Largest acylindrical actions and Stability in hierarchically hyperbolic groups

By Carolyn Abbott, Jason Behrstock, and Matthew Gentry Durham, With an appendix by Daniel Berlyne and Jacob Russell

Abstract

We consider two manifestations of non-positive curvature: acylindrical actions (on hyperbolic spaces) and quasigeodesic stability. We study these properties for the class of hierarchically hyperbolic groups, which is a general framework for simultaneously studying many important families of groups, including mapping class groups, right-angled Coxeter groups, most –manifold groups, right-angled Artin groups, and many others.

A group that admits an acylindrical action on a hyperbolic space may admit many such actions on different hyperbolic spaces. It is natural to try to develop an understanding of all such actions and to search for a “best” one. The set of all cobounded acylindrical actions on hyperbolic spaces admits a natural poset structure, and in this paper we prove that all hierarchically hyperbolic groups admit a unique action which is the largest in this poset. The action we construct is also universal in the sense that every element which acts loxodromically in some acylindrical action on a hyperbolic space does so in this one. Special cases of this result are themselves new and interesting. For instance, this is the first proof that right-angled Coxeter groups admit universal acylindrical actions.

The notion of quasigeodesic stability of subgroups provides a natural analogue of quasiconvexity which can be considered outside the context of hyperbolic groups. In this paper, we provide a complete classification of stable subgroups of hierarchically hyperbolic groups, generalizing and extending results that are known in the context of mapping class groups and right-angled Artin groups. Along the way, we provide a characterization of contracting quasigeodesics; interestingly, in this generality the proof is much simpler than in the special cases where it was already known.

In the appendix, it is verified that any space satisfying the a priori weaker property of being an “almost hierarchically hyperbolic space” is actually a hierarchically hyperbolic space. The results of the appendix are used to streamline the proofs in the main text.

1. Introduction

Hierarchically hyperbolic groups were recently introduced by Behrstock, Hagen, and Sisto Reference BHS17b to provide a uniform framework in which to study many important families of groups, including mapping class groups of finite type surfaces, right-angled Coxeter groups, most –manifold groups, right-angled Artin groups and many others. A hierarchically hyperbolic space (HHS) consists of: a quasigeodesic space, ; a set of domains, , which index a collection of –hyperbolic spaces to which projects; and, some additional information about these projections, including, for instance, a partial order on the domains and a unique largest element in that order, which we denote by (i.e., is comparable to and larger than every other domain in ).

Largest acylindrical actions

The study of acylindrical actions on hyperbolic spaces, as initiated in its current form by Osin Reference Osi16 building on earlier work of Sela Reference Sel97 and Bowditch Reference Bow08, has proven to be a powerful tool for studying groups with some aspects of non-positive curvature. As established in Reference BHS17b, non-virtually cyclic hierarchically hyperbolic groups admit non-elementary acylindrical actions when the –hyperbolic space associated to the maximal element in has infinite diameter, a property which holds in all the above examples except for those that are direct products.

Any given group with an acylindrical action may actually admit many acylindrical actions on many different spaces. A natural question is to try and find a “best” acylindrical action. There are different ways that one might try to optimize the acylindrical action. For instance, the notion of a universal acylindrical action, for a given group , is an acylindrical action on a hyperbolic space such that every element of which acts loxodromically in some acylindrical action on some hyperbolic space, must act loxodromically in its action on . As established by Abbott, there exist finitely generated groups which admit acylindrical actions, but no universal acylindrical action Reference Abb16; we also note that universal actions need not be unique Reference ABO19.

In Reference ABO19, Abbott, Balasubramanya, and Osin introduce a partial order on cobounded acylindrical actions which, in a certain sense, encodes how much information the action provides about the group. When there exists an element in this partial ordering which is comparable to and larger than all other elements it is called a largest action. By construction, any largest action is necessarily a universal acylindrical action and unique.

In this paper we construct a largest action for every hierarchically hyperbolic group. Special cases of this theorem recover some recent results of Reference ABO19, as well as a number of new cases. For instance, in the case of right-angled Coxeter groups (and more generally for special cubulated groups), even the existence of a universal acylindrical action was unknown. Further, outside of the relatively hyperbolic setting, our result provides a single construction that simultaneously covers these new cases as well as all previously known largest and universal acylindrical actions of finitely presented groups. The following summarizes the main results of Section 5 (where there are also further details on the background and comparison with known results).

Theorem A (HHG have actions that are largest and universal).

Every hierarchically hyperbolic group admits a largest acylindrical action. In particular, the following admit acylindrical actions which are largest and universal:

(1)

Hyperbolic groups.

(2)

Mapping class groups of orientable surfaces of finite type.

(3)

Fundamental groups of compact three-manifolds with no Nil or Sol component in their prime decomposition.

(4)

Groups that act properly and cocompactly on a special CAT(0) cube complex, and more generally any cubical group which admits a factor system. This includes right-angled Artin groups, right-angled Coxeter groups, and many other examples as in Reference HS16.

We use this construction of a largest action to characterize stable subgroups (Theorem B) and contracting elements (Corollary 5.5) of hierarchically hyperbolic groups, and to describe random subgroups of hierarchically hyperbolic groups (Theorem E).

Stability in hierarchically hyperbolic groups

One of the key features of a Gromov hyperbolic space is that every geodesic is uniformly Morse, a property also known as (quasigeodesic) stability; that is, any uniform quasigeodesic beginning and ending on a geodesic must lie uniformly close to it. In fact, any geodesic metric space in which each geodesic is uniformly Morse is hyperbolic.

In the context of geodesic metric spaces, the presence of Morse geodesics has important structural consequences for the space; for instance, any asymptotic cone of such a space has global cut points Reference DMS10. Moreover, quasigeodesic stability in groups is quite prevalent, since any finitely generated acylindrically hyperbolic group contains Morse geodesics Reference Osi16Reference Sis16.

There has been much interest in developing alternative characterizations Reference DMS10Reference CS15Reference ACGH17Reference ADT17 and understanding this phenomenon in various important contexts Reference Min96Reference Beh06Reference DMS10Reference DT15Reference ADT17. This includes the theory of Morse boundaries, which encode all Morse geodesics of a group Reference CS15Reference Cor17Reference CH17Reference CD19Reference CM19. In Reference DT15, Durham and Taylor generalized the notion of stability to subspaces and subgroups.

In this paper, we obtain a complete characterization of stability in hierarchically hyperbolic groups.

Let be an HHS. We say that a subset has –bounded projections when for all non-maximal ; when the constant does not matter, we simply say the subset has uniformly bounded projections.

Theorem B (Equivalent conditions for subgroup stability).

Any hierarchically hyperbolic group admits a hierarchically hyperbolic group structure such that for any finitely generated , the following are equivalent:

(1)

is stable in ;

(2)

is undistorted in and has uniformly bounded projections;

(3)

Any orbit map is a quasi-isometric embedding, where is the –maximal element in .

Theorem B generalizes some previously known results. In the case of mapping class groups: Reference Beh06 proved that 2 implies 1 for cyclic subgroups; Reference DT15 proved equivalence of 1 and 3; equivalence of 2 and 3 follows from the distance formula; moreover, Reference KL08Reference Ham yield that these conditions are also equivalent to convex cocompactness in the sense of Reference FM02. The case of right-angled Artin groups was studied in Reference KMT17, where they prove equivalence of 1 and 3.

Section 6 contains a more general version of Theorem B, as well as further applications, including Theorem 6.6, which concerns the Morse boundary of hierarchically hyperbolic groups and proves that all hierarchically hyperbolic groups have finite stable asymptotic dimension.

On purely loxodromic subgroups

In the mapping class group setting Reference BBKL20 proved that the conditions in Theorem B are also equivalent to being undistorted and purely pseudo-Anosov. Similarly, in the right-angled Artin group setting, it was proven in Reference KMT17 that 1 and 3 are each equivalent to being purely loxodromic.

Subgroups of right-angled Coxeter groups all of whose elements act loxodromically on the contact graph were studied in the recent preprint Reference Tra, Theorem 1.4, which proved that property is equivalent to 3. Since there often exist Morse elements in a right-angled Coxeter group which do not act loxodromically on the contact graph (which plays the role of in the standard HHG structure on the group), his condition is not equivalent to 1. It is the ability to change the hierarchically hyperbolic structure as we do in Theorem 3.7, discussed below, which allows us to prove our more general result which characterizes all stable subgroups, not just the ones acting loxodromically on the contact graph.

Mapping class groups and right-angled Artin groups have the property that in their standard hierarchically hyperbolic structure they admit a universal acylindrical action on , where is the hyperbolic space associated to the –maximal domain . On the other hand, right-angled Coxeter groups often don’t admit universal acylindrical actions on in their standard structure. Accordingly, we believe the following questions are interesting. The first item would generalize the situation in the mapping class group as established in Reference BBKL20, and the second item for right-angled Artin groups would generalize results proven in Reference KMT17, and for right-angled Coxeter groups would generalize results in Reference Tra. If the second item is true for the mapping class group, this would resolve a question of Farb–Mosher Reference FM02. See also Reference ADT17, Question 1.

Question C.

Let be a hierarchically hyperbolic group which admits a universal acylindrical action on , where is the –maximal element in . Let be a finitely generated subgroup of .

Are the conditions in Theorem B also equivalent to: is undistorted and acts purely loxodromically on ?

Under what hypotheses on , are the conditions in Theorem B also equivalent to: acts purely loxodromically on ?

Note that in the context of Question C, an element acts loxodromically on if and only if it has positive translation length. This holds since the action is acylindrical and thus each element either acts elliptically or loxodromically.

In an early version of this paper, we asked if the second part of Question C held for all hierarchically hyperbolic groups. In the general hierarchically hyperbolic setting, however, the undistorted hypothesis is necessary, as pointed out to us by Anthony Genevois with the following example. The necessity is shown by Brady’s example of a torsion-free hyperbolic group with a finitely presented subgroup which is not hyperbolic Reference Bra99. This subgroup is torsion-free and thus purely loxodromic. But, a subgroup of a hyperbolic group is stable if and only if it is quasiconvex. Thus, since this subgroup is not quasiconvex, we see that being purely loxodromic is strictly weaker than the conditions of Theorem B.

New hierarchically hyperbolic structures

In order to establish the above results, we provide some new structural theorems about hierarchically hyperbolic spaces.

One of the key technical innovations in this paper is provided in Section 3. There we prove Theorem 3.7 which allows us to modify a given hierarchically hyperbolic structure by removing for some and, in their place, enlarging the space . For instance, this is how we construct the space on which a hierarchically hyperbolic group has its largest acylindrical action.

Another important tool is Theorem 4.4 which provides a simple characterization of contracting geodesics in a hierarchically hyperbolic space.

The following is a restatement of that result in the case of groups (see Theorem 4.4 for the precise statement):

Theorem D (Characterization of contracting quasigeodesics).

Let be a hierarchically hyperbolic group. For any there exist uniform constants depending only on and such that the following holds for every –quasigeodesic : the quasigeodesic is uniformly contracting if and only if has uniformly bounded projections (in any structure with unbounded products, e.g., in one as provided by Corollary 3.8).

Since the presence of a contracting geodesic implies the group has at least quadratic divergence, an immediate consequence of Theorem D is that any hierarchically hyperbolic group has quadratic divergence whenever projects to an infinite diameter subset of .

As a sample application of Theorem D and using work of Taylor–Tiozzo Reference TT16, we prove the following in Section 6.4 as Theorem 6.8.

Theorem E (Random subgroups are stable).

Let be an HHS for which has infinite diameter, where is the –maximal element, and consider which acts properly and cocompactly on . Then any –generated random subgroup of stably embeds in via the orbit map.

We note that one immediate consequence of this result is a new proof of a theorem of Maher–Sisto: any random subgroup of a hierarchically hyperbolic group which is not the direct product of two infinite groups is stable Reference MS19. The mapping class group and right-angled Artin group cases of this result were first established in Reference TT16.

Finally, at the end of the paper we discuss a technical condition on hierarchically hyperbolic structures, called having clean containers. While in Proposition 7.2 this hypothesis is shown to hold for many groups, it does not hold in all cases. This condition was used in earlier versions of this paper in which it was assumed for the proof of Theorem 3.7, and then the general result was bootstrapped from there. In light of Theorem A.1 in the Appendix, this property is no longer required for this paper. We keep the contents of this section in the paper nonetheless, since they have found independent interest and already been used elsewhere, e.g., Reference BRReference HS16Reference Rus20, as well as in several papers in progress.

2. Background

We begin by recalling some preliminary notions about metric spaces, maps between them, and group actions. Given metric spaces , we use to denote the distance functions in , respectively. A map is –Lipschitz if there exists a constant such that for every , ; it is –coarsely Lipschitz if The map is a –quasi-isometric embedding if there exist constants and such that for all ,

If, in addition, is contained in the –neighborhood of , then is a –quasi-isometry. For any interval , the image of an isometric embedding is a geodesic and the image of a –quasi-isometric embedding is a –quasigeodesic.

If any two points in can be connected by a –quasigeodesic, then we say is a –quasigeodesic space. If , we may simply say that is a –quasigeodesic space. A subspace is –quasi-convex if there exists a constant such that any geodesic in connecting points in is contained in the –neighborhood of . For all of the above notions, if the particular constants are not important, we may drop them and simply say, for example, that a map is a quasi-isometry.

Throughout this paper, we will assume that all group actions are by isometries. The action of a group on a metric space , which we denote by , is proper if for every bounded subset , the set is finite. The action is cobounded (respectively, cocompact) if there exists a bounded (respectively, compact) subset such that . If a group acts on metric spaces and , we say a map is –equivariant if for every and every we have . A quasi-action of on associates to each a quasi-isometry of with uniform quasi-isometry constants, such that is within uniformly bounded distance of .

2.1. Hierarchically hyperbolic spaces

In this section we recall the basic definitions and properties of hierarchically hyperbolic spaces as introduced in Reference BHS17bReference BHS19.

Definition 2.1 (Hierarchically hyperbolic space).

A –quasigeodesic space is said to be hierarchically hyperbolic if there exists , an index set , and a set of –hyperbolic spaces , such that the following conditions are satisfied:

(1)

(Projections.) There is a set of projections sending points in to sets of diameter bounded by some in the various . Moreover, there exists so that each is –coarsely Lipschitz and is –quasiconvex in .

(2)

(Nesting.) is equipped with a partial order , and either or contains a unique –maximal element which is larger than all other elements; when , we say is nested in . For each , we denote by the set of such that . Moreover, for all with there is a specified subset with . There is also a projection .

(3)

(Orthogonality.) has a symmetric and anti-reflexive relation called orthogonality: we write when are orthogonal. Also, whenever and , we require that . Finally, we require that for each and each for which , there exists , so that whenever and , we have ; we say is a container associated with and . Finally, if , then are not –comparable.

(4)

(Transversality and consistency.) If are not orthogonal and neither is nested in the other, then we say are transverse, denoted . There exists such that if , then there are sets and each of diameter at most and satisfying:

for all .

For satisfying and for all , we have:

Finally, if , then whenever satisfies either or and .

(5)

(Finite complexity.) There exists , the complexity of (with respect to ), so that any set of pairwise––comparable elements has cardinality at most .

(6)

(Large links.) There exist and such that the following holds. Let and let . Let . Then there exists such that for all , either for some , or . Also, for each .

(7)

(Bounded geodesic image.) For all , all , and all geodesics of , either or .

(8)

(Partial realization.) There exists a constant with the following property. Let be a family of pairwise orthogonal elements of , and let . Then there exists so that:

for all ,

for each and each with , we have , and

if for some , then .

