}(\ell)\cup\Lambda_{v}^{\operatorname{ver}}(\ell^{\prime})\bigr{)}\leq\lambda\bigl{|}\ell-\ell^{\prime}\bigr{|}_{Y_{v}}+\lambda
Proof.

We first prove (2). Choose the fixed basepoints o=(o1,o2)o=(o_{1},o_{2}) in YvY_{v} and o=(o1,o2)o^{\prime}=(o_{1}^{\prime},o_{2}^{\prime}) in Y¯v×𝔣𝔩(v)\overline{Y}_{v}\times\mathfrak{fl}(v) so that their projections into the factor Y¯v\overline{Y}_{v} are the same: o1=o1Y¯vo_{1}=o^{\prime}_{1}\in\overline{Y}_{v}. Take any point p=(a,t)p=(a,t) in Yv=Y¯v×Y_{v}=\overline{Y}_{v}\times\mathbb{R}, so πv(p)=a\pi_{v}(p)=a. By our definition of the GvG_{v}–equivariant quasi-inverse YvGvY_{v}\to G_{v}, there exists a group element gGvg\in G_{v} so that |gop|Yvλ|go-p|_{Y_{v}}\leq\lambda for some uniform constant λ\lambda. We write g=(h,n)g=(h,n) in Hv×H_{v}\times\mathbb{Z}. Note that GvG_{v} acts on Y¯v×𝔣𝔩(v)\overline{Y}_{v}\times\mathfrak{fl}(v) diagonally, thus the image of the group element g=(h,n)g=(h,n) under the composition map

GvY¯v×𝔣𝔩(v)Y¯vG_{v}\to\overline{Y}_{v}\times\mathfrak{fl}(v)\to\overline{Y}_{v}

is ho1h\cdot o_{1}, where the first one is the orbital map and the second one is the projection map. If YvY_{v} is equipped with L1L^{1}-metric, it follows that |ho1a|Y¯v|gop|Yvλ|ho_{1}-a|_{\overline{Y}_{v}}\leq|go-p|_{Y_{v}}\leq\lambda. As the map Λv\Lambda_{v} descends to the map Y¯vY¯v\overline{Y}_{v}\to\overline{Y}_{v} sending aa to h(o1)h(o_{1}). Our claim is confirmed.

|Λvhor(x)Λvhor(y)|Y¯vλdh(x,y)λ+|Λvhor(x)Λvhor(y)|Y¯v\bigl{|}\Lambda_{v}^{\operatorname{hor}}(x)-\Lambda_{v}^{\operatorname{hor}}(y)\bigr{|}_{\overline{Y}_{v}}-\lambda\leq d^{h}(x,y)\leq\lambda+\bigl{|}\Lambda_{v}^{\operatorname{hor}}(x)-\Lambda_{v}^{\operatorname{hor}}(y)\bigr{|}_{\overline{Y}_{v}}

For the part (1), as there are only finitely many isometric types of YvY_{v} of XX, we only need to prove that diam(Λvver())λ\operatorname{diam}(\Lambda_{v}^{\operatorname{ver}}(\ell))\leq\lambda for one given YvY_{v}. Indeed, recall that Λvver:Yv𝔣𝔩(v)\Lambda_{v}^{\operatorname{ver}}:Y_{v}\to\mathfrak{fl}(v) factors through YvXv=Y¯v×𝔣𝔩(v)Y_{v}\to X_{v}=\overline{Y}_{v}\times\mathfrak{fl}(v) as the natural projection Xv𝔣𝔩(v)X_{v}\to\mathfrak{fl}(v). The latter agrees with the quasi-homomorphism φ:Gv\varphi:G_{v}\to\mathbb{R} up to a bounded error in the proof of Lemma 5.1, vanishing on the center Z(Gw)Z(G_{w}). If B(a,λ)B(a,\lambda) denotes the ball at some element a𝔣𝔩(v)a\in\mathfrak{fl}(v) with radius λ\lambda, it follows that Z(Gw)oY¯v×B(a,λ)Z(G_{w})o\subset\overline{Y}_{v}\times B(a,\lambda). Every fiber line \ell in YwY_{w} lies in a uniform neighborhood of the orbit of a Z(Gw)Z(G_{w})-coset. Our second claim is thus verified.

The part (3) is clear from our construction. ∎

5.3. Projection maps

Recall T0=𝒱1𝒱2T^{0}=\mathcal{V}_{1}\cup\mathcal{V}_{2} where 𝒱i\mathcal{V}_{i} consists of vertices in TT with pairwise even distances. Denote 𝔽1={𝔣𝔩(v):v𝒱1}\mathbb{F}_{1}=\{\mathfrak{fl}(v):v\in\mathcal{V}_{1}\} and 𝔽2={𝔣𝔩(w):w𝒱2}\mathbb{F}_{2}=\{\mathfrak{fl}(w):w\in\mathcal{V}_{2}\}. It remains to define a family of projection maps for them.

Definition 5.3 (Projection maps in 𝔽i\mathbb{F}_{i}).

Let e1=[v,w]e_{1}=[v,w], e2=[w,v2]e_{2}=[w,v_{2}] denote the first two (oriented) edges in [v,v][v,v^{\prime}]. Let Fe1=YvYwF_{e_{1}}=Y_{v}\cap Y_{w} and Fe2=Yv2YwF_{e_{2}}=Y_{v_{2}}\cap Y_{w} be the two boundary planes of YwY_{w}. Let 𝒮e1e2\mathcal{S}_{e_{1}e_{2}} be the strip in YwY_{w} joining two boundary plane Fe1F_{e_{1}} and Fe2F_{e_{2}} of YwY_{w} (see Section 2.3.1 for the definition of strips). We note that 𝒮e1e2Fe1\mathcal{S}_{e_{1}e_{2}}\cap F_{e_{1}} is a line in Fe1F_{e_{1}} that is parallel to a fiber in YwY_{w}. We then define projection from 𝔣𝔩(v)\mathfrak{fl}(v^{\prime}) into 𝔣𝔩(v)\mathfrak{fl}(v) to be

Π𝔣𝔩(v)(𝔣𝔩(v):=Λvver(𝒮e1e2Fe1)\Pi_{\mathfrak{fl}(v)}(\mathfrak{fl}(v^{\prime}):=\Lambda_{v}^{\operatorname{ver}}\bigl{(}\mathcal{S}_{e_{1}e_{2}}\cap F_{e_{1}}\bigr{)}

where Λvver\Lambda_{v}^{\operatorname{ver}} defined in (11) is the vertical projection to the quasi-line in Xv=Y¯v×𝔣𝔩(v).X_{v}=\overline{Y}_{v}\times\mathfrak{fl}(v).

Lemma 5.4.

Let λ>0\lambda>0 be the constant given by Lemma 5.2. Let a,b,c{a},{b},{c} be distinct vertices in 𝒱i\mathcal{V}_{i} with i=1,2i=1,2. If dT(a,[b,c])2d_{T}(a,[b,c])\geq 2 then Π𝔣𝔩(a)(𝔣𝔩(c))=Π𝔣𝔩(a)(𝔣𝔩(b))λ\Pi_{\mathfrak{fl}(a)}(\mathfrak{fl}(c))=\Pi_{\mathfrak{fl}(a)}(\mathfrak{fl}(b))\leq\lambda

Proof.

Let [b,a][b,a] and [c,a][c,a] be the geodesics in the tree TT connecting bb and cc to ww respectively. Let us denote eee\cdot e^{\prime} be the last two edges in [b,a][b,a] (that is also the last two edges in [c,a][c,a]). Let 𝒮ee\mathcal{S}_{ee^{\prime}} be the strip in Ye+Y_{e_{+}} connecting two boundary planes FeF_{e} and FeF_{e^{\prime}} of Ye+Y_{e_{+}} By our definition of projection maps, we have that Π𝔣𝔩(a)(𝔣𝔩(c))=Π𝔣𝔩(a)(𝔣𝔩(b))=Λaver(𝒮eePe)λ\Pi_{\mathfrak{fl}(a)}(\mathfrak{fl}(c))=\Pi_{\mathfrak{fl}(a)}(\mathfrak{fl}(b))=\Lambda_{a}^{\operatorname{ver}}(\mathcal{S}_{ee^{\prime}}\cap P_{e^{\prime}})\leq\lambda. ∎

5.4. Projection axioms

We are now going to verify that 𝔽i\mathbb{F}_{i} (i=1,2i=1,2) with the above-defined projection maps in Definition 5.3 satisfy the projection axioms (see Definition 2.24). For each vertex vTv\in T, let 𝕃v\mathbb{L}_{v} be the collection of boundary lines in the hyperbolic space Y¯v\overline{Y}_{v} defined at the beginning of Section 4. Let 1,2\ell_{1},\ell_{2}, and 3\ell_{3} be three distinct boundary lines in 𝕃v\mathbb{L}_{v}. We denote

d1(2,3)=diam(π1(2)π1(3))d_{\ell_{1}}(\ell_{2},\ell_{3})=\operatorname{diam}\bigl{(}\pi_{\ell_{1}}(\ell_{2})\cup\pi_{\ell_{1}}(\ell_{3})\bigr{)}

where πi(j)\pi_{\ell_{i}}(\ell_{j}) is the shortest projection of j\ell_{j} to i\ell_{i} in the CAT(0) hyperbolic space Y¯v\overline{Y}_{v} (note that Y¯v\overline{Y}_{v} is a hyperbolic space since HvH_{v} acts geometrically on Y¯v\overline{Y}_{v} and HvH_{v} is a non-elementary hyperbolic group). Recall that

d𝔣𝔩(v1)(𝔣𝔩(v2),𝔣𝔩(v3)):=diam(Π𝔣𝔩(v1)(𝔣𝔩(v2))Π𝔣𝔩(v1)(𝔣𝔩(v3))).d_{\mathfrak{fl}(v_{1})}(\mathfrak{fl}(v_{2}),\mathfrak{fl}(v_{3})):=\operatorname{diam}\bigl{(}\Pi_{\mathfrak{fl}(v_{1})}(\mathfrak{fl}(v_{2}))\cup\Pi_{\mathfrak{fl}(v_{1})}(\mathfrak{fl}(v_{3}))\bigr{)}.
Lemma 5.5.

There exists a uniform constant λ>0\lambda>0 such that the following holds. Let v1,v2,v3v_{1},v_{2},v_{3} be distinct vertices in 𝒱1\mathcal{V}_{1} such that v1,v2,v3v_{1},v_{2},v_{3} are in Lk(o)\operatorname{Lk}(o) for some vertex oo in 𝒱2\mathcal{V}_{2}. Let eie_{i} denote the edge [vi,o][v_{i},o] with i=1,2,3i=1,2,3 and let FeiF_{e_{i}} be the plane in XX associated to eie_{i}. For each i=1,2,3i=1,2,3, let i\ell_{i} denote the boundary line of Y¯o\overline{Y}_{o} that is the projection of FeiF_{e_{i}} into Y¯o\overline{Y}_{o}. Then

1λd1(2,3)λd𝔣𝔩(v1)(𝔣𝔩(v2),𝔣𝔩(v3))λd1(2,3)+λ\frac{1}{\lambda}d_{\ell_{1}}(\ell_{2},\ell_{3})-\lambda\leq d_{\mathfrak{fl}(v_{1})}(\mathfrak{fl}(v_{2}),\mathfrak{fl}(v_{3}))\leq\lambda d_{\ell_{1}}(\ell_{2},\ell_{3})+\lambda
Proof.

Let 𝒮e1e2\mathcal{S}_{e_{1}e_{2}} and 𝒮e1e3\mathcal{S}_{e_{1}e_{3}} be the strips in YoY_{o} connecting the planes Fe1F_{e_{1}} to Fe2F_{e_{2}} and Fe1F_{e_{1}} to Fe3F_{e_{3}} respectively. We denote the line 𝒮e1e2Fe1\mathcal{S}_{e_{1}e_{2}}\cap F_{e_{1}} by \ell and denote the line 𝒮e1e3Fe1\mathcal{S}_{e_{1}e_{3}}\cap F_{e_{1}} by \ell^{\prime}. Note that both lines \ell and \ell^{\prime} are fibers in YoY_{o}. Recall that by our definition of projection maps, we have Π𝔣𝔩(v1)(𝔣𝔩(v2))=Λv1ver()\Pi_{\mathfrak{fl}(v_{1})}(\mathfrak{fl}(v_{2}))=\Lambda^{\operatorname{ver}}_{v_{1}}(\ell) and Π𝔣𝔩(v1)(𝔣𝔩(v3))=Λv1ver()\Pi_{\mathfrak{fl}(v_{1})}(\mathfrak{fl}(v_{3}))=\Lambda^{\operatorname{ver}}_{v_{1}}(\ell^{\prime}). By part (3) of Lemma 5.2, for some λ>0\lambda>0, we have that 1λ||λdiam(Π𝔣𝔩(v1)(𝔣𝔩(v2))Π𝔣𝔩(v1)(𝔣𝔩(v3)))λ||+λ\frac{1}{\lambda}\bigl{|}\ell-\ell^{\prime}\bigr{|}-\lambda\leq\operatorname{diam}\bigl{(}\Pi_{\mathfrak{fl}(v_{1})}(\mathfrak{fl}(v_{2}))\cup\Pi_{\mathfrak{fl}(v_{1})}(\mathfrak{fl}(v_{3}))\bigr{)}\leq\lambda\bigl{|}\ell-\ell^{\prime}\bigr{|}+\lambda. Note that ||=d1(2,3)\bigl{|}\ell-\ell^{\prime}\bigr{|}=d_{\ell_{1}}(\ell_{2},\ell_{3}) (indeed, let α\alpha and β\beta be the shortest geodesics joining 2\ell_{2} to 1\ell_{1} and 3\ell_{3} to 1\ell_{1} respectively, then \ell and \ell^{\prime} are the product α+×\alpha_{+}\times\mathbb{R} and β+×\beta_{+}\times\mathbb{R} of endpoints of α\alpha and β\beta with the \mathbb{R} direction in Yo=Y¯o×Y_{o}=\overline{Y}_{o}\times\mathbb{R} respectively). Combining the above inequalities, we obtain a constant λ=λ(λ)>0\lambda^{\prime}=\lambda^{\prime}(\lambda)>0 still denoted by λ\lambda such that 1λd1(2,3)λdiam(Π𝔣𝔩(v1)(𝔣𝔩(v2))Π𝔣𝔩(v1)(𝔣𝔩(v3)))λd1(2,3)+λ\frac{1}{\lambda}d_{\ell_{1}}(\ell_{2},\ell_{3})-\lambda\leq\operatorname{diam}\bigl{(}\Pi_{\mathfrak{fl}(v_{1})}(\mathfrak{fl}(v_{2}))\cup\Pi_{\mathfrak{fl}(v_{1})}(\mathfrak{fl}(v_{3}))\bigr{)}\leq\lambda d_{\ell_{1}}(\ell_{2},\ell_{3})+\lambda. The lemma is proved. ∎

We are now going to prove the following.

Lemma 5.6.

There exists a constant ξ>0\xi>0 such that for each i{1,2}i\in\{1,2\}, the collection 𝔽i\mathbb{F}_{i} with projection maps π𝔣𝔩(v)\pi_{\mathfrak{fl}(v)}’s satisfies the projection axioms with projection constant ξ\xi.

