Croke-Kleiner admissible groups: Property (QT) and quasiconvexity
Abstract.
Croke-Kleiner admissible groups firstly introduced by Croke-Kleiner in [CK02] belong to a particular class of graph of groups which generalize fundamental groups of –dimensional graph manifolds. In this paper, we show that if is a Croke-Kleiner admissible group, acting geometrically on a CAT(0) space , then a finitely generated subgroup of has finite height if and only if it is strongly quasi-convex. We also show that if is a flip CKA action then is quasi-isometric embedded into a finite product of quasi-trees. With further assumption on the vertex groups of the flip CKA action , we show that satisfies property (QT) that is introduced by Bestvina-Bromberg-Fujiwara in [BBF19].
1. Introduction
In [CK02], Croke and Kleiner study a particular class of graph of groups which they call admissible groups and generalize fundamental groups of –dimensional graph manifolds and torus complexes (see [CK00]). If is an admissible group that acts geometrically on a Hadamard space then the action is called Croke-Kleiner admissibe (see Definition 2.1) termed by Guilbault-Mooney [GM14]. The CKA action is modeling on the JSJ structure of graph manifolds where the Seifert fibration is replaced by the following central extension of a general hyperbolic group:
(1) |
However, CKA groups can encompass much more general class of groups and can actually serve as one of simplest algebraic means to produce interesting groups from any finite number of hyperbolic groups.
Let be a finite graph with vertices, each of which are associated with a hyperbolic group . We then pick up an independent set of primitive loxodromic elements in which crossed with are the edge groups . We identify in adjacent ’s by flipping and loxodromic elements as did in flip graph manifolds by Kapovich and Leeb [KL98]. These are motivating examples of flip CKA groups and actions, for the precise definition of flip CKA actions, we refer the reader to Section 4.2.
The class of CKA actions has manifested a variety of interesting features in CAT(0) groups. For instance, the equivariant visual boundaries of admissible actions are completely determined in [CK02]. Meanwhile, the non-homeomorphic visual boundaries of torus complexes were constructed in [CK00] and have sparked an intensive research on boundaries of CAT(0) spaces. So far, the most of research on CKA groups is centered around the boundary problem (see [GM14], [Gre16]). In the rest of Introduction, we shall explain our results on the coarse geometry of Croke-Kleiner admissible groups and their subgroups.
1.1. Proper actions on finite products of quasi-trees
A quasi-tree is a geodesic metric space quasi-isometric to a tree. Recently, Bestvina, Bromberg and Fujiwara [BBF19] introduced a (QT) property for a finitely generated group: acts properly on a finite product of quasi-trees so that the orbital map from with word metrics is a quasi-isometric embedding. This is a stronger property of the finite asymptotic dimension by recalling that a quasi-isometric embedding implies finite asdim of . It is known that Coxeter groups have property (QT) (see [DJ99]), and thus every right-angled Artin group has property (QT) (see Induction 2.2 in [BBF19]). Furthermore, the fundamental group of a compact special cube complex is undistorted in RAAGS (see [HW08]) and then has property (QT). As a consequence, many 3-manifold groups have property (QT), among which we wish to mention chargeless (including flip) graph manifolds [HP15] and finite volume hyperbolic 3-manifolds [Wis20]. In [BBF19], residually finite hyperbolic groups and mapping class groups are proven to have property (QT). It is natural to ask which other groups have property (QT) rather than these groups above.
The main result of this paper adds flip CKA actions into the list of groups which have property (QT). The notion of an omnipotent group is introduced by Wise in [Wis00] and has found many applications in subgroup separability. We refer the reader to Definition 5.10 for its definition and note here that free groups [Wis00], surfaces groups [Baj07], and the more general class of virtually special hyperbolic groups [Wis20] are omnipotent.
Theorem 1.1.
Let be a flip admissible action where for every vertex group the central extension (1) has omnipotent hyperbolic quotient group. Then acts properly on a finite product of quasi-trees so that the orbital map is a quasi-isometric embedding.
Remark 1.2.
It is an open problem whether every hyperbolic group is residually finite. In [Wis00, Remark 3.4], Wise noted that if every hyperbolic group is residually finite, then any hyperbolic group is omnipotent.
Remark 1.3.
As a corollary, Theorem 1.1 gives another proof that flip graph manifold groups have property (QT). This was indeed one of motivations of this study (without noticing [HP15]).
In [HS13], Hume-Sisto prove that the universal cover of any flip graph manifold is quasi-isometrically embedded in the product of three metric trees. However, it does not follow from their proof that the fundamental group of a flip graph manifold has property (QT).
We now give an outline of the proof of Theorem 1.1 and explain some intermediate results, which we believe are of independent interest.
Proposition 1.4.
Let be a flip CKA action. Then there exists a quasi-isometric embedding from to a product of two hyperbolic spaces.
If for every vertex and for every edge , then there exists a subgroup of finite index at most such that the above Q.I. embedding is -equivariant.
Let us describe briefly the construction of . By Bass-Serre theory, acts on the Bass-Serre tree with vertex groups and edge groups . Let be one of the two sets of vertices in with pairwise even distance. Note that is the central extension of a hyperbolic group by , so acts geometrically on a metric product where acts geometrically on and acts by translation on -lines. Roughly, the space is obtained by isometric gluing of the boundary lines of ’s over vertices in the link of every . In proving Proposition 1.4, the main tool is the construction of a class of quasi-geodesic paths called special paths between any two points in . See Section 3 for the details and related discussion after Theorem 1.6 below.
To endow an action on , we pass to an index at most 2 subgroup preserving and the stabilizer in of is by Lemma 4.6. Under the assumptions on and ’s, acts by isometry on and the Q.I. embedding is -equivariant.
To prove Theorem 1.1, we exploit the strategy as [BBF19] to produce a proper action on products of quasi-trees. By Lemma 4.17, we first produce enough quasi-lines
for the hyperbolic space where is a -finite set of quasi-lines in so that the so-called distance formula follows in Proposition 4.20. On the other hand, the “crowd” quasi-lines in may fail to satisfy the projection axioms in [BBF15] with projection constant required from the distance formula. Thus, we have to partition into finite sub-collections of sparse quasi-lines: , for a uniform constant .
Using local finiteness, we can partition quasi-lines without respecting group action and prove the following general result. This generalize the results of Hume-Sisto in [HS13] to flip CKA actions.
Theorem 1.5.
Let be a flip CKA action. Then is quasi-isometric embedded into a finite product of quasi-trees.
However, the difficulty in establishing property (QT) is to partition all quasi-lines so that each is -invariant and sparse. In §5.1, we cone off the boundary lines of ’s so that ’s from different pieces are “isolated”. The gives the coned-off space with new distance formula in Proposition 5.9 so that is quasi-isometric embedded into the product of with a quasi-tree from the boundary lines. See Proposition 5.7.
The goal is then to find a finite index subgroup so that each orbit in is sparse. This is done in the following two steps:
By residual finiteness of , we first find a finite index subgroup whose orbit in is sparse. This follows the same argument in [BBF19]. Secondly, we need to reassemble those finite index subgroups as a finite index group so that the orbit in is sparse. This step uses crucially the omnipotence, with details given in §5.4. The projection axioms thus fulfilled for each -orbit produce a finite product of actions on quasi-trees, and finally, the distance formula finishes the proof of Theorem 1.1.
1.2. Strongly quasi-convex subgroups
The height of a finitely generated subgroup in a finitely generated group is the maximal such that there are distinct cosets such that the intersection is infinite. The subgroup is called strongly quasi-convex in if for any , there exists such that every –quasi-geodesic in with endpoints in is contained in the –neighborhood of . We note that strong quasiconvexity does not depend on the choice of finite generating set of the ambient group and it agrees with quasiconvexity when the ambient group is hyperbolic. In [GMRS98], the authors prove that quasi-convex subgroups in hyperbolic groups have finite height. It is a long-standing question asked by Swarup that whether or not the converse is true (see Question 1.8 in [Bes]). Tran in [Tra19] generalizes the result of [GMRS98] by showing that strongly quasi-convex subgroups in any finitely generated group have finite height. It is natural to ask whether or not the converse is true in this setting (i.e, finite height implies strong quasiconvexity). If the converse is true, then we could characterize strongly quasi-convex subgroup of a finitely generated group purely in terms of group theoretic notions.
In [NTY], the authors prove that having finite height and strong quasiconvexity are equivalent for all finitely generated –manifold groups except the only ones containing the Sol command in its sphere-disk decomposition, and the graph manifold case was an essential case treated there. More precisely, Theorem 1.7 in [NTY] states that finitely generated subgroups of the fundamental group of a graph manifold are strongly quasi-convex if and only if they have finite height. The second main result of this paper is to generalize this result to Croke-Kleiner admissible action .
Theorem 1.6.
Let be a CKA action. Let be a nontrivial, finitely generated infinite index subgroup of . Then the following are equivalent.
-
(1)
is strongly quasi-convex.
-
(2)
has finite height in .
-
(3)
is virtually free and every infinite order elements are Morse.
-
(4)
Let be the action of on the associated Bass-Serre tree. is virtually free and the action of on the tree induces a quasi-isometric embedding of into .