(9)

(Uniqueness.) For each , there exists such that if and , then there exists such that .

Notation 2.2.

Note that below we will often abuse notation by simply writing or to refer to the entire package of an hierarchically hyperbolic structure, including all the associated spaces, projections, and relations given by the above definition.

Notation 2.3.

When writing distances in for some , we often simplify the notation slightly by suppressing the projection map , i.e., given and we write for and for . Note that when we measure distance between a pair of sets (typically both of bounded diameter) we are taking the minimum distance between the two sets. For distance/diameter, if the space in which the measurement is being made is not clear from the context, we will denote it by a subscript. Given and we let denote .

Remark 2.4.

In the setting of hierarchically hyperbolic spaces, we often encounter maps which are well-defined only up to uniformly bounded error, in the following sense. Given a map between quasi-geodesic spaces , there may be multiple possible points in that one could define as for a particular . If the diameter of such possible points is uniformly bounded in over all , then we say that the map is coarsely well-defined, since we could arbitrarily make a choice for each and the map would be well-defined up to uniformly bounded error. For example, gives a coarsely well-defined map .

An important consequence of being a hierarchically hyperbolic space is the following distance formula, which relates distances in to distances in the hyperbolic spaces for . The notation means include in the sum if and only if .

Theorem 2.5 (Distance formula for HHS; Reference BHS19).

Let be a hierarchically hyperbolic space. Then there exists such that for all , there exist so that for all ,

We now recall an important construction of subspaces in a hierarchically hyperbolic space called standard product regions introduced in Reference BHS17b, Section 13 and studied further in Reference BHS19. First we define a consistent tuple, which will be used to define the two factors in the product space.

Definition 2.6 (Consistent tuple).

Fix , and let be a tuple such that for each , the coordinate is a subset of with . The tuple is –admissible if for all . The –admissible tuple is –consistent if, whenever ,

and whenever ,

Definition 2.7 (Nested partial tuple ()).

Recall . Fix and let be the set of –consistent tuples in .

Definition 2.8 (Orthogonal partial tuple () ).

Let , where is a –minimal element such that for all (note that exists by the container axiom for an HHS, i.e., Definition 2.1.3). Fix , let be the set of –consistent tuples in .

Definition 2.9 (Product regions in ).

Given and , there is a coarsely well-defined map which restricts to coarsely well-defined maps . Indeed, for each , and each , the projection is defined as follows. If , then . If , then . If , then . Finally, if , and , let . The tuple is –consistent (see Reference BHS19, Construction 5.10), and therefore Reference BHS19, Theorem 3.1 provides a point such that for all . Moreover, the point is coarsely unique in the sense that the set of all which satisfy for each has diameter at most in . We define ; the coarse uniqueness of shows that this map is coarsely well-defined. Fixing any yields a map , and is defined analogously. We refer to as a product region, which we denote .

We often abuse notation slightly and use the notation , and to refer to the image in of the associated set. In Reference BHS19, Construction 5.10 it is proven that these standard product regions have the property that they are “hierarchically quasiconvex subsets” of . We leave out the definition of hierarchical quasiconvexity, because its only use here is that product regions have “gate maps,” as given by the following in Reference BHS19, Lemma 5.5:

Lemma 2.10 (Existence of coarse gates; Reference BHS19, Lemma 5.5).

If is –hierarchically quasiconvex and non-empty, then there exists a gate map for , i.e., for each there exists such that for all , the set (uniformly) coarsely coincides with the projection of to the –quasiconvex set . The point is called the gate of in .

Remark 2.11 (Surjectivity of projections).

As one can always change the hierarchical structure so that the projection maps are coarsely surjective Reference BHS19, Remark 1.3, we will assume that is such a structure. That is, for each , if is not surjective, then we identify with .

We also need the notion of a hierarchy path, whose existence was proven in Reference BHS19, Theorem 4.4 (although we use the word path, since they are quasi-geodesics, typically we consider them as discrete sequences of points):

Definition 2.12.

For , a path in is a –hierarchy path if

(1)

is a –quasigeodesic,

(2)

for each , is an unparametrized –quasigeodesic. An unbounded hierarchy path is a hierarchy ray.

We call a domain relevant to a pair of points, if the projections to that domain are larger than some fixed (although possibly unspecified) constant depending only on the hierarchically hyperbolic structure. We say a domain is relevant for a particular quasi-geodesic if it is relevant for the endpoints of that quasi-geodesic.

Proposition 2.13 (Reference BHS19, Proposition 5.17).

There exists such that for all , all with relevant for , and all –hierarchy paths joining to , there is a subpath of with the following properties:

(1)

;

(2)

is coarsely constant on for all , i.e., it is a uniformly bounded distance from a constant map.

Remark 2.14.

Let , and suppose is relevant for . As and consist of –consistent tuples (for a fixed ) and is only coarsely well-defined, by appropriately increasing to accomodate for the chosen constant in Proposition 2.13, we may assume that is actually a subset of .

It is often convenient to work with equivariant hierarchically hyperbolic structures, we now recall the relevant structures for doing so. For details see Reference BHS19.

Definition 2.15 (Hierarchically hyperbolic groups).

Let be a hierarchically hyperbolic space. An automorphism of consists of a map , together with a bijection and, for each , an isometry so that the following diagrams commute up to uniformly bounded error whenever the maps in question are defined (i.e., when , are not orthogonal):

and

Two automorphisms , are equivalent if and for all we have . The set of all such equivalence classes forms the automorphism group of , denoted . A finitely generated group is said to be a hierarchically hyperbolic group (HHG) if there is a hierarchically hyperbolic space and a group homomorphism so that the induced uniform quasi-action of on is metrically proper, cobounded, and contains finitely many –orbits. Note that when is a hyperbolic group then, with respect to any word metric, it inherits a hierarchically hyperbolic structure.

2.2. Acylindrical actions

We recall the basic definitions related to acylindrical actions; the canonical references are Reference Bow08 and Reference Osi16. We also discuss a partial order on these actions which was recently introduced in Reference ABO19.

Definition 2.16 (Acylindrical).

The action of a group on a metric space is acylindrical if for any there exist such that for all with ,

Recall that given a group acting on a hyperbolic metric space , an element is loxodromic if the map defined by is a quasi-isometric embedding for some (equivalently any) . However, an element of may be loxodromic for some actions and not for others. Consider, for example, the free group on two generators acting on its Cayley graph and acting on the Bass-Serre tree associated to the splitting . In the former action, every non-trivial element is loxodromic, while in the latter action, no powers of and are loxodromic.

Definition 2.17 (Generalized loxodromic).

An element of a group is called generalized loxodromic if it is loxodromic for some acylindrical action of on a hyperbolic space.

Remark 2.18.

By Reference Osi16, Theorem 1.1, every acylindrical action of a group on a hyperbolic space either has bounded orbits or contains a loxodromic element. By Reference Osi16, Theorem 1.4.(L4) and Sisto Reference Sis16, Theorem 1, every generalized loxodromic element is Morse, i.e., every quasi-geodesic with endpoints on the axis of the element lies uniformly close to that axis (see Definition 2.22). Therefore, if a group does not contain any Morse elements, it does not contain any generalized loxodromics, and thus must have bounded orbits in every acylindrical action on a hyperbolic space. This is the case when, for example, is a non-trivial direct product, that is, a direct product of two infinite groups.

Definition 2.19 (Universal acylindrical action).

An acylindrical action of a group on a hyperbolic space is a universal acylindrical action if every generalized loxodromic element is loxodromic. Such an action is sometimes called a loxodromically universal action.

Notice that if every acylindrical action of a group on a hyperbolic space has bounded orbits, then does not contain any generalized loxodromic elements, and the action of on a point (which is acylindrical) is a universal acylindrical action.

The following notions are discussed in detail in Reference ABO19. We give a brief overview here. Fix a group . Given a (possibly infinite) generating set of , let denote the word metric with respect to , and let be the Cayley graph of with respect to the generating set . Given two generating sets and , we say is dominated by and write if

Note that when , the action provides more information about the group than , and so, in some sense, is a “larger” action. The two generating sets and are equivalent if and ; when this happens we write .

Let be the set of equivalence classes of generating sets of such that is hyperbolic and the action is acylindrical. We denote the equivalence class of by . The preorder induces an order on , which we also denote .

Definition 2.20 (Largest).

We say an equivalence class of generating sets is largest if it is the largest element in under this ordering.

Given a cobounded acylindrical action of on a hyperbolic space , a Milnor–Svarc argument gives a (possibly infinite) generating set of such that there is a –equivariant quasi-isometry between and . By a slight abuse of language, we will say that a particular cobounded acylindrical action on a hyperbolic space is largest, when, more precisely, it is the equivalence class of the generating set associated to this action through the above correspondence, , that is the largest element in .

Remark 2.21.

By definition, every largest acylindrical action is a universal acylindrical action. To see this, notice that if , then the set of loxodromic elements in must be a subset of the set of loxodromic elements in .

2.3. Stability

Stability is strong coarse convexity property which generalizes quasiconvexity in hyperbolic spaces and convex cocompactness in mapping class groups Reference DT15. In the general context of metric spaces, it is essentially the familiar Morse property generalized to subspaces, so we begin there.

Definition 2.22 (Morse/stable quasigeodesic).

Let be a metric space. A quasigeodesic is called Morse (or stable) if there exists a function such that if is a –quasigeodesic in with endpoints on , then

We call the stability gauge for and say is –stable if we want to record the constants.

We can now define a notion of stable embedding of one metric space in another which is equivalent to the one introduced by Durham and Taylor Reference DT15:

Definition 2.23 (Stable embedding).

We say a quasi-isometric embedding between quasigeodesic metric spaces is a stable embedding if there exists a stability gauge such that for any quasigeodesic constants and any –quasigeodesic , then is an –stable quasigeodesic in . We say a subset is stable if it is undistorted and the inclusion map is a stable embedding.

The following generalizes the notion of a Morse quasigeodesic to subgroups:

Definition 2.24 (Subgroup stability).

Let be a subgroup of a finitely generated group . We say is a stable subgroup of if some (equivalently, any) orbit map of into some (any) Cayley graph (with respect to a finite generating set) of is a stable embedding.

If for some , is stable, then we call stable. Such elements are often called Morse elements.

Stability of a subset is preserved under quasi-isometries. Note that stable subgroups are undistorted in their ambient groups and, moreover, they are quasiconvex with respect to any choice of finite generating set for the ambient group.

3. Altering the hierarchically hyperbolic structure

The goal of this section is to prove that any hierarchically hyperbolic space satisfying a technical assumption—the bounded domain dichotomy—admits a hierarchically hyperbolic structure with unbounded products, i.e., every non-trivial product region in the ambient space has unbounded factors; see Theorem 3.7 below.

In particular, this establishes that all hierarchically hyperbolic groups admit a hierarchically hyperbolic group structure with unbounded products. It is for this reason that our complete characterization of the contracting property in spaces with unbounded products in Section 4 yields a characterization of the contracting property for all hierarchically hyperbolic groups, as stated in Theorem D.

3.1. Unbounded products

Fix a hierarchically hyperbolic space .

Let and let be the set of domains such that there exists and satisfying: , , and .

Recall that a set of domains is closed under nesting if whenever and , then .

Lemma 3.1.

For any , the set is closed under nesting.

Proof.

Let and . By definition of , there exists with and satisfying: and there exists such that . Since , it follows that , as desired.

Definition 3.2 (Bounded domain dichotomy).

We say has the –bounded domain dichotomy if there exists such that any with satisfies . If the value of is not important, we simply refer to the bounded domain dichotomy.

Recall that for every hierarchically hyperbolic group , the set of domains contains finitely many –orbits and each induces an isometry for each (see Definition 2.15). It thus follows that every hierarchically hyperbolic group has the bounded domain dichotomy. (Also, note that this property implies the space is “asymphoric” as defined in Reference BHS17c.)

Definition 3.3 (Unbounded products).

We say that a hierarchically hyperbolic space has unbounded products if it has the bounded domain dichotomy and the property that if has , then .

3.2. Almost hierarchically hyperbolic spaces

In this section we introduce a tool for verifying a space is hierarchically hyperbolic.

The following is a weaker version of the orthogonality axiom:

(Bounded pairwise orthogonality) has a symmetric and anti-reflexive relation called orthogonality: we write when are orthogonal. Also, whenever and , we require that . Moreover, if , then are not –comparable. Finally, the cardinality of any collection of pairwise orthogonal domains is uniformly bounded by .

By Reference BHS19, Lemma 2.1, the orthogonality axiom (Definition 2.1, (3)) for an hierarchically hyperbolic structure implies axiom . However, the converse does not hold; that is, the last condition of does not directly imply the container statement in (3), and thus this is an a priori strictly weaker assumption. However, as is proven in the appendix in Theorem A.1, this weakened version of the axiom is enough to produce a hierarchically hyperbolic structure.

We now introduce the notion of an almost hierarchically hyperbolic space:

Definition 3.4 (Almost HHS).

If satisfies all axioms of a hierarchically hyperbolic space except (3) and additionally satisfies axiom , then is an almost hierarchically hyperbolic space.

In the appendix, Berlyne and Russell prove Theorem A.1, establishing that if a space is almost hierarchically hyperbolic, then the associated structure can be modified to obtain a hierarchically hyperbolic structure on the original space. This result is used in our proof of Theorem 3.7.

3.3. A new hierarchically hyperbolic structure

In this section we describe a new hierarchically hyperbolic structure on hierarchically hyperbolic spaces with the bounded domain dichotomy. We first describe the hyperbolic spaces that will be part of the new structure.

Let be a hierarchically hyperbolic space with the –bounded domain dichotomy. Recall that we define to be the set of such that there exists with for which there exists a satisfying . For each , define similarly.

Remark 3.5 (Factored spaces).

As defined in Reference BHS17a, given and the factored space is the space obtained from by coning-off each for all and all . Sometimes we abuse language slightly and refer to this as the factored space obtained from by collapsing . In particular, when is the –maximal element of , then can be taken to be the space , which is obtained from by coning-off for all and all .

We often consider the case of a fixed and and then apply this construction to the hierarchically hyperbolic structure . For this application, note that is quasi-isometric to , by Reference BHS17a, Corollary 2.9, and thus so is , by Remark 2.11.

Lemma 3.6.

Let be a hierarchically hyperbolic space and consider which is closed under nesting. Let be a hierarchy path in . Then, the path obtained by including is an unparametrized quasi-geodesic. Moreover, if for each which is a relevant domain for and for each , we modify the path through by removing all but the first and last vertex of the hierarchy path which passes through , then the new path obtained, is a hierarchy path for .

Proof.

The proof is by induction on complexity. Consider all the nest-minimal elements which are relevant for ; by Proposition 2.13 and Remark 2.14 for each such there is a subpath of which passes through a collection of slices within the product region associated to . By Reference BHS19, Lemma 2.14 there is a bounded (in terms of ) coloring of with the property that all the domains of a given color are pairwise transverse. Starting from , we take one color at a time, together with all the domains nested inside domains of that color, and create the factored space by coning off those domains. At each step, we obtain a new hierarchically hyperbolic space with the property that in this space the relevant domains for are exactly the original ones except for those in the colors we have coned off thus far. Since this path still travels monotonically through each of the relevant domains, it is an unparametrized quasi-geodesic in the new factored space. Thus the path is a parametrized quasi-geodesic and thus a hierarchy path in the new factored space (with constants depending only on the constants for the original hierarchy path). Once the colors of are exhausted, repeat one step up the nesting lattice. Since the complexity of the hierarchically hyperbolic structure and the coloring are both bounded, this will terminate after finitely many steps. Finally we cone off any domains in which are not relevant for to obtain the space . Through this final step remains a uniform quality hierarchy path since it is still a quasigeodesic.