Proof.

We verify in order the projection axioms (see Definition 2.24) for the projection maps defined on 𝔽1\mathbb{F}_{1}. The case for 𝔽2\mathbb{F}_{2} is symmetric. The constant ξ\xi will be defined explicitly during the proof.

Axiom 1: Let λ>0\lambda>0 be the constant given by Lemma 5.2. Since 𝒮e1e2Fe1\mathcal{S}_{e_{1}e_{2}}\cap F_{e_{1}} is a fiber line in YwY_{w}, it follows from Lemma 5.2 that diamΛvver(𝒮e1e2Fe1)λ\operatorname{diam}\Lambda^{\operatorname{ver}}_{v}\bigl{(}\mathcal{S}_{e_{1}e_{2}}\cap F_{e_{1}}\bigr{)}\leq\lambda. Thus diam(Π𝔣𝔩(v)(𝔣𝔩(v)))λ\operatorname{diam}\bigl{(}\Pi_{\mathfrak{fl}(v)}(\mathfrak{fl}(v^{\prime}))\bigr{)}\leq\lambda. Axiom 1 in Definition 2.24 is verified.

Axiom 2: Let u,v,w{u},{v},{w} be distinct vertices in 𝒱1\mathcal{V}_{1}. We will show that there exists ξ0\xi\geq 0 sufficiently large satisfying the following property: if d𝔣𝔩(w)(𝔣𝔩(u),𝔣𝔩(v))>ξd_{\mathfrak{fl}(w)}(\mathfrak{fl}(u),\mathfrak{fl}(v))>\xi, then d𝔣𝔩(u)(𝔣𝔩(w),𝔣𝔩(v))ξd_{\mathfrak{fl}(u)}(\mathfrak{fl}(w),\mathfrak{fl}(v))\leq\xi or d𝔣𝔩(v)(𝔣𝔩(w),𝔣𝔩(u))ξd_{\mathfrak{fl}(v)}(\mathfrak{fl}(w),\mathfrak{fl}(u))\leq\xi. The constant ξ\xi will be defined explicitly during the proof. Since d𝔣𝔩(w)(𝔣𝔩(u),𝔣𝔩(v))>ξd_{\mathfrak{fl}(w)}(\mathfrak{fl}(u),\mathfrak{fl}(v))>\xi, it follows from Lemma 5.4 that there is some restriction on ww, i.e, ww is either lies on [u,v][u,v] or dT(w,[u,v])=1d_{T}(w,[u,v])=1.

Case 1: ww lies on [u,v][u,v]. Since u,w,v𝒱1u,w,v\in\mathcal{V}_{1}, we have dT(u,[w,v])2d_{T}(u,[w,v])\geq 2 and dT(v,[u,w])2d_{T}(v,[u,w])\geq 2. Axiom 2 thus follows from Lemma 5.4.

Case 2: dT(w,[u,v])=1d_{T}(w,[u,v])=1. Without loss of generality, we can assume that u,v,wu,v,w lie in the same link Lk(o)\operatorname{Lk}(o) for some vertex oo in 𝒱2\mathcal{V}_{2}. Indeed, let o[u,v]o\in[u,v] be adjacent to ww and u,vLk(o)[u,v]u^{\prime},v^{\prime}\in Lk(o)\cap[u,v]. It is clear by definition that π𝔣𝔩(u)(𝔣𝔩(v))=π𝔣𝔩(u)(𝔣𝔩(v))\pi_{\mathfrak{fl}(u)}(\mathfrak{fl}(v^{\prime}))=\pi_{\mathfrak{fl}(u)}(\mathfrak{fl}(v)) and π𝔣𝔩(v)(𝔣𝔩(u))=π𝔣𝔩(v)(𝔣𝔩(u))\pi_{\mathfrak{fl}(v)}(\mathfrak{fl}(u^{\prime}))=\pi_{\mathfrak{fl}(v)}(\mathfrak{fl}(u)). As a result, we can thus assume that u=uu=u^{\prime} and v=vv=v^{\prime} lie in the link Lk(o)\operatorname{Lk}(o).

Recall that Y¯o\overline{Y}_{o} is a δ\delta-hyperbolic space whose boundary lines 𝕃o\mathbb{L}_{o} satisfy the projection axioms for a constant ξ0\xi_{0} [Sis13]. We claim that ξ=ξ0\xi=\xi_{0} is the desired constant for Axiom 2.

Refer to caption
Figure 2. Verification of Axiom 2

Denote e=[w,o]e=[w,o], e1=[u,o],e2=[v,o]e_{1}=[u,o],e_{2}=[v,o]. Let e,e1,e2\ell_{e},\ell_{e_{1}},\ell_{e_{2}} be the corresponding boundary lines of Y¯o\overline{Y}_{o} to the oriented edges e,e1,e2e,e_{1},e_{2}. By Lemma 5.5, we have

1λde(e1,e2)λd𝔣𝔩(w)(𝔣𝔩(u),𝔣𝔩(v))λde(e1,e2)+λ\frac{1}{\lambda}d_{\ell_{e}}(\ell_{e_{1}},\ell_{e_{2}})-\lambda\leq d_{\mathfrak{fl}(w)}(\mathfrak{fl}(u),\mathfrak{fl}(v))\leq\lambda d_{\ell_{e}}(\ell_{e_{1}},\ell_{e_{2}})+\lambda

As 𝕃o\mathbb{L}_{o} satisfies the projection axioms, we see that if de(e1,e2)>ξ0d_{\ell_{e}}(\ell_{e_{1}},\ell_{e_{2}})>\xi_{0}, then de1(e,e2)ξ0d_{\ell_{e_{1}}}(\ell_{e},\ell_{e_{2}})\leq\xi_{0}. Using Lemma 5.5 again, we have that

1λde1(e,e2)λd𝔣𝔩(u)(𝔣𝔩(w),𝔣𝔩(v))λde1(e,e2)+λ\frac{1}{\lambda}d_{\ell_{e_{1}}}(\ell_{e},\ell_{e_{2}})-\lambda\leq d_{\mathfrak{fl}(u)}(\mathfrak{fl}(w),\mathfrak{fl}(v))\leq\lambda d_{\ell_{e_{1}}}(\ell_{e},\ell_{e_{2}})+\lambda

Let ξ\xi be a constant such that ξ>λξ0+λ\xi>\lambda\xi_{0}+\lambda. It follows from the above inequalities that

d𝔣𝔩(u)(𝔣𝔩(w),𝔣𝔩(v))=diam(Π𝔣𝔩(u)(𝔣𝔩(w))Π𝔣𝔩(u)(𝔣𝔩(v)))ξd_{\mathfrak{fl}(u)}(\mathfrak{fl}(w),\mathfrak{fl}(v))=\operatorname{diam}\bigl{(}\Pi_{\mathfrak{fl}(u)}(\mathfrak{fl}(w))\cup\Pi_{\mathfrak{fl}(u)}(\mathfrak{fl}(v))\bigr{)}\leq\xi

so Axiom 2 is verified.

Axiom 3: For uv𝒱1u\neq v\in\mathcal{V}_{1}, the set

{w𝒱1:d𝔣𝔩(w)(𝔣𝔩(u),𝔣𝔩(v))>ξ}\{w\in\mathcal{V}_{1}:d_{\mathfrak{fl}(w)}(\mathfrak{fl}({u}),\mathfrak{fl}({v}))>\xi\}

is a finite set.

Indeed, by Lemma 5.4, such ww is either contained in the interior of [u,v][u,v] or d(w,[u,v])=1d(w,[u,v])=1. The first case yields only (d(u,v)1)(d(u,v)-1) choices for ww. We now consider the case d(w,[u,v])=1d(w,[u,v])=1. Since u,v,wu,v,w have pairwise even distance, there exists o𝒲[u,v]0o\in\mathcal{W}\cap[u,v]^{0} and two vertices u,vu^{\prime},v^{\prime} on [u,v][u,v] adjacent to oo so that u,v,wLk(o)u^{\prime},v^{\prime},w\in Lk(o). By the projection axioms of boundary lines 𝕃o\mathbb{L}_{o} of Y¯o\overline{Y}_{o}, the set of ww satisfying d𝔣𝔩(w)(𝔣𝔩(u),𝔣𝔩(v))>ξd_{\mathfrak{fl}(w)}(\mathfrak{fl}({u}),\mathfrak{fl}({v}))>\xi is finite. Thus, in both cases, the set of such ww’s is finite. Axiom 3 is verified. ∎

Lemma 5.7.

For each i=1,2i=1,2, the collection 𝔽i={𝔣𝔩(v:v𝒱i}\mathbb{F}_{i}=\{\mathfrak{fl}({v}:v\in\mathcal{V}_{i}\} admits an action of the group GG so that

Πg𝔣𝔩(v)(g𝔣𝔩(u))=gΠ𝔣𝔩(v)(𝔣𝔩(u))\Pi_{g\mathfrak{fl}(v)}(g\mathfrak{fl}({u}))=g\Pi_{\mathfrak{fl}({v})}(\mathfrak{fl}({u}))

for any v,u𝒱iv,u\in\mathcal{V}_{i} and any gGg\in G.

Proof.

Firstly, let us recall some discussion in the beginning of Section 5.2. Recall that {v1,v2,,vn}T\{v_{1},v_{2},\cdots,v_{n}\}\subset T be the full set of vertex representatives of TT and for each representative vertices v1,v2,vnv_{1},v_{2},\cdots v_{n} of TT, the quasi-line 𝔣𝔩(vj)\mathfrak{fl}(v_{j}) is the Cayley graph Cay(Gvj,Svj)\operatorname{Cay}(G_{v_{j}},S_{v_{j}}) for some generating set SvjS_{v_{j}} of GvjG_{v_{j}} (see Lemma 5.1). Let vv be an arbitrary vertex in TT, then v=gviv=gv_{i} for some gGg\in G and i{1,2,,n}i\in\{1,2,\cdots,n\}. The quasi-line 𝔣𝔩(v)\mathfrak{fl}(v) is given by g𝔣𝔩(vi)=gCay(Gvi,Svi)g\mathfrak{fl}(v_{i})=g\operatorname{Cay}(G_{v_{i}},S_{v_{i}}), and the action of Ggvi=gGvig1G_{gv_{i}}=gG_{v_{i}}g^{-1} on g𝔣𝔩(vi)g\mathfrak{fl}(v_{i}) is induced from the action of GviG_{v_{i}} on 𝔣𝔩(vi)\mathfrak{fl}(v_{i}). We are now going to show that

Πg𝔣𝔩(v)(g𝔣𝔩(u))=gΠ𝔣𝔩(v)(𝔣𝔩(u))\Pi_{g\mathfrak{fl}(v)}(g\mathfrak{fl}(u))=g\Pi_{\mathfrak{fl}(v)}(\mathfrak{fl}(u))

Recall that the family of maps Λgvver:Ygv=gYvg𝔣𝔩(v)\Lambda_{gv}^{\operatorname{ver}}\colon Y_{gv}=gY_{v}\to g\mathfrak{fl}(v) are GG–equivariant: Λgvver(gx)=gΛvver(x)\Lambda_{gv}^{\operatorname{ver}}(gx)=g\Lambda_{v}^{\operatorname{ver}}(x) for all xYvx\in Y_{v}. Let e1e_{1} and e2e_{2} be the first two edges in the geodesic [v,u][v,u] with v=(e1)v=(e_{1})_{-} and (e1)+=(e2)(e_{1})_{+}=(e_{2})_{-}. By Definition 5.3 of projection map, we have that Π𝔣𝔩(gv)(𝔣𝔩(gu))=diam(Λgvver(𝒮ge1ge2Fge1)=diam(Λgvver(g(𝒮e1e2Fe1)))=diam(gΛvver(𝒮e1e2Fe1))=gdiam(Λvver(𝒮e1e2Fe1))=gΠ𝔣𝔩(v)(𝔣𝔩(u))\Pi_{\mathfrak{fl}(gv)}(\mathfrak{fl}(gu))=\operatorname{diam}\bigl{(}\Lambda_{gv}^{\operatorname{ver}}(\mathcal{S}_{ge_{1}ge_{2}}\cap F_{ge_{1}}\bigr{)}=\operatorname{diam}\bigr{(}\Lambda_{gv}^{\operatorname{ver}}(g(\mathcal{S}_{e_{1}e_{2}}\cap F_{e_{1}}))\bigr{)}=\operatorname{diam}\bigl{(}g\Lambda_{v}^{\operatorname{ver}}(\mathcal{S}_{e_{1}e_{2}}\cap F_{e_{1}})\bigr{)}\quad\text{}=g\operatorname{diam}\bigl{(}\Lambda_{v}^{\operatorname{ver}}(\mathcal{S}_{e_{1}e_{2}}\cap F_{e_{1}})\bigr{)}=g\Pi_{\mathfrak{fl}(v)}(\mathfrak{fl}(u)) for any gGg\in G. ∎

Definition 5.8.

Let ξ>0\xi>0 be the projection constant given by Lemma 5.6, so the collection of quasi-lines 𝔽i={𝔣𝔩(v):v𝒱i}\mathbb{F}_{i}=\{\mathfrak{fl}(v):v\in\mathcal{V}_{i}\} with i=1,2i=1,2 satisfies the projection axioms. For any fixed K>4ξK>4\xi, we obtain the unbounded quasi-trees of metric spaces 𝒞K(𝔽1)\mathcal{C}_{K}(\mathbb{F}_{1}) and 𝒞K(𝔽2)\mathcal{C}_{K}(\mathbb{F}_{2}) (see Section 2.4). Combining Lemma 5.7 with [BBF15, Theorem 4.4], the spaces 𝒞K(𝔽1)\mathcal{C}_{K}(\mathbb{F}_{1}) and 𝒞K(𝔽2)\mathcal{C}_{K}(\mathbb{F}_{2}) are quasi-trees and admit unbounded isometric actions G𝒞K(𝔽1)G\curvearrowright\mathcal{C}_{K}(\mathbb{F}_{1}) and G𝒞K(𝔽2)G\curvearrowright\mathcal{C}_{K}(\mathbb{F}_{2}). The quasi-trees 𝒞K(𝔽1)\mathcal{C}_{K}(\mathbb{F}_{1}) and 𝒞K(𝔽2)\mathcal{C}_{K}(\mathbb{F}_{2}) are called vertical quasi-trees hereafter.

6. Distance formulas in the CKA space XX

Let 𝒞K(𝔽1)\mathcal{C}_{K}(\mathbb{F}_{1}) and 𝒞K(𝔽2)\mathcal{C}_{K}(\mathbb{F}_{2}) be the vertical quasi-trees in Definition  5.8. Let 𝒳1˙,𝒳2˙\dot{\mathcal{X}_{1}},\dot{\mathcal{X}_{2}} be the coned-off spaces defined in Section 4.1. According to the outline of the proof of Theorem 1.3 in Section 4.4, the last step to prove property (QT) of GG is to show that there is a GG-equivariant quasi-isometric embedding

Φ:Go𝒞K(𝔽1)×𝒞K(𝔽2)×𝒳1˙×𝒳2˙×T\Phi\colon Go\to\mathcal{C}_{K}(\mathbb{F}_{1})\times\mathcal{C}_{K}(\mathbb{F}_{2})\times\dot{\mathcal{X}_{1}}\times\dot{\mathcal{X}_{2}}\times T

This section is devoted to constructing such a desired map Φ\Phi and verifying it is a quasi-isometric embedding.