We prove Theorem 1.6 by showing that and . Similarly as in [NTY], the heart part of Theorem 1.6 is the implication . We briefly review ideas in the proof of Theorem 1.7 in [NTY]. Suppose that is a finitely generated finite height subgroup of where is a graph manifold. Let be the covering space of corresponding to . The authors in [NTY] prove that is strongly quasi-convex in by using Sisto’s notion of path system in the universal cover of , and prove that the preimage of the Scott core of in is –contracting in the sense of Sisto. In this paper, the strategy of the proof of Theorem 1.6 is similar to the proof of Theorem 1.7 in [NTY] where we still use Sisto’s path system in but details are different. Sisto’s construction of special paths are carried out only in flip graph manifolds. Our construction of relies on the work of Croke-Kleiner [CK02] and applies to any admissible space (so any nonpostively curved graph manifold). We then construct a subspace on which acts geometrically and show that is contracting in with respect to the path system . As a consequence, is strongly quasi-convex in .
To conclude the introduction, we list a few questions and problems.
Quasi-isometric classification of graph manifolds has been studied by Kapovich-Leeb [KL98] and a complete quasi-isometric classification for fundamental groups of graph manifolds is given by Behrstock-Neumann [BN08]. Kapovich-Leeb prove that for any graph manifold , there exists a flip graph manifold such that their fundamental groups are quasi-isometric. We would like to know that whether or not such a result holds for admissible groups.
Question 1.7.
Let be an admissible group such that each vertex group is the central extension of a omnipotent hyperbolic CAT(0) group by . Does there exist flip CKA action so that and are quasi-isometric?
Question 1.8 (Quasi-isometry rigidity).
Let be a flip CKA action, and be a finitely generated group which is quasi-isometric to . Does there exist a finite index subgroup such that is a flip CKA group?
With a positive answer to the above questions, we hope one can try to follow the strategy described in [BN08] to attack the following.
Problem 1.9.
Under the assumption of Theorem 1.1, give a quasi-isometric classification of admissible actions.
In [Liu13], Liu showed that the fundamental group of a non-positively curved graph manifold is virtually special (the case was also obtained independently by Przytycki–Wise [PW14]). Thus, it is natural to ask the following.
Question 1.10.
Let be a CKA action where vertex groups are the central extension of a virtually special hyperbolic group by the integer group. Is virtually special?
As above, a positive answer to the question (with virtual compact specialness) would give an other proof of Theorem 1.1 under the same assumption.
Overview
In Section 2, we review some concepts and results about Croke-Kleiner admissible groups. In Section 3, we construct special paths in admissible spaces and give some results that will be used in the later sections. The proof of Theorem 1.5 and Proposition 1.4 is given in Section 4. We prove Theorem 1.1 and Theorem 1.6 in Section 5 and Section 6 respectively.
Acknowledgments
We would like to thank Chris Hruska, Hongbin Sun and Dani Wise for helpful conversations. W. Y. is supported by the National Natural Science Foundation of China (No. 11771022).
2. Preliminary
Admissible groups firstly introduced in [CK02]. This is a particular class of graph of groups that includes fundamental groups of –dimensional graph manifolds (i.e, compact –manifolds are obtained by gluing some circle bundles). In this section, we review admissible groups and their properties that will used throughout in this paper.
Definition 2.1.
A graph of group is admissible if
-
(1)
is a finite graph with at least one edge.
-
(2)
Each vertex group has center , is a non-elementary hyperbolic group, and every edge subgroup is isomorphic to .
-
(3)
Let and be distinct directed edges entering a vertex , and for , let be the image of the edge homomorphism . Then for every , is not commensurable with , and for every , is not commensurable with .
-
(4)
For every edge group , if are the edge monomorphism, then the subgroup generated by and has finite index in .
A group is admissible if it is the fundamental group of an admissible graph of groups.
Definition 2.2.
We say that the action is an Croke-Kleiner admissible (CKA) if is an admissible group, and is a Hadamard space, and the action is geometrically (i.e, properly and cocompactly by isometries)
Examples of admissible actions:
-
(1)
Let be a nongeometric graph manifold that admits a nonpositively curve metric. Lift this metric to the universal cover of , and we denote this metric by . Then the action is a CKA action.
-
(2)
Let be the torus complexes constructed in [CK00]. Then is a CKA action.
-
(3)
Let and be two torsion-free hyperbolic groups such that they act geometrically on spaces and respectively. Let (with ), then acts geometrically on the space . A primitive hyperbolic element in gives a totally geodesic torus in the quotient space . Choose a basis on each torus . Let be a flip map. Let be the space obtained by gluing to along the homemorphism . We note that there exists a metric on such that with respect to this metric, is a locally space. Then is a CKA action.
Let be an admissible action, and let be the action of on the associated Bass-Serre tree. Let and be the vertex and edge sets of . For each , we let be the stabilizer of . For each vertex , let and for every edge we let . We note that the assignments and are –equivariant with respect to the natural actions.
The following lemma is well-known.
Lemma 2.3.
If for some then splits isometrically as a metric product so that acts trivially on and as a translation lattice on . Moreover, acts cocompactly on .
As a corollary, we have
-
(1)
acts co-compactly on and acts by translation on the –factor and trivially on where is a Hadamard space.
-
(2)
acts co-compactly on where is a compact Hadamard space.
-
(3)
if then generates a finite index subgroup of .
We summarize results in Section 3.2 of [CK02] that will be used in this paper.
Lemma 2.4.
Let be an CKA action. Then there exists a constant such that the following holds.
-
(1)
. We define and for all , .
-
(2)
If and then .
Strips in admissible spaces: (see Section 4.2 in [CK02]). We first choose, in a –equivariant way, a plane for each edge . Then for every pair of adjacent edges , . we choose, again equivariantly, a minimal geodesic from to ; by the convexity of , , this geodesic determines a Euclidean strip (possibly of width zero) for some geodesic segment . Note that is an axis of . Hence if , are distinct vertices, then the angle between the geodesics and is bounded away from zero.
Remark 2.5.
-
(1)
We note that it is possible that is just a point. The lines and are axes of .
-
(2)
There exists a uniform constant such that for any edge , the Hausdorff distance between two spaces and is no more than this constant.
Remark 2.6.
There exists a –equivariant coarse –Lipschitz map such that for all . The map is called indexed map. We refer the reader to Section 3.3 in [CK02] for existence of such a map .
Definition 2.7 (Templates, [CK02]).
A template is a connected Hadamard space obtained from disjoint collection of Euclidean planes (called walls) and directed Euclidean strips (a direction for a strip is a direction for its –factor ) by isometric gluing subject to the following conditions.
-
(1)
The boundary geodesics of each strip , which we will refer to as singular geodesics, are glued isometrically to distinct walls in .
-
(2)
Each wall is glued to at most two strips, and the gluing lines are not parallel.
Notations: We use the notion if the exists such that , and we use the notion if and . Also, when we write we mean that .
Denote by and the –length and –length of a path in a metric product space . These two lengths are equal for a path if it is parallel to a factor; in general, they are bilipschitz.
3. Special paths in CKA action
Let be a CKA action. In this section, we are going to define special paths (see Definition 3.6) in that will be used on the latter sections. Roughly speaking, each special path in is a concatenation of geodesics in consecutive pieces ’s of and they are uniform quasi-geodesic in the sense that there exists a constant such that every special path is –quasi-geodesic.
We first introduce the class of special paths in a template which shall be mapped to special paths in up to a finite Hausdorff distance.
3.1. Special paths in a template
Definition 3.1.
Let be the template given by Definition 2.7. A (connected) path in is called special path if is a concatenation of geodesics such that each lies on the strip adjacent to and .
Remark 3.2.
By the construction of the template, the endpoints of must be the intersection points of singular geodesics on walls .
Lemma 3.3.
Assume that the angles between the singular geodesics on walls are between and for a universal constant . There exists a constant such that any special path is a –quasi-geodesic.
Proof.
Let be a special path with endpoints . We are going to prove that for a constant . Since any subpath of a special path is special, this proves the conclusion.
Let be the unique CAT geodesic between and . By the construction of the template, if does not pass through the intersection point of the singular geodesics on a wall , then it passes through a point on one singular geodesic and then a point on the other singular geodesic . Recall that the angle between the singular geodesics on walls are uniformly between and . There exists a constant depending on only such that . We thus replace by for every possible triangle on each wall . The resulted path then connects consecutively the points on the walls in the order of their intersection with , so it is the special path from to satisfying the following inequality
Thus, we proved that is a –quasi-geodesic. ∎
We are going to define a template associated to a geodesic in the Bass-Serre tree as the following.
Definition 3.4 (Standard template associated to a geodesic ).
Let be a geodesic segment in the Bass-Serre tree . We begin with a collection of walls and an isometry for each edge . For every pair , of adjacent edges of , we let be a strip which is isometric to if the width of is at least , and isometric to otherwise; we let be an affine map which respects product structure ( is an isometry if the width of is greater than or equal to and compresses the interval otherwise). We construct by gluing the strips and walls so that the maps and descend to continuous maps on the quotient, we denote the map from by .
The following lemma is cited from Lemma 4.5 and Proposition 4.6 in [CK02].
Lemma 3.5.
-
(1)
There exists such that the following holds. For any geodesic segment , the angle function satisfies .
-
(2)
There are constants such that the following holds. Let be a geodesic segment in , and let be the map given by Definition 3.4. Then is a –quasi-isometric embedding. Moreover, for any , there exists a continuous map such that .
3.2. Special paths in the admissible space
In this subsection, we are going to define special paths in .
Recall that we choose a –equivariant family of Euclidean planes . For every pair of planes so that , a minimal geodesic between in determines a strip for some geodesic . It is possible that is trivial so the width of the strip is zero. Let and an edge with an endpoint . The minimal geodesic from to (possibly not belong to ) also define a strip where the geodesic is the projection to of the intersection of this minimal geodesic with . Thus, is possibly not in the strip but within its -neighborhood by Lemma 2.4.
Definition 3.6 (Special paths in ).