The next result uses the above spaces to obtain a hierarchically hyperbolic structure with particularly nice properties from a given hierarchically hyperbolic structure.

Theorem 3.7.

Every hierarchically hyperbolic space with the bounded domain dichotomy admits a hierarchically hyperbolic structure with unbounded products.

Proof.

Let be a hierarchically hyperbolic space. Let denote the –maximal element together with the subset of consisting of all with both and unbounded.

We begin to define our new hierarchically hyperbolic structure on by taking as our index set. For each we set the associated hyperbolic space to be . For the top-level domain, , we obtain a hyperbolic space, , as follows. By Lemma 3.1, is closed under nesting and hence is a hierarchically hyperbolic space. Moreover, since this hierarchically hyperbolic space has the property that no pair of orthogonal domains both have diameter larger than , by Reference BHS17c, Corollary 2.16 it is hyperbolic for some constant depending only on and ; we call this space .

To avoid confusion, we use the notation for distance in and the notation for distance in .

When , the projections are as defined in the original hierarchically hyperbolic space. We take the projection to be the factor map . If and , then the relative projections are defined as in . For the remaining cases the relative projections are as follows: is defined to be and is defined to be the image of under the factor map .

We now check the axioms to verify that is an almost hierarchically hyperbolic space (i.e., all the conditions of a hierarchically hyperbolic space except for a weakened version of the orthogonality axiom). Once these axioms have been verified, we can then invoke Theorem A.1 to conclude that the almost hierarchically hyperbolic structure can be modified to yield an actual hierarchically hyperbolic space. By construction, satisfies the hypothesis of Corollary A.8, and therefore the associated modified hierarchically hyperbolic structure will have unbounded products, as desired.

Projections: The only case to check is for the top-level domain . Since is a factor map, it is coarsely Lipschitz and coarsely surjective.

Nesting: The partial order and projections are given by construction. The diameter bound in the case of nesting projections is immediate from the bound from , except in the case of for . The bound on the diameter of follows from the construction of as a factor space and the fact that .

Orthogonality: We now verify axiom is satisfied by this new structure. The first three conditions are clear, since and thus they are inherited from the hierarchically hyperbolic structure . For the last condition, any collection of pairwise orthogonal domains in is also a collection of pairwise orthogonal domains in and thus by Reference BHS19, Lemma 2.2 has uniformly bounded size, verifying the axiom.

Transversality and consistency: This axiom only involves domains which are not nest-maximal, and hence holds using the original constants from the hierarchically hyperbolic structure on .

Partial realization: This axiom only involves domains which are not nest-maximal, and hence holds using the original constants from the hierarchically hyperbolic structure on .

Finite complexity: This clearly holds by construction.

Large link axiom: Let and be the constants from the large link axiom for , let , and let . Consider the set provided by the large link axiom for . Since , it follows that is unbounded for each . Let . If , it follows that is unbounded. Furthermore, , whence for some by the large link axiom for . Therefore is unbounded, and so . The result follows.

Bounded geodesic image: For all domains in , the corresponding hyperbolic spaces are unchanged from those in the original structure and thus the axiom holds in these cases.

Hence the only case which it remains to check is when . Suppose is a geodesic in , and such that . The partial realization axiom implies that there exists a hierarchy path whose end-points project under to the end-points of . This projected path is a quasigeodesic by Lemma 3.6. Since is hyperbolic, the projected path lies uniformly close to . By Reference BHS19, Proposition 5.17 we can replace by an appropriate subpath for which the only relevant domains are all nested in ; thus . By definition, there is a bounded distance between and ; thus (and hence ) is a bounded distance from , as needed.

Uniqueness: Let . We can take , where is the original constant from the uniqueness axiom for . Then if with , then uniqueness for implies there exists with . Either or and is bounded. We are done in the first case. In the second case, by construction the factor space of obtained by collapsing is quasi-isometrically embedded in and there is a –Lipschitz map from to . Thus the lower bound on distance in provides a lower bound on the distance in , which, in turn, provides a lower bound in , as desired.

Corollary 3.8.

Every hierarchically hyperbolic group admits a hierarchically hyperbolic group structure with unbounded products.

Proof.

Recall that every hierarchically hyperbolic group has the bounded domain dichotomy. Accordingly, if we start with a hierarchically hyperbolic group, , then Theorem 3.7 yields a hierarchically hyperbolic structure with unbounded products, , where is the structure from the proof of Theorem 3.7 with the additional “dummy domains” added as provided at the end of that proof via Theorem A.1. It remains only to show that this is a hierarchically hyperbolic group structure. The action of on itself, by left multiplication, is clearly metrically proper and cobounded, and thus it only remains to show that contains finitely many –orbits. If but , then either or must be bounded. Then for each , the same will be true for or , which shows that . Thus . The result now follows from the fact that has only finitely many –orbits and that any dummy domains added fall into only finitely many orbits, as noted in Remark A.7.

4. Characterization of contracting geodesics

For this section, fix a hierarchically hyperbolic space with the bounded domain dichotomy; denote the –maximal element .

Definition 4.1 (Bounded projections).

Let and . We say that has –bounded projections if for every , we have .

Definition 4.2 (Contracting).

A subset in a metric space is said to be contracting if there exist a map and constants satisfying:

(1)

For any , we have ;

(2)

If with , then ;

(3)

For all , if we set , then .

In this section, we will focus our attention to the case of Definition 4.2 where is a quasigeodesic. In Section 6 we will consider results about arbitary subsets with the contracting property.

We note that sometimes authors refer to any quasigeodesic satisfying 3 as contracting. Nonetheless, for applications one also needs to assume the coarse idempotence and coarse Lipschitz properties given by (1) and (2), so for convenience we combine them all in one property.

A useful well-known fact is stability of contracting quasigeodesics. Two different proofs of the following occur as special cases of the results Reference MM99, Lemma 6.1 and Reference Beh06, Theorem 6.5; this explicit statement is also in Reference DT15, Section 4.

Lemma 4.3.

If is a –contracting –quasigeodesic in a metric space , then is –stable for some depending only on and .

The following result and argument both generalize and simplify the analogous result for mapping class groups in Reference Beh06.

Theorem 4.4.

Let be a hierarchically hyperbolic space. For any and there exists a depending only on and such that the following holds for every –quasigeodesic . If has –bounded projections, then is –contracting. Moreover, if has the bounded domain dichotomy, then admits a hierarchically hyperbolic structure with unbounded products where, additionally, we have that if is –contracting, then has –bounded projections.

Proof.

First suppose that has –bounded projections. It follows immediately from the definition that is a hierarchically quasiconvex subset of . Hierarchical quasiconvexity is the hypothesis necessary to apply Reference BHS17a, Lemma 5.5 (see Lemma 2.10), which then yields existence of a coarsely Lipschitz gate map , i.e., for each , the image has the property that for all the set is a uniformly bounded distance from the projection of to .

We will use as the map to prove is contracting. Gate maps satisfy condition (1) of Definition 4.2 by definition and condition (2) since they are coarsely Lipschitz. Hence it remains to prove that condition (3) of Lemma 4.3 holds.

Fix a point with and let be any point with for constants and as determined below.

Since is a gate map and has –bounded projections, for all we have . Thus, by taking a threshold for the distance formula (Theorem 2.5) larger than , we have

for uniform constants . Thus it suffices to prove that is bounded by some uniform constant . We also choose to be larger than the constants in Definition 2.1.4.

By Definition 2.1.1, the maps are Lipschitz with a uniform constant. Taking sufficiently large, it follows that there exists so that . By choosing to be sufficiently small, and applying the distance formula to the pairs and , the fact that the projections are Lipschitz implies that the sum of the terms in the distance formula associated to is much greater than the sum of those associated to . Having chosen , we have . Thus, there exists for which .

If , then having (where we enlarge if necessary) would already show that the –geodesic between and was disjoint from and then hyperbolicity of would yield a uniform bound on the .

Otherwise, we may assume . By the triangle inequality, we have . Further, since, as noted above, the projections between and are uniformly bounded, by choosing large enough and small enough, we also have .

By the bounded geodesic image axiom (Definition 2.1.7), any geodesic in either has bounded projection to or satisfies for any . For any geodesic from to (or from to ), the above argument implies that the first condition doesn’t hold for . Thus, in both cases, we know that any such geodesic must pass uniformly close to . Hence the hyperbolicity of implies is contracting, and the first implication holds.

We prove the second implication by contradiction. By Theorem 3.7, we obtain a new structure which has unbounded products. For every we have that both and are unbounded, hence every yields a non-trivial product region which is uniformly quasi-isometrically embedded in .

Suppose is contracting but doesn’t have –bounded projections. Then we obtain a sequence with . Thus there is a sequence of pairs of points , so that , with . For each , let be a –hierarchy path between . By Reference BHS19, Proposition 5.17, there exists depending only on and , such that

Since is contracting, it is uniformly stable by Lemma 4.3. Since is uniformly stable and the are uniform quasigeodesics, it follows that each is contained in a uniform neighborhood of . Hence arbitrarily long segments of are uniformly close to the product regions . This contradicts the assumption that is contracting and completes the proof.

5. Universal and largest acylindrical actions

The goal of this section is to show that for every hierarchically hyperbolic group the poset has a largest element. Recall that the action associated to such an element is necessarily a universal acylindrical action.

We prove the following stronger result which, in addition to providing new largest and universal acylindrical actions for cubulated groups, gives a single construction that recovers all previously known largest and universal acylindrical actions of finitely presented groups that are not relatively hyperbolic.

The following is Theorem A of the introduction:

Theorem 5.1.

Every hierarchically hyperbolic group admits a largest acylindrical action.

Before giving the proof, we record the following result which gives a sufficient condition for an action to be largest. This result follows directly from the proof of Reference ABO19, Theorem 4.13; we give a sketch of the argument here. Recall that an action is elliptic if has bounded orbits.

Proposition 5.2 (Reference ABO19).

Let be a group, a finite collection of subgroups of , and be a finite subset of such that generates . Assume that:

(1)

is hyperbolic and the action of on it is acylindrical.

(2)

Each is elliptic in every acylindrical action of on a hyperbolic space.

Then is the largest element in .

Proof.

First notice that by assumption (1), is an element of . Let be a cobounded acylindrical action of on a hyperbolic space, , and fix a basepoint . Then there exists a bounded subspace such that . By assumption (2), the orbit is bounded for all , …, . Since , we know and thus

is finite. Let , and let

The standard Milnor-Svarc Lemma argument shows that is an infinite generating set of and there exists a –equivariant quasi-isometry . It is clear that contains , as well as for all , …,  and thus . The result follows.

Proof of Theorem 5.1.

Let be a hierarchically hyperbolic group with finite generating set . By Corollary 3.8, there is a hierarchically hyperbolic group structure with unbounded products. Recall that is the –maximal element of with associated hyperbolic space . The action on is acylindrical by Reference BHS17b, Theorem K.

Moreover, the action of on is cobounded, so let be a fundamental domain for and

Notice that will contain at most one representative from each –orbit of domains, and so must be a finite set. Indeed, for a hierarchically hyperbolic group, this follows from the fact that the action of on is cofinite.

Let be the stabilizer of for each . By a standard Milnor-Svarc argument (see Reference ABO19 for details) there is a –equivariant quasi-isometry between and , where . Therefore condition (1) of Proposition 5.2 is satisfied.

By definition, each sits inside a non-trivial direct product in , the product region associated to each . It follows that must be elliptic in every acylindrical action of on a hyperbolic space (see Remark 2.18), satisfying condition (2).

Therefore, by Proposition 5.2, the action is largest.

Remark 5.3.

The proof of Theorem 5.1 can be extended to treat a number of groups which are hierarchically hyperbolic spaces, but not hierarchically hyperbolic groups. For example, it was shown in Reference BHS19, Theorem 10.1 that every fundamental group of a compact –manifold with no Nil or Sol in its prime decomposition admits a hierarchically hyperbolic space structure, which is constructed by first putting an HHS structure on each geometric piece in the prime decomposition. However, as explained in Reference BHS19, Remark 10.2 it is likely that such fundamental groups don’t all admit hierarchically hyperbolic group structures. Nonetheless, the proof of the above theorem works in this case by replacing the use of the fact that the action of on is cofinite, with the fact that for , the set is precisely the set of –maximal domains in the hierarchically hyperbolic structure on each of the Seifert-fibered components of the prime decomposition of , and so is finite.

Remark 5.4.

There is an instructive direct proof of the universality of the above action, using the characterization of contracting quasigeodesics in Section 4, which we now give. We call an infinite order element contracting if its orbit is a contracting quasigeodesic in the Cayley graph. Now, let be an infinite order element and consider the geodesic in .

If is contracting in , then by Theorem 4.4 all proper projections are bounded, and thus by the distance formula, is loxodromic for the action on .

If is not contracting in , then there exists some such that is unbounded. Thus for any increasing sequence of constants with , there are sequences of pairs of points such that as and . For each , let be an –hierarchy path between and . By definition, is a uniform quasigeodesic. Then by Reference BHS19, Proposition 5.17, there exists depending only on and such that . If is a generalized loxodromic, then is stable, by Reference Sis16, and so the subgeodesic stays within a uniform bounded distance of . Thus arbitrarily long subgeodesics of stay within a uniformly bounded distance of a product region, . This contradicts being Morse, and therefore is not a generalized loxodromic element.

This remark directly implies that the action on is a universal acylindrical action. (The universality of the action can also be proven using the classification of elements of described in Reference DHS17.)

Another immediate consequence of the above remark is the following, which for hierarchically hyperbolic groups strengthens a result obtained by combining Reference Osi16, Theorem 1.4.(L4) and Reference Sis16, Theorem 1, which together prove that a generalized loxodromic element in an acylindrically hyperbolic group is quasi-geodesically stable.

Corollary 5.5.

Let be a hierarchically hyperbolic group. An element is generalized loxodromic if and only if is contracting.

The next result provides information about the partial ordering of acylindrical actions. Of the groups listed below, the largest and universal acylindrical action of the class of special CAT(0) cubical groups is new; the other cases were recently established to be largest in Reference ABO19.

Corollary 5.6.

The following groups admit acylindrical actions that are largest (and therefore universal):

(1)

Hyperbolic groups.

(2)

Mapping class groups of orientable surfaces of finite type.

(3)

Fundamental groups of compact three-manifolds with no Nil or Sol in their prime decomposition.

(4)

Groups that act properly and cocompactly on a special CAT(0) cube complex, and more generally any cubical group which admits a factor system. This includes right-angled Artin groups, right-angled Coxeter groups, and many other examples as in Reference HS16.

Proof.

With the exception of (3), by Reference BHS17bReference BHS19Reference HS16 the above are all hierarchically hyperbolic groups and therefore have the bounded domain dichotomy. In case (3), where is the fundamental group of a compact three-manifold with no Nil or Sol in its prime decomposition, then while is not always known to be a hierarchically hyperbolic group, it has a hierarchically hyperbolic structure . To see this, we use the fact that there is a group which is quasi-isometric to and has a hierarchically hyperbolic structure with all of the associated hyperbolic spaces infinite Reference BHS19, Theorem 10.1 & Remark 10.2; thus by quasi-isometric invariance of hierarchically hyperbolic structures Reference BHS19, Proposition 1.10, does as well. Since all of the associated hyperbolic spaces are infinite, has the bounded domain dichotomy, so the result follows.

We give an explicit description of these actions for each hierarchically hyperbolic group in the corollary, in the sense that we describe the set of domains which are removed from the standard hierarchical structure of the group and whose associated hyperbolic space is infinite diameter. Recall that the space is constructed from by coning off all elements of which consists of those components of whose associated product regions have both factors with infinite diameter. Coning off all of yields a space which is is quasi-isometric to the space obtained by just coning off .