We list here notations that will be used in the rest of this section.

  • We fix an edge [v0,w0][v_{0},w_{0}] in the Bass-Serre tree TT such that v0𝒱1v_{0}\in\mathcal{V}_{1}. Let oXo\in X be a base point in the common boundary plane F[v0,w0]F_{[v_{0},w_{0}]} between two pieces Yv0Y_{v_{0}} and Yw0Y_{w_{0}}.

  • Assume that x=oYv0x=o\in Y_{v_{0}} and y=goYv2ny=go\in Y_{v_{2n}} for some gGg\in G and v2n=gov_{2n}=go. We list the vertices on the geodesic [v0,v2n][v_{0},v_{2n}] by {v0,v1,,v2n}\{v_{0},v_{1},\ldots,v_{2n}\} where v2i𝒱1v_{2i}\in\mathcal{V}_{1} and v2i+1𝒱2v_{2i+1}\in\mathcal{V}_{2}. Denote ei+1=[vi,vi+1]e_{i+1}=[v_{i},v_{i+1}] the oriented edge towards vi+1v_{i+1}. By definition of special paths, let pi:=𝒮ei1ei𝒮eiei+1p_{i}:=\mathcal{S}_{e_{i-1}e_{i}}\cap\mathcal{S}_{e_{i}e_{i+1}} be the intersection of two strips with p0:x=op_{0}:x=o and p2n+1=y=gop_{2n+1}=y=go.

  • Set

    α~:=e0αe2n+1\widetilde{\alpha}:=e_{0}\cup\alpha\cup e_{2n+1}

    where e0=[w0,v0]e_{0}=[w_{0},v_{0}] and e2n+1=[v2n,w1]e_{2n+1}=[v_{2n},w_{1}]. It is possible that e0=e¯1e_{0}=\overline{e}_{1} and e2n+1=e¯2ne_{2n+1}=\overline{e}_{2n}, i.e, α~\widetilde{\alpha} contains backtracking at e0e_{0} and e2ne_{2n}.

6.1. Construction of the desired map Φ\Phi

It is a product of the following four maps with the index map ρ\rho in Definition 2.19.

  • We define ϑ1:Go𝒞K(𝔽1)\vartheta_{1}\colon Go\to\mathcal{C}_{K}(\mathbb{F}_{1}) as follows. Recall that each quasi-line 𝔣𝔩(v)\mathfrak{fl}(v) for v𝒱1v\in\mathcal{V}_{1} embeds as a convex subset into 𝒞K(𝕃1)\mathcal{C}_{K}(\mathbb{L}_{1}) and Λvver:Gvo𝔣𝔩(v)\Lambda_{v}^{\operatorname{ver}}:G_{v}o\to\mathfrak{fl}(v) is a GvG_{v}-equivariant map. For every gGg\in G, we set ϑ1(go):=Λgv0ver(go)=gΛv0ver(o)\vartheta_{1}(go):=\Lambda_{gv_{0}}^{\operatorname{ver}}(go)=g\Lambda_{v_{0}}^{\operatorname{ver}}(o). The second equality follows by GvG_{v}-equivariance.

  • Similarly, define ϑ2:Go𝒞K(𝔽2)\vartheta_{2}\colon Go\to\mathcal{C}_{K}(\mathbb{F}_{2}) by ϑ2(go):=Λgw0ver(go)=gΛw0ver(o)\vartheta_{2}(go):=\Lambda_{gw_{0}}^{\operatorname{ver}}(go)=g\Lambda_{w_{0}}^{\operatorname{ver}}(o) for every gGg\in G.

  • Define ϑ3(o):=πY¯v0(o)\vartheta_{3}(o):=\pi_{\overline{Y}_{v_{0}}}(o) and extend the definition by equivariance so that ϑ3(go):=gϑ3(o)\vartheta_{3}(go):=g\vartheta_{3}(o) for any gGg\in G. We thus obtain a GG–equivariant map ϑ3:Go𝒳1˙\vartheta_{3}\colon Go\to\dot{\mathcal{X}_{1}}.

  • Choose ϑ4(o)\vartheta_{4}(o) to be the cone point of the hyperbolic cone attached to the boundary line [v0,w0]\ell_{[v_{0},w_{0}]} of Y¯w0\overline{Y}_{w_{0}}. We then extend ϑ4(go)=gϑ4(o)\vartheta_{4}(go)=g\vartheta_{4}(o) for any gGg\in G so that gϑ4(o)g\vartheta_{4}(o) is the corresponding cone point to [gv0,gw0]\ell_{[gv_{0},gw_{0}]} of Y¯gw0\overline{Y}_{gw_{0}}. We thus obtain a GG–equivariant map ϑ4:Go𝒳2˙\vartheta_{4}\colon Go\to\dot{\mathcal{X}_{2}}.

We then define

(\clubsuit) Φ:Go𝒞K(𝔽1)×𝒞K(𝔽2)×(𝒳1˙,d𝒳1˙K)×(𝒳2˙,d𝒳2˙K)×T\Phi\colon Go\to\mathcal{C}_{K}(\mathbb{F}_{1})\times\mathcal{C}_{K}(\mathbb{F}_{2})\times(\dot{\mathcal{X}_{1}},d^{K}_{\dot{\mathcal{X}_{1}}})\times(\dot{\mathcal{X}_{2}},d^{K}_{\dot{\mathcal{X}_{2}}})\times T

by

Φ:=ϑ1×ϑ2×ϑ3×ϑ4×ρ\Phi:=\vartheta_{1}\times\vartheta_{2}\times\vartheta_{3}\times\vartheta_{4}\times\rho

where 𝒳i˙\dot{\mathcal{X}_{i}} for i=1,2i=1,2 are equipped with the KK-thick distance d𝒳i˙Kd^{K}_{\dot{\mathcal{X}_{i}}} (not genuine distance) defined in (9), and the other three spaces are with length metric. By abuse of language, we call the sum of the distances over the factors as L1L^{1}-metric on the product space.

The remainder of this section is to verify the following.

Proposition 6.1.

The map Φ\Phi in (\clubsuit6.1) is a GG–equivariant quasi-isometric embedding.

Idea of the proof of Proposition 6.1:

Since the orbital map of any isometric action is Lipschitz (e.g. see [BH99, Lemma I.8.18]), we will only need to give a linear upper bound on |xy|X|x-y|_{X}. Recall from (2), for any x,yXx,y\in X,

|xy|X|xy|Xhor+|xy|Xver|x-y|_{X}\sim\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}+\bigl{|}x-y\bigr{|}_{X}^{\operatorname{ver}}

where |xy|Xhor=i=02n|pipi+1|Yvihor\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}=\sum_{i=0}^{2n}\bigl{|}p_{i}-p_{i+1}\bigr{|}_{Y_{v_{i}}}^{\operatorname{hor}} and |xy|Xver=i=02n|pipi+1|Yviver\bigl{|}x-y\bigr{|}_{X}^{\operatorname{ver}}=\sum_{i=0}^{2n}\bigl{|}p_{i}-p_{i+1}\bigr{|}_{Y_{v_{i}}}^{\operatorname{ver}}.

Recall from Section 5.2, we build a new geometric model XvX_{v} of YvY_{v} for each vertex vv in the Bass-Serre tree TT. Namely, we have a GvG_{v}–equivariant quasi-isometric map Λv:Yv=Y¯v×Xv=Y¯v×𝔣𝔩(v)\Lambda_{v}\colon Y_{v}=\overline{Y}_{v}\times\mathbb{R}\to X_{v}=\overline{Y}_{v}\times\mathfrak{fl}(v). For x,yGox,y\in Go, we shall accordingly replace |xy|Xver\bigl{|}x-y\bigr{|}_{X}^{\operatorname{ver}} by the following quantity

(12) V(x,y):=0i2n|Λviver(pi)Λviver(pi+1)|𝔣𝔩(vi)\displaystyle V(x,y):=\sum_{0\leq i\leq 2n}\bigl{|}\Lambda_{v_{i}}^{\operatorname{ver}}(p_{i})-\Lambda_{v_{i}}^{\operatorname{ver}}(p_{i+1})\bigr{|}_{\mathfrak{fl}(v_{i})}

To be precise, we first prove in Lemma 6.2 the following

|xy|Xϵ(|ρ(x)ρ(y)|T+|xy|Xhor+V(x,y))|x-y|_{X}\leq\epsilon\left(|\rho(x)-\rho(y)|_{T}+\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}+V(x,y)\right)

We then find suitable upper bounds of V(x,y)V(x,y) (see Proposition 6.5) and |xy|Xhor\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}} (see Lemma 6.6).

6.2. Verifying Φ\Phi is a quasi-isometric embedding

In this section, we will verify that the map Φ\Phi in (\clubsuit6.1) is a quasi-isometric embedding.

6.2.1. Upper bound of the distance |xy|X|x-y|_{X} on XX

Lemma 6.2.

Let x,yGox,y\in Go. The exists a constant ϵ>0\epsilon>0 such that

(13) |xy|Xϵ(|ρ(x)ρ(y)|T+|xy|Xhor+V(x,y))|x-y|_{X}\leq\epsilon\left(|\rho(x)-\rho(y)|_{T}+\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}+V(x,y)\right)
Proof.

Recall that p0=xp_{0}=x and p2n+1=yp_{2n+1}=y. Using the triangle inequality we have

|xy|Xi=02n|pipi+1|Yvi|x-y|_{X}\leq\sum_{i=0}^{2n}|p_{i}-p_{i+1}|_{Y_{v_{i}}}

Note that 2n=|ρ(x)ρ(y)|T2n=|\rho(x)-\rho(y)|_{T} and |xy|Xhor=i=02n|pipi+1|Yvihor\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}=\sum_{i=0}^{2n}\bigl{|}p_{i}-p_{i+1}\bigr{|}_{Y_{v_{i}}}^{\operatorname{hor}}. The proof is then completed by summing over 0i2n0\leq i\leq 2n the following inequality (14).

Claim.

There exists a uniform constant ϵ>0\epsilon^{\prime}>0 such that for any i{0,1,,2n}i\in\{0,1,\cdots,2n\},

(14) |pipi+1|Yviϵ+ϵ|pipi+1|Yvihor+ϵ|Λviver(pi)Λviver(pi+1)|𝔣𝔩(vi).|p_{i}-p_{i+1}|_{Y_{v_{i}}}\leq\epsilon^{\prime}+\epsilon^{\prime}\,\bigl{|}p_{i}-p_{i+1}\bigr{|}_{Y_{v_{i}}}^{\operatorname{hor}}+\epsilon^{\prime}\,\bigl{|}\Lambda^{\operatorname{ver}}_{v_{i}}(p_{i})-\Lambda^{\operatorname{ver}}_{v_{i}}(p_{i+1})\bigr{|}_{\mathfrak{fl}(v_{i})}.
Proof of the Claim.

Indeed, since Λvi:YviXvi\Lambda_{v_{i}}:Y_{v_{i}}\to X_{v_{i}} is a quasi-isometry by Lemma 5.2, we then have

|pipi+1|Yviλ|Λvi(pi)Λvi(pi+1)|Xvi\displaystyle|p_{i}-p_{i+1}|_{Y_{v_{i}}}\sim_{\lambda}\bigl{|}\Lambda_{v_{i}}(p_{i})-\Lambda_{v_{i}}(p_{i+1})\bigr{|}_{X_{v_{i}}}

Using part (2) of Lemma 5.2 we have that

|Λvihor(pi)Λvihor(pi+1)|Y¯viλ|pipi+1|Yvihor\bigl{|}\Lambda^{\operatorname{hor}}_{v_{i}}(p_{i})-\Lambda^{\operatorname{hor}}_{v_{i}}(p_{i+1})\bigl{|}_{\overline{Y}_{v_{i}}}\sim_{\lambda}\bigl{|}p_{i}-p_{i+1}\bigr{|}_{Y_{v_{i}}}^{\operatorname{hor}}

It implies that

|Λvi(pi))Λvi(pi+1)|Xvi\displaystyle\quad\quad\bigl{|}\Lambda_{v_{i}}(p_{i}))-\Lambda_{v_{i}}(p_{i+1})\bigr{|}_{X_{v_{i}}}
2|Λvihor(pi)Λvihor(pi+1)|Y¯vi+|Λviver(pi)Λviver(pi+1)|𝔣𝔩(vi)\displaystyle\sim_{\sqrt{2}}\;\bigl{|}\Lambda^{\operatorname{hor}}_{v_{i}}(p_{i})-\Lambda^{\operatorname{hor}}_{v_{i}}(p_{i+1})\bigl{|}_{\overline{Y}_{v_{i}}}+\bigl{|}\Lambda^{\operatorname{ver}}_{v_{i}}(p_{i})-\Lambda^{\operatorname{ver}}_{v_{i}}(p_{i+1})\bigr{|}_{\mathfrak{fl}(v_{i})}
λ|pipi+1|Yvihor+|Λviver(pi)Λviver(pi+1)|𝔣𝔩(vi)\displaystyle\sim_{\lambda}\;\;\bigl{|}p_{i}-p_{i+1}\bigr{|}_{Y_{v_{i}}}^{\operatorname{hor}}+\bigl{|}\Lambda^{\operatorname{ver}}_{v_{i}}(p_{i})-\Lambda^{\operatorname{ver}}_{v_{i}}(p_{i+1})\bigr{|}_{\mathfrak{fl}(v_{i})}

where the first coarse equality holds by definition of Λvihor\Lambda_{v_{i}}^{\operatorname{hor}} and Λviver\Lambda_{v_{i}}^{\operatorname{ver}}. Hence there exists a uniform constant ϵ>0\epsilon^{\prime}>0 such that the inequality (14) holds. The above claim is proved. ∎

The lemma is proved. ∎

6.2.2. Preparation for upper bounds of V(x,y)V(x,y) and |xy|Xhor|x-y|_{X}^{\operatorname{hor}}

Fix K4ξK\geq 4\xi where the constant ξ>λ\xi>\lambda is given by Lemma 5.6. Let 𝒞K(𝔽1)\mathcal{C}_{K}(\mathbb{F}_{1}) and 𝒞K(𝔽2)\mathcal{C}_{K}(\mathbb{F}_{2}) be the vertical quasi-trees given by Remark 5.8. With i{1,2}i\in\{1,2\}, Proposition 2.27 gives the distance formula

(\circledast) |ϑi(x)ϑi(y)|𝒞K(𝔽i)K𝔣𝔩(w)𝔽i[d𝔣𝔩(w)(ϑi(x),ϑi(y))]K\bigl{|}\vartheta_{i}(x)-\vartheta_{i}(y)\bigr{|}_{\mathcal{C}_{K}(\mathbb{F}_{i})}\;\sim_{K}\;\sum_{\mathfrak{fl}(w)\in\mathbb{F}_{i}}[d_{\mathfrak{fl}(w)}(\vartheta_{i}(x),\vartheta_{i}(y))]_{K}

To give an appropriate upper bound of V(x,y)V(x,y), we need the following two technical lemmas (Lemma 6.3 and Lemma 6.4).

Lemma 6.3.