Let be the indexed map given by Remark 2.6. Let and be two points in . If then we define a special path in connecting to is the geodesic . Otherwise, let be the geodesic edge path connecting to and let be the intersection point of the strips and , where and . The special path connecting to is the concatenation of the geodesics
Remark 3.7.
By definition, the special path except the and depends only on the geodesic in and the choice of planes .

3.3. Special paths in the admissible space are uniform quasi-geodesic
In this section, we are going to prove the following proposition.
Proposition 3.8.
There exists a constant such that every special path in is a –quasi-geodesic.
Lemma 3.9.
[CK02, Lemma 3.17] There exists a constant with the following property. Let be a geodesic in with and be the geodesic edge path connecting to . Then there exists a sequence of points such that for any .
Let be a geodesic in with . We apply a minimizing horizontal slide of the endpoint to obtain a point so that is parallel to and the projection of on is orthogonal to .
Lemma 3.10.
Let where are the endpoints of a geodesic in . Then there exists a universal constant depending on such that for each , we have
where and and .
Proof.
We use the notion if there exists such that .
Let be the constant given by Lemma 2.4 and satisfying Lemma 3.9 such that . Without loss of generality, we can assume .
Denote and in this proof. By Lemma 3.9, there exists a point such that . For ease of computation, we will consider the mixed length of the path which satisfies
(2) |
where is the –metric on the metric product .
Note that the Euclidean plane for contains two non-parallel lines and . So we can apply a minimizing horizontal slide of the endpoint of in to a point on . On the one hand, since the line on is –factor of , this slide decreases the -distance by for a constant depending on hyperbolicity constant of . On the other hand, by the triangle inequality, this slide increases by at most . Hence, we obtain
Similarly, by a minimizing horizontal slide of the endpoint of in to ,
yielding
Together with (2) this completes the proof of the lemma. ∎
Proof of Proposition 3.8.
Let be the special path from to for so that ; otherwise it is a geodesic, and thus there is nothing to do. Let be the geodesic in from to . With notations as above (see Definition 3.6),
By Lemma 3.10, there exists a constant such that
Denoting , it remains to give a linear bound on in terms of .
By Lemma 3.5, there exists a –template for the such that is a –quasi-isometric map from the template to the union of the planes with the strips . Moreover, sends walls and strips of to the –neighborhood of planes and strips of accordingly. Hence, maps the intersection point on of the singular geodesics of two strips in to a finite –neighborhood of . Since the map is affine on strips and isometric on walls of , we conclude that there exists a special path in such that is sent to a finite neighborhood of the special path . Lemma 3.3 then implies that is a –quasi-geodesic for some so is a –quasi-geodesic for some depending on . The proof is complete. ∎
4. Quasi-isometric embedding of admissible groups into product of trees
A quasi-tree is a geodesic metric space quasi-isometric to a tree. In this section, we are going to prove Theorem 1.5 that states if is a flip CKA action (see Definition 4.1) then is quasi-isometric embedded into a finite product of quasi-trees. The strategy is that we first show that the space is quasi-isometric embedded into product of two hyperbolic spaces , (see Subsection 4.2). We then show that each hyperbolic space is quasi-isometric embedded into a finite product of quasi-trees (see Subsection 4.4).
4.1. Flip CKA actions and constructions of two hyperbolic spaces
Let be a CKA action. Recall that each decomposes as a metric product of a hyperbolic Hadamard space with the real line such that admits a geometric action of . Recall that we choose a –equivariant family of Euclidean planes .
Definition 4.1 (Flip CKA action).
If for each edge , the boundary line is parallel to the –line in , then the CKA action is called flip in sense of Kapovich-Leeb.
Let be the set of boundary lines of which are intersections of with for all edges issuing from . Thus, there is a canonical one-to-one correspondence between and the link of denoted by .
Definition 4.2.
A flat link is the countable union of (closed) flat strips of width 1 glued along a common boundary line called the binding line.
Construction of hyperbolic spaces and : We first partition the vertex set of the Bass-Serre tree into two disjoint class of vertices and such that if and are in then is even.
Given , we shall build a geodesic (non-proper) hyperbolic space by glueing for all along the boundary lines via flat links.
Consider the set of vertices in such that their pairwise distance in equals 2. Equivalently, it is the union of the links of every vertex . For any , the edges and determine two corresponding boundary lines and which are the intersections of with for respectively. There exists a canonical identification between and so that their –coordinates equal in the metric product .
Note the link determines a flat link so that the flat strips are one-to-one correspondence with . In equivalent terms, it is a metric product , where is parallel to the binding line.
For each , the set of hyperbolic spaces where are glued to the flat links along the boundary lines of flat strips and of hyperbolic spaces with the identification indicated above. Therefore, we obtain a metric space from the union of and flat links .
Remark 4.3.
By construction, and are disjoint in for any two vertices with . Endowed with induced length metric, is a hyperbolic geodesic space but not proper since each is glued via flat links with infinitely many ’s where .
Definition 4.4.
Let be an element in . The translation length of is defined to be . Let . If and then is called elliptic. If and , it is called loxodromic (or hyperbolic).
Remark 4.5.
We note that for any . If is loxodromic, is isometric to , and acts on as translation by .
Lemma 4.6.
There exists a subgroup of index at most 2 in so that preserves and respectively and for any .
Proof.
Observe first that if for some and , then holds for any . Indeed, if is elliptic and thus rotates about a point , the geodesic for any is contained in the union and thus has even length. Otherwise, must be a hyperbolic element and leaves invariant a geodesic acted upon by translation. By a similar reasoning, if moves the points on with even distance, then for any .
Consider now the set of elements such that for any . Using the tree again, if , then for any . Thus, is a group of finite index 2. ∎
Let be the subgroup of given by the lemma. By Bass-Serre theory, it admits a finite graph of groups where the underlying graph is bipartite with vertex sets and where , and the vertex groups are isomorphic to those of .
Lemma 4.7.
The space is a –hyperbolic Hadamard space where only depends on the hyperbolicity constants of ().
If for every , and for each edge , then the subgroup given by Lemma 4.6 acts on with the following properties:
-
(1)
for each , the stabilizer of is isomorphic to and acts geometrically on , and
-
(2)
for each , the flat link admits an isometric group action of so that acts by translation on the line parallel to the binding line and on the set of flat strips by the action on the link .
Proof.
On one hand, acts on the boundary line through . On the other hand, acts on the boundary line of the flat strip corresponding to the edge . Since these two actions are compatible with glueing of ’s where , we can extend the actions on ’s and flat links ’s to get the desired action of on . ∎
4.2. Q.I. embedding into the product of two hyperbolic spaces
Proposition 4.8.
Let be a flip CKA action and the subgroup in of index at most given by Lemma 4.6. Let () be the hyperbolic space constructed in Section 4.1 with respect to . Then there exists a quasi-isometric embedding map from to .
If for every , and for each edge , then the above map can be made -equivariant.
Proof.
Let be the indexed map given by Remark 2.6. Choose a vertex and a point such that . Note that is a –equivariant, hence . Without loss of generality, we can assume that . Let be the orbit of in .
For any then for some . We remark here that in general it is possible that belong to several ’s for some . However, recall that we have the given indexed map . This indexed function will tell us exactly which space we should project into, i.e, should project to .
We recall that has finite index in and it preserves and .
Step 1 : Construct a quasi-isometric embedding map
We are going to define the map where . We first define a map .
For any then for some , and thus . Since we assume that and preserves , it follows that . We define where is the projection of to the factor . We define to be the point on the binding line of the flat link so that its –coordinate is the same as that of in the metric product .
Step 2: Verifying is a quasi-isometric embedding.
We are now going to show that is a quasi-isometric embedding. Before getting into the proof, we clarify here an observation that will be used later on.
Observation: By the tree-like construction of , any geodesic in with endpoints crosses for alternating vertices in their order appearing in the interval, where is the geodesic in the tree . Using the convexity of boundary lines in a hyperbolic CAT(0) space, we see that the intersection connects two boundary lines in so that the projection is uniformly close to . We can then construct a quasi-geodesic in with the same endpoints as so that connects to .
Claim: There exists a constant such that for any .
Indeed, let and be two elements such that and . Note that and . We consider the following cases.
Case 1: . In this case, it is easy to see the claim holds since
Case 2: . Note that they both belong to , so is even. Let . We write
as the edge paths in . Let be the initial vertex of with and be the terminal vertex of . We note that and .
With notations in in Definition 3.6, the special path between decomposes as the concatenation of geodesics:
Denote and . By the above observation, we connect to by a quasi-geodesic in so that whenever passes through , it is orthogonal to the boundary lines. In this way, we can write as the concatenation of geodesic segments , where are maximal segments contained in the flat links. The first and last may have overlap with boundary lines, and the other are orthogonal to the boundary lines of .
Similarly, let be a quasi-geodesic from to in as the concatenation of geodesic segments .
We relabel by and relabel by . For each vertex , let and denote the projections of to the factor and respectively.
By the construction of and we note that there exists a constant such that
and
Summing over two equations above, we obtain
(3) |
for some constant .
Since is a –quasi-geodesic connecting two points and (for some uniform constant that does not depend on , ), we have that with . This fact together with formula (3) and the fact give a constant such that . The claim is verified. Therefore, is a quasi-isometric embedding. ∎
4.3. Q.I. embedding into a finite product of trees
In this section, we first recall briefly the work of Bestvina-Bromberg-Fujiwara [BBF15] on constructing a quasi-tree of spaces. The reminder is then to produce a collection of quasi-lines to establish a distance formula for geometric actions of hyperbolic groups, see Lemma 4.17. The result is not new (cf. [BBF19, Prop. 3.3]), but the construction of quasi-lines is new and generalizes to certain non-proper actions, see Lemma 5.5.