(1)

Hyperbolic groups have a canonical simplest hierarchically hyperbolic group structure given by taking , where is the Cayley graph of the group with respect to a finite generating set. For this structure, , and the action on the Cayley graph is clearly largest.

(2)

For mapping class groups, the natural hierarchically hyperbolic group structure is the set of homotopy classes of non-trivial non-peripheral (possibly disconnected) subsurfaces of the surface; the maximal element is the surface itself, and the hyperbolic space is the curve complex of . For this structure . (Note that to form one must remove the nest-maximal collections of disjoint subsurfaces; the hyperbolic space associated to each of these, except , has finite diameter). Additionally, we emphasize that although the new hyperbolic space is not , it is quasi-isometric to , the action on which is known to be universal. Universality of this action was shown by Osin in Reference Osi16, and follows from results of Masur-Minsky and Bowditch Reference Bow08Reference MM99.

(3)

If is a compact –manifold with no Nil or Sol in its prime decomposition and , then is exactly the set of vertex groups in the prime decomposition that are fundamental groups of hyperbolic 3–manifolds (each of which has exactly one domain in its hierarchically hyperbolic structure).

(4)

If is a group that acts properly and cocompactly on a special CAT(0) cube complex , then by Reference BHS17b, Proposition B, has a –equivariant factor system. This factor system gives a hierarchically hyperbolic group structure in which is the closure under projection of the set of hyperplanes along with a maximal element , where is the contact graph as defined in Reference Hag14. In this structure, is the set of indices whose stabilizer in contains a power of a rank one element.

In the particular case of right-angled Artin groups, no power of a rank one element will stabilize a hyperplane, so . In this case, the contact graph is quasi-isometric to the extension graph defined by Reference KK14. That the action on the extension graph is a universal acylindrical action follows from the work of Reference KK14 and the centralizer theorem for right-angled Artin groups. This action is also shown to be largest in Reference ABO19.

We give a concrete example of the situation in the case of a right-angled Coxeter group.

Example 5.7.

Let be the right-angled Coxeter group whose defining graph is a pentagon. Then , and the Cayley graph of is the tiling of the hyperbolic plane by pentagons. We consider the dual square complex to this tiling. To form the contact graph , we start with the square complex and cone off each hyperplane carrier, which is equivalent to coning off the hyperplane stabilizers in the Cayley graph. The result is a quasi-tree. Thus a fundamental domain for the hierarchically hyperbolic group structure of is where is associated to the stabilizer of the hyperplane labeled by and is associated to the contact graph described above.

Consider the hyperplane that is labeled by . Then the stabilizer of is the subgroup generated by the star of the vertex , which is . This subgroup contains the infinite order element . As is a hyperbolic group, all infinite order elements are generalized loxodromic, but is not loxodromic for the action on the contact graph since its axis lies in a hyperplane stabilizer that has been coned-off. Thus the action on the contact graph is not universal.

Let be the element associated to . Then is a product region, and the maximal orthogonal component is bounded. Thus , as is for each vertex of the defining graph. The contact graph associated to is a line, and the element is loxodromic for the action on this space.

Note that once has been removed from , the resulting hierarchically hyperbolic structure is , the canonical hierarchically hyperbolic structure for a hyperbolic group, in which .

6. Characterizing stability

In this section, we will give several characterizations of stability which hold in any hierarchically hyperbolic group. In fact, we will characterize stable embeddings of geodesic metric spaces into hierarchically hyperbolic spaces with unbounded products. One consequence of this will be a description of points in the Morse boundary of a proper geodesic hierarchically hyperbolic space with unbounded products as the subset of the hierarchically hyperbolic boundary consisting of points with bounded projections.

6.1. Stability

While it is well-known that contracting implies stability Reference Beh06Reference DMS10Reference MM99, the converse is not true in general. Nonetheless, in several important classes of spaces the converse holds, including in hyperbolic spaces, CAT(0) spaces, the mapping class group, and Teichmüller space Reference Sul14Reference Beh06Reference DT15Reference Min96. We record the following corollary of Theorem 4.4 which gives a relationship between stability and contracting subsets that holds in a fairly general context.

Corollary 6.1.

Suppose that has unbounded products, is a hyperbolic metric space, and is a –quasi-isometric embedding. Then is –stable if and only if is –contracting, where and determine each other.

Proof.

First assume that is –contracting. Since is a –quasi-isometric embedding, to show that is –stable for some gauge , we need only show that the (quasigeodesic) image of every geodesic in is –stable. Since is –contracting and is hyperbolic, is –contracting for some depending only on , , , and the hyperbolicity constant of . Lemma 4.3 shows that is therefore –stable, with depending only on , as desired. (Note that the assumption on unbounded products is not necessary for this implication.)

For the other direction, the fact that has unbounded products implies that has bounded projections, since otherwise one could find large segments of quasigeodesics contained inside product regions with unbounded factors, contradicting stability. The result now follows from Theorem 4.4.

The following provides a general characterization of stability in HHSs, a special case of which is Theorem B.

Corollary 6.2.

Let be a quasi-isometric embedding from a metric space into a hierarchically hyperbolic space with unbounded products. The following are equivalent:

(1)

is a stable embedding;

(2)

has uniformly bounded projections;

(3)

is a quasi-isometric embedding.

Proof.

That item (2) implies (3) follows from the distance formula and the assumption that is a quasi-isometric embedding.

The hypothesis of item (1) implies that is hyperbolic. Moreover, since (2) implies (3), the hypothesis of (2) also implies that is hyperbolic. Thus items (1) and (2) are equivalent via Corollary 6.1 and Theorem 4.4.

We now prove that (3) implies (2). Suppose for a contradiction that for any integer there exists and satisfying . Now, we consider a hierarchy path between and . Applying the bounded geodesic image axiom (Definition 2.1.(7)) to the associated –geodesic between and it follows that this –geodesic has non-trivial intersection with the ball of radius about the set . Indeed, this yields that there exist points on the geodesic which are both distance at most from ; by Reference BHS19, Lemma 5.17 we can assume that and were chosen so that and also satisfy and . Thus, we have that . The hypothesis in (3) implies that there is a uniform bound on . The distance formula then implies a uniform bound on for any , contradicting the fact that we chose to be large.

6.2. The Morse boundary

In the rest of this section, we turn to studying the Morse boundary and use this to give a bound on the stable asymptotic dimension of a hierarchically hyperbolic space. We begin by describing two notions of boundary.

In Reference DHS17, Durham, Hagen, and Sisto introduced a boundary for any hierarchically hyperbolic space. We collect the relevant properties we need in the following theorem:

Theorem 6.3 (Theorem 3.4 and Proposition 5.8 in Reference DHS17).

If is a proper hierarchically hyperbolic space, then there exists a topological space such that compactifies , and the action of on extends continuously to an action on .

Moreover, if is a hierarchically quasiconvex subspace of , then, with respect to the induced hierarchically hyperbolic structure on , the limit set of of in is homeomorphic to and the inclusion map extends continuously to an embedding .

Building on ideas in Reference CS15, Cordes introduced the Morse boundary of a proper geodesic metric space Reference Cor17, which was then refined further by Cordes–Hume in Reference CH17. The Morse boundary is a stratified boundary which encodes the asymptotic classes of Morse geodesic rays based at a common point. Importantly, it is a quasi-isometry invariant and generalizes the Gromov boundary of a hyperbolic space Reference Cor17.

We briefly discuss the construction of the Morse boundary and refer the reader to Reference Cor17Reference CH17 for details.

Consider a a proper geodesic metric space with a basepoint . Given a stability gauge , define a subset to be the collection of points such that and can be connected by an –stable geodesic in . Each such is –hyperbolic for some depending on and Reference CH17, Proposition 3.2; here, we use the Gromov product definition of hyperbolicity, as need not be connected. Moreover, any stable subset of embeds in for some Reference CH17, Theorem A.V.

The set of stability gauges admits a partial order: if and only if for all constants . In particular, if , then .

Since each is Gromov hyperbolic, each admits a Gromov boundary . Take the direct limit with respect to this partial order to obtain a topological space called the Morse boundary of .

We fix , a hierarchically hyperbolic structure with unbounded products.

Definition 6.4.

We say has bounded projections if for any , there exists such that any –hierarchy path has –bounded projections. Let denote the set of points with bounded projections.

The boundary contains for each , by construction. The next lemma shows that the boundary points with bounded projections are contained in , as a subset of , where is the –maximal element. In general, the set of boundary points with bounded projections may be a very small subset of . For instance, in the boundary of the Teichmüller metric, these points are a proper subset of the uniquely ergodic ending laminations and have measure zero with respect to any hitting measure of a random walk on the mapping class group.

Lemma 6.5.

The inclusion holds for any with unbounded products where is the –maximal element of . Moreover, if is also proper, then for any there exists depending only on and such that if is a sequence with such that has –bounded projections for some and each , then has –bounded projections.

Proof.

Let . If is an –hierarchy path, then has an infinite diameter projection to some , see, e.g., Reference DHS17, Lemma 3.3. As has bounded projections, we must have . Since is a quasigeodesic ray, the first statement follows.

Now suppose that is also proper. For each , let be any –hierarchy path between and in . The Arzela-Ascoli theorem implies that after passing to a subsequence, converges uniformly on compact sets to some –hierarchy path with depending only on and . Hence has –bounded projections for some depending only on and . Moreover, since in , it follows that is asymptotic to in .

If is any other –hierarchy path, it follows from uniform hyperbolicity of the and the definition of hierarchy paths that is uniformly bounded for all . Since has –bounded projections, the distance formula implies that has –bounded projections for some depending only on and , as required.

6.3. Bounds on stable asymptotic dimension

The asymptotic dimension of a metric space is a coarse notion of topological dimension which is invariant under quasi-isometry. Introduced by Cordes–Hume Reference CH17, the stable asymptotic dimension of a metric space is the maximal asymptotic dimension of a stable subspace of .

The stable asymptotic dimension of a metric space is always bounded above by its asymptotic dimension. Behrstock, Hagen, and Sisto Reference BHS17a proved that all proper hierarchically hyperbolic spaces have finite asymptotic dimension (and thus have finite stable asymptotic dimension, as well). The bounds on asymptotic dimension obtained in Reference BHS17a are functions of the asymptotic dimension of the top level curve graph.

In the following theorem, we prove that a hierarchically hyperbolic space has finite stable asymptotic dimension under the assumption that , where is the hyperbolic space associated to the –maximal domain in .

Recall that asymptotic dimension is monotonic under taking subsets. Thus, if is assumed to be proper, so that , then (and therefore its stable subsets) have finite asymptotic dimension by Reference BHS17a. Here, using some geometry of stable subsets we obtain a sharper bound on than .

Theorem 6.6.

Let be a hierarchically hyperbolic space with unbounded products such that has finite asymptotic dimension, where is the –maximal element of . Then . Moreover, if is also proper and geodesic, then there exists a continuous bijection .

Proof.

By Reference CH17, Lemma 3.6, for any stability gauge there exists such that is –stable. Hence, there exists depending only on and such that has –bounded projections. By Corollary 6.2, it follows that the projection is a quasi-isometric embedding with constants depending only on and . Since every stable subset of embeds into some Reference CH17, Theorem A.V, the first conclusion then follows from the definition of stable asymptotic dimension.

Now suppose that is proper.

Since each is stable in , these sets have bounded projections by Corollary 6.2; from this it follows that is hierarchically quasiconvex for each . Hence by Reference DHS17, Proposition 5.8, the canonical embedding extends to an embedding .

By Corollary 6.2 and Lemma 6.5, we have . Let be the direct limit of the . Since it is injective on each stratum, is injective.

To prove surjectivity, let . Let and fix a hierarchy path . Since , has –bounded projections for some . Let be such that in . If is a sequence of geodesics between and , then, by properness, the Arzela–Ascoli theorem, and passing to a subsequence if necessary, there exists a geodesic ray with such that converges on compact sets to . Since each has –bounded projections, it follows that has –bounded projections for some depending only on and . Moreover, by hyperbolicity of and the construction of we have that is uniformly bounded and thus, by the distance formula, so is . Since by construction, it follows that , as required.

Continuity of for each follows from Reference DHS17, Proposition 5.8, as above. This and the definition of the direct limit topology implies continuity of .

The following corollary is immediate:

Corollary 6.7.

If is a hierarchically hyperbolic group, then has finite stable asymptotic dimension.

6.4. Random subgroups

Let be any countable group and a probability measure on whose support generates a non-elementary semigroup. A –generated random subgroup of , denoted is defined to be the subgroup generated by the step of independent random walks on , where . For other recent results on the geometry of random subgroups of acylindrically hyperbolic groups, see Reference MS19.

Following Taylor-Tiozzo Reference TT16, we say a –generated random subgroup of has a property if

Theorem 6.8.

Let be an HHS for which the –maximal element, , has infinite diameter, and consider which acts properly and cocompactly on  via the orbit map. Then any –generated random subgroup of stably embeds in .

Proof.

By Reference BHS17b, Theorem K, acts acylindrically on . Let be generated by independent random walks as above. Now, Reference TT16, Theorem 1.2 implies that a.a.s. quasi-isometrically embeds into , and hence is hyperbolic. Moreover, the distance formula implies that is undistorted in G and any orbit of in has bounded projections by the distance formula. By Theorem 4.4, having bounded projections implies contracting; thus any orbit of in is a.a.s. contracting, which gives the conclusion by Corollary 6.1. (Note that the directions of Theorem 4.4 and Corollary 6.1 used here do not require that has unbounded products.)

In particular, one consequence is a new proof of the following result of Maher–Sisto. This result follows from the above, together with Rank Rigidity for HHG (Reference DHS17, Theorem 9.14) which implies that a hierarchically hyperbolic group which is not a direct product of two infinite groups has infinite diameter.

Corollary 6.9 (Maher–Sisto; Reference MS19).

If is a hierarchically hyperbolic group which is not the direct product of two infinite groups, then any –generated random subgroup of is stable.

7. Clean containers

The clean container property is a condition related to the orthogonality axiom. In Proposition 7.2 this property is shown to hold for many, but not all, hierarchically hyperbolic groups. Unlike earlier versions of this paper, this condition is no longer needed to prove the main theorems of the earlier sections. However, we keep the content in this paper since this property has found independent interest and is used elsewhere.

Definition 7.1 (Clean containers).

In a hierarchically hyperbolic space for each and each with the orthogonality axiom provides a container. If, for each , such a container can be chosen to be orthogonal to , then we say that has clean containers.

We first describe some interesting examples with clean containers. Then we show that this property is preserved under some combination theorems for hierarchically hyperbolic spaces. We refer the reader to Reference BHS19, Sections 8 & 9 and Reference BHS17a, Section 6 for details on the structure in the new spaces.

Proposition 7.2.

The following spaces admit hierarchically hyperbolic structures with clean containers.

(1)

Hyperbolic groups.

(2)

Mapping class groups of orientable surfaces of finite type.

(3)

Special cubical groups, and more generally, any cubical group which admits a factor system.

(4)

, for a compact –manifold with no Nil or Sol in its prime decomposition.

Proof.

Hierarchically hyperbolic structures for these spaces were constructed in Reference BHS17b and Reference BHS19.

(1)

The statement is immediate for hyperbolic groups, as they each admit a hierarchically hyperbolic structure with no orthogonality, and thus the container axiom is vacuous.

(2)

For mapping class groups, in the standard structure, a container for domains orthogonal to a given subsurface is the complementary subsurface, which is orthogonal to .