For any v2i[v0,v2n]v_{2i}\in[v_{0},v_{2n}] with 0in0\leq i\leq n, we have

(15) d𝔣𝔩(v2i)(ϑ1(x),ϑ1(y))λ|Λv2iver(p2i)Λv2iver(p2i+1)|𝔣𝔩(v2i)\displaystyle d_{\mathfrak{fl}(v_{2i})}\bigl{(}\vartheta_{1}(x),\vartheta_{1}(y)\bigr{)}\sim_{\lambda}\bigl{|}\Lambda_{v_{2i}}^{\operatorname{ver}}(p_{2i})-\Lambda_{v_{2i}}^{\operatorname{ver}}(p_{2i+1})\bigr{|}_{\mathfrak{fl}(v_{2i})}

For any 0in10\leq i\leq n-1, we have

(16) d𝔣𝔩(v2i+1)(ϑ2(x),ϑ2(y))λ|Λv2i+1ver(p2i+1)Λv2i+1ver(p2i+2)|𝔣𝔩(v2i+1)\displaystyle d_{\mathfrak{fl}(v_{2i+1})}\bigl{(}\vartheta_{2}(x),\vartheta_{2}(y)\bigr{)}\sim_{\lambda}\bigl{|}\Lambda^{\operatorname{ver}}_{v_{2i+1}}(p_{2i+1})-\Lambda^{\operatorname{ver}}_{v_{2i+1}}(p_{2i+2})\bigr{|}_{\mathfrak{fl}(v_{2i+1})}
Proof.

We first prove (15) for the case 0<i<n0<i<n. The cases i=0i=0 or i=ni=n are similar.

Note that 1:=𝒮e2i1e2iFe2i\ell_{1}:=\mathcal{S}_{e_{2i-1}e_{2i}}\cap F_{e_{2i}} is a fiber line of Yv2i1Y_{v_{2i-1}} containing p2ip_{2i}, and similarly, 2:=𝒮e2i+1e2i+2Fe2i+1\ell_{2}:=\mathcal{S}_{e_{2i+1}e_{2i+2}}\cap F_{e_{2i+1}} contains p2i+1p_{2i+1}. By Definition 5.3 of projection maps, we have

Π𝔣𝔩(v2i)(𝔣𝔩(v0))=Λv2iver(1)\Pi_{\mathfrak{fl}(v_{2i})}(\mathfrak{fl}(v_{0}))=\Lambda_{v_{2i}}^{\operatorname{ver}}\bigl{(}\ell_{1}\bigr{)}

and

Π𝔣𝔩(v2i)(𝔣𝔩(v2n))=Λv2iver(2)\Pi_{\mathfrak{fl}(v_{2i})}(\mathfrak{fl}(v_{2n}))=\Lambda_{v_{2i}}^{\operatorname{ver}}\bigl{(}\ell_{2}\bigr{)}

Let λ>0\lambda>0 be the constant given by Lemma 5.2, so the fiber lines 1,2\ell_{1},\ell_{2} are sent by Λv2iver\Lambda_{v_{2i}}^{\operatorname{ver}} into v2i\mathcal{L}_{v_{2i}} as subsets of diameter at most λ\lambda:

diamΛv2iver(1),diamΛv2iver(2)λ\operatorname{diam}{\Lambda_{v_{2i}}^{\operatorname{ver}}(\ell_{1})},\;\operatorname{diam}{\Lambda_{v_{2i}}^{\operatorname{ver}}(\ell_{2})}\leq\lambda

By definition of ϑ1\vartheta_{1}, we have ϑ1(x)=Λv0ver(x)𝔣𝔩(v0)\vartheta_{1}(x)=\Lambda_{v_{0}}^{\operatorname{ver}}(x)\in\mathfrak{fl}(v_{0}) and ϑ1(y)=Λv2nver(y)𝔣𝔩(v2n)\vartheta_{1}(y)=\Lambda_{v_{2n}}^{\operatorname{ver}}(y)\in\mathfrak{fl}(v_{2n}). Thus

d𝔣𝔩(v2i)(ϑ1(x),ϑ1(y))λd𝔣𝔩(v2i)(𝔣𝔩(v0),𝔣𝔩(v2n))d_{\mathfrak{fl}(v_{2i})}\bigl{(}\vartheta_{1}(x),\vartheta_{1}(y)\bigr{)}\sim_{\lambda}d_{\mathfrak{fl}(v_{2i})}\bigl{(}\mathfrak{fl}(v_{0}),\mathfrak{fl}(v_{2n})\bigr{)}

As p2i1p_{2i}\in\ell_{1} and p2i+12p_{2i+1}\in\ell_{2}, we obtain

d𝔣𝔩(v2i)(𝔣𝔩(v0),𝔣𝔩(v2n))λ|Λv2iver(p2i)Λv2iver(p2i+1)|𝔣𝔩(v2i)d_{\mathfrak{fl}(v_{2i})}\bigl{(}\mathfrak{fl}(v_{0}),\mathfrak{fl}(v_{2n})\bigr{)}\sim_{\lambda}\bigl{|}\Lambda_{v_{2i}}^{\operatorname{ver}}(p_{2i})-\Lambda_{v_{2i}}^{\operatorname{ver}}(p_{2i+1})\bigr{|}_{\mathfrak{fl}(v_{2i})}

completing the proof of (15).

We are now going to prove (16). If w0v1w_{0}\neq v_{1} or 1in11\leq i\leq n-1, the same proof for (15) proves (16). We now consider w0=v1w_{0}=v_{1} and i=0i=0. In this case, we note that e0=e¯1e_{0}=\overline{e}_{1}. By definition, we have that ϑ2(x)=ϑ2(o)=Λw0ver(o)𝔣𝔩(w0)\vartheta_{2}(x)=\vartheta_{2}(o)=\Lambda^{\operatorname{ver}}_{w_{0}}(o)\in\mathfrak{fl}(w_{0}), so we obtain Π𝔣𝔩(v1)(ϑ2(x))=ϑ2(x)\Pi_{\mathfrak{fl}(v_{1})}(\vartheta_{2}(x))=\vartheta_{2}(x). Recall that 𝒮xe1\mathcal{S}_{xe_{1}} is the strip in Yv0Y_{v_{0}} over the shortest arc from xx to Fe1F_{e_{1}} (See construction of special path). As xFe0=Fe1x\in F_{e_{0}}=F_{e_{1}}, we have 1:=𝒮xe1\ell_{1}:=\mathcal{S}_{xe_{1}} is a fiber line of Yv0Y_{v_{0}} that passes through xx and also p1p_{1}. Thus, ϑ2(x)Π𝔣𝔩(v1)(1)\vartheta_{2}(x)\in\Pi_{\mathfrak{fl}(v_{1})}(\ell_{1}).

Recall that 𝒮xe1\mathcal{S}_{xe_{1}} is the strip in Yv0Y_{v_{0}} over the shortest arc from xx to Fe1F_{e_{1}} (See construction of special path). As xFe0=Fe1x\in F_{e_{0}}=F_{e_{1}}, we have 1:=𝒮xe1\ell_{1}:=\mathcal{S}_{xe_{1}} is a fiber line of Yv0Y_{v_{0}} that passes through xx and also p1p_{1}. Thus, ϑ2(x)Π𝔣𝔩(v1)(1)\vartheta_{2}(x)\in\Pi_{\mathfrak{fl}(v_{1})}(\ell_{1}). Let 2=𝒮e2e3Fe2\ell_{2}=\mathcal{S}_{e_{2}e_{3}}\cap F_{e_{2}} be the fiber line on Yv2Y_{v_{2}} that passes through p2p_{2}. If w1=v1w_{1}=v_{1}, then α=[v0,v1][v1,v2]\alpha=[v_{0},v_{1}][v_{1},v_{2}] and yFe2y\in F_{e_{2}}. By the same reason, 2\ell_{2} passes through yy, so ϑ2(y)Π𝔣𝔩(v2)(2)\vartheta_{2}(y)\in\Pi_{\mathfrak{fl}(v_{2})}(\ell_{2}). If w1v1w_{1}\neq v_{1}, the projection Πv1(ϑ2(y))\Pi_{\mathcal{L}_{v_{1}}}(\vartheta_{2}(y)) must be contained in Π𝔣𝔩(v1)(2)\Pi_{\mathfrak{fl}(v_{1})}(\ell_{2}). In both cases, we have

d𝔣𝔩(v1)(ϑ2(x),ϑ2(y))λdiam(Λv1ver(1)Λv1ver(2))d_{\mathfrak{fl}(v_{1})}(\vartheta_{2}(x),\vartheta_{2}(y))\sim_{\lambda}\operatorname{diam}(\Lambda^{\operatorname{ver}}_{v_{1}}(\ell_{1})\cup\Lambda^{\operatorname{ver}}_{v_{1}}(\ell_{2}))

where we use diam(Λv1ver(1)),diam(Λv1ver(2))λ\operatorname{diam}\bigl{(}\Lambda^{\operatorname{ver}}_{v_{1}}(\ell_{1})\bigr{)},\;\operatorname{diam}\bigl{(}\Lambda^{\operatorname{ver}}_{v_{1}}(\ell_{2})\bigr{)}\leq\lambda by Lemma 5.2. For p11p_{1}\in\ell_{1} and p22p_{2}\in\ell_{2}, we obtain

diam(Λv1ver(1)Λv1ver(2))λ|Λv1ver(p1)Λv1ver(p2)|𝔣𝔩(v1)\operatorname{diam}(\Lambda^{\operatorname{ver}}_{v_{1}}(\ell_{1})\cup\Lambda^{\operatorname{ver}}_{v_{1}}(\ell_{2}))\sim_{\lambda}\bigl{|}\Lambda^{\operatorname{ver}}_{v_{1}}(p_{1})-\Lambda^{\operatorname{ver}}_{v_{1}}(p_{2})\bigr{|}_{\mathfrak{fl}(v_{1})}

concluding the proof of (16). ∎

Let us recall the notation from Section 2.4. Let x𝔣𝔩(v),z𝔣𝔩(u)𝔽ix\in\mathfrak{fl}(v),z\in\mathfrak{fl}(u)\in\mathbb{F}_{i}.

If 𝔣𝔩(v)𝔣𝔩(u)𝔣𝔩(w)\mathfrak{fl}(v)\neq\mathfrak{fl}(u)\neq\mathfrak{fl}(w) then

d𝔣𝔩(w)(x,z):=d𝔣𝔩(w)(𝔣𝔩(v),𝔣𝔩(u))d_{\mathfrak{fl}(w)}(x,z):=d_{\mathfrak{fl}(w)}(\mathfrak{fl}(v),\mathfrak{fl}(u))

If 𝔣𝔩(w)=𝔣𝔩(v),𝔣𝔩(w)𝔣𝔩(u)\mathfrak{fl}(w)=\mathfrak{fl}(v),\mathfrak{fl}(w)\neq\mathfrak{fl}(u), define d𝔣𝔩(w)(x,z):=diam(π𝔣𝔩(w)(x,𝔣𝔩(u)))d_{\mathfrak{fl}(w)}(x,z):=\operatorname{diam}(\pi_{\mathfrak{fl}(w)}(x,\mathfrak{fl}(u))).

If 𝔣𝔩(v)=𝔣𝔩(u)=𝔣𝔩(w)\mathfrak{fl}(v)=\mathfrak{fl}(u)=\mathfrak{fl}(w), let d𝔣𝔩(w)(x,z)d_{\mathfrak{fl}(w)}(x,z) be the distance in 𝔣𝔩(w)\mathfrak{fl}(w).

Lemma 6.4.

Let ϑ2\vartheta_{2} be the map given by Section 6.1. Let vv be a vertex in vLk(v2i)α~v\in\operatorname{Lk}(v_{2i})-\widetilde{\alpha} and let e=[v,v2i]e=[v,v_{2i}]. Let e\ell_{e}, e2i\ell_{e_{2i}}, and e¯2i+1\ell_{\overline{e}_{2i+1}} be the boundary lines of F~v2i\widetilde{F}_{v_{2i}} associated to distinct edges e,e2ie,e_{2i} and e¯2i+1\overline{e}_{2i+1} respectively. Then we have

(17) d𝔣𝔩(v)(ϑ2(x),ϑ2(y))λde(e2i,e¯2i+1)\displaystyle d_{\mathfrak{fl}(v)}(\vartheta_{2}(x),\vartheta_{2}(y))\sim_{\lambda}d_{\ell_{e}}(\ell_{e_{2i}},\ell_{\overline{e}_{2i+1}})
Proof.

Note that e\ell_{e}, e2i\ell_{e_{2i}}, and e¯2i+1\ell_{\overline{e}_{2i+1}} are the projection of planes FeF_{e}, Fe2iF_{e_{2i}}, Fe2i+1F_{e_{2i+1}} of Yv2iY_{v_{2i}} into the factor Yv2iY_{v_{2i}}. We prove (17) case by case, according to the configuration of e0,e2n+1e_{0},e_{2n+1} with α\alpha.

Case 1. 0<i<n0<i<n. By Definition 5.3 of projection maps, the projection of ϑ2(x)=Λw0ver(o)𝔣𝔩(w0)\vartheta_{2}(x)=\Lambda^{\operatorname{ver}}_{w_{0}}(o)\in\mathfrak{fl}(w_{0}) to 𝔣𝔩(v)\mathfrak{fl}(v) is the same as that of 𝔣𝔩(v1)\mathfrak{fl}(v_{1}) to 𝔣𝔩(v)\mathfrak{fl}(v), and the projection of ϑ2(y)𝔣𝔩(w1)\vartheta_{2}(y)\in\mathfrak{fl}(w_{1}) to 𝔣𝔩(v)\mathfrak{fl}(v) is the same as that of 𝔣𝔩(v2n1)\mathfrak{fl}(v_{2n-1}) to 𝔣𝔩(v)\mathfrak{fl}(v). That is to say, d𝔣𝔩(v)(ϑ2(x),ϑ2(y))=d𝔣𝔩(v)(𝔣𝔩(v1),𝔣𝔩(v2n1))d_{\mathfrak{fl}(v)}\bigl{(}\vartheta_{2}(x),\vartheta_{2}(y)\bigr{)}=d_{\mathfrak{fl}(v)}\bigl{(}\mathfrak{fl}(v_{1}),\mathfrak{fl}(v_{2n-1})\bigr{)}. Hence, the Equation (17) follows by Lemma 5.5: d𝔣𝔩(v)(𝔣𝔩(v1),𝔣𝔩(v2n1))λde(e2i,e¯2i+1)d_{\mathfrak{fl}(v)}\bigl{(}\mathfrak{fl}(v_{1}),\mathfrak{fl}(v_{2n-1})\bigr{)}\sim_{\lambda}d_{\ell_{e}}(\ell_{e_{2i}},\ell_{\overline{e}_{2i+1}}) for any vLk(v2i)α~.v\in\operatorname{Lk}(v_{2i})-\widetilde{\alpha}.

Case 2. i=0i=0 or i=ni=n. We only consider the case i=0i=0 and analyze the configuration of w0w_{0} with α\alpha. The analyze for the case for i=ni=n and w1w_{1} is symmetric.