We shall make use of the work of . Their theory applies to any collection of spaces equipped with a family of projection maps
satisfying the so-called projection axioms with projection constant . The precise formulation of projection axioms is irrelevant here. We only mention that their results apply to a collection of quasi-lines with bounded projection property in a (not necessarily proper) hyperbolic space , where the projection maps for are shortest point projections to in . Then satisfies projection axioms with projection constant (see [BBF19, Proposition 2.4]).
Fix . Following [BBF15], a quasi-tree of spaces is constructed with a underlying quasi-tree (graph) structure where every vertex represents a quasi-line in and two quasi-lines are connected by an edge of length 1 from to . If , then is a unbounded quasi-tree.
If admits a group action of so that for any and , then acts by isometry on .
By [BBF15, Lemma 4.2, Corollary 4.10], every quasi-line with induced metric from is totally geodesically embedded in and the shortest projection maps from to in the quasi-tree coincides with the projection maps up to uniform finite Hausdorff distance.
By abuse of language, for both and , we denote
where the projections in the right-hand are understood in and accordingly. The above discussion implies that the two projections gives the same value up to a uniform bounded error.
Proposition 4.9.
Let be a collection of quasi-lines in a –hyperbolic space . If there is such that for all , then satisfies the projection axioms with projection constant depending only on . Moreover, for any ,
for all .
Remark 4.10.
In [BBF19, Proposition 2.4], the above formula is stated for with the strong projection axioms. However, by [BBF19, Theorem 2.2], a modification of projection maps within finite Hausdorff distance can always be done so that satisfies the strong projection axioms. Thus, the same formula still holds with original projection maps.
As a corollary, the distance formula still works when the points are perturbed up to bounded error.
Corollary 4.11.
Proof.
If then there exists a constant such that for any . Assuming then we see that
yielding the desired formula. ∎
Definition 4.12 (Acylindrical action).
By [Bow08], any nontrivial isometry of acylindrical group action on a hyperbolic space is either elliptic or loxodromic. A -quasi-geodesic for some is referred to as a quasi-axis for a loxodromic element , if have (uniform) finite Hausdorff distance.
The following property in hyperbolic groups is probably known to experts, but is referred to a more general result [Yan19, Lemma 2.14] since we could not locate a precise statement as follows. A group is called non-elementary if it is neither finite nor virtually cyclic.
Lemma 4.13.
Let be a non-elementary group admitting a co-bounded and acylindrical action on a –hyperbolic space . Fix a basepoint . Then there exist a set of three loxodromic elements and with the following property.
For any there exists so that is a loxodromic element and the bi-infinite path
is a –quasi-geodesic.
Sketch of the proof.
This follows from the result [Yan19, Lemma 2.14] which applies to any isometric action of on a metric space with a set of three pairwise independent contracting elements (loxodromic elements in hyperbolic spaces). If denotes the set of –translated quasi-axis of all elements in , the pairwise independence condition is equivalent (defined) to be the bounded projection property of . Thus, the existence of such is clear in a proper action of a non-elementary group. For acylindrical actions, this is also well-known, see [BBF19, Proposition 3.4] recalled below. ∎
Proposition 4.14.
[BBF19] Assume that a hyperbolic group acts acylindrically on a hyperbolic space . For a loxodromic element , consider the set of all -translates of a fixed -quasi-axis of for given . Then there exist and such that for any , the set
consists of at most double -cosets.
In particular, there are at most distinct pairs satisfying up to the action of .
The following corollary can be derived from the “in particular” statement using hyperbolicity.
Corollary 4.15.
Under the assumption of Proposition 4.14, for any , there exist constants such that for any geodesic segment of length , we have
Convention 4.16.
When speaking of quasi-lines in hyperbolic spaces with actions satisfying Lemma 4.13, we always mean –quasi-geodesics where depend on and .
Lemma 4.17.
Let be a non-elementary group admitting a proper and cocompact action on a –hyperbolic space . Assume that is a –finite collection of quasi-lines. Then for any sufficiently large , there exist a –finite collection of quasi-lines in and a constant , such that for any , the following holds
Remark 4.18.
The statement of Lemma 4.17 (with torsion allowed here) is a re-package of Proposition 3.3 and Theorem 3.5 in [BBF19]. Our proof is different and generalizes to certain co-bounded and acylindrical actions coming from relatively hyperbolic groups on their relative Cayley graphs. See Lemma 5.5 for details.
Proof.
Fixing a point , the co-bounded action of on gives a constant such that . By hyperbolicity, if is a –quasi-geodesic, then there exists a constant such that implies
(4) |
Let be the constant given by Corollary 4.15 for the geometric (so acylindrical) action of on .
Fix and denote . Since the action is proper, the set is finite. Let us consider the set of loxodromic elements where and is provided by Lemma 4.13. Note that . Let be the set of all –translated axis of . It is possible that since two elements in may be conjugate.
Assume that . Consider a geodesic from to and choose points on for such that for and where . Since , there exists so that . It implies that , and thus we have for . Noting that is contained in a -translated axis of some loxodromic element in , we thus obtain axis (with possible multiplicities: for ) satisfying so that
which yields
where the constant bounds the length of the last segment .
By Equation (4), and then . Thus, we obtain
which implies the upper bound over . Of course, the upper bound holds as well after adjoining into .
The remainder of the proof is to prove the lower bound. Let be the set of quasi-lines satisfying . Note that the set of axis obtained as above is included into .
By the proper action of on , the –finite has bounded intersection so does . Thus, there is so that for any . In particular, different ’s have overlaps bounded above by .
4.4. Q.I. embedding into a finite product of trees
This subsection is devoted to the proof of Theorem 1.5. The results obtained here are not used in other places, and so can be skipped if the reader is interested in the stronger conclusion, the property (QT), under stronger assumption.
We start by explaining the choice of the constants and the collection of quasi-lines in that will be used in the rest of this subsection.
The constants and and : Let and be two –hyperbolic spaces given by Lemma 4.7 where depends on the hyperbolicity constants of .
Note that each for is isometrically embedded into and thus –hyperbolic. We follow the Convention 4.16 on the quasi-lines which are –quasi-geodesics in and .
By the –hyperbolicity of , there exist constants depending on (and also ) such that if any (–)quasi-lines have a distance at least then .
We then obtain the projection constant by Proposition 4.9.
The collection of quasi-lines in : Fix any sufficiently large number depending on , where is the collection of boundary lines of . By Lemma 4.17, there exist a locally finite collection of quasi-lines in and a constant such that
(5) |
for any . Since there are only finitely many –orbits of we assume furthermore if for . Then
is a locally finite collection of quasi-lines in , preserved by the group .
We use the following lemma in the proof of Proposition 4.20 that gives us a distance formula for .
Lemma 4.19.
There exists a constant depending only on with the following properties.
-
(1)
For any , we have .
-
(2)
For any and , there are at most quasi-lines in such that .
Proof.
Since there are only finitely many ’s up to isometry, and if , the union of quasi-lines containing is uniformly locally finite: any ball of a fixed radius in intersects a uniform number of quasi-lines depending only on the radius. By the hyperbolicity of , the local finiteness implies the bounded projection property, so gives the desired constant in the assertion (1).
By the construction of , the shortest projection of a point to for has to pass through a boundary line of , so is contained in the projection of to . By the assertion (1) we have . If for and with , then by the above defining property of and . By local finiteness, there are at most quasi-lines with this property, proving the assertion (2). ∎
Proposition 4.20 (Distance formula for ).
For any , there exists a constant such that
(6) |
Proof.
Since the –neighborhood of the union is , assume for simplicity and where . By the construction of , a geodesic travels through and then flat links , where and appear alternatively on . Thus, .
Let us denote the exit point on the boundary line of and entry point on of by and respectively for where and by convention. Thus,
(7) |
Therefore, we shall derive (6) from (7) which requires to apply the formula (5) for . To that end, we need the following estimates. Recall that means the equality holds up to a multiplicative constant depending on .
Claim 1.
-
(1)
If there is such that then
(8) -
(2)
. If , then
Proof of the Claim 1.
If there is such that we then have
where Lemma 4.19 is applied, and after taking the cutoff function ,
This in turn implies (8).
Recall that is contained in the union of two flat strips with width in a flat link, and is from one boundary line to the other . Thus, . If is assumed, then and . The assertion (2) follows similarly as above. ∎
Recalling , the assertion (2) of the Claim 1 implies a constant such that
(9) |
Lemma 4.21.
The collection can be written as a union (possibly non-disjoint) with the following properties for each :
-
(1)
for any two quasi-lines we have ,
-
(2)
the –neighborhood of the union contains ,
-
(3)
for any the quasi-tree of quasi-lines is a quasi-tree.
Proof.
Since acts geometrically on for and is –finite, there exists a constant such that the –neighborhood of the union contains for each . Since is locally finite and –invariant, the –neighborhood of any quasi-line in intersects quasi-lines from for some .
We can now write as the (possibly non-disjoint) union with the following two properties for each :
-
(1)
for any two quasi-lines we have ,
-
(2)
the –neighborhood of the union contains .
Indeed, by definition of , any ball of radius intersects a quasi-line so for each , there exists such that . Starting from a quasi-line , we inductively choose the quasi-lines which intersect the –neighborhood of the already chosen ones, and by the axiom of choice, a collection of quasi-lines containing is obtained so that the properties (1) and (2) are true. The other collections for is obtained similarly from the other quasi-lines intersecting the –neighborhood of . The property (2) guarantees from the choice of . We do allow , but any would appear at most once in each .