(3)

The statement follows immediately from Reference BHS17b, Proposition B and Reference HS16, Corollary 3.4.

(4)

Given a geometric 3–manifold of the above form, is quasi-isometric to a (possibly degenerate) product of hyperbolic spaces, and so has clean containers by Proposition 7.3. Given an irreducible non-geometric graph manifold , the hierarchically hyperbolic structure comes from considering as a tree of hierarchically hyperbolic spaces with clean containers and hence has clean containers by Proposition 7.5. Finally, the general case of a non-geometric –manifold follows immediately from Proposition 7.4 and the fact that is hyperbolic relative to its maximal graph manifold subgroups.

Proposition 7.3.

The product of two hierarchically hyperbolic spaces which both have clean containers has clean containers.

Proof.

Let and be hierarchically hyperbolic spaces with clean containers. In the hierarchically hyperbolic structure given by Reference BHS19, Theorem 8.27 there are two types of containers, those that come from one of the original structures and those that do not. Containers of the first type are clean, as both original structures have clean containers.

The second type of container consists of new domains obtained as follows. Given a domain , a new domain is defined with the property that it contains under nesting any domain in which is orthogonal to and also any domain in . Thus, by construction is a container for everything orthogonal to . As , the result follows.

Proposition 7.4.

If is hyperbolic relative to a collection of hierarchically hyperbolic spaces which all have clean containers, then is a hierarchically hyperbolic space with clean containers.

Proof.

That is a hierarchically hyperbolic space follows from Reference BHS19, Theorem 9.1. In the hierarchically hyperbolic structure on , no new orthogonality relations are introduced, and thus all containers are containers in the hierarchically hyperbolic structure of one of the peripheral subgroups. As each of these structures have clean containers, it follows that does, as well.

The following example relies on the combination theorem Reference BHS19, Theorem 8.6. We provide this as another example of hierarchically hyperbolic spaces with clean containers, but since we don’t rely on this elsewhere in the paper, we refer to that reference for the relevant definitions. Nonetheless, we include a full proof for the expert, since it is short. (We note that after this paper was circulated, Berlai and Robbio proved a combination theorem under weaker conditions than Reference BHS19, Theorem 8.6 and, in the process, also proved that if all the vertex spaces have clean containers, then so does the combined space, see Reference BR, Theorem A.)

Proposition 7.5.

Let be a tree of hierarchically hyperbolic spaces satisfying the hypotheses of Reference BHS19, Theorem 8.6, so that is hierarchically hyperbolic. If for each , the hierarchically hyperbolic space has clean containers, then so does .

Proof.

This follows immediately from the proof of Reference BHS19, Theorem 8.6 and the fact that edge-hieromorphisms are full and preserve orthogonality. In the notation from that result, we note that, if has clean containers for each , then the domain described in the proof also has the property that . Therefore, as edge-hieromorphisms are full and preserve orthogonality, .

The following uses the notion of hierarchically hyperbolically embedded subgroups introduced in Reference BHS17a; see also Reference DGO17 for the related notion of hyperbolically embedded subgroups.

Proposition 7.6.

Let be a hierarchically hyperbolic group with clean containers, and let be a hierarchically hyperbolically embedded subgroup of . Then there exists a finite set such that for all with and is hyperbolic, the group , obtained by quotienting by the normal closure, is a hierarchically hyperbolic group with clean containers.

Proof.

Recall that in the hierarchically hyperbolic structure obtained in Reference BHS17a, Theorem 6.2 (and in the notation used there), two domains satisfy (respectively ) if there exists a linked pair with and such that (respectively ). Let and with . To prove the container axiom, we consider domains such that , and for all , and such that any pair is a linked pair. Then the orthogonality axiom for provides a domain such that and . As has clean containers, we also have that . This implies that and are coarsely equal by Reference DHS17, Lemma 1.5, and so is a linked pair. Therefore, .

Appendix A. Almost HHSs are HHSs By Daniel Berlyne and Jacob Russell

The main result in this appendix is that every almost HHS structure can be promoted to an HHS structure. Recall that, as introduced in Section 3.2, an almost HHS is a space which satisfies all the axioms of an HHS except for the orthogonality axiom, which is instead replaced by a weaker axiom without a container requirement. In Theorem A.1, we show that an almost HHS structure can be made into an actual HHS structure by adding appropriately chosen “dummy domains” to serve as the containers. This result provides a useful method for producing an HHS structure while only needing to verify the weaker axioms of an almost HHS. This method is used in the main text in the proof of Theorem 3.7, where it is shown that every hierarchically hyperbolic space with the bounded domain dichotomy admits an HHS structure with unbounded products.

Theorem A.1.

Let be an almost HHS. There exists an HHS structure for so that , and if then the associated hyperbolic space for is a single point.

To prove Theorem A.1, we will need to collect three additional tools about almost HHSs. Each of these tools was proved in the setting of hierarchically hyperbolic spaces, but they continue to hold in the almost HHS setting. Indeed, the only use of the containers in their proofs is Reference BHS19, Lemma 2.1, which proves that the cardinality of any collection of pairwise orthogonal domains is uniformly bounded by the complexity of the HHS.

The first tool says the relative projections of orthogonal domains coarsely coincide. Note, and are both defined when or and or .

Lemma A.2 (Reference DHS17, Lemma 1.5).

Let be an almost HHS. If with , and with and both defined, then where is the constant from the consistency axiom of .

The second tool we will need is the realization theorem for almost HHSs. The realization theorem characterizes which tuples in the product are coarsely the image of a point in . Essentially, it says if a tuple satisfies the consistency inequalities of an almost HHS (see Definition 2.6), then there exists a point such that is uniformly close to for each .

Theorem A.3 (The realization of consistent tuples, Reference BHS19, Theorem 3.1).

Let be an almost HHS. There exists a function so that if is a –consistent tuple, then there exists so that for all .

The last result we need is that the relative projections of an almost HHS also satisfy the inequalities in the consistency axiom.

Lemma A.4 (–consistency, Reference BHS19, Proposition 1.8).

Let be an almost HHS structure for and . Suppose or and or . Then we have the following, where is the constant from the consistency axiom of .

(1)

If , then .

(2)

If , then .

We are now ready to prove that every almost HHS is an HHS (Theorem A.1). If is an almost HHS, then the only HHS axiom that is not satisfied is the container requirement of the orthogonality axiom. The most obvious way to address this is to add an extra element to every time we need a container. That is, if with and there exists some with , then we add a domain to serve as the container for in , i.e., every nested into and orthogonal to will be nested into . However, this approach is perilous as once a domain is nested into , we may now need a container for in ! To avoid this, we add domains where is a pairwise orthogonal set of domains nested into ; that is, contains all domains that are nested into and orthogonal to all . This allows for all the needed containers to be added at once, avoiding an iterative process.

Proof of Theorem A.1.

Let be an almost HHS and let be the maximum of all the constants in . Let denote a non-empty set of pairwise orthogonal elements of and let . We say the pair is a container pair if the following are satisfied:

for all , ;

there exists such that for all .

Let denote the set of all container pairs. We will denote a pair by .

Let . We will prove has a hierarchically hyperbolic space structure with index set . Since is an almost HHS, we can continue to use the spaces, projections, and relations for elements of . Thus we only define new projections, relative projections, and relations when elements of are involved. If , then the associated hyperbolic space, , will be a single point.

Projections: For , the projection map is just the constant map to the single point in .

Nesting: Let and .

Define if in and for all .

Define if in .

Define if in and for all either or there exists with .

These definitions ensure is still a partial order and maintain the –maximal element of as the –maximal element of .

Since the hyperbolic spaces associated to elements of are points, define for every and with . The downwards relative projection can be defined arbitrarily.

If and with , then in for each . Thus we define . Lemma A.2 ensures that has diameter at most . In this case, we define as the constant map to the single point in .

Finite complexity: First consider a nesting chain of the form .

Claim A.5.

The length of is bounded above by .

Proof.

For each , we have and hence . As for each , every element of must therefore be nested into an element of . Denote the elements of by , …, . Since each is a pairwise orthogonal subset of , we have for each by the bounded pairwise orthogonality axiom of an almost HHS (Definition 3.4). We define a –nesting chain to be a maximal chain of the form for some and , with . Since the elements of are pairwise orthogonal for each , if is the –minimal element of a –nesting chain, then is nested into exactly one element of for each . This implies that each –nesting chain is determined by its –minimal element. Further, the set of –minimal elements of –nesting chains is pairwise orthogonal. By the bounded pairwise orthogonality axiom of an almost HHS, this implies there exist at most –nesting chains.

In order for , either or there exists , such that . Thus, every step up the chain results in either a strict decrease in (the cardinality of ) to (the cardinality of ) or a strict step down one of the –nesting chains. Note that may increase when we encounter a strict step down one of the –nesting chains, since multiple elements of may be nested into the same element of . Such an increase in corresponds to the nesting chain branching into multiple chains, which may only happen at most times, as there are at most –nesting chains. Hence, the length of is bounded by plus the total number of times a strict decrease can occur across all of the –nesting chains.

Each –nesting chain contains at most distinct elements of by the finite complexity of . Bounded pairwise orthogonality implies there are at most different –nesting chains, thus the number of steps of the chain where there is a strict decrease within one of the –nesting chains is at most . This bounds the length of by .

We now consider a nesting chain of the form . In this case, , but not all of these nestings must be proper. Let be the minimal subset of such that if , then . Thus , and by finite complexity of . Claim A.5 established that , so , that is, any –chain of elements of has length at most .

Finally, since any –chain of elements of can be partitioned into a –chain of elements of and a –chain of elements of , any –chain in has length at most .

Orthogonality: Two elements are orthogonal if in . Let and . Define if, in , either or for some . These definitions, plus the definition of nesting, imply for all , if and , then . We now verify that satisfies the container requirements of the orthogonality axiom.

Let with and , i.e., is a container pair. In this case, the container of in for is . We now show containers exist for situations involving elements of . We split this into three subcases.

Case 1 ( and with ).

Since is a container pair, there exists with and for all . Suppose that requires a container in , that is, there is an element of that is orthogonal to and nested in . We verify that is a container pair and is a container of in .

If with and , then or for some . In either case, we have , so is a container pair and . If with and , then and . Since , this implies . Therefore, is again a container pair, and .

Case 2 ( where ).

Since is a container pair, there exists so that and for all . Since , it follows that for all , either or there exists so that . In both cases, . Thus is a pairwise orthogonal collection of elements of . Suppose that requires a container in , that is, there is an element of that is orthogonal to and nested in . We verify that is a container pair and is a container for in .

If satisfies and , then either or for some . In both cases, . Further, we must have and for each as . Thus is a container pair and . On the other hand, if satisfies and , then , , and for each either or there exists with . Since is a container pair, there exists such that and for all . Since , we also have as and . For each , either or there exists with . In both cases, . Thus, is orthogonal to all elements of and moreover , so is a container pair. Furthermore, since and . We have therefore shown that is a container for in .

Case 3 ( and with ).

This implies is a pairwise orthogonal set of elements of . Further, suppose that requires a container in , that is, there is an element of that is orthogonal to and nested in . We verify that is a container pair and is a container for in .

Suppose there exists with and . Then and is orthogonal to all the elements of . Thus is a container pair, so exists and . Now suppose there exists such that . Since is a container pair, there exists with and orthogonal to each element of . As , for each either or there exists such that . In both cases, . Further, as , we have or for some . In both cases, . Therefore is orthogonal to every element of , and moreover since . Thus is a container pair and . Now, for each , either or for some . Since and , this implies . Thus, is a container pair and is a container for in .

Transversality: An element of is transverse to an element of whenever it is not nested or orthogonal. Since the hyperbolic spaces associated to elements of are points, we only need to define the relative projections from an element of to an element of . Let and and suppose . This implies and . We define based on the –relation between and the elements of .

If for all , then as would imply . Thus we must have , so we define .

If or for some , then exists and we define to be the union of all the for with or . Lemma A.2 ensures has diameter at most in this case.

If for some , then which contradicts , so this case does not occur.

Consistency: Since the only elements of whose associated spaces are not points are in , the first two inequalities in the consistency axiom for imply the same two inequalities for . To verify the final clause of the consistency axiom, we need to check that if such that with and both defined, then is uniformly bounded in terms of . We can assume as has diameter zero otherwise. We can further assume at least one of and is an element of , as we already have the consistency axiom for elements of .

Case 1 ( ).

Assume and . Fix . Since and , we have . Since , Lemma A.2 says .

Assume and . Fix . In this case, since . Since , we have . Thus, the consistency axiom for says .

Assume and . Thus and consistency in implies . Fix and . Consistency in also implies and . Since and , we have

Case 2 (, , and ).

In this case we have either or .

Assume and . Since is transverse to we cannot have for any (this would imply ). If for all , then (as shown in the proof of transversality) and . Since , we have and consistency in implies . If instead there exists so that or , then . Since , and Lemma A.2 gives .

Assume and . As before, for all . First assume there exists so that or . This occurs when either or and not every element of is orthogonal to . In both cases, and consistency in implies because . Now assume for all . This can only occur when is transverse to . In this case, and . Since , consistency in implies .

Assume and . As before, for all . If , then we have the first case of transversality, that is, and for all . Thus, if , then the result reduces to the previous bullet, replacing with . We can therefore assume , meaning we have the second case of transversality where there exists so that is either transverse to or properly nested into .

Suppose too. This implies there also exists so that is either transverse to or properly nested into . Furthermore, and . Now, implies or is nested into an element of . If , then and Lemma A.2 implies . If is nested into an element of , then either or since is a pairwise orthogonal subset of . By applying consistency in when or Lemma A.2 when , we have .

Now suppose . Then implies or is nested into . Applying Lemma A.2 if , or consistency in if , we again obtain .

Uniqueness, bounded geodesic image, large links: Since the only elements of whose associated spaces are not points are in , these axioms for follow from the fact that they hold in .

Partial realization: Let , …, be pairwise orthogonal elements of , and let for each . Without loss of generality, assume , …, and , …, where . If (resp. ), then each (resp. ).

For , let and let be any point in . Since , …, are pairwise orthogonal, it follows that , …, are pairwise orthogonal too, and for each , is either nested into an element of some or orthogonal to all , …, . Without loss of generality, assume that , …, are nested into elements of and , …, , , …, are pairwise orthogonal, where , , and . If , then and each is orthogonal to every . Otherwise, for each , is nested in some for . In both cases, , …, , , …, are pairwise orthogonal elements of . We can therefore use the partial realization axiom in on the points , …, , , …, to produce a point with the following properties:

(1)

for ;

(2)

for ;

(3)

for all if or , then ;

(4)

for all if or , then .

Now, for , define as follows. Let and . If , then define to be any point in . Since is a collection of pairwise orthogonal elements of , the diameter of is at most by Lemma A.2. If either for some or for all then define . Since is a collection of pairwise orthogonal elements of , these two cases encompass all elements of .

Claim A.6.

The tuple is –consistent.

Proof.

Let . Recall that if and , then the –consistency inequalities for and are satisfied by the consistency axiom of . Thus we can assume that there exists so that either or . Fix so that . We need to verify the consistency inequalities when , , and .

Consistency when : Assume . If , , or then either Lemma A.2 or consistency in implies . Since , we have . Now suppose so that is non-empty. In this case, and so is within of . Now, if , then . Thus –consistency (Lemma A.4) implies . It follows that by the triangle inequality.

Consistency when : Assume . As before, if , , or then and we have . Thus, we can assume so that is within of . Now, if , then , and –consistency implies . However, this implies since and .