Case 2.1: w0v1w_{0}\neq v_{1}. In this case e0αe_{0}\cdot\alpha is a geodesic from w0w_{0} to v2nv_{2n}. By Definition 5.3 of projection maps, no matter whether e¯2n+1=e2n\overline{e}_{2n+1}=e_{2n} holds or not, the projection of ϑ2(x)𝔣𝔩(w0)\vartheta_{2}(x)\in\mathfrak{fl}(w_{0}) to 𝔣𝔩(v)\mathfrak{fl}(v) is the same as that of 𝔣𝔩(w0)\mathfrak{fl}(w_{0}) to 𝔣𝔩(v)\mathfrak{fl}(v), and the projection of ϑ2(y)𝔣𝔩(w1)\vartheta_{2}(y)\in\mathfrak{fl}(w_{1}) to 𝔣𝔩(v)\mathfrak{fl}(v) is the same as that of 𝔣𝔩(v2n1)\mathfrak{fl}(v_{2n-1}) to 𝔣𝔩(v)\mathfrak{fl}(v). By Lemma 5.5, we have d𝔣𝔩(v)(𝔣𝔩(w0),𝔣𝔩(v2n1))λde(e2i,e¯2i+1)d_{\mathfrak{fl}(v)}\bigl{(}\mathfrak{fl}(w_{0}),\mathfrak{fl}(v_{2n-1})\bigr{)}\sim_{\lambda}d_{\ell_{e}}(\ell_{e_{2i}},\ell_{\overline{e}_{2i+1}}).

Case 2.2: w0=v1w_{0}=v_{1}. No matter whether w0=w1w_{0}=w_{1} or not, we have d𝔣𝔩(v)(ϑ2(x),ϑ2(y))Π𝔣𝔩(v)(𝔣𝔩(w0))ξd_{\mathfrak{fl}(v)}(\vartheta_{2}(x),\vartheta_{2}(y))\leq\Pi_{\mathfrak{fl}(v)}(\mathfrak{fl}(w_{0}))\leq\xi where ξ\xi is the projection constant given by Lemma 5.6. On the right side of (17), we have de(e2i,e¯2i+1)d_{\ell_{e}}(\ell_{e_{2i}},\ell_{\overline{e}_{2i+1}}) is bounded above by ξ\xi for i=0i=0 (as e0=e¯1e_{0}=\overline{e}_{1}). Thus (17) holds as well in this case. ∎

6.2.3. Upper bound of V(x,y)V(x,y)

Let ϑ1\vartheta_{1} and ϑ2\vartheta_{2} be the maps defined in Section 6.1. We now have prepared all ingredients for the proof of the following result.

Proposition 6.5.

Let x,yGox,y\in Go and α:=[ρ(x),ρ(y)]\alpha:=[\rho(x),\rho(y)] be the geodesic in TT. Then

(18) V(x,y)Kj=1,2(vα𝒱j[d𝔣𝔩(v)(ϑj(x),ϑj(y))]K)+dT(ρ(x),ρ(y))V(x,y)\preceq_{K}\sum_{j=1,2}\left(\sum_{v\in\alpha\cap\mathcal{V}_{j}}[d_{\mathfrak{fl}(v)}(\vartheta_{j}(x),\vartheta_{j}(y))]_{K}\right)+d_{T}(\rho(x),\rho(y))
Proof.

The goal is to recover the sum on the right side of (12), that is

V(x,y)=0i2n|Λviver(pi)Λviver(pi+1)|𝔣𝔩(vi)V(x,y)=\sum_{0\leq i\leq 2n}\bigl{|}\Lambda^{\operatorname{ver}}_{v_{i}}(p_{i})-\Lambda^{\operatorname{ver}}_{v_{i}}(p_{i+1})\bigr{|}_{\mathfrak{fl}(v_{i})}

via the maps ϑ1\vartheta_{1} and ϑ2\vartheta_{2}. By Lemma 6.3, we have desired inequalities (15) for even indices v2i[v0,v2n]𝒱1v_{2i}\in[v_{0},v_{2n}]\cap\mathcal{V}_{1} with 0in0\leq i\leq n, that is

d𝔣𝔩(v2i)(ϑ1(x),ϑ1(y))λ|Λv2iver(p2i)Λv2iver(p2i+1)|𝔣𝔩(v2i)d_{\mathfrak{fl}(v_{2i})}\bigl{(}\vartheta_{1}(x),\vartheta_{1}(y)\bigr{)}\sim_{\lambda}\bigl{|}\Lambda_{v_{2i}}^{\operatorname{ver}}(p_{2i})-\Lambda_{v_{2i}}^{\operatorname{ver}}(p_{2i+1})\bigr{|}_{\mathfrak{fl}(v_{2i})}

By Lemma 6.4, the inequalities (16) recover the odd indices v2i+1[v0,v2n]𝒱2v_{2i+1}\in[v_{0},v_{2n}]\cap\mathcal{V}_{2} with 0in10\leq i\leq n-1 in (12), that is,

d𝔣𝔩(v2i+1)(ϑ2(x),ϑ2(y))λ|Λv2i+1ver(p2i+1)Λv2i+1ver(p2i+2)|𝔣𝔩(v2i+1)d_{\mathfrak{fl}(v_{2i+1})}\bigl{(}\vartheta_{2}(x),\vartheta_{2}(y)\bigr{)}\sim_{\lambda}\bigl{|}\Lambda^{\operatorname{ver}}_{v_{2i+1}}(p_{2i+1})-\Lambda^{\operatorname{ver}}_{v_{2i+1}}(p_{2i+2})\bigr{|}_{\mathfrak{fl}(v_{2i+1})}

Plugging the inequalities (15) (16) into (12), and using the term |ρ(x)ρ(y)|T|\rho(x)-\rho(y)|_{T} to count the additive errors in this process completes the proof of the desired inequality (18). Applying then the KK-cutoff function []K[\cdot]_{K} does not affect the inequalities. ∎

6.2.4. Upper bound of |xy|Xhor|x-y|_{X}^{\operatorname{hor}}

The horizontal distance dhd^{h} defined in (2) of the special path γ\gamma from xx to yy records the totality of the projected distances to the base hyperbolic spaces Y¯v\overline{Y}_{v}:

|xy|Xhor=|xp1|Yv1hor+|p1p2|Yv2hor++|p2ny|Yv2nhor=|ϑ3(x)Fe1|Yv0+i=12n1|FeiFei+1|Yvi+|Fe2nϑ3(y)|Yv2n\begin{array}[]{rl}\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}&=\bigl{|}x-p_{1}\bigr{|}_{Y_{v_{1}}}^{\operatorname{hor}}+\bigl{|}p_{1}-p_{2}\bigr{|}_{Y_{v_{2}}}^{\operatorname{hor}}+\cdots+\bigl{|}p_{2n}-y\bigr{|}_{Y_{v_{2n}}}^{\operatorname{hor}}\\ &\\ &=|\vartheta_{3}(x)-F_{e_{1}}|_{Y_{v_{0}}}+\displaystyle\sum_{i=1}^{2n-1}|F_{e_{i}}-F_{e_{i+1}}|_{Y_{v_{i}}}+|F_{e_{2n}}-\vartheta_{3}(y)|_{Y_{v_{2n}}}\end{array}

where the map ϑ3\vartheta_{3} defined in Section 6.1 sends a point in Yv=Y¯v×Y_{v}=\overline{Y}_{v}\times\mathbb{R} to the hyperbolic base Y¯v\overline{Y}_{v}.

Before moving on, let us introduce more notations to represent the horizontal distance. Let x0=ϑ3(x),y0Fe1x_{0}=\vartheta_{3}(x),y_{0}\in F_{e_{1}} and x2nFe2n,y2n=ϑ3(y)x_{2n}\in F_{e_{2n}},y_{2n}=\vartheta_{3}(y) so that [x0,y0][x_{0},y_{0}] is orthogonal to Fe1F_{e_{1}}, and [x2n,y2n][x_{2n},y_{2n}] to Fe2nF_{e_{2n}}. Choose xiFei,yiFei+1x_{i}\in F_{e_{i}},y_{i}\in F_{e_{i+1}} so that [xi,yi][x_{i},y_{i}] is a geodesic in Y¯vi\overline{Y}_{v_{i}} orthogonal to FeiF_{e_{i}} and Fei+1F_{e_{i+1}}. Thus,

(19) |xy|Xhor=i=02n|xiyi|Y¯vi.\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}=\sum_{i=0}^{2n}|x_{i}-y_{i}|_{\overline{Y}_{v_{i}}}.

Recall that 𝒳1˙,𝒳2˙\dot{\mathcal{X}_{1}},\dot{\mathcal{X}_{2}} are the coned-off spaces defined in Section 4.1. By Definition 4.3 of the KK-thick distance of 𝒳j˙\dot{\mathcal{X}_{j}} for any K>0K>0 and the Remark after it, we have

(20) i=02n|xiyi|Y˙viK=|ϑ3(x)ϑ3(y)|𝒳1˙K+|ϑ4(x)ϑ4(y)|𝒳2˙K\sum_{i=0}^{2n}\bigl{|}x_{i}-y_{i}\bigr{|}_{\dot{Y}_{v_{i}}}^{K}=\bigl{|}\vartheta_{3}(x)-\vartheta_{3}(y)\bigr{|}_{\dot{\mathcal{X}_{1}}}^{K}+\bigl{|}\vartheta_{4}(x)-\vartheta_{4}(y)\bigr{|}_{\dot{\mathcal{X}_{2}}}^{K}

where ||Y˙viK|\cdot|_{\dot{Y}_{v_{i}}}^{K} defined in (3) is the KK-thick distance on the coned-off space Y˙vi\dot{Y}_{v_{i}}. The map ϑ4\vartheta_{4} defined in Section 6.1 sends a point gogo in GoGo to the hyperbolic cone point to the boundary line g[v0,w0]\ell_{g[v_{0},w_{0}]} (recall that oo is chosen on a common boundary plane F[v0,w0]F_{[v_{0},w_{0}]}).

Hence, the KK–thick distance (20) differs from the horizontal distance (19) by the amount coned-off on boundary lines. The purpose of this subsection is to recover the loss in the coned-off from the projection system of fiber lines.

Lemma 6.6.

For any x,yGox,y\in Go, we have

|xy|XhorKi=02n|xiyi|Y˙viK+i=02nwLk(vi)α[d𝔣𝔩(w)(ϑj(x),ϑj(y))]K\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}\preceq_{K}\sum_{i=0}^{2n}\bigl{|}x_{i}-y_{i}\bigr{|}_{\dot{Y}_{v_{i}}}^{K}+\sum_{i=0}^{2n}\sum_{w\in\operatorname{Lk}(v_{i})-\alpha}[d_{\mathfrak{fl}(w)}(\vartheta_{j}(x),\vartheta_{j}(y))]_{K}

where the index j=1j=1 is chosen if ii is odd, otherwise j=2j=2.

Proof.

We consider the equation (19) for the horizontal distance |xy|Xhor\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}. Let 𝕃vi\mathbb{L}_{v_{i}} be the set of boundary lines of Y¯vi\overline{Y}_{v_{i}} corresponding to the set of oriented edges eSt(vi)e\in St(v_{i}) (i.e., the collection {FeY¯vi:eSt(vi)}\{F_{e}\cap\overline{Y}_{v_{i}}:e\in St(v_{i})\}). By Lemma 3.1, for each 0i2n0\leq i\leq 2n, we have

(21) |xiyi|Y¯viK|xiyi|Y˙viK+e𝕃vi[de(xi,yi)]K\displaystyle\bigl{|}x_{i}-y_{i}\bigr{|}_{\overline{Y}_{v_{i}}}\sim_{K}\bigl{|}x_{i}-y_{i}\bigr{|}_{\dot{Y}_{v_{i}}}^{K}+\sum_{\ell_{e}\in\mathbb{L}_{v_{i}}}[d_{\ell_{e}}(x_{i},y_{i})]_{K}

for any sufficiently large K0K\gg 0.

Let e=[w,vi]St(vi)e=[w,v_{i}]\in St(v_{i}) and e𝕃vi\ell_{e}\in\mathbb{L}_{v_{i}} be the corresponding boundary line of Y¯vi\overline{Y}_{v_{i}}. Set j=1j=1 if ii is odd, otherwise j=2j=2.

If e=eie=e_{i} or e=e¯i+1e=\overline{e}_{i+1} for 1i2n11\leq i\leq 2n-1, then

de(xi,yi)ξd_{\ell_{e}}(x_{i},y_{i})\leq\xi

since [xi,yi][x_{i},y_{i}] is orthogonal to e\ell_{e}.

We remark that when i=0i=0 (the case i=2ni=2n is similar), it is possible that [x0,y0][x_{0},y_{0}] may not be perpendicular to e\ell_{e}. However, we have

de(x0,y0)d𝔣𝔩(w0)(ϑ2(x),ϑ2(y))d_{\ell_{e}}(x_{0},y_{0})\preceq d_{\mathfrak{fl}(w_{0})}(\vartheta_{2}(x),\vartheta_{2}(y))

Otherwise, if eeie\neq e_{i} and ee¯i+1e\neq\overline{e}_{i+1} for 1i2n11\leq i\leq 2n-1, we have eαe\notin\alpha for which the following holds by Lemma 6.4 for j=2j=2 and by Lemma 5.5 for j=1j=1 ,

d𝔣𝔩(w)(ϑj(x),ϑj(y))de(xi,yi).d_{\mathfrak{fl}(w)}(\vartheta_{j}(x),\vartheta_{j}(y))\sim d_{\ell_{e}}(x_{i},y_{i}).

Note that AλB+CA\leq\lambda B+C with BKCB\geq K\geq C implies [A]KK[B]K[A]_{K}\preceq_{K}[B]_{K}. Thus, for each 0i2n0\leq i\leq 2n, we deduce from (21) the following

(22) |xiyi|Y¯viK|xiyi|Y˙viK+wLk(vi)α[d𝔣𝔩(w)(ϑj(x),ϑj(y))]K\displaystyle\bigl{|}x_{i}-y_{i}\bigr{|}_{\overline{Y}_{v_{i}}}\sim_{K}\bigl{|}x_{i}-y_{i}\bigr{|}_{\dot{Y}_{v_{i}}}^{K}+\sum_{w\in Lk(v_{i})-\alpha}[d_{\mathfrak{fl}(w)}(\vartheta_{j}(x),\vartheta_{j}(y))]_{K}

for any K0K\gg 0, where j=1j=1 if ii is odd, and j=2j=2 otherwise. We sum up (22) over viαv_{i}\in\alpha to get the horizontal distance dh(x,y)d^{h}(x,y) in (19):

|xy|Xhor\displaystyle\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}} =i=02n|xiyi|Y¯vi\displaystyle=\sum_{i=0}^{2n}\bigl{|}x_{i}-y_{i}\bigr{|}_{\overline{Y}_{v_{i}}}
K\displaystyle\preceq_{K} i=02n|xiyi|Y˙viK+i=02nwLk(vi)α[d𝔣𝔩(w)(ϑj(x),ϑj(y))]K\displaystyle\sum_{i=0}^{2n}\bigl{|}x_{i}-y_{i}\bigr{|}_{\dot{Y}_{v_{i}}}^{K}+\sum_{i=0}^{2n}\sum_{w\in\operatorname{Lk}(v_{i})-\alpha}[d_{\mathfrak{fl}(w)}(\vartheta_{j}(x),\vartheta_{j}(y))]_{K}

We now have prepared all ingredients in the proof of Proposition 6.1.