Proof of Theorem 1.5.
Let () be the hyperbolic space constructed in Section 4.1 with respect to . By Proposition 4.8, the admissible group admits a quasi-isometric embedding into . Thus, to complete the proof of Theorem 1.5, we only need to show that each hyperbolic space is quasi-isometric embedded into a finite product of quasi-trees. We give the proof for and the proof for is symmetric.
Let be the indexed map given by Remark 2.6. Let be the collection of quasi-lines given by Lemma 4.21.
Let . Since the –neighborhood of is , it suffices define a quasi-isometric embedding map from to a finite product of quasi-trees.
We now define a map
where is the Bass-Serre tree of .
Let and assume . By the property (2) of Lemma 4.21, we choose a point for each such that . Denote . Let .
We now verify that is a quasi-isometric embedding. Since and , let be given by Corollary 4.11 so that for , the following distance formula holds
for each . Therefore,
Recall that is a possibly non-disjoint union, but any quasi-line with in the above sum appears at most once in each . We thus obtain
which together with distance formula (6) for concludes the proof of Theorem. ∎
5. Proper action on a finite product of quasi-trees
In this section, under a stronger assumption on vertex groups as stated in Theorem 1.1, we shall promote the quasi-isometric embedding to be an orbital map of an action of the admissible group on a (different) finite product of quasi-trees.
By [BBF19, Induction 2.2], if has finite index and acts on a finite product of quasi-trees, then so does . We are thus free to pass to finite index subgroups in the proof.
Recall that where consists of vertices in with pairwise even distances, and is the hyperbolic space constructed from . By Lemma 4.6, let be the subgroup of index at most preserving and .
5.1. Construct cone-off spaces: preparation
In this preparatory step, we first introduce another hyperbolic space which is the cone-off of the previous hyperbolic space over boundary lines of flat links. We then embed into a product of and a quasi-tree built from the set of binding lines from the flat links.
Definition 5.1 (Hyperbolic cones).
[BH99, Part I, Ch. 5] For a line and a constant , a hyperbolic –cone denoted by is the quotient space of by collapsing as a point called apex. A metric is endowed on so that it is isometric to the metric completion of the universal covering of a closed disk of radius in the real hyperbolic plane punctured at the center.
A hyperbolic multicones of radius is the countable wedge of hyperbolic –cones with apex identified.
If is an isometry on then extends to a natural isometric action on the hyperbolic cone which rotates around the apex and sends the radicals to radicals.
Similar to the flat links, the link for determines a hyperbolic multicones of radius denoted by so that the set of hyperbolic cones is bijective to the set of vertices adjacent to . And acts on so that the center of rotates each hyperbolic cone around the apex and permutes the set of hyperbolic cones by the action of on .
Construction of the cone-off space . Let and . Let be the disjoint union of and hyperbolic multicones glued by isometry along the boundary lines of and those of hyperbolic multicones .
Remark 5.2.
We note that is obtained from by replacing each flat links by hyperbolic multicones. However, the identification in between boundary lines of and multicones is only required to be isometric, while the –coordinates of the boundary lines in constructing have to be matched up.
We now give an alternative way to construct the cone-off space , which shall be convenient in the sequel.
Relatively hyperbolic structure of cone-off spaces. Let be the disjoint union of and hyperbolic cones glued along boundary lines . If denotes the stabilizer in of the boundary , then is virtually cyclic and almost malnormal. Since is -finite by conjugacy, choose a complete set of representatives. By [Bow12], is hyperbolic relative to peripheral subgroups .
Let be the coned-off Cayley graph (after choosing a finite generating set) by adding a cone point for each peripheral coset of and joining the cone point by half edges to each element in the peripheral coset. The union of two half edges with two endpoints in a peripheral coset shall be referred to as an peripheral edge. See the relevant details in [Far98].
By [Bow08, Lemma 3.3], [Osi16, Prop. 5.2], the action of on is acylindrical. There exists a -equivariant quasi-isometry between the coned-off Cayley graph and the coned-off space which sends peripheral coset of to . Thus, the action of on is co-bounded and acylindrical as well.
Alternatively, the cone-off space could be obtained from the disjoint union of by identifying the apex of hyperbolic cones from the same link where .
Since has the bounded intersection property, by [DGO17, Corollary 5.39], for a sufficiently large constant , the space is a hyperbolic space with constant depending only on the original one. Thus, the space is a hyperbolic space.
By Lemma 4.6, a subgroup of index at most 2 in leaves invariant and . The following lemma is proved similarly as Lemma 4.7.
Lemma 5.3.
Fix a sufficiently large . The space is a –hyperbolic space where only depends on the hyperbolicity constants of ().
If for every and for every edge , then a subgroup of index at most 2 in acts on with the following properties:
-
(1)
for each , the stabilizer of is isomorphic to and acts coboundedly on , and
-
(2)
for each , the stabilizer of the apex of is isomorphic to so that acts on the set of hyperbolic cones by the action on the link and on acts by rotation on each hyperbolic cone.
From now on and until the end of the next subsection, we assume that for every and for every edge . After passage to a finite index subgroup of , this assumption will be guaranteed by Corollary 5.13 below, where are assumed to omnipotent.
The remainder of this section is to relate the metric geometry of and . This is achieved by a series of lemmas, accumulating in Lemma 5.6.
Let denote the shortest projection to a quasi-line in or in . To be clear, we use (resp. ) to denote the -diameter (resp. -diameter) of the projection of the points to .
Let be the set of all binding lines in the flat links over . Note that is disjoint with the union , although the binding lines in flat links are parallel to the boundary lines of (the flat strips and) for .
We first give an analogue of Lemma 4.17 for acylindrical action on coned-off spaces .
Up the equivariant quasi-isometry mentioned above, we shall identify with the Cayley graph of , and with the coned-off Cayley graph , and with the collection of left peripheral cosets of .
A geodesic edge path in the coned-off Cayley graph is -bounded for if every peripheral edge has two endpoints within -distance at most .
By definition, a geodesic can be subdivided into maximal -bounded non-trivial segments () separated by peripheral edges () where . It is possible that : consists of only peripheral edges.
Define
which sums up the lengths of -bounded subpaths of length at least . It is possible that , so . Define to be the maximum of over all relative geodesics between . Thus, is -invariant. We remark that the “thick” distance is inspired by the corresponding [BBF19, Def. 4.8] in mapping class groups.
Lemma 5.4.
For any sufficiently large ,
(10) |
Proof.
Let be a geodesic between in so that . We obtain the lifted path by replacing each peripheral edge with endpoints in a peripheral coset by a geodesic in with same endpoints. The well-known fact ([DS05], [GP16]) that is a uniform quasi-geodesic gives the so-called distance formula [Sis13, Theorem 0.1]: for any ,
Indeed, the additional ingredient (to quasi-geodesicity of ) is the following fact: if and only if contains a peripheral edge with two endpoints in with , where depends on and vice versa.
By definition, differs from in the sum of lengths of at most segments and edges , where and . Therefore, we can replace with by worsening the multiplicative constant in the above distance formula. This shows (10). ∎
Lemma 5.5.
For any sufficiently large , there exist a –finite collection of quasi-lines in and a constant , such that for any two vertices , the following holds
(11) |
Proof.
For simplicity, we suppress indices from in the proof.
Let be the set of elements so that and some geodesic in is -bounded. The definition of -boundedness implies that is a finite set, since and the word metric is proper.
Let be a geodesic path between so that . Let be the set of maximal -bounded segments of in . Then
(12) |
We now follow the argument in the proof of Lemma 4.17 to produce the collection of quasi-lines to approximate .
The argument below considers every in the above sum of (12), but for simplicity, denote with endpoints .
Since the action of on is acylindrical, consider the set of loxodromic elements on where and is provided by Lemma 4.13. Since is finite, is finite as well. Let be the set of all –translated axis of .
If , we subdivide further , so that subsegments have -length and the last one have -length less than . Each is -bounded since the -boundedness is preserved by taking subpaths. Similarly as in Lemma 4.17, a set of axes of loxodromic elements is then produced to give the upper bound of . The lower bound uses a certain local finiteness of due to the proper action in Lemma 4.17, which is now guaranteed by Corollary 4.15 in an acylindrical action. Thus,
(13) |
where the constant depends on .
Recall that each is a uniform quasi-line, i.e.: a -quasi-geodesic in . Since is a subsegment of the geodesic , the hyperbolicity of implies that and are contained in a -neighborhood of and respectively, where depends on . Thus, we have
(14) |
for every . So for every , we obtained (13) and (14), and the conclusion thus follows from (12). ∎
Recall that a geodesic in is the union of a sequence of maximal geodesics in ’s. As is related with , we define
(15) |
Since is -invariant, we see that is -invariant (even though it is not a metric anymore).
Lemma 5.6.
There exists such that for any two points and , we have
(16) |
Recall that the notation means that equals up to multiplicative and additive constants depending on .
Proof.
Let be a -geodesic with endpoints . Let us consider the generic case that ; the case is much easier and left to the reader.
Assume that are not contained in any hyperbolic -cones, up to moving with a distance at most . We can then write the geodesic as the following union
(17) |
where is a geodesic in the cone-off space with endpoints on boundary lines of , and is contained in the hyperbolic multicones passing the apex. Thus, has length .
We replace by an -geodesic with the same endpoints : is contained in the corresponding flat links, whose binding line is denoted by . Thus, . This replacement becomes non-unique when different ’s have overlap (in the subspace ). However, the bounded intersection of gives a uniform upper bound on the overlap. Let be any constant sufficiently bigger than this bound. We then number those subpaths of with in a fixed order (eg. from left to right): . Up to bounded modifications, we obtain a well-defined notion of lifted path with same endpoints of .