Consistency when : Assume . If is orthogonal to all elements of , then implies which contradicts the assumption that or . On the other hand, if there exists so that , then either (if ) or (if ). But this implies either or , both of which give a contradiction if or . There must therefore be an element of that is either properly nested in or transverse to , and we can repeat the same argument as in the previous case, switching the roles of and .

Let be the point produced by applying the realization theorem (Theorem A.3) in to the tuple . We claim is a partial realization point for , …, in . Since is a single point, satisfies the first requirement of the partial realization axiom in for , …, . For , is either nested into an element of or orthogonal to all , …, . This implies is either nested into an element of or orthogonal to all elements of . In both cases , and we have that is uniformly close to , which is in turn –close to .

Now, let with or for some . We verify is uniformly bounded when and separately.

Assume , so that . If and , then is bounded by item 3. If and , then for some and is either orthogonal to or nested into . If then by Lemma A.2. If then by consistency. The result then follows from the triangle inequality since is uniformly close to .

Now assume , so that . If , then for all . Since is within of any for , this bounds uniformly. On the other hand, if , then either for all or there exists so that or . In the latter case, and we are finished since is within of , giving a uniform bound on the distance from to . In the former case, we must have and is equal to . If then we are done by item 4. Otherwise, there exists so that or and . Since , it follows that is within of . Thus , and hence , is uniformly close to . This concludes the proof of Theorem A.1.

Remark A.7.

We say is an almost HHG if there exists an almost HHS such that and satisfy the definition of a hierarchically hyperbolic group where ‘HHS’ is replaced with ‘almost HHS’. The above proof shows that if is an almost HHG, then the structure from Theorem A.1 is an HHG structure for .

The following corollary gives criteria for the HHS structure from Theorem A.1 to have unbounded products. This is the version of Theorem A.1 that is applied in Theorem 3.7 to prove that every hierarchically hyperbolic space with the bounded domain dichotomy admits an HHS structure with unbounded products.

Corollary A.8.

Let be an almost HHS with the bounded domain dichotomy. If for every non––maximal domain , there exist so that , , and , then the HHS structure obtained by applying Theorem A.1 to has unbounded products.

Proof.

Assume for every non––maximal domain , there exist so that , and . Let be the HHS structure obtained from using Theorem A.1. If and is not –maximal, then the above property implies that and are both infinite diameter. Thus, we need only verify unbounded products for elements of . Using the notation of Theorem A.1, let and assume . Now, for all , and by construction of , there exists so that and . Since , this implies . Therefore is an HHS with unbounded products.

Acknowledgments

The authors thank Mark Hagen and Alessandro Sisto for lively conversations about hierarchical hyperbolicity. The second author thanks Chris Leininger for an interesting conversation which led to the formulation of Question C. The authors thank Daniel Berlyne, Ivan Levcovitz, Jacob Russell, Davide Spriano, and the anonymous referee for helpful feedback. The authors thank Anthony Genevois for resolving a question asked in an early version of this article.

Table of Contents

  1. Abstract
  2. 1. Introduction
    1. Largest acylindrical actions
    2. Theorem A (HHG have actions that are largest and universal).
    3. Stability in hierarchically hyperbolic groups
    4. Theorem B (Equivalent conditions for subgroup stability).
    5. On purely loxodromic subgroups
    6. Question C.
    7. New hierarchically hyperbolic structures
    8. Theorem D (Characterization of contracting quasigeodesics).
    9. Theorem E (Random subgroups are stable).
  3. 2. Background
    1. 2.1. Hierarchically hyperbolic spaces
    2. Definition 2.1 (Hierarchically hyperbolic space).
    3. Theorem 2.5 (Distance formula for HHS; BHS19).
    4. Definition 2.6 (Consistent tuple).
    5. Definition 2.7 (Nested partial tuple ()).
    6. Definition 2.8 (Orthogonal partial tuple () ).
    7. Definition 2.9 (Product regions in ).
    8. Lemma 2.10 (Existence of coarse gates; BHS19, Lemma 5.5).
    9. Definition 2.12.
    10. Proposition 2.13 (BHS19, Proposition 5.17).
    11. Definition 2.15 (Hierarchically hyperbolic groups).
    12. 2.2. Acylindrical actions
    13. Definition 2.16 (Acylindrical).
    14. Definition 2.17 (Generalized loxodromic).
    15. Definition 2.19 (Universal acylindrical action).
    16. Definition 2.20 (Largest).
    17. 2.3. Stability
    18. Definition 2.22 (Morse/stable quasigeodesic).
    19. Definition 2.23 (Stable embedding).
    20. Definition 2.24 (Subgroup stability).
  4. 3. Altering the hierarchically hyperbolic structure
    1. 3.1. Unbounded products
    2. Lemma 3.1.
    3. Definition 3.2 (Bounded domain dichotomy).
    4. Definition 3.3 (Unbounded products).
    5. 3.2. Almost hierarchically hyperbolic spaces
    6. Definition 3.4 (Almost HHS).
    7. 3.3. A new hierarchically hyperbolic structure
    8. Lemma 3.6.
    9. Theorem 3.7.
    10. Corollary 3.8.
  5. 4. Characterization of contracting geodesics
    1. Definition 4.1 (Bounded projections).
    2. Definition 4.2 (Contracting).
    3. Lemma 4.3.
    4. Theorem 4.4.
  6. 5. Universal and largest acylindrical actions
    1. Theorem 5.1.
    2. Proposition 5.2 (ABO19).
    3. Corollary 5.5.
    4. Corollary 5.6.
    5. Example 5.7.
  7. 6. Characterizing stability
    1. 6.1. Stability
    2. Corollary 6.1.
    3. Corollary 6.2.
    4. 6.2. The Morse boundary
    5. Theorem 6.3 (Theorem 3.4 and Proposition 5.8 in DHS17).
    6. Definition 6.4.
    7. Lemma 6.5.
    8. 6.3. Bounds on stable asymptotic dimension
    9. Theorem 6.6.
    10. Corollary 6.7.
    11. 6.4. Random subgroups
    12. Theorem 6.8.
    13. Corollary 6.9 (Maher–Sisto; MS19).
  8. 7. Clean containers
    1. Definition 7.1 (Clean containers).
    2. Proposition 7.2.
    3. Proposition 7.3.
    4. Proposition 7.4.
    5. Proposition 7.5.
    6. Proposition 7.6.
  9. Appendix A. Almost HHSs are HHSs By Daniel Berlyne and Jacob Russell
    1. Theorem A.1.
    2. Lemma A.2 (DHS17, Lemma 1.5).
    3. Theorem A.3 (The realization of consistent tuples, BHS19, Theorem 3.1).
    4. Lemma A.4 (–consistency, BHS19, Proposition 1.8).
    5. Corollary A.8.
  10. Acknowledgments

Mathematical Fragments

Theorem A (HHG have actions that are largest and universal).

Every hierarchically hyperbolic group admits a largest acylindrical action. In particular, the following admit acylindrical actions which are largest and universal:

(1)

Hyperbolic groups.

(2)

Mapping class groups of orientable surfaces of finite type.

(3)

Fundamental groups of compact three-manifolds with no Nil or Sol component in their prime decomposition.

(4)

Groups that act properly and cocompactly on a special CAT(0) cube complex, and more generally any cubical group which admits a factor system. This includes right-angled Artin groups, right-angled Coxeter groups, and many other examples as in Reference HS16.

Theorem B (Equivalent conditions for subgroup stability).

Any hierarchically hyperbolic group admits a hierarchically hyperbolic group structure such that for any finitely generated , the following are equivalent:

(1)

is stable in ;

(2)

is undistorted in and has uniformly bounded projections;

(3)

Any orbit map is a quasi-isometric embedding, where is the –maximal element in .

Question C.

Let be a hierarchically hyperbolic group which admits a universal acylindrical action on , where is the –maximal element in . Let be a finitely generated subgroup of .

Are the conditions in Theorem B also equivalent to: is undistorted and acts purely loxodromically on ?

Under what hypotheses on , are the conditions in Theorem B also equivalent to: acts purely loxodromically on ?

Theorem D (Characterization of contracting quasigeodesics).

Let be a hierarchically hyperbolic group. For any there exist uniform constants depending only on and such that the following holds for every –quasigeodesic : the quasigeodesic is uniformly contracting if and only if has uniformly bounded projections (in any structure with unbounded products, e.g., in one as provided by Corollary 3.8).

Theorem E (Random subgroups are stable).

Let be an HHS for which has infinite diameter, where is the –maximal element, and consider which acts properly and cocompactly on . Then any –generated random subgroup of stably embeds in via the orbit map.

Definition 2.1 (Hierarchically hyperbolic space).

A –quasigeodesic space is said to be hierarchically hyperbolic if there exists , an index set , and a set of –hyperbolic spaces , such that the following conditions are satisfied:

(1)

(Projections.) There is a set of projections sending points in to sets of diameter bounded by some in the various . Moreover, there exists so that each is –coarsely Lipschitz and is –quasiconvex in .

(2)

(Nesting.) is equipped with a partial order , and either or contains a unique –maximal element which is larger than all other elements; when , we say is nested in . For each , we denote by the set of such that . Moreover, for all with there is a specified subset with . There is also a projection .

(3)

(Orthogonality.) has a symmetric and anti-reflexive relation called orthogonality: we write when are orthogonal. Also, whenever and , we require that . Finally, we require that for each and each for which , there exists , so that whenever and , we have ; we say is a container associated with and . Finally, if , then are not –comparable.

(4)

(Transversality and consistency.) If are not orthogonal and neither is nested in the other, then we say are transverse, denoted . There exists such that if , then there are sets and each of diameter at most and satisfying:

for all .

For satisfying and for all , we have:

Finally, if , then whenever satisfies either or and .

(5)

(Finite complexity.) There exists , the complexity of (with respect to ), so that any set of pairwise––comparable elements has cardinality at most .

(6)

(Large links.) There exist and such that the following holds. Let and let . Let . Then there exists such that for all , either for some , or . Also, for each .

(7)

(Bounded geodesic image.) For all , all , and all geodesics of , either or .

(8)

(Partial realization.) There exists a constant with the following property. Let be a family of pairwise orthogonal elements of , and let . Then there exists so that:

for all ,

for each and each with , we have , and

if for some , then .

(9)

(Uniqueness.) For each , there exists such that if and , then there exists such that .

Theorem 2.5 (Distance formula for HHS; Reference BHS19).

Let be a hierarchically hyperbolic space. Then there exists such that for all , there exist so that for all ,

Definition 2.6 (Consistent tuple).

Fix , and let be a tuple such that for each , the coordinate is a subset of with . The tuple is –admissible if for all . The –admissible tuple is –consistent if, whenever ,

and whenever ,

Lemma 2.10 (Existence of coarse gates; Reference BHS19, Lemma 5.5).

If is –hierarchically quasiconvex and non-empty, then there exists a gate map for , i.e., for each there exists such that for all , the set (uniformly) coarsely coincides with the projection of to the –quasiconvex set . The point is called the gate of in .

Remark 2.11 (Surjectivity of projections).

As one can always change the hierarchical structure so that the projection maps are coarsely surjective Reference BHS19, Remark 1.3, we will assume that is such a structure. That is, for each , if is not surjective, then we identify with .

Proposition 2.13 (Reference BHS19, Proposition 5.17).

There exists such that for all , all with relevant for , and all –hierarchy paths joining to , there is a subpath of with the following properties:

(1)

;

(2)

is coarsely constant on for all , i.e., it is a uniformly bounded distance from a constant map.

Remark 2.14.

Let , and suppose is relevant for . As and consist of –consistent tuples (for a fixed ) and is only coarsely well-defined, by appropriately increasing to accomodate for the chosen constant in Proposition 2.13, we may assume that is actually a subset of .

Definition 2.15 (Hierarchically hyperbolic groups).

Let be a hierarchically hyperbolic space. An automorphism of consists of a map , together with a bijection and, for each , an isometry so that the following diagrams commute up to uniformly bounded error whenever the maps in question are defined (i.e., when , are not orthogonal):

and

Two automorphisms , are equivalent if and for all we have . The set of all such equivalence classes forms the automorphism group of , denoted . A finitely generated group is said to be a hierarchically hyperbolic group (HHG) if there is a hierarchically hyperbolic space and a group homomorphism so that the induced uniform quasi-action of on is metrically proper, cobounded, and contains finitely many –orbits. Note that when is a hyperbolic group then, with respect to any word metric, it inherits a hierarchically hyperbolic structure.

Remark 2.18.

By Reference Osi16, Theorem 1.1, every acylindrical action of a group on a hyperbolic space either has bounded orbits or contains a loxodromic element. By Reference Osi16, Theorem 1.4.(L4) and Sisto Reference Sis16, Theorem 1, every generalized loxodromic element is Morse, i.e., every quasi-geodesic with endpoints on the axis of the element lies uniformly close to that axis (see Definition 2.22). Therefore, if a group does not contain any Morse elements, it does not contain any generalized loxodromics, and thus must have bounded orbits in every acylindrical action on a hyperbolic space. This is the case when, for example, is a non-trivial direct product, that is, a direct product of two infinite groups.

Definition 2.22 (Morse/stable quasigeodesic).

Let be a metric space. A quasigeodesic is called Morse (or stable) if there exists a function such that if is a –quasigeodesic in with endpoints on , then

We call the stability gauge for and say is –stable if we want to record the constants.

Lemma 3.1.

For any , the set is closed under nesting.

Definition 3.4 (Almost HHS).

If satisfies all axioms of a hierarchically hyperbolic space except (3) and additionally satisfies axiom , then is an almost hierarchically hyperbolic space.

Lemma 3.6.

Let be a hierarchically hyperbolic space and consider which is closed under nesting. Let be a hierarchy path in . Then, the path obtained by including is an unparametrized quasi-geodesic. Moreover, if for each which is a relevant domain for and for each , we modify the path through by removing all but the first and last vertex of the hierarchy path which passes through , then the new path obtained, is a hierarchy path for .

Theorem 3.7.

Every hierarchically hyperbolic space with the bounded domain dichotomy admits a hierarchically hyperbolic structure with unbounded products.

Corollary 3.8.

Every hierarchically hyperbolic group admits a hierarchically hyperbolic group structure with unbounded products.

Definition 4.2 (Contracting).

A subset in a metric space is said to be contracting if there exist a map and constants satisfying:

(1)

For any , we have ;

(2)

If with , then ;

(3)

For all , if we set , then .

Lemma 4.3.

If is a –contracting –quasigeodesic in a metric space , then is –stable for some depending only on and .

Theorem 4.4.

Let be a hierarchically hyperbolic space. For any and there exists a depending only on and such that the following holds for every –quasigeodesic . If has –bounded projections, then is –contracting. Moreover, if has the bounded domain dichotomy, then admits a hierarchically hyperbolic structure with unbounded products where, additionally, we have that if is –contracting, then has –bounded projections.

Theorem 5.1.

Every hierarchically hyperbolic group admits a largest acylindrical action.

Proposition 5.2 (Reference ABO19).

Let be a group, a finite collection of subgroups of , and be a finite subset of such that generates . Assume that:

(1)

is hyperbolic and the action of on it is acylindrical.

(2)

Each is elliptic in every acylindrical action of on a hyperbolic space.

Then is the largest element in .

Corollary 5.5.

Let be a hierarchically hyperbolic group. An element is generalized loxodromic if and only if is contracting.

Corollary 6.1.

Suppose that has unbounded products, is a hyperbolic metric space, and is a –quasi-isometric embedding. Then is –stable if and only if is –contracting, where and determine each other.

Corollary 6.2.

Let be a quasi-isometric embedding from a metric space into a hierarchically hyperbolic space with unbounded products. The following are equivalent:

(1)

is a stable embedding;

(2)

has uniformly bounded projections;

(3)

is a quasi-isometric embedding.