Proof of Proposition 6.1.

Since ρ\rho, ϑi\vartheta_{i} (with i{1,2,3,4}i\in\{1,2,3,4\}) are GG-equivariant maps, it follows that Φ\Phi is a GG-equivariant map. Since the orbital map of any isometric action is Lipschitz (e.g. see [BH99, Lemma I.8.18]), it suffices to give an upper bound on d(x,y)d(x,y).

Let ϵ>0\epsilon>0 be the constant given by Lemma 6.2, so that

|xy|Xϵ(|ρ(x)ρ(y)|T+|xy|Xhor+V(x,y))|x-y|_{X}\leq\epsilon\left(|\rho(x)-\rho(y)|_{T}+\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}+V(x,y)\right)

Appropriate upper bounds of the vertical distance V(x,y)V(x,y) and the horizontal distance |xy|Xhor\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}} have been already treated in Proposition 6.5 and Lemma 6.6 respectively. They are

V(x,y)Kj=1,2(vα𝒱j[d𝔣𝔩(v)(ϑj(x),ϑj(y))]K)+|ρ(x)ρ(y)|TV(x,y)\preceq_{K}\sum_{j=1,2}\left(\sum_{v\in\alpha\cap\mathcal{V}_{j}}[d_{\mathfrak{fl}(v)}(\vartheta_{j}(x),\vartheta_{j}(y))]_{K}\right)+|\rho(x)-\rho(y)|_{T}

and

|xy|XhorKi=02n|xiyi|Y˙viK+i=02nwLk(vi)α[d𝔣𝔩(w)(ϑj(x),ϑj(y))]K\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}\preceq_{K}\sum_{i=0}^{2n}\bigl{|}x_{i}-y_{i}\bigr{|}_{\dot{Y}_{v_{i}}}^{K}+\sum_{i=0}^{2n}\sum_{w\in\operatorname{Lk}(v_{i})-\alpha}[d_{\mathfrak{fl}(w)}(\vartheta_{j}(x),\vartheta_{j}(y))]_{K}

where the index jj depends on ii: j=1j=1 if ii is odd, otherwise j=2j=2. The above two inequalities yield:

|xy|Xhor+V(x,y)K|ρ(x)ρ(y)|T+i=02n|xiyi|Y˙viK+i=02nwLk(vi)[d𝔣𝔩(w)(ϑj(x),ϑj(y))]K\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}+V(x,y)\preceq_{K}|\rho(x)-\rho(y)|_{T}+\sum_{i=0}^{2n}|x_{i}-y_{i}|_{\dot{Y}_{v_{i}}}^{K}+\sum_{i=0}^{2n}\sum_{w\in\operatorname{Lk}(v_{i})}[d_{\mathfrak{fl}(w)}(\vartheta_{j}(x),\vartheta_{j}(y))]_{K}

By (\circledast6.2.2), we have

i=02nwLk(vi)[d𝔣𝔩(w)(ϑj(x),ϑj(y))]KK|ϑ1(x)ϑ1(x)|𝒞K(𝔽1)+|ϑ2(x)ϑ2(x)|𝒞K(𝔽2)\sum_{i=0}^{2n}\sum_{w\in\operatorname{Lk}(v_{i})}[d_{\mathfrak{fl}(w)}(\vartheta_{j}(x),\vartheta_{j}(y))]_{K}\preceq_{K}\bigl{|}\vartheta_{1}(x)-\vartheta_{1}(x)\bigr{|}_{\mathcal{C}_{K}(\mathbb{F}_{1})}+\bigl{|}\vartheta_{2}(x)-\vartheta_{2}(x)\bigr{|}_{\mathcal{C}_{K}(\mathbb{F}_{2})}

It follows that

|xy|Xhor+V(x,y)K|ρ(x)ρ(y)|T+i=02n|xiyi|Y˙viK+i=12|ϑi(x)ϑi(x)|𝒞K(𝔽i)\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}+V(x,y)\preceq_{K}|\rho(x)-\rho(y)|_{T}+\sum_{i=0}^{2n}|x_{i}-y_{i}|_{\dot{Y}_{v_{i}}}^{K}+\sum_{i=1}^{2}\bigl{|}\vartheta_{i}(x)-\vartheta_{i}(x)\bigr{|}_{\mathcal{C}_{K}(\mathbb{F}_{i})}

Plugging the thick distance formula (20) into the above inequality, we obtain

|xy|Xhor+V(x,y)\displaystyle\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}+V(x,y) K|ρ(x)ρ(y)|T+|ϑ3(x)ϑ3(y)|𝒳1˙K+|ϑ4(x)ϑ4(y)|𝒳2˙K\displaystyle\preceq_{K}|\rho(x)-\rho(y)|_{T}+\bigl{|}\vartheta_{3}(x)-\vartheta_{3}(y)\bigr{|}_{\dot{\mathcal{X}_{1}}}^{K}+\bigl{|}\vartheta_{4}(x)-\vartheta_{4}(y)\bigr{|}_{\dot{\mathcal{X}_{2}}}^{K}
+|ϑ1(x)ϑ1(x)|𝒞K(𝔽1)+|ϑ2(x)ϑ2(x)|𝒞K(𝔽2)\displaystyle+\bigl{|}\vartheta_{1}(x)-\vartheta_{1}(x)\bigr{|}_{\mathcal{C}_{K}(\mathbb{F}_{1})}+\bigl{|}\vartheta_{2}(x)-\vartheta_{2}(x)\bigr{|}_{\mathcal{C}_{K}(\mathbb{F}_{2})}

As |xy|Xϵ(|ρ(x)ρ(y)|+|xy|Xhor+V(x,y))|x-y|_{X}\leq\epsilon(|\rho(x)-\rho(y)|+\bigl{|}x-y\bigr{|}_{X}^{\operatorname{hor}}+V(x,y)), it is a consequence from the above inequality that the map Φ=ϑ1×ϑ2×ϑ3×ϑ4×ρ\Phi=\vartheta_{1}\times\vartheta_{2}\times\vartheta_{3}\times\vartheta_{4}\times\rho in (\clubsuit6.1) is a GG–equivariant quasi-isometric embedding from XX to 𝒞K(𝔽1)×𝒞K(𝔽2)×𝒳1˙×𝒳2˙×T\mathcal{C}_{K}(\mathbb{F}_{1})\times\mathcal{C}_{K}(\mathbb{F}_{2})\times\dot{\mathcal{X}_{1}}\times\dot{\mathcal{X}_{2}}\times T. The proof of Proposition is complete. ∎

7. Proof of Theorem 1.3

Let GXG\curvearrowright X be a CKA action such that for every vertex group the central extension (1) has omnipotent hyperbolic quotient group. Let G˙<G\dot{G}<G be the subgroup of the index at most 22 preserving 𝒱1\mathcal{V}_{1} and 𝒱2\mathcal{V}_{2} given by Lemma 2.23. Upon passing to further finite index subgroups in Lemma 4.8, we obtain a finite index subgroup GG^{\prime} of G˙\dot{G} such that the results in Section 5 and Section 6 hold for GG^{\prime}. We caution the reader that at the beginning of Section 5 we assume that each vertex group of GG is a direct product, this assumption may not hold for the original GG, but holds in the finite index subgroup GG^{\prime} of GG.

As GG^{\prime} is a subgroup of G˙\dot{G}, it follows from Proposition 4.7 that there exists a GG^{\prime}–equivariant quasi-isometric embedding

η:(𝒳1˙×𝒳2˙×T,d𝒳1˙K×d𝒳2˙K×dT)T1×T2×Tn×T\eta\colon(\dot{\mathcal{X}_{1}}\times\dot{\mathcal{X}_{2}}\times T,d^{K}_{\dot{\mathcal{X}_{1}}}\times d^{K}_{\dot{\mathcal{X}_{2}}}\times d_{T})\to T_{1}\times T_{2}\cdots\times T_{n}\times T

Applying Proposition 6.1 to GG^{\prime}, we have a GG^{\prime}–equivariant quasi-isometric embedding

Φ:Go𝒞K(𝔽1)×𝒞K(𝔽2)×(𝒳1˙,d𝒳1˙K)×(𝒳2˙,d𝒳2˙K)×T\Phi\colon G^{\prime}o\to\mathcal{C}_{K}(\mathbb{F}_{1})\times\mathcal{C}_{K}(\mathbb{F}_{2})\times(\dot{\mathcal{X}_{1}},d^{K}_{\dot{\mathcal{X}_{1}}})\times(\dot{\mathcal{X}_{2}},d^{K}_{\dot{\mathcal{X}_{2}}})\times T

It implies that (id𝒞K(𝔽1)×id𝒞K(𝔽2)×η)Φ(id_{\mathcal{C}_{K}(\mathbb{F}_{1})}\times id_{\mathcal{C}_{K}(\mathbb{F}_{2})}\times\eta)\circ\Phi is a GG^{\prime}–equivariant quasi-isometric embedding from GoG^{\prime}\cdot o to the finite product of quasi-trees 𝒞K(𝔽1)×𝒞K(𝔽1)×T1×T2×Tn×T\mathcal{C}_{K}(\mathbb{F}_{1})\times\mathcal{C}_{K}(\mathbb{F}_{1})\times T_{1}\times T_{2}\cdots\times T_{n}\times T. Thus GG^{\prime} has property (QT), implying GG has property (QT).

8. Applications: Property (QT) of 3-manifold groups

In this section, we apply results obtained in previous sections to give a complete characterization of property (QT) of all finitely generated 3-manifold groups (Theorem 1.1). Note that property (QT) is a commensurability invariant. Hence, we can always assume that all 3-manifolds are compact, orientable (by taking Scott’s compact core and double cover).

Let MM be a compact, connected, orientable, irreducible 3-manifold with empty or tori boundary. MM is called geometric if its interior admits geometric structures in the sense of Thurston, that are S3S^{3}, 𝔼3\mathbb{E}^{3}, 3\mathbb{H}^{3}, S2×S^{2}\times\mathbb{R}, 2×\mathbb{H}^{2}\times\mathbb{R}, SL(2,)~\widetilde{SL(2,\mathbb{R})}, Nil and Sol. If MM is not geometric, then MM is called a nongeometric 33–manifold. By geometric decomposition of 33–manifolds, there is a nonempty minimal union 𝒯M\mathcal{T}\subset M of disjoint essential tori and Klein bottles, unique up to isotopy, such that each component of M\𝒯M\backslash\mathcal{T} is either a Seifert fibered piece or a hyperbolic piece. MM is called graph manifold if all the pieces of M\𝒯M\backslash\mathcal{T} are Seifert pieces, otherwise it is a mixed manifold.

We remark here that the geometric decomposition is slightly different from the torus decomposition, but they are closely related (if MM has no decomposing Klein bottle, then these two decompositions agree with each other). Such a difference can be got rid of by passing to some finite cover of MM. Since we are only interested in virtual properties of 33–manifolds in this paper, we can always assume that these two decompositions agree with each other (on some finite cover of MM).

8.1. Property (QT) of geometric 3-manifolds

Proposition 8.1.

The fundamental group π1(M)\pi_{1}(M) of a geometric 33–manifold MM has property (QT) if and only if MM does not support Sol and Nil geometry.

Proof.

We are going to prove the necessity. Assume that π1(M)\pi_{1}(M) has property (QT). By Lemma 2.5, π1(M)\pi_{1}(M) does not contain any distorted element, while the fundamental group of a 3-manifold with Nil geometry or Sol geometry contains quadratically/exponentially distorted elements (for example, see [NS20, Proposition 1.2]). Hence, MM does not support Sol or Nil geometry.

Now, we are going to prove sufficiency. If MM supports geometry 𝔼3,S3,S2×\mathbb{E}^{3},S^{3},S^{2}\times\mathbb{R}, then π1(M)\pi_{1}(M) is virtually abelian so has property (QT). If the geometry of MM is 2×\mathbb{H}^{2}\times\mathbb{R} then MM is virtually covered by Σ×S1\Sigma\times S^{1} for some hyperbolic surface Σ\Sigma. Note that π1(Σ)\pi_{1}(\Sigma) is a residually finite hyperbolic group so it has property (QT) by [BBF19, Theorem 1.1]. Hence, π1(Σ)×\pi_{1}(\Sigma)\times\mathbb{Z} has property (QT). Since π1(Σ)×\pi_{1}(\Sigma)\times\mathbb{Z} is a finite index subgroup of π1(M)\pi_{1}(M), it follows that π1(M)\pi_{1}(M) has property (QT) by Lemma 2.3. If MM supports geometry 3\mathbb{H}^{3}, π1(M)\pi_{1}(M) is virtually compact special by deep theorems of Agol and Wise (see [Ago13] [Wis20]), thus π1(M)\pi_{1}(M) has property (QT).

Finally, we need to show that if MM supports SL(2,)~\widetilde{SL(2,\mathbb{R})} geometry then π1(M)\pi_{1}(M) has property (QT). To see this, by passing to a finite cover if necessary, we could assume that MM is a nontrivial circle bundle over a closed surface Σ\Sigma with χ(Σ)<0\chi(\Sigma)<0. Let 1Kπ1(M)π1(Σ)11\to K\to\pi_{1}(M)\to\pi_{1}(\Sigma)\to 1 be the short exact sequence associated with the circle bundle where KK is the normal cyclic subgroup of π1(M)\pi_{1}(M) generated by a fiber. Let π:π1(M)π1(Σ)\pi\colon\pi_{1}(M)\to\pi_{1}(\Sigma) be the surjective homomorphism in the above short exact sequence. Note that the short exact sequence does not split since MM is supporting SL(2,)~\widetilde{SL(2,\mathbb{R})} geometry. According to the first paragraph in the proof of [HRSS22, Corollary 4.3], there exists a generating set 𝒮\mathcal{S} of G=π1(M)G=\pi_{1}(M) so that :=Cay(G,𝒮)\mathcal{L}:=\operatorname{Cay}(G,\mathcal{S}) is a quasi-line. Moreover, the diagonal action of GG on π1(Σ)×\pi_{1}(\Sigma)\times\mathcal{L} is metrically proper and cobounded, and thus its orbital map is a quasi-isometry. Since π1(Σ)\pi_{1}(\Sigma) is a residually finite hyperbolic group, it follows from [BBF19] that π1(Σ)\pi_{1}(\Sigma) has property (QT). Hence there exists a finite product of quasi-trees i=1nTi\prod_{i=1}^{n}T_{i} such that π1(Σ)i=1nTi\pi_{1}(\Sigma)\curvearrowright\prod_{i=1}^{n}T_{i} so that its orbital map is a quasi-isometric embedding. It is easy to see that the orbital map of the diagonal action Gi=1nTi×G\curvearrowright\prod_{i=1}^{n}T_{i}\times\mathcal{L} of GG on the product i=1nTi×\prod_{i=1}^{n}T_{i}\times\mathcal{L} is a quasi-isometric embedding. Therefore π1(M)\pi_{1}(M) has property (QT). ∎

8.2. Property (QT) of nongeometric 3-manifolds

In this section, we are going to prove Theorem 1.2. Recall that a nongeometric 3-manifold is either a graph manifold or a mixed manifold.