Observe that is a uniform quasi-geodesic. This is well-known and one proof proceeds as follows: since the set of binding lines has bounded intersection in , the above construction implies that is an efficient semi-polygonal path in the sense of [Bow12, 6, Section 7]. The observation then follows as a consequence of [Bow12, Lemma 7.3].
If so , then follows from the quasi-geodesicity of and the hyperbolicity of . We incorporate the ’s in Eq. (17) with into the multiplicative constant, and thus the following holds
Combining the formula (10) and (15), we get
Using the local finiteness and bounded intersection , for each with , there are only finitely many such that . Hence, by worsening the multiplicative constant, we have
Recall that from the quasi-geodesic . Then the desired (16) follows. ∎
Recall that is the set of all binding lines in over . Since has bounded projection property in , it satisfies the projection axioms and we can construct the quasi-tree for , equipped with the length metric .
By the following result, we shall be reduced to embed the action of on into finite product of quasi-trees. We warn the reader that is not a metric on . However, we can still talk about the product space equipped with the symmetric non-negative function . The quasi-isometric embedding of into the product only means the coarse bilipschitz inequality.
Proposition 5.7.
For any , there exists a –equivariant quasi-isometric embedding from to the product .
Proof.
We first define the map .
If define in . To define , choose a (non-unique) closest quasi-line in to and define . The choice of is not important and the important thing is the distance from to any such chosen is uniformly bounded, thanks to co-compact action of on . Extend by –equivariance for all .
If lies in the flat links define to be the apex of and where is the binding line of .
5.2. Construct the collection of quasi-lines in
The goal of this subsection is to introduce a collection of quasi-lines so that a distance formula holds for .
Recall that is the hyperbolic cone-off space constructed from . According to Proposition 5.7, we are working in the cone-off space endowed with length metric . In particular, all quasi-lines are understood with this metric, and boundary lines of are of bounded diameter so are not quasi-lines anymore in .
First of all, let us fix the constants used in the sequel.
Recall that for every , the cone-off space admits a co-bounded and acylindrical action of . Thus, when talking about quasi-lines, we follow the Convention 4.16: quasi-lines are –quasi-geodesics in and ’s (isometrically embedded into the former), where are given by Lemma 4.13 applied to those actions of on .
If is a quasi-line in , denote by the –diameter of the shortest projection of to in . By Lemma 5.3, is –hyperbolic for a constant . The coning-off construction is crucial to obtain the uniform constant in the next lemma.
Lemma 5.8.
There exists a constant depending on (and also ) with the following property: for any (–)quasi-lines in and in with we have .
Proof.
By the construction of , any geodesic from to has to pass through the apex between and , and thus the shortest projection is contained in the projection of the apex to . By hyperbolicity, there exists a constant depending only on such that the diameter of the projection of any point to every quasi-line is bounded above by . The conclusion then follows. ∎
Fix . For each , there exist a -finite collection of quasi-lines in and a constant such that (11) holds.
Since preserves and , by Lemma 5.3, there are only finitely many –orbits in , so we can assume furthermore if for . Then the collection is a –invariant collection of quasi-lines in .
Recall that is the radius of the multicones in constructing for , and is the Bass-Serre tree for admissible group .
Proposition 5.9.
For any , the following holds
(18) |
Proof.
The proof proceeds similarly as that of Proposition 4.20, so only the differences are spelled out. Assume that are not in any hyperbolic multicones. We can then write the geodesic as in (17) and keep notations there.
5.3. Reassembling finite index vertex groups
By Bass-Serre theory, the finite index subgroup from Lemma 4.6 acts on the Bass-Serre tree of and can be represented as a finite graph of groups where the vertex subgroups are isomorphic to those of .
Let be an oriented edge in from to (it is possible that because could be a loop) and be the oriented edge with reversed orientation. A collection of finite index subgroups is called compatible if whenever , we have
We shall make use of [DK18, Theorem 7.51] to obtain a finite index subgroup in from a compatible collection of finite index subgroups. For this purpose, we assume the quotient of each vertex group for is omnipotent in the sense of Wise.
In a group two elements are independent if they do not have conjugate powers. (see Definition 3.2 in [Wis00]).
Definition 5.10.
A group is omnipotent if for any set of pairwise independent elements () there is a number such that for every choice of positive natural numbers there is a finite quotient such that has order for each .
Remark 5.11.
If is hyperbolic, two loxodromic elements are usually called independent if the collection of –translated quasi-axis of has the bounded projection property. When is torsion-free, it is equivalent to the notion of independence in the above sense.
Let be a loxodromic element in a hyperbolic group and be the maximal elementary group containing . By [BH99, Ch. II Theorem 6.12], contains a subgroup intersecting trivially with so that the direct product is a finite index subgroup. Thus, the image of in is of finite index in and acts geometrically on hyperbolic spaces .
The following result will be used in the next subsection to obtain desired finite index subgroups.
Lemma 5.12.
Let be a collection of finite index subgroups. Then there exists a compatible collection of finite index subgroups such that is of finite index in for each , where is of finite index in .
Proof.
Let be an oriented edge in from to (it is possible that ) and be the oriented edge with reversed orientation.
If , then the abelian group is a nontrivial cyclic group contained in a maximal elementary in where is a primitive loxodromic element. Similarly, for , let be a primitive loxodromic element in containing . Then and preserve two lines respectively which are orthogonal in the Euclidean plane and thus generate an abelian group of rank 2 so that is of finite index.
Consider the collection of all oriented edges in such that . Let be the set of primitive loxodromic elements in obtained as above in correspondence with . Note that are pairwise independent in . This follows from the item (3) in Definition 2.1 of admissible group, otherwise and would be commensurable for .
By the finite index of in , there exists a set of powers of ’s in denoted by . Since are still pairwise independent in , the omnipotence of gives the constant by Definition 5.10. Let
Define with . By the omnipotence of and restricting to , there is a finite index subgroup of such that .
For each vertex in and for each edge in , we define
and
To conclude the proof, it remains to note the collection is compatible. It is obvious that , so . Conversely, . ∎
Corollary 5.13.
There is a finite index admissible group in the sense of Definition 2.1 where every vertex group are direct products of a hyperbolic group and .
Proof.
By [DK18, Theorem 7.50], the compatible collection of finite index subgroups from Lemma 5.12 determines a finite index group . Indeed, is the fundamental group of a finite covering space which are obtained from finite many copies of finite coverings in correspondence to glueeing along edge spaces in correspondence to . Thus, splits over edge groups as a finite graph of groups where the vertex groups are conjugates of . In view of Definition of 2.1, it suffices to certify the non-commensurable edge groups adjacent to the same vertex group. This follows from the above proof of Lemma 5.12, where the edge groups are direct products of with pairwise independent loxodromic elements. Thus, different edge groups are not commensurable. ∎
5.4. Partition into sub-collections with good projection constants: completion of the proof
With purpose to prove Theorem 1.1, it suffices to prove property (QT) for a finite index subgroup of . By Corollary 5.13, we can assume that the CKA flip action of on satisfies that every vertex group are direct products where is of finite index in . Since is omnipotent and then residually finite, without loss of generality we can assume that is torsion-free.
Since the assumption of Lemma 5.3 is fulfilled, the results in Sections 5.1 and 5.2 hold: a finite index at most 2 subgroup acts on the cone-off spaces for with distance formula.
Let . Let us recall the data we have now:
- (1)
- (2)
The first step is passing to a further finite index subgroup of so that is partitioned into -invariant sub-collections with projection constants . It follows closely the argument in [BBF19] which is presented below for completeness.
The constants and : The constant is chosen so that it satisfies Proposition 4.14 and Lemma 5.8 simultaneously. Then is given by Proposition 4.9.
Lemma 5.14.
Let be a -finite collection of quasi-lines obtained as above by Lemma 4.17. Then there exists a finite index subgroup such that any two distinct quasi-lines in the same -orbit have -bounded projection.
Proof.
By construction, the quasi-lines in are quasi-axis of loxodromic elements whose maximal elementary groups are virtually cyclic. Recalling that is torsion-free, the maximal elementary group is cyclic and thus is the centralizer of . By [BBF19, Lemma 2.1], since is residually finite, then the centralizer of any element is separable, i.e. the intersection of all finite index subgroups containing .
The next step is re-grouping appropriately the collections of quasi-lines in Lemma 5.14.
By [DK18, Theorem 7.51], the compatible collection of finite index subgroups from Lemma 5.12 determines a finite index group such that and for every vertex and edge .
By Bass-Serre theory, acts on the Bass-Serre tree of with finitely many vertex orbits. To be precise, let be the full set of vertex representatives.
Since for each , is of finite index, Lemma 5.14 implies that consists of finitely many -orbits, say ,
each of which satisfies projection axioms with projection constant .
Recall that acts on and . We now set so we have
We summarize the above discussion as the following.
Proposition 5.15.
For each , there exists a finite partition where such that for each , is –invariant and satisfies projection axioms with projection constant .
We are now ready to complete the proof of Theorem 1.1.
Proof of Theorem 1.1.
By [BBF19, Induction 2.2], if a finite index subgroup of has property (QT) then so does . Thus it suffices to show that has property (QT) where is the finite index subgroup of given by Corollary 5.13. By abuse of notations, we denote by , and we remark here that for the rest of the proof, results in Section 4 and Section 5 will apply for , but not for the original .
By Proposition 4.8, acts on the product so that the orbital map is quasi-isometrically embedded. Furthermore, there exists a –equivariant quasi-isometric embedding of each into the product of the cone-off space and a quasi-tree by Proposition 5.7. Therefore, it suffices to establish a –equivariant quasi-isometric embedding of into a finite product of quasi-trees.