Lemma 6.5.

The inclusion holds for any with unbounded products where is the –maximal element of . Moreover, if is also proper, then for any there exists depending only on and such that if is a sequence with such that has –bounded projections for some and each , then has –bounded projections.

Theorem 6.6.

Let be a hierarchically hyperbolic space with unbounded products such that has finite asymptotic dimension, where is the –maximal element of . Then . Moreover, if is also proper and geodesic, then there exists a continuous bijection .

Theorem 6.8.

Let be an HHS for which the –maximal element, , has infinite diameter, and consider which acts properly and cocompactly on  via the orbit map. Then any –generated random subgroup of stably embeds in .

Proposition 7.2.

The following spaces admit hierarchically hyperbolic structures with clean containers.

(1)

Hyperbolic groups.

(2)

Mapping class groups of orientable surfaces of finite type.

(3)

Special cubical groups, and more generally, any cubical group which admits a factor system.

(4)

, for a compact –manifold with no Nil or Sol in its prime decomposition.

Proposition 7.3.

The product of two hierarchically hyperbolic spaces which both have clean containers has clean containers.

Proposition 7.4.

If is hyperbolic relative to a collection of hierarchically hyperbolic spaces which all have clean containers, then is a hierarchically hyperbolic space with clean containers.

Proposition 7.5.

Let be a tree of hierarchically hyperbolic spaces satisfying the hypotheses of Reference BHS19, Theorem 8.6, so that is hierarchically hyperbolic. If for each , the hierarchically hyperbolic space has clean containers, then so does .

Theorem A.1.

Let be an almost HHS. There exists an HHS structure for so that , and if then the associated hyperbolic space for is a single point.

Lemma A.2 (Reference DHS17, Lemma 1.5).

Let be an almost HHS. If with , and with and both defined, then where is the constant from the consistency axiom of .

Theorem A.3 (The realization of consistent tuples, Reference BHS19, Theorem 3.1).

Let be an almost HHS. There exists a function so that if is a –consistent tuple, then there exists so that for all .

Lemma A.4 (–consistency, Reference BHS19, Proposition 1.8).

Let be an almost HHS structure for and . Suppose or and or . Then we have the following, where is the constant from the consistency axiom of .

(1)

If , then .

(2)

If , then .

Claim A.5.

The length of is bounded above by .

Case 2 (, , and ).

In this case we have either or .

Assume and . Since is transverse to we cannot have for any (this would imply ). If for all , then (as shown in the proof of transversality) and . Since , we have and consistency in implies . If instead there exists so that or , then . Since , and Lemma A.2 gives .

Assume and . As before, for all . First assume there exists so that or . This occurs when either or and not every element of is orthogonal to . In both cases, and consistency in implies because . Now assume for all . This can only occur when is transverse to . In this case, and . Since , consistency in implies .

Assume and . As before, for all . If , then we have the first case of transversality, that is, and for all . Thus, if , then the result reduces to the previous bullet, replacing with . We can therefore assume , meaning we have the second case of transversality where there exists so that is either transverse to or properly nested into .

Suppose too. This implies there also exists so that is either transverse to or properly nested into . Furthermore, and . Now, implies or is nested into an element of . If , then and Lemma A.2 implies . If is nested into an element of , then either or since is a pairwise orthogonal subset of . By applying consistency in when or Lemma A.2 when , we have .

Now suppose . Then implies or is nested into . Applying Lemma A.2 if , or consistency in if , we again obtain .

Uniqueness, bounded geodesic image, large links: Since the only elements of whose associated spaces are not points are in , these axioms for follow from the fact that they hold in .

Partial realization: Let , …, be pairwise orthogonal elements of , and let for each . Without loss of generality, assume , …, and , …, where . If (resp. ), then each (resp. ).

For , let and let be any point in . Since , …, are pairwise orthogonal, it follows that , …, are pairwise orthogonal too, and for each , is either nested into an element of some or orthogonal to all , …, . Without loss of generality, assume that , …, are nested into elements of and , …, , , …, are pairwise orthogonal, where , , and . If , then and each is orthogonal to every . Otherwise, for each , is nested in some for . In both cases, , …, , , …, are pairwise orthogonal elements of . We can therefore use the partial realization axiom in on the points , …, , , …, to produce a point with the following properties:

(1)

for ;

(2)

for ;

(3)

for all if or , then ;

(4)

for all if or , then .

Now, for , define as follows. Let and . If , then define to be any point in . Since is a collection of pairwise orthogonal elements of , the diameter of is at most by Lemma A.2. If either for some or for all then define . Since is a collection of pairwise orthogonal elements of , these two cases encompass all elements of .

Remark A.7.

We say is an almost HHG if there exists an almost HHS such that and satisfy the definition of a hierarchically hyperbolic group where ‘HHS’ is replaced with ‘almost HHS’. The above proof shows that if is an almost HHG, then the structure from Theorem A.1 is an HHG structure for .

Corollary A.8.

Let be an almost HHS with the bounded domain dichotomy. If for every non––maximal domain , there exist so that , , and , then the HHS structure obtained by applying Theorem A.1 to has unbounded products.