8.2.1. property (QT) of graph manifolds

Let MM be a graph manifold. Since property (QT) is preserved undertaking finite index subgroups (see Lemma 2.3), we only need to show that a finite cover of MM has property (QT). By passing to a finite cover, we can assume that each Seifert fibered piece in the JSJ decomposition of MM is a trivial circle bundle over a hyperbolic surface of genus at least 22, and the intersection numbers of fibers of adjacent Seifert pieces have absolute value 11 (see[KL98, Lemma 2.1]). Also we can assume that the underlying graph of the graph manifold MM is bipartite since any non-bipartite graph manifold is double covered by a bipartite one.

We note that π1(M)\pi_{1}(M) is an admissible group in the sense of Definition 2.12. However, it is not always true that π1(M)\pi_{1}(M) can acts geometrically on a CAT(0) space so property (QT) in this case does not follow immediately from Theorem 1.3. Indeed, if MM is a graph manifold with nonempty boundary then it always admits a Riemannian metric of nonpositive curvature (see [Lee95]). In particular, π(M)M~\pi(M)\curvearrowright\tilde{M} is a CKA action, and thus property (QT) of π1(M)\pi_{1}(M) follows from Theorem 1.3. However, many closed graph manifolds are shown to not support any Riemannian metric of nonpositive curvature (see [Lee95]).

We remark here that the CAT(0) metric on the CKA space XX in Section 5 and Section 6 is not really essential in the proofs. Below we will make certain modifications on some steps to run the proof of Theorem 1.3 for closed graph manifolds.

Metrics on MM:

We now are going to describe a convenient metric on MM introduced by Kapovich–Leeb [KL98]. For each Seifert component Mv=Fv×S1M_{v}=F_{v}\times S^{1} of MM, we choose a hyperbolic metric on the base surface FvF_{v} so that all boundary components are totally geodesic of unit length, and then equip each Seifert component Mv=Fv×S1M_{v}=F_{v}\times S^{1} with the product metric dvd_{v} such that the fibers have length one. Metrics dvd_{v} on MvM_{v} induce the product metrics on M~v\tilde{M}_{v} which by abuse of notations is also denoted by dvd_{v}.

Let MvM_{v} and MwM_{w} be adjacent Seifert components in the closed graph manifold MM, and let TMvMwT\subset M_{v}\cap M_{w} be a JSJ torus. Each metric space (T~,dv)(\tilde{T},d_{v}) and (T~,dw)(\tilde{T},d_{w}) is a Euclidean plane. After applying a homotopy to the gluing map, we may assume that at each JSJ torus TT, the gluing map ϕ\phi from the boundary torus TMv\overleftarrow{T}\subset M_{v} to the boundary torus TMw\overrightarrow{T}\subset M_{w} is affine in the sense that the identity map (T~,dv)(T~,dw)(\tilde{T},d_{v})\to(\tilde{T},d_{w}) is affine. We now have a product metric on each Seifert component Mv=Fv×S1M_{v}=F_{v}\times S^{1}. These metrics may not agree on the JSJ tori but the gluing maps are bilipschitz (since they are affine). The product metrics on the Seifert components induce a length metric on the graph manifold MM denoted by dd (see [BBI01, Section 3.1]) for details). Moreover, there exists a positive constant LL such that on each Seifert component Mv=Fv×S1M_{v}=F_{v}\times S^{1} we have

1Ldv(x,y)d(x,y)Ldv(x,y)\frac{1}{L}\,d_{v}(x,y)\leq d(x,y)\leq L\,d_{v}(x,y)

for all xx and yy in MvM_{v}. (See [Pau05, Lemma 1.8] for detailed proof of the last claim.) Metric dd on MM induces metric on M~\tilde{M}, which is also denoted by dd (by abuse of notations). Then for all xx and yy in M~v\tilde{M}_{v} we have

1Ldv(x,y)d(x,y)Ldv(x,y)\frac{1}{L}\,d_{v}(x,y)\leq d(x,y)\leq L\,d_{v}(x,y)
Remark 8.2.

Note that the space (M~,d)(\tilde{M},d) may not be a CAT(0) space but π1(M)\pi_{1}(M) acts geometrically on (M~,d)(\tilde{M},d) via deck transformations.

In Section 2.3.2, we define special paths on a CAT(0) space XX. In this section, although (M~,d)(\tilde{M},d) is no longer a CAT(0) space, we are still able to define special paths in (M~,d)(\tilde{M},d). The construction is similar to Section 2.3.2 with slight changes.

Special paths on M~\tilde{M}:

Lift the JSJ decomposition of the graph manifold MM to the universal cover M~\tilde{M}, and let TT be the tree dual to this decomposition of M~\tilde{M} (i.e., the Bass-Serre tree of π1(M)\pi_{1}(M)). For every pair of adjacent edges e1e_{1}, e2e_{2} in TT, let vv be the common vertex of e1e_{1} and e2e_{2}. Let \ell and \ell^{\prime} be two boundary lines of F~v\tilde{F}_{v} corresponding to the edges e1e_{1} and e2e_{2} respectively. Let γe1e2\gamma_{e_{1}e_{2}} be the shortest geodesic joining \ell and \ell^{\prime} in (M~v,dv)(\tilde{M}_{v},d_{v}). This geodesic determines an Euclidean strip 𝒮e1e2:=γe1e2×\mathcal{S}_{e_{1}e_{2}}:=\gamma_{e_{1}e_{2}}\times\mathbb{R} in (M~v,dv)(\tilde{M}_{v},d_{v}). Let xx be a point in (M~v,dv)(\tilde{M}_{v},d_{v}) and ee be an edge with an endpoint vv. The minimal geodesic from xx to the plane FeF_{e} also define a strip 𝒮xe:=γxe×\mathcal{S}_{xe}:=\gamma_{xe}\times\mathbb{R} in (M~v,dv)(\tilde{M}_{v},d_{v}) where γxeF~v\gamma_{xe}\subset\tilde{F}_{v} is the projection to F~v\tilde{F}_{v} of this minimal geodesic.

Now, let xx and yy be any two points in the universal cover M~\tilde{M} of MM such that xx and yy belong to the interiors of pieces M~v\tilde{M}_{v} and M~v\tilde{M}_{v}^{\prime} respectively. If v=vv=v^{\prime} then we define a special path in XX connecting xx and yy to be the geodesic [x,y][x,y] in (M~,d)(\tilde{M},d). Otherwise, let e1ene_{1}\cdots e_{n} be the geodesic edge path connecting vv and vv^{\prime}. For notational purpose, we write e0:=xe_{0}:=x and en+1:=ye_{n+1}:=y. 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}}. The special path connecting xx and yy is the concatenation of the geodesics

[x,p1][p1,p2][pn,y][x,p_{1}]\cdot[p_{1},p_{2}]\cdots[p_{n},y]

We label p0:=xp_{0}:=x and pn+1:=yp_{n+1}:=y.

Proposition 8.3.

If MM is a graph manifold, then π1(M)\pi_{1}(M) has property (QT).

Proof.

If MM is a non-positively curved graph manifold (for example, when MM has nonempty boundary) then the fact π1(M)\pi_{1}(M) has property (QT) is followed from Theorem 1.3. The only case that does not follow directly from Theorem 1.3 is when MM is a closed graph manifold (recall many closed graph manifolds are non-positively curved but many are not). Since the metric dd on M~\tilde{M} restricted to each piece M~v\tilde{M}_{v} is LL–bilipschitz equivalent to dvd_{v}, so the inequalities in Section 6 are slightly changed by a uniform multiplicative constant. For example, the statement aKba\asymp_{K}b (or aKba\preceq_{K}b) in Section 6 will be changed to aKba\asymp_{K^{\prime}}b (or aKba\preceq_{K}^{\prime}b) for some constant KK^{\prime} depending on KK. Thus, the proof, in this case, is performed along lines with the proof of Theorem 1.3. ∎

8.2.2. property (QT) of mixed 3-manifolds

Recall that a non-geometric 3-manifold with empty or tori boundary is either a graph manifold or a mixed 3-manifold. The case of graph manifold has been addressed in Section 8.2.1. In this section, we address the mixed 3-manifold case.

Proposition 8.4.

The fundamental group of a mixed 3-manifold has property (QT).

The fundamental group of a mixed 3-manifold has a natural relatively hyperbolic structure as follows: Let M1,,MkM_{1},\cdots,M_{k} be the maximal graph manifold pieces, isolated Seifert fibered components of the JSJ-decomposition of MM, and S1,,SlS_{1},\cdots,S_{l} be the tori in MM not contained in any MiM_{i}. The fundamental group G=π1(M)G=\pi_{1}(M) is hyperbolic relative to the set of parabolic subgroups

𝒫={π1(Mp):1pk}{π1(Sq):1ql}\mathcal{P}=\{\pi_{1}(M_{p}):1\leq p\leq k\}\cup\{\pi_{1}(S_{q}):1\leq q\leq l\}

(see [BW13], [Dah03]).

The following lemma provides many separable subgroups in π1(M)\pi_{1}(M), generalizing [Sun21, Lemma 3.3]. The proof uses a recent result of the second author and Sun in [NS20] where the authors show that separability and distortion of subgroups in 3-manifold groups are closely related.

Lemma 8.5.

Let MM be a compact, orientable, irreducible, 3-manifold with empty or tori boundary, with nontrivial torus decomposition and does not support the Sol geometry. If HH is a finitely generated, undistorted subgroup of π1(M)\pi_{1}(M), then HH is separable in π1(M)\pi_{1}(M).

Proof.

Let MHM_{H} be the covering space of MM corresponding to Hπ1(M)H\leq\pi_{1}(M). Generalizing a notion called “almost fiber part” in [Liu17], an embedded (possibly disconnected) subsurface Φ(H)\Phi(H) in MHM_{H} called “almost fiber surface” is introduced in [Sun20]. Sun shows in [Sun20, Theorem 1.3] that all information about the separability of HH can be obtained by examining the almost fibered surface.

In [NS20], the authors introduce a notion called “modified almost fibered surface” (denoted Φ^(H)\hat{\Phi}(H)) that modify slightly the original definition of almost fibered surface in [Sun20] and show that information about the distortion of HH in GG can be also obtained by examining the “modified almost fibered surface”. We refer the reader to [Sun20] for the definition of the almost fiber surface and to [NS20] for the definition of the modified almost fiber surface. The precise definitions are not needed here, so we only state here some facts from [NS20] that will be used later in the proof.

The torus decomposition of MM induces the torus decomposition of Φ(H)\Phi(H). Let Φ(H)\Phi(H) and Φ^(H)\hat{\Phi}(H) be the almost fiber surface and modified almost fiber surface of HH respectively.

  1. (1)

    Both the almost fiber surface Φ(H)\Phi(H) and the modified almost fiber surface Φ^(H)\hat{\Phi}(H) are (possibly disconnected) subsurfaces of MHM_{H}.

  2. (2)

    The almost fiber surface Φ(H)\Phi(H) has some piece that is homeomorphic to the annulus and parallel to the boundary of Φ(H)\Phi(H). We delete these annulus pieces from Φ(H)\Phi(H) to get the modified almost fiber surface, and we denote it by Φ^(H)\hat{\Phi}(H).

The surface Φ(H)\Phi(H) (resp. Φ^(H)\hat{\Phi}(H)) has a natural graph of spaces structure with the dual graph denoted by ΓΦ(H)\Gamma_{\Phi(H)} (resp. ΓΦ^(H)\Gamma_{\hat{\Phi}(H)}). By [NS20, Theorem 1.4], every component SS of the modified almost fiber surface Φ^(H)\hat{\Phi}(H) must contain only one piece (otherwise, the distortion of HH in π1(M)\pi_{1}(M) is at least quadratic, this contradicts the fact that HH is undistorted in π1(M)\pi_{1}(M)). This fact combined with (2) implies that the graph ΓΦ(H)\Gamma_{\Phi(H)} is a union of trees. By [Sun20, Theorem 1.3] (or see also [Sun21, Theorem 3.2] for a statement) tells us that whenever ΓΦ(H)\Gamma_{\Phi(H)} does not contain a simple cycle then HH is separable. As shown above, we are in this case, hence we conclude that the subgroup HH is separable in π1(M)\pi_{1}(M). ∎

We now give the proof of Proposition 8.4.

Proof of Proposition 8.4.

Let M1,,MkM_{1},\cdots,M_{k} be the collection of maximal graph manifold components and Seifert fibered pieces in the geometric decomposition of MM. Let S1,,SS_{1},\dots,S_{\ell} be the tori in the boundary of MM that bound a hyperbolic piece, and let T1,,TmT_{1},\dots,T_{m} be the tori in the JSJ decomposition of MM that separate two hyperbolic components of the JSJ decomposition. Then π1(M)\pi_{1}(M) is hyperbolic relative to

={π1(Mp)}p=1k{π1(Sq)}q=1{π1(Tr)}r=1m\mathbb{P}=\{\pi_{1}(M_{p})\}_{p=1}^{k}\cup\{\pi_{1}(S_{q})\}_{q=1}^{\ell}\cup\{\pi_{1}(T_{r})\}_{r=1}^{m}

(see [BW13], [Dah03]).

We are going to show that G=π1(M)G=\pi_{1}(M) satisfies all conditions in Theorem 1.5.

Claim 1: π1(M)\pi_{1}(M) induces the full profinite topology on each P𝒫P\in\mathcal{P}. Indeed, it is well-known that the fundamental groups of all compact 3-manifolds are residually finite, thus π1(M)\pi_{1}(M) is residually finite. Since each peripheral subgroup PP is undistorted in π1(M)\pi_{1}(M), it follows from Lemma 8.5 that PP is separable in π1(M)\pi_{1}(M). Again, by Lemma 8.5, each finite index subgroup of PP is also separable in π1(M)\pi_{1}(M). By [Rei18, Lemma 2.8], π1(M)\pi_{1}(M) induces the full profinite topology on PP.

Claim 2: For each peripheral subgroup PP\in\mathbb{P}, there exists a finite index subgroup PP^{\prime} of PP acting isometrically on a finite number of quasi-trees so that the diagonal action of PP^{\prime} on the finite product of these quasi-trees induces quasi-isometric embeddings on orbital maps. Indeed, if P=π1(Tr)P=\pi_{1}(T_{r}) or P=π1(Sq)P=\pi_{1}(S_{q}) for some r,qr,q then π1(P)=2\pi_{1}(P)=\mathbb{Z}^{2}, we denote P:=PP^{\prime}:=P. If P=π1(Mj)P=\pi_{1}(M_{j}) for some Seifert component Mj=Fj×S1M_{j}=F_{j}\times S^{1} then P=π1(Fj)×P=\pi_{1}(F_{j})\times\mathbb{Z}. In this case, as FjF_{j} is a hyperbolic surface with nonempty boundary, we have π1(Fj)\pi_{1}(F_{j}) is a free group, hence we choose P=PP^{\prime}=P as π1(Fj)\pi_{1}(F_{j}) is a quasi-tree. The last case we must consider is that P=π1(Mj)P=\pi_{1}(M_{j}) where MjM_{j} is a maximal graph manifold component. Passing to an appropriate finite cover MjMjM^{\prime}_{j}\to M_{j} we could assume that π1(Mj)\pi_{1}(M^{\prime}_{j}) acts on a finite number of quasi-trees (but they are not quasi-lines) T1,T2,,TnT_{1},T_{2},\cdots,T_{n}’s so that the orbital map induced from the diagonal action π1(Mj)1=1nTi\pi_{1}(M_{j})\curvearrowright\prod_{1=1}^{n}T_{i} is a quasi-isometric embedding (see Proposition 8.3). Claim 2 is confirmed. We then repeat the proof of Theorem 3.5 (the second and third paragraph) to show that PP satisfies the hypothesis of Theorem 1.5.