By construction, each is –invariant and acts co-boundedly on , so there exists some independent of so that the union of quasi-lines in is –cobounded in .
Let and , we choose a point for such that . By –equivariance we define for any .
6. Finite height subgroups in a CKA action
In this section, we are going to prove Theorem 1.6 that basically says having finite height and strongly quasiconvexity are equivalent to each other in the context of CKA actions, and both properties can be characterized in term of their group elements. The heart of the proof of this theorem belongs to the implication where we use Sisto’s notion of path systems ([Sis18]). We first review some concepts finite height subgroups, strongly quasi-convex subgroups as well as some terminology in [Sis18].
Definition 6.1.
Let be a group and a subgroup of . We say that conjugates are essentially distinct if the cosets are distinct. We call has height at most in if the intersection of any essentially distinct conjugates is finite. The least for which this is satisfied is called the height of in .
Definition 6.2 (Strongly quasiconvex, [Tra19]).
A subset of a geodesic space is called strongly quasiconvex if for every there is some such that every –quasi–geodesic with endpoints on is contained in the –neighborhood of .
Let be a finitely generated group and a subgroup of . We say is strongly quasiconvex in if is a strongly quasi-convex subset in the Cayley graph for some (any) finite generating set . A group element in is Morse if is of infinite order and the cyclic subgroup generated by is strongly quasiconvex.
Remark 6.3.
The strong quasiconvexity of a subgroup does not depend on the choice of finite generating sets, and this notion is equivalent to quasiconvexity in the setting of hyperbolic groups. It is shown in [Tra19] (see Theorem 1.2) that strongly quasi-convex subgroups of a finitely generated group are finitely generated and have finite height.
The following proposition is cited from Proposition 2.3 and Proposition 2.6 in [NTY].
Proposition 6.4.
-
(1)
Let be a group such that the centralizer of is infinite. Let be a finite height infinite subgroup of . Then must have finite index in
-
(2)
Assume a group is decomposed as a finite graph of groups that satisfies the following.
-
(a)
For each vertex of each finite height subgroup of vertex group must be finite or have finite index in .
-
(b)
Each edge group is infinite.
Then, if is a finite height subgroup of G of infinite index, then is finite for each vertex group and each group element . In particular, if is torsion free, then is a free group.
-
(a)
Definition 6.5 (Path system, [Sis18]).
Let be a metric space. A path system in is a collection of –quasi-geodesic for some such that any subpath of a path in is in , and all pairs of points in can be connected by a path in .
Definition 6.6 (–contracting, [Sis18]).
Let be a metric space and let be a path system in . A subset of is called –contracting if there exists and a map such that
-
(1)
For any , then
-
(2)
For any such that then for any path in connecting to then and .
The map will be called –projection on with constant .
Lemma 6.7.
[Sis18, Lemma 2.8] Let be a –contracting subset of a metric space , then is strongly quasi-convex.
Theorem 6.8.
Let be a CKA action. Let be the collection of all special paths defined in Definition 3.6. Then is a path system.
Proof.
The proof follows from Proposition 3.8. ∎
For the rest of this section, we fix a CKA action and the action of on the associated Bass-Serre tree. We also fix the path system in Theorem 6.8.
To get into the proof of Theorem 1.6, we need several lemmas. The following lemma tells us that finite height subgroups in the CKA action are virtually free.
Lemma 6.9.
Let be a nontrivial finitely generated infinite index subgroup of . If has finite height in , then is finite for any and . In particular, is virtually free.
Proof.
Suppose that has finite height in . Since the centralizer each each vertex group is isomorphic to , it follows from Proposition 6.4 that for any and , the intersection is finite. Thus, acts properly on the tree and the stabilizer in of each vertex in is finite. It follows from [DK18, Theorem 7.51] that is virtually free. ∎
Remark 6.10.
Let be a nontrivial finitely generated infinite index subgroup of . Suppose that is a free group of finite rank and every nontrivial element in is not conjugate into any vertex group. Then there exists a subspace of such that acts geometrically on with respect to the induced length metric on . The subspace is constructed as the following.
Fix a vertex in , and fix a point in such that . Let be a generating set of . For each , let . Let be the geodesic in connecting to with . Let be the union of segment where varies elements of and .
Lemma 6.11.
Let be a –hyperbolic Hadamard space. Let and be two geodesic lines of such that . Let be a minimal geodesic segment between to . Then any geodesic segment running from to will pass within distance of both endpoints of . Moreover, when then we may take .
Lemma 6.12.
Given a constant , there exists a constant such that the following holds. Let and be two consecutive edges in with a common vertex . Let be a subset of such that . Suppose that and .
Let be the shortest path joining two lines to where and . For any and , let and be the projections of and into the lines and respectively. Then and .
Proof.
We recall that and acts properly and cocompactly on . Since is a nonelementary hyperbolic group, it follows that is a –hyperbolic space for some . Let be the maximum of the hyperbolicity constants of the ’s.
Let be the constant given by Lemma 6.11. Let . Since and , , it follows from Lemma 6.11 that there exist such that and . By the triangle inequality, we have . Since and are projection points of and into the factor of respectively, it follows that . Since and , it follows that . Hence . Thus, and .
By Lemma 6.11, we note that whenever the distance between two lines and is at least then we can define . We remark here that module there are only finitely many cases . Thus there are only finitely many up to the action of . Let be the maximum of these constants. ∎
Lemma 6.13.
Proof.
By the construction of , we note that there exists a constant such that for any vertex . Let be the constant given by Lemma 6.12.
Recall that we choose a –equivariant family of Euclidean planes . Let be the collection of strips in given by Section 2. For any three consecutive edges in the tree , two lines and in the plane determine an angle in . However, there are only finitely many angles shown up. We denote these angles by .
Let be the constant given by Lemma 2.4 such that for every vertex . Let
and
Let and be the initial and terminal points of . We note that and . We consider the following cases:
Case 1: . In this case, the special path is the geodesic in connecting to . Since and , it follows that . Thus, .
Case 2: . Since for any vertex , hence without losing of generality, we can assume that and . We recall the construction of the path from Definition 3.6. Let be the geodesic edge path connecting to and let be the intersection point of the strips and , where and . Then
Let and . In order to prove that , we only need to show that with .
The proofs for the cases and and for the cases are similar, so we only need to give the proofs for the cases .
Proof of case :
Let be the initial vertex of (with ), and be the terminal vertex of . We recall that two lines and in the plane are perpendicular. Since , we choose a point .
Claim: .
Proof of the claim.
Let be the projection of into the line . Let be the projection of into the line . By Lemma 6.12, we have
(19) |
(we note that ). Since and belong to and , it follows that . Let be the projection of into the factor of . We have that . Since is the minimal distance from to the line and we have that . Using the triangle inequality for three points , , and , we have
(20) |
Let be the projection of into the line . Using formula (19), we have
(21) |
Using the facts and we have . Since , it follows that .
Proof of the case :
Since , we choose a point . Since belong to and , we have . By a similar argument as in the proof of the claim of the case , we can show that . Thus, Since , it is easy to see that . ∎
We recall that an infinite order element in a finitely generated group is Morse if the cyclic subgroup generated by is strongly quasi-convex.
Lemma 6.14.
If an infinite order element in is More, then it is not conjugate into any vertex group of .
Proof.
Since is Morse, it follows that the infinite cyclic subgroup generated by is strongly quasi-convex in . We would like to show that is not conjugate into any vertex group. Indeed, by way of contradiction, we assume that for some and for some vertex group . Hence, the cyclic subgroup generated by is strongly quasi-convex in . Since is undistorted in (as acts geometrically on and is undistorted in ), it follows from Proposition 4.11 in [Tra19] that is strongly quasi-convex in . By Theorem 1.2 in [Tra19], has finite height in . Since the centralizer of is isomorphic to , it follows from Proposition 6.4 that has finite index in . This contradicts to the fact that is not virtually cyclic group. Therefore is not conjugate into any vertex group of . ∎
Proposition 6.15.
Let be a finitely generated free subgroup of of infinite index such that all nontrivial elements in are Morse in . Choose a vertex in a minimal –invariant subtree of . Let be the subspace of given by Remark 6.10 with respect to a generating set of . Then is contracting in . As a consequence, is strongly quasi-convex in .
Proof.
We first recall the construction of from Remark 6.10. We first fix a point in such that . For each , let . Let be the geodesic in connecting to with . Then is the union of segment where varies elements of and . Since is a free subgroup of and all nontrivial elements in are Morse in , it follows from Lemma 6.14 that every nontrivial element in is not conjugate into any vertex group . Hence, acts freely on the Bass-Serre tree . To show that (we note that ) is contracting in , we need to define a –projection satisfying conditions (1) and (2) in Definition 6.6.
Step 1: Constructing –projection on .
Let be the indexed -map given by Remark 2.6, which is coarsely -lipschitez. Let be the nearest point projection from to the minimal -invariant subtree . Since acts on the minimal tree cocompactly, it follows that there exists a constant such that .
Let be any point in . Choose an element such that . We define .
Step 2: Verifying the condition (1) in Definition 6.6. Recall that is the geodesic in connecting to . Let . Let be a constant such that for any vertex . Let be the constant given by Lemma 6.13. Since acts cocompactly both on and , there exists a constant such that for any and in then
(22) |
Claim 1: Let be a sufficiently large constant such that . Then for any .
Indeed, since , there exists such that . Recalling that is a -equivariant, coarsely -lipschitez map, and , we have
The nearest point projection of implies . Hence, . It implies that . Putting the above inequalities together with formula (22), we have
Claim 1 is confirmed.