References

Reference [Abb16]
Carolyn R. Abbott, Not all finitely generated groups have universal acylindrical actions, Proc. Amer. Math. Soc. 144 (2016), no. 10, 4151–4155, DOI 10.1090/proc/13101. MR3531168,
Show rawAMSref \bib{Abbott:notuniversal}{article}{ label={Abb16}, author={Abbott, Carolyn R.}, title={Not all finitely generated groups have universal acylindrical actions}, journal={Proc. Amer. Math. Soc.}, volume={144}, date={2016}, number={10}, pages={4151--4155}, issn={0002-9939}, review={\MR {3531168}}, doi={10.1090/proc/13101}, }
Reference [ABO19]
Carolyn Abbott, Sahana H. Balasubramanya, and Denis Osin, Hyperbolic structures on groups, Algebr. Geom. Topol. 19 (2019), no. 4, 1747–1835, DOI 10.2140/agt.2019.19.1747. MR3995018,
Show rawAMSref \bib{AbbottBalasubramanyaOsin:extensions}{article}{ label={ABO19}, author={Abbott, Carolyn}, author={Balasubramanya, Sahana H.}, author={Osin, Denis}, title={Hyperbolic structures on groups}, journal={Algebr. Geom. Topol.}, volume={19}, date={2019}, number={4}, pages={1747--1835}, issn={1472-2747}, review={\MR {3995018}}, doi={10.2140/agt.2019.19.1747}, }
Reference [ACGH17]
Goulnara N. Arzhantseva, Christopher H. Cashen, Dominik Gruber, and David Hume, Characterizations of Morse quasi-geodesics via superlinear divergence and sublinear contraction, Doc. Math. 22 (2017), 1193–1224. MR3690269,
Show rawAMSref \bib{ACGH}{article}{ label={ACGH17}, author={Arzhantseva, Goulnara N.}, author={Cashen, Christopher H.}, author={Gruber, Dominik}, author={Hume, David}, title={Characterizations of Morse quasi-geodesics via superlinear divergence and sublinear contraction}, journal={Doc. Math.}, volume={22}, date={2017}, pages={1193--1224}, issn={1431-0635}, review={\MR {3690269}}, }
Reference [ADT17]
Tarik Aougab, Matthew Gentry Durham, and Samuel J. Taylor, Pulling back stability with applications to and relatively hyperbolic groups, J. Lond. Math. Soc. (2) 96 (2017), no. 3, 565–583, DOI 10.1112/jlms.12071. MR3742433,
Show rawAMSref \bib{aougab2016middle}{article}{ label={ADT17}, author={Aougab, Tarik}, author={Durham, Matthew Gentry}, author={Taylor, Samuel J.}, title={Pulling back stability with applications to ${\mathrm {Out}}(F_n)$ and relatively hyperbolic groups}, journal={J. Lond. Math. Soc. (2)}, volume={96}, date={2017}, number={3}, pages={565--583}, issn={0024-6107}, review={\MR {3742433}}, doi={10.1112/jlms.12071}, }
Reference [BBKL20]
Mladen Bestvina, Kenneth Bromberg, Autumn E. Kent, and Christopher J. Leininger, Undistorted purely pseudo-Anosov groups, J. Reine Angew. Math. 760 (2020), 213–227, DOI 10.1515/crelle-2018-0013. MR4069890,
Show rawAMSref \bib{BBKL}{article}{ label={BBKL20}, author={Bestvina, Mladen}, author={Bromberg, Kenneth}, author={Kent, Autumn E.}, author={Leininger, Christopher J.}, title={Undistorted purely pseudo-Anosov groups}, journal={J. Reine Angew. Math.}, volume={760}, date={2020}, pages={213--227}, issn={0075-4102}, review={\MR {4069890}}, doi={10.1515/crelle-2018-0013}, }
Reference [Beh06]
Jason A. Behrstock, Asymptotic geometry of the mapping class group and Teichmüller space, Geom. Topol. 10 (2006), 1523–1578, DOI 10.2140/gt.2006.10.1523. MR2255505,
Show rawAMSref \bib{Behrstock:asymptotic}{article}{ label={Beh06}, author={Behrstock, Jason A.}, title={Asymptotic geometry of the mapping class group and Teichm\"{u}ller space}, journal={Geom. Topol.}, volume={10}, date={2006}, pages={1523--1578}, issn={1465-3060}, review={\MR {2255505}}, doi={10.2140/gt.2006.10.1523}, }
Reference [BHS17a]
Jason Behrstock, Mark F. Hagen, and Alessandro Sisto, Asymptotic dimension and small-cancellation for hierarchically hyperbolic spaces and groups, Proc. Lond. Math. Soc. (3) 114 (2017), no. 5, 890–926, DOI 10.1112/plms.12026. MR3653249,
Show rawAMSref \bib{BehrstockHagenSisto:asdim}{article}{ label={BHS17a}, author={Behrstock, Jason}, author={Hagen, Mark F.}, author={Sisto, Alessandro}, title={Asymptotic dimension and small-cancellation for hierarchically hyperbolic spaces and groups}, journal={Proc. Lond. Math. Soc. (3)}, volume={114}, date={2017}, number={5}, pages={890--926}, issn={0024-6115}, review={\MR {3653249}}, doi={10.1112/plms.12026}, }
Reference [BHS17b]
Jason Behrstock, Mark F. Hagen, and Alessandro Sisto, Hierarchically hyperbolic spaces, I: Curve complexes for cubical groups, Geom. Topol. 21 (2017), no. 3, 1731–1804, DOI 10.2140/gt.2017.21.1731. MR3650081,
Show rawAMSref \bib{BehrstockHagenSisto:HHS_I}{article}{ label={BHS17b}, author={Behrstock, Jason}, author={Hagen, Mark F.}, author={Sisto, Alessandro}, title={Hierarchically hyperbolic spaces, I: Curve complexes for cubical groups}, journal={Geom. Topol.}, volume={21}, date={2017}, number={3}, pages={1731--1804}, issn={1465-3060}, review={\MR {3650081}}, doi={10.2140/gt.2017.21.1731}, }
Reference [BHS17c]
Jason Behrstock, Mark F Hagen, and Alessandro Sisto. Quasiflats in hierarchically hyperbolic spaces. arXiv:1704.04271, 2017. To appear in Duke Mathematical Journal.
Reference [BHS19]
Jason Behrstock, Mark Hagen, and Alessandro Sisto, Hierarchically hyperbolic spaces II: Combination theorems and the distance formula, Pacific J. Math. 299 (2019), no. 2, 257–338, DOI 10.2140/pjm.2019.299.257. MR3956144,
Show rawAMSref \bib{BehrstockHagenSisto:HHS_II}{article}{ label={BHS19}, author={Behrstock, Jason}, author={Hagen, Mark}, author={Sisto, Alessandro}, title={Hierarchically hyperbolic spaces II: Combination theorems and the distance formula}, journal={Pacific J. Math.}, volume={299}, date={2019}, number={2}, pages={257--338}, issn={0030-8730}, review={\MR {3956144}}, doi={10.2140/pjm.2019.299.257}, }
Reference [Bow08]
Brian H. Bowditch, Tight geodesics in the curve complex, Invent. Math. 171 (2008), no. 2, 281–300, DOI 10.1007/s00222-007-0081-y. MR2367021,
Show rawAMSref \bib{Bowditch:tight}{article}{ label={Bow08}, author={Bowditch, Brian H.}, title={Tight geodesics in the curve complex}, journal={Invent. Math.}, volume={171}, date={2008}, number={2}, pages={281--300}, issn={0020-9910}, review={\MR {2367021}}, doi={10.1007/s00222-007-0081-y}, }
Reference [BR]
Federico Berlai and Bruno Robbio, A refined combination theorem for hierarchically hyperbolic groups, Groups Geom. Dyn. 14 (2020), no. 4, 1127–1203, DOI 10.4171/ggd/576. MR4186470,
Show rawAMSref \bib{BerlaiRobbio:combinationHHG}{article}{ label={BR}, author={Berlai, Federico}, author={Robbio, Bruno}, title={A refined combination theorem for hierarchically hyperbolic groups}, journal={Groups Geom. Dyn.}, volume={14}, date={2020}, number={4}, pages={1127--1203}, issn={1661-7207}, review={\MR {4186470}}, doi={10.4171/ggd/576}, }
Reference [Bra99]
Noel Brady, Branched coverings of cubical complexes and subgroups of hyperbolic groups, J. London Math. Soc. (2) 60 (1999), no. 2, 461–480, DOI 10.1112/S0024610799007644. MR1724853,
Show rawAMSref \bib{Brady:branchedcoverings}{article}{ label={Bra99}, author={Brady, Noel}, title={Branched coverings of cubical complexes and subgroups of hyperbolic groups}, journal={J. London Math. Soc. (2)}, volume={60}, date={1999}, number={2}, pages={461--480}, issn={0024-6107}, review={\MR {1724853}}, doi={10.1112/S0024610799007644}, }
Reference [CD19]
Matthew Cordes and Matthew Gentry Durham, Boundary convex cocompactness and stability of subgroups of finitely generated groups, Int. Math. Res. Not. IMRN 6 (2019), 1699–1724, DOI 10.1093/imrn/rnx166. MR3932592,
Show rawAMSref \bib{cordes2016boundary}{article}{ label={CD19}, author={Cordes, Matthew}, author={Durham, Matthew Gentry}, title={Boundary convex cocompactness and stability of subgroups of finitely generated groups}, journal={Int. Math. Res. Not. IMRN}, date={2019}, number={6}, pages={1699--1724}, issn={1073-7928}, review={\MR {3932592}}, doi={10.1093/imrn/rnx166}, }
Reference [CH17]
Matthew Cordes and David Hume, Stability and the Morse boundary, J. Lond. Math. Soc. (2) 95 (2017), no. 3, 963–988, DOI 10.1112/jlms.12042. MR3664526,
Show rawAMSref \bib{cordes2016stability}{article}{ label={CH17}, author={Cordes, Matthew}, author={Hume, David}, title={Stability and the Morse boundary}, journal={J. Lond. Math. Soc. (2)}, volume={95}, date={2017}, number={3}, pages={963--988}, issn={0024-6107}, review={\MR {3664526}}, doi={10.1112/jlms.12042}, }
Reference [CM19]
Christopher H. Cashen and John M. Mackay, A metrizable topology on the contracting boundary of a group, Trans. Amer. Math. Soc. 372 (2019), no. 3, 1555–1600, DOI 10.1090/tran/7544. MR3976570,
Show rawAMSref \bib{cashen2017metrizable}{article}{ label={CM19}, author={Cashen, Christopher H.}, author={Mackay, John M.}, title={A metrizable topology on the contracting boundary of a group}, journal={Trans. Amer. Math. Soc.}, volume={372}, date={2019}, number={3}, pages={1555--1600}, issn={0002-9947}, review={\MR {3976570}}, doi={10.1090/tran/7544}, }
Reference [Cor17]
Matthew Cordes, Morse boundaries of proper geodesic metric spaces, Groups Geom. Dyn. 11 (2017), no. 4, 1281–1306, DOI 10.4171/GGD/429. MR3737283,
Show rawAMSref \bib{cordes2015morse}{article}{ label={Cor17}, author={Cordes, Matthew}, title={Morse boundaries of proper geodesic metric spaces}, journal={Groups Geom. Dyn.}, volume={11}, date={2017}, number={4}, pages={1281--1306}, issn={1661-7207}, review={\MR {3737283}}, doi={10.4171/GGD/429}, }
Reference [CS15]
Ruth Charney and Harold Sultan, Contracting boundaries of spaces, J. Topol. 8 (2015), no. 1, 93–117. MR3339446,
Show rawAMSref \bib{charneysultan}{article}{ label={CS15}, author={Charney, Ruth}, author={Sultan, Harold}, title={Contracting boundaries of $\mathrm {CAT}(0)$ spaces}, journal={J. Topol.}, volume={8}, date={2015}, number={1}, pages={93--117}, issn={1753-8416}, review={\MR {3339446}}, }
Reference [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 (2017), no. 1156, v+152, DOI 10.1090/memo/1156. MR3589159,
Show rawAMSref \bib{DGO}{article}{ label={DGO17}, author={Dahmani, F.}, author={Guirardel, V.}, author={Osin, D.}, title={Hyperbolically embedded subgroups and rotating families in groups acting on hyperbolic spaces}, journal={Mem. Amer. Math. Soc.}, volume={245}, date={2017}, number={1156}, pages={v+152}, issn={0065-9266}, isbn={978-1-4704-2194-6}, isbn={978-1-4704-3601-8}, review={\MR {3589159}}, doi={10.1090/memo/1156}, }
Reference [DHS17]
Matthew Gentry Durham, Mark F. Hagen, and Alessandro Sisto, Boundaries and automorphisms of hierarchically hyperbolic spaces, Geom. Topol. 21 (2017), no. 6, 3659–3758, DOI 10.2140/gt.2017.21.3659. MR3693574,
Show rawAMSref \bib{DurhamHagenSisto:HHS_boundary}{article}{ label={DHS17}, author={Durham, Matthew Gentry}, author={Hagen, Mark F.}, author={Sisto, Alessandro}, title={Boundaries and automorphisms of hierarchically hyperbolic spaces}, journal={Geom. Topol.}, volume={21}, date={2017}, number={6}, pages={3659--3758}, issn={1465-3060}, review={\MR {3693574}}, doi={10.2140/gt.2017.21.3659}, }
Reference [DMS10]
Cornelia Druţu, Shahar Mozes, and Mark Sapir, Divergence in lattices in semisimple Lie groups and graphs of groups, Trans. Amer. Math. Soc. 362 (2010), no. 5, 2451–2505, DOI 10.1090/S0002-9947-09-04882-X. MR2584607,
Show rawAMSref \bib{DrutuMozesSapir:Divergence}{article}{ label={DMS10}, author={Dru\c {t}u, Cornelia}, author={Mozes, Shahar}, author={Sapir, Mark}, title={Divergence in lattices in semisimple Lie groups and graphs of groups}, journal={Trans. Amer. Math. Soc.}, volume={362}, date={2010}, number={5}, pages={2451--2505}, issn={0002-9947}, review={\MR {2584607}}, doi={10.1090/S0002-9947-09-04882-X}, }
Reference [DT15]
Matthew Gentry Durham and Samuel J. Taylor, Convex cocompactness and stability in mapping class groups, Algebr. Geom. Topol. 15 (2015), no. 5, 2839–2859, DOI 10.2140/agt.2015.15.2839. MR3426695,
Show rawAMSref \bib{DurhamTaylor:stable}{article}{ label={DT15}, author={Durham, Matthew Gentry}, author={Taylor, Samuel J.}, title={Convex cocompactness and stability in mapping class groups}, journal={Algebr. Geom. Topol.}, volume={15}, date={2015}, number={5}, pages={2839--2859}, issn={1472-2747}, review={\MR {3426695}}, doi={10.2140/agt.2015.15.2839}, }
Reference [FM02]
Benson Farb and Lee Mosher, Convex cocompact subgroups of mapping class groups, Geom. Topol. 6 (2002), 91–152, DOI 10.2140/gt.2002.6.91. MR1914566,
Show rawAMSref \bib{FarbMosher}{article}{ label={FM02}, author={Farb, Benson}, author={Mosher, Lee}, title={Convex cocompact subgroups of mapping class groups}, journal={Geom. Topol.}, volume={6}, date={2002}, pages={91--152}, issn={1465-3060}, review={\MR {1914566}}, doi={10.2140/gt.2002.6.91}, }
Reference [Hag14]
Mark F. Hagen, Weak hyperbolicity of cube complexes and quasi-arboreal groups, J. Topol. 7 (2014), no. 2, 385–418, DOI 10.1112/jtopol/jtt027. MR3217625,
Show rawAMSref \bib{Hagen:quasi_arboreal}{article}{ label={Hag14}, author={Hagen, Mark F.}, title={Weak hyperbolicity of cube complexes and quasi-arboreal groups}, journal={J. Topol.}, volume={7}, date={2014}, number={2}, pages={385--418}, issn={1753-8416}, review={\MR {3217625}}, doi={10.1112/jtopol/jtt027}, }
Reference [Ham]
U. Hamenstädt, Word hyperbolic extensions of surface groups, Preprint, arXiv:math/0505244.
Reference [HS16]
Mark F. Hagen and Tim Susse, On hierarchical hyperbolicity of cubical groups, Israel J. Math. 236 (2020), no. 1, 45–89, DOI 10.1007/s11856-020-1967-2. MR4093881,
Show rawAMSref \bib{HagenSusse:HHScubical}{article}{ label={HS16}, author={Hagen, Mark F.}, author={Susse, Tim}, title={On hierarchical hyperbolicity of cubical groups}, journal={Israel J. Math.}, volume={236}, date={2020}, number={1}, pages={45--89}, issn={0021-2172}, review={\MR {4093881}}, doi={10.1007/s11856-020-1967-2}, }
Reference [KK14]
Sang-Hyun Kim and Thomas Koberda, The geometry of the curve graph of a right-angled Artin group, Internat. J. Algebra Comput. 24 (2014), no. 2, 121–169, DOI 10.1142/S021819671450009X. MR3192368,
Show rawAMSref \bib{KimKoberda:curve_graph}{article}{ label={KK14}, author={Kim, Sang-Hyun}, author={Koberda, Thomas}, title={The geometry of the curve graph of a right-angled Artin group}, journal={Internat. J. Algebra Comput.}, volume={24}, date={2014}, number={2}, pages={121--169}, issn={0218-1967}, review={\MR {3192368}}, doi={10.1142/S021819671450009X}, }
Reference [KL08]
Richard P. Kent IV and Christopher J. Leininger, Shadows of mapping class groups: capturing convex cocompactness, Geom. Funct. Anal. 18 (2008), no. 4, 1270–1325, DOI 10.1007/s00039-008-0680-9. MR2465691,
Show rawAMSref \bib{KentLeininger:shadows}{article}{ label={KL08}, author={Kent, Richard P., IV}, author={Leininger, Christopher J.}, title={Shadows of mapping class groups: capturing convex cocompactness}, journal={Geom. Funct. Anal.}, volume={18}, date={2008}, number={4}, pages={1270--1325}, issn={1016-443X}, review={\MR {2465691}}, doi={10.1007/s00039-008-0680-9}, }
Reference [KMT17]
Thomas Koberda, Johanna Mangahas, and Samuel J. Taylor, The geometry of purely loxodromic subgroups of right-angled Artin groups, Trans. Amer. Math. Soc. 369 (2017), no. 11, 8179–8208, DOI 10.1090/tran/6933. MR3695858,
Show rawAMSref \bib{koberda2014geometry}{article}{ label={KMT17}, author={Koberda, Thomas}, author={Mangahas, Johanna}, author={Taylor, Samuel J.}, title={The geometry of purely loxodromic subgroups of right-angled Artin groups}, journal={Trans. Amer. Math. Soc.}, volume={369}, date={2017}, number={11}, pages={8179--8208}, issn={0002-9947}, review={\MR {3695858}}, doi={10.1090/tran/6933}, }
Reference [Min96]
Yair N. Minsky, Quasi-projections in Teichmüller space, J. Reine Angew. Math. 473 (1996), 121–136, DOI 10.1515/crll.1995.473.121. MR1390685,
Show rawAMSref \bib{Minsky:quasiproj}{article}{ label={Min96}, author={Minsky, Yair N.}, title={Quasi-projections in Teichm\"{u}ller space}, journal={J. Reine Angew. Math.}, volume={473}, date={1996}, pages={121--136}, issn={0075-4102}, review={\MR {1390685}}, doi={10.1515/crll.1995.473.121}, }
Reference [MM99]
Howard A. Masur and Yair N. Minsky, Geometry of the complex of curves. I. Hyperbolicity, Invent. Math. 138 (1999), no. 1, 103–149, DOI 10.1007/s002220050343. MR1714338,
Show rawAMSref \bib{MasurMinsky:I}{article}{ label={MM99}, author={Masur, Howard A.}, author={Minsky, Yair N.}, title={Geometry of the complex of curves. I. Hyperbolicity}, journal={Invent. Math.}, volume={138}, date={1999}, number={1}, pages={103--149}, issn={0020-9910}, review={\MR {1714338}}, doi={10.1007/s002220050343}, }
Reference [MS19]
Joseph Maher and Alessandro Sisto, Random subgroups of acylindrically hyperbolic groups and hyperbolic embeddings, Int. Math. Res. Not. IMRN 13 (2019), 3941–3980, DOI 10.1093/imrn/rnx233. MR4023758,
Show rawAMSref \bib{maher2017random}{article}{ label={MS19}, author={Maher, Joseph}, author={Sisto, Alessandro}, title={Random subgroups of acylindrically hyperbolic groups and hyperbolic embeddings}, journal={Int. Math. Res. Not. IMRN}, date={2019}, number={13}, pages={3941--3980}, issn={1073-7928}, review={\MR {4023758}}, doi={10.1093/imrn/rnx233}, }
Reference [Osi16]
D. Osin, Acylindrically hyperbolic groups, Trans. Amer. Math. Soc. 368 (2016), no. 2, 851–888, DOI 10.1090/tran/6343. MR3430352,
Show rawAMSref \bib{Osin:acyl}{article}{ label={Osi16}, author={Osin, D.}, title={Acylindrically hyperbolic groups}, journal={Trans. Amer. Math. Soc.}, volume={368}, date={2016}, number={2}, pages={851--888}, issn={0002-9947}, review={\MR {3430352}}, doi={10.1090/tran/6343}, }
Reference [Rus20]
Jacob Russell, From hierarchical to relative hyperbolicity, Int. Math. Res. Not. IMRN, 2020. To appear.
Reference [Sel97]
Z. Sela, Acylindrical accessibility for groups, Invent. Math. 129 (1997), no. 3, 527–565, DOI 10.1007/s002220050172. MR1465334,
Show rawAMSref \bib{Sela:acylindrical}{article}{ label={Sel97}, author={Sela, Z.}, title={Acylindrical accessibility for groups}, journal={Invent. Math.}, volume={129}, date={1997}, number={3}, pages={527--565}, issn={0020-9910}, review={\MR {1465334}}, doi={10.1007/s002220050172}, }
Reference [Sis16]
Alessandro Sisto, Quasi-convexity of hyperbolically embedded subgroups, Math. Z. 283 (2016), no. 3-4, 649–658, DOI 10.1007/s00209-016-1615-z. MR3519976,
Show rawAMSref \bib{Sisto:qconvex}{article}{ label={Sis16}, author={Sisto, Alessandro}, title={Quasi-convexity of hyperbolically embedded subgroups}, journal={Math. Z.}, volume={283}, date={2016}, number={3-4}, pages={649--658}, issn={0025-5874}, review={\MR {3519976}}, doi={10.1007/s00209-016-1615-z}, }
Reference [Sul14]
Harold Sultan, Hyperbolic quasi-geodesics in CAT(0) spaces, Geom. Dedicata 169 (2014), 209–224, DOI 10.1007/s10711-013-9851-4. MR3175245,
Show rawAMSref \bib{SultanhypquasiCAT0}{article}{ label={Sul14}, author={Sultan, Harold}, title={Hyperbolic quasi-geodesics in CAT(0) spaces}, journal={Geom. Dedicata}, volume={169}, date={2014}, pages={209--224}, issn={0046-5755}, review={\MR {3175245}}, doi={10.1007/s10711-013-9851-4}, }
Reference [Tra]
Hung Cong Tran, Malnormality and join-free subgroups in right-angled Coxeter groups arXiv:1703.09032.
Reference [TT16]
Samuel J. Taylor and Giulio Tiozzo, Random extensions of free groups and surface groups are hyperbolic, Int. Math. Res. Not. IMRN 1 (2016), 294–310, DOI 10.1093/imrn/rnv138. MR3514064,
Show rawAMSref \bib{TaylorTiozzo:randomqi}{article}{ label={TT16}, author={Taylor, Samuel J.}, author={Tiozzo, Giulio}, title={Random extensions of free groups and surface groups are hyperbolic}, journal={Int. Math. Res. Not. IMRN}, date={2016}, number={1}, pages={294--310}, issn={1073-7928}, review={\MR {3514064}}, doi={10.1093/imrn/rnv138}, }

Article Information

MSC 2020
Primary: 20F55 (Reflection and Coxeter groups (group-theoretic aspects)), 20F65 (Geometric group theory), 20F67 (Hyperbolic groups and nonpositively curved groups)
Author Information
Carolyn Abbott
Department of Mathematics, Columbia University, New York, New York
abbott@math.columbia.edu
MathSciNet
Jason Behrstock
Department of Mathematics, Lehman College and The Graduate Center, CUNY, New York, New York
jason.behrstock@lehman.cuny.edu
ORCID
MathSciNet
Matthew Gentry Durham
Department of Mathematics, University of California, Riverside, Riverside, California
matthew.durham@ucr.edu
MathSciNet
Contributor Information
Daniel Berlyne
ORCID
Jacob Russell
ORCID
MathSciNet
Additional Notes

The authors were supported in part by NSF grant DMS-1440140 while at the Mathematical Sciences Research Institute in Berkeley during Fall 2016 program in Geometric Group Theory. The first author was supported by the NSF RTG award DMS-1502553 and NSF award DMS-1803368. The second author was supported by NSF award DMS-1710890, and the third author was supported by NSF RTG award DMS-1045119 and NSF award DMS-1906487.

Journal Information
Transactions of the American Mathematical Society, Series B, Volume 8, Issue 3, ISSN 2330-0000, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , , , and published on .
Copyright Information
Copyright 2021 by the authors under Creative Commons Attribution-Noncommercial 3.0 License (CC BY NC 3.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/btran/50
  • MathSciNet Review: 4215647
  • Show rawAMSref \bib{4215647}{article}{ author={Abbott, Carolyn}, author={Behrstock, Jason}, author={Durham, Matthew}, title={Largest acylindrical actions and Stability in hierarchically hyperbolic groups}, journal={Trans. Amer. Math. Soc. Ser. B}, volume={8}, number={3}, date={2021}, pages={66-104}, issn={2330-0000}, review={4215647}, doi={10.1090/btran/50}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.