In summary, we have verified the hypotheses in Theorem 1.5 for G=π1(M)G=\pi_{1}(M), so mixed 3-manifold groups have property (QT). ∎

We now give the proof of Theorem 1.2.

Proof of Theorem 1.2.

Let MM be a compact orientable irreducible 3-manifold with empty or tori boundary, with nontrivial torus decomposition, and that does not support the Sol geometry. Such a 3-manifold MM is either a graph manifold or a mixed manifold. The graph manifold case and mixed manifold case have been addressed in Proposition 8.3 and Proposition 8.4 respectively, and hence the theorem is proved. ∎

8.3. Property (QT) of finitely generated 3-manifolds

Proposition 8.6.

Let MM be a compact, orientable, irreducible, \partial–irreducible 3-manifold such that it has a boundary component of genus at least 22. Then π1(M)\pi_{1}(M) has property (QT).

Proof.

We consider the following two cases:

Case 1: MM has trivial torus decomposition. In this case, MM supports a geometrically finite hyperbolic structure with infinite volume. We paste hyperbolic 3-manifolds with totally geodesic boundaries to MM to get a finite volume hyperbolic 3-manifold NN. By the Covering Theorem (see [Can96]) and the Subgroup Tameness Theorem (see [Ago04], [CG06]), a finitely generated subgroup of the finite volume hyperbolic 3-manifold NN is either a virtual fiber surface subgroup or undistorted. By the construction of NN, the subgroup π1(M)π1(N)\pi_{1}(M)\leq\pi_{1}(N) could not be a virtual fiber surface subgroup, and thus π1(M)\pi_{1}(M) must be undistorted in π1(N)\pi_{1}(N). Since π1(N)\pi_{1}(N) has property (QT), it follows that π1(M)\pi_{1}(M) has property (QT) (see Lemma 2.3).

Case 2: We now assume that MM has nontrivial torus decomposition. By [Sun20, Section 6.3], we paste hyperbolic 3-manifolds with totally geodesic boundaries to MM to get a 3-manifold NN with empty or tori boundary. The new manifold NN satisfies the following properties.

  1. (1)

    MM is a submanifold of NN with incompressible tori boundary.

  2. (2)

    The torus decomposition of MM also gives the torus decomposition of NN.

  3. (3)

    Each piece of MM with a boundary component of genus at least 22 is contained in a hyperbolic piece of NN.

In particular, it follows from (2) and (3) that NN is a mixed 3-manifold. The subgroup π1(M)\pi_{1}(M) sits nicely in π1(N)\pi_{1}(N). By the proof of Case 1.2 in the proof of [NS20, Theorem 1.3]), we have that π1(M)\pi_{1}(M) is undistorted in π1(N)\pi_{1}(N) (even more than that, π1(M)\pi_{1}(M) is strongly quasiconvex in π1(N)\pi_{1}(N) (see [NTY21]). Note that π1(N)\pi_{1}(N) has property (QT) by Proposition 8.4. Since π1(M)\pi_{1}(M) is undistorted in π1(N)\pi_{1}(N) and π1(N)\pi_{1}(N) has property (QT), it follows that π1(M)\pi_{1}(M) has property (QT). ∎

We now give the proof of Theorem 1.1 which gives a complete characterization of property (QT) for finitely generated 3-manifolds groups.

Proof of Theorem 1.1.

Since MM is a compact, orientable 33–manifold, it decomposes into irreducible, \partial–irreducible pieces M1,,MkM_{1},\dots,M_{k} by the sphere-disc decomposition. In particular, π1(M)\pi_{1}(M) is the free product π1(M1)π1(M2)π1(Mk)Fr\pi_{1}(M_{1})*\pi_{1}(M_{2})*\cdots*\pi_{1}(M_{k})*F_{r} for some free group FrF_{r}. We remark here that π1(M)\pi_{1}(M) is hyperbolic relative to the collection ={P1,,Pk,Fr}\mathbb{P}=\{P_{1},\cdots,P_{k},F_{r}\} where Pi:=π1(Mi)P_{i}:=\pi_{1}(M_{i}).

We are going to prove the necessity. Assume that π1(M)\pi_{1}(M) has property (QT). Since π1(Mi)\pi_{1}(M_{i}) is undistorted in π1(M)\pi_{1}(M), it follows that π1(Mi)\pi_{1}(M_{i}) has property (QT) (see Lemma 2.3). By Proposition 8.1, MiM_{i} does not support Sol and Nil geometry.

Now, we are going to prove sufficiency. Assume that there is no piece MiM_{i} that supports either Sol or Nil geometry. We would like to show that π1(M)\pi_{1}(M) has property (QT). In this case, we observe that each peripheral subgroup PP\in\mathbb{P} has property (QT). Indeed, a free group P=FrP=F_{r} of course has property (QT), so let us now assume that P=π1(Mi)P=\pi_{1}(M_{i}) for some 1ik1\leq i\leq k. If MiM_{i} has a boundary component of genus at least 22 then property (QT) of π1(Mi)\pi_{1}(M_{i}) follows from Proposition 8.6. Otherwise, MiM_{i} has empty or tori boundary. Then the property (QT) of π1(Mi)\pi_{1}(M_{i}) follows from Proposition 8.1 for geometric manifolds, Proposition 8.3 for graph manifolds, and Proposition 8.4 for mixed graph manifolds.

We are going to show that G=π1(M)G=\pi_{1}(M) satisfies all conditions in Theorem 1.5. The proof is similar to the proof of Proposition 8.4 with minor changes.

Claim 1: π1(M)\pi_{1}(M) induces the full profinite topology on each Pi𝒫P_{i}\in\mathcal{P}. It is well-known that the fundamental groups of all compact 3-manifolds are residually finite, thus π1(M)\pi_{1}(M) is residually finite and its finite index subgroups are residually finite as well. Any finite index subgroup HH of Pi=π1(Mi)P_{i}=\pi_{1}(M_{i}) is separable in the free product G=P1P2PkFrG=P_{1}*P_{2}*\cdots*P_{k}*F_{r} by [Bur71, Theorem 1.1]. Hence it follows from [Rei18, Lemma 2.8 ] that GG induces the full profinite topology on PiP_{i}.

Claim 2: For each peripheral subgroup PP\in\mathbb{P}, there exists a finite index subgroup PP^{\prime} of PP acting isometrically on a finite number of quasi-trees, so that the diagonal action of PP^{\prime} on the finite product of these quasi-trees induces quasi-isometric embeddings on orbital maps. Indeed, the claim obviously holds for P=FrP=F_{r} or P=2P=\mathbb{Z}^{2}. The claim also holds for P=π1(Mi)P=\pi_{1}(M_{i}) where MiM_{i} is a geometric 3-manifold. The case of graph manifolds is proved in the Claim 2 of the proof of Proposition 8.4. The only case left is when MiM_{i} is a mixed 3-manifold or MiM_{i} has a boundary component with genus at least 22. It has been shown in Proposition 8.6 that if MiM_{i} has a boundary component with genus at least 22 then it is an undistorted subgroup in a mixed 3-manifold. Therefore it suffices to consider only the mixed 3-manifold case. Recall that in the proof of Proposition 8.4, we show that there exists a finite index subgroup of π1(Mi)\pi_{1}(M_{i}) such that it is a relatively hyperbolic group, satisfying conditions in Theorem 1.5, and thus Claim 2 is confirmed.

With Claim 1 and Claim 2, we use the same argument as in the proof of Theorem 3.5 (see the second and third paragraph) to find a finite index normal subgroup GG^{\prime} of GG such that GG^{\prime} is hyperbolic relative to a collection of subgroups satisfying the hypotheses in Theorem 1.5, and thus GG^{\prime} has property (QT). Therefore, π1(M)\pi_{1}(M) has property (QT) since GG^{\prime} is a finite index subgroup of π1(M)\pi_{1}(M) and GG^{\prime} does have property (QT). ∎

References

  • [ABO19] Carolyn Abbott, Sahana H. Balasubramanya, and Denis Osin. Hyperbolic structures on groups. Algebr. Geom. Topol., 19(4):1747–1835, 2019.
  • [Ago04] Ian Agol. Tameness of hyperbolic 3-manifolds. arXiv Mathematics e-prints, page math/0405568, May 2004.
  • [Ago13] I. Agol. The virtual Haken conjecture. Doc. Math., 18:1045–1087, 2013. With an appendix by Agol, Daniel Groves, and Jason Manning.
  • [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.
  • [BBFS19] Mladen Bestvina, Ken Bromberg, Koji Fujiwara, and Alessandro Sisto. Acylindrical actions on projection complexes. Enseign. Math., 65(1-2):1–32, 2019.
  • [BBI01] D. Burago, Y. Burago, and S. Ivanov. A course in metric geometry, volume 33 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2001.
  • [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.
  • [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.
  • [Bur71] R. G. Burns. On finitely generated subgroups of free products. J. Austral. Math. Soc., 12:358–364, 1971.
  • [But20] J. O. Button. Groups acting purely loxodromically on products of hyperbolic graphs, 2020.
  • [BW13] Hadi Bigdely and Daniel T. Wise. Quasiconvexity and relatively hyperbolic groups that split. Michigan Math. J., 62(2):387–406, 2013.
  • [Can96] Richard D. Canary. A covering theorem for hyperbolic 33-manifolds and its applications. Topology, 35(3):751–778, 1996.
  • [CCMT15] Pierre-Emmanuel Caprace, Yves Cornulier, Nicolas Monod, and Romain Tessera. Amenable hyperbolic groups. J. Eur. Math. Soc. (JEMS), 17(11):2903–2947, 2015.
  • [CG06] Danny Calegari and David Gabai. Shrinkwrapping and the taming of hyperbolic 3-manifolds. J. Amer. Math. Soc., 19(2):385–446, 2006.
  • [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.
  • [Con00] Gregory R. Conner. Discreteness properties of translation numbers in solvable groups. J. Group Theory, 3(1):77–94, 2000.
  • [Dah03] François Dahmani. Combination of convergence groups. Geom. Topol., 7:933–963, 2003.
  • [dC08] Yves de Cornulier. Dimension of asymptotic cones of Lie groups. J. Topol., 1(2):342–361, 2008.
  • [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.
  • [FL08] Thomas Foertsch and Alexander Lytchak. The de Rham decomposition theorem for metric spaces. Geom. Funct. Anal., 18(1):120–143, 2008.
  • [GP16] Victor Gerasimov and Leonid Potyagailo. Quasiconvexity in relatively hyperbolic groups. J. Reine Angew. Math., 710:95–135, 2016.
  • [Ham01] Emily Hamilton. Abelian subgroup separability of Haken 3-manifolds and closed hyperbolic nn-orbifolds. Proc. London Math. Soc. (3), 83(3):626–646, 2001.
  • [HO13] Michael Hull and Denis Osin. Induced quasicocycles on groups with hyperbolically embedded subgroups. Algebr. Geom. Topol., 13(5):2635–2665, 2013.
  • [HP15] Mark F. Hagen and Piotr Przytycki. Cocompactly cubulated graph manifolds. Israel J. Math., 207(1):377–394, 2015.
  • [HRSS22] Mark Hagen, Jacob Russell, Alessandro Sisto, and Davide Spriano. Equivariant hierarchically hyperbolic structures for 3-manifold groups via quasimorphisms, 2022.
  • [HW08] Frédéric Haglund and Daniel T. Wise. Special cube complexes. Geom. Funct. Anal., 17(5):1551–1620, 2008.
  • [KL96] Michael Kapovich and Bernhard Leeb. Actions of discrete groups on nonpositively curved spaces. Math. Ann., 306(2):341–352, 1996.
  • [KL98] M. Kapovich and B. Leeb. 33-manifold groups and nonpositive curvature. Geom. Funct. Anal., 8(5):841–852, 1998.
  • [Lee95] Bernhard Leeb. 33-manifolds with(out) metrics of nonpositive curvature. Invent. Math., 122(2):277–289, 1995.
  • [Liu17] Yi Liu. A characterization of virtually embedded subsurfaces in 3-manifolds. Trans. Amer. Math. Soc., 369(2):1237–1264, 2017.
  • [Man05] Jason Fox Manning. Geometry of pseudocharacters. Geom. Topol., 9:1147–1185, 2005.
  • [Man06] J. F. Manning. Quasi-actions on trees and property (QFA). J. London Math. Soc. (2), 73(1):84–108, 2006. With an appendix by N. Monod and B. Rémy.
  • [MM00] H. A. Masur and Y. N. Minsky. Geometry of the complex of curves. II. Hierarchical structure. Geom. Funct. Anal., 10(4):902–974, 2000.
  • [MS13] John M. Mackay and Alessandro Sisto. Embedding relatively hyperbolic groups in products of trees. Algebr. Geom. Topol., 13(4):2261–2282, 2013.
  • [NS20] Hoang Thanh Nguyen and Hongbin Sun. Subgroup distortion of 3-manifold groups. Trans. Amer. Math. Soc., 373(9):6683–6711, 2020.
  • [NTY21] Hoang Thanh Nguyen, Hung Cong Tran, and Wenyuan Yang. Quasiconvexity in 3-manifold groups. Math. Ann., 381(1-2):405–437, 2021.
  • [NY] Hoang Thanh Nguyen and Wenyuan Yang. Croke-kleiner admissible groups: Property (qt) and quasiconvexity. To appear in Michigan Math. J.
  • [Osi16] D. Osin. Acylindrically hyperbolic groups. Trans. Amer. Math. Soc., 368(2):851–888, 2016.
  • [Pau05] G. Paulik. Gluing spaces and analysis. Bonner Mathematische Schriften [Bonn Mathematical Publications], 372. Universität Bonn, Mathematisches Institut, Bonn, 2005. Dissertation, Rheinische Friedrich-Wilhelms-Universität Bonn, Bonn, 2005.
  • [PW18] Piotr Przytycki and Daniel T. Wise. Mixed 3-manifolds are virtually special. J. Amer. Math. Soc., 31(2):319–347, 2018.
  • [Rei18] Alan W. Reid. Profinite rigidity. In Proceedings of the International Congress of Mathematicians—Rio de Janeiro 2018. Vol. II. Invited lectures, pages 1193–1216. World Sci. Publ., Hackensack, NJ, 2018.
  • [Sis13] Alessandro Sisto. Projections and relative hyperbolicity. Enseign. Math. (2), 59(1-2):165–181, 2013.
  • [Sun20] Hongbin Sun. A characterization on separable subgroups of 3-manifold groups. J. Topol., 13(1):187–236, 2020.
  • [Sun21] Hongbin Sun. All finitely generated 3-manifold groups are Grothendieck rigid. arXiv e-prints, page arXiv:2103.00547, February 2021.
  • [Tid18] Joseph Tidmore. Cocompact cubulations of mixed 3-manifolds. Groups Geom. Dyn., 12(4):1429–1460, 2018.
  • [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.