Claim 2: Let be the constant given by Claim 1. Then the projection satisfies condition (2) in Definition 6.6 with respect to this constant .
Let and be two points in such that . Let be a special path in connecting to . We would like to show that and .
We recall that for any vertex . Thus we assume, without loss of generality that and . Recall that and where such that and are in the balls and respectively. We have . Hence . Since and , we have .
Choose vertices , in the geodesic such that and are the smallest integers bigger than or equal to . Thus . Similarly, we have .
Let be a special path in connecting to . We remark here that there exist a subpath of the special path and a subpath of such that and connect some point in to some point in . By Remark 3.7, we have that for any vertex in the geodesic .
Let be the constant given by Lemma 6.13. It follows from Lemma 6.13( see Case 1 and Case 2 in this lemma) that . Choose and such that . Since , and and , it follows that . Hence,
Similarly, we can show that . Thus Claim 2 is established.
Combining Step 1, Step 2, and Step 3 together, we conclude that is contracting in . By Lemma 6.7, is strongly quasi-convex in . We conclude that is strongly quasi-convex in . The proposition is proved. ∎
Proposition 6.15 has the following corollary that characterize Morse elements and contracting element in the admissible group .
Corollary 6.16.
-
(1)
An infinite order element in is Morse if and only if is not conjugate into any vertex group .
-
(2)
An element of is contracting with respect to if and only if its acts hyperbolically on the Bass-Serre tree .
Proof.
(1). By Lemma 6.14, if is Morse in then it is not conjugate into vertex group of . Conversely, if is not conjugate into any vertex group of then acts hyperbolically on the Bass-Serre tree . If acts hyperbolically on the Bass-Serre tree . Let be the infinite cyclic subgroup generated by . Since acts hyperbolically on the Bass-Serre tree , it is not conjugate into any vertex subgroup. Fix a point in , by Proposition 6.15 the orbit space is contracting in (because is contracting in ). Thus is a contracting element with respect to . By Lemma 2.8 in [Sis18], must be Morse.
(2). If acts hyperbolically on the Bass-Serre tree . Let be the infinite cyclic subgroup generated by . Since acts hyperbolically on the Bass-Serre tree , it is not conjugate into any vertex subgroup. Fix a point in , by Proposition 6.15 the orbit space is contracting in (because is contracting in ). Thus is a contracting element with respect to .
Conversely, if is contracting with respect to , then is Morse. By the assertion (1), is not conjugate into any vertex group. Thus it acts hyperbolically on the Bass-Serre tree . ∎
We now ready to prove Theorem 1.6.
Proof of Theorem 1.6.
We are going to prove the following implications: , , , and .
: The implication just follows from Theorem 1.2 in [Tra19].
: By Lemma 6.9, has a finite index subgroup such that is a free group of finite rank. Let be an infinite order element in . By way of contradiction, suppose that is not Morse in . By Corollary 6.16, is conjugate into a vertex group of . In other words, for some vertex group and for some . Hence that is finite by Proposition 6.4. So, has finite order that contradicts to our assumption that is an infinite order element. Since any infinite order element in has a power that belongs to , the implication is verified.
: Let be a finite index subgroup of such that is free. It follows from Proposition 6.15 that is strongly quasi-convex in . Since is a finite index subgroup of , it follows that is also strongly quasi-convex in .
: Let be a finite index subgroup of such that is free. Let be the minimal subtree of that contains . Since the map given by is quasi-isometric and the inclusion map is also a quasi-isometric embedding, it follows that the composition is a quasi-isometric embedding. Since is a finite index subgroup of , it follows that the map given by is a quasi-isometric embedding.
: By way of contradiction, suppose that there exists an infinite order element such that is not Morse in . It follows from Corollary 6.16 that is conjugate into a vertex group, hence fixes a vertex of . By our assumption, there is a vertex in such that the map given by is a quasi-isometric embedding. It implies that the map given by is also a –quasi-isometric embedding for some . Choose an integer large enough such that . We have . Hence . This contradicts to our choice of . ∎
References
- [Baj07] Jitendra Bajpai. Omnipotence of surface groups. Master’s thesis, McGill University, 2007.
- [BBF15] Mladen Bestvina, Ken Bromberg, and Koji Fujiwara. Constructing group actions on quasi-trees and applications to mapping class groups. Publ. Math. Inst. Hautes Études Sci., 122:1–64, 2015.
- [BBF19] Mladen Bestvina, Kenneth Bromberg, and Koji Fujiwara. Proper actions on finite products of quasi-trees. arXiv e-prints, page arXiv:1905.10813, May 2019.
- [Bes] Mladen Bestvina. Questions in geometric group theory. M. Bestvina’s home page, 2004.
- [BH99] Martin R. Bridson and André Haefliger. Metric spaces of non-positive curvature, volume 319 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1999.
- [BN08] Jason A. Behrstock and Walter D. Neumann. Quasi-isometric classification of graph manifold groups. Duke Math. J., 141(2):217–240, 2008.
- [Bow08] Brian H. Bowditch. Tight geodesics in the curve complex. Invent. Math., 171(2):281–300, 2008.
- [Bow12] B. H. Bowditch. Relatively hyperbolic groups. Internat. J. Algebra Comput., 22(3):1250016, 66, 2012.
- [CK00] Christopher B. Croke and Bruce Kleiner. Spaces with nonpositive curvature and their ideal boundaries. Topology, 39(3):549–556, 2000.
- [CK02] C. B. Croke and B. Kleiner. The geodesic flow of a nonpositively curved graph manifold. Geom. Funct. Anal., 12(3):479–545, 2002.
- [DGO17] F. Dahmani, V. Guirardel, and D. Osin. Hyperbolically embedded subgroups and rotating families in groups acting on hyperbolic spaces. Mem. Amer. Math. Soc., 245(1156):v+152, 2017.
- [DJ99] A. Dranishnikov and T. Januszkiewicz. Every Coxeter group acts amenably on a compact space. In Proceedings of the 1999 Topology and Dynamics Conference (Salt Lake City, UT), volume 24, pages 135–141, 1999.
- [DK18] Cornelia Druţu and Michael Kapovich. Geometric group theory, volume 63 of American Mathematical Society Colloquium Publications. American Mathematical Society, Providence, RI, 2018. With an appendix by Bogdan Nica.
- [DS05] Cornelia Druţu and Mark Sapir. Tree-graded spaces and asymptotic cones of groups. Topology, 44(5):959–1058, 2005. With an appendix by Denis Osin and Mark Sapir.
- [Far98] Benson Farb. Relatively hyperbolic groups. Geometric and Functional Analysis, 8:810–840, 1998.
- [GM14] Craig R. Guilbault and Christopher P. Mooney. Boundaries of Croke–Kleiner-admissible groups and equivariant cell-like equivalence. J. Topol., 7(3):849–868, 2014.
- [GMRS98] Rita Gitik, Mahan Mitra, Eliyahu Rips, and Michah Sageev. Widths of subgroups. Trans. Amer. Math. Soc., 350(1):321–329, 1998.
- [GP16] Victor Gerasimov and Leonid Potyagailo. Quasiconvexity in relatively hyperbolic groups. J. Reine Angew. Math., 710:95–135, 2016.
- [Gre16] Sebastian Grensing. Virtual boundaries of Hadamard spaces with admissible actions of higher rank. Math. Z., 284(1-2):1–22, 2016.
- [HP15] Mark F. Hagen and Piotr Przytycki. Cocompactly cubulated graph manifolds. Israel J. Math., 207(1):377–394, 2015.
- [HS13] David Hume and Alessandro Sisto. Embedding universal covers of graph manifolds in products of trees. Proc. Amer. Math. Soc., 141(10):3337–3340, 2013.
- [HW08] Frédéric Haglund and Daniel T. Wise. Special cube complexes. Geom. Funct. Anal., 17(5):1551–1620, 2008.
- [KL98] M. Kapovich and B. Leeb. -manifold groups and nonpositive curvature. Geom. Funct. Anal., 8(5):841–852, 1998.
- [Liu13] Yi Liu. Virtual cubulation of nonpositively curved graph manifolds. J. Topol., 6(4):793–822, 2013.
- [NTY] Hoang Thanh Nguyen, Hung Cong Tran, and Wenyuan Yang. Quasiconvexity in 3-manifold groups. arXiv:1911.07807. To appear in Mathematische Annalen.
- [Osi16] D. Osin. Acylindrically hyperbolic groups. Trans. Amer. Math. Soc., 368(2):851–888, 2016.
- [PW14] Piotr Przytycki and Daniel T. Wise. Graph manifolds with boundary are virtually special. J. Topol., 7(2):419–435, 2014.
- [Sis13] Alessandro Sisto. Projections and relative hyperbolicity. Enseign. Math. (2), 59(1-2):165–181, 2013.
- [Sis18] Alessandro Sisto. Contracting elements and random walks. J. Reine Angew. Math., 742:79–114, 2018.
- [Tra19] Hung Cong Tran. On strongly quasiconvex subgroups. Geom. Topol., 23(3):1173–1235, 2019.
- [Wis00] Daniel T. Wise. Subgroup separability of graphs of free groups with cyclic edge groups. Q. J. Math., 51(1):107–129, 2000.
- [Wis20] Daniel T. Wise. The Structure of Groups with a Quasiconvex Hierarchy, volume AMS-209. Annals of Mathematics Studies, 2020.
- [Yan19] Wen-yuan Yang. Statistically convex-cocompact actions of groups with contracting elements. Int. Math. Res. Not. IMRN, (23):7259–7323, 2019.